16treatise 9_04 geochemistry of mercury in the environment w_ f_ fitzgerald and c_ h_ lamborg

42
9.04 Geochemistry of Mercury in the Environment W. F. Fitzgerald University of Connecticut, CT, USA and C. H. Lamborg Woods Hole Oceanographic Institution, MA, USA 9.04.1 INTRODUCTION 108 9.04.1.1 The Global Mercury Cycle 110 9.04.2 FUNDAMENTAL GEOCHEMISTRY 112 9.04.2.1 Solid Earth Abundance and Distribution 112 9.04.2.2 Isotopic Distributions 113 9.04.2.3 Minable Deposits 113 9.04.2.4 Occurrence of Mercury in Fossil Fuels 114 9.04.3 SOURCES OF MERCURY TO THE ENVIRONMENT 114 9.04.3.1 Volcanic Mercury Emissions 114 9.04.3.2 Mercury Input to the Oceans via Submarine Volcanism 117 9.04.3.3 Low-temperature Volatilization 118 9.04.3.4 Anthropogenic Sources 119 9.04.3.5 Mining 119 9.04.3.6 Biomass Burning, Soil and Oceanic Evasion—Mixed Sources 120 9.04.3.7 Watersheds and Legacy Mercury 120 9.04.4 ATMOSPHERIC CYCLING AND CHEMISTRY OF MERCURY 121 9.04.5 AQUATIC BIOGEOCHEMISTRY OF MERCURY 124 9.04.5.1 Environmental Mercury Methylation 127 9.04.5.1.1 Near-shore regions 128 9.04.5.1.2 Open-ocean mercury cycling 129 9.04.5.1.3 Open-ocean mercury profiles 130 9.04.6 REMOVAL OF MERCURY FROM THE SURFICIAL CYCLE 132 9.04.7 MODELS OF THE GLOBAL CYCLE 134 9.04.8 DEVELOPMENTS IN STUDYING MERCURY IN THE ENVIRONMENT ON A VARIETY OF SCALES 136 9.04.8.1 Acid Rain and Mercury Synergy in Lakes 136 9.04.8.2 METAALICUS 136 9.04.8.3 Ambient Isotopic Studies of Mercury—Biological Fractionation of Mercury Isotopes 137 9.04.8.4 Tracing Atmospheric Mercury with 210 Pb and Br 137 9.04.8.5 Mercury and Organic Matter Interactions 138 9.04.8.6 Bioreporters—A New Technique for Ultra-trace Determinations of Mercury 139 9.04.9 SUMMARY 139 ACKNOWLEDGMENTS 140 REFERENCES 140 107

Upload: sgutier

Post on 11-May-2017

216 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

9.04Geochemistry of Mercuryin the EnvironmentW. F. Fitzgerald

University of Connecticut, CT, USA

and

C. H. Lamborg

Woods Hole Oceanographic Institution, MA, USA

9.04.1 INTRODUCTION 1089.04.1.1 The Global Mercury Cycle 110

9.04.2 FUNDAMENTAL GEOCHEMISTRY 1129.04.2.1 Solid Earth Abundance and Distribution 1129.04.2.2 Isotopic Distributions 1139.04.2.3 Minable Deposits 1139.04.2.4 Occurrence of Mercury in Fossil Fuels 114

9.04.3 SOURCES OF MERCURY TO THE ENVIRONMENT 1149.04.3.1 Volcanic Mercury Emissions 1149.04.3.2 Mercury Input to the Oceans via Submarine Volcanism 1179.04.3.3 Low-temperature Volatilization 1189.04.3.4 Anthropogenic Sources 1199.04.3.5 Mining 1199.04.3.6 Biomass Burning, Soil and Oceanic Evasion—Mixed Sources 1209.04.3.7 Watersheds and Legacy Mercury 120

9.04.4 ATMOSPHERIC CYCLING AND CHEMISTRY OF MERCURY 121

9.04.5 AQUATIC BIOGEOCHEMISTRY OF MERCURY 1249.04.5.1 Environmental Mercury Methylation 127

9.04.5.1.1 Near-shore regions 1289.04.5.1.2 Open-ocean mercury cycling 1299.04.5.1.3 Open-ocean mercury profiles 130

9.04.6 REMOVAL OF MERCURY FROM THE SURFICIAL CYCLE 132

9.04.7 MODELS OF THE GLOBAL CYCLE 134

9.04.8 DEVELOPMENTS IN STUDYING MERCURY IN THE ENVIRONMENT ON A VARIETY OF SCALES 1369.04.8.1 Acid Rain and Mercury Synergy in Lakes 1369.04.8.2 METAALICUS 1369.04.8.3 Ambient Isotopic Studies of Mercury—Biological Fractionation of Mercury Isotopes 1379.04.8.4 Tracing Atmospheric Mercury with 210Pb and Br 1379.04.8.5 Mercury and Organic Matter Interactions 1389.04.8.6 Bioreporters—A New Technique for Ultra-trace Determinations of Mercury 139

9.04.9 SUMMARY 139

ACKNOWLEDGMENTS 140

REFERENCES 140

107

Page 2: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

9.04.1 INTRODUCTION

Mercurial, the metaphor for volatile unpredict-able behavior, aptly reflects the complexities ofone of the most insidiously interesting andscientifically challenging biogeochemical cyclesat the Earth’s surface. Elemental mercury isreadily recognized as a silvery liquid at roomtemperature. Its gas phase is important geochemi-cally, since mercury and some of its compoundshave relatively high vapor pressures. Mercury(Hg, from the Latin hydrargyrum or “waterysilver”) is sulfur loving (i.e., chalcophilic) andextremely active biologically. It is mobilizedtectonically, and significant deposits are found inmineralized regions characterized by subductionzones and deep-focus earthquakes (Schluter,2000). Many of the major deposits are shown inFigure 1 (Kesler, 1994).

The remarkable and useful qualities of mercuryand its major mineralized form (cinnabar, HgS)have been well known for thousands of years. TheAlmaden mine in Spain, for example, the “richestknown single source of cinnabar and quicksilver,”has been active for more than 2,500 yr (Goldwater,1972). Mercury products (e.g., thermostats,batteries, switches, fluorescent lighting) and appli-cations (e.g., chlor-alkali production, dentistry,pharmaceuticals, catalysis) have been a practicalpart of modern life, while the wastes have beenquite detrimental. Today, mercury emissionsassociated with fossil fuel burning, especiallycoal, and high-temperature combustion processes(e.g., municipal waste incineration) represent theprimary sources of pollutant mercury to theenvironment on a global scale. As a result, mercuryemissions since the mid-nineteenth century appearto have been in step with increases in emissions ofCO2 (Lamborg et al., 2002a). During the pastdecade, emissions, discharges, and nonpointsource inputs of mercury appear to have peakedand may be diminishing in many developedcountries. Unfortunately, and on a global scale,declines are countered by increases in anthropo-genic mercury releases from developing nations,particularly in Asia (E. G. Pacyna and J. M. Pacyna,2002). Additionally, environmental gains throughpollution controls, remediation, and regulations aretempered by the large reservoir of historic mercury,the pollution “legacy,” residing in watershedsand sediments of many terrestrial, freshwater, andmarine environments.

Although mercury has been used in “springtonics,” as a “cure” for syphilis, and a panacea forother afflictions, it is now recognized as a highlytoxic trace metal that concentrates in aquatic foodwebs. According to Clarkson (1997), the principalhuman exposure to inorganic mercury species isfrom elemental mercury vapor, which is derivedprincipally from industries such as gold and silver

mining and chlor-alkali plants, and from dentalamalgams. Deleterious health effects (e.g., “MadHatters Disease”) have been known since ancienttimes, and as Clarkson states, “severe exposureresults in a triad of symptoms, erethism, tremor,and gingivitis.” Today, however, the principalmercury-related human health concern is associ-ated with exposure to the highly neurotoxicorganomercury species, monomethylmercury,MMHg. This exposure is due almost entirely toconsumption of fish and fish products (Fitzgeraldand Clarkson, 1991; National Research Council,2000). Inorganic mercury, whether natural orpollution derived, can be readily methylated inaquatic systems. Mercury methylation appears tobe predominately biotic, although some abioticproduction is likely in natural waters (Benoit et al.,2003; Gardfeldt et al., 2003). In situ methylationof “reactive” or bioavailable mercury by sulfate-reducing bacteria has been documented to result inthe accumulation of MMHg in freshwater food-webs and fish (e.g., Westoo, 1966; Wiener et al.,1990a; Gilmour and Henry, 1991; Watras andBloom, 1994; Watras et al., 1994; Hall et al.,1998). This linkage is quite evident in the elevatedMMHg in fish from reservoirs created by damconstruction and subsequent flooding of land-scapes (e.g., Cox et al., 1979; Bodaly et al., 1984;Tremblay et al., 1998). A similar process isthought to occur in the marine water column,possibly first through the formation of dimethylHg(DMHg) followed by decomposition to MMHg(Mason and Fitzgerald, 1993; Mason et al., 1998;Mason and Sullivan, 1999). Microbially mediatedmethylation continues to amplify the insidious-ness of current and historic mercury pollution andhealth risks to wildlife and humans. Indeed,toxicologically, methylation of inorganic mercuryis the most important transformation affecting thebehavior and fate of mercury in aquatic systems.Its accumulation in freshwater and marine fish canreach levels that not only pose a threat to humanhealth (Grandjean et al., 1997; Davidson et al.,2000), but can reduce the reproductive success ofpiscivorous wildlife (e.g., Scheuhammer, 1991;special section of Environmental Toxicologyand Chemistry, 1998, 17/2: 137–227, 12 papers,M. W. Meyer, editor) and the fish themselves(Wiener et al., 1990b; Wiener and Spry, 1996;Hammerschmidt et al., 2002). Wiener et al.(2002) presented an up-to-date review of mercurytoxicology in a number of different animalspecies.

MMHg poisoning is known as “Minamatadisease.” Unfortunately, and between 1950 and1975, major industrially related mass poisonings,severe debilitation, and many deaths occurred inMinamata and Niigata, Japan and in Iraq. TheJapanese poisonings resulted from consumptionof locally caught fish and seafood that had been

Geochemistry of Mercury in the Environment108

Page 3: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Figure 1 The global Hg belt (source Kesler, 1994).

Page 4: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

contaminated principally by MMHg dischargedwith wastewater from factories making acet-aldehyde (Chisso Co. Ltd and Showa Denko Co.Ltd.). The MMHg was synthesized abiotically as aby-product during the production of acetaldehyde(inorganic mercury was used as a catalyst). In theIraqi tragedy, the source was contaminated bread,which had been made with flour, unknowinglymilled from wheat treated with MMHg as afungicide (Bakir et al., 1973). There is anextensive scientific, medical, and general litera-ture, as well as news accounts of these tragedies.The following works on the Japanese poisoningsprovide an overview and useful starting point for amore in-depth examination: Smith (1975), JapanPublic Health Association (2001), The SocialScientific Study Group on Minamata Disease(2001) and George (2001). There are otherexamples of severely mercury contaminated sitesand the interested reader is directed to thevolume entitled Mercury Contaminated Sites—Characterization, Risk Assessment and Remedia-tion, edited by Ebinghaus et al. (1998), which alsocontains descriptions of the Minamata situation.

In the mid-twentieth century, Goldschmidt(1954) cryptically summarized knowledge of theenvironmental cycling of mercury as, “not muchinformation is available concerning the geochem-istry of mercury.” Moreover, and as he empha-sized for that period, “most of the modernanalytical data are due to A. Stock and co-workers,” which were derived from their pre-World War II studies (e.g., Stock and Cucuel,1934). In contrast, and at the beginning of thetwenty-first century, dozens of environmentallyrelated mercury publications appear yearly. Thereis an abundance of distributional data, anenhanced knowledge of the biogeochemicalcycling of mercury and the impact of anthropo-genic activities, a fuller appreciation of the utility,dangers, and complexities characterizing themobilization, interactions, and fate of this bio-logically active element, and an awareness ofdaunting challenges inherent in studying a metalthat includes a gas phase as a major feature ofits biogeochemistry at the Earth’s surface. Thepotential risks of human exposure to MMHg,especially prenatally, and the potential deleteriousecological consequences from localized to global-scale mercury pollution have given much impetusto mercury studies and regulatory activitiesinternationally. There have been six internationalconferences on “Mercury as a Global Pollutant,”since 1990. The city of Minamata, the venue forthe sixth conference in 2001, provided confereeswith poignant evidence of the tragic legacy ofmercury contamination. The abstracts and publi-cations from these broadly based meetings, whichinclude basic biogeochemistry, environmental andpollution-related studies, ecological and human

toxicology, and analytical developments, chro-nicle the rapid worldwide expansion of mercuryresearch and knowledge.

This chapter focuses principally on the low-temperature environmental biogeochemistry ofmercury. We are presenting current understandingof mercury cycling at the Earth’s surface (soils,sediments, natural waters, and the atmosphere).Our coverage, as appropriate, will include theanthropogenic interferences (i.e., mercury pol-lution) and biological mediation, which affectsignificantly the speciation, behavior, and fate ofmercury in the environment. The reader will bereferred to other chapters for complementaryinformation on mercury geochemical cycling inand among the various earthly reservoirs.

9.04.1.1 The Global Mercury Cycle

Major features of the global mercury cycle havebeen illustrated using the relatively simple three-reservoir mass balance developed by Mason,Fitzgerald, and Morel in 1994 (“MFM”). Refine-ments are considered in Section 9.04.7. The MFMmodel provides estimates of natural and anthro-pogenic fluxes and an assessment of impact fromhuman-related mercury emissions on the naturalcycle for 1990 (Figure 2(a)). The pre-industrialsituation is shown in Figure 2(b). It is evident thatthe atmosphere and oceans play important roles inthe distribution and redistribution of mercury at theEarth’s surface. Indeed, atmospheric mercurydeposition to the oceans greatly exceeds riverineinputs. The MFM model also suggests that thenatural cycle of mercury has been severelyperturbed. Human-related mercury emissionsdominate the cycle such that most of the mercuryin the atmosphere and surface oceans is anthro-pogenic. Moreover, the integrated estimate fortotal loadings from globally dispersed, anthropo-genic mercury emissions between 1890 and 1990 is1,000 Mmol (200 kt). The MFM simulation pre-dicts that ,95% of this mercury is sequestered interrestrial systems, with the remainder in the oceansand atmosphere. The MFM analysis suggestedthat about half of the modern pollution-relatedemissions (20 Mmol yr21) enter the global cyclewhile the other half is deposited near sources. TheMFM estimate for the globally dispersed anthro-pogenic contribution is similar to the most recentestimate of 9.6 Mmol yr21 (for 1995) reported byPacyna and Pacyna (2002).

Elemental mercury, a monatomic gas, is thedominant atmospheric form and has a longresidence time in the troposphere (.1 yr;Fitzgerald et al., 1981; Slemr et al., 1981;Lindqvist et al., 1991; Lamborg et al., 2000,2002a). Such longevity allows emissions ofmercury to the atmosphere from natural and

Geochemistry of Mercury in the Environment110

Page 5: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

anthropogenic sources to be dispersed widelyacross the Earth’s surface. The redox couple ofHg0 with the stable mercuric ion (Hg0/Hg2þ;E 8 ¼ 0.85 V) is similar to that of the Fe(II)/Fe(III)couple, thereby providing the potential fordynamic oxidation and reduction cycling in therange of common environmental redox (i.e., pe)conditions. A unique and important example ofthis is the biologically and abiotically mediatedreduction of Hg2þ resulting in widespread super-saturation of Hg0 in freshwater and saltwater andsubsequent evasion (Kim and Fitzgerald, 1986;Mason and Fitzgerald, 1993; Amyot et al., 1997;Rolfhus, 1998). Field and laboratory observationsin aqueous systems have documented the import-ant influence of photochemical and bacterialreactions on the in situ reduction of Hg2þ

(Mason and Fitzgerald, 1993; Amyot et al.,1997; Rolfhus, 1998; Rolfhus and Fitzgerald,2001), and therefore link Hg0 production andevasion from the ocean to the complex biogeo-chemical cycles of carbon, nitrogen, phosphorus,and sulfur. Through evasion, the ocean prolongs

the residence of mercury at the Earth’s surface andexacerbates the human perturbation of the globalmercury cycle. The MFM analysis shows suchoceanic emission of Hg0 to be substantial(10 Mmol) and comparable to the atmosphericinput of mercury to the oceans.

Although the marine biogeochemical cyclingof mercury is greatly simplified in the MFMsimulation, the importance of the ocean as both asink for atmospheric deposition as well as a sub-stantial source of mercury to the atmosphere (Kimand Fitzgerald, 1986; Nriagu, 1989; Lindqvistet al., 1991; Mason et al., 1994) is consistent withmodels that use more realistic physical andbiogeochemical oceanic dynamics (Hudson et al.,1995; Mason et al., 2001; Lamborg et al., 2002a).

The oceans actively and significantly partici-pate in the transport and transformation of thistoxic metal. Thus, there is an obvious need toincrease knowledge and understanding concern-ing the biogeochemical cycling of mercury andMMHg and the impact of anthropogenicallyrelated inputs in the marine environment. Acursory examination of papers presented at theinternational mercury conferences is sufficient toshow that there has been much interest in andsupport for studies examining linkages betweenthe cycling of mercury in the atmosphere,anthropogenic emissions/discharges, depositionto terrestrial systems, and the bioaccumulation ofMMHg in freshwaters (e.g., Mercury in Temper-ate Lakes Program—Watras et al., 1994;METAALICUS Project; USEPA Mercury StudyReport to Congress, 1997; Ebinghaus and Kruger,1996; Petersen et al., 1995; The Nordic Network—Lindqvist et al., 1991; Iverfeldt, 1991a). Incontrast, mercury cycling in the oceans hasreceived scant attention. This is unfortunate asthere is compelling evidence, for example, that themercury content in fish has been increasing overthe past several decades in the North Atlantic(e.g., Monteiro and Furness, 1997). Moreover, in2001, growing recognition of human exposure toMMHg from marine fish and fish productsprompted the US Food and Drug Administration(USFDA) to place four pelagic marine fish(tilefish, king mackerel, swordfish, and shark) ontheir consumer advisory list (USFDA ConsumerAdvisory, 2001). The advisory emphasizes theneed for women of childbearing age to limit theirconsumption of marine fish to 12 ounces perweek.

A more stringent fish consumption advisoryreference dose (Schober et al., 2003) was issuedby the US Environmental Protection Agency(USEPA) in 2001. Women of childbearing agewere cautioned to limit their MMHg intaketo 0.1 mg kg21 of body weight per day. It isilluminating to translate this recommendation intofish consumption. A summary of the MMHg

Anthropogenicemissions20

10

1

5

5

3

3

1

5

10

1015

Wet and drydeposition

Naturalemissions

All fluxes in Mmol yr–1

Naturalemissions

Wet and drydeposition

0.3

Mixed layer54 Mmol

Mixed layer18 Mmol

Particleremoval

10Oceanicevasion

Air 25 Mmol

Hgp Hg0

Hg0 CH3Hg+

Hg2+ Hgp

98% Hg0

2% Hgp

Deposition Hg2+

Air 8 MmolHg

p Hg098% Hg0

2% Hgp

Deposition Hg2+

100 m

100m

Hg0 CH3Hg+

Hg2+ Hgp

Oceanicevasion

Particleremoval

(a)

(b)

0.3

Figure 2 Fluxes of Hg on a global scale. a) Thecurrent condition and (b) the pre-industrial condition.All fluxes in Mmol of Hg yr21. Note the balance ofdeposition to and evasion from the ocean, making soilsthe primary sink on decade/century timescales. Oneresult of human activity has been a tripling of Hg in theatmosphere and surface ocean (after Mason et al.,

1994).

Introduction 111

Page 6: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

content in some major marine fish and shellfish ispresented in Table 1. Using fresh tuna with anaverage MMHg concentration of 0.32 mg g21 wetweight (Table 1) as an example, the recommendedweekly consumption for a 50–70 kg woman ofchildbearing age would be 35–49 mg or theequivalent of ,110–155 g (,4–5.5 oz, half atypical can) of fresh tuna. This is much smallerthan the USFDA recommendation, which isequivalent to 0.5 mg kg21 of body weight perday (however, a maximum of 12 oz of fish perweek is recommended). The current provisionaladvisory from the Joint Food and AgriculturalOrganization/World Health Organization ExpertCommittee on Food Additives (JECFA, 2000)is 0.5 mg kg21 of body weight per day. Forfurther information regarding MMHg andfish consumption, the reader is referred to theuseful and informative publication from the USNational Academy of Sciences entitled “Toxico-logical Effects of Methylmercury” (NAS, 2000).The examining committee from the NationalResearch Council of the NAS reached thefollowing consensus: “the value of EPA’s currentreference dose for MeHg, 0.1mg kg21 per day isscientifically justifiable for the protection ofpublic health.”

In summary, mercury entering the marineenvironment may continue to actively participatein aquatic chemistry, while much of the mercurydeposited from the atmosphere to terrestrialsystems is sequestered. Given the importance ofthe oceans as a whole in global mercury cycling,and international concerns and issues regardinghuman exposure to MMHg through marine fish andseafood consumption, the current situation is oneof insufficient study and undersampling. Indeed,there have been relatively few open-ocean cruisesto examine mercury biogeochemistry (e.g., Kimand Fitzgerald, 1986; Gill and Bruland, 1987;Gill and Fitzgerald, 1988; Mason and Fitzgerald,1993; Dalziel and Yeats, 1985; Dalziel, 1992;

Cossa et al., 1997; Mason et al., 1998; Mason andSullivan, 1999; Lamborg et al., 1999; Leermakerset al., 2001), and almost no longer-term, seasonallyoriented mid-ocean studies. Even in the moreaccessible near-shore zones, the biogeochemistryof mercury is understudied. In this regard, estuariesand adjacent coastal waters, as regions of highbiological productivity, MMHg production andfishery activity, are of special interest. They aremajor repositories for natural and pollutantmercury (see Section 9.04.6).

9.04.2 FUNDAMENTAL GEOCHEMISTRY

In this section, we consider aspects of thefundamental geochemistry and biogeochemistryof mercury used later in the chapter. These topicsinclude: (i) solid Earth abundance and distri-bution, (ii) isotopic composition and recentadvances in mercury cosmochemistry, (iii) theformation and distribution of minable mercurydeposits, and (iv) mercury in fossil fuels.

9.04.2.1 Solid Earth Abundance and Distribution

A summary of mercury data from the report ofTurekian and Wedepohl (1961) is shown inTable 2. Due to the chalcophilic nature of itsassociations, mercury is found in higher abun-dances in intrusive magmatic rocks and locationsof subaerial and submarine volcanism. Peakconcentrations in these rocks may be as high asseveral percent in ore-grade minerals (e.g., 35%mercury in sphalerite; Ozerova, 1996). Also as aresult of this association, the distribution ofhighest mercury concentrations in rocks at thesurface and near surface mirrors regions of currentand past tectonic activity and has been describedas the “global mercury belt” (e.g., Gustin et al.,2000). As noted in greater detail in Section 9.04.3,mercury concentrations in soils weathered from

Table 2 Mercury content of selected rocks andsediments.

Rock type Hg content(ppm)

Ultrabasic igneous 0.0Xa

Basaltic rocks 0.09High- and low-Ca granites 0.08Syenites 0.0XShales 0.4Sandstones 0.03Carbonates 0.04Deep-sea carbonate 0.0XSedimentDeep-sea clays 0.X

Source: Turekian and Wedepohl (1961).a The X notation indicates order of magnitude estimate.

Table 1 Levels of total mercury (mg g21 wet weight)in seafood (USFDA, 2001). Most (.95%) of the total

mercury in edible fish tissue is MMHg).

Fish species Mean (range) na

Tilefish 1.45 (0.65–3.73) 60Swordfish 1.00 (0.10–3.22) 598King Mackerel 0.73 (0.30–1.67) 213Shark 0.96 (0.05–4.54) 324Tuna (fresh or frozen) 0.32 (ND–1.30) 191Tuna (canned) 0.17 (ND–0.75) 248Atlantic cod 0.19 (ND–0.33) 11Pollock 0.20 (ND–0.78) 107Mahi mahi 0.19 (0.12–0.25) 15American lobster 0.31 (0.05–1.31) 88

Sources: Grieb et al. (1990), Bloom (1992), and Hammerschmidt et al.(1999). ND denotes that the mercury level was not detectable.a Number of samples analyzed.

Geochemistry of Mercury in the Environment112

Page 7: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

this material can be very high as well (e.g.,Steamboat Springs, NV, USA: 1.2–14.6 ppm;Gustin et al., 2000) and represent a potentiallysignificant source of mercury to the atmosphere ata variety of spatial scales through low-temperaturevolatilization. These rocks and their weatheredproducts are rich in other metals as well, andemission of mercury from soils and rock has beenused as a tool for large-scale ore and petroleumexploration as well as an indicator of tectonicactivity (e.g., McCarthy, 1968; Varekamp andBuseck, 1983; Klusman and Jaacks, 1987). Thesegeneral characterizations are significant then, aswe later consider the cause for concentrationsof mercury in soils removed from the mercurybelt that are also elevated above the crustalabundances.

Also indicated in Turekian and Wedepohl’scompilation is a relatively high concentration ofmercury in sedimentary material rich in organiccarbon, such as shales. Mercury associations withorganic matter are considered in later sections(Sections 9.04.5 and 9.04.8). Here, we stress thatsuch associations can lead not only to higherconcentrations of mercury in these types of rockunits but also in the transport of mercury awayfrom sites of sediment accumulation as a resultof petroleum movement (White, 1967; Fein andWilliams-Jones, 1997). The mercury content ofmajor rock types has not been systematicallyrevisited since Turekian and Wedepohl’s report,and there is evidence that some of their valuesmay be overestimates. As an example, the recentrecovery and analysis of glacial till in Glacier BayNational Park by Engstrom and Swain (1997)indicated much lower concentrations of mercury(,10 ppb). It is reasonable to assume, however,that mercury concentration trends across rocktypes suggested by Turekian and Wedepohl andthe geochemistry they suggest are valid.

