Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Quaternary Research 75 (2011) 24–35

Contents lists available at ScienceDirect

Quaternary Research
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / y q r e s

The application of foraminifera to reconstruct the rate of 20th century sea level rise,
Morbihan Golfe, Brittany, France
Veronica Rossi a,⁎, Benjamin P. Horton b, D. Reide Corbett c,d, Eduardo Leorri c,
Lucia Perez-Belmonte e, Bruce C. Douglas f
a
Dipartimento di Scienze della Terra e Geologico-Ambientali, Università di Bologna, Via Zamboni 67, 40126 Bologna, Italy
b
Department of Earth and Environmental Science, University of Pennsylvania, Philadelphia, PA 19104, USA
c
Department of Geological Sciences, East Carolina University, Greenville, NC 27858, USA
d
Institute for Coastal Science and Policy, East Carolina University, Greenville, NC 27858, USA
e
Université de Bretagne Sud, Vannes, France
f
International Hurricane Center, Florida International University, Miami, FL 33199, USA

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

Article history: Foraminiferal assemblages preserved within salt-marsh sediment can provide an accurate and precise means to
Received 23 December 2009 reconstruct relative sea level due to a strong relationship with elevation, which can be quantified using a transfer
Available online 12 October 2010 function. We collected a set of surface samples from two salt marshes in the Morbihan Golfe, France to determine
foraminiferal distribution patterns. Dominant taxa included Jadammina macrescens, Trochammina inflata,
Keywords:
Haplophragmoides spp. and Miliammina fusca. We developed a foraminifera-based transfer function using a
Foraminifera
Transfer function
modern training set of 36 samples and 23 species. The strong relationship between observed and predicted values
Sea level (r2jack = 0.7) indicated that foraminiferal distribution is primarily controlled by elevation with respect to the tidal
Salt marshes frame and precise reconstructions of former sea level are possible (RMSEPjack = 0.07 m). The application of the
transfer function to a short salt-marsh core (0.32 m) allowed the reconstruction of former sea levels, which were
placed in a chronological framework using short-lived radionuclides (210Pb and 137Cs). The agreement between
the foraminifera-based sea level curve and the Brest tide-gauge record confirms the reliability of transfer function
estimates and the validity of this methodology to extend sea level reconstructions back into the pre-instrumental
period. Both instrumental and microfossil records suggest an acceleration of sea level rise during the 20th century.
© 2010 University of Washington. Published by Elsevier Inc. All rights reserved.

Introduction With recent advances in sea level research, instrumental records


can be augmented with high-resolution relative sea level (RSL)
Accurate estimates of sea level rise in the pre-satellite era are records derived from salt-marsh sedimentary sequences. Indeed, salt-
needed in order to provide a context for the 21st century-estimates marsh studies in North America (Gehrels et al., 2002, 2005; van de
and calibrate climate models (Douglas, 2008). Good quality tide- Plassche et al., 2003; Donnelly et al., 2004; Edwards et al., 2004; Kemp
gauge data provide direct evidence for sea level accelerations and et al., 2009b), Europe (Gehrels et al., 2006; Leorri et al., 2008b; Vicente
decelerations (e.g., Douglas, 1991, 1997; Peltier, 2001; Church and et al., 2008) and elsewhere (Gehrels et al., 2008) support the inference
White, 2006; Jevrejeva et al., 2008; Woodworth et al., 2008). During that modern rates of relative sea level rise (last 100 years) may be
the 20th century the global rate of sea level rise recorded by tide- more rapid than the long-term rate of rise (between 800 and
gauges was 1.7 to 1.8 mm/yr (Douglas, 1991, 1997; Peltier, 2001; 1000 years ago), and that the timing of the 20th century acceleration
Church and White, 2006), which is much greater than the long-term may be indicative of a link with human-induced climate change.
rate of rise that has been inferred from late Holocene (last 4000 years) This paper aims to reconstruct RSL from the salt marshes of the
geological records (Peltier, 1996; Shennan and Horton, 2002; Morbihan Golfe, Brittany, France and to validate these reconstructions
Donnelly et al., 2004; Engelhart et al., 2009). However, the limited using the nearby long tide-gauge record from Brest. High-resolution
number, distribution and duration of tide-gauges precludes efforts to reconstructions of RSL may be developed using transfer functions,
robustly test climate vs. sea level hypotheses and establish the driving which quantify the relationship between salt-marsh microfossils and
mechanisms responsible for such changes (Woodworth et al., 2008). tide level (e.g., Horton et al., 1999; Gehrels et al., 2001; Horton and
Edwards, 2006; Leorri et al., 2008a,b; Woodroffe, 2009; Hawkes et al.,
2010). The most precise sea level reconstructions are those derived from
⁎ Corresponding author. Fax: + 39 051 2094522. microfossils occupying the uppermost part of the intertidal zone
E-mail address: veronica.rossi4@unibo.it (V. Rossi). (Gehrels, 2007; Kemp et al., 2009b). We first describe the modern

0033-5894/$ – see front matter © 2010 University of Washington. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.yqres.2010.07.017
V. Rossi et al. / Quaternary Research 75 (2011) 24–35 25

distribution of salt-marsh foraminifera from two salt marshes and ramosissima (purple glasswort). Algal mats characterize the tidal flat
subsequently employ the relationship between foraminifera and located close to the Kréagan creek. St. Goustan marsh (Fig. 1b) is located
elevation to reconstruct RSL from fossil assemblages preserved within close to the apex of Noyalo River Estuary. The floral species of St. Goustan
a 20th-century salt-marsh sedimentary sequence. are very similar to Kréagan Marsh. The extensive high marsh is dominated
by Phragmites spp. and Juncus geradi, with a small low marsh that has a
Study area notable presence of Spartina spp. Sins Marsh (Fig. 1b) is located about
1 km to the south of St. Goustan Marsh. Sins Marsh consists of an extensive
The Morbihan Golfe is a natural harbor in Northwest France high-marsh area covered by Juncus geradi with freshwater shrubs at the
(southern Brittany), sheltered from the Atlantic Ocean (Bay of Biscay) landward boundary. A middle marsh of Spartina spp.-dominated
by the peninsula of Rhuys, Belle Ile Island and the structural arc vegetation is present. A bank is present at the seaward edge of Sins
formed by peninsula of Quiberon, Houat and Hoedic islands (Fig. 1). It Marsh, which precludes the collection of data from a low-marsh or tidal-
is a shallow basin (ca. 11,500 ha) with a maximum water depth value flat environment.
of 30 m and is connected to the sea by a 900 m wide and 30 m deep
tidal inlet (Fig. 1). Across this entrance strong tidal currents affect the Materials and methods
basin with water speeds that can reach 4.6 m s− 1 at spring tides
(Marcos et al., 1995; Henocque, 2003). Morbihan Golfe has a semi- We selected Kréagan and Sins marshes to produce a modern training
diurnal tidal regime (tidal range ca. 3 m at spring tide) with set of salt-marsh foraminiferal data. The sampling stations were placed
freshwater input from the Auray, Vincin and Noyalo rivers. at uniform elevation intervals. At Kréagan Marsh we collected 21 surface
Since the mid Holocene, the Morbihan area has experienced strong samples along a 50 m transect that was orientated perpendicular to the
human influence. There is an extraordinary range of megalithic tidal channel across the intertidal zone from the tidal flat to the
monuments preserved around the Golfe, the oldest of which is ca. Phragmites spp. and Juncus spp. high marsh. At Sins Marsh we collected
3300 years BC, suggesting that the first stable human settlements date 21 samples from two salt-marsh transects (81 m long transect I and
back to the Neolithic. Extensive reclamation for agricultural use began 32 m long transect II), which included high marsh/upland boundary.
during the middle ages and continued until the 19th century (e.g., The elevation of each modern sample was determined using a total
Billaud, 1984; Astill and Langdon, 1997). Thereafter, much of the station, referenced to the local tide gauge of the Auray and Noyalo
reclaimed salt marshes were abandoned and subsequently returned rivers, and expressed relative to the local ordnance datum (IGM69). We
to intertidal environments (e.g., Brigand et al., 1992). selected the St. Goustan marsh for stratigraphical analysis to retrieve a
We selected three modern salt-marsh sites (Kréagan, St. Goustan and core (STG6) to reconstruct former sea level after investigations of
Sins marshes) that lie on the east side of Auray and Noyalo estuarine satellite imagery, air photographs, archaeological and topographical
system (Fig. 1). The sites lie in proximity of minor tidal channels and are maps. The stratigraphy suggested that the sediments of the St Goustan
located near to a local tide gauge to provide information regarding tidal were from a high salt-marsh depositional environment.
datums. Kréagan Marsh (Fig. 1a) lies in the upper-middle part of Auray
River Estuary, which is the largest river entering the Morbihan Golfe. The Foraminiferal data
marsh vegetation consists of halophytic plants such as Phragmites spp.
(common reed) and Juncus geradi (saltmeadow rush), which dominate Foraminifera samples of approximately 10 cm3 were collected by
the high marsh. Spartina spp. (saltmarsh cordgrass) is dominant in middle pressing down a hard plastic ring into the surface layer (Scott et al.,
and low-marsh environments with the notable presence of Salicornia 2001). The top 1 cm of sediment was preserved in ethanol and stained

