Oceanography - Lamont-Doherty Earth Observatory

Regional Oceanography
of the Philippine Archipelago
B y A r no l d L . G o r don , J a n e t S p r i n ta l l , a nd A m y F f i e l d
Photo courtesy of Leilani Solera
| Vol.24, No.1
Oceanography
14
This article has been published in Oceanography, Volume 24, Number 1, a quarterly journal of The Oceanography Society. © 2011 by The Oceanography Society. All rights reserved. Permission is granted to copy this article for use in teaching and research. Republication, systemmatic reproduction,
or collective redistirbution of any portion of this article by photocopy machine, reposting, or other means is permitted only with the approval of The Oceanography Society. Send all correspondence to: info@tos.org or Th e Oceanography Society, PO Box 1931, Rockville, MD 20849-1931, USA.
P h i l i pp i n e S t r a i t s D y n a m i c s E x p e r i m e n t
Abstr ac t. Confined by the intricate configuration of the Philippine Archipelago,
forced by the monsoonal climate and tides, responding to the remote forcing from the
open Pacific and adjacent seas of Southeast Asia, the internal Philippine seas present a
challenging environment to both observe and model. The Philippine Straits Dynamics
Experiment (PhilEx) observations reported here provide a view of the regional
oceanography for specific periods. Interaction with the western Pacific occurs by way
of the shallow San Bernardino and Surigao straits. More significant interaction occurs
via Mindoro and Panay straits with the South China Sea, which is connected to the
open Pacific through Luzon Strait. The Mindoro/Panay throughflow reaches into the
Sulu Sea and adjacent Bohol and Sibuyan seas via the Verde Island Passage and Tablas
and Dipolog straits. The deep, isolated basins are ventilated by flow over confining
topographic sills that causes upward displacement of older resident water, made more
buoyant by vertical mixing, which is then exported to surrounding seas to close the
overturning circulation circuit.
Introduc tion
Following multiple pathways, waters of
the western Pacific enter the complex,
multidimensional array of seas and
straits that form the impressive archipelago stretching some 3400 km from
Southeast Asia to Australia. The northern
segment of this system is the Philippine
Archipelago (Figure 1a), where the North
Equatorial Current bifurcation near 14°N
(Nitani, 1972; Toole et al., 1990; Qiu and
Lukas, 1996; Qu and Lukas, 2003) forms
the western boundary for the equatorward-flowing Mindanao Current and
the nascent poleward-flowing Kuroshio.
Pacific water seeps into the Sibuyan and
Bohol (Mindanao) seas by way of the
shallow San Bernardino and Surigao
straits, respectively, and in greater
volume through the 2200-m-deep Luzon
Strait into the South China Sea (Metzger
and Hurlburt, 1996, 2001; Centurioni
et al., 2004; Qu et al., 2006). From the
South China Sea, the flow enters into the
Sulu Sea through Mindoro and Panay
straits, and eventually into the western
Bohol (Mindanao) Sea through Dipolog
Strait, and perhaps into the Sibuyan Sea
by way of the Verde Island Passage and
Tablas Strait. The South China Sea also
has access to the southern Sulu Sea via
Balabac Strait. The Sibutu Passage links
the southern Sulu Sea to the Sulawesi Sea.
Once within the confines of the
Philippine Archipelago, circulation and
stratification are subjected to monsoonal
winds that are textured by passages
between island morphology (Pullen
et al., 2008, 2011; May et al., 2011),
by sea-air heat and freshwater fluxes
including river outflow, and by regions
with strong tidal currents. Overflow
across < 500-m-deep topographic sills
ventilates the depths of isolated basins,
the Sulu Sea, and the smaller Bohol
and Sibuyan seas.
The Office of Naval Research sponsored the Philippine Straits Dynamics
Experiment (PhilEx) with a goal
of exploring the oceanography and
dynamics in the narrow straits and deep
basins of the Philippine Archipelago
using both observations and model
output. During PhilEx fieldwork,
conductivity, temperature, depth, and
dissolved oxygen measurements were
obtained, and lowered acoustic Doppler
current profiler (CTD-O2/LADCP)
Oceanography
| March 2011
15
b
122°E
126°E
16°N
119°E
120°E
121°E
Luzon
122°E
100
124°E
Pacific Ocean
6000
0
10
10
0
6000
2000
100
100
00
10
Mindanao
100
0
Regional IOP: Mar 2009
Regional IOP: Jan 2008
Joint: Nov/Dec 2007
Exploratory: Jun 2007
4000
7°N
119°E
8°N
0
ADCP Moorings
126°E
9°N
100
10
400
15
0
100
Bohol Sea
10
0
Sulu Sea
4000
Sibutu
Passage
10
0
1000
100
10
0
10 500
00
100
500
20
00
1500
50
0
1500
10
0
500
124°E
6°N
10°N
10
00
100
1000
0
10
2000
500
2000
00
10
0
10
0
0
10
40
8°N
120°E
11°N
0
10
Sulawesi Sea
122°E
12°N
100
120°E
ch
Ar
2000
118°E
lu
Su
0
4000
0
100
00
1000
Borneo
0
10
0
100
50
00
40
6°N
Bohol
Negros
8°N
Mindanao
500
go
la
ipe
00
Samar
Leyte
100
500
100
100
00
15
00
te
Cebu
10°N
9°N
1000
0
400
20
10
ba
Panay
100
1000
Bohol Sea
Dipolog Strait
1500
13°N
100
0
50
10°N
an
w
la
Pa
2000
1000
0
0
10
Surigao
Strait
Bohol
Negros
Sulu Sea
00
0
00
as
10
00
10
00
4000
11°N
4000
0
200
1500
10
1500
40
12°N
1000
20
00
40
Balabac
Strait
Leyte
Gulf
Sea
1500
100
8°N
Leyte
Camotes
50
0
00
500
10
1500
100
0
00
Panay
1000
100
Mindoro
0
1000
Visayan
Sea
100
an
aw
Pacific Ocean
100
10
10
Panay
Strait
6000
500
0
0
South
10
Sibuyan
00
Sea
500
126°E
0
M
12°N
125°E
10
0
50
San Bernardino
0
Strait
10
Tablas
Strait
100
100
10
l
e
500
15
00
Pa
500
Sibuyan Sea
100
10
10°N
sag
0
50
Mindoro
10
124°E
14°N
Pas
100
0
1000
1500
Mindoro
Strait
50
land
100
00
40
e Is
500
South China Sea
12°N
Luzon
Verd
00
40
2000
Lamon Bay
14°N
13°N
100
South China
Sea
1000
00
100
4000
15
1000
500
123°E
1000
120°E
100
118°E
100
a
16°N
121°E
122°E
123°E
10
00
124°E
7°N
MMP Moorings
125°E
126°E
Figure 1. (a) Bathymetry of the seas and straits of the Philippine Archipelago from Smith and Sandwell (1997) and http://topex.ucsd.edu/marine_topo.
