Vous êtes sur la page 1sur 13

Applied Geochemistry 25 (2010) 11071119

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: www.elsevier.com/locate/apgeochem

Geochemistry of highly acidic mine water following disposal into a natural lake
with carbonate bedrock
Christian Wisskirchen a,*, Bernhard Dold a,b, Kurt Friese c, Jorge E. Spangenberg a, Peter Morgenstern d,
Walter Glaesser e
a

Institute of Mineralogy and Geochemistry, University of Lausanne, CH-1015 Lausanne, Switzerland


Instituto de Geologa Economica Aplicada, Universidad de Concepcin, Concepcin, Chile
UFZ Helmholtz Centre for Environmental Research, Department of Lake Research, D-39114 Magdeburg, Germany
d
UFZ Helmholtz Centre for Environmental Research, Department of Analytical Chemistry, D-04318 Leipzig, Germany
e
Institute of Geophysics and Geology, University of Leipzig, D-04211 Leipzig, Germany
b
c

a r t i c l e

i n f o

Article history:
Received 21 November 2008
Accepted 26 April 2010
Available online 29 April 2010
Editorial handling by B. Wang

a b s t r a c t
Acid mine drainage (AMD) from the ZnPb(AgBiCu) deposit of Cerro de Pasco (Central Peru) and
waste water from a Cu-extraction plant has been discharged since 1981 into Lake Yanamate, a natural
lake with carbonate bedrock. The lake has developed a highly acidic pH of 1. Mean lake water chemistry
was characterized by 16,775 mg/L acidity as CaCO3, 4330 mg/L Fe and 29,250 mg/L SO4. Mean trace element concentrations were 86.8 mg/L Cu, 493 mg/L Zn, 2.9 mg/L Pb and 48 mg/L As, which did not differ
greatly from the discharged AMD. Most elements showed increasing concentrations from the surface to
the lake bottom at a maximal depth of 41 m (e.g. from 3581 to 5433 mg/L Fe and 25,609 to 35,959 mg/L
SO4). The variations in the H and O isotope compositions and the element concentrations within the
upper 10 m of the water column suggest mixing with recently discharged AMD, shallow groundwater
and precipitation waters. Below 15 m a stagnant zone had developed. Gypsum (saturation index,
SI  0.25) and anglesite (SI  0.1) were in equilibrium with lake water. Jarosite was oversaturated
(SI  1.7) in the upper part of the water column, resulting in downward settling and re-dissolution in
the lower part of the water column (SI  0.7). Accordingly, jarosite was only found in sediments from
less than 7 m water depth. At the lake bottom, a layer of gel-like material (90 wt.% water) of pH 1 with
a total organic C content of up to 4.40 wet wt.% originated from the kerosene discharge of the Cu-extraction plant and had contaminant element concentrations similar to the lake water. Below the organic layer
followed a layer of gypsum with pH 1.5, which overlaid the dissolving carbonate sediments of pH 5.37.
In these two layers the contaminant elements were enriched compared to lake water in the sequence
As < Pb  Cu < Cd < Zn = Mn with increasing depth. This sequence of enrichment was explained by the
following processes: (i) adsorption of As on Fe-hydroxides coating plant roots at low pH (up to
3326 mg/kg As), (ii) adsorption at increasing pH near the gypsum/calcite boundary (up to 1812 mg/kg
Pb, 2531 mg/kg Cu, and 36 mg/kg Cd), and (iii) precipitation of carbonates (up to 5177 mg/kg Zn and
810 mg/kg Mn; all data corrected to a wet base). The inltration rate was approximately equal to the discharge rate, thus gypsum and hydroxide precipitation had not resulted in complete clogging of the lake
bedrocks.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Management of acid mine drainage (AMD) is a major problem
for the present day mining industry. Sulde-rich deposits tend to
produce highly acidic waters, if no carbonate minerals neutralize
the efuents (Plumlee and Logsdon, 1999). Pit lakes have become
a subject of increasing importance in environmental geochemistry,
* Corresponding author. Present address: Golder associates S.A., Av.11 de
Septiembre 2353 Piso 2, Providencia, Santiago, Chile. Tel.: + 56 2 594 2024;
Fax: +56 2 594 2001.
E-mail address: ChristianWisskirchen@web.de (C. Wisskirchen).
0883-2927/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apgeochem.2010.04.015

due to the closure of open pit mines (Castro and Moore, 2000;
Geller et al., 1998; Snchez Espaa et al., 2008). The main sources
of acidity and metals in pit lakes include oxidation of suldes from
coal and metalliferous deposits or backll, water from interconnected ooded underground mines or dissolution of sulfates by
water ushes (Bowell and Parshley, 2005; Castendyk et al., 2005;
Denimal et al., 2005; Knller et al., 2004; Pellicori et al., 2005). In
most pit lakes, the sources of contaminants are diffuse and difcult
to apportion. In pit lakes with carbonate wall rocks or carbonatecontaining veins the system is buffered to slightly basic pH,
limiting the solubility of most metals (Shevenell et al., 1999).

1108

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

Conclusive interpretation by modeling of data sets for a number of


pit lakes has suggested control of aqueous element concentrations
by mineral equilibrium (Eary, 1999). Addition of lime is still the
only successful method to increase the alkalinity of highly acidic
pit lakes. Addition of nutrients (phosphate or organic matter) to
stimulate the microbiological activity in order to increase the pH
has only been applied with good results for pit lakes with pH >3
(Koschorreck et al., 2007).
For AMD neutralization, carbonate minerals are an alternative
to lime, because of their lower cost. Several studies have been
carried out in recent years to investigate the interactions of carbonates with mine efuents, and their application to AMD neutralization and metal retention (Komnitsas et al., 2004; Simon et al.,
2005; Webb and Sasowsky, 1994). For low mineralized AMD,
anoxic and oxic limestone drains have been used successfully
(Cravotta and Trahan, 1999; Santomartino and Webb, 2007). However, the problem of armoring of the carbonate grains and clogging
of the limestone drains by precipitation of Fe- and Al-hydroxides
and gypsum still limits the life-span of these treatment systems
(e.g. Cravotta, 2008; Rose et al., 2004).
In this article, results are presented of a detailed geochemical
study of a natural lake within a carbonate bedrock used as an
AMD storage facility for the polymetallic ZnPb(AgBiCu) deposit Cerro de Pasco (Central Peru). The AMD with a pH 1 has
been discharged without treatment into Lake Yanamate (3 km2)
since 1981. The formation of AMD at Cerro de Pasco and its handling has been studied with regard to the Excelsior waste rock
dump and tailings partly underlying it (Dold et al., 2009; Smuda
et al., 2007). Additionally, spoil waters with high concentrations
of kerosene from a solvent extraction-electrowinning (SX-EW)
plant were discharged into the lake until August 2001. The geochemical approach used in this rst case study of a discharge pond
in carbonate rocks combines major and trace element concentration of waters and sediments, H and O stable isotopes of waters,
and mineralogical data. The code PHREEQC V2 (Parkhust and
Appelo, 1999) was used to model element speciation and mineral
precipitation or dissolution in the water column and at the

watersediment interface. The new data set and geochemical modeling provided insights into the lake chemistry, in particular the
formation of secondary minerals, the interactions in the AMD-carbonate system, and the migration of contaminant elements.
2. Site description
Lake Yanamate is located 5 km south of the Cerro de Pasco deposit at the northern end of the Puna plateau between the Western
and Eastern Cordillera of Peru (Fig. 1A and B). This lake is part of
the mining facilities of the Cerro de Pasco mine, at present operated by Volcan Corporation S.A.A. and previously (until 1999) by
the former state company Centromin S.A. Currently the deposit is
exploited by means of both underground and open pit mining. Cerro de Pasco is one of the largest polymetallic resources of the
world. Accumulated past production and known reserves include
175 Mt @ 7% Zn, 2% Pb and 3 oz/t Ag. Further, 1.2 billon oz of
Ag and 2 million oz of Au, and an equivalent of 50 Mt of Cu have
been produced, largely before 1950 (Baumgartner et al., 2008). The
deposit is dominated by a 1800  300 m quartz-pyrite body
formed by replacement of Triassic limestone, representing
90 vol.% of the entire ore body (Enaudi, 1977). Different types of
pyrite have been described, some of them containing up to
10 wt.% As (Petersen, 1965; Ward, 1961). The main ore minerals
are sphaleritegalena, pyrrhotite, enargite, luzonite, tennantite,
marmantite, arsenopyrite and chalcopyrite (Enaudi, 1977). Final
hydrothermal and hypogene activities resulted in mineral associations dominated by pyritehematiterealgar and covellitebornite,
respectively. Supergene processes formed an enrichment blanket
of chalcocite and covellite (Petersen, 1965). The sources of mine
water were: (1) groundwater of the Cerro de Pasco fault zone inltrating into the open pit and the underground mine and (2) rainwater accumulating in the open pit, that subsequently inltrated
into underlying mine works (Luna Bernal and Delgado Venero,
1985). Old mine works on Cu minerals were used for in situ leaching with highly acidic solutions. The Cu was extracted in a SX-EW
plant using kerosene as a solvent with addition of 4 vol.% Acorga

Fig. 1. (A) Location map of the Cerro de Pasco region in Peru. (B) Map of the local geology, the open pit, discharge pipe, and Lake Yanamate. (C) Map of Lake Yanamate and
Lake Huaygacocha with sampling points. (D) Cross-section of the lake basin from west to east. The two lakes were connected during the sampling period due to the water
level being 14 m higher than before discharge began in 1981.

