SciELO - Scientific Electronic Library Online

 
vol.48Underwater vessel noise in a commercial and tourist bay complex in the Mexican Central PacificTrophic ecology of neonates and juveniles of the silky shark, Carcharhinus falciformis, off the coast of Guerrero, Mexico author indexsubject indexsearch form
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • Have no similar articlesSimilars in SciELO

Share


Ciencias marinas

Print version ISSN 0185-3880

Cienc. mar vol.48  Ensenada Jan./Dec. 2022  Epub Nov 17, 2023

https://doi.org/10.7773/cm.y2022.3265 

Articles

Factors determining the ocean-atmosphere CO2 flux variability in 5 coastal zones of the Gulf of California

Pedro Morales-Urbina1 
http://orcid.org/0000-0003-3414-0469

T Leticia Espinosa-Carreón1  * 
http://orcid.org/0000-0002-0003-7757

Saúl Álvarez-Borrego2 
http://orcid.org/0000-0002-7586-8678

José Martín Hernández-Ayón3 
http://orcid.org/0000-0001-6869-6225

Luz de Lourdes Aurora Coronado-Álvarez3 
http://orcid.org/0000-0001-5572-3247

Lorena Flores-Trejo1  3 
http://orcid.org/0000-0001-5918-9942

Cecilia Chapa-Balcorta4 
http://orcid.org/0000-0001-8305-0844

1Instituto Politécnico Nacional-Centro Interdisciplinario de Investigación para el Desarrollo Integral Regional, Unidad Sinaloa, 81101 Guasave, Sinaloa, Mexico.

2Independent researcher. Ensenada, Baja California, Mexico.

3Instituto de Investigaciones Oceanológicas, Universidad Autónoma de Baja California, 22860 Ensenada, Baja California, Mexico.

4Universidad del Mar, Puerto Ángel, Oaxaca, Mexico.


Abstract.

The Gulf of California (GC) features many oceanographic processes. It communicates with the Pacific Ocean via a surface water outflow (0-200 m) with relatively low dissolved inorganic carbon (DIC) values and a water inflow (200-600 m) with high DIC values. Data on the marine carbon system in the GC are scarce and most have been taken from the Midriff Islands region, in the central part of the gulf. We explored possible forcing agents that control the ocean-atmosphere CO2 flux (fCO2) variability in 5 coastal zones of the GC. We carried out 6 oceanographic cruises in 5 regions: off northern Sinaloa in September 2016 (NAV2016) and in March 2017 (NAV2017), in the Guaymas Basin (central gulf) in September 2016 (GUA2016), in Concepción Bay (Baja California Sur) in July 2017 (BC2017), in Mulegé (Baja California Sur) in July 2017 (MUL2017), and off Mazatlán (southern gulf) in July 2017 (MAZ2017). We measured temperature, salinity, DIC, and total alkalinity and calculated the surface water partial pressure of CO2 and fCO2. We also used sea surface height anomaly with geostrophic flow, sea surface temperature, and chlorophyll concentration data from satellite imagery to generate composites for the sampling days. The lowest temperature, highest DIC, and negative fCO2 were registered in NAV2017. NAV2016, GUA2016, and BC2017 showed the highest temperatures; and MUL2017 and MAZ2017, intermediate temperatures. The most contrasting fCO2 values occurred in GUA2017 (0.56 ± 0.46 mmol C·m-2·d-1) and MAZ2017 (-2.26 ± 1.85 mmol C·m-2·d-1). In general, fCO2 is determined by the oceanographic conditions of each study area.

Key words: CO2 fluxes; upwelling; forcing agents; coastal regions; Gulf of California

Resumen.

El golfo de California (GC) presenta diversos procesos oceanográficos. Tiene comunicación con el océano Pacífico mediante un flujo de salida de agua superficial (0-200 m) con valores relativamente bajos de carbono inorgánico disuelto (CID) y un flujo de entrada de agua (200-600 m) con valores altos de CID. Los datos sobre el sistema de carbono marino en el GC son escasos, y la mayoría proviene de la región de las islas grandes, en el centro del golfo. Se exploraron los posibles agentes forzantes que controlan la variabilidad del flujo de CO2 océano-atmósfera (fCO2) en 5 zonas costeras del GC. Se realizaron 6 cruceros oceanográficos en 5 regiones: frente al norte de Sinaloa en septiembre de 2016 (NAV2016) y marzo de 2017 (NAV2017), en la cuenca de Guaymas (centro del golfo) en septiembre de 2016 (GUA2016), en bahía Concepción (Baja California Sur) en julio de 2017 (BC2017), en Mulegé (Baja California Sur) en julio de 2017 (MUL2017) y frente a Mazatlán (golfo sur) en julio de 2017 (MAZ2017). Se midió la temperatura y la salinidad, se estimó el CID y la alcalinidad total y se calculó la presión parcial de CO2 superficial y el fCO2. Se utilizaron imágenes de satélite para generar compuestos de la anomalía del nivel del mar con flujo geostrófico, la temperatura superficial del mar y la concentración de clorofila en los días de muestreo. La temperatura más baja, el CID más alto y el fCO2 negativo se registraron en NAV2017. NAV2016, GUA2016 y BC2017 registraron las temperaturas más altas, y MUL2017 y MAZ2017, temperaturas intermedias. Los mayores contrastes de fCO2 ocurrieron en GUA2017 (0.56 ± 0.46 mmol C·m-2·d-1) y MAZ2017 (-2.26 ± 1.85 mmol C·m-2·d-1). En general, el fCO2 está determinado por las condiciones oceanográficas de cada zona de estudio.

Palabras clave: flujos de CO2; surgencia; agentes forzantes; regiones costeras; golfo de California

INTRODUCTION

Recent studies have focused on the role coastal zones play in the CO2 flux ocean-atmosphere (fCO2), given how scarce information is for these zones (Laruelle et al. 2014, Jiménez-López et al. 2019). Coastal zones play a considerable role in the global carbon cycle due to high primary production and coastal processes such as upwelling (Gattuso et al. 1998). The fCO2 depends on complex interactions between heating/cooling of the water column, photosynthesis, respiration, wind mixing, internal waves, tidal processes, advection, and thermocline vertical shifts (Takahashi et al. 2002, Wanninkhof 2014, Coronado-Álvarez et al. 2017). In general, these interactions have been classified into thermal and non-thermal effects (these represent biological activity and include changes in alkalinity, advection, and sea-air exchange) (Takahashi et al. 2002).