Figure 1. Location map of Morbihan Golfe (Brittany, France), showing the three marsh sites reported in the text. Two salt marshes were sampled along modern transects (black lines)
at Kréagan (a) and Sins (b) sites. A 0.32 m long-core (STG6) was extruded from St Goustan marsh (b).
26 V. Rossi et al. / Quaternary Research 75 (2011) 24–35

with rose Bengal to differentiate specimens that were living at the the coefficient of determination (r2). These statistical coefficients are
time of collection (Walton, 1952; Murray and Bowser, 2000). representative of the predictive abilities of the function (RMSEP) and the
Several authors (Edwards et al., 2004; Horton and Edwards, 2006; strength of the relationship between observed and inferred values (r2).
Leorri et al., 2008a,b; Kemp et al., 2009b) suggest that dead assemblages Both parameters were estimated through cross-validation (“jack-knifing”;
are a better analogue for paleoenvironmental reconstruction compared RMSEPjack and r2jack). Jack-knifing measures are reliable indicators of the
to total or living assemblages, because they are less sensitive to seasonal true predictive ability of the transfer functions as they are less biased by
(Murray, 1991) or post-depositional (Horton and Edwards, 2006) sample re-substitution (Dixon, 1993). During the calibration process to
changes. Thus, we assume that dead, surficial (0–1 cm) assemblages are the fossil core a statistical evaluation of the sample-specific error of
representative of the modern distribution of foraminifera and represent prediction was quantified (10,000 cycles; Birks, 1995).
an appropriate model for RSL reconstructions. We assume the influence A firm understanding of the ecology and taphonomy of salt-marsh
of infaunal foraminifera (e.g., Goldstein and Harben, 1993; Hippensteel foraminifera must remain central to the development and application of
et al., 2000) in the minerogenic marshes of Brittany is negligible. the transfer functions (Horton and Edwards, 2006). However, we also
Although infaunal distributions have been reported from the French employed the modern analogue technique (MAT) to test the reliability
Atlantic coast (Duchemin et al., 2005), analyses from the United of paleo-marsh elevations obtained through the calibration of the
Kingdom (Horton and Edwards, 2006) and Plentzia (Southern Bay of transfer function. Using the chi-square distance dissimilarity coefficient,
Biscay) and Tagus (Portugal) estuaries (unpublished data) suggest that we evaluated the level of analogy between core foraminiferal
infaunal populations do not significantly enrich downcore populations. assemblages and the foraminiferal associations recorded within the
Samples were wet-sieved through sieves of 500 μm and 63 μm to modern samples. The largest minimum dissimilarity coefficient
remove plant macrofossils and fine-grained material, respectively. (MinDc) of the modern training set was used as a threshold to identify
The residual fraction was split into eight subsamples that were core samples with a good modern analogue (Woodroffe, 2009).
observed under a stereoscopic binocular microscope using reflected
light. Identifications of foraminifera were confirmed by comparison Chronology
with type and figured specimens lodged at the Smithsonian
Institution, Washington, D.C. Where possible, a minimum amount of A chronology was developed for the 0.32 m long core (STG6) extruded
250 dead specimens were picked for each modern sample. Species from the St Goustan marsh based upon sedimentation rates estimated by
210
relative abundances are presented in Appendix A. Pb (t1/2 =22.3 years) and 137Cs (t1/2 =30.2 years). Samples were
We followed the same procedure described for the modern analyzed for 137Cs activity by direct gamma counting on a low-
environment samples to prepare and analyze fossil samples (2 cm3 of background, high-efficiency Germanium detector coupled with a multi-
sediment for each sample) collected at 1 cm intervals from the 0.32 m channel analyzer. Samples were packed into standardized vessels and
long core STG6. More than 6000 foraminifers were counted from these were counted for approximately 24 h. Detectors were calibrated using
samples and their relative abundances are reported in Appendix B. natural matrix standards (IAEA-300, 312, 314) at the energy of interest
(661 keV) in the standard counting geometry for the associated detector.
Foraminifera-based transfer function Total 210Pb was measured by alpha spectroscopy following the
methodology of Nittrouer et al. (1979). Approximately 1.5 g of sediment
We developed a transfer function using the modern training dataset was spiked with 209Po, as a yield determinant, and then was partially
from Kréagan and Sins marshes to quantify the relationship between digested with 8 N nitric acid (HNO3) by microwave heating. Polonium-
salt-marsh foraminiferal assemblages and tidal elevation in the 209 and 210Po in solution was then electroplated onto nickel planchets
Morbihan Golfe. Only samples with counts exceeding 250 individuals/ in a dilute acid solution (modified from Flynn, 1968). Excess 210Pb was
10 cm3 were included for statistical analyses (Patterson and Fishbein, determined by subtracting the 210Pb activity supported by 226Ra from
1989; Fatela and Taborda, 2002). To improve the predictive capability of the total 210Pb activity, where the supported 210Pb activity for a given
a transfer function which will be calibrated to fossil foraminiferal core was assumed to be equal to the uniform background activity found
assemblages within salt-marsh sediments of core STG6 we excluded at depth (Nittrouer et al., 1979). A sediment accumulation rate was
tidal flat samples (e.g. Gehrels et al., 2005). To account for variations in calculated using the constant flux–constant sedimentation (CF–CS)
tidal range between the two modern marshes and the fossil study site, model (Appleby and Oldfield, 1992). Downcore 137Cs activities were
sample elevations were converted into Standardized Water Level Index used to substantiate the 210Pb-determined accumulation rates.
(SWLI), following Horton and Edwards (2006):
Results
SWLIn = 1004 ðEn −MLLW Þ = ðMHHW−MLLW Þ ð1Þ
Modern foraminiferal distribution of Morbihan marshes
where SWLIn is the standardized water level index for sample n, En
represents the elevation of sample n (m IGM69), MLLW is the mean We identified 23 dead foraminifera from 42 samples of Kréagan
low low-water elevation (m IGM69) and MHHW is the mean higher and Sins marshes (Appendix A). The average concentration was 4511
high-water elevation (m IGM69). individuals/10 cm3.
We performed detrended canonical correspondence analysis (DCCA) Kréagan Marsh was dominated by agglutinated foraminiferal
to determine the most appropriate response model (unimodal or linear) species, which represent more than 70% of the entire assemblage
for a foraminifera-based transfer function for the training set of Kréagan (Fig. 2a). In the high marsh (2.45–2.20 m IGM69) Jadammina macrescens
and Sins marshes (CANOCO; ter Braak and Šmilauer, 2002). A gradient (21–65%) and Trochammina inflata (7–39%) are the main species in
length of 0.67 SD units suggested that the ecological behavior of association with Arenoparella mexicana, Miliammina fusca and Haplo-
foraminifera is linear with respect to elevation (Birks, 1995). We phragmoides spp. There is also a notable occurrence of Quinqueloculina in
developed a partial least squares (PLS) transfer function using the one station (Station 5; 34%). The change in elevation across the intertidal
computer program C2 (version 1.4.3., Juggins, 2006). PLS is a linear-based zone into the middle (2.32–2.11 m IGM69) and low marsh (2.23–2.12 m
technique widely used for paleoenvironmental reconstructions, which IGM69) corresponds to a decrease in abundance of J. macrescens and T.
maximizes the covariance with the response variable and gives a lower inflata to be replaced by M. fusca and calcareous species such as
prediction error than other linear regressive methods (Stone and Brooks, Elphidium williamsoni. The tidal flat (2.05–1.52 m IGM69) environment
1990). The performance of the transfer function was assessed by the is further dominated by Elphidium and Quinqueloculina genera, although
statistical measures, Root Mean Square Error of Prediction (RMSEP) and there is still a notable agglutinated occurrence.
V. Rossi et al. / Quaternary Research 75 (2011) 24–35
Figure 2. Relative dead abundances of the most common benthic foraminifera taxa from the modern samples collected along the Kréagan (a) and Sins (b) transects. MLHW: mean lower high water; MHHW: mean higher high water. Elevation
along the transects is also reported (m IGM69).

