You are on page 1of 39

Contents

1 Wind Driven Flows 3


1.1 Friction or Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Important Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Ekman Dynamics 9
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Ekman Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Ekman Spiral and Ekman Transport . . . . . . . . . . . . . . . . . . 14
2.4 Ekman Surface Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Ekman pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.6 Summary and Important Concepts . . . . . . . . . . . . . . . . . . . 21

3 Wind-Driven Oceanic Circulation 23


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 The Primitive Equations . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3 Scale Analysis and Approximations . . . . . . . . . . . . . . . . . . . 25
3.4 Homogeneous Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.5 The Sverdrup Balance . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.6 A Simple Model of Midlatitude Circulation . . . . . . . . . . . . . . . 30
3.7 Summarizing Sverdrup Theory . . . . . . . . . . . . . . . . . . . . . . 37

4 References 39

1
2 CONTENTS
Chapter 1

Wind Driven Flows

1.1 Friction or Viscosity


Newtonian fluids, such as air and water, have molecular viscosity given by:
1
F = µ∇2 v + µ∇(∇ · v) (1.1)
3
where µ is the molecular viscosity coefficient. For sea water, µ = 1.075 ×
10 kg m−1 s−1 , at 200 C.
−3

This kind of friction occurs due to random movement of fluid molecules. Two
questions can be raised here: (1) How can this molecular effect be so important for
large scale circulation? (2) How to consider this molecular effect to describe a large
scale movement?
The answer to the first question lies in the concept of Energy Cascade. This
was first addressed by Richardson in the 1930s, and was subsequently taken up also
by the mathematician Kolmogorov in the following decade. Richardson conceived
of turbulence as being composed of ‘eddies’ of different sizes. Richardson then had
the idea that large eddies are prone to break into smaller eddies, which break up
into smaller eddies, etc. etc. In each breakup step, the larger eddy transfers its
energy to the smaller ones, without dissipation, meaning that the energy transfer
process in turbulence is inviscid. This makes perfect sense, since turbulence requires
high Reynolds numbers, which in turn mean that viscous forces are negligible (with
respect to inertial forces).
But energy cannot just be transfered from large scales to smaller scales forever.
Energy dissipation is a viscous process, and energy transfer process is inviscid, so
energy is only taken out of the system at the smallest lenght scales. The diagram
below represents the energy cascade process.

Energy Dissipation ←− Energy Transfer ←− Energy Production


Small Scale ←− Large Scale
An interesting analogy can be made in a system of two parallel plates, between
which sit a bunch of rocks with characteristic size roughly equal to the distance
between the plates (fig. 1.1). Suppose we allow the plates to move tangentially and
in opposite directions to each other, generating a shear on the rocks. Eventually the
big rocks break to form smaller ones, until finally you’re left with dust that can’t be
made any smaller.

3
4 CHAPTER 1. WIND DRIVEN FLOWS

Figure 1.1: Rock breakup process to illustrate the cascade energy.

Another way to interpret the interaction between different scales in a fluid parcel
moving in the Eulerian system. From the Newton’s 2nd Law, the total variation of
momentum is due to the local term plus the advective term:

∂~p X
+ (v · ∇)~p = F~ (1.2)
∂t
Note that the advective term is nonlinear, which implies that the movement
described by the above equation cannot coexist independently. In other words,
movement in different scales interact between them. Turbulent fluctuations embed-
ded into the large scale motion drain energy of the larger scales and transfer it to
smaller scales, where they are finally dissipated.
The answer to the second question lies in the parameterizations. In order to
show how this term is parameterized, let’s consider the total velocity field in the
fluid expressed as the sum of a large-scale flow < v > plus a term representing the
smaller-scale turbulence v0 ,
v =< v > + v0 . (1.3)
Consider now the momentum equation for an incompressible homogeneous fluid
rotating about the z-axis. Using the decomposition (1.3), the component of momen-
tum equation in the x-direction, for example, takes the form:

 
∂ 0 ∂ 0 ∂ 0 ∂
+ (< u > + u ) + (< v > + v ) + (< w > + w ) (< u > + u0 )
∂t ∂x ∂x ∂z
0 1 ∂ < p > 1 ∂p0
−f (< v > + v ) = − − + Fx , (1.4)
ρ ∂x ρ ∂x
where Fx is the x-component of the viscous force (1.1). The average of (1.4) yields
an equation for < v >:

∂<u> ∂<u> ∂<u> ∂<u>


+<u> +<v> +<w>
∂t ∂x ∂x ∂z
1∂ < p >
−f < v >= − + µ∇2 < u > (1.5)
ρ ∂x
∂ < u 0 u0 > ∂ < v 0 u0 > ∂ < w 0 u 0 >
− − −
∂x ∂y ∂z
where the condition of incompressibility (∇ · v = 0) has been used. We also
assumed that for a relatively long time the average of v0 vanishes, i.e. < v0 >= 0.
1.1. FRICTION OR VISCOSITY 5

Note that, although u0 , v 0 and w0 have zero average (< u0 >=< v 0 >=< w0 >= 0),
the momentum flux of the fluctuations, which is quadractic in the fluctuation veloc-
ities, need not vanish when averaged. This is analogous to the nonzero momentum
flux due to random motion of molecules in a fluid as in Figure 1.2.

Figure 1.2: Molecules colliding with the wall and with each other transfer momentum
from the fluid to the wall, slowing the fluid velocity.

Molecules in a fluid close to a solid boundary sometime strike the boundary


and transfer momentum to it. Molecules further from the boundary collide with
molecules that have struck the boundary, further transferring the change in momen-
tum into the interior of the fluid. This transfer of momentum is molecular viscosity.
Molecules, however, travel only micrometers between collisions, and the process is
very inefficient for transferring momentum even a few centimeters. Molecular vis-
cosity is important only within a few millimeters of a boundary. It is represented
by the ratio of the stress τx tangential to the boundary of a fluid and the shear of
the fluid at the boundary. So, for example, the stress in the x-direction due to the
small-scale motions across the wall z has the form:
∂u
τxz = ρµ ≡ −ρ < w0 u0 > . (1.6)
∂z

Therefore, the momentum flux due to the small scale motion, as far as the large-
scale motion is concerned, is equivalent to a stress τxz on the large scale flow. In
terms of this stress field the averaged equations for the large-scale flow are:

∂<u> ∂<u> ∂<u> ∂<u>


+<u> +<v> +<w> − f < v >=
∂t ∂x ∂x
 ∂z 
1 ∂ < p > 1 ∂τxx ∂τyx ∂τzx
− + + + + µ∇2 < u > (1.7)
ρ ∂x ρ ∂x ∂y ∂z
∂<v> ∂<v> ∂<v> ∂<v>
+<u> +<v> +<w> + f < u >=
∂t ∂x ∂x
 ∂z 
1 ∂ < p > 1 ∂τyx ∂τyy ∂τyx
− + + + + µ∇2 < v > (1.8)
ρ ∂y ρ ∂x ∂y ∂z
∂<w> ∂<w> ∂<w> ∂<w>
+<u> +<v> +<w> =
∂t ∂x  ∂x  ∂z
1 ∂ < p > 1 ∂τzx ∂τzx ∂τzz
− + + + + µ∇2 < w > (1.9)
ρ ∂z ρ ∂x ∂y ∂z

So, the frictional forces per unit mass for an incompressible fluid (∇ · v = 0) are:
6 CHAPTER 1. WIND DRIVEN FLOWS

 
∂τxx ∂τyx ∂τzx
Fx = ρ + +
∂x ∂y ∂z
 
∂τyx ∂τyy ∂τyx
Fy = ρ + + (1.10)
∂x ∂y ∂z
 
∂τzx ∂τzx ∂τzz
Fz = ρ + +
∂x ∂y ∂z
We can make the following association, using eq. 1.6:

