Acessibilidade / Reportar erro

Carbon regeneration in the Cariaco Basin, Venezuela

Abstract

The carbon regeneration in the water column of the Cariaco Basin (Venezuela) was investigated using a regression model of total alkalinity (TA) and the concentration of total inorganic carbon (TCO2). Primary productivity (PP) was determined from the inorganic carbon fraction assimilated by phytoplankton and the variation of the 22 and 23ºC isotherm was used as an indicator of coastal upwelling. The results indicate that CO2 levels were lowest (1962 µmol/kg) at the surface and increased to 2451 µmol/kg below the oxic-anoxic redox interface. The vertical regeneration distribution of carbon was dominated (82%) by organic carbon originating from the soft tissue of photosynthetic organisms, whereas 18% originated from the dissolution of biogenic calcite. The regeneration of organic carbon was highest in the surface layer in agreement with the primary productivity values. However, at the oxic-anoxic interface a second more intense maximum was detected (70-80%), generated by chemotrophic respiration of organic material by microorganisms. The percentages in the anoxic layers were lower than in the oxic zone because aerobic decomposition occurs more rapidly than anaerobic respiration of organic material because more labile fractions of organic carbon have already been mineralized in the upper layers.

Descriptors:
Cariaco Basin; Carbon regeneration; Water column; Coastal upwelling

Resumo

A regeneração de carbono na coluna de água da Bacia de Cariaco (Venezuela) foi investigada utilizando um modelo de regressão de alcalinidade total (TA) e a concentração de carbono inorgânico total (TCO2). Produtividade primária (PP) foi determinada a partir da fracção de carbono inorgânico assimilável pelo fitoplâncton e a variação da isotérmica de 22 e 23ºC foi utilizada como um indicador de ressurgência costeira. Os resultados indicam que os níveis de CO2 eram os mais baixos (1962 µmol/kg) na superfície e aumentado para 2451 µmol/kg redox abaixo da interface ico-anóxica. A distribuição vertical de regeneração de carbono foi dominada (82%) por carbono orgânico originário do tecido mole de organismos fotossintéticos, enquanto 18% foram originados da dissolução de calcite biogênica. A regeneração do carbono orgânico foi maior na camada superficial de acordo com os valores de produtividade primária. No entanto, na interface ico-anóxica foi detectado um segundo máximo mais intenso (70-80%), gerado pela respiração quimiotrófica de material orgânico por microrganismos. As percentagens nas camadas anóxicas foram menores porque a decomposição aeróbica ocorre mais rapidamente do que a respiração anaeróbica de matérias orgânicas, pois as fracções mais lábeis de carbono orgânico já foram mineralizadas nas camadas superiores.

Descritores:
Bacia de Cariaco; Regeneração de carbono; Coluna de água; Afloramento costeiro

INTRODUCTION

Oceans play a central role in modulating climate through the absorption and exchange of carbon dioxide from the atmosphere and its sequestration into the oceans' interior. At the same time, the mechanisms of exchange and assimilation of heat make the oceans important reservoirs of CO2 that can influence climatic changes (STEINBERG et al., 2001STEINBERG, D. K.; CARLSON, C. A.; BATES, N. R.; JOHNSON, R. J.; MICHAELS, A. F.; KNAP, A. H. Overview of the US JGOFS Bermuda Atlantic Time series Study (BATS): a decade-scale look at ocean biology and biogeochemistry. Deep Sea Res. Part II: Top. Stud. Oceanogr., v. 48, n. 8-9, p. 1405-1447, 2001.; SITCH et al., 2015SITCH, S.; FRIEDLINGSTEIN, P.; GRUBER, N.; JONES, S. D.; MURRAY-TORTAROLO, G.; AHLSTRÖM, A.; LEVIS, S.; LEVY P. E.; LOMAS M.; POULTER B.; VIOVY N.; ZAEHLE S.; ZENG N.; ARNETH A.; BONAN G.; BOPP L.; CANADELL J. G.; CHEVALLIER F.; CIAIS P.; ELLIS R.; GLOOR M.; PEYLIN P.; PIAO S. L.; LE QUÉRÉ C.; SMITH B.; ZHU Z.; MYNENI R. Recent trends and drivers of regional sources and sinks of carbon dioxide. Biogeosciences, v. 12, n. 3, p. 653-679, 2015.; MCKINLEY et al., 2016MCKINLEY, G. A.; PILCHER, D. J.; FAY, A. R.; LINDSAY, K.; LONG, M. C.; LOVENDUSKI, N. S. Timescales for detection of trends in the ocean carbon sink. Nature, v. 530, p. 469-472, 2016.). The magnitude of the oceanic CO2 sink is influenced by the biological pump, that is, the sequestration of atmospheric CO2 by phytoplankton during primary production and the subsequent export by the sinking of particulates out of the mixed layer. Only a fraction of the sinking particles reach the sea floor; ~1% of organic carbon that sinks out of the surface mixed layer reaches the sediment (BERGER et al., 1988BERGER, W. H.; FISCHER, K.; LAI, C.; WU, G. Ocean carbon flux: Global maps of primary production and export production. In: AGEGIAN, C. (Ed.). Biogeochemical cycling and fluxes between the deep euphotic zone and other oceanic realms. Rockville: U.S. Dept. of Commerce, National Oceanic and Atmospheric Administration, Oceanic and Atmospheric Research, Office of Undersea Research, 1988. p. 131-176.; THUNELL et al., 2000THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000.; SMOAK et al., 2004SMOAK, J. M.; BENÍTEZ-NELSON, C.; MOORE, W. S.; THUNELL, R.; ASTOR, Y.; MULLER-KARGER, F. Radionuclide fluxes and particle scavenging in Cariaco Basin. Cont. Shelf Res., v. 24, n. 13-14, p. 1451-1463, 2004.; THUNELL et al., 2007THUNELL, R. C.; BENITEZ-NELSON, C.; VARELA, R.; ASTOR, Y.; MULLER-KARGER, F. E. Particulate organic carbon fluxes along upwelling-dominated continental margins: Rates and mechanisms. Global Biogeochem. Cycles, v. 21, n. 1, p. 1-12, 2007.).

The continental margins are sites of active physical and biogeochemical processes that are relevant to the global carbon cycle (LIU et al., 2000LIU, K. K.; ATKINSON, L.; CHEN, C. T. A.; GAO, S.; HALL, J.; MACDONALD, R. W.; MCMANUS, L. T.; QUINONES, R. Exploring continental margin carbon fluxes in the global context. Eos, Trans. Amer. Geophys. Union., v. 81, n. 52, p. 641- 644, 2000.; ASTOR et al., 2003ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003.; ASTOR et al., 2013ASTOR, Y. M.; LORENZONI, L.; THUNELL, R.; VARELA, R.; MULLER-KARGER, F.; TROCCOLI, L.; TAYLOR, G. T.; SCRANTON, M. I.; TAPPA, E; RUEDA-ROA, D. Interannual variability in sea surface temperature and fCO2 changes in the Cariaco Basin. Deep Sea Res. Part II Top. Stud. Oceanogr., v. 93, p. 33-43, 2013.; BAUER et al., 2013BAUER, J. E.; WEI-JUN, C.; RAYMOND, P. A.; BIANCHI, T. S.; HOPKINSON, C. S.; REGNIE, P. A. G. The changing carbon cycle of the coastal ocean. Nature, v. 504, n. 7478, p. 61-70, 2013.). These include regions of coastal upwelling where intermediate and deep water masses rise to the surface, degas, and generate a flux of CO2 into the atmosphere (SMITH, 1994SMITH, S. Phosphorus versus nitrogen limitation in the marine enviroment. Limnol. Oceanogr. v. 29, n.6, p 1149-1160,1994.; COBO-VIVEIROS et al., 2013COBO-VIVEROS, A. M.; PADIN, X. A.; OTERO, P.; DE LA PAZ, M.; RUIZ-VILLARREAL, M.; RÍOS, A. F.; PÉREZ, F. F. Short-term variability of surface carbon dioxide and sea-air CO2 fluxes in the shelf waters of the Galician coastal upwelling system. Sci. Mar., v. 77, n. 1, p. 37-48, 2013.). In addition, limiting nutrients for the biota are transported by the upwelling waters to the surface layers and help sustain primary productivity (HUTCHINGS et al., 1994; IANSON, 2003IANSON, D.; ALLENA, S. E.; HARRISA, S. L.; ORIANSA, K. J.; VARELA, D. E.; WONG, C. S. The inorganic carbon system in the coastal upwelling region west of Vancouver Island, Canada. Deep Sea Res. Part I: Oceanogr. Res. Pap. v. 50, n. 8, p. 1023-1042, 2003.; CAPONE; HUTCHINS, 2013CAPONE, D. G.; HUTCHINS, D. A. Microbial biogeochemistry of coastal upwelling regimes in a changing ocean. Nat. Geosci., v. 6, p. 711-717, 2013.; SCRANTON et al.; 2014SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.). The transport of biological material to the ocean's interior is the main driver of the non-zero gradient of many dissolved species in seawater and an important regulator of atmospheric CO2 on millennial timescales (LOUANCHI; NAJJAR, 2000LOUANCHI, F.; NAJJAR, R. G. A global monthly climatology of phosphate, nitrate, and silicate in the upper ocean: Spring-summer export production and shallow remineralization. Global Biogeochem. Cycles, v. 14, n. 3, p. 957-977, 2000.; SIPPO et al.; 2016SIPPO, J. Z.; MAHER, D. T.; TAIT, D. R.; HOLLOWAY, C.; SANTOS, I. R. Are mangroves drivers or buffers of coastal acidification? Insights from alkalinity and dissolved inorganic carbon export estimates across a latitudinal transect. Global Biogeochem. Cycles, v. 30, n. 5, p. 753-766, 2016.). In the open ocean, the fluxes of carbon through the water column decease exponentially with water depth due to aerobic respiration by microorganisms (PACE et al., 1987PACE, M.; KNAUER, G. A.; KARL, D. M.; MARTIN, J. H. Primary production, new production and vertical flux in the eastern Pacific Ocean. Nature, v. 325, p. 803-804, 1987.; PACKARD et al., 2015PACKARD, T. T.; OSMA, N.; FERNÁNDEZ-URRUZOLA, I.; CODISPOTI, L. A.; CHRISTENSEN, J. P.; GÓMEZ, M. Peruvian upwelling plankton respiration: calculations of carbon flux, nutrient retention efficiency, and heterotrophic energy production. Biogeosciences, v. 12, n. 9, p. 2641-2654, 2015.).

