Chile reveals persistent PDO (Pacific Decadal Oscillation) impact

Chile reveals persistent PDO (Pacific Decadal Oscillation) impact

Journal Pre-proof A new 20th century lake sedimentary record from the Atacama Desert/Chile reveals persistent PDO (Pacific Decadal Oscillation) impact...

2MB Sizes 0 Downloads 33 Views

Journal Pre-proof A new 20th century lake sedimentary record from the Atacama Desert/Chile reveals persistent PDO (Pacific Decadal Oscillation) impact Mauricio Cerda, Heitor Evangelista, Jorge Valdés, Abdelfettah Siffedine, Hugues Boucher, Juliana Nogueira, Aguinaldo Nepomuceno, Luc Ortlieb PII:

S0895-9811(19)30361-X

DOI:

https://doi.org/10.1016/j.jsames.2019.102302

Reference:

SAMES 102302

To appear in:

Journal of South American Earth Sciences

Received Date: 14 July 2019 Revised Date:

1 August 2019

Accepted Date: 1 August 2019

Please cite this article as: Cerda, M., Evangelista, H., Valdés, J., Siffedine, A., Boucher, H., Nogueira, J., Nepomuceno, A., Ortlieb, L., A new 20th century lake sedimentary record from the Atacama Desert/ Chile reveals persistent PDO (Pacific Decadal Oscillation) impact, Journal of South American Earth Sciences (2019), doi: https://doi.org/10.1016/j.jsames.2019.102302. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier Ltd.

1 2

A new 20th century lake sedimentary record from the Atacama Desert/Chile reveals persistent PDO (Pacific Decadal Oscillation) impact

3 4 5

MAURICIO CERDA1,2,3*, HEITOR EVANGELISTA4,7, JORGE VALDÉS3,5, ABDELFETTAH SIFFEDINE5,6, HUGUES BOUCHER6, JULIANA NOGUEIRA4,5,7, AGUINALDO NEPOMUCENO1& LUC ORTLIEB6

6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

1

21

Abstract

22 23 24 25 26 27 28 29 30 31 32 33 34 35

The Atacama Desert lakes located at high altitudes with very low anthropogenic impacts preserve great potential to recover regional climate changes and environmental processes at a decadal to centennial time scales. Herein, we present a high-resolution analysis of the stratigraphy and metal composition of Inka Coya Lake sediment core, in order to clarify the main forcings acting on sedimentation rates and both eolian and precipitation patterns at Atacama Desert environment. The results suggest that terrigenous inputs can be mainly related to the PDO (Pacific Decadal Oscillation), while laminations preserve a direct association with El Nino-Southern Oscillation – ENSO (4.2 to 5.3 yr. cyclicity). Additionally, the Cu mining activities in Chile are recorded in the lake. The sedimentation rates and precipitation patterns were associated with ENSO, highlighting the relation between La Niña and wetter periods at Atacama. Finally, high-resolution analyses of the geochemical composition provided additional criteria to identify the erosion processes at the basin, suggesting different sources of sediments, which fluxes into the lake covariated with PDO at a decadal scale, especially during the 20th century.

Programa de Biologia Marinha e Ambientes Costeiros Universidade Federal Fluminense (PBMAC-UFF) Outeiro de São João Batista s/n, Centro Niterói Brazil. 2 Centro de Investigación e Innovación para el Cambio Climático (CiiCC), Facultad de Ciencias Universidad Santo Tomás Santiago Chile 3 Laboratorio de Sedimentología y Paleoambientes, Instituto Alexander von Humboldt. Facultad de Ciencias del mar y de recursos biológicos, Universidad de Antofagasta, Casilla 170, Antofagasta, Chile 4 LARAMG/IBRAG/UERJ, Pavilhão Haroldo L. da Cunha, Subsolo, Rua São Francisco Xavier 524, Maracanã, Rio de Janeiro, 20550-013, Brazil 5 LMI PALEOTRACES, IRD-UFF-UANTOF-UPCH. 6 LOCEAN - IPSL UMR 7159, IRD-Sorbonne Universités (Université P. et M. Curie, Paris 06)CNRS/UPMC/IRD, IRD France-Nord 32, av. Henri Varagnat,FR-93143 BONDY cedex 7 Programa de Pós-graduação em Geoquímica, Universidade Federal Fluminense/Instituto de Química/Departamento de Geoquímica/Outeiro de São João Batista s/n, Centro, Niterói, Brazil.

36 37

∗ Corresponding author

38

E-mail addresses: *[email protected]

39 40 41

Keywords: Atacama, PDO, lake sedimentology, elemental composition, Copper record

42

1.

Introduction

43

The Andes plateau at Northern Chile created by the subduction of the Nazca Plate beneath

44

the South America Plate houses a series of lakes (Risacher et al. 2003). They are believed

45

to provide conditions for reconstructing the regional past climatic changes and the past

46

aridity conditions (Anderson et al., 2011; Ritter et al., 2018). The aridity of the Atacama

47

desert may have started in the Eocene (from ~ 55 to 36 million years) and evolved into the

48

current aridity since the mid-Miocene (from 23 to 5.3 million years) (Clarke, 2006). The

49

extreme aridity observed today is due to a combination of factors that includes the influence

50

of the Subtropical High-pressure belt, the Ecuador-advection, the Humboldt coastal cold

51

marine current, and the orographic aspects of the Andes Cordillera around. According to

52

Schwalb (2003), and considering the above aspects, the Altiplano is a region very sensitive

53

to changes in the moisture advection, evaporation and precipitation dynamics. Precipitation

54

at Atacama Desert is mainly controlled by two climatic mechanisms: the northern monsoon

55

(Bolivian Winter) and the southern frontal systems. Precipitation is minimum within the

56

area that corresponds to the boundary between both mechanisms, 18-28°S, where their

57

effects are minimized. It is believed that the El Niño–Southern Oscillation (ENSO), may

58

regionally influence the surface winds as a result of the sea surface temperatures over the

59

Pacific Ocean. Its two related phases (El Niño and La Niña) control the longitudinal

60

intensity of the Walker circulation that ultimately may bring more or less ocean influence

61

towards the inland. According to literature, ENSO events have contrasted effects on

62

precipitation regimes at the Andean region with intense summer rain at the Western side

63

and droughts at the Eastern side (Lavado-Casimiro and Espinoza, 2014).