τxx = −ρ < u0 u0 >


τyy = −ρ < v 0 v 0 >
τzz = −ρ < w0 w0 >
τxy = τyx = ρ < u0 v 0 >
τxz = τzx = ρ < u0 w0 >
τyz = τzy = ρ < v 0 w0 >

or

< u 0 u 0 > < u0 v 0 > < u 0 w 0 >


 

τR = −ρ  < v 0 u0 > < v 0 v 0 > < v 0 w0 >  (1.11)


< w 0 u0 > < w 0 v 0 > < w 0 w 0 >
These stresses, which appear as a natural result of averaging each momentum
equation, are called the Reynolds stresses. The Reynolds stresses appear only
because of we splitted the velocity field into large-scale and turbulent small-scale
flow.
The problem now is how to specify the Reynolds stresses in terms of only the
large-scale velocities. To consider the molecular effects on the momentum equation is
impossible in practice due to the different spatial and temporal time scales compared
to the macro phenomena in oceanography.
Let’s consider a simpler 2D example, with wind in the x-direction at the ocean
surface. Wind blowing over the sea surface produces a force per unit area on the
sea surface called wind stress. The stress due to the wind is transmitted below the
ocean surface by viscous forces, resulting in a horizontal momentum transfered to
the vertical direction, as illustrated in Figure 1.3.
For small dz, the tensor is proportional to the velocity difference between z+dz
and z+dz/2 and can be represented by:

∂u
τxz = ρAz , (1.12)
∂z
where Az is the “eddy viscosity” which replaces the molecular viscosity µ in
equation 1.1. In a general case, however, there is also horizontal variation of the
vertical velocity, thus τxz is better represented as:

∂w ∂u
τxz = ρAx + ρAz ≡ ρ < w 0 u0 > (1.13)
∂x ∂z
1.1. FRICTION OR VISCOSITY 7

Figure 1.3: Momentum tranfered vertically by wind stress.

Note that the Reynolds stresses are assumed to depend in a linear way on the
spatial derivatives of the large-scale flow velocity. Similarly to (1.11), the three
components of stresses can be written as:

Ax ∂u + Ax ∂u ∂v
Ax ∂x + Ax ∂u Ax ∂w + Az ∂u
 
∂x ∂x ∂y ∂x ∂z
τ = ρ  Ay ∂u
∂y
∂v
+ Ax ∂x ∂v
Ay ∂y ∂v
+ Ay ∂y Ay ∂w
∂y
+ Az ∂v
∂z
 (1.14)
Az ∂u
∂z
+ Ax ∂w
∂x
Az ∂v
∂z
+ Ay ∂w
∂y
Az ∂w
∂z
+ Az ∂w
∂z

In analogy to (1.10), the frictional forces per unit mass is represented as:

∂ 2 ui
 
Fi = ρ AH ∇2horiz ui + AV . (1.15)
∂z 2

where Ax = Ay = AH and Az = AV are the horizontal and vertical turbulent


viscosity coefficients respectively. There is no a priori reason why AH and AV should
be identical. The assumption that an eddy viscosity A can be used to relate the
Reynolds stress to the mean flow was described in 1925 by Ludwig Prandtl, who
introduced the concept of a boundary layer. It is important to realize that this
parameterization oversimplies the problem of molecular friction. Since the details
of the turbulent flow are ignored, AH and AV cannot be obtained by theory.
In summary, turbulent fluctuations produce stresses which can be “parameter-
ized” by merely increasing the size of the viscosity from its molecular value to
account for the greater efficiency of smoothing of momentum on a large scale by
the transport of great chuncks of fluid, rather than molecules, across the averaged
momentum gradient.
Estimates of AH and AV in the atmosphere and the ocean vary enormously, see
table below.

AV (cm2 /s) AH (cm2 /s)


Near the earth’s surface 105 107
3
Ocean estimate 1 to 10 10 to 108
5
8 CHAPTER 1. WIND DRIVEN FLOWS

1.2 Important Points


• Oceanographers assume that turbulence influences flows over distances greater
than a few centimeters in the same way that molecular viscosity influences flow
over much smaller distances.

• The influence of turbulence leads to Reynolds stress terms in the momentum


equation.

• Instability in the ocean leads to mixing. Mixing across surfaces of constant


density is much smaller than mixing along such surfaces.

• Horizontal eddy diffusivity in the ocean is much greater than vertical eddy
diffusivity.

• Measurements of eddy diffusivity indicate water is mixed vertically near oceanic


boundaries such as above seamounts and mid-ocean ridges.
Chapter 2

Ekman Dynamics

2.1 Introduction

The end of 19th century was marked by several expeditions that gave important
contributions to oceanography. For instance, the Norwegian oceanographer Fridtjof
Nansen, during the Fram expedition (1893-1896), was the first to note the deflection
of surface currents. As part of a polar expedition in the late 1890s, Nansen froze
his ship Fram into the ice north of Spitzbergen Island and allowed it to drift for
more than two years. During the expedition he noticed that the drift of the boat
was generally to the right of the wind. Nansen proposed that this motion was the
result of the Coriolis force, which causes objects to veer to the right in the northern
hemisphere and to the left in the southern hemisphere. The matematical formal-
ism of this effect was only developed in 1905 by the Swedish oceanographer Vagn
Ekman. With his theoretical model of the effect on water of wind blowing over the
ocean, Ekman derived the so-called Ekman layer, Ekman spiral, Ekman transport
and Ekman pumping.
In this chapter we see the dynamical equations of Ekman’s theory.

2.2 Ekman Layer

The actual physics of the wind stress is complex and involves the generation and
breaking of wind waves. The effects of the wind stress are transmitted down into
the water column by the action of turbulent eddies generated by breaking waves and
boundary shear stresses.
The role of friction is closely related to the structure of the frictional layer which
appears on a rigid surface perpendicular to the rotation vector. The fundamental
character of the friction region, called Ekman layer, is illustrated by the following
example.
Consider the motion of a homogenous, incompressible fluid in a rotating frame
with angular velocity ω, as in Fig. 2.2. The governing equations of motion includes
then the pressure gradient force, Coriolis effect, gravity, and turbulent viscosity, as:

9
10 CHAPTER 2. EKMAN DYNAMICS

 2
∂ u ∂2u
  2 
∂u ∂u ∂u ∂u 1 ∂p ∂ u
+u +v +w − fv = − + AH 2
+ 2 + AV
∂t ∂x ∂y ∂z ρ ∂x ∂x ∂y ∂z 2
 2
∂ v ∂2v ∂2v
  
∂v ∂v ∂v ∂v 1 ∂p
+u +v +w + fu = − + AH + + AV (2.1)
∂t ∂x ∂y ∂z ρ ∂y ∂x2 ∂y 2 ∂z 2
 2
∂ w ∂2w
  2 
∂w ∂w ∂w ∂w 1 ∂p ∂ w
+u +v +w =− − g + AH 2
+ 2
+ AV
∂t ∂x ∂y ∂z ρ ∂z ∂x ∂y ∂z 2
∂u ∂v ∂w
+ + =0
∂x ∂y ∂z
Since Prandtl and his theory of boundary layers, we know that in that case the
fluid system exhibits two distinct behaviors: At some distance from the boundaries
(called the interior ) friction is usually negligible, whereas, near the wall (called the
boundary layer ), friction acts to bring the finite interior velocity to zero at the wall.
So, let’s assume now a steady, homogeneous, horizontal flow of velocity U far from
the bottom and zero velocity in the bottom (at z = 0), i.e., the velocity vanishes in
the bottom (Figure 2.2).

