You are on page 1of 11

Earth and Planetary Science Letters 299 (2010) 447457

Contents lists available at ScienceDirect

Earth and Planetary Science Letters


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / e p s l

Andean uplift and climate evolution in the southern Atacama Desert deduced from
geomorphology and supergene alunite-group minerals
Thomas Bissig ,1, Rodrigo Riquelme
Departamento de Ciencias Geolgicas, Universidad Catlica del Norte, Avenida Angamos 0610, Antofagasta, Chile

a r t i c l e

i n f o

Article history:
Received 5 February 2010
Received in revised form 21 September 2010
Accepted 21 September 2010
Available online 18 October 2010
Editor: T.M. Harrison
Keywords:
Atacama Desert
Andes
uplift
climate evolution
supergene alunite
El Salvador
Potrerillos
geochronology
stable isotopes
Chile

a b s t r a c t
Supergene alunite group minerals from the Late Eocene El Salvador porphyry Cu district, the El Hueso
epithermal gold deposit and the Coya porphyry Au prospect located in the Precordillera of Northern Chile
(~ 26 to 26 30 Lat. S) have been dated by the 40Ar/39Ar method and analyzed for stable isotopes. These data
support published geomorphologic and sedimentologic evidence suggesting that the Precordillera in the
Southern Atacama Desert had been uplifted as early as the late Eocene and, thus, signicantly prior to the
Altiplano which attained its high elevation only in the late Miocene.
The oldest supergene alunite from the Damiana exotic deposit at El Salvador was dated at 35.8 1 Ma and
yielded a D (VSMOW) value of 74 which indicates elevations of the Precordillera near El Salvador of at
least 3000 m in the Late Eocene. In contrast, Miocene supergene alunite from El Salvador, El Hueso, and Coya
have less negative D signatures reaching values as high as 23 to 25 at El Hueso and El Salvador between
about 8.2 and 14 Ma. Late Miocene to Holocene supergene alunite, jarosite and natroalunite ages are restricted
to El Hueso and Coya located near 4000 m above sea level in the Precordillera, roughly 1000 m higher than the
present elevation of El Salvador. The D values of samples younger than ~ 5 Ma vary between 57 and 97.
The complex evolution of the D signatures suggests that meteoric waters recorded in supergene alunite
group minerals were variably affected by evaporation and provides evidence for climate desiccation and onset
of hyper arid conditions in the Central Depression of the southern Atacama Desert after 15 Ma, which agrees
well with published constraints from the Atacama Desert at 2324 Lat. S. Our data also suggest that wetter
climatic conditions than at present prevailed in the latest Miocene and early Pliocene in the Precordillera.
The new and previously published age constraints for El Salvador indicate that supergene mineralization at
the Damiana exotic Cu deposit occurred periodically over 23 Ma in a locally exceptionally stable
paleohydrologic and geomorphologic conguration.
2010 Elsevier B.V. All rights reserved.

1. Introduction
The Andean uplift history, its causes and effects on the climate
have been subject of signicant research in recent years (e.g.,
Garzione et al., 2008; Lamb and Davis, 2003; Schlunegger et al.,
2006). Much of this work has been concentrated in the northern Chile
and Altiplano transects (~ 1820 Lat. S, Fig. 1). Faras et al. (2005) and
Victor et al. (2004) suggest that up to 2600 m of uplift occurred in the
late middle Miocene and was accommodated by high-angle west
verging faults in the western Cordillera. Geomorphologic (Garcia and
Hrail, 2005; Hoke et al. 2007; Schlunegger et al., 2006; Thouret et al.
2007) and stable isotope evidence (Garzione et al., 2008) places the
major uplift which gave rise to the present day high elevations of the
Altiplano in the late Miocene. The southern Atacama Desert (~ 2627
Corresponding author.
E-mail address: tbissig@eos.ubc.ca (T. Bissig).
1
Mineral Deposit Resarch Unit, Department of Earth and Ocean Sciences, University
of British Columbia, 6339 Stores Road, Vancouver, B.C., V6T 1Z4, Canada.
0012-821X/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2010.09.028

Lat. S, Fig. 1) has received comparatively less recent attention, but


available evidence indicates that the uplift history was fundamentally
distinct, irrespective of the controversies on the exact timing of
Altiplano uplift. For example, no signicant high angle west verging
faults active during the Miocene have been documented for the
southern Atacama Desert. In addition, geomorphologic, apatite ssion
track and sedimentological evidence (Nalpas et al., 2005; Riquelme
et al., 2007) suggest that in the southern Atacama Desert the
Precordillera had attained considerable elevations in the Oligocene
or earlier, which greatly precedes the Altiplano uplift. We herein
assess the uplift and climate evolution in an oblique transect across
the Precordillera at 2626 30 Lat. S (Figs. 1, 2) on the basis of the well
established geomorphologic framework (Bissig and Riquelme, 2009;
Nalpas et al. 2008; Riquelme et al., 2003, 2007, 2008) and eleven new
40
Ar/39Ar ages and corresponding stable isotope data for supergene
alunite group minerals from the El Salvador porphyry Cu district (e.g.,
Gustafson et al., 2001), the El Hueso epithermal Au deposit (Marsh
et al., 1997; Thompson et al., 2004) and the Coya porphyry Au
prospect (Rivera et al., 2004), all located in the southern Atacama

448

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

69 W

70 W

18 S

Altiplano Segment

ARICA

SouthAmerica

O C E A N

IQUIQUE

P A C I F I C

TRENCH

WC

1
CALAMA

CB
CD

23 S

PD

CC

3 SP
4
5

COPIAPO
0

50

100 km

PC

CCG

PER
CHAARAL

Southern Atacama Desert


(Puna Segment)

CHILE

ANTOFAGASTA

TALTAL

50

68 W

27 S

Fig. 1. Map of the western Andean slope of northern Chile. The study region is outlined
(Fig. 2). Dotted lines indicate physiographic boundaries from Riquelme et al. (2007).
Abbreviations: CC: Coastal Cordillera; CD: Central Depression; PC: Precordillera; PD:
Preandean Depression; SP: Salar de Pedernales; WC: Western Cordillera; CB: Calama
basin; CCG: Cordillera Claudio Gay. Ore deposits and prospects relevant for this paper
are 1: Chuquicamata, 2: Escondida; 3: El Salvador; 3: Potrerillos/El Hueso/Coya; 5: La
Coipa.

Desert of Chile (Fig. 2). Our new age constraints complement


published data for El Hueso and El Salvador (Marsh et al., 1997;
Mote et al., 2001, respectively). Our results conrm the notion that
important differences exist between the timing of uplift in the
Altiplano region and the Southern Atacama Desert and provide new
insights into climate evolution across the Precordillera.

