• Keine Ergebnisse gefunden

More on Geostrophic Flows

N/A
N/A
Protected

Academic year: 2021

Aktie "More on Geostrophic Flows"

Copied!
36
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

More on Geostrophic Flows

Chapter 4 (continued)

(2)

So far we have assumed a homogeneous, incompressible fluid:

no buoyancy forces

continuity equation   u = 0

Consider now the additional effects of having an inhomogeneous fluid; i.e., variable .

Unless the density is a function of height only, buoyancy forces must be included in the analysis.

The momentum equation becomes, assuming geostrophy,

2 1 p b

     

uk

to be defined

The effects of stratification

(3)

For an incompressible fluid we can still use the simple form of the continuity equation

   u 0 under certain circumstances.

We make the Boussinesq approximation which assumes that density variations are important only:

- inasmuch as they give rise to buoyancy forces and

- that variations in density as they affect the fluid inertia or continuity can be ignored.

may be regarded as an average density over the whole

The Boussinesq approximation

(4)

The neglect of density variations with height requires strictly that D/H

s

<< 1.

1 0

 D

Dt     u

D is the flow depth

H

S

the density height scale

The assumption is that 



/

<< 1

(z) is the average density at height z



is the maximum difference in 

(z )

The full continuity equation for an inhomogeneous compressible fluid is

in the Boussinesq approximation

(5)

The Boussinesq approximation is an excellent one in the oceans where relative density differences nowhere exceed more than one or two percent.

It is not very accurate in the atmosphere, except for motions in shallow layers (1-2 km deep).

The reason is that air is compressible under its own weight to a degree that the density at the height of tropopause, say 10 km, is only about one quarter the density at sea level.

For motions which occupy the whole depth of the troposphere, D ~ H .

Validity of the Boussinesq approximation

(6)

At any height in the atmosphere departures of  from (z) are small and an accurate form of the full continuity

equation appropriate to deep atmospheric motions is

1 0

0

u     

0

u or   (  0 u )  0

The inclusion of 

(z) - the so-called anelastic

approximation - complicates the mathematics without leading to new insights.

We shall use the Boussinesq approximation in our study of atmospheric motions.

The assumption is quite adequate for acquiring an understanding of the dynamics of these motions.

The anelastic approximation

(7)

2(    ) u    k b

Now the Taylor-Proudman theorem no longer holds.

In component form with z vertical and in the

direction of  as before, the thermal wind equation becomes

u v w b b

2 , , , ,0

z z z y x

 

    

 

               

To explore the effects of stratification we take again the curl of the momentum equation to obtain the thermal wind equation

The thermal wind equation

(8)

u v w b b

2 , , , ,0

z z z y x

 

    

 

               

Again, w = 0 at z = 0  w = 0 in the entire flow.

Later we shall show that for finite, but small Ro, w is not exactly zero, but is formally of order Ro.

Here we are considering the limit Ro0.

(9)

T

0

= T

0

(z)

0 0

0 0

0

( ) ( )

g g ,

b (T T ) (T T )

g g ,

T T

     

  

  

      



With the Boussinesq approximation, the buoyancy force can be approximated, either in terms of density or temperature as follows:

T

*

is a constant temperature analogous to 

*

(10)

cold

warm u

y

z x T 0

y

 

z

u(z)

 10 km

troposphere stratosphere

(Northern hemisphere)

   

 u

z

g T

T

o

y Thermal wind equation 2

A simple zonal flow in thermal wind balance

(11)

January

(12)

July

(13)

2      b

u k z

2(    ) u    k b

u

h

 1 k  

2 p

Compare with

The vertical wind gradient is parallel with the isotherms at any height and has low temperature on the left in the

northern hemisphere and on the right in the southern hemisphere.

The vertical wind gradient is proportional to the magnitude

of the temperature gradient.

(14)

So far we assumed that the geostrophic wind and thermal wind are in the same direction.

This happens if the isotherms have the same direction at all heights.

In general this is not the case and we consider now the situation in which the geostrophic wind blows at an angle to the isotherms.

General case

(15)

Suppose that the geostrophic wind at height z blows towards high temperature.

The geostrophic wind at height z + z, (z small), can be written

u u u

( z z ) ( ) z ( )

z z z

     

  0  2

z

u(z)

u(z + z)

z z + z

 

  uu

z z

warm

cool

(16)

Turning of the geostrophic wind with height as a result of thermal wind effects (northern hemisphere case).

