16
ORIGINAL PAPER Quantitative climate reconstruction linking meteorological, limnological and XRF core scanner datasets: the Lake Sanabria case study, NW Spain S. Giralt M. T. Rico-Herrero J. C. Vega B. L. Valero-Garce ´s Received: 22 September 2010 / Accepted: 18 February 2011 / Published online: 1 March 2011 Ó Springer Science+Business Media B.V. 2011 Abstract Monthly limnological monitoring in Lake Sanabria (Spain) since 1986 provided a unique oppor- tunity to test relationships among climate, hydrology and lake dynamics and how they are recorded in the lake sediments. Four datasets were employed: (1) meteorological (monthly maximum and minimum air temperature and total precipitation), (2) limnological (Secchi disk, water temperature, conductivity, pH, dissolved oxygen, nitrate, silicon, total and reactive phosphorus, and total chlorophylls and chlorophyll a), (3) hydrological (Tera River water input and output), and (4) XRF core scanner measurements carried out in short cores. Linear models between the different dataset variables allowed us to characterize the climate signal transmission from one to the other and cross- correlation analyses permitted us to identify the different response times (if any) between them. Principal Component Analyses (PCA) of the limno- logical and geochemical datasets allowed us to identify the main processes that link lake dynamics, primarily nutrient supply and organic productivity, with some sedimentological processes, e.g. organic matter and phosphorus accumulation. Sediment chronology was established by gamma spectrometry ( 210 Pb). Water input to Lake Sanabria is controlled mostly by the Tera River input and is linked directly to precipitation. Response of the Lake Sanabria water budget to climate oscillations is immediate, as the strongest correlation between these two datasets occurs with no lag time. PCA of the limnological dataset indicated that most of the variance is related to nutrient input, and compar- ison with the Tera River water discharge shows that nutrient input was controlled mainly by oscillations in the hydrological balance. The lag time between the hydrological and limnological datasets is 1 month. The PCA of the XRF core scanner dataset showed that the principal process that controls the chemical com- position of the Lake Sanabria sediments is related to sediment and nutrient delivery from the Tera River and organic productivity. Comparison of the nutrient input reconstructed using the limnological dataset and the XRF core scanner data indicated that the sediments act as a low-pass filter, smoothing the climate signal. It was, however, possible to establish the link between these datasets, and obtain a quantitative reconstruction of precipitation for the 1959–2005 AD period that captures the regional variability. This quantitative S. Giralt (&) Institute of Earth Sciences Jaume Almera (ICTJA-CSIC), Lluı ´s Sole ´ i Sabarı ´s s/n, 08028 Barcelona, Spain e-mail: [email protected] M. T. Rico-Herrero B. L. Valero-Garce ´s Pyrenean Institute of Ecology (IPE-CSIC), Av. Montan ˜ana 1005, 50059 Zaragoza, Spain e-mail: [email protected] B. L. Valero-Garce ´s e-mail: [email protected] J. C. Vega Laboratory of Limnology of the Lago de Sanabria Natural Park, Eras 1, 49300 Puebla de Sanabria, Zamora, Spain e-mail: [email protected] 123 J Paleolimnol (2011) 46:487–502 DOI 10.1007/s10933-011-9509-x

Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

ORIGINAL PAPER

Quantitative climate reconstruction linking meteorological,limnological and XRF core scanner datasets: the LakeSanabria case study, NW Spain

S. Giralt • M. T. Rico-Herrero • J. C. Vega •

B. L. Valero-Garces

Received: 22 September 2010 / Accepted: 18 February 2011 / Published online: 1 March 2011

� Springer Science+Business Media B.V. 2011

Abstract Monthly limnological monitoring in Lake

Sanabria (Spain) since 1986 provided a unique oppor-

tunity to test relationships among climate, hydrology

and lake dynamics and how they are recorded in the

lake sediments. Four datasets were employed: (1)

meteorological (monthly maximum and minimum air

temperature and total precipitation), (2) limnological

(Secchi disk, water temperature, conductivity, pH,

dissolved oxygen, nitrate, silicon, total and reactive

phosphorus, and total chlorophylls and chlorophyll a),

(3) hydrological (Tera River water input and output),

and (4) XRF core scanner measurements carried out in

short cores. Linear models between the different

dataset variables allowed us to characterize the climate

signal transmission from one to the other and cross-

correlation analyses permitted us to identify the

different response times (if any) between them.

Principal Component Analyses (PCA) of the limno-

logical and geochemical datasets allowed us to identify

the main processes that link lake dynamics, primarily

nutrient supply and organic productivity, with some

sedimentological processes, e.g. organic matter and

phosphorus accumulation. Sediment chronology was

established by gamma spectrometry (210Pb). Water

input to Lake Sanabria is controlled mostly by the Tera

River input and is linked directly to precipitation.

Response of the Lake Sanabria water budget to climate

oscillations is immediate, as the strongest correlation

between these two datasets occurs with no lag time.

PCA of the limnological dataset indicated that most of

the variance is related to nutrient input, and compar-

ison with the Tera River water discharge shows that

nutrient input was controlled mainly by oscillations in

the hydrological balance. The lag time between the

hydrological and limnological datasets is 1 month.

The PCA of the XRF core scanner dataset showed that

the principal process that controls the chemical com-

position of the Lake Sanabria sediments is related to

sediment and nutrient delivery from the Tera River and

organic productivity. Comparison of the nutrient input

reconstructed using the limnological dataset and the

XRF core scanner data indicated that the sediments act

as a low-pass filter, smoothing the climate signal. It

was, however, possible to establish the link between

these datasets, and obtain a quantitative reconstruction

of precipitation for the 1959–2005 AD period that

captures the regional variability. This quantitative

S. Giralt (&)

Institute of Earth Sciences Jaume Almera (ICTJA-CSIC),

Lluıs Sole i Sabarıs s/n, 08028 Barcelona, Spain

e-mail: [email protected]

M. T. Rico-Herrero � B. L. Valero-Garces

Pyrenean Institute of Ecology (IPE-CSIC), Av.

Montanana 1005, 50059 Zaragoza, Spain

e-mail: [email protected]

B. L. Valero-Garces

e-mail: [email protected]

J. C. Vega

Laboratory of Limnology of the Lago de Sanabria Natural

Park, Eras 1, 49300 Puebla de Sanabria, Zamora, Spain

e-mail: [email protected]

123

J Paleolimnol (2011) 46:487–502

DOI 10.1007/s10933-011-9509-x

Page 2: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

precipitation reconstruction suggests it is possible to

obtain accurate climate reconstructions using non-

laminated sediments.

Keywords Quantitative precipitation

reconstruction � Lacustrine sediments � Iberian

Peninsula � Statistical modeling � Multiproxy

approach

Introduction

The sediment record of lakes is one of the best

continental sensors for qualitative reconstruction of

past environmental and climate changes. Using a

variety of techniques and proxies, recent lacustrine

sediments have been used to quantitatively infer, with

high-temporal resolution, past precipitation (Kalugin

et al. 2007; Nichols et al. 2009; Tonello et al. 2009)

and temperature (Pla and Catalan 2005; Francis et al.

2006; Blass et al. 2007; Larocque and Finsinger

2008), as well as to qualitatively derive land manage-

ment (Hyodo et al. 2008; Djamali et al. 2009;

Striewski et al. 2009) and lake environmental or

limnological conditions such as pH (Schwalb and

Dean 2002; Battarbee et al. 2005), salinity (Chen et al.

