You are on page 1of 11

Contents lists available at ScienceDirect

Continental Shelf Research


journal homepage: www.elsevier.com/locate/csr

Western South Atlantic seasonal variability recorded in a mid-deglacial


bivalve from the outer Uruguayan continental shelf
André Klicpera a, b, h, Alvar Carranza c, *, Cristiano M. Chiessi d, Paula Franco-Fraguas e,
Till J.J. Hanebuth f, Hildegard Westphal a, b, g
a
Leibniz Centre for Tropical Marine Research (ZMT), Bremen, Germany
b
Department of Geosciences, MARUM, University of Bremen, Bremen, Germany
c
Departamento de Ecología y Gestión Ambiental, Centro Universitario Regional Este (CURE), Universidad de la República, Maldonado, Uruguay
d
School of Arts, Sciences and Humanities, University of São Paulo (USP), São Paulo, Brazil
e
Ciencia y Tecnología Marina (CINCYTEMA), Universidad de la República, Montevideo, Uruguay
f
Department of Marine Science, Coastal Carolina University (CCU), Conway, SC, USA
g
Physical Science and Engineering Division, King Abdullah University of Science and Technology (KAUST), Thuwal, Kingdom of Saudi Arabia
h
Área Biodiversidad y Conservación, Museo Nacional de Historia Natural, Montevideo, Uruguay

A R T I C L E I N F O A B S T R A C T

Keywords: The oceanographic dynamics on the continental shelf off southeastern South America are primarily controlled by
Retrotapes exalbidus the southward-flowing warm Brazil Current, converging with the northward-directed cold Malvinas (Falkland)
Brazil-Malvinas Confluence Current, and interacting with the continental discharge of the Plata River. The seasonally reversing regional wind
Subtropical shelf front
field together with the seasonal cycle of riverine discharge, determines which of these three components provides
Río de la plata estuary
Deglacial
the dominant forcing. The Uruguayan shelf is thus located in a transitional zone that extends from the region
influenced by the Brazil-Malvinas Confluence (BMC) in the open ocean to the Subtropical Shelf Front (STSF) on
the continental shelf. Understanding how the resulting oceanic seasonal variability responded to different cli­
matic boundary conditions may shed light on its future behavior. This study presents the first reconstruction of
mid-deglacial seasonal hydrographic variability on the continental shelf off southeastern South America in
seasonal resolution based on stable oxygen and carbon isotopes (δ18O, δ13C) from a thick-walled shell of a long-
living bivalve. The mid-deglacial (14.3 cal ka BP; Bølling-Allerød interstadial) Retrotapes exalbidus bivalve shows
a mean δ18O of 3.27 ± 0.42‰ (2.50 ± 0.42‰ when corrected for changes in global ice volume) and a seasonal
δ18O amplitude of 1.69‰ for raw isotopic excursions. Moreover, the δ13C exhibits abrupt negative peaks coin­
cident with more negative δ18O values that indicate seasons of elevated freshwater discharge. Finally, the growth
rate of the bivalve suggests that the specimen was closer to the metabolically optimum than modern individual of
this species from southern South America. Combining biogeographic and ecologic information with these iso­
topic data, the results point to colder waters and a slightly lower mid-deglacial seasonal amplitude in temper­
ature compared with modern conditions at this shelf site. Because of the northward-displaced Plata River mouth
during deglacial times, negative δ13C peaks are expected to reflect an influence of non-point freshwater sources
in the form of small fluvial distributaries along the paleo-coast. Most of this signal may, however, be driven by
seasonal metabolic effects associated with low ambient water temperatures related to a shallow-water envi­
ronment located closer to the respective paleo-coastline due to the low sea level at those times.

1. Introduction zone, the Brazil-Malvinas (Falkland) Confluence (BMC), formed by the


convergence of the Malvinas (Falkland) Current (MC) and the Brazil
The southeastern South American (SESA) continental shelf offshore Current (BC) (Olson et al., 1988; Lumpkin and Garzoli, 2011); (b) the
northeastern Argentina and Uruguay (Fig. 1) is of particular interest for vast Plata River estuary (RdlP), which supplies freshwater from the
oceanographic studies because it hosts: (a) a major oceanic confluence second largest drainage system in South America (Piola et al., 2005;

* Corresponding author..
E-mail address: alvardoc@fcien.edu.uy (A. Carranza).

https://doi.org/10.1016/j.csr.2023.105014
Received 1 February 2023; Received in revised form 25 April 2023; Accepted 7 May 2023
A. Klicpera et al.

Campos et al., 2008); and (c) a dynamic atmospheric system that is 2. Study area and settings
influenced by the South American monsoon and seasonally reversing
wind fields with the dominance of southwestern winds in the austral The Uruguayan continental shelf is more than 150 km wide and hosts
winter, and of southeastern winds in the austral summer (Matano et al., a mosaic of markedly diverse submarine topographic elements and types
1993; Lumpkin and Garzoli, 2011). of sediment, where current-controlled coarse-grained shelf deposits may
The seasonal latitudinal and offshore displacement of shelf waters is occur in close vicinity to relief-confined mud depocenters (Urien et al.,
traceable through physico-chemical seawater properties along the con­ 1980; Violante and Parker, 2004; Brazeiro et al., 2003). The inner shelf
tinental shelf and the associated ecosystems (Piola et al., 2018, and is dominated by a bathymetric depression with late Pleistocene sedi­
references therein). Recent studies suggested that the BMC, including its ments (Fig. 1b), and a zone of partly carbonate-rich sand ridges adjacent
shallow-water extension in the form of a vertical hydrographic boundary to this depression (Lantzsch et al., 2014). Numerous local autochtho­
known as the Subtropical Shelf Front (STSF), has shifted southwards nous and par-autochthonous shell beds are found on the mid and outer
over the past century and particularly over the past few decades as a shelf that consist of fine-to coarse-grained calcareous detritus (Violante
response to current climatic change (Lumpkin and Garzoli, 2011; Bender and Parker, 2004). Towards the shelf break, large areas of the shelf are
et al., 2013; Combes and Matano, 2014; de Souza et al., 2019). Under­ polished by strong confluence-related bottom currents, and deposition is
standing how this seasonal variability changed under different boundary restricted to confined locations (Violante and Parker, 2004; Matano
conditions may shed light on its future behavior. Therefore, et al., 2010).
seasonally-resolved studies comparing the modern situation with past From an oceanographic perspective, the SESA margin is currently
conditions are instructive. Previous paleoceanographic studies from the characterized by the confluence of the warm and saline BC (>20 ◦ C,
western South Atlantic reconstructed changes in the BMC and the STSF >36‰) flowing from the North, and the colder and less saline MC
with millennial-scale resolution (Bender et al., 2013; Lantzsch et al., (<15 ◦ C, <34‰) from the South (Fig. 1). These two currents converge
2014; Voigt et al., 2015; Boretto et al., 2020; Gu et al., 2019), but the between 36 and 40◦ S, where they form the BMC (Fig. 1; Piola et al.,
seasonal variability remained elusive, mainly due to the lack of appro­ 2018, and references therein).
priate archives. The STSF is a marked vertical hydrographic boundary located at
Sclerochronological archives from large bivalve species known for 50–100 m water depth offshore Uruguay (32◦ S to 36◦ S), which separates
their long individual life spans exceeding several decades provide the colder Subantarctic Shelf Water (SASW, <11 ◦ C, <34.5‰), a branch
extremely valuable paleoceanographic information based on high- of the MC, from the warm and more saline Subtropical Shelf Water
resolution analyses (e.g., Dettman and Lohmann, 1995; Jones and (STSW, >16 ◦ C, >34.8‰) carried by the BC (Brandini et al., 2000; Piola
Quitmyer, 1996; Ivany and Runnegar, 2010; Lomovasky et al., 2002b; et al., 2000, 2008). Both the STSF and the BMC migrate latitudinally by
Schöne et al., 2005). Here, we micro-sampled the multi-annual shell about 3◦ due to the reversing wind regime, reaching northern locations
increment succession of a fossil specimen of Retrotapes exalbidus, a in the austral winter, and southern ones in the austral summer (Schmid
long-living marine bivalve. We performed δ13C and δ18O analyzes in and Garzoli, 2009). The shelf surface water (<20 m water depth) un­
sub-seasonal to daily resolution, in order to gain insights into the dy­ dergoes large seasonal fluctuations in sea surface temperature (SST)
namics of: (1) the intra and inter-annual variability in oceanographic with amplitudes of about 10 ◦ C at 33◦ S and 3 ◦ C at 23◦ S (Piola et al.,
conditions; (2) the latitudinal displacement of the STSF driven by the 2000). Sea-surface salinity (SSS) shows only minimal seasonal vari­
competition between the BC and the MC; and (3) the intensity of the ability (33.6–33.7‰) south of 37◦ S. Between 37◦ S and 35◦ S, in contrast,
Plata River freshwater signal related to the changing rainfall intensity a significant surface water freshening (<25–30.0‰) is related to the
over SESA. The new data provide key information for understanding the large Plata River plume (Emilsson, 1961; Piola et al., 2000). This low
long-term regional climate and ocean circulation over the continental salinity plume is transported northwards close to the coast by the Bra­
shelf. Finally, the isotopic information stored in the shell was compared zilian Coastal Current (BCC) (de Souza and Robinson, 2004).
with the observed patterns in modern oceanographic stations along the The Plata River is fed by the Paraná and Uruguay Rivers covering
study region. 20% (3.2*106 km2) of the South American continent (Campos et al.,
2008). The mean discharge of the Paraná River (over the years