9.04.2.2 Isotopic Distributions

Mercury has a relatively even distribution ofits seven stable isotopes (196, 0.15%; 198, 10.0%;199, 16.7%; 200, 23.2%; 201, 13.2%; 202, 29.8%;204, 6.8%; Friedlander et al., 1981; Lauretta et al.,2001). This pattern presented cosmochemistswith a formidable task when mercury isotopicdistributions in meteorites were examined (e.g.,Jovanovic and Reed, 1976; Thakur and Goel,1989). Analytical difficulties apparently resulted ininaccurate determinations of the bulk abundanceand isotopic composition of some meteorites,leading to the so-called “mercury problem”;examined meteorites did not show the same bulkabundance and isotopic distribution as terrestrialmaterial (Grevesse, 1970; Lauretta et al., 1999).Subsequent advances in mass spectrometry, andespecially the development of multi-collectors,

have shown that the isotopic distributionsof mercury in terrestrial and extraterrestrialmaterial are very similar (e.g., Allende meteorite:0.03–0.3 ppm; Lauretta et al., 2001).

With so many isotopes from which to choose,one might expect examination of mercury isotopicfractionation in natural media to be a profitablearea of research as it has been, for example, withlead, light metals, and nonmetals (e.g., Allemanet al., 2001; Richter et al., 1992; Hoefs, 1980). Todate, this has not been widely explored, but thereis little indication thus far of either biologicalfractionation in aquatic systems useful in tracingbiogeochemistry or of geogenic fractionationleading to isotopic signatures of particular ores(Klaue et al., 2000; Evans et al., 2001). The widerange of stable isotopes is being used in deliberateaddition experiments ranging from benchscaleto whole-watershed-scale (e.g., Hintelmann andEvans, 1997; Hintelmann et al., 2002). Theseadvances will be highlighted in Section 9.04.8.

The radioisotope 203Hg has played an importantrole in laboratory investigations of mercurybiogeochemistry (e.g., Gilmour and Riedel,1995; Stordal and Gill, 1995; Costa and Liss,1999). Continued production of 203Hg-enrichedmaterial has been curtailed recently, and thusfuture mechanistic studies will instead feature theuse of stable isotopes.

9.04.2.3 Minable Deposits

As mentioned, higher mercury concentrationsin rock and soil are associated with the globalmercury belt. However, the occurrence of minabledeposits is not continuous along this belt. Inaddition to Almaden (Spain), the most productivemercury mines included Idrija (Slovenia), NewAlmaden (California, USA), and Huancavelica(Peru). Very high concentrations of mercury havebeen reported to be associated with oceanichydrothermal sulfide chimneys and their weath-ered remains (e.g., Koski et al., 1994; Ozerova,1996; Stoffers et al.,1999). In all of these cases,mercury occurs almost exclusively as cinnabar(red HgS), with smaller amounts of metacinnabar(black HgS) and elemental mercury (often asinclusion with HgS in minerals such as sphaleriteand chalcopyrite). HgS is extremely insoluble(log Kso (cinnabar) ¼ 236.8; Martell et al., 1998)under typical surface water conditions, and thustransport of mercury from these mineral richenvironments at the Earth’s surface generallyinvolves sediment transport (e.g., Ganguli et al.,2000). The fate of cinnabar in anoxic sediments isaddressed in Section 9.04.5. Transport of mercuryto and from ore bodies invariably involveshydrothermal systems in the subsurface, withHgS solubility strongly controlled by fluidpH, temperature, chloride, sulfide, and organic

Fundamental Geochemistry 113

Page 8: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

carbon contents (White, 1967; Varekamp andBuseck, 1984).

9.04.2.4 Occurrence of Mercury in Fossil Fuels

Even with the stated stability of cinnabar,mercury is mobile in the surface environment.This is indicated by the relatively high concen-tration of mercury in organic-rich deposits, suchas fossil fuels and shales. As we will explorelater, mercury has high affinities for organiccarbon as well as sulfide. Interest in fossil fuelrecovery and burning as a source of mercury tothe environment has led to a few publishedstudies in this area and we summarize some ofthese data in Table 3. With notable exceptions,concentrations in coal appear higher than thosein oil, suggesting preferential burial of mercuryin terrestrial and coastal systems rather than inpelagic marine environments. Though mercuryconcentrations are not as high in various refinedpetroleum materials as in coal, the oil andgas recovery process often liberates largeamounts of mercury leading to localized con-tamination of marine sediments (e.g., Grieb et al.,2001). The concentration of mercury is suffi-ciently high in some crude petroleum materialsto also pose a corrosion hazard to the drillingand transportation apparatus and represents asignificant engineering problem (e.g., Wilhelm,1999; Bloom, 2000).

9.04.3 SOURCES OF MERCURY TO THEENVIRONMENT

There have been few well-designed studies toconstrain mercury emission estimates from natural

sources and allow global extrapolations. Indeed,some work has been extraordinarily misleading. Forexample, in flawed studies based on the accumu-lation of mercury and other metals in glacial ice,Jaworowski et al. (1981) estimated the annualmercury flow into the atmosphere at 1,000 Mmolwith an anthropogenic contribution at 50 Mmol!Low- and high-temperature volatilization processeswere offered by Jaworowski et al. as a potentialexplanation for the huge natural fluxes of mercury.Such inaccuracies as well as the paucity of reliableinvestigations begs the question “What is theappropriate flux range for global mercury emissionsfrom sources such as volcanism, biomass burning,and low-temperature volatilization from naturalwaters and soils?” The MFM simulation (Masonet al., 1994) of the global mercury cycle estimatesnatural terrestrial emissions at 5 Mmol annually.Such emissions would include inputs from subaerialvolcanism under erupting and non-erupting con-ditions, and the pre-industrial mercury fluxes frommineralized regions, forest fires, biological acti-vities, and natural waters. Today, volcanism andvolatilization from mineralized regions may be theonly purely natural sources of mercury, because, andas illustrated in Figure 2, anthropogenic mercurycontamination is present throughout the atmos-phere, biosphere, the terrestrial realm, and thehydrosphere. Thus, a portion of the emissionsfrom these secondary reservoirs represents recycledpollutant mercury; this component has often beenoverlooked when the source strengths from “naturalsources” have been compared and assessed relativeto modern anthropogenic mercury inputs.

9.04.3.1 Volcanic Mercury Emissions

The determination of global volcanicmercury fluxes from direct measurements is atbest a hazardous, expensive, labor-intensive,and, perhaps, impossible task. Reasonablywell-constrained global estimates, however, canbe achieved through use of elemental ratios and ageochemical indexing approach. Sulfur providesan appropriate index, since it is a major com-ponent of volcanic emanations, and there isagreement on its global flux to the atmosphere.An example of this approach is the work ofPatterson and Settle (1987), who combined Pb/S,Tl/S, and Bi/S ratios measured in volcanic gasescollected under eruptive and quiescent (fumarolic)conditions with the global volcanic sulfur fluxto “approximate global volcanic emissions ofthese three metals to the atmosphere.” One of thepresent authors (WFF), measured mercury duringthe Patterson and Settle study. Results from thisresearch were presented at two conferences(Fitzgerald, 1981, 1996), and Fitzgerald’s esti-mate of 40 t yr21 appears in the paper by Patterson

Table 3 Mercury content of fossil fuels.

Sample type Total Hg(ng g21)

CoalChina

a220

Std. Ref. Mat.b

77.4–433.2Global estimate

c20–1,000

Unrefined petroleumd

Crude oil ,d.l. to .7,000Condensate ,d.l. to .12,000

Refined petroleumd

Light distillates 1 ^ 3Utility fuel oil 1 ^ 1Asphalt 0.3 ^ 0.3Gasoline 0.2–3Diesel 0.4–3Kerosene 0.04Heating oil 0.59Naphtha 3–60Petroleum coke 0–250

a Wang et al. (2000). b Long and Kelly (2002). c Pacyna andPacyna (2002). d Wilhelm (2001) and the references therein.

Geochemistry of Mercury in the Environment114

Page 9: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

and Settle (1987). Given the potential importanceof volcanic mercury inputs, and the lack ofinvestigations, a detailed description of themercury study is included here.

Patterson, Settle, Buat-Menard (University ofBordeaux), Fitzgerald, and colleagues (seeAcknowledgments section) evaluated volcanicmetal fluxes using an experimental design basedon the hypothesis that metal volatilization wouldbe dependent on temperature, sulfur and halogencomposition of magma and mobilization byvolcanic gases. Therefore, volcanoes and fuma-roles were selected to provide a range of tempera-tures and S/Cl ratios. The characteristics of thevolcanic gas samples are tabulated in Table 4. Thefour volcanoes examined were Kilauea (Hawaii,USA), Mt. Etna (Sicily, Italy), Vulcano (AeolianIslands, Italy), and White Island (New Zealand).

Plumes were sampled for mercury using asimple gas train consisting of a pre-blanked(pyrolyzed) glass fiber prefilter stage forparticulate phases, followed by two gold trapsarranged in tandem to collect gaseous mercury(Fitzgerald and Gill, 1979; Bloom and Fitzgerald,1988). The fumarolic collections were made witha gas sampling train designed by Clair Pattersonand modified for mercury (Patterson and Settle,1987; shown here in Figure 3). Any mercuryassociated with aerosols or the gas phase thatescaped the two cold traps was collected by a gas-sampling train analogous to the plume samplers.The gas was pumped through this system at,1 L min21.

In general, and as illustrated for the 1977 studyof 100 8C fumarole at Kilauea, essentially all ofthe mercury and the other metals are trapped in the08 condensate (Table 5). Thus, fumarolic gascollectors for metal studies can be simplified. Theobserved Hg/S weight ratio in the Kilaueacondensate was 0.9 £ 1026. The Hg/S ratiosfrom the other investigations are summarized inTable 4. Values range from 0.7 to 14 £ 1026 witha suggestion that Hg/S may increase as the S/Cldecreases. Dedeurwaerder et al. (1982) were part

of the Mt. Etna study and their results for theBocca Nuevo plume are included in the summary.Their plume-sampling apparatus was based on theFitzgerald technique and similar to that describedabove. The sulfur-flux weighted mean volcanicHg/S from all locations is 5 £ 1026.

Estimates of global volcanic sulfur emissionsare summarized in Table 6. We have chosen a valueof 9 £ 106 t S yr21 as representative of the recentestimates. Therefore, by applying the determinedHg/S ratio, a global mercury flux from subaerialvolcanism is estimated to be ,45 t yr21, or0.23 Mmol annually. These average emissionsare only 5% of the natural flux of 5 Mmol yr21

estimated by Mason et al. (1994). Thus, and underlong-term mean conditions, other types of terres-trial volatilization processes for mercury woulddominate. Given this conclusion, it is important toplace additional constraints on the validity of the45 t yr21 estimate for subaerial volcanic mercuryemissions.

First, Varekamp and Buseck (1986) reported anaverage Hg/S weight ratio of 7.4 £ 1026 foremissions under non-erupting conditions fromMts. Hood, Shasta, and St. Helens in the UnitedStates, and from Mt. Etna. In 1981, theseinvestigators reported a very high Hg/S weight

Gas sampling cannister

Glass fiber filter

Switching valves

Au-coatedsand trap

Teflon tubing

Quartz tubinginlet

Volcanic gas

Flowmeters

Pump

Condensate0 ºC

Condensate–70 ºC

Figure 3 Hg sampling apparatus used to collectfumarole gas samples (after Patterson and Settle, 1987).

Table 4 Volcano sampling locations and gas characteristics.

Volcano Date Type of gas Temp.(8C)

S(wet gas)(mg S m3)

S/Cl(weight)

Cl/F(weight)

Hg/S(weight)

Hg/Bi(weight)

Kilauea Caldera 6/23/77 Fumarole 100 22,800 40 2 0.9Kilauea Pu’u O’o 5/12/84 Eruptive plume 1,140 110 9 0.7 0.0019Mt. Etna Bocca Nuova

a6/7/80 Eruptive plume 1,100 4.3 0.7 8 11–14

Mt. Etna SE Crater 6/8/80 Eruptive plume 1,100 2.5 0.7 4 ,13Volcano Site A 6/11/80 Fumarole 280 5,500 0.3 90 0.9White Island Site A 7/22/83 Fumarole 180 33,000 45 30 0.9White Island Site B 7/22/83 Fumarole 590 54,000 ,0.44 0.024White Island Site C 7/22/83 Fumarole

b650 180 0.0015

After Patterson and Settle, 1987.a Dedeurwaerder et al. (1982). b Collected using plume filter device.

Sources of Mercury to the Environment 115

Page 10: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

ratio of 6,000 £ 1026 for Mt. St. Helens undererupting conditions (Varekamp and Buseck,1981). However, the mercury and sulfur measure-ments were not measured simultaneously and,according to Varekamp (personal communi-cation), should not be used for a global-scaleanalysis of this kind. Thus, we suggest that a meanHg/S ratio of 5 £ 1026 and the annual emissionestimate of 45 t is of the appropriate magnitude.Nriagu and Becker (2003), using the same Hg/Sratio and a more detailed volcanic SO2 emissionsinventory, derived much the same conclusion(,95 t yr21; 60% eruptive, 40% degassing).

Second, the scale of volcanic mercury fluxes canbe approximated indirectly. For example, using theHg/Bi ratios listed in Table 4 as well as the Lambertet al. (1988) estimate for annual global bismuthemissions of 1,500 t, we obtain a range for theglobal volcanic mercury flux of 2–36 t yr21.Hinkley et al. (1999) suggests that the Lambertet al. values are a factor of 3 or 4 too high.

A comparable value of 18 t annually is obtainedusing Olafsson’s (1975) estimate of mercuryemissions (7 £ 105 g Hg/6 £ 1014 g ejecta) forthe volcanic eruption at Heimay, Iceland, and anestimate of 6 km3 (,15 £ 1015 g) for annual,subaerial lava production (,20% of total annuallava production being subaerial; Crisp, 1984;Varekamp, personal communication).

Thus, annual global volcanic mercury emissionsestimated or measured in several ways are less than0.5 Mmol (100 t), and our work yielded an averagevalue of 0.23 Mmol (45 t). This scaling serves twopurposes: (i) it provides a framework for further,needed studies of mercury cycling associated withvolcanism, and (ii) it provides a reasonably wellconstrained estimate for global emissions formodeling and assessment of human perturbationsof the natural mercury cycle. It is clear that averagevolcanic mercury inputs are small relative toestimates for modern anthropogenic mercuryfluxes (,9.6 Mmol; Pacyna and Pacyna, 2002) tothe global mercury cycle. Indeed, and using anatmospheric residence time of 1.25 yr (Lamborget al., 2002a), the average yearly terrestrialdeposition of mercury from volcanism would be,0.07 mg m22. Since volcanic eruptions vary intime and space, large-scale eruptions might beapparent in natural archives such as lake sediments(Chapter 9.03), and ice cores. Carefully collectedand dated lake sediments, however, do not revealunusually high accumulations in strata coincidentwith large explosive aperiodic volcanic eruptions(e.g., Pinatubo in 1991, Krakatau in 1883 orTambora in 1815). In contrast, and accordingto Schuster et al. (2002), volcanic mercurydepositional signals may be evident in a recentstudy of the Fremont Glacier in Wyoming, USA.

Table 5 Mercury and sulfur concentrations in volcanic gas at sulfur fumarole site, Kilauea caldera.

Collection Hg concentration Total Hg(ng)

Total S(g)

Hg/S ratio(weight)

0 8C condensate 120 ng g21 59,200 4.0 14.8 £ 1026

270 8C condensate 10 ng g21 150 0.1 1.5 £ 1026

Volcanic gas free of water 0.88 ng g21 230 60 3.9 £ 1029

Total 59,600 64 9.3 £ 1027

Ambient air 10.2 ng m23

Table 6 Estimates of the global annual sulfur flux from volcanic activity.

Method S Flux(106 t yr21)

References

Rate of S loss per volcano 5.0 Stoiber and Jepson (1973)3.8 Cadle (1975)4–5 Friend (1973)

S measurements from selected 9.4 Stoiber et al. (1987)volcanoes Nonexplosive (4.5)

Explosive (4.9)

Literature review 9 Spiro et al. (1992)

Satellite survey of SO2 emissions 6.5 Bluth et al. (1993)Nonexplosive (4.5;

Stoiber et al., 1987)Explosive (2.0)

Satellite and modeling 7.5–10.5 Halmer et al. (2002)

Geochemistry of Mercury in the Environment116

Page 11: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

For example, their results suggest an averageglobal mercury depositional peak increase of16 mg m22 yr21 from the 1883 Krakatau eruption,which released ,21 km3 of volcanic material(Rampino and Self, 1984). As the pre-eruptionmercury deposition was ,2 mg m22 yr21, theKrakatau event is extraordinarily prominent at,8 £ the background. In contrast, a mercurysignal associated with the June 1991 MountPinatubo eruption in the Philippines, whichreleased ,5 km3 ash and pumice and,50 £ 1012 g SO2 or 25 £ 1012 g S (USGS FactSheet 113-97), is not evident in the FremontGlacier ice core. If Mt. Pinatubo were assumed tobe analogous to Krakatau according to the Fremontice core, then the predicted mercury depositionwould be ,4 £ less or ,4 mg m22 yr21. Theanthropogenically enhanced background mercurydeposition for 1991 in the ice core was consider-ably larger at ,11 mg m22 yr21. Thus, and giventhe large pollution-derived mercury deposition, theuncertainty inherent in reconstructed fluxes and theassumptions, the lack of a mercury signal fromMt. Pinatubo could reasonably be expected.

We suggest, that the work of Schuster et al.(2002), while somewhat speculative and not wellconstrained, is provocative and stimulating. More-over, there is reasonable coherence with the Hg/Sfor volcanic emissions shown in Table 4. Forexample, the Hg/S of 5 £ 1026 when combinedwith the 25 £ 1012 g sulfur emission estimates forthe Mt. Pinatubo eruption yields a mercury inputof 125 £ 106 g mercury or ,0.6 Mmol. This isjust 6% of the 10 Mmol anthropogenic mercuryemissions estimated by the MFM simulation forthe global contribution in 1990. Therefore, themeasured Hg/S ratio suggests the Mt. Pinatuboeruption would not be detectable as well. TheKrakatau eruption would be predicted to be,2.4 Mmol, which would result in an averageenhancement of global mercury deposition of only1 mg m22 yr21. It must be noted, however, thatwork from long peat cores has suggested that largeaperiodic volcanic eruptions may have leftdetectable signals in that archiving medium aswell (e.g., Martınez-Cortizas et al., 1999;Roos-Barraclough et al., 2002). The lack ofidentifiable volcanic signals in lacustrine sedi-ments is likely due to the “smoothing” effect that aseveral-year residence time of mercury withina lake and its watershed would exert on suchshort-duration signals.

In summary, it has been demonstrated thatHg/S ratios measured for a variety of volcanicplumes and fumaroles, when indexed to estimatesof global sulfur emissions from volcanism, yielda mean volcanic mercury flux of 0.23 Mmol(45 t), which is consistent with other estimatesand observations. Accordingly, average yearlymercury emission from volcanoes is small

relative to natural terrestrial fluxes to theatmosphere (5 Mmol) and modern pollutionmercury entering the global cycle (9.6 Mmol).Over the 100 yr time period used in the MFMsimulation, anthropogenic mercury inputs to theglobal atmosphere were 1,000 Mmol, whileaverage mercury emissions from volcanoeswould be 23 Mmol. Periodic large eruptions,such as Tambora, Krakatau, Mt. St. Helens, andMt. Pinatubo, would add significantly to this fluxbut for very short periods.

9.04.3.2 Mercury Input to the Oceans viaSubmarine Volcanism

The few attempts to determine mercuryconcentrations in oceanic hydrothermal fluidshave not been successful. As a consequence, wemust estimate the fluxes of mercury associatedwith submarine volcanic activity indirectly. First,we assume that the Hg/S of 5 £ 1026 establishedfor subaerial volcanism (Table 4) can be appliedto the hydrothermal inputs associated withsubmarine tectonic activity, and second, thatoceanic lava production is ,5 £ as large as theamounts formed subaerially (e.g., Crisp, 1984).Accordingly, the submarine inputs would be,1.3 Mmol annually. An upper estimate of1.8 Mmol yr21 was proposed by Fitzgerald et al.(1998), who used oceanic mercury profiles and anestimate for the rate of vertical mixing in thewater column. The agreement for theseestimates negates the extraordinarily large fluxes(20—40 £ ) suggested by heat flow calculations(e.g., Rasmussen, 1994), and provides additionalsupport for an average Hg/S ratio for volcanicemissions of ,5 £ 1026. An input of1–2 Mmol yr21 is significant and comparable toworldwide river input as estimated by MFM(1994). It is probable, however, that only a smallfraction of the tectonically associated marinemercury inputs mixes with bulk ocean water. It islikely that mercury will be removed from thehydrothermal fluids through co-precipitation ofmetal sulfides and scavenging by precipitatinghydrous oxides of manganese and iron ashydrothermal fluids mix with the oxygenatedseawater near their entry points. Moreover,elevated levels of mercury are present in thehydrothermally derived metal-rich deposits foundon the East Pacific Rise (Bostrom and Fisher,1969) and the Gorda Ridge (Koski et al., 1994).Stoffers et al. (1999) observed some nascentelemental mercury around the sulfide chimneysof the White Island (New Zealand) complex, butgiven the scaling arguments above it wouldappear that submarine volcanism does notrepresent a significant source of mercury to theglobal ocean.

Sources of Mercury to the Environment 117

Page 12: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

9.04.3.3 Low-temperature Volatilization

As noted in the previous section, direct low-temperature weathering inputs from mineralizedmercury deposits to aquatic environments occurprimarily through sediment transport of cinnabar-containing material. Volatilization is an additionalform of low-temperature weathering in whichmercury is unparalleled by other metals. Thevolatility of elemental mercury is well documen-ted, and to the extent that mercury-containingmaterials possess some fraction of their burden inthe elemental form, weathering by volatilizationwill occur. Other mercury species are somewhatvolatile as well (Table 7), but most are less so byorders of magnitude than elemental mercury.Volatilization of mercury from soils and rock tothe atmosphere has only recently received signifi-cant attention. Unlike air–water gas exchange,air–soil/rock gas exchange has not been describedin theoretical terms; thus, all that is known ofmercury volatilization is from direct flux measure-ments and the results of soil manipulationexperiments. Several others (e.g., Poissant andCasimir, 1998 and references therein) have notedfrom measured fluxes and their temperaturedependence, the similarity of estimated andtheoretical activation energies of vaporization(,60 kJ mol21 and 85 kJ mol21, respectively).These observations hint that a more rigoroustheoretical description of this process may beforthcoming.

Measurements using flux chambers and micro-meteorological techniques are the most numerous(see special section of J. Geophys. Res. 104(D17),1999). As noted in Section 9.04.2, Klusman and

colleagues (e.g., Klusman and Jaacks, 1987)attempted to develop a tracer approach based onmeasurements of 222Rn/He/Hg to estimate the fluxof mercury from soils indexed to the fluxes of theother gases. Their work, however, has not beenextended beyond small-scale applications.Finally, isotope addition experiments includingthose of Schluter (2000) using radioactive 203Hgadditions and those of Lindberg and colleaguesusing stable isotopes in the METAALICUSprogram (see Section 9.04.8) are proving veryinsightful.

The results from some volatilization measure-ments over a number of substrates are shown inTable 8, and vary widely. In general terms, thevarious studies indicate that higher concentrationsof mercury in the soil/rock substrates lead to higherevasional fluxes. Other factors are strongly influ-ential as well. These include temperature, light,wind speed, and soil moisture (e.g., Gustin et al.,1999). It is clear that evasion of mercury frommineralized areas can be significant; however, theresults from other substrates are currently limitedby the large uncertainties and variability inherentin making such difficult measurements. Gustin andcolleagues have made efforts to scale up theirmeasurements, made primarily from Nevada, tothe western US and Mexico (10 Mg yr21; Gustinet al., 2000). Thus, this important area of researchis still developing and should be active in thefuture.

In his review of soil volatilization experiments,Schluter (2000) also highlighted the importance ofthe dissolved organic carbon concentration in thesoil fluids, with higher concentrations of fulvicacids, for instance, leading to an enhancement ofmercury reduction and evasion by generating Hg(I)and then aiding the disproportionation reaction of(2HgðIÞ ¼ Hgð0Þ þ HgðIIÞ) through sequestrationof Hg(II). The source of the reducing equivalents insoils appears to be species generated indirectlythrough photoreduction of some kind (e.g., organiccarbon and Fe(II); Schluter, 2000). The flux fromnonenriched soils, though, is substantially lowerthan that from the mineralized areas and mayaverage around 0.2 mg m22 yr21 (Schluter, 2000).

Combining estimates for volcanic and low-temperature inputs of mercury from mineralized

Table 7 Henry’s law constants for selected mercuryspecies (at STP).

Equilibrium H(M atm21)

Hg0(g) $ Hg0

(aq) 0.11Hg(OH)2(g) $ Hg(OH)2 (aq) 1.2 £ 104

HgCl2(g) $ HgCl2 (aq) 1.4 £ 106

(CH3)2Hg(g) $ (CH3)2Hg(aq) 0.13CH3HgCl(g) $ CH3HgCl(aq) 2.2 £ 103

Sources: Sanemasa, (1975), Iverfeldt and Lindqvist (1982),and Lindqvist and Rodhe (1985).

Table 8 Some examples of measured fluxes over natural soils.

Location Method(s) Soil conc.(ng g21)

Evasional flux(ng m22 h21)

References

Sweden FC NA 22 to 2 Xiao et al. (1991)Tennessee, USA FC 61–469 21.81 to 54.94 Carpi and Lindberg (1998)Quebec, Canada FC NA 0.62–8.29 Poissant and Casimir (1998)Nevada, USA FC and MM 1,200214,600 50–600 Gustin et al. (1999)Nova Scotia, Canada FC NA 21.4 to 4.3 Boudala et al. (2000)

FC ¼ flux chamber; MM ¼ micrometeorology (Bowen ratio).

Geochemistry of Mercury in the Environment118

Page 13: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

areas and non-enriched soils to the atmosphereallows an estimate of the total amount of naturalterrestrial emissions to be made. The volcanowork benefits from the existence of tracing speciessuch as sulfur that make tractable the scaling ofindividual measurements to the global scale. In thecase of low-temperature volatilization, however,no such index has yet been developed. Therefore,translating values such as those of Table 8 intoglobal fluxes is difficult.

Using the data from Nevada (Gustin et al.,2000; 0.011 Mmol yr21; 1.8 £ 1011m2 area) anemission rate of ,10 mg m22 yr21 for the globalmercury belt areas can be estimated. Furtherassuming that these enriched areas represent nomore than ,15% of the continental area suggests amaximum contribution for volatilization fromthese areas of ,5.6 Mmol yr21. The addition ofthe small volcanic contribution suggests thatnatural emissions of mercury to the atmosphereare ,5.8 Mmol yr21 and that subaerial and sub-marine emissions combined are ,7.1 Mmol yr21.