2010), or quantify nutrient concentrations (Bennion

et al. 2005; Bigler and Hall 2002). The reliability of

paleoenvironmental reconstructions improves if they

are validated with instrumental data (Fritz 1990;

Kattel et al. 2008). Although quantitative climate and

environmental reconstructions constitute essential

robust data for Global Climate Models (GCM), few

quantitative high-resolution records are available

(Trachsel et al. 2008). Scarcity of such records is

due, in part, to the lack of good, well-calibrated proxy

datasets of high resolution and spatial distribution, as

well as a lack of long-term ([10 years) lake monitor-

ing data to validate environmental reconstructions. In

most regions there are few long, reliable limnological

and/or meteorological datasets, which precludes con-

version of qualitative climate reconstructions into

quantitative ones.

The Iberian Peninsula (IP) is an excellent site to

conduct quantitative climate reconstructions owing to

its location between the Eurosiberian and Mediterra-

nean biogeographic/climatic regions (Carrion et al.

2010). Due to its geographic position, climate across

most of the IP is controlled by the North Atlantic

Oscillation (NAO), the position of the Azores High

and the Intertropical Convergence Zone (ITCZ),

though other climate modes such as the El Nino

Southern Oscillation (ENSO) also exert some influ-

ence (Rodo et al. 1997; Cacho et al. 2001; Romero-

Viana et al. 2008; Martın-Puertas et al. 2008).

There have been several attempts to reconstruct

climate variability using lake sediment records, with

variable degrees of success (Battarbee et al. 2002,

2005; Catalan et al. 2009; Pla and Catalan 2005;

Trachsel et al. 2008; Tonello et al. 2009). Only a few

tried to calibrate the sediment record using instru-

mental or monitoring data. Recently, the link between

measured physical–chemical variables and sedi-

ments has been established using marine sediments

(Abrantes et al. 2009).

In Spain, Lakes Sanabria and Estany Redon have

some of the longest monitoring records. Physical and

biological variables in the lakes were measured

monthly since 1986 (23 years) and 1984 (25 years),

respectively (De Hoyos 1996, http://www.ceab.csic.

es/obser1.htm). In the Spanish Central Range there

are also monthly data for Penalara Lake (Sierra de

Guadarrama, Madrid) since 1995 (14 years) and

several years between 1989–1993 (Toro et al. 2006)

and Cimera Lake (Sierra de Gredos, Avila) for the

1996–1999 and 2006–2008 periods (Granados and

Toro 2000). The aim of this study was to explore how

the climate signal is transferred from the atmosphere

to the lake, and ultimately to the sediments. Estab-

lishment of such links allowed us to infer quantita-

tively the pattern of precipitation over 46 years

(1959–2005 AD) using the chemical composition of

recent sediments in Sanabria Lake.

Site description

Lake Sanabria (Zamora) is located on the northwestern

Iberian Peninsula (IP) (42�070 N–06�430 W), at

1,000 m a.s.l., within a protected area called ‘‘Parque

Natural del Lago de Sanabria y Alrededores’’

(Fig. 1). The basement of the catchment consists of

granitic rocks (gneiss and granodiorite) that origi-

nated during the Variscan Orogeny (Martınez-Garcıa

1973; Vega and Aldasoro 1994), and of Quaternary

deposits of glacial origin. The lake is located in a

valley bounded by a terminal moraine that was

deposited as ice retreated at the end of the Last

488 J Paleolimnol (2011) 46:487–502

123

Page 3: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

Glacial Maximum (Vega et al. 1991; Cowton et al.

2009). The long axis of the lake runs W–E, coincid-

ing with the longitudinal direction of the pre-glacial

fluvial valley. From the climate perspective, the lake

lies on the boundary between the relatively maritime

north coast of Iberia and the arid central Meseta

altiplano. Monthly surveys since 1986 indicate that

river discharge, lake productivity and phytoplankton

succession are greatly influenced by the North

Atlantic Oscillation (NAO), highlighting the strong

Atlantic influence on the entire area (De Hoyos 1996;

Luque and Julia 2002).

Lake Sanabria is the largest glacial lake (368 ha) in

the IP (Aldasoro et al. 1991; Vega et al. 1992; De

Hoyos 1996). The lake is approximately 3,160 m long

by 1,530 m wide and has a volume of 96 hm3 as

determined by topographic survey (Vega et al. 2005).

The lake bathymetry displays two subbasins. The

western subbasin has a maximum depth of 46 m,

whereas the eastern subbasin has a zmax of 51 m. The

topographic high that separates the subbasins is

located 20 m below the water surface. Transverse

sections show a typical glacial morphology, with steep

north and south basin margins and a flat bottom. The

east and west margins show gentle gradients. The lake

lies in an exorheic basin (127.3 km2). The Tera River

is the main source of water and sediment input to the

lake. The hydrologic contribution of this river to Lake

Sanabria ranges between 60 and 100% (mean 85%) of

the total water input, depending on the month and year.

Upstream there are small dams used to produce

electric power, to regulate flow, and provide water to

the lake during summer. Nevertheless, their capacity is

limited and they are repeatedly emptied within the

year, depending on water demands downstream. They

discharge through an overflow channel during strong

storms. Their influence on the hydrology of the Tera

River can therefore, be considered negligible. A

hydroelectric power plant on the Tera River is located

at Moncabril, approximately 1.5 km upstream of Lake

Sanabria (Fig. 1). Furthermore, there are small, semi-

permanent input streams located around the lake.

Fig. 1 a Location of Lake Sanabria on the Iberian Peninsula.

b Bathymetric map of the lake and location of the core sites

and of the monthly limnological sampling locations (red filleddots). The bathymetric contour interval is 1.70 m (modified

from Vega et al. 2005). The locations of the old village of

Ribadelago, the new Ribadelago de Franco village constructed

after the 1959 AD flood that destroyed the old village, and the

Moncabril waste treatment plant are also shown. The distance

between the waste treatment plant and the new Ribadelago de

Franco village is about 1.5 km

J Paleolimnol (2011) 46:487–502 489

123

Page 4: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

The most important one is Seoane Creek, located in

the NW, and those found along the south shore of the

lake. These small streams and creeks are mostly active

during autumn, when most of the annual precipitation

in this area occurs. Water input into the lake is strongly

linked to precipitation in the watershed and limnolog-

ical studies proved that the lake is highly sensitive to

the rainfall regime (De Hoyos 1996). Lake Sanabria is

oligotrophic and warm monomictic, remaining ther-

mally stratified from March/April to mid November.

The lake water has pH values between 6 and 7.3,

conductivity is *13 lS/cm and Total Dissolved

Solids (TDS) are between 7.5 and 13 mg/l (De Hoyos

1996).

Lake Sanabria experienced a catastrophic event

when the dam on the Vega de Tera Reservoir, located

upstream of the Tera River, failed on the night of 9

January 1959. Intense rainfall and poor dam con-

struction provoked the collapse and a flash flood that

killed 144 inhabitants of Ribadelago, a small village

located between the lake and the dam. Rapid input of

about 8 9 103 m3 of water, sediment and debris into

the lake, raised the lake water level about 2.6 m and

left a clear signal (a clastic layer) in the sediment

record (Luque 2003). The village was reconstructed

downstream of the Tera River and it is known as

Ribadelago de Franco (Fig. 1).