The Gulf of California (GC) has a net vertical exchange of water with the adjacent Pacific Ocean, with an outflow of surface water toward the Pacific (0-200 m) and an inflow of subsurface water (200-600 m) (Marione 2003). The surface waters that outflow from the GC to the Pacific are relatively poor in dissolved inorganic carbon (DIC), and the water that enters at depth has higher DIC concentrations (Rodríguez-Ibáñez et al. 2013). Within the eastern Pacific area with very low dissolved oxygen values, the GC is located in the minimum dissolved oxygen zone (Levin 2002, Paulmier and Ruiz-Pino 2009). The extremely low dissolved oxygen values cause denitrifying bacteria respiration, which results in a nitrate deficit with respect to the Redfield ratios for NO3-PO4 (Thomas 1966) and NO3-DIC (Rodríguez-Ibáñez et al. 2013). Due to this, and to the dissolution of calcium carbonate exoskeletons at depth (Peterson 1966), GC upwelling waters have an excess of DIC relative to nitrate concentrations; furthermore, when photosynthesis proceeds and nitrate is consumed, excess DIC could cause CO2 to leak into the atmosphere.

Data on the partial pressure of CO2 in surface waters (pCO2W) of the GC are scarce and most come from the region of the Midriff Islands (Tiburón and Ángel de la Guarda) (Hernández-Ayón et al. 2013). The GC has intense upwelling events on the eastern coast in winter (December to May) (Álvarez-Borrego 2010) and very weak upwelling events on the western coast in summer (July to October) (Santamaría-del-Ángel et al. 1999). Heating of upwelling water (with high values of pCO2W) decreases the solubility of gases and contributes to producing CO2 release into the atmosphere (Coronado-Álvarez et al. 2017). This implies that upwelling areas in the GC are sources of CO2 to the atmosphere, at least for long-term averages, for example, the average of 1 year. However, fCO2 can revert to negative values in upwelling relaxation events due to the effect of primary production. This study presents the variability of fCO2 in 5 coastal zones of the GC and analyzes the influence of controlling processes on fCO2 and the emission and absorption capacity of CO2 in these zones.

MATERIALS AND METHODS

Oceanographic campaigns

Oceanographic campaigns were carried out off the Mexican coast between September 2016 and July 2017. In September 2016, 2 samplings were done aboard the R/V Altaír, which belongs to the Secretaría de Marina Armada de México (SEMAR); the first was done in the marine-coastal zone of Navachiste (off the north of Sinaloa) (NAV2016) and the second in the coastal zone off Guaymas (Sonora) (GUA2016). In March 2017, samplings were carried out in the marine-coastal zone of Navachiste (NAV2017) on board a boat with an outboard motor. In June 2017, 3 areas were sampled aboard the R/V Río Tecolutla of the SEMAR: off Mulegé (MUL2017), in Concepción Bay (BC2017), and off Mazatlán (MAZ2017) (Fig. 1).

Figure 1 Study area in the Gulf of California. Sampled subregions are shown in black squares: Navachiste (a), Guaymas (b), Concepción Bay (c), Mulegé (d), Mazatlán (e), Midriff Islands (f). 

Temperature and salinity were measured using a Sea-Bird 19 Plus CTD, equipped with a 5-L Niskin bottle rosette. Discrete samples of surface water were collected to analyze pH, DIC, and total alkalinity (TA) in 350-mL sodium borosilicate bottles with 100 µL of a saturated solution of mercuric chloride used as a preservative to stop biological activity. Samples were taken at 8 stations at NAV2016, 5 at GUA2016, 9 at NAV2017, 4 at BC2017, 9 at MUL2017, and 7 at MAZ2017. The bottles were sealed with Apiezón grease to prevent evaporation and contact with the atmosphere.

Laboratory analysis

We used an Apollo SciTech AS-C3 instrument to analyze the DIC. We used 6% phosphoric acid to convert all the DIC to CO2, and we quantified this with a Li-COR 7000 high-precision infrared analyzer (Dickson and Goyet 1994). TA was analyzed with an Apollo SciTech AS-ALK2 titration system and 0.1 M hydrochloric acid solution (the 3 methods are detailed in Hernández-Ayón 1995). We verified the precision and accuracy of the methods and took care not to exceed 5 µmol·kg‒1 (0.25% error) in the difference of DIC and TA with respect to reference values. The reference substandards were developed at the Instituto de Investigaciones Oceanológicas of the Universidad Autónoma de Baja California (Mexico).

The calculation of pCO2W was performed with the CO2SYS program (Lewis and Wallace 1998) using surface data of DIC, TA, temperature, salinity, and the dissociation constants of Lueker et al. (2000). We estimated fCO2 using the Liss and Merlivat (1986) equation: F = KwK0 (∆pCO2), where Kw is the CO2 transfer coefficient as a function of wind speed at 10 m above sea level and K0 is the solubility coefficient of CO2 as a function of temperature and salinity. The gas transfer coefficient of Ho et al. (2006) and the CO2 solubility coefficient of Weiss (1974) were used. The partial pressure of CO2 in the atmosphere (pCO2A) was obtained from the internet portal of the Earth System Research Laboratories (ESRL) of the National Oceanic and Atmospheric Administration (NOAA, USA) for each sampling day, and the values corresponded to the Mauna Loa station, Hawaii (NOAA 2017). Due to failures in the ship’s anemometer, the wind data for the different areas were obtained from the MeteoEarth electronic portal (http://www.meteoearth.com), from the model of the European Center for Medium-Range Weather Forecasts; this model was used because the spectral space is larger than the physical parameterizations in a reduced grid (ECMWF 2017).

To better understand the variation of the sea surface temperature (SST) and chlorophyll concentration (Chlsat) in the GC during the different cruises, 5-day satellite averages of the Aqua-Modis sensor were used for the approximate dates of each campaign. For SST, daytime data were used with radiance of 11 µm wavelength and pixel size of 1.1 × 1.1 km2. For Chlsat, images with processing level 3 and the same pixel size as those of the SST were used. In addition, daily sea surface height anomaly (SSHa) images were obtained, with spatial resolution of 0.25° and L4 level data, produced by AVISO and CMEMS. Geostrophic velocity was calculated according to Pond and Pickard (2013). All images were processed in Matlab.

