dynamic oceanography

107
DYNAMIC OCEANOGRAPHY JAN ERIK WEBER Department of Geosciences Section for Meteorology and Oceanography University of Oslo 09.11.04

Upload: ayanbanerjee1

Post on 07-Mar-2015

914 views

Category:

Documents


5 download

TRANSCRIPT

Page 1: Dynamic Oceanography

DYNAMIC OCEANOGRAPHY

JAN ERIK WEBER

Department of Geosciences

Section for Meteorology and Oceanography

University of Oslo

09.11.04

Page 2: Dynamic Oceanography

2

CONTENTS

1. SHALLOW-WATER THEORY, QUASI-HOMOGENEOUS OCEAN

1.1 Inviscid motion, potential vorticity………………………………………….3

1.2 Linear waves in the absence of rotation……………………………………..8

1.3 Effect of rotation, geostrophic adjustment…………………………………13

1.4 Sverdrup waves, inertial waves and Poincarè waves………………………16

1.5 Kelvin waves at a straight coast……………………………………………23

1.6 Amphidromic systems……………………………………...........................26

1.7 Equatorial Kelvin waves…………………………………………………...29

1.8 Topographically trapped waves……………………………………………31

1.9 Planetary Rossby waves……………………………………………………35

1.10 Topographic Rossby waves………………………………………………..41

1.11 Barotropic instability………………………………………………………42

1.12 Barotropic flow over an ocean ridge………………………………………45

2. WIND-DRIVEN CURRENTS AND OCEAN CIRCULATION

2.1 Equations for the mean motion………………………………......................51

2.2 The Ekman elementary current system……………………………………..57

2.3 The Ekman transport………………………………………………………..61

2.4 Divergent Ekman transport and forced vertical motion(Ekman suction)…..62

2.5 Variable vertical eddy viscosity…………………………………………….65

2.6 Equations for the volume transport…………………………………………68

2.7 The Sverdrup transport……………………………………………………...72

2.8 Theories of Stommel and Munk (western intensification)…………….........75

3. BAROCLINIC MOTION

3.1 Two-layer model………………………………………………………........81

3.2 Continuously stratified fluid………………………………………………..87

3.3 Free internal waves with rotation ………………………………………….88

3.4 Internal response to wind-forcing, upwelling at a straight coast…………...94

3.5 Baroclinic instability………………………………………………............100

Page 3: Dynamic Oceanography

3

1. SHALLOW-WATER THEORY, QUASI-HOMOGENEOUS

OCEAN

1.1 Inviscid motion, potential vorticity

Governing equations

We study motion in an inviscid fluid with densityρ. The fluid is rotating about the

z-axis with constant angular velocity Ωsinϕ, where ϕ is the latitude of our site of

observation. Furthermore, (x, y) are horizontal coordinate axes along the undisturbed

sea surface, and the z-axis is directed upwards. The position of the free surface is

given by ),,( tyxz η= , where η is referred to as the surface elevation. The atmospheric

pressure at the surface is denoted by PS(x, y, t). The bottom topography does not vary

with time, and is given by ),( yxHz −= ; see the sketch in Fig. 1.1.

Fig 1.1. Definition sketch.

The equation of motion for a fluid particle of unit mass can now be written

kgpvkfdt

vD rrrr

−∇−=×+ρ1 , (1.1.1)

where D/dt ≡ ∂/∂t + ∇⋅vr is the total derivative following a fluid particle, and f =

2Ωsinϕ is the Coriolis parameter.

We take that the mass of fluid within a geometrically fixed volume only can

change as a result of advection of fluid particles. The basic assumption of

conservation of mass for a fluid can be stated mathematically as

Page 4: Dynamic Oceanography

4

vdtD r

⋅∇−= ρρ . (1.1.2)

This is often referred to as the continuity equation. In this section we will assume that

the density of a particle is conserved (Dρ/dt = 0). Thus, the continuity equation

reduces to

0=⋅∇ vr . (1.1.3)

Furthermore we will assume that the density is the same for all particles

(homogeneous fluid). We orientate our coordinate system such that the x-axis is

tangential to a latitudinal circle and the y-axis is pointing northwards; see Fig. 1.2.

Fig. 1.2. Orientation of coordinate axes.

In our reference system f is only a function of y. We may then write approximately

that

yfydydfff β+=⎟⎟

⎞⎜⎜⎝

⎛+= 0

00 , (1.1.4)

where

( ) ⎪⎭

⎪⎬⎫

Ω=Ω=

Ω=

.cos2sin21,sin2

0

00

0ϕϕ

ϕβ

ϕ

ϕ Rdd

R

f (1.1.5)

This is called the beta-plane approximation. If f is approximately constant in an (x, y)-

area, we say that the motion occurs on an f-plane.

The kinematic and dynamic boundary conditions at the surface can be written,

respectively, as

Page 5: Dynamic Oceanography

5

,0)( =−ηzdtD ),,( tyxz η= , (1.1.6)

,SPp = ),,( tyxz η= . (1.1.7)

The kinematic boundary condition at the bottom can be expressed as

( ) ,0=+∇⋅ Hzvr ),( yxHz −= . (1.1.8)

We now integrate the continuity equation (1.1.3) from z = −H to z = η(x, y, t):

∫ ∫∫− −−

−−=η ηη

H Hyx

Hz dzvdzudzw . (1.1.9)

Utilizing (1.1.6) and (1.1.8), we find that

∫ ∫− −∂

∂−

∂∂

−=η η

ηH H

t dzvy

dzux

. (1.1.10)

The basic assumption in shallow water theory is that the pressure distribution in the

vertical direction is hydrostatic. By utilizing that the density is constant, and applying

(1.1.7), this leads to

),,()( tyxPzgp S+−= ηρ . (1.1.11)

This means, when we return to the vertical component in (1.1.1), that the vertical

acceleration Dw/dt must be so small that it does not noticeably alter the hydrostatic

pressure distribution. We will return to the validity of this in section 1.2. The

horizontal components of (1.1.1) can thus be written

Sxx PgfvdtDu

ρη 1

−−=− (1.1.12)

SyPgfudtDv

y ρη 1

−−=+ (1.1.13)

Throughout this text we will alternate between writing partial derivatives in full, and

(for economic reasons) as subscripts.

We realize that the right-hand sides of (1.1.12) and (1.1.13) are independent of z.

By utilizing that v ≡ Dy/dt and f = f0 + βy, (1.1.12) can be written

Sxx PgyyfudtD

ρηβ 1

21 2

0 −−=⎟⎠⎞

⎜⎝⎛ −− (1.1.14)

From (1.1.14) it follows that dtyyfuD /)2/( 20 β−− is independent of z. Thus, this is

also true for )2/( 20 yyfu β−− , and thereby also for u, if u and v were independent of

Page 6: Dynamic Oceanography

6

z at time t = 0. Similarly, from (1.1.13) we find that v is independent of z. We can

accordingly write

⎭⎬⎫

==

).,,(),,,(

tyxvvtyxuu

(1.1.15)

Furthermore, it now follows from (1.1.3) that wz is independent of z. Hence, by

integrating in the vertical:

),,()( tyxCzvuw yx ++−= . (1.1.16)

The function C is obtained by applying the boundary condition (1.1.8) at the ocean

bottom. The vertical velocity can thus be written

yxyx vHuHHzvuw −−++−= ))(( . (1.1.17)

Since u and v are independent of z, the integrations in (1.1.10) are easily performed.

We then obtain the following nonlinear, coupled set of equations for the horizontal

velocity components and the surface elevation

Sxxyxt Pgfvvuuuuρ

η 1−−=−++ , (1.1.18)

Syyyxt Pgfuvvuvvρ

η 1−−=+++ , (1.1.19)

0)()( =++++ yxt HvHu ηηη . (1.1.20)

To solve this set of equations we require three initial conditions, e.g. the distribution

of u, v, and η in space at time t = 0. If the fluid is limited by lateral boundaries (walls),

we must in addition ensure that the solutions satisfy the requirements of no flow

through impermeable walls.

Potential vorticity

We define the vertical component of the relative vorticity in our coordinate

system, e.g. Fig. 1.2, by

yx uv −=ζ . (1.1.21)

In addition, every particle in this coordinate system possesses a planetary vorticity f,

arising from solid body rotation with angular velocity ϕsinΩ . Hence, the absolute

vertical vorticity for a particle becomes ζ+f . We shall derive an equation for the

absolute vorticity. It is obtained by differentiating the equations (1.1.18) and (1.1.19)

by − y∂∂ / and x∂∂ / , respectively, and then add the resulting equations.

Page 7: Dynamic Oceanography

7

Mathematically, this means to operate the curl on the vector equation to eliminate the

gradient terms. Since f is independent of time, we find that

))(()( yx vuffdtD

++−=+ ζζ . (1.1.22)

By using that H is independent of time (1.1.20) can be written

))(()( yx vuHHdtD

++−=+ ηη . (1.1.23)

Here, H + η is the height of a vertical fluid column. We define the potential vorticity

Q by

ηζ++

≡HfQ . (1.1.24)

By eliminating the horizontal divergence between (1.1.22) and (1.1.23), we find for Q

that

0=dt

DQ . (1.1.25)

This equation expresses the fact that a given material vertical fluid column always

moves in such a way that its potential vorticity is conserved.

Alternatively, we can apply Kelvin’s circulation theorem for an inviscid fluid to

derive this important result. Kelvin’s theorem states that the circulation of the

absolute velocity around a closed material curve (always consisting of the same fluid

particles) is conserved. For a material curve Γ in the horizontal plane, Kelvin’s and

Stokes’ theorems yield

∫ ∫∫Γ

=×∇⋅=⋅ .const)(σ

δσδ absabs vkrv rrrr , (1.1.26)

where σ is the area inside Γ. Furthermore, in the surface integral:

ζ+=×∇⋅ fvk abs )( rr. (1.1.27)

When the surface area σ in (1.1.26) approaches zero, we have

const.)( =+ δσζf (1.1.28)

In addition, the mass of a vertical fluid column with base δσ must be conserved, and

hence

const. )( =+ δσηρ H (1.1.29)

This is valid for all times, since a vertical fluid column will remain vertical; see

(1.1.15). In our case the fluid is homogeneous and incompressible, i.e. ρ is the same

Page 8: Dynamic Oceanography

8

for all particles. Thus, by eliminating δσ between (1.1.28) and (1.1.29), we find as

before that

const.=++

=ηζ

HfQ , (1.1.30)

or, equivalently, 0/ =dtDQ .

In the ocean we usually have that |ζ| << f and |η| << H. For stationary flow,

assuming that |∇H| >> |∇η| and |f ∇H| >> |H ∇ζ|, (1.1.25) yields approximately that

0)/( =∇⋅ Hfvr . (1.1.31)

On an f-plane, this equation reduces to

0=∇⋅ Hvr . (1.1.32)

Accordingly, the flow in this case follows the lines of constant H (i.e. the bottom

contours). This phenomenon is called topographic steering. On a beta-plane the flow

will follow the contours of the function f/H; see (1.1.31).

Take that the motion is approximately geostrophic. For constant surface pressure,

we then obtain from (1.1.18) and (1.1.19) that

)( η∇×= kfgv

rr . (1.1.33)

For stationary flow the theorem of conservation of potential vorticity reduces

to 0=∇⋅ Qvr . Combined with (1.1.33), this leads to

0)( =∇×∇⋅ Qk ηr

. (1.1.34)

This means that the surfaces of constant η and constant Q coincide. Accordingly, Q

and η are uniquely related, or

)(Qηη = . (1.1.35)

For the special case where we have topographic steering on an f-plane (when Q varies

only with H), we find from (1.1.35) that

)(Hηη = . (1.1.36)

Accordingly, for stationary, geostrophic motion with a free surface, the iso-lines for

the surface elevation are parallel to the depth contours.

1.2 Linear waves in the absence of rotation

We assume small disturbances from the state of equilibrium in the ocean, two-

dimensional motion (∂/∂y = 0, v = 0), and constant depth. Furthermore, we take the

Page 9: Dynamic Oceanography

9

atmospheric pressure to be constant along the surface. The set of equations (1.1.18)-

(1.1.20) can then be linearized, reducing to

⎭⎬⎫

−=−=

.,

xt

xt

Hugu

ηη

(1.2.1)

Eliminating the horizontal velocity, we find

0=− xxtt gHηη . (1.2.2)

This equation is called the wave equation. By inserting a wave solution of the type

)(exp( 0tcxik ±∝η , representing a complex Fourier component, we find from (1.2.2)

that 2/10 )(gHc = , i.e. all wave components move with the same phase speed

regardless of wavelength. Such waves are called non-dispersive. We need not limit

ourselves to consider one single Fourier component. From (1.2.2) we realize

immediately that a general solution can be written

)()( 0201 tcxFtcxF −++=η . (1.2.3)

If, at time t = 0, the surface elevation was such that η = F(x), and ηt = 0, it is easy to

see that the solution becomes

)()(21

00 tcxFtcxF −++=η . (1.2.4)

From (1.2.1) and (1.2.4) we find for the acceleration

)(')('2 00 tcxFtcxFggu xt −++−=−= η , (1.2.5)

where ξξ ddFF /)(' = . Hence, the horizontal velocity is given by

)()(2 00

0

tcxFtcxFcgu −−+−= . (1.2.6)

From (1.2.4) we can display the evolution of an initially bell-shaped surface elevation

F(x) with typical width L; see the sketch in Fig. 1.3.

Page 10: Dynamic Oceanography

10

Fig. 1.3. Evolution of a bell-shaped surface elevation.

We note that the initial elevation splits into two identical pulses moving right and left

with velocity c0 = (gH)1/2. In a deep ocean (H = 4000 m), the phase speed is c0 ≈ 200

m s−1, while in a shallow ocean (H = 100 m) we have c0 ≈ 30 m s−1. If the maximum

initial elevation in this example is h, i.e. F(0) = h, we find from (1.2.6) that the

velocity in the ocean directly below peak of the right-hand pulse can be written

02c

ghu = , (1.2.7)

when t >> L/c0, that is after the two pulses have split. If we take h = 1 m as a typical

value, the deep ocean example yields u ≈ 2.5 cm s−1, while for the shallow ocean we

find u ≈ 17 cm s−1.

As a second example we consider an initial step function:

⎪⎩

⎪⎨

<−

>=

.0,21

,0,21

)(xh

xhxF (1.2.8)

In this case, the velocity and amplitude development becomes as sketched in Fig. 1.4.

Page 11: Dynamic Oceanography

11

Fig. 1.4. Evolution of a surface step function.

It is obvious that we in an example like this (with a step in the surface at t = 0) must

be careful when using linear theory, which requires small gradients. In a more

realistic example where differences in height occurs, the initial elevation will have a

final (an quite small) gradient around x = 0. Qualitatively, however, the solution

becomes as discussed above.

It is easy to show that the solution for the step problem satisfies the mass and

energy conservation laws: Choose a geometrically fixed area –D ≤ x ≤ D, where D

≥ c0t. From Fig. 1.4 it is obvious that the mass of fluid within this area is constant.

We choose z = −h/2 as the level where the potential energy is zero. The initial

potential energy can thus be written

2

21

0gDhMghE Gp ρ== , (1.2.9)

where M = ρDh is the fluid mass above the zero level, and hG = h/2 is the vertical

distance to the centre of mass of this fluid. The initial kinetic energy is zero, since we

start the problem from rest. At a later time t = D/c0, the potential energy becomes

2

41

4gDhhMgE p ρ== , (1.2.10)

while the kinetic energy can be written

Page 12: Dynamic Oceanography

12

22

41)2(

21 gDhuDHE k ρρ == . (1.2.11)

Thus, from (1.2.9)-(1.2.11) we find that 0ppk EEE =+ , i.e. the total energy is

conserved. This is of course also true for the bell-shaped elevation in the first

example.

The hydrostatic approximation

Finally, we address the validity of the hydrostatic approximation in the case of

waves in a non-rotating ocean. We rewrite the pressure as a hydrostatic part plus a

deviation:

')( pPzgp S ++−= ηρ , (1.2.12)

where 'p is the non-hydrostatic deviation. The vertical component of (1.1.1) becomes

to lowest order:

zt pw ′−=ρ1 , (1.2.13)

while the horizontal component can be written

xxt pgu ′−−=ρ

η 1 . (1.2.14)

The hydrostatic assumption implies that

tx up <<′ρ1 . (1.2.15)

If the typical length scales in the x- and z-directions are L and H, respectively, we

obtain from the continuity equation that

wHLu ~ , (1.2.16)

where ~ means order of magnitude. From (1.2.13) we then find

tuL

Hp 2

~'ρ

. (1.2.17)

Utilizing this result, the condition (1.2.15) reduces to

1/ 22 <<LH . (1.2.18)

Thus, we realize that the assumption of a hydrostatic pressure distribution in the

vertical requires that the horizontal scale L of the disturbance must be much larger

Page 13: Dynamic Oceanography

13

than the ocean depth. For a wave, L is associated with the wavelength; for a single

pulse, L corresponds to the characteristic pulse width.

1.3 Effect of rotation, geostrophic adjustment

We now consider the effect of the earth’s rotation upon wave motion. Linear

theory still applies, and we take the depth and the surface pressure to be constant.

Furthermore, we assume that f is constant. Equations (1.1.18)-(1.1.20) then reduces to

xt gfvu η−=− , (1.3.1)

yt gfuv η−=+ , (1.3.2)

0)( =++ yxt vuHη . (1.3.3)

We compute the vertical vorticity and the horizontal divergence, respectively, from

(1.3.1) and (1.3.2). By utilizing (1.1.3), we then obtain

0)( =−− ttyx Hfuv η , (1.3.4)

and

)()( yyxxyxtt guvf

Hηηη

+=−+ . (1.3.5)

The vorticity equation can be integrated in time, i.e.

000 ηηHfuv

Hfuv yxyx −−=−− , (1.3.6)

where sub-zeroes denote initial values. We assume that the problem is started from

rest, which means that there are no velocities or velocity gradients at t = 0. Thus

)( 0 ηη −−=−Hfuv yx . (1.3.7)

Inserting for the vorticity in (1.3.5), we find that

0222

0 )( ηηηηη ffc yyxxtt =++− , (1.3.8)

where gHc =20 , and η0 is a known function of x and y (the surface elevation at t = 0).

The solution to (1.3.8) can be written as a sum of a transient (free) part and a

stationary (forced) part

),(),,(~ yxtyx ηηη += , (1.3.9)

where η~ and η fulfil, respectively

0~)~~(~ 220 =++− ηηηη fc yyxxtt , (1.3.10)

Page 14: Dynamic Oceanography

14

0222

0 )( ηηηη ffc yyxx =++− . (1.3.11)

Equation (1.3.10) for the transient, free solution is called the Klein-Gordon equation

and occurs in many branches in physics. Here, it describes long surface waves that are

modified by the earth’s rotation (Sverdrup waves). These waves will be discussed in

the next section. Notice that the initial conditions for the free solution are

),(),()0,,(~0 yxyxyx ηηη −= and 0~ =tη .

As an example of a stationary solution of (1.3.8), we return to the problem in the

last section where the surface elevation initially was a step function:

⎩⎨⎧

<−>

=,0,2,0,2

)(0 xhxh

xη (1.3.12)

or, for simplicity,

.0,0

,1,1

)sgn(),sgn(21)(0 <

>

⎩⎨⎧−

==xx

xxhxη (1.3.13)

We assume that the motion is independent of the y-coordinate. From (1.3.11) we then

obtain

)sgn(21 22 xhaaxx

−− −=− ηη , (1.3.14)

where fca /0= . It is easy to show that the solution of (1.3.14) is given by

)sgn()/exp(121 xaxh −−=η . (1.3.15)

We have sketched this solution in Fig. 1.5.

Fig. 1.5 Geostrophic adjustment of a free surface.

Page 15: Dynamic Oceanography

15

The distance a = c0/f = (gH)1/2/f is called the Rossby radius of deformation, or the

Rossby radius for short. A typical value for f at mid latitudes is 10−4s−1. For a deep

ocean (H = 4000 m), we find that a ≈ 2000 km, while for a shallow ocean (H = 100

m), a ≈ 300 km.

From (1.3.1) and (1.3.2) we find the velocity distribution for this example, i.e.

xgvf η= , (1.3.16)

and

0=u . (1.3.17)

We note from (1.3.16) that we have a balance between the Coriolis force and the

pressure-gradient force (geostrophic balance) in the x-direction. From (1.3.15) and

(1.3.16) the corresponding geostrophic velocity in the y-direction can be written

)/exp(2 0

axc

ghv −= . (1.3.18)

This is a “jet”-like stationary flow in the positive y-direction. Although the

geostrophic adjustment occurs within the Rossby radius, we notice from (1.3.18) that

the maximum velocity in this case is independent of the earth’s rotation. By

comparison with (1.2.7), we see that our maximum velocity it is the same as the

velocity below a moving pulse with height h/2, or as the velocity in the non-rotating

step-problem.

Let us compute the kinetic and the potential energy within a geometrically fixed

area DxD ≤≤− for the stationary solution (1.3.15)-(1.3.18), which is valid when t

→ ∞. The kinetic energy becomes

∫ ∫−

−=⎟⎟⎠

⎞⎜⎜⎝

⎛=

D

D

aD

Hk eaghdxdzvE )1(

81

21 /222 ρρ

η

, (1.3.19)

where we have used the fact that η>>H . For the potential energy we find

)43(81

21'' /2/22

2/

0

aDaDD

D

h

p eeaghDghdxdzzgE −−

+

−+−+=⎟⎟⎠

⎞⎜⎜⎝

⎛= ∫ ∫ ρρρ

η

, (1.3.20)

where we have taken z = −h/2 as the level of zero potential energy, and introduced

2/' hzz += . Initially, the total mechanical energy within the considered area equals

the potential energy, or

DghEE p2

0 21

0ρ== . (1.3.21)

Let us choose D >> a. We then notice from (1.3.19)-(1.3.21) that

Page 16: Dynamic Oceanography

16

0EEE pk <+ . (1.3.22)

Thus, when t → ∞, the total mechanical energy inside the considered area is less than

it was at t = 0. The reason is that energy in the form of free Sverdrup waves (solutions

of the Klein-Gordon equation) has “leaked” out of the area during the adjustment

towards a geostrophically balanced steady state. We will consider these waves in

more detail in the next section.

