ENSO and short-term variability of the South Equatorial Current

advertisement
1
1
ENSO and short-term variability of the
2
South Equatorial Current entering the Coral Sea
3
To be submitted to the Journal of Physical Oceanography
4
5
William S. Kessler
6
NOAA Pacific Marine Environmental Laboratory, Seattle WA USA
7
8
Sophie Cravatte
9
Institut de Recherche pour le Développement, Nouméa, New Caledonia
10
11
Abstract
12
Historical section data extending to 1985 are used to estimate the interannual variability
13
of transport entering the Coral Sea between New Caledonia and the Solomon Islands.
14
Typical magnitudes of this variability are ±5-8 Sv on a mean of about 30 Sv. Transport
15
increases a few months after an El Niño event, and decreases following a La Niña. Most
16
of the interannual transport variability is well-simulated by a reduced-gravity long
17
Rossby wave model. Vigorous westward-propagating mesoscale eddies can yield
18
substantial aliasing (under- or over-estimates comparable to the mean transport) on
19
individual ship or glider surveys. Since the transport variability is surface-intensified and
20
highly correlated with satellite-derived surface geostrophic currents, a simple index of
21
South Equatorial Current transport based on satellite altimetry is developed.
22
2
23
1. Introduction
24
The South Equatorial Current carries the westward limb of the South Pacific
25
subtropical gyre to the Coral Sea where it bifurcates into poleward and equatorward
26
branches (REFs). It has been long known, and is expected from simple dynamical ideas
27
(REFs), that the incoming transport varies annually (Kessler and Gourdeau 20NN) and
28
with the ENSO cycle (REFs), but developing time series of this transport variability has
29
been difficult because much of the historical data consists of occasional ship surveys that
30
are hard to put in context. With the advent of the Argo program (REF), this problem will
31
be rectified for the future, but Argo sampling of the Coral Sea became sufficiently dense
32
only in 2005 or 2006. The purpose of this paper is to make use of older data to extend the
33
transport time series back before the Argo era. We would also like to develop indices that
34
could allow satellite altimetry to be used as a proxy for interannual transport variations;
35
the [rapid and regular] satellite sampling gives the opportunity to evaluate high-frequency
36
signals that could alias ship surveys.
37
Several estimates of the transport entering the Coral Sea based on ship surveys
38
have been published, and these vary by about a factor of two (REFs), confusing the
39
interpretation of these data. There are several factors that cause such estimates to differ:
40
sampling locations (especially obtaining adequate resolution of the narrow zonal jets and
41
boundary currents), sampling depths, choices of reference level, timing relative to the
42
seasonal or ENSO cycles, and short-term eddy aliasing. By extending the existing time
43
series back before Argo, and making use of weekly satellite altimetry, we hope to
44
evaluate these estimates.
3
45
46
2. Data and methods
Three sources of geostrophic current information provide time series of the transport
47
entering the Coral Sea: First, the SIO/ORSTOM Volunteer Observing Ship XBT program
48
(REF Donguy-Meyers) deployed XBTs along a regularly-repeated merchant ship track
49
from Auckland, New Zealand to Solomon Strait, passing just west of New Caledonia
50
(Fig.AAA), from 1985 through 2001. 3917 temperature profiles, usually to 400m, were
51
made on this track and quality controlled by the Coriolis Data Centre; only profiles with
52
quality flags indicating good data were considered. 1346 of these profiles were within the
53
20°S-11°S region considered here, for an average of about 1.5 profiles per degree latitude
54
per month. For each XBT temperature value, a mean T-S relation from the CARS gridded
55
compilation (Dunn and Ridgway 2002; http://www.cmar.csiro.au/cars) was used to
56
associate a salinity. The resulting temperatures and salinities were interpolated to regular
57
5m depth intervals, then gridded in (y,t) using a Gaussian mapping procedure to a 0.5°
58
latitude by 3-month grid, using mapping scales 0.75° latitude and 3 months. Dynamic
59
height relative to 400m was found from these, and crosstrack (perpendicular,
60
approximately southwestward) geostrophic velocity estimated by centered differencing.
