mass-independent fractionation of ercury isotopes … · deposits (jensen and jensen 1990,...

143
MASS-INDEPENDENT FRACTIONATION OF MERCURY ISOTOPES IN FRESHWATER SYSTEMS by Carla Hendrika Rose A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Geology University of Toronto © Copyright by Carla H. Rose (2010)

Upload: others

Post on 13-May-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

MASS-INDEPENDENT FRACTIONATION OF

MERCURY ISOTOPES IN FRESHWATER

SYSTEMS

by

Carla Hendrika Rose

A thesis submitted in conformity with the requirements for the degree of Master of Applied Science

Graduate Department of Geology University of Toronto

© Copyright by Carla H. Rose (2010)

Page 2: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

ii

Mass-Independent Fractionation of Mercury Isotopes in Freshwater Systems

Carla H. Rose

Master of Applied Science

Graduate Department of Geology

University of Toronto

2010

ABSTRACT Mass-independent fractionation (MIF) of Hg isotopes has the potential to track the

environmental transport and fate of Hg. Herein we demonstrate that reducing both the

frequency and intensity of light have a large effect on the expression and magnitude of

MIF. This strongly supports the magnetic isotope effect as the mechanism behind MIF

observed during aqueous photo-reduction of Hg(II) and MeHg. The ratios of MIF,

199Hg/201Hg, were 1.00 ± 0.04 (2SE) for Hg(II) and 1.35 ± 0.16 (2SE) for MeHg

respectively and did not change as incident radiation energy and magnitude of MIF

diminished, suggesting the respective MIF pathways remained constant regardless of

experimental conditions. Comparable amounts of total photo-reduction were shown to

coincide with different magnitudes of MIF depending the wavelength light available for

photo-reduction. This confirms there are multiple pathways for photo-reduction in

freshwater reservoirs and indicates that quantitatively relating photo-reduction and MIF

will be challenging.

Page 3: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

iii

ACKNOWLEDGMENTS

I would like to thank my advisor, Prof. Bridget Bergquist, without whom this thesis would

have been impossible. I have learned a great deal from her over the past two years, and

deeply appreciate both her scientific insight and her personal concern for my success.

I would also like to thank Prof. Joel Blum, Marcus Johnson and the members of the U.M.

BEIGL laboratory for assistance and guidance in obtaining the isotope measurements

presented in this thesis and for their warm welcome and discussions when I was a guest in

their lab.

The members of my supervisory committee, Prof. Barbara Sherwood-Lollar and Prof. Grant

Ferris both welcomed me when I came to them with questions and broadened my perspective

of isotopes and biogeochemical cycling and I appreciate their help very much.

Dr. Sanghamitra Ghosh has been a wonderful colleague over the last couple of years and

offered help, discussions, moral support and generally a great attitude. Priyanka Chandan has

also been a great colleague and I have enjoyed working with her. Dr. Georges Lacrampe-

Couloume has been extremely helpful over the last two years and I am grateful for his

assistance and support.

Finally, Mom and Dad, I don’t know how I would have come this far without your

encouragement and support. Thank you a hundred times over and I love you very much.

Page 4: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

iv

TABLE OF CONTENTS ABSTRACT……………………………………………………………………….… ii AKNOWLEDGEMENTS……………………………………………………………… iii TABLE OF CONTENTS……………………………………………………………… iv LIST OF TABLES…………………………………………………………………… vi LIST OF FIGURES……………………………………………………………………vii CHAPTER 1: INTRODUCTION INTRODUCTION…………………………………………………………………… 1 RESEARCH OBJECTIVE……………………………………………………………… 7 THESIS OUTLINE…………………………………………………………………… 8 STATEMENT OF AUTHORSHIP AND PUBLICATION STATUS………………………… 10 LITERATURE CITED………………………………………………………………… 11 CHAPTER 2: OVERVIEW OF STABLE ISOTOPE GEOCHEMISTRY, FRESHWATER MASS

TRANSFER PROCESSES AND SPECIES TRANSFORMATIONS 2.1. MASS-DEPENDENT FRACTIONATION (CLASSICAL EFFECT)…………………… 17

2.1.1. Mechanisms of mass-dependent fractionation………………………… 18 2.1.2. Mercury isotopes and mass-dependent fractionation………………… 19

2.2. MASS-INDEPENDENT FRACTIONATION………………………………………… 25 2.2.1. Mercury isotopes and mass-independent fractionation……………… 25 2.2.2. Mechanisms of mass-independent fractionation……………………… 28

2.3. FRESHWATER MERCURY MASS TRANSFER PROCESSES AND SPECIES

TRANSFORMATIONS………………………………………………………… 32 2.3.1. Concentrations………………………………………………………… 32 2.3.2. Sources………………………………………………………………… 33 2.3.3. Sinks…………………………………………………………………… 34

2.3.3.1. Outflow…………………………………………………………… 34 2.3.3.2. Sedimentation…………………………………………………… 35 2.3.3.3. Bioaccumulation………………………………………………… 35 2.3.3.4. Degradation of MeHg…………………………………………… 36 2.3.3.5. Volatilization………………………………………………………38

2.3.4. Light Penetration……………………………………………………… 43 2.3.5. Organic Matter Complexation………………………………………… 45

LITERATURE CITED………………………………………………………………… 47

Page 5: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

v

CHAPTER 3: EFFECTS OF ULTRAVIOLET RADIATION ON THE MAGNETIC ISOTOPE

EFFECT IN FRESHWATER PHOTO-REDUCTION OF ORGANIC AND INORGANIC

MERCURY

3.1. INTRODUCTION…………………………………………………………………58 3.2. METHODS …………………………………………………………………… 62

3.2.1. Cleaning equipment…………………………………………………… 62 3.2.2. Photo-reduction experiments…………………………………………… 62 3.2.3. Hg concentration analysis ……………………………………………… 65 3.2.4. Matrix transfer in preparation for isotope analysis…………………… 65 3.2.5. MC-ICP-MS analysis…………………………………………………… 67 3.2.6. Reporting analytical uncertainty ……………………………………… 68 3.2.7. Calculations…………………………………………………………… 69

3.3. RESULTS……………………………………………………………………… 71 3.3.1. Concentrations ………………………………………………………… 71 3.3.2. Mass-dependent fractionation………………………………………… 73 3.3.3. Mass-independent fractionation………………………………………… 75

3.4 DISCUSSION …………………………………………………………………… 78 3.4.1. Relationship between MIF and light available ………………………… 78 3.4.2. MIF and Hg(II) ………………………………………………………… 78 3.4.3. MIF and MeHg………………………………………………………… 83 3.4.4. Mass-dependent fractionation………………………………………… 87 3.4.5. Implications for the natural world ……………………………………… 87

3.5 CONCLUSIONS ………………………………………………………………… 92 FIGURES…………………………………………………………………………… 93 TABLES…………………………………………………………………………… 102 LITERATURE CITED ……………………………………………………………… 110 CHAPTER 4: MASS-INDEPENDENT FRACTIONATION DURING CHEMICAL REDUCTION

OF AQUEOUS INORGANIC MERCURY IN THE ABSENCE OF LIGHT 4.1. INTRODUCTION…………………………………………………………………116 4.2. METHODS…………………………………………………………………… 117 4.3. RESULTS……………………………………………………………………… 118 4.4. DISCUSSION…………………………………………………………………… 118 4.5. CONCLUSION………………………………………………………………… 121 FIGURES…………………………………………………………………………… 122 TABLE……………………………………………………………………………… 125 LITERATURE CITED ………………………………………………………………… 126 CHAPTER 5: CONCLUSIONS 4.1. CONCLUSIONS………………………………………………………………… 129 LITERATURE CITED ………………………………………………………………… 131 APPENDIX A.1. TABLE OF ALL PHOTOEXPERIMENTS AND ANALYTICAL STATUS…………… 133

Page 6: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

vi

LIST OF TABLES Table 2.1. Isotopic abundances of mercury ………………………………………21 Table 2.2. Typical Hg concentrations in natural, unpolluted freshwaters………… 32 Table 2.3. Example depths of 99% attenuation of solar radiation at various

DOC concentrations………………………………………………… 44 Table 3.1. Summary of solar radiation measurements…………………………… 102 Table 3.2. Isotopic Data for Hg(II) aqueous photo-reduction with UV filtration…103 Table 3.3. Summary of subsrate losses through photo-reduction, isotopic

signatures and fractionation factors…………………………………… 105 Table 3.4. Isotopic Data for MeHg aqueous photo-reduction with UV filtration…106 Table 3.5. Summary of substrate losses through photo-reduction, isotopic

signatures and fractionation factors…………………………………… 108 Table 3.6. Sample recoveries after matrix transfer, all experiments…………… 109 Table 3.7. Calculated properties of CH3HgL complexes, from Tossell 1998…… 85 Table 4.1. Isotopic Data for Hg(II) dark reduction with SnCl2 in the presence

of organic matter……………………………………………………… 125

Page 7: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

vii

LIST OF FIGURES Figure 3.1. Schematic of the photochemical reaction apparatus………………… 93 Figure 3.2. Schematic of the phase separator…………………………………… 94 Figure 3.3. Mercury concentration in the substrate reservoir over time for

Hg(II) and MeHg. …………………………………………………… 95 Figure 3.4. Mass-dependent isotopic fractionation (202Hg) of the substrate

reservoir for Hg(II) and MeHg……………………………………… 96 Figure 3.5. Mass-independent isotopic fractionation (MIF, 199Hg), observed

in odd isotopes only, for Hg(II) photo-reduction experiments……… 97 Figure 3.6. Mass-independent isotopic fractionation (MIF, 199Hgas a

function of Hg(II) remaining………………………………………… 98 Figure 3.7. Mass-independent isotopic fractionation (MIF, 199Hg), observed

in odd isotopes only, for MeHg photo-reduction experiments……… 99 Figure 3.8. Mass-independent isotopic fractionation (MIF, 199Hg), observed

in odd isotopes only, as a function of MeHg remaining…………… 100 Figure 3.9. 201Hg versus 199Hg for photochemical reduction of Hg(II)

and MeHg…………………………………………………………… 101 Figure 4.1. Mercury concentration over time for dark chemical reduction

of Hg(II) by SnCl2……………………………………………………122 Figure 4.2. Mass-dependent isotopic fractionation (202Hg) as a function of Hg

remaining for dark chemical reduction of Hg(II) by SnCl2………… 123 Figure 4.3. Mass-independent isotopic fractionation (199Hg), as a function of

Hg(II) remaining for dark chemical reduction of Hg(II) by SnCl2 … 124

Page 8: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

1

CHAPTER 1: INTRODUCTION

This thesis is principally concerned with understanding and quantifying the effects of

frequency and intensity of sunlight on the magnitude and signature of mass-independent

fractionation (MIF) that occurs through photo-reduction of mercury (Hg). MIF in Hg

isotopes has been observed in both laboratory demonstrations (Bergquist and Blum 2007,

Estrade et al. 2009, Zheng and Hintelmann 2009, 2010a, 2010b, Malinovsky et al. 2010;

Wiederhold et al., 2010; Sherman et al., 2010) and natural samples (Bergquist and Blum

2007, Bergquist and Blum 2009, Ghosh et al. 2008, Biswas et al. 2008, Jackson et al.

2008, Gantner et al. 2009, Laffont et al. 2009, Sherman et al. 2009). Mass-independent

effects, and stable isotope behavior of Hg in general, have great potential to improve our

understanding of biogeochemical cycling of Hg in nature.

Mercury has always circulated in the atmosphere, oceans and terrestrial surface. Natural

emissions originate from volcanic activity (as Hg(II)P and Hg(0)g, Patterson and Settle

1987, Zambardi et al. 2009), geothermal activity (Hg(0) from fumaroles and hot springs,

Gustin et al. 2008), halos of Hg enrichment in deposits surrounding hot springs and

sulfide ore bodies, (Varekamp and Buseck 1984) and emissions from land areas that are

naturally enriched in Hg (see review Fitzgerald 2007). The majority of surface Hg

deposits are in the form of red and black cinnabar (HgS), which is poorly soluble in water

under standard conditions and mobilized chiefly by erosion. Based on sediment cores

taken from remote seepage lakes (Engstrom and Swain 1997 and references therein), peat

deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores

(Vandal et al. 1993, Schuster et al. 2002), anthropogenic emissions have increased the Hg

that circulates in the atmosphere, oceans and terrestrial surface by a factor of

approximately 3 since the industrial revolution. Mercury use in commercial products

(thermostats, batteries, fluorescent lighting) and processes (cement, metals, iron & steel

and chlor-alkali production) has peaked and is declining in much of the western world;

currently the largest source category of anthropogenic emissions worldwide is stationary

combustion of coal (and to a lesser extent fossil fuels) in which it is present as a trace

Page 9: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

2

element (Pacyna et al. 2010). Emissions in many of these categories have been growing

in Asia with China now being the largest global emitter due to its reliance on coal for

power and fuel (Jaffe and Strode 2008, Pacyna et al. 2010). Burial in sediments, in

particular deep-ocean sediments, is the ultimate sink for Hg but this process is far too

slow to counterbalance anthropogenic inputs (Fitzgerald and Lamborg 2007, Selin 2009).

The actively cycling lifetime of Hg in the atmosphere, terrestrial surface and ocean

surface is difficult to assess. Although the atmosphere contains less than one percent of

actively cycling Hg, it is the major avenue of global distribution. There is a great deal of

uncertainty around inputs to the atmosphere. Volcanic emissions, for example, are

inherently difficult to monitor and are spatially and temporally variable. They are

estimated by applying the few Hg/SO2 mass ratios available to SO2 concentrations

measured for a given eruption (Gustin et al. 2008). This leads to estimates that vary by

over an order of magnitude, and the portion of global emissions recently assigned to

volcanoes ranges from less than five to over 20 percent (Zambardi et al. 2009, Selin

2009). The most recent inventory of primary anthropogenic emissions is by Pacyna et al.

2010, for the year 2005. They cite an uncertainty in emissions by continent of at best

27% (North Amercia) and at worst 50% (Africa and South America). Mercury is emitted

to the atmosphere dominantly in three forms: volatile Hg(0), reactive gaseous Hg(II)

(RGM), and particulate Hg(II) (HgP). The atmospheric residence times of RGM and HgP

are days to weeks (Selin 2009), whereas the residence time for highly volatile Hg(0) is on

the order of 0.75 y (the uncertainty in this estimate is a factor of 2, Lindberg et al. 2007).

Most Hg in the lower atmosphere is therefore Hg(0) and well-mixed. Thus, a given

region may have important Hg contributions from distant sources as well as local ones.

For example, Sunderland et al. 2008 estimate that approximately half of the atmospheric

deposition to the Bay of Fundy area in the late 1990s was from global sources.

Deposition to terrestrial and ocean surfaces is thought to follow Hg(0) oxidation to Hg(II)

and scavenging by water droplets and/or particulate matter. Measurement networks exist

in some regions for wet deposition (snow and rain), but nowhere for dry deposition and

this is a significant gap in our understanding of Hg cycling (Lindberg et al. 2007, Selin

Page 10: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

3

2009). Vegetation and soils contain about 80% of actively cycling Hg. Some portion of

the Hg newly deposited to land is promptly revolatilized (“prompt recycling”, 5 to 60%

depending on the surface, Selin 2009 and references therein). The remainder is

incorporated into the soil and vegetation pools where it is released over longer timescales

(decades to centuries; “legacy mercury”) by chemical transformations or physical erosion

and runoff. It has been found that a portion of Hg newly deposited to water bodies

(which contain about 20% of actively cycling Hg) also revolatilizes promptly (Garcia, et

al. 2005, Poulain et al. 2006). To complicate the problem, revolatilization depends on

substrate concentration, matrix composition, light, temperature, precipitation, and for

soils, moisture content (Gustin et al. 2008). Assessing the relative importance of natural

versus anthropogenic sources of mercury in soils, water bodies and the atmosphere, and

also the fluxes between them, are therefore fundamental problems in understanding the

global Hg cycle. In addition, it is unlikely that Hg mobilized by natural biotic and abiotic

processes can be measured without interference of previously deposited Hg (“legacy

mercury”) of anthropogenic origin (Lindberg et al. 2007).

Although high or repeated exposure to different forms of Hg can have serious health

consequences, the most important toxicity risk that Hg poses to humans is as

methylmercury (MeHg) and exposure is mainly through consumption of fish. MeHg is a

neurotoxin and bioaccumulates, particularly in aquatic food webs. There is at least a

hundred-fold variation in Hg concentrations among species of fish and shellfish, but

concentrations in piscivorous fish can reach levels one million times greater than in the

surrounding water (Lindqvist et al. 1991, Mergler et al. 2007). The sites where MeHg

production is most intense in nature are wetlands, lake sediments and anoxic bottom

waters (Rudd 1995, Eckley and Hintelmann 2006). Since Hg is delivered to these areas

mainly as particulate and/or ionic species, biogeochemical cycling is a principle

modulator of mercury toxicity (Barkay and Wagner-Dobler 2005). Methylation appears

to be largely mediated by sulfur-reducing and iron-reducing bacteria (Compeau and

Bartha 1985, Gilmour et al. 1992, Selin 2009) and although Hg can be methylated

abiotically (Akagi and Takabatake 1973, Weber 1993), microbial production is sufficient

to account the amount of MeHg that is bioaccumulated (Fitzgerald and Lamborg 2007).

Page 11: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

4

Both the amount of bioavailable inorganic Hg and physiochemical factors that stimulate

microbial activity affect MeHg production (Barkay and Wagner-Dobler 2005 and

references therein). Characterizing the methylation and degradation capacities of natural

systems will clarify the relationship between continued anthropogenic loading and

contamination of food webs, including fish harvested for human consumption.

Stable isotope fractionation has been successfully used to trace biogeochemical processes

at both human and geologic timescales. The recent development of an analytical

technique using cold vapor multicollector inductively coupled plasma mass spectrometry

(CV-MC-ICP, Lauretta et al. 2001) now allows accurate measurements of Hg isotope

composition, and the reported range of isotopic signatures imparted by mass-dependent

fractionation (MDF) already exceeds 7‰ (typical uncertainty of ± 0.2 ‰, Bergquist and

Blum 2009, Sherman et al. 2010). In order to use the isotopes of Hg as a tool to track its

transport and fate in the environment, however, it is necessary to characterize source

signatures and, in particular, processes that induce fractionation.

Mass-independent fractionation (MIF) of Hg has been observed in some environmental

samples, such as snow (Sherman et al., 2010), fish and food webs (Bergquist and Blum

2007, Epov et al. 2008, Jackson et al. 2008, Gantner et al. 2009, Laffont et al. 2009),

mosses and lichens (Ghosh et al. 2008, Carignan et al. 2009, Bergquist and Blum 2009),

sediments, soils and peats (Foucher and Hintelmann 2006, Biswas al. 2008, Gehrke et al.

2009, Feng et al. 2010) and hydrothermal ores (Sherman et al. 2009; 0.13 ± 0.06‰) but

not in others, for example hydrothermal deposits and ores (Hintelmann and Lu 2003,

Smith et al. 2005, 2008, Sherman et al. 2009), chondrites (Lauretta et al. 2001), volcanic

emissions (Zambardi et al. 2009), and fly ash from waste incineration (Estrade et al.

2009). MIF has also been associated with some abiotic transformative processes, such as

photo-reduction of Hg(II) (Bergquist 2007, Zheng 2009, 2010b, Sherman 2010) and

photo-degradation of MeHg (Bergquist 2007, Malinovsky 2010) and volatilization from a

liquid phase (Estrade 2009). It has been observed not to occur in other processes, such as

hydrothermal transport (Smith 2005, 2008), volatilization from an aqueous phase (Zheng

et al. 2007), oxidative condensation of Hg(0) to Hg(II)P in volcanic plumes (Zambardi

Page 12: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

5

2009) and biological transformations such as demethylation and Hg(II) reduction (Kritee

et al. 2007, 2008, 2009). The Hg samples displaying large MIF to date (>1‰) are snow,

fish, food web components and moss and may be assumed to have undergone previous

photochemical transformation. For those sediments, hydrothermal deposits, soils and

peats that display small but significant MIF, previous photochemical transformation of

the Hg that is preserved in them is likely, although other MIF pathways might need to be

considered.

Reliable measurements of Hg isotopic composition are only a decade old with the first

report published in 2001 (Lauretta et al. 2001). Thus Hg isotope is a young field, and

there are many processes left to investigate and many natural archives that have yet to be

sampled. With the evidence available to date, the emerging picture is that MIF (greater

than approximately 0.2 ‰) is particular to photochemical transformations. MIF should

not be altered by mass-dependent fractionation (MDF). This is not to say that MIF

signatures are unsusceptible to alteration; they can be changed by other MIF processes or

diluted by the mixing of Hg pools bearing different MIF signatures (Bergquist and Blum

2009). Understanding the factors that control MIF will permit assessment of which

signatures are susceptible to alteration by successive transformations or mixing and

whether the various contributions to MIF can be deciphered. This is a prerequisite for

using MIF signatures to shed light on the transformations and fate of environmental Hg.

The geochemical processes investigated in this thesis are photochemical reduction of

inorganic Hg(II) and photochemical degradation of MeHg in freshwater aqueous

reservoirs. Two plausible mechanisms, the nuclear volume effect (NVE) and the

magnetic isotope effect (MIE), have been put forward to explain mass-independent

anomalies in Hg isotopes. The most probable mechanism behind the large MIF

signatures observed in Hg is the MIE, which can influence the rates of chemical reactions

that are electron spin selective. Photo-reduction and photo-degradation have been

demonstrated to produce relatively large MIF (> 2 ‰) in lab experiments, and the

photochemistry involved is characterized by radical intermediates. Radical pair reactions

are spin-selective and provide a venue for the manifestation of the MIE. Investigating the

Page 13: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

6

factors that affect MIF in photo-reduction and photo-degradation will thus shed light on

MIF processes in general. In addition, photo-reduction represents a significant pathway

by which the Hg burden of aqueous reservoirs is lost. Likewise, photo-degradation of

MeHg represents a significant portion of the demethylation capacity of ecosystems. If

MIF or some aspect of it is particular to these pathways among others, then it could

provide a method of identifying and quantifying them absolutely or relative to other

reduction and demethylation processes respectively. The extent to which MIF is

characterized by photo-reduction and photo-demethylation respectively will determine

how useful it is in quantifying these processes.

Page 14: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

7

RESEARCH OBJECTIVES:

While the MIE is the most probable cause of the large kinetic MIF observed in the

experimental results of Bergquist and Blum 2007, the only direct test of the mechanism

to date has been by Malinovsky et al. 2009. These investigators found that MIF during

photo-demethylation of MeHg was suppressed in solutions with ascorbic acid and with

ammonia, attributed to suppression or scavenging of the radical pair during

decomposition. Based on the explanation above, the efficiency of the MIE should be

affected by several aspects of the reaction pathway. The frequency of light required to

transform a molecule or complex into a radical pair will depend on the nature and energy

of the bond that needs to be broken. This means there will be an energy threshold below

which radiation cannot effect electronic transition to a radical pair. Bond energy is

affected by the nature of the ligand, and by solution properties such as pH and ionic

strength. The chemical composition of the solution can affect the formation and fate of

the radicals, as demonstrated by Malinovsky et al. Varying the intensity of radiation may

also affect the magnitude of MIF expression

One of the more straightforward tests of the mechanism to perform is elimination of the

photo-generated radical pair. If the higher energy wavelengths of the solar spectrum are

removed gradually, then at some point the incident radiation will not have enough energy

to excite an Hg complex to form a radical pair. This closes the avenue of expression for

the MIE.

Objectives:

To determine whether removing the UVA or the UVB portions of the electromagnetic

spectrum affects the expression and magnitude of the MIF observed during the aqueous

photo-reduction of MeHg and Hg2+ respectively.

Does the MIF allow us to distinguish and quantify photoreductive processes from other

process that reduce Hg in freshwater aqueous environments?

Page 15: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

8

THESIS OUTLINE

CHAPTER 1: INTRODUCTION. This chapter introduces some of the outstanding problems in mercury research and the

potential of stable isotope geochemistry to elucidate them. In particular, it introduces the

importance of a mechanistic understanding of mass-independent fractionation (MIF) as a

prerequisite for using mass-independent isotopic signatures to characterize fluxes and

transformations of mercury in the environment.

CHAPTER 2: OVERVIEW OF MERCURY ISOTOPES AND FRESHWATER MASS TRANSFER

PROCESSES AND PHOTOCHEMISTRY This chapter provides background information on mercury stable isotopes as well as on

processes of physical transport and species transformations of Hg in freshwaters.

CHAPTER 3: EFFECTS OF ULTRAVIOLET RADIATION ON THE MAGNETIC ISOTOPE

EFFECT IN FRESHWATER PHOTO-REDUCTION OF ORGANIC AND INORGANIC MERCURY This chapter addresses the objective of this thesis. It describes experiments in which

quartz reactors containing Hg(II) or MeHg in aqueous solutions with organic matter

were simultaneously exposed to three different light regimes to examine the effect of

photo-reduction on the isotopic fractionation of the respective Hg species. It demonstrates

that limiting the energy of incident radiation has a pronounced effect on MIF, supporting

the magnetic isotope effect (MIF) as the mechanism behind the MIF observed during

aqueous photo-reduction of Hg species. It also demonstrates that comparable amounts of

total photo-reduction may coincide with different amounts of MIF depending on the

wavelengths of light available for photo-reduction. This will make quantitatively relating

photo-reduction and MIF difficult.

CHAPTER 4: MASS-INDEPENDENT FRACTIONATION DURING CHEMICAL REDUCTION OF

AQUEOUS INORGANIC MERCURY IN THE ABSENCE OF LIGHT This chapter describes an experiment where Hg(II) is chemically reduced in the dark. It

is not directly related to the objectives of this thesis and experimental conditions were

significantly different. It demonstrates that during the process of dark chemical reduction

of aqueous Hg(II) by SnCl2 (in the presence of organic matter) and subsequent evasion of

the Hg(0), mass-independent anomalies in isotopic composition are produced.

Page 16: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

9

Attributing these effects to a particular mechanism or combination of mechanisms is not

possible without further investigation.

Page 17: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

10

STATEMENT OF AUTHORSHIP AND PUBLICATION STATUS

CHAPTER 1: INTRODUCTION Contributors: Carla Rose, Bridget Bergquist. Contributions: Research and writing was carried out by CR with guidance and editing by BB. Publication Status: Not submitted for publication. CHAPTER 2: OVERVIEW OF MERCURY ISOTOPES AND FRESHWATER

PHOTOCHEMISTRY CHAPTER 3: EFFECTS OF ULTRAVIOLET RADIATION ON THE MAGNETIC ISOTOPE

EFFECT IN FRESHWATER PHOTO-REDUCTION OF ORGANIC AND INORGANIC MERCURY Contributors (in alphabetical order other than first author): Carla Rose, Bridget Bergquist and Joel Blum. We thank Marcus Johnson, Sanghamitra Ghosh and Georges Lacrampe-Couloume for discussions and experimental assistance. Contributions: Photoexperiments were carried out by CR. Isotopic analysis were carried out both by MJ and by CR under close supervision of MJ. These analyses were funded by an NSF Grant to JB. Isotopic analysis of additional dark controls will be carried under supervision of SG. Data interpretation and writing was done by CR with input from BB. Publication Status: Not submitted for publication CHAPTER 4: MASS-INDEPENDENT FRACTIONATION DURING CHEMICAL REDUCTION OF

AQUEOUS INORGANIC MERCURY IN THE ABSENCE OF LIGHT Contributors (in alphabetical order other than first author): Carla Rose, Bridget Bergquist, Joel Blum and Marcus Johnson. We thank Sanghamitra Ghosh for discussions. Contributions: Photoexperiments were carried out by CR. Isotopic analysis were carried out by MJ and funded by an NSF Grant to JB. Data interpretation and writing was done by CR with input from BB. Publication Status: Not submitted for publication.

Page 18: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

11

LITERATURE CITED Akagi, H. and Takabatake, E., 1973. Photochemical formation of methylmercuric

compounds from mercuric acetate. Chemosphere 3, 131-133.

Barkay, T. and Wagner-Dobler, I., 2005. Microbial transformations of mercury: Potentials, challenges, and achievements in controlling mercury toxicity in the environment, Advances in Applied Microbiology, Vol 57.

Bergquist, B. A. and Blum, J. D., 2007. Mass-dependent and -independent fractionation of Hg isotopes by photoreduction in aquatic systems. Science 318, 417-420.

Bergquist, R. A. and Blum, J. D., 2009. The Odds and Evens of Mercury Isotopes: Applications of Mass-Dependent and Mass-Independent Isotope Fractionation. Elements 5, 353-357.

Biswas, A., Blum, J. D., Bergquist, B. A., Keeler, G. J., and Xie, Z. Q., 2008. Natural Mercury Isotope Variation in Coal Deposits and Organic Soils. Environmental Science & Technology 42, 8303-8309.

Carignan, J., Estrade, N., Sonke, J. E., and Donard, O. F. X., 2009. Odd Isotope Deficits in Atmospheric Hg Measured in Lichens. Environmental Science & Technology 43, 5660-5664.

Compeau, G. C. and Bartha, R., 1985. SULFATE-REDUCING BACTERIA - PRINCIPAL METHYLATORS OF MERCURY IN ANOXIC ESTUARINE SEDIMENT. Applied and Environmental Microbiology 50, 498-502.

Eckley, C. S. and Hintelmann, H., 2006. Determination of mercury methylation potentials in the water column of lakes across Canada. Science of the Total Environment 368, 111-125.

Engstrom, D. R. and Swain, E. B., 1997. Recent declines in atmospheric mercury deposition in the upper Midwest. Environmental Science & Technology 31, 960-967.

Epov, V. N., Rodriguez-Gonzalez, P., Sonke, J. E., Tessier, E., Amouroux, D., Bourgoin, L. M., and Donard, O. F. X., 2008. Simultaneous determination of species-specific isotopic composition of Hg by gas chromatography coupled to multicollector ICPMS. Analytical Chemistry 80, 3530-3538.

