You are on page 1of 33

Progress in

Oceanography
Progress in Oceanography 69 (2006) 285–317
www.elsevier.com/locate/pocean

Primary production in the eastern tropical Pacific: A review


J. Timothy Pennington a, Kevin L. Mahoney a, Victor S. Kuwahara a,
Dorota D. Kolber a, Ruth Calienes b, Francisco P. Chavez a,*
a
Monterey Bay Aquarium Research Institute, P.O. Box 628, 7700 Sandholdt Road, Moss Landing, CA 95039-0628, United States
b
Instituto del Mar del Perú, Esquina de Gamarra y General Valle S/N Chucuito, Callao, Peru

Abstract

The eastern tropical Pacific includes 28 million km2 of ocean between 23.5N and S and Central/South America and
140W, and contains the eastern and equatorial branches of the north and South Pacific subtropical gyres plus two equa-
torial and two coastal countercurrents. Spatial patterns of primary production are in general determined by supply of
macronutrients (nitrate, phosphate) from below the thermocline. Where the thermocline is shallow and intersects the
lighted euphotic zone, biological production is enhanced. In the eastern tropical Pacific thermocline depth is controlled
by three interrelated processes: a basin-scale east/west thermocline tilt, a basin-scale thermocline shoaling at the gyre mar-
gins, and local wind-driven upwelling. These processes regulate supply of nutrient-rich subsurface waters to the euphotic
zone, and on their basis we have divided the eastern tropical Pacific into seven main regions. Primary production and its
physical and chemical controls are described for each.
Enhanced rates of macronutrient supply maintains levels of primary production in the eastern tropical Pacific above
those of the oligotrophic subtropical gyres to the north and south. On the other hand lack of the micronutrient iron
limits phytoplankton growth (and nitrogen fixation) over large portions of the open-ocean eastern tropical Pacific,
depressing rates of primary production and resulting in the so-called high nitrate-low chlorophyll condition. Very high
rates of primary production can occur in those coastal areas where both macronutrients and iron are supplied in abun-
dance to surface waters. In these eutrophic coastal areas large phytoplankton cells dominate; conversely, in the open-
ocean small cells are dominant. In a ‘shadow zone’ between the subtropical gyres with limited subsurface ventilation,
enough production sinks and decays to produce anoxic and denitrified waters which spread beneath very large parts
of the eastern tropical Pacific.
Seasonal cycles are weak over much of the open-ocean eastern tropical Pacific, although several eutrophic coastal areas
do exhibit substantial seasonality. The ENSO fluctuation, however, is an exceedingly important source of interannual var-
iability in this region. El Niño in general results in a depressed thermocline and thus reduced rates of macronutrient supply
and primary production. The multi-decadal PDO is likely also an important source of variability, with the ‘El Viejo’ phase
of the PDO resulting in warmer and lower nutrient and productivity conditions similar to El Niño.
On average the eastern tropical Pacific is moderately productive and, relative to Pacific and global means, its produc-
tivity and area are roughly equivalent. For example, it occupies about 18% of the Pacific Ocean by area and accounts for
22–23% of its productivity. Similarly, it occupies about 9% of the global ocean and accounts for 10% of its productivity.

*
Corresponding author. Tel.: +1 831 775 1709.
E-mail address: chfr@mbari.org (F.P. Chavez).
URL: http://www.mbari.org/staff/chfr (F.P. Chavez).

0079-6611/$ - see front matter  2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pocean.2006.03.012
286 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

While representative, these average values obscure very substantial spatial and temporal variability that characterizes the
dynamics of this tropical ocean.
 2006 Elsevier Ltd. All rights reserved.

Keywords: Eastern tropical Pacific; Upwelling; Thermocline; Nutricline; Nitrate; Iron; Phytoplankton; Primary production; Chlorophyll

1. Introduction

1.1. Physical and biological coupling

Patterns of primary production in the world ocean reflect patterns of thermocline topography. The ther-
mocline is that portion of the water column where temperature decreases maximally with increasing depth
(reviewed by Fiedler and Talley, 2006). This thermocline/productivity linkage occurs because the thermo-
cline almost invariably coincides with the nutricline, defined as that portion of the water column where phy-
toplankton nutrients increase maximally with depth (Barber and Chavez, 1983; Barber and Chavez, 1991;
Fiedler and Talley, 2006). Areas of high primary productivity occur where nutrient supply to the euphotic
zone is enhanced by mixing and a shallow thermocline (Wyrtki, 1966; Guillén, 1966; Calienes and Guillén,
1981; Guillén and Calienes, 1981a; Barber and Chavez, 1983; Turk et al., 2001; McClain et al., 2002).
Because temperature is much easier to measure than nutrients, both from ships and satellites, variations
in thermocline depth and even surface temperature are often associated with variations in phytoplankton
production.
Two patterns of enhanced nutrient supply occur in the ocean: (1) At high latitudes, typically in winter, deep
mixing erodes the thermocline downwards while nutrients are mixed upwards. In these areas phytoplankton
often bloom after the winter mixing season when stratification sets in and sunlight in the euphotic zone
becomes sufficient; (2) At low latitudes, dynamics of the wind-driven surface flow can cause the thermocline
to ‘lift’ or ‘shoal’ towards or into the euphotic zone (reviewed by Barber and Smith, 1981; Barber and Chavez,
1983). The supply of nutrients to phytoplankton is consequently increased, and rates of primary production
are elevated. In general, winter mixing and thermocline erosion dominate nutrient supply processes at high
latitudes, whereas at low latitudes, such as in the eastern tropical Pacific where seasonality is often weak, ther-
mocline shoaling and upwelling become the main supply processes (Chavez and Toggweiler, 1995). Within the
subtropical gyres where neither process operates effectively, the thermocline and nutricline are stable and deep
(200–300 m) and phytoplankton in the euphotic zone remain nutrient-impoverished. These biological/chemi-
cal/physical relationships have been long-recognized and remain an important paradigm in biological ocean-
ography (Sverdrup et al., 1942; Sverdrup, 1955; Brandhorst, 1958). Fleming (1957) developed a global map of
primary production based on upwelling, mixing and depth of the permanent thermocline (Fig. 1A) that is sim-
ilar to maps created today from satellite imaging of phytoplankton color (Fig. 1B; Behrenfeld and Falkowski,
1997a,b; McGowan, 2004), and mathematical models today predict patterns of primary production based in
part on the underlying thermocline dynamics.

1.2. Eastern tropical Pacific heterogeneity

The eastern tropical Pacific is herein defined as reaching from the coast of Central and South America to
140W, about halfway across the Pacific, and between the Tropics of Cancer and Capricorn at 23.5N and S,
respectively. As such, this ‘box’ includes 28 million square kilometers, about three times the area of the United
States, and accounts for about 23% of Pacific and 10% of global oceanic primary production (Table 1).
Because thermocline depth controls nutrient supply and thus biological production, it becomes important
to understand the main processes which control eastern tropical Pacific thermocline depth, all of which are
related to the wind. In the north [or south] Pacific subtropical gyre, the wind blows clockwise [or counterclock-
wise] around the basin to drive the surface flow of the gyral circulation (see Fig. 3). First, because these cur-
rents export water and accumulated heat from the eastern to the western tropical Pacific, the thermocline tilts
down towards the western side of the basin and is much shallower in the east (Barber and Chavez, 1983;
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 287

Fig. 1. Historical and modern analyses of primary productivity of the world oceans (g C m2 y1): (A) Fleming’s (1957) predictions based
on thermocline depth; and (B) results of Behrenfeld and Falkowski’s (1997b) VGP Model initialized with SeaWiFS chlorophyll and sea
surface temperature.

Tomczak and Godfrey, 1994). Second, the currents interact with the rotation of the Earth to lift the thermo-
cline at the gyre edges and depress it in the gyre centers, shoaling the thermocline over substantial portions of
the eastern tropical Pacific (see Fig. 2C). Third, in some areas, local winds drive additional thermocline shoal-
ing via wind-driven upwelling. The first two of these processes operate on basin-scales while the third is
locally-driven, but all three vary spatially within the eastern tropical Pacific. Fig. 2 shows the nutrient
response.
As a consequence biological productivity is heterogeneous and cannot be understood without an under-
standing of the region’s atmospheric forcing, circulation and hydrography. These physical dynamics are
briefly introduced in this paper (see Section 3.2, below, for further discussion of thermocline topography)
and are reviewed in detail elsewhere in this volume as follows. The NE and SE trade winds, and the
Inter-Tropical Convergence Zone (ITCZ) between them, are reviewed by Amador et al. (2006). The eastern
boundary currents (Peru and California), their subsurface countercurrents, and the equatorial currents
(North and South Equatorial Currents, the Equatorial Undercurrent and the North Equatorial Countercur-
rent; hereafter NEC, SEC, EUC and NECC) are reviewed by Kessler (2002, 2006). A poorly ventilated,
sub-thermocline ‘shadow zone’ develops between the American coast and the NEC and SEC, resulting in
the largest zone of low oxygen in the world’s ocean (reviewed by Fiedler and Talley, 2006).
288 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Table 1
Primary production depth-integrated over the euphotic zone for the regions (bold) and subregions (italics) shown in Fig. 3
Integrated primary Ship-collected VGP model Area
production
Regional % of % of Regional % of % of % of Regional % of
(Gt C yr1) Pacific Global (Gt C yr1) ETP Pacific Global (·1012 m2) ETP
Coastal boundary 0.9 3.5 1.6 0.9 16.1 3.5 1.6 2.0 6.4
Peruvian coastal upwelling 0.5 1.9 0.8 0.3 5.0 1.1 0.5 0.4 1.1
Tehuantapec 0.1 0.2 0.1 0.1 1.1 0.2 0.1 0.2 0.5
Papagayo 0.1 0.3 0.1 0.2 3.6 0.8 0.4 0.5 1.7
Panama 0.03 0.1 0.05 0.04 0.7 0.2 0.1 0.1 0.3
Eastern boundary current 0.7 3.0 1.3 1.2 21.4 4.7 2.1 5.4 17.2
Equatorial upwelling 0.9 3.6 1.6 0.7 13.6 3.0 1.3 3.8 11.9
Galapagos 0.05 0.2 0.1 0.02 0.5 0.1 0.0 0.1 0.2
South Equatorial Current 1.3 5.1 2.3 1.0 17.5 3.8 1.7 6.2 19.7
South Pacific Subtropical Gyre 1.1 4.5 2.0 0.6 11.4 2.5 1.1 6.4 20.1
North Equatorial Countercurrent 0.4 1.5 0.7 0.5 8.8 1.9 0.9 3.2 9.9
10N Thermocline Ridge 0.1 0.6 0.2 0.2 4.3 0.9 0.4 1.6 5.0
North Equatorial Current 0.2 1.0 0.4 0.4 6.9 1.5 0.7 3.1 9.8
Totals
Eastern Tropical Pacific 6 23 10 5 – 22 10 32 –
Entire Pacific – – – 25 – – 45 181 –
Global – – – 56 – – – 340 –
Production is expressed as metric gigatons (1015 g) carbon per year.
The ship-collected and VGP Model data are described in caption to Table 2; the two methods yield similar eastern tropical Pacific (ETP)
productivity estimates with regional differences. Regional ship-collected percentages of Pacific and global primary production are cal-
culated relative to VGP Model output.

Continental geometry is also important. It positions the confluence of the NE and SE trade winds (the
‘Inter-Tropical Convergence Zone’, or ITCZ) and the NECC not at the geographic equator, but at about
10N, so that both subtropical gyres are shifted to the north (Philander et al., 1996). As a consequence, the
SE trade winds and the SEC penetrate the northern hemisphere to about 5N. Divergence of Ekman transport
on the equator then creates the famed open-ocean equatorial upwelling system (reviewed by Kessler, 2006).
Also as a consequence of continental geometry, the eastern Pacific warm pool above the thermocline lies
entirely north of the equator, and the below-thermocline oxygen minimum is much more pronounced to
the north (Fiedler and Talley, 2006). The above currents, thermocline dynamics and coastal features all affect
euphotic zone nutrient supply and hence biological productivity within the eastern tropical Pacific.

1.3. Purpose and outline

This paper provides an overview of phytoplankton production and its controls in the eastern tropical
Pacific. The purpose is to review previous work and to provide region-specific estimates of primary pro-
ductivity, chlorophyll biomass, and nutrient concentrations. The discussion is based on a compilation of
ship-collected physical, chemical and biological data which are compared to satellite chlorophyll and mod-
eled primary production estimates (SeaWiFS and the VGP Model of Behrenfeld and Falkowski, 1997b,
respectively; analyses detailed in Appendix). We emphasize spatial and temporal patterns of primary pro-
duction and their relationship to atmospheric forcing (Amador et al., 2006), hydrography (Fiedler and
Talley, 2006) and circulation (Kessler, 2006). The focus is on near-surface waters and the sunlit euphotic
zone.
Based on the spatial heterogeneity described above, we have divided the eastern tropical Pacific into seven
biogeographic regions (Fig. 4). These are (1) an Eastern Margin, (2) an Equatorial Upwelling region (EU), (3–
4) two southern and (5–7) three northern hemisphere open ocean regions. The Eastern Margin and EU regions
have been additionally subdivided where particular features require (see Fig. 4). Below we define the regions,
introduce their hydrography and nutrient dynamics, and describe the spatial and seasonal patterns of phyto-
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 289

Fig. 2. Nitrate concentrations from the World Ocean Database showing basin-scale nutricline shoaling: (A) Global 100 m nitrate. The
white lines indicate the positions of panels B and C. 100 m nitrate is near zero in the subtropical gyres and at the western boundaries, but
higher at the eastern boundaries, at high latitudes, and in the eastern tropical Pacific; (B) Vertical section of nitrate along the equator. The
equatorial thermocline tilt causes near-surface nitrate to be higher in the east; (C) Vertical section of nitrate on 110W (vertical white line
on panel A) showing basin-scale nutricline shoaling across the equator associated with the subtropical gyre circulation, equatorial
upwelling, and the ridge of the nutricline at 10N .

plankton biomass and productivity based on ship-collected data. Because the subtropical gyres flow towards
the equator near the continents and then westward across the Pacific, the regions are presented in that order,
beginning in the northern hemisphere. In a final ‘synthesis’ section we reconnect the regions, compare ship-
collected data to satellite and model results, and highlight processes that regulate primary production in
the eastern tropical Pacific. During assembly of this paper it became clear that many aspects of biological pro-
duction in the region are poorly understood and can only be sketched in the broadest outline.
290
Table 2
Means for biological, physical and chemical properties for the regions (bold) and subregions (italics) shown in Fig. 3
A. Surface Chlorophyll Primary production Temperature (C) Salinity (PSS) NO3 (lM) PO4 (lM) SIO2 (lM)
Ship-collected SeaWIFS Ship-collected
(mg m3) (mg C m3 d1)
Coastal boundary 1.63 1.27 39.5 22.0 34.2 5.7 1.0 11.6
Peruvian coastal upwelling 2.55 2.27 142.8 19.4 35.0 8.0 1.2 9.2
Tehuantapec 0.33 0.89 13.7 27.2 33.9 2.9 0.2 10.1
Papagayo 0.26 1.05 12.7 26.7 33.7 3.5 0.5 4.0