Statistical analysis and data visualization

Principal component analysis (PCA) was performed to determine which factors influenced the variability of pCO2W in the different areas of the GC. Beforehand, the corresponding analysis of variance was performed to determine the normality of the data; when they were not normal, the PCA was carried out. Data of the SST, salinity, DIC, TA, pCO2W, pCO2A, and Chlsat were normalized; the Broken-stick criterion was used according to Peres-Neto et al. (2003), and the correlation and significance of all variables were obtained. All of the above was done with the help of the Statistica software.

Maps of the salinity, DIC, pCO2W, and fCO2 were made with the free distribution Ocean Data View software v.5.0 (Schlitzer 2016). Map interpolations were done using the data-interpolating variational analysis (DIVA) of the Ocean Data View to ensure the conservation of structures and better visualize results.

RESULTS

Oceanographic conditions inferred from satellite images

Satellite composites of SSHa, SST, and Chlsat showed contrasting conditions in September 2016 and March 2017, and transition conditions in June 2017 (Fig. 2). In September 2016 anticyclonic eddies were recorded along the GC. In NAV2016, geostrophic flow (~16 cm·s-1) towards the south was observed; and in GUA2016, geostrophic flow (~18 cm·s-1) towards the coast was observed as part of a small anticyclonic eddy (~90 km) with a SSHa of ~24 cm (Fig. 2a). Negative anomalies of the sea level were observed in March 2017 in most of the GC, with the exception of the zone adjacent to La Paz Bay, and an anticyclonic eddy was recorded at the entrance of the gulf (~17 cm·s-1 and height of 15 cm). Part of a cyclonic eddy (-10 cm) was observed in NAV2017 (Fig. 2b). In June 2017, an intense water inflow occurred in the eastern region of the GC and an intensification of the current with a high sea level (>25 cm·s-1 and height of 25 cm) was observed in Guaymas, which extended to the region of the Midriff Islands (>30 cm·s-1 and height of 22-36 cm). In MUL2017, we observed the influence of a cyclonic eddy with flow towards the coast in the northern part (~12 cm·s-1). In MAZ2017, we recorded the entry of water into the GC with a speed of ~25 cm·s-1 and a cyclonic circulation at the mouth of the GC (Fig. 2c).

Figure 2 Sea surface height anomalies (SSHa) and geostrophic flow (cm·s-1) (a-c), sea surface temperature (SST, d-f), and chlorophyll (Chlsat, g-i) in September 2016 (squares: red, NAV2016; blue, GUA2016), March 2017 (red square, NAV2017), and June 2017 (squares: red, MAZ2017; blue, MUL2017). 

In September 2016, in practically the entire GC, SST was >29 °C (Fig. 2d) and Chlsat values were <0.5 mg·m-3. In NAV2016, a coastal strip with Chlsat values >1.5 mg·m-3 was observed; in GUA2016, lower values (~0.8 mg·m-3) were observed (Fig. 2g). In March 2017, SST values were lower than in September throughout the gulf, with differences greater than 8 °C, and Chlsat was up to ˃1 mg·m-3 higher than in September, with different mesoscale structures in concordance with SSHa data. In NAV2017, SST values were <22 °C and Chlsat was >1.5 mg·m-3 (Fig. 2e, h). Due to the patterns recorded in June, this month was considered a transition period, with intermediate values of SST and Chlsat between September and March. MUL2017 and MAZ2017 showed SST values <29 °C and Chlsat values >0.5 mg·m-3 (Fig. 2f, i).

Variability of salinity, dissolved inorganic carbon, pCO2W, and fCO2

Salinity distribution showed that NAV2016 (Fig. 3a) had lower values (34.70 ± 0.21, mean and standard deviation, respectively, here and henceforth) than NAV2017 (35.29 ± 0.29) (Fig. 3c). There was little variation (35.14 ± 0.01) in GUA2016 (Fig. 3b). Homogeneous values were also recorded (35.94 ± 0.13) in BC2017 (Fig. 3d). Values were 35.28 ± 0.13 in MUL2017 (Fig. 3e), and the central region had the maximum values. MAZ2017 also had little variation in salinity (35.02 ± 0.05) (Fig. 3f).

Figure 3 Horizontal distribution of salinity and dissolved inorganic carbon (DIC) for the different subregions: NAV2016 (a, g), GUA2016 (b, h), NAV2017 (c, i), BC2017 (d, j), MUL2017 (e, k), and MAZ2017 (f, l). The figures were confectioned using Ocean Data View. 

DIC and pCO2W values showed contrasting conditions in NAV2016 (1,994 ± 3.52 µmol·kg-1 and 428 ± 12.21 µatm, respectively; Figs. 3g, 4a) and NAV2017 (2,076 ± 25.08 µmol·kg-1 and 363 ± 46.94 µatm, respectively, Figs. 3i, 4c). In NAV2017, variations of DIC and pCO2W were similar; values of both properties were relatively low at the northernmost station (~2,070 µmol·kg-1 and ~315 µatm, respectively), increased in the central area (~2,120 µmol·kg-1 and ~420 µatm), and decreased towards the south (~2,080 µmol·kg-1 and ~330 µatm). GUA2016 showed little variation of DIC and pCO2W, although with higher values towards the south of the area (2,015 ± 16.22 µmol·kg-1 and 434 ± 29.06 µatm, respectively) (Figs. 3h, 4b). In BC2017, little variation of DIC and slightly more marked variation of pCO2W were recorded (2,017 ± 14.69 µmol·kg-1 and 377 ± 29.45 µatm, respectively) (Figs. 3j, 4d). In MUL2017 there was greater variability; DIC had low values towards the central zone in the open sea and high values towards the edges (2,058 ± 15.05 µmol·kg-1), and pCO2W had higher values to the north of the study area (414 ± 37.79 µatm) (Figs. 3k, 4e). DIC and pCO2W had the lowest values (1,934 ± 5.94 µmol·kg-1 and 305 ± 12.64 µatm, respectively) in MAZ2017, although pCO2W increased in the center of the area (Figs. 3l, 4f). In general, DIC and pCO2W had the highest average values in NAV2017 and GUA2016, respectively, whereas both variables had the lowest values in MAZ2017. All the above information is summarized in Table 1.

Table 1 Mean values (standard deviation) of salinity, dissolved inorganic carbon (DIC), seawater CO2 partial pressure (pCO2W), and ocean-atmosphere CO2 flux (fCO2) during the 6 cruises. 