27
28 V. Rossi et al. / Quaternary Research 75 (2011) 24–35

The high (2.02–1.86 m IGM69) and middle (1.92–1.76 m IGM69) We applied the transfer function to the 33 foraminiferal samples from
marsh floral environments of Sins Marsh (Fig. 2b) are characterized by the core to estimate paleo-marsh elevation (with respect to IGM69) with
agglutinated species; no calcareous species were found in any sample-specific errors ranging from 0.08 m to 0.09 m (Fig. 4 and Table 2).
samples. J. macrescens (N40% in all samples, with the exception of The paleo-marsh elevations throughout the core are approximate to local
sample S.II 8 where the abundance of J. macrescens is 30%) in asso- mean high high water at St. Goustan (1.98 m IGM69) and are consistent
ciation with subordinate T. inflata, Haplophragmoides spp. and M. fusca with the lithostratigraphic and biostratigraphic features of the fossil
dominate the assemblages. Further, monospecific assemblages with succession. The highest paleo-marsh elevation (2.06 ± 0.08 m above
low abundance (b120 individuals/10 cm3) of J. macrescens are found IGM69) occurs at 30 cm core depth, where high abundances of high to
at the high marsh/upland boundaries of both transects. middle marsh taxa such as J. macrescens (40%), T. inflata (26%) and H.
manilianensis (22%) are dominant. The lowest elevation (1.83 ±0.08 m
above IGM69) is found at 20 cm core depth where the maximum
Development of a foraminifera-based transfer function
abundance of the low-marsh taxa M. fusca (41%) is recorded.
The reliability of these paleo-marsh elevation estimates is supported
We have developed a Morbihan Golfe foraminifera-based transfer
by the strong similarity existing between the fossil assemblages and the
function from a screened training set of 36 salt-marsh samples and 23
foraminiferal content of the modern training set. Indeed, MAT analysis
species of Kréagan and Sins transects. In addition to the removal samples
shows that 31 of 33 core samples have a good modern analogue
from tidal-flat environments and samples with low abundance (e.g.
(Table 2). The exceptions are samples collected at 27 cm and 29 cm core
monospecific assemblage of J. macrescens at the landward edge of Sins
depth, characterized by very high percentages of H. manilaensis (ca. 70%,
transects I and II), allothonous species (high abundance of Quinquelo-
see Appendix B and Fig. 4). In the modern training set the maximum
culina genera in high marsh of Kréagan) were excluded.
abundance of H. manilaensis is much lower (about 12%, see Appendix A).
The PLS transfer function produced results for five components,
Thus, the paleo-marsh elevations derived from these samples should be
the choice of component is based upon the prediction statistics and
treated with caution.
the principle of parsimony (the lowest component that gives an
acceptable model is usually chosen). Thus, we have chosen compo-
Discussion
nent two of the transfer function because it performs significantly
better than component one when jack-knifed errors are considered:
Intertidal foraminiferal distribution along the European Atlantic coast
prediction errors (RMSEPjack = 0.07 m) were lower and squared
correlations (r2jack = 0.7) were higher (Table 1). Despite the small
The foraminiferal analysis performed on the Kréagan and Sins
gradient length of the training set (Woodroffe, 2009), the transfer
marshes has provided useful information about the spatial distribution
function shows a strong performance between observed and
patterns of modern intertidal foraminifera in the Morbihan Golfe. We
predicted SWLIs (Fig. 3), although the residuals do suggest structure
have identified three main foraminiferal assemblages along the modern
with the model under predicting elevations from the landward limit
transects on the basis of the relative species abundances (Figs. 2a–b). The
of the transects.
high and middle marsh assemblages were dominated by two aggluti-
nated species, J. macrescens and T. inflata, with the secondary occurrence
Core foraminiferal assemblages and calibration of the transfer function of Haplophragmoides spp., M. fusca and A. mexicana. A similar assemblage
has been found within the high marsh of Plougoumelen (Duchemin et
The STG6 core is 0.32 m long and consists of a brown silty peat with al., 2005), located in the western part of the Morbihan Golfe close to the
abundant macrofossils (Fig. 4). No evidence of bioturbation was found in Kréagan site (see Fig. 1), where a noticeable presence of Quinqueloculina
core succession. The surface altitude of the core was 2.03 m IGM69. Nine spp. was also encountered. Jadammina macrescens and T. inflata also co-
agglutinated species of benthic foraminifers were found within core dominated the high and middle marsh environments of the Basque
STG6, analogous to the modern foraminiferal assemblages of the intertidal zones from the southern Bay of Biscay (Cearreta et al., 2002;
Morbihan Golfe, indicating a salt-marsh depositional environment Leorri et al., 2008a,b). A similar high-marsh faunal zone has been
(Appendix B). The assemblages were dominated by Haplophragmoides identified in eastern England (Coles, 1977; Coles and Funnell, 1981;
spp., J. macrescens and M. fusca. From the base of the core to 15 cm core Funnell and Boomer, 1998; Gehrels et al., 2001; Horton and Edwards,
depth, Haplophragmoides spp. (22–72%) and J. macrescens (usually 2006). Murray (1991) distinguished J. macrescens and T. inflata
N20%) are the dominant taxa with variable amounts of M. fusca (0–42%) associations in high marsh environments of Europe based on the work
and T. inflata (0–26%). Very high abundances of J. macrescens (33–51%) of Le Campion (1970), Phleger (1970) and Pujos (1976). Further, Sins
and considerable percentages of Miliammina fusca (13–27%) and Marsh exhibits a monospecific assemblage of J. macrescens at the
Haplophragmoides spp. (19–35%) characterize the interval between landward limit of the high marsh. Similar monospecific faunal
15–9 cm core depth, where relatively low amounts of T. inflata are assemblages have been documented in the United Kingdom (Horton
found. From 9 to 0 cm core depth there was an increase of and Edwards, 2006) and are characteristic of marshes of northeastern
Haplophragmoides spp., which becomes the dominant taxa (33–55%) North America (Scott and Medioli, 1978; Gehrels, 1994; Edwards et al.,
along with J. macrescens (20–40%), paralleled by a decrease of 2004; Kemp et al., 2009a).
Miliammina fusca (usually b10%). Allochthonous foraminiferal content The Kréagan low-marsh assemblages are characterized by relatively
is almost null throughout the core succession. high percentages of M. fusca. Notable counts of Quinqueloculina and
Elphidium species are also reported from the Kréagan site. Leorri et al.
Table 1
(2008a,b) similarly found an increase in the abundances of M. fusca and
Statistics of PLS transfer functions developed using the modern database from two hyaline foraminifers at the transition to the low-marsh environment. In
salt marshes of the Morbihan Golfe. The shaded row corresponds to component 2, the United Kingdom a middle to low-marsh zone around or below mean
selected for calibration purposes. high-water spring tide dominated by M. fusca in association with low
frequencies of calcareous species has been identified (Coles, 1977; Coles
and Funnell, 1981; Horton, 1999; Horton and Edwards, 2006).
Finally, we identified a relatively high-diversity foraminiferal fauna
within the tidal flat of the Kréagan transect, consisting of J. macrescens,
Ammotium salsum and calcareous species (Quinqueloculina spp. and
Haynesina germanica). Similar intertidal studies from Europe also
V. Rossi et al. / Quaternary Research 75 (2011) 24–35 29

