late holocene environmental changes inferred from diatom, chironomid, and pollen assemblages in an...

19
MOUNTAIN LAKES Late Holocene environmental changes inferred from diatom, chironomid, and pollen assemblages in an Andean lake in Central Chile, Lake Laja (36°S) Roberto Urrutia Alberto Araneda Laura Torres Fabiola Cruces Caterina Vivero Fernando Torrejo ´n Ricardo Barra Nathalie Fagel Burkhard Scharf Published online: 28 April 2010 Ó Springer Science+Business Media B.V. 2010 Abstract A sediment core encompassing the last 2000 years was extracted from Lake Laja, Chile, (36°54 0 S, 71°05 0 W) using an Uwitec drilling plat- form. The sediment was subsampled for loss on ignition, nutrients, biogenic silica, and biological proxies (diatoms, chironomids, pollen). The sedimen- tary profile was characterized by several coarse volcanic layers. Loss on ignition, nutrients, and biogenic silica showed an increasing trend that suggests a recent shift to a higher trophic status. Diatom assemblages also suggested higher nutrient content with increased abundances of Aulacoseira granulata, A. distans, and Asterionella formosa. At the same time, a marked change in the benthic and facultative planktonic taxa may be associated with cooling. This period of change coincides with the European Little Ice Age (LIA). The chironomid profile showed four key zones distinguished largely by changes in the abundance of Tanytarsini, Para- chironomus, and Macropelopia. Like diatoms, chir- onomids also seemed to reflect a shift to higher trophic conditions in the upper part reflected by increasing abundance of taxa such as Tribelos/ Phaenopsectra, Cricotopus/Orthocladius, and Abla- besmyia. The most striking feature in the chironomid assemblage is the abundance of Podonominae, Parapsectrocladius, and Limnophyes/Compterosmit- tia, which could be associated with a cold-dry period between 1500 and 1900 AD in Lake Laja (the period of the European LIA). Pollen assemblages indicated fluctuations in humidity through changes in Nothof- agus dombeyi-type, Poaceae, and Ephedra, and we inferred a strong human impact over the last 100 years from the appearance of Plantago and increased levels of Poaceae and Asteraceae subf. Cichorioidae. Finally, the three proxies showed the occurrence of a cold-dry event in Lake Laja (*1550– 1900 AD), which roughly coincides with the Guest editors: Hilde Eggermont, Martin Kernan & Koen Martens / Global change impacts on mountain lakes R. Urrutia (&) A. Araneda L. Torres F. Cruces C. Vivero F. Torrejo ´n R. Barra Group of Paleolimnological Studies (GEP), Aquatic Systems Research Unit, Environmental Sciences Center EULA-Chile, University of Concepcion, Barrio Universitario s/n, Concepcion, Chile e-mail: [email protected] L. Torres Department of Basic Sciences, Education School, Campus Los Angeles, University of Concepcion, Juan Antonio Coloma 0201, Los Angeles, Chile F. Cruces Faculty of Natural and Oceanographic Sciences, Department of Botany, University of Concepcion, Concepcion, Chile N. Fagel Clays and Paleoclimate Research Unit, Department of Geology, University of Liege, Liege, Belgium B. Scharf Centre for Environmental Research, UFZ-Centrum, Mardeburg, Germany 123 Hydrobiologia (2010) 648:207–225 DOI 10.1007/s10750-010-0264-1

Upload: roberto-urrutia

Post on 15-Jul-2016

214 views

Category:

Documents


2 download

TRANSCRIPT

MOUNTAIN LAKES

Late Holocene environmental changes inferred from diatom,chironomid, and pollen assemblages in an Andean lakein Central Chile, Lake Laja (36�S)

Roberto Urrutia • Alberto Araneda • Laura Torres • Fabiola Cruces •

Caterina Vivero • Fernando Torrejon • Ricardo Barra • Nathalie Fagel •

Burkhard Scharf

Published online: 28 April 2010

� Springer Science+Business Media B.V. 2010

Abstract A sediment core encompassing the last

2000 years was extracted from Lake Laja, Chile,

(36�540S, 71�050W) using an Uwitec drilling plat-

form. The sediment was subsampled for loss on

ignition, nutrients, biogenic silica, and biological

proxies (diatoms, chironomids, pollen). The sedimen-

tary profile was characterized by several coarse

volcanic layers. Loss on ignition, nutrients, and

biogenic silica showed an increasing trend that

suggests a recent shift to a higher trophic status.

Diatom assemblages also suggested higher nutrient

content with increased abundances of Aulacoseira

granulata, A. distans, and Asterionella formosa. At

the same time, a marked change in the benthic and

facultative planktonic taxa may be associated with

cooling. This period of change coincides with the

European Little Ice Age (LIA). The chironomid

profile showed four key zones distinguished largely

by changes in the abundance of Tanytarsini, Para-

chironomus, and Macropelopia. Like diatoms, chir-

onomids also seemed to reflect a shift to higher

trophic conditions in the upper part reflected by

increasing abundance of taxa such as Tribelos/

Phaenopsectra, Cricotopus/Orthocladius, and Abla-

besmyia. The most striking feature in the chironomid

assemblage is the abundance of Podonominae,

Parapsectrocladius, and Limnophyes/Compterosmit-

tia, which could be associated with a cold-dry period

between 1500 and 1900 AD in Lake Laja (the period

of the European LIA). Pollen assemblages indicated

fluctuations in humidity through changes in Nothof-

agus dombeyi-type, Poaceae, and Ephedra, and we

inferred a strong human impact over the last

100 years from the appearance of Plantago and

increased levels of Poaceae and Asteraceae subf.

Cichorioidae. Finally, the three proxies showed the

occurrence of a cold-dry event in Lake Laja (*1550–

1900 AD), which roughly coincides with the

Guest editors: Hilde Eggermont, Martin Kernan & Koen

Martens / Global change impacts on mountain lakes

R. Urrutia (&) � A. Araneda � L. Torres �F. Cruces � C. Vivero � F. Torrejon � R. Barra

Group of Paleolimnological Studies (GEP), Aquatic

Systems Research Unit, Environmental Sciences Center

EULA-Chile, University of Concepcion, Barrio

Universitario s/n, Concepcion, Chile

e-mail: [email protected]

L. Torres

Department of Basic Sciences, Education School,

Campus Los Angeles, University of Concepcion,

Juan Antonio Coloma 0201, Los Angeles, Chile

F. Cruces

Faculty of Natural and Oceanographic Sciences,

Department of Botany, University of Concepcion,

Concepcion, Chile

N. Fagel

Clays and Paleoclimate Research Unit, Department

of Geology, University of Liege, Liege, Belgium

B. Scharf

Centre for Environmental Research, UFZ-Centrum,

Mardeburg, Germany

123

Hydrobiologia (2010) 648:207–225

DOI 10.1007/s10750-010-0264-1

European LIA. However the data from this research,

does not prove neither rejects the existence of the

occurrence of the MWP in the central Andes.

Keywords Late Holocene � MWP �LIA � Chilean Andes � Lake sediments �Multiproxy approach

Introduction

High resolution paleoclimate reconstructions of the

last 1000 years have become very important in recent

decades. According to Esper et al. (2005), under-

standing the correct amplitude of major climatic

episodes in the past millennium is ‘‘critical for

predicting future trends.’’ This is because evidence

showing that the amplitudes of major historical

climate episodes were as great as or even greater

than twentieth century global warming would indi-

cate the continued importance of the role played by

natural drivers in forcing temperature changes.

Among the climate events of the last millennium,

the Medieval Warm Period (MWP), or Medieval

Climate Anomaly (Bradley et al., 2003) which

occurred from *900 to 1300 AD and the Little Ice

Age (LIA; *1350–1850 AD) have gained much

attention. However, despite the considerable research

effort, the magnitude and timing of these events are

still debated, particularly the timing and synchronic-

ity (Lamb, 1977; Bradley & Jones, 1993; Mann et al.,

1998; Grove, 2001; Ogilvie & Jonsson, 2001; Mann,

2002; Soon et al., 2003; Goosse et al., 2005), and

some researchers even doubt the global expression of

such events (Winkler, 2004; McKinzey et al., 2004).

Hughes & Diaz (1994) indicated that, in some

areas of the world, temperatures were relatively high

during the MWP. However, other studies have shown

that temperatures during this period were slightly

above current levels (Huang & Pollack, 1997; Loehle,

2007), and temperatures at other sites such as

southeastern US and southern Europe did not differ

from those of previous or subsequent centuries

(Goosse et al., 2005). Crowley & Lowery (2000)

pointed out an asynchrony in the worldwide expres-

sion of this warm event. Moreover, Bradley et al.

(2003) rejected the idea of a warm event during

medieval times, indicating the need to disentangle

natural climate variability from anthropogenic

forcing and, thus, the necessity for more reconstruc-

tions of the last millennium.

High resolution climatic studies of Late Holocene

climate events such as the LIA and MWP are

common for the Northern Hemisphere (Jones et al.,

2001; Soon et al., 2003; Goosse et al., 2005; Torres

et al., 2008). Comparatively little research has been

published about these events in South America (e.g.,

Villalba, 1994; Lamy et al., 2001; Koch & Kilian,

2005; Bertrand et al., 2005; Araneda et al., 2007;

Moy et al., 2008), and this limits our ability to

understand and predict future global climate changes

(Villalba, 1990; Bradley & Jones, 1993; Nesje &

Dahl, 2003).

Several different approaches have been used to try

to identify the LIA in the Southern Hemisphere.

Thompson et al. (1986) used the Quelccaya ice core

(Southern Peru) to place the LIA between ca. AD

1530 and 1900, with a peak around AD 1800–1820.

Villalba (1994) relied on tree-ring records from

Northern Patagonia to identify two cold periods

(AD 900–1070 and 1270–1660) either side of a warm

period (AD 1080–1250) that was coincident with the

MWP.

Cioccale (1999) used documentary data to deter-

mine that the LIA in central Argentina comprised two

cold pulses interrupted by an intermediate period.

The first pulse lasted from the beginning of the

fifteenth to the end of the sixteenth century, and the

second and most important pulse from the early

eighteenth until the start of the nineteenth century.

