Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Arch. Hydrobiol.

165 3 289–312 Stuttgart, March 2006

Hydrological and land-use changes in the Cuzco


region (Cordillera Oriental, South East Peru) during
the last 1200 years: a diatom-based reconstruction
Mieke Sterken1 *, Koen Sabbe1, Alex Chepstow-Lusty2,
Michael Frogley3, Koenraad Vanhoutte1, Elie Verleyen1,
Andy Cundy3 and Wim Vyverman1

With 10 figures and 1 table

Abstract: A quantitative diatom analysis was carried out on a sediment core from the
small infilled lake basin of Marcacocha and was compared with other proxy data (e. g.,
pollen and plant macroremains) in order to reconstruct environmental changes between
800 and 1850 AD in the Cuzco region of SE Peru. Five stratigraphical zones were dis-
tinguished by means of constrained cluster analysis. Very few diatoms were present
between 790 and 1070 AD, possibly reflecting dry and cool conditions, at a time when
anthropogenic impact was limited around the basin. These conditions were probably
interrupted by several flooding events, causing short-term macrophyte-dominated sta-
ges in the lake. The transition at ca. 1070 AD represented the most prominent change in
the diatom record and corresponded with a marked shift towards higher temperatures,
as deduced from plant macroremains and the pollen record. This transition further
coincided with a sudden climate shift recorded elsewhere in tropical South America.
Between 1070 and 1650 AD, our proxy data indicate the existence of a stable, shallow
lake as evidenced by increased diatom production and high concentrations of charo-
phyte oospores. The transition in the diatom data at ca. 1650 AD, with peaks in diatom
abundance centered on 1700 AD, lags the start of the “Little Ice Age” (LIA; around
1490–1530 AD), though matches a major cool and dry phase of the LIA at the end of
the 17th and early 18th centuries as recorded in Peruvian ice cores. This could be caused
by a threshold that was passed after the lake level had lowered sufficiently, due to a
cooling climate causing a drier lake environment, combined with infilling processes.

1
Authors’ addresses: Ghent University, Department of Biology, Section Protistology
& Aquatic Ecology, Krijgslaan 281-S8, 9000 Gent, Belgium.
2
Département Paléoenvironnements, Institut des Sciences de l’Evolution, Université
de Montpellier II (case postale 061), Place Eugène Bataillon, 34095 Montpellier cedex
05, France.
3
Centre for Environmental Research, University of Sussex, Brighton BN19QJ, UK.
* Corresponding author; E-mail: Mieke.Sterken@UGent.be

DOI: 10.1127/0003-9136/2006/0165-0289 0003-9136/06/0165-0289 $ 6.00


 2006 E. Schweizerbart’sche Verlagsbuchhandlung, D-70176 Stuttgart
290 Mieke Sterken et al.

Further transitions in the diatom community may be interpreted less in terms of cli-
matic change, and more as increasing sensitivity to local environmental changes, such
as lake level decrease due to infilling processes and increased anthropogenic activity in
the catchment area. The final stage of infilling occurred after ca. 1845 AD, with com-
plete colonization of the remaining lake surface by Juncaceae and the ongoing ac-
cumulation of peat.

Key words: Andes, diatoms, human impact, Inca civilization, “Little Ice Age”,
“Medieval Warm Period”, El Niño, palaeolimnology, Peru.

Introduction
Continuous long-term climate records are needed in order to reveal the impact
of past climate change on human societies (e. g., Hodell et al. 1995, Brenner
et al. 2001, Weiss & Bradley 2001), to expand our knowledge of the global
climate system and to calibrate and improve recently-developed general circu-
lation models (Markgraf et al. 2002, Smol & Cumming 2000). Lake sedi-
ments contain the remains of biotic and abiotic proxies that can provide a basis
for reconstructing past environmental conditions (Lotter et al. 1995). Among
the available palaeolimnological proxies, diatoms are considered as reliable
palaeo-indicators because of their short generation times (Denys 1993, Fritz
et al. 1999), abundance in different types of aquatic environments (Round et
al. 1990), good preservation and narrow species-specific niche-boundaries
(Smol & Cumming 2000). It is, however, inherent to palaeolimnological re-
cords that the signals obtained represent not only climatic conditions, but are
also affected by non-climatic factors such as anthropogenic activities and vol-
canic events (e. g., Telford 1998, Barker et al. 2000), and autogenic proces-
ses of natural infilling and lake succession (Gasse & Van Campo 2001, Gai-
ser et al. 2001). The combined effect of these processes on the sedimentary re-
cord should therefore generally be critically evaluated. Although some studies
exist on the role of autogenic versus allogenic processes in the development of
peat bogs (e. g., Arlen-Pouliot & Bhiry 2005, Muller et al. 2003, Vardy
et al. 2005), few studies have focused explicitly on the interaction between pa-
laeoclimatic signals and autogenic processes in lake systems (e. g., Singer et
al. 1996, Pokorny & Jankowska 2000). For example, the long term evolu-
tion of lake infilling at Portage Marsh (Indiana, USA) was stated to occur on
time scales of 100 to 1000 years, while at shorter time scales no successional
changes were observable. It was concluded that autogenic infilling over the
last 10,000 years might have influenced the vegetation response to climate
change, but that it did not serve as the primary forcing mechanism (Singer et
al. 1996). Less attention has been paid to the influence of autogenic processes
Hydrological and land-use changes in the Cuzco region 291

during the final stages of infilling, when rapid changes in ecosystem function-
ing might be observed.
In this paper we present a diatom stratigraphy of a sediment core from
Marcacocha (13˚ 13′ S, 72˚ 12′ W), a small infilled lake in the South Central Pe-
ruvian Andes (3355 m above sea level). The research focuses on the period 790
to 1845 AD, which is important in terms of both anthropogenic activity in the
region (including the expansion of the Inca civilization and the subsequent
Spanish occupation), and short-term climatic fluctuations (e. g., the “Medieval
Warm Period” and the “Little Ice Age”). More specifically, the global extent
and even existence of a “Medieval Warm Epoch” (MWE; Lamb 1965) or “Me-
dieval Climate Anomaly” (MCA; Stine 1994) is still highly debated, high-
lighting the importance for further palaeoclimatic reconstructions, especially
from the Southern Hemisphere and the tropics. In addition, we aim to clearly
distinguish between autogenic and allogenic processes by comparing the di-
atom record with other proxies (e. g., pollen, macrophyte remains and geo-
chemical markers; Chepstow-Lusty et al. 1998, 2003), as well as with other
palaeoclimate records in the region.

Study area
The infilled lake of Marcacocha is situated in the South Central Andes (Cordillera
Oriental, Peru) (Fig. 1). This small lake is bordered towards the north (with a minimal
distance of 95 m), west and south by the Patacancha River (Fig. 2), a tributary of the
Urubamba River, which joins the latter in the Inca town of Ollantaytambo, 12 km to
the south. According to Chepstow-Lusty et al. (2003), Marcacocha may have formed
as a nivation hollow. It has a diameter of 40 m, and its bedrock is assumed to be at
8.25 m depth, as observed in the core lithology (Chepstow-Lusty et al. 2003). Ac-
cording to local sources, Marcacocha was a sacred lake (Chepstow-Lusty et al. 1996,
2003); today it is filled in and covered by several Juncaceae species. The flat region
adjacent to the basin (Figs 2 and 3) is currently colonised by sedges, provides minor
potato cultivation, and is important for local grazing of cattle (Chepstow-Lusty et al.
2003); some herring-bone shaped drainage furrows are present between the lake and
the Patacancha River to the south (Fig. 3). The surrounding landscape consists of aban-
doned or underused terraces of Inca or pre-Inca origin. An archaeological site, the pro-
montory of Yuchuy Aya Orqo, is found at 95 m to the west of the infilled lake, contain-
ing stratified archaeological deposits that range from ca. 800 BC to ca. 1500 AD
(Early 1995). The location and particular morphology of the former Laguna Marca-
cocha, combined with the proximity of numerous archaeological remains makes it a
sensitive site for recording local palaeoclimatic and (agri)cultural changes.
The climate in the high Central Andes is monsoonal, with wet summers (Decem-
ber – February) and generally with a marked dry season between April – October (Gar-
reaud et al. 2003). In the highlands, El Niño events mostly correspond with dry con-
ditions in December – February, although some moderate to low strength El Niño
292 Mieke Sterken et al.