Figure 2.1: Representation of a uniform flow U in a rotating frame with angular


velocity ω, (f = 2ωsinθ), and zero flow at z=0.

Figure 2.2: Frictional influence of a flat bottom on a uniform flow. Source:


Cushman-Roisin (1994).

The fact that the velocity vanishes at z=0 in our idealized homogenous ocean
2.2. EKMAN LAYER 11

implies that there is friction in the bottom. In the absence of frictional forces, a
exact solution of (2.1) would be:

1 ∂p
U =−
ρf ∂y
v=0
w=0
which is in geostrophic balance in the y-direction. However, close to the bot-
tom, in the presence of friction, the flow must significantly depart from geostrophic
balance.
Problem: Determine u(z), v(z) and w(z) that satisfies

 
u = U = const
 u = 0

for z → ∞ v = 0 for z → 0 v = 0 (2.2)
 
w=0 w=0
 

Under these conditions, the continuity equation yields

∂w
= 0. (2.3)
∂z
The momentum equations (2.1), with AH = 0, are written as:

1 ∂p ∂2u
−f v = − + AV 2 (2.4)
ρ ∂x ∂z
1 ∂p ∂2v
fv = − + AV 2 (2.5)
ρ ∂y ∂z
1 ∂p
g=− (2.6)
ρ ∂z

Since the fluid is homogeneous, ρ = const,


 
∂ ∂p ∂p
, =0 (2.7)
∂z ∂x ∂y
So that the horizontal pressure gradient must be independent of z, and thus for
z → ∞, the following two equations must also hold for the entire fluid column:

1 ∂p
0=− (2.8)
ρ ∂x
1 ∂p
fU = − (2.9)
ρ ∂y

This means that the horizontal pressure gradient is determined only by the U
flux far from the boundary.
However, near the bottom, the boundary wall and friction must induce changes
in the velocity field. Therefore, we can write the horizontal velocity field as the sum
of the basic state plus departures from geostrophic flow:
12 CHAPTER 2. EKMAN DYNAMICS

(
u = ũ + U
(2.10)
v = ṽ + 0

This allows (2.4) and (2.5) to be written as:

1 ∂p ∂ 2 ũ
−f (ṽ + 0) = − + AV 2 (2.11)
ρ ∂x ∂z
1 ∂p ∂ 2 ṽ
f (ũ + U ) = − + AV 2 (2.12)
ρ ∂y ∂z
Since (2.7), we have a balance between rotation and friction, representing the
model idealized by Ekman:

∂ 2 ũ
−f (ṽ + 0) = AV (2.13)
∂z 2
∂ 2 ṽ
f (ũ + U ) = AV 2 (2.14)
∂z
D2
Taking Dz 2
of (2.13) and replacing (2.14) to eliminate ṽ, we have:

D4 ũ f2
+ ũ = 0 (2.15)
Dz 4 A2V
The general solution is the sum of the four independent homogeneous solutions:

ũ = C1 e(1+i)z/δE + C2 e(1−i)z/δE + C3 e−(1+i)z/δE + C4 e−(1−i)z/δE (2.16)


where δE is called the Ekman layer thickness and is given by
s
AV
δE = . (2.17)
|f |/2

Note that the depth of the Ekman layer depends only on AV and |f |−1/2 and
not on U . As seen before, AV is an oversimplification. Unless AV is dependent on
the applied stress, δ is independent of the stress forcing strenght, which lead us to
question how realistic is this.
Using the boundary conditions,

C1 = C2 = 0 (2.18)
U
C3 = C4 = − (2.19)
2
and the solution reduces to:

ũ = −U e−z/δE cos(z/δE ) (2.20)


ṽ = U e−z/δE sin(z/δE ) (2.21)
2.2. EKMAN LAYER 13

or

u = U [1 − e−z/δE cos(z/δE )] (2.22)


−z/δE
v = Ue sin(z/δE ) (2.23)

Note that the solution satisfies the case when z → ∞,

(
u→U
(2.24)
v → 0.

The Ekman layer is the region directly affected by the friction. It increases with
increasing AV , but decreases as increasing rotation. The stronger the rotation, the
narrower the region directly affected by viscosity and thus the thinner the Ekman
layer.

Figure 2.3: Profile of velocity components in the direction of the geostrophic velocity
U. The vertical axis is in units of the Ekman layer thickness.

Figure 2.2 shows that the flux is geostrophic in the x-direction far from the bot-
tom. When the bottom tends to approach, the retarding effect of friction decreases
u. However, the pressure gradient in the y-direction is independent of z. As u de-
creases, the Coriolis force weakens, and in the presence of the pressure force in the
y-direction, a velocity v in that direction must be produced, flowing from high to
low pressure, retarted only by fluid friction.
The geostrophic imbalance produces a flow across the isobars, from the high to
low pressure, which implies that work is being done on the fluid in the Ekman layer
by the pressure force. The rate, W , at which work is done by the pressure force in
the Ekman layer is given by:
Z ∞
DW ∂p U2
= −v dz = −ρ (2AV f )1/2 (2.25)
Dt 0 ∂y 2
We can write the kinetic energy of a geostrophic layer of fluid of depth D, moving
with velocity U, as:
14 CHAPTER 2. EKMAN DYNAMICS

U2
K=ρ
D. (2.26)
2
Unless externally maintained, the geostrophic flow will decay through viscous
dissipation in a time given by:
 
K D D
τ = DW = f −1 = √ . (2.27)
Dt
δE 2AV f
The natural decay time τ is also refered to as the “spin-down” time or “spinup”
time, that measures the effectiveness of friction. Considering the ratio between this
decay time and the inertial rotation period, we have:
1/2
D2 f

τ D −1/2
= = ≡ EV (2.28)
f −1 δE 2AV
or
AV
EV = 2
, (2.29)
f D2
which is called the Ekman Number. Note that this is a nondimensional num-
ber. We can see that for f and D constants,

EV ↑ −→ F riction ↑
EV ↓ −→ F riction ↓
In addition,

EV << 1 −→ Friction doesn’t have a strong influence on large-scale flow


EV >> 1 −→ Friction does have a strong influence on large-scale flow

2.3 Ekman Spiral and Ekman Transport


Due to the appearance of the velocity in the y-direction, the velocity vector
changes direction within the Ekman layer. As each layer of fluid is frictionally
retarded by the layer beneath it, the response of the velocity of the upper layer, due
to the planetary rotation, is to turn slowly, producing a spiral as z decreases to zero.
This spiral is termed as Ekman spiral.
Figure 2.4 shows the velocity vector as a function of z/δE . As the bottom is
approached,
v
= 1,
lim
z→0 u

so, that the velocity vector has turned 450 to the left of the geostrophic velocity.
The mass transport per unity of area in the Ekman layer can be calculated by
Z ∞ Z ∞
δE
Mx = ũdz = − U e−z/δE cos(z/δE )dz = −U (2.30)
0 0 2
δE
My = U (2.31)
2
2.3. EKMAN SPIRAL AND EKMAN TRANSPORT 15

Figure 2.4: The twist of the velocity vector within the bottom Ekman layer. The
deflection is to the left of the current above the layer.

which implies that


U
M = δE (−i + j). (2.32)
2
The stress exerted by the rigid surface at z = 0 is:
 
∂u ∂v AV
τ = −ρAV i |z=0 + j |z=0 = − ρU (i + j). (2.33)
∂z ∂z δE
Taking the vectorial product between the stress τ and the unit vector in the
z-directon:

AV ρU
τ × k = τy i − τx j = (−i + j). (2.34)
δE
Multiplying the above term by δE /δE , we have:

AV ρU δE
τ ×k= (−i + j), (2.35)
2AV /f
which gives (from equation (2.32))

τ ×k
ME = . (2.36)
ρf
As k is perpendicular to the bottom and τ is paralel to the bottom, M is a vector
paralel to the bottom and perpendicular to τ . Therefore, the Ekman transport
(ME )
- is 900 to the right of the frictional stress, in the Northern Hemisphere,
where f > 0.
- is 900 to the left of the frictional stress, in the Southern Hemisphere, where
f < 0.
Note that ME depends only on τ and not on the details of the turbulent stress
parameterization (see the absence of AV in eq. 2.36).
16 CHAPTER 2. EKMAN DYNAMICS

2.4 Ekman Surface Layer


Let’s consider the Ekman dynamics for the frictional layer generated at a free
surface in response to the wind stress, as in Fig. 2.4.