Atmospheric carbon dioxide (CO2) has increased from 278 to 400 parts per million (ppm) over the industrial period and, together with the increase of other greenhouse gases, has driven a series of major environmental changes (GATTUSO et al., 2015GATTUSO, J. P.; MAGNAN, A.; BILLÉ, R.; CHEUNG, W.; HOWES, E. L.; JOOS, F.; ALLEMAND, D.; BOPP, L.; COOLEY, S. R.; EAKIN, C. M.; HOEGH-GULDBERG, O.; KELLY, R. P.; PÖRTNER, H. O.; ROGERS, A.; BAXTER, J. M.; LAFFOLEY, D.; OSBORN, D.; RANKOVIC, A.; ROCHETTE, J.; SUMAILA, U. R.; TREYER, S.; TURLE, C. Contrasting futures for ocean and society from different anthropogenic CO2 emissions scenarios. Science., v. 349, n. 6243, p. aac4722, 2015.). Approximately 30% of the anthropogenic carbon emitted from fossil fuels is removed from the atmosphere by the ocean, which may lead to an uncertain future for the planet's climate (OCCC, 2004OCCC. 2004. Ocean carbon climate change. Scott. C. Doney editor. USA.108 pp.; JOHN et al., 2007JOHN, D.; WANG, Z. A.; LIU, X.; BYRNE, R. H.; CORREDOR, J.; LÓPEZ, J. M.; CABRERA, A.; BRONK, D. A.; TABITA, F. R.; PAUL, J. H. Phytoplankton carbon fixation gene (RuBisCO) transcripts and air-sea CO2 flux in the Mississippi River plume. ISME J., v. 1, n. 6, p. 517-531, 2007.). Biogeochemical ocean circulation models suggest that the ocean absorbed approximately 37 Pg C of anthropogenic C between 1994 and 2010, increasing the inventory from 118±20 Pg C to 155±31 Pg C. This is equivalent to a mean annual uptake rate of approximately 2.3 Pg C year, or approximately 27% of the total anthropogenic C emissions over this time period (KHATIWALA et al., 2013KHATIWALA, S.; TANHUA, T.; MIKALOFF FLETCHER, S.; GERBER, M.; DONEY, S. C.; GRAVEN, H. D.; GRUBER, N.; MCKINLEY, G. A.; MURATA, A.; RÍOS, A. F.; SABINE, C. L. Global ocean storage of anthropogenic carbon. Biogeosciences, v. 10, p. 2169-2179, 2013.).

The present investigation was carried out within the framework of the Cariaco Time-Series Program (Carbon Retention in a Colored Ocean) with the objective of estimating the percentages of organic and inorganic regenerated carbon in the Cariaco Basin.

MATERIAL AND METHODS

Study area

The Cariaco Basin (Figure 1) is the largest anoxic and truly marine pelagic system (MADRID et al., 2001MADRID, VANESSA, M., et al. Phylogenetic diversity of bacterial and archaeal communities in the anoxic zone of the Cariaco Basin. Applied and environmental microbiology, v. 67, n. 4, p. 1663-1674, 2001.). It forms a depression ca. 160 km long and 70 km wide on the Venezuelan continental shelf with an approximate maximum water depth of 1400 m (MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.; ASTOR et al., 2013ASTOR, Y. M.; LORENZONI, L.; THUNELL, R.; VARELA, R.; MULLER-KARGER, F.; TROCCOLI, L.; TAYLOR, G. T.; SCRANTON, M. I.; TAPPA, E; RUEDA-ROA, D. Interannual variability in sea surface temperature and fCO2 changes in the Cariaco Basin. Deep Sea Res. Part II Top. Stud. Oceanogr., v. 93, p. 33-43, 2013.). A pronounced seasonality can be observed in water column physical parameters at the start of the year caused by the seasonal displacement of the Inter-Tropical Convergence Zone (ITCZ) that generates significant variability in wind regime, currents and river discharge (RICHARDS, 1960RICHARDS, F. Some chemical and hydrographic observations along the north coast of South America - I. Cabo Tres Puntas to Curacao, including the Cariaco Trench and the Gulf of Cariaco. Deep Sea Res, 1960, v. 7, n. 3, p. 163-182, 1960.; HERRERA; FEBRES ORTEGA, 1975HERRERA, L. E. FEBRES-ORTEGA, G. Kinematics of the windgenerated velocity field in the surface waters off eastern Venezuela, Caribbean Sea. Bol. Inst. Oceanogr. Univ. Oriente, 1975, vol. 14, no 2, p. 165-186.; MULLER-KARGER; APARICIO, 1994MULLER-KARGER, F. E; APARICIO, R. Mesoscale Processes Affecting Phytoplankton Abundance in the Southern Caribbean Sea. Cont. Shelf Res., v. 14, n. 2-3, p. 199-221, 1994.; ASTOR et al., 1998ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.; RUEDA-ROA; MULLER-KARGER, 2013). These changes lead to an increase in primary production that exceeds 500 gCm-2 yr-1 (MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.). The surface water layers within the basin also experience a decrease in temperature, mainly due to the influence of coastal upwelling (MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.). Below 200 m, the water mass is isolated and is characterized by near-constant temperature, salinity and total anoxia (ASTOR et al., 1998ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.; SCRANTON et al., 2002SCRANTON, M., et al. Controls on temporal variability of the geochemistry of the deep Cariaco Basin. Deep Sea Res.I: Oceanogr. Res. Papers, 2001, v. 48, n.7, p. 1605-1625, 2002, ASTOR et al., 2013ASTOR, Y. M.; LORENZONI, L.; THUNELL, R.; VARELA, R.; MULLER-KARGER, F.; TROCCOLI, L.; TAYLOR, G. T.; SCRANTON, M. I.; TAPPA, E; RUEDA-ROA, D. Interannual variability in sea surface temperature and fCO2 changes in the Cariaco Basin. Deep Sea Res. Part II Top. Stud. Oceanogr., v. 93, p. 33-43, 2013.). Due to the geomorphology of the basin, advection is confined to the surface layers of the water column, from which biogenic material and detritus rich in opal, organic carbon and lithogenic material are exported to the deeper waters (THUNELL et al., 2000THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000.; MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.). The laminated sediments underlying the anoxic water mass serve as an important repository for past variability of inter- and extra-tropical climate (REUER et al., 2003REUER, M.; BOYLE, K., EDWARD A.; COLE, JULIA, E. A mid-twentieth century reduction in tropical upwelling inferred from coralline trace element proxies. Earth Planet SC Lett, v. 210, n. 3, p. 437-452, 2003.; MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.; SCRANTON et al., 2014SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.).

Figure 1
Northeastern Venezuela showing the Cariaco Basin. The sampling station (CARIACO) is indicated (10º30' N, 64º40' W).