The volatilization estimates are crude extra-polations, as they are based on assumptions of soilconcentration distributions and understanding ofdriving forces behind volatilization. They dosuggest however, that the emissions measuredand estimated in some of the work cited areconsistent in the first order with Mason et al.(1994) and that natural land-based sources ofmercury to the atmosphere are consistent andlikely to be ,5 Mmol yr21. It has also been notedthat a flux 5 Mmol yr21 for natural sources isconsistent with the rate of atmospheric depositionin the pre-industrial past indicated by analysisof lake sediments (e.g., Lamborg et al., 2002b).Finally, it must be noted that emissions from soilsremoved from natural enrichments likely contain asignificant fraction of mercury initially mobilizedby anthropogenic activities that was subsequentlydeposited to soils (see the section on mixedsources below).

9.04.3.4 Anthropogenic Sources

The human-related sources of mercury to theenvironment are numerous and widespread. Mostdirect inputs of mercury from point sources toaquatic systems have largely been containedin most developed countries. Inputs of mercuryto the environment via the atmosphere are of thegreatest concern. These emissions, coupled withlong-distance transport of elemental mercury,have resulted in elevated concentrations ofmercury in fish from locations that are removedfrom anthropogenic sources (e.g., open-ocean, andsemi-remote regions in the United States, Canada,Scandinavia; Wiener et al., 2002). A summary ofthe fluxes from major sources (for 1995) is shown

in Table 9 (total of 9.6 Mmol yr21). Hightemperature processes, principally coal andmunicipal waste burning, dominate anthropogenicinputs of mercury to the atmosphere. The emissionof anthropogenic mercury is higher in the northernhemisphere, as a result of greater industrialactivity and population density. Between 1990and 1995 the emissions from developed econo-mies in North America and Europe have declinedsubstantially. Unfortunately, they have almostbeen completely replaced by emissions fromcountries, especially in Asia, that have rapidlydeveloping economies that are coal-driven.Accordingly, Asian sources of mercury currentlyconstitute 56% of all anthropogenic emissions.Based on the Pacyna and Pacyna inventory and thenatural source strength suggested by Mason et al.(1994), human activity contributes ,2/3 of themercury emitted from land-based sources eachyear. Similarly, these estimates suggest that theemission and deposition fluxes of mercury arecurrently 3 £ what they were in the pre-humanenvironment. Such estimates are now widelysupported by the reconstruction of mercurydeposition from remote lakes worldwide (morebelow; e.g., Fitzgerald et al., 1998).

9.04.3.5 Mining

Mining has been a long-standing and continu-ing source of environmental mercury contami-nation. Indeed, a partial analog to the alchemist’squest to transmute base metals into gold iscontained in the patio process in which naturallyoccurring but trace amounts of gold and silver areamalgamated (concentrated) using large amountsof liquid mercury. The dense amalgam can beseparated from the crushed, parent rock or fromplacer and alluvial deposits, often with much lossof mercury to air and aquatic systems. The gold orsilver is recovered by heating the amalgam andvaporizing the mercury. This technology has been

Table 9 Major classes of anthropogenic emissions ofmercury to the atmosphere in 1995.

Source typea

1995 flux(Mmol yr21)

Stationary combustion 7.4Non-ferrous metal production 0.8Cement production 0.7Waste disposal 0.6Pig iron and steel production 0.1

Total 9.6

After: Pacyna and Pacyna, 2002.a Stationary combustion includes fossil fuel burning power plants,while waste disposal includes municipal waste combustion.

Sources of Mercury to the Environment 119

Page 14: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

employed broadly and often crudely since itsintroduction by Bartolome de Medina in 1557(Nriagu, 1979). Historically, uses of mercury ingold and silver mining were especially significantin the Americas from the mid-1500s to the turn ofthe twentieth century. This unhealthy and ecolo-gically damaging practice continues today, and ona large scale in many countries (e.g., China,Brazil, Philippines, Kenya, and Tanzania). In theirreview of current gold mining, Lacerda andSalomons (1998) found that environmentallosses of mercury are large, 1–1.7 kg per kggold recovered. Much of the pollutant mercuryaccumulates in the surrounding lands, watersheds,waterways, and mine tailings, and the associatedenvironmental and human-health concerns areprimarily local and regional. However, there areglobal worries as well, because a portion of themercury is emitted to the atmosphere (Porcellaet al., 1997).

Mercury losses occur not only with the proces-sing and recovery of gold and silver, but in themining and production of mercury. For example,the nineteenth century “gold rush” in the westernUnited States was fueled by mercury miningin California. Egleston (1887) reports that bet-ween 1850 and 1889, the mercury yield fromCalifornia mines, especially from the NewAlmaden operation (85%), was 1,518,380 flasks(,34.5 kg/flask). This was comparable to thecombined output of the two other major mines,the Almaden (Spain) and Idrija, Austria (nowSlovenia), which produced 1,291,636 and 347,586flasks respectively over the same time period.Moreover, and as Egleston emphasizes, “accord-ing to the best California authorities, the loss in thebest constructed furnaces as near as can beapproximated is not less than 15–20%, and inmany of the works the losses will probablyamount to double that.” Mercury mining activitiescontinue today in Spain though with a highersensitivity to preventing mercury releases to theenvironment. Despite this, mercury mines remainsignificant sources of mercury to watersheds andcoastal marine systems including inoperative sitessuch as Idrija and Clear Lake, CA that supplymercury from abandoned tailings.

9.04.3.6 Biomass Burning, Soil and OceanicEvasion—Mixed Sources

There are three prominent processes that releasemercury of mixed natural and anthropogenicorigin to the atmosphere. These three includebiomass burning (deliberate and natural) and theevasion of mercury from soils and the ocean. Thegeneral factors controlling emission of mercuryfrom soils have been discussed in the section onlow-temperature volatilization. The mercury

released from soils that are not naturally enriched(unlike some of the mineralized substratesdescribed) is mercury derived principally fromatmospheric deposition and is released from theupper horizon pool (Schluter, 2000 and referencestherein). As the mercury that is deposited from theatmosphere is of mixed origin, so is the mercuryemitted to the atmosphere from these soils.Therefore, though mercury may be released fromcompletely natural and undisturbed soils asregulated by ambient biogeochemical processes,the current flux is not entirely natural. In the caseof non-enriched soils, this is not very significant asthese materials are net sinks of atmosphericmercury deposition (see Section 9.04.6). In thecase of biomass burning and especially oceanicevasion, however, the fluxes to the atmospheremay be very important in the global mercurycycle. As with soils though, these processesmobilize both natural and anthropogenic mercuryand represent sources of mixed origin. In this way,these media act to recycle mercury in theenvironment, extending the residence time ofmercury at the Earth’s surface.

Recent measurements of mercury in biomassburning plumes from research aircraft suggest thatthis process releases substantial amounts ofmercury. Brunke et al. (2001) and Friedli et al.(2001) used CO and CO2 as indexing species toestablish fluxes of mercury of ,1–5 Mmol yr21.A first-order estimate, based on the relativestrengths of truly anthropogenic emissions andtruly natural emissions, suggests that some two-thirds of the mercury released by biomassburning was initially released by human activi-ties (i.e., 0.67–3.4 Mmol yr21 anthropogenic;0.33–1.6 Mmol yr21 natural).

Oceanic evasion is a major component of themercury cycle. As described in greater detail inSection 9.04.5 below, there are a number ofprocesses that may lead to evasion of elementalmercury from the ocean. Mason et al. (1994)found that evasion from the ocean had tripled inmagnitude in concert with the increase in anthro-pogenic activities. Therefore, as with biomassburning, nearly two-thirds of the mercury cur-rently evading from the ocean is anthropogenic.

9.04.3.7 Watersheds and Legacy Mercury

Watersheds are sources of mercury to theaquatic environment. However, and similar tobiomass burning and evasion, the mercuryreleased from watersheds is of mixed origin.Because the residence time of mercury withinwatersheds is fairly long (see Section 9.04.6), thepotential for the buildup of “legacy” mercuryexists. This feature is relevant when consideringhow rapidly a system might respond to decreased

Geochemistry of Mercury in the Environment120

Page 15: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

mercury loadings. Therefore, though decreases inmercury deposition to a watershed may occur, thewatershed will contribute more mercury to itsreceiving waters than enters it each year. Legacymercury, however, was released to the environ-ment as a result of human activity and should beviewed as an anthropogenic source term.

9.04.4 ATMOSPHERIC CYCLING ANDCHEMISTRY OF MERCURY

Mercury is found in the atmosphere in both gasand particle phases. Greater than 95% of mercuryexists as gas-phase elemental mercury (Fitzgeraldand Gill, 1979; Bloom and Fitzgerald, 1988;Iverfeldt, 1991a). There is growing evidence tosuggest that some chemical form of Hg2þ alsoexists in the gas phase (the so-called “reactivegaseous mercury” or RGM; Stratton andLindberg, 1995; Sheu and Mason, 2001; Landiset al., 2002). Concentrations of total gaseousmercury (TGM; including elemental, ionic andgaseous alkylated forms such as dimethylHg) inremote areas are typically in the range of 1–2 ng(as mercury) m23. Concentrations below 1 ng m23

are to be found under certain conditions (morebelow) and higher values are often observed inurban/suburban locations. Some selected concen-tration and deposition data are shown in Table 10.Particle-phase atmospheric mercury appears to belargely Hg2þ and comprises a few percent of totalatmospheric mercury in the troposphere (Iverfeldt,1991a; Fitzgerald et al., 1991). There is only onepublished report of mercury in the stratosphere toour knowledge and no measurements in otherregions of the upper atmosphere (Murphy et al.,1998). Not surprisingly, the authors found theconcentration of particulate mercury increasedabove the tropopause as a result of enhancedoxidation of elemental mercury by ozone and thecondensation of the less volatile Hg2þ ontoambient particles.

The vertical profile of mercury in the tropo-sphere has been determined in a few cases (Banicet al., 2003; Landis and Stevens, 2001). In mostsituations, it appears that there is little change inthe mercury mixing ratio with altitude, indicatingthorough vertical mixing and an atmosphericresidence time that is long enough to make thispossible. There have been some suggestions thatelemental mercury decreases with altitude, whilereactive gaseous mercury increases, creating agradient for atmospheric deposition on largescales of the more soluble and surface activeHg2þ. For example, Landis and Stevens (2001)have suggested that the vertical RGM gradient ison the order of 400 pg m23 over the troposphere(6,340 m, isobaric). This is a rather large gradient,representing ,25% of the total atmosphericmercury. Applying a vertical eddy diffusivity of1 m2 s21 (Seinfeld, 1986) to this gradient providesan estimate of the potential rate of removal ofRGM on a large scale of 2 mg m22 yr21 (or,10–30% of the observed flux at most locations).It must be noted that these datasets are only nowbeing developed and therefore the flux estimatemade above is highly speculative.

The situation of horizontal profiling is betterdeveloped. Figure 4 illustrates data from a numberof sampling locations worldwide, showing a smallbut discernible interhemispheric gradient in TGM.Values for TGM in the northern hemisphere (NH;,1.7 ng m23) are larger than in the south(,1.2 ng m23) as a result of the NH representing agreater proportion of land-based natural and anthro-pogenic emissions. Horizontal gradients on smallerscales (i.e., plumes) have also been observedincluding continental-scale, urban plumes, andsingle industries (Fitzgerald, 1995 and the refer-ences therein; Lamborg et al., 2002a).

Assuming little vertical variation in the mixingratio of total mercury in the troposphere and usingthe available horizontal surface-based measure-ments, Mason et al. (1994) estimated the total

Table 10 Atmospheric deposition concentration data.

Location TGM(ng m23)

HgT in precip.(ng L21)

Deposition(mg m22 yr21)

Calculatedlifetime

(yr)

Reference(s)

Florida, USA 1.4–3.1 13–23 15–28 0.3–0.7 Guentzel et al. (1995) andGill et al. (1995)

Tennessee, USA 5.8 ^ 3.6 3 30 1.2 Lindberg et al. (1992)Michigan, USA 2.0 10 9 ^ 3 1.4 Hoyer et al. (1995)S. Atlantic Ocean 1.4 4 6 1.5 Lamborg et al. (1999)Wisconsin, USA 1.6 ^ 0.4 6 7 1.5 Lamborg et al. (1995)Alert, CAN 1.2 ,15 5 1.5 Schroeder et al. (1998) and

Schroeder, pers. comm.Global average 1.6 NA 5.6 1.8 Lamborg et al. (2002a)Eq. Pacific 1.3 3 4 2 Mason and Fitzgerald (1993)Sweden 2.9 ^ 0.7 10 13 ^ 12 2 ^ 1 Lindqvist et al. (1991)

After Lamborg et al. (2001).

Atmospheric Cycling and Chemistry of Mercury 121

Page 16: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

atmospheric burden of mercury to be 25 Mmol(5 kt; Figure 2(a)). Accordingly, the averagetropospheric air column mercury burden is,10 mg m22. If the emissions of mercury asoutlined in Section 9.04.3 equal 20 Mmol yr21,then the average residence time of mercury in theatmosphere is ,1.25 yr. Such a residence time isrelatively long on atmospheric timescales (e.g.,the mixing time for the hemispheres is ,1.3 yr;Geller et al., 1997), and thus we should expect tofind mercury reasonably well mixed vertically andhorizontally as we do. This global-scale value iswithin the range of similar estimates made fromvarious specific locations (Table 10).

It was quickly realized that the mercury speciesto be found in greatest abundance in precipitationwas ionic mercury (e.g., Fogg and Fitzgerald,1979). Some typical values of total mercury inprecipitation are shown in Table 10. Extensivedatabases of precipitation mercury concentrationsare available from monitoring networks in the US,Canada, and Nordic countries (e.g., US MercuryDeposition Network: http://nadp.sws.uiuc.edu/mdn). The discrepancy between the dominantgas and precipitation phase species implied aprocess of oxidation of elemental mercury in theatmosphere and its subsequent scavenging asbeing a major component of the mercury cycle.Since the initial work, and partially in response to

increased governmental interest in long-rangeatmospheric transport of pollutant mercury, therehas been an extraordinary increase in research onthe atmospheric chemistry of mercury. Manymechanisms for elemental mercury oxidation inthe atmosphere have been proposed and a fewhave been studied in detail through laboratoryexperiments (e.g., Munthe, 1992; Hall et al., 1995;Tokos et al., 1998; Lin and Pehkonen, 1999;Sommar et al., 2001). These include homo-geneous gas-phase and heterogeneous-phase reac-tions occurring in cloud-water/precipitation andaerosols. The principal constraint on gas-phaseoxidation is that the overall reaction rate must besimilar to the residence time of mercury. The rateconstants for oxidation by ozone and hydroxylradical are consistent with this constraint. Inheterogeneous phase reactions, oxidation may bepartially balanced by reduction, leading to com-plex cycling of mercury species within cloud-water or aerosols that includes influences bysorbent surfaces such as soot (Pleijel and Munthe,1995). Some of the proposed mechanisms areshown in Figure 5. The ongoing challenge forthose studying atmospheric mercury is to identifywhich of these mechanisms is actually influentialin the atmosphere and under what conditions.Lamborg et al. (2002a), based on strong rainwatercorrelations between Hg and 210Pb, havesuggested that a mechanism including gas-phaseoxidation followed by particle and precipitationscavenging is the dominant overall process for theremoval of mercury from the global atmosphere.The atmospheric chemistry embedded in thetransport and deposition model of Shia et al.(1999), which included many of these reactions,estimated the atmospheric residence of mercuryto be ,1.7 yr. Their simulation highlighted theimportance of aqueous-phase reduction reactions

Figure 4 Total gaseous Hg in the atmosphere atseveral locations. Notice the discernible interhemi-spheric gradient, resulting from greater emissions ofHg to the atmosphere in the more industrializednorthern hemisphere (data from Fitzgerald(1995)—Mid-Pacific (filled circles); Slemr and Langer(1992)—Mid-Atlantic (open triangles); Lamborget al. (1999)—Mid-Atlantic (shaded triangles); Masonet al. (1998)—North Atlantic (open diamonds);Schroeder et al. (1998)—Alert, N.W.T. (open hexa-gons); Lamborg et al. (1995)—rural Wisconsin, USA(filled diamonds); Lindberg et al. (2002)—PointBarrow, Alaska, USA (filled hexagons); Ebinghauset al. (2002)—Cape Town, R.S.A. (shaded invertedtriangles) and Antarctica (filled inverted triangles);Fitzgerald (1989)—Ninety Mile Beach, NZ (open

circles); figure from Lamborg et al., 2002a).

Hg0 (g)

Hg0 (aq) Hg(II)(aq)

HgX(aq)

Gas-phase

Wet and drydepositionHg (II)(g)

O3 or OH0

O3

HO20

S(IV)

Hg-S(IV)aq)

Hg(II) (particle)

Soot and mineral

Cr, OH0,CO3

2–

Br or org.Wet and drydeposition

Precipitation and aerosol phase

Emissions

Figure 5 Summary of some of the important physicaland chemical transformations of mercury in theatmosphere figure style adapted from Shia et al.,1999; reactions from Shia et al. (1999) and Lin and

Pehkonen (1999) and the references therein).

Geochemistry of Mercury in the Environment122

Page 17: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

in controlling the atmospheric residence time ofmercury. In more urbanized environments andclose to sources, dry deposition of particulatemercury emitted from sources or forming shortlyafter emission is likely to be the dominant removalmechanism (e.g., Keeler et al., 1995; Chiaradiaand Cupelin, 2000).

There are, however, new findings from theArctic and from mid-ocean that suggest this is notthe entire story. One of the more dramaticobservations in mercury biogeochemistry inrecent years is the so-called “spring timedepletion” of mercury in high latitudes. Schroederet al. (1998) were the first to observed thisphenomenon at Alert on the northern tip of

Ellesmere Island (Canada). Total gaseous mercuryat this location shows a fairly steady value of1.5 ng m23, typical of the northern hemispheric“background.” However, at the advent of polarsunrise and lasting for several weeks, the concen-tration begins to fluctuate between the baselinevalue and near zero, with the depletion episodeslasting hours to days. There is also a highcorrelation between depletion of mercury andozone, thus forging a connection between mercuryand the chemistry and physics of Arctic hazeformation and breakdown (Figure 6; Barrie andHoff, 1985; MacDonald et al., 2000). Furtherwork by Bill Schroeder and his colleagues as wellas others in the Arctic (e.g., Lu et al., 2001;

5.0Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

20

0

–20

–40

–60

2.5

2

1.5

1.0

0.5

00 10 20 30 40 50

4.5

4.0

3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0

60

50

40

30

20

10

0001 031 061 091 122 152 182 212 243 273 303 333 364

Julian day

TG

M

Ozone

Tem

p. (

ºC)

Month

TG

M (

ng m

–3)

Ozo

ne (

ppb)

Temp

TGM

y = 0.04 x + 0.39; R2 = 0.80

Figure 6 Depletion of Hg8 in the atmosphere during Arctic spring as observed by Schroeder et al. (1998) at Alert.The traces, from top to bottom, are surface air temperature, total gaseous mercury, and ozone. The inset illustrates the

correlation between mercury and ozone during the spring depletion period (source Schroeder et al., 1998).

Atmospheric Cycling and Chemistry of Mercury 123

Page 18: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Munthe and Berg, 2001 and references therein;Lindberg et al., 2002) and Antarctic (Ebinghauset al., 2002) confirms this phenomenon asrecurrent, seasonal, and occurring in both polarregions (though perhaps more dramatically in theArctic). Working at mid-ocean, Mason et al.(2001) have observed a rapid oxidation of mercurynear the air–sea interface resulting from reactionswith sea-spray halogens. As Lindberg et al. (2002)and others have observed, the reaction in theArctic is also coincident with a buildup of reactivebromine compounds in the polar atmosphere,pointing to a possible mechanistic similaritybetween the Arctic and mid-ocean phenomena.In both cases, the generation of significantconcentrations of oxidized mercury in the gas-phase may lead to significant dry depositionalfluxes in addition to fluxes associated withprecipitation. The assessment made by Schroederet al. (1998) when describing the depletion eventsinitially was that perhaps 0.5 Mmol of mercurywas deposited to the Arctic as a result of thisphenomenon (a boundary layer of 500 m over2 £ 105 km2 containing 1.84 ng m23 emptied ofits mercury 5 times: 4.6 mg m22 yr21). This issignificant but not an enormous unanticipated sinkfor mercury on a global-scale. However, the fluxcould still be quite significant for delicate Arcticecosystems and their human inhabitants. Thisbears further scrutiny. The polar and mid-oceanworks also imply that such a mechanism should belooked for in other atmospheric environments andmay be of broad significance.

Research on pathways of mercury dry depo-sition in addition to particle phase and RGMsuggests that plants (especially trees) may take upelemental mercury from the atmosphere into theirleaves above a certain “compensation point”concentration (Hanson et al., 1995; Beneschet al., 2001; Rea et al., 2002). Elemental mercuryabsorbed in this way could then be deposited tosoils in the form of litterfall (e.g., Johnson andLindberg, 1995; Grigal et al., 2000; Lee et al.,2000; St. Louis et al., 2001). Forest foliage mayalso act as a particle interceptor, effectivelyincreasing the dry deposition velocity of particles(throughfall; Iverfeldt, 1991b). Throughfall andlitterfall studies suggest that the removal ofmercury from the atmosphere in forestsmight be as much as 3 £ more than “open-field”deposition.

Long-term monitoring datasets of mercury in thetroposphere are now being developed at severallocations. Some studies have suggested secularincreases in TGM of as much as 1.6% yr21, thoughthis has not been widely reported (Slemr andLanger, 1992; Ebinghaus et al., 2001; Baker et al.,2002). Such research efforts are being performed ina social context of decreased mercury emissionsfrom a number of countries (see Section 9.04.3).

It is therefore difficult to gauge the current directionof secular change of mercury in the atmosphere.However, as suggested, it may be that the reducedemissions from relatively uncontrolled sources inEastern Europe, for example, will be more thanmade up for by increased emissions associated withdeveloping economies such as China.

However, assuming for the moment that theatmosphere is near a steady state betweenemissions and deposition (Mason et al., 1994),the emissions estimates of Section 9.04.3 can beused to gauge the magnitude of the depositionalflux (20 Mmol yr21). This flux is not uniform,with low latitudes receiving more mercury perunit area than high latitudes and continentalregions more than oceanic areas. These generaltrends are the result of several factors. If wetdeposition tends to be more important in theremoval of mercury than dry processes, wetterregions such as the tropics can be expected to havehigher overall fluxes of mercury. This trend isevident even on regional scales, as for instancethe flux of mercury to lakes on the wet side ofNew Zealand’s South Island are much higher thanon the dry side (Lamborg et al., 2002b). Thedifference between mid-continent and mid-oceanfluxes of mercury from the atmosphere appears tobe one driven principally by transport from thecontinents, where sources to the atmosphere arestrong, a process recently traced using 222Rn(Lamborg et al., 1999).

There remains an intriguing inconsistencybetween experiments related to the mechanismsfor mercury removal. Many lab, field, and modelefforts indicate that the lifetime of mercury in theatmosphere must be 1–2 yr, but there exist anumber of plausible removal mechanisms (such asfoliar mercury uptake followed by litterfall) thatsuggest the flux from the atmosphere is moreconsistent with lifetimes that are less than 1 yr.The likely resolution of this problem lies in theobservation that majority of the Earth’s surface iscovered by areas that are not temperate or borealforests, including the open ocean and tropicalregions. The deposition to the ocean is consistentwith an atmospheric residence time in excess of1 yr, while the mercury cycling within tropicalforests is understudied.

9.04.5 AQUATIC BIOGEOCHEMISTRYOF MERCURY

A brief overview of the marine biogeochemicalcycle of mercury was presented in the introduc-tion. Here, a broader picture of the reactions andspecies-specific interactions involving mercury innatural waters appears in Figure 7. This mecha-nistic scheme, taken from Fitzgerald and Mason(1997), is derived in part from the simulation of

Geochemistry of Mercury in the Environment124

Page 19: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

the biogeochemistry of mercury in temperate lakes(Hudson et al., 1994). Hudson and colleagues(1994) as part of the successful and scientificallyinfluential Mercury in Temperate Lakes Program(MTL) conducted in northern Wisconsin, USA,developed a mercury cycling model (MCM).Comparable models have been developed forother freshwater environments such as the FloridaEverglades (Beals et al., 2002) and OnondagaLake, a highly mercury-contaminated and USEPAdesignated “Superfund” site (Gbondo-Tugbawaand Driscoll, 1998). We anticipate, as informationincreases, analogous biogeochemical MCMs willbe developed and applied for marine systems.Accordingly, we have chosen to illustrate majorfeatures of the aquatic cycling of mercury usingthe near-shore environment as a generalizedworking analogue for aquatic systems.

Although organic and inorganic ligands andorganisms differ in fresh and salty environments,much of the biogeochemical processing andmovement of mercury are expected to be similar.Representative distribution and speciation datafor mercury in natural waters are presented inTable 11. MMHg was first determined in fresh-waters by N. Bloom (Bloom, 1989; Watras et al.,1994), and is now well documented (e.g., Vertaand Matilainen, 1995; Mason and Sullivan,

1997). In 1990, Mason and Fitzgerald reportedfinding methylated mercury species, includingDMHg, in the open-ocean waters of theequatorial Pacific Ocean (Mason and Fitzgerald,1990, 1993). Their presence has been confirmedfor the Atlantic Ocean, Mediterranean Sea andestuaries (Cossa et al., 1994; Mason et al., 1998;Mason and Sullivan, 1999). Likewise, theproduction and supersaturation of Hg0 is welldocumented in fresh and salt waters since theinitial papers were published by Vandal et al.,1991 for lakes and Kim and Fitzgerald, 1986 forthe equatorial Pacific Ocean. As Figure 7 shows,the cycling of Hg0 and MMHg is intimatelylinked by their competing and critical roles inthe aquatic biogeochemistry of mercury (sub-strate hypothesis). Notice that, in general, thespeciation, transformation pathways, reactionsand processes can be connected to reactivemercury. This reactant or “substrate” should beviewed broadly to encompass labile inorganicand organically associated mercury species.Sources include atmospheric deposition/exchange, watersheds, riverine inputs, sewageand other human-related discharges.

Using Table 11 as background and Figure 7 as aguide, important features associated with mercurycycling in all natural waters but especially

Figure 7 Generalized view of mercury biogeochemistry in the aquatic environment. Prominent processes arelabeled (Hudson et al., 1994).