Legend:
City
N
Marcacocha
W E Palaeolimnological record
Ice core record
S Volcano
Huascaran

Junin
0 250 500 km
Lima Paca
Ayacucho Cuzco
Quelccaya
Lake Titicaca
Arequipa
Huaynaputina
Tacna

N Legend:
W E River
Macchu Picchu Road or railway
S
Marcacocha Town
Marcacocha
Ollantaytambo Urubamba Important
archaeological sites
Pisac
Ri
o
Ur
ub

Cuzco
am
ba

Pikillacta
0 10 20 km

Fig. 1. Location of Marcacocha in the Cuzco-region, Peru. Map adapted from Olivera
(1996).

events appear to have normal or even slightly wetter than average wet seasons (Gar-
reaud et al. 2003). The most severe El Niño events however appear to correspond
with marked droughts in the high Central Andes (Thompson et al. 1992, Ortlieb &
Macharé 1993).
Hydrological and land-use changes in the Cuzco region 293

Fig. 2. Location of Marcacocha, the archaeological site of Yuchuy Aya Orqo and the
Patacancha River. Map adapted from Early (1995).

Fig. 3. General view northwards up the Patacancha Valley over the in-filled lake of
Marcacocha (1) surrounded by Inca and pre-Inca terraces (2). The promontory on
which Huchuy Aya Orqo (3) is found lies behind Marcacocha, projecting from the
eastern side of the valley towards the Patacancha River (4), creating a small ravine.
The herring-boned ridges are still clearly visible today (5).
294 Mieke Sterken et al.

Table 1. Bulk radiocarbon and 210Pb dates for Marcacocha. Radiocarbon dates were
calibrated using the SHCal04 dataset (McCormac et al. 2004) in conjunction with the
CALIB 5.0 calibration program (Stuiver et al. 1993).
14
Depth Dating Laboratory C age Calibrated age Calendar
(cm) Type Reference (yr BP)* (cal. yr BP)** age***
210
0–2 Pb AD 1991
210
10 – 12 Pb AD 1958
210
18 – 20 Pb AD 1926
210
27 – 29 Pb AD 1918
210
39 – 41 Pb AD 1907
210
47 – 49 Pb AD 1905
210
50 – 51 Pb AD 1845
210
52 – 53 Pb AD 1813
101 – 102 14
C Beta-190482 400 ± 40 410 ± 80 AD 1540
115 – 123 14
C Q-2917 620 ± 50 590 ± 50 AD 1360
210 – 218 14
C Q-2918 1,460 ± 50 1,320 ± 60 AD 630
310 – 318 14
C Q-2919 1,805 ± 50 1,670 ± 120 AD 280
478 – 486 14
C Q-2920 2,245 ± 50 2,190 ± 150 BC 240
610 – 618 14
C Q-2921 3,650 ± 60 3,910 ± 200 BC 1960
* Years before present (AD1950).
** Calibrated years before present (AD1950).
*** Calendar ages are cited without errors.

Methods
Two overlapping series of cores reaching a depth of 8. 25 m were taken using a Living-
stone corer from the centre of the infilled lake of Marcacocha in 1993 (Figs 2 and 3).
Extensive descriptions of the cores and physical properties are provided in Chepstow-
Lusty et al. (1996, 1998, 2003). The original chronology for the sequence, spanning
4200 cal. yr, was based on five bulk 14C dates from samples taken at regular intervals,
and analysed at the Godwin Laboratory, Cambridge. The dates have been published
previously (Chepstow-Lusty et al. 2003). In the top interval (0–1.89 m, representing
1200 cal. yr) an additional AMS date was obtained, and calibrated using the SHCal04
data set (McCormac et al. 2004), as well as ten 210Pb dates to constrain the chronology
of the uppermost sediments, suspected as having rapidly accumulated during lake in-
filling (Table 1, Fig. 4). 210Pb activity was determined by a proxy method, through
alpha spectrometric measurement of its grand-daughter nuclide 210Po (method details
are given in Murray et al. 2003). Dates were calculated using the CRS (Constant Rate
of Supply) model (Appleby & Oldfield 1992).
Samples were taken at 1cm intervals below 50 cm and every 4 cm above 50 cm, and
were analysed for macrophyte remains, macrocharcoal and other proxies, including C-
and N-content. Samples were also taken every 4 cm for pollen analysis (for full details
see Chepstow-Lusty et al. 2003).
Hydrological and land-use changes in the Cuzco region 295

2200
2000
1800 1.794 mm yr -1
1600
Age (yr AD)

0.921 mm yr -1
1400
1200
1.266 mm yr -1
1000
800
600
400
200
0
0 40 80 120 160 200 240
Depth (cm)
Fig. 4. Age-depth model for the Marcacocha core sediments between 1 and 214 cm
depth. From the lowermost 210Pb date (52.5 cm) onwards mean sedimentation rates
were calculated and ages were linearly interpolated between the dates.

Diatom analysis
For diatom analysis, samples were investigated every 3 – 4 cm, and were weighed after
drying at 60 ˚C for 24 hours. The dried samples were oxidized using hydrogen per-
oxide (30 % volume), as described by Renberg (1990), in order to remove organic ma-
terial. The oxidized material was then mounted in Naphrax medium (refractive index
1.710). Droplets with a known volume of diatom-suspension were scanned at a 1000 ×
magnification and the surface area examined was calculated.
The concentration of diatoms per g dry weight was subsequently calculated from
the ratio of the scanned surface to the total surface of the diatom-suspension droplets.
In samples with sufficiently high diatom concentrations, a minimum of 400 valves was
counted, as well as additional thecamoebal plates and chrysophyte cysts. From sam-
ples with lower diatom concentrations, a complete droplet of 5 µl was scanned. Di-
atoms were identified to species level where possible using Round et al. (1990), Rum-
rich et al. (2000), Barber & Haworth (1981), Cox (1996), Krammer & Lange-
Bertalot (1986–1991), Manguin (1964), Frenguelli (1942), Krasske (1939), Sims
(1983), Metzeltin & Lange-Bertalot (1998) and Metzeltin & García-Rodri-
guez (2003). Difficult-to-identify diatom species were photographed, described, and
separated into morphological groups/taxa, to which a genus name and number was at-
tributed.
296 Mieke Sterken et al.

Macrofossil and total carbon and nitrogen analysis


The diatom data were complemented by data from pollen, macrofossil remains, includ-
ing macrocharcoal particles, and total carbon and nitrogen content. A detailed descrip-
tion of the pollen preparation techniques can be found in Chepstow-Lusty et al.
(2003).
For macrofossil analyses, fresh volumetric sub-samples (1cm3) were disaggregated
in deionised water, and seeds, charophyte oospores, oribatid mites and other faunal re-
mains were then hand-picked under a low-power dissecting microscope.
To calculate macrocharcoal particle abundance, volumetric sub-samples (1 cm3)
were oxidized with 30 % hydrogen peroxide in test tubes placed in a hot water bath at
60 ˚C for 2 – 4 hours. The residue was then sieved and charcoal particles > 100 µm
were counted using a low-power dissecting microscope.
For C/N analysis, sediment samples were treated with 5 % HCl overnight to re-
move inorganic carbon, and the residue was washed several times, dried at 50 ˚C and
ground to a fine powder. Well-mixed sub-samples were then combusted in a Carlo
Erba 1500 on-line to a VG TripleTrap and Optima dual-inlet mass spectrometer. C/N
ratios were calibrated using an acedanalid standard; replicate analyses indicated a pre-
cision smaller than ± 0. 1.