Kreutz et al. (1997) found evidence of a synchronous

onset of the LIA when comparing results from studies

of West Antarctic ice cores with the Greenland Ice

Sheet Project Two (GISP2).

In Chile, Lamy et al. (2001) used humidity proxies

in marine sediments to show the existence of the

MWP and LIA in southern Chile (41�S). Bertrand

et al. (2005) studied sedimentological proxies in Lake

Puyehue (40�S) to reveal wet climate conditions

between AD 1490 and 1700; they associated these

with the onset of the European LIA. Araneda et al.

(2007) analyzed historical data and found evidence of

a cold period lasting from AD 1766 to 1898 that

peaked between 1857 and 1871; this would have been

the last pulse of the LIA in Northern Patagonia.

However, these authors did not record a warm event

corresponding to the MWP. Recently, Moy et al.

(2008) used stable isotopes to identify two periods of

208 Hydrobiologia (2010) 648:207–225

123

enhanced evaporation (900–550 and 400–50 cal yr

BP), the first one coincident with the MWP and the

second broadly agreeing with the timing of the LIA.

As indicated above, most information about the

occurrence of LIA and MWP in Chile stems from the

southern part of the country, with relatively few

studies undertaken in central Chile. Furthermore,

most of these studies relied on a single proxy.

Accordingly, the aim of this research was to assess

the occurrence, timing, and expression of the LIA and

MWP in the Andes of central Chile using a multipr-

oxy approach including diatoms, pollen, chironomids,

and chemical parameters in the sediment record of

Lake Laja.

Materials and methods

Study area

Lake Laja (36�540S; 71�050W; Fig. 1) is located at

1360 m a.s.l. in the Chilean Andes. It has a maximum

depth of 135 m, a surface area of 87 km2, an approx-

imate length of 35 km from north (Punta Chillan) to

south (El Pino Bay), and a maximum width of 7 km in

the area of Antuco Volcano. The lake was created by a

volcanic avalanche that dammed the natural valley of

the Laja River around 9700 ± 600 yr BP (Thiele et al.,

1998). The average annual precipitation is 2172 mm,

occurring mostly between May and August (fall-

winter) and falling mainly as snow. The driest months

are January, February, and March (summer). The

lowest temperatures occur in June (on average,

-0.3�C), and the highest in January (on average,

13.6�C, Torres et al., 2008).

Current vegetation in the Lake Laja basin is

composed of Austrocedrus chilensis (D. Don) Pic.

Ser. et Bizz, Lomatia hirsuta (Lam.) Diels ex Macbr,

Maytenus boaria Mol., Schinus polygamus (Cav.)

Cabrera, Colletia ulicina Gill. et Hook., Aristotelia

chilensis (Mol.) Stuntz, Orites myrtoidea (Poepp. and

Endl.) Benth. and Hooker, and Ephedra chilensis K.

Presl (Donoso, 1982; Hoffmann, 1982; Veblen &

Schlegel, 1982; Torres et al., 2008).

Sampling and sedimentological analyses

Before selecting Lake Laja as the sampling site,

bathymetric and seismological studies were

performed using a 300 J sparker equipment from

the Renard Center of Marine Geology, University of

Ghent, Belgium. This allowed us to select an

appropriate site for sediment coring. A 522-cm core

was taken from the deepest part of the lake (Fig. 1)

using an Uwitec platform, but only the section

covering the last 2000 years was analyzed (402 cm).

Grain size measurements, performed on bulk

sediment using a laser diffraction particle analyzer

Malvern Mastersizer 2000, detected a range between

0.02 and 2000 lm. Samples were introduced into a

100-ml deionized water tank free of additive disper-

sants, mixed with a 2000 rpm stirrer, and crumbled

with ultrasonic waves. The grain size was classified

according to Wentworth (1922). Organic matter and

carbonates were determined by loss on ignition. For

this, approximately 1 g of sediment was dried for

24 h at 105�C and then was weighed after heating to

550�C to determine carbon content (Boyle, 2002).

After this, the samples were heated to 1000�C for 2 h

and then weighed again in order to obtain the

carbonates present in the sediment.

Biogenic silica was determined in two stages.

First, the organic matter was eliminated and the

compound was extracted from the sediment matrix

Fig. 1 Study area showing the core sampling site and the

bathymetric map of Lake Laja

Hydrobiologia (2010) 648:207–225 209

123

with chlorhydric acid 1 N, hydrogen peroxide 10%,

and sodium carbonate 2 M. Then, the dissolved silica

was quantified according to the modified Mortlock &

Froelich (1989) molybdenum blue method, determin-

ing absorbance with molecular spectrophotometry at

810 nm wavelength.

Samples for determination of total phosphorous

were prepared with sulfuric acid 30% and potassium

peroxydisulfate solution. In this determination, PO4

was quantified with the colorimetric molybdenum

blue method (Jackson, 1964) using a molecular

absorption spectrophotometer at 890 nm wavelength.

Geochronology

The first 26 cm of the sediment core was analyzed for210Pb and 137Cs activity; ages were determined using

the Constant Rate of Supply model (CRS; Appleby &

Oldfield, 1978). In order to date the entire sediment

sequence from Lake Laja, four samples were ana-

lyzed by AMS radiocarbon dating and the calibrated

ages were obtained using the program CALIB 5.0.2

(McCormac et al., 2004). A detailed description of

the dating models applied to 210Pb and AMS dates

can be found in Torres et al. (2008).

Biological proxies

Diatoms

Diatoms were analyzed according to Battarbee

(1986), 0.1 g of dry sediment was oxidised with

H2O2 and permanent slides were mounted with Hyrax

resin (I.R. = 1.7). The diatom concentration was

estimated by adding a microsphere solution to the

samples (Battarbee & Kneen, 1982), and about 500–

600 diatom valves were counted and identified for

each sample using standard taxonomies (Rivera,

1970, 1974; Rivera et al., 1973, 1982; Krammer &

Lange-Bertalot, 1991–2000).

Chironomids

The chironomid analysis consisted of deflocculating

4 ml of wet sediment in KOH 10% for 15 min at

70�C and passing it through a 90-lm sieve. The

remains were then transferred to a Bogorov counting

tray where head capsules were picked out with

entomological forceps. Each head capsule was

dehydrated in 80 and 100% ethanol and then mounted

in Euparal, ventral side up. Chironomids were

identified using a Zeiss microscope at magnifications

of 25, 40, or 1009, and the keys of Hofmann (1971),

Rieradevall & Brooks (2001), Wiederholm (1983),

Epler (2001), and Paggi (2001).

Pollen

The pollen analysis was performed following Erdt-

man (1960). The sediment samples were processed

using concentrated HF and acetolysis. Permanent

samples were mounted in Gelatin–Glycerin, and 350

pollen grains were counted for each centimeter.

Pollen was identified following Marticorena (1968)

and Heusser (1971). The results are shown in

percentage diagrams; the terrestrial pollen estimate

includes arboreal and non-arboreal taxa, excluding

fern spores.

Statistical methods

To distinguish different associations along the profile,

a stratigraphically constrained sum-of-squares cluster

analysis (CONISS) was applied to the percentage

values of the different proxies using Tilia and Tilia

Graph (Grimm, 1987) programs. Diagrams were also

constructed using Tilia, Tilia Graph (Grimm, 1991).

A detrended correspondence analysis (DCA) was also

applied to the chironomid, pollen, and diatom

assemblages in order to assess the compositional

structure and taxa turnover throughout the profile

(Birks, 1998; Langdon et al., 2004) with the program

R and the vegan and rioja packages (Juggins, 2009;

Oksanen et al., 2005).

Results

Geochronology

The activity of 210Pb in the upper section of Lake

Laja was discussed thoroughly in Quiroz et al.

(2005), but, in general terms, the activity of unsup-

ported 210Pb shows important fluctuations along the

sediment profile. It was low in the upper part, had an

important peak around 5 cm, and then generally

tended to decline toward the deepest layers of the

sediment. Despite these important fluctuations, it was

210 Hydrobiologia (2010) 648:207–225

123

possible to use the CRS model to obtain a chronol-

ogy. The estimate of 210Pb dating and the peak of137Cs activity in Lake Laja correlated well, confirm-

ing, to a degree, the dating model. A detailed

description of 210Pb dating is founded elsewhere

(Torres et al., 2008).

Radiocarbon ages were calibrated to calendar

years using Calib 4.2 (Stuiver & Reimer, 1993). A

chronological model was developed for both 210Pb

and 14C dating by using a second-degree polynomial

equation (Villa-Martınez et al., 2004) in order to

determine calendar ages in different stratigraphic

levels along the core (Fig. 2). A detailed description

of radiocarbon dating is presented in Torres et al.

(2008).

Stratigraphy and physical–chemical analyses

The strata in the 402 cm sediment column considered

were highly different in terms of their composition

and granulometric characteristics (Fig. 3). From the

surface to 125 cm depth, the predominant fine-

sediment stratum was interrupted by tephra deposits

of various magnitudes. A tephra stratum followed

from 130 to 179 cm, after which the proportion of silt

increased. Another layer of tephra was found between

231 and 253 cm, followed by a stratum of vegetal

remains from 253 to 259 cm. A specific microsonde

analysis (Cameca SX50) performed in the Centre

d’Analyse par Microsonde pour les Sciences de

la Terre, Louvain-la-Neuve University, Belgium

(CAMST), indicated that the tephra contained a high

proportion of basaltic andesite (54.3% SiO2 and 4.3%

Na2O ? K2O), practically identical in composition to

that of the tephra reported for Lake Galletue (Urrutia

et al., 2007).

Physical–chemical parameters analyzed in the

sediment column are given in Fig. 4. Some param-

eters (e.g., organic content, biogenic silica, CO3, total

phosphorous) are clearly related to the stratigraphy of

the core. The organic content fluctuated considerably

(Fig. 4), from more than 10% in some levels (cm 325

and 255) to half that within the tephra layer (5%).

Despite these variations, organic matter clearly

increased in the last part of the sedimentary column,

which represents the last 500 years.

