ITF VAR Pujiana 2013

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

JOURNAL OF GEOPHYSICAL RESEARCH: OCEANS, VOL. 118, 2023–2034, doi:10.1002/jgrc.

20069, 2013

Intraseasonal Kelvin wave in Makassar Strait


K. Pujiana,1,3 A. L. Gordon,1 and J. Sprintall2
Received 28 May 2012; revised 17 December 2012; accepted 19 December 2012; published 18 April 2013.

[1] Time series observations during 2004–2006 reveal the presence of 60–90 days
intraseasonal events that impact the transport and mixing environment within Makassar
Strait. The observed velocity and temperature fluctuations within the pycnocline reveal the
presence of Kelvin waves including vertical energy propagation, energy equipartition, and
nondispersive relationship. Two current meters at 750 and 1500 m provide further evidence
that the vertical structure of the downwelling Kelvin wave resembles that of the second
baroclinic wave mode. The Kelvin waves derive their energy from the equatorial Indian
Ocean winds, including those associated with the Madden-Julian oscillations, and
propagate from Lombok Strait to Makassar Strait along the 100-m isobath. The northward
propagating Kelvin waves within the pycnocline reduce the southward Makassar Strait
throughflow by up to 2 Sv and induce a marked increase of vertical diffusivity.
Citation: Pujiana, K., A. L. Gordon, and J. Sprintall (2013), Intraseasonal Kelvin wave in Makassar Strait, J. Geophys.
Res. Oceans, 118, 2023–2034, doi:10.1002/jgrc.20069.

1. Introduction the Indo-Pacific region and proposed that the equatorial


Indian Ocean wind-forced CTKWs also accounted for intra-
[2] The Makassar Strait (Figure 1) throughflow is a seasonal variability in the ITF passages. These modeling
prominent component of the Indonesian Throughflow studies emphasized the ITF reversal event as an indicator of
(ITF) [Gordon and Fine, 1996; Gordon et al., 2003; Gordon, the CTKW passage in Makassar Strait but did not elaborate
2005]. Makassar Strait is not only the main inflow conduit for on the mechanism of the CTKW transmission from Lombok
the ITF [Gordon et al., 2008, 2010] but is also part of a wave- Strait to Makassar Strait.
guide extending along the equatorial Indian Ocean to the [3] Observations during December 1996 to June 1998
southwestern coasts of Indonesia then to Makassar Strait, from two current meters (Figure 1) at depths of 250 and
via Lombok Strait (Figure 1) [Wijffels and Meyers, 2004].
350 m in Makassar Strait revealed that the subinertial flow
Previous modeling studies have discussed the role of the
was marked with northward (positive) along-strait flow in
waveguide in connecting wind energy in the equatorial
May 1997, during which strong intraseasonal northward
Indian Ocean to the Makassar Strait throughflow variability
flow completely reversed the southward lower-frequency
with timescales varying from semiannual to intraseasonal
background flow (Figure 2). In addition to the May–June
[Qiu et al., 1999; Sprintall et al., 2000; Schiller et al.,
1997 event, intraseasonal variability’s most pronounced
2010]. The study of Sprintall et al. [2000] argued that semi- impact to attenuate the southward Makassar Strait through-
annual variations (180 days) in the equatorial Indian Ocean flow was observed in December 1996, September, October,
zonal winds induced a response in the South Java Current and December 1997, and January, February, March, and
through coastally trapped Kelvin waves (CTKWs). The April 1998 (Figure 2).
semiannual CTKWs propagate farther east along the [4] Analyzing along-strait flow data within the pycnocline
southern coasts of the Indonesian archipelago into the Savu in Makassar Strait and Lombok Strait observed during the
Sea, but part of the wave energy is transmitted through the International Nusantara Stratification and Transport (IN-
narrow Lombok Strait, which then move equatorward to STANT) 2004–2006 program [Gordon et al., 2008; Gordon
force the northward Makassar Strait throughflow. At intra- et al., 2010; Sprintall et al., 2009] and zonal current data from
seasonal timescales (20–90 days), Qiu et al. [1999] and a mooring in the eastern equatorial Indian Ocean, Pujiana et al.
Schiller et al. [2010] modeled the intraseasonal variation in [2009] found a coherent signal at intraseasonal timescales prop-
1
agating from the equatorial Indian Ocean to Makassar Strait
Lamont-Doherty Earth Observatory, Columbia University, Palisades, through Lombok Strait with the speed consistent with that of a
New York, USA.
2
Scripps Institution of Oceanography, University of California San baroclinic wave. Drushka et al. [2010], focusing their analyses
Diego, La Jolla, California, USA. on the vertical structure of along-strait flow at the outflow pas-
3
Faculty of Earth Sciences and Technology, Bandung Institute of sages of the ITF, indicated that the intraseasonal along-strait
Technology, Bandung, Indonesia. motions at Lombok Strait had a vertical structure of a remotely
Corresponding author: K. Pujiana, Lamont-Doherty Earth Observatory, wind-forced Kelvin wave.
Columbia University, 61 Route 9W, Palisades, NY 10964, USA. [5] In this study, we utilize the 2004–2006 Makassar
(kandaga@ldeo.columbia.edu) Strait dataset to investigate intraseasonal features consistent
©2013. American Geophysical Union. All Rights Reserved. with the theoretical Kelvin wave characteristics, specifically
2169-9275/13/10.1002/jgrc.20069 those of the downwelling Kelvin waves forcing the

2023
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

90oE 105oE 120oE 135oE


o
12
(a)
South-China Pacific Ocean
Sea
ITF
Sulawesi
Sea -8o
Pn

SU

t
trai
KALIMANTAN

M
o
0

AT
Pw

rS
Equatorial BALI Pe

R
SULAWESI

ssa
Kelvin wave C Lw Le
o
Ke ast

it
LOMBOK

Stra
ka
lv ally Java Sea

Ma
in -

bok
Indian Ocean w Tra
av p

Lom
JAVA o
e pe -9
d Ps
Savu Sea
o o
115.0 E 116.5 E
-12o
A B
-50
(c) (b)
Mak-West Mak-East
-500 Mak-West Mak-East
+ + +
A B
Depths(m)

o
-1000 -3
19 km
Labani Channel
-1500
-2000
-500
-1500
-2000
0 2 -2 2
: ADCP
N (s ) : CM +
:T&S +
-2500
o
118.7oE
o
118.3 E 118.5 E

