You are on page 1of 14

PUBLICATIONS

Global Biogeochemical Cycles


RESEARCH ARTICLE
10.1002/2013GB004780
Key Points:
Hg deposition did not vary with
past precipitation, temperature,
and volcanism
Maximum Holocene Hg uxes
occurred ~3 thousand years ago
Modern Hg uxes are 3 times greater
than natural uxes

Supporting Information:
Readme
Table S1
Table S2
Table S3
Table S4
Table S5
Figure S1
Figure S2
Figure S3
Figure S4
Figure S5
Correspondence to:
S. A. Beal,
samuel.a.beal.gr@dartmouth.edu

Citation:
Beal, S. A., M. A. Kelly, J. S. Stroup, B. P.
Jackson, T. V. Lowell, and P. M. Tapia
(2014), Natural and anthropogenic variations in atmospheric mercury deposition
during the Holocene near Quelccaya Ice
Cap, Peru, Global Biogeochem. Cycles, 28,
doi:10.1002/2013GB004780.
Received 26 NOV 2013
Accepted 27 MAR 2014
Accepted article online 31 MAR 2014

Natural and anthropogenic variations in atmospheric


mercury deposition during the Holocene
near Quelccaya Ice Cap, Peru
Samuel A. Beal1, Meredith A. Kelly1, Justin S. Stroup1, Brian P. Jackson1, Thomas V. Lowell2,
and Pedro M. Tapia3
1

Department of Earth Sciences, Dartmouth College, Hanover, New Hampshire, USA, 2Department of Geology, University of
Cincinnati, Cincinnati, Ohio, USA, 3Department of Biological Sciences, Universidad Peruana Cayetano Heredia, Lima, Peru

Abstract Mercury (Hg) is a toxic metal that is transported globally through the atmosphere. Emissions
of Hg from mineral reservoirs and recycling between soil/biomass, oceans, and the atmosphere are
fundamental to the global Hg cycle, yet past emissions from anthropogenic and natural sources are not
fully constrained. We use a sediment core from Yanacocha, a headwater lake in southeastern Peru, to study
the anthropogenic and natural controls on atmospheric Hg deposition during the Holocene. From 12.3 to
3.5 ka, Hg uxes in the record are relatively constant (mean 1: 1.4 0.6 g m 2 a 1). Past Hg deposition
does not correlate with changes in regional temperature and precipitation or with most large volcanic
events that occurred regionally (~300400 km from Yanacocha) and globally. In 1450 B.C. (3.4 ka), Hg uxes
abruptly increased and reached the Holocene-maximum ux (6.7 g m 2 a 1) in 1200 B.C., concurrent with
a ~100 year peak in Fe and chalcophile metals (As, Ag, Tl) and the presence of framboidal pyrite.
Continuously elevated Hg uxes from 1200 to 500 B.C. suggest a protracted mining-dust source near
Yanacocha that is identical in timing to documented pre-Incan cinnabar mining in central Peru. During
Incan and Colonial time (A.D. 14501650), Hg deposition remains elevated relative to background levels
but lower relative to other Hg records from sediment cores in central Peru, indicating a limited spatial
extent of preindustrial Hg emissions. Hg uxes from A.D. 1980 to 2011 (4.0 1.0 g m 2 a 1) are 3.0 1.5 times
greater than preanthropogenic uxes.

1. Introduction
Rapidly rising anthropogenic emissions of mercury (Hg) to the atmosphere during the past decade are
superimposed on a longer-term increasing trend since the industrial revolution [Streets et al., 2011]. Hg is
transported globally as gaseous Hg0 [e.g., Mason et al., 1994], deposited to the land/water surface as Hg2+,
and rapidly transferred to biota as extremely toxic methyl-Hg [Harris et al., 2007], posing a great risk to human
and ecosystem health. An accurate understanding of the global Hg cycle is required to assess the role of
anthropogenic emissions on current and future Hg deposition. Information on the biogeochemical cycling of
Hg primarily comes from reconstructions of Hg deposition over time in sedimentary archives (i.e., lake
sediment, peat, and ice) and from global Hg models. A wealth of lake sediment records from around the
world provide direct evidence for an average 3.5-fold increase in Hg deposition since ~ A.D. 1850 [Biester
et al., 2007], but very few records extend earlier in time. A recent model of global Hg cycling, forced with
estimates of anthropogenic Hg emissions from 2000 B.C. to A.D. 2008 and constant natural emissions, yields a
similar amount of increase (2.6 times) since A.D. 1840 but a much larger increase (7.5 times) since 2000 B.C.
[Amos et al., 2013]. The apparent importance of anthropogenic emissions before ~ A.D. 1850 (i.e., during
preindustrial time) and the assumption of constant natural emissions require independent validation with
geophysical evidence, such as Hg contained in sedimentary archives.
Natural variations in Hg emissions to the atmosphere can be caused by changes in volcanism, low-temperature
volatilization, and external factors which affect exchanges between surface Hg reservoirs (soil/biomass, ocean,
and atmosphere) [Fitzgerald and Lamborg, 2007]. Terrestrial volcanic Hg sources are somewhat constrained
[Nriagu and Becker, 2003; Pyle and Mather, 2003], but large uncertainties remain in estimates of the inputs from
submarine volcanism [Lamborg et al., 2006] and low-temperature volatilization [Gustin et al., 2000] due to
limited observational data. A number of factors are thought to affect the exchange of Hg between surface

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles


a.)

80W

75W

70W

65W

10.1002/2013GB004780

b.)

5S

Yanacocha

Peru

Quelccaya
Ice Cap

10S
Huancavelica
Negrilla

Yanacocha

15S

MU
H

200 400

800
km

To Potos

4
km

c.)

Figure 1. (a) Digital elevation model of northwestern South America with the locations of the study lake (Yanacocha), Laguna Negrilla, the mining center of
Huancavelica, and the late Holocene-active volcanoes El Misti (M), Ubinas (U), Huaynaputina (H), and Yucamane (Y) in the Andean CVZ. Black arrows represent
NCEP/NCAR reanalysis V1 annual average vector wind at 500 mb from A.D. 1948 to 2012 [Kalnay, 1996]. (b) False-color Landsat 7 image of Quelccaya Ice Cap (light
blue) and the location of Yanacocha. Bedrock ridges are apparent in gray-red colors, whereas glacially carved valleys with vegetation are green. (c) 180 panoramic
image of the Yanacocha basin looking toward the east/northeast. The headwall is approximately 100 m above lake surface at its highest point. Red marker denotes
approximate coring location for the YC1 and YANA11 cores.

reservoirs including biomass burning [Friedli et al., 2003], permafrost thaw/freeze [Rydberg et al., 2010], and
oceanic evasion [Strode et al., 2007].
Use of Hg by humans began as early as 1500 B.C. in Egypt and Peru and continued later in parts of Asia and
the Roman Empire [Nriagu, 1979; Cooke et al., 2009]. This early use primarily consisted of extracting the
common mineral form cinnabar (HgS) as the bright red pigment vermilion, although there are also early
accounts of metal amalgamation using liquid Hg0 [Nriagu, 1979]. Anthropogenic Hg emissions increased
dramatically in the late sixteenth century when Hg amalgamation for silver extraction was introduced to
South and Central America [Nriagu, 1993]. Hg was emitted during the smelting of cinnabar to form liquid Hg0,
which occurred extensively in Huancavelica, central Peru, and during the heating of silver amalgams, which
occurred throughout the Andes but most notably in Potos, Bolivia (Figure 1) [Robins and Hagan, 2012].
Estimates of preindustrial Hg emissions are based on historical records and anecdotes of past metal use
coupled with assumed emission factors, and they are subject to high uncertainty [Nriagu, 1993; Streets et al.,
2011]. In addition, the spatial distribution of Hg emissions from preindustrial mining remains uncertain. There
is strong evidence for local deposition in highly enriched soils and sediments near mining sites [Cooke et al.,
2009; Robins et al., 2012], limited evidence for regional (~200500 km) transport [Beal et al., 2013; Cooke et al.,
2013], and no evidence for an impact of preindustrial Hg emissions on a global scale [Lamborg et al., 2002].
In this study, we reconstruct atmospheric Hg deposition during the Holocene in a sediment core from a
headwater lake in southeastern Peru near Quelccaya Ice Cap (QIC). Past Hg deposition is recorded reliably in
lake sediments and is not affected by diagenetic changes [e.g., Biester et al., 2007; Rydberg et al., 2008]. We use
this continuous record of atmospheric Hg deposition and coregistered proxies for paleoenvironmental
change to (1) assess natural variability in Hg deposition by comparing the Hg record to local and regional
paleoclimate conditions and major volcanic eruptions, (2) evaluate the impact of preindustrial anthropogenic
emissions on Hg deposition in the study lake by examining the Hg record during periods of known

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

preindustrial metal use, and (3) quantify the extent of anthropogenic modication to natural Hg cycling by
calculating an atmospheric deposition Hg ux ratio using modern and preanthropogenic Hg uxes.