Fig. 1. Modern schematic oceanographic setting of


the study area on the southeastern South American
continental margin and location of the sediment core
taken on the outer shelf (A). The glacial paleo-valley
of the Plata River (white dashed arrow) (Lantzsch
et al., 2014) during the shelf exposure and the hy­
pothesized northward displacement of the
Brazil-Malvinas Confluence are illustrated in (B).
RdlP– Rio de la Plata plume BC – Brazil Current; BMC
– Brazil-Malvinas Confluence; SASW – Subantarctic
Shelf Water; STSF – Subtropical Shelf Front; STSW –
Subtropical Shelf Water.
A. Klicpera et al.

1884–1975) is 17,000 m3s-1 with a maximum of 22,000 and a minimum seafloor on the inner Uruguayan shelf and forms a major, coast-parallel
of 8000 m3s-1. The mean discharge of the Uruguay River (1916–1975) is buoyant tongue transported northwards by the BCC (Fig. 2; Wells and
4700 m3s-1 with a maximum of 14,300 m3s-1 and a minimum of 800 Daborn, 1997). Conductivity and temperature through depth (CTD) data
m3s-1. The discharge of these rivers peaks at different times during the (Antonov et al., 2010; Locarnini et al., 2010) indicate that the PPW
year, resulting in only minor seasonal variability in the Plata River occupies the upper 10 m of the water column in austral summer
discharge to the ocean overall (5% of the annual mean; Burrage et al., (January–March), and in the upper 20 m in austral winter (July–Sep­
2008). The present-day annual mean Plata River discharge is about 25, tember), flowing along the inner Brazilian shelf as far north as 26◦ S
000 m3s-1 (Wells and Daborn, 1997; Piola et al., 2018, and references (Fig. 2; Piola et al., 2000).
therein) and leads to the formation of the Plata Plume Water (PPW; This oceanographic setting controls the regional biogeographic pat­
salinity <31‰; Depetris and Griffin, 1968). The PPW detaches from the terns observed in benthic mollusks. Currently, the Uruguayan shelf hosts

Fig. 2. Modern oceanographic conditions off southeastern South America for austral summer (A, B, E, F) and austral winter (C, D, G, H). Salinity (A to D) and
temperature (E to H) are shown for both seasons. The latitudinal shelf sections shown in panels A, C, E and G (left column) follow a southeast-northeast transect
marked as dashed line in panel B. SASW – Subantarctic Shelf Water; PPW –Plata Plume Water; STSW – Subtropical Shelf Wate; STSF – Subtropical Shelf Front. Data:
World Ocean Atlas 2010 (Antonov et al., 2010; Locarnini et al., 2010).
A. Klicpera et al.

a malaco-faunal transition zone between the magellanic (to the South) 3.3. Carbonate mineralogy of inner and outer shell layers
and subtropical (to the North) benthic biogeographic provinces (Olivier
and Scarabino, 1972; Kaiser, 1977; Floeter and Soares-Gomes, 1999; The mineralogy of shell carbonate powder samples was determined
Scarabino, 2003). In addition, a separate benthic mollusk assemblage at the Central Laboratory for Crystallography and Applied Material
occurs along the inner shelf under the influence of the Plata River, where Sciences (ZEKAM, University of Bremen) by means of X-ray diffraction
its diversity strongly correlates with the salinity gradient. The outer shelf (XRD) (Fig. 3a). A Philips X’Pert Pro multipurpose diffractometer
hosts both magellanic and subtropical elements, and benthic species equipped with a Cu-tube (k" 1.541, 45 kV, 40 mA), a secondary Ni-Filter
turnover is associated with latitude (Carranza et al., 2008). The steep and an X’Celerator detector system were used for continuous scans from
meridional temperature gradient that characterizes both BMC and STSF 3 to 85◦ 2θ with a calculated step size of 0.016◦ 2θ. Minerals were
may act as a biogeographic barrier preventing an exchange between identified with the Philips software X’Pert HighScore and the X-ray
Antarctic cold-water and subtropical warm-water species, as evidenced diffraction interpretation software MacDiff version 4.25.
for at least eight cold-water bivalves exhibiting their northernmost
distribution boundary in the area; associated with offshore sub-Antarctic 3.4. Radiocarbon dating
waters (Scarabino et al., 2015). Thus the occurrence of a given fossil or
subfossil mollusk species can be highly informative of the biogeographic The shell was sampled for accelerator mass spectrometry radio­
setting in which they lived, informing about the paleo-distribution of carbon (AMS-14C) dating in two positions, namely the juvenile umbonal
water masses in the study area. shell (dorsal) and near the shell margin (ventral) (Fig. 3b, d). Powder
samples were micro-drilled along marginal transects of 10 mm in total
3. Material and methods length, set parallel to the growth bands. The resulting samples of 30–50
mg comprised less than two growth increments. AMS-14C dating was
3.1. Bivalve shell and taxonomic identification carried out at the Poznan Radiocarbon Laboratory, Poland. The age of
the R. exalbidus specimen was calibrated with the CALIB v. 8.2 software
During research cruise M78 Leg 3a performed with the German R/V and the Marine20 dataset, using the automatically included conserva­
METEOR in 2009 (Krastel et al., 2012), a large, articulated, intact tive global reservoir age of 550 years (Stuiver et al., 2020; Heaton et al.,
bivalve shell was recovered from the base of Vibracore GeoB13802-2 2020).
(345 cm recovery; bivalve sampling depth at 335–341 cm), taken near
the shelf break at 141 m water depth (Fig. 1b; Table 1). The bivalve 3.5. Stable oxygen and carbon isotope analyses
specimen was identified to species level as Retrotapes exalbidus (Dillwyn,
1817) aka Eurhomalea exalbida (Dillwyn, 1817) (the “Ring Clam”) A MicroMill system (New Wave Research) was used for high-
through comparison with reference material housed at the Museo resolution sampling, which is designed to recover 40–50 μg of powder
Nacional de Historia Natural (MNHN), Montevideo, Uruguay (deter­ for isotopic analysis (Dettman and Lohmann, 1995). Only
mined by F. Scarabino). R. exalbidus is a cold-water species that lives well-preserved shell areas were drilled while altered areas were omitted.
within the shallow sub-littoral zone. With a maximum life span of about For sampling the shells, 50-μm thin sections were cut and polished.
70 years and massive valves that reach up to 9 cm length, this bivalve is Horizontal drilling speed was adjusted to 700 μm/s, milled in eight
considered a useful climate archive (Gordillo et al., 2015). Species consecutive passes of 35 μm depth. The sampling track distance was 100
nomenclature used in this study is based on MolluscaBase Editorial μm. Digital images of the shell sections were taken with a Leica DFC-280
Board (2022). The description of the inner bivalve structures (e.g., digital camera attached to a Leica binocular microscope (Fig. 3b). The
shell-layer configuration) follows th e revised terminology of Schöne bivalve was sampled along a 2.25 cm long transect encompassing six
et al. (2011). growth increments (AI-1 to AI-6) with 256 single tracks (T1 to T256;
Fig. 3B). For obtaining sufficient carbonate material, up to three drilling
3.2. Microscopy of shell preservation status and shell microstructures tracks in the outer shell layer were combined in one sample, resulting in
a sampling resolution of more than 10 samples per year and 66 samples
The suitability of a specimen for isotopic analysis depends on its for stable isotope analysis (Fig. 3e: RE1 to RE66). Additional samples
preservation status. Possible impacts by bioerosion, recrystallization, or were drilled from the inner shell layer (n = 23) placed in increment I–C
physical abrasion were assessed visually in petrographic thin sections (Fig. 3e: REA1 to REA13) and increment I-D (Fig. 3e: REA14 to REA23).
using reflected and transmitted light microscopy. Thin sections were The powder micro-samples were analyzed for δ13Cshell and δ18Oshell
oriented along the maximum growth axis of the shell valve and groun­ using a Kiel III automated carbonate preparation device connected on­
ded to a slide thickness of 50 μm. The thin sections were stained with line to a ThermoFinnigan MAT 252 isotope-ratio mass-spectrometer
Feigl’s Solution to confirm its original aragonitic mineralogy (Leitmeier coupled to a GasBench II (Kim et al., 2007). All isotopic values are re­
and Feigl, 1934). Growth increment width was measured on the petro­ ported in the conventional δ-notation in parts per mil (‰) relative to the
graphic thin sections across the outer shell (Fig. 3a) and the hinge sec­ Vienna Pee Dee Belemnite (VPDB) using the NBS19 standard (δ18O =
tion (chondrophore; Fig. 4b) using the image analysis software ImageJ − 2.20‰). Reproducibility was evaluated by replicate analysis of labo­
1.45s. The microstructure was investigated with a Tescan Vega 3 XMU ratorial standards to ±0.03‰ 1σ for δ13Cshell and ±0.05‰ 1σ for
scanning electron microscope (SEM, 20.0 kV, secondary electron mode) δ18Oshell.
at the Leibniz Center for Tropical Marine Research (ZMT), Bremen,
Germany. Samples for SEM analysis were cut from those umbonal and 3.6. Modern environmental data integration
central shell positions (Fig. 3a), where growth lines and microstructures
were found in well-preserved conditions. The samples were etched with To capture the potential influence of both continental water
hydrochloric acid (0.25% HCl) for 10–20 s to better reveal the growth discharge and alternating shelf water masses on the stable isotopic sig­
lines, and gold-sputtered afterwards. nals recorded in the R. exalbidus shell, we obtained the modern oceanic