Page 19: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

12

Estrade, N., Carignan, J., Sonke, J. E., and Donard, O. F. X., 2009. Mercury isotope fractionation during liquid-vapor evaporation experiments. Geochimica Et Cosmochimica Acta 73, 2693-2711.

Feng, X. B., Foucher, D., Hintelmann, H., Yan, H. Y., He, T. R., and Qiu, G. L., 2010. Tracing Mercury Contamination Sources in Sediments Using Mercury Isotope Compositions. Environmental Science & Technology 44, 3363-3368.

Fitzgerald, W. F. and Lamborg, C. H., 2007. Geochemistry of Mercury in the Environment. In: Holland, H. D. and Turekian, K. K. Eds.)Treatise on Geochemistry. Elsevier.

Foucher, D. and Hintelmann, H., 2006. High-precision measurement of mercury isotope ratios in sediments using cold-vapor generation multi-collector inductively coupled plasma mass spectrometry. Analytical and Bioanalytical Chemistry 384, 1470-1478.

Gantner, N., Hintelmann, H., Zheng, W., and Muir, D. C., 2009. Variations in Stable Isotope Fractionation of Hg in Food Webs of Arctic Lakes. Environmental Science & Technology 43, 9148-9154.

Garcia, E., Amyot, M., and Ariya, P. A., 2005. Relationship between DOC photochemistry and mercury redox transformations in temperate lakes and wetlands. Geochimica Et Cosmochimica Acta 69, 1917-1924.

Gehrke, G. E., Blum, J. D., and Meyers, P. A., 2009. The geochemical behavior and isotopic composition of Hg in a mid-Pleistocene western Mediterranean sapropel. Geochimica Et Cosmochimica Acta 73, 1651-1665.

Ghosh, S., Xu, Y. F., Humayun, M., and Odom, L., 2008. Mass-independent fractionation of mercury isotopes in the environment. Geochemistry Geophysics Geosystems 9.

Gilmour, C. C., Henry, E. A., and Mitchell, R., 1992. SULFATE STIMULATION OF MERCURY METHYLATION IN FRESH-WATER SEDIMENTS. Environmental Science & Technology 26, 2281-2287.

Gustin, M. S., Lindberg, S. E., and Weisberg, P. J., 2008. An update on the natural sources and sinks of atmospheric mercury. Applied Geochemistry 23, 482-493.

Hintelmann, H. and Lu, S. Y., 2003. High precision isotope ratio measurements of mercury isotopes in cinnabar ores using multi-collector inductively coupled plasma mass spectrometry. Analyst 128, 635-639.

Page 20: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

13

Jackson, T. A., Whittle, D. M., Evans, M. S., and Muir, D. C. G., 2008. Evidence for mass-independent and mass-dependent fractionation of the stable isotopes of mercury by natural processes in aquatic ecosystems. Applied Geochemistry 23, 547-571.

Jaffe, D. and Strode, S., 2008. Sources, fate and transport of atmospheric mercury from Asia. Environmental Chemistry 5, 121-126.

Jensen, A. and Jensen, A., 1990. Historical deposition rates of mercury in Scandinavia estimated by dating and measurement of mercury in cores of peat bogs. Water Air Soil Pollution 56, 769-777.

Kritee, K., Barkay, T., and Blum, J. D., 2009. Mass dependent stable isotope fractionation of mercury during mer mediated microbial degradation of monomethylmercury. Geochimica Et Cosmochimica Acta 73, 1285-1296.

Kritee, K., Blum, J. D., and Barkay, T., 2008. Mercury Stable Isotope Fractionation during Reduction of Hg(II) by Different Microbial Pathways. Environmental Science & Technology 42, 9171-9177.

Kritee, K., Blum, J. D., Johnson, M. W., Bergquist, B. A., and Barkay, T., 2007. Mercury stable isotope fractionation during reduction of Hg(II) to Hg(0) by mercury resistant microorganisms. Environmental Science & Technology 41, 1889-1895.

Laffont, L., Sonke, J. E., Maurice, L., Hintelmann, H., Pouilly, M., Bacarreza, Y. S., Perez, T., and Behra, P., 2009. Anomalous Mercury Isotopic Compositions of Fish and Human Hair in the Bolivian Amazon. Environmental Science & Technology 43, 8985-8990.

Lauretta, D. S., Klaue, B., Blum, J. D., and Buseck, P. R., 2001. Mercury abundances and isotopic compositions in the Murchison (CM) and Allende (CV) carbonaceous chondrites. Geochimica Et Cosmochimica Acta 65, 2807-2818.

Lindberg, S., Bullock, R., Ebinghaus, R., Engstrom, D., Feng, X. B., Fitzgerald, W., Pirrone, N., Prestbo, E., and Seigneur, C., 2007. A synthesis of progress and uncertainties in attributing the sources of mercury in deposition. Ambio 36, 19-32.

Lindqvist, O., Johansson, K., Aastrup, M., Andersson, A., Bringmark, L., Hovsenius, G., Hakanson, L., Iverfeldt, A., Meili, M., and Timm, B., 1991. MERCURY IN THE SWEDISH ENVIRONMENT - RECENT RESEARCH ON CAUSES, CONSEQUENCES AND CORRECTIVE METHODS. Water Air and Soil Pollution 55, R11-&.

Page 21: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

14

Malinovsky, D., Latruwe, K., Moens, L., and Vanhaecke, F., 2010. Experimental study of mass-independence of Hg isotope fractionation during photodecomposition of dissolved methylmercury. Journal of Analytical Atomic Spectrometry 25, 950-956.

Martinez-Cortizas, A., Pontevedra-Pombal, X., Garcia-Rodeja, E., Novoa-Munoz, J. C., and Shotyk, W., 1999. Mercury in a Spanish peat bog: Archive of climate change and atmospheric metal deposition. Science 284, 939-942.

Mergler, D., Anderson, H. A., Chan, L. H. M., Mahaffey, K. R., Murray, M., Sakamoto, M., and Stern, A. H., 2007. Methylmercury exposure and health effects in humans: A worldwide concern. Ambio 36, 3-11.

Pacyna, E. G., Pacyna, J. M., Sundseth, K., Munthe, J., Kindbom, K., Wilson, S., Steenhuisen, F., and Maxson, P., 2010. Global emission of mercury to the atmosphere from anthropogenic sources in 2005 and projections to 2020. Atmospheric Environment 44, 2487-2499.

Patterson, C. C. and Settle, D. M., 1987. MAGNITUDE OF LEAD FLUX TO THE ATMOSPHERE FROM VOLCANOS. Geochimica Et Cosmochimica Acta 51, 675-681.

Poulain, A. J., Orihel, D. M., Amyot, M., Paterson, M. J., Hintelmann, H., and Southworth, G. R., 2006. Relationship to aquatic between the loading rate of inorganic mercury ecosystems and dissolved gaseous mercury production and evasion. Chemosphere 65, 2199-2207.

Rudd, J. W. M., 1995. SOURCES OF METHYL MERCURY TO FRESH-WATER ECOSYSTEMS - A REVIEW. Water Air and Soil Pollution 80, 697-713.

Schuster, P. F., Krabbenhoft, D. P., Naftz, D. L., Cecil, L. D., Olson, M. L., Dewild, J. F., Susong, D. D., Green, J. R., and Abbott, M. L., 2002. Atmospheric mercury deposition during the last 270 years: A glacial ice core record of natural and anthropogenic sources. Environmental Science & Technology 36, 2303-2310.

Selin, N. E., 2009. Global Biogeochemical Cycling of Mercury: A Review. Annual Review of Environment and Resources 34, 43-63.

Sherman, L. S., Blum, J. D., Johnson, K. P., Keeler, G. J., Barres, J. A., and Douglas, T. A., 2010. Mass-independent fractionation of mercury isotopes in Arctic snow driven by sunlight. Nature Geoscience 3, 173-177.

Page 22: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

15

Sherman, L. S., Blum, J. D., Nordstrom, D. K., McCleskey, R. B., Barkay, T., and Vetriani, C., 2009. Mercury isotopic composition of hydrothermal systems in the Yellowstone Plateau volcanic field and Guaymas Basin sea-floor rift. Earth and Planetary Science Letters 279, 86-96.

Smith, C. N., Kesler, S. E., Blum, J. D., and Rytuba, J. J., 2008. Isotope geochemistry of mercury in source rocks, mineral deposits and spring deposits of the California Coast Ranges, USA. Earth and Planetary Science Letters 269, 398-406.

Smith, C. N., Kesler, S. E., Klaue, B., and Blum, J. D., 2005. Mercury isotope fractionation in fossil hydrothermal systems. Geology 33, 825-828.

Sunderland, E. M., Cohen, M. D., Selin, N. E., and Chmura, G. L., 2008. Reconciling models and measurements to assess trends in atmospheric mercury deposition. Environmental Pollution 156, 526-535.

Vandal, G. M., Fitzgerald, W. F., Boutron, C. F., and Candelone, J. P., 1993. VARIATIONS IN MERCURY DEPOSITION TO ANTARCTICA OVER THE PAST 34,000 YEARS. Nature 362, 621-623.

Varekamp, J. C. and Buseck, P. R., 1984. THE SPECIATION OF MERCURY IN HYDROTHERMAL SYSTEMS, WITH APPLICATIONS TO ORE DEPOSITION. Geochimica Et Cosmochimica Acta 48, 177-185.

Weber, J. H., 1993. REVIEW OF POSSIBLE PATHS FOR ABIOTIC METHYLATION OF MERCURY(II) IN THE AQUATIC ENVIRONMENT. Chemosphere 26, 2063-2077.

Wiederhold, J. G., Cramer, C. J., Daniel, K., Infante, I., Bourdon, B., and Kretzschmar, R., 2010. Equilibrium Mercury Isotope Fractionation between Dissolved Hg(II) Species and Thiol-Bound Hg. Environmental Science & Technology 44, 4191-4197.

Yang, L. and Sturgeon, R., 2009. Isotopic fractionation of mercury induced by reduction and ethylation. Analytical and Bioanalytical Chemistry 393, 377-385.

Zambardi, T., Sonke, J. E., Toutain, J. P., Sortino, F., and Shinohara, H., 2009. Mercury emissions and stable isotopic compositions at Vulcano Island (Italy). Earth and Planetary Science Letters 277, 236-243.

Zheng, W., Foucher, D., and Hintelmann, H., 2007. Mercury isotope fractionation during volatilization of Hg(0) from solution into the gas phase. Journal of Analytical Atomic Spectrometry 22, 1097-1104.

Page 23: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

16

Zheng, W. and Hintelmann, H., 2009. Mercury isotope fractionation during photoreduction in natural water is controlled by its Hg/DOC ratio. Geochimica Et Cosmochimica Acta 73, 6704-6715.

Zheng, W. and Hintelmann, H., 2010a. Nuclear Field Shift Effect in Isotope Fractionation of Mercury during Abiotic Reduction in the Absence of Light. Journal of Physical Chemistry A 114, 4238-4245.

Zheng, W. and Hintelmann, H., 2010b. Isotope Fractionation of Mercury during Its

Photochemical Reduction by Low-Molecular-Weight Organic Compounds. Journal of Physical Chemistry A 114, 4246-4253.

Page 24: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

17

CHAPTER 2: OVERVIEW OF MERCURY

ISOTOPE GEOCHEMISTRY, FRESHWATER

MASS TRANFER PROCESSES AND SPECIES

TRANSFORMATIONS

2.1. MASS-DEPENDENT FRACTIONATION (CLASSICAL

EFFECT) The discovery and early study of multiple stable isotopes of elements happened in the

first decades of the twentieth century and is reviewed by Urey 1947. Some of the earliest

demonstrations of variable isotopic compositions of elements were of differences in the

vapor pressures of neon and of hydrogen (Keesom and van Dijk 1931, Urey et al. 1932).

The theoretical laws that describe how much of isotopic fractionation happens (i.e., the

mass-dependent laws) were also developed early (Urey 1947, Biegleisen 1949) and gave

rise to the field of traditional stable isotope geochemistry. This early isotope work

focused on lighter isotopic systems where the relative mass differences, and thus the

expected fractionation of the isotopes, were large. It also focused on elements that were

amenable to gas source mass spectrometry (i.e., H, C, N, O and S). The varying isotopic

composition of different reservoirs provided insight into, for example, historical climates

(the oxygen isotope paleotemperature scale, Urey et al. 1948), metabolism and reaction

mechanisms (the relative abundance of carbon isotopes in plants, Wickman 1952) and

fluxes in the meteoric water cycle (Craig 1961). With recent analytical developments and

improvements in precision, the isotopic variations of heavier elements (including Mg, Ca,

Fe, Cu, Zn, Mo, and Hg) is now being explored and has led to the development of a new

field called non-traditional stable isotope geochemistry.

Isotopic compositions in stable isotope geochemistry are generally reported as the per mil

(‰) deviation of the ratio of two isotopes from the same ratio observed in a standard

Page 25: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

18

reference material. This deviation is referred to as delta () notation. The lowest-mass

major isotope is usually used as the denominator:

otherX (‰) = (other/lowest majorRsampleother/lowest majorRstandard) – 1) x 1000

Relative differences in isotopic ratios can be determined with far greater precision than

absolute differences.

2.2.1. Mechanisms of mass dependent fractionation

Much of the observed variations in isotope ratios found in nature follow mass dependent

fractionation (MDF) laws. These were the first laws used to predict and describe isotope

fractionation (Urey 1947, Bigeleisen and Mayer 1947, Bigeleisen 1949). Mass-

dependent fractionation laws have also been derived in detail, as in Hulston and Thode

1965, Mariotti et al. 1981 and Young et al. 2002. Mass dependent fractionation occurs

in both equilibrium and kinetic reactions.

Equilibrium effects arise from the tendency of systems to adjust to minimize their energy.

The equilibrium constant of a chemical reaction is the difference in the standard free

energies of the reactants and product divided by RT. For an isotopic exchange reaction,

this is a small difference between large numbers that cannot be well-characterized by

classical models. The equilibrium constant for an isotopic exchange reaction can be

determined through a quantum mechanical consideration of two equilibrium constants:

those for each of the two isotopic molecules to dissociate into atoms. The magnitude of

these dissociation constants depends on the translational, rotational and vibrational

energies of the molecules. It is differences in the vibrational energy of the two molecules

at the quantum mechanical level that dominantly leads to equilibrium isotopic

fractionation, such that the heavy isotope is more stable in the molecule with stronger

bonds. The lighter isotope will preferentially be found as individual atoms and in weaker

bonds. The vibrational energy differences decrease as a function of temperature (1/T2) so

that at higher temperatures, isotope fractionation disappears (Bigeleisen and Mayer

1947). It is worth mentioning that chemical equilibrium does not guarantee isotopic

Page 26: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

19

equilibrium because the times required to achieve isotopic and chemical equilibria

respectively are not necessarily related. As an example, the chemical equilibrium

between CO2(g) and water is established more quickly that oxygen istotope equilibrium

between the two (Fritz and Fontes 1980).

Kinetic isotope effects are associated with processes that are fast, unidirectional or

incomplete (such as evaporation, diffusion and enzymatic reactions). Zero point energies

(the energy at which a molecule vibrates in its ground state) are lower for heavier

isotopes, so the energy gap between the ground state of a heavier isotope and its bond

dissociation energy is larger. Lighter isotopes require less energy to break chemical

bonds and frequently display slightly lower chemical stability. Lighter isotopes may also

have slightly greater translational velocities and diffusion coefficients. This also leads to

kinetic fractionation where transfer between compartments/reservoirs is incomplete or

unidirectional.

For both equilibrium and kinetic mass-dependent isotope effects, the relationship

between the mass difference of the isotopes (m) being studied and the magnitude of

fractionation is approximately linear (Hulston and Thode 1965, Bigeleisen 1949, Young

et al. 2002). For this reason, MDF effects are the largest for hydrogen and tend to

decrease. For the first few rows of the periodic table, m/m is generally 5 to 10% or more

but by row six (containing Hg), m/m is only 1 to 3%. The magnitude of isotopic

fractionation, however, does not decrease proportionately. Smith et al. 2005 note that all

elements from Fe to Hg showing unexpectedly large fractionation have multiple

oxidation states.

2.1.2. Mercury isotopes and mass-dependent fractionation.

Separation of Hg isotopes was achieved as early as 1920 by Bronsted and von Hevesy by

evaporation and subsequent condensation on a cold surface. In 1950, Alfred Nier

reported the relative isotopic abundances of a laboratory sample of Hg on a gas-source,

electron bombardment mass spectrometer using sample-standard bracketing and

calibration of machine mass discrimination using 36Ar and 40Ar. These values became

Page 27: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

20

the isotopic composition of Hg accepted by IUPAC until Zadnik et al. repeated the

measurements in 1989.

In the late 20th century, there was considerable interest in the cosmochemical behavior of

Hg and several investigations of Hg in the Muchison and Allende carbonaceous

chondrites (for example) were undertaken between 1970 and 2001. However these

studies were conducted by neutron activation analysis (NAA) and the measured

variations have been shown to be analytical artifacts (Nier and Schlutter 1986). Nier and

Schlutter 1986 obtained samples of the Allende meteorite and measured all Hg isotopes

relative to 202Hg using an electron bombardment gas-source mass spectrometer. They

failed to find anomalous abundance ratios relative to a laboratory standard for 196/202Hg

(uncertainty of ~ 20‰) or any of the other isotopes (uncertainty of ~ 1‰).

Lauretta et al. 2001 undertook measurements of both the Allende and Murchison

chondrites on a multi-collector inductively coupled plasma mass spectrometer (MC-ICP-

MS). The samples were reduced to Hg(0) online and introduced through a gas-liquid

separator (cold vapor generation; this produces a steady signal for isotope measurement),

along with a Tl aerosol to correct for machine mass bias and bracketed with a reference

standard. They found no significant variations in isotopic abundances between the

chondrites within their precision of 0.2 – 0.5‰, with the exception of 200/202Hg in the

Murchison which fell only slightly outside the analytical uncertainty.

The CV-MC-ICP-MS technique of Lauretta 2001 has become the standard method of

determining Hg isotope ratios. Blum and Bergquist 2007 proposed NIST SRM 3133 as a

universal reference standard and the method described above for reporting ratios, XXXHg/198Hg where XXX is the isotope of interest. These have generally been adopted,

and investigation of the isotopic variation in terrestrial Hg samples has been underway

for the last few years. Mercury has seven stable isotopes which span a relative mass

difference of 4% as summarized in Table 2.1.

Page 28: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

21

Table 2.1 Isotopic abundances of mercury

Atomic Mass Number

Relative Abundance (%)

196 0.16 198 10.0 199 16.9 magnetic 200 23.1 201 13.2 magnetic 202 29.7 204 6.8

The lowest-mass isotope (196Hg) has by far the lowest abundance (0.15 %) and is

generally not measured. Thus 198Hg is used as the denominator, and the accepted

reference standard has come to be NIST Standard Reference Material 3133 (Blum and

Bergquist 2007):

XXXHg (‰) = (XXX/198Rsample/XXX/198RSRM 3133) – 1) x 1000.

The other five isotopes can be measured conveniently by MC-ICP-MS. Isotopic shifts

are not as large as for lighter elements such as C, N and O, but with typical uncertainties

for 202Hg of 0.1 to 0.3 ‰ (2SD, Bergquist et al. 2009), isotopic differences in samples

can be readily resolved.

Mercury’s active redox chemistry, tendency to form covalent bonds, biological cycling,

phase changes and volatile form (Hg0) gives rise to many circumstances in which isotope

fractionation based on nuclear mass and diffusion coefficients may be expressed. A large

range of mass-dependent isotopic signatures, 202Hg, has been reported for hydrothermal

systems and ore (from -4 ‰ to nearly 3‰ , Hintelmann and Lu 2003, Smith et al. 2005,

2008, Sherman et al. 2009). Sherman et al. 2009 conclude that multiple processes

contribute to the variation in isotopic composition of such systems. Smith et al. 2005

identify the boiling of hydrothermal fluids as an important fractionating process. For

example, Hg mobilized from magma or country rock by epithermal systems travels

principally as Hg(0)aq in reduced, low sulfur fluids. Lighter isotopes of Hg(0) partition

preferentially into the vapor phase as the fluid ascends and boils. Isotopically heavy Hg

Page 29: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

22

is deposited in solid solution at depth, whereas isotopically light Hg(0) can mix with and

be oxidized by groundwater at the surface and precipitate with reduced sulfur as

isotopically light cinnabar. The range in 202Hg signatures resulting from hydrothermal

fractionation is much wider (for example, -3.5 ± 0.1 ‰ to 2.1 ± 0.1 ‰, Smith et al.

2005) than the -0.6 ± 0.2 ‰ that is typical of crustal rocks investigated so far (Bergquist

and Blum 2009, Smith et al. 2008).

Phase changes have been demonstrated to produce MDF in Hg. Estrade et al. 2009

investigated both equilibrium and kinetic volatilization of pure liquid mercury. In the

equilibrium experiments, the Hg vapor was slightly depleted in the heavier isotopes

(202Hgvapor was 0.20 ‰ less negative that 202Hgliquid) and was independent of

temperature in the range 2 to 22 ºC. Kinetic volatilization under a 10-5 bar vacuum

showed an increase in residual liquid 202Hg of 7.23 ± 0.22‰ for a fraction of Hg

remaining (fR ) = 0.3, a dramatic enrichment in the heavier isotopes. Zheng et al. 2007

tested free volatilization of aqueous Hg(0) (unassisted by sparging) by reducing a

solution of inorganic Hg(II) SnCl2 in a reactor while flushing the headspace with N2.

They report that volatilization was subject to Rayleigh-type mass-dependent fractionation

only, resulting in enrichment of the unvolatilized substrate in 202Hg by 1.5 ± 0.1 ‰

when fR had dropped to 0.04.

Reductive process, both biological and abiotic, have been demonstrated to result in mass-

dependent fractionation of Hg. Biological reduction of Hg(II) (Kritee et al., 2007, 2008)

has been reported to enrich the heavier isotopes in the unreacted substrate. Kritee et al.

2008 observed that MDF during reduction of Hg(II) to Hg(0) by various strains of

bacteria initially followed a Rayleigh distillation model. As the reaction proceeded,

however, MDF was suppressed to different degrees according to the microbial strain

studied. The authors hypothesized that with unlimited bioavailable Hg(II), reduction of

Hg(II) by MerA is the rate-limiting and single fractionation-determining step. As cell

densities grow and the Hg(II) supply shrinks, non-fractionating process such as Hg(II)

diffusion and cell uptake become rate-limiting. Reduction by MerA is then quantitiative

and fractionation is suppressed. Biological demethylation and reduction of MeHg also

Page 30: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

23

induces MDF. Kritee et al. 2009 observed preferential degradation of the lighter isotopes

of MeHg during the initial phases of mer-mediated microbial degradation. Complete

dampening of fractionation was observed as the reactions progressed, seemingly after cell

densities exceeded a certain level. The authors suggest that, as breakage of the organo-

Hg bond by MerB happens prior to and much more slowly than reduction of Hg(II) by

MerA, that catalysis by MerB was initially rate-limiting and fractionation determining but

that at higher cell concentrations, uptake of MeHg by the cell becomes rate-determining

and causes no significant fractionation. These examples demonstrate that the magnitude

of MDF during enzymatic reaction depends the reaction conditions. If the fractionating

step is rate-limiting, biological processes can produce significant changes in Hg isotopic

composition. If a previous step becomes rate-limiting, however, the fractionating step

will quantitatively transform all isotopes and no MDF will be observed.

Depending on reaction conditions, the product of reduction, Hg(0), may be isotopically

light and can be lost from the reaction medium through volatilization. As mentioned

above, light isotopes preferentially volatilize from an aqueous phase. Abiotic reduction

of aqueous Hg(II) and MeHg has also been shown to enrich the unreacted substrate in

202Hg. This includes photochemical processes in the presence of organic matter

(Bergquist and Blum 2007, Zheng and Hintelmann 2009, 2010b). Malinovsky et al. 2010

report MDF during photo-degradation of MeHg under UVC radiation in the absence of

organic matter, but it seems doubtful that this would occur in nature (see, for example,

Zhang and Hsu-Kim 2010). Abiotic reduction by organic matter occurs in the dark as

well, and has been shown to enrich the unreacted substrate in 202Hg (Bergquist and

Blum, 2007; Zheng and Hintelmann 2009). Chemical derivitization of Hg species (dark

chemical ethylation of Hg(II) by NaBEt4 , Yang and Sturgeon 2009) and biological

methylation (Rodriguez-Gonzalez et al. 2009) have both been shown to induce MDF.

Overall, it might be expected that Hg remaining in a cell, the reaction medium or the

aqueous phase generally will be isotopically heavy.

In contrast, oxidative condensation of Hg(0) as Hg(II) on particulates and acidic aerosols

in volcanic plumes shows the oxidized Hg(II)P is enriched in the heavy isotopes

Page 31: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

24

(Zambardi et al. 2009). If this process is representative of atmospheric oxidation of

Hg(0) in general, it may be expected that the Hg(II) deposed as Hg(II)P and Hg(II)aq will

be isotopically heavy and Hg(0) remaining in the atmosphere will be isotopically light.

Assessing the relative importance of reductive and oxidative processes, as well as of

phase changes, on Hg isotopes in nature is difficult at this early stage of research.

Natural archives show both positive and negative signatures. Zambardi et al. 2009 report

202Hg from -1.74 to 0.01‰ for fumarolic gasses. Biswas et al. 2008 report a range of -

2.8 to 0‰ in 202Hg for 30 coal deposits from the United States, China and Russia-

Kazakhstan. Lichen and moss show negative MDF to about -2 ‰ (Ghosh et al. 2008,

Carignan et al. 2009, Bergquist and Blum 2009). Soils, peat and sediments show larger

negative range than coal (as low as -4 ‰) but also exhibit positive signatures (with

202Hg up to 1 ‰ or more; Foucher and Hintelmann 2006, Biswas et al. 2008, Ghosh et

al. 2008, Foucher et al. 2009, Gehrke et al. 2009.) Fish and food web 202Hg signatures

range from -3 ‰ to nearly 2 ‰ (Bergquist and Blum 2007 and 2009, Epov et al. 2008,

Laffont et al. 2009, Gantner et al. 2009). Sherman et al. 2010 report signatures in Arctic

snow ranging from -0.49 to 0.70 ± 0.16‰.

Differentiating between natural and anthropogenic Hg emissions based on isotopic

compositions is difficult at present because of the limited measurements, the likely source

variability and also fractionation induced by processes after the point source (Bergquist

and Blum 2009). It is worth mentioning, however, that the various coal deposits

sampled by Biswas et al. 2008 were generally isotopically distinguishable. In addition,

Foucher et al. 2009 demonstrated that there is a well-resolved difference between

background Hg isotopic composition of Adriatic sea sediment (depleted in 202Hg) and

Hg derived originally from the Idrija region (depleted in 202Hg), which drains into the

Adriatic. One of the largest Hg mines in the world is located in the Idrija region, and

isotopic compositions can be used to track, differentiate and quantify its contributions to

Hg in the Adriatic. This is a well-constrained system with one dominant Hg source that

did not undergo much fractionation during transport, however, and the general usefulness

of MDF to track the transport and fate of anthropogenic Hg in the environment requires

further investigation.

Page 32: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

25

2.2. MASS-INDEPENDENT FRACTIONATION

Mass-independent fractionation (MIF) was first reported by Thiemens and Heidenreich in

1983, in the ozone produced by molecular oxygen in an electric plasma. It was

subsequently observed in other kinds of reactions, such as UV-photolysis (Thiemens &

Jackson 1987) and in ion-molecule encounters (Griffith & Gellene 1992). Mass-

independent isotopic signatures have since been helpful in problems such as deciphering

the reaction pathways of ozone (Schueler et a. 1990) and determining the historical

oxidative capacity of Earth (see review: Thiemens 2006). Small MIF (much less than 0.5

‰) has also been observed in several other elements, including Ti, Cr, Zn, Sr, Mo, Ru,

Cd, Sn, Te, Ba, Nd, Sm, Gd, Yb and U (see review: Fujii 2009).

MIF is the difference in the measured isotope ratio from the ratio predicted by mass-

dependent fractionation laws. It is reported in capital delta notation, and can be

approximated (for 5 ‰) by

XXXHg = XXXHg – (202Hg*XXX)

where is the kinetic or equilibrium factor applicable to the isotope of interest (see

Younge et al. 2002; see also the methods section of Chapter 3 of this document).

2.2.1. Mercury isotopes and mass-independent fractionation

MIF of Hg isotopes is observed in many natural samples and is only ever seen in the odd

isotopes. The most dramatic deviations are positive anomalies in food webs, with 199Hg

signatures in fish tissue reaching nearly 5‰ (Bergquist and Blum 2007, 2009, Epov et al.

2008, Jackson et al. 2008, Gantner et al. 2009, Laffont et al. 2009) and negative

anomalies, with 199Hg signatures lower than -5 ‰, in arctic snow (Sherman et al.

2010). When measuring MIF, typical analytical uncertainties are lower than 0.1 ‰,

2(Bergquist and Blum 2009) To date, O and S are the only other elements that have

displayed large degrees ( >1‰) of MIF in natural samples (Bergquist and Blum 2009).

Page 33: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

26

Tin, however, has been reported to undergo MIF during laboratory experiments, with

117Sn reaching nearly 5 ‰ (Malinovsky et. al 2009).

Hydrothermal deposits, ore, host rock and volcanic emissions rarely display MIF

(Hintelmann and Lu 2003, Smith et al. 2005, 2008, Sherman et al. 2009, Zambardi et al.