2. Study Site
The study lake informally known as Yanacocha is located in the South Fork valley on the western side of
Quelccaya Ice Cap (QIC) in the Cordillera Vilcanota of southeastern Peru (13.945S, 70.875W, 4910 m above
sea level; Figure 1). Yanacocha is a tarn that occupies 0.036 km2 in a catchment of 0.11 km2. The catchment is
composed of a sparsely vegetated and gently sloping colluvial apron that extends radially ~100 m from the
edge of the lake, beyond which a near-vertical ~100 m high ignimbrite bedrock headwall surrounds the
north, east, and south sides of the lake (Figure 1). Inows are limited to surface runoff from the catchment,
and a single outow on the west side of the lake is active only during the wet season. During the eld season
in June 2011, the lake exhibited constant pH (~8), temperature (~6 C), and conductivity (~10 S) with depth
(Figure S1 and Table S1), characteristic of a holomictic lake.
Situated near the eastern edge of the Andes at 4910 m above sea level, Yanacocha likely receives most of its
precipitation from easterly middle-upper troposphere ows in Austral Summer that bring moisture from the
Amazon Basin [e.g., Garreaud et al. [2003]]. Precipitation and atmospheric conditions at the study site have
likely changed with the position of the Intertropical Convergence Zone and El NioSouthern Oscillation,
with drier conditions during modern-day El Nio and wetter conditions during modern-day La Nia [e.g.,
Garreaud et al., 2003].
An expanded QIC prior to ~12.8 ka (kiloannum; dened here as thousands of years before A.D. 1950) had a
terminus position ~2 km downvalley from Yanacocha, covering the lake with glacial ice [Kelly et al., 2012].
Retreat of QIC began ~12.3 ka, leaving the Yanacocha catchment by at least 11.6 ka and remaining ~3 km
upvalley of Yanacocha during the Holocene [Kelly et al., 2012]. The bedrock of the headwall surrounding
Yanacocha prevented inows of QIC meltwater from entering the lake during the Holocene. Therefore, any
material transported to the lake occurred either by surface runoff within the relatively small catchment or
atmospheric deposition.
Yanacocha is removed from major development. The nearest major population center is Cusco, located
~130 km away. Present-day land use in the vicinity of Yanacocha is limited to sparse livestock grazing. We
are not aware of any mining near the margins of QIC, and although there are currently no large-scale
mining operations in the region, a large silver-lead-zinc mine is in planning stages ~25 km northwest of
Yanacocha. Small-scale and artisanal gold mining is prevalent in the Amazon basin ~120 km away, but this
mining was shown not to be a major contributor of Hg to high-elevation lakes in southeastern Peru [Beal
et al., 2013].

3. Methods
3.1. Core Collection and Processing
We collected a long (4 m) sediment core, YANA11, near the center of Yanacocha and at its greatest water depth
(5.5 m) in June 2011. We used a Bolivian coring system from a oating platform to retrieve ~1 m drives of
sediment into polycarbonate tubes, collecting two adjacent cores offset by ~50 cm. Core tubes were capped,
kept unfrozen in the eld, and then shipped from Cusco to the National Lacustrine Core Facility (LacCore) at the
University of Minnesota. At LacCore, we split the polycarbonate core tubes and took high-resolution core
images. Working halves of each core drive were shipped to Dartmouth College for subsequent analyses, and
archive halves are stored at the LacCore repository.
We also collected a short (40 cm) sediment core, YC1, adjacent to the YANA11 core using a gravity corer
that preserves the sediment-water interface. This core was collected prior to YANA11 to avoid disturbance
of the sediment-water interface. YC1 was extruded in the eld at 1 cm intervals and stored in Whirlpak
bags [Beal et al., 2013].
3.2. Geochemical Analyses
We sampled the YANA11 core at continuous 1 cm intervals using acid-clean polystyrene spoons. The samples
from YANA11 and YC1 were freeze-dried in new polypropylene centrifuge tubes, homogenized in an agate

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

mortar and pestle, and subsampled for loss on ignition (LOI), biogenic silica (BSi), major and trace metals, and
heavy mineral separations.
3.2.1. Loss on Ignition and Biogenic Silica
We performed LOI in three stages: 110C overnight, 550C for 4 h, and 1000C for 2 h. BSi was determined at
Northern Arizona University by molybdate-blue reaction and spectrophotometry following Mortlock and
Froelich [1989]. Bulk density was calculated based on water content and assumed densities for the organic
(1.4 g cm 3), carbonate (2.7 g cm 3), and inorganic (2.0 g cm 3) components determined by LOI.
3.2.2. Major and Trace Metals
We determined total Hg using a Milestone DMA-80 on ~50 mg subsamples. One of the Standard Reference
Materials (SRMs) IAEA-SL-1 (lake sediment), STSD-1 and STSD-2 (stream sediment), and NIST-1547 (peach
leaves) was run every 10 samples. Measured SRM concentrations (Table S2) were within their published 95%
condence intervals. Sample replicates were run every 10 samples with typical precision (relative percent
difference for n = 2, relative standard deviation for n 3) of less than 10%. We also extracted ~200 mg
subsamples by strong acid (9:1 HNO3:HCl) in open microwave vessels at 90C and analyzed the leachates for
metal concentrations (henceforth referred to as Mext) by quadrupole ICP-MS (Agilent 7700x), running
calibration checks and blanks every 10 samples. Typical precision on replicate samples for detectable analytes
was less than 10%. All concentrations are expressed as mass of metal per mass of dry sediment. In addition, total
metals were measured at 0.5 cm resolution on archive core halves by ITRAX core-scanning XRF at the University
of Minnesota Duluth with a dwell time of 30 s [Croudace et al., 2006].
3.2.3. Heavy Mineral Separation and Analysis
We separated the heavy mineral fraction of selected samples by mixing ~500 mg of freeze-dried sediment
with 10 ml of sodium polytungstate adjusted to a density of 2.8 g cm 3, placing the mixtures in an ultrasonic
bath for 30 min and centrifuging the mixtures for 90 min at 4500 rpm. This separation procedure
accommodates a theoretical minimum cinnabar (8.1 g cm 3) particle diameter of 65 nm following the
equation in Plathe et al. [2013]. We rinsed the heavy fraction by following the above ultrasonic and
centrifugation steps with 10 ml of deionized H2O, repeated 3 times. We digested and analyzed selected heavy
fraction samples for metal concentrations (henceforth referred to as Mhvy) using the same methods
described above for bulk samples, while accounting for contamination by the heavy liquids with one
procedural blank for every ve samples. For certain nondigested samples, we dried the heavy fractions and
studied them using a scanning electron microscope (SEM; Hitachi TM3000) with energy-dispersive X-ray
spectroscopy (EDS).
3.3. Chronology
The composite record (henceforth the Yanacocha record) includes YC1 from 0 to 27 cm depth and
YANA11 from 27 to 333 cm depth. We correlated the offset drives from YANA11 based on visual
stratigraphy and then correlated YC1 to YANA11 using LOI550 (R = 0.90, p < 0.001; Figure S2). Age control
in the Yanacocha record consists of 12 210Pb ages from YC1 (details previously published in Beal et al.
[2013]) and 10 AMS 14C ages on macrofossils isolated from discrete sediment layers throughout YANA11
(Table S3). The age-depth model uses 210Pb ages from 0 to 12 cm, linear interpolation from 12 to 27 cm,
Bayesian age modeling from 27 to 303 cm using the program Bacon [Blaauw and Christen, 2011], and
linear interpolation from 303 to 333 cm (Figure 2). The following nondefault parameters were used in
Bacon: mean accumulation rate 0.0167 cm yr 1, memory strength 20, and memory mean 0.1. All ages
were calibrated in Bacon using SHCal04 [McCormac et al., 2004], except for the basal age which was
calibrated using IntCal09 [Reimer et al., 2011] because its 14C age exceeds SHCal04. Linear interpolation
was used from 303 to 333 cm because of deviations in Bacons t near the base of cores where
accumulation rates decrease rapidly.

