Multiple thermal fronts near the Patagonian shelf break



Similar documents
Cross-front variations in adult abundance and recruitment of Patagonian scallop (Zygochlamys patagonica) at the SW Atlantic Shelf Break Front

Brazil Malvinas Frontal System as seen from 9 years of advanced very high resolution radiometer data

NOTES AND CORRESPONDENCE. Interaction of Mesoscale Variability with Large-Scale Waves in the Argentine Basin

Daily High-resolution Blended Analyses for Sea Surface Temperature

Present Status of Coastal Environmental Monitoring in Korean Waters. Using Remote Sensing Data

Interannual to decadal changes in the western South Atlantic s surface circulation

Temporal variation in snow cover over sea ice in Antarctica using AMSR-E data product

Intra-seasonal and Annual variability of the Agulhas Current from satellite observations

CLIMATOLOGICAL DATA FROM THE WESTERN CANADIAN ODAS MARINE BUOY NETWORK

ENSO Cycle: Recent Evolution, Current Status and Predictions. Update prepared by Climate Prediction Center / NCEP 9 May 2011

Underwater gliders reveal rapid arrival of El Niño effects off California s coast

Climate Extremes Research: Recent Findings and New Direc8ons

Dynamics IV: Geostrophy SIO 210 Fall, 2014

ALMOFRONT 2 cruise in Alboran sea : Chlorophyll fluorescence calibration

Temporal characterization of the diffuse attenuation coefficient in Abrolhos Coral Reef Bank, Brazil

Approaches to biogeographic classification of the world s oceans. Marjo Vierros United Nations University Institute of Advanced Studies

The concepts developed in this standard include the following: Oceans cover about 70% of the surface of the Earth.

How to analyze synoptic-scale weather patterns Table of Contents

South Africa. General Climate. UNDP Climate Change Country Profiles. A. Karmalkar 1, C. McSweeney 1, M. New 1,2 and G. Lizcano 1

CTD Oceanographic Tags

ATMS 310 Jet Streams

Original DOI: /2008JC005018

Long-term hydrographic variation in the Skagerrak based on the section Torungen Hirtshals

Radiative effects of clouds, ice sheet and sea ice in the Antarctic

ENVIRONMENTAL STRUCTURE AND FUNCTION: CLIMATE SYSTEM Vol. II - Low-Latitude Climate Zones and Climate Types - E.I. Khlebnikova

Huai-Min Zhang & NOAAGlobalTemp Team

Near Real Time Blended Surface Winds

The relationships between Argo Steric Height and AVISO Sea Surface Height

2008 Global Surface Temperature in GISS Analysis

VOCALS-CUpEx: The Chilean Upwelling Experiment

Hurricane-forced upwelling and chlorophyll a enhancement within cold-core cyclones in the Gulf of Mexico

Chapter Overview. Seasons. Earth s Seasons. Distribution of Solar Energy. Solar Energy on Earth. CHAPTER 6 Air-Sea Interaction

Coastal Southern Ocean: A strong anthropogenic CO 2 sink

Canadian Prairie growing season precipitation variability and associated atmospheric circulation

Chapter 15. Hydrology of the Atlantic Ocean

What Causes Climate? Use Target Reading Skills

How Do Oceans Affect Weather and Climate?

Temporal and spatial evolution of the Antarctic sea ice prior to the September 2012 record maximum extent

Report EU BASIN Kickoff Meeting, Copenhagen, Denmark

Develop a Hybrid Coordinate Ocean Model with Data Assimilation Capabilities

Ocean Floor Continental Slope Begins at the Continental Shelf Very sharp drop to depths over 2 miles Covered with thick layers of sand, mud, and rocks

Lecture 4: Pressure and Wind

Mechanisms of an extraordinary East Asian summer monsoon event in July 2011

Plotting Earthquake Epicenters an activity for seismic discovery

Real-time Ocean Forecasting Needs at NCEP National Weather Service

Step 2: Learn where the nearest divergent boundaries are located.

