SciELO - Scientific Electronic Library Online

 
vol.42 número3Modelización del impacto bentónico de los residuos particulados provenientes de una granja de doradas (Sparus aurata) durante un ciclo completo de cultivo índice de autoresíndice de assuntospesquisa de artigos
Home Pagelista alfabética de periódicos  

Serviços Personalizados

Journal

Artigo

Indicadores

Links relacionados

  • Não possue artigos similaresSimilares em SciELO

Compartilhar


Ciencias marinas

versão impressa ISSN 0185-3880

Cienc. mar vol.42 no.3 Ensenada Set. 2016

https://doi.org/10.7773/cm.v42i3.2651 

Articles

Seasonal air-sea CO 2 fluxes in the southern California Current off the Baja California Peninsula (Mexico)

Manuel Mariano-Matías1 

Gilberto Gaxiola-Castro1 

Martín Efraín De la Cruz-Orozco1 

Francisco P Chavez2 

Translation:

Christine Harris*

1 Departamento de Oceanografía Biológica, Centro de Investigación Científica y de Educación Superior de Ensenada, Carretera Ensenada-Tijuana, no. 3918, Zona Playitas, CP 22860, Ensenada, Baja California, México.

2 Monterey Bay Aquarium Research Institute, 7700 Sandholdt Road, Moss Landing, California 95039, USA.


Abstract:

Seasonal climatology of air-sea CO2 exchange was estimated from in situ sea surface CO2 partial pressure, temperature, salinity, and satellite wind data obtained from 2004 to 2011 in the southern region of the California Current, off the Baja California Peninsula (Mexico). The average annual CO2 flux indicates that the study area is a source of CO2 to the atmosphere (0.65 mmol m-2 d-1). It changes from being a source in summer (2.33 mmol m-2 d-1) and autumn (0.92 mmol m-2 d-1) to acting as a sink in winter (-0.26 mmol m-2 d-1) and spring (-0.37 mmol m-2 d-1). The area to the north of latitud 28°N (off Punta Eugenia) is a CO2 sink (-0.42 mmol m-2 d-1), whereas the area to the south of this latitude is a source of CO2 to the atmosphere (1.80 mmol m-2 d-1), mostly due to thermodynamic effects. The northern coastal zone is a permanent CO2 sink (-1.29 mmol m-2 d-1). During the 2004 El Niño event the whole area contributed 2.00 mmol m-2 d-1 of CO2 to the atmosphere, but during the 2011 La Niña the ocean absorbed 5.30 mmol m-2 d-1 as a result of physical and biological dynamics. The seasonal cycle is dominated by temperature rather than biological effects, except in the northern coastal area. It is necessary to continue with in situ measurements of the CO2 system to have solid foundations to estimate the effect of the long-term increase in dissolved inorganic carbon on marine organisms.

Key words: CO2 fluxes; interannual variability; El Niño; La Niña; California Current

Resumen:

La climatología estacional del flujo de CO2 entre el mar y la atmósfera en la región sur de la corriente de California, frente a la península de Baja California, fue calculada con datos in situ de la presión parcial de CO2, temperatura y salinidad recolectados de 2004 a 2011. Según el promedio anual del flujo de CO2 (0.65 mmol m-2 d-1), la zona de estudio es una fuente de CO2 a la atmósfera, y cambia de ser fuente en verano (2.33 mmol m-2 d-1) y otoño (0.92 mmol m-2 d-1) a ser sumidero en invierno (-0.26 mmol m-2 d-1) y primavera (-0.37 mmol m-2 d-1). La zona al norte de la latitud 28°N (frente a punta Eugenia) presenta condiciones de sumidero de CO2 (-0.42 mmol m-2 d-1), mientras que la zona al sur de esta latitud se manifiesta como una fuente de CO2 a la atmósfera (1.80 mmol m-2 d-1), lo cual es atribuible mayormente al efecto termodinámico. La zona costera norte es un sumidero permanente de CO2 (-1.29 mmol m-2 d-1). Durante el evento de El Niño 2004 el océano aportó 2.00 mmol m-2 d-1 de CO2 hacia la atmósfera, y durante el evento de La Niña 2011 el océano incorporó 5.30 mmol m-2 d-1 debido a un efecto combinado de factores físicos y biológicos. En el ciclo estacional predominaron los efectos de la temperatura en vez de los efectos biológicos, excepto en la zona costera norte. Es necesario continuar con las mediciones in situ del sistema del CO2 para contar con bases sólidas que permitan estimar el efecto del aumento a largo plazo del carbono inorgánico disuelto del océano sobre los organismos.

Palabras clave: flujos de CO2; variabilidad interanual; El Niño; La Niña; corriente de California

Introduction

Over the past 2 centuries human activities have increased the concentration of CO2 in the atmosphere by ~400 petagrams (1 Pg = 1015 g), that is, by ~40% (Zeebe 2012). Atmospheric CO2 content has changed from 280 μatm before the industrial revolution to 407 μatm in recent times (NOAA 2016), and could reach 850 μatm in 2100 (Hoegh-Guldberg et al. 2014). Oceans absorb about 2 Pg C yr-1 of the CO2 emitted through human activities (Takahashi et al. 2009), but the absorption rate is not high enough to prevent the increase of CO2 in the atmosphere. Since CO2 is a greenhouse gas, global mean temperature has risen by 0.85 °C and an increase of ~4 °C relative to the global mean estimated from 1980 to 1999 has been reported for 2100 (IPCC 2014).

On average, oceans act as a net CO2 sink but the direction and magnitude of the exchange of CO2 with the atmosphere varies considerably over space and time. Hales et al. (2012) characterized the region of the Pacific Ocean between 22°N and 50°N (30 to 370 km offshore) as a source of atmospheric CO2 of 0.60 mol C m-2 yr-1, a value 3 times higher than that estimated by Chavez et al. (2007) for the same area. De La Cruz-Orozco et al. (2010) studied air-sea CO2 exchange off the Baja California Peninsula from October 2004 to October 2005, and on average it acted as a source of CO2, with an annual value of 0.40 mol m-2. The average annual CO2 partial pressure (pCO2) data collected from 1993 to 2001 by ships of opportunity off Baja California revealed that the area is a source of CO2, but that it acts as a sink in winter and spring and is a source in summer and autumn (Hernández-Ayón et al. 2010).

