The Oxygen Cycle of The Terrestrial Planets Insigh

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/228666216

The Oxygen Cycle of the Terrestrial Planets: Insights into the Processing and
History of Oxygen in Surface Environments

Article in Reviews in Mineralogy and Geochemistry · January 2007


DOI: 10.2138/rmg.2008.68.16

CITATIONS READS

21 2,763

2 authors:

James Farquhar David Thomas Johnston


University of Maryland, College Park Harvard University
277 PUBLICATIONS 14,738 CITATIONS 152 PUBLICATIONS 9,157 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by David Thomas Johnston on 25 December 2013.

The user has requested enhancement of the downloaded file.


Reviews in Mineralogy & Geochemistry
Vol. 68, pp. XXX-XXX, 2007 16
Copyright © Mineralogical Society of America

The Oxygen Cycle of the Terrestrial Planets:


Insights into the Processing and
History of Oxygen in Surface Environments
James Farquhar* and David T. Johnston
Department of Geology and ESSIC
University of Maryland, College Park
College Park, Maryland, 20742, U.S.A.
jfarquha@essic.umd.edu

ABSTRACT
Physical and chemical processes that operate at low temperature in Earth’s surface
environments and in other planetary settings have left an indelible record of the evolution and
interactions of planetary surface and near-surface reservoirs. This chapter reviews the chemistry
of oxygen (elemental and isotopic) in lunar, planetary, and Earth surface reservoirs. The discussion
begins with a brief review of the relative abundances of oxygen isotopes in planetary materials,
including lunar samples, the SNC (Martian) meteorites, and components that are inferred to
come from Martian surface pools. This is followed by a brief, largely historical account of the
development of oxygen isotope geochemistry in Earth-surface materials. The discussion then
transitions, using an Earth systems perspective, into an overview of developments in oxygen
isotope geochemistry of atmospheric compounds. This treatment includes recent studies that
(1) use high-precision measurements of δ17O and δ18O to characterize atmospheric oxygen;
(2) apply information about isotopically substituted species (isotopologs) for characterizing
the position-dependent isotopic fractionations of some oxygen-bearing atmospheric species;
and (3) investigate non mass-dependent isotopic fractionations in ozone and other atmospheric
species. The discussion then examines the evidence and models that point to change in the
oxygen cycle and oxidation chemistry in Earth surface environments over the course of geologic
time. We use the subdivisions of the Earth systems - the hydrosphere, geosphere and biosphere,
as a basis for examining these changes. This record is explored in the context of hypotheses and
data that relate to the nature of the changes and transformations in the oxidation state of these
different components of the Earth’s surface environments. We conclude with a brief accounting
of areas of research where scientific inquiry is presently very active.

INTRODUCTION
Before consideration can be given to the consequences of oxygen chemistry, it is valuable
to understand why terrestrial planets have their complement of this element and its three stable
isotopes (16O, 17O and 18O). Production of 16O began in the first generation of stellar helium
burning that occurred after the Big Bang. The establishment of the carbon-nitrogen-oxygen
(CNO) cycle of nucleosynthesis in subsequent generations of stars, however, has provided the
predominant source of oxygen (16O, 17O, and 18O) in the terrestrial planets. The most abundant
isotope of oxygen, 16O, has a particularly stable nuclear configuration, with 8 protons, 8
neutrons, and filled nuclear shells. On Earth it makes up ~99.76% of the total oxygen. The
next most abundant isotope, 18O, accounts for approximately 0.2%, and has an even number
of neutrons (10), while the rarest stable oxygen isotope, 17O, makes up the remaining ~0.04%
and has an odd number of neutrons (9).

1529-6466/07/0068-0016$05.00 DOI: 10.2138/rmg.2007.68.16

16_Farquhar_Johnston.indd 1 12/11/2007 11:16:12 PM


2 Farquhar & Johnston

Oxygen is one of the principal constituents of silicate minerals and, as a result, composes
a significant fraction of the mantles and crusts of the terrestrial planets. It is also a stable
component of many gases, liquids, and ices that may be present on the terrestrial planets,
including molecular oxygen (O2), carbon dioxide (CO2), water (H2O) and ozone (O3), and it is
an important component of the atmospheres, cryospheres, and hydrospheres of these planets.
The reactivity, or more specifically, the electronegativity of oxygen makes it a significant player
in determining the oxidation state of the surface environments of the terrestrial planets. The
proportion of molecular oxygen in the atmospheres of the terrestrial planets varies widely.
Mercury has only a tenuous atmosphere and is not considered here, and while the atmospheres
of Mars and Venus are substantial, they have only trace amounts of oxygen. On Mars, molecular
oxygen is produced by photolysis of carbon dioxide, leading to an atmospheric abundance of
approximately 0.13% (cf. Yung and Demore 1999; Lodders and Fegley 1998). On Venus, the
oxygen in the atmosphere is consumed by reactions with reduced species, and it is present in
only trace amounts (cf. Yung and Demore 1999; Lodders and Fegley 1998). On Earth, molecular
oxygen gas (O2) constitutes approximately 21% of dry air and is also dissolved into its oceans
at concentrations that vary from a few tens to a few hundred ppm. This abundance of oxygen
is a direct reflection of Earth’s biosphere. A prebiotic Earth would have had oxygen levels
controlled by reactions with reduced species, as they are on present-day Venus.
In this chapter, we review the chemistry, evolution, and isotopic chemistry and composition
of oxygen in the surface reservoirs (atmospheric, oceanic, biologic, and cryogenic). We begin
with a brief historical account of the background of oxygen isotopic studies of atmospheric and
oceanic oxygen with a focus on molecular oxygen, atmospheric species produced by oxygen
photochemistry, and water. We then examine the evidence and models that point to change
in the oxygen cycle and chemistry in Earth surface environments throughout geologic time,
and we discuss the oxidation states of the atmospheres and surface environments of Mars and
Venus. We use the principal subdivisions of the Earth systems- the hydrosphere, geosphere,
biosphere, and atmosphere as a basis for discussion of the oxygen isotopic variations of other
terrestrial planets (and Earth’s Moon) because it provides a framework that can be used to
compare and contrast the distinct behavior of oxygen isotopes on these different bodies.
We conclude with a brief listing of features of the oxygen isotope system that remain to be
documented and better understood.

ISOTOPIC VARIATIONS AMONG TERRESTRIAL MATERIALS


The variations in the relative abundances of oxygen isotopes in planetary materials arise as
a result of chemical, physical, and biological processes that occur in planetary environments and
from variations in the compositions of the original sources of oxygen. Resolvable differences in
the oxygen isotopic compositions of planetary materials have been revealed by measurements
of terrestrial samples, lunar samples, and meteorites. These variations occur for δ18O and Δ17O†,
and a fully consistent reconciliation of both parameters has been devised that captures the
genetic relationships between these different oxygen reservoirs (e.g., Lodders 2000), although
the reader is referred to studies of Lodders and Fegley (1997) and Sanloup et al. (1999) for
models that use oxygen isotopes as constraints on planetary accretion models for Mars. The
Earth and Moon lie within error of a single mass-fractionation array, known as the terrestrial
fractionation line (Clayton and Mayeda 1975; Wiechert et al. 2001), and are inferred to have

( O O) ( O O) ( O O)
† ⎛ 18 ⎞ ⎛ 17 ⎛ 18 ⎞
0.5247 ⎞
⎜ 16 ⎟ ⎜ 16 ⎜ 16 ⎟ ⎟
δ18O = 1000 × ⎜ − 1⎟ , Δ17 O = 1000 × ⎜ −⎜ ⎟ ⎟ (Miller 2002)
sample sample sample




( O O)
18
16
reference




⎜⎜

( O O)
17
16
reference



( O O)
18
16


reference ⎠

⎟⎟

16_Farquhar_Johnston.indd 2 12/11/2007 11:16:28 PM


Oxygen Cycle of Terrestrial Planets 3

sampled the same primordial oxygen reservoir. The isotopic heterogeneities for oxygen in
lunar materials define a significantly smaller range than those for terrestrial materials, and this
reflects the different types of physical and chemical processes that have acted on lunar oxygen
compared to terrestrial oxygen. The active hydrosphere and biosphere on Earth produce a large
and variable range of oxygen isotope fractionations compared to lunar materials (e.g., Onuma et
al. 1970, Epstein and Taylor 1970, Clayton and Mayeda 1975; Wiechert et al. 2001).
Analyses of SNC (shergottite-nakhlite-chassignite) meteorites, which are inferred
to have come from Mars on the basis of their trapped atmospheric gases, young geologic
ages compared to other meteorites, the unique Δ17O of silicate minerals they possess, and
dynamical arguments associated with the delivery of material from other evolved planets
(Ashwal et al. 1982; Bogard and Johnson 1983; McSween 1984, 1994), yield information
about the oxygen cycle of Mars. Significant differences exist between the oxygen isotopic
composition of minerals from SNC meteorites and terrestrial minerals. Measurements of
silicate minerals from SNC meteorites indicate that the silicate parts of Mars lie on a different
mass-fractionation line than the one that is defined for terrestrial silicate minerals (Clayton and
Mayeda 1983; Clayton and Mayeda 1996; Franchi et al. 1999). SNC silicate minerals have
higher 17O/16O and lower 18O/16O ratios than average Earth values. Unlike low-temperature
environments on Earth, analyses of water extracted from hydrous minerals (Karlsson et al.
1992; Romanek et al. 1998), oxygen in carbonate minerals (Farquhar et al. 1998; Farquhar and
Thiemens 2000), and in sulfate (Farquhar and Thiemens 2000) indicate that low-temperature
Martian minerals do not lie on the same mass-fractionation line as their silicate counterparts.
These oxygen isotopic variations have been interpreted to reflect the lack of plate tectonics;
the effects of liquid water; and a different (more active) role for atmospheric chemistry and
atmosphere-surface exchange in Martian surface environments. Microscale variations in δ18O
from carbonate minerals in the SNC meteorite ALH 84001 have been interpreted to reflect
low-temperature conditions for carbonate formation.
The oxygen isotopic compositions of Mercury and Venus remain to be determined. This
information will provide important constraints that may help to reconcile and understand the
δ18O and Δ17O variations among the Earth, Moon, Mars, and differentiated meteorites.
The most significant differences between the different planetary oxygen cycles are
summarized in Figure 1.1 The isotopic consequences of these differences can be summarized as
follows. First, materials from the Moon exhibit the smallest range of variation in δ18O values,
and essentially no variation of Δ17O values with the possible exception of implanted oxygen
from the solar wind. These variations reflect high-temperature oxygen isotope partitioning.
Second, materials from Earth preserve a significantly wider range of variation of δ18O values,
reflecting low-temperature processing of oxygen, fractionation effects associated with the
hydrologic cycle, and atmospheric processing of oxygen isotopes. The atmospheric processing
also imparts a significant range of variation in Δ17O values of atmospheric species and a small
Fig. 22). The
range of values for oxygen-bearing ions that readily exchange oxygen with water (Fig.
largest water reservoir on Earth, the ocean, has an isotopic composition that is determined on
long time-scales by exchange with the crust as a result of hydrothermal circulation at divergent
plate boundaries (a link to tectonic processes), with a superimposed, short time-scale response
resulting from evaporation and precipitation, and most notably to changes in ice volume. Third,
Fig. 3)
materials from Mars also preserve a large range of oxygen isotopic variations (Fig. 3 because
of low-temperature fractionation processes, but the absence of active tectonic processes and
an active hydrologic cycle leads to the efficient transfer of nonzero Δ17O values to products of
weathering such as carbonate minerals and secondary iddingsite.
Historical account of oxygen isotopic variations of terrestrial reservoirs
Due to its abundance, low mass, and large isotopic fractionations in the geologic and
meteoritic records, oxygen has inspired great interest, extending back to the 1930’s and 1940’s.

16_Farquhar_Johnston.indd 3 12/11/2007 11:16:28 PM


4

16_Farquhar_Johnston.indd 4
Farquhar & Johnston

Figure 1. Cartoons of the oxygen cycle on Earth, Mars, and the Moon. These sketches illustrate the sources and sinks of oxygen as well as the factors that determine the
isotopic variations that are observed in materials from each of these three planetary environments. The principal differences are the presence or absence of low-temperature
chemistry (atmospheric and hydrologic cycles), and the presence or absence of an active tectonic cycle. Low-temperature geochemistry determines the absolute magnitude
of mass-dependent and mass-independent fractionation effects. Tectonic cycling determines the long-term controls on the surface reservoirs through exchange with deep
planetary oxygen reservoirs.

12/11/2007 11:16:28 PM
Oxygen Cycle of Terrestrial Planets 5

Figure 2. Schematic plot of δ17O vs. δ18O for various terrestrial oxygen pools. Atmospheric pools are in grey
and non-atmospheric pools are in black. Pools that possess evidence for large isotopic fractionation effects by
mass-dependent processes include terrestrial precipitation, rocks and minerals, and atmospheric constituents
such as molecular oxygen and tropospheric carbon dioxide. Mass-independent fractionation effects are
observed in atmospheric ozone, nitrate, sulfate, carbon monoxide, hydrogen peroxide, nitrogen dioxide,
and, to a lesser extent, molecular oxygen and carbon dioxide. The expression of these different isotopic
fractionation effects in terrestrial materials is a direct reflection of the richness of Earth’s oxygen cycle. Note
that ocean water plots at the origin and is the standard to which all other materials are compared.

