thiemans 2011 isotop..

advertisement

Mass Independent Isotopic Composition of Terrestrial and Extraterrestrial

Materials

M.H. Thiemens and Robina Shaheen

Department of Chemistry and Biochemistry 0356

University of California, San Diego, CA, USA

4.06.1 General Introduction

Stable isotope ratio measurements have been utilized for more than 60 years as a tool in an resolving extended range of chemical and natural processes. This includes, for example, igneous and sedimentary thermometry, paleothermometry, hydrological cycle events (evaporation, condensation), biogeochemical processes, atmospheric and planetary chemistry and solar system evolution. The theoretical basis for calculating the position of equilibrium in isotopic exchange between two molecules was first developed from calculation of isotopic reduced partition

function ratios (Bigeleisen and Mayer, 1947; Urey, 1947)

. It was shown that the position of isotope exchange equilibrium between two differing isotopically substituted molecules (such as H

2

18 O and C 16 O

2

) may be calculated as a function of temperature at very high precision from knowledge of the isotopic reduced partition functions. The difference in chemical behavior for isotopically substituted molecules in this specific instance arises from quantum and statistical mechanical effects and the characteristics of the vibrational frequencies of the isotopically substituted molecules. It is well established that a vibrating molecule is energetically represented as a near classical harmonic oscillator. Thus, in the case of isotopically substituted molecules, the quantized energy levels vary, with the heavier isotopic species possessing lower vibrational frequencies. As a consequence, heavy isotopically substituted molecules possess slightly stronger bond strengths, and the relative reactivities of the isotopically substituted species may be precisely determined. The predominant parameter is the vibrational frequency, which are measured spectroscopically at high resolution facilitating calculation of the isotope exchange values typically in the tenths of per mil level. As a result, knowledge of the partition function temperature dependency and, measurement of the isotopic compositions of the two molecular species precisely fixes the temperature of equilibrium basic physical

chemical (Van Wyngarden et al., 2007) principle forms the cornerstone of igneous and paleo-thermometry. The theoretical developments of Urey and Bigeleisen (Bigeleisen and Mayer, 1947; Urey, 1947) serendipitously

coincided with the development of the isotope-ratio mass spectrometer (Nier, 1947). Prior to this isotope ratios

could not be measured at a sufficient precision to apply to natural systems because of the small range of isotope ratios. The isotope ratio mass spectrometer permitted isotope ratio measurements at a precision and accuracy of natural samples allowed development of paleothermometry and the use of oceanic planktonic foraminifera to measure ocean temperature changes over long time periods. Subsequently, a broad diversity of isotopic fractionation processes have been resolved in chemical reaction kinetics, molecular atmospheric diffusion (Earth and Mars), evaporation–condensation (e.g. the hydrological cycle), mineral crystallization temperatures, and biologic processes

1

(e.g., photosynthesis, respiration, nitrogen fixation, sulfate reduction, and transpiration). The application of relevant isotopic measurements requires basic understanding of physico-chemical principles associated with these transformations. In this Chapter new varieties of isotopic phenomena and how their chemistry is applied towards understanding a variety of natural processes, terrestrial and extra-terrestrial, present and past, are discussed.

First Applications of Mass Independent Isotopic Observations

Although, conventional isotope effects vary widely in the physicochemical basis origin, they all are ultimately dependent upon relative isotopic mass differences for the isotopically substituted molecules. It was shown

that isotope effects alter isotope ratios in a relation dependent upon relative mass differences (Hulston and Thode,

1965b). For example, the changes in

33 S and

34 S (or the changes in the 33 S/ 32 S, 34 S/ 32 S ratios respectively, expressed in parts per thousand variation with respect to a standard, which for sulfur is Canon Diablo troilite) are shown to be highly correlated such that the change in

33 S is half that of the concomitant

34 S value. The mass range in the

33 S value is 1 amu (33 – 32 amu) and for

34 S is 2 amu (34 – 32 amu). Thus the relation:

δ 33 S = 0.5δ 34 S (1) is observed in terrestrial geological samples. Equation (1) arises as a consequence of the reliance of chemical and physical processes known at that time, from the relative isotopic mass differences of the sulfur bearing molecules.

The actual coefficient varies slightly vary from 0.516 to 0.524, depending upon the molecular mass and the

fractionation process and has been discussed in detail in several reviews (Matsuhisa et al., 1978; Thiemens, 2006;

Weston, 1999). The ability to measure the coefficient to three significant has facilitated understanding of

atmospheric, geological, and biological processes not achievable otherwise. Development of the concept of mass dependence as an isotopic marker was originally conceived to resolve nuclear (nucleosynthetic or spallogenic) from

non-nuclear components in meteorites (Hulston and Thode, 1965a, b). A highly anomalous oxygen isotopic

composition (

17 O

 

18 O) was observed in the high-temperature calcium–aluminum inclusions (CAIs) present in

the carbonaceous chondritic meteorite Allende and attributed to nuclear process (Clayton et al., 1973) as all known

physicochemical processes produced expected mass-dependent

17 O

0.5

18 O. Figure 1 displays the slope one line

(termed Allende CAI Line) in a three isotope plot of oxygen. The slope ½ line represents the solid terrestrial oxygen isotopic reservoir of the bulk earth and moon, with the origin at Standard Mean Ocean Water (SMOW) and which expresses the normal mass dependent isotopic fractionation processes.

A three isotope plot displaying the oxygen isotopic composition of ozone produced from molecular oxygen

at liquid nitrogen temperature (Thiemens and Heidenreich, 1983) is shown in Fig. 2. The slope of this line is 1.0,

thus reproducing the slope of the CAI line depicted in Figure 1, thought to be a nuclear process. The starting molecular isotopic composition is at δ 18 O = δ 17 O = 0.

2

-50

50

Stratospheric Ozone

(70-110, 70-90)

25

Mass-Fractionation Line

(Slope = 0.5)

Sedimentary Rocks

-30

10

H Chondrites

Enstatite Chondrites

Meteoric Water

-10

-10

L and LL

Chondrites

Atmospheric Oxygen

Basaltic / Lunar Rocks

CM Chondrites

10

SMOW

C3 Chondrites

30

 18 O (‰)

50

A lle nde

C

AI

L ine

(-40, -40)

-25

-50

Figure 1. Three isotope oxygen isotopic composition of meteoritic and some terrestrial reservoirs. The point at (-40,-40) at the terminus of the slope one line reflects the general calcium-aluminum inclusion (CAI) line first observed by Clayton et al (1973).

It was suggested that this anomalous isotopic composition must derive from a nuclear, rather than chemical process as physical or chemical processes would move along the slope of ½ line, in either direction depending upon the relevant fractionating process. Arithmetically, a

17 O =

18 O relation may arise in two ways, either by alteration of 17 O and 18 O by equal amounts or, by addition/subtraction of pure 16 O. Existing models for supernova processes at that time demonstrated that under certain nuclear reaction conditions and within certain stellar zones, nearly pure

16

O is produced. It was suggested (Clayton et al., 1973) that it is unlikely that

17 O and 18 O would be equally altered in any nuclear reaction sequence , and a 16 O supernova event was deemed the most plausible process to account for the Allende isotopic observations. However, once the basic assumption that chemical reactions must be mass dependent was found to be incorrect when mass independent oxygen isotope fractionation processes was

demonstrated during ozone formation (Thiemens and Heidenreich, 1983) from molecular oxygen. The ozone

produced was equally enriched in 17 O and 18 O (

17 O =

18 O) with respect to its precursor molecular oxygen. Figure 2 schematically displays the results of those experiments and this lead to the reconsideration of the existing enigmatic theories of the possible source of oxygen isotopic anomaly.

3

60

Product Ozone

Molecular oxygen

Slope = 1.0

40

20 Slope = 0.52

-80 -60 -40 -20

0

0

-20

20 40

 18 O (‰)

60 80

-40

-60

-80

Figure 2. A three isotope plot displaying the oxygen isotopic composition of ozone produced from

molecular oxygen at liquid nitrogen temperature (Thiemens and Heidenreich, 1983). The slope of this line is 1.0,

thus reproducing the slope of the CAI line depicted in Figure 1. The starting molecular isotopic composition is at

δ 18 O = δ 17 O = 0.

The observations displayed in Figure 2 have several consequences. First, the assumption that only a nuclear process may produce mass-independent isotopic compositions is invalid. Though nucleosynthetic components are present in meteorites, the distinction between chemical and nuclear is not straightforward.Secondly, the

17 O =

18 O relation is identical to that for Allende CAIs. Whether this feature accounts for the observed meteoritic oxygen

isotopic compositions remains unresolved, though nucleosynthetic contributions (Clayton et al., 1973) have now

been abandoned and most likely derive from chemical processes (Thiemens and Heidenreich, 1983)

.

The source of

the meteoritic oxygen anomalies remains one of the largest unanswered questions in cosmochemistry (Thiemens,

2006) The existing theories for the production of the oxygen anomalies is discussed in detail in an extended volume

on oxygen in the solar system, including the factors controlling the distribution of oxygen (MacPherson, 2008).

Recent advances in the ability to measure isotopic distributions within single mineral grains is possible due to the development of secondary ion mass spectroscopy (SIMs), nano-SiMS, and continuous-flow isotope-ratio mass spectrometry and have facilitated deeper understanding of nebular processes associated with production and secondary alteration of oxygen isotopic compositions. The ability to measure isotopic distributions on tens of nano meter scale lengths requires that enhanced theoretical understanding of the physical chemistry of how isotopic

4

distributions result at these sizes be achieved. An outcome of the development of oxygen isotopic analytical capabiliitiesfor single grain oxygen isotopic measurement was a key factor in the abandonment of the nuclear model for the meteorite oxygen-isotopic compositions. The prediction of 16 O carrier grains or, a direct correlation with

other supernova-produced isotopes (Clayton et al., 1973) was never observed after searches at the highest spatial

resolution scales. It is well documented that there are grains of highly anomalous oxygen present in meteorites that

are definitively nucleosynthetic (Nittler, 2003), however they have neither the requisite material amount to explain

the observed bulk level meteoritic oxygen isotopic compositions, or the required isotopic composition. As a consequence, chemical and/or photochemical processes are the preferred mechanism for production of meteoritic oxygen isotopic compositions.

The first theory for chemical production of meteoritic oxygen isotopic components by photochemical

isotopic self-shielding by an early active Sun (Clayton, 2002, 2003; Lyons and Young, 2005; MacPherson, 2008;

Thiemens and Heidenreich, 1983), or from symmetry dependent chemical reactions (Heidenreich and Thiemens,

1986; Marcus, 2004; Thiemens, 2006) are all mechanisms under active consideration. The goal of these models is to

describe the mechanism by which the bulk oxygen isotopic composition of meteorites is acquired as well as the CAI features. Several recent models have been developed to describe the evolution of oxygen isotopes in the solar nebula

(Krot et al., 2005a; Krot et al., 2006). Figure 3 depicts some bulk level oxygen isotopic compositions of meteorites

that lie above and below the bulk Earth, Mars and the moon fractionation line. The size of the isotopic anomaly and the fact that oxygen is the most abundant element of stony object in the solar system requires that the responsible process is dominant in the evolution of the solar system.

5

Figure 3. A three isotope plot of the oxygen isotopic composition of meteoritic material. The symbols denote:

Terrestrial Fractionation Line (TFL), CV3,CO3,CM, CR are classes of carbonaceous chondrites, enstatite chondrites

(E), ordinary chondrites (H,L,LL), Rumaruti Chondrites (R), and calcium aluminum inclusions (CAI).

Experiments have now tested the nebular self shielding models of carbon monoxide (Chakraborty et al.,

2008) . The work is requisite for understanding in general, the isotope effects associated with basic physical

chemical processes such as photodissociation molecules. Furthermore, such analysis are required to model many natural processes, not only nebular but also planetary atmospheric. To perform self shielding experiments relevant to the pre solar nebula; CO photodissociation must be done at short wavelengths (< 110 nm) and, at relevant CO columnar densities. Shielding requires isotopically discrete spectral lines in the region between 90 and 110 nm and under such restrictions, and in spectral regions where molecular hydrogen does not absorb. As such, 105.17 nm is the wavelength where it must occur. The results of these experiments are shown in Figure 4.

Fig. 4. Three isotope plot showing anomalous oxygen isotopic composition of CO

2

produced during CO photolysis at different wavelengths. CAI line is shown for comparison. Adapted from Chakraborty et al, (2008)

The data in Fig. 4 show that there is a massive mass independent isotope effect in the process of CO photodissociation. The similarity in slope of wavelengths where self shielding should occur and where it may not was considered demonstration that self shielding does not occur and rather, there is a large isotope effect associated

with the actual photodissociation process (Chakraborty et al., 2008). Recent theoretical calculations (Muskatel et al.,

2011) have shown that in general, due to resonance effects between excited states there is a highly energy selective

effect in photolysis and as a result, a highly sensitive wavelength dependency resulting during photodissociation.

Such considerations are not taken into account in self shielding models and it is assumed that there is no isotope effect during photolysis. The interpretation of the CO work is a matter of active discussion at present and further

experiments and modeling efforts are needed (Chakraborty et al., 2009; Federman and Young, 2009; Lyons et al.,

6

2009b; Smith et al., 2009; Yin et al., 2009). To understand the origin of the meteoritic oxygen isotopic anomalies in

CAI two models, self shielding and chemically produced mass independent chemical fractionation processes, were

reviewed (Lyons et al., 2009a). It was concluded that the measurements of the solar wind by the Genesis spacecraft

would provide a test of self shielding and the oxygen isotopic composition should resemble CAI’s. The self shielding model predict a

17 O of -25 per mil and at first, apparently consistent with the early return solar wind samples from the Genesis space craft, which had a

17 O=

18 O=-56-57 per mill, as predicted by models. It is shown that the Genesis solar wind, however, has a value at

17 O =-79,

18

O = -99 per mil (McKeegan et al., 2011), far from

that predicted by self shielding models(Lyons et al., 2009a). A further argument against self shielding was

developed on the basis that the solar wind may not necessarily be relevant to understanding meteoritic oxygen

isotopes (Ozima et al., 2007). From a statistical analysis of the oxygen isotopic (

17 O) composition of approximately 400 meteorites, a systematic trend was demonstrated revealing that the values statically scatter about zero per mil (terrestrial). The scatter for the mean of

17 O = O decreases with representative parent body size. The trend captures the effect of accretion of smaller bodies and the values consequently should approach the implied solar value. From their analysis it was concluded that solar oxygen is the major reservoir of oxygen and isotopically is the same as terrestrial, but different from CAI. This implies then that the solar wind samples do not reflect bulk solar and, that isotopic self-shielding did not produce meteoritc anomalies.

An important aspect of the early solar system oxygen issue is that they are concerned with the basic isotopic effect associated with photodissociation of CO. Investigations of such basic processes are relevant to the present atmosphere as understanding them is needed for interpretation of terrestrial and Martian atmospheric isotopic compositional measurements. Quantum numeric simulations have indicated that a strong and selective isotope effect from UV excitation of molecular nitrogen occurs and these selective isotope effects may be observed

during photodissociation (Muskatel et al., 2011). During the photolysis of nitrogen, where all relevant potential

energy surfaces are known at high spectral resolution, an additional isotope effect due to the coupling of diabatic electronic states of different bonding nature occurs. The effect arise where there are two or more excited states of the same, or nearly same energy. Numerical solutions of the time dependent Schrödinger equation of both the electronic and nuclear modes are included in the calculations. Population of excited valence and Rydberg states prior to dissociation were observed with these numeric simulations, and the process is highly wavelength dependent. When the isotope effect associated with the wavelength dependent population of crossing states is convolved with the structured solar spectrum, large isotope effects are apparent, especially at the CIII solar emission line (97.03 nm).

Similar calculations are needed for CO to provide greater details of nebular photochemistry.

Aside from photophysical isotopic effects, understanding new physical phenomena that alter isotopic compositions are presently being developed to understand both meteoritic and atmospheric processes. In accounting for the observed meteoritic oxygen isotopic anomalies non photochemical models have been suggested. MIF in the formation of ozone was observed and attributed to the symmetry of the oxygen molecule. Interestingly the experimentally produced O-isotope fractionation line of the ozone and left over oxygen reservoir was equal in slope

to the observed carbonaceous chondrite mixing line (Heidenreich and Thiemens, 1983, 1986; Thiemens and

7

Heidenreich, 1983)A physical-chemical model based upon isotopic symmetry and dynamical effects in CAI

formation was developed to account for observed oxygen isotopic compositions (Marcus, 2004). It is hypothesized

that on the surface of pre solar metal oxide grains, reactions of oxygen atoms with metallic oxides (e.g. SiO, CaO) lead to symmetry driven isotope effects. Due to a high entropic concentration factor on mineral surfaces, the enrichment leads to a reaction rate increase of many orders of magnitude over gas phase volumetric reactions. On grain surfaces, reactions such as MO

(ads)

+ O

(ads)

→MO

2

*, with MO

2

* being a vibrationally excited molecule and M may be atomic species such as Si, Al, Ca, or Ti, readily occur. The MO

2

* is equally enriched in 17 O, 18 O as a consequence of the symmetry of the molecules and associated mass independent isotopic factors considered in the reactions. The excited molecule evaporates into the gas phase enriching the residual precursor MO and O species in

16 O leading to growth of solid grains of high 16 O enrichment, thus resembling CAI. This model has the advantage in that it produces the slope one configuration in a three isotope plot and accounts for the observation that CAI possess isotopic compositions clustering at about

17 O=

18 O around -50 to -60 per mil. In this consideration only one reservoir is required, a significant advantage over the requirement for multiple reservoir mixing in self shielding models. The present limitation is that there are no relevant experiments to explore the fractionations associated with metal oxide reactions under nebular conditions.

Recently, a heterogeneous chemical reaction model was reported as a new mechanism for production of the origin of the various 16

O reservoirs of the solar system such as those displayed in Figure 5 (Dominguez, 2010). In this

model, heterogeneous chemical reactions on the surfaces of interstellar (IS) dust grains in dense molecular clouds, such as our solar systems parent molecular cloud, are capable of producing oxygen reservoirs from reactions involving water ice. In such environments, accretion of atomic species (H, C, N, and O) on cold IS grain surfaces react with water to produce a variety of molecular species, particularly on 10 5 -10 6 year time scales. A surprising outcome of the ice model is that though the process occurs in interstellar molecular cloud environments of high molecular hydrogen mixing levels, surface grain reactions lead to an ozone intermediary with a large mass independent isotopic anomaly and

17 O=

18 O which is subsequently transferred to different reservoirs. In the model with ozone assisted ice reactions, 16 O enriched nebular gas reservoir and H

2

0-ice, 16 O depleted reservoirs are created along the slope one fractionation line over time periods of 10 6 -10 7 years. Figure 5 schematically represents this process in the parent molecular cloud and the interaction between residual gas and ice grains.

8

Fig. 5. Oxygen in the solar system on a triple-oxygen isotope plot (δ 17 O vs. δ 18 O). Calcium-aluminum-rich inclusions are shown in green, chondrules in white, and SMOW is at the origin by definition. The TFL with the slope ½ line is shown (dashed). The oxygen isotopic composition of Mars (Δ 17 O 0.3‰) is shown in red, asteroidal H

2

O, as inferred by Choi et al. ( 1998

) from studies of magnetite (Δ 17 O

5‰), is shown in blue and molecular cloud is indicated by the gray region.