(b) Conductivity, temperature, depth, and dissolved oxygen, and lowered acoustic Doppler current profiler (CTD-O2/LADCP) stations obtained by the
four PhilEx cruises identified within the figure legend. PhilEx mooring positions are indicated.
stations were made during the exploratory cruise of June 2007 and two
regional Intensive Observational Period
cruises in the winters of January 2008
(IOP-08) and March 2009 (IOP-09;
Figure 1b). CTD casts were also undertaken during the Joint US/Philippines
Cruise of November and December
2007. All PhilEx cruises were conducted
from R/V Melville, whose underway sea
surface water system provides sea surface
temperature (SST) and salinity (SSS) at
high resolution, as well as other meteorological and oceanographic parameters.
The hull-mounted shipboard 75- and
150-kHz ADCP system measures circulation of the upper hundreds of meters
of the water column along the ship
track. The cruise data provide nearly
synoptic views of regional water-column
circulation and stratification. Moorings
(Figure 1b) provide time series of the
currents within major straits. Data from
MacLane Labs moored profilers (MMPs;
Figure 1b) characterize the internal
wave environments (Girton et al., 2011).
Satellite remote sensing provides a
variety of regional observations, such as
Arnold L. Gordon (agordon@ldeo.columbia.edu) is Professor, Earth and Environmental
Sciences and Associate Director, Ocean and Climate Physics, Lamont-Doherty Earth
Observatory of Columbia University, Palisades, NY, USA. Janet Sprintall is Research
Oceanographer, Scripps Institution of Oceanography, University of California,
San Diego, La Jolla, CA, USA. Amy Ffield is Senior Scientist, Earth & Space Research,
Upper Grandview, NY, USA.
16
Oceanography
| Vol.24, No.1
SST, sea level, ocean color, and wind. The
Panay Island-based high-frequency (HF)
radar array provides high-resolution sea
surface current information in Panay
Strait. In situ measurements of ocean
currents and properties are provided
by towed instrumentation and freefloating sensors, such as surface drifters
(Ohlmann, 2011), profilers (Girton,
2011), and gliders. Biological parameters
closely related to the ocean physical
processes are also a component of PhilEx
(Cabrera et al., 2011; Jones et al., 2011).
Model studies (Hurlburt et al., 2011;
Arango et al., 2011; May et al., 2011)
complete the suite of methods employed
by PhilEx researchers to investigate the
oceanographic conditions and processes
within the Philippine Archipelago.
The coupled ocean/atmosphere is
characterized by variability across a
GPCP Precipitation (mm/day)
b)
2007.0
2007.5
2008.0
2008.5
2009.0
120˚ 123˚ 126˚
2009.5
10
9
14˚
9
8
12˚
8
10˚
7
7
8˚
6
6
5
5
4
4
3
3
2
2
1
1
2
1
0
-1
-2
2006.5
2007.0
2007.5
2008.0
2008.5
2009.0
-2
-1
0
1
2
2009.5
Nino 3 (dashed)
2006.5
a) 10
Anomaly
wide range of spatial and temporal
scales. Beyond tidal forcing and daily
weather conditions are intraseasonal
Madden Julian Oscillations, seasonal
monsoon forcing, and interannual
El Niño-Southern Oscillation (ENSO),
plus decadal and longer fluctuations. A
recurring question arises when dealing
with observational data, which inherently have gaps in spatial and temporal
coverage: how typical are the observations of the entire region, and of the
longer-term “climate” conditions?
Observations can be “leveraged” by
being used to evaluate model output,
or directly assimilated into the models,
thus allowing for a reliable model-based
glimpse of this fuller spectrum of events.
In addition to a change in the winds,
the monsoon brings a change in precipitation to the Philippine Archipelago,
which is expected to be reflected in
surface layer thermohaline seasonal
stratification. Although there are localized variations due to the interaction
of the wind with orographic lifting
processes and subsequent river runoff,
in general, over the ocean waters, the
winter monsoon brings a time of less
rainfall and the summer more rainfall
(Figure 2a). However, comparison of the
actual rainfall for a specific year to the
average annual cycle reveals precipitation anomalies (Figure 2b), which may
be expected to induce anomalies in the
surface layer thermohaline stratification. During the eight months prior to
the June 2007 exploratory cruise, there
was an anomalous dry period, an accumulative response to the El Niño of the
preceding year, whereas the IOP-08 and
IOP-09 cruises were conducted during
anomalously wet periods.
The objective of this paper is to
Figure 2. (a) The averaged precipitation time series (green line), mean-annual time series
(purple dashed line), and mean (thin horizontal black line) are shown for data from locations near/within the Philippine seas. (b) The anomaly time series (green=wet, yellow=dry)
is shown, along with Nino3 (blue dashed line) for comparison. In (a), the four precipitation
data locations are indicated by red squares on the map inset. On both (a) and (b), the three
PhilEx cruise periods are indicated by vertical red bars. The precipitation data are from the
GPCP Satellite-Gauge (SG), One-Degree Daily (1DD), Version 1.1 data set that is produced
by optimally merging estimates computed from microwave, infrared, and sounder data, and
precipitation gauge analyses from October 1996 to June 2009 (http://lwf.ncdc.noaa.gov/oa/
wmo/wdcamet-ncdc.html#version2).
provide an introduction to the regional
setting in order to place in context the
information presented in the collection of studies described in this special
PhilEx issue of Oceanography. Each topic
included in this regional introduction
has been covered by PhilEx researchers
(e.g., published to date are Han et al.,
2008; Rypina et al., 2010; Tessler et al.,
2010), is presented in this special issue
of Oceanography, or will be further
developed in future publications. Here,
we use data from the regional PhilEx
hydrographic cruises and moored time
series measurements to describe the
flow pattern and water mass distribution within the Philippine Archipelago.