Finally we discuss in a quantitative way when it is possible to neglect the

effect of earth’s rotation on the motion. For this to be possible, we must have that

vkfvtrrr

×>> . (1.3.23)

Accordingly, the typical timescale T for the motion must satisfy

f

T π2<< . (1.3.24)

At mid latitudes we typically have 2π/f ≈ 17 hours. If the characteristic horizontal

scale of the motion is L and the phase speed is c ~ (gH)1/2, we find from (1.3.24) that

the effect of earth’s rotation can be neglected if

aL << , (1.3.25)

where a = c/f is the Rossby radius of the problem. In the open ocean L will be

associated with the wavelength, while in a fjord or canal, L will be the width.

Oppositely, when

aL ≥ , (1.3.26)

the effect of the earth’s rotation on the fluid motion can not be neglected.

1.4 Sverdrup waves, inertial waves and Poincaré waves

We consider long surface waves in a rotating ocean of unlimited horizontal

extent. Such waves are often called Sverdrup waves (Sverdrup, 1927). They are

solutions of the Klein-Gordon equation (1.3.10). Actually, Sverdrup’s name is usually

related to friction-modified, long gravity waves, but here we will use it also for the

frictionless case. In literature long waves in an inviscid ocean are often called

Poincaré waves. However, this term will be reserved for a particular combination of

Sverdrup waves that can occur in canals with parallel walls.

Sverdrup waves

A surface wave component in a horizontally unlimited ocean can be written

Page 17: Dynamic Oceanography

17

))(exp( tlykxiA ωη −+= . (1.4.1)

This wave component is a solution of (1.3.10) if

)( 2220

22 lkcf ++=ω . (1.4.2)

Here k and l are real wave numbers in the x- and y-direction, respectively, and ω is the

wave frequency. Equation (1.4.2) is the dispersion relation for inviscid Sverdrup

waves. From this relation we note that the Sverdrup wave must always have a

frequency that is larger than (or equal to) the inertial frequency f.

For simplicity we let the wave propagate along the x-axis, i.e. l = 0. The phase

speed now becomes

2/1

22

2

0 41 ⎟⎟

⎞⎜⎜⎝

⎛+==

ac

kc

πλω , (1.4.3)

where λ = 2π/k is the wavelength and a = c0/f is the Rossby radius. We note that the

waves become dispersive due to the earth’s rotation, i.e. the phase speed depends on

the wavelength (here: increases with increasing wave length). The group velocity

becomes

2/1

22

2

0

41

dd

⎟⎟⎠

⎞⎜⎜⎝

⎛+

==

a

ck

cg

πλ

ω . (1.4.4)

We notice that the group velocity decreases with increasing wavelength. From (1.4.3)

and (1.4.4) we realize that ccg = c02, i.e. the product of the phase and group velocities

are constant. From (1.4.2), with l = 0, we can sketch the dispersion diagram for

positive wave numbers; see Fig. 1.6.

Fig. 1.6. The dispersion diagram for Sverdrup waves.

Page 18: Dynamic Oceanography

18

For k << a−1 (i.e. λ >> a) we have that ω ≈ f. This means that the motion is reduced to

inertial oscillations in the horizontal plane. For k >> a−1 gravity dominates, i.e. ω ≈

c0k, and we have surface gravity waves that are not influenced by the earth’s rotation.

Contrary to gravity waves in a non-rotating ocean, the Sverdrup waves discussed

here do possess vertical vorticity. For a wave solution ( )exp( tiω∝ ), (1.3.4) yields

ηζHf

= . (1.4.5)

If we still assume that ∂ /∂y = 0, we obtain from (1.4.5) and (1.3.2) that

⎪⎪⎭

⎪⎪⎬

−=

=

.1

,

t

x

vf

uHfv η

(1.4.6)

With η in (1.4.1) depending only on x and t, we find, when we let the real part

represent the physical solution:

⎪⎪⎪

⎪⎪⎪

−⎟⎠⎞

⎜⎝⎛ +

=

−=

−=

−=

).sin(

),sin(

),cos(

),cos(

0

0

0

0

tkxH

Hzw

tkxkH

fv

tkxkH

u

tkx

ωωη

ωη

ωωηωηη

(1.4.7)

Here the vertical velocity w has been obtained from (1.1.17). Since f≥ω for

Sverdrup waves, we must have that vu ≥ . Furthermore, from (1.4.7) we find that

( ) ( )

1)/()/( 2

0

2

20

2

=+kHf

vkH

uηωη

. (1.4.8)

This means that the horizontal velocity vector describes an ellipsis where the ratio of

the major axis to the minor axis is f/ω . From (1.4.7) it is easy to see that the

velocity vector turns clockwise, and that one cycle is completed in time 2π/ω.

Sverdrup (1927) demonstrated that the tidal waves on the Siberian continental

shelf were of the same type as the waves discussed here. In addition, they were

modified by the effect of bottom friction, which leads to a damping of the wave

amplitude as the wave progresses. Furthermore, friction acts to reduce of the phase

speed, and it causes a phase displacement between maximum current and maximum

surface elevation.

Page 19: Dynamic Oceanography

19

Inertial waves

When discussing the dispersion relation (1.4.2), we realized that for very long

wavelengths the Sverdrup wave was reduced to an inertial wave. For the inertial wave

phenomenon, the effect of gravity is unimportant. Hence, to study such waves we may

take that the surface is horizontal at all times, i.e. η ≡ 0. Furthermore, for infinitely

long waves we must put 0/,0/ ≈∂∂≈∂∂ yx in the governing equations (1.3.1)-

(1.3.3). They then reduce to

⎭⎬⎫

=+=−

.0,0

fuvfvu

t

t (1.4.9)

Accordingly,

0),(22

2

=⎟⎟⎠

⎞⎜⎜⎝

⎛+ vuf

dtd , (1.4.10)

where we have changed to ordinary derivatives since u and v now are only functions

of time. With initial conditions u = u0, v = 0, the solution becomes

⎭⎬⎫

−==

.sin,cos

0

0

ftuvftuu

(1.4.11)

We notice from (1.4.11) that 20

22 uvu =+ , i.e. the magnitude of the velocity is u0

everywhere in the ocean, and that the velocity vector turns clockwise in the northern

hemisphere (where f > 0). The trajectory of a single fluid particle (xp, yp) can be found

from the relations u = dxp /dt and v = dyp /dt, where u and v are given by (1.4.11).

Accordingly

⎪⎪⎭

⎪⎪⎬

=−

=−

,cos

,sin

00

00

ftf

uyy

ftf

uxx

p

p

(1.4.12)

or

2

202

02

0 )()(fuyyxx pp =−+− . (1.4.13)

This means that a fluid particle moves in a closed circle around the point (x0, y0),

which is the mean position of the particle in question. The radius of the circle is r =

u0/f, which is called the inertial radius. If we use the typical values u0 = 10 cm s−1 and

f = 10−4s−1, we find that r = 1 km. The inertial period T = 2π/f, which is the time

Page 20: Dynamic Oceanography

20

needed for a particle to make one closed loop, is about 17 hours in this example. The

particle moves in a clockwise direction in the northern hemisphere. Since the vertical

component of the earth’s rotation at a location with latitude ϕ is Ωsinϕ, we can

introduce the pendulum day TΩ defined by

ϕ

πsin2

Ω=ΩT . (1.4.14)

We notice that the inertial period is half a pendulum day. It turns out that inertial

oscillations are quite common in the ocean. Most ocean current spectra show a local

amplitude or energy maximum at the inertial frequency.

Poincaré waves

We consider waves in a uniform canal along the x-axis with depth H and width B.

Such waves must satisfy the Klein-Gordon equation (1.3.10). But now the ocean is

laterally limited. At the canal walls, the normal velocity must vanish, i.e. v = 0 for y =

0, B. By inspecting (1.4.7), we realize that no single Sverdrup wave can satisfy these

conditions. However, if we superimpose two Sverdrup waves, both propagating at

oblique angles (α and −α, say) with respect to the x-axis, we can construct a wave

which satisfies the required boundary conditions. The velocity component in the y-

direction must then be of the form

,..3,2,1)),(exp()sin(0 =−= ntkxiB

ynvv ωπ (1.4.15)

Since the wave number Bnl /π= in the y-direction now is discrete due to the

boundary conditions, the dispersion relation (1.4.2) becomes

...,3,2,1,2/1

2

2222

02 =⎥

⎤⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛++±= n

Bnkcf πω (1.4.16)

We notice from (1.4.15) that the spatial variation in the cross-channel direction is

trigonometric. Such trigonometric waves in a rotating channel are called Poincaré

waves. They can propagate in the positive as well as the negative x-direction. We shall

see that this is in contrast to coastal Kelvin waves, which we discuss later in section 1.

In general, the derivation of the complete solution for Poincaré waves is too lengthy

to be discussed in this text. For a detailed derivation; see for example LeBlond and

Mysak (1978), p. 270.

Page 21: Dynamic Oceanography

21

Energy considerations

By utilizing the solution (1.4.7), we can compute the mechanical energy

associated with Sverdrup waves. The mean potential energy per unit area of a fluid

column can be written

20

/2

0 0 41)(

/21 ρηρωπ

ωπ η

gdtdzzgEp == ∫ ∫ . (1.4.17)

The mean kinetic energy per unit area becomes

( ) 2022

22/2

0

0222

/1/1

41

21

/21 η

ωωρρ

ωπ

ωπ

⎥⎦

⎤⎢⎣

⎡−+

=++= ∫ ∫− f

fgdtdzwvuEH

k , (1.4.18)

where we have utilized that kH << 1. We see that in a rotating ocean (f ≠ 0), the mean

potential and the mean kinetic energy in the wave motion are no longer equal

(compare with the results (1.2.10) and (1.2.11) for the non-rotating case, where we

have an equal partition between the two). The dominating part of the mean energy is

now kinetic. The total mean mechanical energy becomes

20

220 /

21 ccgEEE pk ηρ=+= . (1.4.19)

Energy flux and group velocity

The formula for the group velocity, dkdcg /ω= , used in (1.4.4), arises from

purely kinematic considerations. By superimposing two progressive wave trains with

the same amplitude and with wave numbers k and k +∆k, and frequencies ω and ω

+∆ω, respectively, one finds that the envelope containing the wave crests and wave

troughs, defined as the wave group, will propagate with the speed k∆∆ /ω . However,

the group velocity has also a very important dynamical interpretation. Consider a

Sverdrup wave that propagates along x-axis. This wave induces a net transport of

energy in the x-direction. The energy transport through a vertical cross-section can be

found by computing the mean work done by the fluctuating pressure at that section.

The mean work F , per unit time and unit area, which is performed on the fluid to the

right of a vertical cross-section, can be written

∫ ∫−

=ωπ η

ωπ

/2

0

)(/2

1

H

dtdzpuF , (1.4.20)

where p is the fluctuating part of the pressure in the fluid. Inserting from (1.1.11) and

(1.4.7), it follows that

Page 22: Dynamic Oceanography

22

cgF 202

1 ηρ= . (1.4.21)

For a regular, infinitely long wave train the mean energy cannot accumulate in space.

Hence there must be a transport of mean total energy E through the considered

cross-section. Denote this transport velocity by ce. Energy balance then requires that

EcF e= . (1.4.22)

Equation (1.4.22) yields, by inserting from (1.4.19) and (1.4.21), that

ge cccc == /20 , (1.4.23)

where the last equality follows from (1.4.4). Accordingly, the mean energy in the

wave motion propagates with the group velocity.

The work done by the fluctuating pressure per unit mass and unit time is often

referred to as the energy flux, and the total energy per unit mass as the energy

density. These quantities vary in space and time. Both concepts follow naturally from

the energy equation for the fluid. With no variation in the y-direction, the linearized

equations (1.3.1)-(1.3.3) reduce to

⎪⎭

⎪⎬

−==+−=−

.,0

,

xt

t

xt

Hufuv

gfvu

η

η (1.4.24)

By multiplying the two first equations by u and v, respectively, and then adding, we

obtain

ηη xgugux

vut

+∂∂

−=⎥⎦⎤

⎢⎣⎡ +

∂∂ )()(

21 22 . (1.4.25)

Obviously, the Coriolis force does not perform any work since it acts perpendicular

to the displacement (or velocity). By inserting from the continuity equation that

Hu tx /η−= into the last term, (1.4.25) becomes

0)(21 222 =

∂∂

+⎥⎦

⎤⎢⎣

⎡⎟⎠⎞

⎜⎝⎛ ++

∂∂ ηη gu

xHgvu

t. (1.4.26)

We write this equation

0=∂∂

+∂∂

fd ex

et

, (1.4.27)

where the energy density ed and the energy flux ef are defined, respectively, as

)(21 222 η

Hgvued ++= , (1.4.28)

Page 23: Dynamic Oceanography

23

ηgue f = . (1.4.29)

The mean values for a vertical fluid column become

⎪⎪⎭

⎪⎪⎬

==

==

∫ ∫

∫ ∫

−ωπ η

ωπ η

ωηωπ

ωηωπ

/2

0

20

/2

020

2

220

./21)(

/21

,21)(

/21

Hff

Hdd

kgdtdzee

ckgdtdzee

(1.4.30)

By comparing with (1.4.19) and (1.4.21) we realize that ρ/Eed = and ρ/Fe f = ,

and hence dgf ece = . Even though we have only shown this to be valid for Sverdrup

waves, it is a quite general result, and valid for all kinds of wave motion.

1.5 Kelvin waves at a straight coast

We consider an ocean that is limited by a straight coast. The coast is situated at y

= 0; see Fig. 1.7.

Fig. 1.7. Definition sketch.

Furthermore, we assume that the velocity component in the y-direction is zero

everywhere, i.e. v ≡ 0. With constant depth and constant surface pressure (1.4.24)

becomes

xt gu η−= , (1.5.1)

ygfu η−= , (1.5.2)

xt Hu−=η . (1.5.3)

Page 24: Dynamic Oceanography

24

We take that the Coriolis parameter is constant, and eliminate u from the problem.

Equations (1.5.1) and (1.5.2) yield

0=− xyt fηη , (1.5.4)

while (1.5.2) and (1.5.3) yield

00 =− xyt ac ηη . (1.5.5)

Here 2/10 )(gHc = and a = c0/f. We assume a solution of the form

),()( txFyG=η . (1.5.6)

By inserting into (1.5.5), we find

G

aGFc

F

x

t '

0

= , (1.5.7)

where dydGG /'= . The left-hand side of (1.5.7) is only a function of x and t, and the

right-hand side is only a function of y. Thus, for (1.5.7) to be valid for arbitrary values

of x, y, and t, both sides must equal to the same constant, which we denote by γ (γ ≠ 0

for a non-trivial solution). Hence

).(

),/exp('

00

tcxFFFc

F

ayGG

aG

x

t γγ

γγ

+=⇒=

=⇒= (1.5.8)

By inserting from (1.5.8) into (1.5.4), we find that

1±=γ . (1.5.9)

Accordingly, from (1.5.8), we have solutions of the form

)()/exp( 0tcxFay +=η , (1.5.10)

and

)()/exp( 0tcxFay −−=η . (1.5.11)

If we have a straight coast at y = 0 and an unlimited ocean for y > 0, as depicted in

Fig. 1.7, the solution (1.5.10) must be discarded. This is because η must be finite

everywhere in the ocean, even when y → ∞. The solution for the surface elevation and

the velocity distribution in this case then become

).()/exp(

),()/exp(

0

0

tcxFayfagu

tcxFay

−−=

−−=η (1.5.12)

This type of wave is called a single Kelvin wave (double Kelvin waves will be treated

in section 1.8). It is trapped at the coast within a region determined by the Rossby

Page 25: Dynamic Oceanography

25

radius. It is therefore also referred to as a coastal Kelvin wave. The Kelvin wave

propagates in the positive x-direction with velocity c0, like a gravity wave without

rotation. The difference from the non-rotating case, however, is that now we do not

have the possibility of a wave in the negative x-direction. This is because the Kelvin

wave solution requires geostrophic balance in the direction normal to the coast; see

(1.5.2). This is impossible for a wave in the negative x-direction in the northern

hemisphere. In general, if we look in the direction of wave propagation (along the

wave number vector), a Kelvin wave in the northern hemisphere always moves with

the coast to the right, while in the southern hemisphere (f < 0), it moves with the coast

to the left; see the sketch in Fig. 1.8 for a single Fourier component in the northern

hemisphere.

Fig. 1.8. Propagation of Kelvin waves along a straight coast when f > 0.

Since the wave amplitude is trapped within a region limited by the Rossby radius, the

wave energy is also trapped in this region. The energy propagation velocity (the group

velocity) is here 00 /)(/ cdkkcddkdcg === ω , and the energy is propagating with the

coast to the right in the northern hemisphere. We note that for Kelvin waves the

frequency has not a lower limit (for Sverdrup waves ω ≥ f).

The oceanic tide may in certain places manifest itself as coastal Kelvin waves of

the type studied here. We will discuss this further in connection with amphidromic

points (points where the tidal height is always zero). From (1.5.12) we notice that the

surface elevation and velocity are in phase, i.e. maximum high tide coincides with

maximum current. It turns out from measurements that the maximum tidal current at a

given location occurs before maximum tidal height. This is due to the effect of

friction at the ocean bottom, which we have neglected so far.

Page 26: Dynamic Oceanography

26

The effect of bottom friction will also modify the Kelvin wave solution (1.5.12)

in various other ways. In particular, it turns out that the lines describing a constant

phase (the co-tidal lines) are no longer directed perpendicular to the coast, but are

slanting backwards relative to the direction of wave propagation (Martinsen and

Weber, 1981). This situation is sketched in Fig. 1.9.

Fig. 1.9. Coastal Kelvin waves influenced by friction.

Other frictional effects on Kelvin waves are reduced phase speed (c < c0), and an

amplitude decrease as the wave progresses, which is similar to how friction affects

Sverdrup waves.

1.6 Amphidromic systems

Wave systems, where the lines of constant phase, or the co-tidal lines, form a

star-shaped pattern, are called amphidromies. They are wave interference phenomena,

and in the ocean they usually originate due to interference between Kelvin waves. Let

us study wave motion in an ocean with width B; see the sketch in Fig. 1.10.

Fig. 1.10. Ocean with parallel boundaries (infinitely long canal).

Page 27: Dynamic Oceanography

27

Since the ocean now is limited in the y-direction, both Kelvin wave solutions (1.5.10)

and (1.5.11) can be realized. Because we are working with linear theory, the sum of

two solutions is also a solution, i.e.

)()/exp()()/exp( 00 tcxFaytcxFay −−++=η . (1.6.1)

In general the F-functions in (1.6.1) can be written as sums of Fourier components. It

suffices here to consider two Fourier components with equal amplitudes:

( ))sin()/exp()sin()/exp( tkxaytkxayA ωωη −−++= , (1.6.2)

where ω = c0k. Along the x-axis, i.e. for y = 0, (1.6.2) reduces to

tkxA ωη cossin2= . (1.6.3)

This constitutes a standing oscillation with period T = 2π/ω. Zero elevation (η = 0)

occur when

...,2,1,0, == nk

nx π (1.6.4)

At the locations given by (nπ/k, 0), the surface elevation is zero at all times. These

nodal points are referred to as amphidromic points.

We consider the shape of the co-phase lines, and choose a particular phase, e.g. a

wave crest (or trough). At a given time the spatial distribution of this phase is given

by 0=tη ; i.e. a local extreme for the surface elevation. Partial differentiation (1.6.2)

with respect to time yields that the co-phase lines are given by the equation

0)cos()/exp()cos()/exp( =−−−+ tkxaytkxay ωω . (1.6.5)

We notice right away that the co-phase lines must intersect at the amphidromic points

0,/ == yknx π for all times. As an example, we consider the amphidromic point at

the origin. In a sufficiently small distance from origin, x and y are so small that we

can make the approximations kxkxkxayay ≈≈±≈± sin,1cos,/1)/exp( . Equation

(1.6.5) then yields

xtkay )tan( ω= . (1.6.6)

This means that the co-phase lines are straight lines in a region sufficiently close to

the amphidromic points. Since tωtan is a monotonically increasing function of time in

the interval 0=t to )2/( ωπ=t , we see that a co-phase line revolves around the

amphidromic point in a counter-clockwise direction in this example (f > 0). It turns

out that, as a main rule, the co-phase lines of the amphidromic systems in the world

oceans rotate counter-clockwise in the northern hemisphere and clockwise in the

Page 28: Dynamic Oceanography

28

southern hemisphere. We notice from (1.6.6) that if we have high tide along a line in

the region x > 0, y > 0 at some time t, we will have high tide along the same line in the

region x < 0, y < 0 at time ωπ /+t , or half a period later.

We now consider the numerical value of η along a co-phase line. Close to an

amphidromic point, (here the origin), we can use (1.6.2) to express the elevation as

)sincos(2 taytkxA ωωη += . (1.6.7)

From (1.6.6) we find that )/(tan kaxyt =ω along a co-phase line. By eliminating the

time dependence between this expression and (1.6.7), we find for the magnitude of the

surface elevation along a co-tidal line:

( ) 2/12222 /2 ayxkA +=η . (1.6.8)

The lines for a given difference between high and low tide are called co-range lines.

These curves are given by (1.6.8), when η is put equal to a constant, i.e.

.const/ 2222 =+ ayxk (1.6.9)

We thus see that the co-range lines close to the amphidromic points are ellipses.

In Fig. 1.11 we have depicted co-phase lines (solid curves) and co-range lines

(broken curves) resulting from the superposition of two oppositely travelling Kelvin

waves, both with periods of 12 hours and amplitudes of 0.5 m. The wavelength is 800

km, the width of the channel is 400 km, the depth is 40 m, and the Coriolis parameter

is 10−4s−1. The Rossby radius becomes 198 km in this example. Hence the right-hand

side of the channel is dominated by the upward-propagating Kelvin wave (the one

with minus sign in the phase), and the left-hand side is dominated by the downward

propagating Kelvin wave.

Page 29: Dynamic Oceanography

29

Fig. 1.11. Amphidromic system in an infinitely long canal.

1.7 Equatorial Kelvin waves

Close to equator we have that f0 ≈ 0. The Coriolis parameter in this region can

then be approximated by

yf β= , (1.7.1)

where the y-axis is directed northwards; see Fig. 1.12.

Fig. 1.12. Sketch of the co-ordinate axes near the equator.