T

T + T u(z) u(z + z)

u

T

T + T u(z) u(z + z)

u

Warm air advection Cold air advection

(17)

When the wind direction turns clockwise, or anticyclonic, with height in the northern hemisphere we say that the wind veers with height.

If the wind turns cyclonically with height we say it backs with height.

In the southern hemisphere 'cyclonic' and 'anticyclonic' have reversed senses, but what is confusing is that the terms "veering" and "backing" still mean turning to the right or left respectively.

Thus cyclonic means in the direction of the earth's

rotation in the particular hemisphere (counterclockwise in the northern hemisphere, clockwise in the southern

Veering and backing

(18)

In general, any air mass will have horizontal temperature gradients within it and the isotherms will be oriented

differently at different heights.

Therefore, unless the wind blows in a direction parallel with the isotherms, there will be local temperature

changes at any point simply due to advection.

If the temperature of fluid parcels is conserved during horizontal displacement, we may express this

mathematically by the equation DT/Dt = 0.

Then the local rate of change of temperature at any point,

T/t, is given by

 T

t    u T

Thermal advection

(19)

 T

t    u T

called the thermal advection

If warmer air flows towards a point u 

T < 0, and

T/t > 0.

We call this warm air advection.

It follows that there is a connection between thermal

advection and the turning of the geostrophic wind vector with height.

In the northern (southern) hemisphere, the wind veers (backs) with height in conditions of warm air advection.

It backs (veers) with height when there is cold air

(20)

It is a diagnostic equation and as such is useful, in

checking analyses of the observed wind and temperature fields for consistency.

Secondly, the z component of the thermal wind equation is 0 = 

*-1

 p/  z + b, which shows that the density-, or buoyancy field is in hydrostatic equilibrium.

Finally, the thermal wind constraint is important also in ocean current systems wherever there are horizontal

density contrasts.

Some notes concerning the thermal wind equation

(21)

When vertical motions are present, the equation DT/Dt = 0

may be inaccurate since ascent or subsidence is associated also with a thermal tendency.

D Dt

  0

However, when diabatic processes such as radiative

heating and cooling can be neglected, and provided that condensation or evaporation does not occur, the potential temperature , of an air parcel is conserved, even when the parcel ascends or subsides.

This is expressed mathematically by the formula .

The thermodynamic equation

This formula encapsulates the first law of

(22)

This is consistent with   u = 0 and .

For a Boussinesq fluid, i. e. one for which the Boussinesq approximation is satisfied, density is conserved following a fluid parcel, i.e., D

Dt

  0

In terms of the buoyancy force b, may be written in the form

Db

2

N w 0 Dt  

where is the square of the Brunt-

Väisälä frequency or the buoyancy frequency of the motion.

N

2

  ( / g 

) ( d 

0

/ dz )

1 0

 D

Dt     u D

Dt

  0

The thermodynamic equation for a Boussinesq liquid

(23)

N

2

 ( / g 

) ( d 

0

/ dz )

(z) is the basic state potential temperature distribution.

In a shallow atmosphere, the thermodynamic equation

reduces to the same form as

D Dt

  0

with b given by g(  

)/

and with N

2

replaced by

Db

2

N w 0

Dt  

(24)

This equation represents the change in buoyancy force experienced by a fluid parcel as it moves around and ascends or descends.

Db

2

N w 0

Dt  

Interpretation of

b   N

2

stably-stratified fluid

If N

2

= constant Db

2

D

2

N N w

Dt Dt

    

(25)

Some authors, including Holton, use a coordinate system in which pressure is used instead of the vertical coordinate z.

This has certain advantages:

(i) pressure is a quantity measured directly in the global meteorological data network and upper air data is

normally presented on isobaric surfaces: i.e. on surfaces p = constant rather than z = constant;

(ii) the continuity equation has a much simpler form in pressure coordinates.

A major disadvantage of pressure coordinates is that the surface boundary condition analogous to, say, w = 0 at z = 0

The use of pressure coordinates

(26)

The simplifications of the pressure coordinate systems disappear in the case of nonhydrostatic motion.

The following comparison is for the hydrostatic system of equations only.