J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317


Panama 0.31 1.06 22.5 25.7 31.2 2.1 0.6 29.4
Eastern boundary current 0.26 0.3 12.2 23.7 34.5 3.6 0.5 4.3
Equatorial upwelling 0.21 0.22 13.2 24.3 34.7 6.0 0.7 4.9
Galapagos 0.65 0.63 92.0 22.3 34.5 4.9 0.6 4.8
South Equatorial Current 0.18 0.16 11.7 25.5 35.4 5.7 0.7 3.7
South Pacific Subtropical Gyre 0.1 0.08 8.4 24.6 36.1 1.3 0.4 1.6
North Equatorial Countercurrent 0.15 0.18 7.4 26.7 34.0 1.5 0.3 2.5
10N Thermocline Ridge 0.14 0.18 4.9 27.4 33.7 0.3 0.3 1.9
North Equatorial Current 0.09 0.1 2.6 24.4 34.3 0.1 0.2 3.7
B. Euphotic zone/0–100 m Chlorophyll Primary production Temperature (C) Salinity (PSS) NO3 (lM) PO4 (lM) SIO2 (lM)
Ship-collected Ship-collected VGP model
(mg m3) (mg C m3 d1)
Coastal boundary 59 1179 1183 17.7 34.7 14.0 1.7 19.7
Peruvian coastal upwelling 111 3580 2087 16.5 35.0 16.2 1.9 17.2
Tehuantapec 30 802 928 19.5 34.5 18.3 1.6 42.5
Papagayo 22 365 997 18.1 34.5 20.3 1.6 16.8
Panama 25 691 1028 20.4 33.7 9.9 1.2 32.1
Eastern boundary current 22 376 589 19.2 34.7 12.0 1.1 9.8
Equatorial upwelling 25 642 537 20.7 34.9 11.5 1.0 9.6
Galapagos 42 1887 979 18.0 35.0 14.0 1.3 11.5
South Equatorial Current 19 564 420 23.0 35.4 8.0 0.9 5.6
South Pacific Subtropical Gyre 14 482 268 23.8 36.1 1.1 0.4 1.5
North Equatorial Countercurrent 19 327 417 22.9 34.4 7.5 0.8 6.5
10N Thermocline Ridge 17 236 398 21.9 34.2 10.3 0.9 7.7
North Equatorial Current 14 210 334 22.2 34.3 1.1 0.4 3.8
Ship-collected and SeaWIFS data span the periods 1980–2005 and September 1997–May 2005, respectively: (A) Surface values are 0–5 m ship-collected data and SeaWIFS chlorophyll
also as in Fig. 3; (B) Values depth-integrated across the euphotic zone are ship-collected data (chlorophyll, primary production) and primary production estimated with 8-day estimates
using the Behrenfeld and Falkowski (1997b) Vertically Generalized Productivity (VGP) Model using SST, irradiance and SeaWIFS chlorophyll data for the period January 1998–
December 2002. Ship-collected (0–100 m) mean values are included for comparison with the integrated values of the biological properties.
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 291

2. Biogeographic regions

2.1. Eastern margin (23S–23N, 0–1000 km from shore)

Eastern boundary currents are equatorward-flowing portions of the subtropical gyres that transition
between the coast in the east and central gyre waters in the west, and are underlain by subsurface countercur-
rents that flow poleward along the continental slope. Eastern boundary currents are slow and broad – on the
order of 1000 km across. Here we subdivide the eastern margin into a Coastal Boundary from the coast to
250 km offshore and an Eastern Boundary Current region from 250 to 1000 km from shore (see Fig. 4; Tables
1 and 2). In the southern hemisphere portion of the eastern tropical Pacific, a ‘traditional’ eastern boundary
current region is formed by the Peru Current. In the northern hemisphere, the California Current leaves the
boundary off Baja California at the northern edge of the eastern tropical Pacific. As Fig. 3 shows, the waters
of the eastern Pacific warm pool off southern Mexico and Central America are not penetrated by the subtrop-
ical gyre circulation. As described below, this creates significant north–south asymmetries in primary
productivity.
The Coastal Boundary (Fig. 5A) is characterized by high standing stocks of phytoplankton and enhanced
primary productivity rates (Strickland et al., 1969; Chavez and Barber, 1987; Chavez, 1995; Fiedler, 2002a).
The high productivity is supported by nutrients brought into the euphotic zone as described above by thermo-
cline shoaling and local wind-driven upwelling (see also Section 3.2, below; Barber and Chavez, 1983). The
upwelled water often derives from the subsurface countercurrent. Regional differences in productivity are evi-
dent in south to north vertical sections of chlorophyll (Fig. 5A and B). The southern hemisphere Peruvian
Coastal Upwelling (PCU) region is by far the richest in terms of chlorophyll concentration and this biomass
extends out to 1000 km from shore. In contrast, the Gulfs of Panama, Papagayo and Tehuantepec are impres-
sive in satellite images of chlorophyll (Fig. 4), but are barely noticeable in averaged in situ chlorophyll data
(Fig. 5A). Oceanographers have known of the high concentrations of chlorophyll off the coast of Peru for

Fig. 3. Average sea surface temperature, wind stress and overview of major surface currents in the eastern tropical Pacific. The eastern
tropical Pacific is defined by the equatorial confluence of the north and South Pacific subtropical gyres and the eastern Pacific warm pool.
Heavy black lines denote major currents, thin black lines are a wind stress climatology (1999–2003 FSU/COAPS Tropical Pacific
pseudostress), and colors are a six-year Aqua-MODIS sea surface temperature climatology (1997–2003, 9 km, Level-3 Standard Mapped
Image product).
292 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Fig. 4. Eastern tropical Pacific SeaWIFS chlorophyll and locations of ship-collected measurements (chlorophyll in color, mg Chl m3,
September 1997–May 2005 average; stations as grey dots). We subdivided the area into seven fairly natural regions and treat their
oceanography individually (white boxes, regions defined in text). The Eastern Boundary Current and Equatorial Upwelling regions
additionally contain subregions that have also been described separately (Tehuantepec, Papagayo and Panama, Peruvian Coastal
Upwelling, Galapagos.

decades (e.g., Ryther, 1969), but the apparent richness off Central America was not appreciated until ocean
color remote sensing became available (Fiedler, 1994; McClain et al., 2002). These coastal subregions are con-
sidered separately below because of their proximity to humans and importance to fisheries.
The Eastern Boundary Current region (Fig. 5B) is in general cool and contains some surface nutrients due
to (1) input from high latitudes, (2) gyre-scale thermocline shoaling, and (3) exchange with the coastal upwell-
ing region. As a consequence, the offshore Eastern Boundary is characterized by moderate standing stocks of
phytoplankton and rates of primary productivity (Strickland et al., 1969; Chavez and Barber, 1987; Chavez,
1995; Fiedler, 2002a). An exception is the eastern Pacific warm pool off Central America, which is not pene-
trated by the subtropical gyre circulation (Fig. 3). The mean primary production for the Eastern Boundary
Current region is 376 mg C m2 d1.

2.1.1. Gulf of Tehuantepec, Gulf of Papagayo and the Costa Rica Dome, and Gulf of Panama
The Gulfs of Tehuantepec, Papagayo, and Panama are seen in satellite chlorophyll images as areas of
enhanced biological production (Figs. 4 and 6). This is a result of nutrient supply by wind-driven upwelling
and near-suface mixing. During boreal winter, strong but intermittent wind blows across Central America
from the Gulf of Mexico and the Caribbean into each of the three Gulf regions (see reviews in Amador
et al., 2006; Willett et al., 2006; Kessler, 2006). Upon reaching the Pacific, these offshore-directed wind jets
fan out, forming velocity gradients perpendicular to the jet axes. Such velocity gradients produce frictional
gradients on the sea surface known as ‘wind stress curl’ (Fig. 6; also Fig. 1 of Kessler, 2006). In the northern
hemisphere, positive curl (decreasing velocity downwind to the left) causes surface divergence and upwelling,
while negative curl (decreasing velocity downwind to the right) causes surface convergence and downwelling
(see Fig. 4 of Willett et al., 2006). This process occurs wherever the surface ocean is exposed to wind gradients
and is called ‘Ekman pumping’. Where surface currents diverge, the pumping is upward and the thermocline
shoals; where the currents converge, Ekman pumping is downward and the thermocline is depressed (Kessler,
2006; see Figs. 6 and 21 in Amador et al., 2006). The curl and pumping typically develops eddies to either side
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 293

Fig. 5. North to south sections of ship-collected chlorophyll samples (station locations on Fig. 4): (A) Near-surface chlorophyll within the
Coastal Boundary (250 km of the coast) is highest in the Peruvian Coastal Upwelling subregion south of the equator; (B) Chlorophyll
within the Eastern Boundary Current region 250–1000 km of shore follows a similar pattern but is lower; (C) Surface chlorophyll between
110 and 120W is highest around the equator and in a subsurface maximum 60–90 m deep; (D) Chlorophyll in the central Pacific, 130–
140W is similar to (C), but is more symmetrically distributed around the equator.

Fig. 6. Physical and biological dynamics associated with the Tehuantepec, Papagayo and Panama wind jets in the Eastern Pacific Warm
Pool: (A) Wind jets from the Gulf of Mexico and the Caribbean blow through mountain passes, creating wind stress curl and upwelling
(upwelling-favorable positive curl is pink, downwelling-favorable negative curl is blue; 4-year QuickSCAT average from Chelton et al.,
2004); (B) Wind stress (arrows) and SST (colors), again from Chelton et al. (2004); (C) QuickSCAT wind velocity (arrows) and SeaWIFS
chlorophyll (colors, 17–24 January 2000, from McClain et al., 2002).
294 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

of the wind jet, ranging from 100 to 500 km in diameter (Willett et al., 2006). Thus each wind jet creates (1)
offshore transport and strong mixing under the jet axis, (2) a counterclockwise-rotating upwelling eddy to the
left of the jet axis, and (3) a clockwise-rotating downwelling eddy to the right of the jet axis (Fig. 6A). Bio-
logical productivity is coupled to this wind forcing (Fig. 6) with areas of high chlorophyll associated with mix-
ing, upwelling, and eddy transport (Willett et al., 2006). Although the effects of the gulf region wind jets,
mixing, eddies and the Costa Rica Dome have not been studied in detail, all these features enhance euphotic
zone nutrient supply and biological productivity in what otherwise would be a near-oligotrophic eastern Paci-
fic warm pool.
Satellite observations show surface chlorophyll concentrations reaching as high as 2 mg Chl m3 along
the Central American coast (Lluch-Cota et al., 1997; McClain et al., 2002). This is a considerable amount
of surface biomass considering the stratified warm-pool conditions of the region (Fiedler and Talley, 2006).
These regions have very high SSTs (Table 2A), but the shallow thermocline leads to much cooler depth-aver-
aged temperatures (Table 2B; Fig. 7). However, ship-collected data (Fig. 8, Table 2A) from the gulf regions
suggest that the chlorophyll concentration is considerably lower than observed with SeaWiFS. Two possible
explanations are (1) the satellite estimates may be contaminated by in-water or atmospheric substances
(CDOM or dust), or (2) because the ship-collected data were collected irregularly over many years, they
may be spatially or temporally biased, missing important blooms or periods.
The wind jets are highly seasonal in all three gulf regions with strong cross-isthmus winds during the boreal
winter (December through April), which is the Central American dry season. In general, the Tehuantepec jet

Fig. 7. Mean profiles (0–200 m) of physical, nutrient and chlorophyll values from oceanographic stations (see Fig. 4) in the coastal regions
and subregions.
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 295

Fig. 8. Comparison between monthly mean ship-collected and SeaWiFS surface chlorophyll for eastern tropical Pacific regions and
subregions. Chlorophyll scales for the different regions vary. The two data sources are in general agreement, except that SeaWiFS
overestimates chlorophyll in the Tehuantepec, Papagayo and Panama subregions.

appears first in early winter and is then followed by the Papagayo and Panama jets (Chelton et al., 2000; Chel-
ton et al., 2004). The jets do not occur after May. The ship-collected and satellite chlorophyll results for Tehu-
antepec and Panama are in phase with the respective winds, with highest values during winter (Fig. 8B and D),
and with Tehuantepec having a longer and earlier maximum in chlorophyll. Particular features of each of the
Gulf regions are reviewed below.
At the Isthmus of Tehuantepec, the average 0–100 m NO3 concentration (Table 2B) is very high at
18.3 lmol l1, second only to the Gulf of Papagayo at 20.3 lmol l1, indicating that a large supply of nutrients
is available near the surface (Fig. 7). Mean primary productivity levels are 802 mg C m2 day1 (Table 2B),
and SeaWiFS chlorophyll data show an area of enhanced chlorophyll centered near 15N, 95W (Figs. 4
and 6C), along the wind jet axis. Downwelling occurs on the right flank of the wind jet due to negative curl
northwest of the Gulf of Tehuantepec (blue shading in Figs. 6A and 9A, D), contributing to a large thermo-
cline depression termed the ‘Tehuantepec Bowl’ by Kessler (2006). Satellite chlorophyll levels are lower in this
area (Fig. 6C).
The Gulf of Papagayo lies southeast of Tehuantepec, between Nicaragua and Costa Rica (Figs. 4 and 6).
The Papagayo wind jet affects nutrient supply as described above for Tehuantepec, but also interacts and
merges with a permanent offshore feature called the Costa Rica Dome (Wyrtki, 1964). The Gulf of Papagayo
and the Costa Rica Dome are treated together here even though they result, at least in part, from separate
dynamic processes. The Costa Rica Dome was first described by Cromwell (1958) and represents the eastern
terminus of the 10N thermocline ridge (Figs. 9C and D; see Section 2.7, below). It appears to be formed in
winter by the wind stress curl (Papagayo wind jet and westerly remnants of the SE trades) and is maintained
during summer by the left-handed turning of the NECC as it reaches the Central American coast to become
296 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Fig. 9. Climatological wind-induced upwelling/downwelling at the base of the surface Ekman layer in the eastern tropical Pacific.
Upwelling is positive and downwelling is negative. Computed from the Comprehensive Ocean–Atmosphere Data Set monthly wind stress
data, http://www.ocgy.ubc.ca/projects/clim.pred/Upwell/ (Xie and Hsieh, 1995). January (A), April (B), July (C) and October (D) were
chosen to emphasize seasonal differences.

the Costa Rica Coastal Current (Fig. 9). These physics result in an offshore doming of the thermocline and
enhanced nutrient supply to the euphotic zone. The dome is centered at 9N, 90W, with a diameter ranging
between 100 and 900 km (Hoffmann et al., 1981; Fiedler, 2002a; Kessler, 2002; Kessler, 2006), and 0–100 m
macronutrient concentrations are high and similar to the Tehuantepec region (Table 2A and B). The Papa-
gayo region, including the Costa Rica Dome, has average integrated chlorophyll of 22 mg Chl m2 and pri-
mary productivity of 365 mg C m2 day1 (Table 2B). These values appear somewhat low and may appear
because, as defined in Fig. 4, the Papagayo subregion is larger and includes more lower productivity offshore
water than the Tehuantepec and Panama subregions. Autotrophic picoplankton dominate the production in
the Costa Rica Dome (Li et al., 1983). In contrast to the Tehuantepec and Panama subregions, Papagayo
maximum chlorophyll levels occur in late spring and summer instead of winter (Fig. 8B–D). This maximum
is apparently caused, not by cross-isthmus winds which are weak or absent in summer, but by the seasonal
shoaling of the 10N ridge and the Costa Rica Dome (Fig. 9C and D), which in summer loses its connection
to the coast and expands westward along the 10N ridge (Fiedler, 2002a; Kessler, 2006). During late fall/early
winter the Costa Rica Dome thermocline is at its maximum depth and Papagayo subregion chlorophyll values
are at their lowest (Fig. 8C).
The Gulf of Panama wind jet is similar to the Tehuantepec jet, producing offshore transport, mixing, and
curl-induced upwelling and downwelling eddies as described above (Fig. 6A). The Panama region has the low-
est 0–100 m salinity (33.7) and the highest 0–100 m silicate (32.1 lmol l1) of the entire eastern tropical Pacific
(Table 2B), implying that this region receives more terrestrial runoff than the other two gulf regions. The Gulf
of Panama subregion produces an average 25 mg chlorophyll m2 and 691 mg C m2 day1 (Table 2B).