We will find that it is possible to have equatorially trapped gravity waves, analogous

to the trapping at a straight coastline. Assume that the velocity component in the y-

direction is zero everywhere, i.e. we assume geostrophic balance in the direction

Page 30: Dynamic Oceanography

30

perpendicular to equator. With constant depth, the equations (1.5.1)-(1.5.3) are

unchanged, but now yf β= in (1.5.2). By assuming a solution of the form

),()( txFyG=η as before, (1.5.7) becomes

γβ

==yGGc

FcF

x

t '0

0

. (1.7.2)

Accordingly:

⎪⎭

⎪⎬

⎟⎟⎠

⎞⎜⎜⎝

⎛=

+=

.2

exp

),(2

0

0

yc

G

tcxFFβγγ

(1.7.3)

By inserting into (1.5.4), we find

1±=γ . (1.7.4)

From (1.7.3) we realize that to have finite solution when y → ± ∞, we must choose γ

= −1 in (1.7.4). The solution thus becomes

⎪⎭

⎪⎬⎫

−−=

−−=

),()/exp(

),()/exp(

022

0

022

tcxFaycgu

tcxFay

e

eη (1.7.5)

where the equatorial Rossby radius is defined by

2/10 )/2( βcae = . (1.7.6)

We note that the solution (1.7.5), referred to as an equatorial Kelvin wave, is valid at

both sides of equator and that it propagates in the positive x-direction, i.e. eastwards

with phase speed c0 = (gH)1/2. The energy also propagates eastwards with the same

velocity, since we have no dispersion. At the equator β is approximately

2×10−11m−1s−1. For a deep ocean with H = 4000 m, we find from (1.7.6) that the

equatorial Rossby radius becomes about 4500 km.

Equatorial Kelvin waves are generated by tidal forces, and by wind stress and

pressure distributions associated with storm events with horizontal scales of thousands

of kilometres. When such waves meet the eastern boundaries in the ocean (the west

coast of the continents), part of the energy in the wave motion will split into a

northward propagating coastal Kelvin waves in the northern hemisphere, and a

southward propagating coastal Kelvin wave in the southern hemisphere. Some of the

energy will also be reflected in the form of planetary Rossby waves. This type of

wave will be discussed in section 1.9.

Page 31: Dynamic Oceanography

31

1.8 Topographically trapped waves

We have seen that gravity waves can be trapped at the coast or at the equator due

to the effect of the earth’s rotation. Trapping of wave energy in a rotating ocean can

also occur in places where we have changes in the bottom topography. In this case,

however, the wave motion is fundamentally different from that due to Kelvin waves.

While the velocity field induced by Kelvin waves is always zero in a direction

perpendicular to the coast, or equator, it is in fact the displacement of particles

perpendicular to the bottom contours that generates waves in a region with sloping

bottom. We call these waves escarpment waves, and they arise as a consequence of

the conservation of potential vorticity.

Rigid lid

The escarpment waves are essentially vorticity waves. The motion in these waves

is a result of the conservation of potential vorticity. More precisely, the relative

vorticity for a vertical fluid column changes periodically in time when the column is

stretched or squeezed in a motion back and forth across the bottom contours. To study

such waves in their purest form, we will assume that the surface elevation is zero at

all times, i.e. we apply the rigid lid approximation. In this way the effect of gravity is

eliminated from the problem. Let us assume that the bottom topography is as sketched

in Fig. 1.13.

Fig. 1.13. Bottom topography for escarpment waves.

The continuity equation (1.1.20) can now be written as

0)()( =+ yx HvHu . (1.8.1)

Page 32: Dynamic Oceanography

32

Accordingly, we can define a stream function ψ satisfying

⎭⎬⎫

=−=

.,

x

y

HvHu

ψψ

(1.8.2)

When linearizing, we obtain from the theorem of conservation of potential vorticity

(1.1.25) that

0=⎟⎠⎞

⎜⎝⎛+⎟

⎠⎞

⎜⎝⎛+

yx

t

Hfv

Hfu

Hζ . (1.8.3)

We here assume that f is constant. Furthermore, we take that )(yHH = . By inserting

ψ from (1.8.2), we can write (1.8.3) as

0)(2 =+′

−∇ xytt fHH ψψψ , (1.8.4)

where dydHH /=′ . We assume a wave solution of the form

))(exp()( tkxiyF ωψ −= . (1.8.5)

By inserting into (1.8.4), this yields

02

2

=⎟⎟⎠

⎞⎜⎜⎝

⎛ ′−−

′⎟⎠⎞

⎜⎝⎛ ′

FHHkf

Hk

HF

ω. (1.8.6)

This equation has non-constant coefficients and is therefore problematic to solve for a

general form of H(y). We shall not make any attempts to do so here. Instead, we

derive solutions for two extreme types of bottom topography. One of these cases,

where the bottom exhibits a weak exponential change in the y-direction, will be dealt

with in section 10 in connection with topographic Rossby waves. The other extreme

case, where the slope tends towards a step function, will be analysed here; see the

sketch in Fig. 1.14. The escarpment waves relevant for this topography are often

called double Kelvin waves.

Fig. 1.14. The bottom configuration for double Kelvin waves.

Page 33: Dynamic Oceanography

33

For trapped waves, the solutions of (1.8.6) in areas (1) and (2) are, respectively

⎭⎬⎫

−==

).exp(),exp(

22

11

kyAFkyAF

(1.8.7)

We note that these waves are trapped within a distance of one wavelength on each

side of the step. At the step itself (y = 0), the volume flux in the y-direction must be

continuous, i.e.

0,2211 == yHvHv . (1.8.8)

This means that ψx (and thereby also ψ) must be continuous for y = 0, i.e. A1 = A2 = A

in (1.8.7). Furthermore, the pressure in the fluid must be continuous for y = 0. The

pressure is obtained from the linearized x-component of (1.1.1), i.e.

( )fHvHuH

fvup ttx −−=−−= )()( ρρ . (1.8.9)

Writing ))(exp()( tkxiyPp ω−= , and applying (1.8.2) and (1.8.5), we find that

⎟⎠⎞

⎜⎝⎛ +

′−=

HFf

kHFP ωρ . (1.8.10)

By inserting from (1.8.7), with A1 = A2, into (1.8.10), continuity of the pressure at y =

0 yields the dispersion relation

⎥⎦

⎤⎢⎣

⎡+−

=12

12

HHHH

fω . (1.8.11)

We note that we always have that f<ω , and that the wave propagates with shallow

water to the right in the northern hemisphere, i.e. ω > 0 when f > 0. These two

properties are generally valid for escarpment waves, even though we have only shown

it for double Kelvin waves with a rigid lid on top.

In the case where the escarpment represents the transition between a continental

shelf of finite width and the deep ocean, this type of waves are often called

continental shelf waves. This kind of bottom topography is found outside the coast of

Western Norway. Here, numerical results show the existence of continental shelf

waves in the area close to the shelf-break, e.g. Martinsen, Gjevik and Røed (1979).

The topographic trapping of long waves near the shelf-break and the currents

associated with these waves, interact with the wind-generated surface waves, which

tend to make the sea state here particularly rough. This is a well-known fact among

fishermen and other sea travellers that frequent this region.

Page 34: Dynamic Oceanography

34

The effect of gravity

In general, we must allow the sea surface to move vertically. Let us consider a

wave solution of the form

))(exp(),,( tkxivu ωη −∝ . (1.8.12)

For such waves, the linear versions of (1.1.18) and (1.1.19), with constant surface

pressure, yield

( )

( )⎪⎪⎭

⎪⎪⎬

+−

=

+−

−=

.

,

22

22

y

y

kff

igv

fkf

gu

ωηηω

ηωηω (1.8.13)

We write the surface elevation as

))(exp()( tkxiyG ωη −= . (1.8.14)

Inserting into (1.1.20), we find

0)( 222

=⎥⎦

⎤⎢⎣

⎡′+−

−+′′ GHkfHk

gfGH

ωω . (1.8.15)

For 22 f<<ω , i.e. quasi-geostrophic motion, we revert to the gravity-modified

escarpment wave. For f = 0 and yH )(tanα= , this equation yields the so-called edge

waves. These are long gravity waves trapped at the coast. The trapping in this case is

not due to the earth’s rotation, but is caused by the fact that the local phase speed

gH increases with increasing distance from the coast. If we represent the wave by a

ray, which is directed along the local direction of energy propagation, i.e.

perpendicular to the co-phase lines, the ray will always be gradually refracted towards

the coast. At the coast, the wave is reflected, and the refraction process starts all over

again. The total wave system thus consists of a superposition between an incident and

a reflected wave in an area near the coast. The width of this area depends on the angle

of incidence with the coast for the ray in question. Outside this region, the wave

amplitude decreases exponentially. The edge waves can propagate in the positive as

well as in the negative x-direction. The lowest possible wave frequency is given by

αω tan2 gk= . (1.8.16)

There are also a number of higher, discrete frequencies for this problem; see LeBlond

and Mysak (1978), p. 221. If we take the earth’s rotation into account (f ≠ 0), the

Page 35: Dynamic Oceanography

35

frequencies for the edge waves in the positive and negative x-directions will be

slightly different.

1.9 Planetary Rossby waves

Like topographic waves, planetary Rossby waves are a result of the conservation

of potential vorticity. However, in this case the relative vorticity of a vertical fluid

column changes periodically in time when the planetary vorticity changes as the

column moves back and forth in the north-south direction. We realize then, that this

motion occurs on a beta-plane. The motion is essentially horizontal. As for

escarpment waves, we can eliminate gravity completely from the problem by

assuming that the surface is horizontal for all times, i.e. η ≡ 0. The hydrostatic

assumption then yields for the pressure

),,( tyxpgzp +−= ρ . (1.9.1)

The linearized equations thus become

⎪⎪⎪

⎪⎪⎪

=+

−=+

−=−

.0)()(

,1

,1

yx

yt

xt

HvHu

pfuv

pfvu

ρ

ρ

(1.9.2)

First, we take that H is constant, and that f = f0+βy. From the continuity equation, we

can define a stream function ψ such that

xy vu ψψ =−= , . (1.9.3)

By eliminating the pressure from (1.9.2) we find

02 =+∇ xtH βψψ , (1.9.4)

where 22222 // yxH ∂∂+∂∂=∇ is the horizontal Laplacian operator. This equation also

follows directly from the linearized potential vorticity conservation equation (1.8.3),

when H is constant and fy = β. We assume wave solutions of the form

))(exp( tlykxi ωψ −+∝ . (1.9.5)

Inserting (1.9.5) into (1.9.4) yields for the frequency

22 lkk+

−=βω . (1.9.6)

Page 36: Dynamic Oceanography

36

We shall discuss this dispersion relation in more detail, but first we repeat some

elements of general wave kinematics.

We introduce the wave number vector κr defined by

332211 ikikikrrrr

++=κ , (1.9.7)

and a radius vector rr , where

332211 iririrrrrrr

++= . (1.9.8)

A plane wave of the type (1.9.5) can now be written

exp))(exp( 2 ⎟⎠⎞

⎜⎝⎛ −⋅=−⋅= triAtriA

κκωκωκψr

rrrr. (1.9.9)

Accordingly, the vectorial phase speed cr can be defined by

23

22

21

22 , kkkc ++== κ

κκω r

r . (1.9.10)

The components of the vectorial group velocity gcr are given by

⎪⎭

⎪⎬

∂∂=∂∂=∂∂=

./,/,/

3)3(

2)2(

1)1(

kckckc

g

g

g

ωωω

(1.9.11)

In vector notation this becomes

ωκ∇=gcr , 3

32

21

1 ki

ki

ki

∂∂

+∂∂

+∂∂

≡∇rrr

κ . (1.9.12)

If the frequency ω only is a function of the magnitude of the wave number vector, i.e.

)(κωω = , we refer to the system as isotropic. If we cannot write the dispersion

relation in this way, the system is anisotropic. We now consider the surface in wave

number space given by Ckkk == ),,( 321ωω , where C is a constant; see Fig. 1.15,

where we display a two-dimensional example.

Fig. 1.15. Constant- frequency surface in wave number space.

Page 37: Dynamic Oceanography

37

The gradient ∇κω is always perpendicular to the constant frequency surface. From

(1.9.12) we note that this means that the group velocity is always directed along the

surface normal, as depicted in Fig. 1.15. Since the phase velocity is directed along the

wave number vector, e.g. (1.9.10), we realize that if the phase speed and group

velocity should become parallel, then the constant frequency surface must be a sphere

in wave number space. Mathematically, this means that )(κωω = , i.e. we have an

isotropic system.

We now return to the discussion of the dispersion relation (1.9.6) for Rossby

waves. Since ),( κωω k= , the system is anisotropic. Hence, the group velocity and the

vectorial phase speed are not parallel. Equations (1.9.6), (1.9.10) and (1.9.11) yield

for the components of the phase speed and group velocity:

⎪⎭

⎪⎬

−=≡

−=≡

,

,

42)(

4

2

2)(

κβ

κω

κβ

κω

kllc

kkc

y

x

(1.9.13)

and

⎪⎭

⎪⎬

=∂∂

−=

∂∂

.2

,)(

4)(

4

22)(

κβωκ

βω

kll

c

lkk

c

yg

xg

(1.9.14)

We note that the x-component of the phase speed is always directed westwards (the x-

axis is assumed to be parallel to the equator). However, the energy propagation can

have an eastward or westward component, depending on the ratio l/k. For waves

propagating along latitudinal circles (l = 0), the energy always propagates eastward.

The magnitudes of the phase speed and the group velocity are easily obtained from

(1.9.13) and (1.9.14):

⎪⎭

⎪⎬

=

=

.

,

2

2

κβκκ

β

gc

kc

r

r

(1.9.15)

We note that || gcc rr≤ . For a typical Rossby wave propagating along equator we take

λ = 2π/k = 1000 km. With β = 11102 −× m−1s−1, the phase speed and period become c ≈

− 0.5 m s−1 and T = |2π/ω| ≈ 23 days, respectively. We realize that such Rossby

waves are long-periodic phenomena.

Page 38: Dynamic Oceanography

38

Since ω is negative (we have chosen k > 0), we can define a positive quantity γ

by:

ωβγ

21

−= . (1.9.16)

Equation (1.9.6) can then be written as

( ) 222 γγ =+− lk . (1.9.17)

For a prescribed constant frequency ω, the wave number components k and l are

obtained as points on a circle in the wave number plane. This circle has a radius γ and

is centred at (γ, 0); see Fig. 1.16. The figure also depicts the phase speed and the

group velocity obtained from (1.9.10) and (1.9.12).

Fig. 1.16. Sketch of the relation between the phase speed and the group velocity for

Rossby waves.

We now consider the reflection of Rossby waves from straight coasts in the

north-south direction. Here, we take that the east coast is situated at x = L, where the

x-axis is parallel to the equator; see Fig. 1.17.

Fig. 1.17. Reflection of Rossby waves from a straight coast.

Page 39: Dynamic Oceanography

39

As far as wave propagation is concerned, it is the group velocity that represents signal

velocity. Accordingly, for a receiver, or a coast, to experience a physical impulse, it

must have a component of the group velocity directed against it. Accordingly, the

incoming wave has a group velocity with an eastward x-component. This wave is

labelled with subscript 1, i.e.

))(exp( 1111)1( tylxkiA ωψ −+= . (1.9.18)

The reflected wave, labelled with subscript 2, must have a westward-directed group

velocity component. The reflected wave is written as

))(exp( 2222)2( tylxkiA ωψ −+= . (1.9.19)

At the east coast, the normal component of the velocity must vanish, i.e.

Lxyy ==+ ,0)2()1( ψψ , (1.9.20)

or

0)(exp)exp()(exp)exp( 2222211111 =−+− tyliLikAiltyliLikAil ωω . (1.9.21)

If this is to be valid for all y and all t, we must have that

⎪⎭

⎪⎬

−===

).exp()exp(,

,

2211

21

21

LikALikA

llωω (1.9.22)

Hence, the frequency and the wave number component in the y-direction are

conserved during the reflection process. This means that the reflected wave number

vector must describe the same circle as the incident one, with the same l; see Fig.

1.18.

Fig. 1.18. Sketch of the relationship between wave characteristics of incident and

reflected Rossby waves.

Page 40: Dynamic Oceanography

40

We notice that the group velocity satisfies the law of reflection in ray theory; that is

the angle with respect to the normal of the reflecting plane is the same for the incident

and for the reflected wave. We also note that an incident short wave is reflected from

an east coast as a long wave, while reflection from a west coast (interchange

subscripts 1 and 2 in Fig. 1.18), results in the transformation from a long wave to a

short wave. For this reason the east coast is said to be a source of long waves, while

the west coast is a source of short waves. Since short waves are subject to stronger

dissipation than long waves (dissipation is proportional to the square of the velocity

gradients), this may lead to an increased transition of mean momentum from waves to

currents at the west coasts of the oceans. This has been used (Pedlosky, 1965) as basis

for explaining why the ocean currents are more intense in such regions (western

intensification). We shall return to that question in section 2.

To see how the effect of a free surface modifies the Rossby wave that we have

studied up to now, we consider the linearized version of the vorticity equation

(1.1.22):

0)( =+++ vvuf yxt βζ . (1.9.23)

Eliminating the horizontal divergence yx vu + in this equation by using the linearized

form of (1.1.20), we find

0)( =+−− vHfuv ttyx βη . (1.9.24)

We now simplify the problem by assuming that the time scale for the Rossby wave is

so large (several days) that the pressure gradient force and the Coriolis force always

have time to adjust to a state of geostrophic balance. This means that the velocity field

in (1.9.24) can be approximated by

⎪⎭

⎪⎬⎫

−==

,/,/fgu

fgv

y

x

ηη

(1.9.25)

where f is a mean (constant) value of f. This relation is often called the quasi-

geostrophic approximation and requires that 22 f<<ω . By replacing f with f in

(1.9.24) and inserting from (1.9.25), we obtain

022 =−+∇ −txtH a ηηβη , (1.9.26)

where fgHa /)( 2/1= . By assuming solutions of the form

Page 41: Dynamic Oceanography

41

))(exp( tlykxi ωη −+∝ , (1.9.27)

the dispersion relation is obtained from (1.9.26):

222 −++−=

alkkβω . (1.9.28)

If the wavelength is much smaller than the Rossby radius ),( 222 −>> alk , the effect of

the surface elevation can be neglected, and we are back to our rigid lid case (1.9.6).

However, if the wavelength is large ),( 222 −<< alk , we find from (1.9.10) that

24

2

22

22)( )( ak

lkakc x κ

κββ

−=+

−≈ , (1.9.29)

2422

2)( )( akl

lkklac y κ

κββ

−=+

−≈ . (1.9.30)

In this case, the group velocity components become

2)( ac xg β−≈ , (1.9.31)

44

)( )(2 aklc yg κ

κβ

≈ . (1.9.32)

Since κa now is a small parameter, we realize that the energy mainly propagates

westward in this case. If the wave number vector is parallel to the equator (l = 0) we

have non-dispersive waves, i.e. 2)()( acc xg

x β−== , and 0)()( == yg

y cc .

1.10 Topographic Rossby waves

Let us assume that somewhere the relative vorticity is zero. From the theorem of

conservation of potential vorticity (1.1.25) with η ≡ 0, we find that a displacement

northwards, where f is increasing, generates negative (anti-cyclonic) relative vorticity.

However, we realize that the same effect can be achieved by a northward

displacement if f is constant and the depth H decreases northward. This gives rise to

the so-called topographic Rossby waves. Of course, the existence of such waves does

not require that the bottom does slope in one particular direction.

Topographic Rossby waves are only a special case of escarpment waves when the

bottom has a very weak exponential slope everywhere in the ocean. By letting

)exp(0 yHH α= , equation (1.8.6) reduces to

02 =⎟⎠⎞

⎜⎝⎛ −−′−′′ FfkkFF

ωαα . (1.10.1)

By assuming

Page 42: Dynamic Oceanography

42

)exp( yiAF κ= , (1.10.2)

insertion into (1.10.1) yields the complex dispersion relation

022 =−++ωακκα fkki . (1.10.3)

In general we may allow for a very weak change of wave amplitude in the direction

normal to the coast, i.e. we take κ in (1.10.2) to be complex:

γκ il += . (1.10.4)

By insertion into (1.10.3), the imaginary part leads to γ = −α/2 (when l ≠ 0). From the

real part of (1.10.3) we then obtain

4/222 α

αω++

=lk

fk . (1.10.5)

We note that the along-shore phase speed kc x /)( ω= is positive. This means that the

wave propagation in the along-shore direction is such that we have shallow water to

the right (in the northern hemisphere).

For a bottom that slopes gently compared to the wavelength (k >> α), we see

from the (1.10.5) that these waves are similar to short planetary waves propagating in

an ocean of constant depth. The expressions for the frequency ω are identical, e.g.

(1.9.6), if

fαβ −= . (1.10.6)

This similarity is often used in laboratory experiments in order to simulate planetary

effects. When l = 0, the equations are satisfied for γ = 0, i.e. constant amplitude

waves. Such waves propagate along the bottom contours with shallow water to the

right, and mimic planetary Rossby waves along latitudinal circles in an ocean of

constant depth.

1.11 Barotropic instability

Rossby waves in the ocean can be generated by atmospheric forcing, as will be

shown in section 2. They can also develop due to energy transfer from unstable ocean

currents. Assume that we have a stationary mean flow along a latitudinal circle

iyUvrr )(= , (1.11.1)

and that this initial state is in geostrophic balance, i.e. UfPy ρ−= . We disturb

(perturbate) the initial state such that the velocities and the pressure can be written

Page 43: Dynamic Oceanography

43

⎪⎭

⎪⎬

+==+=

).,,(~),,,(~

),,,(~

tyxpPptyxvv

tyxuUu (1.11.2)

Here we have assumed that the motion is two-dimensional, i.e. independent of z. If the

perturbations (denoted by ~) are sufficiently small, the equations governing the

motion can be linearized. With w = 0, i.e. constant depth and horizontal surface in the

model depicted in Fig. 1.1, equations (1.1.1) and (1.1.3) reduce to

⎪⎪⎪

⎪⎪⎪

=+

−=++

−=−′++

.0~~,~1~~~

,~1~~~~

yx

yxt

xxt

vu

pufvUv

pvfUvuUu

ρ

ρ

(1.11.3)

By introducing the stream function ψ for the perturbation, defined by

xy vu ψψ =−= ~,~ , and eliminating the pressure from the equation above, we find

0)()( 22 =′′−+∇+∇ xxHtH UyU ψβψψ . (1.11.4)

For U ≡ 0, (1.11.4) reduces to the equation for Rossby waves (1.9.4). We assume that

the solutions can be written as

))(exp()( ctxiky −=ϕψ . (1.11.5)

Furthermore, the ocean is restricted by two straight coasts, defining a channel in the x-

direction; see Fig. 1.19.

