Sie sind auf Seite 1von 44

spe500-08 1st pgs

page 1

The Geological Society of America Special Paper 500 2013

18 8 8

20 013

CELEBRATING ADVANCES IN GEOSCIENCE

Smaller, better, more: Five decades of advances in geochemistry


Clark M. Johnson* Department of Geoscience, University of Wisconsin, Madison, Wisconsin 53706, USA, and National Aeronautics and Space Administration (NASA) Astrobiology Institute, Wisconsin Team Scott M. McLennan Department of Geoscience, State University of New York at Stony Brook, Stony Brook, New York 11794, USA Harry Y. McSween Department of Earth and Planetary Sciences, University of Tennessee, Knoxville, Tennessee 37996, USA Roger E. Summons Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA, and National Aeronautics and Space Administration (NASA) Astrobiology Institute, MIT Team

ABSTRACT Many of the discoveries made in geochemistry over the last 50 yr have been driven by technological advances that have allowed analysis of smaller samples, attainment of better instrumental precision and accuracy or computational capability, and automation that has provided many more data. These advances occurred during development of revolutionary concepts, such as plate tectonics, which has provided an overarching framework for interpreting many geochemical studies. Also, spacecraft exploration of other planetary bodies, including analyses of returned lunar samples and remote sensing of Mars, has added an additional dimension to geochemistry. Determinations of elemental compositions of minerals and rocks, either through in situ analysis by various techniques (e.g., electron microprobe, secondary ion mass spectrometry [SIMS], synchrotron X-ray uorescence [XRF], laser ablation) or bulk analysis (e.g., XRF, inductively coupled plasmaatomic emission spectrometry [ICP-AES], inductively coupled plasmamass spectrometry [ICP-MS]), have become essential approaches to many geochemical studies at levels of sensitivity and spatial resolution undreamed of ve decades ago. Although major-element distributions in igneous rocks have been understood at a basic level for some time, advances using major-element abundances to understand sedimentary provenance and processes have been especially noteworthy during the past half-century. The great diversity of trace elements in terms of geochemical behavior (e.g., lithophile, siderophile, etc.) has made them invaluable to many studies, providing unique constraints on redox conditions, mineral-melt and mineral-uid reactions, and planetary differentiation.

*clarkj@geology.wisc.edu Johnson, C.M., McLennan, S.M., McSween, H.Y., and Summons, R.E., 2013, Smaller, better, more: Five decades of advances in geochemistry, in Bickford, M.E., ed., The Web of Geological Sciences: Advances, Impacts, and Interactions: Geological Society of America Special Paper 500, p. 144, doi:10.1130/2013.2500(08). For permission to copy, contact editing@geosociety.org. 2013 The Geological Society of America. All rights reserved.

spe500-08 1st pgs 2 Johnson et al. Signicant advances in microanalytical techniques have markedly improved experimental determinations of trace-element partitioning among phases and in characterizing elemental distributions in rocks and minerals using two-dimensional (2-D) and three-dimensional (3-D) mapping. Rare earth elements, in particular, have proved to be invaluable tracers of magmatic, sedimentary, aqueous, redox, and cosmochemical processes, and siderophile trace elements form a basis for modeling many aspects of planetary accretion and early evolution. An anomalous amount of iridium at Mesozoic-Cenozoic boundary has revolutionized our view of one of Earths most important biologic extinctions. Isotopic variations, whether produced by stable or radiogenic isotopes, provide a third dimension to the Periodic Table of Elements, and tremendous advances in instrumentation since the early 1960s have greatly broadened this eld of geochemistry. Early work outlined the stable H and O isotope ngerprints of natural waters and water-rock interactions, and stable C and S isotope studies dened the biological fractionations that occur by photosynthesis and microbial sulfate reduction, respectively, topics that have since been applied to problems relating to the evolution of life and Earths atmosphere. Recent work on stable O isotopes has documented the likelihood that liquid water existed >4 b.y. ago on Earth, which profoundly affects our view of Earths evolution. New work on nontraditional stable isotopes has investigated redox cycling over Earths history, as has study of non-mass-dependent stable isotope variations. New approaches using stable isotopes as paleothermometers include exploiting the unique energetics of bonds between rare stable isotopes. Early work on the radiogenic Rb-Sr and U-Th-Pb isotope systems documented the key distinctions between continental crust and mantle, setting the stage for later tracing of mass uxes via plate tectonics, as well as documenting the great antiquity of continental crust formation and mantle differentiation on Earth. The Sm-Nd and Lu-Hf isotope systems provided a temporal context for earlier studies of rare earth element variations in nature, including new constraints on crustal growth rates and mechanisms extending back earlier than 4 Ga. The siderophile Re-Os isotope system has been used to study the accretion of planetary bodies, core-mantle interaction, and the nature of the ancient lithospheric mantle. The branch of geochemistry that deals with fossilized organic molecules had its origins in elucidating the processes and pathways that led to petroleum formation. As awareness of the richness and diversity of organic compounds that can be preserved in sedimentary rocks grew, this gave way to the broader endeavor of molecular paleobiology. Despite great challenges in tying specic biomolecules to groups of organisms, or to metabolic processes, as well as issues of preservation mechanisms, molecular paleobiology remains a prime approach for studying the history of microorganisms, which have been the dominant life form for most of Earths history and yet are rarely preserved in the fossil record. Work on molecular biomarkers has produced numerous paleoenvironmental proxies for the chemistry and redox state (euxinia, anoxic, oxic) of the ancient oceans, as well as new paleoclimate records. The biochemical diversity of relatively simple life forms, including bacteria and archaea, has provided a wealth of lipid biomarkers that inform us about the evolution of metabolisms over Earth history, including oxygenic and anoxygenic photosynthesis, methanogenesis, and methanotrophy, and these records have been tied into stable isotope variations of many individual chemical elements (C, H, N, O, S, Fe, Mo, etc.), which provide a broad view of the biogeochemical evolution and biologically catalyzed redox cycling of Earth, and, potentially, other planetary bodies. Although many geochemists focus exclusively on terrestrial problems, research over the past ve decades has been intimately linked to the chemistry of other solar system bodies and the universe beyond. We routinely rely on meteorite falls, interplanetary dust particles, and Moon rocks for a baseline for comparison to Earth, which has been extensively differentiated and repeatedly resurfaced. Sophisticated

page 2

spe500-08 1st pgs Five decades of advances in geochemistry remote-sensing capabilities based on past and current spacecraft missions are enabling active study of other planetary bodies such as the Moon, Mercury, and Mars. Ideas about nucleosynthesis within stars are tested by reference to the measured isotopic compositions of tiny presolar grains extracted from chondrites. Short-lived radionuclides in meteorites provide a detailed record of the condensation, mixing, and differentiation history of the earliest solar system. Mass-independent oxygen isotope fractionation in extraterrestrial samples may identify photochemical processes in the early solar nebula. More broadly, the temperature stabilities of elements and minerals constrain the sequence of nebular condensation, which provides a rst-order explanation for the bulk composition of the terrestrial planets relative to the planets of the outer solar system. Organic compounds from space inform us on the delivery of complex organic molecules to the early Earth, which likely inuenced the earliest organic chemistry reactions, which in turn must have affected the origin and evolution of life. Chemical characterizations of samples of the Moon from the Apollo missions have provided the key data to recognize the Moons formation by impact of a Mars-size object with Earth and the likelihood that both bodies solidied from magma oceans. The individual subelds in geochemistry are becoming increasingly integrated, where systems are now viewed in a more holistic fashion, such as multi-element or multi-isotopic studies of biogeochemical cycles. Such approaches seem likely to continue in the future, and they offer a comprehensive way to test multiple hypotheses and address geologic questions that continue to be important as we use geochemistry to better understand the geologic history of Earth and the solar system.

page 3 3

INTRODUCTION Fifty years ago, in the early 1960s, modern geochemistry had begun to take shape, and from efforts during the rst half of the twentieth century, many of its fundamental questions had come into focus. Even biogeochemistry, arguably the most modern branch of geochemistry, had been pioneered years earlier by Vernadsky and Baas Becking. The rst journal specically devoted to the eld, Geochimica et Cosmochimica Acta, began publication in 1950. Two years later, the rst English-language geochemistry text was published by Mason (1952), building on the pioneering efforts of his mentor Goldschmidt (whose classic tome was published posthumously two years later; Goldschmidt, 1954). The year 1952 also witnessed Nobel Prizewinning chemist Harold Urey publish his seminal book, The Planets, widely considered a pivotal event in what would become the eld of cosmochemistry. In 1961, the American president John F. Kennedy laid down the challenge of sending humans to the Moon, and within a decade, intense competition for the opportunity to study returned lunar samples would result in major advances in geochemical techniques and thought. On the other hand, the technical aspects of modern geochemistry were very much in their infancy 50 yr ago. Three examples illustrate this in the period leading up to our 50 yr review time frame. High-pressure, high-temperature experimental methods, pioneered at the Geophysical Laboratory by Norman Bowen during the rst half of the twentieth century, were still only capable of exploring conditions relevant to the upper few kilometers of Earth by the early 1960s. In the early 1950s, distribution of the rst international geochemical rock standards, G-1 and W-1,

revealed a disturbing lack of agreement among the major geochemical laboratories of the world (e.g., Fairbairn et al., 1951), and analytical geochemists continued to struggle with data quality until the demanding standards of lunar sample analysis seeped through the broader geochemical community. Eventually, major strides were made as geochemical analyses moved away from techniques based on wet chemistry to those based on X-ray or mass spectra, allowing a marked increase in sample throughput. The use of chemical modeling in geochemistry has taken place almost entirely within the past 50 yr (Bethke, 2008). Garrels and Thompson (1962) rst modeled the speciation of seawater, but the use of computers in geochemical modeling was not introduced until the work of Helgeson (1968). Accordingly, our judgment is that developments in analytical, experimental, and modeling methods have largely facilitated the major advances of modern geochemistry during the last ve decades. These developments have followed three simple but fundamental themes: smaller, better, and more. By smaller, we mean the capability to analyze increasingly smaller samples and thus to evaluate geochemical problems at increasingly ner scales. For example, analytical capabilities now often achieve subfemtogram (10-15 g) sensitivities, permitting study of vanishingly small samples, such as individual interstellar dust particles, or isotopic analysis of micron-sized regions of individual mineral grains. By better, we mean that modern geochemical methods permit everincreasing precision, accuracy, and resolution, greater control on experimental conditions, and more stringent constraints on geochemical modeling. Thus, it is possible to measure many isotopic ratios with precisions approaching parts per million, and Moores law has witnessed >106 improvement in computing capabilities

spe500-08 1st pgs 4 Johnson et al.

page 4

over the past 50 yr. By more, we mean that with modern methods, it is possible to study increasing numbers of samples and everwidening geological conditions. For example, high-throughput mass spectrometers allow for construction of high-resolution paleoclimate proxy records; improvements in high-pressure experimental equipment (e.g., diamond anvils) allow for experiments to be carried out at conditions equivalent to the centers of terrestrial planets; improvements in chemical models permit evaluation of increasingly high ionic strength aqueous conditions. Importantly, these advances in instrumentation occurred during, and after, the emergence of plate tectonics, and in terms of providing a framework for terrestrial geochemical studies, plate tectonics permeated work in the eld. Another conceptual advance in terrestrial geochemistry has been examination of Earth within a planetary perspective, with much insight coming from studies of meteorites and rocks from the Moon and Mars. Exploration of the solar system by spacecraft also occurred during this time, prodding the expansion of geochemical methods into remote sensing. Our charge has been to summarize the remarkable progress that has been made in geochemistry during the last 50 yr in a journal-length chapter. We therefore begin with an apology to our colleagues for the large body of important work that must be omitted by the constraints imposed. Here, we are able to highlight only a small sample of the remarkable advances in geochemistry research in the last ve decades. This contribution is intended to be useful for geochemists who desire a glimpse of progress in areas of research beyond their own, as well as generalists who are curious about what has been going on in those laboratories down the hall. GEOCHEMISTRY OF THE ELEMENTS Major- and trace-element compositions of rocks, minerals, and natural uids have long been understood to be among the most fundamental data of geochemistry. Systematic evaluations of the abundances of the individual elements, and the basic laws governing their distributions in rocks and minerals were largely established by the 1960s, pioneered by the likes of Frank Clarke, Victor Goldschmidt, Louis Ahrens, and Ted Ringwood. Early analytical methods relied mainly on classical gravimetry and X-ray spectrographs to determine element compositions (Mason, 1992). During the 1950s, these laborious techniques began to give way to rapid and sensitive spectrophotometric methods, with the development and availability of instruments such as arc-source optical emission spectrographs and ame photometry, and chemical complexing agents (e.g., EDTA, ethylenediaminetetraacetic acid) that could be used for colorimetry. Advances in bulk chemical analyses over the past 50 yr have witnessed astonishing developments of high-sensitivity instrumentation and methods (e.g., X-ray uorescence, plasma-emission spectroscopy, neutron activation, chromatography, thermal ionization and plasmasource mass spectrometry), allowing for increasingly rapid and precise measurements of increasingly smaller samples (e.g., Gill,

1997; Sutton et al., 2006). A variety of microbeam methods further allows for determination of major- and trace-element (and isotope) compositions of very small volumes of minerals, and to spatially accumulate such data into two-dimensional (2-D) maps and three-dimensional (3-D) tomographic images of the compositions of rocks and minerals at micron to tens of micron scales of resolution. Major-Element Geochemistry Although early chemical analyses were time consuming, by the mid-twentieth century enough major-element data had been accumulated from igneous rocks to provide a basic appreciation of the ways in which fundamental magmatic processes (e.g., partial melting, equilibrium, and fractional crystallization) controlled observed variations (e.g., Harker and alkali-Fe-Mg [or AFM] diagrams). Because igneous and metamorphic processes are largely controlled by equilibrium, as shown, for example, by Norman Bowens studies of igneous systems, they are amenable to high-temperature, high-pressure experimental investigations, and experimental petrology has provided the fundamental background for interpreting chemical compositions. This work continues with the ability to carry out experiments at increasingly extreme conditions, for example, using diamond anvils that can attain pressures approaching the center of Earth (>300 GPa; Mao et al., 1990). One area where fundamental advances in major-element geochemistry have been made over the past 50 yr is in microanalysis. The electron microprobe was developed during the rst half of the twentieth century, but commercial probes only became available in the 1960s (Long, 1995). In order for these instruments to provide quantitative analyses of complex rock-forming minerals, improved understanding of the inuences of atomic number, X-ray absorption and uorescence (so-called ZAF corrections), and various matrix corrections (e.g., Bence and Albee, 1968; Reed, 1995) was required. Applications of rapid quantitative analysis of minerals are legion. Among the pioneering studies were evaluation of phase relations in high-pressure, hightemperature experiments, studies of diffusion from elemental proles through minerals, quantitative study of coexisting mineral equilibria to constrain geothermometry, geobarometry, and oxygen fugacity (e.g., Andersen et al., 1993), and determination of mineral-melt partition coefcients in rocks and experiments allowing for quantitative geochemical modeling (see following). In addition to the electron microprobe, other microbeam techniques have improved spatial resolution and sensitivity (i.e., detection limits) and allowed for elemental mapping (Jansen and Slaughter, 1982), including 3-D tomographic imaging (Jerram and Higgins, 2007). Examples include analytical electron microscopy, secondary ionization mass spectrometry, laserablation sources for emission and mass spectrometers, and synchrotron X-ray uorescence microprobes. Figure 1 provides one example of modern geochemical mapping using a synchrotron X-ray uorescence microprobe.

spe500-08 1st pgs Five decades of advances in geochemistry Sedimentary Geochemistry and the Chemical Index of Alteration (CIA) Concept Unlike igneous systems, where major-element distributions have been studied in great detail for decades, interpretations of the elemental geochemistry of sedimentary systems until recently lagged far behind. The main reason for this is that siliciclastic sediments (sandstones, shales) are mostly physical mixtures and have been affected by multiple episodes of kinetically dominated water-rock interaction prior to, during, and after sedimentation. Accordingly, there was no simple measure, analogous to partition coefcients or Harker diagrams, that quantitatively linked bulk chemistry to process. Well into the 1970s, textbooks routinely tabulated chemical analyses of sedimentary rocks (e.g., Pettijohn, 1975) but provided little in the way of quantitative interpretation. Weathering processes were understood but focused primarily on gains/losses in soil proles to evaluate weathering intensity as an index of paleoclimates. Experiments and eld investigations made signicant progress in evaluating the kinetics and time scales of weathering (Brantley and Lebedeva, 2011; White and Brantley, 1995). Garrels and Mackenzie (1971) rst began to describe the relations between low-temperature aqueous geochemistry (e.g., mineral aqueous stability diagrams) and the bulk composition of sedimentary rocks. Accordingly, a major advance in sedimentary geochemistry was development of the so-called chemical index of alteration (CIA) concepts during the 1980s and 1990s, in a series of papers by Wayne Nesbitt and coworkers (summarized in Nesbitt, 2003).

page 5 5

Developed rst to evaluate weathering in soil and weathering proles (Nesbitt and Young, 1984), this approach also allowed for quantitative, petrologically based understanding of the majorelement composition of siliciclastic sediments. Over the past several decades, few papers reporting the chemical composition of siliciclastics have omitted geochemical concepts in interpreting the data. One especially inuential diagram is the feldspar ternary diagram (in mole fraction), Al2O3(CaO* + Na2O)K2O, or A-CN-K (Fig. 2), where CaO* is CaO in silicate minerals only (i.e., corrected for carbonates and phosphates). The A-CN-K ternary diagram captures most of the major-element (and mineralogical) changes observed in weathering of igneous rocks (i.e., alteration of feldspar and glass to clay minerals). The CIA scale (CIA = 100 Al2O3/[Al2O3 + CaO* + Na2O + K2O]) is shown on the left side of Figure 2. This diagram also proves useful for interpreting the geochemistry of sedimentary rocks for several reasons: Minerals most relevant to sedimentary rocks plot at well-separated locations on the apices or along joins, and twocomponent mixing/unmixing relationships plot as straight lines and follow the lever rule. Accordingly, these relations have been used to quantify or constrain (1) the degree of weathering affecting sedimentary source regions, which typically has paleoclimatological implications; (2) mineral sorting and simple two-component mixing; (3) average provenance and mixing of provenance/mineral components; and (4) diagenesis. Variants on this diagram are the A-CNK-FM and A-CNKM-F diagrams (F-FeOT; M-MgO), which have less thermodynamic/kinetic foundation (Nesbitt and Young, 1984) but are especially useful in basaltic systems (Nesbitt and Wilson, 1992), and thus have been used in planetary applications (Hurowitz and McLennan, 2007; McSween and Keil, 2000). Trace-Element Geochemistry During the past ve decades, some of the most signicant advances in elemental chemistry have had to do with quantitative interpretations of trace-element distributions in rocks and minerals (as pioneered by Louis Ahrens, Paul Gast, Larry Haskin, Denis Shaw, and Ross Taylor), and these advances have naturally mirrored developments in analytical geochemistry. Partition Coefcients One such advance was the determination of partition coefcients (solid-liquid, solid-gas; expressed as Kd) from experiments and natural rock systems over a broad range of pressuretemperature-composition, beginning in earnest during the 1970s and 1980s. These data provided the foundation to the eld of trace-element modeling of igneous processes (see reviews in Allgre and Minster [1978] and Green [1994]). Early studies were limited by the detection limits of the electron microprobe used to measure trace-element abundances in minerals and melts. Experiments were highly constrained by the necessary balancing act between having trace-element abundances high enough to be

Figure 1. Synchrotron X-ray uorescence continuous scan mapping of a Jurassic crinoid stem taken at the Brookhaven National Synchrotron Light Source Beamline X-26A. Image is 6.1 mm by 5.3 mm, collected with a pixel size of 10 mm at a 100 ms/pixel collection rate. Three maps of Ca (green), Fe (blue), and Sr (red) are superimposed, with green being suppressed in this image. Areas of high strontium content (red) are magnesian calcite, whereas the blue areas are ferroan calcite inll. Image is courtesy of Steve Sutton and Tony Lanzirotti (University of Chicago/Brookhaven) and Troy Rasbury (State University of New York at Stony Brook).

spe500-08 1st pgs 6 Johnson et al.

page 6

Figure 2. Ternary diagrams plotting molar proportions of Al2O3(CaO* + Na2O)K2O, or A-CN-K. CaO* refers to calcium in the silicate components only (i.e., corrected for carbonates and phosphates). The chemical index of alteration (CIA) scale is shown on the left side of the diagram. Also plotted are selected igneous and sedimentary minerals, average compositions of major igneous rock types, average shale, and the range for most river waters. The horizontal line connecting plagioclase and K-feldspar separates the lower part of the diagram dominated by primary igneous minerals (i.e., unweathered) from the upper part dominated by clay minerals (i.e., weathered). (A) Typical weathering trends where the arrows schematically represent the general pathways observed for increasing degrees of weathering for various rock types, based on eld studies of weathering proles and on thermodynamic-kinetic modeling. (B) Plot of suspended sediment from many of the worlds major rivers, illustrating how the overall effects of weathering are preserved within the composition of sediments.

detected by the microprobe but not exceeding the Henrys law limit of remaining a trace element within the system (i.e., ai = kxi as xi 0 in some phase where i = element of interest, a = activity, x = molar concentration, and k = Henrys law constant). The development of highly sensitive modern microbeam techniques (e.g., ion microprobe, synchrotron X-ray uorescence, laser-ablation mass spectrometry) now allows for precise determination of a wide range of partition coefcients, including those for highly incompatible elements (i.e., Kdmineral/melt < 10-3), under conditions that represent nature (e.g., Frei et al., 2009a). Rare Earth Elements The rare earth elements (REEs; LaLu, Y1) are the most studied and inuential group of trace elements, and they have provided crucial evidence for a wide variety of geochemical and cosmochemical processes. In addition to elemental abundances, the REEs are used for several important radiogenic isotope systems (e.g.,
1

Scandium may also be considered a REE but typically is not included with the others in geochemical discussions due to its smaller size and differing partitioning behavior.

Sm-Nd, Lu-Hf), discussed later herein. REEs have been used to address problems ranging from constraining the earliest history of the solar nebula based on variations in condensation temperatures, coupled with measurements of their abundances in Ca-Alrich inclusions and minerals in meteorites (Mason and Taylor, 1982), to tracing the movement of water masses through the oceans based on their short residence times (Piepgras and Wasserburg, 1980). The past ve decades have also witnessed developments in the use of REEs as components for a variety of high-technology applications, such that these elements are now considered to be strategic metals, thus providing further impetus for future geochemical research (Haxel et al., 2005). Modern research on the REEs dates from the development of efcient separation methods and high-precision instrumental analytical techniques in the early 1960s that resolved, in the afrmative, the long-standing question of whether or not these elements could be fractionated during formation of Earths crust (Haskin and Gehl, 1962). By geochemical standards, REEs are an extremely coherent group in terms of size (ionic radius), charge, mineral cation site coordination, lithophile behavior, and aqueous complexing and speciation characteristics. From a planetary perspective, REEs

spe500-08 1st pgs Five decades of advances in geochemistry do not fractionate signicantly during planetary formation from the solar nebula, and, accordingly, average chondritic meteorite abundances serve as a useful reference for examining planetary and geological processes. Under geological conditions, REEs are trivalent, except for the distinctive redox chemistries of europium (Eu3+ or Eu2+) and cerium (Ce3+ or Ce4+), which result in unique insights into magmatic and aqueous processes, respectively. Reduction of Eu (17% increase in ionic radius) occurs under highly reducing conditions, which, with rare exception, exist only within magmatic environments. On the other hand, oxidation of Ce (15% decrease in ionic radius) is common under surcial aqueous conditions, such as those encountered during weathering. The REEs have proven particularly valuable for constraining magmatic processes because their mineral/melt partition coefcients, which (apart from Eu2+) vary smoothly as a function of atomic number, vary over several orders of magnitude among the common rock-forming minerals (Fig. 3). For example, the ionic radius of Eu2+ is virtually identical to that of Sr2+ and readily substitutes for Ca2+ in plagioclase. Accordingly, the presence of Eu anomalies in magmatic rocks commonly results from fractionation of plagioclase during partial melting or crystallization. Since plagioclase is stable only up to ~1 GPa pressure (~40 km depth on Earth), the presence of Eu enrichments or depletions in magmatic rocks generally indicates relatively shallow conditions. The ubiquitous presence of chondrite-normalized negative Eu anomalies in sedimentary rocks is thus interpreted to indicate that intracrustal igneous differentiation processes dominated formation of the upper continental crust (the source of sedimentary REEs). In another example, the presence of very steep REE patterns that are depleted in heavy REEs (high Gd/Yb) in magmatic rocks is taken to indicate fractionation of garnet, a mineral only stable at pressures higher than ~1 GPa in ultramac systems and thus indicative of mantle origins. Accordingly, the origin of steep REE patterns in the ubiquitous Archean tonalite-trondhjemitegranodiorite suites has been central to development of models for the origin of Archean continental crust, and determination of whether or not the modern style of plate tectonics was operating at that time (Taylor and McLennan, 2009). Although all REEs are cosmochemically refractory elements (50% Tcondensation > 1300 K at 10 Pa), slight differences in their condensation temperatures lead to remarkably complex REE patterns in 4.567 Ga refractory Ca-Alrich inclusions, the oldest material preserved in certain chondritic meteorites. In addition to unusual and highly variable overall shapes, the REE patterns of these objects also have both positive and negative anomalies involving the least refractory REEs (Ce, Eu, and Yb). These patterns thus provide persuasive evidence for very complex but relatively localized evaporation-condensation processes in the early solar nebula (Mason and Taylor, 1982). The REEs have proven to be extremely useful tracers for understanding a wide range of aqueous processes (Byrne and Sholkovitz, 1996), primarily because they tend not to partition into the aqueous uid (Duid/solid <<1) during uid-rock interactions. Accordingly, REEs have been shown to be resistant to

page 7 7

remobilization beyond the mineralogical scale during weathering, diagenesis, and metamorphism, except under conditions of very high uid/rock ratios. REEs are particle-reactive and tend to be efciently scavenged from seawater by sediment particles, resulting in very low residence times, ranging from ~50 yr (Ce) to ~2900 yr (Lu). In seawater, Ce is readily oxidized on Mn-oxide particle surfaces, and insoluble ceric oxides and hydroxides are then preferentially removed from seawater compared to the other trivalent REEs, resulting in extreme negative Ce anomalies. The stability constants for most REE complexes tend to increase with increasing atomic number, resulting in the light REEs (LaSm) being preferentially scavenged over GdLu. The overall effect of these characteristics is that REEs in natural uids vary by many orders of magnitude in absolute abundances, with highly variable chondrite-normalized patterns, providing a very sensitive tracer.