3.4. Flux Calculations


Hg uxes were calculated as the product of concentration and dry mass sedimentation rate. Dry mass
sedimentation rate (g m 2 a 1) was calculated as the product of bulk density (g cm 3) and accumulation rate
(cm a 1). Because Bacon assumes unrealistic changes in accumulation rate between adjacent sections
[Jacobson et al., 2012], a second-order polynomial t was used to calculate accumulation rates (Figure S3).
BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

4. Results
4.1. Stratigraphy and Age-Depth Model
The composition of the Yanacocha record is a
diatomaceous gyttja from the top of the core to
a depth of 333 cm (Figure 2). Below 333 cm, the
lithology is uniformly silt and clay. A macrofossil
just above this abrupt transition from silt and
clay to gyttja dates to 12.3 ka and likely marks
the termination of meltwater input caused by
recession of QIC behind the bedrock ridge
surrounding Yanacocha. This basal age is older
than two previously reported 14C ages (both
11.2 ka) in Yanacocha basal sediments from
slightly above the transition in another core
[Kelly et al., 2012]. The Yanacocha record
exhibits constant sedimentation throughout
the Holocene with no evidence for a hiatus in
either the age-depth model or the stratigraphy.
4.2. Holocene Sedimentology
Organic matter (LOI550) and BSi, proxies for
productivity in the lake, each comprise between
~20 and 60% of Yanacocha sediments and are
signicantly inversely correlated throughout
the Holocene (Figure 3). BSi is high (~3855%)
and LOI550 is low (~1030%) from 12.3 ka to
6.5 ka (Figure 4), followed by relatively low BSi
(~2638%) and high LOI550 (~3150%) from 6.5
to 4.7 ka. Subsequent to 4.7 ka, BSi and LOI550
remain within their early Holocene values,
except for a brief reversal from 1.1 to 0.6 ka
when BSi is low and LOI550 is high. Total Ti, a
proxy for total lithogenic input, is relatively high
in the early Holocene from ~12.3 to 9 ka,
followed by lower values from ~7 to 5 ka.
Higher than average Ti persists from ~4.8 to
3 ka and then is variable from ~3 ka through
the late Holocene.
Figure 2. Composite image for the YANA11 core and the age-depth
14
210
Pb
model including calibrated C age ranges (blue points) and
ages (red points).

4.3. Hg Variability During the Holocene

Hg concentrations in the Yanacocha record


range from a minimum of 13 g kg 1 at
1
8.9 ka to a maximum of 115 g kg at 3.2 ka (Figure 4). Pre-3.5 ka Hg concentrations are relatively stable
(mean 1: 32 9 g kg 1), except from ~10 to 9 ka when Hg concentrations are relatively elevated
(~4060 g kg 1). An abrupt increase in Hg concentration occurs at 3.4 ka and reaches the Holocene
maximum concentration at 3.2 ka, followed by a steady decline to pre-3.5 ka values by ~2.5 ka. Slightly
elevated Hg concentrations (~45 g kg 1) persist from 1.5 to 0.5 ka. An abrupt increase beginning
in ~ A.D. 1480 is followed by consistently elevated concentrations (4675 g kg 1) until the most recent
sediment in A.D. 2011.
The record of Hg ux is largely a reection of the record of Hg concentration, as it is the product of Hg
concentration and sedimentation rate (Figure S4). Pre-3.5 ka Hg uxes are ~1.01.5 g m 2 a 1, compared to a
maximum of 6.7 g m 2 a 1 at 3.2 ka and average post-A.D. 1980 uxes of ~4.1 g m 2 a 1. The main deviation
of Hg ux from concentration occurs from ~1.5 to 0.5 ka, concurrent with increased LOI550 (Figure 4). Because
BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

60
R2 = 0.5317
p < 0.001

BSi (%)

50
40
30
12 ka

20
10
0

10

20

30

40

50

60

estimates of Hg ux are subject to high


uncertainty, particularly in older records such as
this one that are dependent upon a limited
number of ages, we use Hg concentrations to
determine secular changes in Hg deposition and
time-averaged Hg uxes to calculate an Hg
ux ratio.
4.4. Heavy Mineral Characterization

LOI550(%)

The composition and morphology of minerals


contained in the heavy fraction of sediment
Figure 3. Correlation between organic matter (LOI550) and BSi
(>2.8 g cm 3) provide insight into the role of
content in the Yanacocha record from 12 to 0 ka.
sulde minerals in Hg deposition. We analyzed 14
heavy fraction samples for metal composition and
six heavy fraction samples by SEM, with a particular focus on the period 3.33.2 ka that is characterized by a
peak in Feext concentrations and Holocene-maximum Hg concentrations (Figure 5b). We did not identify Hg
suldes in any of the six samples analyzed by SEM, but we found abundant framboidal pyrite in one sample
from 3.3 ka with diameters of 1015 m (Figure 5a) and Fe, S, and C spectral peaks identied by EDS. A
pronounced one-sample peak in concentrations of Fehvy and Shvy at 3.2 ka (Figure 5) has a molar Fe:S ratio of
1:1.79 similar to observed framboidal pyrite and highly elevated concentrations of Ashvy, Aghvy, and Tlhvy
[Large et al., 2001]. Although the Hghvy concentration is relatively elevated in this sample, the percent of Hg in
the heavy fraction (%Hghvy) is not relatively elevated (Figure 5b).

5. Discussion
5.1. Depositional Pathway
We rst test the hypothesis that atmospheric deposition is the primary source of Hg to Yanacocha by comparing
Hg concentrations and sedimentology in the Yanacocha record during the entire record (12.3 to 0 ka) and just

Figure 4. Hg deposition (concentration and ux) and coregistered proxies of environmental conditions (LOI550 and BSi) and
lithogenic input (Ti) in the Yanacocha record from 12.3 ka to A.D. 2011. Dashed line represents Holocene average total Ti.

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

a.)

10 m

b.)

10 m

80
40

5
0.2
0.6
0.4

10

Cuhvy

(g kg ) As

(mg kg ) Cu

0.2
10

2.0

Ashvy
1.0

5
0

4.0

(%)

10

Fehvy

(%) Fe /Fe

0.4

Shvy

(%) Cu /Cu

7
6
5
4
3
2
0.5
0.4
0.3
0.2
0.1
0.6

As /As

(mg kg ) Fe

(g kg )

(g kg )

Fe

(g kg )

Hg (g kg )

120

the preanthropogenic period (dened and


used herein as 12.3 to 3.5 ka, based on
previous records and historical information
[Nriagu, 1979; Martnez-Cortizas et al., 1999;
Cooke et al., 2009]). Previous millennial-scale
Hg records in lake sediments relate changes
in Hg uxes to groundwater level [Jacobson
et al., 2012], lithogenic input from weathering
within the catchment [Thevenon et al., 2011],
and mobilization of Hg from soils [Cannon
et al., 2003]. However, the lack of correlations
of Hg concentration with LOI550 and with Ti
(Figure 6) shows that organic matter and
lithogenic input, respectively, do not have
signicant effects on Hg deposition in
Yanacocha. The only statistically signicant
correlation is between Hg and LOI550 during
the preanthropogenic period, but this
correlation has a very weak effect
(R2 = 0.049). The absence of increased Hg
deposition when lithogenic input was
relatively high during the lakes early stage
(~12.3 to 11 ka; Figure 4) indicates that
weathering of surrounding bedrock is not a
signicant source of Hg. Based on these
correlations, the small catchment area, and the
lack of stream inputs, we conclude that
atmospheric Hg deposition is the primary
source of Hg to Yanacocha sediments.

