Acessibilidade / Reportar erro

Benthic aerobic respiration and nutrient fluxes in Cananéia-Iguape Estuarine-Lagoon complex along a salinity gradient

Abstract

Estuaries are important features for global biogeochemical understanding, due to their highly coupled interaction between sediment and water. In those places, the land-sea transect sometimes has a gradient of physical and chemical characteristics, influencing the availability of nutrients and consequently primary production. This study aimed to observe the benthic aerobic respiration and sediment-water nutrient fluxes in a freshwater to marine transect in the Cananéia-Iguape Estuarine-Lagoon Complex (CIELC) on the southeastern Brazilian Coast. This area contains the most pristine mangrove at the limit between tropical and subtropical zones, providing an ideal observatory for ecological research programs. Intact sediment cores were incubated in laboratory to determine fluxes of O2, TCO2, and dissolved nutrients. Three different sites with different salinities were sampled during the four seasons of the year. Sediment characteristics of the sites were compared, showing higher organic matter in the lowest saline (LS) site and higher phytopigments in the highest saline (HS) site, as expected. Benthic aerobic respiration, O2 and CO2 fluxes, ranged from −0.4 to −3.2 mmol m−2 h−1 and 0.1 to 1.5 mmol m−2 h−1, respectively, and exhibited statistically significant variations between seasons and in a salinity gradient. Dissolved inorganic nitrogen and silicate, ranging from −228.7 to 544.8 μmol m−2 h−1 and −205.8 to 4,173.5 μmol m−2 h−1, respectively, were generally released from the sediment, whereas phosphate ranged from −25.2 to 29.6 μmol m−2 h−1 with more variation in time and space. The LS site was characterized as a nitrogen sink and a silicate and phosphate releaser, and the HS site was characterized as a nitrogen producer and a phosphorus consumer. However, seasonal and spatial variations were observed, and the interaction between space and time factors was always highly significant, indicating that the metabolic behavior of the benthic compartment depends on both trophic and physicochemical conditions.

Keywords:
Benthic-pelagic coupling; Benthic metabolism; Estuary; Subtropical climate

INTRODUCTION

Coastal and estuarine waters are characterized by exceptionally high rates of primary production (Boynton et al., 1982Boynton, W. R., Kemp, W. M. & Keefe, C. W. 1982. A comparative analysis of nutrients and other factors influencing estuarine phytoplankton production. In: Kennedy, V. S. (ed.), Estuarine comparisons (pp. 69–90). New York: Academic Press.; Cloern et al., 2014Cloern, J., Foster, S. & Kleckner, A. 2014. Phytoplankton primary production in the world’s estuarine-coastal ecosystems. Biogeosciences, 11(9), 2477–2501. DOI: https://doi.org/10.5194/bg-11-2477-2014
https://doi.org/10.5194/bg-11-2477-2014...
). In these shallow systems, nutrient fluxes across the sediment-water interface and benthic respiration are important links between benthic and pelagic environments (Cowan and Boynton, 1996Cowan, J. & Boynton, W. 1996. Sediment-water oxygen and nutrient exchanges along the longitudinal axis of Chesapeake Bay: seasonal patterns, controlling factors and ecological significance. Estuaries, 19(3), 562–580. DOI: https://doi.org/10.2307/1352518
https://doi.org/10.2307/1352518...
). The sediment may support 15–32% of the N and 17–100% of the P required by phytoplankton (Bonaglia et al., 2014Bonaglia, S., Deutsch, B., Bartoli, M., Marchant, H. & Brüchert, V. 2014. Seasonal oxygen, nitrogen and phosphorus benthic cycling along an impacted Baltic Sea estuary: regulation and spatial patterns. Biogeochemistry, 119(1–3), 139–160. DOI: https://doi.org/10.1007/s10533-014-9953-6
https://doi.org/10.1007/s10533-014-9953-...
; Carstensen et al., 2014Carstensen, J., Conley, D., Bonsdorff, E., Gustafsson, B., Hietanen, S., Janas, U., Jilbert, T., Maximov, A., Norkko, A., Norkko, J., Reed, D., Slomp, C., Timmermann, K. & Voss, M. 2014. Hypoxia in the Baltic Sea: Biogeochemical Cycles, Benthic Fauna, and Management. Ambio, 43(1), 26–36. DOI: https://doi.org/10.1007/s13280-013-0474-7
https://doi.org/10.1007/s13280-013-0474-...
; Boynton et al., 2018Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
https://doi.org/10.1007/s12237-017-0275-...
). Nevertheless, estuarine sediments can play a dual role, in which benthic nutrient inputs can prevent an estuarine ecosystem from seasonal nutrient limitation and estuarine sediments can act as a nutrient trap (Hellemann et al., 2017Hellemann, D., Tallberg, P., Bartl, I., Voss, M. & Hietanen, S. 2017. Denitrification in an oligotrophic estuary: a delayed sink for riverine nitrate. Marine Ecology Progress Series, 583, 63–80. DOI: https://doi.org/10.3354/meps12359
https://doi.org/10.3354/meps12359...
) by reducing the excess carbon and nutrients load that could otherwise imbalance the delicate estuarine ecosystem (Zhang et al., 2010Zhang, J., Gilbert, D., Gooday, A., Levin, L., Naqvi, W., Middelburg, J., Scranton, M., Ekau, W., Pena, A., Dewitte, B., Oguz, T., Monteiro, P., Urban, E., Rabalais, N., Ittekkot, V., Kemp, W., Ulloa, O., Elmgren, R., Escobar-Briones, E. & Van Der Plas, A. 2010. Natural and human-induced hypoxia and consequences for coastal areas: synthesis and future development. Biogeosciences, 7, 1443–1467. DOI: https://doi.org/10.5194/bgd-6-11035-2009
https://doi.org/10.5194/bgd-6-11035-2009...
; Evans and Scavia, 2013Evans, M. & Scavia, D. 2013. Exploring estuarine eutrophication sensitivity to nutrient loading. Limnology and Oceanography, 58(2), 569–578. DOI: https://doi.org/10.4319/lo.2013.58.2.0569
https://doi.org/10.4319/lo.2013.58.2.056...
). The supply of organic matter (OM) to the benthic compartment, the macromolecular quality of this OM, the availability of oxygen, the salinity, and the water temperature are considered primary environmental and biological factors that control the direction and magnitude of solute fluxes (Banta et al., 1995Banta, G., Giblin, A., Hobbie, J. & Tucker, J. 1995. Benthic respiration and nitrogen release in Buzzards Bay, Massachusetts. Journal of Marine Research, 53(1), 107–135. DOI: https://doi.org/10.1357/0022240953213287
https://doi.org/10.1357/0022240953213287...
; Hopkinson et al., 2001Hopkinson, C., Giblin, A. & Tucker, J. 2001. Benthic metabolism and nutrient regeneration on the continental shelf of Eastern Massachusetts, USA. Marine Ecology Progress Series, 224, 1–19. DOI: https://doi.org/10.3354/meps224001
https://doi.org/10.3354/meps224001...
; Smith et al., 2012Smith, J., Burford, M., Revill, A., Haese, R. & Fortune, J. 2012. Effect of nutrient loading on biogeochemical processes in tropical tidal creeks. Biogeochemistry, 108(1–3), 359–380. DOI: https://doi.org/10.1007/s10533-011-9605-z
https://doi.org/10.1007/s10533-011-9605-...
). These factors can operate on a variety of temporal and spatial scales (Twilley et al., 1999Twilley, R., Rick, S., Bond, D. & Baker, J. 1999. Benthic nutrient fluxes in selected estuaries in the Gulf of Mexico. In: Bianchi, T. R. & Twilley, R. R. (eds.), Biogeochemistry of Gulf of Mexico estuaries (pp. 163-207). New York: Wiley.) and can exhibit a seasonal pattern of sediment fluxes of nutrients, normally high values in summer and low values in winter (Kemp and Boynton, 1984Kemp, M. & Boynton, W. 1984. Spatial and temporal coupling of nutrient inputs to estuarine primary production: the role of particulate transport and decomposition. Bulletin of Marine Science, 35(3), 522–535.; Cowan et al., 1996Cowan, J. L. W., Pennock, J. & Boynton, W. 1996. Seasonal and interannual patterns of sediment-water nutrient and oxygen fluxes in Mobile Bay, Alabama (USA): regulating factors and ecological significance. Marine Ecology Progress Series, 141, 229–245. DOI: https://doi.org/10.3354/meps141229
https://doi.org/10.3354/meps141229...
). Estimating nutrient fluxes across the sediment-water interface is significant to understand the ecosystem functioning and assess the biogeochemical cycling of nutrients, sedimentary environment, trophic state, and quality of an ecosystem (Zhang et al., 2019Zhang, L., Xiong, L., Zhang, J., Jiang, Z., Zhao, C., Wu, Y., Liu, S. & Huang, X. 2019. The benthic fluxes of nutrients and the potential influences of sediment on the eutrophication in Daya Bay, South China. Marine Pollution Bulletin, 149, 110540. DOI: https://doi.org/10.1016/j.marpolbul.2019.110540
https://doi.org/10.1016/j.marpolbul.2019...
). The estuarine biogeochemistry and ecosystem functioning can be largely controlled by benthic-pelagic coupling by which estuaries remain pristine, productive, and efficient ecosystem service providers (Barbier et al., 2011Barbier, E., Hacker, S., Kennedy, C., Koch, E., Stier, A. & Silliman, B. 2011. The value of estuarine and coastal ecosystem services. Ecological Monographs, 81(2), 169–193. DOI: https://doi.org/10.1890/10-1510.1
https://doi.org/10.1890/10-1510.1...
; Thrush et al., 2013Thrush, S. F., Townsend, M., Hewitt, J. E., Davies, K., Lohrer, A. M., Lundquist, C. & Cartner, K. 2013. The many uses and values of estuarine ecosystems. In: Dymond, J. R. (ed.), Ecosystem services in New Zealand: conditions and trends (pp. 226–237). New Zealand: Manaaki Whenua Press.).

Despite the abundant literature describing benthic nutrient fluxes, most studies have been conducted in temperate estuaries. Boynton et al. (2018Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
https://doi.org/10.1007/s12237-017-0275-...
) specifically mentioned that studies such as this one on benthic biogeochemistry of tropical and subtropical estuaries are rare. Tropical and subtropical estuaries characteristically differ from the temperate ones, owing to their contrasting hydrology, solar radiation, temperature, terrestrial input, turbidity, and type of fringing vegetation (Eyre and Balls, 1999Eyre, B. & Balls, P. 1999. A Comparative Study of Nutrient Behavior along the Salinity Gradient of Tropical and Temperate Estuaries. Estuaries, 22(2), 313–326. DOI: https://doi.org/10.2307/1352987
https://doi.org/10.2307/1352987...
). Higher temperature makes the microbe-driven respiration processes faster (Helder and Vries, 1983Helder, W. & De Vries, R. 1983. Estuarine nitrite maxima and nitrifying bacteria (Ems-Dollard estuary). Netherlands Journal of Sea Research, 17(1), 1–18. DOI: https://doi.org/10.1016/0077-7579(83)90002-9
https://doi.org/10.1016/0077-7579(83)900...
; Eyre and Balls, 1999Eyre, B. & Balls, P. 1999. A Comparative Study of Nutrient Behavior along the Salinity Gradient of Tropical and Temperate Estuaries. Estuaries, 22(2), 313–326. DOI: https://doi.org/10.2307/1352987
https://doi.org/10.2307/1352987...
). The type of vegetation (mangrove versus coniferous vegetation) changes the quality and quantity of organic carbon, and the flux rates and mineralization of tropical and subtropical estuaries significantly vary from those in temperate estuaries (Hopkinson et al., 1999Hopkinson, C., Giblin, A., Tucker, J. & Garritt, R. 1999. Benthic Metabolism and Nutrient Cycling along an Estuarine Salinity Gradient. Estuaries, 22(4), 863–881. DOI: https://doi.org/10.2307/1353067
https://doi.org/10.2307/1353067...
; Humborg et al., 2003Humborg, C., Danielsson, Å., Sjöberg, B. & Green, M. 2003. Nutrient land–sea fluxes in oligothrophic and pristine estuaries of the Gulf of Bothnia, Baltic Sea. Estuarine, Coastal and Shelf Science, 56(3–4), 781–793. DOI: https://doi.org/10.1016/s0272-7714(02)00290-1
https://doi.org/10.1016/s0272-7714(02)00...
). For example, the high amount of OM from terrestrial sources (Pradhan et al., 2014Pradhan, U., Wu, Y., Shirodkar, P., Zhang, J. & Zhang, G. 2014. Sources and distribution of organic matter in thirty five tropical estuaries along the west coast of India-a preliminary assessment. Estuarine, Coastal and Shelf Science, 151, 21–33. DOI: https://doi.org/10.1016/j.ecss.2014.09.010
https://doi.org/10.1016/j.ecss.2014.09.0...
; Sarma et al., 2014Sarma, V., Krishna, M., Prasad, V., Kumar, B., Naidu, S., Rao, G., Viswanadham, R., Sridevi, T., Kumar, P. & Reddy, N. 2014. Distribution and sources of particulate organic matter in the Indian monsoonal estuaries during monsoon. Journal of Geophysical Research: Biogeosciences, 119(11), 2095–2111. DOI: https://doi.org/10.1002/2014jg002721
https://doi.org/10.1002/2014jg002721...
) and the organic-rich sediments below a thin top oxic layer can be strongly reducing (Pratihary et al., 2009Pratihary, A., Naqvi, S., Naik, H., Thorat, B., Narvenkar, G., Manjunatha, B. & Rao, V. 2009. Benthic fluxes in a tropical Estuary and their role in the ecosystem. Estuarine, Coastal and Shelf Science, 85(3), 387–398. DOI: https://doi.org/10.1016/j.ecss.2009.08.012
https://doi.org/10.1016/j.ecss.2009.08.0...
; Gomez-Ramirez et al., 2019Gomez-Ramirez, E. H., Corzo, A., Garcia-Robledo, E., Bohorquez, J., Agüera-Jaquemet, A., Bibbo-Sanchez, F.,Soria-Píriz, S., Jiménez-Arias, J. L., Morales, A. & Papaspyrou, S. 2019. Benthic-pelagic coupling of carbon and nitrogen along a tropical estuarine gradient (Gulf of Nicoya, Costa Rica). Estuarine, Coastal and Shelf Science, 228, 1-11. DOI: https://doi.org/10.1016/j.ecss.2019.106362
https://doi.org/10.1016/j.ecss.2019.1063...
). On the other hand, due to high temperature and consequent high microbial and faunal activity, OM mineralization tends to be faster than temperate estuarine sediments (Helder and Vries, 1983Helder, W. & De Vries, R. 1983. Estuarine nitrite maxima and nitrifying bacteria (Ems-Dollard estuary). Netherlands Journal of Sea Research, 17(1), 1–18. DOI: https://doi.org/10.1016/0077-7579(83)90002-9
https://doi.org/10.1016/0077-7579(83)900...
). Additionally, information is limited about benthic fluxes in varying salinity environments that have strong gradients of environmental parameters (Boynton et al., 2018Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
https://doi.org/10.1007/s12237-017-0275-...
).