Figure 2.5: Friction layer at the surface under an applied wind stress.

Figure 2.6: The surface Ekman layer generated by a wind stress on the ocean.
Source: Cushman-Roisin (1994).

The dynamical balance from equations 2.10 to 2.12 also holds from the Ekman
layer at the ocean surface. The Ekman equations for the non-geostrophic compo-
nents only are then:

∂ 2 ũ
−f ṽ = AV (2.37)
∂z 2
∂ 2 ṽ
f ũ = AV 2 (2.38)
∂z

Differentiating twice eq. (2.38) in relation to z and substituting it in eq. 2.37,


we obtain:
2.4. EKMAN SURFACE LAYER 17

2
∂ 4 ũ

f
+ ũ = 0 . (2.39)
∂z 4 AV
Seeking a general solution of the type ũ = A eλz , we find that λ obeys the
equation above with λ = ± (1 ± i) δ1E . The general solution is given then by:

ũ(z) = A1 e(1+i)z/δE + A2 e(1−i)z/δE + A3 e−(1+i)z/δE + A4 e−(1−i)z/δE . (2.40)

The constants A1 to A4 are obtained using the boundary conditions:

1. At the surface (z = 0):

∂ ũ
τzx = −ρAV |z=0
∂z
∂ṽ
τzy = −ρAV |z=0 (2.41)
∂z
2. Below the Ekman layer (z → −∞):

ũ(z → −∞) → 0
ṽ(z → −∞) → 0 (2.42)

Using the boundary conditions (2.41) and (2.42) and assuming τzy = 0, it can
be shown that A2 = A4 = 0, A3 = iA1 , and

 
τzx δE 1−i
A1 = . (2.43)
2ρAV 2
Substituting the constants back in eq. (2.40), we have:

 
τzx δE z/δE z π
ũ(z) = √ e sin + (2.44)
2ρAV δE 4
 
τzx δE z/δE z π
ṽ(z) = − √ e sin + (2.45)
2ρAV δE 4
Therefore at the surface (z = 0), equations 2.44 and 2.45 give us:


2
ũ(z = 0) = V0 (2.46)
√2
2
ṽ(z = 0) = − V0 (2.47)
2
where,

p τzx δE τzx
V0 = u2 |z=0 + v 2 |z=0 = √ = √ , (2.48)
2ρAV ρ f AV
18 CHAPTER 2. EKMAN DYNAMICS

with δE as the Ekman depth, given by eq. (2.17). As a typical eddy viscosity is
of order 10−2 m2 s−1 and f is about 10−4 s−1 , δE is of order 10m. The depth range
over which the Ekman balance can apply is very limited, of order only tens of meters
below the sea surface.
Note that the velocity vector V0 = iu|z=0 + jv|z=0 exerced by the wind stress at
the top of the Ekman upper layer is 450 to the right (left) of the wind direction
in the Northern Hemisphere (Southern Hemisphere). Therefore, the wind
stress on a ocean surface produces a surface current of a |450 | angle to the wind
direction, as shown in Fig. 2.7. This fact explains why icebergs, which have mass
that is mostly underwater, systematically drift to the right of the wind in the North
Atlantic.
For a wind stress of 0.1P a and an assumed AV of 10−2 m2 s−1 that gives rise to
a O(10m) Ekman depth, we get

V0 ∼ (0.1)(10−3 )(10−2 .10−4 )−1/2 = 0.1m s−1 or 10cm s−1 .

Since a stress of 0.1 Pa is produced by a roughly 10 m/s wind, this calculation


suggests surface wind driven Ekman currents are typically order(100) times smaller
than the wind speed. This result also shows that the same wind produces a different
maximum surface current at different latitudes.

Figure 2.7: Ekman current generated by a 10ms−1 wind at 350 N .

Similarly to the Ekman bottom layer, we can also calculate the integrated trans-
port in the surface layer:

τ ×k
M= . (2.49)
ρf
Therefore, M is proportional to τ intensity and it is oriented perpendicu-
lar to the applied wind stress, to the right (left) in the Northern Hemisphere
(Southern Hemisphere).
Note that the magnitude of the mass flux in the surface friction layer does not
depend on the turbulent viscosity, but only on the applied stress and the Coriolis
parameter.
2.5. EKMAN PUMPING 19

Figure 2.8: Ekman spiral and Ekman transport.

2.5 Ekman pumping


Integrating the continuity equation from the bottom to the top of the Ekman surface
layer, we have:
∂Mx ∂My
+ = w0 − wE , (2.50)
∂x ∂y
where w0 is the vertical velocity at the ocean surface and wE is the vertical
velocity at the bottom of the Ekman layer. Scale analysis allows us to conclude
that w0 is neglegible compared to wE , making that the horizontal divergence of the
integrated transport in the Ekman layer is compensated by the vertical velocity at
the bottom of this layer. For this reason, wE is commonly refered to as Ekman
pumping velocity. Using the relations between M and τ , we can demonstrate
that
τ
wE = k · ∇ × . (2.51)
f
Note again that the velocity pumped into the upper Ekman layer is independent
of the small-scale turbulence.
In the Northern Hemisphere (f¿0), a clockwise wind pattern (negative curl) gen-
erates a downwelling, whereas a counterclockwise wind pattern causes upwelling.
The directions are opposite in the Southern Hemisphere.
Fig. 2.9 summarizes the relation between the integrated Ekman transport, the
pumping velocity and the wind stress. Consider the stress field in the x-direction,
but varying in intensity in the y-direction. The Ekman transport is oriented perpen-
dicular to the right of the applied wind stress. Because the wind stress intensity is
varying, the transport also varies and causes convergence and divergence in certain
regions. To conserve mass, fluid is sucked into the Ekman layer at a rate propor-
tional to the divergence of the Ekman flux (wE > 0). The inverse occurs in the
convergence regions. In the absence of an applied surface stress, the upper Ekman
layer, though it may exist, has zero intensity, i.e., the correction to the geostrophic
flow is zero.
20 CHAPTER 2. EKMAN DYNAMICS

Figure 2.9: Transport integrated in the Ekman layer due to the wind stress in the
x-direction, varying in the y-direction. Divergence in M produces upwelling in the
bottom Ekman layer (wE > 0). Convergence implies subsidence (wE < 0).

In the globe we have 3 major wind systems in each hemisphere (Figure 2.10),
i.e., the polar easterlies (trade winds) and the westerlies (midlatitude winds).

Figure 2.10: Representation of winds and current gyres.

The consequence of the Ekman transport is a net movement of water away from
equator, since ME is to the right in the Northern Hemisphere and to the left in
the Southern Hemisphere. As a consequence, there is a positive Ekman pumping
2.6. SUMMARY AND IMPORTANT CONCEPTS 21

that causes upwelling in the equator. On the other hand, where the midlatitude
westerlies and the trades coalesce, the net transport generates convergence in the
sutroipics (subtropical gyres). The dynamical consequence is an upward bulge in
the sea surface (and a downward depression of the thermocline). These bulges are
very slight, i.e. changes of order O(1m) over distances of 100’s of km. Negative
Ekman pumping wE is generated from mass conservation, and downwelling occurs
in the centre of the gyres.