The physical chemistry of equilibrium isotopic fractionation was described in a series of papers
beginning with Urey and Greif (1935), and its introduction to earth sciences took place over the
next fifteen years. The development of high-precision mass spectrometric techniques by Nier
and others (Nier 1947; Nier et al. 1947), McKinney et al. (1950), and the coincident publication
of studies by Urey (1947) and Bigeleisen and Mayer (1947) introduced oxygen isotopes to
geochemistry and described the fundamental statistical thermodynamics of isotopic partitioning
in equilibrium processes. Previously published studies had described the different isotopic
compositions of oxygen in water and of atmospheric oxygen (Dole 1935) and attributed them
to biological cycling of oxygen by bacterial photosynthesis (Dole et al. 1947). Starting in the
1950’s, isotopic studies saw a rapid expansion in the number and types of measurements as well
as in the understanding of the oxygen isotopic compositions of geologically relevant materials.
The first high-precision measurements of the oxygen isotopic compositions of seawater were
undertaken (Epstein and Mayeda 1953), as were measurements of precipitation and groundwater,
meteorites, a wide variety of minerals (carbonates, silicates, and oxides), and rocks (igneous,
metamorphic, hydrothermally altered) (e.g., Emiliani 1955; Clayton and Epstein 1961; Taylor
and Epstein 1962; Clayton et al. 1973b; Taylor 1974; Knauth and Epstein 1976; Muehlenbachs
and Clayton 1976; Gregory and Taylor 1981). Oxygen isotope measurements were later

16_Farquhar_Johnston.indd 5 12/11/2007 11:16:29 PM


6 Farquhar & Johnston

Figure 3. Plot of Δ17O vs. δ18O for materials from SNC meteorites and for lunar samples. Data from Wiechert
et al. (2001) for lunar samples – black circles; from Franchi et al. (1999) for SNC meteorite silicates – black
diamonds; from Karlsson et al. (1992) for SNC water- open symbols; and from Farquhar and Theimens
(2000) and Farquhar et al. (1998) for SNC silicates, carbonates, and sulfate – open symbols). Only the
highest-temperature water extracts of Karlsson et al. (1992) are plotted here. The dashed grey lines connect
water and carbonate data points from the same meteorites and have slopes that are consistent with a mass-
dependent relationship between these materials. This plot illustrates the very small range of values observed
for lunar materials and the clear evidence for an atmospheric component in sulfate, carbonate, and water
from SNC meteorites. The transfer and preservation of these relationships has been interpreted to reflect
the structure of the oxygen cycle of Mars, which includes transfer of an atmospheric signal to carbonate
as a result of small amounts of water in the regolith, and the lack of a process that would exchange oxygen
between water and silicate, such as hydrothermal activity associated with plate tectonics.

extended to investigate the relationship between Pleistocene glaciations, ocean water volume,
and the oxygen isotopic composition of seawater, indicating a tie between the Earth’s climate
and the Sun (Emiliani 1955; Hays et al. 1976). Likewise, connections between high-temperature
hydrothermal systems and the long-term oxygen isotopic composition of seawater were made
and used to describe large-scale, long-term oxygen exchange between the hydrosphere and the
geosphere (Muelenbachs and Clayton 1976).
Clayton et al. (1973a) published measurements of the three oxygen isotopes for meteoritic
samples, forming a new focus for oxygen isotope research. In a series of studies that followed,
Clayton and coworkers documented anomalous oxygen isotopic compositions for a large
number of meteoritic phases (Clayton and Mayeda 1983, 1984, 1996, 1999; Clayton et al. 1976,
1991). While oxygen isotopic compositions of terrestrial minerals and water had been found
to fall on the terrestrial mass-dependent fractionation line (TFL: Δ17OSMOW ≡ 0, where SMOW
is Standard Mean Ocean Water), the isotopic compositions of many components of meteorites
were found to have different proportions of 16O, 17O and 18O, deviating significantly from this
mass-fractionation array. This work laid the foundation for oxygen isotopic studies of meteorites
by describing variations as a combination of mass-dependent fractionation arrays (of which the
TFL is but one) that are superimposed on a series of tightly constrained non-mass dependent
fractionations, where δ17OSMOW ≈ δ18OSMOW + Β (where B is a constant). With the exception of

16_Farquhar_Johnston.indd 6 12/11/2007 11:16:29 PM


Oxygen Cycle of Terrestrial Planets 7

interstellar grains and some trace species in Earth’s atmosphere, the high-temperature minerals
in the carbonaceous chondrites are the most strongly 16O-depleted materials in this array, and
the low-temperature minerals in some carbonaceous chondrites are the most 16O-enriched. Note
that the description of variations of 16O enrichments was originally chosen because it described
both the variations in the isotopic ratios and because it was consistent with a model to account
for these variations that included an 16O-enriched carrier phase. Other models that have since
been proposed, and are preferred by many workers, would imply correlated effects for 17O and
18
O rather than effects for 16O. The descriptions herein are intended to be model-independent
and to recognize the evolution of thought on this important subject.
The work of Clayton and coworkers on the oxygen isotopic variations in meteorites also
forms one of the fundamental bases of meteorite classification. A decade later, Thiemens and
Heidenreich (1983) published the results of an experimental study of ozone that demonstrated
a chemical process capable of producing non mass-dependent isotopic fractionations in
Fig. 4).
atmospheric gases (Fig. 4 In additional studies that followed, Thiemens and coworkers
documented non mass-dependent oxygen isotopic compositions (non-zero Δ17O) in a variety
of terrestrial atmospheric species, and ultimately provided evidence for the transfer of this
signature to minerals (especially salts) in soils, building crusts, glacial ice and snow (Thiemens
et al. 1995, 2001; Savarino and Thiemens 1999; Bao et al. 2000; Savarino et al. 2000, 2003;
Alexander et al. 2001; Lee and Thiemens 2001; Michalski et al. 2003; Thiemens 2006).
Observation of both mass-dependent and non mass-dependent fractionations in geologic and
planetary processes provides a new and exciting application of the geochemistry of oxygen to
models of planet formation and to its cycling in planetary environments.
In the past five years, there have been several new developments in oxygen isotope
research. These stem from our ability to make highly precise measurements of δ17O and
δ18O, and from our interests in developing applications for isotopically substituted species
(isotopologs) and for characterizing the position-dependent isotopic fractionations of some

Figure 4. Plot of δ17O vs. δ18O for


data from the ozone-formation
experiments of Thiemens and
Heidenreich (1983). Residual
oxygen (diamonds) possesses a
non mass-dependent depletion
of 17O and 18O and product
ozone (squares) possesses a non
mass-dependent enrichment of
17
O and 18O compared with the
starting composition (δ17O =
δ18O = 0 on this plot). The mass-
dependent fractionation line for
this system is plotted as the solid
black line.

16_Farquhar_Johnston.indd 7 12/11/2007 11:16:30 PM


8 Farquhar & Johnston

oxygen-bearing atmospheric species. Some recent studies have used highly precise and
accurate measurement of the three oxygen isotopes to revisit questions about the cycling
of oxygen between the atmosphere, biosphere, and hydrosphere (e.g., Luz et al. 1999;
Blunier et al. 2002; Angert et al. 2003; Hendricks et al. 2004; Luz and Barkan 2005). These
studies focus on the added information provided by three-isotope measurements of oxygen
to evaluate the transfer of small, non mass-dependent anomalies from ozone to carbon
dioxide, as well as other atmospheric species such as molecular oxygen (Luz et al. 1999);
the different mass-dependent fractionation laws that govern different chemical and physical
processes (e.g., Young et al. 2002); and the way that conservation of mass in multiple-step
processes can result in small deviations in Δ17O (e.g., Matsuhisa et al. 1978, 1979). The
conservation of mass effect arises because mixing between pools and removal of material
from pools follows a strictly linear relationship for δ17O and δ18O, while the distribution of
isotopes between pools and for the mass-fractionation line follow exponential laws. This
specific terminology has not been used to describe these effects for oxygen, but we use it here
to parallel descriptions of analogous effects that have been made for sulfur isotope effects
(Ono et al. 2006a, Farquhar et al. 2007a). A brief explanation of this effect is given in Figure
5. Below, we briefly review the literature that describes the isotopic chemistry of a variety
of terrestrial atmospheric species, with a focus on recent discoveries and new directions of
research. The discussion is not intended to be comprehensive, and the reader is referred to
excellent reviews on these and related topics (e.g., Thiemens 1999, 2005; Mook 2000; Alley
and Cuffey 2001; Brenninkmeijer et al. 2003). Here, we focus on specific aspects of the
isotopic chemistry of oxygen that have attracted recent attention, starting with a discussion
of molecular oxygen and ozone.

Figure 5. A plot of theoretical fractionations and mixing in δ17O and δ18O. The “linear mixing line” is
defined by fractional interaction of “A” to “B.” When the bulk composition is tracked (A+B), a linear
mixing line is defined connecting A and B. When the composition of the residual components is tracked,
the compositions extend towards more enriched or depleted values than the initial composition. The
fractionation array was calculated using an exponential definition. This figure demonstrates that regardless
of choice in Δ17O definition (linear or exponential), when mass mixing and fractionation are present, non-
zero Δ17O can be produced.

16_Farquhar_Johnston.indd 8 12/11/2007 11:16:30 PM


Oxygen Cycle of Terrestrial Planets 9

Molecular oxygen
The biogeochemistry and isotopic chemistry of the atmospheric oxygen cycle of the Earth
have been reviewed by Bender et al. (1994b), Thiemens (1999, 2005), Brenninkmeijer et al.
(2003), and Petsch et al. (2005). The oxygen cycle can be viewed from the perspective of atmo-
spheric molecular oxygen, the chemistry of other oxygen-bearing compounds, and the sources
and sinks of atmospheric oxygen. The largest single pool of atmospheric oxygen is molecular
O2, at approximately 21% of the atmosphere. Luz et al. (1999) report the isotopic composition
of atmospheric molecular oxygen to be 23.228‰ and 11.848‰ for δ18OSMOW and δ17OSMOW,
respectively. This reservoir is generally considered to be well-mixed due to turbulent mixing
within the atmosphere. Historical changes in the abundance of oxygen have been reported,
however, on seasonal, regional, and annual scales (e.g., Keeling and Shertz 1992; Battle et al.
1996), and reflect changes in its sources and sinks. The principal source of atmospheric mo-
lecular oxygen is photosynthesis by higher- and lower-level plants, whereas sinks for molecular
oxygen include respiration, oxidation of volcanic gases, mineral oxidation reactions, and fossil
fuel burning. The isotopic composition of molecular oxygen reflects a balance between oxy-
gen production by photosynthesis and sinks associated with respiration and the production of
ozone. Photosynthesis contributes a source of approximately 2-3 × 1016 moles of O2 per year
(Keeling and Shertz 1992, Bender et al. 1994b, Luz and Barkan 2005) with an isotopic com-
position that is thought to be very close to that of the water involved in the reaction (Guy et al.
1993). The principal control on the composition of atmospheric oxygen is associated with sink
reactions involving respiration, which preferentially removes the light isotope of oxygen and
follows a mass fractionation law of δ17O = 1000×[(1+δ18O/1000)0.52 – 1] (Angert et al. 2003;
Luz and Barkan 2005). This measurement is also interesting in light of recent work to identify
the source of the isotopic fractionation in the enzymatic process of oxygen respiration by glu-
cose oxidase (Roth et al. 2004).
In addition, recent work has used geochemical box models to evaluate the triple isotopic
composition of oxygen in air, gas bubbles trapped in ice cores, and water in the context of the
rates of gross and net oxygen production and consumption by respiration and other sinks (cf.
Luz et al. 1999; Blunier et al. 2002; Hendricks et al. 2005; Reuer et al. 2007). This work has
highlighted the new types of information that can be obtained from determination of both δ18O
and δ17O with high precision and is a particularly rich area of research that is providing new
constraints on global productivity. These constraints can be obtained because the isotopic com-
position of oxygen reflects the balance between the few processes that affect the Δ17O and δ18O
differently. Figure 6 illustrates the short-term processes that control the isotopic composition
of molecular oxygen, which include the production of ozone in the stratosphere, the cycling of
oxygen by photosynthesis and respiration in the biosphere, and the transfer of oxygen isotopic
anomalies to carbon dioxide, coupled with CO2-H2O oxygen isotope exchange in the biosphere.
In the longer term, oxygen concentrations are controlled by the balance between burial of or-
ganic carbon (a source) and the consumption of oxygen by reaction with reduced volcanic
gases, organic matter, and reduced minerals (principally iron, which rusts) in the crust (Claire
et al. 2006).
Ozone
While molecular oxygen (O2) is the principal oxygen-bearing species in the atmosphere,
another oxygen-bearing substance, ozone, has drawn considerable interest from isotope geo-
chemists and atmospheric chemists. Ozone is involved in a number of atmospheric reactions
involving trace constituents, and it has the potential to influence their isotopic compositions
through isotopic exchange and through the direct transfer of oxygen atoms. It has also been
known for over twenty years that the chemical reactions that form ozone involve large and
anomalous isotopic fractionations (Thiemens and Heidenreich 1983). Recent advances in our
theoretical understanding of the physical chemistry of ozone formation (cf. Gao and Marcus

16_Farquhar_Johnston.indd 9 12/11/2007 11:16:30 PM


10 Farquhar & Johnston

Figure 6. Cartoon of short-term controls on the isotopic composition of tropospheric oxygen (after Luz
et al. 1999). The isotopic composition of atmospheric oxygen reflects the balance between processes that
rapidly cycle the oxygen between the atmosphere and the biosphere and through ozone chemistry in the
stratosphere. High-altitude transfer of the isotopic signal from ozone to carbon dioxide by way of atomic
oxygen and subsequent exchange between carbon dioxide and water in the biosphere closes the short-term
oxygen cycle. CO3* indicates CO3 in an excited state.