Table 1
Sampling location. GeoB – Bremen/MARUM Core Repository; VC – vibracore.
Sediment core Latitude (◦ S) Longitude (◦ W) Coring gear Water depth (m) Recovery (cm) Species

GeoB13802-2 36 05.30
◦ ′
53 20.72
◦ ′
VC 141 341 Retrotapes exalbidus
A. Klicpera et al.

Fig. 3. (A) Photo-micrograph of the investigated


deglacial specimen of Retrotapes exalbidus, (B) Close-
up of growth increments sampled (AI-1 to AI-6) and
MicroMill transects T1 to T256 on the outer shell; (C)
Schematic illustration of sampling scheme. Accre­
tionary increments were sampled starting at an age of
nine years onward in order to avoid metabolically
driven fractionation effects during early ontogeny
(AI-1 to AI-6 in panels B and C). (D) At a higher
resolution, each annual growth increment provides a
record of several (monthly) sub-increments, which
(E) at highest resolution show daily micro-growth
increments of 30–80 μm in individual width. ISL –
inner shell layer; OSL – outer shell layer; iOSL –
innermost outer shell layer; oOSL – outermost outer
shell layer.

We used averaged monthly temperature and salinity data from the


World Ocean Atlas 2010(Antonov et al., 2010; Locarnini et al., 2010) for
water masses identification (subtropical and subantartic water masses)
and an estimation of δ18Osw and δ18Oeq. For the estimation of δ18Osw
(VSMOW) based on salinity, the Tropical Atlantic (Equation (1);
LeGrande and Schmidt, 2006) and the South Atlantic (Equation (2);
LeGrande and Schmidt, 2006) equations were applied concerning the
subtropical and subantarctic water masses, respectively. An alternative
equation, calibrated for the shallow shelf and related to subantarctic
waters (Equation (3); Colonese et al., 2012), was used for comparison
reasons. A paleo-temperature equation for aragonite (Equation (4);
Grossman and Ku, 1986) was used for estimating δ18Oeq (VPDB).

Δ18Osw = 0.15 * S – 4.61 (1)


Fig. 4. (A) Shell height versus age plot for the investigated deglacial Retrotapes
exalbidus specimen (black line) in comparison to previously reported modern Δ18Osw = 0.51 * S – 17.4 (2)
specimen growth records (grey area; Lomovasky et al., 2002a) indicates a later
onset of maturity for the investigated specimen. (B) Full growth record of the
18
Δ Osw = 0.28 * S – 10.79 (3)
chondrophore (hinge section) of the investigated Retrotapes exalbidus is pre­ ◦ 18 18
T( C) = 20.6–4.34 * (δ Oeq – (δ Osw (SMOW) – 0.27)) (4)
sented as a mosaic of ten thin sections.
18
The seasonal range (amplitude) in δ Oeq estimated for each virtual
conditions (temperature, salinity), and estimated both stable oxygen oceanographic station was compared with the ice volume-corrected
isotopic composition of seawater (δ18Osw) and corresponding δ18O δ18Oivc seasonal range of the R. exalbidus fossil specimen (δ18Oivc-shell).
values precipitated in equilibrium (δ18Oeq) at virtual stations located During deglacial low-stand sea level conditions (Guilderson et al., 2000;
along inner and outer shelf. The selected five “coastal” stations on the Lambeck and Chappell, 2001) δ18Osw was ca. 0.77‰ (0.0081‰/m,
inner shelf (C1 to C5) were distributed close to the 50 m isobath, and five Fairbanks and Matthews, 1978; Lambeck and Chappell, 2001) higher
outer shelf stations (O1 to O4) located close to the 150 m isobaths, than modern δ18Osw. Hence, 0.77‰ was extracted from δ18Oshell before
including the particular site where the R. exalbidus shell was recovered comparison with estimated δ18Oeq. Since there is to our knowledge no
(E Core) (Fig. 7). species-specific paleotemperature equation for R. exalbidus we are
A. Klicpera et al.

unable to directly compare absolute δ18Oeq values to δ18Oivc-shell values. throughout the shell record ranged from 23.11 mm for the earliest ju­
However, because the slopes of different paleotemperature equations venile increment to 0.23 mm for the latest adult increment, with an
are similar (e.g., Grossman and Ku, 1986; Böhm et al., 2000; Kim et al., average increment width of 4.03 mm (N = 25).
2007) we can compare the seasonal range of the δ18Oeq to the seasonal Second-order subordinate increments (herein called micro-
range of the δ18Oivc-shell. increments) were obvious in the R. exalbidus shell as well (Fig. 3c, d,
e). In the collected specimen, the second-order increment widths ranged
3.7. Temporal alignment of Retrotapes exalbidus δ18Oshell from 9 to 164 μm (average = 46 μm), being narrowest towards the
annual growth lines and widest in the center of a first-order increment.
In order to compare the estimated δ18Oeq time series obtained from The analysis of the first 20 years, i.e., during juvenile growth gives a
modern stations that are provided as monthly values together with the mean growth rate of 3.96 mm yr− 1 (10.85 μm day− 1).
δ18Oivc-shell time series, both data series have to be plotted against the
same time axis. Importantly, modern WOA09 (Antonov et al., 2010; 4.2. Age control and stable oxygen and carbon isotopes micro-transects
Locarnini et al., 2010) data are monthly averaged over several decades,
whereas each shell value analyzed here represents those environmental The R. exalbidus specimen was dated around a median average age of
conditions that prevailed during a specific month or a short series of 14,275 cal ka BP (Table 2; Fig. 3d). This radiocarbon age corresponds to
month. Thus, shell raw data are assumed to record absolute values, mid-deglacial times and coincides with the Bølling-Allerød interstadial
while modern data are smoothed, showing average conditions and, thus, and the end of Meltwater Pulse 1A (Hanebuth et al., 2000). The
underrepresenting the true amplitude. To reduce this bias, we deter­ high-resolution δ18Oshell profile (N = 66, Fig. 5a) from the oOSL of
mined the average minimum and maximum of δ18Oeq within an annual R. exalbidus reveals a mean δ18Oshell(oOSL) value of 3.27 ± 0.42‰, and,
cycle. We found that δ18Oeq maxima, averaged across the region, occur after correction for ice volume, a value of 2.50 ± 0.42‰ (data provided
in September and minima in April. Then, we assigned the maximum in the supplementary table). The uncorrected δ18Oshell(oOSL) amplitude is
δ18Oivc-shell value within each annual increment (distinct micro­ 1.69‰ with a maximum value of 3.98‰ and a minimum of 2.29‰. The
structurally units that form over a period of several months each year) to temporally unaligned δ18Oshell(oOSL) signature (Fig. 5c; Suppl. Table 1)
match with September values in the δ18Oeq time series and, conversely, exhibits successive cycles of rapidly increasing values occurring shortly
the minimum δ18Oivc-shell to match April in the δ18Oeq time series. after growth-line formation, and more gradually decreasing values until
Missing values in the δ18Oivc-shell time series were reconstructed using a the newt growth line. As a result, the δ18Oshell(oOSL) signature describes a
moving average between consecutive values of δ18Oivc-shell. Thus, we characteristic asymmetric cyclic pattern. The time series shows most
obtained a monthly time series for the δ18Oivc-shell that allowed us to negative δ18Oshell(oOSL) values (high water temperatures) associated with
compare δ18Oeq and δ18Oivc-shell in equal statistical footing (i.e., 12 annual growth lines, while the most positive δ18Oshell(oOSL) values (low
values each per year; averaging monthly values for the six annual in­ water temperatures) appears shortly after the first third of growth in­
crements). This procedure reduces the variability of δ18Oivc-shell. Further, terval (Fig. 5a). No significant differences were found between the mean
one way Analysis of Variance (ANOVA) was performed to evaluate values of the six annual increments (ANOVA; p > 0.05), and thus no
differences in mean values between the six annual increments sampled significant ontogenetic trend was found (see Table 3).
in the analyzed R. exalbidus specimen. Significant differences may The δ13Cshell(oOSL) signal shows a mean value of 0.41 ± 0.49‰. The
13
indicate the differential action of vital effects during the observed δ Cshell(oOSL) amplitude range is 2.76‰ with a maximum of 1.02‰ and
interval. a minimum of − 1.74‰. The unaligned δ13Cshell(oOSL) signature exhibits a
slight trend towards lighter values per annual growth interval, inter­
4. Results rupted by sharp cyclic minima, being increasingly pronounced with
ontogeny. These negative excursions are closely associated with the five
4.1. Species determination, shell growth patterns, mineralogy and growth lines (Fig. 5c) and thus indicate a distinct positive correlation
preservation status with δ18Oshell(oOSL) minimum values.