Carbonates were closely related to the organic

matter content, showing the same behavior during

tephra deposits and increasing toward the surface of

the column. The concentration of biogenic silica in

the sediment column also exhibited this tendency

(Fig. 4). This parameter was relatively constant from

the deepest part of the core to 100 cm depth, where

the concentration was at its lowest. From 100 cm to

the surface, although some fluctuations were regis-

tered, biogenic silica clearly increased. The same

trend was observed for phosphorous.

Diatom assemblages

A total of 45 diatom taxa were identified in the Lake

Laja sediment core. Most of these were well preserved

and could be identified to species level, with some

exceptions that were identified to the genus level.

The most abundant species (more than 10% relative

abundance) were Staurosira construens Ehrenberg,

Staurosira pinnata Ehrenberg, Cyclotella aff. glomer-

ata, Cyclotella stelligera (Cleve & Grunow) Van

Heurck, Aulacoseira grannulata (Ehrenberg) Simon-

sen, Aulacoseira distans (Ehrenberg) Simonsen, and

Asterionella formosa Hassal.

The cluster analysis performed with the diatom

assemblages defined four zones in the sediment

column: Zone 1 (402–179 cm), Zone 2 (179–

91 cm), Zone 3 (91–25 cm), and Zone 4 (25–0 cm)

(Fig. 5). These represented the main environmental

changes recorded in the lacustrine system over the

last millennia.

Fig. 2 Chronologic model for the Lake Laja record, combin-

ing both 210Pb and 14C dating by using a second-degree

polynomial equation

Hydrobiologia (2010) 648:207–225 211

123

Zone D-I (402–179 cm)

The diatom assemblages are dominated by Staurosira

pinnata, S. construens, and C. stelligera (Fig. 5), with

maximum abundances of 39, 33, and 30%, respec-

tively. Diatom valves were absent in the sediment

between 253 and 231 and 299–296 cm; these strata

correspond to tephra deposits. Diatom concentrations

fluctuate between 3.13 9 105 and 1.72 9 106 valves

per gram of dry sediment.

Zone D-II (175–91 cm)

This zone is dominated by the same species as D-I

and a thick layer of tephra lies between 179 and

130 cm. The total diatom concentration increases

immediately after the tephra layer and peaked at the

end of this zone. The diatom concentration in this

section ranges from 3.16 9 105 to 4.42 9 106 valves

per gram of dry sediment.

Zone D-III (91–25 cm)

The diatom composition changes significantly within

this zone, which is dominated by the centric species

C. stelligera and C. aff. glomerata, with maximum

abundances of 60% and 50%, respectively (Fig. 5).

The abundance of the Staurosira genera declines

drastically, and Aulacoseira distans appears in the

sediment record for the first time, reaching a max-

imum abundance of 11.6%. The total diatom con-

centration continues to increase, fluctuating between

1.32 9 106 and 5.83 9 106 valves per gram of dry

sediment.

Fig. 3 Stratigraphy and grain size (Phi) of the sedimentary column

212 Hydrobiologia (2010) 648:207–225

123

Zone D-IV (25–0 cm)

In this zone diatom assemblages change significantly.

First, Aulacoseira distans (80.6%) and then Asterio-

nella formosa (85.1%) displace the dominant species

from Zone D-III. Abundances of A. granulata

(23.9%) remain high (Fig. 5). C. aff. glomerata and

C. stelligera abundances decrease sharply, and the

former returns to the abundance levels found in zones

D-I and D-II. Synedra species (S. rumpens (Kutzing)

Carlson, S. ulna (Nitzsch) Lange-Bertalot y S. ulna

var. acus (Kutzing) Lange-Bertalot), initially very

scarce, increases during this period. Diatom concen-

trations which peaked at 8.36 9 106 valves per gram

of dry sediment at the beginning of this zone start to

decrease from 15 cm upwards and reached

9.06 9 105 valves per gram of dry sediment at the

surface.

Fig. 4 Physical–chemical

characterization of the

sediment column. The

major sand layers are also

shown (light-grey color)

Fig. 5 Diatom stratigraphy of Lake Laja showing the dominant taxa, abundances (%), total concentration, and biogenic silica in the

different zones identified by the cluster analysis

Hydrobiologia (2010) 648:207–225 213

123

Chironomid assemblages

A total of 5599 chironomid head capsules from Lake

Laja were sorted into 30 different taxa. The most

abundant subfamily was Chironominae (54.2%),

composed of Tanytarsini (30%) and Chironomini

(24%), followed by Orthocladiinae (23.4%), Tany-

podinae (19.7%), and finally Podonominae and

Diamesinae (both at 1.4%); unidentified remains

made up 4% of the total. Two morphotypes of the

Tanytarsini tribe were identified according to the

shape of their antennal pedestal (types A and B).

Type A was defined by a pedestal with a short,

rounded spur, whereas type B presented a more

pronounced spur, generally at an obtuse angle with

respect to the antennal pedestal. Figure 6 shows the

chironomid assemblages of Lake Laja. The CONISS

analysis identified four zones in the assemblage: Zone

C-I (402–248 cm), Zone C-II (247–180 cm), Zone C-

III (180–98 cm), and Zone C-IV (98–0 cm).

Zone C-I (402–175 cm)

Ablabesmyia abundances were high during this

period, especially toward the beginning of the zone,

exceeding 20% total abundance before declining

toward the end of the zone. Cricotopus/Orthocladius

and Dicrotendipes taxa were very abundant exceed-

ing 18% of the total in some layers but the

Tanytarsini as a group were the most abundant in

this zone. Similarly, Podonominae abundances are

also high in this zone, and abundances of Limno-

phyes/Compterosmittia, Macropelopia, Parapsectro-

cladius, and Tanytarsina type B were important. The

number of head capsules per gram of wet sediment

remained around five.

Zone C-II (175–98 cm)

At the beginning of this zone Ablablesmyia is absent

but it increases toward the upper part of this zone.

Other taxa abundant in the previous zone such as

Cricotopus/Orthocladius and Diamesinae, decrease

to zero abundance. Toward the end of this zone other

taxa also recover (e.g., Parapsectrocladius, Tanytar-

sini group). This zone is characterized by the thickest

tephra layer of the all profile.

Zone C-III (98–20 cm)

In this zone there is a marked change in comparison

with to the previous zone (C-II), with an increase in

taxa that previously had very low abundances such as

Macropelopia and Parachironomus, which exhibit

their highest abundances throughout the profile within

this zone. Other taxa that also increase in abundance

(but with a lower magnitude) are Tribelos/Phaenop-

sectra, Podonominae, Polypedilum, Limnophyes,

Djalmabista. Clearly the increase in the abundance

of these taxa could be related to the re-establishment

of previous lake conditions after the deposition of the

tephra layer together with a possible increase in

nutrient inputs. Other taxa record a decrease in

abundance compared to the previous zone (Tanytar-

sina type B, Parapsectrocladius, Ablabesmyia).

Zone C-IV (20–0 cm)

In this zone there are important changes in taxa that

previously exhibited very high abundances. Macro-

pelopia and Parachironomus which had abundances

of almost 60% in the zone C-III both decreased until

absent within this zone. Other taxa decreasing to zero

abundances include Monodiamesa, Limnophyes/

Compterosmittia, Parapsectrocladius, and Djalmaba-

tista. The top of the core shows important increases in

Ablabesmyia, Apedilum, Cricotopus/Orthocladius,

and Tribelos/Phaenopsectra, the latter reaching its

highest abundance.

Pollen

The pollen assemblages were divided into four zones

(Fig. 7). Figure 7 shows the total lack of arboreal

(AP) and non-arboreal pollen (NAP) in some strata.

Those sections probably correspond to sediments that

arrived suddenly at the lake bottom (landslides/tephra

falls) triggered by seismic or volcanic events, which

would explain the absence of pollen.

Zone P-I (402–120 cm)

In this zone, the assemblages were characterized by

significant proportions of AP (mean = 60.7%), with

the most important taxa being N. dombeyi-

type (mean = 36.3%), N. obliqua-type (mean =

17.6%), and E. chilensis (mean = 21.7%). Poaceae

214 Hydrobiologia (2010) 648:207–225

123

and Austrocedrus chilensis are also present but at low

abundances. Within this zone Caryophyllaceae

reached its highest abundance throughout the profile.

In the upper part of the zone the tephra layer is

clearly distinguishable and is characterized by the

complete absence of pollen. The total pollen concen-

tration fluctuated between 93,285 pollen grains g-1 at

340 cm and 22,875 pollen grains g-1 at 234 cm.

Zone P-II (120–63 cm)

This zone was dominated by N. dombeyi-type

(mean = 37.3%), E. chilensis (mean = 23.1%), and

N. obliqua-type (mean = 12.1%). The assemblage

was dominated by AP (mean = 55.3%). This period

was characterized by a substantial rise in Poaceae

(12.9%) and a decline in E. chilensis in the upper part

of the zone. The total pollen concentration ranged

between 137,402 pollen grains g-1 (101 cm) and

17,560 pollen grains g-1 (77 cm).

Zone P-III (63–41 cm)

This zone was characterized by significant proportions

of AP (mean = 50.8%). N. dombeyi-type decreased

(mean = 32.8%) and N. obliqua-type averaged 5%.

The Poaceae increased to 20%, whereas E. chilensis

decreased to 18.2%. The arboreal taxa Austrocedrus

chilensis exhibited an increase toward the upper part of

the zone. The highest total pollen concentration value

was 39,314 pollen grains g-1 (53 cm) and the lowest

was 7,554 pollen grains g-1 (43 cm).

Zone P-IV (41–0 cm)

This zone was dominated by AP (mean = 57.3%),

specifically N. dombeyi-type (mean = 35%), N. ob-

liqua-type (mean = 12.5%), and A. chilensis

(mean = 9.4%). Poacaea and E. chilensis declined

subsequently in the upper part of the zone. Total

pollen concentration values increased with respect to

the previous zone, reaching a maximum of 184,770

pollen grains g-1 at 20 cm and a minimum of 8,347

pollen grains g-1 at 2 cm.

