1664
JOURNAL OF PHYSICAL OCEANOGRAPHY
VOLUME 33
Erosion of a Surface Vortex by a Seamount
STEVEN HERBETTE
YVES MOREL
AND
EPSHOM/CMO, Brest, France
MICHEL ARHAN
LPO, CNRS/IFREMER/UBO, Plouzane, France
(Manuscript received 1 July 2002, in final form 12 December 2002)
ABSTRACT
Numerical experiments are carried out on the f plane, using a shallow-water isopycnal model, to analyze the
behavior of a surface-intensified anticyclonic vortex when it encounters an isolated seamount. The advection
by the vortex of deep fluid parcels across the isobaths is known to generate deep anticyclonic and cyclonic
circulations above and near the bathymetry, respectively. These circulations are shown to exert a strong shear
on the upper layers, which causes an erosion of the initial vortex by filamentation. The erosion often results in
a subdivision of the eddy. While the eroded original structure forms a dipole with the deep cyclone and is
advected away, the filaments torn off from the original core aggregate into a new eddy above the seamount.
Splitting in more than two structures is sometimes observed. The erosion process is quantified by the bulk
volume integral of the eddy potential vorticity anomaly. A sensitivity study to different parameters of the
configuration (distance between vortex and seamount, vortex radius, seamount radius, seamount height, or
stratification) shows that the intensities of the deep anticyclonic and cyclonic circulations and the vortex erosion
are governed both by the reservoir of positive potential vorticity associated with the seamount and by the strength
of the cross-isobath flow induced by the eddy.
1. Introduction
This study was initially motivated by observations of
Agulhas ring subdivisions caused by seamounts, shortly
after their formation. Agulhas rings are surface-intensified large anticyclonic vortices that are spawned from
the Agulhas current retroflection and then propagate
northwestward in the South Atlantic. Because they are
associated with high sea surface height perturbations,
one can take advantage of the data provided by satellite
altimetry to locate and track them (e.g., Gordon and
Haxby 1990). Using this technique, Arhan et al. (1999)
back-tracked some rings that they had previously studied at sea. They then suggested that two of them resulted
from a single structure that had split when passing over
isolated seamounts, at the exit of the Agulhas retroflection region. Schouten et al. (2000), by also analyzing
altimetric data, showed evidence that such splitting
events were not exceptional and confirmed that subdivision was linked to the eddy propagation over isolated
seamounts. Those authors pointed out that of 21 rings
formed between 1993 and 1996, 6 were subdivided at
least once. Therefore, the subdivision process, which
Corresponding author address: Steven Herbette, EPSHOM, CMO/
RE, 13, rue du Chatellier, B.P. 30316, Brest CEDEX 29603, France.
E-mail:
[email protected]
q 2003 American Meteorological Society
influences the paths of the structures and their rates of
dissipation, is of significant importance, at least for
Agulhas rings and likely for other vortex species.
So far, many studies have focused on the displacement
of oceanographic vortices over a bottom slope. There
actually exist analogies between the planetary b effect
and the effect of constant bottom slopes (named topographic b effect): both effects are similar for barotropic
flows. When stratification is taken into account, some
differences exist, but the influence of both processes on
the dynamics of vortices relies on their associated potential vorticity (hereinafter referred to as PV) gradient
(McWilliams and Flierl 1979; Nof 1983; Smith and
O’Brien 1983; Mory 1985; Sutyrin and Flierl 1994; Sutyrin and Morel 1997; Morel and McWilliams 1997).
This PV gradient generally induces a westward1 displacement for all vortices, with an additional northward
or southward component for cyclonic or anticyclonic
vortices, respectively.
For dissipation, because an oceanic vortex can be
scattered by Rossby waves on the planetary b plane
(McWilliams and Flierl 1979), a similar effect is expected for the topographic b. Kamenkovich et al.
(1996), LaCasce (1998), and Thierry and Morel (1999)
1
For the topographic b effect, north is toward shallow depths.
AUGUST 2003
1665
HERBETTE ET AL.
showed that a vortex propagating over a steep bottom
slope evolves toward a compensated state in which there
is no motion in the lower layers. Dispersion also strongly depends on the vortex size: for large vortices, topographic waves have a strong signature in the surface
layers also and contribute to the dispersion of the whole
structure, even in the upper layers (Thierry and Morel
1999).
The encounter of an eddy with an isolated topographic
obstacle is different and more complicated than with a
constant slope. Carnevale et al. (1991) have carried out
numerical and tank experiments for small barotropic
vortices. They show that, when approaching an isolated
seamount, the path of small vortices is downhill for
anticyclones whereas cyclones climb up the seamount
with trajectories describing spirals. For larger vortices,
their experiments led them to the conclusion that the
trajectories are then very sensitive to the initial conditions, because of the formation of topographic vortices.
In fact, Huppert (1975), Huppert and Bryan (1976), Verron and Le Provost (1985), and Smith (1992) showed
how a large-scale current over an isolated seamount can
lead to the formation of a cyclone downstream and an
anticyclone upstream of the seamount. In the experiments of Carnevale et al. (1991), these newly formed
eddies are generated by the original vortex and can in
turn interact with the latter. However, Carnevale et al.
(1991) have not explored the consequences of such interactions. Van Geffen and Davies (2000) have more
recently considered larger barotropic vortices on the b
plane. They clearly show that the formation of topographic eddies can strongly affect the original vortex
path. An approaching anticyclone can either be deflected
from its planetary b trajectory or form a coherent dipole
with the cyclonic eddy. This modon (cyclone–anticyclone vortex pair) is found to be able to self-propagate
away from the topography in a direction opposite to the
initial incoming path of the vortex. Nevertheless, neither
of those studies mentioned a splitting event or gave
information about the dissipation associated with eddy–
seamount encounters. This aspect has been studied in
the case of collision of a vortex with an island (Wang
and Dewar 2003; Simmons and Nof 2000; Cenedese
2002; Dewar 2002). In this case, as the incoming vortex
hits the sidewall of the obstacle, part of the eddy mass
situated on the edge forms a jet along the island, which
erodes the vortex. When the jet emerges on the other
side of the seamount, it leads to the formation of a new
eddy. Dewar (2002) recently showed that an isolated
seamount could have drastic consequences on the dynamics of a vortex constituted of opposite-sign PV
anomalies initially vertically aligned. As in Morel and
McWilliams (1997), a hetonic structure usually emerges
and modifies the propagation of the structure, but it is
also shown that the lower-layer PV pole can be trapped
by the seamount and can separate from the upper-layer
vortex. The erosion of the latter has, however, not been
studied.