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

P-5100 or M-5615, both being Cu-selective extractants (USEPA,


1994). The industrial spoil solutions were pumped into the lake
from 1981 until August 2001.
Discharge of AMD and spoil solutions resulted in an estimated lake volume increase from an original 3 to 25  106 m3
and the inundation of a swell, which formerly separated Lake
Huaygacocha from Lake Yanamate (Fig. 1C and D). The bedrock
of Lake Huaygacocha consists of limestone conglomerate with
a calcite matrix. The bedrock of Lake Yanamate consists of thick
bedded limestone with layers of dolomite, sandy limestone and a
few shale interbeds (Jenks, 1951). The two formations are separated by the Cerro de Pasco-fault. The rocks are fractured due to
faulting, and are affected by the formation of karst. Glacial activity eroded loose sediments, allowing an almost direct contact of
the lake water with the carbonates. An end moraine forms the
natural dam SW of the lake and was heightened by the construction of an articial dam, to prevent drainage to the marshland
around Lake Cuchis. The lake covered an area of 3 km2 within
a hydrologic basin of 10 km2. The south-eastern part of the lake
receives surface runoff from a marsh.
The climate of the Cerro de Pasco area is characterized by cold
and dry austral winters (June to August) and warmer austral summers with heavy rain events (October to March). The annual mean
precipitation of 1025 mm/a is nearly equaled by the annual mean
evaporation of 988 mm/a. The annual mean temperature is
4.2 C. The main wind direction is NE to SW.

3. Methods
3.1. Sampling and eld methods
During sampling in June/July 2003 (dry and cold season) the
AMD discharge rate was 9.5 m3/min. The lake water table was at
4359 m above sea level (m a.s.l.) (14 m higher than in 1981), and
had a maximum depth of 41 m. The discharged AMD was sampled
from a retention tank at the pumping station. From the lake, a total
number of 45 water samples were taken at 11 sampling stations in
proles of 0.5 m, 5 m, and continued in 5 m steps down to the lake
bottom (Fig. 1C). Samples were taken with a peristaltic pump and
ltered through 0.45 lm in-line lters, into lled 60 mL HDPE bottles, sub-samples for cation analysis were acidied with HNO3
(Merck suprapure), and frozen immediately until analysis. Temperature, pH, Eh (calculated to a standard H electrode), and electrical
conductivity (EC) were measured directly in the eld with two
closed ow-through cells and a set of WTW electrodes and meters.
Hot acidity (H+ + hydrolysable metals) was measured within the
next 12 h by titration to pH 8.3 with 1.6 N NaOH after addition of
H2O2 and boiling, and reported as mg/L CaCO3 (Ficklin and Mosier,
1999).
Lake sediments were sampled by gravity coring at the same stations where water samples were taken. The 12 cores were divided
into 44 sub-samples. Nine cores of the soils of the inundated zone
(due to the increased lake water level) were taken by percussion
drilling, and were divided into 41 sub-samples. The cores were described and divided on the basis of color and grain size. Pore water
pH was measured with a Sentix SP electrode and a WTW meter.
Samples were dried at 30 C and their weight was determined before and after drying to calculate water content. Two samples of a
soil prole on Pucar limestone and two samples of the Pucar
limestone were taken as reference for background concentrations
of contaminant elements.
To study the direct interaction of the lake water with the underlying bedrock, fresh hand specimens of Pucar limestone taken in
the surroundings of the lake were placed in the lake at 0.5 m water
depth for 25 days and subsequently studied.

1109

3.2. Analytical methods


3.2.1. Isotope analysis of d2H and d18O in water
The stable H and O isotope analyses of the water samples
were performed at the Stable Isotope Laboratory of the University of Lausanne. The stable H isotope analyses were performed
using a ThermoFisher H-Device connected to a Delta S isotope ratio mass spectrometer (IRMS). For this method, the H2 gas was
produced by reduction of a volume of 1.2 lL water over hot
(840 C) Cr within a reactor connected to the dual inlet system
of the IRMS. Oxygen isotope analyses were made through equilibration of 0.5% CO2 in He with 1.2 mL of water for 24 h at room
temperature followed by extraction in a continuous He ow
using a ThermoFisher GasBench II connected to a Delta Plus XL
IRMS. The stable H and O isotope ratios are reported in the d
notation as the deviation relative to the Vienna Standard
Mean Ocean Water (VSMOW). The reproducibility, assessed in
terms of the within-run replicate analyses of laboratory standards, was better than 0.3 and 0.1 (1r) for d2H and d18O values, respectively.
3.2.2. Hydrogeochemical analyses
All the water analyses were performed at the Helmholtz Centre
for Environmental Research (UFZ) Magdeburg, Germany. Water
samples were analyzed by inductively coupled plasma-optical
emission spectrometry (ICP-OES; Perkin Elmer Optima 3000) and
with inductively coupled plasma-mass spectroscopy (ICP-MS; Agilent 7500 C) for major and trace elements. ICP-OES measurements
of S were used to calculate SO4 concentrations in water. Dissolved
P was determined photometrically with ammonium-molybdate
(German standard method DIN EN 1189). Reproducibility and
accuracy of all methods was better than 5%.
3.2.3. Chemical analyses of solid samples
The major and trace element analyses of sediment and soil samples were performed at the Helmholtz Centre for Environmental
Research (UFZ), Dept. of Analytical Chemistry in Leipzig, Germany.
Aliquots of the sample material were dried at 105 C, ground with
an agate ball mill, mixed with stearine wax, pressed to pellets and
analyzed for trace elements by means of the energy dispersive
X-ray uorescence spectrometer XLAB2000 (Spectro Instruments)
running the software package X-Lab Pro 2.2. The calibration functions for the individual analytes were based on the combination of
the measured response and the certied concentration data of
about 50 reference materials including soils, river- and lake sediments and samples rich in Fe (Morgenstern et al., 2001). Samples
strongly enriched with metals were diluted with SiO2 powder in
order to t their analyte concentrations to the working ranges of
the available calibrations. To achieve optimum spectrometer operating conditions, the samples were excited by both polarized -and
monochromatic X-radiation. An aliquot of the powdered sample
was diluted with Li2B4O7 to prepare glass discs for analyzing major
elements. The measurements were performed with the wavelength dispersive X-ray uorescence spectrometer (WDXRF) S4
Pioneer (Bruker-axs). The calibrations of the WDXRF-spectrometer
for the major elements were adjusted using the certied reference
materials, CANMET-LKSD1-LKSD4 (lake sediments), CANMETSTSD1-STSD4 and GBW07309-11 (stream sediments), NISTSRM2689 and NIST-SRM2691(coal y ashes). The relative precision
of the measurements was better than 3% for the traces and better
than 1% for the major components.
Total C (TC) and total organic C (TOC) were determined with a
Coulomat Strhlein 702 CS in the Centre of Mineral Analysis at
the University of Lausanne, Switzerland. To determine TOC, samples were moistened with ethanol and fumed-off overnight at
120 C after addition of HCl. Total inorganic C (TIC) was calculated

1110

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

by difference. Calibration was checked every eight samples


with pure CaCO3, and reproducibility and accuracy was better
than 3%.
Results of XRF and C determination were corrected for water
content to obtain element concentrations for non-dried samples,
in order to allow the direct comparison of element concentrations
in sediments and water.
3.2.4. Mineralogical analysis
Solid samples were analyzed for their mineralogical composition by means of X-ray diffraction (Seifert XRD) and optical microscopy at the Institute of Mineralogy, Crystallography and Material
Science of the University of Leipzig. Scan settings were 0.05 2h
(range 5100) and 2 s. counting time per step with monochromated Cu Ka X-radiation with secondary graphite monochromator.
Polished thin sections were prepared using epoxy resin to embed
loose grain samples, and analyzed by optical microscopy.
3.3. Geochemical modeling
The geochemical modeling code PHREEQC V2 (Parkhust and
Appelo, 1999) and the thermodynamic database minteq from
MINTEQA2 (Allison et al., 1991) were used to calculate speciation
of elements in water and mineral saturation indexes. Ionic
strengths were between 0.2 and 0.7, which is in the upper range
where the ion-association model and Debye Hckel expressions
for activity correction are still applicable (Parkhust and Appelo,
1999).
4. Results and discussion
4.1. Isotope geochemistry of lake water
The water of the discharged AMD had a d18O value of 14.7
and a d2H value of 105.7 and plots close to the Local Meteoric
Water Line (LMWL) of La Paz (Fig. 2A). Spring waters feeding Lake
Junin, located 30 km south of Cerro de Pasco at 4100 m a.s.l., had
comparable values with d18O from 15.8 to 13.5 and d2H
from 116.0 to 102.1 (Flusche et al., 2005). This indicates a
local origin of the discharged mine water.
The d18O values of the lake waters ranged from 11.7 to
8.7 and the d2H values ranged from 93.9 to 79.7. All
lake water samples were isotopically heavier than the discharged
AMD, and followed an evaporation line (d2H = 4.5 d18O  40). Evaporation seems to be the main process controlling the variation of
the d2H and d18O values in the lake. The slope of 4.5 indicates a