Figure 3. Performance of the foraminifera-based transfer function derived from Kréagan and Sins modern database. The relationship between observed and estimated elevation
(SWLI) and residuals versus observed elevation (SWLI) are shown.

identified tidal flat faunal zones with a notable presence of calcareous the relative sea level trend of the Morbihan Golfe region during the
species (Phleger, 1970; Pujos, 1976; Coles, 1977; Coles and Funnell, past ~ 150 years (Fig. 5).
1981; Murray, 1991; Boomer, 1998; Funnell and Boomer, 1998; Horton To test the reliability and the spatial representativity of foraminifera-
and Edwards, 2006). An assemblage dominated by H. germanica and based RSL reconstructions, we choose to compare the salt-marsh records
Ammonia tepida occurred within the surficial samples collected from six against the Brest tide-gauge data obtained from the Permanent Service for
different estuaries in the southern Bay of Biscay (Redois and Debenay, Mean Sea Level. The Brest tide gauge has been shown to be representative
1996; Leorri and Cearreta, 2009). of the Bay of Biscay region (Leorri et al., 2010). Further, the Brest tide-
gauge data have been included in most sea level rise studies (e.g., Douglas,
Establishing a geochronology 2001; Church and White, 2006; Jevrejeva et al., 2008; Leorri et al., 2008b;
Grinsted et al., 2009) due to its completeness and accuracy since AD 1807
The CF:CS modeled downcore 210Pb activity and 137Cs peak provided (Wöppelmann et al., 2006). Wöppelmann et al. (2006) distinguished
an age–depth model for the STG6 core (Fig. 4). There has been recent linear trends of +1.3±0.15 mm yr− 1 (between AD 1890 and AD 1980)
debate about the validity of salt-marsh chronologies derived solely from and +3.0±0.5 mm yr− 1 (between AD 1980 and AD 2004). Using the
210
Pb dating (e.g., Yang et al., 2001) and the use of independent regression analysis, the Morbihan Golfe salt-marsh derived data we obtain
chronological markers such as pollen chrono-horizons, bomb spike 14C comparable values of +1.6±0. 5 mm yr− 1 from AD 1890 to AD 1980 and
and spheroid carbonaceous particles as a means to corroborate age–depth an increase in the rate of rise to +4.7±1.6 mm yr− 1 from AD 1980 to AD
relationships inferred from 210Pb dating has been widely suggested. We 2004. The general agreement between the records of RSL change confirms
used a peak in 137Cs activity (related to the global fallout dated AD 1963) the validity of this methodology and its potential to extend RSL estimates
as an independent time-stratigraphic marker to develop our age–depth back into the pre-instrumental period. Comparable acceleration in sea
model (Robbins and Edgington, 1975; Ritchie and McHenry, 1990). level rise between the 19th and 20th century has been suggested from
other salt-marsh proxy record located along the Atlantic Coasts of North
Reconstruction of former sea levels in the Morbihan Golfe by the application America (Gehrels et al., 2005; Kemp et al., 2009a) and Northwestern
of foraminifera-based transfer function Europe (Leorri et al., 2008b).
However, it is clear from the Morbihan Golfe that the reconstruction
We applied the transfer function to the STG6 fossil sequence to of recent RSL in this region remains a work in progress. Ample scope
reconstruct the paleo-marsh surface elevation with an associated error. exists for improving the reliability, accuracy and precision of the transfer
We obtain an estimate of former sea levels by subtracting the paleo- function reconstructions through the expansion of the modern salt-
marsh elevation of each sample from its present elevation using: marsh dataset. Further, considerable temporal errors influence the
earlier period of the RSL record, reducing the accuracy of past sea level
RSLi = Ei −PMEi ð2Þ estimates from AD 1860 to AD 1950. Developments in transfer functions
will need to be supported by improvements in the dating of sediments
where RSLi is the relative sea level for sample i, Ei and PMEi are the and the methods used to interpolate between dated horizons (van de
elevation and the paleo-marsh elevation value expressed in m IGM69 Plassche et al., 2001; Edwards, 2004; Gehrels et al., 2008; Kemp et al.,
for sample i. 2009b) and the collection of multiple salt-marsh cores to assess the
Whilst paleo-marsh elevations are dependent on the accumulation reproducibility of the sea level reconstructions (Kemp et al., 2009b).
history at the borehole location, RSL estimates should be spatially
representative in the absence of complicating factors such as differential
sediment autocompaction and depositional hiatuses. We believe core Conclusions
STG6 is of compaction-free because bulk density down downcore remains
relatively consistent (Fig. 4; Gehrels et al., 2008). Similarly, no evidences We developed a foraminifera-based transfer function using a
of depositional hiatuses are recorded within the fossil succession. modern training set of 36 samples and 23 species from two salt
Combining the sample-specific error for RSL estimates with the marshes of the Morbihan Golfe area, Brittany, France, to reconstruct
age uncertainties introduced by the core age model, we reconstruct former sea levels during the 20th century.
30
V. Rossi et al. / Quaternary Research 75 (2011) 24–35
Figure 4. Lithology, chronostratigraphy and foraminiferal assemblages of STG6 core. Paleo-marsh elevation estimates produced by foraminifera-based transfer function are also presented with associated errors (estimates from core samples
without a modern analogue are shown by dashed lines). The main benthic foraminifera taxa found within the core samples are reported as relative abundances. The age–depth model developed for the STG6 core was based on the 210Pb and
137
Cs activities. The errors associated with the age–depth model are also indicated.
V. Rossi et al. / Quaternary Research 75 (2011) 24–35 31

Table 2
Downcore paleo-marsh elevations with associated bootstrapped error obtained by the application of foraminifera-based transfer function to 33 core samples. Fossil dissimilarity
coefficient (MinDC) is also reported for each sample. Core samples without a good modern analogue are highlighted. The critical threshold used to distinguish samples without a
good modern analogue is the largest dissimilarity coefficient (62.06) obtained from the modern training set.