61
Second, the Argo Atlas (Roemmich and Gilson 2009) provides gridded temperature
62
and salinity profiles to 2000m on a global 1° by 1° monthly grid. The Argo Atlas fields
63
were sampled along the XBT track at 1° latitude intervals to produce a time series
64
comparable to the XBT data from 2004 through 2011. Crosstrack and zonal geostrophic
65
currents were found relative to 2000m or the deepest common level.
4
66
Third, weekly AVISO satellite altimetric sea surface height (SSH) data (REF) on a
67
roughly 1/3° grid from August 1992 through 2011, was sampled along the same track and
68
used to find surface geostrophic velocity anomalies.
69
In all three cases, anomalies were found as differences from the average annual cycle
70
over the record length of each time series, then filtered with a 7-month triangle. This
71
filter has a half-power point of about 8 months, and passes more than 80% of the
72
variability with timescales longer than 13 months; the result is here referred to as
73
"interannual".
74
The South Equatorial Current (SEC) can have ambiguous and fluctuating boundaries
75
(REFs). In this work, however, we are interested in anomalies to the flow entering the
76
Coral Sea, not to a specifically-defined SEC. Thus we have chosen a simple metric:
77
transport perpendicular to the XBT line between 20°S and 11°S, which approximates the
78
transport entering the Coral Sea between New Caledonia and the Solomon Islands
79
(Fig.1). The anomalous transports are not sensitive to the actual boundaries of the current
80
at a given time because interannual dynamic height anomalies are highly correlated over
81
latitude in this range.
82
83
3. Results
84
3.1 Mean transports and representativeness of shallow transport anomalies
85
Although here we are concerned with anomalies, especially since we consider
86
disparate measures of flow, one motivation is to place historical cruise data into its ENSO
87
context, so total transport values are relevant. Additional benefits of doing this are to
88
show what is and is not sampled by the 400m XBT profiles, to compare perpendicular
5
89
transport across the slanted XBT track (Fig.1) to meridional transport sections, and to
90
evaluate the magnitudes of ENSO and other anomalies relative to mean transport values.
91
Crosstrack speed across the XBT track (Fig.BBBa) shows two well-known westward
92
branches of the SEC: the narrow North Caledonian Jet (NCJ; Gourdeau et al. 2007;
93
REFs) at 19°S, with a mean speed relative to 400m of about 1 cm s–1 and transport of
94
about 0.5 Sv, and the North Vanuatu Jet (NVJ; REFs) from about 15°S to 9.5°S with
95
speeds to about 6 cm s–1 and transport of about 7 Sv. Between them the shallow eastward
96
Coral Sea Countercurrent (CSCC; Qiu et al. 2010) is centered at 15.5°S. Comparable
97
mean crosstrack speed relative to 400m from the Argo Atlas is very similar, with slightly
98
weaker velocities (Fig.BBBb).
99
Relative to the 2000m reference level available from the Argo Atlas data, the NVJ is
100
found to be stronger, with a similar structure, but the NCJ extends much deeper and its
101
speeds are at least five times larger (Fig.BBBc), as has been previously described
102
(Gourdeau et al. 2007). The CSCC is weaker and narrower. NCJ transport relative to
103
2000m from the Argo Atlas is about 11 Sv, and NVJ transport about 15 Sv. Thus the
104
relative-to-400m transports omit most of the mean transport of the NCJ, and about half
105
the mean transport of the NVJ.
106
The Argo Atlas provides a further check on the representativeness of transport
107
variability relative to 400m. Time series of interannual SEC transport relative to 400m
108
and 1000m are very highly correlated with that relative to 2000m (r = 0.98 and 0.93
109
respectively) . The magnitude of transport variance relative to 1000m is about 46% larger
110
than to 400m, and that relative to 2000m is about 54% larger. Regression of the 0/400m
111
onto 0/2000m transport gives a slope of 0.61 and intercept of 0.02 Sv (essentially zero).
