You are on page 1of 16

Progress in Oceanography 119 (2013) 32–47

Contents lists available at SciVerse ScienceDirect

Progress in Oceanography
journal homepage: www.elsevier.com/locate/pocean

Land–ocean gradient in haline stratification and its effects on plankton


dynamics and trophic carbon fluxes in Chilean Patagonian fjords
(47–50°S)
H.E. González a,b,c,d,⇑, L.R. Castro b,c,e, G. Daneri b,c,d, J.L. Iriarte b,c,d,f, N. Silva g, F. Tapia b,c,e, E. Teca a,
C.A. Vargas h
a
Instituto de Ciencias Marinas y Limnológicas, Universidad Austral de Chile, Valdivia, Chile
b
Centro de Investigación Oceanográfica en el Pacífico Sur-Oriental (COPAS), Universidad de Concepción, Concepción, Chile
c
COPAS Sur Austral (PFB-31-2007), Concepción, Chile
d
Centro de Investigación en Ecosistemas de la Patagonia (CIEP), Coyhaique, Chile
e
Departamento de Oceanografía, Universidad de Concepción, Concepción, Chile
f
Instituto de Acuicultura, Universidad Austral de Chile, Puerto Montt, Chile
g
Pontificia Universidad Católica de Valparaíso, Valparaíso, Chile
h
Unidad de Sistemas Acuáticos, Centro de Ciencias Ambientales (EULA-Chile), Universidad de Concepción, Concepción, Chile

a r t i c l e i n f o a b s t r a c t

Article history: Patagonian fjord systems, and in particular the fjords and channels associated with the Baker/Pascua
Available online 20 June 2013 Rivers, are currently under conspicuous natural and anthropogenic perturbations. These systems dis-
play very high variability, where limnetic and oceanic features overlap generating strong vertical and
horizontal physicochemical gradients. The CIMAR 14-Fiordos cruise was conducted in the Chilean fjords
located between 47° and 50°S during the spring (October–November) of 2008. The main objectives
were to study vertical and horizontal gradients in physical, chemical and biological characteristics of
the water column, and to assess plankton dynamics and trophic carbon fluxes in the fjords and chan-
nels of central-south Patagonia.
The water column was strongly stratified, with a pycnocline at ca. 20 m depth separating a surface
layer of silicic acid-rich freshwater discharged by rivers, from the underlying nitrate- and orthophos-
phate-rich Subantarctic waters. The outflows from the Baker and Pascua Rivers, which range annually
between 500 and 1500 m3 s1, generate the strong land–ocean gradient in salinity (1–32 psu) and inor-
ganic nutrient concentrations (2–8 and 2–24 lM in nitrate and silicic-acid, respectively) we observed
along the Baker Fjord. The POC:chl-a ratio fluctuated from 1087 near the fjord’s head to 175 at its oce-
anic end in the Penas Gulf. This change was mainly due to an increase in diatom dominance and a con-
current decrease in allochthonous POC towards the ocean.
Depth-integrated net primary production (NPP) and bacterial secondary production (BSP) fluctuated
between 49 and 1215 and 36 and 150 mg C m2 d1, respectively, with higher rates in oceanic waters.
At a time series station located close to the Baker River mouth, the average NPP was lower (average
360 mg C m2 d1) than at more oceanic stations (average 1063 mg C m2 d1), and numerically domi-
nated (45%) by the picoplankton (<2 lm) and nanoplankton (2–20 lm) size fractions. The high average
vertical carbon flux (234 mg m2 d1) and high export production (65% of the NPP) support the idea
that Patagonian fjords may behave as a net sink for CO2 during the productive (spring) season. Trophic
fluxes near the head of the fjords, with oligotrophic low-salinity waters, were dominated by heterotro-
phic nanoflagellates (HNF) and small copepods (52 mg C m2 d1, each), suggesting that the microbial
food web is the main trophic pathway in these environments.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction

Chilean Patagonia constitutes a large (>1600 km from 41° to


55°S), extensive (240,000 km2) and unique ecosystem consti-
⇑ Corresponding author at: Instituto de Ciencias Marinas y Limnológicas,
tuted by semienclosed highly haline-stratified areas. The region
Universidad Austral de Chile, Valdivia, Chile. Tel.: +56 (63) 221559; fax: +56 (63)
221455.
is made up mainly of fjords and channels, characterized by
E-mail address: hgonzale@uach.cl (H.E. González). intertangled geomorphologies where water inputs from terrestrial

0079-6611/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.pocean.2013.06.003
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 33

and marine ecosystems overlap and mix. These highly vulnerable and fjords down to 50°S during the austral spring of 2008 (30 Octo-
ecosystems have been intensively used, providing socio-economic ber–24 November) on board the AGOR ‘Vidal Gormáz’ of the Chil-
(aquaculture, tourism, hydroenergy, etc.) and other less conspicu- ean Navy (Fig. 1). Three research approaches were implemented.
ous ecosystem services, such as the CO2 uptake capacity during
the productive season and carbon export to the deep sea. Patago- 2.1. Time series stations
nian systems play an important role in biological productivity,
with highly seasonal production that results in efficient carbon ex- Data on the volume of freshwater discharged by the Baker and
port to the sediments in spring (González et al., 2010). This also re- Pascua Rivers were obtained from the Chilean General Water Direc-
sults in the transport and exchange of important amounts of torate (DGA; www.dga.cl) for 2008–2010. A string of temperature
organic matter between terrestrial and marine systems, being the and light intensity sensors was moored at ca. 20 m depth near the
main contributors to carbon flux in coastal oceans (Walsh, 1991). Baker River mouth. Light intensity was recorded at 6 m depth,
The potential impacts of local (anthropogenic) as well as remote whereas four thermistors were located at 0, 5, 10 and 15 m depths
(global warming) perturbations on these ecosystem services need (Fig. 2). Both variables were measured at 10 m intervals.
to be taken into consideration for future sustainable uses. The average spring–summer distribution of freshwater along
The fjords and channels of Patagonia receive Subantarctic Water the Baker Basin was characterized using satellite-derived data on
(SAAW) with high loads of nitrate and phosphate from the ocean, total suspended sediments gathered by the MERIS-ENVISAT mis-
and freshwater with high loads of silicic acid from land. The surface sion, during the rare clear-sky days encountered between October
freshwater layer that is formed by river discharges, high precipita- 2005 and January 2011, to make a composite of 25 MERIS images
tion, and glacier melting, gradually mixes with the deeper and salty (Fig. 3).
SAAW layer through estuarine circulation. The strong pycnocline
thus formed may in some cases act as a barrier for nutrient exchange
and zooplankton vertical migration (Silva et al., 1998; Sievers and 2.2. Oceanographic transects
Silva, 2008; Sánchez et al., 2011). Events of mixing across this den-
sity barrier may be caused by fluctuations in freshwater input, and Five oceanographic transects were conducted from the Penas
may enhance new primary production (Gargett et al., 2003) and fa- Gulf (47°S) to Trinidad Gulf (50°S). (1) The Penas Gulf–Martinez
vor diatom blooms due to a reduction in macro-nutrient limitations Fjord transect comprised 9 stations (St. 92, 2, 93, 4, 96, 97, 14,
within the thin photic zone (Goebel et al., 2005). Such enhancement 13, 12 and 16) arranged in an E–W orientation, and will be referred
in nutrient conditions seems to compensate for the reduction in light to as the ‘‘Martinez transect’’. (2) The Baker Fjord transect included
penetration that is caused by the often high loads of particulate and 6 stations (St. 5, 6, 7, 8, 9 and 10), arranged in an E–W orientation
dissolved organic and inorganic matter associated to river dis-
charges. It has been shown that changes in near-surface salinity
may explain a large fraction of temporal variability in light extinc-
tion in fjord environments (Gibbs, 2001; Goebel et al., 2005).
Physical and chemical conditions along the Baker Basin
(47°470 S; 73°370 W) are strongly influenced by allochthonous mate-
rial, including organic matter, discharged into the Martinez and Ba-
ker Fjords by the Baker and Pascua Rivers, respectively. These
rivers have outflows that range annually between 500 and
3000 m3 s1, and originate from the large General Carrera
(46°310 S; 71°440 W) and O0 Higgins (48°550 S; 73°070 W) lakes,
respectively. This basin is located in between the two large fresh-
water reserves made up by the Northern (46–47°S) and the South-
ern (48–52°S) Ice Fields, from which glacier-melt freshwater flows
into the Steffen Fjord and Pascua River, respectively. Most proba-
bly, a great fraction of the organic matter carried by these freshwa-
ter inputs originates from the extensive evergreen and deciduous
forests, dominated by species of the family Nothofagaceae (Rodrí-
guez et al., 2008; V. Sandoval, pers. comm.), and subsequent trans-
port to the estuary due to intensive run-off driven by an active
precipitation regime that ranges between 1000 and 7000 mm y1
(National Water Directorate; www.dga.cl). Recent studies focused
on stable-isotope signatures have shed light on the contribution
of terrestrial versus marine carbon inputs in this region (Sepúlveda
et al., 2011; Silva et al., 2011; Vargas et al., 2011).
The aims of this contribution were: (1) to describe the vertical
and horizontal gradients in physical, chemical and biological fea-
tures, and (2) to assess the plankton dynamics, trophic carbon
fluxes and carbon export in the highly haline-stratified Baker Basin,
focusing on the Martinez, Baker and Eyre Fjords, as well as the
Trinidad and Messier Channels.