The data presented in this study originate from samples that were collected monthly from January 2006 to September 2007 at the Cariaco sampling station (Figure 1) at 18 water depths from the surface down to 1310 m (1, 7, 15, 25, 35, 55, 75, 100, 130, 160, 200, 250, 300, 350, 400, 500, 750, 1310 m). Water samples were recovered with a CTD (SBE 25) rosette equipped with 12 Niskin bottles (8 L) and dissolved oxygen, fluorescence and light attenuation sensors. The upper layers of the basin present the most pronounced changes in temperature due to the influence of coastal upwelling and lateral advection of cooler water masses from the Caribbean Sea (ASTOR et al., 1998ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.; MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.; ASTOR et al., 2003ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003.). Vertical mixing below 100 m is inhibited by the presence of a strong pycnocline (SCRANTON et al., 1987; SCRANTON et al., 2014SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.). Temperatures were only used down to 200 meters. Below 200 m, the waters are isolated and acquire anoxic characteristics with almost constant temperature and salinity (ASTOR et al., 1998ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.; ASTOR et al., 2013ASTOR, Y. M.; LORENZONI, L.; THUNELL, R.; VARELA, R.; MULLER-KARGER, F.; TROCCOLI, L.; TAYLOR, G. T.; SCRANTON, M. I.; TAPPA, E; RUEDA-ROA, D. Interannual variability in sea surface temperature and fCO2 changes in the Cariaco Basin. Deep Sea Res. Part II Top. Stud. Oceanogr., v. 93, p. 33-43, 2013.; SCRANTON et al., 2014SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.). Primary productivity (PP) was determined from the inorganic carbon fraction assimilated by phytoplankton using 14C bicarbonate following the method of STEEMAN-NIELSEN (1952)STEEMAN-NIELSEN, E. The use of radioactive carbon (14C) for measuring organic production in the sea. J. Cons. Perm. Int. Explor. Mer., v. 18, n. 2, p. 117-140, 1952. with modifications by KNAP et al. (1997)KNAP, A. H.; MICHAELS, A. F.; STEINBERG, D.; BAHR, F.; BATES, N.; BELL, S.; COUNTWAY, P.; CLOSE, A.; DOYLE, A.; DOW, R.L.; HOWSE, F.; GUNDERSEN, K.; JOHNSON, R.; KELLY, R.; LITTLE, R.; ORCUTT, K.; PARSONS, R.; RATHBUN, C.; SANDERSON, M.; STONE, S. BATS methods manual. Vol. 4. Woods Hole: U.S. JGOFS Planning Office, Woods Hole, 1997. 136 p. and UNESCO (1994)UNESCO/Intergovernmental Oceanographic Commission. Protocols for the Joint Global Ocean flux Study (JGOFS). Core measurements. Manual and Guides, v. 29. Paris: UNESCO, 1994. p. 128-134.. The procedure is derived from that employed in BATS (Bermuda Atlantic Time-series Study, KNAP et al., 1997KNAP, A. H.; MICHAELS, A. F.; STEINBERG, D.; BAHR, F.; BATES, N.; BELL, S.; COUNTWAY, P.; CLOSE, A.; DOYLE, A.; DOW, R.L.; HOWSE, F.; GUNDERSEN, K.; JOHNSON, R.; KELLY, R.; LITTLE, R.; ORCUTT, K.; PARSONS, R.; RATHBUN, C.; SANDERSON, M.; STONE, S. BATS methods manual. Vol. 4. Woods Hole: U.S. JGOFS Planning Office, Woods Hole, 1997. 136 p.) adjusted for the elevated productivity of the sampling station. The minimum incubation time of the samples was between 4 and 5 h, which detects C fixation rates of 0.05 to 100 mg C m-3 h-1. Analysis of total alkalinity (TA) was performed following BRELAND and BYRNE (1993)BRELAND, J. A.; BYRNE, R. H. Spectrophotometric procedures for determination of sea water alkalinity using bromocresol green. Deep Sea Res. Part I: Oceanogr. Res. Pap., v. 40, n. 3, p. 629-641, 1993.. Certified reference materials (supplied by A. Dickson) were used. Water samples for the determination of TA were siphoned from the Niskin bottles into dark 250 ml borosilicate containers and fixed with 50 ml saturated HgCl2 (equivalent to 50 µM Hg). The samples were stored at 4 ºC until analysis. In the laboratory, samples were acidified by titration to the inflexion point using bromocresol indicator. TA concentrations were calculated using absorbance readings at 444 nm and 616 nm, with a precision of ± 4 µmol kg-1. Samples for the determination of pH were taken directly from the Niskin bottles and measured within 1 h by spectrophotometry (Ocean Optics SD-1000) using 10 cm cells (CLAYTON; BYRNE, 1993CLAYTON, T. D.; BYRNE, R. H. Spectrophotometric seawater pH measurements: total hydrogen ion concentration scale calibration of m-cresol purple and at-sea results. Deep Sea Res. Part I: Oceanogr. Res. Pap., v. 40, n. 10, p. 2115-2129, 1993.; DICKSON; GOYET, 1994DICKSON, A. G.; GOYET, C. (Eds). DOE. Handbook of methods for the analysis of the various parameters of the carbon dioxide system in sea water, v. 2. ORNL/CDIAC-74. Washington: U. S. Department of Energy, 1994. 187 p.). The procedure measures the absorbance at 434 nm and 578 nm before and after the addition of the purple indicator m-cresol. The cells were placed in a water bath at 25 ºC for 15 minutes prior to measurement. This methodology is used for the CARIACO time-series project and more details are given in ASTOR et al. (2005)ASTOR, Y. M.; SCRANTON, M. I.; MULLER-KARGER, F.; BOHRERD, R.; GARCIA, J. fCO2 variability at the CARIACO tropical coastal upwelling time series station. Mar. Chem., v. 97, n. 3-4, p. 245-261, 2005.. The precision was ± 0.003 pH units.

TCO2 concentrations were calculated using the program CO2SYS (LEWIS; WALLACE, 1995LEWIS, E.; WALLACE, D. Program developed for CO2 system calculations. 1995. Available:<http://cdiac.ornl.gov/oceans/co2rprt.html>. Accessed: 2016/12/12.
http://cdiac.ornl.gov/oceans/co2rprt.htm...
) from pH and TA using the equilibrium constants of MERHBACH et al. (1973)MEHRBACH, C.; CURBERSON, C.; HAWLEY, J.; PYTKOWICZ, T. Measurement of the apparent dissociation constants of carbonic acid in seawater at atmospheric pressure. Limnol. Oceanogr., v. 18, p. 897-907, 1973. and MILLERO (1995)MILLERO, F. Thermodynamics of the carbon dioxide system in the oceans. Geochim Cosmochim Acta, v. 59, n. 4, p. 661-677, 1995.. This program allows the components of the carbon dioxide system to be determined using two variables only. The precision of this approach was 5.6 µmol/ kg (ASTOR et al., 2005ASTOR, Y. M.; SCRANTON, M. I.; MULLER-KARGER, F.; BOHRERD, R.; GARCIA, J. fCO2 variability at the CARIACO tropical coastal upwelling time series station. Mar. Chem., v. 97, n. 3-4, p. 245-261, 2005.; ASTOR et al., 2013ASTOR, Y. M.; LORENZONI, L.; THUNELL, R.; VARELA, R.; MULLER-KARGER, F.; TROCCOLI, L.; TAYLOR, G. T.; SCRANTON, M. I.; TAPPA, E; RUEDA-ROA, D. Interannual variability in sea surface temperature and fCO2 changes in the Cariaco Basin. Deep Sea Res. Part II Top. Stud. Oceanogr., v. 93, p. 33-43, 2013.).

The regeneration of carbon in the water column was calculated from TA and TCO2 using equations (1) and (2) in BROECKER and PENG (1982)BROECKER, W.; PENG, T. H. Tracers in the sea. New York: Eldigio Press, 1982. p. 690.. This method calculates the levels of organic and inorganic carbon regenerated in the water column from the slope of the TA vs TCO2 regression where TCO2=∑CO2. The model considers that, in the case of dissolution of calcareous skeletons (CaCO3), TA increases twice as much as ∑CO2. Similarly, when biogenic tissue is mineralized, TA decreases by one unit for each mole of nitrate formed by the nitrification of organic nitrogen with a corresponding increase in ∑CO2 set by the C:N ratio of the organic material (RÍOS et al., 1989RÍOS, A. F.; FRAGA, F.; PÉREZ, F. F. Estimation of coefficients for the calculation of ''NO'', ''PO'' and ''CO'', starting from the elemental composition of natural phytoplankton. Sci. Mar., v. 53, n. 4, p. 779-784, 1989.). Based on equations used by (BROECKER; PENG,1982BROECKER, W.; PENG, T. H. Tracers in the sea. New York: Eldigio Press, 1982. p. 690.) on inorganic carbon derived from calcareous test and organic carbon from soft tissue (REDFIELD, 1963REDFIELD, A.C., KETCHUM, B.H., RICHARDS, F.A. The influence of organisms on the composition of seawater. In: Hill, M.N. Ed.., The Sea, v. 2, p. 26-77, 1963.), is the slope of the TA vs TCO2 relationship (RÍOS et al., 1995RÍOS, A. F.; ANDERSON, T. R.; PÉREZ, F. F. The carbonic system distribution and fluxes in the NE Atlantic during spring 1991. Progr. Oceanogr., v. 35, n. 4, p. 295-314, 1995.).