6
112
The high correlations are consistent with the idea that ENSO variability is focused in the
113
thermocline, and suggests that the XBT 400m transport gives a good representation of the
114
temporal variance pattern, but the interannual magnitude of 20°S-10°S transport
115
variability is larger by about 50-60% than measured by the XBT line.
116
The representativeness of perpendicular velocities across the slanted section can be
117
checked using a meridional section at 160°E from the Argo Atlas. Relative to 2000m, the
118
three currents on the purely meridional section have an almost identical structure to that
119
on the track-sampled field, with slightly stronger velocities and total (20°S-10°S) zonal
120
transport of 28 Sv (Fig.BBBd). This correspondence is expected since the vector flow is
121
nearly zonal so the crosstrack velocity component gives the same integral. A similar
122
calculation at 160°E from the CARS climatology relative to 2000m gives a transport of
123
just under 30 Sv.
124
125
126
3.2 Internannual transport anomalies
Unfortunately the XBT and Argo time series do not overlap so no direct comparison
127
can be made. We note that the 1985-2001 period covered by the XBT data included the
128
strong El Niño events of 1986-87, 1991-92, 1994-95 and 1997-98, as well as the strong
129
La Niña of 1988, while the Argo time series has been made during a period of generally
130
weaker El Niño events. Two other ENSO-relevant time series span these in situ data
131
sources: The Southern Oscillation Index (SOI;
132
http://www.cpc.ncep.noaa.gov/data/indices/soi), and crosstrack surface geostrophic
133
currents derived from AVISO satellite SSH (Fig.CCC). All the estimates of the SEC are
134
well-correlated with the SOI with a few-month lag consistent with Rossby wave
7
135
propagation from the mid-basin focus of ENSO wind variability (REFs). The 16-year
136
XBT transport correlates with the SOI at 0.88 with a 4-month lag, the 7-year Argo
137
transport at 0.67 with a 2-month lag, and the 19-year AVISO surface velocity at 0.75 with
138
a 3-month lag. The shorter recent lag times could reflect the change from the east-Pacific-
139
centered El Niños during the 1980s and 1990s sampled by the XBT and AVISO data, to
140
the late 2000s sampled by Argo when central Pacific “Modoki” El Niños (Ashok et al.
141
2007) produced wind anomalies further west.
142
The ENSO signal in crosstrack velocity is clearly seen in examples of El Niños and
143
La Niñas on the track. Anomalies of the zonal flow were strongest at the surface and
144
toward the north end of the New Caledonia-Solomons gap (Fig.DDD). Near-surface NVJ
145
speeds increased by about 50% following the El Niño of 2009-10, and decreased about
146
the same following the La Niña of 2008-09, while NCJ changes were much smaller
147
(Fig.DDD). Very similar anomalies were seen during the other ENSO events, both in the
148
XBT and Argo data. This weakening of interannual variance with depth is similar to that
149
seen in glider data across the mouth of the Solomon Sea (Davis et al. 2012).
150
Since ENSO anomalies are surface-intensified, it is not surprising that AVISO surface
151
geostrophic velocities are highly correlated with transport anomalies: the correlation of
152
AVISO speeds (averaged along the XBT line from 20°S to 11°S) with XBT transport is
153
0.93, and with Argo transport is 0.94, as is obvious from Fig.CCC. Thus AVISO could be
154
used as a proxy for interannual transport anomalies, with the factor of about 1.3 Sv
155
(transport relative to 400m) per cm/s (surface speed) given by the regression slope. To
156
approximate Argo transport variability relative to 2000m, the factor would be about 2.1.