In the California Current, corrosive water (pH less than 7.80) has been detected near the surface, a condition that was not expected to occur for some 50 years (Feely et al. 2008, Alin et al. 2012). Though some of the corrosive characteristics of these waters are the result of respiration processes in subsurface waters, the accumulation of atmospheric CO2 continues, modifying the natural dynamics of the CO2 cycle in coastal upwelling zones (Feely et al. 2008).

Large-scale events such as El Niño/Southern Oscillation (ENSO) affect the air-sea flux of carbon (FCO2). During the 1997-1998 El Niño event, in the central California upwelling system (to the north of our study area), the FCO2 fluctuated between -0.30 and -0.70 mol C m-2 yr-1, with inputs of 1.50 to 2.20 mol C m-2 yr-1 under 1999 La Niña conditions (Friederich et al. 2002). Off the tropical Pacific coast of Mexico (to the south of our study area), the FCO2 ranged from -0.40 to 2.50 mmol C m-2 d-1 during the 2009 El Niño event and from -4.40 to 3.30 mmol C m-2 d-1 during the 2010 La Niña event (Franco et al. 2014). In the southern region of the California Current off the Baja California Peninsula, our area of interest, we would expect fluxes from the ocean to the atmosphere under El Niño conditions and from the atmosphere to the ocean under La Niña conditions.

Analysis of high-frequency (3-h) pCO2 data obtained by sensors on an autonomous buoy moored off the northern coast of Baja California revealed that this coastal zone is a net source of CO2 (~1.00 mol C mm-2 yr-1) to the atmosphere (Reimer et al. 2013b, Muñoz-Anderson et al. 2015). Reimer et al. (2013b) concluded that 91% of the positive FCO2 (source) was contributed during coastal upwelling events, and Muñoz-Anderson et al. (2015) found that the system acts as a slight sink (0.06 mol C m-2 yr-1) during non-upwelling periods.

Few studies have been published on the seasonal variability of air-sea CO2 exchange in the southern region of the California Current off Baja California. In this paper we present an analysis of pCO2 and FCO2 calculated with in situ data from 22 oceanographic cruises. This database allows us to affirm that the results are robust for the study area and contribute significantly to the knowledge of the spatial and seasonal variability of pCO2 and FCO2 in this region. The data also indicate that the temporal variations and intrinsic controls are subject to dynamic spatial variations and should be examined individually in each spatial domain.

Materials and methods

The Pacific Ocean off the Baja California Peninsula presents high spatial and temporal variability caused mainly by the physical dynamics derived from changes in the intensity of flow of the California Current (Durazo and Baumgartner 2002); coastal upwelling events (Pérez-Brunius et al. 2007); the presence of fronts, meanders, and eddies (Barocio-León et al. 2007); and seasonal and interannual variability (Aguirre-Hernández et al. 2004, Gaxiola-Castro et al. 2008). Because of its seasonal variability, the study area comprising the southern region of the California Current has been divided into northern and southern portions at latitude 28°N (Durazo 2015).

In the southern region of the California Current, the IMECOCAL (Spanish acronym for Mexican Research of the California Current) program undertakes oceanographic cruises 4 times a year (winter, spring, summer, and autumn) off the Baja California Peninsula, between ~30 and ~220 km offshore. The sampling grid consists of 90 stations, separated by ~37 km, arranged in lines perpendicular to the coast, with a distance of ~70 km between hydrographic lines (Fig. 1a). In this paper we analyze the hydrographic and surface pCO2 data obtained during 22 cruises carried out between 2004 and 2011 (Fig. 1b).

Figure 1 (a) Zonification of the study area and sampling sites of the IMECOCAL program. Clear circles and squares indicate the stations in the northern oceanic zone and southern oceanic zone, respectively. Black circles and squares indicate the stations in the northern coastal zone and southern coastal zone, respectively. (b) Distribution of the number of hydrographic stations occupied during the IMECOCAL cruises conducted from 2004 to 2011. 

During the oceanographic cruises, casts were made with a Sea-Bird Electronics 911 plus CTD (calibrated by the manufacturer), coupled to a rosette fitted with 5-L Niskin bottles. The CTD contained sensors to measure temperature, salinity, and dissolved oxygen. Water samples were taken with the Niskin bottles at discrete depths (0, 10, 20, 50, 100, 150, and 200 m) to analyze dissolved oxygen and phyto-plankton chlorophyll a.

The samples for the dissolved oxygen analysis were collected in 125-mL glass bottles and were analyzed on board by the micro-Winkler method (Helm et al. 2009). The linear correlation coefficient between the discrete dissolved oxygen values obtained with the micro-Winkler method and those derived by the CTD was 0.96 (n = 346, P < 0.05). Based on the CTD oxygen data, the percentage of oxygen saturation (%OS) was estimated with the Weiss (1970) equations.

The water aliquots for the chlorophyll a analysis were collected in 1-L dark plastic bottles, and the concentration was determined using Turner Designs 10-AU-05 and Trilogy (model 7200-000) fluorometers by the method proposed by Holm-Hansen et al. (1965) and modified by Venrick and Hayward (1984). Further details of the collection methods and analysis used by the IMECOCAL program can be found at http://imecocal.cicese.mx.

Sea surface pCO2 (pCO2sea) was analyzed using an automated continuous flow LI-COR 6262 system that measures the mole fraction of CO2 in dry air (xCO2). Seawater was pumped from an intake in the ship's hull (below the water-line, ~2.50 m depth). It flowed to a thermo-salinometer and then to the LI-COR 6262 analyzer, both located in the dry laboratory. The accuracy of the LI-COR 6262 analyzer for the pCO2sea measurements is ±1 |iatm and it autocalibrated every 2 h against commercial standard CO2 obtained from the National Institute of Standards and Technology and the National Oceanic Atmospheric Administration Climate Monitoring Diagnostic Laboratory (Friederich et al. 2002).

Further details of how the pCO2 data were obtained and processed can be found in Reimer et al. (2013a). Since the pCO2sea data were obtained continuously during navegation, it was necessary to obtain an average during the time (~40 min) that the boat remained at each sampling station in order to associate them with the discrete temperature, salinity, %OS, and chlorophyll a data.

