You are on page 1of 22

REMOTE SENSING OF THE OCEAN

(extracted from Santos, A. M. P., P.B. Machado and P. Relvas (2009). Applications of satellite remote sensing and GIS to oceanography and fisheries: Case studies in the Western Iberia. Chapter 4 in Geoinformatics for Natural Resource Management, P.K. Joshi, P. Pani, S.N. Mohapartra & T.P. Singh, eds, Nova Science Publisher, New York, USA.)

STATE-OF-THE-ART
Satellite Remote Sensing
Although satellite oceanography history is already almost four decades long, a wide variety of improved and innovative sensors have been launched in the last decade on board of satellites. The recognition of the potential of applying satellite RS to oceanography started in the 1960s with Earth photographs taken by the first astronauts. These photographs revealed an amazing amount of details in ocean structures. Figure 1 is an example of the kind of images that can be obtained by astronauts. Although these pictures had created enormous expectations among the scientific community, it is only in the end of the 1970s-beginning of the 1980s with the installation of the AVHRR (Advanced Very High Resolution Radiometer) (hereafter see Appendix A for acronyms list) on board the second generation of NOAA satellites that remotely-sensed satellite data started to play a major role in oceanography, namely in the determination of sea surface temperature. Since then, AVHRR has been carried on the meteorological satellites of the NOAA series until now, providing regular and continuous operational global observations of SST (Sea Surface Temperature) four times per day using two satellites for almost 30 years. This allow to produce consistent and reliable satellitederived SST time series, namely for assess climate change (Emery et al., 1995; Kilpatrick et al., 2001). AVHRR is planned to continue to flow on future NOAA-EUMETSAT (European Organisation for the Exploitation of Meteorological Satellites) missions, namely the MetOp satellites series planned to be operational at least until 2015. The ATSR (Along Track Scanning Radiometer) and AATSR (Advanced Along Track Scanning Radiometer) are more recent instruments carried by the ESA's satellites (ERS-x and ENVISAT), which the main purpose is to make more accurate measurements of global SST (Stricker et al., 1995; Minnett, 1995a; b; Smith et al., 2001; Merchant et al., 2007).

Figure 1. - Space Shuttle photograph of the Oyashio Current in the Bering Sea on March 1992. In the image it is clear the detail and complexity of the features that are present the ocean (Image courtesy of the Image Science and Analysis Laboratory, NASA Johnson Space Center "The Gateway to Astronaut Photography of Earth." http://earth.jsc.nasa.gov/sseop/efs/printinfo.pl?PHOTO=STS045-79-N)

The first satellite specifically designed for ocean applications was SeaSat-A launched in 1978 (Borne et al., 1979). Although it last only about 3 months due to spacecraft malfunction, its data clearly demonstrated the capabilities of microwave sensors for ocean research (e.g., circulation; surface winds, wave heights, tides, storm surges). After the demonstration of the potential of this kind of sensors for the monitoring of the global ocean, subsequent satellites missions carrying

similar instruments were launched, such as GEOSAT that flown from 1985 to 1990 and subsequent follow-on missions, ESA's ERS series (1991-present) and ENVISAT (2002-present) satellites, the French-American TOPEX/Poseidon (1992-present) and Jason (2001-present), the Japanese ADEOS (Advanced Earth Observing Satellites, called Midori in Japanese) series (1996-1997 and 20022003), NASA's QuickSCAT (1999-present), and the Canadian RADARSAT (1995-present) (Pettersson et al., 1995; Heimbach and Hasselmann, 2000; Fu, 2001). One of the great advantages of the microwave sensors in relation to the visible and infrared ones is the capacity of provide an allweather coverage of the ocean because they penetrates through clouds, rain and snow. Figure 1 - Space Shuttle photograph of the Oyashio Current in the Bering Sea on March 1992. In the image it is clear the detail and complexity of the features that are present the ocean (Image courtesy of the Image Science and Analysis Laboratory, NASA Johnson Space Center "The Gateway to Astronaut Photography of Earth." http://earth.jsc.nasa.gov/sseop/efs/printinfo.pl?PHOTO=STS045-79-N) Another important milestone in ocean remote sensing was the launched in 1978 of CZCS (Coastal Zone Color Scanner) flown on NIMBUS-7 satellite that allows an 8 years data set widely used in ocean colour (biology and optics) research. Despite of this immense potential for biogeochemical, as well as physical oceanography studies, the ocean colour community had to wait for about 10 years for a new ocean colour mission, namely the Japanese ADEOS OCTS (Ocean Color and Temperature Scanner on board ADEOS satellites) and POLDER (POlarization and Directionality of the Earth's Reflectances - French sensor on board ADEOS satellites) (1996-1997), the German MOS on board the Indian satellite IRS-P3 (1996-2004) and the US SeaWiFS flown on Orbview-2 satellite (1997-present). Presently there are 10 ocean colour sensors in operation and the continuity colour measurements from space are guaranteed at least for the next decade with 6-7 more missions planned (Dickey et al., 2006; IOCCG, 2007). A good source of information about satellite ocean colour is the web site of the International Ocean-Colour Coordinating Group (IOCCG) at http://www.ioccg.org/index.html. Salinity is an ocean state variable that partially controls sea water density and has important implications for example in climate, ocean circulation and ecology. However, their global monitoring remains poor because of the scarcity of observations in large portions of the ocean. The objective of the future ESA's SMOS (Soil Moisture and Ocean Salinity - schedule for launch in 2008) and NASA/CONAE's [Comisin Nacional de Actividades Espaciales (Space Agency of Argentina)] Aquarius (2009) missions, is to measure surface salinity systematically at a global scale. Some more information about ocean remote sensing of salinity could be found in Lagerloef (2000; 2001), Le Vine et al. (2000) and Lagerloef and Schmitt (2006). Sensor calibration and data processing, a central problem in the remote sensing initial stages, has known a tremendous evolution. Consequently, the range of parameters observed from space, or derived from space observations, with positive results has broadened. The oceanographic scales resolved by remote sensing have decreased, and research on satellite observations of sub-mesoscale features is underway. Application of ocean RS that illustrated this evolution can be found in several reviews and books published during the last years, namely Ikeda and Dobson (1995), Nihoul et al. (1998), Halpern (2000), Cheney (2001), Fu (2001), Liu and Katsaros (2001), Liu and Wu (2001), McClain (2001), Minnett (2001), Parkinson (2001) and Plant (2001). To our knowledge the last review about the use of satellite and airborne remote sensing methods in fisheries was published by Santos (2000), in which several references to previous reviews about the subject could be found. Nowadays, many countries use satellite remote sensing technology for operational fisheries forecasting services, namely Japan and the US (Santos, 2000). Platt et al. (2007) discuss the role of biological oceanography in fisheries management and the importance of satellite RS as a tool complementary of ship observation for studying ocean processes of relevance at appropriate time and space scales. Future missions planned until 2015 that include instruments for ocean monitoring are presented in Figure 2. ESA also plans to launch the Sentinel satellite series in 2011-2012 in the frame of the Global Monitoring for Environment and Security (GMES) programme that will monitor the marine

environment (e.g., ocean circulation, sea-level, chlorophyll concentration, sediments, oil spills), sea ice and icebergs among other things (Atterna et al., 2007; Aguirre et al., 2007).

Figure 2 - Present and future satellite oceanography missions (extracted from the Web site of the NOAA-NESDIS-Center for Satellite Applications and Research (STAR) at http://www.orbit.nesdis.noaa.gov/star/socdr_driverstrends_programs.php).