8
157
Published estimates of the transport at or near 160°E from synoptic sampling were
158
coincidentally taken at times when the ENSO cycle was nearly neutral, and thus do not
159
provide a useful test of the SEC response to ENSO. Scully-Power (1973) found 30 Sv
160
from a cruise at 162°E in Aug-Sep 1970, while Gourdeau et al. (2008) found 32 Sv
161
absolute transport above 600m from glider data and 28 Sv from cruise data based on
162
geostrophy relative to 2000m, both in Aug-Oct 2005. These values are all
163
indistinguishable from the mean transport from the Argo Atlas and CARS climatological
164
means (28 and 30 Sv respectively). A somewhat larger estimate, again from Scully-
165
Power (1973) based on data taken during mid-1968 (very weak La Niña conditions), was
166
37 Sv.
167
There is one outlier to these values: Sokolov and Rintoul’s (2000) SEC transport
168
estimate of 54 Sv westward, relative to 2000db, from the WOCE P11S cruise data taken
169
during June-July 1993, a value well above any others noted before or since. This cruise
170
took place following the relatively weak El Niño of 1993, and the AVISO proxy index
171
suggests that westward transport would have therefore been about 6-7 Sv greater than
172
normal during this period; not nearly enough to account for the large P11S value. Since
173
this apparently cannot be explained based on ENSO conditions, we use the weekly
174
AVISO data to ask if higher-frequency variability could have aliased the sampling.
175
176
3.3 An example of eddy aliasing?
177
The standard deviation of high-passed AVISO crosstrack velocity averaged along the
178
XBT section is about 1.4 cm/s, which is about 60% the size of the interannual RMS, and
179
about 3 times larger than the annual cycle RMS. Inspection of weekly maps shows that
9
180
much of this variability is due to westward-moving mesoscale eddies with typical
181
diameters of 100-300km and anomalous surface geostrophic velocities of 20 cm/s or
182
more (e.g. Fig.EEE). These features translate west across the Coral Sea with speeds
183
decreasing from about 17 cm/s at 12°S to 12 cm/s at 20°S; that is, with somewhat smaller
184
meridional speed gradient than would be expected from linear Rossby wave propagation
185
(Chelton; Maharaj et al. 2009; REF). At these speeds, an individual eddy affects a section
186
line for 15-25 days. However, most of the time such an eddy will produce oppositely-
187
directed velocity anomalies on its flanks that do not result in a correspondingly-large
188
transport anomaly when integrated over a section as we consider here.
189
Nevertheless, during the WOCE P11S cruise in June-July 1993, such mesoscale
190
eddies lined up so as to produce eastward surface velocity anomalies within 100km of the
191
coast of Papua New Guinea, where the mean flow is the eastward boundary current (the
192
Gulf of Papua Current; GPC), and westward anomalies across the background SEC from
193
13°S to 19°S, thus temporarily augmenting both mean currents. In the GPC, geostrophic
194
crosstrack surface currents relative to 2000db measured by P11S were about 50 cm/s, but
195
the short-term eddy anomalies seen by AVISO at this time were about 20 cm/s. Across
196
the SEC, P11S surface currents averaged about 13 cm/s (westward), while AVISO
197
anomalies were about 8 cm/s, thus more than half the westward surface velocity
198
measured during P11S was transient. We note that if the cruise had been conducted 2
199
months earlier, the AVISO anomalies along the P11S section would have had the
200
opposite sign, emphasizing the short-term nature of these eddy signals.
201
It is impossible to determine the depth extent of the eddy anomalies, but it is
202
reasonable to think that they represent fluctuations of the upper thermocline, and an
10
203
exponential decay scale of 400m would be a conservative estimate of the eddy vertical
204
structure. If that is the case, the SEC eddy anomalies on the P11S section amount to about
205
23 Sv of transient westward transport. Thus, the background SEC seen in the P11S cruise
206
could then be estimated at about 31 Sv, a value consistent with other samples and
207
climatologies. This has an important implication for the redistribution of incoming flow
208
from the SEC. Sokolov and Rintoul (2000) noted that the 54 Sv SEC transport across
209
P11S implied that about 29 Sv of the arriving westward flow turns south into the East
210
Australian Current. If the typical transport is instead only about 30 Sv, then the required
211
southward flow would be small, and the bulk of the SEC would be found to turn north
212
into the Solomon Sea, as most other studies have found (REFs).