2009). Where they do, it is quite small. For example, Sherman et al. 2009 report small

but significant MIF at a hydrothermal site in Yellowstone (199Hg = 0.13 0.06 ‰) that

likely had some exposure to sunlight. Whereas, hydrothermal samples from a deep sea

floor rift showed no MIF. The authors suggested that a portion of the Hg in the

hydrothermal fluids from Yellowstone had been photochemically reduced in its

geological history. Biswas et al. 2009 report 199Hg in coal from 0.3 to -0.6 ± 0.04 ‰. In

contrast to the samples above, natural samples that are part of the surface pools of Hg that

undergo active biogeochmical cycling including photochemistry do display MIF. Soils,

sediments and peats display small negative MIF (less than 1 ‰, Foucher and Hintelmann

2006, Biswas et al. 2008, Ghosh et al. 2008, Jackson et al. 2008, Foucher et al. 2009,

Gehrke et al. 2009). Moss and Lichen display similar small, negative MIF (Ghosh et al.

2008, Carignan et al. 2009, Bergquist and Blum 2009).

The first demonstration of a process that induced MIF was photochemical reduction of

methylmercury (MeHg) and Hg(II) in the presence of dissolved organic matter

(Bergquist and Blum, 2007). Odd isotopes were preferentially retained in the unreacted

substrate, resulting in mass-independent anomalies of over 2 ‰. Examination of

199Hg/201Hg revealed that the ratios of each process were distinct: 1.00 ± 0.02 (2SE)

for of Hg(II) and 1.36 ± 0.03 (2SE) for of MeHg. The authors suggested that the ratio

might be unique to the pathway, which raised the possibility of using MIF signatures to

identify and quantify photochemical transformations in nature. They suggested the

mechanism causing this MIF was the magnetic isotope effect (MIE), which will be

discussed hereafter. MIF during photochemical reduction of Hg has also been reported

by Yang and Sturgeon 2009, Zheng and Hintelmann 2009, 2010b. Malinovsky et al.

2010 report MIF during photochemical reduction of MeHg, with a 199Hg/201Hg ratio of

1.28 ± 0.03 (2SE).

Page 34: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

27

Recently, Sherman et al. 2010 reported MIF during photochemical reduction of snow.

The direction of MIF is opposite aqueous photochemical reduction, with the odd isotopes

being depleted in the substrate snow. They note a 199Hg/201Hg ratio of 1.07 ± 0.04 ‰.

They are unsure about the mechanism, but suggest multi-step heterogeneous redox

reactions in the quasi-liquid layer on snow crystal surfaces as a starting point for

investigation.

While equilibrium process do not produce the large MIF signatures that kintetic photo-

processes do, there have been a few demonstrations of small (< 0.2 ‰) Hg MIF during

equilibrium processes. Equilbration of Hg metal with its vapor phase has been reported

to cause small MIF with 199Hg ranging from +0.10 to 0.14 ± 0.03 ‰ in the vapor phase

(Estrade et al. 2009) and from +0.13 to 0.31 ± 0. 05‰ (Ghosh et a., in prepartation). This

indicates the non-volatilized substrate was slightly depleted in the odd isotopes. Estrade

et al. obtained a ratio 199Hg/201Hg of 2.0 ± 0.6 and Ghosh et al. a ratio of 1.62 ±

0.06(2SE), which are distinct from the ratios obtained for photo-reductive processes by

Bergquist and Blum 2007 and Malinovsky et al., 2010. Equilibrium fractionation is

attributed to the nuclear volume effect (NVE), which will be discussed hereafter. Estrade

et al. also observed 199Hg ranging from -0.12 to -0.01 ± 0.03 ‰ in residual liquid

phase for kinetic volatilization of pure Hg. Weiderhold et al. 2010 report small MIF (<

0.1‰) during equilibration binding of aqueous Hg(II) with various thiols.

While volatilization of aqueous Hg(0) has been shown not to produce MIF in Hg (Zheng

and Hintelmann 2007), Zheng and Hintelmann 2010a report that dark chemical reduction

of Hg(II) with SnCl2 results in small but significant and consistent negative MIF in the

unreacted substrate. In contrast, dark chemical reduction of Hg(II) in the presence of

organic matter was not observed to induce MIF in the unreacted substrate, as reported by

Bergquist and Blum 2007 and Zheng and Hintelmann 2009. Zheng and Hintelmann 2009

do report a small, positive MIF anomaly in volatilized Hg trapped after dark abiotic

reduction, indicating the unreacted substrate was slightly depleted in the odd isotopes. It

should be noted that dark reduction in the presence of organic matter is a slow process,

Page 35: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

28

and investigators were considering reactions with a maximum of 20% loss of initial Hg.

The SnCl2 reduction, on the other hand, could be pushed to completion.

Several of the processes demonstrated to cause MDF do not cause MIF in Hg. Smith et

al. 2005, 2008 and Sherman et al., 2009, conclude that hydrothermal processes do not

induce MIF. None of the microbially mediated reduction or degradation processes

investigated by Kritee et al. 2007, 2008 or 2009 show any MIF and Kritee et al. 2009

present a detailed discussion of why MIF is unlikely to be induced by dark enzymatic

pathways. Yang and Sturgeon 2009 report that dark ethylation of Hg(II) by NaBEt4

shows only strong MDF. Neither do Malinovsky et al. 2009 see MIF in dark methylation

of inorganic tin by methylcobalamin. They do see MIF if they expose the solution to UV

light, but only for acidic and semineutral pH. Oxidative condensation of Hg(0) as Hg(II)

on particulates and acidic aerosols in volcanic plumes shows no MIF (Zambardi et al.

2009). In general, MIF signatures larger than approximately 0.2‰ seem likely to be

associated with a small group of reaction pathways and the pathways seem to be

identifiable by their 199Hg/201Hg ratio (Bergquist and Blum 2009). MIF has the

potential to highlight these few pathways of Hg transformation among the many

pathways that result in MDF and this potential increases the utility of Hg in tracking the

transport and fate of Hg in the environment.

2.2.2. Mechanisms of mass-independent fractionation

There are two plausible mechanisms that have been proposed to account for the MIF

observed in Hg isotopes: The nuclear volume effect (NVE) and the magnetic isotope

effect (MIE). The NVE was postulated by Bigeleisen (1996) to explain the even-odd

MIF observed for uranium (Fujii et al. 1989a, 1989b) and is relevant for heavy elements

in general. Schauble 2007 gives an extensive theoretical discussion of its equilibrium

effects on the isotopes of mercury and thallium. As atoms get larger, nuclear volume and

nuclear charge radii do not increase linearly with mass. This effect becomes significant

for heavy elements such as U, Hg and Tl, but has been observed for lighter elements such

as Cr as well (see review: Fuji et al. 2009). It is predicted to have the largest deviations

Page 36: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

29

from MDF for the odd isotopes resulting in an even-odd pattern for isotopic fractionation.

NV predicts that the volume of the odd isotopes (ex. 199Hg) is closer to that of the next

lower even nucleus rather than being intermediate between the volumes of its even

neighbours (ex. 198Hg, 200Hg). This effects the shielding of electrons and thus bond

strengths. Schauble 2007 calculated expected NV contributions to Hg MIF based on

mean squares of nuclear charge radii, which he retrieved from calculations by Angeli

2004, and predicted mass deviations for 196Hg, 199Hg, 201Hg and 204Hg. Mass deviations

are observed, however, only in 199Hg and 201Hg. Ghosh et al. 2008 pointed out that the

experimentally determined 204Hg charge radius (Hahn et al. 1979) differed from the

Angeli et al.’s calculations; after revising Schauble’s scaling factor they found no mass-

independent anomalies were expected for 204Hg.

The smaller relative size and larger relative charge density of the odd nuclei mean that s

orbitals, which have a higher electron density at the nucleus, will be more strongly bound

to the odd nuclei. P, d and f electronic orbitals have a smaller electron density at the

nucleus. When orbiting an odd nucleus, these three will be more effectively screened by

the s electrons and therefore more weakly bound to odd nuclei. The two major oxidation

states of mercury (Hg0 and Hg2+) have zero or two electrons in the 6s orbital respectively.

At equilibrium, therefore, the even isotopes will be more stable in a state where they have

a lower electron density around the nucleus (Hg2+) because their valence orbitals (5d)

will be more strongly bound. The odd isotopes will be more stable in a state with a

higher electron density around the nucleus (Hg0) because their valence orbitals (6s) will

be more strongly bound (Bigeleisen 1996 and Schauble 2007). If all isotopes are in the

form of elemental Hg, the odd isotopes will form weaker covalent bonds and be more

volatile.

The second mechanism that is likely to produce MIF in mercury is the magnetic isotope

effect, first observed for 13C in the photodecomposition of dibenzyl ketone (Buchachenko

et al. 1976) and proposed to apply to natural mercury isotopic fractionation by Bergquist

and Blum 2007. This effect is a result of interactions between the magnetic moments of

nuclei (only odd isotopes have non-zero nuclear magnetic moments) and the spin

Page 37: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

30

magnetic moments of electrons (any paramagnetic species will have net spin angular

momentum and therefore a magnetic moment). The interactions are called hyperfine

coupling, and are so small that they have no effect on the equilibrium characteristics of

chemical systems (Turro1983, Bigeleisen 1996, Engel and Reid 2006). They can,

however, produce a kinetic effect by allowing isotopes with different nuclear spins to

react at different rates.

Consider as an example an aqueous photo-dissociation reaction that involves a pair of

radical intermediates. The two radical electrons could be spin-paired in a singlet state:

one electron has spin + ½ and the other - ½ such that their spins are precisely opposite in

three-dimensional space. Within the solvent cage formed by water molecules, such a

radical pair is highly reactive and likely to recombine or disproportionate (Turro 1983

and Engel and Reid 2006). The two radical electrons could also be in a triplet state: the

electrons are unpaired and have spins of + ½, + ½ or - ½,- ½ or + ½, - ½. In this last

case, the projections of the two spins on a reference axis add to zero, but the spins are not

opposite in three-dimensional space.

Recombination (or disproportionation) requires the radical electrons to finish in a spin-

paired singlet state, but this path is unavailable to the electrons in the triplet state because

it does not conserve their total spin angular momentum. Within a solvent cage, the triplet

radicals could undergo chemical rearrangement resulting in a structurally different radical

pair or they could diffuse out of the solvent cage and become free (uncorreleated)

radicals. A third possibility is triplet-singlet conversion, also called intersystem crossing

(ISC), to produce a set of singlet radicals. If circumstances are favorable, hyperfine

coupling can induce ISC through a simultaneous exchange of electron-nuclear spin (total

spin angular momentum is conserved). Re-phasing or reversal of an electron spin results.

A spin-selective process such as ISC can sort isotopes because the (odd) magnetic nuclei

provide a mechanism for electrons to undergo ISC that the (even) non-magnetic nuclei do

not. Radicals with magnetic nuclei will undergo triplet-singlet spin conversion more

quickly and accumulate preferentially in the recombination (or disproportionation)

products, whereas the non-magnetic nuclei will be more likely to diffuse out of the

Page 38: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

31

solvent cage or undergo structural rearrangement (see reviews: Turro 1983, Buchachenko

2001).

In order for hyperfine coupling to effectively induce ISC, the radical fragments must

separate to a physical distance where the energy gap between the singlet and triplet states

is on the same order or smaller than the hyperfine interaction. At such a distance, the

original chemical bond will have completely dissociated and one or more solvent

molecules may separate the radicals. The probability of the newly-singlet radical

fragments re-encountering each other before separating completely is high because they

will “bounce” against the surrounding solvent molecules and each other several times

before finding avenues through the solvent to become fully separated (Turro et al. 2009).

Example reaction schemes for the photo-reduction of Hg2+ and the photo-degradation of

MeHg have been proposed (Bergquist and Blum 2007, 2009; Buchachenko 2009;

schematic courtesy of B. Bergquist) and are presented below for Hg2+:

X Hg L hv X Hg

L

T 198,200202,204 L

Hg

dissociation tofree radicals

199,201

X Hg

L

S

Hg0

and for MeHg:

CH3HgXodd

hv

X Hg

C

H3

even Hg0

T-S conversion kTS(odd) > kTS(even)

Page 39: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

32

2.3. FRESHWATER MERCURY MASS TRANSFER

PROCESSES AND SPECIES TRANSFORMATIONS

2.3.1. Concentrations

Aqueous freshwater Hg species are usually grouped into three categories: total mercury

(HgT), dissolved gaseous mercury (DGM) and methyl mercury (MeHg). Representative

concentrations are displayed in Table 2.2 below.

Table 2.2. Typical Hg concentrations in natural, unpolluted freshwaters

Species low

concentration high

concentration Comment Reference a few hundred

pg/L

Arctic: 3 lakes and 1 wetland

Amyot et al. 1997a

same 100 high-altitude

U.S. lakes Krabbenhoft et al. 2002

a few ng/L 23 northern

Wisconsin lakes Watras et al. 1995a

total mercruy (HgT)

same 6 temperate

Canadian lakes Amyot et al. 1997b

HgT*10-3 or less high altitude lakes Krabbenhoft et al. 2002

0.1*HgT to 0.01*HgT

Wisconsin Lakes Watras et al. 1995a methyl

mercury (MeHg)

0.05 ng/L 3.0 ng/L Experimental Lake Area, NW Ontario

Sellers et al. 1996

HgT*10-5 HgT*10-2 Wisconsin lakes;

Great Lakes Vandal et al. 1991, Jeremiason et al. 2009

106-215 pg/L maximum summer

concentrations Garcia et al. 2005a

dissolved gaseous mercury (DGM)

17-126 pg/L maximum summer

concentrations Amyot et al. 1997b

Both DGM and MeHg can be measured directly. HgT is generally the portion in an

acidified sample that can be reduced by SnCl2, purged from the solution and collected on

a gold trap. Mercury remaining after subtraction of DGM and MeHg from the total is

assumed to be Hg(II) of some variety. Approximately a quarter of both Hg(II) and MeHg

is estimated to be particulate bound in freshwaters (Watras et al., 1995a, Rolfhus et al.

2003). These estimates are based on the difference in the total Hg (HgT and MeHg)

detected in unfiltered versus filtered water samples (Watras et al. 1995a) or multiplying

the particulate concentration by suspended particulate matter concentration (Rolfhus et al.

Page 40: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

33

2003). There remain, however, substantial uncertainties in how Hg species partition

between the aqueous phase and particulate matter, as well as in the chemical lability of

particulate-bound Hg (Lindberg et al. 2007). Lamborg et al. 2003 conclude that over

99% of Hg in freshwaters is complexed to organic ligands. Some authors see no

significant seasonal variation in either MeHg or HgT when measuring in spring and fall

(ex. Watras et al. 1995a) but others report seasonal variation by a factor of two or three in

both HgT and MeHg (Krabbenhoft et al. 2002, Selvendiran et al. 2009),

2.3.2. Sources

2.3.2.1. Deposition

For precipitation-dominated lakes, atmospheric deposition (rain, snow and dry

deposition) is the dominant source of Hg (Watras et al. 1995a, Watras and Morrison

2008). Although ambient tropospheric Hg is principally Hg(0) (98% or more), the

majority of Hg in rain is derived from atmospheric scavenging of particulate Hg.

Fitzgerald et al. 1991 report that HgR at their Wisconson study site averaged only 19%

(2.0 0.3 ng/L) of the total Hg content in rain and 52% (3.1 2.2 ng/L) in snow. They

conclude that atmospheric deposition could easily account for the total mass of Hg

detected in fish, water and sediments in their study lake. MeHg in precipitation is

sometimes detectable in trace amounts (tens to hundreds of pg/L, Fitzgerald et al. 1991,

Hammerschmidt et al. 2007).

2.3.2.2. Runoff

In drainage lakes, runoff can be a large source of HgT. Runoff is a significant source of

MeHg in water bodies, in particular runoff from wetland areas (Watras et al. 1995a,

Sellers et al. 2001, Selvendiran et al. 2009). Terrestrial organic carbon is an important

transport vector for HgT and MeHg. Events causing an export of catchment soils

(agriculture, forestry, urbanization) to water bodies can contribute substantially to

waterborn Hg loads. Export to lakes from watersheds generally lags far behind

atmospheric deposition, but is expected to continue for decades or longer after

atmospheric deposition subsides (Harris et al. 2007, Watras and Morrison 2008, Mills et

al. 2009).

Page 41: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

34

2.3.2.3. In-situ production of MeHg

Inorganic methylation of Hg has been both argued (Weber 1993) and demonstrated

(Akagi and Takabatake 1973, Gardfeldt et al., 2003). In general, however, most mercury

methylation is considered a biologically mediated process (Fitzgerald and Lamborg 2007,

Selin 2009). While sulfate reducing bacteria (SRB) are assumed to play the largest role,

iron-reducing bacteria have also been implicated (as in Kerin et al. 2006) and

methanogens have not been ruled out (Barkay and Wagner-Dobler 2005). For example,

mechanisms involving corrinoid-containing proteins have been proposed (Barkay and

Wagner-Dobler 2005 and references therein). Cinnabar may serve as a substrate,

although uncharged soluble forms of mercuric sulfide are the more likely source for

methylation. Mercury bound to organic ligands may also be methylated (Barkay and

Wagner-Dobler 2005), however, and Schaefer and Morel 2009 demonstrate that mercury

methylation is greatly enhanced by the formation of Hg-cysteine complexes.

Methylation of mercury can occur both in anoxic sediments (Gilmour et al. 1992) and

anoxic hypolimnia of lakes (Watras et al. 1995b, Eckley and Hintelmann 2006).

Methylation in the water column has been associated with bacterial communities

migrating to the oxic-anoxic interface that can move from the sediment surface to higher

in the water column during periods of stratification. While methylation rates are more

intense in the surficial sediments, methylation zones in the water column can represent a

larger volume (depending on the depth of the hypolimnion, which changes seasonally)

and also generate significant quantities MeHg. MeHg produced in the water column may

also account for more of the MeHg that is bioaccumulated.

2.3.3. Sinks

2.3.3.1. Outflow

Lake and watershed morphology, precipitation, groundwater discharge and recharge and

other variables that influence the hydraulic residence time of water in a lake affect

whether a given water body will be an overall source or sink for mercury. The relative

magnitude of inflow and outflow of a particular Hg species varies from case to case (for

Page 42: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

35

example, seepage versus drainage lakes) and from season to season (Krabbenhoft et al.

2002, Selvendrian et al. 2009).

2.3.3.2. Sedimentation

Sediments represent both the largest repository of MeHg and HgT (Fitzgerald et al. 1991)

and also an important removal pathway. Mercury sequestration through sedimentation

can be calculated using annual sediment deposition rates and Hg concentration in

sediment material (Krabbenhoft et al. 1992, Mills et al. 2009, Rolfhus et al. 2003).

Unfortunately, estimates of sediment accumulation rates are lacking in many studies

(Jeremiason et al. 2009) and researchers frequently estimate sedimentation rates by the

discrepancies of other mass components (ex. Sellers et al. 1996, Watras and Morrison

2008, Selvendiran et al. 2009).

2.3.3.3. Bioaccumulation

The relationship between environmental Hg loading and bioaccumulation is the subject

of ongoing research. Both MeHg and and Hg(II) species can passively diffuse across a

lipid membrane in the form of small, uncharged complexes. Trophic transfer, however,

is more efficient for MeHg. MeHg accumulates in the cell cytoplasm of phytoplankton

and is assimilated more efficiently (a factor of 4) by zooplankton than Hg(II), which

remains bound to phytoplankton membranes and is more likely to be excreted by

zooplankton (Mason et al. 1996). Biota can respond quickly to alterations in deposition

rates. A recent and striking example of this has been noted during the METAALICUS1

project, involving the addition (or spike) of three different highly enriched stable isotopes

of Hg (198Hg, 200Hg and 202Hg) separately to a lake, its watershed and a large wetland that

drains into the lake. The methylated lake spike was detected in the anoxic bottom waters

three days after the first surface addition. In sediments, zooplankton and benthos it was

detected after one month and in several fish species after 2 months. Spiking increased the

Hg loading of the lake by approximately 120% annually. By the end of the third year of

consecutive spiking, MeHg in water and biota was 30 to 40% higher than they would

1 Mercury Experiment to Assess Atmospheric Loading in Canada and the United States, conducted at Lake 658 in the Experimental Lakes Area of northwestern Ontario.

Page 43: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

36

have been if the spike had not been added (Harris et al. 2007). Concentrations of Hg in

the water column and in short-lived organisms such as periphyton, zooplankton and water

mites can respond quickly to reductions in atmospheric deposition and to MeHg

availability (Orihel et al. 2008). Hg concentrations in older individuals of long-lived

species, especially those participating in benthic food webs,are likely to be much slower

to respond (Orihel et al. 2008). Reductions in fish Hg concentration (50% or more) have

been reported, however, in both freshwater and marine environments subjected to point-

source contamination within ten years of remedial actions. Hg levels of fish in remote

waterbodies have likewise been observed to systematically decline when a decline in wet

or bulk Hg deposition is also documented (see review: Munthe et al. 2007). Total Hg

concentrations in sediments, however, are slow to respond to decreases in loading (ex.

Orihel et al. 2008). Recent experimental manipulations such as the labeled ecosystem

additions in the METAALICUS project are designed to increase our understanding of the

relationships between Hg loading and the response of Hg in organisms. Reports on these

manipulations will be forthcoming once enough time has passed for adequate observation

of the recovery phase.

2.3.3.4. Degradation of MeHg

Methyl mercury can be degraded by both biotic and abiotic processes. Two microbial

degradations mechanisms are known to exist. Oxidative demethylation (OD) results

principally in CO2 and in Hg(II), which becomes available for re-methylation. Small

amounts of CH4 are generated. Pathways of oxidative demethylation are not known and

pure cultures have not been characterized (Barkay and Wagner-Dobler 2005 and

references therein). Reductive demethylation (RD) produces only CH4 and Hg(0), which

may escape the aqueous reservoir by evasion. The best known path of RD is mediated by

microorganisms that carry Hg resistance (mer) operons, which include the enzymes

organomercury lyase and merucuric reductase. Two dominant factors that control which

of the two pathways is followed are redox conditions and levels of mercury

contamination. RD is favored when mercury concentrations exceed thousands of ng/g

sediment in anaerobic conditions (or hundreds of ng/g sediment in aerobic conditions),

whereas OD dominates at low mercury concentrations and anoxic conditions. Overall,

Page 44: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

37

MeHg is degraded reductively in highly contaminated environments but in less

contaminated conditions, OD dominates.

It does not appear that dark abiotic processes result in significant degradation of MeHg.

Aqueous solutions containing MeHg and organic matter have been kept strictly in the

dark under laboratory conditions. Concentrations have been monitored for days to weeks

at a time, with no reported decrease (Bergquist and Blum 2007, Zheng and Hsu-Kim

2010). In sediments and bottom waters, microbial processes most likely dominate MeHg

degradation, whereas in light-exposed environments, photo-degradation may be the major

mechanism (Barkay and Wagner-Dobler 2005).

Photodemethylation was reported as an important sink for MeHg in an Ontario lake by

Sellers et al. 1996. The investigators estimated that it represented approximately twice

the sum of MeHg inputs, which necessitated significant in-situ MeHg production. In a

more detailed mass-balance analysis, Sellers et al. 2001 concluded that over the course of

a single year, photo-degradation represented ~78% of MeHg losses from the lake under

study. Other studies support the significance of MeHg photo-degradation. Krabbenhoft

et al. 2001 found photo-degradation to be active in the top meter of a high altitude lake

and to be a MeHg sink of comparable importance to sedimentation and stream flow.

Hammerschmidt and Fitzgerald 2006 concluded that photodecomposition could account

for as much as 80% of the MeHg mobilized annually from sediments in Toolik Lake,

Alaska.

UV radiation has been identified as the portion of naturally available radiation resulting

in the highest rates of photo-degradation (Sellers et al. 2001, Lehnherr and St. Louis

2009, Hammerschmidt and Fitzgerald 2006). Hammerschmidt and Fitzgerald 2006 found

that photodecomposition due to visible radiation was important in their Arctic study lake.

Photodecomposition of MeHg was a linear function of PAR (which decreased

exponentially with depth) with the exception of the top few cm of the water column.

Photo-decomposition in these top few centimeters is more intense (see also Sellers 1996),

but Hammerschmidt and Fitzgerald 2006 point out that the penetration depth of visible

Page 45: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

38

radiation is much larger than of UV radiation, and the photodecomposition zone of

visible radiation represents a much larger volume. They did not identify a mechanism of

photodemethylation due to visible light (see Table 2.3 p. 44) but noted that

photosensitization of dissolved organic material (DOM) seemed to be a requisite.

When normalized for initial mercury concentration and incident radiation,

photodecomposition rates for a temperate Ontario lake (Sellers et al. 1996), for controlled

field experiments (Lehnherr and St. Louis 2009) and the Alaskan lake from

Hammerschmidt and Fitzgerald 2006 were similar (0.0018 ng m2 L-1 E-1).

Environmental factors such as pH and DOC were very different in these lakes, and it is

possible that this rate is approximately uniform among freshwater bodies (Lehnherr and

St. Louis 2009). In addition to light intensity and frequency, photo-degradation rates also

depend on the presence of humic substances. Specifically, the relative concentrations of

MeHg and humic acid with thiol-containing ligands has been shown to speed degradation

rates (Zhang and Hsu-Kim 2010), possibly by decreasing electron transition energies

sufficiently for low UV wavelengths to induce demethylation.

2.3.3.5. Volatilization

There are both biotic and abiotic process that carry out the reduction of Hg(II) to Hg(0) in

natural freshwaters. Dissolved gaseous mercury (Hg(0); DGM) is a volatile species that

can diffuse to the lake surface and escape to the atmosphere. DGM is also a product of

both the biotic and abiotic processes that degrade MeHg mentioned above.

Concentrations of MeHg, however, are typically one or more orders of magnitude lower

than those of HgT. (see Table 1.2). Contributions of MeHg to DGM, while representing

an important component of the MeHg aquatic cycle, do not comprise a very significant

portion of DGM produced in lakes. Experimentally, it is more convenient to monitor the

production of dissolved gaseous mercury than the reduction of either Hg(II) or MeHg.

The phrase “DGM production” is somewhat interchangeable with “Hg reduction”. While

DGM production represents the fate of reduced Hg(II) well, it is not indicative of MeHg

reduction in natural waters where a large excess of Hg(II) species are present. The

Page 46: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

39

following discussion of DGM production mechanisms and conditions, therefore, applies

to Hg(II).

Inorganic mercury(II) can be reduced by biotic and abiotic pathways. Both direct and

indirect biotic pathways have been reported. As with reductive demethylation, the most

well-studied pathway of direct biological reduction of Hg involves the mer operon (and

MerA, mercuric reductase), which allows cells to avoid the toxicity of Hg(II) by reducing

it to Hg(0) and removing it from their environment. It is inducible by high environmental

concentrations of Hg(II) (50 pM or more, Siciliano et al. 2002). Actively growing

cultures (105 – 106 cells/mL) in a cell medium contaminated by as much as 20 ppm

Hg(II) can reduce 99% of the Hg in under an hour (Kritee et al. 2007). Between 1 and

10% of culturable microbes taken from natural, uncontaminated environments are likely

to have a functional MerA enzyme; this proportion increases among culturable microbes

that survive in Hg-contaminated sites (Kritee et al. 2007).

Mer-independent microbial reduction has also been observed. For example, it has been

reported for the species Shewanella oneidensis MR-1 under anoxic conditions, which

cannot live at the micromolar concentrations of mercury tolerated by mer-carrying

species (Wiatrowski et al. 2006). Despite not being able to survive at high concentrations

of Hg, it has been observed to reduce Hg(II) more effectively than its mer-containing

transconjugate at nM concentrations (Wiatrowski et al. 2006). Such mer-independent

pathways of microbial reduction may contribute to the removal of Hg(II) in anoxic

aqueous environments, which are important sites for Hg(II) methylation.

Thus there are microbes that do not carry an Hg-specific reductive pathway that

nevertheless appear to mediate the reduction of Hg species. Ben-Bassat and Mayer 1978

showed that illuminated algal suspensions of Chlorella pyrenoidosa volatilize Hg(0) at a

fairly constant rate, given a renewable supply of Hg(II). By alternating light and dark

conditions, they documented tight coupling between illumination and volatilization of

Hg(0) and suggested the reduction occurred in the cell medium as a result of leakage of

some reducing agent from the cells, possibility metabolites produced during

Page 47: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

40

photosynthesis. Poulain et al. 2004 observed peaks in DGM concentration at the bottom

of the metalimnion in two lakes in the Experimental Lake Area (Ontario) at depths where

over 90% of incident radiation had been attenuated, which coincided with maxima in

chlorophyll and phytoplankton biomass. No transcripts of the merA gene could be

detected from RNA extracts.

The following discussion of abiotic reductive pathways will focus on photo-degradation,

but it must be noted that dark abiotic reduction does occur, although at much slower rates

than photo-induced processes. A small increase in DGM concentrations is sometimes

observed in field samples taken early in the morning and incubated, starting before dawn,

in dark bottles (for example Garcia et al. 2005a) . Such dark abiotic reduction is expected

to be a result of humic substances. In laboratory experiments, Alberts et al. 1974 showed

that purified (and sterilized) humic acid from pond sediment would slowly reduce Hg(II)

(added as HgCl2) at room temperature in a way that showed some dependence on pH.

Allard and Arsenie 1991 confirmed that reduction of Hg(II) by humic substances (in

filtered solutions, using sterile equipment) would take place aerobically in the dark. Low

rates of dark reduction relative to photo-reduction have been demonstrated in laboratory

experiments (17% compared to 90% HgT losses over 5 hours, Bergquist and Blum 2007)

and in-situ field experiments (up to 50% of DGM photo-production in a wetland but

negligible in lakes, Garcia et al. 2005a). By definition, dark reduction is not dependent

on sunlight, however, and can continue at all hours of the day.

The production of Hg(0) through the midday hours of summer can represent several

percent of the HgT burden in surface waters, for example up to ~8% over the course of a

day in lakes (Amyot et al. 1997a, Amyot et al. 1997b). Through multiple additions of

isotopically enriched Hg(II) to field mesocosms, Poulain et al. 2006 showed that 30 to

60% of the added mercury was lost to the atmosphere, and that loss was unrelated to the

rate of loading.