Materials and methods

Sediment coring campaigns were carried out in May

2004 and June 2007. We retrieved 5 Kullenberg

cores, up to 8.9 m long, and 13 short gravity cores,

from 29 to 69 cm long, in the two subbasins. All

cores except one were immediately sealed and stored

in a cold room at ?4�C until they were opened for

sampling in the laboratory. One short core (SAN07-

2 M) was extruded and sampled in the field. The

other short gravity cores were transported and stored

vertically to prevent disturbance of the sediment-

water interface. This paper focuses on two short

gravity cores, SAN07-1 M (33 cm) and SAN07-2 M

(64 cm), collected in 2007 from the deepest part of

the east subbasin (Fig. 1).

Magnetic susceptibility and lightness were mea-

sured every 0.5 cm using a GEOTEK multi-sensor

core logger at the Limnological Research Center

(LRC), University of Minnesota (USA). Grain size

was measured on SAN07-2 M at 1-cm intervals with

a LS-13-320 Beckman Coulter Counter in the

Department of Stratigraphy, Paleontology and Marine

Geosciences, University of Barcelona (Spain). X-Ray

Fluorescence (XRF) core scanner data from SAN07-

1 M allowed us to do a high-resolution characteriza-

tion of the chemical composition of the uppermost

sediment infill. XRF measurements were carried out

every mm for the uppermost 13 cm using the ITRAX

XRF core scanner of the Large Lake Observatory

(USA). The XRF settings were 45 kV, 30 mA and

30 s of time exposure with Mo and Cr tubes to obtain

statistically significant results for the maximum

number of chemical elements possible. Results are

expressed as element intensities in counts per second

(cps).

The Total Carbon (TC) content of the two studied

cores was determined every cm using a LECO SC-

144DR available at the Pyrenean Institute of Ecology

(IPE-CSIC), Spain. Results are expressed in percent

TC with respect to the total sample weight.

The SAN07-2 M core was sliced into cm sections

in the field immediately after recovery, and samples

were stored in sterilized, pre-weighed containers. The210Pb activity was measured by gamma spectrometry

in 20 samples between 0 and 64 cm depth. Measure-

ments were carried out at the St. Croix Watershed

Research Station, Science Museum of Minnesota

(USA). Results are expressed in pico-Curies per gram

dry sediment (pCi/g). The age model was obtained

using the CIC (constant initial concentration) model,

which assumes that the initial excess 210Pb activity in

surface sediments is constant and there is no mixing

of sediments (Robbins and Edgington 1975).

Since January 1986, the lake has been monitored

monthly for Secchi disk, water temperature, conduc-

tivity, pH, dissolved oxygen, nutrients (nitrates, sili-

con, total phosphorus, reactive phosphorus, total

chlorophyll and chlorophyll a) and hydrological (Tera

River water input and output, expressed as cubic

meters per month, m3/month) variables (De Hoyos

(1996). Table 1 shows the range of values for these

variables for the period 1986–2005 AD. Conductivity,

pH, temperature and dissolved oxygen were measured

in situ using a WTW multi-parameter liquid analyzer.

Secchi disk measurements were carried out using a

19-cm-diameter black and white disk. Water samples

were obtained using a 5-L PVC Niskin bottle.

Determination of nutrient concentrations generally

490 J Paleolimnol (2011) 46:487–502

123

Page 5: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

followed standard procedures (APHA 1989), though

some methods were modified slightly owing to low

chemical concentrations. Phosphates were determined

using a modified version of the method of Murphy and

Riley (1962), by using a spectrophotometer with a

cuvette containing a 100-mm light path. Nitrates

(Morris and Riley 1963) and chlorophyll (SCOR-

UNESCO 1966) contents were measured using stan-

dard methods, whereas silicon was determined using

the Dienert and Wandenbulcke (1923) method. Mea-

surements were generally carried out between the

second and the third week of every month in the center

of the deepest subbasin at 0, 2.5, 5, 10, 15, 20, 25, 35

and 45 m water depth for nutrients and every 2.5 m

from 0 to 50 for oxygen and temperature. Total and

reactive phosphorus concentrations are different

before and after 1990 owing to the installation of a

wastewater treatment plant on the Tera River just

above where it enters the lake. Total phosphorus

content was not measured systematically before

January 1992. Hence, the limnological dataset

employed in this study ranged from 1992 to 2005

AD, whereas the hydrological dataset comprised data

from January 1986 to December 2005. All variables

were measured on the same day.

Monthly maximum and minimum air temperature

(Celsius degrees), and monthly precipitation (millime-

ters) from the Ribadelago, Paramio de Sanabria and

Puebla de Sanabria meteorological stations cover the

period 1950–2007 and were used to explore relation-

ships between rainfall, Lake Sanabria hydrological

balance, and limnological variables. Meteorological

variables were checked for inconsistencies and quality

control followed the recommended procedure of Brunet

et al. (2006). The most complete period of instrumental

information, i.e. limnological and meteorological data,

spans from January 1992 to December 2007.

We used statistical approaches such as ordination

analyses (Principal Component Analyses), factor

analysis, time series (auto- and cross-correlation

functions) and linear models (lm) using the R software

package (R Development Core Team 2010) to inves-

tigate relationships between the chemical composition

of the sediment core and the limnological dataset.

Factor Analysis (FA) allowed us to define linear

combinations of variables (factors) that represent

underlying fundamental quantities of which the

observed variables are expressions. The idea behind

FA is that a small number of factors might explain aTa

ble

1M

axim

um

and

min

imu

mv

alu

eso

fth

em

on

thly

lim

no

log

ical

par

amet

ers

of

Lak

eS

anab

ria

for

the

Jan

uar

y1

98

6–

Mar

ch2

00

8st

ud

ied

per

iod

Par

amet

erW

ater

dep

th

Ox

yg

en

con

ten

t

Wat

er

tem

per

atu

re

pH

To

tal

ph

osp

ho

rus

Rea

ctiv

e

ph

osp

ho

rus

Nit

rate

sS

ilic

on

To

tal

Ch

loro

ph

yll

Ch

loro

ph

yll

-aS

ecch

i

dis

k

Co

nd

uct

ivit

y

Un

its

mm

g/l

�Clg

P/l

lg

P-P

O4/l

lgN

-NO

3/l

mg

/llg

/ll

g/l

ml

S/c

m

Min

imu

mv

alu

e0

1.5

3.4

5.2

10

.10

.10

.90

.06

0.0

13

.58

.6

Max

imu

mv

alu

e5

01

2.3

25

.28

.41

82

6.0

22

1.0

2.4

9.3

88

.31

12

.21

9.0

J Paleolimnol (2011) 46:487–502 491

123

Page 6: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

large number of measurements (Venables and Ripley

2002). Linear regression models belong to classical

statistics and are the basis for much of the statistical

practice. These statistical models assume that there is a

plausible physical explanation for a trend in a variable

that can be be modeled in a deterministic manner.

These models are part of a larger multivariate statistical

family that has been used widely to develop transfer

functions to reconstruct meteorological variables

quantitatively from biological and inorganic variables

(Guiot and de Vernal 2007 and references therein).

They have been used extensively in paleoclimatic and

paleoenvironmental reconstructions (Johnson and

Ingram 2004; Mangini et al. 2005; Danis et al.

2006; Blass et al. 2007; Guiot and de Vernal 2007).