Estuaries drain continental river basins into coastal waters, and tidal seawater is diluted by riverine freshwater inputs flowing seaward, creating an environmental gradient along the main channel (Barletta and Dantas, 2016Barletta, M. & Dantas, D. V. 2016. Environmental gradient. In: Kennish, M. J. (ed.), Encyclopedia of estuaries. Dordrecht: Springer.). These coastal systems often display strong gradients of salinity, nitrate concentrations, and sedimentary organic carbon sources and characteristics, far from river mouths, and the consequent strong availability of terminal electron acceptors. The quality of organic matter (OM) reaching the sediment is also likely to vary along the salinity gradient. Structurally complex terrestrial detritus with high C:N dominates at the freshwater end, and phytoplankton detritus with low C:N dominates at the marine end (Hopkinson et al., 1998Hopkinson, C. S., Buffam, I., Hobbie, J., Vallino, J., Perdue, M., Eversmeyer, B., Prahl, F., Covert, J., Hodson, R., Moran, M. A., Smith, E., Baross, J., Crump, B., Findlay, S. & Foreman, K. 1998. Terrestrial inputs of organic matter to coastal ecosystems: An intercomparison of chemical characteristics and bioavailability. Biogeochemistry, 43(3), 211–234.; Fellman et al., 2011Fellman, J., Petrone, K. & Grierson, P. 2011. Source, biogeochemical cycling, and fluorescence characteristics of dissolved organic matter in an agro-urban estuary. Limnology and Oceanography, 56(1), 243–256. DOI: https://doi.org/10.4319/lo.2011.56.1.0243
https://doi.org/10.4319/lo.2011.56.1.024...
). The lack of extensive biogeochemical process studies in tropical and subtropical estuaries (Cloern et al., 2014Cloern, J., Foster, S. & Kleckner, A. 2014. Phytoplankton primary production in the world’s estuarine-coastal ecosystems. Biogeosciences, 11(9), 2477–2501. DOI: https://doi.org/10.5194/bg-11-2477-2014
https://doi.org/10.5194/bg-11-2477-2014...
; Boynton et al., 2018Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
https://doi.org/10.1007/s12237-017-0275-...
) and in a salinity gradient leaves a gap in our understanding of global carbon and nutrient cycling. Better understanding of benthic biogeochemical processes in those estuaries is necessary for appropriate ecosystem management.

Many estuaries in the tropical zone are found in Brazil, which is the country with the second largest mangrove area in the world (Basha, 2018Basha, S. K. 2018. An overview on global mangroves distribution. Indian Journal of Geo Marine Sciences, 47(4), 766–772.; Nabeelah et al., 2019Nabeelah, B. S., Fawzi, M. M., Gokhan, Z., Rajesh, J., Nadeem, N., Kannan, R., Albuquerque, R. D. D. G. & Pandian, A. K. 2019. Ethnopharmacology, Phytochemistry, and Global Distribution of Mangroves―A Comprehensive Review. Marine Drugs, 17(4), 231. DOI: https://doi.org/10.3390/md17040231
https://doi.org/10.3390/md17040231...
). However, despite the importance of these systems for understanding the global patterns of biogeochemical changes, the study of the biogeochemical cycles in these environments is still lacking (Boynton et al., 2018Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
https://doi.org/10.1007/s12237-017-0275-...
). Some studies were found about the benthic metabolism and sediment-water nutrient fluxes in an estuarine system in the Amazon region (Matos et al., 2022Matos, C., Berrêdo, J., Machado, W., Metzger, E., Sanders, C., Faial, K. & Cohen, M. 2022. Seasonal changes in metal and nutrient fluxes across the sediment-water interface in tropical mangrove creeks in the Amazon region. Applied Geochemistry, 138, 105217. DOI: https://doi.org/10.1016/j.apgeochem.2022.105217
https://doi.org/10.1016/j.apgeochem.2022...
) and in coastal lagoons in the states of Rio de Janeiro (Machado and Knoppers, 1988Machado, E. & Knoppers, B. 1988. Sediment oxygen consumption in an organic-rich, subtropical lagoon, Brazil. Science of The Total Environment, 75(2–3), 341–349. DOI: https://doi.org/10.1016/0048-9697(88)90045-9
https://doi.org/10.1016/0048-9697(88)900...
; Knoppers et al., 1996Knoppers, B., Souza, W., Souza, M., Gonzalez Rodriguez, E., Landim, E. & Vieira, A. 1996. In situ measurements of benthic primary production, respiration and nutrient fluxes in a hypersaline coastal lagoon of SE Brazil. Revista Brasileira de Oceanografia, 44(2), 155–165. DOI: https://doi.org/10.1590/s1413-77391996000200005
https://doi.org/10.1590/s1413-7739199600...
, 2004Knoppers, B., Machado, E., Brandini, N. & Landim De Souza, W. F. 2004. Environmental Geochemistry in Tropical and Subtropical Environments Science. In: Duursma, E. K. & Abrão, J. J. (eds.) (pp. 253–275). Berlin, Heidelberg: Springer. DOI: https://doi.org/10.1007/978-3-662-07060-4_19
https://doi.org/10.1007/978-3-662-07060-...
) and Paraná (Niencheski and Jahnke, 2002Niencheski, L. & Jahnke, R. A. 2002. Benthic respiration and inorganic nutrient fluxes in the estuarine region of Patos Lagoon (Brazil). Aquatic Geochemistry, 8(3), 135–152.; Knoppers et al., 2004Knoppers, B., Machado, E., Brandini, N. & Landim De Souza, W. F. 2004. Environmental Geochemistry in Tropical and Subtropical Environments Science. In: Duursma, E. K. & Abrão, J. J. (eds.) (pp. 253–275). Berlin, Heidelberg: Springer. DOI: https://doi.org/10.1007/978-3-662-07060-4_19
https://doi.org/10.1007/978-3-662-07060-...
), but no previous study, to our knowledge, has been carried out at the Cananéia-Iguape Estuarine Complex (CIELC).

The CIELC was included in the Ramsar’s List of Wetlands of International Importance in 2017 (https://rsis.ramsar.org/ris/2310) and in the UNESCO Biosphere Reserve. The CIELC contains the most pristine mangrove at the limit between tropical and subtropical zones in the southwestern Atlantic, where sensitivity to climate change is higher. Thus, it is an ideal observatory for ecological research programs. The Cananéia region has great ecological importance since it contains highly diversified environments, fauna and flora biodiversity, and a significant preservation of the biota (Diegues, 1987Diegues, A. C. 1987. Conservação e desenvolvimento sustentado de ecossistemas litorâneos no Brasil. São Paulo: Secretaria do Meio Ambiente.). On the other hand, the artificial channel, Valo Grande, constructed in the 19th century has been responsible for an important introduction of fresh water into the system, and a salinity gradient decreasing from Cananéia to Iguape region can now be observed (Mahiques et al., 2009Mahiques, M., Burone, L., Figueira, R., Lavenére-Wanderley, A., Capellari, B., Rogacheski, C., Barroso, C., Samaritano Dos Santos, L., Cordero, L. & Cussioli, M. 2009. Anthropogenic influences in a lagoonal environment: a multiproxy approach at the valo grande mouth, Cananéia-Iguape system (SE Brazil). Brazilian Journal of Oceanography, 57(4), 325–337. DOI: https://doi.org/10.1590/s1679-87592009000400007
https://doi.org/10.1590/s1679-8759200900...
). The CIELC is in a transitional area between the tropical and subtropical climates and is characterized as a wet subtropical climate, with a very humid spring/summer and a drier autumn/winter, leading to variable seasonal inflow of freshwater from continental drainage of rivers (Mahiques et al., 2009Mahiques, M., Burone, L., Figueira, R., Lavenére-Wanderley, A., Capellari, B., Rogacheski, C., Barroso, C., Samaritano Dos Santos, L., Cordero, L. & Cussioli, M. 2009. Anthropogenic influences in a lagoonal environment: a multiproxy approach at the valo grande mouth, Cananéia-Iguape system (SE Brazil). Brazilian Journal of Oceanography, 57(4), 325–337. DOI: https://doi.org/10.1590/s1679-87592009000400007
https://doi.org/10.1590/s1679-8759200900...
).

Due to the great ecological importance of the region and the lack of studies in tropical and subtropical areas, this study aimed to analyze the benthic oxygen consumption and the nutrients flow through the sediment-water interface in a salinity gradient in the CIELC region. For that, intact sediment cores were incubated ex situ at three different salinities, in four different periods from August 2021 to June 2022.

METHODS

Study Area

The study was carried out at the Cananéia-Iguape Estuarine-Lagoon Complex (CIELC) in the state of São Paulo on the southeast coast of Brazil (Figure 1). The system is composed of an extended coastal plain, with an intricate set of channels and interconnected coastal lagoons, in the municipalities of Iguape (northern sector), Ilha Comprida, and Cananéia (southern sector), as well as Cardoso Island. The islands are isolated from the continent by interconnected channels, the Pequeno, Cubatão, and Cananéia Seas and the Trapandé Bay, and are connected to the Atlantic Ocean through channels at the northern (Barra de Icapara) and southern (Barra de Cananéia) end of Comprida Island (Tessler and Souza, 1998Tessler, M. & Souza, L. 1998. Dinâmica sedimentar e feições sedimentares identificadas na superfície de fundo do sistema Cananéia-Iguape, SP. Revista Brasileira de Oceanografia, 46(1), 69–83. DOI: https://doi.org/10.1590/s1413-77391998000100006
https://doi.org/10.1590/s1413-7739199800...
). In the northeastern portion of the system, the Pequeno Sea channel is strongly influenced by freshwater from the artificial channel of Valo Grande (located in the municipality of Iguape), which connects it to the Ribeira de Iguape River (Italiani and Mahiques, 2014Italiani, D. & Mahiques, M. 2014. O registro geológico da atividade antropogênica na região do Valo Grande, Estado de São Paulo, Brasil. Quaternary and Environmental Geosciences, 5(2), 33–44. DOI: https://doi.org/10.5380/abequa.v5i2.34522
https://doi.org/10.5380/abequa.v5i2.3452...
). Tidal cycles associated with freshwater inflows and atmospheric precipitation cause wide variations of salinity along the CIELC. The Ribeira de Iguape River is the main freshwater source in the complex, and its flow rate responds synchronously to rainfall variations (lowest in June with an average of 218 m−3 s−1, which triples during the rainy season to an average of 634 m−3 s−1) significantly influencing the salinity of the lagoon waters (Departamento de Águas e Energia Elétrica do Estado de São Paulo, 2019Departamento de Águas e Energia Elétrica do Estado de São Paulo. 2019. Banco de dados fluviométricos do Estado de São Paulo. Banco de dados hidrológicos. Accessed: http://www.hidrologia.daee.sp.gov.br/
http://www.hidrologia.daee.sp.gov.br/...
). The southern region has eutrophic conditions to support a primary production and the richness of a preserved area influenced by the marine hydrodynamic and micro-tide regimes (Azevedo and Braga, 2011Azevedo, J.S. & Braga, E.S. 2011. Caracterização hidroquímica para qualificação ambiental dos estuários de Santos-São Vicente e Cananéia. Arquivos de Ciências do Mar, 44(2), 52-61.; Pecoraro et al., 2019Pecoraro, G., Hortellani, M., Hagiwara, Y., Braga, E., Sarkis, J. & Azevedo, J. 2019. Bioaccumulation of Total Mercury (THg) in Catfish (Siluriformes, Ariidae) with Different Sexual Maturity from Cananéia-Iguape Estuary, SP, Brazil. Bulletin of Environmental Contamination and Toxicology, 102(2), 175–179. DOI: https://doi.org/10.1007/s00128-018-2485-3
https://doi.org/10.1007/s00128-018-2485-...
). The general pattern of water circulation depends on tides, which are semidiurnal with a mean of 0.8 m and maximum amplitude of 1.25 m (Bernardes and Miranda, 2001Bernardes, M. & Miranda, L. 2001. Circulação estacionária e estratificação de sal em canais estuarinos: simulação com modelos analíticos. Revista Brasileira de Oceanografia, 49(1–2), 115–132. DOI: https://doi.org/10.1590/s1413-77392001000100010
https://doi.org/10.1590/s1413-7739200100...
). The nutrient dynamics of the lagoon region are influenced by inorganic nutrient runoff from the Tropical Atlantic Forest, dissolved and particulate organic matter input, phytoplankton biomass, and decomposition in tidal creeks of the upper estuary (Schaeffer-Novelli et al., 1990Schaeffer-Novelli, Y., De Souza Lima Mesquita, H., Cintrón-Molero, G. & Cintron-Molero, G. 1990. The Cananéia Lagoon Estuarine System, São Paulo, Brazil. Estuaries, 13(2), 193–203. DOI: https://doi.org/10.2307/1351589
https://doi.org/10.2307/1351589...
). The estuarine channels are surrounded by exuberant mangrove vegetation, and mangrove litter significantly contributes to OM production, where the highest values of chlorophyll and organic carbon are observed (Garcia et al., 2018Garcia, J., Lopes, A., Silvestre, A., Grabowski, R., Barioto, J., Costa, R. & Castilho, A. 2018. Environmental characterization of the Cananéia coastal area and its associated estuarine system (São Paulo state, Brazil): Considerations for three Penaeoidean shrimp species. Regional Studies in Marine Science, 19, 9–16. DOI: https://doi.org/10.1016/j.rsma.2018.02.010
https://doi.org/10.1016/j.rsma.2018.02.0...
).

Figure 1.
Location of the Cananéia-Iguape Estuarine-Lagoon Complex (CIELC) showing High Salinity (1), Intermediate Salinity (2) and Low Salinity (3) sampling site (star).

Sampling and experimental setup

Water and sediment were collected four times from August 2021 to June 2022 in three stations along a salinity gradient in the CIELC: high salinity station (Trapandé, S25°03.988 W47°92.241), intermediate salinity station (Pedrinhas, S25°88.594 W47°79.351), and low salinity station (Valo Grande, S24°16.050, W47°56.273) (Figure 1). At each station, sediments were sampled with minimum disturbance by scuba divers using transparent plexiglass liners (internal diameter = 7 cm, length = 35 cm), to have approximately equal heights (15–17 cm) of sediment and water column. Eight intact sediment cores were taken at each station, and the bottom water was characterized for temperature and salinity. Also, ~100 L of water were collected from each site for maintenance, preincubation, andincubation of the cores. Additionally, two cores were sliced in five sediment layers (0–2, 2–4, 4–6, 6–8, and 8–10 cm depth), homogenized, placed in dark vials, and frozen (−18 °C) for sediment characterization (density, porosity, organic matter content, and phytopigments).

