Mercury Isotope Signatures as Tracers for Hg ... - ACS Publications

May 10, 2013 - minerals (e.g., montroydite, HgO; eglestonite, Hg6Cl3O2H; corderoite, Hg3S2Cl2, and elemental Hg as found by Kim et al.37 and Jew et al...
1 downloads 0 Views 995KB Size
Subscriber access provided by Otterbein University

Article

Mercury isotope signatures as tracers for Hg cycling at the New Idria Hg mine Jan Georg Wiederhold, Robin Sue Smith, Hagar Siebner, Adam Jew, Gordon E. Brown, Bernard Bourdon, and Ruben Kretzschmar Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/es305245z • Publication Date (Web): 10 May 2013 Downloaded from http://pubs.acs.org on May 18, 2013

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Environmental Science & Technology

1

Mercury isotope signatures as tracers for Hg cycling at the

2

New Idria Hg mine

3

Jan G. Wiederhold1,2*, Robin S. Smith1,2, Hagar Siebner3, Adam D. Jew3, Gordon E. Brown, Jr.3,

4

Bernard Bourdon2,4, Ruben Kretzschmar1

5

1

Soil Chemistry Group, Institute of Biogeochemistry and Pollutant Dynamics, ETH Zurich, Switzerland

6

2

Isotope Geochemistry Group, Institute of Geochemistry and Petrology, ETH Zurich, Switzerland

7

3

Department of Geological and Environmental Sciences, Stanford University, Stanford, CA, USA

8

4

Laboratoire de Géologie de Lyon, ENS Lyon, CNRS and UCBL, Lyon, France

9

*

Corresponding author e-mail: [email protected]; phone: +41-44-6336008; fax: +41-44-6331118

10

Abstract

11

Mass-dependent (MDF) and mass-independent (MIF) fractionation of Hg isotopes provides a new tool

12

for tracing Hg in contaminated environments such as mining sites, which represent major point sources

13

of Hg pollution into surrounding ecosystems. Here, we present Hg isotope ratios of unroasted ore waste,

14

calcine (roasted ore), and poplar leaves collected at a closed Hg mine (New Idria, CA, USA). Unroasted

15

ore waste was isotopically uniform with δ202Hg values of -0.09 to 0.16‰ (±0.10‰, 2SD) close to the

16

estimated initial composition of the HgS ore (-0.26‰). In contrast, calcine samples exhibited variable

17

δ202Hg values ranging from -1.91‰ to +2.10‰. Small MIF signatures in the calcine were consistent

18

with nuclear volume fractionation of Hg isotopes during or after the roasting process. The poplar leaves

19

exhibited negative MDF (-3.18 to -1.22‰) and small positive MIF values (∆199Hg of 0.02 to 0.21‰).

20

Sequential extractions combined with Hg isotope analysis revealed higher δ202Hg values for the more

21

soluble Hg pools in calcines compared with residual HgS phases. Our data provide novel insights into

22

possible in-situ transformations of Hg phases and suggest that isotopically heavy secondary Hg phases

23

were formed in the calcine, which will influence the isotope composition of Hg leached from the site.

24 1

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 28

25

Introduction

26

Mercury (Hg) has been listed by the World Health Organization as one of the top ten chemicals of

27

public health concern1 and a global mercury treaty coordinated by the United Nations Environmental

28

Programme will be launched in 2013.2 Mercury poses a serious threat for human and ecosystem health

29

at local and global scales, especially in its methylated form, which bioaccumulates in natural food webs.

30

However, all Hg species are toxic and can be transformed by various abiotic and biotic processes and

31

transported in particulate, dissolved, or gaseous forms in the environment.3,4 The modern global Hg

32

cycle is dominated by anthropogenic influences, and current research is focused on understanding the

33

complex interplay of different natural and anthropogenic sources and processes controlling

34

biogeochemical Hg cycling and its environmental impact on ecosystems worldwide.5

35

Despite its toxicity, Hg has a long history of commercial use for various applications and has been

36

mined since Roman times.6 Mercury is a very rare element in the Earth’s crust (< 0.1 µg g-1), but occurs

37

locally in ore deposits at much higher concentrations mainly in the form of cinnabar (α-HgS).7 These

38

deposits are globally distributed with the two largest being located in Spain (Almaden) and Slovenia

39

(Idrija). The two largest and historically most productive Hg mines in North America are located in the

40

California Coast Range mercury belt8 in the New Almaden and New Idria deposits, which were named

41

after the European mines. All of these Hg mines are closed today, but the legacy of the intense mining

42

activities is still present at the former mining sites and in the surrounding ecosystems. Abandoned Hg

43

mines are major point sources for Hg emissions, mainly due to the presence of large volumes of calcine,

44

the residue of the on-site roasting process, which was deposited in tailings at the mine sites.9 Emissions

45

occur in gaseous form into the atmosphere10 and in particle-bound or dissolved form in the mine

46

drainage11 and leaching of tailing piles12 into aqueous environments13. Despite intense research on Hg

47

emissions from Hg mines worldwide9,14-18, many questions concerning the long-term fate of Hg and its

48

environmental impact still remain unanswered.

2

ACS Paragon Plus Environment

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Environmental Science & Technology

49

Stable Hg isotope analysis offers a new tool for addressing some of these questions by providing

50

characteristic “fingerprint” signatures of sources and processes in biogeochemical Hg cycling. Due to

51

recent analytical advances, especially in multicollector inductively-coupled plasma mass spectrometry

52

(MC-ICPMS), high-precision analysis of many “non-traditional” stable isotope systems19,20 including

53

Hg21,22 have become feasible. The rapid development of stable Hg isotope geochemistry during the last

54

few years was additionally sparked by the discovery of mass-independent fractionation (MIF)23

55

affecting mainly the two odd-mass Hg isotopes

56

fractionation (MDF) influencing all seven stable Hg isotopes (196Hg, 198Hg, 199Hg, 200Hg, 201Hg, 202Hg,

57

and

58

associated with kinetic radical-pair mechanisms in photochemical processes in which the two odd-mass

59

Hg isotopes possessing nuclear spin and magnetic moment can be strongly fractionated. Smaller MIF of

60

Hg isotopes can also originate from nuclear volume effects25 caused by the non-linear increase of

61

nuclear charge radii with mass, again affecting mainly the two odd-mass Hg isotopes. The two effects

62

can be distinguished by their different extent of MIF for

63

MDF signature obtained from the

64

samples. Experimental work and an increasing number of field studies have revealed that Hg isotopes

65

are fractionated by kinetic and equilibrium effects during abiotic and biotic processes and that

66

significant variations exist in the Hg isotope composition of samples from various environments (see

67

recent reviews26-28). These variations form the basis for the application of Hg isotopes as source and

68

process tracers of biogeochemical Hg cycling in nature. Despite the impressive progress over the last

69

years, many processes and mechanisms controlling the distribution of Hg isotopes in natural samples are

70

still poorly understood, and case studies from field systems with well-defined Hg sources coupled with

71

previously studied transformation processes are needed to further develop this novel isotopic tool.