(1) Δ TA Δ Σ CO 2 = 2 Δ C C a C O 3 Δ C C a C O 3 + Δ C Corg Δ C Corg R Δ C C a C O 3 + Δ C Corg
(2) Δ C C a C O 3 Δ C Corg = 0 . 5 Δ TA Δ Σ CO 2 + 1 R Δ TA Δ Σ CO 2

The percentage of organic and inorganic carbon was determined for each month in order to investigate possible seasonal changes in the regeneration of carbon in the regression model. Monthly profiles of TA and TCO2 were regressed to obtain the slope. Thus, all the TA and TCO2 were averaged at each depth and used in the regression model.

To validate the data and determine if monthly and annual variations existed in the percentages of regenerated carbon, multifactor ANOVA (analysis of variance type III) was used at a significance level of 5%. Homogeneity of variance was checked using the Cochran test for 18 months and 306 observations. To identify the homogeneous groups in the analysis, the multiple range LSD test was used at a level of 5% (SOKAL; ROHLF, 1969SOKAL ROBERT, R.; JAMES, ROHLF F. Biometry. The principles and practice of statistics in biological research. 1969., JOHNSON; WICHERN, 1992JOHNSON, R. A.; WICHERN, D. W. Applied Multivariate Statistical Analysis. 3rd ed. New York: Prentice Hall, 1992. 156 p.). The ANOVA analysis was made using the statistical package Statgraphics plus 5.1.

RESULTS

Primary production in the Cariaco Basin presented strong seasonality, with higher rates during the upwelling season that occurs from November until April (Figure 2). At this time, water ascends to the surface from ~ 100-120 m depth and is relatively cool (22 to 23ºC; Figure 3) and enriched in nutrients (ASTOR et al., 1998ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.; MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.; ASTOR et al., 2003ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003.; SCRANTON et al., 2014SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.). Despite the high productivity, TCO2 concentrations were relatively low in the photic zone (< 200 m), with values between 1968 and 2300 µmol kg-1. Below the anoxic chemocline (> 250 m), TCO2 concentrations increased to 2451 µmol kg-1 (Figure 4).The monthly and annual percentages of organic and inorganic carbon regenerated in the Cariaco Basin showed no significant seasonal differences at a significance level of p<0.05. The p-values test the statistical significance of each of the factors. Since no p-values were less than 0.05, none of the factors had a statistically significant effect on organic and inorganic carbonic at the 95% confidence level. The mean percentages of inorganic and organic carbon over the study period were 24% and 76%, respectively (Figure 5).

Figure 2
Distribution of primary productivity (mg C m-3 h-1) at the sampling station in Cariaco Basin from January 2006 to September 2007.

Figure 3
Distribution of temperature (ºC) at the sampling station in Cariaco Basin from January 2006 to September 2007. Gray contours show the variability in the 23ºC isotherm during coastal upwelling.

Figure 4
Distribution of TCO2 (µmol/kg) at the sampling station in Cariaco Basin from January 2006 to September 2007.

Figure 5
Monthly percentages of depth-integrated regenerated carbon (inorganic and organic) at the sampling station in Cariaco Basin from January 2006 to September 2007.

DISCUSSION

Coastal upwelling systems account for approximately half of the global ocean's primary production and contribute disproportionately to biologically driven carbon sequestration (ABRANTES et al., 2016ABRANTES, F.; CERMENO, P.; LOPES, C.; ROMERO, O.; MATOS, L.; VAN IPEREN, J.; RUFINO, M.; MAGALHÃES, V. Diatoms Si uptake capacity drives carbon export in coastal upwelling systems. Biogeosciences, v. 13, p. 4099-4109, 2016.). Diatoms, silica-precipitating microalgae, constitute the dominant phytoplankton in these productive regions, and their abundance and assemblage composition in the sedimentary record is considered one of the best proxies for primary productivity (ABRANTES et al., 2016ABRANTES, F.; CERMENO, P.; LOPES, C.; ROMERO, O.; MATOS, L.; VAN IPEREN, J.; RUFINO, M.; MAGALHÃES, V. Diatoms Si uptake capacity drives carbon export in coastal upwelling systems. Biogeosciences, v. 13, p. 4099-4109, 2016.). In the Cariaco Basin primary production is dominated by diatoms (THUNELL et al., 2000THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000.; ROMERO et al., 2009ROMERO, O. E.; THUNELL, R. C.; ASTOR, Y.; VARELA, R. Seasonal and interannual dynamics in diatom production in the Cariaco Basin, Venezuela. Deep Sea Res. Part I: Oceanogr. Res. Pap., v. 56, n. 4, p. 571-581, 2009.) and is mainly confined to the upper 50 m of the water column (Figure 2). Production is relatively higher in the upper 7 m, with values exceeding 4 mg C m-3 h-1. However, levels of 1 mg C m-3 h-1 were detected between 25 and 50 m throughout the study period. The highest rates measured in this study were 17 mg C m-3 h-1 between January and March. An interesting finding was that in non-upwelling months, for example July 2002, uncharacteristically high values (5 mg C m-3 h-1) were measured. These values are influenced by a secondary upwelling caused by the regional geostrophic flow FEBRES-ORTEGA; HERRERA, 1975HERRERA, L. E. FEBRES-ORTEGA, G. Kinematics of the windgenerated velocity field in the surface waters off eastern Venezuela, Caribbean Sea. Bol. Inst. Oceanogr. Univ. Oriente, 1975, vol. 14, no 2, p. 165-186.; HERRERA; FEBRES-ORTEGA, 1975HERRERA, L. E. FEBRES-ORTEGA, G. Kinematics of the windgenerated velocity field in the surface waters off eastern Venezuela, Caribbean Sea. Bol. Inst. Oceanogr. Univ. Oriente, 1975, vol. 14, no 2, p. 165-186.; MULLER-KARGER et al., 2001MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.; LORENZONI et al., 2013LORENZONI, L.; TAYLOR, G. T.; BENITEZ-NELSON, C.; HANSELL, D. A.; MONTES, E.; MASSERINI, R.; FANNING, K.; VARELA, R.; ASTOR, Y.; GUZMÁN, L.; MULLER-KARGER, F. E. Spatial and seasonal variability of dissolved organic matter in the Cariaco Basin. J. Geophys. Res. Biogeosci., v. 118, p. 1-12, 2013.). The high productivity is reflected in the annually integrated values which exceed 600 g C m-3 h-1. These agree with the rates of 540 to 690 g C m-3 h-1 reported by MULLER-KARGER et al. (2001)MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.. MULLER-KARGER et al. (2001)MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001. and ASTOR et al. (2003)ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003. further indicate that, besides upwelling, ventilation processes in this zone also induce an increase in productivity. Temporal variability in the production and export of diatoms in the Cariaco Basin is primarily controlled by the seasonal cycle of primary production in the surface waters. When the ITCZ reaches its southernmost position during the boreal winter, strong E-NE trade winds enhance the upwelling of nutrients into surface waters, which in turn results in increased fluxes of diatoms and bulk biogenic components to the seafloor. The ITCZ moves northward in the boreal summer, resulting in weakened E-NE trades (ASTOR et al., 2003ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003.), decreased input of nutrients to the surface water and reduced primary production. Diatom and opal fluxes reach their lowest values during boreal summers off Venezuela (ROMERO et al., 2009ROMERO, O. E.; THUNELL, R. C.; ASTOR, Y.; VARELA, R. Seasonal and interannual dynamics in diatom production in the Cariaco Basin, Venezuela. Deep Sea Res. Part I: Oceanogr. Res. Pap., v. 56, n. 4, p. 571-581, 2009.).