Figure 3. Plot of mineral-melt partition coefcients (Kd) for the rare earth elements (REEs) in major minerals in equilibrium with basaltic melts. Data are from a compilation in White (2012).

spe500-08 1st pgs 8 Johnson et al.

page 8

Siderophile and Platinum Group Elements A second group of trace elements that has proven especially useful for evaluating geochemical phenomena is the siderophile trace elements in general, including the refractory and highly siderophile platinum group elements (PGEs = Ru, Rh, Pd, Os, Ir, Pt). The major value of highly siderophile elements is their strong partitioning into metal phases during metal-silicate equilibrium (Dmetal liquid/silicate liquid >> 103; Righter, 2003). Accordingly, their distributions provide insight into the nature of core-forming processes in planets and in the parent bodies of the various meteorite classes, and into the evolution of planetary mantles after extraction of cores. One key discovery, based on PGE abundances, was that the mass extinction at the ca. 65 Ma Mesozoic-Cenozoic (Cretaceous-Paleogene; K-Pg) boundary was related to the impact of an asteroid, thus opening the door to an entirely new way to consider the relations between biologic evolution and geologic history. Alvarez et al. (1980) determined that clay-rich sediment deposited exactly at the K-Pg boundary was highly elevated in iridium, typically by ~20200 times the levels in the enclosing sediments (Fig. 4). Since PGEs (including Ir) occur at very low levels in Earths mantle and crust due to sequestration into the metallic core during planetary differentiation, they argued that the likely source of globally distributed Ir was from impact of a large meteorite. Mass-balance calculations indicated the impactor, assuming a CI chondrite composition, would have been ~10 km in diameter. Much subsequent work, including the presence within the K-Pg sediment of additional PGE enrichments, as well as Re-Os isotope data, in addition to the presence of soot, shocked quartz, stishovite, and impact glass spherules, reinforced this model, and discovery of the buried ~200-km-diameter Chicxulub crater in Mexico with a 65.0 Ma age (e.g., Swisher et al., 1992), has conrmed that a major impact took place at that time. Siderophile element distributions are also important in cosmochemistry (McSween and Huss, 2010). In addition to elemental abundances, siderophile elements possess important radiogenic isotope systems (e.g., Re-Os, Pt-Os, Hf-W) that provide details about planetary differentiation time scales (Kleine et al., 2009; Shirey and Walker, 1998). The family of siderophile elements encompasses a broad range of siderophile tendencies (as measured by metal-silicate partition coefcients under different pressure [P], temperature [T], and oxygen fugacity [fO2] conditions), volatilities (condensation temperatures), and incompatibilities (mineral/silicate melt partition coefcients). Accordingly, they provide insights into core-mantle differentiation and silicate mantle melting in the terrestrial planets, Moon, and meteorite parent bodies (Righter, 2003; Righter and Drake, 1996; Walker, 2009). ISOTOPE GEOCHEMISTRY The eld of isotope geochemistry offers the opportunity to look at the Periodic Table of Elements in a third dimension, where isotopic variations may result from stable isotope fractionation or radioactive decay. Isotope geochemistry provides

an important means with which to trace mass uxes among Earths reservoirs that are independent of concentration effects in the sense of activity coefcients and the thermodynamics of mixing. Historically, stable and radiogenic isotopes were studied by different laboratories, reecting the distinct instrumentation required by various techniques, but such boundaries have now blurred. Our theme of smaller, better, more is embodied in the history of development of mass spectrometers. Modern mass spectrometers capable of high-precision isotopic measurements are based on the Nier (1940) geometry, and by the 1960s, the isotope ratio mass spectrometers (IRMS) available to the stable isotope community had a well-established dual-inlet system for comparing sample and standard gases, as well as simultaneous collection of two isotopes. Later, continuous-ow systems would come online (Hayes et al., 1990), and after that, online laser uorination systems (Sharp, 1990). The thermal ionization mass spectrometers (TIMS) available to the radiogenic isotope community 50 yr ago, however, could not match the precision of dual-inlet, double-collector IRMS instruments. As electronics continued to improve, as well as the addition of multicollection, precisions attainable using TIMS instruments increased markedly in the 1980s (Thirlwall, 1991). Negative ion-capable TIMS later provided a breakthrough for the Re-Os isotope system (Creaser et al., 1991). A major innovation for both stable and radiogenic isotope geochemistry was development of a multicollector, magnetic-sectorbased inductively coupled plasmamass spectrometer (MC-ICP-MS) in the 1990s (Halliday et al., 1998). Finally, efforts have focused on in situ isotopic analysis at the micron scale from the 1990s to present. This includes secondary ion mass spectrometry (SIMS, or ion microprobe), where the current state-of-the-art, large-radius, multicollector instruments can determine d18O precisions, for example, of 0.3 on ~10 m spots (Valley and Kita, 2009). Laser-ablation (LA) coupled to MC-ICP-MS has also emerged as an important technique for in situ isotopic analysis, a technique widely applied, for example, to Hf isotope analysis (Grifn et al., 2000), and new research is currently investigating ultrafast (femtosecond; 10-15 s) lasers (Poitrasson et al., 2003). Next, we touch on some of the major stable and radiogenic isotope systems used in geochemistry in the last ve decades. Space limitations prevent us from covering some important isotopic systems, such as rare gas isotopes, which have seen important applications, for example, to mantle evolution using 3He/4He ratios (Kurz et al., 1982). Other important isotopic systems we have omitted include cosmogenic radionuclides, which, in solid Earth geochemistry, have provided unique constraints on sediment subduction using isotopes such as 10Be (Tera et al., 1986). Stable Isotopes Broadly, stable isotope variations reect partitioning between phases that results from differences in zero-point energies for isotopically substituted species, although in detail, complexities exist that are mass-independent or reect nonstochastic

spe500-08 1st pgs Five decades of advances in geochemistry distribution of rare isotopes. The basic thermodynamic framework for stable isotope geochemistry was laid out by Bigeleisen and Mayer (1947) and Urey (1947). Here, we rst focus on the main players that existed 50 yr ago, H, C, O, and S, followed by a discussion of newer stable isotope systems that have been developed since then. Hydrogen and Oxygen Isotopes The importance of water in the geologic cycle logically placed H and O isotopes as early targets in stable isotope geochemistry (Fig. 5). Signicant accomplishments in the 1960s were numerous. The canonical relation between D/H and 18 O/16O ratios of meteoric waters was determined (Craig, 1961), dening the meteoric water line, a relation that is a key factor to understanding the origin of uids and uid-rock interactions. The kinetic and equilibrium isotopic fractionations associated with meteoric waters, as well as temperature, latitude, and orographic effects, were established (Craig et al., 1963; Friedman

page 9 9

et al., 1964), and especially large seasonal and temperature effects were documented in the d18O values of precipitation in polar regions (Dansgaard et al., 1969; Epstein et al., 1963), setting the stage for use of oxygen isotopes from ice cores to infer paleoclimate. The D-d18O relations of hydrothermal waters that underwent uid-rock interaction were recognized to be different from those produced by evaporation of meteoric waters (Craig, 1966). Initial efforts began to study the d18O values of the ancient oceans when Perry (1967) suggested that the Precambrian oceans had much lower d18O values than the modern ocean based on analysis of cherts. Muehlenbachs and Clayton (1976) proposed the important concept that the d18O value of seawater was buffered near zero by extensive water-rock interactions at mid-ocean ridges (MORs), which would imply that the d18O value of ancient seawater would be roughly invariant as long as plate tectonics operated. A landmark study of the Skaergaard intrusion combined detailed eld studies with numerical modeling of convective heat transport to quantify water-rock interaction at a fossil hydrothermal system (Norton and Knight, 1977; Norton and Taylor, 1979). Studies of ophiolites (Gregory and Taylor, 1981) conrmed that extensive hydrothermal interaction at MORs has

Figure 4. Stratigraphic variations in iridium (Ir) concentrations (parts per billion by mass) in sedimentary rocks at the CretaceousPaleogene (Tertiary in older nomenclature) boundary in the vicinity of Gubbio, Italy. Analyses are for 2 molar nitric acid insoluble residues of limestones and boundary clay. The highly elevated Ir abundances at this stratigraphic boundary were the rst direct evidence that a major meteorite impact was involved with the mass extinctions at the 65 Ma Mesozoic-Cenozoic boundary. Figure is adapted from Alvarez et al. (1980).

Figure 5. D-d18O variations for waters and rocks. Meteoric water line (MWL) and standard mean ocean water (SMOW) are shown for reference, as is box for magmatic waters. Hydrothermal uids are shown for Salton Sea (Craig, 1966) and Lassen (Janik et al., 1983), which are shifted from the MWL, where the arrows mark increasing rock inuence (low water/rock ratios). Sedimentary formation waters may lie along a slope lower than that of the MWL, reecting kinetic effects of evaporation (eld shown for oil-eld brines from California; see Sheppard, 1986). Hydrothermally altered igneous rocks have distinct D-d18O trends compared to waters, and data shown for the Idaho batholith (Criss and Taylor, 1983); arrow marks direction of increasing water inuence (high water/rock ratios). Diagram is adapted from those in Criss (1999).

spe500-08 1st pgs 10 Johnson et al.

page 10

been responsible for xing the d18O value of seawater. If the d18O value of seawater was relatively constant, the low-d18O values of Precambrian cherts could be interpreted to reect very high ocean temperatures (Knauth and Epstein, 1976), although the possibility exists that very old samples have later exchanged with uids that lowered their d18O values after deposition. Oxygen isotope studies of igneous and metamorphic rocks in the 1970s and 1980s focused on (1) use of d18O values as a tracer of the sources of magmas, and (2) application of O isotope thermometry to metamorphism (Valley, 1986) and hydrothermal alteration (Gregory and Criss, 1986). The large decreases in d18O values that occur from meteoric hydrothermal alteration at elevated temperature (e.g., Criss and Taylor, 1983), or the large increases in d18O that occur during weathering (e.g., Savin and Epstein, 1970) produce a wide range in d18O values for uppercrustal rocks that may be incorporated into magmas. Elevated d18O values are characteristic of S-type granites, a term used for granites that contain a sedimentary component, and conrmed by O isotopes (ONeil et al., 1977). In contrast, Friedman et al. (1974) discovered that magmas could attain low-d18O values through interaction with meteoric waters directly, or through assimilation of hydrothermally altered country rocks. Collectively, this discussion points to one of the most important contributions of stable O isotope geochemistry: A rock that has a d18O value that is distinct from that of the mantle probably contains O that was cycled near the surface of Earth in the presence of water (Taylor and Sheppard, 1986). One of the most profound demonstrations of the use of O isotopes as a water tracer is found in the >4 Ga zircons from the Jack Hills, Western Australia, which have elevated d18O values, providing compelling evidence that liquid water existed on Earth over 4 b.y. ago (Cavosie et al., 2005; Mojzsis et al., 2001; Peck et al., 2001). This work, in fact, highlights the importance in isotope geochemistry of moving away from bulk sample analysis to in situ approaches (e.g., Valley and Kita, 2009). Without the ability to make precise, and accurate, O isotope analyses of micron-sized spots, the elevated d18O values found in the Jack Hills zircons would not have been discovered, and the concept of a cool early Earth (Valley et al., 2002) would not have arisen. Carbon Isotopes By 1960, it was already well established that photosynthetic xation of CO2 into organic carbon produced a decrease in d13C values by ~2530 (Park and Epstein, 1960). In the following years, the rst surveys of d13C values for marine carbonates and organic C conrmed that the overall isotopic fractionation between these C reservoirs can be found in natural samples, including those from Precambrian rocks (Hoefs and Schidlowski, 1967; Keith and Weber, 1964). As the database for C isotope compositions of carbonates greatly expanded in the following decade, it became clear that the vast majority of Ca-Mg marine carbonates had a restricted range in d13C values near zero for most of Earth history, which was interpreted to reect a relatively constant balance between the organic and inorganic

C reservoirs (Schidlowski et al., 1975). An important exception was Paleoproterozoic carbonates, which had unusually positive d13C values, rst documented in the Lomagundi Group, Rhodesia (Schidlowski et al., 1975), and later shown to be global and correlative with a major rise in atmospheric oxygen, likely, at least in part, reecting organic C burial (Karhu and Holland, 1996). In contrast to the generally zero d13C values for Ca-Mg carbonates, Fe-rich carbonates from Precambrian banded iron formations (BIFs) have signicantly negative d13C values. Early workers suggested microbial oxidation of organic carbon as an explanation (Becker and Clayton, 1972), but later work favored vertical zonation in d13C values for dissolved inorganic carbon (DIC) in the oceans (Beukes et al., 1990; Winter and Knauth, 1992). More recently, geochemical modeling and studies of shelf-tobasin transects suggest that vertical zonation in d13C values for DIC is unlikely, swinging the interpretation for negative d13C values back to microbial respiration (Beukes and Gutzmer, 2008; Fischer et al., 2009). By the 1970s, there was already a substantial database for C isotope compositions of kerogen from Precambrian shales as old as 3.4 Ga, documenting that d13C values for organic carbon generally lay between -25 and -35 (Oehler et al., 1972). An important exception was discovery of highly negative d13C values for organic carbon in sedimentary rocks between ca. 2.8 and 2.6 Ga in age, down to 60. Such low values are accepted to reect a role for involvement of methane, and Hayes (1983) speculated that aerobic methanotrophy might have been responsible. Follow-up work has conrmed these unusually low d13C values and documented correlations with sedimentary facies that suggest a transition from an anaerobic ecosystem to one supported by oxygenic photosynthesis at the end of the Archean (Eigenbrode and Freeman, 2006). Looking at the oldest known sedimentary rocks, Schidlowski et al. (1979) noted that graphite in metasedimentary rocks from the Isua belt, SW Greenland, did not generally reach the low-d13C values that are characteristically thought to reect photosynthetically xed C, and this was interpreted to reect the effects of metamorphism (Schidlowski, 1987). The importance of obtaining primary d13C values from C from Early Archean rocks, despite their commonly high grade of metamorphism, has generated numerous studies. Mojzsis et al. (1996) documented d13C values for graphite from Akilia Island, SW Greenland, obtained via in situ SIMS analysis, that were signicantly lower than those measured in bulk samples, but this work has come under criticism on a number of fronts, including arguments (1) that the graphite is not photosynthetic in origin, but instead a breakdown product of Fe-bearing carbonates (van Zuilen et al., 2002), (2) that the sample analyzed was not a sedimentary rock but a metasomatic dike (Fedo and Whitehouse, 2002), and (3) that the graphite analyzed was not enclosed in apatite and therefore isolated from the effects of metamorphism (Lepland et al., 2005). The question of when the rst C isotope compositions that suggest photosynthesis appeared on Earth remains an important one, although there is no consensus on the answer.

spe500-08 1st pgs Five decades of advances in geochemistry Recent work in C isotope geochemistry has expanded into in situ isotopic analyses, including studies of organic carbon and individual microfossils (House et al., 2000). Increasingly, it is recognized that the range in d13C values for individual microfossils, or kerogen on submillimeter scales, is much larger than would be suggested by bulk C isotope analysis. Such an approach has affected the range in C isotope fractionations inferred between organic and inorganic C, which in turn may constrain atmospheric CO2 contents (Kaufman and Xiao, 2003), test the indigenous nature of molecular biomarkers (Rasmussen et al., 2008), or distinguish between eukaryotic and cyanobacterial photosynthesis (Williford et al., 2013). Sulfur Isotopes Early studies of S isotope variations in the laboratory and natural environments had documented a large fractionation in 34 32 S/ S ratios during microbial sulfate reduction that was dependent upon the rate of reduction and the abundance of sulfate, and similar ranges in isotopic compositions had been documented in modern marine sediments (Kaplan et al., 1963; Thode et al., 1961). In contrast to the strongly positive d34S values measured for Phanerozoic sulfates, Perry et al. (1971) found that older than 3 Ga barites from South Africa had only slightly positive d34S values, which they interpreted to reect very low seawater sulfate contents, suggesting very low atmospheric O2 contents. Detailed experimental work on microbial reduction and S disproportionation in the 1990s demonstrated that extremely large 34S/32S fractionations could be produced during microbial S cycling that involved S species of intermediate oxidation state such as sulte and thiosulfate (Caneld and Thamdrup, 1994; Caneld and Teske, 1996). This in turn suggested that the exceptionally low d34S values in sedimentary suldes of younger than 1 Ga age reected an increase in the oxidative component of the S cycle, which in turn provided strong evidence for a major rise in atmospheric O2 contents in the Neoproterozoic. The decrease in d34S values of suldes in the Neoproterozoic, relative to earlier time periods, was accompanied by an increase in the estimated d34S values for seawater sulfate (Caneld, 2001). Burdett et al. (1989) proposed that the isotopic compositions of seawater sulfate may be determined through analysis of carbonate-associated sulfate (CAS), an approach that greatly extends the lithologies that may be used to infer ancient seawater S. The CAS proxy proved particularly valuable for the ancient rock record, where the d34S values of ancient seawater sulfate had been previously inferred indirectly based on the maximum values obtained in a suite of sedimentary suldes. Application of the CAS proxy to the Proterozoic has conrmed expectations that seawater sulfate contents were very low in the Paleoproterozoic, before or immediately after the Great Oxidation Event (GOE), but rose substantially in the Neoproterozoic at the time of the second increase in atmospheric oxygen (Kah et al., 2004). The last decade has seen a rapid increase in in-situ S isotope studies, primarily focused on pyrite. These efforts have shown that detrital, authigenic, and hydrothermal components may

page 11 11

be recognized in the same single pyrite grain (Williford et al., 2011). Work on ca. 3.4 Ga rocks from the Pilbara craton, Australia, has documented signicant S isotope variations on the micron scale, which have been generally interpreted to reect biological cycling of S, including S0 disproportionation and sulfate reduction (Philippot et al., 2007; Wacey et al., 2011). A very large range in d34S values, >35, was observed in pillow lava textures in the Barberton greenstone belt, which were interpreted to reect microbial microboring, and which McLoughlin et al. (2012) suggested provides evidence for a Paleoarchean subseaoor biosphere. Mass-Independent Stable Isotopes Most stable isotope variations in terrestrial systems fractionate in a mass-dependent manner. Mass-independent fractionations (MIF) for O isotopes (and others) are commonly observed in extraterrestrial samples (Birck, 2004), and will be discussed later in this review. Sulfur MIF (commonly termed S-MIF) has been studied extensively in ancient sedimentary rocks as a tracer of past atmospheric O2 contents. The most common mechanism called upon to produce S-MIF is photolysis reactions involving SO2 and H2S in the upper atmosphere, based on ultraviolet (UV) radiation experiments (Farquhar et al., 2001, 2000b). Because ozone is a strong absorbent of UV in the atmosphere, its presence would greatly decrease S-MIF in aerosols, which suggests that the S-MIF recorded in Archean and Paleoproterozoic rocks indicates very low atmospheric O2 contents (Farquhar et al., 2000a). Photochemical modeling suggests that transport of S-MIF to the sedimentary cycle requires ambient atmospheric oxygen contents less than 105 of present day (Pavlov and Kasting, 2002). The specic mechanisms and pathways responsible for creating S-MIF in the ancient rock record, however, remain unclear, where initial UV experiments may not have adequately modeled the full spectrum of UV radiation, and the roles of other gases, including methane and inert gases, are issues that have been raised (e.g., Domagal-Goldman et al., 2008; Lyons, 2009; Masterson et al., 2011; Zahnle et al., 2006). Further complexity arises from the nding that thermochemical sulfate reduction by organics can produce mild S-MIF (Watanabe et al., 2009). Nevertheless, most workers accept S-MIF in Archean and Paleoproterozoic sedimentary suldes as indicating essentially anoxic conditions in Earths atmosphere at this time, although research continues on the ancient rock record, as well as the mechanisms to produce MIF. Nontraditional Stable Isotopes Nearly three quarters of the elements on the Periodic Table of Elements have two or more stable isotopes, but beyond the stable isotope systems discussed here, and a few others, most elements have remained relatively unexplored due to analytical barriers or the opinion that the range of isotopic variations in nature is too small to be worth the trouble. Vanguard efforts in the nontraditional stable isotopes include early studies of Li (Chan, 1987), Si (Douthitt, 1982), and Ca (Russell et al., 1978)

spe500-08 1st pgs 12 Johnson et al.

page 12

isotopes. The largest isotopic variations of the nontraditional stable isotopes are observed for Li, where the range in 7Li/6Li in nature exceeds 75 (Tomascak, 2004). Smaller variations are found for intermediate-mass elements, where, for example, 26 Mg/24Mg, 44Ca/40Ca, 53Cr/52Cr, 56Fe/54Fe, 65Cu/63Cu, 80Se/76Se, and 98Mo/95Mo ratios vary by ~510 in natural samples (e.g., Albarde, 2004; Anbar, 2004; Beard and Johnson, 2004a; DePaolo, 2004; Johnson and Bullen, 2004; Young and Galy, 2004). Results from additional stable isotope systems are being regularly reported at meetings and in publications, extending all the way up to mass U (e.g., Weyer et al., 2008). The eld is changing so rapidly that the rst review of the subject in 2004 is now signicantly out of date (Johnson et al., 2004). One of the most intensively studied nontraditional stable isotope systems has been Fe. The large bonding changes that occur between Fe3+ and Fe2+ species in solutions and minerals were recognized early as the major driving force for producing Fe isotope fractionations (Beard et al., 1999; Bullen et al., 2001; Johnson et al., 2002; Polyakov and Mineev, 2000). The vast majority of Fe in the crust has a 56Fe value near zero, including low-C, low-S sedimentary rocks (Beard et al., 2003a, 2003b) and most igneous rocks, although there are small variations in high-temperature rocks (Beard and Johnson, 2004b). The largest range in 56Fe values, however, is restricted to organic-rich shales and BIFs, which vary from 4 to +2 (Johnson et al., 2008a; Rouxel et al., 2005; Yamaguchi et al., 2005). The origin of such large variations is debated, where some workers argue that they reect variable extents of oxidation of aqueous Fe2+ (Anbar and Rouxel, 2007), and others interpret such variations, particularly in Fe-rich rocks, to reect microbial Fe3+ reduction (Johnson et al., 2008b). Recently, multiple stable isotope systems have been used to study redox-driven geochemical cycling, greatly expanding the utility of stable Fe isotopes, including Fe and S isotopes (Archer and Vance, 2006), Cr isotopes and Fe redox changes (Frei et al., 2009b), Fe and C isotopes, including carbonate C (Heimann et al., 2010) and organic C (Czaja et al., 2010), and Fe and Mo isotopes (Czaja et al., 2012). In summary, the eld of nontraditional stable isotopes is growing rapidly, although progress is currently hampered by the relative paucity of stable isotope fractionation factors as compared to other stable isotope systems that have been studied for several decades. Clumped Stable Isotopes Stable isotope compositions generally consider substitution of a single minor isotope, such as 13C16O16O or 12C18O16O in carbon dioxide because multiply substituted minor isotopes, such as 13 18 16 C O O, are very low in abundance in nature. Recently, however, the unique aspects of multiply substituted minor isotopes have been exploited (Wang et al., 2004). Now termed clumpedisotope geochemistry, to describe, for example, the enhanced stability of 13C-18O bonds in CO2 and carbonates relative to a random or stochastic distribution of minor isotopes, this eld of stable isotope geochemistry promises great breakthroughs in paleoclimate studies, which require accurate determination

of paleotemperatures, as well as the study of atmospheric gases (Eiler, 2007). For CO2 and carbonate, the measured enrichment in 13C-18O bonds relative to that expected for a random distribution of 13C-18O bonds is dened as 47, reecting nominal mass 47 for 13C18O16O, and this has been shown to be an inverse function of temperature (Ghosh et al., 2006). Because the enhanced stability, or clumping, of 13C-18O bonds, under equilibrium conditions, reects internal, homogeneous equilibrium that is independent of the bulk d13C or d18O values, 47 provides an internal thermometer that does not require knowledge of the isotopic composition of the uid from which the carbonate precipitated. This critical aspect of clumped-isotope geochemistry obviates the need to know, or assume, the d18O or d13C value of ancient seawater when extracting paleotemperatures from marine carbonates. Clumped stable isotopes therefore hold great promise for resolving debates such as the temperatures of the Archean oceans (Kasting et al., 2006; Knauth, 2005). An important component to using clumped isotope thermometry is the preliminary observation that the vital effects that are known to fractionate 13 C/12C and 18O/16O ratios during biologically catalyzed carbonate formation in marine environments do not apparently affect 47 values (Thiagarajan et al., 2011; Tripati et al., 2010). There seems little doubt that as the very demanding analytical issues of clumped-isotope geochemistry are tackled by more laboratories, and rened with improvements and new directions in instrumentation and experimental studies, this area of stable isotope geochemistry will expand. Radiogenic Isotopes Radiogenic isotopes may be used for geochronologic information, or as a genetic tracer. Here, we will generally cover the latter, as the subject of geochronology is covered in another chapter in this series, but it should be recognized that radiogenic isotope systems often provide both types of information. The following discussion is grouped by isotope system, as is traditionally done, but we note that modern radiogenic isotope studies commonly combine multiple isotopic systems, blurring such grouping of subjects. Rb-Sr The isotope 87Rb decays to 87Sr with a half-life (t1/2) of 49 b.y. Work in the 1960s established that Precambrian continental crust had signicantly higher 87Sr/86Sr ratios than mantle-derived rocks, reecting the higher time-integrated 87Rb/86Sr ratios of continental crust (Faure and Hurley, 1963). Early studies of the mantle documented Sr isotope heterogeneity across tholeiitic and alkaline basalts, but they documented a generally nonradiogenic (low 87Sr/86Sr) isotopic composition that was similar to achondrite meteorites (Engel et al., 1965). In contrast, the rst studies of granitic batholiths showed intermediate 87Sr/86Sr ratios, suggesting they were composed of mixtures of mantle and crustal material (Hurley et al., 1965). A landmark paper by Kistler and Peterman (1973) on the Sierra Nevada batholith documented across-arc

spe500-08 1st pgs Five decades of advances in geochemistry variations in 87Sr/86Sr ratios that correlated with outcrop patterns of Precambrian rocks. Following contouring of the data, they proposed that a line of initial 87Sr/86Sr = 0.706 marked the boundary of the continental crust. The 1980s saw an increased focus on interaction between arc magmas and continental crust, including many studies that combined O and Sr isotopes. Taylor (1980) and DePaolo (1981b) advanced the idea of coupled assimilation and fractional crystallization, drawing upon heat balance arguments put forth decades earlier by Norman Bowen. Numerous studies on the covariation of O and Sr isotopes in granitic batholiths provided constraints on the crustal lithologies that were blended in arc magmas, as well as insights into deep crustal architecture (e.g., Fleck and Criss, 1985; Solomon and Taylor, 1989). There has long been an interest in the Sr isotope composition of seawater, and Faure et al. (1965) made one of the rst measurements of modern seawater, found that it was isotopically homogeneous, and proposed that the isotopic composition reected a mixture of weathering inputs from continental and young volcanic components. This homogeneity was conrmed by later studies, leading to the conclusion that the residence time of Sr in the oceans far exceeds the time scales of water mass mixing. The rst detailed survey of marine carbonates of Phanerozoic age sketched out the broad outlines of seawater 87Sr/86Sr variations, identifying periods of radiogenic (high 87Sr/86Sr) compositions in the Carboniferous and late Cenozoic, and relatively nonradiogenic (low 87Sr/86Sr) compositions in the Cretaceous (Peterman et al., 1970). Archean-age carbonates were found to be very nonradiogenic, which was interpreted to indicate minimal input from continental crust, possibly reecting small continental exposure (Veizer and Compston, 1976). Spooner (1976) made the breakthrough that the nonradiogenic Sr ux was more likely to reect MOR hydrothermal input rather than erosion of exposed young volcanic rocks. As mass spectrometry precision improved, both broad (Burke et al., 1982; Veizer et al., 1983) and detailed (Hess et al., 1986) Sr isotope studies of carbonates, including shells and foraminifera, provided a highly resolved Sr seawater curve. This curve has been used for both stratigraphic chronology and to infer changes due to sea-level variations, tectonic activity, weathering, and continental-scale glaciations in the Phanerozoic (DePaolo, 1986; Eldereld, 1986; Raymo et al., 1988). The Rb-Sr isotope research outlined here provided a broad view of Sr isotope variations in the crust and mantle, and in the last decade, there has been an increase in work on determining 87 Sr/86Sr variations in individual minerals, rather that bulk samples, using in situ approaches, which have included microdrilling, SIMS, and LA-ICP-MS. For example, studies on feldspar phenocrysts in igneous rocks have been used to determine openversus closed-system behavior in magmatic systems, where both magma recharge and crustal assimilation have been documented from intracrystal variations in 87Sr/86Sr ratios (Davidson et al., 2001; Knesel et al., 1999). In addition, Sr isotope proles in phenocrysts potentially provide information on crystal residence times when coupled with diffusion modeling across discontinuities (Davidson et al., 2007). In addition to analysis of individual

page 13 13

minerals, Sr isotope measurements of melt inclusions document very large 87Sr/86Sr variations in olivine-hosted melt inclusions from oceanic basalts, which Jackson and Hart (2006) interpreted to reect mixing of primitive magmas that originated from enriched and depleted mantle reservoirs. U-Th-Pb The U-Th-Pb system involves decay of three parent isotopes to distinct Pb isotopes: 238U-206Pb (t1/2 = 4.5 b.y.), 235U-207Pb (t1/2 = 0.7 b.y.), and 232Th-208Pb (t1/2 = 14 b.y.). Early efforts on the U-Th-Pb isotope system focused on determining the evolution of the continental crust and mantle, given the framework that Pattersons (1956) Pb-Pb age of Earth provided. Broad surveys of galena of various ages conrmed that a single-stage growth curve for the crust can explain some data (Kanasewich and Farquhar, 1965). In contrast, Patterson and Tatsumoto (1964) studied detrital feldspars as a widespread measure of North American continental crust and found that the relatively high abundance of 207 Pb, which can only have been produced early in Earths history, required a two-stage growth curve that included an early U/Pb differentiation event between 3.5 and 2.5 Ga, which they interpreted to be the primary formation age for the crust. Armstrong (1968) recognized, as did many others, that the average isotopic composition of modern Pb plotted to the right of the geochron (the Pb-Pb array of the solar system, inferred to include the terrestrial planets), that is, at high 206Pb/204Pb on a 206Pb/204Pb207 Pb/204Pb diagram; this was commonly referred to at the time as anomalous or future Pb, and it was recognized that this likely required a multistage history of U/Pb evolution. The scope of the Pb isotope mass-balance problem (sometimes referred to as the Pb paradox; Allgre, 1968) became clear in studies of basaltic rocks from the ocean basins, where most oceanic basalts lay to the right of the geochron, requiring some sort of multistage history (Gast et al., 1964; Tatsumoto, 1966). An important attempt at solving the Pb paradox and the failure of single-stage growth curves came from the Stacey-Kramers average crustal growth curve, used to this day, which is a two-stage growth curve that involved a global U/Pb enrichment event at 3.7 Ga, possibly reecting the rst major differentiation event on Earth (Stacey and Kramers, 1975). The plumbotectonics model put forth in the late 1970s, and rened in following years, attempted to explain the transport of Pb between the major reservoirs of Earth, and offered a possible explanation to the Pb paradox (Doe and Zartman, 1979; Zartman and Doe, 1981). Moving forward, Pb isotope studies, as well as Sr isotopes, became increasingly integrated with Sm-Nd isotope studies, which provided important breakthroughs in understanding of crust-mantle evolution; these are discussed in the next section. Recent work on in situ Pb isotope analysis by SIMS and LA-ICP-MS has been applied to subjects ranging from sedimentary provenance to melt inclusions. Pb isotope analyses of single detrital feldspars in both modern (Alizai et al., 2011) and ancient (Tyrrell et al., 2007) sediments have been used to infer paleodrainage congurations, the extent of sediment recycling, and mineral

spe500-08 1st pgs 14 Johnson et al.