Our goal is to determine the potentially
important first-order processes that
might lead to observed circulation
patterns within the major Philippine
basins of the Sulu and Bohol seas, and
to provide schematic overviews of the
circulation to serve as a guide to the more
detailed studies that are found in this
issue of Oceanography or are to follow.
Regional Str atification
and Circul ation
The Sea Surface Layer
Surface-layer circulation observed by
the shipboard ADCP (Figure 3a,b)
for IOP-08 and IOP-09 brings out the
complexity of the pattern of surface
currents and the relationship to SST.
The direct connection of the
Philippine seas to the western Pacific is
through San Bernardino Strait, with a sill
depth of 92 m near 12°47'N, 124°14'E,
and Surigao Strait, with a sill depth
Oceanography
| March 2011
17
a
14°N
14°N
119°E
120°E
121°E
122°E
123°E
124°E
125°E
126°E
14°N
119°E
120°E
121°E
Cool SST
122°E
123°E
124°E
125°E
126°E
14°N
Luzon
Luzon
Cyclonic
eddy
13°N
Cool SST
Mindoro
yc
Cool SST
Cyclonic eddy
Cool SST
11°N
an
aw
1 m/s
an
aw
1 m/s
l
Pa
10°N
l
Pa
as
ba
Panay
27.2
Cebu
Sulu Sea
9°N
27.2
28.0
Strong tidal
currents
Bohol
11°N
10°N
10°N
Jet
9°N
28.4
R0827.6
SST NB150:
28.025-55m 28.4
11°N
Leyte
Bohol
Bohol
Sulu Sea
27.6
12°N
Strong tidal
currents
Cebu
Negros
9°N
12°N
Samar
te
Leyte
Negros
10°N
13°N
Strong tidal
currentsSamar
te
Panay
Cyclonic eddy
11°N
ba
M
Tablas
Island
dy
ed
12°N
as
Tablas
Island
dy
ed
c
ni
lo
yc
tic
An
SCS
12°N
M
Mindoro
c
ni
lo
13°N
13°N
Strong tidal
currents
tic
An
Cyclonic
SCS
eddy
Bohol Jet
Iligan Bay
Eddy
Cool SST
Mindanao
9°N
Iligan Bay
Mindanao
123°E
124°E
125°E
126°E
Eddy
Cool SST
119°E
120°E
121°E
122°E
123°E
124°E
125°E
126°E
14°N
14°N
119°E
120°E
121°E
122°E
123°E
124°E
125°E
126°E
Luzon
119°E
120°E
121°E
122°E
R08 SST NB150: 25-55m
b
119°E
14°N
120°E
Cool 121°E
SST
Luzon
122°E
Mindoro
SCS
125°E
126°E
flo
ba
w
S
flo
SC
ce
to
rfa
Tablas
Island
Panay
SC
S
Panay
n
1 m/s
n
1 m/s
10°N
Cebu
Negros
9°N
9°N
Sulu Sea
8°N
27.2
Bohol
Strong tidal
currents
Jet
27.6
Bohol Jet
Iligan Bay
Eddy
Mindanao
Iligan Bay
Eddy
Mindanao
28.0
28.4
Cool SST
8°N
27.2
11°N
10°N
10°N
9°N
Cool SST
Sulu Sea
Strong tidal
currents
Bohol
Bohol
Negros
11°N
Leyte
Cebu
wa
la
Pa
12°N
Leyte
wa
la
Pa
12°N
Samar
te
to
11°N
13°N
Strong tidal
Samar
currents
te
M
as
w
Su
12°N
ba
Tablas
Island
ce
12°N
M
as
14°N
13°N
Strong tidal
currents
Mindoro
rfa
SCS
Su
13°N
10°N
124°E
Cool SST
13°N
11°N
123°E
9°N
8°N
8°N
R0927.6
SST OS150:
28.023-55m 28.4
7°N
7°N
R09 SST OS150: 23-55m
7°N
7°N
119°E
120°E
121°E
122°E
123°E
124°E
125°E
126°E
119°E
120°E
121°E
122°E
123°E
124°E
125°E
126°E
Figure 3. Current vectors from the shipboard-mounted 150-KHz ADCP system within
the 25- to 55-m layer color coded by sea surface temperature (SST). A 1 m s-1 arrow is
given for scale. SCS is the South China Sea. Various features are shown in italics. (a) PhilEx
regional cruise January 2008. (b) PhilEx regional cruise March 2009.
18
Oceanography
| Vol.24, No.1
of 58 m near 10°17'N, 125°50'E (both
topographic estimates are from http://
topex.ucsd.edu/marine_topo; Smith
and Sandwell, 1997). A comparison of
the thermohaline water-column profile
within these straits to that of the western
Pacific depicts an environment of
vigorous mixing, with possible upwelling
of subsill-depth western Pacific water
entering into the confines of the straits.
The currents within San Bernardino
and Surigao straits were the strongest
observed during all the PhilEx cruises.
In each strait, the average speed of the
upper 50 m for the one- to two-day
ship occupation was about 0.5 m s-1
from the western Pacific to the interior
seas. However, the strong tidal currents
obscure the nontidal flow and therefore prohibit meaningful estimation of
the nontidal throughflow, other than
suggesting that it is likely that the direction of the throughflow was into the
Philippine interior seas.
The shipboard ADCP and watercolumn temperature/salinity profiles
do not show clear continuity of the
San Bernardino water into the Sibuyan
or Camotes seas. In contrast, the Surigao
Strait characteristics do intrude into the
Bohol Sea within a well-defined surface
current across the northern Bohol Sea,
linking the Surigao Strait throughflow
to Dipolog Strait with export into the
Sulu Sea. South of this “Bohol Jet” is a
cyclonic circulation feature that we call
the Iligan Bay Eddy, found on all of the
PhilEx cruises, near 124°E (Figure 3).
Generation of the Iligan Bay Eddy
is not coupled to a sea surface ocean
color signal, as may be expected from
upwelling usually associated with a
cyclonic circulation pattern (Cabrera
et al., 2011). The Bohol Jet enters the
Sulu Sea, though its path once within
the Sulu Sea is not clear. Relatively cool
SST is observed south of the Bohol Jet
extension into the eastern Sulu Sea, a
consequence of upwelling along the
Zamboanga coast (Villanoy et al., 2011).