Non-parametric estimation of seasonal variations in GNSS-derived time series

BIO-OPTICAL MODELING OF PRIMARY PRODUCTION FROM SEAWIFS OCEAN COLOR DATA FOR THE WESTERN ANTARCTIC PENINSULA REGION

Scholar: Elaina R. Barta. NOAA Mission Goal: Climate Adaptation and Mitigation

Basics of weather interpretation

ENSO: Recent Evolution, Current Status and Predictions. Update prepared by: Climate Prediction Center / NCEP 29 June 2015

Mesoscale variations of sea surface temperature and ocean color patterns at the Mid Atlantic Bight shelfbreak

Queensland rainfall past, present and future

Chapter 3: Weather Map. Weather Maps. The Station Model. Weather Map on 7/7/2005 4/29/2011

Slide 1. Slide 2. Slide 3

Coral Reef Watch A Satellite View. AE Strong. Satellite SST Anomalies. January 2002 May 2003

Operational Monitoring of Mesoscale Upper Layer Circulation Fields with Multi-Satellite Technology in the Southwestern Atlantic Ocean

Extra-Tropical Cyclones in a Warming Climate:

Indian Ocean and Monsoon

Future needs of remote sensing science in Antarctica and the Southern Ocean: A report to support the Horizon Scan activity of COMNAP and SCAR

2. The map below shows high-pressure and low-pressure weather systems in the United States.

Map Skills. Before Coming to Lab: Read Appendix III (pages ) in Trujillo & Thurman, 11 th ed.

The Definition of El Niño

ARI: The water-mass signature and pathways of Greenland Ice Sheet meltwater in the Arctic and North Atlantic as inferred by an inverse method

Supporting Online Material for

OCEANIC CIRCULATION 434 OCEAN CURRENTS

The Oceans Role in Climate

Thompson/Ocean 420/Winter 2005 Tide Dynamics 1

Variability of coastal ocean processes along the west coast of India

IGAD CLIMATE PREDICTION AND APPLICATION CENTRE

CLIMATE CHANGE, HUMAN SYSTEMS, AND POLICY - Vol.I - Effects of Global Warming on Marine Ecosystems - G.G. Matishov

Dynamic topography of the ice covered Arctic Ocean from ICESat

ESCI 107/109 The Atmosphere Lesson 2 Solar and Terrestrial Radiation

Can latent heat release have a negative effect on polar low intensity?

Satellite SST Product Development Proposal

TIDES. 1. Tides are the regular rise and fall of sea level that occurs either once a day (every 24.8 hours) or twice a day (every 12.4 hours).

Chapter 3: Weather Map. Station Model and Weather Maps Pressure as a Vertical Coordinate Constant Pressure Maps Cross Sections

Climate and Global Dynamics National Center for Atmospheric Research phone: (303) Boulder, CO 80307

Bipolar Atlantic Thermohaline Circulation (BIAC) a IPY proposal coordinated by Tor Gammelsrød Geophysical Institute, University of Bergen

James Hansen, Reto Ruedy, Makiko Sato, Ken Lo

Jessica Blunden, Ph.D., Scientist, ERT Inc., Climate Monitoring Branch, NOAA s National Climatic Data Center

WILLOCHRA BASIN GROUNDWATER STATUS REPORT

User s Guide. Zuniga Point. Point Loma

Ocean in Motion 2: What Causes Ocean Currents and How Do We Measure Them?

The Integration of Hydrographic and Oceanographic Data in a Marine Geographic Information System U.S. Hydro 2015

Biological Fronts in the Strait of Georgia, British Columbia, and Their Relation to Recent Measurements of Primary Productivity

California Standards Grades 9 12 Boardworks 2009 Science Contents Standards Mapping

Using LIDAR to monitor beach changes: Goochs Beach, Kennebunk, Maine

Tropical Cyclone Climatology

HIGH RESOLUTION SATELLITE IMAGERY OF THE NEW ZEALAND AREA: A VIEW OF LEE WAVES*

PERSONAL VIEW (editorial)

Regional climate change experiments over southern South America. II: Climate change scenarios in the late twenty-first century