Figure 1. (a) The Makassar Strait location in the maritime continent and the schematic of the waveguide
for the Kelvin wave. Inset is an expanded view of Lombok Strait and the corresponding moorings
(Lw and Le) and shallow pressure gauge sites (Pw and Pe) and location of altimeter measurement
used for along-strait sea level anomaly (Pn and Ps). (b) A blow-up of the blue box in Figure 1a
showing the Labani Channel, the narrowest passage within Makassar Strait where the moorings
(Mak-West and Mak-East) were deployed. Yellow crosses denote CTD stations. (c) A bathymetric cross
section A-B (shown in Figure 1b) in the Labani Channel, mooring configuration, and an averaged
stratification profile obtained from the CTD stations.

northward along-strait flow events, and evaluate the likely


0.50
pathway channeling the intraseasonal energy from Lombok
0.25
to Makassar Straits. Moreover, the Kelvin wave’s impacts
v (m/s)

0
-0.25
on the ITF variability and mixing environment in Makassar
-0.50
Strait are addressed.
-0.75
[6] The paper is organized as follows. The data description
0.50 is given in section 2, followed in section 3 with a discussion of
0.25 the evidence supporting the intraseasonal Kelvin wave
variability. The waveguide pathway from Lombok Strait to
v (m/s)

0
-0.25 Makassar Strait is discussed in section 4. The Kelvin wave’s
-0.50 influence on the Makassar throughflow transport and mixing
-0.75 is evaluated in section 5. A discussion and summary follow
D J F M A M J J A S O N D J F M A M J
in section 6.
1997 1998

Figure 2. Time series of subinertial (dashed line) and intra- 2. Data


seasonal (solid line) along-strait flow at 250 m (upper panel)
and 350 m (lower panel) at the Mak-West site. Positive 2.1. Makassar Strait Mooring in situ Data
velocity indicates northward flow. The data were observed [7] We employ the velocity and temperature datasets
during late November 1996 to mid-June 1998. obtained during the INSTANT 2004–2006 program to

2024
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

identify intraseasonal Kelvin wave events in Makassar variability is less well resolved at the Mak-East mooring, we
Strait. The INSTANT program monitored the ITF variability will only analyze the temperature profile dataset from the
in Makassar Strait at two moorings: 2 51.90 S, 118 27.30 E Mak-West. The sensors sampled temperature and pressure at
[Mak-West] and 2 51.50 S, 118 37.70 E [Mak-East], within 6-min intervals over the entire 3-year INSTANT deployment
the 45-km wide Labani Channel [Gordon et al., 2008] period. The temperature datasets are linearly interpolated onto
(Figure 1). An upward-looking RD Instruments Long a 25-m depth grid for each 2-hour time step to provide the
Ranger 75-kHz Acoustic Doppler Current Profiler (ADCP) gridded temperature data from 150 to 350 m of water column.
at a depth of 300 m and three current meters at 200, 400,
and 750 m on each mooring (the Mak-West mooring had 2.2. Shallow-pressure Gauge and Satellite-derived Data
an additional current meter at 1500 m) recorded the velocity [9] In addition to the moored data in Makassar and Lombok
data with a sampling period of 0.5–2 h. Both moorings fully Straits, we also employ the sea level anomaly (SLA) data from
resolved the vertical structure of the horizontal velocity across two shallow pressure gauges (Pw and Pe) in Lombok Strait
the pycnocline depths of 50–450 m almost continuously (Figure 1) and from the gridded SLA products (merged and
(other than a short gap between recovery and redeploy- delayed time products) of Archiving, Validation, and Interpre-
ment on 7–10 July 2005) for a ~3-year-long period from tation of Satellite Oceanographic Data (AVISO) [Ducet et al.,
January 2004 to November 2006. The horizontal current 2000]. Here, the daily shallow pressure gauge data are used,
vectors are subsequently projected to the along (y) and and more information on the pressure gauge setting can be
across-strait (x) axis of the Labani Channel, which are found in Sprintall et al. [2004]. The satellite-derived SLA
oriented along 10 and 80 (relative to north and posi- has a horizontal resolution of 0.25  0.25 and temporal
tive is clockwise) respectively (Figure 1), to yield gridded resolution of 7 days, and the emphasis of our analysis will
along (v) and across-strait (u) currents. In addition to the be on the SLA variability in Lombok Strait and along the
moored velocity datasets in Makassar Strait, we also use the southeastern coasts of the Indonesian archipelago (Figure 1).
hourly velocity data at pycnocline depth of 50–150 m from Furthermore, we will be focusing on the SLA variability
INSTANT moorings in Lombok Strait (Figure 1; for more during April–June 2004 (Figure 3), one of the events when
information on the Lombok mooring configurations, see both moorings in Makassar and Lombok Straits display a
Sprintall et al. [2009]). strong northward flow. The northward flow observed in
[8] Temperature stratification from 100 to 400 m within the Makassar Strait on 30 May 2004 (Figures 4b, 4c, and 4d)
Makassar Strait pycnocline are resolved by the temperature was preceded by mass piling up (+) against the southwestern
and pressure sensors attached to the Mak-West and Mak-East coasts of Sumatra on 12 May, followed up by southeastward
moorings. The Mak-West mooring provided higher resolution propagation of the + signal so that the southern coasts of Java
with 17 sensors attached, while the Mak-East mooring only was marked with + on 19 May (Figure 3). The occurrence
had 5 sensors. Since the vertical structure of temperature of + signal extended from the southern coasts of Java to

o
10
5 May 26 May
Pacific Ocean
trait

o
ar S

0
ass
Mak

Indian Ocean
o
-10o
10
12 May 2 June

o
0

o
-10o
10
19 May 9 June

o
0

o
-10
100oE 115oE 135oE 100oE 115oE 135oE

-30 -10 0 10 30
η (cm)

Figure 3. Snapshots of satellite altimetry-derived SLA [h] over the maritime continent during 5 May to
9 June 2004. The h data vary at intraseasonal timescales.

2025
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

(a) 4

Index
2

0
(b) 0.25
0

v (m/s)
-0.25
-0.50
250 m
-0.75
(c) 0.25
0
v (m/s)

-0.25
-0.50
350 m
-0.75
(d) 0.25
0
v (m/s)

-0.25
-0.50
450 m
-0.75
F M A M J J A S O N D J F M A M J J A S O N D J F M A M J J A S O N
2004 200 5 200 6
(e) 0.15 3
2004 2005 2005

MJO Index
2
v(m/s)

0
1
-0.15
03 09 15 21 27 02 08 14 03 09 15 21 27 02 08 14 03 09 15 21 27 02 08 14
March April May June September Oct.

