Sie sind auf Seite 1von 11

Bulletin of the Australian Meteorological and Oceanographic Society Vol.

19 page 95
Articles
Surface and sub-surface circulation and water masses off Western Australia
Charitha Pattiaratchi
School of Environmental Systems Engineering, University of Western Australia
Address for Correspondence: Charitha Pattiaratchi, School of Environmental Systems Engineering, University of
Western Australia, 35 Stirling Highway, Crawley, 6009, WA. Australia. Email: chari.pattiaratchi@uwa.edu.au
1. Introduction
During the past two decades the physical
oceanographic processes off the Western Australian
coast have received much attention. This is mainly
due to the observation that, globally, eastern ocean
basins are highly productive ecosystems supporting
high primary productivity and large pelagic finfish
stocks. However, the exception to this rule is off the
West Australian coast where, although the wind
regime is similar to other eastern ocean margins, the
waters are oligotrophic. Thus, the initial studies
concentrated on addressing why the circulation off
Western Australia was different from other eastern
margins, which led to the discovery of the Leeuwin
Current and its dynamics. Later studies addressed
the dynamics of the Leeuwin Current System, which
is a system comprised of three currents: the Leeuwin
Current, Leeuwin Undercurrent, and shelf current
systems consisting of the Ningaloo, Capes and
Cresswell currents (Woo et al., 2006a).
The main contributions to the understanding of the
physical oceanography of this region have come
from Andrews (1977, 1983), Cresswell and Golding
(1980), Hamilton (1986), Smith et al. (1991), Pearce
and Walker (1991), Cresswell and Peterson (1993),
Gersbach et al. (1999), Pearce and Pattiaratchi
(1999), Feng et al. (2003, 2005), Morrow et al.
(2003), Ridgway and Condie (2004), Woo et al.
(2006a), Rennie et al. (2006) and Meuleners et al.
(2006).
The main oceanographic features of the region
between Shark Bay and Esperance primarily consist
of the following surface and subsurface current
systems (Figures 1 and 2):
The West Australian Current
The Leeuwin Current
The Leeuwin Undercurrent
The Flinders Current
Continental shelf current systems:
the Ningaloo, Capes and
Cresswell currents
Here, a short description of each of these current
systems is presented together with a discussion on
the various water masses found within the area.
2. The West Australian Current
Eastern boundary currents (EBC) occur along most
eastern ocean boundaries as slow and broad
equatorward currents (in contrast to the intensified
poleward western boundary currents) forming one
part of the anticyclonic subtropical gyre in each
hemispheres oceans. These gyres result from the
Sverdrup balance: a balance between wind stress,
pressure gradients and Coriolis acceleration. The
EBC regions are characterised by cooler (through
upwelling) water and high primary productivity, thus
supporting major pelagic finfish industries. Western
Australia does not have the level of biological
productivity induced by the Humboldt Current off
Africa or the Benguela Current off Peru because the
Leeuwin Current suppresses the upwelling of cooler,
nutrient-rich water along the continental shelf.
Schott and McCreary (2001) postulated that the
northward current extends westward from the
Western Australian coast up to longitude 60
o
E.
3. The Leeuwin Current
It has been known for several decades that the
circulation off the West Australian (WA) coast is
different from any other western continental margins
(Schott, 1935; Smith et al., 1991). In each of the
main ocean basins, the surface circulation forms a
gyre with poleward flow along the westward
boundary of the basin and equatorward flow along
the eastern margin. In addition, the eastern margins
(off southern America and southern Africa, for
example) are areas of high productivity due to
upwelling. The exception to this rule is off the WA
coast, where the Leeuwin Current transports water
poleward (Figures 1 and 2).
The Leeuwin Current (LC) is a shallow (< 300 m),
narrow band (< 100 km wide) of warm, lower
salinity, nutrient-depleted water of tropical origin,
which flows poleward from Exmouth to Cape
Leeuwin and into the Great Australian Bight (Church
et al., 1989; Smith et al., 1991; Ridgway and Condie,
2004). It is now accepted that the Leeuwin Current
signature extends from North West Cape to
Tasmania as the longest boundary current in the
world (Ridgway and Condie, 2004). Here, we
follow the same definition as Cresswell and Peterson
COVER PAGE

Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 96
(1993) to define the LC as a warm water current of
tropical origin, which, during the summer months, is
augmented by the addition of (salty) water from the
West Australian Current (see Section 2).
Warmer, lower salinity water flows through the
Indonesian Archipelago from the Pacific to the
Indian Ocean and results in lower density water
being present between Australia and Indonesia
compared with the cooler and more saline ocean
waters off south-western Australia. This density
difference results in a change in sea level of about
0.5 m along the WA coast and is the driving force of
the LC. Due to the effect of the earths rotation,
water is entrained from the Indian Ocean into the LC
as it flows southward; thus, the LC becomes stronger
as it flows southward.
Studies undertaken over the past decade have shown
that, along the west coast, the Leeuwin Current is
driven by an alongshore pressure gradient, which
overwhelms the opposing equatorward wind stress
(Thompson, 1984, 1987; Godfrey and Ridgway,
1985; Weaver and Middleton, 1989; Batteen and
Rutherford, 1990; Pattiaratchi and Buchan, 1991).
These investigators have demonstrated that the
pressure gradient (in the upper 250300 m of the
ocean) overcomes the upwelling favourable winds,
inducing an onshore surface flow, which results in
downwelling at the coast. Onshore geostrophic flow
from the central Indian Ocean occurs towards
Western Australia between the approximate latitudes
15
o
S and 35
o
S. Geostrophic inflow in the north
(1528
o
S), augmented by tropical water from the
North-west Shelf, forms the warm, low salinity core
of the Leeuwin Current (Smith et al., 1991; Woo et
al., 2006a).
It has been postulated that south of about 30
o
S the
Leeuwin Current intensifies (increase in the transport
and the velocity) because of the geostrophic inflow
of subtropical water from the southwest especially
during the summer months (Hamilton, 1986;
Cresswell and Peterson, 1993; Cresswell, 1996).
The current continues beyond Cape Leeuwin
eastward into the Great Australian Bight (Ridgway
and Condie, 2004). Here, the dynamics are thought
to be similar to that along the west coast in that the
current is still driven by the alongshore pressure
gradient, the magnitude of which is slightly lower
than that along the west coast (Godfrey et al., 1986).
However, Ridgway and Condie (2004) postulated
that the wind stress played a more dominant role,
compared with the alongshore pressure gradient, in
driving the current along the south coast during the
winter.
The Leeuwin Current exhibits both seasonal and
inter-annual variability: it is generally stronger
during winter and under La Nia conditions and is
weaker during the summer and El Nio conditions.
The transport within the LC is strongest near the
shelf edge in late summer and early autumn and
farther seaward in winter with maxima around 5 Sv
(x10
6
m
3
s
-1
) and weaker during the summer (~2 Sv).
The seasonal variability reflects changes in the local
wind stress: during the summer months the LC is
weaker as it accelerates into the maximum
equatorward wind stress whereas during the winter
months the LC is stronger as both the stress applied
by the equatorward wind stress is weaker and the
pressure gradient is also higher (Godfrey and
Ridgway, 1985). Similarly, along the south coast,
the current is weaker during the summer as the
prevailing winds are from between east and
southeast.
The LC also responds to the El Nio Southern
Oscillation (ENSO) signals which propagate from
the western Pacific ocean to the north of the western
Australian coast and the higher coastal sea levels are
transmitted during La Nia years resulting in a
stronger Leeuwin current and lower sea levels during
El Nio years resulting weaker Leeuwin current
(Pattiaratchi and Buchan, 1991; Feng et al., 2003).
Leeuwin Current eddies
The Leeuwin Current is generally associated with
meso-scale eddies and meanders (Pearce and
Griffiths 1991; Cresswell, 1996; Fang and Morrow
2003; Morrow et al. 2003; Feng et al. 2005; Fieux et
al. 2005; Meuleners et al., 2006; Rennie et al., 2006).
Eddies form at the shelf break and eventually
separate from the current and drift westward. These
eddies are apparent in sea surface temperature
satellite imagery (Griffin et al., 2001) and in
altimeter data (Fang and Morrow, 2003).
Interaction of the LC with changes in the bathymetry
and offshore water of different densities results in
the generation and subsequent offshore transport of
eddiesin particular off Shark Bay, Abrolhos
Islands, Jurien Bay, Rottnest Island and Cape
Leeuwin (cover picture). The main regions of eddy
generation may be summarized as follows:
(1) West of Rankin Bank (2021.5
o
S, 114.5
115.5
o
E). Here, the slope of the continental
slope changes abruptly because of the
presence of the Rankin Bank. One-third of
all long-lived, warm-core eddies were shed
from this region (Fang and Morrow, 2003).
(2) South-west of Shark Bay (2627
o
S, 113
114
o
E) over one-third from 2831
o
S. At
Shark Bay (~25
o
S, cover picture), the coastal
topography undergoes a 90
o
change in
orientation: to the north, flow along isobaths
is directed to the southwest; to the south, the
flow turns abruptly to the southeast (Fang
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 97
and Morrow, 2003; Woo et al., 2006a). Field
data indicate that in this region the strength
of the Leeuwin Current changes in the wider
shelf off Shark Bay; the current speed is
weaker as it is distributed along the wider
shelf. To the south, the current accelerates as
the continental shelf narrows and continental
slope becomes steeper (Woo et al., 2006a).
(3) Western edge of the Abrolhos Island chain
(2829
o
S, 113114
o
E). The instabilities
generated in the Leeuwin Current as it flows
past Shark Bay and accelerates (see (2)
above), together with interaction with the
Leeuwin Undercurrent, results in the
generation of eddies in this region
(Meuleners et al., 2006).
(4) South-west of Jurien Bay (2930
o
S, 114
115
o
E). The eddies generated offshore of the
Abrolhos Islands have a length scale ~200
km (cover picture), and the interaction
between the eddies generated to the north and
the coastline at Jurien Bay results in the