2.6 Summary and Important Concepts

Assumptions of Ekman dynamics:

• No boundaries. This is valid away from coasts.

• Deep water. This is valid if depth >> 200m.

• f-plane. This is valid.

• Homogenous density. This is probably good, except as it effects stability.

• AV constant. This is generally not real.

• Steady state. This is valid in a time scale of a few days. The Ekman balance
is established over several inertial periods, i.e. the balance is not established
instantly when the wind starts blowing.

• Ekman equations is the balance between wind stress and Coriolis.

Characteristics of Ekman theory:

• Steady winds produce a thin boundary layer, the Ekman layer, at the top of
the ocean. Ekman boundary layers also exist at the bottom of the ocean and
the atmosphere. The Ekman layer in the atmosphere above the sea surface is
called the planetary boundary layer.

• Surface current directed 450 to the right(left) of the wind looking downwind
in the Northern Hemisphere (Southern Hemisphere).

• Surface current speed: 1-2.5% of wind speed depending on latitude.

• Current strength decay exponentially from the surface.

• Ekman layer depth: typical values are of order 10 to 30m, but it depends on
latitude and wind velocity.

• Transport is 900 to the right(left) of the wind in the NH (SH).

• Ekman pumping, which is driven by spatial variability of winds, drives a ver-


tical current, which drives the interior geostrophic circulation of the ocean.
22 CHAPTER 2. EKMAN DYNAMICS

• Where Ekman transports converge and diverge they generate pressure gradi-
ents that are in turn balanced by the Coriolis force, and the resulting geostroph-
ically balanced currents form the upper ocean pattern of gyres and western
boundary currents.

Realism of Ekman layer flows:

Two factors are not included in Ekman’s theory, i.e., stratification and turbu-
lence. AV is a function of wind speed at 10m only. It is assumed to be independent
of depth. This is not a good assumption. The mixed layer may be thinner than the
Ekman depth, and AV will change rapidly at the bottom of the mixed layer because
mixing is a function of stability. Mixing across a stable layer is much less than mix-
ing through a layer of a neutral stability. More realistic profiles for the coefficient
of eddy viscosity as a function of depth change the shape of the calculated velocity
profile.
The observed angle between τ and u(z = 0) is more like 50 − 200 rather than
450 . Also, a distinct Ekman velocity spiral is seldom observed in the ocean.

Figure 2.11: Annual average surface wind stress (blue arrows in stress/unit area)
and the average flow at 15 m depth (red arrows in cm/sec) which arises directly in
response to the winds. From: Ralph and Niiler (1999).

Fig. 2.11 depicts the average velocity based on thousands of satellite-tracked sur-
face drifters after the average flow resulting from the ocean’s pressure field (geostrophic
flow) is subtracted out. Hence the red arrows are the ageostrophic component of the
velocity, and correspond to the Ekman response at 15 m. Note that this is velocity,
and not the Ekman transport, which is exactly at right angles to the wind.
Chapter 3

Wind-Driven Oceanic Circulation

3.1 Introduction
Ocean physics was basically descriptive until the middle of 20th century, although
a few important scientists (e.g. Rossby, 1936) had introduced some dynamical for-
malism. The paradigm breakpoint, however, occurred in the late-40’s with Harald
Sverdrup’s (1947) work. In this work, for the first time, it is used a dynamical for-
malism to elaborate a model of the circulation driven by the wind. The main result
was that, in the ocean interior, the transport integrated into the whole column (from
the bottom to the surface) is proportional to the curl of the wind stress. Although
this result is extremely simple, it is still used nowadays as the first approximation
of the oceanic circulation. This result also served as an initial point for many other
studies in the following decade (e.g., Stommel, 1948; Munk, 1950; Fofonoff, 1954;
Charney, 1955).
In 1948, Henry Stommel showed that the circulation in oceanic gyres is asymmet-
ric because the Coriolis force varies with latitude. Two years later, Walter Munk
(1950) added eddy viscosity and calculated the circulation of the upper layers of
the Pacific. Together the three oceanographers laid the foundations for the modern
theory of ocean circulation.
During the 50’s, other several analytical models were proposed to study the
oceanic circulation. Despite their significant contribution to the understanding of
the ocean dynamics, these analytical models progressed a little when compared to
the pioneer works of Sverdrup, Stommel and Munk. These “small” progress is
explained by the high complexity of the fluid dynamics theory. Important processes
as those occurring close to the western boundary of the oceans can only be correctly
simulated when the non-linear terms are included in the momentum equation. The
inclusion of the non-linear terms makes the analytical solutions extremely difficult,
sometimes being only possible to be solved numerically.
In this chapter we see the classical model of the wind driven circulation by
Sverdrup.

3.2 The Primitive Equations


Let’s consider the momentum equation, including the pressure gradient, gravity
and friction forces:

23
24 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

 2
∂2 ∂2
  
∂v 1 ∂
+ (v · ∇)v + f k × v = ∇p − kg + AH + + AZ 2 v (3.1)
∂t ρ ∂x2 ∂y 2 ∂z

Assuming the hidrostatic approximation, the three components of this equation


become:

∂u ∂u ∂u ∂u 1 ∂p ∂2u
+u +v +w − fv = + AH ∇2H u + AZ 2 (3.2)
∂t ∂x ∂y ∂z ρ ∂x ∂z
∂v ∂v ∂v ∂v 1 ∂p ∂2v
+u +v +w + fu = + AH ∇2H v + AZ 2 (3.3)
∂t ∂x ∂y ∂z ρ ∂y ∂z
1 ∂p
0=− −g (3.4)
ρ ∂z

Recalling the mass conservation law for an incompressible fluid:

∂u ∂v ∂w
+ + = 0. (3.5)
∂x ∂y ∂z

Assuming that the density ρ is the sum of a mean value plus its deviation:

ρ = ρ̄ + ρ0 , (3.6)

where |ρ0 | << ρ̄. So, the quocient 1/ρ is

−1
ρ0

1 1 1
= 0
= 1+ (3.7)
ρ ρ̄(1 + ρ /ρ̄) ρ ρ̄

Expanding the last term as a Newton binomial, we have:

−1 "  #
0 0 2
ρ0

ρ ρ
1+ =1− +O . (3.8)
ρ̄ ρ̄ ρ̄

0 ρ0
Because |ρ0 | << ρ̄, terms of order ( ρρ̄ )2 can be neglected, compared to ρ̄
. So,

ρ0
 
1 1
∼ 1− . (3.9)
ρ ρ̄ ρ̄

Let’s consider now the Boussinesq approximation, which the density deviations
can be neglected in the horizontal components, but not in the vertical component
where (ρ̄+ρ0 ) is multiplied by g. This means that ρ can be replaced by ρ̄ in equations
3.2 and 3.3, but equation 3.4 yields:

ρ0
 
1 ∂p
0=− 1− − g. (3.10)
ρ̄ ρ̄ ∂z
3.3. SCALE ANALYSIS AND APPROXIMATIONS 25

Similarly, we can assume the pressure p as the sum of the basic value and its
deviation:

p = p̄ + p0 . (3.11)

Using then the hidrostatic approximation in equation 3.10, we have:

ρ0 ∂ p̄ ∂p0
  
1
0=− 1− + −g
ρ̄ ρ̄ ∂z ∂z
1 ∂ p̄ 1 ∂p0 ρ0 ∂ p̄ ρ0 ∂p0
0=− − + 2 + .
ρ̄ ∂z ρ̄ ∂z ρ̄ ∂z ρ̄2 ∂z

Neglecting the terms of 2nd order and taking into account ∂ p̄/∂z = −ρ̄g, we
have:

∂p0
= −ρ0 g. (3.12)
∂z
This means that the perturbation fields defined by p and ρ are also hydrostatic.