2. Materials and methods


Fig. 1. Study area showing both transects (lines) and stations (dots) between the
Penas Gulf and Trinidad Gulf, Chile. Black star: location of the process station (St.
A research cruise (CIMAR 14 Fiordos) was conducted in the Ba- 14); black square: location of the temperature and solar radiation time series
ker Basin (47°56.00 S, 74°29.60 W), spanning several inner channels stations at Caleta Tortel, close to the Baker and Pascua Rivers.
34 H.E. González et al. / Progress in Oceanography 119 (2013) 32–47

Fig. 2. Discharge of the Baker and Pascua Rivers (top panel), as well as precipitation in the area (middle, red line), obtained from DGA monitoring stations. Light intensity at
ca. 6 m below the surface (middle, black line) and temperature profiles (bottom panel) were measured at a 20-m depth mooring near the Baker River mouth. Black contours
on the bottom panel correspond to 10 °C.

and in the text will be referred to as the ‘‘Baker transect’’. (3) The cess-oriented study. Measurements conducted at St. 14, hereafter
Picton Channel–Penas Gulf transect included 11 stations (St. 80, referred to as ‘process station’, were aimed at establishing the pos-
83, 84, 85, 86, 87, 88, 89, 90, 91 and 91R) along a N–S axis and will sible fate of photosynthetically generated organic carbon through
be referred to as the ‘‘Trinidad transect’’. (4) The Wide Channel– classical or microbial food webs. Measurements were conducted
Messier Channel transect included 10 stations (St. 40, 35, 31, 24, over 5 days (6–10 November, 2008) where primary production
23, 22, 20, 19, 18 and 17), arranged in a N–S orientation and in (PP) and bacterial secondary production (BSP) were estimated on
the text will be referred to as the ‘‘Messier transect’’. (5) The Wide days 1, 3 and 5, whereas micro- and meso-zooplankton grazing
Channel–Eyre Fjord included 7 stations (St. 40, 35, 31, 25, 26, 27 experiments were conducted on days 2, 3 and 5. The vertical flux
and 28), arranged in a N–S orientation and in the text will be re- of particulates was estimated using sediment traps deployed for
ferred to as the ‘‘Eyre transect’’. Stations 40, 35 and 31 were also two periods of ca. 2 days each (days 1–2 and 3–4). Although no
included in the last transect to distinguish the spatial transition to- data on freshwater discharge are available for the Steffen Fjord,
wards a fjord that ends in an extensive ice-field (Fig. 1). At these most of the freshwater input to the area corresponded to glacier
stations, water samples for plankton, dissolved oxygen and inor- meltwater from the Steffen Glacier located at this fjord’s head. This
ganic nutrients were collected at discrete depths (see details be- process station represents average conditions of the inner/middle
low). The physical structure of the water column (temperature, area of the Martinez Fjord, characterized by a high input of both
salinity) was recorded with a Seabird 19 CTD. The salinity sensor freshwater and suspended particulate matter.
was calibrated by measuring salinity with an Autosal salinometer- We estimated size-fractionated PP, BSP, nano-, micro-, and
in discrete water samples. To estimate water column stability at mesozooplankton grazing, and vertical fluxes of particulate mate-
every station, the Brunt-Väisälä frequency (N2) was computed as rial. Additionally, we measured the abundance and carbon-based
N 2 ¼ ½ðg=qÞ  ðDdh =DzÞ (IOC, SCOR and IAPSO, 2010). biomass of the following planktonic components: phytoplankton,
bacteria, heterotrophic and autotrophic nanoflagellates (HNF,
ANF), dinoflagellates and ciliates (see below). Size-fractionated PP
2.3. Process-oriented time-series station and BSP were also estimated at stations 17, 22 and 40, located in
the northernmost, central and southernmostpart of the Messier
This fixed station (St. 14) in the middle of the Martinez Fjord Channel (Fig. 1 and Table 1).
and near the mouth of the Steffen Fjord was occupied for a pro-
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 35

Fig. 3. Average spring–summer distribution of suspended sediments in the study area (left panel) and an enlarged view of the Martinez and Baker Fjords, showing the Pascua
and Baker River mouths and the Steffen Fjord (right panel). Data correspond to a composite of 25 MERIS images gathered by the ENVISAT satellite during clear-sky days
between October 2005 and January 2011.

The different physical, chemical and biological parameters mea- 50 m) in the upper 50 m of the water column. Bacterioplankton
sured/estimated during the cruise are summarized in Table 1, and abundance (cells mL1) was measured on polycarbonate membrane
a detailed description of all measurements is given below. filters (Nuclepore 0.2 lm) stained with the fluorochrome DAPI (Por-
ter and Feig, 1980) and using epifluorescence microscopy. Bacterial
2.4. Size-fractionated chl a and POC biomass was estimated using a factor of 20 fg C cell1 (Lee and Fuhr-
man, 1987). For the enumeration of nanoflagellates, 20 mL sub-sam-
For chl a and phaeopigments, 200 mL of seawater were filtered ples were filtered on a 0.8 lm polycarbonate membrane filter and
(MFS glass fiber filters, 0.7 lm nominal pore size) in triplicate and stained with Proflavine (0.033% w/v in distilled water) according
immediately frozen (20 °C) until later analysis via fluorometry, to Haas (1982) and fixed with glutaraldehyde. Sub-samples for dia-
using acetone (90% v/v) for the pigment extraction (Turner Design tom counts (300 mL) were preserved in an acid Lugol’s solution (1%
TD-700) according to standard procedures (Parsons et al., 1984). final conc.). Sub-samples of 50 mL were placed in settling chambers
Chl a size fractionation was done as for the PP (see below). In addi- for 30 h before analysis under an inverted microscope (Zeiss Axio-
tion, water samples were collected to determine POC and PON con- vert 200, 400 magnification) using the standard methodology
centrations in the water column, including stable isotopes (13C) at (Utermöhl, 1958); a carbon:plasma volume ratio of 0.11 pg C lm3
selected oceanographic stations. These samples (from 0.5 to 1.0 L) was applied for diatom carbon estimations (Edler, 1979). For dino-
were filtered through precombusted MFS filters and stored frozen flagellate and ciliate counts, 10–30 L of seawater from different
until later analysis following standard procedures (von Bodungen depths (surface, 5, 10, 25 m) were filtered gently through a 10 lm
et al., 1991). The autochthonous POC contribution was estimated mesh sieve, then concentrated up to a final volume of 100 mL
as the sum of the phytoplankton, bacterial, and microzooplankton and preserved with buffered formalin (5% final conc.). Car-
carbon. The latter fraction included flagellates (thecate and athe- bon:plasma volume ratios of 0.3 and 0.19 pg C lm3 were used for
cate dinoflagellates), ciliates (loricate and aloricate ciliates) and heavily thecate and athecate dinoflagellate forms, respectively (Les-
crustacean larvae (nauplii). Carbon estimates were based on car- sard unpubl. data fide Gifford and Caron, 2000), and 0.148 pg C lm3
bon:plasma volume ratios (see below). Allochthonous POC was was applied for ciliates (Ohman and Snyder, 1991).
estimated as total POC minus autochthonous POC. Zooplankton samples were collected by oblique tows using a
For dissolved oxygen and nutrient (nitrate, orthophosphate, silicic Tucker trawl net (1 m2 catching area, 300 lm mesh size) within
acid) analyses, water samples were collected using a CTD-rosette the upper 75 m of the water column. Samples were preserved in
equipped with 24 Niskin bottles from the following depths: 1, 5, 10, borax buffered formalin (10% final conc.) for later analyses of the
25, 50, 75, 100, 150, 200, 250, 300, 350 and 400 m, with maximum dominant zooplankton groups (large copepods and euphausiids).
depth of each profile determined by water depth at each station. Dis- Additional tows using a WP-2 net (200 lm mesh size) were con-
solved oxygen was analyzed on board with the Winkler method (Car- ducted to estimate the abundance of small calanoid copepods.
penter, 1965). Nutrient samples (50 mL) were stored at 20 °C in acid-
cleaned high-density polyethylene bottles and analyzed later in a
nutrient autoanalyzer (Mod. Technicon) according to Atlas et al. 2.6. Primary production (PP) and bacterial secondary production (BSP)
(1971).
Water samples for PP determinations were incubated in 100 mL
2.5. Plankton abundance and biomass borosilicate glass bottles (2 clear replicate bottles and 1 dark bottle
for each depth) and placed in a natural-light incubator for 4 h (usu-
Bacterioplankton, protozoan and phytoplankton counts were ally from 10:00 to 14:00). Temperature was controlled by running
done for water samples collected at selected depths (0, 5, 10, 25, surface seawater over the incubation bottles. For incubations
Table 1

36
Parameter measured/estimated during the CIMAR 14 cruise.

Station number Water column depth (m) Lat. and long. Chl (fractionated chlorophyll-a), POC (particulate organic carbon), BB (bacterial biomass), ANF-B (autotrophic nanoflagellate biomass), HNF-B
(heterotrophic nanoflagellate biomass), NUT (nitrate, phosphate, silicic acid), Microzoo-B (microzooplankton biomass), Cop-B (copepod biomass),
Eup-B (euphausiid biomass), HNF-G (heterotrophic nanoflagellate grazing), Microzoo-G (microzooplankton grazing), Cop-G (copepod grazing), Vert-F
(vertical flux of particulates) and PP (primary production)
Chl POC BB ANFB HNFB NUT Microzoo-B Cop-B Eup-B HNFG Microzoo-G Cop-G Vert-F BSP PP
Penas Gulf–Martinez Fjord transect
p p p p p p p p p p
92 200 47°31,190 S; 75°37,410 W
p p p p p p p p p
2 150 47°17,500 S; 75°00,100 W
p p p p p p p p p
93 156 47°22,200 S; 74°38,700 W
p p p p p p p p p
4 550 47°47,000 S; 74°32,000 W
p p p p p p p p p
96 246 47°48,000 S; 74°15,900 W
p p p p p p p p p
97 124 47°48,400 S; 74°03,900 W
p p p p p p p p p p p p p p p
14a 276 47°48,400 S; 73°45,000 W
p p p p p p p p p
13 188 47°49,900 S; 73°40,950 W
p p p p p p p p p
12 94 47°51,400 S; 73°35,900 W
p p p p p p p p p
16 285 47°59,380 S; 73°22,550 W

H.E. González et al. / Progress in Oceanography 119 (2013) 32–47


Baker Fjord transect
p p p p p p p p p
5 505 47°56,000 S; 74°29,600 W
p p p p p p p p p
6 706 48°00,000 S; 74°19,300 W
p p p p p p p p p
7 404 47°59,300 S; 74°04,300 W
p p p p p p p p p
8 841 48°01,800 S; 73°50,100 W
p p p p p p p p p
9 401 48°04,500 S; 73°38,200 W
p p p p p p p p p
10 278 48°12,750 S; 73°26,000 W
Trinidad Channel transect
p p p p p p p p p
80 210 49°57,120 S; 74°58,000 W
p p p p p p p p p
83 296 49°48,750 S; 75°09,800 W
p p p p p p p p p
84 236 49°37,400 S; 75°17,500 W
p p p p p p p p p
85 98 49°28,900 S; 75°24,900 W
p p p p p p p p p
86 290 49°18,500 S; 75°29,000 W
p p p p p p p p p
87 970 49°06,240 S; 75°15,250 W
p p p p p p p p p
88 630 48°56,500 S; 75°02,000 W
p p p p p p p p p
89 352 48°39,200 S; 74°59,200 W
p p p p p p p p p
90 550 48°23,400 S; 75°02,400 W
p p p p p p p p p
91 535 48°04,150 S; 75°14,000 W
p p p p p p p p p
91R 82 47°52,500 S; 75°35,000 W
Messier Channel transect
p p p p p p p p p p p
40 323 50°09,550 S; 74°43,200 W
p p p p p p p p p
35 939 49°58,100 S; 74°25,800 W
p p p p p p p p p
31 733 49°47,800 S; 74°21,900 W
p p p p p p p p p
24 225 49°24,300 S; 74°24,400 W
p p p p p p p p p
23 151 49°11,200 S; 74°22,500 W
p p p p p p p p p p p
22 168 49°02,800 S; 74°25,800 W
p p p p p p p p p
20 345 48°53,100 S; 74°24,700 W
p p p p p p p p p
19 300 48°39,000 S; 74°22,600 W
p p p p p p p p p
18 1173 48°30,000 S; 74°30,000 W
p p p p p p p p p p p
17 1238 48°02,010 S; 74°38,300 W
Eyre Fjord transect
p p p p p p p p
25 504 49°33,500 S; 74°12,600 W
p p p p p p p p
26 532 49°31,600 S; 74°04,300 W
p p p p p p p p
27 318 49°23,100 S; 74°05,700 W
p p p p p p p p
28 162 49°16,100 S; 74°04,000 W
a
The ‘‘process station’’ (St. 14).
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 37