The sea-air CO2 exchange was calculated using the following equation: FCO2 = Kw x K0 x (ΔpCO2), where Kw is the CO2 transfer coefficient as a function of wind speed (Wanninkhof 2014), K0 is the CO2 solubility coefficient as a function of temperature and salinity (Weiss 1974), and ΔpCO2 is the difference between the pCO2 in seawater (pCO2sea) and in the atmosphere (pCO2air). By convention, negative FCO2 values indicate fluxes from the atmosphere to the ocean and positive values indicate fluxes from the ocean to the atmosphere. Atmospheric CO2 is well mixed and its partial pressure varies within a very narrow range relative to its mean value (Sarmiento and Gruber 2006); hence, the daily pCO2air data were obtained from http://scrippsco2.ucsd.edu/data/atmospheric_co2. These data correspond to measurements taken at Mauna Loa Observatory, Hawaii, from 2004 to 2011. Based on this information, pCO2air averages were obtained for the period close to each oceanographic campaign and used to calculate an average pCO2air value for each of the 22 cruises analyzed. Monthly ocean surface wind speed data (spatial resolution of 0.25° x 0.25°) were obtained from http://podaac.jpl.nasa.gov/DATACATALOG/ccmpinfo.html; they are the product of cross calibration of several multiplatforms (Atlas et al. 2011).

The seasonal climatology of temperature, salinity, chlorophyll, %OS, pCO2, and FCO2 was calculated for the period under study (5 winters, 6 springs, 5 summers, and 6 autumns). Since the sampling grid was not always entirely covered, the averages were calculated with the number of data available for each cruise using an arithmetic mean. To minimize measurement errors, the values that fell outside the range of 3 standard deviations were eliminated. To determine the response of pCO2 and FCO2 to the ENSO cycle in the region, we analyzed the effects during its warm phase (El Niño) in autumn 2004 and during its cold phase (La Niña) in spring 2011, both events identified according to the Multivariate ENSO Index (http://www.cpc.ncep.noaa.gov/products/analysismonitoring/ensostuff/ensoyears.shtml).

Results

The seasonal climatology (2004-2011) of sea surface temperature showed a latitudinal gradient, with higher values in the southern portion (below line 120) and lower values in the northern portion (above line 120) of the IMECOCAL grid (Fig. 2). Surface temperature showed 2 different conditions: lower values in winter (16.50 °C; Fig. 2a) and spring (16.30 °C; Fig. 2b), and higher values in summer (19.90 °C; Fig. 2c) and autumn (20.20 °C; Fig. 2d). A longitudinal pattern was persistent in the northern portion, with lower values close to shore and higher values in the open sea. A latitudinal pattern predominated in the southern portion, with isotherms perpendicular to the coast, particularly in summer and autumn (Fig. 2c, d).

Figure 2 Seasonal climatology (2004-2011) of sea surface temperature (SST, °C) in (a) winter, (b) spring, (c) summer, and (d) autumn. 

The seasonal climatology of sea surface salinity showed a latitudinal gradient, with isohalines perpendicular to the coast (Fig. 3a, c). A longitudinal pattern was observed only in spring in the northern coastal zone (NCZ), from Punta Eugenia to Ensenada (Fig. 3b), with isohalines parallel to the coast and extending ~50 km offshore. The mean salinity values were higher in autumn (33.65; Fig. 3d) and winter (33.62; Fig. 3a), and lower in spring (33.53; Fig. 3b) and summer (33.57; Fig. 3c).

Figure 3 Seasonal climatology (2004-2011) of sea surface salinity in (a) winter, (b) spring, (c) summer, and (d) autumn. 

Chlorophyll and %OS showed an inshore-offshore gradient, with higher values close to shore and lower values towards the open sea. Mean chlorophyll concentration was higher in spring (0.88 mg m-3; Fig. 4b) and the isolines of 0.50 mg m-3 extended ~100 km offshore. A plume of chlorophyll values exceeding 0.50 mg m-3 that extends from Punta Eugenia to ~140 km offshore (Fig. 4a, c) can be observed in all the seasons. The relative chlorophyll maximum values occurred in the NCZ. During winter and spring, chlorophyll concentrations ranged from 0.50 to 1.50 mg m-3 in the southern coastal zone (SCZ). The longitudinal pattern of %0S distribution was more pronounced in summer (Fig. 5c), with isolines of 100%0S parallel to the peninsula between the coast and ~100 km offshore, particularly in the NCZ. Saturation values below 100% predominated in winter (97%0S; Fig. 5a), indicating a predominance of the thermodynamic effect. In spring (Fig. 5b), the area off Ensenada had saturation values above 110%, with isolines parallel to the peninsula between the coast and ~100 km offshore. In summer, the >100%OS conditions persisted off Ensenada, but the isolines parallel to the peninsula covered an area between the coast and ~30 km offshore.

Figure 4 Seasonal climatology (2004-2011) of sea surface chlorophyll a (mg m 3) in (a) winter, (b) spring, (c) summer, and (d) autumn. 

Figure 5 Seasonal climatology (2004-2011) of sea surface dissolved oxygen saturation (%0S) in (a) winter, (b) spring, (c) summer, and (d) autumn. 

In general, the seasonal climatology of pCO2sea showed a latitudinal pattern, with higher values in the southern portion and lower in the northern portion of the study area (Fig. 6); these values are associated with spatial changes in sea surface temperature. The 380 μatm isoline perpendicular to the coast was situated at ~29°N in autumn (Fig. 6d) and moved to ~27°N in winter (Fig. 6a). In spring (Fig. 6b), the coastal area to the south of Ensenada stands out, with pCO2sea values below 350 μatm and isolines parallel to the peninsula between the coast and ~80 km offshore. In summer (Fig. 6c), the area covered by the isoline with values below 350 μatm decreased (from the coast to ~40 km offshore), and values higher than 400 μatm predominated in the oceanic fringe at 29°N and to the south of the study area. The coastal zone to the north of Punta Eugenia consistently had low pCO2 values (372 μatm), indicating that it acted as a possible CO2 sink during the seasonal cycle. The pCO2 climatology revealed a temporal pattern similar to that of temperature: lower values in winter (378 μatm; Fig. 6a) and spring (388 μatm; Fig. 6b) and higher in summer (410 μatm; Fig. 6c) and autumn (395 μatm; Fig. 6d).

Figure 6 Seasonal climatology (2004-2011) of sea surface partial pressure of CO2 (pCO2, μatm) in (a) winter, (b) spring, (c) summer, and (d) autumn. 