The regeneration of the monthly organic carbon fraction varied between 60 and 82% (Figure 5), between January and June 2006 and then decreased to 68% in July 2006. Thereafter, the variability was low and was maintained between 70% and 83% during the following months. The variation oscillated between 18% and 40% for inorganic carbon, with decreases from 40% to 18% from January to June 2006 and a secondary maximum of 32% in July 2006. The decrease in the percentage of regenerated organic carbon, although lower in the first months of the year -particularly in 2006, coincides with upwelling intensity, which was less intense in 2006 than 2007 (Figure 3). In July 2006, the decrease was lower than in July 2007 due to a stronger increase of the 23 ºC isothermal. FEBRES-ORTEGA (1974)FEBRES ORTEGA, G. Circulación de las aguas superiores de la Fosa de Cariaco en Abril de 1974. Bol. Inst. Oceanogr. Univ. Oriente, v. 13, n. 1-2, p. 79-86, 1974. indicated that the geostrophic flow in the Cariaco Basin has a notable influence in the upwelling of cold water to the surface. The upwelling events are characteristic of the central and eastern coast of Venezuela, originating when the Trade winds blow from NE-E-SE to NW-W-SW (MULLER-KARGER; APARICIO, 1994MULLER-KARGER, F. E; APARICIO, R. Mesoscale Processes Affecting Phytoplankton Abundance in the Southern Caribbean Sea. Cont. Shelf Res., v. 14, n. 2-3, p. 199-221, 1994.; ASTOR et al., 1998ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.; ASTOR et al., 2003ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003.; ALVERA-AZCARATE et al., 2009ALVERA-AZCÁRATE, A.; BARTH, A.; WEISBERG, R. A nested model of the Cariaco Basin (Venezuela): description of the basin's interior hydrography and interactions with the open ocean. Ocean. Dynam. v. 59, n.1, p. 97-120, 2009.; RUEDA-ROA; MULLER- KARGUER, 2014RUEDA-ROA, D.; MULLER-KARGER, F. The southern Caribbean upwelling system: Sea surface temperature, wind forcing and chlorophyll concentration patterns. Deep. Sea. Res.I. v. 78, p. 102-114, 2014.). However, lateral advection also has an influence, transporting water masses from the regions of greatest upwelling, such as the north coast of Sucre State (FEBRES-ORTEGA, 1974FEBRES ORTEGA, G. Circulación de las aguas superiores de la Fosa de Cariaco en Abril de 1974. Bol. Inst. Oceanogr. Univ. Oriente, v. 13, n. 1-2, p. 79-86, 1974.; RUEDA-ROA; MULLER-KARGUER, 2014RUEDA-ROA, D.; MULLER-KARGER, F. The southern Caribbean upwelling system: Sea surface temperature, wind forcing and chlorophyll concentration patterns. Deep. Sea. Res.I. v. 78, p. 102-114, 2014.). The regeneration of organic carbon is a source of marine dissolved organic matter (DOM), which is one of the largest reservoirs of reduced carbon in the ocean. Thus small changes in the cycling of DOM can have dramatic impacts on the magnitude of CO2 released or sequestered by marine systems (~662 Pmol C) (HANSELL et al., 2009HANSELL, D. A.; CARLSON, C. A.; REPETA, D. J.; SCHLITZER, R. Dissolved organic matter in the ocean: A controversy stimulates new insights. Oceanography, v. 22, n. 4, p. 202-211, 2009.; LORENZONI et al., 2013LORENZONI, L.; TAYLOR, G. T.; BENITEZ-NELSON, C.; HANSELL, D. A.; MONTES, E.; MASSERINI, R.; FANNING, K.; VARELA, R.; ASTOR, Y.; GUZMÁN, L.; MULLER-KARGER, F. E. Spatial and seasonal variability of dissolved organic matter in the Cariaco Basin. J. Geophys. Res. Biogeosci., v. 118, p. 1-12, 2013.).

The slope of the regression curve of the TA vs TCO2 (Eq. 2) was 0.22 (Figure 6). Accordingly, when the variation of organic carbon produced by the decomposition of soft tissue (∆Corg) is equal to 1, the variation of inorganic carbon generated by the dissolution of hard or calcareous skeletons (∆CCaCO3) is 0.22. The sum of both fractions is 1.22; therefore the percentage of regenerated inorganic carbon is equal to (0.22/1+0.22) × 100 = 18%. Similarly, by difference, the regeneration of organic carbon, dominated by soft or organic material, is 82%.

Figure 6
TA versus TCO2 at the sampling station in Cariaco Basin from January 2006 to September 2007.

These values are similar to those reported by BROECKER and PENG (1982)BROECKER, W.; PENG, T. H. Tracers in the sea. New York: Eldigio Press, 1982. p. 690., who obtained 1 mol CaCO3 per 4 moles Corg using a slope of ∑CO2 versus TA with surface data and Antarctic bottom water. The results in this investigation are in agreement with reports by RÍOS et al. (1995)RÍOS, A. F.; ANDERSON, T. R.; PÉREZ, F. F. The carbonic system distribution and fluxes in the NE Atlantic during spring 1991. Progr. Oceanogr., v. 35, n. 4, p. 295-314, 1995. in the northeastern Atlantic, which indicated that the regeneration of carbon in water layers between 0 and 1000 m and > 2300 m is controlled by the fraction of soft tissue in the surface layer (86%), with the contribution of inorganic or hard calcareous tests being only 14%.

The vertical distribution of regenerated organic carbon in the Cariaco Basin (Figure 7) shows highest variability in the photic zone, with two maxima of 81.1% and 84.1% at 1m and 55 m, respectively. The 81.1% value coincides with the maximum in primary productivity. However, the 55 m peak suggests mineralization of organic material. Other mechanisms that cannot be ruled out are contributions from the vertical migration of mesozooplankton and nekton such as Bregmaceros cantori which migrates between the oxic water layers to anoxic layers deeper than 500 m (MARÍN et al., 1989MARÍN, B.; GÓMEZ, A.; UROSA, L. Taxonomía y distribución del Ictioplancton en la fosa de Cariaco, Venezuela. Bol. Inst. Oceanogr. Venezuela, v. 28, n. 1-2, p. 15-27. 1989.), as well as lateral advection that produces anomalies in the sedimentation of particulate material (TAYLOR et al., 2000). These migrations, along with the microbial production in situ, have important consequences in the anomalies of the vertical flux of carbon in the Cariaco Basin, and may enrich the carbon fluxes below 275 m, in particular in the proximity of 455 m water depth (TAYLOR et al., 2000). On the other hand, at depths between 7 and 35 m the values of regenerated organic carbon oscillated between approximately 44% and 64%, respectively. Between 55 m and 250 m there was little variation, with values between 52% and 64%. Below the anoxic threshold, located between 250 and 350 m, a third maximum of 70 - 80% was detected, which is lower than those in the photic zone. Most previous studies have found the oxic/anoxic interface in the Cariaco Basin (usually defined by the appearance of sulfide) to be at between 250 m and 300 m water depth (MCPARLAND et al., 2013MCPARLAND, E., BENITEZ-NELSON, C., TAYLOR, G., THUNELL. G., ROLLINS, A., LORENZONI, L. Cycling of suspended particulate phosphorus in the redoxcline of the Cariaco Basin. Mar. Chem., 2015, v. 176, p. 64-74, 2013; SCRANTON et al., 2014SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.). REIMERS and SUESS (1983)REIMERS, C.; SUESS, E. The partitioning of organic carbon fluxes and sedimentary organic matter decomposition rates in the ocean. Mar. Chem., v. 13, n. 2, p. 141-168, 1983., EMERSON (1985)EMERSON, S. Organic carbon preservation in marine sediments. In: SUNQUIST, E. T.; BROECKER, W. S. (Ed.). The carbon cycle and atmospheric CO2: Natural variations archean to present. New York: American Geophysical Union, 1985. p. 78-87. and THUNELL et al. (2000)THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000. note that the degradation and preservation of organic carbon in marine ecosystems is commonly associated with the availability of dissolved oxygen, whereby decomposition under aerobic conditions is usually more efficient than under anaerobic conditions (BLAIR et al., 2012BLAIR, N. E.; ALLER, R. C. The fate of terrestrial organic carbon in the marine environment. Ann. Rev. Mar. Sci., v. 4, p. 401-423, 2012.; MEHRBACH et al., 1973MEHRBACH, C.; CURBERSON, C.; HAWLEY, J.; PYTKOWICZ, T. Measurement of the apparent dissociation constants of carbonic acid in seawater at atmospheric pressure. Limnol. Oceanogr., v. 18, p. 897-907, 1973., KEELLER et al., 2002KEELLER, K.; SLATER, R.; BENDER, M.; KEY, R. Possible biological or physical explanations for decadal scale trends in North Pacific nutrient concentration and oxygen utilization. Deep. Sea. Res. II., v. 49, p. 325-362. 2002.). This is corroborated by the proportions of regenerated organic C determined in the oxic and anoxic layers of the Cariaco Basin. This appears to be different from observations made in the Black Sea, where mineralization is a sink for organic matter and DOC, even under anoxic conditions, obviates any associated DOC accumulation. Some workers have also suggested that concentrations are high relative to oceanic levels because, unlike the ocean, the Black Sea receives a proportionately large amount of terrigenous DOC from rivers such as the Danube (MARGOLIN et al., 2016MARGOLIN, A.; GERRINGA, L. J. A.; HANSELL, D. A.; RIJKENBERG, M. J. A. Net removal of dissolved organic carbon in the anoxic waters of the Black Sea. Mar. Chem., v. 183, p. 13-24, 2016.).

Figure 7
Vertical distribution of the regeneration of organic and inorganic carbon at the sampling station in Cariaco Basin from January 2006 to September 2007.