offshore movement of water, resulting in the
generation of eddies (cover picture).
(5) Perth Canyon (32
o
S, 115
o
E). The Perth
Canyon is the major topographic feature
along the continental slope and has the effect
of trapping eddies within the canyon. The
influence of the Leeuwin Undercurrent in the
formation of eddies has been documented by
Rennie et al., 2006).
(6) South-west of Capes Naturaliste and Leeuwin
(34.5
o
S, 114.5
o
E). At Cape Leeuwin (cover
picture), the coastal topography undergoes a
90
o
change in orientation: to the north, flow
along isobaths is directed to the south; to the
south, the flow turns abruptly to the east.
(7) South of Albany (35.5
o
S, 118
o
E). Here, the
coastal topography also undergoes a change
in orientation: to the west, flow along
isobaths is directed to the southeast; to the
east, the flow is directed to the east-north-
east.
(8) South of Esperance (35
o
S, 123
o
E). Similar to
Albany, the changes in bathymetry and the
locations of the numerous islands (Recherche
Archipelago) result in the generation of
eddies.
4. The Leeuwin Undercurrent
The Leeuwin Undercurrent (LU) has received the
least attention in the literature. Studies by
Thompson (1984, 1987) indicated that there was an
equatorward undercurrent flowing beneath the
Leeuwin Current. Current meter data from the
LUCIE experiment (Smith et al., 1991) confirmed
the observations of Thompson (1987) and indicated
that the equatorward undercurrent was narrow and
situated between 250 m and 450 m depth contours,
adjacent to the continental slope. The Undercurrent
transports 5 Sv of higher salinity (> 35.8), oxygen-
rich, nutrient-depleted water at a rate of 0.320.40
ms
-1
northward (Thompson, 1984). Measurements
indicate the current is stronger during November to
January (Thompson 1984; Smith et al. 1991).
The LU is driven by an equatorward geopotential
gradient located at the depth of the Undercurrent
(Thompson, 1984). The Leeuwin Undercurrent is
apparent in the geopotential anomaly data
500db/3000db (Wyrtki, 1971) and 450db/1300db
(Godfrey and Ridgway, 1985) with subsurface slopes
of 0.4 x 10
-7
and 0.2 x 10
-7
reported by Thompson
(1984) and Smith et al. (1991), respectively. Woo et
al. (2006a), through analysis of field along the 1000-
m contour along the west coast, estimated a slope of
1 x 10
-7
, which is the driving force of the LU
equatorward enhanced by an east-west subsurface
slope generated by the downwelling of the Leeuwin
Current.
The LU is closely associated with the subantarctic
mode water (SAMW) (see Section 6). A feature of
this water mass, resulting from convection to the
region south of Australia, is high, dissolved oxygen
concentration; thus a cross-section of the LU core
can be identified from the dissolved oxygen
distribution: the core of the current consists of a
dissolved oxygen maximum (252 M/L) centred at a
depth of approximately 400 m (Figure 3).
The Leeuwin Undercurrent may be considered an
extension of the Flinders Current northwards along
the west coast (see section 5).
5. Flinders Current
The Flinders Current (FC)the only northern
boundary current in the Southern Hemisphereis
the dominant feature along the southern coast of
Australia and extends from Tasmania to Cape
Leeuwin (Bye, 1972). The current is driven by a
positive wind stress curl (in both summer and
winter), which results in a northward transport (due
to the Sverdrup balance), centred along 135E, which
is deflected to the west to satisfy vorticity dissipation
and mass conservation (Middleton and Cirano,
2004). The current is present in both summer and
winter, and the westward transport, between latitudes
37S and 39S, varies between 8 Sv and 17 Sv and is
stronger during the summer months. The current
extends through the water column to a maximum
depth of 800 m, with maximum currents up to 0.5
ms
-1
in the offshore region (Figure 4). The FC
interacts with the Leeuwin Current (LC) at the shelf
break and slope. Here, the LC is observable near the
shelf break/slope as a surface current flowing
eastward, whereas the FC, located as a subsurface
current, is flowing westward (Figure 4). Comparison
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 98
of temperature and salinity data collected along the
south and west coasts indicated the Flinders Current
was the source of the subantarctic mode water
(Section 7), which forms the core of the Leeuwin
Undercurrent (Section 4).
6. Coastal currents off Western Australia
The structure of the continental shelf circulation
during the summer months, along the west coast of
Australia, has been addressed using field data and
satellite imagery in several recent studies (Cresswell
et al., 1989; Cresswell and Peterson, 1993; Pearce
and Pattiaratchi, 1997; Pearce and Pattiaratchi, 1999;
Gersbach et al., 1999; Woo et al., 2006a,b; Hanson
et al., 2005a,b). These studies have shown the
existence of a cooler, northward current on the
continental shelf (the Capes and Ningaloo currents)
with the southward flowing Leeuwin Current, in
general, located farther offshore.
Capes Current
Pearce and Pattiaratchi (1999) defined the Capes
Current (CC) as a cool, inner shelf current,
originating from the region between Capes Leeuwin
(34S) and Naturaliste, which moves equatorward
along the southwestern Australian coast in summer
(Figure 5). It has been postulated the CC may
extend as far north as the Abrolhos Islands (28-
29S), and this has been confirmed through field data
by Woo et al. (2006a). The current is high in salinity
(35.3735.53) and cooler (21.0C21.4C) than the
Leeuwin Current.
The Capes Current appears to be well established
around November when winds in the region become
predominantly southerly because of the strong sea
breezes (Pattiaratchi et al., 1997) and continues until
about March when the sea breezes weaken.
Gersbach et al. (1999) showed that the source water
of the Capes Current was through upwelling between
Capes Leeuwin and Naturaliste, which was
augmented by water from the south, to the east of
Cape Leeuwin.
Gersbach et al. (1999) described the dynamics of the
Capes Current, off Cape Mentelle. Here, the
southerly wind stress overcomes the alongshore
pressure gradient. This results in the surface layers
