You are on page 1of 15

RESEARCH ARTICLE Seasonal Changes in Carbonate Saturation State and Air‐

10.1029/2019JG005028
Sea CO2 Fluxes During an Annual Cycle in a Stratified‐
Key Points:
• North Patagonian fjords are
Temperate Fjord (Reloncaví Fjord, Chilean Patagonia)
aquaculture‐heavy used
environments, and there are not
Maximiliano J. Vergara‐Jara1,2 , Michael D. DeGrandpre3 , Rodrigo Torres2,4, Cory M. Beatty3,
available data for understanding L. Antonio Cuevas5,6 , Emilio Alarcón2,4, and José Luis Iriarte2,7,8
major biogeochemical process like
1
air‐sea CO2 fluxes and aragonite Programa de Doctorado en Ciencias de la Acuicultura, Universidad Austral de Chile, Puerto Montt, Chile, 2Centro de
saturation Investigación Dinámica de Ecosistemas Marinos de Altas Latitudes, Universidad Austral de Chile, Valdivia, Chile,
• A high‐frequency time series of the 3
Department of Chemistry and Biochemistry, University of Montana, Missoula, MT, USA, 4Laboratorio de química del
annual cycle of CO2, pH, and
dissolved oxygen in a north
carbonato, Centro de Investigación en Ecosistemas de la Patagonia, Coyhaique, Chile, 5Centro para el Estudio de
Patagonian fjord is evaluated Forzantes Múltiples sobre Sistemas Socio‐Ecológicos Marinos (MUSELS), Concepción, Chile, 6Facultad de Ciencias
• The fjord inorganic carbon cycle is Ambientales, Universidad de Concepción, Concepción, Chile, 7COPAS‐Sur Austral, Centro de Investigación
very dynamic, primarily driven by Oceanográfica en el Pacífico Sur‐Oriental, Universidad de Concepción, Concepción, Chile, 8Instituto de Acuicultura,
varying contributions from primary
Universidad Austral de Chile, Puerto Montt, Chile
production and respiration of
allochthonous organic carbon

Supporting Information:
Abstract Changes may be occurring in the carbonate chemistry of fjords due to natural and anthropogenic
• Supporting Information S1 disturbance of major freshwater sources. We present a high‐frequency time series study of seasonal pH and
CO2 partial pressure (pCO2) in a north Patagonian fjord with a focus on changes in freshwater inflows and
biological processes. To do this, we monitored pH and pCO2 in situ, along with river streamflow, salinity,
Correspondence to: temperature, and dissolved oxygen (DO) in the Reloncaví Fjord (41.5°S) for a full year (January to December
J. L. Iriarte,
2015). Strong seasonal variability was observed in the pCO2, pH, and DO of the fjord's surface waters. During
jiriarte@uach.cl
the summer, pCO2 reached its annual minimum (range: 187–571 μatm) and pH its maximum (range:
7.98–8.24), coinciding with lower freshwater inflows (204–307 m3/s) and high DO (280–378 μmol/kg), as well as
Citation:
aragonite saturation states (ΩArag) higher than 1. In contrast, in winter, pCO2 ranged from 461–1,008 μatm and
Vergara‐Jara, M. J., DeGrandpre, M. D.,
Torres, R., Beatty, C. M., Cuevas, L. A., pH from 7.57–8.03, coinciding with high freshwater inflows (1,049–1,402 m3/s), lower oxygen (216–348
Alarcón, E., & Iriarte, J. L. (2019). μmol/kg), and constant undersaturation of ΩArag. Reloncaví Fjord had an annual air‐water CO2 flux of 0.716 ±
Seasonal changes in carbonate
2.54 mol·m−2·year−1 during 2015 and thus acted as a low emission system. The annual cycle was mainly
saturation state and air‐sea CO2 fluxes
during an annual cycle in a stratified‐ governed by seasonal changes in biological processes that enhanced the shift from a CO2 sink in late spring and
temperate fjord (Reloncaví Fjord, summer, caused by high primary production rates, to a CO2 source during the rest of the year caused by high
Chilean Patagonia). Journal of
community respiration due to allochthonous organic carbon inputs.
Geophysical Research: Biogeosciences,
124, 2851–2865. https://doi.org/
10.1029/2019JG005028

1. Introduction
Received 23 FEB 2019
Fjords are areas of special importance for research due to the magnitude of the biogeochemical processes
Accepted 26 JUL 2019
Accepted article online 23 AUG 2019 that are present there, for example, they are important areas of organic carbon burial (Hinojosa et al.,
Published online 9 SEP 2019 2014; Skei, 1983; Smith et al., 2015), and their role in ecosystem services such as aquaculture (Brattegard,
1980). The characterization of the CO2 system variability is particularly important in Patagonian fjords
because many of them are important nursery areas for marine calcifiers, including species utilized for inten-
sive aquaculture (e.g., mussels). Early life stages of calcifiers invertebrates may be particularly vulnerable to
perturbation in the carbonate system induced by high levels of CO2 (Duarte, 2015; Ellis et al., 2016; Navarro
et al., 2013, 2016; Wang et al., 2015). The CO2 system of coastal waters can respond to a myriad of factors and
processes (natural or anthropogenically driven) in a wide range of spatiotemporal scales. However, due
latitudinal distribution of fjord systems, seasonal biogeochemical cycle is a dominant source of variability
but modulated by long‐term fluctuation including those associated to the global change (e.g., ocean
acidification, perturbation in the hydrological cycles, and global warming).
Highly seasonal primary production (PP) of Patagonian fjords has been attributed to a combination of phy-
sical (e.g., stratification and solar irradiation) and chemical (e.g., nutrient input) processes (Iriarte et al.,
2007; Jacob et al., 2014). PP is dominated by highly silicified, chain‐forming diatoms (Iriarte & González,
©2019. American Geophysical Union.
2008) supported by nutrients and resources of marine and continental origin; the latter is particularly true
All Rights Reserved. in fjords with direct rich freshwater inflows like Reloncaví Fjord (Silva, 2008; Torres et al., 2014;

VERGARA‐JARA ET AL. 2851


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Vandekerkhove et al., 2015). Patagonia archipelago inner waters have been suggested to be net sinks of
atmospheric CO2 during the warm season (Torres et al., 2011); a small pCO2 data during the cold season sug-
gest that this tendency could be reduced or inverse during winter time (Torres et al., 2011); however, the low
frequency of those measurements preclude any annual balance. These authors argue that the haline‐
stratified fjords of Patagonia can have a particularly strong capacity to remove atmospheric CO2 since haline
stratification not only enhance light and continental nutrients (e.g., silicic acid and iron) availability for sur-
face water phytoplankton but also enhance its efficiency (in term of nutrients) to drop pCO2, due the low
alkalinity (low buffer capacity) of surface waters. In other hand, they pointed out that while a strong halo-
cline precludes the vertical fluxes of respiration products (CO2) from below the halocline, it cannot prevent
the rain of particulate organic matter from surface to subhalocline waters, therefore intensifying the effects
of the vertical segregation between productivity (low surface water pCO2) and respiration (high subsurface
water pCO2).
Surface water pCO2 data display considerable variability in high‐latitude coastal regions, with relatively high
surface pCO2 occurring in riverine‐influenced coastal areas, probably associated to organic matter dis-
charges (net metabolism) to this coastal regions; however, areas of sea ice melt coincided with low‐surface
pCO2 (Atamanchuk et al., 2015; Burgers et al., 2017; Mørk et al., 2014). Fjords in the Northern
Hemisphere have been more intensively studied (Andersson et al., 2017; Atamanchuk et al., 2015; Borges
et al., 2004; Feely et al., 2010; Meire et al., 2015; Murray et al., 2015; Omar et al., 2016; Reisdorph &
Mathis, 2014; Reum et al., 2014; Ruiz‐Halpern et al., 2010; Rysgaard et al., 2012). In most of these studies,
seasonal changes due to PP, ice formation, ice melting, and freshwater inputs have been pointed out as
the main drivers for changes in air‐sea CO2 fluxes and the carbonate system.
Patagonian fjords occurred from 41°S to 56°S have been an interesting location for studying the carbonate
system; however, just a few studies have focused on the CO2 cycle within this vast region (Alarcón et al.,
2015; Torres et al., 2011).
The Reloncaví Fjord is the northern most fjord of Patagonia and has many freshwater sources like the Puelo
River (León‐Muñoz et al., 2013) that has been reported having a high interannual variability, driven by sea-
sonal rainfall and snowmelt regimes. In the fjord, the magnitude of the PP during seasons when blooms
occur has been associated with interannual changes in streamflow patterns (Iriarte et al., 2016). The produc-
tivity of Reloncaví Fjord has an strong seasonal variability, with relatively high PP in spring–summer (~170
mmol C·m−2·day−1) and very low in winter (<8 mmol C·m−2·day−1; González et al., 2010) likely playing a
major role in driving the seasonal carbonate system in this fjord.
Furthermore, as the northernmost fjord in South America, Reloncaví does not experience some other phy-
sical processes like ice formation or melting from glaciers at the head of the fjord (Iriarte et al., 2014). This
makes it a unique and interesting system offering a clear view of the impacts of freshwater changes on the
surface chemistry and biology of the fjord system. Importantly, recent climatic predictions for the
Patagonian fjord ecosystem lead us to expect decreasing freshwater supply to the fjord's surface water as a
result of decreased atmospheric precipitation in the coming decades (Boisier et al., 2016; Garreaud
et al., 2013).
The aim of this study was to describe the annual fluctuations of surface water inorganic carbon chemistry
and air‐sea CO2 fluxes in the fjord, through the deployment of high performance sensors and field sampling
campaigns. We discuss the role of the biological and physical processes driving the inorganic carbon chem-
istry and CO2 fluxes and their dependence on the local hydrological cycle.