2001; Babikov et al. 2003) have come about through a significant amount of high quality ex-
perimental work that used isotopically labeled compounds (Fig.
Fig. 7)
7 (Mauersberger et al. 1999).
These experiments, summarized in Figure 7, illustrated that the rates of formation for some
isotopic forms of ozone (isotopomers) obey an approximate mass-dependent relationship, while
rates involving the formation of other isotopomers do not. In short, there can be significant
differences in reaction rates of formation connected to molecular symmetry that can lead to
significant enrichments of some isotopomers over others.
The isotopic signature of ozone’s unusual chemistry is not only observed in laboratory
experiments (by UV, photolysis and spark discharge), but is also observed in the atmosphere.
There is also evidence for the transfer of this unique signature to other atmospheric species
(e.g., to atmospheric nitrate during the cycling of nitrogen species and to sulfate by oxidation
reactions), through both oxidation and exchange reactions. The oxygen isotopic compositions
of stratospheric ozone have been measured by a variety of different spectroscopic and mass
spectrometric techniques (Mauersberger 1981, 1987; Morton et al. 1989; Krankowsky et al.
1995), yielding compositions that range from ~65 to ~410‰ for δ18O, and from ~65 to ~130‰
for δ17O (note that some of the extreme measurements were for δ18O only). Collection of tropo-
spheric ozone for isotopic measurement is more difficult due to low natural abundances, though
two studies have reported that tropospheric ozone varies on a smaller scale than its strato-
spheric counterpart, with δ17O ranging between 66 and 78‰ and δ18O values from 82 to 90‰
(Krankowsky et al. 1995; Johnston and Thiemens 1997). One of the most useful applications of
the ozone measurements reported in the Johnston and Thiemens (1997) study was that the data
from Pasadena, CA differed significantly from that for any other site, and that this was a direct
consequence of the involvement and alteration of the reaction network by high NO number
densities. This finding leads to a direct ozone reaction and a secondary fractionation.

16_Farquhar_Johnston.indd 10 12/11/2007 11:16:30 PM


Oxygen Cycle of Terrestrial Planets 11

Figure 7. Illustration of the results of Mauersberger et al. (1999) for the relative isotopic fractionation
of ozone produced by different reaction pathways. The fractionations are normalized to the rates for
production of 16O3, and fractionation produced during formation of 17O3 (777) is approximately ½ the
fractionation produced during formation of 18O3 providing one indication of mass-dependency in the
relative rates. Mauersberger et al. (1999) describe other ways that the relative rates possess mass-dependent
relationships while also illustrating the striking non mass-dependent effects for species with different
isotopic substitutions. The numbers next to the bars indicate the identities of the isotopologues, by the last
digits of the three constituent O-isotopes.

Other oxygen-bearing atmospheric species with nonzero Δ17O


A number of other atmospheric species have been documented with nonzero Δ17O, and in
many cases this signal has been attributed to transfer from ozone. In others it has been attributed
either to a primary photochemical origin or to transfer from other species carrying nonzero Δ17O.
Several thorough reviews have discussed this subject (Thiemens 1999, 2005; Brenninkmeijer et
al. 2003). The oxygen isotopic compositions of oxygen-bearing species, including sulfate and
nitrate, have been studied using models of atmospheric chemistry and aerosol formation (e.g.,
Lyons 2001; Michalski et al. 2003; Alexander et al. 2005). Measurements of the triple-oxygen
isotope composition of carbon dioxide have been combined with concentration data for other
atmospheric trace species by Boering et al. (2004) to trace the origin and age of stratospheric
air masses. Michalski et al. (2003) documented large-magnitude Δ17O in atmospheric nitrate
that they attribute to nuclidic transfer between ozone and nitrogen-oxygen species during
the oxidation processes leading to the formation of nitrate. Observations of nonzero Δ17O in
sulfate of atmospheric origin have been studied by a number of researchers interested in the
reconstruction of atmospheric oxidation processes in past atmospheres (e.g., Alexander et al.
2003; Bao et al. 2003; Savarino et al. 2003; McCabe et al. 2007; Savarino et al. 2007).
Multiply substituted molecular species
An exciting group of new studies has focused on extracting information from the
distribution of isotopes between different nuclidic sites within given molecules in atmospheric
and non-atmospheric species. These studies have focused on the information that can be gained
by measurements of multiply substituted molecular species. Although the physical chemistry
of these effects has a history that extends to Bigeleisen and Friedman (1950) and McCrea
(1951), the insight and analytical capabilities required to apply them to planetary cycling of

16_Farquhar_Johnston.indd 11 12/11/2007 11:16:31 PM


12 Farquhar & Johnston

oxygen have only recently been realized. Yoshida and colleagues have developed novel ways
to use the intramolecular site preference for nitrogen isotopes in nitrous oxide to study its
sources (e.g., Yoshida and Toyoda 2000).
Eiler and colleagues have focused on the substitution of carbon and oxygen isotopes in
CO2 and have demonstrated that the abundance of multiply-substituted species depends upon
some of the same parameters, such as temperature, that determine singly substituted species,
and that the inclusion of these measurements with more traditional isotope ratios provides
new and independent constraints on the conditions of equilibration. These measurements have
provided some of the first single-molecule and single-mineral isotopic thermometers and have
the potential to circumvent some of the traditional assumptions that have been made in isotopic
thermometry. In addition, this approach expands the potential applications of thermometry. For
example, a recent application of this approach by Came et al. (2007) provides an independent
reconstruction of sea surface temperatures and suggests relatively stable δ18O of seawater
during the middle Paleozoic, and future studies with this approach promise data that may
resolve longstanding controversies about the long-term controls on the oxygen isotopic
composition of seawater. The origin and interpretation of these types of isotopic variations
continues to be a subject of active and exciting new research, promising new insights into the
role of oxygen in the modern system (Affek and Eiler 2006; Gosh et al. 2006; Schauble et al.
2006) and new applications in the analysis of atmosphere-ocean-biosphere systems.

EVOLUTION OF OXYGEN IN EARTH’S SURFACE ENVIRONMENTS


The present-day abundance of oxygen in Earth’s surface environments reflects a balance of
factors that include biological production by photosynthesis and consumption by respiration,
as well as longer-term sinks attributed to oxidative weathering and reaction with volcanogenic
reduced gases. The different response times for these sources and sinks allows the abundance
of oxygen in terrestrial environments to vary over a wide range of temporal and spatial scales.
Measurements of oxygen in surface environments point to short-term, as well as long-term,
variability in isotopic composition and concentration. Work on the oxygen concentrations in
the present-day atmosphere (described briefly above) by Keeling, Bender, and coworkers had
documented the small-magnitude, but profound, changes in atmospheric oxygen content that
are complementary to the historic rise in atmospheric carbon dioxide, and also reflects seasonal
changes in biological production and consumption of oxygen (e.g., Keeling and Shertz 1992;
Bender et al. 1994a; Battle et al. 1996).
The chemistry of the atmosphere, and as a consequence that of atmospheric oxygen, is also
influenced by incoming solar radiation and by numerous oxidation reactions. These oxidation
reactions also play a critical role in defining the concentrations of many trace atmospheric
species (e.g., carbon monoxide, methane, dimethyl sulfide, hydrogen sulfide). The types of
atmospheric reactions that involve oxygen have changed throughout Earth’s history as a result
of changes in the amounts of oxygen in the atmosphere. Some of these changes, however, are
only indirectly coupled to the absolute amount of molecular oxygen in the atmosphere. For
instance, while the atmospheric oxidation state is related to the atmospheric concentrations
of molecular oxygen, this relationship is not directly linear, with some models suggesting
that oxygen levels lower than the present one can, in fact, have a greater oxidizing potential.
This occurs because the oxidizing potential of the bulk atmosphere reflects the abundances
of other oxidized species, such as the OH radical, ozone and peroxides, in addition to O2.
The abundances of these oxygen-bearing compounds (most significantly ozone) affect UV
penetration into the atmosphere and the photochemistry that results. Such considerations may
be relevant when considering the evolution of Earth’s atmosphere, since it is widely accepted
that concentrations of molecular oxygen were lower on the ancient Earth, and the production

16_Farquhar_Johnston.indd 12 12/11/2007 11:16:31 PM


Oxygen Cycle of Terrestrial Planets 13

of oxidized species by interactions of other atmospheric gases with more intense, unshielded
solar deep UV (190-300 nm) may have controlled the atmospheric oxidation state.
On Earth, biological processes, global circulation, and air-sea exchange control the
chemistry of molecular oxygen in the oceans. The processes associated with hydrothermal
activity and submarine weathering also play important roles in the oceanic oxygen budget. The
typical concentration profile of oxygen in Earth’s oceans at low- and mid-latitudes exhibits high
concentrations in the surface water, diminishing oxygen at mid-depths, and relatively oxygen-
rich deep waters. Ocean-surface oxygen contents are controlled by photosynthetic oxygen
production in the photic zone. Between surface- and deep-ocean reservoirs resides an oxygen
minimum, thought to be due to the decay of sinking organic matter. At depth, deep water
formed at high latitudes and in open exchange with the atmosphere prior to sinking results
in relatively O2-rich deep waters. Thus, the distribution of oxygen in today’s oceans reflects
photosynthesis, respiration, and high-latitude air-sea transfer. This scenario may be unique
to the recent geologic past, however, as some models of the Archean and Proterozoic oceans
include stages during which deep water masses would have been anoxic. This is suggested in
light of inferred low atmospheric O2 concentrations during those periods, which would result
in formation of oxygen-poor deep waters, in addition to hydrothermal and biogenic sources of
reductants capable of removing the down-welled oxygen.
While life on Earth has had a profound impact on the abundance of oxygen in the
atmosphere and oceans, molecular oxygen and the products of oxygen chemistry play a direct
role in a number of Earth’s geological, geochemical, and geobiological systems. Oxygen is
essential for some metabolic processes, and thermodynamics illustrate that aerobic respiration
is much more efficient, generating thirty-eight ATP molecules for every two generated by its
anaerobic counterpart. The potency of oxygen is also reflected in its toxicity to many forms
of life, its interference with some key enzymatic processes (e.g., nitrogen fixation), and in the
generation of reactive oxygen products during aerobic respiration (e.g., peroxide, superoxide,
and other radical species). In addition to these biological relationships, oxygen also plays an
indirect role in making Earth habitable. Ozone, which is one of the products of atmospheric
chemistry involving oxygen, plays a critical role in Earth’s energy balance (the greenhouse
effect) and an essential role in the absorption of ultraviolet radiation that is harmful to many
forms of life. Ozone and other products of oxygen photochemistry also determine the lifetimes
of many atmospheric gases, such as methane, which can have a direct effect on the planetary
energy balance. The evolution of oxygen on Earth, and possibly on other planets, is one of
the fundamental factors that have determined how Earth’s surface environments, climate, and
biosphere have evolved.
Planetary processing of oxygen
While some of the factors that determined the abundances and distribution of oxygen in
the past are similar to those that determine them in today’s atmosphere and oceans, others may
have been markedly different. Oxygenic photosynthesis and oxygen respiration have played
a central role in controlling the oxygen levels of Earth’s atmosphere and oceans. It is also
clear that changes in the intensity of hydrothermal circulation and in the amounts of volcanic
activity may have had a profound impact on the composition of the atmosphere and oceans and
on the distribution and abundance of oxygen within them. Recent measurements of oxygen
and sulfur isotopes in ice cores (e.g., Alexander et al. 2002, 2003; Savarino et al. 2003) point
to changes in oxidation pathways for atmospheric sulfur and for other atmospheric species
from atmospheric chemistry, with the energy balance dictated by incoming solar radiation.
Chemical weathering reactions are thought to have played as important a role in the past as
they do today. The principal controls on atmospheric oxygen variability are thus biological
activity and geological processes, since these control the addition and removal of oxygen in
Earth’s surface environments. In addition to these, certain basic feedbacks are also inferred to

16_Farquhar_Johnston.indd 13 12/11/2007 11:16:31 PM


14 Farquhar & Johnston

have played important roles in stabilizing the concentrations and distribution of oxygen among
different components of the Earth system.
In today’s atmosphere, short-term variations in atmospheric oxygen content have been
reported on daily, seasonal, and yearly timescales (Keeling and Shertz 1992; Bender et al.
1994a; Keeling et al. 1996). These are attributed to variations in biological activity and to
burning of fossil fuels. In the longer term, variations in the concentrations of oxygen have been
attributed to geological processes such as variations in volcanic sources, weathering sinks,
and biological innovations such as the colonization of land by plants and the evolution of
oxygenic photosynthesis. In the following discussion, we will examine the geological and
geochemical evidence for changes in the content and distribution of oxygen in the atmosphere
and oceans, starting with the Archean and proceeding through the records of the Paleo-, Meso-
, and NeoProterozoic, into the Phanerozoic. We will examine the hypotheses about these times
in Earth history, and discuss some of the areas where incompatible hypotheses exist.