Air-sea FCO2 showed a latitudinal distribution pattern, with higher values in the southern portion and lower in the northern portion (Fig. 7), similar to the distribution of pCO2 (Fig. 6). In winter (Fig. 7a), the coverage of the area with negative values extended to ~26°N in the northern oceanic zone (N0Z) and to 28°N in the NCZ. Further south, the values were predominantly positive, on average 0.85 mmol m-2 d-1. In spring (Fig. 7b), the NCZ stands out, with FCO2 values above -6.00 mmol m-2 d-1 and isolines parallel to the coast and extending ~80 km offshore. An area with positive FCO2 values (1.25 mmol m-2 d-1) extended from ~28.5°N in the N0Z towards Punta Eugenia and covered practically all the southern portion of the study area. In summer (Fig. 7c), the coverage of the negative flows decreased in the NCZ (extending ~40 km offshore) and the positive flows predominated. A fringe with positive FCO2 values between 1.20 and 1.80 mmol m-2 d-1 occurred in the NOZ (31°N) and extended southwards covering the southern oceanic zone (S0Z) and the SCZ, with values reaching 6.00 mmol m-2 d-1. In autumn (Fig. 7d), the area to the north of 28°N showed negative fluxes of up to -3.00 mmol m-2 d-1 (Fig. 7d), but to the south of this latitude, the values were positive, of up to 3.50 mmol m-2 d-1. The NCZ consistently had negative FCO2 values, indicating that it acts as a CO2 sink during the seasonal cycle. The FCO2 climatology revealed a similar pattern to that of pCO2: negative fluxes in winter (-0.26 mmol m-2 d-1; Fig. 7a) and spring (-0.37 mmol m-2 d-1; Fig. 7b) and positive in summer (2.33 mmol m-2 d-1; Fig. 7c) and autumn (0.92 mmol m-2 d-1; Fig. 7d).

Figure 7 Seasonal climatology (2004-2011) of air-sea CO2 flux (FCO2, mmol m 2 d 1) in (a) winter, (b) spring, (c) summer, and (d) autumn. 

In the 2004-2011 time series for the northern portion of the study area, mean sea surface temperature ranged from 14.5 to 20.0 °C, the lowest values occurring in winter and spring and the highest in summer and autumn (Fig. 8a). A decreasing trend (by 0.10 °C) occurred after the summer of 2007. Chlorophyll concentration was higher in spring. There was a period of low chlorophyll concentration (0.42 mg m-3) between 2004 and mid-2007 (Fig. 8b), after which it increased. The average pCO2 tended to decrease after 2007, with values below the atmospheric value as of 2009 (Fig. 8c). This same tendency was observed for FCO2, with negative values as of 2009 (Fig. 8d). In the spring of 2011 the northern portion had the maximum CO2 sink value of all the time series (-6.00 mmol C m-2 d-1).

Figure 8 Time series (2004-2011) of average (a) temperature (°C), (b) chlorophyll (Chl, mg m 3), (c) partial pressure of CO2 (pCO2, μatm), and (d) air-sea CO2 flux (FCO2, mmol m-2 d-1) in the northern portion of the study area (above IMECOCAL line 120). The bars indicate the standard error. The continuous line in (c) indicates the atmospheric pCO2 values. The dashed line in (d) indicates the equilibrium between the atmosphere and ocean. 

In the southern portion of the study area, mean sea surface temperature ranged from 15.80 to 23.50 °C, the lowest values occurring in winter and spring and the highest in summer and autumn (Fig. 9a). Mean chlorophyll concentration (0.48 mg m-3) was slightly lower than in the northern portion (0.62 mg m-3) and showed an increasing trend from the summer of 2006 (Fig. 9b) to the end of the study period. In general, pCO2 was higher than the atmospheric value (Fig. 9c), except in 2011, when it was lower, and a decreasing trend occurred from 2007 until the end of the study period. The southern portion acted mainly as a source of CO2 to the atmosphere, with an annual average of 1.56 mmol m-2 d-1 (Fig. 9d) and values close to equilibrium in winter, except in 2011 when it acted as a sink.

Figure 9 Time series (2004-2011) of average (a) temperature (°C), (b) chlorophyll (Chl, mg m 3), (c) partial pressure of CO2 (pCO2, μatm), and (d) air-sea CO2 flux (FCO2, mmol m-2 d-1) in the southern portion of the study area (IMECOCAL line 120 and below). The bars indicate the standard error. The continuous line in (c) indicates the atmospheric pCO2 values. The dashed line in (d) indicates the equilibrium between the atmosphere and ocean. 

During the study period (2004-2011), the southern region of the California Current was under the influence of El Niño (2004, 2007, 2009-2010) and La Niña (2008, 2009, 2011) events (Bjorkstedt et al. 2012) (http://www.cpc.ncep.noaa.gov/products/analysis_monitoring/ensostuff/ensoyears.shtml). For all the region, the average pCO2 during the 2004 El Niño (Fig. 10a) was 44 μatm higher than during the 2011 La Niña (Fig. 10b). During the 2004 El Niño, the spatial distribution of pCO2 showed a latitudinal pattern, with higher values (414 μatm) in the southern portion and lower (399 μatm) in the northern portion; in the latter, a longitudinal pattern was observed, with lower pCO2 values (394 μatm) in the coastal zone. During the 2011 La Niña, the average pCO2 was 360 μatm, lower than the value (393 μatm) obtained for pCO2air. During the 2004 El Niño, the ocean off Baja California acted as a source of CO2 to the atmosphere (on average, 2.00 mmol m-2 d-1; Fig. 10c), whereas during the 2011 La Niña it acted as a sink (on average, 5.30 mmol m-2 d-1; Fig. 10d). Note the marked latitudinal gradient in pCO2 and FCO2 under La Niña conditions, with low (high) values of 355 μatm (-6.40 mmol m-2 d-1) in the northern portion and high (low) values of 366 μatm (-4.00 mmol m-2 d-1) in the southern portion.

Figure 10 Spatial distribution of (a, b) the partial pressure of CO2 (pCO2, μatm) and (c, d) air-sea CO2 flux (FCO2, mmol m 2 d-1) under El Niño conditions (autumn 2004; a, c) and La Niña conditions (spring 2011; b, d). 

Discussion

Northern oceanic zone (NOZ)

In the NOZ, the low temperature and salinity values and high %OS and chlorophyll values in winter and spring are associated with the maximum equatorward flow of the California Current (Durazo 2015), which transports low-temperature and low-salinity subarctic water. The low salinity values reflect the presence of subarctic water that covers the northern portion of the study area all year long. Simple linear correlations were performed to identify the variables that regulate the variability of pCO2. The variability of pCO2 in the NOZ is mainly regulated by temperature, with maximum correlation between both variables in autumn (r = 0.67, P < 0.05, n = 30). This is confirmed by estimating the relative effect of temperature minus the effect of biological processes (ET - EB) during a seasonal cycle, as proposed by Takahashi et al. (2002). Positive values indicate that the physical effect predominates, whereas negative values indicate that the biological effect predominates. In the NOZ, the ET - EB index was 7.60, which indicates a predominance of the physical effect on pCO2 during the seasonal cycle.