3.3 Scale Analysis and Approximations


In order to derive the equations for the large-scale oceanic circulation, we need
to consider some approximations. To do that, we need to write the momentum
equation in its dimensionless form. Let’s consider an ocean current as in Fig. 3.1,
with the following characteristic scales:
Length: U (horizontal), H (vertical)
Velocity: U (horizontal), W (vertical)
Frequency: ω0

Figure 3.1: Idealized ocean current and its characteristics scales.

Start variables are nondimensional.


26 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

∂ ∂
= ω0 ∗
∂t ∂t
∗ ∗ ∗
u = U u , xy = L(x , y )
w = W w∗ , z = Hz ∗
 
∂ ∂ 1 ∂ ∂
=
∂x ∂y L ∂x∗ ∂y ∗
From the continuity equation, we have:

∂w ∂u W U
∼ → ∼ , (3.13)
∂z ∂x H L
or recalling the aspect ratio δ = H/L,

W = δU (3.14)
w = (δU )w∗ (3.15)

For the large scale circulation, we assume δ << 1, and therefore |w| << |u|, |v|.
We also should assume that in the first order approximation, the system is in
∂p
geostrophic balance, i.e. ∂x ∼ f ρ̄u. This implies that

p = (f LρU )p∗ . (3.16)

Now we can write the nondimensional form of component x (equation 3.2), for
instance, as:

∂u∗ U 2 ∗ ∗ ∗
 
∗ ∂u ∗ ∂u ∗ ∂u
w0 U ∗ + u +u +w − f U v∗ = (3.17)
∂t L ∂x∗ ∂y ∗ ∂z ∗
∂p∗ U ∂ 2 u∗ ∂ 2 u∗ U ∂ 2 u∗
 
−f U ∗ + AH 2 + + A V
∂x L ∂x∗2 ∂y ∗2 H 2 ∂z ∗2
Dividing all member of the above equation by f U , and recalling:

U
Rossby Number ≡ R0 = (3.18)
fL

(
horizontal : EH = fALH2
Ekman Number (3.19)
vertical : EV = fAHV2

we have:

w0 ∂u∗ ∗ ∗ ∗
∂p∗ ∂ 2 u∗
 
∗ ∂u ∗ ∂u ∗ ∂u ∗ ∗2 ∗
+ R 0 u + u + w − v = − + E ∇
H H u + EV .
f ∂x∗ ∂x∗ ∂y ∗ ∂z ∗ ∂x∗ ∂z ∗2
(3.20)
3.4. HOMOGENEOUS MODELS 27

3.4 Homogeneous Models


As we are interested in large scale circulation, we apply 2 approximations:

Large spatial scale: R0 << 1


Large temporal scale: ω0 << f

The dimensionless form of equations 3.2 for the x component and similarly for
the y direction is:

∂p∗ ∂ 2 u∗
−v ∗ = − + EH ∇ ∗2 ∗
H u + EV (3.21)
∂x∗ ∂z ∗2

∂p ∂ 2v∗
u∗ = − ∗ + EH ∇∗2 H v ∗
+ E V (3.22)
∂y ∂z ∗2

Bringing back to its dimensional form:

∂p ∂2u
−ρf v + = ρAH ∇2H u + ρAV 2 (3.23)
∂x ∂z
∂p ∂2v
ρf u + + ρAH ∇2H v + ρAV 2 (3.24)
∂y ∂z

and
∂p
= −ρg (3.25)
∂z
∂u ∂v ∂w
+ + = 0. (3.26)
∂x ∂y ∂z

Let’s integrate the previous equations from the bottom to the ocean surface, as
shown in Fig. 3.2. To do this, we introduce the following definitions:
Z η
Mx = ρudz Transport integrated in the x-direction (3.27)
−H
Z η
My = ρvdz Transport integrated in the y-direction (3.28)
−H
Z η Z η
∂P ∂P ∂η ∂H
P = pdz −→ = dz + Pη − P−H . (3.29)
−H ∂x −H ∂x ∂x ∂x

Notice that Mx has dimensions of density.velocity.depth = kg s−1 m−1 .


Assuming that the products of surface pressure generated by the horizontal gra-
dient of elevation are negligible compared to the first term of the equation 3.29 and
that the bottom of the ocean is flat, then:

Z η
∂P ∂P
∼ dz (3.30)
∂x ∂x
Z−H
η
∂P ∂P
∼ dz (3.31)
∂y −H ∂y
28 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

Figure 3.2: Idealized ocean with depth H and elevation η.

The last term of the equation 3.23 vertically integrated results:


Z η Z η
∂2u
 
∂ ∂u ∂u η
ρAV 2 dz = ρ AV dz = ρAV . (3.32)
−H ∂z −H ∂z ∂z ∂z −H
However, ρAV ∂u
∂z
is the wind stress component of the u velocity in the z-direction.
Therefore,

η
∂2u
Z
ρAV dz = τxη − τx−H (3.33)
−H ∂z 2
Similarly,

η
∂2v
Z
ρAV dz = τyη − τy−H (3.34)
−H ∂z 2
The continuity equation results:

Z η Z η Z η
∂u ∂v ∂w
dz + dz + dz = 0. (3.35)
−H ∂x −H ∂y −H ∂z
Multiplying by ρ and using definitions of integrated transport (equations 3.27
and 3.28):

∂Mx ∂My
+ = −W (z = η) + W (z = −H). (3.36)
∂x ∂y
But W (z = η) = W (−H) = 0, and

∂Mx ∂My
+ = 0. (3.37)
∂x ∂y
Finally, we can write the integrated form of the horizontal components of the
momentum equation:

∂P
−f My + = τxη − τx−H + AH ∇2H Mx (3.38)
∂x
∂P
−f Mx + = τyη − τy−H + AH ∇2H My (3.39)
∂y
3.5. THE SVERDRUP BALANCE 29

Taking ∂/∂x from the 2nd equation, subtracting from the ∂/∂y of the 1st and
using β = Df /Dy and ∂M
∂x
x
+ ∂M
∂y
y
= 0, we have:

βMy = ∇ × τ 0 − ∇ × τ −H + AH ∇2 ∇ × M, (3.40)
where
∂τy−η ∂τx−η
∇ × τ0 = − (3.41)
∂x ∂y
−H
∂τy ∂τ −H
∇ × τ −H = − x (3.42)
∂x ∂y
∂My ∂Mx
∇×M = − (3.43)
∂x ∂y
The equation (3.40) represents the balance between the diverse contributions for
the vorticity of the system:
I - βMy −→ Advection of planetary vorticity through the mass field;
II - ∇ × τ 0 −→ Vorticity transferred to the ocean due to the wind stress;
III - ∇ × τ −H −→ Vorticity destroyed by the bottom friction;
IV - AH ∇2 ∇ × M −→ Vorticity dissipated by the lateral friction.

The balance between these terms characterizes the classic linear models:

I II III IV
Sverdrup X X
Stommel X X X
Munk X X X

3.5 The Sverdrup Balance


The terms III and IV can be considered small compared to the first other two, as
the vorticity induced by the surface winds is much more vigorous than those caused
by bottom and lateral friction. Sverdrup (1947) used this first approximation to
explain the large-scale circulation. Sverdrup’s regime is characterized then by the
balance between the advection of planetary vorticity and the vorticity transfered by
the wind stress, i.e.