72

Mining environments offer an ideal field system for this purpose because the dominant Hg species and

73

pathways are relatively well-known, in addition to the high Hg concentrations and the environmental

204

199

Hg and

201

Hg, in addition to mass-dependent

Hg). Large MIF of Hg isotopes can be caused by magnetic isotope effects,24 probably mainly

202

199

Hg and

201

Hg,26 which, together with the

Hg/198Hg ratio, provides an “isotopic fingerprint” of Hg in natural

3

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

74

relevance of these sites as detailed above. The pioneering study of Stetson et al.29 investigating Hg

75

isotope ratios of samples from Hg mines in Nevada and Texas (USA) has already revealed that

76

significant differences exist between different materials, which were at least partly caused by the ore

77

processing. Calcine samples, although exhibiting a relatively wide range of Hg isotope compositions,

78

were found to be mainly enriched in heavy isotopes relative to the respective cinnabar ore samples.

79

Interesting isotopic variations were also reported for leachates (single step at pH 5) of calcine samples

80

and for museum specimens of several Hg-containing minerals, which could occur as trace constituents

81

in calcine. Recently, Yin et al.30 reported isotopic differences between unprocessed Hg ores and calcine

82

materials from the Wanshan Hg mine (SW China) with an enrichment in heavy Hg isotope in the

83

calcine samples of on average +0.8‰ in δ202Hg, which was interpreted as being caused by the roasting

84

process. Our study builds upon the data reported in these earlier studies by presenting Hg isotope data of

85

different materials collected at the New Idria Hg mine, California (USA). The specific objectives of the

86

study were (1) to investigate the Hg isotope composition of unroasted ore waste and roasted calcine, (2)

87

to apply sequential extraction methods to assess the isotope signatures of individual Hg pools, thereby

88

allowing us to gain insights into in-situ transformations of Hg species in calcine during or after the

89

roasting process, and (3) to report Hg isotope ratios of poplar leaf samples collected at the site, which

90

are presumably influenced by the isotope composition of Hg emitted to the atmosphere.

91 92

Materials and Methods

93

Description of Sampling Site. The New Idria Hg mine, which is a silica-carbonate type Hg deposit in

94

San Benito County, California, USA, was in operation between 1854 and 1972 and was the second

95

largest Hg mine in North America producing a total of ~20,000 tons of elemental Hg.31-35 The HgS-rich

96

rock was mined in underground tunnels, crushed to smaller pieces (generally < 5 cm), and roasted

97

on-site in rotary furnaces or retorts at ~600-700°C, producing gaseous elemental Hg (HgS(s) + O2(g) →

98

Hg0(g) + SO2(g)) which was trapped in condenser coils and bottled for sale (e.g., used in gold mining and 4

ACS Paragon Plus Environment

Page 5 of 28

1 2 3 4 99 5 6 100 7 8 101 9 10 11102 12 13103 14 15104 16 17 18105 19 20106 21 22107 23 24 108 25 26 27109 28 29110 30 31 111 32 33 34112 35 36113 37 38114 39 40 41115 42 43116 44 45117 46 47 118 48 49 50119 51 52120 53 54 121 55 56 57122 58 59123 60

Environmental Science & Technology

various industrial applications). The Hg removal process was incomplete and the roasted calcine, which turned reddish due to the formation of ferric iron oxide minerals such as hematite, was piled on-site in tailings that still contained high Hg concentrations. As a result of the mining activities, an estimated volume of 0.38 to 1.5 million m3 of waste rock and calcine materials is spread today over an area of ~162,000 m2 at the New Idria site.35 The site was placed on the National Priority List of the US Environmental Protection Agency (“Superfund list”) in September 2011 and remediation actions with the goal to minimize the environmental impact (e.g., acid mine drainage from open adit) are currently underway.36 However, at the time of sampling for this study, the site was still in a state similar to the situation after the mine closure in 1972. Previous studies have investigated various aspects of Hg cycling at New Idria12,13,37,38, and Hg isotope ratios from a limited number of samples from New Idria were already reported in compilations of signatures for cinnabar ores,22 rocks from the California coast ranges,39 and potential Hg sources for San Francisco Bay. 40 Sampling and Sample Preparation. Unroasted ore waste from seven locations (O1 to O7) and calcine from nine locations (C1 to C9) were collected as grab samples at New Idria between March 2009 and April 2011 (Supporting Information, SI Fig. S1). At some locations, multiple samples were collected, resulting in a total number of ten unroasted ore waste (grey) and fourteen calcine (red) samples. Six leaf samples of poplar trees (Populus spec.) were collected 1 to 1.75 m above ground from different locations at New Idria (SI Fig. S1) in April 2008 and May and October 2009, washed thoroughly using deionized water, freeze-dried, ground, homogenized, and stored in a desiccator before further treatment. A sample of the original HgS ore from the New Idria mine was obtained from a Stanford University Research Mineral Collection. Many of the larger calcine samples exhibited a pronounced internal layer structure with a black core, a grey rim, and a red rind at the surface that must have formed during the roasting process. A detailed small-scale investigation on these layered calcine materials, including Hg isotope variations between the individual layers will be presented in a separate paper. In the present study, all calcine samples represent bulk materials that were homogenized by grinding. Elemental 5

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 124 5 6 125 7 8 126 9 10 11127 12 13128 14 15129 16 17 18130 19 20131 21 22132 23 24 133 25 26 27134 28 29135 30 31 136 32 33 34137 35 36138 37 38139 39 40 41140 42 43141 44 45142 46 47 143 48 49 50144 51 52145 53 54 146 55 56 57147 58 59148 60

Page 6 of 28

concentrations in the ore waste and calcine samples were determined by energy-dispersive X-ray fluorescence spectrometry (Spectro-X-Lab 2000). For the ore waste and calcine samples, ~500 mg of powdered material was digested with aqua regia and BrCl. For the poplar leaf samples, 1 g of freezedried, ground material was digested with H2SO4/HNO3 (see SI for details). For the sequential extractions, ~1 g of powdered ore waste or calcine material was leached first with 25 mL of a 0.1 M acetic acid / 0.01 M HCl solution (F1: simulated “stomach acid”-step targeting more soluble Hg forms) and placed on an end-over-end shaker overnight before centrifugation and filtration with 0.45 µm filters. The second extraction step was performed in an analogous manner using 12 M HNO3 (F2: targeting all remaining non-sulfidic Hg species), before the residual material was digested with aqua regia and BrCl as described above (F3: containing the remaining stable HgS phases). This three-step sequential extraction procedure corresponds to the extraction steps F2, F4, and F5 of the established five-step method published by Bloom et al..41 Step F1 (water extract) was omitted to yield sufficiently high Hg concentrations for Hg isotope analysis in the first extraction step (thus containing F1+F2 after Bloom et al.). Step F3 of Bloom’s procedure (1 M KOH) was omitted as we assumed there was no significant organically bound Hg pool in the investigated ore waste and calcine samples. Concentration and Isotope Analyses. Total Hg concentrations in all digest and extract solutions were measured by cold vapor atomic fluorescence spectrometry (CV-AFS, Millennium Merlin, PS Analytical). All sample and standard solutions were diluted to 20 µg L-1 Hg prior to isotope analysis, except for the poplar leaf digests which were diluted to 10 µg L-1 Hg due to their lower Hg contents. Stable Hg isotope analysis was performed by MC-ICPMS (Nu Plasma, Nu instruments) with cold-vapor introduction (HGX-200, Cetac) and online Tl addition with a desolvating nebulizer for mass-bias correction in addition to standard-bracketing with NIST-3133 following our published methods42,43 which are based on previous method developments44,45. The results are reported in permil (‰) as δ202Hg = [(202Hg/198Hg)sample / (202Hg/198Hg)NIST-3133] - 1 for mass-dependent fractionation (MDF) and