Principles, Technologies and Methodologies

Satellite Remote Sensing


Remote sensing is the acquisition of information about an object or phenomenon without physical contact with it. In general, the term remote sensing is used to describe the measurement and analysis of information that is obtained with a sensor installed on board satellites and airplanes, although we can consider also as remote sensing the observations of vessel-mounted sonars and ecosounders. Thus, ocean remote sensing allows the measurement of oceanographic parameters and the knowledge of their spatial and temporal variability without physical contact with the sea. It is based on the measurement, processing and analysis of the electromagnetic radiation emitted by the ocean, or reflected by the sea surface from the incident solar radiation (passive remote sensing) or from emitting sources on board of the satellite (active remote sensing). Typically all the row of satellite remote sensing of the sea includes the sensor calibration, atmospheric correction, geo-location, sampling the sea truth conditions, image processing and applications of the satellite remote sensing. Next we will detail each step. All sensors employed on

ocean observing satellites use electromagnetic radiation to view the sea. Sensors are calibrated before launching, but no periodic recalibration can be executed. Some sensors use onboard reference targets, with constant optical characteristics, to detect instrumental drifts that can be corrected during data processing. Only certain wavelengths of the electromagnetic radiation are fully or partly transmitted through the atmosphere and the transmittance varies with the composition and state of the atmosphere. Several strategies can be applied to deal with the atmospheric effect, ranging from detailed modelling of the atmospheric conditions during data acquisition to simple calculations based solely on the image data. However, for satellite imagery in visible and infrared spectral bands, the cloud cover detection is the main difficulty. Analysis of the pixel reflectance by comparison with a preset threshold or with the surrounding pixels is used to map the cloud contamination. The geo-location consists in the attribution of a geographical location to each pixel. Often control points in the ground are used, but this is difficult far from coastal regions. In recent years the positional identification process has taken advantage of the GPS satellites signals. More sophisticated methods, based on Doppler effect for orbit determination, are used in radar-altimeter sensors due to the required accuracy. The choice of the strategy for the oceanographic sampling of the sea true is very important, and is closely related to the spatial resolution of the remote sensing data when compared with the spatial variability of the measured parameter. The samples shall span as large spectrum of data values as possible, keeping in mind that the value measured in a point may not be representative of the average parameter within the whole pixel sampled by the satellite. Oceanographic remote sensing data are presented in several levels of processing that tend to be standardized. The level-0 contains the raw data as transmitted by radio-signal and received by the ground station. Level-1 adds calibration data, navigation data, and instrument and spacecraft details to the previous level. Level-2 product presents already oceanographic parameters derived from the application of algorithms (e.g. surface temperature, surface chlorophyll concentration, etc.), atmospherically corrected and geo-located in sensor coordinates. Level-3 includes oceanographic parameters sampled during a certain period and interpolated on a geographical grid. Level-4 means images representing ocean variables averaged within each grid cell as a result of data analysis and/or modelling. These products can contain data merged from different sources, including in situ measurements, and the gaps where no observations are available filled by optimal interpolation schemes. Finally the multiple applications of the satellite remote sensing depend on the sensor and/or sensor configuration. For oceanographic purposes the important remote sensing devices are the infrared radiometers, visible wavelength sensors, passive microwave radiometers, active microwave sensing and satellite altimetry. Next we will briefly describe each device. Visible wavelength or ocean colour sensors operate in the visible part of the electromagnetic spectrum, measuring solar radiation returned by the ocean in this band. The sunlight that penetrates the ocean surface is selectively absorbed, reflected and scattered by the suspended material in the upper layers. Thus, the ocean colour reflects the suspended matter in the top layer of the ocean, with depths of 50 meter or more in the case of colour radiance in the blue-green, in contrast with the infrared observations that sample only a surface film of the ocean. From the satellite remote sensing of the ocean colour it is possible to estimate, with variable accuracy, a set of biological and optical parameters and processes, including chlorophyll-a (phytoplankton) concentration, water-leaving radiation, diffuse attenuation coefficient, algal blooms and particulate matter as some of the most common. The most significant ocean colour sensors are the CZCS (Coastal Zone Color Scanner) that operated between 1978 and 1986, SeaWiFS (Sea-viewing Wide Field-of-view Sensor), since 1997, and MODIS (Moderate Resolution Imaging Spectroradiometer) on board of Terra satellite, launched in late 1999, and on board of Aqua satellite launched in mid 2002. Infrared scanning radiometers infer the sea surface temperature (SST) from near-infrared and infrared sensors measuring the electromagnetic radiation within the band 1-30 m, emitted by the ocean surface. Processing is based on the fact that all surfaces emit radiation, the strength of which

depends on the surface temperature. Stefan-Boltzman law states the total emitted energy is proportional to the fourth power of the temperature. Sea surface temperature is difficult to interpret because the upper ocean (approx. top 10 meters) has a complex and variable vertical temperature structure that is related to ocean turbulence and air-sea fluxes of heat, moisture and momentum. The thickness of the layer whose temperature is remotely sensed varies approximately between 10-20 m depths. It is called skin SST and it measurement is subject to a large potential diurnal cycle including cool skin layer effects (especially at night under clear skies and low wind speed conditions) and warm layer effects in the daytime. In contrast, the measured in situ SST (called also bulk SST) corresponds to few centimetres or more, depending on surface waves. The SST measurements on buoys and ships may be anything between 0.5 and 3 m deep. The presence of surface films, like transient oil slicks, may also affect the difference between skin and bulk SSTs. The most important SST sensors are the Advanced Very High Resolution Radiometer (AVHRR) on NOAA near polar earth orbiting satellites since 1978, the ATSR (Along Track Scanning Radiometer) on board ESA satellites, MODIS and some others. Passive microwave radiometers operate at electromagnetic wavelengths 1.5300 mm (frequency 1200 GHz). They are not sensitive to scattering by the atmosphere or aerosols, haze, dust, or small water particles in clouds, due to the relatively long wavelength. So, the microwave sensors can survey in all weather conditions. This principle advantage is countered by the fact that thermal emission is very weak at these longer wavelengths. To overcome noise levels a large field of view must be received, that results in low spatial resolution (25150 km). These observations are used for studies of heat balance of the ocean. The emissivity of the sea at microwave frequencies varies with the dielectric properties of sea water (including salinity) and the surface roughness. Hence, the development of this technique in future may enable the measurements of surface salinity. Active microwave radiometers, with oblique viewing, measure the sea surface roughness based on the Bragg scattering. Oblique viewing of a smooth surface with active radar give virtually no return, but if the surface is rough significant backscatter occurs, which is perceived by the microwave scatterometer. The Synthetic Aperture Radar (SAR) is an active microwave device, based on the comprehensive analysis of contribution from individual points to the signal received when the sensor is at a particular point. The result has very high resolution. SAR images enable the analysis of small-scale and mesoscale eddies, internal waves, river plumes, oil slicks, ice packs, etc. Sea surface roughness is also the elemental parameter for the estimation of the wind over the ocean. The SeaWinds instrument on the QuikSCAT (Quick Scatterometer) satellite is a specialized microwave radar lofted in mid 1999 to measure the near surface wind speed and direction under all weather and cloud conditions over the global ocean. The derived operational product presently gives vector winds with 25 25 kilometers resolution. Remote sensing of the sea surface topography is carried out by satellite altimeters radars, which permanently transmit signals at high frequency toward the earth beneath them. The return time of the signal after reflection at the earth's surface is measured, and this yields the height of the satellite. From all sensors carried on satellites, the altimeter is the most dependent upon its orbit to achieve successful calibration and interpretation. Sea surface is rough rather than flat and each individual return signal is very noisy. The radar altimeter is able to remove the effect of the ocean waves by averaging many successive pulses. The most important are TOPEX/Poseidon, ERS and Jason satellites.

CASE STUDIES IN THE IBERIAN PENINSULA


Visible Remote Sensing

Mesoscale Features (Eddies, Filaments, Etc.)


One of the most frequent applications of the ocean colour remote sensing is the analysis of phytoplankton concentrations. Since remote sensing covers large areas on a regular mode, it has proven to be a precious aid in the description of the spatial structure and temporal variability of the near surface phytoplankton concentrations in the Iberian region on a seasonal to inter-annual basis. For instance, summer coastal upwelling is clearly identified by the large pigment values over the continental shelf and upper slope. Less obvious is the continuous band of very high concentrations along most of the continental shelf seen in the ocean colour imagery during the winter, which almost vanishes during the transitions seasons (spring and autumn). Such enhanced primary production in winter may be attributed to the nutrient input from the river runoff (Peliz and Fiza, 1999). The poleward current, a characteristic mesoscale feature that flows along the shelf break and upper slope off western Iberia from autumn till spring, is captured in the ocean colour imagery as a large intrusion of low pigment concentration progressing northward, limiting the pigment rich coastal waters over the shelf and the moderate pigment concentrations offshore. The frontal region between the rich costal waters and the oligotrophic offshore waters is populated during the summer by mesoscale features, such as filaments of upwelled water stretching seaward and detached eddies. The ocean colour images reveal the biological richness of these features, which increase the dispersion of the pigments into the relatively poor open ocean. The location of the filaments recurrently coincides with protrusions in the coastal morphology, such as prominent capes, and topographic features, such as submarine ridges. These well developed phytoplankton structures are generally related to moderate or intense offshore transport, whereas the absence of filaments correspond to either weak offshore transport or coastal convergence (Sousa and Bricaud, 1992). Thus, filaments and jet-like features are found to be the main transport mechanism for the shelf-ocean exchange of phytoplankton and associated carbon fluxes. More recently attention has been drawn to the relatively high chlorophyll concentrations derived from ocean colour (SeaWiFS) images over the shelf and slope, stretching to large distances from the coast and associated with winter upwelling events. Such surface concentrations, high for wintertime (up to 3.5 mg m-3 against typical values of 4-5 mg m-3 in summer), are attributed to the retention of phytoplankton in a shallow buoyant layer of waters of inland origin, that spreads offshore under the influence of upwelling favourable winds (Ribeiro et al., 2005). An extreme winter event was identified through SeaWiFS derived chlorophyll concentration off southwest Iberia in February 2001 (Peliz et al., 2004). An extraordinary long filament transported coastal rich waters as far as 400 km offshore (Figure 5). Costal fresh water river plumes, following a particularly intense period of rainfalls, provided buoyancy and nutrients for an extensive phytoplankton production that was dragged till far from the coast by the offshore eddy dynamics.