Page 48: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

41

The link between solar radiation and in situ freshwater DGM production was

demonstrated by Amyot et al. 1994. Samples of lake water (at approximately 5 mg/L

DOC) incubated in transparent bottles yielded DGM concentrations 2 to 9 times higher

than those kept in black bottles. Blocking UVB radiation (280-320 nm) by wrapping

bottles in mylar film did not affect DGM production significantly. DGM concentrations

showed a clear diel pattern, with the lowest concentrations (~ 0.2 pM) before sunrise and

the highest at noon (0.6 to 0.7 pM). Comparison of midday concentrations from August,

September and November 1993 showed large seasonal differences as well (the November

maximum DGM was less than 0.2 pM compared with 0.7 pM in August).

UV radiation on the whole has been observed to contribute more to DGM production

than visible radiation. Investigators find that visible radiation alone generates far less

DGM (over 70% decrease, Garcia et al. 2005a) or none at all (O’Driscoll et al. 2006).

In some cases, it has been shown that UVB makes a significant contribution to DGM

production apart from UVA (Amyot et al. 1997a, 1997b, Garcia et al. 2005a). Other

studies have not shown that UVB is more important than UVA (Amyot et al. 1994,

1997b).

An additional important influence on photo-reduction is organic matter. Amyot 1997a

proposed that high DOC concentrations in some lakes would absorb most of the UV

radiation at the lake surface, making it unavailable for direct reduction of Hg complexes.

In contrast, increased penetration in clear lakes would allow UV radiation to affect DGM

production significantly. Garcia et al. 2005a also concluded that UVB contributions to

photo-reduction were more important in clearer lakes. They found a strong negative

correlation between DGM production and DOC concentration for samples exposed to

UVB radiation (r = 0.99) and a strong positive correlation between DGM production and

DOC concentration under UVA and visible range (400-700 nm) exposure (r=0.99). They

remark that incoming photons (presumably in the UVA range) could be photosensitizing

the chromophoric dissolved organic carbon. It should be mentioned that some lakes

show no DGM production at all, for example the arctic lake Amituk (Amyot et al. 1997a)

Page 49: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

42

and Lake Eerie (Amyot et al. 1997b); in both cases the authors suggested low levels of

photoreducible complexes as an explanation.

In general, DGM levels incubated in temperate lake water samples under sunlight reach a

plateau after a few hours (about 3, Amyot et al. 1997b) and drop slightly thereafter.

DGM levels in samples incubated in black bottles show slow decay. By incubating pre-

purged samples (14 hours in quartz bottles under cloudless conditions followed by

flushing with Hg-free air in an effort to exhaust the photoreducible Hg), Garcia et al.

2005a showed that plateaus were not likely due to limited HgR because DGM production

resumed upon exposure to daylight the following day. The authors suggest that plateaus

are more likely a balance between oxidation and reduction.

Photo-oxidation was suggested as a simultaneous competitive process with photo-

reduction by Lalonde et al. 2001 to explain why lake water samples spiked with Hg(0)

(introduced by bubbling with a gas) and subjected to UVB radiation showed a decrease in

Hg(0) over time. Garcia et al. 2005b gathered evidence of competition between photo-

reduction and photo-induced oxidation in natural freshwater systems by collecting

surface water samples in batches and monitoring DGM concentration over time under

different incubation conditions. They noted initial DGM concentration, DGM after one

hour of dark and light incubation respectively and DGM after two hours of dark and light

incubation respectively. Samples incubated in the light showed an increase in DGM

while samples incubated in the dark showed a decrease. The greatest decrease was

observed in samples collected at 12:50 pm, when oxidation was equivalent to 77% of the

DGM production under sunlight exposure. The authors concluded that both processes

occur simultaneously in sunlight with the leading process depending on light intensity: at

high irradiance, photo-reduction dominates but under cloudy conditions and towards

sunset, photo-induced oxidation dominates.

To investigate photo-reduction alone, therefore, it is necessary to find a way to decouple

it from photo-oxidation. One way to do this is to remove the DGM as it is formed so that

photo-oxidation has no substrate (i.e., via sparging with Hg-free gas), as in Allard and

Page 50: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

43

Arsenie 1991 and O’Driscoll et al. 2006. The photochemical reactions can be

generalized as follows (as in O’Driscoll et al. 2004):

Hgreducible + photo-reductants DGM + photo-oxidants

Thus removal of DGM pushes this equation toward reduction.

In order to estimate the losses of Hg via reduction to Hg(0) in an environmental system, it

is necessary to quantify the relative importance of all three (biotic, dark aboitic and

photochemical) reduction pathways. Distinguishing between these three processes can be

difficult in the environment. All three types of reduction are drawn on by Poulain et al.

2004 to explain the DGM depth profiles they obtained for two lakes in the Experimental

Lakes Area (sampled in summer, fall and under ice in the late winter). DGM

concentrations were always highest in the top two meters of the water column (~ 2 pM

L-1), but a second, smaller peak in concentration was frequently observed at the limit of

the euphotic zone, which was at the bottom of the metalimnion (at depths of 3.5 to 5.5 m,

depending on the lake and sampling occasion). A third, small peak (up to 0.3 pM L-1)

was observed at the sediment-water interface when the bottom of the hypolimnion was

anoxic. They proposed that DGM production at the surface was photo-chemically

mediated, since DGM production stopped in the dark and filtration to remove organisms

did not affect the production rates. DGM production in the metalimnion was associated

with phytoplankton biomass. As discussed above, and the authors drew on Ben-Bassat

and Mayer’s (year) observations of reduction of Hg(II) by illuminated algal suspensions

to explain it. No transcripts of the merA gene could be detected from RNA extracts. For

the peaks observed at the sediment-water interface, they propose anoxic reactions

involving humic substances or microbial activities. Similar rationales can be applied to

other DGM depth profiles, for example Vandal et al. 1991 who report that DGM

concentrations reach a maximum at a depth of 7 m before decreasing and then rising

slightly at the sediment-water interface.

2.3.4. Light penetration

Page 51: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

44

When considering photo-reduction, light penetration is an important variable needed to

quantify and understand the amount of photo-reactions possible in a natural water body.

Because photo-reduction is a major source of DGM, high DGM production rates drop off

quickly with depth. There have been several studies that have attempted to quantify light

and UV penetration in freshwater systems. Scully and Lean 1994 undertook systematic

investigations of UV attenuation in temperate freshwater lakes (n = 20, DOC from 0.50 to

7.80 mg/L). They found that light attenuation coefficients, generally calculated as:

z

EK depth

depth

ln

Where E is downward irradiance (in W*area-1*nm-1) and

z is distance below the water surface

varied considerably across survey lakes. DOC was the principal attenuating substance of

UVB and UVA respectively, and from their data Scully and Lean derived integrated

(over wavebands) attenuation coefficients that were power functions of DOC

concentration. Morris et al. 1995 also investigated UV attenuation in freshwater lakes (n

= 59; 30 lakes with DOC below 2 mg/L DOC). They offer a modification to Scully and

Lean’s model for calculating attenuation coefficients in low DOC environments.

Examples depths of 99% attenuation of various radiation regions are given in Table 2.3:

Table 2.3. Example depths of 99% attenuation of solar radiation at various DOC concentrations. Depths were calculated from reported measured attenuation coefficients and not generalized equations.

Depth of 99% attenuation (m) Lake DOC (mg/L)

UVB UVA Visible Reference

Gutierrez 0.32 8.5 15 38 Morris et al. 1995 Giles 1.16 11.5 17.5 35.5 Morris et al. 1995 Hamilton Harbor 4.90 0.43 0.8 1.7 Scully and Lean 1994 Sharpes Bay, Ontario 5.60 0.48 0.76 2.1 Scully and Lean 1994 Toolik 5.07 n/a 0.83 10 Morris et al. 1995 Red Rock 10.11 0.15 0.19 1.9 Morris et al. 1995

Page 52: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

45

Crump et al. 1999 find that the equations in Scully and Lean 1994 and Morris et al. 1995

were poor predictors for UV attenuation in high DOC environments (n = 11 ponds, DOC

from 8 to 16 mg/L). They found the most useful predictor for attenuation coefficients

with depth to be absorption coeffecients measured on filtered water using a laboratory

spectrophotometer (r2 = 0.94). Ninety-nine percent of the UVB was attenuated in 12-60

cm while the depth of UVA penetration was twice as large. They also noted that the

percentage of DOC that acted to absorb UVB radiation changed throughout the year, with

the highest attenuation per unit DOC occurring in May and the lowest in September.

Thus light penetration is important in predicting the amount of Hg photo-reduction that

occurs in nature, but is a difficult variable to use to quantify photo-reduction.

2.3.5. Organic matter complexation

Natural organic matter also plays an important role in determining the amount of Hg that

is reduced to DGM in the environment. Trace metals, including mercury, are usually

bound to the acid sites of natural organic matter (NOM). Carboxyllic acids and phenols

comprise over 90% of these sites. However Hg is expected to bind preferentially to

reduced sulfur sites, which are thought to play a role disproportionate to their abundance

in the complexation of “soft” trace metals in the environment due to their high binding

coefficients (Xia et al. 1998, 1999, Ravichandran 2004 and references therein).

Sulfur is a minor component of DOM, but Hg(II) stability constants for compounds

containing reduced sulfur are usually much stronger than stability compounds with other

ligands. Since the amount of reduced sulfur available for binding in natural freshwaters

(up to a few ppt) is a few orders of magnitude larger than Hg concentration in

uncontaminated sites, S-containing moieties are predicted to dominate complexation

(Hesterberg et al. 2001, Haitzer et al. 2002). It has been suggested that the reactivity of

thiol-complexed Hg differs from the reactivity of Hg complexed to oxygen species

(Zheng and Hsu-Kim 2010, Zhang and Hintelmann 2009).

There are large variations in Hg isotopic compositions in natural samples. Many

environmentally relevant transformations have been shown to cause MDF in Hg isotopes,

Page 53: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

46

and distinguishing source MDF signatures from MDF imparted during transport and

transformation in the environment will be challenging. There are fewer processes that

induce MIF signatures in Hg samples, and large MIF ( > than 0.2‰) seem to be

associated with photochemical transformations. MIF therefore has the potential to

highlight and track photochemical transformations among the many processes that affect

the environmental fate of Hg. For example, using both mass-dependent and mass-

independent Hg isotope signatures, along with 15N and 13C signatures, Senn et al. 2010

build a strong argument that coastal and oceanic fish in the Gulf of Mexico derive their

MeHg from different sources. The strong possibility that benthic and pelagic food webs

incorporate MeHg from different sources has important implications for future toxicity

mitigation and remediation efforts. As discussed in the previous pages, photochemical

reduction is a significant pathway whereby water bodies release their burden of Hg back

to the atmosphere, and photo-degradation is an important sink (and detoxification

mechanism) for the MeHg in water bodies. These processes are strongly influenced by

the intensity and frequency of sunlight available, and the kind and quantity of dissolved

organic matter in the water column. If MIF is to be used as a tool to understand the

photochemical transformations of Hg, it is important to investigate whether and how the

frequency and intensity of light and the kind and quantity of organic matter affect MIF.

Chapter 3 of this thesis is devoted to the first of these questions: the manner in which

frequency and intensity of available sunlight affects the expression and magnitude of MIF

during photo-reduction of Hg(II) and MeHg respectively.

Page 54: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

47

LITERATURE CITED Akagi, H. and Takabatake, E., 1973. Photochemical formation of methylmercuric

compounds from mercuric acetate. Chemosphere 3, 131-133.

Alberts, J. J., Schindle.Je, Miller, R. W., and Nutter, D. E., 1974. ELEMENTAL MERCURY EVOLUTION MEDIATED BY HUMIC ACID. Science 184, 895-896.

Allard, B. and Arsenie, I., 1991. ABIOTIC REDUCTION OF MERCURY BY HUMIC SUBSTANCES IN AQUATIC SYSTEM - AN IMPORTANT PROCESS FOR THE MERCURY CYCLE. Water Air and Soil Pollution 56, 457-464.

Amyot, M., Lean, D., and Mierle, G., 1997a. Photochemical formation of volatile mercury in high Arctic lakes. Environmental Toxicology and Chemistry 16, 2054-2063.

Amyot, M., Mierle, G., Lean, D., and McQueen, D. J., 1997b. Effect of solar radiation on the formation of dissolved gaseous mercury in temperate lakes. Geochimica Et Cosmochimica Acta 61, 975-987.

Amyot, M., Mierle, G., Lean, D. R. S., and McQueen, D. J., 1994. SUNLIGHT-INDUCED FORMATION OF DISSOLVED GASEOUS MERCURY IN LAKE WATERS. Environmental Science & Technology 28, 2366-2371.

Angeli, I., 2004. A consistent set of nuclear rms charge radii: properties of the radius surface R (N, Z). Atomic Data and Nuclear Data Tables 87, 185-206.

Barkay, T. and Wagner-Dobler, I., 2005. Microbial transformations of mercury: Potentials, challenges, and achievements in controlling mercury toxicity in the environment, Advances in Applied Microbiology, Vol 57.

Benbassat, D. and Mayer, A. M., 1978. LIGHT-INDUCED HG VOLATILIZATION AND O2 EVOLUTION IN CHLORELLA AND EFFECT OF DCMU AND METHYLAMINE. Physiologia Plantarum 42, 33-38.

Bergquist, B. A. and Blum, J. D., 2007. Mass-dependent and -independent fractionation of Hg isotopes by photoreduction in aquatic systems. Science 318, 417-420.

Bergquist, R. A. and Blum, J. D., 2009. The Odds and Evens of Mercury Isotopes: Applications of Mass-Dependent and Mass-Independent Isotope Fractionation. Elements 5, 353-357.

Page 55: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

48

Bigeleisen, J., 1949. THE RELATIVE REACTION VELOCITIES OF ISOTOPIC MOLECULES. Journal of Chemical Physics 17, 675-678.

Bigeleisen, J., 1996. Nuclear size and shape effects in chemical reactions. Isotope chemistry of the heavy elements. Journal of the American Chemical Society 118, 3676-3680.

Bigeleisen, J. and Mayer, M. G., 1947. CALCULATION OF EQUILIBRIUM CONSTANTS FOR ISOTOPIC EXCHANGE REACTIONS. Journal of Chemical Physics 15, 261-267.

Biswas, A., Blum, J. D., Bergquist, B. A., Keeler, G. J., and Xie, Z. Q., 2008. Natural Mercury Isotope Variation in Coal Deposits and Organic Soils. Environmental Science & Technology 42, 8303-8309.

Blum, J. D. and Bergquist, B. A., 2007. Reporting of variations in the natural isotopic composition of mercury. Analytical and Bioanalytical Chemistry 388, 353-359.

Bronsted, J. and Hevesy, G., 1920. The Separation of the Isotopes of MercuryNature.

Buchachenko, A. L., 2001. Magnetic isotope effect: Nuclear spin control of chemical reactions. Journal of Physical Chemistry A 105, 9995-10011.

Buchachenko, A. L., Galimov, E. M., Ershov, V. V., Nikiforov, G. A., and Pershin, A. D., 1976. ISOTOPIC ENRICHMENT INDUCED BY MAGNETIC-INTERACTIONS IN CHEMICAL-REACTIONS. Doklady Akademii Nauk Sssr 228, 379-381.

Carignan, J., Estrade, N., Sonke, J. E., and Donard, O. F. X., 2009. Odd Isotope Deficits in Atmospheric Hg Measured in Lichens. Environmental Science & Technology 43, 5660-5664.

Castillo, A., Rodriguez-Gonzalez, P., Centineo, G., Roig-Navarro, A. F., and Alonso, J. I. G., 2010. Multiple Spiking Species-Specific Isotope Dilution Analysis by Molecular Mass Spectrometry: Simultaneous Determination of Inorganic Mercury and Methylmercury in Fish Tissues. Analytical Chemistry 82, 2773-2783.

Craig, H., 1961. ISOTOPIC VARIATIONS IN METEORIC WATERS. Science 133, 1702-&.

Crump, D., Lean, D., Berrill, M., Coulson, D., and Toy, L., 1999. Spectral irradiance in pond water: Influence of water chemistry. Photochemistry and Photobiology 70, 893-901.

Page 56: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

49

Eckley, C. S. and Hintelmann, H., 2006. Determination of mercury methylation potentials in the water column of lakes across Canada. Science of the Total Environment 368, 111-125.

Engel, T. and Reid, P., 2006. Physical Chemistry. Pearson Education, Toronto.

Epov, V. N., Rodriguez-Gonzalez, P., Sonke, J. E., Tessier, E., Amouroux, D., Bourgoin, L. M., and Donard, O. F. X., 2008. Simultaneous determination of species-specific isotopic composition of Hg by gas chromatography coupled to multicollector ICPMS. Analytical Chemistry 80, 3530-3538.

Fitzgerald, W. F., Mason, R. P., and Vandal, G. M., 1991. ATMOSPHERIC CYCLING AND AIR-WATER EXCHANGE OF MERCURY OVER MIDCONTINENTAL LACUSTRINE REGIONS. Water Air and Soil Pollution 56, 745-767.

Foucher, D. and Hintelmann, H., 2006. High-precision measurement of mercury isotope ratios in sediments using cold-vapor generation multi-collector inductively coupled plasma mass spectrometry. Analytical and Bioanalytical Chemistry 384, 1470-1478.

Foucher, D., Ogrinc, N., and Hintelmann, H., 2009. Tracing Mercury Contamination from the Idrija Mining Region (Slovenia) to the Gulf of Trieste Using Hg Isotope Ratio Measurements. Environmental Science & Technology 43, 33-39.

Fritz, P. and Fontes, J. C., 1980. Handbook of Environmental Isotope Geochemistry. Elsevier, New York.

Fujii, T., Moynier, F., and Albarede, F., 2009. The nuclear field shift effect in chemical exchange reactions. Chemical Geology 267, 139-156.

Fujii, Y., Nomura, M., Onitsuka, H., and Takeda, K., 1989a. ANOMALOUS ISOTOPE FRACTIONATION IN URANIUM ENRICHMENT PROCESS. Journal of Nuclear Science and Technology 26, 1061-1064.

Fujii, Y., Okamoto, M., Kadotani, H., and Kakihana, H., 1989b. EQUILIBRIUM TIME AND CRITICALITY CONSIDERATIONS IN URANIUM ENRICHMENT BY THE CHEMICAL-EXCHANGE PROCESS. Nuclear Technology 86, 282-288.

Gantner, N., Hintelmann, H., Zheng, W., and Muir, D. C., 2009. Variations in Stable Isotope Fractionation of Hg in Food Webs of Arctic Lakes. Environmental Science & Technology 43, 9148-9154.

Page 57: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

50

Garcia, E., Amyot, M., and Ariya, P. A., 2005a. Relationship between DOC photochemistry and mercury redox transformations in temperate lakes and wetlands. Geochimica Et Cosmochimica Acta 69, 1917-1924.

Garcia, E., Poulain, A. J., Amyot, M., and Ariya, P. A., 2005b. Diel variations in photoinduced oxidation of Hg-0 in freshwater. Chemosphere 59, 977-981.

Gardfeldt, K., Munthe, J., Stromberg, D., and Lindqvist, O., 2003. A kinetic study on the abiotic methylation of divalent mercury in the aqueous phase. Science of the Total Environment 304, 127-136.

Gehrke, G. E., Blum, J. D., and Meyers, P. A., 2009. The geochemical behavior and isotopic composition of Hg in a mid-Pleistocene western Mediterranean sapropel. Geochimica Et Cosmochimica Acta 73, 1651-1665.

Ghosh, S., Xu, Y. F., Humayun, M., and Odom, L., 2008. Mass-independent fractionation of mercury isotopes in the environment. Geochemistry Geophysics Geosystems 9.

Gilmour, C. C., Henry, E. A., and Mitchell, R., 1992. SULFATE STIMULATION OF MERCURY METHYLATION IN FRESH-WATER SEDIMENTS. Environmental Science & Technology 26, 2281-2287.

Griffith, K. S. and Gellene, G. I., 1992. SYMMETRY RESTRICTIONS IN DIATOM DIATOM REACTIONS .2. NONMASS-DEPENDENT ISOTOPE EFFECTS IN THE FORMATION OF O-4(+). Journal of Chemical Physics 96, 4403-4411.

Hahn, A. A., Miller, J. P., Powers, R. J., Zehnder, A., Rushton, A. M., Welsh, R. E., Kunselman, A. R., and Roberson, P., 1979. EXPERIMENTAL-STUDY OF MUONIC X-RAY TRANSITIONS IN MERCURY ISOTOPES. Nuclear Physics A 314, 361-386.

Hammerschmidt, C. R. and Fitzgerald, W. F., 2006. Photodecomposition of methylmercury in an arctic Alaskan lake. Environmental Science & Technology 40, 1212-1216.

Hammerschmidt, C. R., Lamborg, C. H., and Fitzgerald, W. F., 2007. Aqueous phase methylation as a potential source of methylmercury in wet deposition. Atmospheric Environment 41, 1663-1668.

Harris, R. C., Rudd, J. W. M., Amyot, M., Babiarz, C. L., Beaty, K. G., Blanchfield, P. J., Bodaly, R. A., Branfireun, B. A., Gilmour, C. C., Graydon, J. A., Heyes, A., Hintelmann, H., Hurley, J. P., Kelly, C. A., Krabbenhoft, D. P., Lindberg, S. E., Mason, R. P., Paterson, M. J., Podemski, C. L., Robinson, A., Sandilands, K. A.,

Page 58: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

51

Southworth, G. R., Louis, V. L. S., and Tate, M. T., 2007. Whole-ecosystem study shows rapid fish-mercury response to changes in mercury deposition. Proceedings of the National Academy of Sciences of the United States of America 104, 16586-16591.

Hintelmann, H. and Lu, S. Y., 2003. High precision isotope ratio measurements of mercury isotopes in cinnabar ores using multi-collector inductively coupled plasma mass spectrometry. Analyst 128, 635-639.

Hulston, J. R. and Thode, H. G., 1965. VARIATIONS IN S33 S34 AND S36 CONTENTS OF METEORITES AND THEIR RELATION TO CHEMICAL AND NUCLEAR EFFECTS. Journal of Geophysical Research 70, 3475-&.

Jackson, T. A., Muir, D. C. G., and Vincent, W. F., 2004. Historical variations in the stable isotope composition of mercury in Arctic lake sediments. Environmental Science & Technology 38, 2813-2821.

Jeremiason, J. D., Kanne, L. A., Lacoe, T. A., Hulting, M., and Simcik, M. F., 2009. A comparison of mercury cycling in Lakes Michigan and Superior. Journal of Great Lakes Research 35, 329-336.

Keesom, W. H. and van Dijk, H., 1931. On the possibility of separating neon into its isotopic components by rectification. Proceedings of the Koninklijke Akademie Van Wetenschappen Te Amsterdam 34, 42-50.

Kerin, E. J., Gilmour, C. C., Roden, E., Suzuki, M. T., Coates, J. D., and Mason, R. P., 2006. Mercury methylation by dissimilatory iron-reducing bacteria. Applied and Environmental Microbiology 72, 7919-7921.

Krabbenhoft, D. P., Olson, M. L., Dewild, J. F., Clow, D. W., Striegl, R. G., Dornblaser, M. M., and Vanmetre, P., 2002. Mercury loading and methylmercury production and cycling in high-altitude lakes from the western united states. Water, Air, and Soil Pollution: Focus, 233-249.

Kritee, K., Barkay, T., and Blum, J. D., 2009. Mass dependent stable isotope fractionation of mercury during mer mediated microbial degradation of monomethylmercury. Geochimica Et Cosmochimica Acta 73, 1285-1296.

Kritee, K., Blum, J. D., and Barkay, T., 2008. Mercury Stable Isotope Fractionation during Reduction of Hg(II) by Different Microbial Pathways. Environmental Science & Technology 42, 9171-9177.

Page 59: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

52

Kritee, K., Blum, J. D., Johnson, M. W., Bergquist, B. A., and Barkay, T., 2007. Mercury stable isotope fractionation during reduction of Hg(II) to Hg(0) by mercury resistant microorganisms. Environmental Science & Technology 41, 1889-1895.

Laffont, L., Sonke, J. E., Maurice, L., Hintelmann, H., Pouilly, M., Bacarreza, Y. S., Perez, T., and Behra, P., 2009. Anomalous Mercury Isotopic Compositions of Fish and Human Hair in the Bolivian Amazon. Environmental Science & Technology 43, 8985-8990.

Lalonde, J. D., Amyot, M., Kraepiel, A. M. L., and Morel, F. M. M., 2001. Photooxidation of Hg(0) in artificial and natural waters. Environmental Science & Technology 35, 1367-1372.

Lamborg, C. H., Tseng, C. M., Fitzgerald, W. F., Balcom, P. H., and Hammerschmidt, C. R., 2003. Determination of the mercury complexation characteristics of dissolved organic matter in natural waters with "reducible Hg" titrations. Environmental Science & Technology 37, 3316-3322.

Lauretta, D. S., Klaue, B., Blum, J. D., and Buseck, P. R., 2001. Mercury abundances and isotopic compositions in the Murchison (CM) and Allende (CV) carbonaceous chondrites. Geochimica Et Cosmochimica Acta 65, 2807-2818.

Lehnherr, I. and Louis, V. L. S., 2009. Importance of Ultraviolet Radiation in the Photodemethylation of Methylmercury in Freshwater Ecosystems. Environmental Science & Technology 43, 5692-5698.

Lindberg, S., Bullock, R., Ebinghaus, R., Engstrom, D., Feng, X. B., Fitzgerald, W., Pirrone, N., Prestbo, E., and Seigneur, C., 2007. A synthesis of progress and uncertainties in attributing the sources of mercury in deposition. Ambio 36, 19-32.

Malinovskiy, D., Moens, L., and Vanhaecke, F., 2009. Isotopic Fractionation of Sn during Methylation and Demethylation Reactions in Aqueous Solution. Environmental Science & Technology 43, 4399-4404.

Malinovsky, D., Latruwe, K., Moens, L., and Vanhaecke, F., 2010. Experimental study of mass-independence of Hg isotope fractionation during photodecomposition of dissolved methylmercury. Journal of Analytical Atomic Spectrometry 25, 950-956.

Mariotti, A., Germon, J. C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, A., and Tardieux, P., 1981. EXPERIMENTAL-DETERMINATION OF NITROGEN KINETIC ISOTOPE FRACTIONATION - SOME PRINCIPLES - ILLUSTRATION FOR THE DENITRIFICATION AND NITRIFICATION PROCESSES. Plant and Soil 62, 413-430.

Page 60: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

53

Mason, R. P., Reinfelder, J. R., and Morel, F. M. M., 1996. Uptake, toxicity, and trophic transfer of mercury in a coastal diatom. Environmental Science & Technology 30, 1835-1845.

Mills, R. B., Paterson, A. M., Blais, J. M., Lean, D. R. S., Smol, J. P., and Mierle, G., 2009. Factors influencing the achievement of steady state in mercury contamination among lakes and catchments of south-central Ontario. Canadian Journal of Fisheries and Aquatic Sciences 66, 187-200.

Morris, D. P., Zagarese, H., Williamson, C. E., Balseiro, E. G., Hargreaves, B. R., Modenutti, B., Moeller, R., and Queimalinos, C., 1995. The attentuation of solar UV radiation in lakes and the role of dissolved organic carbon. Limnology and Oceanography 40, 1381-1391.

Munthe, J., Bodaly, R. A., Branfireun, B. A., Driscoll, C. T., Gilmour, C. C., Harris, R., Horvat, M., Lucotte, M., and Malm, O., 2007. Recovery of mercury-contaminated fisheries. Ambio 36, 33-44.

Nier, A. O., 1950. A REDETERMINATION OF THE RELATIVE ABUNDANCES OF THE ISOTOPES OF CARBON, NITROGEN, OXYGEN, ARGON, AND POTASSIUM. Physical Review 77, 789-793.

Nier, A. O. and Schlutter, D. J., 1986. MASS-SPECTROMETRIC STUDY OF THE MERCURY ISOTOPES IN THE ALLENDE METEORITE. Journal of Geophysical Research-Solid Earth and Planets 91, E124-E128.

O'Driscoll, N. J., Lean, D. R. S., Loseto, L. L., Carignan, R., and Siciliano, S. D., 2004. Effect of dissolved organic carbon on the photoproduction of dissolved gaseous mercury in lakes: Potential impacts of forestry. Environmental Science & Technology 38, 2664-2672.

O'Driscoll, N. J., Siciliano, S. D., Lean, D. R. S., and Amyot, M., 2006. Gross photoreduction kinetics of mercury in temperate freshwater lakes and rivers: Application to a general model of DGM dynamics. Environmental Science & Technology 40, 837-843.

Orihel, D. M., Paterson, M. J., Blanchfield, P. J., Bodaly, R. A., Gilmour, C. C., and Hintelmann, H., 2008. Temporal changes in the distribution, methylation, and bioaccumulation of newly deposited mercury in an aquatic ecosystem. Environmental Pollution 154, 77-88.

Poulain, A. J., Amyot, M., Findlay, D., Telor, S., Barkay, T., and Hintelmann, H., 2004. Biological and photochemical production of dissolved gaseous mercury in a boreal lake. Limnology and Oceanography 49, 2265-2275.

Page 61: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

54

Poulain, A. J., Orihel, D. M., Amyot, M., Paterson, M. J., Hintelmann, H., and Southworth, G. R., 2006. Relationship to aquatic between the loading rate of inorganic mercury ecosystems and dissolved gaseous mercury production and evasion. Chemosphere 65, 2199-2207.

Rodriguez-Gonzalez, P., Epov, V. N., Bridou, R., Tessier, E., Guyoneaud, R., Monperrus, M., and Amouroux, D., 2009. Species-Specific Stable Isotope Fractionation of Mercury during Hg(II) Methylation by an Anaerobic Bacteria (Desulfobulbus propionicus) under Dark Conditions. Environmental Science & Technology 43, 9183-9188.