2.1.2. Peruvian coastal upwelling (4–15S; 0–250 km offshore)


The Peruvian Coastal Upwelling (PCU) region has been relatively well studied, primarily because it sup-
ports the largest-tonnage fishery in the world, yielding 6–12 million metric tons of anchoveta annually
(FAO, 1993). Oceanographers have known of the high concentrations of chlorophyll off the coast of Peru
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 297

for decades (Ryther, 1969), and links between oceanographic processes and fisheries fluctuations have been
investigated since the 1960s (Guillén, 1966; Guillén and de Rodán, 1968; Guillén et al., 1969; Zuta and Guil-
lén, 1970; Jordán, 1971). Because anchoveta are one of the few fish that feed on phytoplankton (Ryther, 1969;
Hutchings, 1992), they efficiently move energy up the food chain but are directly and dramatically impacted by
changes in primary production.
The PCU region is embedded in the Coastal Boundary region of the Peru Current and is defined by low
surface temperature and high rates of primary production, most strongly developed from 4 to 16S and to
100 km offshore (Figs. 3 and 4). Coastal upwelling occurs where alongshore winds produce offshore Ekman
transport of surface water (Fig. 9), which, because of the continental boundary, must be replaced by subsur-
face water, typically from 60 m deep (Barber and Smith, 1981). The thermocline shoals to <40 m and very
often breaks the surface, resulting in very low sea surface temperatures for these latitudes (SST; see Fig. 7 of
Fiedler and Talley, 2006). The PCU region has both the lowest SST (19.42 C) and depth-integrated temper-
ature (16.51 C) of all the regions of the eastern tropical Pacific (Table 2). Upwelling proceeds across a wide
band of coastal ocean almost year-round because of Peru’s low latitude and persistent winds (Fig. 9; Chavez,
1995). Upwelling also occurs farther south off southern Peru and Chile, but is less intense.
Upwelling brings nutrients into the euphotic zone so that nearshore phytoplankton experience optimal
nutrient conditions. As discussed below, denitrification consumes nitrate in the upwelling source waters, so
once at the surface, nitrogen becomes exhausted before phosphate (Harrison et al., 1981; Guillén and Calienes,
1981a,b). Farther offshore, however, silicate has often been suggested to be the limiting macronutrient for dia-
toms (Dugdale, 1972; Guillén et al., 1977; Dugdale et al., 1979). Offshore, and particularly in regions of nar-
row continental shelves, PCU phytoplankton are also limited by the micronutrient iron (see Section 2.4 below;
Hutchins et al., 2002; Bruland et al., in press). Iron is probably supplied to the PCU region surface water via
interaction of the upwelling circulation with continental shelf sediments (see Johnson et al., 1999). Early
researchers noted that phytoplankton seem to be affected by the source of upwelled water (Strickland
et al., 1969; Barber et al., 1971), with less production occurring in water upwelled offshore (Barber et al.,
1971); this observation can probably now be explained by different iron concentrations in source waters.
Regardless of when and where PCU region nutrients are limiting, mean surface nitrate and phosphate are
the highest of all of the Table 2 regions (see also Figs. 7 and 10), and luxuriant phytoplankton growth is sup-
ported with the highest rates of depth-integrated primary productivity (3580 mg C m2 day1; Table 2B) and
the highest average surface chlorophyll concentration in the eastern tropical Pacific (2.55 mg Chl m3; Fig. 4
and Table 2A). Mean 0–50 m chlorophyll ranges from 1 to 3 mg Chl m3 (Fig. 5A). These high values are dri-
ven by ‘new’ upwelled nitrate rather than by nitrogen recycled in the euphotic zone (see Section 3.4, below, for
a discussion of new versus recycled nitrogen sources). Large centric diatoms dominate the biomass of the PCU
region (Chavez, 1987; Chavez and Barber, 1987; Chavez et al., 1996a; Rojas de Mendiola, 1981).
In spite of their proximity to the equator, PCU region phytoplankton exhibit surprisingly strong seasonal-
ity, with the highest levels of chlorophyll and primary production occurring during the austral summer and fall
(Fig. 8E; Calienes et al., 1985; Chavez, 1995). This seasonal cycle is perplexing given that the highest upwelling
rates occur earlier, during winter and spring (marginally visible in Fig. 9C). Low light may suppress produc-
tion in winter due to increased mixed layer depth or fog (Calienes et al., 1985; Chavez, 1995), or it could be
that strong offshore advection and mixing essentially dilute phytoplankton faster than it can grow during peak
winter and spring upwelling. Given that iron is rapidly depleted from newly upwelled waters (via drawdown,
scavenging and photo-oxidation), enhanced winter and spring upwelling may also result in a large HNLC
region offshore during these seasons (see Section 3.5 for further discussion of HNLC). Farther south off Chile,
the seasonal cycle of winds reverses, with strongest upwelling during summer, in phase with biological
production.
Satellite comparisons of primary production for the world’s major coastal upwelling regions suggest that
production off Peru is significantly greater than off California, but is similar or even lower than that in the
Benguela and Canary systems (Carr, 2002; Carr and Kearns, 2003). This is surprising given that fish produc-
tion for Peru is several fold that for the other coastal upwelling areas. We note that the in situ estimates of
production from Peru are almost twice the satellite estimates (Table 2).
A portion of PCU region phytoplankton production sinks to deeper layers either as ungrazed phyto-
plankton or feces. This material is ‘‘captured’’ in the poleward-flowing Peru Undercurrent where it decays.
298 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Fig. 10. Mean profiles (0–200 m) of physical, nutrients and chlorophyll values at stations in the oceanic regions (see Fig. 3).

The decay is mediated by bacteria and is termed ‘remineralization’ – the process consumes oxygen and con-
verts organic compounds back into inorganic nutrients. As the undercurrent flows poleward, the export of
organic carbon from the overlying euphotic zone produces a large oxygen minimum layer beneath the ther-
mocline (Fiedler and Talley, 2006). A similar anoxic layer forms beneath the eastern Pacific warm pool and
the Gulf subregions described above (Fig. 3). The sub-thermocline waters of the PCU and eastern Pacific
warm pool are poorly ventilated and their source waters are low in oxygen, preconditions for full consump-
tion of oxygen (see Section 3.3, below). Once the oxygen is depleted, anoxic bacteria use nitrate instead of
oxygen as a terminal electron acceptor in a process called ‘denitrification’ (Codiposti et al., 1986; Gruber
and Sarmiento, 1997). The rates of denitrification off Peru are among the highest in the world (Codiposti
et al., 1986). The resulting anoxic and nitrate-deficient (relative to phosphate) waters spread poleward
and westward, with important biogeochemical consequences (see Section 3.3). The upwelling of nitrate-defi-
cient water also indirectly increases the flux of CO2 from PCU region surface waters to the atmosphere
given that there is higher residual CO2, once nitrate is consumed by phytoplankton, relative to waters that
are not denitrified (Fig. 11A).

2.2. Equatorial upwelling (3N–3S, 90–140W)

Water masses of the eastern tropical Pacific were first documented at 140W by the Carnegie in 1929 (Sverd-
rup et al., 1942), and a maximum in biological activity was noted to be associated with the equatorial diver-
gence and upwelling (Graham, 1941; more recent observations by King and Demond, 1953; Hasle, 1959;
Blackburn et al., 1970). The EU region has received considerable modern study because of its impact on
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 299

Fig. 11. Global mean annual distributions of CO2 flux, N*, iron dust deposition and N2 fixation. The panels result from heavy
interpolation of few data and so are not precise representations of spatial pattern; white boxes enclose the eastern tropical Pacific: (A)
Equatorial upwelling brings high-CO2 waters to the surface where they warm and CO2 vents to the atmosphere (figure modified from
Takahashi et al., 1997); (B) Distribution of N* on the 26.5 isopycnal. Positive N* values indicate greater than Redfield ratios of nitrate to
phosphate; negative values indicate lower ratios (figure modified from Gruber, 2004); (C) Open ocean fallout of atmospheric iron dust
reflecting continental dust sources and wind (data from Ginoux et al., 2001); (D) N2-supported new production, as mol C m2 y1 (figure
modified from Lee et al., 2002).

the carbon cycle in relation to global warming and because of the equator’s role as an oceanographic ‘wave-
guide’ or ‘teleconnector’ during El Niño (Chavez and Barber, 1987; Murray et al., 1994; Murray et al., 1995;
Chavez et al., 1999; Chai et al., 2002). Below we introduce EU region hydrography and nutrient chemistry as
300 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

they relate to phytoplankton dynamics. The Galapagos Islands lie within the EU region at about 92W, and
discussion of waters surrounding the archipelago is also included here.
The South Equatorial Current (SEC) is the trade wind-driven, northern branch of the South Pacific sub-
tropical gyre which flows northwestwards and across the equator to about 5N, then westward across the east-
ern tropical Pacific (Fig. 5B of Kessler, 2006). An enormous, open-ocean upwelling system is embedded in this
current (see Kessler, 2006). This upwelling is created by the reversal of Coriolis acceleration to either side of
the equator: Ekman transport north [or south] of the equator is to the right [or left] of the westward compo-
nent of the SE trade winds (Fig. 9). This divergence of flow causes upwelling along the equator. The band of
cool surface water which results, from about 3N to 3S and across the eastern and central tropical Pacific
(Wyrtki, 1981; Fiedler and Talley, 2006), is called the equatorial ‘cold tongue’ (see Fig. 3) and is known
for fertilizing the euphotic zone with macronutrients (Chavez and Barber, 1987). Equatorial upwelling is
centered on the equator but is strongest in the central Pacific where the trade winds are most easterly
(Fig. 3; marginally visible in Fig. 9; see Kessler, 2006).
The SEC and the NEC export massive amounts of surface water and accumulated heat from the eastern
into the western Pacific (Wyrtki and Wenzel, 1984). The thermocline consequently tilts downwards to the west
(40 ! 150 m) and the shallow thermocline in the east, in combination with the southeasterly trade winds and
the merging of cool Peru current waters with the SEC, cause SST to be colder in the east and south of the
equator (Fig. 3; and Kessler, 2006). These EU region dynamics occur within the SEC and thus are part of
South Pacific subtropical gyre hydrography and circulation. At 3–5N, the transition between the EU region
and the NECC is marked by the equatorial front (see Section 2.5.1; Fig. 3).
Warm water accumulates in the western Pacific where it raises sea level by 30–50 cm relative to the east
(Wyrtki and Wenzel, 1984). The elevated sea level in turn produces an eastward-directed pressure gradient
that drives the Equatorial Undercurrent (EUC) back towards South America (Kessler, 2006). The EUC flows
within the equatorial thermocline and thus shoals from west to east until it approaches the Galapagos Islands
(Fig. 5B of Kessler, 2006), and is important biologically as the source of nutrients upwelled along the
equator.
Due to upwelling, EU region nitrate and phosphate concentrations are high relative to the surrounding oce-
anic regions (Figs. 2 and 10 Table 2A). As with the thermocline, the EU nutricline deepens towards the west
(Fig. 2B). EU region surface nitrate concentrations are maximal just south of the equator (Fig. 2C; Fiedler and
Talley, 2006), and surface nitrate is in general higher south rather than north of the equator (Fig. 2C) in the
cooler eastern SEC. EU region surface silicate concentration is moderate but iron concentrations are very low
due to lack of upwind dust sources (surface values near 5 lmol- and 1 nmol, respectively; Fig. 10; Coale et al.,
1996b). The EUC, however, has been suggested to transport iron-enriched waters from Indonesia into the
eastern Pacific (Wells et al., 1999). EUC waters are well oxygenated, but concentration decreases in the east;
rates of denitrification in sub-thermocline EU waters are nevetheless low (Gruber and Sarmiento, 1997). The
EUC also splits the eastern Pacific oxygen minimum and denitrified zone into its northern and southern hemi-
sphere tongues (Fig. 11B). Equatorial upwelling brings CO2 to the surface, and the EU region is the largest
global oceanic source of CO2 to the atmosphere, at about 0.7–1 · 1015 g C y1 (see Section 3.5.1; Fig. 11A;
Tans et al., 1990; Takahashi et al., 1997, 2002, 2003; Feely et al., 1999).
Relative to the oligotrophic subtropical gyres (surface chlorophyll <0.1 mg Chl m3), the concentration of
surface chlorophyll across the EU region is high (averaging 0.21 mg Chl m3, Table 2A). This band of high
biomass extends from the Galapagos to well beyond 140W (Fig. 4). Phytoplankton biomass is generally
maximal on or near the equator. Biomass is enhanced south of the equator in the far eastern Pacific
(<95W) due to influx of PCU region nutrients, phytoplankton and the cool/higher-nitrate SEC (Figs. 2, 4
and 5). Farther west, biomass is more symmetrically-distributed even though there is considerable asymmetry
in the physics (see Kessler, 2006). The subsurface chlorophyll maximum (0.2–0.3 mg Chl m3; Figs. 5D and
10C) is shallow as compared to that in the gyres and occurs in the upper thermocline (see Fiedler and Talley,
2006; Kessler, 2006). The EU region deep chlorophyll maximum slopes downwards from 30–40 m in the east
to 50–60 m in the west, and chlorophyll values are higher near the equator than to the north or south (Fig. 5).
EU region upwelling is strongest in the central Pacific and during the austral winter (>100W, Fig. 9); EU
phytoplankton however only exhibit weak seasonality, with a surface chlorophyll maximum also in winter
(0.25–0.30 mg Chl m3; Figs. 8 and 12).
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 301

Fig. 12. Seasonal fluctuation of surface chlorophyll based on 8-years of SeaWiFS data (1997–2005). January (A), April (B), July (C) and
October (D) were chosen to emphasize seasonal differences, as was the color scale (different than in Fig. 4).

Depth-integrated primary productivity of the EU region averages 642 mg C m2 day1, substantially
higher than in the other open-ocean regions, but only 18% of that of the PCU region on a per m2 basis (Table
2A). The EU region nevertheless contributes nearly twice as much to global primary production as does the
PCU region because of its size (Table 1; Chavez and Barber, 1987). Early studies of EU region primary pro-
ductivity suggested levels of about 400 mg C m2 d1 (Koblentz-Mishke et al., 1970). In the 1980s, contami-
nation of incubation bottles was suggested to have depressed early values and mean EU production was found
to be nearer 600 mg C m2 d1 (Chavez and Barber, 1987). In 1992–1993, results of in situ incubations per-
formed during JGOFS cruises indicated that on-deck (a.k.a. ‘simulated in situ’) incubations may have under-
estimated productivity by another 30–40% (Barber et al., 1996; Chavez et al., 1996), increasing EU production
to about 900 mg C m2 d1. This in situ/on-deck bias was not found, however, in coastal waters. There is still
some debate over this issue and the values in Table 2 are from on-deck incubation data that have not been
revised upwards.
Productive regions, particularly eutrophic coastal areas, are typically dominated by large phytoplankton
cells – in particular diatoms and dinoflagellates. The EU region is exceptional is this regard. While productive,
the region is dominated by the small solitary cells characteristic of the subtropical gyres, primarily prochloro-
phytes, Synechococcus spp., eukaryotic picoplankton, haptophytes, small autotrophic dinoflagellates and pen-
nate diatoms (Chavez, 1989; Chavez et al., 1991,1996; Landry and Kirchman, 2002).
Nutrient and phytoplankton interaction within the EU region remains a topic of considerable interest.
Given the extremely low nitrate uptake rates observed on the equator, if upwelling stopped it would still take
over 6 months to deplete surface nitrate (6–12 lM; Dugdale et al., 1992; McCarthy et al., 1996; Aufdenkampe
and Murray, 2002; Aufdenkampe et al., 2002). Upwelled macronutrients, particularly nitrate, are never fully
drawn down and chlorophyll and productivity do not reach the levels that seemingly ‘should’ be supported by
the available nutrients (Barber and Chavez, 1991). This high-nitrate low-chlorophyll condition (HNLC; see
302 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Section 3.4) develops because iron limitation restricts production to small cells that often use ammonia rather
than nitrate as a nitrogen source (e.g. Prochlorococcus; Rocap et al., 2003). Silicate may secondarily limit dia-
tom growth rate in the EU region (Dugdale et al., 1995, 2002; Dugdale and Wilkerson, 1998).

2.2.1. Galapagos Islands (1N–2S, 90–92W)


The Galapagos archipelago lies within the eastern margin of the EU region (Figs. 3 and 4; see also Fig. 3 in
Fiedler and Lavı́n, this volume). Nutrients are supplied to surface waters surrounding the Galapagos by (1)
equatorial upwelling as described above and (2) mixing and upwelling where the EUC impinges on the western
side of the islands (Chavez and Brusca, 1991). EUC water is mixed both to the surface and to depth where it
collides with the archipelago (see Fig. 2B), and eventually flows eastwards past the islands to merge with the
NECC to the north and the Peruvian Undercurrent to the south (Wyrtki, 1966; Kessler, 2006). As a result
near-surface temperatures are lower and macronutrient concentrations are higher than in the remainder of
the EU or any other open-ocean region (Table 2, Fig. 10). Galapagos surface waters are also iron-fertilized
by the islands themselves (Martin et al., 1994; Coale et al., 1998).
With abundant nutrients, waters surrounding the Galapagos support at least twice the phytoplankton bio-
mass and primary production as in the remainder of the EU or any of the open-ocean regions (Table 2B),
although the PCU region is still about twice as productive on a per m2 basis (Table 2B). Associated with this
high production near the Galapagos, denitrification can occur at very shallow depths, most notably at Darwin
Bay on the small northeastern island of Genovesa (Cline and Richards, 1972). Diatoms and other large phy-
toplankton cells are important contributors to Galapagos productivity as is typical of other eutrophic coastal
areas (Chavez and Smith, 1995).