Fig. 1.19. Horizontal shear flow in an ocean with parallel boundaries.

Page 44: Dynamic Oceanography

44

At y = 0 and y = D the normal velocity must be zero, i.e. ψx = 0. Since this is

valid for all x along the boundaries, we must accordingly have that ψ = 0 at y = 0, D.

Insertion from (1.11.5) into (1.11.4) yields

02 =⎟⎠⎞

⎜⎝⎛ +

−−′′

−′′ ϕβϕ kcU

U , (1.11.6)

with boundary conditions

Dy ,0,0 ==ϕ . (1.11.7)

If β ≡ 0, equation (1.11.6) is reduced to the well-known Rayleigh equation for two-

dimensional perturbations of shear flow in a homogeneous, inviscid fluid.

We assume that the wave number k in (1.11.5) is real, while the phase speed c

and the amplitude function ϕ are generally complex, i.e.

⎭⎬⎫

+=+=

.,

ir

ir

iicccϕϕϕ

(1.11.8)

From (1.11.5) we see that if ci > 0, ψ will grow exponentially in time. The initial state

is then said to be unstable, and we have an energy transfer from the mean flow to the

disturbance, which in this case is a Rossby wave.

Since equation (1.11.6) for the perturbation amplitude ϕ has non-constant

coefficients and a singularity for U = c (the term containing the highest derivative

vanishes at locations in the fluid where U = c), it is in general quite complicated to

solve. We will not attempt to do so here. Instead of a detailed investigation of the

stability conditions, we will be content with the derivation of a necessary condition

for the occurrence of instability. First, we make the denominator in (1.11.6) real, i.e.

( )( )( )

0222 =⎟⎟

⎞⎜⎜⎝

⎛+

+−+−−′′

−′′ ϕβϕ kccU

iccUU

ir

ir . (1.11.9)

Now we multiply this equation with the complex conjugate amplitude function

ir iϕϕϕ −=∗ , and note that 0222 ≥=+=∗ ϕϕϕϕϕ ir . By integrating the resulting

equation from y = 0 to y = D, and applying the boundary conditions 0== ∗ϕϕ at y =

0, D, we find that

( )( )

( )( )( )

00

22

2

022

2222 =

+−−′′

+⎟⎟⎠

⎜⎜⎝

+−−−′′

++′ ∫∫D

iri

D

ir

r dyccU

Uicdy

ccUcUU

kϕβϕβ

ϕϕ . (1.11.10)

For this equation to be satisfied, both real and imaginary parts must be zero. It

suffices here to consider only the imaginary part, i.e.

Page 45: Dynamic Oceanography

45

( )( )

00

22

2

=+−

−′′∫D

iri dy

ccUU

cϕβ

. (1.11.11)

If ci > 0, i.e. the basic flow is unstable, the integral in (1.11.11) must be zero.

However, the integrand is positive, possibly except for the factor β−′′U in the

numerator. Hence, for the integral to become zero, β−′′U must change sign in the

fluid, which means that there must be at least one place where β=′′U . This

expresses a necessary (but not sufficient) condition for instability of the basic flow.

An equivalent statement is that the absolute vorticity of the basic flow, Ufabs ′−=ζ ,

must have an extreme somewhere in the ocean (where 0/ =dyd absζ ). This kind of

instability is related to the energy transfer from the kinetic energy of the mean flow

(via the velocity shear) to the perturbations, and is called barotropic instability.

Perhaps even more common in the ocean is baroclinic instability, where energy is

transferred from the potential energy of the basic state to the perturbations via the

density stratification. We will return to discuss baroclinic instability in section 3.

Finally, we mention that on an f-plane (β = 0), the necessary condition for

instability is that the basic flow profile must have a point of inflexion ( 0=′′U )

somewhere in the fluid. Also, the numerical value of the vorticity U ′ must have a

maximum where 0=′′U (the Rayleigh-Fjørtoft criterion). The last condition can be

found by applying that the real part of (1.11.10) is zero, together with (1.11.11) and ci

≠ 0.

1.12 Barotropic flow over an ocean ridge

We consider an ocean current that crosses a sub-sea ridge. For simplicity, we

assume that the ridge is infinitely long in the y-direction, and that the depth far away

from the ridge is constant (= H0); see Fig. 1.20,

Page 46: Dynamic Oceanography

46

Fig. 1.20. Model sketch.

The surface far upstream is horizontal and the constant upstream flow U0 is

geostrophic, i.e.

yP

fU s

∂∂

−=ρ1

0 . (1.12.1)

On an f-plane, which we assume here, this means that the surface pressure Ps of the

basic flow must vary linearly with y.

We study the effect of the ridge by applying the potential vorticity theorem

(1.1.25). In a steady state we have

0=∇⋅ Qvr . (1.12.2)

A possible solution of this equation is const.=Q , in the (x, y)-plane. We assume that

the effect of the ridge is not felt far upstream, i.e. 0/ HfQ = for all values of y when

−∞→x . Since Q is a conserved quantity, we then must have

0H

fHfQ =++

≡ηζ (1.12.3)

We introduce the shape of the ridge q(x), i.e.

)(0 xqHH −= . (1.12.4)

The typical width of the ridge is 2L; see Fig. 1.20. Inserting (1.12.4) into (1.12.3), and

assuming that 0/ =∂∂ y , we obtain

)(0

qHfvx −= η . (1.12.5)

Page 47: Dynamic Oceanography

47

In the literature one often finds discussions of this problem where the surface is

modeled as a rigid lid, i.e. 0≡η . In this case we notice from (1.12.5) that anti-

cyclonic vorticity is generated over the entire ridge. By integration we obtain

∫−

−=x

L

dqHfv ξξ )(

0

. (1.12.6)

For Lx −≤ , we find that v = 0. For Lx ≥ , we obtain

0

)(HfAvv −=∞= , (1.12.7)

where A is the cross-section of the ridge, given by

∫ ∫∞

∞− −

==L

L

dqdqA ξξξξ )()( . (1.12.8)

The corresponding streamlines are sketched in Fig. 1.21.

Fig. 1.21. Deflection of flow over a ridge when the surface is approximated by a rigid

lid.

The deflection angle θ of the streamlines on the lee side of the ridge is

⎟⎟⎠

⎞⎜⎜⎝

⎛=⎟⎟

⎞⎜⎜⎝

⎛ ∞=

000

arctan)(arctanHU

fAU

vθ . (1.12.9)

The solution (1.12.6) may not be realistic, since it (through geostrophic balance in the

x-direction) implies that xp ∂∂ / must be constant far downstream. This means that the

perturbation pressure becomes infinitely large when ∞→x . However, the solution

obtained here may be of interest close to the ridge.

In most geophysical applications the surface can move freely in the vertical

direction. This means that a fluid column can be stretched, and not just be equal to, or

Page 48: Dynamic Oceanography

48

shorter than the upstream value, as in our previous example. From the linearized

versions of (1.1.18)-(1.1.19), we obtain for the surface elevation that is induced by the

ridge:

vgfv

gfU

xxx +=20η . (1.12.10)

By insertion into (1.12.5), we find

xxx qHfv

avF

02

20

1)1( −=−− . (1.12.11)

Here 2/1000 )/(gHUF = is the Froude number, and fgHa /)( 2/1

0= is the Rossby

radius. In the ocean F0 << 1, while we typically have that a ∼ 103 km. It is then

reasonable to assume that L << a. Hence, on the left-hand side of (1.12.11), the first

term dominates. Accordingly, a particular solution, when F0 << 1, can be written

210

)( CxCqdxHfv p ++−= ∫ . (1.12.12)

The solution of the homogeneous equation is:

)/exp()/exp( 43)( axCaxCv h −+= . (1.12.13)

The coefficients C1, C2, C3, and C4 are determined from the boundary conditions. We

assume that the ridge is symmetric about x = 0, and we require that the velocity and

the pressure are continuous at x = 0. Furthermore, we assume that −∞→→ xv ,0 ,

and that the perturbation pressure becomes finite when ∞→x . The last condition is

equivalent to require vanishing v at infinity (in the case of no waves behind the ridge).

The solution becomes:

⎪⎪

⎪⎪

≥−−−

≤−

=

∫∞−

x

x

xdqHfax

HfA

xdqHfax

HfA

v

000

00

.0,)()/exp(12

,0,)()/exp(2

ξξ

ξξ (1.12.14)

Here A is the cross-sectional area defined by (1.12.8). This solution yields 0)0( =v .

For x < −L, we find that 0)2/()/exp( 0 >= HaxfAv , while for x > L we have that

0)2/()/exp( 0 <−−= HaxfAv . For a symmetrical ridge, v becomes anti-symmetric

about the centerline of the ridge. The streamlines in this case are sketched in Fig.

1.22.

Page 49: Dynamic Oceanography

49

Fig. 1.22. Deflection of flow over a symmetric ridge when the surface is free to move.

In this case we only have a local disturbance of the current in the area close to the

ridge. This is also obvious from the conservation of potential vorticity (1.12.3):

00 H

fqH

f=

+−+

ηζ . (1.12.15)

We have stretching of a fluid column, and an associated generation of cyclonic

vorticity just before, and just after the ridge. Over the higher parts of the ridge the

column becomes compressed, and here anti-cyclonic vorticity is generated; see the

schematic sketch in Fig. 1.23.

Fig. 1.23. Conservation of potential vorticity for a fluid column when the upstream

fluid depth H0 is independent of the y-coordinate.

In the present case we will always have an elevation of the surface over the ridge, as

sketched in Fig. 1.23. By inserting from (1.12.5) into (1.12.14), we find that

02

)0( >=aAη . (1.12.16)

Page 50: Dynamic Oceanography

50

The elevation computed here is caused by the effect of the earth’s rotation. In the

general case where the Froude number is not small, we find from (1.12.5) and

(1.12.10) that

qFa

qF

FFa xxxx )1(

11)1(

12

022

0

20

20

2 −−

−−=

−− ηη . (1.12.17)

Solutions of this equation may yield surface elevations or surface depressions above

the ridge, depending of the form of q, and the magnitudes of a and F0. We will not

discuss the general solution here. However, for a non-rotating fluid (f = 0, i.e. a→∞),

we find from (1.12.17) that

)(1 2

0

20 xqF

F−

−=η . (1.12.18)

This is a well-known result in fluid mechanics. It states that for sub-critical currents

(F0 < 1), we have 0)0( <η above a ridge. For super-critical currents (F0 > 1), we find

0)0( >η . Examples of both these cases are found for flow over bottom rocks in

shallow rivers.

Stationary Rossby waves behind ridges

For large-scale motion over wide ridges we must take into account the fact that

the Coriolis parameter varies with the latitude, i.e. yff β+= 0 . Then stationary, or

standing, planetary Rossby waves can be generated by the presence of the ridge. In

general, the condition for the existence of a standing wave is that it must have a phase

speed that is equal in magnitude, and oppositely directed to the basic current. The

direction of the energy flow (basic current plus group velocity), then decides on which

side of the ridge we find the standing wave, because non-damped wave motion cannot

exist without continuous supply of energy to the wave region. In our case, the phase

of the Rossby wave moves westward, while the energy moves eastward. This apply

to planetary Rossby waves where the wavelength λ and the Rossby radius a satisfy

the relation aπλ 2< ; see the dispersion relation (1.9.28). We then realize

immediately that standing Rossby waves may occur on the lee-side of the ridge (here:

for x > 0), if the current is directed from the west towards the east. Furthermore, it is

obvious that we cannot have standing Rossby waves at all when the current is directed

towards the west. In the atmosphere, standing, mountain-generated Rossby waves in

Page 51: Dynamic Oceanography

51

the westerlies may be responsible for observed distributions of pressure ridges and

pressure troughs; see Charney and Eliassen (1949).

Alternatively, on an f-plane, we may find standing topographic Rossby waves on

the lee-side of the ridge in our example (Fig. 1.20) if the ocean bottom slopes

upwards in the y-direction (remember: decreasing H gives the same effect as

increasing f in the potential vorticity, e.g. sec. 1.10). If )exp(0 yHH α= , where α < 0,

the frequency for short Rossby waves is 0/ <= kfαω , e.g. (1.10.5). As mentioned

before, a standing wave requires a phase speed that is equal, and oppositely directed

to the basic flow, i.e.

0Uk=

ω . (1.12.19)

Accordingly, the wavelength of standing topographic Rossby waves becomes

2/1

02 ⎟⎟⎠

⎞⎜⎜⎝

⎛=

fUα

πλ . (1.12.20)

This scale is relevant for laboratory simulations of Rossby waves in rotating tank

experiments.

When ocean currents are flowing over isolated sub-sea mountains or banks, anti-

cyclonic vorticity may be generated over the top of the mount. If the vorticity is

sufficiently strong, as it can be for high mounts, a permanent vortex, or a Taylor

column (Taylor, 1922) may be formed. Measurements from the Halten Bank west of

Norway (Eide, 1979), suggest the presence of such vortexes, or eddies. This means

that the water mass above the bank may have a rather long residential time. But bank

regions are important for the early growth of fish larvae. Therefore, in the case of

accidental oil spills, or other kind of pollution, the long residence time of the water

mass means that the fish larvae will stay in contact with the polluted water for quite

some time.

2. WIND-DRIVEN CURRENTS AND OCEAN CIRCULATION

2.1 Equations for the mean motion

Mass and momentum

The velocity introduced in Section 1 was related to the displacement of individual

Page 52: Dynamic Oceanography

52

fluid particles. However, many oceanic flow phenomena are turbulent, which means

that it is difficult to keep track of individual particles. But turbulent flows are not

without structure. Even if individual particles move in a chaotic way, they may in an

average sense define an orderly flow. In the ideal case the mean flow has a much

larger characteristic time scale than the rapid fluctuations of individual fluid particles.

Then we can define mean velocities by a time averaging process:

∫∆+

∆−∆=

2/

2/

1 Tt

Tt

dtvT

v rr . (2.1.1)

Here ∆T represents a period that is large compared to the characteristic scale of the

turbulence, but small compared to the typical time scale of the mean motion we intend

to study. The same type of averaging must be done for the pressure and the density.

Accordingly we write

⎪⎭

⎪⎬

′+=

′+=

′+=

.,

,

ppp

vvvρρρ

rrr

(2.1.2)

where v ′r , ρ′ , p′ represent the fluctuating (turbulent) part of the motion.

The continuity equation (1.1.2) expresses the conservation of mass in a fluid, and

is a fundamental law in non-relativistic mechanics. Obviously, molecular processes

like heat and salt diffusion may change the density of a fluid particle. This effect is

accounted for by the application of the equation of state for the fluid. The change of

density has a very important dynamical implication in that it changes the buoyancy of

a fluid particle, which in turn may lead to thermohaline circulations in the system.

Equation (1.1.2) is derived under the assumption that only advective fluxes (transport

of mass by the velocity field) can change the mass within a fixed, small geometrical

volume. This assumption is essential for the definition of the velocity field in the

fluid.

It is a fact that for most low-frequency phenomena in the ocean, the individual

change of density for a fluid particle is very small, and can be neglected in the

continuity equation. Hence, we take that

0=⋅∇ vr , (2.1.3)

i.e. the velocity field in the ocean is approximately non-divergent. By averaging this

equation and applying (2.1.2), we find that

Page 53: Dynamic Oceanography

53

⎭⎬⎫

=′⋅∇=⋅∇

.0,0

vvr

r

(2.1.4)

By separating the variables as in (2.1.2), and applying (2.1.4) we find from averaging

the viscous Navier-Stokes equation in a reference system rotating with constant

angular velocity f/2 about the vertical axis, that

[ ]vvvkgpvkfvvvt ′′−∇⋅∇+−∇−=×+∇⋅+rrrrrrrrr ν

ρ1 . (2.1.5)

We have neglected ρ′ compared to ρ in the pressure term, since typically in the ocean 310~/ −′ ρρ . Furthermore, ν is the kinematic molecular viscosity coefficient (ν

≈10−2 cm2s−1). The term vr∇ν is usually much smaller than the turbulent momentum

transport per unit mass, vv ′′−rr , and can therefore be neglected. The term vv ′′−

rrρ is

usually referred to as the turbulent Reynolds stress tensor.

We first consider the case of isotropic turbulence, i.e. turbulence that has no

preferred directions, and introduce a constant kinematic eddy viscosity A0. Analogous

to the stress tensor for a Newtonian fluid, we assume for the turbulent Reynolds stress

tensor per unit density:

3,2,1,,32

0 =−⎥⎥⎦

⎢⎢⎣

∂+

∂∂

=′′− jiKxv

xvAvv ij

i

j

j

iji δ . (2.1.6)

Here

⎩⎨⎧

≠=

=,,0,,1

jiji

ijδ and ∑=

′′=3

121

iiivvK .

The quantity K is the kinematic energy per unit mass of the turbulent fluctuations. The

sum of the diagonal, turbulent Reynolds stress components (i = j) must always equal

2K. We note from the right-hand side of (2.1.6) that this is true for our choice of

parameterisation, since 0=⋅∇ vr . The term 3/2 Kρ formally acts as an “isotropic”

pressure. This term can be combined with the mean pressure. In the momentum

equation we then write

*1)

32(1

321 pKpKp ∇=+∇=∇+∇

ρρ

ρρ, (2.1.7)

neglecting the variation of ρ in connection with K. If we assume that A0 is much

larger than the molecular viscosity coefficient, insertion from (2.1.6) and (2.1.7) into

(1.2.5) yields

Page 54: Dynamic Oceanography

54

kgvApvkfvvvt

rrrrrrr−∇+∇−=×+∇⋅+ 2

0*1ρ

. (2.1.8)

Here 2222222 /// zyx ∂∂+∂∂+∂∂=∇ is the three-dimensional Laplacian operator.

For large-scale ocean circulation it is usual to assume that the turbulence is

dependent on direction, and we define AH and A as turbulent eddy viscosity

coefficients in the horizontal and the vertical directions, respectively. The surfaces of

constant density are mainly horizontal in the ocean, and the stratification is stable.

Hence the turbulence activity in the vertical direction will be smaller than in the

horizontal direction. Therefore one usually assumes that AH >> A. For such motion zw

is often negligible, i.e. 0≈+ yx vu . Furthermore, we typically have 222 wvu ′>>′+′ .

Accordingly, we can assume for the normal stresses:

⎪⎭

⎪⎬

≈′′−−=′′−−=′′−

,0,2,2

wwKvAvvKuAuu

yH

xH

(2.1.9)

while the tangential stresses become:

⎪⎭

⎪⎬

+=′′−=′′−+=′′−=′′−+=′′−=′′−

.,

),(

yHz

xHz

yxH

wAvAvwwvwAuAuwwu

uvAuvvu (2.1.10)

Furthermore, we assume that AH is constant, while we permit that )(zAA = . The

horizontal component of (2.1.5) can then be written

zzHHHt qAqApqkfqvq )(1 2*

rrrrrrr+∇+∇−=×+∇⋅+

ρ, (2.1.11)

where jviuqrrr

+≡ , 22222 // yxH ∂∂+∂∂=∇ , and Kpp ρ+=* .

For large-scale ocean circulation we can assume that the vertical pressure

distribution is approximately hydrostatic. This means that we can replace the z-

component in (2.1.5) by

∫ +=η

ρz

S tyxPdzgp ),,( , (2.1.12)

where PS is the atmospheric pressure at the sea surface.

Heat and salt

The internal energy per unit mass of sea water, e, can be written approximately as

Page 55: Dynamic Oceanography

55

Tce v= , (2.1.13)

where cv is the specific heat at constant volume, and T the absolute temperature. By

considering a geometrically fixed volume τ with surface σ, we find for the change of

internal energy with time within τ that

∫∫∫ ⋅−⋅−=∂∂

σστ

σδσδρδτρ rrrrTFvee

t)( , (2.1.14)

where the vectorial area σδ r is normal to the surface and points outwards from the

considered fluid volume. The two terms on the right-hand side represent the advective

and diffusive surface heat fluxes, respectively, into the volume. We assume that the

diffusive energy flux, TFr

, is given by Fourier's law:

TkF TT ∇−=r

. (2.1.15)

Here kT is the coefficient of heat conduction. Applying Gauss’ theorem, we obtain

from (2.1.14)

0)()()( =⎥⎦⎤

⎢⎣⎡ ∇⋅∇−⋅∇+∂∂

∫ δτρρτ

TkvTcTct Tvv

r . (2.1.16)

This equation is valid for an arbitrary volume τ in the fluid. Accordingly, the

integrand must be zero everywhere. Furthermore, by taking that ρcv and kT are

constants, which are reasonable assumptions, and using (2.1.3), we find from (2.1.16):

TTvtT

T2∇=∇⋅+

∂∂ κr . (2.1.17)

Here )/( vTT ck ρκ = is called the thermal diffusivity. Equation (2.1.17) is often

referred to as the heat equation.

Equivalently, the change of salt with time within a geometrically fixed volume τ

is given by

∫ ∫ ∫ ⋅−⋅−=∂∂

τ σ σ

σδσδρδτρ rrrrSFvSS

t)( , (2.1.18)

where the terms on the right-hand side represent the advective and diffusitive surface

salt fluxes, respectively, into the fluid volume. We assume that the diffusitive salt

flux, SFr

, is given by Fick’s law

SkF SS ∇−=r

, (2.1.19)

where kS is the diffusion coefficient for salt in water. As for heat, we now obtain for

salt from (2.1.18)

Page 56: Dynamic Oceanography

56

0)()()( =⎥⎦⎤

⎢⎣⎡ ∇⋅∇−⋅∇+∂∂

∫ δτρρτ

SkvSSt S

r . (2.1.20)

If we assume that ρ and kS are approximately constants in this equation, and apply

(2.1.3), we find that

SSvtS

S2∇=∇⋅+

∂∂ κr . (2.1.21)

Here ρκ /SS k= is called the salt diffusivity. Equation (2.1.21) is often referred to as

the salt equation.