Fundamentals of Climate Change (PCC 587): Water Vapor

SIXTH GRADE WEATHER 1 WEEK LESSON PLANS AND ACTIVITIES

RESPONSE OF DOMINANT SPECIES IN COASTAL AND OCEANIC REGIONS IN PERU

Southern AER Atmospheric Education Resource

Examining the Recent Pause in Global Warming

Gravitational potential

Chapter Overview. Bathymetry. Measuring Bathymetry. Echo Sounding Record. Measuring Bathymetry. CHAPTER 3 Marine Provinces

Transport of Antarctic krill (Euphausia superba) across the Scotia Sea. Part I: Circulation and particle tracking simulations

Transcription:

Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L02607, doi:10.1029/2007gl032066, 2008 Multiple thermal fronts near the Patagonian shelf break Bárbara C. Franco, 1,2 Alberto R. Piola, 1,2 Andrés L. Rivas, 3,4 Ana Baldoni, 5 and Juan P. Pisoni 2,3 Received 17 September 2007; revised 9 November 2007; accepted 3 December 2007; published 22 January 2008. [1] Eighteen year (1985 2002) sea surface temperature (SST) data are used to study the intraseasonal variability of the Patagonian shelf break front (SBF) in the SW South Atlantic Ocean between 39 and 44 S. The cross-shelf break SST gradients reveal distinct, previously undocumented thermal fronts located both, offshore and inshore of the SBF. Throughout the year the main SBF, identified as a band of negative SST gradient maxima (relatively strong offshore temperature decrease), forms a persistent feature located closed to the 200 m isobath, while two distinct negative gradient maxima are located inshore and offshore of this location. Daily SST images reveal the presence of three branches of cold waters whose edges delineate the above mentioned fronts. The two offshore branches closely follow lines of constant potential vorticity ( f/h) and appear to be associated with the Malvinas Current, while a third branch, located further onshore, is not steered by the bottom topography. South of 40 S the onshore branch forms a quasi permanent front parallel to the SBF. Citation: Franco, B. C., A. R. Piola, A. L. Rivas, A. Baldoni, and J. P. Pisoni (2008), Multiple thermal fronts near the Patagonian shelf break, Geophys. Res. Lett., 35, L02607, doi:10.1029/2007gl032066. 1. Introduction 1 Departamento Oceanografía, Servicio de Hidrografía Naval (SHN), Buenos Aires, Argentina. 2 Also at Departamento de Ciencias de la Atmósfera y los Océanos, FCEyN, Universidad de Buenos Aires, Buenos Aires, Argentina. 3 Centro Nacional Patagónico (CENPAT-CONICET), Puerto Madryn, Argentina. 4 Also at Facultad de Ciencias Naturales, Universidad Nacional de la Patagonia, Puerto Madryn, Argentina. 5 Instituto Nacional de Investigación y Desarrollo Pesquero (INIDEP), Mar del Plata, Argentina. Copyright 2008 by the American Geophysical Union. 0094-8276/08/2007GL032066$05.00 [2] Downstream of Drake Passage, the northernmost branch of the Antarctic Circumpolar Current, describes a sharp anticyclonic turn and penetrates about 1800 km into the western Argentine Basin forming the Malvinas Current [e.g., Peterson and Whitworth, 1989]. This is the only location in the Southern Hemisphere where a permanent injection of cold, nutrient-rich subpolar waters extends beyond 40 S [Orsi et al., 1995, Figure 11]. The intrusion creates unique oceanographic and environmental conditions in the southwest South Atlantic. Near 38 S the Malvinas Current (MC) collides with the southward flowing Brazil Current creating one of the most energetic regions of the world ocean [Gordon, 1981; Chelton et al., 1990]. The region is characterized by numerous oceanographic fronts, which are generally associated with high biological productivity [Longhurst, 1998; Saraceno et al., 2005], enhanced exchanges between the ocean and the atmosphere, and intense vertical circulation. The focus of this work is on the shelf break front (SBF) which marks the transition between the Patagonia continental shelf and the Malvinas Current. [3] North of about 50 S the cross-shelf hydrographic structure at the outer Patagonian shelf presents a persistent, surface-to-bottom front intersecting the sea bottom near the shelf break and the surface some tens of km offshore [Romero et al., 2006]. The SBF represents the transition between diluted subantarctic shelf waters and cold, salty, and relatively nutrient-rich waters of the MC. The SBF inner boundary is located between the 90 and 100 m isobaths with a W E extension of around 80 km at the surface and 40 km at the bottom [Bogazzi et al., 2005]. In summer the SBF presents strong thermal gradients (>0.08 C/km) [e.g., Martos and Piccolo, 1988; Saraceno et al., 2005], and weak salinity gradients (0.002 psu/km) [e.g., Romero et al., 2006]. Around 38 39 S the front location varies seasonally, displacing offshore during summer and onshore during spring and autumn [Carreto et al., 1995]. [4] In situ [Hubold, 1980a, 1980b; Lutz and Carreto, 1991; Carreto et al., 1995], and remote sensing measurements [Saraceno et al., 2005; Romero et al., 2006] show that the SBF is associated with a band of high chlorophyll-a (chl-a), which is indicative of high phytoplankton concentration. The region of high chl-a forms a quasi-continuous band located close to the shelf break during austral spring and summer [Podestá, 1988; Longhurst, 1998]. According to Saraceno et al. [2005], local SST gradient and chl-a maxima correspond in time and space and are located at the shelf break, emphasizing the strong topographic control on the frontal system. In addition, other studies suggest that surface chl-a blooms at the shelf break are located inshore of the thermal front [Romero et al., 2006]. The SBF coincides spatially with aggregations of zooplankton, scallops, fishes and mammals [Brunetti et al., 1998; Thomson et al., 2001; Acha et al., 2004; Bogazzi et al., 2005; Ciocco et al., 2006; Campagna et al., 2007]. High phytoplankton biomass associated with the SBF is attributed to nutrient input by upwelling processes along the shelf break [Carreto et al., 1995; Romero et al., 2006; R. Matano and E. Palma, personal communication, 2007]. [5] The thermal expression of the SBF extends approximately between 39 and 44 S and its intensity presents sharp seasonal and somewhat lower interannual variability [Saraceno et al., 2004]. The study also found the SBF closely follows the 300 m isobath. Despite the strong topographic control of the front, zonal displacements have been reported near its northern boundary. Carreto et al. [1995] report zonal displacements of the SBF near 38 L02607 1of6