38 N

3 S Chl-a (mg/m )

37 N

SV

36 N

13 W

11 W

9 W

7 W

Figure 5 - SeaWiFS-derived chlorophyll-a distribution in SW Iberia in 12 February 2001. The southwestward filament is about 400 km long. S is Cape Sines and SV is Cape So Vincente (adapted from Santos et al., 2007).

Seawifs, The Western Iberian Buoyant Plumeand the Survival Of Sardine Larvae
In the Western Iberia, the transport/dispersal of sardine larvae is affected by factors such as the wind driven transport, the structure and circulation of the Western Iberia Buoyant Plume-WIBP (Western Iberia Buoyant Plume), and the slope circulation associated with the Iberian Poleward Current (IPC) (Santos et al., 2004). The WIBP is a lens of water of 'low' salinity (< 35.8) fed by the winter discharges of several rivers of the NW coast of the Iberian Peninsula where phytoplankton productivity is enhanced by the availability of nutrients and stratification conditions (Ribeiro et al., 2005). For these reasons, the WIBP could be studied using satellite-derived chlorophyll-a distributions (Ribeiro et al., 2005). The spatial patterns captured by in situ measurements during an oceanographic survey off Western Iberia and by SeaWiFS were very similar (Figure 6). Furthermore, the locations of high chlorophyll-concentrations related with sub-surface maxima measured in situ with a fluorometer where also clearly seen in the SeaWiFS-derived distributions (Figure 6). One of these maxima are related with a convergence zone located in the shelf break and it is interested to note that such frontal zone is not visible in the thermal AVHRR image but present in SeaWiFS data for the same day. Thus, the use of ocean colour RS can be an important tool in the study of frontal structures of this nature and also for studying processes related to the variability of recruitment of sardine (Sardina pilchardus) because the WIBP is a suitable environment for larval retention and survival (Santos et al., 2004; 2006a). The study of the evolution in time of the WIBP using sequential SeaWiFS images revealed that eastward frontal velocities could be of the order of 0.34 m s-1 (Ribeiro et al., 2005) and this is another interesting application of ocean colour RS that could be used to estimated the role of the WIBP in the drift of larval sardine during winter upwelling events off Western Iberia (Santos et al., 2004).

Figure 6 - SeaWiFS-derived chlorophyll-a distribution of NW Iberia in 19 February 2000 (middle image). In the right plot is presented the in situ surface distribution of chlorophyll-a and in the left plot a cross-slope section obtained during an oceanographic survey cruise contemporaneous to the satellite image. The white crosses in the satellite image indicate the location of the oceanographic stations and the red arrow the crossslope section presented in the right plot. The shaded red areas in the left and right plots represent the locations of the maxima and features seen in the satellite image (adapted from Ribeiro et al., 2005).

Infrared Remote Sensing

Coastal Circulation
Infrared imagery is the most popular remote sensing source used by coastal oceanographers. The conspicuous contrast of the sea surface temperature in most of the coastal regions, induced by the strong mesoscale activity, makes infrared sensors an ideal tool for monitoring and track mesoscale features there. In fact, the present recognition that the ocean circulation is dominated by meanders and eddies is itself a consequence of the satellite oceanography advent. Off Western Iberia the first applications of infrared satellite imagery to oceanography came from the late seventies (Santos et al., 2006b). There, the coastal circulation is dominated by mesoscale features, some of them associated to the upwelling regime that dominates the oceanography of the region during a substantial part of the year. The upwelling season is roughly defined from March to September (Wooster et al., 1976). The region is poorly sampled by in situ devices, in particular permanently moored instruments for long term observations and monitoring. Therefore, most of the present knowledge of the regional oceanography was built upon model simulations along with satellite imagery analysis, infrared imagery ahead. In situ observations have been limited to short periods (1 to 3 weeks) on board of research vessels, the majority dedicated to specific oceanographic features limited in space, rather than to the coastal circulation at the basin scale. Infrared satellite imagery, when transmitted on board on almost real time, has proofed to be a powerful instrument for the guidance and optimization of many research cruises in the region. Some main results obtained via infrared remote sensing are summarized next.

Box 1 What is upwelling? The Ekman mechanism

According to the Ekman theory the surface current lies 45 to the right (left) from the wind direction in the Northern (Southern) Hemisphere caused by the rotation of the Earth (Coriolis effect). However, that angle of deviation increase with depth but their amplitude decay exponentially until a depth where the frictional influence of the wind is null (Ekman depth) forming the so-called Ekman spiral in the layer directly influence by the frictional

action of the wind (Ekman layer). The net movement of surface water in the integrated Ekman layer (Ekman transport) is about 90 to the right (left) of the wind direction in the Northern (Southern) Hemisphere. Thus, coastal upwelling occurs where Ekman transport moves surface waters away from the coast and is created by alongshore wind blowing across the ocean surface and pushing surface waters offshore (Ekman transport) perpendicular to the wind direction due to the Coriolis force. The displaced surface waters are replaced by upwelled cold and nutrient-rich subsurface waters that allowed the development of phytoplankton (primary producers) at the surface, which are the base of the food web. Thus, these ecosystems are some of the most productive of the world and maintained important populations of fish, marine mammals and seabirds. Examples of such systems are the four major Eastern Boundary Current Systems of the World Oceans: the Canary, California, Benguela and Humboldt Current systems. They account for about 50% of the total worldwide catches of marine species, while representing less than 3% of the ocean surface.
The thermal front that separates the coastal cold nutrient rich upwelled waters from the more oligotrophic offshore waters is much more complicated than a simple contorted border parallel to the coast. During the upwelling season the frontal region is populated with meanders and filaments. These are narrow contorted tongues of cooler upwelled water extending hundreds of kilometers seaward from the coastal zone. Filaments export a much larger mass along their principal axis than expected by the purely wind-driven Ekman circulation, being an important mechanism of exchange between coastal and open ocean waters, with obvious implications in the ecosystem functioning. The existence of such structures off western Iberia was first identified trough infrared remote sensing, and significant advances in the understanding of the filaments behaviour were done, and still are, through remote sensing analysis. The number of filaments along the coast and the recurrent location of their roots in the major topographic features like prominent capes, their maximum extension and width, the development and decaying time scales, the frequency of occurrence and seasonal variability, all is presently known (Haynes et al., 1993). The perception that the offshore eddy field could drag and modulate the filaments offshore is also a result from infrared imagery (Peliz et al., 2004). The dynamics of the filament formation is closely related to the wind regime and possibly with the sea surface topography off the continental shelf, since offshore eddies result in sea level anomalies. The frequency of satellite passes (2 a day for NOAA-AVHRR) and the relatively sparse cloud cover over western Iberia during the upwelling season, make possible to follow the evolution of the coastal filaments and relate it with the forcing variables. Cape So Vicente, the culminating point of southwest Iberia, is the root of a major cold filament. Infrared imagery was fundamental for the definition of it recurrent pattern and a valuable tool in the understanding of the process of it growth and decay (Relvas and Barton, 2002). Recently, a research cruise guided by SST imagery transmitted on board on quasi real time, sampled several transects perpendicular to the filament axes. Thus, it was possible to infer the exchange between the nutrient rich coastal waters and the oligotrophic offshore waters promoted by the upwelling filament. The recent evolution of the remote sensing of the sea level anomalies and wind over the ocean trough on board microwave radars, along with the improvement in their accessibility, makes the study of filaments and other costal features highly remote sensing dependent in regions, like western Iberia, where the in situ sampling is poor.