Folding and thrusting in the Precordillera took place during the late
Eocene Incaic Orogeny and is evident in the Potrerillos area (Niemeyer
and Munizaga, 2008; Tomlinson et al., 1994). This orogenic phase led to
uplift, exhumation and supergene oxidation of the El Salvador porphyry
Cu district as early as 36 Ma (see below; Mote et al., 2001; Nalpas et al.,
2005). In the Oligocene, following the Incaic orogeny, a deeply incised
drainage network developed in the Precordillera and valleys formed at
that time were as deep as 2 km below the highest neighbouring
summits, indicating that the Precordillera was already uplifted and
reached altitudes of at least 2000 m (Riquelme et al., 2007). No
signicant movement has been documented on the principal Incaic
faults, which includes the Sierra Castillo fault (Fig. 2) representing the
local segment of the extensive Domeyko Fault system, since the late
Oligocene (Cornejo and Mpodozis, 1996; Niemeyer and Munizaga,
2008). At that time, the focus of thrusting shifted east to the western
edge of the Cordillera Occidental (Cordillera Claudio Gay: Mpodozis and
Clavero, 2002). This shift in the locus of deformation led to the present
day conguration of the internally drained Preandean depression
hosting the Salar de Pedernales (Figs. 1, 2).
The deeply incised Oligocene valleys in the western Precordillera
were lled with continental clastic sediments with a minimum age of
16.3 Ma at their base, as constrained by the oldest intercalated tuff
layers (Nalpas et al., 2008). Inlling of the incised landscape of the
western Precordillera was probably accompanied by pediment
formation as represented by the early Miocene Sierra Checo del
Cobre surface in the Coastal Cordillera (Mortimer, 1973). Low relief
surfaces are present above El Hueso and La Coya in the eastern
Precordillera and are tentatively assigned to the Sierra Checo del
Cobre surface (Fig. 2; Bissig and Riquelme, 2009). A pediplain with a
local base-level in the Salar de Pedernales incised the Sierra Checo del
Cobre surface in the early to middle Miocene (Asientos pediplain:
Bissig and Riquelme, 2009). Later landscape evolution was largely the
result of slow tilting of the Precordillera and Central Depression that
began in the middle Miocene. A relatively low tilting rate resulted in
the middle Miocene alluvial fan backlling in the Central Depression
and the formation of the Atacama Pediplain in the western
Precordillera (Riquelme et al., 2007; Sillitoe et al., 1968). The El
Salvador porphyry Cu deposit is situated at the back-scarp of the
Atacama Pediplain (Fig. 3). The Atacama Pediplain is composite in
nature and likely formed over several stages between ~14 and 10 Ma
(Bissig and Riquelme, 2009). Minimum age constraints for this surface
are given by an ignimbrite deposit covering the pediment surface
dated between 9 and 10 Ma (Clark et al., 1967; Cornejo et al., 1993;
Riquelme et al., 2007), which is in good agreement with the
radiogenic nuclide exposure age of 9 Ma reported by Niishizumi
et al. (2005). A change from alluvial fan backlling to incision of the
Asientos and El Salado canyons into the relict Atacama pediplain has
been interpreted as being the result of slightly increased tilting rates
which allowed the transition from a depositional to erosional regime.
This led to incision of the Salado in the Central Depression and
moderate (b800 m) uplift of the Precordillera in the Late Miocene
(Mortimer, 1973; Riquelme et al., 2007).
3. The use of supergene alunite group minerals

2. Tectonic history and landscape evolution of the southern


Atacama Desert
The present day geomorphologic conguration of the fore-arc
region of northern Chile (1828S) is dominated by extensive
pediplain surfaces which are the result of interaction between
climatic and tectonic evolution during the late Cenozoic (e.g. Lamb
and Davis, 2003; Mortimer, 1973; Rieu, 1975; Riquelme et al., 2003).
These relict pediplains have resisted signicant modication through
erosion for exceptionally long periods of time in some areas (e.g.,
Dunai et al., 2005). The tectonic and geomorphologic evolution of the
studied transect at 262630 S Lat is summarized in the following.

Supergene alunite group minerals (e.g., alunite, natroalunite, jarosite)


are weathering products of porphyry Cu or epithermal deposits and are
found in the leached caps of porphyry Cu deposits (Sillitoe, 2005),
commonly in paleospring settings under acidic uid conditions and
upstream from exotic Cu deposits (Mote et al., 2001). Oxidation of suldes
in porphyry Cu deposits is controlled by the uctuations of the water table
which in turn depends on tectonic and geomorphologic processes as well
as climate (Sillitoe, 2005). Supergene alunite can be dated by the 40Ar/39Ar
method (Vasconcelos, 1999) and although minor recoil loss of 39Ar may
occur in some cases reasonable age dates are usually obtained
(Vasconcelos and Conroy, 2003). Alunite group minerals can also be

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

70

7030

449

6930
C. Doa Ins

A
Not mapped

Q.
ue

Salar de Pedernales

rq
Tu
sa

El Salvador

El S

ald
a

o Ca

nyo

Damiana

26

26

As
ie
nt
os
C
an
yo
n

Jernimo

Potrerillos

El Hueso
2615

2615
C. El Hueso
Coya

Pediment surfaces

Sierra Checos
del Cobre
Asientos

10 km

Early Atacama
(> 14?)

Cerros Bravos

Atacama >10 Ma
Atacama < 10 Ma
Pediment
backscarp
Sierra Castillo Fault
(approx. trace)

A`
El Salado Canyon
Asientos Canyon

2630

Asientos Canyon
Cerros Bravos

m a.s.l. A

A`

El Hueso
Potrerillos Coya

5000
El Salvador

4000

Sierra Checos del Cobre


Asientos

3000
2000

Early Atacama
Atacama pediplain

SCF

20

40

80

80 Km

Fig. 2. Map and cross section of the Precordillera showing principal landscape elements, locations of ore deposits and other features mentioned in the text. A and A' indicate the end
points of the cross section on the map. Cross section is slightly angled at Potrerillos. Modied from Bissig and Riquelme (2009).

analyzed for stable oxygen and hydrogen isotopes to potentially constrain


the paleo meteoric water at the time of its formation (Arehart et al., 1992;
Rye et al., 1992). Since the isotopic composition of meteoric water
depends on elevation (e.g., Poage and Chamberlain, 2001), supergene
alunite has the potential to record uplift histories (Taylor et al., 1997).
However, the isotopic composition of meteoric waters in arid climates
may also be inuenced by evaporation (e.g., Godfrey et al., 2003) and the

relative importance of the latter may be assessed if the tectonic and


geomorphologic framework of a region is independently constrained.
4. Samples and analytical methods
Supergene alunite, natroalunite and jarosite, ranging from powdery to porcellaneous, white to slightly greenish to yellowish veins,

450

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

6937 W

1km

.T

ur

qu

21.5
14.4

es

14.8
16.3
22.9
21.4

El Salvador
townsite
13.6, 13.5, 13.2
13.0, 12.9, 12.0
35.8, 15.3,
14.2, 13.8

2615 S

na

ia

Da

35.4
25.3
13.9
11.1

19.4

Landscape elements

Alunite age
(this study in bold)