39 S. Coastal-trapped waves were suggested as a possible mechanism leading to the peaks in SST and chl-a variability observed at intraseasonal frequencies along the SBF [Saraceno et al., 2005]. Given the ecological importance of the SBF and the evidence of its seasonal zonal displacements near 39 S, in this article we discuss the seasonal and intraseasonal frontal variability based on the analysis of 18 years of monthly mean SST data. 2. Data and Methods [6] Satellite-derived SST data were used to locate the SBF climatological monthly mean position and its variability. As the SBF develops at the transition from warm shelf waters to colder MC waters, it is associated with negative maximum cross-shelf break SST gradients near the shelf break. Zonal (g x ) and meridional (g y ) SST gradients were computed using a centered difference scheme based on 18 years (1985 2002) of monthly mean satellite data from the NOAA/NASA Pathfinder SST [Vazquez et al., 1998]. Cross-shelf break gradients (gsst) were then calculated projecting those components in a direction 125 from true north. The rotation is based on the mean shelf break direction between 39 and 44 S in an Equidistant Cylindrical projection. Thus, gsst = g x. cos(35 ) g y. sin(35 ). [7] We use Pathfinder best-sst values daytime data with 9.28 km resolution. The estimated average accuracy of the Pathfinder SST from most daytime matchups is 0.00 ± 0.24 C [Kearns et al., 2000]. Because in austral summer the MC advects waters much colder than those present on the continental shelf [Rivas and Piola, 2002], the highest SST gradient intensity is reached in that season. In contrast, in winter the temperature difference between the continental shelf and the MC decreases, therefore the SST gradients are significantly lower than in summer. Therefore, to analyze the intraseasonal variability of the SBF, the minimum jgsstj observed between 39 and 44 S during the climatological winter was selected as a threshold value (A. Rivas and J. P. Pisoni, unpublished data, 2007). Although the measurement accuracy is larger than the frontal threshold adopted, fronts can be effectively detected based on gradient algorithms when applied on monthly mean SSTs [Ullman and Cornillon, 1999; Hickox et al., 2000]. 3. Results and Discussion [8] Climatological (1985 2002) monthly means of satellite-derived gsst were analyzed in the area between 39 and 44 S near the shelf break at 1 latitude intervals (Figure 1). In this region the shelf break is located between 110 and 165 m depth [Parker et al., 1997]. The analysis shows that the SBF is characterized by a negative gsst maximum located near the 200 m isobath. The front is persistent throughout the year. The only exception is at 40 S, where the SBF is displaced inshore, close to the 100 m isobath. Similar inshore locations of the SBF near 40 S are apparent in the analysis of Saraceno et al. [2004]. Negative gsst along the front reach values < 0.05 C/km mainly from November to May; while gradients slightly lower than 0.02 C/km are persistent along the front during austral winter (July September). Regardless of the high gradient variability, depicted by the gsst standard deviations, mainly during austral autumn (April June) and spring (October December), the strongest gradients are located around the 200 m isobath (Figure 1). Maximum (negative) gsst are located away from the 200 m isobath mainly in the northern (39 S) and southern (42 44 S) regions, where stronger thermal fronts are apparent at other locations. At 39 S, a negative gsst maximum is observed onshore (offshore) from the mean position of the SBF mainly during May (January) (Figure 1). Previous studies [e.g., Carreto et al., 1995; C. Mauna, personal communication, 2007] have reported similar seasonal displacements of the SBF at this location. However, our analysis of gsst suggests that these gradient maxima are not associated with the main thermal band of the SBF around this latitude (Figure 2). In autumn, between 39 and 44 S, relatively intense gradients (gsst < 0.02 C/km) are observed 40 km onshore from the SBF (Figure 1). A persistent and continuous front parallel to the SBF appears from spring to autumn south of 40 S, while in winter a distinct thermal front develops offshore from the SBF. During December, January, and February at 44 S this feature appears to be displaced further east (Figure 1). This offshore front is associated with well-defined gsst peak at 43 S and 44 S during July and August (Figure 1), and is also apparent as a distinct band of negative gsst separated from the SBF by non-significant (jgsstj < 0.005 C/km) or weakly positive gradients during austral spring and summer (November February) (Figure 2). [9] The analysis of cross-shelf break gradients (gsst), rather than the SST gradient magnitude, (g x 2 +g y 2 ) 1/2, used in previous studies, reveals distinct thermal fronts around the SBF that have not been previously described in the literature. The core of the MC closely follows the 1000 m isobath [Vivier and Provost, 1999a], its equivalent-barotropic structure [Vivier and Provost, 1999b] is likely responsible for locking the frontal structure to lines of constant potential vorticity (f/h). For instance, Saraceno et al. [2004] found that the SBF approximately follows the line where f/h = 2.10 7 m 1.s 1, which lies close to the 300 m isobath. Other studies suggested that the MC bifurcates in the western Argentine Basin [Piola and Gordon, 1989]. Their quasi-synoptic density distribution from winter 1980 suggests that at about 43 S the MC upper layer splits in two branches: an offshore branch describes a sharp cyclonic loop and returns southward, while the westernmost branch continues northward along the continental slope [Piola and Gordon, 1989, Figure 3]. Our analyses of gsst reveal the surface expressions of the fronts associated with these two branches of the MC. During spring and early summer (October February), between 42 and 44 S, the thermal front between the MC branches is more evident as low, or positive gradients (gsst < 0.005 C/km) are observed between (negative) gsst maxima, associated with the SBF and the offshore front (Figure 2). During winter the offshore front is located closer to the SBF between 41 and 44 S and gsst maxima are observed at 43 and 44 S (Figure 1). [10] Inspection of daily SST images reveals the complex thermal structure and further suggests a multi-branch MC. Figure 3a illustrates a situation in which three distinct branches of relatively cold waters are observed (A, B, and C). The thermal fronts fa and fc shown in Figure 3b 2of6

