PDO Notes ch3
PDO Notes ch3
3 Fundamental tools 3
3.1 Some basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.1.1 Shear stresses and Newton’s law of viscosity . . . . . 3
3.1.2 Pressure in a fluid . . . . . . . . . . . . . . . . . . . . 3
3.1.3 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.1.4 (Gauss’) Divergence theorem . . . . . . . . . . . . . . 3
3.1.5 Stokes’ theorem . . . . . . . . . . . . . . . . . . . . . . 4
3.2 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.2.1 The material Derivative . . . . . . . . . . . . . . . . . 6
3.2.2 Streamlines and streamfunctions . . . . . . . . . . . . 7
3.2.3 Strain rates . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2.4 Vorticity and circulation . . . . . . . . . . . . . . . . . 8
3.2.5 Relative motion near a point . . . . . . . . . . . . . . 8
3.3 Equations of motion . . . . . . . . . . . . . . . . . . . . . . . 9
3.3.1 Motion in a rotating frame of reference . . . . . . . . 11
3.3.2 Thin shell approximation . . . . . . . . . . . . . . . . 13
3.3.3 The β-plane . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4 Vorticity and Circulation . . . . . . . . . . . . . . . . . . . . . 17
3.5 Kinematical and dynamical approximations . . . . . . . . . . 18
3.5.1 Hydrostatic balance . . . . . . . . . . . . . . . . . . . 18
3.5.2 Hydrostatic approximation . . . . . . . . . . . . . . . 20
3.5.3 Shallow water approximation . . . . . . . . . . . . . . 20
3.5.4 Boussinesq approximation . . . . . . . . . . . . . . . 20
3.5.5 Rigid lid approximation . . . . . . . . . . . . . . . . . 20
3.6 Rossby number . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.7 Geostrophic and Thermal Wind Balance . . . . . . . . . . . . 20
3.8 The Rossby radius . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.9 The shallow-water equations . . . . . . . . . . . . . . . . . . 30
1
Page 2
Chapter 3
Fundamental tools
3.1.3 Tensors
3
Tbis very useful theorem relates a volume integral to a surface integral. Let V be a
volume bounded by a closed surface A . Consider an infinitesimal surface element
dA, whose outward unit normal is n (Figure 2. IO). The vector n d A has a magnitude
d A and direction n, and we sball write d A to mean the same thing. Let Q(x) be a
scalar, vector, or tensor field of any order. Gauss’ theorem states that
(2.30)
ndA- dA
Z Z
∂Q
dV = dAi Q (3.1)
v ∂xi A
For a vector Q:
Z Z
∂Qi
dV = dAi Q , (3.2)
v ∂xi A
Z Z
∇ · Q dV = dA · Q (3.3)
v A
The theorem relates a surface over an open surface to a line integral. Con-
sider an open surface A, with bounding curve C. Let dr be an element of
the bounding curve whose direction is that of the tangent.
Page 4
Thcn the theorem stales that
(2.34)
which signifies that thc surface integral of the curl of a vector field u is equal to the
line integral of u along thc bounding curve.
The line integral of a vector u around a closed curve C (as in Figure 2.1 1) is called
the “circulation of u about C.” This can be used to define the curl o€ a vector through
I$
the surface integral of the curl of a vector field F is equal to the line integral
of F along the bounding curve. The line integral of a vector around a
closed curve C is the circulation of the field about C.
Page 5
3.2 Kinematics
• In the LAGRANGIAN description of motion, one essentially follows
the history of an individual particle. A flow variable F (r0 , t) and its
velocity is given by ui = d(ri )/dt
• In the Eulerian case, d/dt gives the local rate of change of F at each
point r and is not the total rate of change seen by a fluid particle ...
DF ∂F ∂F
= + ui (3.5)
Dt ∂t ∂xi
Page 6
3.2.2 Streamlines and streamfunctions
The streamline
• At t = t0 , streamlines are curves that are tangent to direction of flow.
• For unsteady flows, streamlines change with time.
Let ds = (dx, dy, dz) be an element of arc length along a streamline,
and let u = (u, v, w) be the local velocity vector along that streamline,
then dx/u = dy/v = dz/w.