Stratigraphically-constrained cluster analysis was

employed to define homogenous geochemical zones.

Results

Relationships between Lake Sanabria water input

and meteorological variables

Because the Tera River contributes 60–100% (mean

85%) of the total water input to Lake Sanabria,

depending on the month and year, this implies that

secondary streams and creeks on the south and NW

shores have a limited influence on the lake water

budget. Initially, the Tera River water discharge was

considered as the only water input to the lake.

Figure 2 displays the Tera River discharge entering

Lake Sanabria, total annual precipitation and the

maximum air temperature for the period between

January 1986 and December 2004. The Tera River

water discharge and precipitation show a similar

pattern, with maximum values between October and

April and minimum ones between May and September.

Maximum temperature values are inversely related to

both precipitation and Tera River water discharge.

Maximum temperatures are usually recorded between

May and October, whereas minimum values occur

during the rest of the year.

A linear model (lm) for the 1986–2004 period

shows that monthly meteorological variables rainfall

and temperature are strong predictors of annual

oscillations of Tera River water discharge. The two

meteorological variables explain [76% of the total

variance of the Tera River water discharge:

water discharge ¼ 8234454þ 39199 � precipitation

� 337594 � temperature

ð1Þ

where water discharge is expressed in m3/month,

precipitation in mm and temperature in degrees

Celsius. The best correlation between Tera River

water discharge and the meteorological variables

occurs when no lag time is used, indicating that

regional climate fluctuations have an immediate

effect on the total amount of water entering the lake.

Precipitation is the variable that contributes most

to oscillations in river discharge (r = 0.85, p \0.001), though temperature also has an effect

(r = -0.55, p \ 0.001). The lowest Tera river water

discharge values usually occur during the summer

months, when the temperature (precipitation) reach

Fig. 2 Monthly

instrumental maximum

temperature (blue solidline), monthly precipitation

(green solid line) and

monthly Tera River water

discharge (red solid line)

for the period between

January 1986 and

December 2004

492 J Paleolimnol (2011) 46:487–502

123

Page 7: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

its highest (lowest) values, indicating that evapora-

tion in the catchment also plays a role in the Lake

Sanabria water budget.

As precipitation is the instrumental variable that

explains the largest variability of the Tera River

water discharge to Lake Sanabria, water discharge

was modeled only using this meteorological variable:

water discharge ¼ 3852606þ 44518 � precipitation

ð2aÞThis equation explains 71% of the total variance of

Tera River water discharge, which indicates that

precipitation values can be inferred from water

discharge with accuracy. Reorganizing Eq. 2a, we

can write:

precipitation ¼ �86:54þ 2:25� 10�5

� water discharge ð2bÞ

Relationships between the Lake Sanabria limnological

parameters

Relationships between mean monthly limnological

variables in the eastern subbasin were explored using

Principal Component Analysis (PCA). The first two

eigenvectors (EV) explain 45% of the total variance.

The first one accounts for[24%, whereas the second

one explains 20.8% (Fig. 3).

The first EV is associated with dissolved oxygen,

reactive phosphorus and pH at the positive end, and

by water temperature, conductivity, and to lesser

extent, nitrates, chlorophyll a and total chlorophyll at

the negative end. On the other hand, the second EV is

related to Secchi disk and water temperature at the

positive end, and by pH, chlorophyll a and total

chlorophyll at the negative end (Fig. 3).

Lithology

SAN07-1 M and SAN07-2 M consist of homoge-

nous, massive, dark brown silty clays. Two erosion

surfaces were identified in core SAN07-1 M by

changes in sediment color, magnetic susceptibility

and total carbon (TC) values (Fig. 4). This interme-

diate layer, between 18.5 and 7 cm core depth and

identified in Fig. 4 by a gray band, is composed of

coarser silts, with higher magnetic susceptibility and

lower TC content. Mean grain size between 22 and

11 cm in the SAN07-2 M core ranges from about

30–60 lm, whereas sediments located above and

beneath have a mean grain size of about 20 lm. The

coarse layer is visible in both cores (gray bands of

Fig. 4) and displays a coarsening upwards sequence

with lighter, less organic sediment towards the top.

This distinctive layer corresponds to sediments

deposited during the Tera River dam failure in 1959

AD and is recognizable all over the lake (Luque

2003).

Chronological model

The 210Pb activity profile in SAN07-2 M shows a

general decreasing trend from the top of the core

(18.2 pCi/g) to 30 cm depth (0.43 pCi/g), below

which it is relatively constant to the base of the core

(Fig. 5). The 210Pb fluctuation between 11 and 22 cm

depth corresponds to the coarser, lighter-color sedi-

ment layer deposited during the Tera River dam

collapse. Lower 210Pb activities were a consequence

of dilution of the 210Pb by rapid accumulation of

coarser sediment after the dam collapse. Sedimenta-

tion of this layer is considered instantaneous, and it

was used as a chronological marker. There is

considerable excess 210Pb in the 10–11 cm interval,

which suggests gradual deposition of sediments over

time following the rapid deposition of the coarser

dam-related sediments. Application of the CIC model

Fig. 3 Plot of the plane defined by the first two eigenvectors

obtained by Principal Component Analysis (PCA) of the

monthly instrumental limnological dataset for the 1992–2005

AD period. Numbers refer to the years. Recent years are

indicated by large numbers

J Paleolimnol (2011) 46:487–502 493

123

Page 8: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

gives an date of 1961.4 ± 5.7 AD at 13 cm depth,

which is statistically indistinguishable from the date

of dam failure in 1959. Therefore, the uppermost

11 cm of core SAN07-2 M were deposited from 1959

to 2005 AD.

Geochemical composition of the sediments

XRF core scanner analyses of SAN07-1 M evaluated

17 chemical elements (Al, Si, P, S, K, Ca, Ti, Mn, Fe,

Ni, Zn, As, Se, Rb, Sr, Zr and Pb).The ratio between

the incoherent (Compton) and coherent (Raleigh)

X-ray scatter intensities was used as an indicator of

organic matter content (Fig. 6), as has been done for

other lake cores (Saez et al. 2009). The excellent

covariance between the measured TC and this ratio,

and the higher measurement resolution of the latter,

allowed us to use the inc/coh ratio in the statistical

model as indicator of the organic matter content.

Stratigraphically-constrained cluster analysis of

the uppermost 13 cm of XRF core scanner data from

core SAN07-1 M revealed that there are two geo-

chemically homogenous zones. Zone A extends from

13 to 6.35 cm depth and the inc/coh ratio is very low.

Zone B runs from 6.35 cm to the core top, where

there is a progressive upward increase in the inc/coh

ratio and a decrease in lithogenic elements.

Total Carbon (TC) profiles in sediments deposited

after the Tera River dam failure display the same

trends in both cores (Fig. 4). Above 6.35 cm in

SAN07-1 M and above 11 cm in SAN07-2 M, the

TC content generally rises, despite minor fluctua-

tions. Previous works carried out in other cores from

Lake Sanabria showed that the total inorganic carbon

(TIC) content is negligible (Luque 2003). Therefore,

for Lake Sanabria, TC represents the total organic

carbon (TOC) content.