This model also has the advantage that only one initial reservoir is required and, is consistent with astronomical observations of IS grains and their spectral features. Future models and relevant ice experiments will be of importance in determining the role of such chemistry in the early solar system. In addition, the role of surface chemistry such as this is of relevance in the Earth’s present day atmosphere since of the most poorly quantified

aspects of atmospheric chemistry is the role of heterogeneous chemistry. The ice model (Dominguez, 2010)

underscores the role of ices, which bears similarity to the well-known polar stratospheric clouds that exert a strong mediating factor in the chemistry of the winter polar atmospheres. As will be discussed, recent atmospheric and potentially Martian meteoritic measurements have shown the importance of grain surface nano film layer chemical

reactions in heterogeneous chemical transformational processes (Shaheen et al., 2010).

4.0. Isotopic anomalies in Extra Terrestrial Atmospheres and Environments

Nebular sulfur, like oxygen has the possibility of production of symmetry, photochemical, and self shielding isotope effects and consequently bears many similarities to oxygen. Sulfur however possesses multiple valence states and has a high volatility rendering it less well suited to maintain its original nebular isotopic signature as a result of secondary exchange reactions. It has the advantage though of possession of four stable isotopes. The only meteoritic class that has the best opportunity to record nebular sulfur isotopic signatures are the ureilites. These

9

meteorites have a combined low sulfur content that prevent secondary alterations and are isotopically anomalous at the bulk level. These are achondritic meteorites, composed mostly of olivines and pyroxenes and possess graphite,

diamond, sulfides, metal, and minor silicates (Mittlefehldt, 1998; Mittlefehldt and Lindstrom, 1998). The source of

these meteorites has remained enigmatic, and their bulk isotopic composition possess highly unusual oxygen

isotopic compositions (Clayton and Mayeda, 1988; Goodrich et al., 1987). Mass independent sulfur isotopic

compositions of these meteorites, with slight 33

S excesses has also been reported (Farquhar et al., 2000a). The

33 S enrichment may potentially derive from a combination of chemical, photochemical, and nuclear processes and stepwise chemical extractions have shown that several achondritic meteorites possess mass independent sulfur

isotopic components (Rai et al., 2005). Four achondritic meteorite groups are observed to possess mass independent

isotopic compositions. Figure 6 is a four isotope sulfur plot of their data and capital delta is a measure of the deviation from mass dependence, and a zero value denotes purely mass dependent compositions. Consequently, the figure displays mass independent correlations of multi isotopes.and associated isotope fractionation processes. The precision and accuracy of measurement of sulfur isotopes is significantly better than for oxygen because of the use of sulfur hexafluoride as the analyte gas. The source of the isotopic anomalies in the meteorite was suggested as

being the same as for the Archean and present day atmospheres, deriving from photochemical reactions (Rai et al.,

2005).

0.15

0.10

0.05

KrF laser

(CS

2

) x

(248

nm)

-Solar

HED

A rF

SO la ser

(1

93

2

nm)

ph ot ol ys is

(Xe la mp)

(CS

2

) x

-UV

0.00

Oldhamite

Ureilites

Aca-Lod

(Norton County)

CDT

Aubrite

-0.05

H

2

S photolysis

-0.3

-0.2

-0.1

0.0

0.1

0.2

 36

S

CDT

(‰)

Figure 6. A plot of the deviation from mass independence for the four stable isotopes of sulfur. In this plot, mass dependent isotopic compositions have

S values of zero. The reference lines indicate isotopic fractionation trajectories for laboratory based photochemistry experiments, as discussed in the text.

10

There are no other chemical processes known in laboratory experiments or nature that produce this array. It was

suggested (Rai et al., 2005) that the actual process was a gas phase irradiation of sulfur species by an early active

sun, such as an X-wind. Figure 7 displays a potential model of an embedded early UV active sun. In this model, the early phase of solar evolution produces copious amounts of short wavelength UV light as the sun undergoes ignition and proceeds towards the main sequence. During this time, there is convective mixing in the solar nebula which may result in the photochemistry of sulfur bearing molecules and produce the observed mass independent sulfur isotopic anomalies. The sulfur anomaly is produced in a reducing environment, consistent with its observation in the more reduced meteoritic classes. It is also possible that during this photolytic event, oxygen was also photochemically reacted, though at present there is no evidence for a linkage between the two elements.

Figure 7. An illustration indicating how gas phase photolysis by the protosun may produce mass independent sulfur isotopic anomalies.

Short wavelength UV photolysis may produce the sulfur isotopic anomalies observed in meteorites and, indicates that the sun was active prior to planet body condensation. The results are the first direct evidence of photochemistry in the proto-nebular environment and has implications for the synthesis of organic molecules. Mass independent sulfur isotopic compositions in organic molecules extracted and isolated from the Murchison meteorite (methyl,

propyl, and ethyl sulfonic acids) have been reported (Cooper et al., 1997a; Cooper et al., 1997b). It was suggested

that these molecular species were created by photochemistry in the nebula, rather than on planetary bodies. This implies that the organic photochemical process was prior to formation of the planetary bodies. If true, then the

11

delivery of organics to the pre biotic Earth would be at least partially governed by photochemical processes and future studies of meteoritic organic molecules will further resolve processes of the origin and evolution of the

Earth’s atmosphere.

It is now well known that there exist preserved interstellar grains in meteoritic material (Nittler, 2003;

Nittler and Alexander, 1999; Wasserburg et al., 2006) . There is an extensive literature on this subject and will not

be reviewed here. The link between short-lived nuclei and their potential nucleosynthetic sources, such as AGB stars has been discussed in details

(Wasserburg et al., 2006). The analysis establishes chronologies and the recognition of

intervention of specific astrophysical processes in the earliest history of the solar system. In the case of potential oxygen nucleosynthetic contributions there are two key points. First, it is no longer considered likely that the source

of the bulk meteoritic oxygen isotopic anomalies is nucleosynthetic (Clayton, 2002) Second, there are recognizable

nucleosynthetic preserved grains and the concentration of these grains is typically parts per million or less, which is insufficient anomalous oxygen isotopic atoms to produce the observed bulk level meteorite compositions.

4.06.1.1 The Physical Chemistry of Mass-independent Isotope Effects

The first observations of a chemically produced mass independent isotopic fractionation (MIF) process

(Thiemens and Heidenreich, 1983)

revealed that there existed no known physical–chemical mechanism that can account for the ozone isotopic enrichments. Since the recognition of the ozone isotopic anomaly in laboratory experiments and in the atmosphere, there have been extensive studies directed towards understanding the MIF

processes of ozone and other oxygen containing molecules, including sulfur tetra isotopes (Babikov et al., 2003a;

Babikov et al., 2003c; Brenninkmeijer et al., 2003; Gao and Marcus, 2001; Gao and Marcus, 2007; Thiemens, 1999,

2002, 2006; Weston, 1999) Recently, various aspects of the physical chemistry of gas-phase mass-independent

processes and the photolysis of long-lived predissociative sulfur bearing molecules has been identified (Lyons,

2008; Mauersberger et al., 2005). The observations of mass-independent isotopic compositions (both oxygen and

sulfur) in various solid reservoirs of Earth and Mars, including theoretical development of the MIF processes have been reviewed

(Thiemens, 2006).

In the first observation of a chemically produced oxygen anomaly, a mechanism

based upon optical self-shielding was proposed (Heidenreich and Thiemens, 1983; Thiemens and Heidenreich,

1983) . Subsequently, self shielding was apparently ruled out (Navon and Wasserburg, 1985) and differential

chemistry for asymmetric 16 O 16 O 17 O, 16 O 16 O 18 O versus symmetric 16 O 16 O 16 O species for the observed mass-

independent effect was suggested (Heidenreich and Thiemens, 1986). In this mechanism, an equal

17 O,

18 O fractionation occurs as a result of the identical chemistry for the asymmetric isotopomeric species with respect to symmetric 16 O 16 O 16 O, leading to heavy isotope enrichment in the product ozone. It was the rotational symmetry properties that originally led to the discovery of oxygen isotopes. With the sun as the background irradiance source, it was observed that there are extra rotational absorption lines for 16 O 18 O and 16 O 17 O as compared to 16 O 16 O

demonstrating the existence of the heavy oxygen isotopes in atmosphere oxygen (Giauque and Johnston, 1929a, b, c) . This arises because of the well-known line doubling in rotational spectra for asymmetric species. As discussed in

those papers, besides the importance of the discovery of oxygen isotopes, the work was also a significant

confirmation of early predictions of quantum theory. It was suggested (Heidenreich and Thiemens, 1986) that the

12

rotational isotopic symmetry and the associated properties may produce a longer lifetime for the asymmetric isotopomer and consequently a greater probability of stabilization to the ground state ozone molecule. This was based upon the following. 1) The probability of an excited O

*

3

intermediate energetically quenching via collisions is determined by the product of its lifetime (

) and its collisional frequency. 2). The collisional frequencies derive from mass-dependent effects, such as velocity, and the source of the mass independence may not arise from the collisional terms for stabilization. 3).The lifetime of the 17 O and 18 O species ( 16 O 16 O 17 O, 16 O 16 O 18 O) is equal and longer than symmetric 16 O 16 O 16 O leading to a product equally enriched in 17 O and 18 O, i.e.,

17 O =

18 O.

The exact mechanism arises in the process of reaction from the intermediate excited state. Descriptions of such

reactions are discussed in texts (Helene Lefebvre-Brion, 2004; Herzberg, 2008). During atom–molecule collisions,

the reactants interact with one another subject to the relevant potential energy surface. The lifetime of the excited intermediate is on the order of molecular vibrational periods, or ~10

–13 s. The lifetime is a complex function of the chemical reaction dynamics, which in turn depends on the number of available rotational or vibrational states. In this specific instance, there is dependence for the isotopically substituted species. Ozone of pure 16 O has a C

2v

symmetry and half the rotational complement of the asymmetric isotopomers. As a result, it was suggested that the extended lifetime for the asymmetric species leads to a greater probability of stabilization. While these assumptions are valid for a gas phase molecular reaction, they do not account for the totality of the experimental ozone isotopic observations.

At present, the most complete model for the ozone effect is that of Markus and colleagues (Gao and Marcus,

2007; Hathorn and Marcus, 1999, 2000, 2001) in accounting for many of the experimental observations. RRKM

theory is employed in its development, which quantitatively describes the energetics of gas phase atom–molecular encounters and the relevant parameters leading to either stabilization and product formation or re-dissociation to atomic and molecular species, with the energetic intermediate being a critical species. These models determine individual rate constants for isotopically and positionally substituted species of ozone. In a theoretical RRKM approach, all parameters are mass dependent, such as collisional frequency, bond strength, vibrational, and rotational energies and symmetry dependence. During the stabilization process of the excited intermediate (

*

O ), there is a

3 partial dependence upon the rate at which excess energy is dispersed. If this does not occur on a sufficiently short time scale stabilization does not occur and the excited

*

O

re-dissociates to the original O and O

3

2

reactants. Part of this energy dispersal process depends upon symmetry factors though the largest contributions are the massdependent processes. It is this symmetry factor that is the source of the mass independence. In the model, there is an inclusion of a “

effect,” which is a modest deviation from the statistical density of states for symmetric versus

asymmetric species(Yi Qin and Marcus, 2002; Yi Qin et al., 2002). There exists a partitioning of energy that derives

from very slight differences in zero-point energies for the exit channels for dissociation of the asymmetric ozone

isotopomers. These exit channels do not appear in isotopically enriched experiments (Hathorn and Marcus, 1999,

2000, 2001; Yi Qin and Marcus, 2002). For the source of the mass-independent isotope effect, this leaves the “

effect,” and this may be regarded as symmetry driven (Yi Qin and Marcus, 2002; Yi Qin et al., 2002). An important

advancement has been the utilization of isotopically enriched reactants to determine the explicit rate constants of

13

individual isotopomeric reactions. The relative rate for 16 O + 16 O 16 O is observed to be slower by a factor 50% than reaction of 16 O + 18 O 18

O, consistent with models(Anderson et al., 1997; Gao and Marcus, 2001) that predict a greater

rate for the asymmetric species. Incorporation of the “ 

effect” quantitatively accounts for the difference in

magnitude for the different measured rate constants. Further experiments(Clayton et al., 2005; Janssen et al., 1999;

Janssen et al., 2001; Mauersberger et al., 1999)

reported the rate differences for several isotopically substituted reactions and have resolved the effect of pressure and third-body composition on the isotopic fractionation step

(Guenther et al., 1999; Krot et al., 2005b) Mass independent isotope effects in the formation of ozone on a quantum

basis have been modeled (Babikov et al., 2003b; Babikov et al., 2003c) and the role of unusual, non statistical,

metastable ozone states formed subsequent to the oxygen atom-molecule collision calculated. In this approach the metastable states are treated as independent species and their stabilization steps are treated separately. The buildup in the population of the states and their ultimate leakage through the different stabilization exit channels are computationally followed. It is found that the stabilization process and the density of these states below the

ZPE are highly dependent upon the isotopic structure of the ozone metastable state. The actual calculation employs different metastable O

3

*(E i

) individual states and their lifetimes obtained utilizing a full dimensional quantum scattering study and a coupled-channel approach with hyperspherical coordinates. Ab initio techniques were

employed to determine precise potential energy surfaces (Babikov et al., 2003a, b) and fractionation factors of the

proper sign and approximate magnitude for the isotopomeric formation rates were calculated from this model. Later, treatment of the intermediate metastable ozone states was simplified, reducing potential energy surfaces to two dimensions while maintaining the ability to account for the anomalous isotopic behavior associated with ozone

formation (Jiang and Babikov, 2009). These treatments (Babikov et al., 2003b; Babikov et al., 2003c; Jiang and

Babikov, 2009) differ significantly from that of Marcus and co-workers (Gao and Marcus, 2007; Hathorn and

Marcus, 2000, 2001; Marcus, 2001, 2002, 2004; Yi Qin and Marcus, 2001) and at present the subject continues to

receive theoretical and experimental consideration. For example, the temperature dependence for the individual isotopomeric reactions needs to be determined to treat the details of the stabilization process at the highest resolution.

The general issue of mass independent chemistry is a rapidly expanding area and there are many

experiments that address aspects of the theoretical development. Laser induced fluorescence was utilized (Michalski et al., 2005b; Michalski et al., 2004b) to examine dissociation energies of the isotopologues of NO

2

and evaluate their zero point energies and the effect of symmetry. Isotopic fractionation during photodissociation of ozone has

been investigated in the laboratory (Bhattacharya and Thiemens, 1988; Chakraborty and Bhattacharya, 2003b; Wen and Thiemens, 1991) and few semi analytical models of the isotopic fractionation of ozone as well as N

2

O has been

developed (Liang et al., 2004; Morgan et al., 2004). Further measurements and models are needed in order to

provide the best interpretation of atmospheric ozone isotopic analysis. As will be discussed, atmospheric ozone serves as the initiator in a number of atmospheric reactions and its use as a tracer relies upon firm understanding of its formation process.

14

4.06.2 Atmospheric Observations of Mass Independent Isotopic Compositions

There is extensive observational basis of mass-independent isotopic compositions in nature and especially in the atmosphere of earth. When the first laboratory measurements of the mass-independent isotope effect were reported

(Thiemens and Heidenreich, 1983), their occurrence in nature was not expected, except for the meteoritic CAI data.

It is now known that most oxygen-bearing molecules in the atmosphere (except water) possess mass-independent

isotopic compositions (Thiemens, 2006). These molecules include O

2

, O

3

, CO

2

, CO, N

2

O, H

2

O

2

, ClO

4

, and aerosol carbonate, nitrate and sulfate. Figure 8 displays the mass independent isotopic composition of most of the molecules that have mass independent isotopic compositions associated with them.

Figure 8. Three isotope plot depicting mass independent oxygen isotopic compositions of various oxygen carrying molecules in terrestrial atmosphere.

Mass-independent sulfur isotopic compositions are also observed in aerosol (solid) sulfates and nitrates and sulfide and sulfate minerals from the Precambrian, Miocene and present day aerosols, volcanic sulfates, Antarctica dry valley sulfates, Namibian Gypretes, and Chilean nitrates. In addition, Martian (SNC meteorites) carbonates and sulfates possess both mass-independent sulfur and oxygen isotopic compositions. In each example cited above, insights into terrestrial atmospheric or Martian cycles there are features that could not have been detected by concentration or single isotope ratio measurements

4.06.2.1 Stratospheric and Tropospheric Ozone

The importance of ozone in the Earth’s atmosphere is well established. Its presence in the stratosphere serves as a shield of UV light, vital for sustenance of life, particularly land-based. Tropospheric ozone serves as a source of electronically excited atomic oxygen O ( 1 D), which occurs via:

15

 

H O

2

2OH

The reaction is responsible for creation of tropospheric hydroxyl radicals (OH) which serves as the lifetime regulator of most reduced molecular species in the troposphere, and consequently may be regarded, in part, as a controlling agent of the oxidative capacity of the atmosphere.

A large enrichment of 18 O in stratospheric ozone were observed using in situ mass spectrometry

(Mauersberger, 1981). Only

18 O was measured and it was not recognized that ozone was mass independently

fractionated until after the laboratory experiments (Thiemens and Heidenreich, 1983). Later atmospheric

measurements demonstrated that stratospheric ozone was mass-independently fractionated for the first time though the magnitude of the

18

O was significantly less than the original measurements (Mauersberger, 1987). Return ozone

isotopic analysis further demonstrated that stratospheric ozone possessed an isotopic composition consistent with

laboratory observations (Schueler et al., 1990). The isotopic composition of ozone collected from four balloon

flights in the 22–33 km range and the details of the sampling device and procedure for collection have been

reported (Krankowsky et al., 2000; Stehr et al., 1996).

The Smithsonian astrophysical far-infrared spectrometer was employed on seven balloon flights to independently measure stratospheric ozone isotopic compositions and it was reported that they are consistent with other in situ ,

and return mass spectrometric and laboratory observations (Johnson et al., 2000). These measurements provide

structural information, namely, the asymmetric to symmetric ratio: 50 O

3

(or 18 O 16 O 16 ) / 49 O

3

( 17 O 16 O 16 O). The IR instrument (FIRS-2) is a remote sensing Fourier transform spectrometer that detects molecular thermal emission in the atmosphere. The spectrometer resolved wavelength at 0.004 cm

–1

, permitting resolution of 16 O 16 O 16 O,

17 O 16 O 16 O, 16 O 17 O 16 O, 18 O 16 O 16 O, and 16 O 18 O 16 O species. Over the altitudinal range 25–35 km, the average enhancements for symmetric, asymmetric, and total 50 O

3

are 61 ± 18‰, 122 ± 10‰, and 102 ± 9‰, respectively.

