23
Some examples of tropical waves Tropical planetary waves and the shallow water equations Goal: Develop the physical and mathematical foundations of tropical waves

Tropical Wave Dynamics [Lectures 11,12]

  • Upload
    lamkiet

  • View
    219

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Tropical Wave Dynamics [Lectures 11,12]

•  Some examples of tropical waves •  Tropical planetary waves and the

shallow water equations

Goal: Develop the physical and mathematical foundations of

tropical waves

Page 2: Tropical Wave Dynamics [Lectures 11,12]

Previously: MCS Hovmoller

Propagating w/ wave velocity

From Chen et al (1996)

Page 3: Tropical Wave Dynamics [Lectures 11,12]

Precipitable Water Anomalies

Page 4: Tropical Wave Dynamics [Lectures 11,12]

Kelvin Wave

Page 5: Tropical Wave Dynamics [Lectures 11,12]

Rossby Wave

Page 6: Tropical Wave Dynamics [Lectures 11,12]

!

d! v dt

= "2! # $! v " 1

%&p +

! g +! F r

Dry adiabatic primitive equations

Vector momentum equation in rotating coordinates

Continuity:

!

" #! v = 0

To derive mathematical expressions of the various types of large-scale tropical waves, we invoke a simplified form of these equations–the shallow water equations [SWE]. The approach here follows Matsuno 1966.

Recall from earlier lectures:

Page 7: Tropical Wave Dynamics [Lectures 11,12]

SWE: Principal Assumptions

•  The atmosphere [or ocean] is approximated by two layers, both of which are horizontally and vertically homogeneous at rest.

•  In the upper layer of the atmosphere, both pressure and density are horizontally invariant; in the lower layer, only density is constant.

•  The fluid is hydrostatic.

Page 8: Tropical Wave Dynamics [Lectures 11,12]

! p!z

= !"g

SWE Derivation (I) From the hydrostatic assumption:

with the last equality following from the horizontal invariance of density in either layer. The last equality implies that the lower level horizontal pressure gradient is also invariant with height.

It is convenient to replace pressure by the depth of fluid, h, in the lower layer. That is:

Taking the horizontal gradient of the hydrostatic equation in either layer gives:

!H! p!z"

#$

%

&'=!H (!g( ) = 0 = !

!z!H p( )

!!z0

h

! "H p( )dz ="H! p!z

dz0

h

! = #"H "gdz0

h

! = #"g"Hh

Page 9: Tropical Wave Dynamics [Lectures 11,12]

SWE Derivation (II) Consider the set-up as illustrated schematically on the right. Here, we take h(x,y,t) = H + η(x,y,t). We need to estimate horizontal variations in pressure in the lower layer.

Thus:

For points 1 and 2 we have [considering, w/o loss of generality, x only]: pl1 = p+!p1 = p+ "ug!z

!u!t+!vH !"Hu# fv = #

1!l

" pl"x

= # $g "#"x

h Hρl

ρu

!v!t+!vH !"Hv+ fu = #

1!l

" pl"y

= # $g "#"y

Here, gʹ′ is reduced gravity:

!

" g = g #l $ #u

#l

p p δz

δx

1 2

pl2 = p+!p2 = p+ "lg!z

lim!x!0

p+ !lg!z( )" (p+ !ug!z)!x

#

$%

&

'(= g !l " !u( ) lim

!x!0

!h /!x( )!x!x

= g(!l " !u )!h!x

The horizontal momentum equations in the lower layer become:

Page 10: Tropical Wave Dynamics [Lectures 11,12]

!w!z

= !"H #!vH

SWE Derivation (III) From the continuity equation:

The vertical velocity vanishes at the surface. On the other hand, w(h) is the rate at which the interfacial height changes, i.e.,

Thus, using h(x,y,t) = H + η(x,y,t), gives:

Integrating from the surface up to h, and noting that the horizontal velocity is independent of height [since the pressure gradient is height independent] yields:

!w!z

dz = w(h)!w(0) =0

h

" ! #H $ vH dz =0

h

" ! h#H $ vH

w(h) = !!t+!vH !"H

#

$%

&

'(h

!"!t

= !!vH "#H! ! (H +!)#H "

!vH

Page 11: Tropical Wave Dynamics [Lectures 11,12]

!

uv"

#

$

% % %

&

'

( ( (

=

U + ) u V + ) v * + ) "

#

$

% % %

&

'

( ( (

Linearizing the SWE We want to solve the SWE in perturbation form, i.e., the wave solutions are calculated with respect to a background state (U, V, and Ν):

A further assumption, appropriate close to the equator, is the equatorial beta-plane approximation, i.e., we can write f in Cartesian coordinates as:

For a motionless background state (U=V=Ν=0) and ignoring terms in quadratic in perturbation quantities gives:

!

" # u "t

= f # v $ "# %

"x

!

" # v "t

= $ f # u $ " # % "y

! !"!t

= " !g H#H $!vH!