The TC content, grain size, magnetic susceptibility

and lithological features of the deposits suggest that

the sediments of zone A, 22-11 cm in core SAN07-

2 M and 18.5-6.35 cm in core SAN07-1 M, were

deposited after the main flooding event, but nonethe-

less during a period of unusual sedimentation in the

lake caused by the Vega de Tera dam collapse. Zone

B corresponds to sediments deposited in Lake

Sanabria after a return to baseline lake sedimentation

Fig. 4 Main lithological features of cores SAN07-1 M and

SAN07-2 M. Core photographs, lightness values and magnetic

susceptibility correspond to the SAN07-1 M core, whereas the

mean and median grain size curves were obtained from core

SAN07-2 M. Total Carbon (TC) profiles were obtained in both

cores. The gray bands identify sediments deposited due to the

Vega de Tera dam collapse in 1959 AD

494 J Paleolimnol (2011) 46:487–502

123

Page 9: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

Fig. 5 Right. Unsupported 210Pb profile obtained by gamma

spectrometry on SAN07-2 M sediments. Left. Age model

derived form the application of the CIC (Constant Initial

Concentration) model. Horizontal error bars represent 95%

confidence interval. In both cases, the gray, thick vertical baridentifies the sediments correspondent to the Tera River dam

failure

Fig. 6 XRF-core scanner

profiles of the main

chemical elements in

SAN07-1 M sediments.

Stratigraphically-

constrained cluster analysis

allowed us to identify the

sediments related to the

Tera River dam collapse.

Distances are expressed as

squared Euclidean distances

J Paleolimnol (2011) 46:487–502 495

123

Page 10: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

conditions, following the catastrophic event. Princi-

pal Component Analysis (PCA) of the XRF dataset

was only conducted on samples from Zone B (Fig. 7).

The PCA results of the XRF dataset show that the

first two EVs account for more than 53% of the total

variance. The first EV explains 36.1%, whereas the

second one only accounts for 17.1%. The first eigen-

vector is mainly tied to the inc/coh ratio, an indicator of

the organic matter content as explained above, and to a

lesser extent, to S, P and Se at the positive end and K,

Ti, Si, Ni and Fe at the negative end. The second

eigenvector (EV2) is controlled by Pb, Zr and Sr, and to

a lesser extent by Ca, Rb, Mn and S at the positive end

and by Al and As at the negative end (Fig. 7). Factor

Analysis (FA) of this dataset found that only EV1 is

significant (w2 = 755.85 on 44 degrees of freedom,

p \ 0.0001).

Discussion

Chronological model

The chronological model for core SAN07-2 M has an

associated dating uncertainty (standard deviation)

that ranges between 2.46 and 4.66 years for the

period 2005–1959 AD. This uncertainty confounds

age control when the chronological model from core

SAN07-2 M is transferred to core SAN07-1 M. In

spite of this problem, good agreement between the

TC profiles in the two cores permits good strati-

graphic correlation between the profiles and justifies

accurate projection of 210Pb modeled ages from dated

core SAN07-2 M to depths in core SAN07-1 M,

which was analyzed by XRF (Fig. 8). Sediments

corresponding to the Tera River dam collapse are

found in core SAN07-1 M between about 18 and

7 cm depth and the entire interval corresponds to

1959 AD (Fig. 4). Consequently, the uppermost

6.5 cm of the sediment in core SAN07-1 M contains

the environmental history from 1959 to 2005 AD.

Environmental interpretation of the limnological

eigenvectors

We focused on the first EV because it explains the

largest percent of the total variance (24.2%). Oppo-

sition of dissolved oxygen and reactive phosphorus,

to water temperature and conductivity, and to lesser

extent nitrates, in the EV1 of the PCA, is thought to

be related to periodic oscillations in nutrient input to

the lake linked to water input (Fig. 9). During

summer, when the water input to the lake is low

and the lake water column is thermally stratified,

which implies warm epilimnetic waters, nitrogen-fixing

Fig. 7 Plot of the first two eigenvectors obtained by the

Principal Component Analysis (PCA) of the topmost 6 cm in

the SAN07-1 M XRF-core scanner dataset. Numbers represent

the location of the samples with respect to their core depth.

Large numbers indicate samples located close to the bottom of

the core, whereas small numbers point to samples near the top

Fig. 8 Comparison of Total Carbon (TC) variations in cores

SAN07-1 M (light dashed line) and SAN07-2 M (dark solidline) for the period 1960–2006. Note that despite the different

sampling intervals applied to both cores, the TC contents

display the same long-term trend

496 J Paleolimnol (2011) 46:487–502

123

Page 11: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

cyanobacteria proliferate in the metalimnion. As

summer progresses, Secchi disk measurements

increase and reach their maximum values during

autumn, due to water column stability, feeding

activity of zooplankton (rotifers) and to the low

productivity of phytoplankton in the epilimnion. The

major water input to the lake, however, occurs during

autumn and winter. During this period, cold, well-

oxygenated water enters the lake, along with

allochthonous suspended material and nutrients. Fur-

thermore, the thermocline disappears, provoking

mixing of the lake water column. Mixing oxygenates

the bottom waters and recirculates phosphorus that

accumulated in deeper water layers during the period

of stratification. The input of external nutrients and

regeneration of available phosphorus triggers the

growth of diatoms during this period (De Hoyos

1996; Luque 2003).

The first EV of the PCA on the monthly limno-

logical dataset can therefore, be interpreted as an

indicator of wet/dry season conditions (pplimno).

Positive values of the EV1 indicate wet conditions

and are linked to the rainy season (spring and

autumn), whereas negative values of this eigenvector

mark dry-season conditions related to summer. A plot

of the x-coordinates of every sample, calculated from

the vector plane defined by the first two eigenvectors

of the PCA, with respect to their core depth, allowed

us to qualitatively reconstruct wet/dry-season condi-

tions (pplimno) through time.

Periodic oscillations of the pplimno are clearly

linked to fluctuations of the Tera River water inputs

to the lake (Fig. 9). These calculated wet/dry-season

conditions do not track the Tera River water input

exactly. The best correlation between the two vari-

ables is obtained with a one-month lag time

(r = 0.66, p \ 0.001).

Good correlation between the instrumental Tera

River water discharge and the reconstructed wet/dry-

season conditions allowed us to model the first

variable as a function of the second one. The equation

that defines the relationship between these two

variables, taking into account the one-month response

(lag) time is:

water discharge ¼ 9633827þ 3026090 � pplim no

ð3Þ

where water discharge is expressed as m3/month and

pplimno does not have units.

In this case, the equation explains 43.3% of the

total variance (p \ 0.0001), suggesting that the lake

water column smoothes the meteorological signal due

to its buffer effect.

Environmental interpretation

of the sedimentological eigenvectors

Similar to the limnological data set, we only focused

on the environmental significance of the first EV

because it explains the largest variance of the PCA.

The opposite behavior of organic matter (inc/coh

ratio) versus K, Ti, Si, Ni and Fe content (Fig. 7) can

be interpreted in terms of oscillations in primary

productivity (ppxrf). Higher inc/coh values indicate

Fig. 9 Comparison of the first EV obtained in the PCA

applied to the limnological dataset (dark solid line) with Tera

River water discharge to Lake Sanabria (light solid line) for the

period 1992–2006. Note the excellent agreement between both

curves when a one-month lag is applied to the first dataset

J Paleolimnol (2011) 46:487–502 497

123

Page 12: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

higher organic matter in the sediments. Accumulation

of organic matter in the sediments is a consequence

of high biotic activity in the lake water column,

mainly chlorophyceans and diatoms, because they are

the most dominant phytoplankton in the lake. On the

other hand, terrigenous input to the lake is directly

related to the Tera River water input. Autumn rainfall

provokes an increase in catchment runoff and mobi-

lization of topsoil particles that are delivered to the

Tera River at the same time that Tera River water

discharge to Lake Sanabria increases. Therefore,

positive values of the ppxrf correspond to periods of

enhanced primary productivity and organic matter

accumulation, whereas negative ones represent peri-

ods of low primary productivity.