Figure 1. Climatological (1985 2002) monthly mean satellite-derived cross-shelf break SST gradients (gsst, C/km) across the Patagonian shelf break at 39, 40, 41, 42, 43, and 44 S. Contour interval is 0.01 C/km and contours lower than 0.04 C/km are shown in white. Standard deviations are shown below each gsst panel, with sign inverted to ease comparison. The solid black lines indicate the locations of the 100 and 200 m isobaths. Only the 200 m isobath is shown at 44 S. The location of the selected area is depicted in Figure 3. indicate the negative peaks in gsst formed by the cold branches A and C. The two offshore temperature minima (B and C), located in the vicinity and east of the 200 m isobath, appear to be associated with the MC and are similar to the MC branches described by Piola and Gordon [1989]. The SBF, the thermal front associated with branch B, closely follows the line where f/h = 5.10 7 m 1.s 1 and the 200 m isobath (Figures 1 and 3a). However, the inshore branch (A) that appears to originate in the outer shelf near 51 S, and extends northward beyond 42 S, and the associated thermal front fa, do not seem to be effectively steered by a particular f/h contour. Both present displacements within 6.10 7 m 1.s 1 < f/h <10 6 m 1.s 1. North of 41 S fais located onshore from the 100 m isobath (Figure 1), where the 10 6 m 1.s 1 f/h contour deflects about 100 km inshore (Figure 3b). Similarly, the SBF is displaced inshore near 3of6