Aquatic Biogeochemistry of Mercury 125

Page 20: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Table 11 Mercury species concentrations in a variety of natural waters. All data in pM, except where noted.

Location Dissolved totalHg

Particulate totalHg

Dissolved reactiveHg

DissolvedMMHg

ParticulateMMHg

DissolvedDMHg

DissolvedHg0

Reference(s)

FreshwatersLake Michigan, USA 1.6 0.58 NA 0.025–0.05 0.01–0.015 NA 0.140 ^ 0.085 Mason and Sullivan

(1997)Lake Superior,USA/Canada

ca. 0.5–5a

0.04–0.43b

0.08–0.57 0.008–0.064 0.00005–3.9b

NA 0.03–0.17 Rolfhus et al. (2003)

Lake Hoare, Antarcticaa

2.7–6.8 NA 0.4–1.2 ,0.4–1.2 NA NA NA Vandal et al. (1998)Everglades 5–10

aNA 0.15–0.5

a0.25–2.5

aNA NA 0.025–0.225 Hurley et al. (1998)

Wisconsin Lakes, USA 3–6 1–2 NA 0.1–0.9 0.15–0.35 ,0.003 0.035–0.355 Watras et al. (1994)and Fitzgeraldet al. (1991)

Estuaries/coastalSan Francisco Bay, USA 0.4–174 0.3–439 NA 0–1.6 0–1.92 NA 0.043–9.8 Conaway et al. (2003)Long Island Sound, USA 1.6–13.1 ,0.1–24.1 ,0.1–7.6 0–3.3 ,0.01–2.91 NA 0.037–0.89 Vandal et al. (2002)

and Rolfhus andFitzgerald (2001)

North Sea and ScheldtEstuary

0.5–14 0.1–6b

NA 0.05–1.37 0.0009–0.0435b

NA 0.06–0.8 Baeyans andLeermakers (1998)and Leermakerset al. (2001)

Siberian Estuaries 0.7–17 0.15–9.4 NA NA NA NA NA Coquery et al. (1995)Loire and Seine Estuaries 1–6 0.42–13.3

b,0.4–2.1 NA ,0.0015–0.0296b NA ,0.05–0.454 Coquery et al. (1997)

Chesapeake Bay, USAa ,3–40 NA NA ,0.05–0.8 NA NA ,0.1 Mason et al. (1999)

Pettaquamscutt R., USA ,1–25 ,0–18 0.4–8a

,0.05–4 ,0.05–6.88 NA ,0.025–0.4 Mason et al. (1993)Brazilian Lagoons 18.5–55.2 18–230 0.18–0.43 NA NA NA NA Lacerda and

Goncalves (2001)

Open oceanMediterranean Sea 0.8–6.4

aNA ,0.2–0.97

a,0.15

aNA ,0.13–0.29 ,0.02–0.39 Cossa et al. (1997)

Eq. Pacific Oceana

NA 0.11–5.87 0.4–6.9 ,0.05–0.58 NA ,0.005–0.67 0.015–0.69 Mason and Fitzgerald(1993)

North Atlantic 2.4 ^ 1.6 0.035 ^ 0.02 0.8 ^ 0.44 1.04 ^ 1.08 NA 0.08 ^ 0.07 0.48 ^ 0.31 Mason et al. (1998)South Atlantic 2.9 ^ 1.7

a0.1 ^ 0.05 1.7 ^ 1.2

a,0.05–0.15 NA ,0.01–0.1 1.2 ^ 0.8 Mason and Sullivan

(1999)

a These samples were unfiltered. NA ¼ not available. b Units of nmol Hg gm21 of suspended material, dry weight.

Page 21: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

seawater are reemphasized and highlighted in thefollowing summary:

(i) The principal source of the toxic species,MMHg, in marine and many freshwater aquaticsystems is in situ biologically mediated conver-sion of labile reactive mercury. As discussed,sulfate-reducing bacteria (SRB) have been impli-cated as the primary synthesizers (Compeau andBartha, 1985; Winfrey and Rudd, 1990; Gilmourand Henry, 1991). The bioamplification of MMHgin the aquatic food chain yields concentrations infish that are often more than a million timesgreater than its levels in water. Many freshwatersystems also receive significant inputs of MMHgfrom their watersheds, particularly wetlands (e.g.,Hultberg et al., 1994; Hurley et al., 1995; St. Louiset al., 1996). Salt marshes can also be prolificgenerators of MMHg, but do not appear to belarger sources than coastal sediments (Langeret al., 2001). They are, however, importantnurseries for many aquatic organisms and couldrepresent locations of significant MMHg accumu-lation in the early life stages of some fish species.

(ii) Hg0 is an important species in air and water,and in situ direct reduction of labile reactivemercury by biotic (i.e., bacterial) and abiotic(i.e., photochemical) means is a principal pathwayfor its aqueous production (Amyot et al., 1994,1997; Rolfhus, 1998; Costa and Liss, 1999);biological demethylation mechanisms (seeFigure 7) yield small amounts of Hg0 (Masonet al., 1993). The mechanisms of reduction areunclear and the focus of current study. The reversereaction, oxidation of Hg0, also occurs (e.g., Amyotet al., 1997; Lalonde et al., 2001). Thus, ambientHg0 concentrations can be expected to vary inspace and time in response to changes in the forcesthat drive the reduction and oxidation reactions(e.g., bacterial activity, light, temperature, DOC,total mercury). Examples of diel and seasonalvariations in Hg0, consistent with this view, arebecoming more common in the literature (e.g.,Lindberg et al., 2000; Rolfhus and Fitzgerald,2001; Amyot et al., 2001; O’Driscoll et al., 2003;Tseng et al., 2003; Tseng et al., in press).

(iii) Reiterating, in situ Hg0 production (naturalwaters are generally supersaturated) and emis-sions to the atmosphere are major processes; theyexert a first-order (primary) control on the overallbiogeochemistry and bioavailability of mercury inaqueous systems, and the water–air fluxes must beconsidered in global/regional atmospheric andaquatic biogeochemical models of the mercurycycle; the reduction reactions (leading to aqueousemissions of Hg0) are recycling mercury derivedfrom both natural and anthropogenic sources ofmercury, and thereby, extending the lifetime ofpollutant and natural mercury in active reservoirs.

(iv) The aqueous production of Hg0 competesfor reactant (i.e., labile reactive mercury) with the

in situ biological synthesis of MMHg; thus, waterbodies with a large production of Hg0 will haveless bioavailable mercury, smaller amounts ofMMHg in biota and reduced mercury accumu-lation in the sediment (Wiener et al., 1990b; Radaand Powell, 1993; Fitzgerald et al., 1991).

(v) Given the affinity of Hg2þ for sulfur (i.e.,sulfhydryl groups) and its ability to form verystable organomercury chelates, organic complexa-tion will exert an important control on thebioavailability of mercury.

(vi) Inorganic complexation with sulfur is aprimary reaction in reducing environments. Thisreaction may occur in oxygenated waters where,for example, microenvironments develop suchthat oxygen is depleted and sulfate reduction takesplace. This is one possible explanation for thepresence of MMHg in ocean surface waters. Thereis a triad of competing ligands for free mercury inmost aqueous systems. Organic ligands competewith chloride in saltwater and with hydroxide infreshwater, while sulfur becomes especiallycompetitive as oxygen levels decline, and SRBactivity increases.

As a point of analytical and environmentalinterest, Hg0 is more readily measured in naturalwaters than MMHg. Since the in situ production ofMMHg and Hg0 is proportional to the supply ofreactive mercury, a comprehensive understandingof the aqueous Hg0 cycle and its temporal andspatial patterns may provide a means to constrainand improve predictive models for the aquatic andatmospheric biogeochemistry of mercury andMMHg in natural waters. For a sense of thepotential geochemical benefits from automatedHg0 measurements, the reader is referred to somerecent field studies of Hg0 (e.g., Lindberg et al.,2000; Amyot et al., 2001; Balcom et al., 2000).

9.04.5.1 Environmental Mercury Methylation

Given the importance of in situ synthesis ofMMHg through conversion of less toxic mercuryspecies and its prominent role in the aquaticcycling of mercury (Figure 7), aqueous mercurymethylation merits added consideration. As out-lined, bacterial mediation enhances the rates atwhich mercury, a “soft acid,” can form alkylatedspecies in aquatic environments. This extraordi-nary interaction and its potential consequenceshave provided the rationale for much environmen-tally related mercury research over the past threedecades. Indeed, the biologically mediated syn-thesis of alkylated mercury species can readilyaccount for most of the MMHg accumulating inbiota, especially large fish, in most marine andfreshwaters (Wiener et al., 1990a; Fitzgerald andWatras, 1989; Watras et al., 1994; Rolfhusand Fitzgerald, 1995; Benoit et al., 2003).

Aquatic Biogeochemistry of Mercury 127

Page 22: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

9.04.5.1.1 Near-shore regions

The near-shore environment provides a usefulbiogeochemical framework for outlining currentknowledge regarding mercury methylation inaqueous systems. Mechanistically, recent work infreshwaters and near-shore sediments has not onlypointed to SRB as methylating agents, buttransition regions between oxygenated and anoxicconditions (e.g., low oxygen/hypoxic) as theprincipal sites of MMHg production (e.g., Gilmourand Henry, 1991; Watras et al., 1994; Langer et al.,2001). While mercury methylation does occur inthe water column and is especially importantthroughout most of the oceans (i.e., pelagicregions, e.g., Topping and Davies, 1981), themajor sites for production are associated withparticles and depositional environments such aslake and coastal/estuarine sediments, wetlands,and marshes (Watras et al., 1994; Gilmour et al.,1998; Langer et al., 2001; Hammerschmidt et al.,in preparation). Microbial production of MMHg insediment is influenced by a number of environ-mental factors that affect either the activity ofmethylating organisms (i.e., SRB) or the avail-ability of inorganic mercury for methylation. Forexample, recent studies of MMHg levels in bulksurface sediment (e.g., Benoit et al., 1998a;Gilmour et al., 1998; Krabbenhoft et al., 1999;Mason and Lawrence, 1999; Hammerschmidtet al., in preparation) have shown dependencieson inorganic mercury, organic matter, and sulfide.In marine and estuarine sediments, where seawaterprovides ample sulfate, rates of sulfate reductionare influenced mostly by availability of organicmatter and temperature (Skyring, 1987). King et al.(1999, 2000, 2001) have demonstrated that the rateof mercury methylation is closely related to that ofsulfate-reduction.

Estuarine/marine systems that are highly pro-ductive or receive autochthonous inputs of organicmatter are prime locales for enhanced rates ofmercury methylation and ecosystem exposure toMMHg. However, recent studies have illustratedthat although significant mercury methylationoccurs in such environs, production of MMHg isattenuated by accumulation of sulfide, the metabolicby-product of sulfate reduction (Figure 8; Gilmourand Henry, 1991). In estuarine and marine sedi-ments, where activity of SRB is high and largelyindependent of sulfate, sulfide inhibition of mercurymethylation is clearly demonstrated by the inverserelationship with sulfate. In contrast, mercurymethylation in freshwater systems is directly relatedto sulfate, which limits SRB metabolism. Hence,maximum mercury methylation occurs in sedimentswhere organic matter and sulfate are sufficientlyhigh as to stimulate SRB metabolism, but not sohigh as to cause accumulation of sulfide thatinhibits the availability of mercury for methylation

(Gilmour and Henry, 1991, Figure 8, the “GilmourCurve”).

A mechanism by which sulfide affects methyl-ation of mercury was recently proposed by Benoitet al. (1999a,b, 2001a,b). Sulfide affects thechemistry of inorganic mercury in sediments byprecipitating it as solid mercuric sulfide andforming dissolved mercury-sulfide complexes,including HgS0, HgS2

22, and HgHS22. HgS0 is a

major dissolved mercury-sulfide complex whensulfide is less than 1025 M and charged com-plexes, mainly as HgHS2

2, are dominant at greaterlevels (Figure 9; Benoit et al., 1999a). Themechanism for uptake of inorganic mercury bymethylating bacteria is not known, though theresearch of Benoit et al. points to diffusion ofneutrally charged HgS0 through the cellularmembrane as the key factor. As a result, maximumrates of mercury methylation occur in sedimentswhere SRB activity is significant but the accumu-lation of sulfide is minimized, thereby favoringspeciation of dissolved Hg–S complexes as HgS0.

Sulfide oxidation occurs both microbially andabiotically. In coastal sediments that are not subjectto water column anoxia, burrowing animals mixthe upper few centimeters of sediment (i.e.,bioturbation), homogenizing the sedimentarysolid-phase and pore water constituents (e.g.,Gerino et al., 1998). In doing so, underlying anoxic(i.e., sulfidic) sediments are mixed with overlyingoxic sediments; thereby minimizing accumulationof sulfide via dilution and abiotic and microbially

Maximum methylation

Sulfideinhibited

Sulfatelimited

Max

0–6 –5 –4 –3 –2 –1

log [sulfate (M)]

Freshwater Estuarine Marine

% H

g ad

ded

met

hyla

ted

per

day

Figure 8 Sulfate/Sulfide controls on mercury methyl-ation in aquatic environments — the “Gilmour curve.” Atrelatively low sulfate concentrations (most freshwaters),methylation of mercury is limited by the rate of sulfatereduction. At higher sulfate concentrations (saltwaters),sulfide buildup from relatively high rates of sulfatereduction results in decreased bioavailability of mercury(figure from Langer et al. (2001); after Gilmour and

Henry (1991).

Geochemistry of Mercury in the Environment128

Page 23: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

mediated oxidation reactions. Sulfide-oxidizingbacteria (SOB) are chemolithotrophs that usesulfide as a source of energy and reducing power.Bioturbation also can stimulate SRB activity bytranslocating organic matter from surface sedi-ments to depth (Hines and Jones, 1985; Skyring,1987; Gerino et al., 1998). Hence, bioturbation ofestuarine sediments likely stimulates mercurymethylation by both enhancing SRB activity andminimizing accumulation of sulfide. Biologicallymediated reworking of coastal/estuarine sedi-ments, in general, keeps some portion of thehistoric (buried) inventory of anthropogenicmercury (“mercury pollution legacy”) active.Given that legacy mercury can be methylated andmobilized in the coastal zone (unlike in lakes), asignificant delay is likely between reductions inmodern loadings and expected declines in MMHgin the fish stock. This unfortunate expectation formarine systems must be emphasized when con-sidering the expected and observable benefits from“zero mercury use” environmental legislation andremediation efforts.

In sediments that are less bioturbated, SOBpromote mercury methylation by minimizingaccumulation of sulfide. These bacteria prolifer-ate in redox transition zones overlying SRB. Byconsuming sulfide, SOB minimize its accumu-lation, promoting speciation of mercury-sulfidecomplexes as HgS0 and facilitating uptake ofinorganic mercury by proximal SRB. The well-defined relationships between redox transitionzones, rates of mercury methylation, andMMHg distributions in salt marsh sedimentsare illustrated in Figure 10 (Langer et al.,2001). These results are consistent with thosepredicted by the hypotheses of Benoit et al.(1999a,b). Clearly, the diverse chemistry andmicrobiology of the redox transition zonemakes it an important location for MMHg

synthesis in sediments. Although SRB appear tothe principal methylators, there is also evidencefor mercury methylation by iron reducingbacteria (Gilmour, personal communication).

Demethylation in the water column and sedi-ments is receiving increasing attention. Bothabiotic (e.g., Sellers et al., 1996, 2001) and biotic(e.g., Pak and Bartha, 1998; Marvin-Dipasqualeand Oremland, 1998; Marvin-Dipasquale et al.,2000; Hintelmann et al., 2000) processes areimplicated. The result is that MMHg accumu-lation in aquatic systems represents a balancebetween methylation, bioaccumulation, and thedemethylation processes. In sediments, MMHgdecomposition is particularly important, and it ispossible that some sediments represent net sinks,rather than net sources, for MMHg in the watercolumn.

9.04.5.1.2 Open-ocean mercury cycling

In contrast to the nearshore, mercury methyl-ation in the water column is the primary source ofMMHg in the open ocean, which represents ,90%of the marine environment. As shown in Table 11,DMHg is found in seawater, but has not beenobserved in common freshwaters (Mason et al.,1993; Mason and Fitzgerald, 1993; Mason andSullivan, 1999). Indeed, MMHg is the predomi-nant alkylated species in temperate lakes, whileDMHg, and (to a lesser extent) MMHg arecommon constituents of the dissolved mercurypool in ocean waters. The unique presence ofDMHg in seawater prompted Mason andFitzgerald (1993, 1996) to propose the followingreaction sequence:

Hg2+ DMHg MMHg

Hg0

part.MMHg

where DMHg would be the principal productfrom the methylation of inorganic mercury withMMHg and Hg0 derived from the decompositionreactions. The primary source of Hg0 in aqueoussystems, however, remains the in situ directreduction of labile reactive mercury by biotic(i.e., bacterial) and abiotic (i.e., photochemical)processes (Figure 7). Decomposition is theprincipal loss term for DMHg, while particulatescavenging and decomposition are important sinksfor dissolved MMHg. This view of mercurycycling in the ocean is very speculative. DGMformation has been little studied, and the orga-nisms responsible for methylation in the openocean are not known (nor have the pathways beenelucidated).

HgS0

Hg(HS)20

HgHS2

Hg0HgS2

2–

–8

–9

–10

–11

–12

–13

–14–7 –6 –5 –4 –3

Dissolved sulfide (log M)

Hg

spec

ies

(log

M)

Figure 9 Dissolved mercury speciation in sedimentpore waters as a function of sulfide concentration. Notethat the most bioavailable form, HgS8, is the dominantchemical form at log S , , , (24.7) (source Benoit

et al., 1999a).

Aquatic Biogeochemistry of Mercury 129

Page 24: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

9.04.5.1.3 Open-ocean mercury profiles

An appreciation of the challenges of ultra-tracemetal investigations can be gained from aconsideration of mercury speciation and distribu-tional information for open ocean. A classicvertical distributional profile for mercury in thenortheast Pacific Ocean is shown in Figure 11(Gill and Bruland, 1987). As expected with abiogeochemically active element, mercury showsa nonconservative distribution, one that is notgoverned by simple mixing of ocean waters. Totalmercury concentrations range from 1.8 pM at thesurface to 0.3 pM in the upper ocean minimum.The atmospheric mercury enrichment in the near-surface waters was captured due to recent rainsand short-term stratification. The minimum isindicative of mercury removal via particulatescavenging processes that are biologicallymediated. The gradual increase in mercury atdepth may reflect regeneration processes. Today,and given the analytical improvements since the

2 2

2

4

6

8

100 50 100 150 200 250 300

0 0

2

4

6

8

0 10 2010

2

0

2

4

6

8

10

2

0

2

4

6

8

0 1 210

0 50 100 150 200 250 300

(a1)

(a2)

(b1)

(b2)

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

Dep

th in

to s

edim

ent (

cm)

Dep

th in

to s

edim

ent (

cm)

Dep

th in

to s

edim

ent (

cm)

Dep

th in

to s

edim

ent (

cm)

Methylation rate (%Hg methylated per day) Oxygen concentration (µM)

Sulfide concentration (µM)

Figure 10 Vertical profiles of methylation rates (a1, a2) and sulfide/oxygen concentrations (b1, b2) in the sedimentsof sandy (1) and muddy (2) sites in a salt marsh (Barn Island, Connecticut, USA). Maximum methylation occurs in thetop 2 cm of these sediments and is coincident with the redox transition zone. Note also that the rates are an order of

magnitude faster in the sandy sediments (source Langer et al., 2001).

0

1,000

2,000

3,000

4,0000.0 0.5 1.0 1.5 2.0

Hg (pM)

Dep

th (

m)

Figure 11 Vertical profile of total dissolved mercuryin the northeast Pacific Ocean (after Gill and Bruland(1987), from the Vertex VII cruise, station T7, August

6–10, 1987).

Geochemistry of Mercury in the Environment130

Page 25: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Gill and Bruland report, a more comprehensivesuite of mercury species can be determined, alongmore insightful biogeochemical reaction-basedexplanations for the distributional patterns inspace and time. A summary of such extensivemercury measurements appears in Table 11.

An example of recent data comes from theThird IOC Baseline Trace Metal Cruise, whichtook place during May – June of 1996 inthe equatorial and South Atlantic (Mason andSullivan, 1999). We selected vertical profiles formercury species for one station (#8 at 178 S,258 W) from among six others examined as partof the Mason and Sullivan Program. These dataare presented in Figure 12. DMHg, Hg0 (totaldissolved gaseous mercury—the small contri-bution of DMHg), and total mercury are plottedversus depth. In addition, and as a reference, thedistributions of dissolved silicon and salinity areshown. First, the results confirm the presence ofmethylated mercury species, especially DMHg,and Hg0 in oxygenated ocean waters. Here,however, and in contrast to the North Atlantic(Table 11), MMHg levels were at the detectionlevel (0.05 pM). This suggests that MMHg iseither decomposing or scavenged more rapidlythan it is formed. The prominence of Hg0 (.50%)

relative to the total mercury present (average of2.4 ^ 1.4 pM) is a most striking feature. Itsabundant presence at depth is consistent with thehypothesis outlined above where MMHg, which isproduced as DMHg decomposes. A portion of theMMHg is scavenged by particulate matter andHg0 produced as the relatively stable product(under dark conditions) of the decomposition ofthe MMHg.

Most DMHg is produced in the near surfacewaters, but, as illustrated in Figure 12, little isfound in the euphotic zone because DMHg isreadily decomposed photochemically. DMHgaccumulates in the intermediate depths (seeprofile) above 1,500 m. Below 1,500 m, smallbut significant concentrations (0.02–0.03 pM)occur, but they are considerably smaller thanvalues (0.16 ^ 0.8 pM) in the source region ofthe North Atlantic Deep Water (NADW; Masonet al., 1998). While the DMHg decline in themodestly advecting NADW (southward traveltime ,100 yr to reach the study regions of SouthAtlantic and equatorial Atlantic) is significant,decomposition rate estimates for DMHg inadvecting deep ocean waters imply that sufficientproduction must be occurring to yield measur-able concentrations (Mason and Sullivan, 1999).

0 2 4 6 0.0 0.1 0.2 32 360

1,000

2,000

3,000

4,000

5,000

0 30 60 90 120

Conc. (pM)

Station 8 (17º S, 25º W)

Dimethyl Hg (pM) Salinity

Diss. gaseous HgTotal Hg Silica (µM)

Dep

th (

m)

Figure 12 Vertical profiles of total dissolved mercury and mercury species in the South Atlantic, from station 8 ofthe IOC Baseline South Atlantic Cruise. MonomethylHg was below detection at all depths, while measurabledimethylHg was found within and below the thermocline. Dissolved gaseous mercury, dominated by Hg8, represents

the majority of mercury at many depths at this location (source Mason and Sullivan, 1999).

Aquatic Biogeochemistry of Mercury 131

Page 26: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

This production is presumably fueled by the verysmall transport of carbon from the surface regionsto below 1,500 m.

In summary, mercury speciation studies inoceanic systems are revealing the complexinteractions of this biologically active and reactiveelement. However, and as emphasized, the spatialand temporal coverage is sparse and many ofthe biogeochemical insights and hypothesesattempting to explain the behavior and fate ofin-marine systems are speculative. Indeed, thechallenging and complicated marine biogeochem-istry of mercury beckons the curious and innova-tive. There is much to discover.

9.04.6 REMOVAL OF MERCURY FROMTHE SURFICIAL CYCLE

The work summarized in Section 9.04.3notwithstanding, most studies indicate that soilsand terrestrial sediments act as net sinks ofmercury on timescales of centuries. Such astatement is supported by a review of the mercurycontent of soils from around the world andsupporting data that allow determinations ofmercury inventories in various soil horizons(Table 12). For comparison, we have estimatedthe “excess mercury” inventories resulting fromanthropogenic activities using the integrated

Table 12 Concentrations of mercury in soils and terrestrial sediments and estimates of the anthropogenic mercuryinventory (xsHg).

Location(s) Soil/Sed. Hg Conc.(ng g21)

Soil/Sed.Hg Inventory

(mg m22)

References

SoilsCedar Creek, MN, USAOrganic layer

140 ^ 30 0.3 ^ 0.2 Grigal et al. (1994)

0–10 cm 36 ^ 7 3.4 ^ 0.510–50 cm 11 ^ 4 7 ^ 4

Fr. Guiana (rain forest)0–30 cm

122–318 Roulet and Lucotte (1995)

Sweden O Horizon 320 ^ 10 ,1.5 Alriksson (2001)B Horizon 43 ^ 3C Horizon 13 ^ 1Nevada, USA 100–15,000 Gustin et al. (1999)

Lake sedimentsNorthern Quebec, Canada 25–450 Lucotte et al. (1995)N.S. and N.Z. lakes 10–300 0.32–0.95

aLamborg et al. (2002b)

Ice coresFremont glacier, USA 0.002–0.035 ,1

aSchuster et al. (2002)

Model estimateEst. avg. anthrop. signal 0.6–2.6

athis work; Mason et al. (1994)

Coastal sedimentsLong Island Sound, USA ,30–.600 ,30–170

aVarekamp et al. (2000)

Northern Adriatic Sea 20–230 ,13a

Fabbri et al. (2001)Chesapeake Bay, USA 60–.1,000 NA Mason et al. (1999)Gulf of Cadiz, Spain ,50–250 ,75

aCossa et al. (2001)

Gulf of Trieste,Adriatic Sea

100–23,300 ,12,000a

Covelli et al. (2001)

Florida Bay, USA ,10–236 ,1.6a

Kang et al. (2000)San Francisco Bay, USA 20–700 (MMHg: 0–3.5) NA Conaway et al. (2003)S. Baltic Sea 2–340 1.2

aPempkowiak et al. (1998)

S. China Sea, Malaysia 20–127 NA Kannan and Falandysz (1998)Anadyr est., Bering Sea,Russia

77–2,100 NA Kannan and Falandysz (1998)

Mouth of St. Lawrence est.,Canada

,50–100 ,4.5a

Gobeil and Cossa (1993)

Greenland ,6–275 NA Asmund and Nielsen (2000)Santa Barbara Basin, USA 60–160 NA Young et al. (1973)

Pelagic sedimentsArctic Ocean 10–116 NA Gobeil et al. (1999)Kara Sea 75–2045 NA Seigel et al. (2001)W. Mediterranean ,80 NA Cossa et al. (1997)Laptev Sea 25–140 NA Cossa et al. (1996)

a Includes just the anthropogenic, or excess Hg.