The two maxima at the surface are in agreement with the high productivity which attains values of 17 mgC m-3 h-1 (Figure 2). The third peak at 300 m is associated with intense chemotrophic microbial activity that utilizes reduced sulfur species and methane as energy sources (TAYLOR et al., 2001TAYLOR, G. T.; IABICHELLA, M.; HO, T. Y., SCRANTON, M. I.; TUNNEL, R. C.; MULLER-KARGER, F.; VARELA, R. Chemoautrotophy in the redox transition zone of the Cariaco Basin: A significant midwater source of organic carbon production. Limnol. Oceanogr., v. 46, n. 1, p. 148-163, 2001.). Stratified water columns, disposed to anoxia, contribute to the biological production of carbon at the redoxcline (SOROKIN, 1972SOROKIN, Y. I. The bacterial population and the process of hydrogen sulphide oxidation in the Black Sea. J. Cons. Perm. Inter. Explor. Mer., v. 34, p. 423-455, 1972.; INDREBØ et al., 1979INDREBØ, G.; PENGERUD, B.; DUNSA, I. Microbial activities in a permanently stratified estuary. I. Primary production and sulfate reduction. Mar. Biol., v. 51, n. 4, p. 295-304, 1979.; JØRGENSEN et al., 1991JØRGENSEN, B. B.; FOSSING, H.; WIRSEN, C. O.; JANNASCH, H. W. Sulfide oxidation in the anoxic Black Sea chemocline. Deep Sea Res. Part I: Oceanogr. Res. Pap. v. 38, Suppl. 2, p. S1083-S1103, 1991.; SOROKIN et al., 1995SOROKIN, Y. I.; SOROKIN, P. Y.; AVDEEV, V. A.; SOROKIN, D. Y.; ILCHENKO, S. V. Biomass, production and activity of bacteria in the Black Sea, with special reference to chemosynthesis and the sulfur cycle. Hydrobiology, v. 308, n. 1, p. 61-76, 1995.; WARDS, 2008WARD, B. B. Nitrification in marine systems. In: CAPONE, D. G.; BRONK, D. A.; MULHOLLAND, M. R.; CARPENTER, E. J. Nitrogen in the marine environment. Burlington: Academy Press, 2008, p. 199-261.; LORENZONI et al., 2013LORENZONI, L.; TAYLOR, G. T.; BENITEZ-NELSON, C.; HANSELL, D. A.; MONTES, E.; MASSERINI, R.; FANNING, K.; VARELA, R.; ASTOR, Y.; GUZMÁN, L.; MULLER-KARGER, F. E. Spatial and seasonal variability of dissolved organic matter in the Cariaco Basin. J. Geophys. Res. Biogeosci., v. 118, p. 1-12, 2013.). Chemical gradients of electron donors and acceptors at the redox interface are ultimately sustained by the anaerobic mineralization of organic material exported from the photic zone. THUNELL et al. (2000)THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000. indicate that the proportion of carbon mineralized in the oxic layer increases in parallel with increasing primary and that decreased degradation in the anoxic zone may also arise because more labile fractions of organic carbon have already been mineralized in the upper layers. Despite this, THUNELL et al. (2000)THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000. note that sulfate reduction in the anoxic zone of Cariaco Basin can also be efficient as aerobic respiration for the mineralization of organic material. The vertical distribution of inorganic carbon (Figure 6) is inverse to that observed for organic carbon, with three minima of 19%, 16% and 29% located at the surface, 55 m and 300 m, respectively, and two maxima of 56% and 52% located from 15 to 25 m and at 400 m. An important feature is that from 300 m to the seafloor (1310 m), the variations in the regeneration of organic and inorganic carbon are close to 50%, ranging from 48 to 55% for organic carbon and inorganic carbon. BROECKER and PENG (1982)BROECKER, W.; PENG, T. H. Tracers in the sea. New York: Eldigio Press, 1982. p. 690. observed similar results, and indicated that for deep waters the proportion of both fractions can be 50%. RÍOS et al. (1995)RÍOS, A. F.; ANDERSON, T. R.; PÉREZ, F. F. The carbonic system distribution and fluxes in the NE Atlantic during spring 1991. Progr. Oceanogr., v. 35, n. 4, p. 295-314, 1995. report that in the NE Atlantic the fraction regenerated from hard calcareous skeletons increases to 51% at depths >2300 m due to the increase in TCO2 and decrease in temperature. In the Cariaco Basin, despite the temperature at the surface's reaching 28 ºC and decreasing gradually from 17 ºC at 250 m to the seafloor (Figure 2), no increases in inorganic carbon are observed.

Primary productivity in the Cariaco Basin is high throughout the year and shows a seasonal trend with maximum values of 17 mgC m-3 h-1. Productivity is influenced by the upwelling of cold nutrient rich waters beginning in January. Dissolved CO2 increases with depth from 1968 - 2300 µmol kg-1 at the surface to 2451 µmol kg-1 at the sea floor due to the degradation of organic material. An influence of alkalinity was observed on the dissociation and regeneration of carbon dioxide, leading to percentages in the inorganic and organic fractions regenerated in the water column of 18 and 82%, respectively. No seasonal variation in the regeneration of carbon in Cariaco Basin was observed. Consequently, regenerated carbon in the Cariaco Basin was dominated by the soft tissues of planktonic organisms. The regeneration of organic carbon below the oxic-anoxic interface partly originates from chemotrophic respiration pathways.

ACKNOWLEDGEMENTS

This study was financially supported by the Cariaco Project (Carbon Retention in a Colored Ocean; Nº 2000001702), Fonacit, the Research Council of the University of Oriente, Venezuela (Project CI-5-1801-1079/02) and the Marine Research Station (EDIMAR-FLASA), Punta de Piedras, Nueva Esparta. We wish, further, to thank the captain and crew of R/V Hermano Gines (EDIMAR-FLASA, Venezuela) for their assistance during the fieldwork. This study was part of the Doctoral Thesis of Aristide Márquez of the Oceanographic Institute of Venezuela.