In this paper, we extend the Carnevale et al. (1991)
and Van Geffen and Davies (2000) studies to a stratified
ocean. We focus on the erosion of a surface-intensified
anticyclonic vortex as it encounters a seamount, and, in
particular, we analyze the processes leading to filamentation and splitting of the anticyclone. We limit the experiments to the f plane, after observing from a few
trials that this was a useful first step toward the understanding of the more complex b-plane flows. The f plane simplification also makes it easier to analyze the
influence of several eddy or seamount parameters on
the erosion process. In section 2, we describe the numerical model and present the setup of the numerical
experiments. Section 3 is dedicated to a basic experiment in which we show that a vortex can split as it
encounters a seamount, and we discuss the mechanisms
leading to this splitting event. The sensitivity of the eddy
erosion to several parameters (distance between vortex
and seamount, vortex radius, seamount radius, seamount
height, and stratification) is studied in section 4, before
a brief conclusion.
2. Model and initial configuration
a. Equations and numerical model
We consider the primitive equations written in isopycnal coordinates (Bleck and Boudra 1986; Bleck and
Smith 1990; Bleck et al. 1992), which, when discretized
vertically, boil down to the well-known shallow-water
equations. In this study, we consider an ocean constituted of three layers, and the equations are then for each
layer k 5 1 to 3:
] t u k 1 (u k · =)u k 2 f y k 5 2] x M k 1 Fx ,
] t y k 1 (u k · =)y k 1 f u k 5 2] y M k 1 Fy ,
] t h k 1 div(h k u k ) 5 0.
and
(1)
Here u 5 (u, y ) is the horizontal velocity field, f is the
constant Coriolis frequency, h is the thickness of an
isopycnal layer, and M is the Montgomery potential
(pressure along an isopycnal surface), which can be related to h:
O gh 1 O r 2r r gh ,
i53
Mk 5
i5k21
i
k
i
i51
i
i51
(2)
o
where r k is the density of layer k, ro is a reference
density, and g is the earth gravity. Last, F 5 (F x , F y )
represents isopycnal diffusion of momentum and is generally associated with a harmonic or biharmonic operator. When the latter is neglected, PV is conserved for
each particle of the flow (see Ertel 1942; Pedlosky
1987). For each layer (bounded by isopycnal surfaces),
PV is written
PV 5
z1 f
,
h
where z 5 ] x y 2 ] y u is the relative vorticity. It is also
1666
JOURNAL OF PHYSICAL OCEANOGRAPHY
convenient to define another quantity that we will refer
to as PV anomaly (PVA):
PVA 5 H
1
2
[
]
z1 f
f
H
f (h 2 H )
2
5 z2
,
h
H
h
H
(3)
where H is constant and represents the layer thickness
at rest. Because PVA is a linear function of PV it has
the same conservation properties. In the following, PVA
will thus be used as a tracer, in particular to evaluate
the erosion of the vortex core. Also notice that in the
present configuration PVA filters out the value of the
PV at rest and is thus a good indicator of the strength
of the (geostrophic) circulation (or vortex strength here).
Last, in the limit of small layer-depth variations, PVA
reduces to z 2 f (h 2 H)/H, which is the well-known
expression of PV in the quasigeostrophic framework,
with z representing the relative vorticity and 2 f (h 2
H)/H the stretching effect.
Equations (1) are solved numerically using the Miami
Isopycnal Coordinate Ocean Model (MICOM: Bleck
and Boudra 1986; Bleck and Smith 1990; Bleck et al.
1992). MICOM has, however, been modified to include
a fourth-order scheme for the nonlinear advective terms
in the momentum equations and a biharmonic diffusion
operator for F (Morel 2001):
Fx 5 ] x (n ] x 3 u) 1 ] y (n ] y 3 u)
and
Fy 5 ] x (n ] x 3 y ) + ] y (n ] y 3y),
(4)
Here n depends on the local velocity modulus and the
deformation rate:
n 5 max(C 1sDx 3 | u | , C 2sDx 2 s),
(5)
where Dx represents the grid step, s 5 [(] x u 2 ] y y ) 2
1 (] x y 1 ] y u) 2 ]1/2 is the deformation tensor, and C 1,2
s
are nondimensional coefficients. Discretization is associated with truncation errors that depend on the chosen scheme. For the fourth-order scheme used here, the
leading-order correction is a dispersive term (velocity
multiplied by the fifth derivative of the velocity), and
a coefficient C 1s 5 1/32 has been chosen so as to overcome its effects for all wavenumbers that the grid can
handle. Here C 2s is more subtle to choose because its
purpose is to overcome the formation of noise when
discontinuities of the velocity field (shocks) develop
(e.g., when frontal configurations are studied or when
a layer disappears). Experiments with dam-breaking
configurations showed that a value C 2s 5 1 was sometimes necessary to keep noise from developing. However, in this study, such values were not necessary, and
a coefficient C 2s 5 0.05 was adopted (notice that sensitivity experiments have been carried out with C 2s 5
0.05–0.1 to check that diffusion played a modest role
here). Last, the original MICOM schemes rely on a timesplitting technique that separates the fast barotropic
modes from the slower baroclinic modes with different
Courant–Friedrichs–Lewy (CFL) conditions for each
VOLUME 33
resolution. The time stepping of the baroclinic part uses
a second-order leapfrog scheme, and the barotropic
mode is based on an Euler scheme that does not conserve second-order accuracy for the slow motion (the
barotropic part of the slow motion is also actually calculated with an Euler time-stepping scheme). In the version of the code we have used, the time-stepping scheme
was thus also modified so as to be of second order for
the slow motions (Y. Morel 2001, personal communication).2
b. Initial configuration
For most of the study, we consider a fixed background
stratification with layer thicknesses at rest of H1 5 300
m, H 2 5 500 m, and H 3 5 4000 m and densities r1 5
1000 kg m 23 , r 2 5 1000.85 kg m 23 , and r 3 5 1001.5
kg m 23 . These thickness values and density differences
between layers were chosen with reference to Agulhas
ring observations in Arhan et al. (1999). The size of the
domain is chosen to be large enough to avoid side effects. Numerical simulations are performed in a closed
square domain of 1500 3 1500 km 2 with a grid step
Dx 5 10 km. The size of the domain has been doubled
for a few experiments and no qualitative or quantitative
differences have been found with the original configuration, which proves that boundaries do not influence
the processes studied here. The Coriolis parameter is f
5 7 3 10 25 s 21 .