corresponding to the sub-surface chl a maximum, 40% and 2% of animals each were placed in an incubator rack on deck for approx-
surface light intensity was simulated using screens to attenuate imately 19–25 h. The seawater incubation was mixed by hand
light at depths where the water was collected. Sodium bicarbonate every hour and, to some extent, by the ship’s motion. Initial control
(40 lCi NaH14CO3) was added to each bottle. Following incubation, bottles were immediately preserved with 2% acid Lugol’s solution,
samples were manipulated under subdued light conditions prior to and a sub-sample was also preserved in glutaraldehyde (as above).
and after the incubation periods. The contents were filtered At the end of the incubation, sub-samples were taken from all the
according to the fractionation procedure described below. Filters bottles and preserved in glutaraldehyde for nanoflagellate counts
(0.7 and 2.0 lm) were placed in 20 mL borosilicate scintillation (20 mL) and acid Lugol’s solution (60 mL, as above) for cell concen-
vials and kept at 20 °C until reading. To remove excess inorganic trations. Ingestion rates, measured as cell removal, were calculated
carbon, the filters were treated with HCl fumes for 4 h. A cocktail according to Frost (1972), as modified by Marín et al. (1986).
(10 mL, Ecolite) was added to the vials, and radioactivity was
determined in a scintillation counter (Beckmann). Depth-inte-
2.8. Vertical flux of particulates
grated values of PP (mg C m2 h1) were calculated using trapezoi-
dal integration throughout the euphotic zone; the 4 depths
The vertical flux of POC was estimated by using surface-teth-
considered were the surface, sub-surface chl a maximum, and
ered, cylindrical sediment traps (122 cm2 area and 8.3 aspect ratio)
40% and 2% surface irradiance depths. Integrated production rates
deployed at 50 m water depth at the ‘process station’ (St. 14) for
per hour were multiplied by daily light hours for the Baker Fjord.
periods of 1–1.5 days. Sub-samples were analyzed through micros-
Phytoplankton size-fractionation was performed post-incuba-
copy using standard techniques (Utermöhl, 1958); for POC and
tion in 3 sequential steps: (1) for the nanoplankton fraction (from
PON, 0.5–1.0 L were filtered through precombusted MFS filters
2.0 to 20 lm), seawater (125 mL) was pre-filtered using a 20 lm
and stored frozen until later analysis.
Nitex mesh and collected on a 2.0 lm Nuclepore; (2) for the pico-
plankton fraction (from 0.7 to 2.0 lm), seawater (125 mL) was pre-
filtered using a 2.0 lm Nuclepore and collected on a 0.7 lm MFS 3. Results
glass fiber filter; and (3) for the whole phytoplankton community,
seawater (125 mL) was filtered through a 0.7 lm MFS glass fiber 3.1. Environmental conditions
filter. The microphytoplankton fraction was obtained by subtract-
ing the production estimated in steps 1 and 2 from the production The Baker and Pascua Rivers provide average freshwater dis-
estimated in step 3. charges of 400–1500 m3 s1 each (DGA, www.dga.cl, 2008–2010),
BSP was estimated using water samples from the same oceano- reaching highest levels in summer and lowest levels from late win-
graphic bottle as for PP experiments. The incorporation of L- ter through early spring. Upon the occurrence of GLOF’s (Glacial
[14C(U)]-leucine into proteins (Simon and Azam, 1989) was used Lake Outburst Floods), driven by changes in hydrometeorological
to estimate BSP through the increment of biomass. Incubations conditions and resulting in the recurrent emptying of lakes such
were conducted at the in situ temperature and in darkness for as the Cachet (ca. 100 km upriver from the Baker’s mouth), river
1 h, filtered over 0.22 lm membrane filters and extracted with cold outflow may reach peak levels that range between 2000 and
5% trichloroacetic acid (TCA). Samples were frozen on board and 2500 m3 s1 (Fig. 2, upper panel).
counted in a liquid scintillation counter in the laboratory. From At the Baker estuary mooring station, the light penetration fol-
each experiment, water samples were taken for bacteria cell num- lowed an inverse pattern with respect to freshwater discharge,
bers and biomass estimations (for a complete description see Mon- with highest sub-surface light intensities (100–1000 lx) during late
tero et al., 2011). winter–early spring, and lowest intensities (<10 lx) during sum-
The nanoflagellate and microzooplankton grazing experiments mer–autumn months. Temperature in the water column showed
were performed using the size-fractionation method (Kivi and Set- strong temporal variability over the 2008–2010 period, with con-
ala, 1995; Sato et al., 2007). Water samples were collected from the spicuous seasonal changes and marked differences among years.
fluorescence maximum depth with a clean 10 L Go-Flo bottle-ro- Summer temperatures were warm (10–14 °C) through most of
sette system. After collection, seawater was size-fractioned by re- the water column in the 2008–2009 summer (December–March),
verse filtration into 3 fractions: (1) <2 lm (i.e. mostly bacteria whereas the following year they ranged from 8 to 13 °C within
and cyanobacteria), (2) <10 lm (i.e. mostly bacteria, cyanobacteria, the upper 10 m of the water column. During winter–spring, an in-
ANF, and HNF), and (3) <115 lm (i.e. the whole photoheterotrophic verse thermal stratification was observed, with cold water (5–8 °C)
community). Grazing rates were estimated by comparing prey in the top 8 m of the water column and warmer waters (9–11 °C)
growth rates in the presence and absence of predators selected near the bottom (Fig. 2).
by reverse filtration as follows: for HNF grazing, by comparing The spring–summer distribution of suspended sediments along
(1) and (2); and for microzooplankton (ciliates + dinoflagellates) the Baker Basin (i.e. Martínez and Baker Fjords) showed that max-
grazing on nanoflagellates (both PNF and HNF), by comparing (2) imum concentrations (50–100 g m3) occurred at the Baker and
and (3) (Gifford, 1993). Further descriptions of procedures and Pascua River mouths, while steadily decreasing towards the ocean,
methods are fully described in Vargas et al. (2008). reaching <1 g m3 at the Penas Gulf. The process station 14 was lo-
cated in the Martinez Fjord, close to the Steffen Fjord mouth
2.7. Copepod grazing (Fig. 3), and also showed a very high concentration of suspended
particles (100 g m3).
For copepod grazing estimations, animals were collected by At all transects, temperatures ranged from 8 to 10 °C. At the
slow vertical hauls in the upper 20 m of the water column using Martinez and Baker transects, tongues of relatively cold (8 °C)
a WP-2 net (mesh size 200 lm) with a large non-filtering cod water projected from the mouths of the Baker and Pascua Rivers;
end (40 L). Undamaged copepods were placed in 500 mL (from water was even colder (6.5 °C) at the station close to the Southern
4 to 8 small copepods) or 1000 mL (from 2 to 3 large copepods) Ice Fields (St. 28) on the Eyre transect. Below the surface layer, a
acid-washed polycarbonate bottles. These bottles were filled with vertical quasi-homogeneous thermal distribution (8–9 °C) was
ambient water loaded with natural food assemblages of micro- found (Fig. 4, upper panel).
plankton pre-screened through a 200 lm net to remove most graz- Along the two E–W transects (Martinez and Baker), a strong
ers. Three control bottles without animals and 3 bottles with 2–4 horizontal salinity gradient was evident with values ranging from
38 H.E. González et al. / Progress in Oceanography 119 (2013) 32–47

1 psu at the Baker and Pascua River mouths to 32 psu at the Penas face tongue that extended up to St. 4 along the Martinez Fjord and
Gulf. Instead of a major river, the SW–NE Eyre section has a glacier along the entire Baker Basin. The N–S transects showed low surface
(Pio XI) at its head; melt water from this glacier, however, does not concentrations of nitrate (2–4 lM) and orthophosphate (0.2–
set horizontal gradient in salinity or silicate (Figs. 4 and 5), as in the 0.4 lM), which increased below 300 m depth to 24 and 2 lM,
Baker or Pascua Rivers. The N–S oriented transects showed a more respectively (Fig. 5). On the other hand, the distributional pattern
homogeneous structure, with salinity values between 24 and of silicic acid showed highest concentrations within the freshwater
26 psu at the surface (Fig. 4, middle panel). The vertical structure plume, with values up to 26 lM, which decreased down to 4–8 lM
showed a strong halocline (from 20 to 32.5 psu) along the N–S in sub-surface waters down to 50 m depth, and then increased
transects with differences of up to 30 psu in the upper 25 m of again in deeper waters. The N–S transect showed low values close
the E–W transects at stations close to the river mouth. In addition, to the surface (4–8 lM) and a steady increase below 100 m depth
a quasi-homohaline (33–34 psu) distribution was found from 25 m (Fig. 5). The distribution of inorganic nutrients along the Eyre sec-
depth down to the bottom. A strong halocline was set at the head tion showed features similar to the Martinez and Baker sections,
of the Martinez and Baker Fjords, which in turn generates a thin except for the lack of high surface concentrations near the fjord’s
layer of high stability in the water column (N2 = 0.02– head (Fig. 5).
0.06 (rad s1)2) at depths of ca. 20–30 m. This layer diminished in
intensity toward the fjords mouths, where it almost vanished
3.2. Chl-a and POC concentrations
(N2 < 0.001 (rad s1)2), in parallel with the halocline gradient
(Fig. 6). As already mentioned, in the Eyre section the Pio XI glacier
Depth-integrated (upper 25 m water column) chl a concentra-
does not release enough fresh water from ice melting to generate a
tions were very low, with average values of total chl a ranging from
high stabilty layer at its head. Nevertheless, it shows a surface sta-
10 to 17 mg chl a m2 at the Martinez Fjord and Messier Channel,
bility peak at its mid section (St. 35; 0.02 (rad s1)2) due to the
respectively. Only the Trinidad Channel displayed a slight increase
higher fresh water input from the Europa Fjord (Fig. 6).
in chla a concentration, with an average of 33 mg m2. The domi-
The dissolved oxygen concentration showed a similar
nant chl a size fraction was <5 lm contributing with 55% of total
vertical structure in all transects, with high values close to the sur-
chl a, while the size fraction of 5–20 and >20 lm accounted for
face (7 mL L1) that steadily decreased towards the bottom
18.5% and 26.5%, respectively (Fig. 7A). The size-fractionated chl
(3.5 mL L1). Slightly elevated dissolved oxygen concentrations
a was dominated by small phytoplankton (autotrophic flagellates
(>8 mL L1) were found in association with the river discharges
<5 lm), with some increase in diatom abundance at the Trinidad
into the Martinez and Baker Fjords (Fig. 4, lower panel). The dis-
Channel and Eyre Fjord (3.3  105 cells L1), and thecate dino-
solved oxygen concentrations in the Eyre section showed similar
flagellates at the Baker Fjord and Messier Channel (1.1  105 -
characteristics to the Martinez and Baker sections, although near-
cells L1). Interestingly, one of the key species was the diatom
bottom concentrations were not as low (Fig. 4).
Guinardia delicatula, which contributed with >50% of the large,
The nitrate and orthophosphate concentrations showed sharp
chain-forming diatoms at St. 14, 17 and 40. Only the oceanic SAAW
land–ocean gradients, with very low values in the low-salinity sur-
of the Penas Gulf (St. 92 and 91R), exhibited a more conspicuous