Main Atacama

Copper wad age

specimen, where possible using a micro-drill tool. In these cases the


sample for geochronology was extracted rst, followed by the sample
for stable isotope analysis. Sample material for XRD was extracted last
due to the larger amount required. Most 40Ar/39Ar analyses were
performed at the Noble Gas Laboratory, Pacic Centre for Isotopic and
Geochemical Research (PCIGR), University of British Columbia,
Vancouver, BC, Canada, but samples CTB43, CTB46, CTB48 and
CTB49 were dated at the 40Ar/39Ar facility at the Geophysical Institute
at the University of Alaska at Fairbanks (UAF). At PCIGR, the samples
were step-heated at increasing laser powers in the defocused beam of
a 10-W CO2 laser. The ux monitor used was Fish Canyon Tuff
sanidine, 28.02 Ma (Renne et al., 1998). For further details on
analytical methods refer to Bissig et al. (2008). At the UAF, an 8 W
Ar laser was used and the ux monitor was TCR-2 with an age of
27.87 Ma (Lanphere and Dalrymple, 2000); the analytical methods
are described in Layer (2000). All ages are reported with the analytical
errors at the 2 level and represent statistically relevant plateau ages
unless indicated otherwise. The reported plateau ages are all within
error of the corresponding inverse isochron ages. All 40Ar/39Ar data
are included in digital appendices.
The 34S, 18OSO4, D values for alunite were determined at the
Queen's University facility for Isotope Research using a method
modied from Arehart et al. (1992) and Wasserman et al. (1992).
Sulfur was extracted online with continuous-ow technology, using a
Finnigan MAT 252 isotope-ratio mass spectrometer. Sulfate oxygen
was extracted using the technique of Clayton and Mayeda (1963) and
hydrogen was extracted from alunite by pyrolysis. All values are
reported in units of per mil (), and were corrected using NIST
standards 8556 for sulfur, and 8557 for sulfur and oxygen and NIST
8535 for hydrogen. Sulfur is reported relative to Canyon Diablo
Troilite (CDT), oxygen and hydrogen relative to Vienna Standard
Mean Ocean Water (V-SMOW). Analytical precision for both 34S and
18OSO4 values is 0.3 and for D 5.

Early Atacama (>14 Ma)

Exotic Cu deposit

Inselbergs

Primary Cu deposit
(El Salvador)

5. Episodes of supergene mineralization


5.1. El Salvador

B
El Salvador
Town

Damiana

Fig. 3. Environment of exotic mineralization at El Salvador. A) Map of El Salvador and


associated exotic Cu deposits. The principal geomorphologic elements are shown and
approximate locations of sample sites for supergene alunite and copper wad are shown.
Age data from Mote et al. (2001) and this study are indicated, the latter in bold letters
(see Table 1 for more details). B) Photograph taken from upstream of the Damiana
exotic deposit, looking W. The original pediment surface hosting Damiana was
disturbed by mining.

were sampled from surface outcrops. The mineralogy was conrmed


by X-ray diffraction and no signicant contaminating phases (except
for some kaolinite in sample STB012A-2) were identied. Both 40Ar/
39
Ar geochronology and D/H, O and S stable isotope analyses have
been performed on the same samples. The supergene nature of the
alunite was conrmed by S isotope analyses and only samples with
34 S (CDT) between 1.8 and + 3 were considered supergene.
Where the alunite was porcellaneous and not powdery, the analyzed
material was extracted from specic locations within the hand

Supergene mineralization at El Salvador is principally represented by


two exotic deposits, Damiana and Quebrada Turquesa (Figs. 2, 3). Mote
et al. (2001) established an overall age range of 35.4 to 11.1 Ma for
supergene activity mostly on the basis of Mn-oxide ages in the Damiana
exotic deposit. In this study we obtained 6 additional supergene alunite
ages (Fig. 4, Table 1) which conrm the overall age range at El Salvador.
However, at the outcrop scale, the published ages were not reproducible. At Quebrada Riolita, upstream form the Damiana exotic deposit
(Fig. 3, see also Fig. 6 in Mote et al., 2001) two alunite samples extracted
from a horizontal vein were dated (Figs. 3, and 4, Table 1): sample
STB12A-1 represents homogeneous porcellaneous alunite from the
central part of the vein and yielded an 40Ar/39Ar age of 14.22 0.16 Ma.
Sample STB12A-2 represents alunite completely replacing the feldspars
and groundmass from a rhyolitic wall rock clast within the porcellaneous vein and was dated at 35.82 0.95 Ma. Both of our new ages are
considerably older than the 12.89 0.06 to 13.02 0.06 Ma age range
obtained by Mote et al. (2001) from a subhorizontal vein from the same
outcrop. Two additional samples were dated from the brecciated inll of
a steeply dipping fault exposed in the Quebrada Riolita outcrop. The
alunite is porcellaneous and occurs as white to pale yellowish
subangular breccia clasts of less than 1 cm in diameter (Sample
STB12B-1), as well as white to pale greenish alunite groundmass
(Sample STB12B-2), which suggests that alunite was emplaced in at
least two stages separated by fault movement. Alunite extracted from a
clast was dated at 15.31 0.63 Ma whereas alunite form the groundmass yielded an age of 13.83 0.23 Ma. Mote et al. (2001) obtained
younger ages ranging from 13.22 0.12 to 13.61 0.06 Ma from a sub
vertical vein in the same outcrop.

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

451

Fig. 4. 40Ar/39Ar age spectra and inverse isochron diagrams for supergene alunite from El Salvador dated in this study. Samples STB12A-1, 2 and STB12B-1,2 are from Quebrada
Riolita, samples STB22 and STB26 were collected upstream from Quebrada Turquesa.

In the El Salvador district, supergene alunites outcropping


upstream from the Quebrada Turquesa exotic deposit were collected.
Sample STB026 from a powdery white alunite vein yielded a plateau
age of 16.31 0.12 Ma (Figs. 3, 4); an additional sample (STB022)
yielded, in two separate analytical runs, reproducible age spectra with
stepwise increasing ages from ~9 to 14 Ma albeit without attaining a
plateau. This sample is interpreted as a mixture of two or more
generations of ne grained alunite. Mote et al. (2001) reported one
alunite age of 14.8 0.16 Ma as well as supergene Mn oxide ages from

22.9 to 14.4 Ma for Quebrada Turquesa. The geochronological results


suggest that exotic mineralization processes at Damiana apparently
outlasted those at Quebrada Turquesa.
5.2. El Hueso/Potrerillos
Late Miocene supergene activity at El Hueso led to the precipitation
of powdery white alunite within a fracture outcropping on the
uppermost bench of the open pit at 3940 m a.s.l. near the pre-mining

452

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

Table 1
List of new ArAr data.

40

Ar/39Ar data.

Sample

Mineral

Location

STB12A-1
STB12A-2
STB12B-1
STB12B-2
STB22

Alunite
Alunite
Alunite
Alunite
Alunite

ES, Qebr.
ES, Qebr.
ES, Qebr.
ES, Qebr.
ES, Qebr.