OBSERVATIONS RELEVANT TO THE EVOLUTION OF


OXYGEN IN THE ATMOSPHERE AND OCEANS
Hypotheses about the levels of oxygen in Earth’s early environments
The prevailing view of atmospheric evolution includes the hypothesis that geologic
evidence (described below) points to low abundances of atmospheric oxygen early in Earth’s
history. The following discussion will focus on this view.
The formation of Earth’s mantle is thought to have established a redox state for the
early atmosphere that would not have been strongly reducing (hydrogen-rich), but would
instead have been neutral (neither reducing nor oxidizing), with low amounts of oxygen and a
balance consisting predominantly of nitrogen and carbon dioxide, with lower abundances of
volcanogenic methane. Gases that are subject to oxidation in the present atmosphere would
have had longer lifetimes in an oxygen-poor early atmosphere unless they were subject to
more rapid photodissociation in an atmosphere that did not shield deep ultraviolet radiation.
The hydrogen content and the presence of hydrogen-bearing compounds is thought to have
been higher under these conditions than at present, and a number of models of atmospheric
evolution suggest that hydrogen loss to space would have led to a gradual shift in the proportions
of reduced components in the atmosphere-ocean system (Catling and Claire 2005; Tian et al.
2005) – irreversible oxidation.
Arguments for an oxygen-poor early atmosphere take several different forms. Some
arguments have been made on the basis of the sources and sinks of oxygen to the atmosphere-
ocean system. These include arguments that before the advent of oxygenic photosynthesis,
the free oxygen in Earth’s surface environments would have been titrated by reactions with
volcanic gases (Holland 2002) and other reduced species in the environment and ocean. The
sinks for oxygen in such an environment have led to the suggestion that oxygen contents in the
early atmosphere would have stabilized at levels 10-12 to 10-13 of present atmospheric levels
(PAL) (Kasting et al. 1993; Catling and Claire 2005).
The geological evidence for low levels of oxygen early in Earth’s history includes
evidence for loss of iron and other redox-sensitive elements from paleosols (e.g., Holland
1984; Rye and Holland 1998). A systematic study by Rye and Holland (1998) argued that
soil profiles older than 2.2 Ga exhibit iron depletions that are consistent with their formation
under oxygen-poor conditions. The basis of the Rye and Holland (1998) argument, and of
similar arguments that preceded theirs (Cloud 1968; Holland 1984) is that in low-oxygen
conditions, reduced iron can be leached from soils in the aqueous phase, while in oxygen-rich
conditions, oxidized iron is fixed in the soils as hematite, Fe2O3. Iron mobility is enhanced

16_Farquhar_Johnston.indd 14 12/11/2007 11:16:32 PM


Oxygen Cycle of Terrestrial Planets 15

in low-oxygen environments because Fe2+ is significantly more soluble in water than Fe3+.
Similar arguments link the formation of banded iron formations (BIF) early in Earth’s history
to low oxygen levels (e.g., Cloud 1968). Iron formations formed during specific stages in Earth
history, and with only a few exceptions, are older than 1.8 Ga. The frequency and amount of
iron formations peak at about 2.5 billion years ago and their occurrence in the geologic record
is attributed to an ocean that contained abundant soluble reduced iron (e.g., Holland 1984).
It was recognized that widespread iron solubility in the oceans would have been required
to allow for the mobilization and concentration of iron necessary to form BIF. Another line
of geological evidence for low oxygen contents is the preservation of detrital pyrite (FeS2)
and uraninite (UO2) in deposits of the Witwatersrand. This occurrence, in Archean sediments
(older than 2.5 by) is taken as an indication that oxidative weathering was suppressed during
the Archean. In modern environments, pyrite oxidizes (rusts) to form hematite and sulfate,
and uraninite oxidizes to form a water-soluble uranium complex (UO22+). Isolated pyrite and
uraninite pebbles occur in the present oxygenated surface environment, but are not common.
The disappearance of banded iron formations at 1.8 Ga has also been attributed to a change
in ocean chemistry that resulted in the titration of reduced iron by biogenic sulfide. Greater
production of biogenic sulfide by sulfate reducers is thought to have occurred because of
higher oceanic sulfate concentrations that were a direct consequence of oxidative weathering
reactions. The connection between weathering, the atmosphere, and the oceans has been used
to suggest that this change also reflects a change in atmospheric oxidation state.
One line of geochemical evidence that has been used to study atmospheric oxygen
contents may, at first, appear to be a little strange. The sulfur isotope record of sedimentary
rocks preserves a strong and clear mass-independent signal throughout much of the Archean
(Fig.
Fig. 8).
8 This signal has been attributed to atmospheric chemistry because correlations between

10

4
33S

-2

-4
0 1000 2000 3000 4000
Age (Millions of years ago)

Figure 8. A compilation of published values of Δ33S versus age for terrestrial sedimentary rocks. The
record has been divided into three stages. Stage 1 extends from 4 Ga to 2.4 Ga, and defined by significant
non-zero Δ33S. Stage 2 (2.4-2.0 Ga) represents a transition period between preservation of large non-zero
Δ33S and near-zero Δ33S (Stage 3: 2.0-present). We highlight the structure within Stage 1, with the largest
effect between 2.4-2.8 Ga., small but resolvable non-zero values from 2.8-3.0 Ga., and a middle range of
values for ages > 3.0 Ga. Data in this compilation are taken from Farquhar et al. (2000, 2002); Mojzsis
et al. (2003); Hu et al. (2003); Ono et al. (2003, 2006a, 2006b); Bekker et al. (2004); Papineau (2005);
Whitehouse (2005); Ohmoto et al. (2006); and Johnston et al. (2005, 2006).

16_Farquhar_Johnston.indd 15 12/11/2007 11:16:32 PM


16 Farquhar & Johnston

Δ33S and Δ36S suggest that it is not a magnetic (hyperfine) isotope effect, but a gas-phase
effect. Hyperfine effects are a class of isotope effects that occur for odd-mass isotopes because
chemical reaction pathways that are not allowed because the electronic spin state is incompatible
can be preferentially opened or ‘allowed’ as a result of magnetic coupling between nuclear
and electronic spin. These effects occur in photochemical liquid-phase chemistry when the
time scales for spin-coupling are on the order of the timescales for transport. Timescales for
transport in the gas phase are generally too short for these effects to be prevalent. The non
mass-dependent isotope effects that occur for even-mass isotopologs are only known to occur
in the gas phase and derive from symmetry-selective chemistry, or chemistry associated with
rovibrational overlap of excited states.
The specific chemical reaction that produced the effect observed in Archean samples
remains unknown, but experiments with sulfur dioxide suggest it is a likely candidate (Farquhar
et al. 2000b, 2001; Farquhar and Wing 2003). The connection to atmospheric oxygen content
is argued using two separate lines of reasoning. If one accepts that chemistry associated
with sulfur dioxide photolysis is the source of the effect, then the deep ultraviolet radiation
necessary for this photochemistry would be required. For this to be the case, the UV-shielding
ozone layer could not be part of the atmospheric system. The link to oxygen contents arises
because the ozone shield requires a minimum oxygen content in order to exist. A second line of
reasoning is independent of the specific source of the mass-independent sulfur isotope effect,
but would still require it to originate in the atmosphere. Pavlov and Kasting (2002) argued
that the presence of a large atmospheric signal in the rock record necessitates a process that
delivers sulfur to the surface by at least two distinct pathways, with deposition of elemental
sulfur and reduced sulfur species defining one pathway, and sulfate and oxidized sulfur species
populating the other pathway. They also noted that the number of pathways available for
atmospheric deposition of sulfur to the surface depends on the atmospheric oxygen content.
This arises because oxygen-bearing species (atomic oxygen, molecular oxygen, carbon
dioxide) react with and consume the neutral sulfur species by oxidation reactions, and limit
the depositional pathways above a threshold atmospheric oxygen content, but allow multiple
pathways to open up below this threshold. The level that Pavlov and Kasting (2002) suggest
is approximately 10−5 of present atmospheric levels of oxygen. This upper limit is well above
the values of 10−12 to 10−13 predicted by models (e.g., Kasting 1993; Catling and Claire 2005),
but also provides one of the most sensitive constraints on paleoatmospheric oxygen that we
presently have. These measurements of sulfur isotopes provide one of the strongest and best
lines of evidence that atmospheric oxygen contents were low during much of the Archean
(Farquhar et al. 2000a; Farquhar and Wing 2003, 2005).
The sulfur isotope record presented in Figure 8 shows that the largest excursions occurred
in the late Archean and were preceded by a dampening of the signal for the interval between
2.8 and 3.0 Ga. The signal before 3.0 Ga is large in magnitude, but not as large as the late
Archean signal. Despite recent arguments that the 2.8-3.0 Ga isotope record indicates increases
in atmospheric oxygen levels and the establishment of a mass-dependent sulfur cycle (Ohmoto
et al. 2006), re-evaluation of the δ34S versus Δ33S co-variation suggests that they are, in fact,
mass-independent and consistent with low oxygen levels throughout the Archean (Fig. Fig. 9).
9 Low
atmospheric oxygen contents at that time are further supported by localities of similar age in
the Witwatersrand basin (e.g., Ono et al. 2006). We conclude that the sulfur isotope data from
Ohmoto et al. (2006) carry a non-mass-dependent signal, but also that there are intervals within
the Archean where the magnitude of the sulfur isotope signal varies. Recently published data
for samples from this interval for Δ36S (Farquhar et al. 2007b) suggest that its relationships
with Δ33S vary in this Mesoarchean interval and have been interpreted as further evidence
for non mass-dependent signals and low oxygen. This may reflect changes in atmospheric
chemistry, either through changes in the UV spectrum available to drive the chemistry, or
through changes in the reaction pathways brought about by changes in the relative abundances

16_Farquhar_Johnston.indd 16 12/11/2007 11:16:33 PM


Oxygen Cycle of Terrestrial Planets 17

33S (‰)
1.5

0.5

34S (‰)

-60 -40 -20 20 40 60

-0.5

-1
Data for samples younger than 2.0 Ga (mass-dependent)
Data for samples older than 2.4 Ga (non mass-dependent component)
Data for Hardey Fm (Ohmoto et al., 2006) (non mass-dependent)
Data for Mosquito Creek Fm (Ohmoto et al., 2006) (non mass-dependent)

Figure 9. A compilation of published values for Δ33S versus δ34S. The dataset has been divided into four
subgroups. The solid grey circles represent samples younger than 2.0 Ga, and illustrate the field that is
attributed to mass-dependent isotopic fractionation effects. Note that measurable variations in Δ33S are
observed for this subgroup, but that the variations are correlated with δ34S. The second subgroup (open
circles) consists of samples older than 2.45 Ga and is interpreted as reflecting non-mass-dependent
chemical fractionation effects associated with atmospheric reactions in a low-oxygen atmosphere. The
third and fourth groups, black squares and black diamonds, are from the 2.76 Ga Hardey Fm. and the 2.92
Ga Mosquito Creek Fm., respectively. These data form arrays that do not overlay the mass-dependent
phase space and exhibit the same sense of variation as the other pre 2.45 Ga samples (grey circles). On
consideration of both Δ33S and δ34S, these data are more consistent with a non mass-dependent interpretation
than with the post 2.0 Ga mass-dependent subset. Note that Ohmoto et al. (2006) interpreted these data as
mass-dependent, but did so without consideration of the relationships between Δ33S and δ34S Data in this
compilation taken from Farquhar et al. (2000, 2002); Mojzsis et al. (2003); Hu et al. (2003); Ono et al.
(2003, 2006a, 2006b); Bekker et al. (2004); Papineau (2005); Whitehouse (2005); Ohmoto et al. (2006);
and Johnston et al. (2005, 2006).

of one or more atmospheric species. The role played by oxygen in the atmospheric sulfur cycle
makes it a prime candidate for further investigation.
The record of large-magnitude Δ33S values for sedimentary samples terminates abruptly
at approximately 2.4 Ga. The timing of the disappearance of this signal also correlates with the
first appearance of glacial diamictites in the early Paleoproterozoic and points to the possibility
of a connection between oxygen abundance and climate. There are several different glacial
events at ~2.4 Ga and it is not known whether closer examination of the record will support
correlation of oxygen concentrations with glaciations. The lifetime of the residual non-zero
Δ33S is unknown. Regardless, the change from an oxygen–poor atmosphere to an oxygen-rich
one was a significant event in Earth’s history and fundamentally changed its atmospheric,
oceanic, and biological character.
Whereas the evidence from sulfur isotopes, banded iron formations, paleosols, and detrital
minerals all indicate that heightened levels of atmospheric oxygen did not become established

16_Farquhar_Johnston.indd 17 12/11/2007 11:16:33 PM


18 Farquhar & Johnston

until ~2.4 Ga, evidence exists for locally oxygenated conditions preceding the global signature.
The strongest evidence for local oxidation comes from biomarkers, with evidence for sterol
biosynthesis that would have required O2 at ~2.7 Ga (Summons et al. 1999; Brocks et al. 2003).
Summons et al. (2006) argue that eleven molecules of O2 are required for the synthesis of one
molecule of cholesterol. Experimental studies have shown that bacterial sulfate reduction yields
large δ34S fractionations when sulfate concentrations exceed a threshold of a few hundred mi-
cromolar (Habicht et al. 2002), and the observation of large δ34S fractionations in the geologic
record has been interpreted to reflect higher oceanic sulfate concentrations, and therefore a
source of oxidized sulfur. These observations have been interpreted by some to indicate local
oxygen oases. The size and scale of such oases may have been very limited (e.g., to a few cen-
timeters or less in a microbial mat, or to particular embayments), but they may, at some point,
have extended to portions of the surface ocean containing modest levels of oxygen, produced
locally by oxygenic photosynthesis. The presence of BIF through this period suggests that iron
remained soluble (and thus reduced) on the global scale, suggesting that any oxygenation of the
surface environment must have been restricted. Taken together, these lines of evidence suggest
that considerations of oxygen must now move beyond the simple atmosphere-ocean perspec-
tive, and begin to consider the different horizons within these reservoirs. A fundamental unan-
swered question is whether the 300 million-year lag between the evidence for oxygenic biosyn-
thesis and an oxygenated atmosphere, or the almost one billion-year lag between evidence for
local sulfate oases and an oxygenated atmosphere are real, and if so, why these lags occurred.
A possible resolution might be that the production of oxygen alone cannot tip the scale toward
atmospheric oxygenation, but instead, that a change in the total oxygen in the surface system
effected by both production and introduction from biological sources and also destruction and
loss to sinks would have been required (see Catling and Claire 2005; Tian et al. 2005).
The transition from a low-oxygen atmosphere to a high oxygen atmosphere
The advent of oxygenic photosynthesis is thought to have resulted in striking changes
in the atmosphere, oceans, and biosphere of the early Earth. The energy yield of oxygenic
photosynthesis and oxygen respiration are considerably greater than for anoxygenic
photosynthesis and anaerobic respiration (Madigan et al. 2000), and the difference in metabolic
yield is thought to have had a significant impact on the total terrestrial biomass and also on
the total mass of reduced organic carbon. Oxygenic photosynthesis also provides a significant
source of oxygen to the atmosphere. It has been pointed out that present concentrations of
atmospheric oxygen will only be maintained while photosynthetic oxygen sources continue, and
that the lifetime of oxygen in the atmosphere (against oxidative weathering and reaction with
reduced gases) is on the order of 106 yr (cf. Claire et al. 2006). The introduction of even small
amounts of oxygen to the system would have had a profound effect on the lifetimes of reduced
species in Earth’s early atmosphere, and once these gases were themselves titrated by oxygen,
instead of vice versa, their relative abundances would fall to levels that were determined by their
lifetime against oxidation. It has been pointed out that some of these gases (e.g., methane) may
have played a significant role in stabilizing climatic conditions on early Earth, and as a result,
the advent of atmospheric oxygen may have had a significant effect on Earth’s greenhouse
(Bekker et al. 2003).
How the oxygen content of the atmosphere rose is another question that has occupied re-
searchers of Earth’s early environments. The current thinking is that atmospheric oxygen levels
reflect a steady-state balance between sources and sinks, and that some atmospheric states are
more stable than others. One of these states, described above, is an atmosphere in which oxygen
levels are controlled by the oxidation state of volcanic gases. Another state for the atmosphere
is thought to exist when oxygen sources outpace the reductive capacity of volcanic gases, and
instead the levels are controlled by mineral oxidation reactions. This second state is thought to
be the state of the present atmosphere-ocean system. Other states have been proposed, but have