moving offshore, colder water upwelling onto the
continental shelf, and the Leeuwin Current migrating
offshore (Figure 6). Numerical model results have
shown that a wind speed of 8 ms
-1
is sufficient to
overcome the alongshore pressure gradient on the
inner continental shelf (Gersbach et al., 1999).
The Capes Current is sourced from shallow
upwelling of water from the bottom of the Leeuwin
Current (~100 m) (Gersbach et al. 1999; Pearce and
Pattiaratchi 1999; Hanson et al., 2005b). This water
mostly derives from the region between Capes
Naturaliste and Leeuwin.
Ningaloo Current
The northward flowing Ningaloo Current (NC) is
located along the inner shelf between the LC and the
coast. It is driven by a strong southerly wind stress
(Taylor & Pearce, 1999; Woo et al. 2006b), similar
to the Capes Current along the southwestern
Australian coastline (see above). Early observations
from aerial surveys (Taylor & Pearce, 1999)
identified the NC moving northward along the
Ningaloo reef front, forming a distinct line in the
water and separating the coastal waters from the
southward flowing LC some 2 km offshore; these
observations were subsequently verified using
satellite imagery (Taylor & Pearce, 1999), field data
(Woo et al., 2006a; Hanson et al., 2005a) and
numerical modelling (Woo et al., 2006b). These
studies indicated the Ningaloo Current was present
inshore of the 50-m isobath, extending from Shark
Bay (cover picture) to North West Cape and beyond.
At Point Cloates, under low wind conditions, the
current bifurcates, with an anticyclonic circulation
pattern located to the southwest of Point Cloates.
Here, part of the Ningaloo Current moves across the
shelf and then flows southward parallel to the coast;
the LC, although part of the current, flows northward
adjacent to the Ningaloo reef (Woo et al., 2006b).
Field data indicated that the NC consisted of colder
(< 23C), more saline (34.92) water when compared
with offshore waters, but had a similar T/S signature
to the LC water (Woo et al., 2006a). Off North-West
Cape, the NC exhibited a clear upwelling signal with
upsloping isotherms, higher nutrient concentrations,
higher phytoplankton biomass and maximum
regional primary production rates (Hanson et al.,
2005a).
Cresswell Current
The dynamics of coastal circulation in the southern
region are largely unknown. Upwelling of cold,
deep water in this region was observed in studies
from the Recherch Archipelago and adjacent waters
(van Hazel, 2001). It has been postulated that the
wind-driven coastal current, the Cresswell Current,
moving east to west in response to the strong,
summer, southeasterly winds south of Western
Australia, similar to the Capes Current and Ningaloo
Current along the west coast, is the major driving
force of the current and upwelling (Figure 6). The
westward flowing current will move the Leeuwin
Current offshore during the summer.
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 99
Water mass Temperature range Salinity range Dissolved oxygen
range
Tropical surface water (TSW) 2224.5
0
C 34.735.1 200220 M/L
South Indian central water (SICW) 1222
0
C 35.135.9 220245 M/L
Subantarctic mode water (SAMW) 8.512
0
C 34.635.1 245255 M/L
Antarctic intermediate water
(AAIW)
4.58.5
0
C 34.434.6 115245 M/L
North west Indian intermediate
(NWII) water
5.56.5
0
C ~ 34.6 100110 M/L
Table 1: Characteristics of the water masses found along the 1000-m isobath along the WA coastline.
7. Water mass characteristics
Analysis of temperature data from
bathythermographs or CTDs reveals the general
structure of the water column. Usually, a well-mixed
layer (WML) exists at the surface, which is produced
by turbulent mixing, e.g. by surface wind stress, and
also the presence of the Leeuwin Current. The WML
is deeper within the Leeuwin Current than onshore
or offshore. The variation in the WML can be greater
than 100 m. The mixed layer depth varies with
season and with ENSO events (Feng et al. 2003).
Below the WML the thermocline usually descends to
around 400 m, although sub-layers may exist in this
depth range. The seasonal changes include a cooler
sea surface temperature (SST) corresponding to the
austral winter, and warmer SST in the austral
summer, with the deepest mixed layer depths
occurring during winter (Hamilton 1986; Feng et al.
2003).Woo et al. (2006a) identified five different
water mass types in the upper Indian Ocean along
the West Australian coast (see Table 1) that
correspond with accepted classical water masses of
the Indian Ocean (Wyrtki, 1971; Warren, 1981).
These water masses were observed in the vertical
distribution of salinity and dissolved oxygen as
interleaving layers of salinity and dissolved oxygen.
In terms of increasing depth these water masses
were:
(i) Lower salinity tropical surface water
(TSW)
(ii) Higher salinity south Indian central water
(SICW)
(iii) Higher oxygen subantarctic mode water
(SAMW)
(iv) Lower salinity Antarctic intermediate
water (AAIW)
(v) Lower oxygen northwest Indian
intermediate (NWII) water.
Figure 1. Schematic of surface and subsurface currents along the continental shelf and slope off WA.
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 100
Figure 2. Schematic of the major surface currents in the southeast Indian Ocean.
Figure 3. Cross-section of dissolved oxygen concentration along 29
o
S showing the presence of a > 252
microM/L core at 400-m depth, which is interpreted as the core of the Leeuwin Undercurrent.
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 101
Figure 4. A cross-shelf transect of the currents offshore Albany obtained using a shipborne Acoustic
Doppler profiler showing the eastward flowing Leeuwin Current on the surface and at the shelf break and
upper slope with the westward flowing Flinders Current located beneath and offshore of the LC
(units are ms
-1
)
.
The location of each of the above five water masses
and their position relative to each other can be
identified for the whole length of the coastline from
North West Cape (21
o
S) to Cape Leeuwin (35
o
S)
using both salinity and oxygen (Figure 7). In the
following sections, the characteristics of each of the
water masses are discussed in detail.
Tropical surface water (TSW)salinity minimum
A layer of lower salinity (< 35.1), warmer (> 22