STB26
HTB04
CTB43
CTB46
CTB48

Alunite
Alunite
Alunite
Jarosite
Natroalunite

ES, Qebr. Turquesa


El Hueso
Coya, Plateau
Coya Maya
Coya Maya

CTB49

Natroalunite Coya Maya

Riolita
Riolita
Riolita
Riolita
Turquesa

Coord. UTM;
elevation (m)

Plateau age Plateau/39Ar %


(Ma)

443.038/7096.252; 2700
443.038/7096.252; 2700
443.038/7096.252; 2700
443.038/7096.252; 2700
443.881/7096.874; 2770

14.220.16
35.820.95
15.310.63
13.830.23
N/A

10 of 10 steps 100% of 39Ar


5 of 8 steps, 83% of 39Ar
9 of 9 steps, 100% of 39Ar
9 of 9 steps, 100% of 39Ar
N/A

14.31 0.36
36.3 1.4
14.7 1.5
13.64 0.4
N./A

443.701/7096.691;
460.300/7069.153;
461.189/7064.792;
460.554/7065.736;
460.713/7065.450;

16.310.12
8.19 0.1
20.090.14
N/A
0.07 0.6

7 of 9 steps, 81.5 % of 39Ar


7 of 9 steps, 62.8% of 39Ar
3 of 14 steps 83% of 39Ar
N/A
9 of 26 steps 76% 39Ar

15.92 0.31
8.31 0.39
20.11 0.4
4.29 0.12
0.39 1.4

N/A

4.83 0.5

2860
3940
3800
3600
3690

460.482/ 7065.289; 3710 N/A

paleosurface. The alunite yielded an age of 8.19 0.1 Ma (Fig. 5,


Table 1), which is slightly younger than the youngest supergene alunite
age reported by Marsh et al. (1997). These authors report 40Ar/39Ar ages
for supergene alunite from El Hueso of 26 1.4, 12.0. 0.5, 9.6 0.9,
and 9.1 0.5 Ma, plus an additional jarosite age of 6.3 0.5 Ma.
5.3. Coya
At Coya, a porphyry Au prospect 4 km to SE from El Hueso (Figs. 2, 6),
supergene alunite from a fracture inll collected at 3800 m elevation on
the north edge of a prominent plateau assigned to the Sierra Checos
del Cobre surface (Fig. 6; Bissig and Riquelme, 2009) yielded an age
of 20.09 0.14 Ma (Fig. 5). Three samples collected about 500 m N
on a separate hill (Coya Maya, Fig. 6) have also been dated. Sample
CTB-46 represents a jarosite veinlet exposed at an elevation of
3600 m. No statistically signicant plateau age was obtained and the
age spectra from two analytical runs reveal possible 39Ar recoil loss
(Fig. 5). A pseudo-plateau containing only two analytical fractions
yielded an age around 4.4 Ma, which is within error of the inverse
isochron age of 4.29 0.12 Ma obtained from both aliquots (Fig. 5,
Table 1). The latter is taken as the preferred age. A similar age was
obtained for sample CTB-49, which, based on the 40Ar/39Ar and XRD
analysis, consists of natroalunite mixed with minamiite (Na,Ca,K)Al3
(SO 4 ) 2 (OH) 6 ). This sample is also from Coya Maya (3690 m
elevation) and yielded an isochron age of 4.83 0.56 Ma on the
basis of two aliquots, but similar to sample CTB-46, the age spectra
may be affected by 39Ar recoil loss (Fig. 5). Thus, neither of the
aliquots provides a statistically signicant age spectrum, but run 1
yielded a pseudo-plateau age of 5.8 0.8 Ma when the errors are
increased to two sigma on the individual heating steps. Due to the
evidence for recoil effects we prefer the inverse isochron age. An
additional sample of natroalunite (CTB-48) was dated from Coya
Maya. Scanning Electron Microscope energy dispersive analysis
determined the presence of sufcient K for 40Ar/39Ar dating. This
sample, like the other samples from Coya Maya, exhibits evidence for
recoil effects but two analytical runs yielded an age not signicantly
different from zero (Fig. 5, Table 1).
6. Stable isotope constraints
The alunites dated in this study have all been analysed for 34 S,
18OSO4 and D isotopic composition (Fig 7; Table 2). The 34 S values
serve to conrm the supergene nature of the alunite. D values of
hydroxyl groups in the alunite directly reect the meteoric water
compositions at the time of supergene processes, because the
hydrogen isotopic fractionation between water and alunite or
natroalunite is minimal at surface temperatures (Bird et al., 1989;

Inv. isochron Preferred


(Ma)
age

Comment

14.22 0.16
35.82 0.95
15.31 0.63
13.83 0.23
9 to 14 Ma
Mix between 2
or more ages
16.31 0.12
8.19 0.1
20.09 0.14
4.29 0.12
Excess Argon
0
Age based on two aliquots,
excess Ar in spectrum
4.83 0.5
Age based on two aliquots,
excess Ar in spectrum

Rye et al., 1992) and the D of water in equilibrium with alunite is


within the analytical uncertainty from the latter. 18O values on the
sulphate oxygen in the supergene alunite occupy a wide range due to
the incorporation of oxygen both from the water as well as the
atmosphere (Rye et al., 1992).
The late Eocene alunite from Quebrada Riolita yielded a D value of
73 whereas the other alunites from the same location exhibit a
marked increase in D from 61 at 15.4 Ma to 50 at 13.8 Ma
(Fig. 7). Alunites from the headwaters of Quebrada Turquesa exhibit
signicantly higher D values of 34 to 23 at ages younger than
16.3 Ma. The D composition of the 8.2 Ma alunite sample from El
Hueso is at 25, similar to those from Quebrada Turquesa.
At Coya, the early Miocene alunite has a D value of 53,
whereas the early Pliocene natroalunite and jarosite yielded strongly
negative D values of 88 and 97 respectively. The most recent
supergene natroalunite has at 57 a less negative D composition.
7. Discussion
7.1. Chronology of supergene oxidation
As documented for an outcrop near the Damiana exotic deposit,
ages of supergene alunite vary widely within a single outcrop or vein,
indicating that uids from which these supergene minerals precipitate exploit the same permeability network periodically over
extended periods of time. Although this has been known on a
porphyry district scale (Sillitoe, 2005), our data, combined with
published data (Mote et al., 2001) suggest that this is also the case at a
local scale at the Damaina exotic deposit. Here, both within the exotic
deposit as well as at the corresponding paleo spring setting ages range
from about 36 to 13 Ma, indicating that exotic mineralization
processes operated periodically over 23 Ma in an individual ore
forming system. Thus, the permeability network exploited by
supergene uids remained active over an extended period of time
and implies that the local geomorphologic conguration has not
changed substantially. Although the pediment hosting Damiana has
likely experienced modications and was shaped most recently
during the formation of the multi-stage Atacama pediplain, erosion
was never substantial enough to strip the gravels down to the
supergene ore.
The timing of the cessation of supergene activity in the Central
Depression and western Precordillera, proposed at ca. 13 Ma (Mote
et al., 2001), has been roughly conrmed. The respective youngest
supergene ages of Damiana and Quebrada Turquesa correspond to the
inferred relative ages of the pediment surfaces hosting these two
exotic deposits (Figs. 3, 8), indicating a potential link between local
pediment formation and exotic mineralization. The cessation of

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

30

12
HTB4 Alunite

10

Age (Ma)

453

CTB43 Alunite

8.19 0.1 Ma

25
20

8
20.09 0.14 Ma

15
6

10

4
2

5
0

20

40

60

80

Cumulative 39Ar percent

100

20

40

60

80

Cumulative 39Ar percent

100

.004

.0030

HTB4 Alunite

CTB43 Alunite

36Ar/40Ar

.0026
.003

Inverse isochron
8.31 0.39 Ma

.0022

Inverse Isochron
20.36 0.95 Ma

.002
.0014
.001
.0010
.0006
0.2

0.4

0.3

0.6

0.7

0.8

.002

39Ar/40Ar

30

25

CTB46 Jarosite, run 1

Age (Ma)