16_Farquhar_Johnston.indd 18 12/11/2007 11:16:36 PM


Oxygen Cycle of Terrestrial Planets 19

not been explored in detail. These alternate states include one where the deep oceans act as a
sink to maintain the atmosphere at an oxygen level between that maintained by photosynthesis
and volcanic gases, and that maintained by photosynthesis and oxidative weathering. Although
it is conceivable that such a state might exist for a sulfidic or iron-rich deep ocean (Canfield
1998), we are unaware of studies that have explored the stability of such a system.
Another question that has interested researchers relates to the transition between states of
different atmospheric and oceanic oxygen levels. Generally, these transitions are thought to
be nearly geologically instantaneous, or are considered a step-wise function, because once the
reducing capacity of one sink is exceeded, the oxygen abundance should increase until the next
sink balances the oxygen production and establishes a new steady-state. Although the transitions
may have been unidirectional and not reversed, the possibility also exists that feedbacks in the
atmosphere-ocean-biosphere-climate system may have caused it to toggle between two or more
states. In this context, recent studies by Anbar et al. (2007) and Kaufman et al. (2007) have
suggested that an initial “whiff” of oxygen appeared in the atmosphere and oceans and preceded
the great oxygenation event (GOE) at ~2.4 Ga.
Arguments have also been made that a change in the nature of the sinks for oxygen in
Earth’s surface oxygen cycle may have changed the oxidation state of the surface environments
(e.g., Kump et al. 2001; Holland 2002; Catling and Claire 2005; Holland 2006). One such
change would be a shift in the oxidation state of the mantle. This has proven difficult to confirm
because of absence of evidence for a large-scale change in the redox state of the mantle (e.g.,
Canil 2002; Li and Lee 2004), and it has been difficult to refute because the magnitude of
change in redox state of the volcanic gases required to drive a change in atmospheric oxidation
state is small (Kump et al. 2001; Holland 2002). Other workers have argued that a change in
the atmospheric oxidation state occurred as a result of a combination of a gradual shift in the
inventory of reduced and oxidized compounds in the atmosphere-ocean system that occurred
because of long-term loss of hydrogen (e.g., Catling and Claire 2005; Tian et al. 2005), and the
production of oxygen by photosynthesis. A geochemical test to select the best hypothesis has
yet to be devised but is a significant target for future work.
Into the Paleoproterozoic and Mesoproterozoic
The change in the oxygen content of the atmosphere at ~2.4 Ga, and the associated
glaciations, ushered in a new period of Earth history, one marked by changes in ocean
chemistry and by the eventual rise of large animals. Studies of the Proterozoic (2.5 to 0.5
Ga) have focused on the geochemistry of shallow and deep oceans and the links between
ocean chemistry, ecology, and the oxidation of Earth’s surface environments. The geochemical
approaches employed to study the chemistry and ecology of the oceans during this period of
time have centered around isotopic systems (carbon, sulfur, molybdenum, and iron) as well as
work focusing on the extraction and interpretation of molecular biomarkers.
Pervasive oxygenation of the Earth’s atmosphere was a new feature for the surface of the
Earth only after the ~2.4 Ga transitions. The period when oxygen rose was coincident with
glaciations recorded in the Huronian deposits of Canada, the Transvaal of South Africa, and in
sequences in Australia. The aftermath of these glaciations saw an ocean much richer in oxidized
compounds than before, and significant increases in seawater sulfate contents are inferred on
the basis of larger δ34S fractionations. Current thinking points to increases in sulfate concentra-
tions that accompanied the oxygenation of Earth’s surface environments, and inferences from
experiments with cultures of sulfate reducers provide constraints used to suggest that sulfate
concentrations rose from < 0.20 mM before the oxygenation to > 0.20 mM in the post-oxygen-
ation ocean (Habicht et al. 2002). The sulfur isotope record of sedimentary sulfides (Canfield
2001, 2005), in combination with measurements of carbonate-associated sulfate (e.g., Strauss
1993; Canfield 1998), has provided a record of sulfur isotope fractionations that has been used

16_Farquhar_Johnston.indd 19 12/11/2007 11:16:36 PM


20 Farquhar & Johnston

to study the oceanic cycles of sulfur and to provide insight into the oxidation state of the shallow
and deep oceans. It is inferred that changes in the concentration of sulfate in seawater mirror
changes in other oxidized species, such as nitrate and phosphate, and that higher concentrations
of oxidized species would have had an impact on the metabolisms of the time, not only by pro-
viding abundant pools of oxidants to fuel metabolisms, but also by challenging the biological
machinery (such as O2-sensitive nitrogenase) that had dominated the largely anoxic pre-2.4 Ga
world. While some anaerobes may have adapted to a changing world, oxygen-tolerant microor-
ganisms were provided with new ecosystems in which they could diversify and flourish.
Oxygen and Proterozoic carbon cycle
One of the traditional links between the oxygen cycle and the carbon cycle is through
burial of organic carbon. The production of oxygen by photosynthesis leads to the produc-
tion of reduced carbon, and the burial of this carbon leads to both the build-up of oxygen and
a change in δ13C of the carbon remaining in Earth’s atmosphere-ocean-biosphere system. A
series of significant excursions to positive δ13C in the early parts of the Paleoproterozoic has
been interpreted to have heralded a significant oxygen pulse (Karhu and Holland 1996; Bekker
et al. 2003). Traditionally, the level of atmospheric oxygen is thought to be modulated by the
relative proportion of organic carbon (OC) burial. When large proportions of organic material
are buried, the complementary oxidizing potential, generally taken as atmospheric O2, is al-
lowed to accumulate. This mechanism is of special interest in the Paleoproterozoic, a period for
which multiple lines of evidence suggest widespread oxidation occurred, which was followed,
rather than initiated, by large-scale variability in the carbon isotope record over the ensuing
400 Ma. Interest in isotopic records of the Paleoproterozoic was spurred by the observation of
13
C-enriched carbonates (Karhu and Holland 1996). These authors suggested that large propor-
tions of organic carbon must have been buried to satisfy the isotopic enrichments observed,
and when coupled with other lines of evidence, pointed to a Paleoproterozoic “great oxidation
event” (Holland 1984, 2002). In the years since, additional work has illustrated that more than
one isotopic excursion can be identified between 2.4 and 2.0 Ga, raising questions regarding the
authenticity of the signature, and possible alternative contributions to the record (Buick et al.
1998; Melezhik et al. 1999; Shields and Viezer 2002; Bekker et al. 2003, 2004, 2006; Kopp et
al. 2005). It has also been suggested that the recycling of this carbon as a result of metamorphic
and igneous activity would have reversed the net effect of the carbon burial (Catling and Claire
2005). What can be said is that the carbon isotope record experienced a turbulent phase in the
hundreds of millions of years following the initial oxygenation of the atmosphere. This ob-
served variability is likely related to the response of the carbon cycle to a more O2-rich surface
environment, but the details of this relationship remain unknown. Strong evidence for active
oxygenic photosynthesis prior to this period (Hayes 1983; Summons et al. 1999; Brocks et al.
2003; Eigenbrode 2004) would certainly point to a biological component as having contributed.
Hayes and Waldbauer (2006) provide a clear review and contemporary thoughts on the Precam-
brian carbon cycle, especially the interval from 2.47 to 2.0 Ga that also addresses these issues.
The presence of banded iron formations (BIF) at ~1.8 Ga presents an observation that
may be inconsistent with surface oxygenation at ~2.4 Ga. These 1.8 Ga BIF may suggest
a decoupling of the surface from the deep ocean and would reflect the upwelling of deep
iron-rich waters onto the continental shelves, where they interacted with a more oxidizing
environment (or bacteria that catalyzed the oxidation), resulting in the formation of the iron
oxides observed in 1.8 Ga BIF (Konhauser et al. 2002; Kappler et al. 2005). Fe-speciation
evidence (Poulton et al. 2004) from the Animike Basin of North America suggests that the
increasing dominance of the sulfur cycle and the subsequent establishment of sulfidic deep-
water conditions brought BIF deposition in the Paleoproterozoic to a close. Implicit in this
argument is that Fe-rich deep waters persisted through the initial rise in atmospheric O2. In
this post-Fe oceanic setting, Canfield and colleagues have argued that the deep ocean may

16_Farquhar_Johnston.indd 20 12/11/2007 11:16:36 PM


Oxygen Cycle of Terrestrial Planets 21

have contained free hydrogen sulfide, indicating that the abundance of available sulfur was
high enough to overwhelm the available iron and begin to accumulate (Canfield 1998, 2005;
Shen et al. 2002, 2003). This interpretation gained further support from a growing body of
work on redox-sensitive elements, such as molybdenum (Arnold et al. 2004) and iron (Rouxel
et al 2005). Some researchers are also investigating the timing of the initial onset of sulfidic
conditions, with preliminary evidence suggesting that, at least at the basin scale, sulfidic deep
waters may extend back to ~2.2 Ga. The presence of reducing conditions in the deep ocean has
direct consequences for calibrating (at this point qualitatively) the state of atmospheric oxygen.
The oxygen content of the deep oceans would reflect the oxygen content of high-latitude cold
water, where deep water formation is thought to occur. It has been recognized that the oxygen
content of these waters would be a function of the concentrations of oxygen in the atmosphere,
temperature, salinity, and the consumption of oxygen along the flow path. The persistence of
anoxic or even sulfidic deep ocean waters would have occurred when oxygen delivery to the
deep ocean was insufficient to titrate the reductants, namely sulfide, present at depth.
Current estimates suggest that sulfidic deep ocean conditions may have characterized a
quarter of Earth’s history, marked by a period of progressive oxygenation that extends from the
end of Paleoproterozoic banded iron formation deposition, throughout the Mesoproterozoic and
well into the Neoproterozoic. The lines of evidence that suggest that progressive oxygenation
occurred include a gradual enrichment in 13C of carbonate (Frank et al.1997); the increased de-
position of evaporite minerals such as gypsum (cf. Kah et al. 2004); and evidence for an active
oxidative sulfur cycle (Johnston et al. 2005) between 1.2 and 1.45 Ga. In addition, increases in
the magnitude of fractionation seen in the sulfur isotope record of the early Neoproterozoic may
suggest a rise in oceanic seawater sulfate concentrations and reflect a second atmospheric oxi-
dation event. The evidence for such increases in atmospheric oxygen during the Neoproterozoic
is preserved within a geologic record of large-scale, potentially global glaciations (Hoffman et
al. 1998). Evidence from molecular biomarkers (Olcott et al. 2005) points to a resilient oxygen-
ated surface ocean. The proliferation of life that is associated with the Precambrian/Cambrian
boundary (the Cambrian explosion) may have been related to the disappearance of anoxic deep
waters. The connection between these two events (biological radiation and oxidation of the
deep ocean) remains the focus of much research and will likely be a rich area of research for
decades to come.
Oxygen concentration variations since the end of the Proterozoic
Studies of the evolution of oxygen since the end of the Precambrian have been undertaken
predominantly using models based on carbon and sulfur isotope covariations. These efforts have
been championed by a series of models produced by Berner (1994) and Berner and Kothavale
(2001). Carbon and sulfur are target systems for tracking atmospheric oxygen because the
mass-balance relationship between the reduced and oxidized forms of these elements would
control the levels of atmospheric oxygen. Berner and Kothavale (2001) suggest that oxygen
levels may have been 25-50% of PAL between 543 and 400 Ma but rose to 175% of PAL by
about 300 Ma. During the Permo-Carboniferous (~ 300 Ma), a time when abundant organic-
rich carbon (coal) was buried, vertebrates invaded land and giant insects (dragonflies) lived.
Giant insects would have required higher oxygen concentrations than PAL because they do not
possess active circulation systems, and respire by diffusion only. The high oxygen levels of the
Permo-Carboniferous reconstructions have not been universally accepted (Lenton 2001).
For more recent parts of the geologic record (e.g., the last 500,000 yr), variations in oxygen
levels have been studied by measuring the isotopic composition of oxygen gas trapped in ice
cores. Ice core data have been used to reconstruct oxygen concentration variations on orbital
timescales associated with the ~20,000 year precession (wobble) of Earth’s rotational axis and
orbit (Petit et al. 1999). These variations are thought to reflect changes in biomass and carbon
burial that were caused by orbitally-induced changes in solar insulation (Milankovich cycles).