The corresponding 17 O enrichment ( 49 O

3

) values are 16 ± 76‰, 80 ± 52‰, and 73 ± 43‰, respectively, in agreement with measurements by other techniques. A solar occultation spectrometer was also utilized to provide vertical profiles for both 16 O 18 O 16 O and 16 O 16 O 18 O and reported a significant difference between symmetric and asymmetric stratospheric species. The average enhancement for the asymmetric over symmetric species is a factor

of ~1.7, consistent with other filed and laboratory measurements (Irion et al., 1996; Janssen et al., 1999, 2003;

Thiemens and Heidenreich, 1983)

Tropospheric ozone measurements from various locations, rural, urban, and maritime dominated regions

(Johnston and Thiemens, 1997; Krankowsky et al., 1995), though difficult to perform also revealed equal

enrichment for both 17 O and 18 O (70–90‰ with respect to air molecular oxygen). The tropospheric observations have shown that these measurements may provide a mechanism to resolve the complex tropospheric ozone cycle, particularly the complex NO x

interactions. Recent measurements of tropospheric sulfate and nitrate aerosols demonstrate that the complexities of the NO x

–SO x

–O

3

cycle (chemical transformation and transport) are significantly better understood from

17 O,

18 O measurements and that has opened up the possibility that the isotopic composition of polar ice nitrate and sulfate could provide a unique means by which paleo-ozone and

oxidative levels may be quantified (Deshler et al., 2003; Johnson et al., 2001), and will be discussed.

Ozone isotopic

16

measurements have been reviewed which indicated heavy-isotope enrichment extends between 70‰ and 110‰ for

stratosphere, somewhat greater than the reported tropospheric range (Krankowsky et al., 2007; Mauersberger et al.,

2005; Thiemens, 2006). In conclusion, with respect to atmospheric ozone isotopes, there have been significant

advancements in measurement techniques though details of global ozone vertical and seasonal isotopic variability is needed. The ozone enrichment process cascades through other atmospheric molecular species and this unique feature of the ozone isotopic signatures makes it an ideal tracer of numerous atmospheric physical, chemical, and photochemical processes and will be presented in the ensuing sections.

4.06.2.2 Stratospheric Carbon Dioxide

Prior to 1991, numerous isotopic measurements of stratospheric carbon dioxide were made, though only for

18 O and

13 C. A mass-independent oxygen isotopic composition was first observed in CO

2

collected from 26 km to

35.5 km (Thiemens et al., 1991). It has been established that the mass independence arises from isotopic exchange

between electronically excited atomic oxygen and CO

2

:

Q( 1 D) + CO

2

↔ CO

2

Q* ↔ COQ+ O( 3 P)

Here Q denotes heavy isotope of ozone. In this process, the 17 O, 18 O enrichment of atomic oxygen, inherited from stratospheric ozone is passed on to CO

2

via exchange with excited oxygen atom (Thiemens et al., 1991; Thiemens et

al., 1995; Yung et al., 1991; Yung et al., 1997). A key facet of the isotope exchange process is that it proceeds via a

short-lived transition state,

*

CO .

and this process has been studied extensively in various laboratories (Chakraborty

3

and Bhattacharya, 2003a; Johnston et al., 2000; Shaheen et al., 2007; Wen and Thiemens, 1993). Photochemical

isotope equilibrium experiments with a mixture of CO

2

-O

3

and CO

2

-O

2

in the presence of UV light indicated transfer of ozone isotopic anomaly from O

3

to CO

2

depends on the pressure and O

2

/CO

2

ratios (Shaheen et al.,

2007). These laboratory experiments yielded

17 O/

18 O ~ 1 unlike observation with

17 O/

18 O ranging from 1.5 to

1.7 (Lammerzahl et al., 2002; Thiemens et al., 1995). Cross molecular beam studies with labeled C

16 O

2 and O

3 provided first experimental evidence that isotope exchange can occur through a long-lived CO *

3

intermediate

without subsequent crossing to the triplet surface (Perri et al., 2004; Van Wyngarden et al., 2004). On the basis of

existing observations and models, it is still not possible to totally account for all of the important aspects of the exchange at present though the major features are understood.

Laboratory and field observations have provided a new, quantitative measure of atomic oxygen, a driving species of upper atmospheric chemistry. Methane and nitrous oxide are both primarily removed from the stratosphere by reaction with O ( 1 D). Recent measurements of carbon dioxide from the Tibetan plateau have shown that there is large isotopic heavy isotope enrichment in CO

2

which appears to coincide with stratospheric down

welling (Liang et al., 2008b). Though all three oxygen isotopes were not measured, the finding is consistent with the

modeled stratosphere-troposphere exchange using the O-isotope anomaly in CO

2

and predicted that at steady state, a mass independent tropospheric CO

2

component should be observable (Hoag et al., 2005). If verified by triple oxygen

isotope measurements it would provide an independent measure of carbon reservoir interactions and their time scales, an independent means by which the kinetics of biogeochemical cycle interactions may be obtained.

17

Other measurements have shown that CO

2

triple isotope measurements are a new tracer of stratosphere– troposphere mixing. Stratospheric CO

2

is mass-independently fractionated due to its coupling with ozone and at higher altitudes, Lyman-

photolysis of O

2

(Liang et al., 2004) that provide another source of isotopically

anomalous oxygen atoms. Tropospheric CO

2

is mass dependent because of its equilibrium with water (Luz and

Barkan, 2000) thus appearance of any mass independent component in the troposphere is stratospheric (or

mesospheric) in origin. Oxygen triple isotopic composition of the stratospheric CO

2

and concentrations of SF

6

,

CCl

3

F (CFC-11), CCl

2

F

2

(CFC-12) and Cl

2

FCClF

2

(CFC-113) were measured within the Arctic polar vortex in an

altitude range extending from the 12 to 21km range were measured (Alexander et al., 2001). An important new

aspect is that a CO

2

oxygen isotopic anomaly correlation with SF

6

number density was observed. Sulfur hexafluoride has a lifetime in excess of a thousand years and is only removed in the mesosphere and above by highly energetic processes such as photodecomposition and/or electron attachment. This long lifetime renders SF

6

as an ideal conservative tracer of stratospheric air mass movement. The observation of an inverse relation between the magnitude of mass independence and SF

6

concentration provides a measure of air mass age plus the integrated odd oxygen chemical activity. Combined measurement of fluorocarbons and CO

2

isotopes have been measured on samples acquired by the NASA ER-2 aircraft in the lower stratosphere and measured for 17 ΔO of CO

2

and N

2

O mixing ratios. The measurements display a tight correlation between the 17 O isotopic anomaly in carbon dioxide and

nitrous oxide mixing ratio facilitating a precise value for stratosphere-troposphere mixing flux (Boering et al., 2004).

Such measures are important parameters for determination of the gross carbon exchange between the atmospheric reservoirs.

Simultaneous Measurements of CO

2

and O

3

isotope ratios from eight balloon flights from Kiruna, Sweden and Aire-sur-l’Adour, France provided the first simultaneous correlation between

17 O and

18 O of stratospheric

CO

2

and ozone (Mauersberger et al., 2003). The observed

17 O/

18 O ratio in carbon dioxide is 1.71 ± 0.03, independent of altitudinal range. The observations include both CO

2

and O

3

isotopes and aid in providing a quantitative model for the isotopic relation between stratospheric CO

2

and O

3

. In the stratosphere the oxygen isotopic composition of the CO

2

can be understood by the exchange of O 1 D from ozone photolysis. Higher in the atmosphere, photolysis of 16 O 17 O and 16 O 18 O by solar Lyman-

becomes a significant source of heavy isotopic enrichments in CO

2

not accounted for in previous models (Liang et al., 2007; Liang et al., 2008a). Laboratory and

mesospheric measurements are needed to further resolve this model and to amplify understanding of upper atmospheric oxygen chemistry

The multi oxygen isotopic measurements have been shown to be a powerful tool in atmospheric chemistry and there is a major need for global measurements, including observation of temporal and altitudinal variations. A significant limitation is the actual measurement techniques which are difficult and time consumptive. Original and

still utilized method involves a fluorination technique (Thiemens et al., 1991) that quantitatively converts CO

2

to

CF

4

and O

2.

In this technique CO

2

is reacted with BrF

5

at 800 °C for 48 h in nickel tubes that catalyze the reaction of fluorine with CO

2

. At the termination of the reaction, O

2

is separated from CF

4

by gas chromatographic techniques to remove trace amounts of CF

4

that produce error in the

17 O measurements. Since the level of precision required is at the 0.1 per mil level, the purification step is vital. Conversion to O

2

is required because measurement of CO

2

has

18

isobaric and uncorrectable interferences from the contribution at mass 45 ( 13 C 16 O 16 O or 12 C 16 O 17 O) from 13 C in carbon dioxide. A second technique converts CO

2

to methane and water, with subsequent chemical conversion of water to HF and O

2

by reaction with fluorine (Brenninkmeijer and Rockmann, 1998). Another technique based

solely on CO

2

isotope ratio mass spectrometry has been developed to measure oxygen isotopic anomaly (Assonov and Brenninkmeijer, 2001) with high precision. Following initial

13 C,

18 O measurement, CO

2

is exchanged with solid CeO

2

thus exchanging the oxygen isotopes with the adsorbed oxygen on the CeO

2

while carbon isotopic composition remaining constant. Since the 13 C/ 12 C ratio does not alter, the

17 O may be determined by calculation of the difference in the two measurements. This technique has several advantages: (i) the mass 44, 45, 46 ratios are commonly measured using standard isotope ratio mass spectrometers; (ii) it is fast and safe, not requiring Br

F

5. In all measurement protocols, the CO

2

isotopic measurements are less time consuming compared to single isotope ratio measurements.

The use of multi oxygen isotope ratio measurements of atmospheric CO

2

has provided higher resolution details of a variety of process, including stratosphere-troposphere mixing, stratospheric-mesospheric oxidative chemical processes and, is also linked to its interaction with the molecular oxygen cycle. Future atmospheric and laboratory measurements along with modeling efforts will provide greater resolution into these atmospheric phenomena and as will be discussed, of gross primary productivity measurements and Martian atmospheric processes.

4.06.3 Atmospheric Aerosol Sulfate: Present Earth’s Atmosphere

The importance of sulfur in the Earth's atmosphere and environment is well established. Aerosol sulfate is known to alter atmospheric radiative processes due to its role in increasing the Earth’s albedo and as a cloud condensation nucleus (CCN) and is a component of the sulfur cycle and environmental acidification. While these important roles are recognized, there remain significant gaps in understanding the role of aerosols in these processes.

An important role of isotopic analysis is for source identification, transport and mechanistic definition of in situ chemical transformational processes. Sulfate and nitrate aerosols are known agents in increasing the incidence of cardiovascular disease and definition of surface characteristics may be of relevance to such studies but, are notoriously difficult to measure. Nitrate is an agent in alteration of terrestrial biodiversity when sufficient quantities are added to soils and streams. Apportionment of nitrate sources is difficult because of the multitude of vectors; consequently, new techniques to aid in source recognition are welcome. Single isotope ratio measurements have been of limited utility in addressing these issues in the past due to limited specificity of single isotope ratio measurements such as

15 N,

34 S, or

18 O.

It is known that 70–80% of atmospheric sulfur species in the northern hemisphere are anthropogenic

(Rasch et al., 2000). A gap in strict quantification of the atmospheric sulfur cycle partially arises from the inability

to adequately define the oxidation processes that govern sulfur dioxide oxidation and chemical transformational

chemistry (Kasibhatla et al., 1997; Lelieveld et al., 1997). Model calculations underestimate the observed sulfate

concentrations in northern latitudes, and it is plausible that heterogeneous chemical reactions are of importance, but inadequately recognized. Measurements of oxygen and sulfur mass-independent isotopic compositions of sulfate

19

aerosols have been developed to provide independent means to study issues of atmospheric sulfur. Mass-

independent isotopic compositions are observed in atmospheric aerosol sulfate (Lee et al., 2002; Lee et al., 2001;

Lee and Thiemens, 2001; Savarino et al., 2003). The measurements of the

17 O,

18 O isotopic composition of aerosol sulfate provide a new means to identify homogeneous vs. heterogeneous oxidative pathways, especially coupled with isotopic measurements of atmospheric H

2

O

2

and ozone (Alexander et al., 2004a; Alexander et al.,

2002; Alexander et al., 2003; Savarino and Thiemens, 1999a, b). Characterization of reaction rate constants, pH reaction rate dependence, and reaction isotopic fractionation factors (Savarino et al., 2000) allow the homogenous

and heterogeneous reaction pathways to be quantified (purely gas phase vs. liquid phase oxidation). The foundation for the model is the recognition that the exclusive gas phase oxygenation pathway for SO

2

oxidation is by reaction with OH radical, known to be strictly mass dependent. Liquid phase oxidation proceeds via reaction with ozone and hydrogen peroxide, both mass independent. If the isotopic composition of oxidizing molecules are known and associated atmospheric variables such as pH, temperature and ambient concentration of the oxidants, the relative

oxidation pathways may be accurately quantified from isotopic measurements and models (Lee et al., 2001;

Savarino et al., 2000). The precise attribution of these oxidative pathways is important for understanding particle

formation processes and transportation. Advanced modeling efforts (Alexander et al., 2005; Kunasek et al., 2010)

using GEOS-CHEM global models have permitted more specific reaction features to be identified, such as the role of aerosol particle surfaces above the Indian Ocean. This work identified a new aerosol generation process previously not considered in aerosol models. Using multi oxygen isotopic measurements of sulfate collected at

Trinidad Head, California, specific reaction characteristics, such as ozone driven oxidation of SO

2

in sea-spray and

small particle coagulatory processes and specific sources, such as China were identified (Patris et al., 2007).

Measurements of

17 O,

18 O, and

34 S of atmospheric sulfates in Baton Rouge, LA for a near two year period have

shown the ability to resolve secondary aerosol sulfate (SAS) formation processes (Davis et al., 2001) . Contrary to

coastal California, the Baton Rouge region is dominated by H

2

O

2 oxidation rather than OH. The work confirmed model predictions for atmospheric chemical reactions among liquid water, gas, and the role of temperature, pH and aerosols in a region such as Baton Rouge.

The use of multi oxygen isotopes has been applied towards reaction resolution in chemically anomalous regions.

Measurements of sulfate collected in Alert, Canada in the Arctic have shown that unlike other geographic regions studied, Arctic oxidation processes are dominated by SO

2

(S(IV)) reaction with O

2

transition metal catalysis, rather than OH and H

2

O

2

(Alexander et al., 2009). A global chemical network transport model was used to interpret the

data and quantify the role of Fe (III) and Mn (II) catalyzed oxidation of S (IV) by O

2

. The inclusion of these reactions and isotope measurements revealed that the solubility and oxidation state of the metals is set by cloud water content, chemical source and sunlight. Inclusion of the metals characteristic of the Arctic in the sampling time period was able to account for the isotope measurements and define the unusual role of O

2

catalyzed reactions in the

Arctic.

Besides model considerations of oxidative mechanisms, single source characterization has been achievable.

Both sulfur and oxygen isotopic measurements of sulfates produced in controlled combustion processes provided

new means to characterize individual source and oxidation mechanism of sulfates in the atmosphere (Lee et al.,

20

2002). These experimental observations were conducted at the Centre de Recherches Atmospheriques in

Lannemezan, France in a dark combustion chamber to isotopically characterize aerosol and gas emissions from fuel combustion and emission. The chamber has a volume of ~160 m 3 , which effectively minimizes wall effects and maximizes mixing. The collections were done in the dark to reduce potential photochemical effects. Fossil fuel and vegetation burns were performed, including Savanna grasses, Lamto grass, rice grass, hay, diesel fuel, and charcoal.

All of these processes produce sulfur and oxygen isotopic compositions that are strictly mass dependent Recent atmospheric oxygen isotopic sulfate measurements at a coastal La Jolla, California location and were able to define

contributions from maritime shipping; the largest poorly regulated global sulfur source (Dominguez et al., 2008). In

this work the ship point sources of primary ship sulfate contributions were measured and revealed it to be mass dependent and equal to tropospheric O

2

(

18 O of ~23 per mil). Based upon the precise three isotope source characterization and back trajectory calculations it was demonstrated that ships contribute between 10% and 44% of the non-sea-salt sulfate found in the fine diameter (< 1.5 micron) particulate material in the southern coastal

California region. As concluded, these findings are relevant for public health issues, especially with the predicted increase in ship traffic.

In sum, multi oxygen isotope ratio analysis of present day aerosol sulfate samples have revealed source and chemical mechanistic details in many cases, such as oxidative pathway quantification that are new and quantitative and supplement existing concentration and modeling efforts. Similarly, source apportionment methodologies are also further amplified with the new multi isotopic measurements.

4.06.3.1 Mass-independent Oxygen Isotopic Composition of Non Present Day

Sulfates

It is well established that sulfate oxygen isotopic compositions are stable over long time periods due to the extreme non lability of the S-O bond within sulfate preventing secondary isotopic exchange reactions. For example, sulfates extracted from carbonaceous chondritic meteorites preserve the sulfur and oxygen isotopic record of

hydrothermal events on the meteorite parent body for multi-billion year time scales (Airieau et al., 2005; Farquhar et al., 2000b; Gao and Thiemens, 1993a, b). As a result of this preservation ability it is possible to extract sulfate from

ice core samples to document the change of atmospheric processes, especially the alteration of oxidants over time such as O

3

, OH, H

2

O

2

using model and measurements of

17

O anomaly (Alexander et al., 2002; Alexander et al.,

2003; Savarino et al., 2003). A specific example of the ability of the technique was provided by Alexander et al

(2004) who measured the past three centuries of sulfate from samples obtained from Greenland. In this record, the effect of elevated biomass burning in the Northern American continent during the industrial revolution time period of the late 19 th early 20 th century was isotopically characterized.

17 O values of sulfate from Greenland and

Antarctica reveal that OH gas phase oxidation dominated over O

3

, H

2

O

2

liquid phase reactions during the last glacial

than during the Eemian and pre-industrial Holocene (Alexander et al., 2003) . The primary cause is the reduction of

water vapor content during glacial time periods, though other parameters must be considered, such as perturbations in the OH and O

3

levels as well as NOx, CO, and CH

4 chemistry. Recent

17 O,

34 S,

33 S, and

36 S measurements

21

(Kunasek et al., 2010) of sulfate for the past 230 years of sulfate in a West Antarctic ice core are consistent with pre-

industrial to industrial time period increases of as much as 50% and decreases in OH by ≈ 20% if accompanied by

H

2

O

2

increases of 50%. The work has demonstrated that the isotopic measurements combined with global transport models are capable of high resolution of global atmospheric chemical changes, such as oxidation chemistry.

Measurements of the oxygen isotopic composition of sulfates have provided atmospheric records for time periods

beyond those obtainable from the ice core record (Bao et al., 2000a; Bao et al., 2000b).

17 O anomalies in continental sulfate deposits, desert varnishes and massive (3 × 10 4 km 2 ) gypsum (CaSO

4

·2H

2

O) deposits of the Namibian desert

have been reported by Bao and coworkers ((Bao and Marchant, 2006; Bao et al., 2003a; Bao et al., 2000a; Bao et al.,

2000b; Bao et al., 2001). The source of the oxygen isotopic anomalies in Namibian sulfates was attributed to

atmospheric photo-oxidation of dimethyl sulfide from nearby oceanic upwelling and inland transport. The measurements have established a new paleo-oceanic record of biologic activity variation on tens of millions of year time scales. The oxygen isotopic composition of sulfates extracted from vertical profiles of Antarctic dry valley

were utilized to identify their source regions, which has historically been elusive to determine (Bao et al., 2000a).