25
20

50

CTB48, Natroalunite, run 1

20

40

15

30

10

20

10

CTB49, Natroalunite, run 1

3 steps at ~5.8 +/- 0.8 Ma

15
2 steps at ~4.4 Ma

10
5
0

20

40

60

Cumulative

39Ar

80

100

20

percent

40

60

Cumulative

30

39Ar

80

20

100

60

80

100

50
CTB48, Natroalunite, run 2

25

40

Cumulative 39Ar percent

percent

30
CTB46 Jarosite, run 2

Age (Ma)

.008

.006
39Ar/40Ar

CTB49, Natroalunite, run 2

24

40

18

30

12

20

10

20
15
10
5
0

20

40

60

Cumulative
.004

39Ar

80

20

.001

40

60

Cumulative
.004

Inverse isochron
4.29 +/- 0.12 Ma

36Ar/40Ar

percent

CTB46 Jarosite, 2 runs

.003

100

39Ar

80

.004

Reference zero age line

.005

.01

.015

39Ar/40Ar

.02

.025

60

80

100

CTB49, Natroalunite, 2 runs

Inverse Isochron
4.83 +/- 0.28 Ma

.002

.002

.001

.001
(calculated age excluding large error fractions)

0
0

40

.003

(arrows denote fractions used in


age calculation)

20

Cumulative 39Ar percent

percent

CTB48, Natroalunite, 2 runs

.003

100

.005

.01

.015

39Ar/40Ar

.02

.025

.001

.002

.003

.004

.005

39Ar/40Ar

Fig. 5. 40Ar/39Ar age spectra and inverse isochron diagrams for supergene alunite group minerals from El Hueso and Coya dated in this study. Sample HTB04 is from El Hueso, the
remainder of samples are from Coya.

454

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

Cordillera Claudio Gay

Relict Asientos pediplain


n

s Canyo

Asiento

Coya

a: 20.09 +/- 0.14 Ma

n: ~4.8 Ma
and zero age
j: ~4.3 Ma

Coya Maya

Fig. 6. View from Cerro El Hueso (see Fig. 2 for location) towards the E showing the Coya prospect with sample locations and supergene alunite (a), natroalunite (n) and jarosite
(j) ages. The horizon is represented by the Cordillera Claudio Gay. Gravel covered relics of the Asientos Pediplain are indicated.

supergene alunite precipitation at El Salvador occurred at a similar


time as in porphyry Cu and epithermal districts farther North (e.g.,
Arancibia et al., 2006; Bouzari and Clark, 2002; Hartley and Rice, 2005;
Sillitoe and McKee, 1996), which, together with other paleoclimatic
evidence (Alpers and Brimhall, 1988; Rech et al., 2006) indicates
climate desiccation in the middle Miocene (Fig. 8).
The periods of most intense supergene activity in the late
Oligocene and Middle Miocene originally dened for northern Chile
and southern Peru (Sillitoe and McKee, 1996) become more blurry as
more geochronological data become available (Hartley and Rice,
2005) and recent studies suggest a continuous period of intense
supergene processes lasting from the late Eocene to the early late
Miocene in Northern Chile (Arancibia et al., 2006). Our results are
consistent with a prolonged history of supergene mineralization for
the El Salvador district.
In the eastern Precordillera at El Hueso and Coya, at elevations
approximately 10001200 m higher than at El Salvador, 40Ar/39Ar
constraints, admittedly still limited, indicate that supergene processes occurred in the late Oligocene and early Miocene as well
as from the late Miocene to early Pliocene and may still be occurring at the present day (Fig. 8). Contrasting with El Salvador, supergene oxidation in the eastern Precordillera appears to have been
limited throughout the middle Miocene. While the late Oligocene
and early Miocene ages roughly coincide with the incision of the
Sierra Checos del Cobre and Asientos pediplains (Fig. 8) and supergene oxidation may have been related to these erosive processes, we
interpret the Late Miocene and younger oxidation to be controlled
by uplift to elevations sufcient to capture increased precipitation
combined with the incision of deep canyons into the previous planar
landscape (Bissig and Riquelme, 2009). This would lead to
depression of the water table, but increased availability of meteoric
water in the vadose zone, generating conditions favorable for sulde
oxidation.
7.2. D through time
The Late Eocene meteoric water at El Salvador was at D = 73
similar to the present day precipitation at ~ 3500 m a.s.l. when
calculated using the empirical relationship for South America from
Poage and Chamberlain (2001). The estimated elevation for the Late

Eocene would be no more than 500 m lower if the long term oxygen
isotopic variations in seawater (Zachos et al., 2001) are considered.
Miocene meteoric waters are considerably less deuterium depleted
and the least negative D values of 23 to 34 were obtained for
samples between 8.2 and 16.3 Ma from both El Hueso and El Salvador.
Early Pliocene waters at Coya were at D = 88 to 97 similar to
present day precipitation around the 38004000 m elevation at which
Coya is presently situated (Poage and Chamberlain, 2001). The most
recent sample yielded a less negative D value of 57. Our data
starkly contrast earlier work (Taylor et al., 1997) which suggests
sharply decreasing D values from ~15 in the Late Oligocene to as
much as 60 in the middle to late Miocene which they interpret as
evidence for a marked uplift pulse in the Middle Miocene. The
discrepancy between the two datasets can probably be explained by
the different scales of the two studies. Taylor et al. (1997) analyzed
alunite samples from 20 to 27 S Lat S (see also Sillitoe and McKee,
1996) which likely reect signicant along strike variations in
geomorphology, uplift history and climate. North of about 23 Lat. S,
there is no Preandean Depression (Fig. 1) and independent evidence
suggests that much of the uplift of the Altiplano has occurred in the
middle or late Miocene (e.g., Gregory-Wodzicki, 2000; Hoke et al.,
2007). In the southern Atacama Desert, the Precordillera attained
elevations of at least 2000 m in the early Oligocene (Riquelme et al.,
2007) and our stable isotope data suggest a Late Eocene elevation of
3000 m a.s.l. or more for the Precordillera near El Salvador. These high
elevations may be attributed to intense folding and thrusting
(Niemeyer and Munizaga, 2008) and crustal thickening (e.g., Haschke
et al., 2002) affecting the region in the late Eocene.
The increasing D values throughout the middle Miocene are
contrary to the trend expected for an uplifting mountain range.
However, the isotopic composition of meteoric water is not only
controlled by orographic effects, but also by evaporation and recycling
of meteoric water (Godfrey et al., 2003). Thus, we interpret the higher
than expected Miocene D values largely as an effect of evaporation.
Bird et al. (1989) and Sillitoe (2005) suggested that high rates of
evaporation are conducive for supergene alunite formation, providing
support to our interpretation. Thus, the least negative D values
would coincide with the most intense evaporation and hyper arid
conditions which likely persisted between about 15 and 8 Ma. The
timing of the onset of hyper-arid conditions is also recorded by a

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

A
Nwa Chile
te r
M
lin e t e
e
or

ic

0
-20

-60
Altiplano
~4500 m
a.s.l.