Discussion

Diatoms

Both diatom assemblages and the chemical parame-

ters of Lake Laja have shown important variations

Fig. 6 Chironomid stratigraphy from Lake Laja showing the abundances (%) of dominant taxa across zones identified by cluster

analysis. Te timing of the LIA is shown between dashed lines and major sand layers in light-grey

Hydrobiologia (2010) 648:207–225 215

123

over the last 2000 years. These changes may be

related to different anthropogenic and natural

forcings.

The elevated abundances of Aulacoseira distans

and Aulacoseira granulata in the upper part of the

sediment record reflect increased lake productivity in

recent decades as evidenced by the increase in S. ulna

and S. rumpens. Bracco et al. (2005) suggest that A.

distans and A. granulata appear when the trophic

condition of a lake increases (e.g., increased organic

matter or phosphorous). Furthermore, the species

belonging to this genus are considered to be oppor-

tunistic (Willen, 1991), which accounts for their

higher abundances in recent times, where important

reductions in the lake level have occurred (Mardones

& Vargas, 2005; Cruces et al., 2006), which in turn

increases the nutrient concentration. The occurrence

and increase of A. formosa and Stephanodiscus sp. in

Zone IV also indicate elevated nutrients in the lake

(e.g., phosphorous, silica), as both species have been

recorded in environments with medium to high

nutrient levels (Anderson, 1990; Alefs & Muller,

1999; Clerk et al., 2000; Merilainen et al., 2000;

Bradbury et al., 2004).

Together, the mineralogical characteristics of the

sediment and the diatom assemblages indicate that

three strata (130–179, 231–253, 296–299) with pre-

dominantly minerogenic environments were depos-

ited in a short time. These sediments would have

originated from the volcanic activity of the Antuco-

Sierra Velluda complex located to the south west of

the lake. Even thought Cruces et al. (2006) indicate

that at least 11 volcanic events have been historically

recognized for this complex over the last 400 years,

not all of these were recorded in the sediment,

probably due to variations in the wind direction at the

time of each eruption. Hence in the sediment of Lake

Laja we can identify seven volcanic events with the

thickest below the 130 cm.

The low concentrations of organic matter, carbon-

ates, and biogenic silica in the tephra layers of the

profile result from the dilution generated by the

sudden input of volcanic material. The peak in organic

content (253–259 cm) was generated by the vegeta-

tion remains in this stratum from the watershed.

The volcanic events that affected Lake Laja are

likely to have affected the diatom assemblages.

According to different authors (Hickman & Schweger,

1991; Hickman & Reasoner, 1998) both Staurosira

construens and Staurosira pinnata have been cited as

the first colonizers following strong perturbations to

lakes but are subject to changes in abundances after

volcanic events. However, unlike the findings of

Cruces et al. (2006) at Lake Galletue, the occurrence

of these both species in Lake Laja seems to be

unaffected by the volcanic eruption. Rather, the high

abundances of these benthic species in zones I and II

suggests lower nutrient levels in the water column, as

Fig. 7 Pollen percentage diagram. AP arboreal pollen, NAP non-arboreal pollen. LIA temporal window is shown by the area of

dashed lines, and major sand layers in light-grey

216 Hydrobiologia (2010) 648:207–225

123

indicated by Brugam et al. (1998) for Lake Michigan.

This idea is backed by the low phosphorous concen-

trations found in these two zones.

Climate fluctuations may also have been associ-

ated with more abundant Staurosira species. Previous

studies have indicated that such species could reflect

more humid, colder conditions (Smol, 1988; Laing

et al., 1999). However, the strong volcanic influence

on Lake Laja hampered our ability to disentangle the

response of the diatoms to climate fluctuations from

their response to other forcings (e.g., volcanic and

anthropogenic activity). Furthermore, the efficacy of

diatoms for reconstructing past temperatures is con-

founded by changes in pH or nutrients (Wolin &

Duthie, 1999). Even though volcanic eruptions can

cause pH changes in the lake due to sulfur deposition

(Birks & Lotter, 1994), in the case of Lake Laja,

diatom assemblages did not show any effect. Unfor-

tunately, no calibration data set has been developed

for pH reconstruction in Chile, and it is not possible

to infer any change in this variable. Rather, diatoms

seem to be more sensitive to nutrient fluctuations.

Changes in assemblages within Zone III resulted

mainly from changes in the water body itself,

probably due to heightened nutrient concentrations

triggered by a reduction in the lake level. This

generated a shift from the previously dominant diatom

taxa to Cyclotella, which reached its maximum

abundances in this zone. According to some studies

(Gasse et al., 1989; Marciniak, 1990; Owen et al.,

1982), changes in the planktonic/non-planktonic ratio

can be used to indicate water level fluctuations.

Benthic and epiphytic diatoms increase when lake

levels are low whereas planktonic species are more

numerous at high water levels. However, depending

on the lake bathymetry, it is possible to find an inverse

relationship between depth and planktonic species

abundance (Wolin & Duthie, 1999). In lakes with

steeply shelving shores such as Lake Laja (Melnick

et al., 2006), a drop in lake level reduces the coastal

zones and, thereby, the representation of taxa from

this type of environment (e.g., Staurosira), at the same

time augmenting planktonic species.

Finally, in the upper part of the sediment record

(Zone IV), the increase in benthic-littoral species

coincided with a decline in the abundance of

planktonic taxa. These changes are the result of

recent hydrological changes in the lake, which is now

used to generate hydroelectric power (Mardones &

Vargas, 2005). In this same period diatom species

compositions suggest a rise in lake productivity (e.g.,

A. formosa, A. granulata, and Stephanodiscus sp.), as

reflected by the increase of organic matter, phospho-

rous and biogenic silica.

Chironomids

The Lake Laja sediment record revealed a 2000-year

sequence of environmental changes that are reflected

in the chironomid assemblages. CONISS analyses

produced four zones determined largely by Tanytar-

sina A and B, Cricotopus/Orthocladius, Parachiron-

omus, Parapsectrocladius, Macropelopia, and

Ablabesmyia.

In Chile, little information is available on the

ecology of these groups (e.g., Brundin, 1966; Arenas,

1995; Andersen, 1996), reducing the utility of these

organisms for reconstructing past environmental

changes. Nevertheless, some researchers have made

paleoenvironmental inferences by transferring the

ecological typology of holarctic chironomids to local

representatives of the same genera (Verschuren &

Eggermont, 2006).

The predominance of Parachironomus in the most

recent part of the record suggests a higher trophic

status of the lake for that period. According to Lotter

et al. (1998), Parachironomus is very abundant in

eutrophic and hypereutrophic lakes in the Alps.

Toward the end of the record here, levels of

Ablabesmyia, Cricotopus/Ortocladius, and Tribelos/

Phaenopsectra are also greater. Rae (1989) found

higher proportions of Ablabesmyia and Tribelos in

fluvial environments with elevated organic loads in

Ohio (USA), whereas Brooks et al. (2001) deter-

mined a slight increase of Cricotopus toward the end

of the Betton Pool (England) record, which was

congruent with an increase of total phosphorous in

the same record. This information, combined with the

chemical parameters of the sediment (e.g., organic

matter, carbonates, and total phosphorous all increas-

ing toward the end of the core), suggest that Lake

Laja has been more productive during the last

*450 years, and particularly over the last

*80 years.

All previous climate reconstruction studies under-

taken in Chile using chironomids as a paleoclimate

proxy have focused on the Late-glacial period

(Massaferro & Brooks, 2002; Massaferro et al.,

Hydrobiologia (2010) 648:207–225 217

123

2005, 2009). Although some records extend to the

present, they lack the temporal resolution required

to analyze changes in the last millennium such as

the LIA. During the LIA, lower temperatures and

glacier advances were registered across the world

(McDermott et al., 2001; Esper et al., 2002; Soon

et al., 2003; Rabatel et al., 2005; Polissar et al., 2006)

and, despite some differences in timing, it is gener-

ally accepted that colder conditions predominated

between 1550 and 1850 in the North Atlantic region

(Jones & Bradley, 1992; Bradley, 2000).

Parachironomus and Macropelopia, which were

previously related to an increase in the lake’s trophic

status, were very abundant in Lake Laja during the

LIA. In the same period, Parapsectrocladius, Podo-

nominae, Limnophyes/Compterosmittia and to a

lesser degree, Parakiefferiella and Monodiamesa also

increased slightly. Brundin (1966), Massaferro &

Brooks (2002), and Massaferro et al. (2005) indicated

that the presence of Podonominae is associated with

cold, glacier-fed waters whereas Massaferro & Corley

(1998) and Porinchu & Cwynar (2002) noted that

Parakiefferiella is related to cold, well-oxygenated

waters. High Limnophyes abundances, associated with

a higher proportion of littoral zone, were indicative of

water level fluctuations (Hofmann, 1998). Hence, the

occurrence of this taxon within the LIA allows us to

infer colder and probably drier conditions. This

conclusion was further reinforced by the occurrence

of taxa typical of more productive conditions, since

nutrient concentrations can increase in a lake with a

lower volume, thereby making it more productive

(Massaferro & Brooks, 2002).

Although chironomids are a useful tool for recon-

structing past climate changes in the Late-glacial,

their utility as a proxy for climate changes during the

Holocene is less clear as the magnitude of compo-

sitional changes in the Holocene are not as great.

However, some studies have shown that chironomids

can also record low-amplitude climate changes.

Hence, Brooks & Birks (2001) used a chironomid-

based temperature reconstruction to indicate a cool-

ing period of about 1.5�C corresponding with to the

LIA in Scotland. Brooks & Birks (2004) also reported

evidence of environmental changes in Norway that

could be associated with the LIA. Likewise, Velle

et al. (2005) found increased abundances of typical

cold-environment taxa such as Heterotrissocladius

and Sergentia in a Holocene sediment sequence from

south central Norway, providing some evidence of

the LIA in that area.

Pollen

The palynological records from Lake Laja show

important changes during the Late Holocene. Accord-

ing to the pollen record, N. dombeyi-type, N. obliqua-

type, E. chilensis, and Poaceae were the most

important taxa over the last 2000 cal yr BP, suggest-

ing alternating drier and wetter conditions during this

period. The presence of Plantago and increasing

levels of Poaceae indicate an important human

influence during the first half of the twentieth century

(Fig. 7).