2.3. South equatorial current (3–13S, 87/90–140W)

The EU region described above represents a surface divergence zone within the SEC, to either side of the
equator (Fig. 3). In the present section, we describe productivity patterns in the off-equatorial SEC from 3 to
13S (Fig. 4). At these latitudes the SEC transitions from the EU region in the north to the high salinity South
Pacific subtropical gyre waters in the south. The eastern SEC is cooled by influx of Peru Current water (Fig. 3),
and SEC region surface waters must also be cooled and receive nutrients from the EU region. At 110W, SEC
region surface nitrate averages almost 6 lM (Fig. 10D; Table 2A), considerably higher than at similar latitudes
in the northern hemisphere (Fig. 2C). The South Pacific is an area of extremely low iron dust input and iron sup-
ply decreases to the south (Fig. 11C). As a result the SEC region is one of the largest and most extreme HNLC
areas in the world’s oceans (see Section 3.4; Behrenfeld and Kolber, 1999). The combination of (1) warming of
upwelled water as it flows away from the equator and westward across the region and (2) low rate of photosyn-
thetic uptake of CO2 due to iron limitation results in net CO2 flux from the ocean into the atmosphere (Fig. 11A).
SEC region chlorophyll levels are intermediate between those of the EU region to the north and the SPSG
to the south (Table 2B). Surface values average 0.16 mg Chl m3 (Table 2A) and a subsurface maximum
0.2 mg Chl m3 descends from 50–75 m in the east (90–120W) to 60–110 m deep in the west (130–
140W; Fig. 5C). Production is likely supported in part by nutrients advected from the EU region and Peru
Current into the eastern SEC, as discussed above (Fig. 5C). SEC region rate of integrated primary production
is also intermediate between that in the EU and SPSG regions and is 16% of that in the PCU region (Table 2).
Although there are few measurements from this region, rates of N2-fixation in the SEC region are thought to
be low, again due to iron limitation (Fig. 11D). Subsurface waters of the SEC region have undergone substan-
tial denitrification, probably reflecting the spread of sub-thermocline Peru Current waters, as indicated by neg-
ative values of N* (Fig. 11B). N* is a Redfield-based N:P index that has been adjusted such that positive values
indicate that nitrogen fixation dominates, whereas negative values indicate denitrification dominates
(N* = 0.87 [NO3  16 PO4 + 2.9] lM kg1, Gruber and Sarmiento, 1997; see also Section 3.3 below).

2.4. South Pacific subtropical gyre (13–23S, 80/87–140W)

The South Pacific subtropical gyre (SPSG) region has the highest surface and 0–100 m salinities of the east-
ern tropical Pacific regions (Table 2A and B), and is defined by the ‘subtropical surface water’ of Fiedler and
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 303

Talley (see their Fig. 2; 2006). The permanent thermocline is substantially deeper (250 m) than the seasonally
variable mixed layer (80 m, Fiedler and Talley, 2006), severely reducing supply of nutrients to the euphotic
zone (<200 m). The SPSG region as defined in this paper (Fig. 4) includes some spillover of equatorial and
eastern boundary waters. As a consequence 0–100 m nutrients levels are not zero as would be found in the
SPSG proper but are still very low (Table 2B), with silicate lower than in all other regions at all depths. Inter-
estingly, surface and mean 0–100 m nitrate levels are similar (1.3 versus 1.1 lM, Tables 2A and B), apparently
due to the deep thermocline and the HNLC/iron-limited waters in the northern part of this region (see Beh-
renfeld and Kolber, 1999).
The SPSG region thus has low surface chlorophyll and productivity values, only 5–6% of those in the PCU
region (Table 2A). These values are similar to those for the open-ocean north of the EU region (Table 2A).
The ship-collected integrated productivity mean value seems high, particularly when compared to the VGP
model results derived from satellite data (see Section 3.1.2 below and Table 2B), which may be due to the small
number of ship-collected samples mostly from the northern, richer, part of the region (Fig. 4). Ship-collected
surface chlorophyll values are similarly biased. A deep chlorophyll maximum occurs over a broad depth
range, from 90 to 190 m (0.1 mg Chl m3; Fig. 7C and D). Given the SPSG region’s near-oligotrophy, pho-
tosynthetic bacteria are likely the dominant microbial group (e.g., Prochlorochoccus spp.; see Landry and Kir-
chman, 2002).

2.5. North equatorial countercurrent (3–9N, 91/105–140W)

As introduced in Section 2.2, the NECC flows towards Central America between the NEC and the SEC
against the weak winds characteristic of the ITCZ. As a surface current transporting water from the western
Pacific, the NECC is warm but salinity decreases towards the east because of heavy ITCZ rainfall (Table 2,
Fig. 10; see also Fiedler and Talley, 2006). The ITCZ migrates north during boreal summer and south during
boreal winter, following the sun. This seasonal migration causes NECC flow to be strongest in the fall and
weak or absent in spring (see Kessler, 2006).
NECC region surface nutrient levels are low and comparable to those in the subtropical gyres (Table
2A; Fig. 14B in Fiedler and Talley, 2006), but the nutricline is much shallower (Figs. 2 and 10; Table 2B)
so that NECC euphotic zone may receive some nutrients from depth in addition to lateral input across the
equatorial front and from the 10N thermocline ridge described below. Iron may be deposited along the
ITCZ by the enhanced rainfall. Perhaps as a consequence, NECC integrated chlorophyll is 36% higher
than in the SPSG region, although only 17% of that in the PCU region (Table 2B). As in the EU and
SEC regions, a deep chlorophyll maximum descends from 50 to 75 m from east to west (0.2–
0.3 mg Chl m3, Fig. 5). Surface chlorophyll appears to be slightly higher in the winter and spring months
when the NECC is weakest (Figs. 8 and 12). Integrated productivity is 9% of that for the PCU region
(Table 2B).

2.5.1. Equatorial front and tropical instability waves


At 2–5N, an abrupt transition between the eastward-flowing NECC and the westward-flowing SEC of
the EU region is called the ‘Equatorial Front’ (Fig. 3; Wyrtki, 1966; Pak and Zaneveld, 1974; Archer et al.,
1997; Kessler, 2006). Particularly in the east, the equatorial front can be very strong, with SEC tempera-
tures much lower (Fig. 3; Table 2) and salinity, nutrients, CO2 and chlorophyll values higher than in the
NECC (Pak and Zaneveld, 1974; Chavez and Smith, 1995; Archer et al., 1997). Strong shear between the
opposing currents produces cusp-like meanders in the front known as tropical instability waves (TIWs; see
Fig. 2 of Willett et al., 2006). Although not visible in the mean fields of Figs. 3 and 4, TIWs enhance mix-
ing and in sequential satellite images appear to propagate westward (Legeckis, 1977; Philander, 1978;
Yoder et al., 1994; Archer et al., 1997; Strutton et al., 2001; Ryan et al., 2002). TIWs enhance nutrient
input, productivity, and CO2 flux to the atmosphere (Foley et al., 1997; Friedrichs and Hofmann, 2001;
Le Borgne et al., 2002). The convergent Equatorial Front accumulates buoyant organic material that
can be seen from space (Yoder et al., 1994) and attracts foraging seabirds and other higher trophic level
species (Spear et al., 2001).
304 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

2.6. 10N thermocline ridge (9–13N, 105/111–140W)

On its northern margin at about 10N, the NECC flows eastward in opposition to the NEC (Fig. 3; Wyrtki,
1966; Kessler, 2006). Unlike at the Equatorial Front, this is a region of thermocline shoaling (Fig. 9 of Fiedler
and Talley, 2006), caused by positive wind stress curl cross the ITCZ (Fig. 9; and Fig. 1 of Kessler, 2006). This
shoaling forms the ‘10N thermocline ridge’ between the thermocline trough associated with the NECC and
equatorial front to the south and the deepening thermocline of the NEC and the North Pacific subtropical
gyre to the north (Fig. 7 of Fiedler and Talley, 2006). This zonal feature is termed the ‘countercurrent ther-
mocline ridge’ by Fiedler and Talley (2006).
The 10 ridge is dramatically visible in the meridional nitrate section of Fig. 2C and can also be discerned in
the temperature, salinity and nutrient profiles of Fig. 10, such that the 0–100 m nutrients are higher than in the
NEC or NECC regions to the north and south, respectively (Table 2B). However, the nutricline appears very
sharp (Fig. 2B) and the uplifted nutrients do not appear to support substantially enhanced phytoplankton
growth, as 10N ridge mean chlorophyll and productivity values are on average low – only 7–15% of those
in the PCU region – and similar to those in the NEC and NECC regions (Table 2B). Surface chlorophyll
and productivity values are even lower at 5% and 4%, respectively, of those in the PCU region (Table 2B),
and the subsurface chlorophyll maximum is not uplifted or enhanced in this region (Fig. 5B and C). However,
SeaWiFS imagery does show an increase in pigment along this ridge (Fig. 4) that is not apparent in the ship-
collected data. In addition, SeaWiFS chlorophyll estimates seem to show a seasonal cycle, with higher values
in the boreal winter (0.2 versus 0.1 mg Chl m3; Figs. 8L and 12). This seasonality seems odd in that the
NECC (and thus the thermocline ridging) is strongest (Kessler, 2006) and chlorophyll of the Costa Rica Dome
is maximal during summer and fall. In spite of the 10N ridge’s apparent near-oligotrophy (Table 2), a major
yellowfin tuna fishery is conducted in the region (see Fig. 2 in Ballance et al., 2006, this volume). Important
aspects of the biological productivity of this region remain unclear, or at least are not well-resolved by our
monthly climatologies and fixed regions.
Denitrified waters spread below the thermocline from the eastern Pacific warm pool into the 10N ridge
region and northwards (Fig. 11B). Little nitrogen fixation is thought to take place in EU and NECC surface
waters, but nitrogen fixation occurs at low but significant rates in the 10N ridge and NEC regions (Fig. 11D).
Few direct measurements of nitrogen fixation have been made in this area.

2.7. North equatorial current (13–23N, 110/115–140W)

The NEC flows southwest from off Baja California to 10–20N where it flows west across the eastern trop-
ical Pacific as the southern branch of the North Pacific subtropical gyre (Fig. 3). The NEC is the northern
counterpart of the SEC, but as discussed in Section 1.2 above and by Kessler (2006), their flows are separated
by the NECC and are asymmetrically distributed around the geographic equator: the NEC remains north of
the NECC and the thermocline ridge discussed above (3–13N), whereas the SEC penetrates the northern
hemisphere to 5N and contains the EU region. Both currents are, however, part of their respective subtrop-
ical gyres, and in both cases geostrophic balance results in increasing thermocline depth away from the equa-
tor, towards the gyre centers (Kessler, 2006).
Here we consider only that portion of the NEC that falls within our eastern tropical Pacific ‘box’ (Fig. 3).
In this NEC region, the depth of the permanent thermocline increases from about 60 m at the 10N ther-
mocline ridge to 100–160 m at 20N (Fig. 9A of Fiedler and Talley, 2006). This deepening permanent ther-
mocline results in decreasing euphotic zone nutrient supply. NEC surface NO3 concentration is the lowest of
all our regions at 1% of that in the PCU region, and the 0–100 m NEC and SPSG averages are equal at
only 7% of that for the PCU region (Table 2A and B, Fig. 10D and E). As a consequence, average
NEC region surface [and integrated] chlorophyll and productivity values are as low or lower than in any
of the regions, at 4% [and 13%] of those for the PCU region (Table 2A and B; Fig. 10C). Seasonality is
weak (Fig. 8M). NEC region surface chlorophyll decreases to the west as in the SEC, but the depth of
the NEC deep chlorophyll maximum is fairly constant at 80 m, unlike in the SEC where it is considerably
deeper in the west (Fig. 5C and D). NEC chlorophyll values are higher than those at similar latitudes in the
SPSG region (Fig. 5C and D). This difference may reflect (1) increased iron supply via atmospheric dust and
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 305

consequent N2-fixation, which is thought to occur at low but significant rates in NEC region surface waters
(Fig. 11D; Karl et al., 1997, 2002), or (2) simply that due to the subtropical gyre asymmetry around the
equator, the truly oligotrophic waters (surface chlorophyll <0.1 mg chlorophyll m3) of both gyres are
shifted northward. Prochlorococcus spp. are a dominant microbial group in the NEC region (Landry and
Kirchman, 2002).
The euphotic zone of the NEC region overlies denitrified waters (N*  0.5; Fig. 11B) that apparently
spread from the eastern Pacific warm pool far into the western Pacific.

3. Synthesis

3.1. Biological productivity

3.1.1. Contribution to Pacific and global production


The eastern tropical Pacific as defined in this paper accounts for 23% of Pacific Ocean primary production
while comprising 18% of its area (Table 1). Similarly, it accounts for 10% of global oceanic primary produc-
tion while comprising 9% of the ocean’s area. Within the eastern tropical Pacific an area of moderately high
chlorophyll extending from about 15N to 15S is sandwiched between the subtropical gyres, with even higher
coastal chlorophyll values (Fig. 4). Surface chlorophyll averages 0.3 mg Chl m3 (data from Table 2B),
whereas that in the oligotrophic northern and southern subtropical gyres is less than 0.1 mg Chl m3. On aver-
age, new production in the EU region is estimated to be five times greater than that in the oligotrophic western
tropical Pacific (Peña et al., 1994); including the eastern margin would make this contrast even stronger.

3.1.2. Ship-collected versus SeaWiFS and VGP model results


The descriptions of regional biological production in Section 2 above are based on samples collected by
shipboard personnel over 1980–2005 and represent many months of labor. These data are nevertheless sparse
and unevenly distributed, particularly when compared to that collected by satellites (see Fig. 4) or generated
by models. To compare ship- and satellite-collected chlorophyll data, we assembled available SeaWiFS data
for each month (1997–2005), averaged the pixel values within each of the regions and subregions, and plotted
the ship- and satellite-collected values by month (see Appendix). Discounting several outliers in the ship-col-
lected data, likely due to low sample size, the two data sets agree remarkably well for the open-ocean regions
(Fig. 8F–G and I–M; Table 2A). For the coastal areas, however, substantial discrepancies occur. In the PCU
and Galapagos subregions, SeaWiFS appears to slightly underestimate chlorophyll (Fig. 8E and H; Table 2A).
In contrast, SeaWIFS appears to greatly overestimate chlorophyll in the Central American gulf regions
(Fig. 8B–D; Table 2A). These differences between ship-collected and satellite chlorophyll estimates in the Gulf
regions need to be resolved. At the time of writing a newly processed SeaWiFS data set was made available
which has lower estimates for the Gulf regions.
We have similarly compared the ship-collected primary production data with that produced by the Verti-
cally Generalized Production (VGP) model of Behrenfeld and Falkowski (1997b). As discussed in Section 2.2,
measurements of primary production based on ship-collected incubations of 14C have been subject to consid-
erable uncertainty and have been revised upwards over the years. The VGP model combines SeaWiFS chlo-
rophyll data as above with climatological SSTs and light to estimate depth-integrated total primary
production (see Appendix). The two methods show remarkably good agreement in the eastern tropical Pacific,
with the VGP model only 4% lower than the ship-collected results (calculated from Table 2B). However, as
with the chlorophyll comparison, there are considerable spatial differences in the results. VGP model produc-
tivities are substantially higher than ship-collected for the Central American gulf regions, and substantially
lower than ship-collected PCU and Galapagos values (Table 2B). These differences presumably propagate into
the model from the satellite chlorophyll discrepancies discussed above. In addition, VGP model production is
similar in (1) open-ocean North Pacific regions combined and (2) EU plus South Pacific regions (408 versus
383 mg C m2 d1; regional data averaged from Table 2B). In contrast, the ship-collected data suggest pro-
ductivity is much lower in the north than in the EU plus South Pacific regions (258 versus 564 mg C m2 d1;
also calculated from Table 2B), which makes sense because the productive EU should contribute to a higher
average productivity to the south. It is not clear why the VGP model results do not confirm this difference.
306 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

3.2. Basin-scale thermocline tilt and shoaling, wind-driven upwelling, and nutrient supply

As introduced in Section 1, thermocline and nutricline depth play a major role in regulating primary pro-
ductivity. In the eastern tropical Pacific, three interrelated processes act to lift the nutricline, the first two on
the basin-scale and the third on local scales:

(1) The trade winds drive warming equatorial surface water from the eastern into the western Pacific – so
much so that the sea level off Indonesia is 30–50 cm higher than off Peru. In the west, as water and heat
accumulate, the thermocline deepens to 200 m, while in portions of the eastern equatorial Pacific the
thermocline is as shallow as 50 m (Barber and Chavez, 1983; Tomczak and Godfrey, 1994). This east–
west thermocline tilt enhances euphotic zone nutrient supply in the east.
(2) In the north [or south] Pacific subtropical gyre, the wind blows clockwise [or counterclockwise] around
the basin, and its frictional stress on the ocean drives surface currents downwind. However because of
the Earth’s rotation, the surface currents are deflected 90 to the right of the wind in the northern hemi-
sphere, and 90 to the left in the southern hemisphere. These Ekman currents converge in the centers of
the subtropical gyres, depressing the thermocline there, while lifting the thermocline toward the gyre
edges. This shoaling of the thermocline near the gyre edges increases the supply of macronutrients over
substantial portions of the eastern tropical Pacific.
(3) Wind-driven upwelling can further exaggerate the thermocline shoaling described above. Where equator-
ward wind blows alongshore, Ekman transport is offshore and waters from below the thermocline and
nutricline are necessarily drawn to the surface to maintain mass balance. This process is called ‘coastal
upwelling’. In the open ocean, the direction of Ekman transport changes to either side of the equator
such that flow is to the right of the wind north of the equator, and to the left of the wind south of
the equator, producing divergent flow right on the equator. Sub-thermocline waters are thus drawn to
the surface and euphotic zone nutrient supply is enhanced (see e.g., Fig. 4.1 of Tomczak and Godfrey,
1994). This process is called ‘equatorial upwelling’. Ekman pumping, as described in Section 2.1.1, can
also result in wind-driven upwelling.