Sulu Sea surface layer circulation
(Figure 3) displays much mesoscale
activity at horizontal scales of ~ 100 km,
with varied SST and SSS, rather than a
clear basin-scale gyre. However, these
synoptic “snapshots” offered by the
shipboard ADCP mapping do not allow
for conclusions about the general circulation pattern of the Sulu Sea as each
cruise covered different segments of the
this sea. The warmer SST observed in
March 2009 relative to January 2008 may
be partly the normal seasonal signal,
though in 2009 the survey extended
further to the south, and 2008 tended
to have stronger winds and surface
currents. The January 2008 surface
circulation of the Sulu Sea east of 120°E
is cyclonic, with northward flow along
the eastern boundary that continues
into Panay Strait and southward flow
west of 121.3°E. The Sulu circulation was
anticyclonic in June 2007 (not shown).
The March 2009 cruise data covers
different parts of the Sulu Sea so that
direct comparison with January 2008
is not practical, though again we find
northward flow in the eastern Sulu Sea
and within Panay Strait, albeit relatively
subdued relative to January 2008.
In Mindoro and Panay straits, the
January 2008 currents form energetic
eddies in response to complex wind
stress curl (Rypina et al., 2010; May
et al., 2011; Pullen et al., 2011), though
on average, the surface currents are
directed toward the South China Sea.
In June 2007, the surface flow was
weak and toward the South China Sea.
In March 2009, the surface flow in
Panay and Mindoro straits, with less
mesoscale activity, was also toward
the South China Sea. In January 2008,
there was a strong cyclonic eddy in the
South China Sea adjacent to Mindoro
summer SST was about 2° warmer than
in January 2008 or March 2009. The
SSS range between cruises amounts to
1.0 psu, with the lowest SSS in January
2008. The highest SSS is in June 2007, a
consequence of the normal dry season
compounded by the drier conditions
“
PhilEx did much to advance our
understanding the waters of the
Philippine Archipelago.
Strait (Pullen et al., 2008). The surface
current in Tablas Strait displayed a weak
cyclone (note westward flow at 121.8°E
between Mindoro Island and Tablas
Island, Figure 3), though the tendency
in March 2009 was for flow out of the
Sibuyan Sea into the Panay-Mindoro
corridor. The throughflow of the relatively cool SST of the Verde Island
Passage was weak, slightly toward the
west in January 2008, and toward the
east in March 2009.
Water-Column Stratification Upper 400-m Profiles
With descent into the water column,
the warm, low-salinity surface water
above the top of the thermocline at
50–70 m rapidly gives way to cooler,
saltier water at 200 m (Figure 4, upper
panels). The potential temperature (θ)
drops by ~ 10°C over only 100 m,
from 70 to 170 m, coinciding with an
intense pycnocline in which density
increases by nearly three sigma-0 units,
from 22.5 to 25.2. The June 2007 early
”
of an El Niño (Figure 2). The lowest
SSS was observed in January 2008 (as
well as during the Joint Cruise of late
2007), a consequence of the phasing out
of the previous wet season, whereas in
March 2009, nearly two more months
into the dry season, SSS was slightly
more elevated. The lowest SSSs (< 33.4)
are observed in the Bohol Sea and the
South China Sea entrance to Mindoro
Strait during the winter regional
cruises, perhaps a consequence of the
delayed river runoff from their respective larger neighboring landmasses of
Mindanao and Luzon. The Sibuyan
and Camotes seas’ SSSs are between
33.4 and 33.6. The seasonal influence
determined by comparing the salinity
differences between the PhilEx cruises
is found to reach to about 130 m, into
the mid thermocline.
The relatively warm thermocline data
(Figure 4) are from the western Pacific
adjacent to San Bernardino and Surigao
straits. In the salinity profiles, these
stations display a pronounced salinity
Oceanography
| March 2011
19
a
Potential Temperature (oC)
8
0
12
16
20
Salinity
24
28
32
33.2
33.6
34.0
34.4
34.8
0
50
50
100
erm
Th
Pressure (db)
150
line
100
oc
Pacific
200
250
200
250
Expo-07: Jun 2007
IOP-08: Jan 2008
IOP-09: Mar 2009
300
350
300
350
400
400
21
20
30
σ0
22
Surface layer
23
Pacific
Subtropical4
S-max 2
25
20
26
uth
12
380-440 m
.4
26
800-1200 m
Sulu Sea
10
Panay Sill
480-580 m
South China Sea
ina
Ch
26
a
Se
15
σ0
IOP-08
25
So
Potential Temperature (oC)
150
.8
26
8
27
10
Pacific
NPIW
34.42
5
34.44
34.46
33.6
34.0
34.4
34.50
6
Salinity
28
33.2
34.48
34.8
Salinity
b
Camotes Sea
Camotes Sea
400
S. Sibuyan Sea
800
1200
Sibuyan Sea
1600
S. Sibuyan Sea
Sibuyan Sea
1200
Panay Strait
Pressure (db)
800
Tablas Strait
400
Bohol Sea
Sulu
Sea
2000
1600
Bohol Sea
2000
Sulu Sea
9
10
11
12
Potential Temperature (oC)
13
14
0.0
0.4
0.8
1.2
1.6
Oxygen (ml/L)
Figure 4. CTD temperature and salinity obtained on the June 2007 PhilEx exploratory cruise,
the January 2008 regional cruise, and the March 2009 regional cruise. (upper panels) Potential
temperature and salinity profiles with pressure (depth) for the upper 400 m for the three
PhilEx cruises. (lower panels) Potential temperature and salinity scatter plot. The lower right
panel shows the 6° to 13°C strata.
20
Oceanography
| Vol.24, No.1
maximum (s-max) in the 75–250 m,
16° to 28°C interval that marks North
Pacific Subtropical Water. Although
the shallow San Bernardino and
Surigao straits block s-max water,
North Pacific Subtropical Water has
access to the South China Sea via
Luzon Strait. However, at the South
China Sea entrance to Mindoro Strait,
the pronounced s-max core in the
20° to 28°C range is greatly attenuated
with only a weak s-max near 28°C
as observed in the June 2007 cruise
data, or a temperature/salinity (T/S)
flexure near 25°C in data from the
regional 2008 and 2009 cruises, and
a deeper s-max near 16°C (Figure 4).
The processes within the South China
Sea that attenuate the North Pacific
subtropical s-max are beyond the
scope of this paper.