A drier period is suggested by increases of E.

chilensis, an arid-adapted species (Marticorena &

Rodrıguez, 1995; Ickert-Bond et al., 2003), around

660 years AD (Zone I). This dry period was followed

by more humid conditions between 660 AD and 1561

AD, as indicated by the rise of N. dombeyi-type and

decline of Poaceae and E. chilensis.

After 933 AD, when the MWP was evident in

Europe, the Lake Laja pollen assemblage implies the

prevalence of conditions more humid than seen at

present. However, the low pollen resolution for that

period impedes verifying the occurrence of such

humid conditions. The stratigraphy in this zone is

very diverse, possibly reflecting a greater influence of

runoff from the watershed, suggesting higher precip-

itation during this period. Unfortunately, little

research has been undertaken for this period in

central Chile, limiting potential comparisons. None-

theless, Lamy et al. (2001) reported declining rainfall

during the MWP in southern Chile and Haberzettl

et al. (2005) suggesting the same for southeastern

Patagonia.

The higher relative abundance of Poaceae and E.

chilensis (the latter at the beginning of Zone II)

between 1450 and 1766 AD and the pollen concen-

trations of Chenopodiaceae and Asteraceae subf.

asteroidae suggest the beginning of a drier period that

peaked from 1766 to 1894 AD (Zone III and the

beginning of zone IV), with a concomitant decrease

in N. dombeyi-type and N. obliqua-type.

The pollen of N. obliqua-type is associated with

humid conditions (Rodrıguez et al., 1983). When

taken together with the tree growth of N. pumilio in

the Lake Laja watershed (a species that forms part of

218 Hydrobiologia (2010) 648:207–225

123

N. dombeyi-type), this record shows a positive

correlation with annual precipitation under a Medi-

terranean-type climate where water availability is a

major limiting factor (Lara et al., 2001, 2005).

The occurrence of Poaceae is normally associated

with human activity but could also be related to lower

precipitation (Bush, 2000). Given the very low

human population in the area at that time, we believe

that Poaceae responded more to changes in precip-

itation than to anthropogenic factors. Therefore, and

according to previous authors, we think that lower

precipitation levels drove an increase in Poaceae

between 1766 and 1894 AD, while the peak between

1900 and 1950 could be associated with anthropo-

genic influence in Lake Laja. A detailed description

of the human influence in the study area is found

elsewhere (Torres et al., 2008).

This dry period may be associated with the

manifestation of the LIA in central Chile. This was

not a single event but rather was characterised by

humid periods. Evidence of this comes from the high

prevalence of Ephedra chilensis, an arid-adapted

species (Marticorena & Rodrıguez, 1995; Ickert-

Bond et al., 2003), interspersed with a few peaks of

N. dombeyi-type (which according the current vege-

tation, would correspond to N. pumilio), an indicator

species of higher precipitation (Lara et al., 2001,

2005).

According to Jenny et al. (2002), the LIA in

central Chile (32�–38�S) began with a humid period

(1300–1700 AD) and ended with a drier period

(1700–1850 AD). Bertrand et al. (2005), working at

Lake Puyehue (40�S), highlighted a humid period

(1490–1700 AD) followed by a drier period (1700–

1900 AD). Only Lamy et al. (2001) found a different

trend, with heightened rainfall during the LIA around

41�S. These findings are in rough agreement with the

drier conditions reported in our results for the LIA in

Lake Laja, which occurred between ca. 1550 and

1900 AD. As stated earlier, the LIA was probably not

a single homogenous event, but comprised different

cold pulses that were not necessarily synchronous

(Ciocale, 1999; Araneda et al., 2009).

The most important changes in Zone IV could be

related to human impacts. NAP, specifically from

Poaceae and E. chilensis, increased at the beginning

of this period and then declined in the upper part

of this zone (1968–2001 AD). N. dombeyi-type and

N. obliqua-type increased simultaneously, whereas

Plantago appeared in the core for the first time

(Fig. 7).

Between 1938 and 1968, the human impacts on the

Lake Laja basin expanded as a result of hydropower

plant construction, tourism, and recreational activities

(Nardini & Montoya, 1993). We think that the

increase in Poaceae from 1938 to 1968 resulted

mainly from intensified human activity. During these

years, E. chilensis was rare, and it is possible that the

species cannot survive in disturbed ecosystems

(Torres et al., 2008). Finally, the palynological

changes detected in the upper part of the core may

be a response to reduced anthropogenic intervention

related to the end of hydropower plant construction

(Nardini & Montoya, 1993) and the founding of

Parque Nacional Laguna del Laja in 1957.

Combined analysis

One advantage of multiproxy reconstructions is they

offer the possibility of determining inherent climate

variation (Battarbee, 2000; Larocque & Bigler,

2004). Moreover, a multiproxy approach decreases

the introduction of errors associated with the ‘‘over-

simplification’’ of relationships between environmen-

tal variables and organisms, which occurs with

reconstructions based on just one proxy (Ammann

et al., 2000; Birks et al., 2000; Bigler et al., 2002).

All three proxies underwent important changes, as

seen in the record of the last 2000 years from Lake

Laja. During the first thousand years of the record

where the volcanic sediments are more evident, the

three proxies do not show any major change as

reflected in the DCA scores; however, this should be

tested in a quantitative way in a further study (e.g.,

Lotter & Birks, 1993). Conversely, in the last

millennium significant changes occurred in the three

proxies (DCA scores) that coincide with the mani-

festation of LIA in Europe (Fig. 8), while in the last

century, important changes are recognized mostly in

diatoms, which reflects heightened anthropogenic

activity in the basin.

The first part of the record coincides with the

timing of the MWP (*800–1300 AD; Trouet et al.,

2009). However, no responses from the three proxies

were observed that might indicate climatic conditions

associated with the MWP. In Chile, few studies have

described the occurrence of the MWP, and the

existence of this period as a global phenomenon

Hydrobiologia (2010) 648:207–225 219

123

generates some controversy (Bradley et al., 2003).

Nonetheless, Villalba (1994) recognized a warm

period in central Chile that coincided with the

MWP in Europe. Later, Ortlieb et al. (2000), working

in Mejillones Bay, suggested an increase in the sea

surface temperature during the period of the MWP,

and Lamy et al. (2001), working with marine

sediments from southern Chile (41�S), indicated less

humid conditions for this same period.

The lack of evidence for the MWP in Lake Laja

could be due to the fact that this event was not

manifested globally, as some authors indicate (Brad-

ley et al., 2003), or that the effects did not impact on

the organisms studied.

Urrutia et al. (2007) reported a short-lived volcanic

influence on Lake Galletue that allowed a rapid re-

establishment of pre-tephra conditions. However, the

input of volcanic ash to Lake Laja, which is very

close to an active volcanic centre, was much more

intense and constant over time. Lake Galletue is

located approximately 42 km from Lonquimay Vol-

cano and 37 km from Llaima Volcano, whereas Lake

Laja is some 4 km from the cone of Antuco Volcano

and incursions of lava flows into the southern sector

of the lake are evident. Thus, it was expected that the

organisms of Lake Laja would show a marked

response to the volcanic activity. However, the

evidence compiled so far indicated a response very

similar to Lake Galletue.

During the LIA, in contrast to the MWP, important

changes occurred in the three proxies. The planktonic

diatom taxa Cyclotella stelligera and Cyclotella aff.

glomerata increased and the Fragilaria species

decreased drastically. This change could be attributed

to a drop in the lake level caused by colder-drier

conditions that generated an increment in the

Fig. 8 Diagram of the first

DCA score of chironomids,

diatoms, and pollen along

the sediment column of

Lake Laja

220 Hydrobiologia (2010) 648:207–225

123

planktonic species. Normally, a drop in the lake level

would cause these types of diatoms to decline in

abundance and benthic species to increase (Fritz,

1990; Moser et al. 1996; Caballero & Ortega, 1998;

Fritz et al., 1999; Telford et al., 1999). Such changes

have been reported at millennial-decadal scales,

inferring changes in lake levels when planktonic

communities are replaced by benthic and epiphytic

diatoms (Barker et al., 2002; Cumming et al., 2002;

Xue et al., 2003). But as was stated above, the

opposite response seems to had occurred lake Laja as

a result of the morphometry of the lake basin.

Chironomids also showed important changes dur-

ing the LIA, including increased abundances of

Macropelopia and Parachironomus taxa, the latter

being an indicator of greater trophic conditions.

Increased levels of Podonominae and Limnophyes

were also recorded, related to cold temperatures and

lower lake levels, respectively. These variations

denote a prevalence of colder, drier conditions over

the period of the LIA in Lake Laja. This is reflected

by the abundance of taxa which favor higher nutrient

levels, since lower lake levels tend to concentrate

nutrients.

The occurrence of cold, dry conditions was also

corroborated by the pollen record. More pollen was

found from Poaceae, E. chilensis, Chenopodiaceae,

and Asteraceae subf. Asteroidae, all taxa that indicate

dry conditions. At the same time, N. dombeyi-type

and N. obliqua-type decreased; according to Lara

et al. (2001, 2005), these taxa flourish under humid

conditions, again supporting the likelihood that the

lake was characterised by cold-dry conditions during

this period.

Finally, the effect of anthropogenic activity on the

lacustrine ecosystem and its watershed was observed

in the upper layers of the sediment column. Changes

recorded in the aquatic communities (diatoms and

chironomids) after the cold-dry condition (LIA

manifestation) were indicative of an increase in the

productivity of the lake associated with the hydro-

logical changes caused by the use of the lake for the

generation of hydroelectric energy (Cruces et al.,

2006). Human activity was also reflected in the pollen

profile in the form of increased NAP, specifically

Poaceae, and the appearance of Plantago in the first

strata along with a decline in the pollen of N.

dombeyi-type and E. chilensis. The latter species is

sensitive to anthropogenic intervention and has low

survival rates in heavily perturbed ecosystems (Tor-

res et al., 2008).

Conclusions

– The sedimentary record shows a clear volcanic

influence, most evident in the first 1000 years;

however, the organisms studied did not show a

marked response. Thus, the volcanic events are

likely to have had a more instantaneous rather

than long-term effects.