humidity of 60% (Gonatini, 1986), which is lower than the average value of 74% reported in the literature for Cerro de Pasco
(PAMA, 1996). The AMD plots between the evaporation line and
the LMWL of La Paz. The AMD waters originated mainly from
fault-waters of the Cerro de Pasco fault zone and precipitation
inltrating the underground mine.
From 15 m water depth to the surface, the mean d18O value increased from 10.6 to 9.6 and the mean d2H value from
88.7 to 82.5 (Fig. 2B). The increases in d2H and d18O values
from 15 m to 10 m and from 5 m to 0.5 m were nearly equal (0.17
d18O/m and 0.21 d18O/m, respectively). The uppermost waters
(0.5 and 5 m depth) had a mean d18O value shifted 1.6 to a more
negative value. This isotope shift can be explained by mixture of
the lake upper 5 m with recent AMD discharge and precipitation.
The deep waters of the lake (2035 m) had a relatively uniform isotope composition with d18O values 11.1 and d2H values
93.6 (Fig. 2B), indicating a stagnant bottom stratum (termed
monimolimnion), which was excluded from taking part in lake
water mixing and circulation processes.
The isotopically heaviest water (d18O 8.7 and d2H 79.7)
was at CPY14(10 m). This isotopically heavier water suggested a local contribution from the marshes SE of the lake (Fig. 1B).
4.2. Hydrogeochemistry
The AMD discharged into the lake had pH 1.1, Eh 609 mV, and
an EC of 27.4 mS/cm (Table 1). The concentrations of Fe and SO4
were 4440 mg/L and 31,875 mg/L respectively, and acidity was
18,720 mg/L as CaCO3. Copper had a concentration of 125 mg/L,
Zn 748 mg/L, Pb 4.2 mg/L and As 47.8 mg/L. These values were
close to the range reported for AMD from massive pyrite deposits
(Plumlee, 1999) and fall into the class of extremely acidic and extremely mineralized pit lakes, following the classication of Snchez Espaa et al. (2008).
The average hydrogeochemical parameters determined
from the 45 lake water samples were pH 1.0, Eh 653 mV, EC
24.1 mS/cm, 4330 mg/L Fe, 29,250 mg/L SO4, 16,775 mg/L acidity
as CaCO3, 87 mg/L Cu, 493 mg/L Zn, 3 mg/L Pb and 48 mg/L As.
These mean values did not differ greatly from values measured
for the discharged AMD.
The pH and Eh decreased slightly with depth from pH 1.2 and
672 mV at 0.5 m to pH 1.1 and Eh values between 585 and
652 mV at the lake bottom (Fig. 3A). Acidity increased from
15,712 mg/L to 19,175 mg/L as CaCO3 with depth, indicating that
dissolution of carbonate bedrock did not result in higher pH and
lower acidity at the lake bottom. Compared to the value of
27.4 mS/cm for discharged AMD, the upper 15 m of the lake water

Fig. 2. (A) d2H v/s d18O binary plot of lake water samples, discharged AMD, and the Local Meteoric Water Line of La Paz (LMWL La Paz). (B) d18O and d2H in VSMOW of
selected water columns of the lake.

Table 1
Stable isotope data of d2H and d18O, physicochemical parameters, and concentrations of selected elements in the sampled water columns of Lakes Yanamate and Huaygacocha and the Cerro de Pasco AMD. The calculated mean
concentration and standard deviation includes all water samples of both lakes. Abbreviations: EC = electrical conductivity; acy = acidity as CaCO3; n.a. = not analyzed; Stdev = standard deviation.
d2H
()

pH Eh
EC
acy
Na
K
Mg
Ca
Mn
Al
Fe
SO4
Cu
Zn
Pb
As
Si
P
Ti
V
Cr
Co
Ni
Zr
Cd
Bi
U
() (mV) (mS/cm) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (lg/L) (lg/L) (lg/L) (lg/L) (lg/L) (lg/L) (lg/L) (lg/L) (lg/L)

CPY04

0.5
5
10
14.5

9.4
10.2
9.6
10.1

82.0
83.3
82.3
85.8

1.0
1.0
1.0
1.0

CPY05

0.5
2.5

n.a.
n.a.

n.a.
n.a.

CPY07

0.5
6

n.a.
n.a.

CPY09

0.5
5
10
15
20
25

CPY10

n.a.
n.a.
15,840
n.a.

16.2
16.0
17.1
17.4

30.9
29.8
31.1
31.2

221
208
216
214

383
390
419
448

178
170
176
172

248
244
265
279

3920
3950
4430
4990

29,865
29,340
31,380
33,300

477
467
493
487

2.7
2.8
2.5
2.2

48.3
47.2
50.7
56.2

14.0
13.0
14.5
15.3

33
36
36
39

515
480
431
346

1936
1924
2082
2188

181
182
199
214

156
166
189
221

299
297
325
348

66
54
44
38

1235 1708 110


1218 1498 107
1347 1030 120
1426 540 117

1.3 674 20.0


1.5 673 19.9

n.a.
n.a.

16.5
16.5

31.1
30.9

219
218

378
377

179
179

248
247

3950
3940

28,935 90.9 482


28,485 90.3 479

3.0
2.8

52.1
47.9

13.9
13.6

35
35

563
523

2149 204
1926 178

171
159

318
289

49
96

1319 1916 121


1239 1742 112

n.a.
n.a.

1.0 673 18.9


0.9 654 22.2

n.a.
n.a.

16.6
16.2

30.9
29.4

219
201

378
381

179
164

248
244

3950
4050

28,470 90.7 482


2.7
28,005 78.8
n.a. 2.6

47.6
50.1

13.9
12.1

34
36

517
439

1918 180
2068 194

161
182

281
301

58
39

1232 1763 115


1336 1097 120

n.a.
n.a.
n.a.
n.a.
n.a.
n.a.

n.a.
n.a.
n.a.
n.a.
n.a.
n.a.

1.5
1.1
1.2
1.1
1.1
1.1

673
660
657
645
637
637

18.1
21.1
23.2
27.4
29.4
29.5

15,962
16,191
17,228
19,302
19,860
19,701

17.0
17.1
18.4
19.5
19.5
19.6

31.7
31.8
34.5
35.9
35.4
35.4

224
224
233
239
227
228

384
383
397
421
413
399

184
185
196
202
191
192

251
252
278
304
300
303

4020
4050
4540
5330
5480
5520

29,025
29,325
32,340
36,150
37,050
37,050

92.8
96.8
100.0
96.4
86.8
86.8

493
500
534
557
542
549

2.9
3.5
3.0
2.8
2.8
2.8

48.1
48.8
54.3
59.3
61.1
62.4

16.4
14.8
17.0
18.4
19.4
18.3

36
36
38
39
40
39

523
515
541
502
420
454

1952
1981
2121
2296
2333
2390

183
187
192
218
228
224

160
164
180
208
224
237

290
302
316
341
354
355

48
44
41
42
35
35

1276
1266
1362
1474
1509
1516

1803
1925
1930
1475
1110
1123

116
117
128
133
136
140

0.5
5
10
15
18

n.a.
n.a.
n.a.
n.a.
n.a.

n.a.
n.a.
n.a.
n.a.
n.a.

1.0
0.8
0.8
0.7
0.6

673
659
657
651
637

21.0
23.8
25.9
27.8
31.3

14,608
17,600
16,320
16,400
21,440

14.5
18.0
18.4
18.6
19.2

28.0
32.8
34.6
35.0
35.0

201
235
231
235
223

367
389
396
393
405

162
195
194
197
188

221
266
276
282
293

3550
4240
4520
4720
5330

25,515
29,985
32,025
33,075
35,700

81.9
99.2
99.6
102.0
85.7

439
517
527
532
537

2.7
3.3
3.0
3.0
2.8

36.8
51.1
53.9
57.4
59.2

13.3
32.6
15.0
16.2
19.1

32
37
39
40
42

484
554
562
584
429

1772
2050
2114
2251
2276

165
192
195
213
218

144
168
178
189
217

266
305
308
331
343

82
62
48
47
35

1150
1336
1376
1439
1458

1745
1908
1912
1992
1119

105
122
128
131
131

CPY13

0.5
5
10
15

9.3
10.0
10.2
10.3

82.3
85.0
85.8
86.7

1.6
1.2
1.2
1.2

673
660
657
652

18.5
21.1
24.6
25.1

14,800
15,120
16,480
17,040

17.2
15.4
18.9
18.7

31.9
28.2
35.2
35.4

225
203
238
236

385
344
406
397

184
167
200
200

252
226
284
285

4020
3630
4640
4760

28,440
25,635
32,190
33,135

92.8
86.2
102.0
103.0

499
453
547
546

2.8
2.6
3.0
5.3

47.0
33.1
54.1
56.7

14.1
12.3
15.4
16.2

37
30
40
41

521
472
562
539

1922
1747
2147
2215

181
159
201
207

157
140
183
185

289
270
311
321

36
29
40
41

1239
1140
1386
1398

1794
1715
1930
1978

116
104
126
129

CPY14

0.5
5
10
15

9.8
11.3
8.6
9.8

83.6
86.6
79.7
86.7

1.1
1.5
1.4
1.4

673
659
657
654

19.5
22.3
24.2
25.4

15,696
16,928
17,248
18,272

17.1
14.9
9.7
18.6

31.4
26.9
16.5
35.1

226
192
114
239

382
383
362
394

182
159
95
198

250
216
130
281

3990
3520
2270
4680

28,110
25,005
16,365
32,880

92.1
83.0
49.9
103.0

494
432
268
540

2.8
2.7
2.8
2.9

48.1
31.6
19.8
55.8

16.4
13.9
7.1
17.4

36
30
18
33

510
443
260
565

1940 176
1649 152
1022 94
2176 207

158
134
91
186

284
248
160
326

34
23
7
39

1266
1079
676
1383

1801 115
1740 95
1439 60
1960 126

CPY16

0.5
5
10
17

10.2
10.9
10.7
11.7

83.2
86.2
85.0
91.7

1.4
1.3
1.2
n.a.

