You are on page 1of 15

ARTICLE IN PRESS

Deep-Sea Research I 55 (2008) 1090– 1104

Contents lists available at ScienceDirect

Deep-Sea Research I
journal homepage: www.elsevier.com/locate/dsri

Hydrography and water masses off the western Australian coast


Mun Woo, Charitha Pattiaratchi 
School of Environmental Systems Engineering, The University of Western Australia, 35 Stirling Highway, Crawley 6009, Australia

a r t i c l e i n f o abstract

Article history: The water mass characteristics of the eastern Indian Ocean margin between latitudes 211
Received 24 October 2007 and 351S, adjacent to the Western Australian coast, are described using field measure-
Received in revised form ments. The results indicated the presence of five different water masses as interleaving
28 April 2008
layers of salinity and dissolved oxygen concentrations in the upper 1000 m of the ocean.
Accepted 7 May 2008
Available online 21 May 2008
These water masses included (i) lower salinity tropical surface water (TSW), (ii) higher
salinity south Indian central water (SICW), (iii) higher dissolved oxygen subantarctic
Keywords: mode water (SAMW), (iv) lower salinity Antarctic intermediate water (AAIW), and
Hydrography
(v) lower oxygen north-west Indian intermediate (NWII) water. Data collected in 2000
Water masses
and 2003 were compared with historical data (1987), and interannual variability in the
Western Australia
Leeuwin current tropical surface water and subantarctic mode water masses were identified and linked to
Leeuwin undercurrent El Niño–southern oscillation (ENSO) events. In the study region, the circulation pattern,
known as the Leeuwin current system, consists of three major currents: the Leeuwin
current (LC), the Leeuwin undercurrent (LU), and shelf current systems consisting of the
Capes and Ningaloo currents. In the study area’s northern region, geopotential gradients
were found driving the Leeuwin current (LC) and Leeuwin undercurrent (LU), with a
(negative) sea surface slope of 4  10 7 driving the LC poleward, and a (positive) slope of
1 10 7 beneath the LC driving the LU equatorward. Cross-shelf geopotential anomalies
revealed the surface layer (in the upper 300 m) sloped seaward at a gradient of 1.7  10 6,
and the subsurface layer (between depths of 300 and 730 m) sloped coastward at a
gradient of 6.3  10 7. This arrangement of geopotential slopes, together with the
positions of the LU relative to the LC, indicated a dynamical relationship between the LU
and the LC. The data indicated the LU’s core was located at a depth of 400 m and it
transported subantarctic mode water (SAMW) equatorward. The presence of the Leeuwin
current and Leeuwin undercurrent at the continental shelf break and slope, respectively,
influenced the water mass distribution at the continental margin. Leeuwin Current
induced downwelling caused the surface and subsurface water mass (SICW) and the
SAMW’s upper edge to slope downward toward the shelf break while subsurface
upwelling beneath the LU moved the Antarctic intermediate water (AAIW) and the
SAMW’s bottom edge higher in the water column.
& 2008 Elsevier Ltd. All rights reserved.

1. Introduction Western Australian coast, where, although the wind


regime is similar to that of other eastern ocean margins,
Eastern ocean basins are highly productive ecosystems the waters are oligotrophic. Detailed studies of the
that support high primary productivity and large pelagic circulation patterns off the WA coast initially addressed
finfish stocks. An exception to this rule, however, is off the why the circulation off Western Australia was different
from that of other eastern margins, leading to the
discovery of the Leeuwin current (LC) and its dynamics
 Corresponding author. Tel.: +618 64883179; fax: +618 64881015. (Cresswell and Golding, 1980; Thompson, 1987; Smith
E-mail address: chari.pattiaratchi@uwa.edu.au (C. Pattiaratchi). et al, 1991); subsequent studies addressed the LC system’s

0967-0637/$ - see front matter & 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.dsr.2008.05.005
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1091