– No evidence of the occurrence of the MWP was

found, either because the event was not global, or

its magnitude was not sufficient regionally to

cause a change. Therefore, it is not possible to

prove or reject reliably the existence of this event

in the central Andes.

– The three proxies indicate the occurrence of a

cold, dry event in Lake Laja centered around

1550–1900 AD, coinciding roughly with the

occurrence of the LIA in Europe. Pollen and

chironomids were the most sensitive proxies to

this change; the diatom response was less direct

and associated more with changes in the lake

level and/or its trophic state.

– In the most recent part of the record, both diatoms

and chironomids indicate an increased trophic state

of the lake due to greater anthropogenic interven-

tion in the watershed. This is further reflected,

although to a lesser degree, in the pollen profile.

– Finally, we note the importance of using a

multiproxy approach to study past environmental

conditions (climate and others), since this dimin-

ished the bias associated with ‘‘oversimplifying’’

the relationships between the environmental

variables and the organisms.

Acknowledgments This research was funded by Fondecyt

projects No 1070508, No 1080294, and 11080158, and WBI

Wallonie-Chile cooperation project, FNRS, and ULg fundings.

References

Alefs, J. & J. Muller, 1999. Differences in the eutrophication

dynamics of Ammersee and Starnberger See (Southern

Germany), reflected by the diatom succession in varve-

dated sediments. Journal of Paleolimnology 21: 395–407.

Hydrobiologia (2010) 648:207–225 221

123

Ammann, B., H. Birks, S. Brooks, U. Eicher, U. von Grafen-

stein, W. Hofmann, G. Lemdahl, J. Schwander, K. To-

bolski & L. Wick, 2000. Quantification of biotic responses

to rapid climatic changes around the Younger Dryas – a

synthesis. Palaeogeography Palaeoclimatology Palaeoe-

cology 159: 313–347.

Andersen, T., 1996. A new species of Monodiamesa Kieffer,

1922 from Southern Chile (Diptera: Chironomidae: Pro-

diamesinae). Revista Chilena de Entomologıa 23: 43–49.

Anderson, N., 1990. The biostratigraphy and taxonomy of

small Stephanodiscus and Cyclostephanos species (Bac-

illariophyceae) in a eutrophic lake and their ecological

implications. British Phycological Journal 25: 217–235.

Appleby, P. G. & F. Oldfield, 1978. The calculation of lead-

210 dates assuming a constant rate of supply of unsup-

ported 210Pb to the sediment. Catena 5: 1–8.

Araneda, A., F. Torrejon, M. Aguayo, L. Torres, F. Cruces, M.

Cisternas & R. Urrutia, 2007. Historical record of San

Rafael glacier advances (North Patagonian Icefield):

another clue for Little Ice Age timing in Souther Chile.

The Holocene 17: 987–998.

Araneda, A., F. Torrejon, M. Aguayo, I. Alvial, C. Mendoza &

R. Urrutia, 2009. Historical records of Cipreses glacier

(34�S): combining documentary-inferred ‘Little Ice Age’

evidence from Southern and Central Chile. The Holocene

19: 1173–1183.

Arenas, J., 1995. Composicion y distribucion del macrozoo-

bentos del curso principal del rıo Biobıo, Chile. Medio

Ambiente 12: 39–50.

Barker, P., R. Telford, F. Gasse & R. Thevenon, 2002. Late

Pleistocene and Holocene palaeohydrology of Lake Ru-

kwa, Tanzania, inferred from diatom analysis. Palaeoge-

ography, Palaeoclimatology, Palaeoecology 187: 295–

305.

Battarbee, R. W., 1986. Diatom analysis. In Berglund, B. E.

(ed.), Handbook of Holocene Palaeoecology and Pala-

eohydrology. Wiley, Chichester: 527–570.

Battarbee, R. W., 2000. Palaeolimnological approaches to

climate change, with special regard to the biological

record. Quaternary Science Reviews 19: 107–124.

Battarbee, R. & M. Kneen, 1982. The use of electronically

counted microspheres in absolute diatom analysis. Lim-

nology and Oceanography 27: 184–188.

Bertrand, S., X. Boes, J. Castiaux, F. Charlet, R. Urrutia, C.

Espinoza, G. Lepoint, B. Charlier & N. Fagel, 2005.

Temporal evolution of sediment supply in Lago Puyehue

(Southern Chile) during the last 600 yr and its climatic

significance. Quaternary Research 64: 163–175.

Bigler, C., I. Larocque, S. Peglar, H. Birks & R. Hall, 2002.

Quantitative multiproxy assessment of long-term patterns

of Holocene environmental change from a small lake

near Abisko, northern Sweden. The Holocene 12: 481–

496.

Birks, H. J. B., 1998. Numerical tools in palaeolimnology –

progress, potentialities, and problems. Journal of Paleo-

limnology 20: 307–332.

Birks, H. J. B. & A. F. Lotter, 1994. The impact of the Laacher

See Volcano (11, 000 yr B.P.) on terrestrial vegetation

and diatoms. Journal of Paleolimnology 11: 313–322.

Birks, H. J. B., R. Battarbee & H. Birks, 2000. The develop-

ment of the aquatic ecosystem at Krakenes Lake, western

Norway, during the late glacial and early Holocene – a

synthesis. Journal of Paleolimnology 23: 91–114.

Boyle, J. F., 2002. Mineralogical and geochemical indicator

techniques. In Last, W. & J. P. Smol (eds), Tracking

Environmental Change Using Lake Sediments. Springer-

Verlag, New York, USA: 83–141.

Bracco, R., H. Inda, L. del Puerto, C. Castineira, P. Sprech-

mann & F. Garcıa-Rodrıguez, 2005. Relationships

between Holocene sea-level variations, trophic develop-

ment, and climatic change in Negra Lagoon, Southern

Uruguay. Journal of Paleolimnology 33: 253–263.

Bradbury, J., S. Colman & R. Reynolds, 2004. The history of

recent limnological changes and human impact on Upper

Klamath Lake, Oregon. Journal of Paleolimnology 31:

151–165.

Bradley, R. S., 2000. Climate paradigms for the last millen-

nium. PAGES Newsletters 8: 2–3.

Bradley, R. & P. D. Jones, 1993. Little Ice Age summer

temperature variations: their nature and relevance to

recent global warming. The Holocene 3: 367–376.

Bradley, R., M. Hughes & H. Diaz, 2003. Climate in medieval

time. Science 302: 404–405.

Brooks, S. J. & H. J. B. Birks, 2001. Chironomid-inferred

temperatures from Lateglacial and Holocene sites in

north-west Europe: progress and problems. Quaternary

Science Reviews 20: 1723–1741.

Brooks, S. J. & H. J. B. Birks, 2004. The dynamics of Chiro-

nomidae (Insecta: Diptera) assemblages in response to

environmental change during the past 700 years on

Svalbard. Journal of Paleolimnology 31: 483–498.

Brooks, S. J., H. Bennion & H. J. B. Birks, 2001. Tracking lake

trophic history with a chironomid-total phosphorus

inference model. Freshwater Biology 46: 513–533.

Brugam, R. B., K. McKeever & L. Kolesa, 1998. A diatom-

inferred water depth reconstruction for an Upper Peninsula,

Michigan, lake. Journal of Paleolimnology 20: 267–276.

Brundin, L., 1966. Transarctic relationships and their signifi-

cance, as evidenced by chironomid midges, with a

monograph of the subfamilies Podonominae and Aph-

roteniinae and the austral Heptagyiae. Kung Svenska

Vetenskapsakademiens Handinglar 11: 1–472.

Bush, M., 2000. Deriving response matrices from Central

American modern pollen rain. Quaternary Research 54:

132–143.

Caballero, M. & B. Ortega, 1998. Lake levels since about

40000 years ago at Lake Chalco, near Mexico City.

Quaternary Research 50: 69–79.

Cioccale, M. A., 1999. Climatic fluctuations in the central

region of Argentina in the last 1000 years. QuaternaryInternational 62: 35–47.

Clerk, S., R. Hall, R. Quinlan & J. Smol, 2000. Quantitative

inferences of past hypolimnetic anoxia and nutrient levels

from a Canadian Precambrian Shield lake. Journal of

Paleolimnology 23: 319–336.

Crowley, T. J. & T. S. Lowery, 2000. How warm was the

medieval warm period? Ambio 29: 51–54.

Cruces, F., R. Urrutia, O. Parra, A. Araneda, H. Treutler, S.

Bertrand, N. Fagel, L. Torres, R. Barra & L. Chirinos,

2006. Changes in diatom assemblages in an Andean lake

in response to a recent volcanic event. Archiv fur Hy-

drobiologie 165: 23–35.

222 Hydrobiologia (2010) 648:207–225

123

Cumming, B., K. Laird, J. Bennett, J. Smol & A. Salomon,

2002. Persistent millennial-scale shifts in moisture

regimes in western Canada during the past six millennia.

Proceedings of the National Academy of Sciences of the

United States of America 25: 16117–16121.

Donoso, C., 1982. Resena ecologica de los bosques mediter-

raneos de Chile. Bosque 4: 117–146.

Epler, J. H., 2001. Identification Manual for the larval Chiro-

nomidae (Diptera) of North and South Carolina. A Guide

to the Taxonomy of the Midges of the Southeastern

United States, Including Florida. North Carolina Depart-

ment of Environment and Natural Resources, Florida: 526

pp.

Erdtman, G., 1960. The acetolysis method. A revised

description. Svensk Botanisk Tidskrift 54: 561–564.

Esper, J., E. Cook & F. H. Schweingruber, 2002. Low-fre-

quency signals in long tree-ring chronologies for recon-

structing past temperature variability. Science 295: 2250–

2253.

Esper, J., R. J. S. Wilson, D. C. Frank, A. Moberg, H. Wanner

& J. Luterbacher, 2005. Climate: past ranges and future

changes. Quaternary Science Reviews 24: 2164–2166.

Fritz, S., 1990. Twentieth-century salinity and water-level

fluctuations in Devil’s Lake, North Dakota, test of a dia-

tom-based transfer function. Limnology and Oceanogra-

phy 35: 1771–1781.