672
660
656
636

19.1
22.9
24.9
31.3

15,920
16,160
17,792
19,296

17.4
15.2
18.6
19.1

32.0
27.3
35.1
35.1

226
195
234
222

388
391
401
405

186
161
197
187

254
221
280
292

4050
3550
4600
5330

28,620
14,085
14,205
35,550

93.6
83.7
101.0
85.6

502
414
536
532

2.7
2.7
3.0
2.7

47.7
32.0
54.4
61.1

15.1
15.4
14.7
18.8

34
30
38
40

510
440
539
446

1960
1668
2121
2327

183
158
195
227

159
132
180
225

292
249
326
357

30
71
57
39

1255
1057
1371
1462

1801 112
1730 96
1871 125
1132 134

CPY19

0.5
5
10
16

9.8
10.5
9.8
11.2

83.3
87.2
84.5
91.5

0.7
0.6
0.6
0.5

673
660
657
641

18.9
22.1
24.6
28.8

14,480
14,880
16,512
18,320

17.1
17.7
18.5
18.9

31.7
32.6
34.7
35.0

223
232
232
228

384
387
398
412

183
192
195
189

251
263
277
294

4000
4190
4530
5230

28,350
29,490
31,650
35,100

92.6
99.3
100.0
87.6

497
511
532
540

2.8
2.9
2.8
2.8

47.1
50.6
54.4
60.2

13.9
14.1
15.2
17.7

34
35
38
39

507
536
554
482

1904
1997
2159
2314

177
189
199
215

155
163
181
219

282
298
318
343

32
35
41
34

1219
1297
1369
1477

1737
1884
1906
1225

110
118
125
133

CPY20

0.5
5
10
15
20
25
30
35

8.9
10.3
9.6
10.6
11.2
11.2
10.9
11.1

80.9
86.1
85.5
89.7
93.9
93.9
92.8
93.6

1.1
1.1
1.0
1.0
0.9
1.0
1.0
1.0

672
662
657
650
637
631
589
588

19.3
21.0
23.9
26.1
30.1
30.6
30.4
29.5

15,360
15,840
16,800
17,920
18,800
20,240
20,992
20,160

7.4
11.9
19.3
17.7
16.3
16.6
40.1
40.8

11.1
20.2
36.4
33.2
29.1
27.4
35.4
36.4

80.6
148
243
227
204
185
208
214

392
376
419
401
413
403
459
466

64
121
205
191
171
156
188
193

84
163
291
267
242
229
371
384

1440
2670
4760
4710
5030
4760
7010
7180

11,205
19,050
33,255
32,355
32,340
30,375
41,250
42,300

33.5
62.7
105.0
99.6
82.6
74.0
76.1
78.1

180
332
564
529
503
473
728
746

2.8
2.8
2.8
2.9
2.6
2.6
2.8
2.6

12.7
24.4
54.2
53.8
49.0
46.6
63.1
58.1

5.9
12.3
20.5
17.3
17.3
19.4
22.9
21.7

12
22
37
37
32
31
44
33

170
339
524
495
340
329
412
415

686
1286
2104
2074
1871
1833
3251
3018

63
120
200
197
178
176
367
338

59
107
178
195
214
199
342
323

105
195
310
322
320
302
555
524

n.a.
11
33
31
19
14
20
15

450
824
1346
1331
1256
1219
2357
2240

1434
1637
1834
1747
986
956
800
756

38
73
123
118
106
99
133
126

CPY27

0.5
3.5

n.a.
n.a.

n.a.
n.a.

1.8 673 18.9


1.4 673 18.6

16,000 11.9
15,840 13.2

20.5
23.5

148
167

338
327

120
135

164
183

2660
3010

18,945 60.3 328


21,345 69.3 375

2.5
2.4

23.8
27.4

14.1
11.2

22
26

330
381

1265 122
1444 140

102
121

188
207

8
15

798 1392
931 1444

73
83

10.3 86.4 1.1 654 23.8

17,193 17.8

30.9

211

394

176

255

4327

29,250 86.8 493

2.8

48.0

15.8

34

467

1996 190

177

304

39

0.8

1865

5.6

5.4

33

26

29

53

1042

95

0.4

12.0

4.1

89

49

50

73

18

18,720 19.1

40.6

375

391

287

284

4440

31,875 125.0 784

4.2

47.8

24.2

2304 209

156

348

44

4.1

0.3

20

19.5
21.7
23.4
25.9

4.1

14.7 105.7 1.1 609 27.4

91.1
86.1
82.9
68.7

6749 15

0.1

660

427

1296 1555 113


310

388

21

1615 3428 172

1111

MEAN 
lake
Stdev 
lake
AMD


673
661
650
621

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

Column Depth d18O


(m)
()

1112

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

Fig. 3. (A) pH, Eh, electrical conductivity (EC), acidity (Acy). (B,C,D) Element concentration distribution of Na, K, Mg, Ca, Mn, Al, Fe, SO4, Cu, Zn, Pb and As in 11 water columns
sampled from Lake Yanamate and Lake Huaygacocha. (E) Saturation indices (SIs) for gypsum, anglesite, jarosite and strengite, calculated with PHREEQCI V2 and the minteq
database. Vertical dashed lines mark values for inowing AMD.

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

column had mean values of 19 mS/cm, while the bottom waters


had equal or higher values. Below 20 m water depth all the EC values were around 30 mS/cm.
From the top to the bottom of the water body, mean concentrations of Fe increased by 52% from 3581 mg/L to 5433 mg/L and for
SO4 by 40% from 25,609 mg/L to 35,959 mg/L (Fig. 3C). Aluminum
followed a trend similar to that of Fe. Calcium concentrations in the
lake (range 327466 mg/L; mean 397 mg/L) were equal to discharged AMD (391 mg/L). In contrast, Mg was depleted in the lake
(range 81243 mg/L, mean 214 mg/L) compared to discharged
AMD (375 mg/L). Gypsum (CaSO42H2O) had relatively invariable
saturation indices (SI) around 0.25 (Fig. 3E) and controlled the
concentration of Ca in the lake and the AMD. Jarosite
(KFe3
3 OH6 SO4 2 ) was oversaturated (SI = 0.8 to +4.5, mean
1.7) in the upper samples and undersaturated in the deepest samples of the water body (SI = 3.8 to +2.3, mean 0.7) (Fig. 3E).
Potassium was depleted in all lake water samples (range 11
36 mg/L, mean 31 mg/L) compared to 40.6 mg/L K for discharged
AMD, while Na concentrations of the lake waters were close to
the 19 mg/L measured for discharged AMD for most samples, and
only depleted for some samples near the lake surface (Fig. 3B). This
can be explained by partial removal of K from the water column by
jarosite precipitation. Jarosite was found by XRD only in sediments
of water depth 67 m; this supports the hypothesis of element cycling by jarosite (see Section 4.5.). Its formation in the upper part of
the water column and its downward settling and later dissolution
may have resulted in downward element cycling of Fe and SO4.
Redox cycling of Fe in natural lakes is a well-studied subject
(Hamilton-Taylor and Davision, 1995; Hamilton-Taylor et al.,
2005). The formation of schwertmannite (2.75 mol H+ per mol
Fe3+) or jarosite (2 mol H+ per mol Fe3+) liberates acidity, thus these
minerals can control the lake pH, as observed in the Berkeley pit
lake and several pit lakes of the Iberian Pyrite Belt for schwertmannite (Pellicori et al., 2005; Snchez Espaa et al., 2008), e.g.:
3

3Fe3 K 2SO2
4 6H2 O () KFe3 OH6 SO4 2 6H

In Lake Yanamate, however, the assumed dissolution of jarosite


in the lower part of the water had no apparent effect on the water
composition. The acidity consumption was probably obscured by a
higher acidity and mineralization in the lower part of the lake
water, due to its origin from times before the discharge of spoil
solutions from the SX-EW plant was stopped.
The discharged AMD contained almost no dissolved P (0.1 mg/L).
Concentrations of 35 mg/L dissolved P in the lake water originated
from the dung of llamas and alpacas feeding frequently close to the
lake, and entered the lake with the surface runoff (Table 1). Strengite (FePO42H2O) was found to be close to saturation (SI = 0.91.1
(mean 0.1)) in the upper part of the water column and undersaturated in the lower part of the water column (SI = 1.60.4
(mean 0.5)) and may have played a smaller role in element
cycling.
Manganese seemed to have precipitated partially from the
water column as its concentration is lower in the lake (range 64
205 mg/L; mean 176 mg/L) than the 287 mg/L measured in AMD.
Copper showed increasing concentrations from 85.2 mg/L at the
top to 93.0 mg/L at 10 m water depth and decreased below to a
mean value of 87 mg/L for the bottom layer, possibly due to the
presence of the Cu-selective extractant in the mixed discharge
and preferred settlement (Fig. 3D). Zinc and Cd concentrations increased slightly with depth (from 455 to 560 mg/L Zn and from 1.2
to 1.5 mg/L Cd). Lead had a uniform concentration of 2.8 mg/L
(Fig. 3D). Anglesite (PbSO4) had a SI of 0.08 and controlled the concentration of Pb in the lake water, causing its spatially invariable
concentration (Fig. 3E). The element of greatest environmental
concern in the lake was As with a mean of 48 mg/L, showing
increasing mean concentrations with depth from 41 to 61 mg/L.