The seamount is located near the center of the domain
(its exact position is X 5 600 km, Y 5 750 km) and has
a Gaussian shape (Figs. 1, 2): hT (r) 5 HT exp(2r 2 /RT2)
with a maximum height H T and a horizontal length scale
RT (r is the horizontal distance from its center).
We decide to specify our initial vortex as an isolated
PVA in the top two layers (Fig. 1). As already mentioned, PVA is a Lagrangian tracer of the shallow-water
equations. Another advantage of controlling the PVA
of the initial vortex has to do with its stability properties:
because we are interested in the splitting of a surface
vortex when it encounters a seamount, splitting resulting
from intrinsic barotropic or baroclinic instabilities has
to be avoided. We thus carefully chose the PVA profile
of the initial vortex. To avoid instability, we follow
Carton and McWilliams (1989) and chose, in each layer
k 5 1 to 3,
1
PVA k 5 DQ k0 1 2
2 1 2
r2
r2
exp 2 2 ,
2
Ry
Ry
(6)
where r is the distance from the center of the vortex.
Here DQ k0 is the maximum PVA in layer k and measures
the vortex strength, and R y is the vortex radius, which
2
Notice that the fact that part of the slow motion—the barotropic
component—is solved with a Euler time stepping can also cause some
CFL problems, in particular for the diffusion operator.
AUGUST 2003
HERBETTE ET AL.
1667
FIG. 1. Cross section of the initial absolute value of PVA in a typical experiment. A seamount
with a Gaussian shape is located eastward of a surface anticyclone. Notice the vortex negative
PVA (dotted contours) is concentrated in the surface and intermediate layer whereas the bottomlayer parcels of fluid located above the seamount have high positive PVA values (plain contours).
roughly corresponds to the radius of maximum velocity.
In the following we chose DQ10 5 DQ 02 and DQ 30 5 0.
Carton and McWilliams (1989) show that this profile
is only slightly unstable in a quasigeostrophic formalism
and that the vortex core remains coherent (the positive
PVA ring forms two cyclonic poles that do not alter the
latter). Notice that, because there is no PVA in the bottom layer and both upper layers have the same PVA
structure, baroclinic instability is avoided. This profile
has been tested on the f plane in the shallow-water
framework for different vortex radii and strengths and
has been found to be stable.
By assuming cyclogeostrophic balance, the initial velocity and layer depth fields can be calculated. Details
FIG. 2. Top view of the initial setup.
of the PVA inversion method are given in the appendix.
The absence of any kind of b effect or weak mean
advective background current makes the vortex a stationary solution of the nonviscous primitive equations.
In the absence of any topography, the cyclogeostrophic
vortex remains steady and stable and does not move.
Therefore, to trigger its erosion mechanism induced by
an isolated seamount, we must initially place it in the
vicinity of the topography. The surface eddy is initially
located close to the seamount at a distance d (Fig. 2).
c. Governing parameters
It is convenient to consider the shallow-water Eqs.
(1) in nondimensional form using the first internal radius
of deformation, R d 5 33 km, as length scale (for the
chosen stratification) and the inverse of the maximum
vortex PVA as timescale, T 5 1/DQ y0 . Neglecting the
dynamics of gravity waves (Herbette and Morel 2001,
unpublished manuscript), Eqs. (1) have five nondimensional parameters: the Rossby number e 5 1/ f T 5
DQ y0 / f characterizing the ageostrophic effects, the Reynolds number Re 5 n/DQ y0 R d4 characterizing the effect of
viscosity and that will be neglected here, and three Burger numbers Bu1 5 g/ f R d , Bu 2 5 g(r 2 2 r1 )H1 /
r 0 f R d2, and Bu 3 5 g(r 3 2 r 2 )H 2 /r 0 f R d2, characterizing
the effect of stratification (the latter are fixed in most
of the study). In addition to the five nondimensional
parameters, four additional parameters characterize the
present configuration: R y /R d (the nondimensional vortex
radius), d/R d (the nondimensional initial distance between the vortex and the seamount), R T /R d (the non-
1668
JOURNAL OF PHYSICAL OCEANOGRAPHY
dimensional seamount radius), and f H T /H 3DQ y0 (the
nondimensional seamount height; notice the latter represents the ratio between the maximum PVA associated
with the seamount and the vortex).
This number of parameters makes the seamount–vortex encounter a complicated initial-value problem with
nine independent nondimensional parameters. It is obvious that it is not possible to study the whole range of
parameter values, and our aim is to give an overview
of the possible dynamical regimes of eddy–seamount
encounters and to focus on physical interpretations of
the splitting and filamentation processes. In the following, keeping in mind the behavior of Agulhas rings, we
use the dynamical quantities in their dimensional form.
The results can, however, be generalized to other vortex
types with similar values of the above nondimensional
numbers.
3. Reference experiment
a. Parameter choice and results
The parameters that characterize the vortex are first
chosen so as to represent approximately the Agulhas
rings and the characteristics of the seamount they may
encounter: DQ10 5 DQ 20 5 0.7 f (DQ 30 5 0) and R y 5
160 km (notice that R y /R d . 5 k 1). As already mentioned, in the absence of any topography, the cyclogeostrophic vortex remains steady. Numerical viscosity
is therefore the only source of erosion and is insignificant for the resolution we have considered (R y /Dx, y
5 16). We then initially place that vortex in the vicinity
of a seamount with characteristics R T 5 60 km and H T
5 2500 m, at a distance equal to the vortex radius d 5
R y 5 160 km.
Figure 3 shows the PVA evolution in the intermediate
layer: the anticyclonic vortex no longer stays axisymmetric and coherent. It is widely elongated and breaks
into two main cores with many filaments torn apart. One
pole stays trapped over the seamount and takes an axisymmetric form very rapidly. The other core takes an
elliptic shape and experiences strong filamentation. It is
rapidly advected eastward away from the topography.