REFERENCES

  • ABRANTES, F.; CERMENO, P.; LOPES, C.; ROMERO, O.; MATOS, L.; VAN IPEREN, J.; RUFINO, M.; MAGALHÃES, V. Diatoms Si uptake capacity drives carbon export in coastal upwelling systems. Biogeosciences, v. 13, p. 4099-4109, 2016.
  • ALVERA-AZCÁRATE, A.; BARTH, A.; WEISBERG, R. A nested model of the Cariaco Basin (Venezuela): description of the basin's interior hydrography and interactions with the open ocean. Ocean. Dynam. v. 59, n.1, p. 97-120, 2009.
  • ASTOR, I. M.; MERI, J. A.; MULLER-KARGER, F. Variabilidad estacional hidrográfica en la fosa de Cariaco. Mem. Soc. Cien. Nat. La Salle, v. 68, n. 149, p. 61-72, 1998.
  • ASTOR, Y. M.; LORENZONI, L.; THUNELL, R.; VARELA, R.; MULLER-KARGER, F.; TROCCOLI, L.; TAYLOR, G. T.; SCRANTON, M. I.; TAPPA, E; RUEDA-ROA, D. Interannual variability in sea surface temperature and fCO2 changes in the Cariaco Basin. Deep Sea Res. Part II Top. Stud. Oceanogr., v. 93, p. 33-43, 2013.
  • ASTOR, Y. M.; SCRANTON, M. I.; MULLER-KARGER, F.; BOHRERD, R.; GARCIA, J. fCO2 variability at the CARIACO tropical coastal upwelling time series station. Mar. Chem., v. 97, n. 3-4, p. 245-261, 2005.
  • ASTOR, Y.; MULLER-KARGER, F.; SCRANTON, M. I. Seasonal and interannual variation in the hydrographic of the Cariaco Basin: implications for basin ventilation. Cont. Shelf Res., v. 23, n. 1, p.125-144, 2003.
  • BAUER, J. E.; WEI-JUN, C.; RAYMOND, P. A.; BIANCHI, T. S.; HOPKINSON, C. S.; REGNIE, P. A. G. The changing carbon cycle of the coastal ocean. Nature, v. 504, n. 7478, p. 61-70, 2013.
  • BERGER, W. H.; FISCHER, K.; LAI, C.; WU, G. Ocean carbon flux: Global maps of primary production and export production. In: AGEGIAN, C. (Ed.). Biogeochemical cycling and fluxes between the deep euphotic zone and other oceanic realms Rockville: U.S. Dept. of Commerce, National Oceanic and Atmospheric Administration, Oceanic and Atmospheric Research, Office of Undersea Research, 1988. p. 131-176.
  • BLAIR, N. E.; ALLER, R. C. The fate of terrestrial organic carbon in the marine environment. Ann. Rev. Mar. Sci., v. 4, p. 401-423, 2012.
  • BRELAND, J. A.; BYRNE, R. H. Spectrophotometric procedures for determination of sea water alkalinity using bromocresol green. Deep Sea Res. Part I: Oceanogr. Res. Pap, v. 40, n. 3, p. 629-641, 1993.
  • BROECKER, W.; PENG, T. H. Tracers in the sea New York: Eldigio Press, 1982. p. 690.
  • CAPONE, D. G.; HUTCHINS, D. A. Microbial biogeochemistry of coastal upwelling regimes in a changing ocean. Nat. Geosci., v. 6, p. 711-717, 2013.
  • CLAYTON, T. D.; BYRNE, R. H. Spectrophotometric seawater pH measurements: total hydrogen ion concentration scale calibration of m-cresol purple and at-sea results. Deep Sea Res. Part I: Oceanogr. Res. Pap., v. 40, n. 10, p. 2115-2129, 1993.
  • COBO-VIVEROS, A. M.; PADIN, X. A.; OTERO, P.; DE LA PAZ, M.; RUIZ-VILLARREAL, M.; RÍOS, A. F.; PÉREZ, F. F. Short-term variability of surface carbon dioxide and sea-air CO2 fluxes in the shelf waters of the Galician coastal upwelling system. Sci. Mar., v. 77, n. 1, p. 37-48, 2013.
  • DICKSON, A. G.; GOYET, C. (Eds). DOE. Handbook of methods for the analysis of the various parameters of the carbon dioxide system in sea water, v. 2. ORNL/CDIAC-74 Washington: U. S. Department of Energy, 1994. 187 p.
  • EMERSON, S. Organic carbon preservation in marine sediments. In: SUNQUIST, E. T.; BROECKER, W. S. (Ed.). The carbon cycle and atmospheric CO2: Natural variations archean to present. New York: American Geophysical Union, 1985. p. 78-87.
  • FEBRES-ORTEGA, G.; HERRERA, L. E. Interpretacion dinamica de algunas de las caracteristicas oceanograficas de la Fosa de Cariaco, Mar Caribe. Boletín del Instituto Oceanográfico de Venezuela, v. 14, n.1, p. 3-29, 1975.
  • FEBRES ORTEGA, G. Circulación de las aguas superiores de la Fosa de Cariaco en Abril de 1974. Bol. Inst. Oceanogr. Univ. Oriente, v. 13, n. 1-2, p. 79-86, 1974.
  • GATTUSO, J. P.; MAGNAN, A.; BILLÉ, R.; CHEUNG, W.; HOWES, E. L.; JOOS, F.; ALLEMAND, D.; BOPP, L.; COOLEY, S. R.; EAKIN, C. M.; HOEGH-GULDBERG, O.; KELLY, R. P.; PÖRTNER, H. O.; ROGERS, A.; BAXTER, J. M.; LAFFOLEY, D.; OSBORN, D.; RANKOVIC, A.; ROCHETTE, J.; SUMAILA, U. R.; TREYER, S.; TURLE, C. Contrasting futures for ocean and society from different anthropogenic CO2 emissions scenarios. Science, v. 349, n. 6243, p. aac4722, 2015.
  • HANSELL, D. A.; CARLSON, C. A.; REPETA, D. J.; SCHLITZER, R. Dissolved organic matter in the ocean: A controversy stimulates new insights. Oceanography, v. 22, n. 4, p. 202-211, 2009.
  • HERRERA, L. E. FEBRES-ORTEGA, G. Kinematics of the windgenerated velocity field in the surface waters off eastern Venezuela, Caribbean Sea. Bol. Inst. Oceanogr. Univ. Oriente, 1975, vol. 14, no 2, p. 165-186.
  • HUTCHINS DA, F. FU, Y. ZHANG, M.E. WARNER, Y. FENG, K. PORTUNE, P.W. BERNHARDT. & M.R. MULHOLLAND. CO2 control of Trichodesmium N2 fixation, photosynthesis, growth rates, and elemental ratios: Implications for past, present, and future ocean biogeochemistry. Limnol Oceanogr. v. 52 n. 4 p. 1293-1304, 2007.
  • IANSON, D.; ALLENA, S. E.; HARRISA, S. L.; ORIANSA, K. J.; VARELA, D. E.; WONG, C. S. The inorganic carbon system in the coastal upwelling region west of Vancouver Island, Canada. Deep Sea Res. Part I: Oceanogr. Res. Pap. v. 50, n. 8, p. 1023-1042, 2003.
  • INDREBØ, G.; PENGERUD, B.; DUNSA, I. Microbial activities in a permanently stratified estuary. I. Primary production and sulfate reduction. Mar. Biol., v. 51, n. 4, p. 295-304, 1979.
  • JOHN, D.; WANG, Z. A.; LIU, X.; BYRNE, R. H.; CORREDOR, J.; LÓPEZ, J. M.; CABRERA, A.; BRONK, D. A.; TABITA, F. R.; PAUL, J. H. Phytoplankton carbon fixation gene (RuBisCO) transcripts and air-sea CO2 flux in the Mississippi River plume. ISME J., v. 1, n. 6, p. 517-531, 2007.
  • JOHNSON, R. A.; WICHERN, D. W. Applied Multivariate Statistical Analysis 3rd ed. New York: Prentice Hall, 1992. 156 p.
  • JØRGENSEN, B. B.; FOSSING, H.; WIRSEN, C. O.; JANNASCH, H. W. Sulfide oxidation in the anoxic Black Sea chemocline. Deep Sea Res. Part I: Oceanogr. Res. Pap. v. 38, Suppl. 2, p. S1083-S1103, 1991.
  • KEELLER, K.; SLATER, R.; BENDER, M.; KEY, R. Possible biological or physical explanations for decadal scale trends in North Pacific nutrient concentration and oxygen utilization. Deep. Sea. Res. II, v. 49, p. 325-362. 2002.
  • KHATIWALA, S.; TANHUA, T.; MIKALOFF FLETCHER, S.; GERBER, M.; DONEY, S. C.; GRAVEN, H. D.; GRUBER, N.; MCKINLEY, G. A.; MURATA, A.; RÍOS, A. F.; SABINE, C. L. Global ocean storage of anthropogenic carbon. Biogeosciences, v. 10, p. 2169-2179, 2013.
  • KNAP, A. H.; MICHAELS, A. F.; STEINBERG, D.; BAHR, F.; BATES, N.; BELL, S.; COUNTWAY, P.; CLOSE, A.; DOYLE, A.; DOW, R.L.; HOWSE, F.; GUNDERSEN, K.; JOHNSON, R.; KELLY, R.; LITTLE, R.; ORCUTT, K.; PARSONS, R.; RATHBUN, C.; SANDERSON, M.; STONE, S. BATS methods manual Vol. 4. Woods Hole: U.S. JGOFS Planning Office, Woods Hole, 1997. 136 p.
  • LEWIS, E.; WALLACE, D. Program developed for CO2 system calculations 1995. Available:<http://cdiac.ornl.gov/oceans/co2rprt.html>. Accessed: 2016/12/12.
    » http://cdiac.ornl.gov/oceans/co2rprt.html
  • LIU, K. K.; ATKINSON, L.; CHEN, C. T. A.; GAO, S.; HALL, J.; MACDONALD, R. W.; MCMANUS, L. T.; QUINONES, R. Exploring continental margin carbon fluxes in the global context. Eos, Trans. Amer. Geophys. Union., v. 81, n. 52, p. 641- 644, 2000.
  • LORENZONI, L.; TAYLOR, G. T.; BENITEZ-NELSON, C.; HANSELL, D. A.; MONTES, E.; MASSERINI, R.; FANNING, K.; VARELA, R.; ASTOR, Y.; GUZMÁN, L.; MULLER-KARGER, F. E. Spatial and seasonal variability of dissolved organic matter in the Cariaco Basin. J. Geophys. Res. Biogeosci., v. 118, p. 1-12, 2013.
  • LOUANCHI, F.; NAJJAR, R. G. A global monthly climatology of phosphate, nitrate, and silicate in the upper ocean: Spring-summer export production and shallow remineralization. Global Biogeochem. Cycles, v. 14, n. 3, p. 957-977, 2000.