Fig. 4. Vertical distribution of the physical and chemical variables recorded in the study area. The different panels show the distributions of temperature, in °C (upper panel),
salinity, in psu (middle panel) and dissolved oxygen, in mL L1 (bottom panel) throughout the water column in November 2008.
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 39

Fig. 5. Vertical distribution of the macro-nutrients recorded in the study area. The different panels show the distribution of (upper panel) nitrate, in lM, (middle panel)
orthophosphate in lM, and (bottom panel) silicic acid, in lM, throughout the water column in November 2008.

diatom community with Skeletonema spp., Leptocylindrus minimus, (752 ± 52 mg C m2), as compared to the other three transects
Thalassiosira spp., Chaetoceros spp., and Pseudonitzschia spp. (182 ± 79 mg C m2) (Fig. 7D). The ANF biomass was, on average,
Integrated POC concentrations in the upper 25 m of the water 2.6 times higher (289 ± 236 mg C m2) than the HNF
column showed only small changes through the study area, with (112 ± 85 mg C m2).
an average of 6186 ± 484 mg C m2, and a slight increase in the Pe- The overall mean depth-integrated (upper 25 m water column)
nas Gulf (11,370 ± 3265 mg C m2) (Fig. 7B). Surface-layer POC was microzooplankton abundance was 298 mg C m2. The dominant
made up mostly by allochthonous POC, with contributions that group was thecate dinoflagellates (238 ± 178 mg C m2), which
ranged from 82% at St. 17 to 100% at St. 13. At 10 m depth there contributed 70% to total microzooplankton biomass. Crustacean
was a gradient in allochthonous POC, decreased from ca. 89% near nauplii contributed 46 ± 13 mg C m2 or 20% of total biomass,
the Baker and Pascua River mouths (St. 10 and 12) to 49% at the while both loricate and aloricate ciliates contributed with 2% each
northern tip of the Messier Channel (St. 17) (Fig. 8). This feature (Fig. 7E). At St. 22, 23 and 25, large numbers of small (<20 lm) and
was corroborated by the d13C stable isotope signature, which at large (>20 lm) dinoflagellates were present, with carbon biomass
the surface showed a strong terrestrial input near the river mouths 1 g C m2.
(d13C range: 25.0 to 26.5) and a slight increase in marine influ- The abundance of small copepods (<1500 lm prosome length)
ence at St. 17 (23.5). At 10 m depth, on the other hand, a domi- integrated within the upper 20 m was numerically dominated by
nance of marine input was evident (d13C range: 25.0 to 22.5) Paracalanus parvus, copepodites and the genera Acartia and Centro-
(Fig. 8). pages. Among the large copepods (>1500 lm PL), Calanus chilensis
and the genera Neocalanus and Rhincalanus dominated. The
3.3. Plankton biomass (bacterioplankton, nanoflagellates, ciliates, overall mean depth-integrated abundance of copepods was
mesozooplankton) 7928 ± 2327 ind. m2, with maximum values at the Trinidad Chan-
nel (10,939 ind. m2) and at St. 88 (40,919 ind. m2). Euphausiid
Depth-integrated bacterioplankton biomass (upper 25 m) abundance integrated within the upper 25 m was 91 ± 75 ind. m2
showed high variability at the station level (within transects) and throughout the study area and was largely accounted for by Eup-
changed relatively little at the transect level (i.e. mean ± SD: hausia vallentini (>90%), (Fig. 7F). It must be noted that these pat-
217 ± 98; 212 ± 67 and 211 ± 109 mg C m2 in the Martinez Fjord, terns were derived from daytime and nighttime plankton tows,
Trinidad Channel and Eyre Fjord) and a slightly higher biomass which may strongly underestimate abundance due to active verti-
in the Baker Fjord and Messier Channel (303–372 mg C m2) cal migration.
(Fig. 7C). No clear E–W or N–S spatial pattern could be identified,
and bacterial carbon only represented an average of 5% (range 3– 3.4. Productivity, pelagic food web and carbon export flux
6%) of total POC.
The overall integrated biomass of NF was 410 ± 318 mg C m2, At the stations where PP and BSP were estimated, chl a was low
with a 4-fold increase along the Martinez and Trinidad transects within the upper 25 m (5–20 mg m2), and dominated by the
40 H.E. González et al. / Progress in Oceanography 119 (2013) 32–47

Fig. 6. Vertical distribution of the Brunt-Väisälä stability index (N2) and salinity (psu) at three stations along the Martinez (St. 16, 97, 2), Baker (St. 10, 7, 5) and Eyre (St. 40, 35,
27) transects, located at the head, center and mouth of the fjords, respectively.

picoplankton size fraction (<2 lm). The >20 lm fraction was dom- the estimated fluxes were similar, although the C:N ratio was high-
inated by the diatom Guinardia delicatula, with minor contributions er at 50 m (14.9) than at 90 m depth (11.2). The highest grazing
by Skeletonema spp., Leptocylindrus minimus, Pseudo-nitzschia spp., rates were those estimated for HNF and small copepods
Thalassionema nitzschioides and Navicula spp. Only at St. 22 were (52 mg C m2 d1) (Fig. 10). Our results are summarized in
dinoflagellates important (Fig. 9, upper panel). Fig. 10, and presented in the context of carbon fluxes through the
On average, the POC:Chl a ratio was very high at the process pelagic food web. They are further discussed in the following
station (St. 14), with a value (596 ± 422) similar to that found section.
at St. 22 (550). Stations 17 and 40, which receive a lesser load
of suspended sediments, exhibited ratios of 286 and 295, respec- 4. Discussion
tively. The PP:chl a ratio did not display a clear trend, with rela-
tively low values that ranged between 33 and 110 for all stations The rugged coastline of the study area is characterized by (1) a
(Fig. 9, middle panel). hydrographical scenario impinged on by freshwater inflow from
Depth-integrated PP was low at Stn. 14, with an average the Baker and Pascua Rivers, and (2) a geomorphological scenario
of 422 ± 255 mg C m2 d1, while it was twice as high consisting of 5 different channels with different orientations and
(1063 mg C m2 d1) at stations located along the Messier Channel connections between them and with the open ocean.
(St. 17, 22 and 40). On average, picophytoplankton contributed the The oceanic, salty SAAW enter the study area mainly from the
bulk (44.5%) of PP at St. 14, while nanophytoplankton was most north (Penas Gulf) and through the Martinez and Baker Fjords,
important (44.3%) at stations along the Messier Channel. mixing with surface freshwater from rivers (Sievers and Silva,
There were no apparent spatial patterns in estimated BSP rates 2008). Clearly, depending on their geographic orientation (E–W
and bacterioplankton biomass, with an average depth-integrated versus S–N) the Patagonian channels and fjords will be affected
production of 113 ± 495 mg C m2 d1. The PP:BSP ratio was al- in different ways by haline stratification and, therefore, will exhibit
ways >1, with both extreme values (1.4 and 11.6) observed at St. different degrees of water column stablity. Along the E–W Fjords,
14, and a mean rate of 6.2 ± 3.5 for this and the three stations along the horizontal salinity gradients and vertical extent of the upper
the Messier Channel (Fig. 9, lower panel). brackish layer were very conspicuous. The former took the form
Estimates for the vertical flux of particles were highly variable, of a freshwater plume extending from the river mouth to the coast-
with 149 ± 45 mg C m2 d1 (n = 2) during the 6–7 November trap al ocean, whereas in the vertical dimension there were strong
deployment, and 320 ± 10 mg C m2 d1 (n = 2) during 8–9 pycnoclines and layers of high stability between 5 and 30 m depth
November (average of 234 mg C m2 d1). At 50 and 90 m depth, (1–30 psu; 0.02–0.06 (rad s1)2). This highly stable surface layer
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 41

Fig. 7. Integrated values of different parameters measured during November 2008 along the transects from the Penas to the Trinidad Gulfs. (A) Size-fractionated chl a, (B)
particulate organic carbon, (C) bacterial biomass, (D) heterotrophic and autotrophic nanoflagellates (HNF and ANF), (E) microzooplankton and (F) mesozooplankton.
Abundances of the different plankton groups were transformed into carbon biomass by using the following equations: Bacteria = 20 fg C cell1 (Lee and Fuhrman, 1987),
flagellates = 6500 fg C cell1 (Børsheim and Bratbak, 1987), tintinnids = pg C = (lm3)  0.053 + 444.5 (Verity and Langdon, 1984), thecate dinoflagellates = pg C = (lm3)  0.13
(Edler, 1979), copepod nauplii = ng C = 1.51  10–5LC2.94 (Uye et al., 1996), calanoid copepods = 26.9 lg C cop1 (González et al., 2000; Hirst et al., 2003), euphausi-
ids = 3812 lg C euph1 (Sánchez et al., 2011). All values were integrated throughout the upper 25 m of the water column.

became thinner towards the fjord’s mouth, where it almost disap- northern and central Patagonia. From observations made in fjords
peared. This pattern is consistent with the one documented by of central Patagonia during the 1970 HUDSON cruise, Guerrero
Bustos et al. (2008) and Landaeta et al. (2001) in the fjords of (2000) found maximum stability values of 0.02–0.06 (rad s1)2
42 H.E. González et al. / Progress in Oceanography 119 (2013) 32–47