1113

The Eh decreased stronger in Lake Huaygacocha from 673 to


621 mV than in Lake Yanamate. Vertical mixing of lakes is partly
controlled by their relative depths (maximum depth divided by
width) (Castro and Moore, 2000), which is higher for Lake Huaygacocha and can explain the differences between the lakes hydrogeochemistry. Concentrations of metals were slightly lower than in
Lake Yanamate (CPY04 and CPY07, Fig. 3D).
In the austral winter of 2003 Lake Yanamate was a meromictic
lake without complete mixing and its density gradient was dominated by the high mineralization of the water. The monimolimnion
found for CPY09 and CPY20 below 20 m probably represented the
mixed discharge of AMD and the spoil water of the SX-EW plant.
Calculated with the discharge rate of July 2003, the lake water
body has not been completely replaced in the time after the SXEW plant was closed in 2001. Possibly in the austral summer, the
higher precipitation may result in an epilimnion and less vertical
mixing. The low element concentrations for CPY14(10 m) are in
agreement with its heavier isotope signature, interpreted as a local
contribution of near-surface groundwater discharge from the
marshes SE of the lake, resulting in dilution (Fig. 2B and Table 1).
4.3. Characterization of lake sediments
At the end of the discharge pipe, the discharged AMD formed a
2 m deep channel in the limestone, showing the general reactivity
of the bedrock with the acidic water. Parts of this area were covered by tailings. Close to the discharge point (CPY11 and CPY13;
Fig. 1C) pyrite, sphalerite, anglesite, quartz and muscovite settled
out of the mine water in the form of detritus-rich sediment, taken
up from the tailings and transported into the lake (Fig. 4A). In these
sediments, the concentrations of Fe, Cu, Zn, As, Pb and Cd were the
highest found in the lake sediments (Table 2).
In most sediment cores a sequence of gel-like matter above
ne-grained material was found. The layer of gel-like matter was
of varying thickness, between 3 and 32 cm. The color of the gels
varied between gray and brown in Lake Yanamate, and in Lake
Huaygacocha a layer of greenish-gray color was found (CPY23,
CPY24). The mean water content was 89 wt.% in the liquid and
79 wt.% in the rigid gels. The gels contained 1.64.4 (mean 3.0)
wt.% TOC and partly imbedded residues of plants, algae and seed
vessels. Some gels contained calcite and had TIC contents of up
to 1 wt.% (Fig. 4B). The pH of the gels was 11.5 where no calcite
residues were found and 5.66.4 where calcite was detected
(CPY14, Fig. 4B). Gypsum detected by XRD in the dried gels crystallized largely during the drying process of the samples and was
present in aggregates of up to 8 mm in size, which were not found
during sampling.
With the spoil water of the SX-EW plant, kerosene entered the
lake until August 2001 (Cerro de Pasco staff, pers. comm.). Oil
droplets formed through wave action or during the industrial processing can occulate as oil-mineral aggregates (OMA), and settle
down due to an increased bulk density (Lee, 2002; Muschenheim
and Lee, 2002). Visibility in the lake water was around 15 cm
and lter residues were high, thus suspended matter was abundant. The high mineralization of lake water favored the stabilization of the OMA through coagulation with cations (Bragg and
Yang, 1995). The density of the OMA may have allowed its winddriven resuspension from the ooded soils and its redeposition in
the deeper zones of the lake. Aging of low-crystalline gypsum may
explain the presence of the rigid gels below the liquid gels. The gel
layers found accumulated during the 20 a of kerosene discharge.
Microbial decomposition of organic matter is very slow or absent
in low-pH environments (Pelmont, 2005).
Materials of clayish to sandy grain sizes with or without gel cover
on top were gray to brown depending on the organic matter content,
ocher and red in Fe-rich horizons, and gray in the carbonate-rich

1114

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

Fig. 4. Sedimentology, mineralogy, pH, TIC, TOC, water content, and bulk chemical data (corrected to a wet base) of selected sediment cores of Lake Yanamate. For better
visibility, graphs of trace elements have varying scales.

deeper parts of the proles. The mean water content was lower and
varied much more than in the gels, from 14 to 84 (mean 51) wt.%.
TOC was between 1.4 and 3.3 (mean 1.7) wt.% and TIC was only
found in samples in which dolomite or calcite was detected. These
layers appear to originate from a former minor loose sediment layer
on top of the bedrock, because reference samples of carbonate bedrock did not contain any TOC. The pH was between 1.4 and 4.5 where
no carbonate was detected. Gypsum was absent in the detrital-enriched layer at CPY13, and was formed by the inltrating acidic sulfate solutions, with dissolved calcite as a source of Ca. Where
carbonates were present, pH increased and stabilized at around 5
5.6, consistent with the modeled equilibrium pH of 5.3 for average
lake water with calcite (Figs. 4 and 5).
Down to the maximum sampled depth of 69 cm, no unaltered
limestone bedrock was reached. The experiment with limestone
specimens put in the lake for about one month showed that, as a

rst step, the acidic water of the lake inltrated the fractures of
the limestone bedrock. In thin sections, hydroxides or hydroxysulfates were visible in the fractures. Additional to Fe- and possibly
Al-hydroxides, the formation of gypsum (replacing calcite) may
clog fractures, because gypsum occupies twice the volume of calcite and could decrease the inltration rate.
4.4. Characterization of soil sediments
All soils except CPY18 had a root layer on top and only CPY07
and CPY24 (both located in Lake Huaygacocha) were covered by
a 10 or 40 cm thick layer of liquid gel, respectively. Soils were
mainly of argillaceous grain size. Root-containing layers exhibited
a middle brown color; deeper horizons were light brown, or in
some samples ochre or oxide red where high concentrations of
Fe were detected.

Table 2
Element concentrations of some selected sediment and soil proles. Element concentrations were corrected to a wet base. Abbreviations: TIC = total inorganic C; TOC: total organic C; ref soil = reference soil; n.a. = not analyzed;
bdl = below detection limit.
Depth
(cm)

TIC
TOC
H2O
Al2O3 SiO2
CaO
Fe2O3 S
K2O
P2O5
V
Cr
Mn
Ni
Cu
Zn
As
Sr
Zr
Pb
U
Bi
Cd
(wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg) (mg/kg)
0.04
0.49
0.08
0.31
5.99

8.12
2.36
3.61
8.85
1.65

68
40
46
58
43

0.47
0.61
0.41
1.62
0.10

3.28
2.50
2.33
4.85
0.70

0.10
0.11
0.07
0.10
bdl

2.62
15.65
8.07
4.00
27.54

2.11
2.84
9.00
3.34
1.28

1.65
6.45
4.67
2.07
0.18

1509
355
1346
1488
266

5
10
25
21
bdl

8
bdl
8
8
bdl

120
111
103
171
539

4
bdl
4
3
bdl

464
463
346
1391
15

387
368
332
846
2258

1436
989
2157
265
13

58
210
122
77
382

13
4
6
16
3

100
87
51
202
6

3.5
bdl
bdl
4.2
bdl

46.7
21.0
8.1
bdl
bdl

2.0
7.6
2.9
21.8
3.0

CPY07 05
510
1015
1518
1825
3035
4550

n.a.
n.a.
0.04
0.34
0.70
3.37
2.74

n.a.
n.a.
1.88
2.14
3.18
2.37
1.50

92
93
84
51
73
62
75

0.03
0.03
0.05
0.09
0.33
0.02
0.02

0.20
0.14
0.59
0.70
1.25
0.48
0.15

0.03
0.03
0.08
0.14
bdl
bdl
bdl

0.25
0.17
2.73
11.03
7.20
17.37
11.90

0.95
0.87
1.45
4.04
0.59
1.26
0.18

0.50
0.39
1.65
5.69
1.97
0.19
0.08

94
72
218
699
309
263
83

bdl
bdl
2
8
4
bdl
bdl

2
2
2
bdl
5
bdl
bdl

80
57
70
59
79
370
74

1
1
1
2
1
bdl
bdl

47
36
54
127
648
17
9

246
177
188
172
888
890
37

81
79
567
2402
135
13
13

22
20
43
104
88
309
165

3
3
1
1
1
bdl
1

39
45
84
212
128
3
3

bdl
bdl
bdl
bdl
1.2
bdl
bdl

63.0
60.1
41.3
10.3
6.2
bdl
bdl

0.7
0.5
0.8
1.0
18.9
bdl
bdl

CPY10 010
1020
2030
3032
3235
3740
4045

0.05
n.a.
n.a.
n.a.
1.79
n.a.
2.12

1.90
3.64
n.a.
n.a.
1.35
n.a.
3.25

92
91
88
54
16
84
72

0.07
0.07
0.12
0.80
6.26
0.32
0.28

0.36
0.66
0.93
4.72
35.02
0.87
1.99

0.01
0.02
0.03
0.14
1.08
0.03
0.07

0.44
0.15
0.11
9.75
11.92
0.72
9.27

0.71
0.43
0.74
0.77
2.61
1.04
1.26

0.68
0.37
0.50
4.95
3.48
0.69
0.31

50
81
101
990
11,617
650
372

1
1
2
13
56
3
5

3
2
3
10
45
3
3

137
70
135
156
544
154
266

1
1
2
2
16
4
3

132
69
60
391
130
19
13

550
314
510
575
3084
1501
1235

62
40
64
68
263
55
21

7
4
5
112
274
17
125

1
2
3
6
36
3
4

8
3
5
16
22
4
3

0.4
0.7
0.3
bdl
1.8
0.7
0.6

1.9
5.9
51.2
9.7
1.7
0.5
bdl

2.4
0.9
1.4
2.3
3.6
0.8
bdl

CPY11 grabber n.a.

n.a.

n.a.

8.58

47.93

1.43

0.38

18.91

10.68

3800

66

41

937

13

1650

59,220

1630

730

112

11,700

2.2

474.0

231.1

CPY13 010
1220
2030
3040

0.27
0.11
0.02
n.a.

1.64
3.39
2.80
n.a.

63
76
80
81

2.81
0.21
0.14
0.06

18.87
3.16
0.98
0.59

0.52
0.04
0.01
0.01

0.10
4.23
4.00
5.08

4.96
0.57
0.42
0.61

2.85
2.33
2.20
0.37

1032
113
88
240

19
6
4
bdl

25
6
4
bdl

138
77
64
84

7
bdl
1
1

814
948
402
20

10,323
476
983
598

403
67
371
27

303
69
55
82

33
6
1
1

3076
142
58
34

2.0
2.1
0.9
bdl

157.3
87.8
2.0
2.3

38.0
2.3
13.8
0.6

CPY14 07
711
1116
1618
1825
2530

0.09
0.23
0.11
0.06
0.95
0.81

2.52
4.40
3.83
3.54
3.36
3.69

91
89
73
67
79
75

0.09
0.05
0.19
1.65
0.46
1.01

0.35
0.28
1.10
1.67
3.22
6.36

0.01
0.02
0.03
0.04
0.13
0.25

0.22
0.21
5.53
6.43
4.69
4.04

1.03
0.44
0.57
0.76
1.04
1.21

0.75
0.34
2.89
3.52
0.50
0.55

22
55
bdl
3175
561
927

2
4
5
10
5
11

4
3
5
9
5
9

179
67
88
116
131
120

2
1
bdl
3
5
9

151
231
405
638
20
13

720
298
296
427
662
598

80
53
66
1908
81
80

6
5
59
76
80
96

1
1
2
3
6
11

13
8
10
77
4
6

0.3
1.5
1.2
1.1
0.9
1.1

1.8
3.3
1.4
1.3
bdl
bdl

2.3
0.8
1.4
2.3
7.6
0.5

CPY20 07
714
1421

0.06
1.22
n.a.