The cores were stored vertically in the dark and immediately transferred to the laboratory. Different tanks were used for the different stations. During overnight preincubation, the cores were maintained submersed with the top open and under stirring, and the water in the tank was maintained at in situ temperature and at 100% oxygen saturation. The day after sampling, dark incubations began by sealing each core with gas-tight lids. A Teflon-coated magnetic bar driven by an external motor at 40 rpm was put on each core, which were suspended 10 cm above the sediment-water interface. Solute concentrations were measured at the beginning and at the end of the incubation, assuming that their uptake or release rate during the incubation was linear. The incubations were set for 4–5 hours for flux measurements of dissolved gases (total oxygen and carbonic gas) and inorganic nutrients (ammonium, nitrite, nitrate, phosphate, and silicate). Water samples were collected at the beginning and at the end of the incubations using plastic syringes from each core water phase. The dissolved gas in the water samples was analyzed immediately, and samples for nutrients were filtered through Whatman GF/F glass fiber filters into 20 mL plastic vials and frozen (−20 °C) until analysis.

Analytical Procedures

Sediment density and porosity was determined by the weight of a 3 mL sediment aliquot before and after being maintained at 60 °C for 48 hours. Density was calculated by weight of dry sediment per milliliter of sediment (g/mL), and porosity was calculated by water content and density. The OM was measured by combustion in a muffle furnace at 400 °C for 4 hours and expressed in percentage. The concentrations of chlorophyll-a and phaeopigments were determined by the spectrophotometric method by extraction with 90% acetone and measured using a Thermo® Evolution 200 spectrophotometer, following Plante-Cuny (1978Plante-Cuny, M. R. 1978. Pigments photosynthétiques et production primaire des fonds meubles néritiques d’une région tropicale (Nosy-Bé, Madagascar) (Travaux et documents de l’O.R.S.T.O.M. ; no 96). Provence: O.R.S.T.O.M.) recommendations and Lorenzen (1967Lorenzen, C. 1967. Determination of chlorophyll and pheo-pigments: spectrophotometric equations. Limnology and Oceanography, 12(2), 343–346. DOI: https://doi.org/10.4319/lo.1967.12.2.0343
https://doi.org/10.4319/lo.1967.12.2.034...
) equations. The pigment content was calculated in micrograms per gram of dry sediment (mg g−1).

Dissolved O2 was measured using the Winkler method. The TCO2 was calculated by measuring total alkalinity with HCl (0.1 N) titration of the samples by using a Metrohm® Titrando, following the recommendations of Dickson et al. (2007Dickson, A. G., Sabine, C. L. & Christian, J. R. (eds.). 2007. Guide to Best Practices for Ocean CO2 Measurements. Asheville: NCEI.). These analyses were performed between 8 and 12 hours after collection. Dissolved inorganic nutrient concentrations were measured using colorimetric methods. Nitrite and nitrate were measured using a continuous flow analyzer Seal® AutoAnalyzer II, as described by Tréguer and Le Corre (1975Tréguer, P. & Le Corre, P. 1975. Manuel d’analyse des sels nutritifs dans l’eau de mer (utilisation de l’autoanalyzer II Technicon R). Brest: Université de Bretagne Occidentale.) with a 0.02 mmol L−1 precision. The N-ammonium concentration was also determined following the method of Tréguer and Le Corre (1975Tréguer, P. & Le Corre, P. 1975. Manuel d’analyse des sels nutritifs dans l’eau de mer (utilisation de l’autoanalyzer II Technicon R). Brest: Université de Bretagne Occidentale.), using a Thermo® Evolution 200 spectrophotometer with a ± 0.02 μmol L−1 detection limit and a ± 0.01 μmol L−1 precision. Dissolved phosphate and silicate concentrations were determined by using the recommendation of Grasshoff et al. (1983Grasshoff, K., Kremling, K. & Ehrhardt, M. (eds.). 1983. Methods of Seawater Analysis (2nd ed.). Berlin: Wiley.) based on the molybdenum blue complex with a detection limit of 0.01 for phosphate and 0.02 μmol L−1 for silicate and a precision of ±0.01 and ±0.1 μmol L−1, respectively. Fluxes of dissolved gases and nutrients were calculated from the change in concentrations in the cores with time and expressed based on area (μmol m−2 h−1 or mmol m−2 h−1) according to equation of Dalsgaard et al. (2000):

\[ \begin{equation}\mathrm{F=}\frac{\left(\mathrm{Cf-Ci}\right)\mathrm{.\ V} }{\mathrm{A.T}}\end{equation} \]

where F is the flux of measured solutes (mmol/μmol m−2 h−1), Ci is the concentration at time zero (mmol/μmol L−1), Cf is concentration at the end of incubation (mmol/μmol L−1), V is the volume of water in the core (l), A is area of sediment surface in the core (m2), and T is the incubation time (h). Fluxes directed from the sediment to the water column were considered as positive.

Statistical Analysis

Differences between fluxes of dissolved gas and nutrients were analyzed by two-way analysis of variance (ANOVA), with time (T, n = 4; winter, spring, summer, and autumn) and salinity (S, n = 3; high salinity – HS; intermediated salinity – IS; and low salinity – LS) as factors. The normality and homogeneity of variance were checked using the Shapiro-Wilk test and the Levene median test, respectively. In the case of heteroscedasticity, the data were 1/x, ln(x), log(x2), or log(x) transformed. Pearson product-moment correlation matrix was performed to determine relationships between variables (significance level p < 0.05). Principal component analysis (PCA) was performed separately for environmental variables and fluxes. Statistics and graphics were done in Sigma Plot 11.0 and Past 4.03.

RESULTS

General Features And Sediment Characterization

The water and sediment samples were collected on the same day for all the stations, keeping the location and time the same throughout the seasons. (Trapandé around 8:00 AM, Pedrinhas around 10:00 AM., and Valo Grande around 12:00). With this, we tried to observe how the change in tidal dynamics and climatic conditions, such as rainfall and temperature, are influencing the benthic population at this specific place. Water temperature over the survey period followed a typical seasonal pattern, ranging from 26 °C in summer/spring to 20 °C in winter/autumn (Table 1). The salinity gradient (from 0 to 30) was typical and reflected the pattern observed in the CIELC, with lower values in the rainy seasons (summer/spring, Table 1).

Table 1.
Physic-chemical information of stations at the sampling times.

Sediments demonstrated higher porosity and total organic matter (TOM) content and lower density in stations with LS, whereas lower porosity and lower density was observed in IS station at all sampled times (Table 2). Chlorophyll-a and phaeopigments were more variable between sites and showed a temporal variation, with the highest value at the IS station in winter and the lowest at the same station in autumn (Table 2). The HS station showed higher values of phaeopigments at almost all times, except in autumn, when the LS station showed the highest values (Table 2).

Table 2.
Sedimentary features (average and standard error) measured in the upper 10 cm layer of stations in the sampling seasons. HS: high salinity site; IS: intermediate salinity site; LS: low salinity site.

Pearson correlation analysis showed a positive correlation between TOM and porosity (r = 0.780, p = 0.002) and a negative correlation with sediment density (r = −0.731, p = 0.006). Negative correlations were observed between sediment density and porosity (r = −0.889, p = 0.000). Principal Component analysis (PCA) was used to observe the general distribution of the values due to the environmental parameters. The PCA explained 64.5% of the total sample variance in the two first axes and clearly separated it into two distinct groups (Figure 2). The first group (A) comprised all samples at the HS site and the second group (B) included all the LS site. The IS site showed more variable parameters (Figure 2).

Figure 2.
Principal Component Analysis (PCA) was performed with the environmental parameters in the stations and seasons. The two first axes (PC1 and PC2) explained 64.5% of the variation. Two groups A (high salinity stations) and B (low salinity stations) were observed. W: winter; Sp: spring; Su: Summer; A: autumn; H: high salinity; I: intermediate salinity; L: low salinity; DEP: Station depth; DEN: Sediment density; TOM: Total organic matter; CHLO: Chlorophyll-a; Phaeo: phaeopigments; Phyt: Phytopigments (chlorophyll-a + phaeopigments); TEMP: temperature; SAL: Salinity.

Benthic aerobic respiration and nutrient fluxes

Benthic aerobic respiration (oxygen consumption and CO2 efflux) displayed significant variations between seasons and in the salinity gradient (Table 3, Figure 3). Oxygen consumption ranged from 0.4 to 3.2 mmol m−2 h−1 and showed a significant interaction between the factors of salinity and seasonality (Figure 2, Table 3). The IS showed significantly higher values in winter and autumn, whereas oxygen consumption in summer and spring was significant higher in HS station. (Figure 3, Table 3). Generally, oxygen consumption was significantly higher in spring (1.6 ± 0.1 mmol m−2 h−1) and lower in autumn (0.90 ± 0.08 mmol m−2 h−1); the HS site had the highest value (1.30 ± 0.08 mmol m−2 h−1) and LS site the lowest (0.80 ± 0.08 mmol m−2 h−1) (Table 3). The TCO2 liberation ranged from 0.1 to 1.5 mmol m−2 h−1, showing a similar pattern to the oxygen consumption, and a significant interaction between time and spatial factors was observed (Figure 3, Table 3). The marine site showed the greatest seasonal variation with significant higher CO2 efflux in summer and lower liberation in autumn (Figure 3, Table 3). Average TCO2 efflux was significantly higher in the HS site (0.60 ± 0.08 mmol m−2 h−1) and in summer (0.70 ± 0.06 mmol m−2 h−1) than the other seasons and stations (Figure 3, Table 3).

Table 3.
Results of the two-way ANOVA on the effects of the factors sampling season, sampling stations and of their interaction on benthic fluxes. Significant differences in the two-way ANOVA and Tukey HSD tests were set at P < 0.05. Dashed border means similar concentrations. Dissolved Inorganic Nitrogen (DIN)=(NO2- + NO3- + NH4+), H: high salinity site; I: intermediate salinity site; L: low salinity site; Su: Summer; A: Autumn; Sp: spring; W: winter.

Figure 3.
Dark O2 and TCO2 fluxes (mmol m-2 h-2) in the sampling seasons at three salinity stations. Average (n=8) and standard deviation are reported. HS. High salinity, IS. Intermediate salinity, LS. Low salinity

Dissolved N fluxes were high and had a very variated pattern (Figure 4). Nitrite (NO2) and nitrate (NO3) ranged from −4.7 to 17.8 and −251.3 to 585.3 mmol N m−2 h−1, respectively, and showed similar uptake or efflux patterns, with significant variation in space, time, and in the interaction between time and space (Figure 4, Table 3). Both displayed sediment release to the water column in the temporal balance at all salinity values, with significantly higher average efflux at the IS site (8.5 ± 0.8 and 127.5 ± 24 mmol N m−2 h−1 for NO2 and NO3, respectively) and lower efflux at the HS (1.4 ± 0.8 and 1.1 ± 22 mmol N m−2 h−1 for NO2 and NO3, respectively) (Table 3). Significant consumption of both N forms was observed at the LS site in winter and spring (Figure 4, Table 3). On the other hand, at the IS site, NO2 was always released from the sediment and exhibited significant liberation of both forms in winter and summer (Figure 4, Table 3). The HS site showed very low fluxes, which characterized a balance between efflux and influx of both N forms (Figure 4). Ammonium (NH4+) fluxes ranged from −97.1 to 81.3 mmol N m−2 h−1 and displayed significant differences between sites, seasons, and in the interaction of factors (Table 3). In general, NH4+ flux was directed from the sediment to the water column (Figure 3), with statistically higher annual average efflux at the HS and lower at the IS sites (48.7 ± 4 and 7.5 ± 4 mmol N m−2 h−1, respectively) (Figure 3, Table 3). Higher NH4+consumption by the sediment was significant in summer (average −25.9 ± 5 mmol N m−2 h−1) at the IS and LS sites (Figure 4, Table 3). The HS site showed the lowest variation between seasons, with NH4+ fluxes oriented towards the water column (Figure 4). Total dissolved inorganic nitrogen (DIN = NO2 + NO3 + NH4+), which ranged from −228.7 to 544.8 mmol N m−2 h−1, was essentially driven by high NO3 flux (Figure 4). Significant variation in time, space, and interaction of factors was observed (Table 3). In general, N was released from the sediment to the water column, with higher efflux in summer and at the IS site (327.4 ± 26 and 160.7 ± 22 mmol N m−2 h−1, respectively) (Figure 4, Table 3). The HS site had the lowest flux variation in time, whereas the LS site showed the highest variation, with influx of N in winter and spring, and efflux in summer and autumn (Figure 4).

Silicate fluxes (Si(OH)4) ranged from −205.8 to 4,173.5 mmol m−2 h−1 with a generally large release from the sediment to the water column, showing a significative variation in time and space (Figure 5, Table 3). Significantly higher average efflux was observed at the LS site and in summer (1,460 ± 140 and 2,556 ± 0166 mmol m−2 h−1, respectively) (Figure 5, Table 3). The HS site showed less variation in time, with a small efflux of Si to the water. However, the IS and LS sites had a significant seasonal variation with higher efflux in summer and a sediment Si consumption in autumn (Figure 5, Table 3). Phosphate fluxes were very variable, ranging from −25.2 to 29.6 mmol m−2 h−1. A significant difference was observed in time and space, with significant interaction between factors (Figure 6, Table 3). The mean total flux showed an efflux of P from the sediment to the water column with statistically higher values at the LS site and in summer (11.1 ± 1.5 and 15.8 ± 2 mmol m−2 h−1, respectively) (Figure 6, Table 3). Phosphate uptake by the sediment was observed at station HS during most of the samplings (Figure 6). Likewise, in winter and autumn, both the HS and IS stations had a significant uptake of P by the sediment (Figure 6, Table 3), which showed an inversion in spring, when the HS and IS stations showed a release of P whereas the LS showed uptake of P from the water column (Figure 6, Table 3).

Figure 4.
Dark fluxes of N (mmol m-2 h-2) derivate nutrients (NO2-, NO3-, NH4+ and DIN-sun of N fluxes) in the sampling seasons at three salinity stations. Average (n=8) and standard deviation are reported. HS: High salinity; IS: Intermediate salinity; LS: Low salinity.

Figure 5.
Dark Si(OH)4 flux (μmol m-2 h-1) in the sampling seasons at three salinity stations. Average (n=8) and standard deviation are reported. HS: High salinity; IS: Intermediate salinity; LS: Low salinity.

Figure 6.
Dark PO4-3 flux (μmol m-2 h-1) in the sampling seasons at three salinity stations. Average (n=8) and standard deviation are reported. HS: High salinity; IS: Intermediate salinity; LS: Low salinity.

Pearson correlation test showed a positive correlation between NO2− flux and depth, and a negative correlation between depth and CO2 liberation (Table 4). Fluxes of Si(OH)4 were significantly correlated with almost all other nutrient fluxes. Positive correlations were observed between Si, NO2−, NO3−, and DIN (Table 4), whereas the pairs Si and P and NO3 and P were negatively correlated (Table 4).