(422 ± 255 mg C m2 d1) than at stations located along the Mess-
ier Channel (1063 ± 166 mg C m2 d1). Two main factors seem to
constrain PP at St. 14: macro-nutrient concentrations and light
availability. Nitrate and orthophosphate concentrations at St. 14
were low enough to be considered limiting (2 and 0 lM, respec-
tively), whereas light intensity in the upper water column was at
its annual minimum, despite the seasonal increase in solar radia-
tion (Aracena et al., 2011). This is because freshwater discharged
by the Baker and Pascua Rivers enters the fjords with a very high
load of particulates (50–100 g m3 of suspended sediments as
can be seen in Fig. 3), coming mostly from the melting of glaciers
in the Northern and Southern Ice Fields. This combination of nutri-
ent and light limitation should impact the physiological condition
of phytoplankton at St. 14, lowering the PP:chl a ratio (i.e. P/B ra-
tio) which has been suggested as a proxy for healthy physiological
conditions, and dependent on daily maximum irradiance and ni-
trate concentration (González et al., 2009). Indeed, the mean inte-
grated P/B ratio in the upper 25 m at St. 14 (49.3 ± 34.2) was lower
than P/B values found along the Messier Channel (70.3 ± 35.5). In
general, P/B ratios were within the range of springtime values re-
ported for the channels and fjords of the Aysén Region, from 60
in the northern Moraleda Channel (44°S) to 20 in the Aysén Fjord
(45°S) (González et al., 2011). In the upwelling system off Concep-
ción, the P/B ratio ranges from <20 in winter to ca. 100 in spring
(González et al., 2009).
Contrasting with the spatial trends in P/B ratios, the POC:chl a
ratio was higher at St. 14 (595.7 ± 422.0) than at stations along
Fig. 8. Percentage of allochthonous (open bars) and autochthonous (black bars) the Messier Channel (377.2 ± 149.9). Low values of this ratio have
particulate organic carbon and d13C stable isotope at stations located along the
been related to high productivity (108 during late summer in the
Martinez and Baker Fjords at the surface (A) and 10 m depth (B).
Skagerrak, Maar et al., 2004; 29–207 in the coastal region off Con-
cepción, Grunewald et al., 2002), whereas high values are charac-
for the Baker, Steffen, and Europa Fjords, and lower than teristic of unproductive areas or high-river-discharge winter
0.02 (rad s1)2 at Eyre. These values are highly consistent with periods, when the input of allochthonous organic matter (probably
our stability results for the CIMAR 14 Fiordos cruises, conducted loaded with refractory carbon) increases. This seasonal pattern in
38 years after the HUDSON. the POC:chl a ratio has been found in other fjord areas such as
Overall, a layer of freshwater with high concentrations of silicic the Reloncaví Fjord and the Interior Sea of Chiloe, in northern Pat-
acid spanned the top 15 m of the water column at the Martinez and agonia (41°S), which exhibit low average values (87) during spring,
Baker Fjords. Below the pycnocline, a much thicker layer of marine and high values (487) during winter (González et al., 2010).
water loaded with nitrate and orthophosphate lead to the overlap- Global warming trends may affect the stratification regime and
ping of limnetic and marine characteristics in the fjords (Silva and productivity patterns in Patagonia over a variety of spatial and
Neshyba, 1979; Silva et al., 1998), as previously reported for Chil- temporal scales. For instance, the frequency of glacial-lake out-
ean Patagonia (Silva et al., 1998; Palma and Silva, 2004; Silva and burst floods (GLOF’s), such as those documented by Dussaillant
Palma, 2006; González et al., 2010, 2011). The ratio of nitrate et al. (2009) for Lake Cachet 2, may be increasing. While five events
and orthophosphate available for phytoplankton in the study area were reported in 2008–2009 (Dussaillant et al., 2009), six of these
(ca. 10) was lower than the global mean ratio of nitrogen to phos- events that doubled the river outflow (to 2000–2800 m3 s1) in a
phorous (N/P  15) (Deutsch and Weber, 2012), which suggests matter of hours were recorded in 2009–2010 (Fig. 2). The large vol-
that productivity in this area may be nitrate-limited. The N/P ratio ume of freshwater released on each event (ca. 200 million m3, Dus-
may determine overall cellular stoichiometry (Geider and La saillant et al., 2009) may certainly alter stratification and
Roche, 2002), since the organic molecules used for resource acqui- conditions for primary production near the Baker River mouth. An-
sition (proteins and pigments) require substantially more N per other potential impact of these GLOF’s is the increased reservoir
atom of P than does the molecular machinery of growth and repro- sedimentation due to a large sediment load, with unexpected im-
duction, i.e. DNA and RNA (Arrigo, 2005). Thus, high stratification pacts on benthic fauna and biodiversity (Quiroga et al., 2012), as
set up in Patagonian fjords by inputs of oligotrophic freshwater well as on the functioning of the planned hydropower dams in
would strongly impact the N/P ratio and, in turn, the physiological the Pascua and Baker Rivers (Dussaillant et al., 2009; Vince,
adaptation of phytoplankton (Finkel et al., 2006). Potential conse- 2010). Thus, natural and human activities may potentialy impact
quences would be the dominance of different functional groups, Patagonian rivers, increasing sediment load on some, and decreas-
such as small-size phytoplankton (picophytoplankton), more ing the flux in others through damming. Such catchment impacts,
adapted to nutrient acquisition under limitation of micro- and/or together with rapid urbanization on delta plains, accentuate the
macro-nutrients (Iriarte and González, 2004), diatom species with threats posed by natural hazards and climate change (Woodroffe
low N:P ratios (<11:1) (Bertilsson et al., 2003; Weber and Deutsch, and Saito, 2011).
2010), or diazotrophs which are not reliant on the nitrate supplied Phytoplankton was low and dominated by the small pico- and
by SAAW. nano-phytoplankton fractions. A dominance of diatoms was evi-
The presence of a brackish upper layer significantly affects the dent at the oceanic end of the transects, where the continuous sup-
productivity and structure of planktonic communities (reviewed ply of freshwater loaded with silicic acid and the enhanced light
by Wetzel, 1995; Pace et al., 2004). Primary production at availability during spring months seem to favor the development
the process station (St. 14) was substantially lower of large, filamentous diatom blooms (Thalassiosira, Chaetoceros
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 43

Fig. 9. From top to bottom: Distribution of fractionated chl a (<2, 2–20 and >20 lm) and the relative contribution of the dominant species of diatoms and thecate
dinoflagellates. Values of the POC:Chl a, PP:chl a and PP:BSP ratios, and estimated PP and BSP rates at the process station (St. 14) in the Martinez Fjord and the three stations
along the Messier Channel (St. 17, 22, 40).

and Skeletonema). This is a recurrent situation in fjords in both of large-sized cells towards benthic environments (Legendre and
southern (González et al., 2011) and northern (Thompson et al., Le Fevre, 1989) is alternated with equally important periods of
2008) hemispheres. nanoflagellate dominance.
The average bacterioplankton biomass at the Baker Fjord The relatively low integrated variability shown by bacterial
(259 ± 195 mg C m2; n = 41, Fig. 7C) was similar to that found biomass (Fig. 7) and average secondary production
near the Aysén Fjord at 45.3°S (307 mg C m2, González et al., (113 ± 49 mg C m2; n = 7) suggests that bacterial mortality and
2011), and nearly half that of that found at the Reloncaví Fjord at growth rates are, at least, within the same order of magnitude. An-
41.4°S (507 mg C m2, González et al., 2010). In contrast to the re- other factor explaining the relatively low variability in bacterial
duced HNF grazing pressure during spring in the Reloncaví Fjord biomass might be high inputs of DOC, due to a subsidy of dissolved
(3.7 mg C m2), high HNF grazing rates in the Baker Fjord organic matter (DOM) from Patagonian rivers (Perakis and Hedin,
(51.9 mg C m2) could exert a partial control on BSP and growth 2002; Pantoja et al., 2011). Freshwater inputs could mediate an
during this season. The total nanoflagellate biomass was higher important flux of DOM/POM and macro- and oligo-elements into
at stations close to the Penas Gulf, with more oceanic conditions coastal areas (Raymond and Bauer, 2000). These are compartments
than stations well within the fjords and channels. For example, still absent from, or scarcely represented in, global biogeochemical
the average integrated flagellate biomass decreased from the Trin- models. High DOC inputs (1.4 g C m3) were observed at St. 14
idad Channel (789 ± mg C m2) to the Messier Channel (unpublished data), indicating that DOC input from the Baker and
(162 ± mg C m2) and Eyre Fjord (115 ± mg C m2). Flagellates Pascua Rivers into the Martinez and Baker Fjords could fluctuate
were dominated by rounded individuals 2–10 lm in size (domi- between 121 and 259 t C d1. In the upper 25 m of St. 14, the aver-
nance of 5 lm). In marine upwelling environments off Central age integrated DOC (33 g C m2) was one order of magnitude high-
Chile (37°S), this increase in nanophytoplankton has been linked er than POC (3–4 g C m2). DOC/POC ratios vary widely, and our
to lower coastal salinities (33 psu) during periods of high river estimations agree well with values observed for temperate forest
outflow and higher haline stratification (Morales and Anabalón, watersheds (10:1), while for rivers draining, predominantly grass-
2012). Thus, both in upwelling systems and in the more oceanic land areas, ratios may be closer to 1:1 (Schlesinger and Melack,
sector of estuarine systems, nanoplankton has been reported as 1981). In estuarine systems, microbes as well as mesozooplankton
the dominant size fraction, contributing to maintain the system’s might rely on this OM for 5–80% of their production, which in-
production and providing food for higher trophic levels such as creases exponentially with increasing river discharge (Maranger
copepods and fish larvae (Anabalón et al., 2007). This suggests that et al., 2005; Hoffman et al., 2008). Although it is widely accepted
the classical view of upwelling and estuarine systems as exporters that allochthonous OM inputs into estuarine systems are quantita-
44 H.E. González et al. / Progress in Oceanography 119 (2013) 32–47