Figure 2. Climatological (1985 2002) monthly mean of cross-shelf break SST gradients (gsst, C/km) along the Patagonian shelf break between 39 and 44 S. The purple and white lines are the 100 and 200 m isobaths, respectively. 40 S, close to the 100 m isobath (see Figures 1 and 3). The apparently wider variability of the inshore branch is most likely due to the substantially weaker bottom slopes (and potential vorticity gradients), and therefore weaker topographic control over the outer shelf (Figure 3a). The transition between the inshore branch and shelf waters creates the strong negative gsst gradients observed in Figure 3b. South of 40 S the later feature forms a persistent 4of6

Figure 3. (a) MODIS/Aqua derived sea surface temperature (SST) image of 4 km resolution for 30 December 2006. Three branches of cold waters are labeled A, B and C. Heavy black lines are constant f/h (10 7 m 1.s 1 ) contours based on the GEBCO bathymetry (http://www.ngdc.noaa.gov/mgg/gebco/gebco.html). The region between 6.10 7 m 1.s 1 < f/h <10 6 m 1.s 1 is hatched. (b) SST cross-shelf break gradient ( C/km) between 39 and 44 S. Here fa and fc indicate bands of negative gsst formed along the western edge of cold branches A and C. The f/h (10 7 m 1.s 1 ) contours are indicated in purple. front from spring to autumn parallel to the SBF and, during autumn, it is also apparent further north (Figure 1). 4. Summary and Final Remarks [11] The intraseasonal variability of the Patagonian shelf break front was studied based on 18-year (1985 2002) of Pathfinder SST data. The analysis of cross-shelf break SST gradients reveals the presence of distinct thermal fronts in the vicinity of the SBF. The SBF is revealed by (negative) gsst maxima persistent throughout the year, located close to the 200 m isobath. In addition, other local (negative) maxima are located away from the 200 m isobath near 39 S and 42 44 S. In austral autumn between 39 and 44 S gsst lower than 0.02 C/km are observed 40 km onshore from the mean location of the SBF. South of 40 S from spring to autumn this feature is associated with a persistent front parallel to the SBF. The onshore front is close to the mean position of surface chl-a blooms at the shelf break during spring and summer [see Romero et al., 2006]. Daily SST images reveal the presence of three main branches of relatively cold waters. Two SST minima located in the vicinity and offshore from the 200 m isobath appear to be associated with previously described branches of the MC. South of 40 S and onshore from the 200 m isobath we 5of6