The diffusion coefficients κT and κS are determined by the molecular structure of

the fluid. They vary in fact with the temperature and the salinity, but this variation is

small, and we will regard them as constants. Typical values for seawater are κT ≈

10−3cm2s−1 and κS ≈ 10−5cm2s−1.

We define TTT ′+= and SSS ′+= . Furthermore, we introduce turbulent eddy

diffusivities K(T) and K(S) for temperature and salt. The eddy diffusivities are generally

much larger than the molecular diffusivities in (2.1.17) and (2.1.21). Analogous to the

molecular fluxes from Fourier’s and Fick’s laws, we now assume for the turbulent

transports

⎪⎭

⎪⎬⎫

=′′−=′′−=′′−=′′−=′′−=′′−

.,,,,,

)()()(

)()()(

zS

yS

HxS

H

zT

yT

HxT

H

SKSwSKSvSKSuTKTwTKTvTKTu

(2.1.22)

As before, subscript H denotes horizontal direction. We assume that the horizontal

eddy diffusivities are constant. Hence the heat and salt equations become, when we

neglect the effect of molecular diffusion:

zzT

HT

Ht TKTKTvT )( )(2)( +∇=∇⋅+r , (2.1.23)

zzS

HS

Ht SKSKSvS )( )(2)( +∇=∇⋅+r . (2.1.24)

In fully developed turbulent motion the exchange of heat and the exchange of salt

must be approximately similar, i.e.

⎭⎬⎫

≈≈

.,

)()(

)()(

ST

SH

TH

KKKK

(2.1.25)

However, the mechanisms for momentum exchange in turbulent flows may be

different from those related to heat/salt, so that we generally have AH ≠ KH and A ≠ K.

In the upper mixed layer in oceans and fjords the motion sometimes seems to be

quasi-turbulent. In such cases we find that K(T) > K(S) (Aas, private communication).

Page 57: Dynamic Oceanography

57

To get a closed system, the equation of state must be specified. The general form

is

),,( pSTF=ρ . (2.1.26)

For seawater, F is a very complicated function. Usually we can replace the variables

in F with its mean values, i.e.

( )pSTF ,,=ρ . (2.1.27)

We have now seven equations, (2.1.4), (2.1.11), (2.1.12), (2.1.23), (2.1.24), (2.1.27)

for the seven unknowns, STpwvu ,,,,,, ρ , which vary in space and time. When we in

the following use some of these equations to describe the mean motion, the mean sign

will be left out for simplicity. We also replace *p by p , i.e. we neglect the

contribution from the turbulent Reynolds stresses to the mean pressure.

Along the sea surface we have wind stress components )()( , yS

xS ττ and a variable air

pressure ),,( tyxPS that act on the ocean. The wind stress components are related to

the wind speed (U10, V10) at 10 meters through the semi-empirical relations

⎭⎬⎫

+=+=

.)(,)(

102/12

102

10)(

102/12

102

10)(

VVUcUVUc

Day

S

Dax

S

ρτρτ

(2.1.28)

where ρa and cD are the density of air (≈1.2×10−3g cm−3) and the drag coefficient,

respectively. Attempts have been made to assess the value of cD from measurements,

and the results vary considerably. Values between 1×10−3 and 3×10−3 are reported in

the literature, where higher values correspond to increasing wind speeds.

2.2 The Ekman elementary current system

We consider stationary motion in a quasi-homogeneous ocean. By quasi-

homogeneous we mean that the generally varying density in the ocean can be

approximated by a constant value in these studies. The ocean extends to infinity in the

horizontal directions, and has constant depth H. We assume that the ratio between the

convective acceleration and the Coriolis force, defining the Rossby number, R, is

small, i.e.

1/ <<×∇⋅= HH vkfvvR rrrr . (2.2.1)

This allows us to disregard the acceleration term in the momentum equation. We also

assume that horizontal turbulent diffusion of momentum can be neglected. Equation

(2.1.11) for the mean variables (skipping the over-bar) then reduces to

Page 58: Dynamic Oceanography

58

⎪⎪⎭

⎪⎪⎬

+−=

+−=−

.)(1

,)(1

zzy

zzx

Avpfu

Aupfv

ρ

ρ (2.2.2)

By assuming constant pressure gradients and constant A in the entire fluid, (2.2.2) has

an exact solution (e.g. Krauss, 1973, p. 248). However, friction is only important in

relatively thin layers close to the surface and the bottom. These are the so-called

Ekman layers, with thickness δS and δB, respectively. In the following we let the

indices S and B signify the upper and lower Ekman layer, respectively. In these layers

we assume that the eddy viscosities AS and AB are constant, but differing in numerical

value. Furthermore, we assume that H >> δS, δB, implying that we have a large

interior where the effect of friction can be completely neglected. Here we have

geostrophic balance. Assuming that the horizontal pressure gradient is known, i.e.

SHHH Pgp ∇+∇=∇ ηρ from (2.1.12), the geostrophic velocity Gvr is given by

⎪⎪⎭

⎪⎪⎬

−=

−=−

.1

,1

yG

xG

pfu

pfv

ρ

ρ (2.2.3)

We assume here that f is constant. Then we note from (2.2.3) that the geostrophic flow

is non-divergent:

0=+ Gy

Gx vu . (2.2.4)

The total steady flow is subdivided into an Ekman part and a geostrophic part:

),(),,( yxvzyxvv GE rrr+= . (2.2.5)

At the surface the tangential stress in the fluid must equal the wind stress given by

(2.1.28). This leads to

0.

,

)(

)(

=⎪⎭

⎪⎬

=∂∂

=∂∂

z

zvAzuA

ySS

xSS

τρ

τρ. (2.2.6)

In general, we will allow for a slipping motion at the bottom. For simplicity we

assume that the frictional stress at the bottom is proportional to the bottom velocity

(Rayleigh friction). Hence

Page 59: Dynamic Oceanography

59

⎪⎭

⎪⎬

=∂∂

=

=∂∂

=

,

,

)(

)(

BB

By

B

BB

Bx

B

vrz

vA

urz

uA

ρρτ

ρρτ ,Hz −= (2.2.7)

where r is a friction coefficient and GB

EBB vvv rrr+= . Formally, (2.2.7) yields the

extreme conditions “no-slip”, or 0→Bvr in the limit ∞→r , and “free-slip”, or

0/ =∂∂ zvBr

when r = 0.

We start out by studying the Ekman layer at the bottom, and we define a

boundary-layer coordinate z~ such that

BzHz δ/)(~ += , (2.2.8)

where 2/1)/2( fABB πδ = is the thickness of the Ekman layer at the bottom. By

introducing the complex velocity EB

EB ivuW += , the equation for the Ekman current

in the bottom layer becomes

02~~ 2 =− iWzzW π . (2.2.9)

The boundary conditions at bottom can be written

⎪⎪⎭

⎪⎪⎬

+=∂+∂

⎟⎟⎠

⎞⎜⎜⎝

+=∂+∂

⎟⎟⎠

⎞⎜⎜⎝

.~)(

,~)(

GEB

GEB

B

B

GEB

GEB

B

B

vvz

vvrA

uuz

uurA

δ

δ 0~ =z , (2.2.10)

Furthermore, above the bottom layer, we must have that GB vv rr→ . Accordingly, in

our new coordinate:

∞→→ zv EB

~,0r. (2.2.11)

A solution of (2.2.9) that satisfies (2.2.11), can be written

~)1(exp ziCW +−= π , (2.2.12)

where C = C r+ iCi is complex. By insertion into (2.2.10), we find

( )

( ) ⎪⎭

⎪⎬

+−++

=

++++

−=

,)1(122

1

,)1(122

1

2

2

GGi

GGr

vKKuKK

C

KvuKKK

C (2.2.13)

where rfAK B /)2/( 2/1= . Now, the total velocity in the bottom Ekman layer can be

written

( )( ) ⎭

⎬⎫

−−+=−++=

).~exp(~sin~cos),~exp(~sin~cos

zzCzCvvzzCzCuu

riG

B

irG

B

ππππππ

(2.2.14)

Page 60: Dynamic Oceanography

60

For simplicity, we take that 0=Gu , and let r→∞ (or K→0), i.e. we apply a no-slip

condition at the bottom. We then find

⎪⎭

⎪⎬⎫

−−=

−−=

)).~exp(~cos1(),~exp(~sinzzvv

zzvuG

B

GB

ππ

ππ (2.2.15)

From this result we see that the flow in the bottom Ekman layer (the Ekman spiral) is

veering to left of the direction of the geostrophic current as we move towards the

bottom. For the other extreme case at the bottom, when K→∞ (free slip), we find that

Cr = Ci = 0, and accordingly GB vv rr= . We realize that we need a frictional force at the

bottom to obtain an Ekman spiral. By choosing AB = 10 cm2s−1 and AB = 100 cm2s−1 as

examples, we find that δB becomes 14 m and 44 m, respectively. With r = 0.24 cm s−1,

which often is used for shallow oceans, we find that K = 0.1 and 0.3. It could be

realistic to expect that K ≤ 0.3 in the ocean.

Taylor (1916) solved the analogous problem for the atmosphere. He assumed that

the frictional stress at the ground was proportional to the square of the wind speed,

instead of using the simplified model with linear friction, as assumed here.

In the upper Ekman layer we define a boundary-layer coordinate

Szz δ/=′ , (2.2.16)

where 2/1)/2( fASS πδ = . By introducing the complex velocity ES

ES ivuW += for the

Ekman flow, we now find that

02 2 =−′′ iWW zz π . (2.2.17)

The boundary conditions (2.2.6) can now be written

,0,

,

)(

)(

=′

⎪⎪⎭

⎪⎪⎬

=′∂

=′∂

zT

zv

Tzu

yES

xES

(2.2.18)

where

)(2/1

)()(2/1

)( 2,2 yS

S

yxS

S

x

fAT

fAT τ

ρπτ

ρπ

⎟⎟⎠

⎞⎜⎜⎝

⎛=⎟⎟

⎞⎜⎜⎝

⎛= . (2.2.19)

We have assumed that H >> δS. Accordingly, for the Ekman current

−∞→′→ zvu ES

ES ,0, . (2.2.20)

The equations (2.2.17)-(2.2.20) define the traditional Ekman spiral for a deep ocean.

The solution is

Page 61: Dynamic Oceanography

61

[ ][ ] ⎪

⎪⎬

′′+−′−−=

′′−+′+=

).exp(sin)(cos)(21

),exp(sin)(cos)(21

)()()()(

)()()()(

zzTTzTTv

zzTTzTTu

yxyxES

yxyxES

ππππ

ππππ (2.2.21)

This solution describes the oceanic surface Ekman spiral veering to the right (in the

northern hemisphere) compared to the wind direction. The surface current is deflected

45° to the right of the wind. The total current in the surface layer becomes

⎭⎬⎫

+=+=

.,

GESS

GESS

vvvuuu

(2.2.22)

The oceanic flow described in this section is driven by wind stresses and pressure

gradients, and is referred to as Ekman’s elementary current system. The pressure

gradients induce a geostrophic flow, which is felt from the top to the bottom

(independent of z). Close to the ocean bottom this flow is modified by friction in the

Ekman layer. As result, the flow here is veering to the left of the geostrophic flow

direction. The wind stress is only felt in the surface Ekman layer, inducing a current

that is veering to the right of the wind direction.

2.3 The Ekman transport

The volume transport in the lower Ekman layer is defined by

⎪⎪

⎪⎪

==

==

∫∫

∫∫+−

+−

−1

0

1

0

,~

,~

zdvdzvV

zdudzuU

BB

H

HBB

BB

H

HBB

B

B

δ

δ

δ

δ

(2.3.1)

where z~ is defined by (2.2.8). As before, we take 0=Gu . By inserting from (2.2.14)

into (2.3.1), we find

⎪⎪⎭

⎪⎪⎬

⎥⎦

⎤⎢⎣

⎡++

−=

+++

−=

,)122(2

11

,)122(2

12

2

2

GBB

GBB

vKK

V

vKK

KU

δπ

δπ

(2.3.2)

where we have utilized that 1)exp( <<−π . The volume transport is deflected to the

left of the geostrophic flow in the interior. We define a deflection angle θB as

B

BB V

Uarctan=θ . (2.3.3)

For K = 0 (no-slip), θB = 11.3°, while for K = 0.3 (corresponding to AB = 100 cm2s−1),

Page 62: Dynamic Oceanography

62

we find θB = 9.4°. Therefore, it is reasonable to assume that the volume transport in

the bottom Ekman layer is deflected about 10° to the left of the geostrophic flow. We

note that this result is derived by using a constant AB in the Ekman layer.

In the upper layer, the Ekman transport is independent of the variation of the

vertical eddy viscosity. Applying constant pressure in (2.2.2) and integrating over the

boundary layer, while using the boundary conditions (2.2.6), we find that

⎪⎪⎭

⎪⎪⎬

−==

==

)./(

),/(

)(0

)(0

fdzvV

fdzuU

xS

ES

ES

yS

ES

ES

S

S

ρτ

ρτ

δ

δ (2.3.4)

We notice the well-known result that the Ekman transport in the surface layer is

deflected 90° to the right of the wind direction (in the northern hemisphere). The

result (2.3.4) is based on the fact that the viscous stresses in the fluid are

approximately zero for z ≤ −δS, which can be seen to be true from (2.2.21). However,

we must remember that (2.2.21) was derived by assuming a constant eddy viscosity in

the upper layer, while this is not done for the Ekman transport. When the eddy

viscosity varies in the upper layer, a precise definition of δS in (2.3.4) is not entirely

obvious. However, we may take that 2/1)/2( fASS πδ = , where SA is some kind of

average (bulk) value for the upper layer.

2.4 Divergent Ekman transport and forced vertical motion (Ekman suction)

We assume that the geostrophic flow and the wind stress vary slowly in space

and time. By integrating the continuity equation (2.1.4) for the mean flow vertically

across the bottom Ekman layer, we find

y

Vx

Uw BBB ∂

∂−

∂∂

−= , (2.4.1)

where wB is the vertical velocity at BHz δ+−= . Furthermore, we have utilized that

w(−H) = 0 (horizontal bottom). By inserting (2.2.14) into (2.3.1), we can calculate UB

and VB. To simplify, we assume that K = 0. Applying (2.2.4), we then finally obtain

from (2.4.1)

GBGBB vkw ζ

πδ

πδ

2)(

2=×∇⋅=

rr, (2.4.2)

where Gζ is the geostrophic vorticity, and we have neglected exp(−π) in comparison

Page 63: Dynamic Oceanography

63

with 1.

A similar integration of the continuity equation in the upper Ekman layer yields

)(1S

SSS k

fyV

xU

w τρ

rr×∇⋅=

∂∂

+∂∂

= , (2.4.3)

where the final result follows from (2.3.4). Here wS is the vertical velocity at Bz δ−= .

Furthermore, we have neglected any vertical motion of the sea surface, i.e. we have

taken w(0) = 0. The situation is sketched in Fig. 2.1.

Fig. 2.1. Schematic model of Ekman suction.

As an example, we let a wind field with maximum wind speed of 40 knots (≈ 20 m

s−1) vary over a typical distance L = 100 km. From (2.1.28) we then estimate that τS

/(ρL) ~ 10−6cm s−2. With f = 10−4s−1, equation (2.4.3) yields wS ~ 0.01cm s−1. This, in

turn, means that this wind field must be “turned on” for 11½ days to transport a

particle 100 m vertically. In comparison, we can estimate the turbulent vertical

diffusion velocity wturb. For a typical vertical length scale H and a time scale T, the

diffusive terms balance; see for example (2.1.11), when H/T ~ A/H. This defines a

turbulent diffusion velocity scale, i.e. wturb ~ A/H. With A = 100 cm2s−1 and H = 100

m, we find that wturb ~ 10−2 cm s−1, which is the same order of magnitude as for the

forced vertical suction velocity in this example.

Let us now assume that the wind stress is zero. Then wS = 0 from (2.4.3). In the

interior of the fluid, where we can disregard the effect of friction, the linearized

vorticity equation from (1.1.18)-(1.1.19) for non-divergent flow on an f-plane reduces

to

zt wf=ζ . (2.4.4)

Vertical integration in the interior yields

Page 64: Dynamic Oceanography

64

BH

t wfdzS

B

−=∫−

+−

δ

δ

ζ , (2.4.5)

since we have assumed that wS = 0. Because of the vertical transport of fluid particles

in the interior, the geostrophic vorticity in this region changes slowly in time, i.e.

),()( yxtFG ζζ = . (2.4.6)

We assume that ζζζ tGtt F=≈ in equation (2.4.5). By utilizing that H >> δS, δB, we

approximately have from (2.4.5):

GBB

Gt

fwfH ζπδζ

2−=−= , (2.4.7)

where the last result follows from (2.4.2). Inserting from (2.4.6):

FH

fF Bt π

δ2

−= . (2.4.8)

This equation has a simple exponential solution. The slowly-changing geostrophic

vorticity can then be written

)2

exp(),( tH

fyx BG

πδζζ −= . (2.4.9)

We notice that the vorticity decreases in time. This phenomenon is referred to as spin-

down, and is due to the vertical transport of fluid particles caused by divergence in the

bottom Ekman layer. In this way the friction, which only is important in a relatively

thin region close to the bottom, may effectively influence fluid motion in the interior.

If we take AB = 10 cm2s−1 and H = 100 m, we find a spin-down time T = 2πH/(fδB) ≈

6 days from (2.4.9) With A = 10 cm2s−1 in the entire fluid, it will take about T ~ H2/A

~ 100 days before the friction in the interior effectively damps the motion.

The spin-down phenomenon also follows readily from Kelvin's circulation

theorem. In the interior of the fluid, where the effect of friction can be neglected, we

can apply (1.1.26) and (1.1.27), i.e.

∫ ∫∫∫∫Γ

=+=×∇⋅=⋅ const.)()(σσ

δσζδσδ Gabsabs fvkrv rrrr (2.4.10)

where σ is the area within the closed material curve Γ in the horizontal plane. The

vertical flux from the bottom Ekman layer must lead to horizontal motion in the

interior; see the sketch in Fig. 2.2.

Page 65: Dynamic Oceanography

65

Fig. 2.2. A material, closed curve at consecutive times t1 and t2.

In the present example, with a cyclone )0( >Gζ in the interior, leading to

convergence in the Ekman layer, a closed material curve in the interior horizontal

plane must be extended, as depicted in Fig. 2.2. Accordingly, the area σ in (2.4.10)

becomes enlarged, and therefore the geostrophic vorticity must decrease, which

means spin-down of the cyclone.

We realize immediately that if we have an anti-cyclone )0( <Gζ in the interior,

this will result in a divergent mass transport in the bottom layer, and hence

contraction of material curves in the interior. According to (2.4.10), the absolute

vorticity then must increase. This increase is achieved by the (negative) Gζ

approaching zero, i.e. a spin-down of the anti-cyclone.

For a variable wind stress, a balance in the interior can be achieved if wB = wS.

From (2.4.2) and (2.4.3) we then obtain

)(2S

B

G kf

τδρπζ rr

×∇⋅= . (2.4.11)

From our previous example with τS /(ρL) ~ 10−6 cm s−2, we find that this balance

results in ~Gζ 4.5×10−5s−1.

2.5 Variable vertical eddy viscosity

Because of mixing due to the action of wind, breaking of waves etc., the level of

turbulence close to the surface is higher than in the interior of the ocean. This means

that the vertical eddy viscosity is not a constant, but depends on depth. Let us make a

simple study of how the effect of a vertical variation of A influences the Ekman flow

in the surface layer in a homogeneous ocean. For that purpose we consider a linear

depth dependence:

Page 66: Dynamic Oceanography

66

zAA α+= 0 , (2.5.1)

where α can be positive or negative. In this way we model an A that is decreasing or

increasing with depth, respectively. We rescale the depth by

0/~ δzz = , (2.5.2)

where 2/100 )/2( fAπδ = . Accordingly we can write

)~1(0 zAA ε+= . (2.5.3)

Here ε is a dimensionless parameter given by

2/1

0

2⎟⎟⎠

⎞⎜⎜⎝

⎛=

fAαπε . (2.5.4)

From (2.2.2), with p independent of x and y, we find that

02)~1( 2~~~ =−++ iWWWz zzz πεε , (2.5.5)

where W = u + iv. We assume a constant wind stress in the y-direction. The boundary

conditions then become

0~,i~ == zTWz , (2.5.6)

where )/( 00)( AT y

S ρδτ= . In the deep ocean we assume that

−∞→→ zW ~,0 . (2.5.7)

We will assume that ε is a small parameter (|ε| << 1), allowing us to expand the

solution in a power series after ε. Accordingly

K+++= )2(2)1()0( WWWW εε (2.5.8)

We now insert this series into the governing equation (2.5.5) and the boundary

conditions (2.5.6) and (2.5.7). Then we collect terms with the same power of ε. Since

this is an arbitrary small parameter, our equations can only be fulfilled if the

coefficients multiplying each power of ε vanish identically.

The terms multiplying ε0 yield, when set to zero:

⎪⎭

⎪⎬

−∞→→===−

.~,0,0~,

,02

)0(

)0(~

)0(2)0(~~

zWziTW

iWW

z

zz π (2.5.9)

This is the ordinary Ekman problem for constant A, and the solution can be written

zmCW ~exp0)0( = , (2.5.10)

where im 22 2π= and miTC /0 = .

Page 67: Dynamic Oceanography

67

Next, the terms multiplying ε1 yield, when set to zero:

⎪⎭

⎪⎬

−∞→→==

−−=−

.~,0,0~,0

,2

)1(

)1(~

)0(~

)0(~~

)1(2)1(~~

zWzW

WzWiWW

z

zzzzz π (2.5.11)

Here W(0) is given by (2.5.10). It is easy to show that a particular solution can be

written

zmzEzmzDWp~exp~~exp~ 2)1( += . (2.5.12)

By insertion into (2.5.11) and comparing terms with equal powers of z~ , we find that

4/,4/ 00 CmECD −=−= . (2.5.13)

A complete solution of (2.5.11) that satisfies −∞→→ zW ~,0)1( , can then be written

)1(1

)1( ~exp pWzmCW += . (2.5.14)

The application of the boundary condition 0~,0)1(~ == zWz , finally determines the

constant C1. The solution can be written

zmzmzmVW ~exp)~~1(24

220)1( −−=π

, (2.5.15)

where

2/10

)(

0 )( fAV

yS

ρτ

= . (2.5.16)

In particular, let us study the surface current, i.e. we take 0~ =z . The complex

surface flow W(0) can be written in the present approximation as

⎟⎠⎞

⎜⎝⎛ ++=+=

πεε

41

2)0()0()0( 0)1()0( iVWWW . (2.5.17)

From the real and imaginary parts we obtain

.