Figure 4. (a) Time series of Madden-Julian Oscillation (MJO) index. The MJO index is based on the
work of Wheeler and Hendon [2004], defined as the sum of the squares of the first two principal compo-
nents of the combined fields of near equatorially averaged 850 hPa zonal wind, 200 hPa zonal wind, and
satellite-observed outgoing longwave radiation data. Time series of subinertial (dashed line) and intrasea-
sonal (solid line) along-strait flow at (b) 250, (c) 350, and (d) 450 m. The MJO and velocity data were
observed from 20 January 2004 to 26 November 2006. (e) Plots of the intraseasonal along-strait flow
time series (solid line) at 150 m and MJO index (dashed line) during strong MJO phases.

the mooring sites in Makassar Strait via Lombok Strait on [11] Over the course of 2004–2006, there were 17 northward
26 May although there are some concerns on the SLA data (+v0 ) events with a recurring period of 2–3 months (Figures 4b,
quality within the internal Indonesian seas. 4c, and 4d). The +v0 during May 2004 and 2005 was the most
energetic of all +v0 events, while other less energetic +v0 events
3. Intraseasonal Variability in Makassar Strait peaked in February, March, August, and December 2004;
and Lombok Strait January, March, July, September, and November 2005; and
January, April, June, August, September, and November
3.1. Weakened ITF Events at Intraseasonal Timescales 2006. The +v0 episodes demonstrated a significant correlation
[10] Focusing on the weakened or reversed southward with the zonal winds in the eastern equatorial Indian Ocean
flow in Makassar Strait, the v component at the base of (not shown), and those particularly occurring in March and
the pycnocline at 450-m depth (Figure 4d) demonstrates May 2004 and May and September 2005 were preceded by
that southward subinertial flow, vs (low-passed v time series significant Madden-Julian Oscillation (MJO) phases [Pujiana
with cut-off periods of 9 days) reversed to northward et al., 2009] (Figure 4a), with the +v0 events in Makassar Strait
flow five times over the course of 2004–2006: May 2004 lagging the MJO by 19–25 days (Figure 4e). Zhou and
and 2005, September 2005 and 2006, and November Murtugudde [2010] conjectured that the MJO influenced
2006. Although no ITF reversal was observed at depths the intraseasonal variability at Lombok Strait. They further
shallower than 250 m, the northward vs phases in May suggested that the meridional currents were reversed from
2004 and 2005 and in November 2006 were observed as southward to northward at Lombok Strait 15 days after the
shallow as 350 m. One feature consistently revealed within day of a peak MJO index. Thus, the lag of 19–25 days revealed
the along-channel pycnocline flow is the strong influence of by +v0 in Makassar Strait, with respect to a strong MJO event,
intraseasonal variability, v0 (band-passed v time series with suggests a link between the intraseasonal variability in
cut-off periods of 20 and 90 days) on vs: a northward Makassar Strait and MJO episodes. More discussion on
vs event always corresponds with a northward v0 phase the relationship between the +v0 events in Makassar Strait
(Figures 4b, 4c, and 4d). and the MJO significant phases are given in section 6.

2026
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

3.2. Subpycnocline Intraseasonal Variability 3.3. Kelvin Wave Signatures


[12] The v time series at 750 and 1500 m also display [14] As discussed above, intraseasonal motions account for
significant intraseasonal variability with a diminished subi- a significant proportion of the variance of the Makassar Strait
nertial background flow marked with recurring intraseasonal variability (Figures 4 and 5) and are clearly evident in the data-
oscillation with a 60-day period (Figures 5a and 5b). The sets as +v0 . In this subsection, we provide evidence linking the
intraseasonal variation explains ~80% of the total variance +v0 attributes to theoretical Kelvin wave characteristics such as
of the subpycnocline current meters v time series. Aside a vertically propagating wave, normal mode approximation,
from their energetic intraseasonal signatures, the v0 at 750 energy equipartition, and semi-geostrophic balance.
and 1500 m in Makassar Strait also displays a unique char-
acteristic that is not expected to be observed at depths where
the stratification frequency is quasi-homogeneous: the flows 3.3.1. Vertically Propagating Waves
at 750 and 1500 m are out of phase. A lagged correlation [15] Over the 17 occasions when vs turned northward or was
between the v0 data at 750 m and that at 1500 m shows a greatly attenuated during 2004–2006, one consistent feature
statistically significant correlation with r2 = 0.8, and a coher- was that the +v0 profile across the pycnocline indicated upward
ence analysis further indicates that the 60-day oscillation is phase propagation as +v0 at deeper levels led that at shallower
the most coherent among the v0 data at both depths with a levels. The composites of the +v0 data, formed by isolating
phase shift of 180 (not shown). each +v0 event and then averaging over the set of events
[13] Dynamic modes, inferred from the observed observed during 2004–2006, demonstrated weak flow at
Brunt-Väisällä in Makassar Strait (N, Figure 1), suggest that depths shallower than 100 m, with a velocity core in the lower
the deepest zero crossing depths for the second and third bar- thermocline at 200 m at the Mak-East and 350 m at the
oclinic modes are at 850 and 1000 m, respectively, which Mak-West (Figure 6). There is an upward phase tilt (dz/dt)
might explain the 180 phase shift between v0 at 750 m and associated with the +v0 of 50 m/day (Figure 6).
that at 1500 m. The v0 data from 50 to 1500 m, attributed [16] What ocean dynamics might force this upward phase
to the 17 +v0 episodes during 2004–2006 observed in the propagation of +v0 events in Makassar Strait? We suggest
pycnocline depths (Figures 4b, 4c, and 4d), exhibit a verti- that this trait is consistent with the dynamics of vertically
cal structure which signifies a baroclinic structure with propagating Kelvin waves. If a wind-forced Kelvin wave
two zero crossing depths, indicative of a second baroclinic perturbed the surface of a stratified ocean, the energy attrib-
wave mode. Examples of the measured +v0 vertical structure uted to the perturbation would not only propagate horizontally
in May, September, and November of 2005 when the +v0 away from the perturbation point but also downward making a
was maximum are shown in Figure 5c. horizontal slope:

(a) 0.25
v (m/s)

-0.25
(b) 0.25
v (m/s)

-0.25
F M A M J J A S O N D J F M A M J J A S O N D J F M A M J J A S O N
2004 2005 2006
(c) -5 0

-250

-450
Depths (m)

-750

31 May 2005 23 Sep 2005 18 Nov 2005


-1500
-0.1 0 0.1 0.2 0.3 -0.1 0 0.1 0.2 0.3 -0.1 0 0.1 0.2 0.3
v (m/s) v (m/s) v (m/s)

Figure 5. The along-strait flow time series varying at subinertial (dashed line) and intraseasonal time-
scales (solid line) observed at (a) 750 and (b) 1500 m within the Mak-West subpycnocline. (c) Vertical
structure of the intraseasonal along-strait flow data across the pycnocline and at two subpycnocline depths
at three events when the intraseasonal along-strait flow at the pycnocline attains maximum positive value.