(

)

(

)

∆199 Hg = δ 199 Hg − δ 202 Hg × 0.2520 and ∆201 Hg = δ 201 Hg − δ 202 Hg × 0.7520 for mass-independent 6

ACS Paragon Plus Environment

Page 7 of 28

1 2 3 4 149 5 6 150 7 8 151 9 10 11152 12 13153 14 15 154 16 17 18155 19 20156 21 22157 23 24 25158 26 27159 28 29160 30 31 32161 33 34162 35 36163 37 38 164 39 40 41165 42 43166 44 45167 46 47 48168 49 50169 51 52170 53 54 55171 56 57172 58 59173 60

Environmental Science & Technology

fractionation (MIF) of Hg isotopes, following recent recommendations for data reporting.45,46 The reproducibility based on replicate measurements of our in-house standard (ETH Fluka Hg) in the same analytical sessions was ±0.10‰ for δ202Hg and ±0.04‰ for ∆199Hg and ∆201Hg (2SD, n=11). The accuracy of our data was verified by analyses of the UM-Almaden standard and digests of the reference material NIST-2711 (Montana soil) yielding results that were in excellent agreement with published values (SI Table S3). Further information about method details and data quality control measures is provided in the SI.

Results and Discussion. Hg Concentrations. The samples exhibited a wide range of Hg concentrations in the unroasted ore waste (17 to 943 µg g-1) and in the calcine (1.7 to 665 µg g-1) (Fig. 1), but all values were highly elevated relative to background values for Hg in uncontaminated rocks and soils (< 0.1 µg g-1).7 The unroasted ore waste samples, collected from the large grey waste pile close to the mine adit, likely represented host rock material from the underground tunnels which was considered not to be rich enough in Hg for ore processing. The Hg contents of the calcine samples represented the residual amounts of Hg after incomplete removal during thermal processing of the high-grade HgS-rich ore materials. Further geochemical differences between the unroasted ore waste and the calcine samples, in addition to the striking difference in color, can be derived from the element concentrations measured by XRF (SI Fig. S7). As expected, the calcine samples exhibited generally lower S and higher Fe contents compared with the unroasted ore waste samples. The Hg concentrations in all samples did not correlate with the concentrations of major elements but exhibited significant correlations with selected trace elements (e.g., Br, Tl, Se), which suggests that these elements were hosted together in similar mineral phases. The initial Hg concentration of the processed ore material before roasting is unknown and was likely heterogeneous. The New Idria HgS ore sample from the mineral collection exhibited a Hg content of 49% which can be taken as a maximum estimate of the initial ore concentration. According to historic 7

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 174 5 6 175 7 8 176 9 10 11177 12 13178 14 15179 16 17 18180 19 20181 21 22182 23 24 183 25 26 27184 28 29185 30 31 186 32 33 34187 35 36188 37 38189 39 40 41190 42 43191 44 45192 46 47 193 48 49 50194 51 52195 53 54 196 55 56 57197 58 59198 60

Page 8 of 28

sources, the initial high-grade ore material processed at New Idria contained up to 10% Hg, but later averages range from 0.5 to 1%, and it was reported that the New Idria furnaces even allowed profitable mining of ore materials containing only 0.1% Hg.32 In addition, the relative fraction of Hg removal varied during different periods of the mining history at New Idria. Retort furnaces with lower extraction efficiencies were replaced in 1918 by more efficient rotary kilns and some of the previously deposited tailing materials were re-processed for further Hg extraction35. It was impossible, however, to relate the sampled calcine materials to specific processing times or methods. Another rough estimate of the average initial Hg concentrations in the roasted materials can be obtained by comparing the total amount of produced Hg in New Idria during the operation of the mine with the estimated volumes of deposited waste rock and calcine materials at the site35 yielding values of approximately 1-5% Hg (or 10,000 to 50,000 µg g-1), consistent with the other estimates above. In any case, the extraction of Hg during roasting was incomplete, leaving behind significant amounts of residual Hg in calcines with contents ranging between 1 to 1000 µg g-1. Thus, although highly variable, the residual fraction in the calcines likely corresponded to 90% of the total Hg pool from the ore into the elemental form (despite being not fully quantitative), the produced elemental Hg cannot be expected to have been strongly fractionated due to mass balance considerations and presumably exhibited an average Hg isotope composition which was similar to the initial ore value. This scenario does not exclude the possibility that significant Hg isotope fractionation occurred during and after the roasting process, impacting the signatures of the smaller Hg pools which were emitted to the atmosphere or remained in primary or secondary phases in the calcine materials. The variability of Hg isotope signatures in the calcine samples clearly suggests that the roasting process caused significant Hg isotope fractionation. It is important to note, however, that the potential to create an isotopically different residual pool only by preferential removal of light Hg isotopes during roasting is rather limited. This is because the Hg extraction process probably progressed along a moving reaction front into the roasted ore fragments during the calcination process (the ore fragments stayed about 30 min. in the rotary furnaces), where a quantitative conversion to elemental Hg on the outside contrasted with the less-affected inner parts of the ore fragments in which some HgS phases presumably remained more or less intact (“nugget effect”, as also proposed in Stetson et al.29). Thus, a kinetic Rayleigh-type fractionation model, which requires a homogenous, well-mixed substrate, does not adequately describe the evolution of isotope ratios during the calcination process, as also described in other isotopic studies investigating the evaporation of minerals.59 In addition, as already noted above, the observed Hg isotope ratios in the calcine samples did not correlate with the Hg concentrations and exhibited both enrichments and depletions in heavy Hg isotopes (Fig. 1). We propose, based on the data reported here and in agreement with the study by 13

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 324 5 6 325 7 8 326 9 10 11327 12 13328 14 15329 16 17 18330 19 20331 21 22332 23 24 25333 26 27334 28 29335 30 31 32336 33 34337 35 36338 37 38 339 39 40 41340 42 43341 44 45342 46 47 48343 49 50344 51 52345 53 54 55346 56 57347 58 59348 60