A salinity minimum (s-min) is
observed from 350 to 600 m in the
western Pacific stations, marking
North Pacific Intermediate Water,
which also has access to the South
China Sea via Luzon Strait. An
attenuated but still visible s-min
near 10°C is observed in Mindoro
and Panay straits. It is this water that
provides the overflow into the Sulu
Sea (Tessler et al., 2010). The θ/S
structure (Figure 4, lower left panel)
further brings out the stratification
features of the Philippine waters,
particularly the “gap” between the
western North Pacific saline subtropical water and fresher thermocline
water of the Philippine waters, as
well as the attenuated North Pacific
Intermediate Water s-min in Mindoro
and Panay straits.
Two CTD stations (Figure 4, upper
panels) in the southern Sulu Sea from
March 2009 show relatively warm, salty
water between 125 and 150 m, marking
the trough of solitons observed in that
area (Jackson et al., 2011).
Deep Basin Ventilation
The Philippine seas are composed of
numerous deep basins isolated from one
another by topographic barriers. There is
the open deep Pacific Ocean to the east;
there are the relatively large seas to the
west and south of the Philippines (the
Sulu, South China, and Sulawesi seas);
and there are the smaller interior seas,
most notably the Bohol and Sibuyan
seas, and the still smaller Visayan and
Camotes seas (Figure 1). Below roughly
500 m, these seas have marked differences in θ (and S) and oxygen values
from each other and from the source
water column of the open western North
Pacific. The deep Sulu Sea has a potential
temperature of 9.9°C, the deep Bohol Sea
11.6°C, and the Sibuyan Sea 10.4°C; the
southern Sibuyan Sea is slightly warmer
at 10.7°C, with the warmest isolated
basin; and the Camotes Sea has a bottom
potential temperature of 13.2°C.
These marked property differences
are a product of sill depths of the topographic barriers to the neighboring
seas. The isolated deep basins are
ventilated by spillover at the topographic barriers that then descend to
the depths, replacing resident water
made less dense by vertical mixing. The
resident water is lifted upward by denser
overflow water, and is subsequently
exported to the surrounding seas to
close the overturning circulation cell.
As these waters are reduced in oxygen
by the rain of organic material from
the sea surface, their export to neighboring seas can be traced as an oxygen
minimum. For example, the Bohol Sea
oxygen minimum near 12°C is observed
spreading near 300 m throughout
the Sulu Sea, with traces entering
into Panay Strait.
The effective sill depths are found by
matching the bottom temperature with
the temperature profile of the external
source water. The shallower the sill, the
warmer the basin waters. The relationship of source sill depth θ/S to deep
basin water θ/S depends on the mixing
of ~ 500 m. The oxygen minimum of
the Sibuyan Sea serves as the overflow
water source for the South Sibuyan Sea,
which, with further oxygen consumption, accounts for the near zero oxygen
(< 0.3 ml l -1) of the southern Sibuyan Sea
bottom water.
The coldest deep basin is that of
the Sulu Sea. The source of deep Sulu
Sea water is generally considered to be
South China Sea water entering through
Mindoro Strait (Broecker et al., 1986).
“
[PhilEx] points the way toward further,
more quantitative research and illustrates the
need for high spatial and temporal resolution
in both observations and modeling.
environment at the controlling sill and
the mixing/entrainment environment
of the descending plume: the effective
sill depth is less deep than the deepest
passage as determined by sonic surveys.
The warmest deep basin is the interior
Camotes Sea, with a controlling sill
depth of less than 300 m. The oxygen
of the Camotes deep water is near
0 ml l -1, indicating slow ventilation relative to the oxygen consumption. The
oxygen levels of the bottom water of
the southern Sibuyan Sea are also near
0 ml l -1. They are slightly warmer than
the bottom water of the main Sibuyan
Sea. The South Sibuyan Sea has an effective sill depth of ~ 400 m, while the main
Sibuyan Sea has an effective sill depth
”
However, the deep salty bottom water
of the Sulu Sea does not match a sole
South China Sea source (Figure 4, lower
right panel). Quadfasel et al. (1990)
recognized that the density of the
South China Sea source cannot reach
the bottom of the Sulu Sea unless there
is substantial addition of suspended
sediment to make for a denser blend.
Based on evidence from sedimentary
records of the Sulu Sea, Quadfasel et al.
(1990) suggest that episodic turbidity
currents from the South China Sea at
intervals of several decades (on average
50 years) may have played an important
role in plunging dense water toward the
bottom of the Sulu Sea. However, this
still would not explain the salty bottom
Oceanography
| March 2011
21
water of the Sulu Sea.
PhilEx observations of stratification
and currents between June 2007 and
March 2009 reveal a strong overflow
between 400- to 570-m depth from
Panay Strait into the Sulu Sea (Tessler
et al., 2010). The overflow water is
derived from approximately 400-m deep
in the South China Sea. Sulu Sea stratification indicates that the overflow does
not descend below 1250 m in the Sulu
Sea, but rather settles above high-salinity
deep water. The mean observed overflow
transport at the sill is 0.32 Sv, with a
residence time of 11 years in the affected
Sulu layer from 575 to 1250 m.
While Sulu Sea ventilation to
~ 1250 m is drawn from the 570-m-deep
northern sill in Panay Strait, we speculate that the deeper Sulu water may be
derived from the Sulawesi Sea to the
south by way of the Sibutu Passage. The
Sibutu Passage sill is around 350 m, but
delivers denser water into the Sulu Sea
than does the South China Sea. This
is primarily because of the shallower
pycnocline of the Sulawesi Sea compared
to that found in the South China Sea,
and the strong tidal heaving of the Sibutu
Passage pycnocline that can lead to
rather startling solitons within the Sulu
Sea (Apel et al., 1985).
Along-Strait Velocity 150 m
0.4
velocity (m/s)
0.2
0
−0.2
−0.4
−0.6
−0.8
Mindoro
2008 F
M
Panay
A
M
J
Tablas
J
A
S
Dipolog
O
N
D
2009 F
M
D
2009 F
M
Along-Strait Velocity 420 m
0.4
velocity (m/s)
0.2
0
−0.2
−0.4
−0.6
−0.8
2008 F
M
A
M
J
J
A
S
O
N
Figure 5. Along-strait velocity at 150 m and 420 m measured at the moorings
shown in Figure 1.