Table 4.
Pearson product-moment correlation coefficients showing the significant correlated fluxes. Significant correlation is observed when p<0.05. Dissolved Inorganic Nitrogen (DIN) = (NO2- + NO3- + NH4+).

DISCUSSION

General environmental characterization

This work aimed to verify if there were any significant differences among processes measured along the salinity gradient inside the CIELC. Normally, the salinity of an estuary exhibits a decreasing trend from closer to an ocean to up a river channel; thus, the system can be classified according to the salinity variations. Polyhaline are regions where salinity varies from 30 to 35; mesohaline shows salinity variations around 15; and the oligohaline systems ranges from 0 to 5 (Day et al., 1989Day, J. W., Crump, B. C., Kemp, W. M. & Yáñez-Arancibia, A. (eds.). 1989. Estuarine Ecology. New York: John Wiley & Sons, Inc.). At CIELC, tidal cycles associated with freshwater inflows and atmospheric precipitation cause wide amplitudes of salinity variation throughout the estuarine system. The Ribeira de Iguape River is the main freshwater source that reaches the complex and has the most significant influence on the salinity of the estuarine-lagoon waters. The Valo Grande station is in the Ribeira Iguape River and exhibited salinity values from 0 to 1, characterizing a predominantly freshwater system; thus, it was chosen as the oligohaline site in this study. Trapandé Bay, in the southern part of the CIELC, was chosen as the saline site, ranging from 20 to 30 of salinity, in spring and winter, respectively. Despite evidence indicating the relative seasonal persistence of saltwater within the southern estuary and Trapandé Bay (Miyao et al., 1986Miyao, S., Nishihara, L. & Sarti, C. 1986. Características físicas e químicas do sistema estuarino-lagunar de Cananéia-Iguape. Boletim Do Instituto Oceanográfico, 34, 23–26. DOI: https://doi.org/10.1590/s0373-55241986000100003
https://doi.org/10.1590/s0373-5524198600...
), short-term episodes of intense continental runoff may move the saltwater wedge toward the southern area (Conti et al., 2012Conti, L., Araujo, C., Paolo, F., Barcellos, R., Rodrigues, M., Mahiques, M. & Furtado, V. 2012. An integrated GIS for sedimentological and geomorphological analysis of a lagoon environment. Barra de Cananéia inlet region, (Southeastern Brazil). Journal of Coastal Conservation, 16(1), 13–24. DOI: https://doi.org/10.1007/s11852-011-0164-1
https://doi.org/10.1007/s11852-011-0164-...
) mainly in wet seasons, and the average salinity in the region is 22.5, classifying it as a polyhaline water body (Perina and Abessa, 2020Perina, F. & Abessa, D. 2020. Contamination and toxicity in a subtropical Estuarine Protected Area influenced by former mining activities. Ocean and Coastal Research, 68. DOI: https://doi.org/10.1590/s2675-28242020068313
https://doi.org/10.1590/s2675-2824202006...
). The salinity in Pedrinhas station ranged from 2 to 12, in summer and winter, respectively, and is characterized as a mesohaline site. The flow rate of the Ribeira de Iguape River responds synchronously to variations in rainfall, which are lowest in June, when it averages 218 m−3 s−1, and triplicate during the rainy season, when it averages 634 m−3 s−1 (DAAE, 2019Departamento de Águas e Energia Elétrica do Estado de São Paulo. 2019. Banco de dados fluviométricos do Estado de São Paulo. Banco de dados hidrológicos. Accessed: http://www.hidrologia.daee.sp.gov.br/
http://www.hidrologia.daee.sp.gov.br/...
). This characteristic has important implications for the distribution and deposition of fine sediments, organic matter, and contaminants in the CIELC and clearly separates the south (greater marine influence) from the north (greater river influence), as we observed (see Figure 2).

Normally, lithogenic sediments (continental origin) are carried to the ocean through a fluvial flux, and areas closer to the continental water flux are mainly composed of coarser sediments and retain less OM. During this study we observed that the density, porosity, and TOM distribution were significantly correlated, with lower sedimentary density associated with a higher degree of porosity and greater OM accumulation. However, the freshwater station showed the highest porosity and TOM concentration. This was also observed in previous studies in the region, which showed higher OM concentrations closer to the Ribeira de Iguape River (Barcellos et al., 2009Barcellos, R. L., Camargo, P., Galvão, A. & Weber, R. 2009. Sedimentary organic matter in cores of the Cananeia-Iguape Lagoonal-Estuarine System, São Paulo State, Brazil. Journal of Coastal Research, Special Issue 56(II), 1335–1339.). This can be in part explained by the strong tidal wave propagation observed in the high and intermediate salinity sites, which leads to the higher hydrodynamic circulation and greater variation in granulometric and organic matter values within the system (Barcellos et al., 2009Barcellos, R. L., Camargo, P., Galvão, A. & Weber, R. 2009. Sedimentary organic matter in cores of the Cananeia-Iguape Lagoonal-Estuarine System, São Paulo State, Brazil. Journal of Coastal Research, Special Issue 56(II), 1335–1339.). On the other hand, the low salinity site is submitted to a higher sedimentation deposition (1.46 cm year−1)(Saito et al., 2001Saito, R. T., Figueira, R. C. L., Tessler, M. G. & Cunha, I. I. L. 2001. Geochronology of sediments in the Cananeia-Iguape estuary and in southern continental shelf of São Paulo State, Brazil. Czechoslovak Journal of Physics, 53(S1), A75–A81. DOI: https://doi.org/10.1007/s10582-003-0012-0
https://doi.org/10.1007/s10582-003-0012-...
) and has more homogenous OM content (Barcellos et al., 2009Barcellos, R. L., Camargo, P., Galvão, A. & Weber, R. 2009. Sedimentary organic matter in cores of the Cananeia-Iguape Lagoonal-Estuarine System, São Paulo State, Brazil. Journal of Coastal Research, Special Issue 56(II), 1335–1339.). This OM is associated with terrestrial input from the Ribeira de Iguape River and presents higher C:N and C:P values (Barcellos et al., 2009Barcellos, R. L., Camargo, P., Galvão, A. & Weber, R. 2009. Sedimentary organic matter in cores of the Cananeia-Iguape Lagoonal-Estuarine System, São Paulo State, Brazil. Journal of Coastal Research, Special Issue 56(II), 1335–1339.). Except in autumn, this site showed the lowest chlorophyll-a values, and the high salinity station always showed higher phytopigments (chlorophyll-a + phaeopigments) (Table 3). Phytopigment concentrations normally have an inverse relationship to the salinity (Aquino et al., 2012Aquino, E., Figueirêdo, L., Anjos, D., Passavante, J. & Silva-Cunha, M. 2012. Biomassa fitoplanctônica e fatores ambientais em um estuário tropical do Brasil. Tropical Oceanography, 40(1), 17–28. DOI: https://doi.org/10.5914/tropocean.v40i1.5190
https://doi.org/10.5914/tropocean.v40i1....
), mainly in sediment where a lower concentration of suspended particulate matter in water allows greater light intrusion, increasing the production of microphytobenthos.

Benthic respiration and sediment-water fluxes

The combination of dissolved oxygen and CO2 fluxes in the dark is the best proxy for heterotrophic respiration, including diffusive (microbial) and total (microbial + macrofaunal) uptake. In sediment overlayed by oxygen saturated waters, benthic respiration includes both the oxygen consumption by aerobic organisms and microbial mediated oxidation of reduced compounds (Mackin and Swider, 1989Mackin, J. & Swider, K. 1989. Organic matter decomposition pathways and oxygen consumption in coastal marine sediments. Journal of Marine Research, 47(3), 681–716. DOI: https://doi.org/10.1357/002224089785076154
https://doi.org/10.1357/0022240897850761...
), and it closely reflects the rates of oxic and anoxic decomposition in the sediment. The oxygen sedimentary consumption values in this study are within the range observed in similar subtropical estuarine systems Grenz et al. (2010Grenz, C., Denis, L., Pringault, O. & Fichez, R. 2010. Spatial and seasonal variability of sediment oxygen consumption and nutrient fluxes at the sediment water interface in a sub-tropical lagoon (New Caledonia). Marine Pollution Bulletin, 61(7–12), 399–412. DOI: https://doi.org/10.1016/j.marpolbul.2010.06.014
https://doi.org/10.1016/j.marpolbul.2010...
). To our knowledge, no published data regarding nutrient or sediment oxygen flux measurements at the sediment-water interface in CIELC are available, and tracing a pattern of benthic respiration is difficult with a single study. This is due to the large number of variables that control benthic oxygen consumption. Benthic respiration rates depend on factors such as dissolved oxygen content of the overlying water, organic load, benthic biomass, temperature, sediment type, and redox conditions. Normally, temperature combined with OM concentration controls the benthic mineralization, but physical and biological factors mediated by disturbances, including the impact of macro- and meiofauna, also affect the exchange rates between sediment and water column (Kristensen et al., 1992Kristensen, E., Devol, A., Ahmed, S. & Saleem, M. 1992. Preliminary study of benthic metabolism and sulfate reduction in a mangrove swamp of the Indus Delta, Pakistan. Marine Ecology Progress Series, 90, 287–297. DOI: https://doi.org/10.3354/meps090287
https://doi.org/10.3354/meps090287...
). Additionally, the OM quality and quantity in sediment control benthic respiration rates (Jahnke et al., 2005Jahnke, R., Richards, M., Nelson, J., Robertson, C., Rao, A. & Jahnke, D. 2005. Organic matter remineralization and porewater exchange rates in permeable South Atlantic Bight continental shelf sediments. Continental Shelf Research, 25(12–13), 1433–1452. DOI: https://doi.org/10.1016/j.csr.2005.04.002
https://doi.org/10.1016/j.csr.2005.04.00...
; Burdige, 2006Burdige, D. J. 2006. Geochemistry of Marine Sediments. Princeton: Princeton University Press.; Alongi et al., 2011Alongi, D., De Carvalho, N., Amaral, A., Costa, A., Trott, L. & Tirendi, F. 2011. Uncoupled surface and below-ground soil respiration in mangroves: implications for estimates of dissolved inorganic carbon export. Biogeochemistry, 109(1–3), 151–162. DOI: https://doi.org/10.1007/s10533-011-9616-9
https://doi.org/10.1007/s10533-011-9616-...
; Pastor et al., 2011Pastor, L., Deflandre, B., Viollier, E., Cathalot, C., Metzger, E., Rabouille, C., Escoubeyrou, K., Lloret, E., Pruski, A., Vétion, G., Desmalades, M., Buscail, R. & Grémare, A. 2011. Influence of the organic matter composition on benthic oxygen demand in the Rhône River prodelta (NW Mediterranean Sea). Continental Shelf Research, 31(9), 1008–1019. DOI: https://doi.org/10.1016/j.csr.2011.03.007
https://doi.org/10.1016/j.csr.2011.03.00...
).

The main results that emerge from this CIELC data are the large spatial and temporal heterogeneity of sediment-water fluxes, which reflect, in part, the high diversity of habitats, like all large estuarine ecosystems. In general, higher oxygen consumption was observed in spring at the HS site, which coincides with high chlorophyll-a content in the sediment. Normally an increase in primary production and chlorophyll content are followed by an increase in benthic metabolism. The stations with intermediate and low salinity showed similar seasonal variations, with higher values in autumn/winter and lower in spring/summer. Our observations showed highest OM content in these moments, which can explain the higher values of oxygen consumption. The IS cores in winter stood out due to the large presence of tubes on its surface (Figure 7), which after analysis were found to be of Tanaidacea. The bioturbation by macrofauna affects oxygen consumption not only due to respiration but also by increasing the solute exchange by sediment reworking, by building holes and tubes, and by bioirrigation, feeding, and excretion (Kristensen et al., 2012Kristensen, E., Penha-Lopes, G., Delefosse, M., Valdemarsen, T., Quintana, C. & Banta, G. 2012. What is bioturbation? The need for a precise definition for fauna in aquatic sciences. Marine Ecology Progress Series, 446, 285–302. DOI: https://doi.org/10.3354/meps09506
https://doi.org/10.3354/meps09506...
). Furthermore, other estuaries also showed higher oxygen consumption by the sediment at sites characterized by intermediate salinity values (Boynton and Kemp, 1985Boynton, W. & Kemp, W. 1985. Nutrient regeneration and oxygen consumption by sediments along an estuarine salinity gradient. Marine Ecology Progress Series, 23, 45–55. DOI: https://doi.org/10.3354/meps023045
https://doi.org/10.3354/meps023045...
; Buzzelli et al., 2013Buzzelli, C., Chen, Z., Coley, T., Doering, P., Samimy, R., Schlezinger, D. & Howes, B. 2013. Dry season sediment-water exchanges of nutrients and oxygen in two Florida estuaries: Patterns, comparisons, and internal loading. Biological Science, Florida Scientist, 1, 54–79. DOI: https://doi.org/10.5194/bg-10-6721-2013
https://doi.org/10.5194/bg-10-6721-2013...
). The greater variety of habitats and consequently the greater variety of metabolisms can lead to greater metabolic activity of the sediment in this area.

The general pattern of TCO2 efflux was somewhat similar to the O2 consumption, although no correlation between these two parameters was found (Figure 4, Table 4). The relationships between dark O2 uptakes and TCO2 effluxes in benthic systems result from a complex 3D mosaic of biogenic and chemical reactions, and TCO2 is produced by both metabolism and calcium carbonate dissolution/precipitation (Ferguson et al., 2003Ferguson, A.J.P., Eyre, B.D. & Gay, J.M. 2003. Organic matter and benthic metabolism in euphotic sediments along shallow sub-tropical estuaries, northern NSW, Australia. Aquatic Microbiology Ecology, 33:137–154). Therefore, O2 and CO2 are not likely to be well correlated in a dynamic system. The TCO2:O2 rate (Table 5) was calculated to observe how much of the CO2 liberation was due to heterotrophic respiration (TCO2:O2 = 1:1). The ratio 1:1 was observed only on one occasion, and in general the TCO2:O2 flux ratio was less than 1 throughout the study region, indicating that either sulfide-oxidation or nitrification may significantly influence the benthic fluxes. Furthermore, both sulfide-oxidizing and nitrifying bacteria are predominantly chemoautotrophic and may therefore lower the TCO2:O2 flux ratio via carbon fixation, especially in coastal areas where a large OM quantity limits the oxygen to the first centimeters of the sediment column.

Figure 7.
Sediment core of the intermediate salinity station in winter (August 2021) with a high number of Tanaidacea tubes.

Table 5.
Average ratios of O2:TCO2 fluxes at stations and in seasons.