The upwelling regime has been extensively studied trough infrared imagery all over the world and Iberia is no exception. Remote sensing did allow the detection of narrow warm inshore countercurrents that flow poleward over the inner shelf, attached to the coast, whenever the upwelling favourable winds decay. Such countercurrents, about 10-20 km width, lay on a region traditionally not sampled because it is too close from the coast for oceanographers, but is too distant for coastal experts. They have been observed off the northern (Galicia) and central parts of the western Iberia (Sordo et al., 2001; Peliz et al., 2002), and off the southern Iberian coast (Relvas and Barton, 2002). Inner shelf countercurrents are also visible in other upwelling systems, such the California system (Cudaback et al., 2005). Assuming that the temperature is a tracer of the near surface current, objectives methods make possible the estimation of advective surface velocities from sequential infrared satellite images. Velocities of 0.1-0.3 m s-1 were estimated for the countercurrent off southwest Iberia, increasing when the continental shelf narrows. The occurrence of coastal countercurrents reduces the cross-shelf transport, characteristic of the upwelling regimes. This has ecological consequences by preventing the transport between the inner shelf and the outer shelf, so that alongshore dispersion and consequent retention of larvae stages, phytoplankton and detritus prevail over the inner shelf. Satellite-derived sea surface temperature maps have been used to describe the upwelling patterns off western Iberia, owing to the clear thermal contrast between the cold, vertically mixed, upwelled waters typically found over the shelf, and the thermally stratified oceanic waters. Tacking advantage of this thermal contrast, automatic edge-detection methods have been developed to discern the location of such thermal fronts, which are known to be biologically actives. With the set up of satellite remote sensing observations of the ocean in a regular basis during the past decades, we can presently access to archives with relatively long time series of images. The application of the front detection algorithms to this times series did permit the present knowledge of the recurrent location of upwelling fronts, with obvious benefit for fisheries research (Relvas et al., 2007) A long archive of satellite derived sea surface temperature is also a valuable data set for the generation of regional climatologies, useful for examining the ocean variability and climatic change assessments. The currently available time series of satellite data can adequately depict the known patterns of the climatology and seasonal variability of the western Iberia ecosystem and may be used for investigation of longer time-scale variability. The existing data set is already long enough to capture decadal changes that occurred in the ecosystem since mid 1980s. Analysis of the magnitude of the surface temperature gradient between coastal and offshore waters, an effective estimate of large-scale upwelling intensity, revealed a decadal scale shift of upwelling regime intensity from weak upwelling in 1980s to a stronger one in the 1990s. This regime shift may be associated to the abrupt change of the North Atlantic Oscilation (NAO) observed in the earlier 1990s (Santos et al., 2005).

Coastal Upwelling, Fronts, and Swordfish and Tuna Catch


The widest application of satellite remote sensing to fisheries has been using infrared thermal data to detect frontal structures favourable for the aggregation of tuna (Laurs et al., 1984; Laurs, 1989), mainly because sea surface temperature is the oceanographic parameter that has been most successfully measured by satellite RS, is an indicator of important ocean processes (such as, coastal upwelling, fronts and eddies) and also because ocean colour and other sensors has not been available on an operational basis. Although the main mechanism of aggregation of tuna and tuna-like fish near ocean fronts are still in debate, it seems that behavioural mechanisms related with their feeding activity are the most accepted explanation (Fiedler and Bernard, 1987; Brill and Lutcavage, 2001). The fishing success of swordfish (Xiphias gladius), bigeye (Thunnus obesus) and albacore (Thunnus alalunga) tuna off Western Iberia in relation to frontal structures associated with the dynamics of coastal upwelling was studied by Santos et al. (2006b). Higher catches of swordfish were associated to the strong thermal fronts formed offshore between old upwelling waters and

oceanic warmer waters during the relaxation of coastal upwelling. In Figure 7a; b there is not any thermal front in the fishing zone and therefore the fishing sets performed at that time led to low swordfish catches (0.7-1.8 swordfish/1000 hooks), however it is clearly seen in the northwest (upper left) corner of the satellite image of Fig. 7b that oceanic warmer waters were converging towards the coast corresponding to the beginning of the upwelling relaxation phase. After one week (Fig. 7c) a strong frontal structure developed in the fishing zone and separates old upwelled waters from openocean warmer waters advected shoreward meanwhile. This situation resulted in an increased catch about one order of magnitude greater (7.8 and 16.8 swd/1000 hooks) than one week before.
(a) (b) (c)

Figure 7 Longline fishing sets (white lines) off the Portuguese west coast superimposed on nearly contemporary NOAA-11 satellite-derived sea surface temperature (SST) distributions: (a) 16 August 1992 at 15:35 UTC; (b) 17 August 1992 at 15:23 UTC and (c) 23 August 1992 at 15:51 UTC. Swordfish (swd) catches per unit effort are: (1) 17-19 August 1992, 0.7-1.8 swd/1000 hooks; (2) 22 August 1992, 16.7 swd/1000 h.; and (3) 23 August 1992, 7.8 swd/1000 h. The SST scale is expressed in o C, with values increasing from purple to red; black represents clouds and land (adapted from Santos et al., 2006b).

At the same time, the highest catches of tuna observed during this study were located on the warm (offshore) side of the thermal fronts that constitute the edges of mushroom-like thermal structures (Fig. 8) that (often) constitute the extremity of upwelling filaments and are the surface manifestation of submesoscale anticyclonic/cyclonic eddy pairs. Unfortunately, during this study (1990-1992) there is not any ocean colour sensor operational that allow any relationship with ocean productivity. However, Santos et al. (2006b) did not find any preferred temperature range for these large pelagic fishes supporting the hypothesis of others that the reason why they aggregate near surface fronts are not determined by thermo-physiological mechanisms (Carey and Robison, 1981; Laurs, 1983; Podest et al., 1993; Bigelow et al., 1999) but probably related to their behaviour in relation to feeding.

Figure 8 Longline fishing set (white line) off the west coast of Portugal in 3 November 1992 superimposed on a NOAA-11 satellite-derived sea surface temperature (SST) distribution of November 2, 1992, at 05:18 UTC. Bigeye (bye) and albacore (alb) tuna catch per unit effort were 37.8 bye/1000 h. and 12.2 alb/1000 h., respectively. The SST scale is expressed in oC with values increasing from purple to red; black represents clouds and land (adapted from Santos et al., 2006b).

Microwaves Remote Sensing

Altimetry, Sea-Level Height, Surface Winds and Ocean Circulation


The coastal transition zone off Western Iberia, defined as the region where coastal waters interact with the open ocean, is populated by active eddies that play an important role on the crossshelf transport, with obvious bio-geo-chemical consequences. Recent research suggests that such eddies can drag the coastal waters offshore and modulate the thermal front that separates the cold upwelled waters over the shelf from the offshore waters. For instance, the offshore eddy field may play a fundamental role in the cold upwelling filaments formation and development (Sanchez, 2005). Most of the present knowledge of the interaction process between the offshore eddy field and the coastal waters was built upon the analysis of the ocean topography acquired through microwave radar altimeters on board of satellites. Although the large errors associated with the detection of sea level anomalies trough satellite altimetry over shallow waters, off the continental shelf, in deeper waters, eddies leave a clear anomaly imprint in the sea level field. The analysis of the positive and negative sea level anomalies (anticyclonic and cyclonic eddies, respectively), along with infrared SST and/or visible imagery, contribute to clarify the dynamical processes associated to the formation and development of coastal features, such meanders and filaments. The cooperative analysis of satellite altimetry data and in situ measurements, such as lagrangian drifters or data from cruises or moorings, is a powerful procedure to describe the near surface ocean circulation. Altimetry, along with surface drifters, was successfully used to infer the surface circulation in the eastern North Atlantic (Martins et al., 2002). Is satellite remote sensing limited to the sampling of the ocean surface, or can it tell us some aspects of the subsurface structure? The Mediterranean outflow trough the Strait of Gibraltar, along the southern frontier of the Iberian Peninsula, offers a unique feature to test such hypothesis. The Mediterranean water forms a relatively warm and salty tongue that extends westward from the Iberian Peninsula into the North Atlantic. It equilibrium depth is centered at about 1000 meters, with few hundred meters of thickness, except for a less conspicuous shallow core that flows at 400-600 meters depth along the southern Iberia continental slope. The classical view of the spreading of the Mediterranean water in the Atlantic changed dramatically in the early nineties with the discovery of meddies. Initially, little effort was put into the satellite detection of meddies, because it was

supposed that such structures would leave no signal at the surface. However, some studies start to associate surface features with the presence of the Mediterranean water and meddies (Stammer et al., 1991; Pingree and Le Cann, 1993). Nowadays there is a reasonable confidence that some surface eddy structures visible in sea surface temperature satellite imagery off western and southern Iberia represent the signal of underlying meddy structures. Also, the altimetry data reveal positive sea level anomalies, consistent with the clockwise meddy rotation, of about a tenth of meter directly above a meddy (Oliveira et al., 2000). Nevertheless, it is difficult to separate the surface signals of meddies from the intrinsic upper layer dynamics and the tracking of the meddy pathways through remote sensing methods is still distant. Also the dynamics of the propagation of the meddy signal to the surface is still unclear.