The basin-scale thermocline tilt and shoaling described above are general characteristics of the subtropical
gyres and eastern margins in both the Pacific and Atlantic Oceans and are the physical basis for the Coastal
Boundary, EU and 10N ridge regions. Relative to these large scale processes, wind-driven upwelling is an
intermittent, shallow (<100 m) and fairly local process. However, at those locations and periods where these
processes co-occur, levels of nutrient supply and biological productivity can be very high. In the eastern trop-
ical Pacific, thermocline shoaling together with wind-driven upwelling and mixing form the physical basis for
the highly productive PCU, Tehuantepec, Papagayo, Panama and EU regions. Conversely, the 10N ridge
occurs in a low wind environment (the ITCZ; Figs. 3 and 12). Even though the thermocline and nutricline
are shallow (50–75 m; Fig. 10A and D), the lack of wind-driven upwelling and/or mixing may account for
the region’s relatively low productivity and chlorophyll (Table 2A).

3.3. The oxygen minimum and denitrified zone

The largest area of low oxygen in the world lies under the thermocline in the eastern tropical Pacific (Fia-
dero and Strickland, 1968; Codiposti et al., 1986). The area of low oxygen extends from Central and north-
ern South America across the eastern tropical Pacific as tongues to either side of the equator (see Fig. 21 in
Fiedler and Talley, 2006). These tongues are separated by the higher oxygen water of the EUC that flows
eastward, underneath the SEC, on the equator. The oxygen minimum is formed by at least three processes:

(1) Lack of ventilation, isolation from the atmosphere, and low oxygen content of source waters. Galbra-
ith et al. (2004) suggest that the meridional overturning circulation controls rate of ventilation of the
eastern Pacific oxygen minimum, which is low during interglacial periods as at present. A permanently
strong pycnocline and shallow mixed layer (Fiedler and Talley, 2006) further isolate sub-thermocline
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 307

waters from the atmosphere. Toggweiler and Carson (1995) argue that the EUC contributes reduced
oxygen source waters to the oxygen minimum because EUC oxygen levels decrease while crossing the
Pacific.
(2) Long residence time, with cumulative oxygen depletion. The eastern Pacific is an area of generally weak
wind and is not penetrated by the subtropical gyres. As a result, residence time is long and oxygen deple-
tion cumulates.
(3) Decay of biological production sinking from overlying surface waters, depleting oxygen. In surface
waters organic particles are created by photosynthesis, phytoplankton growth and the consumption
of phytoplankton by animals. A fraction of these particles sink and decay below the thermocline. Decay
depletes oxygen, and in the absence of ventilation, oxygen levels can fall to zero.

Beneath the eastern Pacific warm pool, low oxygen source waters and long residence times must be impor-
tant factors in the formation of the oxygen minimum given the relatively low rates of primary production in
overlying waters (Table 2). There is still debate over the relative roles of ventilation, circulation and produc-
tivity off Peru, but the very high rates of PCU region primary production must have a large impact.
Where oxygen is completely depleted, nitrate can serve as a terminal electron acceptor and is converted to
NO2 or N2 in a process called ‘denitrification’ (Brandhorst, 1959; Thomas, 1966; Fiadero and Strickland, 1968;
Cline and Richards, 1972; Codiposti and Richards, 1976). As a result, a nitrate deficit develops relative to phos-
phate, and this deficit is measured by the calculated property ‘N*’ (defined in Section 2.7, above; Gruber and
Sarmiento, 1997). Denitrified waters spread poleward and far to the west of the eastern tropical Pacific and over
a depth range of about 200–1000 m (Fig. 11B; Castro et al., 2001; Gruber and Sarmiento, 1997). The low oxy-
gen waters of the eastern tropical Pacific are an important factor in the Pacific and global nitrogen budgets
(Codiposti and Richards, 1976; Gruber and Sarmiento, 1997; Codispoti, 2002; Montecino et al., in press).

3.4. New production and community structure

Macronutrients occur in a fairly constant molar stoichiometry in cells (the ‘Redfield ratio’, 16N:1P) and are
abundant below the thermocline but often depleted in the euphotic zone (see Broecker and Peng, 1982, and
references therein). There has been perennial discussion as to which of these nutrients (N or P) limits primary
production on global and long-term scales (Tyrrell, 1999). Because of the high denitrification rates described
above, in the eastern tropical Pacific nitrogen is the primary limiting macronutrient. Dugdale and Goering
(1967) proposed that euphotic zone primary production can be divided into that supported by imported or
‘new’ nutrients versus that supported by recycled or ‘regenerated’ nutrients. Several important corollaries fol-
lowed from this distinction: (1) New nutrients have in practice become synonymous with nitrate, and regen-
erated nutrients with ammonia, permitting measurement of the levels and rates of new versus regenerated
production (Dugdale and Goering, 1967); (2) Phytoplankton communities that enjoy high rates of nutrient
input are highly productive and characterized by large-sized phytoplankton which preferentially use nitrate
(in particular diatoms, which in some areas become regulated by silicate instead of nitrate); whereas (3) Com-
munities with low rates of nutrient input are less productive and characterized by small-sized phytoplankton
that use ammonia (pico- and nanoplankton; Chavez, 1989; Probyn, 1992; Goldman, 1993; Chavez and Smith,
1995; Chavez et al., 1996; Dugdale and Wilkerson, 1998; Mackey et al., 2002). The Dugdale and Goering
(1967) construct and its corollaries focus on macronutrient supply and set the conceptual framework for
our understanding of biological productivity in the oceans, a framework recently broadened to include the
micronutrient iron (see Section 2.4, below). In the productive areas of the eastern tropical Pacific, community
structure is determined and primary production maintained at levels above those of the oligotrophic gyres via
macronutrient supply associated with thermocline shoaling and wind-driven upwelling as described in Section
3.2 (see also Kessler, 2006; Fiedler and Talley, 2006).

3.5. HNLC, iron, and nitrogen fixation

The Dugdale and Goering (1967) new versus regenerated production distinction has placed a great deal
of emphasis on nitrogen budgets, in particular on nitrate, and the eastern tropical Pacific is in general a
308 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

nitrogen-limited system (Fiedler et al., 1991). However, in over 20% of the world’s oceans (Martin et al., 1994),
euphotic zone nitrate is not fully utilized by phytoplankton and levels remain high (Table 2). In the eastern trop-
ical Pacific, this high-nitrate low-chlorophyll condition (HNLC) is found in the EU and SEC regions, in offshore
waters of the PCU region, and likely in the Costa Rica Dome (Hutchins et al., 2002; Bruland et al., in press).
In part because of its implications for carbon cycling and global warming, oceanographers have been fas-
cinated by HNLC. If the excess nitrate were used by phytoplankton, an enormous amount of CO2 would also
be ‘fixed’ as organic carbon, potentially altering global biogeochemical budgets and climate (Martin, 1990).
Hypotheses to explain HNLC have included trace metal inhibition of phytoplankton growth (Barber and
Ryther, 1969), grazing (Walsh, 1976; Frost, 1987; Cullen et al., 1992; Landry et al., 1997; Landry and Kirch-
man, 2002), removal of photosynthetic cells from the euphotic zone by the EUC (Sorokin et al., 1977), other
physical processes (Thomas, 1979), species composition (Chavez, 1989), nitrate or silicate regulation (Wilker-
son and Dugdale, 1992; Dugdale et al., 1995, 2002; Dugdale and Wilkerson, 1998), and iron limitation (Soro-
kin et al., 1977; Martin, 1990; Barber and Chavez, 1991).
Iron is a required constituent of phytoplankton cells but is present at picomolar levels (1012 M) in surface
waters of many open ocean regions (Martin et al., 1994; reviewed by Coale et al., 1996b). Until the develop-
ment of trace-metal clean techniques beginning in the 1970s, the role of iron as a regulating ‘micronutrient’
remained obscure. In the 1990s, however, dissolved iron was pumped overboard in experiments that fertilized
patches of ocean (<1000 km2) in the EU and SEC regions. Spectacular increases in phytoplankton biomass
and productivity resulted, demonstrating that insufficient iron is a main factor limiting phytoplankton growth
in HNLC waters (Behrenfeld et al., 1996; Coale et al., 1996a,b, 1998; Gordon et al., 1997). While the non-iron
hypotheses mentioned above do not appear responsible for HNLC, they nevertheless remain important pro-
cesses in the highly interdependent biological dynamics of the eastern tropical Pacific and elsewhere. As an
example, small cells are favored under iron limitation, and many of these use ammonia rather than nitrate
as a nitrogen source (Rocap et al., 2003). These small cells are subject to intense grazing by microzooplankton
with doubling rates that match those of the phytoplankton, preventing blooms that in any case could not be
supported by the available ammonia. These and other factors interact to maintain HNLC conditions.
For the Coastal Boundary region and the Galapagos, iron is presumably not limiting and derives from the
continental shelves and land. For open-ocean regions of the eastern tropical Pacific, iron is often limiting and
the primary source is fallout of dust transported at high altitude from Asia (Duce and Tindale, 1991; Prospero,
1996; Pettke et al., 2002; Gao et al., 2003; Meskhidze et al., 2003). In general, iron supply decreases from north
to south (Fig. 11C; Chavez et al., 1996b; Behrenfeld and Kolber, 1999; Moore et al., 2002), although rainfall
may alter this pattern in the ITCZ. (Prospero, 1996). The EUC has also been suggested to transport iron from
the continental shelves of Indonesia to the eastern tropical Pacific (Wells et al., 1999; Ryan et al., in press).
In a seeming evolutionary paradox, all eukaryotic photosynthetic organisms, including phytoplankton,
require nitrogen but cannot use the N2 gas that is abundant in the atmosphere and dissolved in the oceans
(Zehr and Ward, 2002; Galloway, 2005). Instead, phytoplankton must use ‘biologically reactive’ forms of
nitrogen such as nitrate and ammonia. Such nitrogen is continually lost to biological systems by denitrification
(see Section 3.3, above), and organic nitrogen is lost when buried in sediments. These losses are, however, off-
set by prokaryotes that convert N2 into reactive nitrogen (Karl et al., 1997; Montoya et al., 2004). Such ‘nitro-
gen fixation’ is performed by a diverse group of planktonic prokaryotes including phototrophic,
heterotrophic, and chemolithotrophic bacteria, heterocystous and non-heterocystous cyanobacteria, and
Archaea (Zehr et al., 2001; Karl et al., 2002; Montoya et al., 2004). It is thought that nitrogen fixing organisms
have high iron requirements and are often iron-limited (Gruber and Sarmiento, 1997; Karl et al., 2002), as iron
is a cofactor for the enzyme nitrogenase that splits N2 (Weinberg, 1989; Karl et al., 1997, 2002). If this idea is
correct, nitrogen fixers should thrive in oligotrophic waters with substantial iron input such as the tropical
Atlantic and North Pacific subtropical gyre (Fig. 11C; Karl et al., 1997, 2002), and may over time shift these
systems from nitrogen- to phosphorous-limitation (Karl et al., 1995). For the open-ocean areas of the eastern
tropical Pacific, rates of nitrogen fixation are thought to be low due to lack of iron (Nemi, 1979; Karl et al.,
1997, 2002; Kamykowski et al., 2002). The general trend of decreasing iron supply towards the southeastern
tropical Pacific is paralleled by decreasing nitrogen fixation to the south (Fig. 11C and D; Lee et al., 2002;
Moore et al., 2002). The northeastern tropical Pacific is the only Pacific Ocean region with any significant
nitrogen fixation according to Lee et al. (2002, Fig. 11D).
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 309

3.5.1. Carbon dioxide flux


In the eastern tropical Pacific the flux of carbon dioxide gas (CO2) into and out of the oceans is controlled
by: (1) upwelling of high CO2 water; (2) warming of cooler waters either recently upwelled or advected from
higher latitudes; (3) wind speed; and (4) the biological uptake of CO2 by phytoplankton (Takahashi et al.,
1997). Where high-CO2 water upwells, the partial pressure of CO2 at the surface is above that in the atmo-
sphere and CO2 diffuses from the ocean. Surface warming decreases the solubility of gasses in water and acts
to increase venting to the atmosphere, and wind increases the rate of transfer across the air-water interface.
Conversely, where CO2 is used by phytoplankton, the euphotic zone partial pressure of CO2 can be drawn
below atmospheric levels, and CO2 will diffuse into the ocean. In the EU and PCU regions, upwelling brings
new CO2-rich water to the surface that then diffuses into the atmosphere. This venting to the atmosphere is
reduced by biological uptake of CO2 by phytoplankton. However, widespread iron limitation in the open-
ocean eastern tropical Pacific causes excess CO2 (in addition to nitrate) to remain unused in HNLC waters,
and as a result the region vents between 0.7 and 1 gigatons of carbon (as CO2) annually into the atmosphere
(Fig. 11A). As with phytoplankton production (Fig. 5A and B) there is a north–south asymmetry in the east-
ern margin CO2 fluxes, with high fluxes out of the ocean in the SEC and neutral conditions north of the equa-
tor (Fig. 11A). Similarly, upwelling of nitrate-deficient waters in the PCU region may also restrict biological
production, leaving an excess of CO2 which vents to the atmosphere.

3.6. Temporal fluctuations

We have focused on the processes that drive the spatial variability of phytoplankton biomass and produc-
tivity. However, significant temporal variations also occur. Seasonal fluctuations are present but not strong
in most regions of the eastern tropical Pacific (Figs. 8, 9, 12, and see Fig. 7 in Wang and Fiedler, 2006). Excep-
tions are (1) the Gulf subregions, where the cross-isthmus winds and consequent upwelling show strong
seasonality, (2) the NECC region, where NECC flow and ITCZ latitude vary seasonally and (3) the PCU
region, where, somewhat surprisingly, seasonal cycles in production are out of phase with seasonal upwelling
winds. These seasonal cycles have been discussed in Section 2 above. Below we introduce two sources of long-
term temporal variability in the eastern tropical Pacific.