The articulated shell from the base of core GeoB13802-2 (Fig. 1; 4.3. Environmental setting
Table 1) was identified as Retrotapes exalbidus (Dillwyn, 1817). The
R. exalbidus specimen shows a well-defined shell-layer configuration The investigated R. exalbidus specimen lived during mid-deglacial
with a well-preserved inner shell layer (ISL; Fig. 3a). The shell in­ times (14.5 cal ka BP). Although the regional post-LGM sea-level his­
crements are devoid of bioerosional traces and exhibit a very distinct tory is not yet well-understood (Violante and Parker, 2004, and refer­
growth-line pattern within the hinge plate that forms the chondrophore. ences therein), the eustatic sea-level at 14.5 cal ka BP can be roughly
The ontogenetically youngest/juvenile increments illustrate high calci­ estimated to a position at or lower than 110 m below present-day
fication rates (rapid shell growth) with a pronounced decrease in sea-level (Lambeck and Chappell, 2001; Hanebuth et al., 2009). Given
increment thickness towards ontogenetically older/adult increments. that the articulated R. exalbidus studied here was found at 145 m modern
SEM imagery revealed well-preserved, composite prismatic structures water depth (considering the sample depth below modern seafloor), its
(aragonite) for the outer shell layer (oOSL), subdivided by finer growth living depth can, accordingly, be estimated to a paleo-water depth of not
lines that run parallel to the main growth lines (possibly daily growth more than 50 m, i.e., the nearshore to inner-shelf zone. This reduced
increments; Chateigner, 2000; Gobac et al., 2009). XRD analysis depth is in agreement with Rosenberg (2009) and Häussermann and
revealed an aragonite-dominated mineralogy (Fig. 3e) with an inner Försterra (2009) who reported modern depth ranges for this species as
shell layer (iOSL) of purely aragonitic nature and an oOSL composed of maximum depths of 70 and 150 m.
92% aragonite (8% calcite). Both oOSL and iOSL provided an excellent
quality record for sampling in the highest resolution. 4.4. Modern conditions
Light microscopy revealed 25 paired major first-order growth in­
crements (dark band followed by a filigree white 50–100 μm thick In general, all stations present seasonal and salinity temperature
growth line), subdivided by (sub-annual) second-order growth in­ oscillations with higher values in austral summer/autumn and lower
crements, a pattern that was consistent along the entire oOSL transect ones in austral winter/spring. Salinity shows relatively less pronounced
(Fig. 3a). The same pattern was observed as a compressed record in the seasonal oscillations than temperature (Suppl. Fig. 1). The virtual sta­
chondrophore section (Fig. 4b) and the iOSL (Fig. 3b, d). The width of tions south of 37◦ 30′ S (Stations O1, O2, C1, C2) show reduced ampli­
the first-order growth increments found in the analyzed specimen tudes in δ18Oeq compared with the stations further north (Fig. 6). On the
A. Klicpera et al.

Table 2
Accelerator mass spectrometry 14C ages of dorsal (ontogenetic youngest) and ventral (ontogenetic oldest) samples from the investigated Retrotapes exalbidus specimen
in sediment core GeoB13802-2. GeoB – Bremen/MARUM Core Repository; POZ – Poznan Radiocarbon Laboratory, Poland.
14
Lab ID Species Depth in core (cm) Position in shell Raw C age (yr BP) 1σ age range (cal yr BP) Median age (cal a BP)

Poz-36091 R. exalbidus 341 dorsal 14,200 ± 70 14,340–14,080 14,340


Poz-36092 R. exalbidus 341 ventral 14,060 ± 60 14,530 - 14,250 14,210

Table 3
Statistical parameters of the stable carbon (δ13Cshell) and oxygen (δ18Oshell) isotopic records from the investigated deglacial Retrotapes exalbidus shell. (*) Location of
micro-sample in across-shell profile: oOS – outermost outer shell layer; iOSL – innermost outer shell layer. Ampl.: Total amplitude of isotopic range.
Species Location (*) δ18O (‰, VPDB) δ13C (‰, VPDB)

Range Mean Ampl. Range Mean Ampl.

R. exalbidus oOSL 2.29–3.98 3.27 ± 0.42 1.69 − 1.74–1.02 0.41 ± 0.49 2.76
R. exalbidus iOSL 2.93–3.97 3.44 ± 0.30 1.04 0.25–1.62 1.17 ± 0.37 1.37

Fig. 6. Comparison of the amplitude of the stable oxygen isotopic values


(δ18Oeq) calculated for modern conditions at the virtual stations (please see the
location of the virtual stations in Fig. 7) and monthly averages of the ice volume
corrected δ18Oshell of the analyzed specimen of Retrotapes exalbidus. Standard­
ized data show the relative amplitude of the isotopic signature for inner (C1 to
5) and outer (O1 to O4) shelf stations including the R. exalbidus core site (R
Core). Data are sorted from left to right based on minimum values.
Fig. 5. (A and B) Schematic drawing of the investigated Retrotapes exalbidus
specimen showing annual growth intervals (AI) along the analyzed transect. (C) Holocene when subtropical species even approached the Uruguayan
Ice volume corrected stable oxygen isotopic values (δ18Oivc-shell) (continuous coast, suggesting the oceanic conditions have been warmer than at
line) and stable carbon isotopic values (δ13Cshell) (dashed line) of the investi­ present-day (e.g. Martínez et al., 2006; Aguirre et al., 2011).
gated R. exalbidus shell. The neoaustral genus Retrotapes first appeared in Patagonia and
Antarctica during early Tertiary times and has been described in
outer shelf, δ18Osw mean values increase towards the north from − 0.1‰ Miocene and Holocene records from southern South America (del Rio,
(O1) to 0.8‰ (O4). Mean δ18Oeq values decrease towards the north from 1997; Gordillo, 2006). R. exalbidus is extant in the Magellan Region
ca. 2.9‰ (O1) to 0.7‰ (O4). Comparisons of normalized amplitudes of (Carcelles, 1944; Reid and Osorio, 2000). Its modern northernmost
the δ18Oeq values for all virtual stations are shown in Fig. 6. biogeographic occurrence (ca. 38◦ S; Penchaszadeh et al., 2008) co­
incides with the northernmost position of the MC and the associated
5. Discussion SASW (Figs. 1 and 2). The investigated specimen was recovered at 36◦ S,
i.e., about 220 km north of the modern species distribution limit (38◦ S)
5.1. Biogeographic evidence reported by Penchaszadeh et al. (2008). Scarabino et al. (2015), in
contrast, placed the modern northern boundary of R. exalbidus appear­
In general, our findings are in agreement with the known paleo- ance onto the Uruguayan shelf (North of 38◦ S)based on single detached
biogeographic distribution of faunal elements suggesting colder tem­ but freshly preserved valves that still contained their organic ligament.
peratures during deglacial times off Uruguay. For example, Aguirre et al. These results may point to modern habitat conditions that might be
(2009, 2011) reported that faunal-associations showed typical marginally suitable for such species in the Uruguayan shelf. This
cold-water elements north of their present-day northernmost boundary observation is also congruent with inferred cooler paleoceanographic
prior to the Holocene climatic amelioration, locally coinciding with the conditions in the region during the Bølling-Allerød interstadial and the
BMC (38◦ S) offshore northern Argentina, linked to a stronger Malvinas end of Meltwater Pulse 1A and an inner-shelf environment with water
current. This trend reversed at least for the shallower ocean towards the depth of (Guilderson et al., 2000; Hanebuth et al., 2000) during this time
A. Klicpera et al.