Fritz, S., B. Cumming, F. Gasse & K. Laird, 1999. Diatoms as

indicators of hydrologic and climatic change in saline

lakes. In Stoermer, E. F. & J. P. Smol (eds), The Diatoms:

Applications to the Environmental and Earth Sciences.

Cambridge University Press, Cambridge, MA: 41–72.

Gasse, F., V. Ledee, M. Massault & J.-C. Fontes, 1989. Water

level fluctuations of Lake Tanganika in phase with oce-

anic changes during the last glaciation and deglaciation.

Nature 342: 57–59.

Goosse, H., H. Renssen, A. Timmermann & R. Bradley, 2005.

Internal and forced climate variability during the last

millennium: a model-data comparison using ensemble

simulations. Quaternary Science Reviews 24: 1345–1360.

Grimm, E., 1987. CONISS: a Fortran 77 program for strati-

graphically constrained cluster analysis by the method of

incremental sum of squares. Computational Geosciences

13: 13–35.

Grimm, E., 1991. Tilia and Tilia Graph. Illinois State Museum,

Springfield.

Grove, J. M., 2001. The initiation of the ‘Little Ice Age’ in

regions around the North Atlantic. Climatic Change 48:

53–82.

Haberzettl, T., M. Fey, A. Lucke, N. Maidana, C. Mayr, C.

Ohlendorf, F. Schabitz, G. Schleser, M. Wille & B. Zo-

litschka, 2005. Climatically induced lake level changes

during the last two millennia as reflected in sediments of

Laguna Potrok Aike, southern Patagonia (Santa Cruz,

Argentina). Journal of Paleolimnology 33: 283–302.

Heusser, C., 1971. Pollen and Spores of Chile. Modern types of

the Pteridophytas, Gimnospermae and Angiospermae.

University of Arizona, Arizona, AZ.

Hickman, M. & M. Reasoner, 1998. Late quaternary diatom

responses to vegetation and climate change in a subalpine

lake in Banff National Park, Alberta. Journal of Paleo-

limnology 20: 253–265.

Hickman, M. & C. E. Schweger, 1991. A palaeoenvironmental

study of Fairfax Lake, a small lake situated in the Rocky

Mountain Foothills of west-central Alberta. Journal of

Paleolimnology 6: 1–15.

Hoffmann, A., 1982. Flora silvestre de Chile. Fundacion

Claudio Gay, Zona Araucana (Santiago de Chile).

Hofmann, W., 1971. Zur taxonomie und palokologie subfoss-

iler Chironomiden (Dipt.) in seesedimenten. Ergebnisse

der Limnologie 6: 1–50.

Hofmann, W., 1998. Cladocerans and chironomids as indica-

tors of lake level changes in north temperate lakes. Journal

of Paleolimnology 19: 55–62.

Huang, S. & H. N. Pollack, 1997. Late quaternary temperature

changes seen in world-wide continental heat flow mea-

surements. Geophysical Research Letters 24: 1947–1950.

Hughes, M. K. & H. F. Diaz, 1994. Was there a ‘‘Medieval

warm period’’ and if so where and when? Climatic

Change 26: 109–142.

Ickert-Bond, S., J. Skvarla & W. Chissoe, 2003. Pollen

dimorphism in Ephedra L. (Ephedraceae). Review of

Palaeobotany and Palynology 124: 325–334.

Jackson, M., 1964. Analisis quımico de suelos, Vol. 1. Edici-

ones Omega S.A., Barcelona, Espana: 367 pp.

Jenny, B., B. Valero-Garces, R. Villa-Martinez, R. Urrutia, M.

Geyh & H. Veit, 2002. Early to mid-Holocene aridity in

central Chile and the southern westerlies: the Laguna

Aculeo record (34�S). Quaternary Research 58: 160–170.

Jones, P. D. & R. S. Bradley, 1992. Climatic variations over the

last 500 years. In Bradley, R. S. & P. D. Jones (eds),

Climate since AD 1500. Routledge, London: 649–665.

Jones, P. D., T. J. Osborn & K. R. Briffa, 2001. The evolution of

climate over the last millennium. Science 292: 662–667.

Juggins, S., 2009. Rioja: an R package for the analysis of

quaternary science data, Version 0.5-3. (http://cran.r-project.org/package=rioja).

Koch, J. & R. Kilian, 2005. ‘‘Little Ice Age’’ glacier fluctua-

tions, Gran Campo Nevado, southernmost Chile. The

Holocene 15: 20–28.

Krammer, K. & H. Lange-Bertalot, 1991–2000. Bacillario-

phyceae. In Ettl, H., J. Gerloff, H. Heyning & D. Molle-

nhauser (eds) Subwasserflora von Mitteleuropa, Vol. 2.

Fischer, Stuttgart: 1–4.

Kreutz, K. J., P. A. Mayewski, L. D. Meeker, M. S. Twickler,

S. I. Whitlow & I. I. Pittalwala, 1997. Bipolar changes in

atmospheric circulation during the Little Ice Age. Science

277: 1294–1296.

Laing, T., K. Ruhland & J. Smol, 1999. Past environmental and

climate changes related to tree-line shifts inferred from

fossil diatoms from a lake near the Lena River Delta,

Siberia. The Holocene 9: 547–557.

Lamb, H. H., 1977. Climate: present, past and future, Vol. 2:

Climatic history and the future. Methuen, London.

Lamy, F., D. Hebbeln, U. Rohl & G. Wefer, 2001. Holocene

rainfall variability in southern Chile: a marine record of

latitudinal shifts of the southern westerlies. Earth and

Planetary Science Letters 185: 369–382.

Langdon, P. G., K. E. Barber & S. H. Lomas-Clarke, 2004.

Reconstructing climate and environmental change in

northern England through chironomid and pollen analy-

ses: evidence from Talkin Tarn, Cumbria. Journal of Pa-

leolimnology 32: 197–213.

Hydrobiologia (2010) 648:207–225 223

123

Lara, A., J. Aravena, R. Villalba, A. Wolodarsky-Franke, B.

Luckman & R. Wilson, 2001. Dendroclimatology of high-

elevation Nothofagus pumilio forests at their northern

distribution limit in the central Andes of Chile. Canadian

Journal of Forest Research 31: 925–936.

Lara, A., R. Villalba, A. Wolodarsky-Franke, J. Aravena, B.

Luckman & E. Cuq, 2005. Spatial and temporal variation

in Nothofagus pumilio growth at tree line along its lati-

tudinal range (35�400–55�S) in the Chilean Andes. Journal

of Biogeography 32: 879–893.

Larocque, I. C. & C. Bigler, 2004. Similarities and discrep-

ancies between chironomid- and diatom-inferred temper-

ature reconstructions through the Holocene at Lake 850,

northern Sweden. Quaternary International 122: 109–121.

Loehle, C., 2007. A 2000-year global temperature reconstruc-

tion based on non-tree-ring proxies. Energy and Envi-

ronment 18: 1049–1058.

Lotter, A. F. & H. J. B. Birks, 1993. The impact of the Laacher

See tephra on terrestrial and aquatic ecosystems in the

Black Forest, southern Germany. Journal of Quaternary

Science 8: 263–276.

Lotter, A. F., H. J. B. Birks, W. Hofmann & A. Marchetto,

1998. Modern diatom, cladocera, chironomid, and

chrysophyte cyst assemblages as quantitative indicators

for the reconstruction of past environmental conditions in

the Alps. II. Nutrients. Journal of Paleolimnology 19:

443–463.

Mann, M. E., 2002. Little ice age. In MacCracken, M. C. & J.

S. Perry (eds), The Earth System: Physical and Chemical

Dimensions of Global Environmental Change. Wiley,

New York: 504–509.

Mann, M. E., R. S. Bradley & M. K. Hughes, 1998. Global-

scale temperature patterns and climate forcing over the

past six centuries. Nature 392: 779–787.

Marciniak, B., 1990. Late glacial and Holocene diatoms in

sediments of the Bledowo Lake (Central Poland). In Si-

mola H (ed.) Proceedings of the 10th Diatom Symposium.

Koeltz Scientific Books, Koenigstein: 379–390.

Mardones, M. & J. Vargas, 2005. Efectos hidrologicos de los

usos electrico y agrıcola en la cuenca del rıo Laja (Chile

centro-sur). Revista de Geografıa Norte Grande 33: 89–

102.

Marticorena, C., 1968. Granos de polen de plantas chilenas,

Vol. 17. Gayana Botanica, Chile: 34 pp.

Marticorena, C. & R. Rodrıguez, 1995. Flora de Chile, Vol 1.

Pteridophyta-Gimnospermae. Universidad de Concepcion,

Concepcion, Chile: 351 pp.

Massaferro, J. & S. J. Brooks, 2002. Response of chironomids

to late quaternary environmental change in the Taitao

peninsula, southern Chile. Journal of Quaternary Science

17: 101–111.

Massaferro, J. & J. Corley, 1998. Environmental disturbance

and chironomid paleodiversity: 15 kyr BP of history at

Lake Mascardi, Patagonia, Argentina. Aquatic Conserva-

tion: Marine and Freshwater Ecosystems 8: 315–323.

Massaferro, J., S. J. Brooks & S. G. Haberle, 2005. The

dynamics of chironomid assemblages and vegetation

during the Late quaternary at Laguna Facil, Chonos

Archipelago, southern Chile. Quaternary Science Reviews

24: 2510–2522.

Massaferro, J., P. Moreno, G. Denton, M. Vandergoes & A.

Dieffenbacher-Krall, 2009. Chironomid and pollen evi-

dence for climate fluctuations during the Last Glacial

Termination in NW Patagonia. Quaternary Science

Reviews 28: 517–525.

McCormac, F., A. Hogg, P. Blackwell, C. Buck, T. Higham &

P. Reimer, 2004. SHCal04 southern hemisphere calibra-

tion 0–11.0 cal Kyr BP. Radiocarbon 46: 1087–1092.

McDermott, F., D. Mattey & C. Hawkesworth, 2001. Centen-

nial-scale Holocene climate variability revealed by high-

resolution speleotherm 18O record from SW Ireland.

Science 294: 1328–1331.

McKinzey, K. M., W. Lawson, D. Kelly & A. Hubbard, 2004.