(%)
Hg /Hg

Tl

(mg kg ) Hg

(g kg ) Ag

One exception to this interpretation is the


brief association of increased Feext
2.0
concentrations and framboidal pyrite with
near-maximum Hg concentrations from 3.3 to
15
40
3.2 ka. Framboidal pyrite often contains many
Hghvy
10
heavy metals including Hg [e.g., Schoonen
20
5
[2004]], presumably due to the afnity that Hg
has for S and the large pyrite surface area
0
0.4
afforded by the crystallite subunits within
Tlhvy
0.2
each framboid (e.g., Figure 5a). Chemical
preservation of framboidal pyrite is not
0.0
inuenced by diagenesis in lake sediments
0
1
2
3
4
5
6
7
Age (ka)
[Suits and Wilkin, 1998]. Framboidal pyrite is
formed either in euxinic water columns or
Figure 5. (a) SEM images of framboidal pyrite from the heavy mineral
within upper sediments, near the sedimentfraction of a Yanacocha sediment sample at 3.3 ka. (b) Comparison of
water interface, where anoxic conditions
the timing of the Hg peak at ~3 ka with framboidal pyrite presence
occur [Suits and Wilkin, 1998]. The relatively
(lled square) and absence (open squares), extractable Fe concentrations, and heavy mineral fraction metal concentrations (gray lines with
large diameters of the observed framboids
diamonds) and percentages (black circles). Gray shading highlights the
(1015 m; Figure 5a) and low modern water
~100 year period of elevated Feext concentrations.
sulfate concentration (238 g L 1; Table S1)
are consistent with formation within the
sediment as opposed to within the water column [Wilkin et al., 1996]. Therefore, we hypothesize that framboidal
pyrite was formed within Yanacochas uppermost sediments due to external input of oxidized Fe and S, which
may have sequestered Hg from the lake during the period of elevated Feext concentrations from 3.3 to 3.2 ka.
Aghvy

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles


a.)

80

12.3 - 0 ka 12.3 - 3.5 ka


R = 1E-5 R = 0.049
p = 0.002
p = 0.94
3.5 - 0 ka

60

12.3 - 3.5 ka

Hg (g kg-1)

100

120
12.3 - 0 ka 12.3 - 3.5 ka
100
R = 0.002 R = 0.003
p = 0.45
p = 0.38
80
3.5 - 0 ka

b.)

12.3 - 3.5 ka 60

40

40

20

20

20

40

LOI550 (%)

0.5

1.0

Hg (g kg-1)

120

10.1002/2013GB004780

0
1.5

Ti (kcps)

Figure 6. Correlations between Hg concentrations and (a) organic matter (LOI550) and (b) lithogenic input (Ti) during the
entire Yanacocha record from 12.3 to 0 ka and during preanthropogenic time from 12.3 to 3.5 ka.

Atmospheric Hg deposition to a lake can occur by oxidation of Hg0 vapor in the atmosphere to Hg2+, which then
binds with particles or is dissolved in precipitation, and by atmospheric transport and deposition of Hg-bearing
particles [e.g., Mason et al., 1994]. Either form of deposited Hg likely had a residence time on the order of years due
to Yanacochas small catchment size and surrounding colluvium that lacks developed soil. In the sections below,
we examine the possible inuences of climate (i.e., precipitation and temperature), volcanism, and anthropogenic
activity on atmospheric Hg deposition to Yanacocha.
5.2. Climate and Hg Deposition
We investigate the possible response of atmospheric Hg deposition to past changes in regional precipitation
and temperature during the preanthropogenic period using nearby paleoclimate records and coregistered
proxy data in the Yanacocha record. Long-term monitoring in North America has found that wet deposition
of Hg (Hg bound in precipitation) represents ~5090% of modern total atmospheric Hg deposition and can
vary seasonally [Prestbo and Gay, 2009]. Temperature has been proposed as a controlling factor on Hg in
sedimentary systems either directly, by affecting the thermal stability of Hg in peat [Martnez-Cortizas et al.,
1999], or indirectly, through warming-driven increases in algal productivity that scavenges Hg from the lake
water column [e.g., Outridge et al. [2007]].
5.2.1. Precipitation
Past precipitation in the central Andes is inferred primarily from lake level records and 18O records of
carbonate lake sediments and cave deposits. Lake level records from Lake Titicaca are interpreted to reect
precipitation amounts on the Altiplano (1421 S) and show high lake levels from ~21 to 10 ka, followed by
low lake levels from ~8 to 5 ka, [Baker et al., 2001; Rowe et al., 2002]. From ~4 ka to the present day, Lake
Titicaca levels were near modern and are interpreted to reect relatively high precipitation [Rowe et al., 2002].
Records of precipitation inferred from 18O values of carbonate lake sediments and cave deposits in central
Peru (11 S) show a similar pattern to Lake Titicaca, although they indicate more gradually increasing
precipitation throughout the Holocene [Bird et al., 2011; Kanner et al., 2013]. Precipitation-related changes in
biomass burning, which cause Hg release to the air primarily as Hg0 [Friedli et al., 2003], are unlikely due to the
rarity of natural res in the Amazon Basin during the Holocene [Bush et al., 2007].
We interpret BSi in the Yanacocha record as a proxy for precipitation on the catchment, resulting from runoff
delivery of Si to the lake for diatom frustule formation. An inverse correlation between BSi and LOI550
throughout the Yanacocha record (Figure 3) likely reects the dilution of organic matter deposition with
diatoms. We infer high precipitation at the lake from high BSi (low LOI550) prior to ~7 ka and lower
precipitation from near minimum BSi (maximum LOI550) between ~7 and 4.5 ka (Figure 4). Continuously
below average Ti between ~7 and 4.5 ka supports decreased lithogenic input from runoff during this time.
Subsequent to ~4.5 ka, generally high BSi (low LOI550) and approximately average Ti suggest relatively high
precipitation. The general pattern of precipitation changes interpreted from BSi in Yanacocha is similar to that
inferred from Lake Titicaca [Baker et al., 2001; Rowe et al., 2002] and 18O records in central Peru [e.g., Kanner
et al., 2013]. A shift in hydrology during the appearance of framboidal pyrite at ~3.3 ka is not evident in
the Yanacocha sedimentology, and regional proxy records do not show precipitation changes during this
time in central Peru [Kanner et al., 2013]. Despite millennial-scale changes in precipitation inferred from BSi

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

Year AD/BC
1000

-1000

-2000

100
80
60
40

Hg Flux (g m-2 a-1)

20
5

1
6

4
100
2

75

Hg Flux (g m-2 a-1)

Negrilla

50
25

104
0
103

LY1
LY2

102
10

Huaynaputina
Yucamane
5

Ubinas
El Misti
4
0

CVZ Eruptions (VEI)

Pbext (mg kg-1)

Hg Flux (g m-2 a-1)

Hg Conc. (g kg-1)

2000
120

Age (ka)

in the Yanacocha record, Hg


concentrations and uxes remain
relatively constant during
preanthropogenic time (Figure 4).
5.2.2. Temperature
While Holocene paleotemperature
proxies in the Central Andes are scarce,
some paleotemperature information
has been inferred from past glacier
extents of QIC [Kelly et al., 2012;
Thompson et al., 2013; Stroup et al.,
2014]. Radiocarbon ages of in situ
plants show that QIC was smaller than
at present prior to ~7 ka, suggesting
relatively warm conditions during this
time. A subsequent advance of QIC that
overran and entombed plants dating to
between ~7 and 5 ka [Thompson et al.,
2006, 2013; Buffen et al., 2009], during
relatively dry middle Holocene
conditions (see section 5.2.1), was likely
inuenced by cooling. Relatively
constant Hg concentrations and uxes
from ~8 to 5 ka (Figure 4) suggest that
regional temperatures did not strongly
inuence atmospheric Hg deposition in
Yanacocha. This nding is consistent
with an Hg record from lake sediments
in arctic Canada in which there is no
relationship between Hg deposition
and Holocene temperature changes
[Cooke et al., 2012].
5.3. Volcanism and Hg Deposition