22
Oceanography
| Vol.24, No.1
Time Serie s
The time series of moored velocity observations provide a longer-term context
for the “archipelago-scale” fieldwork
undertaken as part of PhilEx (Figure 1),
such as the synoptic shipborne flow and
property measurements of the regional
survey described above, and the shortterm drifter deployments (Ohlmann
et al., 2011). The high-resolution time
series data from the moorings are also
used as an important metric to test the
veracity of numerical models in the
Philippine region (e.g., Hurlburt et al.,
2011; Pullen et al., 2011).
The current measurements from
the PhilEx moorings (Figure 5) in
Mindoro, Panay, Tablas, and Dipolog
straits (see Figure 1 for locations) reveal
much variability and baroclinity over
the ~ 15-month deployment period.
At 150 m, the strong southward flows
at Mindoro and Panay are evident
as two intraseasonal (30–60 day)
pulses during the northeast monsoon
(December–February). During the
southwest monsoon (April–October)
when the flow is mainly northward at
150 m in Mindoro and Panay, the flow
is more southward at Tablas. Some of
this southward Tablas flow is probably
siphoned off to contribute to the stronger
northward flow observed at Mindoro
compared to Panay during this period.
The along-strait flow at 150 m in Dipolog
Strait exhibits more of a short-period
(15–20 day) signal and is primarily
eastward. At 300-m depth, the flow in
all four passages is also characterized by
short-period variability (not shown). At
420 m, there is high coherence between
the consistently strong southward flow
through Mindoro and Panay straits.
At this depth, the flow is near bottom
in Mindoro Strait and often slightly
stronger than that observed at Panay
Strait, although the deeper, near-bottom
flow in Panay Strait (~520 m) can
reach velocities of > 1 m s-1. This strong
benthic overflow contributes to the
ventilation of the deep Sulu Sea (Tessler
et al., 2010). In Dipolog Strait, the nearbottom 420-m flow is strongly eastward
from the Sulu Sea into the Bohol Sea,
although as in the shallower layers, the
flow here is also characterized by strong
15–20-day variability.
is of prime importance in ventilating
the subsurface layers of the Bohol Sea.
The eastern end of the Bohol Sea is
connected to the Pacific Ocean across
the broad, shallow Leyte Sea through
the 58-m-deep Surigao Strait, where
there appears to be a small net flow of
surface water into the eastern Bohol Sea.
As described above, Surigao inflow
streams across the northern Bohol Sea
to be exported into the Sulu Sea through
Dipolog Strait. The northern Bohol Sea
is connected to San Bernardino Strait by
way of a 330-km-long narrow channel
(based on ADCP and CTD data) in these
channels is negligible, but this channel
may be an effective way to deliver river
runoff from the islands of the central
Philippines, and reduce the SSS of the
northern Bohol Sea.
The LADCP profiles reveal the
highly layered circulation profile within
Dipolog Strait (Figure 6), with two layers
of inflow into the Bohol Sea and two
outflow layers. Near the topographic sill
where the PhilEx mooring was sited, the
flow into the Bohol Sea occurs within the
thermocline from roughly 80 to 200 m
Bohol Se a
that runs through the Camotes Sea to
enter the Bohol Sea both to the east and
west of Bohol Island. This channel has
an 18-m-deep constriction to the northeast of Bohol Island and a 3-km-wide,
though deep (280 m), channel to the
northwest of Bohol Island. PhilEx observations indicate that the throughflow
and in the benthic layer overflow at the
Dipolog sill. The 150-m and 420-m time
series (Figure 5), which show eastward
flow, are consistent with the LADCP
data. The two layers exported into the
Sulu Sea consist of the surface water
of the upper 50 m and a second layer
centered at 300 m. The inflow/outflow
With a sill depth of 504 m, the
47-km-wide (as measured between the
100-m isobaths) Dipolog Strait between
Negros and Mindanao separates the
Bohol Sea from the Sulu Sea and is the
deepest connection of the Bohol Sea to
surrounding seas. The Dipolog Strait
Westward
0
North
Eastward
0
54
55
North
South
56
57
58 54
55
Towards South
Towards Sulu
South
56
57
58
0
Towards North
Out of Bohol Sea
Mainly northern side of strait
100
100
into Bohol Sea
200
100
Towards Bohol
200
200
300
Depth (m)
Depth (m)
Towards Sulu
Dipolog Strait Throughflow
300
300
Out of Bohol Sea
400
123°00'E
Towards Bohol
400
123°30'E
400
9°00'N
Bohol
500
500
Dipolog
Overflow into Bohol Sea
8°30'N
600
-0.3
-0.2
-0.1
0.0
U (m/sec)
0.1
0.2
0.3
0
10
Dipolog
Sulu
LADCP
U
500
Bohol
V
20
30
40
0
10
Distance (km)
20
30
40
m/sec
-0.4
0.0
0.4
Figure 6. Regional 2008 LADCP data from within Dipolog Strait. (left) LADCP profiles of the zonal flow within the deep channel shown in the
map insert. The position of the moorings is shown as a yellow triangle. (right) The LADCP section across Dipolog Strait showing zonal and
meridional currents.
Oceanography
| March 2011
23
cores are tilted across Dipolog Strait
(Figure 6), with the outflow strongest
at the northern side of the strait and
the inflow strongest along the southern
side, consistent with the Coriolis force;
the deep overflow is confined by the
constrictive seafloor topography.
One can envision a double estuary
overturning circulation within the Bohol
Sea (Figure 7). The shallow estuary
circulation is composed of surface water
outflow to the Sulu Sea, compensated
with upwelling by entrainment of thermocline inflow waters into the Bohol Sea,
bolstered by the Surigao throughflow.
The deeper estuary overturning circulation is controlled by dense water overflow
to the depths of the Bohol Sea within
the lower 50–100 m of Dipolog Strait,
with export in the 300–350-m interval
toward the Sulu Sea derived from the
upwardly displaced resident water. This
water is low in oxygen (~ 1.3 ml l-1) and
is the likely source of a low-oxygen core
within the Sulu Sea within that depth
interval. Estimates from the LADCP
and mooring time series suggest that the
deep overturning circulation amounts to
~ 0.2 Sv. The westward transport in the
upper limb of the shallow cell, as estimated from the PhilEx cruises’ LADCP
data across Dipolog Strait, may amount
to ~ 0.5 Sv, part of which is drawn from
Surigao Strait. As the LADCP average
for the lower limb is ~ 0.2 Sv, the Surigao
Strait throughflow is probably around
0.3 Sv, assuming the Bohol Sea river
inflow is negligible.