16_Farquhar_Johnston.indd 21 12/11/2007 11:16:37 PM


22 Farquhar & Johnston

The abundance and isotopic composition of molecular oxygen have also been found to vary
on shorter (diurnal to decadal) timescales, and these variations are attributed to variations in
biological oxygen production and to human activity (Bender et al. 1998). The changes brought
on by man are significantly smaller, however, than those that Earth has experienced throughout
its ~4.5 billion year history.
Conceptual model for oxygenation of Earth surface environments
Our interpretation of evolution of atmospheric and oceanic oxygen concentrations and
of the oxic and anoxic parts of the biosphere is illustrated schematically in Figure 10.
10 This
evolution includes several different periods in Earth history. The first period in Earth history
is thought to have been marked by widespread anoxia in both the atmosphere and the oceans.
We suggest that the anoxia in the surface oceans may have ended before oxygenation of the
atmosphere, and that the oxygenation of the oceans may have differed from the atmosphere
in the following ways. Atmospheric oxygenation is thought to have been geologically rapid,
to have evolved from one state to another (possibly with oscillations) and to reflect changes
in the balance between sources and sinks. This occurs because the oxidative capacity of an
anoxic atmosphere is small in comparison to the oceans. We suggest that oxygenation of the
oceans may have been more progressive because of the significant potential reductive and
oxidative capacity of ions dissolved in the oceans. The growth (lateral growth and deepening)
of an oxidized surface ocean has the potential to modulate the size of a deep ocean reductive
pool and to serve as a “variable oxygen capacitor” that could allow for a more gradual change
in oxidation state. We do not rule out the possibility of oceanic switches that would result
from oxygen content exceeding threshold values (e.g., Canfield 1998), but we add to this the
possibility of variable oceanic oxidation potential. Implicit in this is the suggestion that the
transformations in the surface ocean would lead the atmospheric changes. The oceans are the

atmosphere present level


oxygen level
amospheric

oxic
surface bios
ocean pher
e
depth of oceanic

ano
oxygenation

xic b
iosp
here
deep ocean

sediments
4.0 3.0 2.0 1.0 0
Age (billions of years ago)
Figure 10. A schematic diagram outlining the major changes in the Earth’s oxygen cycle. The principal
features of this diagram are the inferences about atmospheric oxygenation, which is inferred to occur as
geologically instantaneous steps between stable or metastable states. The curve for atmospheric evolution
illustrates the hypotheses about the relative oxygen levels. The curve for the oceanic system (grey curve)
illustrates the depth to the chemocline. The ocean above the curve is inferred to be oxygenated, and the ocean
below the curve is inferred to be anoxic. This grey curve illustrates the progressive oxygenation and growth
of the oxic ecosystems in Earth’s oceans, and also the systematic contraction of the anoxic biosphere.

16_Farquhar_Johnston.indd 22 12/11/2007 11:16:37 PM


Oxygen Cycle of Terrestrial Planets 23

site of active oxygen production through photosynthesis and settling of organic matter. The
early Fe2+-rich oceans also possessed the capacity to consume significant amounts of oxygen.
These two features would have made the oceans the ultimate source of oxygen and one of the
factors that forced the change of atmospheric oxidation state. The oxygenation of the surface
ocean would also reflect a lower bound (oxygen chemocline) that would have the possibility
of migrating within the water column to reflect different oxygen production rates. Progressive
oxygenation of the surface ocean would proceed until the reaching of a tipping point for the
ocean as a whole (e.g., Canfield 1998), which would lead to oxygenation of the deep ocean
at the end of the Proterozoic. This process would be marked by the progressive expansion of
the oxic biosphere and contraction of the anoxic biosphere from the entire oceans to isolated
basins and sedimentary ecosystems.

NEW FRONTIERS
The systems perspective describes the planet Earth, or any terrestrial planet, as a large
system of interacting components, reservoirs, or pools. It provides a ready context for describing
changes in the observed variations of isotopic compositions and also the physical and chemical
connections that control isotopic and mass transfer. For Earth, the largest system components
include the atmosphere, hydrosphere, biosphere, and geosphere, and these components have
subsystems with smaller-scale cycles. This perspective also provides a way to compare and
contrast the distribution of oxygen in the surface environments of the other terrestrial planets.
A number of issues remain to be resolved. These include:
• Deconvolving the oxygen isotope fractionation effects associated with photosynthesis
and respiration. A significant number of new findings have been made and new
applications are opening for studies of bioproductivity and the controls on atmospheric
oxygen content.
• New studies that combine measurements of mass-independent isotope effects with
atmospheric models and other types of atmospheric data to place new constraints on
atmospheric processes and to use the mass-independent signals as independent and
new constraints on atmospheric processes. A number of studies discussed above have
already shown us how this can be done to obtain these new types of information and
emerging research that marries oxygen with models—for instance a recent study of
water isotopes that incorporates isotopes into the NCAR general circulation model
by Lee et al. (2007)—will inevitably yield new insights and new tools for evaluating
atmospheric chemistry and physics.
• Studies of mass-dependent fractionation effects that have the potential to identify new
mass-dependent chemical and physical pathways. Work on different mass-dependent
pathways and on catabolic effects is providing a foundation that future studies will
use to place new constraints on the biological, physical, chemical, and mass-transfer
effects that operate in natural systems.
• Studies of isotope clumping and other site-specific oxygen isotope effects are providing
new tools for understanding the conditions of last equilibration for molecules such
as carbon dioxide and carbonate minerals. We anticipate a considerable amount of
activity in this area in particular as researchers explore other applications and develop
new uses for this approach.
• Although studies of Earth-surface oxygenation have made significant inroads in the
past few years, there are still fundamental questions that remain to be addressed, such
as:
• Whether the rise of oxygen was a singular event or a series of events culminating

16_Farquhar_Johnston.indd 23 12/11/2007 11:16:38 PM


24 Farquhar & Johnston

with an oxygen-rich atmosphere;


• When the rise of oxygen in fact occurred, and what the nature of the record is for
multiple proxies. The identification of a continuous stratigraphic succession that
pinpoints this event will go a long way to providing a target for many workers
using diverse tools to study the same sample set;
• How many stable steady-states exist for the oxygen content of the atmosphere-
ocean biosphere system, and what feedbacks operate to determine the levels of
oxygen for these states;
• What the reactions are that produce the mass-independent sulfur isotope signals
that have been used as a direct proxy of atmospheric oxygen content, and
whether these reactions and the inferences about oxygen content are as robust as
we presently believe they are; and
• How and when biological innovations occurred in Earth history, for instance
in the middle Proterozoic, that may have influenced the state of oxygen in the
atmosphere-ocean system.
• It is not known how the evolution of oxygen occurred in the Martian atmosphere.
The observation of nonzero Δ17O in SNC meteorite carbonates points to active
atmospheric photochemical sources of anomalous oxygen and also to limits on the
amounts of water in the soils from which the carbonates formed, extending to the
time of carbonate formation in ALH 84001.
• There are also some basic, longstanding questions, such as whether samples returned
from Mars will exhibit isotopic systematics that are similar to those observed in SNC
meteorites. It is not known whether the inferences about the role of water, the role
of atmosphere-surface exchange, and the implications of the SNC meteorites for the
oxygen isotope budget of the silicate planet will stand the test of time.

CONCLUDING STATEMENTS
The beginnings of isotope geochemistry were marked by the development of gas-source
isotope ratio mass spectrometry, early studies of oxygen, and the realization that isotopic
fractionations observed in nature could be used to deconvolve the intricacies of physical,
chemical and biological process that occur in planetary environments (Urey and Grief 1935;
Dole 1935; Dole et al. 1947; Urey 1947). The investigations that followed ranged from studies
of the primitive meteoritic materials that tell us about nucleosynthetic processes, to signatures
that have been used to answer questions about animal metabolisms, paleoclimatology, and
Earth systems science. Some especially noteworthy applications of oxygen isotopic studies
include: the observation and description of the latitudinal control of the isotopic composition
of rainwater (Dole effect); the buffering of the isotopic composition of seawater (Muelenbachs
and Clayton 1976); its use in tracking ice volume and in stratigraphic correlation during the
past million years; application to thermometry (Urey 1947; McCrea 1950; Clayton and Epstein
1961; Taylor and Epstein 1962); providing evidence for a common source for the Earth and its
moon (Clayton et al. 1975; Wiechert et al. 2001); and the tracking of atmospheric processes
using non mass-dependent signatures (Thiemens 1999, 2005; Brenninkmeijer et al. 2003).
New findings and approaches are constantly in the making. It will be particularly interesting
to watch the future of isotope clumping studies unfold and to see the types of new information
that oxygen isotope geochemistry has yet to reveal.
The role of oxygen in the Earth surface system has also been a rich area of study. Min-
eralogical, and now isotopic evidence suggests that anoxic atmospheres defined the first half
of Earth history. After the advent of oxygenic photosynthesis, the story began to change as

16_Farquhar_Johnston.indd 24 12/11/2007 11:16:38 PM


Oxygen Cycle of Terrestrial Planets 25

molecular oxygen began to accumulate. The increase in atmospheric oxygen at ~2.4 Ga marks
one of the “big” events in Earth history, yet it also marks only the beginning of a battle between
oxidants and reductants in Earth’s oceans that remains to be understood. Geochemical evidence
suggests that another 1.5 Ga. may have passed before the deepest ocean became oxidizing. The
oxygenation of the deep ocean is thought to have opened the environmental niches necessary
for the rise of large animals. In the subsequent Phanerozoic world, levels of atmospheric oxy-
gen fluctuated by tens of per cent and are thought to be controlled by the fine balance between
photosynthesis, respiration and silicate weathering.
In all, we look forward to the future of oxygen geochemistry and the work on better
understanding of multiple isotope fractionations.

ACKNOWLEDGEMENTS
We would like to acknowledge support from the NASA COS program, the NASA
Astrobiology Institute, the NSF, and the Hanse Wissenschaftskolleg. We thank Andy Masterson
and Lisa Tuit for reading and making comments on early drafts of this manuscript. We thank
Mark Thiemens for a constructive review.

REFERENCES
Affek HP, Eiler JM (2006) Abundance of mass 47 CO2 in urban air, car exhaust, and human breath. Geochim
Cosmochim Acta 70:1-12
Alexander B, Thiemens MH, Farquhar J, Kaufman AJ, Savarino J, Delmas RJ (2003) East Antarctic ice core
sulfur isotope measurements over a complete glacial-interglacial cycle. J Geophys Res-Atmospheres
108(D24) (PAGE RANGE?)
Alexander B, Vollmer MK, Jackson T, Weiss RF, Thiemens MH (2001) Stratospheric CO2 isotopic anomalies
and SF6 and CFC tracer concentrations in the Arctic polar vortex. Geophys Res Lett 28:4103-4106
Alley RB, Cuffey KM (2001) Oxygen- and hydrogen-isotopic ratios of water in precipitation: Beyond
paleothermometry. Stable Isotope Geochem 43:527-553
Anbar AD, Duan Y, Lyons TW, Arnold GL, Kendall B, Creaser RA, Kaufman AJ, Gordon GW, Scott C, Garvin
J, Buick R (2007) A whiff of oxygen before the Great Oxidation Event? Science 317:1903-1906
Angert A, Rachmilevitch S, Barkan E, Luz B (2003) Effects of photorespiration, the cytochrome pathway, the
alternative pathway on the triple isotopic composition of atmospheric O2. Global Biogeochem Cycles
17(1), doi: 10.1029/2002GB001933
Arnold GL, Anbar AD, Barling J, Lyons TW (2004) Molybdenum isotope evidence for widespread anoxia in
mid-proterozoic oceans. Science 304:87-90
Ashwal LD, Warner JL, Wood CA (1982) SNC Meteorites - Evidence against an asteroidal origin. J Geophys
Res 87:A393-A400
Babikov D, Kendrick BK, Walker RB, Pack RT, Fleurat-Lesard P, Schinke R (2003) Metastable states of ozone
calculated on an accurate potential energy surface. J Chem Phys 118:6298-6308
Bao HM, Thiemens MH, Farquhar J, Campbell DA, Lee CCW, Heine K, Loope DB (2000) Anomalous 17O
compositions in massive sulphate deposits on the Earth. Nature 406:176-178
Bao HM, Thiemens MH, Loope DB, Yuan XL (2003) Sulfate oxygen-17 anomaly in an Oligocene ash bed in
mid-North America: Was it the dry fogs? Geophys Res Lett 30(16) (PAGE RANGE?)
Bekker A, Holland HD, Wang PL, Rumble D, Stein HJ, Hannah JL, Coetzee LL, Beukes NJ (2004) Dating the
rise of atmospheric oxygen. Nature 427:117-120
Bekker A, Karhu JA, Kaufman AJ (2006) Carbon isotope record for the onset of the Lomagundi carbon isotope
excursion in the Great Lakes area, North America. Precamb Res 148:145-180
Bekker A, Sial AN, Karhu JA, Ferreira VP, Noce CM, Kaufman AJ, Romano AW, Pimentel MM (2003)
Chemostratigraphy of carbonates from the minas supergroup, quadrilatero ferrifero (iron quadrangle),
Brazil: A stratigraphic record of early proterozoic atmospheric, biogeochemical and climatic change. Am
J Sci 303:865-904
Bender M, Sowers T, Labeyrie L (1994b) The Dole effect and its variations during the last 130,000 years as
measured in the Vostok ice core. Global Biogeochem Cycles 8:363-376
Bender M, Sowers, T, Dickson ML, Orchardo J, Grootes P, Mayewski PA, Meese DA (1994a) Climate
correlations between Greenland and Antarctica during the past 100,000 years. Nature 372:663-666