Geochemistry of Mercury in the Environment132

Page 27: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

human emission inventory over the last 100 yrestimated by Mason et al. (1994; 947 Mmol total,447 Mmol globally uniform, 500 Mmol nearsource). As the continental area of the globe is,1.5 £ 108 km2 and assuming uniform distri-bution and no soil loss, we expect to find from0.6 mg m22 in remote areas to 2.6 mg m22 in“non-remote” areas (taken to be one-third of thecontinental area). Most of the lake sediment datacited fit this range reasonably well. The soil dataalso display agreement depending on the assumeddepth of penetration of the modern mercurysignal. Matilainen et al. (2001) found thatadditions of radioactive 203Hg were quantitativelyretained in the organic humus layer of their testsoils (O-horizon), and thus we might expect littlevertical penetration of the modern signal. Thisleads to the conclusion that most of the mercurythat has been deposited to soils in the last ,100 yris still present in active biogeochemical zones inthese systems.

These estimates suggest that human-relatedcontributions to soil loadings of mercury (theexcess mercury) are a significant if not alwaysdominant contributor to the total soil columnmercury inventory. The natural weathering inputsof mercury to soils may be examined if we assumethat the residence time of mercury within soils issimilar to that of organic carbon, citing the strongassociations between mercury and humic materialsalready mentioned. Estimates of organic carbonresidence within soils is wide ranging and depen-dent on ecological setting, but may average,100 yr. This approach is supported bythe observation that agricultural fields, whichexperience net losses of organic carbon as a resultof human use, show lower mercury inventoriesthan undisturbed soils (Grigal et al., 1994). As100 yr is the approximate timescale over whichmajor human perturbation has taken place, we maysubtract the estimated excess mercury loadingfrom the total inventory to arrive at an estimate ofthe weathering input (Table 12). In most instances,the natural and excess loadings are comparable aswas the ratio of natural to human-related emissionsinventories during that period. Thus, it appears thaton millennial timescales, soils are not net sinks formercury, and release their burden to natural watersthrough runoff and erosion and to the atmosphereby volatilization and from fires.

Such bulk, integrative studies are supported byshort-term studies of mercury loadings to water-sheds and subsequent runoff in surface streams.A number of studies have indicated that 70–95%of the mercury deposited under these conditions isretained within the soils of the watersheds.Retention estimates made in this way could bebiased low due to the slow release of “legacymercury” deposited over the previous decades thatincreases in magnitude as the total watershed

inventory increases (e.g., Mierle, 1990; Aastrupet al., 1991; Swain et al., 1992; Krabbenhoft et al.,1995; Scherbatskoy et al., 1998; Lawson andMason, 2001; Kamman and Engstrom, 2002;Lamborg et al., 2002b).

The flux of mercury from the continents tothe ocean in river runoff has been estimated to be,1–5 Mmol yr21 (Mason et al., 1994; Cossa et al.,1996). While some of this material is volatilizedas outlined in Section 9.04.5, most is buried incoastal sediments (e.g., Fitzgerald et al., 2000).Deeper coastal sediments or remote, surfacelacustrine sediments are of the order of10–50 ng g21 dry weight. Typical surface sedi-ments, even from systems not receiving directinputs from industrial activities can be 10 £ thebackground value. Extreme examples, such as thenear-shore sediments of the Gulf of Trieste (NEAdriatic Sea), which receives the river dischargefrom the mercury mining and mineralized area ofIdrija, are also to be found (Table 12). Thisenrichment of surface coastal sediments is notlikely to be due to sediment diagenesis as the fluxof mass and mercury to the sediment overwhelmsupward diffusion (Gobeil and Cossa, 1993). Thus,the excess mercury associated with the enrich-ments may be used to gauge the impact of humanactivity on the coastal zone, as shown in Table 12.Many of the values estimated here are larger thanthat expected from release of mercury fromwatersheds following atmospheric deposition.This estimate was made by assuming ,25% ofthe continental deposition (250 Mmol) was deliv-ered to 10% of the ocean area (3.6 £ 107 km2)giving 4 mg m22. This difference is the result ofperiods when significant non-atmospheric point-and area-source inputs to coastal systems and theirwatersheds were in effect. Such a condition existscurrently, for example, in the case of relativelyunrestricted gold mining in Amazonia resulting inlarge documented accumulations of mercury inthe river sediments and presumably in the coastalsediments of the western equatorial Atlantic (e.g.,Lechler et al., 2000). The residence time ofmercury in this pool is difficult to estimate givenour relatively sparse data, but is likely to besimilar to the residence time of the sedimentarymaterial. For example, Herut et al. (1996) foundthat mercury contamination of the sediments ofHaifa Bay, Israel, was disappearing in a mannerconsistent with sediment remobilization, andcould have a half-life in excess of 100 yr. Thesesediments then represent a significant sink formercury on the global scale. This pool shouldalso be viewed as an environmental and publichealth concern for the future because thismaterial is concentrated in a relatively smallarea of the ocean and one that is biologicallyproductive and commercially important (seeSection 9.04.5).

Removal of Mercury from the Surficial Cycle 133

Page 28: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Finally, open-ocean sediments receive a portionof the mercury deposited to the ocean. Here, thedata are particularly sparse and only a handful ofstudies are available to guide our discussion.Concentrations of mercury in these materials canbe found in Table 12. Given the slow rate ofaccumulation and rapid recycling of material in theupper ocean, it is to be expected that little of thismercury is of anthropogenic origin, though onestudy claims to observe surface enrichmentsconsistent with pollution inputs in a relativelynearshore and heavily impacted region (Seigelet al., 2001). Gobeil et al. (1999) found evidencefor diagenetic remobilization of mercury alongwith Fe/Mn and Mercone et al. (1999) haveproposed coupled Hg/Se diagenesis and mineralformation under the slow sediment accumulationconditions present in the open ocean. They arguedthat this remobilization could result in surfaceenrichments not associated with anthropogenicinputs. We estimate the total amount of mercuryassociated with marine sediments to be6 £ 107 Mmol (,10 ng Hg g21 of sediments, 1%organic carbon content and 12 £ 1015 t organiccarbon total; Lalli and Parsons, 1993). Withinputs of mercury to open-ocean sediments on theorder of 1 Mmol yr21 (Mason et al., 1994), thesedimentary pool should be expected to representthe primary sink for mercury on million yeartimescales.

9.04.7 MODELS OF THE GLOBAL CYCLE

Secular change in the global mercury cycle as aresult of human activity is one of the major themesin this chapter and a focus of research currently.One of the most insightful research activities inpursuit of this theme has been the development ofhistorical archives of mercury change. As of early2000s, this development effort has focused onthree archives: peat bogs, lake sediments, and icecores. All are used to reconstruct historical chan-ges in the flux of mercury from the atmosphere.

Peat cores are generally collected from bogsdominated by Sphagnum moss species and fromsystems that are ombrotrophic (“rain fed”). Underthese conditions, it is assumed that the moss isonly able to receive mercury from the atmosphereas these systems are isolated from groundwaterand geologic inputs other than dust. Mercurydelivered to the moss is incorporated in the livematerial at the surface and remains associatedeven as the moss continues to grow vertically.Therefore, collection and sectioning of a peatcore, followed by dating and mercury analysis,should reveal any changes in the rate of incorpor-ation of mercury into the moss. As discussed byBenoit et al. (1998b), the source of the mercuryarchived could include mercury associated with

rain and dust as well as elemental mercury takenup from the gas phase by the moss. A number ofstudies have made successful use of the peatarchive (e.g., A. Jensen and A. Jensen, 1990;Norton et al., 1997; Benoit et al., 1998b;Martınez-Cortizas et al., 1999). There are, how-ever, other reports of variable element mobilities,including the dating isotope 210Pb (e.g., Damman,1978; Urban et al., 1990; Norton et al., 1997). Inour recent work in Nova Scotia, we foundevidence for mobility of lead but not mercury(Lamborg et al., 2002a). We therefore submit thatpeat archives, or sections of archives, that aredated with 210Pb may yield inaccurate estimates ofmercury deposition, while other dating techniquesmay result in reliable data. For example, in NovaScotia, we dated surface peat using a co-occurringmoss species Polytrichum that develops annualmarkings on the male gametophytes. The resultsof that analysis when compared to lake sedimentsand rain measurements indicated that dry depo-sition of dust or uptake of elemental mercury bythe Sphagnum may contribute some of themercury archived in peat, but that the contributionwas likely small.

Sediments collected from remote lakes havebeen the most profitable of the archives studied todate. In most cases, lake sediments are free ofdiagenetic mobility that plagues the use of peat,allowing reliable dating and more precise recon-structions (see Chapter 9.03). The lake approachdoes have drawbacks, however, as suitable lakes(small watersheds, simple morphology, littlein/outflow) are often difficult to find in a locationof interest. In the higher latitudes of the northernhemisphere, where extensive glaciation was ineffect, small seepage lakes tend to be morecommon, and thus much of this reconstructionwork has been performed in Scandinavia, Canadaand the northern US (e.g. Swain et al., 1992;Lockhart et al., 1998; Bindler et al., 2001).Suitable lakes are being investigated currently inadditional locations such as in the tropics,temperate, and subpolar southern hemisphere.Extension of lake sediment studies to thesenew regions should bolster the conclusions fromthe northern studies. The general pictureemerging from the use of lake sediments (andpeat) is one of widespread, global-scale 2–4 timesincrease in the amount of mercury delivered fromthe atmosphere since the advent of the industrialrevolution. The similarity in timing and scale ofthe increased mercury deposition across thesevaried systems leaves little doubt that the signal isreal and the result of the substantial impact thathuman activity has had on the global mercurycycle. Application of this information to modelingefforts (see below) represents an important stepforward in our understanding of how the mercurycycle operates.

Geochemistry of Mercury in the Environment134

Page 29: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Ice core archives are receiving increasingattention. In general, these archives are not asuseful as sediments due to the relatively low rateof accumulation (usable ice cores are generally tobe found in dry high latitudes and accumulate atslow rates). Furthermore, recent work by MarcAmyot and colleagues (Lalonde et al., 2002) hassuggested that mercury deposited in snow isphotoactive and may be lost from the mediafollowing reduction to elemental mercury. How-ever, and described in Section 9.04.3, Schusteret al. (2002) analyzed ice core samples from therelatively rapidly accumulating (,70 cm yr21)Fremont Glacier (Wyoming, USA) to reconstructdeposition over the last 270 yr. As indicated inTable 12, their results are consistent with predic-tions made for the inventory of anthropogenicmercury made here based on the model of Masonet al. (1994). Therefore, mountain glaciers mayrepresent an important archive for future develop-ment. We noted that this glacial core alsoapparently preserved signals associated withvolcanic eruptions that have not been found inlake sediment archives, but that these signals aredifficult to interpret as they appear not to be self-consistent. In a more conventional location for icecore studies, Boutron et al. (1998) used snowblocks to estimate changes in mercury depositionto Greenland over the last 40 yr. Their workillustrates other fundamental difficulties associ-ated with studying Hg in ice cores (small samplesizes and exceptionally low levels, ,1 pg g21).Stretching even further back, Vandal et al. (1993)used core material from Dome C in Antarctica toreconstruct mercury accumulation at that locationto before the last glacial maximum (,34 ka Bp).This unique study found evidence that increaseddeposition of mercury to Antarctica during thatlast glacial maximum was attributable toincreased evasion from the ocean and suggeststhat mercury in ice cores may be a usefulpaleoproductivity proxy.

Seabird feathers may provide a proxy foroceanic mercury secular change (Monteiro andFurness, 1997). This unique dataset includesfeathers retrieved from museum specimens ofexclusively pelagic, piscivorous seabirds such asshearwaters and petrels. While there is scatterassociated with the values, as should be expectedfrom such a natural archive, an increase of ,3 £between 1885 and 1994 can be seen in themercury concentration of the feathers. Furness,Monteiro and co-workers have extended theirwork to include studies on the movement ofmercury within living birds and the relationshipbetween prey concentrations of mercury and thoseobserved in the bird’s feathers (Monteiro et al.,1998; Monteiro and Furness, 2001), which aid inthe interpretation of the feather record. Thisapproach has so far been applied to the NE

Atlantic (Azores), and should be extended to otherocean regions.

A number of mathematical models have beendeveloped to test the consistency of a variety ofenvironmental mercury data sets as well as aid ingenerating testable hypotheses concerning thoseaspects of the global mercury cycle that are difficultto observe directly (Lantzy and MacKenzie, 1979;Millward, 1982; Mason et al., 1994; Hudson et al.,1995; Shia et al., 1999; Bergan et al., 1999; Berganand Rodhe, 2001; Lamborg et al., 2002a). Theearliest efforts were limited by the general lack ofhigh-quality data available at the time. Mason et al.(1994; MFM) therefore can be said to be the firstglobal mercury model capable of providing mean-ingful insight into the whole cycle. The findings ofMFM, including their reconstruction of the cycle inits pre-industrial and modern forms have beenillustrated (Figure 2) and discussed in Section9.04.1. As demonstrated, the atmospheric depo-sition predictions from MFM can be compared toanthropogenic mercury inventories in soils andsediments and appear to accurately predict the fluxof mercury from the atmosphere to these archives.Additional insights from MFM include an atmos-phere/ocean system that is nearly balanced,indicating the influential role of oceanic evasionin redistributing and sustaining the lifetime ofmercury at the Earth’s surface. Furthermore, MFMfound that anthropogenic activities should havetripled the current atmospheric deposition ofmercury, resulting in a tripling of the amount ofmercury entering terrestrial and marine ecosys-tems. As indicated by the lake sediment archives,this is what is found. Finally, the MFM modelestimated that only 50% of the mercury emitted tothe atmosphere as a result of human activities isable to participate in the global cycle, and that theremaining mercury is removed from the atmos-phere close to the source. Thus, the MFM treatmentdescribes the current condition of the mercurycycle as one significantly perturbed by humanactivity on local, regional, and global scales.

As with any modeling effort, MFM waspartially supported by a number of assumptionsbased on, in some cases, scant information. SinceMFM, several other modeling efforts have soughtto address particular hypotheses concerning someof these assumptions. For example, Hudson et al.(1995) included additional reservoirs to theirsimulated ocean than were used in MFM in aneffort to gauge the importance of ocean mixing inmodulating mercury increases in the environment.They also introduced a more complex emissionssource function indexed to CO2 that released lessmercury overall than in MFM, and did so in anonlinear fashion. The conclusions from this workwere quite similar to MFM but did indicate thatocean mixing should be considered an importantprocess in the global-scale biogeochemistry of

Models of the Global Cycle 135

Page 30: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

mercury. Perhaps most importantly, ocean mixingincreases the size of the oceanic pool and as aresult increases the predicted residence time ofmercury in the ocean. An increased oceanicresidence time could result in an increased overallrate of mercury methylation, particularly as muchof this time would be spent in the lower oxygen/microbially active region of the permanentoceanic thermocline. Additional consideration ofthe oceanic biogeochemistry of mercury byMason and Fitzgerald (1996) support this modelview, particularly as an explanation for localimbalances in the air–sea exchange of mercury incertain regions such as the equatorial Pacific.Hudson et al. (1995) were also the first authors todirectly compare their results to those provided bythe lake sediments and noted, among otherfindings, that emissions from gold and silvermining prior to 1900 predicted to be significantwere not to be found in the archives.

Three other recent global models (Shia et al.,1999; Bergan et al., 1999; Bergan and Rodhe,2001) focused on the atmospheric aspects of theglobal cycle and have been able to place importantconstraints on which of the several chemicalmechanisms proposed for elemental mercuryoxidation are likely to be important at this scale.Particularly noteworthy are the efforts by Berganand colleagues, who have exploited the data oninterhemispheric concentrations of elementalmercury in the air. Exchange between the twohemispheres is relatively slow (,1.3 yr) due tophysical forcings, but yet the north/south gradientof elemental mercury is relatively weak (Slemrand Langer, 1992; Fitzgerald, 1995; Lamborget al., 1999). As the emissions of mercury in thenorthern hemisphere are estimated to be largerthan in the south, Bergan and colleagues usedgradient and mixing time information to placeconstraints on the atmospheric lifetime of mer-cury. More recently, we have extended this appro-ach through inverse box modeling (Lamborg et al.,2002a), which used the gradient and lake sedimentdata explicitly to constrain both the atmosphericlifetime as well as aspects of the oceanic cycleincluding evasion, particle scavenging, and burial.This particular model (the global/regional inter-hemispheric mercury model, or GRIMM) suggeststhat evasion from the ocean is less than thatestimated in MFM and that the ocean is a net sink.

Our inverse, or data assimilation, modelunderscores the importance of developing newdata sets to improve our understanding of mercurybiogeochemistry. As mentioned here, models suchas Hudson et al., Bergan et al., and Lamborg et al.are principally driven by atmospheric and atmos-pherically related data. As these efforts are usingessentially the same data, it is not surprising thatthe conclusions from these efforts are fairlysimilar. A more rigorous test of the current view

of the global mercury cycle will require whollynew information, particularly from the ocean.

9.04.8 DEVELOPMENTS IN STUDYINGMERCURY IN THE ENVIRONMENTON A VARIETY OF SCALES

9.04.8.1 Acid Rain and Mercury Synergyin Lakes

As noted, sulfate reduction is thought to be thedriving biogeochemical process behind mercurymethylation in many ecosystems. In freshwatersthat are relatively low in sulfate, recent increases insulfate deposition associated with the acid rainphenomenon are hypothesized to result inincreased rates of sulfate reduction and sub-sequently in mercury methylation (e.g., Gilmourand Henry, 1991; see Chapter 9.10, for details onacid rain geochemistry). Thus, not only are lakesreceiving more total mercury than in the pre-industrial past, they may be methylating a greaterproportion of this load. Swain and Helwig (1989),for example, have noted that the concentration ofmercury in Minnesota fish (usually dominated byMMHg), has increased by 10 £ , while depositionof mercury has increased by only 3 £ . Recent workby Branfireun et al. (1999, 2001) have recreatedthis phenomenon through sulfate additions tomicrocosms in the Experimental Lakes Area ofOntario, and there are ongoing efforts to expandthis research to watershed/whole-lake scales(D. R. Engstrom, personal communication).

9.04.8.2 METAALICUS

METAALICUS is a project titled MercuryExperiment To Assess Atmospheric Loading InCanada and the United States. As described in thischapter, the atmosphere is the principal avenue forthe mobilization of mercury in the environmentand anthropogenic mercury emissions to theatmosphere, especially from coal combustion,have contaminated all the major reservoirs. Wenoted the strong interest and support for studiesexamining linkages between the cycling of mer-cury in the atmosphere, anthropogenic emissions,deposition to terrestrial systems, and the bioaccu-mulation of MMHg in freshwaters. We havesuggested that there has been much emphasis onfreshwater fish contamination at the expense ofmarine studies, given that the major exposureof humans to MMHg is through the consumption ofmarine fish and seafood products. Given ourconcerns as a caveat, we note that many scientistsand “stake holders” think that an unequivocalconnection between direct atmospheric mercurydeposition and the levels of MMHg in fish infreshwater or marine systems has not beenestablished. In other words, if mercury deposition

Geochemistry of Mercury in the Environment136

Page 31: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

is reduced, through, for example, controls onmercury emissions from coal-burning powerplants, would MMHg concentrations in fishdecline? The METAALICUS team of scientistsfrom the US and Canada is now attemptingempirically to address questions of linkagesbetween mercury deposition and the bioaccumula-tion of MMHg in fish. They are conducting a wholeecosystem experiment in which stable isotopes ofmercury are added to a lake, its upland watershedand adjoining wetland, and mercury in fish as wellas the food chain are examined. The project, whichbegan in 1999 and is funded through 2004, is beingconducted in the Experimental Lakes region.

The objectives of METAALICUS are as follows(taken from the METAALICUS web page:http://www.umanitoba.ca/institutes/fisheries/METAALICUS1.html):

(i) “To determine the relationship between therate of atmospheric loading of mercury to lakesand the rate of mercury accumulation in fish.

(ii) To determine the response time of mercurylevels in fish to changes in atmospheric loading ofmercury.

(iii) To determine the relative importance ofdifferent sources of mercury to fish—the uplandwatersheds, adjacent wetlands, and the lake sur-face, by the addition of three different stableisotopes of mercury.

(iv) To calibrate and parameterize a mecha-nistic model that will be used to predict reductionsin fish mercury concentrations following controlson mercury emissions.”

The use of stable isotopes of mercury to trackpathways of inputs and uptake is innovative. Theadditions will be increased over time to levelscomparable to the current mercury deposition inthe northeastern US. The work is ongoing and, as of2002, one paper has been published (Hintelmannet al., 2002). We wish to reemphasize that inproductive near-shore regions of marine ecosys-tems, the legacy of pollution derived mercury in thesurficial sediments is likely to predominate over“new mercury” as a substrate for methylation. Theintense bioturbation in coastal marine sedimentscan keep much historical mercury active, relativeto the more quiescent sediments of lakes. Unfortu-nately, the mechanistic predictions for declines infish mercury levels following controls on mercuryemissions, derived from the anticipated successfulMETAALICUS program, will not be applicable tothe marine environment.

9.04.8.3 Ambient Isotopic Studies ofMercury—Biological Fractionation ofMercury Isotopes

With improvements in ICPS-MS detectors,there is great interest in the possibility that there

may be biologically induced fractionation of thestable isotopes of mercury. As noted, there haveonly recently been meaningful measurements ofthe isotopic composition of mercury in mediaother than rocks and meteorites. This is largelydue to the difficulty in measuring isotopic ratioswith suitable accuracy for such a heavy element.However, the range of isotopic masses presentedby the mercury system (196–204) at the extremeoffers a 4% difference in mass which approachesthat of elements whose isotopic composition areroutinely used to study cycling in the environment(e.g., 14N–15N: 7%). It is theoretically possiblethat biological and physical processes couldfractionate mercury by mass.

If fractionation of mercury isotopes does occur,then their possible uses are obviously extraordi-nary. Isotopic variations would provide biogeo-chemical tracers for the mercury cycle. Forexample, the mercury incorporated into MMHgshould be isotopically lighter than the bulk mercuryof sediment pore waters. Watersheds/rivers mayhave a unique isotopic patterns, while wastewatereffluents another. Coals from different regions mayshow variations that would characterize themercury emissions, as has been documented forlead. Perhaps, pre-industrial atmospheric depo-sition will show a mercury isotopic signaturedistinct from the modern distribution. The possi-bilities are many, but evidence for biologicallyinduced fractionation is lacking. Interest wasspurred by a recent controversial report (Jackson,2001a; Hintelmann et al., 2001; Jackson, 2001b)indicating substantial fractionation of mercuryisotopes in a lacustrine food chain. There isongoing research in several laboratories (e.g.,Amouroux et al., 2003; Krupp et al., 2003), butfew publications as we go to press. Indeed, and toreiterate, as of early 2000s, to our knowledge, nostudy has been published that convincinglydemonstrates ambient fractionation of mercury.The instrumental capability is young, and improve-ments in technologies and sensitivity in this area ofresearch will be very active in the future.

9.04.8.4 Tracing Atmospheric Mercury with210Pb and Br

As noted in Section 9.04.4, an ongoing effortamong researchers studying the atmosphericchemistry of mercury is the determination ofwhich reaction mechanisms are most influential inthe field. Tracers are exceptionally powerful toolsin this regard, as they greatly simplify the processof scaling up individual measurements to a widerange of scales. They are often integrative as well,so that discrete measurements of mercury andtracers in a particular medium at a particular timecan be easily generalized to a whole year or area.

Developments in Studying Mercury in the Environment on a Variety of Scales 137

Page 32: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Two tracers have recently shown some promise inunderstanding the large-scale behavior of mercuryin the atmosphere and in complementary ways:210Pb and Br.

Lamborg et al. (1999, 2000) have demonstrateda strong correlation between Hg and 210Pb inprecipitation from samples collected in remotecontinental (Wisconsin) and mid-ocean (tropicalSouth Atlantic) locations. The relationship fromthese collections was very similar and indicatesthat 210Pb may be a useful tracer of mercury inprecipitation. The geochemistry of 210Pb is welldescribed and this tracer has been useful foratmospheric studies of numerous kinds. Thecorrelation between these two species implicates,by analogy, a homogeneous gas-phase oxidationof mercury that is first order as 210Pb is generatedfrom the homogeneous-phase first-order radio-active decay of 222Rn. As the flux of 222Rn to theatmosphere is known and constant, the currentflux of mercury to and from the atmosphere can beestimated from Hg/210Pb ratios, and the univers-ality of the ratio used to examine regionality ofmercury deposition.

Two recent findings suggest that bromine mayalso be an important tracer and clue in theatmospheric chemistry of mercury. Lindberget al. (2002) have noted that the advent of mercurydepletion events in the Arctic is coincident withthe buildup of reactive bromine compounds (suchas BrO) in the polar atmosphere. Halogencompounds are potent oxidizers of mercury,stabilizing the Hg(II) created by forming halogencomplexes. For example, BrCl is routinely used tooxidize solutions for mercury analysis in thelaboratory. It has also been noted that suchreactive compounds can be created in situ fromsea salt in aerosols in the marine boundary layer(e.g., Disselkamp et al., 1999), supporting thehypothesis contained in Mason et al. (2001) of arapid oxidation of elemental mercury over theocean. Roos-Barraclough et al. (2002) recentlypresented findings from long peat cores thatshowed a strong correlation between higherdeposition of mercury coincident with higherdeposition of bromine over relatively long time-scales (centuries). The cause of this correlation isstill unknown and may have as much to do withsources of mercury to the atmosphere as well asthe atmospheric chemistry of mercury.

9.04.8.5 Mercury and Organic MatterInteractions

Mercury, the quintessential soft metal, formsexceptionally strong associations with naturalorganic matter (e.g., Mantoura et al., 1978). Thisbehavior has been recognized as influential in anumber of aspects of mercury biogeochemistry.

For example, lakes have often been shown to havehigher mercury in higher dissolved organic carbon(DOC) waters (e.g., Vaidya et al., 2000). Thisindicates that watersheds (the principal supply ofthe DOC in lakes) can contribute significantamounts of mercury as well, and that the DOCmobilization from uplands mobilizes mercury andMMHg. In addition, higher concentrations of DOCprovide enhanced complexation capacity formercury in the water column of lakes, enhancingmercury solubility and increasing mercury resi-dence within the lake (e.g., Ravichandran et al.,1999). There are also competing effects of DOC onthe reduction of mercury within natural waters.Recently, Amyot et al. (1997), Rolfhus (1998), andCosta and Liss (1999) have indicated that DOCmay act to both enhance and inhibit mercuryreduction (and therefore mercury evasion) in bothfreshwater and salt water depending on theconcentration of organic carbon in the water. Themechanisms for these intricate redox processes arenot well known, but DOC is clearly an importantmaster variable.