Box 2 Meddies Meddies are clockwise rotating eddies of Mediterranean water, with typical diameters less than 100 km and about 200 meters of thickness, which are able to transport water of Mediterranean origin over thousands of kilometres with very little mixing. They are a key factor for the understanding of the salt and heat transport in the Atlantic.
A differently designed active microwave sensor, with oblique viewing, conceived to sample the sea surface roughness and estimate the wind over the ocean, is transported on board of the QuikScat satellite. The wind stress is the main forcing mechanism in the dynamics of upwelling systems, like western Iberia. Traditionally the coastal meteorological stations are the source of the wind data in the upwelling systems studies. Typically they measure reliable wind velocities every hour or three hours, but in sparse locations and on land, subject to orographic effects. Remote sensed wind data are trustworthy, they cover a large area over the ocean and have daily passes over the eastern North Atlantic. They represent a powerful tool in upwelling systems because they allow the computation of wind stress curl over the ocean and, within a certain extent, the evaluation of the wind field inhomogeneity. Off southwest Iberia remote sensing winds retrieved from the QuikScat microwave scatterometer, with 25 km resolution, were compared with the winds measured in land-based coastal stations. Results did show that coastal winds largely fail on representing the offshore wind field. The exception is the station located in the prominent and exposed Cape So Vicente, the southwest tip of the Iberian Peninsula. The wind measured in other stations, located along relatively straight coastline segments, is representative of the wind pattern over a limited area of the coastal ocean only (Sanchez et al., 2007a). The combined analysis of satellite winds and sea surface temperatures, through canonical correlation analysis, did reveal the influence of the wind pattern on the establishment of characteristic patterns of the sea surface temperature at several time scales. The spatial scales of the analysis vary from a vast region centered in the Iberian Peninsula (0-20W; 30-50N), till the local scale around Cape So Vicente (5.5-10.5W; 35.5-39.0N). Due to the satellite remote sensing nature of the data, that covers the studied area in a continuous way, it was possible to conclude about the relation of the wind pattern, and consequently of the sea surface pattern, with the North Atlantic Oscillation (NAO) (Sanchez et al., 2007b).

SAR, Sea Surface Roughness, Submesoscale Phenomena And Primary Production


Synthetic Aperture Radar (SAR) is an active microwave sensor designed to observe the sea surface roughness. It performs well in all weather conditions. The first applications of SAR images off the Iberian Peninsula come from the late seventies, during the short SEASAT mission. The sea surface roughness pattern shown by a SAR image covering the coastal region from 40 to 41 N was interpreted as the surface manifestation of a large quantity of internal wave trains, propagating towards the shore (Alpers, 1985). Internal waves were first sampled from space over the western Iberia shelf through colour photography taken by the US astronauts during the Apollo-Soyuz mission in 1975 (Apel, 1979).

Box 3 Internal waves

Internal waves are waves that propagate within the body of a stratified fluid. During the summer, the surface layer of the ocean (30-50 m) can be up to 10C warmer than the water bellow. The pycnocline formed provide the ideal interface along which such waves can travel. Large scale internal waves with tidal periods are found on the Iberian shelf, as well as in many others shelf edge regions, and are referred as internal tides. Shorter period internal waves, including packets of internal solitary waves, are also frequently observed at the west coast of the Iberian Peninsula. Direct observational evidence of the propagation of internal waves off western Iberia with ship-mounted thermistor chains reveal the appearance of thermocline depressions at different points of the shelf, with vertical displacements as large as 45 meters (Jeans and Sherwin, 2001). Internal tidal waves may increase primary production in the upper pycnocline by increasing the average light intensity experienced by phytoplankton near the pycnocline. Light intensity decreases exponentially with depth (Beers law). A neutrally buoyant or slowly sinking phytoplankton cell, undergoing vertical displacements by internal tides, is exposed to a greater average light intensity than that at its average depth. Assuming that near the pycnocline the photosynthesis is proportional to the total daily irradiance, because of the linear response to the dim light conditions near the euphotic zone, then the vertical motion of internal waves will introduce a significantly enhancement in primary production (Lande and Yentsch, 1988). Short-period internal waves are also important biological factors because of their impact on the development and transport of plankton. Many species are dependent on environmental cross-shore transport mechanisms to reach adult habitats, since they are not autonomous in their early stages and cannot control their cross-shore position. Non-linear internal waves may produce a net transport of in-water particles (phytoplankton, zooplankton and even small fish), which in the upper layer is usually in the

same direction as the internal wave propagation (when the pycnocline displacements are of depression type, as it is the case off the western Iberian coast). Typical distances reached by such horizontal transport are of the order of several kilometres (Lamb, 1997). This mechanism is particularly effective when the transport by internal waves cooperates with the wind drift and plankton swimming (Shanks, 1995). Internal tidal bores have also been identified as an important mechanism of nutrient supply to the near-shore. Larval accumulation can occur at the leading edge of internal tidal bores, causing aggregation of organisms in slicks (Pineda, 1999). Some of these organisms have the ability to remain at the leading edge of the internal motion by swimming against the dowelling currents or because they are sufficiently buoyant, and when the disturbance propagates all the way to the shore, the larvae would be effectively transported onshore, reach the adult habitat, and have an opportunity to complete their life cycle.
A prompt detection of internal waves through remote sensing is important since they are believed to impact the primary production over the shelf, thus in coastal regions, as it will be discussed in the next paragraph. Satellite SAR high resolution images can also provide a functional manner to predict the location of internal waves. A good agreement between predicted and observed locations of internal wave packets was already achieved in an experiment carried out off the northern part of the west coast of the Iberian Peninsula (da Silva et al., 1997). Internal tidal waves may leave a distinct signature in ocean colour remotely sensed images too (da Silva et al., 2002). Bands of enhanced levels of near-surface chlorophyll in the central region of the Bay of Biscay were associated with the uplifting of a subsurface chlorophyll maximum by passing internal tide waves travelling away from the shelf break. Those bands were 30-50 km width and could be detected by the SeaWiFS sensor, with 1 km resolution. Almost contemporaneous SAR overpasses, with resolution of 25 m, did confirm the assumption. Because of its large coverage, independence of the day-night cycle and all-weather capability, space-borne SAR has been proven to be an useful tool for detection of surface films in the ocean, in particular oil spill monitoring and tracking. Research following the Prestige accident off northwest Iberia did show that radar imagery along with other information, such as wind data or in situ observations, integrated into a Geographical Information System (GIS) database, provide a powerful instrument to study the spatial distribution and the evolution of the oil slicks (Palenzuela et al., 2006).