Variation of the

17 O values reflects the source of the anomalous sulfate to the Dry Valleys. The source, particle size dependence, and relative amounts of sulfate from sea salt aerosols and biogenic sources were characterized with the mass independent compositional measurements. Other applications included defining sulfate sources to the Atacama

desert (Bao, 2005),) detection of Oligocene volcanic processes (Bao et al., 2003b) and chemical description of atmospheric rain depositional processes (Bao and Reheis, 2003).

A most striking observation was that of the first negative

17 O values, observed in evaporates and barites

over the past 750 million years (Bao et al., 2008). Based upon the negative isotopic anomalies shown in Figure 9, it

was proposed that these anomalies track the isotopic composition of atmospheric oxygen.

Figure 9. The

17 O of evaporite and barite sulphate over the past 750 million years (adapted from Bao et al., 2008).

22

Based upon the triple isotopic analysis of atmospheric molecular oxygen (Luz and Barkan, 2000) and its connection

to the stratospheric ozone-carbon dioxide cycle it was suggested that the negative values displayed in Figure 8 result from dramatic excursions in atmospheric pCO2 levels, which concomitantly depress the molecular oxygen

17 O value, the source of oxidation oxygen to sulfate (along with water derived oxygen). A peak in pCO

2

occurred in the past 750 million years at a time (~650 million years ago) consistent with the melting of the snowball earth and massive CO

2

and methane release. Depending upon model parameters, the CO

2

levels may have exceeded 25,000 ppm. Measurement of sulfate extracted from carbonate lenses within a Neoproterozoic glacial diamictite suite from

Svalbard (635 million year age) also apparently indicates exceptionally high atmospheric carbon dioxide levels or a severely perturbed oxygen cycle. The mass independent observation of sulfate has allowed constraints of the carbon system during this time period to be determined.

Measurement of atmospheric sulfate oxygen isotopic compositions on time scales from the present to the deep past have provided new insights and records of global and local processes that would escape detection by any other measurement. Coupled with modeling efforts, future measurements will develop our recognition of processes including atmospheric sulfur transformation, transport, source identification, changes in the earth’s atmospheric oxidative capacity, and the snowball earth chemistry. Though not discussed in this chapter, sulfate oxygen isotopic measurements in meteorites have provided a mechanism by which hydrothermal processes on parent bodies may be

modeled and, the distribution of water in the solar system better elucidated (Airieau et al., 2005; Benedix et al.,

2003; Farquhar et al., 2000b).

4.06.4 Atmospheric Mass Independent Molecular Oxygen

It has been known for decades that photosynthesis and respiration predominantly establish the steady-state levels of O

2

and CO

2

in the Earth’s atmosphere. From this standpoint, it might be argued that these processes are among the most important of all global biogeochemical cycles and precise establishment of the rates of photosynthetic and primary productivity in the world's oceans is essential to full characterization of global geochemical cycling. This rate is intimately linked to the primary productivity of the oceans, carbon cycling, and

CO

2

levels in the ocean and atmosphere (Battle et al., 2000; Bender, 2003). Quantification of this linkage is difficult

due to the magnitude of the oxygen reservoir differences and in the lifetimes of different reservoirs. The detection of an atmospheric O

2

17

 anomaly

 provided a new and direct technique to evaluate global primary productivity

(Blunier et al., 2002; Luz et al., 1999). This new technique is based upon isotopic material balance constraints on the

creation of anomalous oxygen isotopic reservoirs in ozone, CO

2

, and O

2

. This negative anomaly in molecular oxygen arises due to the isotope exchange between CO

2

and O

3

via excited oxygen atom in the stratosphere thus enriching stratospheric CO

2

with concomitant depletion of heavy isotopes from the atmospheric oxygen. The magnitude of the O-isotopic anomaly is in the per mil level in stratospheric CO

2

(~400 ppm), but is not quantifiable immediately for O

2

(~21%) as a steady-state effect on the

17 O,

18 O of atmospheric O

2

is not measurable because of the large difference in reservoir sizes and extended lifetime of O

2

compared to CO

2

(a factor of hundreds). The Oisotopic anomaly of CO

2 is lost by photosynthesis and respiration in the upper ocean and leads to an accrual of a

23

small and measurable negative isotopic anomaly in O

2 over extended time period The magnitude of the anomaly is a direct measure of global primary productivity and measurement of the

17 O composition in oceanic O

2

vertical profiles provides a new, quantitative means by which primary productivity may be evaluated. The rate the O

2 anomaly is removed is dependent upon photosynthetic and respiration activity in the oceans and its rate of disappearance is thereby quantitatively linked to primary productivity. Measurement of

17 O,

18 O of O

2

trapped in polar ice is also used as a means of determining global biospheric primary productivity over time. Oxygen triple isotopic measurements of dissolved oxygen in sea water allowed the estimation of integrated oceanic productivity

on a time scale of weeks (Luz and Barkan, 2000). The oxygen isotopic composition of water on earth is controlled

by two major processes, liquid vapor equilibrium isotope fractionation and diffusivity of water vapor into air, a kinetic isotope fractionation. The abundance of 17 O in precipitation has been assumed to carry no additional information to that of 18 O. In contrast to the deuterium excess, the 17 Δ of precipitation is controlled primarily by kinetic effects during evaporation of the initial vapor and is dependent on the temperature at the evaporation (and condensation) site. This makes 17 Δ a unique tracer that complements 18 O and deuterium, and may allow for a decoupling of changes in the temperature of the ocean, that serves as the vapor source, from changes in the relative

humidity above it (Angert et al., 2004).

The 18 O enhancement of atmospheric O

2

is utilized to determine paleo-variations in the relative O

2

contributions of marine and terrestrial sources (Angert et al., 2001; Angert et al., 2003b). Quantification of the

oxygen cycle is mandatory to resolve the tightly coupled carbon cycle and to quantitatively evaluate the components

of the global biogeochemical cycle of oxygen (Angert et al., 2003a; Angert et al., 2001; Angert et al., 2003b).

Global primary productivity has been modeled using

values for the most important processes (photosynthesis and soil respiration) that determine the isotopic composition of molecular oxygen:

  ln(

17

) ln(

18

)

which is a previously defined relation (Mook, 1971; Mook and Vogel, 1968). In this equation,

17

and 18

are single stage isotopic fractionation factors for photosynthetic processes. The term expresses the contribution of the various processes that effect the isotopic composition of atmospheric O

2

. The value of

has been determined for the dark respiration factor, i.e., the cytochrome pathway (0.516 ± 0.001) and its mechanistic alternative (0.514 ± 0.001)

(Angert et al., 2003a). A modestly higher value for diffusion in air (0.521 ± 0.001) was also measured. The precise

determination of these

-factors leads to the ability to determine the steady-state value for diffusion and respiration processes and their contributions to atmospheric O

2 levels. The isotopic composition of O

2

is established primarily by the relative rates of photorespiration and dark respiration. This conclusion was further strengthened by examining its relation to changes in ice core

17

O measurements (Blunier et al., 2002). This work modeled oxygen changes to

understand the variability in primary productivity for the last glacial, interglacial, early Holocene and demonstrated the utility of the high precision multi oxygen isotope approach in resolving the global oxygen biogeochemical cycle.

24

4.06.5

The Atmospheric Aerosol Nitrate and the Nitrogen Cycle

The nitrogen cycle is a cornerstone of all living processes extending from the air we breathe to the food we eat and is an essential participant in synthesis of amino acids and nucleic acids. Anthropogenic activities have significantly altered the nitrogen cycle by addition of nitrates in various chemical forms especially fertilizers and fossil fuel burning. The nitrogen footprint also expresses its importance in a number of vital biogeochemical cycles. These include, for example its role as a major sink for nearly all NO x

species, which produces its impact as an acid agent in the environment. Nitrate also participates as a heterogeneous reactant in polar stratosphere cloud (PSC) chemistry and associated with the destruction of ozone in the Antarctic polar vortex. Nitrogen is a source of fixed nitrogen to key ecosystems, as photosynthetic CO

2

and N

2

fixation are fundamentally important in mediating production

dynamics of intertidal and sub tidal marine microbial mat communities (Paerl, 1993a; Paerl, 1993b). The

recognition that the terrestrial nitrogen footprint needs to be better understood has led to the development of new methodologies to study its behavior, including multi isotopic measurements. Some of the facets requiring further resolution includes i) depositional rates; ii) chemical tropospheric chemical transformation mechanisms; iii) source variability and quantitative identification; and iv) significance of long range transport of nitrates .

It has been shown that atmospheric aerosol nitrate possess one of the largest mass-independent oxygen isotopic

compositions (Fig. 8) observed in nature (Michalski et al., 2005a; Michalski et al., 2004a; Michalski et al., 2002;

Michalski et al., 2003). An important advancement in atmospheric nitrate studies has been the development of the

ability to perform

17 O,

18 O measurements in aerosol nitrates. In the process of conversion of nitrate to O

2

for mass spectrometer isotope ratio analysis, high purity is essential. Ion chromatography is utilized to simultaneously isolate and concentrate nitrate. The nitrate is subsequently reacted under a proper pH regime to convert to HNO

3

. Reaction with Ag

2

O quantitatively converts nitric acid to AgNO

3

, which is filtered and dried. AgNO

3

is thermally decomposed to O

2

, NO

2

, and Ag at a constant chemical and isotopic branching ratio. Following a final purification process, the O

2

is measured for the

17 O,

18 O composition. A variety of standards have been analyzed, including

USGS-35, a sample of NaNO

3

from the Atacama desert of Chile and IAEA-N3, an internationally distributed standard sample KNO

3

, with a range of reported

18

O values (Revesz et al., 1997). Another method now routinely

employed for measurement of both N and O isotopic composition of nitrate at the natural-abundance level. This procedure is based on the isotopic analysis of nitrous oxide (N

2

O) generated from nitrate by denitrifying bacteria that lack N

2

O-reductase activity (Boehlke et al., 2007; Sigman et al., 2001). However, proper correction methods

with international references (USGS32, USGS34 and USGS35) are needed. As a consequence, it is important to realize that the corrected isotope values are derived from a combination of several other measurements with

associated uncertainties (Casciotti et al., 2007; Chmura et al., 2009; Xue et al., 2010)

Atmospheric nitrates (HNO

3

+ particulate NO

3

) are predominantly formed through oxidation of NO x

(NO +

NO

2

), which originate from soils, lightening and fossil fuel combustion. The oxidation of NO x

occurs via O

3

, OH,

HO

2

/ RO

2

in the atmosphere and nitrate is subsequently deposited to the surface through wet and dry deposition.

Nitrate possesses an extraordinarily large mass-independent isotopic composition, second only to ozone in

25

magnitude (Michalski et al., 2002; Michalski et al., 2003) due to its steady state equilibrium between O

3

and the

NO x

cycle. The relative significance of these oxidation pathways in the current atmosphere has been investigated using the

17 O of NO

3

since oxidants transfer different

17 O to nitrates (

17 O (O

3

) ~30‰; ( 

17 O (OH) ~0‰; ( 

17 O

(HO

2

) ~ 1‰) in aerosols samples first collected at La Jolla, CA. The 

17 O (NO

3

) has also been used to infer the role

of biomass burning in preindustrial time from Antarctic ice core samples (Alexander et al., 2004b). Laboratory

experiments suggest post depositional nitrate loss from snow packs due to photolysis may affect the

17 O of NO

3

.

However, post depositional renoxification (conversion of HNO

3

back to photochemically active forms that can regenerate ozone such as HONO and NO

2

) and reoxidation involves local oxidants thus producing a mixed signal of locally produced nitrate (~5-10%) and transported NO

3

(McCabe et al., 2005; McCabe et al., 2007). Size resolved

present-day aerosol nitrate isotopic measurements have added additional insight into the resolution of sources and

atmospheric transformation mechanisms (Patris et al., 2007). Using combined multi-isotope ratio measurements and

size-fractionated collection procedures, it is possible to provide mechanistic details of renoxification of nitrate aerosol and their fate in the atmosphere. Oxygen isotope anomalies in atmospheric nitrate has provided a new means to elucidate source and chemical transformation processes to understand nitrate biogeochemical cycles. For example, the massive mass-independent isotopic signature observed in Chilean desert nitrates uniquely reveals atmospherically derived nitrate deposition since all other sources (by measurement) possess mass-dependent isotopic compositions. Resolution of the source of nitrate to these regions has historically been argued and the isotopic measurements have provided a straight forward answer to source regions. In addition, these measurements, coupled with contemporary aerosol nitrate measurements reveal that the oxygen isotopic signatures are stable on

million year timescales, at least in highly dry or extreme cold environments (Michalski et al., 2004a). This is

particularly valuable, as this permits measurement of nitrate in polar ice samples to examine paleo-variations in nitrate and NO x

chemistry (Kunasek et al., 2008). As discussed in the next section, when linked to sulfate oxygen

isotopic measurements, an entirely new mechanism to study paleo-atmospheric oxidative capacity variation is available.

4.06.7 Mass-Independent Oxygen Isotopic Compositions in Solids to Reflect Atmospheric

Change: Earth and Mars

It is argued that the terrestrial planets, especially Mars and Earth had roughly occupied comparable starting

formational components from the parent solar nebula (MacPherson, 2008). Consequently surface conditions may

have been similar on both the planets (Strasdeit, 2010; Taylor, 2010). This has led to supposition that early Mars

had a CO

2

rich atmosphere, therefore, models of Martian climate history includes carbonate precipitation as a potential sink of atmospheric CO

2

. Based on understanding of the link between atmospheric CO

2

levels and greenhouse warming, high levels of CO

2

could support abundant liquid water thus facilitating rock weathering by carbonic acid to release Ca, Mg and Fe ions to water and to promote the precipitation of solid carbonates. However, massive carbonate deposits are missing on Mars as inferred from rovers and orbital based spectroscopic studies

(Bandfield et al., 2003; Bullock and Moore, 2007; Palomba et al., 2009). Like CO

2

, atmospheric SO

2

is another greenhouse gas whose complex geochemical cycle includes dissolution into surface waters, acid weathering of

26

surface rocks and precipitation as a relatively stable solid (Halevy et al., 2007; Johnson et al., 2008; Tian et al.,

2010). Spectroscopic observations indicated the presence of Mg and CaSO

4

, coexisting with coarse grained

crystalline hematite deposits (Christensen et al., 2000; Gendrin et al., 2005a; Gendrin et al., 2005b). By analogy

with terrestrial geology, this suggests the presence of liquid water on the surface of Mars for a significant period of time. On earth, massive carbonate sediments representing its early atmosphere are found in association with

massive deposits of sulfates (Gendrin et al., 2005b; Holland, 1978). The detection of evaporitic minerals and their

isotopic signatures can play a crucial role in understanding the history of the planet (Bandfield, 2002a; Farquhar and

Thiemens, 2000; Smith and Onstott, 2011; Strasdeit, 2010; van Zuilen, 2008). The Martian meteorites such as SNC

(Shergottite-Nakhlite-Chassignite), Zagami and Lafayatte are igneous rocks with trace quantities of secondary minerals. The multi oxygen isotopic composition of the secondary minerals (sulfate and carbonate) in Martian meteorites as well as geochemical observations have provided a basic understanding of the aqueous geochemistry

of Mars (Bandfield, 2002b; Farquhar and Thiemens, 2000; Farquhar et al., 1998; Jull et al., 1995).

In a study of secondary minerals in Martian (SNC) meteorites Thiemens and coworkers (Farquhar and

Thiemens, 2000; Farquhar et al., 1998) demonstrated both carbonates and sulfates possess mass independent oxygen

isotopic compositions (Fig. 10), thought to arise from the interaction of anomalous atmospheric carbon dioxide with water in the Martian regolith. Martian atmospheric CO

2

acquires it isotopic anomaly from the atmospheric O

3

-O

2 cycle, the same as in the Earth’s present atmosphere and are ultimately removed as carbonates. These measurements afford a new means by which Martian atmospheric-surface interactions could be studied. The O-isotopic composition of carbonates, sulfate and water indicate a range of value and do not lie along terrestrial fractionation line passing through the bulk silicate rock, indicating that the Martian meteorites are formed from distinct water reservoirs and are not in equilibrium with the host rock.

27

Fig. 10. Plot of

17 O versus

18 O of SNC meteorite carbonates, sulfates and silicates. Data for water of Karlsson et al., 1992 is included to illustrate the relationship between

17 O of water and carbonates and Farquhar et al

(2000,1998)

An oxygen isotopic anomaly in terrestrial atmospheric carbonates has been observed recently (Shaheen et al., 2010). Aerosol samples were collected at Scripps pier, La Jolla, California and the oxygen isotopic composition

of the CO

2

released upon acid digestion was measured using reaction with BrF

5

(Farquhar et al., 1998). Control

experiments were conducted to ensure CO

2

released is exclusively from inorganic carbonates. Excess 17 O (

17 O) in atmospheric carbonates ranged from 0.4 to 3.9‰ (Fig. 11) as compared to soil carbonates (

17 O ≈ 0) which also make up a significant fraction of normal aerosol carbonate. These observations indicate heterogeneous chemical transformation occurs on aerosol surfaces and can preserve isotopic signature of the trace gases that interact with particle surfaces.

Fig. 11. Oxygen isotopic anomaly in terrestrial atmospheric carbonates. For comparison soil carbonates are shown.

 ’ i O = 1000 ln (1+

 i O/1000), i O = 17 O and 18 O. 1

SD = 0.1 ‰.

(data and Figure adapted from Shaheen et al,

2010)

Laboratory experiments have been performed to resolve and quantify the mechanism for the production of the observed isotopically anomalous carbonates. It was shown that transfer of the oxygen isotopic anomaly of ozone to carbonates occurs via two different, previously unrecognized pathways (Fig. 12). Mechanism A involves oxygen isotopic exchange reactions on pre-existing carbonates whereas mechanism B involves in-situ formation of

28

carbonates on mineral particles. In both cases, the interaction of ozone with surface adsorbed water is required to produce the carbonate O-isotopic anomaly. Mechanism (A) initiates with the dissociative adsorption reaction of water on mineral surfaces and the formation of a surface complex (M(OH)(HCO

3

)) sc

. The presence of carbonic acid has been confirmed in numerous other studies on CaCO

3

surfaces in the presence of H

2

O vapor (Al-Hosney et al.,

2005; Al-Hosney and Grassian, 2005; Karagulian et al., 2006). The uptake of ozone at the hydrated calcite surface

is a multi- step process consisting of i) gas phase diffusion of the O

3

molecules to the liquid interface, ii) transfer across the interface, iii) diffusion and reaction with water in the condensed phase, and iv) diffusion out and desorption of the residual reaction products. Isotopically anomalous hydrogen peroxide formation via dissociation of

O

3

transfers the isotopic signature in the hydration layer which in turn transfers the isotopic anomaly to the

carbonates (Shaheen et al., 2010) as shown in Fig. 12. Hydrogen peroxide produced on aerosol surfaces was

measured and O-isotopic anomaly was identified for the first time in surface adsorbed peroxide. Mechanism B) involves in-situ formation of carbonates in the presence of CO

2

and water. The oxygen-bearing mineral aerosols

(CaO, MgO, Fe

2

O

3

, ZnO, SiO

2

, and Cu

2

O) are derived from crustal materials, sea spray and anthropogenic sources and may serve as nucleating sites for in-situ carbonate formation. It is known that alkaline metals in solution enhance the uptake of CO

2

. Similarly, the presence of basic oxide and hydroxide aerosols facilitates the interaction between chemisorbed water and CO

2

and promotes in-situ formation of bicarbonates. The surface layer upon drying becomes supersaturated, leading to secondary carbonate formation on mineral surfaces via heterogeneous chemistry.