Sediments play a major role in controlling the cycling and availability of nitrogen in estuaries (Giblin et al., 1997Giblin, A., Hopkinson, C. & Tucker, J. 1997. Benthic Metabolism and Nutrient Cycling in Boston Harbor, Massachusetts. Estuaries, 20(2), 346. DOI: https://doi.org/10.2307/1352349
https://doi.org/10.2307/1352349...
; Hopkinson et al., 1999Hopkinson, C., Giblin, A., Tucker, J. & Garritt, R. 1999. Benthic Metabolism and Nutrient Cycling along an Estuarine Salinity Gradient. Estuaries, 22(4), 863–881. DOI: https://doi.org/10.2307/1353067
https://doi.org/10.2307/1353067...
). Although sediments are generally considered a significant internal source of nutrients in shallow coastal ecosystems, several studies have shown that they may be a net sink of dissolved nitrogen at least during certain times of the year (Sundbäck et al., 2000Sundbäck, K., Miles, A. & Göransson, E. 2000. Nitrogen fluxes, denitrification and the role of microphytobenthos in microtidal shallow-water sediments:an annual study. Marine Ecology Progress Series, 200, 59–76. DOI: https://doi.org/10.3354/meps200059
https://doi.org/10.3354/meps200059...
; Tyler et al., 2003Tyler, A., Mcglathery, K. & Anderson, I. 2003. Benthic algae control sediment-water column fluxes of organic and inorganic nitrogen compounds in a temperate lagoon. Limnology and Oceanography, 48(6), 2125–2137. DOI: https://doi.org/10.4319/lo.2003.48.6.2125
https://doi.org/10.4319/lo.2003.48.6.212...
). Coastal ecosystems have been characterized as nitrate sinks, resulting in gaseous (N2) losses from high rates of denitrification (Galloway et al., 2008Galloway, J., Townsend, A., Erisman, J., Bekunda, M., Cai, Z., Freney, J., Martinelli, L., Seitzinger, S. & Sutton, M. 2008. Transformation of the Nitrogen Cycle: Recent Trends, Questions, and Potential Solutions. Science, 320(5878), 889–892. DOI: https://doi.org/10.1126/science.1136674
https://doi.org/10.1126/science.1136674...
). This study observed that, in general, N was liberated from the sediment to the water column. The efflux of NOX (NO2 and NO3) during most of the sediment incubation period indicated that nitrification was an important process within the CIELC estuarine sediments. Nitrification, which is the oxidation of ammonium to nitrate, is an important process in the water column and in sediments of many estuaries around the world (Wilde and Bie, 2000Wilde, H. P. J. de & Bie, M. J. M. de. 2000. Nitrous oxide in the Schelde estuary: production by nitrification and emission to the atmosphere. Marine Chemistry, 69(3–4), 203–216. DOI: https://doi.org/10.1016/s0304-4203(99)00106-1
https://doi.org/10.1016/s0304-4203(99)00...
; Dai et al., 2008Dai, M., Wang, L., Guo, X., Zhai, W., Li, Q., He, B. & Kao, S.-J. 2008. Nitrification and inorganic nitrogen distribution in a large perturbed river/estuarine system: the Pearl River Estuary, China. Biogeosciences, 5(5), 1227–1244. DOI: https://doi.org/10.5194/bg-5-1227-2008
https://doi.org/10.5194/bg-5-1227-2008...
; Damashek et al., 2016Damashek, J., Casciotti, K. & Francis, C. 2016. Variable Nitrification Rates Across Environmental Gradients in Turbid, Nutrient-Rich Estuary Waters of San Francisco Bay. Estuaries and Coasts, 39(4), 1050–1071. DOI: https://doi.org/10.1007/s12237-016-0071-7
https://doi.org/10.1007/s12237-016-0071-...
). The higher NOx efflux was observed in summer at the IS and LS sites. Despite de absence of significant correlations between NOx and NH4+ fluxes, at this time, the liberation was coupled with a large consumption of NH4+, indicating large nitrification rates.

At the HS site, the most dominant compound in the DIN flux was attributed to NH4+, and the NOx flux was very low at all flux measurements (Figure 3). At this site, NH4+ was always released from the sediment, mainly in spring and summer, when higher oxygen consumption was observed, and when the primary production in the water column was maximum (Barrera-alba et al., 2008Barrera-alba, J.J., Gianesella, S.M.F., Moser, G.A.O. & Saldanha-Correa F.M.P. 2008. Bacterial and 2169 phytoplankton dynamics in a sub-tropical estuary. Hydrobiologia, 598, 229-246.). The dominant NH4+ efflux observed during the study can indicate higher benthic faunal respiration, excretion, and a degradation of sedimentary organic N (Belias et al., 2007Belias, C., Dassenakis, M. & Scoullos, M. 2007. Study of the N, P and Si fluxes between fish farm sediment and seawater. Results of simulation experiments employing a benthic chamber under various redox conditions. Marine Chemistry, 103(3–4), 266–275. DOI: https://doi.org/10.1016/j.marchem.2006.09.005
https://doi.org/10.1016/j.marchem.2006.0...
). Ammonification is an important process adding ammonium in estuarine systems, and it is very active in sediments with abundant OM like estuarine sediments (Li et al., 2015Li, J., Nedwell, D., Beddow, J., Dumbrell, A., Mckew, B., Thorpe, E. & Whitby, C. 2015. amoA Gene Abundances and Nitrification Potential Rates Suggest that Benthic Ammonia-Oxidizing Bacteria and Not Archaea Dominate N Cycling in the Colne Estuary, United Kingdom. Applied and Environmental Microbiology, 81(1), 159–165. DOI: https://doi.org/10.1128/aem.02654-14
https://doi.org/10.1128/aem.02654-14...
). In addition, the dissimilatory nitrate reduction to ammonium (DNRA) could contribute to the accumulation of ammonium in anaerobic sediments (An and Gardner, 2002An, S. & Gardner, W. 2002. Dissimilatory nitrate reduction to ammonium (DNRA) as a nitrogen link, versus denitrification as a sink in a shallow estuary (Laguna Madre/Baffin Bay, Texas). Marine Ecology Progress Series, 237, 41–50. DOI: https://doi.org/10.3354/meps237041
https://doi.org/10.3354/meps237041...
; Gardner et al., 2006Gardner, W., Mccarthy, M., An, S., Sobolev, D., Sell, K. & Brock, D. 2006. Nitrogen fixation and dissimilatory nitrate reduction to ammonium (DNRA) support nitrogen dynamics in Texas estuaries. Limnology and Oceanography, 51, 558–568. DOI: https://doi.org/10.4319/lo.2006.51.1_part_2.0558
https://doi.org/10.4319/lo.2006.51.1_par...
). The DNRA is indicated to replace denitrification in marine sediments due to an inhibition in the activity of denitrifying organisms (Wang et al., 2018Wang, S., Wang, W., Liu, L., Zhuang, L., Zhao, S., Su, Y., Li, Y., Wang, M., Wang, C., Xu, L. & Zhu, G. 2018. Microbial Nitrogen Cycle Hotspots in the Plant-Bed/Ditch System of a Constructed Wetland with N2O Mitigation. Environmental Science & Technology, 52(11), 6226–6236. DOI: https://doi.org/10.1021/acs.est.7b04925
https://doi.org/10.1021/acs.est.7b04925...
). On the other hand, the salinity gradient can also directly affect DIN fluxes at the sediment-water interface (Zhou et al., 2017Zhou, M., Butterbach-Bahl, K., Vereecken, H. & Brüggemann, N. 2017. A meta-analysis of soil salinization effects on nitrogen pools, cycles and fluxes in coastal ecosystems. Global Change Biology, 23(3), 1338–1352. DOI: https://doi.org/10.1111/gcb.13430
https://doi.org/10.1111/gcb.13430...
). Under freshwater or low saline conditions, a greater portion of the pore-water NH4+ pool remained adsorbed in sediment clay minerals and was possibly nitrified, resulting in very low benthic NH4+ release. However, with increasing salinity, NH4+ could be paired with seawater anions, and the adsorption sites could be blocked by seawater cations. Consequently, the pore-water NH4+ largely escapes nitrification, which leads to higher upward diffusion of NH4+ across sediment-water interface (Gardner et al., 1991Gardner, W., Seitzinger, S. & Malczyk, J. 1991. The Effects of Sea Salts on the Forms of Nitrogen Released from Estuarine and Freshwater Sediments: Does Ion Pairing Affect Ammonium Flux? Estuaries, 14(2), 157–166. DOI: https://doi.org/10.2307/1351689
https://doi.org/10.2307/1351689...
; Seitzinger et al., 1991Seitzinger, S., Gardner, W. & Spratt, A. 1991. The Effect of Salinity on Ammonium Sorption in Aquatic Sediments: Implications for Benthic Nutrient Recycling. Estuaries, 14(2), 167–174. DOI: https://doi.org/10.2307/1351690
https://doi.org/10.2307/1351690...
). Other estuarine systems also showed the decrease in benthic NH4+ efflux with decreasing salinity (Rysgaard et al., 1999Rysgaard, Søren, Thastum, P., Dalsgaard, T., Christensen, P. B., Sloth, N. P. & Rysgaard, Soren. 1999. Effects of Salinity on NH 4+ Adsorption Capacity, Nitrification, and Denitrification in Danish Estuarine Sediments. Estuaries, 22(1), 21. DOI: https://doi.org/10.2307/1352923
https://doi.org/10.2307/1352923...
; Weston et al., 2010Weston, N., Giblin, A., Banta, G., Hopkinson, C. & Tucker, J. 2010. The Effects of Varying Salinity on Ammonium Exchange in Estuarine Sediments of the Parker River, Massachusetts. Estuaries and Coasts, 33(4), 985–1003. DOI: https://doi.org/10.1007/s12237-010-9282-5
https://doi.org/10.1007/s12237-010-9282-...
).

Nevertheless, some other factors, such as temperature, sediment organic content, and macroinvertebrate assemblages, also affect sediment DIN flux. For example, the large DIN release observed in the IS station in winter could be due to the high density of Tanaidacea found in the sediment (Figure 7). As previously stated, oxygen consumption and the release of NOx and NH4+ largely increased at this time. Bioturbation affects the transport of nutrients and often increases oxygen penetration into the sediment, which enhances mineralization processes and stimulates the nitrification and denitrification rates (Kristensen et al., 2012Kristensen, E., Penha-Lopes, G., Delefosse, M., Valdemarsen, T., Quintana, C. & Banta, G. 2012. What is bioturbation? The need for a precise definition for fauna in aquatic sciences. Marine Ecology Progress Series, 446, 285–302. DOI: https://doi.org/10.3354/meps09506
https://doi.org/10.3354/meps09506...
; Moraes et al., 2018Moraes, P. C., Zilius, M., Benelli, S. & Bartoli, M. 2018. Nitrification and denitrification in estuarine sediments with tube-dwelling benthic animals. Hydrobiologia, 819(1), 217–230. DOI: https://doi.org/10.1007/s10750-018-3639-3
https://doi.org/10.1007/s10750-018-3639-...
; Bartoli et al., 2020Bartoli, M., Benelli, S., Magri, M., Ribaudo, C., Moraes, P. & Castaldelli, G. 2020. Contrasting Effects of Bioturbation Studied in Intact and Reconstructed Estuarine Sediments. Water, 12(11), 3125. DOI: https://doi.org/10.3390/w12113125
https://doi.org/10.3390/w12113125...
).

Silicate is commonly released from the sediments as a diffusion flux controlled by its concentration gradient at the sediment-water interface (Willey, 1978Willey, J. D. 1978. Release and uptake of dissolved silica in seawater by marine sediments. Marine Chemistry, 7(1), 53–65. DOI: https://doi.org/10.1016/0304-4203(78)90042-7
https://doi.org/10.1016/0304-4203(78)900...
). This study found, on two occasions, an influx of silicate to the sediment (in IS and LS, in autumn). The concentration of the water column silicate in the area were not analyzed but we sampled water from the sites for the experiment. These values were our time zero concentration. We observed that in the moments in which silicates showed an influx to the sediment silicates were in high concentration on the water. These unusual Si(OH)4 influxes related to high water column silicate were also recorded by Niencheski and Jahnke (2002Niencheski, L. & Jahnke, R. A. 2002. Benthic respiration and inorganic nutrient fluxes in the estuarine region of Patos Lagoon (Brazil). Aquatic Geochemistry, 8(3), 135–152.) in the Patos lagoon (Brazil) and by Grenz et al. (2019Grenz, C., Moreno, M., Soetaert, K., Denis, L., Douillet, P. & Fichez, R. 2019. Spatio-temporal variability in benthic exchanges at the sediment-water interface of a shallow tropical coastal lagoon (south coast of Gulf of Mexico). Estuarine, Coastal and Shelf Science, 218, 368–380. DOI: https://doi.org/10.1016/j.ecss.2019.01.012
https://doi.org/10.1016/j.ecss.2019.01.0...
) in Laguna de Términos (Mexico). For all other observations, Si(OH)4 fluxes were directed toward the water column and correlated with the DIN, NOx, and PO4-3 fluxes. Other factors that possibly influence the benthic silicate release are the abundance, distribution, and dissolution rate of biogenic silica (diatom frustules) in the estuarine sediments. High diatom abundance in the estuary during high productivity seasons is common (Barrera-Alba and Moser, 2016Barrera-Alba, J.J. & Moser, G.A. O. 2016. Short-term response of phytoplankton community to over-enrichment of nutrients in a well-preserved subtropical estuary. Brazilian Journal of Oceanography, 64(2), 191-196.) and results in higher availability of diatom frustules in the sediments and their consequent dissolution. Dissolution of biogenic silica in sediments and benthic Si(OH)4 efflux is stimulated by temperature and salinity (chemical composition) of the overlying water (Pratihary et al., 2021Pratihary, A., Naik, R., Karapurkar, S., Gauthankar, M., Khandeparker, R., Manjrekar, S. & Gauns, M. 2021. Benthic exchange along a tropical estuarine salinity gradient during dry season: Biogeochemical and ecological implications. Journal of Sea Research, 177, 102124. DOI: https://doi.org/10.1016/j.seares.2021.102124
https://doi.org/10.1016/j.seares.2021.10...
).