Location Salinity DIC (mmol·kg-1) pCO2W (μatm) fCO2 (mmol C·m-2·d-1)
NAV2016 34.70 (0.21) 1,994 (3.52) 428 (12.21) 0.38 (0.52)
GUA2016 35.14 (0.01) 2,015 (16.22) 434 (29.06) 0.56 (0.46)
NAV2017 35.29 (0.29) 2,076 (25.08) 363 (46.94) -0.40 (0.39)
BC2017 35.94 (0.13) 2,017 (14.69) 377 (29.45) -0.90 (0.87)
MUL2017 35.28 (0.13) 2,058 (15.05) 414 (37.79) -0.05 (0.74)
MAZ2017 35.02 (0.05) 1,934 (5.94) 305 (12.64) -2.26 (1.85)

Figure 4 Horizontal distribution of seawater CO2 partial pressure (pCO2W) and CO2 flux variability (fCO2) for the different subregions: NAV2016 (a, g), GUA2016 (b, h), NAV2017 (c, i), BC2017 (d, j), MUL2017 (e, k), and MAZ2017 (f, l). The figures were confectioned using Ocean Data View. 

Four of the 5 regions studied behaved as CO2 sinks in March and June 2017: NAV2017, BC2017, MUL2017, and MAZ2017. NAV2016 behaved as a light source of CO2 into the atmosphere; the average fCO2 there was 0.38 ± 0.52 mmol C·m-2·d-1, and values ranged from practically in equilibrium with the atmosphere (0.03 mmol C·m-2·d-1) to 1.57 mmol C·m-2·d-1 (Fig. 4g). In GUA2016, the average fCO2 was 0.56 ± 0.46 mmol C·m-2·d-1; the spatial pattern showed equilibrium values at the most coastal station and maximum values of 1.14 mmol C·m-2·d-1 at the most oceanic station (Fig. 4h). The positive values for September 2016 indicate a slight flow from the sea to the atmosphere. NAV2017 behaved differently; 78% of the area had negative fCO2 values, with an average of -0.40 ± 0.39 mmol C·m-2·d-1 and a range of values from -1.03 to 0.39 mmol C·m-2·d-1 (Fig. 4i). BC2017 trended to larger negative values towards the inner bay, with an average CO2 flux of -0.90 ± 0.87 mmol C·m-2·d-1 (Fig. 4j). MUL2017 showed positive values to the north and negative values to the south, with an average of -0.05 ± 0.74 mmol C·m-2·d-1 and a range of values from -1.01 to 1.37 mmol C·m-2·d-1 (Fig. 4k). The entire area of MAZ2017 showed negative values, with an average of -2.26 ± 1.85 mmol C·m-2·d-1 and a range of values from -4.99 to -0.42 mmol C·m-2·d-1 (Fig. 4l).

Statistics

The PCA showed an explained variance of 34% for the first component and 30% for the second (Fig. 5). In the first component, the variables with the greatest weight were DIC, TA, and SST (0.95, 0.80, and -0.74, respectively), and the one with the least contribution was Chlsat (0.57). In the second component, the variables with the greatest weight were pCO2W and pCO2A (0.95 and -0.74, respectively). The zones that were centered throughout the first component were NAV2016, GUA2016, and some MUL2017 stations. NAV2017 and MAZ2017 showed an opposite behavior (Fig. 5). Due to the amount of data recorded per zone, it was difficult to indicate that the weight of the variables suggested that in the primary processes of the carbonate system in NAV2016 and GUA2017 thermal effects predominate and that biological ones do so to a lesser degree. The PCA suggested that biological effects were stronger than physical effects at NAV2017, whereas at MUL2017, BC2017, and MAZ2017, both effects may have contributed similarly.

Figure 5 Principal component analysis (PCA) for Navachiste (NAV2016, NAV2017), Guaymas (GUA2016), Concepción Bay (BC2017), Mulege (MUL2017), and Mazatlán (MAZ2017). 

DISCUSSION

Oceanographic conditions

The circulation pattern in the GC is determined by the presence of eddies at the mouth. In September 2016 (summer, Fig. 2a), we recorded a cyclonic eddy at the mouth, the inflow of water from the western part of the gulf, the presence of anticyclonic eddies, and the outflow of water in the Sinaloa coast. The aforementioned inflow and outflow of water from the gulf coincides with that reported by Portela et al. (2016), but for spring; for summer, they indicated that the water entered through the eastern coast (Sinaloa) and outflowed through the western coast (Baja California). Beier (1997) reported a cyclonic circulation in the GC in summer, which does not coincide with the pattern recorded in this study. The anticyclonic eddies recorded in Guaymas and Sinaloa coincide with that reported by Pegau et al. (2002) from a time series from 1997 to 2001. In March 2017 (winter, Fig. 2b), an anticyclonic eddy appeared at the mouth of the GC, with negative sea level anomalies in practically the entire GC that, despite the geostrophic velocities, were small; in addition, we observed a flow towards the south of the gulf on the peninsular coast, but the influence of the anticyclonic eddy that crossed towards Sinaloa suggests a limited circulation towards the inner gulf. Beier (1997) and Castro et al. (2017) reported an anticyclonic circulation in winter, whereas Castro et al. (2000) reported, in their 1992-1998 study, a cyclonic circulation between winter and spring in the central part of the gulf. This difference in circulation patterns has been reported by Castro et al. (2017), who concluded that mesoscale eddies can modify the seasonal exchange of water at the entrance of the GC. In June 2017 (Fig. 2c) a geostrophic flow towards the inner gulf was recorded off the coast of Sinaloa, indicative of a cyclonic circulation, which coincides with Beier (1997) and Portela et al. (2016). Therefore, this study suggests that the circulation pattern was anticyclonic in the GC in September 2016 (summer) and cyclonic in June 2017 (transition).

The climatic conditions in the GC were contrasting in September 2016 and March 2017, and transitional in June 2017, which coincides with that reported by Roden (1964). September is considered one of the warmest months and March one of the coldest (Soto-Mardones et al. 1999). In September 2016 (Fig. 2d, g), high temperatures (˃30 °C) were associated with low chlorophyll concentrations (˂0.1 mg·m‒3) near the eastern coast, which coincides with what was reported by Santamaría-del-Angel et al. (1999). For March 2017 (Fig. 2e, h), low SSTs (~22 °C) were associated with high Chlsat concentrations (˃2 mg·m‒3) in the eastern coast of the gulf, due to the effect of coastal upwelling in winter. In particular, comparisons of NAV2016 with NAV2017 suggested the presence of coastal upwelling in March and its absence in September (Fig. 2d, e, g, h). In June 2017 (Fig. 2f, i), in the central and southern regions of the GC, SSTs were higher by the mainland than by the peninsula, which coincides with Santamaría-del-Ángel et al. (1999) and Soto-Mardones et al. (1999), mainly for summer (~1 °C difference).