mechanisms may also have influenced the R. exalbidus specimen studied


here.
The trend towards lighter δ13Cshell during ontogenesis has been
observed in a number of studies (e.g., Keller et al., 2002; Gillikin et al.,
2007). Keller et al. (2002) attributed more negative δ13C values in older
shell parts of Chamelea gallina to the utilization of δ13CDIC derived from
pore waters associated with deeper burrowing into the seabed sediment.
Gillikin et al. (2007) correlated this isotopic trend with metabolic
changes during the organism’s life cycle. Yan et al. (2012) also noted
such a trend for R. exalbidus. The existence of an ontogenetically
controlled, fast-growing juvenile phase and a successive reduction
during the adult stage is typical for bivalves and assumed to be related to
the onset of sexual maturity (Gosling, 2003). Modern specimens of
E. exalbida show, for instance, an average growth rate not exceeding
2.50 mm yr− 1 (N = 214; Fig. 4a) and an onset of the adult stage after
15–20 years (Lomovasky et al., 2002a). The R. exalbidus specimen
studied here reveals a higher growth rate for the first 20 years (mean
3.96 mm yr− 1 = 10.85 μm day− 1), followed by a delayed onset of
maturity (25–30 year) compared to modern individuals collected in
Ushuaia Bay (Tierra del Fuego, southernmost South America; Lomova­
sky et al., 2002a). Such an ontogenetic offset of our R. exalbidus spec­
imen towards higher growth rates and later maturity can be attributed to
oceanic conditions closer to its environmental metabolic optimum (i.e.,
colder water, normal salinity, sufficient food supply) than it is the case
today in Usuahia Bay. This can be linked to an exposure to a higher
oceanic alkalinity and dissolved calcium concentration associated with
less marked estuarine conditions during mid deglacial times, when
compared with modern salinity ranges observed in Usuahia bay (Gor­
dillo, 2006). For modern Z. patagonica, growth rates have been shown to
Fig. 7. Location of modern (virtual) stations along the study area. Retrotapes
exalbidus specimen was found in station R CORE.
decrease along a geographic gradient southward of 39◦ 30′ S, while
Uruguayan stocks located between 35◦ 50′ S and 36◦ 50′ S show
decreasing growth rates towards the northern limit of its geographical
at the study site.
range (Gutierrez and Defeo, 2005). From our data, the scenario of an
exposure to a fully marine environment is unlikely, according to the
5.2. Sclerochronology δ13C values (see below). Thus, the sampling site may have been located
closer to the metabolic optimum of the species in comparison with
The internal banding of the early deglacial R. exalbidus specimen modern populations living in the southern limit of its distribution.
provided a record of 25 major growth bands that documents the growth
over a lifespan of 22–25 years. For comparison, R. exalbidus from the 5.3. Regional paleoceanographic and climatic implications
Magellan region is known to live up to 70 years (Gordillo, 2006; Gordillo
et al., 2013a; Lomovasky et al., 2002b; Yan et al., 2012). Additionally, This study presents the first subannually resolved δ18Oshell record
the interpretation of annually precipitated increments in our data set is from the early deglacial off SESA. The δ18Oshell record show most
corroborated by the stable isotopic signature, which provides negative values associated with annual growth lines and negative
well-defined cycles interpreted to be seasonal (Fig. 5). This finding is δ13Cshell peaks. This covariation between δ18O and δ13C values would be
consistent with previous the interpretation of first-order increments as expected under seawater-freshwater mixing (e.g., Wefer, 1985; Ingram
representing annual growth pattern (Lomovasky et al., 2002a; Ivany et al., 1996; Ivany et al., 2008), suggesting short-term increases in the
et al., 2008, and references therein). discharge of fluvial waters into the marine realm. There is no evidence of
The annual growth pattern of modern R. exalbidus shells has been increased rainfall coevally with minima in the δ18Oshell record, since
addressed in a number of studies (Dextraze and Zinsmeister, 1987; summer precipitation related to the South American monsoon system
Lomovasky et al., 2002a, 2002b; Ivany et al., 2008). Yan et al. (2012) most probably was not substantially increased during the
have shown that growth lines in modern shells do not simply reflect a Bølling-Allerød interstadial (Razik et al., 2013; Gu et al., 2018).
winterly cold-water signature, but that other key paramters coupled to According to Urien and Ottman (1971) and Lantzsch et al. (2014),
primary production, such as food availability, may have an influence on the wide SESA continental shelf was largely exposed and characterized
biochemical carbonate precipitation processes and, thus, shell growth by coastal plains during deglacial times, while the Plata River followed a
pattern. Contrary to the metabolism-controlled growth of first-order steeply northward-directed morphological depression right off the
increments, intra-annual growth structures are thought to be resulting modern Uruguayan coastline and was largely isolated from our sampling
from regional environmental parameters that include year-specific site (Fig. 1a). Because of the Plata River paleo-mouth having been
short-term variability in water temperature, food availability (primary located at the latitude of the Rio Grande Cone off southern Brazil during
production) and environmental stress factors (e.g., influencing the deglacial times (e.g., Chiessi et al., 2008; Lantzsch et al., 2014; Gu et al.,
timing of the spawning season; Gordillo et al., 2013). Control factors 2018), negative δ13C peaks should, thus, reflect the influence of
causing environmental stress such as seasonality in temperature, but non-point freshwater sources in the form of small fluvial distributaries
also the reproductive activity, have been shown to influence shell along the paleo-coast (e.g., Violante et al., 2008).
growth, demonstrated, for instance, on the long-lived North Atlantic The amplitude of the raw δ13C signal from our specimen is higher
cold-water Artica islandica and the Patagonian Zygochlamys patagonica (2.76‰) compared with data from Yan et al. (2012). These authors re­
(Jones 1980; Morriconi et al., 2002; Elliot et al., 2003; Schöne et al., ported an average amplitude of 0.73‰ (the highest amplitude recorded
2003, 2005; Lomovasky et al., 2008). It is, thus, likely that such control from an individual was 1.63‰), based on data from four young (4–10
A. Klicpera et al.