Fig. 11. The molecular mechanisms proposed to explain the origin of the oxygen isotopic anomaly in atmospheric carbonates. A). Ozone isotope exchange on existing carbonate aerosols with dissociative adsorption of water and peroxide formation. B). In-situ formation of carbonates and interaction with ozone on particle surfaces. The red circles indicate ozone isotope exchange reaction and probable hydrogen peroxide formation sites. The abbreviation on the side bar are S= solid (MO and MCO

3

) such as CaO, MgO and Fe

2

O

3

, CaCO

3

, MgCO

3

, L= liquid or adsorbed water film, G = gas phase (adapted from Shaheen et al., 2010).

Mars exploration for extinct life is associated with location of bio signatures in secondary minerals, however, simple morphology of microfossils can be mimicked by abiological mineral structures, rendering it difficult to distinguish between true microbial fossils and microscopic pseudofossils. The multi oxygen isotopic values can be a valuable tool to discriminate carbonate fossils from their abiotic counterparts. By analogy with

29

terrestrial carbonate minerals, the O-isotopic signature of fossilized carbonate minerals should follow a standard mass dependent fractionation process (

17 O ≈ 0) if the water is known. The presence of an oxygen isotopic anomaly

(

17 O = 0.4-0.7 ‰) in carbonate minerals in Martian meteorites indicate otherwise, especially since it does not equilibrate with Martian water (Fig. 10). Recent laboratory experiments indicate that heterogeneous chemical transformations on carbonate minerals in the presence of ozone produce excess 17 O in carbonates in the presence of adsorbed water. The reaction intermediate hydrogen peroxide also shown to possesses oxygen isotope anomaly.

These observations collectively suggest that future measurements could be another means by which water levels on

Mars might be quantified. Stepwise extraction of water and hydrogen peroxide with concomitant oxygen isotopic analysis of carbonates and water may help to distinguish surface transformation from biotic and abiotic signatures of carbonate minerals and resolve water levels.

Sulfur is a particularly interesting element on the Martian surface because it is abundant, exists in the atmosphere and regolith, but has a poorly described origin and evolution. Establishing the connection between the reservoirs as well as determining the mechanism of transfer is of interest in the evaluation of potential life on Mars and other planets. It is observed that the SNC meteorites possess a mass independent sulfur isotopic composition and experimental by photochemical studies suggest that the anomalous isotopic compositions are synthesized by UV

photolysis of sulfur dioxide, which was emitted by sporadic volcanogenic activities (Farquhar et al., 2000b;

Farquhar and Thiemens, 2000; Farquhar and Wing, 2003). To produce the sulfur isotope anomalies, SO

2 photolysis must occur at short wavelengths (186-251 nm) producing an excited electronic state of SO

2

(SO

2

*) which subsequently undergoes a reaction sequence leading to production of a stable sulfate product. Present SNC data rule

out hydrothermal oxidative effects (Farquhar et al., 2007a) . The relation of the sulfate

33 S and

36 S apparently requires photochemistry as the origin and definition of the wavelength dependent source of the photolytic isotope effect can allow better understanding of Martian surficial processes. The SNC observations are significant though, in that they provide a clear mechanistic link between the Martian atmosphere and the surface, however, with the availability of an expanding data base of in-situ Martian mineralogical composition and multi isotopic measurements, these results may aid in interpretation of both surface and subsurface environments.

4.06.8

Sulfur in the Earth’s Earliest Atmosphere: The Rise of Oxygen

In the present day terrestrial atmosphere, UV photochemistry of SO

2

does not occur due to removal of the required spectral UV component t by stratospheric ozone absorption and its short tropospheric life time restricting admixture to the stratosphere. Mass independent sulfur isotopic compositions of sulfide and sulfate samples from the

Earth’s pre-Cambrian period were first reported by Farquhar et al (2000). In that work it was shown (Figure 13) that between 2,090 and 2,450 million years ago the Earths atmospheric composition was dramatically different from today’s, particularly with respect to lowered oxygen levels. Without a stratospheric UV filtration layer, ground level

SO

2

photolysis is possible. As for Mars, the Precambrian isotopic anomalies result from ground level, or at least from tropospheric UV photolysis of sulfur dioxide. As shown in the figure 13, both positive and negative mass

30

independent sulfur anomalies are present, with positive

17 O primarily associated with reduced sulfur species and negative with oxidized.

Fig. 12.

A pictogram of Δ 33 S vs. age, with the blue area representing data from various geological samples. The record divides Earth’s history into three stages: Stage I extends from before 3.8 Ga to 2.45 Ga and is characterized by Δ 33 S of variable sign and magnitude; Stage II extends from 2.45 to 2 Ga and displays more subdued Δ 33 S variability; and Stage III extends from 2 Ga until present and is characterized by Δ 33 S variability much less than

±0.2‰. The large filled circle represents hundreds of analyses of samples younger than 2.0 Ga. Stage III is defined on the basis of these measurements, not on the representative samples shown in the figure. The change from Stage I to Stage II is attributed to a change in the Earth’s atmospheric chemistry. Photolysis reactions involving SO

2

and SO in Earth’s early atmosphere, coupled with an efficient transfer of the signature to the Earth’s surface, produced the

Stage I record. The smaller Stage II record may reflect the onset of oxidative weathering or it may reflect stabilization of atmospheric oxygen to intermediate levels (10

−5 –10 −2

PAL).

Tropospheric sulfur dioxide photochemistry by UV light requires low levels of oxygen and ozone present to allow tropospheric light penetration as the lifetime of SO

2

is too short to allow transport to the stratosphere for

UV photochemical reaction. The isotopic wavelength sensitivity of SO

2

was experimentally studied (184.9, 193, 248 nm) and shown that the 193nm photochemistry produces a fractionation factor most similar to that observed in terrestrial sedimentary rock samples older than 2450Ma. This spectral region overlaps with the Schumann-Runge

31

bands of oxygen and the Hartley ozone bands and it is these features that lead to the suggestion that the absence of oxygen and ozone at these time periods leads to the photochemistry of SO

2

and production of the positive and negative anomalies observed in sulfur isotopes in ancient sediments. Based upon the photo absorption crosssection of O

2

, O

3

, and CO

2

and SO

2

lifetime of, it was suggested that the oxygen levels must have been at least a couple of orders of magnitude lower than present values. The observation that the sulfur anomaly disappears at approximately

2459 Ma years ago is interpreted as indicating the rise of oxygen levels towards levels comparable to present. There are at least three stages involved in the change of the oxygen levels and the details of these processes are reviewed

(Farquhar and Wing, 2003). The linkage to the evolution of the global sulfur system (Farquhar et al., 2010) and definition of transitional periods such as the Mesoarchaean interval (Farquhar et al., 2007b) and at the end of the

Paleoproterozoic within a marine basin have been discussed (Johnston et al., 2006) .There are also a range of

models and theoretical considerations that expand understanding of how these processes of atmospheric -chemical

and global biogeochemical cycles interact (Farquhar and Wing, 2003; Kasting, 2001; Mojzsis et al., 2003; Pavlov and Kasting, 2002; Runnegar et al., 2002). The global aspects of the late Archean biosphere oxygenation have been

documented from perturbations in sulfur isotopes simultaneously occurring in Mount McRae Shale of northwestern

Australia and equivalent features of the Gamohaan and Kuruman formations of South Africa (Kaufman et al., 2007).

It is apparent from these measurements that globally controlled processes of the biogeochemical cycle of sulfur are involved in producing controlling the observed isotopic anomalies. The Archean isotopic record of sulfates was

modeled (Halevy et al., 2010)and the role of extra-atmospheric processes, including biological, crustal formation,

and sulfur SO

2

and H

2

S volcanic emissions were defined. These events are particularly significant in accounting for spikes in the MIF isotopic fractionations over time. The use of the MIF compositions of Post-Archean times has also

been shown to be an effective means to detail global sulfur biogeochemical processes in this time period (Ono et al.,

2006a; Ono et al., 2006b) illustrated the benefits attained from detailed understanding of the physical chemistry of

the isotope effects associated with specific reaction steps and for example, the ability to measure multiple sulfur isotopes at the highest precisions and accuracies. The sulfur isotopic (

34 S and

33 S) record of seawater sulfate were utilized utilizing to place limits on pyrite burial and quantify the geological organic sulfur cycle with resolution of

short term organic burial processes (Wu et al., 2010).

Other possible scenarios for generation of the sulfur mass independent signal have been suggested where the anomalous fractionations of sulfur are produced by non photochemical processes, such as thermo chemical

reactions (Watanabe et al., 2009)

. The suggestion is based upon observed experimental production of a very small sulfur isotopic anomaly when mixtures of sulfate and amino acids are thermolyzed. There are, however, numerous issues with this suggestion and experimental interpretations. There also is no chemical mechanism available for the experiments which render extrapolation to a natural environment limited. The observed variation in

33 S is very small and insufficient to account for the sulfur data of the Archean (Figure 12) and most of the

36 S data is normal within experimental error. There is also no likely mechanism by which volcanoes may produce this anomaly in nature to explain the observed isotopic anomalies. Finally, the sulfur isotope anomalies in the ice core record associated with large volcanic events and the time evolution is consistent with photochemistry of an SO

2

plume and,

32

that fractionation pattern is also consistent with the Archean data. At present, without a clear definition of how this anomaly is chemically produced this mechanism is restricted in probable occurrence.

Another model for the observed fluctuations of sulfur isotopes in the Archean has been proposed

(Domagal-Goldman et al., 2008). The presence of an organic haze in the early Earth’s atmosphere could block the

lower atmosphere from UV fluxes responsible for the production of the sulfur anomaly and, simultaneously trigger an anti-greenhouse effect that initiates glacial activity. In this photochemical-geochemical model CH

4

, CO

2

levels mediate the haze level as well as modulates the thickness thereby controlling the evolution of the climate, chemistry, and biology of the Archean. It has also been suggested that based upon the geological record that impact events may be associated with the disappearance of ozone and increase of UV radiation that produce the sulfur anomalies

(Glikson, 2010)

.

The use of the atmospherically generated mass independent sulfur signal has proven to be a tracer of biological processes.

Microscopic isotopic sulfides in 3490-million-year old marine sulfates suggest that the early

microbial activity did not involve sulfate reduction, but rather, dispropotionation of elemental sulfur (Philippot et al.,

2007).

Microscopic pyrites with low 34 S/ 32 S ratios in sedimentary barites (3.47 Gyr old) from North Pole, Australia

have been interpreted as evidence for microbial sulfate reduction (Shen et al., 2009)and concluded that sulfate-

reducing bacteria had evolved by ~3.47 billion years ago, contrary to the hypothesis of Philippot et al (2007).

Quadruple sulfur isotope studies of the 3500 million year old Dressler Formation, Western Australia, also suggested

that sulfate reduction was active (Ueno et al., 2008) and disproportionation reactions were restricted. Thus at

present, the use of mass independent sulfur isotopic compositions have allowed new insights into the earths earliest life and such studies are expanding, allowing details of biogeochemical processes and specific microbial activities

such as phototrophic oxidation of sulfur species by green sulfur bacterium (Zerkle et al., 2008).

The storage of ancient atmospheric samples has proven to be surprisingly robust. Sulfur inclusions from

diamonds have been measured for their sulfur multi-isotopic compositions (Farquhar et al., 2002). Sulfide inclusions

in diamonds from the Orpa kimberlite pipe in the Kaapvaal-Zimbabwe craton of Botswana possess mass independent sulfur isotopic compositions. The anomalous isotopic fractionation pattern was produced via Archean atmospheric photochemistry and subsequent transfer to the Earth’s surface and sequestration into subducting material. The process initiated with volcanic emissions of sulfur bearing molecules in the reducing, low oxygen atmosphere and subsequent UV photochemistry that produce mass independent sulfur isotopic compositions. This also partitions the mass independent sulfur isotopic anomaly between elemental and oxidized redox states with opposing isotopic sign. The photolytic products are transferred to the surface of the Earth and enter the geological reservoir where they eventually convert to either sulfate (oceanic) or reduced sedimentary sulfides, possibly occurring in marginal seas and estuaries. These reservoirs become buried, metamorphose and are added to mantle material. The measured data imply that the material is mixed from the atmosphere to the mantle in the Archean based upon the mass independent isotopic composition of diamonds greater than 2450Ma. Younger diamonds are isotopically normal consistent with photochemistry of the anomalous oxygen in an older, low oxygen environment.

33

Aside from the preservation of an ancient atmospheric record, the diamond measurements offer an independent insight into mantle convective processes.

Sulfur isotopes in other sulfates are reflective of a variety processes. The pre Cambrian results have been a catalyst in developing understanding of isotope effects associated with sulfur gas phase photochemistry. Sulfur

isotopic anomalies are observed in ice core sulfate associated with major volcanic events (Baroni et al., 2008; Cole-

Dai et al., 2009; Savarino et al., 2003). The data have been interpreted as arising from the photolysis of sulfur

dioxide as the plume of volcanic emissions passes polewards (Greenland and Antarctica). The sign change in

33 S is also consistent with model predictions. It has been suggested that the anomalous composition of early earth sulfur

compounds may be a result of self shielding at wavelength less than 220 nm (Lyons, 2007; Lyons, 2008; Lyons,

2009) though these models are inconsistent with mass independent sulfur isotopic compositions of sulfates collected

from Antarctic snow that record the Pinatubo and Agung eruptions (Baroni et al., 2007,2008)). Another hypothesis

involves optical shielding by OCS and SO

2

photolysis that may account for the observed early earth samples based upon their high resolution UV spectroscopic studies of SO

2

(Ueno et al., 2009). Cumulatively, the sulfur (and

oxygen) photochemical studies demonstrate that high resolution interpretation of natural isotopic data require fundamental physical chemical understanding of the relevant chemical processes, but some of the most fascinating problems in geochemistry may be subsequently attacked with these measurements.

4.06.9

Sulfur Isotopic Fractionation Processes in Other Solar System Objects

There are significant examples of mass-independent isotopic fractionations associated with photochemically

initiated gas-solid formation. (Cook et al., 2007). Irradiation (313 nm) of gas phase CS

2

produces a solid aerosol

(CS

2

) x

that is highly mass-independently fractionated, with a 33 S excess of 5–10‰, and 36 S deficit of 61–84‰. An interesting aspect of these observations was that the optical properties of this solid are similar to those observed for aerosols produced by comet Shoemaker–Levy 9(SL9) during collision with the outer atmosphere of Jupiter. It was suggested that the mechanism responsible for the observed sulfur isotopic fractionation process involves the differing rates of nonradiative transfer from the initial absorbing rate to the final reactive state of the electronically excited CS

2

molecule (Colman et al., 1996). The rate of such processes is predominantly dependent upon Franck–

Condon factors that are partially mass dependent and the work experimentally clarified the mechanism for production of the photopolymeric (CS

2

) x

solid from CS

2

photolysis. The sulfur isotopic composition of products from photolysis of 12 CS

2

and 13 CS

2

differ significantly which rules out involvement of a symmetry-dependent fractionation effect. Franck–Condon and vibronic coupling effects were suggested as the source of the observed effect, specifically the nonradiative decay and intersystem crossing rates for the lowest excited states which give rise

to the highly anomalous sulfur isotopic composition observed in the solid product (Zmolek et al., 1999).

This work suggested that atmospheres of planets with CS

2

will give rise to anomalous compositions of sulfur, e.g. the Jupiter produced aerosols. It has also been shown that there exists a new class of mass independent isotopic fractionation process that may occur in extra terrestrial environments.

34

The processes of CO

2

photolysis may be of importance in the Martial upper atmosphere, and definition of relevant fractionation processes are needed to interpret SNC meteorite carbonate and sulfate isotopic data.

Photodecomposition of CO

2

by UV light at (185 nm) and the isotopes of the CO and O

2

products were observed to be highly enriched in 17

O (>100‰) (Bhattacharya et al., 2000). The dissociation arises via a spin-forbidden process

during the singlet to triplet transition. The mass independence derives from the reliance of the reaction rate on essentially resonant spin–orbit coupling of the low-energy vibrational states of the 16 O 12 C 17 O molecule of the singlet with the triplet state. Later studies indicated wavelength dependency (184.9, 123.6, 116.5 nm) and proposed involvement of a hyperfine interaction between nuclear spin and electron spins or orbital motion resulting in preferential 17 O and 13

C photolysis (Mahata and Bhattacharya, 2009). At present, the role of photochemistry of CO

2 in the Martian atmosphere is being evaluated as this photolysis may occur to produce high 17 O enrichments.

4.07 Concluding Comments

Since the discovery of a chemically produced mass-independent isotope effect (Thiemens and Heidenreich,

1983), it has been demonstrated that there exists an extensive range of applications and observations in nature.

Resolution of the actual physical–chemical mechanism has advanced understanding of gas-phase transition states and chemical reactions and fundamental photochemistry. Mass-independent oxygen isotopic compositions have been observed in essentially all atmospheric molecules, except water. This includes: O

2

, O

3

, CO

2

, N

2

O, H

2

O

2

, and aerosol carbonate, sulfate and nitrate. In each specific instance, our understanding of biogeochemical cycle has been enhanced as a result of its specific isotopic signature. Measurements of sulfate and nitrate in ice core samples have provided new details of the past oxidative capacity of the Earth’s atmosphere. Oxygen isotopic compositions of ancient sulfates (e.g. Miocene) have provided new paleo-atmospheric information on the desert and nearby ocean in

Namibia, volcanogenic processes, Antarctic dry valley sources, and desert varnish formation processes. Sulfur massindependent isotopic anomalies in the Earth's oldest rocks (3.8 × 10 9 yr) have provided a new means by which the evolution of oxygen-ozone, and consequently life has evolved until ~2.2 × 10 9 years before present, a parameter sought for decades but only attained from the mass independent sulfur isotopic measurements.

There continue to be new and novel applications of the mass independent isotopic fractionation processes in nature such as the role and potential applicability of 17

O effects in the hydrologic cycle (Angert et al., 2004). The

multi isotopic fractionation associated with precipitation and the temperature independency may render high precision measurements of water an important new and unique tracer of the hydrologic cycle. This application may also extend to measurements of the three isotope composition of ice at high precision may provide important new insight into the relative contributions of kinetic isotope effects, evaporation/condensation, and relative humidity on the hydrologic cycle. This work is an important advancement and is likely to advance understanding of both the hydrologic cycle and climate change.

Measurements of sulfur and oxygen isotopic anomalies in secondary minerals from Martian meteorites have provided new insights into crust–atmosphere interactions on Mars, especially with the observations of a new

35

surface reaction mechanism involving ozone-peroxide-carbonates. The observation of mass independent sulfur isotopic anomalies in the Earth’s Pre Cambrian period has opened an entirely new field for the study of the origin and evolution of oxygen. Finally, chemical mass-independent process appears to be responsible for the production of the anomalous oxygen isotopic compositions observed in meteorites and thus was a major process in the formation of the solar system, though the exact mechanism remains presently unresolved.

There remain many new horizons in mass-independent chemistry and its applications. This includes theory, laboratory experiments, and new applications in nature, extending from Earth throughout the solar system, and in time, from the present to 4.55Ga.