REFERENCES
Afonso-Dias, M.A., Simes, J.M., Pinto, C., and Sousa, P. (2002). Use of Satellite GPS data to map effort and landings of the Portuguese crustacean fleet. In: Final report of the E.U. study project 99/059 GEOCRUST. pp 106. Afonso-Dias, M., Simes, J., and Pinto, C. (2004). A dedicated GIS to estimate and map fishing effort and landings for the Portuguese crustacean trawl fleet. In: Proceedings of the Second International Symposium on GIS/Spatial Analyses in Fishery and Aquatic Sciences (Eds. Nishida, T., Kailola, P.J. and Hollingworth, C.E. Fishery) GIS research Group, Saitama, Japan, pp. 323-340. Aguirre, M., Berruti, B., Bezy, J.-L., Drinkwater, M., Heliere, F., Klein, U., Mavrocordatos, C., Silvestrin, P., Greco, B., and Benveniste, J. (2007). Sentinel-3. The ocean and mediumresolution land mission for GMES operational services. ESA Bulletin 131: 25-29. Atterna, E., Bargellini, Edwards, P., Levrini, G., Lokas, S., Moeller, L., Rosich-Tell, B., Secchi, P., Torres, R., Davidson, M., and Snoeij, P. (2007). Sentinel-1. The radar mission for GMES operational land and sea services. ESA Bulletin 131: 11-17. Alpers, W. (1985). Theory of radar imaging of internal waves. Nature 314: 245-247. Apel, J.R. (1979). Observations of internal wave surface signatures in ASTP photographs. ApolloSoyuz Test Project, Summary Sci. Rep., NASA SP-412, 505-509. Bez, N. (2002). Global fish abundance estimation from regular sampling: the geostatistical transitive method. Canadian Journal of Fisheries and Aquatic Sciences 59: 1921-1931. Bigelow, K.A., Boggs, C.H., and He, X. (1999). Environmental effects on swordfish and blue shark catch rates in the US North Pacific longline fishery. Fisheries Oceanography 8(3): 178-198. Bordalo Machado, P. (2002). Geostatistical analysis of seafloor depth in the Portuguese continental slope off Figueira da Foz its application in GIS. MSc. Thesis in Geographic Information Systems, Technical Superior Institute, Technical University of Lisbon, pp. 79. (in Portuguese) Bordalo-Machado, P. (2006). Fishing effort analysis and its potential to evaluate stock size. Reviews in Fisheries Science 14: 431-455. Bordalo-Machado, P., and Figueiredo, F. (2007). Extraction and classification of longline fishing trips from vessel monitoring systems data with sequential recording gaps. In: The International Archives of the Photogrammetry, Remote Sensing and Spatial Information Science 34(XXX) (Ed. Stein, A.), ISPRS working group II-7, 5th International Symposium on Spatial Data Quality, Enschede, pp. 114118. Bordalo-Machado, P., Figueiredo, I., Silva, C., and Leotte, F. (2004). Spatial and temporal changes in the distribution and abundance of four deep-water crustacean species in the southwestern and southern slopes of mainland Portugal. In: 4th Assembleia Luso-Espanhola de Geodesia e Geofsica SIG session, Fig. da Foz, Portugal, pp. 2. Bordalo-Machado, P., Sousa, R.T., and Matos, J. L. (in press). A software toolkit to analyse Vessel Monitoring System data from longline fisheries. In: Proceedings of the Third International Symposium on GIS/Spatial Analyses in Fishery and Aquatic Sciences, Shanghai, 2005. Ed. Nishida, T., Kailola, P.J. and Hollingworth, C.E. Fishery GIS research Group, Saitama, Japan. Borne, G.H., Dunne, J., A. and Lame, D. B. (1979). SEASAT mission overview. Science 204: 14051406 Breman, J., Ed. (2002). Marine Geography. GIS for the Oceans and Seas. ESRI press, Redlands, USA. pp. 204. (ISBN-1-58948-045-7). Brill, R.W., and Lutcavage, M.E. (2001). Understanding environmental influences on movements and depth distributions of tunas and billfishes can significantly improve population assessments. American Fisheries Society Symposium 25: 179-198. Burrough, P. A., and McDonnell, R. A. (1998). Principles of Geographic Information Systems, 2nd Ed., Clarendon Press. pp. 346 (ISBN 0-19-854592-4). Carey, F.G., and Robison, B.H. (1981). Daily patterns in the activities of swordfish, Xiphias gladius, observed by acoustic telemetry. Fishery Bulletin US 79(2): 277-292.

Cheney, R.E. (2001). Satellite altimetry. In: Encyclopedia of ocean sciences (Vol. 5). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 2504-2510. Chorley, R. (1987). Handling Geographic Information. Report of the Committee of Enquiry. DoE, HMSO, London. Clarke, K. C. (1995). Analytical and Computer Cartography, 2nd Edition Englewood Cliffs, New Jersey. Prentice-Hall. pp. 288 (ISBN-10: 0130334812). Cooke, J.G. (1985). On the relationship between catch per unit effort and whale abundance. Reports of the International Whaling Commission 35: 511-519. Cowen, R.K., Gawarkiewicz, G., Pineda, J., Thorrold, S.R., and Werner, F.E. (2007). Population connectivity in marine systems: an overview. Oceanography 20(3): 14-21. Crowder, L.B., Osherenko, G., Young, O.R., Airam, S., Norse, E.A., Baron, N., Day, J.C., Douvere, F., Ehler, C.N., Halpern, B.S., Langdon, S.J., McLeod, K.L., Ogden, J.C., Peach, R.E., Rosenberg, A.A., and Wilson, J.A. (2006). Resolving mismatches in U.S. ocean governance. Science 313: 617-618. Cudaback, C.N., Washburn, L., and Dever, E. (2005). Subtidal inner-shelf circulation near Point Conception, California. Journal of Geophysical Research 110, C10007, doi:10.1029/2004JC002608. da Silva, J.C.B., Jeans, D.R.G., Robinson, I.S., and Sherwin, T.J. (1997). The application of near real-time ERS-1 SAR data to the prediction of the location of internal waves at sea. International Journal of Remote Sensing 18(16): 3507-3517. da Silva, J.C.B., New, A.L., Srokosz, M.A., and Smyth, T.J. (2002). On the observability of internal tidal waves in remotely-sensed ocean colour data. Geophysical Research Letters 29(12), 10.1029/2001GL013888. Dickey, T., Lewis, M., and Chang, G. (2006). Optical oceanography: Recent advances and future directions using global remote sensing and in situ observations. Reviews of Geophysics 44, RG1001, doi:10.1029/2003RG000148. Emery, W.J., Wick, G.A., and Schluessel, P. (1995). Skin and bulk sea surface temperatures: Satellite measurement and corrections. In: Oceanographic applications of remote sensing. Ed. Ikeda, M. and Dobson, F.W., CRC Press, Boca Raton, USA. pp. 145-165 Fedorov, K.N., and Sklyarov, V.E. (1981). Space oceanography: Hopes and realities. In: Oceanography from space. Ed. Gower, J.F.R., Plenum Press, New York, USA. pp. 35-47. Fiedler, P. C., and Bernard, H. J. (1987). Tuna aggregation and feeding near fronts observed in satellite imagery. Continental Shelf Research 7(8): 871-881. Fu, L.-L. (2001). Ocean circulation and variability from satellite altimetry (Chapter 3.3). In: Ocean circulation and climate. Observing and modelling the global ocean. Ed. Siedler, G., Church, J. and Gould, J., Academic Press, San Diego, USA. pp. 141-172. Gaile, G. L., and Willmott, C. J. (1989) Geography in America. Columbus: Merrill Publishing Co. pp. 840 Goodchild, M.F. (1992) Geographical information science. International Journal of Geographical Information Systems 6(1): 31-45. Goodchild, M.F. (1997). What is Geographic Information Science?, NCGIA Core Curriculum in GIScience, http://www.ncgia.ucsb.edu/giscc/units/u002/u002.html, posted October 7, 1997 Goovaerts, P. (1997). Geostatistics for Natural Resources Evaluation. Oxford University Press. pp. 483 (ISBN 0-19-511538-4). Halpern, D., Ed. (2000). Satellites, oceanography and society. Elsevier Oceanography Series, 63, Elsevier, Amsterdam, The Netherlands. pp. 367. (ISBN-0-444-50501-6). Haynes, R., Barton, E., and Pilling, I. (1993). Development, persistence, and variability of upwelling filaments off the Atlantic coast of the Iberian Peninsula. Journal of Geophysical Research 98(C12): 2268122692. Heimbach, P., and Hasselmann, K. (2000). Development and application of satellite retrievals of ocean wave spectra. In: Satellites, oceanography and society. Ed. Halpern, D., Elsevier Oceanography Series, 63, Elsevier, Amsterdam, The Netherlands. pp. 5-33.