C),
tropical water was found in the surface water in the
northern region and corresponded with the
temperature/salinity characteristics of the Leeuwin
Current water. This water mass is derived from the
Australasian Mediterranean water (AAMW), a
tropical water mass with origins in the Pacific Ocean
central water and formed during transit through the
Indonesian archipelago (Tomczak and Godfrey,
1994). At the North West Cape (21
o
S), the northern
extent of the study region, this water mass extends to
180 m, with a surface salinity < 34.9 (Figure 9). The
depth of the water mass decreases southwards with
the passage of the Leeuwin Current and at ~26
o
S its
salinity signature (< 35.1) disappears because of the
entrainment of cooler, more saline south Indian
central water (see below) from offshore due to
geostrophic inflow (Woo et al., 2006a).
South Indian central water (SICW)salinity
maximum
South Indian central water (SICW) is identified here
as a salinity maximum layer (35.135.9). Along the
1000-m bathymetric contour, shipborne ADCP data
revealed the SICWs core moving northward along
26.8 density (
t
) level, with a maximum speed of 0.3
ms
-1
. However, near the shelf break this same water
mass is part of the Leeuwin Current flowing
southwards (Woo et al., 2006a). Here, ADCP data
indicated that the Leeuwin Current extends up to
300-m water depth, which is the total depth of this
water mass (Figure 9). SICW had a temperature
range of 1222