3.6.1. Interannual variability – El Niño


‘El Niño’ is arguably the most significant source of temporal variability in the eastern tropical Pacific. El
Niño is part of an unstable, coupled, air/sea heat flux interaction in the equatorial Pacific known as ENSO (El
Niño-Southern Oscillation; Philander, 1999). El Niño is that portion of the oscillation wherein changes in
atmospheric circulation cause heat from the warm surface waters of the far western Pacific to be redistributed
eastward into the eastern equatorial Pacific and elsewhere (reviewed in detail by Fiedler, 2002b; Wang and
Fiedler, 2006). For reasons not fully understood, El Niños are triggered by weakening or reversals of the east-
erly trades in the western Pacific. These wind anomalies reduce or eliminate the east-west downwards tilt of the
equatorial thermocline, producing Kelvin waves and associated advective flow that travels eastward along the
equatorial waveguide across the Pacific to the South American coast (see Chavez et al., 1999; McPhaden and
Yu, 1999), where they are deflected and propagate north and south along the continental margins as coastally-
trapped waves (Enfield and Allen, 1980). El Niños of varying strengths occur irregularly every 2–7 years and
last 6–18 months.
El Niño thus affects the eastern tropical Pacific most strongly in the EU and Coastal Boundary regions (see
Wang et al., 2004; Fiedler and Talley, 2006). In the EU region, easterly trade winds weaken or reverse, the ther-
mocline deepens, the SEC weakens, and the EUC weakens or disappears (on the EUC see Feldman et al., 1984).
Similar changes occur along the coastal boundaries in both hemispheres, where equatorward winds weaken or
reverse and poleward flow of the inshore countercurrents is enhanced. These changes are manifested as local
increases in sea surface height, temperature and salinity. As a consequence of the passage of the Kelvin waves
and the advective changes noted above, a deeper thermocline occurs throughout the EU and Coastal Boundary
regions. Most importantly to biological production, this thermocline deepening results in substantial decreases
in euphotic zone nutrient supply, particularly in the far eastern Pacific (Barber and Chavez, 1983). Even in those
regions where wind-driven upwelling continues (EU, PCU, Gulfs of Tehuantepec, Papagayo and Panama), low
310 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

nutrient waters from above the thermocline are often brought to the surface (see for example Huyer
et al., 1987). In the EU region, El Niño results in a decrease in CO2 venting to the atmosphere because sub-
thermocline CO2-rich waters are not upwelled (Wanninkhof et al., 1992; Chavez et al., 1999; Feely et al.,
1997; Feely et al., 2002; Le Borgne et al., 2002).
El Niños have been observed to reduce nutrient supply, and consequently primary production and chloro-
phyll by several-fold in the eastern tropical Pacific (Guillén and Calienes, 1981b,c; Feldman et al., 1984; Bar-
ber and Chavez, 1983, 1986; Fiedler et al., 1991, 1992; Barber et al., 1996; Sanchez et al., 2000; Chavez, 2005).
In coastal regions the phytoplankton community also shifts — from diatoms and larger phytoplankton to a
community of picoplankton more typical of the open-ocean (cyanobacteria and prochlorophytes; Chavez,
2005). Warm-water species, such as rare and solitary oceanic dinoflagellates, are advected onshore and pole-
ward into the normally productive nearshore environments (Avarı́a and Muñoz, 1987; Ochoa and Gomez,
1987). These changes disrupt food webs and lead to death or emigration of higher trophic level animals depen-
dent on phytoplankton production (Chavez, 2005 and references therein). Off Peru and Chile, the oxygen min-
imum layer deepens along with the thermocline, oxygenating the shelf and upper continental slope and thus
altering levels of denitrification and the export flux of organic material (Morales et al., 1999; Montecino et al.,
in press). In the Central American gulf regions, the impact of El Niño is less severe than it is in the EU and
PUC regions (Fiedler, 2002b).
More recently the focus has increased on La Niña, the cool phase of ENSO. Following the 1997–1998 El
Niño, a strong La Niña developed about a month after trade winds and upwelling resumed in May 1998 (Cha-
vez et al., 1999; Strutton and Chavez, 2000; Wilson and Adamec, 2001; Carr et al., 2002; Ryan et al., 2002;
Wang et al., 2005). A bloom developed in the EU that was larger in size than the entire PCU region and must
have had dramatic biological consequences beyond primary production (Ryan et al., in press). This La Niña
continued into 1999 and was the most dramatic in recent history, and resulted in a huge increase in primary
production that contrasted with the low productivity associated with El Niño, both in the eastern tropical
Pacific and globally (Behrenfeld et al., 2001).

3.6.2. Multi-decadal variability – PDO


Multi-decadal variability refers to low frequency physical fluctuations with periods of 40–60 years that affect
biogeochemical cycles and marine populations (Chavez et al., 2003; Takahashi et al., 2003; Mestas-Nuñez and
Miller, 2006). A primary mode of multi-decadal variability is the Pacific Decadal Oscillation (PDO), which
strictly speaking refers to the fluctuations of sea surface temperatures in the North Pacific (Mantua and Hare,
2002). More recently however, Chao et al. (2000) have shown that the PDO affects the entire Pacific with about
25 warm years (‘El Viejo’) followed by 25 cool years (‘La Vieja’; see Chavez et al., 2003; Chavez, 2005). A great
deal of both spatial and temporal variability occurs within these long-term fluctuations (storms, seasons,
ENSO). The strongest multi-decadal effects in the eastern tropical Pacific are found in the EU region (McPhaden
and Zhang, 2002) and PCU (Alheit and Niquen, 2004) where, during El Viejo, SST is warmer, the thermocline is
deeper and the north and southeastern limbs of the gyral circulation slows (Chavez et al., 2003). As with ENSO,
the effects of the PDO are less notable in the eastern Pacific warm pool (Fig. 4 in Mestas-Nuñez and Miller, 2006)
and may reverse sign within the interior of the subtropical gyres.
El Viejo and La Vieja appear analogous to El Niño and La Niña in that both types of fluctuation have
similar effects on basin-wide SST and thermocline depth (Mantua and Hare, 2002; Chavez, 2005), although
it remains uncertain to what extent these similarities are rooted in cause. In the eastern tropical Pacific, the
multi-decadal physical perturbations are of smaller absolute magnitude than ENSO, but perhaps have large
cumulative effects on fish populations because of their long duration. For example, global variations in
anchovy and sardine fisheries have been linked to multi-decadal variability (Sanchez et al., 2000; Chavez
et al., 2003; Alheit and Niquen, 2004; Bertrand et al., 2004; Chavez, 2005). In the PCU region, the warm
El Viejo favors oceanic or picoplankton-dominated phytoplankton communities, and sardines flourish. Dur-
ing La Vieja, cool conditions favor diatom-dominated phytoplankton communities, which in the PCU transfer
up the food chain directly to anchovies. Even though significant progress has been made in understanding
such relationships between phytoplankton, fish and climate, the mechanisms which underlie these linkages
are in many cases unresolved. Until these linkages are understood, it will be impossible to predict how the
eastern tropical Pacific and the global ocean will respond to natural and man-induced change.
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 311

Acknowledgements

The authors acknowledge long standing support from NOAA, NASA, NSF, and the David and Lucile
Packard Foundation. This is a contribution to the scientific agenda of the Eastern Pacific Consortium of
the InterAmerican Institute for Global Change Research. Support was also provided by the Protected Re-
sources Division of NOAA Fisheries, Southwest Fisheries Science Center.

Appendix A. Methods

A.1. Ship-collected samples

The primary database used in this paper has been assembled from a variety of sources including unpub-
lished data by one of the authors (FPC). Ship-collected chlorophyll data from bottle casts that fall within
the eastern tropical Pacific region total 16,700 discreet depth points. The methods for nutrients, chlorophyll
and primary productivity generally follow those described by Pennington and Chavez (2000).
In addition to MBARI observations, we also utilized 49,000 discreet depth chlorophyll measurements from
the National Oceanic and Atmospheric Association (NOAA) data center World Ocean Database (WOD) 1998
CD-ROM data set version 2.0 (Conkright et al., 1998), and 14,000 discreet depth points from the NOAA
National Marine Fisheries Service, Southwest Fisheries Science Center (NMFS-SWFSC) database. NMFS-
SWFSC samples were collected from NOAA research vessels David Starr Jordan, McArthur and Endeavor.

A.2. SeaWiFS chlorophyll

In addition to ship-collected chlorophyll data, analyses were based on SeaWiFS ocean color data. In par-
ticular, monthly global 9 km level 3 standard mapped images (version 4) from January 1999 to December 2003
produced by Ocean Biology Processing Group (OBPG) were utilized. Satellite chlorophyll data were used on
the maps of biogeophysical provinces (Fig. 3), seasonal chlorophyll (Fig. 12), and in the tabulated regional
surface chlorophyll of Table 2A.

A.3. VGP Model

The Vertically Generalized Production (VGP) model of Behrenfeld and Falkowski (1997a,b) was used to
estimate net primary productivity (Fig. 1B and Table 2B). The model was run with SeaWiFS 8 day global
9 km level 3 chlorophyll data from January 1999 to December 2002 (version 4).

References

Alheit, J., Niquen, M., 2004. Regime shifts in the Humboldt Current ecosystem. Progress in Oceanography 60, 201–222.
Archer, D., Aiken, J., Balch, W., Barber, D., Dunne, J., Flament, P., Gardner, W., Garside, C., Goyet, C., Johnson, E., Kirchman, D.,
McPhaden, M., Newton, J., Peltzer, E., Welling, L., White, J., Yoder, J., 1997. A meeting place of great ocean currents: shipboard
observations of a convergent front at 2N in the Pacific. Deep-Sea Research I 44, 1827–1849.
Amador, J.A., Alfaro, E.J., Lizano, O.G., Magaña, V.O., 2006. Atmospheric forcing of the eastern tropical Pacific: a review. Progress in
Oceanography 69 (2–4), 101–142.
Avarı́a, S., Muñoz, P., 1987. Effects of the 1982–1983 El Niño on the marine phytoplankton off northern Chile. Journal of Geophysical
Research 92, 14367–14382.
Aufdenkampe, A.K., McCarthy, J.J., Navarette, C., Rodier, M., Dunne, J., Murray, J.W., 2002. Biogeochemical controls on new
production in the tropical Pacific. Deep-Sea Research II 49, 2619–2648.
Aufdenkampe, A.K., Murray, J., 2002. Controls of new production: the role of iron and physical processes. Deep-Sea Research II 49,
2649–2688.
Ballance, L.T., Pitman, R.L., Fiedler, P.C., 2006. Seabirds, cetaceans and the oceanography of the eastern tropical Pacific: a review.
Progress in Oceanography, this volume.
Barber, R.T., Chavez, F.P., 1983. Biological consequences of El Niño. Science 222, 1203–1210.
Barber, R.T., Chavez, F.P., 1986. Ocean variability in relation to living resources during the 1982/83 El Niño. Nature 319, 279–285.
Barber, R.T., Chavez, F.P., 1991. Regulation of primary productivity rate in the equatorial Pacific Ocean. Limnology and Oceanography
36, 179–186.
312 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Barber, R.T., Dugdale, R.C., MacIsaac, J.J., Smith, R.L., 1971. Variations in phytoplankton growth associated with the source and
conditioning of upwelling water. Investagaciones Pesqueras 35, 171–193.
Barber, R.T., Ryther, J.H., 1969. Organic chelators: factors affecting primary production in the Cromwell Current upwelling. Journal of
Experimental Marine Biology and Ecology 3, 191–199.
Barber, R.T., Sanderson, M.P., Lindley, S.T., Chai, F., Newton, J., Trees, C.C., Foley, D.G., Chavez, F.P., 1996. Primary productivity
and its regulation in the equatorial Pacific during and following the 1991–92 El Niño. Deep-Sea Research II 43, 933–970.
Barber, R.T., Smith, R.L., 1981. Coastal upwelling ecosystems. In: Longhurst, A. (Ed.), Analysis of Marine Ecosystems. Academic Press,
New York, pp. 31–68.
Behrenfeld, M.J., Bale, A.J., Kolber, Z.S., Aiken, J., Falkowski, P.G., 1996. Confirmation of iron limitation of phytoplankton
photosynthesis in the equatorial Pacific Ocean. Nature 383, 508–511.
Behrenfeld, M.J., Falkowski, P.G., 1997a. A consumer’s guide to phytoplankton primary productivity models. Limnology and
Oceanography 42, 1479–1491.
Behrenfeld, M.J., Falkowski, P.G., 1997b. Photosynthetic rates derived from satellite-based chlorophyll concentration. Limnology and
Oceanography 42, 1–20.
Behrenfeld, M.J., Kolber, Z.S., 1999. Widespread iron limitation of phytoplankton in the South Pacific Ocean. Science 283, 840–843.
Behrenfeld, M.J., Randerson, J., McClain, C., Feldman, G., Los, S., Tucker, C., Falkowski, P., Field, C., Frouin, R., Esaias, W., Kolber,
D., Pollack, N., 2001. Biospheric primary production during an ENSO transition. Science 291, 2594–2598.
Bertrand, A., Segura, M., Gutierrez, M., Vasquez, L., 2004. From small-scale habitat loopholes to decadal cycles: a habitat-based
hypothesis explaining fluctuation in pelagic fish populations off Peru. Fish and Fisheries 5, 296–316.
Blackburn, M., Laurs, R., Owen, R., Zeitzschel, B., 1970. Seasonal and areal changes in standing stocks of phytoplankton, zooplankton
and micronekton in the eastern tropical Pacific. Marine Biology 7, 14–31.
Brandhorst, W., 1958. Thermocline topography, zooplankton standing crop, and mechanisms of fertilization in the eastern tropical
Pacific. Journal du Conseil Permanent International pour l’Exploration de la Mer 24, 16–31.
Brandhorst, W., 1959. Nitrification and denitrification in the eastern tropical North Pacific. Journal du Conseil Permanent International
pour l’Exploration de la Mer 25, 187–193.
Broecker, W.S., Peng, T-H., 1982. Tracers in the Sea. Lamont-Doherty Geological Observatory, New York.
Bruland, K.W., Rue, E.L., Smith, G.J., DiTullio, G.R., in press. Iron, macronutrients and diatom blooms in the Peru Upwelling regime:
brown and blue waters of Peru. Marine Chemistry.
Calienes, R., Guillén, O., 1981. Masas de agua y producción primaria en el Perú. Boletin del Instituto del Mar del Perú, vol.
extraordinario, Proy. Intvestigacción Cooperativa de la Anchoveta y su Ecosistema (ICANE), pp. 155–163.
Calienes, R., Guillén, O., Lostaunau, N., 1985. Variabilidad espacio-temporal de clorofila, produccı́on primaria y nutrientes frente a la
costa peruana. Boletin del Instituto del Mar del Perú 10.
Carr, M.E., 2002. Estimation of potential productivity in Eastern Boundary Currents using remote sensing. Deep-Sea Research II 49, 59–
80.
Carr, M.E., Kearns, E.J., 2003. Production regimes in four Eastern Boundary Current systems. Deep-Sea Research II 50, 3199.
Carr, M.E., Strub, P.T., Thomas, A.C., Blanco, J.L., 2002. Evolution of 1996-1999 La Niña and El Niño conditions off the western coast
of South America: a remote sensing perspective. Journal of Geophysical Research 107 (C12), 3236. doi:10.1029/2001/JC001183.
Castro, C.G., Chavez, F.P., Collins, C.A., 2001. Role of the California Undercurrent in the export of denitrified waters from the eastern
tropical Pacific. Global Biogeochememical Cycles 15, 819–830.
Chai, F., Dugdale, R.C., Peng, T.-H., Wilkerson, F.P., Barber, R.T., 2002. One-dimensional ecosystem model of the equatorial Pacific
upwelling system. Part I: model development and silicon and nitrogen cycling. Deep-Sea Research II 49, 2713–2745.
Chao, Y., Ghil, M., McWilliams, J.C., 2000. Pacific interdecadal variability in this century’s sea surface temperatures. Geophysical
Research Letters 27, 2261–2264.
Chavez, F.P., 1987. Ocean variability and phytoplankton community structure. Ph.D. Dissertation. Duke University, Durham, NC,
unpublished.
Chavez, F.P., 1989. Size distribution of phytoplankton in the central and eastern tropical Pacific. Global Biogeochememical Cycles 3, 27–
35.
Chavez, F.P., 1995. A comparison of ship and satellite chlorophyll from California and Peru. Journal of Geophysical Research 100,
24855–24862.
Chavez, F.P., 2005. Biological consequences of interannual to multidecadal variability. In: Robinson, A., Brink, K. (Eds.), The Sea, vol.
13. Harvard University Press, Cambridge, pp. 643–679.
Chavez, F.P., Barber, R.T., 1987. An estimate of new production in the equatorial Pacific. Deep-Sea Research 34, 1229–1243.
Chavez, F.P., Brusca, R.C., 1991. The Galapagos Islands and their relation to oceanographic processes in the tropical Pacific. In: James,
M.J. (Ed.), Galapagos Marine Invertebrates. Plenum Press, New York, pp. 9–33.
Chavez, F.P., Buck, K.R., Coale, K.H., Martin, J.H., DiTullio, G.R., Welschmeyer, N.A., Jacobson, A.C., Barber, R.T., 1991. Growth
rates, grazing, sinking, and iron limitation of equatorial Pacific phytoplankton. Limnology and Oceanography 36, 1816–1833.
Chavez, F.P., Buck, K.R., Service, S.K., Newton, J., Barber, R.T., 1996a. Phytoplankton variability in the central and eastern tropical
Pacific. Deep-Sea Research II 43, 835–870.
Chavez, F.P., Ryan, J., Lluch-Cota, S.E., Niquen, C.M., 2003. From Anchovies to Sardines and Back: multidecadal change in the Pacific
Ocean. Science 299, 217–221.
Chavez, F.P., Service, S.K., Buttrey, S.E., 1996b. Temperature-nitrate relationships in the central and eastern tropical Pacific. Journal of
Geophysical Research 101 (C9), 553–563.
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 313