Acknowledgments

The generous support of the NSF (Atmospheric Chemistry and Polar Programs) and NASA (Cosmochemistry and

Origins of Solar Systems) have allowed and facilitated many of the avenues of research discussed in this Chapter.

Without this support a major portion of this chapters work would not have been accomplished. Drs. S. Chakraborty,

J. Farquhar, H. Bao, G. Dominguez are acknowledged for their help with creation of the figures for this text.

Airieau, S. A., Farquhar, J., Thiemens, M. H., Leshin, L. A., Bao, H., and Young, E., 2005, Planetesimal sulfate and aqueous alteration in CM and CI carbonaceous chondrites: Geochimica Et Cosmochimica Acta, v. 69, no.

16, p. 4166-4171.

Al-Hosney, H. A., Carlos-Cuellar, S., Baltrusaitis, J., and Grassian, V. H., 2005, Heterogeneous uptake and reactivity of formic acid on calcium carbonate particles: a Knudsen cell reactor, FTIR and SEM study:

Physical Chemistry Chemical Physics, v. 7, p. 3587-3595.

Al-Hosney, H. A., and Grassian, V. H., 2005, Water, sulfur dioxide and nitric acid adsorption on calcium carbonate:

A transmission and ATR-FTIR study: Physical Chemistry Chemical Physics, v. 7, no. 6, p. 1266-1276.

Alexander, B., Park, R. J., Jacob, D. J., Li, Q. B., Yantosca, R. M., Savarino, J., Lee, C. C. W., and Thiemens, M.

H., 2005, Sulfate formation in sea-salt aerosols: constraints from oxygen isotopes: Journal of Geophysical

Research-Part D-Atmospheres, v. 110, no. D10, p. 12 pp.

Alexander, B., Park, R. J., Jacob, D. J., and Sunling, G., 2009, Transition metal-catalyzed oxidation of atmospheric sulfur: global implications for the sulfur budget: Journal of Geophysical Research - Part D - Atmospheres, p. D02309 (02313 pp.).

Alexander, B., Savanna, J., Kreutz, K. J., and Thiemens, M. H., 2004a, Impact of preindustrial biomass-burning emissions on the oxidation pathways of tropospheric sulfur and nitrogen: Journal of Geophysical Research, v. 109, no. D8, p. 8 pp.

Alexander, B., Savarino, J., Barkov, N. I., Delmas, R. J., and Thiemens, M. H., 2002, Climate driven changes in the oxidation pathways of atmospheric sulfur: Geophysical Research Letters, v. 29, no. 14, p. 30-34.

Alexander, B., Savarino, J., Kreutz, K. J., and Thiemens, M. H., 2004b, Impact of preindustrial biomass-burning emissions on the oxidation pathways of tropospheric sulfur and nitrogen: Journal of Geophysical Research-

Atmospheres, v. 109, no. D8.

Alexander, B., Thiemens, M. H., Farquhar, J., Kaufman, A. J., Savarino, J., and Delmas, R. J., 2003, East Antarctic ice core sulfur isotope measurements over a complete glacial-interglacial cycle: Journal of Geophysical

Research-Atmospheres, v. 108, no. D24.

Alexander, B., Vollmer, M. K., Jackson, T., Weiss, R. F., and Thiemens, M. H., 2001, Stratospheric CO2 isotopic anomalies and SF6 and CFC tracer concentrations in the Arctic polar vortex: Geophysical Research Letters, v. 28, no. 21, p. 4103-4106.

Anderson, S. M., Hulsebusch, D., and Mauersberger, K., 1997, Surprising rate coefficients for four isotopic variants of O+O-2+M: Journal of Chemical Physics, v. 107, no. 14, p. 5385-5392.

36

Angert, A., Barkan, E., Barnett, B., Brugnoli, E., Davidson, E. A., Fessenden, J., Maneepong, S., Panapitukkul, N.,

Randerson, J. T., Savage, K., Yakir, D., and Luz, B., 2003a, Contribution of soil respiration in tropical, temperate, and boreal forests to the O-18 enrichment of atmospheric O-2: Global Biogeochemical Cycles, v. 17, no. 3.

Angert, A., Cappa, C. D., and DePaolo, D. J., 2004, Kinetic O-17 effects in the hydrologic cycle: Indirect evidence and implications: Geochimica Et Cosmochimica Acta, v. 68, no. 17, p. 3487-3495.

Angert, A., Luz, B., and Yakir, D., 2001, Fractionation of oxygen isotopes by respiration and diffusion in soils and its implications for the isotopic composition of atmospheric O-2: Global Biogeochemical Cycles, v. 15, no.

4, p. 871-880.

Angert, A., Rachmilevitch, S., Barkan, E., and Luz, B., 2003b, Effects of photorespiration, the cytochrome pathway, and the alternative pathway on the triple isotopic composition of atmospheric O-2: Global Biogeochemical

Cycles, v. 17, no. 1.

Assonov, S. S., and Brenninkmeijer, C. A. M., 2001, A new method to determine the O-17 isotopic abundance in

CO2 using oxygen isotope exchange with a solid oxide: Rapid Communications in Mass Spectrometry, v.

15, no. 24, p. 2426-2437.

Babikov, D., Kendrick, B. K., Walker, R. B., Pack, R. T., Fleurat-Lesard, P., and Schinke, R., 2003a, Formation of ozone: Metastable states and anomalous isotope effect: Journal of Chemical Physics, v. 119, no. 5, p. 2577-

2589.

-, 2003b, Metastable states of ozone calculated on an accurate potential energy surface: Journal of Chemical Physics, v. 118, no. 14, p. 6298-6308.

Babikov, D., Kendrick, B. K., Walker, R. B., Schinke, R., and Pack, R. T., 2003c, Quantum origin of an anomalous isotope effect in ozone formation: Chemical Physics Letters, v. 372, no. 5-6, p. 686-691.

Bandfield, J. L., 2002a, Global mineral distributions on Mars: Journal of Geophysical Research-Planets, v. 107, no.

E6.

-, 2002b, Global mineral distributions on Mars: Journal of Geophysical Research, v. 107, no. E6, p. 9-1-20.

Bandfield, J. L., Glotch, T. D., and Christensen, P. R., 2003, Spectroscopic identification of carbonate minerals in the martian dust: Science, v. 301, no. 5636, p. 1084-1087.

Bao, H., and Marchant, D. R., 2006, Quantifying sulfate components and their variations in soils of the McMurdo

Dry Valleys, Antarctica: Journal of Geophysical Research-Part D-Atmospheres, v. 111, no. D16, p. 13 pp.

Bao, H., and Reheis, M. C., 2003, Multiple oxygen and sulfur isotopic analyses on water-soluble sulfate in bulk atmospheric deposition from the southwestern United States: Journal of Geophysical Research, v. 108, no.

D14, p. ACH9-1-9.

Bao, H., Thiemens, M. H., Loope, D. B., and Xun-Lai, Y., 2003a, Sulfate oxygen-17 anomaly in an Oligocene ash bed in mid-North America: was it the dry fogs?: Geophysical Research Letters, v. 30, no. 16, p. HLS1-1-4.

Bao, H. M., 2005, Sulfate in modem playa settings and in ash beds in hyperarid deserts: implication for the origin of

O-17-anomalous sulfate in an Oligocene ash bed: Chemical Geology, v. 214, no. 1-2, p. 127-134.

Bao, H. M., Campbell, D. A., Bockheim, J. G., and Thiemens, M. H., 2000a, Origins of sulphate in Antarctic dryvalley soils as deduced from anomalous O-17 compositions: Nature, v. 407, no. 6803, p. 499-502.

Bao, H. M., Koch, P. L., and Thiemens, M. H., 2000b, Oxygen isotopic composition of ferric oxides from recent soil, hydrologic, and marine environments: Geochimica Et Cosmochimica Acta, v. 64, no. 13, p. 2221-

2231.

Bao, H. M., Lyons, J. R., and Zhou, C. M., 2008, Triple oxygen isotope evidence for elevated CO2 levels after a

Neoproterozoic glaciation: Nature, v. 453, no. 7194, p. 504-506.

Bao, H. M., Michalski, G. M., and Thiemens, M. H., 2001, Sulfate oxygen-17 anomalies in desert varnishes:

Geochimica Et Cosmochimica Acta, v. 65, no. 13, p. 2029-2036.

Bao, H. M., Thiemens, M. H., Loope, D. B., and Yuan, X. L., 2003b, Sulfate oxygen-17 anomaly in an Oligocene ash bed in mid-North America: Was it the dry fogs?: Geophysical Research Letters, v. 30, no. 16.

Baroni, M., Savarino, J., Cole-Dai, J., Rai, V. K., and Thiemens, M. H., 2008, Anomalous sulfur isotope compositions of volcanic sulfate over the last millennium in Antarctic ice cores: Journal of Geophysical

Research-Atmospheres, v. 113, no. D20.

Battle, M., Bender, M. L., Tans, P. P., White, J. W. C., Ellis, J. T., Conway, T., and Francey, R. J., 2000, Global carbon sinks and their variability inferred from atmospheric O-2 and delta C-13: Science, v. 287, no. 5462, p. 2467-2470.

Bender, M., 2003, Climate-biosphere interactions on glacial-interglacial timescales: Global Biogeochemical Cycles, v. 17, no. 3.

37

Benedix, G. K., Leshin, L. A., Farquhar, J., Jackson, T., and Thiemens, M. H., 2003, Carbonates in CM2 chondrites:

Constraints on alteration conditions from oxygen isotopic compositions and petrographic observations:

Geochimica Et Cosmochimica Acta, v. 67, no. 8, p. 1577-1588.

Bhattacharya, S. K., Savarino, J., and Thiemens, M. H., 2000, A new class of oxygen isotopic fractionation in photodissociation of carbon dioxide: Potential implications for atmospheres of Mars and Earth:

Geophysical Research Letters, v. 27, no. 10, p. 1459-1462.

Bhattacharya, S. K., and Thiemens, M. H., 1988, Isotopic Fractionation in Ozone Decomposition: Geophysical

Research Letters, v. 15, no. 1, p. 9-12.

Bigeleisen, J., and Mayer, M. G., 1947, CALCULATION OF EQUILIBRIUM CONSTANTS FOR ISOTOPIC

EXCHANGE REACTIONS: Journal of Chemical Physics, v. 15, no. 5, p. 261-267.

Blunier, T., Barnett, B., Bender, M. L., and Hendricks, M. B., 2002, Biological oxygen productivity during the last

60,000 years from triple oxygen isotope measurements: Global Biogeochemical Cycles, v. 16, no. 3.

Boehlke, J. K., Smith, R. L., and Hannon, J. E., 2007, Isotopic analysis of N and O in nitrite and nitrate by sequential selective bacterial reduction to N2O: Analytical Chemistry, v. 79, no. 15, p. 5888-5895.

Boering, K. A., Jackson, T., Hoag, K. J., Cole, A. S., Perri, M. J., Thiemens, M., and Atlas, E., 2004, Observations of the anomalous oxygen isotopic composition of carbon dioxide in the lower stratosphere and the flux of the anomaly to the troposphere: Geophysical Research Letters, v. 31, no. 3.

Brenninkmeijer, C. A. M., Janssen, C., Kaiser, J., Rockmann, T., Rhee, T. S., and Assonov, S. S., 2003, Isotope effects in the chemistry of atmospheric trace compounds: Chemical Reviews, v. 103, no. 12, p. 5125-5161.

Brenninkmeijer, C. A. M., and Rockmann, T., 1998, A rapid method for the preparation of O-2 from CO2 for mass spectrometric measurement of O-17/O-16 ratios: Rapid Communications in Mass Spectrometry, v. 12, no.

8, p. 479-483.

Bullock, M. A., and Moore, J. M., 2007, Atmospheric conditions on early Mars and the missing layered carbonates:

Geophysical Research Letters, v. 34.

Casciotti, K. L., Boehlke, J. K., McIlvin, M. R., Mroczkowski, S. J., and Hannon, J. E., 2007, Oxygen isotopes in nitrite: Analysis, calibration, and equilibration: Analytical Chemistry, v. 79, no. 6, p. 2427-2436.

Chakraborty, S., Ahmed, M., Jackson, T. L., and Thiemens, M. H., 2008, Experimental test of self-shielding in vacuum ultraviolet photodissociation of CO: Science, v. 321, no. 5894, p. 1328-1331.

-, 2009, Response to Comments on "Experimental Test of Self-Shielding in Vacuum Ultraviolet Photodissociation of CO": Science, v. 324, no. 5934.

Chakraborty, S., and Bhattacharya, S. K., 2003a, Experimental investigation of oxygen isotope exchange between

CO /sub 2/ and O(/sup 1/D) and its relevance to the stratosphere: Journal of Geophysical Research, v. 108, no. D23, p. ACH5-1-15.

-, 2003b, Oxygen isotopic fractionation during UV and visible light photodissociation of ozone: Journal of Chemical

Physics, v. 118, no. 5, p. 2164-2172.

Chmura, W. M., Rozanski, K., Kuc, T., and Gorczyca, Z., 2009, Comparison of two methods for the determination of nitrogen and oxygen isotope composition of dissolved nitrates: Nukleonika, v. 54, no. 1, p. 17-23.

Christensen, P. R., Bandfield, J. L., Smith, M. D., Hamilton, V. E., and Clark, R. N., 2000, Identification of a basaltic component on the Martian surface from Thermal Emission Spectrometer data: Journal of

Geophysical Research-Planets, v. 105, no. E4, p. 9609-9621.

Clayton, R. E., Hudson-Edwards, K. A., Malinovsky, D., and Andersson, P., 2005, Fe isotope fractionation during the precipitation of ferrihydrite and transformation of ferrihydrite to goethite: Mineralogical Magazine, v.

69, no. 5, p. 667-676.

Clayton, R. N., 2002, Solar System - Self-shielding in the solar nebula: Nature, v. 415, no. 6874, p. 860-861.

-, 2003, Oxygen isotopes in the solar system: Space Science Reviews, v. 106, no. 1-4, p. 19-32.

Clayton, R. N., Grossman, L., and Mayeda, T. K., 1973, COMPONENT OF PRIMITIVE NUCLEAR

COMPOSITION IN CARBONACEOUS METEORITES: Science, v. 182, no. 4111, p. 485-488.

Clayton, R. N., and Mayeda, T. K., 1988, FORMATION OF UREILITES BY NEBULAR PROCESSES:

Geochimica Et Cosmochimica Acta, v. 52, no. 5, p. 1313-1318.

Cole-Dai, J., Ferris, D., Lanciki, A., Savarino, J., Baroni, M., and Thiemens, M. H., 2009, Cold decade (AD 1810-

1819) caused by Tambora (1815) and another (1809) stratospheric volcanic eruption: Geophysical

Research Letters, v. 36.

Colman, J. J., Xu, X. P., Thiemens, M. H., and Trogler, W. C., 1996, Photopolymerization and mass-independent sulfur isotope fractionations in carbon disulfide: Science, v. 273, no. 5276, p. 774-776.

38

Cook, D. L., Wadhwa, M., Clayton, R. N., Dauphas, N., Janney, P. E., and Davis, A. M., 2007, Mass-dependent fractionation of nickel isotopes in meteoritic metal: Meteoritics & Planetary Science, v. 42, no. 12, p. 2067-

2077.

Cooper, G. W., Thiemens, M. H., Jackson, T., and Chang, S., 1997a, Oxygen isotopic composition of organic and inorganic compounds of the Murchison meteorite: Meteoritics & Planetary Science, v. 32, no. 4, p. A32-

A32.

Cooper, G. W., Thiemens, M. H., Jackson, T. L., and Chang, S., 1997b, Sulfur and hydrogen isotope anomalies in meteorite sulfonic acids: Science, v. 277, no. 5329, p. 1072-1074.

Davis, A. M., Lugaro, M., Gallino, R., Pellin, M. J., Lewis, R. S., and Clayton, R. N., 2001, Isotopic compositions of heavy elements in presolar grains: new constraints on nucleosynthesis: Memorie della Societa

Astronomica Italiana, v. 72, no. 2.

Deshler, T., Larsen, N., Weissner, C., Schreiner, J., Mauersberger, K., Cairo, F., Adriani, A., Di Donfrancesco, G.,

Ovarlez, J., Ovarlez, H., Blum, U., Fricke, K. H., and Dornbrack, A., 2003, Large nitric acid particles at the top of an Arctic stratospheric cloud: Journal of Geophysical Research-Atmospheres, v. 108, no. D16.

Domagal-Goldman, S. D., Kasting, J. F., Johnston, D. T., and Farquhar, J., 2008, Organic haze, glaciations and multiple sulfur isotopes in the Mid-Archean Era: Earth and Planetary Science Letters, v. 269, no. 1-2, p. 29-

40.

Dominguez, G., 2010, A HETEROGENEOUS CHEMICAL ORIGIN FOR THE (16)O-ENRICHED AND (16)O-

DEPLETED RESERVOIRS OF THE EARLY SOLAR SYSTEM: Astrophysical Journal Letters, v. 713, no. 1, p. L59-L63.

Dominguez, G., Jackson, T., Brothers, L., Barnett, B., Nguyen, B., and Thiemens, M. H., 2008, Discovery and measurement of an isotopically distinct source of sulfate in Earth's atmosphere: Proceedings of the National

Academy of Sciences of the United States of America, v. 105, no. 35, p. 12769-12773.

Farquhar, J., Jackson, T. L., and Thiemens, M. H., 2000a, A S-33 enrichment in ureilite meteorites: Evidence for a nebular sulfur component: Geochimica Et Cosmochimica Acta, v. 64, no. 10, p. 1819-1825.

Farquhar, J., Kim, S.-T., and Masterson, A., 2007a, Implications from sulfur isotopes of the Nakhla meteorite for the origin of sulfate on Mars: Earth and Planetary Science Letters, v. 264, no. 1-2, p. 1-8.

Farquhar, J., Peters, M., Johnston, D. T., Strauss, H., Masterson, A., Wiechert, U., and Kaufman, A. J., 2007b,

Isotopic evidence for Mesoarchaean anoxia and changing atmospheric sulphur chemistry: Nature, v. 449, no. 7163, p. 706-U705.

Farquhar, J., Savarino, J., Jackson, T. L., and Thiemens, M. H., 2000b, Evidence of atmospheric sulphur in the martian regolith from sulphur isotopes in meteorites: Nature, v. 404, no. 6773, p. 50-52.

Farquhar, J., and Thiemens, M. H., 2000, Oxygen cycle of the Martian atmosphere-regolith system: Delta O-17 of secondary phases in Nakhla and Lafayette: Journal of Geophysical Research-Planets, v. 105, no. E5, p.

11991-11997.

Farquhar, J., Thiemens, M. H., and Jackson, T., 1998, Atmosphere-surface interactions on Mars: Delta O-17 measurements of carbonate from ALH 84001: Science, v. 280, no. 5369, p. 1580-1582.

Farquhar, J., and Wing, B. A., 2003, Multiple sulfur isotopes and the evolution of the atmosphere: Earth and

Planetary Science Letters, v. 213, no. 1-2, p. 1-13.

Farquhar, J., Wing, B. A., McKeegan, K. D., Harris, J. W., Cartigny, P., and Thiemens, M. H., 2002, Massindependent sulfur of inclusions in diamond and sulfur recycling on early earth: Science, v. 298, no. 5602, p. 2369-2372.