The magnitude of mercury–organic inter-actions (strength of the complexes and theabundance of the complexing agents) has beenstudied through partitioning experiments of var-ious kinds. Mantoura et al. (1978), a frequentlycited report, made use of a form of size exclusionchromatography to separate free mercury fromorganically complexed mercury. Other importantwork includes that of Hintelmann et al. (1997),who employed dialysis membranes to estimateMMHg partitioning between organic and inor-ganic complexes, and the recent work of Benoitet al. (2001c), who used calibrated octanol–waterpartitioning behavior of mercury species to studymercury-binding ligands in sediment pore watersfrom the Everglades. For other trace metals,electrochemical techniques have often beenused, and one such study has been published formercury (Wu et al., 1997). Watson et al. (1999)have suggested recently that the gold strippingelectrodes used for this determination do notquantitatively release mercury, making thisapproach somewhat suspect. Lamborg et al.(2003), who developed a wet chemical analogueto the electrochemical approach, found affinityresults similar to those of Benoit et al. and the highend of Mantoura et al. from bulk water samples(most of the other studies have been performed onisolated DOC fractions). The general assumptionhas been that it is the reduced sulfur moieties(thiols) in the macromolecules of natural waterDOC that are the sites for binding, and recentspectroscopic evidence supports this (Xia et al.,1999; Hesterberg et al., 2001). However,Lamborg et al. (2003) have found ligandconcentrations, when normalized to DOC, whichsuggest mercury-binding functional group

Geochemistry of Mercury in the Environment138

Page 33: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

abundances on the order of parts per million. Thisis well below the abundance of reduced sulfur,which is generally found to be in the parts perthousand range. It therefore appears thatthe locations for mercury binding are rare inthe DOC pool but still present in 10–1,000 £excess of mercury.

In saltwater, organic matter complexation ofmercury may not compete with the more abundantchloride ion (see Section 9.04.5). In estuaries,ligand exchange was observed by Rolfhus andFitzgerald (2001) and by Tseng et al. (2001). Oneresult of the exchange of mercury from organic toinorganic complex forms is a general increase inthe reactivity of mercury within estuaries. Thechange in complexation can result in dramaticchanges in the reactivity of the mercury as a result(e.g., enhanced Hg8 production, Rolfhus andFitzgerald, 2001).

9.04.8.6 Bioreporters—A New Technique forUltra-trace Determinations of Mercury

Modern biotechnology offers the potential forthe development of highly sensitive reagents andmethods to determine mercury species such asbioreactive or bioavailable mercury and MMHg.In one of the earliest studies on the subject, Barkayand colleagues (Selifonova et al., 1993) suggestedthat mercury-specific luminescent reagents couldbe bioengineered (i.e., biologically synthesizedand amplified with molecular techniques), for usein environmental studies. They reported thesuccessful development of three “biosensors” forHg(II). These were created by combining orfusing the “the well-understood Tn21 mercuryresistance operon (mer) with promoterless lux-CDABE from the marine bacterium, Vibriofischer.” Light production via luxCDABE iscontrolled by the mer regulatory gene andpromoter, which is in turn activated by reactivemercury species. These combined reagents arecommonly referred to as “bioreporters,” where thelight can be detected photonically and quantified.Aspects of the techniques are patented (e.g.,Rosson, 1996, patent states “the lux operoncomplex comprises luxC, luxD, luxA, luxB, andluxE genes but is free of (1) a promoter for thecomplex and (2) an inducible regulatory gene forthe complex.” The Sanseverino et al. (2002) USpatent application indicates that “an exemplarybioreporter is an E. coli that has been modified torespond to mercury II as a result of incorporationof a merRop/Iux gene cassette into its genome.”

The first applications were encouraging. How-ever, the mercury “bioreporters” were semiquan-titative and not sensitive at environmentally usefullevels (i.e., pM to fM). Progress in loweringdetection limits has been swift. For example, in

1993, Selifonova and co-workers could work inthe 1 nM range, and Tescione and Belfort (1993)at 10 nM. Virta et al. (1995) reported a laboratorydetection limit of 0.1 fM for mercury using “fireflyluciferase gene as a reporter, the mer promoterfrom transposon Tn21, and Escherichia coli(E. coli) MC1061 as the host organism.” Today,the methods, in general, are quite advanced andthe sensitivity appears suitable for studies even inpristine ecosystems such as the open ocean andremote lakes (see recent reports by Kelly et al.,2001; Scott et al., 2001; Barrocas et al., 2001).

Quantification, the operational nature of “bio-availability” and the need to add a suspension ofnon-standardized genetically engineered bacteriaas an integral part of the analysis are limitations tobroad application of the “bioreporter” techniquesto environmental studies of mercury in aqueoussystems. However, variations on this bacterialrecombinant DNA technological approach arepromising. That is, mercury activated organicmolecules can be bioengineered, isolated, andpurified to provide reagents for mercury investi-gations. Wylie et al. (1991, 1992) producedmonoclonal antibodies specific for ionic mercuryand applied the immunoassay to water. Theirexperiments were conducted in the nM range,concentrations too high for investigating mercuryin most natural aquatic environments. However, itis likely that the detection limits can be loweredsignificantly, and the mercury-specific antibodyreagent technique optimized for natural waters.

9.04.9 SUMMARY

In this chapter, we have summarized some ofthe gains made in understanding the environmen-tal biogeochemistry of mercury since Gold-schmidt’s groundbreaking work. Much of thisadvancement has come since the early 1970s, andthe growth in mercury research continues atbreakneck pace. This is fortunate as there is aneed for urgency, we believe, in these endeavors.While human activity has perturbed the mercurycycle by a smaller degree than, for example, lead,the implications for continued perturbation onhuman and ecological health are enormous.

The way forward will be a fascinating andchallenging one. As we have summarized, this isbecause the biogeochemistry of mercury operatesat a variety of time and space scales and in manyenvironmental media. Due to the complexity of theprocesses and the minute quantities of materialoften encountered in the environment, future re-search will also require new hypotheses and newinstrumentation. Similarly, and as with so manyenvironmental research efforts, new collaborationsamong scientific disciplines will be required.

Summary 139

Page 34: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

ACKNOWLEDGMENTS

Foremost, we thank the editor, BarbaraSherwood-Lollar. Rob Mason and Gary Gillprovided some graphs, Todd Hinkley and JohanVarekamp provided guidance for the volcanosection. The volcano study was aided greatly byF. LeGuern, CNRS and P. Buat-Menard,University of Bordeaux, France; W. Giggenbach(deceased), DSIR, New Zealand; J. Spedding(deceased), University of Auckland, New Zealand;L.P. Greenland, USGS, C. C. Patterson (deceased)and D. Settle (retired). Chad Hammerschmidtassisted with the sections on methylation. DanEngstrom, Don Porcella, Jim Wiener, and KrisRolfhus graciously read drafts and provided manyuseful suggestions.

REFERENCES

Aastrup M., Johnson J., Bringmark E., Bringmark L., andIverfeldt A. (1991) Occurrence and transport of mercurywithin a small catchments area. Water Air Soil Pollut. 56,155–168.

Alleman L. Y., Church T. M., Veron A. J., Kim G., Hamelin B.,and Flegal A. R. (2001) Isotopic evidence of contaminantlead in the South Atlantic troposphere and surface waters.Deep-Sea Res. II 48(13), 2811–2827.

Alriksson A. (2001) Regional variability of Cd, Hg, Pb and Cconcentrations in different horizons of Swedish forest soils.Water Air Soil Pollut. Focus 1, 325–341.

Amouroux D., Donard O. F., Krupp E., Pecheyran C., BessonT., and Fitzgerald W. F. (2003) Isotopic fractionation ofmercury species in top predators fishes (tuna and whale)from world ocean: preliminary results. In Abstracts of ASLOAquatic Sciences Meeting, Salt Lake City, UT, USA, 8–14February.

Amyot M., Mierle G., Lean D. R. S., and McQueen D. J. (1994)Sunlight-induced formation of dissolved gaseous mercury inlake waters. Environ. Sci. Technol. 28(13), 2366–2371.

Amyot M., Gill G. A., and Morel F. M. M. (1997) Productionand loss of dissolved gaseous mercury in coastal seawater.Environ. Sci. Technol. 31, 3606–3611.

Amyot M., AuClair J. C., and Poissant L. (2001) In situ hightemporal resolution analysis of elemental mercury in naturalwaters. Anal. Chim. Acta 447, 153–159.

Asmund G. and Nielsen S. P. (2000) Mercury in datedGreenland marine sediments. Sci. Tot. Environ. 245,61–72.

Baeyans W. and Leermakers M. (1998) Elemental mercuryconcentrations and formation rates in the Scheldt estuary andthe North Sea. Mar. Chem. 60, 257–266.

Baker P. G. L., Brunke E. G., Slemr F., and Crouch A. M.(2002) Atmospheric mercury measurements at Cape Point,South Africa. Atmos. Environ. 36(14), 2459–2465.

Bakir F., Damluji S. F., Amin-Zaki L., Murtadha M., KhalidiA., al-Rawi N. Y., Tikriti S., Dahahir H. I., Clarkson T. W.,Smith J. C., and Doherty R. A. (1973) Methylmercurypoisoning in Iraq. Science 181(96), 230–241.

Balcom P. H., Fitzgerald W. F., Lamborg C. H., Rolfhus K. R.,Vandal G. M., Langer C. S., and Tseng C.-M. (2000)Mercury cycling in and emissions from Long Island Sound.Presented at 11th International Conference on Heavy Metalsin the Environment. Ann Arbor, MI, USA, August.

Banic C. M., Beauchamp S. T., Tordor R. J., Schroeder W. H.,Steffen A., Anlauf K. A., Wong H. K. T. (2003) Verticaldistribution of gaseous elemental mercury in Canada. J. Geo.Res. 108(D9), Art no. 4264.

Barrie L. A. and Hoff R. M. (1985) Five years of air chemistryobservations in the Canadian Arctic. Atmos. Environ. 19,1995–2010.

Barrocas P. R. G., Landing W. M., and Procter L. M. (2001) Abacterial biosensor for aquatic mercury(II) speciation andbioavailability. Abstracts 6th International Conference onMercury as a Global Pollutant, Minamata, Japan, 130p.

Beals D. I., Harris R. C., and Pollman C. (2002) Predicting fishmercury concentrations in Everglades marshes: Handlinguncertainty in the Everglades mercury cycling model(E-MCM) with a Monte Carlo approach. Abstracts of papersof the American Chemical Society 223: 200-ENVR, Part 1APR 7 2002.

Benesch J. A., Gustin M. S., Schorran D. E., Coleman J.,Johnson D. A., and Lindberg S. E. (2001) Determining therole of plants in the biogeochemical cycling of mercury onan ecosystem level. Presented at the 6th InternationalConference on Mercury as a Global Pollutant. Minamata,Japan, October.

Benoit J. M., Gilmour C. C., Mason R. P., Riedel G. S., andRiedel G. F. (1998a) Behavior of mercury in the PatuxentRiver estuary. Biogeochemistry 40, 249–265.

Benoit J. M., Fitzgerald W. F., and Damman A. W. (1998b)The biogeochemistry of an ombrotrophic bog: evaluation ofuse as an archive of atmosphere mercury deposition.Environ. Res. 78(2), 118–133.

Benoit J. M., Gilmour C. C., Mason R. P., and Heyes A.(1999a) Sulfide controls on mercury speciation and bioavail-ability to methylating bacteria in sediment pore waters.Environ. Sci. Technol. 33, 951–957.

Benoit J. M., Mason R. P., and Gilmour C. C. (1999b)Estimation of mercury-sulfide speciation in sediment porewaters using octanol-water partitioning and implications foravailability to methylating bacteria. Environ. Toxicol. Chem.18, 2138–2141.

Benoit J. M., Gilmour C. C., and Mason R. P. (2001a) Aspectsof bioavailability of mercury for methylation in pure culturesof Desulfobulbus propionicus (1pr3). Appl. Environ. Micro-biol. 67, 51–58.

Benoit J. M., Gilmour C. C., and Mason R. P. (2001b) Theinfluence of sulfide on solid-phase mercury bioavailabilityfor methylation by pure cultures of Desulfobulbus propio-nicus (1pr3). Environ. Sci. Technol. 35, 127–132.

Benoit J. M., Mason R. P., Gilmour C. C., and Aiken G. R.(2001c) Constants for mercury binding by dissolved organicmatter isolates from the Florida Everglades. Geochim.Cosmochim. Acta 65(24), 4445–4451.

Benoit J. M., Gilmour C. C., Heyes A., Mason R. P., and MillerC. L. (2003) Geochemical and biological controls overmethylmercury production and degradation in aquaticecosystems. In Biogeochemistry of Environmentally Import-ant Trace Elements. ACS Symposium Series #835 (eds.Y. Chai and O. C. Braids). American Chemical Society,Washington, DC, pp. 262–297.

Bergan T. and Rodhe H. (2001) Oxidation of elementalmercury in the atmosphere: constraints imposed by globalscale modeling. J. Atmos. Chem. 40, 191–212.

Bergan T., Gallardo L., and Rodhe H. (1999) Mercury in theglobal troposphere: a three-dimensional model study. Atmos.Environ. 33, 1575–1585.

Bindler R., Renberg I., Appleby P. G., Anderson N. J., andRose N. L. (2001) Mercury accumulation rates and spatialpatterns in lake sediments from West Greenland: a coast toice margin transect. Environ. Sci. Technol. 35(9),1736–1741.

Bloom N. (1989) Determination of picogram levels ofmethylmercury by aqueous phase ethylation, followed bycryogenic gas chromatography with cold vapor atomicfluorescence detection. Can. J. Fish. Aquat. Sci. 46(7),1131–1140.

Bloom N. S. (1992) On the chemical form of mercury in ediblefish and marine invertebrate tissue. Can. J. Fish. Aquat. Sci.49, 1010–1017.

Geochemistry of Mercury in the Environment140

Page 35: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Bloom N. S. (2000) Analysis and stability of mercuryspeciation in petroleum hydrocarbons. Fres. J. Anal.Chem. 366, 438–443.

Bloom N. S. and Fitzgerald W. F. (1988) Determination ofvolatile mercury species at the picogram level by lowtemperature gas chromatography with cold vapor atomicfluorescence detection. Anal. Chim. Acta 208, 151–161.

Bluth G. J. S., Schnetzler C. C., Krueger A. J., and Walter L. S.(1993) The contribution of explosive volcanism to globalatmospheric sulfur dioxide concentrations. Nature 366,327–329.

Bodaly R. A., Hecky R. E., and Fudge R. J. P. (1984) Increasesin fish mercury levels in lakes flooded by the Churchill RiverDiversion, northern Manitoba. Can. J. Fish. Aquat. Sci. 41,682–691.

Bostrom K. and Fisher D. E. (1969) Distribution of mercury inEast Pacific sediments. Geochim. Cosmochim. Acta 33(6),743–745.

Boudala F. S., Folkins I., Beauchamp S., Tordon R., Neima J.,and Johnson B. (2000) Mercury flux measurements over airand water in Kejimkujik National Park, Nova Scotia. WaterAir Soil Pollut. 122, 183–202.

Boutron C. F., Vandal G. M., Fitzgerald W. F., and Ferrari C. P.(1998) A forty year record of mercury in central Greenlandsnow. Geophys. Res. Lett. 25(17), 3315–3318.

Branfireun B. A., Roulet N. T., Kelly C. A., and Rudd J. W. M.(1999) In situ sulphate stimulation of mercury methylation ina boreal peatland: toward a link between acid rain andmethylmercury contamination in remote environments.Global Biogeochem. Cycles 13(3), 743–750.

Branfireun B. A., Bishop K., Roulet N. T., Granberg G., andNilsson M. (2001) Mercury cycling in boreal ecosystems:the long-term effect of acid rain constituents on peatlandpore water methylmercury concentrations. Geophys. Res.Lett. 28(7), 1227–1230.

Brunke E. G., Labuschagne C., and Slemr F. (2001) Gaseousmercury emissions from a fire in the Cape Peninsula, SouthAfrica, during January 2000. Geophys. Res. Lett. 28(8),1483–1486.

Cadle R. D. (1975) Volcanic emissions of halides and sulfurcompounds to the troposphere and stratosphere. J. Geophys.Res. 80(12), 1650–1652.

Carpi A. and Lindberg S. E. (1998) Application of a Teflondynamic flux chamber for quantifying soil mercury flux: testand results over background soil. Atmos. Environ. 32(5),873–882.

Chiaradia M. and Cupelin F. (2000) Gas-to-particle conversionof mercury, arsenic and selenium through reactions withtraffic-related compounds (Geneva)? Indications from leadisotopes. Atmos. Environ. 34(2), 327–332.

Clarkson T. W. (1997) The toxicology of mercury. Crit. Rev.Clin. Lab. Sci. 34(4), 369–403.

Compeau G. C. and Bartha R. (1985) Sulfate-reducingbacteria: principal methylators of mercury in anoxicestuarine sediment. Appl. Environ. Microbiol. 50(2),498–502.

Conaway C. H., Squire S., Mason R. P., and Flegal A. R. (2003)Mercury speciation in the San Francisco Bay estuary. Mar.Chem. 80, 199–225.

Coquery M., Cossa D., and Martin J.-M. (1995) Thedistribution of dissolved and particulate mercury in 3Siberian estuaries and adjacent arctic coastal waters. WaterAir Soil Pollut. 80(1–4), 653–664.

Coquery M., Cossa D., and Sanjuan J. (1997) Speciation andsorption of mercury in two macro-tidal estuaries. Mar.Chem. 58, 213–227.

Cossa D., Sanjuan J., and Noel J. (1994) Mercury transport inwaters of the Strait of Dover. Mar. Pollut. Bull. 28(6),385–388.

Cossa D., Coquery M., Gobeil C., and Martin J.-M. (1996)Mercury fluxes at the ocean margins. In Global and RegionalMercury Cycles: Sources, Fluxes and Mass Balances,NATO ASI: Series 2. Environment (eds. W. Baeyens,

R. Ebinghaus, and O. Vasiliev). Kluwer, Boston, vol. 21,pp. 229–248.

Cossa D., Martin J.-M., Takayanagi K., and Sanjuan J. (1997)The distribution and cycling of mercury species in thewestern Mediterranean. Deep-Sea Res. II 44(3 – 4),721–740.

Cossa D., Elbaz-Poulichet F., and Nieto J. M. (2001) Mercuryin the Tinto-Odiel estuarine system (Gulf of Cadiz, Spain):sources and dispersion. Aquat. Geochem. 7, 1–12.

Costa M. and Liss P. S. (1999) Photoreduction of mercury insea water and its possible implications for Hg-0 air–seafluxes. Mar. Chem. 68(1–2), 87–95.

Covelli S., Faganeli J., Horvat M., and Brambati A. (2001)Mercury contamination of coastal sediments as the result oflong-term cinnabar mining activity (Gulf of Trieste, northernAdriatic Sea). App. Geochem. 16, 541–558.

Cox J. A., Carnahan J., Dinunzio J., McCoy J., and Meister J.(1979) Source of mercury in new impoundments. Bull.Environ. Contamin. Toxicol. 23, 779.

Crisp J. A. (1984) Rates of magma emplacement and volcanicoutput. J. Volcanol. Geotherm. Res. 20, 177–211.

Dalziel J. A. (1992) Reactive mercury on the Scotian Shelf andin the adjacent Northwest Atlantic Ocean. Mar. Chem.37(3–4), 171–178.

Dalziel J. A. and Yeats P. A. (1985) Reactive mercury in thecentral North Atlantic Ocean. Mar. Chem. 15(4), 357–361.

Damman A. W. H. (1978) Distribution and movement ofelements in ombrotrophic peat bogs. Oikos 30, 480–495.

Davidson P. W., Palumbo D., Myers G. J., Cox C., ShamlayeC. F., Sloane-Reeves J., Cernichiari E., Wilding G. E., andClarkson T. W. (2000) Neurodevelopmental outcomes ofSeychellois children from the pilot cohort at 108 monthsfollowing prenatal exposure to methylmercury from amaternal fish diet. Environ. Res. 84(1), 1–11.

Dedeurwaerder H., Decadt G., and Baeyens W. (1982)Estimations of mercury fluxes emitted by Mount Etnavolcano. Bull. Volcanol. 45(3), 191–196.

Disselkamp R. S., Chapman E. G., Barchet W. R., Colson S. D.,and Howd C. D. (1999) BrCl production in NaBr/NaCl/HNO3/O3 solutions representative of sea-salt aerosols in themarine boundary layer. Geophys. Res. Lett. 26(4),2183–2186.

Ebinghaus R. and Kruger O. (1996) Emission and localdeposition estimates of anthropogenic mercury in NorthWestern and Central Europe. In Global and RegionalMercury Cycles: Sources, Fluxes and Mass Balances.NATO ASI: Series 2. Environment (eds. W. Baeyens,R. Ebinghaus, and O. Vasiliev). Kluwer, Boston, vol. 21,pp. 135–159.

Ebinghaus R., Turner R. R., Lacerda L. D., Vasiliev B. O.,and Salomons W. (eds.) (1998) Mercury ContaminatedSites—Characterization, Risk Assessment and RemediationSpringer, Regensburg, Germany, 548pp.

Ebinghaus R., Kock H. H., and Schmolke S. R. (2001)Measurements of atmospheric mercury with high timeresolution: recent applications in environmental researchand monitoring. Fres. J. Anal. Chem. 371(6), 806–815.

Ebinghaus R., Kock H. H., Temme C., Einax J. W., Lowe A. G.,Richter A., Burrows J. P., and Schroeder W. H. (2002)Antarctic springtime depletion of atmospheric mercury.Environ. Sci. Technol. 36(6), 1238–1244.

Egleston T. (1887) The Metallurgy of Silver, Gold, andMercury in the United States. Wiley, NY.

Engstrom D. R. and Swain E. B. (1997) Recent declines inatmospheric mercury deposition in the upper Midwest.Environ. Sci. Technol. 31(4), 960–967.

Evans R.D., Hintelmann H. and Dillon P.J. (2001) Naturalvariation in mercury isotope ratios of coals and cinnabars.Proc. 6th Int. Conf. Mercury Glob. Pollut. Minimata, Japan,Oct. 15–19.

Fabbri D., Gabbianelli G., Locatelli C., Lubrano D., TrombiniC., and Vassura I. (2001) Distribution of mercury and other

References 141

Page 36: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

heavy metals in core sediments of the northern Adriatic Sea.Water Air Soil Pollut. 129, 143–153.

Fein J. B. and Williams-Jones A. E. (1997) The role ofmercury-organic interactions in the hydrothermal transportof mercury. Econ. Geo. 92, 20–28.

Fitzgerald W. F. (1981) Volcanic mercury emissions and theglobal mercury cycle. Programs and Abstracts, Symposiumon the Role of the Ocean in Atmospheric Chemistry. IAMAP3rd Scientific Assembly, Hamburg, Germany. August 17–28, 134p.

Fitzgerald W. F. (1989) Atmospheric and oceanic cycling ofmercury. In Chemical Oceanography (eds. J. P. Riley andR. Chester). Academic Press, New York, chap. 57, vol. 10,pp. 151–186.

Fitzgerald W. F. (1995) Is mercury increasing in theatmosphere? The need for an atmospheric mercury network(AMNET). Water Air Soil Pollut. 80, 245–254.

Fitzgerald W. F. (1996) Mercury emissions from volcanos. InAbstr. 4th Int. Conf. on Mercury as a Global Pollutant.Hamburg, Germany, August, pp. 4–7.

Fitzgerald W. F. and Clarkson T. W. (1991) Mercury andmonomethylmercury: present and future concerns. Environ.Hlth. Perspect. 96, 159–166.

Fitzgerald W. F. and Gill G. A. (1979) Subnanogramdetermination of mercury by two-stage gold amalgamationapplied to atmospheric analysis. Anal. Chem. 51, 1714–1720.

Fitzgerald W. F. and Mason R. P. (1997) Biogeochemicalcycling of mercury in the marine environment. Metal IonsBiol. Sys. 34, 53–111.

Fitzgerald W. F. and Watras C. J. (1989) Mercury in surficialwaters of rural Wisconsin lakes. Sci. Tot. Environ. 87(88),223–232.

Fitzgerald W. F., Gill G. A., and Hewitt A. D. (1981) Mercury atrace atmospheric gas. Presented at the symposium on therole of the ocean in atmospheric chemistry, IAMAP 3rdScientific Assembly, Hamburg, Germany, August 17–28.

Fitzgerald W. F., Mason R. P., and Vandal G. M. (1991)Atmospheric cycling and air–water exchange of mercuryover mid-continental lacustrine regions. Water Air SoilPollut. 56, 745–767.

Fitzgerald W. F., Engstrom D. R., Mason R., and Nater E. A.(1998) The case for atmospheric mercury contamination inremote areas. Environ. Sci. Technol. 32(1), 1–7.

Fitzgerald W. F., Vandal G. M., Rolfhus K. R., Lamborg C. H.,and Langer C. S. (2000) Mercury emissions and cycling inthe coastal zone. J. Environ. Sci. 12, 92–101.

Fogg T. R. and Fitzgerald W. F. (1979) Mercury in southernNew England rains. J. Geophys. Res. 84, 6987–6988.

Friedlander G., Kennedy J. W., Macias E. S., and Miller J. M.(1981) Nuclear and Radiochemistry, 3rd edn. Wiley,New York, pp. 640–641.

Friedli H. R., Radke L. R., and Lu J. Y. (2001) Mercury insmoke from biomass fires. Geophys. Res. Lett. 28(17),3223–3226.

Friend J. P. (1973) Global sulfur cycle. In Chemistry of theLower Atmosphere (ed. S. I. Rusool). Plenum, New York,pp. 177–201.

Ganguli P. M., Mason R. P., Abu-Saba K. E., Anderson R. S.,and Flegal A. R. (2000) Mercury speciation in drainage fromthe New Idrija mercury mine, California. Environ. Sci.Technol. 34(22), 4773–4779.

Gardfeldt K., Munthe J., Stromberg D., and Lindqvist (2003) Akinetic study on the abiotic methylation of divalent mercuryin the aqueous phase. Sci. Total Environ. 304, 127–136.

Gbondo-Tugbawa S. and Driscoll C. T. (1998) Application ofthe regional mercury cycling model to predict the fate andremediation of mercury in Onondaga Lake, New York.Water Air Soil Pollut. 105, 417–426.