Rolfhus, K. R., Sakamoto, H. E., Cleckner, L. B., Stoor, R. W., Babiarz, C. L., Back, R. C., Manolopoulos, H., and Hurley, J. P., 2003. Distribution and fluxes of total and methylmercury in Lake Superior. Environmental Science & Technology 37, 865-872.

Schaefer, J. K. and Morel, F. M. M., 2009. High methylation rates of mercury bound to cysteine by Geobacter sulfurreducens. Nature Geoscience 2, 123-126.

Schauble, E. A., 2007. Role of nuclear volume in driving equilibrium stable isotope fractionation of mercury, thallium, and other very heavy elements. Geochimica Et Cosmochimica Acta 71, 2170-2189.

Schueler, B., Morton, J., and Mauersberger, K., 1990. MEASUREMENT OF ISOTOPIC ABUNDANCES IN COLLECTED STRATOSPHERIC OZONE SAMPLES. Geophysical Research Letters 17, 1295-1298.

Scully and Lean, D., 1994. The attenuation of UV radiation in temperate lakes. Arch. Hydrobiol. Beih.

Sellers, P., Kelly, C. A., and Rudd, J. W. M., 2001. Fluxes of methylmercury to the water column of a drainage lake: The relative importance of internal and external sources. Limnology and Oceanography 46, 623-631.

Sellers, P., Kelly, C. A., Rudd, J. W. M., and MacHutchon, A. R., 1996. Photodegradation of methylmercury in lakes. Nature 380, 694-697.

Selvendiran, P., Driscoll, C. T., Montesdeoca, M. R., Choi, H. D., and Holsen, T. M., 2009. Mercury dynamics and transport in two Adirondack lakes. Limnology and Oceanography 54, 413-427.

Senn, D. B., Chesney, E. J., Blum, J. D., Bank, M. S., Maage, A., and Shine, J. P., 2010. Stable Isotope (N, C, Hg) Study of Methylmercury Sources and Trophic Transfer

Page 62: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

55

in the Northern Gulf of Mexico. Environmental Science & Technology 44, 1630-1637.

Sherman, L. S., Blum, J. D., Johnson, K. P., Keeler, G. J., Barres, J. A., and Douglas, T. A., 2010. Mass-independent fractionation of mercury isotopes in Arctic snow driven by sunlight. Nature Geoscience 3, 173-177.

Sherman, L. S., Blum, J. D., Nordstrom, D. K., McCleskey, R. B., Barkay, T., and Vetriani, C., 2009. Mercury isotopic composition of hydrothermal systems in the Yellowstone Plateau volcanic field and Guaymas Basin sea-floor rift. Earth and Planetary Science Letters 279, 86-96.

Smith, C. N., Kesler, S. E., Blum, J. D., and Rytuba, J. J., 2008. Isotope geochemistry of mercury in source rocks, mineral deposits and spring deposits of the California Coast Ranges, USA. Earth and Planetary Science Letters 269, 398-406.

Smith, C. N., Kesler, S. E., Klaue, B., and Blum, J. D., 2005. Mercury isotope fractionation in fossil hydrothermal systems. Geology 33, 825-828.

Thiemens, M. H., 2006. History and applications of mass-independent isotope effects. Annual Review of Earth and Planetary Sciences 34, 217-262.

Thiemens, M. H. and Heidenreich, J. E., 1983. THE MASS-INDEPENDENT FRACTIONATION OF OXYGEN - A NOVEL ISOTOPE EFFECT AND ITS POSSIBLE COSMOCHEMICAL IMPLICATIONS. Science 219, 1073-1075.

Thiemens, M. H. and Jackson, T., 1987. PRODUCTION OF ISOTOPICALLY HEAVY OZONE BY ULTRAVIOLET-LIGHT PHOTOLYSIS OF O-2. Geophysical Research Letters 14, 624-627.

Turro, N., Scaiano, J. C., and Ramamurthy, V., 2008. Principles of Molecular Photochemistry: An Introduction. University Science Books.

Turro, N. J., 1983. INFLUENCE OF NUCLEAR-SPIN ON CHEMICAL-REACTIONS - MAGNETIC ISOTOPE AND MAGNETIC-FIELD EFFECTS (A REVIEW). Proceedings of the National Academy of Sciences of the United States of America-Physical Sciences 80, 609-621.

Urey, H. C., 1947. THE THERMODYNAMIC PROPERTIES OF ISOTOPIC SUBSTANCES. Journal of the Chemical Society, 562-581.

Urey, H. C., Brickwedde, F. G., and Murphy, G. M., 1932. Relative abundance of H-1 and H-2 in natural hydrogen. Physical Review 40, 0464-0465.

Page 63: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

56

Urey, H. C., Epstein, S., McKinney, C., and McCrea, J., 1948. METHOD FOR MEASUREMENT OF PALEOTEMPERATURES. Geological Society of America Bulletin 59, 1359-1360.

Vandal, G. M., Mason, R. P., and Fitzgerald, W. F., 1991. CYCLING OF VOLATILE MERCURY IN TEMPERATE LAKES. Water Air and Soil Pollution 56, 791-803.

Watras, C. J., Bloom, N. S., Claas, S. A., Morrison, K. A., Gilmour, C. C., and Craig, S. R., 1995a. METHYLMERCURY PRODUCTION IN THE ANOXIC HYPOLIMNION OF A DIMICTIC SEEPAGE LAKE. Water Air and Soil Pollution 80, 735-745.

Watras, C. J. and Morrison, K. A., 2008. The response of two remote, temperate lakes to changes in atmospheric mercury deposition, sulfate, and the water cycle. Canadian Journal of Fisheries and Aquatic Sciences 65, 100-116.

Watras, C. J., Morrison, K. A., Host, J. S., and Bloom, N. S., 1995b. CONCENTRATION OF MERCURY SPECIES IN RELATIONSHIP TO OTHER SITE-SPECIFIC FACTORS IN THE SURFACE WATERS OF NORTHERN WISCONSIN LAKES. Limnology and Oceanography 40, 556-565.

Wiatrowski, H. A., Ward, P. M., and Barkay, T., 2006. Novel reduction of mercury(II) by mercury-sensitive dissimilatory metal reducing bacteria. Environmental Science & Technology 40, 6690-6696.

Wickman, F. E., 1952. VARIATIONS IN THE RELATIVE ABUNDANCE OF THE CARBON ISOTOPES IN PLANTS. Geochimica Et Cosmochimica Acta 2, 243-254.

Wiederhold, J. G., Cramer, C. J., Daniel, K., Infante, I., Bourdon, B., and Kretzschmar, R., 2010. Equilibrium Mercury Isotope Fractionation between Dissolved Hg(II) Species and Thiol-Bound Hg. Environmental Science & Technology 44, 4191-4197.

Xia, K., Skyllberg, U. L., Bleam, W. F., Bloom, P. R., Nater, E. A., and Helmke, P. A., 1999. X-ray absorption spectroscopic evidence for the complexation of Hg(II) by reduced sulfur in soil humic substances. Environmental Science & Technology 33, 257-261.

Xia, K., Weesner, F., Bleam, W. F., Bloom, P. R., Skyllberg, U. L., and Helmke, P. A., 1998. XANES studies of oxidation states of sulfur in aquatic and soil humic substances. Soil Science Society of America Journal 62, 1240-1246.

Page 64: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

57

Yang, L. and Sturgeon, R., 2009. Isotopic fractionation of mercury induced by reduction and ethylation. Analytical and Bioanalytical Chemistry 393, 377-385.

Young, E. D., Galy, A., and Nagahara, H., 2002. Kinetic and equilibrium mass-dependent isotope fractionation laws in nature and their geochemical and cosmochemical significance. Geochimica Et Cosmochimica Acta 66, 1095-1104.

Zadnik, M. G., Specht, S., and Begemann, F., 1989. REVISED ISOTOPIC COMPOSITION OF TERRESTRIAL MERCURY. International Journal of Mass Spectrometry and Ion Processes 89, 103-110.

Zambardi, T., Sonke, J. E., Toutain, J. P., Sortino, F., and Shinohara, H., 2009. Mercury emissions and stable isotopic compositions at Vulcano Island (Italy). Earth and Planetary Science Letters 277, 236-243.

Zhang, T. and Hsu-Kim, H., 2010. Photolytic degradation of methylmercury enhanced by binding to natural organic ligands. Nature Geoscience 3, 473-476.

Zheng, W., Foucher, D., and Hintelmann, H., 2007. Mercury isotope fractionation during volatilization of Hg(0) from solution into the gas phase. Journal of Analytical Atomic Spectrometry 22, 1097-1104.

Zheng, W. and Hintelmann, H., 2009. Mercury isotope fractionation during photoreduction in natural water is controlled by its Hg/DOC ratio. Geochimica Et Cosmochimica Acta 73, 6704-6715.

Zheng, W. and Hintelmann, H., 2010a. Isotope Fractionation of Mercury during Its Photochemical Reduction by Low-Molecular-Weight Organic Compounds. Journal of Physical Chemistry A 114, 4246-4253.

Zheng, W. and Hintelmann, H., 2010b. Nuclear Field Shift Effect in Isotope Fractionation of Mercury during Abiotic Reduction in the Absence of Light. Journal of Physical Chemistry A 114, 4238-4245.

Page 65: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

58

CHAPTER 3: EFFECTS OF ULTRAVIOLET

RADIATION ON THE MAGNETIC ISOTOPE

EFFECT IN FRESHWATER PHOTO-REDUCTION OF ORGANIC AND INORGANIC

MERCURY 3.1. INTRODUCTION

The toxicity of various forms of mercury (Hg) is well established. The form that is of

most concern is monomethylmercury (MeHg), which is a potent neurotoxin that

bioaccumulates in aquatic food chains with concentrations in piscivorous fish reaching

levels one million times greater than in the surrounding water (Lindqvist et al., 1991;

Mergler et al., 2007). Humans have increased the amount of mercury that is actively

cycling in the environment by mining, industrial activities and combustion of fossil fuels

by a factor of approximately three since pre-industrial times. Once released to the

environment, Hg undergoes redox, biological, and phase transformations. Its elemental

form is volatile and is transported globally in the atmosphere. Uncertainty about the

origin and magnitude of primary emissions to the atmosphere, deposition to terrestrial

and ocean surfaces and re-emission of previously deposed Hg presents challenges to

scientists and policy-makers alike (Fitzgerald and Lamborg, 2007; Lindberg et al., 2007;

Selin, 2009).

Stable isotope fractionation has been successfully used to trace biogeochemical processes

at both human and geologic timescales. Mercury has seven stable isotopes: 196Hg

(0.16%) 198Hg (10.0%) 199Hg (16.9%), 200Hg (23.1%), 201Hg (13.2%), 202Hg (29.7%) and 204Hg(6.8%). The recent development of an analytical technique using cold vapor

multicollector inductively coupled plasma mass spectrometry (CV-MC-ICP, (Lauretta et

al., 2001) has reduced measurement uncertainty (typically 0.1 to 0.3 ‰, (Bergquist and

Blum, 2009)to a point where variations in the isotope compositions of natural samples are

readily resolved for Hg. The reported range of isotopic signatures imparted by mass-

Page 66: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

59

dependent fractionation (MDF) already exceeds 7‰ (Bergquist and Blum, 2009;

Sherman et al., 2010) and MDF has recently been employed to track sources of Hg in the

environment (Feng et al., 2010; Foucher et al., 2009). Mass-independent fractionation

(MIF) of the odd isotopes of Hg has also been observed, with a reported range of

signatures spanning nearly 10 ‰ (Bergquist and Blum, 2009; Sherman et al., 2010). Such

large MIF is unusual in isotopic systems, with O (during ozone formation) and S (only in

the Archean) being the only other elements showing MIF of comparable magnitude in

natural samples.

Processes that induce MDF for Hg include hydrothermal transport (Sherman et al., 2009;

Smith et al., 2005), biological reduction (Kritee et al., 2009; 2008; 2007) , abiotic redox

transformations (Bergquist and Blum, 2007; Yang and Sturgeon, 2009; Zheng and

Hintelmann, 2009; Zheng and Hintelmann, 2010a; Zheng and Hintelmann, 2010b) and

evaporation (Estrade et al., 2009; Zheng et al., 2007). One challenge to using MDF as a

tool to elucidate the transformation and fate of Hg in the environment is distinguishing

source signatures from MDF imparted during transport and subsequent transformations.

Processes reported to induce MIF are fewer and large MIF (> 0.3‰) has only been

observed in photochemical reduction (Bergquist and Blum, 2007; Malinovsky et al.,

2010; Sherman et al., 2010; Zheng and Hintelmann, 2009; Zheng and Hintelmann,

2010b) and now photo-oxidation (Ghosh). While MIF is not affected by MDF processes,

it could be altered by other MIF processes or through mixing of Hg pools bearing

different MIF signatures (Bergquist and Blum, 2009). Understanding the factors that

control MIF will permit MIF to be utilized to track and quantify specific pathways in

nature. It will also permit assessment of which signatures are susceptible to alteration by

successive transformations or mixing and whether the various contributions to MIF can

be deciphered.

Of the two plausible mechanisms, the nuclear volume effect (NVE) and the magnetic

isotope effect (MIE), put forward to explain mass-independent anomalies in Hg isotopes,

the more probable cause of the large MIF observed in photochemical reactions is the MIE

(Bergquist and Blum, 2007; Bergquist and Blum, 2009; Malinovsky et al., 2010). Only

Page 67: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

60

the odd isotopes are magnetic, which is to say, have non-zero nuclear spin and nuclear

magnetic moments that can interact with electron magnetic moments (hyperfine

coupling). This interaction can become important where the orientation of electron spin

affects the reaction pathway, as in the fate of a geminate excited radical pair. If the

unpaired electrons are in a triplet state (unpaired spins) with respect to each other, then

the forming of a covalent bond is spin-forbidden (total electron spin would not be

conserved) and the radicals are likely to drift apart. Hyperfine coupling can cause

electron spin conversion (intersystem crossing, ISC), such that the unpaired electrons

convert from a triplet state into a singlet state (spin paired). From a singlet state,

recombination of the radical pair in a covalent bond is spin-allowed. The difference in

the rates of spin conversion for the odd versus the even isotopes sorts the nuclei into

different products. Odd nuclei are more likely to be found in the recombination product

due to faster ISC and even nuclei are more likely to dissociate spatially and meet other

fates. Other forces that influence the electron spin orientation, such as coupling between

electron spin and orbital angular momenta (spin-orbital coupling (SOC)), can also affect

expression of the MIE. Another factor that is important in the expression of the MIE is

the lifetime of the radical pair. If the lifetime is very short relative to ISC, the MIE will

be suppressed. If it is very long relative to ISC, spin-conversion may take place by other

processes and nuclear sorting of the odd and even isotopes will not be consistent

(Bergquist and Blum, 2009; Buchachenko, 2001).

This mechanism was first proposed by Bergquist and Blum 2007 to explain MIF of over

2 ‰ observed during the photochemical reduction of aqueous Hg species in the presence

of dissolved organic matter and MIF of over 4 ‰ in Hg isolated from fish. While it has

been widely adopted by subsequent investigators to explain mass-independent anomalies,

the only direct test of the mechanism to date has been by Malinovsky et al. 2009. They

found that MIF during photodemethylation of MeHg was suppressed in solutions with

ascorbic acid and with ammonia, attributed to suppression or scavenging of the radical

pair during decomposition. Expression of the MIE is dependent on many factors, a few

of which are mentioned above. Another test of the mechanism is elimination of the

photo-generated radical pair. If the higher energy wavelengths of the solar spectrum are

Page 68: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

61

removed gradually, the radiation available will eventually not carry enough energy to

effect the electronic transition to an anti-bonding orbital and radical pairs will not be

produced. This closes the avenue of expression for the MIE. The objective of this study

is the following: to determine whether removing the ultraviolet B (290-320 nm) or the

ultraviolet A (320-400 nm) portions of the electromagnetic spectrum affects the

expression and magnitude of MIF during the aqueous photo-reduction of MeHg and

Hg(II) respectively. Other goals were to determine how MIF is related to total photo-

reduction and to determine whether MIF wil allow us to distinguish photo-reduction from

other processes that reduce Hg in aqueous freshwater environment.

Page 69: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

62

3.2. METHODS

3.2.1. Cleaning equipment

All sample containers and analytical vials were glass and cleaned according to the

following procedure:

(1) glass vials were rinsed five times with Milli-Q (18 M) water

(2) glass vials were soaked in cold 10% HCl for 48 hours

(3) glass vials were rinsed five times with Milli-Q water along with the teflon-lined

caps

(4) glass vials were then filled with 2-3% BrCl (acidic and a complexing agent for

Hg), capped and left to stand at least 12 hours right side up and 12 hours inverted to

clean the entire inner surface.

(5) five final Milli-Q water rinses (including caps)

(6) drying in a class 100 laminar workstation (which was equipped with a carbon

filtration unit prior to the HEPA filter to remove Hg(0) from air) followed by capping

and storage in clean, sealed plastic bags.

If the 48-hour cold HCl soak was not feasible, it was substituted with a preliminary BrCl

soak. The quartz photochemical reactors and associated teflon tubing and fittings were

rinsed five times with Milli-Q water, filled with 2-3% BrCl and left to soak for at least 12

hours and then rinsed copiously with Milli-Q water before use.

3.2.2. Photo-reduction experiments

Two sets of three photochemical reduction experiments each were carried out using a

Hg(NO3)2 standard (J.T. Baker lot E04632, hereafter called Hg(II)). Two sets of three

photochemical reduction experiments each were carried out using a CH3HgCl standard

(Alfa Aesar lot 82-082844A, hereafter called MeHg). Photo-reduction took place under

natural sunlight on the roof of the McLennan Physical Building, 60 St. George Street,

Toronto, Ontario. (43 N, 79 W). The initial set for each species was carried out in July,

2009 and the replicate set was carried out in September, 2009

Aqueous mercury solutions

Page 70: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

63

Aqueous solutions of 2 mg/L Suwannee River Fulvic Acid Standard (1S101F) from the

International Humic Substance Society (http://ihss.gatech.edu/ihss2/) were made up in

glass bottles in simulation of natural fresh waters and sampled to verify negligible

mercury content. The fulvic acid solutions were then spiked to a final concentration of

approximately 30 g/L of the Hg species under study, and immediately subsampled to

measure the initial isotopic composition. These experimental Hg concentrations are four

to five orders of magnitude higher than in natural freshwaters, but were chosen as the

minimum concentrations at which isotopic composition could be measured with

confidence.

The Hg-fulvic acid solutions were left to equilibrate overnight in aluminum foil, which

was not removed until the moment of filling the quartz photochemical reactors. For UV-

filtration experiments, the filtration boxes were placed over the reactors immediately after

filling and the solution was sampled to monitor any change in concentration or isotopic

composition that occurred overnight. All subsamples for isotopic analysis were

immediately preserved by spiking with concentrated BrCl to achieve a sample

concentration of 2.5%.

Photoreactor setup

Photochemical reactors were 1 L quartz Erlenmeyer flasks with air input and sample

withdrawal lines running through a teflon cap and an arm for headspace purging, as

depicted in Figure 2.1. Before entering the reaction vessel, the air passed through a 0.2

m filter to remove particulate matter and a gold-covered sand trap to remove any

ambient Hg(0). Experiments were started between 9:30 and 10:30 a.m. and allowed to

run for several hours, through the solar noon (approximately 1 p.m.). Sub-samples were

removed from the aqueous reservoir with a syringe as the reaction progressed and

preserved immediately with Hg-clean 2.5% BrCl. The sum of volumes sampled never

exceeded half the reservoir volume.

Light filtration

Page 71: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

64

Light filtration was accomplished by building boxes, closed on five sides, which stood

over the quartz reactors. Mylar (poly(ethylene terphthalate) sheets were used to

selectively screen the UVB portion of the electromagnetic spectrum (290-320 nm); five

sheets per side) were attached to a 0.5-inch acrylic frame. Mylar’s transmission of

ultraviolet light is negligible at 305 nm and less than 30% at 310 nm but rises quickly to

over 60% at 315 nm (Mitrofanov et al., 2009; Mizuno et al., 2008). It is not completely

transparent to the UVA spectrum, and in removing the majority of the UVB spectrum

(93-95%), some of the UVA spectrum was also lost (320 – 400 nm, 43-45%). A small

amount of the visible spectrum was also lost (400 - 700 nm, 19-22%). To selectively

screen both the UVB and the UVA portions of the spectrum, solid acrylic sheets covered

with a 3M Scotchshield (Ultra S150) window film were used. This consistently blocked

96% of UVB, 98% of the UVA, and a small amount the visible spectrum when the sun

was not directly overhead. Around mid-day, the intensity of visible light was higher

under the acrylic box (by 30% or more) than the intensity of visible light under direct

solar radiation. Measurements of the light reaching each reactor were taken four or five

times throughout the experiment, with the three readings for each electromagnetic region

taken in as rapid a sequence as possible (Table 3.1). Measurements were taken with a

Solar Light PMA2200 Radiometer fitted with detectors specific to the three spectral

regions of interest to us: UVB (PMA2106), UVA (PMA2110) and Visible (PMA2130).

Dark controls

Dark control experiments for Hg2+ and MeHg were performed in the laboratory with the

same reactor configuration except that the quartz reactor was covered with aluminum

foil. Initial dark controls (Number 1 for both species) showed no concentration loss

during 8 h of continuous sparging. Subsequent dark controls for Hg(II) left in a sealed,

foil-wrapped bottle with no sparging over a few weeks showed significant concentration

losses (ex. Hg(II) Dark Control 2, 40 % loss). Dark control experiments for both Hg(II)

and MeHg, with the quartz reactor wrapped in foil and continuous sparging, were

performed outdoors in Michigan by B. Bergquist, as published in Bergquist and Blum

2007.

Page 72: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

65

3.2.3. Hg concentration analysis

Concentrations of Hg in the aqueous samples were measured by a Tekran 2600 Cold

Vapor Atomic Fluorescence Spectrometer (CVAFS). Samples were introduced through a

gas-liquid separator modified by B. Bergquist. The method was an adaptation of US

EPA Method 1631, Revision E. The BrCl (or KMnO4 for trap solutions, discussed

hereafter) preserving the solutions was reduced with hydroxylamine hydrochloride

(prepared fresh monthly) prior to analysis. This reduces the oxidizing matrix promptly,

but takes weeks to reduce the Hg. After uptake, mercury in solution was reduced on-line

by a Sn(II)Cl2 solution. The dissolved Hg(0) that was generated was stripped from the

solution by Hg-free argon through the gas-liquid separator and concentrated on a series of

two gold traps before being released to the detector. To maximize signal stability and

reproducibility, all tubing, connections and the phase-separating surface were routinely

cleaned and changed immediately if any sort of precipitate (including organic matter and

tin oxide) was suspected of accumulating on the interior. Ten minutes of apparatus

washing with a 7% (each) HNO3 and HCl aqueous solution at the end of each analytical

session was found to prolong the useful lifetime of tubing and connections, and to

improve signal stability in general. Instrument operation and maintenance was still being

optimized during the July analyses and typical precisions (± 4.9 %, Hg(II), 2SE and ±

3.76 %, MeHg, 2SE) are poorer than in September (± 2.8 % , Hg(II), 2SE and 2.8 %,

MeHg, 2SE). Calibration accuracy and instrument drift were monitored with a reference

mercury solution prepared from NIST 1641d. Reference standard accuracy varied from

5% to ±1%, 2SD. This accuracy was based on the first two hours of sample run time

only (n = 4 or more). This metric was chosen because the Tekran 2600 could (but did not

always) drift by 5% to 8% during an analytical session. This drift was monitored closely

with the reference standard so that concentrations could be corrected. Drift was not

significant during the first two hours of analysis. Errors shown on graphs are the greater

of average sample 2SE and standard 2SD for one analytical session. Method blanks were

run prior to calibration and periodically throughout the analysis.

3.2.4. Mercury purifcation and matrix transfer in preparation for isotope analysis

Page 73: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

66

Any variability in instrumental mass bias introduced into isotopic analysis through the

matrix effects of natural samples is not consistent enough to be corrected for. To obtain

reliable isotope compositions, therefore, it was necessary to transfer the Hg from complex

sample matrices to an aqueous acidic matrix consisting of 10% (wt) H2SO4 and 1% (wt)

KMnO4. This was done by cold vapor generation in a manner similar to that described

for concentration analysis by reduction with SnCl2 and the use of a gas-liquid separator

(Figure 2.2). The Hg(0) generated was trapped by sparging into the acidic permanganate

solution. Through a process of recovery optimization, it was found that a gas flow rate

(ultrapure N2) of approximately 170 mL/min streaming into 35 mL of trapping solution

yielded the best results. The diameter of the trapping vials was approximately 22 mm.

If less trapping solution was desired in order to concentrate the sample, the gas flow rate

had to be reduced to maintain recoveries above 95%. Trap reagents were tested for

negligible Hg content (less than 0.05% of sample concentrations). The apparatus was

cleared of each successive sample with a minimum of 15 minutes of washout time with

10% (wt) HNO3 followed by 10% (wt) H2SO4. A heat gun was applied to the transfer

lines to desorb Hg(0) stuck to the interior and Teflon lines, and spargers were rinsed

between samples with dilute hydroxylamine hydrochloride, soaked in 2.5 % BrCl for 15-

30 minutes and finally rinsed in 10% H2SO4. Blanks during this matrix transfer were on

the order of 0.25 % of sample concentrations. Matrix transfer process standards using

NIST 3133 reference mercury solution (concentration- and matrix-matched to samples)

were run routinely at the beginning of each session and then throughout the day as

needed. Sample recoveries after matrix transfer were rarely lower than 95% and never

less than 90% (see Table 2.6). While trouble-shooting the apparatus, process standard

recoveries were sometimes lower than this. Process standards with recoveries ranging

from 86% to 100% were analyzed for isotopic composition to monitor how incomplete

recovery affected isotopic fractionation. The lowest recovery (86%) showed small

offsets (less than 0.2 ¥) in mass-dependent fractionation, but no offset in mass-

independent fractionation. Transfer replicates (indeed, overall process replicates) were

achieved by processing R0 from each of the three experiments from both sets. These

were all prepared from the same Hg(II) and MeHg stock respectively and sampled before

Page 74: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

67

Hg-DOM equilibration and transfer to the quartz reactor. Thus they should represent the

experimental reproducibility in terms of isotopic composition.

3.2.5. MC-ICP-MS analysis

Isotopic compositions were measured on a Nu Instruments multiple collector inductively

coupled plasma mass spectrometer (MC-ICP-MS) at the Department of Geological

Sciences, University of Michigan. Samples were introduced to the plasma by the cold

vapor generation technique already described. Instrumental mass bias was corrected in

two ways as recommended in (Blum and Bergquist, 2007). First, the isotopic

composition of a thallium internal standard (NIST 997; reported 205Tl/203Tl ratio of

2.38714) added to the Hg vapor by a CETAC Aridus desolvating nebulizer was

monitored and used to determine the instrumental mass bias factor using an exponential

fractionation law. This factor was used to correct Hg isotope abundances. Second,

sample-standard bracketing (SSB) was carried out with a reference Hg solution of NIST

SRM 3133 that was matched to the samples both in concentration (5 ppb, usually to

within 5%) and in matrix (1% KMnO4 in 10% H2SO4). Prior to cold vapor generation,

the matrix was neutralized with hydroxylamine hydrochloride a minimum of one hour

(but usually much longer) before sampling. On-peak zero corrections were applied to all

masses by measuring a matrix matched blank solution prior to each sample after 10 to 12

minutes of washout. Isobaric interference of 204Pb on 204Hg was corrected by monitoring 206Pb. The Faraday cups were configured to collect masses of 196, 198, 199, 200, 201,

202, 203, 204, 205 and 206 atomic mass units.

Isotopic compositions are reported in delta notation as the permil (‰) deviation from the

NIST SRM 3133 standard. It is determined through SSB as follows, where the

(XXXHg/198Hg)NIST SRM 3133) ratio in the denominator is the average ratio of the two

bracketing standards run prior to and after the sample:

XXXHg = [(XXXHg/198Hg)sample/(XXXHg/198Hg)NIST SRM 3133) – 1] x 1000

Page 75: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

68

where XXX is the mass the Hg isotope of interest from 198 to 204 amu. This convention

uses the lowest-mass major isotope (196Hg has an abundance of 0.15% and cannot be

measured as precisely as 198Hg, with an abundance of 10.0%) in the denominator. Thus

higher values (more positive) will indicate a sample is “isotopically heavier” than the

SRM, as is the convention in other stable isotope systems.

Mass-independent fractionation is reported using capital delta notation (XXXHg) as the

deviation in isotope ratios from the theoretical values predicted by mass-dependent

fractionation laws, also in units of permil (‰). The theoretical isotope ratios are

determined using the 202Hg/198Hg ratio (i.e. 202Hg) and applying the kinetic mass-

dependent fractionation law derived by Bigeleisen 1949 and reviewed by (Young et al.,

2002)as follows:

199Hg = 1000 x { ln[199Hg/1000 + 1] - 0.2520 x ln[202Hg/1000 +1] }

200Hg = 1000 x { ln[200Hg/1000 + 1] - 0.5024 x ln[202Hg/1000 +1] }

201Hg = 1000 x { ln[201Hg/1000 + 1] - 0.7520 x ln[202Hg/1000 +1] }

204Hg = 1000 x { ln[204Hg/1000 + 1] - 1.493 x ln[202Hg/1000 +1] }.

For values lower than 5 ‰, these may be approximated by:

199Hg = 199Hg - (0.2520 x 202Hg)

200Hg = 200Hg - (0.5024 x 202Hg)

201Hg = 201Hg - (0.7520 x 202Hg)

204Hg = 204Hg - (1.493 x 202Hg)

(all equations from (Blum and Bergquist, 2007).