page 14

diagenesis. In situ Pb isotope analyses of feldspar phenocrysts in large caldera-related volcanic systems have shown correlations between the proportion of mantle and crustal components in phenocryst cargos and eruptive frequency and volume (Simon et al., 2007). Initial work on olivine-hosted melt inclusions in oceanic basalt demonstrated very large ranges in isotopic compositions that greatly exceed those measured in bulk samples, and such ranges have been inferred to reect blending of mantle melts from distinct mantle reservoirs, in addition to wall-rock interactions in the plumbing system (Saal et al., 1998). Later work has suggested less extreme ranges in the isotopic compositions of melt inclusions, but it has also highlighted the utility of in situ isotopic analysis in distinguishing between mantle and crustal processes (Paul et al., 2011). Sm-Nd Arrival of the Sm-Nd isotope system on the scene in the 1970s opened new research avenues due to the relatively restricted REE variations in a large variety of rocks, as compared to Rb/Sr and U/Pb. The most widely used decay system has been 147Sm-143Nd (t1/2 = 106 b.y.), although the 146Sm-142Nd (t1/2 = 0.1 b.y.) decay system has also been used to trace early solar system processes. A powerful component of the Sm-Nd isotope system was the uniform nature of chondrite meteorites, which reected generally limited Sm/Nd fractionation (Jacobsen and Wasserburg, 1980), and this led to a reference reservoir, the chondritic uniform reservoir, or CHUR (DePaolo and Wasserburg, 1976), which would be used extensively in studies of terrestrial rocks for inferring the age of differentiation events. Early work documented a broad anticorrelation between 87Sr/86Sr and 143Nd/144Nd ratios in terrestrial samples (DePaolo and Wasserburg, 1976; ONions et al., 1977; Richard et al., 1976), which was recognized as reecting the distinct time-integrated Rb/Sr and Sm/Nd ratios of depleted (high 143Nd/144Nd) and enriched (low 143Nd/144Nd) components. For oceanic lavas, this relation was initially termed the mantle array. In the three components of Sr, Nd, and Pb isotopes, a mantle plane was proposed (Zindler et al., 1982), later modied in multiple component space by the landmark paper by Zindler and Hart (1986), who dened four principal mantle components: a depleted mantle (DMM) anchored by mid-ocean-ridge basalt (MORB), a high 238U/204Pb () and high 206Pb/204Pb component (HIMU), dened by ocean-island basalts (OIBs) such as St. Helena, a moderate 87Sr/86Sr and low 143Nd/144Nd enriched component (EM I) represented by Walvis Ridge, and a high 87Sr/86Sr and low 143 Nd/144Nd enriched component (EM II), represented by Samoa. It is generally accepted that the DMM component reects depletion of the mantle through long-term production of continental crust, and the three enriched components (HIMU, EM I, and EM II) are generally thought to reect lithospheric recycling (Fig. 6). Mantle xenoliths from the subcontinental lithospheric mantle demonstrated that the DMM, EM I, and EM II components can be identied in the subcontinental lithospheric mantle for some isotopic systems such as Sr and Nd (Hawkesworth et al., 1990; Menzies, 1989).

Covariation of Sr and Nd isotopes in orogenic arcs showed that many arc lavas were shifted to high 87Sr/86Sr ratios relative to the Sr-Nd mantle array (Hawkesworth et al., 1979). Coupled with new Sr and Nd isotope data from ophiolites that showed strong shifts in 87Sr/86Sr ratios but invariant 143Nd/144Nd during hydrothermal alteration at MORs (McCulloch et al., 1981), preferential shifts in 87Sr/86Sr in orogenic arcs were interpreted to reect a subduction component derived from the altered slab. A landmark study by Hildreth and Moorbath (1988) proposed that mixing, assimilation, storage, and homogenization of arc magmas extensively occurred in the lower crust during ponding of mantle-derived basaltic magmas near the Moho, a model known as MASH, and one that still provides a framework for arc magmatism. The relative coherence of Sm/Nd ratios during melting provided a means of seeing through recent magmatic events to infer the time when Nd was extracted from the mantle, an approach that led to the concept of Nd model ages (DePaolo, 1981a). In western North America, Nd model age

Figure 6. 143Nd/144Nd-87Sr/86Sr variations in the mantle as inferred from oceanic mac lavas and mantle-derived xenoliths. Box for midocean-ridge basalts (MORB) denes a depleted mantle (DMM) component, and ocean-island basalts (OIB) extend from MORB to lower 143 Nd/144Nd and higher 87Sr/86Sr ratios toward two enriched mantle (EM) components: EM I (low 143Nd/144Nd, moderate 87Sr/86Sr) and EM II (moderate 143Nd/144Nd and high 87Sr/86Sr). A high U/Pb component (HIMU) is subtly dened for Nd-Sr isotopes, but it is prominent for Pb isotopes. These mantle components were discussed in detail in Zindler and Hart (1986) and Hart (1988). Nd-Sr isotope variations for xenoliths derived from the subcontinental mantle greatly extend the range observed for oceanic basalts, reecting isolation from the asthenosphere, but the data are consistent with the DMM, HIMU, EMI, and EM II components (Hawkesworth et al., 1990). Subduction-related volcanic arcs may deviate from the OIB eld, sometimes toward relatively high 87Sr/86Sr ratios, possibly reecting a component from hydrothermally altered oceanic crust (McCulloch et al., 1981). Subduction-related volcanic rocks may also extend to very low 143Nd/144Nd and high 87Sr/86Sr ratios (data shown for Martinique; Davidson, 1983). Diagram is adapted from those in Dickin (1995).

spe500-08 1st pgs Five decades of advances in geochemistry provinces correlated with Archean, Proterozoic, and Phanerozoic crustal boundaries and allowed identication of older Nd components in granitic batholiths (Bennett and DePaolo, 1987; Farmer and DePaolo, 1983). Neodymium isotope data from large caldera complexes from North America, however, identied large proportions of mantle, indicating that such complexes represented new periods of net crustal growth (Johnson, 1991), suggesting that ignimbrites may contain a larger proportion of mantle than granitic rocks, and hence periods of net crustal growth (Johnson, 1993). The Nd model age concept has been extensively applied to ne-grained sedimentary rocks, and for samples that had depositional ages younger than 2 Ga, Nd model ages were usually older than 2 Ga, indicating that most Phanerozoic- and Proterozoicage sedimentary rocks have been recycled from older sources (Allgre and Rousseau, 1984; Goldstein et al., 1984; ONions et al., 1983). An exception is sedimentary rocks that were deposited at the same time as orogenic episodes, which were found to contain a larger proportion of mantle Nd than nonorogenic sediments, reecting net crustal growth (Michard et al., 1985). Massage distributions for the continents calculated from Nd crustal residence times using sedimentary rocks indicated that ~40% of the present-day continental mass formed by 3.8 Ga (Jacobsen, 1988), a conclusion that is similar to that inferred from Pb isotope compositions of detrital feldspars, as discussed earlier (Patterson and Tatsumoto, 1964). A critical component to extracting crustal growth curves from Nd model ages, however, is quantifying the extent of crustal recycling in sediments (Allgre and Rousseau, 1984); as will be seen later herein, signicantly different crustal growth curves have been inferred using Hf isotope data. In contrast to Sr, the residence time of Nd in seawater is very short, on the order of 102 as, and early researchers in Sm-Nd isotopes recognized that this should make the isotopic composition of Nd in seawater a sensitive indictor of local input (Piepgras and Wasserburg, 1980; Piepgras et al., 1979). This early work, which conrmed the isotopic provinciality of Nd in seawater, and hence its use as a sensitive tracer of water masses, was later extended back in time through analysis of Fe-Mn crusts and authigenic marine sediments. Examples include tracing the distinct evolution of the Pacic-Panthalassa and Iapetus Oceans from the Neoproterozoic to present (Keto and Jacobsen, 1988), and closure of the Central American isthmus that restricted water mass communication between the Pacic and North Atlantic Oceans (Burton et al., 1997). In addition to Nd, Pb isotope studies of Fe-Mn sediments demonstrated that Pb isotopes were also provincial, commensurate with the short residence time for Pb in seawater (Abouchami and Goldstein, 1995). Advances in mass spectrometry in the 1990s allowed application of the 146Sm-142Nd isotope system, which requires measurement of 142Nd/144Nd ratios to a very high precision of several parts per million. The short half-life of the 146Sm-142Nd system provides a sensitive tracer of processes in the rst few hundred million years of solar system history. Harper and Jacobsen (1992) demonstrated an average 142Nd/144Nd enrichment of 32 ppm for

page 15 15

rocks from the 3.8 Ga Isua supracrustal belt, Greenland, relative to average terrestrial rocks. Consideration of coupled 142Nd/144Nd and 143Nd/144Nd variations suggested that the 142Nd enrichment recorded Sm/Nd fractionation in the source reservoir(s) of the Isua rocks likely occurred between 4.55 and 4.45 Ga. These results were conrmed by Bennett et al. (2007), who interpreted the 142Nd enrichment to record early mantle differentiation in the rst 30 75 m.y. of Earths history. Boyet and Carlson (2005, 2006) compared high-precision 142Nd/144Nd data from meteorites and a wide variety of terrestrial mantlederived rocks and documented that chondrite meteorites have 142Nd depletions on the order of 20 ppm relative to the average of terrestrial rocks, and they suggested that this records a major mantle differentiation event ~30 m.y. after Earth formation. Moreover, Boyet and Carlson proposed that there is a missing low-142Nd/144Nd reservoir in Earth, possibly located at the base of the mantle, assuming that Earth formed with a chondritic Sm/Nd ratio. Carlson and Boyet (2008) suggested that the missing reservoir, which should have enriched incompatible trace-element contents, reects high-density components that settled into the deep mantle from an early terrestrial magma ocean. Evidence for an incompatible elementenriched mac crust very early in Earth history lies in rocks from the Nuvvuagittuq greenstone belt, Canada, which has relatively low 142Nd/144Nd ratios that overlap those of chondrites, and which has a 146Sm-142Nd age of ca. 4.3 Ga (ONeil et al., 2008). A contrasting view of the missing 142Nd reservoir is offered by Jacobsen et al. (2008), who argued that the difference in 142Nd/144Nd ratios between terrestrial rocks and chondrite meteorites may be explained by isotopic heterogeneity during Earths accretion, rather than invoking a missing reservoir in Earths mantle. Lu-Hf The isotope 176Lu decays to 176Hf with a half-life of 37 b.y. The large Lu/Hf fractionations that are produced by garnet, as well as the high abundance of Hf in zircon, have long spurred interest in this analytically difcult isotope system. A series of pioneering papers in the 1980s outlined the range in Hf isotope compositions of meteorites, mantle-derived rocks, crustal zircons, and sedimentary rocks (Patchett et al., 1981; Patchett and Tatsumoto, 1980; Patchett et al., 1984). Hafnium and Nd isotope compositions for OIB are broadly correlated, but Hf and Nd isotopes for MORB are not correlated. Early work on zircons from igneous and metamorphic rocks as old as 3.7 Ga suggested a chondritic Hf evolution for the crust until ca. 2.8 Ga, at which time high initial 176Hf/177Hf ratios appeared, indicating the presence of a globally differentiated mantle (Patchett et al., 1981). Later work on detrital zircons indicated the presence of a depleted mantle by 3.0 Ga (Stevenson and Patchett, 1990). Detailed studies of the Hf isotope compositions for the oceanic mantle identied HIMU, EM I, and EM II components dened by Sr-Nd-Pb isotope studies (Salters and Hart, 1991), but Hf isotopes require two DMM components, one of which has high 176Hf/177Hf ratios that requires garnet as an important ancient component in the source regions of MORB (Salters, 1996).

spe500-08 1st pgs 16 Johnson et al.

page 16

The late 1990s saw a marked increase in Lu-Hf research as MC-ICP-MS instruments became widespread. High-precision Hf isotope measurements of juvenile crystalline rocks, as well as ancient sediments, initially documented a limited range in initial 176 Hf/177Hf ratios prior to 3.0 Ga, which stood in contrast with Sm-Nd isotope data that suggested very large ranges in initial 143 Nd/144Nd early in Earths history (Vervoort and Blichert-Toft, 1999; Vervoort et al., 1999), although later work, based on detrital zircons (see following), has since expanded the Hf isotope database for Precambrian rocks by several orders of magnitude and documented a much larger range in initial 176Hf/177Hf ratios. High-quality MC-ICP-MS data for OIBs identied subtle variations in Hf-Nd isotope arrays that provided strong support for an ancient subduction of pelagic sediment that was previously hypothesized but difcult to conrm using TIMS data (BlichertToft et al., 1999). A major effort in combined U/Pb geochronology and Hf isotope studies of detrital zircons, using in situ analysis methods, began in the early 2000s and has continued to the present. Much of this work has been aimed at addressing the growth and evolution of continental crust, given the wide variety of crustal growth histories that have been proposed (e.g., Allgre and Rousseau, 1984; Armstrong, 1981; Condie, 1998; Hurley and Rand, 1969; Taylor and McLennan, 1985). U/Pb zircon geochronology has long highlighted major peaks in crystallization ages at ca. 2.7 and ca. 1.8 Ga, as well as other peaks in more detail (e.g., Condie, 1998), leading a number of workers to conclude that crustal growth has been episodic (e.g., Rino et al., 2004). This in turn has led to models where catastrophic events in the mantle have been invoked to explain punctuated periods of very rapid crustal growth. In contrast, the distributions of Hf isotope model ages for detrital zircons do not show the strong peaks recorded in U/Pb ages, which suggests that the U/Pb age record may reect biases in preservation rather than periods of episodic crustal growth (e.g., Belousova et al., 2010; Hawkesworth et al., 2009, 2010; Voice et al., 2011). Recognizing that Hf isotope model ages in detrital zircons may reect mixtures of mantle-derived and recycled components, Kemp et al. (2006) combined U/Pb geochronology and O and Hf isotopes to separate a recycled high-d18O sedimentary component from juvenile, mantle-derived components that reect net addition to the continents. This approach has forced a major change in thinking about crustal growth rates based on detrital zircons, and it suggests that as much as 70% of current continental crustal volume was produced by 3 Ga, followed by slower, but relatively uniform, rates of crustal growth since that time (Dhuime et al., 2012). Looking at the oldest detrital zircons, the Archean and Hadean zircons from the Jack Hills, Australia, have been a logical target of Lu-Hf studies. Early work on mineral separates by Amelin et al. (2000, 1999) highlighted the difculty in interpreting both Lu-Hf and U-Pb data from complex zircons, and they found little evidence for enriched or depleted components at 3.8 Ga, suggesting minimal differentiation on Earth by this time. Later work by Harrison et al. (2005) reported remarkably

positive Hf values up to +13 at 4.3 Ga, obtained by in situ methods using LA-MC-ICP-MS, and they argued that an active platetectonic system existed at this time. Valley et al. (2006) noted that signicant errors could be introduced in calculating initial Hf values given the complexity of the zircons and the disparate volumes involved in U-Pb (SIMS) and Lu-Hf (LA-MC-ICP-MS) measurements. Harrison et al. (2008) combined Pb and Lu-Hf analysis to constrain 207Pb/206Pb ages to the same ablated volume analyzed for Lu-Hf, and this revised approach found only negative Hf values down to 5 at 4.2 Ga. Recently, Kemp et al. (2010) reported LA-MC-ICP-MS Pb and Hf isotope analyses that produced negative Hf values between 3.9 and 4.3 Ga in age (Fig. 7), which they argued reects only small domains of enriched (low Lu/Hf) components in primitive crust, and not widespread melting and depletion of the mantle, nor an active plate-tectonic system prior to 4 Ga. Re-Pt-Os Radiogenic Os isotope studies have generally focused on the 187 Re-187Os system (t1/2 = 42 b.y.), although the 190Pt-186Os system (t1/2 = 489 b.y.) has also been explored. The Re-Pt-Os isotope system differs greatly from the radiogenic systems discussed previously in that these elements are siderophile/chalcophile (Shirey

Figure 7. Hf(t) vs. age relations for zircons from the Jack Hills, Australia, the oldest known terrestrial samples. The Hf(t) value was calculated from initial 176Hf/177Hf ratios and 207Pb/206Pb ages (U-Pb zircon geochronology). Chondrite uniform reservoir (CHUR) reference is shown, as well as the eld for depleted mantle inferred from mid-ocean-ridge basalts (MORB-DM), which spans a range in present-day Hf values from CHUR to highly positive values (Salters and Hart, 1991). The limiting line for an enriched reservoir, such as continental crust, is shown at Lu/Hf = 0. Some previous studies reported a limited number of positive Hf(t) values (Blichert-Toft and Albarde, 2008; Harrison et al., 2005), which would indicate widespread mantle depletion, but other studies found only negative Hf(t) values (Amelin et al., 1999; Harrison et al., 2008; Kemp et al., 2010); such values indicate the presence of an enriched component, possibly tied to crust formation, but this does not provide conclusive proof of widespread mantle depletion and continental crust formation.

spe500-08 1st pgs Five decades of advances in geochemistry and Walker, 1998), and the major inventories exist in Earths core, distantly followed by the mantle (see previous discussion). Work on this analytically challenging system, which began in the 1980s, showed that the mantle has remained relatively nonradiogenic, lying within the range of chondritic meteorites, but the very high Re/Os ratios of continental crust produced very high 187 Os/186Os ratios over time (Allgre and Luck, 1980; Luck and Allgre, 1983). Correlations between Os and O isotopes in OIBs have been interpreted to reect ancient subducted sediments (Lassiter and Hauri, 1998), or assimilation of hydrothermally altered oceanic crust during ascent of magmas to the surface (Gaffney et al., 2005). In contrast, Escrig et al. (2005) interpreted high 187Os/187Os ratios at Fogo Island to record assimilation of lower continental crust during opening of the Atlantic Ocean. The large Re/Os fractionations that are produced during melting of peridotite have been used to trace the evolution of the lithospheric mantle. Walker et al. (1989) rst proposed this approach for cratonic peridotite xenoliths, where they calculated Os model ages that reect Re depletion events that may record the time of stabilization of the lithospheric mantle. Compilations of Os isotope data from both whole-rock samples and suldes from cratonic peridotite xenoliths commonly show very nonradiogenic 187Os/188Os ratios that may be interpreted as Re depletion ages between ca. 2.5 and 3.0 Ga (Carlson et al., 2005; Pearson and Wittig, 2008). Although Os model ages provide insight into the evolution of the lithospheric mantle not available by other radiogenic isotope systems, complexities such as sulde breakdown, metasomatism, and mantle-melt interactions can cloud interpretations (Rudnick and Walker, 2009). In contrast to the nonradiogenic Os isotope compositions that have been used to infer Re depletion histories, Shirey and Richardson (2011) recently reported high 187Os/188Os ratios in sulde inclusions in diamonds of 3 Ga age or younger, which they interpreted to reect eclogites that originated as subducted oceanic crust that became incorporated in the subcontinental lithospheric mantle. Radiogenic (high) 187Os/188Os ratios in orogenic arc lavas have been interpreted in two ways. Some workers have inferred such compositions to reect subducted sediments (Alves et al., 1999, 2002), although other workers have argued that the low Os content of crustal material is unlikely to shift the Os isotope compositions of the subarc mantle, instead interpreting the radiogenic Os isotope compositions as reecting assimilation of young, mac lower crust (Chesley and Ruiz, 1998; Hart et al., 2003; Jicha et al., 2009). The importance of the second interpretation lies in the potential ability of the Re-Os isotope system to trace interaction with arc crust, which is essentially invisible using other isotopic systems such as O, Sr, Nd, Hf, and Pb isotopes, and which has implications for estimating net crustal growth in orogenic arcs. Turning to the very long-lived 190Pt-186Os system, Walker et al. (1995) rst proposed that core material might be identied in OIBs, often suggested to reect mantle plumes that originate at the core-mantle boundary (e.g., Hawkesworth and Schersten, 2007). Brandon et al. (1999, 2000) observed correlated enrich-

page 17 17

ments in 186Os/188Os and 187Os/188Os in modern OIBs that they interpreted to represent a high-190Pt component from the core, and later work on Archean komatiites identied similar features (Puchtel et al., 2005). Alternative proposals for correlated 187 Os/188Os and 186Os/188Os in OIBs include Pt- and Re-rich materials that could reect ancient crustal components, or ancient pyroxenite in the mantle, although Brandon and Walker (2005) suggested that such materials are unlikely to be common components in OIBs. Recently, Luguet et al. (2008) documented 186Os enrichments in pyroxenites, eclogites, suldes, and Pt-rich alloys in peridotites, providing support for the argument that 186Os may not be a unique tracer of core components in OIBs. ORGANIC GEOCHEMISTRY: INVESTIGATING CARBON COMPOUNDS PRESERVED IN ROCKS The past 50 yr have witnessed exponential growth in our appreciation of the ubiquity and diversity of organic compounds that are preserved in sediments and sedimentary rocks. A number of these substances, or their derivatives formed through diagenetic alteration processes, were discovered in sediments before they were recognized as natural products in living organisms (Ourisson and Albrecht, 1992; Ourisson et al., 1979). Technological innovation combined with the quest for an understanding of fossil fuel composition, formation, and occurrence were the initial drivers of research on sedimentary organic matter in the 1960s and 1970s (Kvenvolden, 2006). As the possibility of gaining evolutionary insights by way of molecular paleobiology became more clear, the driving force for organic geochemistry shifted (Eglinton, 1970; Peterson et al., 2007). In more recent times, it has been recognized that organic molecules in the air, in water, and in sediments carry in their chemical structures and stable isotopic compositions a myriad of environmental and climate signals. Investigating these records on both human and geologic time scales has now become the prime focus of research in organic geochemistry (Eglinton and Eglinton, 2008; Gaines et al., 2009). The Quest to Understand the Origins of Petroleum and Other Fossil Fuels The invention of computerized mass spectrometers, in combination with high-resolution capillary gas chromatography, was a critical early technical development, and this led to the discovery that petroleum was composed of thousands of discrete organic compounds. In the 1960s, geochemists working in the fossil fuel industry, in collaboration with academic scientists, also demonstrated that sedimentary kerogen and bitumen, the precursors of petroleum, could be correlated to commercial deposits of oil and gas using combinations of organic compound distributions and isotopic patterns (Hunt et al., 2002; Peters et al., 2005). They also convincingly demonstrated that buried organic matter was predominantly of biological origin (e.g., Hills and Whitehead, 1966), and that progressive changes in its chemical

spe500-08 1st pgs 18 Johnson et al.