-80
-100

jarosite

supergene alunite
sulfate field

wa
te

rd

-120
-140

air dominan
t

om
ina

nt

D (VSMOW)

-40

-160
-20

-15

-10

-5

10

15

18OSO4(VSMOW)

B
-20

ect

2500

ic eff
raph

3000

Dess

Orog

D (VSMOW)

-40

-60

icatio

n tren

455

marked decrease of sediment accumulation in the Central Depression


in the middle Miocene (Nalpas et al., 2008; Riquelme et al., 2007).
Given that the Precordillera attained considerable elevations significantly prior to the middle Miocene hyper arid climate, the uplift of
the mountain range probably does not by itself account for the climate
desiccation in the southern Atacama Desert (see also Lamb and Davis,
2003). However, the eastward migration of the deformation into the
Cordillera Claudio Gay in the late Oligocene (Mpodozis and Clavero,
2002) and the formation of the Preandean depression likely enhanced
aridication of the Central Depression. We suggest that the widening
rather than simply the uplift of the Andes likely has resulted in
increased rain shadow effects at the western Andean slope.
Somewhat wetter conditions probably dominated the early
Pliocene in the Precordillera when compared to the arid middle
Miocene climate. Stable isotope evidence suggests that the Precordillera probably had attained elevations similar to the present and that
evaporation effects were limited. This is interpreted as the result of
increased capture of orographically controlled precipitation at that
time. Moreover, sedimentological evidence in the Calama basin, some
400 km farther north (Fig. 1; Hartley and Chong, 2002), indicates that
semiarid climatic conditions prevailed in the Precordillera and
western Andes between about 6 and 3 Ma.
The present climate and hydrologic conditions in the eastern
Precordillera are potentially still wet enough to permit the formation
of supergene alunite group minerals, but signicant evaporation likely
affects the meteoric waters. Strong evaporation effects have been
documented for meteoric waters in the internally drained basins of
the Salar de Hombre Muerto and Salar de Atacama basins (Godfrey
et al., 2003).

3500

8. Conclusions
-80
4000
-100

10

20
40Ar/39Ar

30

40

age (Ma)

Symbols
El Salvador, Damiana
Other El Salvador

Coya
El Hueso

Fig 7. Stable isotope composition of supergene alunite group minerals. A) D vs.


18so4. Note that all alunite samples fall within the large eld for supergene alunite
but generally closer to the air dominated than water dominated oxygen isotope
composition. The jarosite sample plots immediately right of the air dominated
boundary for supergene alunite. Reference eld for high altitude precipitation for the
Chilean Altiplano is from Herrera et al. (2006). B) D isotopic composition of supergene
alunite group minerals through time. The right vertical axis is labeled with the
elevations corresponding to the D values on the left axis. Values were calculated using
the empirically determined relationship for central and South America (Poage and
Chamberlain, 2001). Interpreted general climatic trends are indicated (see text for
discussion).

Table 2
Stable isotope data.
Sample

Mineral

dD

d34S

d18OSO4

age (Ma)

HTB004
STB-022
STB-026
STB-12A-1
STB-12A-2
STB-12B-1
STB-12B-2
CTB-43
CTB-46
CTB-48
CTB-49

Alunite
Alunite
Alunite
Alunite
Alunite
Alunite
Alunite
Alunite
Jarosite
Natroalunite
Natroalunite

25
23
34
54
74
61
50
53
97
57
88

1.8
1.4
0.5
0.9
0.3
0.1
0.0
0.0
0.8
1.1
3.0

3.7
4.8
6.8
5.3
9.7
4.1
3.9
10.7
11.0
2.8
2.6

8.19 0.1
9 to 14
16.31 0.12
14.22 0.16
35.8 1
15.3 0.6
13.8 0.2
20.1 0.1
4.3 0.1
0
4.8 0.6

Geomorphologic and stable isotope evidence strongly suggests


that the Precordillera in the Southern Atacama Desert has attained
elevations of at least 3000 m a.s.l. already in the early Oligocene
and thus, signicantly prior to the major uplift of the Altiplano.
The climate evolved differently in the western Precordillera and
Central Depression from the eastern Precordillera. The cessation of
supergene processes at El Salvador around 13 Ma has been
conrmed and is attributed to climate desiccation, an interpretation also supported by sedimentological and stable isotope
evidence. However, conditions at Coya and El Hueso in the Eastern
Precordillera, situated near 4000 m present day elevation
remained conducive for at least episodic supergene alunite
formation until the early Pliocene, and possibly up to the present
day. Uplift to elevations near 4000 m a.s.l. have led to increased
capture of moisture and consequently increased availability of
meteoric waters.
The new 40Ar/39Ar age constraints presented herein provide
evidence conrming the previously proposed protracted history
of the Damiana exotic Cu deposit and indicate that the local
geomorphologic and hydrologic conguration has remained
relatively stable over 23 Ma.
Supplementary data to this article can be found online at doi:
10.1016/j.epsl.2010.09.028.

Acknowledgements
This study has been funded by Fondo Nacional de Desarrollo
Cientco y Tecnolgico de Chile (Fondecyt) grant # 11060516. Kerry
Klassen is thanked for the stable isotope analyses whereas Paul Layer
and Tom Ullrich provided the Ar/Ar analyses. Fritz Schlunegger and an
anonymous EPSL reviewer are thanked for their constructive reviews.
This is MDRU publication P-264.

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457

Climate
Literature This study
WP

EP

moderate tilting
in the fore-arc

Pliocene

hyper arid

Atacama gravel deposition

Canyon Incision

Coya

El Hueso

Tectonics

Late
Miocene

10

Other, El Salvador

Q. Turquesa, El Salvador

Damiana, El Salvador

Landscape

Supergene ages

Pediment formation

456

A3
A2
A1

?
3

Oligocene
?

35

40

thrusting and
folding, Potrerillos
Fold and Thrust
belt

Eocene

30

SC

semi-arid

25

thrusting and
uplift, Cordillera
Claudio Gay

semi-arid

20

Early
Miocene

AS

Hyper aridity
(Hartley and Chong, 2002)

Hyper aridity
(Alpers and Brimhall, 1988)

15

slow tilting
in the fore-arc

Middle
Miocene

Fig. 8. Chart integrating landscape chronology, tectonic episodes, ages of supergene minerals and climate. Abbreviations: A1: early stage Atacama pediplain, A2: Main stage Atacama
pediplain; A3: late stage Atacama pediplain; AS: Asientos surface; SC: Sierra Checos del Cobre surface. WP: Western Precordillera; EP: Eastern Precordillera. Supergene ages are
plotted individually (black bars; bold correspond to this study) or as groups of ages (boxes; number of dates indicated). References as follows: supergene ages from El Hueso: Marsh
et al. (1997); supergene ages from El Salvador: Mote et al. (2001); Pediment formation: Mortimer (1973), Sillitoe et al. (1968), Bissig and Riquelme (2010); canyon incision:
Riquelme et al. (2003, 2007); Gravel deposition: Riquelme et al. (2007); Tectonic episodes: Niemeyer and Munizaga (2008), Mpodozis and Clavero (2002), Riquelme et al. (2003,
2007).