Farquhar, J., Wu, N. P., Canfield, D. E., and Oduro, H., 2010, Connections between Sulfur Cycle Evolution, Sulfur

Isotopes, Sediments, and Base Metal Sulfide Deposits: Economic Geology, v. 105, no. 3, p. 509-533.

Federman, S. R., and Young, E. D., 2009, Comment on "Experimental Test of Self-Shielding in Vacuum Ultraviolet

Photodissociation of CO": Science, v. 324, no. 5934.

Gao, X., and Thiemens, M. H., 1993a, Isotopic Composition and Concentration of Sulfur in Carbonaceous

Chondrites: Geochimica Et Cosmochimica Acta, v. 57, no. 13, p. 3159-3169.

-, 1993b, Variations of the Isotopic Composition of Sulfur in Enstatite and Ordinary Chondrites: Geochimica Et

Cosmochimica Acta, v. 57, no. 13, p. 3171-3176.

Gao, Y. Q., and Marcus, R. A., 2001, Strange and unconventional isotope effects in ozone formation: Science, v.

293, no. 5528, p. 259-263.

Gao, Y. Q., and Marcus, R. A., 2007, An approximate theory of the ozone isotopic effects: Rate constant ratios and pressure dependence: Journal of Chemical Physics, v. 127, no. 24.

39

Gendrin, A., Mangold, N., Bibring, J. P., Langevin, Y., Gondet, B., Poulet, F., Bonello, G., Quantin, C., Mustard, J.,

Arvidson, R., and LeMouelic, S., 2005a, Suffates in martian layered terrains: the OMEGA/Mars Express view: Science, v. 307, no. 5715, p. 1587-1591.

Gendrin, A., Mustard, J. F., Hutchison, L., Bibring, J. P., Poulet, F., Zettler, L. A. A., Amils, R., and Fernandez -

Remolar, D., 2005b, Mineral deposits on Mars revealed with the OMEGA visible-infrared imaging spectrometer and their implications for astrobiology: Astrobiology, v. 5, no. 2, p. 301.

Giauque, W. F., and Johnston, H. L., 1929a, An isotope of oxygen, mass 17, in the earth's atmosphere: Journal of the

American Chemical Society, v. 51, p. 3528-3534.

-, 1929b, An isotope of oxygen, mass 18: Nature, v. 123, p. 318-318.

-, 1929c, Isotope of oxygen, mass 18. interpretation of the atmospheric absorption bands: Journal of the American

Chemical Society, v. 51.

Glikson, A., 2010, Archaean asteroid impacts, banded iron formations and MIF-S anomalies: A discussion: Icarus, v. 207, no. 1, p. 39-44.

Goodrich, C. A., Jones, J. H., and Berkley, J. L., 1987, ORIGIN AND EVOLUTION OF THE UREILITE PARENT

MAGMAS - MULTISTAGE IGNEOUS ACTIVITY ON A LARGE PARENT BODY: Geochimica Et

Cosmochimica Acta, v. 51, no. 9, p. 2255-2273.

Guenther, J., Erbacher, B., Krankowsky, D., and Mauersberger, K., 1999, Pressure dependence of two relative ozone formation rate coefficients: Chemical Physics Letters, v. 306, no. 5-6, p. 209-213.

Halevy, I., Johnston, D. T., and Schrag, D. P., 2010, Explaining the structure of the Archean mass-independent sulfur isotope record: Geochimica Et Cosmochimica Acta, v. 74, no. 12, p. A371-A371.

Halevy, I., Zuber, M. T., and Schrag, D. P., 2007, A sulfur dioxide climate feedback on early Mars: Science, v. 318, no. 5858, p. 1903-1907.

Hathorn, B. C., and Marcus, R. A., 1999, An intramolecular theory of the mass-independent isotope effect for ozone.

I: Journal of Chemical Physics, v. 111, no. 9, p. 4087-4100.

-, 2000, An intramolecular theory of the mass-independent isotope effect for ozone. II. Numerical implementation at low pressures using a loose transition state: Journal of Chemical Physics, v. 113, no. 21, p. 9497-9509.

-, 2001, Estimation of vibrational frequencies and vibrational densities of states in isotopically substituted nonlinear triatomic molecules: Journal of Physical Chemistry A, v. 105, no. 23, p. 5586-5589.

Heidenreich, J. E., and Thiemens, M. H., 1983, A Non-Mass-Dependent Isotope Effect in the Production of Ozone from Molecular-Oxygen: Journal of Chemical Physics, v. 78, no. 2, p. 892-895.

-, 1986, A Non-Mass-Dependent Oxygen Isotope Effect in the Production of Ozone from Molecular-Oxygen - the

Role of Molecular Symmetry in Isotope Chemistry: Journal of Chemical Physics, v. 84, no. 4, p. 2129-

2136.

Helene Lefebvre-Brion, R. W. F., 2004, Spectroscopy and Dynamics of Diatomic Molecules, Oxford, UK, Elsevier

Science & Technology.

Herzberg, G., 2008, Atomic Spectra and Atomic Structure, Herzberg Press.

Hoag, K. J., Still, C. J., Fung, I. Y., and Boering, K. A., 2005, Triple oxygen isotope composition of tropospheric carbon dioxide as a tracer of terrestrial gross carbon fluxes: Geophysical Research Letters, v. 32, no. 2, p. 5 pp.

Holland, H. D., 1978, CO2-HCO3(-)-CO3(2-) SYSTEM THROUGH GEOLOGIC TIME AND ITS

IMPLICATIONS FOR EVOLUTION OF ATMOSPHERE: Pure and Applied Geophysics, v. 116, no. 2-3, p. 232-233.

Hulston, J. R., and Thode, H. G., 1965a, COSMIC-RAY-PRODUCED S36 AND S33 IN METALLIC PHASE OF

IRON METEORITES: Journal of Geophysical Research, v. 70, no. 18, p. 4435-&.

-, 1965b, VARIATIONS IN S33 S34 AND S36 CONTENTS OF METEORITES AND THEIR RELATION TO

CHEMICAL AND NUCLEAR EFFECTS: Journal of Geophysical Research, v. 70, no. 14, p. 3475-&.

Irion, F. W., Gunson, M. R., Rinsland, C. P., Yung, Y. L., Abrams, M. C., Chang, A. Y., and Goldman, A., 1996,

Heavy ozone enrichments from ATMOS infrared solar spectra: Geophysical Research Letters, v. 23, no.

17, p. 2377-2380.

Janssen, C., Guenther, J., Krankowsky, D., and Mauersberger, K., 1999, Relative formation rates of O-50(3) and O-

52(3) in O-16-O-18 mixtures: Journal of Chemical Physics, v. 111, no. 16, p. 7179-7182.

-, 2003, Temperature dependence of ozone rate coefficients and isotopologue fractionation in O-16-O-18 oxygen mixtures: Chemical Physics Letters, v. 367, no. 1-2, p. 34-38.

Janssen, C., Guenther, J., Mauersberger, K., and Krankowsky, D., 2001, Kinetic origin of the ozone isotope effect: a critical analysis of enrichments and rate coefficients: Physical Chemistry Chemical Physics, v. 3, no. 21, p.

4718-4721.

40

Jiang, L., and Babikov, D., 2009, A reduced dimensionality model of ozone: Semiclassical treatment of van der

Waals states: Chemical Physics Letters, v. 474, no. 4-6, p. 273-277.

Johnson, D. G., Jucks, K. W., Traub, W. A., and Chance, K. V., 2000, Isotopic composition of stratospheric ozone:

Journal of Geophysical Research-Atmospheres, v. 105, no. D7, p. 9025-9031.

Johnson, M. S., Billing, G. D., Gruodis, A., and Janssen, M. H. M., 2001, Photolysis of nitrous oxide isotopomers studied by time-dependent hermite propagation: Journal of Physical Chemistry A, v. 105, no. 38, p. 8672-

8680.

Johnson, S. S., Mischna, M. A., Grove, T. L., and Zuber, M. T., 2008, Sulfur-induced greenhouse warming on early

Mars: Journal of Geophysical Research-Planets, v. 113, no. E8.

Johnston, D. T., Poulton, S. W., Fralick, P. W., Wing, B. A., Canfield, D. E., and Farquhar, J., 2006, Evolution of the oceanic sulfur cycle at the end of the Paleoproterozoic: Geochimica Et Cosmochimica Acta, v. 70, no.

23, p. 5723-5739.

Johnston, J. C., Rockmann, T., and Brenninkmeijer, C. A. M., 2000, CO2+O(D-1) isotopic exchange: Laboratory and modeling studies: Journal of Geophysical Research-Atmospheres, v. 105, no. D12, p. 15213-15229.

Johnston, J. C., and Thiemens, M. H., 1997, The isotopic composition of tropospheric ozone in three environments:

Journal of Geophysical Research-Atmospheres, v. 102, no. D21, p. 25395-25404.

Jull, A. J. T., Eastoe, C. J., Xue, S., and Herzog, G. F., 1995, ISOTOPIC COMPOSITION OF CARBONATES IN

THE SNC METEORITES ALLAN-HILLS-84001 AND NAKHLA: Meteoritics, v. 30, no. 3, p. 311-318.

Karagulian, F., Santschi, C., and Rossi, M. J., 2006, The heterogeneous chemical kinetics of N2O5 on CaCO3 and other atmospheric mineral dust surrogates: Atmospheric Chemistry and Physics, v. 6, p. 1373-1388.

Kasibhatla, P., Chameides, W. L., and StJohn, J., 1997, A three-dimensional global model investigation of seasonal variations in the atmospheric burden of anthropogenic sulfate aerosols: Journal of Geophysical Research-

Atmospheres, v. 102, no. D3, p. 3737-3759.

Kasting, J. F., 2001, Earth history - The rise of atmospheric oxygen: Science, v. 293, no. 5531, p. 819-820.

Kaufman, A. J., Johnston, D. T., Farquhar, J., Masterson, A. L., Lyons, T. W., Bates, S., Anbar, A. D., Arnold, G.

L., Garvin, J., and Buick, R., 2007, Late Archean biospheric oxygenation and atmospheric evolution:

Science, v. 317, no. 5846, p. 1900-1903.

Krankowsky, D., Bartecki, F., Klees, G. G., Mauersberger, K., Schellenbach, K., and Stehr, J., 1995,

MEASUREMENT OF HEAVY ISOTOPE ENRICHMENT IN TROPOSPHERIC OZONE: Geophysical

Research Letters, v. 22, no. 13, p. 1713-1716.

Krankowsky, D., Lammerzahl, P., and Mauersberger, K., 2000, Isotopic measurements of stratospheric ozone:

Geophysical Research Letters, v. 27, no. 17, p. 2593-2595.

Krankowsky, D., Lammerzahl, P., Mauersberger, K., Janssen, C., Tuzson, B., and Rockmann, T., 2007,

Stratospheric ozone isotope fractionations derived from collected samples: Journal of Geophysical

Research-Atmospheres, v. 112, no. D8.

Krot, A. N., Hutcheon, I. D., Yurimoto, H., Cuzzi, J. N., McKeegan, K. D., Scott, E. R. D., Libourel, G.,

Chaussidon, M., Aleon, J., and Petaev, M. I., 2005a, Evolution of oxygen isotopic composition in the inner solar nebula: Astrophysical Journal, v. 622, no. 2, p. 1333-1342.

Krot, A. N., Yurimoto, H., Hutcheon, I. D., and MacPherson, G. J., 2005b, Chronology of the early Solar System from chondrule-bearing calcium-aluminium-rich inclusions: Nature, v. 434, no. 7036, p. 998-1001.

Krot, A. N., Yurimoto, H., McKeegan, K. D., Leshin, L., Chaussidon, M., Libourel, G., Yoshitake, M., Huss, G. R.,

Guan, Y., and Zanda, B., 2006, Oxygen isotopic compositions of chondrules: Implications for evolution of oxygen isotopic reservoirs in the inner solar nebula: Chemie Der Erde-Geochemistry, v. 66, no. 4, p. 249-

276.

Kunasek, S. A., Alexander, B., Steig, E. J., Hastings, M. G., Gleason, D. J., and Jarvis, J. C., 2008, Measurements and modeling of Delta(17)O of nitrate in snowpits from Summit, Greenland: Journal of Geophysical

Research-Atmospheres, v. 113.

Kunasek, S. A., Alexander, B., Steig, E. J., Sofen, E. D., Jackson, T. L., Thiemens, M. H., McConnell, J. R.,

Gleason, D. J., and Amos, H. M., 2010, Sulfate sources and oxidation chemistry over the past 230 years from sulfur and oxygen isotopes of sulfate in a West Antarctic ice core: Journal of Geophysical Research-

Atmospheres, v. 115.

Lammerzahl, P., Rockmann, T., Brenninkmeijer, C. A. M., Krankowsky, D., and Mauersberger, K., 2002, Oxygen isotope composition of stratospheric carbon dioxide: Geophysical Research Letters, v. 29, no. 12.

Lee, C. C. W., Savarino, J., Cachier, H., and Thiemens, M. H., 2002, Sulfur (S-32, S-33, S-34, S-36) and oxygen (O-

16, O-17, O-18) isotopic ratios of primary sulfate produced from combustion processes: Tellus Series B-

Chemical and Physical Meteorology, v. 54, no. 3, p. 193-200.

41

Lee, C. C. W., Savarino, J., and Thiemens, M. H., 2001, Mass independent oxygen isotopic composition of atmospheric sulfate: Origin and implications for the present and past atmosphere of earth and mars:

Geophysical Research Letters, v. 28, no. 9, p. 1783-1786.

Lee, C. C. W., and Thiemens, M. H., 2001, The delta O-17 and delta O-18 measurements of atmospheric sulfate from a coastal and high alpine region: A mass-independent isotopic anomaly: Journal of Geophysical

Research-Atmospheres, v. 106, no. D15, p. 17359-17373.

Lelieveld, J., Roelofs, G. J., Feichter, J., and Rodhe, H., 1997, Terrestrial sources and distribution of atmospheric sulphur: Philosophical Transactions of the Royal Society B-Biological Sciences, v. 352, no. 1350, p. 149-

157.

Liang, M.-C., Blake, G. A., Lewis, B. R., and Yung, Y. L., 2007, Oxygen isotopic composition of carbon dioxide in the middle atmosphere: Proceedings of the National Academy of Sciences of the United States of America, v. 104, no. 1, p. 21-25.

Liang, M.-C., Blake, G. A., and Yung, Y. L., 2008a, Seasonal cycle of C(16)O(16)O, C(16)O(17)O, and

C(16)O(18)O in the middle atmosphere: Implications for mesospheric dynamics and biogeochemical sources and sinks of CO(2): Journal of Geophysical Research-Atmospheres, v. 113, no. D12.

Liang, M.-C., Tang, J., Chan, C.-Y., Zheng, X. D., and Yung, Y. L., 2008b, Signature of stratospheric air at the

Tibetan Plateau: Geophysical Research Letters, v. 35, no. 20.

Liang, M. C., Blake, G. A., and Yung, Y. L., 2004, A semianalytic model for photo-induced isotopic fractionation in simple molecules: Journal of Geophysical Research-Atmospheres, v. 109, no. D10.

Luz, B., and Barkan, E., 2000, Assessment of oceanic productivity with the triple-isotope composition of dissolved oxygen: Science, v. 288, no. 5473, p. 2028-2031.

Luz, B., Barkan, E., Bender, M. L., Thiemens, M. H., and Boering, K. A., 1999, Triple-isotope composition of atmospheric oxygen as a tracer of biosphere productivity: Nature, v. 400, no. 6744, p. 547-550.

Lyons, J. R., 2007, Mass-independent fractionation of sulfur isotopes by isotope-selective photodissociation of

SO(2): Geophysical Research Letters, v. 34, no. 22.

Lyons, J. R., 2008, Fractionation of Sulfur and Oxygen isotopes during SO(2) photolysis: Geochimica Et

Cosmochimica Acta, v. 72, no. 12, p. A577-A577.

Lyons, J. R., 2009, Atmospherically-derived mass-independent sulfur isotope signatures, and incorporation into sediments: Chemical Geology, v. 267, no. 3-4, p. 164-174.

Lyons, J. R., Bergin, E. A., Ciesla, F. J., Davis, A. M., Desch, S. J., Hashizume, K., and Lee, J. E., 2009a,

Timescales for the evolution of oxygen isotope compositions in the solar nebula: Geochimica Et

Cosmochimica Acta, v. 73, no. 17, p. 4998-5017.

Lyons, J. R., Lewis, R. S., and Clayton, R. N., 2009b, Comment on "Experimental Test of Self-Shielding in Vacuum

Ultraviolet Photodissociation of CO": Science, v. 324, no. 5934.

Lyons, J. R., and Young, E. D., 2005, CO self-shielding as the origin of oxygen isotope anomalies in the early solar nebula: Nature, v. 435, no. 7040, p. 317-320.

MacPherson, G. J., 2008, Oxygen in the solar system, in Rosso, J. J., ed., Reviews in Mineralogy and Geochemistry,

Volume 68: Virginia, USA, The Mineralogical Society of America.

Mahata, S., and Bhattacharya, S. K., 2009, Anomalous enrichment of (17)O and (13)C in photodissociation products of CO(2): Possible role of nuclear spin: Journal of Chemical Physics, v. 130, no. 23.

Marcus, R. A., 2001, Does symmetry drive isotopic anomalies in ozone isotopomer formation? Response: Science, v. 294, no. 5544, p. 951A-951B.

-, 2002, Mass-independent isotope effects: Geochimica Et Cosmochimica Acta, v. 66, no. 15A, p. A484-A484.

-, 2004, Mass-independent isotope effect in the earliest processed solids in the solar system: A possible chemical mechanism: Journal of Chemical Physics, v. 121, no. 17, p. 8201-8211.

Matsuhisa, Y., Goldsmith, J. R., and Clayton, R. N., 1978, MECHANISMS OF HYDROTHERMAL

CRYSTALLIZATION OF QUARTZ AT 250-DEGREES-C AND 15 KBAR: Geochimica Et

Cosmochimica Acta, v. 42, no. 2, p. 173-&.

Mauersberger, K., 1981, MEASUREMENT OF HEAVY OZONE IN THE STRATOSPHERE: Geophysical

Research Letters, v. 8, no. 8, p. 935-937.

-, 1987, OZONE ISOTOPE MEASUREMENTS IN THE STRATOSPHERE: Geophysical Research Letters, v. 14, no. 1, p. 80-83.

Mauersberger, K., Erbacher, B., Krankowsky, D., Gunther, J., and Nickel, R., 1999, Ozone isotope enrichment:

Isotopomer-specific rate coefficients: Science, v. 283, no. 5400, p. 370-372.

Mauersberger, K., Krankowsky, D., and Janssen, C., 2003, Oxygen isotope processes and transfer reactions: Space

Science Reviews, v. 106, no. 1-4, p. 265-279.

42

Mauersberger, K., Krankowsky, D., Janssen, C., and Schinke, R., 2005, Assessment of the ozone isotope effect:

Advances in Atomic, Molecular, and Optical Physics, Vol 50, v. 50, p. 1-54.

McCabe, J. R., Boxe, C. S., Colussi, A. J., Hoffmann, M. R., and Thiemens, M. H., 2005, Oxygen isotopic fractionation in the photochemistry of nitrate in water and ice: Journal of Geophysical Research-Part D-

Atmospheres, v. 110, no. D15, p. 9 pp.