Page 7
The streamfunction
Page 8
3.3 Equations of motion
There can be two kind of forces acting on fluids. Body forces, and we will
restrict our attention, for now, to gravitational force per unit mass
∂( gz)
g = −∇( gz) = −k̂ = −k̂g (3.6)
∂z
and surface forces, which can be normal or tangential to the fluid. Normal
forces will be relate to pressure, whereas tangential forces will be related
to shear stresses.
In order to derive a principle of conservation of momentum we will
start by applying Newton’s law of motion to an infinitesimal element of
fluid. The continuity equation, for an element of fluid of constant density
is
∂ρ
+ ∇ · (ρu) = 0 (3.7)
∂t
and we multiply this by u:
Page 9
which reduces to the Euler equation under the assumption of frictionless
flow
D ui
ρ = ρg − ∇ p (3.15)
Dt
Page 10
3.3.1 Motion in a rotating frame of reference
Eq.3.28 is valid for an inertial or fixed frame of reference. But in GFD we
measure positions and velocities relative to a frame of reference fixed on
the surface of the Earth, which rotates w.r.t. to a frame inertial.
Let’s have a frame of reference (x1 , x2 , x3 ) rotating at a uniform angular
velocity Ω w.r.t. a fixed frame (X1 , X2 , X3 ). Any vector P is represented by
P = P1 i1 + P2 i2 + P3 i3 (3.16)
For a fixed observer, the directions of the rotating unit vectors (i1 , i2 , i3 )
change with time. The time derivatives of P is thus
dP d
= ( P1 i1 + P2 i2 + P3 i3 ) =
dt I dt
(3.17)
dP1 dP2 dP3 di1 di2 di3
i1 + i2 + i3 + P1 + P2 + P3
dt dt dt dt dt dt
For an observer rotating with (x1 , x2 , x3 ) the rate of change of P is equal
to the first three terms in Eq.3.17, and so
dP dP di di di
= + P1 1 + P2 2 + P3 3 (3.18)
dt I dt R dt dt dt
X1
x1 Ω
x2
i1
i2
X2
i3
x3
X3
Page 11
Each unit vector i traces a cone with radius sin α, where α is a constant
angle. i changes in time dt as di = sin αdθ which is the length travelled by
the top of i. The rate fo change is thus
di dθ
= sin α = sin αΩ (3.19)
dt dt
duI d
= (u R + Ω × r) R + Ω × (u R + Ω × r)
dt I dt
(3.25)
duR dr
= +Ω× + Ω × u R + Ω × ( Ω × r)
dt R dt R
The second term of the r.h.s is the Coriolis acceleration and the last term
the centripetal acceleration. This last term is added to the Newtonian grav-
ity as an effective gravity
g = gn + Ω2 r (3.27)
Page 12
The apparent force Ω2 r will be zero at the poles.
The momentum equations are now
Du 1
= g − ∇ p + ν∇2 u − (2Ω × u) (3.28)
Dt ρ
It is clear that the Coriolis force (−2Ω × u) will deflect a particle to the
right of its direction in the northern hemisphere (right-hand rule). As the
Coriolis force constantly acts normal to the fluid path, it will not accelerate
the particle (in fact, Coriolis does not play any role in the energy equation).
Page 13
Ω
z
θ
Ωy Ωz
Figure 3.2: Components of the angular velocity vector for a point on the sphere.
Or
Du 1
+ 2Ω × u = − ∇ p − g + ν∇2 u (3.34)
Dt ρ
where ν = µ/ρ is the kinematic viscosity.
Page 14
Figure 3.3: A cartesian reference system (x,y,z) and its associated spherical sys-
tem (r, θ, ϕ) around the point (a, θ0 , ϕ0 ). The plane z = 0 (or β-plane) is tangent
to the sphere around the point (a, θ0 , ϕ0 ). The approximation tan(θ − θ0 ) ≈
(θ − θ0 ) is well justified for small variations in latitude. On the β-plane, the ro-
tation vector is kΩ sinθ, where sinθ ≈ sinθ0 + (y/a)cosθ0 .