Northern coastal zone (NCZ)

In the NCZ, coastal upwelling occurs all year round but it is more intense in spring and summer (Durazo 2015), as indicated by the average seasonal temperatures (14.80 °C), high chlorophyll concentration (1.80 mg m-3), and OS values of more than 104%. Isotherms parallel to the coast extend 50-100 km offshore, especially in spring when the alongshore wind is more intense (Pérez-Brunius et al. 2007). In spring, Ekman pumping transports subsurface nutrient-rich water to the surface (Perez-Brunius et al. 2007, Durazo 2015), which stimulates phytoplankton productivity and increases the levels of dissolved oxygen, leading to oxygen oversaturation in the coastal zone.

The pCO2 showed a high inverse linear correlation with %OS in all 4 seasons, the maximum value corresponding to autumn (r = -0.79, P < 0.05, n = 25). The negative correlation between pCO2 and %OS allows us to identify the expected patterns of the effect of the biological processes on CO2 uptake and production in the ocean. Low pCO2 values and high %OS (>100%) would indicate the prevalence of photosynthesis (Carrillo et al. 2004) and the inverse situation would indicate greater ecosystem respiration (Schloss et al. 2007). The ET - EB index for the NCZ was -8.26, indicating that the variability of pCO2 in the seasonal cycle is mainly governed by the biological factor.

Southern oceanic zone (SOZ)

When the SOZ is under the influence of the California Current in winter and spring (Durazo and Baumgartner 2002, Durazo 2015), the water has low temperature and salinity values and high %OS and chlorophyll values. In summer and autumn when the SOZ is under the influence of subtropical water, which flows northward and transports warm and salty water (Zaitzev et al. 2014, Durazo 2015), the temperature and salinity values are higher and the %OS and chlorophyll values are lower. In the SOZ, pCO2 was correlated with temperature in summer and autumn; maximum correlation occurred in summer (r = 0.85, P < 0.05, n = 30). The biological factors dominated in winter and spring. The highest correlation was between pCO2 and %OS in spring (r = -0.62, P < 0.05, n = 29). The ET - EB index for this zone was 20.79, indicating that the seasonal cycle was dominated by the effect of temperature on pCO2.

Southern coastal zone (SCZ)

In the SCZ, the latitudinal temperature gradient is maximum in summer, when upwelling occurs in the area (Durazo et al. 2010). Maximum temperature and salinity values are associated with the presence of subtropical subsurface water, which exerts greater influence on the area in summer and autumn (Durazo et al. 2010, Zaitsev et al. 2014, Durazo 2015). In the SCZ, the effect of temperature predominates in autumn and winter, with maximum correlation in autumn (r = 0.64, P < 0.05, n = 25). The direct relation between pCO2 and temperature indicates that the thermodynamic effect derived from the changes in circulation (Hernández-Ayón et al. 2010, Durazo 2015) is one of the main factors regulating pCO2 in the area. Takahashi et al. (1993) indicate that for every one-degree increase in temperature, pCO2 increases by 4.23%; hence, an increase in pCO2 of 50 μatm would be expected, a value slightly higher than that calculated (39 μatm). In spring and summer, pCO2 is affected by biological processes since linear correlation with chlorophyll was highest in summer (r = -0.59, P < 0.05, n = 32). The inverse relationship between pCO2 and chlorophyll shows, as has been reported for other oceanic regions (Mathis et al. 2010), that sea surface pCO2 is governed by phytoplankton productivity, highlighting the photosynthetic process in the oceanic carbon cycle. The ET - EB index was 18.93, indicating that the seasonal variability of pCO2 was governed by temperature effects rather than biological effects during the seasonal cycle.

Based on the above analysis and on the results of the oneway ANOVA performed to verify the existence of statistical differences between zones, when studying pCO2 and FCO2 it would be advisable to divide the study area as has been done here. The ANOVA results showed that there were significant differences between zones, except between NCZ and NOZ (P = 0.97) in winter and between SOZ and SCZ (P = 0.11) in autumn.

Sea-air CO 2 exchange

The average annual FCO2 indicates that the study area is a source of CO2 to the atmosphere (2,043 t). It changes from being a source in summer and autumn to acting as a sink in winter and spring. According to the classification proposed by Paulmier et al. (2008), the study area can be characterized as a weak source of CO2 to the atmosphere since it emits 0.65 mmol m-2 d-1. This average value is 52% lower than the value reported by De la Cruz-Orozco et al. (2010) for the same study area, calculated using data collected during cruises undertaken in 2004 and 2005, a period influenced by El Niño conditions and the anomalous intrusion of a large volume of subarctic water (Espinoza-Carreón et al. 2015). Hernández-Ayón et al. (2010), based on data collected between 1993 and 2001, reported a carbon flux of 2.60 mmol m-2 d-1 for the oceanic region off Baja California, a value 3 times higher than that reported here. Hales et al. (2012), based on model-derived data, reported that the area is a source of CO2 (1.80 mmol m-2 d-1). A reason for the difference between their and our results is that these authors fitted their models to obtain pCO2 using data collected between 1997 and 2005, and they calculated the solubility coefficient with satellite temperature data and did not use the same parameterization to calculate the transfer coefficient.

The NCZ acts as a permanent CO2 sink, with a seasonal average of -5.20 mmol m-2 d-1 (715 t). The NOZ acts as a sink from autumn to spring and as a source in summer (360 t) (Fig. 7). The southern portion is a source of CO2 to the atmosphere throughout the year (1,252 t), the lowest values occurring in winter and spring. This seasonal change from source to sink has been described by other authors; however, De la Cruz-Orozco et al. (2010) used data obtained when the system was influenced by an anomalous intrusion of subarctic water and by El Niño conditions, Hernández-Ayón et al. (2010) used data obtained mainly in the oceanic portion, and Hales et al. (2012) used model-derived data. The results of the present study can therefore be considered the most robust to date.

In spring and winter, when the northern portion acts as a CO2 sink and the lowest flux values correspond to the southern portion, the entire study area is under the influence of subarctic water (Durazo 2015), which transports cold water with a high concentration of dissolved oxygen. This statement is based on the spatial distribution of temperature (mean values of 16.50 °C in winter and 16.30 °C in spring; Fig. 2a, b). On the other hand, %OS is highest in winter and spring (97.14% and 97.70%, respectively; Fig. 5a, b).