Numerical analysis
Cluster analysis (constrained by sample depth) was performed on the relative diatom
abundances using the programs Coniss (Grimm 1987) and Tilia 2.0b4 (Grimm 1991).
Graphs were plotted using Tilia Graph and TG View 1.1.1.1. (Grimm 2001). Ordina-
tions were performed on the relative diatom abundances. We used indirect gradient
analyses, in which samples were plotted along axes that represent the variation in spe-
cies composition (ter Braak & Prentice 1988). Using this approach, samples can be
grouped according to their species composition, without being restricted to sample
depth. An initial Detrended Correspondence Analysis (DCA) was performed on the
relative diatom abundances. In these analyses, detrending was by segments, species
data were log transformed, and down-weighting of rare species was applied. The re-
sults showed that the gradient length of the first axis was short (i. e. 2.377), implying
that linear methods (Principal Components Analysis; PCA) should be used for further
analysis (cf. Ter Braak & Prentice 1988). Ordinations were carried out using the
program CANOCO 4.0 for Windows (ter Braak 1998).
The Simpson dominance index was calculated according to Maguran (1988), in
order to compare the dominance and evenness of the diatom communities in the differ-
ent stratigraphic zones.

Results
Chronology
The 210Pb and 14C dates formed a stratigraphically consistent series (Table 1).
Errors of the CRS-modelled 210Pb dates that were included in the age-depth
Hydrological and land-use changes in the Cuzco region 297

model were smaller than + 8.33 and – 6.59 years. Errors on the 14C dated sedi-
ments did not exceed ± 70 years. From the lowermost 210Pb date (52.5 cm) on-
wards mean sedimentation rates were calculated and ages were linearly inter-
polated between the dates (Fig. 4). Sedimentation rates varied from 1.794 mm
per year between 52.5 and 101.5 cm depth, to 0.921mm per year between 101.5
and 119 cm depth, and finally to 1.266 mm per year between 119 and 214 cm.

Diatoms, pollen, macrofossils and sedimentological proxies


A total of 184 morphological types of diatoms, grouped into 44 morphological
entities, were identified (Figs 5 and 6). Using constrained cluster analysis,
five stratigraphical zones were distinguished (Fig. 6).
Zone MA 1 (189–154 cm; ca. 790–1070 AD) was characterized by very low
diatom abundances (mean number: 21 × 103 valv es per g dry weight). The
highest concentrations in this zone were found at 183 and 155 cm depth (re-
spectively 74 × 103 and 78 × 103 valves per g dry weight). The sample at
183 cm was dominated by the genera Encyonema Kützing (51 %), Frustulia

Fig. 5. LM-photographs of selected diatom forms as identified in this study, as well as


selected chrysophyte cysts and thecamoebal plates. (1) Frustulia cf. neofrenguellii
Lange-Bertalot & Rumrich, (2) Epithemia cf. argus var. alpestris Grunow, (3)
chrysophyte cyst, (4) Thecamoebal plate (Euglypha spp.), (5) Nitzschia sp. 1, (6)
Encyonema sp. 1, (7) Fragilaria cf. pinnata var. lancettula (Schumann) Hustedt, (8)
Anomoeoneis cf. sphaerophora (Ehrenberg) Pfitzer, (9) Fragilaria cf. exigua Gru-
now, (10) Aulacoseira cf. granulata (Ehrenberg) Simonsen.
Fig. 6.
790
870
950
1105
1185
1605
1660
1715
1770
1845

1025
1265
1340
1450
1550

14
14

620 +
400 +
Age (yrs AD)

(± 3 yr)

(± 9 yr)

- 50
1845 AD

1813 AD

- 40

C yrs BP
C yrs BP

210
14
(interpolated) Dates

90
80
70
60
50

190
180
170
160
150
140
130
120
110
100

De
14

pth
Ac h n ant
Pb (yrs AD)

hes
C ( C yrs BP)

spp
Am .
p h o
An ra s
o
0 4 8 00 4

m p
cf. oeon . 1
sph e
aer is
A u
oph
o
lac r a
o
Cra seira
ti spp
En cula s .
c p
En yone p.
c y one ma sp
E n m a p
cyo sp. . 2
En nema 1
c sp
En yone
cyo ma . 2
8 00 24 02 02 0 2 0 0 2 0

nem sp.
Ep
ithe a s 5
mia pp.
spp
.
10
20
Eu
n
Eu otia
Franotia sp. 1
gi sp
Fra laria sp.
gila p
ria . 11
Fra sp.
3
Frugilaria
Go stulia spp.
mp sp
Ha hone . 1
nt m
Ko zsch a sp i p
b a yas a cf. a .
ie l l mp
Na a sp. hioxi
vic
ula 1 s
Na spp
vic
ula .
Nit
zsc sp. 3
Nit hia sp
zsc .1
hia

30 0 0 01 0 2 4 01 0 0 2 0 0 2 4 6 0 2 0 2 01 0 2 4 0
spp
N a .
vic
u
val la bro
ves ken

5
Pin
n
Pin ularia
Pinnular sp.
Pinnularia sp 2
Pin nu ia . 3
nullaria sp. 4
U n
ari sp.
5
def a spp

5 10 15 20 01 0 0 0 0 3 0 4
i .
val ned b
ves rok
en
Rh
o
Se palod
ll i

All values x 10 valves per g dry weight


8 01 01 0
Sp apho a spp
ec. ra s .
4 pp.
MA5

MA3
MA4

MA2

MA1
zone

0
Diatom

84
CONISS

12 16
Total sum of squares
Mieke Sterken et al. 298
Hydrological and land-use changes in the Cuzco region 299

Fig. 7. PCA sample-plot, showing the 1st and 2nd axes. Analysis is based on the relative
abundances of the 44 defined taxa found in the Marcacocha core. The first two axes
explained respectively 22.0 % and 14.7% of the variation in the diatom data.

Rabenhorst (11 %), Pinnularia Ehrenberg (26 %), Craticula Grunow


(6 %), Brachysira Kützing (5 %) and Eunotia Ehrenberg (5 %). The sample
at 155 cm was marked by some low-pH-indicative genera, containing mostly
Frustulia (55 %), Pinnularia (26 %) and Eunotia (6 %). All other samples in
zone MA 1 mainly contained the genera Fragilaria Lyngbye (19 %), Epithe-
mia Kützing (16 %) and Aulacoseira Thwaites (15 %). The samples at 183
and 155 cm were most similar in species composition (Fig. 7), although the ab-
solute abundances of diatoms in this zone were so low that caution is needed
when comparing the diatom compositions. Zone MA 1 was characterized by
high pollen concentrations, reaching a maximum between 175 and 165 cm
depth, generally linked with high abundances of Cyperaceae and Asteraceae
(Fig. 8, Chepstow-Lusty et al. 2003). Ambrosia pollen appeared from 168 cm