page 18

composition correlated with burial depth and geothermal heating rates, thereby providing a tool with which to assess the nature and distribution of coal, petroleum source rocks, and natural gas (Seifert and Moldowan, 1980, 1978). Transport of organic matter in clays opened a new mechanistic understanding to its preservation and accumulation in source rocks, and a way in which organic carbon burial could be understood in the context of the principles of sequence stratigraphy (Creaney and Passe, 1993; Hedges and Keil, 1995). Diagenetic processes, mediated by microbes, water, and reduced sulfur compounds, render complex biomolecules into both simpler hydrocarbons and much more complex material (kerogen) through polymerization and cross-linking (Adam et al., 1993; Kohnen et al., 1989). Integration of geochemical understanding with sedimentary geology and basin modeling ultimately resulted in a more informed understanding of hydrocarbon systems, including resource and risk assessment (Hunt, 1996; Tissot and Welte, 1978). Development of Molecular Paleontology Paleontologists have long recognized that microbial life has dominated life on Earth for most of its history, yet microscopic fossils comprise an innitesimally small fraction of the biomass that has been present on Earth, and the vast majority of microbes leave no visible trace at all. It was recognized early, however, that microbial lipids derived from their cell walls, membranes, and pigments are preserved, and, in fact, form the dominant organic molecules that can be found preserved in sedimentary rocks. In Figure 8, we summarize some of the important organic molecules used in geochemistry, organized according to contemporary molecular phylogeny. The concept of biomarker chemostratigraphy has emerged as a means for using these molecules to chart evolutionary innovation, mass extinctions and radiations, chemical events in the ocean, and climate change (Gaines et al., 2009). For example, the rise of complex multicellular life in the Neoproterozoic, the transitions in ocean plankton through the Phanerozoic, and the advent and radiation of land plants are among a few of the biological innovations that have been, and continue to be studied using biomarkers. After several decades of study, it is now clear that major paleontological mileposts are accompanied by corresponding biosynthetic innovations that are recorded in molecular fossils. Colonization of the continents by plants, for example, was enabled by invention of a range of structural biopolymers, including lignin and cellulose. Plants built defenses against desiccation and predation utilizing cuticles made of waxy hydrocarbons for the former, and an array of sesqui- (C15), di- (C20), and triterpenoids (C30) and resinitic compounds for signaling and for warding off insect predators. Distinct differences in the chemical defenses utilized by conifers (Gymnosperms) and owering plants (Angiosperms) can be recognized through the presence of distinctive diterpenoids in Paleozoic and early Mesozoic rocks and petroleums (e.g., Noble et al., 1985). An increase in the prev-

alence of triterpenoids is now known to be associated with the increased abundance of owering plants in the late Mesozoic and Cenozoic (Moldowan et al., 1994; Murray et al., 1998; Simoneit et al., 1986). Recently, it has become clear that biogeographical aspects of plant occurrence and dispersal are reected in sedimentary hydrocarbons (Dutta et al., 2011). In the marine realm, micropaleontology informs us that the Paleozoic dominance of green and red clades of planktonic algae gave way to the modern dominance of chlorophyll a + c plankton, an observation that is directly correlated with the prevalence of their respective steroidal and acyclic isoprenoidal hydrocarbons in marine sediments and oils (e.g., Knoll et al., 2007; Rampen et al., 2007; Sinninghe Damst et al., 2004; Volkman, 2003). A practical application of the evolution of complex organisms and changes in the composition of ocean plankton through time has been development of age-diagnostic biomarkers that can place constraints on the timing of deposition of petroleum source rocks (e.g., Grantham and Wakeeld, 1988; Holba et al., 1998). As with studies of algae, sensitive GS-MS (gas chromatographymass spectrometry) methods developed in the last few decades have provided the means with which to study biomolecules that are unique to animal phyla. Cholesterol is the predominant membrane sterol of animals, and there is little else in the way of other preservable molecules that could constitute a robust metazoan biomarker. Basal invertebrate phyla, dened as those least removed from their single-celled ancestors, include sponges, cnidarians, and echinoderms, and these are the exceptions in that they have a wealth of complex biochemistries, possibly reecting defenses against predation. Demosponges are well known for their biosynthetic capacity to produce distinctive terpenoids, including sterols that have a rare 24-isopropycholestane skeleton (Bergquist et al., 1991). 24-Isopropylcholesterols have been found in the calcisponges, hexactinellid sponges, or choanoagellates, and these form a unicellular sister group to sponges (Love et al., 2009). Hydrocarbon derivatives of the distinctive C30 sponge sterols are prevalent in sedimentary rocks and petroleums of Cryogenian (Neoproterozoic) to Early Cambrian age, and they are thought to reect an acme in the abundance of demosponges in sedimentary environments of this time (Love et al., 2009; McCaffrey et al., 1994). The molecular fossil record suggests that demosponges made their rst appearance in the Cryogenian, and this is concordant with molecular clock estimates of metazoan divergence times (Peterson et al., 2007). It is now recognized that biomarker hydrocarbons can be diagnostic for certain kinds of paleoenvironmental circumstances (Fig. 9). Water-column redox stratication, elevated salinity, marine versus nonmarine sedimentation, and clastic versus carbonate environments are some of the conditions that can be encoded into fossil hydrocarbon distribution patterns. A prime example lies in the insights provided by organic geochemical proxies into biogeochemical processes occurring during oceanic anoxic events (OAEs), and especially those associated with mass extinction events. The work of Holser and others (Holser, 1977; Holser et al., 1989) rst recognized

spe500-08 1st pgs Five decades of advances in geochemistry shifts in the isotopic compositions of C, O, and S in marine sediments, interpreted to record the disruption of ocean chemistry and consequent biological mass extinction. Biogeochemists have subsequently observed concomitant signals in organic molecules, redox-sensitive trace elements, and other geochemical proxies. One recurring trend is seen in the prevalence of biomarkers derived from the green sulfur bacteria (Chlorobi) in sediments that were deposited during OAEs. Chlorobi are

page 19 19

anoxygenic phototrophs that use sulde as an electron donor for photosynthesis, and these produce distinctive light-harvesting carotenoid and chlorophyll pigments that may be recorded in preservable structural features (Grice et al., 1996; Sinninghe Damst and Koopmans, 1997; Summons and Powell, 1987). Diagenetic reduction and stabilization of these compounds are enhanced under the strongly reducing (euxinic) conditions that are favored by the green sulfur bacteria.

Figure 8. A tree of life, based loosely on the nucleotide sequences of small subunit ribosomal ribonucleic acid (RNA), to illustrate lifes three domains and the diverse structures of lipids that are characteristic of organisms at the domain level and, in some cases, phylum level. For example, most eukaryotes utilize sterols, and, in a few cases such as diatoms, dinoagellates, and demosponges, there are particular sterols that are very specic. The membrane lipids of bacteria and eukaryotes incorporate hydrocarbon chains that are linear or branched in relatively simple ways, and these are mostly linked to glycerol via ester linkages. Archaea, on the other hand, build equivalent structures using isoprenoidal chains, and they are linked to glycerol with ether bonds. Many of these features are preservable and can be observed in organic matter extracted from sedimentary rocks. Improved understanding of the way in which lipid chemistry relates to physiology, phylogeny, and environment, and deduction of the pathways by which organic molecules become preserved in sediments have been prime accomplishments of organic geochemists in the past 50 yr. GDGTglycerol dibiphytanyl glycerol tetraethers.

spe500-08 1st pgs 20 Johnson et al.

page 20

Figure 9. An example of the way in which marine water-column redox structure can be deduced from the preserved remains of photosynthetic pigments. The carotenoid -carotene is an accessory pigment common to oxygen-producing phototrophs such as algae and bacteria and is, therefore, produced in, and diagnostic for, oxygenated environments. In contrast, isorenieratene is produced by brown strains of the Chlorobi, which are obligately anaerobic photoautotrophic bacteria and which thrive in low-light, suldic environments, typically where the oxycline is as deep as 60 m. Chlorobactene is produced by the green forms of the Chlorobi, which require higher light intensities and are generally found where the oxycline shoals to 2040 m depth. Okenone is characteristic of the purple sulfur bacteria (Chromatiaceae), which require sulde and even higher light intensities. This pigment can be diagnostic for highly restricted environments such as hypersaline lagoons and where oxyclines are at 20 m or less. All these pigments can also be found in redox-structured microbial mats. It is probably no coincidence that the hydrocarbon derivatives of these pigments abound in sediments and petroleum from the Mesozoic and other warm intervals of Earths history and particularly when narrow seaways and lack of ice resulted in sluggish circulation and ready maintenance of deep-ocean anoxia.

Geologists and paleontologists have long debated the origin of the Permian-Triassic extinction, both the cause and the biological response, and work in molecular paleontology in the last decade has provided important insights. Studies of a core drilled into the end-Permian type section at Meishan, China, show that Chlorobi carotenoids are particularly abundant throughout the last few million years of the Permian and into the Early Triassic, implying that shallow-water euxinic conditions were protracted at the type section of the greatest mass extinction of the geological record (Cao et al., 2009; Grice et al., 2005). Examination of other Permian-Triassic transition sections conrms the pattern and suggests that euxinic conditions were pervasive globally, as far as can be discerned from continental margin sediments of the Tethys, Panthalassic, and Boreal Oceans (Hays et al., 2007, 2012). These observations are consistent with other isotopic and inorganic proxies for a global and intense oceanic anoxic event during the Permian-Triassic transition. Independent modeling also suggests that toxic hydrogen sulde would have been upwelling from the deep ocean onto continental shelves and entering the atmosphere (Kump et al., 2005), implying that sulde toxicity

was a contributing factor to the biological mass extinction in the marine realm as well as on land. Co-occurring with the evidence for euxinia, molecular fossils shed light on other aspects of this event, including transitions in ocean plankton (Cao et al., 2009), and a massive terrestrial weathering event (French et al., 2012; Sephton et al., 2005; Xie et al., 2007). Although many theories abound, no consensus has been reached on the processes that instigated the end-Permian mass extinction (Erwin, 2006). Evidence that the event had its roots in ocean chemistry is consistent: Anoxia with accompanying effects of hypercapnia, ocean acidication, and sulde toxicity could all have contributed to the loss of marine life, and, arguably, to loss of life on land (Holser, 1977; Knoll et al., 1996; Twitchett et al., 2001; Wignall and Hallam, 1992). Numerous researchers also point to volatile release during emplacement of the Siberian Traps lavas and intrusions (Payne and Clapham, 2012). These authors argued that gases liberated during the volcanism itself, augmented by additional C and S volatiles released from affected sediments, can account for most of the end-Permian paleontological and geochemical observations, and this is consistent with

spe500-08 1st pgs Five decades of advances in geochemistry the timing and tempo indicated by the most recent geochronologic studies (Kamo et al., 2003; Mundil et al., 2010; Shen et al., 2011). Despite continued debate on the ultimate trigger, most authors agree that the geochemical and geologic evidence for a bolide impact (Becker et al., 2001) is unclear (Farley and Mukhopadhyay, 2001). In this context, it is also noteworthy that the organic geochemical signals that accompany the CretaceousPaleogene impact and mass extinction event bear no resemblance to those seen at OAEs (Sepulveda et al., 2009). There are other examples where biomarker research in the last 20 yr has helped us to understand oceanic anoxia. The Chlorobi carotenoid signal is one of several organic geochemical features that are consistently observed during intervals of enhanced black shale deposition during the Paleozoic and Mesozoic Eras (Koopmans et al., 1996; Pancost et al., 2004). Isorenieratane, chlorobactane, and the aryl isoprenoids derived from them, along with C, S, and N isotope and other biomarker anomalies that are indicative of anoxia, euxinia, and water-column stratication, comprise a recurring signal that is seen in both clastic and carbonate sediments deposited during Mesozoic OAEs (Kuypers et al., 2004; Schouten et al., 2000b). Many of these sediments, such as the Kimmeridge Clay Formation, are also prolic source rocks for petroleum deposits that carry the same signal (Van Kaam-Peters et al., 1998a, 1998b). The prevalence of organicrich and petroleum-prone sediments that were deposited during the Mesozoic Era comprises the bulk of the worlds oil-in-place inventory, reecting pervasive development of narrow rift basins and seaways, greenhouse conditions, and sluggish ocean circulation during breakup of the Pangean supercontinent (Klemme and Ulmishek, 1991). Studies of molecular fossils in Precambrian sedimentary rocks have been pursued in earnest since the 1980s, and this work has shown that such rocks exhibit an array of novel molecular and isotopic features that are not seen in younger sedimentary sequences (Grantham and Wakeeld, 1988; Grosjean et al., 2009; Klomp, 1986; Summons et al., 1988). Unprecedented distributions of steroids, for example, can be traced to the proliferation of green and red algae, as well as sponges, and this accords with other geochemical evidence for the Neoproterozoic oxygenation of the atmosphere and ocean (Caneld et al., 2007; Fike et al., 2006) and paleontological evidence of the proliferation of complex life (Erwin et al., 2011; Knoll et al., 2004). Euxinic conditions have been inferred from geochemical studies of Mesoproterozoic strata, and this is reected in carotenoid molecular fossil distributions in the same rocks (Brocks et al., 2005). Steroid and triterpenoid hydrocarbons have also been recovered from much older rocks (Dutkiewicz et al., 2006; George et al., 2008; Waldbauer et al., 2009). Some of this work has been challenged on the basis of C isotope data obtained by nano-SIMS (Rasmussen et al., 2008), suggesting that the recovered biomarkers instead reect contamination. In contrast, studies of fresh drill cores recovered from the Kaapvaal craton that were targeted to address this issue (Waldbauer et al., 2009) suggest that complex terpenoids, and particularly the steroids, could be indigenous to the samples

page 21 21

and represent compounds that were biosynthesized prior to the Great Oxidation Event (GOE) at ca. 2.32.4 Ga. The presence of steroids in older than 2.5 Ga rocks supports a raft of isotopic and trace-element proxies, some of which have been discussed already, which indicate that trace amounts of oxygen were being produced (e.g., Anbar et al., 2007; Czaja et al., 2012) and respired (Eigenbrode and Freeman, 2006) in the surface oceans well in advance of the GOE. Further drilling, using ultraclean drilling protocols and contamination tracers, is currently under way in the Pilbara craton, Australia, and this is expected to shed further light on ocean-atmosphere redox conditions in the Neoarchean. Elucidating Paleoenvironmental Records One of the most important discoveries in organic geochemistry in the last 50 yr has been the recognition of some classes of organic molecules that encode signals for past climate regimes. Of key importance are the proxies for sea-surface temperature (SST) based on the degree of unsaturation in long-chain ketones (C37 alkenones) from marine algae (the Uk37 index; Brassell et al., 1986), and the proportions of isoprenoidal ether lipids (dubbed glycerol dibiphytanyl glycerol tetraethers or GDGT) derived from archaea (the Tex86 index; Schouten et al., 2002). In soils and paleosols, bacterial (non-isoprenoidal) ether lipids have been proposed to provide information on both temperature and pH (Weijers et al., 2006). Signicant effort has been expended in elucidating the specic organisms responsible for these proxies, as well as the physiological basis of the sedimentary temperature records that are encoded by alkenones and ether lipids. In the case of the former, prominent marine haptophyte algae such as Isochrysis galbana, Emiliania huxleyi, and Gephyrocapsa oceanica, living primarily in the surface mixed layer and sometimes extending to the thermocline (Ohkouchi et al., 1999), are widely recognized as the primary biological sources in the oceans (Volkman et al., 1995, 1980). Given the biogeographical trends and temperature preferences of alkenone-producing algae, it appears likely that the alkenones record lipids that were derived from a combination of cold- and warm-water adapted strains. Nutrient status, salinity, growth rate, growth stage, and temperature all inuence the distributions of alkenones in cultured haptophytes, yet empirical calibrations of the Uk37 indices to core tops, and to cultures, appear to be very robust over a wide range of temperatures, environmental variables, and spatial scales (Prahl et al., 2003; Rosell-Mel et al., 1995; Sikes and Volkman, 1993). The C37 alkenones occur in sediments that predate the appearance of contemporary alkenone-producing haptophytes, and, because of their apparently close evolutionary relations and similarity to fossil forms, this has potentially opened application of the proxy to ancient sedimentary records on multi-million-year time scales. Paleotemperature reconstructions into deep time, however, will have to be viewed cautiously until more is known about the physiological function of C37 alkenones in haptophyte algae. What is most impressive about the alkenone SST records is the degree of concordance with independent isotopic records, including

spe500-08 1st pgs 22 Johnson et al.

page 22

O isotope variations recorded in ice cores and foraminifera (e.g., Herbert et al., 2001). It has become clear that multiple approaches are needed to understand and constrain each proxy in terms of the robustness of paleotemperature records and the effects of lateral sediment redistribution (Mollenhauer et al., 2011). In contrast to the alkenones and their sedimentary records as expressed in Uk37 data, the temperature proxy based on archaeal ether lipids (Tex86) is less well constrained with respect to the precise sources of the isoprenoidal GDGTs and the controls on their production. The Tex86 proxy emerged from the discovery of abundant archaeal plankton communities that populate oceans, lakes, and other aquatic environments, in addition to recognition that membrane-spanning GDGTs contain variable numbers of cyclopentane rings that could be isolated from particulate organic matter and the underlying sediments (Delong et al., 1998). A particular problem, however, is that diverse Crenarchaeota and Euryarchaeota that inhabit both the water column and the sediments produce some of the same compounds that comprise the proxy. In addition, other studies show that the prevalence, and production, of GDGT often takes place at great depth (Ingalls et al., 2006; Sinninghe Damst et al., 2002), frequently near or within oxygen minimum zones. Nevertheless, evidence suggests that the distributions of GDGT in core-top sediments are similar to those of (epipelagic) marine Crenarchaeota living in the top 100 m of the water column, an observation consistent with the relatively robust core-top calibrations of SST with Tex86 (Wuchter et al., 2005). Numerous studies point to an origin for archaeal ether lipids in ammonia-oxidizing Marine Group 1 Crenarchaeota (Schouten et al., 2008). Evidence has been presented to show that Tex86 values agree with foraminifer proxies that indicate rapid ocean warming during the late Paleocene thermal maximum (Zachos et al., 2006). Paleotemperature records extending into Mesozoic time are considered possible (Schouten et al., 2003), but there are few avenues available for independent verication. An important consequence that ows from the discovery of isoprenoidal GDGT has been the realization that the Crenarchaeota are exceptionally important components of the marine carbon and nitrogen cycles (Delong, 2009; Ingalls et al., 2006). Using Stable Isotopes to Better Understand the Origins of Biomarkers One of the continuing challenges of organic geochemistry has been to determine the sources of individual biomarker molecules, because few compounds are thought to exclusively derive from a single organism or represent a single biogeochemical process. Sterols, for example, are made or used by almost all eukaryotes, from microbes to mammals. Despite their varied chemical structures, sterols are not as diverse as the taxonomic distributions that they represent. One way to constrain the origins of molecules is through their stable isotope compositions. Initial work used bulk measurements (e.g., Hare et al., 1991), but a desire for improved specicity and precision drove development of tools that allow isotopic measurements at the molecular level using

continuous-ow approaches that were analogous to GC-MS and LC-MS (liquid chromatographymass spectrometry; see also previous discussion). Carbon was the rst and most obvious element targeted for compound-specic isotope analysis (CSIA), and the initial results immediately revealed the diverse origins of sedimentary organic molecules (Freeman et al., 1990; Hayes et al., 1990). Analogous methods for H (Sessions et al., 1999), N (Macko et al., 1997), and S (Amrani et al., 2009) soon followed. It is now common to measure multiple isotopic compositions of individual compounds such as chlorophylls, as well as the geoporphyrins and maleimides that are derived from them (Chikaraishi et al., 2005). In turn, the technical capability to make precise isotopic measurements on these elements exposed the dearth of knowledge about the ways in which these biosynthetic processes, organismic physiology, and environmental parameters inuence the isotopic abundances of individual molecules (Hayes, 2001; Laws et al., 1997, 1995; Popp et al., 1998; Sessions et al., 1999; Smith and Freeman, 2006). The impact of developing and perfecting tools and techniques for compound-specic isotope analyses has been profound. The protocols and logic are now routinely used in archaeology, paleoclimatology, paleohydrology ecology, and forensics, and geochemists have been the instigators of many of these applications (Lichtfouse, 2000). Radiocarbon measurements of individual organic molecules have also led to new insights into the age structure of fossil lipid assemblages, and their transport from source to sedimentary sinks (Pearson et al., 2001). Thus, in contemporary environmental settings, it is possible to discern the age relations of multiple paleoclimate proxies (Mollenhauer et al., 2003), thereby resolving processes such as winnowing and lateral advection of organic matter (Kusch et al., 2010; Mollenhauer et al., 2011). The robustness, scope, and delity of organic SST proxies based on algal alkenones and archaeal ether lipids have been signicantly increased through knowledge gained from compound-specic radiocarbon measurements (Eglinton and Eglinton, 2008). Which Organisms Produce All These Sedimentary Lipids? As interest turned to understanding the origins and diagenetic pathways by which organic molecules became fossilized, there was a great expansion in our knowledge of the diversity of lipids in sediments before, in fact, their biological precursors were identied in living organisms (Ourisson et al., 1979; Rohmer, 2010; Rohmer et al., 1980). The emergence of another analytical tool, high-performance liquid chromatography coupled to mass spectrometry (HPLC-MS), enabled identication of increasingly large and complex polar lipid molecules such that, today, geochemists have focused on tracking the distributions of membrane-spanning lipids (Schouten et al., 2000a), intact polar lipids (Sturt et al., 2004), and complex biohopanoids (Talbot et al., 2008) from organisms through the water column and into sediments. Intact polar lipids can be proxies for living organisms and, often in combination with deoxyribonucleic acid (DNA), have been used to map out biogeochemical processes

spe500-08 1st pgs Five decades of advances in geochemistry in sediments (e.g., methanotrophy; Orphan et al., 2002) and the water column (e.g., Sinninghe Damst et al., 2005), as well as to help dene the extent of the deep subsurface biosphere (Lipp et al., 2008; Rtters et al., 2002). Advances in identication of the diverse types of organic molecules in rocks have been concomitant with efforts to increase our understanding of the ways in which organisms produce these lipids (biosynthetic pathways), the evolution of these pathways, and the array of functions performed by lipid molecules in cells (physiological roles). This is one of the current frontiers of organic geochemistry, because, very often, the taxonomic relations are often ambiguous between molecules of interest in the environment or rock record and their precursor organisms. Based on knowledge derived from cultured taxa, we know that hopanoids, for example, are distributed very unevenly across the bacterial domain, but the reasons for this remain mysterious (Rohmer et al., 1984). Precisely how key compounds are produced is only now coming to light, and efforts to deduce this have led to the discovery of a completely unknown isoprenoid biosynthesis pathway in bacteria (Rohmer, 2003; Rohmer et al., 1993); such ndings can have societal implications because they may lead to new classes of antibiotics. The revolution in genomics has, therefore, given geochemists a new tool for identication of important biosynthetic genes and gene families, thereby enabling culture-independent approaches to be more nuanced in providing an understanding of the sources and function of biomarker lipids. By way of example, we know that pentacyclic hopanoids and tetracyclic steroids are biosynthetically and evolutionarily related (Rohmer, 2010; Summons et al., 2006) because they are derived from the common acyclic precursor squalene. Genomic databases can be queried for both hopanoid and steroid cyclase genes, the sequences of which carry information on their evolutionary heritages (Fischer and Pearson, 2007). No longer are biosynthetic capabilities solely based on studies of cultured organisms because their genomes and natural environmental samples can now be queried directly for the biosynthetic capabilities of community members (Fischer et al., 2005; Pearson et al., 2009). The unexpected occurrence of 2-methylhopanoids in cultures of a purple nonsulfur bacterium has led to a raft of new knowledge about the genes responsible for hopanoid biosynthesis (Welander et al., 2012) and the capacity to construct mutants lacking genes of interest (Welander et al., 2010), pointing to the involvement of these complex molecules in ameliorating stress (Doughty et al., 2009; Welander et al., 2009). In summary, molecular paleontology that is informed by individual and community genomes is providing signicant insights into the sources of organic molecules that may be preserved in the rock record, which in turn provides fundamental constraints on the evolution of the biosynthetic pathways used by life. GEOCHEMISTRY ON OTHER WORLDS The last 50 yr have been a golden age for exploration of our solar system, and geochemistry has played a dominant role.

page 23 23

Meteorites were initially the only extraterrestrial samples available for geochemical analysis, and these have now been supplemented by samples returned from the Moon, a comet, an asteroid, and the solar wind. In addition, new kinds of meteorites, including samples from Mars and the Moon, have been recognized in the last ve decades. Signicant strides have also been made in geochemical analyses by remote sensing. Extinct Radionuclides and Early Solar System Chronology Short-lived radionuclides that were present in the solar nebula have decayed away, but their presence can be inferred from their radiogenic daughter isotopes. The rst extinct radionuclide found in meteorites, 129I, was discovered in 1961, and others followed in the next 50 yr. Additional extinct nuclides are now recognized to have been present in the solar nebula, including 10 Be, 26Al, 41Ca, 53Mn, 60Fe, 107Pd, 146Sm, and 182Hf. In some cases, extinct radionuclides can provide high-resolution chronometers for early solar system events (Nyquist et al., 2009). The 146Sm142 Nd system has already been discussed here in regard to early terrestrial rocks. One of the most useful extinct nuclides is 26Al, which decays 26 to Mg with a half-life of ~730,000 yr. Its former existence was rst proven when Lee et al. (1977) analyzed plagioclase and other minerals in a refractory Ca-Al inclusion (CAI) from the Allende meteorite, and they determined that plagioclase had a large 26Mg excess (relative to nonradiogenic 24Mg). Because CAIs are the oldest solar system materials, based on their 207Pb-206Pb ages, their 26Al contents are taken to reect those at the beginning of the solar system, and other measurements are compared to the 26 Al/27Al ratio and 207Pb-206Pb age of CAIs for chronology. The 26 Al value reects production by stellar nucleosynthesis, and its occurrence in the solar nebula is commonly attributed to seeding by a nearby supernova. Renements in SIMS and ICP-MS techniques have permitted measurements of 26Al-26Mg systematics in samples that have low Al contents, including chondrules and differentiated meteorites. An example is illustrated in Figure 10, which shows different 26Al/27Al ratios for chondrules in several classes of chondrites, and this implies that chondrules formed several million years after CAIs. The 182Hf-182W system, with a half-life of 8.9 m.y., has also proved to be very useful in dating early solar system events. Early measurements by N-TIMS (negative thermal ionization mass spectrometry) were difcult, but this chronometer was made truly accessible when MC-ICP-MS analysis became available (Halliday et al., 1996). Both Hf and W are refractory elements and so usually occur in chondritic proportions. Because, however, Hf is lithophile and W is siderophile, the 182Hf-182W system can date metal-silicate fractionation events such as core formation. Such ages, in turn, constrain the time of planet accretion. For example, the estimated age of Earths accretion (50% mass), based on this chronometer, is 30 to >100 m.y. after CAIs (Halliday, 2004). More rapid accretion ages for asteroidal bodies were determined by Kleine et al. (2009).

spe500-08 1st pgs 24 Johnson et al. Nebular Condensation

page 24

Figure 10. Ion microprobe measurements of initial 26Al/27Al, and corresponding ages relative to Ca-Al inclusion (CAI), in chondrules from primitive meteorites (unequilibrated ordinary chondrites and several classes of carbonaceous chondrites). Figure is from McSween and Huss (2010; used with permission). UOC unequilibrated ordinary chondrite.

Mass-Independent Isotope Fractionation and Early Solar System Conditions One of the rst mass-independent fractionations (MIF) for stable isotopes was observed for O measured in CAIs in meteorites (Clayton et al., 1973), as illustrated in Figure 11. For 17 O/16O-18O/16O variations, mass-dependent fractionation denes a line with a slope of ~0.5. Subsequent analyses of O isotopes in bulk meteorites showed that igneous meteorites (achondrites) dened distinct fractionation lines parallel to that of Earth. Bulk chondrites of different classes also plot in distinct areas of the diagram. As a consequence, O isotopes have become a very useful criterion for meteorite classication, and they place strong constraints on processes that operated early in the history of the solar system (Clayton, 2004). The MIF trend for CAIs in Figure 11 was originally interpreted as a mixing line between solar system gas (plotting on or above the terrestrial fractionation line) and exotic grains that contained pure 16O (plotting on an extension of the CAI line). A newer explanation is that the isotopic variations in CAIs arose from self-shielding during photodissociation of CO, a major nebular gas (Clayton, 2002). Because 16O is abundant, there were large differences between the amounts of C16O, on the one hand, and C17O and C18O on the other. The abundant C16O thus became optically thick far from the Sun; C17O and C18O remained optically thin, and so were dissociated to a greater extent by radiation. This produced an inner zone in the nebula that was enriched in 17 O and 18O, which reacted with H to make H2O. The 16O-depleted water then reacted with dust or condensates to form nebular solids. In this model, the Sun, representing the nebular gas, must be 16 O-rich. The rst measurement of O isotopes in the solar wind returned to Earth by the Genesis mission (McKeegan et al., 2011) is consistent with this model.