The plane z = 0, what will be called the β-plane is tangent to the sphere
in (a, θ0 , ϕ0 ). For small variations in latitude we can approximate tan(θ −
θ0 ) ≈ θ − θ0 . hence, our meridional cartesian coordinate is
y = a ( θ − θ0 )
z = r−a
x = (ϕ − ϕ0 ) a cosθ0 ,
where r is the distance of the fluid from the center of the sphere, θ is the
latitude, ϕ the longitude, and a is the radius of the Earth.
Hence, latitude θ is a linear function of y
y
θ = θ0 + . (3.35)
a
Now, for small variations in latitude we have:
y
sinθ ≈ sinθ0 + cosθ0 , (3.36)
a
as a truncated series around θ0 . And we can express f as the following:
2Ω
f = 2Ωsinθ = 2Ωsinθ0 + cosθ0 y = f 0 + β y. (3.37)
a
Page 15
1.0 3.5
3.0
0.5 2.5
β [*10 −11 ]
f/(2 ∗ Ω)
2.0
0.0
1.5
−0.5 1.0
0.5
−1.0 0.0
90S 60S 30S 0 30N 60N 90N 90S 60S 30S 0 30N 60N 90N
Latitude Latitude
Figure 3.4: The Coriolis parameter f and its meridional gradient β as a function
of latitude.
∂f 1 ∂f 2Ω
( θ = θ0 ) = ( θ = θ0 ) = cosθ0 = β. (3.38)
∂y a ∂θ a
Typical mid-latitude values for f and β are 10−4 s−1 and 10−11 m−1 s−1
(Fig. 3.4).
In conclusion, we have the β-plane approximation as
f = f 0 + βy (3.39)
For relatively large areas, with θ varying over a few tens of degrees, be-
tween mid-latitudes and the equator, the tangent plane approximation is
called β-plane. This approximation is only valid if
βy
βy ≪ f 0 or ≪ 1. (3.40)
f0
f = f 0 = 2Ω sinθ0 . (3.41)
Page 16
3.4 Vorticity and Circulation
Page 17
3.5 Kinematical and dynamical approximations
3.5.1 Hydrostatic balance
The vertical component (the component parallel to the gravitational force,
g) of the momentum equation is
Dw 1 ∂p
=− − g, (3.42)
Dt ρ ∂z
where p a is the sea surface pressure resulting from external forcing (e.g.,
atmospheric loading, sea ice, ...).
Du
+ f × u = −∇ϕ (3.45)
Dt
Dw ∂ϕ
= − + b, (3.46)
Dt ∂z
where ϕ = p/ρ0 and buoyancy b = − gρ/ρ0 . In the case of f = 0 the
horizontal momentum equation reduces to
Du
= −∇ϕ (3.47)
Dt
Page 18
and a scaling for the horizontal equation is
U Φ LU
∼ , or ∼ Φ, or U 2 ∼ Φ. (3.48)
T L T
Using mass conservation to scale vertical velocities we obtain
∂w
∇z · u + = 0. (3.49)
∂z
A scaling of this equation is
U W
+ =0 (3.50)
L H
H
W = U = αU (3.51)
L
where α ≡ HL is the aspect ratio between the typical horizontal and vertical
scales. The advective terms in the vertical momentum equation scale as
Dw W U U H U2 H
∼ = W = ( U) = 2 . (3.52)
Dt T L L L L
Now we can use the scaling for the horizontal and vertical motions, to-
gether with the aspect ratio of their typical scales, to reveal the condition
for hydrostasy.
For hydrostatic balance to hold, the ratio of advective terms to the pres-
sure gradient term in (3.46) must be
| DDwt |
≪1 (3.53)
| ∂ϕ
∂z |
| DDwt | U 2 H/L2 H 2
∼ ∼ ≪ 1. (3.54)
| ∂ϕ U 2 /H L
∂z |
Page 19
3.5.2 Hydrostatic approximation
3.5.3 Shallow water approximation
3.5.4 Boussinesq approximation
3.5.5 Rigid lid approximation
∂u 1
+ (v · ∇)u + f × u = − ∇z p, (3.56)
∂t ρ
U
Ro ≡ (3.57)
fL
Page 20
Figure 3.5: Schematic of a geostrophically balanced flow with a positive
value of the Coriolis parameter f . Flow is parallel to the lines of constant
pressure. Cyclonic flow is anticlockwise around a low pressure region.