This simple idealized numerical simulation has therefore
been able to reproduce the ability, for a seamount located in the vicinity of a surface anticyclonic eddy, to
induce erosion both through filamentation and through
splitting of the surface vortex.
b. Interpretation
The evolution of the PVA in the bottom layer is
shown in Fig. 4. Notice that, in this layer, the seamount
is initially associated with a strong positive PVA of
maximum strength f H T /H 3 . 0.6 f , comparable with
the maximum absolute value of the vortex PVA in the
two upper layers. As the anticyclonic circulation, induced by the negative PVA cores in the upper layers,
VOLUME 33
advects fluid parcels over the seamount, it rapidly deforms the bottom PVA field. Part of the initial positive
PVA pole detaches from the seamount as visualized in
Fig. 4, inducing a cyclonic bottom circulation to the east
of the topography. At the same time, some water from
above the flat part of the bottom is advected over the
seamount, setting up an anticyclonic circulation on top
of it. In Fig. 4, this negative relative vorticity above the
bathymetry is detected through a diminished PVA signal. Coming back to the intermediate layer (Fig. 3), we
see that the deep anticyclone and cyclone, in turn, deform the original vortex, initially into a crescent-shaped
feature and then into a long filament-shaped pattern. The
westernmost tip of the crescent is constituted of negative
PVA advected by the deep anticyclone to the seamount
center, where it accumulates. The other tip of the crescent (the northeastern one in panel ‘‘10 days’’ of Fig.
3) evolves differently, because it is not trapped above
the bathymetry. Its more developed filamentation reflects advection by the deep cyclone. After about 40
days, an eastward advection of the initial vortex itself
(or what remains of it) by the cyclone becomes evident.
The trajectories of the main surface anticyclone and
bottom cyclone are plotted in Fig. 5. They are typical
of a baroclinic dipole (also called heton; Hogg and
Stommel 1985) self-propagating away from the topography.
The above description of Figs. 3 and 4 shows that
the vortex behavior is closely related to the generation
of the deep vortices above the submarine obstacle. The
formation of such deep eddies over an isolated obstacle,
when a large-scale current flows over it, has already
been studied for barotropic flow by Huppert (1975) and
Verron and Le Provost (1985) and for stratified fluids
by Huppert and Bryan (1976). For better evidence of
the role of the topographic eddies, we show in Fig. 6
the velocity and PVA fields of the intermediate layer.
Notice the strong cyclonic and anticyclonic circulations
developing in the initial vortex vicinity. The elliptic
shape of the main PVA core is evidently associated with
this circulation, as well as the generation of filaments
that wrap around the newly formed eddies. As long as
filamentation only is involved, both deep circulations
play a symmetrical role. However, it is the locking of
the deep anticyclone above the seamount that causes the
low PVA filaments entrained by this eddy to accumulate
at this location, resulting, after about 40 days, in a division of the original vortex into two parts.
Filamentation of a vortex when subject to a background shear has been studied for two-dimensional
flows by Legras and Dritschel (1993a,b), Polvani and
Flierl (1989), Mariotti and Legras (1994), and Legras
et al. (2001). They show that the background shear is
responsible for the elliptic shape and can lead to filamentation when it is larger than the vortex vorticity.
Arai and Yamagata (1994) have shown that the presence
of a stagnation point (hyperbolic point at which the
velocity is null) inside the core of the eddy is a necessary
AUGUST 2003
HERBETTE ET AL.
1669
0
0
FIG. 3. Evolution of the PVA in the intermediate layer for the reference experiment: DQk51,2
5 20.7 f , DQk53
5 0, R y 5 160 km, R T 5
60 km, H T 5 2500 m, and d 5 160 km. The contour interval is 0.2 3 10 25 s 21 .
1670
JOURNAL OF PHYSICAL OCEANOGRAPHY
VOLUME 33
FIG. 4. Evolution of the PVA in the bottom layer for the same experiment as in Fig. 3. The contour interval is now 1.0 3 10 25 s 21 .
AUGUST 2003
1671
HERBETTE ET AL.
erosion and splitting are associated with the ability of
the background shear (itself linked to the anticyclonic
and cyclonic topographic eddies) to peel PV filaments
from the vortex core. These filaments then give rise to
new PV poles or vortices such as the one that stays
trapped above the seamount in this reference experiment. We thus expect the erosion and splitting processes
to be strongly sensitive to the strength of the topographic
eddies with respect to the original vortex strength. We
now study the effect of several parameters and their
relative influence.
4. Sensitivity analysis
a. Measurement of the vortex erosion
To measure the vortex erosion, we found it convenient
to define the quantity:
FIG. 5. Trajectory of the bottom topographic cyclone (dashed line)
and main surface anticyclonic pole (plain line). The star locates the
seamount position.
condition to its subdivision. The Okubo–Weiss quantity
(Okubo 1970; Weiss 1991) measures the local influence
of the vorticity and shear/strain:
lok 5 s 2 2 z 2 .
(7)
It allows one to distinguish regions that are dominated
by vorticity in comparison with regions that are dominated by shear/strain effects. In the latter case, lok is
positive, the PV field is generally strongly distorted, and
filamentation occurs.3
In our study, the background shear is associated with
the development of the topographic eddies and is particularly strong in the early stage of the encounter, when
both the anticyclonic and cyclonic bottom eddies interact with the initial vortex. Figure 6 shows that stagnation
points do exist in our simulation between days 10 and
30 (in particular in the region between the seamounttrapped anticyclone and the main core) and are associated with the development of the topographic eddies.
Figure 7 represents the Okubo–Weiss quantity at time
t 5 20 days, with the position of one particular stagnation point marked with a star symbol. The latter and
Okubo–Weiss criterion both reveal the same thing: the
vortex PV field is likely to be strongly deformed, and
filaments are likely to be expelled in the region between
the vortex core and the seamount.
All of these results indicate that the observed vortex
3
Notice that this criterion assumes that the tensor (=u ) is stationary
and that the horizontal divergence is small when compared with the
square root of lok . Lapeyre et al. (1999) have recently derived, in
the framework of two-dimensional flows, a more accurate criterion
that takes into account some Lagrangian variations of the tensor (=u ).
Its accurate calculation, however, requires the gravity waves to be
filtered out.
S k (t) 5
EE
h k (x, y)PVA k (x, y) dx dy,
(8)
V
which is the bulk volume integral of the PVA over
some region V. In this study, the latter is chosen as
the region in which PVA k is smaller than a threshold
value (PVAthreshold 5 0 here) but is also restricted to a
threshold radius around the main PVA pole center (here
Rthreshold 5 1.25R y ). Different threshold PVA and radius
values have been tested, and we found that our results
are little sensitive to the chosen values, provided they
are not too small or too large. Because PVA is a tracer,
it is usually tightly related to the temperature or salinity
anomaly inside the core of vortices. Here S (t) thus
quantifies the tracer content of the vortex core. Because
it is related to the relative vorticity and velocity field
of the vortex, S (t) also measures the evolution of the
vortex strength.
Our numerical results show that splitting and erosion
act rapidly so that the final stage is reached after 50
days or so. To estimate the vortex erosion, we thus follow the most intense PVA pole, calculate S at time t 5
80 days, and compare it to the initial value S(t 5 0).