The SOZ and SCZ are a permanent source of CO2 to the atmosphere (2.05 and 0.10 mmol m-2 d-1, respectively). The highest positive FCO2 values occur in summer and autumn, when the California Current weakens (Durazo 2015) and subtropical water, which is warmer, saltier, and lower in dissolved oxygen, enters the southern portion (Zaitsev et al. 2014). This seasonal change in circulation during summer and autumn results in higher temperatures (Fig. 2c, d) and salinities (Fig. 3a, b), and lower %OS (Fig. 5b, c) in the southern portion.

Air-sea CO 2 exchange and p CO 2 under El Niño and La Niña conditions

During El Niño (autumn 2004), the study area was a source of CO2 to the atmosphere (1.90 mmol m-2 d-1) and during La Niña (spring 2011), it was a sink (-5.30 mmol m-2 d-1). During El Niño events, ocean stratification increases, the nutricline deepens (Huyer and Smith 1985), and phytoplank-ton biomass and productivity decrease (Hernández de la Torre et al. 2004). The increase in temperature generates greater fugacity of CO2 (about 4.23% per one-degree Celsius change in temperature; Takahashi et al. 1993). Moreover, increased stratification of the water column restricts the rising of nutrient-rich water. Conditions are thus not propitious for phytoplankton growth and the biological use of dissolved inorganic carbon diminishes. Conversely, during La Niña events, upwelling-favorable conditions prevail and nutrient-rich water is transported to the surface where the phytoplankton assimilates the CO2 and, consequently, sea surface pCO2 decreases. Average chlorophyll concentration for the entire study area was 0.21 mg m-3 during the 2004 El Niño event but increased to 1.12 mg m-3 during the 2011 La Niña event (Figs. 8b, 9b). Oxygen saturation was lower under El Niño than under La Niña conditions (99.33% and 101.70%, respectively). The one-way ANOVA results indicated significant differences (P < 0.05) for pCO2 and FCO2 under El Niño and La Niña conditions.

The thermodynamic factor is also relevant when explaining the differences observed during the ENSO phases. Mean surface temperature was 20.92 °C during El Niño and 16.03 °C during La Niña (Figs. 8a, 9a). According to Takahashi et al. (1993), this means that just the thermody-namic effect would increase pCO2 by about 70 μatm under El Niño conditions relative to La Niña.

Effects on the epipelagic environment

The response of phytoplankton-the organisms primarily responsible for the sinking of biogenic carbon to greater depths-to an increase in CO2 in the ocean has so far been contradictory. For example, it has been reported that increased oceanic CO2 levels decrease the calcification of organisms such as corals, foraminifera, and coccolithophores (Beaufort et al. 2011) because this increase produces corrosive water (decrease in pH), and evidence of corrosive water has been found in the southern region of the California Current (Feely et al. 2008, Alin et al. 2012). Yet phytoplankton with high rates of calcification, such as the coccolithophore Emiliania huxleyi (Beaufort et al. 2011, Ribero-Calle et al. 2015), which has great adaptive capacity (Lohbeck et al. 2012), have been found in water with a low pH (7.62). These are some examples that show the complexity of explaining the response of the epipelagic ecosystem to increasing CO2 levels. It is necessary to continue with in situ CO2 measurements in order to generate solid information that can be used to estimate the effect on marine organisms of the long-term increase in dissolved inorganic carbon.

Acknowledgments

The National Council for Science and Technology (CONACYT, Mexico) provided funding for the IMECOCAL oceanographic campaigns through projects G35326T, SEP-2003-CO2-42569, 47044, 23947, and 99252. The Mexican Ministry of Environment and Natural Resources (SEMARNAT) and CONACYT provided financial support through project 23804. MMM acknowledges receipt of a CONACYT postgraduate scholarship. GGC was supported by CONACYT through the Sistema Nacional de Investigadores. We thank the crew of the R/V Francisco de Ulloa and the participants of the IMECOCAL 2004-2011 campaigns.

REFERENCES

Aguirre-Hernández E, Gaxiola-Castro G, Nájera-Martínez S, Baumgartner TR, Kahru M, Mitchell GB. 2004. Phytoplankton absorption, photosynthetic parameters, and primary production off Baja California: summer and autumn 1998. Deep-Sea Res. (II) 51(6-9): 799-816. http://dx.doi.org/10.1016/j.dsr2.2004.05.015 [ Links ]

Alin SR, Feely RA, Dickson AG, Hernández-Ayón JM, Juranek LW, Ohman MD, Goericke R. 2012. Robust empirical relationships for estimating the carbonate system in the southern California Current System and application to CalCOFI hydrographic cruise data (2005-2011). J. Geophys. Res. 117(C5): C05033. http://dx.doi.org/10.1029/2011JC007511 [ Links ]

Atlas R, Hoffman RN, Ardizzone J, Leidner SM, Jusem JC, Smith DK, Gombos D. 2011. A cross-calibrated, multiplatform ocean surface wind velocity product for meteorological and oceanographic applications. Bull. Am. Meteorol. Soc. 92(2): 157-174. http://dx.doi.org/10.1175/2010BAMS2946.1 [ Links ]

Barocio-León OA, Millán-Núñez R, Santamaría-del-Ángel E, González-Silvera A. 2007. Phytoplankton primary productivity in the euphotic zone of the California Current System estimated from CZCS imagery = Productividad primaria del fitoplancton en la zona eufótica del Sistema de la Corriente de California estimada mediante imágenes de CZCS. Cienc. Mar. 33(1): 59-72. [ Links ]

Beaufort L, Probert I, de Garidel-Thoron T, Bendif EM, Ruiz-Pino D, Metzl N, Goyet C, Buchet N, Coupel P, Grelaud M, Rost B, Rickaby REM, de Vargas C. 2011. Sensitivity of coccolithophores to carbonate chemistry and ocean acidification. Nature 476(7358): 80-83. http://dx.doi.org/10.1038/nature10295 [ Links ]

Bjorkstedt EP, Goericke R, McClatchie S, Weber E, Watson W, Lo NCH, Peterson WT, Brodeur RD, Auth T, Fisher J, Morgan C, Peterson J, Largier J, Bograd SJ, Durazo R, Gaxiola-Castro G, Lavaniegos BE, Chavez FP, Collins C, Hannah B, Field J, Sakuma K, Satterthwaite W, O'Farrell M, Hayes S, Harding J, Sydeman WJ, Thompson SA, Warzybok P, Bradley R, Jahncke J, Golightly RT, Schneider SR, Suryan RM, Gladics AJ, Horton CA, Kim SY, Meling SR, DeLong RL, Abell J. 2012. State of the California Current 2011-2012: Ecosystems respond to local forcing as La Niña wavers and wanes. CalCOFI Rep. 53: 41-76. [ Links ]