2027
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

−50
0.2
−100

−150

−200

Depths (m)

v (m/s)
−250 0

−300

−350

−400
-0.2
−450
-10 -8 -6 -4 -2 0 2 4 6 8 10 -10 -8 -6 -4 -2 0 2 4 6 8 10
Time from peak (days) Time from peak (days)

Figure 6. The composite of the northward along-strait flow at intraseasonal timescales observed at the
Mak-West (left panel) and Mak-East (right panel) pycnocline. Day 0 was defined as the time of the peak
northward flow at 350 m. The dashed line with a narrow denotes the phase line, and the angle (green curve)
that the line makes with the horizontal dashed line indicates upward phase shift of 50 m/day.

o dz where m is the vertical wavenumber and ro is the back-


θ¼ ¼ (1)
N dx ground density. Assuming density (r0 ) fluctuations can be
represented by temperature (T0 ) fluctuations and vertical dis-
where o is a wave frequency, N is the stratification placements of density0surfaces are solely driven by the vertical
frequency, z is the depth, and x is the horizontal distance. 0 @
r
velocity following @r @t þ w @z , and hydrostatic balance holds
As our composite plots are in depth and time domains, the @r0 0 0 0
slope equation can be equivalently written as @z ¼ r g , it is implied that the highs and lows of T and p
0
are separated by p/2. Since p is in phase with the wave-
dz o:c induced horizontal velocity vectors (u0 , v0 ), it is also implied
¼ (2)
dt N that T0 and (u0 , v0 ) attributed to a theoretical Kelvin wave vary-
ing at a certain frequency would have the same phase shift.
where c is a theoretical phase speed of a baroclinic wave
Plots of the v0 and T0 time series for several depths, which
mode and t is time. Thus, an observer, stationed at a distance
focus on an episode of strong +v0 in late May 2004, demon-
away from where a Kelvin wave is generated, would see the
strate that maximum northward along-strait flow occurs prior
linear wave signature arrive deeper in the water column first.
to maximum temperature oscillations (Figure 7), implying that
[17] To test whether the observed phase slope (Figure 6)
T0 would lead v0 in a space domain. This agrees well with the
inferred from the +v0 composite is consistent with the slope
v0 and T0 relationship required for a theoretical upward propa-
formula, we use an averaged N from several Conductivity, Tem-
gating Kelvin wave. Nevertheless, Figure 7 does not demon-
perature, Depth (CTD) casts in Makassar Strait (Figure 1).
strate a uniform phase shift of p/2 between v0 and T0 across
The phase speed (c) is inferred from the normal mode anal-
depths as one would expect from a vertically propagating
ysis of the N profile. Given o = 2p/(20  90) days, c = 1.2–
Kelvin wave with a specific frequency. Figure 7 shows that
2.5 m/s (the phase speed for the first two wave modes), and
the lag between v0 and T0 increases with depth and is of 15 days
N = 0.0068 s1, thus dz/dt varies within 15–115 m/day, which
at the depth of 375 m, which is equivalent to a phase shift of
agrees fairly well with the phase slope inferred from the v0 com-
p/2 for a 60-day Kelvin wave.
posite (Figure 6). Thus, the pattern of upward phase shift with a
slope of 50 m/day observed in the v0 composite consistently
expresses a linear Kelvin wave signature. 3.3.2. Energy Equipartition
[18] A vertically propagating Kelvin wave can also be [19] As in the oscillation of a pendulum, a theoretical Kel-
identified from how temperature fluctuations and velocity vin wave also expresses equal partition between its
vectors associated with the wave are related. The pressure corresponding kinetic and potential energies [Pedlosky,
perturbation (p0 ) and its corresponding vertical velocity 2003]. We use the v0 and T0 data to examine whether the en-
(w0 ) for a vertically propagating Kelvin wave are ergy equipartition in the wave field also characterizes the
Z intraseasonal motions observed in the Makassar Strait pyc-
0 2
p ðx; y; z; t Þ ¼ ^p cosðmzÞdm (3) nocline. To estimate the potential energy (PE), the tempera-
p
0
ture time series over several depths ranging the pycnocline
0 N 2 @2p of the Mak-West mooring is converted to vertical displace-
w ¼ (4)
ro @z@t ments (0 ) using a heat equation

2028
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

0.2
250 m 325 m 0.6

v (m/s)

T (oC)
0 0

-0.2 -0.6
0.2 0.6
275 m 350 m

T (oC)
v (m/s)
0 0

-0.2 -0.6
0.2
300 m 375 m 0.6
v (m/s)

T (oC)
0 0

-0.2 -0.6
01 07 13 19 25 31 06 12 18 24 30 01 07 13 19 25 31 06 12 18 24 30
May June May June

Figure 7. The along-strait flow (solid line) and temperature (dashed line) time series varying at intraseaso-
nal timescales observed at several depths within the Mak-West pycnocline during May–June 2004. Positive
velocity indicates northward flow.
0
0 T ðz; t Þ -150
 ðz; t Þ ¼ (5)
@ T
@z

in which horizontal advection, diffusion, and heat sources -200


are neglected. We also remove the effect of static stability
on the thermal field through normalizing 0 with the ratio
between N and its corresponding vertical average No -250
0 0 N
 n ðz; t Þ ¼  ðz; t Þ (6)
Depths (m)

No
-300
where z, t, and subscript n denote depth, time, and normal-
ized, respectively. The PE per unit mass is computed as
0:5ro N 2 0 2 , where 0 2 is the variance of vertical displace-
ments averaged over intraseasonal frequency band and ro -350

is the averaged density across pycnocline depths. Kinetic


energy (KE) is inferred from the spectra of v0 and defined
as 0:5ro v0 2 where v0 2 is the variance attributed to v0 averaged -400
over intraseasonal periods. Vertical structures of PE and KE
(Figure 8) demonstrate that PE and KE have similar ampli-
tudes, indicative of equipartition between PE and KE, which -450
further supports our hypothesis that the Kelvin wave field 0 0.4 0.8 1.2
contributes significantly to intraseasonal motion in the Energy per unit mass
Makassar Strait thermocline. (m2s-2)