213
214
215
4. A simple Rossby wave model
The simplest representation of low-frequency wind-driven dynamics is the linear,
216
reduced-gravity long Rossby wave model as has been frequently used to study tropical
217
ocean variability (Meyers 1979; Kessler 1990; Chen and Qiu 2004). The model is:
218
  
h
h
 cr
 Rh  Curl ,
t
x
f 
(1)
219
where h is the pycnocline depth anomaly (positive down). The long Rossby speed is
220
cr=c2/f 2, (c is the internal long gravity wave speed, f is the Coriolis parameter and  its

221
meridional derivative), R is a damping timescale and  is the wind stress (data sources are
222
given below). The two parameters to be chosen are c and R, which we take to be
223
c = 2.5 m s–1 and R = (24 months)–1. In this case, the results are only weakly sensitive to
224
these choices because the largest interannual forcing is in the western half of the basin
11
225
(Fig.FFF, bottom panels), so differences in propagation have little time to manifest
226
themselves. We ignore signals that might emanate from the eastern boundary (Kessler
227
1990; Minobe and Takeuchi 1995; Vega et al. 2003; Fu and Qiu 2002), so solutions
228
represent purely the linear response to interior-ocean wind stress curl.
229
The model was forced with a sequence of interannually-filtered monthly wind
230
stresses: from January 1981 to July 1991 with ship-derived winds from the Center for
231
Ocean-Atmospheric Predictions Studies at Florida State University (Bourassa et al.
232
2005), then until July 1999 with the scatterometer winds from ERS 1 and 2 (European
233
Space Agency Remote Sensing) satellites, then until November 2009 with QuikSCAT
234
winds [REF]. We ignored the effects of islands: the stresses  were linearly interpolated
235
in x across all the islands east of Australia-New Guinea, then we did not consider any
236
effects of island blocking (Island Rule; Firing et al. 1999).
237
Geostrophic zonal transports analogous to the XBT and Argo estimates (Fig.3) were
238
found from the solution h(x,y,t) of (1), using U=(c2/f )dh/dy, then similarly integrated
239
from 20°S to 11°S.
240
The Rossby transport timeseries is very clearly correlated with the 16-year XBT
241
transport (r=0.86), but does less well during the 5-year period when the winds and Argo
242
data overlap (r=0.53; Fig.FFFa). We note that Argo sampling of the southwest Pacific
243
was sparse until at least mid-2005, when the Rossby and Argo transport time series
244
appear dissimilar, which perhaps contributes to the weaker overall relation. The
245
magnitude of the Rossby-predicted transport anomalies is comparable to the observed as
246
well (Fig.FFFa, where the RMS of XBT anomalies is 4.2 Sv, of Argo 3.1 Sv, and of the
247
Rossby model 5.1 Sv). As with the XBT, Argo and Aviso time series, the Rossby
12
248
solution has a strong ENSO signal, with a lag relation to the SOI similar to what was seen
249
for the XBT and Argo transport (section 3.2; Fig.FFFa,b).
250
The genesis of the ENSO transport signals can be seen by comparing average wind
251
stress and curl over peaks of the ENSO cycle: defining La Niña as those times when the
252
SOI was greater than 1, and El Niño when it was less than 1 (blue and red highlighting
253
in Fig.FFFb). Because ENSO changes character after 1999, this means that El Niño
254
mainly samples the period before 1999, while La Niña samples the time after, so this
255
method does not distinguish between interannual and longer-timescale variations.
256
La Niña periods have easterly anomalies spanning the equator to about 10°S west of
257
about 150°W, then northwesterlies south of about 15°S. Downwelling curl (red shades in
258
Fig.FFFc,d) north of 15°S deepens the thermocline there and to the west, while upwelling
259
curl from 15°-25°S shoals it; the result is an anomalous tilt down towards the equator and
260
consequent weakening of the SEC. Similar but opposite-sign anomalies occur during El
261
Niño, with the entire pattern shifted a few degrees south (Fig.FFFd; Harrison and Vecchi
262
1999). The curl patterns south of 20°S suggest that a second reversal of anomalies occurs
263
at 25°-30°S, with anomalous westward flow during La Niña and eastward during El Niño
264
(Fig.FFFc,d); thus La Niña can be seen as shifting the subtropical gyre to the south and El
265
Niño to the north (Wyrtki and Wenzel 1984).