1.77
2.14
n.a.

88
86
88

0.18
0.07
0.06

1.46
1.20
1.49

0.02
0.01
0.01

1.97
4.40
2.91

0.62
0.24
0.09

0.97
0.14
0.08

530
411
319

3
2
bdl

7
bdl
1

107
54
36

2
1
1

186
19
6

964
54
33

102
10
5

34
75
54

1
1
1

157
35
4

0.3
bdl
bdl

10.0
1.0
0.2

4.3
bdl
0.3

CPY23 05
510
1015
1520
2025
2530

0.00
0.01
n.a.
0.05
0.00
0.12

4.10
2.41
1.41
1.11
1.54
1.54

85
86
75
70
61
59

0.11
0.08
0.06
0.14
2.62
4.84

0.47
0.47
0.67
0.90
1.96
2.41

0.01
0.01
0.11
0.06
bdl
bdl

0.83
2.15
4.36
5.28
8.05
6.98

1.21
0.69
3.23
4.37
1.68
2.03

1.21
1.48
2.52
2.97
4.36
4.03

119
bdl
1166
1322
254
bdl

2
bdl
6
16
23
25

5
2
bdl
7
10
7

283
153
86
104
121
437

4
2
4
2
bdl
bdl

252
127
68
144
1728
2531

887
431
252
314
443
3847

109
64
3328
1830
98
18

13
27
79
89
115
104

4
2
1
bdl
bdl
bdl

14
40
181
14
5
8

0.6
0.3
bdl
bdl
5.6
9.8

56.7
41.6
34.6
2.1
bdl
bdl

2.8
1.8
0.9
1.4
6.6
7.2

CPY25 05
510
1015
1520
2027
2735

2.65
2.72
4.39
6.84
7.45
7.78

0.02
0.86
3.70
1.25
0.07
0.00

61
48
49
44
43
40

0.10
0.12
0.08
0.07
0.07
0.10

0.33
0.35
0.44
0.42
0.46
0.70

0.03
bdl
bdl
bdl
bdl
bdl

9.77
14.61
21.63
27.37
27.81
28.21

0.95
0.74
1.05
0.88
1.55
2.26

4.96
6.66
1.10
0.30
0.23
0.26

272
303
742
723
642
512

bdl
bdl
bdl
bdl
bdl
bdl

5
bdl
bdl
bdl
bdl
bdl

78
88
192
294
516
808

bdl
bdl
bdl
bdl
bdl
bdl

96
155
83
39
18
5

197
362
1324
1436
3422
5177

431
289
38
19
29
3

104
145
349
463
503
518

2
2
1
1
2
4

142
66
111
61
14
6

bdl
bdl
bdl
bdl
bdl
1.8

77.4
65.9
6.4
2.6
3.0
bdl

0.8
2.6
1.5
0.9
6.1
1.8

Ref.
1020
soil
2030

6.18

1.12

n.a.

4.50

69.86

1.13

1.46

1.37

0.11

4167

39

37

817

11

67

245

76

105

74

165

2.2

2.0

bdl

3.63

1.15

n.a.

6.69

70.80

1.86

1.67

2.11

0.12

5502

65

43

1173

14

22

303

81

124

93

65

bdl

bdl

bdl

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

CPY06 06
614
1420
2026
2934

1115

1116

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

Fig. 5. Sedimentology, mineralogy, pH, TIC, TOC, water content, and bulk chemical data (corrected to a wet base) of selected soil cores of Lake Yanamate (CPY25 and CPY06)
and of Lake Huaygacocha (CPY07), and one sediment core of Lake Huaygacocha (CPY23). For better visibility, graphs of trace elements have varying scales.

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

CPY06 had comparatively high TOC with 8.1 and 8.9 wt.% in the
upper root-containing horizon and between 21 and 26 cm, where
the brownish color indicated a former top-soil (Fig. 5B). CPY06 is
located in the Pampa area between the two lakes and may have
had dense vegetation before the discharge period. Other soil samples had TOC contents 64 wt.%. Soil layers that were free of calcite
(XRD and TIC results) had a pH of 12. In deeper layers, where
calcite was detected as a residual (CPY06) or dominating mineral
(CPY25), the pH was >5 (Fig. 5).
4.5. Jarosite in solid samples down to 7 m water depth
Jarosite was found in all the 3 soil samples from the smaller
Lake Huaygacocha: in liquid gel of two cores at 7 m water depth
(CPY24 and CPY07) and once in the clay size matrix for a core at
3.5 m water depth (CPY05). In contrast, two cores taken at 14 m
water depth near the center of Lake Huaygacocha did not show
jarosite (CPY04 and CPY23). The cores CPY07 and CPY23 are
shown in Fig. 5. The occurrence of jarosite as detected by XRD
is in accordance with the results of modeling, which showed that
jarosite is stable in the upper part but not in the deeper part of
the water body. It could not be explained by the modeling, why
jarosite was found only in the sediments of the smaller Lake
Huaygacocha. Similar to a Na-jarosite formation in a waste
H2SO4 lake with a pH 0.85 (Schuiling and van Gaans, 1997),
concentrations of contaminant metals and As contained in Lake
Huaygacochas K-jarosite bearing layers were 1001000 times
lower than in jarosites from Zn industries, because the sorption
of metals is positively correlated with pH (Bigham and Nordstrom, 2000).
4.6. Trace metal and arsenic distribution in gels
Samples CPY10(010 cm) and CPY14(07 cm) (Fig. 4) had comparatively low concentrations of Al2O3 and SiO2 in their liquid gel
matrices (0.07 and 0.09 wt.% Al2O3; 0.36 and 0.35 wt.% SiO2,
respectively). Thus detrital input was minimal for these two samples and geochemical data was interpreted as characteristic for the
gel-like matter. Both had similar pH of 1.2 and 1.5, similar water
content of 91 and 92 wt.%, and similar TOC of 1.9 and 2.2 wt.%,
respectively. The absence of detritus allowed the element concentrations in the liquid gels to be compared with those of the overlying water. CPY10(010 cm) and CPY14(07 cm) had 132 and
151 mg/kg Cu, 550 and 720 mg/kg Zn, and 62 and 80 mg/kg As,
respectively, thus they were not signicantly enriched in these elements compared to the overlying water with 85.7 mg/L Cu,
537 mg/L Zn, and 59 mg/L As for CPY10(18 m) bottom water. Lead
was more highly concentrated in the gel than in the overlying
water, with values of 8 and 13 mg/kg, respectively, compared to
2.8 mg/L in CPY10 and 2.9 mg/L Pb in CPY14 for the bottom water,
conrming its settling from the water column in form of anglesite.
With the exception of Pb, none of the contaminant metals, Fe (2482
and 3602 mg/kg), or Mn (137 and 179 mg/kg) were xed in the gel
but were mobile in the gels pore water (water content 92 and
91 wt.%). If metals were xed in any form, the high hydraulic permeability of the gel-like matter would have resulted in diffusion
and downward migration of metals from the overlying lake water
into the pore water and consequently in concentrations higher
than those in the lake water.
The rigid gel of CPY10(1020 and 2030 cm) had values for
major and trace elements similar to those of the overlying liquid
gel (Fig. 4C) while rigid gel of CPY14 did not (Fig. 4B). Prole
CPY14 contained calcite below 18 cm and had a pH of 5.6. Directly
above the calcite a layer with gypsum concretions formed,
where up to 638 mg/kg Cu and 1908 mg/kg As had precipitated
(Fig. 4B).

1117

4.7. Trace metal and arsenic distribution in ne-grained sediments and


soils
The precipitation of Fe and Al was controlled by the pH increase
with depth. Iron was found at lower depth and lower pH than Al.
This is in agreement with investigations at the micro scale for carbonate grains reacting with AMD (Hammarstrom et al., 2003) and
also with eld-size studies of oxic limestone drains (Rtting et al.,
2008), which found a sequence of Fe- and Al-rich layers at the top,
followed by gypsum with carbonate below. Controlling factors for
Fe and Al precipitation and trace element distribution were the pH
gradient and Ca diffusion for gypsum formation (Hammarstrom
et al., 2003).
In CPY25, from the surface to 31 cm depth, concentrations increased from 197 to 5177 mg/kg Zn and from 78 to 808 mg/kg
Mn. This trend was observed in all proles, where calcite was present. These values exceeded the concentration in bedrock (up to
24 mg/kg Zn and up to 74 mg/kg Mn) by 215 times for Zn and 11
times for Mn and implied that the bedrock was not an important
source of Zn and Mn in the overlying layers. Modeling of mean lake
water in equilibrium with calcite resulted in SIs of 0.2 for rhodochrosite (MnCO3) and 0.2 for smithsonite (ZnCO3). Manganese
and Zn were probably xed as carbonate minerals at depths where
calcite was still oversaturated.
Adsorption capacity is controlled by the pH, and increases for
oxyanions at low pH and for cations at high pH. Enrichment in
As was correlated to the presence of plant roots: it was enriched
up to 2402 mg/kg in CPY07(1518 cm) and up to 3328 mg/kg in
CPY23(1015 cm) (Fig. 5C and D). The As concentrations were less
correlated to the TOC than to the visibly detected plant roots. In the
rhizosphere, Fe-hydroxides form plaques near the roots due to
downwards O2(g) uxes (Hansel et al., 2001), and As is adsorbed
on the Fe plaques (Hansel et al., 2002; Voegelin et al., 2007). In
some cases, high As concentrations were directly correlated to
the presence of Fe due to adsorption on Fe-hydroxides, as in
CPY06(1421 cm) with 9.0 wt.% Fe2O3 and 2157 mg/kg As (Dzombak and Morel, 1990). Lead, Cu and Cd were found where the pH
increased. Because Pb was always found above Cd in the proles,
a sorption sequence as found by Dzombak and Morel (1990) can
be assumed.
Trace element migration with depth can be inferred in the sequence As < Pb  Cu < Cd < Zn = Mn. This is in relative agreement
with the migration sequence of trace elements in carbonate soils
after the Aznalcllar tailings spill, which showed low mobility for
As and Pb, moderate mobility for Cu, and high mobility for Cd
and Zn (Dorronsoro et al., 2002). After the cessation of disposal,
metals and particularly As (the element of most environmental
concern in the studied case) may desorb and partly resorb during
pH changes and transformation processes within the gel-like and
gypsum layers, e.g. to goethite (Acero et al., 2006).
4.8. Lessons learned
Whereas jarosite was found to precipitate in the sediments of
various pit lakes (e.g. Bozau et al., 2007; Triantafyllidis and
Skarpelis, 2006), to the present authors knowledge Lake Yanamate
is the rst case where jarosite resulted in element cycling within
the water column. Thus, the analysis of the lake surface water
would not be sufcient to infer deposition of dissolved constituents by jarosite precipitation.
The discharge of kerosene had no detectable inuence on the
solubility of metals and As within the acidic lake water and did
not retain metals in the sediments. Possibly, it lowered the suspended solids within the water column, due to the formation of
OMA. However, this effect does not favor its discharge together
with acidic mine or process waters, because it has to be seen as a