3.2.6. Reporting of analytical uncertainty

External reproducibility of the method was monitored with a standard of different

composition than the bracketing standard and reported as 2SD (since it is the distribution

Page 76: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

69

of measurements about a fixed value). At the University of Michigan Biogeochemistry

and Environmental Isotope Geochemistry Laboratory, this secondary standard solution

was made from metallic Hg mined from Almadén, Spain (UM-Almadén). The data in

this thesis was produced over 8 analytical sessions (October 2009 –July 2010). The mean

Almadén values for the October through January sessions were 202Hg = -0.58 ± 0.13 ‰

and 199Hg value of 0.00 ± 0.07 ‰ (2SD, n = 12). The mean Almadén values for the

June and July sessions were 202Hg = -0.58 ± 0.09 ‰ and 199Hg value of -0.02 ± 0.07 ‰

(2SD, n = 20). These values are within error of the values for UM-Almadén cited in

(Bergquist and Blum, 2007)for an aqueous BrCl sample matrix and for solutions with

higher Hg concentrations (30 ppb): 202HgAlmaden =-0.54 ± 0.08 ‰ (2SD, n = 25) and

199HgAlmaden of -0.01 ± 0.05 ‰ (2SD, n=25). Our uncertainties are likely larger due to

the different matrix and lower concentrations used for analyses (5 ppb).

The mean isotopic values for JTBaker Hg(II), taken from 16 samples of R0 over 5

analytical sessions were 202Hg = -0.69 ± 0.18 (2SD) and 199Hg = 0.03 ± 0.10 ‰

(2SD). The mean isotopic values for Alfa Aesar MeHg, taken from 12 samples of R0

over 4 analytical sessions were 202Hg = -0.17 ± 0.10 ‰ (2SD) and 199Hg = 0.10 ±

0.10‰ (2SD).

3.2.7. Calculations

Photo-reduction of both Hg(II) and MeHg in these experiments appears to follow a

Rayleigh distillation model well. Rayleigh models describe the isotopic composition of a

substrate-reactant system under equilibrium isotope separation or under kinetic isotope

separation where the product is removed instantaneously after its formation (Fritz and

Fontes, 1980). Kinetic fractionation factors were determined from the results of our

experiments using the form of the Rayleigh distillation equation outlined by (Mariotti et

al., 1981) Briefly, with 199Hg as an example, 199Hg is defined in reference to the

abundance of the isotope of interest:

Page 77: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

70

199Hg

d199HgPr oduct

d198HgPr oduct199HgReac tan t198HgReac tan t

.

With as series of substitutions, it can be shown that 199Hg

d198HgR198HgR

d199HgR199HgR

.

Assuming is constant during the reaction and integrating both sides, one obtains

199Hg ln198HgR , f198HgR ,i

ln

199HgR , f199HgR ,i

where i and f refer to the initial abundance and the abundance after some time

respectively. Through another series of substitutions, it can be shown that

i

i

f

f

i

f

i

fHg

Hg

Hg

Hg

Hg

Hg

Hg

Hg

Hg

198

199

198

199

198

198

198

198

199 lnlnln .

In delta notation and assuming that 198Hgf/198Hgi can be approximated by f, the fraction

of substrate remaining, this equation can be re-written for 199Hg (or any other isotope) as

1 ln f ln103substrate, f 1

103substrate,i 1

.

.

The fractionation factor, , can thus be obtained from the slope of a plot of ln(f) against

the right side of the previous equation. Conversely, a value of allows one to predict the

isotopic signature of (in this case) the substrate for a given fraction remaining:

e(-1)lnf

(i + 1000) – 1000 = f.

Fractionation factors reported in this thesis were obtained in this way from experimental

data.

Page 78: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

71

3.3. RESULTS

Quartz reactors containing either Hg(II) or MeHg in aqueous solutions with organic

matter were simultaneously exposed to three different light regimes to examine the effect

of photo-reduction on the isotopic fractionation of the respective Hg species. Initial sets

of three were carried out on July 19th (MeHg) and July 30th (Hg(II)), 2009 and replicate

sets of three two months later on September 1st (MeHg) and September 3rd (HgII), 2009.

Dark controls, wrapped in aluminum foil throughout the experiment, were also

performed. Solar radiation intensity (W/m2) in each of three wavebands, UVB (290-320

nm), UVA (380-400 nm) and Visible (400-700 nm) was measured at intervals throughout

the experiments (see Table 3.1).

3.3.1. Concentrations

Photo-reduction of Hg(II) and MeHg resulted in highly volatile Hg(0), which was

continuously purged from the substrate reservoir by sparging with Hg-free air. The

amount of Hg(II) or MeHg lost via photo-reduction was estimated from the decreasing

concentration of Hg in the substrate reservoir over time (Figure 3.3). Blocking different

wavebands did not always have the same effect on September photo-reduction as it did

on July photo-reduction. Ultraviolet intensities in September were generally lower, for

example, the maximum UVB intensities (recorded near solar noon, which is

approximately 1 pm) are 15 and 30% lower than the corresponding July intensities and

the maximum September UVA intensities are 10 and 20% lower than in July. Visible

waveband intensities were not significantly lower in September than they were in July.

Hg(II)

In July, Hg(II) showed maximum similar losses of 80% in both the experiment exposed

to full sun and the experiment with UVB blocked (Figure 3.3a). Blocking both the UVB

and the UVA suppressed most of the photo-reduction (only 10% loss). In contrast,

blocking both the UVB and the UVA in September did not suppress photo-reduction

nearly as much (40% loss). Maximum September loss (70%) was in the full spectrum

experiment, and blocking of UVB resulted in approximately 40% loss. Dark control

experiments were carried out in the laboratory with the Hg(II)-organic matter solution

Page 79: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

72

shielded with aluminum foil at all times. They showed no detectable losses after 8.5

hours of sparging with Hg-free air (for example, Dark Control 1, Figure 3.3a). Thus, we

conclude that Hg loss due to non-photochemical reduction during the 6-7 hour UV

filtration experiments was negligible.

In contrast, an Hg-organic matter solution kept in a foil-wrapped glass bottle with no

sparging showed 20% loss after 2 days and 40% loss after 8 days (Dark Control 2,

Figure 3.3a), indicating that while non-photochemical pathways can contribute to Hg

volatilization, the kinetics of these are far slower than the kinetics of photochemical

pathways. An additional dark control with Hg(II) in a foil-wrapped photochemical

reactor with continuous sparging was carried out by B. Bergquist (as published in

(Bergquist and Blum, 2007) that showed approximately 20% loss after 5.5 hours.

In the September experiments with UVB and UVA removed (Figure 2.3b), the lowest

reservoir Hg concentrations are recorded after approximately four hours and are followed

by small concentration increases during the last two hours of exposure. It is possible that

this is a result of reservoir evaporation, as the water vapor would have been pumped out

of the reactor headspace along with the Hg(0). It is likely that the UV filtering boxes

increased the ambient temperature as compared with the uncovered, full sun experiment.

No such Hg concentration increase is seen in the September full spectrum trial, which

may reflect a lower ambient temperature. Unfortunately, temperature under the boxes in

contrast to outside the boxes was not monitored.

MeHg

In the July MeHg photoexperiments, neither blocking the UVB waveband only nor

blocking all UV radiation had a significant effect on the total amount of MeHg lost as

Hg(0) (Figure 3.3c). All three experiments show losses of approximately 20% after

seven hours of exposure. Overall concentration losses in the September trials were

smaller, at approximately 10% loss in the full spectrum experiment and the UVB blocked

experiment (Figure 3.3d). The smaller losses observed in September may be due to the

lower intensities of light available in September (see Table 3.1). In contrast to the July

Page 80: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

73

trial, the September trial with both UVB and UVA blocked underwent significantly less

photo-reduction (~ 4%) than its full spectrum counterpart (~ 10% loss). Our method of

concentration measurement, however, cannot be relied upon to resolve differences

smaller than 5% and so it is difficult to determine whether loss in the September no UV

trial is indeed significantly lower than losses in the other two trials.

The dark MeHg control experiment carried out in the laboratory showed no detectable

losses after 8.5 hours of sparging with Hg-free air. We therefore conclude that any losses

due to non-photochemical reduction during the 6-7 hour UV filtration experiments were

negligible. Based on literature reports (Bergquist and Blum, 2007; Zhang and Hsu-Kim,

2010) no loss was expected.

Apparent kinetic plateaus

Figure 3.3 displaying reservoir losses of both Hg(II) and MeHg against time suggests

that photo-experiments were allowed to proceed until a kinetic plateau was reached. This

is somewhat misleading in that the experimental hours were chosen to cover the brightest

six hours of daylight available (10 a.m. to 4 p.m.), and while the rates of reaction appear

to slow considerably after 2 to 3 hours of exposure, it is not clear whether the kinetic

plateaus are due to reaction exhaustion or to a decline in the radiation available. It is

likely that photo-reduction would resume in some or all of the reactors if bright daylight

was restored, as was observed in other photoexperiments left outside for 30 subsequent

hours.

3.3.2. Mass-dependent fractionation

When considering the mass-dependent fractionation (MDF) of Hg, it is sufficient to

examine one isotope ratio that does not display mass-independent anomalies. As

recommended in (Blum and Bergquist, 2007), MDF will be reported using 202Hg (see

Methods). If fractionation follows a Rayleigh distillation model, the fractionation factor

() can relate the change in isotope ratios to the amount of substrate that has reacted.

Photo-reduction of both Hg(II) and MeHg in continuously purged reactors appears to

follow a Rayleigh model well, and in the discussion that follows, comparisons will be

Page 81: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

74

made in terms of the kinetic fractionation factor . An = 1.0000 indicates no

fractionation; fractionation becomes more pronounced as the absolute distance of from

1.0000 increases. Throughout this thesis is defined as Rproduct/Rreactant (see Methods)

so, for example, 202 = 0.9980 for a given process means the substrate is becoming

isotopically heavier as the reaction progresses.

Hg(II)

MDF was observed in all experiments where there was photo-reduction with the heavier

isotopes being preferentially retained in the reactor. Complete isotopic data are displayed

in Table 3.2. In the July experiments, the largest change in 202Hg was 1.88 0.17 ‰

(2SE) in the no UVB experiment. MDF was greater in the experiment with UVA

blocked (202 = 0.9988 0.0001, 2SE) than in the full spectrum experiment (202 =

0.9992 0.0002, 2SE; Figure 3.4a and Table 3.3). In the September experiments, the

largest change in 202Hg was 2.02 0.17 ‰ (2SE) in the full sun experiment. September

MDF was not significantly affected by blocking any of the wavebands. All fractionation

factors 202 fall within error of each other (Figure 3.4b and Table 3.3). The July

experiments do not show greater fractionation than the September experiments, despite

exposure to ultraviolet radiation of higher intensity.

MeHg

MDF was observed in all experiments where there was significant photochemical loss,

with the heavier isotopes being preferentially retained in the reactor. Complete isotopic

data are displayed in Table 3.4. In the July experiments, the largest change in 202Hg is

0.33 0.22 ‰ (2SE) in the no UV experiment. July MDF was not significantly affected

by blocking any of the wavebands; 202 for all three experiments fall within error of each

other (Figure 3.4c and Table 3.5). In the September experiments, the largest change in

202Hg is 0.28 0.13 ‰ (2SD) in the no UVB experiments. Differences in mass-

dependent fractionation between the September full sun experiment and the September

no UVB experiment were not significant; fractionation factors (202 values) were not

meaningful due to the small extent of reaction and are not included in Table 3.5 (see also

Page 82: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

75

Figure 3.4d). The September, no UV experiment shows no significant MDF. This is to

be expected if very little photo-reduction occurred, as the concentration data suggests.

The July experiments do not show greater fractionation than the September experiments,

despite exposure to ultraviolet radiation of higher intensity. The dark control is not

shown in Figure 3.4 because, as no significant loss occurred, there could be no

fractionation.

3.3.3 Mass-independent fractionation

Mass-independent fractionation (MIF) occurred during photo-reduction of both Hg(II)

and MeHg in continuously purged reactors and appears to follow a Rayleigh model well.

Comparisons will be made in terms of the kinetic fractionation factor XXX, where XXX

is 199Hg or 201Hg. It is calculated in the same way as mass-dependent fractionation

factors (see Methods) except that a capital delta () value is used in place of a delta ()

value. An XXX = 1.0000 still indicates no fractionation. Throughout this thesis, is

defined as Rproduct/Rreactant (see Methods) so, for example, 199 = 0.9980 for a given

process means the substrate is becoming enriched in the odd isotopes as the reaction

progresses.

Hg (II)

MIF was observed with the odd isotopes, 199Hg and 201Hg, being preferentially retained in

the reactor as observed in (Bergquist and Blum, 2007). MIF occurred for the odd

isotopes only. An example is shown in Figure 3.5, using the July experiments. In

Figure 3.5a, 199Hg for all three experiments is plotted against 202Hg to demonstrate

odd-isotope deviation from mass-dependent behavior. In contrast, when even isotopes

are plotted against 202Hg (for example, 204Hg in Figure 2.5b), the progressive changes

in even follow the values predicted by kinetic mass-dependent fractionation laws (see

Methods).

In July, the largest MIF signature was 2.45 0.14 ‰ (2SE) in the experiment exposed to

full sun. Complete MIF data are displayed in Table 3.2. Blocking the UVB waveband

reduced MIF significantly (Figure 3.6a) with 199 going from 0.9986 0.0001 (full sun)

Page 83: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

76

to 0.9993 0.0001 (no UVB). Blocking all UV radiation suppressed most of the photo-

reduction and isotopic fractionation, so it is not possible to determine what effect this had

on MIF. Rayleigh models based on estimated MIF fractionation factors for July

experiments are shown in Figure 3.6b. Given the insignificant MIF signature and the

small extent of reaction, the July no UV experiment was not plotted on a Rayleigh model.

In September, the largest MIF signature, 0.69 0.14 ‰ was once again in the experiment

exposed to full sun, and a small but significant decrease was observed in fractionation

when the UVB waveband was blocked (Figure 3.6b and Table 3.3). In contrast to July,

photo-reduction in September was not suppressed by blocking all UV radiation, and

fractionation observed in the no UV experiment was not significantly smaller than

fractionation with only the UVB waveband blocked. MIF in the July experiments was

greater than in the September experiments, corresponding with higher intensity of

ultraviolet radiation. No MIF was observed in Dark Control 2 (left for weeks in a sealed

bottle), despite a 40% substrate loss. Nor was MIF observed in the dark control

performed by B. Bergquist (a foil-wrapped reactor with continuous sparging), despite a

20% substrate loss, as reported by (Bergquist and Blum, 2007).

MeHg

Similar to the Hg(II) experiments, MIF was observed with the odd isotopes, 199Hg and 201Hg, being preferentially retained in the reactor. MIF occurred in the odd isotopes only.

An example is shown in Figure 3.7, using the September experiments. In Figure 3.7a,

199Hg for each of the three experiments is plotted against 202Hg to demonstrate mass-

independent anomalies in the odd isotopes. Even isotopes show no such anomalies:

when they are plotted against 202Hg (for example, 200Hg in Figure 3.7b), isotopic

signatures show mass-dependent behavior as predicted by kinetic mass-dependent

fractionation laws (see Methods).

In the July, the largest MIF signature was 0.55 0.07 ‰ (2SD) in the experiment

exposed to full sun. Complete MIF data are displayed in Table 3.4. Although blocking

the UVB and UVA wavebands did not have an appreciable effect on total losses of MeHg

through photo-reduction, MIF was suppressed in both the experiment with only UVB

Page 84: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

77

blocked and the experiment with all UV blocked (Figure 3.8a and Table 3.5).

Unfortunately, a rather large amount of instrumental noise accompanied the production of

this data (2SE is at times half as large or larger than the measurement) and it is not clear

if MIF was suppressed completely or only partially.

In September, the full sun experiment also displayed the largest MIF signature, 0.42

0.14 ‰. Fractionation was significantly reduced by blocking the UVB waveband

(Figure 3.8b and Table 3.5). The September no UVB experiment shows an MIF

signature only half as large. The effect of blocking all UV on MIF in the September

experiments is not clear, since photo-reduction was almost completely suppressed in this

experiment. Uncertainty associated with the fractionation factors in the July full sun and

the September full sun experiments (Table 3.5) makes it difficult to determine if MIF in

July was greater than in September. No change in MeHg concentration was observed in

the dark control, providing no opportunity for MIF. Similarly, no drop in MeHg

concentration was reported for dark controls by (Bergquist and Blum, 2007).

Relationship between 199Hg and 201Hg

The average ratio 199Hg/201Hg was 1.00 0.04 (2SE) for the Hg(II) photo-reduction

experiments. Figure 3.9a shows this ratio for all samples where MIF was significant

(4SE or larger). Although not all Hg(II) experiments lost substrate Hg through photo-

reduction to the same extent, all samples showing significant MIF fall within error of a

line of slope 1.00 ± 0.04, 2SE. In contrast, the average ratio 199Hg/201Hg for MeHg

samples showing significant MIF (4SE or larger) was 1.35 0.16 (2SE), as shown in

Figure 3.9b. All points showing significant MIF but one fall within error of this line.

Page 85: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

78

3.4. DISCUSSION

3.4.1. We see a relationship between the expression and magnitude of MIF and the

energy of light available. Removing the UVB waveband clearly reduces the magnitude

of MIF in Hg(II). In both the July and September experiments, the most pronounced MIF

is in the experiments exposed to full sun. The July experiments were exposed to more

intense ultraviolet radiation and show correspondingly larger fractionation factors. No

MIF was observed in a sealed dark control with 40% substrate loss, or in a dark control

with continuous sparging (carried out by B. Bergquist and reported in (Bergquist and

Blum, 2007) with a substrate loss of 20%.

Removing the UVB waveband also clearly reduces the magnitude of MIF in MeHg.

Fractionation is largest in the experiments exposed to full sun in both the July and the

September experiments. Uncertainty in the fractionation factors, 199 make it difficult to

asses whether exposure to more intense sunlight in July was accompanied by greater

MIF. No MeHg degradation was observed in the dark control, which agrees with other

published results (Bergquist and Blum, 2007; Zhang and Hsu-Kim, 2010). These results

strongly support the magnetic isotope effect (MIE), expressed through spin-selective

reactions such as geminate radical pair recombination, as the likely mechanism of

MIF in photo-reduction. Radical pair generation is a frequency-dependent process and

irradiation by higher frequencies increases the production of radical pairs, allowing

greater expression of the MIE.

3.4.2. For Hg(II), UVB is the larger contributor to MIF but both the UVA and the

visible portions of the spectrum contribute significantly to MIF. In the July

experiments, blocking the UVB spectrum reduces MIF by 50 ± 10%. In the September

experiments, blocking UVB radiation also reduces MIF significantly (by 30 ± 20%).

Blocking all UV radiation reduced MIF (in September) by a similar amount. In the case

of the July experiments, total photo-reduction in the absence of UVB is not significantly

reduced despite the decrease in MIF. This indicates that there are multiple pathways for

photo-reduction, and that not all of them express the MIF. Multiple pathways for

Page 86: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

79

aqueous photo-reduction have been proposed in the literature and include direct

photolysis, indirect photolysis and also direct followed by indirect photolysis.

Direct photolysis of Hg halides, HgXn2-n in organic solvents under laboratory UVC

radiation has been well studied (Kunkely et al., 1997) and yields Hg(I) species and X·

radicals as primary products. There are far fewer studies of photolysis of Hg compounds

in water under environmentally relevant conditions. One such is (Zepp et al., 1973), who

report that photo-degradation of aqueous phenylmercury salts by broadband wavelengths

greater than 290 nm takes place through homolytic cleavage of the carbon-mercury bond:

L HgC6H5hv L Hg C6H5 (1)

if the ligand is another phenyl group, the organomercury radical will decompose to

Hg(0). The singlet-triplet excitation energy of the phenylmercurials was determined to

be approximately 3.47 eV (the energy of radiation at 357 nm), which allowed them to be

photosensitized by acetone (which has comparable singlet-triplet transition energy). This

is an interesting example because it illustrates photo-dissociation of organically-bound

Hg(II) to a radical pair over a range of environmentally available wavelengths. These are

conditions that would allow the expression of the MIE. In addition, it demonstrates that

in-situ photosensitization by a small organic compound can accelerate photo-

decomposition (for energy transfer to be effective unless, the excited state energy of the

sensitizer must be greater than that of the photoreactive compound, (Montalti, 2006; Zepp

et al., 1973). It is very likely that components of organic matter serve as photosensitizers

of Hg complexes in natural waters. More evidence of this comes from the lack of or very

slow photo-reduction of Hg(II) without organic matter.

If the mercury(I) radical (as in equation 1) does not recombine with the original ligand or

spontaneously decompose to Hg(0), indirect photo-reduction may reduce the radical to

Hg(0). (Han et al., 2007; Zheng et al., 2005) demonstrate that photo-generated reactive

intermediates (such as •H and CO•) produced from small organic molecules

(formaldehyde, methanol, acetic, oxalic and malonic acid) can reduce Hg(II) to Hg(0) in

aqueous solution. Alternatively, photo-generated oxidants such as 1O2 and •OH could

convert the radical back to Hg(II).

Page 87: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

80

Two pathways of direct photolysis were proposed by (Lin and Ariya, 2008) to explain the

reduction of Hg(II) in the presence of oxalic, malonic or succinic acid under UV radiation

(290-400 nm). Three small dicarboxylic acids were chosen because carboxylic acid

functional groups are abundant in natural organic matter (NOM) and important sites of

trace metal complexation. One pathway will be discussed subsequently; the other is the

following:

HgCRCOOROOCHg hv222)( (2)

from where •Hg+ can be reduced by some electron donor (such as the photo-generated

reactive intermediates mentioned above) if it is not re-oxidized by (for example)

dissolved O2. Here also, initial photo-excitation of the complex by environmentally

available wavelengths produces a radical pair and provides an opportunity for expression

of the MIE. While carboxylic acids are ubiquitous in NOM, Hg is expected to be

primarily associated with reduced sulfur complexes. Sulfur is only a minor component of

NOM (usually less than 1%, based on the product analyses of the International Humic

Substance Society, n = 23) but its molar concentrations in natural freshwaters exceeds

those of Hg by several orders of magnitude and the stability constants of Hg-thiol

complexes are particularly high (Ravichandran, 2004) and references therein). Thus it is

predicted that most Hg will be bound to reduced S ligands. Based on ab initio

calculations, (Stromberg et al., 1991) expect the aqueous complexes Hg(SH)2 and

HgS(SH)- to be photolysed by sunlight. In addition, the aqueous photo-reduction of Hg-

cysteine complex has been demonstrated in sunlight (this author, unpublished data) and

under UV irradiation ( > 300 nm, (Zheng and Hintelmann, 2010b). Whether Hg in

freshwater is bound to thiolate or to carboxylate groups, therefore, it is plausible that

radical pairs generated by photo-excitation of organomercury complexes are significant

in nature, particularly if the photosensitizing potential of small organic compounds

associated with NOM is considered.

A second pathway of direct photolysis of Hg-organic complexes is two-electron donation

by the organic ligand to Hg(II). This was invoked by (Gardfeldt and Jonsson, 2003)to

Page 88: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

81

explain aqueous photo-reduction ( > 290 nm) of HgC2O4 complexes after reduction by

HO2•/O2•- had been ruled out:

Hg(C2O4 )n (22n ) hv Hg0 2CO2 (n 1)C2O42 (3)

This two-electron transfer is the second pathway proposed by (Lin and Ariya, 2008) to

explain photo-reduction of Hg(II) by small dicarboxylic acids:

[Hg((OOC)2R)n](2-2n) h/H2O HORCO2H + Hg0 + CO2 + (n-1)R(CO2)2

2- (4)

Here is an avenue of photo-reduction that results in the direct production of Hg(0).

Although the lifetime of an excited molecule can range as widely as 10-14 to 100 seconds

(Turro et al. 2008), it is frequently orders of magnitude shorter than the time required for

triplet-singlet conversion (up to about 0.1 s). The MIE would not necessarily have time

to influence the rate of this reaction and MIF would not be expressed. When their entire

experimental reaction system was modeled, Lin and Ariya, 2008 found that the majority

of Hg(II) reduction took place through intramolecular 2-electron transfer (as in equation

4), but that processes starting with photo-generated radical pairs (as in equation 2) also

contributed.

The modeling results of (Lin and Ariya, 2008) suggest that a major pathway of Hg(II)

photo-reduction in the experiments of this study is by 2-electron transfer from an organic

ligand, which provides one avenue for mercury loss that would not allow expression of

the MIE. Another such mechanism is indirect photolysis, that is, two-step reduction by

the reactive intermediates generated through photolysis of organic molecules. (Zheng et

al., 2005), for example, show that Hg(II) could be reduced to Hg(0) by photolysis

products of aqueous formic acid under ambient laboratory light only (although the rate of

Hg(0) production was far higher under UV radiation). This is one example of an indirect

photolysis mechanism through which Hg(II) can be reduced by a small organic molecule

in the absence of ultraviolet radiation. Here is a second avenue for reservoir mercury to

be lost without expressing the MIE.

Radical pairs are produced through absorption of a photon when a photochemically

excited electron orbits at a higher energy than its usual ground state. Such electronic

Page 89: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

82

promotion usually results in two half-filled orbitals, one of which is bonding and the

other non-bonding. If the two moieties no longer share orbitals of net bonding character,

they can drift apart spatially. The energy difference between the highest occupied

molecular orbital and the lowest unoccupied molecular orbital determines the longest

wavelength of light with sufficient energy to cause photoexcitation. If this difference is

small, then wavelengths in the visible region may suffice. As the difference gets larger,

wavelengths in the UVA, UVB, or higher will be required. It is to be expected that direct

photolysis of Hg-organic complexes resulting in radical pairs would be more effectively

stimulated by higher-energy radiation, and that blocking UVB radiation would limit these

reactions and MIF expression considerably.

Relationship between 199Hg and 201Hg for Hg(II)

In examining MIF during photo-reduction of Hg(II), (Bergquist and Blum, 2007) noted

that the ratio199Hg/201Hg was very consistent (1.00 ± 0.02, 2SE) and distinct from the

ratio199Hg/201Hg resulting from photo-degradation of MeHg. It is also distinct from

199Hg/201Hg ratios obtained from transformations of Hg attributed to the nuclear

volume effect (Estrade et al., 2009; Wiederhold et al., 2010). Bergquist and Blum 2007

suggested the ratio might be indicative of the reductive pathway. A ratio of

approximately 1 is observed in lichens and in coals (Bergquist and Blum, 2009; Biswas et

al., 2008; Carignan et al., 2009; Ghosh et al., 2008). Lichens, in particular, are presumed

to derive their Hg from the atmosphere (they are not reported to fractionate during

assimilation) and suggest that atmospheric mercury may also have a ratio 199Hg/201Hg

of 1.0. In contrast, 199Hg/201Hg from freshwater fish tissue samples (which is

approximately all MeHg) has been reported to be 1.3 ± ~0.1 (Bergquist and Blum, 2007;

Jackson et al., 2008; Laffont et al., 2009).

In all Hg(II) reduction experiments in this thesis, the ratio 199Hg/201Hg falls within

error of 1.0, even when photo-reduction is severely reduced. If the ratio is indeed

indicative of the pathway, then the mechanism of MIF is not changing even as the

magnitude is reduced by screening higher energy radiation. Based on the above

Page 90: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

83

mechanistic explanation, it suggests that direct photolysis of mercury-organic compounds

takes place to some extent even in the range of visible radiation (as in the September, no

UVB experiment).

3.4.3. For methyl mercury, UVB is likely the major contributor to MIF during

photo-reduction. A small contribution is made by other wavelengths. When UVB is

blocked, we see at most a small MIF signature in MeHg during the July experiments, and

a small but significant signature in the September experiment exposed to UVA radiation.

Photo-reduction, however, can take place through processes that do not cause MIF. This

is particularly apparent in the July photo-experiments, where MIF was largely or

completely suppressed when the UVB and UVA wavebands were blocked, but the

overall amount of total photo-reduction was not significantly affected. A considerable

amount has been published on the photo-degradation of MeHg, and considering some of

this work provides insight into the relationship between photo-reduction and MIF.

Organomercury compounds have long been known to decompose photolytically (and

thermally) into radical components; evidence for both one-step and two step-mechanisms

R1HgR2 R1Hg R2 R2 (Hg R1) (6)

is reviewed by (Friswell and Gowenlock, 1965). (Janzen and Blackbur.Bj, 1969)note

that photolysed organomercury compounds (R2Hg, RHgX) tend to cleave on the organic

group rather than the halide or carboxylate group. (Inoko, 1981)invokes a mechanism of

this type to account for the aqueous photochemical decomposition of CH3HgCl:

CH3HgCl hv CH3 HgCl

CH3 HgCl CH3Cl Hg0 (7, 8)

where the second line is a reaction known to take place in solvent cages. This pathway of

direct photolysis involves a geminate radical pair whose fate (recombination or

dissociation to free radicals in solution) could be influenced by the MIE. (Malinovsky et

al., 2010)and co-workers have recently published a study examining isotopic

fractionation of aqueous MeHg during UV photolysis (2010). They observe large MIF

signatures (greater than 0.7‰) in acidic and acidic saline solutions. In contrast, MIF is

suppressed by the addition of a radical scavenger (ascorbic acid) and in alkaline

Page 91: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

84

solutions, which inhibit radical formation. These findings are entirely compatible with a

mechanism where MIF is expressed through radical pair chemistry.

Direct photolysis, however, has been studied principally in laboratories using lamps that

produce radiation of much higher frequency than is available at the earth’s surface.