Volcanic eruptions with a Volcanic


Explosivity Index (VEI) 6 (i.e., Plinian
eruptions that inject volcanic gases into
the stratosphere) are known to have
occurred throughout the Holocene
[Siebert and Simkin, 2002]. Hg records
from peat cores in Switzerland [RoosBarraclough et al., 2002] and ice cores in
Wyoming, United States [Schuster et al., 2002], report short-lived (~100 year for peat, ~110 year for ice) peaks
in Hg deposition, usually manifested as a greater than tripling of Hg ux, that are similar in timing to explosive
volcanic eruptions in both the Northern and Southern Hemispheres. Based on the temporal resolution of the
Yanacocha record (median = 26 years per sample), we would expect to nd Hg peaks during times of known
volcanic eruptions. However, Hg deposition in the Yanacocha record during the preanthropogenic period is
relatively stable, and eruption-related increases in Hg deposition are not distinguishable from the noise
(Figure 4). Continuous volcanic degassing and more frequent smaller eruptions may contribute signicant
amounts of natural Hg to the atmosphere [Pyle and Mather, 2003] but similarly cannot be distinguished in the
Yanacocha record.
Figure 7. The Yanacocha Hg (green and blue) and Pb (red) records compared with Hg ux records from Laguna Negrilla [Cooke et al., 2013] and
two lakes near Huancavelica (LY1 and LY2) [Cooke et al., 2009]. Also shown
are volcanic eruptions with a VEI of 4 in the Andean CVZ during the past
4000 years [Siebert and Simkin, 2002]. Gray shading highlights the early
and later phases of anthropogenic metal use in the Andes as dened by
Cooke et al. [2009].

The Andean Central Volcanic Zone (CVZ) is located ~300400 km from Yanacocha (Figure 1) and has
hosted a number of Plinian eruptions since ~3.5 ka (Figure 7). The VEI 5 eruption of the volcano Yucamane

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

Global Biogeochemical Cycles

10.1002/2013GB004780

(~3270 14C yr B.P.) [Siebert and Simkin, 2002] roughly overlaps in timing with the abrupt increase in Hg and
Feext concentrations and framboidal pyrite appearance from ~3.3 to 3.2 ka. Deposition of volcanic sulfate
and Fe to Yanacocha from this eruption may have provided adequate reactants for framboid formation within
the lakes surface sediments and sequestration of Hg from the water column or volcanic ash. Further evidence
for volcanism at ~3.2 ka comes from highly enriched Ashvy, Aghvy, and Tlhvy concentrations in Yanacocha
sediments (Figure 5b), which, in addition to having an afnity for framboidal pyrite [Schoonen, 2004; Neumann
et al., 2013], are also emitted predominantly from volcanic sources [Kellerhals et al., 2010]. If volcanism were
responsible for the sharp increase in Hg concentration at ~3.3 ka, then the Hg is retained in the less dense
fraction of sediment (< 2.8 g cm 3) or in nanometer-scale particles because of near-constant %Hghvy during
this time (Figure 5b).
The largest eruption in the CVZ during the Holocene was the VEI 6 eruption of the volcano Huaynaputina,
historically dated to 19 February, A.D. 1600. Ashfall from this eruption with particle diameters of ~20 m is
registered in ice cores from QIC [Thompson et al., 1986], and lava ows on Huaynaputina have a similar Fe
content (~36 wt %) to those on Yucamane [Mamani et al., 2008]. Hg concentrations in the Yanacocha record
do not register this volcanic eruption, but instead generally decline between A.D. 1590 and 1730 (Figure 7).
This nding is consistent with a lake sediment record from Southern Chile that shows relatively constant Hg
uxes within and subsequent to visible tephra layers from three separate Holocene eruptions [Hermanns and
Biester, 2013]. In contrast to the peat and ice core records that show volcanic Hg peaks, the overall lack of
volcanic events registered in the Yanacocha Hg record from both regional and global eruptions suggests that
large volcanic events during the Holocene had negligible decadal- to century-scale effects on atmospheric
Hg levels.
5.4. Anthropogenic Activity and Hg Deposition
5.4.1. 1450500 B.C.
An early phase of increased atmospheric Hg deposition in the Yanacocha record began at 1450 B.C. (3.4 ka),
reached a maximum at 1200 B.C. (3.15 ka), and remained elevated until at least 500 B.C. (2.45 ka; Figure 7).
This peak is not associated with a change in any of the other bulk analytes in the Yanacocha record except for
a brief peak in Feext concentrations from 1340 to 1240 B.C. associated with the presence of framboidal pyrite
and discussed above. A mining dust source of Fe and S for framboidal pyrite formation is unlikely due to the
low solubility of most sulde ore minerals. However, increased concentrations of Cuhvy, Cohvy, Nihvy, Mohvy,
and Pbhvy from 1650 to 1500 B.C. (Figures 5b and S5) suggest an early mining dust source to Yanacocha.
These metals are commonly found together within the same sulde deposits and can be accessible at the
surface in areas affected by glaciation in Peru [Petersen, 1965]. This period of enhanced chalcophile
deposition preceded the abrupt increase in Hg deposition at 1400 B.C. and is concurrent with a slight
monotonic increase in Hg concentration and ux. Following the peak in Hg deposition at 1200 B.C. , the
endurance of elevated Hg deposition (5.0 to 6.8 g m 2 a 1) for nearly a millennium implies a persistent
local anthropogenic source of Hg to Yanacocha. Furthermore, the shapes of Hg concentration and ux
peaks, characterized by onsets with abrupt increases and subsequent slow declines to background levels, are
similar to preindustrial anthropogenic peaks found in cores from the headwater lake Laguna Negrilla in Peru
(Figure 7) [Cooke et al., 2013] and a saltwater lagoon in France [Elbaz-Poulichet et al., 2011]. Near-constant %Hghvy
(Figure 5b) suggests that Hg from mining during this time was likely emitted either as ultrane (< 65 nm diameter)
cinnabar particles or as Hg0/Hg2+ that was subsequently bound to less dense materials.
The timing of the early phase of Hg deposition in Yanacocha is identical to precolonial cinnabar mining
registered in the lakes LY1 and LY2 located ~10 km from Huancavelica (Figure 7) [Cooke et al., 2009]. Cooke
et al. [2009] found that the Hg deposited during pre-Incan time was primarily bound as cinnabar, and neither
an increase in Hg uxes nor a distinct change in Hg isotopes was observed during pre-Incan time in a
sediment core from Laguna Negrilla, located ~200 km southeast of Huancavelica (Figure 1) [Cooke et al.,
2013]. This spatial limitation of Hg emissions from Huancavelica would have likely precluded the longer
distance transport to Yanacocha, located ~460 km southeast of Huancavelica (Figure 1), which suggests that
the early phase of Hg deposition in Yanacocha is from pre-Incan metal use near the catchment.
We hypothesize that the early phase of anthropogenic Hg deposition in Yanacocha was due to a three-part
sequence of events. First, mining of a nearby polymetallic sulde deposit provided minimal Hg contributions

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

10

Global Biogeochemical Cycles

10.1002/2013GB004780

Table 1. Time-Weighted Means for Hg Flux and Concentration During Periods Representative of Natural and
Anthropogenic Conditions
Flux (g m
Period
Pre-8 ka
Pre-6 ka
Pre-3.5 ka
Post-A.D. 1980
Post-A.D. 2000