Insights into the dominance of allochthonous versus autochtho-


nous processes in the study area can be obtained through the C/N
ratio of the particulate organic matter fraction. We measured an
average C/N ratio of 6.7 ± 1.6 (n = 160) within the upper 50 m of
the study area, which is well within the generally accepted range
for marine material of 6–8 (Richards et al., 1997; Andrews et al.,
1998) and suggests a dominance of marine processes. On the con-
trary, the stations located close to the Pascua and Baker River
mouths (St. 8, 9, 10 and 12, 13, 15, 16) showed ratios of 9.1 ± 3.0
and 12.0 ± 2.0, respectively, closer to values reported for terrestrial
material (Thornton and McManus, 1994). The highest ratios were
measured in material collected from sediment traps deployed at
50 and 90 m depth, with average values of 13.3 ± 2.3 at St. 14, sug-
gesting a rapid utilization of the labile marine fraction, leaving the
refractory terrestrial material to sink towards the sediments. Bac-
teria are known to preferentially utilize labile organic carbon from
marine diatoms when a mixed carbon pool is available (Dai et al.,
2008). The highly significant correlation between gross primary
production and microbial respiration in the Reloncaví Fjord (Mon-
tero et al., 2011) suggests a strong coupling between the synthesis
of organic matter by Patagonian phytoplankton and its usage by
the microbial heterotrophic community.
Discharging fluids, whether derived from land or composed of
re-circulated seawater, will react with water column and sediment
components. Thus, terrestrially-derived fluids represent a pathway
for new material fluxes and reactions that may substantially in-
crease the concentrations of nutrients, carbon and metals, and will
be a source of biogeochemically important constituents for estua-
rine areas and the coastal ocean (Burnett et al., 2003). In general,
the groundwater flux (estimated with naturally-enriched radium
isotopes) to these estuarine/coastal waters should be ca. 40% of
Fig. 10. Conceptual model of carbon flux at the process station in the Martinez the river water flux (Moore, 1996; Charette et al., 2001). A strong
Fjord, showing the average values of primary production (PP), bacterial secondary terrestrial signature was especially conspicuous at the surface
production (BSP), vertical carbon flux, and grazing rates through heterotrophic
layer of the Martinez and Baker Fjords (>85% of allochthonous car-
nanoflagellates (HNF), microzooplankton (Microzoo), and mesozooplankton (Mes-
ozoo) split into euphausiids, large and small copepods during November 2008. All bon and d13C range of 23.5 to 26.4). The river discharge forms a
values are given in mg C m2 d1 (modified from González et al. (2010)). plume constituted by organic and inorganic sediments that on
average produces an annual inflow of 2.4  106 t y1 (assuming a
tively more important, autochthonous OM is likely to provide more sediment load and water discharge of 50 g m3 and 1500 m3 s1,
readily available substrates for bacteria (Maranger et al., 2005). respectively). Overall, riverine/estuarine DOM/POM is considered
In lakes, grazing and microbial-detrital trophic pathways sup- to be recalcitrant and transported conservatively to the adjacent
port higher consumers. For instance, it has been estimated that coastal areas, such as the Penas Gulf (Mantoura and Woodward,
22–50% of total zooplankton ingestion, in terms of daily body car- 1983). However, the sharp terrestrial-marine gradients in biologi-
bon, is associated with allochthonous terrigenous sources (Pace cal and physico-chemical conditions results in changes in the up-
et al., 2004). The same situation was found in the Baker River estu- take of DOM by bacteria (Zweifel, 1999) and chemical removal
ary, with copepod dependence on allochthonous OM sources processes (Lisitzin, 1995). Furthermore, sedimentary POC and
declining with increasing fjord trophy or microplankton availabil- DOC that is loosely bound to mineral surfaces can be released dur-
ity (Vargas et al., 2011). The same study found that dominant cope- ing resuspension events (Komada and Reimers, 2001). All of these
pods in the Baker Estuary (Acartia tonsa, Paracalanus parvus and processes can contribute to a nonconservative behavior in
Calanus patagoniensis) may switch from low ingestion (1– estuaries.
2 lg C ind.1 d1) of microplankton (diatoms and flagellates of
marine origins) and slightly depleted d13C values (24‰), to high 4.1. Fate of autotrophic production and pelagic–benthic coupling in
microplankton ingestion (4 and 12 lg C ind.1 d1) and enriched the area
d13C values (21‰). Various recent reports suggest that a signifi-
cant fraction of the OM in Chilean fjords is of allochthonous origin, Seasonal and shorter-term changes in estuarine stratification,
and can be channeled either directly to higher trophic levels (zoo- influenced by the annual cycle in ice melting plus precipitation
plankton and fish) through the ingestion of detrital particles, or and events such as GLOF’s, can have profound influences on the
indirectly through microbial food webs (i.e. allochthonous DOC – marine ecosystem. The strong halocline gives rise to a thin layer
bacteria – HNF – Zooplankton) (González et al., 2010, 2011; Vargas of high stability in the water column, defining a natural barrier
et al., 2011). Contrastingly, it has been shown that in shallow estu- with an upper brackish layer dominated by small copepods and
aries with a substantially lower freshwater discharge (e.g. Ythan cladocerans, and a lower layer dominated by marine copepods
Estuary, UK, with mean discharge of 5–9 m3 s1) material of mar- and euphausiids. The grazing impact of mesozooplankton was
ine origin dominates sedimentary POC throughout the year (Zet- mainly due to the small calanoid copepods, which removed
sche et al., 2011). Mass-balance calculations indicate that 14.3% of daily PP at the process station. Large calanoid copepods
terrestrial organic carbon is limited to sediments within a rela- and the euphausiid Euphausia vallentini exerted a minor role
tively short distance from the river mouth, where this distance is (3.8% and 0.2%) (Table 2; Fig. 10). The grazing impact of euphausi-
a function of river outflow (Shultz and Calder, 1976). ids, and probably also large copepods, should be treated with
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 45

caution because their abundance may have been understimated values for the northern Reloncaví Fjord (200–600 mg C m3) but
during daytime trawls, due to the active daily vertical migration. higher than values recorded in the southern Strait of Magellan
In addition, the copepod impact refers only to the few species of (76 mg C m3; Fabiano et al., 1999). This latitudinal decrease ap-
copepods that were studied (Paracalanus parvus, Acartia spp. and pears to parallel the percentage of organic carbon in the sediments
Calanus spp.). Small cyclopod copepods such as Oithona spp. were from the same areas (from 5.3% to 3.8%, respectively, Silva and
not included in this study, despite their numerical dominance in Prego, 2002). These gradients could be due to more efficient pela-
numerous estuarine systems (Nielsen and Andersen, 2002; Maar gic–benthic coupling in northern than in southern Patagonian
et al., 2004), because our zooplankton mesh was too coarse (300 fjords, which may result from higher spring–summer PP in north-
and 200 lm) to adequately collect such small copepods. In the ern than in southern fjords. For example, PP estimates for the
northern Patagonian fjords and channels (e.g. Reloncaví Fjord, Reloncaví (41°), Aysén (46°) and Baker Fjords (48°) were 1893,
Moraleda Channel), chain-forming diatoms dominate the copepod 831 and 360 mg C m2 d1. Another factor is the higher diatom
diet during the productive season (spring–summer months), while abundance in northern than in southern fjords that results in high
nanoflagellates (ANF and HNF) and dinoflagellates, along with tin- vertical POC fluxes. Consistently, the POC vertical flux at our pro-
tinnids and aloricate choeotrich ciliates (i.e. Strombidium sp.), are cess station on the Martinez Fjord was ca. threefold lower than
the main food items for copepods in the much less productive in the Reloncaví Fjord (234 versus 725 mg C m2 d1). In general,
winter months (González et al., 2010). This change highlights the its seems that the POC export depended more on the variable input
trophic plasticity of copepods, which are able to adjust their met- of both particulate and dissolved matter from terrestrial systems
abolic demands to the in situ availability of prey (Vargas et al., and was less coupled with local PP than in coastal areas. This
2008). was evident in the nearby Chiloé and Aysén channels and fjords,
Beyond stratification, several other physical–chemical features where surface sediments contain up to 50% terrestrial material
of Patagonian fjords will be potentially altered in a climate-change (Sepúlveda et al., 2005; Silva et al., 2011). In the Martinez Fjord
scenario. Changes in water temperature, circulation, nutrient in- (Stn. 14), the total seston recorded in the sediment traps was
puts, oxygen content, and acidity may have far-reaching biological 50.2 g m2 d1, from which only 4.2% was organic matter. Further
effects, including the fluxes of carbon through the pelagic food south at 51.5°S, the Almirante Montt Gulf does not receive impor-
web. It has been shown that a 2 °C warming of the upper water col- tant river discharges and indeed is not affected by freshwater in-
umn may produce a 37% increase in the growth rate of some phy- puts. In this system, the total seston flux collected in sediment
toplankton species (Eppley, 1972), whereas increased freshwater traps was one order of magnitude lower than in the Martinez Fjord
discharge and stratification in the Baker Fjord produced adverse (5.47 g m2 d1), and the relative organic matter contribution in-
conditions for ichthyoplankton (Maurolicus parvipinnis and Sprattus creased to 22% (González, unpublished data). In general, one of
fuegensis), resulting in a lower prey consumption per individual the most conspicuous particles collected in sediment traps were
and a reduction of growth rates in low salinity water (Landaeta faecal pellets from planktonic crustaceans (copepods and eup-
et al., 2001). hausiids). Faecal pellets from copepods and krill were loaded with
On average, primary production (PP) was four times higher than fine sediments and minerals of possible lacustrine and riverine ori-
bacterial secondary production (BSP), with BSP:PP ratios of gins. This material may act as ballast to exacerbate settling velocity
0.25 ± 0.22. The ratio was lower (0.14) at the three Messier Channel and POC transport to the sediments (Lewis and Syvitski, 1983).
stations (St. 17, 22 and 40), indicating that PP was seven times Overall, carbon fluxes and fjord system functioning are highly
greater than BSP. Northern Patagonian fjords such as the Reloncavi variable, both spatially (northern to southern fjords and channels)
(41°S) showed low BSP:PP values during spring (0.2) and high val- and seasonally. In winter, community respiration (CR) exceeds PP
ues in winter (3.7) (González et al., 2010), highlighting the high PP in many lakes (Pace et al., 2004) and northern Patagonian fjords,
in northern Patagonia during the productive season. Overall, sea- as the Reloncaví Fjord located at 41°300 S (Montero et al., 2011),
sonal changes in this ratio are related more to sharp spring-to-win- which contrasts with >1 ratios in spring (González et al., 2010).
ter decreases in PP than to changes in bacterial biomass or The metabolism of Patagonian fjords is dominated by prokariotic
production. and protistan heterotrophs, and the strong input of allochthonous
The average POC concentrations in the upper 25 m of the water organic matter from rivers and underground watershed obscure
column (247 mg C m3) were in the lower range of the reported the local productive processes over a gradient from river mouths

Table 2
Average individual grazing rates (mg C ind.1 d1), average integrated (upper 25 m water column) abundance (no. ind. m2), and grazing rates exerted by the integrated
communities (mg C m2 d1) of heterotrophic nanoflagellates (HNF), microzooplankton (MICR), represented by thecate dinoflagellates, small copepods, and euphausiids of the
species Euphausia vallentini at the ‘‘process station’’ (St. 14).

