Geoscience Reference
In-Depth Information
Note that L b is sometimes called the overturning scale
because of its association with the appearance of small-
scale density overturning [e.g., Munk , 1981; Waite and
Bartello , 2006], and it is distinct from the smaller Ozmi-
dov scale (which, confusingly, is sometimes also called the
buoyancy scale). The argument of Billant and Chomaz
[2001] was based on the self-similarity of the equations of
motion when L v
by numerical simulations. Herring and Métais [1989]
attempted to obtain an inverse cascade by applying
small-scale forcing in simulations over a range of Froude
numbers, but they were unsuccessful. Lilly et al. [1998]
considered rotating stratified turbulence with small-scale
forcing; they found an inverse cascade when strong rota-
tion was present, but not for purely stratified turbulence.
Smith and Waleffe [2002] found a direct transfer of energy
from small-scale forcing into a vertically sheared mean
flow, but there was no cascade through intermediate
scales. A theoretical explanation for the lack of an inverse
cascade was given by Waite and Bartello [2004]: The
two-dimensional inverse cascade is a result of the conser-
vation of enstrophy (along with energy) by wave number
triads; the analogous quantity in stratified turbulence
is the potential enstrophy, which is approximately con-
served by triads for Fr v
L b as well as the finding that L b is the
dominant vertical scale of the zigzag instability [ Billant
and Chomaz , 2000]. However, this scaling is also implied
by Lilly [1983], since L b is the vertical scale at which the
Richardson number of layerwise two-dimensional turbu-
lence becomes O( 1 ) and, presumably, Kelvin-Helmholtz
instabilities develop. Waite and Bartello [2004] measured
L v in numerical simulations of stratified turbulence and
confirmed that L v
L b , and subsequent numerical studies
have been consistent with this finding. The inviscid limit-
ing dynamics of (8.13)-(8.16) at Fr h
1. But the relationship between
energy and (potential) enstrophy is weaker in stratified
turbulence than in two-dimensional turbulence, because
gravity waves carry some of the energy but no potential
enstrophy. Even when Fr v
1are
very different from the classical picture of layerwise two-
dimensional turbulence. Such turbulence is three dimen-
sional in the sense that horizontal and vertical advection
have the same order of magnitude, but it is anisotropic
because α
1andFr v
1, layerwise two-dimensional
turbulence eventually leaks energy into gravity waves,
which cascade downscale and destroy the possibility of an
inverse cascade. More recent data analysis after Nastrom
and Gage [1985] has also contributed to the rejection
of the inverse cascade theory, as it shows a downscale
flux of mesoscale energy below scales of around 100 km
[ Lindborg and Cho , 2001].
The lack of an inverse cascade seemed to mark the end
of the stratified turbulence hypothesis for the atmospheric
mesoscale, but it was revived in a very different form by
Lindborg [2006]. The idea that stratified turbulence natu-
rally develops a vertical scale of L b , and hence Fr v
Fr h
1.
8.2.2. Cascade Theories and Application
to Atmospheric Mesoscale
Observations of the atmospheric kinetic energy spec-
trum suggest that it has a double power law form in
horizontal wave number: At synoptic scales, it has a spec-
tral slope of
3, in agreement with QG turbulence theory
[ Charney , 1971], but in the mesoscale the slope shallows to
something resembling
O( 1 ) ,
suggests that it should be anisotropic but three dimen-
sional. Lindborg [2006] proposed a theory for stratified
turbulence with a direct cascade of energy from large
to small horizontal scales. Proceeding on dimensional
grounds along the lines of the Kolmogorov [1941] theory
for three-dimensional turbulence, Lindborg [2006] argued
that the kinetic energy spectrum should have the form
5
3 [ Nastrom and Gage , 1985; Cho
etal. , 1999]. A number of different theories were proposed
in the late 1970s and early 1980s to explain the observed
form of the mesoscale spectrum. Lilly [1983] and Gageand
Nastrom [1986] advanced a stratified turbulence hypoth-
esis: They argued that the layerwise two-dimensional
nature of stratified turbulence with small Fr v might sup-
port an inverse cascade of energy through the mesoscale,
in analogy with two-dimensional turbulence [ Kraichnan ,
1967]. Such a cascade would require a small-scale source
of kinetic energy, which Lilly [1983] speculated could be
due to moist convection at the O( 1 ) km scale. Around the
same time, a different explanation based on a direct cas-
cade of gravity wave energy was proposed [ Dewan , 1979;
VanZandt , 1982]. More recent alternatives to the strati-
fied turbulence hypothesis have included theories based on
QG [ Tung and Orlando , 2003] and surface-QG turbulence
[ Tulloch and Smith , 2009]. Waite and Snyder [2013] have
considered the effect of direct mesoscale forcing by latent
heating.
Although it was an intriguing idea, the inverse cas-
cade theory for stratified turbulence was not supported
2 / 3 k 5 / 3
h
N 2 k 3
v
E K (k h )
,
E K (k v )
,
(8.19)
in horizontal and vertical wave numbers k h and k v . He sug-
gested that the mesoscale kinetic energy spectrum could be
understood as a stratified turbulence direct cascade from
synoptic scales to the microscale. The same claim has also
been advanced by Riley and Lindborg [2008] to explain the
energy spectrum of the ocean sub-mesoscale.
Lindborg [2006] presented numerical simulations that
showed good agreement with (8.19), at least in the hori-
zontal. To minimize viscous effects and capture the strong
anisotropy at small Fr h , he used thin numerical grids
with z
x , along with separate horizontal and verti-
cal hyperviscosity to keep dissipation focused around the
 
Search WWH ::




Custom Search