Hutchings, J.A. (1996). Spatial and temporal variation in the density of northern cod and review of hypotheses for the stock's collapse. Canadian Journal of Fisheries and Aquatic Sciences 53(5): 943-962. Ikeda, M., and Dobson, F.W., Eds. (1995). Oceanographic applications of remote sensing. CRC Press, Boca Raton, USA. pp. 492. (ISBN-0-8493-4525-1). IOCCG (2007). Ocean-Colour Data Merging. Gregg, W. (ed.), Reports of the International OceanColour Coordinating Group, No. 6, IOCCG, Dartmouth, Canada. pp. 68 Jardim, E., and Ribeiro, P.J. Jr. (2007). Geostatistical assessment of sampling designs for Portuguese bottom trawl surveys. Fisheries Research 85: 239247. Jeans, D.R.G., and Sherwin, T.J. (2001). The variability of strongly non-linear solitary internal waves observed during an upwelling season on the Portuguese Shelf. Continental Shelf Research, 21: 1855-1878. Kilpatrick, K.A., Podest, G.P., and Evans, R. (2001). Overview of the NOAA/NASA advanced very high resolution radiometer Pathfinder algorithm for sea surface temperature and associated matchup database. Journal of Geophysical Research 106(C5): 9179-9197. Lagerloef, G. (2000). Recent progress toward satellite measurements of the global sea surface salinity field. In: Satellites, oceanography and society. Ed. Halpern, D., Elsevier Oceanography Series, 63, Elsevier, Amsterdam, The Netherlands. pp. 309-319. Lagerloef, G. (2001). Satellite measurements of salinity. In: Encyclopedia of ocean sciences (Vol. 5). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 2511-2516. Lagerloef, G., and Schnitt, R. (2006). Role of ocean salinity in climate and near-future satellite measurements. Eos Transactions of the American Geophysical Union 87(43): 466. Lamb, K.G. (1997). Particle transport by nonbreaking, solitary internal waves. Journal of Geophysical Research 102(C8): 18641-18660. Lande, R.. and Yentsch, C.S. (1988). Internal waves, primary production and the compensation depth of marine phytoplankton. Journal of Plankton Research 10(3): 565-571. Laurs, R.M. (1983). The North Pacific albacore - an important visitor to California Current waters. CalCOFI Reports 24: 99-106. Laurs, R.M. (1989). Applications de la tldtection satellitaire du germon du Pacifique Nord, Thunnus alalunga (Bonnaterre). In: Tldtection satellitaire et Pcheries Thonires. Ed. Le Gall, J.-Y., FAO Document Technique sur les Pches, No. 302, FAO, Rome, Italy. pp. 87-97. Laurs, R.M., Fiedler, P.C., and Montgomery, D. R. (1984). Albacore tuna catch distribution relative to environmental features observed from satellites. Deep-Sea Research 31(9A): 1085-1099. Le Vine, D.M., Zaitzeff, J.B., D'Sa, E.J., Miller, J.L., Swift, C., and Goodberlet, M. (2000). Sea surface salinity: Towards an operational remote-sensing system. In: Satellites, oceanography and society. Ed. Halpern, D., Elsevier Oceanography Series, 63, Elsevier, Amsterdam, The Netherlands. pp. 321-335. Liu, A.K., and Wu, S.Y. (2001). Satellite remote sensing SAR. In: Encyclopedia of ocean sciences (Vol. 5). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 2563-2573. Liu, W.T., and Katsaros, K.B. (2001). Air-sea fluxes from satellite data (Chapter 3.4). In: Ocean circulation and climate. Observing and modelling the global ocean. Ed. Siedler, G., Church, J. and Gould, J., Academic Press, San Diego, USA. pp. 173-180. Machado, P.B., and Sousa, A.J. (2004). A geoestatstica como ferramenta de estudo da batimetria na vertente continental portuguesa. Nmero temtico de Cartografia e SIG, FINISTERRA, Revista Portuguesa de Geografia. Vol. XXXIX, 76: 95-107(in Portuguese). Manley, T.O., and Tallet, J.A. (1990). Volumetric visualization: An effective use of GIS technology in the field of oceanography. Oceanography, 23-30. Martins, C.S., Hamann, M., and Fiza, A.F.G. (2002). Surface circulation in the eastern North Atlantic, from drifters and altimetry. Journal of Geophysical Research 107(C12), 3217, doi: 10.1029/2000JC000345.

McClain, C. (2001). Ocean color from satellites. In: Encyclopedia of ocean sciences (Vol. 4). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 19461959. Meaden, G.J. (1987). Where Should Trout Farms be in Britain? Fish farmer 10(2). Meaden, G.J. and Do Chi, T. (1996). Geographical Information Systems Application to Marine fisheries. FAO Fisheries Technical Paper, No 356. FAO, Rome. pp. 335. Meaden, G.J., Kapetsky, J.M. (1991). Geographical information systems and remote sensing in inland fisheries and aquaculture. FAO Fisheries Technical Paper, No. 318, FAO, Rome, Italy. pp. 262. Merchant, C.J., Llewellyn-Jones, D., Saunders, R.W., Rayner, N.A., Kent, E.C., Old, C.P., Berry, D., Birks, A.R., Blackmore, T., Corlett, G.K., Embury, O., Jay, V.L., Kennedy, J., Mutlow, C.T., Nightingale, T.J., O'Carroll, A.G., Pritchard, M.J., Remedios, J.J., and Tett, S. (2007). Deriving a sea surface temperature record suitable for climate change research from the alongtrack scanning radiometers. Advances in Space Research doi:10.1016/j.asr.2007.07.041. Minnett, P.J. (1995a). Sea surface temperature from the Along-Track Scanning Radiometer. In: Oceanographic applications of remote sensing. Ed. Ikeda, M. and Dobson, F.W., CRC Press, Boca Raton, USA. pp. 131-143. Minnett, P.J. (1995b). The Along-Track Scanning Radiometer. In: Oceanographic applications of remote sensing. Ed. Ikeda, M. and Dobson, F.W., CRC Press, Boca Raton, USA. pp. 461-472. Minnett, P.J. (2001). Satellite remote sensing of sea surface temperatures. In: Encyclopedia of ocean sciences (Vol. 5). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 2552-2563. Munk, W. (2000). Oceanography before, and after, the advent of satellites. In: Satellites, oceanography and society. Ed. Halpern, D., Elsevier Oceanography Series, 63, Elsevier, Amsterdam, The Netherlands. pp. 1-4. Nihoul, J.C, Strub, P.T. and LaViollete, P.E. (1998). Remote sensing (Chapter 12). In: The Sea. The global coastal ocean. Processes and methods. Ed. Brink, K.H. and Robinson, A.R., John Wiley and Sons. Inc., New York, USA. pp. 329-357. NRC (2001). Marine protected areas: tools for sustaining ocean ecosystem. Committee on the Evaluation, Design, and Monitoring of Marine Reserves and Protected Areas in the United States, Ocean Studies Board, Commission on Geosciences, Environment, and Resources, National Research Council. National Academy Press, Washington, D.C., USA. pp. 272. (ISBN-0-309-07286-7). Oliveira, P.B., Serra, N., Fiza, A.F.G., and Ambar, I. (2000). A study of meddies using simultaneous in situ and satellite observations. In: Satellites, oceanography and society. Ed. Halpern, D., Elsevier Oceanography Series, 63, Elsevier, Amsterdam, The Netherlands. pp. 125148. Palenzuela, J.M.T., Vilas, L.G., and Cuadrado, M.S. (2006). Use of ASAR images to study the evolution of the Prestige oil spill off the Galician coast. International Journal of Remote Sensing 27(9-10): 1931-1950. Parkinson, C.L. (2001). Satellite passive microwave measurements of sea ice. In: Encyclopedia of ocean sciences (Vol. 5). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 2531-2539. Peixoto, J.P., and Oort, H. (1992). Physics of climate. American Institute of Physics, New York, USA. pp. 520. (ISBN-0-88318-712-4). Peliz, A., and Fiza, A.F.G. (1999). Temporal and spatial variability of CZCS-derived phytoplankton pigment concentrations off the western Iberia Peninsula. International Journal of Remote Sensing 20(7): 1363-1403. Peliz, A., Rosa, T.L., Santos, A.M.P., and Pissarra, J.L. (2002). Fronts, jets, and counter-f lows in the Western Iberian upwelling system. Journal of Marine Systems 35: 61-77.