64

Sediment cores from lacustrine environments are widely used to reconstruct temporal

65

variability of natural and anthropogenic processes (e.g.: von Gunten et al., 2009). In this

66

context, many authors use joint analysis of sediment geochemical composition and their

67

laminated structures to provide insights into runoff processes associated with precipitation

68

and dust deposition of chemical species. For Atacama Desert, although the impact of ENSO

69

on the regional climate is well documented, we explore its impact over the sedimentary

70

record measured from dated lacustrine sediment deposits, sedimentary stratigraphy and

71

metal contents using high-resolution XRF analysis. As a complementary aspect, we have

72

also investigated if an imprint of the major regional anthropogenic activity, represented by

73

the Cu mining and tailing does exist. Over the last hundred years, mining and related

74

activities caused a significant reduction in the major watersheds (Romero et al., 2012) and

75

several pieces of evidence indicate that their resuspended materials were potential sources

76

of contamination (Gregor and Gummer, 1989; Rapaport et al., 1985; Swackhamer and

77

Armstrong, 1986).

78

Herein we present high-resolution elemental composition variability from a sediment core

79

retrieved at the high-altitude Inka Coya Lake, Northern Atacama Desert. This lake is

80

characterized by having a relatively low human direct impact and no significant land-use

81

change around it during the last century. The analysis on its 20th century lacustrine

82

sedimentary record sector allowed us to investigate the impact of the PDO (Pacific Decadal

83

Oscillation) over its geochemical and physical characteristics.

84 85

2.

Methods

86

2.1. Study site

87

The Inka Coya Lake (San Francisco de Chiu-Chiu/Antofagasta; 22° 20.300’ S; 068°35.981’

88

W) is located in the Northern Atacama Desert region at an altitude of 2516.4 m a.s.l (Fig.1).

89

It is approximately 48 km away from the city of Calama. It has an asymmetric oval shape

90

and an area of 1.52 km2. The Inka Coya Lake has an average volume of 1.6x106 m3 and is

91

surrounded by a varied xerophytic fauna (Gantz et al., 2009). The local terrain is stony like,

92

with salt lakes (salares), sand and felsic lava. The extreme aridity conditions of the region

93

are mostly due to the cool north-flowing Humboldt ocean current, and the presence of the

94

strong Pacific anticyclone. The lake is situated on the limit among the central valley and the

95

cordillera zone, between two mountain ranges - the Andes and the Chilean Coastal Zone.

96

This orography protects the site against intense moisture advection from either the Pacific

97

or the Amazon basin, contributing to the maintenance of hyper aridity conditions over the

98

Atacama Desert. Geological evidence suggests that the study site is composed mostly of

99

moderately consolidated sedimentary rock from Pliocene-Pleistocene, while its

100

surroundings can be related to old sediments from the late Miocene-Pliocene transition

101

period and sediments as recent as the Pleistocene-Holocene (Fig. 2) (Jordan et al., 2015).

102

103 104 105 106 107 108 109 110 111 112 113 114 115

Figure 1 – The study site: (a) total view of the Inka Coya Lake (b) bathymetry of the Inka Coya Lake showing a central depression and the sediment core sampling point (T1= XLCHT1); (c) Inka Coya Lake location in the context of South America and Northern Atacama orography.

116 117 118 119 120

Figure 2 – Map of the Inka-Coya site main geological facies. Based on Chilean geological map from the National Geology and Mining Service of Chile.

121 122

2.2. Regional climate characterization

123

The small annual precipitation amount that falls on the Atacama is a result of the northeast

124

summer monsoon, southwest frontal precipitation, fog deposition below 800 m a.s.l and

125

also occasional winter rains at coastal areas below 1,000 m. Occasional winter rains (May-

126

October) on Andean slopes South of 25ºS, and summer storms (November-March) that

127

cross the Altiplano, spill over the Andes, and rain out on the Pacific slope north of 25°S

128

(Latorre et al., 2003). According to Köppen climate classification, the predominant climate

129

domain at Chiu-Chiu region is “BWk” – cold desert climate – characterized by dry

130

summers with marked seasonal thermal amplitude. During winter, air temperature reaches

131

an average of 22 ºC, during the day, and 4ºC by night. Clear sky condition occurs almost

132

through all the year, due to a high-pressure system that is generally located over the region

133

20-30°S during the summer and early fall. This system acts as a barrier associated with the

134

low-pressure systems southwards. During summer, temperature reaches 27 ºC, during the

135

day, and 16 ºC by night (with some occasional rain).

136 137

2.3. Sediment coring and samples preparation

138

Two sediment cores, of approximately 50 cm in length, were retrieved from the maximum

139

lake depth (22° 20.300’ S; 068°35.981’ W) in July 2013 (Fig 1b) using a 5 cm diameter

140

gravity core (Wildco, K-B Corer 20’’). Geochemical analysis and dating were employed to

141

one of them, measuring 47 cm, identified as X-LCHT1 (Fig 4a-b). The sediment core was

142

transported to the laboratory in polyethylene tube, longitudinally sectioned and stored at 4

143

°C until analysis. One half was kept closed, without physical or chemical processing, for

144

geochemical analysis at the Institut de recherche pour le développement (IRD)-France

145

laboratory; and the other half was sliced into thin sections of approximately 0.5 cm. These

146

slices were dried, powdered and sealed in plastic flasks for radiometric analysis and dating

147

(non-destructive analyses) at LARAMG Laboratory/Rio de Janeiro State University-Brazil.