2)0(

,4

12

)0(

0

0

Vv

Vu

=

⎟⎠⎞

⎜⎝⎛ +=

πε

(2.5.18)

If we had continued our approximations, the next terms in (2.5.18) would have been

proportional to ε2. If ε is sufficiently small, these terms can be neglected. The surface

current is deflected an angle θ0 to the right of the wind. This angle is given by

⎟⎠⎞

⎜⎝⎛ +==

πεθ

41arctan

)0()0(arctan0 v

u . (2.5.19)

Page 68: Dynamic Oceanography

68

We then observe that if ε < 0 (A increases with depth), we have θ0 < 45°. For ε > 0 (A

decreases with depth), we have θ0 > 45°. We must remember that this result has been

derived under the assumption that |ε| << 1, or

2/1

20

2⎟⎠⎞

⎜⎝⎛<<π

α fA . (2.5.20)

Observations close to the surface often show a tendency for wind-driven ocean

currents (when geostrophic components are removed) to be more in the direction of

the wind than what follows from an Ekman analyses with constant eddy viscosity.

This can partly be due to the fact that some of the transport is caused by surface wind

waves in a direction close to the wind. However, it also suggests that the turbulent

eddy viscosity must increase with depth in the upper part of the surface layer, as

indicated by the present analysis. Since the eddy coefficients are small in the interior

of the ocean, the eddy viscosity must therefore increase to a maximum value at a

certain depth, before it decreases to attain its deep ocean value.

2.6 Equations for the volume transport

At the free surface ),,( tyxz η= , we neglect any turbulent fluctuations. This

means that η can be associated with the mean value defined by (2.1.1). By integrating

the continuity equation 0=⋅∇ vr for the mean motion in the vertical, and applying the

mean versions of the boundary conditions (1.1.6) and (1.1.8), we find

∫∫−− ∂

∂−

∂∂

−=ηη

ηHH

t dzvy

dzux

. (2.6.1)

This is identical to (1.1.10), but the dependent variables here are mean quantities.

When we take into account the effects of friction and a variable density, u and v are

no longer independent of z, as in section 1. We now introduce volume transports per

unit length. In the x- and y-direction they are, respectively

⎪⎪⎭

⎪⎪⎬

=

=

−η

η

H

H

dzvV

dzuU

.

, (2.6.2)

Accordingly, we can write (2.6.2) as

yxt VU −−=η . (2.6.3)

Page 69: Dynamic Oceanography

69

In the momentum equation we apply the Boussinesq approximation, i.e. we

assume that the density changes are only important in connection with the action of

gravity. This means that in (2.1.11) we can use ρ = ρr, where ρr is a constant reference

density. Integrating the acceleration term in (2.1.11), we find

∫∫∫−−− ∂

∂+

∂∂

+=∇⋅+ηηη

HHt

Ht dzuv

ydzu

xUdzuvu 2)( r . (2.6.4)

From the assumption of hydrostatic balance in the vertical, we can write the

pressure as

∫+=η

ρz

S dzgtyxPp ),,( . (2.6.5)

Let us now consider the x-component of (2.1.11). By integrating the pressure term

vertically, and applying the Boussinesq approximation, we find that

xBr

xSr

px

rHx

rx HPPIdzpdzp

ρη

ρρρρ

ηη 11111 )(

H

++−=−≈− ∫∫−−

. (2.6.6)

Here PB and I(p) are the pressure at the bottom and the depth-integrated pressure,

respectively. They are defined by

⎪⎪⎭

⎪⎪⎬

=

+≡

.

,

H

)(η

η

ρ

dzpI

dzgPP

p

HSB

(2.6.7)

The friction term containing the vertical variation becomes

)(1)( )()( xB

xS

rHzz dzAu ττ

ρ

η

−=∫−

, (2.6.8)

where )()()( ηττ xxS = is the wind stress and )()()( Hxx

B −=ττ is the bottom stress in the

x-direction. The integrated friction terms containing the horizontal variation are more

difficult to write in a simple form. For example,

⎥⎥⎦

⎢⎢⎣

⎡−−−−−−

∂∂

= ∫∫−−

)(2)()(2)(2

2

HuHHuHuudzux

AdzuA xxxxxxxxH

HH

xxH ηηηηηη

, (2.6.9)

where we have taken AH to be constant. We will follow Munk (1950), and assume that

the first term dominates on the right hand side, i.e.

xxHH

xxH UAdzuA ≈∫−

η

. (2.6.10)

Page 70: Dynamic Oceanography

70

Quite similar results are obtained for the y-component of (2.1.11). Furthermore, we

assume that the integrated advection terms in (2.6.4) are small compared to the

integrated Coriolis term (i.e., we assume small Rossby numbers). The integrated

momentum equation can then be written

⎪⎭

⎪⎬⎫

∇+−+++−=+∇+−+++−=−

.)(,)(

2)()()(

2)()()(

VAHPPIfUVUAHPPIfVU

HHry

By

SyBySp

ytr

HHrx

Bx

SxBxSp

xtr

ρττηρρττηρ

(2.6.11)

These equations are very useful for describing oceanic transports. Even though we

have applied the Boussinesq approximation, they are valid for an ocean with variable

density, thus including baroclinic motion; see section 3.

In literature one often find these equations written in terms of mass transports

(M(x), M(y)) instead of volume transports, as here. The transition is very simple, since

⎪⎪⎭

⎪⎪⎬

=≈≡

=≈≡

∫∫

∫∫

−−

−−

.

,

)(

)(

VdzvdzvM

UdzudzuM

rH

rH

y

rH

rH

x

ρρρ

ρρρ

ηη

ηη

(2.6.12)

An interesting challenge of great practical importance is to determine how the

surface elevation varies in time and space, when we know the atmospheric surface

pressure and the wind stresses from meteorological prognoses. For this particular

problem, it turns out that we to a good approximation can neglect the entire density

variation, i.e. treat the ocean as a constant density fluid. In that case we obtain from

(2.6.5):

)(),,( zgtyxPp rS −+= ηρ . (2.6.13)

Furthermore, we assume a simple, linear relation between the bottom stress and the

volume flux:

⎭⎬⎫

==

,//,//

)(

)(

HrVHrU

ry

B

rx

B

ρτρτ

(2.6.14)

where r is a constant friction coefficient. This is analogous to (2.2.6) for the Ekman

current at the bottom, except that here we use the mean velocities U/H and V/H, and

not the bottom velocity. By assuming that H<<η , which always is a good

approximation, (2.6.11) reduces to

Page 71: Dynamic Oceanography

71

⎪⎪⎭

⎪⎪⎬

−+−−=+

−+−−=−

.1

,1

)(

)(

VHrPHgHfUV

UHrPHgHfVU

yS

rSy

ryt

xS

rSx

rxt

τρρ

η

τρρ

η (2.6.15)

Here we have disregarded horizontal diffusion of momentum, i.e. taken AH = 0. These

equations, together with (2.6.3), govern the so-called storm surge problem for an

ocean of variable depth. The numerical solution of this system of equations, within

the appropriate geometrical domain, is performed routinely on a daily basis in many

countries bordering the sea to predict storm-induced water level changes along the

coasts.

We will not go into the general solution of this problem, but we mention that for

a storm area of limited extent, the solution will consist of a combination of free waves

“leaking” out of the area (propagating much faster than the storm centre) and a forced

solution like Ekman's elementary current system. This part of the solution has the

same lateral dimension as the storm area, and it follows approximately the motion of

the storm centre.

Equations (2.6.15) also describe how the motion becomes attenuated if the

atmospheric sea surface pressure becomes constant (∇PS = 0), and the wind ceases to

blow ( 0=Sτr ). Let us assume that H is constant. By introducing the mean vorticity

)(1yx UV

H−≡ζ , (2.6.16)

we find on an f-plane that

ζζHr

t −= . (2.6.17)

Accordingly, the integrated motion decreases exponentially with time. This is

analogous to what we found in (2.4.7) for the spin-down of a barotropic vortex due to

Ekman suction. We note that (2.6.17) is in fact correct even if ∇PS ≠ 0, as long as the

wind stress has zero vorticity, i.e. if 0)( =×∇⋅ Sk τrr

.

As a second example of a transient solution of (2.6.15) for constant depth, we

consider planetary motion (β ≠ 0). We now neglect the effect of bottom friction. For

an approximately stationary surface elevation (thereby excluding the presence of

Sverdrup waves), (2.6.3) reduces to

0=+ yx VU . (2.6.18)

Page 72: Dynamic Oceanography

72

We introduce a stream function ϕ for the volume transport such that

⎭⎬⎫

=−=

.,

x

y

VU

ϕϕ

(2.6.19)

By obtaining the vorticity from (2.6.15), and inserting from (2.6.19), we find

)/(2rSxtH k ρτβϕϕ rr

×∇⋅=+∇ (2.6.20)

By comparing with (1.9.4), we realize that the homogenous solution of this equation

yields Rossby waves. We also note that the vorticity of the wind stress acts as a

source term for Rossby waves in the ocean.

2.7 The Sverdrup transport

For the rest of section 2 we are going to study ocean circulation. We simplify,

and assume that the surface pressure PS and the depth H are constants. Furthermore,

we consider stationary motion in the ocean. Then, in terms of the mass transports,

(2.6.3) becomes

0)()( =+ yy

xx MM , (2.7.1)

which naturally leads to the definition of a stream function ψ for the steady mass

transport:

.

,)(

)(

xy

yx

M

M

ψ

ψ

=

−= (2.7.2)

For this problem it proves convenient to derive an equation for the vertical vorticity

associated with the mass transports. With the present assumptions, and utilizing

(2.6.12), we find

)()()( )()(2)( xy

yxHHBS

y MMAkkM −∇+×∇⋅−×∇⋅= ττβ rrrr. (2.7.3)

By substituting the stream function from (2.7.2), we obtain a partial differential

equation for the single variable ψ :

ψττψβ 4)()( HHBSx Akk ∇+×∇⋅−×∇⋅=rrrr

. (2.7.4)

Sverdrup (1947) assumed that the effects of friction could be neglected altogether. In

that case we need not utilize the stream function, since the mass transport in the north-

south direction M(y) is given directly by (2.7.3), i.e.

)()(S

y kM τβ rr×∇⋅= . (2.7.5)

Page 73: Dynamic Oceanography

73

This relation is often referred to as the Sverdrup balance. From (2.7.1) we then obtain

that

⎥⎦

⎤⎢⎣

⎡ ×∇⋅∂∂

−=βτ )()( Sx

xk

yM

rr

. (2.7.6)

Sverdrup applied these equations to the equatorial region and assumed that the wind

(and the wind stress) was purely zonal, i.e.

⎭⎬⎫

==

).(,0

)()(

)(

yxS

xS

yS

τττ

(2.7.7)

Then (2.7.6) for the east-west transport reduces to

2

)(2)( 1

dydM

xSx

β= . (2.7.8)

This equation is only 1. order in x, and accordingly we can only apply one boundary

condition. We take that 0)( =xM at the eastern boundary. If this boundary is located

at x = L, the Sverdrup transports become, respectively:

)(12

)(2)( Lx

dydM

xSx −=

τβ

. (2.7.9)

dy

dMx

Sy)(

)( 1 τβ

−= . (2.7.10)

In Fig. 2.3 we have sketched the wind stress distribution in the equatorial region, and

the associated mass transport from (2.7.9).

Fig. 2.3. Sketch of the wind stress distribution and the mass transport in the

equatorial region.

Page 74: Dynamic Oceanography

74

Since M(x) ≠ 0 for x = 0, this solution is not valid close to the left (western) boundary.

We note from (2.7.9) that the equatorial transport is proportional to the 2. derivative

(the curvature) of the north-south distribution of the wind stress. By inserting an

equatorial easterly wind field (the trade winds) with a local minimum (the doldrums)

slightly north of the equator, Sverdrup demonstrated that we in principle obtain the

observed equatorial current system with westerly flowing currents north and south of

the equator and an easterly flowing counter current sandwiched between them; see the

sketch in Fig. 2.3. In reality, the equatorial flow pattern is more complicated, which is

probably due to the fact that the real wind system is not as simple as the one used

here.

An apparent paradox is revealed when we compare the Sverdrup east-west

transport, which is along the wind, to the Ekman transport, which is directed 90° to

the right of the wind. The solution to this problem lays in the fact that the Sverdrup

solution also contains the integrated pressure I(p), defined by (2.6.7). From the

stationary version of (2.6.11), with )( ySτ = )( y

Bτ = η = AH = 0 and H = const., we find

)()( xpy MfI −= . (2.7.11)

By inserting from (2.7.9):

CLxdy

dfIx

Sp +−−= )()(

)( τβ

, (2.7.12)

where C is a constant. In this expression, we note that x ≤ L. Thus, we see that for a

given x, the integrated pressure I(p) has a maximum where dyd xS /)(τ has its largest

value. This can only be a result of the transport to the right of the wind in the Ekman

layer close to the surface; see Fig. 2.4, where we have assumed a simple sinusoidal

wind profile.

Page 75: Dynamic Oceanography

75

Fig. 2.4. The connection between the wind direction, the Ekman transport and the

Sverdrup transport.

Accordingly, the Sverdrup transport in the east-west direction results from

geostrophic balance in the north-south direction.

Finally, we remark that the Sverdrup balance (2.7.5) expresses the fact that the

advection of planetary vorticity is equal to the vorticity of the surface wind stress. For

an anti-cyclonic wind field, as depicted in Fig. 2.4, we have that )( Sk τrr

×∇⋅ =

0/)( ≤− dyd xSτ . Hence, a fluid column must move in such a way that it has a

southward-directed velocity component, as indicated in the figure.

2.8 Theories of Stommel and Munk (western intensification)

As emphasized in the previous section, Sverdrup’s solution for the east-west

transport does not satisfy the boundary condition at the western boundary. To remedy

this fault, the effect of friction must be taken into account. This was done by Stommel

(1948), and by Munk (1950).

We return to the more general equation (2.7.4). Stommel (1948) solved it with an

idealized wind stress distribution, while taking AH = 0. Munk (1950) neglected the

bottom friction, i.e. assumed 0=Bτr , and solved the equation using realistic wind

data. Munk's solution reproduces most of the features of the large-scale ocean

circulation. However, this solution is quite complicated. We will therefore be content

with Stommel’s simplified approach, which yields the main features of the circulation

Page 76: Dynamic Oceanography

76

pattern. However, we will return to Munk’s model when we discuss the boundary

layer at the western coast, using a simplified wind field.

Stommel modelled the bottom stress1 by assuming

⎭⎬⎫

==

,,

)()(

)()(

yyB

xxB

MKMK

ττ

(2.8.1)

where K is a constant friction coefficient. The ocean is taken to be a rectangular basin

of length L and width D, subject to a zonal wind field of the form:

⎪⎭

⎪⎬⎫

−=

=

.cos

,0

0)(

)(

Dyx

S

yS

πττ

τ (2.8.2)

The ocean geometry and the wind stress are sketched in Fig. 2.5.

Fig. 2.5. Ocean basin and wind field for the Stommel solution.

With the parameterisations (2.8.1) and (2.8.2), and the assumption AH = 0, (2.7.4)

reduces to

Dy

DK xH

πτπβψψ sin02 −=+∇ . (2.8.3)

The boundaries of the model are impermeable, which means that the flow must be

purely tangential at any part of the boundary. From the definition of the stream

function, this means that the boundary is a streamline given by ψ = 0.

We assume that the solution of (2.8.3) can be written

1 Stommel imagined that the lower limit of the vertical integration was taken to be the depth where the current velocity was practically zero. This depth is usually much less than the total depth. In this case, ”the bottom stress” refers to the frictional stress between the dynamically active upper layer, and the passive lower layer.

Page 77: Dynamic Oceanography

77

Dyx πφψ sin)(= . (2.8.4)

This solution satisfies the boundary conditions for y =0, D, i.e. ψ(0) = ψ(D) = 0 for all

x. Inserting (2.8.4) into (2.8.3), we find

DKDK

02

2 τπφπφβφ −=−′+′′ , (2.8.5)

with boundary conditions Lx ,0for,0 ==φ . This equation has a particular solution

KD

p πτ

φ 0= , (2.8.6)

while the homogeneous solution can be written

)exp()exp( 2211 xrCxrCh +=φ . (2.8.7)

Here

2/1

2

2

2

2

21 42, ⎥

⎤⎢⎣

⎡+±−=

DKKrr πββ . (2.8.8)

The constants C1 and C2 are obtained from the boundary conditions. By insertion:

⎭⎬⎫

=+=+

.0)(,0)0(

ph

ph

L φφφφ

(2.8.9)

This yields

⎪⎪⎭

⎪⎪⎬

−−

−=

−−

=

.)exp()exp(

1)exp(

,)exp()exp(

1)exp(

21

102

21

201

LrLrLr

KDC

LrLrLr

KDC

πτ

πτ

(2.8.10)

The complete solution can then be written

⎥⎥⎥

⎢⎢⎢

⎡ −−+−−=

)sinh(

))(sinh(2

exp)sinh()(2

exp1sin0

L

xLxK

xxLK

Dy

KD

α

αβαβπ

πτψ , (2.8.11)

where

2/1

2

2

2

2

4 ⎥⎦

⎤⎢⎣

⎡+=

DKπβα . (2.8.12)

Let us first see what this solution yields on an f-plane, i.e. when β = 0. It is then

convenient to place the origin in the centre of the basin. This is done by defining new

coordinates:

2/,2/ ** DyyLxx −=−= . (2.8.13)

Page 78: Dynamic Oceanography

78

By insertion into (2.8.11), with β = 0, we find that

⎥⎥⎥⎥

⎢⎢⎢⎢

⎡⎟⎠⎞

⎜⎝⎛ ++⎟

⎠⎞

⎜⎝⎛ −

−=sinh

2

sinh2

sinh1cos),(

***0

**

LD

xLD

xLD

Dy

KDyx π

πππ

πτψ . (2.8.14)

We realize straightaway that the streamlines from (2.8.14) are symmetric about the

basin centre. This situation is depicted in Fig. 2.6.

Fig. 2.6. Stream lines when the beta-effect is zero.

This is not in accordance with our knowledge of the large-scale ocean circulation,

which tells us that the currents are much more intense on the western side (the Gulf

Stream, Kuroshio) than on the eastern side. We therefore conclude that the beta-effect

must be significant. By inserting relevant parameters (β ~ 2×10−11 m−1s−1, K ~ 10−6s−1,

D ~ 6×103 km, L ~ 104 km) we find a flow pattern from (2.7.23) as sketched in Fig.

2.7.

Fig. 2.7. The flow pattern when the beta-effect is included.

Page 79: Dynamic Oceanography

79

In this case the solution clearly reproduces the intensified flow on the western side of

the ocean.

The northward transport on the western side of the ocean occurs mainly within a

thin boundary layer, where the thickness is determined by the effect of friction and the

beta-effect. For Stommel’s model, the typical boundary-layer scale δS is

βδ /~ KS . (2.8.15)

The mass transport )( ybM in the boundary layer is sketched in Fig. 2.8. We will return

to this problem at the end of this section.

Fig. 2.8. Sketch of the transport at the western boundary in Stommel’s model.

Finally, by a simple boundary-layer analysis we discuss the northward transport

at the western boundary in the Munk model ( 0=Bτr , AH ≠ 0). By assuming that

yx ∂∂>>∂∂ // in the boundary layer, the leading terms in (2.7.3) becomes

0=− xxxxxHA βψψ . (2.8.16)

Since xyM ψ=)( , we can insert )( y

bM directly into (2.8.16). Furthermore, the

boundary-layer scale δM for this problem is defined by

3/1)/( βδ HM A= . (2.8.17)

By introducing a boundary-layer coordinate Mxx δ/~ = , equation (2.8.16) can be

written

0~)(

3

)(3

=− yb

yb M

xdMd . (2.8.18)

This equation has a solution of the form

)~exp()~exp()~exp( 332211)( xrCxrCxrCM y

b ++= , (2.8.19)

Page 80: Dynamic Oceanography

80

where

2

321,

23

21,1

2/1

3

2/1

21 irirr −−=+−== . (2.8.20)

For the solution of (2.8.18) we must require that )( ybM is finite when ∞→x~ , and that

0)( =ybM when 0~ =x (no-slip). Accordingly, C1 = 0 and C2 = − C3 in (2.8.19). The

solution can then be written as

⎟⎟⎠

⎞⎜⎜⎝

⎛−= xxCM y

b~

23sin)2/~exp()( . (2.8.21)

Here, the constant C must be determined by matching with the solution valid outside

the boundary layer, i.e. the Sverdrup transport. Utilizing the simplified zonal wind

stress (2.8.2), (2.7.10) yields

.,sin1 0)( LxDy

DM M

ySv ≤<⎟

⎠⎞

⎜⎝⎛−= δπτπ

β (2.8.22)

In a closed basin, the total north-south transport must be zero. This constitutes the

required matching condition, i.e.

0~ )(

0

)( =+ ∫∫∞ L

ySvM

yb

M

dxMxdMδ

δ . (2.8.23)

Insertion from (2.8.21) and (2.8.22) allows us to determine C. When we utilize that δM

<< L, the mass transport in the boundary layer becomes

⎟⎠⎞

⎜⎝⎛

⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛−=

Dyxx

DLM

MMM

yb

πδδβδ

τπ sin2

3sin2

exp32 0)( . (2.8.24)

We have sketched this solution in Fig. 2.9.

Fig. 2.9. Sketch of the transport at the western boundary in the Munk model.

Page 81: Dynamic Oceanography

81

The dimensional thickness bx∆ of the layer where the transport is northward, is given

by

Mbx δπ3

2=∆ . (2.8.25)

With AH ~ 103 m2s−1 and β = 2×10−11m−1s−1, we find that bx∆ ~ 134 km. This fits well

with the width of the Gulf Stream.

A similar boundary-layer analysis for the Stommel model yields that

⎟⎠⎞

⎜⎝⎛

⎟⎟⎠

⎞⎜⎜⎝

⎛−=

Dyx

DL

KM

S

yb

πδ

τπ sinexp0)( , (2.8.26)

where δS = K/β. If we choose a typical boundary-layer thickness Sbx δπ=∆ , our

former values for the parameters of the problem gives that bx∆ ~ 160 km.