Fig. 6. Absolute abundances of the diatom operative taxonomical groups that ac-
counted for at least 0.07 % of the total diatom abundances. The five diatom zones
(MArcacocha 1– MArcacocha 5) were defined by the Coniss cluster analysis based on
the relative diatom abundances. For both the Coniss cluster analysis and the ordina-
tions (Fig. 7) the taxonomical groups mentioned in Fig. 6 were used, except for the ‘un-
defined broken valves’, and combined with the following taxa: Aulacoseira sp. 2,
Brachysira sp. 1, Caloneis spp., Cocconeis spp., Cyclotella spp., Eunotia sp. 2, Frustu-
lia spp., Frustulia cf. neofrenguellii, Frustulia sp. 4, Tabellaria flocculosa.
300 Mieke Sterken et al.

n
tio

e
tra

ac c e a

ns

ne
en

en om
ep la )

tio
Po (c d)
D rpo AD

zo
e
nc
lle m)

yp odi

ea
th te

tra
a

nc iat
ia

m
co
te rs

ce

To o
os

co l d
g

to
(in (y

no
ra

ta
n

er

Pl e
br
s

ia
ta
e

z
nu

te

an
he

D
Am
Ag

ai
As
Al

M
C

C
1845 50 MA 5
MA 4
1770 60
1715 70 MA
3
1660 80
1605 90
1550 100
1450 110
MA
1340 120 2
1265 130
1185 140
1105 150
1025 160
MA
950 170 1

870 180
790 190
0 100 200 300 0 2 4 6 0 100 200 300 400 0 2 4 0.5 0 5 10 0.2 1 2 0 2 4 6 8
3 3 6
Pollen: x 10 grains per cm Diatoms: x 10 valves
per g dry weight

Fig. 8. Total pollen concentration (number of pollen grains per cm3) and pollen con-
centrations of Alnus, Ambrosia, Asteraceae, Chenopodiaceae, Cyperaceae, Maize and
Plantago (Chepstow-Lusty et al. 2003). The total diatom concentration is given to
provide comparison with the pollen sequence. Horizontal lines represent diatom zone
boundaries.

and were present throughout the rest of the core. The carbon content was low,
especially between 189 and 183 cm and above 165 cm (Fig. 9). Evidence of
aquatic macrophytes was lacking in this zone, except for Myriophyllum pollen
which had a brief appearance at 180 and more notably at 160 cm depth (Fig. 9).
The transition between zones MA 1 and MA 2 (153 – 83 cm: ca. 1070–1650
AD) represented the largest change in diatom composition and abundances
throughout the core interval. Zone MA 2 was characterized by generally
higher concentrations of diatoms; abundance maxima were observed at 85, 107
and 127 cm depth, with concentrations around 281 × 103 valves per g dry
weight. MA 2 was dominated by Epithemia (75 %), and had lower relative
abundances of the genera Anomoeoneis Pfitzer (4.7%) Fragilaria (4 %), Crati-
cula (3.8 %) and Pinnularia (3.5 %). Depths with lower diatom abundances
were also dominated by Epithemia (78 %), but further contained Fragilaria
(3.2 %), Encyonema (2.6 %) and Nitzschia Hassall (2.3 %). Zone MA 2 was
characterized by the sudden appearance of Alnus pollen (Fig. 8, Chepstow-
Lusty et al. 2003) and charophyte oospores (Fig. 9), and by a significant rise
in Juncaceae seeds at 154 cm depth. At 142 cm depth, the carbon content
reached a minimum of 13.2 %, where the concentration of macrocharcoal par-
ticles showed a peak, followed by another maximum at 123 cm depth. The
3
l cm 3 3
) ) ) ) oa r s m ns
) % (% m ed 3 te lu m ne m tio
a rc s pe rc yl r c zo to a
r
AD ed cm nt (
s lat ( e e nt ch le se m hy es 3 ph pe ia ntr
(y o h t t pe s c io m l d e
n ro rtic s cu er r op por r cm yr en to ta nc
e terp ept con co :N ite ha s e ia
Ag (in D N C C M
ac pa
M Jun p C oo p
M poll D To co
1845 50 MA 5

1770 MA 4
60

1715 70 MA 3
1660 80

1605 90

1550 100

1450 110
MA 2
1340 120
1265 130

1185 140
1105 150
1025 160

950 MA 1
170

870 180
790 190

tions in this zone, particularly between 145 and 125 cm depth.


0 2 40 20 40 60 0 20 40 0 100 200 300 0 20 40 60 80 0 20 40 60 80 100 0 100 200 300 0 1.0 0 4000 8000
% % 6
x 10 valves per
g dry weight
Hydrological and land-use changes in the Cuzco region

the total diatom concentration. Horizontal lines represent diatom zone boundaries.

abundance of mites, Juncaceae and charophyte oospores showed high fluctua-


Fig. 9. N and C content (%), the C/N ratio, the number of macrocharcoal particles, Jun-
caceae seeds, charophyte oospores, mites and Myriophyllum pollen grains per cm3 and
301
302 Mieke Sterken et al.

ns
es

tra :
tio
en s
at

nc cie
pl

s
s

co spe
x
om
om

p.

ne
d)

de
n
o )

s
en ed
rp D

sp
tio

e
en m
ep late

zo
at

in
t
te s A

ci

om of
ia

ok fin
n c to
tra

di
ha

pe
:D

on

m
co dia

at °
br nde
(in (yr

di o n
yp

to
fs
th

ps
ts

ts

ia
l
e

gl

°o
ta

i
m
ys

ys

at
Ag

D
Eu
To

Si
D

R
C

C
1845 50 MA 5
MA 4
1770 60

1715 70 MA
3
1660 80

1605 90

1550 100

1450 110
MA
1340 120 2

1265 130

1185 140

1105 150

1025 160 MA
1
950 170

870 180

790 190
0 4000 8000 0 2000 0 1000 20 40 60 80 20 40 60 0 1 0 20 0 2
x 103 valves per g x 103 per g % x 103 per g %
dry weight dry weight dry weight

Fig. 10. Total diatom and chrysophyte cyst concentration, cyst/diatom ratios (as % of
total diatom abundance), the concentration of Euglypha spp. thecamoebal plates, % of
broken diatom valves not classifiable into taxonomic groups, the Simpson dominance
index, the total number of species per sample, and the ratio of the latter two. Horizon-
tal lines represent diatom zone boundaries.

At the boundary between zones MA 2 and MA 3 (81– 63 cm: ca. 1650–1750


AD), diatom concentrations were minimal, after which they increased to more
than 4000 × 103 valves per g dry weight (75 cm) (Fig. 6). Although Epithemia
was still the dominant genus, zone MA 3 was characterized by a generally
lower Simpson dominance index and a higher number of species than zone
MA 2. Dominant genera were Epithemia (40.5 %), Anomoeoneis (11.0 %), Na-
vicula Bory de St. Vincent (16.2 %), Craticula (10.4 %) and Fragilaria (2.9 %).
Towards the end of zone MA 3 (63 cm), the cyst/diatom ratio reached a max-
imum (Fig. 10). The carbon content was slightly lower than in zones MA 2 and
MA 4, while the N content was high (Fig. 9). Both the concentrations of mac-
rocharcoal particles and charophyte oospores declined towards the end of this
zone, while the mite concentrations peaked at 76 cm (48 mites per cm3) and
65 cm (88 mites per cm3) depth.
Zone MA 4 (61– 55 cm: ca. 1750–1810 AD) contained very few diatoms per
g dry weight (mean concentration: 30 × 103 valves per g dry weight), similar
to zone MA 1. The species composition, as well as the low Simpson domi-
nance index (0.20), are similar between these two zones. Zone MA 4 is charac-
terized by Fragilaria (30.1 %), Aulacoseira (13.8 %), Cyclotella (Kützing)
Hydrological and land-use changes in the Cuzco region 303