dynamics and variability (Pattiaratchi and Buchan, 1991; southerly and has the greatest contribution from sea
Ridgway and Condie, 2004; Feng et al., 2003; Meuleners breezes (Pattiaratchi et al., 1997).
et al., 2007). In this paper, detailed field data—conductivity-tem-
In the absence of detailed data collected from the perature-depth (CTD), dissolved oxygen, and acoustic
continental shelf and slope waters, the region’s hydro- Doppler current profiler (ADCP) data—collected along
graphy has been inferred from transoceanic sections the Western Australian coast (Fig. 1) are used to describe
performed across other parts of the south Indian Ocean, the hydrography and circulation of the upper 1000 m of
including (a) meridional sections along 1101E (Rochford, the water column.
1969a), (b) zonal sections along 181S (Warren, 1981; Field,
1997), and (c) zonal sections along 321S (Wyrtki, 1971; 2. Data sets
Toole and Warren, 1993). Within the study region, the
accepted classical model of water masses for the upper The data reported here were collected during two field
1000 m of ocean consists of south Indian central water campaigns performed in the early austral summer
(SICW) and Antarctic intermediate water (AAIW) (Pinet, (October/November) of 2000 and 2003. Eleven cross-shelf
1992); however, data from other studies (see a–c above) transects were obtained in the northern section of the
have revealed the presence of additional water masses, study region, between 211S and 281S (Fig. 1), between 13
including lower salinity tropical surface water (TSW), and 27 November 2000 using the RV Franklin. The
subantarctic mode water (SAMW), and north-west Indian instruments deployed included a Neil Brown CTD recorder
intermediate (NWII) water. Fieux et al. (2005) also with a 24  5-L bottle Niskin rosette for calibration and
confirmed the presence of the above water masses. water sampling, a 150-kHz RDI acoustic Doppler current
The LC system consists of three major current systems: profiler (ADCP) linked to the global positioning system, a
the Leeuwin current, Leeuwin undercurrent, and shelf Turner Designs fluorometer, a near-surface thermosalino-
current systems consisting of the Ningaloo, and Capes graph, and meteorological sensors. Data from the 14 cross-
currents (Woo et al., 2006a; Rennie et al., 2007). The LC is shelf transects in the southern section were obtained from
an anomalous eastern boundary current, which carries 24 October to 9 November 2003 using the RV Southern
warm, fresh tropical waters poleward along the continen- Surveyor. Here a Sea-Bird CTD was used to collect the CTD
tal shelf break. Extensive observational and modelling data. On both voyages, the cross-shelf stations extended
studies have shown that an alongshore geopotential from 50 to 1000 m water depths. Depending on the width
gradient, which overwhelms the opposing equatorward of the continental shelf, 10–15 stations were occupied at
wind stress, causes the LC (Thompson, 1984, 1987; the 50, 100, 150, 200, 250, 300, 500, 750, and 1000 m
Godfrey and Ridgway, 1985; Weaver and Middleton, depth contours.
1989; Batteen and Rutherford, 1990; Pattiaratchi and Additional historical CTD data1 obtained during
Buchan, 1991; Ridgway and Condie, 2004). The LC is January–March of the 1987 austral summer were re-
weaker (1.5 Sv) as it flows against the southerly (oppos- trieved for comparison purposes from RV Franklin voyages
ing) winds during October–March (summer) and stronger FR87/03 and FR87/04. All the CTD data were calibrated to
(7 Sv) when the southerly winds are weaker during World Ocean Circulation Experiment standards to allow
April–September (winter) (Smith et al., 1991). The mean comparison between years.
volume transport is estimated to be 3.4 Sv (Feng et al.,
2003). 3. Results and discussion
The LU flows northward beneath the LC at depths of
200–400 m, transporting 5 Sv of oxygen-rich, nutrient-
3.1. Water mass characteristics
depleted, higher salinity (435.8) water at a rate of
0.32–0.40 ms 1 northward (Thompson, 1984). An equator-
Five different water mass types corresponding to the
ward undercurrent was also apparent in steric height data
Indian Ocean water masses (Wyrtki, 1971; Warren, 1981;
at 500 dbar/3000 dbar (Wyrtki, 1971) and 450 dbar/
Fieux et al., 2005) were detected in the upper Indian
1300 dbar (Godfrey and Ridgway, 1985), as well as in
Ocean along the Western Australian coast (see Table 1).
current metre data (at 250–450 m) from the LC Inter-
These water mass types were observed as interleaving
disciplinary Experiment (Smith et al., 1991). It has been
layers in the vertical distribution of salinity and dissolved
suggested that an equatorward geopotential gradient at
oxygen. In order of increasing depth, these water masses
the depth of the undercurrent drives the LU (Thompson,
were:
1984) and that the LU’s water is advected northward
underneath the LC, transporting cooler, oxygen-rich,
higher salinity water from the surface of the south Indian (i) lower salinity tropical surface water (TSW),
Ocean (Thompson and Cresswell, 1983). (ii) higher salinity south Indian central water (SICW),
Inshore of the LC, wind-driven, higher-salinity currents (iii) higher oxygen subantarctic mode water (SAMW),
have been recorded at several locations on the Western (iv) lower salinity Antarctic intermediate water (AAIW),
Australian coast (Rochford, 1969b; Cresswell and Golding, (v) lower oxygen north-west Indian intermediate (NWII)
1980; Cresswell et al., 1985, 1989; Smith et al., 1991; water.
Gersbach et al., 1999; Pearce and Pattiaratchi, 1999; Woo
et al., 2006b). Generally, the northward coastal currents 1
CSIRO online database: http://www.marine.csiro.au/maru/marlin_
are strongest in summer when the wind pattern is mostly admin.source_list/.
ARTICLE IN PRESS
1092 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

Fig. 1. Location map of the research area, including the positions of CTD transect lines performed during voyages SS09/2003 and FR10/00 and CTD stations
(over 1000 m isobath) obtained from voyages FR87/03 and FR87/04.