1
My = k · (∇ × τ ). (3.44)
β
Considering the equations with the pressure gradient, Coriolis and wind stress,
we have:

∂P
−f My + = τx (3.45)
∂x
∂P
f Mx + = τy . (3.46)
∂y
30 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

The equations above are basically those we derived for the Ekman dynamics. So,
let’s write:

(
M y = M yE + M yG
(3.47)
M x = M xE + M xG

where the indices G means the geostrophic component of the balance and E is
the ageostrophic component due to Ekman dynamics:

(
MxE = f1 τy
(3.48)
MyE = − f1 τx

(
MxG = − f1 ∂P
∂y
1 ∂P
(3.49)
MyG = f ∂x

Consequently,

     
∂MxE ∂MyE ∂ τy ∂ τx τ
∇ · ME ≡ + = − =∇× (3.50)
∂x ∂y ∂x f ∂y f f
∂τy 1 ∂τx βτx 1 β
− + 2
= ∇ × τ + 2 τx (3.51)
∂x f ∂y f f f
1
∇ · ME = ∇ × τ − βMyE . (3.52)
f
For the geostrophic part of the integrated transport:

∂2P ∂2P
 
1 β
∇ · MG = − + − M yG . (3.53)
f ∂y∂x ∂x∂y f
Since the terms in the brackets vanish and considering that the divergent of the
total transport is igual zero, we re-obtain the vorticity equation:

∇ × τ = β(MyE + MyG ). (3.54)


This means that the Sverdrup balance, i.e. the equilibrium between the vor-
ticity generated by the meridional advection of planetary vorticity trough the mass
field (βMy ) and the vorticity transfered by the wind (∇ × τ ), corresponds to the
balance between the forces due to the pressure gradient, wind stress and the inertial
effect of Earth’s rotation (Coriolis).

3.6 A Simple Model of Midlatitude Circulation


Consider an idealized rectangular oceanic basin aligned with the east-west and
north-south axes, with lateral dimensions of L1 × L2 (O(L) ∼ several thousands of
km), flat bottom, and homogenous water, as in Figure 3.3. Let’s also consider a wind
distribution with maximum intensity in the low and midlatitudes, to represent the
3.6. A SIMPLE MODEL OF MIDLATITUDE CIRCULATION 31

trade winds and the westerlies, respectively (Figure 3.3). So, the wind distribution
can vary only in the meridional direction, as given by the following equation and
represented in Fig. 3.4.

τx (y) = −τ0 cos(πy/L2 ). (3.55)

Figure 3.3: Idealized oceanic basin limited by a east coast at x = 0 and west coast
at x = L1 . Source: Cushman-Roisin (1994).

Figure 3.4: Idealized wind used for the Svedrup model.


32 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

So far, we have the following scenario:

1. Winds blow over the ocean and set up an Ekman layer just below the sea
surface;

2. Because the wind stress has a nonzero curl, cross-wind transports in this up-
per layer of the ocean are convergent or divergent, and a vertical velocity is
generated that is communicated to the waters below;

3. In those water a vertical velocity can be accommodated only be convergence


or divergence of a geostrophic flow under the influence of the beta effect;

4. Thus a horizontal geostrophic flow reaches the bottom where friction sets up
a bottom Ekman layer, which generates an additional vertical velocity, with
which the interior flow must reckon.

Let’s substitute the idealized wind function (equation 3.55) into the Sverdrup
balance to calculate the integrated transport:
∂τy ∂τx ∂ τ0 π
βMy = − = (−τ0 cos(πy/L2 )) = − sin(πy/L2 )
∂x ∂y ∂y L2
τ0 π
∴ My = − sin(πy/L2 ) (3.56)
βL2
Using the fact that the flow is non-divergent (∇·M = 0), we obtain an expression
equivalent to the zonal transport gradient in the x-direction:

 
∂Mx ∂ 1
=− k · (∇ × τ ) . (3.57)
∂x ∂y β

To calculate the mass transport at a point x (where 0 < x < L1 ), let’s integrate
equation 3.57 from L1 to x, recalling that our idealized ocean basin is bounded by
two meridional coasts at x = 0 and x = L1 as in Fig. 3.3. Note that we could also
obtain Mx at x by integrating equation 3.57 from x to 0 instead, but we will see
that this gives us an unrealistic solution.

1 ∂
Mx (x) = Mx (L1 ) − (L1 − x) ∇ × τ
β ∂y
1 ∂ 2 τx
= Mx (L1 ) − (L1 − x) 2 . (3.58)
β ∂y

τ0 π 2
Mx (x) = Mx (L1 ) − cos(πy/L2 )(L1 − x). (3.59)
βL22
The normal flow at each boundary must vanish. So, the solution of equation
3.57 is:

 2
τ0 π
Mx (x) = − cos(πy/L2 )(L1 − x). (3.60)
β L2
3.6. A SIMPLE MODEL OF MIDLATITUDE CIRCULATION 33

As the flow is non-divergent, this allow us to use the streamfunction ψ = ψ(x, y),
such as:
∂ψ
= My (3.61)
∂x
∂ψ
= −Mx (3.62)
∂y
Basically, we have two solutions for ψ: the first when considering the integration
between x and x = L1 , and the second when integrating from x to 0. Figure 3.6
represents both streamline solutions. Here, we derive the realistic solution (Fig.
3.6-left panel). Integrating equation 3.61 between x and x = L1 , yields:

τ0 π
ψ(x, y) = sin(πy/L2 )(L1 − x) + ψ(L1 , y). (3.63)
βL2
But, the normal component of the transport must be zero at x = L1 , then
ψ(L1 , y) = 0.
The consequence of Sverdrup solutions is represented in Figure 3.6. The first
solution (from x and x = L1 ) requires a narrow return current in the western side of
the basin (Fig. 3.6-left panel). On the other hand, if the integration is taken from
x to 0, an opposite sign of streamfunction is found (Fig. 3.6-right panel), implying
a return flow on the eastern side of the ocean basin (Fig. 3.6-left panel).
Figure 3.6 (left panel) indicates that the meridional component of the transport
will be always equatorward while the zonal component is to the east in the superior
half of the basin and to the west the inferior part of the basin. Comparisons with
the observed gyre patterns show behaviour consistent with the slower return flow on
the eastern side of ocean basin and also strong western boundary current northerly
return flow.

Figure 3.5: Streamlines calculated from the Sverdrup solution for the stress dis-
tribution of the indealized wind in Fig. 3.4. The wind vector is paralel to the
streamlines.

Although the second solution (satisfying ψ = 0 at x = 0) is mathematically


correct, it doesn’t correspond to the pattern observed in nature and therefore it is
to be rejected in dynamic grounds. So, in order to obtain good agreement between
the theory and observations Sverdrup arbitrarily chose the transport to satisfy the
boundary condition at the eastern oceanic boundary.
Problem: Note that this solution was derived for the Northern Hemisphere.
How would be the cicrulation pattern for the Southern Hemisphere?
34 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

Figure 3.6: Consequence of Sverdrup solutions: (left panel) boundary current on


the western side, (right panel) boundary current on the eastern side. The latter is
rejected because is non-realistic. Source: Cushman-Roisin (1994).

In the idealized oceanic basin L1 × L2 in the Northern Hemisphere, the Ekman


dynamics yields a southward integrated transport in the Ekman upper layer in the
northern half of the basin and a northward transport in the southern half of the
basin. Consequently divergence occurs along the meridional domain. In the centre
of the basin, however, there is convergence, implying a center of high pressure (sub-
tropical gyre). This can cause horizontal pressure gradients that leads to an outward
movement from the convergence centre. But the pressure gradient is balanced by
the Coriolis force that is in opposite direction. The result is currents flowing in
the direction of the wind. Note that this not simply the wind dragging the
water in the same direction as the wind. See the steps below.