Many, although not all, of the solids that now comprise solar system bodies once condensed from nebular vapor. Early attempts to determine the condensation sequence in a cooling gas of solar composition involved simple calculations, performed long before digital computers were available. The equilibrium condensation behavior of elements in a nebular gas was rigorously modeled by Grossman and Larimer (1974) and other workers since then (e.g., Petaev and Wood, 1998). The canonical condensation sequence, showing condensing phases and the fraction of each element condensed at various temperatures, is illustrated in Figure 12. Based on experimental determinations of entropy, enthalpy, and heat capacity, equations of state describing the thermodynamic stabilities of a host of possible minerals under various conditions can be calculated. Liquids are not stable at the low pressures appropriate for the solar nebula. Some minerals in the condensation sequence do not condense directly, but form or adjust their compositions by reactions of previously condensed phases with the gas. This condensation calculation was done for the 23 elements with the highest cosmic abundances. Generally, thermodynamic data for trace elements are lacking, so chemical analyses of trace elements in high-temperature minerals of chondritic meteorites guide our understanding of their condensation behavior. Accordingly, in these models trace elements are allowed to condense as simple metals, oxides, or suldes, which are assumed to dissolve into appropriate major mineral phases. The validity of condensation calculations has been supported by studies of refractory inclusions (CAIs) in chondrites, which have bulk chemical compositions consistent with those calculated for the rst 5% of condensable matter (Davis and Richter, 2004). Moreover, the same minerals that comprise these inclusions are predicted to have been the earliest condensed phases (Fig. 12). There remains some controversy, however, about whether the CAIs are actually condensates or refractory residues from evaporation (the reverse of condensation). Condensation calculations have also been done at different nebular pressures and using nonsolar gas compositions (Ebel and Grossman, 2000; Wood and Hashimoto, 1993; Yoneda and Grossman, 1995). Parts of the solar nebula may have been enriched in dust, which, when vaporized, could have yielded nonsolar vapor, possibly allowing more reduced phases or even liquids to condense. These condensation calculations could also apply to other stars, e.g., red giants, which can have nonsolar compositions. Organic Matter from Space The molecular and isotopic chemistry of extraterrestrial organic matter is another area where technical innovation has spurred improved appreciation for the nature and diversity of organic compounds formed in the interstellar medium and the processes that alter them prior to their delivery to Earth in meteorites (Derenne and Robert, 2010) and interstellar dust particles (Flynn et al., 2004). The low-molecular-weight compounds

spe500-08 1st pgs Five decades of advances in geochemistry

page 25 25

tein amino acids (Cronin and Pizzarello, 1997). Moreover, the l-enantiomeric excess of some compounds such as isovaline in the Murchison and Orgueil meteorites appears to be related to the extent of aqueous processing, suggesting that it could reect amplication of a small initial isovaline asymmetry. If correct, this would be inconsistent with the theory that ultraviolet (UV) circularly polarized light was the primary source of l-enrichment in amino acids. It seems possible, therefore, that early life on Earth had access to molecular building blocks with the left handedness that characterizes the amino acids of all life today (Engel and Macko, 2001; Glavin and Dworkin, 2009). Such results demonstrate the importance of off-world geochemistry to informing us about terrestrial evolution.
Figure 11. Oxygen isotopes (relative to SMOW) measured in Ca-Al inclusion (CAI) from a carbonaceous chondrite dene a mass-independent fractionation trend, distinct from the terrestrial mass-dependent trend. Figure is modied from Clayton et al. (1973). SMOWstandard mean ocean water.

Lunar Geochemistry The Moon is a geochemical experiment conducted under vastly different conditions than Earth. The geochemical exploration of the Moon began with the return of samples by Apollo astronauts in 1969, which extended over an exciting 4 yr period. Depletions in volatile elements and enrichments in refractory elements measured in these rocks support the giant impact hypothesis for the Moons formation (Taylor et al., 2006). Vaporization of portions of the target (Earth) and the impactor, followed by incomplete condensation in Earth orbit, can account for the fractionation observed in volatile and refractory elements. Measured depletions of siderophile elements in lunar rocks are consistent with models that indicate preferential incorporation of the silicate mantles of the impactor and target into the Moon, accompanied by accretion of the impactors core into Earths core. The revelation that a magma ocean once existed on the Moon (Wood et al., 1970) has profoundly changed planetary science, and now wholesale melting has been proposed for a number of bodies, including early Earth. REE analyses came of age during

identied include many classes of biologically important molecules, for example, amino acids (Kvenvolden et al., 1970), hydroxyacids and dicarboxylic acids (Cronin et al., 1993; Lawless et al., 1974; Peltzer et al., 1984), as well as nucleobases (Martins et al., 2008). Isotopic data and pyrolysis studies suggest that distinct processes are involved in formation of small molecules and macromolecular material (Sephton et al., 2004; Yuen et al., 1984). Perhaps the most profound, enigmatic, and controversial nding was that some amino acids were not racemic, as had long been considered fact (Engel and Macko, 1997; Engel et al., 1990; Engel and Nagy, 1982). This discovery, originally dismissed as due to terrestrial contamination, became accepted once it was demonstrated also to be a feature of some nonpro-

Figure 12. A model of the condensation sequence for a cooling gas of solar composition at 104 atm pressure. Condensed minerals are labeled in italics, and curves show the fraction of each element condensed as a function of temperature. REErare earth element. Figure is modied from Grossman and Larimer (1974).

spe500-08 1st pgs 26 Johnson et al.

page 26

the Apollo program, as discussed earlier, and the complementary REE patterns of anorthosite from the highlands crust and basalts from the maria (Fig. 13) provide the most persuasive evidence for a lunar magma ocean (Taylor and Jakes, 1974). In this model, the thick anorthositic crust was formed by otation of plagioclase, which produced a positive Eu anomaly in the crust, and this was balanced by the Eu depletion that was produced by olivine and pyroxene that accumulated to form the mantle, a characteristic that was inherited by later mare basalt melts. The Eu anomaly in lunar rocks was enhanced by the Moons low oxidation state, which increased the proportion of Eu2+, allowing it to partition into plagioclase. The last dregs of the magma ocean were rich in Fe, Ti, and incompatible trace elements, including K, REEs, and P (the KREEP component), which became sandwiched between the crystallized crust and mantle. Gravitational overturn of this dense, Fe- and Ti-rich layer resulted in mixing of KREEP into magmas that subsequently intruded the anorthositic crust. The discovery of KREEP was initially made through geochemical analyses of Apollo samples (Taylor et al., 2001). Geochemistry by Remote Sensing Geochemical analyses no longer are restricted to the laboratory. Here, we consider two examples of geochemistry by spacecraft that have altered our perceptions of what is possible using remote sensing. The Gamma Ray Spectrometer on the Lunar Prospector orbiter analyzed the abundances of a handful of elements (Fe, Ti, K, Mg, Al, Ca, Si, Th, and H; Lawrence et al., 1998; Prettyman et al., 2006), utilizing characteristic gamma-ray emissions produced by radioactive decay or by reactions initiated by cos-

mic rays. Orbital measurements provide global coverage and thus are particularly useful in understanding geochemical processes at a planetary scale. Global maps of the distribution of Fe and Th, which are particularly sensitive to this technique, have been employed to distinguish lunar terranes based on their geochemical characteristics (Jolliff et al., 2000). The petrology of each terrane has been interpreted unambiguously by comparison with the Fe and Th contents of returned lunar samples. Other results, made possible by measuring the neutron ux from the surface, included the discovery of H at the lunar poles associated with cold, permanently shadowed craters believed to contain water ice (Feldman et al., 2001). The Mars Exploration Rovers Opportunity and Spirit carried Alpha Particle X-Ray Spectrometers (APXS) that analyzed the chemical compositions of hundreds of rocks and soils on the surface of Mars (Brckner et al., 2008; Gellert et al., 2006). By measuring characteristic X-rays produced by alpha particles and X-rays emitted from a radioactive source, the APXS was able to analyze all of the major elements and a few minor and trace elements. These studies were a critical part of the classication of rocks encountered during rover traverses, and they provided constraints on the processes that formed these materials. Opportunity analyzed sedimentary rocks that contained high contents of S, Cl, and Br, interpreted to reect evaporation of salt-laden brines, demonstrating that liquid water was once abundant on Mars (Squyres et al., 2004). On the other side of the planet, Spirit analyzed a variety of ancient basaltic rocks (Fig. 14) that have compositions that are distinct from those of younger Martian basaltic meteorites, buttressing the argument that although heterogeneous, Mars is fundamentally a basalt-covered world (McSween et al., 2009). Stardust in the Laboratory Tiny motes of stardust, condensates that formed around dying stars or farther out in the interstellar medium, were discovered in chondritic meteorites (Lewis et al., 1987), after a long search beginning in the 1960s. These diamond nanoparticles were isolated by dissolving chondrites in a series of harsh acids, and at each step tracking the isotopically anomalous Xe they contained. The grains are thought to have been implanted when the outer layers of giant C-rich stars were sloughed off and condensed as diamond, and were subsequently implanted with distinctive nuclides produced when the star exploded as a supernova. Approximately 20 types of presolar grains, including silicon carbide (Tang and Anders, 1988) and graphite (Amari et al., 1990), as well as oxides, nitrides, and silicates, have now been found. All are distinguished by their exotic isotopic compositions (Tang and Anders, 1988). The challenging elemental and isotopic analyses of stardust grains, which range from a few nanometers to a few micrometers in size, illustrate the great progress made in micro-analytical techniques. An example, using SIMS-analyzed C and N isotopes to ngerprint the sources of presolar silicon carbide grains, is

Figure 13. Chondrite-normalized rare earth element (REE) patterns for lunar anorthosite and mare basalt. The complementary europium anomalies for crust and mantle-derived rocks support the magma ocean hypothesis.

spe500-08 1st pgs Five decades of advances in geochemistry shown in Figure 15. More importantly, the isotopic compositions of presolar grains provide ground truth for stellar nucleosynthesis models, directly linking cosmochemical measurements in the laboratory to astrophysical theory. These data also (1) provide information on capture cross sections for neutrons, the capture of which makes heavier elements, (2) demonstrate where theoretical models are inadequate to describe the internal structures of stars, (3) identify neutron sources for the s-process, and constrain the scale of mixing in supernovae (Nittler, 2003; Zinner, 2004). Presolar grains also fundamentally changed the way we think about the solar systems formation. Their widespread occurrence in primitive meteorites demonstrates that a hot nebula did not vaporize all preexisting solids, overturning a view that prevailed 50 yr ago. INTEGRATING THE PICTURE: GEOCHEMICAL CYCLES The concept of geochemical cycles dates to the late nineteenth century and was well established (i.e., began to appear in textbooks) by the mid-twentieth century (e.g., Goldschmidt, 1954). The modern phase of quantifying global geochemical cycles arguably dates from the pioneering work of Garrels and Mackenzie (1971, 1972), who integrated major-element chemistry (including C, S, and Cl) into a steady-state recycling model for the evolution of the sedimentary mass. As described in greater detail in the following, geochemical cycles are often complex and inter-related, and considerable effort has been directed

page 27 27

toward trying to understand how cycles are linked. Many elemental cycles require monograph-length treatment, and the fact that there is a journal devoted solely to biogeochemical cycling (Global Biogeochemical Cycles) reects the sustained interest in geochemical cycles. In geochemical cycles, the reservoirs (mass, M) and uxes of individual elements (or groups of closely related elements) into and out of the system (mass/unit time, Fin and Fout) are quantitatively accounted for over geological time (t). Accordingly, assessments of geochemical cycles are essentially a mass balance of the element of interest on some appropriate physical scale over some appropriate duration, and to a great extent are the natural consequence of quantifying the overall rock cycle (Gregor, 1992) and hydrological cycle (Fig. 16). Geochemical cycles are considered to be either exogenic, for those operating on or near Earths surface (hydrosphere, atmosphere, biosphere sediments), typically on relatively short time scales (<~104107 yr), or endogenic, for those operating within the interior of Earth (oceanic and continental crust, mantle, core sediments), typically on relatively long time scales (>~104107 yr). The physical interface between the endogenic and exogenic parts of an elements geochemical cycle typically occurs in soils and the sedimentary cover. Characterization of complete global cycles requires integration of both endogenic and exogenic cycles as witnessed, for example, by the relatively recent recognition that the C cycle is

Figure 14. A geochemical classication diagram for volcanic rocks comparing the compositions of Martian rocks and soils derived from them in Gusev crater, analyzed by the Spirit rovers Alpha Particle X-Ray Spectrometers (APXS), with laboratory analyses of Martian meteorites. Figure is modied from McSween et al. (2009).

Figure 15. Carbon and nitrogen isotopic ratios in presolar silicon carbide grains, measured with an ion microprobe. Most grains from stars on the asymptotic giant branch (AGB) of the HurtzsprungRussell diagram plot above the solar composition (dashed lines). Supernovae grains have lower 14N/15N, and rare grains from novae, which are powered by explosive hydrogen burning, plot in the lower-left quadrant. The stellar sources of other grains are not currently known. Figure is modied from Zinner (2004).

spe500-08 1st pgs 28 Johnson et al.

page 28

affected by signicant levels of microbial activity deep within parts of Earths crust (e.g., Hazen et al., 2012), in addition to mantle sources. Depending on the way in which the physical and/or temporal scales are dened, geochemical cycles may be either open or closed, where the distinction is a function of whether or not external inputs and/or outputs take place within the dened reservoirs. Those concepts govern whether or not a geochemical cycle is in steady state (i.e., dM/dt = 0; Fin = Fout), leading to the further concept of residence time ( = M/F), which in turn is equivalent to the inverse of the rst-order rate constant ( = 1/k) for simple linear cycles, thus providing information about the response times (i.e., kinetics) of the system (Lasaga and Berner, 1998). One problem that attracted early attention was the issue of elemental cycling through the oceans. This interest was initiated largely by the pioneering work of Silln (1961), who demonstrated that seawater chemistry could be in a steady-state condition, controlled by equilibrium reactions between atmospheric gases and marine carbonate and silicate minerals. Up until the early 1950s, the conventional wisdom was that the oceans salt had accumulated over the entirety of geological time. During

the 1960s1980s, considerable effort went into determining the apparent mean oceanic residence times of the elements (), a concept rst introduced by Barth (1952), as well as documenting the balance between the masses of the various elements that entered the oceans from rivers, and those exiting the oceans through sedimentation (e.g., Drever et al., 1988). The ensuing research identied important additional sources of elements to the oceans (e.g., basalt-seawater hydrothermal interaction) that had not been previously recognized, additional reservoirs for elements (e.g., pore waters, altered basalts, estuaries, continental shelves), and numerous mineral and biogeochemical reactions, leading to a far more complete understanding of marine chemistry (e.g., Broecker and Peng, 1982). By evaluating geochemistry in the framework of geochemical cycles, the fundamental processes that inuence elemental distributions (geological, geophysical, chemical, biological, temporal) are better addressed (Lerman and Wu, 2007). A good example of the value of this approach in identifying previously unrecognized processes and reservoirs is the well-known missing sink issue for carbon (Broecker, 2012). During the 1990s, the short-term (exogenic) C cycle had been quantied sufciently

Figure 16. Representation of the pre-industrial (A.D. 1750) global carbon cycle (adapted from Sundquist and Visser, 2003). Boxes show major reservoirs, with carbon mass listed in units of Pg (1015 g), and arrows show major uxes of carbon (listed in units of Pg/yr). DICdissolved inorganic carbon.

spe500-08 1st pgs Five decades of advances in geochemistry to recognize that ~15%20% of the CO2 delivered to the atmosphere by fossil fuel combustion could not be accounted for by the recognized major C reservoirs (atmosphere, accounting for ~50%, and seawater, accounting for ~30%35%). This led to a major research effort to identify the missing sink, which in turn resulted in recognition of the terrestrial biosphere (previously considered relatively minor) as a major reservoir for exogenic C. The Biological Connection The importance of biological activity in controlling the cycling of a wide range of elements has been long appreciated (e.g., Vinogradov, 1943). Attempts, however, to both dene and quantify global biogeochemical cycles involving elements that are essential for biological activity have only occurred somewhat recently, though they are of great geochemical interest (e.g., Garrels and Lerman, 1981). Biogeochemical cycles consider some or all of a wide array of biologically relevant elements in an integrated manner, including those that are a signicant part of living tissue and skeletons (C, O, H, N, P, S, Ca, Si), those that may be less involved in living matter but are important in redox processes allowing for energy transfer (Fe, Mn, in addition to N and S), and a long list of minor and trace elements that may be necessary for metabolism (e.g., biolimiting) and/or substitute for major elements of biogeochemical importance (e.g., Fe, Mn, Mg, Ba, Ge, B, Mo, V, Zn, and even the REEs). The C cycle (and its role in biogeochemical cycles) is of intrinsic interest to geochemists due to the role of C in all biological activity and its importance in the geological record as a common rock-forming constituent (limestone, dolomite, carbonaceous sediment) and as a natural resource (fossil fuels). The fact, however, that C has received by far the most attention of any of the geochemical cycles, and indeed may be the most intensely studied geochemical problem of the past 50 yr and more (Berner, 2004; Broecker, 2012; Des Marais, 2001), is mainly due to the observation that atmospheric CO2 concentrations have systematically risen since the industrial revolution at a rate that is unprecedented for the Phanerozoic, due mainly to the burning of fossil fuels (Fig. 17). It is now widely recognized that perturbations within the C cycle (temporal variations in atmospheric CO2) represent a dominant control on the changes of Earths current climate, as well as both short-term and long-term paleoclimate. The recent abrupt increase in atmospheric levels of this important greenhouse gas, due mainly to human activity, is also now understood to play a central role in inuencing recent global climate change and increases in mean global surface temperatures (Berner, 2003). Redox Processes in Biogeochemical Cycles The recent focus on element cycling, including biogeochemical cycles, is primarily driven by redox chemistry, both abiologic or biologic, due to its importance in controlling the composition and evolution of the ocean-atmosphere system (e.g., Raiswell

page 29 29

and Caneld, 2012). As discussed already, C is certainly the most studied redox-sensitive element that is cycled between reduced and oxidized forms today on Earth, driven largely by photosynthesis. This in turns drives atmospheric O2 contents through the extent of C burial and thus plays a central role in the long-term O cycle. This inuences, for example, the long-term cycling of Fe, one of the most important redox elements in rocks, given its high abundance in the terrestrial planets. The redox couple with Fe was studied extensively in the 1970s, with a major focus on banded iron formations (BIFs), and later, the relation between the Fe biogeochemical and marine evolutions. Evaluation of the geochemical cycles of redox-sensitive trace elements such as Cr and Mo rose in prominence in the 1980s and 1990s. One of the major issues addressed in studies of redox biogeochemical cycling over the past 50 yr has been the history of atmospheric O2 over geological time (Holland, 1962). As briey noted already, a prominent approach to understanding the history of atmospheric O2 on Earth has been study of both massdependent and mass-independent fractionation (MIF) of stable S isotopes, and these records are compared, along with C and Fe isotopes, in Figure 18. The range in d34S values for suldes in marine sedimentary rocks generally increases with decreasing age, an observation that has also been long recognized (e.g., Caneld, 2001). The general increase in the highest d34S values measured for suldes is broadly taken to record an increase in the d34S values of seawater sulfate, and a recent compilation of this can be found in Caneld and Farquhar (2009). It is almost universally accepted that the range in d34S values for suldes

Figure 17. Monthly mean concentration of carbon dioxide, measured as parts per million molecules of CO2 in total molecules of dry air, measured at the Mauna Loa Observatory, Hawaii (data from National Oceanic and Atmospheric Administration [NOAA] Web site (http://www.esrl.noaa.gov/gmd/ccgg/trends/), the so-called Keeling curve. The annual cycle observed in the record is a seasonal effect due to the much greater land mass (and vegetation) in the Northern Hemisphere, resulting in greater CO2 uptake during plant growth in the northern summer. These relations were rst recognized by Charles Keeling (Keeling, 1960).

spe500-08 1st pgs 30 Johnson et al.

page 30

Figure 18. Temporal variations in (A) banded iron formation (BIF) deposition and atmospheric oxygen, (B) d13C values of Ca-Mg carbonates, (C) d34S values for suldes, (D) mass-independent S isotope fractionation, expressed as D33S, and (E) 56Fe values; 56Fe values are broken out as black shales (high-S, high-C contents), Ca-Mg carbonates, and BIF samples. Green band represents period of maximum BIF deposition and immediately predates the increase in atmospheric oxygen contents. BIF deposition is from Bekker et al. (2010). Atmospheric O2 curve is from Catling and Claire (2005), shifted to older ages based on disappearance of mass-independent S isotope fractionation (Farquhar et al., 2010). Early pulses of O2 time band are based on previous studies (Anbar et al., 2007; Czaja et al., 2012; Duan et al., 2010; Kendall et al., 2010; Voegelin et al., 2010). The d13C data are from Shields and Veizer (2002). The d34S data for suldes and seawater sulfate are from Caneld and Farquhar (2009), and D33S values for suldes are from same source. The 56Fe values are from sources cited in Johnson et al. (2008b), with additional data sources (Czaja et al., 2010, 2012; Heimann et al., 2010; Hofmann et al., 2009; Planavsky et al., 2009; Steinhoefel et al., 2009; Tsikos et al., 2010; Valaas-Hyslop et al., 2008; Von Blanckenburg et al., 2008). All isotopic data reect bulk sample analyses.

spe500-08 1st pgs Five decades of advances in geochemistry reects various extents of bacterial sulfate reduction (Caneld, 2001, 2005). Turning to S-MIF, nonzero D33S values for marine sedimentary rocks are restricted to samples of ca. 2450 Ma age and older, and most studies interpret this to indicate very low atmospheric oxygen contents (originally discovered by Farquhar et al. [2000a], and recently reviewed by Farquhar et al. [2010]), although alternative explanations have been proposed (discussed earlier herein). The transition from large D33S values to zero D33S values at ca. 2450 Ma correlates with an increase in the range of d34S values, consistent with an increase in seawater sulfate contents and development of free oxygen in the atmosphere, which in turn would enhance rates of bacterial sulfate reduction (Caneld, 2005; Farquhar et al., 2011). As discussed earlier, the d13C values for Ca-Mg carbonates of Archean and Proterozoic age largely scatter closely about zero, with the exception of the 2.32.0 Ga Lomagundi excursion (Fig. 18). Increased organic C burial seems the most likely explanation for the increase in d13C values for carbonates at this time, which in turn would drive further increase in atmospheric O2. This would tend to increase seawater sulfate contents, providing opportunities to increase the 34S/32S fractionations produced by microbial sulfate reduction due to excess sulfate; this would increase the inventory of sedimentary suldes that have very negative d34S values (Fig. 18). Accompanying the rise in atmospheric O2 would be a loss of S-MIF, reected in a shift toward zero D33S values for suldes younger than 2.3 or 2.4 Ga in age (Fig. 18). The largest Fe isotope excursion known in the rock record occurs in the Neoarchean and Paleoproterozoic (Fig. 18). Because the vast majority of Fe in the crust has a 56Fe value near zero, including low-C, low-S sedimentary rocks that have Fe contents similar to those of the average crust (e.g., Johnson et al., 2008b), deviations in the 56Fe values from zero are signicant and generally rare in terms of the Fe mass balance of Earth. The zero to positive 56Fe values for rocks older than 3.5 Ga in age, which to date mainly include BIFs, are generally accepted to reect partial oxidation of marine hydrothermal Fe2+aq, suggesting that the amount of oxidant was limited (Dauphas et al., 2004), and recent Fe isotope work suggests that photic zone O2 levels were <0.001% of present day at this time (Czaja et al., 2013). More controversial, however, is the strong decrease in 56Fe values between ca. 3 and 2.5 Ga. Rouxel et al. (2005) and Anbar and Rouxel (2007) proposed that oxidation of marine hydrothermal Fe2+aq during BIF genesis, or oxide precipitation on continental shelves, produced negative 56Fe values in seawater, which was directly incorporated into sulde-rich marine sedimentary rocks. As recently discussed by Czaja et al. (2012), extensive oxidation of Fe2+aq is a likely explanation for the low 56Fe values of Neoarchean Ca-Mg carbonates and, in fact, provides one of several lines of evidence for intermittent oxygenation of surface environments in the time leading up to the Great Oxidation Event (GOE) at ca. 2.3 or 2.4 Ga (Anbar et al., 2007; Kendall et al., 2010; Voegelin et al., 2010). Oxidation of Fe2+aq, however, cannot easily explain the spread in 56Fe values for Fe-rich rocks such as BIFs, and Yamaguchi et al. (2005) and Johnson et al. (2008a, 2008b) do

page 31 31

not generally interpret the Fe isotope compositions of such lithologies to be a direct proxy for seawater, instead favoring microbial Fe cycling as an explanation for the Fe isotope variability. The relative paucity of highly negative 56Fe values after the GOE (Fig. 18) has been interpreted to reect a decrease in the footprint of microbial Fe3+ reduction as microbial sulfate reduction increased in its impact on redox cycling. Sulfate reduction would produce abundant sulde, which readily reacts with Fe3+ oxides, decreasing Fe3+ availability to microbial reduction. The increase in 56Fe values that generally corresponds with a decrease in BIF deposition (Fig. 18) is consistent with a transition from an Fe-dominated redox cycle before the GOE to a S-dominated redox cycle after the GOE. Collectively, the temporal record of BIF deposition and C, S, and Fe isotopes displays remarkable coherence that can be well explained by expected changes in microbial redox cycling during the transition from an anoxic to oxic world. We close by considering redox cycling on another planetary body. The recent phase of Mars exploration has led to recognition of a long-lived dynamic sedimentary cycle on that planet. This has generated interest in characterizing geochemical cycles and evaluating the implications for surface processes (e.g., Grotzinger et al., 2011), an interest that undoubtedly will increase with the successful landing of the Mars Science Laboratory. Mars may represent a valuable end member for understanding near-surface geochemical cycles in general due to the absence of plate tectonics. There appears to be a long-term evolution of sedimentary mineralogy, where a common occurrence of Noachian (older than 3.7 Ga) clay minerals gives way to a sulfate-rich mineralogy in the Hesperian (ca. 3.73.2 Ga), and nally to anhydrous ferric oxides in the Amazonian (younger than 3.2 Ga; Bibring et al., 2006). Recent work on the ALH84001 meteorite has provided support for clay minerals on the surface of Mars prior to 4.0 Ga (Beard et al., 2013). As on Earth, where sedimentary mineralogy (e.g., BIF) provides a proxy for geochemical evolution (Hazen et al., 2008), this secular change in sedimentary mineralogy has been interpreted to represent acidication, oxidation, and desiccation of the Martian surface over geological time. Although the current atmosphere is very thin, most models call for a more substantial early Martian atmosphere that was dominated by CO2. Sulfur, however, is also enriched in Mars surcial deposits, and it now appears that some form of a S-cycle may have strongly inuenced surcial processes over much of geological time, leading to widespread low pH conditions (Halevy et al., 2007; King and McLennan, 2010; McLennan, 2012). CONCLUDING REMARKS We are not so bold as to predict what advances will occur in geochemistry in the next 50 yr, as we are quite sure that many of the analytical and computational gains that have been key factors to a number of geochemical advances were not envisioned by geochemists in the early 1960s. It was also probably not clear to these geochemists what the impact of plate tectonics would

spe500-08 1st pgs 32 Johnson et al.

page 32

be, nor of obtaining samples from the Moon, or of in situ analyses of rocks and soils on Mars. In a shorter time frame than ve decades, however, it seems likely that the trend toward interdisciplinary research, where multiple large data sets are brought to bear on geologic problems, is likely to expand. In addition, there will likely be huge gains in computational and data storage capacity, although analytical advances may be more muted, given the fact that many analyses are now limited by counting statistics rather than electronics, although there is certainly room for large improvements in terms of sampling efciency of many instruments (what gets to the detector versus what is consumed). A major unknown for the future of geochemistry, and of science in general, is the commitment to research by broader society, and such issues will undoubtedly affect what is written in a similar paper in the early 2060s. ACKNOWLEDGMENTS We thank the organizing committee of the 125th anniversary celebration of the Geological Society of America for the opportunity to write this summary of progress in geochemistry. Johnson and Summons acknowledge support from the National Aeronautics and Space Administration (NASA) Astrobiology Institute during the time this review was prepared. McLennan acknowledges support from NASA. McSween acknowledges support from NASAs Cosmochemistry Program. We thank Richard Carlson, Robert Clayton, John Hayes, Scott Samson, John Valley, and Richard Walker for helpful reviews of the manuscript. REFERENCES CITED
Abouchami, W., and Goldstein, S.L., 1995, A lead isotopic study of circumAntarctic manganese nodules: Geochimica et Cosmochimica Acta, v. 59, no. 9, p. 18091820, doi:10.1016/0016-7037(95)00084-D. Adam, P., Schmid, J.C., Mycke, B., Strazielle, C., Connan, J., Huc, A., Riva, A., and Albrecht, P., 1993, Structural investigation of nonpolar sulfur crosslinked macromolecules in petroleum: Geochimica et Cosmochimica Acta, v. 57, no. 14, p. 33953419, doi:10.1016/0016-7037(93)90547-A. Albarde, F., 2004, The stable isotope geochemistry of copper and zinc: Reviews in Mineralogy and Geochemistry, v. 55, p. 409427, doi:10.2138/ gsrmg.55.1.409. Alizai, A., Clift, P.D., Giosan, L., Van Laningham, S., Hinton, R., Tabrez, A.R., and Danish, M., 2011, Pb isotopic variability in the modernPleistocene Indus River system measured by ion microprobe in detrital K-feldspar grains: Geochimica et Cosmochimica Acta, v. 75, p. 47714795, doi:10.1016/j.gca.2011.05.039. Allgre, C.J., 1968, Comportement des systemes U-Th-Pb dans le manteau superieur et modele devolution de ce dernier au cours des temps geologiques: Earth and Planetary Science Letters, v. 5, p. 261269, doi:10.1016/ S0012-821X(68)80050-0. Allgre, C.J., and Luck, J.-M., 1980, Osmium isotopes as petrogenetic and geological tracers: Earth and Planetary Science Letters, v. 48, p. 148154, doi:10.1016/0012-821X(80)90177-6. Allgre, C.J., and Minster, J.F., 1978, Quantitative models of trace element behavior in magmatic processes: Earth and Planetary Science Letters, v. 38, p. 125, doi:10.1016/0012-821X(78)90123-1. Allgre, C.J., and Rousseau, D., 1984, The growth of the continent through geological time studied by Nd isotope analysis of shales: Earth and Planetary Science Letters, v. 67, p. 1934, doi:10.1016/0012-821X(84)90035-9. Alvarez, L.W., Alvarez, W., Asaro, F., and Michel, H.V., 1980, Extraterrestrial cause for the Cretaceous-Tertiary extinction: Science, v. 208, p. 1095 1108, doi:10.1126/science.208.4448.1095.