Table 1
The different characteristics of each of the water masses found in the 1-km-deep water column defined

Water mass Temperature range (1C) Salinity range Dissolved oxygen range (mmol/L)

Tropical surface water (TSW) 22–24.5 34.7–35.1 200–220


South Indian central water (SICW) 12–22 35.1–35.9 220–245
Subantarctic mode water (SAMW) 8.5–12 34.6–35.1 245–255
Antarctic intermediate water (AAIW) 4.5–8.5 34.4–34.6 115–245
North-west Indian intermediate (NWII) water 5.5–6.5 34.6 100–110

This table combines data from voyages FR10/00 (21.3–27.91S) and SS09/2003 (28.1–351S).
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1093

When the temperature–salinity diagram and tempera- mass was associated with lower nutrient (near zero) and
ture–oxygen diagram obtained from measurements from higher dissolved oxygen concentrations. (The reader is
the northern study region were compared, it was evident referred to Woo et al. (2006a), for detailed discussions on
the salinity and dissolved oxygen distributions had an the different coastal surface water types and their
inverse relationship (Fig. 2); this was highlighted when dynamics.)
the spatial distribution of properties was examined. At North West Cape (211S)—the study region’s north-
A north–south cross-section along the 1000 m isobath ern extreme—this water mass extended to 180 m (Fig. 4)
line, parallel to the shore, is shown in Fig. 3. The salinity with a surface salinityo34.9. The water mass depth
distribution along the cross-section indicated a salinity decreased southward along the path of the Leeuwin
maximum and two minima (Fig. 3a); the dissolved oxygen Current. The LC dynamics caused the salinity signature
distribution indicated an oxygen maximum and two (o35.1) of the water mass to disappear at 261S. An
minima, but located at different depths from the salinity alongshore geopotential gradient drives the Leeuwin
maximum and minima (Fig. 3b). current, and geostrophic inflow entraining cooler, more
The locations of the five water masses and their saline SICW (see below) from offshore is a LC feature (Woo
positions relative to each other were identified using et al., 2006b).
salinity and oxygen (Fig. 4) for the entire length of the
coast, from North West Cape (211S) to Cape Leeuwin
(351S) (see Fig. 1 inset). In the following sections, the 3.3. SICW—salinity maximum
characteristics of each of the water masses are discussed
in detail. SICW was identified here as a salinity maximum layer
(35.1–35.9). Along the 1000 m bathymetric contour, ADCP
3.2. TSW—salinity minimum data revealed the SICW’s core moving northward along
26.8sT with a maximum speed of 0.3 ms 1. Near the shelf
A layer of warmer (422 1C), lower salinity (o35.1) break, however, this same water mass was part of the LC
tropical water, which corresponded with the tempera- flowing southward (Woo et al., 2006b). Here, ADCP data
ture–salinity characteristics of the LC water, was found in indicated the LC extended to 300 m in depth, which was
the surface water in the northern region. This water mass the total depth of this water mass (Fig. 4). The SICW had a
was derived from the Australasian Mediterranean temperature range of 12–22 1C and was associated with
water—a tropical water mass with origins in the Pacific weak minima of dissolved nitrate, silica, and phosphate; it
Ocean central water and formed during transit through was found at the surface south of 29.01S, and the salinity
the Indonesian archipelago (Tomczak and Godfrey, 1994; maximum depth increased northward from the surface at
Wijffels et al., 2002). Field data revealed this surface water 29.01S to 245 m at 21.51S.

Fig. 2. Temperature–salinity and temperature–oxygen diagrams showing interleaving positions of property extrema.
ARTICLE IN PRESS
1094 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

Fig. 3. Three-dimensional ‘blocks’ of ocean depicting the cross-shelf (across transect J, the southernmost transect obtained from FR10/00) and along-shelf
(along 1000 m isobath) distribution of (a) salinity extrema and (b) dissolved oxygen extremes.

In the study region’s northern section, SICW salinity maximum agreed with the results of Wyrtki
was subducted underneath the TSW derived from Aus- (1971), who found higher salinity water across the
tralasian Mediterranean water. The observed surface breadth of the surface Indian Ocean at latitudes ranging
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1095

Fig. 4. Major water masses observed at the 1000 m isobath along the Western Australian shelf. Asterisks on the surface indicate CTD station positions.
This chart combines data from voyages FR10/00 (21.3–27.91S) and SS09/2003 (28.1–35.21S).

Fig. 5. Sparse CTD data indicating the major water masses observed in 1987 at the 1000 m isobath. Subantarctic mode water (SAMW) was less ventilated
in 1987 than in 2000/3 (Fig. 4). Asterisks on the surface indicate CTD station positions.

from 25 to 351S. At these latitudes, an excess of This is then subducted below the surface water (Karstensen
evaporation over precipitation forms the higher salinity and Tomczak, 1997), extending northward to 12–161S
water at the sea surface (Baumgartner and Reichel, 1975). (Church et al., 1989), where it meets the lower salinity
ARTICLE IN PRESS
1096 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

Australasian Mediterranean water flowing westward from Tomczak, 1997); southern subtropical surface water
the Indonesian archipelago in the South Equatorial current (Muromtsev, 1959); tropical surface waters (Ivanenkov
(Sharma, 1972; Tomczak and Godfrey, 1994). and Gubin, 1960); and subtropical surface water (Wyrtki,
In the literature, SICW has been identified with 1973; Fieux et al., 2005).
different terminology but having the same TS character-
istics. Here, we follow the terminology of Rochford 3.4. SAMW—oxygen maximum
(1969a) and Webster et al. (1979) and identified it as
SICW. The same water mass, with a salinity maximum, Beneath the SICW, a water mass with high dissolved
was termed the Indian central water (Karstensen and oxygen concentrations of 245–255 mM/L and a core