1118

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119

remediation target on its own during closure planning. The organic


rich gel-like matter precipitated from the water column had a
water content of 90 wt.% and was not able to inhibit the inltration of AMD into the bedrock.
After 22 a of continuous interaction of highly acidic AMD with
limestone formations, the higher acidity at the lake bottom relative
to the lake surface water indicates that in Lake Yanamate AMD
neutralization of lake water by limestone was limited.
However, water balance modeling for the years 19962002 suggested inltration of 910 Mio. m3/a, exceeding discharge of 79
Mio. m3/a (Volcan, 2002). Although a sequence (Fe- and Al-hydroxides followed by gypsum, covering calcite) similar to precipitates
in oxic limestone drains developed, these layers did not stop the
inltration of lake water into the subsurface. Why the system
was not suffering from complete clogging (as expected from the
experiment with hand specimens) needs further study, and could
be related to the small grain size of the carbonate soil material,
or possible retention of CO2 within the lake bottom sediments,
increasing calcite dissolution (Arakaki and Mucci, 1995). Although,
hydraulically conductive faults and different forms of karst exist in
the lake basin, contamination of surface waters in the area surrounding Lake Yanamate is not reported. Possibly, the hydraulic
depression of the open pit and its intersection by the Cerro de Pasco-fault caused preferential groundwater ow from the lake towards the open pit.

form of carbonates (Zn and Mn). The layers of OMA and more
importantly gypsum lowered the contact of AMD solutions with
the carbonate bedrock and thus hindered the neutralization of
the lake water. No distinct crust was found that could have acted
as a self-sealing protection mechanism, such as found by Schuiling
and van Gaans (1997) and inltration into the bedrock was equal
to discharge of AMD. Clogging and armoring of bedrock fractures
has not resulted in a considerable decrease in inltration of lake
water.
Acknowledgements
We thank the mining company Volcan Compaa Mineria S.A.A.
for logistic and nancial support during the eldwork, especially V.
Gobitz, F. Grimaldo, L. Osorio and W. Heredia. Further, we wish to
thank SVS Ingenieros S.A., in particular Ing. A. Samaniego for providing eld equipment and M. Torres for his help during sampling.
We thank W. Schmitz, G. Kommischau, and P. Schreck, University
of Leipzig, for analytical support and helpful discussions and the
team of the Centre danalyse minrale, University of Lausanne, particularly H.-R. Pfeifer and J. Smuda, for help during eldwork, discussions and improvement of the manuscript. Financial support
from the German Academic Exchange Service (DAAD) and from
the Swiss National Science Foundation (SNSF) is gratefully
acknowledged. The comments of two anonymous reviewers are
appreciated.

5. Conclusions
At Lake Yanamate, a combination of hydrogeochemical, stable
isotope, bulk chemical and mineralogical methods were used to
study the controls on processes of the lake water chemistry and
the watersediment interaction at the lake bottom.
The mean lake water composition did not differ signicantly
from the discharged AMD and was dominated by high Fe and
SO4 concentrations and acidity. The lake showed an increase of
electrical conductivity and the concentrations of most elements
with depth. Stable isotope results for d2H and d18O indicate a local
origin of discharged AMD. Trends of stable isotopes within the lake
water body showed the inuence of evaporation for the upper
10 m and the local inltration of near-surface groundwater for
the eastern part of the lake. Below 15 m a stagnant zone with
near-equal values for stable isotopes and element concentrations
had developed. Modeling suggested that gypsum, anglesite, strengite and jarosite to some extent controlled the lake water chemistry.
Anglesite caused a homogeneous concentration of 2.8 mg/L Pb in
the lake, whereas jarosite formed within the upper part of the lake,
settled, and dissolved in the deeper part of the water column,
resulting in downward element transport. This cycling was conrmed by the presence of jarosite being limited to cores from
67 m water depth. The low pH was the limiting factor in mineral
precipitation. Reaction of lake water with carbonate bedrock did
not cause lower acidity or higher pH in the bottom layer of the
lake.
Discharged kerosene most likely caused the occulation of Oil
mineral agglomerates (OMA) that settled and formed a gel-like
layer on top of the lake bottom. Metal and As concentrations recalculated to a wet base showed no enrichment within this material
compared to the overlying water, thus the OMA did not adsorb
these elements. In the sediments below, a layer of gypsum with
pH 11.5 was found on top of dissolving calcite with pH 5.37.
In many samples the pore water pH was around 5.3, in accordance
to the modeled pH for lake water in equilibrium with calcite. For
the sediments, a migration sequence of As < Pb  Cu < Cd <
Zn = Mn was found. This sequence is explained by adsorption on
Fe-hydroxides at low pH (As), adsorption at increasing pH near
the gypsum/calcite boundary (Pb, Cu and Cd), and xation in the

References
Acero, P., Ayora, C., Torrent, C., Nieto, J.-M., 2006. The behavior of trace elements
during schwertmannite precipitation and subsequent transformation into
goethite and jarosite. Geochim. Cosmochim. Acta 70, 41304139.
Allison, J.D., Brown, D.S., Novo-Gradac, K.J., 1991. MINTEQA2, A geochemical
assessment data base and test cases for environmental systems: Vers. 3.0 users
manual, EPA, Athens.
Arakaki, T., Mucci, A., 1995. A continuous and mechanistic representation of calcite
reaction-controlled kinetics in dilute solutions at 25 C and 1 atm total
pressure. Aquat. Geochem. 1, 105130.
Baumgartner, R., Fontbot, L., Vennemann, T., 2008. Mineral zoning and
geochemistry of epithermal polymetallic ZnPbAgCuBi mineralization at
Cerro de Pasco. Peru. Econ. Geol. 103, 493537.
Bigham, J.M., Nordstrom, D.K., 2000. Iron and aluminum hydroxysulfates from acid
sulfate waters. In: Alpers, C.N., Jambor, J.L., Nordstrom, D.K. (Eds.), Sulfate
Minerals Crystallography, Geochemistry, and Environmental Signicance.
Mineralogical Society of America/Geochemical Society, Washington, DC, pp.
351403.
Bowell, R.J., Parshley, J.V., 2005. Control of pit-lake water chemistry by secondary
minerals, Summer Camp pit, Getchell mine. Nevada. Chem. Geol. 215, 373385.
Bozau, E., Bechstedt, T., Friese, K., Frmmichen, R., Herzsprung, P., Koschorreck, M.,
Meier, J., Vlkner, C., Wendt-Potthoff, K., Wieprecht, M., Geller, W., 2007.
Biotechnological remediation of an acidic pit lake: Modelling the basic
processes in a mesocosm experiment. J. Geochem. Explor. 92, 212221.
Bragg, Y.R., Yang, S.H., 1995. Clayoil occulation and its effects on the rate of
natural cleansing in Prince William Sound following the Exxon Valdez oil spill.
In: Wells, P.G., Butler, J.N., Hughes, J.S. (Eds.), Exxon Valdez Oil Spill Fate and
Effects in Alaskan Waters. American Society of Testing and Materials,
Philadelphia, pp. 178214.
Castendyk, D.N., Mauk, J.L., Webster, J.G., 2005. A mineral quantication method for
wall rocks at open pit mines, and application to the Martha AuAg mine, Waihi,
New Zealand. Appl. Geochem. 20, 135156.
Castro, J.M., Moore, J.N., 2000. Pit lakes; their characteristics and the potential for
their remediation. Environ. Geol. 39, 12541260.
Cravotta III, C.A., 2008. Laboratory and eld evaluation of a ushable oxic limestone
drain for treatment of net-acidic drainage from a ooded anthracite mine,
Pennsylvania, USA. Appl. Geochem. 12, 34043422.
Cravotta III, C.A., Trahan, M.K., 1999. Limestone drains to increase pH and remove
dissolved metals from acidic mine drainage. Appl. Geochem. 14, 581606.
Denimal, S., Bertrand, C., Mudry, J., Paquette, Y., Hochart, M., Steinmann, M., 2005.
Evolution of the aqueous geochemistry of mine pit lakes Blanzy-Montceaules-Mines coal basin (Massif Central, France): origin of sulfate contents; effects
of stratication on water quality. Appl. Geochem. 20, 825839.
Dold, B., Wade, C., Fontbot, L., 2009. Water management for acid mine drainage
control at the polymetallic ZnPb(AgBiCu) deposit Cerro de Pasco, Peru. J.
Geochem. Explor. 100, 133141.
Dorronsoro, C., Martin, F., Ortiz, I., Garca, I., Simn, M., Fernndez, E., Aguilar, J.,
Fernndez, J., 2002. Migration of trace elements from pyrite tailings in
carbonate soils. J. Environ. Qual. 31, 829835.