Concentration (g kg

Mean

Mean

1.0
1.2
1.4
4.0
3.4

0.4
0.5
0.6
1.0
0.3

31
31
32
68
70

11
10
9
4
1

58
91
189
5
2

to Yanacocha from 1650 to 1450 B.C. Second, a combination of nearby mining emissions, potential volcanic
emissions, and/or an afnity of Hg for framboidal pyrite caused the abrupt increase in Hg deposition from
1450 to 1200 B.C. Third, ongoing nearby mining supplied decreasing amounts of Hg to Yanacocha from
1200 B.C. to at least 500 B.C.
5.4.2. A.D. 14802011
A later phase of enhanced atmospheric Hg deposition in Yanacocha is registered from ~ A.D. 1480 to 2011.
Increased Hg concentrations (5573 g kg 1) from ~ A.D. 1480 to 1640 (Figure 7) may reect both cinnabar
mining in Huancavelica, rst by the Inca from ~ A.D. 1450 and then by the Spanish from A.D. 1564 onward
[Cooke et al., 2009], and the concurrent growth of Ag rening using Hg amalgamation throughout the Andes
beginning ~ A.D. 1570 [Robins and Hagan, 2012]. A simultaneous peak in Pbext concentrations from ~ A.D.
1500 to 1670 (Figure 7) is similar in timing to the initial use of Pb for smelting Ag ores [Guerrero, 2012]. If Hg
was cotransported with aerosol-based smelting emissions, it must either reside as cinnabar with particle
diameters less than 65 nm (because %Hghvy does not change substantially (Figure 5b)) or as Hg adsorbed to
less dense aerosols. Atmospheric transport of Hg from Huancavelica to Laguna Negrilla between ~ A.D. 1450
and 1650 is supported by a pronounced increase in Hg uxes (~10 fold increase, up to 82 g m 2 a 1;
Figure 7) and a shift in the mass-dependent fractionation of Hg isotopes [Cooke et al., 2013]. The relatively
small increase in Hg deposition in Yanacocha compared to Laguna Negrilla suggests that Hg emissions from
Huancavelica were, at least during the time of Inca control (~A.D. 14501564), predominantly in the solid
phase and decreased in spatial extent with distance from Huancavelica.
The shift to elemental Hg production for silver mining between A.D. 1564 and 1810 likely inuenced more
globally distributed Hg emissions [Nriagu, 1993; Robins and Hagan, 2012]. Decreased Hg concentrations and
uxes in the Yanacocha record from ~ A.D. 1650 to 1750 are followed by a general increase coincident in
timing with estimated maximum Hg0 emissions in South and Central America from ~ A.D. 1750 to 1810
[Nriagu, 1993]. However, increasing Hg uxes are not evident during this period in Laguna Negrilla (Figure 7)
or in two lakes ~65 km west of Yanacocha [Beal et al., 2013]. The spatially inconsistent signal of Hg uxes in
this region suggests that mining dust continued to contribute signicant amounts of Hg to certain lakes and
that any increase in Hg deposition due to anthropogenic Hg0 emissions was relatively negligible during the
preindustrial period. A more localized distribution of preindustrial Hg emissions is consistent with new
chemical modeling by Guerrero [2012] that shows solid calomel (Hg2Cl2) comprised up to 90% of Hg losses
from Ag rening in the Hispanic New World. Post-industrial increases in Hg deposition in the Yanacocha
record were likely caused by global Hg0 emissions.
5.4.3. Modern Flux Ratio
The extent of anthropogenic modication to natural Hg cycling is typically represented by an Hg ux ratio,
which is the ratio of recent Hg uxes to background uxes that occurred at some earlier time (i.e., from A.D.
1800 to 1850 in most sediment records). Table 1 lists mean Hg concentrations and uxes in the Yanacocha
record for key time periods during the Holocene, weighted on the length of time each sample represents.
Because of the evidence for signicant pre-A.D. 1850 anthropogenic deposition in the Yanacocha record,
natural Hg uxes are likely only represented prior to 3.5 ka in this record. Whereas Hg concentrations remain
remarkably constant from 12.3 to 3.5 ka, Hg uxes gradually decrease with increasing age (Table 1). This is
likely an artifact of the age-depth model. We therefore calculate a best approximation of the Hg ux ratio in
the Yanacocha record as time-weighted mean post-A.D. 1980 uxes (4.0 g m 2 a 1) over pre-3.5 ka uxes
(1.4 g m 2 a 1), yielding a ux ratio of 3.0 1.5. This ux ratio, which accounts for total anthropogenic
BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

11

Global Biogeochemical Cycles

10.1002/2013GB004780

modication to the global Hg cycle during the Holocene, is in good agreement with sediment records that
use the period A.D. 1800 to 1850 as background uxes from two other lakes in southeastern Peru (i.e.,
4.0 1.0) [Beal et al., 2013] and from lakes around the world (i.e., on average 3.5) [Biester et al., 2007]. The
discrepancy between our Holocene Hg ux ratio (3.0 1.5) and the modeled 7.5-fold enrichment since
2000 B.C. by Amos et al. [2013] indicates that preindustrial Hg emissions were either not as globally
distributed as assumed in the model or were not as persistent in labile surface reservoirs. Revised
accounting for losses of Hg0 to the atmosphere from preindustrial mining may improve the accuracy of
global Hg models and help reconcile them with sedimentary records.

6. Conclusions
During the preanthropogenic period, atmospheric Hg deposition recorded in Yanacocha was relatively
constant and did not vary with changes in local and regional climate. Holocene volcanic eruptions are
generally not registered in the Hg record despite a number of Plinian eruptions that occurred both globally
and within the Andean CVZ. An early phase of enhanced Hg deposition in Yanacocha began in 1450 B.C.
(3.4 ka) likely due to a combination of nearby mining emissions and volcanic input of Fe and S that led to
framboidal pyrite formation and possible Hg sequestration between ~ 1340 and 1240 B.C. The endurance of
this early phase of enhanced Hg deposition until 500 B.C. is coincident with known pre-Incan cinnabar mining
in Huancavelica. The limited spatial distribution of Hg emissions from Huancavelica and the magnitude of Hg
uxes during this early phase, which are greater than modern uxes, indicate a separate and nearby mining
source of Hg to Yanacocha, likely from within the Cordillera Vilcanota. Increased concentrations of Hg and
Pbext from ~ A.D. 1480 to 1640 suggest sources of Hg to the lake rst from Incan cinnabar mining and then
from colonial Hg production and Ag rening. The agreement of the Holocene ux ratio determined from the
Yanacocha record with ux ratios determined from post-industrial lake sediment records suggests that
preindustrial Hg emissions either were not well distributed globally or did not have a long-lasting impact on
the global atmospheric Hg burden.
Acknowledgments
This research was supported by NSF
Awards EAR-1003460 to Kelly and EAR1003072 to Lowell, and a LacCore visiting
graduate student award to Beal. We
thank Colby Smith, Hannah Baranes,
Yves and Elena Chemin, and the
Crispin Family for eld work and
logistics; Amy Myrbo, Devon Renock,
and Jenny Howley for lab assistance;
and Colin Cooke and David Pyle for
providing constructive reviews of this
manuscript. The primary data for this
paper can be accessed for free in the
supporting information.

BEAL ET AL.