C and was associated with weak


minima of dissolved nitrate, silica, and phosphate; it
was found at the surface south of 29.0

S, and the
depth of the salinity maximum increased northward
from the surface at 29.0

S to 245 m at 21.5

S. In the
northern latitudes of the study region, the water mass
subducted underneath the tropical surface water
derived from the Australasian Mediterranean water
(AAMW).
Subantarctic mode water (SAMW)oxygen
maximum
Beneath the south Indian central water (SICW), a
water mass with high dissolved oxygen
concentrations of 245255 M/L can be identified as
subantarctic mode water (SAMW), whose core
occurred at 400510 m. The data revealed that
SAMW consisted of water with a temperature range
of 8.512

C and salinity range of 34.635.1.


SAMW is formed by deep winter convection at 40

50

S in the zone between the subtropical


convergence and the Subantarctic Front to the south
of Australia (Wong, 2005). As SAMW is formed by
deep convection rather than subduction, newly
formed SAMW penetrates to a greater depth than the
newly subducted SICW (thus, is comparatively
better ventilated) and then moves northward from its
formation region. Due to its high oxygen content, the
SAMW plays an important role in ventilating the
lower thermocline of the southern hemisphere
subtropical gyres (McCartney, 1982).
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 102
Figure 5. (a) Sea surface temperature and (b) Modis ocean colour images off southwestern Australia
showing upwelling of cold water onto the Capes Current and the associated high chlorophyll concentration.
Figure 6. Schematic of the steady-state summer current regime off southwestern Australia (Gersbach et al.,
1999).
Figure 7. Major water masses observed at the 1000-m isobath along the Western Australian shelf. Asterisks
on the surface indicate CTD station positions.
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 103
Antarctic intermediate water (AAIW)salinity
minimum
Below the SAMW, a salinity minimum (34.434.6)
was observed, indicating the presence of Antarctic
intermediate water (AAIW) along the coast. The
water was cold (4.5

C) and the position of its core


became shallower northward (core depth of 875 m at
27.5

S and 520 m at 21.5

S). It has been reported the


AAIW extends northward from the Antarctic Polar
Front to latitudes 10

15

S, and is thought to flow


more slowly than the oxygen maximum layer above
it (Warren, 1981).
Northwest Indian intermediate (NWII) water
oxygen minimum
An oxygen minimum signature of < 110 M/L in the
northern region (21.3

24.5

S) indicated the presence


of northwest Indian intermediate (NWII) water
immediately beneath the AAIW. This water mass
originated from the Red Sea and Persian Gulf
outflows. Occupying depths of 8001175 m, its
orientation implied southward deepening, and
evidence of the water mass has been found within
the Perth Canyon at 30
o
S (Rennie, 2005). The
temperature of the NWII water was < 5