observed an additional branch of cold water. The later branch is also observed north of 40 S during austral autumn. [12] The SBF plays a strong role on the life cycle of a variety of species [e.g., Acha et al., 2004]. For instance, recent studies suggest that zonal displacements of the SBF may determine the cross-shelf extension of Patagonian scallop beds (C. Mauna, personal communication, 2007). Given the cross-shelf beds extension (40 km) is close to the distance between the mean locations of the inshore front (A in Figure 3) and the main SBF (Figure 1), the branch of cold waters located onshore of the 200 m isobath might also play a significant role on the ecology of marine species. [13] Acknowledgments. This work was supported by the Inter-American Institute for Global Change Research grant CRN2076, which is supported by the US National Science Foundation (Grant GEO- 0452325). Additional support was provided by Agencia Nacional de Promoción Científica y Tecnológica grant PICT04 25533 and Glaciar Pesquera, Argentina. SST data were obtained from NASA Physical Oceanography Distributed Active Center at the Jet Propulsion Laboratory. References Acha, E. M., H. W. Mianzan, R. A. Guerrero, M. Favero, and J. Bava (2004), Marine fronts at the continental shelves of austral South America: Physical and ecological processes, J. Mar. Syst., 44, 83 105. Bogazzi, E., A. Baldoni, A. Rivas, P. Martos, R. Reta, J. M. Orensanz, M. Lasta, P. Dell Arciprete, and F. Werner (2005), Spatial correspondence between areas of concentration of Patagonian scallop (Zygochlamys patagonica) and frontal systems in the southwestern Atlantic, Fish. Oceanogr., 14, 359 376. Brunetti, N. E., B. Elena, G. R. Rossi, M. L. Ivanovic, A. Aubone, R. Guerrero, and H. Benavides (1998), Summer distribution, abundance and population structure of Illex argentinus on the Argentine shelf in relation to environmental features, S. Afr. Mar. Sci., 20, 175 186. Campagna, C., A. R. Piola, M. R. Marin, M. Lewis, U. Zajaczkovski, and T. Fernández (2007), Deep divers in shallow seas: Southern elephant seals on the Patagonian shelf, Deep Sea Res., Part I, 54, 1792 1814, doi:10.1016/j.dsr.2007.06.006. Carreto, J. I., V. A. Lutz, M. O. Carignan, A. D. Cucchi Colleoni, and S. G. De Marco (1995), Hydrography and chlorophyll-a in a transect from the coast to the shelf break in the Argentinean Sea, Cont. Shelf Res., 15, 315 336. Chelton, D. B., M. G. Schlax, D. L. Witter, and J. G. Richman (1990), Geosat altimeter observations of the surface circulation of the Southern Ocean, J. Geophys. Res., 95, 17,877 17,903. Ciocco, N. F., M. Lasta, M. Narvarte, C. Bremec, E. Bogazzi, J. Valero, and J. M. Orensanz (2006), Fisheries and aquaculture: Argentina, in Scallops: Biology, Ecology and Aquaculture, 2nd ed., edited by S. E. Shumway and G. J. Parsons, pp. 1251 1283, Elsevier Sci., Amsterdam. Gordon, A. L. (1981), South Atlantic thermocline ventilation, Deep Sea Res., Part A, 28, 1239 1264. Hickox, R., I. Belkin, P. Cornillon, and Z. Shan (2000), Climatology and seasonal variability of ocean fronts in the East China, Yellow and Bohai seas from satellite SST data, Geophys. Res. Lett., 27, 2945 2948. Hubold, G. (1980a), Hydrography and plankton off southern Brazil and Rio de la Plata, August November 1977, Atlantica, 4, 1 22. Hubold, G. (1980b), Second report on hydrography and plankton off southern Brazil and Rio de la Plata; Autumn cruise: April June 1978, Atlantica, 4, 23 42. Kearns, E. J., J. A. Hanafin, R. H. Evans, P. J. Minnett, and O. B. Brown (2000), An independent assessment of Pathfinder AVHRR sea surface temperature accuracy using the Marine