16_Farquhar_Johnston.indd 25 12/11/2007 11:16:38 PM


26 Farquhar & Johnston

Bender ML, Battle M, Keeling RF (1998) The O2 balance of the atmosphere: A tool for studying the fate of
fossil-fuel CO2. Ann Rev Energy Environ 23:207-223
Berner RA (1994) 3GEOCARB-II - A revised model of atmospheric CO2 over Phanerozoic time. Am J Sci
294:56-91
Berner RA, Kothavala Z (2001) GEOCARB III: A revised model of atmospheric CO2 over phanerozoic time.
Am J Sci 301:182-204
Bigeleisen J, Friedman L (1950) The Infra-Red Spectra of N15N14O16 and N14N15O16 - Some thermodynamic
properties of the isotopic N2O molecules. J Chem Phys 18:1656-1659
Bigeleisen J, Mayer MG (1947) Calculation of equilibrium constants for isotopic exchange reactions. J Chem
Phys 15:261-267
Blunier T, Barnett B, Bender ML, Hendricks MB (2002) Biological oxygen productivity during the last
60,000 years from triple oxygen isotope measurements. Global Biogeochem Cycles 16(3), doi:
10.1029/2001GB001460
Boering KA, Jackson T, Hoag KJ, Cole AS, Perri MJ, Thiemens M, 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. Geophys Res Lett 31(3) (PAGE RANGE?)
Bogard DD, Johnson P (1983) Martian gases in an Antarctic meteorite. Science 221:651-654
Brenninkmeijer CAM, Janssen C, Kaiser J, Rockmann T, Rhee TS, Assonov SS (2003) Isotope effects in the
chemistry of atmospheric trace compounds. Chem Rev 103:5125-5161
Brocks JJ, Buick R, Summons RE, Logan GA (2003) A reconstruction of Archean biological diversity based
on molecular fossils from the 2.78 to 2.45 billion-year-old Mount Bruce Supergroup, Hamersley Basin,
Western Australia. Geochim Cosmochim Acta 67:4321-4335
Buick IS, Uken R, Gibson RL, Wallmach T (1998) High-delta C-13 Paleoproterozoic carbonates from the
Transvaal Supergroup, South Africa. Geology 26:875-878
Came RE, Eiler JM, Veizer J, Azmy K, Brand U, Weidman CR (2007) Coupling of surface temperatures and
atmospheric CO2 concentrations during the Paleozoic era. Nature 449:198-201
Canfield DE (1998) A new model for Proterozoic ocean chemistry. Nature 396:450-453
Canfield DE (2001) Biogeochemistry of sulfur isotopes. Stable Isotope Geochem 43:607-636
Canfield DE (2005) The early history of atmospheric oxygen: Homage to Robert A Garrels. Ann Rev Earth
Planet Sci 33:1-36
Canil D (2002) Vanadium in peridotites, mantle redox and tectonic environments: Archean to present. Earth
Planet Sci Lett 195:75-90
Catling DC, Claire MW (2005) How Earth’s atmosphere evolved to an oxic state: A status report. Earth Planet
Sci Lett 237:1-20
Claire MW, Catling DC, Zahnle KJ (2006) Biogeochemical modeling of the rise in atmospheric oxygen.
Geobiology 4:239-269
Clayton RN, Epstein S (1961) The use of oxygen isotopes in high-temperature geological thermometry. J Geol
69:447-452
Clayton RN, Grossman L, Mayeda TK (1973a) Component of primitive nuclear composition in carbonaceous
meteorites. Science 182:485-488
Clayton RN, Mayeda TK (1975) Genetic relations between the moon and meteorites. Proc Lunar Sci Conf
6:1761-1769
Clayton RN, Mayeda TK (1984) The oxygen isotope record in Murchison and other carbonaceous chondrites.
Earth Planet Sci Lett 67:151-161
Clayton RN, Mayeda TK (1996) Oxygen isotope studies of achondrites. Geochim Cosmochim Acta 60:1999-
2017
Clayton RN, Mayeda TK (1999) Oxygen isotope studies of carbonaceous chondrites. Geochim Cosmochim
Acta 63:2089-2104
Clayton RN, Mayeda TK, Goswami JN, Olsen EJ (1991) Oxygen isotope studies of ordinary chondrites.
Geochim Cosmochim Acta 55:2317-2337
Clayton RN, Mayeda TK, Hurd JM (1973b) Oxygen isotopic fractionation within ultrabasic clasts of lunar
breccia 15445. J Geol 81:227-228
Clayton RN, Mayeda, TK (1983) Oxygen isotopes in eucrites, shergottites, nakhlites, and chassignites. Earth
Planet Sci Lett 62:1-6
Clayton RN, Onuma N, Mayeda TK (1976) Classification of meteorites based on oxygen isotopes. Earth Planet
Sci Lett 30:10-18
Cloud PE (1968) Atmospheric and hydrospheric evolution on primitive Earth. Science 160:729-736
Cloud PE (1972) Working model of primitive Earth. Am J Sci 272:537-548
Cloud PE (1973) Paleoecological significance of banded iron-formation. Econ Geol 68:1135-1143
Dole M (1935) The relative atomic weight of oxygen in water and in air. J Amer Chem Soc 57:2731-2731

16_Farquhar_Johnston.indd 26 12/11/2007 11:16:38 PM


Oxygen Cycle of Terrestrial Planets 27

Dole M, Hawkings RC, Barker HA (1947) Bacterial fractionation of oxygen isotopes. J Amer Chem Soc 69:226-
228
Eigenbrode JL, Freeman KH, Love GD, Snape CE, Summons RE (2004) Hydropyrolytic release of hopane and
sterane biomarkers from 2.7 Ga kerogens. Geochim Cosmochim Acta 68:A240
Emiliani C (1955) Pleistocene temperatures. J Geol 63:538-578
Epstein S, Mayeda T (1953) Variation of 18O content of waters from natural sources. Geochim Cosmochim Acta
4:213-224
Epstein S, Taylor HP (1970) 18O/16O, SI-30/SI-28, D-H, C-13/C-12 studies of lunar rocks and minerals. Science
167:533-535
Farquhar J, Bao HM, Thiemens M (2000a) Atmospheric influence of Earth’s earliest sulfur cycle. Science
289:756-758
Farquhar J, Johnston DT, Wing BA (2007a) Implications of conservation of mass effects on mass-dependent
isotope fractionations: Influence of network structure on sulfur isotope phase space of dissimilatory sulfate
reduction. Geochim Cosmochim Acta doi:10.1016/j.gca.2007.08.028
Farquhar J, Peters M, Johnston DT, Strauss H, Masterson A, Wiechert U, Kaufman AJ (2007b) Isotopic evidence
for Mesoarchaean anoxia and changing atmospheric sulphur chemistry. Nature 449:706-709
Farquhar J, Savarino J, Jackson TL, Thiemens, MH (2000b) Evidence of atmospheric sulphur in the martian
regolith from sulphur isotopes in meteorites. Nature 404:50-52
Farquhar J, Thiemens MH (2000) Oxygen cycle of the Martian atmosphere-regolith system: Delta 17O of
secondary phases in Nakhla and Lafayette. J Geophys Res-Planets 105:11991-11997
Farquhar J, Thiemens MH, Jackson TL (1998) Atmosphere-surface interactions on Mars: Delta 17O measurements
of carbonate from ALH 84001. Science 280:1580-1582
Farquhar J, Wing BA (2003) Multiple sulfur isotopes and the evolution of the atmosphere. Earth Planet Sci Lett
213:1-13
Farquhar J, Wing BA (2005) The terrestrial record of stable sulphur isotopes: a review of the implications for
evolution of Earth’s sulphur cycle. Geol Soc London Spec Pub 248:167–77
Farquhar J, Wing BA, McKeegan KD, Harris JW, Cartigny P, Thiemens MH (2002) Mass-independent sulfur of
inclusions in diamond and sulfur recycling on early earth. Science 298:2369-2372
Franchi IA, Wright IP, Sexton AS, Pillinger CT (1999) The oxygen-isotopic composition of Earth and Mars.
Meteorit Planet Sci 34:657-661
Frank TD, Kah LC, Lyons TW (2003) Changes in organic matter production and accumulation as a mechanism
for isotopic evolution in the Mesoproterozoic ocean. Geol Mag 140:397-420
Gao YQ, Marcus RA (2001) Strange and unconventional isotope effects in ozone formation. Science 293:259-
263
Ghosh P, Garzione CN, Eiler JM (2006) Rapid uplift of the Altiplano revealed through C-13-18O bonds in
paleosol carbonates. Science 311:511-515
Gregory RT, Taylor HP (1981) An oxygen isotope profile in a section of cretaceous oceanic crust, Samail
ophiolite, Oman: Evidence for 18O buffering of the oceans by deep (>5 km) seawater-hydrothermal
circulation at mid-ocean ridges. J Geophys Res 86:2737-2755.
Guy RD, Fogel ML, Berry JA (1993) Photosynthetic fractionation of the stable isotopes of oxygen and carbon.
Plant Phys 101:37-47
Habicht KS, Gade M, Thamdrup B, Berg P, Canfield DE (2002) Calibration of sulfate levels in the Archean
Ocean. Science 298:2372-2374
Hayes JM (1983) Geochemical evidence bearing on the origin of aerobiosis: A speculative hypothesis. In:
Earth’s Earliest Biosphere: Its Origin and Evolution. Schopf JW (ed) Princeton University Press, Princeton
New Jersey, p. 291-301
Hayes JM, Waldbauer JR (2006) The carbon cycle and associated redox processes through time. Phil Trans R
Soc B-Biol Sci 361:931-950
Hays JD, Imbrie J, Shackleton NJ (1976) Variations in Earth’s orbit - Pacemaker of ice ages. Science 194:1121-
1132
Hendricks MB, Bender ML, Barnett BA (2004) Net and gross O2 production in the Southern Ocean from
measurements of biological O2 saturation and its triple isotope composition. Deep-Sea Research Part I.
Oceanograph Res Papers 51:1541-1561
Hendricks MB, Bender ML, Barnett BA, Strutton P, Chavez FP (2005) Triple oxygen isotope composition of
dissolved O2 in the equatorial Pacific: A tracer of mixing, production, respiration. J Geophys Res Oceans
110(C12) (PAGE RANGE?)
Hoffman PF, Kaufman AJ, Halverson GP, Schrag DP (1998) A Neoproterozoic snowball earth. Science
281:1342-1346
Holland HD (1984) The Chemical Evolution of the Atmosphere and Oceans. xii. Princeton University Press,
Princeton New Jersey

16_Farquhar_Johnston.indd 27 12/11/2007 11:16:39 PM


28 Farquhar & Johnston

Holland HD (2002) Volcanic gases, black smokers, the Great Oxidation Event. Geochim Cosmochim Acta
66:3811-3826
Holland HD (2006) The oxygenation of the atmosphere and oceans. Phil Trans R Soc B-Biol Sci 361:903-915
Hu GX, Rumble D, Wang PL (2003) An ultraviolet laser microprobe for the in situ analysis of multisulfur isotopes
and its use in measuring Archean sulfur isotope mass-independent anomalies. Geochim Cosmochim Acta
67:3101-3118
Johnston DT, Poulton SW, Fralick PW, Wing BA, Canfield DE, Farquhar J (2006) Evolution of the oceanic
sulfur cycle at the end of the Paleoproterozoic. Geochim Cosmochim Acta 70:5723-5739
Johnston DT, Wing BA, Farquhar J, Kaufman AJ, Strauss H, Lyons TW, Kah LC, Canfield DE (2005) Active
microbial sulfur disproportionation in the Mesoproterozoic. Science 310:1477-1479
Kah LC, Lyons TW, Frank TD (2004) Low marine sulphate and protracted oxygenation of the proterozoic
biosphere. Nature 431:834-838
Kappler A, Pasquero C, Konhauser KO, Newman DK (2005) Deposition of banded iron formations by
anoxygenic phototrophic Fe(II)-oxidizing bacteria. Geology 33:865-868
Karhu JA, Holland HD (1996) Carbon isotopes and the rise of atmospheric oxygen. Geology 24:867-870
Karlsson KR, Clayton RN, Gibson EK, Mayeda TK (1992) Water in SNC meteorites - Evidence for a Martian
hydrosphere. Science 255:1409-1411
Kasting JF (1993) Earth’s early atmosphere. Science 259:920-926
Kaufman AJ, Johnston DT, Farquhar J, Masterson AL, Lyons TW, Bates S, Anbar AD, Arnold GL, Garvin J,
Buick R (2007) Late Archean biospheric oxygenation and atmospheric evolution. Science 317:1900-1903
Keeling RF, Piper SC, Heimann M (1996) Global and hemispheric CO2 sinks deduced from changes in
atmospheric O2 concentration. Nature 381:218-221
Keeling RF, Shertz SR (1992) Seasonal and interannual variations in atmospheric oxygen and implications for
the global carbon-cycle. Nature 358:723-727
Knauth LP, Epstein S (1976) Hydrogen and oxygen isotope ratios in nodular and bedded cherts. Geochim
Cosmochim Acta 40:1095-1108
Konhauser KO, Hamade T, Raiswell R, Morris RC, Ferris FG, Southam G, Canfield, DE (2002) Could bacteria
have formed the Precambrian banded iron formations? Geology 30:1079-1082
Kopp RE, Kirschvink JL, Hilburn IA, Nash CZ (2005) The paleoproterozoic snowball Earth: A climate disaster
triggered by the evolution of oxygenic photosynthesis. Proc Nat Acad Sci USA 102:11131-11136
Krankowsky D, Bartecki F, Klees GG, Mauersberger K, Schellenbach K, Stehr J (1995) Measurement of heavy
isotope enrichment in tropospheric ozone. Geophys Res Lett 22:1713-1716
Kump LR, Kasting JF, Barley ME (2001) Rise of atmospheric oxygen and the “upside-down” Archean mantle.
Geochem Geophys Geosys 2 (PAGE RANGE?)
Lee CCW, Thiemens MH (2001) The delta 17O and delta 18O measurements of atmospheric sulfate from a coastal
and high alpine region: A mass-independent isotopic anomaly. J Geophys Res-Atmospheres106:17359-
17373
Lee JE, Fung I, DePaolo DJ, Henning CC (2007) Analysis of the global distribution of water isotopes using the
NCAR atmospheric general circulation model. J Geophys Res-Atmospheres 112 (D16): Art. No. D16306,
doi:10.1029/2006JD007657
Lenton TM (2001) The role of land plants, phosphorus weathering and fire in the rise and regulation of
atmospheric oxygen. Global Change Biology 7:613-629
Li ZXA, Lee, CTA (2004) The constancy of upper mantle fO2 through time inferred from V/Sc ratios in basalts.
Earth Planet Sci Lett 228:483-493
Lodders K (2000) An oxygen isotope mixing model for the accretion and composition of rocky planets. Space
Sci Rev 92:341-354
Lodders K, Fegley B (1997) An oxygen isotope model for the composition of Mars. Icarus 126:373-394
Lodders K, Fegley B (1998) Presolar silicon carbide grains and their parent stars. Meteor Planet Sci 33:871-
880
Luz B, Barkan E (2005) The isotopic ratios 17O/16O and 18O/16O in molecular oxygen and their significance in
biogeochemistry. Geochim Cosmochim Acta 69:1099-1110
Luz B, Barkan E, Bender ML, Thiemens MH, Boering KA (1999) Triple-isotope composition of atmospheric
oxygen as a tracer of biosphere productivity. Nature 400:547-550
Lyons JR (2001) Transfer of mass-independent fractionation in ozone to other oxygen-containing radicals in the
atmosphere. Geophys Res Lett 28:3231-3234
Madigan MT, Martinko JM, Parker J (2000) Brock Biology of Microorganisms. XIX. Prentice Hall, Upper
Saddle River New Jersey
Matsuhisa Y, Goldsmith JR, Clayton RN (1978) Mechanisms of hydrothermal crystallization of quartz at 250-
Degrees-C and 15 KBar. Geochim Cosmochim Acta 42:173-182
Matsuhisa Y, Goldsmith JR, Clayton RN (1979) Oxygen isotopic fractionation in the system Quartz-Albite-
Anorthite-Water. Geochim Cosmochim Acta 43:1131-1140