C and its
salinity ranged between 34.55 and 34.65. NWII
water was associated with maxima of dissolved
nitrate, silica, and phosphate.
Acknowledgements
This is a shorter version of an article prepared for the
National Oceans Office on the physical
oceanography of Western Australia between Shark
Bay and Esperance. The more recent works
described here are based on a series of research
voyages undertaken using the National Facility
research vessels RV Franklin (November 2000) and
RV Southern Surveyor (October 2003 and April
2006). The author would like to thank Dr Mun Woo
and Dell Mohd Akhir for help with the preparation
of the figures and Ruth Gongora-Mesas for proof
reading.
References
Andrews, J.C. (1977). Eddy structure and the West
Australian Current. Deep-Sea Research, 24, 1133-
1148.
Andrews, J.C. (1983) Ring structure in the poleward
boundary current off Western Australia. Australian
Journal of Marine and Freshwater Research, 34,
547-561.
Batteen M.L. and Rutherford M.J. (1990) Modelling
studies of eddies in the Leeuwin Current: the role of
thermal forcing. J. Physical Oceanography,
20:1484-1520.
Bye, J. A. T., (1972) Ocean Circulation South of
Australia, in Antarctic Oceanology II: The
Australian -New Zealand Sector, edited by D. E.
Hayes, A.G.U., Antarctic Res. Series, Vol 19, 95-
100.
Church J.A., Cresswell G.R. and Godfrey J.S. (1989)
The Leeuwin current. In: Neshyba, S., Mooers
C.N.K. Smith R.L. and Barber R.T. (eds). Poleward
flow along eastern ocean boundaries. Lecture notes
on Coastal and Estuarine studies 34. Springer-
Verlag, 230-252.
Cresswell, G.R. (1996). The Leeuwin Current near
Rottnest Island, Western Australia. Marine and
Freshwater Research, 47,483-487.
Cresswell G., Boland F.M., Peterson J.L. and Wells G.S.
(1989). Continental Shelf current near Abrolhos
Islands, Western Australia. Australian Journal
Marine Freshwater Research, 40, 113-128.
Cresswell, G.R. and Golding, T.J. (1980). Observations
of a south-flowing current in the southeastern Indian
Ocean. Deep-Sea Research, 27A, 449-466.
Cresswell G. and Peterson J. (1993). The Leeuwin
Current south of Western Australia. Australian
Journal Maine Freshwater Research, 44:285-303.
Fang, F. and R. Morrow (2003). Evolution, movement
and decay of warm-core Leeuwin Current eddies.
Deep Sea Research II. 50: 2245-2261.
Feng, M., G. Meyers, A. F. Pearce and S. Wijffels
(2003). Annual and inter-annual variations of the
Leeuwin Current at 32S. Journal of Geophysical
Research, 108, 3355.
Feng, M., S. Wijffels, J. S. Godfrey and G. Meyers
(2005). Do Eddies Play a Role in the Momentum
Balance of the Leeuwin Current? Journal of
Physical Oceanography, 35, 964-975.
Fieux, M., R. Molcard and R. Morrow (2005). Water
properties and transport of the Leeuwin Current and
Eddies off Western Australia. Deep-Sea Research I,
52, 1617-1635
Gersbach G., Pattiaratchi C.B., Ivey G. and Cresswell.
(1999). Upwelling on the southwest coast of
Australiasource of the Capes Current? Continental
Shelf Research, 19: 363-400
Griffin, D. A., J. L. Wilkin, A. F. Pearce and C. F.
Chubb (2001). Sea surface temperatures and currents
off Western Australia, 1993-2000, CSIRO Marine
Research.
Godfrey J.S. and Ridgway K.R. (1985). The large-
scale environment of the poleward- flowing Leeuwin
current, Western Australia: Longshore steric height
gradients, wind stresses and geostrophic flow. J.
Physical Oceanography, 15 :481-495.
Godfrey, J. S., Vaudrey, D. J. and Hahn, S. D. (1986).
Observations of the shelf-edge current south of
Australia, winter 1982. Journal of Physical
Oceanography, 16, 668-79.
Bulletin of the Australian Meteorological and Oceanographic Society Vol.19 page 104
Hamilton L.J. (1986) Statistical features of the
oceanographic area off southwestern Australia,
obtained from bathythermograph data. Australian
Journal Maine Freshwater Research, 37:421-436.
Hanson C.E., Pattiaratchi C.B. and Waite A.M. (2005a).
Sporadic upwelling on a downwelling coast:
phytoplankton responses to spatially variable
nutrient dynamics off the Gascoyne region of
Western Australia. Continental Shelf Research, 25,
15611582.
Hanson, C.E., Pattiaratchi C.B. and Waite A.M.
(2005b). Seasonal production regimes off
southwestern Australia: Influence of the Capes and
Leeuwin Currents on phytoplankton dynamics.
Marine and Freshwater Research, 56, 1011-1026.
McCartney M. (1982). The subtropical re-circulation