Fig. 6. Differences between observations from 1987 (Fig. 5) and 2000/3 (Fig. 4) of (a) dissolved oxygen and (b) salinity. The shaded areas indicate values
were greater in 1987 than in 2000/3.
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1097

occurring at 400–510 m was identified as SAMW. Comparing the data sets obtained over the different
The data revealed the SAMW had a temperature range of years revealed the SAMW’s characteristics changed
8.5–12 1C and a salinity range of 34.6–35.1; its s T value with the prevailing atmospheric conditions. The CTD data
ranged between 28.9 and 29.5. from RV Franklin voyages FR03/87 and FR04/87 (Fig. 1)
Deep winter convection at 40–501S in the zone were used to obtain a composite image of water masses
between the subtropical convergence and the subantarctic following the 1000 m isobath along the Western Austra-
front to the south of Australia forms the SAMW (Wyrtki, lian coastline (Fig. 5). Although of low resolution (com-
1973; Colborn, 1975; McCartney, 1977; Toole and Warren, pared with Fig. 4) because of the scarcity of data, the
1993; Karstensen and Tomczak, 1997; Sallée et al., 2006). figure, which shows data from 1987, clearly shows
As a deep convection rather than subduction forms the change in the SAMW; its dissolved oxygen signa-
SAMW, newly formed SAMW penetrates to a greater ture was 5–10 mmol/L higher in 2000 than it was in 1987
depth than newly subducted SICW (thus is comparatively
better ventilated) and moves northward from its forma-
tion region. The SAMW is important because its high
oxygen content ventilates the lower thermocline of
the southern hemisphere subtropical gyres (McCartney,
1982). The SAMW, which is formed to the south of
Australia (Wong, 2005) and transported westward by the
Flinders current, forms the source water for the LU (see
below).
The SAMW also corresponds to the Indian Ocean
central water (ICW), as defined by Sverdrup et al. (1942).
SAMW and ICW often have similar temperatures and
salinities; consequently, SAMW has been thought to
contribute to the ICW’s depth range (Karstensen and
Tomczak, 1997). However, Karstensen and Tomczak (1997)
showed the source characteristics of the SAMW differed
from region to region depending on the prevailing atmo-
spheric conditions when the SAMW was formed. The data
sets presented here, which were collected during non-El
Niño–southern oscillation (ENSO) years (2000 and 2003), Fig. 8. Schematic diagram illustrating the general flow patterns at the
with data from 1987, an ENSO year, confirmed this. continental margin.

Fig. 7. 1000 m isobath cross-sections of (a) dissolved oxygen (mmol/L) and (b) fluorescence (uncalibrated units). The white line in both plots shows the
core of the band of shallow oxygen minimum water.
ARTICLE IN PRESS
1098 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

(Fig. 6a). The SICW’s lower boundary also overlapped with directly influenced the water mass formation, changes
the top of the SAMW (Fig. 5); this interaction lowered the were seen in the surface water; that is, the surface
salinity at that interface (Fig. 6b). salinity minimum was less intense, and surface water
Overall, all the major water masses’ positions remained with such low salinity did not seem to extend as far south
unchanged between the years, and there were no in 1987 as it did in 2000. The LC would have also
changes to the core-defining signature of any other contributed to these changes since it transports less low
subsurface water mass. Because atmospheric conditions salinity water southward during ENSO years than during

0.1

100
0
200
-0.1
300

400 -0.2
depth m

500 -0.3

600
-0.4
700
-0.5
800

900 -0.6

1000
111.5 112 112.5 113
longitude E

Fig. 9. Transect I cross-section of ADCP alongshore velocities (ms 1) showing an equatorward Leeuwin undercurrent (LU) and coastal current and a
poleward Leeuwin current (LC). The trace lines show the LC and LU positions, as Meuleners et al. (2007) numerically modelled them.

0 0
-0.1 -0.1
0

100 0.1
2
-0.
0

200 0.2 0.4


0.3

0.6
0.9 .8
0.5

300
1
0
0.7

400
depth m

500

600

700

800

900

1000
111.5 112 112.5 113
longitude °E
1
Fig. 10. Transect I cross-section of geostrophic flow (ms ) relative to the surface showing the Leeuwin undercurrent flowing equatorward at a depth of
400 m.
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1099