C. Wisskirchen et al. / Applied Geochemistry 25 (2010) 11071119


Dzombak, D.A., Morel, F.M.M., 1990. Surface Complexation Modeling Hydrous
Ferric Oxides. Wiley, New York.
Eary, L.E., 1999. Geochemical and equilibrium trends in mine pit lakes. Appl.
Geochem. 14, 963987.
Enaudi, M.T., 1977. Environment of ore deposition at Cerro de Pasco, Peru. Econ.
Geol. 72, 893924.
Ficklin, W.H., Mosier, E.L., 1999. Field methods for sampling and analysis of
environmental samples for unstable and selected stable constituents. In:
Plumlee, G.S., Logsdon, M.J. (Eds.), The Environmental Geochemistry of
Mineral Deposits. Society of Economic Geologists, Washington, pp. 249264.
Flusche, M.A., Seltzer, G., Rodbell, D., Siegel, D., Samson, S., 2005. Constraining water
sources and hydrologic processes from the isotopic analysis of water and
dissolved strontium, Lake Junin, Peru. J. Hydrol. 312, 113.
Geller, W., Klapper, H., Salamons, W., 1998. Acidic Mining Lakes Acid Mine
Drainage, Limnology and Reclamation. Springer, Berlin, Heidelberg, New York.
Gonatini, R., 1986. Environmental isotopes in lake sediments. In: Fritz, P., Fontes, J.
(Eds.), Handbook of Environmental Geochemistry, The Terrestrial Environment,
vol. 2. B. Elsevier, New York, pp. 113168.
Hamilton-Taylor, J., Davision, W., 1995. Redox-driven cycling of trace elements in
lakes. In: Lerman, A., Imboden, D., Gat, J. (Eds.), Physics and Chemistry of Lakes.
Springer, Berlin, Heidelberg, New York, pp. 217263.
Hamilton-Taylor, J., Smith, E.J., Davison, W., Sugiyama, M., 2005. Resolving and
modeling the effects of Fe and Mn redox cycling on trace metal behavior in a
seasonally anoxic lake. Geochim. Cosmochim. Acta 69, 19471960.
Hammarstrom, J.M., Sibrell, P.L., Belkin, H.E., 2003. Characterization of limestone
reacted with acid-mine drainage in a pulsed limestone bed treatment system at
the Friendship Hill National Historical Site, Pennsylvania, USA. Appl. Geochem.
18, 17051721.
Hansel, C.M., Fendorf, S., Sutton, S., Newville, M., 2001. Characterization of Fe plaque
and associated metals on the roots of mine-waste impacted aquatic plants.
Environ. Sci. Technol. 35, 38633868.
Hansel, C.M., La Force, M.J., Fendorf, S., Sutton, S., 2002. Spatial and temporal
association of As and Fe species on aquatic plant roots. Environ. Sci. Technol. 36,
19881994.
Jenks, W.F., 1951. Triassic and tertiary stratigraphy near Cerro de Pasco, Peru. Geol.
Soc. Am. Bull. 62, 203220.
Knller, K., Fauville, A., Mayer, B., Strauch, G., Friese, K., Veizer, J., 2004. Sulfur
cycling in an acid mining lake and its vicinity in Lusatia, Germany. Chem. Geol.
204, 303323.
Komnitsas, K., Bartzas, G., Paspaliaris, I., 2004. Efciency of limestone and red mud
barriers: laboratory column studies. Miner. Eng. 17, 183194.
Koschorreck, M., Bozau, L., Frmmichen, R., Geller, W., Herzsprung, P., WendtPotthoff, K., 2007. Processes at the sediment water interface after addition of
organic matter and lime to an acid mine pit lake mesocosm. Environ. Sci.
Technol. 41, 16081614.
Lee, K., 2002. Oil-particle interactions in aquatic environments: inuence on the
transport, fate, effect and remediation of oil spills. Spill Sci. Technol. Bull. 8, 38.
Luna Bernal, R., Delgado Venero, O., 1985. Origin and treatment of underground
waters from Cerro de Pasco mine, Peru. In: Proc. Mine Water Congress Granada,
Spain, pp. 2740.
Morgenstern, P., Friese, K., Wendt-Potthoff, K., Wennrich, R., 2001. Bulk chemistry
analysis of sediments from acid mine lakes by means of wavelength dispersive
X-ray uorescence. Mine Water Environ. 20, 105113.
Muschenheim, D.K., Lee, K., 2002. Removal of oil from the sea surface through
particulate interactions: review and prospectus. Spill Sci. Technol. Bull. 8, 918.
PAMA, 1996. Programa de Adecuacion y Manejo Ambiental (PAMA) UDP: Cerro de
Pasco, Empresa Minera del Cobre del Peru.
Parkhust, D.L., Appelo, C.A.J., 1999. Users guide to PHREEQC (Version 2) a
computer program for speciation, batch-reaction, one-dimensional transport,

1119

and inverse geochemical calculations. US Geol. Surv. Water-Resour. Invest. Rep.,


pp. 994259.
Pellicori, D.A., Gammons, C.H., Poulson, S.R., 2005. Geochemistry and stable isotope
composition of the Berkeley pit lake and surrounding mine waters, Butte,
Montana. Appl. Geochem. 20, 21162137.
Pelmont, J., 2005. Biodgradations et Mtabolismes. Collection Grenoble Sciences,
Grenoble.
Petersen, U., 1965. Regional geology and mayor ore deposits of Central Peru. Econ.
Geol. 60, 407476.
Plumlee, G.S., 1999. The environmental geology of mineral deposits. In: Plumlee,
G.S., Logsdon, M.J. (Eds.), The Environmental Geochemistry of Mineral Deposits.
Part A: Processes, Techniques and Health Issues. Reviews Econ. Geol. 6A, 71
116.
Plumlee, G.S., Logsdon, M.J. (Eds.), 1999. The Environmental Geochemistry of
Mineral Deposits. Part A: Processes, Techniques and Health Issues. Reviews
Econ. Geol. 6A, 1111.
Rose, A.W., Bisko, D., Daniel, A., Bower, M.A., Heckman, S., 2004. An autopsy of the
failed Tangaskootack #1 vertical ow pond, Clinton Co., Pennsylvania. In: Joint
Conf. 21st Ann. Meetings Am. Soc. Mining and Reclamation and 25th West
Virginia Surface Mine Drainage Task Force Symp., Morgantown, WV, pp. 1580
1594.
Rtting, T.S., Caraballo, M.A., Serrano, J.A., Ayora, C., Carrera, J., 2008. Field
application of calcite Dispersed Alkaline Substrate (calcite-DAS) for passive
treatment of acid mine drainage with high Al and metal concentrations. Appl.
Geochem. 23, 16601674.
Snchez Espaa, J., Lpez Pamo, E., Santomia Pastor, E., Diez Ercilla, M., 2008. The
acidic mine pit lakes of the Iberian Pyrite Belt: an approach to their physical
limnology and hydrogeochemistry. Appl. Geochem. 23, 12601287.
Santomartino, S., Webb, J.A., 2007. Estimating the longevity of limestone drains in
treating acid mine drainage containing high concentrations of iron. Appl.
Geochem. 22, 23442361.
Schuiling, R.D., van Gaans, P.F.M., 1997. The waste sulfuric acid lake of the TiO2plant at Armyansk, Crimea, Ukraine, Part I. Self-Sealing as an environmental
protection mechanism. Appl. Geochem. 12, 181186.
Shevenell, L., Connors, K.A., Henry, C.D., 1999. Controls on pit lake water quality at
sixteen open-pit mines in Nevada. Appl. Geochem. 14, 669687.
Simon, M., Martn, F., Garca, I., Bouza, P., Dorronsoro, C., Aguilar, J., 2005.
Interaction of limestone grains and acidic solutions from the oxidation of
pyrite tailings. Environ. Pollut. 135, 6572.
Smuda, J., Dold, B., Friese, K., Morgenstern, P., Glaesser, W., 2007. Mineralogical and
geochemical study of element mobility at the sulde-rich Excelsior waste rock
dump from the polymetallic ZnPb(AgBiCu) deposit, Cerro de Pasco, Peru. J.
Geochem. Explor. 92, 97110.
Triantafyllidis, S., Skarpelis, N., 2006. Mineral formation in an acid pit lake from a
high-suldation ore deposit: Kirki, NE Greece. J. Geochem. Explor. 88, 6871.
USEPA, 1994. Technical Resource Document. Extraction and beneciation of ores
and minerals, Copper, vol. 4. Environmental Protection Agency, Washington.
Voegelin, A., Weber, F.A., Kretzschmar, R., 2007. Distribution and speciation of
arsenic around roots in a contaminated riparian oodplain soil: Micro-XRF
element mapping and EXAFS spectroscopy. Geochim. Cosmochim. Acta 71,
58045820.
Volcan, S.A.A., 2002. Control Hidrologico de la Laguna Yanamate. Volcan Compaa
Minera S.A.A., Unidad Economica Administrativa Cerro de Pasco. Internal
Report.
Ward, H.J., 1961. The pyrite body and copper orebodies, Cerro de Pasco mine,
Central Peru. Econ. Geol. 56, 402422.
Webb, J.A., Sasowsky, I.D., 1994. The interaction of acid mine drainage with a
carbonate terrain: evidence from the Obey River, north-central Tennessee. J.
Hydrol. 161, 327346.

Vous aimerez peut-être aussi