years old) modern specimens of R. exalbidus sampled at 51◦ 40′ S (salin­ amplitude of modern temperature conditions on the R Core site, sug­
ities averaging 33.68 ± 0.12). This observation reinforces the hypoth­ gesting only slightly lower mid-deglacial seasonal amplitude in tem­
esis of more pronounced variation in salinity at the site of collection of perature compared to present day conditions. Future research may focus
our bivalve specimen during the mid-deglaciation. As expected, our on (i) producing a species-specific paleo-temperature equation for R .
modern virtual oceanographic stations on the inner shelf reflect riverine Exalbidus, and (ii) dating and analyzing fossil specimens found on the
freshwater dilution of shelf water masses. This dilution effect is partic­ Brazilian shelf ca. 34◦ S.
ularly evident in intermediate stations close to the Plata River mouth
and reflects the highest continental discharge in the austral winter/­
spring (Simionato et al., 2001; Moller et al., 2001). Declaration of competing interest
However, according to our temporally aligned δ18Oshell data, the
seasonality of the environment experienced by the R. exalbidus specimen The authors declare that they have no known competing financial
differed only slightly from what is observed today in the sampling site, interests or personal relationships that could have appeared to influence
with the mid-deglacial specimen probably exposed to a reduced intra- the work reported in this paper.
annual variability (Fig. 6). Standardized data clearly shows a lat­
itudinal trend in modern amplitudes, with stations located south of Acknowledgements
36◦ 30’ (C1, C2, O1 and O2) exhibiting narrower amplitudes compared
with the remaining stations. In this comparison, data from the core site Chief Scientist Sebastian Krastel-Gudegast and the participants and
(R Core) showed a very similar amplitude to the isotopic signal recov­ crew of the Meteor Cruise M78/3a are gratefully acknowledged for
ered from the shell. Further, modern data from our virtual stations collecting the sample material and for providing sample background
(Fig. 7) indicate that the northernmost stations are bathed by relatively data. Sebastian Flotow (ZMT Bremen, Germany) is thanked for prepar­
warm and salty water masses (i.e., STSW) while the southernmost sta­ ing the thin sections. Matthias López Correa (GeoZentrum Nordbayern,
tions represent relatively cold and fresh water masses (i.e., SASW). In­ Germany) is thanked for support with the MicroMill sampling. Michael
termediate stations are located at the confluence of both shelf water Joachimski, Daniele Lutz (both GeoZentrum Nordbayern, Germany)
masses (i.e., near the STSF). Over the Uruguayan shelf, below 100 m performed stable oxygen and carbon isotopes analyses. We also thank
water depth, seasonal changes in water temperatures are driven by the Felipe García-Rodríguez (Facultad de Ciencias, UdelaR Montevideo,
relative dominance of STSW over SASW in autumn, whereas spring is Uruguay) and Fabrizio Scarabino (Museo Nacional de Historia Natural
characterized by a dominance of SASW (Ortega and Martínez, 2007). In in Montevideo, Uruguay) for their local support and helpful discussions.
contrast, the general seasonal temperature oscillation observed within Leonardo Ortega is acknowledged for his help with the oceanographic
water masses likely reflects the cooling of the upper layer during cold analyses. This project was funded through the DFG-Research Center/
seasons (Piola et al., 2000). Taken together, the evidence points out to a Cluster of Excellence “The Ocean in the Earth System” Project SD-2. C.M.
more pronounced dominance of SASW over shallower coastal areas C. acknowledges the financial support from FAPESP (grants 2018/
during Bølling-Allerød interstadial. 15123–4, and 2019/24349–9), CNP (grant 312458/2020–7), CAPES
(grant 88881.313535/201901) and the Alexander von Humboldt
6. Conclusions Foundation.

Although we recognise the limited value of the information retrieved Appendix A. Supplementary data
from a single specimen, this study presents the first high-resolution
reconstruction of seasonal oceanic variability on the continental shelf Supplementary data to this article can be found online at https://doi.
off southeastern South America (SESA) during mid-deglacial times. We org/10.1016/j.csr.2023.105014.
analyzed the 6-years spanning δ18O and δ13C stable isotope record of an
in-situ preserved, long-living bivalve specimen. Allowing for a compar­
References
ative approach, the modern bathymetric and biogeographic distribution
pattern is known and stable isotopic data exist for modern R. exalbidus, Aguirre, M., Richiano, S., Álvarez, M.F., Eastoe, C., 2009. Malacofauna cuaternaria del
although the modern environmental conditions experienced by inves­ litoral norte de santa cruz (Patagonia, Argentina). Geobios (Lyon) 42, 411–434.
Aguirre, M.L., Donato, M., Richiano, S., Farinati, E.A., 2011. Pleistocene and Holocene
tigated individual organisms may not be fully comparable to those
interglacial molluscan assemblages from Patagonian and Bonaerensian littoral
environmental conditions experienced by the collected fossil specimen (Argentina, SW Atlantic): palaeobiodiversity and palaeobiogeography. Palaeogeogr.
that lived about 14.5 ka ago. Palaeoclimatol. Palaeoecol. 308, 277–292.
An accurate paleoceanographic reconstruction of ocean water pa­ Antonov, J.I., Seidov, D., Boyer, T.P., Locarnini, R.A., Mishonov, A.V., Garcia, H.E.,
Baranova, O.K., Zwerg, M.M., Johnson, D.R., 2010. World Ocean Atlas 2009 Volume
rameters in the absence of species-specific paleo-temperature equations 2: Salinity. S. Levitus Ed. NOAA Atlas NESDIS 69. U.S. Gov. Printing Office,
is a challenging approach. Our new data help to characterize the Washington, D.C., p. 184
participating processes that were once active on the SESA continental Bender, V.B., Tjj, Hanebuth, Chiessi, C.M., 2013. Holocene shifts of the Subtropical Shelf
Front off southeastern South America controlled by high and low latitude
shelf. These data pointed out to colder waters during deglacial condi­ atmospheric forcing. Paleoceanography 28, 1–10.
tions, and δ13C suggests a transition to a more variable salinity regime in Böhm F, Joachimski M, Dullo W-C, Eisenhauer A, Lehnert H, Reitner J, Wörheide G (200)
nearshore areas during the Bølling-Allerød interstadial. The δ18O and Oxygen isotope fractionation in marine aragonite of coralline sponges. Geochem.
Cosmochim. Acta, 64, 1695–1703.
δ13C record from the R. exalbidus specimen presented here shows coin­ Boretto, G., Zanchetta, G., Consoloni, I., Baneschi, I., Guidi, M., Isola, I., Bini, M.,
cident and abrupt negative peaks that indicate pulses of freshwater Ragaini, L., Terrasi, F., Regattieri, E., Dallai, L., 2020. Stable oxygen and carbon
discharge. Further, since both series shows coincident peaks, the tem­ isotope composition of Holocene mytilidae from the camarones coast (chubut,
Argentina): palaeoceanographic implications. Water 12 (12), 3464.
perature dependence of isotopic fractionation processes cannot be Brandini, F.P., Boltovskoy, D., Piola, A., Kocmur, S., Röttgers, R., Cesar Abreu, P.,
excluded as a driver of the observed pattern. Despite the challenge of Mendes Lopes, R., 2000. Multiannual trends in fronts and distribution of nutrients
identifying the exact mechanism responsible for these negative δ13C and chlorophyll in the southwestern Atlantic (30–62◦ S). Deep-Sea Res. Pt. I -
Oceanographic Research Paper 47 (6), 1015–1033.
excursions, the existence of minor fluvial distributaries with their river
Brazeiro, A., Acha, E.M., Mianzan, H., Gómez-Erache, M., Fernández, V., 2003. Aquatic
mouths located all along the coast might offer a plausible explanation priority areas for the conservation and management of the ecological integrity of the
for diverting the isotopic signals. Since the shell collection site was Rio de la Plata and its maritime front, p. 81. Technical Report PNUD Project/GEF
located in shallow waters and closer to the respective mid-deglacial RLA/99/G31.
Burrage, D., Wesson, J., Martinez, C., Pérez, T., Möller, O., Piola, A., 2008. Patos Lagoon
coastline, this hypothesis is worth exploring in the future. Impor­ outflow within the Río de la Plata plume using an airborne salinity mapper:
tantly, the overall amplitude of the δ18Oivc-shell record is similar to the observing an embedded plume. Continent. Shelf Res. 28, 1625–1638.
A. Klicpera et al.