Page 14 of 28

Stetson et al.,29 that the Hg isotope variations observed in the calcine samples were dominantly caused by the formation of isotopically fractionated secondary Hg phases during and/or after the roasting process. This inference is corroborated by the results of the sequential extraction analyses. If the preferential removal of light Hg isotopes from HgS phases during the roasting process were primarily responsible for the observed isotopically heavy calcine signatures, then positive δ202Hg values should be found mainly in the residual HgS phases probed by the F3 step of the sequential extraction. However, this fraction exhibited the smallest fractionations relative to the initial ore composition and was always isotopically lighter than the bulk calcine samples. The isotopically heavy calcine values are thus dominantly caused by the signatures of the more soluble Hg phases which dissolved in the first two extraction steps and most likely represent secondary mineral phases formed during or after the calcination process. Secondary “by-product” minerals (e.g., montroydite, HgO; eglestonite, Hg6Cl3O2H; corderoite, Hg3S2Cl2, and elemental Hg as found by Kim et al.37 and Jew et al.38) exhibit a much higher solubility compared with HgS phases and contain Hg in the mercuric (HgII) or mercurous (HgI) oxidation state, which would be expected to be isotopically heavy relative to elemental Hg vapor.25,42 Thus, the isotopically heavy signatures of the first extracts (especially F1) are consistent with the concept that secondary Hg phases formed in reactions with the large elemental Hg pool (unfractionated relative to the initial ore due to mass balance constraints) potentially during the cooling of the calcine pieces after leaving the furnaces. The exact geochemical conditions and pathways of Hg in the furnaces are impossible to reconstruct in detail, but it is likely that small Hg pools, which were not extracted from the calcine pieces and transported in gaseous elemental form to the condenser coils, were strongly affected by physical transport limitations within the cm-sized calcine fragments (e.g., diffusion controlled transport). In addition, re-condensation and re-oxidation reactions during and after passage through the rotary furnaces (~30 min) and the subsequent cooling of the fragments, which occurred from the outside to the inside, might have influenced the system. Thus, it is not surprising that these 14

ACS Paragon Plus Environment

Page 15 of 28

1 2 3 4 349 5 6 350 7 8 351 9 10 11352 12 13353 14 15 354 16 17 18355 19 20356 21 22357 23 24 25358 26 27359 28 29360 30 31 32361 33 34362 35 36363 37 38 39364 40 41365 42 43366 44 45 367 46 47 48368 49 50369 51 52 370 53 54 55371 56 57372 58 59373 60

Environmental Science & Technology

residual Hg phases in calcine exhibit large variations in their Hg isotope composition considering that multiple redox changes, changes in the coordination environment, and incomplete physical separations of Hg pools by transport processes can all be expected to exert an influence on the Hg isotope distribution. Furthermore, our calcine data indicate that the Hg isotope fractionation related to the roasting process can shift δ202Hg values in both directions. A recent study reported only enrichments of heavy Hg isotopes in calcine relative to unroasted ore.30 However, Stetson et al.29 also found light Hg isotope ratios in some calcine samples presumably influenced by different fractionation mechanisms and Hg transport pathways, which are, as in our study, likely to vary between samples considering the complexity and heterogeneity of the geochemical conditions during and after the roasting processes as discussed above.The discovery of small MIF signatures in calcine relative to the unfractionated ore and unroasted ore waste suggests that at least some of the fractionating processes were accompanied by a small mass-independent component. Considering the magnitude of the observed effects and the “opposite trend” of MDF and MIF signatures (Fig. 2 for bulk samples, Fig. 4 and SI Fig. S5 for extracts), it is likely that nuclear volume fractionation (NVF), for instance during the reaction between elemental Hg and secondary oxidized Hg phases, was involved, which would be consistent with theoretical predictions.25,42 An accurate determination of the slope in the ∆199Hg/∆201Hg plot (SI Fig. S6) allowing further insight into the underlying MIF mechanisms is impeded by the small magnitude of the effects and the relatively large scatter in the data. The data points were in general agreement with a slope of 1.65 characteristic for NVF,42,58 but a potential influence of magnetic isotope effects causing slopes between 1.0 and 1.323 cannot be completely excluded, although there is no evidence for the involvement of radical-pair reactions or photochemical processes in the history of the studied calcine samples. Hg Isotope Signatures in Poplar Leaves. The Hg concentrations of the poplar leaves collected in New Idria (sampling locations see SI Fig. S1) were about an order of magnitude higher than the control sample of poplar leaves collected several kilometers away from the mine site. It was not possible with 15

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 374 5 6 375 7 8 376 9 10 11377 12 13378 14 15 379 16 17 18380 19 20381 21 22382 23 24 25383 26 27384 28 29385 30 31 32386 33 34387 35 36388 37 38 389 39 40 41390 42 43391 44 45392 46 47 48393 49 50394 51 52395 53 54 55396 56 57397 58 59398 60

Page 16 of 28

our available methods to analyze the isotopic composition of the control sample due to its low Hg concentration. However, the concentration difference indicated that the poplar leaves collected at the mine site were strongly influenced by the presence of the Hg mine. The observed strong enrichment of light Hg isotopes in poplar leaves from New Idria (-3.18 to -1.22‰ in δ202Hg) can be explained by different potential pathways. Fractionation during uptake through the root system and/or preferential translocation of light isotopes to the leaves has been demonstrated for other metal isotope systems60 such as Ca61, Fe62, Cu and Zn63. However, plant leaves absorb Hg under field conditions predominantly in gaseous form from the atmosphere, whereas upward transport from the root system to the leaves is generally believed to be of minor importance.48 The dominance of atmospheric uptake pathways for Hg in plant leaves has been clearly demonstrated in many studies.64-68 Poplars are deciduous trees and thus, the Hg in our leaf samples originated from recent emissions during the last growing season. Gaseous elemental Hg, which has been generated by reduction and/or volatilization processes, can be expected to be enriched in light Hg isotopes based on experimental studies.26-28 Thus, our poplar leaf data could be explained by the uptake of elemental Hg from the near-surface atmosphere at New Idria, likely enriched in light Hg isotopes by evasion processes from the mine tailings. However, Hg isotope fractionation could occur during different steps of the Hg transport from the mine tailings through the atmosphere into the poplar leaves. The observed mass-dependent enrichment of light Hg isotopes in the poplar leaves appeared to correlate with decreasing Hg concentrations in the leaves (R2 = 0.69) (Fig. 1). Such a relationship could be caused for instance by changes in the isotopic composition of the atmospheric Hg pool, fractionation during uptake of atmospheric Hg, or mixing of different Hg pools in the leaves. In addition, we cannot exclude the possibility that re-emission processes further influenced the Hg isotope composition of the poplar leaves. The first Hg isotope data from samples of higher plants were published by Ghosh et al.69 who reported large negative MIF effects for Tillandsia samples (“Spanish moss”), but their analytical method involving an internal normalization did not allow the extraction of the MDF signature. Recently, 16

ACS Paragon Plus Environment

Page 17 of 28

1 2 3 4 399 5 6 400 7 8 401 9 10 11402 12 13403 14 15 16404 17 18405 19 20406 21 22 23407 24 25408 26 27409 28 29 30410 31 32411 33 34412 35 36 413 37 38 39414 40 41415 42 43416 44 45 46417 47 48418 49 50 419 51 52 53420 54 55421 56 57 422 58 59 60423