148

After that, the granulometry aspect of these samples was analyzed using a standard sieve

149

and Cilas®1180 Laser Particle Analyzer (Ziervogel and Bohling, 2003) at the Coastal

150

Environments and Marine Biology Program Laboratory (PBMAC-UFF). For grain size

151

analysis, samples were dried, homogenized, and carbonates and organic matter were

152

extracted by adding aliquotes of HClconc and H2O2 solution (30 %), respectively, until stop

153

reactions (Smith, 2000). Dispersion was performed with Sodium Hexametaphosphate (40 g

154

L-1) followed by sonication (35 W for 2 min). Cilas®’s analysis was performed on 0.2 g of

155

samples and particle size range spanned from 0.04 to 2500 µm. For the other half of the

156

core, we first performed the X-ray (X-LCHT1) image acquisition, with a voltage of 50 kV

157

(Axelsson, 1983). At the Central medical facility in Antofagasta-Chile (Fig 4b), the image

158

allowed the reconnaissance of internal stratigraphy of the core and processes related to

159

bioturbation. Over the X-ray image of the sediment core we have drawn three continuous

160

longitudinal parallel lines from the top to the bottom of the image to obtain the average

161

grey scale variability along the depth of the core using the Image-J software (Fig 4c). This

162

spatial database was converted to temporal scale based on the sedimentation rate obtained

163

by the 210Pb dating method (item 2.5). From these data we have applied FFT (Fast Fourier

164

Transformation) to search for core lamination structure periodicities (24 – 47 cm).

165 166

167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192

Figure 3- The meteorological database for Calama Station (22° 29.734’ S; 68°54.350 W; https://climatologia.meteochile.gob.cl); (a) Seasonal precipitation at Chiu-Chiu; (b) local wind direction; (c) MEI index from www.esrl.noaa.gov/psd/enso/mei/ spanning the period 1976-2010. MEI index is the multivariated ENSO Index, defined based on six main observed variables over the tropical Pacific: sea-level pressure, zonal wind (U). Meridional wind (V), sea surface temperature, surface air temperature and total cloudiness fraction of the sky (d) wind intensity variability for the three main wind component directions. Arrows point more intense NIÑO events in the above period.

193

2.4. Elemental composition analysis

194

Insights of the natural environmental changes and anthropogenic impacts recorded in the

195

sediment layers of Inka Coya lake were investigated by the high-resolution X-ray

196

fluorescence (XRF) technique (ARTAX® device from the Bruker™). This scanning

197

technology allows high-resolution, non-destructive elemental analysis at sub-millimeter

198

scales (Croudace et al., 2006; Francus et al., 2009). Elemental compositions analysis here

199

corresponded to the dated section of the core, that is, the top 24 cm from the total 47 cm-

200

long sequence. The XRF analysis performed at the ALYSES platform (IRD/Bondy/France)

201

provides geochemical area mapping of surfaces up to 100cm x 5cm. This technique has the

202

potential to correspond to both short-duration events and the long-term variability.

203

Measurements were made using a Chromium (Cr) X-Ray tube with the electrical

204

parameters: 35 kV and 1142 µA. A spatial resolution of 1mm was used for both X (core

205

length) and Y (core width) axes, using a counting time of 30 s per measured point.

206

Elemental area mapping results of nearly 50cm x 4cm were then cumulated along the Y-

207

axis in order to obtain line scans along the X-axis. This technique was used to characterize

208

the geochemical variability of main terrigenous elements as Titanium (Ti), Iron (Fe),

209

Potassium (K), Rubidium (Rb), Silicon (Si), Zinc (Zn), Zirconium (Zr), Calcium (Ca),

210

Phosphorus (P), Ytterbium (Yb), Arsenic (As), Strontium (Sr), Aluminum (Al), Manganese

211

(Mn), Sulfur (S), Chlorine (Cl) and Cooper (Cu). XRF analysis values are expressed in

212

counts that refer to the specific peak area of each element.

213

2.5. Dating of the sediment core

214

Dating method employed in this work used the radionuclides 210Pb and 226Ra from the 238U

215

natural series. The core from Inka Coya Lake was analyzed radiometrically by a high-

216

resolution gamma spectrometry making use of an extended energy range co-axial

217

hyperpure germanium (HPGe) detector (model GX5021 – Canberra). Equipment relative

218

efficiency is 50% and resolution of 2.1 keV (FWHM) at the 60Co (1.33 MeV) energy peak.

219

This detector was installed inside a very low background lead shielding of an average

220

thickness of 15 cm, internally coated with pure Cu. Detector efficiency curve for the

221

sediment samples was performed using a liquid solution containing a cocktail of

222

radionuclides - NIST (serial number HV951). The cocktail used in this study included the

223

radionuclides 133Ba, 57Co, 139Ce, 85Sr, 137Cs, 54Mn, 88Y, and 65Zn, which enclose the energy

224

range of interest. The total counting time for each sample was 24 h.

225

The geochronological model used in this work was the Constant Rate of Supply (CRS). In

226

CRS model, the initial 210Pb concentration in the sediment is constant and the influx rate of

227

sediment is variable (Goldberg 1963; Appleby and Oldfield, 1978). The model itself

228

considers that no post-depositional mixing occurs. Therefore, bioturbation process is a

229

source of error in the sedimentation rates estimates through the time. Another assumption is

230

that the transport parameterization is independent of the sedimentation rate. Chronology is

231

based on the unsupported fraction of the total 210Pb in each strata as 210Pbunsuported = 210Pbtotal

232



233

but also from dust storms (in the case of a desert site) in which it is assumed to be in

234

equilibrium with 226Ra. In this model, the mass sediment accumulation rates are a measure

235

of sedimentation that changes with depth. Sedimentation rates are determined by plotting

236

the log of the excess

237

calculating the sedimentation rates over intervals of constant slope. According to Appleby

238

and Oldfieldz (1983), the age of sediment layers can be obtained as equation 1.