HNF MICR (thecate dinoflagellates) Small copepods Large copepods Euphausiids


Grazing rate (mg C ind.1 d1)
9.8  109a 1.9  109b 0.00294c 0.00945d 0.0787e
2
Integrate abundance (no. ind. m )
5.3  109 9.1  106f 17,520g 1458h 8.4i
Integrated grazing rate (mg C m2 d1)
51.9 0.018 51.5 13.8 0.66
a
Range: 20–33 bact. HNF1 h1 (or 9–16  109 mg C ind.1 d1).
b
2.3  106–1.7  106 lg C ind.1 d1.
c
0.7–7.3 lg C ind.1 d1.
d
1.9–18.8 lg C ind.1 d1.
e
Average of E. vallentini juveniles and adults 78.7 lg C ind.1 d1 (Sánchez et al., 2011).
f
Thecate dinoflagellates >20 lm (mainly Protoperidinium spp.) and copepod nauplii.
g
Paracalanus spp. and Acartia tonsa.
h
Calanus chilensis and Rhyncalanus spp.
i
Average of E. vallentini juveniles and adults.
46 H.E. González et al. / Progress in Oceanography 119 (2013) 32–47

to the coastal ocean. Thus, basin and seasonal scales seem to be Finkel, Z.V., Quigg, A., Raven, J.A., Reinfelder, J.R., Schofield, O.E., Falkowski, P.G.,
2006. Irradiance and the elemental stoichiometry of marine phytoplankton.
adequate for comparative quantitative analyses of material and en-
Limnology and Oceanography 51, 2690–2701.
ergy fluxes at all trophic and organizational levels in Patagonian Frost, B.W., 1972. Effects of size and concentration of food particles on the feeding
fjords. The suggested annual framework (see Wetzel, 1995) would behaviour of the marine planktonic copepod Calanus pacificus. Limnology and
allow the analysis of seasonal variability (i.e. freshwater discharges Oceanography 17, 805–815.
Gargett, A.E., Stucchi, D., Whitney, F., 2003. Physical processes associated with high
and solar radiation among others) in the context of processes and primary production in Saanich Inlet, British Columbia. Estuarine, Coastal and
mechanisms from Patagonian fjords and channels such as CO2 Shelf Science 56, 1141–1156.
sequestration during the productive season, and the relevance of Geider, R.J., La Roche, J., 2002. Redfield revisited: variability of C:N:P in marine
microalgae and its biochemical basis. European Journal of Phycology 37, 1–17.
the microbial loop in allochthonous versus autochthonous organic Gibbs, M.T., 2001. Aspects of the structure and variability of the low-salinity-layer
matter mineralization. in Doubtful Sound, a New Zealand fjord. New Zealand Journal of Marine and
Freshwater Research 35, 59–72.
Gifford, D.J., 1993. Consumption of protozoa by copepod feeding on natural
Acknowledgements microplankton assemblages. In: Kemp, P.F., Sherr, B.F., Sherr, E.B., Cole, J.J. (Eds.),
Handbook of Methods in Aquatic Microbial Ecology. Lewis Publishers, London,
pp. 723–737.
We thank the entire crew and all the scientists of the CIMAR 14- Gifford, D.J., Caron, D.A., 2000. Sampling, preservation, enumeration and biomass of
2008 oceanographic cruise. We thank María José Calderón, Cesar marine protozooplankton. In: Harris, R.P., Wiebe, P.H., Lenz, J., Skjoldal, H.R.,
Huntley, M. (Eds.), ICES Zooplankton Methodology Manual. Academic Press, NY,
Barrales, Paola Reinoso, Francisco Gallardo, Eduardo Menschel
pp. 193–221.
and María Inés Muñoz for collaboration during the sampling and Goebel, N.L., Wing, S.R., Boyd, P.W., 2005. A mechanism for onset of diatom blooms
data analysis and Leticia Cisternas for euphausiid identification in a fjord with persistent saline stratification. Estuarine, Coastal and Shelf
and counting. We also thank Eduardo Menschel, Nicolás Sánchez Science 64, 546–560.
González, H.E., Sobarzo, M., Figueroa, D., Nöthig, E.M., 2000. Composition, biomass
and Ricardo Giesecke for stimulating discussions on plankton and potential grazing impact of the crustacean and pelagic tunicates in the
dynamics in Patagonian systems. This study was funded by the northern Humboldt Current area off Chile: differences between El Niño and
Projects CIMAR14-Fiordos (CONA-C14F 08-05 and CONA-CF14 non-El Niño years. Marine Ecology Progress Series 195, 201–220.
González, H.E., Daneri, G., Iriarte, J.L., Yannicelli, B., Menschel, E., Barría, C., Pantoja,
08-13). Additional support is also acknowledged from COPAS Grant S., Lizárraga, L., 2009. Carbon fluxes within the epipelagic zone of the Humboldt
FONDAP-15010007 and COPAS Sur-Austral Grant PFB-31/2007. Current System off Chile: the significance of euphausiids and diatoms as key
functional groups for the biological pump. Progress in Oceanography 83, 217–
227.
References González, H.E., Castro, L., Daneri, G., Iriarte, J.L., Silva, N., Vargas, C., Giesecke, R.,
Sánchez, N., 2011. Seasonal plankton variability in Chilean Patagonia fjords:
carbon flow through the pelagic food web of the Aysen Fjord and plankton
Anabalón, V., Morales, C.E., Escribano, R., Varas, M.A., 2007. The contribution of
dynamics in the Moraleda Channel basin. Continental Shelf Research 31, 225–
nano- and micro-plankton assemblages in the surface layer (0–30 m) under
243.
different hydrographic conditions in the upwelling area off Concepción, central
González, H.E., Calderón, M.J., Castro, L., Clement, A., Cuevas, L., Daneri, G., Iriarte,
Chile. Progress in Oceanography 75, 396–414.
J.L., Lizárraga, L., Martínez, R., Menschel, E., Silva, N., Carrasco, C., Valenzuela, C.,
Andrews, J.E., Greenway, A.M., Dennis, P.F., 1998. Combined carbon isotope and C/N
Vargas, C.A., Molinet, C., 2010. Primary production and its fate in the pelagic
ratios as indicators of source and fate of organic matter in a poorly flushed,
food web of the Reloncaví Fjord and plankton dynamics of the Interior Sea of
tropical estuary: Hunts Bay, Kingston Harbour, Jamaica. Estuarine Coastal and
Chiloé, Northern Patagonia, Chile. Marine Ecology Progress Series 402, 13–30.
Shelf Sciences 46, 743–756.
Grunewald, A.C., Morales, C.E., González, H.E., Silvester, C., Castro, L., 2002. Grazing
Aracena, C., Lange, C.B., Iriarte, J.L., Rebolledo, L., Pantoja, S., 2011. Latitudinal
impact of copepod assemblages and gravitational flux in coastal and oceanic
patterns of export production recorded in surface sediments of the Chilean
waters off Central Chile during two contrasting seasons. Journal of Plankton
Patagonian fjords (41–55°S) as a response to water column productivity.
Research 24, 55–67.
Continental Shelf Research 31, 340–355.
Guerrero, Y., 2000. Distribución de temperatura, salinidad y oxígeno disuelto en las
Arrigo, K.R., 2005. Marine microorganisms and the global nutrient cycle. Nature
aguas interiores de la zona de los canales australes, entre el golfo de Penas y el
437, 349–355.
seno Almirantazgo. Tesis para obtener el Título de Oceanógrafo, Escuela de
Atlas, E.S., Hager, S., Gordon, L., Park, P., 1971. A Practical Manual for Use of the
Ciencias del Mar, UCV, 96 pp.
Technicon Autoanalyser in Sea Water Nutrient Analyses. Technical Report.
Haas, L.W., 1982. Improved epifluorescence microscopy for observing planktonic
Department of Oceanography, Oregon State University, pp. 1–215.
microorganisms. Annales de ÍInstitut Oceanographique, Paris 58, 261–266.
Bertilsson, S., Berglund, O., Karl, D.M., Chisholm, S.W., 2003. Elemental composition
Hirst, A.G., Roff, J.C., Lampitt, R.S., 2003. A synthesis of growth rates in marine
of marine Prochlorococcus and Synechococcus: implications for the ecological
epipelagic invertebrate zooplankton. Advances in Marine Biology 44, 1–142.
stoichiometry of the sea. Limnology and Oceanography 48, 1721–1731.
Hoffman, J.C., Bronk, D.A., Olney, J.E., 2008. Organic matter sources supporting lower
Børsheim, K.Y., Bratbak, G., 1987. Cell volume to cell carbon conversion factor for a
food web production in the tidal freshwater portion of the York River Estuary,
bacterivorous Monas sp. enriched from seawater. Marine Ecology Progress
Virginia. Estuaries and Coasts 31, 898–911.
Series 36, 171–175.
IOC, SCOR, IAPSO, http://www.TEOS-10.org, 2010. The International
Burnett, W.C., Bokuniewicz, H., Huettel, M., Moore, W.S., Taniguchi, M., 2003.
Thermodynamic Equation of Seawater – 2010: Calculation and Use of
Groundwater and pore water inputs to the coastal zone. Biogeochemistry 66, 3–
Thermodynamic Properties. Intergovernmental Oceanographic Commission,
33.
Manuals and Guides No. 56, UNESCO (English), 196 pp.
Bustos, C.A., Landaeta, M.F., Balbontín, F., 2008. Spawning and early nursery areas of
Iriarte, J.L., González, H.E., 2004. Phytoplankton size structure during and after the
anchoveta Engraulis ringens Jenyns, 1842 in fjords of southern Chile. Revista de
1997/98 El Niño in a coastal upwelling area of the northern Humboldt Current
Biología Marina y Oceanografía 43 (2), 381–389.
System. Marine Ecology Progress Series 269, 83–90.
Charette, M.A., Buesseler, K.O., Andrews, J.E., 2001. Utility of radium isotopes for
Kivi, K., Setala, O., 1995. Simultaneous measurement of food particle selection and
evaluating the input and transport of groundwater-derived nitrogen to a Cape
clearance rates of planktonic oligotrich ciliates (Ciliophora: Oligotrichina).
Cod estuary. Limnology and Oceanography 46, 465–470.
Marine Ecology Progress Series 119, 125–137.
Carpenter, J., 1965. The Chesapeake Bay Institute Technique for the Winkler
Komada, T., Reimers, C.E., 2001. Resuspension-induced partitioning of organic
dissolved oxygen method. Limnology and Oceanography 10, 141–143.
carbon between solid and solution phases from a river–ocean transition. Marine
Dai, J., Sun, M., Culp, R.A., Noakes, J.E., 2008. A laboratory study on biochemical
Chemistry 76, 155–174.
degradation and microbial utilization of organic matter comprising a marine
Landaeta, M.F., López, G., Suárez-Donoso, N., Bustos, C.A., Balbontín, F., 2001. Larval
diatom, land grass, and salt marsh plant in estuarine ecosystems. Aquatic
fish distribution, growth and feeding in Patagonian fjords: potential effects of
Ecology. http://dx.doi.org/10.1007/s10452-008-9211-x.
freshwater discharge. Environmental Biology of Fishes. http://dx.doi.org/
Deutsch, C., Weber, T., 2012. Nutrient ratios as a tracer and driver of ocean
10.1007/s10641-011-9891-2.
biogeochemistry. Annual Review of Marine Science 4, 113–141.
Lee, S., Fuhrman, J., 1987. Relationship between biovolume and biomass of
Dussaillant, A., Benito, G., Buytaert, W., Carling, P., Meier, C., Espinoza, F., 2009.
naturally-derived marine bacterioplankton. Applied and Environmental
Repeated glacial-lake outburst floods in Patagonia: an increasing hazard?
Microbiology 53, 1298–1303.
Natural Hazards 54, 469–481. http://dx.doi.org/10.1007/s11069-009-9479-8.
Legendre, L., Le Fevre, J., 1989. Hydrodynamical singularities as controls of recycled
Edler, L., 1979. Recommendations for marine biological studies in the Baltic Sea.
versus export production in oceans. In: Berger, W.H., Smetacek, V.S., Wefer, G.
Baltic Marine Biological Publication 5, 11–38.
(Eds.), Productivity of the Oceans: Present and Past. John Wiley and Sons, pp.
Eppley, R., 1972. Temperature and phytoplankton growth in the sea. Fisheries
49–63.
Bulletin 70, 1063–1085.
Lewis, A.C., Syvitski, J.P.M., 1983. The interaction of plankton and suspended
Fabiano, M., Povero, P., Danovaro, R., Misic, C., 1999. Particulate organic matter
sediment in fjords. Sedimentary Geology 36, 81–92.
composition in a semi-enclosed Periantarctic system: the Straits of Magellan.
Lisitzin, A.P., 1995. The marginal filter of the ocean. Oceanology 34, 671–682.
Scientia Marina 63, 89–98.
H.E. González et al. / Progress in Oceanography 119 (2013) 32–47 47