Environmental Science & Technology

Demers et al.70 reported a δ202Hg range of -2.53 to -1.89‰ for leaves of aspen trees (belonging to the same genus as poplar) from a forest ecosystem experiment in Wisconsin (USA), which lies within the range we observed for poplar leaves in this study. The authors were able to analyze the Hg isotope composition of total gaseous Hg in the atmosphere (0.48 to 0.93‰) and concluded that the uptake step of Hg from the atmosphere into the aspen leaves resulted in a large shift in δ202Hg of about -2.9‰. Another recent study by Yin et al.71 presented Hg isotope data of rice plants from the Wanshan Hg mining district in China. The rice leaves exhibited δ202Hg values of -2.88 to -3.38‰ compared with atmospheric total gaseous Hg (-1.85 to -2.32‰) and lichens (-1.83 to -2.32‰), suggesting a fractionation of about -1.0‰ during uptake of atmospheric Hg into the rice leaves, but not during uptake into the lichens. In previous studies, Hg isotope data of lichen samples exhibiting negative MDF and MIF signatures have been reported and used as an estimate of the atmospheric composition72 and for air pollution tracing73,74. Our poplar leaf data are generally consistent with the concept that the local atmospheric Hg in New Idria is enriched in light Hg isotopes, but we are not able to provide a definite answer to this issue, due to the lack of direct atmospheric Hg isotope data and the high probability of superimposed fractionation effects during uptake into the leaves.70,71 Considering the potential pathways of Hg into the poplar leaves at the New Idria site, it is very likely though that the evaporation of gaseous elemental Hg from tailing piles represented the most important source. However, this interpretation should be confirmed in future studies which will have to investigate in more detail the Hg isotope systematics of uptake into plant leaves and potential re-emission processes. The relatively small magnitude of the MIF signatures (∆199Hg of 0.02 to 0.21‰) measured in our poplar leaf samples does not indicate a major influence of photochemical processes involving magnetic isotope effects on the Hg evasion and atmospheric transformation processes at New Idria. No correlation with Hg concentrations was found for the MIF signatures (SI Fig. S4). Interestingly, all our poplar leaf samples exhibited positive ∆199Hg values, which constrasts to the mostly negative MIF values reported for leaf and lichen samples in previous studies69-74. This difference could be related to atmospheric 17

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 424 5 6 425 7 8 426 9 10 11427 12 13428 14 15429 16 17 18430 19 20431 21 22432 23 24 433 25 26 27434 28 29435 30 31 436 32 33 34437 35 36438 37 38439 39 40 41440 42 43441 44 45442 46 47 443 48 49 50444 51 52445 53 54 446 55 56 57447 58 59448 60

Page 18 of 28

processes or other site-specific parameters in our field site, but we are not able to provide a clear explanation based on the available data. Environmental Implications. The Hg isotope data reported in this study represent an important step towards understanding Hg isotope variations in mining environments. The variable Hg isotope signatures of calcines were shown to be caused mainly by secondary Hg phases that formed during or after the roasting process. In contrast, the preferential removal of light isotopes during the incomplete extraction of the HgS ore likely played only a minor role in causing the calcine isotope composition. Overall, our data indicate that Hg isotopes can indeed provide important new insights about Hg cycling in contaminated environments such as mining sites. The analysis of Hg pools separated by sequential extractions, performed in parallel with bulk analyses to verify the isotopic mass balance, allowed, for the first time, the resolution of in-situ differences between different components of the Hg isotope signature of a bulk sample. We suggest that this is a promising approach for future Hg isotope studies in other environmental systems. Obviously, sufficiently high Hg concentrations are required to allow the separation into different pools, but current analytical advances and the development of pre-enrichment steps will broaden the range of sequential extractions in future Hg isotope studies. The finding that the more soluble Hg pools of the calcine samples were enriched in heavy Hg isotopes compared with the bulk composition suggests that Hg emissions in dissolved forms caused by drainage of calcine tailings will most likely be strongly enriched in heavy Hg isotopes, consistent with the conclusion of Stetson et al.29 that the signature of leached Hg from mining environment can be expected to differ significantly from the initial ore signature. Finally, our poplar leaf data suggest that vegetation taking up atmospheric Hg at contaminated sites is characterized by strong mass-dependent enrichments in lighter Hg isotopes and slightly positive MIF values, which is probably related to the isotope composition of gaseous elemental Hg evading from tailing piles as well as potential fractionation effects during uptake into the leaves, but this interpretation requires confirmation and mechanistic clarification by future studies.

18

ACS Paragon Plus Environment

Page 19 of 28

1 2 3 4 449 5 6 450 7 8 451 9 10 11452 12 13453 14 15454 16 17 18455 19 20456 21 22457 23 24 458 25 26 27459 28 29460 30 461 31 32462 33463 34464 35 36465 37466 38467 39468 40 41469 42470 43471 44 472 45 46473 47474 48475 49 476 50 51477 52478 53479 54 55480 56481 57482 58483 59 484 60

Environmental Science & Technology

Acknowledgements We thank Martin Jiskra for help in the MC-ICPMS lab and various additional support, Patrick Früh and Kurt Barmettler for support in the soil chemistry lab, and Joel Blum and two anonymous reviewers for helpful comments. This study was supported in part by ETH Zurich (Grant No. ETH-15-09-2), by the Stanford-NSF Environmental Molecular Science Institute (NSF-CHE—0431425), and by the NSF Center for Environmental Implications of Nanotechnology (NSF Cooperative Agreement EF-0830093). Supporting Information The Supporting Information (SI) contains a map of the sampling site, additional method descriptions, figures, and data tables. This material is available free of charge via the Internet at http://pubs.acs.org.

References (1) WHO Mercury and health, Fact Sheet N°361. World Health Organization, April 2012. (2) UNEP Minamata convention agreed by nations. United Nations Environment Programme. Press Release, 19 January 2013 (http://www.unep.org). (3) Liu, G., Cai, Y., O'Driscoll, N. J., Eds. Environmental Chemistry and Toxicology of Mercury. John Wiley & Sons, New York, USA, 2012. (4) Bank, M. S., Ed. Mercury in the environment: pattern and process. University of California Press, Berkeley, USA, 2012. (5) Selin, N. E. Global biogeochemical cycling of mercury: A review. Annu. Rev. Environ. Resour. 2009, 34: 43-63. (6) Brooks, W. E. Industrial use of mercury in the ancient world. In Mercury in the environment: pattern and process. Bank, M.S. (ed.) University of California Press, Berkeley, USA, 2012, pp 19-24. (7) Fitzgerald, W. F., Lamborg, C. H. Geochemistry of mercury in the environment. Treatise on Geochemistry 2003, 9: 107-149. (8) Rytuba, J. J. Mercury from mineral deposits and potential environmental impact. Environ. Geol. 2003, 43: 326-338. (9) Rytuba, J. J. Geogenic and mining sources of mercury to the environment. In: Mercury, Sources, Measurements, Cycles and Effects. Parsons, M. B., Percival, J. B., Eds.; Min. Ass. of Canada, Short Course Series, Vol. 34., 2005. (10) Coolbaugh, M. F.; Gustin, M. S.; Rytuba, J. J. Annual emissions of mercury to the atmosphere from natural sources in Nevada and California. Env. Geol. 2002, 42: 338–349. (11) Rytuba, J. J. Mercury mine drainage and processes that control its environmental impact. Sci. Tot. Environ. 2000, 260: 57-71. (12) Lowry, G. V.; Shaw, S.; Kim, C. S.; Rytuba, J. J.; Brown, Jr., G. E. Macroscopic and microscopic observations of particle-facilitated mercury transport from New Idria and Sulphur Bank mercury mine tailings. Environ. Sci. Technol. 2004, 38: 5101-5111. 19