Sulu Sea
‘Double’ Estuary Pattern
Mindoro and Panay
Throughflow
Mindoro and Panay straits connect
the Sulu Sea with the South China Sea.
These straits exhibit much variability
in depth and width. Apo Reef near
12.66°N represents a significant obstacle
within Mindoro Strait (Ohlmann, 2011).
Between Mindoro and Panay straits
south of the Semirara islands (11°N,
121°30'E), there is an east-west offset
of the these straits, where Tablas Strait
connects Mindoro and Panay straits with
the Sibuyan interior sea. The sea between
Mindoro and Panay straits may be
considered a triple junction of connective passages. This region received much
PhilEx attention in the form of process
studies. That analysis is under way, and
Bohol Sea
Low salinity surface water export
Shallow Estuary
“Pervasive” O2-min
2
Sulu O2-max, 400-600 m derived from South China Sea
Base of O2-max ~600 m
Figure 7. Schematic representation
of the water exchange between the
Bohol Sea and the Sulu Sea through
Dipolog Strait. The depiction is
based on the CTD-O2/LADCP data
from the PhilEx cruises shown in
Figures 4 and 6.
24
Oceanography
| Vol.24, No.1
Dipolog Sill
500 m
47 km wide at 100 m
Deep Estuary
outside of the scope of this regional
oceanography view.
The 150-m and 450-m mooring time
series (Figure 5) reveal much along-strait
flow variability in Panay and Mindoro
straits. The mean throughflow in Panay
Strait at 150 m is northward, albeit
weak from April through October, with
strong southward flow during the winter
months. The Panay Strait January 2008
and March 2009 LADCP profiles near
the mooring site (Figure 8a) show a net
transfer of surface water above ~ 100 m
from the Sulu Sea into the South China
Sea, with flow toward the Sulu Sea
associated with the s-max below 150 m,
and stronger flow below 400 m feeding
the overflow into the Sulu Sea (Tessler
et al., 2010). The South China Sea s-max
-20
30
-10
0
10
is found throughout the Sulu Sea, and
enters into the Bohol Sea as part of a
deep limb of the shallow overturning
circulation cell.
The Mindoro throughflow (Figures 5
and 8b) is similar to the Panay structure.
The flow above 150 m is toward the
South China Sea, that is, to the northwest (the along-strait orientation at the
mooring site), with reversals during the
winter months (Figure 5); below 300 m,
the flow is toward Panay Strait with weak
intervening flow.
A schematic of the Mindoro/Panay
throughflow (Figure 9) provides a sense
of the mean throughflow conditions.
However, wind-induced energetic eddies
as observed in January–February 2008
induce much intraseasonal activity in
20
12°00'N
IOP-09: 92-94
30
20
20
10
-40
-30
-20
-10
0
10
20
IOP-09: 115-117
Panay
20
this region (Pullen et al., 2008, 2011)
that can obscure the mean, longer-term
conditions. In the upper 150 m, there
is net flow toward the South China
Sea. Eddies are generated as this flow
encounters Apo Reef (Ohlmann, 2011).
At and below 150 m, the flow is toward
the Sulu Sea. Above ~ 500 m, this water
spreads at a similar depth into the Sulu
Sea, marking an s-max near 300 m and
an oxygen maximum near 500 m, traces
of which enter into the Bohol Sea. Spill
over topographic sills occurs into the
Semirara Sea (the isolated 1300-m-deep
basin south of the Semirara islands) and
over the Panay sill to depths of 1200 m in
the Sulu Sea (Figure 4a, lower left panel;
Tessler et al., 2010).
10
11°30'N
10 cm/sec
10
121°30'E
122°00'E
10
0
0
121°00'E
0
Mindoro
10 cm/sec
0
0
-100
-10
12°00'N
0
-300
-100
-10
-10
-200
-300
-20
-20
-400
-30
-10
-20
0
10
-20
-400
-40
-30
-20
-10
-200
-30
-20
-10
0
10
Figure 8. LADCP vectors color coded for depth in Panay and Mindoro straits
from January 2008 data. For Panay Strait stations, 92, 93, and 94 are averaged;
for Mindoro Strait, 115, 116, and 117 are averaged. The map inset for each
panel shows LADCP station positions; the green triangle marks the mooring
site in each strait.
20
Oceanography
| March 2011
25
South China Sea
Mindoro/Panay Straits
Apo reef
Western intensified
Center channel flow
Western intensified
Sulu Sea
Low salinity Sulu surface water 0-150 m 20-50 cm/s
SCS S-min/S-max layering, ~150-350 m ~10 cm/s, weakens as summer monsoon evolves
Benthic layer spill-over into Sulu
Mindoro Strait
~500 m
O2 -max ~500 m Sulu Sea
10-60 cm/s
Panay Strait
~600 m
entrainment
Semirara
Sea
Figure 9. Schematic representation of the water exchange between the South China Sea and the Sulu Sea through Mindoro
and Panay straits. The depiction is based on the CTD-O2/LADCP data from the PhilEx regional 2008 and 2009 cruises.
Conclusion
The region stretching 3400 km from
Australia to Southeast Asia, which may
be referred to as a “mega” archipelago,
separating the western Pacific and
eastern tropical Indian oceans, represents a complex, yet fascinating oceanic
environment. The western Pacific “sees”
a porous western boundary, representing
a challenge for both observational
and model research to unravel. The
regional circulation responds to strong
monsoonal winds textured by mountainous islands, and to complex ocean
bottom morphology—all amidst a
multitude of isolated deep basins within
a network of interconnecting straits.
Their actions modify the thermohaline
26
Oceanography
| Vol.24, No.1
stratification and impact both climate
and the marine ecosystem.
To provide context for the studies
included in this collection of PhilEx
results, this article presented a brief overview of a selection of topics depicting
the regional oceanography of Philippine
Archipelago waters. More thorough
analysis of each topic has been done elsewhere or is presently in preparation.
PhilEx did much to advance our
understanding of the waters of the
Philippine Archipelago. It points the
way toward further, more quantitative
research and illustrates the need for high
spatial and temporal resolution in both
observations and modeling.