Variability of salinity, dissolved inorganic carbon, pCO2W, and fCO2

The lowest average salinity was recorded in NAV2016, and the highest salinities were recorded in NAV2017 and BC2017 (Fig. 3a, c, d). In GUA2016 (Fig. 3b), surface salinity values >35 indicated the presence of the Gulf of California Water (GCW), which had already been described by Delgadillo-Hinojosa et al. (2001). In NAV2016 (Fig. 3a), values <35 were a result of the presence of transition water and GCW (Portela et al. 2016). The salinity recorded in GUA2016 (Fig. 3b) represents, according to Bray (1988), the general circulation of the GC. In NAV2017 (Fig. 3c), in agreement with what was reported by Castro et al. (2000), salinity values were a combined product of coastal upwelling water and GCW flowing south into the gulf. Concepción Bay (Fig. 3d) is a semi-enclosed bay where solar radiation in summer could have caused water evaporation and increased salinity. In MAZ2017 (Fig. 3f), salinity was homogeneous throughout the area, which suggests that transitional water entered the gulf (Portela et al. 2016). Therefore, in the sampling months (September 2016, March and June 2017), according to salinity records, GCW and transitional water were located at the surface.

In NAV2016, we recorded the highest temperature and Chlsat ( x- = 29.94 °C and x- = 1.23 mg·m-3, respectively), and pCO2W ( x- = 428 μatm) indicated a supersaturation of CO2 and a positive fCO2 ( x- = 0.38 mmol C·m-2·d-1). In GUA2016, we also recorded a SST of x- = 29.94 °C, but low concentrations of Chlsat ( x- = 0.23 mg·m-3), a pCO2W ( x- = 434 μatm) above the average for the pCO2A (402 μatm), and a positive fCO2 of x- = 0.56 mmol C·m-2·d-1. In both areas there were summer conditions, anticyclonic eddies, CO2 supersaturation, and a slight CO2 release into the atmosphere, with pCO2 values above equilibrium and those reported by Rodríguez-Ibáñez et al. (2013) for summer, which indicated that the GC is almost in equilibrium with the atmosphere. The water mass that predominates in summer is GCW (Álvarez-Borrego and Schwartzlose 1979). Portela et al. (2016) reported, for this same period, the presence of Superficial Tropical Water and GCW, which suggests that both masses of water are present in NAV2016 and GUA2016. The contrast in the concentration of Chlsat between both zones could be explained by differences in the intensity of grazing by zooplankton; it is important to consider that Chlsat concentrations only reflect the first optical depth and that, due to the anticyclonic eddy, the subsurface maximum of chlorophyll of GUA2016 could be below this optical depth, and thus cause low concentrations to be observed in satellite images. This behavior was recorded by Espinosa-Carreón et al. (2012) for the southern region of the California Current. Flores-Trejo et al. (2019) reported surface DIC values of 2,055 μmol·kg-1 associated with the GCW for Punta Lobos, Sonora (south of Guaymas), in October 2018, which were slightly higher than those reported in this work for the GUA2016 area (2,015 μmol·kg-1). The ocean is so dynamic that the fCO2 can vary over the course of days. Morales-Urbina et al. (2017) reported data from an oceanographic buoy anchored in the NAV2016 area during the summer of 2016, and their average fCO2 value for 10 days was 2.70 mmol C·m‒2·d‒1, higher than the average fCO2 for NAV2016 recorded in the present study, which suggests an increase in the contribution of CO2 to the atmosphere.

The highest DIC concentration was recorded in NAV2017 ( x- = 2,076 μmol·kg-1), which suggests that upwelling conditions could have occurred. However, based on the values obtained for pH ( x- = 8.08, data not shown) and fCO2 ( x- = -0.40 mmol C·m-2·d-1), said event was in a relaxation phase. The negative fCO2 indicated that the GC behaved as a sink, and the Chlsat values ( x- = 1.09 mg·m-3) indicated that the phytoplankton was increasing and consuming DIC. Coronado-Álvarez et al. (2017) reported similar conditions for the upwelling area off the northwest of the Baja California Peninsula, and they indicated that when upwelling relaxes, nutrients are used more effectively, which causes DIC and pCO2W to decrease; changes in these variables were associated with factors ranging from semidiurnal variabilities (sea breezes), upwelling events, to interannual variabilities (El Niño/Southern Oscillation). DIC values found in this study were lower than those reported by Linacre et al. (2010) (~2,215 μmol·kg-1) for the upwelling zone off the western coast of Baja California at a depth of ~80 m.

In BC2017, fCO2 values ranged from -2.15 to -0.16 mmol C·m‒2·d‒1 (Fig. 4j). Chlsat values (data not shown) indicated that much of the area was mesotrophic to eutrophic, with possible high photosynthetic activity in the days prior to sampling, which could have caused the negative fCO2 values. However, there are spatiotemporal variations between the absorption and contribution of CO2 in the study area that should be considered in subsequent research to describe the variability of fCO2.

In the MUL2017 region, the average DIC concentration was 2,058 μmol·kg-1, lower than that reported by Hernández-Ayón et al. (2013), who indicated that there was a relationship between the variables of the carbonate system and the water masses in the region of the Midriff Islands and reported values of ~2,080 μmol·kg-1 for the summer of 2004. The spatial variability of DIC in this region could be due to the cyclonic eddy found in this area, with differences in the intensity of the geostrophic current (Fig. 2c); this eddy could have influenced the pCO2W values to be below equilibrium to the south of the area and to be positive to the north of the area, probably due to the shearing of the water, which promotes the release of CO2 into the atmosphere. Therefore, the fCO2 value of -0.05 mmol C·m-2·d-1 (Fig. 4k) was probably due to combined physical and biological effects.

The flow of water into the GC along the eastern coast in MAZ2017 carried properties that resulted in the lowest concentration records of Chlsat, DIC, pCO2W, and fCO2 ( x- = 0.18 mg·m-3, 1,934 μmol·kg-1, 305 μatm, -2.26 mmol C·m-2·d-1, respectively) (Fig. 2c). Franco et al. (2014) reported, for the area south of Cabo Corrientes in August 2010, DIC concentrations of ~1,980 μmol·kg-1 and minimum fCO2 values of -4.4 mmol C·m-2·d-1, values which are higher than those of the present study for MAZ2017. Trucco-Pignata et al. (2019) reported, for June 2015, transition water at the entrance of the GC that traveled at the surface (0-115 m) with a DIC concentration of ~2,050 μmol·kg-1 (higher than that reported in the present study). Therefore, in MAZ2017, the advection of water and its productivity determined that the region can act as a source or sink of CO2 (Franco et al. 2014).