Carrillo CJ, Smith RC, Karl DM. 2004. Processes regulating oxygen and carbon dioxide in surface waters west of the Antarctic Peninsula. Mar. Chem. 84(3-4): 161-179. http://dx.doi.org/10.1016/j.marchem.2003.07.004 [ Links ]

Chavez FP, Takahashi T, Cai WJ, Friederich GE, Hales B, Wanninkhof R, Feely RA. 2007. Coastal oceans. In: King AWL, Dilling ZGP, Marland RA, Rose AZ, Wilbanks TJ (eds.), The First State of the Carbon Cycle Report (SOCCR): North American Carbon Budget and Implications for the Global Carbon Cycle. National Oceanic and Atmospheric Administration, National Climatic Data Center, Asheville, NC, USA, pp. 157-163. [ Links ]

De La Cruz-Orozco ME, Valdez-Holguín JE, Gaxiola-Castro G, Mariano-Matías M, Espinosa-Carreón TL. 2010. Flujos de CO2 océano-atmósfera. In: Gaxiola-Castro G, Durazo R (eds.), Dinámica del ecosistema pelágico frente a Baja California 1997-2007. SEMARNAT, INE, CICESE, UABC, México, DF, pp. 165-179. [ Links ]

Durazo R . 2015. Seasonality of the transitional region of the California Current System off Baja California. J. Geophys. Res. 120(2): 1173-1196. http://dx.doi.org/10.1002/2014JC010405 [ Links ]

Durazo R , Baumgartner TR . 2002. Evolution of oceanographic conditions off Baja California: 1997-1999. Prog. Oceanogr. 54(1-4): 7-31. http://dx.doi.org/10.1016/S0079-6611(02)00041-1 [ Links ]

Durazo R, Ramírez-Manguilar AM, Miranda LE, Soto-Mardones LA. 2010. Climatología de variables hidrográficas. In: Gaxiola-Castro G, Durazo R (eds.), Dinámica del ecosistema pelágico frente a Baja California 1997-2007. SEMARNAT, INE, CICESE, UABC, México, DF, pp. 25-57. [ Links ]

Espinoza-Carreón TL, Gaxiola-Castro G , Durazo R , De La Cruz-Orozco ME , Norzagaray-Campos M, Solana-Arellano E. 2015. Influence of anomalous subarctic water intrusion on phytoplankton production off Baja California. Cont. Shelf Res. 92: 108-121. http://dx.doi.org/10.1016/j.csr.2014.10.003 [ Links ]

Feely RA , Sabine CL, Hernández-Ayón JM , Ianson D, Hales B . 2008. Evidence for upwelling of corrosive "acidified" water onto the continental shelf. Science 320: 1490-1492. http://dx.doi.org/10.1126/science.1155676 [ Links ]

Franco AC, Hernández-Ayón JM , Beier E, Garcon V, Maske H, Paulmier A, Farber-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. 119(5): 2923-2937. http://dx.doi.org/10.1002/2013JC009337 [ Links ]

Friederich GE, Walz PM, Burczynski MG, Chavez FP. 2002. Inorganic carbon in the central California upwelling system during the 1997-1999 El Niño-La Niña event. Prog. Oceanogr. 54(1-4): 185-203. http://dx.doi.org/10.1016/S0079-6611(02)00049-6 [ Links ]

Gaxiola-Castro G , Durazo R , Lavaniegos BE , De La Cruz-Orozco M, Millán-Núñez R , Soto-Mardones L, Cepeda-Morales J. 2008. Pelagic ecosystem response to interannual variability off Baja California = Respuesta del ecosistema pelágico a la variabilidad interanual del océano frente a Baja California. Cienc. Mar. 34(2): 263-270. [ Links ]

Hales B, Strutton PG, Saraceno M, Letelier R, Takahashi T, Feely R, Sabine C, Chavez F. 2012. Satellite-based prediction ofpCO2 in coastal waters of the eastern North Pacific. Prog. Oceanogr. 103: 1-15. http://dx.doi.org/10.1016/j.pocean.2012.03.001 [ Links ]

Helm I, Jalukse L, Vilbaste M, Leito I. 2009. Micro-Winkler titration method for dissolved oxygen concentration measurement. Anal. Chim. Acta 648(2): 167-173. http://dx.doi.org/10.1016/j.aca.2009.06.067 [ Links ]

Hernández de la Torre B, Gaxiola-Castro G , Nájera-Martínez S . 2004. ENSO effects on primary production off Baja California = Efectos del ENSO en la producción primaria frente a Baja California. Cienc. Mar. 30(3): 427-441. [ Links ]

Hernández-Ayón JM, Gaxiola-Castro G, Chavez F, Takahashi T, Feely D, Sabine CL, Hales B, Lara-Lara R. 2010. Variabilidad espacial y temporal del flujo de CO2 océano-atmósfera. In: Gaxiola-Castro G , Durazo R (eds.), Dinámica del ecosistema pelágico frente a Baja California 1997-2007 . SEMARNAT, INE, CICESE, UABC, México, DF, pp. 197-207. [ Links ]

Holm-Hansen O, Lorenzen CJ, Holmes RW, Strickland JDH. 1965. Fluorometric determination of chlorophyll. J. Cons. Int. Explor. Mer 30(1): 3-15. http://dx.doi.org/10.1093/icesjms/30.L3 [ Links ]

Hoegh-Guldberg O, Cai R, Poloczanska ES, Brewer PG, Sundby S, Hilmi K, Fabry VJ, Jung S. 2014. The Ocean. In: Barros VR, et al. (eds.), Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part B: Regional Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York, NY, USA, pp. 1655-1731. [ Links ]

Huyer A, Smith RL. 1985. The signature of El Niño off Oregon, 1982-1983. J. Geophys. Res. 90(C4): 7133-7142. http://dx.doi.org/10.1029/JC090iC04p07133 [ Links ]

ENT#91;IPCCENT#93; Intergovernmental Panel on Climate Change. 2014. Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change ENT#91;Core Writing Team, RK Pachauri and LA Meyer (eds.)ENT#93;. IPCC, Geneva, Switzerland, 151 pp. [ Links ]