de Brébisson (11.6 %) and Nitzschia (10.3 %). Zone MA 4 was characterized


by high C/N ratios, and notable peaks in macrocharcoal particle abundances
and Juncaceae seeds (Fig. 9). Charophyte oospores were absent and Cypera-
ceae pollen concentrations were markedly suppressed (Figs 8 and 9).
In zone MA 5 (53 – 50 cm: ca. 1810–1845 AD) the concentration of diatom
valves increased to 7564 × 103 valves per g dry weight. In contrast to zone MA
3, this zone showed a low Simpson dominance index (0.10), indicating higher
evenness of the communities (Fig. 10). Dominant taxa were Encyonema
(17.2 %), Navicula (16.7 %) and Epithemia (16.7 %). Other genera were Nitz-
schia (8.2 %), Pinnularia (7.5 %), Achnanthes (7.4 %), Kobayasiella (4.4 %)
and Fragilaria (3.8 %). In contrast, Anomoeoneis and Craticula, which were
very abundant in zone MA 3, were virtually absent from zone MA 5. Encyo-
nema was more abundant in zone MA 5, and Frustulia sp. 4, Pinnularia sp. 2
and Pinnularia spp. were present in zone MA 5, but absent in zone MA 3.
Zone MA 5 was further characterized by a maximal concentration of theca-
moebal plates (Euglypha spp.) (Fig. 10) and high Cyperaceae pollen concentra-
tions (Fig. 8). Total N % and C %, and concentrations of Juncaceae seeds,
charophyte oospores, mites and macrocharcoal particles were low in this zone
(Fig. 9).
PCA plots show the rate of change between consecutive strata and the rela-
tionship between the zones defined by the constrained cluster analysis, as sam-
ple points are plotted nearer to each other when having a more similar diatom
composition (ter Braak 1994). In the PCA sample-plot (Fig. 7), most samples
were grouped per diatom zone defined by the Coniss cluster analysis (Fig. 6).
The first and second axes explained respectively 22.0 % and 14.7% of the total
variance in the diatom data. Samples belonging to MA 1 and MA 4 were
grouped together on the right side of the diagram as a result of their similarity
in diatom community composition, which is not dominated by Epithemia, Na-
vicula, Craticula or Anomoeoneis. The samples from MA 3 and MA 5 are sit-
uated on the left side of the plot (Fig. 7).

Discussion
Marcacocha zone MA 1: 189–155 cm: ca. 790–1070 AD
Zone MA 1 is characterized by very low abundances of diatoms, with the ap-
pearance of slightly higher diatom concentrations only at 185 and 155 cm
depth. The taxa occurring at these depths (Encyonema spp., Fragilaria spp.,
Pinnularia sp. 3, Frustulia sp. 1, Frustulia sp. 2, Frustulia sp. 4, Eunotia sp. 1,
Brachysira sp. 1, Tabellaria flocculosa) generally indicate oligotrophic soft
water environments that are often acidic and sub-aerated. The suppressed di-
atom abundances in zone MA 1 might reflect low productivity, possibly in
304 Mieke Sterken et al.

combination with taphonomical loss, as the preservation of diatoms is known


to be poor in either acid or alkaline environments (Beyens 1985, Denys
1993). The low diatom productivity in this zone might result from very low
lake levels, during generally dry climate conditions, periodically interrupted
by flash floods of the Patacancha River. These flood events are evidenced by
the high amount of broken versus unbroken valves present in this zone
(Fig. 10), as well as a lower C content indicating increased inorganic input
(Fig. 9). The occurrence of unstable lacustrine conditions is further confirmed
by the lack of macrophyte remains in this zone, with the notable exception of
the samples at 160 and 180 cm depth, in which Myriophyllum pollen concen-
trations peak towards respectively 765 and 139 pollen grains per cm3 (Fig. 9).
This aquatic plant is a typical colonizer of shallow water conditions (Haber
1997). The brief appearance of Myriophyllum might thus have mirrored two
short-term phases of a macrophyte-dominated shallow water system, possibly
caused by flooding events. The high C/N ratio and low C content, indicative of
increased inorganic input, might – besides by flash floods – alternatively be
caused by anthropogenic activities, grazing and increased regional dust pro-
duction during dry events. Our multi-proxy results, which indicate that the pe-
riod between ca. 790 and 1070 AD was probably characterized by relatively
dry and cold climate conditions, are in agreement with previous findings (e. g.,
Chepstow-Lusty et al. 2003), and with ice core data from the Peruvian An-
des, that were characterized by high dust deposition, attributed to either anth-
ropogenic impact, a dry climate or more possibly a combination of both, in the
period between 830 and 1060 AD (Thompson et al. 1988).

Marcacocha zone MA 2: 153 – 83 cm: ca. 1070–1650 AD


A clear hydrological change occurred around 1070 AD, resulting in an increase
in total diatom abundances and a shift towards an Epithemia-dominated di-
atom community (Fig. 6). Most of the genera found in this zone are benthic or
epiphytic (e. g., Epithemia spp.) and indicative of oligotrophic, shallow fresh-
water conditions. It appears that at this stage, Marcacocha was a shallow lake
with relatively stable lake levels. This is confirmed by the sudden appearance
of charophyte oospores, as these submerged macrophytes favour clear-water-
conditions (Scheffer 1998). The marked rise in Alnus pollen from 1070 AD
(Chepstow-Lusty et al. 2003; Fig. 9) further indicates warmer, though still
generally dry, conditions (Weng et al. 2004). In many other palaeoclimatic
studies in the central-Peruvian Andes and the Bolivian Altiplano, an important
climate shift has also been observed around 1070 AD, which may be equiva-
lent to the period known in the northern hemisphere as the “Medieval Warm
Hydrological and land-use changes in the Cuzco region 305

Period” (“MWP”). For example, Thompson et al. (1988, 1995), Binford et al.
(1997) and Abbott et al. (1997) recorded a dry period commencing at ca.
1000–1100 AD and continuing to 1300–1500 AD. Instead of a dry interval at
Marcacocha, our proxy data point to the establishment of shallow lake condi-
tions and a higher water table, which might be the result of increased melt-
water runoff via the Patacancha River. The sudden character of the shift in di-
atom composition and abundance around 1070 AD, and its correspondence
with different palaeoclimatic data from the central tropical Andes, indicate ab-
rupt climatic and hydrological changes, instead of autogenic processes or local
disturbances in the vicinity of the lake.
According to Chepstow-Lusty et al. (2003) recurrent phases of increas-
ing anthropogenic landscape management occurred from 1070 AD. The low
abundance of C % at 1150 AD may be linked to a major phase of terrace build-
ing, although probably prior to Inca expansion from Cuzco. However, this
event at 1150 AD was not reflected in the diatom stratigraphy, though pro-
nounced changes in macrocharcoal (e. g., around 1140 and 1290 AD), oribatid
mites, Juncaceae seeds and charophyte oospores (which all show short term
maxima between 1425 and 1530 AD; Fig. 9) could partially reflect changing
anthropogenic activities. The collapse of the Inca Empire (1538 AD) may be
reflected by a small reduction in N %, macrocharcoal and oribatid mites
around 1500–1550 AD through the abandonment of sites in the Patacancha
valley. Following the Spanish Conquest, gradual long term increases in mac-
rocharcoal particles and N content are detected, probably reflecting renewed
anthropogenic activity in the area. This zone in the Marcacocha sediment core
thus represents one of the most important time periods for human and climate
history in the region and merits further study.
The lack of prominent changes in the diatom record at the start of the
“Little Ice Age” (“LIA”), as defined between 1490 and 1530 AD in tropical
South America (Thompson et al. 1986, 1995), might indicate that the lake was
buffered and stable enough to retain its limnological characteristics; a trend
probably reinforced by the gradual character of the cooling at the beginning of
the “LIA” in tropical South America (Thompson et al. 1986).