References
Amos, H. M., D. J. Jacob, D. G. Streets, and E. M. Sunderland (2013), Legacy impacts of all-time anthropogenic emissions on the global mercury
cycle, Global Biogeochem. Cycles, 27, 410421, doi:10.1002/gbc.20040.
Baker, P. A., G. O. Seltzer, S. C. Fritz, R. B. Dunbar, M. J. Grove, P. M. Tapia, S. L. Cross, H. D. Rowe, and J. P. Broda (2001), The history of South
American tropical precipitation for the past 25,000 years, Science, 291(5504), 640643, doi:10.1126/science.291.5504.640.
Beal, S. A., B. P. Jackson, M. A. Kelly, J. S. Stroup, and J. D. Landis (2013), Effects of historical and modern mining on mercury deposition in
southeastern Peru, Environ. Sci. Technol., 47(22), 12,71512,720, doi:10.1021/es402317x.
Biester, H., R. Bindler, A. Martinez-Cortizas, and D. R. Engstrom (2007), Modeling the past atmospheric deposition of mercury using natural
archives, Environ. Sci. Technol., 41(14), 48514860, doi:10.1021/es0704232.
Bird, B. W., M. B. Abbott, D. T. Rodbell, and M. Vuille (2011), Holocene tropical South American hydroclimate revealed from a decadally
18
resolved lake sediment O record, Earth Planet. Sci. Lett., 310(34), 192202, doi:10.1016/j.epsl.2011.08.040.
Blaauw, M., and J. A. Christen (2011), Flexible paleoclimate age-depth models using an autoregressive gamma process, Bayesian Anal., 6(3),
457474.
Buffen, A. M., L. G. Thompson, E. Mosley-Thompson, and K. I. Huh (2009), Recently exposed vegetation reveals Holocene changes in the
extent of the Quelccaya Ice Cap, Peru, Quat. Res., 72(2), 157163, doi:10.1016/j.yqres.2009.02.007.
Bush, M. B., M. R. Silman, M. B. de Toledo, C. Listopad, W. D. Gosling, C. Williams, P. E. de Oliveira, and C. Krisel (2007), Holocene re and
occupation in Amazonia: Records from two lake districts, Philos. Trans. R. Soc. B Biol. Sci., 362(1478), 209218, doi:10.1098/rstb.2006.1980.
Cannon, W. F., W. E. Dean, and J. H. Bullock (2003), Effects of Holocene climate change on mercury deposition in Elk Lake, Minnesota: The
importance of eolian transport in the mercury cycle, Geology, 31(2), 187190, doi:10.1130/0091-7613(2003)031<0187:EOHCCO>2.0.CO;2.
Cooke, C. A., P. H. Balcom, H. Biester, and A. P. Wolfe (2009), Over three millennia of mercury pollution in the Peruvian Andes, Proc. Natl. Acad.
Sci. U. S. A., 106(22), 88308834, doi:10.1073/pnas.0900517106.
Cooke, C. A., A. P. Wolfe, N. Michelutti, P. H. Balcom, and J. P. Briner (2012), A Holocene perspective on algal mercury scavenging to sediments
of an arctic lake, Environ. Sci. Technol., doi:10.1021/es3003124.
Cooke, C. A., H. Hintelmann, J. J. Ague, R. Burger, H. Biester, J. P. Sachs, and D. R. Engstrom (2013), Use and legacy of mercury in the Andes,
Environ. Sci. Technol., 47(9), 41814188, doi:10.1021/es3048027.
Croudace, I. W., A. Rindby, and R. G. Rothwell (2006), ITRAX: Description and evaluation of a new multi-function X-ray core scanner, in New
Techniques in Sediment Core Analysis, edited by R. G. Rothwell, pp. 5163, Geological Society of London, London.
Elbaz-Poulichet, F., L. Dezileau, R. Freydier, D. Cossa, and P. Sabatier (2011), A 3500-year record of Hg and Pb contamination in a Mediterranean
sedimentary archive (The Pierre Blanche Lagoon, France), Environ. Sci. Technol., 45(20), 86428647, doi:10.1021/es2004599.
Fitzgerald, W. F., and C. H. Lamborg (2007), Geochemistry of mercury in the environment, in Treatise on Geochemistry, edited by H. D. Holland
and K. K. Turekian, pp. 147, Pergamon, Oxford, U. K.
Friedli, H. R., L. F. Radke, J. Y. Lu, C. M. Banic, W. R. Leaitch, and J. I. MacPherson (2003), Mercury emissions from burning of biomass from temperate North American forests: Laboratory and airborne measurements, Atmos. Environ., 37(2), 253267, doi:10.1016/S1352-2310(02)00819-1.
Garreaud, R., M. Vuille, and A. C. Clement (2003), The climate of the Altiplano: Observed current conditions and mechanisms of past changes,
Palaeogeogr. Palaeoclimatol. Palaeoecol., 194(13), 522, doi:10.1016/S0031-0182(03)00269-4.