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 485 5 486 6 487 7 488 8 9 489 10490 11491 12492 13 14493 15494 16495 17 496 18 19497 20498 21499 22 23500 24501 25502 26503 27 28504 29505 30506 31507 32 508 33 34509 35510 36 511 37 38512 39513 40514 41 42515 43516 44517 45518 46 519 47 520 48 49521 50522 51 523 52 524 53 54525 55526 56527 57 528 58 59529 60530

531

Page 20 of 28

(13) Ganguli, P. M.; Mason, R. P.; Abu-Saba, K. E.; Anderson, R. S.; Flegal, A. R. Mercury speciation in drainage from the New Idria mercury mine, California. Environ. Sci. Technol. 2000, 34: 4773-4779. (14) Gray, J. E.; Hines, M. E.; Higueras, P. L.; Adatto, I.; Lasorsa, B. K. Mercury speciation and microbial transformations in mine wastes, stream sediments, and surface waters at the Almaden Mining District, Spain. Environ. Sci. Technol. 2004, 38: 4285-4292. (15) Navarro, A. Review of characteristics of mercury speciation and mobility from areas of mercury mining in semi-arid environments. Rev. Environ. Sci. Biotechnol. 2008, 7:287–306. (16) Lin, Y.; Larssen, T.; Vogt, R.D.; Feng, X. Identification of fractions of mercury in water, soil and sediment from a typical Hg mining area in Wanshan, Guizhou province, China. Appl. Geochem. 2010, 25: 60-68. (17) Gosar, M.; Tersic, T. Environmental geochemistry studies in the area of Idrija mercury mine, Slovenia. Environ. Geochem. Health 2012, 34: 27-41. (18) Kocman, D.; Horvat, M.; Pirrone, N.; Cinnirella, S. Contribution of contaminated sites to the global mercury budget. Environ. Res. in press, doi: 10.1016/j.envres.2012.12.011. (19) Johnson, C. M.; Beard, B. L.; Albarède, F. Geochemistry of non-traditional stable isotopes. Rev. Mineral. Geochem. 2004, 55. (20) Baskaran, M., Ed. Handbook of Environmental Isotope Geochemistry. Springer, Heidelberg, Germany, 2012. (21) Lauretta, D. S.; Klaue, B.; Blum, J. D.; Buseck, P. R. Mercury abundances and isotopic compositions in the Murchison (CM) and Allende (CV) carbonaceous chondrites. Geochim. Cosmochim. Acta 2001, 65: 2807-2818. (22) Hintelmann, H.; Lu, S. High precision isotope ratio measurements of mercury isotopes in cinnabar ores using multi-collector ICP/MS. Analyst 2003, 128: 635-639. (23) Bergquist, B. A.; Blum, J. D. Mass-dependent and -independent fractionation of Hg isotopes by photoreduction in aquatic systems. Science 2007, 318: 417-420. (24) Buchachenko, A. L. Mercury isotope effects in the environmental chemistry and biochemistry of mercury-containing compounds. Russ. Chem. Rev. 2009, 78: 319-328. (25) Schauble, E. A. Role of nuclear volume in driving equilibrium stable isotope fractionation of mercury, thallium, and other very heavy elements. Geochim. Cosmochim. Acta 2007, 71: 2170-2189. (26) Blum, J. D. Applications of stable mercury isotopes to biogeochemistry. In Handbook of Environmental Isotope Geochemistry. Baskaran, M. (ed.) Springer, Springer, Heidelberg, Germany, 2012, pp 229-245. (27) Hintelmann, H., Zheng, W. Tracking geochemical transformations and transport of mercury through isotope fractionation. In Environmental Chemistry and Toxicology of Mercury, Liu, G., Cai, Y., O'Driscoll, N. J, . Eds.; John Wiley & Sons, New York, USA, 2012, pp 293-327. (28) Hintelmann, H. Use of stable isotopes in mercury research. In Mercury in the environment: pattern and process. Bank, M.S., Ed.; University of California Press, Berkeley, USA, 2012, pp 55-71. (29) Stetson S. J.; Gray J. E.; Wanty R. B.; Macalady, D. L. Isotopic variability of mercury in ore, mine-waste calcine, and leachates of mine-waste calcine from areas mined for mercury. Environ. Sci. Technol. 2009, 43: 7331-7336. (30) Yin, R.; Feng, X.; Wang, J.; Li, P.; Liu, J.; Zhang, Y.; Chen, J.; Zheng, L.; Hu, T. Mercury speciation and mercury isotope fractionation during ore roasting process and their implication to source identification of downstream sediment in the Wanshan mercury mining area, SW China. Chem. Geol. 2013, 336: 72-79. (31) Eckel, E. B.; Meyers, W. B. Quicksilver deposits of the New Idria District, San Benito and Fresno counties, California. Calif. Jour. Mines and Geol. 1946, 42: 81-124. 20

ACS Paragon Plus Environment

Page 21 of 28

1 2 3 4 532 5 533 6 534 7 535 8 9 536 10537 11 538 12 13539 14540 15541 16 542 17 18543 19544 20545 21546 22 23547 24548 25549 26550 27 28551 29552 30553 31554 32 555 33 34556 35557 36558 37559 38 39560 40561 41562 42 563 43 564 44 45565 46566 47 48567 49568 50569 51570 52 53571 54572 55573 56574 57 575 58 59576 60577