Chavez, F.P., Smith, S.L., 1995. Biological and chemical consequences of open ocean upwelling. In: Summerhayes, C.P., Emeis, K.-C.,
Angel, M.V., Smith, R.L., Zeitschel, B. (Eds.), Upwelling in the Ocean: Modern Processes and Ancient Records. Wiley, Chichester, pp.
149–169.
Chavez, F.P., Strutton, P.G., Friederich, G.E., Feely, R.A., Feldman, G.C., Foley, D.G., McPhaden, M.J., 1999. Biological and chemical
response of the equatorial Pacific Ocean to the 1997-98 El Niño. Science 286, 2126–2131.
Chavez, F.P., Toggweiler, J.R., 1995. Physical estimates of global new production: the upwelling contribution. In: Summerhayes, C.P.,
Emeis, K.-C., Angel, M.V., Smith, R.L., Zeitschel, B. (Eds.), Upwelling in the Ocean: Modern Processes and Ancient Records. Wiley,
Chichester, pp. 313–320.
Chelton, D.B., Freilich, M.H., Esbensen, S.K., 2000. Satellite observations of the wind jets off the Pacific coast of Central America. Part I:
Case studies and statistical characteristics. Monthly Weather Review 128, 1993–2018.
Chelton, D.B., Schlax, M.G., Freilich, M.H., Milliff, R.F., 2004. Satellite measurements reveal persistent small-scale features in ocean
winds. Science 303, 978–983.
Cline, J.D., Richards, F.A., 1972. Oxygen deficient conditions and nitrate reduction in the eastern tropical North Pacific Ocean.
Limnology and Oceanography 17, 885–900.
Coale, K.H., Fitzwater, S.E., Gordon, R.M., Johnson, K.S., Barber, R.T., 1996a. Control of community growth and export production by
upwelled iron in the equatorial Pacific. Nature 379, 621–624.
Coale, K.H., Johnson, K.S., Fitzwater, S.E., Gordon, R.M., Tanner, S., Chavez, F.P., Ferioli, L., Sakamoto, C., Rogers, P., Millero, F.,
Steinberg, P., Nightingale, P., Cooper, D., Cochlan, W., Landry, M., Constantinou, J., Rollwagen, G., Transvina, A., Kudela, R.,
1996b. A massive phytoplankton bloom induced by an ecosystem-scale iron fertilization experiment in the equatorial Pacific Ocean.
Nature 383, 495–501.
Coale, K.H., Johnson, K.S., Fitzwater, S.E., Blain, S.P.G., Stanton, T.P., Coley, T.L., 1998. IronEx-I, an in situ iron-enrichment
experiment: experimental design, implementation and results. Deep-Sea Research II 45, 919–945.
Codispoti, L.A., 2002. Is the ocean losing nitrate? Nature 376. doi:10.1038/376724a0.
Codiposti, L.A., Friederich, G.E., Packard, T.T., Glover, H.E., Kelly, P.J., Spinrad, R.W., Barber, R.T., Elkin, J.W., Ward, B.B.,
Lipschultz, F., Lostaunau, N., 1986. High nitrate levels off northern Peru: a signal of instability in the marine denitrification rate.
Science 233, 1200–1202.
Codiposti, L.A., Richards, F.A., 1976. An analysis of the horizontal regime of denitrification in the eastern tropical North Pacific.
Limnology and Oceanography 21, 379–388.
Conkright, M.E., Levitus, S., O’Brien, T., Boyer, T.P., Stephens, C., Johnson, D., Stathoplos, L., Baranova, O., Antonov, J., Gelfeld, R.,
Burney, J., Rochester, J., Forgy, C., 1998. World Ocean Database 1998: Documentation and Quality Control, Version 1.2. Internal
Report 14, National Oceanographic Data Center.
Cromwell, T., 1958. Thermocline topography, horizontal currents and ‘‘ridging’’ in the eastern tropical Pacific. Bulletin of the Inter-
American Tropical Tuna Commission 111, 135–164.
Cullen, J.J., Lewis, M.R., Davis, C.O., Barber, R.T., 1992. Photosynthetic characteristics and estimated growth rates indicate grazing is
the approximate control of primary production in the equatorial Pacific. Journal of Geophysical Research 97, 639–654.
Duce, R.A., Tindale, N.W., 1991. Atmospheric transport of iron and its deposition in the ocean. Limnology and Oceanography 36, 1715–1726.
Dugdale, R.C., 1972. Chemical oceanography and primary productivity in upwelling regions. Geoforum 2, 47–61.
Dugdale, R.C., Goering, J.J., 1967. Uptake of new and regenerated forms of nitrogen in primary productivity. Limnology and
Oceanography 12, 196–206.
Dugdale, R.C., Jones, B.H., MacIsaac, J.J., Goering, J.J., 1979. Interactions of primary nutrient and carbon uptake in the Peruvian
current: a review and synthesis of laboratory and field results. Coastal Upwelling Ecosystems Analysis Technical Report 51, 40.
Dugdale, R.C., Wilkerson, F.P., 1998. Silicate regulation of new production in the equatorial Pacific upwelling. Nature 391, 270–273.
Dugdale, R.C., Wilkerson, F.P., Barber, R.T., Chavez, F.P., 1992. Estimating new production in the equatorial Pacific Ocean at 150W.
Journal of Geophysical Research 97, 681–686.
Dugdale, R.C., Wilkerson, F.P., Minas, H.J., 1995. The role of the silicate pump in driving new production. Deep-Sea Research 42, 697–
719.
Dugdale, R.C., Wischmeyer, A.G., Wilkerson, F.P., Barber, R.T., Chai, F., Jiang, M.S., Peng, T.H., 2002. Meridional asymmetry of
source nutrients to the equatorial Pacific upwelling ecosystem and its potential impact on ocean–atomosphere CO2 flux; a data and
modeling approach. Deep-Sea Research II 49, 2513–2531.
Enfield, D.B., Allen, J.S., 1980. On the structure and dynamics of monthly mean sea level anomalies along the Pacific coast of North and
South America. Journal of Physical Oceanography 10, 557–578.
FAO, 1993. FAO Yearbook of Fishery Statistics 1991Yearbook of Fishery Statistics, vol. 73. Food and Agricultural Organization of the
United Nations, Rome.
Feely, R.A., Boutin, J., Cosca, C.E., Dandonneau, Y., Etcheto, J., Inoue, H.Y., Ishii, M., Le Quere, C., Mackey, D.J., McPhaden, M.J.,
Metzl, N., Poisson, A., Wanninkhof, R., 2002. Seasonal and interannual variability of CO2 in the equatorial Pacific. Deep-Sea
Research II 49, 2443–2469.
Feely, R., Wanninkhof, A.R., Goyet, C., Archer, D.E., Takahashi, T., 1997. Variability of CO2 distributions and sea-air fluxes in the
central and eastern equatorial Pacific during the 1991–1994 El Niño. Deep-Sea Research II 44, 1851–1867.
Feely, R.A., Wanninkhof, R., Takahashi, T., Tans, P., 1999. Influence of El Niño on the equatorial Pacific contribution to atmospheric
CO2 accumulation. Nature 398, 597–601.
Feldman, G., Clark, D., Halpen, D., 1984. Satellite color observations of the phytoplankton distribution in the eastern equatorial Pacific
during the 1982–1983 El Niño. Science 226, 1069–1071.
314 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Fiadero, M., Strickland, J.D.H., 1968. Nitrate reduction and the occurrence of a deep nitrite maximum in the ocean off the west coast of
South America. Deep-Sea Research 26, 187–201.
Fiedler, P.C., 1994. Seasonal and interannual variability of coastal zone color scanner phytoplankton pigments and winds in the eastern
tropical Pacific. Journal of Geophysical Research – Oceans 99, 18371–18384.
Fiedler, P.C., 2002a. The annual cycle and biological effects of the Costa Rica Dome. Deep-Sea Research I 49, 321–338.
Fiedler, P.C., 2002b. Environmental change in the eastern tropical Pacific Ocean: review of ENSO and decadal variability. Marine
Ecological Progress Series 244, 265–283.
Fiedler, P.C., Chavez, F.P., Behringer, D.W., Reilly, S.B., 1992. Physical and biological effects of Los Niños in the eastern tropical Pacific,
1986-1989. Deep-Sea Research 39, 199–219.
Fiedler, P.C., Philbrick, V., Chavez, F.P., 1991. Oceanic upwelling and productivity in the eastern tropical Pacific. Limnology and
Oceanography 36, 1834–1850.
Fiedler, P.C., Talley, L.D., 2006. Hydrography of the eastern tropical Pacific: a review. Progress in Oceanography 69 (2–4), 143–180.
Fleming, R.H., 1957. General features of the ocean. In: Hedgepeth, J.W. (Ed.), Treatise on Marine Ecology and Paleoecology, vol. 1. The
Geological Society of America, pp. 87–108.
Foley, D.G., Dickey, T.D., McPhaden, M.J., Bidigare, R.R., Lewis, M.R., Barber, R.T., Lindley, S.T., Garside, C., Manov, D.V.,
McNeil, J.D., 1997. Longwaves and primary productivity variations in the equatorial Pacific at 0, 140W. Deep-Sea Research II 44,
1801–1826.
Friedrichs, M.A.M., Hofmann, E.E., 2001. Physical control of biological processes in the central equatorial Pacific Ocean. Deep-Sea
Research I 48, 1023–1069.
Frost, B.W., 1987. Grazing control of phytoplankton stock in the open subarctic Pacific Ocean: a model assessing the role of
mesozooplankton, particularly the calanoid Neocalanus spp. Marine Ecology Progress Series 39, 49–68.
Galbraith, E.D., Kienast, M., Pedersen, T.F., Calvert, S.E., 2004. Glacial–interglacial modulation of the marine nitrogen cycle by oxygen
supply to intermediate waters. Paleoceanography 19. doi:10.1029/2003PA001000.
Galloway, J.N., 2005. The global nitrogen cycle. In: Schlesinger, W.H. (Ed.), Biogeochemistry. Elsevier, Amsterdam, pp. 557–583.
Gao, Y., Fan, S.-M., Sarmiento, J.L., 2003. Aeolian iron input to the ocean through precipitation scavenging: a modeling perspective and
its implication for natural iron fertilization in the ocean. Journal of Geophysical Research 108. doi:10.1029/2002JD002420.
Ginoux, P., Chin, M., Tegen, J., Prospero, J.M., Holben, B., Dubovik, O., Lin, S.J., 2001. Sources of distributions of dust aerosols
simulated with the GOCART model. Journal of Geophysical Research 106, 20255–20274.
Goldman, J.C., 1993. Potential role of large oceanic diatoms in new primary production. Deep-Sea Research 40, 159–168.
Gordon, R.M., Coale, K.H., Johnson, K.S., 1997. Iron distribution in the equatorial Pacific: implications for a new production.
Limnology and Oceanography 42, 419–431.
Graham, H.W., 1941. Plankton production in relation to character of water in the open ocean Pacific. Journal of Marine Research 4, 189–197.
Gruber, N., 2004. The dynamics of the marine nitrogen cycle and its influence on atmospheric CO2. In: Follows, M., Oguz, T. (Eds.), The
Ocean Carbon Cycle and Climate, NATO ASI Series. Kluwer Academic, Dordrecht, pp. 97–148.
Gruber, N., Sarmiento, J.L., 1997. Global patterns of marine nitrogen fixation and denitrification. Global Biogeochemical Cycles 11, 235–
266.
Guillén, O., 1966. Variación de los fosfotas en el callao como medida de la producción primaria. Memoria del I Seminario Latino-
Americano sobre el Océano Pacı́fico Oriental, Universidad Nacional Mayor de San Marcos, pp. 192–198.
Guillén, O., Calienes, R., 1981a. Productividad y afloramiento frente a las aguas costeras peruanas. Boletin del Instituto del Mar del Perú,
vol. extraordinario, Proy. Investigación Cooperativa de la Anchoveta y su Ecosistema (ICANE), pp. 130–143.
Guillén, O., Calienes, R., 1981b. Upwelling off Chimbote. In: Richards, F. (Ed.), Coastal Upwelling. American Geophysical Union,
Washington, DC, pp. 312–326.
Guillén, O., Calienes, R., 1981c. Biological productivity and El Niño. In: Glantz, M., Thompson, D. (Eds.), Resource Management and
Environmental Uncertainty, Lessons from Coastal Upwelling Fisheries. Wiley, New York, pp. 255–282.
Guillén, O., Calienes, R., de Rodán, R.I., 1969. Contribución al estudio del ambrinte de la anchoveta (Engraulis ringens J.). Boletin del
Instituto del Mar del Perú 2, 52–74.
Guillén, O., Calienes, R., de Rodán, R.I., 1977. Medio ambiente y produccı´on primaria frente al área Pimentel-Chimbote. Boletin del
Instituto del Mar del Perú, Proy. OEA/IMARPE, vol. 3, pp. 107–159.
Guillén, O., de Rodán, R.I., 1968. Producción primaria dle las aquas costeras del Peru en al año 1964. Boletin del Instituto del Mar del
Perú 1, 352–376.
Harrison, W.G., Platt, T., Calienes, R., Ochoa, N., 1981. Photosynthetic parameters and primary production of phytoplankton
populations off the nothern coast of Peru. In: Richards, F. (Ed.), Coastal Upwelling. American Geophysical Union, Washington, DC,
pp. 303–311.
Hasle, G.R., 1959. A quantitative study of phytoplankton from the equatorial Pacific. Deep-Sea Research 6, 38–59.
Hoffmann, E.E., Busalacchi, A.J., O’Brien, J.J., 1981. Wind generation of the Costa Rica Dome. Science 214, 552–554.
Hutchings, L., 1992. Fish harvesting in a variable, productive environment: searching for rules or searching for exception? In: Payne,
A.I.L., Mann, K.H., Hilborn, R. (Eds.), Benguela Trophic Functioning, South African Journal of Marine Science, 12, pp. 297–318.
Hutchins, D.A., Hare, C.E., Weaver, R.S., Zhang, Y., Firme, G.F., DiTullio, G.R., Alm, M.B., Riseman, S.F., Maucher, J.M., Geesey,
M.E., Trick, C.G., Smith, G.J., Rue, E.L., Conn, J., Bruland, K.W., 2002. Phytoplankton iron limitation in the Humboldt Current and
Peru Upwelling. Limnology and Oceanography 47, 997–1011.
Huyer, A., Smith, R.L., Paluszkiewicz, T., 1987. Coastal upwelling off Peru during normal and El Niño times, 1981–1984. Journal of
Geophysical Research 92 (C13), 14297–14307.
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 315