  • MADRID, VANESSA, M., et al. Phylogenetic diversity of bacterial and archaeal communities in the anoxic zone of the Cariaco Basin. Applied and environmental microbiology, v. 67, n. 4, p. 1663-1674, 2001.
  • MARGOLIN, A.; GERRINGA, L. J. A.; HANSELL, D. A.; RIJKENBERG, M. J. A. Net removal of dissolved organic carbon in the anoxic waters of the Black Sea. Mar. Chem., v. 183, p. 13-24, 2016.
  • MARÍN, B.; GÓMEZ, A.; UROSA, L. Taxonomía y distribución del Ictioplancton en la fosa de Cariaco, Venezuela. Bol. Inst. Oceanogr. Venezuela, v. 28, n. 1-2, p. 15-27. 1989.
  • MCKINLEY, G. A.; PILCHER, D. J.; FAY, A. R.; LINDSAY, K.; LONG, M. C.; LOVENDUSKI, N. S. Timescales for detection of trends in the ocean carbon sink. Nature, v. 530, p. 469-472, 2016.
  • MCPARLAND, E., BENITEZ-NELSON, C., TAYLOR, G., THUNELL. G., ROLLINS, A., LORENZONI, L. Cycling of suspended particulate phosphorus in the redoxcline of the Cariaco Basin. Mar. Chem., 2015, v. 176, p. 64-74, 2013
  • MEHRBACH, C.; CURBERSON, C.; HAWLEY, J.; PYTKOWICZ, T. Measurement of the apparent dissociation constants of carbonic acid in seawater at atmospheric pressure. Limnol. Oceanogr, v. 18, p. 897-907, 1973.
  • MILLERO, F. Thermodynamics of the carbon dioxide system in the oceans. Geochim Cosmochim Acta, v. 59, n. 4, p. 661-677, 1995.
  • MULLER-KARGER, F. E; APARICIO, R. Mesoscale Processes Affecting Phytoplankton Abundance in the Southern Caribbean Sea. Cont. Shelf Res., v. 14, n. 2-3, p. 199-221, 1994.
  • MULLER-KARGER, F.; VALERA, R.; THUNNEL, R.; SCRANTON, M.; BOHRER, R.; TAYLOR, G.; CAPELO, J.; ASTOR, I.; TAPPA, E.; HO, T.; IABICHELLA, M.; WALSH, J. Annual cycle of primary production in the Cariaco Basin: Response to upwelling and implications for vertical export. J. Geophys. Res., v. 106, n. 3, 4527-4542, 2001.
  • OCCC. 2004. Ocean carbon climate change. Scott. C. Doney editor. USA.108 pp.
  • PACE, M.; KNAUER, G. A.; KARL, D. M.; MARTIN, J. H. Primary production, new production and vertical flux in the eastern Pacific Ocean. Nature, v. 325, p. 803-804, 1987.
  • PACKARD, T. T.; OSMA, N.; FERNÁNDEZ-URRUZOLA, I.; CODISPOTI, L. A.; CHRISTENSEN, J. P.; GÓMEZ, M. Peruvian upwelling plankton respiration: calculations of carbon flux, nutrient retention efficiency, and heterotrophic energy production. Biogeosciences, v. 12, n. 9, p. 2641-2654, 2015.
  • REDFIELD, A.C., KETCHUM, B.H., RICHARDS, F.A. The influence of organisms on the composition of seawater. In: Hill, M.N. Ed.., The Sea, v. 2, p. 26-77, 1963.
  • REIMERS, C.; SUESS, E. The partitioning of organic carbon fluxes and sedimentary organic matter decomposition rates in the ocean. Mar. Chem., v. 13, n. 2, p. 141-168, 1983.
  • RICHARDS, F. Some chemical and hydrographic observations along the north coast of South America - I. Cabo Tres Puntas to Curacao, including the Cariaco Trench and the Gulf of Cariaco. Deep Sea Res, 1960, v. 7, n. 3, p. 163-182, 1960.
  • RÍOS, A. F.; ANDERSON, T. R.; PÉREZ, F. F. The carbonic system distribution and fluxes in the NE Atlantic during spring 1991. Progr. Oceanogr., v. 35, n. 4, p. 295-314, 1995.
  • RÍOS, A. F.; FRAGA, F.; PÉREZ, F. F. Estimation of coefficients for the calculation of ''NO'', ''PO'' and ''CO'', starting from the elemental composition of natural phytoplankton. Sci. Mar., v. 53, n. 4, p. 779-784, 1989.
  • ROMERO, O. E.; THUNELL, R. C.; ASTOR, Y.; VARELA, R. Seasonal and interannual dynamics in diatom production in the Cariaco Basin, Venezuela. Deep Sea Res. Part I: Oceanogr. Res. Pap., v. 56, n. 4, p. 571-581, 2009.
  • REUER, M.; BOYLE, K., EDWARD A.; COLE, JULIA, E. A mid-twentieth century reduction in tropical upwelling inferred from coralline trace element proxies. Earth Planet SC Lett, v. 210, n. 3, p. 437-452, 2003.
  • RUEDA-ROA, D.; MULLER-KARGER, F. The southern Caribbean upwelling system: Sea surface temperature, wind forcing and chlorophyll concentration patterns. Deep. Sea. Res.I v. 78, p. 102-114, 2014.
  • SCRANTON, M., et al. Controls on temporal variability of the geochemistry of the deep Cariaco Basin. Deep Sea Res.I: Oceanogr. Res. Papers, 2001, v. 48, n.7, p. 1605-1625, 2002
  • SCRANTON, M. I.; TAYLOR, G. T.; THUNELL, R. C.; BENITEZ-NELSON, C. R.; MULLER-KARGER, F.; FANNING, K.; LORENZONI, L.; MONTES, E.; VARELA, R.; ASTOR, Y. Interannual and subdecadal variability in the nutrient geochemistry of the Cariaco Basin. Oceanography, v. 27, n. 1, p. 148-159, 2014.
  • SIPPO, J. Z.; MAHER, D. T.; TAIT, D. R.; HOLLOWAY, C.; SANTOS, I. R. Are mangroves drivers or buffers of coastal acidification? Insights from alkalinity and dissolved inorganic carbon export estimates across a latitudinal transect. Global Biogeochem. Cycles, v. 30, n. 5, p. 753-766, 2016.
  • SITCH, S.; FRIEDLINGSTEIN, P.; GRUBER, N.; JONES, S. D.; MURRAY-TORTAROLO, G.; AHLSTRÖM, A.; LEVIS, S.; LEVY P. E.; LOMAS M.; POULTER B.; VIOVY N.; ZAEHLE S.; ZENG N.; ARNETH A.; BONAN G.; BOPP L.; CANADELL J. G.; CHEVALLIER F.; CIAIS P.; ELLIS R.; GLOOR M.; PEYLIN P.; PIAO S. L.; LE QUÉRÉ C.; SMITH B.; ZHU Z.; MYNENI R. Recent trends and drivers of regional sources and sinks of carbon dioxide. Biogeosciences, v. 12, n. 3, p. 653-679, 2015.
  • SMITH, S. Phosphorus versus nitrogen limitation in the marine enviroment. Limnol. Oceanogr. v. 29, n.6, p 1149-1160,1994.
  • SMOAK, J. M.; BENÍTEZ-NELSON, C.; MOORE, W. S.; THUNELL, R.; ASTOR, Y.; MULLER-KARGER, F. Radionuclide fluxes and particle scavenging in Cariaco Basin. Cont. Shelf Res., v. 24, n. 13-14, p. 1451-1463, 2004.
  • SOKAL ROBERT, R.; JAMES, ROHLF F. Biometry. The principles and practice of statistics in biological research. 1969.
  • SOROKIN, Y. I. The bacterial population and the process of hydrogen sulphide oxidation in the Black Sea. J. Cons. Perm. Inter. Explor. Mer, v. 34, p. 423-455, 1972.
  • SOROKIN, Y. I.; SOROKIN, P. Y.; AVDEEV, V. A.; SOROKIN, D. Y.; ILCHENKO, S. V. Biomass, production and activity of bacteria in the Black Sea, with special reference to chemosynthesis and the sulfur cycle. Hydrobiology, v. 308, n. 1, p. 61-76, 1995.
  • STEEMAN-NIELSEN, E. The use of radioactive carbon (14C) for measuring organic production in the sea. J. Cons. Perm. Int. Explor. Mer., v. 18, n. 2, p. 117-140, 1952.
  • STEINBERG, D. K.; CARLSON, C. A.; BATES, N. R.; JOHNSON, R. J.; MICHAELS, A. F.; KNAP, A. H. Overview of the US JGOFS Bermuda Atlantic Time series Study (BATS): a decade-scale look at ocean biology and biogeochemistry. Deep Sea Res. Part II: Top. Stud. Oceanogr., v. 48, n. 8-9, p. 1405-1447, 2001.
  • TAYLOR, G. T.; IABICHELLA, M.; HO, T. Y., SCRANTON, M. I.; TUNNEL, R. C.; MULLER-KARGER, F.; VARELA, R. Chemoautrotophy in the redox transition zone of the Cariaco Basin: A significant midwater source of organic carbon production. Limnol. Oceanogr., v. 46, n. 1, p. 148-163, 2001.
  • THUNELL, R. C.; BENITEZ-NELSON, C.; VARELA, R.; ASTOR, Y.; MULLER-KARGER, F. E. Particulate organic carbon fluxes along upwelling-dominated continental margins: Rates and mechanisms. Global Biogeochem. Cycles, v. 21, n. 1, p. 1-12, 2007.
  • THUNELL, R. C.; VARELA, R.; LLANO, M.; COLLISTER, J., MULLER-KARGER, F.; BOHRER, R. Organic carbon flux in an anoxic water column: sediment trap results from the Cariaco Basin. Limnol. Oceanogr., v. 45, n. 2, p. 300-308, 2000.
  • UNESCO/Intergovernmental Oceanographic Commission. Protocols for the Joint Global Ocean flux Study (JGOFS). Core measurements. Manual and Guides, v. 29 Paris: UNESCO, 1994. p. 128-134.
  • WARD, B. B. Nitrification in marine systems. In: CAPONE, D. G.; BRONK, D. A.; MULHOLLAND, M. R.; CARPENTER, E. J. Nitrogen in the marine environment Burlington: Academy Press, 2008, p. 199-261.

Publication Dates

  • Publication in this collection
    Jan-Mar 2017

History

  • Received
    07 Feb 2016
  • Accepted
    26 Oct 2016
Universidade de São Paulo, Instituto Oceanográfico Praça do Oceanográfico, 191 , 05508-120 Cidade Universitária, São Paulo - SP - Brasil, Tel.: (55 11) 3091-6501, Fax: (55 11) 3032-3092 - São Paulo - SP - Brazil
E-mail: io@usp.br