210

Pbsupported. Where

210

Pbsupported is delivered from lake borders due to erosion from rain,

210

Pb activities against cumulative dry sediment weights and

239

t= 240

1  AO ln  λ  A X

   (1)

241 242

where A0 is the total unsupported 210Pb activity in the sediment column and Ax is the total

243

unsupported 210Pb activity in the sediment column beneath depth x and λ is the radioactive

244

decay constant of 210Pb, this is, 0.03114 year-1.

245

2.6 Statistical analysis

246

We have established 95 and 99 confidence intervals for FFT-Fast Fourier Transform to

247

detect statistically significant periodicities at visible lamination structures (rhythmites) in

248

the sediment core X-ray image. Power function was obtained with PAST Software.

249

Principal Components Analysis (PCA) was employed as an attempt to infer factors that

250

may control the core geochemistry variability. To improve the detection of non-linear

251

relationships, cubic-root or log (x + 1) transformations were applied to the data (Legendre

252

and Legendre, 1998). The total number of samples was 44, and 20 variables were included

253

in the analysis (Ti, Fe, K, Rb, Si, Zn, Zr, Ca, P, Yb, As, Sr, Al, Mn, Cu, Clay, Silte, Sand

254

and particle diameter). The results of the PCA are presented in the form of a plot diagram.

255

The quantitative environmental variables are indicated by arrows. This analysis was

256

performed by PAST software Hammer Ø (Hammer et al., 2001).

257 258

3.

259

Results of the time series analysis of wind speed direction for 43 years (Calama

260

meterological station) (Fig 3b) revealed the influence of a large-scale atmospheric system

261

and local to regional effects in Andean plateau of the Atacama Desert determine the

262

circulation of the predominant SW-W-E winds during the whole year, with stronger

263

intensity during the spring and summer season. As one may observe from Fig 3c, surface

264

winds coming from the Pacific Sector, W-SW, were affected by the two strongest El Niño

265

Events during the 1980s and 1990s decades, while the Eastern component, coming from the

266

Andean Cordillera sector, did the episodes of 82/83 and 97/98 wind intensities changed

267

from a typical values around 6 m/s to 8 m/s, suggesting a higher probability for local dust

268

resuspension during strong ENSO phases. On the other hand, the analysis of the temporal

269

series of precipitation and wind speed direction (Calama meteorological station) revealed

270

three distinct patterns seasonal precipitation: (1) a wet season during May-June-July with

271

median of 34 mm, (2) a prolonged dry season from August to December with median of 2.3

272

mm and; (3) an intermediate season from January to August with median of 9 mm. No rain

273

was observed in this period for November and December (Fig 3a).

274 275 276 277 278 279 280

Results and Discussion

281

3.1. Sediment core stratigraphy, geochronology and relation with ENSO

282

In this study of the stratigraphic record, the X-LCHT1 technique exhibited no obvious signs

283

of physical or biological mixture and some redistribution of 210Pb may have occurred only

284

on the uppermost part of the core as revealed by the radionuclide vertical behavior. 210Pbexc

285

data showed a general tendency to decline with depth (Fig 3d). However, the profile had

286

some fluctuations that superimpose to the exponential trend of decay. The chronology for

287

the Inka-Coya Lake record, by using the

288

sampling) to approximately 1886 AD, which corresponded to the top 24 cm from the 47 cm

289

collected. Our estimate for the mean sedimentation rate in this interval is 0.19 cm/yr.

290

The sedimentation rate along the 20th century increased steadily from the 1970s decade,

291

reaching a maximum at the beginning of the 1990s following a decrease until the year 2000

292

(Fig 4d). In general, the behavior of the sedimentation rate follows approximately the trend

293

of the accumulated Southern Oscillation Index (SOI) within a lagged response. SOI is an

294

index related to the intensity of El Niño (negative SOI) and La Niña (positive SOI) events,

295

here calculated from the year 1876 to 2010. This suggests a potential relationship of ENSO

296

phenomena and the sedimentary regime at the Northern Chilean desert environment, at

297

least over the last approximately 150 years. Like the cumulative SOI, the sedimentation rate

298

was quite stable up to approximately 1945 and then it increases until the 80’s decade, a

299

period dominated by La Niña events, then decreases until the 90’s when El Niño was

300

dominant. Between 23 and 25 cm, we found a layer characterized by a very low

301

sedimentation rate, of blackish texture that marked a transition between a well-defined

302

laminated phase (rythmites) and a non-laminated phase. Herein we will concentrate our

303

efforts on analyzing the well-dated sector correspondent with the non-laminated phase of

304

the core.

305 306 307 308 309

210

Pb dating, spanned a period of 2013 (year of

310

311 312 313 314

Figure 4 - (a) Sediment core of Lake Inka-Coya (b) corresponding X-ray image; (c) profile of gray values; (d) profile of 210Pbexc in the core; and (e) sedimentation rates inferred from CRS model.

315 316

We have investigated the nature of these laminations by the periodicity they occur. For that

317

we employed the greyscale variability along the longitudinal length of the core. From the

318

top to the bottom of the core we have arranged a total of 400 pixels, each one having

319

distinct greyscale intensity. Over the greyscale time series, we have applied an FFT

320

technique to search for periodicities (cyclic patterns) within statistically significant levels of

321

95% and 99%. This allowed obtaining values of significant frequency bands as depicted in

322

Fig 5. Periods were obtained as P=1/f. Periodicities of pixels were converted to thicknesses

323

of the light and dark layers considering that the total of 400 pixels is equivalent to the total

324

core length of 47 cm. Then by using the average sedimentation rate value of 0.19 cm/year

325

we converted to period. From this approach, we obtained two sets of periodicities; (1)

326

4.2yr, 4.7yr and 5.3yr which are near the periodicity band of ENSO (~ 4-7yr) and (2) 8.4yr