Campos, E., Piola, A., Matano, R., Miller, J., 2008. PLATA: a synoptic characterization of Ivany, L.C., Runnegar, B., 2010. Early Permian seasonality from bivalve 18O and
the southwest Atlantic shelf under influence of the Plata River and Patos Lagoon implications for the oxygen isotopic composition of seawater. Geology 38,
outflows. Continent. Shelf Res. 28, 1551–1555. 1027–1030.
Carcelles, A., 1944. Catálogo de los moluscos marinos de Puerto Quequén, vol. 3. Revista Ivany, L.C., Lohmann, K.C., Hasiuk, F., Blake, D.B., Glass, A., Aronson, R.B., Moody, R.
del Museo Argentino de Ciencias Naturales, Sección Zoología, pp. 233–309. M., 2008. Eocene climate record of a high southern latitude continental shelf:
Carranza, A., Scarabino, F., Brazeiro, A., Ortega, L., Martínez, S., 2008. Assemblages of seymour Island, Antarctica. Geol. Soc. Am. Bull. 120, 659–678.
megabenthic gastropods from Uruguayan and northern Argentinean shelf: spatial Jones, D.S., 1980. Annual cycle of shell growth increment formation in two continental
structure and environmental controls. Continent. Shelf Res. 28, 788–796. shelf bivalves and its paleoecologic significance. Paleobiology 6, 331–340.
Chateigner, D., 2000. Mollusc shell microstructures and crystallographic textures. Jones, D.S., Quitmyer, I.R., 1996. Marking time with bivalve shells. Palaios 11, 340–346.
J. Struct. Geol. 22, 1723–1735. Kaiser, P., 1977. Beiträge zur Kenntnis der Voluten (Mollusca) in argentinisch-
Chiessi, C.M., Mulitza, S., Paul, A., Pätzold, J., Groeneveld, J., Wefer, G., 2008. South brasilianischen Gewässern (mit der Beschreibung zweier neuer Arten), vol. 74.
Atlantic interocean exchange as the trigger for the Bølling warm event. Geology 36, Mitteilungen Hamburg Zoologisches Museum Institute, pp. 11–26.
919–922. Keller, N., Del Piero, D., Longinelli, A., 2002. Isotopic composition, growth rates and
Colonese, A.C., Verdún-Castelló, E., Álvarez, M., i Godino, I.B., Zurro, D., Salvatelli, L., biological behaviour of Chamelea gallina and Callista chione from the Bay of Trieste
2012. Oxygen isotopic composition of limpet shells from the Beagle Channel: (Italy). Mar. Biol. 140, 9–15.
implications for seasonal studies in shell middens of Tierra del Fuego. J. Archaeol. Kim, S.T., O’Neil, J.R., Hillaire-Marcel, C., Mucci, A., 2007. Oxygen isotope fractionation
Sci. 1;39 (6), 1738–1748. between synthetic aragonite and water: influence of temperature and Mg2+
Combes, V., Matano, R.P., 2014. Trends in the Brazil/Malvinas confluence region. concentration. Geochem. Cosmochim. Acta 71 (19), 4704–4715.
Geophys. Res. Lett. 41, 8971–8977. Krastel, S., Wefer, G., cruise participants, 2012. Report and preliminary results of RV
de Souza, R.B., Robinson, I.S., 2004. Lagrangian and satellite observations of the METEOR Cruise M78/3. Sediment transport off Uruguay and Argentina: from the
Brazilian Coastal Current. Continent. Shelf Res. 24, 241–262. shelf to the deep sea. 19.05.2009 – 06.07.2009. Montevideo (Uruguay) – Montevideo
de Souza, M.M., Mathis, M., Pohlmann, T., 2019. Driving mechanisms of the variability (Uruguay). Meteor Fahrtberichte 285, 79.
and long-term trend of the Brazil–Malvinas confluence during the 21st century. Clim. Lambeck, K., Chappell, J., 2001. Sea level change through the last glacial cycle. Science
Dynam. 53, 6453–6468. 292, 679–686.
del Rio, C.J., 1997. Cenozoic biogeographic history of the eurythermal genus Retrotapes, Lantzsch, H., Tjj, Hanebuth, Chiessi, C.M., Schwenk, T., Violante, R., 2014. The high-
new genus (Subfamily Tapetinae) from Southern South America and Antarctica. supply, current-dominated continental margin of southeastern South America during
Nautilus 110, 77–93. the late Quaternary. Quat. Res. 81, 339–354.
Depetris, P.J., Griffin, J.J., 1968. Suspended load in the Rio de La Plata drainage basin. LeGrande, A.N., Schmidt, G.A., 2006. Global gridded data set of the oxygen isotopic
Sedimentology 11, 53–60. composition in seawater. Geophys. Res. Lett. 33, 1–5.
Dettman, D.L., Lohmann, K.C., 1995. Microsampling carbonates for stable isotope and Leitmeier, H., Feigl, F., 1934. Eine einfache Reaktion zur Unterscheidung von Calcit und
minor element analysis: physical separation of samples on a 20-micrometer scale. Aragonit. Zeitschrift für Kristallographie, Mineralogie und Petrographie 45,
J. Sediment. Res. 65, 566–569. 447–456.
Dextraze, B.L., Zinsmeister, W.J., 1987. A study of the internal annual growth lines of the Locarnini, R.A., Mishonov, A.V., Antonov, J.I., Boyer, T.P., Garcia, H.E., Baranova, O.K.,
late Eocene mollusk Eurhomalea antarctica. Antarct. J. U. S. 22, 14–15. Zweng, M.M., Johnson, D.R., 2010. In: Levitus, S. (Ed.), World Ocean Atlas 2009,
Elliot, M., DeMenocal, P.B., Linsley, B.K., Howe, S.S., 2003. Environmental controls on Volume 1: Temperature, NOAA Atlas NESDIS 68. U.S. Government Printing Office,
the stable isotopic composition of Mercenaria mercenaria: potential application to Washington, D.C., p. 184
paleoenvironmental studies. G-cubed 4, 1056. Lomovasky, B.J., Brey, T., Morriconi, E., Calvo, J., 2002a. Growth and production of the
Emilsson, I., 1961. The Shelf and Coastal Waters off Southern Brazil, vol. 11. Boletim do venerid bivalve Eurhomalea exalbida in the Beagle Channel, Tierra del Fuego. J. Sea
Instituto Oceanográfico da Universidade de São Paulo, pp. 102–112. Res. 48, 209–216.
Fairbanks, R.G., Matthews, R.K., 1978. The marine oxygen isotope record in Pleistocene Lomovasky, B.J., Morriconi, E., Brey, T., Calvo, J., 2002b. Individual age and connective
coral, Barbados, West Indies. Quat. Res. 10, 181–196. tissue lipofuscin in the hard clam Eurhomalea exalbida. J. Exp. Mar. Biol. Ecol. 276,
Floeter, S.R., Soares-Gomes, A., 1999. Biogeographic and species richness patterns of 83–94.
gastropoda on the southwestern atlantic. Rev. Bras. Biol. 59, 567–575. Lomovasky, B.J., Lasta, M., Valiñas, M., Bruschetti, M., Ribeiro, P., Campodónico, S.,
Gobac, Z.Z., Posilovic, H., Bermanec, V., 2009. Identification of biogenic calcite and Iribarne, O., 2008. Differences in shell morphology and internal growth pattern of
aragonite using SEM. Geol. Croat. 9611, 201–206. the Patagonian scallop Zygochlamys patagonica in the four main beds across their
Gordillo, S., 2006. Pleistocene Retrotapes de Rió, 1997 (Veneridae, Bivalvia) from Tierra SW Atlantic distribution range. Fish. Res. 89 (3), 266–275.
del Fuego, Argentinia. Ameghiniana 43, 757–761. Lumpkin, R., Garzoli, S., 2011. Interannual to decadal changes in the western South
Gordillo, S., Brey, T., Beyer, K., Lomovasky, B., 2013. Retrotapes exalbidus from southern Atlantic’s surface circulation. J. Geophys. Res.: Oceans 116 (C1).
South America: are fossil shells reliable proxy archives for Holocene climate. In: 3rd Martínez, S., Rojas, A., Ubilla, M., Verde, M., Perea, D., Piñeiro, G., 2006. Molluscan
International Sclerochronology Conference. Caernarfon, Wales, UK, p. 162. assemblages from the marine Holocene of Uruguay: composition, geochronology and
Gordillo, S., Brey, T., Beyer, K., Lomovasky, B.J., 2015. Climatic and environmental paleoenvironmental signals. Ameghiniana 43 (2), 385–397.
changes during the middle to late Holocene in southern South America: a Matano, R.P., Schlax, M.G., Chelton, D.B., 1993. Seasonal variability in the southwestern
sclerochronological approach using the bivalve Retrotapes exalbidus (Dillwyn) from Atlantic. J. Geophys. Res.: Oceans 98, 18027–18035.
the Beagle Channel. Quat. Int. 377, 83–90. Matano, R.P., Palma, E.D., Piola, A.R., 2010. The influence of the Brazil and Malvinas
Gosling, E., 2003. Bivalve Molluscs: Biology, Ecology and Culture. Blackwell Publishing currents on the southwestern atlantic shelf circulation. Ocean Sci. 6, 983–995.
Ltd, Oxford, UK, p. 443. Moller, O.O., Castaing, P., Salomon, J.-C., Lazure, P., 2001. The Influence of local and
Grossman, E., Ku, T.L., 1986. Oxygen and carbon isotope fractionation in biogenic non-local forcing effects on the subtidal circulation of Patos Lagoon. Estuaries 24,
aragonite: temperature effects. Chem. Geol. 59, 59–74. 297.
Guilderson, T.P., Burckle, L., Hemming, S., Peltier, W.R., 2000. Late Pleistocene sea level MolluscaBase Editorial Board, 2022. MolluscaBase. Eurhomalea cossmann, 1920.
variations derived from the Argentine Shelf. G-cubed 1, 2000GC000098. Accessed through: World Register of Marine Species at: https://www.marinespecies.
Gillikin, D.P., Lorrain, A., Meng, L., Dehairs, F., 2007. A large metabolic carbon org/aphia.php?p=taxdetails&id=492474 on 2022-12-16.
contribution to the δ13C record in marine aragonitic bivalve shells. Geochem. Morriconi, E., Lomovasky, B.J., Calvo, J., Brey, T., 2002. The reproductive cycle of
Cosmochim. Acta 71 (12), 2936–2946. Eurhomalea exalbida (chemnitz, 1795) (Bivalvia: veneridae) in Ushuaia bay (54
Gu, F., Chiessi, C.M., Zonneveld, K.A., Behling, H., 2018. Late Quaternary environmental degrees 50′ S), beagle channel (Argentina). Invertebr. Reprod. Dev. 42, 61–68.
dynamics inferred from marine sediment core GeoB6211-2 off southern Brazil. Olivier, S.R., Scarabino, V., 1972. Distribución ecológica de algunos moluscos recogidos
Palaeogeogr. Palaeoclimatol. Palaeoecol. 496, 48–61. por la expedición del “Walther Herwig” (R.F.A.) al Atlántico sudoccidental (1966).
Gu, F., Chiessi, C.M., Zonneveld, K.A., Behling, H., 2019. Shifts of the Brazil-falklands/ J. Cell Biol. 322, 235–247.
Malvinas confluence in the western South Atlantic during the latest Olson, D.B., Podestá, G.P., Evans, R.H., Brown, O.B., 1988. Temporal variations in the
pleistocene–holocene inferred from dinoflagellate cysts. Palynology 43, 483–493. separation of Brazil and Malvinas currents. Deep sea research Part A. Oceanographic
Gutierrez, N.L., Defeo, O., 2005. Spatial patterns in population dynamics of the scallop Research Papers 35, 1971–1990.
Psychrochlamys patagonica at the northern edge of its range. J. Shellfish Res. 24 (4), Ortega, L., Martínez, A., 2007. Multiannual and seasonal variability of water masses and
877–882. fronts over the Uruguayan shelf. J. Coast Res. 23, 618–629.
Hanebuth, T., Stattegger, K., Grootes, P.M., 2000. Rapid flooding of the Sunda Shelf: a Penchaszadeh, P.E., Pastorino, G., Brogger, M.I., 2008. Atlas de sensibilidad ambiental
late-glacial sea-level record. Science 288 (5468), 1033–1035. del mar y de la costa. www.atlas.ambiente.gov.ar.
Hanebuth, T.J., Stattegger, K., Bojanowski, A., 2009. Termination of the last glacial Piola, A.R., Campos, E.J.D., Möller, O.O., Charo, M., Martinez, C., 2000. Subtropical shelf
maximum sea-level lowstand: the sunda-shelf data revisited. Global Planet. Change front off eastern South America. J. Geophys. Res. 105, 6565–6578.
66 (1–2), 76–84. Piola, A.R., Matano, R.P., Palma, E.D., Ller, O.O.M., Campos, E.J.D., 2005. The influence
Häussermann, V., Försterra, G., 2009. Marine Benthic Fauna of Chilean Patagonia. of the Plata River discharge on the western South Atlantic shelf. Geophys. Res. Lett.
Nature in Focus, p. 1000. Santiago. 32, L01603.
Heaton, T.J., Köhler, P., Butzin, M., Bard, E., Reimer, R.W., Austin, W.E.N., Bronk Piola, A.R., Romero, S.I., Zajaczkovski, U., 2008. Space-time variability of the Plata
Ramsey, C., Hughen, K.A., Kromer, B., Reimer, P.J., Adkins, J., Burke, A., Cook, M.S., plume inferred from ocean color. Continent. Shelf Res. 28, 1556–1567.
Olsen, J., Skinner, L.C., 2020. Marine20 – the marine radiocarbon age calibration Piola, A.R., Palma, E.D., Bianchi, A.A., Castro, B.M., Dottori, M., Guerrero, R.A.,
curve (0-55,000 cal BP). Radiocarbon 62, 779–820. Marrari, M., Matano, R.P., Oojr, Möller, Saraceno, M., 2018. Physical oceanography
Ingram, B.L., Conrad, M.E., Ingle, J.C., 1996. Stable isotope and salinity systematics in of the SW atlantic shelf: a review. Plankton ecology of the Southwestern Atlantic
estuarine waters and carbonates: san Francisco Bay. Geochem. Cosmochim. Acta 60 37–56.
(3), 455–467.
A. Klicpera et al.