2. Materials and Methods


2.1. Study Area
Reloncaví Fjord (Figure 1), located between 41°22′S and 41°44′S, is approximately 55 km long by 3 km wide
and is divided into three basins. The maximum depths are 450 m at the mouth and 50 m at the head. The
fjord is dominated by fresh estuarine surface waters and modified Sub‐Antarctic Water with salinities ran-
ging from 31–33 in the subsurface (Castillo et al., 2016; Silva et al., 2009). It presents a relatively shallow
(<5 m), continuously stratified buoyant layer and has a tidal regime that is mainly semidiurnal, with a tidal
range between 6 and 7 m during spring tides decreasing to ~1 m at neap tides (Valle‐Levinson et al., 2007).

VERGARA‐JARA ET AL. 2852


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Figure 1. (A) South America. (B) Location of the fjord area in the South American Continent. (C) Reloncaví Fjord area
map. The buoy was anchored adjacent to the river mouth (North Patagonia Buoy: http://portal.goa‐on.org/Explorer).
The legend indicates the position of the instruments used in the study. The samples were taken at exactly the same place as
the buoy.

The fjord has a three‐layer vertical circulation pattern in which the surface (<5 m) and deepest (>100 m)
layers tend to move toward the mouth (outflow layers), whereas the intermediate inflow layer (>5 and
100 m) moves toward the head of the fjord (Castillo et al., 2016; Valle‐Levinson et al., 2007). There is
exchange of dissolved inorganic nutrients between the more saline (29–32), nutrient‐rich (N and P) subsur-
face layer, and the low‐salinity (2–20), low‐nutrient (except for silicic acid) surface layer. The surface layer
rarely extends below 5 m, but there is a marked seasonal variability in its inorganic nutrient load between
spring and winter (Castillo et al., 2016; González et al., 2010). Specifically, the upper 5 m of the surface layer
is low in NO3− (<5 μM) and PO43− (<0.7 μM) in summer through autumn but high in silicic acid (>100 μM);
while in winter, silicic acid, NO3− and PO43− concentrations are all high (30–138, 10–22, and >1 μM, respec-
tively; González et al., 2010).
The sampling site was located close to the mouth of the Puelo River (Figure 1; 41°35′S, 72°20′W). The Puelo
River is the primary source of riverine freshwater inflow into the Pacific Ocean in northern Patagonia
(Dávila & Figueroa, 2002; León‐Muñoz et al., 2013) with a mean discharge of 600 m3/s. Therefore, this sta-
tion offered a representative view of the total inflow contributed by the river basin.
2.2. Field Sampling and Measurements
Samples were taken from the upper part of the surface layer (1‐, 5‐, 10‐, and 15‐m depths) from January to
December 2015 during eight field campaigns. Water samples for total alkalinity (AT) and pH determination
were collected with a 10‐L Go‐Flo bottle and stored in the dark at low temperature (<7 °C) in a 250‐ml gas‐
tight container. Seawater samples for AT analysis were poisoned with HgCl2; pH was analyzed after collec-
tion (<12 hr) using impure m‐cresol purple as an indicator at 25.0 °C with an OceanOptics STS‐Vis (350–800
nm) spectrophotometer (Byrne et al., 1988). When possible, potentiometric pH was determined at the same
time as a comparison, calibrated by using standard pH tris buffer (pH = 8.089 at 25.0 °C; Dickson et al., 2007;
Riebesell et al., 2010; DOE, 1994). All pH samples were measured in duplicate giving a reproducibility better
than 0.001 pH units when the spectrophotometric methods was used. Total alkalinity (AT) was determined
at the laboratory using an automatic potentiometric titration system (Haraldsson et al., 1997). We used cer-
tified reference material supplied by Andrew Dickson (Scripps Institution of Oceanography) to verify AT
accuracy or correct AT values; reproducibility was typically less than 2 μmol/kg.
For inorganic dissolved nutrients, 500 ml of seawater were collected from the upper layer (surface to 15 m)
during the synoptic campaigns. The water was filtered through glass fiber filters (Whatman GF/F) and

VERGARA‐JARA ET AL. 2853


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

stored frozen (−20 °C) until analysis (Parsons et al., 1984). The analysis of nutrients was done in a certified
laboratory (CERAM of the Universidad Austral de Chile at Puerto Montt). For the autotrophic biomass
(chlorophyll‐a), 250 ml of seawater were filtered through a 0.7‐μm glass fiber filter (Micro Filtration
System), extracted with 90% acetone, and measured with a fluorometer (Turner P700) as recommended by
Parsons et al. (1984).
The discrete AT, pH, inorganic dissolved nutrients, and chlorophyll‐a measurements are available in the sup-
porting information Table S1.

2.3. Sensor‐Based Measurements


High‐resolution pCO2, pH, depth, temperature, conductivity, and dissolved O2 (DO) measurements were
recorded simultaneously in situ using autonomous Submersible Autonomous Moored instrument (SAMI)‐
CO2, SAMI‐pH sensors (DeGrandpre et al., 1995; Seidel et al., 2008, Sunburst Sensors, LLC), and SBE 37
MicroCAT CTD‐ODO (SeaBird Electronics), respectively. Measurements started in the austral summer
(January 2015). All the sensors were placed at exactly the same depth mounted in one single steel frame with
sensors water intakes at same vertical position. The absolute measurement depth of the pressure sensor var-
ied between 1.0 and 3.5 m below the surface during the entire period due to the extreme Puelo River flow and
the high tidal variation. DO values were used to compute the apparent oxygen utilization (Emerson &
Hedges, 2008). All sensors recorded data hourly throughout 2015, with the exception of January and
February when data were recorded every 20 min. Gaps of 2 or 3 weeks in the time series were due to routine
maintenance and calibrations. Sensors were cleaned at 2, 7, and 10 month after deployment to prevent bio-
fouling formation, and the intake tubings were cleaned with deionized water. At month 10 postdeployment
we run an extended sensor maintenance and calibration using sensor's manufactures recommended proto-
cols. The SAMI‐CO2 sensor was calibrated using standard CO2 concentration gas tanks up to 800 ppm of
CO2. Values above the calibration range are likely to have an error (up to 5% at 1,500 ppm) due to nonlinear-
ity and insensitivity of the response at these high pCO2 levels (DeGrandpre et al., 1999). SAMI‐pH instru-
ments use an accuracy test instead of a calibration procedure (Seidel et al., 2008) with a tris pH buffer
sample using purified m‐Cresol Purple at 25 °C, giving an accuracy of ~0.005 pH units (Sunburst Sensors,
LLC). However, there could be inaccuracy in the indicator equilibrium constant over the broad pH and sali-
nity range found in the fjord. We assume that the pH values obtained at lower salinities are subject to an
error of up to ±0.02 pH units at S < 5, and pH values above 7 (Mosley et al., 2004).
The salinity and AT values from the bottle sampling presented a linear relation (r2 = 0.976; supporting infor-
mation Figure S1), and this regression was used to predict surface AT based on salinity (ATsal). The SAMI‐pH
data were compared with discrete bottle data in order to evaluate in situ sensor performance. SAMI‐pH and
ATsal data were used to compute the other carbonate parameters at the ambient temperature and salinity,
because that pair of carbonate system parameters has shown good accuracy (Cullison‐Gray et al., 2011).
The pCO2 computed from the SAMI‐pH and ATsal was compared with the in situ pCO2 data. All inorganic
carbon parameters were calculated with CO2SYS_v2.1‐2018 software (Orr et al., 2015; Pierrot et al., 2006)
modified using the K1 and K2 constants in Table 5 of Dickson and Millero (1987) for salinity = 0–40.
Nutrient data were not included in the computations, due to the lack of continuous measurements during
the study. The calculated pCO2 changes by <10 μatm when the highest observed levels of total phosphate
(2 μM) and silicate (100 μM) are included in CO2SYS.
Previous studies have shown that SAMI‐pH and SAMI‐CO2 sensors are stable over a time period of several
months, showing no significant drift during a deployment period (Omar et al., 2016), although the pCO2 sen-
sors require in situ validation during deployment (Cullison‐Gray et al., 2011; DeGrandpre et al., 1995). The
differences between discrete and sample pCO2 and pH, along with the comparison of in situ and calculated
values, are shown and discussed below. In order to address the differences obtained between the estimated
values (linear regression for ATsal) and computed values from CO2SYS using measured pH and pCO2, we
plot the delta AT correlated to salinity (Figure S4). As is known that enclosed coastal areas with high influ-
ence of major freshwater sources, like Reloncaví Fjord and the Puelo River, the dissolved organic matter
could act as organic bases enhancing AT (Cai et al., 1998; Hernández‐Ayon et al., 2007; Yang et al., 2015);
there is no empirical evidence to suggest that the Reloncaví Fjord could have allochthonous organic

VERGARA‐JARA ET AL. 2854


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

alkalinity addition. However S‐AT relationship have been found to be highly correlated in Patagonian waters
(Alarcón et al., 2015; Torres et al., 2011), suggesting that most of AT variability is driven by mixing in this
estuarine systems.