Malinovsky 2010 use a mercury vapor lamp emitting radiation at 254 nm as a light

source (in addition, Hg from the lamp may directly photo-excite Hg in solution), and

Inoko, 1981 presents data obtained with wavelengths of 206 nm. There has been debate

over whether a direct photolysis mechanism is applicable to natural systems. (Takizawa

et al., 1981), for example, find that MeHg was not degraded by a black light lamp

(maximum wavelength 366 nm). Since photo-degradation of MeHg in natural fresh

waters has been clearly demonstrated (Hammerschmidt and Fitzgerald, 2006; Lehnherr

and Louis, 2009; Sellers et al., 1996), other mechanisms have been proposed. (Chen et

al., 2003) investigate the degradation of aqueous MeHg by hydoxyl radicals (•OH)

generated at a steady state by photolysis of nitrate (absorption maximum at 302 nm) at

naturally relevant wavelengths. Degradation of MeHg followed pseudo-first-order

kinetics and among the mechanistic paths proposed were both stepwise and simultaneous

decomposition:

CH3HgCl OH CH3OH HgCl CH3OH Hg0 Cl (9)

In addition, methyl radicals (generated by natural organic matter, NOM) could react with

HgCl radicals to produce elemental Hg:

CH3 HgCl Hg0 CH3Cl (10)

Here is an encounter between a radical pair. Whether such an encounter could allow the

expression of the MIE requires an understanding of whether spin coupling between a

magnetic nucleus and its electrons affects random radical pair interactions, as well as

geminate radical pair reactions. The environmental significance of this reaction depends

on the steady state concentration of •OH in environmental systems and also on competing

photo-degradation reactions.

Field reports of MeHg photo-degradation in fresh water have been from water bodies

with dissolved natural organic matter (NOM), which Chen et al. 2003 do not add to their

Page 92: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

85

experiments. (Zhang and Hsu-Kim, 2010) report that aqueous MeHg is readily degraded

by both sunlight and a UVA lamp, but only in the presence of NOM and only when there

was a molar excess of reduced sulfur relative to MeHg. Theoretical studies offer some

explanation of the increased photo-reactivity of MeHg with thiol complexes in particular.

The highest-energy radiation that reaches the surface of the earth can induce an electronic

transition of approximately 4.4 electron volts (eV; this is ~ 282 nm). Through ab initio

calculations,(Tossell, 1998) estimates singlet to triplet-state excitation energies (ES-T)

for various for various MeHg-ligand complexes (see Table 3.7).

Table 3.7. Calculated properties of CH3HgL complexes, from Tossell 1998

ESinglet-Triplet in eV ESinglet-Triplet in nm

CH3HL' Hartree-Frock

Moller-Plesset

Hartree-Frock

Moller-Plesset

changes upon S-T excitation

CH3HgOH2+ 3.97 4.97 312 249 CH3 leaves

CH3HgOH ~5.9 210 CH3 and OH leave CH3HgCl 5.41 6.04 229 205 CH3 leaves CH3HgSH 5.16 4.98 240 249 CH3 leaves CH3Hg(SH)2 5.47 4.26 227 291 one SH leaves

Most of these ES-T transitions are greater than the high-energy boundary for solar

radiation. Singlet-singlet transitions energies are always approximately 2-3 eV (400-600

nm) higher than S-T energies and lie well into the UVC range. Electronic excitation

through direct photon absorption, therefore, is most likely to take place through the

lowest energy triplet state. Depending on pH, the MeHg-thiol complexes appear to have

the lowest ES-T excitation energies, and are therefore the most likely to have absorption

tails that extend to the longer wavelengths reaching earth’s surface. Table 2.7 above

shows there is a strong tendency toward dissociation in the lowest energy triplet state of

the complex, with methyl radicals being generated. This tendency is because, while the

highest occupied molecular orbitals (HOMO) are bonding, the lowest unoccupied

molecular orbitals (LUMO) are anti-bonding. In a triplet state, the antibonding orbital

becomes half-occupied and the bonding orbital occupancy decreases to half. The Hg-C

bond strength is greatly reduced and •CH3 moiety dissociates during geometry

optimization.

Page 93: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

86

The larger MIF signatures observed under exposure to full sun (and hence to higher

energy UVB and UVA) in both July and September could be due particularly to

dissociation of •CH3 and •Hg-thiol radical intermediates following singlet-triplet

excitation of a MeHg-thiol complexes. This would allow an opportunity for the MIE to

be expressed. Radical pairs generated from odd-isotope complexes would have a higher

probability of recombining to form the starting products before complete dissociation. If

the radical pair does dissociate, the fate of radical Hg(I) species (both odd and even) may

be re-oxidation (Zhang and Hsu-Kim 2010 present strong evidence for 1O2; •OH is also a

possibility) or reduction (by •H or CO• generated by irradiation of low molecular weight

organic compounds as in Zheng et al., 2005 and Han et al 2007). At lower wavelengths,

mechanisms such as the one involving •OH proposed by Chen et al. 2003 could induce

photo-reduction through pathways that do not involve radical pairs, reducing the

possibility for expression of the MIE.

Relationship between 199Hg and 201Hg for MeHg

Bergquist and Blum 2007 noted that the ratio199Hg/201Hg obtained during the photo-

reduction of MeHg was very consistent (1.36 ± 0.02, 2SE) and distinct from the

199Hg/201Hg ratio associated with photo-degradation of Hg(II). It is also distinct from

the ratio of 199Hg/201Hg ratios obtained from transformations of Hg attributed to the

nuclear volume effect (2.0 ± 0.6, Estrade et al., 2009, and 1.62 ± 0.06 (2SE), Ghosh et

al., in preparation and 1.54 ± 0.22 (2SE), Weiderhold et al., 2010). The average ratio

199Hg/201Hg for all MeHg experiments in this thesis was 1.35 ± 0.16 (2SE). The

consistency of this ratio among experiments of the same Hg species suggest that the

mechanism(s) of photo-reduction is/are the same for a given species regardless of

changes in the radiation available. The significant difference of this ratio between MeHg

and Hg(II) suggest that the pathway(s) of photo-reduction are not the same for the

organic Hg species as for the inorganic one.

Malinkovsky et al. 2010, report a ratio 199Hg/201Hg for experimental photo-reduction

of MeHg under UVC radiation (1.28 ± 0.03, 1SD) that is similar to that of Bergquist and

Blum 2007, which was obtained experimentally under natural sunlight. It is also within

Page 94: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

87

error of the ratio obtained in this study under natural sun. This supports the possibility

that the mechanism leading to MIF under exposure to radiation of wavelength 254 nm

(leading to radical pairs through direct photolysis) is similar to the mechanism leading to

MIF under exposure to natural sunlight. The 199Hg/201Hg ratio obtained from samples

of temperate lake fish tissue (1.28 ± 0.03, 2SE , Bergquist and Blum 2007, 2009),

tropical lake fish tissue (1.28 ± 0.12, 2SE , Laffont et al. 2009) and arctic lake fish tissue

(1.32 ± 0.06, 2SE for 9 out of 10 lakes, (Gantner et al., 2009) support the possibility that

MIF signatures carried by MeHg in organic tissue was imparted by the same mechanism,

prior to entry into the food web.

3.4.4. Mass-Dependent Fractionation

It is interesting that MDF in the July Hg(II) experiments increases significantly as higher-

energy wavebands are blocked. This effect is not observed in September, however,

where the mass-dependent fractionation factors of the three experiments are

indistinguishable. As noted in Table 3.1, visible radiation in July at mid-day was more

intense under the UV-filtering box than it was under full solar exposure. It is likely that

the July no UVB and the no UV experiments underwent more evaporation than the July

full spectrum experiment. In this case, 202 will reflect a combination of reduction and

evaporation effects, and should not be over-interpreted. In September, the MeHg

experiment showing the largest MDF signature is the one shielded from UVB.

Generally, the effects of UV filtration on MDF during photo-reduction of Hg species are

not amenable to interpretation.

3.4.5. Implications for the natural world

In nature, photo-reductive processes operate in competition with photo-oxidative

processes, and the rates of photo-reduction measured during daylight in natural systems

usually reflect the net rather than the total photo-reduction (Garcia et al., 2005). In our

experiments, continuous sparging with Hg-free air removed most (if not all) of the Hg(0)

from the reaction vessel as it formed, likely eliminating the photo-oxidative reaction

pathway. For this reason, and also the fact that experimental concentrations of Hg were

Page 95: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

88

several orders of magnitude higher than concentrations of Hg in natural fresh waters, the

rates of photo-reduction that we observe are not representative of natural rates.

Mass-dependent fractionation

The range of Hg(II) mass-dependent fractionation factors in this thesis, using 202Hg as the

example isotope, is from 0.9992 ± 0.0002 (July full sun) to 0.9983 ± 0.0003 (September

no UVB). Zheng and Hintelmann report on photoexperiments similar to the ones in this

thesis investigating the effects of the ratio of Hg to DOM as the variable of interest

(2009) and the effects of sulfur-less versus sulfur-containing ligands (2010b).

Fractionation factors range from 0.9986 ± 0.0001 to 0.9996 ± 0.0007 with the least

significant fractionation at the highest Hg/DOM ratio and most significant associated

with an intermediate ratio. Dark controls reached a maximum of 0.9981 ± 0.0004. Thus

intensity, frequency, and Hg/DOM ratio all affect MDF over the same range and one

would need to have additional information about the ecosystem of interest to interpret

MDF values. Additionaly, the variability of 202 with available radiation and with kind

and proportion of organic matter falls into the same range as dark abiotic reduction

fractionation factors and does not appear useful in distinguishing photochemical from

dark abiotic reduction.

MDF does not appear to be helpful in distinguishing photochemical from biological

reductive processes either. Kritee et al., 2007; 2008 investigated fractionation of Hg(II)

during dark reduction by microorganisms, using both aerobic and facultative anearobes

that reduce Hg by the mer pathway and one facultative anaerobe the employs an

undetermined, non-mer mechanism. They report fractionation factors ranging from

0.9982 ± 0.0003 to 0.9988 ± 0.0001. Kritee et al., 2009 report an MDF fractionation

factor of 0.0096 ± 0.0002 for the mer-mediated degradation of MeHg. The narrowness of

this range is interesting in view of the bacterial diversity, but falls squarely into the range

of MDF fractionation factors generated by abiotic photochemical and dark reduction.

Mass-independent fractionation: Quantifying Hg(II) photo-reduction

Page 96: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

89

Large mass-independent fractionation of Hg(II) (greater than approximately 0.2 ‰)

remains associated with photochemical reduction. No MIF was reported by Bergquist

and Blum 2007 during reduction in dark abiotic (organic matter) controls or by Kritee et

al. 2008 for any of the strains of Hg-reducing bacteria. Zheng and Hintelmann 2010a

report 199Hg as large as -0.6‰ after nearly complete dark chemical reduction of Hg(II)

by SnCl2 with continuous removal of the product Hg(0), although it is not clear whether

this represents any natural process. Relating MIF to photo-reduction quantitatively,

however, will be difficult because, as this thesis has demonstrated, comparable amounts

of photo-reduction may coincide with different amounts of MIF depending on the

frequency of light available. In addition, it has been argued that there are multiple

simultaneous pathways for photo-reduction of Hg(II) in natural fresh waters, and that not

all of them permit the expression of the MIE.

Other factors influencing MIF during aqueous photo-reduction of Hg(II)

Of interest is the dramatic effect Hg/DOM ratios have on the magnitude of MIF. Zheng

and Hintelmann (2009) investigated this variable during aqueous photo-reduction and

reported maximum MIF in a trial with 10 gL-1 Hg/12 mgL-1 DOM exposed to sunlight

(199 = 0.9943 ± 0.0008). This max MIF occurred at an intermediate value of Hg/DOM

that they investigated (they varied Hg/DOM ratios from 34.6 to 8330). The observed

MIF at 10 gL-1 Hg/12 mgL-1 DOM is a much larger effect than is observed by blocking

portions of the electromagnetic spectrum and was obtained under conditions of poorer

light penetration of the reactor. As the Hg/DOM ratio continues to diminish, however,

the MIF effects become less pronounced and the lowest ratio experiment (2 ppb Hg in 60

mg DOM/L) has an 199 = 0.9973 ± 0.0006 which is not that different from the July, full

spectrum experiment of this thesis with an 199 of 0.9986 ± 0.0001. However, even the

smaller fractionation seen at this low Hg/DOM is larger than any of the fractionation

factors observed in this thesis. It should be noted that important variables such as the

kind of organic matter, the light intensity (also the light source) were not the same for this

thesis as for the work of Zheng and Hintelmann, 2009. Zheng and Hintelmann 2009

investigate Hg/DOM ratios over three orders of magnitude. It is not clear if variability

demonstrated in their experiments can be expected to be relevant to MIF signatures in

Page 97: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

90

nature, where aqueous concentrations of reduced sulfur (to say nothing of organic matter

as a whole) exceed aqueous Hg concentrations by orders of magnitude and Hg/DOM

ratios are still considerable smaller than what they investigated.

Also of interest are photoexperiments conducted by Zheng and Hintelmann, 2010b on the

effects of a sulfur-containing ligand cysteine and its sulfur-less counterpart serine. Using

cysteine, they acheive a molar concentration of reduced sulfur three orders of magnitude

larger than that of Hg (natural ratios are 3 to 5 orders larger), which may create a more

realistic model of Hg-NOM bonds in freshwater environments. Large, consistently

negative MIF was observed in the reactant reservoir indicating that the odd isotopes were

preferentially reduced and purged from the reactant reservoir. This effect is opposite the

MIF associated with photo-reduction in the presence of organic matter reported in

Bergquist and Blum, 2007, Zheng and Hintelmann, 2009 and in this study. The

explanation was put forth as follows by the authors: the original radical pair produced

through photo-exitation of a complex could start in an excited singlet state. Photolysis

products from an excited singlet state would be inhibited by recombination to the starting

material, but hyperfine coupling between a magnetic (odd) nucleus and an electron could

invert or rephase the spin of the electron such that the pair becomes triplet.

Recombination would no longer be possible and the triplet pair would be likely to

dissociate spatially into independent radicals, resulting in a preferential depletion of the

odd isotopes in the starting material. Whether photo-excitation produces a singlet state or

a triplet state radical pair depends on the nature of the chemical bond broken and the

reaction conditions, such as temperature and exciting wavelength. This is an interesting

complication and raises the possibility that MIF signatures in nature (as Zheng and

Hintelmann 2010b argue) will be the net result of competing photo-processes.

Mass-independent fractionation: Quantifying MeHg photo-reduction

Mass-independent fractionation is associated with photochemical reduction of MeHg, but

it is not associated with dark biological reduction (Kritee et al., 2009). It is not clear

under what conditions MeHg undergoes dark abiotic reduction; none was observed in the

dark controls of this study. Relating MIF to photo-reduction quantitatively, however,

Page 98: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

91

will be difficult because, as this thesis has demonstrated, comparable amounts of photo-

reduction may coincide with different amounts of MIF depending on the

wavelengths of light available for photo-reduction. In addition, it has been argued that

there are multiple simultaneous pathways for photo-reduction of MeHg in natural fresh

waters, and that not all of them permit the expression of the MIE.

Other factors influencing MIF during aqueous photo-reduction of MeHg

There is some evidence that Hg/DOM ratios may have an important effect on photo-

reduction of MeHg as well. Bergquist and Blum 2007 undertook photoexperiments

similar to the ones in this thesis, with aqueous MeHg in 2 and in 20 mg/L NOM

(Suwannee River, of the same origin as the NOM in this study). Their data show an

199 = 0.9967 ± 0.0006 for the 2 mg NOM/L experiment and an 199 = 0.9921 ± 0.0013

for the 20 mg NOM/L experiment with fR = 0.80 respectively. The magnitude of the

fractionation factor 199 obtained under full sun with 20 mg/L NOM is surprisingly large

compared to the changes in 202 effected by varying UV exposure in this experiment.

The increase in fractionation appears to be associated with decreasing the ratio of Hg to

DOM in solution. The mechanistic discussion presented earlier of MeHg-thiol

complexes enabling photo-generation of radical pairs at naturally available wavelengths

is compatible with these results. However Bergquist and Blum also conducted these

experiments on different days, and variations in solar radiation intensity and frequency

cannot be ruled out as a driver of the differences observed for the two experiments.

Page 99: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

92

3.5. CONCLUSIONS

3.5.1. Blocking the UVB spectrum significantly reduces MIF during photochemical

reduction of both MeHg and Hg(II). This is evidence of a mechanism that is stimulated

by photo-excitation and has a certain energy threshold. The magnetic isotope effect,

expressed through photo-excitation of a solvated complex to a radical pair, is such a

mechanism and our experiments solidly support it as the cause of large MIF signatures

during photochemical reduction of Hg species.

3.5.2. UVA radiation can contribute to MIF during photochemical reduction of both

MeHg and Hg(II). Electronic excitation energies are influenced by many factors such as

the atoms in the complex, the chemical composition of the solution and the ambient

temperature. A mechanism where the primary step is electronic excitation of a solvated

complex is compatible with a range of excitation energies. Visible light can contribute

to MIF during photochemical reduction of Hg(II) as well.

3.5.3. There are multiple pathways for photo-reduction of both MeHg and Hg(II) in

freshwater reservoirs. Some of these are spin-selective (since they express the MIE) and

some are not. Pathways that express the MIE have spin-selective processes at or before

their rate-limiting steps. This knowledge can be used to elucidate the photochemistry of

Hg in natural waters.

3.5.4. Quantitatively relating total photo-reduction in a freshwater system and the

magnitude of MIF will not be straightforward for either Hg(II) or MeHg. For both

species, we have demonstrated cases where comparable amounts of total photo-reduction

are accompanied by significantly different MIF.

Page 100: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

93

FIGURES

Figure 3.1. Taken from Bergquist 2007. Schematic of the photochemical

reaction apparatus with Hg(II) as an example species.

Page 101: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

94

Figure 3.2. Courtesy of B. Bergquist. Phase separator. Hg(II) in solution is reduced online with

SnCl2 to produce Hg(0), a vapor. The solution carrying dissolved Hg vapor runs in a thin film

down a frosted tip. The Hg0 is stripped out by a clean gas flowing upwards and out an arm at

the top of the phase separator case. From there, the Hg can be either trapped in an acidic

solution or directed into an analytical machine.

Page 102: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

95

Figure 3.3. Mercury concentration in the substrate reservoir over time for replicate photo-

reduction experiments on two Hg species, inorganic Hg(II) and organic MeHg, in the presence of

natural organic matter and under varying exposure to solar radiation. Error in time elapsed is 1

minute, which is smaller than the width of the symbols used. Maxima in solar irradiation were

recorded at approximately 1 pm, Eastern Standard Time, which was found to coincide with solar

noon in Toronto, Ontario. Effective blocking of the UVB spectrum necessitated partial blocking

of the UVA spectrum (by 30 to 52%). Complete radiation measurements are shown in Table

3.1. a) Dark Control 1 was carried out in a foil-wrapped reactor with continuous sparging.

Dark Control 2 was carried out in a sealed, foil-wrapped bottle with no sparging.

Page 103: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

96

Figure 3.4. Mass-dependent isotopic fractionation (202Hg) of the substrate

reservoir as a function of Hg remaining for replicate photo-reduction experiments

on two Hg species, inorganic Hg(II) and organic MeHg with dissolved natural

organic matter under varying exposure to solar radiation. Error in fraction of

substrate Hg remaining is set at 0.05 or less (see Reporting analytical

uncertainty, p. 68), but not shown on plots (c) and (d) for clarity. Rayleigh

distillation models based on estimated fractionation factors (202) are shown

where there was significant loss in concentration. Solid black lines: full sun;

broken grey lines: UVB spectrum blocked; broken black lines: UVB and UVA

spectra blocked. Complete isotope data are displayed in Table 3.2 and Table 3.4.

Mass-dependent fractionation factors for all isotopes and associated uncertainty

are displayed in Table 3.3.

Page 104: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

97

Figure 3.5. Mass-independent isotopic fractionation (MIF, 199Hg), observed in

odd isotopes only, for Hg(II) photo-reduction experiments with dissolved natural

organic matter under varying exposure to solar radiation. (a) Deviation from

theoretical mass-dependent 199Hg values (black line) based on measured 202Hg

and the kinetic mass-dependent fractionation law derived from transition state

theory (Young et al. 2002). (b) Adherence to theoretical mass-dependent 204Hg

values (black line). Complete isotope data are displayed in Table 3.2

Page 105: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

98

Figure 3.6. Mass-independent isotopic fractionation (MIF, 199Hg), observed in

odd isotopes only, as a function of inorganic Hg(II) remaining for replicate photo-

reduction experiments with dissolved natural organic matter under varying

exposure to solar radiation. No MIF was observed in dark controls. Rayleigh

distillation models based on estimated mass-independent fractionation factors are

shown where MIF was significant. Solid black lines: full sun; broken grey

lines: UVB spectrum blocked; broken black lines: UVB and UVA spectra

blocked. Complete isotope data are displayed in Table 3.2 Mass-independent

fractionation factors for both odd isotopes and associated uncertainty are

displayed in Table 3.3.

Page 106: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

99

Figure 3.7. Mass-independent isotopic fractionation (MIF, 199Hg), observed in

odd isotopes only, for MeHg photo-reduction experiments with dissolved natural

organic matter under varying exposure to solar radiation. (a) Deviation from

theoretical mass-dependent 199Hg values (black line) based on measured 202Hg

and the kinetic mass-dependent fractionation law derived from transition state

theory (Young et al. 2002). (b) Adherence to theoretical mass-dependent 200Hg

values (black line). Complete isotope data are displayed in Table 3.4. Mass-

independent fractionation factors for both odd isotopes and associated uncertainty

are displayed in Table 3.5.

Page 107: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

100

Figure 3.8. Mass-independent isotopic fractionation (MIF, 199Hg), observed in

odd isotopes only, as a function of organic MeHg remaining for replicate photo-

reduction experiments with dissolved natural organic matter under varying

exposure to solar radiation. MeHg was not significantly reduced in dark controls.

A Rayleigh distillation model based on an estimated mass-independent

fractionation factor is shown where significant MIF took place: Solid black line:

full sun. Complete isotope data are displayed in Table 3.4. Mass-independent

fractionation factors for both odd isotopes and associated uncertainty are

displayed in Table 3.5.

Page 108: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

101

Figure 3.9. 201Hg versus 199Hg for photochemical reduction of inorganic

Hg(II) and organic MeHg in the presence of natural organic matter under varying

exposure to solar radiation. Dark controls all plot at the origin. Only samples

with significant MIF were plotted, so R0 (pre-exposure) and R1 (taken

immediately after first exposure to solar radiation) do not appear for any

experiment. (a) For Hg(II), the average ratio 201Hg/199Hg is 1.00 0.04, 2SE,

n = 21 (black line). (b) for MeHg, the average ratio 201Hg/199Hg is 1.35 0.16,

2SE, n = 11 (black line).

Page 109: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

102

TABLES

Page 110: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

103

Page 111: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

104

Page 112: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

105

Page 113: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

106

Page 114: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

107

Page 115: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

108

Page 116: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

109

Page 117: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

110

LITERATURE CITED Bergquist, B. A. and Blum, J. D., 2007. Mass-dependent and -independent fractionation

of Hg isotopes by photoreduction in aquatic systems. Science 318, 417-420.

Bergquist, R. A. and Blum, J. D., 2009. The Odds and Evens of Mercury Isotopes: Applications of Mass-Dependent and Mass-Independent Isotope Fractionation. Elements 5, 353-357.

Biswas, A., Blum, J. D., Bergquist, B. A., Keeler, G. J., and Xie, Z. Q., 2008. Natural Mercury Isotope Variation in Coal Deposits and Organic Soils. Environmental Science & Technology 42, 8303-8309.

Blum, J. D. and Bergquist, B. A., 2007. Reporting of variations in the natural isotopic composition of mercury. Analytical and Bioanalytical Chemistry 388, 353-359.

Buchachenko, A. L., 2001. Magnetic isotope effect: Nuclear spin control of chemical reactions. Journal of Physical Chemistry A 105, 9995-10011.

Carignan, J., Estrade, N., Sonke, J. E., and Donard, O. F. X., 2009. Odd Isotope Deficits in Atmospheric Hg Measured in Lichens. Environmental Science & Technology 43, 5660-5664.

Chen, J., Pehkonen, S. O., and Lin, C. J., 2003. Degradation of monomethylmercury chloride by hydroxyl radicals in simulated natural waters. Water Research 37, 2496-2504.

Estrade, N., Carignan, J., Sonke, J. E., and Donard, O. F. X., 2009. Mercury isotope fractionation during liquid-vapor evaporation experiments. Geochimica Et Cosmochimica Acta 73, 2693-2711.

Feng, X. B., Foucher, D., Hintelmann, H., Yan, H. Y., He, T. R., and Qiu, G. L., 2010. Tracing Mercury Contamination Sources in Sediments Using Mercury Isotope Compositions. Environmental Science & Technology 44, 3363-3368.

Fitzgerald, W. F. and Lamborg, C. H., 2007. Geochemistry of Mercury in the Environment. In: Holland, H. D. and Turekian, K. K. Eds.)Treatise on Geochemistry. Elsevier.

Foucher, D., Ogrinc, N., and Hintelmann, H., 2009. Tracing Mercury Contamination from the Idrija Mining Region (Slovenia) to the Gulf of Trieste Using Hg Isotope Ratio Measurements. Environmental Science & Technology 43, 33-39.

Page 118: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

111

Friswell and Gowenlock, 1965. Advances in Free Radical Chemistry 1, 39.

Fritz, P. and Fontes, J. C., 1980. Handbook of Environmental Isotope Geochemistry. Elsevier, New York.

Gantner, N., Hintelmann, H., Zheng, W., and Muir, D. C., 2009. Variations in Stable Isotope Fractionation of Hg in Food Webs of Arctic Lakes. Environmental Science & Technology 43, 9148-9154.

Garcia, E., Poulain, A. J., Amyot, M., and Ariya, P. A., 2005. Diel variations in photoinduced oxidation of Hg-0 in freshwater. Chemosphere 59, 977-981.

Gardfeldt, K. and Jonsson, M., 2003. Is bimolecular reduction of Hg(II) complexes possible in aqueous systems of environmental importance. Journal of Physical Chemistry A 107, 4478-4482.

Ghosh, S., In preparation.

Ghosh, S., Xu, Y. F., Humayun, M., and Odom, L., 2008. Mass-independent fractionation of mercury isotopes in the environment. Geochemistry Geophysics Geosystems 9.

Hammerschmidt, C. R. and Fitzgerald, W. F., 2006. Photodecomposition of methylmercury in an arctic Alaskan lake. Environmental Science & Technology 40, 1212-1216.

Han, C. F., Zheng, C. B., Wang, J., Cheng, G. L., Lv, Y., and Hou, X. D., 2007. Photo-induced cold vapor generation with low molecular weight alcohol, aldehyde, or carboxylic acid for atomic fluorescence spectrometric determination of mercury. Analytical and Bioanalytical Chemistry 388, 825-830.

Inoko, M., 1981. STUDIES ON THE PHOTOCHEMICAL DECOMPOSITION OF ORGANOMERCURIALS-METHYLMERCURY(II) CHLORIDE. Environmental Pollution Series B-Chemical and Physical 2, 3-10.

Jackson, T. A., Whittle, D. M., Evans, M. S., and Muir, D. C. G., 2008. Evidence for mass-independent and mass-dependent fractionation of the stable isotopes of mercury by natural processes in aquatic ecosystems. Applied Geochemistry 23, 547-571.

Janzen, E. G. and Blackbur.Bj, 1969. DETECTION AND IDENTIFICATION OF SHORT-LIVED FREE RADICALS BY ELECTRON SPIN RESONANCE TRAPPING TECHNIQUES (SPIN TRAPPING) - PHOTOLYSIS OF

Page 119: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

112

ORGANOLEAD, -TIN, AND -MERCURY COMPOUNDS. Journal of the American Chemical Society 91, 4481-&.

Kritee, K., Barkay, T., and Blum, J. D., 2009. Mass dependent stable isotope fractionation of mercury during mer mediated microbial degradation of monomethylmercury. Geochimica Et Cosmochimica Acta 73, 1285-1296.

Kritee, K., Blum, J. D., and Barkay, T., 2008. Mercury Stable Isotope Fractionation during Reduction of Hg(II) by Different Microbial Pathways. Environmental Science & Technology 42, 9171-9177.

Kritee, K., Blum, J. D., Johnson, M. W., Bergquist, B. A., and Barkay, T., 2007. Mercury stable isotope fractionation during reduction of Hg(II) to Hg(0) by mercury resistant microorganisms. Environmental Science & Technology 41, 1889-1895.

Kunkely, H., Horvath, O., and Vogler, A., 1997. Photophysics and photochemistry of mercury complexes. Coordination Chemistry Reviews 159, 85-93.

Laffont, L., Sonke, J. E., Maurice, L., Hintelmann, H., Pouilly, M., Bacarreza, Y. S., Perez, T., and Behra, P., 2009. Anomalous Mercury Isotopic Compositions of Fish and Human Hair in the Bolivian Amazon. Environmental Science & Technology 43, 8985-8990.

Lauretta, D. S., Klaue, B., Blum, J. D., and Buseck, P. R., 2001. Mercury abundances and isotopic compositions in the Murchison (CM) and Allende (CV) carbonaceous chondrites. Geochimica Et Cosmochimica Acta 65, 2807-2818.

Lehnherr, I. and Louis, V. L. S., 2009. Importance of Ultraviolet Radiation in the Photodemethylation of Methylmercury in Freshwater Ecosystems. Environmental Science & Technology 43, 5692-5698.

Lin, S. and Ariya, P., 2008. Reduction of oxidized mercury species by dicarboxylic acids (C-2-C-4): Kinetic and product studies. Environmental Science & Technology 42, 5150-5155.

Lindberg, S., Bullock, R., Ebinghaus, R., Engstrom, D., Feng, X. B., Fitzgerald, W., Pirrone, N., Prestbo, E., and Seigneur, C., 2007. A synthesis of progress and uncertainties in attributing the sources of mercury in deposition. Ambio 36, 19-32.