The parameter
Rfc 5
S (t 5 80 days)
3 100
S (t 5 0 days)
(9)
thus measures the final PVA content of the vortex, the
missing part being located in the filaments and the newly
formed PVA pole(s).
b. Distance between the vortex and the topography
Assuming a vortex and a seamount with the same
0
characteristics as in the reference experiment (DQ k51,2
5 20.7 f , R y 5 160 km, H T 5 2500 m, and R T 5 60
km), we vary the initial distance d between the vortex
and the topography from 0 to approximately 2 times the
vortex radius. Figure 8 represents the dependence of Rfc
1672
JOURNAL OF PHYSICAL OCEANOGRAPHY
VOLUME 33
FIG. 6. Same as Fig. 3 except the velocity field is superimposed and we have zoomed in on the area near the vortex.
AUGUST 2003
HERBETTE ET AL.
1673
FIG. 7. Okubo–Weiss quantity at time t 5 20 days in the intermediate layer for the reference
experiment. Plus marks are stagnation points in regions where lok is negative; the star is a
stagnation point in a region where lok is positive and therefore represents a filamentation region.
on the initial vortex–seamount distance for both the upper and second layer. Erosion is very efficient when the
distance between the vortex and the seamount is close
to the vortex radius (d . R y ), otherwise it remains weak.
In fact, we have seen that the strength of the topographic
eddies is related to the strength of the vortex-induced
FIG. 8. Anticyclone final PVA content ratio Rfc as a function of d
for the reference experiment. The surface layer is associated with the
dash–dotted line with circles, and the intermediate layer is associated
with the plain line with stars.
cross-isobath flow above the seamount. When d is small,
the circulation associated with the vortex approximately
follows the isobaths and no or few fluid parcels are
detached from above the seamount. As a result, the topographic eddies are weak and do not have much effect
on the upper-layer anticyclone. We have verified that no
cyclone is formed and that the surface vortex stays
trapped above the seamount. When d increases, the
cross-isobath flow also increases and is maximum when
the vortex edge is just above the seamount, that is to
say, when d . R y . Strong cyclonic and anticyclonic
topographic eddies are then generated, which erode the
initial vortex as seen in the reference experiment. When
d is further increased, the cross-isobath flow associated
with the original vortex rapidly diminishes again as the
velocity field decreases exponentially beyond the vortex
radius. A cyclone detaches from the seamount, but it is
too weak to deform the original vortex and generate
filamentation. However, it still interacts with the latter,
forming a baroclinic dipole that is advected away from
the topography. The topographic anticyclone trapped
above the seamount is too far from the vortex to interact
with it. As seen in the reference experiment, the vortex
splits into two main negative PVA poles when erosion
is maximum. One stays trapped above the seamount
while the other, forming a baroclinic dipole with the
bottom cyclone, is advected away. For values of d slightly below R y , the main PVA pole (with the largest Rfc )
1674
JOURNAL OF PHYSICAL OCEANOGRAPHY
VOLUME 33
FIG. 9. Sensitivity of the intermediate-layer anticyclone final PVA
content ratio Rfc 2 to the vortex radius. The solid line with squares
corresponds to the reference experiment and is the same as in Fig.
8. The dotted line with diamonds corresponds to R y 5 50 km, the
dashed line with triangles corresponds to R y 5 100 km, and the
dashed–dotted line with circles corresponds to R y 5 200 km.
FIG. 10. Sensitivity of the intermediate-layer anticyclone final PVA
content ratio Rfc 2 to the seamount height. The characteristics of the
experiments are the same as in the reference configuration except for
the seamount height (circles: H T 5 3300 m, squares: H T 5 2500 m,
triangles: H T 5 1500 m, diamonds: H T 5 1000 m, and stars: H T 5
500 m).
is the one above the seamount, and for values of d
slightly above R y , it is the one in the baroclinic dipole.
d. Seamount height
c. Vortex radius
We now study how varying the surface vortex radius
may affect its erosion as it encounters a seamount. Considering intense vortices (DQ10 5 DQ 02 5 20.7 f ) located
in the vicinity of our basic seamount (R T 5 60 km and
H T 5 2500 m), we show in Fig. 9 the final PVA content
ratio Rfc of the intermediate layer for eddy radii R y of
50, 100, 160, and 200 km. For each vortex radius, d is
varied from 0 to approximately 2 times the vortex radius. Notice the maximum erosion always happens when
d . Ry .
Figure 9 shows that small surface vortices only weakly dissipate and appear not to feel the seamount, at least
when the latter has the characteristics we have chosen.
Two effects contribute to decrease the strength of the
cross-isobath flow associated with small vortices. First
the maximum velocity strongly depends on the ‘‘reservoir’’ of negative PVA associated with the initial eddy
and thus diminishes with R y . Second, the influence of
a PVA pole on adjacent layers also depends on the vortex radius (it is measured by the Burger numbers) and
decreases with R y also. Except for numerical viscosity
effects that cut down the PVA content by about 15%,
there is almost no erosion for R y 5 50 km, and therefore
no filamentation or splitting. For larger radii, the vortex
erosion is effective and may lead to a final main PVA
pole that contains only about 30% of its initial PVA.
The rest has been expelled into filaments and new PVA
poles.
In the following experiments, we keep the reference
0
values for the vortex structure (DQ k51,2
5 20.7 f ,
0
DQ k53 5 0, and R y 5 160 km) and vary the seamount
height. For each value of H T , we test the influence of
the eddy–seamount initial distance d from 0 to approximately 2 times the eddy radius. The results (Fig. 10)
show Rfc in the intermediate layer, after 80 days, for H T
values of 500, 1000, 1500, 2500, and 3300 m. Whatever
is the seamount height, maximum erosion coincides with
an eddy–seamount distance roughly equal to the vortex
radius (d . R y ), that is to say, when the induced crossisobath flow in the bottom layer is also maximum. Nevertheless, when d . R y , erosion is low for H T 5 500
m and is significantly larger for the other tested heights.
This result is expected because the strength of the
emerging cyclone is linked to the reservoir of positive
PVA associated with the seamount, which obviously
increases with the seamount height. Yet, it is surprising
to see that erosion is maximum for H T . 1500 m and
decreases for higher seamounts. Analysis of the PVA
evolution (not shown) reveals that in all experiments
with d . R y a cyclone detaches from the topography.