2.4. Meteorological Data


A meteorological station (HOBO‐U30; 41°41′S; 71°23′W) close to the Puelo River mouth measured air tem-
perature, solar radiation, wind speed and direction, rain, and barometric pressure every 5 min. This informa-
tion was synchronized with the buoy data in order to make air‐water CO2 flux estimates. For the
atmospheric pCO2, values of atmospheric measurements from the Earth System Research Laboratory data-
base (National Oceanic and Atmospheric Administration Marine Boundary Layer Reference 53.1°S to
17.5°S; www.esrl.noaa.gov/gmd/ccgg/mbl/data.php) were interpolated for the year 2015 and corrected with
local barometric pressure. These values assume clean marine air, and therefore, a possible error is created
due to terrestrial effects on pCO2. The error is not readily quantified because no regional pCO2 data
are available.
Streamflow information for the Puelo River was obtained from the Carrera Basilio hydrological station, the
station closest to the mouth of the river (41.6°S; 72.2°W), run by Dirección General de Aguas de Chile
(http://snia.dga.cl/BNAConsultas/reportes; Figure 1). The data consisted of hourly streamflows for the
2015 hydrological year (January to December). Because the Puelo is one of Patagonia's major rivers, this sta-
tion offered a representative view of the variability in the freshwater contributions from the Reloncaví
Fjord basin.

2.5. Flux Determinations


Air‐sea flux was calculated using the diffusive boundary layer model, from the bulk flux equation expressed
in terms of CO2 partial pressure (equation (1)):

F ¼ k*K 0 ðpCO2w –pCO2a Þ; (1)

where F is the air‐water flux (moles per area per time) and k is the gas transfer velocity (length per time) that
accounts for gas diffusion at the air‐sea boundary and was estimated using a wind speed relationship
adjusted for in situ conditions using the updated equation in Wanninkhof (2014):

k ¼ 0:251<U 2 >ðSc=660Þ−1=2 ; (2)

where <U2> is average squared wind speed adjusted to 10‐m height above sea level and Sc is the Schmidt
number, which accounts for differences in molecular diffusivity between gases. A positive F value represents
a flux from the ocean to the atmosphere. K0 is the solubility of the gas expressed in units of
concentration/partial pressure, and pCO2w and pCO2a are the partial pressures of CO2 in the surface water
and in equilibrium with the overlying air, respectively, as described in Wanninkhof (2014). Because different
months of the year have different total amounts of data points, the monthly average flux was determined for
each month and these values were used to calculate the annual mean flux. The values for atmospheric pCO2
obtained from interpolation (section 2.4) were used as input data. Atmospheric pCO2 in wet air was then cal-
culated by including water vapor pressure according to the following formula (Dickson et al., 2007):

pCO2 ðat sea surfaceÞ ¼ pCO2 ðin dry airÞ×ð1−Pw Þ; (3)

where Pw is the water vapor of seawater at in situ salinity and temperature of equilibration (Forstner &
Gnaiger, 1983).

2.6. Temperature Effect on pCO2


Because seawater pCO2 is strongly affected by changes in temperature due to the temperature dependence of
CO2 equilibria and solubility, we used the equation of Takahashi et al. (2009); equation (4)) to determine the
effect of seasonal cooling/warming on CO2 variability in the fjord.

VERGARA‐JARA ET AL. 2855


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

pCO2 at T obs ¼ ðpCO2mean Þ*exp½0:0433ðT obs –T mean Þ; (4)

where the annual mean temperature (Tmean) was 13.33 °C, the annual mean pCO2 (pCO2mean) was 643
μatm, and Tobs is the measured temperature in degrees Celsius. This equation uses the temperature coeffi-
cient for pCO2 to determine to what extent the pCO2 will vary from a fixed pCO2 (in this case the mean of
the entire time series) over the observed range of temperatures.

3. Results
3.1. Hydrographic Conditions
The Puelo River had highly contrasting flows: streamflow was at its lowest (monthly means: 204–307 m3/s)
in summer–autumn (January–April) and highest (monthly means: 713–1,402 m3/s) in winter (May–
August). During the year, the river presented several massive freshwater pulses, reaching up to >4,000
m3/s in early June (Figure 2d). Accordingly, salinity fluctuated significantly at the buoy mooring point
due to the effect of freshwater and the mooring line changing depth from tides and current (Figure 2e). In
summer the variation of salinity was due to a signal in the tidal cycle, while in winter the precipitous decline
in salinity corresponded with the massive freshwater pulses from the Puelo River (Figure 2d). During sum-
mer, salinity values fluctuated dramatically with the tidal cycle, likely because sensors were sampling
through the stratified layer during different tidal states due to the proximity of the mooring to the river
mouth. During the annual cycle the atmospheric temperature varied from −0.8 to 29 °C, while the sea sur-
face temperature was more stable and varied from 8.7 up to 19.4 °C (Figure 2f). Seawater temperature
changes occurred gradually, with the highest monthly average sea surface temperature in February and
the lowest in August (Table 1).

3.2. Data Quality and Carbonate System Consistency


The full‐year time series obtained by the pH sensors showed a good match with the field sample data
(Figure 2a), with a mean difference of 0.05 pH units (SD ± 0.059, salinity range 23.9–32.2, n = 7) using
the 5‐m depth as contrast. As stated above, the relationship between total alkalinity and salinity was deter-
mined using discrete samples collected throughout the period for the surface layer (top 15 m). The resulting
linear model for the observations (ATsal = 53.107 × salinity + 385.97 in micromoles per kilogram; R2 = 0.976,
n = 46, Table S2 and Figure S1) is an ideal tool for assessing the total alkalinity based only on salinity values.
The intercept (386 ± 34 μmol/kg) is a reasonably good match with the Puelo River end member, which had
an average AT of ~340 μmol/kg (n = 3, salinity of 0.3 PSU) in the spring.
During the annual cycle, pH and pCO2 were correlated in a negative exponential relationship (Figure 3),
with the most corrosive (lowest pH) conditions occurring in winter (Figure 3b). Data collected in higher sali-
nity water seem to present a better fit to the exponential model than the data collected in low salinity water
(Figure 3a); points that fall well below the fit are those found during winter (Figure 3b), specifically during
the months when there were massive freshwater flood pulses from the Puelo River, with correspondingly
low salinities (Figure 3a). The steep relationship between pH and pCO2 for the low salinity (<7) data can
be reproduced in CO2SYS using low constant AT (e.g., 500 μmol/kg), whereas seawater conditions fall
around the general exponential trend (Figure S3). Therefore, the large scatter shown in Figure 3 makes sense
for the large range in salinity observed during the study. Note also that the vertical or horizontal deviations
in pH or pCO2, respectively, could be fouling of the SAMI‐pH or SAMI‐CO2. Regarding pH, same steep trend
is observed in December and January (Figure 3b) and follow the same trend that modeled data from Figure
S3.
The ATsal can be used with in situ pH data to provide an in situ data evaluation (Cullison‐Gray et al., 2011).
In this case, the in situ pCO2 data were compared with pCO2 computed from in situ pH and ATsal using
CO2SYS with in situ temperature and salinity. Although there is a high linear correlation between the com-
puted and measured values observed for pH (R2 = 0.92; Figure S4a) and pCO2 (R2 = 0.91; Figure S4b), the
slope of the pH model was closer to 1 (0.98) than the lower slope obtained for the pCO2 model (0.78), show-
ing lower calculated values. The mean difference between measured pCO2 and computed pCO2 was 76 μatm
+ 131 μatm (n = 10,227), while the mean difference between measured pH and computed pH (in situ pH −
computed pH) was 0.051 ± 0.078 pH units.

VERGARA‐JARA ET AL. 2856


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Figure 2. Annual dynamics of the primary measured parameters at the Reloncaví buoy (Figure 1). Time series of pH (a),
pCO2 (b), dissolved oxygen (c), Puelo River streamflow (d), salinity (e), and air (black) and sea surface (blue) temperatures
(f). A shows data sampled at different discrete depths; the horizontal red line in (b) shows the pCO2 atmospheric value.
Depth variation from (e) is explain above.

The deviations obtained from the CO2 system calculations appear large, but they form only a small (<4%)
portion of the observed pH (~1.3 pH unit) and pCO2 (~2,000 μatm) ranges. Deviations may be caused by sev-
eral factors such as biofouling, deviation of ATsal from the linear ratio, organic sources of AT, systematic
errors at high pCO2 due to nonlinear calibration, inaccuracy of the calculated pH and CO2 due to uncertainty
of the different pKa and equilibrium constants at low salinity, and errors due to rapid changes and extreme
stratification around the individual sensors on the buoy mooring so that they measure different water due to
slightly different measurement times (Cullison‐Gray et al., 2011; Lai et al., 2016).