16_Farquhar_Johnston.indd 28 12/11/2007 11:16:39 PM


Oxygen Cycle of Terrestrial Planets 29

Mauersberger K (1981) Measurement of heavy ozone in the stratosphere. Geophys Res Lett 8:935-937
Mauersberger K (1987) Ozone isotope measurements in the stratosphere. Geophys Res Lett 14:80-83
Mauersberger K, Erbacher B, Krankowsky D, Gunther J, Nickel R (1999) Ozone isotope enrichment: Isotopomer-
specific rate coefficients. Science 283:370-372
McCabe JR, Thiemens MH, Savarino J (2007) A record of ozone variability in South Pole Antarctic snow: Role
of nitrate oxygen isotopes. J Geophys Res – Atmospheres 112:D12303 JUN 16 2007
McCrea JM (1950) On the isotopic chemistry of carbonates and a paleotemperature scale. J Chem Phys 18:849-
857
McCrea, JM (1951) Intramolecular isotopic fractionation - the example of nitrous oxide. J Chem Phys 19:48-
49
McKinney CR, McCrea JM, Epstein S, Allen HA, Urey HC (1950) Improvements in mass spectrometers for the
measurement of small differences in isotope abundance ratios. Rev Sci Instrum 21:724-730
McSween HY (1984) SNC Meteorites; are they Martian rocks? Geology 12:3-6
McSween HY (1994) What we have learned about Mars from SNC meteorites. Meteoritics 29:757-779
Michalski G, Scott Z, Kabiling M, Thiemens MH (2003) First measurements and modeling of Delta 17O in
atmospheric nitrate. Geophys Res Lett 30 (16) (PAGE RANGE?)
Mojzsis SJ, Coath CD, Greenwood JP, McKeegan KD, Harrison TM (2003) Mass-independent isotope effects in
Archean (2.5 to 3.8 Ga) sedimentary sulfides determined by ion microprobe analysis. Geochim Cosmochim
Acta 67:1635-1658
Mook WG (2001) Environmental Isotopes in the Hydrological Cycle: Principles and Applications. Paris:
UNESCO Publishing
Morton J, Schueler B, Mauersberger K (1989) Oxygen fractionation of ozone isotopes 48O3 through 54O3. Chem
Phys Lett 154:143-145
Muehlenbachs K, Clayton RN (1976) Oxygen isotope composition of oceanic-crust and its bearing on seawater.
J Geophys Res 81:4356-4369
Nier AO (1947) A mass spectrometer for isotope and gas analysis. Rev Sci Instrum 18:398-411
Nier AO, Ney EP, Inghram MG (1947) A null method for the comparison of 2 ion currents in a mass spectrometer.
Rev Sci Instrum 18:294-297
Ohmoto H, Watanabe Y, Ikemi H, Poulson SR, Taylor BE (2006) Sulphur isotope evidence for an oxic Archaean
atmosphere. Nature 442:908-911
Olcott AN, Sessions AL, Corsetti FA, Kaufman AJ, de Oliviera TF (2005) Biomarker evidence for photosynthesis
during Neoproterozoic glaciation. Science 310:471-474
Ono S, Beukes NJ, Rumble D, Fogel ML (2006b) Early evolution of atmospheric oxygen from multiple-sulfur
and carbon isotope records of the 2.9 Ga Mozaan Group of the Pongola Supergroup, Southern Africa.
South African J Geol 109: 97-108
Ono S, Eigenbrode JL, Pavlov AA, Kharecha P, Rumble D, Kasting JF, Freeman KH (2003) New insights into
Archean sulfur cycle from mass-independent sulfur isotope records from the Hamersley Basin, Australia.
Earth Planet Sci Lett 213:15-30
Ono S, Wing B, Johnston D, Farquhar J, Rumble D (2006a) Mass-dependent fractionation of quadruple stable
sulfur isotope system as a new tracer of sulfur biogeochemical cycles. Geochim Cosmochim Acta 70:2238-
2252
Onuma N, Clayton RN, Mayeda TK (1970) Oxygen isotope fractionation between minerals and an estimate of
temperature of formation. Science 167:536-538
Papineau D, Mojzsis SJ, Coath CD, Karhu JA, McKeegan KD (2005) Multiple sulfur isotopes of sulfides from
sediments in the aftermath of Paleoproterozoic glaciations. Geochim Cosmochim Acta 69:5033-5060
Pavlov AA, Kasting JF (2002) Mass-independent fractionation of sulfur isotopes in Archean sediments: Strong
evidence for an anoxic Archean atmosphere. Astrobiology 2:27-41
Petit JR, Jouzel J, Raynaud D, Barkov NI, Barnola JM, Basile I, Bender M, Chappellaz J, Davis M, Delaygue G,
Delmotte M, Kotlyakov VM, Legrand M, Lipenkov VY, Lorius C, Pepin L, Ritz C, Saltzman E, Stievenard
M (1999) Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica.
Nature 399:429-436
Petsch ST, Edwards KJ, Eglinton TI (2005) Microbial transformations of organic matter in black shales and
implications for global biogeochemical cycles. Palaeogeo Palaeoclim Palaeoecol 219:157-170
Poulton SW, Fralick PW, Canfield DE (2004) The transition to a sulphidic ocean similar to 1.84 billion years
ago. Nature 431:173-177
Reuer MK, Barnett BA, Bender ML, Falkowski, PG, Hendricks MB (2007) New estimates of Southern Ocean
biological production rates from O2/Ar ratios and the triple isotope composition of O2. Deep Sea Research
Part I: Oceanograph Res Pap 54:951-974
Romanek CS, Perry EC, Treiman AH, Socki RA, Jones JH, Gibson EK (1998) Oxygen isotopic record of
silicate alteration in the Shergotty-Nakhla-Chassigny meteorite Lafayette. Meteor Planet Sci 33:775-
784

16_Farquhar_Johnston.indd 29 12/11/2007 11:16:39 PM


30 Farquhar & Johnston

Roth JP, Wincek R, Nodet G, Edmondson DE, McIntire WS, Klinman JP (2004) Oxygen isotope effects on
electron transfer to O2 probed using chemically modified flavins bound to glucose oxidase. J Am Chem
Soc 126:15120-15131
Rouxel OJ, Bekker A, Edwards KJ (2005) Iron isotope constraints on the Archean and Paleoproterozoic ocean
redox state. Science 307:1088-1091
Rye R, Holland HD (1998) Paleosols and the evolution of atmospheric oxygen: A critical review. Am J Sci
298:621-672
Sanloup C, Jambon A, Gillet P (1999) A simple chondritic model of Mars. Phys Earth Planet Interiors 112:43-
54
Savarino J, Bekki S, Cole-Dai JH, Thiemens MH (2003) Evidence from sulfate mass independent oxygen
isotopic compositions of dramatic changes in atmospheric oxidation following massive volcanic eruptions.
J Geophys Res -Atmospheres 108(D21) PAGE RANGE?
Savarino J, Kaiser J, Morin S, Sigman DM, Thiemens MH (2007) Nitrogen and oxygen isotopic constraints on
the origin of atmospheric nitrate in coastal Antarctica. Atmos Chem Phys 7:1925-1945
Savarino J, Lee, CCW, Thiemens MH (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. J Geophys Res-Atmospheres 105:29079-29088
Savarino J, Thiemens MH (1999) Analytical procedure to determine both δ18O and δ17O of H2O2 in natural
water and first measurements. Atmos Environ 33:3683-3690
Schauble EA, Ghosh P, Eiler JM (2006) Preferential formation of C-13-18O bonds in carbonate minerals,
estimated using first-principles lattice dynamics. Geochim Cosmochim Acta 70:2510-2529
Shen Y, Knoll AH, Walter MR (2003) Evidence for low sulphate and anoxia in a mid-Proterozoic marine basin.
Nature 423:632-635
Shen YN, Canfield DE, Knoll AH (2002) Middle proterozoic ocean chemistry: Evidence from the McArthur
Basin, northern Australia. Am J Sci 302:81-109
Shields G, Veizer J (2002) Precambrian marine carbonate isotope database: Version 1.1. Geochem Geophys
Geosys 3:1031
Strauss H (1993) The sulfur isotopic record of precambrian sulfates - New data and a critical-evaluation of the
existing record. Precambrian Res 63:225-246
Summons RE, Bradley AS, Jahnke LL, Waldbauer JR (2006) Steroids, triterpenoids and molecular oxygen.
Phil Trans R Soc B-Biol Sci 361:951-968
Summons RE, Jahnke LL, Hope JM, Logan GA (1999) 2-Methylhopanoids as biomarkers for cyanobacterial
oxygenic photosynthesis. Nature 400:554-557
Taylor HP (1974) Application of oxygen and hydrogen isotope studies to problems of hydrothermal alteration
and ore deposition. Econ Geol 69:843-883
Taylor HP, Epstein S (1962) Relationship between 18O/16O ratios in coexisting minerals of igneous and
metamorphic rocks; Part 2, Application to petrologic problems. Geol Soc Am Bull 73:675-693
Thiemens MH (1999) Atmosphere science - Mass-independent isotope effects in planetary atmospheres and
the early solar system. Science 283:341-345
Thiemens MH (2006) History and applications of mass-independent isotope effects. Ann Rev Earth Planet Sci
34:217-262
Thiemens MH, Heidenreich JE (1983) The mass-independent fractionation of oxygen - A novel isotope effect
and its possible cosmochemical implications. Science 219:1073-1075
Thiemens MH, Jackson TL, Brenninkmeijer CAM (1995) Observation of a mass-independent oxygen isotopic
composition in terrestrial stratospheric CO2, the link to ozone chemistry, the possible occurrence in the
Martian atmosphere. Geophys Res Lett 22:255-257
Thiemens MH, Savarino J, Farquhar J, Bao, HM (2001) Mass-independent isotopic compositions in terrestrial
and extraterrestrial solids and their applications. Acc Chem Res 34:645-652
Tian F, Toon OB, Pavlov AA, De Sterck H (2005) A hydrogen-rich early Earth atmosphere. Science 308:1014-
1017
Urey HC (1947) The thermodynamic Properties of Isotopic Substances. J Am Chem Soc MAY:562-581
Urey HC, Greiff LJ (1935) Isotopic exchange equilibria. J Am Chem Soc 57:321-327
Whitehouse MJ, Kamber BS, Fedo CM, Lepland A (2005) Integrated Pb- and S-isotope investigation of
sulphide minerals from the early Archaean of southwest Greenland. Chem Geol 222:112-131
Wiechert U, Halliday AN, Lee DC, Snyder GA, Taylor LA, Rumble D (2001) Oxygen isotopes and the moon-
forming giant impact. Science 294:345-348
Yoshida N, Toyoda S (2000) Constraining the atmospheric N2O budget from intramolecular site preference in
N2O isotopomers. Nature 405:330-334
Yung YL, Demore WB (1999) Photochemistry of Planetary Atmospheres. Oxford University Press, USA

16_Farquhar_Johnston.indd 30 12/11/2007 11:16:39 PM

View publication stats

You might also like