327

and 10.5yr that is close to Schwalbe solar cycle of 11 yr. The sunspot cycle has been

328

already recognized in laminated lake sediments from many regions although the scientific

329

climate basis is not completely described (Muñoz et al., 2002). To better understand the

330

impact of solar activity in the desert environment, we compared the sunspot number with

331

precipitation data of the closest meteorological station. Lacustrine varve layers can be of

332

annually resolved or produced in resolutions associated with longer hydrological cycles

333

(Hernández et al., 2010). A varve layer can be produced when particles are washed into the

334

sedimentary basin because the strengthened water flow can carry coarser material, followed

335

by a finer material deposition occurring during the subsequent phase when energy for

336

transporting materials diminishes. It is probable that salts from the Atacama Desert

337

contribute to coagulate the clay into coarse grains. This process may form a rhythmites

338

structure of corse-fine texture, more visible at x-Ray image. From Fig 3a and Fig 5b it is

339

clear that the events of higher precipitation during the austral winter are mostly followed by

340

the driest period at Atacama. Our comparison between solar activity and precipitation

341

shows a coherent variability between these two parameters since the 1960s decade that

342

corresponds to our longer regional database. Smoothed precipitation pattern shown at Fig

343

5b varied at the same phase of solar activity and therefore the decadal varved signal could

344

be related to changes in solar output energy. During the same period, ENSO variability

345

changed at frequencies between 3 and 7 years while the PDO (Pacific Decadal Oscillation)

346

varied in frequencies much higher than the ~a variability (Valdés-Pineda et al., 2018). The

347

total extent on ENSO control upon precipitation patterns hence the hydro-climatology at

348

Atacama Desert is an issue not fully portrayed (Vuille and Keimig, 2004). Nevertheless,

349

early evidence suggests that El Niño may induce droughts in the Altiplano during both

350

summer and winter seasons (Houston, 2006).

351

352 353 354 355 356

Figure 5- (a) Periodicities detected by FFT (Fourier Fast Transformation) based on greyscale series of Inka-Coya Lake sediment core. Confidence curves are depicted; (b) sunspot number (smoothed time series) corresponding to precipitate database.

357 358

3.2. Sediment core elemental composition and grain size variability

359

Total elemental composition of Inka Coya sediment core record is displayed in Fig 6. From

360

all detected elements we identified 3 distinct groups: (a) Ti, Fe, K, Rb, Si, Mn and S named

361

hereafter as Group 1 (G1) (Fig 6 – upper part); (b) P, Sr, Yb, Cl, Ca, As and Al named

362

hereafter as Group 2 (G2) (Fig 6 – lower part); (c) and Cu (whose time series will be

363

presented later in view of the mining process present in the region). G1 group presented

364

decreases in concentrations during the periods 1908-1968 and 1994-2013 while increases

365

occurred during 1886-1908 and 1968-1994. The inverse behavior is reported for the G2

366

group. This suggests two allochthonous terrigenous sources that may control the total input

367

of lithological material into the lake and that may be associated with the light and dark

368

pattern observed. This is especially evident for Ca and Sr, which are commonly associated

369

with authigenic carbonate minerals or biogenic calcium carbonates from super-arid and

370

limestone/carbonate environments. Also for the Atacama Desert, Stivaletta et al. (2012),

371

have reported a common abundance of fossil deposits of halite (NaCl) as part of the

372

composition from the Salar Grande salt. This deposit presents accessory amounts of

373

gypsum (CaSO4. 2H2O) and glauberite (Na2Ca(SO4)2), which explain the association of Cl

374

and Ca in the same elemental group. Elements in G2 suggest a possible composite mixture

375

of both marine spray and eolian salar materials as sulfate salt sources based on Sr and Cl.

376

On the other hand, the G1 group is characterized by typical lithological material that could

377

be provided by erosion from Andes Cordillera.

378 379 380 381 382

Figure 6- Elemental composition along Inca-Coya sediment core as revealed by XRF analysis. Upper part, the set of elements that defined the group 1 (G1) due to their covariability; and lower part, the same for group 2 (G2). Dotted lines separate major changes in geochemistry.

383

The distinction of the aforementioned groups of elements was clarified by the Principal

384

Component Analysis (PCA) (Fig 7), with Component 1 and Component 2, accounting for

385

57.6% and 19.2% of the total variance, respectively. PCA evidenced an association of G1

386

elements (the terrigenous, or crustal component) with the silt (0.05 – 0.002 mm) and clay

387

(< 0.002 mm) content, that represents the finer fraction of the sediment while G2 elements

388

to sand (0.05 – 2.0 mm), the coarse fraction. This means that group G1 represents the fine

389

mineral dust material being transported and deposited into the lacustrine system, probably

390

through eolic erosion. Pre-cordillera and Coastal Cordillera, where Inka-Coya is located,

391

are Pre-Mesozoic basement with scattered metamorphic outcrops of Precambrian age where

392

surface rock geochemistry contains Fe, Mn, and Zn, besides Co, Cr and V (Tapia et al.,

393

2018). These elements reflect the geochemistry of volcanic and intrusive rocks, soils,

394

regoliths and sediments while Fe–Ti oxides depict the importance of the Andean

395

magmatism. Group G2 may represent a multi-elemental source, since As points to the

396

influence of its potential natural sources as volcanic rocks, hot-springs and mineral deposits

397

(Bull et al., 2016). On the other hand, Ca may derive from the abundant carbonates and

398

gypsum (a sulfate mineral composed of calcium sulfate dihydrate), that are major

399

components in marginal mudflat, ephemeral-stream and alluvia-fan of the Salar de Atacama

400

basin (Boschetti et al., 2007). These materials occur associated with the course mode of the

401

sediment core being associated with the sand fraction. The PCA also depicts the Cu as

402

probably derived from an independent emission source. We attribute that to the extensive

403

open-pit copper mining at the vicinity region of Antofagasta. With respect the P, it is

404

associated with deposits of phosphatic guano that typically contain up to 20% of P2O3, with

405

several occurrences at the northern coast of Chile and associated with Ca when occurring at

406

vein phosphate rocks (Tapia et al., 2019).