3.3. Year‐Round System Dynamics


The carbonate system exhibited a large range of variability over the year‐long time series (Figures 2a, 2b, and
3 and Table 1). The pH presented maximum values during midsummer (>8.0). January was the month with
the highest average pH (pH = 8.24), which then started to decrease during the autumn and reached its mini-
mum in early winter (June average pH = 7.58). The opposite is observed for pCO2 where the lowest values
were recorded in midsummer (January average pCO2 = 200 μatm); pCO2 then increased in autumn up to its
highest point in May (May average pCO2 = 1,312 μatm; Figure 2b and Table 1). This trend is followed by the
seasonal changes in DO that occur during spring–summer, where the fjord system drives high DO values in
the warmer months (>16 °C, >350 μmol/kg). During early fall and spring, DO had the highest seasonal

VERGARA‐JARA ET AL. 2857


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Table 1 correlation found for pH and pCO2 (Table S4). Toward winter, the DO
Monthly Mean Statistics of the Carbonate System (pH and pCO2), Salinity,
values gradually decreased in the surface layer (<200 μmol/kg; Figure 2
Temperature, Dissolved Oxygen (DO), and CO2 Flux in 2015
c and Table 1); in fact, the DO peak and lowest point coincide with the
DO Flux CO2 extreme values for pH recorded at the end member of the Reloncaví
pCO2 Temp (μmol/ (mmol·m Wind
−2 −1 Fjord system (Figures 2a and 2c, red arrows). The pH and DO minima
Month pH Salinity (μatm) (°C) kg) ·hr ) (m/s)
occurred at the same time that the temperature was low and river dis-
Jan 8.24 20.41 187 16.3 378 −0.422 4.17 charge was high (Figures 2d–2f). An abrupt reduction of DO (from 300
Feb 8.19 21.65 243 16.9 335 −0.296 3.66
to 180 μmol/kg; Figure 2c), an increase in pCO2 (to 500–1,000 μatm;
Mar 7.98 28.42 571 14.4 280 0.104 2.41
Apr 7.81 29.85 912 12.5 229 0.259 1.66 Figure 2b), and pH dropping consistently below 7.8 (Figure 2a) were asso-
May 7.63 29.35 1,308 11.7 176 0.446 1.86 ciated with the lowest observed river streamflow (early May, Figure 2d,
Jun 7.57 23.37 1,008 11.2 216 0.521 2.44 red arrow, and Table 1). Salinity and DO shows significant correlation
Jul 7.72 27.32 970 10.6 223 0.358 1.77 values with pCO2 during winter months (Table S4), while temperature
Aug 7.75 27.66 805 10.6 236 0.171 1.80
Sep 8.03 25.98 461 10.9 348 0.069 2.25
does not show a clear patter, and at the same time changes in temperature
Oct 8.09 28.05 471 11.9 339 −0.045 2.47 (equation (2)) explained less than 15% of the annual pCO2 variability
Nov 8.10 25.03 330 13.1 317 −0.094 2.73 (not shown).
Dec 8.16 20.15 297 14.7 348 −0.185 3.57
Max 8.24 29.85 1,308 16.9 378 0.521 4.17 Aragonite and calcite saturation are important variables of the carbonate
Min 7.57 20.15 187 10.6 176 −0.422 1.66 system within the Reloncaví Fjord, particularly because it is an area of
Mean 7.94 25.60 630 12.9 285 0.074 2.57 intense aquaculture production—principally mussel farming (Molinet
SD 0.23 3.45 361 2.2 67 0.295 0.83 et al., 2017). The aragonite saturation state (ΩArag) was computed using
Median 8.01 26.65 521 12.2 299 0.086 2.44
the pH and ATsal in combination with the measured temperature and sali-
Note. Annual values are calculated from monthly mean data. nity. There is a clear seasonal trend (Figure 4) with a marked decrease in
the saturation state from summer to autumn and then a period of consis-
tent undersaturation during late autumn and winter. Puelo River AT values at zero salinity are approxi-
mately 380 μmol/kg based on the linear regression of alkalinity data at S = 0, as stated above. The near‐
zero values of ΩArag in late autumn mark the system minimum; at the same time, the undersaturated values
found in winter, even in high salinity waters (Figure 2e), are clearly driven by low pH and high CO2 water
(Figures 2a and 2b and Table S1). There was a considerable period of time when the system was undersatu-
rated even at high salinities. ΩArag moved above the saturation line again in the spring to remain saturated
most of the time and at high salinity.

3.4. Air‐Sea CO2 Fluxes


The air‐sea CO2 fluxes varied widely throughout the entire deployment period, with values ranging from
−2.1 mmol·m−2·hr−1 (net uptake) in summer to 9.3 mmol·m−2·hr−1 in winter (Figure 5 and Table 1).
CO2 uptake started in the spring (October–December: monthly mean flux of −0.013 to −0.13 mmol·m
−2
·hr−1) and became more intense during the summer (January–February: monthly mean flux of −0.44
and −0.29 mmol·m−2·hr−1). Outgassing occurred during the rest of the year (March–September: monthly
mean flux of 0.27 mmol·m−2·hr−1; Figure 5 and Table 1). The data exhibit prevalent outgassing activity in

Figure 3. The relationship for pH and pCO2 for the data from Figure 2 colored by salinity (a) and date (b).

VERGARA‐JARA ET AL. 2858


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Figure 4. Annual in situ aragonite saturation (ΩArag) computed from total alkalinity (ATsal) and in situ pH. The solid line
shows an aragonite saturation threshold of 1, while the black arrows indicate the evolution of the system from saturated to
undersaturated and vice versa.

the fjord during the year, with a marked shift in the flux behavior in early autumn, from sink to source, and
again, in spring, suggesting that the study area behaves as a net source for atmospheric CO2 (Figure 5). The
central portion of Reloncaví Fjord had an annual flux of 0.71 mol·m−2·year−1 ± 2.48 mol·m−2·year−1.

4. Discussion
In coastal areas, the carbonate system of surface waters is modulated mainly by heating and cooling (e.g.,
DeGrandpre et al., 2002), production and respiration (Burgers et al., 2017; Cox et al., 2015; Feely et al.,
2010; Gattuso et al., 1998; Hales et al., 2005), freshwater runoff (Harris et al., 2013; Jacquet et al., 2017;
Krauss et al., 2018), transport of allochthonous nutrients and organic matter (Bhatia et al., 2013; Silva
et al., 2011; Tiwari et al., 2018), surface stratification, and advective processes such as upwelling (Beitzel
Barriquand et al., 2015; Burgers et al., 2017). Some of these factors play important roles in fjord systems
(Jolivet et al., 2015).
In Reloncaví Fjord, the increase to high dissolved oxygen values observed during spring were due to high PP,
being consistent with the higher pH and lower pCO2 values observed during this study (Table S4). High PP
rates and chlorophyll‐a (up to 3.8 g C·m−2·day−1) have been reported for the spring–summer period in this
region (González et al., 2010; Montero et al., 2011). The decrease in CO2 is associated with the negative CO2
fluxes observed in Reloncaví Fjord (Figure 5). If we take into account only the CO2 flux in summer and win-
ter, it appears that the two seasons almost cancel each other out. During the spring, pH increased and pCO2
decreased, which based on the DO correlation was due to the onset of spring–summer phytoplankton
blooms in the region (Figure 2). In the northern section of the Patagonian marine system, the classic
spring–summer diatom blooms account for the greater part (80%) of the total annual PP (Iriarte et al.,
2007). Diatoms assemblages are efficient at absorbing and exporting carbon and therefore capable of contri-
buting to the seasonal decline of pCO2 levels in the surface layers of Patagonian fjords. Recently, experimen-
tal approaches through microcosms experiments have shown high growth rates coincident with high pH
values (~8.0) during an artificially triggered phytoplankton bloom in north Patagonian waters (Iriarte
et al., 2013; Olsen et al., 2014). Following the low pCO2 values in summer, pCO2 start a marked and steady
increase to levels above the atmospheric equilibrium in autumn (March–May), which persisted through late
winter and early spring (June–August; Figure 2b).
On the other hand, the combination of lower pH and higher pCO2, as well as the drop in DO observed during
the winter months, could be attributed to a number of processes: (a) enhanced postbloom heterotrophic pro-
cesses (Montero et al., 2011; Olsen et al., 2014), (b) the mixing of oxygen‐poor deep water from below the
halocline up to the surface layer, as has been suggested for Reloncaví Fjord during winter months
(Castillo et al., 2016; González et al., 2010; León‐Muñoz et al., 2013), and (c) allochthonous organic matter
transported to the fjord (Rebolledo et al., 2015) by runoff from heavy rain events during winter months.
Evidence of short‐term variations in pCO2 and O2 have revealed that strong winds can overcome stratifica-
tion in semienclosed basins, mixing water types with different pCO2 levels (Atamanchuk et al., 2015; Bates
et al., 1998; Turk et al., 2013). In Reloncaví Fjord, we noted that periods of maximum winds have an

VERGARA‐JARA ET AL. 2859


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Figure 5. Annual dynamics of the air‐sea CO2 flux in Reloncaví Fjord.