The spatial distribution pattern of benthic foraminifers from lived radionuclides (210Pb and 137Cs). The resulting Morbihan sea level
Morbihan salt-marsh environments is mainly controlled by the trend is in general agreement with the Brest tide-gauge record, showing
elevation with respect to the tidal frame. This inference is supported a rate of rise of +1.6 ± 0. 5 mm yr− 1 from AD 1890 to AD 1980 and an
by the strong relationship between observed and predicted values increase to +4.7 ± 1.6 mm yr− 1 from AD 1980 to AD 2004. These data
obtained from the transfer function (r2jack = 0.7). Moreover, the transfer suggest acceleration in sea level rise during the 20th century, already
function suggests that precise reconstructions of former sea level are observed in observational and other salt-marsh proxy records located
possible (RMSEPjack = 0.07 m). In order to reconstruct former sea levels along the Atlantic Coasts of North America and Northwestern Europe.
we applied the transfer function to a short salt-marsh core (0.32 m) and Thus, our data confirm that salt-marsh foraminifera can be used to
placed RSL estimates in a chronological framework using two short- develop reliable estimates of former sea levels.

Figure 5. Relative sea level curve for the Morbihan Golfe area derived by the integration of elevation estimates produced by foraminifera-based transfer function with the STG6 core
age–depth model. Core samples without modern analogue are plotted as filled symbols. Brest tide-gauge sea level curve is also shown for comparison.
32
Appendix A
Relative dead abundances of foraminifera taxa found within the modern samples collected from Kréagan-K. (one transect) and Sins-S.I and S.II (two transects) marshes. The elevation (expressed in m IGM69) of each sample is also reported.

Transect sample K. 1 K. 2 K. 3 K. 4 K. 5 K. 6 K. 7 K. 8 K. 9 K. 10 K. 11 K. 12 K. 13 K. 14 K. 15

TF sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Elevation (m IGM69) 2.45 2.35 2.27 2.28 2.20 2.33 2.35 2.35 2.38 2.37 2.32 2.22 2.18 2.11 2.23
Ammobaculites crassus 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammobaculites sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammutium salsum 0.00 0.00 0.00 0.00 0.00 0.40 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Arenoparella mexicana 18.73 17.82 17.24 10.55 8.87 4.76 9.25 5.13 6.64 5.20 5.43 7.06 8.19 9.57 8.49
Haplophragmoides canariensis 0.60 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Haplophragmoides manilaensis 6.65 7.92 3.88 5.50 3.94 12.30 8.81 8.12 4.27 6.00 1.36 0.39 0.00 0.00 0.00
Haplophragmoides wilberti 2.42 0.99 0.00 0.00 0.00 0.00 0.00 0.00 0.95 0.40 0.90 0.00 0.00 0.00 0.00
Haplophragmoides sp. 1 1.51 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.47 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina fusca 9.06 13.37 12.93 20.64 2.96 5.95 3.96 5.13 4.74 6.00 5.88 7.84 6.47 13.88 14.62
Miliammina petila 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina subrotunda 0.60 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.90 0.00 0.00 0.00 0.00
Tiphotrocha comprimata 0.00 0.00 0.00 1.83 0.49 0.79 0.88 0.00 2.37 1.20 0.00 0.00 0.00 0.00 0.00
Trochammina inflata 38.97 27.72 15.52 12.39 7.39 10.32 10.13 17.52 28.91 34.80 25.79 27.84 13.36 13.40 23.58
Jadammina macrescens 21.45 29.21 46.98 46.79 41.87 64.68 61.67 55.56 43.13 46.00 57.01 48.63 43.53 44.98 48.11
Reophax sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Cornuspira involvens 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.43 0.00 0.00 0.00 0.78 1.29 0.00 0.00

V. Rossi et al. / Quaternary Research 75 (2011) 24–35


Quinqueloculina agglutinans 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.40 0.00 0.00 0.00 0.00 0.00
Quinqueloculina sp. 0.00 2.97 3.02 1.83 34.48 0.79 4.41 8.12 8.53 0.00 2.71 7.45 25.43 14.83 3.77
Ammonia sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 1.72 2.39 0.47
Elphidium incertum 0.00 0.00 0.00 0.00 0.00 0.00 0.44 0.00 0.00 0.00 0.00 0.00 0.00 0.48 0.00
Elphidium excavatum 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium williamsoni 0.00 0.00 0.43 0.00 0.00 0.00 0.44 0.00 0.00 0.00 0.00 0.00 0.00 0.48 0.94
Haynesina germanica 0.00 0.00 0.00 0.46 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Forams counted per 10cc 15888 6464 14848 10464 6496 16128 10896 14976 5064 12000 7072 8160 2784 3344 3392

Appendix A
Relative dead
Appendix abundances of foraminifera taxa found within the modern samples collected from Kréagan-K. (one transect) and Sins-S.I and S.II (two transects) marshes. The elevation (expressed in m IGM69) of each sample is also reported.
A (continued)

Transect sample K. 16 K. 17 K. 18 K. 19 K. 20 K. 21 S.II 1 S.II 2 S.II 3 S.II 4 S.II 5 S.II 6 S.II 7 S.II 8 S.II 9

TF sample 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30

Elevation (m IGM69) 2.13 2.12 2.16 2.05 1.92 1.52 2.02 1.89 1.98 1.94 1.86 1.86 1.87 1.90 1.92
Ammobaculites crassus 0.00 0.48 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammobaculites sp. 0.00 0.00 0.00 0.00 0.63 1.69 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammutium salsum 0.00 1.92 0.00 0.00 3.80 6.21 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Arenoparella mexicana 5.71 0.48 4.09 1.29 1.90 4.52 0.00 0.79 0.00 0.00 0.46 2.23 0.00 2.13 0.76
Haplophragmoides canariensis 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Haplophragmoides manilaensis 1.90 2.40 0.45 2.59 0.00 1.13 3.33 0.39 0.00 0.00 0.00 0.00 0.00 2.13 9.09
Haplophragmoides wilberti 0.48 0.96 0.00 0.00 0.00 0.56 0.00 1.18 0.77 0.00 0.00 0.00 0.00 0.00 2.27
Haplophragmoides sp. 1 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina fusca 33.33 8.65 1.82 6.47 1.90 1.69 0.00 0.39 0.00 0.00 3.23 1.86 4.63 5.11 9.85
Miliammina petila 0.00 0.00 0.00 0.43 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina subrotunda 0.48 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Tiphotrocha comprimata 0.48 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Trochammina inflata 8.10 12.50 33.18 28.88 12.66 1.69 0.00 2.36 15.33 1.34 14.29 44.98 36.11 60.43 27.27
Jadammina macrescens 36.67 22.60 59.55 28.02 22.15 26.55 96.67 94.88 83.91 98.66 82.03 50.93 59.26 30.21 50.76
Reophax sp. 0.00 0.48 0.00 0.00 0.00 0.56 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Cornuspira involvens 0.00 0.00 0.00 2.59 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Quinqueloculina agglutinans 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Quinqueloculina sp. 4.76 0.00 0.91 25.86 28.48 25.42 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammonia sp. 0.00 1.92 0.00 1.72 0.00 0.56 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium incertum 1.90 10.10 0.00 0.86 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium excavatum 0.00 0.96 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium williamsoni 6.19 36.54 0.00 0.43 28.48 25.42 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Haynesina germanica 0.00 0.00 0.00 0.86 0.00 3.95 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Forams counted per 10cc 1440 289 2347 1237 68 139 120 1016 2088 1015 1488 1076 1728 3760 12672
Appendix A
Relative
Appendixdead abundances of foraminifera taxa found within the modern samples collected from Kréagan-K. (one transect) and Sins-S.I and S.II (two transects) marshes. The elevation (expressed in m IGM69) of each sample is also reported.
A (continued)