407 408 409

Figure 7- Components of the Principal Component Analysis (PCA) for sediment core elemental composition and grain size distribution.

410 411

Based on the anti-phase behavior of the two metal sets in Figure 6, we have defined an

412

index of variability for G1 and G2. To perform that, we have standardized each element

413

database using a Gaussian variable [z=(value-mean)/standard deviation] and averaged the

414

z-values of all elements in the same period. Also considering that the core chronology

415

spans nearly a century we have compared the G1 and G2 values with the Pacific Decadal

416

Oscillation (PDO). PDO is defined as the leading pattern of sea surface temperature

417

anomalies in the North Pacific basin (Zhou et al., 2015). PDO resembles the ENSO

418

spatial/temporal pattern where the largest distinction is the timescales, since ENSO is

419

primarily an interannual phenomenon; the PDO is decadal in scale. Positive PDO values are

420

associated with periods of predominantly El Niño events, while negative PDO values are

421

associated with La Niña events (Hoyos et al., 2013). Our comparisons in Figure 8 show the

422

anti-phase pattern between G1 and G2 and that PDO covariates satisfactorily with G1clay

423

and silt. During El Niño phases an increase on the wind intensity is detected, therefore

424

sustaining mobilization of soil particles under the effect of surface wind stress. This

425

indicates that G1 elements apportionment in the lake could be related to regional aeolian

426

dust deposition from close sources areas such as soils from the Pliocene-Pleistocene (PPl1l)

427

formations that surround the lake. This formation type is associated with sedimentary

428

lacustrine sequences, displaying silts and clays with intercalations of calcareous levels,

429

conglomerates or pyroclastic material. Likewise, G2 elements which are related to the sand

430

component – sediment coarse fraction with diameters larger than 0.06 mm up to 2 mm –

431

could be associated with a regional upper Miocene-Pliocene formation (MP1l),

432

characterized by lacustrine sedimentary sequences, partly fluvial and alluvial limestones

433

and volcanic ash. The G2 elements occurrence can be linked with precipitation events

434

during La Niña associated dry conditions on the coast of Chile, and wet conditions in the

435

Altiplano (Williams et al., 2008).

436 437 438

Figure 8 - (a) PDO time series from 1900 and 2013 and running average; (b) silt and clay data; (c) normalized G1 data; (d) sand data; and (e) normalized G2 data.

439 440

3.3. The anthropogenic signal in the Inka-Coya record

441 442

Copper mining in Chile started initially at scale before the 20th century. By the 1910s,

443

Chile was already one of the main copper mining countries of the World, together with

444

Spain and the USA. Nevertheless, only after the First World War mining and associated

445

industry have developed in Chile. The first significant mining production occurred from

446

1914 to 1918 (Collier, 2018) and increased through the 1920s and the 1930s decades. The

447

Great Depression of 1929 in the USA has affected prominently the Copper production in

448

Chile and a decrease occurred until 1933. Copper production and industry were recovered

449

by the 1940s and 1960s decades and due to the nationalization of copper mines in Chile,

450

Copper production has increased again by mid-1970s (Schwanck et al., 2016). Investments

451

in mineral exploration in Chile reached a maximum in 1997 and declined from there on to

452

2003. After this period, Copper production in Chile presented a wider variability, following

453

the World’s demand. Figure 9a shows historical Copper variability at Inca-Koya sediment

454

core and comparison with mining production and percent of Chilean production with

455

respect to the World’s market (Fig 9c-d). In general Cu variability in the sediment core

456

represents the mining history activities consistently along the 20th century in Northern

457

Chile, depicting the initial phase of Cu mining development until the end of 1920 decade

458

and main shifts in the World’s demand. One point to raise here in that Cu variability in the

459

lake does not covariate with the production data after the 70’s decade, which has exhibited

460

a steady increase. This pattern depicts a higher similarity to the Illimani copper record

461

retrieved from the ice core at 16°37’S, 067° 46’W, in Bolivia at the very high altitude of

462

6350 m a.s.l. (Fig 9b). We attribute the discrepancy between the sedimentary and

463

glaciological record to the fact that Ilimani data capture a regional-to-global emission from

464

the observation of both curves “b” and “d” from Figure 9. On the other hand, the Inka-Coya

465

sediment core records the local production. Cu values in Figure 8a, are not consistent with

466

the mining production amount along the 20th century since maximum values observed

467

during the 1930s decade are comparable to modern concentrations after the 1990s. We

468

attribute that to the different mining emission factors since during the early stage of the

469

mining process almost no control in the dust release to the environment existed. During the

470

last decades, although larger Cu ore volumes are remobilized at the open-pit mines, actions

471

of dust emissions mitigation have caused a decrease of aeolian Cu deposition.

472 473 474 475 476 477

Figure 9 - (a) Copper variability in the Inka-Coya sediment core by XRF technique; (b) Chilean copper production (%) in the world market from 1890 to 2008 (modified from; COCHILCO, 2015); and (c) Chilean production in kilotons from 1900 to 2008 (d) Illimani and Inka-Coya copper. (Source: USGS, 2014; COCHILCO, 2015). See that scale is interrupted for better observation.

478 479

4. Conclusion

480

The high resolved elemental composition provided by the XRF scanning technique over a

481

sediment core of the Inka-Coya Lake at Andes plateau/Northern Chile together with grain

482

size analysis allowed recognizing different sedimentation patterns related to PDO

483

variability. Among several elements analyzed, three groups were identified: one composed

484

by Ti, Fe, K, Rb, Si, Zn, Zr, Mn and S, a second by P, Sr, Yb, Cl, Ca, As and Al, and Cu

485

presenting an independent pattern. The first two groups varied in opposite phase,

486

suggesting different sources of sediments apportioning the lake. These changes, together

487

with the grain size variability, showed a clear influence of the PDO at the decadal scale.