important role in the air‐sea pCO2 exchange, especially in winter when the dominant northerlies (>5 m/s)
affect the surface layer CO2 flux dynamic (Castillo et al., 2016). During our study, the yearly mean wind
speed was 2.6 m/s (from monthly mean values), with a maximum monthly average wind speed in
January (4.29 m/s) and minimum in August (1.76 m/s). The highest wind speed was found to occur
during a winter storm in early June (13.4 m/s).
Reloncaví Fjord, at the study area, was found to be a smaller CO2 source than fjords in the Northern
Hemisphere at higher latitudes, like Roskilde Fjord, North Zealand, Denmark, which has been reported
to be an annual CO2 source (3.9 mol CO2·m−2·year−1; Mørk et al., 2014), and Koljo Fjord, Sweden, which
also showed a marked transition in pCO2 dynamics from autumn to spring (Atamanchuk et al., 2015). In
Friday Harbor in the glacially formed San Juan Archipelago (Washington, USA), pCO2 ranged from ~300
to ~1,100 μatm over a 2‐year period, mostly with high positive ΔpCO2, suggesting that the sea is a source
for atmospheric CO2 in this area (Murray et al., 2015). On the other hand, Kongsfjorden and
Tempelfjorden Arctic fjords (Svalbard, Norway) appeared to be less variable systems (in terms of the carbo-
nate dynamics and saturation state of ΩArag) and they apparently act as CO2 sink systems (Fransson et al.,
2016, 2017). In the case of the Godthåbsfjord system (SW Greenland), the input of fresh glacial meltwater
produces a strong CO2 uptake giving a net negative CO2 flux within this fjord system (Meire et al., 2015).
Because of the high heterogeneity between these fjords‐estuary ecosystems and their middle‐ to high‐
latitude occurrence, is hard to establish a direct CO2 source‐sink gradient, in more cases the autotrophic‐
heterotrophic nature of them and the proximity to densely populated areas could play a major role, particu-
larly under the present scenario of high and increasing atmospheric CO2 (Feely et al., 2010; Gattuso et al.,
1998; Vieira Borges et al., 2004).
A year‐long, high‐temporal resolution record of parameters such as Ωara, ATsal, pH, pCO2, and CO2 fluxes
provides a valuable tool for present and future assessment of ocean acidification in studies focused on pro-
cesses involved in the natural variability of the carbonate system. It is therefore especially necessary to
obtain ΩArag saturation values, given the importance of this mineral in the life cycles of calcifying organisms
in natural habitats, as well as in intensive mussel aquaculture zones like Reloncaví Fjord. The dynamics and
changing conditions of the carbonate system can affect different aspects of organisms, from their ecological
behavior to their physiological functioning, especially in fisheries and aquaculture development (Broitman
et al., 2017; Haigh et al., 2015; Vargas et al., 2017), the two most economically important industries occurring
in Patagonian coastal waters.
The saturation state of seawater in terms of Aragonite (ΩArag) in Reloncaví Fjord was maximum in summer
and minimum in winter consistently with previously for the adjacent Reloncavi sound (Alarcón et al., 2015).
ΩArag can be highly variable in coastal areas (Harris et al., 2013; Xu et al., 2017), some regions at the open
ocean (Hauri et al., 2015; Jiang et al., 2015), and inland seas like fjords and bays from different regions
(Feely et al., 2010; Fransson et al., 2015, 2016; Jantzen et al., 2013; Kapsenberg et al., 2015), driving challen-
ging conditions for calcifying species inhabiting those systems. Water undersatured in calcium carbonate in
the form of aragonite (ΩArag < 1) can have deleterious effects growth and survival implications for organisms
dwelling near the surface across a broad spectrum of taxa (Cornwall et al., 2013; Duarte et al., 2013; Gattuso

VERGARA‐JARA ET AL. 2860


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

et al., 2015; Turley & Gattuso, 2012). Experimental work in the context of ocean acidification, using the mus-
sel Mytilus chilensis, suggest that under high pCO2 conditions the scope for growth (SfG is the subtraction of
energy spent in feces and metabolism from the energy consumed, Naylor et al., 1989), metabolism, and cal-
cification of this mussel decreases (Duarte et al., 2014; Navarro et al., 2016, 2013). Therefore, changes in this
inorganic carbon speciation can have effects on the aquaculture industry and coastal communities by
reduced SfG. Additionally, the net rate of calcium deposition and total weight has been found to be nega-
tively affected by high pCO2 to low pH (Navarro et al., 2016). Particularly, the mussel industry in Chile
depends on the harvesting of seeds from the natural environment, where the Reloncaví Fjord plays a very
important role by providing most of the seeds (≈80% of all seeds collectors; Molinet et al., 2017), at the same
time, as harvesting extends its period from February up to August (middle winter; Viviana Videla unpub-
lished data, 2019), exist a broad time frame were critical production activities (e.g. mussel seed harvesting)
have place under harsh acidic conditions (Ωara < 1), situation that has not been studied and may be of crucial
interest for mussels farmers. Considering that it is projected to have extended periods of up to 6 months/year
of Ωara undersaturation in surrounding seas by the end of the century (Hauri et al., 2015), we should expect
to see an expansion in the winter Ωara undersaturated period, highlighting the urgency to count with stable
and continuous environmental monitoring of carbonate system parameters in order to have the ability to
properly study the scope from the different drivers and stressors (local and global). In the other hand, the
most important aquaculture resource by value in Chile, the Atlantic salmon (Salmo salar L), the heavily
farmed salmon in Chile, is also seriously impacted by high pCO2 levels by decreased growth and deteriora-
tion of important physiological parameters such as reduction in plasma chloride (Fivelstad et al., 2015).
These conditions should be of concern to the fjord region of northern Patagonia, considering the extent of
mussel and salmon farms in this large marine system.

5. Conclusions
Acknowledgments These time series data enabled us to make an initial estimate of the CO2 flux over seasonal periods with sig-
This research was funded by nificantly contrasting biological and hydrological processes. A low CO2 efflux (0.65 mol·m−2·year−1) was
CONICYT‐FONDECYT 1141065 (J. L.
Iriarte) and is partially included in the estimated, characterizing this fjord system as a net source. The gradual changes and marked trends of
framework of Research Program 1 of pCO2, pH, ΩArag, and O2 can mostly be accounted for by biological processes, especially PP in the spring–
the IDEAL Center (Grant 15150003).
summer seasons and respiration in the autumn–winter seasons. Temperature was not an important driver
Partial funding was provided by
CONICYT‐FONDECYT 1140385 (R. within the carbonate system. Phytoplankton production was certainly a significant and major factor in the
Torres), Ocean Certain EU‐FP7 603773 low pCO2 observed in spring–summer, as evidenced by the high oxygen and pH values. During the study
(J. L. Iriarte), and DID‐UACh. Special
period, hydrological‐meteorological phenomena were responsible for the rapid, abrupt changes in the fjord
thanks to Manuel Díaz and Dr. Carlos
Molinet for making the map and system, especially regarding the massive freshwater pulses from the Puelo River (up to 4,000 m3/s). Recent
providing the meteorological data. Data climatic predictions for the Patagonian fjord ecosystem lead us to expect decreasing freshwater supply to the
presented are part of the PhD Thesis of
M. V. J. at UACh. During this study, M.
fjord's surface water as a result of decreased atmospheric precipitation in the coming decades (Boisier et al.,
V. J. was receiving financial support 2016; Garreaud et al., 2013). Furthermore, Reloncaví Fjord is under intensive pressure from an aquaculture
from a CONICYT Scholarship (Beca industry that depends entirely on the ecosystem services that the fjord provides (natural mussel seeds collec-
Doctorado Nacional 2015 # 21150285).
M. DeGrandpre received funding from
tion site, protected and brackish coast that helps with the sea lice and gives an ideal smoltification environ-
the U.S. National Science Foundation ment for salmonids; Soto et al., 2019).
(Grant OCE‐1459255). L. A. Cuevas was
supported by Millennium Nucleus
Project MUSELS funded by MINECON
NC1200286. The authors gratefully
References
acknowledge the insightful comments Alarcón, E., Valdés, N., & Torres, R. (2015). Calcium carbonate saturation state in an area of mussels culture in the Reloncaví Sound,
and suggestions of two anonymous Northern Patagonia, Chile. Latin American Journal of Aquatic Research, 43(2), 277–281. https://doi.org/10.3856/vol43‐issue2‐fulltext‐1
reviewers that helped to improve this Andersson, A., Falck, E., Sjöblom, A., Kljun, N., Sahlée, E., Omar, A. M., & Rutgersson, A. (2017). Air‐sea gas transfer in high Arctic fjords.
manuscript. All sensor data (SBE 37 Geophysical Research Letters, 44, 2519–2526. https://doi.org/10.1002/2016GL072373
MicroCAT CTD‐ODO, SAMI‐pH, and Atamanchuk, D., Kononets, M., Thomas, P. J., Hovdenes, J., Tengberg, A., & Hall, P. O. J. (2015). Continuous long‐term observations of the
SAMI‐CO2) used for this study are carbonate system dynamics in the water column of a temperate fjord. Journal of Marine Systems, 148, 272–284. https://doi.org/10.1016/j.
publicly available and can be accessed jmarsys.2015.03.002
at https://figshare.com/articles/Puelo_ Bates, N. R., Takahashi, T., Chipman, S. W., & Knap, A. H. (1998). Variability of pCO2 on diel to seasonal timescales in the Sargasso Sea
Bouy/7754258. Finally, as this is M. V. near Bermuda. Journal of Geophysical Research, 103(C8), 15567–15585. https://doi.org/10.1029/98JC00247
J.'s first research paper, he would like to Beitzel Barriquand, T., Bouruet‐Aubertot, P., Cuypers, Y., Vivier, F., Lourenço, A., & Le Goff, H. (2015). The impacts of stratification on
dedicate it to his wife Leslie and sons high latitude ocean mixing: A case study of internal waves in Storfjorden, Svalbard. Continental Shelf Research, 110, 162–182. https://doi.
Camilo and Nacho and his marine org/10.1016/j.csr.2015.10.001
biologist father, Toño, and Pepa, his Bhatia, M., Kujawinski, E. B., Das, S. B., Breier, C. F., Henderson, P. B., & Charette, M. A. (2013). Greenland meltwater as a significant and
mother, in gratitude for their support. potentially bioavailable source of iron to the ocean. Nature Geoscience, 6(4), 274–278. https://doi.org/10.1038/NGEO1746