Transient start-up problem:

1. Water begins to flow in wind direction;

2. After first few hours, Coriolis force deflects motion to the right;

3. After first few days, net Ekman transport is to the right of the wind, confined
to the top tenths meters of the ocean;

4. A bulge of water builds up near the centre of gyre causing pressure gradient
(N-S);

5. The horizontal pressure gradient is balanced by Coriolis force;

6. The net result is a flow in the direction of the wind!

Note, however, that the currents do not necessary flow in the wind direction for
all latitudes. In a real world, with a different wind field, the flow can be oppositely
∂2
directed to τ . Looking back to equation 3.58 we see that Mx depends on ∂y 2 and not
∂ 2
only on τx . Because τx do not necessarily vanishes at the same latitudes that ∂y 2 , the

resulting flow can be in a different direction of the blowing wind. Figure 3.6 depicts
a more realistic meridional wind profile and the consequence for Ekman transport
and geostropic currents. It shows the subtropical converge centered at 300 N and
the associated high sea level that in turn induces negative Ekman pumping and
downwelling to the ocean interior. Conversely, the wind pattern at lower latitudes
3.6. A SIMPLE MODEL OF MIDLATITUDE CIRCULATION 35

causes divergence south of 200 , leading to low sea level. The mean north-south pres-
sure gradients associated with the highs and lows are balanced by the Coriolis force
of east-west geostrophic currents in the upper ocean producing a upwind current.
This means that Sverdrup theory doesn’t only ellucitates the midlatitude circula-
tion in the subtropical gyres, but also explains the existence of the observed surface
countercurrent in Pacific (and Atlantic) at a latitude between 50 and 100 N (North
Equatorial Countercurrent - NECC), where it flows against the trade winds.

Figure 3.7: Depth-integrated Sverdrup transport applied globally using the wind
stress from Hellerman and Rosenstein (1983). Contour interval is 10 Sverdrups.
From Tomczak and Godfrey (1994).

Figure 3.8: An example of how winds produce geostrophic currents running up-
wind. Ekman transports due to winds in the north Pacific (left) lead to Ek-
man pumping (center), which sets up north-south pressure gradients in the up-
per ocean. The pressure gradients are balanced by the Coriolis force due to
east-west geostrophic currents (right). Horizontal lines indicate regions where
the curl of the zonal wind stress changes sign. AK: Alaskan Current, NEC:
North Equatorial Current, NECC: North Equatorial Counter Current. Source:
http : //oceanworld.tamu.edu/resources/ocng textbook/chapter12/ .
36 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

Note that both Sverdrup solutions are incomplete because they cannot explain
what happens in the proximities of the boundaries. The Sverdrup balance only
describes the north-south component of flow, and doesn’t say anything about the
east-west flow. It satisfies the flux condition on oceanic eastern boundaries but the
western boundary currents would be necessary to satisfy mass conservation. This
conundrum indicates that considering the balance only between the terms I and II of
the vorticity equation 3.40 is not suficient to explain the western boundary currents.

Figure 3.9: Streamfunction at midlatitudes from Stommel’s solution. Source:


Cushman-Roisin (1994).

Stommel (1949) provided a closed model for the circulation in the subtropical
gyres and answered for the first time why there is an intensification of the western
boundary currents. The existence of a boundary layer implies the local breakdown
of the dynamical balances of the Sverdrup relation, i.e., that terms in the vortic-
ity equation which are negligible in the open ocean become important near the
boundaries. Based on Sverdrup theory, Stommel basically added the term III to
the vorticity balance (equation 3.40), i.e., considered the existence of the bottom
friction. Stommel showed that western boundary currents are required for flow to
circulate around an ocean basin when the Coriolis parameter varies with latitude.
This results an increase of the downstream in the west of the ocean basin that
has been termed western intensification by Stommel. With his theory, Stommel
proposed a correct explanation for the existence of the Gulf Stream, for example.
The graphical representation of Stommel solution shown in Figure 3.6 exhibits
an evident asymmetric gyre, with a slow southward flow (the Sverdrup transport)
occupying most of the domain and a swift boundary-layer current on the western
side that returns the water masses northward (for the Northern Hemisphere). This
current is to be identified with the Gulf Stream of the North Atlantic Ocean and the
respective western boundary currents in the other ocean basins (e.g., Kuroshio in
the North Pacific). Stommel’s theory proposed a reasonable theory for the existence
of the Gulf Stream.
In 1950, Walter Munk assumed the vorticity balance including the lateral dissipa-
tion term (IV in equation 3.40) due to horizontal viscosity. With Munk’s theory, the
Sverdup balance flow pattern that corresponds to the observed mean zonal (west-
east) winds in the North Pacific was computed with an astonish agreement with
observations.
3.7. SUMMARIZING SVERDRUP THEORY 37

Figure 3.10: Mass streamfunction from Munk’s solution.

3.7 Summarizing Sverdrup Theory

Assumtions of Sverdrup dynamics:

• AV ≡ eddy viscosity δ = H/L << 1 R0 << 1 ω0 << f .

• Hydrostatic approximation.

• Beta plane approximation.

• The internal flow in the ocean is geostrophic.

• There is a uniform depth of no motion. The Sverdrup solution was derived


without needing to consider any details about how the oceanic density field
arranges itself.

• We integrated momentum equations vertically over the whole water column


from the surface to the level of no motion.

• We kept the Coriolis, pressure gradient, and wind stress terms in the momen-
tum equations.

• The assumed dynamics is that there is a steady state geostrophic balance to


the net influence of the Ekman pumping.

• Ekman’s transport is correct.

Characteristics of Sverdrup theory:


38 CHAPTER 3. WIND-DRIVEN OCEANIC CIRCULATION

• Sverdrup showed that the curl of the wind stress drives a northward mass
transport, and that this can be used to calculate currents in the ocean away
from western boundary currents.

• The Sverdrup transport is the combination of geostrophic and Ekman trans-


ports together.

• The direction of the total Sverdrup=Ekman+Geostrophic depends on the sign


of the wind stress curl

• The Sverdrup transport result still holds for a continuously stratified ocean.

How real is the Sverdrup theory?

• The Sverdrup solutions only give us information about the depth-integrated


velocities. No information is available on the vertical structure of the currents.
We know, that currents may well be faster in the upper ocean. Therefore the
theory is not wrong, but is incomplete.

• The observed circulation in the ocean is very turbulent. Many years of obser-
vations may need to be averaged together to obtain a stable map of the mean
flow.

• The Gulf Stream is a region of baroclinic instability in which turbulence ac-


celerates the stream. This creates a Gulf Stream recirculation. Transports in
the recirculation region are much larger than transports calculated from the
Sverdrup-Munk theory.

• In the boundary current the Sverdrup balance doesn’t hold, but we do know
from the principle of mass conservation that the gyre scale Sverdrup transports
tells us the total mass transport of the (equal and opposite) western boundary
current.
Chapter 4

References

References for this note are:

• Pedlosky, J. 1987. Geophysical Fluid Dynamics, 2nd Edition. Springer, 710pp


+ xiv.

• Cushman-Roisin, B. 1994. Introduction to Geophysical Fluid Dynamics, Prentice-


Hall Inc., USA.

• Campos, E. Course Notes of “Oceanic Models” from University of São Paulo,


USP, Brazil.

• England, M. Course Notes from University of New South Wales, Sydney, Aus-
tralia.

• http : //marine.rutgers.edu/dmcs/ms451/2004/N otesW ilkin

• http : //oceanworld.tamu.edu/resources/ocngt extbook/chapter11/chapter110 1.htm

• http : //www − pord.ucsd.edu/ ltalley/sio210/W indf orcing/

39

You might also like