488

Cooper variability reflected the evolution of the regional mining activity along the 20th

489

century.

490 491

Acknowledgements

492 493

This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de

494

Nível Superior - Brasil (CAPES) - Finance Code 001. We would like to thank Universidad

495

de Antofagasta for the initiation scholarship for young Researchers granted to the

496

corresponding author (Vice-rectory UA 2015), to Laboratory of Sedimentology and

497

Paleoenvironments (LASPAL-UA), Laboratory Radioecology and Climatic Changes

498

(LARAMG- UERJ) and Laboratory of Oceanographic and Climate (LOCEAN-IRD) for

499

respectively XRF analysis. Dr. Jorge Valdés was favored with a stay project in Spain, by

500

the DFI program of the Ministry of Education, Chile. Thanks are also due to Dr. Marcos

501

Guiñez and Dr. Alexis Castillo for assistance in the field work. In Memorian Luc Ortlieb

502

(1948-2016).

503 504

References

505 506

Anderson, D.G., Maasch, K., Sandweiss, D.H., 2011. Climate Change and Cultural Dynamics: A Global Perspective on Mid-Holocene Transitions. Elsevier Science.

507 508

Appleby, P.G., Oldfieldz, F., 1983. The assessment of 210Pb data from sites with varying sediment accumulation rates. Hydrobiologia 103, 29-35.

509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525

Axelsson, V., 1983. The use of X-ray radiographic methods in studying sedimentary properties and rates of sediment accumulation. Hydrobiologia 103, 65-69. Clarke, J.D.A., 2006. Antiquity of aridity in the Chilean Atacama Desert. Geomorphology 73, 101-114. Boschetti, T., Cortecci, G., Barbieri, M., Mussi, M., 2007. New and past geochemical data on fresh to brine waters of the Salar de Atacama and Andean Altiplano, northern Chile. Geofluids 7, 33-50. Bull, A.T., Asenjo, J.A., Goodfellow, M., Gómez-Silva, B., 2016. The Atacama Desert: Technical Resources and the Growing Importance of Novel Microbial Diversity. Annual Review of Microbiology 70, 215-234. Croudace, I.W., Rindby, A., Rothwell, R.G., 2006. ITRAX: description and evaluation of a new multi-function X-ray core scanner. Geological Society, London, Special Publications 267, 51-63.

526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572

Francus, P., Lamb, H., Nakagawa, T., Brown, E., 2009. The potential of high-resolution Xray fluorescence core scanning: Applications in paleolimnology. Collier, S., 2018. Historia de Chile : 1808-1994 / Simon Collier, William F. Sater. SERBIULA (sistema Libr. 2.0). Gantz, A., Rau, J., Couve, E., 2009. Ensambles de aves en el desierto de Atacama, norte grande de Chile. Gayana (Concepción) 73, 172-179. Gregor, D.J., Gummer, W.D., 1989. Evidence of atmospheric transport and deposition of organochlorine pesticides and polychlorinated biphenyls in Canadian Arctic snow. Environmental Science & Technology 23, 561-565. Hakanson, L., 1980. An ecological risk index for aquatic pollution control.a sedimentological approach. Water Research 14, 975-1001. Hammer, O., Harper, D., Ryan, P., 2001. PAST: Paleontological Statistics Software Package for Education and Data Analysis. Hernández, A., Giralt, S., Bao, R., Sáez, A., Leng, M.J., Barker, P.A., 2010. ENSO and solar activity signals from oxygen isotopes in diatom silica during late glacial-Holocene transition in Central Andes (18°S). Journal of Paleolimnology 44, 413-429. Houston, J., 2006. Evaporation in the Atacama Desert: An empirical study of spatiotemporal variations and their causes. Journal of Hydrology 330, 402-412. Hoyos, N., Escobar, J., Restrepo, J.C., Arango, A.M., Ortiz, J.C., 2013. Impact of the 2010–2011 La Niña phenomenon in Colombia, South America: The human toll of an extreme weather event. Applied Geography 39, 16-25. Jordan, T., Lameli, C.H., Kirk-Lawlor, N., Godfrey, L., 2015. Architecture of the aquifers of the Calama Basin, Loa catchment basin, northern Chile. Geosphere 11, 1438-1474. Latorre, C., Betancourt, J.L., Rylander, K.A., Quade, J., Matthei, O., 2003. A vegetation history from the arid prepuna of northern Chile (22–23°S) over the last 13 500 years. Palaeogeography, Palaeoclimatology, Palaeoecology 194, 223-246. Lavado-Casimiro, W., Espinoza, J.C., 2014. Impactos de El Niño y La Niña en las lluvias del Perú (1965-2007). Revista Brasileira de Meteorologia 29, 171-182. Legendre, P., Legendre, L., 1998. Preface, In: Pierre, L., Louis, L. (Eds.), Developments in Environmental Modelling. Elsevier, pp. xi-xv. Muñoz, A., Ojeda, J., Sánchez-Valverde, B., 2002. Sunspot-like and ENSO/NAO-like periodicities in lacustrinelaminated sediments of the Pliocene Villarroya Basin (La Rioja, Spain). Journal of Paleolimnology 27, 453-463.