Alves, S., Schiano, P., and Allgre, C.J., 1999, Rhenium-osmium isotopic investigation of Java subduction zone lavas: Earth and Planetary Science Letters, v. 168, p. 6577, doi:10.1016/S0012-821X(99)00050-3. Alves, S., Schiano, P., Capmas, F.O., and Allgre, C.J., 2002, Osmium isotope binary mixing arrays in arc volcanism: Earth and Planetary Science Letters, v. 198, p. 355369, doi:10.1016/S0012-821X(02)00524-1. Amari, S., Anders, E., Virag, A., and Zinner, E., 1990, Interstellar graphite in meteorites: Nature, v. 345, p. 238240, doi:10.1038/345238a0. Amelin, Y., Lee, D.C., Halliday, A.N., and Pidgeon, R.T., 1999, Nature of the Earths earliest crust from hafnium isotopes in single detrital zircons: Nature, v. 399, p. 252255, doi:10.1038/20426. Amelin, Y., Lee, D.C., and Halliday, A.N., 2000, Early-Middle Archaean crustal evolution deduced from Lu-Hf and U-Pb isotopic studies of single zircon grains: Geochimica et Cosmochimica Acta, v. 64, p. 42054225, doi:10.1016/S0016-7037(00)00493-2. Amrani, A., Sessions, A.L., and Adkins, J.F., 2009, Compound-specic 34S analysis of volatile organics by coupled GC/multicollector-ICPMS: Analytical Chemistry, v. 81, p. 90279034, doi:10.1021/ac9016538. Anbar, A.D., 2004, Molybdenum stable isotopes: Observations, interpretations and directions: Reviews in Mineralogy and Geochemistry, v. 55, p. 429 454, doi:10.2138/gsrmg.55.1.429. Anbar, A.D., and Rouxel, O., 2007, Metal stable isotopes in paleoceanography: Annual Review of Earth and Planetary Sciences, v. 35, p. 717746, doi:10.1146/annurev.earth.34.031405.125029. Anbar, A.D., Duan, Y., Lyons, T.W., Arnold, G.L., Kendall, B., Creaser, R.A., Kaufman, A.J., Gordon, G.W., Scott, C., Garvin, J., and Buick, R., 2007, A whiff of oxygen before the Great Oxidation Event?: Science, v. 317, p. 19031906, doi:10.1126/science.1140325. Andersen, D.J., Lindlsey, D.H., and Davidson, P.M., 1993, QUILFA Pascal program to assess equilibria among Fe-Mg-Mn-Ti oxides, pyroxenes, olivine, and quartz: Computers & Geosciences, v. 19, p. 13331350, doi:10.1016/0098-3004(93)90033-2. Archer, C., and Vance, D., 2006, Coupled Fe and S isotope evidence for Archean microbial Fe(III) and sulfate reduction: Geology, v. 34, p. 153 156, doi:10.1130/G22067.1. Armstrong, R.L., 1968, A model for Sr and Pb isotopic evolution in a dynamic Earth: Reviews of Geophysics, v. 6, p. 175199, doi:10.1029/ RG006i002p00175. Armstrong, R.L., 1981, Radiogenic isotopes: The case for crustal recycling on a steady-state no-continental-growth Earth: Philosophical Transactions of the Royal Society, ser. A, v. 301, p. 443472, doi:10.1098/rsta.1981.0122. Barth, T.W., 1952, Theoretical Petrology: New York, John Wiley, 387 p. Beard, B.L., and Johnson, C.M., 2004a, Fe isotope variations in the modern and ancient Earth and other planetary bodies: Reviews in Mineralogy and Geochemistry, v. 55, p. 319357. Beard, B.L., and Johnson, C.M., 2004b, Inter-mineral Fe isotope variations in mantle-derived rocks and implications for the Fe geochemical cycle: Geochimica et Cosmochimica Acta, v. 68, p. 47274743, doi:10.1016/j .gca.2004.04.023. Beard, B.L., Johnson, C.M., Cox, L., Sun, H., Nealson, K.H., and Aguilar, C., 1999, Iron isotope biosignatures: Science, v. 285, p. 18891892, doi:10.1126/science.285.5435.1889. Beard, B.L., Johnson, C.M., Skulan, J.L., Nealson, K.H., Cox, L., and Sun, H., 2003a, Application of Fe isotopes to tracing the geochemical and biological cycling of Fe: Chemical Geology, v. 195, p. 87117, doi:10.1016/ S0009-2541(02)00390-X. Beard, B.L., Johnson, C.M., Von Damm, K.L., and Poulson, R.L., 2003b, Iron isotope constraints on Fe cycling and mass balance in oxygenated Earth oceans: Geology, v. 31, p. 629632, doi:10.1130/00917613(2003)031<0629:IICOFC>2.0.CO;2. Beard, B.L., Ludois, J.M., Lapen, T.J., and Johnson, C.M., 2013, Pre4.0 billion year weathering on Mars constrained by Rb-Sr geochronology on meteorite ALH84001: Earth and Planetary Science Letters, v. 361, p. 173182, doi:10.1016/j.epsl.2012.10.021. Becker, L., Poreda, R.J., Hunt, A.G., Bunch, T.E., and Rampino, M., 2001, Impact event at the Permian-Triassic boundary: Evidence from extraterrestrial noble gases in fullerenes: Science, v. 291, p. 15301533, doi:10.1126/science.1057243. Becker, R.H., and Clayton, R.N., 1972, Carbon isotopic evidence for the origin of a banded iron-formation in Western Australia: Geochimica et Cosmochimica Acta, v. 36, p. 577595, doi:10.1016/0016-7037(72)90077-4.

spe500-08 1st pgs Five decades of advances in geochemistry


Bekker, A., Slack, J.F., Planavsky, N., Krapez, B., Hofmann, A., Konhauser, K.O., and Rouxel, O.J., 2010, Iron formation: The sedimentary product of a complex interplay among mantle, tectonic, oceanic, and biospheric processes: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 105, no. 3, p. 467508, doi:10.2113/gsecongeo.105.3.467. Belousova, E.A., Kostitsyn, Y.A., Grifn, W.L., OReilly, S.Y., and Pearson, N.J., 2010, The growth of the continental crust: Constraints from zircon Hf-isotope data: Lithos, v. 119, p. 457466, doi:10.1016/j .lithos.2010.07.024. Bence, A.E., and Albee, A.L., 1968, Empirical correction factors for the electron microanalysis of silicates and oxides: The Journal of Geology, v. 76, p. 382403, doi:10.1086/627339. Bennett, V.C., and DePaolo, D.J., 1987, Proterozoic crustal history of the western United States as determined by neodymium isotopic mapping: Geological Society of America Bulletin, v. 99, p. 674685, doi:10.1130/0016 -7606(1987)99<674:PCHOTW>2.0.CO;2. Bennett, V.C., Brandon, A.D., and Nutman, A.P., 2007, Coupled 142Nd-143Nd isotopic evidence for Hadean mantle dynamics: Science, v. 318, p. 1907 1910, doi:10.1126/science.1145928. Bergquist, P.R., Karuso, P., Cambie, R.C., and Smith, D.J., 1991, Sterol composition and classication of the Porifera: Biochemical Systematics and Ecology, v. 19, p. 1724, doi:10.1016/0305-1978(91)90109-D. Berner, R.A., 2003, The long-term carbon cycle, fossil fuels and atmospheric composition: Nature, v. 426, p. 323326, doi:10.1038/nature02131. Berner, R.A., 2004, The Phanerozoic Carbon Cycle: Oxford, UK, Oxford University Press, 158 p. Bethke, C.M., 2008, Geochemical and Biogeochemical Reaction Modeling: Cambridge, UK, Cambridge University Press, 543 p. Beukes, N.J., and Gutzmer, J., 2008, Origin and paleoenvironmental signicance of major iron formations at the Archean-Paleoproterozoic boundary: Society of Economic Geologists (SEG) Reviews, v. 15, p. 547. Beukes, N.J., Klein, C., Kaufman, A.J., and Hayes, J.M., 1990, Carbonate petrography, kerogen distribution, and carbon and oxygen isotope variations in an Early Proterozoic transition from limestone to iron-formation deposition, Transvaal Supergroup, South-Africa: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 85, p. 663690, doi:10.2113/gsecongeo.85.4.663. Bibring, J.-P., Langevin, Y., Mustard, J. F., Poulet, F., Arvidson, R., Gendrin, A., Gondet, B., Mangold, N., Pinet, P., Forget, F., et al., 2006, Global mineralogical and aqueous Mars history derived from OMEGA/Mars Express data: Science, v. 312, no. 5772, p. 400404, doi:10.1126/science.1122659. Bigeleisen, J., and Mayer, M.G., 1947, Calculation of equilibrium constants for isotopic exchange reactions: The Journal of Chemical Physics, v. 15, p. 261267, doi:10.1063/1.1746492. Birck, J.L., 2004, An overview of isotopic anomalies in extraterrestrial materials and their nucleosynthetic heritage: Reviews in Mineralogy and Geochemistry, v. 55, p. 2564, doi:10.2138/gsrmg.55.1.25. Blichert-Toft, J., and Albarde, F., 2008, Hafnium isotopes in Jack Hills zircons and the formation of the Hadean crust: Earth and Planetary Science Letters, v. 265, p. 686702, doi:10.1016/j.epsl.2007.10.054. Blichert-Toft, J., Frey, F.A., and Albarde, F., 1999, Hf isotope evidence for pelagic sediments in the source of Hawaiian basalts: Science, v. 285, p. 879882, doi:10.1126/science.285.5429.879. Boyet, M., and Carlson, R.W., 2005, 142Nd evidence for early (>4.53 Ga) global differentiation of the silicate Earth: Science, v. 309, p. 576581, doi:10.1126/science.1113634. Boyet, M., and Carlson, R.W., 2006, A new geochemical model for the Earths mantle inferred from 146Sm-142Nd systematics: Earth and Planetary Science Letters, v. 250, p. 254268, doi:10.1016/j.epsl.2006.07.046. Brandon, A.D., and Walker, R.J., 2005, The debate over core-mantle interaction: Earth and Planetary Science Letters, v. 232, p. 211225, doi:10.1016/j .epsl.2005.01.034. Brandon, A.D., Norman, M.D., Walker, R.J., and Morgan, J.W., 1999, 186Os187Os systematics of Hawaiian picrites: Earth and Planetary Science Letters, v. 174, p. 2542, doi:10.1016/S0012-821X(99)00251-4. Brandon, A.D., Snow, J.E., Walker, R.J., Morgan, J.W., and Mock, T.D., 2000, 190Pt-186Os and 187Re-187Os systematics of abyssal peridotites: Earth and Planetary Science Letters, v. 177, p. 319335, doi:10.1016/S0012 -821X(00)00044-3. Brantley, S.L., and Lebedeva, M., 2011, Learning to read the chemistry of regolith to understand the critical zone: Annual Review of Earth and Planetary Sciences, v. 39, p. 387416, doi:10.1146/annurev-earth-040809-152321.

page 33 33

Brassell, S.C., Eglinton, G., Marlowe, I.T., Paumann, U., and Sarnthein, M., 1986, Molecular stratigraphy: A new tool for climatic assessment: Nature, v. 320, p. 129133, doi:10.1038/320129a0. Brocks, J.J., Love, G.D., Summons, R.E., Knoll, A.H., Logan, G.A., and Bowden, S.A., 2005, Biomarker evidence for green and purple sulphur bacteria in a stratied Palaeoproterozoic sea: Nature, v. 437, p. 866870, doi:10.1038/nature04068. Broecker, W., 2012, The carbon cycle and climate change: Memoirs of my 60 years in science: Geochemical Perspectives, v. 1, p. 221339, doi:10.7185/ geochempersp.1.2. Broecker, W.S., and Peng, T.-H., 1982, Tracers in the Sea: Palisades, New York, Lamont-Doherty Geological Observatory, 690 p. Brckner, J., Dreibus, G., Gellert, R., Squyres, S.W., Wnke, H., Yen, A., and Zipfel, J., 2008, Mars Exploration Rovers: Chemical composition by the APXS, in Bell, J., III, ed., The Martian Surface: Composition, Mineralogy, and Physical Properties: Cambridge, UK, Cambridge University Press, p. 58101. Bullen, T.D., White, A.F., Childs, C.W., Vivit, D.V., and Schulz, M.S., 2001, Demonstration of signicant abiotic iron isotope fractionation in nature: Geology, v. 29, p. 699702, doi:10.1130/0091-7613(2001)029<0699 :DOSAII>2.0.CO;2. Burdett, J.W., Arthur, M.A., and Richardson, M., 1989, A Neogene seawater sulfur isotope age curve from calcareous pelagic microfossils: Earth and Planetary Science Letters, v. 94, p. 189198, doi:10.1016/0012 -821X(89)90138-6. Burke, W.H., Denison, R.E., Hetherington, E.A., Koepnick, R.B., Nelson, H.F., and Otto, J.B., 1982, Variation of seawater 87Sr/86Sr throughout Phanerozoic time: Geology, v. 10, p. 516519, doi:10.1130/0091-7613 (1982)10<516:VOSSTP>2.0.CO;2. Burton, K.W., Ling, H.-F., and ONions, R.K., 1997, Closure of the Central American isthmus and its effect on deep-water formation in the North Atlantic: Nature, v. 386, p. 382385, doi:10.1038/386382a0. Byrne, R.H., and Sholkovitz, E.R., 1996, Marine chemistry and geochemistry of the lanthanides, in Gschneidner, K.A., Jr., and Eyring, L., eds., Handbook on the Physics and Chemistry of Rare Earths, Volume 23: Amsterdam, Elsevier, p. 497593. Caneld, D.E., 2001, Biogeochemistry of sulfur isotopes: Reviews in Mineralogy and Geochemistry, v. 43, p. 607636, doi:10.2138/gsrmg.43.1.607. Caneld, D.E., 2005, The early history of atmospheric oxygen: Homage to Robert Garrels: Annual Review of Earth and Planetary Sciences, v. 33, p. 136, doi:10.1146/annurev.earth.33.092203.122711. Caneld, D.E., and Farquhar, J., 2009, Animal evolution, bioturbation, and the sulfate concentration of the oceans: Proceedings of the National Academy of Sciences of the United States of America, v. 106, no. 20, p. 81238127, doi:10.1073/pnas.0902037106. Caneld, D.E., and Teske, A., 1996, Late Proterozoic rise in atmospheric oxygen concentration inferred from phylogenetic and sulphur-isotope studies: Nature, v. 382, p. 127132, doi:10.1038/382127a0. Caneld, D.E., and Thamdrup, B., 1994, The production of 34S-depleted sulde during bacterial disproportionation of elemental sulfur: Science, v. 266, p. 19731975, doi:10.1126/science.11540246. Caneld, D.E., Poulton, S.W., and Narbonne, G.M., 2007, Late-Neoproterozoic deep-ocean oxygenation and the rise of animal life: Science, v. 315, p. 9295, doi:10.1126/science.1135013. Cao, C., Love, G.D., Hays, L.E., Wang, W., Shen, S., and Summons, R.E., 2009, Biogeochemical evidence for euxinic oceans and ecological disturbance presaging the end-Permian mass extinction event: Earth and Planetary Science Letters, v. 281, p. 188201, doi:10.1016/j.epsl .2009.02.012. Carlson, R.W., and Boyet, M., 2008, Composition of the Earths interior: The importance of early events: Philosophical Transactions of the Royal Society, ser. A, v. 366, p. 40774103, doi:10.1098/rsta.2008.0166. Carlson, R.W., Pearson, D.G., and James, D.E., 2005, Physical, chemical, and chronological characteristics of continental mantle: Reviews of Geophysics, v. 43, RG1001, p. 124, doi:10.1029/2004RG000156. Catling, D.C., and Claire, M.W., 2005, How Earths atmosphere evolved to an oxic state: A status report: Earth and Planetary Science Letters, v. 237, no. 12, p. 120, doi:10.1016/j.epsl.2005.06.013. Cavosie, A.J., Valley, J.W., Wilde, S.A., and E.I.M.F, 2005, Magmatic 18O in 44003900 Ma detrital zircons: A record of the alteration and recycling of crust in the Early Archean: Earth and Planetary Science Letters, v. 235, p. 663681, doi:10.1016/j.epsl.2005.04.028.

spe500-08 1st pgs 34 Johnson et al.

page 34

Chan, L.H., 1987, Lithium isotope analysis by thermal ionization mass spectrometry of lithium tetraborate: Analytical Chemistry, v. 59, p. 2662 2665, doi:10.1021/ac00149a007. Chesley, J.T., and Ruiz, J., 1998, Crust-mantle interaction in large igneous provinces: Implications from the Re-Os isotope systematics of the Columbia River ood basalts: Earth and Planetary Science Letters, v. 154, p. 111, doi:10.1016/S0012-821X(97)00176-3. Chikaraishi, Y., Matsumoto, K., Ogawa, N.O., Suga, H., Kitazato, H., and Ohkouchi, N., 2005, Hydrogen, carbon and nitrogen isotopic fractionations during chlorophyll biosynthesis in C3 higher plants: Phytochemistry, v. 66, p. 911920, doi:10.1016/j.phytochem.2005.03.004. Clayton, R.N., 2002, Self-shielding in the solar nebula: Nature, v. 415, p. 860 861, doi:10.1038/415860b. Clayton, R.N., 2004, Oxygen isotopes in meteorites, in Davis, A.M., ed., Treatise on Geochemistry, Volume 1: Meteorites, Comets, and Planets: Oxford, UK, Elsevier, p. 192142. Clayton, R.N., Grossman, L., and Mayeda, T.K., 1973, A component of primitive nuclear composition in carbonaceous meteorites: Science, v. 182, p. 485488, doi:10.1126/science.182.4111.485. Condie, K.C., 1998, Episodic continental growth and super-continents: A mantle avalanche connection?: Earth and Planetary Science Letters, v. 163, p. 97108, doi:10.1016/S0012-821X(98)00178-2. Craig, H., 1961, Isotopic variations in meteoric waters: Science, v. 133, p. 17021703, doi:10.1126/science.133.3465.1702. Craig, H., 1966, Isotopic composition and origin of the Red Sea and Salton Sea geothermal brines: Science, v. 154, p. 15441548, doi:10.1126/science .154.3756.1544. Craig, H., Gordon, L.I., and Horibe, Y., 1963, Isotopic exchange effects in the evaporation of water: Journal of Geophysical Research, v. 68, p. 5079 5087, doi:10.1029/JZ068i017p05079. Creaney, S., and Passe, Q., 1993, Recurring patterns of total organic carbon and source rock quality within a sequence stratigraphic framework: American Association of Petroleum Geologists Bulletin, v. 77, p. 386401. Creaser, R.A., Papanastassiou, D.A., and Wasserburg, G.J., 1991, Negative thermal ion mass spectrometry of osmium, rhenium and iridium: Geochimica et Cosmochimica Acta, v. 55, p. 397401, doi:10.1016/0016 -7037(91)90427-7. Criss, R.E., 1999, Principles of Stable Isotope Distribution: New York, Oxford University Press, 254 p. Criss, R.E., and Taylor, H.P.J., 1983, An 18O/16O and D/H study of Tertiary hydrothermal systems in the southern half of the Idaho batholith: Geological Society of America Bulletin, v. 94, p. 640663, doi:10.1130/0016 -7606(1983)94<640:AOADSO>2.0.CO;2. Cronin, J.R., and Pizzarello, S., 1997, Enantiomeric excesses in meteoritic amino acids: Science, v. 275, p. 951955, doi:10.1126/science.275.5302.951. Cronin, J.R., Pizzarello, S., Epstein, S., and Krishnamurthy, R.V., 1993, Molecular and isotopic analyses of the hydroxy acids, dicarboxylic acids, and hydroxicarboxylic acids of the Murchison meteorite: Geochimica et Cosmochimica Acta, v. 57, p. 47454752, doi:10.1016/0016-7037(93)90197-5. Czaja, A.D., Johnson, C.M., Beard, B.L., Eigenbrode, J.L., Freeman, K.H., and Yamaguchi, K.E., 2010, Iron and carbon isotope evidence for ecosystem and environmental diversity in the 2.7 to 2.5 Ga Hamersley Province, Western Australia: Earth and Planetary Science Letters, v. 292, p. 170 180, doi:10.1016/j.epsl.2010.01.032. Czaja, A.D., Johnson, C.M., Roden, E.E., Beard, B.L., Voegelin, A.R., Naggler, T.F., Beukes, N.J., and Wille, M., 2012, Evidence for free oxygen in the Neoarchean ocean based on coupled iron and molybdenum isotope fractionation: Geochimica et Cosmochimica Acta, v. 86, p. 118137, doi:10.1016/j.gca.2012.03.007. Czaja, A.D., Johnson, C.M., Beard, B.L., Roden, E.E., Li, W., and Moorbath, S., 2013, Biological Fe oxidation controlled deposition of banded iron formation in ca. 3770 Ma Isua Supracrustal Belt (West Greenland): Earth and Planetary Science Letters, v. 363, p. 192203, doi:10.1016/j .epsl.2012.12.025. Dansgaard, W., Johnsen, S.J., Muller, J., and Langway, C.C., 1969, One thousand centuries of climatic record from Camp Century on the Greenland Ice Sheet: Science, v. 166, no. 3903, p. 377380, doi:10.1126/science .166.3903.377. Dauphas, N., van Zuilen, M., Wadhwa, M., Davic, A.M., Marty, B., and Janney, P.E., 2004, Clues from Fe isotope variations on the origin of Early Archean BIFs from Greenland: Science, v. 306, p. 20772080, doi:10.1126/ science.1104639.