Marcacocha zone MA 3: 81– 63 cm: ca. 1650–1750 AD


The number of diatom species and total concentration of diatoms per g dry
weight increased at 81 cm depth (Figs. 6 and 10). The most abundant genera
(i. e., Epithemia, Anomoeoneis, Navicula, Craticula, Fragilaria and Nitzschia)
contain epiphytic and benthic species that are tolerant or even typical of high
conductivity waters (Cox 1996, Round et al. 1990, Cholnoky 1968). Anom-
oeoneis sphaerophora is a benthic species (Kociolek & Herbst 1992) often
occurring in pools prone to temporary drying, and sometimes brackish or even
306 Mieke Sterken et al.

saline waters (Blinn et al. 1994, Cholnoky 1968, Cox 1996, Rakowska
1997). Many of the Craticula specimens observed in this section were found in
their ‘Heribaudii-form’, a form in which the diatoms build internal valves with
very broad transapical costae when the organisms are under osmotic pressure
(Patrick & Reimer 1966, Cox 1996). Together, these data suggest fluctuating
water levels. A threshold, resulting in an enhanced shallowing of the lake-
level, was possibly exceeded, due either to autogenic gradual infilling or to a
cool-dry environment, caused by a reduction of melt-water transported by the
Patacancha River as cooling prevailed in the high Andes, or a combination of
both. These data are in agreement with a prolonged cooling trend at the end of
the 17th and early 18th centuries as recorded in Peruvian ice cores (Thompson
et al. 1986).
The end of zone MA 3 (66 – 63 cm depth) is marked by a maximal chryso-
phyte cyst to diatom ratio. The diatom concentrations are very low and the
cyst abundances reach a maximum at 63 cm, which could indicate oligotroph-
ication (Smol 1985). This short episode (approx. 1730–1750 AD) coincides
with the coldest and driest phase of the LIA in the tropical Peruvian Andes
(Thompson et al. 1986, 1995). The dominance of epiphytes and species typical
of electrolyte-rich waters are therefore probably the result of a negative pre-
cipitation evaporation balance, possibly in combination with natural infilling
processes.

Marcacocha zone MA 4: 61– 55 cm: ca. 1750–1810 AD


Zone MA 4 resembles zone MA1 mainly in the low abundances of siliceous
microfossils and also in its diatom composition (Fig. 6). Together with the ab-
sence of aquatic macrophytes (e. g., charophytes), these data could indicate the
continuation of a terrestrialization process. This hypothesis is supported by the
lower C and N content and the high C/N ratio (Fig. 9), which indicate the input
of more allochthonous material into the lake basin. The high peak of mites at
the end of zone MA 3 (at 67 cm), followed by a sudden maximum in macro-
charcoal and Juncaceae seeds (63 – 55 cm), similarly suggests terrestrialization
and the development of a Juncaceae wetland community, colonizing inwards
from the margin of the lake, associated with increased grazing and anthropo-
genic burning. It is possible that the fish-bone shaped furrows between the
lake and the Patacancha River (Fig. 3) were constructed in this time interval
(1750–1810 AD), increasing the drainage from Marcacocha, particularly dur-
ing periods of low river discharge. In cold and dry episodes, reduced precipita-
tion and less meltwater is likely to result in a lower river discharge. This
source of water currently appears to feed the lake, largely by infiltration, as
shown by a flooded 0.5 m deep archaeological test-pit, dug 80 m to the west
during the height of the dry season (Early 1995). The occurrence of such a
Hydrological and land-use changes in the Cuzco region 307

cold and dry period between 1750 and 1810 AD is in close agreement with the
Quelccaya ice core record, which indicates a cold and wet climate during the
major part of the LIA (1500–1720 AD), dry between 1720–1860 AD, and ex-
tremely cold between 1800–1820 AD (Thompson et al. 1986, 1988, 1995).
Historical data (Ortlieb 1994, Quinn et al. 1987) similarly indicate a higher
frequency of El Niño events, which currently often cause droughts in the Cuz-
co-region, in the 19th and early 20th centuries (Thompson et al. 1984).

Marcacocha zone MA 5: 53 – 50 cm: ca. 1810–1845 AD


The diatom species composition and high total abundance at 50 cm indicate
the brief re-establishment of shallow lake conditions at Marcacocha around
1845 AD, possibly during an episode of enhanced rainfall, or rising temperatu-
res leading to increased meltwater input. This would cause inorganic material
deposition in Marcacocha, during rapid run-off or flooding of the Patacancha
River, as evidenced by the low carbon content (Fig. 9). The short-lived lake
was probably shallower (due to further infilling), yet more stable than in the
period of 1650–1750 AD, as evidenced by the high abundances of Epithemia,
Navicula and Achnanthes (Fig. 6). The high number of species in this zone
could have resulted from the appearance of new niches in the shallow lake or
wetland during continued infilling and the compression of Juncaceae vegeta-
tion into peats. Anomoeoneis and Craticula are less dominant than in zone
MA 3, probably indicating a very shallow but more stable water table, and a
decreased nutrient concentration. The low numbers of chrysophyte cysts and
the maximum abundance of Euglypha spp. thecamoebal plates (Fig. 10) prob-
ably indicate that the lake was evolving into a swamp or peat-like environment
(Bobrov et al. 1999). Euglypha species are commonly found on Sphagnum
mosses in peat bogs (Beyens & Chardez 1984), but also in shallow circum-
neutral lakes (Beyens et al. 1995). The presence of these plates at this depth is
thus not surprising as this zone represents the final infilling stage of the lake,
and the upper 50 cm of the core consists of peats (Chepstow-Lusty et al.
2003).

Conclusions
It can be concluded that the diatom stratigraphy of Marcacocha reflects the
ecological changes of the former lake which, at an earlier stage, could be inter-
preted in terms of a regional climatic shift around 1070 AD, when a period-
ically interrupted dry and cooler period ended, and relatively stable lacustrine
conditions were established. As the lake became gradually shallower due to
autogenic infilling processes and increased anthropogenic activity in the catch-
308 Mieke Sterken et al.

ment area, it became more sensitive to small-scale climate shifts and local dis-
turbances between 1070 and 1845 AD. Ca. 1750 AD, the lake experienced a
pronounced cold and dry phase, as indicated by the virtual absence of diatoms.
This was possibly enhanced by the construction of herring-bone shaped fur-
rows for drainage near the lake at this time. Around 1845 AD, a very brief wet
phase was observed in the Marcacocha sequence, possibly caused by an epi-
sode of enhanced rainfall or rising temperatures leading to increased melt-
water input, marking an early end to the “LIA”. Subsequently, Marcacocha
became a wetland with continuous Juncaceae peat deposition.

Acknowledgements
Keith Bennett is gratefully acknowledged for technical expertise and initiating the
work at Marcacocha. Funding was provided by NERC, the Royal Society, the British
Ecological Society and the Montgomery Trust. Ann Kendall and the Cusichaca
Trust archaeological group provided support in visits to Marcacocha. Tino Aucca
Chutas, Sandra Echegaray, Alfredo Tupayachi Herrera and Washington
Galliano also provided invaluable assisance in Peru. Melanie Leng (NERC Isotope
Geosciences Laboratory) is warmly acknowledged for generously providing the indis-
pensable C% and N% data. Elie Verleyen is funded as a Senior Research Assistant
by the Fund for Scientific Research (Belgium). We also thank Dr. A. Robertson, Prof.
Dr. F. García-Rodriguez and an anonymous reviewer for their valuable comments.