Lindqvist, O., Johansson, K., Aastrup, M., Andersson, A., Bringmark, L., Hovsenius, G., Hakanson, L., Iverfeldt, A., Meili, M., and Timm, B., 1991. MERCURY IN THE SWEDISH ENVIRONMENT - RECENT RESEARCH ON CAUSES,

Page 120: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

113

CONSEQUENCES AND CORRECTIVE METHODS. Water Air and Soil Pollution 55, R11-&.

Malinovsky, D., Latruwe, K., Moens, L., and Vanhaecke, F., 2010. Experimental study of mass-independence of Hg isotope fractionation during photodecomposition of dissolved methylmercury. Journal of Analytical Atomic Spectrometry 25, 950-956.

Mariotti, A., Germon, J. C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, A., and Tardieux, P., 1981. EXPERIMENTAL-DETERMINATION OF NITROGEN KINETIC ISOTOPE FRACTIONATION - SOME PRINCIPLES - ILLUSTRATION FOR THE DENITRIFICATION AND NITRIFICATION PROCESSES. Plant and Soil 62, 413-430.

Mergler, D., Anderson, H. A., Chan, L. H. M., Mahaffey, K. R., Murray, M., Sakamoto, M., and Stern, A. H., 2007. Methylmercury exposure and health effects in humans: A worldwide concern. Ambio 36, 3-11.

Mitrofanov, A. V., Karban, O. V., Sugonyako, A., and Lubomska, M., 2009. Investigation of the Surface of Poly(ethylene terephthalate) Films Modified by Vacuum Ultraviolet Irradiation in Air. Journal of Surface Investigation-X-Ray Synchrotron and Neutron Techniques 3, 519-526.

Mizuno, K., Gan, B. K., Kondyurin, A., Bilek, M. M. M., and McKenzie, D. R., 2008. Reducing Water Permeability while Maintaining Transparency of PET: A Plasma Immersion Ion Implantation Study. Plasma Processes and Polymers 5, 834-839.

Montalti, M. e. a., 2006. Handbook of Photochemistry. CRC/Taylor & Francis, Boca Raton, Fla.

Ravichandran, M., 2004. Interactions between mercury and dissolved organic matter - a review. Chemosphere 55, 319-331.

Selin, N. E., 2009. Global Biogeochemical Cycling of Mercury: A Review. Annual Review of Environment and Resources 34, 43-63.

Sellers, P., Kelly, C. A., Rudd, J. W. M., and MacHutchon, A. R., 1996. Photodegradation of methylmercury in lakes. Nature 380, 694-697.

Sherman, L. S., Blum, J. D., Johnson, K. P., Keeler, G. J., Barres, J. A., and Douglas, T. A., 2010. Mass-independent fractionation of mercury isotopes in Arctic snow driven by sunlight. Nature Geoscience 3, 173-177.

Page 121: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

114

Sherman, L. S., Blum, J. D., Nordstrom, D. K., McCleskey, R. B., Barkay, T., and Vetriani, C., 2009. Mercury isotopic composition of hydrothermal systems in the Yellowstone Plateau volcanic field and Guaymas Basin sea-floor rift. Earth and Planetary Science Letters 279, 86-96.

Smith, C. N., Kesler, S. E., Klaue, B., and Blum, J. D., 2005. Mercury isotope fractionation in fossil hydrothermal systems. Geology 33, 825-828.

Stromberg, D., Stromberg, A., and Wahlgren, U., 1991. RELATIVISTIC QUANTUM CALCULATIONS ON SOME MERCURY SULFIDE MOLECULES. Water Air and Soil Pollution 56, 681-695.

Takizawa, Y., Minagawa, K., and Hisamatsu, S., 1981. Studies on mercruty behaviour in man's environment: (report V) photodegradation of methylmercury in the atmosphere by ultraviolet rays with sterilization. Japanese Journal of Public Health, 313-320.

Tossell, J. A., 1998. Theoretical study of the photodecomposition of methyl Hg complexes. Journal of Physical Chemistry A 102, 3587-3591.

Turro, N., Scaiano, J. C., and Ramamurthy, V., 2008. Principles of Molecular Photochemistry: An Introduction. University Science Books.

Wiederhold, J. G., Cramer, C. J., Daniel, K., Infante, I., Bourdon, B., and Kretzschmar, R., 2010. Equilibrium Mercury Isotope Fractionation between Dissolved Hg(II) Species and Thiol-Bound Hg. Environmental Science & Technology 44, 4191-4197.

Yang, L. and Sturgeon, R., 2009. Isotopic fractionation of mercury induced by reduction and ethylation. Analytical and Bioanalytical Chemistry 393, 377-385.

Young, E. D., Galy, A., and Nagahara, H., 2002. Kinetic and equilibrium mass-dependent isotope fractionation laws in nature and their geochemical and cosmochemical significance. Geochimica Et Cosmochimica Acta 66, 1095-1104.

Zepp, R. G., Wolfe, L., and Gordon, J. A., 1973. Photodecomposition of Phenylmercury Compounds in Sunlight. Chemosphere, 93-99.

Zhang, T. and Hsu-Kim, H., 2010. Photolytic degradation of methylmercury enhanced by binding to natural organic ligands. Nature Geoscience 3, 473-476.

Zheng, C. B., Li, Y., He, Y. H., Ma, Q., and Hou, X. D., 2005. Photo-induced chemical vapor generation with formic acid for ultrasensitive atomic fluorescence

Page 122: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

115

spectrometric determination of mercury: potential application to mercury speciation in water. Journal of Analytical Atomic Spectrometry 20, 746-750.

Zheng, W., Foucher, D., and Hintelmann, H., 2007. Mercury isotope fractionation during volatilization of Hg(0) from solution into the gas phase. Journal of Analytical Atomic Spectrometry 22, 1097-1104.

Zheng, W. and Hintelmann, H., 2009. Mercury isotope fractionation during photoreduction in natural water is controlled by its Hg/DOC ratio. Geochimica Et Cosmochimica Acta 73, 6704-6715.

Zheng, W. and Hintelmann, H., 2010a. Nuclear Field Shift Effect in Isotope Fractionation of Mercury during Abiotic Reduction in the Absence of Light. Journal of Physical Chemistry A 114, 4238-4245.

Zheng, W. and Hintelmann, H., 2010b. Isotope Fractionation of Mercury during Its Photochemical Reduction by Low-Molecular-Weight Organic Compounds. Journal of Physical Chemistry A 114, 4246-4253.

Page 123: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

116

CHAPTER 4: MASS-INDEPENDENT

FRACTIONATION DURING CHEMICAL

REDUCTION OF AQUEOUS INORGANIC

MERCURY IN THE ABSENCE OF LIGHT 4.1. INTRODUCTION Over the last few years, large mass-independent fractionation (MIF) of mercury (Hg)

isotopes has been reported for kinetic photo-reduction (Bergquist and Blum 2007, Zheng

and Hintelmann 2009, 2010a, 2010b, Sherman et al. 2010, this thesis) and now also

photo-oxidation (Ghosh et al., in preparation). Kinetic, reductive biological processes

have been investigated in the absence of light and found not to generate MIF in Hg

(Kritee et al. 2007, 2008, 2009). Bergquist and Blum 2007 investigated dark, kinetic

abiotic reduction of aqueous Hg(II) in the presence of dissolved organic matter (DOM)

and observed no MIF. Zheng and Hintelmann 2010a, however, report that small MIF

signatures are observed during the dark reduction of aqueous Hg(II) in the presence of

organic matter (actual -values are not reported). A significant difference in

experimental design is that Zheng and Hintelmann 2010a allowed their reaction to

continue for a few weeks, while Bergquist and Blum 2007 monitored their reaction for a

few hours. Zheng and Hintelmann 2010a went on to push the dark reductive reaction

forward by adding Sn(II)Cl2 (without DOM) to their aqueous mercury solutions. This

resulted in MIF with the odd isotopes being preferentially lost from the substrate

reservoir (199Hg reaching -0.6 ± 0.05 ‰, 2SD). This is opposite the MIF generated by

photochemical reduction (odd isotopes are preferentially retained in the substrate

reservoir). Apparently the process causing MIF is associated with reduction and not with

phase change, since volatilization of Hg from aqueous solution was shown not to induce

MIF by (Zheng et al. 2007). The objective of this investigation was to see if dark

chemical reduction of aqueous Hg(II) in the presence of organic matter produces mass-

independent isotopic anomalies. In addition, it provided a way to investigate non-

photochemical contributors to MIF in the experiments described in Chapter 2.

Page 124: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

117

4.2. METHODS Equipment, sample vials and analytical containers were cleaned according to the process

described in Chapter 2 of this document. An aqueous solution of 2 mg/L Suwannee

River Fulvic Acid Standard (1S101F) from the International Humic Substance Society

(http://ihss.gatech.edu/ihss2/) was made up in a glass bottle and then spiked to a final

concentration of approximately 31 g/L (ppb) of Hg(II) in the form of Hg(NO3)2 (J.T.

Baker lot E04632, hereafter called Hg(II)). This solution was quickly subsampled to

measure the initial isotopic composition and then wrapped in aluminum foil to protect it

from light. It was left to stand and equilibrate for 3.5 hours. It was then poured into the

photochemical reactor described in Chapter 2 and in Figure 3.1, except that the reactor

was completely wrapped in aluminum foil at all times. Continuous sparging with Hg-free

air removed the Hg(0) from the reaction vessel as it formed. Once inside the reactor, the

Hg-fulvic acid solution was initially subsampled again to verify that no change in

isotopic composition occurred on standing. Ten mL of (get moles) SnCl2 were then

injected into the reactor through the sample withdrawal line. Subsamples were removed

from the aqueous reservoir with a syringe as the reaction progressed. All subsamples for

isotopic analysis were immediately preserved by spiking with concentrated BrCl.

Concentrations were analysed as in Chapter 2. Overall uncertainty in concentration was

set at ± 5%, the 2SD of the reference standard for that analytical session. Matrix

separation and transfer for isotopic analysis was carried out as in Chapter 2. Sample

recoveries after matrix transfer were mostly 98% and 100% respectively (± 3%, 2SD).

However, sample recoveries for R2, R3 and R4 were only 83 ± 3%. Two attempts were

made to trap R3 with no improvement in recovery despite the fact that it was bracketed

by NIST 3133 processes standards at 97-100 % recovery. On the second attempt, an

effort was made to simultaneously trap sample Hg running down the drain. Nine percent

of the original sample Hg was caught from the drain, indicating incomplete reduction of

the sample. Isotopic analysis, carried out as in Chapter 2, included process standards

with recoveries as low as 86%. This lowest recovery (86%) showed small offsets

associated with this incomplete reduction (less than 0.2 ¥) in mass-dependent

Page 125: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

118

fractionation, but no offset in mass-independent fractionation was observed. Analytical

uncertainty is determined as in Chapter 2 and given in Table 4.1.

4.3. RESULTS An aqueous solution of Hg(II) with 2 mg/L fulvic acid was poured into a photo-reactor

covered with aluminum foil. SnCl2, a reducing agent, was added subsequently while the

solution was sparged with Hg-free air to remove any Hg(0) generated. The amount of

Hg(II) lost via dark chemical reduction was estimated from the decreasing concentration

of Hg(0) in the reactor over time (Figure 4.1). Sampling was discontinued after loss in

the reactor reached 78 ± 5%. Mass-dependent fractionation (MDF) was observed with

the heavier isotopes being preferentially retained in the reactor (Figure 4.2). Complete

isotopic data are displayed in Table 4.1. MDF followed a Rayleigh distillation model,

with 202 = 0.9997 ± 0.0002, R2 = 0.85.

Mass-independent fractionation (MIF) was observed (Figure 4.3) with the odd isotopes

being preferentially retained in the reactor during the earlier part of the reaction, similar

to photo-reduction, but opposite what Zheng et al., 2010 observed. Maximum MIF was

observed in the subsample taken soonest after SnCl2 addition (201Hg = 0.16 ± 0.05 ‰

and 199Hg = 0.24 ± 0.07 ‰, Table 4.1). MIF decreased in magnitude thereafter with the

final data point (fraction remaining = 0.22 ± 0.05) showing 201Hg = 0.01 ± 0.05 ‰ and

199Hg = 0.10 ± 0.07 ‰). The ratio 199Hg/201Hg for both data points showing

significant MIF (4SE or more) is 1.50 ± 0.86, 2SE. The point at fR = 0.22 did not show

MIF of 4SE or more, and so the 199Hg/201Hg ratio was not calculated.

4.4. DISCUSSION

During the process of dark chemical reduction of aqueous Hg(II) by SnCl2 (in the

presence of organic matter) and subsequent evasion of the Hg(0), mass-independent

anomalies in isotopic composition are produced, as displayed in Figure 4.3. The

mechanism of MIF in this process is not immediately clear. Two plausible mechanisms

have been put forward to explain the MIF observed in odd isotopes of Hg. The

Page 126: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

119

mechanism normally associated with photochemical reduction is likely the magnetic

isotope effect (MIE). It is caused by hyperfine coupling between electrons and the

nuclear magnetic moment present in odd nuclei, which can speed up triplet-singlet

conversion in a geminate radical pair. From a singlet state, the radical pair can

recombine to form the starting products anew. Triplet-singlet conversion of a non-

magnetic radical pair is slower, so non-magnetic pairs are more likely to dissociate and

form reaction products (Bergquist and Blum 2009, Buchachenko 2001). Since the

reaction took place in the dark, there could not have been photo-generated radical pairs.

Investigation of other possible radical reactions under these experimental conditions is

beyond the scope of this thesis.

The other mechanism suggested to produce MIF in odd isotopes of mercury is the nuclear

volume effect (NVE). The NVE arises because the nuclear volume and charge radius do

not scale linearly with mass. It produces even/odd staggering of mass-independent

anomalies because the odd isotopes have ground-state energies closer to those of the

adjacent lower-mass even isotope (Bigeleisen 1996, Schauble 2007). Estimates of the

relationship 199Hg/Hg produced by the NVE vary, depending on how the

(experimentally determined) values for nuclear charge radii are chosen. Theoretical

estimations range from 1.65 (Wiederhold et al. 2010) to 2.7 (Schauble 2007) for

equilibrium conditions. Experimentally determined relationships include 2.0 ± 0.6

(Estrade et al. 2009) and 1.62 ± 0.06, (2SE, Ghosh et al. in preparation) for equilibrium

volatilization of pure Hg metal, and 1.54 ± 0.44, 2SE for equilibrium binding of aqueous

Hg(II) to various thiol groups (Weiderhold et al., 2010) and 1.60 ± 0.06, 2SE for kinetic

reduction of Hg(II) by SnCl2 in the dark (attributed principally to the NVE, Zheng and

Hintelmmann 2010a).

Different 199Hg/Hg ratios have been associated with particular chemical pathways,

for example 1.28 ± ~0.12 (2SE) with photo-degradation of methylmercury (Bergquist

and Blum 2007, Malinovsky et al. 2010), and 1.0 ± ~0.1 (2SE) for photo-reduction of

Hg(II) (Bergquist and Blum 2007). These ratios have the potential to identify processes

that transform Hg in nature (Bergquist and Blum 2009). Of the ratios mentioned above,

Page 127: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

120

the 199Hg/Hg observed in this study is closer to those associated with the NVE than

with the ratios associated with the MIE if the more recent estimates (Weiderhold et al.,

2010; Ghosh et al., in prep, Zheng and Hintelmann 2010a) of NV 199Hg/Hg (1.6)

are considered.

In Hg however, the NVE causes a concentration of the odd isotopes in the chemical

species with higher electron density at the nucleus, which is to say, odd isotopes will

preferentially be found as Hg(0) and Hg(II) will be enriched in even isotopes. In the

experiment described above, Hg(0) was continuously removed from the substrate

reservoir through sparging with Hg-cleaned ambient air. If the odd isotopes were

reacting preferentially to form Hg(0) then the substrate reservoir should have been

depleted in the odd isotopes. This is the opposite of what was observed. In addition, our

observations are contrary to what is found by Zheng and Hintelmann 2010a, who use the

NVE to explain their data for SnCl2 reduction. One difference in experimental design is

that Zheng and Hintelmann 2010a do not use both SnCl2 and organic matter in the same

reactor. A second significant difference is that they add the reducing agent, SnCl2 in tiny

increments such that a molar excesss (by several orders of magnitude) of Hg is always

maintained. In contrast, the initial injection of SnCl2 in this experiment created a vast ( x

106) molar excess of Sn(II) with respect to Hg. SnCl2 reduction of Hg(II) is prompt (as

demonstrated by the matrix transfer process in Chapter 2: Methods). It is likely that the

entirety of Hg(II) was reduced immediately and loss of Hg(0) from the reactor was

initially limited by diffusion velocity. This would result in an abundance of Hg(0) in the

substrate reservoir and increased time for it to react further in the reservoir before exiting.

Initial quantitative reduction of Hg(II) entails complete mass transfer to Hg(0) and would

not induce isotopic fractionation of any kind during the reduction step. The pattern of

data points (Figure 4.3) suggests that a large positive MIF signature is imparted early in

the progress of the reaction and then slowly erased by a different processes that depletes

odd isotopes preferentially in the reservoir. The NVE might be invoked to explain the

decreasing MIF signature over time, but the question remains of how the positive MIF

signature was imparted in the first place. It is unclear what if any implications these

Page 128: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

121

findings have for non-photochemical contributors to MIF in the experiments described in

Chapter 2. Because of quite different experimental conditions, they are not likely

comparable.

4.5. CONCLUSION

Mass-independent anomalies in isotopic composition are observed during dark chemical

reduction of aqueous Hg(II) by SnCl2 in the presence of organic matter. Whether they

may be attributed to the MIE, the NVE, to some combination of the two or to another

mechanism entirely requires further investigation.

Page 129: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

122

FIGURES

Figure 4.1. Mercury concentrationin the substrate reservoir over time for dark chemical

reduction of Hg(II) by SnCl2 in the presence or organic matter. Error in time elapsed is ±

1 minute. Error in concentration is ± 0.05, which is 2SD of the reference standard for one

analytical session.

Page 130: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

123

Figure 4.2. Mass-dependent isotopic fractionation (202Hg) of the substrate reservoir as a

function of Hg remaining for dark chemical reduction of Hg(II) by SnCl2 in the presence

of organic matter. Line shown is a Rayleigh distillation model estimated from the data.

202 = 0.9997 ± 0.0001

Page 131: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

124

Figure 4.3. Mass-independent isotopic fractionation (199Hg), observed in the odd

isotopes only, as a function of inorganic Hg(II) remaining for dark chemical reduction of

Hg(II) by SnCl2 in the presence of organic matter.

Page 132: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

125

Page 133: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

126

Page 134: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

127

LITERATURE CITED

Bergquist, B. A. and Blum, J. D., 2007. Mass-dependent and -independent fractionation of Hg isotopes by photoreduction in aquatic systems. Science 318, 417-420.

Bergquist, R. A. and Blum, J. D., 2009. The Odds and Evens of Mercury Isotopes: Applications of Mass-Dependent and Mass-Independent Isotope Fractionation. Elements 5, 353-357.

Bigeleisen, J., 1996. Nuclear size and shape effects in chemical reactions. Isotope chemistry of the heavy elements. Journal of the American Chemical Society 118, 3676-3680.

Buchachenko, A. L., 2001. Magnetic isotope effect: Nuclear spin control of chemical reactions. Journal of Physical Chemistry A 105, 9995-10011.

Carignan, J., Estrade, N., Sonke, J. E., and Donard, O. F. X., 2009. Odd Isotope Deficits in Atmospheric Hg Measured in Lichens. Environmental Science & Technology 43, 5660-5664.

Estrade, N., Carignan, J., Sonke, J. E., and Donard, O. F. X., 2009. Mercury isotope fractionation during liquid-vapor evaporation experiments. Geochimica Et Cosmochimica Acta 73, 2693-2711.

Kritee, K., Barkay, T., and Blum, J. D., 2009. Mass dependent stable isotope fractionation of mercury during mer mediated microbial degradation of monomethylmercury. Geochimica Et Cosmochimica Acta 73, 1285-1296.

Kritee, K., Blum, J. D., and Barkay, T., 2008. Mercury Stable Isotope Fractionation during Reduction of Hg(II) by Different Microbial Pathways. Environmental Science & Technology 42, 9171-9177.

Kritee, K., Blum, J. D., Johnson, M. W., Bergquist, B. A., and Barkay, T., 2007. Mercury stable isotope fractionation during reduction of Hg(II) to Hg(0) by mercury resistant microorganisms. Environmental Science & Technology 41, 1889-1895.

Laffont, L., Sonke, J. E., Maurice, L., Hintelmann, H., Pouilly, M., Bacarreza, Y. S., Perez, T., and Behra, P., 2009. Anomalous Mercury Isotopic Compositions of Fish and Human Hair in the Bolivian Amazon. Environmental Science & Technology 43, 8985-8990.

Malinovsky, D., Latruwe, K., Moens, L., and Vanhaecke, F., 2010. Experimental study of mass-independence of Hg isotope fractionation during photodecomposition of

Page 135: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

128

dissolved methylmercury. Journal of Analytical Atomic Spectrometry 25, 950-956.

Sherman, L. S., Blum, J. D., Johnson, K. P., Keeler, G. J., Barres, J. A., and Douglas, T. A., 2010. Mass-independent fractionation of mercury isotopes in Arctic snow driven by sunlight. Nature Geoscience 3, 173-177.

Yang, L. and Sturgeon, R., 2009. Isotopic fractionation of mercury induced by reduction and ethylation. Analytical and Bioanalytical Chemistry 393, 377-385.

Zheng, W., Foucher, D., and Hintelmann, H., 2007. Mercury isotope fractionation during volatilization of Hg(0) from solution into the gas phase. Journal of Analytical Atomic Spectrometry 22, 1097-1104.

Zheng, W. and Hintelmann, H., 2009. Mercury isotope fractionation during photoreduction in natural water is controlled by its Hg/DOC ratio. Geochimica Et Cosmochimica Acta 73, 6704-6715.

Zheng, W. and Hintelmann, H., 2010a. Isotope Fractionation of Mercury during Its Photochemical Reduction by Low-Molecular-Weight Organic Compounds. Journal of Physical Chemistry A 114, 4246-4253.

Zheng, W. and Hintelmann, H., 2010b. Nuclear Field Shift Effect in Isotope Fractionation of Mercury during Abiotic Reduction in the Absence of Light. Journal of Physical Chemistry A 114, 4238-4245.

Page 136: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

129

CHAPTER 5: CONCLUSIONS Mass-independent fractionation (MIF) of Hg isotopes has the potential to highlight and

track photochemical transformations among the many processes that affect the

environmental fate of Hg. A solid mechanistic understanding of MIF in nature is

necessary, however, before it can be applied as a tool. Two plausible mechanisms have

been proposed to cause MIF in Hg isotopes: The magnetic isotope effect (MIE) and the

nuclear volume effect (NVE). The MIE is the most likely mechanism behind large (> 0.2

‰) MIF observed during photochemical reduction of Hg species (Bergquist and Blum,

2007; Malinovsky et al., 2010; Zheng and Hintelmann, 2009; 2010b). Photochemical

reduction is a significant pathway whereby water bodies release their burden of Hg back

to the atmosphere, and photo-degradation is an important sink (and detoxification

mechanism) for the MeHg in water bodies. These processes are strongly influenced by

the intensity and frequency of sunlight available. This thesis investigated whether and

how the frequency and intensity of sunlight available affected MIF in photo-reduction

and clarified the potential of MIF to document these processes. It also a tested the

hypothesis that the MIE was the mechanism behind the large MIF signatures observed

during photo-reduction of Hg species.

Experiments were conducted to determine whether removing the UVA or the UVB

portions of the electromagnetic spectrum affected the expression and magnitude of the

MIF observed during the aqueous photoreduction of MeHg and Hg(II) respectively.

Quartz reactors containing either Hg(II) or MeHg in aqueous solutions with organic

matter were simultaneously exposed to three different light regimes to examine the effect

on photo-reduction and on the isotopic fractionation of the respective Hg species.

Replicate sets of experiments were carried out for each species, with the first groups in

July and the second groups in September. Blocking UVB radiation had no effect on total

photoreduction of MeHg. It had some effect on total reduction of Hg(II) in the

September replicates (~70% loss under full spectrum and ~40% loss with UVB blocked)

but no effect in the July replicates. For both MeHg and Hg(II), however, blocking UVB

radiation suppressed MIF significantly (by ~30 to 50% for Hg(II); by ~70% or more for

Page 137: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

130

MeHg). In addition, the lower light intensity in September was accompanied by a

significant decrease in the magnitude of MIF in Hg(II).

This is evidence of a mechanism that is stimulated by photo-excitation and has a certain

energy threshold. The magnetic isotope effect, expressed through photo-excitation of a

solvated complex to a radical pair, is such a mechanism and our experiments solidly

support it as the cause of large MIF signatures during photochemical reduction of Hg

species. Exposure to UVA radiation produced significant MIF in both Hg(II) and MeHg.

Electronic excitation energies are influenced by many factors such as the atoms in the

complex, the chemical composition of the solution and the ambient temperature. A

mechanism where the primary step is electronic excitation of a solvated complex is

compatible with a range of excitation energies. Visible light can contribute slightly to

MIF during photochemical reduction of Hg(II) as well.

For both species, we have demonstrated cases where comparable amounts of total photo-

reduction are accompanied by MIF of significantly different magnitude. This confirms

there are multiple pathways of photo-reduction for both MeHg and Hg(II) in freshwater

reservoirs. Some of these pathways are spin-selective (since they express the MIE) and

some are not. Pathways that express the MIE have spin-selective processes at or before

their rate-limiting steps. This knowledge can be used to elucidate the photochemistry of

Hg in natural waters. The observation that MIF in both Hg(II) and MeHg can be

significantly reduced while total photo-reduction remains constant means that

quantitatively relating MIF and photo-reduction in natural freshwater systems will be

challenging.

In all Hg(II) reduction experiments in this thesis, the ratio 199Hg/201Hg falls within

error of 1.0, even when photo-reduction and MIF are severely reduced. If this ratio is

indicative of the reductive pathway, then the mechanism of MIF is not changing even as

the magnitude is reduced by screening higher energy radiation. If the mechanism of

Hg(II) MIF was constant for all experimental conditions, this suggests that direct

photolysis of mercury(II)-organic compounds takes place to some extent even in the

Page 138: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

131

range of visible radiation (as in the September, no UVB experiment). The average ratio

199Hg/201Hg for all MeHg experiments in this thesis was 1.35 ± 0.16 (2SE), regardless

of whether or not UVB radiation was blocked. The consistency of this ratio among

MeHg experiments suggests the mechanism of photo-reduction was the same regardless

of changes in the radiation available. The significant difference of this ratio between

MeHg and Hg(II) suggest that the pathway(s) of photo-reduction are not the same for the

organic Hg species as for the inorganic one.

The ratio 199Hg/201Hg for Hg(II) in this thesis is consistent with other experimental

results for aqueous photo-reduction of Hg(II) (1.00 ± 0.02, 2SE,(Bergquist and Blum,

2007). The ratio 199Hg/201Hg for MeHg in this thesis is consistent with other

experimental results for aqueous photoreduction of MeHg (1.36 ± 0.02, 2SE, Bergquist

and Blum, 2007, 1.28 ± 0.06, 2SD, Malinovsky et al., 2010). The ratios for aqueous

photoreduction of both species (in this thesis) are significantly different from

experimentally determined ratios 199Hg/201Hg for other transformative processes where

MIF has been attributed to the NVE. These process include the equilibrium

volatilization of pure Hg metal (2.0 ± 0.6, Estrade et al., 2009 and 1.62 ± 0.06 (2SE),

Ghosh et al., in preparation) and also equilibrium binding of aqueous Hg(II) to various

thiol groups (1.54 ± 0.22, 2SE, Weiderhold et al., 2010). This contributes to the evidence

suggesting that the ratios of the mass-independent isotopic signatures are unique to the

reaction pathways.

In a separate experiment not directly related to the main research objectives of this thesis,

and under different experimental conditions, it was observed that the process of dark

chemical reduction of aqueous Hg(II) by SnCl2 (in the presence of organic matter) and

subsequent evasion of the Hg(0), produced mass-independent anomalies in Hg isotopic

compositions. These anomalies are not immediately attributable to either the MIE or the

NVE and illustrate that much remains to be discovered about the processes that

fractionate Hg isotopes.

Page 139: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

132

LITERATURE CITED Bergquist, B. A. and Blum, J. D., 2007. Mass-dependent and -independent fractionation

of Hg isotopes by photoreduction in aquatic systems. Science 318, 417-420.

Estrade, N., Carignan, J., Sonke, J. E., and Donard, O. F. X., 2009. Mercury isotope fractionation during liquid-vapor evaporation experiments. Geochimica Et Cosmochimica Acta 73, 2693-2711.

Ghosh, S., In preparation.

Malinovsky, D., Latruwe, K., Moens, L., and Vanhaecke, F., 2010. Experimental study of mass-independence of Hg isotope fractionation during photodecomposition of dissolved methylmercury. Journal of Analytical Atomic Spectrometry 25, 950-956.

Wiederhold, J. G., Cramer, C. J., Daniel, K., Infante, I., Bourdon, B., and Kretzschmar, R., 2010. Equilibrium Mercury Isotope Fractionation between Dissolved Hg(II) Species and Thiol-Bound Hg. Environmental Science & Technology 44, 4191-4197.

Zheng, W. and Hintelmann, H., 2009. Mercury isotope fractionation during photoreduction in natural water is controlled by its Hg/DOC ratio. Geochimica Et Cosmochimica Acta 73, 6704-6715.

Zheng, W. and Hintelmann, H., 2010b. Isotope Fractionation of Mercury during Its Photochemical Reduction by Low-Molecular-Weight Organic Compounds. Journal of Physical Chemistry A 114, 4246-4253.

Page 140: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

133

Page 141: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

134

Page 142: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

135

Page 143: MASS-INDEPENDENT FRACTIONATION OF ERCURY ISOTOPES … · deposits (Jensen and Jensen 1990, Martinez-Cortizas et al. 1999) and glacial ice cores (Vandal et al. 1993, Schuster et al

136