Atmosphere Emitted Radiance Interferometer (MAERI), Bull. Am. Meteorol. Soc., 81, 1525 1536. Longhurst, A. (1998), Ecological Geography of the Sea, 398 pp., Elsevier, New York. Lutz, V. A., and J. I. Carreto (1991), A new spectrofluorometric method for the determination of chlorophylls and degradation products and its application in two frontal areas of the Argentine Sea, Cont. Shelf Res., 11, 433 451. Martos, P., and M. C. Piccolo (1988), Hydrography of the Argentine continental shelf between 38 and 42 S, Cont. Shelf Res., 8, 1043 1056. Orsi, A. H., T. Whitworth III, and W. D. Nowlin Jr. (1995), On the meridional extent and fronts of the Antarctic Circumpolar Current, Deep Sea Res., Part I, 42, 641 673. Parker, G., C. M. Paterlini, and R. A. Violante (1997), El Fondo Marino, in El Mar Argentino y sus Recursos Pesqueros, vol. 1, edited by E. Boschi, pp. 65 87, Inst. Nac. de Invest. y Desarrollo Pesquero, Mar del Plata, Argentina. Peterson, R. G., and T. Whitworth III (1989), The Subantarctic and Polar fronts in relation to deep water masses through the southwestern Atlantic, J. Geophys. Res., 94, 10,817 10,838. Piola, A. R., and A. L. Gordon (1989), Intermediate waters in the southwest South Atlantic, Deep Sea Res., Part A, 36, 1 16. Podestá, G. P. (1988), Migratory pattern of Argentine hake (Merluccius hubbsi) and oceanic processes in the southwestern Atlantic Ocean, Fish. Bull., 88, 167 177. Rivas, A. L., and A. R. Piola (2002), Vertical stratification on the shelf off northern Patagonia, Cont. Shelf Res., 22, 1549 1558. Romero, S. I., A. R. Piola, M. Charo, and C. A. E. Garcia (2006), Chlorophyll a variability off Patagonia based on SeaWiFS data, J. Geophys. Res., 111, C05021, doi:10.1029/2005jc003244. Saraceno, M., C. Provost, A. R. Piola, A. Gagliardini, and J. Bava (2004), Brazil Malvinas Frontal System as seen from 9 years of advanced very high resolution radiometer data, J. Geophys. Res., 109, C05027, doi:10.1029/2003jc002127. Saraceno, M., C. Provost, and A. R. Piola (2005), On the relationship between satellite retrieved surface temperature fronts and chlorophyll a in the western South Atlantic, J. Geophys. Res., 110, C11016, doi:10.1029/2004jc002736. Thomson, G. A., A. A. Alder, and D. Boltovskoy (2001), Tintinnids (Ciliophora) and other net microzooplankton (>30 mm) in southwestern Atlantic shelf break waters, Mar. Ecol., 22, 343 355. Ullman, D. S., and P. C. Cornillon (1999), Surface temperature fronts of the east coast of North America from AVHRR imagery, J. Geophys. Res., 104, 23,459 23,478. Vazquez, J., K. Perry, and K. Kilpatrick (1998), NOAA/NASA AVHRR oceans Pathfinder sea surface temperature data set user s reference manual, version 4.0, 10 April 1998, JPL Techn. Rep. D14070, Jet Propul. Lab., Pasadena, Calif. (Available at http://podaac.jpl.nasa.gov/sst/.) Vivier, F., and C. Provost (1999a), Direct velocity measurements in the Malvinas Current, J. Geophys. Res., 104, 21,083 21,103. Vivier, F., and C. Provost (1999b), Volume transport of the Malvinas Current: Can the flow be monitored by TOPEX/POSEIDON?, J. Geophys. Res., 104, 21,105 21,122. A. Baldoni, Instituto Nacional de Investigación y Desarrollo Pesquero (INIDEP), Paseo V. Ocampo N 1, Mar del Plata, B7602HSA, Argentina. B. C. Franco and A. R. Piola, Departamento Oceanografía, Servicio de Hidrografía Naval (SHN), Av. Montes de Oca 2124, Buenos Aires, C1270ABV, Argentina. (ocebcf@furg.br) J. P. Pisoni, Centro Nacional Patagónico (CENPAT-CONICET), Boulevard Brown s/n, Puerto Madryn, 9120, Argentina. A. L. Rivas, Centro Nacional Patagónico (CENPAT-CONICET), Boulevard Brown s/n, Puerto Madryn, 9120, Argentina. 6of6