Correlations

The PCA grouped, on the one hand, NAV2016 and GUA2016, where anticyclonic eddies predominated and the variables were similar, except Chlsat. On the other hand, it grouped NAV2017 and MAZ2017, both with the lowest fCO2 values; in the first, the negative sea level anomaly predominated, and in the second, the coastal flow on the eastern coast. Therefore, the fCO2 is determined by the oceanographic conditions of the area, which acts as a source or sink at different times.

It is important to consider that the GC is part of the great oceanographic dynamics of the eastern Pacific. According to Rodríguez-Ibáñez et al. (2013), the net contribution of nutrients and DIC is not transported homogeneously throughout the GC, due to regional differences in physical dynamics, which promotes variations in CO2 fluxes. These authors reported that, in a one-year average, the GC was a light emitter. In the present study, the average of the samplings carried out in 2016 and 2017 in the 5 coastal zones was -0.42 mol C·m-2·d-1, which suggests that the GC behaved as a light CO2 sink. Laruelle et al. (2014) classified GC as a CO2 emitter (0-1 mol C·m-2·y-1). In regions dominated by coastal upwelling on the eastern edges, when there are changes in coastal currents, and these occur in short time scales, the physical and biological processes are modified, as well as the ocean-atmosphere CO2 equilibrium conditions (Cai et al. 2020). Concentrations of DIC, pCO2W, and, therefore, the fCO2 are associated with the oceanographic dynamics of the area and, among others, with the presence of mesoscale processes, coastal upwelling, advection, and biological processes. Thus, oceanographic conditions in the GC (such as cyclonic eddies, anticyclonic eddies, and coastal upwelling) and its biology, which vary throughout it and in time, influenced changes in pCO2W and fCO2. Based on the information obtained in this study in the 5 zones, we describe possible responses to understand the mechanisms by which a positive or negative ocean-atmosphere CO2 exchange occurs; nevertheless, substantial efforts are needed for more information at the spatial and temporal level to further the understanding of CO2 flux dynamics in the GC.

Coronado-Álvarez et al. (2017) created time series of SST, salinity, and pCO2W with high-frequency point data (every 3 h) recorded by a MAPCO2 system (for more details see Sutton et al. 2014) anchored to an oceanographic buoy. With the Observatorio del Monitoreo Costero (Coastal Monitoring Observatory), these authors obtained a 7-year time series, off the northwest of Baja California, in an area of intense upwelling. Coronado-Álvarez et al. (2017) observed significant changes in these properties and in fCO2 that lasted for periods as short as hours and as long as interannual segments. Their spectral analysis showed significant semidiurnal, diurnal, ~15-d, and ~28-d variance components. Their description is very limited because they only used data for one geographic point, but their results indicate that those of the present study should be taken as preliminary results because they are instantaneous for each cruise. Continued efforts are needed to understand and more appropriately piece together the puzzle of the role GC plays in fCO2.

ACKNOWLEDGMENTS

This study was financed by the Instituto Politécnico Nacional (IPN, Mexico) with the support of projects SIP20164820, 20170983, 20195181, and 2020716. PMU and LFT had a postgraduate scholarship from the Consejo Nacional de Ciencia y Tecnología (CONACYT, Mexico) and a BEIFI scholarship from the IPN. We thank the Secretaría de Marina de México and the crews of the R/V Altair and Río Tecolutla for their support in sample taking. Composite satellite images of SST and Chlsat were provided by M Kahru of Scripps-University of California in San Diego and Copernicus Marine Service. CCB provided support through the Semarnat-2016-C01-278637 project. We thank the anonymous reviewers for their valuable comments and suggestions.

REFERENCES

Álvarez-Borrego, S. 2010. Physical, Chemical and Biological Oceanography of the Gulf of California. In: Brusca, R. (ed.), The Gulf of California: Biodiversity and Conservation. Arizona-Sonora Desert Museum Studies in Natural History. Tucson (AZ): The University of Arizona Press and ASDM. p. 24-48. [ Links ]

Álvarez-Borrego, S., Schwartzlose, R.A. 1979. Masas de agua del Golfo de California = Water masses of the Gulf of California. Cienc Mar. 6(1):43-63. http://dx.doi.org/10.7773/cm.v6i1.350 [ Links ]

Beier, E. 1997. A numerical investigation of the annual variability in the Gulf of California. J Phys Ocean. 27(5):615-632. https://doi.org/10.1175/1520-0485(1997)027<0615:ANIOTA>2.0.CO;2 [ Links ]

Bray, N.A. 1988. Water mass formation in the Gulf of California. J Geophys Res. 93(C8):9223-9240. https://doi.org/10.1029/jc093ic08p09223 [ Links ]

Cai, W.J., Xu, Y.Y., Feely, R.A., Wanninkhof, R., Jönsson, B., Alin, S.R., Barbero, L., Coross, J.N., Azetsu-Scott, K., Fassbender, A., et al. 2020. Controls on surface water carbonate chemistry along North American ocean margins. Nature. 11(1):2691. https://doi.org/10.1038/s41467-020-16530-z [ Links ]

Castro, R., Collins, C.A., Rago, T.A., Margolina, T., Navarro-Olache, L.F. 2017. Currents, transport, and thermohaline variability at the entrance to the Gulf of California (19-21 April 2013) = Corrientes, transportes y variabilidad termohalina en la entrada al golfo de California (19-21 de abril de 2013). Cienc Mar. 43(3):173-190. https://doi.org/10.7773/cm.v43i3.2771 [ Links ]

Castro, R., Mascarenhas, A.S., Durazo, R., Collins, C.A. 2000. Variación estacional de la temperatura y salinidad en la entrada del golfo de California, México = Seasonal variation of the temperature and salinity at the entrance to the Gulf of California, Mexico. Cienc Mar. 26(4):561-583. http://dx.doi.org/10.7773/cm.v26i4.621 [ Links ]