266
267
268
5. Conclusion
Historical section data extending back to 1985 are used to estimate the interannual
269
variability of transport entering the Coral Sea between New Caledonia and the Solomon
270
Islands. Typical magnitudes of this variability are ±5-8 Sv on a mean of about 30 Sv.
13
271
Westward transport increases a few months after an El Niño event, and decreases
272
following a La Niña. A linear, reduced-gravity Rossby model reproduces the observed
273
transport variability well, especially during the period before 2001, and shows that much
274
of the ENSO transport anomalies are due to wind changes extending south to at least
275
20°S. These fluctuations are consistent with ENSO transport anomalies measured across
276
the mouth of the Solomon Sea by Davis et al. (2012, submitted). Interannual fluctuations
277
are about 50% larger than the annual cycle transport variability on a similar section
278
estimated from the CARS climatology by Kessler and Gourdeau (2007; their Fig.7).
279
Vigorous westward-propagating mesoscale eddies produce strong transient current
280
fluctuations on scales of 100-300km, but usually these average out over a long section as
281
considered here. Occasionally these eddies transiently be aligned to yield substantial
282
aliasing (under- or over-estimates comparable to the mean transport) on individual ship or
283
glider surveys.
284
Since the transport is highly correlated with AVISO surface geostrophic currents, and
285
its variability is surface-intensified, a simple index of SEC transport based on AVISO
286
currents predicts its anomalies to be about 1.3 Sv (transport relative to 400m) per cm/s
287
(section-average surface speed). A factor of 2.1 Sv per cm/s estimates the total (relative
288
to 2000m) transport.
289
290
Acknowledgements
291
This paper was possible because of the data collection, compilation and quality control
292
efforts of several international programs. The quality-controlled XBT profiles were made
293
available by the Coriolis Data Centre of IFREMER (http://www.coriolis.eu.org). The
14
294
Argo profiling float data were collected and made freely available by the International
295
Argo Program and the national programs that contribute to it (http://www.argo.ucsd.edu).
296
The Aviso altimeter products were produced by Ssalto/Duacs and distributed by Aviso,
297
with support from CNES (http://www.aviso.oceanobs.com/duacs/). Winds: COAPS/FSU,
298
ERS, Quikscat.
299
300
References
301
Ashok, K, S.K. Behara, S.A. Rao, H. Weng and T. Yamagata, 2007: El Niño Modoki and
302
its possible teleconnection. J.Geophys.Res., 112,C11007,doi:10.1029/2006JC003798.
303
304
305
306
307
308
Bourassa, M. A., R. Romero, S. R. Smith, and J, J. O'Brien, 2005: A new FSU winds
climatology, J. Clim., 18, 3692-3704.
Chelton, D.B., M.G. Schlax and R.M. Samelson, 2011: Global observations of nonlinear
mesoscale eddies. Prog.Oceanogr., 91(2), 167-216.
Chen, S. and B. Qiu, 2004: Seasonal variability of the South Equatorial Countercurrent.
J.Geophys.Res., 109, C08003, doi:10.1029/2003JC002243.
309
Davis, R. E., W. S. Kessler and J. Sherman, 2012:
310
Donguy and Meyers (XBT program)
311
Dunn J. R., and K. R. Ridgway, 2002: Mapping ocean properties in regions of complex
312
topography, Deep Sea Research I: Oceanographic Research, 49(3), 591-604
313
Firing, E., Qiu, B. and W. Miao, 1999: Time-dependent Island Rule and its application to
314
the time-varying North Hawaiian Ridge Current. J. Phys. Oceanogr., 29, 2671-2688.
15
315
Fu, L.-L. and B. Qiu, 2002: Low-frequency variability of the North Pacific Ocean: The
316
roles of boundary- and wind-driven baroclinic Rossby waves. J.Geophys.Res.,
317
107(C12), 3220, doi:10.1029/2001JC001131.