578

Environmental Science & Technology

(32) Linn, R. K. New Idria Mining District, In Ore Deposits of the United States 1933-1967. AIME: New York, 1968, Part 11, Chapter 78, pp 1623-1649. (33) Holmes, G. H. Mercury in California. In Mercury Potential of the United States. U.S. Department of the Interior, Bureau of Mines, Information Circular 8252, Washington, DC, 1965, pp 87-206. (34) Boctor, N. Z.; Shieh, Y. N.; Kullerud, G. Mercury ores from the New Idria mining district, California: geochemical and stable isotope studies. Geochim. Cosmochim. Acta 1987, 51: 1705-1715. (35) EPA Expanded Site Inspection Report. New Idria Mercury Mine, San Benito County, California. United States Environmental Protection Agency, 2010, EPA ID No.: CA0001900463. (36) EPA New Idria Mercury Mine. EPA finalizes site on the NPL. Begins interim cleanup actions. United States Environmental Protection Agency, Region 9, Public information, October 2011. (37) Kim, C. S.; Rytuba, J. J.; Brown, Jr., G. E. Geological and anthropogenic factors influencing mercury speciation in mine wastes: an EXAFS spectroscopy study. Appl. Geochem. 2004, 19: 379– 393. (38) Jew, A. D.; Kim, C. S.; Rytuba, J. J.; Gustin, M. S.; Brown, Jr. G. E. New technique for quantification of elemental Hg in mine wastes and its implications for mercury evasion into the atmosphere. Environ. Sci. Technol. 2011, 45: 412-417. (39) Smith, C. N.; Kesler, S. E.; Blum, J. D.; Rytuba, J. J. Isotope geochemistry of mercury in source rocks, mineral deposits and spring deposits of the California Coast Ranges, USA. Earth Planet. Sci. Lett. 2008, 269: 398–406. (40) Gehrke G. E.; Blum J. D.; Marvin-DiPasquale M. Sources of mercury to San Francisco Bay surface sediment as revealed by mercury stable isotopes, Geochim. Cosmochim. Acta 2011, 75: 691705. (41) Bloom N.; Preus E.; Katon J.; Hiltner M. Selective extractions to assess the biogeochemically relevant fractionation of inorganic mercury in sediments and soils. Anal. Chim. Acta 2003, 479: 233– 248. (42) Wiederhold, J. G.; Cramer, C. J.; Daniel, K.; Infante, I.; Bourdon, B.; Kretzschmar, R. Equilibrium mercury isotope fractionation between dissolved Hg(II) species and thiol-bound Hg. Environ. Sci. Technol. 2010, 44: 4191–4197. (43) Jiskra, M.; Wiederhold, J. G.; Bourdon, B.; Kretzschmar, R. Solution speciation controls mercury isotope fractionation of Hg(II) sorption to goethite. Environ. Sci. Technol. 2012, 46, 6654−6662. (44) Foucher, D.; Hintelmann, H. High-precision measurement of mercury isotope ratios in sediments using cold-vapor generation multi-collector inductively coupled plasma mass spectrometry. Anal. Bioanal. Chem. 2006, 384: 1470-1478. (45) Blum, J. D.; Bergquist, B. A. Reporting of variations in the natural isotopic composition of mercury. Anal. Bioanal. Chem. 2007, 388: 353-359. (46) Coplen, T. B. Guidelines and recommended terms for expression of stable-isotope-ratio and gasratio measurement results. Rapid Commun. Mass Spectrom. 2011, 25 (17), 2538−2560. (47) Schuette, C. N. Stack-loss determination, New Idria, Feb. 1931. Letter to W.R. Moorehead. Manager, New Idria Quicksilver Mines, Inc., 1931. (48) Gustin, M.S. Exchange of mercury between the atmosphere and terrestrial ecosystems. In Environmental Chemistry and Toxicology of Mercury, Liu, G., Cai, Y., O'Driscoll, N. J., Eds.; John Wiley & Sons, 2012, pp 423-451. (49) Navarro, A.; Cardellach, E.; Corbella, M. Mercury mobility in mine waste from Hg-mining areas in Almería, Andalusia (SE Spain). J. Geochem. Explor. 2009, 101: 236–246. (50) Gray, J. E.; Plumlee, G. S.; Morman, S. A.; Higueras, P. L.; Crock, J. G.; Lowers, H. A.; Witten, M. L. In vitro studies evaluating leaching of mercury from mine waste calcine using simulated human body fluids. Environ. Sci. Technol. 2010, 44, 4782–4788. 21

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 579 5 580 6 581 7 582 8 583 9 10584 11585 12586 13 587 14 15588 16589 17590 18 19591 20592 21593 22594 23 595 24 25596 26597 27598 28 29599 30600 31601 32602 33 603 34 35604 36605 37606 38 607 39 40608 41609 42610 43 44611 45612 46613 47614 48 615 49 616 50 51617 52618 53619 54 55620 56621 57622 58 623 59 60624

625

Page 22 of 28

(51) Yin, R.; Feng, X.; Wang, J.; Bao, Z.; Yu, B.; Chen, J. Mercury isotope variations between bioavailable mercury fractions and total mercury in mercury contaminated soil in Wanshan Mercury Mine, SW China. Chem. Geol. 2013, 336: 80-86. (52) Wiederhold, J. G.; Kraemer, S. M.; Teutsch, N.; Borer, P. M.; Halliday, A. N.; Kretzschmar, R. Iron isotope fractionation during proton-promoted, ligand-controlled, and reductive dissolution of goethite. Environ. Sci. Technol. 2006, 40: 3787-3793. (53) Wiederhold, J. G.; Teutsch, N.; Kraemer, S. M.; Halliday, A. N.; Kretzschmar, R. Iron isotope fractionation in oxic soils by mineral weathering and podzolization. Geochim. Cosmochim. Acta 2007, 71: 5821-5833. (54) Issaro, N.; Abi-Ghanem, C.; Bermond, A. Fractionation studies of mercury in soils and sediments: A review of the chemical reagents used for mercury extraction. Anal. Chim. Acta 2009, 631: 1-12. (55) Bacon, J. R.; Davidson, C. M. Is there a future for sequential chemical extraction? Analyst 2008, 133: 25-46. (56) Hall, G. E. M.; Pelchat, P.; Percival, J. B. The design and application of sequential extractions for mercury, Part 1. Optimization of HNO3 extraction for all non-sulphide forms of Hg. Geochem. Explor. Environ. Anal. 2005, 5: 107-113. (57) Estrade, N.; Carignan, J.; Sonke, J. E.; Donard, O. F. X. Mercury isotope fractionation during liquid-vapor evaporation experiments. Geochim. Cosmochim. Acta 2009, 73: 2693-2711. (58) Ghosh, S.; Schauble, E. A;, Lacrampe Couloume, G.; Blum, J. D.; Bergquist, B. A. Estimation of nuclear volume dependent fractionation of mercury isotopes in equilibrium liquid–vapor evaporation experiments. Chem. Geol. 2013, 336: 5-12. (59) Young, E. D.; Nagahara, H.; Mysen, B. O.; Audet, D. M. Non-Rayleigh oxygen isotope fractionation by mineral evaporation: Theory and experiments in the system SiO2. Geochim. Cosmochim. Acta 1998, 62: 3109-3116. (60) von Blanckenburg, F.; von Wirén, N.; Guelke, M.; Weiss, D. J.; Bullen, T. D. Fractionation of metal stable isotopes by higher plants. Elements 2009, 5, 375-380. (61) Hindshaw, R. S.; Reynolds, B. C.; Wiederhold, J. G.; Kiczka, M.; Kretzschmar, R.; Bourdon, B. Calcium isotope fractionation in alpine plants. Biogeochem. 2013, 112: 373–388. (62) Kiczka, M.; Wiederhold, J. G.; Kraemer, S. M.; Bourdon, B.; Kretzschmar, R. Iron isotope fractionation during Fe uptake and translocation in alpine plants. Environ. Sci. Technol. 2010, 44: 6144–6150. (63) Jouvin, D.; Weiss, D. J.; Mason, T. F. M.; Bravin, M.; Louvat, P.; Zhao, F. J.; Ferec, F.; Hinsinger, P.; Benedetti, M. F. Stable isotopes of Cu and Zn in higher plants: evidence for Cu reduction at the root surface and two conceptual models for isotope fractionation processes. Environ. Sci. Technol. 2012, 46, 2652−2660. (64) Lindberg, S. E.; Jackson, D. R.; Huckabee, J. W.; Janzen, S. A.; Levin, M. J.; Lund, J. R. Atmospheric emission and plant uptake of mercury from agricultural soils near the Almaden mercury mine. J. Environ. Qual. 1979, 8: 572-578. (65) Ericksen, J. A.; Gustin, M. S.; Schorran, D. E.; Johnson, D. W.; Lindberg, S. E.; Coleman, J. S. Accumulation of atmospheric mercury in forest foliage. Atmos. Environ. 2003, 37: 1613-1622. (66) Millhollen, A. G.; Gustin, M. S.; Obrist, D. Foliar mercury accumulation and exchange for three tree species. Environ. Sci. Technol. 2006, 40: 6001-6006. (67) Stamenkovic, J.; Gustin, M. S. Nonstomatal versus stomatal uptake of atmospheric mercury. Environ. Sci. Technol. 2009, 43: 1367-1372. (68) Lodenius, M. Use of plants for biomonitoring of airborne mercury in contaminated areas. Environ. Res. in press, doi: 10.1016/j.envres.2012.10.014. 22