Coronado-Álvarez, L.L.A., Álvarez-Borrego, S., Lara-Lara, J.R., Solana-Arellano, E., Hernández-Ayón, J.M., Zirino, A. 2017. Temporal variations of water pCO2 and the air-water CO2 flux at a coastal location in the southern California Current System: diurnal to interannual scales = Variaciones temporales de pCO2 del agua y flujos de aire-agua de CO2 en una localidad costera en el sur del Sistema de la Corriente de California: de la escala diurna a la interanual. Cienc Mar. 43(3):137-156. http://dx.doi.org/10.7773/cm.v43i3.2707 [ Links ]

Delgadillo-Hinojosa, F., Macías-Zamora, J.V., Segovia-Zavala, J.A., Torres-Valdés, S. 2001. Cadmium enrichment in the Gulf of California. Mar Chem. 75(1-2):109-122. https://doi.org/10.1016/S0304-4203(01)00028-7 [ Links ]

Dickson, A.G., Goyet, C. 1994. Handbook of methods for the analysis of the various parameters of the carbon dioxide system in sea water. Version 2. United States of America: Oak Ridge National Laboratory. http://dx.doi.org/10.2172/10107773 [ Links ]

[ECMWF] European Centre for Medium-Range Weather Forecasts. 2017. Advancing global NWP through international collaboration. England: ECMWF; accessed 2020 Jul 14. https://www.ecmwf.int/en/about/contact-us/location . [ Links ]

Espinosa-Carreón, T.L., Gaxiola-Castro, G., Beier, E., Strub, P.T., Kurczyn, J.A. 2012. Effects of mesoscale processes on phytoplankton chlorophyll off Baja California. J Geophys Res. 117(C4):C04005. https://doi.org/10.1029/2011JC007604 [ Links ]

Flores-Trejo, L., Espinosa-Carreón, T.L., De-la-Cruz-Ruíz, A.I, Hernández-Ayón, J.M., Chapa-Balcorta, C. 2019. Dinámica del sistema del carbono en la columna de agua en octubre 2018 en Punta Lobos, Sonora. In: Paz, F., Velázquez, A., Rojo, M. (eds.), Estado Actual del Conocimiento del Ciclo De Carbono y sus Interacciones en México: Síntesis 2019. Texcoco (México): [Programa Mexicano del Carbono]. p. 304-310. [ Links ]

Franco, A.C., Hernández-Ayón, J.M., Beier, E., Garçon, V., Maske, H., Paulmier, A., Färber-Lorda, J., Castro, R., Sosa-Ávalos, R. 2014. Air-sea CO2 fluxes above the stratified oxygen minimum zone in the coastal region off Mexico. J Geophys Res Oceans. 119(5):2923-2937. https://doi.org/10.1002/2013JC009337 [ Links ]

Gattuso, J.P., Frankignoulle M, Wollast R. 1998. Carbon and carbonate metabolism in coastal aquatic ecosystems. Annu Rev Ecol Syst. 29(1):405-434. https://doi.org/10.1146/annurev.ecolsys.29.1.405 [ Links ]

Hernández-Ayón, J.M. 1995. Desarrollo de un sistema automático, sencillo y preciso de medición de CO2 total, alcalinidad y pH [MSc thesis]. [Ensenada (Mexico)]: UABC. 67 p. [ Links ]

Hernández-Ayón, J.M., Chapa-Balcorta, C., Delgadillo-Hinojosa, F., Camacho-Ibar, V.F., Huerta-Diaz, M.A., Santamaría-del-Angel, E., Galindo-Bect, S., Segovia-Zavala, J.A. 2013. Dynamics of dissolved inorganic carbon in the Midriff Islands region of the Gulf of California: Influence of water masses = Dinámica del carbono inorgánico disuelto en la región de las grandes islas del golfo de California: Influencia de las masas de agua. Cienc Mar. 39(2):183-201. http://doi.org/10.7773/cm.v39i2.2243 [ Links ]

Ho, D.T., Law, C.S., Smith, M.J., Schlosser, P., Harvey, M., Hill, P. 2006. Measurements of air‐sea gas exchange at high wind speeds in the Southern Ocean: Implications for global parameterizations. Geophys Res Lett. 33(16). https://doi.org/10.1029/2006GL026817 [ Links ]

Jiménez-López, D., Sierra, A., Ortega, T., Garrido, S., Hernández-Puyuelo, N., Sámchez-Leal, R., Forja, J. 2019. pCO2 variability in the surface waters of the eastern Gulf of Cádiz (SW Iberian Peninsula). Ocean Sci. 15:1225-1245. https://doi.org/10.5194/os-15-1225-2019. [ Links ]

Laruelle, G.G., Lauerwald, R., Pfeil, B., Regnier, P. 2014. Regionalized global Budget of the CO2 exchange at the air-water interface in continental shelf seas. Glob Biogeochem Cycles. 28(11):1199-1214. https://doi.org/10.1002/2014GB004832 [ Links ]

Levin, L.A. 2002. Deep-Ocean life where oxygen is scarce. Am Sci. 90(5):436-444. [ Links ]

Lewis, E., Wallace, D. 1998. Program developed for the CO2 systems calculations. Oak Ridge (TN): Carbon Dioxide Information Analysis Center. Report ORNL/CDIAC-105. [ Links ]

Linacre, L., Durazo, R., Hernández-Ayón, J.M., Delgadillo-Hinojosa, F., Cervantes-Díaz, G., Lara-Lara, J.R., Camacho-Ibar, V., Siqueiros-Valencia, A., Bazán-Guzmán, C. 2010. Temporal variability of the physicochemical water characteristics at a coastal monitoring observatory: Station ENSENADA. Cont Shelf Res. 30(16):1730-1742. https://doi.org/10.1016/j.csr.2010.07.011 [ Links ]

Liss, P.S., Merlivat, L. 1986. Air-Sea gas exchange rates: Introduction and synthesis. In: Buat-Ménard, P. (ed.), The Role of Air-Sea Exchange in Geochemical Cycling. Dordrecht (Netherlands): Springer. (NATO ASI Series; vol.185). p. 113-127. https://doi.org/10.1007/978-94-009-4738-2_5 [ Links ]

Lueker, T.J., Dickson, A.G., Keeling, C.D. 2000. Ocean pCO2 calculated from dissolved inorganic carbon, alkalinity, and equations for K1 and K2: validation based on laboratory measurements of CO2 in gas and seawater at equilibrium. Mar Chem. 70(1-3):105-119. https://doi.org/10.1016/s0304-4203(00)00022-0 [ Links ]