Acknowledgements
This work was supported by Office
of Naval Research grant N00014-091-0582 to Lamont-Doherty Earth
Observatory of Columbia University,
ONR-N00014-06-1-0690 to Scripps
Institution of Oceanography, and
ONR- N0001406C0578 to ESR. LamontDoherty Earth Observatory contribution
number 7419.
Reference s
Apel, J.R., J.R. Holbrook, J. Tsai, and A.K. Liu.
1985. The Sulu Sea internal solitons experiment. Journal of Physical Oceanography
15(12):1,625–1,651.
Arango, H.G., J.C. Levin, E.N. Curchitser, B. Zhang,
A.M. Moore, W. Han, A.L. Gordon, C.M. Lee,
and J.B. Girton. 2011. Development of a
hindcast/forecast model for the Philippine
Archipelago. Oceanography 24(1):58–69.
Broecker, W., W. Patzert, R. Toggweiler, and
M. Stuiver. 1986. Hydrography, chemistry,
and radioisotopes in the Southeast Asian
basins. Journal of Geophysical Research
91:14,345–14,354.
Cabrera, O.C., C.L. Villanoy, L.T. David, and
A.L. Gordon. 2011. Barrier layer control of
entrainment and upwelling in the Bohol Sea,
Philippines. Oceanography 24(1):130–141.
Centurioni, L.R., P.P. Niiler, and D.K. Lee. 2004.
Observations of inflow of Philippine Sea surface
water into the South China Sea through the
Luzon Strait. Journal of Physical Oceanography
34:113–121.
Girton, J.B., B.S. Chinn, and M.H. Alford. 2011.
Internal wave climates of the Philippine seas.
Oceanography 24(1):100–111.
Han, W., A.M. Moore, E. Di Lorenzo, A.L. Gordon,
and J. Lin. 2008. Seasonal surface ocean
circulation and dynamics in the Philippine
Archipelago region during 2004–2008.
Dynamics of Atmospheres and Oceans 47:114–
137, doi:10.1016/j.dynatmoce.2008.10.007.
Hurlburt, H.E., E.J. Metzger, J. Sprintall,
S.N. Riedlinger, R.A. Arnone, T. Shinoda,
and X. Xu. 2011. Circulation in the
Philippine Archipelago simulated by 1/12°
and 1/25° global HYCOM and EAS NCOM.
Oceanography 24(1):28–47.
Jackson, C.R., Y. Arvelyna, and I. Asanuma.
2011. High-frequency nonlinear
internal waves around the Philippines.
Oceanography 24(1):90–99.
Jones, B.H., C.M. Lee, G. Toro-Farmer, E.S. Boss,
M.C. Gregg, and C.L. Villanoy. 2011.
Tidally driven exchange in an archipelago
strait: Biological and optical responses.
Oceanography 24(1):142–155.
May, P.W., J.D. Doyle, J.D. Pullen, and L.T. David.
2011. Two-way coupled atmosphere-ocean
modeling of the PhilEx Intensive Observational
Periods. Oceanography 24(1):48–57.
Metzger, E.J., and H.E. Hurlburt. 1996. Coupled
dynamics of the South China Sea, the Sulu Sea,
and the Pacific Ocean. Journal of Geophysical
Research 101:12,331–12,352.
Metzger, E.J., and H.E. Hurlburt. 2001. The nondeterministic nature of Kuroshio penetration and
eddy shedding in the South China Sea. Journal
of Physical Oceanography 31:1,712–1,732.
Nitani, H. 1972. Beginning of the Kuroshio.
Pp. 129–163 in Kuroshio: Its Physical Aspects.
H. Stommel and K. Yoshida, eds, University of
Tokyo Press.
Ohlmann, J.C. 2011. Drifter observations of
small-scale flows in the Philippine Archipelago.
Oceanography 24(1):122–129.
Pullen, J., J.D. Doyle, P. May, C. Chavanne,
P. Flament, and R.A. Arnone. 2008. Monsoon
surges trigger oceanic eddy formation and
propagation in the lee of the Philippine Islands.
Geophysical Research Letters, 35, L07604,
doi:10.1029/2007GL033109.
Pullen, J.D., A.L. Gordon, J. Sprintall,
C.M. Lee, M.H. Alford, J.D. Doyle, and
P.W. May. 2011. Atmospheric and oceanic
processes in the vicinity of an island strait.
Oceanography 24(1):112–121.
Qu, T., and R. Lukas. 2003. The bifurcation of the
North Equatorial Current in the Pacific. Journal
of Physical Oceanography 33(1):5–18.
Qu, T., Y. Du, and H. Sasaki. 2006. South China Sea
throughflow: A heat and freshwater conveyor.
Geophysical Research Letters, 33, L23617,
doi:10.1029/2006GL028350.
Quadfasel, D., H. Kudrass, and A. Frische. 1990.
Deep-water renewal by turbidity currents in the
Sulu Sea. Nature 348:320–322.
Qiu, B., and R. Lukas. 1996. Seasonal and interannual variability of the North Equatorial Current,
the Mindanao Current and the Kuroshio
along the Pacific western boundary. Journal of
Geophysical Research 101:12,315–12,330.
Rypina, I.I., L.J. Pratt, J. Pullen, J. Levin, and
A.L. Gordon. 2010. Chaotic Advection in an
Archipelago. Journal of Physical Oceanography
40:1,988–2,006, doi:10.1175/2010JPO4336.1.
Smith, W.H.F., and D.T. Sandwell. 1997. Global
seafloor topography from satellite altimetry and
ship depth soundings. Science 277:1,957–1,962,
http://topex.ucsd.edu/marine_topo.
Tessler, Z.D., A.L. Gordon, L.J. Pratt, and
J. Sprintall. 2010. Transport and dynamics
of the Panay sill overflow in the Philippine
seas. Journal of Physical Oceanography,
doi:10.1175/2010JPO4395.1.
Toole, J., R. Millard, Z. Wang, and S. Pu. 1990.
Observations of the Pacific North Equatorial
Current bifurcation at the Philippine coast.
Journal of Physical Oceanography 20:307–318.
Villanoy, C.L., O.C. Cabrera, A. Yñiguez,
M. Camoying, A. de Guzman, L.T. David,
and P. Flament. 2011. Monsoon-driven
coastal upwelling off Zamboanga Peninsula,
Philippines. Oceanography 24(1):156–165.
Oceanography
| March 2011
27