Here, φʹ′ is:

!

" # = " g " h

!

f " #y; # =2$Re

Page 12: Tropical Wave Dynamics [Lectures 11,12]

!

" u " v " #

$

%

& & &

'

(

) ) )

=

ˆ u (y)ˆ v (y)ˆ # (y)

*

+

, , ,

-

.

/ / /

ei(kx01t )

Solving the linearized SWE (I) Consider separable solutions of the form:

Solving the first equation above for and substituting into the remaining two, and then eliminating , yields a single equation in :

Then:

!

"i#ˆ u = $y ˆ v " ik ˆ %

!

"i# ˆ v = "$y ˆ u " %ˆ &

%y

!

"i# ˆ $ = " % g H ik ˆ u + &ˆ v

&y'

( )

*

+ ,

!

ˆ u

!

" 2 ˆ v "y 2 +

# 2

$ g H% k 2 %

k#&

'

( )

*

+ , %

& 2y 2

$ g H

-

. /

0

1 2 ̂ v = 0

!

ˆ "

!

ˆ v

Page 13: Tropical Wave Dynamics [Lectures 11,12]

Solving the linearized SWE (II)

This equation is an example of an eigenvalue equation. Recall that an eigenvalue equation can be expressed as:

!

" 2 ˆ v "y 2 +

# 2

$ g H% k 2 %

k#&

'

( )

*

+ , %

& 2y 2

$ g H

-

. /

0

1 2 ̂ v = 0

H (y)!l (y) = "l!l (y)

where H is an operator, is an ψl eigenfunction, and is an λl eigenvalue. The subscript l indicates that there may be more than 1 solution. Note that each of the eigenfunctions is orthogonal (and can be made orthonormal), i.e.,

A general solution to this equation can be expressed as a linear combination of eigenfunctions:

!l*(y)!m (y)

!"

"

# = "lmHere * denotes the “complex conjugate”

! = al!l (y)l=1

"

#

Page 14: Tropical Wave Dynamics [Lectures 11,12]

Solving the linearized SWE (II)

The eigenfunctions are of the form:

v̂l (! ) = v0Hl (! )e!! 2 /2 ; ! =

""g H

#

$%%

&

'((

1/2

y!

" g H#

$ 2

" g H% k 2 % k

$#

&

' (

)

* + = 2l +1; l = 0,1,2,...

l is the meridional wavenumber

Here, the function Hl(ξ) is the lth Hermite polynomial. The first three Hermite polynomials are: H0 =1; H1 = 2ξ; H2 = 4ξ2 - 2.

Note that the eigenfunctions decay as |y|→∞. This requirement follows from the equatorial beta-plane approximation which is not valid at high latitudes, so solutions must be “equatorially trapped”.

For our equation, the eigenvalues are given by the following (“dispersion”) relationship:

Page 15: Tropical Wave Dynamics [Lectures 11,12]

Gravity/Rossby waves Since the meridional dispersion relationship is cubic in ω, it has 3 solutions. These are interpreted as (i) eastward and (ii) westward propagating equatorially trapped inertia- gravity waves and (iii) westward propagating Rossby waves. The case for l=0, i.e., v is a Gaussian centered on the equator, must be treated separately from l > 0. In particular, the dispersion relationship can be written as:

!

"# g H

$%"$ k

&

' (

)

* +

"# g H

+ k&

' (

)

* + = 0

westward eastward

The root coming from the second term in parentheses is not permitted, as this term is required not to vanish by the derivation. The remaining two roots correspond to an eastward propagating gravity wave and a westward propagating mode that asymptotes to Rossby wave-like behavior in the short wavenumber limit.

k

ω

Inertia-gravity

Rossby

Rossby-gravity

Page 16: Tropical Wave Dynamics [Lectures 11,12]

Equatorial Kelvin wave (I) The equatorial Kelvin mode has no meridional velocity perturbations, so

On the other hand, eliminating between the first and second equations gives:

!

"i#ˆ u = "ik ˆ $

!

0 = "#y ˆ u " $ˆ %

$y

!

"i# ˆ $ = " % g H(ik ˆ u )

!

ˆ "

Eliminating between the first and third equations gives:

!

ˆ "

!

" 2 # $ g Hk 2( ) ˆ u = 0% " 2

k 2 = c 2 = $ g H

!

0 = "#y ˆ u " c $ˆ u $y

Integrating this equation gives:

!

ˆ u = u0e("#y 2 / 2c )

Note that for this solution to decay away from the equator, c>0, i.e., the Kelvin wave mode propagates eastward.

Page 17: Tropical Wave Dynamics [Lectures 11,12]

Equatorial Kelvin wave (II) The structure of the equatorial Kelvin mode looks like:

The “e-folding” decay width of the Kelvin wave structure is given by:

!

YK =2c"

#

$ %

&

' (

1/ 2

For c = 30 ms-1, YK = 1600 km.