It leads to only slight filamentation for H T 5 500 m and
to stronger filamentation and splitting for larger values,
particularly for H T . 1500 m. In fact, increasing the
seamount height also steepens the bottom slope. A
strong topographic slope is also associated with a strong
topographic b effect, which makes it more difficult for
fluid parcels to move across isobaths. This effect counteracts the PVA reservoir increase because it can keep
the cyclonic eddy from detaching and limits the
strengths of the topographic eddies. Figure 11 shows
AUGUST 2003
HERBETTE ET AL.
1675
up in breaking of the original surface eddy. Nevertheless, as expected, the final erosion is different for each
configuration: it is weak for the smaller seamount (Rfc
. 80%) and increases with larger R T . Indeed, in this
case, there are no counteracting effects: increasing R T
increases the topographic PVA reservoir and diminishes
the topographic b effect. It is interesting to notice that,
whereas in the reference experiment the original vortex
splits into two negative PVA poles, five are created for
R T 5 90 km including the one trapped on the seamount
(Fig. 12), and four poles are created for R T 5 120 km.
The merging of some of the poles makes it difficult to
define a rule for the number of poles created, which
eventually depends on the ability of the cyclone to
stretch the original vortex, but the possibility to create
more than two eddies is illustrated.
FIG. 11. Strength of the bottom cyclone as a function of the seamount height (here d 5 160 km is fixed). Note that the maximum is
obtained for an intermediate seamount height.
the nondimensional value of the content of the positive
PVA pole that leaves the topography in the bottom layer
when d 5 160 km. It is calculated with Eq. 8 and is
then scaled by the initial anticyclonic vortex PVA content. We note that the number of positive PVA particles
leaving the topography is indeed maximum for H T .
1500 m and slightly decreases for larger values. Notice
finally that, in the region between the bottom anticyclone and cyclone, both eddies accumulate their effects
on the surface intensified eddy. Besides, because the
topographic anticyclone roughly has the same strength
as the cyclone,4 the cumulative effect of the topographic
eddies can be evaluated by multiplying the cyclone PVA
content by 2. Figure 11 then also shows that, to achieve
strong filamentation and splitting, the cumulated PVA
content associated with the topographic eddies must be
higher than that of the original vortex. This condition
is necessary for the circulation associated with the former vortices to be strong enough to create a stagnation
point.
e. Seamount radius
The next parameter we have chosen to study is the
seamount radius R T . Like H T , variations of this parameter modify both the seamount PVA reservoir and the
seamount steepness. We thus resume the reference experiment with several values of R T : 30, 60, 90, and 120
km. The simulation with R T 5 30 km is conducted with
a thinner grid (Dx, y 5 5 km). All four experiments end
4
The cyclone and anticyclone are formed because fluid parcels are
exchanged between two regions with different PVA: the topography,
with high PVA values, and the flat bottom region that corresponds
to PVA 5 0. Except for their sign, exchanged particles thus correspond to the same anomaly when calculated with respect to their new
environment.
f. Stratification
In going back to Fig. 8, we notice that the erosion
and splitting processes lead to a vortex with a modified
vertical structure, because the negative PVA pole is
more eroded in the intermediate layer than in the upper
layer: for d 5 R y , the final PVA content of the surface
and second layer vortex, respectively, represents about
40% and 50% of the initial value. In fact, the influence
of the topographic eddies decreases upward, because
the latter are associated with PVA in the bottom layer
only. Therefore, the shear induced by these eddies is
larger and filamentation is more effective in the intermediate layer than in the surface layer. This effect is
expected to be sensitive to the stratification. We thus
resume the reference experiment, keeping the same densities for the upper and intermediate layer and modifying
r 3 . Two different values are studied, one with a smaller
density r 3 5 1001.1 kg m 23 and one with a higher
density r 3 5 1001.6 kg m 23 (r 3 5 1001.5 kg m 23 in
the reference experiment). The corresponding deformation radii are 26 and 36 km, respectively (and 33 km
for the reference experiment). Figure 13 shows the final
PVA content Rfc for each experiment and for each layer
in comparison with the reference simulation. The plainline curves correspond to the reference experiment, the
dashed-line curves correspond to r 3 5 1001.6 kg m 23 ,
and the dash–dotted-line curves correspond to r 3 5
1001.1 kg m 23 . Circles are associated with the surface
layer, and stars are associated with the intermediate layer.
We first note that the shapes of the erosion curves
remain similar to those of the reference experiment: the
erosion is maximum for d . R y and, in this case, splitting occurs. However, as expected, erosion is stronger
when D r 2–3 is weaker. When the stratification between
the bottom and intermediate layers D r 2–3 is increased,
the circulation, induced in the upper layers by the topographic eddies, decreases. The result of the eddy–
seamount encounter therefore becomes less erosive. On
the contrary, when D r 2–3 is decreased, the upper-layer
1676
JOURNAL OF PHYSICAL OCEANOGRAPHY
VOLUME 33
FIG. 12. Evolution of the PVA in the intermediate layer for an experiment with the same characteristics as in the reference experiment
except for the seamount radius: R T 5 90 km. Note the many small vortices generated here.
AUGUST 2003
HERBETTE ET AL.
FIG. 13. The surface-layer (circles) and intermediate-layer (stars)
anticyclone final PVA content ratios Rfc (k 5 1, 2) are plotted as a
function of the eddy–seamount distance for several ambient stratifications. Weak stratification (r 3 5 1001.1 kg m 23 ): dashed–dotted
lines; reference experiment (r 3 5 1001.5 kg m 23 ): plain lines; strong
stratification (r 3 5 1001.6 kg m 23 ): dashed lines.
signature of the bottom eddies becomes stronger. Thus,
more filamentation is expected. In the latter case, the
difference between vortex erosion in surface and intermediate layers increases. For a critical eddy–seamount
distance d 5 R y and a weak stratification (r 3 5 1001.1
kg m 23 ), the top surface eddy keeps 49% of its original
PVA content after 80 days, whereas the intermediate
layer only keeps 22%. Tracers have thus been much
more dispersed in the intermediate layer than in the
upper layer, and the resulting eddy is more surface intensified than originally.
5. Summary
In this study, we have analyzed the mechanisms intervening at the encounter of a strong anticyclonic surface-intensified vortex with an isolated seamount. The
result of this encounter can be very erosive and can lead
to the splitting of the original vortex with expelling of
PV filaments from the main core. The mechanism is
different from the one already proposed by Simmons
and Nof (2000) and Cenedese (2002), who addressed
the collision of an eddy with an island. In our simulations, erosion stems from the advection of deep water
across the isobaths of the topography. The deep crossisobath flow generates bottom-intensified vortices of either sign, which in turn deform, stir, and often subdivide
the initial eddy.