It is worth noticing that in the strong shear in the western boundary current (see

Fig. 2.9), the Rossby number )/(~ vkfvvR rrrr×∇⋅ is of the same order of magnitude

as the horizontal Ekman number )/(~ 2 vkfvAE HHrrr

×∇ . This means that also the

nonlinear terms in the momentum equation may be important in this region.

3. BAROCLINIC MOTION

3.1 Two-layer model

We now proceed to study the effect of vertical density stratification in the ocean.

In many situations the density is approximately constant in a layer close to the

surface, while the density in the deeper water is also are constant (and larger). The

transition zone between the two layers is called the pycnocline. Thin pycoclines are

typically found in many Norwegian fjords. In extreme cases we can imagine that the

pycnocline thickness approaches zero, resulting in a two-layer model with a jump in

the density across the interface between the layers.

We start out by studying such a model. For simplicity we describe the motion in

reference system as shown in Fig. 3.1, where the x-axis is situated at the undisturbed

interface between the layers. The constant density in each layer is ρ1 and ρ2,

respectively, where ρ2 > ρ1.

Page 82: Dynamic Oceanography

82

Fig. 3.1. Model sketch of the two-layer system.

We assume hydrostatic pressure distribution in each layer. By applying that the

pressure is PS along the surface, and continuous at the interface, i.e. p1(z = ξ) = p2(z =

ξ), we find that

⎭⎬⎫

+++−+−=+++−=

.)()(,)(

111222

1111

S

S

PHgggzpPHggzp

ηρξρρρηρρ

(3.1.1)

We average the motion in the upper and lower layer:

∫+

ξ

1

),(1)ˆ,ˆ( 111

11

H

dzvuh

vu , (3.1.2)

∫−

2

),(1)ˆ,ˆ( 222

22H

dzvuh

vu , (3.1.3)

Here, ξη −+= 11 Hh and 22 Hh += ξ are the total depths of the upper and lower

layers, respectively.

We assume that our equations can be linearized, i.e. we neglect the convective

accelerations. Furthermore, we will disregard the effect of the horizontal eddy

viscosity. This means that AH = 0 in each layer. By introducing volume transports

⎭⎬⎫

==

),ˆ,ˆ(),(),ˆ,ˆ(),(

22222

11111

vuhVUvuhVU

(3.1.4)

the momentum equation for the upper layer becomes:

Page 83: Dynamic Oceanography

83

⎪⎪⎭

⎪⎪⎬

−+−−=+

−+−−=−

.11

,11

)(

1

)(

11

1111

)(

1

)(

11

1111

yi

ySySyt

xi

xSxSxt

PhghfUV

PhghfVU

τρ

τρρ

η

τρ

τρρ

η (3.1.5)

Here, ( )(xSτ , )( y

Sτ ) are the wind stresses along the surface, and ( )( xiτ , )( y

iτ ) are the

internal frictional stresses between the layers.

Equivalently, for the lower layer we find

⎪⎪⎭

⎪⎪⎬

−+−−−=+

−+−−−=−

,11

,11

)(

2

)(

22

22*2

2

122

)(

2

)(

22

22*2

2

122

yB

yiySyyt

xB

xixSxxt

PhhgghfUV

PhhgghfVU

τρ

τρρ

ξηρρ

τρ

τρρ

ξηρρ

(3.1.6)

where ( )(xBτ , )( y

Bτ ) are the frictional stresses at the bottom. Furthermore, we have

defined

gg ⎟⎟⎠

⎞⎜⎜⎝

⎛ −≡

2

12* ρ

ρρ , (3.1.7)

which is referred to as the reduced gravity, because the fraction 112 /)( ρρρ − is small

for typical ocean conditions.

By integrating the continuity equation 0=++ zyx wvu in each layer, we find,

without any linearization of the boundary conditions, that

⎭⎬⎫

−−=−−=−

.,

22

11

yxt

yxtt

VUVU

ξξη

(3.1.8)

Barotropic response

Assume that the mean velocities in each layer are approximately equal, i.e.

21 ˆˆ uu ≈ , 21 ˆˆ vv ≈ . This leads to

2

2

1

1

2

2

1

1 ,HV

HV

HU

HU

≈≈ , (3.1.9)

when we assume that 21,, HH<<ξη . For simplicity, we also take that the lower

layer has a constant depth. From (3.1.8) we then obtain

tyxtt HHVU

HH ξξη

2

122

2

1 )( =+−=− , (3.1.10)

or

Page 84: Dynamic Oceanography

84

ηξ21

2

HHH+

= . (3.1.11)

Here the integration constant must be zero when we consider wave motion. We note

from (3.1.11) that ξ and η are in phase, and that |ξ | < |η |.

By neglecting the effect of the earth’s rotation, assuming constant surface

pressure and neglecting frictional effects in (3.1.5), equations (3.1.9) and (3.1.11)

yield

0)( 21 =+− xxtt HHg ηη , (3.1.12)

when 0/ =∂∂ y . The solution is

)()( 0201 tcxFtcxF ++−=η , (3.1.13)

where )( 2120 HHgc += . The expression (3.1.13) describes long surface waves

propagating in a non-rotating canal with depth H1 + H2. This is the solution we would

have found if we, as a starting point, had neglected the density difference between the

layers; see the one-layer model in section 1. Such a solution (a free wave), which is

not influenced by the small density difference between the layers, is often referred to

as the barotropic response.

The original meaning of the word “barotropic” is related to the field of mass, and

expresses the fact that the pressure is constant along the density surfaces, i.e. the

isobaric and isopycnal surfaces coincide. Mathematically this can be expressed as

0=∇×∇ ρp . This was the case for the free waves in section 1, where the density was

constant everywhere, and the pressure was constant along the sea surface. For the

two-layer model this would mean that the pressure should be constant along the

interface between the two layers. This is only approximately satisfied here, but

nonetheless it has become customary to denote the response in this case as the

barotropic response.

Baroclinic response

We now assume that tt ξη << . Then, from (3.1.8):

0)()( 2121 =+++ yx VVUU . (3.1.14)

For simplicity we take the bottom to be flat. A particular solution of (3.1.14) can be

written

Page 85: Dynamic Oceanography

85

⎭⎬⎫

−=−=

,,

21

21

VVUU

(3.1.15)

i.e. the volume fluxes are equal, but oppositely directed in each layer. By taking the

surface pressure to be constant, and neglecting the effect friction, summation of

(3.1.5) and (3.1.6) yields

ηρρρρξ ⎥

⎤⎢⎣

⎡−+

−=212

2112

)( HHH , (3.1.16)

where, as in the barotropic case, the integration constant must be zero. Furthermore,

we have used that h1 ≈ H1, h2 ≈ H2. The difference between ρ1 and ρ2 is quite small,

which allows us to use the approximations ρ2 −ρ1 = ∆ρ, and ρ1 ~ ρ2 ≈ ρ. Thus,

equation (3.1.16) can be rewritten as

ηρρξ ⎥

⎤⎢⎣

⎡ +∆

−=2

21

HHH . (3.1.17)

We note that ξ and η are oppositely directed, and that |ξ | >> |η |, as initially assumed.

Assuming that 0/ =∂∂ y , we obtain from (3.1.6), (3.1.8), and (3.1.17) that

021 =− xxtt c ξξ , (3.1.18)

where

21

21*

21 HH

HHgc+

= . (3.1.19)

The solution of (3.1.18) can be written

)()( 1211 tcxFtcxF ++−=ξ . (3.1.20)

This represents internal gravity waves propagating with phase speed c1 along the

interface between the layers; see the sketch in Fig. 3.2.

Fig. 3.2. Internal wave in a two-layer model.

Page 86: Dynamic Oceanography

86

The solution to (3.1.20) is often called the baroclinic response. As for barotropic, the

term “baroclinic” is linked to the mass field. In a baroclinic mass field the constant

pressure surfaces and the constant density surfaces intersect, i.e. 0≠∇×∇ ρp .

Equation (3.1.17) shows that this is the case here, since when ξ > 0, then η < 0.

Accordingly, the pressure varies along the interface, which is a constant density

surface.

Let us assume that the lower layer is very deep, i.e. H2 >> H1. This is the most

common configuration in the ocean. From (3.1.6) we find for the x-component in the

lower layer

⎥⎦

⎤⎢⎣

⎡+−−+−−= 22

)(

2

)(

222*

2

1 1111 VfUh

Pgg tx

Bx

ixSxx τρ

τρρ

ξηρρ . (3.1.21)

For the baroclinic case, U2 and V2 are finite when h2 → ∞, and so are the frictional

stresses. Accordingly, for this limit, (3.1.21) reduces to

xSxx Pgg1

*1

2 1ρ

ξρρη −−= . (3.1.22)

In the same way we find for the y-component

ySyy Pgg1

*1

2 1ρ

ξρρη −−= . (3.1.23)

By inserting (3.1.22) and (3.1.23) into (3.1.5) for the upper layer, we find that

⎪⎪⎭

⎪⎪⎬

−+=+

−+=−

.11

,11

)(

1

)(

1*1

1

211

)(

1

)(

1*1

1

211

yi

ySyt

xi

xSxt

ghUfV

ghVfU

τρ

τρ

ξρρ

τρ

τρ

ξρρ

(3.1.24)

For the baroclinic case, the depth of the upper layer can be written

ξηξ −≈+−= 111 HHh . Furthermore, we apply that ρρρ ∆=− 12 , and ρ1 ≈ ρ2 = ρ.

By linearizing the pressure term (the first term on the right-hand side), equations

(3.1.24) and (3.1.8) yield

⎪⎪⎪

⎪⎪⎪

−−=

−+−=+

−+−=−

.

,11

,11

111

)()(11*11

)()(11*11

yxt

yi

ySyt

xi

xSxt

VUh

hHgUfV

hHgVfU

τρ

τρ

τρ

τρ

(3.1.25)

These equations for the baroclinic response in the upper layer are formally identical to

the equations describing the storm surge problem for a homogeneous ocean; see

Page 87: Dynamic Oceanography

87

(2.6.15), when the upper layer thickness replaces the surface elevation, and the gravity

g is replaced by *g . The set of equations (3.1.25) describes what is often referred to as

a reduced gravity model for the volume transport in the upper layer. Even though the

numerical values for the volume fluxes in the lower layer are of the same order of

magnitude as in the upper layer, the mean velocity in the lower layer is negligible,

since H2 → ∞. Therefore, we usually say that the lower layer has no motion in this

approximation.

We immediately realize from (3.1.25) that transient phenomena such as

Sverdrup-, Kelvin- and planetary Rossby waves in a rotating ocean of constant

density have their internal (baroclinic) counter-parts in a two-layer model. The

analysis for the internal response is identical to the analysis is section 1. It often

suffices to replace g with *g and H with H1 in the solution for the barotropic

response.

Analogous to the barotropic case we can define a length scale a1 that

characterizes the significance of earth’s rotation. We write

fca /11 = , (3.1.26)

where 1*21 Hgc = . The length scale a1 is called the internal, or baroclinic, Rossby

radius. Typical values for c1 in the ocean are 2-3 m s−1. Hence, a1 ≈ 20-30 km, which

is much less than the typical barotropic Rossby radius. Therefore, the effect of earth’s

rotation will be much more important for the baroclinic response than for the

barotropic one with the same horisontal scale, or wavelength.

3.2 Continuously stratified fluid

We now turn to the more general problem of a continuous density stratification,

and start by investigating the stability of a stratified incompressible fluid under the

influence of gravity. The equilibrium values are:

⎪⎭

⎪⎬

+−====

∫ .const)(),(

,0

00

0

dzgzppz

v

ρρρ

r

(3.2.1)

We introduce small perturbations (denoted by primes) from the state of equilibrium,

writing the velocity, density, and pressure as

Page 88: Dynamic Oceanography

88

⎪⎭

⎪⎬

′+=

′+=

′=

).,,,()(),,,,()(

),,,,(

0

0

tzyxpzpptzyxz

tzyxvvρρρ

rr

(3.2.2)

We assume that the density is conserved for a fluid particle. Furthermore, we take that

the perturbations are so small that we can linearize our problem, i.e. neglect terms that

are products of perturbation quantities. The equations for the conservation of

momentum, density, and mass then reduce to

)(0 ))(( x

zxt pfvuz τρ +−=− , (3.2.3)

)(0 ))(( y

zyt pfuvz τρ +−=+ , (3.2.4)

gpwz zt ρρ −−=)(0 , (3.2.5)

00 =+ wdz

dt

ρρ , (3.2.6)

0=++ zyx wvu . (3.2.7)

Here we have for simplicity left out the primes that mark the perturbations. In (3.2.3)

and (3.2.4) )( xτ and )( yτ represent the tangential stresses in the fluid. We have also

assumed that the vertical variations of these stresses are much larger than the

corresponding horisontal variations. Furthermore, we have neglected the effect of

friction in vertical component of the momentum equation (3.2.5).

3.3 Free internal waves in a rotating ocean

We start by disregarding completely the effect of friction on the fluid motion, i.e.

we take 0)()( == yx ττ in (3.2.3) and (3.2.4). Furthermore, we introduce the Brunt-

Väisälä frequency (or the buoyancy frequency) N, defined by

dz

dgzN 0

0

2 )( ρρ

−= . (3.3.1)

We are here going to study motion in a stably stratified incompressible fluid. In this

case we must have that 0/0 <dzdρ , meaning that N is real and positive. Equation

(3.2.6) can then be written

002 =− wNg t ρρ . (3.3.2)

By differentiating (3.2.5) with respect to time, and utilizing (3.3.2), we find that

tztt pwNw −=+ )( 20ρ . (3.3.3)

Page 89: Dynamic Oceanography

89

From this equation we note that the time scale for pure vertical motion ( 0=ztp ) is

1−N . Elimination of the gradient from (3.2.3)-(3.2.4), yields the vorticity equation. On

an f-plane we obtain

ztyx fwuv =− )( , (3.3.4)

where we have applied (3.2.7). Forming the horizontal divergence from the same two

equations, and applying (3.2.7), we find

puvfwz Hyxzt2

0 )()( ∇=−+ρ . (3.3.5)

Elimination of the vorticity from the equations above, yields

[ ] tHzttz pwfw 220 ∇=+ρ . (3.3.6)

Finally, by eliminating the pressure between (3.3.3) and (3.3.6), we obtain

0)(1)(10

0

2220

0

2 =+∇+⎥⎦

⎤⎢⎣

⎡+∇ zzH

ttzzH wfwNww ρ

ρρ

ρ. (3.3.7)

We simplify (3.3.7) by assuming that ρ0(z) varies slowly over the typical vertical scale

for w, i.e.

zzzz ww ≈)(10

0

ρρ

. (3.3.8)

This is Boussinesq approximation for internal waves. By introducing the Brunt-

Väisälä frequency (3.3.1), we can write

zzzzz wwg

Nw +−=2

00

)(1 ρρ

. (3.3.9)

We realize from (3.3.8) that the Boussinesq approximation implies that

zzz wwg

N<<

2

. (3.3.10)

If d is a typical vertical scale for the motion, the above equation yields

dgN /2 << , (3.3.11)

where Hd ≤max . For a shallow ocean we typically have that g/H ~ 10−1s−2, while for

a deep ocean (H = 4000 m), the corresponding value becomes ×41~/ Hg 10−2s−2.

Measurements in the ocean show that 2N ~ 10−4 − 10−6s−2, so (3.3.11) is usually very

well satisfied. We will therefore utilize the Boussinesq approximation in the future

analysis of this problem. Equation (3.3.7) then reduces to

0)( 2222 =+∇+∇ zzHtt wfwzNw , (3.3.12)

Page 90: Dynamic Oceanography

90

where ∇2 is the three-dimensional Laplacian operator. We derive the same equation

by letting ρ0(z) ≈ ρr on the left-hand sides of (3.2.3)-(3.2.5), where ρr is a constant

reference density. Then dzdgN r /)/( 02 ρρ−= . This latter approach is probably the

most common one when applying the Boussinesq approximation.

We assume that the ocean is unlimited in the horizontal direction, and consider a

wave solution of the form

))(exp()( tkxizWw ω−= . (3.3.13)

Here the x-axis is directed along the wave propagation direction. From (3.3.12) we

then obtain

022

222 =⎥

⎤⎢⎣

⎡−−

+′′ Wf

NkWω

ω , (3.3.14)

where a prime denote differentiation with respect to z.

Constant Brunt-Väisälä frequency

Later on we shall allow N to vary with z. In this treatment, we simplify, and

assume that N is constant. Typical values for N and f in the ocean (and atmosphere)

are N ~ 10−2s−1 and f ~ 10−4s−1, i.e. N >> f. From equation (3.3.14) we then note that

we have wave solutions in the z-direction if f < ω < N, while for ω < f or ω > N, the

solutions must be of exponential character in the z-direction.

Let us assume that f < ω < N. We then take

)exp( zimW ∝ , (3.3.15)

where m is a wave number in the vertical direction. By insertion into (3.3.14), we

obtain the dispersion relation

22

22222

mkfmNk

++

=ω . (3.3.16)

We realize that we have an-isotropic system, since ω cannot be expressed solely as a

function of the magnitude of the wave number vector.

Analogous to the discussion of planetary Rossby waves in section 1.9, we can

define a wave number vector as

),( mk=κr . (3.3.17)

Then the phase speed and group velocity, become, respectively

2/κκω rr=c , (3.3.18)

Page 91: Dynamic Oceanography

91

and

ωκ∇=gcr . (3.3.19)

From the last relation we notice that the group velocity is always directed towards

increasing values of ω. When ω is constant, we find from (3.3.16) that this gives iso-

lines that are straight lines in the wave number space; see also the sketch in Fig. 3.3.

Fig. 3.3. Lines of constant frequency for internal waves with rotation in the two-

dimensional wave number space.

Along the m-axis (where k = 0), we have ω = f (small). Along the k-axis (where m =

0), we have ω = N (large). Hence, the direction of the group velocity becomes as

shown in the figure, while the phase speed from (3.3.18) is directed along the wave

number vector.

If we imagine that the wave number m is given, we can plot ω as a function of k,

as depicted in Fig. 3.4.

Fig. 3.4. Dispersion diagram for internal waves with rotation.

Page 92: Dynamic Oceanography

92

We can define a Rossby radius of deformation for internal motion with vertical wave

number m by

mfNai = . (3.3.20)

For 1−<< iak , the effect of rotation dominates (compare with Fig. 1.6 for the

barotropic case).

In the ocean, the wave number m cannot be chosen arbitrarily, since the vertical

distance is limited by the depth. If we, for simplicity, disregard the surface elevation

and assume a constant depth, we must have that w = 0 for z = 0, H; see Fig. 3.5.

Fig. 3.5. Internal waves in an ocean with horizontal surface and horizontal bottom.

A solution of (3.3.14), which satisfies the upper boundary condition, is

zmCW sin= , (3.3.21)

where

⎥⎦

⎤⎢⎣

⎡−−

= 22

2222

fNkmω

ω . (3.3.22)

For the solution to satisfy the boundary condition at Hz −= , we must require

...,2,1, == nHnm π (3.3.23)

This means that vertical wave number must form a discrete (but infinite) set. Equation

(3.3.22) then yields for the frequency

2/1

2222

222222

//

⎥⎦

⎤⎢⎣

⎡++

=Hnk

HnfNkπ

πω . (3.3.24)

Page 93: Dynamic Oceanography

93

We see that, for a disturbance with a given wave number in the horizontal direction,

the system (ocean) responds with a discrete number of eigenfrequencies.

The solution for w in this case can be written

( ))(exp)(exp2

)(exp)sin(

tmzkxitmzkxii

CtkxizmCw

ωω

ω

−−−−+=

−= (3.3.25)

The latter expression can be interpreted as the superposition of waves in a horizontal

layer consisting of an incoming, obliquely upward propagating wave, and an

obliquely downward reflected wave, where m must attain the value (3.3.23) for the

wave system to satisfy the boundary condition at the bottom.

We now consider the case where the motion is mainly horizontal. This allows us

to disregard the vertical acceleration in the momentum equation, i.e. we apply the

hydrostatic approximation. Accordingly, in (3.3.3) we take that

wNw tt2<< , (3.3.26)

which leads to

tzpwN0

2 1ρ

−≈ . (3.3.27)

From (3.3.26) we realize that the hydrostatic approximation leads to the condition

22 N<<ω . (3.3.28)

From Fig. 3.4 we note that this implies that k << m, i.e. the horizontal scale of motion

is much larger than the vertical scale. Since the depth H yields the upper limit for the

vertical scale, disturbances with wavelength λ >> H will satisfy the hydrostatic

condition. This requirement applies to barotropic surface waves as well as baroclinic

internal waves.

Applying the hydrostatic approximation, (3.3.16) reduces to

2/1

2

222

⎥⎦

⎤⎢⎣

⎡+=

mNkfω . (3.3.29)

This is the frequency for internal Sverdrup waves. For an ocean with depth H and a

horizontal surface, i.e. m = nπ/H as in equation (3.3.23), we can write

2/1222 )( kcf n+=ω . (3.3.30)

Here

...,3,2,1),/( == nnHNcn π (3.3.31)

which is the phase speed for long internal waves in the non-rotating case. Since

Page 94: Dynamic Oceanography

94

ncNHnm // == π , (3.3.32)

equation (3.3.20) yields the internal (baroclinic) Rossby radius

...,3,2,1,/ === nfcaa nni (3.3.33)

We note that this is analogous to the definition of the barotropic Rossby radius

appearing in (1.3.14). For one single internal mode, i.e. a two-layer structure, this is

similar to (3.1.26).

3.4 Internal response to wind forcing; upwelling at a straight coast

We apply the set of equations (3.2.3)-(3.2.7), and utilize the Boussinesq

approximation and the hydrostatic approximation, i.e.