VERGARA‐JARA ET AL. 2861


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Boisier, J. P., Rondanelli, R., Garreaud, R. D., & Muñoz, F. (2016). Anthropogenic and natural contributions to the Southeast Pacific pre-
cipitation decline and recent megadrought in central Chile. Geophysical Research Letters, 43, 413–421. https://doi.org/10.1002/
2015GL067265
Borges, A. V., Delille, B., Schiettecatte, L.‐S., Gazeau, F., Abril, G., & Frankignoulle, M. (2004). Gas transfer velocities of CO2 in three
European estuaries (Randers Fjord, Scheldt, and Thames). Limnology and Oceanography, 49(5), 1630–1641. https://doi.org/10.4319/
lo.2004.49.5.1630
Brattegard, T. (1980). Why biologists are interested in fjords. In H. J. Freeland, D. M. Farmer, & C. Levings (Eds.), Fjord Oceanography, (pp.
53–60).
Broitman, B. R., Halpern, B. S., Gelcich, S., Lardies, M. A., Vargas, C. A., Vásquez‐Lavín, F., et al. (2017). Dynamic interactions among
boundaries and the expansion of sustainable Aquaculture. Frontiers in Marine Science, 4, 15. https://doi.org/10.3389/fmars.2017.00015
Burgers, T. M., Miller, L. A., Thomas, H., Else, B. G. T., Gosselin, M., & Papakyriakou, T. (2017). Surface water pCO2 variations and sea‐air
CO2 fluxes during summer in the eastern Canadian Arctic. Journal of Geophysical Research: Oceans, 122, 9663–9678. https://doi.org/
10.1002/2017JC013250
Byrne, R. H., Robert‐Baldo, G., Thompson, S. W., & Chen, C. T. A. (1988). Seawater pH measurements: an at‐sea comparison of spectro-
photometric and potentiometric methods. Deep Sea Research Part A. Oceanographic Research Papers, 35(8), 1405–1410. https://doi.org/
10.1016/0198‐0149(88)90091‐x
Cai, W. J., Wang, Y., & Hodson, R. E. (1998). Acid‐base properties of dissolved organic matter in the estuarine waters of Georgia, USA.
Geochimica et Cosmochimica Acta, 62(3), 473–483. https://doi.org/10.1016/S0016‐7037(97)00363‐3
Castillo, M. I., Cifuentes, U., Pizarro, O., Djurfeldt, L., & Caceres, M. (2016). Seasonal hydrography and surface outflow in a fjord with deep
sill: The Reloncavi Fjord, Chile. Ocean Science, 12(2), 533–544. https://doi.org/10.5194/os‐12‐533‐2016
Cornwall, C. E., Hepburn, C. D., McGraw, C. M., Currie, K. I., Pilditch, C. A., Hunter, K. A., et al. (2013). Diurnal fluctuations in seawater
pH influence the response of a calcifying macroalga to ocean acidification. Proceedings of the Royal Society, 280(1772). https://doi.org/
10.1098/rspb.2013.2201
Cox, T. J. S., Maris, T., Soetaert, K., Kromkamp, J. C., Meire, P., & Meysman, F. (2015). Estimating primary production from oxygen time series:
A novel approach in the frequency domain. Limnology and Oceanography: Methods, 13(10), 529–552. https://doi.org/10.1002/lom3.10046
Cullison‐Gray, S., DeGrandpre, M. D., Moore, T. S., Martz, T. R., Friederich, G. E., & Johnson, K. S. (2011). Applications of in situ pH
measurements for inorganic carbon calculations. Marine Chemistry, 125(1–4), 82–90. https://doi.org/10.1016/j.marchem.2011.02.005
Dávila, P. M., Figueroa, D., & Müller, E. (2002). Freshwater input into the coastal ocean and its relation with the salinity distribution off
austral Chile (35‐55°S). Continental Shelf Research, 22(3), 521–534. https://doi.org/10.1016/S0278‐4343(01)00072‐3
DeGrandpre, M. D., Baehr, M. M., & Hammar, T. R. (1999). Calibration‐free optical chemical sensors. Analytical Chemistry, 71(6),
1152–1159. https://doi.org/10.1021/ac9805955
DeGrandpre, M. D., Hammar, T. R., Smith, S. P., & Sayles, F. L. (1995). In situ measurements of seawater pCO2. Limnology &
Oceanography, 40(5), 969–975. https://doi.org/10.4319/lo.1995.40.5.0969
DeGrandpre, M. D., Olbu, G. J., Beatty, C. M., & Hammar, T. R. (2002). Air–sea CO2 fluxes on the US middle Atlantic bight. Deep‐Sea
Research II, 49(20), 4355–4367. https://doi.org/10.1016/S0967‐0645(02)00122‐4
Dickson, A. G., & Millero, F. J. (1987). A comparison of the equilibrium constants for the dissociation of carbonic acid in seawater media.
Deep Sea Research Part A, 34(10), 1733–1743. https://doi.org/10.1016/0198‐0149(87)90021‐5
Dickson, A. G., Sabine, C. L., & Christian, J. R. (2007). Guide to best practices for ocean CO2 measurements, (Vol. 3, p. 191). PICES Special
Publication. https://doi.org/10.1039/9781847550835
DOE (1994). SOP 6: Determination of the pH of sea water using a glass /reference electrode cell., & SOP 7: Determination of the pH of sea
water using the indicator dye m‐cresol purple. In A. G. Dickson & C. Goyet (Eds.), Handbook of methods for the analysis of the various
parameters of the carbon dioxide system in sea water; version 2.1, ORNL/CDIAC‐74 (p. 187). U. S. Department of Energy.
Duarte, C., Navarro, J. M., Acuña, K., Torres, R., Manríquez, P. H., Lardies, M. A., et al. (2015). Intraspecific variability in the response of
the edible mussel Mytilus chilensis (Hupe) to ocean acidification. Estuaries and Coasts, 38(2), 590–598. https://doi.org/10.1007/s12237‐
014‐9845‐y
Duarte, C., Navarro, J. M., Acuña, K., Torres, R., Manríquez, P. H.,Lardies, M. A., et al. (2014). Combined effects of temperature and ocean
acidification on the juvenile individuals of the mussel Mytilus chilensis. Journal of Sea Research, 85, 308–314. https://doi.org/10.1016/j.
seares.2013.06.002
Duarte, C. M., Hendriks, I. E., Moore, T. S., Olsen, Y. S., Steckbauer, A., Ramajo, L., et al. (2013). Is Ocean acidification an open‐ocean
syndrome? Understanding anthropogenic impacts on seawater pH. Estuaries and Coasts, 36(2), 221–236. https://doi.org/10.1007/
s12237‐013‐9594‐3
Ellis, R. P., Urbina, M. A., & Wilson, R. W. (2016). Lessons from two high CO2 worlds—Future oceans and intensive aquaculture. Global
Change Biology, 23(6), 2141–2148. https://doi.org/10.1111/gcb.13515
Emerson, S., & Hedges, J. (2008). Chemical oceanography and the marine carbon cycle, (p. 475). Cambridge University Press. https://doi.
org/10.1017/CBO9780511793202
Feely, R. A., Alin, S. R., Newton, J., Sabine, C. L., Warner, M., Devol, A., et al. (2010). The combined effects of ocean acidification, mixing,
and respiration on pH and carbonate saturation in an urbanized estuary. Estuarine, Coastal and Shelf Science, 88(4), 442–449. https://
doi.org/10.1016/j.ecss.2010.05.004
Fivelstad, S., Kvamme, K., Handeland, S., Fivelstad, M., Olsen, A. B., & Hosfeld, C. D. (2015). Growth and physiological models for Atlantic
salmon (Salmo salar L.) parr exposed to elevated carbon dioxide concentrations at high temperature. Aquaculture, 436, 90–94. https://
doi.org/10.1016/j.aquaculture.2014.11.002
Forstner, H., & Gnaiger, E. (1983). Calculation of Equilibrium Oxygen Concentration. Polarographic Oxygen Sensors, 321–333. https://doi.
org/10.1007/978‐3‐642‐81863‐9_28
Fransson, A., Chierici, M., Hop, H., Findlay, H. S., Kristiansen, S., & Wold, A. (2016). Late winter‐to‐summer change in ocean acidification
state in Kongsfjorden, with implications for calcifying organisms. Polar Biology, 39(10), 1841–1857. https://doi.org/10.1007/s00300‐016‐
1955‐5
Fransson, A., Chierici, M., Nomura, D., Granskog, M. A., Kristiansen, S., Martma, T., & Nehrke, G. (2015). Effect of glacial drainage water
on the CO2 system and ocean acidification state in an arctic tidewater‐glacier fjord during two contrasting years. Journal of Geophysical
Research: Oceans, 120, 2413–2429. https://doi.org/10.1002/2014JC010320
Fransson, A., Chierici, M., Skjelvan, I., Olsen, A., Assmy, P., Peterson, A. K., et al. (2017). Effects of sea‐ice and biogeochemical processes
and storms on under‐ice water fCO2 during the winter‐spring transition in the high Arctic Ocean: Implications for sea‐air CO2 fluxes.
Journal of Geophysical Research: Oceans, 122, 5566–5587. https://doi.org/10.1002/2016JC012478