A revised Little Ice Age chronology of the Franz Josef

Glacier, Westland, New Zealand. Journal of the Royal

Society of New Zealand 34: 381–394.

Melnick, D., F. Charlet, H. P. Echtler & M. De Batist, 2006.

Incipient axial collapse of the Main Cordillera and strain

partitioning gradient between the central and Patagonian

Andes, Lago Laja, Chile. Tectonics 25: 1–22.

Merilainen, J., J. Hynynen, A. Teppo, A. Palomaki, K. Gran-

berg & P. Reinikainen, 2000. Importance of diffuse

nutrient loading and lake level changes to the eutrophi-

cation of an originally oligotrophic boreal lake: a palae-

olimnological diatom and chironomid analysis. Journal of

Paleolimnology 24: 251–270.

Mortlock, R. & P. Froelich, 1989. A simple method for the

rapid determination of biogenic opal in pelagic marine

sediments. Deep Sea Research 36: 1415–1426.

Moser, K., G. MacDonald & J. Smol, 1996. Applications of

freshwater diatoms to geographical research. Progress in

Physical Geography 20: 21–52.

Moy, C., R. Dunbar, P. Moreno, J.-P. Francois, R. Villa-

Martınez, D. Mucciarone, T. Guilderson & R. Garreaud,

2008. Isotopic evidence for hydrologic change related to

the westerlies in SW Patagonia, Chile, during the last

millennium. Quternary Science Reviews 27: 1335–1349.

Nardini, A. & D. Montoya, 1993. Planteamiento de un modelo

decisional para la gestion integrada del sistema lago Laja-

rıo Laja (con respecto al proyecto ‘‘Canal Laja-Dig-

uillın’’), Vol. 3. Monografıa cientıficas, Centro Eula-

Chile.

Nesje, A. & S. O. Dahl, 2003. The ‘‘Little ice Age’’ – only

temperature? The Holocene 13: 139–145.

Ogilvie, A. E. J. & T. Jonsson, 2001. ‘Little Ice Age’ research:

a perspective from Iceland. Climate Change 48: 9–52.

Oksanen, J., R. Kindt, P. Legendre & R. B. O’Hara, 2005.

Vegan: community ecology package version 1.15-3.

http://cc.oulu.fi/*jarioksa/.

Ortlieb, L., R. Escribano, R. Follegati, O. Zuniga, I. Kong, L.

Rodriguez, J. Valdes, N. Guzman & P. Iratchet, 2000.

Recording of ocean-climate changes during the last 2,

000 years in a hypoxic marine environment off northern

Chile (23�S). Revista Chilena de Historia Natural 73(2):

221–242.

Owen, R., J. Barthelme, R. Renaut & A. Vincens, 1982. Pa-

leolimnology and archaeology of Holocene deposits

north-east of Lake Turkana, Nenya. Nature 298: 523–529.

Paggi, A., 2001. Diptera: chironomidae. In Fernandez, H. R. &

E. Domınguez (eds), Guıa para la determinacion de los

224 Hydrobiologia (2010) 648:207–225

123

artropodos bentonicos Sudamericanos. Editorial Univers-

itaria de Tucuman, Tucuman, Argentina: 167–194.

Polissar, P. J., M. B. Abbott, A. P. Wolfe, M. Bezada, V. Rull

& R. S. Bradley, 2006. Solar modulation of Little Ice Age

climate in the tropical Andes. Proceedings of the National

Academy of Sciences of the United States of America

103: 8937–8942.

Porinchu, D. F. & L. C. Cwynar, 2002. Late-Quaternary history

of midge communities and climate from a tundra site near

the lower Lena River, Northeast Siberia. Journal of Pa-

leolimnology 27: 59–69.

Quiroz, R., P. Popp, R. Urrutia, C. Bauer, A. Araneda, H.-C.

Treutler & R. Barra, 2005. PAH fluxes in the Laja Lake of

south central-Chile Andes over the last 50 years: evidence

from a dated sediment core. Science of the Total Envi-

ronment 349: 150–160.

Rabatel, A., V. Jomelli, P. Naveau, B. Francou & D. Grancher,

2005. Dating of Little Ice Age glacier fluctuations in the

tropical Andes: Charquini glaciers, Bolivia, 16�S.

Comptes Rendur Geoscience 337: 1311–1322.

Rae, J. G., 1989. Chironomid midges as indicators of organic

pollution in the Scioto River basin, Ohio. Ohio Journal of

Science 1: 5–9.

Rieradevall, M. & S. J. Brooks, 2001. An identification guide

to subfossil Tanypodinae larvae (Insecta: Diptera: Chiro-

nomidae) based on cephalic setation. Journal of Paleo-

limnology 25: 81–99.

Rivera, P., 1970. Diatomeas de los lagos Ranco, Laja y Laguna

Chica de San Pedro Vol. 20. Gayana Botanica, Chile: 23

pp.

Rivera, P., O. Parra & M. Gonzalez, 1973. Fitoplancton del

Estero Lenga, Vol. 23. Gayana Botanica, Chile: 93 pp.

Rivera, P., 1974. Diatomeas de agua dulce de Concepcion y

alrededores Vol. 28. Gayana Botanica, Chile: 134 pp.

Rivera, P., O. Parra, M. Gonzalez, V. Dellarossa & M. Orell-

ana, 1982. Manual taxonomico del fitoplancton de aguas

continentales. Editorial Universidad de Concepcion,

Concepcion: 97.

Rodrıguez, R., O. Matthei & M. Quezada, 1983. Flora Arborea

de Chile. Editorial de la Universidad de Concepcion,

Concepcion, Chile.

Smol, J., 1988. Paleoclimate proxy from freshwater arctic

diatoms. Verhandlungen der Internationalen Vereinigung

fur Limnologie 23: 837–844.

Soon, W., S. Baliunas, C. Idso, S. Idso & D. Legates, 2003.

Reconstructing climatic and environmental changes of the

past 1000 years: a reappraisal. Energy and Environment

14(2–3): 233–299.

Stuiver, M. & P. J. Reimer, 1993. Extended 14C database and

revised CALIB radiocarbon calibration program. Radio-

carbon 35: 215–230.

Telford, R. J., H. F. Lamb & U. Mohammed, 1999. Diatom-

derived paleoconductivity estimates from Lake Awassa,

Ethiopia: evidence for pulsed inflows of saline ground-

water? Journal of Paleolimnology 21: 409–421.

Thiele, R., H. Moreno, S. Elgueta, A. Lahsen, S. Rebolledo &

M. Petit-Breuilh, 1998. Evolucion geologico-geomor-

fologica cuaternaria del tramo superior del valle del rıo

Laja. Revista Geologica de Chile 25(2): 229–253.

Thompson, L. G., E. Mosley-Thompson, W. Dansgaard & P.

M. Grootes, 1986. The Little Ice Age as recorded in the

stratigraphy of the tropical Quelccaya ice cap. Science

234: 361–364.

Torres, L., O. Parra, A. Araneda, R. Urrutia, F. Cruces & L.

Chirinos, 2008. Vegetational and climatic history during

the late Holocene in Lake Laja basin (central Chile)

inferred from sedimentary pollen record. Reviews of

Palaeobotany and Palynology 149: 18–28.

Trouet, V., J. Esper, N. Graham, A. Baker, J. Scourse & D.

Frank, 2009. Persistent positive North Atlantic oscillation

mode dominated the medieval climate anomaly. Science

324: 78–80.

Urrutia, R., A. Araneda, F. Cruces, L. Torres, L. Chirinos, H.-

C. Treutler, N. Fagel, S. Bertrand, I. Alvial, R. Barra & E.

Chapron, 2007. Changes in diatom, pollen, and chirono-

mid assemblages in response to a recent volcanic event in

Lake Galletue0 (Chilean Andes). Limnologica 37: 49–62.

Veblen, T. & F. Schlegel, 1982. Resena ecologica de los

bosques del sur de Chile. Bosque 4(2): 73–115.

Velle, G., J. Larsen, W. Eide, S. M. Peglar & H. J. Birks, 2005.

Holocene environmental history and climate of Ratasjøen,

a low-alpine lake in south-central Norway. Journal of

Paleolimnology 33: 129–153.

Verschuren, D. & H. Eggermont, 2006. Quaternary paleo-

ecology of aquatic Diptera in tropical and Southern

Hemisphere regions, with special reference to the Chiro-

nomidae. Quaternary Science Reviews 25: 1926–1947.

Villalba, R., 1990. Climatic fluctuations in Northern Patagonia

during the last 1000 years as inferred from tree-ring

records. Quaternary Research 34: 346–360.

Villalba, R., 1994. Tree-ring and glacial evidence for the

medieval warm epoch and the little ice age in Southern

South America. Climatic Change 26: 183–197.

Villa-Martınez, R., C. Villagran & B. Jenny, 2004. Pollen

evidence for late-Holocene climatic variability at Laguna

de Aculeo, Central Chile (lat. 34�S). The Holocene 14:

361–367.

Wentworth, C. K., 1922. A scale of grade and class terms for

clastic sediments. Journal of Geology 30: 377–392.

Wiederholm, T., 1983. Chironomidae of the holartic region.

Keys and diagnoses. Part I – Larvae. Entomologica

Scandinavica Supplement 19: 1–457.

Willen, E., 1991. Planktonic diatoms-an ecological review.

Algological Studies 62: 69–106.

Winkler, S., 2004. Lichenometric dating of the ‘Little Ice Age’

maximum in Mt Cook National Park, Southern Alps, New

Zealand. The Holocene 14: 911–920.

Wolin, J. & H. Duthie, 1999. Diatoms as indicators of water

level change in freshwater lakes. In Stoermer, E. F. & J. P.

Smol (eds), The Diatoms: Applications to the Environ-

mental and Earth Sciences. Cambridge University Press,

Cambridge: 183–204.

Xue, B., W. Qu, S. Wang, Y. Ma & M. D. Dickman, 2003.

Lake level changes documented by sediment properties

and diatom of Hulun Lake, China since the late Glacial.

Hydrobiologia 498: 133–141.

Hydrobiologia (2010) 648:207–225 225

123