573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617

Rapaport, R.A., Urban, N.R., Capel, P.D., Baker, J.E., Looney, B.B., Eisenreich, S.J., Gorham, E., 1985. “New” DDT inputs to North America: Atmospheric deposition. Chemosphere 14, 1167-1173. Risacher, F., Alonso, H., Salazar, C., 2003. The origin of brines and salts in Chilean salars: a hydrochemical review. Earth-Science Reviews 63, 249-293. Ritter, B., Binnie, S.A., Stuart, F.M., Wennrich, V., Dunai, T.J., 2018. Evidence for multiple Plio-Pleistocene lake episodes in the hyperarid Atacama Desert. Quaternary Geochronology 44, 1-12. Romero, H., Méndez, M., Smith, P., 2012. Mining Development and Environmental Injustice in the Atacama Desert of Northern Chile. Environmental Justice 5, 70-76. Schwalb, A., 2003. Lacustrine ostracodes as stable isotope recorders of late-glacial and Holocene environmental dynamics and climate. Journal of Paleolimnology 29, 265-351. Schwanck, F., Simões, J.C., Handley, M., Mayewski, P.A., Bernardo, R.T., Aquino, F.E., 2016. Anomalously high arsenic concentration in a West Antarctic ice core and its relationship to copper mining in Chile. Atmospheric Environment 125, 257-264. Smith, K.A., 2000. Soil and Environmental Analysis: Physical Methods, Revised, and Expanded. Taylor & Francis. Stivaletta, N., Barbieri, R., Billi, D., 2012. Microbial Colonization of the Salt Deposits in the Driest Place of the Atacama Desert (Chile). Origins of Life and Evolution of Biospheres 42, 187-200. Swackhamer, D.L., Armstrong, D.E., 1986. Estimation of the atmospheric and nonatmospheric contributions and losses of polychlorinated biphenyls to Lake Michigan on the basis of sediment records of remote lakes. Environmental Science & Technology 20, 879-883. Tapia, J., González, R., Townley, B., Oliveros, V., Álvarez, F., Aguilar, G., Menzies, A., Calderón, M., 2018. Geology and geochemistry of the Atacama Desert. Antonie van Leeuwenhoek 111, 1273-1291. Tapia, J., Murray, J., Ormachea, M., Tirado, N., Nordstrom, D.K., 2019. Origin, distribution, and geochemistry of arsenic in the Altiplano-Puna plateau of Argentina, Bolivia, Chile, and Perú. Science of The Total Environment 678, 309-325. Valdés-Pineda, R., Cañón, J., Valdés, J.B., 2018. Multi-decadal 40- to 60-year cycles of precipitation variability in Chile (South America) and their relationship to the AMO and PDO signals. Journal of Hydrology 556, 1153-1170.

618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636

von Gunten, L., Grosjean, M., Beer, J., Grob, P., Morales, A., Urrutia, R., 2009. Age modeling of young non-varved lake sediments: methods and limits. Examples from two lakes in Central Chile. Journal of Paleolimnology 42, 401-412. Vuille, M., Keimig, F., 2004. Interannual Variability of Summertime Convective Cloudiness and Precipitation in the Central Andes Derived from ISCCP-B3 Data. Journal of Climate 17, 3334-3348. Williams, A., Santoro, C.M., Smith, M.A., Latorre, C., 2008. The impact of ENSO in the Atacama Desert and Australian arid zone: exploratory time-series analysis of archaeological records. Chungará (Arica) 40, 245-259. Zhou, X., Sun, Y., Huang, W., Smol, J.P., Tang, Q., Sun, L., 2015. The Pacific decadal oscillation and changes in anchovy populations in the Northwest Pacific. Journal of Asian Earth Sciences 114, 504-511. Ziervogel, K., Bohling, B., 2003. Sedimentological parameters and erosion behaviour of submarine coastal sediments in the south-western Baltic Sea. Geo-Marine Letters 23, 4352.

637 638 639 640

Figure captions:

641 642 643 644

Figure 1 – The study site: (a) total view of the Inka Coya Lake (b) bathymetry of the Inka Coya Lake showing a central depression and the sediment core sampling point (T1= XLCHT1); (c) Inka Coya Lake location in the context of South America and Northern Atacama orography.

645 646

Figure 2 – Map of the Inka-Coya site main geological facies. Based on Chilean geological map from the National Geology and Mining Service of Chile.

647 648 649 650 651 652 653 654

Figure 3- The meteorological database for Calama Station (22° 29.734’ S; 68°54.350 W; https://climatologia.meteochile.gob.cl); (a) Seasonal precipitation at Chiu-Chiu; (b) local wind direction; (c) MEI index from www.esrl.noaa.gov/psd/enso/mei/ spanning the period 1976-2010. MEI index is the multivariated ENSO Index, defined based on six main observed variables over the tropical Pacific: sea-level pressure, zonal wind (U). Meridional wind (V), sea surface temperature, surface air temperature and total cloudiness fraction of the sky (d) wind intensity variability for the three main wind component directions. Arrows point more intense NIÑO events in the above period.

655 656 657

Figure 4 - (a) Sediment core of Lake Inka-Coya (b) corresponding X-ray image; (c) profile of gray values; (d) profile of 210Pbexc in the core; and (e) sedimentation rates inferred from CRS model.

658 659 660

Figure 5- (a) Periodicities detected by FFT (Fourier Fast Transformation) based on greyscale series of Inka-Coya Lake sediment core. Confidence curves are depicted; (b) sunspot number (smoothed time series) corresponding to precipitate database.

661 662 663 664

Figure 6- Elemental composition along Inca-Coya sediment core as revealed by XRF analysis. Upper part, the set of elements that defined the group 1 (G1) due to their covariability; and lower part, the same for group 2 (G2). Dotted lines separate major changes in geochemistry.

665 666

Figure 7- Components of the Principal Component Analysis (PCA) for sediment core elemental composition and grain size distribution.

667 668

Figure 8 - (a) PDO time series from 1900 and 2013 and running average; (b) silt and clay data; (c) normalized G1 data; (d) sand data; and (e) normalized G2 data.

669 670 671 672 673

Figure 9 - (a) Copper variability in the Inka-Coya sediment core by XRF technique; (b) Chilean copper production (%) in the world market from 1890 to 2008 (modified from; COCHILCO, 2015); and (c) Chilean production in kilotons from 1900 to 2008 (d) Illimani and Inka-Coya copper. (Source: USGS, 2014; COCHILCO, 2015). See that scale is interrupted for better observation.

674 675 676 677 678 679 680 681 682