2014. American Geophysical Union. All Rights Reserved.

12

Global Biogeochemical Cycles

10.1002/2013GB004780

Guerrero, S. (2012), Chemistry as a tool for historical research: Identifying paths of historical mercury pollution in the hispanic new world, Bull.
Hist. Chem., 37(2), 6170.
Gustin, M. S., S. E. Lindberg, K. Austin, M. Coolbaugh, A. Vette, and H. Zhang (2000), Assessing the contribution of natural sources to regional
atmospheric mercury budgets, Sci. Total Environ., 259(13), 6171, doi:10.1016/S0048-9697(00)00556-8.
Harris, R. C., et al. (2007), Whole-ecosystem study shows rapid sh-mercury response to changes in mercury deposition, Proc. Natl. Acad. Sci.
U. S. A., 104(42), 16,58616,591, doi:10.1073/pnas.0704186104.
Hermanns, Y.-M., and H. Biester (2013), A 17,300-year record of mercury accumulation in a pristine lake in southern Chile, J. Paleolimnol.,
49(4), 547561, doi:10.1007/s10933-012-9668-4.
Jacobson, G. L., S. A. Norton, E. C. Grimm, and T. Edgar (2012), Changing climate and sea level alter Hg mobility at Lake Tulane, Florida, U.S,
Environ. Sci. Technol., doi:10.1021/es302138n.
Kanner, L. C., S. J. Burns, H. Cheng, R. L. Edwards, and M. Vuille (2013), High-resolution variability of the South American summer monsoon
over the last seven millennia: Insights from a speleothem record from the central Peruvian Andes, Quat. Sci. Rev., 75(0), 110, doi:10.1016/j.
quascirev.2013.05.008.
Kellerhals, T., L. Tobler, S. Brtsch, M. Sigl, L. Wacker, H. W. Gggeler, and M. Schwikowski (2010), Thallium as a tracer for preindustrial volcanic
eruptions in an ice core record from Illimani, Bolivia, Environ. Sci. Technol., 44(3), 888893, doi:10.1021/es902492n.
Kelly, M. A., T. V. Lowell, P. J. Applegate, C. A. Smith, F. M. Phillips, and A. M. Hudson (2012), Late glacial uctuations of Quelccaya Ice Cap,
southeastern Peru, Geology, 40(11), 991994, doi:10.1130/G33430.1.
Lamborg, C. H., W. F. Fitzgerald, A. W. H. Damman, J. M. Benoit, P. H. Balcom, and D. R. Engstrom (2002), Modern and historic atmospheric
mercury uxes in both hemispheres: Global and regional mercury cycling implications, Global Biogeochem. Cycles, 16(4), 1104,
doi:10.1029/2001GB001847.
Lamborg, C. H., K. L. Von Damm, W. F. Fitzgerald, C. R. Hammerschmidt, and R. Zierenberg (2006), Mercury and monomethylmercury in uids
from Sea Cliff submarine hydrothermal eld, Gorda Ridge, Geophys. Res. Lett., 33, L17606, doi:10.1029/2006GL026321.
Large, D. J., N. J. Fortey, A. E. Milodowski, A. G. Christy, and J. Dodd (2001), Petrographic observations of iron, copper, and zinc zuldes in
freshwater canal sediment, J. Sediment. Res., 71(1), 6169, doi:10.1306/052600710061.
Mamani, M., A. Tassara, and G. Wrner (2008), Composition and structural control of crustal domains in the central Andes, Geochem. Geophys.
Geosyst., 9, Q03006, doi:10.1029/2007GC001925.
Martnez-Cortizas, A., X. Pontevedra-Pombal, E. Garca-Rodeja, J. C. Nvoa-Muoz, and W. Shotyk (1999), Mercury in a Spanish peat bog:
Archive of climate change and atmospheric metal deposition, Science, 284(5416), 939942, doi:10.1126/science.284.5416.939.
Mason, R. P., W. F. Fitzgerald, and F. M. M. Morel (1994), The biogeochemical cycling of elemental mercury: Anthropogenic inuences,
Geochim. Cosmochim. Acta, 58(15), 31913198, doi:10.1016/0016-7037(94)90046-9.
McCormac, F., A. Hogg, P. Blackwell, C. Buck, T. Higham, and P. Reimer (2004), SHCal04 Southern Hemisphere calibration, 011.0 cal kyr BP,
Radiocarbon, 46(3), 10871092.
Mortlock, R. A., and P. N. Froelich (1989), A simple method for the rapid determination of biogenic opal in pelagic marine sediments, Deep
Sea Res. Part A Oceanogr. Res. Pap., 36(9), 14151426, doi:10.1016/0198-0149(89)90092-7.
Neumann, T., F. Scholz, U. Kramar, M. Ostermaier, N. Rausch, and Z. Berner (2013), Arsenic in framboidal pyrite from recent sediments of a
shallow water lagoon of the Baltic Sea, Sedimentology, 60(6), 13891404, doi:10.1111/sed.12031.
Nriagu, J. O. (1979), Production and uses of mercury, in The Biogeochemistry of Mercury in the Environment, pp. 2340, Elsevier/North-Holland
BiomedicalPress, Amsterdam, Netherlands.
Nriagu, J. O. (1993), Legacy of mercury pollution, Nature, 363(6430), 589589.
Nriagu, J., and C. Becker (2003), Volcanic emissions of mercury to the atmosphere: Global and regional inventories, Sci. Total Environ.,
304(13), 312, doi:10.1016/S0048-9697(02)00552-1.
Outridge, H. S., H. Stern, and F. Goodarzi (2007), Evidence for control of mercury accumulation rates in Canadian High Arctic lake sediments
by variations of aquatic primary productivity, Environ. Sci. Technol., 41(15), 52595265, doi:10.1021/es070408x.
Petersen, U. (1965), Regional geology and major ore deposits of central Peru, Econ. Geol., 60(3), 407476, doi:10.2113/gsecongeo.60.3.407.
Plathe, K. L., F. von der Kammer, M. Hassellv, J. N. Moore, M. Murayama, T. Hofmann, and M. F. Hochella Jr. (2013), The role of nanominerals
and mineral nanoparticles in the transport of toxic trace metals: Field-ow fractionation and analytical TEM analyses after nanoparticle
isolation and density separation, Geochim. Cosmochim. Acta, 102(0), 213225, doi:10.1016/j.gca.2012.10.029.
Prestbo, E. M., and D. A. Gay (2009), Wet deposition of mercury in the U.S. and Canada, 19962005: Results and analysis of the NADP mercury
deposition network (MDN), Atmos. Environ., 43(27), 42234233, doi:10.1016/j.atmosenv.2009.05.028.
Pyle, D. M., and T. A. Mather (2003), The importance of volcanic emissions for the global atmospheric mercury cycle, Atmos. Environ., 37(36),
51155124, doi:10.1016/j.atmosenv.2003.07.011.
Reimer, P., et al. (2011), IntCal09 and Marine09 radiocarbon age calibration curves, 050,000 Years cal BP, Radiocarbon, 51(4), 11111150.
Robins, N. A., and N. A. Hagan (2012), Mercury production and use in colonial Andean silver production: Emissions and health implications,
Environ. Health Perspect., 120(5), 627631.
Robins, N. A., et al. (2012), Estimations of historical atmospheric mercury concentrations from mercury rening and present-day soil concentrations of total mercury in Huancavelica, Peru, Sci. Total Environ., 426(0), 146154, doi:10.1016/j.scitotenv.2012.03.082.
Roos-Barraclough, F., A. Martinez-Cortizas, E. Garca-Rodeja, and W. Shotyk (2002), A 14 500 year record of the accumulation of atmospheric
mercury in peat: Volcanic signals, anthropogenic inuences and a correlation to bromine accumulation, Earth Planet. Sci. Lett., 202(2),
435451, doi:10.1016/S0012-821X(02)00805-1.
Rowe, H. D., R. B. Dunbar, D. A. Mucciarone, and G. O. Seltzer (2002), Insolation, moisture balance and climate change on the South American
Altiplano since the Last Glacial Maximum, Clim. Change, 52(12), 175.
Rydberg, J., V. Glman, I. Renberg, R. Bindler, L. Lambertsson, and A. Martnez-Cortizas (2008), Assessing the stability of mercury and
methylmercury in a varved lake sediment deposit, Environ. Sci. Technol., 42(12), 43914396, doi:10.1021/es7031955.
Rydberg, J., J. Klaminder, P. Rosn, and R. Bindler (2010), Climate driven release of carbon and mercury from permafrost mires increases
mercury loading to sub-arctic lakes, Sci. Total Environ., 408(20), 47784783, doi:10.1016/j.scitotenv.2010.06.056.
Schoonen, M. A. A. (2004), Mechanisms of sedimentary pyrite formation, Geol. Soc. Am. Spec. Pap., 379, 117134, doi:10.1130/0-81372379-5.117.
Schuster, P. F., D. P. Krabbenhoft, D. L. Naftz, L. D. Cecil, M. L. Olson, J. F. Dewild, D. D. Susong, J. R. Green, and M. L. Abbott (2002), Atmospheric
mercury deposition during the last 270 years: A glacial ice core record of natural and anthropogenic sources, Environ. Sci. Technol., 36(11),
23032310, doi:10.1021/es0157503.
Siebert, L., and S. Simkin (2002), Volcanoes of the world: An illustrated catalogue of Holocene volcanoes and their eruptions. Smithsonian
Institution, Global Volcanism Program Digital Information Series, GVP-3, 2002, Washington, D. C.

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

13

Global Biogeochemical Cycles

10.1002/2013GB004780

Streets, D. G., M. K. Devane, Z. Lu, T. C. Bond, E. M. Sunderland, and D. J. Jacob (2011), All-time releases of mercury to the atmosphere from
human activities, Environ. Sci. Technol., 45(24), 10,48510,491, doi:10.1021/es202765m.
Strode, S. A., L. Jaegl, N. E. Selin, D. J. Jacob, R. J. Park, R. M. Yantosca, R. P. Mason, and F. Slemr (2007), Air-sea exchange in the global mercury
cycle, Global Biogeochem. Cycles, 21, GB1017, doi:10.1029/2006GB002766.
Stroup, J. S., M. A. Kelly, T. V. Lowell, P. J. Applegate, and J. A. Howley (2014), Late Holocene uctuations of Qori Kalis outlet glacier Quelccaya
Ice Cap, Peruvian Andes, Geology, doi:10.1130/G35245.1.
Suits, N. S., and R. T. Wilkin (1998), Pyrite formation in the water column and sediments of a meromictic lake, Geology, 26(12), 10991102,
doi:10.1130/0091-7613(1998)026<1099:PFITWC>2.3.CO;2.
Thevenon, F., S. Gudron, M. Chiaradia, J.-L. Loizeau, and J. Pot (2011), (Pre-) historic changes in natural and anthropogenic heavy metals
deposition inferred from two contrasting Swiss Alpine lakes, Quat. Sci. Rev., 30(12), 224233.
Thompson, L. G., E. Mosley-Thompson, W. Dansgaard, and P. M. Grootes (1986), The Little Ice Age as recorded in the stratigraphy of the
tropical Quelccaya Ice Cap, Science, 234(4774), 361364.
Thompson, L. G., E. Mosley-Thompson, H. Brecher, M. Davis, B. Len, D. Les, P.-N. Lin, T. Mashiotta, and K. Mountain (2006), Abrupt tropical
climate change: Past and present, Proc. Natl. Acad. Sci. U. S. A., 103(28), 10,53610,543, doi:10.1073/pnas.0603900103.
Thompson, L. G., E. Mosley-Thompson, M. E. Davis, V. S. Zagorodnov, I. M. Howat, V. N. Mikhalenko, and P.-N. Lin (2013), Annually resolved ice
core records of tropical climate variability over the past ~1800 years, Science, 340(6135), 945950, doi:10.1126/science.1234210.
Wilkin, R. T., H. L. Barnes, and S. L. Brantley (1996), The size distribution of framboidal pyrite in modern sediments: An indicator of redox
conditions, Geochim. Cosmochim. Acta, 60(20), 38973912, doi:10.1016/0016-7037(96)00209-8.

BEAL ET AL.

2014. American Geophysical Union. All Rights Reserved.

14

You might also like