VERGARA‐JARA ET AL. 2862


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Garreaud, R., Lopez, P., Minvielle, M., & Rojas, M. (2013). Large‐scale control on the Patagonian climate. American Meteorological Society,
26(1), 215–230. https://doi.org/10.1175/JCLI‐D‐12‐00001.1
Gattuso, J.‐P., Frankignoulle, M., & Wollast, R. (1998). Carbon and carbonate metabolism in coastal aquatic ecosystems. Annual Review of
Ecology and Systematics, 29(1), 405–434. https://doi.org/10.1146/annurev.ecolsys.29.1.405
Gattuso, J.‐P., Magnan, A., Billé, R., Cheung, W. W. L., Howes, E. L., Joos, F., et al. (2015). Contrasting futures for ocean and society from
different anthropogenic CO2 emissions scenarios. Nature, 349(6243), aac4722. https://doi.org/10.1126/science.aac4722
González, H. E., Calderón, M. J., Castro, L., Clement, A., Cuevas, L. A., Daneri, G., et al. (2010). Primary production and plankton dynamics
in the Reloncaví Fjord and the interior sea of Chiloé, Northern Patagonia, Chile. Marine Ecology Progress Series, 402, 13–30. https://doi.
org/10.3354/meps08360
Haigh, R., Ianson, D., Holt, C. A., Neate, H. E., & Edwards, A. M. (2015). Effects of ocean acidification on temperate coastal marine eco-
systems and fisheries in the Northeast Pacific. PLoS ONE, 10(2). https://doi.org/10.1371/journal.pone.0117533
Hales, B., Takahashi, T., & Bandstra, L. (2005). Atmospheric CO2 uptake by a coastal upwelling system. Global Biogeochemical Cycles, 19,
GB1009. https://doi.org/10.1029/2004GB002295
Haraldsson, C., Anderson, L. G., Hassellöv, M., Hulth, S., & Olsson, K. (1997). Rapid, high‐precision potentiometric titration of
alkalinity in ocean and sediment pore waters. Deep Sea Research Part I, 44(12), 2031–2044. https://doi.org/10.1016/S0967‐
0637(97)00088‐5
Harris, K. E., DeGrandpre, M. D., & Hales, B. (2013). Aragonite saturation state dynamics in a coastal upwelling zone. Geophysical Research
Letters, 40, 2720–2725. https://doi.org/10.1002/grl.50460
Hauri, C., Friedrich, T., & Timmermann, A. (2015). Abrupt onset and prolongation of aragonite undersaturation events in the Southern
Ocean. Nature Climate Change, 6(2), 172–176. https://doi, https://doi.org/10.1038/nclimate2844
Hernández‐Ayon, J., Zirino, A., Dickson, A., Camiro‐Vargas, T., & Valenzuela, E. (2007). Estimating the contribution of organic bases from
microalgae to the titration alkalinity in coastal seawaters. Limnology and Oceanography Methods, 5(7), 225–232. https://doi.org/10.4319/
lom.2007.5.225
Hinojosa, J. L., Moy, C. M., Stirling, C. H., Wilson, G. S., & Eglinton, T. I. (2014). Carbon cycling and burial in New Zealand's fjords.
Geochemistry, Geophysics, Geosystems, 15, 4047–4063. https://doi.org/10.1002/2014GC005433
Iriarte, J. L., & González, H. E. (2008). Phytoplankton bloom ecology of the inner sea of Chiloé, Southern Chile. Nova Hedwigia, 133, 67–79.
Iriarte, J. L., González, H. E., Liu, K. K., Rivas, C., & Valenzuela, C. (2007). Spatial and temporal variability of chlorophyll and primary
productivity in surface waters of Southern Chile (41.5‐43° S). Estuarine, Coastal and Shelf Science, 74(3), 471–480. https://doi.org/
10.1016/j.ecss.2007.05.015
Iriarte, J. L., León‐Muñoz, J., Marcé, R., Clément, A., & Lara, C. (2016). Influence of seasonal freshwater streamflow regimes on phyto-
plankton blooms in a Patagonian fjord. New Zealand Journal of Marine and Freshwater Research, 51(2), 304–315. https://doi.org/
10.1080/00288330.2016.1220955
Iriarte, J. L., Pantoja, S., & Daneri, G. (2014). Oceanographic processes in Chilean fjords of Patagonia: From small to large‐scale studies.
Progress in Oceanography, 129, 1–7. https://doi.org/10.1016/j.pocean.2014.10.004
Iriarte, J. L., Pantoja, S., González, H. E., Silva, G., Paves, H., Labbé, P., et al. (2013). Assessing the micro‐phytoplankton response to nitrate
in Comau Fjord (42°S) in Patagonia (Chile), using a microcosms approach. Environmental Monitoring Assessment, 185(6), 5055–5070.
https://doi.org/10.1007/s10661‐012‐2925‐1
Jacob, B. G., Tapia, F. J., Daneri, G., Iriarte, J. L., Montero, P., Sobarzo, M., & Quiñones, R. A. (2014). Springtime size‐fractionated primary
production across hydrographic and PAR‐light gradients in Chilean Patagonia (41‐50 S). Progress in Oceanography, 129, 75–84. https://
doi.org/10.1016/j.pocean.2014.08.003
Jacquet, J., McCoy, S. W., McGrath, D., Nimick, D. A., Fahey, M., O'kuinghttons, J., et al. (2017). Hydrologic and geomorphic changes
resulting from episodic glacial lake outburst floods: Rio Colonia, Patagonia, Chile. Geophysical Research Letters, 44, 854–864. https://doi.
org/10.1002/2016GL071374
Jantzen, C., Häussermann, V., Försterra, G., Laudien, J., Ardelan, M., Maier, S., & Richter, C. (2013). Occurrence of a cold‐water coral along
natural pH gradients (Patagonia, Chile). Marine Biology, 160(10), 2597–2607. https://doi.org/10.1007/s00227‐013‐2254‐0
Jiang, L.‐Q., Feely, R. A., Carter, B. R., Greeley, D. J., Gledhill, D. K., & Arzayus, K. M. (2015). Climatological distribution of aragonite
saturation state in the global oceans. Global Biogeochemical Cycles, 29, 1656–1673. https://doi.org/10.1002/2015GB005198
Jolivet, A., Asplin, L., Strand, Ø., Thebault, J., & Chauvaud, L. (2015). Coastal upwelling in Norway recorded in Great Scallop shells.
Limnology & Oceanography, 60(4), 1265–1275. https://doi.org/10.1002/lno.10093
Kapsenberg, L., Kelley, A. L., Shaw, E. C., Martz, T. R., & Hofmann, G. E. (2015). Near‐shore Antarctic pH variability has implications for
the design of ocean acidification experiments. Scientific Reports, 5(1), 9638. https://doi.org/10.1038/srep09638
Krauss, K. W., Noe, G. B., Duberstein, J. A., Conner, W. H., Stagg, C. L., Cormier, N., et al. (2018). The role of the upper tidal estuary in
wetland blue carbon storage and flux. Global Biogeochemical Cycles, 32, 817–839. https://doi.org/10.1029/2018GB005897
Lai, C.‐Z., DeGrandpre, M. D., Wasser, B. D., Brandon, T. A., Clucas, D. S., & Jaqueth, E. J. (2016). Spectrophotometric measurement of
freshwater pH with purified meta‐cresol purple and phenol red. Limnology & Oceanography: Methods, 14(12), 864–873. https://doi.org/
10.1002/lom3.10137
León‐Muñoz, J., Marcé, R., & Iriarte, J. L. (2013). Influence of hydrological regime of an Andean river on salinity, temperature and oxygen
in a Patagonia fjord, Chile. New Zealand Journal of Marine and Freshwater Research, 47(4), 515–528. https://doi.org/10.1080/
00288330.2013.802700
Meire, L., Søgaard, D. H., Mortensen, J., Meysman, F. J. R., Soetaert, K., Arendt, K. E., et al. (2015). Glacial meltwater and primary pro-
duction are drivers of strong CO2 uptake in fjord and coastal waters adjacent to the Greenland ice sheet. Biogeosciences, 12(8), 2347–2363.
https://doi.org/10.5194/bg‐12‐2347‐2015
Molinet, C., Díaz, M., Marín, S. L., Astorga, M. P., Ojeda, M., Cares, L., & Asencio, E. (2017). Relation of mussel spatfall on natural and
artificial substrates: Analysis of ecological implications ensuring long‐term success and sustainability for mussel farming. Aquaculture,
467, 211–218. https://doi.org/10.1016/j.aquaculture.2016.09.019
Montero, P., Daneri, G., González, H. E., Iriarte, J. L., Tapia, F. J., Lizárraga, L., et al. (2011). Seasonal variability of primary production in a
fjord ecosystem of the Chilean Patagonia: Implications for the transfer of carbon within pelagic food webs. Continental Shelf Research,
31(3‐4), 202–215. https://doi.org/10.1016/j.csr.2010.09.003
Mørk, E. T., Sørensen, L. L., Jensen, B., & Sejr, M. K. (2014). Air‐sea CO2 gas transfer velocity in a shallow estuary. Boundary‐Layer
Meteorology, 151(1), 119–138. https://doi.org/10.1007/s10546‐013‐9869‐z
Mosley, L. M., Husheer, S. L. G., & Hunter, K. A. (2004). Spectrophotometric pH measurement in estuaries using thymol blue and m‐cresol
purple. Marine Chemistry, 91(1‐4), 175–186. https://doi.org/10.1016/j.marchem.2004.06.008