Because the influence of the bottom eddies decreases
upward as a function of the ambient stratification, the
process is found to be sensitive to the stratification, and
the eroded eddy is always more surface intensified than
the original one. The sensitivity of the interaction to the
1677
vortex and seamount characteristics (vortex radius, seamount radius, and seamount height) is found to be governed by the reservoir of positive PVA associated with
the seamount in the bottom layer, by the topography
steepness, and by the initial cross-isobath velocity associated with the vortex. Seamounts with small radii or
heights have a reduced PVA reservoir, which can limit
the strength of the topographic eddies, and the latter
will not be able to tear apart strong incoming vortices.
In the case of steep seamount slopes, on the other hand,
the topographic b effect can keep fluid parcels from
being expelled from above the bathymetry. This is another limiting parameter for the generation of topographic eddies sufficiently strong to erode the initial
vortex. Another key parameter is naturally the distance
between the vortex and the seamount: erosion is maximum when the distance is close to the vortex radius,
that is to say, when the cross-isobath flow induced by
the vortex is maximum. When the vortex is initially far
from the seamount, the deep cyclone generated by the
interaction, though weak, may form a baroclinic dipole
with the original vortex and advect it away from the
topography before significant erosion can occur.
One may then wonder if a vortex propagating toward
a seamount can get close enough to the latter so that
topographic eddies can first be generated and then induce a critical shear on the approaching eddy. The bplane experiments carried out by Van Geffen and Davies
(2000) have apparently led not to strong erosion but
only to deviations of the vortex trajectories. It is clear
that more ingredients than considered here would be
required for a more realistic approach. For instance, the
planetary b effect, or the presence of a mean flow, might
help the eddy to move sufficiently close to the seamount
to trigger the processes described above. In a similar
way, although subdivisions of the initial vortex in more
than two structures have been observed, one may wonder which additional process could move away from the
seamount the pole that always remains trapped on it.
The planetary b effect might also play this part. Moreover, the dynamics of oceanic vortices, in particular of
large structures such as Agulhas rings, are generally
influenced by the planetary b effect. The presence of b
can lead to their substantial erosion and therefore can
modify the result of their encounter with topography.
Nevertheless, the f -plane experiments that we have carried out constitute a first step toward the understanding
of the physical mechanisms that govern an eddy–seamount encounter and will provide a useful insight of
the dynamics for more realistic studies.
For the Agulhas rings that we presented as oceanic
examples of eddy subdivisions, we may mention the
recent increased interest in the cyclonic structures that
often are observed to accompany these eddies, particularly in their possible formation processes (Penven et
al. 2001). The formation of baroclinic dipoles that we
have seen to result from the eddy and seamount en-
1678
JOURNAL OF PHYSICAL OCEANOGRAPHY
counter should also probably be considered in this context.
Acknowledgments. The authors are grateful to Drs.
Rainer Bleck, Linda Smith, and Eric Chassignet for the
use of MICOM and to Dr. Xavier Carton for stimulating
discussions and comments. Support for this study has
been provided by the Service Hydrographique et Océanographique de la Marine for Yves Morel and by the
Institut Français de Recherche pour l’Exploitation de la
Mer for Michel Arhan. Steven Herbette was supported
by a grant from the Délégation Générale de l’Armement
during his Ph.D thesis.
APPENDIX
Initialization of a Stable Vortex
We consider an N-layer flat-bottom ocean and an axisymmetric vortex in cyclogeostrophic balance. This vortex is associated with a potential vorticity anomaly profile DQ(r). To initialize our shallow-water model, we
need to invert the potential vorticity equation and calculate the layer thicknesses and azimuthal velocities associated with the PVA field. In polar coordinates, the
system to solve is
y u2k
1 f y uk 5 ]r M k ,
r
O Dh ,
r
5 O
gDh 1 O gDh ,
r
M1 5 g
i5N c
i
i51
i5k21
Mk
i5N c
i
i
i51
zk 2 f
i
1
and
i5k
k
2
Dh k
Dh k
5 DQ k 1 1
,
Hk
Hk
(A1)
where z k 5 1/r] r (ry u k ) is the relative vorticity in polar
coordinates, k is the layer subscript, y u k is the azimuthal
velocity, M k is the Montgomery potential, and Dh k 5
h k 2 H k is the layer thickness variation. This system of
partial derivative equations is not linear because of the
quadratic term associated to the centrifugal force y u2/r
and does not admit trivial analytical solutions. We therefore use a numerical iterative method that consists in
approaching the quadratic term at each iteration (n) by
its value at iteration (n 2 1). The first step of the method
(n 5 1) supposes geostrophic equilibrium, and the quadratic term is simply absent from the equations. To guarantee that our vortex is isolated, we have to ensure that
the bulk volume integral of PVA is zero (Morel and
McWilliams 1997). To overcome this constraint, we
have chosen to fix F k 5 DQ k (1 1 Dh k /H k ) instead of
DQ k .
The system we invert is the following:
(M¹ 2 Dh n ) k 2 f 2
VOLUME 33
Dh kn
1
5 Fk 1 ] r (y u2k ) n21 ,
Hk
r
y u2k 5
and
(y u2k ) n21
1
(M] r Dh) k 2
,
f
r
(A2)
where M is the matrix that links the Montgomery potential M to Dh.
The boundary conditions are
1O Dh 2 5 0,
k5N
lim (] r Dh k ) 5 0,
r→1`
lim
r→1`
k
and
k51
y u (r 5 0) 5 0.
(A3)
Equations (A2) with the boundary conditions Eqs. (A3)
form a linear system of partial derivative equations. The
latter is solved by calculating the eigenvalues and eigenvectors of the matrix M. Iterations are stopped when
the difference in Dh between two successive iterations
is less than a small value e (typically e . 1 mm):
max( | Dh n 2 Dh n21 | ) # e.
(A4)
REFERENCES
Arai, M., and T. Yamagata, 1994: Asymmetric evolution of eddies
in rotating shallow water. Chaos, 4, 163–175.
Arhan, M., H. Mercier, and J. R. E. Lutjeharms, 1999: The disparate
evolution of three Agulhas rings in the South Atlantic Ocean.
J. Geophys. Res., 104, 20 987–21 005.
Bleck, R., and D. Boudra, 1986: Wind-driven spin-up in eddy-resolving ocean models formulated in isopycnic and isobaric coordinates. J. Geophys. Res., 91, 7611–7621.