ACS Paragon Plus Environment

Page 23 of 28

1 2 3 4 626 5 627 6 628 7 629 8 9 630 10631 11632 12 13633 14634 15635 16636 17 18637 19638 20639 21640 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Environmental Science & Technology

(69) Ghosh, S.; Xu, Y. F.; Humayun, M.; Odom, L. Mass-independent fractionation of mercury isotopes in the environment. Geochem. Geophys. Geosyst. 2008, 9: Q03004. (70) Demers, J. D.; Blum, J. D.; Zak, D. R. Mercury isotopes in a forested ecosystem: implications for air-surface exchange dynamics and the global mercury cycle. Global Biogeochem. Cycles 2013, 27, 117. (71) Yin, R;, Feng, X.; Meng, B. Stable Hg isotope variation in rice plants (Oryza sativa L.) from the Wanshan Hg mining district, SW China. Environ Sci Technol. 2013, 47: 2238-2245. (72) Carignan, J.; Estrade, N.; Sonke, J. E.; Donard, O. F. X. Odd isotope deficits in atmospheric Hg measured in lichens. Environ. Sci. Technol. 2009, 43: 5660–5664. (73) Estrade, N.; Carignan, J.; Donard, O. F. X. Isotope tracing of atmospheric mercury sources in an urban area of northeastern France. Environ. Sci. Technol. 2010, 44: 6062–6067. (74) Blum, J. D.; Johnson, M. W.; Gleason, J. D.; Demers, J. D.; Landis, M. S.; Krupa, S. Mercury concentration and isotopic composition of epiphytic tree lichens in the Athabasca oil sands region. Developments in Environmental Science 2012, 11: 373-390.

23

ACS Paragon Plus Environment

Environmental Science & Technology

2.5 unroasted ore waste [μg g-1]

2.0

calcine [μg g-1]

1.5

poplar leaves [ng g-1]

1.0 0.5

δ202Hg [‰]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38641 39 40 41642 42 43 643 44 45 46644 47 48645 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 28

0.0 -0.5 -1.0 -1.5 -2.0 -2.5 -3.0 -3.5 0

200

400

600 800 Hg concentration [µ µg g-1 or ng g-1]

1000

Figure 1: Total Hg concentrations vs. δ202Hg in unroasted ore waste (grey diamonds), calcine (red squares) and poplar leaves (green triangles) from New Idria. Please note that concentrations for unroasted ore waste and calcine are in µg g-1 and for poplar leaves in ng g-1. Isotope data are reported relative to NIST-3133 and error bars for δ202Hg of ±0.10‰ (2SD) are smaller than symbol sizes.

24

ACS Paragon Plus Environment

Page 25 of 28

0.25 HgS ore

0.20

unroasted ore waste calcine

0.15

poplar leaves

0.10

∆199Hg [‰]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33646 34 35 36647 37 38 648 39 40 41649 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Environmental Science & Technology

0.05 0.00 -0.05 -0.10 -0.15 -4

-3

-2

-1

0

1

2

3

δ202Hg [‰] Figure 2: Hg isotope signatures in samples from New Idria with mass-dependent fractionation (δ202Hg) and mass-independent fractionation (∆199Hg) relative to NIST-3133. Error bars plotted for HgS ore indicate 2SD reproducibility for all samples.

25

ACS Paragon Plus Environment

Environmental Science & Technology

unroasted ore waste

calcine

1.0

relative fraction of total Hg

0.8

0.6

F3 F2 F1

0.4

0.2

C1 C2 C3 C4 C5a C5b C6

0.0

O1g O1r O2 O3 O4 O5a O5b O6 O7a O7b

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38650 39 40 41651 42 43652 44 45 46653 47 48654 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 28

Figure 3: Relative distribution of Hg pools in unroasted ore waste and selected calcine samples from New Idria based on operationally defined sequential extraction: F1 (0.1 M HAc / 0.01 M HCl), F2 (12 M HNO3), and F3 (aqua regia). Total Hg concentrations of the samples are given in the SI (Table S1).

26

ACS Paragon Plus Environment

C1

C3

C4

C6

C4 F1 C4 F2 C4 F3 C4 total calc. total

C6 F1 C6 F2 C6 F3 C6 total calc. total

O4

C3 F1 C3 F2 C3 F3 C3 total calc. total

3.0 2.5 2.0

δ202Hg [‰]

1.5 1.0 0.5 0.0 -0.5 -1.0

O4 F1 O4 F2 O4 F3 O4 total calc. total

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32655 33 34 35656 36 37657 38 39658 40 41 42659 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Environmental Science & Technology

C1 F1 C1 F2 C1 F3 C1 total calc. total

Page 27 of 28

Figure 4: Hg isotope signatures (MDF) in sequential extracts of one unroasted ore waste sample (O4) and four calcine samples (C1, C3, C4, C6) from New Idria. Total digest data (total) are plotted for comparison. The isotope mass balance (calc. total) is based on the relative pool sizes of the three fractions. Error bars indicate 2SD reproducibility (for calc. total with error propagation).

27

ACS Paragon Plus Environment

Environmental Science & Technology

1 2 3 4 660 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31661 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 28

TOC/Abstract Art:

elemental Hg0

Hg isotopes: MDF + NVF secondary Hg(II) phases

Hg for sale

HgS ore

Hg0

calcination

HgS

~700 °C

incomplete Hg removal

28

ACS Paragon Plus Environment