Comparison of the first limnological

and sedimentological eigenvectors

The first eigenvector of the PCA applied to the

limnological dataset (pplimno) responds to changes in

the wet/dry-season conditions, whereas the first

eigenvector from the XRF core scanner dataset

(ppxrf) responds to oscillations in lake primary

productivity. Therefore, they must show the same

temporal responses, as they must be related. During

wet-season conditions, higher input of nutrient-rich

and well-oxygenated waters increases lake primary

productivity. Conversely, during the dry season, there

is lower input of nutrients, waters are poorly

oxygenated, and primary productivity is lower.

Figure 10 compares the first eigenvector obtained

in the PCA applied to the limnological dataset with

that obtained in the PCA applied to the XRF core

scanner dataset for the period 1992–2005 AD. In

spite of differences in smoothness, due to the

different temporal sampling intervals, and a temporal

offset, the two curves show similar patterns. The

pplimno displays an annual pattern linked to the

seasonality of rainfall. The ppxrf shows a pluriannual

oscillation pattern, possibly related to two factors: (1)

low sedimentation rate in Lake Sanabria and (2) the

‘‘low-pass filter’’ nature of most non-varved lakes,

filtering out the high-frequency climate oscillations.

Varved lakes are known to capture inter-annual

climate varibility because of their limnological

(seasonally stratified, anoxic hypolimnion), sedi-

mentological (non-bioturbated sediments, marked

seasonality, etc.) and geomorphological (deep and

wind-protected lakes) peculiarities. These features

are lacking in almost all non-varved sediments. This

probably accounts for the fact that this inter-annual

climate variability is smoothed out and therefore, not

recorded in the sediments. This ‘‘low-pass filter’’

nature of non-varved lakes has been recognized in

other water bodies such as Lake Gallocanta (Rodo

et al. 1997). Another explanation is that the sampling

step in the XRF measurements was too large to

capture higher-frequency environmental events.

The temporal offset between 1992 and 2005 may

be related to uncertainties in the age model. Despite

the excellent agreement between the TC curves in the

Fig. 10 Comparison of the

first eigenvector obtained in

the PCA applied to the

limnological dataset (darkthin solid line) with that

obtained in the PCA applied

to the XRF core scanner

dataset (light thick solidline)

498 J Paleolimnol (2011) 46:487–502

123

Page 13: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

two cores, which allowed projection of 210Pb dates

from one core to the other, small correlation errors

may have been made. Furthermore, the 210Pb age

model plot (Fig. 5) shows that age uncertainties

become larger with core depth, which could have led

to the observed temporal offset. Both factors have a

positive feedback. Nevertheless, the maximum tem-

poral offset identified for the 13 years of the studied

period (1992–2005 AD) is about 1 year, i.e. *1.3

mm of sediment thickness.

This long-term fit between the pplimno and ppxrf

curves also permits us to model the first curve as a

function of the second one. The equation that relates

both reconstructions is:

pplim no ¼ 0:23þ 0:92 � ppxrf ð4Þ

where both variables do not have units. The percent

variance explained by Eq. (4) is low (9%, p \ 0.001),

but this could be related to the low-pass filter nature

of the XRF core scanner dataset. When comparing

the pplimno values derived from Eq. (3) with those

obtained by applying Eq. (4), there is a notable

difference in the range of the values. The values of

the pplimno curve from Eq. (3) range from -4 to 4,

whereas those from reconstructed pplimno using

the Eq. (4) vary from -1 to 1, suggesting that the

depositional processes in the lake integrate the

climate signal, but smooth out high-frequency

climate events (Fig. 10).

Quantitative reconstruction of the precipitation

of the Lake Sanabria region

Equations (2a, 2b, 3) and (4) show that it is possible

to trace how the climate signal from the atmosphere

is transferred to the lake water mass, and from there

to the sediments. This should allow us to invert such

relations and reconstruct quantitatively precipitation

oscillations using the first score of the PCA carried

out with XRF core scanner values. Substituting (4) in

(3) and the subsequent Eq. in (2b) permits us to

obtain an equation relating the instrumental precip-

itation values to the ppxrf. This general equation is:

precipitation ¼ 145:27þ 62:63 � ppfrx ð5Þ

where ppfrx does not have units and precipitation is

expressed in mm. Figure 11 shows precipitation

inferred using the EV1 from the PCA on the XRF core

scanner dataset (ppfrx) for the period 1959–2005 AD.

The comparison between reconstructed and instru-

mental rainfall shows similar patterns and although the

reconstructed precipitation is lower, most of the rainier

months are reflected in the reconstructed curve. The

reconstructed precipitation follows the long-term

pattern of the measured rainfall, smoothing out its

high-frequency components, but capturing the annual

variability. This is apparent if the reconstructed

precipitation is compared with the 5-month-smoothed

instrumental precipitation (Fig. 11).

Fig. 11 Quantitative reconstructed precipitation curve obtained

from the XRF core scanner values for the 1959–2007 AD period

(dark thick line). This quantitative reconstruction is compared

with the raw (dashed thin line) and 5-month-smoothed

(continuous thin line) instrumental meteorological rainfall

curves for the 1959–2005 AD period. The periods with no

instrumental meteorological data are highlighted with the grey

bands

J Paleolimnol (2011) 46:487–502 499

123

Page 14: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

Conclusions

The [20-years monitoring dataset for Sanabria Lake

provided a unique opportunity to investigate the

relationship between recent depositional dynamics in

the lake and climate. The comparison of limnolog-

ical, hydrological and meteorological datasets with

the chemical composition of short cores from Lake

Sanabria illustrates the potential and pitfalls of

climate reconstructions based on high-resolution

sediment analyses. The Tera River water discharge

to Lake Sanabria is directly controlled by rainfall, and

to lesser extent by temperature. The best correlation

between the hydrological and meterological datasets

is obtained when no lag is introduced, indicating that

the hydrologic response of Lake Sanabria to rainfall

and temperature is immediate. PCA of the limnolog-

ical dataset showed that the largest variance is linked

to changes in nutrient input. Comparison of the first

eigenvector (EV) of this PCA with the Tera River

water discharge for the period 1992–2005 AD

demonstrates that the pattern of the wet/dry-season

conditions follows fluctuations in the hydrological

balance, and therefore, oscillations in precipitation.

The lag time between these two datasets is around

1 month. The PCA of the XRF core scanner

measurements showed that the main process control-

ling the chemical composition of the Lake Sanabria

sediments is related to primary productivity, and

therefore, to river water input. Comparison of the

precipitation reconstructed using the limnological

data with that obtained from the XRF core scanner

dataset indicated that the sediments act as a low-pass

filter, smoothing the climate (rainfall) signal.