References
Abbott, M. B., Seltzer, G. O., Kelts, K. R. & Southon, J. (1997): Holocene paleo-
hydrology of the tropical Andes from lake records. – Quatern. Res. 47: 70 – 80.
Appleby, P. G. & Oldfield, F. (1992): Applications of 210Pb to sedimentation studies.
– In: Ivanovich, M. & Harmon, R. S. (eds): Uranium-series disequilibrium. Ap-
plications to earth, marine and environmental sciences. 2nd Edition, – Oxford Sci-
ence, Oxford.
Arlen-Pouliot, Y. & Bhiry, N. (2005): Palaeoecology of a palsa and a filled thermo-
karst pond in a permafrost peatland, subarctic Quebec, Canada. – Holocene 15:
408 – 419.
Barber, H. G. & Haworth, E. Y. (1981): A guide to the morphology of the diatom
frustule. Scientific publication no 44. – Freshwater Biological Association, Amble-
side, UK, 112 pp.
Barker, P., Telford, R., Merdaci, O., Williamson, D., Taieb, M., Vincens, A. &
Gibert, E. (2000): The sensitivity of a Tanzanian crater lake to catastrophic tephra
input and four millennia of climate change. – Holocene 10: 303 – 310.
Beyens, L. (1985): On the subboreal climate of the Belgian Campine as deduced from
diatom and testate amoebae analyses. – Rev. Palaeobot. Palynol. 46: 9 – 31.
Beyens, L. & Chardez, D. (1984): Testate amoebae (Rhizopoda, testaceae) from
southwest Ireland. – Arch. Protistenkunde 128: 109–126.
Beyens, L., Chardez, D., Debaere, D. & Verbruggen, C. (1995): The aquatic testate
amoebas fauna of the Stromness Bay Area, South-Georgia. – Antarct. Sci. 7: 3 – 8.
Hydrological and land-use changes in the Cuzco region 309

Binford, M. W., Kolata, A. L., Brenner, M., Janusek, J. W., Seddon, M. T., Ab-
bott, M. & Curtis, J. H. (1997): Climate variation and the rise and fall of an An-
dean civilisation. – Quatern. Res. 47: 235 – 248.
Blinn, D. W., Hevly, R. H. & Davis, O. K. (1994): Continuous Holocene record of
diatom stratigraphy, paleohydrology, and anthropogenic activity in a spring-
mound in southwestern United-States. – Quatern. Res. 42: 197– 205.
Bobrov, A. A., Charman, D. J. & Warner, B. G. (1999): Ecology of testate amoebae
(Protozoa: Rhizopoda) on peatlands in western Russia with special attention to
niche separation in closely related taxa. – Protist 150: 125–136.
Brenner, M., Hodell, D. A., Curtis, J. H., Rosenmeier, M. F., Binford, M. W. &
Abbott, M. B. (2001): Abrupt climate change and pre-Columbian cultural col-
lapse. – In: Markgraf, V. (ed.): Interhemispheric climate linkages. – Academic
press, San Diego, pp. 87–104.
Chepstow-Lusty, A. J., Bennett, K. D., Fjeldså, J., Kendall, A., Galiano, W. &
Herrera, A. T. (1998): Tracing 4,000 years of environmental history in the Cuzco
area, Peru, from the pollen record. – Mt. Res. Dev. 18: 159–172.
Chepstow-Lusty, A. J., Bennett, K. D., Switsur, V. R. & Kendall, A. (1996):
4000 years of human impact and vegetation change in the central Peruvian Andes
– with events paralleling the Maya-record? – Antiquity 70: 824 – 833.
Chepstow-Lusty, A. J., Frogley, M. R., Bauer, B. S., Bush, M. B. & Herrera, A.
T. (2003): A late Holocene record of arid events from the Cuzco region, Peru. – J.
Quaternary Sci. 18: 491– 502.
Cholnoky, B. J. (1968): Die Ökologie der Diatomeen in Binnengewässern. – J. Cra-
mer, Lehre. 699 pp.
Cox, E. J. (1996): Identification of Freshwater Diatoms from Live Material. – Chap-
man & Hall, London, 158 pp.
Denys, L. (1993): Paleoecologisch diatomeeënonderzoek van de Holocene afzettingen
in de westelijke Belgische kustvlakte. – Unpublished PhD research, UIA, Biology
department, Antwerp, 655 pp.
Early, R. (1995): Excavation at Juchuy Aya Orqo, Marcacocha Basin, Patacancha Val-
ley. – Unpublished Archaeological Report, The Cusichaca Trust: Witney, UK.
Frenguelli, J. (1942): Diatomeas del Neuquén (Patagonia). – Revista del Museo de
la Plata, Sección Botanica 5: 73 – 219.
Fritz, S. C., Cumming, B. F., Gasse, F. & Laird, K. R. (1999): Diatoms as indicators
of hydrologic and climatic change in saline lakes. – In: Stoermer, E. F. & Smol,
J. P. (eds): The diatoms: applications for the environmental and Earth Sciences. –
Cambridge University Press, Cambridge, pp. 41–71.
Gaiser, E. E., Taylor, B. E. & Brooks, M. J. (2001): Establishment of wetlands on
the southeastern Atlantic Coastal Plain: paleolimnological evidence of a mid-Hol-
ocene hydrologic threshold from a South Carolina pond. – J. Paleolimnol. 26:
373 – 391.
Garreaud, R., Vuille, M. & Clement, A. C. (2003): The climate of the Altiplano:
observed current conditions and mechanisms of past changes. – Palaeogeogr. Pa-
laeoclimat. 194: 5 – 22.
Gasse, F. & Van Campo, E. (2001): Late Quaternary environmental changes from a
pollen and diatom record in the southern tropics (Lake Tritrivakely, Madagascar).
– Palaeogeogr. Palaeoclimat. 167: 287– 308.
310 Mieke Sterken et al.