318
319
320
321
322
323
324
325
326
Gourdeau. L., W.S. Kessler, R.E. Davis, J.Sherman, C. Maes and E. Kesternare, 2008:
Zonal jets entering the Coral Sea. J.Phys.Oceanogr., 38(3), 715-725.
Harrison, D.E. and G.A. Vecchi, 1999: On the termination of El Niño.
Geophys.Res.Lett., 26(11), 1593-16596.
Kessler, W.S., 1990: Observations of long Rossby waves in the northern tropical Pacific.
J.Geophys.Res., 95(C4), 5183-5217.
Kessler, W.S. and L. Gourdeau, 2007: The annual cycle of circulation in the southwest
subtropical Pacific, analyzed in an ocean GCM. J.Phys.Oceanogr., 37(6), 1610-1627.
Maharaj, A.M., N.J. Holbrook and P. Cipollini, 2009: Multiple westward-propagating
327
signals in South Pacific sea level anomalies. J.Geophys.Res.-O., 114, C12016, doi:
328
10.1029/2008JC004799.
329
330
331
332
333
Meyers, G., 1979: The annual Rossby waves in the tropical North Pacific Ocean.
J.Phys.Oceanogr., 9(4), 663-674.
Minobe, S. and K. Takeuchi, 1995: Annual period equatorial waves in the Pacific Ocean.
J.Geophys.res.-O., 100(C9), 18379-18392.
Roemmich, D. and J. Gilson, 2009: The 2004-2008 mean and annual cycle of
334
temperature, salinity and steric height in the global ocean from the Argo program.
335
Prog.Oceanogr., 82(2), 81-100.
336
337
Scully-Power, P.D., 1973. Coral Sea flow budgets during winter. Aust. J. Mar.
Freshwater Res., 24, 203-215.
16
338
Sokolov, S., and Rintoul, S.R., 2000: Circulation and water masses of the southwest
339
Pacific: WOCE section P11, Papua New Guinea to Tasmania. J.Mar.Res., 58, 223-
340
268.
341
Vega, A., Y. du-Penhoat, B. Dewitte and O. Pizarro, 2003: Equatorial forcing of
342
interannual Rossby waves in the eastern South Pacific. Geophys.Res.Lett., 30(5),
343
1197, doi: 10.1029/2002GL015886.
344
345
346
347
Wyrtki, K. and J. Wenzel, 1984: Possible gyre-gyre interaction in the Pacific Ocean.
Nature, 309, 538-540.
17
348
Figure captions
349
Figure 1. IFREMER XBT profiles in the southwest Pacific, showing the selected track
350
from Auckland to Solomon Strait.
351
Figure 2. Mean crosstrack geostrophic speed on the XBT track. Positive values (red)
352
cross the track to the east. a) XBT, b) Argo 400m, c) Argo 2000m, d) Meridional section
353
at 160°E .
354
Figure 3. Time series of interannual anomalies on the XBT track between 20°S and 11°S
355
from the XBT data (green, Sv), Argo data (red, Sv), Aviso surface speed (black, cm s–1).
356
The SOI is overlaid in blue dots (scale at right).
357
Figure 4. Example of crosstrack currents during mean (top), El Niño (middle) and La
358
Niña (bottom) conditions. Positive values (red) cross the track to the east.
359
Figure 5. AVISO sea level anomalies (colors) and surface geostrophic velocity during
360
the WOCE P11S cruise in June-July 1993. The P11S shiptrack is shown in green, roughly
361
along 154.5°E.
362
Figure 6. a) Time series of transport anomalies from the Rossby model (black), XBT
363
data (green) and Argo (red). The XBT and Argo time series are the same as in Fig.3.
364
b) Southern Oscillation Index, where red and blue shading indicates El Niño (SOI less
365
than −1) and La Niña (SOI greater than +1) respectively. c) Mean wind stress and curl for
366
the La Niña periods shown in panel (b). Units given at bottom. d) As panel c but for El
367
Niño periods.
Download