3.3.3. Dispersion Relation Figure 8. Vertical structure of potential (solid line) and
kinetic (dashed line) energy across the lower pycnocline.
[20] The dispersion relation for a Kelvin wave in Makassar The potential and kinetic energy are inferred from tempera-
Strait, a North-South-oriented channel, would be o = lc, where ture and along-strait flow varying at intraseasonal timescales
l and c are wavenumber in the y direction and wave phase observed at the Mak-West.
speed, signifying a constant phase speed or nondispersive
relationship. We employ the v0 datasets at Lombok and
Makassar Straits to investigate whether the coherent Lombok Strait and Makassar Strait are not only statistically co-
oscillation transmitting from Lombok Strait to Makassar Strait herent but also show a linear trend of decreasing phase differ-
reported by Pujiana et al. [2009] exhibits a dispersion ence with period (Figure 9), indicative of a propagating feature
relationship consistent with a Kelvin wave. with a quasi-constant phase speed of 1.2  0.01 m/s from
[21] The v0 time series at the upper 200 m of Lombok Strait Lombok Strait to Makassar Strait (a distance of 1240 km,
and that at depths below 75 m of Makassar Strait are statistically the 100-m isobath length connecting the moorings at Lom-
coherent above the 95% significance level across the period bok and Makassar Straits, is used in the computation). Thus,
band of 45–90 days, and the corresponding phase differences the relationship between o and l inferred from the coher-
linearly decrease with period. For example, v0 at 100 m in ence analysis exemplifies a nondispersive wave.

2029
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

Period (days)
90 50 33 25 20
1.0
1.6

coherence2
95% confidence level
0.5
1.4

Lombok vs Makassar

ve
ω (s-1)
0
1.2

wa
180

lv in
Ke
1.0
phase

0.8

-180
0.011 0.020 0.033 0.040 0.050 0.5 1.0 1.5 2.0
Frequency (cycle/day) (m-1)

Figure 9. (Left) Coherence amplitudes and phases between the intraseasonal along-strait flow time series
at 100 m in Lombok Strait (Lombok-East/Le) and those at 100 m in Makassar Strait (Mak-West). The
phase 95% confidence limits are shaded. (Right) Dispersion relation diagram inferred from the phase shift
plot shown in the left panel.
3.0
[22] In addition to the relation between o and l, the relation
between l and m (vertical wavenumber) is also useful to deter-
mine the wave classification. Substituting a free wave solution
of the form, p0 (x,y,z,t) = f(x)Re{exp[i(ot  ly  mz)]}, to 2.0

the following equation of motion for the pressure and its m (cycle/ 2000 m) n=1
corresponding boundary condition [Romea and Allen, 1983]:
1.5
 0 0
@ @2p 2 @ 1 @p
þ f ; (7)
@t @x2 @z N 2 @z
@ @p
0
@p
0
1.0 n=2
þf ¼ 0; (8)
@t @x @y
would yield 0.5

@ f
2
 ðfm=No Þ2 f ¼ 0 (9) n=-1
@x2
0
@f -8 -6 -4 -2 0 2
 ðfl=oÞf ¼ 0 (10) (cycle/6000 km)
@x
@f Figure 10. Dispersion diagram of horizontal versus zonal
f; < ; x ! e: (11)
@x e wavenumber at v = 2p60 day1. The markers indicate the
estimated values from the data, the solid black line demon-
[23] Substituting the solution to equation (9), f = exp(flx/o), strates dispersion relation for a theoretical Kelvin wave,
results in the following dispersion relation: and the grey lines are the dispersion relation for the first
o ¼ lNo =m: (12) two modes of theoretical Rossby waves.
Kelvin wave attributes measured at Lombok Strait: semi-
[24] The estimated l and m obtained from coherence anal- geostrophic balance and reflected Kelvin wave.
yses of v0 at Lombok and Makassar Straits also approximate [26] A simple wave equation and its boundary condition
well the Kelvin wave dispersion curve (Figure 10). applicable for simplified Lombok Strait, a channel bounded
by vertical walls, are
3.3.4. Semi-geostrophic Balance and Increased Decay 0  2 
Scale @2 o  f 2
þ  l 2
¼0 (13)
@x2 c2
[25] Observations and numerical experiments agree that the
0  
Kelvin wave energy originating in the equatorial Indian Ocean @ lf 0
significantly influences the intraseasonal variability at Lom-   ¼ 0; x ¼ 0; W (14)
@x o
bok Strait [Arief and Murray, 1996; Qiu et al., 1999; Schiller
et al., 2010; Drushka et al., 2010]. Drushka et al. [2010] pro- where 0 is the sea level anomaly across the strait, l is the
posed that the vertical structure of the observed Lombok Strait wavenumber in the along-strait axis direction, and W is
throughflow at intraseasonal timescales was best explained by the strait width. The general solution, with o = lc, is
0 fx
a baroclinic Kelvin wave. We will be discussing two other  ðx; t Þ ¼ o e c cosðly þ ot Þ , where c denotes Kelvin

2030
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

0.1
wave phase0 speed. Given the 0 corresponding momentum (a)
equation is v ðf 2  o2 Þ  gf @ @ 2  ¼ 0, we can then obtain
0

@x þ g @x@y

dη(m)
0
0 g @ 0
v ¼ (15)
f @x
which implies a semi-geostrophic balance. -0.1
[27] The u0 , v0 obtained from moorings and 0 from the pres- (b) 0.8
sure gauges and altimeter are employed to examine the Kelvin
wave signatures in Lombok Strait. We focus on the May 2004 0.4

v(m/s)
event during which the instruments were all operational.
Moorings and pressure gauges in Lombok Strait demonstrated 0
that +v0 and +0 characterized the Lombok Strait intraseasonal
variability during May 2004 (Figure 11). We also identify that -0.4
the across-strait SLA slopes down toward the east (@ 0 /@ x), (c) 0.15
with higher sea level at Pw than Pe, when v0 > 0 during May
2004 (Figure 11), indicating a balance between @ 0 /@ x and