[from Vallis (2006)]
dominates the local time derivative. The only term that can then balance
the rotation term is the pressure term, leaving us with
1 ∂p
fv ≈ (3.58)
ρ ∂x
1 ∂p
fu ≈ − . (3.59)
ρ ∂y
1 ∂p 1 ∂p
f ug = − f vg = (3.60)
ρ ∂y ρ ∂x
Page 21
the slope balanced by the Coriolis force up the slope. In the northern hemi-
sphere, the flow is anticlockwise round a region of low pressure and clock-
wise around a region of high pressure.
Consider now a plane horizontal flow in which density does not vary
along the fluid path (the Boussinesq approximation). In this case the con-
tinuity equation reduces to
∂u ∂v
+ = 0. (3.61)
∂x ∂y
∂ψ
u ≡ − , (3.62)
∂y
∂ψ
v ≡ , (3.63)
∂x
and Eq.3.61 is thus satisfied, and this is called a streamfunction.
Returning to our geostrophic balance, if the Coriolis force is constant
and if density does not vary in the horizontal, the geostrophic flow is hor-
izontally non-divergent
∂u g ∂v g
∇ z · ug = + = 0, (3.64)
∂x ∂y
p
and we may define a geostrophic streamfunction, ψg , by ψg ≡ fρ, and
∂ψ ∂ψ
ug ≡ − , vg ≡ . (3.65)
∂y ∂x
Thermal wind
Thermal wind balance arises when combining the geostrophic and hydro-
static approximations. They are useful in elucidating how temperature
differences in the horizontal can lead to vertical variations in geostrophic
velocities, hence the term thermal wind equations.
Taking the vertical derivative of the geostrophic equations for a Boussi-
nesq fluid
∂p
ρ0 f ∂ z u = − ∂ z (3.66)
∂y
∂p
ρ0 f ∂ z v = ∂ z . (3.67)
∂x
Page 22
y
f>0
ρ1
ρ2
ρ3
ρ0 f ∂ z u = g ∂ y ρ (3.68)
ρ0 f ∂z v = − g ∂ x ρ. (3.69)
These equations represent the thermal wind balance, and the vertical deriva-
tive of the geostrophic wind is the ‘thermal wind’. Thermal wind balance
says that the geostrophic velocity has a vertical thermal wind shear in
case where density has a horizontal gradient.
In general, zonally averaged ocean temperature decrease poleward due
to the differential heating received from solar radiation. Neglecting salin-
ity effects on density, this poleward reduction in temperature corresponds
to a poleward increase in density. Also, for a stably stratified fluid, density
increases with depth. In a zonally-averaged flow, ∂ x ρ = 0, and so thermal
wind reduces to
g
∂z u = ∂y ρ (3.70)
ρ0 f
This equation is telling us that, if temperature falls in the poleward direc-
Page 23
tion, ∂y ρ > 0, then the zonally-averaged thermal wind is eastward. Wind
shear also increases as we move upward in the ocean, ∂z u > 0, which
yields a surface intensified zonal velocity field. Thermal wind, although
diagnostic, represents a valid steady state balance of a frictionless rotat-
ing fluid. That is, in the presence of rotation, a flow can exist in steady
state with nonflat isopycnals. Vertical integration of the thermal wind re-
lation, along with knowledge of the geostrophic velocity at a point along
the integration path, allows for determination of the full geostrophic ve-
locity in terms of density. However, the baroclinic density field (with a
horizontal gradient) is related to the baroclinic component of the velocity
field through thermal wind balance. The barotropic flow component has
zero vertical shear.