Geller L. S., Elkins J. W., Lobert J. M., Clarke A. D., HurstD. F., Butler J. H., and Myers R. C. (1997) TroposphericSF6: observed latitudinal distribution and trends, derivedemissions and interhemispheric exchange time. Geophys.Res. Lett. 24(6), 675–678.

George T. S. (2001) Minamata: Pollution and the Struggle forDemocracy in Postwar Japan (Harvard East Asian Mono-graphs, No 194). Harvard University Press, Cambridge, MA,385pp.

Gerino M., Aller R. C., Lee C., Cochran J. K., Aller J. Y.,Green M. A., and Hirschberg D. (1998) Comparison ofdifferent tracers and methods used to quantify bioturbationduring a spring bloom: 234-thorium, luminophores, andchlorophyll a. Estuar. Coast. Shelf Sci. 46, 531–547.

Gill G. A. and Bruland K. W. (1987) Mercury in the northeastPacific. EOS 68, 1763.

Gill G. A. and Fitzgerald W. F. (1988) Vertical mercurydistributions in the oceans. Geochim. Cosmochim. Acta 52,1719–1728.

Gill G. A., Guentzel J. L., Landing W. M., and Pollman C. D.(1995) Total gaseous mercury measurements in Florida:The FAMS project (1992–1994). Water, Air Soil Poll. 80,235–244.

Gilmour C. C. and Henry E. A. (1991) Mercury methylation inaquatic systems affected by acid deposition. Environ. Pollut.71(2–4), 131–169.

Gilmour C. C. and Riedel G. S. (1995) Measurement of Hgmethylation in sediments using high specific-activity 203Hgand ambient incubation. Water Air Soil Pollut. 80, 747–756.

Gilmour C. C., Riedel G. S., Ederington M. C., Bell J. T.,Benoit J. M., Gill G. A., and Stordal M. C. (1998)Methylmercury concentrations and production rates acrossa trophic gradient in the northern Everglades. Biogeochem-istry 40, 326–346.

Gobeil C. and Cossa D. (1993) Mercury in sediments andsediment pore water in the Laurentian Trough. Can. J. Fish.Aquat. Sci. 50, 1794–1800.

Gobeil C., MacDonald R. W., and Smith J. N. (1999) Mercuryprofiles in sediments of the Arctic Ocean basins. Environ.Sci. Technol. 33, 4194–4198.

Goldschmidt V. M. (1954) Geochemistry. Oxford UniversityPress, Oxford, pp. 274–280.

Goldwater L. J. (1972) Mercury. In A History of Quicksilver.York Press, Baltimore, 318pp.

Grandjean P., Weihe P., White R. F., Debes F., Araki S.,Yokoyama K., Murata K., Sorensen N., Dahl R., andJorgensen P. J. (1997) Cognitive deficit in 7-year-oldchildren with prenatal exposure to methylmercury. Neuro-tox. Terat. 19(6), 417–428.

Grevesse N. (1970) Solar and meteoritic abundances ofmercury. Geochim. Cosmochim. Acta. 34, 1129–1130.

Grieb T. M., Driscoll C. T., Gloss S. P., Schofield C. L., BowieG. L., and Porcella D. B. (1990) Factors affecting mercuryaccumulation in fish in the Upper Michigan peninsula.Environ. Toxicol. Chem. 9, 919–930.

Grieb T. M., Petcharuttana Y., Roy S., Bloom N., and BrownK. (2001) Mercury studies in the central Gulf of Thailand.Proc. 6th Int. Conf. Mercury Global Pollut., Minimata,Japan, Oct. 15–19.

Grigal D. F., Nater E. A., and Homann P. S. (1994) Spatialdistribution patterns of mercury in an east-central Minnesotalandscape. In Mercury Pollution: Integration and Synthesis(eds. C. Watras and J. Huckabee). Lewis Pubs., Ann Arbor,MI, chap. III. 2, pp. 305–312.

Grigal D. F., Kolka R. K., Fleck J. A., and Nater E. A. (2000)Mercury budget of an upland-peatland watershed. Biogeo-chemistry 50(1), 95–109.

Guentzel J. L., Landing W. M., Gill G. A., and Pollman C. D.(1995) Atmospheric deposition of mercury in Florida: theFAMS project (1992–1994). Water, Air, Soil Poll. 80,393–402.

Gustin M. S., Lindberg S., Marsik F., Casimir A., EbinghausR., Edwards G., Hubble-Fitzgerald C., Kemp R., Kock H.,Leonard T., London J., Majewski M., Montecinos C., OwensJ., Pilote M., Poissant L., Rasmussen P., Schaedlich F.,Schneeberger D., Schroeder W., Sommar J., Turner R., VetteA., Wallschlaeger D., Xiao Z., and Zhang H. (1999) NevadaSTORMS project: measurement of mercury emissions from

Geochemistry of Mercury in the Environment142

Page 37: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

naturally enriched surfaces. J. Geo. Res. 104(D17),21831–21844.

Gustin M. S., Lindberg S. E., Austin K., Coolbaugh M., VetteA., and Zhang H. (2000) Assessing the contribution ofnatural sources to regional atmospheric mercury budgets.Sci. Tot. Environ. 259, 61–71.

Hall B., Schager P., and Ljungstrom E. (1995) An experimen-tal-study on the rate of reaction between mercury-vapor andgaseous nitrogen-dioxide. Water Air Soil Pollut. 81(1–2),121–134.

Hall B. D., Rosenberg D. M., and Wiens A. P. (1998) Methylmercury in aquatic insects from an experimental reservoir.Can. J. Fish. Aquat. Sci. 55, 2036–2047.

Halmer M. M., Schmincke H. U., and Graf H. F. (2002) Theannual volcanic gas input into the atmosphere, in particularinto the stratosphere: a global data set for the past 100 years.J. Volcanol. Geotherm. Res. 115(3–4), 511–528.

Hammerschmidt C. R., Sandheinrich M. B., Wiener J. G., andRada R. G. (2002) Effects of dietary methylmercury onreproduction of fathead minnows. Environ. Sci. Technol.36(5), 877–883.

Hanson P. J., Lindberg S. E., Tabberer T., Owens J. G., andKim K.-H. (1995) Foliar Exchange of Mercury Vapor:evidence for a Compensation Point. Water Air Soil Pollut.80, 373–382.

Herut B., Hornung H., Kress N., and Cohen Y. (1996)Environmental relaxation in response to reduced contami-nant input: the case of mercury pollution in Haifa Bay, Israel.Mar. Pollut. Bull. 32(4), 366–373.

Hesterberg D., Chou J. W., Hutchison K. J., and Sayers D. E.(2001) Bonding of Hg(II) to reduced organic sulfur in humicacid as affected by S/Hg ratio. Environ. Sci. Technol. 35(13),2741–2745.

Hines M. E. and Jones G. E. (1985) Microbial biogeochemistryin the sediments of Great Bay, New Hampshire. Estuar.Coast. Shelf Sci. 20, 729–742.

Hinkley T. K., Lamothe P. J., Wilson S. A., Finnegan D. L., andGerlach T. M. (1999) Metal emissions from Kilauea, and asuggested revision of the estimated worldwide metal outputby quiescent degassing of volcanoes. Earth Planet. Sci. Lett.170(3), 315–325.

Hintelmann H. and Evans R. D. (1997) Application of stableisotopes in environmental tracer studies-Measurement ofmonomethylmercury (CH3Hgþ) by isotope dilution ICP-MSand detection of species transformation. Fres. J. Anal. Chem.358, 378–385.

Hintelmann H., Welbourn P. M., and Evans R. D. (1997)Measurement of complexation of methylmercury(II) com-pounds by freshwater humic substances using equilibriumdialysis. Environ. Sci. Technol. 31(2), 489–495.

Hintelmann H., Keppel-Jones K., and Evans R. D. (2000)Constants of mercury methylation and demethylation ratesin sediments and comparison of tracer and ambient mercuryavailability. Environ. Tox. Chem. 19(9), 2204–2211.

Hintelmann H., Dillon P., Evans R. D., Rudd J. W. M., andBodaly R. A. (2001) Comment: variations in the isotopecomposition of mercury in a freshwater sediment sequenceand food web. Can. J. Fish. Aquat. Sci. 58(11), 2309–2311.

Hintelmann H., Harris R., Heyes A., Hurley J. P., Kelly C. A.,Krabbenhoft D. P., Lindberg S., Rudd J. W. M., Scott K. J.,and St. Louis V. L. (2002) Reactivity and mobility of newand old mercury deposition in a boreal forest ecosystemduring the first year of the METAALICUS study. Environ.Sci. Technol. 36(23), 5034–5040.

Hoefs J. (1980) Stable Isotope Geochemistry. Springer,Heidelburg.

Hoyer M., Burke J., and Keeler G. (1995) Atmosphericsources, transport and deposition of mercury in Michigan:two years of event precipitation. Water Air Soil Pollut. 80,192–208.

Hudson R. J. M., Gherini S. A., Watras C. J., and Porcella D. B.(1994) Modeling the biogeochemical cycle of mercury inlakes: the mercury cycling model (MCM) and its application

to the MTL study lakes. In Mercury Pollution: Integrationand Synthesis (eds. C. Watras and J. Huckabee). LewisPublishers, Ann Arbor, MI, chap. V.1, pp. 473–526.

Hudson R. J. M., Gherini S. A., Fitzgerald W. F., andPorcella D. B. (1995) Perturbation of the global mercurycycle: a model-based analysis. Water Air Soil Pollut. 80,192–208.

Hultberg H., Iverfeldt A., and Lee Y.-H. (1994) Methylmer-cury input/output and accumulation in forested catchmentsand critical loads for lakes in southwestern Sweden.In Mercury Pollution: Integration and Synthesis (eds.C. Watras and J. Huckabee). Lewis Publishers, Ann Arbor,MI, chap. III. 3, pp. 313–322.

Hurley J. P., Benoit J. M., Babiarz C. L., Shafer M. M., AndrenA. W., Sullivan J. R., Hammond R., and Webb D. A. (1995)Influences of watershed characteristics on mercury levels inWisconsin rivers. Environ. Sci. Technol. 29(7), 1867–1875.

Hurley J. P., Krabbenhoft D. P., Cleckner L. B., Olson M. L.,Aiken G. R., and Rawlik P. S., Jr. (1998) System controls onthe aqueous distribution of mercury in the northern FloridaEverglades. Biogeochemistry 40, 293–310.

Iverfeldt A. (1991a) Occurrence and turnover of atmosphericmercury over the Nordic countries. Water Air Soil Pollut. 56,251–265.

Iverfeldt A. (1991b) Mercury in forest canopy throughfallwater and its relation to atmospheric deposition. Water AirSoil Pollut. 56, 553–564.

Iverfeldt A. and Lindqvist O. (1982) Distribution equilibriumof methyl mercury chloride between water and air. Atmos.Environ. 16, 2917–2925.

Jackson T. A. (2001a) Variations in the isotope composition ofmercury in a freshwater sediment sequence and food web.Can. J. Fish. Aquat. Sci. 58(1), 185–196.

Jackson T. A. (2001b) Reply: variations in the isotopecomposition of mercury in a freshwater sediment sequenceand food web. Can. J. Fish. Aquat. Sci. 58(11), 2312–2316.

Japan Public Health Association (2001) MethylmercuryPoisoning in Minamata and Niigata, Japan (eds. Okizawaand Osame). Tokyo, 154p.

Jaworowski Z., Bysiek M., and Kownacka L. (1981) Flow ofmetals into the global atmosphere. Geochim. Cosmochim.Acta 45(11), 2185–2199.

JECFA (2000) Safety Evaluation of Certain Food Additivesand Contaminants WHO Food Additives Series: 44;Methylmercury. Prep. by 53rd Joint FAO/WHO ExpertCommittee on Food Additives (JECFA), World HealthOrganization, Geneva, Switzerland. available at: http://www. inchem. org/documents/jecfa/jecmono/v44jec13.htm

Jensen A. and Jensen A. (1990) Historical deposition rates ofmercury in Scandinavia estimated by dating and measure-ment of mercury in cores of peat bogs. Water Air Soil Pollut.56, 769–777.

Johnson D. W. and Lindberg S. E. (1995) The biogeochemicalcycling of Hg in forests—alternative methods for quantify-ing total deposition and soil emission. Water Air Soil Pollut.80(1–4), 1069–1077.

Jovanovic S. and Reed G. W., Jr. (1976) Interrelations amongisotopically anomalous mercury fractions from meteoritesand possible cosmological inferences. Science 193,888–891.

Kamman N. C. and Engstrom D. R. (2002) Historical andpresent fluxes of mercury to Vermont and New Hampshirelakes inferred from 210Pb dated sediment cores. Atmos.Environ. 36, 1599–1609.

Kang W. J., Trefry J. H., Nelsen T. A., and Wanless H. R.(2000) Direct atmospheric inputs versus runoff fluxes ofmercury to the lower Everglades and Florida Bay. EnvironSci. Technol. 34(19), 4058–4063.

Kannan K. and Falandysz J. (1998) Speciation and concen-trations of mercury in certain coastal marine sediments.Water Air Soil Pollut. 103, 129–136.

Keeler G., Glinsorn G., and Pirrone N. (1995) Particulatemercury in the atmosphere—its significance, transport,

References 143

Page 38: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

transformation and sources. Water Air Soil Pollut. 80(1–4),159–168.

Kelly C. A., Rudd J. W. M., Golding G., Scott K., Holoka M.,Barkay T., and Bloom N. (2001) The use of genetically-engineered bioreporters to study mercury availability at traceconcentrations. In Abstr. 6th Int. Conf. on Mercury as aGlobal Pollutant, Minamata, Japan. 129p.

Kesler S. E. (1994) Mineral Resources, Economics and theEnvironment. MacMillan, NY.

Kim J. P. and Fitzgerald W. F. (1986) Sea-air partitioning ofmercury over the equatorial Pacific Ocean. Science 231,1131–1133.

King J. K., Saunders F. M., Lee R. F., and Jahnke R. A. (1999)Coupling mercury methylation rates to sulfate reductionrates in marine sediments. Environ. Toxicol. Chem. 18,1362–1369.

King J. K., Kostka J. E., Frischer M. E., and Saunders F. M.(2000) Sulfate-reducing bacteria methylate mercury atvariable rates in pure culture and in marine sediments.Appl. Environ. Microbiol. 66, 2430–2437.

King J. K., Kostka J. E., Frischer M. E., Saunders F. M., andJahnke R. A. (2001) A quantitative relationship thatdemonstrates mercury methylation rates in marine sedimentsare based on the community composition and activity ofsulfate-reducing bacteria. Environ. Sci. Technol. 35,2491–2496.

Klaue B., Kesler S. E., Blum J. D. (2000) Investigation of naturalfractionation of stable mercury isotopes by inductivelycoupled plasma mass spectrometry. In Proc. Int. Conf.Heavy Metals Environ. Ann Arbor, MI, August 6–10.

Klusman R. W. and Jaacks J. A. (1987) Environmentalinfluences upon mercury, radon and helium concentrationsin soil gases at a site near Denver, Colorado. J. Geochem.Explor. 27, 259–280.

Koski R. A., Benninger L. M., Zierenberg R. A., andJonasson I. R. (1994) Composition and growth history ofhydrothermal deposits in Escanaba Trough, southern GordaRidge. In Geologic, Hydrothermal, and Biologic Studies atEscanaba Trough, Gorda Ridge, Offshore NorthernCalifornia. US Geological Survey Bulletin 2022 (eds.J. L. Morton, R. A. Zierenberg, and C. A. Reiss),pp. 293–324.

Krabbenhoft D. P., Benoit J. M., Babiarz C. L., Hurley J. P.,and Andren A. W. (1995) Mercury cycling in the AllequashCreek watershed, northern Wisconsin. Water Air Soil Pollut.80, 425–433.

Krabbenhoft D. P., Wiener J. G., Brumbaugh W. G., OlsonM. L., DeWild J. F., and Sabin T. J. (1999) A national pilotstudy of mercury contamination of aquatic ecosystems alongmultiple gradients. In US Geological Survey Toxic Sub-stances Hydrology Program, Proceeding of the TechnicalMeeting, Volume 2 (eds. D. W. Morganwalp and H. T.Buxton). Contamination of Hydrologic Systems and RelatedEcosystems: USGS, Water-resources Investigations Report99-4018B. United States Geological Survey, Reston, VA,pp. 147–160.

Krupp E. M., Pecheyran C., and Donard O. F. X. (2003)Investigation on the mercury isotope distribution of mercuryspecies (MMHg and Hg2þ) in different fish tissue samples.In Abstracts of the 2003 European Winter Conference onPlasma Spectrochemistry, Garmisch – Partenkirchen,Germany, January, 12–17.

Lacerda L. D. and Goncalves G. O. (2001) Mercurydistribution and speciation in waters of the coastal lagoonsof Rio de Janeiro, SE Brazil. Mar. Chem. 76, 47–58.

Lacerda L. D. and Salomons W. (1998) Mercury from Gold andSilver Mining: a Chemical Time Bomb? Springer, New York,146pp.

Lalli C. M. and Parsons T. R. (1993) Biological Oceanogra-phy: An Introduction. Heinemann, Oxford, UK, 147p.

Lalonde J. D., Amyot M., Kraepiel A. M. L., and Morel F. M. M.(2001) Photooxidation of Hg(0) in artificial and naturalwaters. Environ. Sci. Technol. 35, 1367–1372.

Lalonde J. D., Poulain A. J., and Amyot M. (2002) The role ofmercury redox reactions in snow on snow-to-air mercurytransfer. Environ. Sci. Technol. 36(2), 174–178.

Lambert G., Le Cloaree M.-F., and Pennisi M (1988) Volcanicoutput of SO2 and trace metals: a new approach. Geochim.Cosmochim. Acta 52(1), 39–42.

Lamborg C. H., Fitzgerald W. F., Vandal G. M., and RolfhusK. R. (1995) Atmospheric mercury in northern Wisconsin:sources and species. Water Air Soil Pollut. 80, 198–206.

Lamborg C. H., Rolfhus K. R., Fitzgerald W. F., and Kim G.(1999) The atmospheric cycling and air-sea exchange ofmercury species in the south and equatorial Atlantic Ocean.Deep-Sea Res. II. 46(5), 957–977.

Lamborg C. H., Fitzgerald W. F., Graustein W. C., andTurekian K. K. (2000) An examination of the atmosphericchemistry of mercury using 210Pb and 7Be. J. Atmos. Chem.36, 325–338.

Lamborg C. H., Fitzgerald W. F., O’Donnell J., and TorgersenT. (2002a) A non-steady-state compartmental model ofglobal-scale mercury biogeochemistry with interhemi-spheric atmospheric gradients. Geochim. Cosmochim. Acta66(7), 1105–1118.

Lamborg C. H., Fitzgerald W. F., Damman A. W. H., BenoitJ. M., Balcom P. H., and Engstrom D. R. (2002b) Modernand historic atmospheric mercury fluxes in both hemi-spheres: global and regional mercury cycling implications.Glob. Biogeochem. Cycles. 16(4), 1104 doi: 10.1029/2001GB001847.

Lamborg C. H., Tseng C.-M., Fitzgerald W. F., Balcom P. H.,and Hammerschmidt C. R. (2003) Determination of themercury complexation characteristics of dissolved organicmatter in natural waters with “reducible Hg” titrations.Environ. Sci. Technol. 37, 3316–3322.

Landis M. S. and Stevens R. K. (2001) Preliminary results fromthe USEPA mercury speciation network and aircraftmeasurements campaigns. Presented at: 6th InternationalConference on Mercury as a Global Pollutant, Minamata,Japan, October.

Landis M. S., Stevens R. K., Schaedlich F., and Prestbo E.(2002) Development and characterization of an annulardenuder methodology for the measurement of divalentinorganic reactive gaseous mercury in ambient air. Environ.Sci. Technol. 36(13), 3000–3009.

Langer C. S., Fitzgerald W. F., Vandal G. M., and VisscherP. T. (2001) Biogeochemical cycling of methylmercury atBarn Island Salt Marsh, Stonington, CT, USA. WetlandsEcol. Manage. 9, 295–310.

Lantzy R. J. and MacKenzie F. T. (1979) Atmospheric tracemetals: global cycles and assessment of man’s impact.Geochim. Cosmochim. Acta 43, 511–525.

Lauretta D. S., Devouard B., and Buseck P. R. (1999) Thecosmochemical behavior of mercury. Earth Planet. Sci. Lett.171, 35–47.

Lauretta D. S., Klaue B., Blum J. D., and Buseck P. R. (2001)Mercury abundances and isotopic compositions in theMurchison (CM) and Allende (CV) carbonaceous chon-drites. Geochim. Cosmochim. Acta 65(16), 2807–2818.

Lawson N. M. and Mason R. P. (2001) Concentration ofmercury, methylmercury, cadmium, lead, arsenic, andselenium in the rain and stream water of two contrastingwatersheds in western Maryland. Water Res. 35(17),4039–4052.

Lechler P. J., Miller J. R., Lacerda L. D., Vinsond D.,Bonzongo J.-C., Lyons W. B., and Warwick J. J. (2000)Elevated mercury concentrations in soils, sediments, water,and fish of the Madeira River basin, Brazilian Amazon:a function of natural enrichments? Sci. Tot. Environ. 260,87–96.

Lee Y. H., Bishop K. H., and Munthe J. (2000) Do conceptsabout catchment cycling of methylmercury and mercury inboreal catchments stand the test of time? Six years ofatmospheric inputs and runoff export at Svartberget, north-ern Sweden. Sci. Tot. Environ. 260(1–3), 11–20.

Geochemistry of Mercury in the Environment144

Page 39: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Leermakers M., Galletti S., De Galan S., Brion N., and BaeyensW. (2001) Mercury in the southern North Sea and Scheldtestuary. Mar. Chem. 75(3), 229–248.

Lin C. J. and Pehkonen S. O. (1999) Aqueous phase reactionsof mercury with free radicals and chlorine: implications foratmospheric mercury chemistry. Chemosphere 38(6),1253–1263.

Lindberg S. E., Meyers T. P., Taylor G. E., Jr., Turner R. R.,and Schroeder W. H. (1992) Atmospheric surface exchangeof mercury in a forest: results of modeling and gradientapproaches. J. Geol. Res. 97(2), 2519–2528.

Lindberg S. E., Vette A. F., Miles C., and Schaedlich F. (2000)Mercury speciation in natural waters: measurement ofdissolved gaseous mercury with a field analyzer. Biogeo-chemistry 48, 237–259.

Lindberg S. E., Brooks S., Lin C. J., Scott K. J., Landis M. S.,Stevens R. K., Goodsite M., and Richter A. (2002) Dynamicoxidation of gaseous mercury in the Arctic troposphere atpolar sunrise. Environ. Sci. Technol. 36(6), 1245–1256.

Lindqvist O. and Rodhe H. (1985) Atmospheric mercury-areview. Tellus 27B, 136–159.

Lindqvist O., Johansson K., Aastrup M., Andersson A.,Bringmark L., Hovsenius G., Haakanson L., Iverfeldt A.,Meili M., and Timm B. (1991) Mercury in the Swedishenvironment—recent research on causes, consequences andcorrective methods. Water Air Soil Pollut. 55(1–2), 1–261.

Lockhart W. L., Wilkinson P., Billeck B. N., Danell R. A.,Hunt R. V., Brunskill G. J., Delaronde J., and St. Louis V.(1998) Fluxes of mercury to lake sediments in central andnorthern Canada inferred from dated sediment cores.Biogeochemistry 40(2–3), 163–173.

Long S. E. and Kelly W. R. (2002) Determination of mercuryin coal by isotope dilution cold-vapor generation inductivelycoupled plasma mass spectrometry. Anal. Chem. 74,1477–1483.

Lu J. Y., Schroeder W. H., Barrie L. A., Steffen A., WelchH. E., Martin K., Lockhart L., Hunt R. V., Boila G., andRichter A. (2001) Magnification of atmospheric mercurydeposition to polar regions in springtime: the link totropospheric ozone depletion chemistry. Geophys. Res.Lett. 28(17), 3219–3222.

Lucotte M., Mucci A., Hillaire-Marcel C., Pichet P., andGrondin A. (1995) Anthropogenic mercury enrichment inremote lakes of northern Quebec (Canada). WASP 80,467–476.

MacDonald R. W., Barrie L. A., Bidleman T. F., DiamondM. L., Gregor D. J., Semkin R. G., Strachan W. M. J., LiY. F., Wania F., Alaee M., Alexeeva L. B., Backus S. M.,Bailey R., Bewers J. M., Gobeil C., Halsall C. J., Harner T.,Hoff J. T., Jantunen L. M. M., Lockhart W. L., Mackay D.,Muir D. C. G., Pudykiewicz J., Reimer K. J., Smith J. N.,Stern G. A., Schroeder W. H., Wagemann R., and YunkerM. B. (2000) Contaminants in the Canadian Arctic: 5 yearsof progress in understanding sources, occurrence andpathways. Sci. Tot. Environ. 254, 93–234.

Mantoura R. F. C., Dickson A., and Riley J. P. (1978) Thecomplexation of metals with humic materials in naturalwaters. Estuar. Coast. Mar. Sci. 6, 387–408.

Martell A. E., Smith R. M., and Motekaitis R. J. (1998) NISTCritically Selected Stability Constants of Metal ComplexesData Base. NIST Stand. Ref. Database no. 46. US Dept. ofCommerce, Gaithersburg, MD.

Martınez-Cortizas A., Potevedra-Pombal X., Garcıa-Rodeja E.,Novoa Munoz J. C., and Shotyk W. (1999) Mercury in aSpanish peat bog: archive of climate change and atmosphericmetal deposition. Science 284, 939–942.

Marvin-Dipasquale M. and Oremland R. S. (1998) Bacterialmethylmercury degradation in Florida Everglades peatsediment. Environ. Sci. Technol. 32, 2556–2563.

Marvin-Dipasquale M., Agee J., McGowan C., Oremland R. S.,Thomas M., Krabbenhoft D., and Gilmour C. C. (2000)Methyl-mercury degradation pathways: a comparison

among three mercury impacted ecosystems. Environ. Sci.Technol. 34, 4908–4916.

Mason R. P. and Fitzgerald W. F. (1990) Alkylmercury speciesin the equatorial Pacific. Nature 347, 457–459.

Mason R. P. and Fitzgerald W. F. (1993) The distribution andbiogeochemical cycling of mercury in the equatorial PacificOcean. Deep-Sea Res. 40, 1897–1924.