of Mode waters. Journal of Marine Research, 40,
427-464.
Meuleners, M.J., Pattiaratchi, C.B. and Ivey, G.N.
(2006). Numerical modelling of the mean flow
characteristics of the Leeuwin Current System.
Deep Sea Research II, (in press).
Middleton, J. F. and M. Cirano (2002). A Northern
Boundary Current along Australias Southern
Shelves: the Flinders Current. Journal of
Geophysical Research, 107, 3129,
doi:10.1029/2000JC000701.
Morrow, R., F. Fang, M. Fieux and Molcard R. (2003).
Anatomy of three warm-core Leeuwin Current
eddies. Deep Sea Research II, 50: 2229-2243.
Pattiaratchi C.B., Hegge B., Gould J. and Eliot I. (1997).
Impact of sea-breeze activity on nearshore and
foreshore processes in southwestern Australia.
Continental Shelf Research, 17(13), 15391560.
Pattiaratchi C.B. and Buchan S. (1991) Implications of
long-term climate change for the Leeuwin Current.
Journal of Royal Society of Western Australia, 74:
133-140.
Pearce, A. F., and Griffiths, R. W. (1991). The
mesoscale structure of the Leeuwin Current: a
comparison of laboratory models and satellite
imagery. Journal of Geophysical Research, 96,
16739-47.
Pearce A. and Pattiaratchi C.B. (1999). The Capes
Current: a summer counter-current flowing past
Cape Leeuwin and Cape Naturaliste, Western
Australia. Continental Shelf Res. 19, 401-420.
Pearce A. and Pattiaratchi C.B. (1997). Applications of
satellite remote sensing to the marine environment in
Western Australia. Journal of Royal Society of
Western Australia, 80, 1-14.
Pearce A.F. and Walker D. (1991) The Leeuwin
Current: an influence on the coastal climate and
marine life of Western Australia. Journal of Royal
Society of Western Australia, 74: 1-140.
Rennie S. (2005). Oceanographic processes in the
Perth Canyon and their impact on productivity.
Unpubl. PhD Thesis. Curtin University of
Technology, 174pp.
Rennie S., Pattiaratchi C.B. and McCauley R. (2006).
Eddy formation through the interaction between the
Leeuwin Current, Leeuwin Undercurrent and
topography. Deep Sea Research II, (in press).
Ridgway K.R. and Condie S.A. (2004). The 5500-km-
long boundary flow off western and southern
Australia. Journal of Geophysical Research, 109,
doi:10.1029/2003JC001921.
Schott G. (1935). Geographie des Indischen und Stillen
Ozeans. Verlag von Boysen, Hamburg, 413pp.
Schott, F. and McCreary, J.P. (2001). The monsoon
circulation of the Indian Ocean. Progress in.
Oceanography, 51, 1123
Smith R., Huyer A., Godfrey S. and Church J. (1991).
The Leeuwin Current off Western Australia, 1986-
1987. Journal of Physical Oceanography, 21:323-
345.
Taylor, J.G. and Pearce, A.F. (1999). Ningaloo Reef
Current observations and implications for biological
systems: Coral spawn dispersal, zooplankton and
whale shark abundance. Journal of the Royal Society
of Western Australia, 82, 57-65.
Thompson R.O.R.Y. (1984). Observations of the
Leeuwin Current off Western Australia. Journal of
Physical Oceanography, 14:624-628.
Thompson R.O.R.Y. (1987). Continental-shelf scale
model of the Leeuwin Current. Journal of Marine
Research, 45:813-827.
Tomczak, M. and Godfrey, J.S. (1994). Regional
oceanography: An introduction, Pergamon, London.
Van Hazel J. (2001). Physical Oceanography of the
Recherche Archipelago/ Fitzgerald Biosphere area,
Unpubl. Hons Thesis, School of Water Research,
The University of Western Australia.
Warren, B.A. (1981). Trans Indian hydrographic section
at Lat 18
o
S: Property distributions and circulation in
the South Indian Ocean. Deep-Sea Research, 28A,
759-788.
Weaver A.J. and Middleton J.H. (1989) On the
dynamics of the Leeuwin Current. Journal of
Physical Oceanography, 19:626-648.
Wong A. P.S. (2005). Subantarctic model water and
Antarctic Intermediate water in the south Indian
Ocean based on profiling float data 2000-2004.
Journal of Marine Research, 63, 789-812.
Woo. M., Pattiaratchi C.B. and Schroeder W.W.
(2006a). Summer Surface Circulation on the
Gascoyne Continental Shelf, Western Australia.
Continental Shelf Research, 26, 132-152.
Woo M., Pattiaratchi C.B. and Schroeder W. (2006b).
Dynamics of the Ningaloo Current off Point Cloates,
Western Australia. Marine and Freshwater
Research, 57, 291301.
Wyrtki, K. (1971). Oceanographic Atlas of the
International Indian Ocean Expedition. National
Science Foundation, Washington, D.C., 531 pp.

Das könnte Ihnen auch gefallen