Page 24
computed on a 1-degree global grid from observations (Fig. 3.7) as follows
Z 0
1
cn ∼ N dz (3.73)
nπ −H
Page 25
Figure 3.7: A global map of the first baroclinic gravity wave phase speed
and its zonal mean. [data from Chelton et al., 1998]
Page 26
Figure 3.8: A global map of the first baroclinic Rossby radius of deforma-
tion and its zonal mean. [data from Chelton et al., 1998]
Page 27
R. Hallberg / Ocean Modelling 72 (2013) 92–103 93
Fig. 1. The horizontal resolution needed to resolve the first baroclinic deformation radius with two grid points, based on a 1/8! model on a Mercator grid (Adcroft et al., 2010)
Figure 3.9: The oceanic resolution needed to resolve the Rossby Radius of
on Jan. 1 after one year of spinup from climatology. (In the deep ocean the seasonal cycle of the deformation radius is weak, but it can be strong on continental shelves.) This
model uses a bipolar Arctic cap north of 65!N. The solid line shows the contour where the deformation radius is resolved with two grid points at 1! and 1/8! resolutions.
deformation in an ocean model [from Hallberg et al., 2013].
of spin-up from climatology. At the coarse resolution that is typical 2. The test configuration and model
of the ocean components of CMIP5 coupled climate models (nom-
inally 1! resolution), an ocean model only resolves the deformation Phillips (1954) analyzed the baroclinic instability that arises in
radius in deep water in a narrow band within a few degrees of the a simple two-layered quasigeostrophic model of a geostrophically
equator; any important extratropical eddy effects will need to be sheared flow in a reentrant channel. This problem has the advan-
parameterized. At a much higher resolution, such as a 1/8! Merca- tage that many of the properties of the eddies, including necessary
tor grid, the deformation radius is resolved in the deep ocean in the conditions for the growth of instabilities, the growth rate, energet-
tropics and mid-latitudes, but even in this case eddies are not re- ics and vertical structure of the exponentially growing linear
solved on the continental shelves or in weakly stratified polar lat- modes can be calculated analytically, as has been documented in
itudes. An unstructured and adaptive grid ocean model could help many textbooks on geophysical fluid dynamics (e.g. Pedlosky,
to address this issue, but such models are not yet in widespread 1987; Vallis, 2006).
use for global ocean climate modeling, and even then computa- This study examines instabilities of a stacked shallow water
tional speed may dictate the use of models that do not resolve variant of the Phillips problem, which is described by the momen-
mesoscale eddies everywhere. tum and continuity equations:
In this paper, a series of numerical simulations of a variant of $ %
@un " #
^ % r & un & un ¼ 'r M n þ 1 kun k2
the Phillips (1954) model of baroclinic instability are used to þ f þk
@t 2
examine the effects of resolution on a numerical model’s ability
' r % T ' dn2 cD ku2 ku2 ; ð1Þ
to exhibit the net overturning circulation driven by mesoscale ed-
dies. The effects of a commonly used parameterization of eddy ef- h " # " #i
fect, both on the models’ explicitly resolved eddies and on the net @hn
þ r % ðhn un Þ ¼ ð3 ' 2nÞ c g3=2 x ' g3=2;Ref ' r % K h rg3=2 :
overturning, are examined. Based on these results, a simple pre- @t
scription is offered for the typical situation in global ocean mod- ð2Þ
els, where eddies are resolved in only part of the domain and in Here un is the horizontal velocity in layer n, where n = 1 for the
that portion it is desired that the model be allowed to explicitly top layer and n = 2 for the bottom layer. hn ¼ gn'1=2 ' gnþ1=2 is the
simulate their effects, but in the remainder of the domain that thickness of layer n, which is bounded above and below by inter-
eddies be entirely parameterized. Specifically, the eddy diffusivi- faces at heights gn'1=2 and gnþ1=2 . These equations are solved in a
ties should be multiplied by a ‘‘resolution function’’, ranging from 2000 m deep channel that is 1200 km long and reentrant in the
Figure 3.10: The same ocean model at different horizontal resolutions, in-
0 to 1, of the ratio of the baroclinic deformation radius to the
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x-direction, and 1600 km wide in the y-direction with vertical
model’s effective grid spacing, D e ¼ ðDx2 þ Dy2 Þ=2. The resolu- walls at the northern and southern boundaries. The Coriolis param-
creasing from right to left.