Johnson, K.S., Chavez, F.P., Friederich, G.E., 1999. Continental-shelf sediment as a primary source of iron for coastal phytoplankton.
Nature 398, 697–700.
Jordán, R., 1971. Distribution of anchoveta (Engravlis ringens J.) in relation to the environment. Investagaciones Pesqueras 35, 113–126.
Kamykowski, D., Zentara, S.-J., Morrison, J.M., Switzer, A.C., 2002. Dynamic global patterns of nitrate, phosphate, silicate, and iron
availability and phytoplankton community composition from remote sensing. Global Biogeochemical Cycles 16. doi:10.1029/
2001GB001640.
Karl, D., Letelier, R., Hebel, D., Tupas, L., Dore, J., Christian, J., Winn, C., 1995. Ecosystem changes in the North Pacific Subtropical
Gyre attributed to the 1991–1992 El Niño. Nature 373, 230–234.
Karl, D., Letelier, R., Tupas, L., Dore, J., Christian, J., Hebel, D., 1997. The role of nitrogen fixation in the biogeochemical cycling in the
subtropical North Pacific Ocean. Nature 388, 533–538.
Karl, D., Michaels, A., Bergman, B., Capone, D., Carpenter, E., Letelier, R., Lipschultz, F., Paerl, H., Sigman, D., Stal, L., 2002.
Dinitrogen fixation in the world’s oceans. Biogeochemistry 57/58, 47–98.
Kessler, W.S., 2002. Mean three-dimensional circulation in the northeast tropical Pacific. Journal of Physical Oceanography 32, 2457–
2471.
Kessler, W.S., 2006. The circulation of the eastern tropical Pacific: a review. Progress in Oceanography 69 (2–4), 181–217.
King, J.E., Demond, J., 1953. Zooplankton abundance in the central Pacific. Fisheries Bulletin 54, 111–144.
Koblentz-Mishke, O.I., Volkovinsky, V.V., Kabanova, J.G., 1970. Plankton primary production of the world ocean. In: Wooster, W.S.
(Ed.), Scientific Exploration of the South Pacific. National Academy of Sciences, Washington, DC, pp. 183–193.
Landry, M.R., Barber, R.T., Bidigare, R.R., Chai, F., Coale, K.H., Dam, H.G., Lewis, M.R., Lindley, S.T., McCarthy, J.J., Roman,
M.R., Stoecker, D.K., Verity, P.G., White, J.R., 1997. Iron and grazing constraints on primary production in the central equatorial
Pacific: an EqPac synthesis. Limnology and Oceanography 42, 405–418.
Landry, M.R., Kirchman, D.L., 2002. Microbial community structure and variability in the tropical Pacific. Deep-Sea Research II 49,
2669–2693.
Le Borgne, R., Feely, R.A., Mackey, D.J., 2002. Carbon fluxes in the equatorial Pacific: a synthesis of the JGOFS programme. Deep-Sea
Research 49, 2425–2442.
Lee, K., Karl, D.M., Wanninkhof, R., Zhang, J.-Z., 2002. Global estimates of net carbon production in the nitrate-depleted tropical and
subtropical oceans. Geophysical Research Letters 29. doi:10.1029/2001GL014198.
Legeckis, R., 1977. Long waves in the eastern equatorial Pacific Ocean: a view from a geostationary satellite. Science 197, 1179–1181.
Li, W.K.W., Subba Rao, D.V., Harrison, W.G., Smith, J.C., Cullen, J.J., Irwin, B., Platt, T., 1983. Autotrophic picoplankton in the
tropical ocean. Science 219, 292–295.
Lluch-Cota, S.E., Alvarez-Borrego, S., Santamaria-del Angel, E.M., Muller-Karger, F.E., Hernandez-Vazquez, S., 1997. The Gulf of
Tehuantepec and adjacent areas: spatial and temporal variation of satellite-derived photosynthetic pigments. Ciencias Marinas 23,
329–340.
Mackey, D.J., Blanchot, J., Higgins, H.W., Neveux, J., 2002. Phytoplankton abundances and community structure in the equatorial
Pacific. Deep-Sea Research II 49, 2561–2582.
Mantua, N.J., Hare, S.R., 2002. The Pacific decadal oscillation. Journal of Oceanography 58, 35–44.
Martin, J.H., 1990. Glacial-interglacial CO2 change: the iron hypothesis. Paleoceanography 5, 1–13.
Martin, J.H., Coale, K.H., Johnson, K.S., Fitzwater, S.E., Gordon, R.M., Tanner, S.J., Hunter, C.N., Elrod, V.A., Nowicki, J.L., Barber,
R.T., Lindley, S., Watson, A.J., Van Skoy, K., Law, C.S., Liddicoat, M.I., Ling, R., Stanton, T., Stockel, J., Collins, C., Anderson, A.,
Bidigare, R., Ondrusek, M., Latatsa, M., Millero, F., Lee, K., Yao, W., Zhang, J., Frederich, G., Sakamoto, C., Chavez, F., Buck, K.,
Kolber, Z., Greene, R., Falkowski, P., Chisolm, S.W., Hoge, F., Swift, R., Yungel, J., Turner, S., Nightingale, P.D., Hatton, A., Liss,
P., Tindale, N.W., 1994. Testing the iron hypothesis in ecosystems of the equatorial Pacific Ocean. Nature 371, 123–129.
McCarthy, J.J., Garside, C., Nevins, J.L., Barber, R.T., 1996. New production along 140W in the equatorial Pacific during and following
the 1992 El Niño event. Deep-Sea Research II 4–6, 1065–1093.
McClain, C.R., Christian, J.R., Signorini, S.R., Lewis, M.R., Asanuma, I., Turk, D., Dupouy-Douchement, C., 2002. Satellite ocean-color
observations of the tropical Pacific Ocean. Deep-Sea Research II 49, 2533–2560.
McGowan, J.A., 2004. Sverdrup’s biology. Oceanography 17, 106–112.
McPhaden, M.J., Yu, X., 1999. Equatorial waves and the 1997–1998 El Niño. Geophysical Research Letters 26, 2961–2964.
McPhaden, M.J., Zhang, D.X., 2002. Slowdown of the meridional overturning circulation in the upper Pacific Ocean. Nature 415, 603–
608.
Meskhidze, N., Chameides, W.L., Nenes, A., Chen, G., 2003. Iron mobilization in mineral dust: can anthropogenic SO2 emissions affect
ocean productivity? Geophysical Research Letters 30, 2085. doi:10.1029/2003GL018035.
Mestas-Nuñez, A.M., Miller, A., 2006. Interdecadal variability and climate change in the eastern tropical Pacific: a review. Progress in
Oceanography 69 (2–4), 267–284.
Montecino, V., Strub, T., Chavez, F.P., Thomas, A., Tarazona, J., Baumgartner, T., in press. Bio-Physical interactions off Western South-
America. In: Robinson, A., Brink, K. (Eds.), The Sea, vol. 14. Harvard University Press, Cambridge.
Montoya, J.P., Holl, C.M., Zehr, J.P., Hansen, A., Villareal, T.A., Capone, D.G., 2004. High rates of N2 fixation by unicellular
diazotrophs in the oligotrophic Pacific Ocean. Nature 430, 1027–1031.
Moore, J.K., Doney, S.C., Glover, D.M., Fung, I.Y., 2002. Iron cycling and nutrient-limitation patterns in surface waters of the World
Ocean. Deep-Sea Research II 49, 463–507.
Morales, C.E., Hormazabal, S.E., Blanco, J.L., 1999. Interannual variabllity in the mesoscale distribution of the depth of the upper
boundary of the oxygen minimum layer off northern Chile (18-24S). Journal of Marine Research 57, 909–932.
316 J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317

Murray, J.W., Barber, R.T., Bacon, M.P., Roman, M.R., Feely, R.A., 1994. Physical and biological controls on carbon cycling in the
equatorial Pacific. Science 266, 58–65.
Murray, J.W., Johnson, E., Garside, C., 1995. A US JGOFS Process Study in the Equatorial Pacific (EqPac): introduction. Deep-Sea
Research II 42, 275–293.
Nemi, A., 1979. Blue-green algal blooms and N:P ratios in the Baltic Sea. Acta Botanica Fennica 110, 57–61.
Ochoa, N., Gomez, O., 1987. Dinoflagellates as indicators of water masses during El Niño, 1982–1983. Journal of Geophysical Research
92, 14355–14367.
Pak, H., Zaneveld, J.R.V., 1974. Equatorial front in the eastern Pacific Ocean. Journal of Physical Oceanography 4, 570–578.
Peña, M.A., Lewis, M.R., Cullen, J.J., 1994. New production in the warm waters of the tropical Pacific Ocean. Journal of Geophysical
Research 99, 14255–14268.
Pennington, J.T., Chavez, F.P., 2000. Seasonal fluctuations of temperature, salinity, nitrate, chlorophyll and primary production at station
H3/M1 over 1989–1996 in Monterey Bay, California. Deep Sea Research Part II 47, 947–974.
Pettke, T., Lee, D.-C., Halliday, A.N., Rea, D.K., 2002. Radiogenic Hf compositions of continental eolian dust from Asia, its variability
and its implications for seawater Hf. Earth and Planetary Science Letters 202, 453–464.
Philander, S.G.H., 1978. Instabilities of zonal equatorial currents 2. Journal of Geophysical Research 83, 3679–3682.
Philander, S.G.H., 1999. A review of tropical ocean–atmosphere interactions. Tellus 51, 71–90.
Philander, S.G.H., Gu, D., Halpern, D., Lambert, G., Lau, N.-C., Li, T., Pacanowski, R.C., 1996. Why the ITCZ is mostly north of the
equator. Journal of Climate 9, 2958–2972.
Probyn, T.A., 1992. The inorganic nitrogen nutrition of phytoplankton in the southern Benguela: new production, phytoplankton size and
implications for pelagic foodwebs. South African Journal of Marine Science 12, 411–420.
Prospero, J.M., 1996. The atmospheric transport of particles to the ocean. In: Ittekkot, V., Schafer, P., Honjo, S., Depetris, P.J. (Eds.),
Particle Flux in the Oceans. Wiley, Chinchester, pp. 19–52.
Rocap, G., Larimer, F., Lamerdin, J., Malfatti, S., Chain, P., Ahlgren, N., Arellano, A., Coleman, M., Hauser, L., Hess, W., Johnson, Z.,
Land, M., Lindell, D., Post, A., Regala, W., Shah, M., Shaw, S., Steglich, C., Sullivan, M., Ting, C., Tolonen, A., Webb, E., Zinser, E.,
Chisholm, S., 2003. Genome divergence in two Prochlorococcus ecotypes reflects oceanic niche differentiation. Nature 424, 1042–1047.
Rojas de Mendiola, B., 1981. Seasonal phytoplankton distribution along the Peru coast. In: Richards, F.A. (Ed.), Coastal Upwelling
Research. American Geophysical Union, Washington, DC, pp. 345–356.
Ryan, J.P., Polito, P.S., Strutton, P.G., Chavez, F.P., 2002. Unusual large-scale phytoplankton blooms in the equatorial Pacific. Progress
in Oceanography 55, 263–285.
Ryan, J., Ueki, I., Chao, Y., Zang, H., Polito, P., Chavez, F.P., in press. Western Pacific modulation of large phytoplankton blooms in the
central and eastern equatorial Pacific. Journal of Geophysical Research.
Ryther, J.H., 1969. Photosynthesis and fish production in the sea. Science 166, 71–76.
Sanchez, G., Calienes, R., Zuta, S., 2000. The 1997–1998 El Niño and its effects on the coastal marine ecosystem off Peru. California
Cooperative Fisheries Investigations Reports 41, 62–86.
Sorokin, Y.I., Sukhanova, I.N., Konovalova, G.V., Pavelyeva, E.B., 1977. Primary production and the phytoplankton of the equatorial
region in the eastern Pacific Ocean. Polish Archives of Hydrobiology 24, 145–162.
Spear, L.B., Balance, L.T., Ainley, D.G., 2001. Response of seabirds to thermal boundaries in the tropical Pacific: the thermocline versus
the Equatorial Front. Marine Ecological Progress Series 219, 275–289.
Strickland, J.D.H., Eppley, R.W., De Mendiola, B.R., 1969. Phytoplankton populations, nutrients and photosynthesis in Peruvian coastal
waters. Boletin del Instituto del Mar del Perú 2, 37–45.
Strutton, P.G., Chavez, F.P., 2000. Primary productivity in the equatorial Pacific during the 1997–1998 El Niño. Journal of Geophysical
Research 105 (C11), 26089–26101.
Strutton, P.G., Ryan, J.P., Chavez, F.P., 2001. Enhanced chlorophyll associated with tropical instability waves in the equatorial Pacific.
Geophysical Research Letters 28, 2005–2008.
Sverdrup, H.U., 1955. The place of physical oceanography in oceanographic research. Journal of Marine Research 14, 287–294.
Sverdrup, H.U., Johnson, M.W., Fleming, R.H., 1942. The Oceans. Prentice Hall, Englewood Cliffs, NJ.
Takahashi, T., Feely, R.A., Weiss, R.F., Wanninkhof, R.H., Chipman, D.W., Sutherland, S.C., Takahashi, T.T., 1997. Global air–sea flux
of CO2: an estimate based on measurements of sea–air pCO2 difference. Proceedings of the National Academy of Science, USA 94,
8292–8299.
Takahashi, T., Sutherland, S.C., Feely, R.A., Cosca, C.E., 2003. Decadal variation of the surface water pCO2 in the western and central
equatorial Pacific. Science 302, 852–855.
Takahashi, T., Sutherland, S.C., Sweeney, C., Poisson, A., Metzl, N., Tilbrook, B., Bates, N., Wanninkhof, B., Feely, R.A., Sabine, C.,
Olafsson, J., Nojiri, Y., 2002. Global sea-air CO2 flux based on climatological surface ocean pCO2, and seasonal biological and
temperature effects. Deep-Sea Research II 49, 1601–1622.
Tans, P., Fung, I., Takahashi, T., 1990. Observational constraints on the global atmospheric CO2 budget. Science 247, 1431–1438.
Thomas, W.H., 1966. On denitrification in the northeastern tropical Pacific Ocean. Deep-Sea Research 13, 1109–1114.
Thomas, W.H., 1979. Anomalous nutrient chlorophyll interrelationships in the offshore eastern tropical Pacific Ocean. Journal of Marine
Research 37, 327–335.
Toggweiler, J.R., Carson, S., 1995. What are upwelling systems contributing to the ocean’s carbon and nutrient budgets. In:
Summerhayes, C.P., Emeis, K.-C., Angel, M.V., Smith, R.L., Zeitschel, B. (Eds.), Upwelling in the Ocean: Modern Processes and
Ancient Records. Wiley, Chichester, pp. 337–360.
Tomczak, M., Godfrey, J.S., 1994. Regional Oceanography: An Introduction. Daya Publishing House, Delhi, 390 pp.
J.T. Pennington et al. / Progress in Oceanography 69 (2006) 285–317 317

Turk, D., McPhaden, M.J., Busalacchi, A.J., Lewis, M.R., 2001. Remotely sensed biological production in the equatorial Pacific. Science
293, 471–474.
Tyrrell, T., 1999. The relative influences of nitrogen and phosphorus on oceanic primary production. Nature 400, 525–531.
Walsh, J.J., 1976. Herbivory as a factor in patterns of nutrient utilization in the sea. Limnology and Oceanography 21, 1–13.
Wang, C., Fiedler, P.C., 2006. ENSO variability in the eastern tropical Pacific: a review. Progress in Oceanography 69 (2–4), 239–266.
Wang, O., Fukumori, I., Lee, T., Johnson, G.C., 2004. Eastern equatorial Pacific Ocean T-S variations with El Niño. Geophysical
Research Letters 31, L04305. doi:10.1029/2003GL019087.
Wang, X., Christian, J.R., Murtugudde, R., Busalacchi, A.J., 2005. Ecosystem dynamics and export production in the central and eastern
equatorial Pacific: a modeling study of impact of ENSO. Geophysical Research Letters 32. doi:10.1029/2004GL021538.
Wanninkhof, R., Feely, R.A., Chen, H., Cosca, C., Murphy, P., 1992. Surface water fCO2 in the eastern equatorial Pacific during the
1992–1993 El Niño. Journal of Geophysical Research 101, 16333–16343.
Weinberg, E.D., 1989. Cellular regulation of iron assimilation. Quarterly Review of Biology 64, 261–290.
Wells, M.L., Vallis, G.K., Silver, E.A., 1999. Tectonic processes in Papua New Guinea and past productivity in the eastern equatorial
Pacific Ocean. Nature 398, 601–604.
Wilkerson, F.P., Dugdale, R.C., 1992. Measurements of nitrogen productivity in the equatorial Pacific at 150W. Journal of Geophysical
Research 97, 669–680.
Willett, C.S., Leben, R., Lavı´n, M.F., 2006. Eddies and mesoscale processes in the eastern tropical Pacific: a review. Progress in
Oceanography 69 (2–4), 218–238.
Wilson, C., Adamec, D., 2001. Correlations between surface chlorophyll and sea surface height in the tropical Pacific during the 1997 –
1999 El Niño-Southern Oscillation event. Journal of Geophysical Research 106, 31175–31188.
Wyrtki, K., 1964. Upwelling in the Costa Rica Dome. Fishery Bulletin 63, 355–372.
Wyrtki, K., 1966. Oceanography of the eastern equatorial Pacific Ocean. Oceanography and Marine Biology Annual Review 4, 33–68.
Wyrtki, K., 1981. An estimate of equatorial upwelling in the Pacific. Journal of Physical Oceanography 11, 1205–1214.
Wyrtki, K., Wenzel, J., 1984. Possible gyre–gyre interactions in the Pacific Ocean. Nature 309, 538–540.
Xie, L., Hsieh, W.W., 1995. The global distribution of wind-induced upwelling. Fisheries Oceanography 4, 52–67.
Yoder, J.A., Ackleson, S.G., Barber, R.T., Flament, P., Balch, W.M., 1994. A line in the sea. Nature 6499, 689–692.
Zehr, J.P., Ward, B.B., 2002. Nitrogen cycling in the ocean: new perspectives on processes and paradigms. Applied and Environmental
Microbiology 68, 1015–1024.
Zehr, J.P., Waterbury, J.B., Turner, P.J., Montoya, J.P., Omoregie, E., Steward, G.F., Hansen, A., Karl, D.M., 2001. Unicellular
cyanobacteria fix N2 in the subtropical North Pacific Ocean. Nature 412, 635–638.
Zuta, S., Guillén, O., 1970. Oceanografı́a de las aguas costeras del Perú. Boletin del Instituto del Mar del Perú 2, 159–323.

You might also like