Peliz, A., Santos, A.M.P., Oliveira, P., and Dubert, J. (2004). Extreme cross-shelf transport induced by eddy interactions southwest of Iberia in winter 2001. Geophysical Research Letters 31, L08301, doi:10.1029/2004GL019618. Petitgas, P., Poulard, J.C., and Biseau, A. (2003). Comparing commercial and research survey catch per unit of effort: megrim in the Celtic Sea. ICES Journal of Marine Science 60: 6676. Pettersson, L.H., Samuel, P., and Drottning, . (1995). Radar altimeter sensors and methods. In: Oceanographic applications of remote sensing. Ed. Ikeda, M. and Dobson, F.W., CRC Press, Boca Raton, USA. pp. 427-441. Pineda, J. (1999). Circulation and larvae distribution in internal tidal bore warm fronts. Limnology and Oceanography 44(6): 1400-1414. Pingree, R.D., and Le Cann, B. (1993). A shallow meddy (a smeddy) from the secondary Mediterranean salinity maximum. Journal of Geophysical Research 98(C11): 20169-20185. Plant, W.J. (2001). Satellite remote sensing microwave scatterometers. In: Encyclopedia of ocean sciences (Vol. 5). Ed. Steele, J.H., Turekian, K.K. and Thorpe, S.A., Academic Press, San Diego, USA. pp. 2539-2551. Platt, T., Sathyendranath, S., and Fuentes-Yaco, C. (2007). Biological oceanography and fisheries management: perspective after 10 years. ICES Journal of Marine Science 64(5): 863-869. Podest, G.P., Browder, J.A., and Hoey, J.J. (1993). Exploring the association between swordfish catch rates and thermal fronts on U. S. longline grounds in the western North Atlantic. Continental Shelf Research 13(2-3): 253-277. Raheja, N. (2003). GIS-based Software Applications for Environmental Risk Management. In: Map India Conference 2003. pp. 9. Relvas, P., and Barton, E. D. (2002). Mesoscale patterns in the Cape So Vicente (Iberian Peninsula) upwelling region. Journal of Geophysical Research 107(C10), 3164, doi:10.1029/2000JC000456. Relvas, P., Barton, E.D., Dubert, J., Oliveira, P.B., Peliz, A., da Silva, J.C.B., and Santos, A.M.P. (2007). Physical Oceanography of the Western Iberia, Ecosystem: latest views and challenges. Progress in Oceanography 74(23): 149173. Ribeiro, A.C., Peliz, A., and Santos, A.M.P. (2005). A study of the response of chlorophyll-a biomass to a winter upwelling event off Western Iberia using SeaWiFS and in situ data. Journal of Marine Systems 53: 87-107. Snchez, R. F. (2005). The regional oceanography off Cape So Vicente: from the large-scale to the upwelling filament. Ph.D. thesis. Universidade do Algarve, Faro, Portugal. Snchez, R.F., Relvas P., and Pires, H.O. (2007a). Comparisons of ocean scatterometer and anemometer winds off the southwestern Iberian Peninsula. Continental Shelf Research 27: 155175, doi:10.1016/j.csr.2006.09.007. Snchez, R.F., Relvas, P., and Delgado, M. (2007b). Coupled ocean wind and sea surface temperature patterns off the western Iberian Peninsula. Journal of Marine Systems, 68: 103-127, doi:10.1016/j.jmarsys.2006.11.003. Santos, A.M.P. (2000). Fisheries oceanography using satellite and airborne remote sensing methods: a review. Fisheries Research, 49(1): 1-20. Santos, A.M.P., Chcharo, A., Dos Santos, A., Moita, T., Oliveira, P.B., Peliz; A., and R, P. (2007). Physical-biological interactions in the life history of small pelagic fish in the Western Iberia Upwelling Ecosystem. Progress in Oceanography 74(2-3): 192-209, doi:10.1016/j.pocean.2007.04.008. Santos, A.M.P., Fiza, A.F.G., and Laurs, R.M. (2006b). Influence of SST on catches of swordfish and tuna in the Portuguese domestic longline fishery. International Journal of Remote Sensing 27(15): 3131-3152. Santos, A.M.P., Kazmin, A.S., and Peliz, A. (2005). Decadal changes in the Canary upwelling system as revealed by satellite observations: their impact on productivity. Journal of Marine Research 63: 359-379.

Santos, A.M.P., Peliz, A., Dubert, J., Oliveira, P.B., Angelico, M.M., and R, P. (2004). Impact of a winter upwelling event on the distribution and transport of sardine eggs and larvae off western Iberia: a retention mechanism. Continental Shelf Research 24: 149-165. Santos, A.M.P., R, P., Dos Santos, A., and Peliz, A. (2006a). Vertical distribution of the European sardine (Sardina pilchardus) larvae and its implications for their survival. Journal of Plankton Research 28(3): 1-10. Shanks, A.L. (1995). Oriented swimming by megalopae of several eastern North Pacific crab species and its potential role in their onshore migration. Journal of Experimental Marine Biology and Ecology, 186: 1-16. Smith, D.L., Delderfield, J., Drummond, D., Edwards, T., Mutlow, C.T., Read, P.D., and Toplis, G.M. (2001). Calibration of the AATSR instrument. Advances in Space Research 28(1): 31-39. Sordo, I., Barton, E.D., Cotos, J.M., and Pazos, Y. (2001). An inshore poleward current in the NW of the Iberian Peninsula detected from satellite images, and its relation with G. catenatum and D. acuminata blooms in the Galician Rias. Estuarine, Coastal and Shelf Science 53: 787-799. Sousa, F. M., and Bricaud, A. (1992). Satellite-derived phytoplankton pigment structures in the Portuguese upwelling area. Journal of Geophysical Research, 97(C7): 11343-11356. Stammer, D., Hinrichsen, H.H., and Kase, R.H. (1991). Can meddies be detected by satellite altimetry? Journal of Geophysical Research 96(C4): 7005-7014. Star, J., and Estes, J. (1990). Geographic Information Systems: An Introduction. Prentice-Hall, New York, USA. pp. 665 (ISBN-10: 0133511235). Stricker N.C.M., Hahne, A., Smith, D.L., Delderfield, J., Oliver, M.B., and Edwards, T. (1995). ATSR-2: The evolution in its design from ERS-1 to ERS-2. ESA Bulletin 83: 32-37. UNESCO (1999). Module A, Introduction to GIS. Training Module on the Applications of Geographic Information Systems (GIS) for On-line Governance and Accessing Public Domain Information. http://gea.zvne.fer.hr/module/module_a/module_a1.html, Last update: 15-07-1999. Wooster, W.S., Bakun, A., and McLain, D. R. (1976). The seasonal upwelling cycle along the eastern boundary of the North Atlantic. Journal of Marine Research 34(2): 131-141. Wright, D.J. (2000). Down to the Sea in Ships: The Emergence of Marine GIS. In: Marine and Coastal Geographical Information Systems (Research Monographs in Geographic Information Systems). Ed. Wright, D.J. and Bartlett, D.J. Taylor & Francis, London Wright, D.J., Ed. (2002). Undersea with GIS. ESRI press, Redlands, USA. pp. 253 + CD-ROM. (ISBN-1-58948-016-3) Wright, D.J., and Goodchild, M.F. (1997), Data from the deep: Implications for the GIS community. International Journal of Geographical Information Science 11(6): 523-528.

QUESTION BANK
True or False
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. Spatial processes are not important in marine ecosystems. Earth photographs taken by astronauts could be used to study the ocean. AVHRR is the most used sensor in ocean remote sensing. AVHRR is mostly used to estimate chlorophyll-a concentration. CZCS, ATSR and SeaWiFS are the most important ocean colour sensors. Visible wavelength is the same as ocean colour. Infrared wavelength is use to estimate sea surface temperature. Infrared wavelength penetrates 10-20 m deep in the ocean. Visible and infrared sensor could not penetrate through clouds, rain and snow. Salinity could not be measured by remote sensing. Surface winds could be measured by remote sensing. Level-1 processing contains the raw data.

Short Answer Questions


1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. What is ocean remote sensing? Describe the main steps in ocean remote sensing. Why is geo-location difficult in the open ocean? What are the main limitations of the use of visible and infrared satellite imagery to study the ocean? Why are infrared sensor data suitable for study coastal upwelling? What is the great advantage of the microwave sensors in relation to the visible and infrared ones? Is satellite remote sensing limited to the sampling of the ocean surface, or can it be used to study subsurface structures? Name 3 biological or optical parameters that could be estimate by ocean colour remote sensing. Name some ocean features that could be analysed by SAR images. Explain why the Western Iberia Buoyant Plume (WIBP) could be study using ocean colour remote sensing. Explain why ocean colour remote sensing could be important in explain the distribution of tuna and tuna-like species (e.g., swordfish). How can internal tidal waves increase primary production in the surface layers?

You might also like