Contours show lower tropospheric pressure with positive (negative) anomalies denoted by solid (dashed) lines. The contour interval is one-fourth the maximum amplitude of the anomaly, and the zero contour is not shown. Anomalies of convergence (divergence) that are greater than two-thirds the maximum amplitude are shaded dark (light) gray. From Majda & Stechmann 2009.

In the Kelvin wave, the meridional force balance is exact geostrophic balance between the meridional pressure gradient and zonal velocity; the equatorial Kelvin mode owes its existence to the change in sign of the Coriolis force at the equator.

Page 18: Tropical Wave Dynamics [Lectures 11,12]

Kelvin vs. Rossby propagation

Contours show lower tropospheric pressure with positive (negative) anomalies denoted by solid (dashed) lines. The contour interval is one-fourth the maximum amplitude of the anomaly, and the zero contour is not shown. Anomalies of convergence (divergence) that are greater than two-thirds the maximum amplitude are shaded dark (light) gray. From Majda & Stechmann 2009.

H L C D D

P [equator] rising P [equator] falling

H L

H L

C D D

C D D

P [off equator] rising P [off equator] falling

P [equator] falling

P [off equator] falling

Page 19: Tropical Wave Dynamics [Lectures 11,12]

Rossby wave phase speeds The (rearranged) meridional dispersion relationship for Rossby and inertia-gravity waves:

Let’s examine this relationship more closely to estimate the phase speed of the associated waves. Recall that Rossby waves exist in the low-frequency limit, i.e., ω→0. Thus, in the dispersion relationship, we can ignore the term in ω3, so:

!

" 3

c02 #" k 2 +

(2l +1)$c0

%

& '

(

) * = k$; l = 0,1,2,...

!

"k

=#$

k 2 + (2l +1)($ /c0)

= #c0

(2l +1) + (c0 /$)k2

It’s immediately obvious that the phase speed of Rossby waves is negative, or westward. For the “gravest” Rossby wave mode (l = 1), we see, in the limit of small zonal wavenumber (k→0), that the magnitude of the phase speed is c0/3, or 1/3 the Kelvin wave phase speed, with higher order Rossby wave mode having smaller phase speeds. The difference in Kelvin and Rossby wave phase speeds plays an key role in the evolution of El Niño conditions.

!

"k

Page 20: Tropical Wave Dynamics [Lectures 11,12]

Inertia-gravity wave phase speeds

Inertia-gravity waves exist in the high-frequency limit, i.e., ω→∞. In this case, we can ignore the kβ term. Thus, for the nontrivial roots:

!

" 3

c02 #" k 2 +

(2l +1)$c0

%

& '

(

) * = k$; l = 0,1,2,...

!

"k

#

$ %

&

' ( 2

= c02 1+

(2l +1))k 2c0

*

+ ,

-

. /

0"k

= ±c0 1+(2l +1))k 2c0

*

+ ,

-

. /

1/ 2

Thus, the phase speed of inertia gravity waves is either positive or negative [and in this limit, equal in magnitude]. Let’s briefly revisit the SWE, in 1D (say x) and in the limit of β→0. Then:

!

" # u "t

= $" # % "x

!

" # $ "t

= % # g H " # u "x

This is a wave equation for shallow water gravity waves with phase speed c0, which we see immediately from the dispersion relationship with β→0.

Page 21: Tropical Wave Dynamics [Lectures 11,12]

Defining equivalent depth Starting from the full 3D equations for an incompressible atmosphere with constant static stability and assuming separable solutions, we can derive a system of equations like the SWE but with vertical structure dictated by an equation of the form:

!

d2"n

dz2+ µn

2"n = Fn

Here µn is the vertical wavenumber, defined as:

!

µn2 =

1HnT0

dT0dz

+gcP

"

# $

%

& ' (

14Hs

2

=N 2

gHn

(14Hs

2

Here Hs is the atmospheric scale height (g/RdT). The equivalent depth Hn is a function of wave mode. As an example, for a dry tropical wave with vertical wavelength of 5 x 103 m, with a scale height of 8 x 103 m:

!

Hn =N 2

g[µn2 + (4Hs

2)"1]#N 2

gµn2

The equivalent depth is then ~10 m. The wavespeed would be (gHn)1/2=10 ms-1

Page 22: Tropical Wave Dynamics [Lectures 11,12]

Summary of tropical wave speeds and spatial scales

*Moist wave speeds, associated with coupling to moist convection, reduces phase speed through latent heat release ahead of the wave, making the environment effectively more buoyant [i.e., lowering N].

Dry wave speed (ms-1)

Moist* wave speed (ms-1)

Spatial scale (km)

Kelvin 30-60 12-25 5-10 x 103

(zonal); 2x103 (meridional)

Rossby 10-20 5-7 4-10 x 103

(zonal); 1x103 (meridional)

Rossby-gravity 15-20 8-10 1-4 x 103

(zonal)

Page 23: Tropical Wave Dynamics [Lectures 11,12]

Wheeler-Kiladis diagram