Davidson, J.P., 1983, Lesser Antilles isotopic evidence for the role of subducted sediment in island arc magma genesis: Nature, v. 306, p. 253256, doi:10.1038/306253a0. Davidson, J.P., Tepley, F., Palacz, Z., and Meffan-Main, S., 2001, Magma recharge, contamination and residence times revealed by in situ laser ablation isotopic analysis of feldspar in volcanic rocks: Earth and Planetary Science Letters, v. 184, p. 427442, doi:10.1016/S0012-821X (00)00333-2. Davidson, J.P., Morgan, D.J., Charlier, B.L.A., Harlou, R., and Hora, J.M., 2007, Microsampling and isotopic analysis of igneous rocks: Implications for the study of magmatic systems: Annual Review of Earth and Planetary Sciences, v. 35, p. 273311, doi:10.1146/annurev.earth.35.031306.140211. Davis, A.M., and Richter, F.M., 2004, Condensation and evaporation of solar system materials, in Davis, A.M., ed., Treatise on Geochemistry, Volume 1: Meteorites, Comets and Planets: Oxford, UK, Elsevier, p. 407430. Delong, E.F., 2009, The microbial ocean from genomes to biomes: Nature, v. 459, p. 200206, doi:10.1038/nature08059. Delong, E.F., King, L.L., Massana, R., Cittone, H., Murray, A., Schleper, C., and Wakeham, S.G., 1998, Dibiphytanyl ether lipids in nonthermophilic crenarchaeotes: Applied and Environmental Microbiology, v. 64, p. 11331138. DePaolo, D.J., 1981a, Neodymium isotopes in the Colorado Front Range and crust-mantle evolution in the Proterozoic: Nature, v. 291, p. 193197, doi:10.1038/291193a0. DePaolo, D.J., 1981b, Trace element and isotopic effects of combined wallrock assimilation and fractional crystallization: Earth and Planetary Science Letters, v. 53, p. 189202, doi:10.1016/0012-821X(81)90153-9. DePaolo, D.J., 1986, Detailed record of the Neogene Sr isotopic evolution of seawater from DSDP Site 590B: Geology, v. 14, p. 103106, doi:10.1130/0091-7613(1986)14<103:DROTNS>2.0.CO;2. DePaolo, D.J., 2004, Calcium isotopic variations produced by biological, kinetic, radiogenic and nucleosynthetic processes: Reviews in Mineralogy and Geochemistry, v. 55, p. 255288, doi:10.2138/gsrmg.55.1.255. DePaolo, D.J., and Wasserburg, G.J., 1976, Nd isotopic variations and petrogenetic models: Geophysical Research Letters, v. 3, p. 249252, doi:10.1029/GL003i005p00249. Derenne, S., and Robert, F., 2010, Model of molecular structure of the insoluble organic matter isolated from Murchison meteorite: Meteoritics & Planetary Science, v. 45, p. 14611475, doi:10.1111/j.1945-5100.2010.01122.x. Des Marais, D.J., 2001, Isotopic evolution of the biogeochemical carbon cycle during the Precambrian: Reviews in Mineralogy and Geochemistry, v. 43, p. 555578, doi:10.2138/gsrmg.43.1.555. Dhuime, B., Hawkesworth, C.J., Cawood, P.A., and Storey, C.D., 2012, A change in the geodynamics of continental growth 3 billion years ago: Science, v. 335, p. 13341336, doi:10.1126/science.1216066. Dickin, A.P., 1995, Radiogenic Isotope Geology: Cambridge, UK, Cambridge University Press, 452 p. Doe, B.R., and Zartman, R.E., 1979, Plumbotectonics: I. The Phanerozoic, in Barnes, H.L., ed., Geochemistry of Hydrothermal Ore Deposits: New York, Wiley Interscience, p. 2270. Domagal-Goldman, S., Kasting, J.F., Johnston, D.T., and Farquhar, J., 2008, Organic haze, glaciations and multiple sulfur isotopes in the Mid-Archean Era: Earth and Planetary Science Letters, v. 269, p. 2940, doi:10.1016/j .epsl.2008.01.040. Doughty, D.M., Hunter, R.C., Summons, R.E., and Newman, D.K., 2009, 2-methylhopanoids are maximally produced in akinetes of Nostoc punctiforme: Geobiological implications: Geobiology, v. 7, p. 524532, doi:10.1111/j.1472-4669.2009.00217.x. Douthitt, C.B., 1982, The geochemistry of the stable isotopes of silicon: Geochimica et Cosmochimica Acta, v. 46, p. 14491458, doi:10.1016/0016 -7037(82)90278-2. Drever, J.I., Li, Y.-H., and Maynard, J.B., 1988, Geochemical cycles: The continental crust and the oceans, in Gregor, C.B., Garrels, R.M., Mackenzie, F.T., and Maynard, J.B., eds., Chemical Cycles in the Evolution of the Earth: New York, John Wiley, p. 1753. Duan, Y., Anbar, A.D., Arnold, G.L., Lyons, T.W., Gordon, G.W., and Kendall, B., 2010, Molybdenum isotope evidence for mild environmental oxygenation before the Great Oxidation Event: Geochimica et Cosmochimica Acta, v. 74, no. 23, p. 66556668, doi:10.1016/j.gca.2010.08.035. Dutkiewicz, A., Volk, H., George, S.C., Ridley, J., and Buick, R., 2006, Biomarkers from Huronian oil-bearing uid inclusions: An uncontaminated record of life before the Great Oxidation Event: Geology, v. 34, p. 437 440, doi:10.1130/G22360.1.

spe500-08 1st pgs Five decades of advances in geochemistry


Dutta, S., Tripathi, S.M., Mallick, M., Mathews, R.P., Greenwood, P.F., Rao, M.R., and Summons, R.E., 2011, Eocene out-of-India dispersal of Asian dipterocarps: Review of Palaeobotany and Palynology, v. 166, p. 6368, doi:10.1016/j.revpalbo.2011.05.002. Ebel, D.S., and Grossman, L., 2000, Condensation in dust-enriched systems: Geochimica et Cosmochimica Acta, v. 64, p. 339366, doi:10.1016/ S0016-7037(00)00550-0. Eglinton, G., 1970, Chemical Fossils: San Francisco, California, W.H. Freeman. Eglinton, T.I., and Eglinton, G., 2008, Molecular proxies for paleoclimatology: Earth and Planetary Science Letters, v. 275, p. 116, doi:10.1016/j .epsl.2008.07.012. Eigenbrode, J.L., and Freeman, K.H., 2006, Late Archean rise of aerobic microbial ecosystems: Proceedings of the National Academy of Sciences of the United States of America, v. 103, p. 15,75915,764, doi:10.1073/ pnas.0607540103. Eiler, J.M., 2007, Clumped-isotope geochemistry: The study of naturallyoccurring, multiply-substituted isotopologues: Earth and Planetary Science Letters, v. 262, p. 309327, doi:10.1016/j.epsl.2007.08.020. Eldereld, H., 1986, Strontium isotope stratigraphy: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 57, p. 7190, doi:10.1016/0031-0182 (86)90007-6. Engel, A.E.J., Engel, C.G., and Havens, R.G., 1965, Chemical characteristics of ocean basalts and the upper mantle: Geological Society of America Bulletin, v. 76, p. 719734, doi:10.1130/0016-7606(1965)76[719 :CCOOBA]2.0.CO;2. Engel, M.H., and Macko, S.A., 1997, Isotopic evidence for extraterrestrial nonracemic amino acids in the Murchison meteorite: Nature, v. 389, p. 265 268, doi:10.1038/38460. Engel, M.H., and Macko, S.A., 2001, The stereochemistry of amino acids in the Murchison meteorite: Precambrian Research, v. 106, p. 3545, doi:10.1016/S0301-9268(00)00123-6. Engel, M.H., and Nagy, B., 1982, Distribution and enantiomeric composition of amino acids in the Murchison meteorite: Nature, v. 296, p. 837840, doi:10.1038/296837a0. Engel, M.H., Macko, S.A., and Silfer, J.A., 1990, Carbon isotope composition of individual amino acids in the Murchison meteorite: Nature, v. 348, p. 4749, doi:10.1038/348047a0. Epstein, S., Sharp, R.P., and Goddard, I., 1963, Oxygen isotope ratios in Antarctic snow, rn and ice: The Journal of Geology, v. 71, p. 698720, doi:10.1086/626950. Erwin, D.H., 2006, Extinction: How Life on Earth Nearly Ended 250 Million Years Ago: Princeton, Princeton University Press, 320 p. Erwin, D.H., Laamme, M., Tweedt, S.M., Sperling, E.A., Pisani, D., and Peterson, K.J., 2011, The Cambrian conundrum: Early divergence and later ecological success in the early history of animals: Science, v. 334, p. 10911097, doi:10.1126/science.1206375. Escrig, S., Doucelance, R., Moreira, M., and Allgre, C.J., 2005, Os isotope systematics in Fogo Island: Evidence for lower continental crust fragments under the Cape Verde southern islands: Chemical Geology, v. 219, p. 93113, doi:10.1016/j.chemgeo.2005.02.011. Fairbairn, H.W., Schlecht, W.G., Stevens, R.E., Dennen, W.H., Ahrens, L.H., and Chayes, F., 1951, A Co-Operative Investigation of Precision and Accuracy in Chemical, Spectrochemical and Model Analysis of Silicate Rocks: U.S. Geological Survey Bulletin 980, p. 171. Farley, K.A., and Mukhopadhyay, S., 2001, An extraterrestrial impact at the Permian-Triassic boundary?: Science, v. 293, p. 2343, doi:10.1126/ science.293.5539.2343a. Farmer, G.L., and DePaolo, D.J., 1983, Origin of Mesozoic and Tertiary granite from the western United States and implications for pre-Mesozoic crustal structure: I. Nd and Sr isotopic studies in the geocline of the northern Great Basin: Journal of Geophysical Research, v. 88, p. 33793401, doi:10.1029/JB088iB04p03379. Farquhar, J., Bao, H., and Thiemens, M., 2000a, Atmospheric inuence of Earths earliest sulfur cycle: Science, v. 289, p. 756758, doi:10.1126/ science.289.5480.756. Farquhar, J., Savarino, J., Jackson, T.L., and Thiemens, M.H., 2000b, Evidence of atmospheric sulphur in the Martian regolith from sulphur isotopes in meteorites: Nature, v. 404, p. 5052, doi:10.1038/35003517. Farquhar, J., Savarino, J., Airieau, S., and Thiemens, M.H., 2001, Observation of wavelength-sensitive mass-independent sulfur isotope effects during SO2 photolysis: Implications for the early atmosphere: Journal of Geophysical Research, v. 106, p. 32,82932,839, doi:10.1029/2000JE001437.

page 35 35

Farquhar, J., Wu, N.P., Caneld, D.E., and Oduro, H., 2010, Connections between sulfur cycle evolution, sulfur isotopes, sediments, and base metal sulde deposits: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 105, no. 3, p. 509533, doi:10.2113/gsecongeo.105.3.509. Farquhar, J., Zerkle, A.L., and Bekker, A., 2011, Geological constraints on the origin of oxygenic photosynthesis: Photosynthesis Research, v. 107, no. 1, p. 1136, doi:10.1007/s11120-010-9594-0. Faure, G., and Hurley, P.M., 1963, The isotopic composition of strontium in oceanic and continental basalts: Application to the origin of igneous rocks: Journal of Petrology, v. 4, p. 3150, doi:10.1093/petrology/4.1.31. Faure, G., Hurley, P.M., and Powell, J.L., 1965, The isotopic composition of strontium in surface water from the North Atlantic Ocean: Geochimica et Cosmochimica Acta, v. 29, p. 209220, doi:10.1016/0016 -7037(65)90018-9. Fedo, C.M., and Whitehouse, M.J., 2002, Metasomatic origin of quartzpyroxene rock, Akilia, Greenland, and implications for Earths earliest life: Science, v. 296, p. 14481452, doi:10.1126/science.1070336. Feldman, W.C., Maurice, S., Lawrence, D.J., Little, R.C., Lawson, S.L., Gasnault, O., Wiens, R.C., Barraclough, B.L., Elphic, R.C., Prettyman, T.H., Steinberg, J.T., and Binder, A.B., 2001, Evidence for water ice near the lunar poles: Journal of Geophysical Research, v. 106, p. 23,23123,251, doi:10.1029/2000JE001444. Fike, D.A., Grotzinger, J.P., Pratt, L.M., and Summons, R.E., 2006, Oxidation of the Ediacaran ocean: Nature, v. 444, p. 744747, doi:10.1038/ nature05345. Fischer, W.W., and Pearson, A., 2007, Hypotheses for the origin and early evolution of triterpenoid cyclases: Geobiology, v. 5, p. 1934. Fischer, W.W., Summons, R.E., and Pearson, A., 2005, Targeted genomic detection of biosynthetic pathways: Anaerobic production of hopanoid biomarkers by a common sedimentary microbe: Geobiology, v. 3, p. 3340, doi:10.1111/j.1472-4669.2005.00041.x. Fischer, W.W., Schroeder, S., Lacassie, J.P., Beukes, N.J., Goldberg, T., Strauss, H., Horstmann, U.E., Schrag, D.P., and Knoll, A.H., 2009, Isotopic constraints on the Late Archean carbon cycle from the Transvaal Supergroup along the western margin of the Kaapvaal craton, South Africa: Precambrian Research, v. 169, p. 1527, doi:10.1016/j.precamres.2008.10.010. Fleck, R.J., and Criss, R.E., 1985, Strontium and oxygen isotopic variations in Mesozoic and Tertiary plutons of central Idaho: Contributions to Mineralogy and Petrology, v. 90, p. 291308, doi:10.1007/BF00378269. Flynn, G.J., Keller, L.P., Jacobsen, C., and Wirick, S., 2004, An assessment of the amount and types of organic matter contributed to the Earth by interplanetary dust: Advances in Space Research, v. 33, p. 5766, doi:10.1016/j.asr.2003.09.036. Freeman, K.H., Hayes, J.M., Trendel, J.M., and Albrecht, P., 1990, Evidence from carbon isotope measurements for diverse origins of sedimentary hydrocarbons: Nature, v. 343, p. 254256, doi:10.1038/343254a0. Frei, D., Liebscher, A., Franz, G., Wunder, B., Klemme, S., and Blundy, J., 2009a, Trace element partitioning between orthopyroxene and anhydrous silicate melt on the lherzolite solidus from 1.1 to 3.2 GPa and 1,230 to 1,535C in the model system Na2O-CaO-MgO-Al2O3-SiO2: Contributions to Mineralogy and Petrology, v. 157, p. 473490, doi:10.1007/ s00410-008-0346-5. Frei, R., Gaucher, C., Poulton, S.W., and Caneld, D.E., 2009b, Fluctuations in Precambrian atmospheric oxygenation recorded by chromium isotopes: Nature, v. 461, p. 250253, doi:10.1038/nature08266. French, K.L., Tosca, N.J., Cao, C., and Summons, R.E., 2012, Diagenetic and detrital origin of moretane anomalies through the Permian-Triassic boundary: Geochimica et Cosmochimica Acta, v. 84, p. 104125, doi:10.1016/j .gca.2012.01.004. Friedman, I., Redeld, A.C., Schoen, B., and Harris, J., 1964, The variation of the deuterium content of natural waters in the hydrologic cycle: Reviews of Geophysics, v. 2, p. 177224, doi:10.1029/RG002i001p00177. Friedman, I., Lipman, P.W., Obradovich, J.D., Gleason, J.D., and Christiansen, R.L., 1974, Meteoric water in magmas: Science, v. 184, p. 10691072, doi:10.1126/science.184.4141.1069. Gaffney, A.M., Nelson, B.K., Reisberg, L., and Eiler, J., 2005, Oxygen-osmium isotope systematics of West Maui lavas: A record of shallow-level magmatic processes: Earth and Planetary Science Letters, v. 239, p. 122139, doi:10.1016/j.epsl.2005.07.027. Gaines, S.M., Eglinton, G., and Rullkotter, J., 2009, Echoes of Life: What Fossil Molecules Reveal about Earth History: New York, Oxford University Press, 367 p.

spe500-08 1st pgs 36 Johnson et al.

page 36

Garrels, R.M., and Lerman, A., 1981, Phanerozoic cycles of sedimentary carbon and sulfur: Proceedings of the National Academy of Sciences of the United States of America, v. 78, p. 46524656. Garrels, R.M., and Mackenzie, F.T., 1971, Evolution of Sedimentary Rocks: New York, Norton, 397 p. Garrels, R.M., and Mackenzie, F.T., 1972, A quantitative model for the sedimentary rock cycle: Marine Chemistry, v. 1, p. 2741. Garrels, R.M., and Thompson, M.E., 1962, A chemical model for sea water at 25C and one atmosphere total pressure: American Journal of Science, v. 260, p. 5766, doi:10.2475/ajs.260.1.57. Gast, P.W., Tilton, G.R., and Hedge, C., 1964, Isotopic composition of lead and strontium from Ascension and Gough Islands: Science, v. 145, p. 1181 1185, doi:10.1126/science.145.3637.1181. Gellert, R., Rieder, R., Brckner, J., Clark, B.C., Dreibus, G., Klingelhfer, G., Lugmair, G., Ming, D.W., Wnke, H., Yen, A., Zipfel, J., and Squyres, S.W., 2006, The Alpha Particle X-ray Spectrometer (APXS): Results from Gusev crater and calibration report: Journal of Geophysical Research, v. 111, no. E2, doi:10.1029/2005JE002555. George, S.C., Volk, H., Dutkiewicz, A., Ridley, J., and Buick, R., 2008, Preservation of hydrocarbons and biomarkers in oil trapped inside uid inclusions for 2 billion years: Geochimica et Cosmochimica Acta, v. 72, p. 844870, doi:10.1016/j.gca.2007.11.021. Ghosh, P., Adkins, J., Affek, H., Balta, B., Guo, W., Schauble, E.A., Schrag, D., and Eiler, J.M., 2006, 13C18O bonds in carbonate minerals: A new kind of paleothermometer: Geochimica et Cosmochimica Acta, v. 70, p. 14391456, doi:10.1016/j.gca.2005.11.014. Gill, R., 1997, Modern Analytical Geochemistry: An Introduction to Quantitative Chemical Analysis for Earth, Environmental and Materials Scientists: Singapore, Addison Wesley Longman, 329 p. Glavin, D.P., and Dworkin, J.P., 2009, Enrichment of the amino acid l-isovaline by aqueous alteration on CI and CM meteorite parent bodies: Proceedings of the National Academy of Sciences of the United States of America, v. 106, p. 54875492, doi:10.1073/pnas.0811618106. Goldschmidt, V.M., 1954, Geochemistry: Oxford, UK, Oxford University Press, 730 p. Goldstein, S.L., ONions, R.K., and Hamilton, P.J., 1984, A Sm-Nd isotopic study of atmospheric dusts and particulates from major river systems: Earth and Planetary Science Letters, v. 70, p. 221236. Grantham, P.J., and Wakeeld, L.L., 1988, Variations in the sterane carbon number distributions of marine source rock derived crude oils through geological time: Organic Geochemistry, v. 12, p. 6173, doi:10.1016/0146-6380(88)90115-5. Green, T.H., 1994, Experimental studies of trace-element partitioning applicable to igneous petrogenesisSedona 16 years later: Chemical Geology, v. 117, p. 136, doi:10.1016/0009-2541(94)90119-8. Gregor, B., 1992, Some ideas on the rock cycle: 17881988: Geochimica et Cosmochimica Acta, v. 56, p. 29933000, doi:10.1016/0016-7037(92)90285-Q. Gregory, R.T., and Criss, R.E., 1986, Isotopic exchange in open and closed systems: Reviews in Mineralogy, v. 16, p. 91127. Gregory, R.T., and Taylor, H.P.J., 1981, An oxygen isotope prole in a section of Cretaceous oceanic crust, Samail ophiolite, Oman: Journal of Geophysical Research, v. 86, p. 27372755, doi:10.1029/JB086iB04p02737. Grice, K., Gibbison, R., Atkinson, J.E., Schwark, L., Eckardt, C.B., and Maxwell, J.R., 1996, Maleimides (1H-pyrrole-2,5-diones) as molecular indicators of anoxygenic photosynthesis in ancient water columns: Geochimica et Cosmochimica Acta, v. 60, p. 39133924, doi:10.1016/0016 -7037(96)00199-8. Grice, K., Cao, C., Love, G.D., Bottcher, M.E., Twitchett, R.J., Grosjean, E., Summons, R.E., Turgeon, S.C., Dunning, W., and Jin, Y., 2005, Photic zone euxinia during the Permian-Triassic superanoxic event: Science, v. 307, p. 706709, doi:10.1126/science.1104323. Grifn, W.L., Pearson, N.J., Belousova, E., Jackson, S.E., van Achterbergh, E., OReilly, S.Y., and Shee, S.R., 2000, The Hf isotope composition of cratonic mantle: LAM-MC-ICPMS analysis of zircon megacrysts in kimberlites: Geochimica et Cosmochimica Acta, v. 64, p. 133147, doi:10.1016/ S0016-7037(99)00343-9. Grosjean, E., Love, G.D., Stalvies, C., Fike, D.A., and Summons, R.E., 2009, Origin of petroleum in the NeoproterozoicCambrian South Oman Salt Basin: Organic Geochemistry, v. 40, p. 87110, doi:10.1016/j .orggeochem.2008.09.011. Grossman, L., and Larimer, J.W., 1974, Early chemical history of the solar system: Reviews of Geophysics and Space Physics, v. 12, p. 71101, doi:10.1029/RG012i001p00071.

Grotzinger, J., Beaty, D., Dromart, G., Gupta, S., Harris, M., Hurowitz, J., Kocurek, G., McLennan, S., Milliken, R., Ori, G.G., and Sumner, D., 2011, Sedimentary geology: Key concepts and outstanding questions: Astrobiology, v. 11, p. 7787, doi:10.1089/ast.2010.0571. Halevy, I., Zuber, M.T., and Schrag, D.P., 2007, A sulfur dioxide climate feedback on early Mars: Science, v. 318, p. 19031907, doi:10.1126/ science.1147039. Halliday, A.N., 2004, The origin and earliest history of the Earth, in Davis, A.M., ed., Treatise on Geochemistry, Volume 1: Meteorites, Comets, and Planets: Oxford, UK, Elsevier, p. 509557. Halliday, A.N., Rehkemper, M., Lee, D.-C., and Yi, W., 1996, Early evolution of the Earth and Moon: New constraints from Hf-W isotope geochemistry: Earth and Planetary Science Letters, v. 142, p. 7589, doi:10.1016/0012 -821X(96)00096-9. Halliday, A.N., Lee, D.-C., Christensen, J.N., Rehkemper, M., Yi, W., Luo, X., Hall, C.M., Ballentine, C.J., Pettke, T., and Stirling, C., 1998, Applications of Multiple Collector-ICPMS to cosmochemistry, geochemistry, and paleoceanography: Geochimica et Cosmochimica Acta, v. 62, p. 919940, doi:10.1016/S0016-7037(98)00057-X. Hare, E.P., Fogel, M.L., Stafford, J.T.W., Mitchell, A.D., and Hoering, T.C., 1991, The isotopic composition of carbon and nitrogen in individual amino acids isolated from modern and fossil proteins: Journal of Archaeological Science, v. 18, p. 277292, doi:10.1016/0305-4403(91)90066-X. Harper, C.L., and Jacobsen, S.B., 1992, Evidence from coupled 147Sm-143Nd and 146Sm-142Nd systematics for very early (4.5-Gyr) differentiation of the Earths mantle: Nature, v. 360, p. 728732, doi:10.1038/360728a0. Harrison, T.M., Blichert-Toft, J., Miller, W., Albarde, F., Holden, P., and Mojzsis, S.J., 2005, Heterogeneous Hadean hafnium: Evidence of continental crust at 4.4 to 4.5 Ga: Science, v. 310, p. 19471950, doi:10.1126/ science.1117926. Harrison, T.M., Schmitt, A.K., McCulloch, M.T., and Lovera, O.M., 2008, Early (4.5 Ga) formation of terrestrial crust: Lu-Hf, 18O, and Ti thermometry results for Hadean zircons: Earth and Planetary Science Letters, v. 268, p. 476486, doi:10.1016/j.epsl.2008.02.011. Hart, G.L., Johnson, C.M., Hildreth, W., and Shirey, S.B., 2003, New osmium isotope evidence for intracrustal recycling of crustal domains with discrete ages: Geology, v. 31, p. 427430, doi:10.1130/0091-7613 (2003)031<0427:NOIEFI>2.0.CO;2. Hart, S.R., 1988, Heterogeneous mantle domains: Signatures, genesis and mixing chronologies: Earth and Planetary Science Letters, v. 90, p. 273296, doi:10.1016/0012-821X(88)90131-8. Haskin, L., and Gehl, M.A., 1962, The rare-earth distribution in sediments: Journal of Geophysical Research, v. 67, p. 25372541, doi:10.1029/ JZ067i006p02537. Hawkesworth, C., and Schersten, A., 2007, Mantle plumes and geochemistry: Chemical Geology, v. 241, p. 319331, doi:10.1016/j.chemgeo .2007.01.018. Hawkesworth, C.J., ONions, R.K., and Arculus, R.J., 1979, Nd and Sr isotope geochemistry of island arc volcanics, Grenada, Lesser Antilles: Earth and Planetary Science Letters, v. 45, p. 237248, doi:10.1016/0012 -821X(79)90126-2. Hawkesworth, C.J., Kempton, P.D., Rogers, N.W., Ellam, R.M., and van Calsteren, P.W., 1990, Continental mantle lithosphere, and shallow level enrichment processes in the Earths mantle: Earth and Planetary Science Letters, v. 96, p. 256268, doi:10.1016/0012-821X(90)90006-J. Hawkesworth, C., Cawood, P., Kemp, T., Storey, C., and Dhuime, B., 2009, A matter of preservation: Science, v. 323, p. 4950, doi:10.1126/ science.1168549. Hawkesworth, C.J., Dhuime, B., Pietranik, A.B., Cawood, P.A., Kemp, A.I.S., and Storey, C.D., 2010, The generation and evolution of the continental crust: Journal of the Geological Society of London, v. 167, p. 229248, doi:10.1144/0016-76492009-072. Haxel, G.B., Hedrick, J.B., and Orris, G.R., 2005, Rare Earth Elements Critical Resources for High Technology: U. S. Geological Survey Fact Sheet 08702, 4 p. Hayes, J.M., 1983, Geochemical evidence bearing on the origin of aerobiosis, a speculative interpretation, in Schopf, J.W., ed., The Earths Earliest Biosphere: Its Origin and Evolution: Princeton, New Jersey, Princeton University Press, p. 291301. Hayes, J.M., 2001, Fractionation of carbon and hydrogen isotopes in biosynthetic processes: Reviews in Mineralogy and Geochemistry, v. 43, p. 225 277, doi:10.2138/gsrmg.43.1.225.