Comparison of the equations in height and

pressure coordinates

(27)

horizontal momentum equations D

Dt w

z f p

h h h

h h

u u

   k u    

 

1

D

Dt

p h

u u p

h

f

h p

k u

      

 

is essentially gz

 is called the geopotential.

z the height of an isobaric surface

 plays the role of w in p-coordinates

  

 Dp  p     

p w p D p

gw

T T

u

T h T

.

(28)

vertical momentum equations

1 p

T

1 p

g or b

z z

    

   



p

RT

  p

continuity equation

 

h

z

h

  z z w ( ( ) )   ( ( ) )

 

0

u

0

0

 

p h

  u  p

 0

(29)

In the pressure coordinate system, p is the total pressure p

T

.

In most situations, |D

h

p

T

/Dt| << |gw| so that sgn() =  sgn(w).

Thus  negative (positive) signifies ascending (descending) motion.

In the pressure coordinate system, the subscripts p on the

operators D

p

/Dt and 

p

signify that derivatives are computed with p held fixed.

Because the isobaric surfaces are very close to horizontal, there is no practical difference between D

h

u

h

/Dt and D

p

u

h

/Dt.

Some notes

(30)

In pressure coordinates, the geostrophic equation is f k u

h

  with solution u h     f k

The thermal wind equation is frequently used in the form

D

Dt

p h

u u p

h

f

h p

    k u   

 

         

u u 500 u 1000 1 k 500 1000 1 k

mb mb

f (   ) f h '

geopotential thickness between 500 mb and 1000 mb is proportional to the mean

temperature between the two pressure surfaces.

 

h RT ln 2

(31)

Thickness charts are charts showing contours of equal thickness.

They are used by weather forecasters, inter alia, to locate regions of cold and warm air in the lower troposphere.

In Australia, the thickness isopleths are given in metres;

that is, contours of h'/g are plotted.

They are similar to 500 mb charts, because often

Thickness charts

u

500 mb

>> u

1000 mb

in magnitude.

(32)

Because of the relationship between in pressure coordinates we may obtain a thickness tendency equation

h and T

 T

t    u T The thermal tendency equation is

 h

t '

h

h

  u  '

where is a measure, in some sense, of the mean wind between 1000 mb and 500 mb.

u

h

Thickness advection

(33)

is approximately the surface wind plus a weighted measure of the thermal wind, or wind difference between 1000 mb and 500 mb.

Write where u

s

is the surface geostrophic wind, u

a

is a measure of the horizontal ageostrophic wind and  is constant.

u

h

u

s

  u u ' 

a

In many circumstances, |u

a

| is small compared with |u

s

| and |u

| ,

u h

   

     u h ' f

1

k h ' h ' 0

h ' h

   u s '

Since

(34)

Under circumstances where the temperature field is merely advected (this is not always the case), the

thickness tendency is due entirely to advection by the surface wind field.

Accordingly the surface isobars are usually displayed on thickness charts so that u

s

can be deduced readily in relation to h'.

A typical thickness chart is shown in the next figure.

 h

t ' h

   u s '

(35)

H H

H

A typical thickness chart

(36)

End of Chapter 4

Referenzen

ÄHNLICHE DOKUMENTE

Additional file 1: Predicted ground speed as a function of wind support and cross wind derived from models based on different methods (indicated by color) of determining

From the wind velocity and mean temperature profiles measured at Climate Change Tower CCT since 2009 (at heights of 33, 10, 5 and 3 m) a selection of well controlled cases has

Lower turbine prices have driven reductions in reported installed project costs Berkeley Lab also compiles data on the total installed cost of wind power projects in the United

a trajectory in the love space (see Figure 4) representing the predicted time evolution of the involvement of Scarlett and Rhett, eleven chronologically ordered short segments of

In the coastal region, wind turbines constructed directly within municipalities are negatively related and wind turbines in municipalities’ vicinities are positively related to

(2009) report that wind turbines displace on average 0.74 acres of land per MW of installed capacity; Reategui and Hendrickson (2011) reference a 2008 DOE report that found that

mation about the coastal influence on the surface wind field is obtained by calculating ageostrophic ratios of observed surface wind to analysed geostrophic wind

The negative wind stress case (Fig. 16) reveals a strongly nonlinear response of the model to the wind forcing. A westward transport, however, establishes that the strength is