——, and L. T. Smith, 1990: A wind-driven isopycnic coordinate
model of the north and equatorial Atlantic Ocean. 1. Model
development and supporting experiments. J. Geophys. Res., 95,
3273–3285.
——, C. Rooth, D. Hu, and L. T. Smith, 1992: Salinity-driven thermocline transients in a wind- and thermohaline-forced isopycnic
coordinate model of the North Atlantic. J. Phys. Oceanogr., 22,
1486–1505.
Carnevale, G., R. Kloossterziel, and G. Van Heijst, 1991: Propagation
of barotropic vortices over topography in a rotating tank. J. Fluid
Mech., 233, 119–139.
Carton, X., and J. McWilliams, 1989: Barotropic and baroclinic instabilities of axisymmetric vortices in a qg model. Mesoscale/
Synoptic Coherent Structures in Geophysical Turbulence, Elsevier, 225–244.
Cenedese, C., 2002: Laboratory experiments on mesoscale vortices
colliding with a sea-mount. J. Geophys. Res., 107, 3053, doi:
10.1029/2000JC000599.
Dewar, W. K., 2002: Baroclinic eddy interaction with isolated topography. J. Phys. Oceanogr., 32, 2789–2805.
Ertel, H., 1942: Ein neuer hydrodynamischer Wirbelsatz (A new scalar to characterize hydrodynamic eddies). Meteor. Z., 59, 271–
281.
Gordon, A. L., and W. F. Haxby, 1990: Agulhas eddies invade the
South Atlantic: Evidence from Geosat altimeter and shipboard
conductivity–temperature–depth survey. J. Geophys. Res., 95,
3117–3125.
Hogg, N., and H. Stommel, 1985: The heton, an elementary interaction between discrete baroclinic geostrophic vortices, and its
implication concerning eddy heat-flow. Proc. Soc. London, 370,
1–20.
AUGUST 2003
HERBETTE ET AL.
Huppert, H., 1975: Some remarks on the initiation of inertial Taylor
columns. J. Fluid Mech., 67, 397–412.
——, and K. Bryan, 1976: Topographically generated eddies. DeepSea Res., 23, 655–679.
Kamenkovich, V. M., Y. P. Leonov, D. A. Nechaev, D. A. Byrne, and
A. L. Gordon, 1996: On the influence of bottom topography on
the Agulhas eddy. J. Phys. Oceanogr., 26, 138–140.
LaCasce, J. H., 1998: A geostrophic vortex over a slope. J. Phys.
Oceanogr., 28, 2362–2381.
Lapeyre, G., P. Klein, and B. Hua, 1999: Does the tracer gradient
vector align with the strain eigen-vectors in 2d turbulence? Physica D, A11, 3729–3737.
Legras, B., and D. Dritschel, 1993a: A comparison of the contour
surgery and pseudo-spectral methods. J. Comput. Phys., 104,
287–302.
——, and ——, 1993b: Vortex stripping and the generation of high
vorticity gradients in two-dimensional flows. Appl. Sci. Res., 51,
445–455.
——, ——, and P. Caillol, 2001: The erosion of a distributed twodimensional vortex in a background straining flow. J. Fluid
Mech., 441, 369–398.
Mariotti, A., and B. Legras, 1994: Vortex stripping and the erosion
of coherent structures in two-dimensional flows. Phys. Fluids,
6, 3954–3962.
McWilliams, J. C., and G. R. Flierl, 1979: On the evolution of isolated, nonlinear vortices. J. Phys. Oceanogr., 9, 1155–1181.
Morel Y., and J. McWilliams, 1997: Evolution of isolated interior
vortices in the ocean. J. Phys. Oceanogr., 27, 727–748.
Mory, M., 1985: Integral constraints on bottom and surface isolated
eddies. J. Phys. Oceanogr., 15, 1433–1438.
Nof, D., 1983: The translation of isolated cold eddies on a sloping
bottom. Deep-Sea Res., 30, 171–182.
Okubo, A., 1970: Horizontal dispersion of floatable particles in the
vicinity of velocity singularities such as convergences. DeepSea Res., 17, 445–454.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. Springer-Verlag,
710 pp.
View publication stats
1679
Penven, P., J. Lutjeharms, P. Marchesiello, C. Roy, and S. Weeks,
2001: Generation of cyclonic eddies by the Agulhas current in
the lee of the Agulhas bank. Geophys. Res. Lett., 27, 1055–1058.
Polvani, L., and G. Flierl, 1989: Filamentation of unstable vortex
structures via separatrix crossing: A quantitative estimate of onset time. Phys. Fluids, A1, 181–184.
Schouten, M. W., W. P. M. De Ruijter, P. J. Van Leeuwen, and J. R.
E. Lutjeharms, 2000: Translation, decay and splitting of Agulhas
rings in the southeastern Atlantic Ocean. J. Geophys. Res., 105,
21 913–21 925.
Simmons, H., and D. Nof, 2000: Islands as eddy splitters. J. Mar.
Res., 58, 919–956.
Smith, D. C., IV, and J. J. O’Brien, 1983: The interaction of a twolayer isolated mesoscale eddy with bottom topography. J. Phys.
Oceanogr., 13, 1681–1697.
Smith, L. T., 1992: Numerical simulations of stratified rotating flow
over finite amplitude topography. J. Phys. Oceanogr., 22, 686–
696.
Sutyrin, G. G., and G. R. Flierl, 1994: Intense vortex motion on the
beta plane: Development of the beta gyres. J. Atmos. Sci., 51,
773–790.
——, and Y. Morel, 1997: Intense vortex motion in a stratified fluid
on the beta-plane: An analytical theory and its validation. J.
Fluid Mech., 336, 203–220.
Thierry, V., and Y. Morel, 1999: Influence of a strong bottom slope
on the evolution of a surface-intensified vortex. J. Phys. Oceanogr., 28, 911–924.
Van Geffen, J. H. G. M., and P. A. Davies, 2000: A monopolar vortex
encounters an isolated topographic feature on a b-plane. Dyn.
Atmos. Oceans, 32, 1–26.
Verron, J., and C. Le Provost, 1985: A numerical study of quasigeostrophic flow over topography. J. Fluid Mech., 154, 231–
252.
Wang, G., and W. Dewar, 2003: Meddy–seamount interactions: Implication for the Mediterranean salt tongue. J. Phys. Oceanogr.,
in press.
Weiss, J., 1991: The dynamics of enstrophy transfer in two-dimensional hydrodynamics. Physica D, 48, 273–294.