tion function that works best for the cases presented here rapidly eter, f, varies linearly in the y-direction between 6.49 & 10'5 s'1
makes a transition from 1 when this ratio is greater than a value and 9.69 & 10'5 s'1, following the common b-plane approxima-
of about 2 (the exact value is not very important and can be cho- tion. The horizontal stress tensor, T, is parameterized with a shear
sen to be higher) to 0 for larger values. In the idealized case pre- and resolution dependent Smagorinsky biharmonic viscosity (Grif-
sented here, this prescription is found to give a reasonable fies and Hallberg, 2000). The Montgomery potentials, Page 28
representation of the net eddy-driven overturning over a wide M n ¼ p=q0 þ gz, in the two layers are given by a vertical integration
range of resolutions. of the hydrostatic equation, so that
Let’s now go a little ahead of ourselves. Consider that the Corio-
lis parameter is not constant and is actually a function of latitude
f (y). The nondivergent condition ∇ · ( f u) = 0 is satisfied by the
geostrophically balanced flow. Cross-differentiating Eq.3.60 gives
∂f
v g + f ∇ z · ug = 0 (3.74)
∂y
∂w
βv g = f , (3.75)
∂z
∂f
where β ≡ ∂y . This is a geostrophic vorticity balance, also called
Sverdrup balance. In a Sverdrup balance, the vertical velocity re-
sults from an external agent, most notably wind stress. It states that
the vertical shear in the vertical velocity balances a meridional cur-
rent, with the Coriolis parameter f and the planetary vorticity gra-
dient β determining the sense and strength of the meridional flow.
A vertical velocity shear arises when there is a nonzero curl in the
wind stress acting on the ocean surface. Vorticity is then transferred
to the ocean via frictional effects causing Ekman pumping or suction.
These effects alter the vertical structure of the vertical velocity and,
through Sverdrup balance, induce a meridional flow.
Page 29
3.9 The shallow-water equations
To describe large-scale oceanic, and atmospheric, motions, where the hor-
izontal scale is much larger than the vertical scale, we can use a set of
simplified equations that retain the necessary ingredients of the fluid mo-
tion but use some useful approximations. We will thus consider a fluid
in hydrostatic balance of constant density and, for simplicity, we will also
consider a flat bottom. The necessary condition of the shallow-water equa-
tions is that the horizontal length scale must be much larger than the ver-
tical scale over which the fluid develops so that L >> H.
If the fluid is in hydrostatic balance
∂p
= −ρg. (3.76)
∂z
Then the total pressure will be
p( x, y, z, t) = −ρgz + p′ . (3.77)
p = p0 + p ′ = 0 (3.78)
and at z = η we have
p′ = ρgη (3.79)
Our total pressure will then be
p( x, y, z, t) = ρg(η ( x, y) − z) (3.80)
This means that the horizontal gradient of pressure, and the flow, is inde-
pendent of depth
∇ p = ρg∇η (3.81)
and the horizontal momentum equations reduce to
Du 1
= − ∇ p = − g∇η (3.82)
Dt ρ
We can now easily add rotation to our shallow-water momentum equa-
tions
Du 1
+ f × u = − ∇ p = − g∇η (3.83)
Dt ρ
The continuity equation is obtained by the mass balance within an in-
finitesimal column of fluid. The mass flux passing through a section of the
Page 30
η
Fm
h H
Z=0 δy
δx
column is Fm = ρu( H + η )δy and the difference between the fluxes into
and out of the section is given by
∂
δxδy [ρu( H + η )] (3.84)
∂x
Considering the total volume, the net rate of change is
∂h ∂ ∂
+ [u( H + η )] + [v( H + η )] = 0 (3.85)
∂t ∂x ∂y
∂h ∂ ∂
+ (uh) + (vh) = 0 (3.86)
∂t ∂x ∂y
∂h
+ ∇ · (uh) = 0 (3.87)
∂t
and if the perturbation is small and H is constant, mass continuity reduces
to the linear equation
∂η
+ H∇ · u = 0 (3.88)
∂t
If there is flux by advection this is balanced by a net increase in mass and
an increase in height, giving rise to a vertical velocity, so that the mass
convergence is balanced by the increase in height allowing for a dynami-
cal surface elevation. This will be the basis for the propagation of waves
within the rotating shallow-water system.
Page 31
Exercices
1. Use ϕ = p/ρ0 and the definition of buoyancy b = − gρ/ρ0 to rewrite
the hydrostatic balance and thermal wind equations.
Page 32
Bibliography
33