η(m)
zonal coriolis force. Meanwhile, the along-strait sea surface tilt 0
(@ 0 /@ y), inferred from satellite-derived SLA measured at two
locations in Lombok Strait (Pn and Ps; Figure 1), is not bal-
anced by f in the y direction as the sea surface slopes up toward -0.15
01 05 09 13 17 21 25 29 02 06 10 14
the south (@ 0 /@ y < 0) when v0 > 0 (Figures 11b, 11c, and May June
12a), indicative of ageostrophy. Moreover, the estimated geo-
strophic +v0 inferred from @ 0 /@ x does not differ significantly Figure 12. (a) Across-strait gradient of SLA (hPw–hPe,
from +v0 measured from moorings (Figure 12b), which con- solid) computed from the shallow pressure gauges and along-
firms a partial geostrophic balance in the x direction. strait gradient of satellite altimetry-derived SLA (hPn–hPs,
[28] The two shallow pressure gauges at the zonal bound- dashed) in Lombok Strait. (b) Geostrophic flow inferred from
aries of Lombok Strait also indicated enhanced across-strait the moorings (solid line) and pressure gauges (dashed line).
decay scale, another Kelvin wave attribute. The estimated 0 (c) The observed hPw (black line) and hPe (grey line), and
variability at Pe the estimated hPe = hPwe w/R (dashed grey line). Figures 12a,
0 0 12b, and 12c demonstrate data from May to mid-June 2004
 Pe ¼  Pw eW =R ; (16) which vary at intraseasonal timescales.
with W = 0.3R the distance between the two pressure gauges
and R is the Rossby radius of deformation, is less than the
(a)
0.15
η(m)

Pw Pe
-0.15

(b) 0.8 20
currents (m/s)

0.4 10
ηi(m)

0 0

-0.4 -10
Lw 100 m Le 100 m
-0.8 -20

(c)
0.8 20
currents (m/s)

0.4 10
ηi(m)

0 0

-0.4 -10

-0.8 Lw 125 m Le125 m -20


01 05 09 13 17 21 25 29 02 06 01 05 09 13 17 21 25 29 02 06
May June May June

Figure 11. (a) Time series of SLA at intraseasonal timescales during May to 6 June 2004 from two
shallow pressure gauges deployed in Lombok Strait, Pw and Pe. (b, c) Time series of along-strait flow
(solid black), across-strait flow (solid grey), and vertical isopycnal displacements (dashed gray) varying
at intraseasonal timescales observed at Lombok moorings.

2031
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

observed 0 Pe (Figure 12c). The discrepancy between the reduced amplitude to propagate along the southern Kalimantan
measured and estimated 0 Pe implies increased across-strait coasts into the Makassar Strait. In contrast, the gap is too
decay scale. Durland and Qiu [2003] connected a low-fre- wide for the observed baroclinic Kelvin wave with a defor-
quency Kelvin wave passage in a strait with enhanced mation radius of less than 100 km to jump over.
across-strait decay scale, mainly forced by the formation of [30] A likely Kelvin wave pathway from Lombok Strait to
a reflected Kelvin wave whose direction was 180 out of Makassar Strait is along the isobaths, across the Sunda conti-
phase with the incident wave. When a Kelvin wave in nental slope, and at depths greater than 50 m (Figure 13a),
Lombok Strait reaches an entrance into a larger basin, it does where a slope-trapped baroclinic Kelvin wave may propagate
not contain the required energy to force a wave of equal towards Makassar Strait with shallower water to the left. The
amplitude in that basin, which leads to the genesis of a propagation of a baroclinic Kelvin wave along an isobath
reflected wave in order to preserve the continuity of pressure has been studied through both laboratory experiments and
at the strait mouth. Since the maximum amplitude of the observations [Codiga et al., 1999; Hallock et al., 2009].
reflected wave is on the opposite side of the strait to the Figure 13a depicts the 100-m isobath, which extends for
maximum amplitude of the incident wave, the reflection ~1240 km and links the moorings in Lombok and Makassar
subtracts more from the minimum amplitude of the incident Straits. A baroclinic Kelvin wave with a speed of 0.8–1.2 m/s,
wave than it does from the maximum amplitude. This likely equivalent with a speed of the theoretical second and third
results in an enhanced across-strait pressure gradient such as baroclinic mode Kelvin waves inferred from stratification,
observed in Lombok Strait (Figure 12c). would take 12–18 days to reach Makassar Strait, which is
in good agreement with the observations (12–16 days,
Figure 13b). We thus speculate that the Kelvin wave propa-
4. Waveguide from Lombok Strait to Makassar gates as the second and third baroclinic modes along the
Strait 100-m isobath directly linking Lombok to Makassar Strait.
[29] We have demonstrated that observations at Makassar
and Lombok Straits reveal intraseasonal motions with 5. Transport and Mixing Associated with Kelvin
Kelvin wave signatures, which signify propagation from Wave in Makassar Strait
Lombok Strait to Makassar Strait. Discontinuous coastlines
separate the moorings at both straits, raising the question 5.1. Transport
of the pathway guiding the wave from Lombok Strait to [31] We have provided supportive evidence for Kelvin
Makassar Strait. Given the averaged depth across section A wave propagation from the velocity vectors, temperature,
(Figure 13a) is 145 m, a barotropic Kelvin wave with a and stratification datasets. We now address a question of
deformation radius of 1900 km would overcome the gap of the relevance of Kelvin waves to ITF studies; more specifi-
~475 km and force another barotropic Kelvin wave of cally, what are the transport anomalies associated with the

(a)

trait
ar S

Kalimantan
ass
Mak

Sumatra
-3°

Sunda Shelf
section A

-6° 475 km

Sunda Strait Java


Lombok Strait
Bali
Bali Strait
-9°
100°E 104°E 108°E 112°E 116°E 120°E
(b) 0.20
v(m/s)

2004 2005 2006


-0.15
01 07 13 19 25 31 06 12 01 07 13 19 25 01 07 13 15 21 27 02 08 14 20 26 01
May June Sep Oct Aug Sep

Figure 13. (a) Potential pathway for a Kelvin wave propagating from Lombok Strait to Makassar Strait,
which extends for 1271 km along the 100-m isobath (dashed line). The black boxes show the mooring
sites in Makassar and Lombok Straits, and the thin dashed contour depicts the 100-m isobath. (b) The
intraseasonal along-strait flow data observed at 150 m of Lombok Strait (solid line) and Makassar Strait
(dashed line).