non-ENSO years (Cresswell et al., 1989; Pattiaratchi and with the atmosphere for long, presumably because of
Buchan, 1991). much slower overall horizontal flow at such depths,
produced the low oxygen values (Warren, 1981).
3.5. AAIW—salinity minimum
3.7. Shallow oxygen minimum
A salinity minimum (34.4–34.6) was observed beneath
the SAMW, indicating the presence of AAIW along the Although not associated with a particular water mass,
coast. The water was cold (4.5–8 1C), and its core became an oxygen minimum layer was observed beneath the
shallower northward (core depth of 875 m at 27.51S and surface layer throughout the study region (Fig. 4). It was
520 m at 21.51S); its sT values spanned 30.3–31.0. associated with maxima in nitrate, silica, and phosphate
The AAIW has been reported to extend northward from concentrations. Its core followed the 26.1sT level, reaching
the Antarctic polar front to latitudes 10–151S and to flow depths of 100–200 m. The oxygen minimum value
more slowly than the oxygen maximum layer above it increased northward from 204 mmol /L at 27.51S to
(Warren, 1981). 175 mmol/L at 21.51S.
Rochford (1969a), Webster et al. (1979), Warren (1981),
3.6. NWII water—oxygen minimum and Church et al. (1989) reported similar oxygen mini-
mum layers. Rochford (1969a) also found a salinity
An oxygen minimum signature ofo110 mmol/L in the minimum associated with the oxygen minimum layer,
northern region (21.3–24.51S) indicated the presence of and proposed that since the oxygen minimum strength-
NWII water immediately beneath the AAIW. Occupying ened southward, a southward advection of lower oxygen,
depths of 800–1175 m, with sT values of 31.8–32.4, its lower salinity tropical water had formed the layer. The LC
orientation implied southward deepening; as such, it may southward motion, which dominated the flow at the shelf
extend farther south into the deeper ocean. The recorded break and slope, likely caused the southward motion
NWII water temperature was less than 5 1C, and its Rochford (1969a) recorded closer to the shelf edge.
salinity ranged between 34.55 and 34.65. The NWII water In contrast to Rochford (1969a), the data collected
was associated with maxima of dissolved nitrate, silica, during this study did not indicate the shallow oxygen
and phosphate. minimum layer was associated with a lower salinity layer.
Rochford (1961), Newell (1974), Webster et al. (1979), The oxygen minimum layer was actually associated with
Warren (1981), and Toole and Warren (1993) observed a higher salinity south of 231S and lower salinity to the
similar water mass of Red Sea origin (Rochford, 1964) in north. This was because the depth of the shallow oxygen
other Indian Ocean regions. In situ consumption of minimum’s core remained constant at around 160 m
dissolved oxygen in water that had not been in contact throughout the study region (Fig. 4). Depending on the

Fig. 11. Cross-section of dissolved oxygen levels for transect I showing the presence of the 4252 mmol/L core at 400 m depth corresponding to the
Leeuwin undercurrent core.
ARTICLE IN PRESS
1100 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

latitude at which observations were made, the oxygen forms beneath the DCM regardless of the type of water
minimum layer was present in two different water mass present.
masses. To the north it existed within the surface tropical
water while to the south it was associated with the higher 4. Surface and subsurface current systems
salinity SICW sometimes within its core (i.e. at 26–311S).
The data indicated the shallow oxygen minimum layer
Along the Western Australian continental shelf margin,
was closely related to a deep chlorophyll maximum
there are three main current systems, collectively defined
(DCM), centred at around 80 m depth, which was found
here as the LC system (Fig. 8):
throughout the study area (Hanson et al., 2007). The
shallow oxygen minimum layer was located directly
beneath the DCM at each station (Fig. 7). This finding (1) the Capes and Ningaloo currents—wind-driven equa-
was important when coupled with the increased levels of torward surface currents on the continental shelf
nitrate, silica, and phosphate, as well as an increasing (usually in depthso50 m) present mainly in the
minimum northward, as it provided evidence to reject summer under strong southerly wind stress;
Rochford’s (1969a) hypothesis. Warren (1981) proposed (2) the Leeuwin current (LC)—a poleward surface current
the water could be northward flowing, with an intensify- located along the shelf break;
ing oxygen minimum occurring when sinking detritus (3) the Leeuwin undercurrent (LU)—an equatorward
decayed and released nutrients. Photosynthetic organisms subsurface undercurrent on the continental slope.
of the DCM live at depths with an optimal balance of
radiance from the surface and nutrients from below The dynamics of the Capes, Ningaloo, and LC were
(Hanson et al., 2005); hence the oxygen minimum layer discussed in detail in Church et al. (1989), Smith et al.

Fig. 12. Transect I cross-sections of (a) salinity and (b) temperature. Circles along the surface indicate CTD station positions.
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1101

(1991), Gersbach et al. (1999), Feng et al. (2003), Ridgway was identified from the dissolved oxygen distribution; the
and Condie (2004), and Woo et al. (2006a, b). Here, we current’s core consisted of a dissolved oxygen maximum
examine the LU dynamics, which have received little (252 mmol/L) centred at around 400 m depth (Fig. 11).
attention in the literature. Although the LU may contain some SICW in the upper
The three current systems, associated with the LC layers, the data presented here showed the LU consisted of
system, were identified from a cross-shore ADCP transect SAMW, and not a tongue of higher salinity SICW, as
obtained along transect I (Figs. 1 and 9). Because of the Thompson (1984) reported.
ADCP’s depth limitation (to about 300 m), only the LU’s It is generally accepted that an alongshore geopotential
upper portion could be identified; nonetheless, the data gradient drives the LC (Thompson, 1984; Godfrey and
indicated an undercurrent at greater depth. The ADCP data Ridgway, 1985; Smith et al., 1991), which through
were closely correlated with numerical modelling results geostrophy results in onshore flow and downwelling. This
(Meuleners et al., 2007; Rennie et al., 2007) which downwellling was evident as downward-sloping iso-
indicated the LU’s core was centred at a depth of 400 m therms and isohalines at the shelf break (Fig. 12) and
(Fig. 9). Northward flow in the surface layer between caused the surface water, together with the SICW and the
111.51E and 1121E (Fig. 9) is due to the presence of an eddy SAMW’s upper edge, to be depressed downward ap-
(see Woo et al., 2006b). proaching the shelf break (Fig. 3). Because the LU flows
Estimating the geostrophic flow obtained from the CTD northward, the opposite occurs, resulting in concomitant
data confirmed the depth of the LU’s core (Fig. 10). The LU upwelling. This upwelling was seen as upward-sloping
is closely associated with the subantarctic mode water isotherms and isohalines below 400 m (Fig. 12); hence the
(SAMW). Along transect I—the longest cross-shore CTD AAIW and NWII water and the SAMW’s bottom edge were
transect (Fig. 1)—a complete cross-section of the LU’s core drawn upward approaching the shelf slope (Fig. 3).