Transect sample S.I 1 S.I 2 S.I 3 S.I 4 S.I 5 S.I 6 S.I 7 S.I 8 S.I 9 S.I 10 S.I 11 S.I 12

TF sample 31 32 33 34 35 36 37 38 39 40 41 42

Elevation (m IGM69) 2.00 1.96 1.97 1.97 1.94 1.93 1.95 1.86 1.89 1.82 1.88 1.76
Ammobaculites crassus 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammobaculites sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.49
Ammutium salsum 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.38 0.00 0.98
Arenoparella mexicana 0.00 0.00 3.95 5.33 6.63 8.24 1.91 4.96 0.84 1.53 3.35 3.92
Haplophragmoides canariensis 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Haplophragmoides manilaensis 0.00 10.29 0.00 1.78 0.60 1.96 0.96 0.00 0.00 0.76 2.09 4.41
Haplophragmoides wilberti 0.00 19.12 0.00 3.11 0.60 3.53 0.00 0.00 0.00 0.00 0.00 0.49
Haplophragmoides sp. 1 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina fusca 0.00 7.35 17.98 15.11 27.11 9.41 4.31 6.61 8.79 26.34 1.26 6.86
Miliammina petila 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina subrotunda 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Tiphotrocha comprimata 0.00 0.00 0.44 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 3.43
Trochammina inflata 0.00 17.65 32.89 28.44 24.10 24.71 19.62 42.98 1.26 22.14 5.86 13.24
Jadammina macrescens 100.00 45.59 44.74 46.22 40.96 52.16 73.21 45.45 89.12 48.85 87.45 66.18
Reophax sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Cornuspira involvens 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

V. Rossi et al. / Quaternary Research 75 (2011) 24–35


Quinqueloculina agglutinans 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Quinqueloculina sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Ammonia sp. 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium incertum 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium excavatum 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Elphidium williamsoni 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Haynesina germanica 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Forams counted per 10cc 20 136 995 1800 5312 2720 2508 2151 1093 1143 1912 1224

Appendix B
Relative abundances of foraminifera taxa found within the samples collected from the STG6 core. The elevation (expressed in m IGM69) of each core sample is also reported.

Core depth (cm) 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

Elevation core sample (m IGM69) 2.03 2.02 2.01 2.00 1.99 1.98 1.97 1.96 1.95 1.94 1.93 1.92 1.91 1.90 1.89 1.88 1.87
Arenoparella mexicana 11.59 13.17 7.89 7.19 3.05 4.15 7.64 7.77 2.15 2.24 1.23 2.06 1.64 0.45 0.00 0.00 0.42
Haplophragmoides manilaensis 52.17 32.93 28.95 41.32 37.20 40.41 31.85 35.23 37.34 32.84 25.31 16.05 19.67 18.47 25.97 55.74 43.64
Haplophragmoides wilberti 1.45 5.99 7.02 5.99 4.88 5.18 1.27 0.52 3.86 1.49 0.00 0.41 1.64 0.90 4.87 0.55 1.69
Haplophragmoides sp. 1 1.45 0.00 2.63 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Miliammina fusca 7.25 10.78 5.26 5.99 5.49 9.33 5.10 10.88 10.30 13.06 24.07 21.40 22.95 24.77 26.62 7.10 4.66
Miliammina subrotunda 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
Tiphotrocha comprimata 0.00 1.20 0.88 1.80 1.22 1.55 5.10 6.22 15.02 10.82 6.17 0.82 0.00 0.45 1.95 1.09 13.14
Trochammina inflata 5.80 8.98 11.40 9.58 7.93 10.88 7.64 8.29 7.30 6.72 2.47 12.76 3.28 4.95 2.94 7.65 8.90
Jadammina macrescens 20.29 26.95 35.96 28.14 40.24 28.50 41.40 31.09 24.03 32.84 40.74 46.50 50.82 50.00 37.65 27.87 27.54
Total count 69 167 114 167 164 193 157 193 204 268 162 243 61 222 308 183 236

33
34 V. Rossi et al. / Quaternary Research 75 (2011) 24–35

Acknowledgments

34.78

17.39

34.78
1.71
4.35

0.00
0.00

0.00
8.70
0.00
32

46
We thank the National Science Foundation award SGER 0713332.
This paper is a contribution to IGCP Project 588, “Preparing for Coastal

1.72
2.82

0.00
0.00

0.00
0.00
25.35

21.13

18.31
32.39
Change” and INQUA 1001: “Quaternary Coastal Change and Records of
31

71
Extreme Marine Inundation on Coastal Environments.”

22.26

25.81
40.00
1.73
5.16

0.65
0.00
4.84
0.00
1.29

310
30

References
Appleby, P.G., Oldfield, F., 1992. Application of Pb-210 to sedimentation studies, In:
69.57

26.09
1.74
0.00

0.00
0.00
0.00
0.00
2.17
2.17
Ivanovich, M., Harmon, R.S. (Eds.), Uranium-Series Disequilibrium: Applications to
29

46
Earth, Marine, and Environmental Problems, second ed. Claredon Press, Oxford, UK,
pp. 731–778.
Astill, G.G., Langdon, J., 1997. Medieval Farming and Technology: The Impact of
1.75
2.86

0.95
0.00

0.00
5.71
28.57

14.29

14.29
33.33
Agricultural Change in Northwest Europe. Brill, Boston. 321 pp.
Billaud, J.P., 1984. Marais Poitevin, Rencontres de la terre et de l'eau. L'Harmanttan, Paris.
28

84
Birks, H.J.B., 1995. Quantitative palaeoenvironmental reconstructions. In: Maddy, D.,
Brew, J.S. (Eds.), Statistical Modeling of Quaternary Science Data. Technical Guide 5.
Quaternary Research Association, Cambridge, pp. 161–254.
71.74
1.76
4.35

0.43
0.00
9.13
0.00
2.61
7.39
4.35
of foraminifera taxa found within the samples collected from the STG6 core. The elevation (expressed in m IGM69) of each core sample is also reported.

Boomer, I., 1998. The relationship between meiofauna (Ostracoda, Foraminifera) and tidal
230
27

levels is modern intertidal environments of North Norfolk: a tool for palaeoenviron-


mental reconstruction. Bulletin of the Geological Society of Norfolk 46, 17–29.
Brigand, L., Bioret, F., Le Démezet, M., 1992. Landscapes and environments on the island
41.67

15.48

40.48
1.77
0.00

2.38
0.00

0.00
0.00
0.00

of Ouessant, Brittany, France: from traditional maintenance to the management of


105

abandoned areas. Environmental Management 16, 613–618.


26

Cearreta, A., Irabien, M.J., Ulibarri, I., Yusta, I., Croudace, I.W., Cundy, A.B., 2002. Recent
salt marsh development and natural regeneration of reclaimed areas in the Plentzia
estuary, N. Spain. Estuarine. Coastal and Shelf Science 54, 863–886.
1.78
9.96

0.40
0.00
4.78
0.00
0.40
27.09

23.11
34.26

Church, J.A., White, N.J., 2006. A 20th century acceleration in global sea level rise.
251

Geophysical Research Letters 33, L01602. doi:10.1029/2005GL024826.