Sediment-water HPO42− fluxes were directed both into and out of sediments, with values in the range of estuarine systems (Boynton et al., 2018Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
https://doi.org/10.1007/s12237-017-0275-...
). The HPO42− flux varied greatly between the sediment and the water column, always with the inverse pattern between the HS and LS stations. Whereas the first presented phosphorus consumption by the sediment in winter and autumn and release in summer and spring, the second had an inverse pattern, with liberation from the sediment in winter and autumn and consumption in spring and summer. Often, when HPO42− efflux is observed in estuarine systems, it is accompanied by reduced dissolved oxygen in the water column (Cowan et al., 1996Cowan, J. L. W., Pennock, J. & Boynton, W. 1996. Seasonal and interannual patterns of sediment-water nutrient and oxygen fluxes in Mobile Bay, Alabama (USA): regulating factors and ecological significance. Marine Ecology Progress Series, 141, 229–245. DOI: https://doi.org/10.3354/meps141229
https://doi.org/10.3354/meps141229...
; Didonato et al., 2006Didonato, G., Lores, E., Murrell, M., Smith, L. & Caffrey, J. 2006. Benthic Nutrient Flux in a Small Estuary in Northwestern Florida (USA). Gulf and Caribbean Research, 18(1), 15–26. DOI: https://doi.org/10.18785/gcr.1801.02
https://doi.org/10.18785/gcr.1801.02...
). No low oxygen concentration were observed in the CIELC in this study or in the consulted literature. Normally, an estuary with high saline sediments retains very little regenerated P (10%), whereas freshwater sediments can retain 80 to 90% (Roden and Edmonds, 1997Roden, E. & Edmonds, J. 1997. Phosphate mobilization in iron-rich anaerobic sediments: microbial Fe (III) oxide reduction versus iron-sulfide formation. Archiv Für Hydrobiologie, 139(3), 347–378. DOI: https://doi.org/10.1127/archiv-hydrobiol/139/1997/347
https://doi.org/10.1127/archiv-hydrobiol...
; Gächter and Müller, 2003Gächter, R. & Müller, B. 2003. Why the phosphorus retention of lakes does not necessarily depend on the oxygen supply to their sediment surface. Limnology and Oceanography, 48(2), 929–933. DOI: https://doi.org/10.4319/lo.2003.48.2.0929
https://doi.org/10.4319/lo.2003.48.2.092...
), and increasing salinity can mobilize the sedimentary HPO42− facilitating its release to the overlying water (Caraco et al., 1990Caraco, N., Cole, J. & Likens, Genee. 1990. A comparison of phosphorus immobilization in sediments of freshwater and coastal marine systems. Biogeochemistry, 9(3), 277–290. DOI: https://doi.org/10.1007/bf00000602
https://doi.org/10.1007/bf00000602...
; Gunnars and Blomqvist, 1997Gunnars, A. & Blomqvist, S. 1997. Phosphate exchange across the sediment–water interface when shifting from anoxic to oxic conditions—an experimental comparison of freshwater and brackish-marine systems. Biogeochemistry, 37, 203-226.; Jordan et al., 2008Jordan, T., Cornwell, J., Boynton, W. & Anderson, J. 2008. Changes in phosphorus biogeochemistry along an estuarine salinity gradient: The iron conveyer belt. Limnology and Oceanography, 53(1), 172–184. DOI: https://doi.org/10.4319/lo.2008.53.1.0172
https://doi.org/10.4319/lo.2008.53.1.017...
). Still, SO42− reduction in anoxic sediments can increase alkalinity and pH, which can inhibit HPO42− adsorption onto Fe-oxide minerals (Caraco et al., 1989Caraco, N., Cole, J. & Likens, G. 1989. Evidence for sulphate-controlled phosphorus release from sediments of aquatic systems. Nature, 341(6240), 316–318. DOI: https://doi.org/10.1038/341316a0
https://doi.org/10.1038/341316a0...
), and Fe-oxide minerals are known to adsorb/co-precipitate PO43− (Krom and Berner, 1980Krom, M. & Berner, R. 1980. Adsorption of phosphate in anoxic marine sediments 1. Limnology and Oceanography, 25(5), 797–806. DOI: https://doi.org/10.4319/lo.1980.25.5.0797
https://doi.org/10.4319/lo.1980.25.5.079...
; Boström et al., 1988Boström, B., Andersen, J. M., Fleischer, S. & Jansson, M. 1988. Exchange of Phosphorus Across the Sediment-Water Interface. In: Persson, G. & Jansson, M. (eds.), Phosphorus in Freshwater Ecosystems. Developments in Hydrobiology. Dordrecht: Springer.; Gunnars and Blomqvist, 1997Gunnars, A. & Blomqvist, S. 1997. Phosphate exchange across the sediment–water interface when shifting from anoxic to oxic conditions—an experimental comparison of freshwater and brackish-marine systems. Biogeochemistry, 37, 203-226.). As in the water column, in this system, the sediment remains well oxidized throughout the year, surface sediments are not likely to ever become anoxic, and the benthic flux of HPO42− could be limited and coupled strongly with Fe-oxyhydroxide.

CONCLUSION

Benthic respiration and nutrient exchange rates varied considerably along the salinity gradient and seasonally at Cananéia-Iguape Estuarine-Lagoon Complex (CIELC), São Paulo, Brazil. Benthic respiration was low but comparable with those found in estuarine sites at the same latitude, although few studies showed similar conditions of system preservation. The CIELC is a pristine environment with very low human impact, and it is part of a large preservation area. Normally, studies found in the literature refer to estuaries with high human impact, different from the condition of the studied area. We observed a dominance of nitrate efflux that suggest a big importance of benthic nitrification in the region. In general, the sediment was a net DIN source to the water column, releasing large quantities of nitrate (NO3) and ammonium (NH4+). While the estuarine sediments behaved as both HPO4−2 sink and HPO4−2 source to the water column, showing that this nutrient must not be limited to the primary production, benthic silicate was always liberated in large quantities to the estuary water, showing that the sediment is a Si(OH)4 source for the water column. Benthic faunal activity possibly enhanced the benthic exchange, causing the IS in winter to show a very particular behavior. Overall, this study showed that benthic exchange played an important role in the biogeochemistry and ecology of the CIELC, as well as seasonal and spatial variations. The interaction between space and time factors was always highly significant, showing that the metabolic behavior of the benthic compartment depends on both trophic and physicochemical conditions. In general, the freshwater site was characterized as a nitrogen sink and a silicate and phosphate release, despite some seasonal variations. The sediment of the marine site was characterized, in general, as a nitrogen producer and a phosphorus consumer. However, further studies on compound-specific isotopic composition, water column, and sediment analysis are necessary to provide a more precise understanding of benthic metabolism and N cycling, its control of benthic nutrient release, and its impact on the pelagic biogeochemistry of the estuary.

ACKNOWLEDGMENTS

The authors thank the São Paulo Research Foundation “Fundação de Amparo à Pesquisa do Estado de São Paulo” (FAPESP) for supporting the Project BIOGEOQUIS-ECOS – Proc. 2020/16485-7 and for the Postdoctoral fellowship 2018/24119-0. We gratefully acknowledge the support from the Research Base of Cananéia “Dr João de Paiva Carvalho” employees and the support from the team of LABNUT-IOUSP. We also thank André C. Souza and Guilherme W.P. Leite for sampling assistance. We thank the editor and the two reviewers for the corrections and suggestions.