Marinone, S.G. 2003. A three‐dimensional model of the mean and seasonal circulation of the Gulf of California. J Geophys Res. 108(C10):2-17. https://doi.org/10.1029/2002jc001720 [ Links ]

Morales-Urbina, P., Espinosa-Carreón, T.L., Álvarez-Borrego, S., Hernández-Ayón, J.M., Flores-Trejo, L., Coronado-Álvarez, L.L.A. 2017. Flujo de CO2 océano-atmósfera en la zona de surgencias frente al norte de Sinaloa. In: Paz, F., Torres, R., Velázquez, A. (eds.), Estado Actual del Conocimiento del Ciclo del Carbono y sus Interacciones en México: Síntesis a 2017. Serie Síntesis Nacionales. Texcoco (Mexico): Programa Mexicano del Carbono, CICESE, UABC. p. 178-183 [ Links ]

[NOAA] National Oceanic and Atmospheric Administration. 2017. Trends in Atmospheric Carbon Dioxide. [Broadway (United States): NOAA]; accessed 2020 Jul 14. https://www.esrl.noaa.gov/gmd/ccgg/trends/graph.html . [ Links ]

Paulmier, A., Ruiz-Pino, D. 2009. Oxygen minimum zones (OMZs) in the modern ocean. Prog Oceanogr. 80(3-4):113-128. https://doi.org/10.1016/j.pocean.2008.08.001 [ Links ]

Pegau, W.S., Boss, E., Martínez, A. 2002. Ocean color observations of eddies during the summer in the Gulf of California. Geophy Res Lett. 29(9):1-6. https://doi.org/10.1029/2001gl014076 [ Links ]

Peres-Neto, P.R., Jackson, D.A., Somers, K.M. 2003. Giving meaningful interpretation to ordination axes: assessing loading significance in principal component analysis. Ecology. 84(9):2347-2363. https://doi.org/10.1890/00-0634 [ Links ]

Peterson, M.N.A. 1966. Calcite: Rates of dissolution in a vertical profile in the central Pacific. Science. 154(3756):1542-1544.https://doi.org/10.1126/science.154.3756.1542 [ Links ]

Portela, E., Beier, E., Barton, E.D., Castro, R., Godínez, V., Palacios- Hernández, E., Fiedler, P.C., Sánchez-Velasco, L., Trasviña, A. 2016. Water masses and circulation in the Tropical Pacific off Central Mexico and surrounding Areas. J Phys Oceanogr. 46(10):3069-3081. https://doi.org/10.1175/JPO-D-16-0068.1 [ Links ]

Pond, S., Pickard, G.L. 2013. Introductory Dynamical Oceanography. 22nd ed. [Oxford]: Bultterworth-Heineman. 329 p. ISBN. 9780750624961. [ Links ]

Roden, G.I. 1964. Oceanographic aspects of the Gulf of California. In: van-Andel, T.H., Shor, G.G. (eds.), Marine Geology of the Gulf of California. Tulsa (OK): Am Assoc Petr Geol Mem 3. p. 30-58. [ Links ]

Rodríguez-Ibáñez, C., Álvarez-Borrego, S., Marinone, S., Lara-Lara, R. 2013. The Gulf of California is a source of carbon dioxide to the atmosphere = El golfo de California es una fuente de bióxido de carbono hacia la atmósfera. Cienc Mar. 39(2):137-150. https://doi.org/10.7773/cm.v39i2.2190 [ Links ]

Santamaría-del-Ángel, E.M., Álvarez-Borrego, S., Millán-Nuñez, R., Muller-Karger, F.E. 1999. Sobre el efecto débil de las surgencias de verano en la biomasa fitoplanctónica del Golfo de California = On the weak effect of summer upwelling on the phytoplankton biomass of the Gulf of California. Rev Soc Mex Hist Nat. 49:207-212. [ Links ]

Schlitzer, R. 2016. Ocean Data View User’s Guide. [Germany]: [publisher unknown]; accessed 2020 Jul 14. http://odv.awi.deLinks ]

Soto-Mardones, L., Marinone, S.G., Parés-Sierra, A. 1999. Variabilidad espaciotemporal de la temperatura superficial del mar en el golfo de California = Time and spatial variability of sea surface temperature in the Gulf of California. Cienc Mar. 25(1):1-30. https://doi.org/10.7773/cm.v25i1.658 [ Links ]

Sutton, A.J., Sabine, C.L., Maenner-Jones, S., Lawrence-Slavas, N., Meinig, C., Feely, R.A., Mathis, J.T., Musielewicz, S., Bott, R., McLain, P.D., et al. 2014. A high-frequency atmospheric and seawater pCO2 data set from 14 open-ocean sites using a moored autonomous system. Earth Sys Sci Data. 6:353-366. https://doi.org/10.5194/essd-6-353-2014 [ Links ]

Takahashi, T., Sutherland, S.C., Sweeny, C., Poisson, A., Metzl, N., Tilbrook, B., Bates, N., Wanninkhof, R., Feely, R.A., Sabine, C., et al. 2002. Global sea-air CO2 flux bases on climatological surface ocean pCO2, and seasonal biological and temperature effects. Deep-Sea Res PT II. 49(9-19):1601-1622. https://doi.org/10.1016/s0967-0645(02)00003-6 [ Links ]

Thomas, W.H. 1966. On denitrification in the northeastern tropical Pacific Ocean. Deep-Sea Res. 13(6):1109-1114. https://doi.org/10.1016/0011-7471(66)90702-9 [ Links ]

Trucco-Pignata, P.N., Hernández-Ayón, J.M., Santamaria-del-Angel, E., Beier, E., Sánchez-Velasco, L., Godínez, V.M., Norzagaray, O. 2019. Ventilation of the upper oxygen minimum zone in the coastal region off Mexico: Implications of El Niño 2015-2016. Front Mar Sci. 6:459. https://doi.org/10.3389/fmars.2019.00459 [ Links ]

Wanninkhof, R. 2014. Relationship between wind speed and gas exchange over the ocean revisited. Limnol Ocean Methods. 12(6):351-362. https://doi.org/10.4319/lom.2014.12.351 [ Links ]

Weiss, R.F. 1974. Carbon dioxide in water and seawater: the solubility of a non-ideal gas. Mar Chem. 2(3):203-215. https://doi.org/10.1016/0304-4203(74)90015-2 [ Links ]

Received: March 08, 2021; Accepted: August 28, 2021

*Corresponding author. E-mail: leticiaesp@gmail.com.

English translation by Claudia Michel-Villalobos.

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License