Maar, M., Nielsen, T.G., Gooding, S., Tönnesson, K., Tiselius, P., Zervoudaki, S., of a southern Chilean fjord (44°S). Estuarine, Coastal and Shelf Science 65, 587–
Christou, E., Sell, A., Richardson, K., 2004. Trophodynamic function of copepods, 600.
appendicularians and protozooplankton in the late summer zooplankton Sepúlveda, J., Pantoja, S., Hughen, K.A., 2011. Sources and distribution of organic
community in the Skagerrak. Marine Biology 144, 917–933. matter in northern Patagonia fjords, Chile (44–47°S): a multi-tracer approach
Mantoura, R.F.C., Woodward, E., 1983. Conservative behavior of riverine dissolved for carbon cycling assessment. Continental Shelf Research 31, 315–329.
organic matter in the Severn estuary. Geochimica et Cosmochimica Acta 47, Shultz, D.J., Calder, J.A., 1976. Organic carbon 13C/12C variations in estuarine
1293–1309. sediments. Geochimica et Cosmochimica Acta 40, 381–385.
Maranger, R.J., Pace, M.L., del Giorgio, P.A., Caraco, N.F., Cole, J.J., 2005. Longitudinal Sievers, A.H., Silva, N., http://www.cona.cl/, 2008. Water masses and circulation in
spatial patterns of bacterial production and respiration in a large river-estuary: austral Chilean channels and fjords. In: Silva, N., Palma, S. (Eds.), Progress in the
implications for ecosystem carbon consumption. Ecosystems 8, 318–330. Oceanographic Knowledge of Chilean Inner Waters, from Puerto Montt to Cape
Marín, V., Huntley, M.E., Frost, B., 1986. Measuring feeding rates of pelagic Horn. Comité Oceanográfico Nacional – Pontificia Universidad Católica de
herbivores: analysis of experimental design and methods. Marine Biology 93, Valparaíso, Valparaíso, Chile, pp. 53–58.
49–58. Silva, N., Neshyba, S., 1979. On the southernmost extension of the Perú-Chile
Montero, P., Daneri, G., González, H.E., Iriarte, J.L., Tapia, F.J., Lizárraga, L., Sánchez, undercurrent. Deep-Sea Research 26, 1387–1393.
N., Pizarro, O., 2011. Seasonal variability of primary production in a fjord Silva, N., Palma, S., 2006. Avances en el conocimiento oceanográfico de las aguas
ecosystem of the Chilean Patagonia: implications for the transfer of carbon interiores chilenas. Puerto Montt a Cabo de Hornos. Comité Oceanográfico
within pelagic food webs. Continental Shelf Research 31, 202–215. Nacional–Pontificia Universidad Católica de Valparaíso, Valparaíso, pp. 162.
Moore, W.S., 1996. Large groundwater inputs to coastal waters revealed by 226Ra Silva, N., Prego, R., 2002. Carbon and nitrogen spatial segregation and stoichiometry
enrichments. Nature 380, 612–614. in the surface sediments of southern Chilean inlets (41–56°S). Estuarine, Coastal
Morales, C.-M., Anabalón, V., 2012. Phytoplankton biomass and microbial and Shelf Science 55, 763–775.
abundances during the spring upwelling season in the coastal area off Silva, N., Calvete, C., Sievers, H.A., 1998. Masas de agua y circulación general para
Concepión, central-southern Chile: variability around a time series station. algunos canales australes entre Puerto Montt y Laguna San Rafael, Chile
Progress in Oceanography 92–95, 81–91. (Crucero Cimar-Fiordo 1). Ciencia y Tecnología del Mar 21, 17–48.
Nielsen, T.G., Andersen, C.M., 2002. Plankton community structure and production Silva, N., Vargas, C.A., Prego, R., 2011. Land–ocean distribution of allochthonous
along a freshwater-influenced Norwegian fjord system. Marine Biology 141, organic matter in surface sediments of the Chiloé and Aysén interior seas
707–724. (Chilean Northern Patagonia). Continental Shelf Research 31, 330–339.
Ohman, M.D., Snyder, R.A., 1991. Growth kinetics of the omnivorous oligotrich Simon, M., Azam, F., 1989. Protein content and protein synthesis rates of planktonic
ciliate Strombidium sp.. Limnology and Oceanography 36, 922–935. marine bacteria. Marine Ecology Progress Series 51, 201–213.
Palma, S., Silva, N., 2004. Distribution of siphonophores, chaetognaths, euphausiids Thompson, R.J., Deibel, D., Redden, A.M., McKenzie, C.H., 2008. Vertical flux and fate
and oceanographic conditions in the fjords and channels of southern Chile. of particulate matter in a Newfoundland fjord at sub-zero water temperatures
Deep Sea Research Part II 51, 513–535. during spring. Marine Ecology Progress Series 357, 33–49.
Pace, M.L., Cole, J.J., Carpenter, S.R., Kitchell, J.F., Hodgson, J.R., Van de Bogert, M.C., Thornton, S.F., McManus, J., 1994. Application of organic carbon and nitrogen stable
Bade, D.L., Kritzberg, E.S., Bastvlken, D., 2004. Whole-lake carbon-13 isotope and C/N ratios as sources indicators of organic matter provenance in
additions reveal terrestrial support of aquatic food webs. Nature 427, estuarine systems: evidence from the Tay estuary, Scotland. Estuarine, Coastal
240–243. and Shelf Science 38, 219–233.
Pantoja, S., Iriarte, J.L., Gutiérrez, M., Calvete, C., 2011. Subpolar margin: Southern Utermöhl, H., 1958. Zur Vervollkommnung der quantitativen Phytoplankton–
Chile. In: Liu, K.K., Atkinson, L., Quiñones, R., Talaue-McManus, L. (Eds.), Carbon Methodik. Mitteilungen Internationale Vereinigung Theorie Angewandte
and Nutrient Fluxes in Continental Margins: A Global Synthesis. Global Change. Limnologie 9, 1–39.
The IGBP Series. Springer Verlag, pp. 265–272. Uye, S.I., Nagano, N., Hidenori, T.H., 1996. Geographical and seasonal variations in
Parsons, T.R., Maita, R., Lalli, C.M., 1984. Counting, Media and Preservation. A abundance, biomass and estimated production rates of microzooplankton in the
Manual of Chemical and Biological Methods for Seawater Analysis. Pergamon Inland Sea of Japan. Japanese Journal of Oceanography 52, 689–703.
Press, Toronto. Vargas, C.A., Martínez, R., González, H.E., Silva, N., 2008. Contrasting trophic
Perakis, S.S., Hedin, L.O., 2002. Nitrogen loss from unpolluted South American interactions of microbial and copepod communities in a fjord ecosystem
forests via dissolved organic compounds. Nature 415, 416–419. (Chilean Patagonia). Aquatic Microbial Ecology 53, 227–242.
Porter, K.G., Feig, Y.S., 1980. The use of DAPI for identifying and counting aquatic Vargas, C.A., Martínez, R.A., San Martin, V., Aguayo, M., Silva, N., Torres, R., 2011.
microflora. Limnology and Oceanography 25, 943–948. Allochthonous subsidies of organic matter across a lake-river-fjord ladscape in
Quiroga, E., Ortiz, P., Gerdes, D., Reid, B., Villagrán, S., 2012. Organic enrichment and the Chilean Patagonia: Implications for marine zooplankton in inner fjord areas.
structure of macrobenthic communities in the glacial Baker Fjord, Northern Continental Shelf Research 31, 187–201.
Patagonia, Chile. Journal of the Marine Biological Association of the United Verity, P., Langdon, C., 1984. Relationship between lorica volume, carbon, nitrogen,
Kingdom 92, 73–83. and ATP content of tintinnids in Narragansett Bay. Journal of Plankton Research
Raymond, P.A., Bauer, J.E., 2000. Bacterial consumption of DOC during transport 6, 859–868.
through a temperate estuary. Aquatic Microbial Ecology 22, 1–12. Vince, G., 2010. Dams for patagonia. Science 329, 382–385.
Richards, P., Riera, P., Galois, R., 1997. Temporal variations in the chemical and von Bodungen, B., Wunsch, M., Fürderer, H., 1991. Sampling and analysis of
carbon isotope composition of marine and terrestrial organic inputs in the Bay suspended and sinking particles in the northern North Atlantic. In: Marine
of Marrenes-Oléron (France). Journal of Coastal Research 13, 879–889. particles: analysis and characterization. AGU Geophysical Monography 63, 47–
Rodríguez, R., Marticorena, A., Teneb, E., 2008. Vascular plants of Baker and Pascua 56.
Rivers, Aysén Region, Chile. Gayana Botánica 65, 39–70 (in Spanish). Walsh, J.J., 1991. Importance of continental margins in the marine biogeochemical
Sánchez, N., González, H.E., Iriarte, J.L., 2011. Trophic interactions of pelagic cycling of carbon and nitrogen. Nature 350, 53–55.
crustaceans in Comau Fjord, Chilean Patagonia: their role in food web structure. Weber, T.S., Deutsch, T.S., 2010. Ocean nutrient ratios governed by plankton
Journal of Plankton Research 33, 1212–1229. biogeography. Nature 467, 550–554.
Sato, M., Yoshikawa, T., Takeda, S., Furuya, K., 2007. Application of the size- Wetzel, R.G., 1995. Death, detritus, and energy flow in aquatic ecosystems.
fractionation method to simultaneous estimation of clearance rates by Freshwater Biology 33, 83–89.
heterotrophic flagellates and ciliates of pico- and nanophytoplankton. Journal Woodroffe, C.D., Saito, Y., 2011. River-dominated coasts. Treatise on Estuarine and
of Experimental Marine Biology and Ecology 349, 334–343. Coastal Science 3, 117–135.
Schlesinger, W.H., Melack, J.M., 1981. Transport of organic carbon in the world’s Zetsche, E., Thornton, B., Midwood, A.J., Witte, U., 2011. Utilization of different
rivers. Tellus 33, 172–187. carbon sources in a shallow estuary identified through stable isotope
Sepúlveda, J., Pantoja, S., Hughen, K., Lange, C., González, F., Muñoz, P., Rebolledo, L., techniques. Continental Shelf Research 31, 832–840.
Castro, R., Contreras, S., Avila, A., Rossel, P., Lorca, G., Salamanca, M., Silva, N., Zweifel, U.L., 1999. Factors controlling accumulation of labile dissolved organic
2005. Fluctuations in export productivity over the last century from sediments carbon in the Gulf of Riga. Estuaries 48, 357–370.

You might also like