2032
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

(a) 4

Transport (Sv)
2
0
-2
-4
-6
(b) -150 0.3
Depths (m) -200
-250 0.2

Ri
-300
0.1
-350
-400 0
FM A M J J A S O N D J F M A M J J A S O N D J FM A M J J A S O N
2004 2005 2006

Figure 14. (a) The along-strait transport magnitudes varying at intraseasonal (solid line) and longer
(>90 days, dashed line) timescales estimated across depths of 125–450 within the Makassar Strait
pycnocline. (b) Time series of Ri obtained from the ratio between the vertical shear of intraseasonal
along-strait flow (dv0 /dz) and N in the Makassar Strait pycnocline. The dv0 /dz and N data are from the
Mak-West mooring. Ri > 0.25 is given in white.

intraseasonal baroclinic Kelvin waves in Makassar Strait? nondiffusive, and stably stratified flow, small disturbances
To deduce the estimated transport anomalies, we use the v0 are stable provided the Richardson number, Ri = N2/(dv0 /dz)2,
datasets at depths of 150–450 m from the Mak-West and is greater than 1/4 everywhere in the fluid. The relationship
Mak-East moorings. At each depth of z, we fit the v0 time between a Kelvin wave and Kelvin-Helmholtz instability
series from both moorings using an exponential function arises from the Kelvin’s wave perturbation to v0 and T0 .
0 0
Thus, if a Kelvin wave were important to govern the v0 varia-
v ¼ v Mak-west ex=b (17) tion across the pycnocline depths in Makassar Strait, we
would expect that the energetic wave episodes correspond
which results in the across-strait structure of v0 , v 0 (x,z,t),
with small Ri.
where b is the horizontal scale coefficient, and x = 0:
[34] The Ri variability (Figure 14b) across the Makassar
L (x = 0 and x = L correspond with the Mak-West and Mak-East
Strait pycnocline is computed using the observed vs and N
mooring sites, respectively). The transport is then computed as
at the Mak-West site, where N was computed by assuming
Z z¼150 Z x¼L r is equivalent to T. Figure 14b demonstrates that the likeli-
0
v ðx; z; t Þdxdz (18) hood of unstable stratification within the lower pycnocline
z¼450 x¼0
depths is greater during the period when observed Kelvin
wave signatures are stronger in Makassar Strait (e.g., late
[32] The transport anomalies observed during 2004–2006 May and December 2004, May, June 2005, and November
across the pycnocline (Figure 14a) indicate that a downwelling 2005). Furthermore, the corresponding eddy viscosity
intraseasonal Kelvin wave passage could drive northward coefficient is parameterized using Kv = 5.6  108 Ri8.2, a
mass transport of 0.75  0.4 Sv (mean  standard deviation) formula appropriate in a region where density variations
across depths of 150–450 m within the Makassar Strait pycno- are dominated by those of temperature [Thorpe, 2004] such
cline with a maximum northward transport of ~2 Sv recorded as in Makassar Strait, and the vertical diffusivity during
in late May 2005 (Figure 14a). Figure 14a also demonstrates strong Kelvin wave events varies from 1 to 5  105 m2 s1.
that intraseasonal Kelvin waves in Makassar Strait did not
reverse the southward ITF due to smaller northward transport
6. Discussion and Summary
anomalies that the waves induce. It thus concluded that the
northward transport, attributed to a downwelling Kelvin wave [35] We have described the characteristics of the 2- to 3-month
in Makassar Strait, contributed to weaken the southward ITF oscillations measured in Makassar Strait during 2004–2006,
transport during 2004–2006. marking the weakened ITF transport across the pycnocline
and the vertically sheared v0 at the subpycnocline. A compos-
5.2. Mixing ite of the 17 +v0 events found within the pycnocline depths
[33] Ffield and Gordon [1992], using an archive of CTD exhibits an upward phase tilt, which implies a vertically prop-
datasets and the 1-D advection and diffusion equation, agating wave attribute with a non-local origin. The v0 and T0
characterized tidal currents as the main force for the data across the pycnocline show a quarter cycle of phase shift,
strong vertical mixing with vertical diffusivity in excess which favors downward energy propagation, and also the
of 104 m2 s1 in the Makassar Strait thermocline. At kinetic-potential energy equipartition. Warmer pycnocline
intraseasonal timescales, Figures 4b, 4c, 4d, 5, and 6 demon- occurring during the +v0 episodes is also evident from the
strate that a Kelvin wave episode in Makassar Strait is marked relation between v0 and T0 . Furthermore, a normal mode
with the vertically sheared v0 , which potentially yields unstable approximation is also applicable to explain the link between
stratification through Kelvin-Helmholtz instability and contri- the vertical structure of v0 and stratification. We suggest that
butes to the vertical mixing strength. In a steady, inviscid, these characteristics pertinent to the 2- to 3-month variability