REFERENCES

  • Alongi, D., De Carvalho, N., Amaral, A., Costa, A., Trott, L. & Tirendi, F. 2011. Uncoupled surface and below-ground soil respiration in mangroves: implications for estimates of dissolved inorganic carbon export. Biogeochemistry, 109(1–3), 151–162. DOI: https://doi.org/10.1007/s10533-011-9616-9
    » https://doi.org/10.1007/s10533-011-9616-9
  • An, S. & Gardner, W. 2002. Dissimilatory nitrate reduction to ammonium (DNRA) as a nitrogen link, versus denitrification as a sink in a shallow estuary (Laguna Madre/Baffin Bay, Texas). Marine Ecology Progress Series, 237, 41–50. DOI: https://doi.org/10.3354/meps237041
    » https://doi.org/10.3354/meps237041
  • Aquino, E., Figueirêdo, L., Anjos, D., Passavante, J. & Silva-Cunha, M. 2012. Biomassa fitoplanctônica e fatores ambientais em um estuário tropical do Brasil. Tropical Oceanography, 40(1), 17–28. DOI: https://doi.org/10.5914/tropocean.v40i1.5190
    » https://doi.org/10.5914/tropocean.v40i1.5190
  • Azevedo, J.S. & Braga, E.S. 2011. Caracterização hidroquímica para qualificação ambiental dos estuários de Santos-São Vicente e Cananéia. Arquivos de Ciências do Mar, 44(2), 52-61.
  • Banta, G., Giblin, A., Hobbie, J. & Tucker, J. 1995. Benthic respiration and nitrogen release in Buzzards Bay, Massachusetts. Journal of Marine Research, 53(1), 107–135. DOI: https://doi.org/10.1357/0022240953213287
    » https://doi.org/10.1357/0022240953213287
  • Barbier, E., Hacker, S., Kennedy, C., Koch, E., Stier, A. & Silliman, B. 2011. The value of estuarine and coastal ecosystem services. Ecological Monographs, 81(2), 169–193. DOI: https://doi.org/10.1890/10-1510.1
    » https://doi.org/10.1890/10-1510.1
  • Barcellos, R. L., Camargo, P., Galvão, A. & Weber, R. 2009. Sedimentary organic matter in cores of the Cananeia-Iguape Lagoonal-Estuarine System, São Paulo State, Brazil. Journal of Coastal Research, Special Issue 56(II), 1335–1339.
  • Barletta, M. & Dantas, D. V. 2016. Environmental gradient. In: Kennish, M. J. (ed.), Encyclopedia of estuaries. Dordrecht: Springer.
  • Barrera-alba, J.J., Gianesella, S.M.F., Moser, G.A.O. & Saldanha-Correa F.M.P. 2008. Bacterial and 2169 phytoplankton dynamics in a sub-tropical estuary. Hydrobiologia, 598, 229-246.
  • Barrera-Alba, J.J. & Moser, G.A. O. 2016. Short-term response of phytoplankton community to over-enrichment of nutrients in a well-preserved subtropical estuary. Brazilian Journal of Oceanography, 64(2), 191-196.
  • Bartoli, M., Benelli, S., Magri, M., Ribaudo, C., Moraes, P. & Castaldelli, G. 2020. Contrasting Effects of Bioturbation Studied in Intact and Reconstructed Estuarine Sediments. Water, 12(11), 3125. DOI: https://doi.org/10.3390/w12113125
    » https://doi.org/10.3390/w12113125
  • Basha, S. K. 2018. An overview on global mangroves distribution. Indian Journal of Geo Marine Sciences, 47(4), 766–772.
  • Belias, C., Dassenakis, M. & Scoullos, M. 2007. Study of the N, P and Si fluxes between fish farm sediment and seawater. Results of simulation experiments employing a benthic chamber under various redox conditions. Marine Chemistry, 103(3–4), 266–275. DOI: https://doi.org/10.1016/j.marchem.2006.09.005
    » https://doi.org/10.1016/j.marchem.2006.09.005
  • Bernardes, M. & Miranda, L. 2001. Circulação estacionária e estratificação de sal em canais estuarinos: simulação com modelos analíticos. Revista Brasileira de Oceanografia, 49(1–2), 115–132. DOI: https://doi.org/10.1590/s1413-77392001000100010
    » https://doi.org/10.1590/s1413-77392001000100010
  • Bonaglia, S., Deutsch, B., Bartoli, M., Marchant, H. & Brüchert, V. 2014. Seasonal oxygen, nitrogen and phosphorus benthic cycling along an impacted Baltic Sea estuary: regulation and spatial patterns. Biogeochemistry, 119(1–3), 139–160. DOI: https://doi.org/10.1007/s10533-014-9953-6
    » https://doi.org/10.1007/s10533-014-9953-6
  • Boström, B., Andersen, J. M., Fleischer, S. & Jansson, M. 1988. Exchange of Phosphorus Across the Sediment-Water Interface. In: Persson, G. & Jansson, M. (eds.), Phosphorus in Freshwater Ecosystems. Developments in Hydrobiology. Dordrecht: Springer.
  • Boynton, W., Ceballos, M., Bailey, E., Hodgkins, C., Humphrey, J. & Testa, J. 2018. Oxygen and Nutrient Exchanges at the Sediment-Water Interface: a Global Synthesis and Critique of Estuarine and Coastal Data. Estuaries and Coasts, 41(2), 301–333. DOI: https://doi.org/10.1007/s12237-017-0275-5
    » https://doi.org/10.1007/s12237-017-0275-5
  • Boynton, W. & Kemp, W. 1985. Nutrient regeneration and oxygen consumption by sediments along an estuarine salinity gradient. Marine Ecology Progress Series, 23, 45–55. DOI: https://doi.org/10.3354/meps023045
    » https://doi.org/10.3354/meps023045
  • Boynton, W. R., Kemp, W. M. & Keefe, C. W. 1982. A comparative analysis of nutrients and other factors influencing estuarine phytoplankton production. In: Kennedy, V. S. (ed.), Estuarine comparisons (pp. 69–90). New York: Academic Press.
  • Burdige, D. J. 2006. Geochemistry of Marine Sediments. Princeton: Princeton University Press.
  • Buzzelli, C., Chen, Z., Coley, T., Doering, P., Samimy, R., Schlezinger, D. & Howes, B. 2013. Dry season sediment-water exchanges of nutrients and oxygen in two Florida estuaries: Patterns, comparisons, and internal loading. Biological Science, Florida Scientist, 1, 54–79. DOI: https://doi.org/10.5194/bg-10-6721-2013
    » https://doi.org/10.5194/bg-10-6721-2013
  • Caraco, N., Cole, J. & Likens, G. 1989. Evidence for sulphate-controlled phosphorus release from sediments of aquatic systems. Nature, 341(6240), 316–318. DOI: https://doi.org/10.1038/341316a0
    » https://doi.org/10.1038/341316a0
  • Caraco, N., Cole, J. & Likens, Genee. 1990. A comparison of phosphorus immobilization in sediments of freshwater and coastal marine systems. Biogeochemistry, 9(3), 277–290. DOI: https://doi.org/10.1007/bf00000602
    » https://doi.org/10.1007/bf00000602
  • Carstensen, J., Conley, D., Bonsdorff, E., Gustafsson, B., Hietanen, S., Janas, U., Jilbert, T., Maximov, A., Norkko, A., Norkko, J., Reed, D., Slomp, C., Timmermann, K. & Voss, M. 2014. Hypoxia in the Baltic Sea: Biogeochemical Cycles, Benthic Fauna, and Management. Ambio, 43(1), 26–36. DOI: https://doi.org/10.1007/s13280-013-0474-7
    » https://doi.org/10.1007/s13280-013-0474-7
  • Chau, K. 2002. Field measurements of SOD and sediment nutrient fluxes in a land-locked embayment in Hong Kong. Advances in Environmental Research, 6(2), 135–142. DOI: https://doi.org/10.1016/s1093-0191(00)00075-7
    » https://doi.org/10.1016/s1093-0191(00)00075-7
  • Cloern, J., Foster, S. & Kleckner, A. 2014. Phytoplankton primary production in the world’s estuarine-coastal ecosystems. Biogeosciences, 11(9), 2477–2501. DOI: https://doi.org/10.5194/bg-11-2477-2014
    » https://doi.org/10.5194/bg-11-2477-2014
  • Conti, L., Araujo, C., Paolo, F., Barcellos, R., Rodrigues, M., Mahiques, M. & Furtado, V. 2012. An integrated GIS for sedimentological and geomorphological analysis of a lagoon environment. Barra de Cananéia inlet region, (Southeastern Brazil). Journal of Coastal Conservation, 16(1), 13–24. DOI: https://doi.org/10.1007/s11852-011-0164-1
    » https://doi.org/10.1007/s11852-011-0164-1
  • Cowan, J. & Boynton, W. 1996. Sediment-water oxygen and nutrient exchanges along the longitudinal axis of Chesapeake Bay: seasonal patterns, controlling factors and ecological significance. Estuaries, 19(3), 562–580. DOI: https://doi.org/10.2307/1352518
    » https://doi.org/10.2307/1352518
  • Cowan, J. L. W., Pennock, J. & Boynton, W. 1996. Seasonal and interannual patterns of sediment-water nutrient and oxygen fluxes in Mobile Bay, Alabama (USA): regulating factors and ecological significance. Marine Ecology Progress Series, 141, 229–245. DOI: https://doi.org/10.3354/meps141229
    » https://doi.org/10.3354/meps141229
  • Dai, M., Wang, L., Guo, X., Zhai, W., Li, Q., He, B. & Kao, S.-J. 2008. Nitrification and inorganic nitrogen distribution in a large perturbed river/estuarine system: the Pearl River Estuary, China. Biogeosciences, 5(5), 1227–1244. DOI: https://doi.org/10.5194/bg-5-1227-2008
    » https://doi.org/10.5194/bg-5-1227-2008
  • Damashek, J., Casciotti, K. & Francis, C. 2016. Variable Nitrification Rates Across Environmental Gradients in Turbid, Nutrient-Rich Estuary Waters of San Francisco Bay. Estuaries and Coasts, 39(4), 1050–1071. DOI: https://doi.org/10.1007/s12237-016-0071-7
    » https://doi.org/10.1007/s12237-016-0071-7
  • Day, J. W., Crump, B. C., Kemp, W. M. & Yáñez-Arancibia, A. (eds.). 1989. Estuarine Ecology. New York: John Wiley & Sons, Inc.
  • Departamento de Águas e Energia Elétrica do Estado de São Paulo. 2019. Banco de dados fluviométricos do Estado de São Paulo. Banco de dados hidrológicos. Accessed: http://www.hidrologia.daee.sp.gov.br/
    » http://www.hidrologia.daee.sp.gov.br/
  • Dickson, A. G., Sabine, C. L. & Christian, J. R. (eds.). 2007. Guide to Best Practices for Ocean CO2 Measurements. Asheville: NCEI.
  • Didonato, G., Lores, E., Murrell, M., Smith, L. & Caffrey, J. 2006. Benthic Nutrient Flux in a Small Estuary in Northwestern Florida (USA). Gulf and Caribbean Research, 18(1), 15–26. DOI: https://doi.org/10.18785/gcr.1801.02
    » https://doi.org/10.18785/gcr.1801.02
  • Diegues, A. C. 1987. Conservação e desenvolvimento sustentado de ecossistemas litorâneos no Brasil. São Paulo: Secretaria do Meio Ambiente.
  • Evans, M. & Scavia, D. 2013. Exploring estuarine eutrophication sensitivity to nutrient loading. Limnology and Oceanography, 58(2), 569–578. DOI: https://doi.org/10.4319/lo.2013.58.2.0569
    » https://doi.org/10.4319/lo.2013.58.2.0569
  • Eyre, B. & Balls, P. 1999. A Comparative Study of Nutrient Behavior along the Salinity Gradient of Tropical and Temperate Estuaries. Estuaries, 22(2), 313–326. DOI: https://doi.org/10.2307/1352987
    » https://doi.org/10.2307/1352987
  • Fellman, J., Petrone, K. & Grierson, P. 2011. Source, biogeochemical cycling, and fluorescence characteristics of dissolved organic matter in an agro-urban estuary. Limnology and Oceanography, 56(1), 243–256. DOI: https://doi.org/10.4319/lo.2011.56.1.0243
    » https://doi.org/10.4319/lo.2011.56.1.0243
  • Ferguson, A.J.P., Eyre, B.D. & Gay, J.M. 2003. Organic matter and benthic metabolism in euphotic sediments along shallow sub-tropical estuaries, northern NSW, Australia. Aquatic Microbiology Ecology, 33:137–154
  • Ferguson, A. & Eyre, B. 2013. Interaction of benthic microalgae and macrofauna in the control of benthic metabolism, nutrient fluxes and denitrification in a shallow sub-tropical coastal embayment (western Moreton Bay, Australia). Biogeochemistry, 112(1–3), 423–440. DOI: https://doi.org/10.1007/s10533-012-9736-x
    » https://doi.org/10.1007/s10533-012-9736-x
  • Gächter, R. & Müller, B. 2003. Why the phosphorus retention of lakes does not necessarily depend on the oxygen supply to their sediment surface. Limnology and Oceanography, 48(2), 929–933. DOI: https://doi.org/10.4319/lo.2003.48.2.0929
    » https://doi.org/10.4319/lo.2003.48.2.0929
  • Galloway, J., Townsend, A., Erisman, J., Bekunda, M., Cai, Z., Freney, J., Martinelli, L., Seitzinger, S. & Sutton, M. 2008. Transformation of the Nitrogen Cycle: Recent Trends, Questions, and Potential Solutions. Science, 320(5878), 889–892. DOI: https://doi.org/10.1126/science.1136674
    » https://doi.org/10.1126/science.1136674
  • Garcia, J., Lopes, A., Silvestre, A., Grabowski, R., Barioto, J., Costa, R. & Castilho, A. 2018. Environmental characterization of the Cananéia coastal area and its associated estuarine system (São Paulo state, Brazil): Considerations for three Penaeoidean shrimp species. Regional Studies in Marine Science, 19, 9–16. DOI: https://doi.org/10.1016/j.rsma.2018.02.010
    » https://doi.org/10.1016/j.rsma.2018.02.010
  • Gardner, W., Mccarthy, M., An, S., Sobolev, D., Sell, K. & Brock, D. 2006. Nitrogen fixation and dissimilatory nitrate reduction to ammonium (DNRA) support nitrogen dynamics in Texas estuaries. Limnology and Oceanography, 51, 558–568. DOI: https://doi.org/10.4319/lo.2006.51.1_part_2.0558
    » https://doi.org/10.4319/lo.2006.51.1_part_2.0558
  • Gardner, W., Seitzinger, S. & Malczyk, J. 1991. The Effects of Sea Salts on the Forms of Nitrogen Released from Estuarine and Freshwater Sediments: Does Ion Pairing Affect Ammonium Flux? Estuaries, 14(2), 157–166. DOI: https://doi.org/10.2307/1351689
    » https://doi.org/10.2307/1351689
  • Giblin, A., Hopkinson, C. & Tucker, J. 1997. Benthic Metabolism and Nutrient Cycling in Boston Harbor, Massachusetts. Estuaries, 20(2), 346. DOI: https://doi.org/10.2307/1352349
    » https://doi.org/10.2307/1352349
  • Gomez-Ramirez, E. H., Corzo, A., Garcia-Robledo, E., Bohorquez, J., Agüera-Jaquemet, A., Bibbo-Sanchez, F.,Soria-Píriz, S., Jiménez-Arias, J. L., Morales, A. & Papaspyrou, S. 2019. Benthic-pelagic coupling of carbon and nitrogen along a tropical estuarine gradient (Gulf of Nicoya, Costa Rica). Estuarine, Coastal and Shelf Science, 228, 1-11. DOI: https://doi.org/10.1016/j.ecss.2019.106362
    » https://doi.org/10.1016/j.ecss.2019.106362
  • Grasshoff, K., Kremling, K. & Ehrhardt, M. (eds.). 1983. Methods of Seawater Analysis (2nd ed.). Berlin: Wiley.
  • Grenz, C., Denis, L., Boucher, G., Chauvaud, L., Clavier, J., Fichez, R. & Pringault, O. 2003. Spatial variability in Sediment Oxygen Consumption under winter conditions in a lagoonal system in New Caledonia (South Pacific). Journal of Experimental Marine Biology and Ecology, 285–286, 33–47. DOI: https://doi.org/10.1016/s0022-0981(02)00518-x
    » https://doi.org/10.1016/s0022-0981(02)00518-x
  • Grenz, C., Denis, L., Pringault, O. & Fichez, R. 2010. Spatial and seasonal variability of sediment oxygen consumption and nutrient fluxes at the sediment water interface in a sub-tropical lagoon (New Caledonia). Marine Pollution Bulletin, 61(7–12), 399–412. DOI: https://doi.org/10.1016/j.marpolbul.2010.06.014
    » https://doi.org/10.1016/j.marpolbul.2010.06.014
  • Grenz, C., Moreno, M., Soetaert, K., Denis, L., Douillet, P. & Fichez, R. 2019. Spatio-temporal variability in benthic exchanges at the sediment-water interface of a shallow tropical coastal lagoon (south coast of Gulf of Mexico). Estuarine, Coastal and Shelf Science, 218, 368–380. DOI: https://doi.org/10.1016/j.ecss.2019.01.012
    » https://doi.org/10.1016/j.ecss.2019.01.012
  • Gunnars, A. & Blomqvist, S. 1997. Phosphate exchange across the sediment–water interface when shifting from anoxic to oxic conditions—an experimental comparison of freshwater and brackish-marine systems. Biogeochemistry, 37, 203-226.
  • Helder, W. & De Vries, R. 1983. Estuarine nitrite maxima and nitrifying bacteria (Ems-Dollard estuary). Netherlands Journal of Sea Research, 17(1), 1–18. DOI: https://doi.org/10.1016/0077-7579(83)90002-9
    » https://doi.org/10.1016/0077-7579(83)90002-9
  • Hellemann, D., Tallberg, P., Bartl, I., Voss, M. & Hietanen, S. 2017. Denitrification in an oligotrophic estuary: a delayed sink for riverine nitrate. Marine Ecology Progress Series, 583, 63–80. DOI: https://doi.org/10.3354/meps12359
    » https://doi.org/10.3354/meps12359
  • Hopkinson, C., Giblin, A. & Tucker, J. 2001. Benthic metabolism and nutrient regeneration on the continental shelf of Eastern Massachusetts, USA. Marine Ecology Progress Series, 224, 1–19. DOI: https://doi.org/10.3354/meps224001
    » https://doi.org/10.3354/meps224001
  • Hopkinson, C., Giblin, A., Tucker, J. & Garritt, R. 1999. Benthic Metabolism and Nutrient Cycling along an Estuarine Salinity Gradient. Estuaries, 22(4), 863–881. DOI: https://doi.org/10.2307/1353067
    » https://doi.org/10.2307/1353067
  • Hopkinson, C. S., Buffam, I., Hobbie, J., Vallino, J., Perdue, M., Eversmeyer, B., Prahl, F., Covert, J., Hodson, R., Moran, M. A., Smith, E., Baross, J., Crump, B., Findlay, S. & Foreman, K. 1998. Terrestrial inputs of organic matter to coastal ecosystems: An intercomparison of chemical characteristics and bioavailability. Biogeochemistry, 43(3), 211–234.