25

Coles, B.P.L., 1977. The Holocene foraminifera and palaeogeography of central


Broadland. Unpublished Ph.D. Dissertation, University of East Anglia.
Coles, B.P.L., Funnell, B.M., 1981. Holocene paleoenvironments of Broadland, England.
1.79
5.75

0.57
0.00
5.75
0.00
0.00
43.10

18.39
26.44

Special Publication of the International Association of Sedimentologists 5, 123–131.


174
24

Dixon, P.M., 1993. The bootstrap and the jack-knife: describing the precision of
ecological niches. In: Scheiner, S.M., Gurevitch, J. (Eds.), Design and Analysis of
Ecological Experiments. Chapman and Hall, New York, pp. 290–318.
1.80
4.21

5.90
0.00
4.49
0.00
1.97
50.84

15.45
17.13

Donnelly, J.P., Cleary, P., Newby, P., Ettinger, R., 2004. Coupling instrumental and
356

geological records of sea level change: evidence from southern New England of an
23

increase in the rate of sea level rise in the late 19th century. Geophysical Research
Letters 31, L05203. doi:10.1029/2003GL018933.
Douglas, B.C., 1991. Global sea level rise. Journal of Geophysical Research 96, 6981–6992.
1.81
0.00

1.92
0.00

0.00
0.00
0.00
43.27

22.12

32.69

Douglas, B.C., 1997. Global sea (level) rise: a redetermination. Surveys in Geophysics 18,
104
22

279–292.
Douglas, B.C., 2001. Sea level change in the era of the recording tide gauge. In: Douglas,
B.C., Kearney, M.S., Leatherman, S.P. (Eds.), Sea Level Rise; History and
Consequences. Academic Press, San Diego, pp. 37–64.
1.82
0.00

6.38
0.00

0.71
5.67
3.55
39.72

24.11

19.86

Douglas, B.C., 2008. Concerning Evidence for Fingerprints of Glacial Melting. Journal of
141
21

Coastal Research 24, 218–227.


Duchemin, G., Jorissen, F.J., Redois, F., Debaney, J.P., 2005. Foraminiferal microhabitats
in a high marsh: consequences for reconstructing past sea levels. Palaeogeography,
1.83
0.00

2.44
0.00

0.20
0.81
1.83
33.33

41.46

19.92

Palaeoclimatology, Palaeoecology 226, 167–185.


492

Edwards, R.J., 2004. Constructing chronologies of sea level change from salt-marsh
20

sediments. In: Buck, C.E., Millard, A.R. (Eds.), Tools for constructing Chronologies:
Crossing Discipline Boundaries. Springer Verlag, London, pp. 189–211.
Edwards, R.J., van de Plassche, O., Gehrels, W.R., Wright, A.J., 2004. Assessing sea level
1.84
4.76

4.08
0.00

2.04
4.76
7.48
25.85

29.93

21.09

data from Connecticut, USA, using a foraminiferal transfer function for tide level.
147
19

Marine Micropaleontology 51, 239–255.


Engelhart, S.E., Horton, B.P., Douglas, B.C., Peltier, W.R., Tornqvist, T.E., 2009. Spatial
Variability of Late Holocene and 20th Century Sea Level Rise along the US Atlantic
Coast. Geology 37, 1115–1118.
1.85
7.29

3.13
0.00
8.85
1.56
4.17
9.90
39.58

25.52

Fatela, F., Taborda, R., 2002. Confidence limits of species proportions in microfossil
192
18

assemblages. Marine Micropaleontology 45, 169–174.


Flynn, W.W., 1968. The determination of low levels of polonium-210 in environmental
materials. Analytica Chimica Acta 43, 221–227.
44.00

10.50

28.50
1.86
4.00

3.00
0.00

0.50
2.00
7.50

Funnell, B.M., Boomer, I., 1998. Microbiofacies tidal-level and age deduction in
200

Holocene saltmarsh deposits on the North Norfolk Coast. Bulletin of the Geological
17

Society of Norfolk 46, 31–55.


Gehrels, W.R., 1994. Determining relative sea level change from saltmarsh foraminifera
and plant zones on the coast of Maine, USA. Journal of Coastal Research 10, 990–1009.
Elevation core sample (m IGM69)

Gehrels, W.R., 2007. Sea level studies: microfossil reconstructions. In: Elias, S. (Ed.),
Haplophragmoides manilaensis

Encyclopedia of Quaternary Sciences. Elsevier, pp. 3015–3023.


Gehrels, W.R., Roe, H.M., Charman, D.J., 2001. Foraminifera, testate amoebae and
Haplophragmoides wilberti

diatoms as sea level indicators in UK salt marshes: a quantitative multi-proxy


Tiphotrocha comprimata
Miliammina subrotunda
Haplophragmoides sp. 1

Jadammina macrescens
Arenoparella mexicana

approach. Journal of Quaternary Science 16, 201–220.


B (continued)

Trochammina inflata

Gehrels, W.R., Belknap, D.F., Black, S., Newnham, R.M., 2002. Rapid sea level rise in the
Relative abundances

Miliammina fusca
Core depth (cm)

Gulf of Maine, USA, since AD 1800. Holocene 12, 383–389.


Gehrels, W.R., Kirby, J.R., Prokoph, A., Newnham, R.M., Achterberg, E.P., Evans, E.H.,
Total count

Black, S., Scott, D.B., 2005. Onset of recent rapid sea level rise in the western Atlantic
Appendix B

Ocean. Quaternary Science Reviews 24, 2083–2100.


Appendix

Gehrels, W.R., Marshall, W.A., Gehrels, M.J., Larsen, G., Kirby, J.R., Eiriksson, J.,
Heinemeier, J., Shimmield, T., 2006. Rapid sea level rise in the North Atlantic
Ocean since the first half of the 19th century. Holocene 16, 948–964.
V. Rossi et al. / Quaternary Research 75 (2011) 24–35 35