2033
PUJIANA ET AL.: KELVIN WAVES IN MAKASSAR STRAIT

observed in the Makassar Strait are in agreement with the RMM1RMM2.74toRealtime.txt). This is Lamont-Doherty contribution
number 7651.
theoretical downwelling Kelvin wave signatures.
[36] The 0 , u0 , and v0 data in Lombok Strait also demon-
strate two Kelvin wave attributes in the semi-geostrophic References
balance between v0 and d0 /dx and in enhanced d0 /dx, which Arief, D., and S. P. Murray (1996), Low frequency fluctuations in the Indonesian
corroborates those reported by Drushka et al. [2010]. The v0 throughflow through Lombok Strait, J. Geophys. Res., 101, 12455–12464,
doi:10.1029/96JC00051.
data in Lombok Strait and those in Makassar Strait are Codiga, D. L., D. P. Renourad, and A. Fincham (1999), Experiments on
coherent and yield a Kelvin wave dispersion relationship waves trapped over the continental slope and shelf in a continuously strati-
between the wavenumber and wave frequency. The waves fied rotating ocean, J. Mar. Res., 57, 585–612.
propagate from Lombok Strait to Makassar Strait with a Drushka, K., J. Sprintall, S. Gille, and I. Brodjonegoro (2010), Vertical
structure of Kelvin waves in the Indonesian Throughflow exit passages,
phase speed of ~1.2 m/s, a second mode baroclinic Kelvin J. Phys. Oceanogr., 40, 1965–1987, doi:10.1175/2010JPO4380.1.
wave speed. The second baroclinic wave mode inferred Ducet, N., P. Y. Le Traon, and G. Reverdin (2000), Global high-resolution
from the measured Kelvin wave phase speed affirms the mapping of ocean circulation from TOPEX/POSEIDON and ERS-1 and -2,
baroclinic vertical structure attributed to the +v0 events in J. Geophys. Res., 105, 19477–19498, doi:10.1029/2000JC900063.
Durland, T., and B. Qiu (2003), Transmission of subinertial Kelvin waves
Makassar Strait which exhibits two zero crossing points through a strait, J. Phys. Oceanogr., 33, 1337–1350.
over depths. Through the comparison of the +v0 episodes Ffield, A., and A. L. Gordon (1992), Vertical mixing in the Indonesian
in Lombok Strait and Makassar Strait, we suggest that the thermocline, J. Phys. Oceanogr., 22, 184–195.
Gordon, A. L., and R. A. Fine (1996), Pathways of water between the
Kelvin waves navigate along the 100-m isobath and Pacific and Indian Oceans in the Indonesian seas, Nature, 379, 146–149.
extends for ~1240 km to transmit its energy from Lombok Gordon, A. L., and R. D. Susanto, and K. Vranes (2003), Cool Indonesian
Strait to Makassar Strait. throughflow as a consequence of restricted surface layer flow, Nature,
425, 824–828.
[37] Drushka et al. [2010] demonstrated that the model v0 Gordon, A. L. (2005), Oceanography of the Indonesian seas and their
data in Lombok Strait, simulated by a simple wind-forced throughflow, Oceanography, 18, 14–27.
model forced by zonal winds in the eastern equatorial Indian Gordon, A. L., and R. D. Susanto, A. Ffield, B. A. Huber, W. Pranowo,
Ocean and along-shore winds along the Sumatra and Java S. Wirasantosa (2008), Makassar Strait throughflow 2004 to 2006,
Geophys. Res. Lett., 35, L24605, doi:10.1029/2008GL036372.
coasts, showed a good agreement with observation. Because Gordon, A. L., and J. Sprintall, H. M. Van Aken, R. D. Susanto, S. Wijffels,
the intraseasonal motions in Makassar Strait and Lombok R. Molcard, A. Ffield, W. Pranowo, S. Wirasantosa (2010), The
Strait are significantly correlated, the intraseasonal Kelvin Indonesian Throughflow during 2004-2006 as observed by the INSTANT
program, Dyn. Atmos. Oceans, 50, 115–128, doi:10.1016/j.
waves in Makassar Strait derive their energy from the same dynatmoce.2009.12.002.
southeastern Indian Ocean winds. Hallock, Z. R., W. J. Teague, and E. Jarosz (2009), Subinertial slope-
[38] Significant MJO phases preceded the Kelvin wave trapped waves in the northeastern Gulf of Mexico, J. Phys. Oceanogr.,
39, 1474–1485, doi:10.1175/2009JPO3925.1.
episodes in Makassar Strait occurring in March and May Pedlosky, J. (2003), Waves in the ocean and atmosphere, Springer, New
2004 and May and September 2005 with the maximum +v0 York.
at 150 m associated with the Kelvin wave trailing the peak Pujiana, K., and A. L. Gordon, J. Sprintall, R. D. Susanto (2009), Intrasea-
of MJO by 19–25 days. Meanwhile, the MJO leads +v0 at sonal variability in the Makassar Strait thermocline, J. Mar. Res.,
67, 757–777, doi:10.1357/002224009792006115.
150 m in Lombok Strait by 6–11 days, implying that the B. Qiu, M. Mao, and Y. Kashino (1999), Intraseasonal variability in the
MJO’s footprint in Makassar Strait lags that in Lombok Indo-Pacific throughflow and the regions surrounding the Indonesian
Strait by 13–14 days, consistent with the number of days a seas, J. Phys. Oceanogr., 29, 1599–1618.
Romea, R. D., and J. S. Allen (1983), On vertically propagating coastal
baroclinic Kelvin wave requires to propagate from the Kelvin waves at low latitudes, J. Phys. Oceanogr., 13, 1241–1254.
Lombok Strait moorings to the Makassar Strait moorings. Schiller, A., S. E. Wijffels, J. Sprintall, R. Molcard, and P. R. Oke (2010), Path-
Unlike the study of Zhou and Murtugudde [2010], which ways of intraseasonal variability in the Indonesian throughflow region, Dyn.
Atmos. Oceans, 50, 174–200, doi:10.1016/j.dynatmoce.2010.02.003.
suggested that the MJO affected the intraseasonal variability Sprintall, J., A. L. Gordon, R. Murtugudde, and R. D. Susanto (2000), A
only at the ITF outflow passages such as Lombok Strait, we semiannual Indian Ocean forced Kelvin wave observed in the Indonesian
propose that the MJO-related oceanic Kelvin wave is also seas in May 1997, J. Geophys. Res., 105, 17217–17230, doi:10.1029/
observed in Makassar Strait. 2000JC900065.
Sprintall, J., S. Wijffels, A. L. Gordon, A. Ffield, R. Molcard, D. Susanto,
[39] The downwelling Kelvin wave passages affect the I. Soesilo, J. Sopaheluwakan, Y. Surachman, and H. Van Aken (2004),
ITF transport anomalies, in which a wave episode results A new international array to measure the Indonesian throughflow, EOS
in an averaged northward anomaly of ~0.75 Sv across the Trans AGU, 85, 39, doi:10.1029/2004EO390001.
Sprintall, J., S. Wijffels, R. Molcard, and I. Jaya (2009), Direct estimates of
pycnocline depths of 150–450 m. The wave passages are Indonesian throughflow entering the Indian Ocean: 2004-2006, J. Geophys.
also commensurate with the small Richardson number Res., 114, C07001, doi:10.1029/2008JC005257.
phases, during which dv0 /dz is intensified, amplifying the Thorpe, S. A. (2004), Recent developments in the study of ocean turbulence,
likelihood of Kelvin-Helmholtz instability resulting in a Annu. Rev. Earth Planet. Sci., 91–109, doi:10.1146/annurev.
earth.32.071603.152635.
vertical mixing with Kv of 1–5  105 m2 s1. Wijffels, S., and G. Meyers (2004), An intersection of oceanic waveguides:
Variability in the Indonesian throughflow region, J. Phys. Oceanography,
[40] Acknowledgments. We appreciate the helpful comments provided 34, 1232–1253.
by Andreas Thurnherr, Hsien-Wang Ou, and four anonymous reviewers. This Wheeler, M. C. and H. H. Hendon (2004), An all-season real-time multivariate
work is supported by the National Science Foundation grants OCE-0219782 MJO index: Development of an index for monitoring and prediction,
and OCE-0725935 (LDEO) and OCE-0725476 (SIO). The altimeter data were Mon. Weather. Rev., 132, 1917–1932.
produced by SSALTO/DUACS and distributed by AVISO. The MJO index Zhou, L., and R. Murtugudde (2010), Influences of Madden-Julian oscillations
was obtained from http://cawcr.gov.au/staff/mwheeler/maproom/RMM/ on the eastern Indian Ocean and the maritime continent, Dyn. Atmos.
Oceans, 50, 257–274, doi:10.1016/j.dynatmoce.2009.12.003.

2034

You might also like