Lohbeck KT, Riebesell U, Collins S, Reusch TBH. 2012. Functional genetic divergence in high CO2 adaptated Emiliana huxleyi populations. Evolution 67(7): 1892-1900. http://dx.doi.org/10.1111/j.1558-5646.2012.01812.x [ Links ]

Mathis JT, Cross JN, Bates NR, Bradley Moran S, Lomas MW, Mordy CW, Stabeno PJ. 2010. Seasonal distribution of dissolved inorganic carbon and net community production on the Bering Sea shelf. Biogeosciences 7(5): 1769-1787. http://dx.doi.org/10.5194/bg-7-1769-2010 [ Links ]

Muñoz-Anderson MA, Lara-Lara JR, Álvarez-Borrego S, Bazán Guzmán C, De La Cruz-Orozco M . 2015. Water-air carbon fluxes in the coastal upwelling zone off northern Baja California = Flujos de carbono agua-aire en la zona de surgencias costeras frente al norte de Baja California. Cienc. Mar. 41(2): 157-168. http://dx.doi.org/10.7773/cm.v41i2.2484 [ Links ]

ENT#91;NOAAENT#93; National Oceanic and Atmospheric Administration. 2016. NOAA/ESRL Trends in atmospheric carbon dioxide ENT#91;accessed 16 June 2016ENT#93;: NOAA/ESRL Trends in atmospheric carbon dioxide ENT#91;accessed 16 June 2016ENT#93;: http://www.esrl.noaa.gov/gmd/ccgg/trends/ . [ Links ]

Paulmier A, Ruíz-Pino D, Garcon V . 2008. The oxygen minimum zone (OMZ) off Chile as intense source of CO2 and N2O. Cont. Shelf. Res. 28(20): 2746-2756. http://dx.doi.org/10.1016/j.csr.2008.09.012 [ Links ]

Pérez-Brunius P, López M, Parés-Sierra A, Pineda J. 2007. Comparison of upwelling indices off Baja California derived from three different wind data sources. CalCOFI Rep. 48: 204-214. [ Links ]

Reimer J, De la Cruz Orozco ME, Gaxiola-Castro G , Bazán Guzmán MC, Lara-Lara JR . 2013a. Procesamiento de datos de pCO2 superficial del océano adquiridos por un sistema de medición continua a bordo del barco. CICESE, Ensenada, México, 13 pp. [ Links ]

Reimer J J, Vargas R, Smith SV, Lara-Lara R , Gaxiola-Castro G , Hernández-Ayón JM , Castro A, Escoto-Rodríguez M, Martínez-Osuna J. 2013b. Air-sea CO2 fluxes in the near-shore and intertidal zones influenced by the California Current. J. Geophys. Res. Oceans 118(10): 4795-4810. http://dx.doi.org/10.1002/jgrc.20319 [ Links ]

Ribero-Calle S, Gnanadesikan A, Del Castillo CE, Balch WM, Guikema SD. 2015. Multidecadal increase in North Atlantic coccolithophores and the potential role of rising CO2. Science 350(6267): 1533-1537. http://dx.doi.org/10.1126/science.aaa8026 [ Links ]

Sarmiento JL, Gruber N. 2006. Ocean Biogechemical Dynamics. Princeton University Press, Princeton. [ Links ]

Schloss IR, Ferreyra GA, Ferrario ME, Almandoz GO, Codina R, Bianchi AA, Balestrini CF, Ochoa HA, Pino DR, Poisson A. 2007. Role of plankton communities in sea-air variations in pCO2 in the SW Atlantic Ocean. Mar. Ecol. Prog. Ser. 332: 93-106. http://dx.doi.org/10.3354/meps332093 [ Links ]

Takahashi T , Olafsson J, Goddard JG, Chipman DW, Sutherland SC. 1993. Seasonal variation of CO2 and nutrients in the high-latitude surface: A comparative study. Global Biogeochem. Cycles 7(4): 843-878. http://dx.doi.org/10.1029/93GB02263 [ Links ]

Takahashi T , Sutherland SC , Sweeney C, Poisson A , Metzl N , Tilbrook B, Bates N, Wanninkhof R , Feely RA , Sabine C , Olafsson J , Nojiri Y. 2002. Global sea-air CO2 flux based on climatological surface ocean pCO2, and seasonal biological and temperature effects. Deep-Sea Res. (II) 49(9-10): 1601-1622. http://dx.doi.org/10.1016/S0967-0645(02)00003-6 [ Links ]

Takahashi T , Sutherland SC , Wanninkhof R , Sweeney C, Feely RA , Chipman DW , Hales B , Friederich G, Chavez F , Sabine C , Watson A, Bakker DCE, Schuster U, Metzl N, Yoshikawa-Inoue H, Ishii M, Midorikawa T, Nojiri Y , Körtzinger A, Steinhoff T, Hoppema M, Olafsson J , Arnarson TS, Tilbrook B , Johannessen T, Olsen A, Bellerby R, Wong CS, Delille B, Bates NR , de Baar HJW. 2009. Climatological mean and decadal change in surface ocean pCO2, and net sea-air CO2 flux over the global oceans. Deep-Sea Res. (II) 56(8-10): 554-577. http://dx.doi.org/10.1016/j.dsr2.2008.12.009 [ Links ]

Venrick E, Hayward TL. 1984. Determining chlorophyll on the 1984 CalCOFI surveys. CalCOFI. Rep. 25: 74-79. [ Links ]

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

Weiss RF. 1970. The solubility of nitrogen, oxygen and argon in water and seawater. Deep-Sea Res. 17: 721-735. [ Links ]

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

Zaitsev O, Trasviña-Castro A , Linero-Cueto J, Gaxiola-Castro G , Cepeda-Morales J. 2014. Oceanographic conditions over the continental shelf off Magdalena Bay (Mexico) in 2011-2012 = Condiciones oceanográficas en la plataforma continental frente a bahía Magdalena (México) en 2011-2012. Cienc. Mar. 40(2): 89-112. http://dx.doi.org/10.7773/cm.v40i2.2314 [ Links ]

Zeebe RE. 2012. History of seawater carbonate chemistry, atmospheric CO2, and ocean acidification. Annu. Rev. Earth Planet. Sci. 40(1): 141-165. http://dx.doi.org/10.1146/annurev-earth-042711-105521 [ Links ]

Received: April 2016; Accepted: July 2016

Creative Commons License Este es un artículo publicado en acceso abierto bajo una licencia Creative Commons