References
Alpers, C.N., Brimhall, G.H., 1988. Middle Miocene climatic change in the Atacama
Desert, northern Chile: evidence from supergene mineralization at La Escondida.
Geol. Soc. Am. Bull. 100, 16401656.
Arancibia, G., Matthews, S.J., De Arce, C.P., 2006. KAr and 40Ar/39Ar geochronology of
supergene processes in the Atacama Desert, Northern Chile: tectonic and climatic
relation. J. Geol. Soc. 163, 107118.
Arehart, G.B., Kesler, S.E., Oneil, J.R., Foland, K.A., 1992. Evidence for the supergene
origin of alunite in sediment-hosted micron gold deposits, Nevada. Econ. Geol. 87,
263270.
Bird, M.I., Andrew, A.S., Chivas, A.R., Lock, D.E., 1989. An isotopic study of surcial
alunite in Australia 1: hydrogen and sulphur isotopes. Geochim. Cosmochim. Acta
53, 32233237.
Bissig, T., Riquelme, R., 2009. Contrasting landscape evolution and development of
supergene enrichment in the El Salvador porphyry Cu and Potrerillos-El Hueso CuAu
districts, Northern Chile. In: Titley, S. (Ed.), Society of Economic Geologists Special
Publication No. 14, Supergene Environments, Processes and Products, pp. 5968.
Bissig, T., Ullrich, T.D., Tosdal, R.M., Friedman, R., Ebert, S., 2008. The timespace
distribution of Eocene to Miocene magmatism in the Central Peruvian polymetallic
province and its metallogenetic implications. J. S. Am. Earth Sci. 26, 1635.
Bouzari, F., Clark, A.H., 2002. Anatomy, evolution, and metallogenic signicance of the
supergene orebody of the Cerro Colorado porphyry copper deposit; I Region,
northern Chile. Econ. Geol. 97, 17011740.
Clark, A.H., Mayer, A.E.S., Mortimer, C., Sillitoe, R.H., Cooke, R.U., Snelling, N.J., 1967.
Implications of the isotopic ages of ignimbrite ows, southern Atacama Desert,
Chile. Nature 215, 723724.

Clayton, R.N., Mayeda, T.K., 1963. The use of bromine pentauoride in the extraction of
oxygen from oxides and silicates for isotopic analysis. Geochim. Cosmochim. Acta
27, 4352.
Cornejo, P., Mpodozis, C., 1996. Geologa de la Region de Sierra Exploradora (Cordillera
de Domeyko, 2526S). Servicio Nacional de Geologa y Mineria-CODELCO,
Informe Registrado IR-96-09, 330 p. Santiago.
Cornejo, P., Mpodozis, C., Ramirez, C.F., Tomlinson, A.J., 1993. Estudio Geolgico de la
Regin de Potrerillos y El Salvador (26-27 Lat.S). Servicio Nacional de Geologa y
Minera-CODELCO, Informe Registrado IR-93-01, 12 cuadrngulos escala 1:50.000 y
vol.texto 258 p. Santiago.
Dunai, T.J., Gonzlez, G., Juez-Larr, J., 2005. OligoceneMiocene age of aridity in the
Atacama Desert revealed by exposure dating of erosion-sensitive landforms.
Geology 33, 321324.
Faras, M., Charrier, R., Compte, D., Martinod, J., Hrail, G., 2005. Late Cenozoic
deformation and uplift of the western ank of the Altiplano: evidence from the
depositional, tectonic, and geomorphologic evolution and shallow seismic activity
(northern Chile at 1930S). Tectonics 24. doi:10.1029/2004TC001667.
Garcia, M., Hrail, G., 2005. Fault-related folding, drainage network evolution and
valleyincision during the Neogene in the Andean Precordillera of Northern Chile.
Geomorphology 65, 279300.
Garzione, C.N., Hoke, G.D., Libarkin, J.C., Withers, S., MacFadden, B., Eiler, J., Prosenjit, G.,
Mulch, A., 2008. Rise of the Andes. Science 320, 13041307.
Godfrey, L.V., Jordan, T.E., Lowenstein, T.K., Alonso, R.L., 2003. Stable isotope constraints
on the transport of water to the Andes between 22 and 26S during the last glacial
cycle. Paleogeocgraphy, Paleoclimatology, Paleoecology 194, 299317.
Gregory-Wodzicki, K., 2000. Uplift history of central and northern Andes: a review.
Geol. Soc. Am. Bull. 112, 10911105.

T. Bissig, R. Riquelme / Earth and Planetary Science Letters 299 (2010) 447457
Gustafson, L.B., Orquera, W., McWilliams, M., Castro, M., Olivares, O., Rojas, G.,
Malvenda, J., Mendez, M., 2001. Multiple Centers of Mineralization in the Indio
Muerto District, El Salvador, Chile. Econ. Geol. 96, 325350.
Hartley, A.J., Chong, G., 2002. Late Pliocene age for the Atacama Desert: implications for
the desertication of western South America. Geology 30, 4346.
Hartley, A.J., Rice, C.M., 2005. Controls on supergene enrichment of porphyry copper
deposits in the Central Andes: a review and discussion. Mineralium Deposita 40,
515525.
Haschke, M., Siebel, W., Gnther, A., Scheuber, E., 2002. Repeated crustal thickening and
recycling during the Andean orogeny in north Chile (21_26_S). Journal of
Geoplysical Research 107 B1:ECV6-1ECV6-18.
Herrera, C., Pueyo, J.J., Saez, A., Valero-Garces, B.L., 2006. Relacin de aguas superciales
y subterrneas en el rea del lago Chungar y lagunas de Cotacotani, norte de Chile:
un estudio isotpico. Rev. Geol. Chile 33, 299325.
Hoke, G.D., Isacks, B.L., Jordan, T.E., Blanco, N., Tomlinson, A.J., Ramezani, J., 2007.
Geomorphic evidence for post 10 Ma uplift of the western ank of the central
Andes 183022S TC5021, Tectonics 26. doi:10.1029/2006TC002082.
Lamb, S., Davis, P., 2003. Cenozoic climate change as a possible cause for the rise of the
Andes. Nature 425, 792797.
Lanphere, M.A., Dalrymple, G.B., 2000. First-principles calibration of 38Ar tracers:
implications for the ages of 40Ar/39Ar uence monitors. U.S. Geological Survey
Professional Paper 1621. 10 p.
Layer, P.W., 2000. 40Argon/39Argon age of the El'gygytgyn impact event, Chukotka,
Russia. Meteroitics and Planatary Science. 35, 591599.
Marsh, T.M., Einaudi, M.T., Mcwilliams, M., 1997. 40Ar/39Ar geochronology of CuAu and
AuAg mineralization in the Potrerillos district. Chile: Economic Geology 92, 784806.
Mortimer, C., 1973. The Cenozoic history of the southern Atacama Desert, Chile. J. Geol.
Soc. Lond. 129, 505526.
Mote, T.I., Becker, T.A., Renne, P., Brimhall, G.H., 2001. Chronology of exotic mineralization
at El Salvador, Chile by 40Ar/39Ar dating of copper wad and Supergene Alunite. Econ.
Geol. 96, 351366.
Mpodozis, C., Clavero, J., 2002. Tertiary tectonic evolution of the southwestern edge of
the Puna Plateau : Cordillera Claudio Gay (2627 S), Northern Chile. Andean
geodynamics: extended abstracts: Paris/Toulouse: Institut de recherche pour le
dveloppement. IRD Universit Paul Sabatier. Toulouse, France, pp. 445448.
Nalpas, T., Hrail, G., Mpodozis, C., Riquelme, R., Clavero, J., Dabard, M.P., 2005.
Thermochronologicals data and denudation history along a transect between
Chaaral and Pedernales (~ 26S), north Chilean Andes: orogenic implications. 6th
International Symposion on Andean Geodynamics (ISAG), Barcelona. Extended
Abstracts, Spain, pp. 548551.
Nalpas, T., Dabard, M.-P., Ruffet, G., Vernon, A., Mpodozis, C., Loi, A., Hrail, G., 2008.
Sedimentation and preservation of the Miocene Atacama Gravels in the Pedernales
Chaaral area, Northern Chile: climatic or tectonic control? Tectonophysics 459,
161173.
Niemeyer, H., Munizaga, R., 2008. Structural control of the emplacement of the
Portrerillos porphyry copper, central Andes of Chile. J. S. Am. Earth Sci. 26, 261270.
Nishiizumi, K., Caffee, M.W., Finkel, R.C., Brimhall, G., Mote, T., 2005. Remnants of a fossil
alluvial fan landscape of Miocene age in the Atacama Desert of northern Chile using
cosmogenic nuclide exposure age dating. Earth Planet. Sci. Lett. 237, 499507.
Poage, M.E., Chamberlain, C.P., 2001. Empirical relationship between elevation and the
stable isotope composition of precipitation and surface waters: considerations for
studies on paleoelevation change. Am. J. Sci. 301, 115.
Rech, J.A., Currie, B.S., Michalski, G., Cowan, A.M., 2006. Neogene climate change and
uplift in the Atacama Desert, Chile. Geology 34, 761764.
Renne, P.R., Swisher III, C.C., Deino, A.L., Karner, D.B., Owens, T., DePaolo, D.J., 1998.
Intercalibration of standards, absolute ages and uncertainties in 40Ar/39Ar dating.
Chem. Geol. 145, 117152.
Rieu, M., 1975. Les formations sedimentaires de la Pampa del Tamarugal et le Ro Loa
(Norte Grande du Chili). Fr. Off. Rech. Sci. Tech. Outre-Mer, Cah., Ser. Geol. 7 (2),
145164.