Grimm, E. C. (1987): CONISS, a FORTRAN–77 program for stratigraphically con-


strained cluster analysis by the method of incremental sum of squares. – Comput.
Geosci. 13: 13 – 35.
– (1991): Tilia version 2.0 b4. – Springfield: Illinois State Museum, Illinois.
– (2001): Tilia Graph and TGView version 1. 1. 1. 1. – Springfield: Illinois State Mu-
seum, Illinois.
Haber, E. (1997): Eurasian watermilfoil fact sheet, invasive exotic plants of Canada,
fact sheet no 10 National Botanical Services, Ottawa, ON, Canada. –
http.//infoweb.magi.com/~ehaber/factfoil.html (15-07-2002).
Hodell, D. A., Curtis, J. H. & Brenner, M. (1995): Possible role of climate in the
collapse of classic Maya civilization. – Nature 375: 391– 394.
Kociolek, J. P. & Herbst, D. B. (1992): Taxonomy and distribution of benthic di-
atoms from Mono Lake, California, USA. – T. Amer. Microsc. Soc. 111: 338 –
355.
Krammer, K. & Lange-Bertalot, H. (1986–1991): Bacillariophyceae: 1– 4. – In:
Ettl, H., Gerloff, J., Heynig, H. & Mollenhauer, D. (eds): Süßwasserflora
von Mitteleuropa, 2 (1), – Gustav Fischer Verlag, Jena.
Krasske, G. (1939): Zur Kieselalgenflora Sudchiles. – Arch. Hydrobiol. 35: 349 – 468.
Lamb, H. H. (1965): The early medieval warm epoch and its sequel. – Palaeogeogr.
Palaeoclimat. 1: 13 – 37.
Lotter, A. F., Birks, H. J. B. & Zolitschka, B. (1995): Late-glacial pollen and di-
atom changes in response to two different environmental perturbations – volcanic
eruption and Younger Dryas cooling. – J. Paleolimnol. 14: 23 – 47.
Magurran, A. E. (1988): Ecological diversity and its measurement. – Princeton Uni-
versity Press, 168 pp.
Manguin, E. (1964): Contribution à la connaissance des Diatomées des Andes du
Pérou. – Mém. Mus. Nat. d’Histoire Naturelle (Série B. Botanique) 12: 41– 98.
Markgraf, V., Baumgartner, T. R., Bradbury, J. P., Diaz, H. F., Dunbar, R. B.,
Luckman, B. H., Seltzer, G. O., Swetnam, T. W. & Villalba, R. (2002): Pa-
leoclimate reconstruction along the Pole-Equator-Pole transect of the Americas
(PEP 1). – Quatern. Sci. Rev. 19: 125–140.
McCormac, F. G., Hogg, A. G., Blackwell, P. G., Buck, C. E., Higham, T. F. G. &
Reimer, P. J. (2004): SHCal04 Southern Hemisphere Calibration 0 – 1000 cal BP.
– Radiocarbon 46: 1086–1092.
Metzeltin, D. & García-Rodriguez, F. (2003): Las diatomeas Uruguayas. – DI. R.
A. C. – Facultad de Ciencias, Montevideo, 207 pp.
Metzeltin D. & Lange-Bertalot, H. (1998): Tropical Diatoms of South America I.
– Iconogr. Diatomol. 5: 1– 695.
Muller, S. D., Richard, P. J. H. & Larouche, A. C. (2003): Holocene development
of a peatland (southern Quebec): a spatio-temporal reconstruction based on pachy-
metry, sedimentology, microfossils and macrofossils. – Holocene 13: 649 – 664.
Murray, J. W., Alve, E. & Cundy, A. B. (2003): The origin of modern agglutinated
foraminiferal assemblages: evidence from a stratified fjord. – Estuar., Coast. Shelf
S. 58: 677– 697.
Olivera, F. (1996): Spatial Hydrology of the Urubamba River System in Peru using
Geographic Information Systems (GIS) –
www.ce.utexas.edu/prof/olivera/peru/peru.htm
Hydrological and land-use changes in the Cuzco region 311

Ortlieb, L. (1994): Major historical rainfalls in central Chile and the chronology of
ENSO events during the 16th –19th centuries. – Rev. Chil. Hist. Nat. 67: 463 – 485.
Ortlieb, L. & Macharé, J. (1993): Former El-Niño events – records from western
South-America. – Global Planet. Change 7: 181– 202.
Patrick, R. & Reimer, C. W. (1966): The diatoms of the United States, exclusive of
Alaska and Hawaii. Volume 1. – Academy of Natural Sciences of Philadelphia,
Philadelphia, 688 pp.
Pokorny, P. & Jankovska, V. (2000): Long-term vegetation dynamics and the infill-
ing process of a former lake (Svarcenberk, Czech Republic). – Folia Geobot. 35:
433 – 457.
Quinn, W. H., Neal, V. T. & Antunez de Mayola, S. E. (1987): El Niño occurren-
ces over the past four and a half centuries. – J. Geophys. Res. 92: 14,449–14,461.
Rakowska, B. (1997): Diatom communities in a salt spring at Pelczyska (Central Po-
land). – Biologia 52: 489 – 493.
Renberg, I. (1990): A procedure for preparing large sets of diatom slides from sedi-
ment cores. – J. Paleolimnol. 4: 87– 90.
Round, F. E., Crawford, R. M. & Mann, D. G. (1990): The diatoms: biology and
morphology of the genera. – Cambridge University Press, Cambridge, 747p.
Rumrich, U., Lange-Bertalot, H. & Rumrich, M. (2000): Diatoms of the Andes,
from Venezuela to Patagonia/Tierra del Fuego, and two additional contributions. –
A. R. G. Gantner Verlag, Königstein, 673 pp.
Scheffer, M. (1998): Ecology of shallow lakes. – Chapman & Hall, London, 357 pp.
Sims, P. A. (1983): A taxonomic study of the genus Epithemia with special reference to
the type species E. turgida (Ehrenb.) – Kutz. Bacillaria 6: 211– 235.
Singer, D. K., Jackson, S. T., Madsen, B. J. & Wilcox, D. A. (1996): Differentiating
climatic and successional influences on long-term development of a marsh. –
Ecology 77: 1765–1778.
Smol, J. P. (1985): The ratio of diatom frustules to chrysophycean statospores: A use-
ful paleolimnological index. – Hydrobiologia 123: 199 – 208.
Smol, J. P. & Cumming, B. F. (2000): Tracking long-term changes in climate using
algal indicators in lake sediments. – J. Phycol. 36: 986–1011.
Stuiver, M., Reimer, P. J. & Reimer, R. W. (2005): CALIB 5. 0 (WWW program and
documentation).
Stine, S. (1994): Extreme and persistent drought in California and Patagonia during
mediaeval time. – Nature 369: 546 – 549.
Telford, R. J. (1998): Diatom stratigraphies of Lakes Awassa and Tilo, Ethiopia: Hol-
ocene records of groundwater variability and climate change. – Thesis submitted
in fulfillment of the award of the degree of Doctor of Philosophy at the University
of Wales, Aberystwyth.
ter Braak, C. J. F. (1994): Canonical community ordination. Part I: Basic theory and
linear methods. – Ecoscience 1: 127–140.
– (1998): CANOCO for Windows 4.0. – Wageningen: Center for biometry Wagenin-
gen.
ter Braak, C. J. F. & Prentice, I. C. (1988): A theory of gradient analysis. – Adv.
Ecol. Res. 18: 271– 317.
Thompson, L. G., Mosley-Thompson, E. & Arnao, B. M. (1984): El Niño-Southern
Oscillation events recorded in the stratigraphy of the tropical Quelccaya ice cap,
Peru. – Science 226: 50 – 52.
312 Mieke Sterken et al.

Thompson, L. G., Davis, M. E., Mosley-Thompson, E. & Liu, K. B. (1988): Pre-


Incan agricultural activity recorded in dust layers in two tropical ice cores. – Na-
ture 336: 763–765.
Thompson, L. G., Mosley-Thompson, E., Dansgaard, W. & Grootes, P. M. (1986):
The Little Ice Age as recorded in the stratigraphy of the Tropical Quelccaya Ice
Cap. – Science 234: 361– 364.
Thompson, L. G., Mosley-Thompson, E., Davis, M. E., Lin, P. N., Henderson, K.
A., Coledai, J., Bolzan, J. F. & Liu, K. B. (1995): Late-glacial stage and Holo-
cene tropical ice core records from Huascarán, Peru. – Science 269: 46 – 50.
Thompson, L. G., Mosley-Thompson, E. & Thomson, P. A. (1992): Reconstructing
interannual climate variability from tropical and sub-tropical ice cores. – In: Diaz,
H. F. & Markgraf, V. (eds): El Niño, Historical and Paleoclimatic Aspects of the
Southern Oscillation, – Cambridge University Press, Cambridge, pp. 295 – 322.
Vardy, S. R., Warner, B. G. & Asada, T. (2005): Holocene environmental change in
two polygonal peatlands, south-central Nunavut, Cananda. – Boreas 34: 324 – 334.
Weiss, H. & Bradley, R. S. (2001): What drives societal collapse? – Science 291:
609 – 610.
Weng, C., Bush, M. B. & Chepstow-Lusty, A. J. (2004): Holocene changes of An-
dean alder (Alnus acuminata) in highland Ecuador and Peru. – J. Quat. Sci. 19:
685 – 691.

Submitted: 8 June 2005; accepted: 21 December 2005.

You might also like