Mason R. P. and Fitzgerald W. F. (1996) Sources, sinks andbiogeochemical cycling of mercury in the ocean. In Globaland Regional Mercury Cycles: Sources, Fluxes andMass Balances. NATO ASI: Series 2. Environment(eds. W. Baeyens, R. Ebinghaus, and O. Vasiliev). Kluwer,Boston, vol. 21, pp. 249–272.

Mason R. P. and Lawrence A. L. (1999) Concentration,distribution, and bioavailability of mercury and methylmer-cury in sediments of Baltimore Harbor and Chesapeake Bay,Maryland, USA. Environ. Toxicol. Chem. 18, 2438–2447.

Mason R. P. and Sullivan K. A. (1999) Mercury in LakeMichigan. Environ. Sci. Technol. 31, 942–947.

Mason R. P. and Sullivan K. A. (1999) The distribution andspeciation of mercury in the south and equatorial Atlantic.Deep-Sea Res. II. 46, 937–956.

Mason R. P., Fitzgerald W. F., Hurley J. P., Hanson A. K., Jr.,Donaghay P. L., and Sieburth J. M. (1993) Mercurybiogeochemical cycling in a stratified estuary. Limnol.Oceanogr. 38, 1227–1241.

Mason R. P., Fitzgerald W. F., and Morel F. M. M. (MFM)(1994) The biogeochemical cycling of elemental mercury:anthropogenic influences. Geochim. Cosmochim. Acta 58,3191–3198.

Mason R. P., Rolfhus K. R., and Fitzgerald W. F. (1998)Mercury in the North Atlantic. Mar. Chem. 61, 37–53.

Mason R. P., Lawson N. M., Lawrence A. L., Leaner J. J., LeeJ. G., and Sheu G.-R. (1999) Mercury in Chesapeake Bay.Mar. Chem. 65, 77–96.

Mason R. P., Lawson N. M., and Sheu G. R. (2001) Mercury inthe Atlantic Ocean: factors controlling air–sea exchange ofmercury and its distribution in the upper waters. Deep-SeaRes. II. 48(13), 2829–2853.

Matilainen T., Verta M., Korhonen H., Uusi-Rauva A., andNiemi M. (2001) Behavior of mercury in soil profiles:impact of increased precipitation, acidity, and fertilizationon mercury methylation. Water Air Soil Pollut. 125,105–119.

McCarthy J. H. (1968) Experiments with mercury in soil gasand air applied to mineral exploration. Min. Eng. 20, 46.

Mercone D., Thomson J., Croudace I. W., and Troelstra S. R.(1999) A coupled natural immobilization mechanism formercury and selenium in deep-sea sediments. Geochim.Cosmochim. Acta 63(10), 1481–1488.

Mierle G. (1990) Aqueous inputs of mercury to Precambrianshield lakes in Ontario. Environ. Toxicol. Chem. 9, 843–851.

Millward G. E. (1982) Nonsteady state simulations of theglobal mercury cycle. J. Geol. Res. 87, 8891–8897.

Monteiro L. R. and Furness R. W. (1997) Accelerated increasein mercury contamination in North Atlantic mesopelagicfood chains as indicated by time series of seabird feathers.Environ. Toxicol. Chem. 16(12), 2489–2493.

Monteiro L. R. and Furness R. W. (2001) Kinetics, dose-response, and excretion of methylmercury in free-living adultCory’s shearwaters. Environ. Sci. Technol. 35(4), 739–746.

Monteiro L. R., Granadeiro J. P., and Furness R. W. (1998)Relationship between mercury levels and diet in Azoresseabirds. Mar. Ecol. Prog. Ser. 166, 259–265.

Munthe J. (1992) The aqueous oxidation of elemental mercuryby ozone. Atmos. Environ. 26A(8), 1461–1468.

Munthe J., and Berg T. (2001) Reply to comment on“Atmospheric mercury species in the European Arctic:measurement and modeling” by Berg et al. Atmos. Environ.35(31), 5379–5380.

Murphy D. M., Thomson D. S., and Mahoney M. J. (1998) Insitu measurements of organics, meteoritic material, mercury,

References 145

Page 40: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

and other elements in aerosols at 5 to 19 kilometers. Science282, 1664–1669.

National Academy of Sciences (NAS) (2000) In ToxicologicalEffects of Methylmercury. Committee on the toxicologicaleffects of methylmercury (ed. R. A. Goyer, chair). NationalAcademy Press, Washington, DC, 344pp.

National Research Council Committee on the ToxicologicalEffects of Methylmercury (2000) Toxicological Effects ofMethylmercury. National Academy Press, Washington,DC.

Nriagu J. O. (1979) Production and uses of mercury. InThe Biogeochemistry of Mercury in the Environment (ed.J. O. Nriagu). Elsevier/North-Hollan Biomedical Press,Amsterdam, pp. 23–40.

Nriagu J. O. (1989) A global assessment of natural sources ofatmospheric trace metals. Nature 338, 47–49.

Nriagu J. O. and Becker C. (2003) Volcanic emissions ofmercury to the atmosphere: global and regional inventories.Sci. Tot. Environ. 304, 3–12.

Norton S. A., Evans G. C., and Kahl J. S. (1997) Comparison ofHg and Pb fluxes to hummocks and hollows of ombrotrophicBig Heath Bog and to nearby Sargent Mt. Pond, Maine,USA. Water Air Soil Pollut. 100, 271–286.

O’Driscol N. J., Siciliano S. D., and Lean D. R. S. (2003)Continuous analysis of dissolved gaseous mercury infreshwater lakes. Sci. Tot. Environ. 304, 285–294.

Olafsson J. (1975) Volcanic influence on sea water at Heimaey.Nature 255, 138–141.

Ozerova N. A. (1996) Mercury in geological systems. InGlobal and Regional Mercury Cycles: Sources, Fluxes andMass Balances. NATO ASI: Series 2. Environment (eds.W. Baeyens, R. Ebinghaus, and O. Vasiliev). Kluwer,Boston, vol. 21, pp. 463–474.

Pacyna E. G. and Pacyna J. M. (2002) Global emission ofmercury from anthropogenic sources in 1995. Water Air SoilPollut. 137, 149–165.

Pak K.-R. and Bartha R. (1998) Mercury methylation anddemethylation in anoxic lake sediments and by strictlyanaerobic bacteria. Appl. Environ. Microbiol. 64(3),1013–1017.

Patterson C. C. and Settle D. M. (1987) Magnitude of lead fluxto the atmosphere from volcanoes. Geochim. Cosmochim.Acta 51(3), 675–681.

Pempkowiak J., Cossa D., Sikora A., and Sanjuan J. (1998)Mercury in water and sediments of the southern Baltic Sea.Sci. Tot. Environ. 213, 185–192.

Petersen G., Iverfeldt A., and Munthe J. (1995) Atmosphericmercury species over central and northern Europe: modelcalculations and comparison with the observations from theNordic air and precipitation network for 1987 and 1988.Atmos. Environ. 29(1), 47–67.

Pleijel K. and Munthe J. (1995) Modeling the atmosphericmercury cycle-chemistry in fog droplets. Atmos. Environ.29(12), 1441–1457.

Poissant L. and Casimir A. (1998) Water–air and soil–airexchange rate of total gaseous mercury measured atbackground sites. Atmos. Environ. 32(5), 883–983.

Porcella D. B., Ramel C., and Jernelov A. (1997) Globalmercury pollution and the role of gold mining: an overview.Water Air Soil Pollut. 97(3–4), 205–207.

Rada R. G. and Powell D. E. (1993) Whole-lake burdens andspatial distribution of mercury in surficial sediments inWisconsin seepage lakes. Can. J. Fish. Aquatic. Sci. 50(4),865–873.

Rampino M. R. and Self S. (1984) Sulfur-rich volcaniceruptions and stratospheric aerosols. Nature 310, 677–679.

Rasmussen P. (1994) Current methods of estimating atmos-pheric mercury fluxes in remote areas. Environ. Sci. Technol.28(13), 2233–2241.

Ravichandran M., Aiken G. R., Ryan J. N., and Reddy M. M.(1999) Inhibition of precipitation and aggregation ofmetacinnabar (mercuric sulfide) by dissolved organic matter

isolated from the Florida Everglades. Environ. Sci. Technol.33, 1418–1423.

Rea A. W., Lindberg S. E., Scherbatskoy T., and Keeler G. J.(2002) Mercury accumulation in foliage over time in twonorthern mixed-hardwood forests. Water Air Soil Pollut.133(1–4), 49–67.

Richter F. M., Rowley D. B., and DePaolo D. J. (1992)Sr isotope evolution of seawater: the role of tectonics.Earth Planet. Sci. Lett. 109, 11–23.

Rolfhus K. R. (1998) The production and distribution ofelemental Hg in a coastal marine environment. PhDDissertation, University of Connecticut.

Rolfhus K. R. and Fitzgerald W. F. (1995) Linkages betweenatmospheric mercury deposition and the methylmercurycontent of marine fish. Water Air Soil Pollut. 80, 291–297.

Rolfhus K. R. and Fitzgerald W. F. (2001) The evasion andspatial/temporal distribution of mercury species in LongIsland Sound, CT–NY. Geochim. Cosmochim. Acta 65(3),407–418.

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 inLake Superior. Environ. Sci. Technol. 37(5), 865–872.

Roos-Barraclough F., Martinez-Cortizas A., Garcia-Rodeja E.,and Shotyk W. (2002) A 14,500 year record of theaccumulation of atmospheric mercury in peat: volcanicsignals, anthropogenic influences and a correlation tobromine accumulation. Earth Planet. Sci. Lett. 202(2),435–451.

Rosson R. (1996) Luciferase operon-containing plasmidengineered for detecting mercury in water. US PatentApplication #5571722.

Roulet M. and Lucotte M. (1995) Geochemistry of mercury inpristine and flooded ferralitic soils of a tropical rain forest inFrench Guiana, South America. WASP 80, 1079–1088.

Sanemasa I. (1975) The solubility of elemental mercury vaporin water. Bull. Chem. Soc. Japan. 48, 1975–1978.

Sanseverino J., Sayler G. S., and Ripp S. A. (2002)Bioluminescent methods for direct visual detection ofenvironmental compounds (US patent # 0214551). PCTInt. Appl. 51pp.

Scherbatskoy T., Shanley J. B., and Keeler G. J. (1998) Factorscontrolling mercury transport in an upland forested catch-ment. Water Air Soil Pollut. 105, 427–438.

Scheuhammer A. M. (1991) Effects of acidification on theavailability of toxic metals and calcium to wild birds andmammals. Environ. Pollut. 71, 329–375.

Schluter K. (2000) Review: evaporation of mercury from soils.An integration and synthesis of current knowledge. Environ.Geol. 39(3–4), 249–271.

Schober S. E., Sinks T. H., Jones R. L., Bolger P. M.,McDowell M., Osterloh J., Garrett E. S., Canady R. A.,Dillon C. F., Sun Y., Joseph C. B., and Mahaffey K. R.(2003) Blood mercury levels in US children and women ofchildbearing age, 1999–2000. J. Am. Med. Assoc. 289,1667–1674.

Schroeder W. H., Anlauf K. G., Barrie L. A., Lu J. Y., SteffenA., Schneeberger D. R., and Berg T. (1998) Arctic spring-time depletion of mercury. Nature 394, 331–332.

Schuster P. F., Krabbenhoft D. P., Naftz D. L., Cecil D., OlsonM. L., DeWild J. F., Susong D. D., Green J. R., and AbbottM. L. (2002) Atmospheric mercury deposition during thelast 270 years: a glacial ice core record of natural andanthropogenic sources. Environ. Sci. Technol. 36,2303–2310.

Scott K. J., Hudson R. J. M., Kelly C. A., Rudd J. W. M., andBarkay T. (2001) Effects of chemical speciation on thebioavailability of Hg(II) in defined media and natural water asmeasured with a mer-lux bioreporter. In Abstr. 6th Int. Conf.on Mercury as a Global Pollutant, Minamata, Japan. 130p.

Seigel F. R., Kravitz J. H., and Galasso J. J. (2001) Arsenic andmercury contamination in 31 cores taken in 1965, St. Anna

Geochemistry of Mercury in the Environment146

Page 41: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Trough, Kara Sea, Arctic Ocean. Environ. Geol. 40(4–5),528–542.

Seinfeld J. H. (1986) Atmospheric Chemistry and Physics ofAir Pollution. Wiley-Interscience, New York, pp. 594–597.

Selifonova O., Burlage R., and Barkay T. (1993) Biolumines-cent sensors for detection of bioavailable Hg(II) in theenvironment. Appl. Environ. Microbiol. 59(9), 3083–3090.

Sellers P., Kelly C. A., Rudd J. W. M., and MacHutchon A. R.(1996) Photodegradation of methymercury in lakes. Nature380, 694–697.

Sellers P., Kelly C. A., and Rudd J. W. M. (2001) Fluxes ofmethylmercury to the water column of a drainage lake: therelative importance of internal and external sources. Limnol.Oceanogr. 46(3), 623–631.

Sheu G.-R. and Mason R. P. (2001) An examination of methodsfor the measurements of reactive gaseous mercury in theatmosphere. Environ. Sci. Technol. 35, 1209–1216.

Shia R.-L., Seigneur C., Pai P., Ko M., and Sze N. D. (1999)Global simulation of atmospheric mercury concentrationsand deposition fluxes. J. Geol. Res. 104(D19),23747–23760.

Skyring G. W. (1987) Sulfate reduction in coastal ecosystems.Geomicrobiol. J. 5, 295–373.

Slemr F. and Langer E. (1992) Increase in global atmosphericconcentrations of mercury inferred from measurements overthe Atlantic Ocean. Nature 355, 434–437.

Slemr F., Seiler W., and Schuster G. (1981) Latitudinaldistribution of mercury over the Atlantic Ocean. J. Geol.Res. 80, 1159.

Smith W. E. (1975) Minamata. Henry Holt and Company,New York.

Sommar J., Gardfeldt K., Stromberg D., and Feng X.-B. (2001)A kinetic study of the gas-phase reaction between thehydroxyl radical and atomic mercury. Atmos. Environ.35(17), 3049–3054.

Spiro P. A., Jacob D. J., and Logan J. A. (1992) Globalinventory of sulfur emissions with 18 £ 18 resolution.J. Geophys. Res. 97/D5, 6023–6036.

St. Louis V. L., Rudd J. W. M., Kelly C. A., Beaty K. G., FlettR. J., and Roulet N. T. (1996) Production and loss ofmethylmercury and loss of total mercury from boreal forestcatchments containing different types of wetlands. Environ.Sci. Technol. 30(9), 2719–2729.

St. Louis V. L., Rudd J. W. M., Kelly C. A., Hall B. D., RolfhusK. R., Scott K. J., Lindberg S. E., and Dong W. (2001)Importance of the forest canopy to fluxes of methyl mercuryand total mercury to boreal ecosystems. Environ. Sci.Technol. 35(15), 3089–3098.

Stock A. and Cucuel F. (1934) The quantitative determinationof microamounts of mercury. Naturwissenchaften 22, 390.

Stoffers P., Hannington M., Wright I., Herzig P., and de RondeC. (1999) Elemental mercury at submarine hydrothermalvents in the Bay of Plenty, Taupo volcanic zone,New Zealand. Geology 27(10), 931–934.

Stoiber R. E. and Jepsen A. (1973) Sulfur dioxide contributionsto the atmosphere by volcanoes. Science 182, 577–578.

Stoiber R. E., Williams S. N., and Huebert B. (1987) Annualcontribution of sulfur dioxide to the atmosphere byvolcanoes. J. Volcanol. Geotherm. Res. 33(1–3), 1–8.

Stordal M. C. and Gill G. A. (1995) Determination of mercurymethylation rates using a 203-Hg radiotracer technique.Water Air Soil Pollut. 80, 725–734.

Stratton W. J. and Lindberg S. E. (1995) Use of a refluxing mistchamber for measurement of gas-phase mercury(II) speciesin the atmosphere. Water Air Soil Pollut. 80, 1269–1278.

Swain E. B. and Helwig D. D. (1989) Mercury in fish fromnortheastern Minnesota lakes: historical trends, environ-mental correlates, and potential sources. J. Minn. Acad. Sci.55, 103–109.

Swain E. B., Engstrom D. R., Brigham M. E., Henning T. A.,and Brezonik P. L. (1992) Increasing rates of atmosphericmercury deposition in midcontinental North America.Science 257, 784–787.

Tescione L. and Belfort G. (1993) Construction and evaluationof a metal ion biosensor. Biotechnol. Bioeng. 42(8), 945–952.

Thakur A. N. and Goel P. S. (1989) Huge variations in theisotopic ratio 196Hg/202Hg in some acid-insoluble resides ofSikhote Alin and other iron meteorites. Earth Planet. Sci.Lett. 96, 235–246.

The Social Scientific Study Group on Minamata Disease(2001) In the Hope of Avoiding Repetition of a Tragedy ofMinamata Disease (What we have learned from theexperience). National Institute for Minamata Disease,Minamata, Japan, 140p.

Tokos J. J. S., Hall B., Calhoun J. A., and Prestbo E. (1998)Homogeneous gas-phase reactions of Hg8 with H2O2, O3,CH3I and (CH3)2S: implications for atmospheric Hg cycling.Atmos. Environ. 32(5), 823–828.

Topping G. and Davies I. M. (1981) Methylmercury productionin the marine water column. Nature 290, 243–244.

Tremblay A., Lucotte M., and Schetagne R. (1998) Totalmercury and methylmercury accumulation in zooplankton ofhydroelectric reservoirs in northern Quebec (Canada). Sci.Tot. Environ. 213(1–3), 307–315.

Tseng C. M., Amouroux D., Abril G., Tessier E., Etcheber H.,and Donard O. F. X. (2001) Speciation of mercury in a fluidmud profile of a highly turbid macrotidal estuary (Gironde,France). Environ. Sci. Technol. 35(13), 2627–2633.

Tseng C.-M., Balcom P. H., Lamborg C. H., and Fitzgerald W. F.(2003) Dissolved elemental mercury investigations in LongIsland Sound using on-line Au amalgamation-flow injectionanalysis. Environ. Sci. Technol. 37, 1183–1188.

Tseng C.-M., Lamborg C. H., Fitzgerald W. F., and EngstromD. R. Cycling of dissolved elemental mercury in arcticAlaskan lakes. Geochim. Cosmochim. Acta (in press).

Turekian K. K. and Wedepohl K. H. (1961) Distribution of theelements in some major units of the Earth’s crust. Geol. Soc.Am. Bull. 72, 175–192.

US Environmental Protection Agency (1997) Mercury StudyReport to Congress. EPA 452/R-97-003 available at: http://www. epa. gov/oar/mercury. html

USAD Consumer Advisory (2001) Mercury levels in seafoodspecies. Center for Food Safety and Applied Nutrition,Office of Seafood.

USFDA (2001) Mercury levels in seafood species. Center forFood Safety and Applied Nutrition, Office of Seafood.

Urban N. R., Eisenreich S. J., Grigal D. F., and Schurr K. T.(1990) Mobility and diagenesis of Pb and 210Pb in peat.Geochim. Cosmochim. Acta 54, 3329–3346.

Vaidya O. C., Howell G. D., and Leger D. A. (2000) Evaluationof the distribution of mercury in lakes in Nova Scotia andNewfoundland (Canada). Water Air Soil Pollut. 117(1–4),353–369.

Vandal G. M., Fitzgerald W. F., and Mason R. P. (1991)Cycling of volatile mercury in temperate lakes. Water AirSoil Pollut. 56, 791–803.

Vandal G. M., Fitzgerald W. F., Boutron C. F., and CandeloneJ. P. (1993) Variations in mercury deposition to the Antarcticover the last 34,000 years. Nature 362, 621–623.

Vandal G. M., Mason R. P., McKnight D., and Fitzgerald W.(1998) Mercury speciation and distribution in a polar desertlake (Lake Hoare, Antarctica) and two glacial meltwaterstreams. Sci. Tot. Environ. 213, 229–237.

Vandal G. M., Fitzgerald W. F., Rolfhus K. R., Lamborg C. H.,Langer C. S., and Balcom P. H. (2002) Sources and cyclingof mercury and methylmercury in Long Island Sound. FinalReport to CTDEP. 105pp.

Varekamp J. C. and Buseck P. R. (1981) Mercury emissionsfrom Mount St. Helens during September 1980. Nature 293,555–556.

Varekamp J. C. and Buseck P. R. (1983) Mercury anomalies insoils: a geochemical exploration method for geothermalareas. Geothermics 12(1), 29–47.

Varekamp J. C. and Buseck P. R. (1984) The speciation ofmercury in hydrothermal systems, with applications to oredeposition. Geochim. Cosmochim. Acta 48(1), 177–185.

References 147

Page 42: 16Treatise 9_04 Geochemistry of Mercury in the Environment W_ F_ Fitzgerald and C_ H_ Lamborg

Varekamp J. C. and Buseck P. R. (1986) Global mercury fluxfrom volcanic and geothermal sources. Appl. Geochem. 1(1),65–73.

Varekamp J. C., Buchholdtz ten Brink M. R., Mecray E. L., andKreulen B. (2000) Mercury in Long Island Sound sediments.J. Coast. Res. 16(3), 613–626.

Verta M. and Matilainen T. (1995) Methylmercury distributionand partitioning in stratified Finnish forest lakes. Water AirSoil Pollut. 80(1–4), 585–588.

Virta M., Lampinen J., and Karp M. (1995) A luminescence-based mercury biosensor. Anal. Chem. 67, 667–669.

Wang Q., Shen W., and Ma Z. (2000) Estimation of mercuryemission from coal combustion in China. Environ. Sci.Technol. 34, 2711–2713.

Watras C. J. and Bloom N. S. (1994) The vertical distributionof mercury species in Wisconsin lakes: accumulation inplankton layers. In Mercury Pollution: Integration andSynthesis, Chap. I.11 (eds. C. Watras and J. Huckabee).Lewis Publishers, Ann Arbor, MI, pp. 137–152.

Watras C. J., Bloom N. S., Hudson R. J. M., Gherini S.,Munson R., Claas S. A., Morrison K. A., Hurley J., WienerJ. G., Fitzgerald W. F., Mason R., Vandal G., Powell D.,Rada R., Rislov L., Winfrey M., Elder J., Krabbenhoft D.,Andren A. W., Babiarz C., Porcella D. B., and HuckabeeJ. W. (1994) Sources and fates of mercury and methylmer-cury in Wisconsin lakes. In Mercury Pollution: Integrationand Synthesis, chap. I.12 (eds. C. Watras and J. Huckabee).Lewis Publishers, Ann Arbor, MI, pp. 153–177.

Watson C. M., Dwyer D. J., Andle J. C., Bruce A. E., and BruceM. R. M. (1999) Stripping analyses of mercury using goldelectrodes: irreversible adsorption of mercury. Anal. Chem.71, 3181–3186.

Westoo G. (1966) Determination of methylmercury com-pounds in foodstuffs: I. Methylmercury compounds in fish,identification and determination. Acta Chem. Scand. 20(8),2131–2137.

White D. E. (1967) Mercury and base-metal depositswith associated thermal and mineral waters. InGeochemistry of Hydrothermal Ore Deposits (ed. H. L.Barnes). Holt, Rinehart and Winston, New York, chap. 13,pp. 575–631.

Wiener J. G. and Spry D. J. (1996) Toxicological significanceof mercury in freshwater fish. In Environmental Contami-nation of Wildlife (eds. W. N. Beyer, G. H. Heinz, and A. W.Redmon-Norwood). Lewis Publication, Boca Raton, FL,USA. pp. 297–339.

Wiener J. G., Fitzgerald W. F., Watras C. J., and Rada R. G.(1990a) Partitioning and bioavailability of mercury in anexperimentally acidified Wisconsin lake. Environ. Toxicol.Chem. 9(7), 909–918.

Wiener J. G., Martini R. E., Sheffy T. B., and Glass G. E.(1990b) Factors influencing mercury concentrations inwalleyes in northern Wisconsin lakes. Trans. Am. Fisher.Soc. 119, 862–870.

Wiener J. G., Krabbenhoft D. P., Heinz G. H., andScheuhammer A. M. (2002) Ecotoxicology of Mercury.In Handbook of Ecotoxicology, 2nd edn. (eds. D. J.Hoffman and B. A. Rattner). Lewis Publishers, BocaRaton, FL, USA.

Wilhelm S. M. (1999) Design mercury removal systems forliquid hydrocarbons. Hydrocarb. Process. 78(4), 61.

Wilhelm S. M. (2001) Estimate of mercury emissions to theatmosphere from petroleum. Environ. Sci. Technol. 35(24),4704–4710.

Winfrey M. R. and Rudd J. W. M. (1990) Environmental factorsaffecting the formation of methylmercury in low pH lakes.Environ. Toxicol. Chem. 9(7), 853–869.

Wu Q., Apte S. C., Batley G. E., and Bowles K. C. (1997)Determination of the mercury complexation capacity ofnatural waters by anodic stripping voltammetry. Anal. Chim.Acta 350, 129–134.

Wylie D. E., Carlson L. D., Carlson R., Wagner F. W., andSchuster S. M. (1991) Detection of mercuric ions in water byELISA with a mercury-specific antibody. Anal. Biochem.194, 381–387.

Wylie D. E., Lu D., Carlson L. D., Carlson R., Babacan K. F.,Schuster S. M., and Wagner F. W. (1992) Monoclonalantibodies specific for mercuric ions. Proc. Natl. Acad. Sci.USA 89, 4104–4108.

Xia K., Skyllberg U. L., Bleam W. F., Bloom P. R., NaterE. A., and Helmke P. A. (1999) X-ray absorption spectro-scopic evidence for the complexation of Hg(II) by reducedsulfur in soil humic substances. Environ. Sci. Technol. 33,257–261.

Xiao Z. F., Munthe J., Schroeder W. H., and Lindqvist O.(1991) Vertical fluxes of volatile mercury over forest soiland lake surfaces in Sweden. Tellus 43B, 267–279.

Young D. R., Johnson J. N., Soutar A., and Isaacs J. D.(1973) Mercury concentrations in dated varved marinesediments collected off southern California. Nature 244,273–275.

q 2003, Elsevier Ltd. All rights reservedNo part of this publication may be reproduced, stored in a retrieval system ortransmitted in any form or by any means, electronic, mechanical, photocopying,recording or otherwise, without prior written permission of the Publisher.

Treatise on GeochemistryISBN (set): 0-08-043751-6

Volume 9; (ISBN: 0-08-044344-3); pp. 107–148

Geochemistry of Mercury in the Environment148