VERGARA‐JARA ET AL. 2863


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Murray, J. W., Roberts, E., Howard, E., O'Donnell, M., Bantam, C., Carrington, E., et al. (2015). An inland sea high nitrate‐low chlorophyll
(HNLC) region with naturally high pCO2. Limnology & Oceanography, 60(3), 957–966. https://doi.org/10.1002/lno.10062
Navarro, J. M., Duarte, C., Manríquez, P. H., Lardies, M. A., Torres, R., Acuña, K., et al. (2016). Ocean warming and elevated carbon
dioxide: Multiple stressor impacts on juvenile mussels from southern Chile. ICES Journal of Marine Science, 73(3), 764–771. https://doi.
org/10.1093/icesjms/fsv249
Navarro, J. M., Torres, R., Acuña, K., Duarte, C., Manriquez, P. H., Lardies, M., et al. (2013). Impact of medium‐term exposure to elevated
pCO2 levels on the physiological energetics of the mussel Mytilus chilensis. Deep Sea Research Part II, 90(3), 1242–1248. https://doi.org/
10.1016/j.chemosphere.2012.09.063
Naylor, C., Maltby, L., & Calow, P. (1989). Scope for growth in Gammarus pulex, a freshwater benthic detritivore. Hydrobiologia, 188(189),
517–523.
Olsen, L. M., Hernández, K. L., van Ardelan, M., Iriarte, J. L., Sánchez, N., González, H. E., et al. (2014). Responses in the microbial food
web to increased rates of nutrient supply in a southern Chilean fjord: possible implications of cage aquaculture. Aquaculture
Environment Interactions, 6(1), 11–27. https://doi.org/10.3354/aei00114
Omar, A. M., Skjelvan, I., Erga, S. R., & Olsen, A. (2016). Aragonite saturation states and pH in western Norwegian fjords: Seasonal cycles
and controlling factors, 2005‐2009. Ocean Science, 12(4), 937–951. https://doi.org/10.5194/os‐12‐937‐2016
Orr, J. C., Epitalon, J. M., & Gattuso, J. P. (2015). Comparison of ten packages that compute ocean carbonate chemistry. Biogeosciences,
12(5), 1483–1510. https://doi.org/10.5194/bg‐12‐1483‐2015
Parsons, T. R., Maita, Y., & Lalli, C. M. (1984). Section 8: Counting, media and preservatives. In A manual of chemical and biological
methods for seawater analysis (p. 173). Toronto: Pergamon Press.
Pierrot, D., Lewis, E., & Wallace, D.W.R. (2006). MS Excel program developed for CO2 system calculations. ORNL/CDIAC‐105a. Carbon
Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, Tennessee. https://doi.
org/10.3334/CDIAC/otg.CO2SYS_XLS_CDIAC105a
Rebolledo, L., Lange, C. B., Bertrand, S., Muñoz, P., Salamanca, M., Lazo, P., et al. (2015). Late Holocene precipitation variability recorded
in the sediments of Reloncaví Fjord (41°S, 72°W), Chile. Quaternary Research, 84(1), 21–36. https://doi.org/10.1016/j.yqres.2015.05.006
Reisdorph, S. C., & Mathis, J. T. (2014). The dynamic controls on carbonate mineral saturation states and ocean acidification in a glacially
dominated estuary. Estuarine, Coastal and Shelf Science, 144, 8–18. https://doi.org/10.1016/j.ecss.2014.03.018
Reum, J. C. P., Alin, S. R., Feely, R. A., Newton, J., Warner, M., & McElhany, P. (2014). Seasonal carbonate chemistry covariation with
temperature, oxygen, and salinity in a fjord estuary: implications for the design of ocean acidification experiments. PLoS ONE,
9(2). https://doi.org/10.1371/journal.pone.0089619
Riebesell, U., Fabry, V.J., Hansson, L., & Gattuso, J.‐P. (2010). Guide to best practices in ocean acidification research and data reporting.
Report EUR 24328 EN. https://doi.org/10.2777/58454
Ruiz‐Halpern, S., Sejr, M. K., Duarte, C. M., Krause‐Jensen, D., Dalsgaard, T., Dachs, J., & Rysgaard, S. (2010). Air‐water exchange and vertical
profiles of organic carbon in a subarctic fjord. Limnology & Oceanography, 55(4), 1733–1740. https://doi.org/10.4319/lo.2010.55.4.1733
Rysgaard, S., Mortensen, J., Juul‐Pedersen, T., Sørensen, L. L., Lennert, K., Søgaard, D. H., et al. (2012). High air‐sea CO2 uptake rates in
nearshore and shelf areas of Southern Greenland: temporal and spatial variability. Marine Chemistry, 128‐129, 26–33. https://doi.org/
10.1016/j.marchem.2011.11.002
Seidel, M. P., DeGrandpre, M. D., & Dickson, A. G. (2008). A sensor for in situ indicator‐based measurements of seawater pH. Marine
Chemestry, 109(1‐2), 18–28. https://doi.org/10.1016/j.marchem.2007.11.013
Silva, N. (2008). Dissolved oxygen, pH, and nutrients in the Austral Chilean channels and fjords. Progress in the Oceanography Knowledge
of Chilean Interior Waters, from Puerto Montt to Cape Horn. Valparaíso.
Silva, N., Rojas, N., & Fedele, A. (2009). Water masses in the Humboldt Current System: Properties, distribution, and the nitrate deficit as a
chemical water mass tracer for Equatorial Subsurface Water off Chile. Deep‐Sea Research II, 56(16), 1004–1020. https://doi.org/10.1016/
j.dsr2.2008.12.013
Silva, N., Vargas, C. A., & Prego, R. (2011). Land–ocean distribution of allochthonous organic matter in surface sediments of the Chiloé and
Aysén interior seas (Chilean Northern Patagonia). Continental Shelf Research, 31(3‐4), 330–339. https://doi.org/10.1016/j.
csr.2010.09.009
Skei, J. (1983). Why sedimentologists are interested in fjords. Sedimentary Geology, 36(2‐4), 75–80. https://doi.org/10.1016/0037‐
0738(83)90002‐7
Smith, R. W., Bianchi, T. S., Allison, M., Savage, C., & Galy, V. (2015). High rates of organic carbon burial in fjord sediments globally.
Nature Geoscience, 8(6), 450–453. https://doi.org/10.1038/ngeo2421
Soto, D., León‐Muñoz, J., Dresdner, J., Luengo, C., Tapia, F. J., & Garreaud, R. (2019). Salmon farming vulnerability to climate change in
southern Chile: Understanding the biophysical, socioeconomic and governance links. Reviews in Aquaculture, 11(2), 354–374. https://
doi.org/10.1111/raq.12336
Takahashi, T., Sutherland, S. C., Wanninkhof, R., Sweeney, C., Feely, R. A., Chipman, D. W., et al. (2009). Climatological mean and decadal
change in surface ocean pCO2, and net sea‐air CO2 flux over the global oceans. Deep‐Sea Research Part II, 56(8–10), 554–577. https://doi.
org/10.1016/j.dsr2.2008.12.009
Tiwari, T., Sponseller, R. A., & Laudon, H. (2018). Extreme climate effects on dissolved organic carbon concentrations during snowmelt.
Journal of Geophysical Research: Biogeosciences, 123, 1277–1288. https://doi.org/10.1002/2017JG004272
Torres, R., Pantoja, S., Harada, N., González, H. E., Daneri, G., Frangopulos, M., et al. (2011). Air‐sea CO2 fluxes along the coast of Chile:
From CO2 outgassing in central northern upwelling waters to CO2 uptake in Southern Patagonian fjords. Journal of Geophysical
Research, 116, C09006. https://doi.org/10.1029/2010JC006344
Torres, R., Silva, N., Reid, B., & Frangopulos, M. (2014). Silicic acid enrichment of subantarctic surface water from continental inputs along
the Patagonian Archipelago Interior Sea (41–56°S). Progress in Oceanography, 129, 50–61. https://doi.org/10.1016/j.pocean.2014.09.008
Turk, D., Book, J. W., & McGillis, W. R. (2013). pCO2 and CO2 exchange during high bora winds in the Northern Adriatic. Journal of Marine
Systems, 117‐118, 65–71. https://doi.org/10.1016/j.jmarsys.2013.02.010
Turley, C., & Gattuso, J.‐P. (2012). Future biological and ecosystem impacts of ocean acidification and their socioeconomic‐policy impli-
cations. Current Opinion in Environmental Sustainability, 4(3), 278–286. https://doi.org/10.1016/j.cosust.2012.05.007
Valle‐Levinson, A., Sarkar, N., Sanay, R., Soto, D., & León, J. (2007). Spatial structure of hydrography and flow in a Chilean fjord, Estuario
Reloncaví. Estuaries and Coasts, 65(3), 513–525. https://doi.org/10.1016/j.ecss.2005.06.016
Vandekerkhove, E., Bertrand, S., Reid, B., Bartels, A., & Charlier, B. (2015). Sources of dissolved silica to the fjords of northern Patagonia
(44‐48°S): The importance of volcanic ash soil distribution and weathering. Earth Surface Processes and Landforms, 41(4), 499–512.
https://doi.org/10.1002/esp.3840

VERGARA‐JARA ET AL. 2864


21698961, 2019, 9, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2019JG005028 by Cochrane Chile, Wiley Online Library on [28/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Journal of Geophysical Research: Biogeosciences 10.1029/2019JG005028

Vargas, C. A., Lagos, N. A., Lardies, M. A., Duarte, C., Manríquez, P. H., Aguilera, V. M., et al. (2017). Species‐specific responses to ocean
acidification should account for local adaptation and adaptive plasticity. Nature Ecology & Evolution, 1(4), 0084. https://doi.org/10.1038/
s41559‐017‐0084
Wang, Y., Li, L., Hu, M., & Lu, W. (2015). Physiological energetics of the thick shell mussel Mytilus coruscus exposed to seawater acidifi-
cation and thermal stress. Science of the Total Environment, 514, 261–272. https://doi.org/10.1016/j.scitotenv.2015.01.092
Wanninkhof, R. (2014). Relationship between wind speed and gas exchange over the Ocean. Limnology & Oceanography: Methods, 97(C5),
7373–7362. https://doi.org/10.1029/92JC00188
Xu, Y.‐Y., Cai, W.‐J., Gao, Y., Wanninkhof, R., Salisbury, J., Chen, B., et al. (2017). Short‐term variability of aragonite saturation state in the
central Mid‐Atlantic Bight. Journal of Geophysical Research: Oceans, 122, 4274–4290. https://doi.org/10.1002/2017JC012901
Yang, B., Byrne, R. H., & Lindemuth, M. (2015). Contributions of organic alkalinity to total alkalinity in coastal waters: A spectrophoto-
metric approach. Marine Chemistry, 176, 199–207. https://doi.org/10.1016/j.marchem.2015.09.008

VERGARA‐JARA ET AL. 2865

You might also like