Fig. 13. Geopotential anomaly in m2 s 2 plotted versus latitude from CTD stations recorded along the 1000 m isobath with straight line fits calculated
between (a) the surface relative to 300 dbar level and (b) 300 dbar relative to 600 dbar level.
ARTICLE IN PRESS
1102 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

The alongshore geopotential gradients estimated from the subsurface layer to a maximum depth of 730 m to plot
the data along the 1000 m isobath (Fig. 13) indicated a sea a line with sufficient data points. To counter the scarcity of
surface slope of 4  10 7 (from north to south) at the data points, calculations were repeated for transects E, F,
surface 0–300 dB level, which is the driving force of the LC G, and H (Fig. 1, in Woo et al., 2006b), and the average
(see above). In the 300–1000 dB layer, the slope is values for geopotential slope were obtained later (other
reversed, with a value of 1 10 7, the driving force of transects were excluded, as their data were unsuitable
the LU equatorward. These values were comparable with for calculations because of insufficient measurement
previous studies; for example, Hamon (1965) found that depths or interference from jets and eddies). It was found
the surface slope was 2.8  10 7 whilst Thompson that the surface layer (6–300 m depth) sloped seaward at
(1984) and Smith et al. (1991) reported it as 3.6  10 7 a gradient of 1.7  10 6, and the subsurface layer
and 2.6  10 7 respectively. Similarly, for the Leeuwin (300–730 m depth) sloped landward at a gradient of
undercurrent, Thompson (1984) and Smith et al. (1991) 6.3  10 7. In geostrophic balance, this arrangement of
reported subsurface slopes of 0.4  10 7 and 0.2  10 7, geopotential slopes contributed to the poleward flow in
respectively. These results confirmed that the alongshore the surface layer and the equatorward flow in the
geopotential gradient remained almost constant through subsurface layer.
seasonal and inter-annual timescales (Hamon, 1965; The LC remains adjacent to the coast because of the
Godfrey and Ridgway, 1985). Coriolis force, but the dynamics that control the LU and its
The cross-slope geopotential anomalies (Fig. 14) in- position along the continental slope are unclear. Thomp-
dicated a thicker surface layer landward and a thicker son (1984) suggested the presence of the continental
subsurface layer seaward. We limited the calculations for slope allowed the constraint of the earth’s rotation to be

Fig. 14. Geopotential anomaly in m2s 2 plotted versus longitude from CTD stations recorded along the 1000-m isobath with straight line fits calculated
between (a) the surface relative to 300 dbar level and (b) 300 dB relative to 730 dbar level.
ARTICLE IN PRESS
M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104 1103