Our analyses demonstrate that the climate signal is

transmitted from the atmosphere to the lake water

mass and, from there to the sediments. The climate

signal, however, is smoothed out by the sedimenta-

tion process and uncertainties in the age model cause

temporal offsets between measured and reconstructed

rainfall. This study highlights that two main limita-

tions might hamper application of the proposed

method: (1) lack of long and reliable limnological

datasets, and (2) lack of a robust core chronology that

minimizes temporal uncertainties related to the

reconstructed meteorological variable. In spite of

these problems, it was possible to establish links

between these datasets, and to obtain a quantitative

reconstruction of rainfall during the period

1959–2005 AD, which captured the main temporal

variability. This quantitative precipitation reconstruc-

tion suggests that it may be possible to obtain

accurate climate reconstructions using non-laminated

sediments.

Acknowledgments The Spanish Ministry of Science and

Innovation funded the research at Lake Sanabria through

the projects LIMNOCLIBER (REN2003-09130-C02-02/CLI),

CALIBRE (CGL2006-13327-C04/CLI), IBERLIMNO (CGL2004-

20236-E) and CONSOLIDER-GRACCIE (CSD2007-00067). The

Limnological Research Center and the Large Lakes Observatory

(University of Minnesota, USA) are acknowledged for technical

assistance with the ITRAX XRF-core scanner and the GEOTEK,

and Daniel R. Engstrom from St. Croix Watershed Research

Station (Science Museum of Minnesota, USA) for radiometric

dating of the SAN07-2 M core. We thank Javier Sigro and

Manola Brunet (Centre for Climate Change, University Rovira i

Virgili, Tarragona, Spain) for providing the precipitation and

temperature data. We are grateful to the ‘‘Parque Natural del Lago

de Sanabria y Alrededores’’ of the Consejerıa de Medio Ambiente

y Ordenacion del Territorio de la Junta de Castilla y Leon

(Environmental Council of Castilla and Leon Autonomous

Spanish Region) and owner of the Laboratorio de Limnologıa

del Parque Natural (Laboratory of Limnology of the Natural Park)

for the field and administrative facilities. Dr. Mark Brenner, Dr.

Pere Anadon and an anonymous referee are acknowledged for

their exhaustive reviews and very constructive comments that

greatly helped to improve this manuscript.

References

Abrantes F, Lopes C, Rodrigues T, Gil I, Witt L, Grimalt J,

Harris I (2009) Proxy calibration to instrumental data set:

implications for paleoceanographic reconstructions. Geo-

chem Geophy Geosy 10. doi: 10.1029/2009GC002604

Aldasoro JJ, De Hoyos C, Vega JC (1991) El lago de Sanabria

(Estudio Limnologico). Publicaciones Monograficas de la

Red de Espacios Naturales de Castilla y Leon. Consejerıa

de Medio Ambiente y Ordenacion del Territorio, 2. Di-

reccion General del Medio Ambiente, Valladolid (Spain)

APHA, AWWA, WPCF (1989) Standard methods for the

examination of water and wastewater, 17th edn. USA

Battarbee RW, Grytnes JA, Thompson R, Appleby PG, Catalan

J, Korhola A, Birks HJB, Heegaard E, Lami A (2002)

Comparing palaeolimnological and instrumental evidence

of climate change for remote mountain lakes over the last

200 years. J Paleolimnol 28:161–179

Battarbee RW, Monteith DT, Juggins S, Evans CD, Jenkins A,

Simpson GL (2005) Reconstructing pre-acidification pH

for an acidified Scottish loch: a comparison of palaeo-

limnological and modelling approaches. Environ Pollut

137:135–149

Bennion H, Johnes P, Ferrier R, Phillips G, Haworth E (2005)

A comparison of diatom phosphorous transfer functions

and export coefficient models as tools for reconstructing

lake nutrient histories. Freshw Biol 50:1651–1670

500 J Paleolimnol (2011) 46:487–502

123

Page 15: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

Bigler C, Hall RI (2002) Diatoms as indicators of climatic and

limnological change in Swedish Lapland: a 100-lake

calibration set and its validation for paleoecological

reconstructions. J Paleolimnol 27:97–115

Blass A, Bigler C, Grosjean M, Sturm M (2007) Decadal-scale

autumn temperature reconstruction back to AD 1580

inferred from the varved sediments of Lake Silvaplana

(southeastern Swiss Alps). Quat Res 68:184–195

Brunet M, Saladie O, Jones P, Sigro J, Aguilar E, Moberg A,

Walther A, Lister D, Lopez D, Almarza C (2006) The

development of a new daily adjusted temperature dataset

for Spain (1850–2003). Int J Climatol 26:1777–1802

Cacho I, Grimalt JO, Canals M, Sbaffi L, Shackleton NJ,

Schonfeld J, Zahn R (2001) Variability of the western

Mediterranean Sea surface temperature during the last

25,000 years and its connection with the Northern Hemi-

sphere climatic changes. Paleoceanography 16:40–52

Carrion JS, Fernandez S, Gonzalez-Samperiz P, Gil-Romera G,

Badal E, Carrion-Marco Y, Lopez-Merino L, Lopez-Saez

JA, Fierro E, Burjachs F (2010) Expected trends and

surprises in the Lateglacial and Holocene vegetation his-

tory of the Iberian Peninsula and Balearic Islands. Rev

Palaeobot Palynol 162:458–475

Catalan J, Pla S, Garcia J, Camarero L (2009) Climate and CO2

saturation in an alpine lake throughout the Holocene.

Limnol Oceanogr 54:2542–2552

Chen FH, Chen JH, Holmes J, Boomer I, Austin P, Gates JB,

Wang NL, Brooks SJ, Zhang JW (2010) Moisture changes

over the last millennium in arid central Asia: a review,

synthesis and comparison with monsoon region. Quat Sci

Rev 29:1055–1068

Cowton T, Hughes PD, Gibbard PL (2009) Paleoglaciation of

Parque Natural Lago de Sanabria, northwest Spain. Geo-

morphology 108:282–291

Danis PA, Masson-Delmotte V, Stievenard M, Guillemin MT,

Daux V, Naveau Ph, von Grafenstein U (2006) Recon-

struction of past precipitation d18O using tree-ring cellu-

lose d18O and d13C: a calibration study near Lac

d’Annecy, France. Earth Planet Sci Lett 243:439–448

De Hoyos C (1996) Limnologıa del Lago de Sanabria: vari-

abilidad interanual del fitoplancton. Unpublished Ph.D.

Thesis. Universidad de Salamanca, 438 pp

Dienert F, Wandenbulcke F (1923) Sur le dosage de la silice

dans les eaux. C R Acad Sci 176:1478–1480

Djamali M, de Beaulieu JL, Andrieu-Poniel V, Berberian M,

Miller NF, Gandouin E, Lahijani H, Shah-Hosseini M,

Ponel P, Salimian M, Guiter F (2009) A late Holocene

pollen record from Lake Almalou in NW Iran: evidence

for changing land-use in relation to some historical events

during the last 3,700 years. J Archaeol Sci 36:1364–1375

Francis DR, Wolfe AP, Walker IR, Miller GH (2006) Inter-

glacial and Holocene temperature reconstructions based

on midge remains in sediments of two lakes from Baffin

Island, Nunavut, Arctic Canada. Palaeogeogr Palaeocli-

matol Palaeoecol 236:107–124

Fritz SC (1990) Twentieth-century salinity and water-level

fluctuations in Devils Lake, North Dakota: test of a dia-

tom-based transfer function. Limnol Oceanogr 35:

1771–1781

Granados I, Toro M (2000) Recent warming in a high mountain

lake (Laguna Cimera, Central Spain) inferred by means of

fossils chironomids. In: Lami A, Cameron N, Korhola A

(eds) Paleolimnology and ecosystem dynamics at remote

European Alpine lakes. J Limnol 59(Suppl 1):109–119

Guiot J, de Vernal A (2007) Tranfer functions methods for

quantitative paleoceanography based on microfossils. Dev

Mar Geol 1:523–563

Hyodo F, Tsugeki N, Azuma J, Urabe J, Nakanishi M, Wada E

(2008) Changes in stable isotopes, lignin-derived phenols,

and fossil pigments in sediments of Lake Biwa, Japan:

implications for anthropogenic effects over the last

100 years. Sci Total Environ 403:139–147

Johnson KR, Ingram BL (2004) Spatial and temporal vari-

ability in the stable isotope systematics of modern pre-

cipitation in China: implications for paleoclimate

reconstructions. Earth Planet Sci Lett 220:365–377

Kalugin I, Daryin A, Smolyaninova L, Andreev A, Diekmann

B, Khlystov O (2007) 800-yr long records of annual air

temperature and precipitation over southern Siberia

inferred from Teletskoye Lake sediments. Quat Res 67:

400–410

Kattel GR, Battarbee RW, Mackay AW, Birks HJB (2008)

Recent ecological change in a remote Scottish mountain

loch: an evaluation of a Cladocera-based temperature

transfer-function. Palaeogeogr Palaeoclimatol Palaeoecol

259:51–76

Larocque I, Finsinger W (2008) Late-glacial chironomid-based

temperature reconstructions for Lago Piccolo di Avigliana

in the southwestern Alps (Italy). Palaeogeogr Palaeocli-

matol Palaeoecol 257:207–223

Luque JA (2003) El Lago de Sanabria: un sensor de las osci-

laciones climaticas del Atlantico Norte durante los ulti-

mos 6.000 anos. El registro de la Pequena Edad del Hielo

(LIA), los eventos Dansgaard-Oeschger (D-O), y el ciclo

de Bond. Unpublished Ph.D. Thesis. University of Bar-

celona, 446 pp

Luque JA, Julia R (2002) Lake sediment response to land-use

and climate change during the last 1000 years inthe oli-

gotrphic Lake Sanabria (northwest of Iberian Peninsula).

Sediment Geol 148:343–355

Mangini A, Spotl C, Verdes P (2005) Reconstruction of tem-

perature in the Central Alps during the past 2000 yr from a

d18O stalagmite record. Earth Planet Sci Lett 235:741–751

Martınez-Garcıa E (1973) Deformacion y metamorfismo en la

zona de Sanabria. Studia Geologica Salmanticensia 5:

7–106

Martın-Puertas C, Valero-Garces BL, Mata MP, Gonzalez-

Samperiz P, Bao R, Moreno A, Stefanova V (2008) Arid

and humid phases in southern Spain during the last

4000 years: the Zonar Lake record, Cordoba. Holocene 18:

907–921

Morris AW, Riley JP (1963) The determination of nitrate in

sea-water. Anal Chim Acta 29:272–279

Murphy J, Riley IP (1962) A modified single solution method

for the determination of phosphate in natural waters. Anal

Chim Acta 27:31–36

Nichols JE, Walcott M, Bradley R, Pilcher J, Huang Y (2009)

Quantitative assessment of precipitation seasonality and

summer surface wetness using ombrotrophic sediments

from an Arctic Norwegian peatland. Quat Res 72:443–451

Pla S, Catalan J (2005) Chrysophyte cysts from lake sediments

reveal the submillennial winter/spring climate variability

J Paleolimnol (2011) 46:487–502 501

123

Page 16: Consejo Superior de Investigaciones Científicas - Quantitative …diapiro.ictja.csic.es/~sgiralt/articles/sgiralt_et_al... · 2020. 1. 16. · Laboratory of Limnology of the Lago

in the northwestern Mediterranean region throughout the

Holocene. Clim Dyn 24:263–268

R Development Core Team (2010) R: a language and envi-

ronment for statistical computing. R Foundation for Sta-

tistical Computing, Vienna

Robbins A, Edgington DN (1975) Determination of recent

sedimentation rates in Lake Michigan using Pb-210 and

Cs-137. Geochim Cosmochim Acta 39:285–304

Rodo X, Baert E, Comın FA (1997) Variations in seasonal

rainfall in Southern Europe during the present century:

relationships with the North Atlantic Oscillation and the

El Nino-Southern Oscillation. Clim Dyn 13:275–284

Romero-Viana L, Julia R, Camacho A, Vicente E, Miracle MR

(2008) Climate signal in varve thickness: Lake La Cruz

(Spain), a case study. J Paleolimnol 40:703–714

Saez A, Valero-Garces BL, Giralt S, Moreno A, Bao R, Pueyo

JJ, Hernandez A, Casas D (2009) Glacial to Holocene

climate changes in the SE Pacific. The Raraku Lake

sedimentary record (Easter Island, 27�S). Quat Sci Rev

28:2743–2759

Schwalb A, Dean WE (2002) Reconstruction of hydrological

changes and response to effective moisture variations

from North-Central USA lake sediments. Quat Sci Rev

21:1541–1554

SCOR-UNESCO (1966) Determination of photosynthetic pig-

ments in sea water. UNESCO Monographs on Oceano-

graphic Methodology 1. Paris

Striewski B, Mayr C, Flenley J, Naumann R, Turner G, Lucke

A (2009) Multi-proxy evidence of late Holocene human-

induced environmental changes at Lake Pupuke, Auck-

land (New Zealand). Quat Int 202:69–93

Tonello MS, Mancini MV, Seppa H (2009) Quantitative

reconstruction of Holocene precipitation changes in

southern Patagonia. Quat Res 72:410–420

Toro M, Granados I, Robles S, Montes C (2006) High moun-

tain lakes of the Central Range (Iberian Peninsula):

regional limnology & environmental changes. Limnetica

25:217–252

Trachsel M, Eggenberger U, Grosjean M, Blass A, Sturm M

(2008) Mineralogy-based quantitative precipitation and

temperature reconstructions from annually laminated lake

sediments (Swiss Alps) since AD 1580. Geophys Res Lett

35:L13707. doi:10.1029/2008GL034121

Vega JC, Aldasoro JJ (1994) Geologıa de Sanabria. In: Mo-

nografıas de la Red de Espacios Naturales de Castilla y

Leon. Consejerıa de Medio Ambiente y Ordenacion del

Territorio (Junta de Castilla y Leon), Valladolid

Vega JC, De Hoyos C, Aldasoro JJ (1991) Estudio del sistema

de lagunas de las sierras Segundera y Cabrera. In: Mo-

nografıas de la Red de Espacios Naturales de Castilla y

Leon, Consejerıa de Medio Ambiente y Ordenacion del

Territorio (Junta de Castilla y Leon), Valladolid

Vega JC, De Hoyos C, Aldaroso JJ (1992) The Sanabria Lake:

the largest natural freshwater lake in Spain. Limnetica

8:49–57

Vega JC, De Hoyos C, Aldasoro JJ, Fraile H (2005) Nuevos

datos morfometricos para el lago de Sanabria. Limnetica

24:115–122

Venables WN, Ripley BD (2002) Modern applied statistics

with S. Springer Science ? Business Media, LLC, New

York

502 J Paleolimnol (2011) 46:487–502

123