457

Riquelme, R., Martinod, J., Hrail, G., Darrozes, J., Charrier, R., 2003. A geomorphological
approach to determining the Neogene to Recent tectonic deformation in the
Coastal Cordillera of northern Chile (Atacama). Tectonophysics 361, 255275.
Riquelme, R., Hrail, G., Martinod, J., Charrier, R., Darrozes, J., 2007. Late Cenozoic
geomorphologic signal of forearc deformation and tilting associated with the uplift
and climate changes of the Andes, Southern Atacama Desert (26S28S).
Geomorphology 86, 283306.
Riquelme, R., Darrozes, J., Maire, E., Hrail, G., Soula, J.C., 2008. Long-term denudation
rates from the Central Andes (Chile) estimated from a Digital Elevation Model
using the Black Top Hat function and Inverse Distance Weighting: implications for
the Neogene climate of the Atacama Desert. Rev. Geol. Chile 35, 105121.
Rivera, S.L., Vila, T., Osorio, J., 2004. Geologic characteristics and exploration
signicance of gold-rich porphyry copper deposits in the El Salvador region,
Northern Chile. In: Sillitoe, R.H., Perell, J., Vidal, C.E. (Eds.), Andean Metallogeny:
New Discoveries, Concepts, and Updates. Society of Economic Geologists Special
publication, 11, pp. 97111.
Rye, R.O., Bethke, P.M., Wasserman, M.D., 1992. The stable isotope geochemistry of acid
sulfate alteration. Econ. Geol. 87, 225262.
Schlunegger, F., Zeilinger, G., Kounov, A., Kober, F., Hsser, B., 2006. Scale of relief
growth in the forearc of the Andes of NorthernChile (Arica latitude, 18_S). Terra
Nova 18, 217223.
Sillitoe, R.H., 2005. Supergene oxidized and enriched porphyry copper and related
deposits. In: Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.),
Economic Geology 100th Anniversary volume, pp. 723768.
Sillitoe, R.H., McKee, E.H., 1996. Age of supergene oxidation and enrichment in the
Chilean porphyry copper province. Econ. Geol. 91, 164179.
Sillitoe, R.H., Mortimer, C., Clark, A.H., 1968. A chronology of landform evolution and
supergene mineral alteration, southern Atacama Desert, Chile. Institution of Mining
and Metallurgy Transactions, section B. 77, 166169.
Taylor, B.E., McKee, E.H., Sillitoe, R.H., 1997. D and 18O maps of South American
meteoric waters: contrasts between the present and Tertiary and their
implications for Andean uplift. Geological Association of Canada/Mineralogical
Association of Canada annual meeting, Ottawa, Canada, abstracts with programs,
A-146.
Thompson, J.F.H., Gale, V.G., Tosdal, R.M., Wright, W.A., 2004. Characteristics and
Formation of the Jernimo carbonate-replacement gold deposit, Potrerillos district,
Chile. In: Sillitoe, R.H., Perell, J., Vidal, C.E. (Eds.), Andean Metallogeny: New
Discoveries, Concepts, and Updates. Society of Economic Geologists Special
publication, 11, pp. 7595.
Thouret, J.-C., Woerner, G., Gunnell, Y., Singer, G., Zhang, X., Souriot, T., 2007.
Geochronologic and stratigraphic constraints on canyon incision and Miocene
uplift of the Central Andes in Peru. Earth Planet. Sci. Lett. 263, 151166.
Tomlinson, A.J., Mpodozis, c, Cornejo, P., Ramirez, C.F., Dumitru, T., 1994. El Sistema de
fallas Sierra Castillo-Agua Amarga: transpresion sinistral Eocena en la precordillera
de Potrerillos-El Salvador. 7 Congreso Geolgico Chileno, actas 14591463.
Vasconcelos, P.M., 1999. 40Ar39Ar geochronology of supergene processes in ore
deposits. In: Lambert, D.D., Ruiz, J. (Eds.), Reviews in Economic Geology 12,
Application of Radiogenic Isotopes to Ore Deposit Research and Exploration,
pp. 73113.
Vasconcelos, P.M., Conroy, M., 2003. Geochronology of weathering and landscape
evolution, Dugald River valley, NWQueensland, Australia. Geochim. Cosmochim.
Acta 67, 29132930.
Victor, P., Oncken, O., Glodny, J., 2004. Uplift of the western Altiplano plateau: evidence
from the Precordillera between 20 and 21S (northern Chile). Tectonics 23.
doi:10.1029/2003TC001519 TC4004.
Wasserman, M.D., Rye, R.O., Bethke, P.M., Arribas Jr., A., 1992. Methods for
separation and total stable isotope analysis of alunite. U. S. Geol. Surv. OpenFile Rep, pp. 9299.
Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, K., 2001. Trends, rhythms, and
aberrations in global climate 65 Ma to present. Science 292, 686693.

You might also like