Batteen, M.L., Rutherford, M.J., 1990. Modeling studies of eddies in the


Leeuwin current: the role of thermal forcing. Journal of Physical
Oceanography 20, 1484–1520.
Baumgartner, A., Reichel, E., 1975. The World Water Balance: Mean
Annual Global, Continental and Maritime Precipitation, Evaporation
and Run-off. Elsevier, Amsterdam.
ρ1
Church, J.A., Cresswell, G.R., Godfrey, J.S., 1989. The Leeuwin current. In:
Neshyba, S.J., Mooers, C.N.K., Smith, R.L., Barber, R.T. (Eds.), Poleward
Flows Along Eastern Ocean Boundaries. Springer, New York, pp.
230–254.
Colborn, J.G., 1975. The Thermal Structure of the Indian Ocean. University
ρ2 Press of Hawaii, Honolulu.
Cresswell, G.R., Golding, T.J., 1980. Observations of a south-flowing
current in the southeastern Indian Ocean. Deep-Sea Research Part A:
Flow out of the paper
Oceanographic Research Papers 27, 449–466.
Cresswell, G.R., Golding, T.J., Boland, F.M., Wells, G.S., 1985. Current
Flow in to the paper measurements from sites near Abrolhos and Rottnest Islands,
Western Australia, 1973–75. Report. Division of Fisheries and
Oceanography, CSIRO Australia.
Fig. 15. A schematic of the Margules equation showing the interface Cresswell, G.R., Boland, F.M., Peterson, J.L., Wells, G.S., 1989. Continental
slopes and flow directions as applied to the Leeuwin current and shelf currents near the Abrolhos Islands, Western Australia. Aus-
Leeuwin undercurrent (adapted from Knauss, 1997). tralian Journal of Marine and Freshwater Research 40, 113–128.
Feng, M., Meyers, G., Pearce, A., Wijffels, S., 2003. Annual and interannual
variations of the Leeuwin current at 321S. Journal of Geophysical
Research—Oceans 108, 3355.
broken, although no theoretical explanation was given. Field, A., 1997. GRL special section: WOCE Indian Ocean expedition.
Based on the geopotential gradients discussed above, Geophysical Research Letters 24 (21), 2539–2540.
however, a mechanism based on the Margules equation Fieux, M., Molcard, R., Morrow, R., 2005. Water properties and transport
of the Leeuwin current and eddies off Western Australia. Deep-Sea
can be used (Fig. 15). At the shelf edge, the LC had an Research Part I: Oceanographic Research Papers 52, 1617–1635.
onshore component of flow, which depressed the Gersbach, G.H., Pattiaratchi, C.B., Ivey, G.N., Cresswell, G.R., 1999.
lower layer, especially beneath the LC. The subsurface Upwelling on the south-west coast of Australia—source of the Capes
current? Continental Shelf Research 19, 363–400.
layer’s cross-shore gradient thus increased, causing an
Godfrey, J.S., Ridgway, K.R., 1985. The large-scale environment of the
associated LU flow close to the continental slope. This poleward-flowing Leeuwin current, Western Australia: longshore
implied that wherever along the coast the LC flow caused steric height gradients, wind stresses and geostrophic flow. Journal
downwelling, an undercurrent was induced to flow of Physical Oceanography 15, 481–495.
Hamon, B.V., 1965. Geostrophic currents in the south-eastern Indian
beneath it. Ocean. Australian Journal of Marine and Freshwater Research 16,
Along Australia’s southern shelf, an undercurrent along 255–271.
the continental slope—the Flinders current—has also Hanson, C.E., Pattiaratchi, C.B., Waite, A.M., 2005. Sporadic upwelling on
a downwelling coast: phytoplankton responses to spatially variable
been identified beneath the eastward flowing LC. Here nutrient dynamics off the Gascoyne region of Western Australia.
the Flinders current flows westward (Middleton and Continental Shelf Research 25, 1561–1582.
Cirano, 2002). Similar to the LU, the Flinders current is Hanson, C.E., Pesant, S., Waite, A.M., Pattiaratchi, C.B., 2007. Assessing the
magnitude and significance of deep chlorophyll maxima of the
also found along the shelf slope and centred at a depth of
coastal eastern Indian Ocean. Deep Sea Research II: Topical Studies in
400 m (Middleton and Cirano, 2002). The LU is a Oceanography 54 (8–10), 884–901.
continuation of the Flinders current, transporting sub- Ivanenkov, V.N., Gubin, F.A., 1960. Water masses and hydrochemistry of
the western and southern parts of the Indian Ocean. Transactions of
antarctic mode water (SAMW) from its generation region
the Marine Hydrophysical Institute, Academy of Sciences of the
northward, as shown by Akhir and Pattiaratchi (2006). USSR, Physics of the Sea, Hydrology 22, 29–99.
Karstensen, J., Tomczak, M., 1997. Ventilation processes and water mass
ages in the thermocline of the southeast Indian Ocean. Geophysical
Research Letters 24 (22), 2777–2780.
Acknowledgements Knauss, J.A., 1997. Introduction to Physical Oceanography. Prentice-Hall
Inc., Englewood cliffs, NJ.
McCartney, M.S., 1977. Subantarctic mode water. In: Angel, M.V. (Ed.),
We thank the captains, crew, and staff of the RV A Voyage of Discovery: George Deacon 70th Anniversary Volume.
Franklin and FRV Southern Surveyor for successfully Supplement to Deep-Sea Research. Pergamon Press, Oxford,
pp. 103–119.
executing cruises FR10/00 and SS09/2003. We acknowl- McCartney, M.S., 1982. The subtropical recirculation of mode waters.
edge discussions with Will Schroeder and Ruth Gongora- Journal of Marine Research 40 (Suppl.), 427–464.
Mesas for the assistance in the preparation of the final Meuleners, M.J., Pattiaratchi, C.B., Ivey, G.N., 2007. Numerical modelling
of the mean flow characteristics of the Leeuwin current system.
manuscript. A University of Western Australia research
Deep-Sea Research Part II: Topical Studies in Oceanography 54,
grant, Vice Chancellor’s discretionary grant, and university 837–858.
postgraduate scholarship awarded to Mun Woo funded Middleton, J.F., Cirano, M., 2002. A northern boundary current along
this work. This contribution is School of Environmental Australia’s southern shelves: the Flinders current. Journal of
Geophysical Research 107, 3129.
Systems Engineering reference SESE 094. Muromtsev, A.M., 1959. Osnovnye Cherty Gidrologii Indiiskogo Okeana
(Principal Features of the Hydrology of the Indian Ocean). Gidrome-
References teorologicheskoe Izdatelstvo, Leningrad.
Newell, B.S., 1974. Distribution of oceanic water types off south-eastern
Tasmania, 1973. Report no. 59. Division of Fisheries and Oceano-
Akhir, M.F., Pattiaratchi, C.B., 2006. Summer physical processes along the graphy, CSIRO Australia.
continental shelf and slope off southern Western Australia. Proceed- Pattiaratchi, C.B., Buchan, S.J., 1991. Implications of long-term climate
ings of the Sixth International Symposium On Stratified Flows, Perth, change for the Leeuwin current. Journal of the Royal Society of
Western Australia, pp. 239–244. Western Australia 74, 133–140.
ARTICLE IN PRESS
1104 M. Woo, C. Pattiaratchi / Deep-Sea Research I 55 (2008) 1090–1104