Razik, S., Chiessi, C.M., Romero, O.E., von Dobeneck, T., 2013. Interaction of the South Simionato, C.G., Nuñez, M.N., Engel, M., 2001. The salinity front of the Río de la Plata - a
American monsoon system and the southern westerly wind belt during the last 14 numerical case study for winter and summer conditions. Geophys. Res. Lett. 28 (13),
kyr. Palaeogeogr. Palaeoclimatol. Palaeoecol. 374, 28–40. 2641–2644.
Reid, D.G., Osorio, C., 2000. The shallow-water marine mollusca of the estero elefantes Stuiver, M., Robinson, S.W., Yang, I.C., 2020. 14C dating to 60,000 Years BP with
and laguna san rafael, southern Chile. Bulletin of the Natural History Museum, proportional counters. In: Radiocarbon Dating. University of California Press,
London (Zoology) 66, 109–146. pp. 202–215.
Rosenberg, G., 2009. Malacolog 4.1.1: A Database of Western Atlantic Marine Mollusca Urien, C.M., Ottman, F., 1971. Historie du Rio de la Plata au Quaternaire, vol. XIV.
[WWW database (version 4.1.1)] URL. http://www.malacolog.org. Quaternaria, pp. 51–59.
Scarabino, F., 2003. Lista Sistematica de los Bivalvos Marinos Y Estuarios de Uruguay. Urien, C.M., Martins, L.R., Martins, I.R., 1980. Evolução geológica do Quaternário do
Comun. Soc. Malacol. Urug. 8, 227–258. litoral atlântico uruguaio, plataforma continental e regiões vizinhas. Notas Técnicas
Scarabino, F., Zelaya, D.G., Orensanz, J.L., Ortega, L., Defeo, O., Schwindt, E., 3, 7–43.
Carranza, A., Zaffaroni, Gastón Martínez, J.C., Scarabino, V., García-Rodríguez, F., Violante, R., Parker, G., 2004. The post-last glacial maximum transgression in the de la
2015. Cold, warm, temperate and brackish: bivalve biodiversity in a complex Plata River and adjacent inner continental shelf, Argentina. Quat. Int. 114, 167–181.
oceanographic scenario (Uruguay, southwestern Atlantic). Am. Malacol. Bull. 33, Violante, R.A., Cavallotto, J.L., Kandus, P., 2008. Río de la Plata y Delta del Paraná.
284–301. Sitios de interés geológico de la República Argentina. Servicio Geológico Minero
Schmid, C., Garzoli, S.L., 2009. New observations of the spreading and variability of the Argentino, vol. 46. Instituto de Geología y Recursos Minerales. Anales, Buenos Aires,
antarctic intermediate water in the atlantic. J. Mar. Res. 67 (6), 815–843. pp. 461–475.
Schöne, B.R., Tanabe, K., Dettman, D.L., Sato, S., 2003. Environmental controls on shell Voigt, I., Chiessi, C.M., Prange, M., Mulitza, S., Groeneveld, J., Varma, V., Henrich, R.,
growth rates and δ18O of the shallow-marine bivalve mollusk Phacosoma japonicum 2015. Holocene shifts of the southern westerlies across the South Atlantic.
in Japan. Mar. Biol. 142, 473–485. Paleoceanography 30, 39–51.
Schöne, B.R., Houk, S., Freyre Castro, A., Fiebig, J., Oschmann, W., Kroncke, I., Wefer, G., 1985. Die verteilung stabiler isotope in kalkschalen mariner organismen. Geol.
Dreyer, W., Gosselck, F., 2005. Daily growth rates in shells of Arctica islandica: Jahrbuch Reihe 82, 114.
assessing sub-seasonal environmental controls on a long-lived bivalve mollusk. Wells, P.G., Daborn, G.R., 1997. The Río de la Plata: an environmental overview; an
Palaios 20, 78–92. EcoPlata Project background report. Dalhousie University, Halifax, NS, CA.
Schöne, B.R., Radermacher, P., Zhang, Z., Jacob, D.E., 2011. Crystal fabrics and element Yan, L., Schöne, B.R., Arkhipkin, A., 2012. Eurhomalea exalbida (Bivalvia): a reliable
impurities (Sr/Ca, Mg/Ca, and Ba/Ca) in shells of Arctica islandica—implications for recorder of climate in southern South America? Palaeogeogr. Palaeoclimatol.
paleoclimate reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol. 373, 50–59. Palaeoecol. 350–352, 91–100.

You might also like