McCabe, J. R., Thiemens, M. H., and Savarino, J., 2007, A record of ozone variability in South Pole Antarctic snow: role of nitrate oxygen isotopes: Journal of Geophysical Research-Part D-Atmospheres, p. 1-9.

McKeegan, K. D., Kallio, A. P. A., Heber, V. S., Jarzebinski, G., Mao, P. H., Coath, C. D., Kunihiro, T., Wiens, R.

C., Nordholt, J. E., Moses, R. W., Jr., Reisenfeld, D. B., Jurewicz, A. J. G., and Burnett, D. S., 2011, The

Oxygen Isotopic Composition of the Sun Inferred from Captured Solar Wind: Science, v. 332, no. 6037, p.

1528-1532.

Michalski, G., Bockheim, J. G., Kendall, C., and Thiemens, M., 2005a, Isotopic composition of Antarctic Dry

Valley nitrate: implications for NO/sub y/ sources and cycling in Antarctica: Geophysical Research Letters, v. 32, no. 13, p. 4 pp.

-, 2005b, Isotopic composition of Antarctic Dry Valley nitrate: Implications for NOy sources and cycling in

Antarctica: Geophysical Research Letters, v. 32, no. 13.

Michalski, G., Bohlke, J. K., and Thiemens, M., 2004a, Long term atmospheric deposition as the source of nitrate and other salts in the Atacama Desert, Chile: New evidence from mass-independent oxygen isotopic compositions: Geochimica Et Cosmochimica Acta, v. 68, no. 20, p. 4023-4038.

Michalski, G., Jost, R., Sugny, D., Joyeux, M., and Thiemens, M., 2004b, Dissociation energies of six NO2 isotopologues by laser induced fluorescence spectroscopy and zero point energy of some triatomic molecules: Journal of Chemical Physics, v. 121, no. 15, p. 7153-7161.

Michalski, G., Savarino, J., Bohlke, J. K., and Thiemens, M., 2002, Determination of the total oxygen isotopic composition of nitrate and the calibration of a Delta O-17 nitrate reference material: Analytical Chemistry, v. 74, no. 19, p. 4989-4993.

Michalski, G., Scott, Z., Kabiling, M., and Thiemens, M. H., 2003, First measurements and modeling of Delta /sup

17/O in atmospheric nitrate: Geophysical Research Letters, v. 30, no. 16, p. ASC14-11-14.

Mittlefehldt, D. W., 1998, Non-chondritic meteorites from asteroidal bodies, in Papike, J. J., ed., Planetary

Materials, Volume 36, p. D1-D195.

Mittlefehldt, D. W., and Lindstrom, M. M., 1998, Petrology and geochemistry of large metaigneous inclusions from

L chondrites: Meteoritics & Planetary Science, v. 33, no. 4, p. A110-A111.

Mojzsis, S. J., Coath, C. D., Greenwood, J. P., McKeegan, K. D., and Harrison, T. M., 2003, Mass-independent isotope effects in Archean (2.5 to 3.8 Ga) sedimentary sulfides determined by ion microprobe analysis:

Geochimica Et Cosmochimica Acta, v. 67, no. 9, p. 1635-1658.

Mook, W. G., 1971, PALEOTEMPERATURES AND CHLORINITIES FROM STABLE CARBON AND

OXYGEN ISOTOPES IN SHELL CARBONATE: Palaeogeography Palaeoclimatology Palaeoecology, v.

9, no. 4, p. 245-&.

Mook, W. G., and Vogel, J. C., 1968, ISOTOPIC EQUILIBRIUM BETWEEN SHELLS AND THEIR

ENVIRONMENT: Science, v. 159, no. 3817, p. 874-&.

Morgan, C. G., Allen, M., Liang, M. C., Shia, R. L., Blake, G. A., and Yung, Y. L., 2004, Isotopic fractionation of nitrous oxide in the stratosphere: Comparison between model and observations: Journal of Geophysical

Research-Atmospheres, v. 109, no. D4.

Muskatel, B. H., Remacle, F., Thiemens, M. H., and Levine, R. D., 2011, On the strong and selective isotope effect in the UV excitation of N(2) with implications toward the nebula and Martian atmosphere: Proceedings of the National Academy of Sciences of the United States of America, v. 108, no. 15, p. 6020-6025.

Navon, O., and Wasserburg, G. J., 1985, SELF-SHIELDING IN O-2 - A POSSIBLE EXPLANATION FOR

OXYGEN ISOTOPIC ANOMALIES IN METEORITES: Earth and Planetary Science Letters, v. 73, no. 1, p. 1-16.

Nier, A. O., 1947, A MASS SPECTROMETER FOR ISOTOPE AND GAS ANALYSIS: Review of Scientific

Instruments, v. 18, no. 6, p. 398-411.

Nittler, L. R., 2003, Presolar stardust in meteorites: recent advances and scientific frontiers: Earth and Planetary

Science Letters, v. 209, no. 3-4, p. 259-273.

Nittler, L. R., and Alexander, C. M. O., 1999, Can stellar dynamics explain the metallicity distributions of presolar grains?: Astrophysical Journal, v. 526, no. 1, p. 249-256.

43

Ono, S., Beukes, N. J., Rumble, D., and Fogel, M. L., 2006a, Early evolution of atmospheric oxygen from multiplesulfur and carbon isotope records of the 2.9 Ga Mozaan Group of the Pongola Supergroup, Southern

Africa: South African Journal of Geology, v. 109, no. 1-2, p. 97-108.

Ono, S., Wing, B., Johnston, D., Farquhar, J., and Rumble, D., 2006b, Mass-dependent fractionation of quadruple stable sulfur isotope system as a new tracer of sulfur biogeochemical cycles: Geochimica Et Cosmochimica

Acta, v. 70, no. 9, p. 2238-2252.

Ozima, M., Podosek, F. A., Higuchi, T., Yin, Q. Z., and Yamada, A., 2007, On the mean oxygen isotope composition of the Solar System: Icarus, v. 186, no. 2, p. 562-570.

Paerl, H. W., 1993a, EMERGING ROLE OF ATMOSPHERIC NITROGEN DEPOSITION IN COASTAL

EUTROPHICATION - BIOGEOCHEMICAL AND TROPHIC PERSPECTIVES: Canadian Journal of

Fisheries and Aquatic Sciences, v. 50, no. 10, p. 2254-2269.

Paerl, H. W., 1993b, Interaction of nitrogen and carbon cycles in the marine environment: Aquatic microbiology: An ecological approach, p. 343-381.

Palomba, E., Zinzi, A., Clouds, E. A., D'Amore, M., Grassi, D., and Maturilli, A., 2009, Evidence for Mg-rich carbonates on Mars from a 3.9 mu m absorption feature: Icarus, v. 203, no. 1, p. 58-65.

Patris, N., Cliff, S. S., Quinn, P. K., Kasem, M., and Thiemens, M. H., 2007, Isotopic analysis of aerosol sulfate and nitrate during ITCT-2k2: Determination of different formation pathways as a function of particle size:

Journal of Geophysical Research-Atmospheres, v. 112, no. D23.

Pavlov, A. A., and Kasting, J. F., 2002, Mass-independent fractionation of sulfur isotopes in Archean sediments:

Strong evidence for an anoxic Archean atmosphere: Astrobiology, v. 2, no. 1, p. 27-41.

Perri, M. J., Van Wyngarden, A. L., Lin, J. J., Lee, Y. T., and Boering, K. A., 2004, Energy dependence of oxygen isotope exchange and quenching in the O(D-1)+CO2 reaction: A crossed molecular beam study: Journal of

Physical Chemistry A, v. 108, no. 39, p. 7995-8001.

Philippot, P., Van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J., and Van Kranendonk, M. J., 2007, Early

Archaean microorganisms preferred elemental sulfur, not sulfate: Science, v. 317, no. 5844, p. 1534-1537.

Rai, V. K., Jackson, T. L., and Thiemens, M. H., 2005, Photochemical mass-independent sulfur isotopes in achondritic meteorites: Science, v. 309, no. 5737, p. 1062-1065.

Rasch, P. J., Barth, M. C., Kiehl, J. T., Schwartz, S. E., and Benkovitz, C. M., 2000, A description of the global sulfur cycle and its controlling processes in the National Center for Atmospheric Research Community

Climate Model, Version 3 (vol 105, pg 1367, 2000): Journal of Geophysical Research-Atmospheres, v.

105, no. D5, p. 6783-6783.

Revesz, K., Bohlke, J. K., and Yoshinari, T., 1997, Determination of delta O-18 and delta N-15 in nitrate: Analytical

Chemistry, v. 69, no. 21, p. 4375-4380.

Runnegar, B., Coath, C. D., Lyons, J. R., and McKeegan, K. D., 2002, Mass-independent and mass-dependent sulfur processing throughout the Archean: Geochimica Et Cosmochimica Acta, v. 66, no. 15A, p. A655-A655.

Savarino, J., Lee, C. C. W., and Thiemens, M. H., 2000, Laboratory oxygen isotopic study of sulfur (IV) oxidation:

Origin of the mass-independent oxygen isotopic anomaly in atmospheric sulfates and sulfate mineral deposits on Earth: Journal of Geophysical Research-Atmospheres, v. 105, no. D23, p. 29079-29088.

Savarino, J., Slimane, B., Cole-Dai, J., and Thiemens, M. H., 2003, Evidence from sulfate mass independent oxygen isotopic compositions of dramatic changes in atmospheric oxidation following massive volcanic eruptions:

Journal of Geophysical Research, v. 108, no. D21, p. ACH7-1-6.

Savarino, J., and Thiemens, M. H., 1999a, Analytical procedure to determine both delta O-18 and delta O-17 of

H2O2 in natural water and first measurements: Atmospheric Environment, v. 33, no. 22, p. 3683-3690.

-, 1999b, Mass-independent oxygen isotope (O-16, O-17, O-18) fractionation found in H-x, O-x reactions: Journal of Physical Chemistry A, v. 103, no. 46, p. 9221-9229.

Schueler, B., Morton, J., and Mauersberger, K., 1990, MEASUREMENT OF ISOTOPIC ABUNDANCES IN

COLLECTED STRATOSPHERIC OZONE SAMPLES: Geophysical Research Letters, v. 17, no. 9, p.

1295-1298.

Shaheen, R., Abramian, A., Horn, J., Dominguez, G., Sullivan, R., and Thiemens, M. H., 2010, Detection of oxygen isotopic anomaly in terrestrial atmospheric carbonates and its implications to Mars: Proceedings of the

National Academy of Sciences of the United States of America, v. 107, no. 47, p. 20213-20218.

Shaheen, R., Janssen, C., and Rockmann, T., 2007, Investigations of the photochemical isotope equilibrium between

O-2, CO2 and O-3: Atmospheric Chemistry and Physics, v. 7, p. 495-509.

Shen, Y., Farquhar, J., Masterson, A., Kaufman, A. J., and Buick, R., 2009, Evaluating the role of microbial sulfate reduction in the early Archean using quadruple isotope systematics: Earth and Planetary Science Letters, v.

279, no. 3-4, p. 383-391.

44

Sigman, D. M., Casciotti, K. L., Andreani, M., Barford, C., Galanter, M., and Bohlke, J. K., 2001, A bacterial method for the nitrogen isotopic analysis of nitrate in seawater and freshwater: Analytical Chemistry, v. 73, no. 17, p. 4145-4153.

Smith, J. A., and Onstott, T. C., 2011, "Follow the Water": Steve Squyres and the Mars Exploration Rovers: Journal of the Franklin Institute-Engineering and Applied Mathematics, v. 348, no. 3, p. 446-452.

Smith, R. L., Pontoppidan, K. M., Young, E. D., Morris, M. R., and van Dishoeck, E. F., 2009, HIGH-PRECISION

C(17)O, C(18)O, AND C(16)O MEASUREMENTS IN YOUNG STELLAR OBJECTS: ANALOGUES

FOR CO SELF-SHIELDING IN THE EARLY SOLAR SYSTEM: Astrophysical Journal, v. 701, no. 1, p.

163-175.

Stehr, J., Krankowsky, D., and Mauersberger, K., 1996, Collection and analysis of atmospheric ozone samples:

Journal of Atmospheric Chemistry, v. 24, no. 3, p. 317-325.

Strasdeit, H., 2010, Chemical evolution and early Earth's and Mars' environmental conditions: Palaeodiversity, no.

3, Suppl. S, p. 107-116.

Taylor, F. W., 2010, Planetary atmospheres: Meteorological Applications, v. 17, no. 4, p. 393-403.

Thiemens, M. H., 1999, Atmosphere science - Mass-independent isotope effects in planetary atmospheres and the early solar system: Science, v. 283, no. 5400, p. 341-345.

-, 2002, Mass-independent isotope effects and their use in understanding natural processes: Israel Journal of

Chemistry, v. 42, no. 1, p. 43-54.

-, 2006, History and applications of mass-independent isotope effects: Annual Review of Earth and Planetary

Sciences, v. 34, p. 217-262.

Thiemens, M. H., and Heidenreich, J. E., 1983, The Mass-Independent Fractionation of Oxygen - a Novel Isotope

Effect and Its Possible Cosmochemical Implications: Science, v. 219, no. 4588, p. 1073-1075.

Thiemens, M. H., Jackson, T., Mauersberger, K., Schueler, B., and Morton, J., 1991, Oxygen Isotope Fractionation in Stratospheric Co2: Geophysical Research Letters, v. 18, no. 4, p. 669-672.

Thiemens, M. H., Jackson, T., Zipf, E. C., Erdman, P. W., and Vanegmond, C., 1995, Carbon-Dioxide and Oxygen-

Isotope Anomalies in the Mesosphere and Stratosphere: Science, v. 270, no. 5238, p. 969-972.

Tian, F., Claire, M. W., Haqq-Misra, J. D., Smith, M., Crisp, D. C., Catling, D., Zahnle, K., and Kasting, J. F., 2010,

Photochemical and climate consequences of sulfur outgassing on early Mars: Earth and Planetary Science

Letters, v. 295, no. 3-4, p. 412-418.

Ueno, Y., Johnson, M. S., Danielache, S. O., Eskebjerg, C., Pandey, A., and Yoshida, N., 2009, Geological sulfur isotopes indicate elevated OCS in the Archean atmosphere, solving faint young sun paradox: Proceedings of the National Academy of Sciences of the United States of America, v. 106, no. 35, p. 14784-14789.

Ueno, Y., Ono, S., Rumble, D., and Maruyama, S., 2008, Quadruple sulfur isotope analysis of ca. 3.5 Ga Dresser

Formation: New evidence for microbial sulfate reduction in the early Archean: Geochimica Et

Cosmochimica Acta, v. 72, no. 23, p. 5675-5691.

Urey, H. C., 1947, THE THERMODYNAMIC PROPERTIES OF ISOTOPIC SUBSTANCES: Journal of the

Chemical Society, no. MAY, p. 562-581.

Van Wyngarden, A. L., Mar, K. A., Boering, K. A., Lin, J. J., Lee, Y. T., Lin, S.-Y., Guo, H., and Lendvay, G.,

2007, Nonstatistical behavior of reactive scattering in the O-18+O-32(2) isotope exchange reaction: Journal of the American Chemical Society, v. 129, no. 10, p. 2866-2870.

Van Wyngarden, A. L., Perri, M. J., Mebel, A. M., Lin, J. J., Lee, Y. T., and Boering, K. A., 2004, Dynamics of

CO2+O(D-1) and implications for isotope exchange between O-3 and CO2: Abstracts of Papers of the

American Chemical Society, v. 228, p. U222-U222. van Zuilen, M., 2008, Stable isotope ratios as a biomarker on mars: Space Science Reviews, v. 135, no. 1-4, p. 221-

232.

Wasserburg, G. J., Busso, M., Gallino, R., and Nollett, K. M., 2006, Short-lived nuclei in the early Solar System:

Possible AGB sources: Nuclear Physics A, v. 777, p. 5-69.

Watanabe, Y., Farquhar, J., and Ohmoto, H., 2009, Anomalous Fractionations of Sulfur Isotopes During

Thermochemical Sulfate Reduction: Science, v. 324, no. 5925, p. 370-373.

Wen, J., and Thiemens, M. H., 1991, Experimental and Theoretical-Study of Isotope Effects on Ozone

Decomposition: Journal of Geophysical Research-Atmospheres, v. 96, no. D6, p. 10911-10921.

-, 1993, Multi-Isotope Study of the O((1)D)+Co2 Exchange and Stratospheric Consequences: Journal of

Geophysical Research-Atmospheres, v. 98, no. D7, p. 12801-12808.

Weston, R. E., 1999, Anomalous or mass-independent isotope effects: Chemical Reviews, v. 99, no. 8, p. 2115-

2136.

45

Wu, N., Farquhar, J., Strauss, H., Kim, S.-T., and Canfield, D. E., 2010, Evaluating the S-isotope fractionation associated with Phanerozoic pyrite burial: Geochimica Et Cosmochimica Acta, v. 74, no. 7, p. 2053-2071.

Xue, D., De Baets, B., Botte, J., Vermeulen, J., Van Cleemput, O., and Boeckx, P., 2010, Comparison of the silver nitrate and bacterial denitrification methods for the determination of nitrogen and oxygen isotope ratios of nitrate in surface water: Rapid Communications in Mass Spectrometry, v. 24, no. 6, p. 833-840.

Yi Qin, G., and Marcus, R. A., 2001, Strange and unconventional isotope effects in ozone formation: Science, v.

293, no. 5528, p. 259-263.

-, 2002, On the theory of the strange and unconventional isotopic effects in ozone formation: Journal of Chemical

Physics, v. 116, no. 1.

Yi Qin, G., Wei-Chen, C., and Marcus, R. A., 2002, A theoretical study of ozone isotopic effects using a modified abinitio potential energy surface: Journal of Chemical Physics, v. 117, no. 4.

Yin, Q.-Z., Shi, X., Chang, C., and Ng, C.-Y., 2009, Comment on "Experimental test of self-shielding in vacuum ultraviolet photodissociation of CO": Science (New York, N.Y.), v. 324, no. 5934, p. 1516.

Yung, Y. L., Demore, W. B., and Pinto, J. P., 1991, ISOTOPIC EXCHANGE BETWEEN CARBON-DIOXIDE

AND OZONE VIA O(1D) IN THE STRATOSPHERE: Geophysical Research Letters, v. 18, no. 1, p. 13-

16.

Yung, Y. L., Lee, A. Y. T., Irion, F. W., DeMore, W. B., and Wen, J., 1997, Carbon dioxide in the atmosphere:

Isotopic exchange with ozone and its use as a tracer in the middle atmosphere: Journal of Geophysical

Research-Atmospheres, v. 102, no. D9, p. 10857-10866.

Zerkle, A. L., Farquhar, J., Johnston, D. T., Cox, R. P., and Canfield, D. E., 2008, Fractionation of multiple sulfur isotopes during phototrophic S oxidation: Geochimica Et Cosmochimica Acta, v. 72, no. 12, p. A1075-

A1075.

Zmolek, P., Xu, X. P., Jackson, T., Thiemens, M. H., and Trogler, W. C., 1999, Large mass independent sulfur isotope fractionations during the photopolymerization of (CS2)-C-12 and (CS2)-C-13: Journal of Physical

Chemistry A, v. 103, no. 15, p. 2477-2480.

46

Download