⎪⎪⎭

⎪⎪⎬

+−=+

+−=−

,11

,11

)(

)(

yz

Sy

St

xz

Sx

St

pfuv

pfvu

τρρ

τρρ (3.4.1)

gpz ρ−= . (3.4.2)

Here the constant surface density ρS is used as a reference density. Furthermore, we

introduce the vertical displacement ),,,( tzyxξ of a material surface, so that w = Dξ/dt

in the fluid. Linearly, this becomes

tw ξ= . (3.4.3)

The conservation of density (3.2.6) then yields for the density perturbation

ξρ

ρ 2Ng

S= . (3.4.4)

where we have assumed that ρ = ξ = 0 at t = 0. Inserting into (3.4.2), we obtain

ξρ 2Np Sz −= , (3.4.5)

while the continuity equation can be written

yxtz vu −−=ξ . (3.4.6)

In general, we take that N = N(z), and we write the solutions to our problem as infinite

series. For simplicity, we assume that the depth is constant, and that the surface is

horisontal at all times. Accordingly:

.0,,0 Hz −==ξ (3.4.7)

In principle, it is also possible to allow the surface to vary in time and space.

However, the solution shows that the internal response can be acchieved, to a good

Page 95: Dynamic Oceanography

95

approximation, by assuming a horisontal surface (the rigid lid approximation); see

Gill and Clark (1975). According to our adopted approach, we write the solutions as

⎪⎪⎪⎪

⎪⎪⎪⎪

=

′=

′=

′=

=

=

=

=

1

1

1

1

),(),,(

),(),,(

,)(),,(

),(),,(

nnn

nnnSn

nnn

nnn

ztyx

ztyxpp

ztyxvv

ztyxuu

φξξ

φρ

φ

φ

(3.4.8)

where the primes denote differentiation with respect to z. By inserting the solutions

into (3.4.5), we find

∑∞

=

=⎥⎦

⎤⎢⎣

⎡+′′

1

2 0)()(),,(n n

nnn zN

pztyxp φξφ . (3.4.9)

For the variables to separate, we must require

2

1const.nn

n

cp==

ξ. (3.4.10)

Furthermore, for (3.4.9) to be satisfied for all x, y and t, we must have that

02

2

=+′′ nn

n cN φφ . (3.4.11)

The boundary conditions (3.4.7) yield

0,,0 Hzn −==φ . (3.4.12)

Equation (3.4.11) and the boundary conditions (3.4.12) define an eigenvalue problem,

i.e. for given N = N(z) we can, in principle, determine the constant eigenvalues cn, and

the eigenfunctions )(znφ , which appear in the series (3.4.8).

It is easy to demonstrate that the differentiated eigenfunctions nφ′ constitute an

orthogonal set. Since (3.4.11) is valid for arbitrary numbers n and m, we can write

⎭⎬⎫

=+′′=+′′

,0,0

22

22

mmm

nnn

NcNcφφφφ

(3.4.13)

where m ≠ n. We multiply the upper and lower equations by φm and φn, respectively.

By subtracting and integrating from z = −H to z = 0, utilizing (3.4.12), we find

∫−

=′′−0

22 0)(H

mnmn dzcc φφ . (3.4.14)

Accordingly, for n ≠ m, i.e. cn ≠ cm, we must have that

Page 96: Dynamic Oceanography

96

mndzmn ≠=′′∫−

,00

H

φφ , (3.4.15)

which proves the orthogonality. Since the eigenfunctions are known, apart from

multiplying constants (as for all homogeneous problems), we can normalize them by

assuming, for example, that

2

02 Hdz

Hn =′∫

φ . (3.4.16)

This procedure is generally valid for N = N(z). To exemplify, and discuss explicit

solutions in a simple way, we assume that N is constant. Then the eigenfunctions

become

⎟⎟⎠

⎞⎜⎜⎝

⎛= z

cNA

nnn sinφ , (3.4.17)

which satisfies equation (3.4.11) and the upper boundary condition. The requirement

φn(−H) = 0 yields the eigenvalues:

,πnc

HNn

= (3.4.18)

or )/( πnHNcn = , which is identical to (3.3.31). Finally, the normalization condition

(3.4.15) gives An = H/(nπ).

We now insert the series (3.4.8) into (3.4.1) and (3.4.6), and multiply each

equation with 1φ ′ , 2φ′ , 3φ′ , etc. By integrating from z = −H to z = 0, and applying the

orthogonality condition (3.4.15), we finally obtain

⎪⎪⎪

⎪⎪⎪

∂∂

−∂∂

−=∂∂

+∂∂

−=+∂∂

+∂∂

−=−∂∂

,

,

,

)(2

)(2

yv

xu

t

ycfu

tv

xcfv

tu

nnn

yn

nnn

n

xn

nnn

n

ξ

τξ

τξ

(3.4.19)

where

⎪⎪⎭

⎪⎪⎬

′∂∂

=

′∂∂

=

−0

H

)()(

0

H

)()(

.2

,2

dzzH

dzzH

n

y

S

yn

n

x

S

xn

φτρ

τ

φτρ

τ (3.4.20)

We notice from (3.4.19) that this set of equations is formally identical to the equations

for the barotropic volume transports driven by surface winds, e.g. (2.6.15).

Page 97: Dynamic Oceanography

97

The problem in the baroclinic case is related to the form of the horizontal shear

stresses )( xτ and )( yτ in the fluid. In principle, these are unknown, and depending on

the fluid motion. However, we shall simplify the problem by assuming that we know

the vertical gradients of these stresses in the fluid.

Assume that a constant wind is blowing along a straight coast, so that the surface

wind stresses become )( xSτ ≠ 0, )( y

Sτ = 0; see the sketch in Fig. 3.6. The model is

situated in the northern hemisphere, i.e. f > 0.

Fig. 3.6. Model sketch of upwelling/downwelling at a straight coast.

We assume that the shear stresses are only felt in a relatively thin layer close to the

surface, i.e. the mixed layer, with a thickness d << H. Here the stresses vary linearly

with depth:

⎪⎪⎭

⎪⎪⎬

≤≤−=

⎪⎩

⎪⎨⎧

−≤≤−

≤≤−⎟⎠⎞

⎜⎝⎛ +

=

.0,0,,0

,0,

)(

)()(

zHdzH

zdd

dz

y

xSx

τ

ττ (3.4.21)

With this variation in z, (3.4.20) yields

⎪⎭

⎪⎬

=

−−=

.0

),(2

)(

)()(

yn

nS

xSx

n dHd

τ

φρττ

(3.4.22)

We assume that the solutions are independent of the along-shore coordinate x, i.e.,

from (3.4.19):

Page 98: Dynamic Oceanography

98

⎪⎪⎪

⎪⎪⎪

∂∂

−=∂∂

∂∂

−=+∂∂

=−∂∂

.

,

,

2

)(

yv

t

ycfu

tv

fvt

u

nn

nnn

n

xnn

n

ξ

ξ

τ

(3.4.23)

These equations have a particular solution where vn is independent of time. By

assuming that 0/ =∂∂ tvn , and eliminating un and ξn from the equations above, we

find

)(222

2 1 xn

nn

n

n

cfv

ayv

τ=−∂∂

, (3.4.24)

where fca nn /= is the Rossby radius for internal waves. By requiring that

⎭⎬⎫

∞→==

,finite,,0,0

yvyv

n

n (3.4.25)

the solution of (3.4.25) becomes

( ))/exp(1)(

n

xn

n ayf

v −−−=τ . (3.4.26)

From (3.4.23) we then obtain

⎪⎭

⎪⎬⎫

−=

−=

)./exp(

),/exp()(

)(

nn

xn

n

nx

nn

ayaf

taytu

τξ

τ (3.4.27)

Thus, un and ξn increase linearly in time during the action of the wind. From the

derived solution we see that a wind parallel to the coast results in upwelling or

downwelling within an area limited by the coast and the baroclinic Rossby radius.

Within this area we also notice the presence of a jet-like flow un parallel to the coast.

This flow is geostrophically balanced; see the second equation in (3.4.23) with

0/ =∂∂ tvn .

We now discuss our solution in some more details. For this purpose the first term

in the series (3.4.8) for v and ξ suffice:

⎭⎬⎫

+=+′=

...)(

...)(

11

11

zzvv

φξξφ

(3.4.28)

To simplify, we again take that N is constant. Then, from (3.4.17), (3.4.18) and

(3.4.22):

Page 99: Dynamic Oceanography

99

⎪⎪

⎪⎪

⎟⎠⎞

⎜⎝⎛=

=

⎟⎠⎞

⎜⎝⎛=

.sin2,/

,sin

)()(

1

1

1

Hd

d

HNcH

zH

S

xSx ππρττ

π

ππ

φ

(3.4.29)

By inserting into (3.4.28), we find

( )

⎪⎪⎭

⎪⎪⎬

+⎟⎠⎞

⎜⎝⎛−⎟

⎠⎞

⎜⎝⎛==

+⎟⎠⎞

⎜⎝⎛−−⎟

⎠⎞

⎜⎝⎛−=

.....sin)/exp(sin2

...cos)/exp(1sin2

1

)(

1

)(

Hzay

Hd

dNw

Hzay

Hd

dfv

S

xS

t

S

xS

πππρτξ

πππρτ

(3.4.30)

Here –H ≤ z ≤ 0 and f > 0. For wind in the negative x-direction ( )(xSτ < 0), we find that

w ≥ 0 in the region limited by baroclinic Rossby radius. Accordingly, the Ekman

surface layer transport away from the coast leads to a compensating flow from below

(upwelling). This is consistent with the sign of v in (3.4.30), since v is positive close

to the surface and negative near the bottom; see the sketch in Fig. 3.7.

Fig.3.7. Sketch of an upwelling situation.

We finally mention that since the x-component u and the vertical displacement ξ

increase linearly in time, the theory developed here is only valid as long as the

nonlinear terms in the equations remain small.

We return for a moment to the two-layer reduced gravity model to find out what

this would yield under similar conditions. By assuming )( xiτ = )( y

iτ = )( ySτ = V1t = 0 in

(3.1.26), we find, analogous to (3.4.24):

21

)(

121

11

cfV

aV

xS

yy ρτ

=− , (3.4.31)

Page 100: Dynamic Oceanography

100

when we take that 0/ =∂∂ x . The solution becomes

( )

⎪⎪⎪

⎪⎪⎪

−=

−=

−−−=

),/exp(

),/exp(

,)/exp(1

11

)(

1

1

)(

1

1

)(

1

aytc

h

aytU

ayf

V

xS

xS

xS

ρτρτρτ

(3.4.32)

where c1 = ( *g H1)1/2 and a1 = c1/f. We may define an upwelling velocity, when )(xSτ <

0, as

)/exp( 11

)(

11 ayc

hwx

St −−=−=

ρτ . (3.4.33)

We can now compare with the case of continuous stratification. First, we assume that

the layer of frictional influence is thin, i.e. d << H. Furthermore, we insert for z =

−H/2 to obtain the maximum vertical velocity. From (3.4.19) we then obtain

)/exp(2)2/( 11

)(

ayc

HzwS

xS −−=−=ρπτ . (3.4.34)

Here c1 = HN/π from (3.4.29). By comparing with (3.4.33), we see that the upwelling

velocities are remarkably similar, even though (3.4.34) is obtained from the first term

in a series expansion.

We will not go into further details of this problem. However, it is appropriate to

emphasize that this phenomenon is important for marine life. The water that upwells

is coming from depths below the mixed layer, and is rich in nutrients. Hence, the

upwelling process brings colder, nutrient-rich water to the euphotic zone, where there

is sufficient light to support growth and reproduction of plant algae (phytoplankton).

This means that upwelling areas are rich in biologic activity. Some of the world’s

largest catches of fish are made in such areas, e.g. off the coasts of Peru and Chile.

3.5 Baroclinic instability

When studying barotropic instability, we found that, under certain circumstances,

small perturbations in a homogeneous fluid could grow in time by extracting energy

from the kinetic energy of the basic flow. In a stratified, rotating fluid, where the

isobaric and the isopycnal surfaces are not parallel, there is another mechanism for

instability. In this case, small perturbations can grow by extracting energy from the

potential energy of the basic state. This instability is called baroclinic instability.

Page 101: Dynamic Oceanography

101

We consider the simple model described by Eady (1949). The geometry is

depicted in Fig. 3.8, and the isopycnals are assumed to be straight lines.

Fig. 3.8. Eady’s model.

The density ρ0 in the basic state can be written

)1(0 zyr γαρρ −+= , (3.5.1)

where α, γ > 0 and ρr is a constant reference density. We assume that the basic state is

geostrophically balanced, and the vertical pressure distribution is hydrostatic.

Utilizing the Boussinesq approximation, the velocity- and pressure distributions can

be written

if

zgiUvrrr α

==0 , (3.5.2)

)( 221

0 zzyzgpp rr γαρ −+−= , (3.5.3)

respectively, where pr is a constant reference pressure. We introduce small

perturbations to this basic state, i.e. we write

⎪⎭

⎪⎬

′+=

′+=

′+=

).,,,(),(),,,,(),(

),,,,()(

0

0

0

tzyxpzypptzyxzy

tzyxvzvvρρρ

rrr

(3.5.4)

Again, we assume that the (primed) perturbations are small enough for the equations

to be linearized. Furthermore, we disregard the effect of friction. The components of

the momentum equation can thus be written, respectively, as

xr

xt pfvdzdUwUuu

ρ1

−=−++ , (3.5.5)

yr

xt pfuUvvρ1

−=++ , (3.5.6)

Page 102: Dynamic Oceanography

102

gpUwwr

zr

xt ρρ

ρ−−=+

1 . (3.5.7)

Here, for simplicity, we have skipped the primes associated with the perturbations.

Furthermore, we assume that the density is conserved for a particle, i.e. 0/ =dtDρ .

Linearly, this means that

000 =∂∂

+∂∂

++z

wy

vU xtρρ

ρρ . (3.5.8)

The continuity equation then becomes

0=++ zyx wvu . (3.5.9)

We assume that f is constant. Equations (3.5.5) and (3.5.6) thus yield

0=−−⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

zy wfwf

gx

Ut

αζ , (3.5.10)

where yx uv −=ζ . Here, we have also applied (3.5.2) and (3.5.9). In studying this

problem we assume that the development in time is slow, and that the velocities and

the velocity gradients are small. We may then assume that the horizontal velocity

perturbations are quasi-geostrophic, i.e.

⎪⎪⎭

⎪⎪⎬

−≈

.1

,1

xr

yr

pf

v

pf

u

ρ

ρ (3.5.11)

Furthermore, we assume that the perturbation pressure is approximately hydrostatic in

the vertical:

zpg1

−≈ρ . (3.5.12)

When we apply the geostrophic assumption (3.5.11), the vorticity equation (3.5.10)

reduces to

02 =⎟⎟⎠

⎞⎜⎜⎝

⎛+−∇⎟⎟

⎞⎜⎜⎝

⎛∂∂

+∂∂

zyrH fwwf

gfpx

Ut

αρ . (3.5.13)

As far as the last two terms are concerned, we have that

Lf

ULf

fHgfw

wf

g

z

ymax/~ =

αα

. (3.5.14)

Page 103: Dynamic Oceanography

103

In applying a quasi-geostrophic approximation, we have implicitly assume that the

Rossby number R0 = U/(fL) is much smaller than one. Thus, we can neglect the third

term in (3.5.13) compared to the fourth term. This equation then reduces to

022 =−∇⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

zrH wfpx

Ut

ρ . (3.5.15)

Inserting for v and ρ from (3.5.11) and (3.5.12) into (3.5.8), yields

0=+⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

zrzz wgpx

Ut

γρ , (3.5.16)

where we also have applied (3.5.1) and (3.5.2). Finally, by eliminating wz between

(3.5.15) and (3.5.16), we obtain a single equation for the perturbation pressure:

022

2

=⎥⎦

⎤⎢⎣

⎡∇+⎟⎟

⎞⎜⎜⎝

⎛∂∂

+∂∂ p

fNp

xU

t Hzz . (3.5.17)

Here we have used that the Brunt-Väisälä frequency N can be written

γρ

ρg

zgN

r

=∂∂

−≡ 02 . (3.5.18)

We consider a wave solution in x and t. Then, in general, the differential operator

in (3.5.17) yields a non-zero factor. Accordingly, for (3.5.17) to be fulfilled for an

arbitrary wave component, we must have that

022

2

=∇+ pfNp Hzz . (3.5.19)

At the boundaries we assume that:

⎭⎬⎫

====

.,0,0,,0,0

HzwLyv

(3.5.20)

These are pure kinematic conditions. For this problem, they must be related to the

perturbation pressure at the boundaries, in order to solve (3.5.19). At the sidewalls, a

vanishing v in the geostrophic approximation (3.5.11) yields that

Lyp ,0,0 == . (3.5.21)

By inserting a vanishing w at the top and bottom boundaries into the density

conservation equation (3.5.8), we find:

Hzpf

gpx

Ut xz ,0,0 ==−⎟

⎠⎞

⎜⎝⎛

∂∂

+∂∂ α . (3.5.22)

Here we have also utilized (3.5.1), (3.5.11) and (3.5.12).

A wave solution that satisfies (3.5.21) can be written

Page 104: Dynamic Oceanography

104

)(exp)sin()( ctxiklyzPp −= , (3.5.23)

where Lnl /π= for any integer n. We take that k is real, while c can be complex. By

inserting this solution into (3.5.19), we find

04 22

2

=− PqPdzd , (3.5.24)

where

2/122 )(2

lkf

Nq += . (3.5.25)

The solution of (3.5.24) of can be written

)2cosh()2sinh()( qzBqzAzP += , (3.5.26)

where A and B are constants of integration. By applying (3.5.22) at z = 0, we obtain

that AgqcfB )/2( α−= , i.e.

⎥⎦

⎤⎢⎣

⎡−= )2cosh(2)2sinh()( qz

gqcfqzAzPα

(3.5.27)

By inserting (3.5.27) into (3.5.22) for z = H, we obtain an equation for the complex

phase speed c:

[ ] 01)2coth(244 222

=−+− qHqHf

gHcqcg

fq αα

. (3.5.28)

After some algebra, we find that the solution to this equation can be written

( )( ) ⎥⎦

⎤⎢⎣

⎡−−±= 2/1

*****

max 1tanh1coth112

qqqqq

Uc , (3.5.29)

where *q = qH.

Let ir iccc += in (3.5.23). Accordingly, if ci > 0, the perturbation solution will

grow exponentially in time, which means that the basic state is unstable. Oppositely,

the basic state is stable if ci < 0. Since always ** /1coth qq > for 0* >q , it is obvious

from (3.5.29) that a root where ci > 0, requires that

2.1i.e.1tanh **** ≈<< cqqqq . (3.5.30)

Inserting from (3.5.25), the condition for instability can be written

76.54)( 2*

222

22

≈<+ cqlkfHN . (3.5.31)

Unstable waves that satisfy (3.5.31), are usually referred to as Eady waves.

Page 105: Dynamic Oceanography

105

It is generally accepted that observed eddies in current systems like the Gulf

Stream and the Norwegian Coastal Current (NCC) may be caused by baroclinic

instability. Let us use NCC as an example. In a lecture note, Aas (1985) summarizes

some of the known facts about NCC. With reference to the parameters in the present

stability analysis, we can assume that H = 100 m and L = 40 km. Furthermore, we

take that the surface value of σt is 23 close to the coast, and 27 where the coastal water

borders the Atlantic inflow. This yields α = 10−7m−1 in equation (3.5.1). Insertion into

(3.5.2), with f = 1.3×10−4s−1, yields a surface current Umax= 0.76 m s−1, i.e. a mean

northward flow of 0.38 m s−1. Furthermore, if σt varies from 23 to 27 over a vertical

distance of 100 m, then N = 2×10−2s−1 in (3.5.18). We consider the first mode (n = 1)

in the y-direction, which means that l = π/L. We calculate the non-dimensional growth

rate fkci / from (3.5.29) as a function of the wavelength the λ in the x-direction by

using the values assessed here. The result is depicted in Fig. 3.9.

Fig. 3.9 Non-dimensional growth rate as a function of the wavelength for Eady waves.

We observe from the figure that maximum growth occurs for waves with λ ∼ 80 km.

This scale fits well with observed length scales for meanders in NCC (Aas, 1985).

For the most unstable Eady wave with λ = 80 km, (3.5.29) yields for the complex

phase speed in this example:

)17.05.0(max iUc ±= . (3.5.32)

Page 106: Dynamic Oceanography

106

The time Te it takes before the amplitude of the disturbance increases by a factor e

(the e-folding time) is given by

i

e kcT 1

= . (3.5.33)

The numerical value of the e-folding time for the most unstable wave in this example

is Te ≈ 27 hrs. The period for this wave, )/(2 rkcT π= , is here approximately 58 hrs.

We note that these time scales are sufficiently large to justify the application of the

quasi-geostrophic approximation, which requires that the time scale of the motion

should be considerably larger than the inertial period 2π/f (≈13 hrs in this example).

The growth rate of the disturbances calculated here is comparable with the observed

time scales for the development of meanders in NCC (Aas, 1985).

REFERENCES

Articles

Charney, J. G., and Eliassen, A.: 1949 Tellus, 1, 38.

Eady, E. A.: 1949 Tellus, 1, 33.

Eide, L. I.: 1979 Deep-Sea Res., 26, 601.

Gill, A. E., and Clarke, A. J.: 1974 Deep-Sea Res., 21, 325.

Martinsen, E. A, Gjevik, B.,

and Røed, L. P.: 1979 J. Phys. Oceanogr., 9, 1126.

Martinsen, E. A.,

and Weber, J. E.: 1981 Tellus, 33, 402.

Munk, W.: 1950 J. Met., 7, 79.

Pedlosky, J.: 1965 J. Mar. Res., 23, 207.

Stommel, H.: 1948 Trans. Am. Geophys. Un., 29, 202.

Sverdrup, H. U.: 1927 Geophys. Publ., 4, 75.

Sverdrup, H. U.: 1947 Proc. Nat. Acad. Sci., Wash., 33, 318.

Taylor, G. I.: 1916 Proc. Roy. Soc., Ser. A, XC 11, 196.

Taylor, G.I.: 1922 Proc. Roy. Soc., Ser. A, 102, 180.

Aas, E.: 1985 Lecture note, 27 pp.

Page 107: Dynamic Oceanography

107

Books

Defant, A.: 1961 Physical Oceanography, Vol. I & II. Pergamon Press 1961.

Gill, A. E.: 1982 Atmosphere-Ocean Dynamics. Academic Press 1982.

Krauss, W.: 1973 Methods and Results of Theoretical Oceanography. Vol. I

Gebrüder Borntraeger 1973.

LeBlond, P. H., and

Mysak, L. A.: 1978 Waves in the Ocean. Elsevier 1978.

Pedlosky, J.: 1987 Geophysical Fluid Dynamics, 2. ed. Springer 1987.