Gehrels, W.R., Hayward, B.W., Newnham, R.M., Southall, K.E., 2008. A 20th century sea Patterson, R.T., Fishbein, E., 1989. Re-examination of the statistical methods used to
level acceleration in New Zealand. Geophysical Research Letters 35, L02717. determine the number of point counts needed for micropaleontological quantita-
Goldstein, S.T., Harben, E.B., 1993. Taphofacies implications of infaunal foraminiferal tive research. Journal of Paleontology 63, 245–248.
assemblages in a Georgia saltmarsh, Sapelo Island. Micropaleontology 39, 55–62. Peltier, W.R., 1996. Mantle viscosity and ice-age ice sheet topography. Science 273,
Grinsted, A., Moore, J.C., Jevrejeva, S., 2009. Reconstructing sea level from paleo and 1359–1364.
projected temperatures 200 to 2100 AD. Climate Dynamics 34, 461–472. Peltier, W.R., 2001. Global glacial isostatic adjustment and modern instrumental
Hawkes, A.D., Horton, B.P., Nelson, A.R., Hill, D.F., 2010. The application of intertidal records of relative sea level history. In: Douglas, B.C., Kearney, M.S., Leatherman,
foraminifera to reconstruct coastal subsidence during the giant Cascadia earth- S.P. (Eds.), Sea Level Rise: History and Consequences. Academic Press, San Diego,
quake of AD 1700 in Oregon, USA. Quaternary International 221, 116–140. California, pp. 65–93.
Henocque, Y., 2003. Development of process indicators for coastal zone management Phleger, F.B., 1970. Foraminiferal populations and marine marsh processes:. Limnology
assessment in France. Ocean and Coastal Management 46, 363–379. and Oceanography 15, 522–534.
Hippensteel, S.P., Martin, R.E., Nikitina, D., Pizzuto, J.E., 2000. The formation of Holocene Pujos, M., 1976. Ecologie des foraminifères benthiques et des thécamoebiens de la
marsh foraminiferal assemblages, middle Atlantic coast, USA: implications for the Gironde et au plateau continental Sud-Gascogne. Application à la connaissaance du
Holocene sea level change. Journal of Foraminiferal Research 30, 272–293. Quaternaire terminal de la région Ouest-Gironde. Memoires de l'Institut de
Horton, B.P., 1999. The contemporary distribution of intertidal foraminifera of Cowpen Géologie du Bassin d'Aquitaine 8, 1–274.
Marsh, Tees Estuary, UK: implications for studies of Holocene sea level changes. Redois, F., Debenay, J.P., 1996. Influence du confinement sur la répartition des
Palaeogeography, Palaeoclimatology, Palaeoecology 149, 127–149 Special Issue. foraminifères benthiques: exemple del'estran d'une ria mésotidale de Bretagne
Horton, B.P., Edwards, R.J., 2006. Quantifying Holocene sea level change using intertidal méridionale. Revue de Paléobiologie 15, 243–260.
foraminifera: lessons from the British Isles. Cushman Foundation for Foraminiferal Ritchie, J.C., McHenry, J.R., 1990. Application of radioactive fallout Cesium-137 for
Research Special Publication 40. measuring soil erosion and sediment accumulation rates and patterns: a review.
Horton, B.P., Edwards, R.J., Lloyd, J.M., 1999. A foraminiferal-based transfer function: Journal of Environmental Quality 19, 215–233.
implications for sea level studies. Journal of Foraminiferal Research 29, 117–129. Robbins, J.A., Edgington, D.N., 1975. Determination of recent sedimentation rates in
Jevrejeva, S., Moore, J.C., Grinsted, A., Woodworth, P.L., 2008. Recent global sea level Lake Michigan using Pb-2 10. Geochimica et Cosmochimica Acta 39, 285–304.
acceleration started over 200 years ago? Geophysical Research Letters 35, L08715. Scott, D.B., Medioli, F.S., 1978. Vertical zonation of marsh foraminifera as accurate
doi:10.1029/2008GL033611. indicators of former sea levels. Nature 272, 528–531.
Juggins, S., 2006. C2, Version 1.4.3: Newcastle University, UK. http://www.campus.ncl. Scott, D.B., Medioli, F.S., Schafer, C.T., 2001. Monitoring in Coastal Environments using
ac.uk/staff/Stephen.Juggins/index.html2006. Foraminifera and Thecamoebian Indicators. Cambridge University Press, Cam-
Kemp, A.C., Horton, B.P., Culver, S.J., 2009a. Distribution of modern salt-marsh bridge. 177 pp.
foraminifera in the Albemarle – Pamlico Estuarine System of North Carolina, Shennan, I., Horton, B.P., 2002. Relative sea level changes and crustal movements of the
USA: Implications for sea level research. Marine Micropaleontology 72, 222–238. UK. Journal of Quaternary Science 16, 511–526.
Kemp, A.C., Horton, B.P., Culver, S.J., Corbett, D.R., van de Plassche, O., Gehrels, W.R., Stone, M., Brooks, R., 1990. Continuum regression: cross-validated sequentially
Douglas, B.C., 2009b. The timing and magnitude of recent accelerated sea level rise constructed prediction embracing ordinary least squares, partial least squares,
(North Carolina, USA). Geology 37, 1035–1038. and principal components regression. Journal of the Royal Statistical Society 52,
Le Campion, J., 1970. Contribution à l'étude des foraminiféres du Bassin d'Arcachon et 237–269.
du proche océan. Bulletin de l'Institute Géologique du Bassin Aquitaine 8, 3–98. ter Braak, C.J.F., Šmilauer, P., 2002. CANOCO Reference Manual and User's Guide to
Leorri, E., Cearreta, A., 2009. Quantitative assessment of the salinity gradient within the CANOCO for Windows: Software for Canonical Community Ordination, Microcom-
estuarine systems in the southern Bay of Biscay using benthic foraminifera. puter Power, Ithaca, NY, Version 4.5.
Continental Shelf Research 29, 1226–1239. van de Plassche, O., Edwards, R.J., van der Borg, K., De Jong, A.F.M., 2001. 14C wiggle-
Leorri, E., Cearreta, A., Horton, B.P., 2008a. A foraminifera-based transfer function as a match dating in high-resolution sea level research. Radiocarbon 43, 391–402.
tool for sea level reconstructions in the southern Bay of Biscay. Geobios 41, van de Plassche, O., van der Schrier, G., Weber, S.L., Gehrels, W.R., Wright, A.J., 2003. Sea
787–797. level variability in the northwest Atlantic during the past 1500 years: a delayed
Leorri, E., Horton, B.P., Cearreta, A., 2008b. Development of a foraminifera-based response to solar forcing? Geophysical Research Letters 30, 1921.
transfer function in the Basque marshes. N. Spain: implications for sea level studies Vincente, I., Leorri, E., Cearreta, A., Gehrels, R., Horton, B.P., 2008. Salt marsh response to
in the Bay of Biscay. Marine Geology 251, 60–74. recent sea level acceleration in the southern Bay of Biscay. Geo-Temas 10, 659–662.
Leorri, E., Gehrels, W.R., Horton, B.P., Fatela, F., Cearreta, A., 2010. Distribution of Walton, W.R., 1952. Techniques for recognition of living foraminifera. Journal of
foraminifera in salt marshes along the Atlantic coast of SW Europe: tools to Foraminiferal Research Special Publication 3, 56–60.
reconstruct past sea level variations. Quaternary International 221, 104–115. Woodroffe, S.A., 2009. Recognising subtidal foraminiferal assemblages: implications for
Marcos, F., Janin, J.M., Le Saux, J.M., 1995. Modélisation hydrodynamique du golfe du quantitative sea level reconstructions using a foraminifera-based transfer function.
Morbihan. Note technique. Conseil Général du Morbihan-EDF. . 47 pp. Journal of Quaternary Science 24, 215–223.
Murray, J.W., 1991. Ecology and Paleoecology of Benthic Foraminifera. Longman, Woodworth, P.L., White, N.J., Jevrejeva, S., Holgate, S.J., Church, J.A., Gehrels, W.R., 2008.
Harlow. 576 pp. Evidence for the accelerations of sea level on multi-decade and century timescales.
Murray, J.W., Bowser, S.S., 2000. Mortality, protoplasm decay rate, and reliability of International Journal of Climatology 29, 777–789.
staining techniques to recognize “living” foraminifera: a review. Journal of Wöppelmann, G., Pouvreau, N., Simon, B., 2006. Brest sea level record: a time series
Foraminiferal Research 30, 66–70. construction back to the early eighteenth century. Ocean Dynamics 56, 487–497.
Nittrouer, C.A., Sternberg, R.W., Carpenter, R., Bennett, J.T., 1979. The use of Pb-210 Yang, H., Rose, N.L., Battarbee, R.W., 2001. Dating of recent catchment peats using
geochronology as a sedimentological tool: application to the Washington spheroid carbonaceous particle (SCP) concentration profiles with particular
Continental Shelf. Marine Geology 31, 297–316. reference to Lochnagar, Scotland. Holocene 11, 593–597.

You might also like