  • Humborg, C., Danielsson, Å., Sjöberg, B. & Green, M. 2003. Nutrient land–sea fluxes in oligothrophic and pristine estuaries of the Gulf of Bothnia, Baltic Sea. Estuarine, Coastal and Shelf Science, 56(3–4), 781–793. DOI: https://doi.org/10.1016/s0272-7714(02)00290-1
    » https://doi.org/10.1016/s0272-7714(02)00290-1
  • Italiani, D. & Mahiques, M. 2014. O registro geológico da atividade antropogênica na região do Valo Grande, Estado de São Paulo, Brasil. Quaternary and Environmental Geosciences, 5(2), 33–44. DOI: https://doi.org/10.5380/abequa.v5i2.34522
    » https://doi.org/10.5380/abequa.v5i2.34522
  • Jahnke, R., Richards, M., Nelson, J., Robertson, C., Rao, A. & Jahnke, D. 2005. Organic matter remineralization and porewater exchange rates in permeable South Atlantic Bight continental shelf sediments. Continental Shelf Research, 25(12–13), 1433–1452. DOI: https://doi.org/10.1016/j.csr.2005.04.002
    » https://doi.org/10.1016/j.csr.2005.04.002
  • Jordan, T., Cornwell, J., Boynton, W. & Anderson, J. 2008. Changes in phosphorus biogeochemistry along an estuarine salinity gradient: The iron conveyer belt. Limnology and Oceanography, 53(1), 172–184. DOI: https://doi.org/10.4319/lo.2008.53.1.0172
    » https://doi.org/10.4319/lo.2008.53.1.0172
  • Kemp, M. & Boynton, W. 1984. Spatial and temporal coupling of nutrient inputs to estuarine primary production: the role of particulate transport and decomposition. Bulletin of Marine Science, 35(3), 522–535.
  • Knoppers, B., Machado, E., Brandini, N. & Landim De Souza, W. F. 2004. Environmental Geochemistry in Tropical and Subtropical Environments Science. In: Duursma, E. K. & Abrão, J. J. (eds.) (pp. 253–275). Berlin, Heidelberg: Springer. DOI: https://doi.org/10.1007/978-3-662-07060-4_19
    » https://doi.org/10.1007/978-3-662-07060-4_19
  • Knoppers, B., Souza, W., Souza, M., Gonzalez Rodriguez, E., Landim, E. & Vieira, A. 1996. In situ measurements of benthic primary production, respiration and nutrient fluxes in a hypersaline coastal lagoon of SE Brazil. Revista Brasileira de Oceanografia, 44(2), 155–165. DOI: https://doi.org/10.1590/s1413-77391996000200005
    » https://doi.org/10.1590/s1413-77391996000200005
  • Kristensen, E., Devol, A., Ahmed, S. & Saleem, M. 1992. Preliminary study of benthic metabolism and sulfate reduction in a mangrove swamp of the Indus Delta, Pakistan. Marine Ecology Progress Series, 90, 287–297. DOI: https://doi.org/10.3354/meps090287
    » https://doi.org/10.3354/meps090287
  • Kristensen, E., Penha-Lopes, G., Delefosse, M., Valdemarsen, T., Quintana, C. & Banta, G. 2012. What is bioturbation? The need for a precise definition for fauna in aquatic sciences. Marine Ecology Progress Series, 446, 285–302. DOI: https://doi.org/10.3354/meps09506
    » https://doi.org/10.3354/meps09506
  • Krom, M. & Berner, R. 1980. Adsorption of phosphate in anoxic marine sediments 1. Limnology and Oceanography, 25(5), 797–806. DOI: https://doi.org/10.4319/lo.1980.25.5.0797
    » https://doi.org/10.4319/lo.1980.25.5.0797
  • Li, J., Nedwell, D., Beddow, J., Dumbrell, A., Mckew, B., Thorpe, E. & Whitby, C. 2015. amoA Gene Abundances and Nitrification Potential Rates Suggest that Benthic Ammonia-Oxidizing Bacteria and Not Archaea Dominate N Cycling in the Colne Estuary, United Kingdom. Applied and Environmental Microbiology, 81(1), 159–165. DOI: https://doi.org/10.1128/aem.02654-14
    » https://doi.org/10.1128/aem.02654-14
  • Liu, W.-C., Chen, W.-B. & Kimura, N. 2009. Measurement of Sediment Oxygen Demand to Simulate Dissolved Oxygen Distribution: Case Study in the Main Danshuei River Estuary. Environmental Engineering Science, 26(12), 1701–1711. DOI: https://doi.org/10.1089/ees.2009.0132
    » https://doi.org/10.1089/ees.2009.0132
  • Lorenzen, C. 1967. Determination of chlorophyll and pheo-pigments: spectrophotometric equations. Limnology and Oceanography, 12(2), 343–346. DOI: https://doi.org/10.4319/lo.1967.12.2.0343
    » https://doi.org/10.4319/lo.1967.12.2.0343
  • Machado, E. & Knoppers, B. 1988. Sediment oxygen consumption in an organic-rich, subtropical lagoon, Brazil. Science of The Total Environment, 75(2–3), 341–349. DOI: https://doi.org/10.1016/0048-9697(88)90045-9
    » https://doi.org/10.1016/0048-9697(88)90045-9
  • Mackin, J. & Swider, K. 1989. Organic matter decomposition pathways and oxygen consumption in coastal marine sediments. Journal of Marine Research, 47(3), 681–716. DOI: https://doi.org/10.1357/002224089785076154
    » https://doi.org/10.1357/002224089785076154
  • Mahiques, M., Burone, L., Figueira, R., Lavenére-Wanderley, A., Capellari, B., Rogacheski, C., Barroso, C., Samaritano Dos Santos, L., Cordero, L. & Cussioli, M. 2009. Anthropogenic influences in a lagoonal environment: a multiproxy approach at the valo grande mouth, Cananéia-Iguape system (SE Brazil). Brazilian Journal of Oceanography, 57(4), 325–337. DOI: https://doi.org/10.1590/s1679-87592009000400007
    » https://doi.org/10.1590/s1679-87592009000400007
  • Matos, C., Berrêdo, J., Machado, W., Metzger, E., Sanders, C., Faial, K. & Cohen, M. 2022. Seasonal changes in metal and nutrient fluxes across the sediment-water interface in tropical mangrove creeks in the Amazon region. Applied Geochemistry, 138, 105217. DOI: https://doi.org/10.1016/j.apgeochem.2022.105217
    » https://doi.org/10.1016/j.apgeochem.2022.105217
  • Miyao, S., Nishihara, L. & Sarti, C. 1986. Características físicas e químicas do sistema estuarino-lagunar de Cananéia-Iguape. Boletim Do Instituto Oceanográfico, 34, 23–26. DOI: https://doi.org/10.1590/s0373-55241986000100003
    » https://doi.org/10.1590/s0373-55241986000100003
  • Moraes, P. C., Zilius, M., Benelli, S. & Bartoli, M. 2018. Nitrification and denitrification in estuarine sediments with tube-dwelling benthic animals. Hydrobiologia, 819(1), 217–230. DOI: https://doi.org/10.1007/s10750-018-3639-3
    » https://doi.org/10.1007/s10750-018-3639-3
  • Nabeelah, B. S., Fawzi, M. M., Gokhan, Z., Rajesh, J., Nadeem, N., Kannan, R., Albuquerque, R. D. D. G. & Pandian, A. K. 2019. Ethnopharmacology, Phytochemistry, and Global Distribution of Mangroves―A Comprehensive Review. Marine Drugs, 17(4), 231. DOI: https://doi.org/10.3390/md17040231
    » https://doi.org/10.3390/md17040231
  • Nielsen, L. P., Brotas, V., Viaroli, P., Underwood, G., Nedwell, D. B., Sundbäck, K., Rysgaard, S., Miles, A., Bartoli, M., Dong, L., Thornton, D. C. O., Ottosen, L. D. M., Castaldelli, G. & Risgaard-Petersen, N. 2000. Protocol handbook for NICE – Nitrogen Cycling in Estuaries: a project under the EU research programme: Marine Science and Technology (MAST III). (Dalsgaard, T., ed.). Silkeborg: National EnvironmentalResearch Institute.
  • Niencheski, L. & Jahnke, R. A. 2002. Benthic respiration and inorganic nutrient fluxes in the estuarine region of Patos Lagoon (Brazil). Aquatic Geochemistry, 8(3), 135–152.
  • Pastor, L., Deflandre, B., Viollier, E., Cathalot, C., Metzger, E., Rabouille, C., Escoubeyrou, K., Lloret, E., Pruski, A., Vétion, G., Desmalades, M., Buscail, R. & Grémare, A. 2011. Influence of the organic matter composition on benthic oxygen demand in the Rhône River prodelta (NW Mediterranean Sea). Continental Shelf Research, 31(9), 1008–1019. DOI: https://doi.org/10.1016/j.csr.2011.03.007
    » https://doi.org/10.1016/j.csr.2011.03.007
  • Pecoraro, G., Hortellani, M., Hagiwara, Y., Braga, E., Sarkis, J. & Azevedo, J. 2019. Bioaccumulation of Total Mercury (THg) in Catfish (Siluriformes, Ariidae) with Different Sexual Maturity from Cananéia-Iguape Estuary, SP, Brazil. Bulletin of Environmental Contamination and Toxicology, 102(2), 175–179. DOI: https://doi.org/10.1007/s00128-018-2485-3
    » https://doi.org/10.1007/s00128-018-2485-3
  • Perina, F. & Abessa, D. 2020. Contamination and toxicity in a subtropical Estuarine Protected Area influenced by former mining activities. Ocean and Coastal Research, 68. DOI: https://doi.org/10.1590/s2675-28242020068313
    » https://doi.org/10.1590/s2675-28242020068313
  • Plante-Cuny, M. R. 1978. Pigments photosynthétiques et production primaire des fonds meubles néritiques d’une région tropicale (Nosy-Bé, Madagascar) (Travaux et documents de l’O.R.S.T.O.M. ; no 96). Provence: O.R.S.T.O.M.
  • Pradhan, U., Wu, Y., Shirodkar, P., Zhang, J. & Zhang, G. 2014. Sources and distribution of organic matter in thirty five tropical estuaries along the west coast of India-a preliminary assessment. Estuarine, Coastal and Shelf Science, 151, 21–33. DOI: https://doi.org/10.1016/j.ecss.2014.09.010
    » https://doi.org/10.1016/j.ecss.2014.09.010
  • Pratihary, A., Naik, R., Karapurkar, S., Gauthankar, M., Khandeparker, R., Manjrekar, S. & Gauns, M. 2021. Benthic exchange along a tropical estuarine salinity gradient during dry season: Biogeochemical and ecological implications. Journal of Sea Research, 177, 102124. DOI: https://doi.org/10.1016/j.seares.2021.102124
    » https://doi.org/10.1016/j.seares.2021.102124
  • Pratihary, A., Naqvi, S., Naik, H., Thorat, B., Narvenkar, G., Manjunatha, B. & Rao, V. 2009. Benthic fluxes in a tropical Estuary and their role in the ecosystem. Estuarine, Coastal and Shelf Science, 85(3), 387–398. DOI: https://doi.org/10.1016/j.ecss.2009.08.012
    » https://doi.org/10.1016/j.ecss.2009.08.012
  • Roden, E. & Edmonds, J. 1997. Phosphate mobilization in iron-rich anaerobic sediments: microbial Fe (III) oxide reduction versus iron-sulfide formation. Archiv Für Hydrobiologie, 139(3), 347–378. DOI: https://doi.org/10.1127/archiv-hydrobiol/139/1997/347
    » https://doi.org/10.1127/archiv-hydrobiol/139/1997/347
  • Rysgaard, Søren, Thastum, P., Dalsgaard, T., Christensen, P. B., Sloth, N. P. & Rysgaard, Soren. 1999. Effects of Salinity on NH 4+ Adsorption Capacity, Nitrification, and Denitrification in Danish Estuarine Sediments. Estuaries, 22(1), 21. DOI: https://doi.org/10.2307/1352923
    » https://doi.org/10.2307/1352923
  • Saito, R. T., Figueira, R. C. L., Tessler, M. G. & Cunha, I. I. L. 2001. Geochronology of sediments in the Cananeia-Iguape estuary and in southern continental shelf of São Paulo State, Brazil. Czechoslovak Journal of Physics, 53(S1), A75–A81. DOI: https://doi.org/10.1007/s10582-003-0012-0
    » https://doi.org/10.1007/s10582-003-0012-0
  • Sarma, V., Krishna, M., Prasad, V., Kumar, B., Naidu, S., Rao, G., Viswanadham, R., Sridevi, T., Kumar, P. & Reddy, N. 2014. Distribution and sources of particulate organic matter in the Indian monsoonal estuaries during monsoon. Journal of Geophysical Research: Biogeosciences, 119(11), 2095–2111. DOI: https://doi.org/10.1002/2014jg002721
    » https://doi.org/10.1002/2014jg002721
  • Schaeffer-Novelli, Y., De Souza Lima Mesquita, H., Cintrón-Molero, G. & Cintron-Molero, G. 1990. The Cananéia Lagoon Estuarine System, São Paulo, Brazil. Estuaries, 13(2), 193–203. DOI: https://doi.org/10.2307/1351589
    » https://doi.org/10.2307/1351589
  • Seitzinger, S., Gardner, W. & Spratt, A. 1991. The Effect of Salinity on Ammonium Sorption in Aquatic Sediments: Implications for Benthic Nutrient Recycling. Estuaries, 14(2), 167–174. DOI: https://doi.org/10.2307/1351690
    » https://doi.org/10.2307/1351690
  • Smith, J., Burford, M., Revill, A., Haese, R. & Fortune, J. 2012. Effect of nutrient loading on biogeochemical processes in tropical tidal creeks. Biogeochemistry, 108(1–3), 359–380. DOI: https://doi.org/10.1007/s10533-011-9605-z
    » https://doi.org/10.1007/s10533-011-9605-z
  • Sundbäck, K., Miles, A. & Göransson, E. 2000. Nitrogen fluxes, denitrification and the role of microphytobenthos in microtidal shallow-water sediments:an annual study. Marine Ecology Progress Series, 200, 59–76. DOI: https://doi.org/10.3354/meps200059
    » https://doi.org/10.3354/meps200059
  • Tessler, M. & Souza, L. 1998. Dinâmica sedimentar e feições sedimentares identificadas na superfície de fundo do sistema Cananéia-Iguape, SP. Revista Brasileira de Oceanografia, 46(1), 69–83. DOI: https://doi.org/10.1590/s1413-77391998000100006
    » https://doi.org/10.1590/s1413-77391998000100006
  • Thrush, S. F., Townsend, M., Hewitt, J. E., Davies, K., Lohrer, A. M., Lundquist, C. & Cartner, K. 2013. The many uses and values of estuarine ecosystems. In: Dymond, J. R. (ed.), Ecosystem services in New Zealand: conditions and trends (pp. 226–237). New Zealand: Manaaki Whenua Press.
  • Tréguer, P. & Le Corre, P. 1975. Manuel d’analyse des sels nutritifs dans l’eau de mer (utilisation de l’autoanalyzer II Technicon R). Brest: Université de Bretagne Occidentale.
  • Twilley, R., Rick, S., Bond, D. & Baker, J. 1999. Benthic nutrient fluxes in selected estuaries in the Gulf of Mexico. In: Bianchi, T. R. & Twilley, R. R. (eds.), Biogeochemistry of Gulf of Mexico estuaries (pp. 163-207). New York: Wiley.
  • Tyler, A., Mcglathery, K. & Anderson, I. 2003. Benthic algae control sediment-water column fluxes of organic and inorganic nitrogen compounds in a temperate lagoon. Limnology and Oceanography, 48(6), 2125–2137. DOI: https://doi.org/10.4319/lo.2003.48.6.2125
    » https://doi.org/10.4319/lo.2003.48.6.2125
  • Wang, S., Wang, W., Liu, L., Zhuang, L., Zhao, S., Su, Y., Li, Y., Wang, M., Wang, C., Xu, L. & Zhu, G. 2018. Microbial Nitrogen Cycle Hotspots in the Plant-Bed/Ditch System of a Constructed Wetland with N2O Mitigation. Environmental Science & Technology, 52(11), 6226–6236. DOI: https://doi.org/10.1021/acs.est.7b04925
    » https://doi.org/10.1021/acs.est.7b04925
  • Weston, N., Giblin, A., Banta, G., Hopkinson, C. & Tucker, J. 2010. The Effects of Varying Salinity on Ammonium Exchange in Estuarine Sediments of the Parker River, Massachusetts. Estuaries and Coasts, 33(4), 985–1003. DOI: https://doi.org/10.1007/s12237-010-9282-5
    » https://doi.org/10.1007/s12237-010-9282-5
  • Wilde, H. P. J. de & Bie, M. J. M. de. 2000. Nitrous oxide in the Schelde estuary: production by nitrification and emission to the atmosphere. Marine Chemistry, 69(3–4), 203–216. DOI: https://doi.org/10.1016/s0304-4203(99)00106-1
    » https://doi.org/10.1016/s0304-4203(99)00106-1
  • Willey, J. D. 1978. Release and uptake of dissolved silica in seawater by marine sediments. Marine Chemistry, 7(1), 53–65. DOI: https://doi.org/10.1016/0304-4203(78)90042-7
    » https://doi.org/10.1016/0304-4203(78)90042-7
  • Zhang, J., Gilbert, D., Gooday, A., Levin, L., Naqvi, W., Middelburg, J., Scranton, M., Ekau, W., Pena, A., Dewitte, B., Oguz, T., Monteiro, P., Urban, E., Rabalais, N., Ittekkot, V., Kemp, W., Ulloa, O., Elmgren, R., Escobar-Briones, E. & Van Der Plas, A. 2010. Natural and human-induced hypoxia and consequences for coastal areas: synthesis and future development. Biogeosciences, 7, 1443–1467. DOI: https://doi.org/10.5194/bgd-6-11035-2009
    » https://doi.org/10.5194/bgd-6-11035-2009
  • Zhang, L., Xiong, L., Zhang, J., Jiang, Z., Zhao, C., Wu, Y., Liu, S. & Huang, X. 2019. The benthic fluxes of nutrients and the potential influences of sediment on the eutrophication in Daya Bay, South China. Marine Pollution Bulletin, 149, 110540. DOI: https://doi.org/10.1016/j.marpolbul.2019.110540
    » https://doi.org/10.1016/j.marpolbul.2019.110540
  • Zhou, M., Butterbach-Bahl, K., Vereecken, H. & Brüggemann, N. 2017. A meta-analysis of soil salinization effects on nitrogen pools, cycles and fluxes in coastal ecosystems. Global Change Biology, 23(3), 1338–1352. DOI: https://doi.org/10.1111/gcb.13430
    » https://doi.org/10.1111/gcb.13430

Publication Dates

  • Publication in this collection
    04 Dec 2023
  • Date of issue
    2023

History

  • Received
    14 Oct 2022
  • Accepted
    09 May 2023
Instituto Oceanográfico da Universidade de São Paulo Praça do Oceanográfico 191, CEP: 05508-120, São Paulo, SP - Brasil, Tel.: (11) 3091-6501 - São Paulo - SP - Brazil
E-mail: diretoria.io@usp.br