Pattiaratchi, C.B., Hegge, B., Gould, J., Eliot, I., 1997. Impact of sea-breeze Thompson, R.O.R.Y., 1984. Observations of the Leeuwin current off
activity on nearshore and foreshore processes in southwestern Western Australia. Journal of Physical Oceanography 14, 623–628.
Australia. Continental Shelf Research 17 (13), 1539–1560. Thompson, R.O.R.Y., 1987. Continental shelf scale model of the Leeuwin
Pearce, A., Pattiaratchi, C.B., 1999. The Capes current: a summer current off Western Australia. Journal of Marine Research 45,
countercurrent flowing past Cape Leeuwin and Cape Naturaliste, 813–827.
Western Australia. Continental Shelf Research 19, 401–420. Thompson, R.O.R.Y., Cresswell, G.R., 1983. The Leeuwin current and
Pinet, P.R., 1992. Oceanography, an Introduction to the Planet Oceanus. undercurrent. Tropical Ocean–Atmosphere Newsletter 19, 10–11.
West Publishing Company, New York. Tomczak, M., Godfrey, J.S., 1994. Regional Oceanography: An Introduc-
Rennie, S.J., Pattiaratchi, C.B., McCauley, R.D., 2007. Eddy formation tion. Pergamon Press, London.
through the interaction between the Leeuwin current, Leeuwin Toole, J.M., Warren, B.A., 1993. A hydrographic section across the
undercurrent and topography. Deep-Sea Research Part II: Topical subtropical south Indian Ocean. Deep-Sea Research Part 1: Oceano-
Studies in Oceanography 54, 818–836. graphic Research Papers 40, 1973–2019.
Ridgway, K.R., Condie, S.A., 2004. The 5500-km-long boundary flow off Warren, B.A., 1981. Trans-Indian hydrographic section at Lat 181S:
western and southern Australia. Journal of Geophysical Research— property distributions and circulation in the south Indian Ocean.
Oceans 109 (C04017). Deep-Sea Research 28A, 759–788.
Rochford, D.J., 1961. Hydrology of the Indian Ocean. I. The water masses Weaver, A.J., Middleton, J.H., 1989. On the dynamics of the Leeuwin
in intermediate depths of the south-east Indian Ocean. Australian current. Journal of Physical Oceanography 19, 626–648.
Journal of Marine and Freshwater Research, 12, 129–149. Webster, I., Golding, T.J., Dyson, N., 1979. Hydrological features of the
Rochford, D.J., 1964. Hydrology of the Indian Ocean. III. Water masses of near shelf waters off Fremantle, Western Australia, during 1974.
the upper 500 m of the south-east Indian Ocean. Australian Journal Report no. 106. Division of Fisheries and Oceanography, CSIRO
of Marine and Freshwater Research 15, 25–55. Australia.
Rochford, D.J., 1969a. Seasonal variations in the Indian Ocean along Wijffels, S., Sprintall, J., Fieux, M., Bray, N., 2002. The JADE and WOCE
1101E. I. Hydrological structure of the upper 500 m. Australian I10/IR6 throughflow sections in the southeast Indian Ocean. Part 1:
Journal of Marine and Freshwater Research 20, 1–50. water mass distribution and variability. Deep-Sea Research Part II:
Rochford, D.J., 1969b. Seasonal interchange of high and low salinity Topical Studies in Oceanography 49, 1341–1362.
surface waters off south-west Australia. Technical Paper. Division of Wong, A.P.S., 2005. Subantarctic mode water and Antarctic intermediate
Fisheries and Oceanography. CSIRO Australia, no. 29. water in the south Indian Ocean based on profiling float data
Sallée, J-B., Wienders, N., Speer, K., Morrow, R., 2006. Formation of 2000–2004. Journal of Marine Research 63, 789–812.
subantarctic mode water in the southeastern Indian Ocean. Ocean Woo, M., Pattiaratchi, C., Schroeder, W., 2006a. Dynamics of the Ningaloo
Dynamics 56, 525–542. current off point Cloates, Western Australia. Marine and Freshwater
Sharma, G.S., 1972. Water characteristics at 200 cl/t in the intertropical Research 57, 291–301.
Indian Ocean during the southwest monsoon. Journal of Marine Woo, M., Pattiaratchi, C., Schroeder, W., 2006b. Summer surface
Research 30, 102–111. circulation along the Gascoyne continental shelf, Western Australia.
Smith, R.L., Huyer, A., Godfrey, J.S., Church, J.A., 1991. The Leeuwin current Continental Shelf Research 26, 132–152.
off Western Australia, 1986–1987. Journal of Physical Oceanography Wyrtki, K., 1971. Oceanographic Atlas of the International Indian Ocean
21, 323–345. Expedition. National Science Foundation, Washington, DC.
Sverdrup, H.U., Johnson, M.W., Fleming, R.H., 1942. The Oceans: Their Wyrtki, K., 1973. Physical oceanography of the Indian Ocean. In:
Physics, Chemistry, and General Biology. Prentice-Hall, Englewood Zeitzschel, B. (Ed.), Ecological Studies: Analysis and Synthesis, vol.
Cliffs NJ. 3. Springer, Berlin, pp. 18–36.

You might also like