spe500-08 1st pgs Five decades of advances in geochemistry


Hayes, J.M., Freeman, K.H., Popp, B.N., and Hoham, C.H., 1990, Compoundspecic isotopic analyses: A novel tool for reconstruction of ancient biogeochemical processes: Organic Geochemistry, v. 16, p. 11151128, doi:10.1016/0146-6380(90)90147-R. Hays, L.E., Beatty, T., Henderson, C.M., Love, G.D., and Summons, R.E., 2007, Evidence for photic zone euxinia through the end-Permian mass extinction in the Panthalassic Ocean (Peace River Basin, western Canada): Palaeoworld, v. 16, p. 3950, doi:10.1016/j.palwor.2007.05.008. Hays, L.E., Grice, K., Foster, C.B., and Summons, R.E., 2012, Biomarker and isotopic trends in a PermianTriassic sedimentary section at Kap Stosch, Greenland: Organic Geochemistry, v. 43, p. 6782, doi:10.1016/j .orggeochem.2011.10.010. Hazen, R.M., Papineau, D., Bleeker, W., Downs, R.T., Ferry, J.M., McCoy, T.J., Sverjensky, D.A., and Yang, H., 2008, Mineral evolution: The American Mineralogist, v. 93, p. 16931720, doi:10.2138/am.2008.2955. Hazen, R.M., Hemley, R.L., and Mangum, A.J., 2012, Carbon in Earths interior: Storage, cycling, and life: Eos (Transactions, American Geophysical Union), v. 93, p. 1718, doi:10.1029/2012EO020001. Hedges, J.I., and Keil, R.G., 1995, Sedimentary organic matter preservation: An assessment and speculative synthesis: Marine Chemistry, v. 49, p. 81115, doi:10.1016/0304-4203(95)00008-F. Heimann, A., Johnson, C.M., Beard, B.L., Valley, J.W., Roden, E.E., Spicuzza, M.J., and Beukes, N.J., 2010, Fe, C, and O isotope compositions of banded iron formation carbonates demonstrate a major role for dissimilatory iron reduction in ~2.5 Ga marine environments: Earth and Planetary Science Letters, v. 294, p. 818, doi:10.1016/j.epsl.2010.02.015. Helgeson, H.C., 1968, Evaluation of irreversible reactions in geochemical processes involving minerals and aqueous solutions: I. Thermodynamic relations: Geochimica et Cosmochimica Acta, v. 32, p. 853877, doi:10.1016/0016-7037(68)90100-2. Herbert, T.D., Schuffert, J.D., Andreasen, D., Heusser, L., Lyle, M., Mix, A., Ravelo, A.C., Stott, L.D., and Herguera, J.C., 2001, Collapse of the California Current during glacial maxima linked to climate change on land: Science, v. 293, p. 7176, doi:10.1126/science.1059209. Hess, J., Bender, M.L., and Schilling, J.-G., 1986, Evolution of the ratio of strontium-87 to strontium-86 in seawater from Cretaceous to present: Science, v. 231, p. 979984, doi:10.1126/science.231.4741.979. Hildreth, W., and Moorbath, S., 1988, Crustal contributions to arc magmatism in the Andes of central Chile: Contributions to Mineralogy and Petrology, v. 98, p. 455489, doi:10.1007/BF00372365. Hills, I.R., and Whitehead, E.V., 1966, Triterpanes in optically active petroleum distillates: Nature, v. 209, p. 977979, doi:10.1038/209977a0. Hoefs, J., and Schidlowski, M., 1967, Carbon isotope composition of carbonaceous matter from the Precambrian of the Witwatersrand system: Science, v. 155, p. 10961097, doi:10.1126/science.155.3766.1096. Hofmann, A., Bekker, A., Rouxel, O., Rumble, D., and Master, S., 2009, Multiple sulphur and iron isotope compositions of detrital pyrite in Archaean sedimentary rocks: A new tool for provenance analysis: Earth and Planetary Science Letters, v. 286, no. 34, p. 436445, doi:10.1016/j .epsl.2009.07.008. Holba, A.G., Dzou, L.I.P., Masterson, W.D., Hughes, W.B., Huizinga, B.J., Singletary, M.S., Moldowan, J.M., Mello, M.R., and Tegelaar, E., 1998, Application of 24-norcholestanes for constraining source age of petroleum: Organic Geochemistry, v. 29, p. 12691283, doi:10.1016/S0146 -6380(98)00184-3. Holland, H.D., 1962, Model for the evolution of the Earths atmosphere, in Engel, A.E.J., James, H.L., and Leonard, B.F., eds., Petrologic Studies: A Volume in Honor of A.F. Buddington: New York, Geological Society of America, p. 447477. Holser, W.T., 1977, Catastrophic chemical events in the history of the ocean: Nature, v. 267, p. 403408, doi:10.1038/267403a0. Holser, W.T., Schonlaub, H.-P., Attrep, M., Boeckelmann, K., Klein, P., Magaritz, M., Orth, C.J., Fenninger, A., Jenny, C., Kralik, M., Mauritsch, H., Pak, E., Schramm, J.-M., Stattegger, K., and Schmoller, R., 1989, A unique geochemical record at the Permian/Triassic boundary: Nature, v. 337, p. 3944, doi:10.1038/337039a0. House, C.H., Schopf, J.W., McKeegan, K.D., Coath, C.D., Harrison, T.M., and Stetter, K.O., 2000, Carbon isotopic composition of individual Precambrian microfossils: Geology, v. 28, p. 707710, doi:10.1130/0091 -7613(2000)28<707:CICOIP>2.0.CO;2. Hunt, J.M., 1996, Petroleum Geochemistry and Geology: New York, W.H. Freeman, 743 p.

page 37 37

Hunt, J.M., Philp, R.P., and Kvenvolden, K.A., 2002, Early developments in petroleum geochemistry: Organic Geochemistry, v. 33, p. 10251052, doi:10.1016/S0146-6380(02)00056-6. Hurley, P.M., and Rand, J.R., 1969, Predrift continental nuclei: Science, v. 164, p. 12291242, doi:10.1126/science.164.3885.1229. Hurley, P.M., Bateman, P.C., Fairbairn, H.W., and Pinson, W.H.J., 1965, Investigation of initial 87Sr/86Sr ratios in the Sierra Nevada plutonic province: Geological Society of America Bulletin, v. 76, p. 165174, doi:10.1130/0016-7606(1965)76[165:IOISSR]2.0.CO;2. Hurowitz, J.A., and McLennan, S.M., 2007, A 3.5 Ga record of water-limited, acidic weathering conditions on Mars: Earth and Planetary Science Letters, v. 260, p. 432443, doi:10.1016/j.epsl.2007.05.043. Ingalls, A.E., Shah, S.R., Hansman, R.L., Aluwihare, L.I., Santos, G.M., Druffel, E.R.M., and Pearson, A., 2006, Quantifying archaeal community autotrophy in the mesopelagic ocean using natural radiocarbon: Proceedings of the National Academy of Sciences of the United States of America, v. 103, p. 64426447, doi:10.1073/pnas.0510157103. Jackson, M.G., and Hart, S.R., 2006, Strontium isotopes in melt inclusions from Samoan basalts: Implications for heterogeneity in the Samoan plume: Earth and Planetary Science Letters, v. 245, p. 260277, doi:10.1016/j .epsl.2006.02.040. Jacobsen, S.B., 1988, Isotopic constraints on crustal growth and recycling: Earth and Planetary Science Letters, v. 90, p. 315329, doi:10.1016/0012 -821X(88)90133-1. Jacobsen, S.B., and Wasserburg, G.J., 1980, Sm-Nd isotopic evolution of chondrites: Earth and Planetary Science Letters, v. 50, p. 139155, doi:10.1016/0012-821X(80)90125-9. Jacobsen, S.B., Ranen, M.C., Petaev, M.I., Remo, J.L., OConnell, R.J., and Sasselov, D.D., 2008, Isotopes as clues to the origin and earliest differentiation history of the Earth: Philosophical Transactions of the Royal Society, ser. A, v. 366, p. 41294162, doi:10.1098/rsta.2008.0174. Janik, C.J., Nehring, N.L., and Truesdell, A.H., 1983, Stable isotope geochemistry of thermal uids from Lassen Volcanic National Park, California: Geothermal Resources Council Transactions, v. 7, p. 295300. Jansen, W., and Slaughter, M., 1982, Elemental mapping of minerals by electron microprobe: The American Mineralogist, v. 67, p. 521533. Jerram, D.A., and Higgins, M.D., 2007, 3D analysis of rock textures: Quantifying igneous microstructures: Elements, v. 3, p. 239245, doi:10.2113/ gselements.3.4.239. Jicha, B.R., Johnson, C.M., Hildreth, W., Beard, B.L., Hart, G.L., Shirey, S.B., and Singer, B.S., 2009, Discriminating assimilants and decoupling deepvs. shallow-level crystal records at Mount Adams using 238U-230Th disequilibria and Os isotopes: Earth and Planetary Science Letters, v. 277, p. 3849, doi:10.1016/j.epsl.2008.09.035. Johnson, C.M., 1991, Large-scale crust formation and lithosphere modication beneath middle to late Cenozoic calderas and volcanic elds, western North America: Journal of Geophysical ResearchSolid Earth and Planets, v. 96, p. 13,48513,507, doi:10.1029/91JB00304. Johnson, C.M., 1993, Mesozoic and Cenozoic contributions to crustal growth in the southwestern United States: Earth and Planetary Science Letters, v. 118, p. 7589, doi:10.1016/0012-821X(93)90160-B. Johnson, C.M., Skulan, J.L., Beard, B.L., Sun, H., Nealson, K.H., and Braterman, P.S., 2002, Isotopic fractionation between Fe(III) and Fe(II) in aqueous solutions: Earth and Planetary Science Letters, v. 195, p. 141153, doi:10.1016/S0012-821X(01)00581-7. Johnson, C.M., Beard, B.L., and Albarde, F., 2004, Geochemistry of nontraditional stable isotopes, in Rosso, J.J., ed., Reviews in Mineralogy and Geochemistry, Volume 55: Washington, D.C., Mineralogical Society of America and Geochemical Society, 454 p.. Johnson, C.M., Beard, B.L., Klein, C., Beukes, N.J., and Roden, E.E., 2008a, Iron isotopes constrain biologic and abiologic processes in banded iron formation genesis: Geochimica et Cosmochimica Acta, v. 72, p. 151169, doi:10.1016/j.gca.2007.10.013. Johnson, C.M., Beard, B.L., and Roden, E.E., 2008b, The iron isotope ngerprints of redox and biogeochemical cycling in the modern and ancient Earth: Annual Review of Earth and Planetary Sciences, v. 36, p. 457493, doi:10.1146/annurev.earth.36.031207.124139. Johnson, T.M., and Bullen, T.D., 2004, Mass-dependent fractionation of selenium and chromium isotopes: Reviews in Mineralogy and Geochemistry, v. 55, p. 289317, doi:10.2138/gsrmg.55.1.289. Jolliff, B.L., Gillis, J.J., Haskin, L.A., Korotev, R.L., and Weiczorek, M.A., 2000, Major lunar crustal terranes: Surface expressions and crust-mantle

spe500-08 1st pgs 38 Johnson et al.

page 38

origins: Journal of Geophysical Research, v. 105, p. 41974216, doi:10.1029/1999JE001103. Kah, L.C., Lyons, T.W., and Frank, T.D., 2004, Low marine sulphate and protracted oxygenation of the Proterozoic biosphere: Nature, v. 431, p. 834 838, doi:10.1038/nature02974. Kamo, S.L., Czamanske, G.K., Amelin, Y., Fedorenko, V.A., Davis, D.W., and Tromov, V.R., 2003, Rapid eruption of Siberian ood-volcanic rocks and evidence for coincidence with the Permian-Triassic boundary and mass extinction at 251 Ma: Earth and Planetary Science Letters, v. 214, p. 7591, doi:10.1016/S0012-821X(03)00347-9. Kanasewich, E.R., and Farquhar, R.M., 1965, Lead isotope ratios from the Cobalt-Noranda area, Canada: Canadian Journal of Earth Sciences, v. 2, p. 361384, doi:10.1139/e65-029. Kaplan, I.R., Emery, K.O., and Rittenbebg, S.C., 1963, The distribution and isotopic abundance of sulphur in recent marine sediments off southern California: Geochimica et Cosmochimica Acta, v. 27, p. 297331, doi:10.1016/0016-7037(63)90074-7. Karhu, J.A., and Holland, H.D., 1996, Carbon isotopes and the rise of atmospheric oxygen: Geology, v. 24, p. 867870, doi:10.1130/0091 -7613(1996)024<0867:CIATRO>2.3.CO;2. Kasting, J.F., Howard, M.T., Wallman, K., Veizer, J., Shields, G.A., and Jaffres, J., 2006, Paleoclimates, ocean depth, and the oxygen isotopic composition of seawater: Earth and Planetary Science Letters, v. 252, p. 8293, doi:10.1016/j.epsl.2006.09.029. Kaufman, A.J., and Xiao, S., 2003, High CO2 levels in the Proterozoic atmosphere estimated from analyses of individual microfossils: Nature, v. 425, p. 279282, doi:10.1038/nature01902. Keeling, C.D., 1960, The concentration and isotopic composition of carbon dioxide in the atmosphere: Tellus, v. 12, p. 200203, doi:10.1111/j.2153-3490.1960.tb01300.x. Keith, M.L., and Weber, J.N., 1964, Carbon and oxygen isotopic composition of selected limestones and fossils: Geochimica et Cosmochimica Acta, v. 28, p. 17871816, doi:10.1016/0016-7037(64)90022-5. Kemp, A.I.S., Hawkesworth, C.J., Paterson, B.A., and Kinny, P.D., 2006, Episodic growth of the Gondwana supercontinent from hafnium and oxygen isotopes in zircon: Nature, v. 439, p. 580583, doi:10.1038/nature04505. Kemp, A.I.S., Wilde, S.A., Hawkesworth, C.J., Coath, C.D., Nemchin, A., Pidgeon, R.T., Vervoort, J.D., and DuFrane, S.A., 2010, Hadean crustal evolution revisited: New constraints from Pb-Hf isotope systematics of the Jack Hills zircons: Earth and Planetary Science Letters, v. 296, p. 4556, doi:10.1016/j.epsl.2010.04.043. Kendall, B., Reinhard, C.T., Lyons, T., Kaufman, A.J., Poulton, S.W., and Anbar, A.D., 2010, Pervasive oxygenation along Late Archaean ocean margins: Nature Geoscience, v. 3, no. 9, p. 647652, doi:10.1038/ngeo942. Keto, L.S., and Jacobsen, S.B., 1988, Nd isotopic variations of Phanerozoic paleoceans: Earth and Planetary Science Letters, v. 90, p. 395410, doi:10.1016/0012-821X(88)90138-0. King, P.L., and McLennan, S.M., 2010, Sulfur on Mars: Elements, v. 6, p. 107 112, doi:10.2113/gselements.6.2.107. Kistler, R.W., and Peterman, Z.E., 1973, Variations in Sr, Rb K, Na, and initial 87Sr/86Sr Mesozoic granitic rocks and intruded wall rocks in central California: Geological Society of America Bulletin, v. 84, p. 34893512, doi:10.1130/0016-7606(1973)84<3489:VISRKN>2.0.CO;2. Kleine, T., Touboul, M., Bourdon, B., Nimmo, F., Mezger, K., Palme, H., Jacobsen, S.B., Yin, Q.-Z., and Halliday, A.N., 2009, Hf-W chronology of the accretion and early evolution of asteroids and terrestrial planets: Geochimica et Cosmochimica Acta, v. 73, p. 51505188, doi:10.1016/j .gca.2008.11.047. Klemme, H.D., and Ulmishek, G.F., 1991, Effective petroleum source rocks of the world: Stratigraphic distribution and controlling depositional factors: American Association of Petroleum Geologists Bulletin, v. 75, p. 18091851. Klomp, U.C., 1986, The chemical structure of a pronounced series of isoalkanes in South Oman crudes: Organic Geochemistry, v. 10, p. 807814, doi:10.1016/S0146-6380(86)80017-1. Knauth, L.P., 2005, Temperature and salinity history of the Precambrian ocean: Implications for the course of microbial evolution: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 219, p. 5369, doi:10.1016/j.palaeo .2004.10.014. Knauth, L.P., and Epstein, S., 1976, Hydrogen and oxygen isotope ratios in nodular and bedded cherts: Geochimica et Cosmochimica Acta, v. 40, p. 10951108, doi:10.1016/0016-7037(76)90051-X.

Knesel, K.M., Davidson, J.P., and Dufeld, W.A., 1999, Evolution of silicic magma through assimilation and subsequent recharge: Evidence from Sr isotopes in sanidine phenocrysts, Taylor Creek Rhyolite, NM: Journal of Petrology, v. 40, p. 773786, doi:10.1093/petroj/40.5.773. Knoll, A.H., Bambach, R.K., Caneld, D.E., and Grotzinger, J.P., 1996, Comparative Earth history and Late Permian mass extinction: Science, v. 273, p. 452457, doi:10.1126/science.273.5274.452. Knoll, A.H., Walter, M.R., Narbonne, G.M., and Christie-Blick, N., 2004, A new period for the geologic time scale: Science, v. 305, p. 621622, doi:10.1126/science.1098803. Knoll, A.H., Summons, R.E., Waldbauer, J.R., and Zumberge, J., 2007, The geological succession of primary producers in the oceans, in Falkwoski, P., and Knoll, A.H., eds., The Evolution of Primary Producers in the Sea: Boston, Massachusetts, Elsevier, p. 133163. Kohnen, M.E.L., Sinninghe Damst, J.S., Ten Haven, H.L., and De Leeuw, J.W., 1989, Early incorporation of polysulphides in sedimentary organic matter: Nature, v. 341, p. 640641, doi:10.1038/341640a0. Koopmans, M.P., Kaster, J., Van Kaam-Peters, H.M.E., Kenig, F., Schouten, S., Hartgers, W.A., de Leeuw, J.W., and Sinninghe Damst, J.S., 1996, Diagenetic and catagenetic products of isorenieratene: Molecular indicators for photic zone anoxia: Geochimica et Cosmochimica Acta, v. 60, p. 44674496, doi:10.1016/S0016-7037(96)00238-4. Kump, L.R., Pavlov, A., and Arthur, M.A., 2005, Massive release of hydrogen sulde to the surface ocean and atmosphere during intervals of oceanic anoxia: Geology, v. 33, p. 397400, doi:10.1130/G21295.1. Kurz, M.D., Jenkins, W.J., Schilling, J.G., and Hart, S.R., 1982, Helium isotopic variations in the mantle beneath the central North Atlantic Ocean: Earth and Planetary Science Letters, v. 58, p. 114, doi:10.1016/0012 -821X(82)90099-1. Kusch, S., Eglinton, T.I., Mix, A.C., and Mollenhauer, G., 2010, Timescales of lateral sediment transport in the Panama Basin as revealed by radiocarbon ages of alkenones, total organic carbon and foraminifera: Earth and Planetary Science Letters, v. 290, p. 340350, doi:10.1016/j.epsl.2009.12.030. Kuypers, M.M.M., van Breugel, Y., Schouten, S., Erba, E., and Sinninghe Damst, J.S., 2004, N2-xing cyanobacteria supplied nutrient N for Cretaceous oceanic anoxic events: Geology, v. 32, p. 853856, doi:10.1130/G20458.1. Kvenvolden, K., 2006, Organic geochemistryA retrospective of its rst 70 years: Organic Geochemistry, v. 37, p. 111, doi:10.1016/j.orggeochem.2005.09.001. Kvenvolden, K., Lawless, J., Pering, K., Peterson, E., Flores, J., Ponnamperuma, C., Kaplan, I.R., and Moore, C., 1970, Evidence for extraterrestrial amino-acids and hydrocarbons in the Murchison meteorite: Nature, v. 228, p. 923926, doi:10.1038/228923a0. Lasaga, A.C., and Berner, R.A., 1998, Fundamental aspects of quantitative models for geochemical cycles: Chemical Geology, v. 145, p. 161175, doi:10.1016/S0009-2541(97)00142-3. Lassiter, J.C., and Hauri, E.H., 1998, Osmium-isotope variations in Hawaiian lavas: Evidence for recycled oceanic lithosphere in the Hawaiian plume: Earth and Planetary Science Letters, v. 164, p. 483496, doi:10.1016/ S0012-821X(98)00240-4. Lawless, J.G., Zeitman, B., Pereira, W.E., Summons, R.E., and Dufeld, A.M., 1974, Dicarboxylic acids in the Murchison meteorite: Nature, v. 251, p. 4042, doi:10.1038/251040a0. Lawrence, D.J., Feldman, W.C., Barraclough, N.L., Binder, A.B., Elphic, R.C., Maurice, S., and Thomsen, D.R., 1998, Global elemental maps of the Moon: The Lunar Prospector gamma-ray spectrometer: Science, v. 281, p. 14841489, doi:10.1126/science.281.5382.1484. Laws, E.A., Popp, B.N., Bidigare, R.R., Kennicutt, M.C., and Macko, S.A., 1995, Dependence of phytoplankton carbon isotopic composition on growth rate and [CO2]aq: Theoretical considerations and experimental results: Geochimica et Cosmochimica Acta, v. 59, p. 11311138, doi:10.1016/0016-7037(95)00030-4. Laws, E.A., Bidigare, R.R., and Popp, B.N., 1997, Effect of growth rate and CO2 concentration on carbon isotopic fractionation by the marine diatom Phaeodactylum tricornutum: Limnology and Oceanography, v. 42, p. 15521560, doi:10.4319/lo.1997.42.7.1552. Lee, T., Papanastassiou, D.A., and Wasserburg, G.J., 1977, Aluminum-26 in the early solar system: Fossil or fuel?: Astrophysical Journal Letters, v. 211, p. L107L110, doi:10.1086/182351. Lepland, A., van Zuilen, M.A., Arrhenius, G., Whitehouse, M.J., and Fedo, C.M., 2005, Questioning the evidence for Earths earliest life Akilia revisited: Geology, v. 33, p. 7779, doi:10.1130/G20890.1.

spe500-08 1st pgs Five decades of advances in geochemistry


Lerman, A., and Wu, L., 2007, Kinetics of global geochemical cycles, in Brantley, S.L., Kubicki, J.D., and White, A.F., eds., Kinetics of Water-Rock Interaction: New York, Springer, p. 655736. Lewis, R.S., Tang, M., Wacker, J.F., Anders, E., and Steel, E., 1987, Interstellar diamonds in meteorites: Nature, v. 326, p. 160162, doi:10.1038/326160a0. Lichtfouse, E., 2000, Compound-specic isotope analysis: Application to archaeology, biomedical sciences, biosynthesis, environment, extraterrestrial chemistry, food science, forensic science, humic substances, microbiology, organic geochemistry, soil science and sport: Rapid Communications in Mass Spectrometry, v. 14, p. 13371344, doi:10.1002/1097-0231(20000815)14:15<1337::AID -RCM9>3.0.CO;2-B. Lipp, J.S., Morono, Y., Inagaki, F., and Hinrichs, K.-U., 2008, Signicant contribution of Archaea to extant biomass in marine subsurface sediments: Nature, v. 454, p. 991994, doi:10.1038/nature07174. Long, J.V.P., 1995, Microanalysis from 1950 to the 1990s, in Potts, P.J., Bowles, J.F.W., Reed, S.J.B., and Cave, M.R., eds., Microprobe Techniques in the Earth Sciences: London, Chapman & Hall, p. 148. Love, G.D., Grosjean, E., Stalvies, C., Fike, D.A., Grotzinger, J.P., Bradley, A.S., Kelly, A.E., Bhatia, M., Meredith, W., Snape, C.E., Bowring, S.A., Condon, D.J., and Summons, R.E., 2009, Fossil steroids record the appearance of Demospongiae during the Cryogenian period: Nature, v. 457, p. 718721, doi:10.1038/nature07673. Luck, J.M., and Allgre, C.J., 1983, 187Re-187Os systematics in meteorites and cosmological consequences: Nature, v. 302, p. 130132, doi:10 .1038/302130a0. Luguet, A., Pearson, D.G., Nowell, G.M., Dreher, S.T., Coggon, J.A., Spetsius, Z.V., and Parman, S.W., 2008, Enriched Pt-Re-Os isotope systematics in plume lavas explained by metasomatic suldes: Science, v. 319, p. 453 456, doi:10.1126/science.1149868. Lyons, J.R., 2009, Atmospherically-derived mass-independent sulfur isotope signatures and incorporation into sediments: Chemical Geology, v. 267, p. 164174, doi:10.1016/j.chemgeo.2009.03.027. Macko, S.A., Uhle, M.E., Engel, M.H., and Andrusevich, V., 1997, Stable nitrogen isotope analysis of amino acid enantiomers by gas chromatography/ combustion/isotope ratio mass spectrometry: Analytical Chemistry, v. 69, p. 926929, doi:10.1021/ac960956l. Mao, H.K., Wu, Y., Chen, L.C., Shu, J.F., and Jephcoat, A.P., 1990, Static compression of iron to 300 GPa and Fe0.8Ni0.2 alloy to 260 GPa: Implications for composition of the core: Journal of Geophysical Research, v. 95, p. 21,73721,742, doi:10.1029/JB095iB13p21737. Martins, Z., Botta, O., Fogel, M.L., Sephton, M.A., Glavin, D.P., Watson, J.S., Dworkin, J.P., Schwartz, A.W., and Ehrenfreund, P., 2008, Extraterrestrial nucleobases in the Murchison meteorite: Earth and Planetary Science Letters, v. 270, p. 130136, doi:10.1016/j.epsl.2008.03.026. Mason, B., 1952, Principles of Geochemistry: New York, John Wiley, 276 p. Mason, B., 1992, Victor Moritz Goldschmidt: Father of Modern Geochemistry: The Geochemical Society Special Publication 4, 184 p. Mason, B., and Taylor, S.R., 1982, Inclusions in the Allende Meteorite: Washington, D. C., Smithsonian Institution, Smithsonian Contributions to Earth Science, 30 p. Masterson, A.L., Farquhar, J., and Wing, B.A., 2011, Sulfur mass-independent fractionation patterns in the broadband UV photolysis of sulfur dioxide: Pressure and third body effects: Earth and Planetary Science Letters, v. 306, p. 253260, doi:10.1016/j.epsl.2011.04.004. McCaffrey, M.A., Michael Moldowan, J., Lipton, P.A., Summons, R.E., Peters, K.E., Jeganathan, A., and Watt, D.S., 1994, Paleoenvironmental implications of novel C30 steranes in Precambrian to Cenozoic age petroleum and bitumen: Geochimica et Cosmochimica Acta, v. 58, p. 529532, doi:10.1016/0016-7037(94)90481-2. McCulloch, M.T., Gregory, R.T., Wasserburg, G.J., and Taylor, H.P.J., 1981, Sm-Nd, Rb-Sr, and 18O/16O isotopic systematics in an oceanic crustal section: Evidence from the Samail ophiolite: Journal of Geophysical Research, v. 86, p. 27212735, doi:10.1029/JB086iB04p02721. McKeegan, K.D., Kallio, A.P.A., Heber, V.S., Jarzebinski, G., Mao, P.H., Coath, C.D., Kunihiro, T., Wiens, R.C., Nordholt, J.E., Moses, R.W., Reisenfeld, D.B., Jurewicz, A.J.G., and Burnett, D.S., 2011, The oxygen isotopic composition of the Sun inferred from captured solar wind: Science, v. 332, p. 15281532, doi:10.1126/science.1204636. McLennan, S.M., 2012, Geochemistry of sedimentary processes on Mars, in Grotzinger, J.P., and Milliken, R.E., eds., Sedimentary Geology on Mars: Society for Sedimentary Geology Special Publication 102, p. 119138.

page 39 39

McLoughlin, N., Grosch, E.G., Kilburn, M.R., and Wacey, D., 2012, Sulfur isotope evidence for a Paleoarchean subseaoor biosphere, Barberton, South Africa: Geology, v. 40, p. 10311034, doi:10.1130/G33313.1. McSween, H.Y.J., and Huss, G.R., 2010, Cosmochemistry: Cambridge, UK, Cambridge University Press, 568 p. McSween, H.Y.J., and Keil, K., 2000, Mixing relationships in the Martian regolith and the composition of globally homogeneous dust: Geochimica et Cosmochimica Acta, v. 64, p. 21552166, doi:10.1016/S0016 -7037(99)00401-9. McSween, H.Y.J., Taylor, G.J., and Wyatt, M.B., 2009, Elemental composition of the Martian crust: Science, v. 324, p. 736739, doi:10.1126/ science.1165871. Menzies, M.A., 1989, Cratonic, circumcratonic and oceanic mantle domains beneath the western United States: Journal of Geophysical Research, v. 94, p. 78997915, doi:10.1029/JB094iB06p07899. Michard, A., Gurriet, P., Soudant, M., and Albarde, F., 1985, Nd isotopes in French Phanerozoic shales: External vs. internal aspects of crustal evolution: Geochimica et Cosmochimica Acta, v. 49, p. 601610, doi:10.1016/0016-7037(85)90051-1. Mojzsis, S.J., Arrhenius, G., McKeegan, K.D., Harrison, T.M., Nutman, A.P., and Friend, C.R., 1996, Evidence for life on Earth before 3,800 million years ago: Nature, v. 384, p. 5559, doi:10.1038/384055a0. Mojzsis, S.J., Harrison, T.M., and Pidgeon, R.T., 2001, Oxygen-isotope evidence from ancient zircons for liquid water at the Earths surface 4,300 Myr ago: Nature, v. 409, p. 178181, doi:10.1038/35051557. Moldowan, J.M., Dahl, J., Huizinga, B.J., Fago, F.J., Hickey, L.J., Peakman, T.M., and Taylor, D.W., 1994, The molecular fossil record of oleanane and its relation to angiosperms: Science, v. 265, p. 768771, doi:10.1126/ science.265.5173.768. Mollenhauer, G., Eglinton, T.I., Ohkouchi, N., Schneider, R.R., Miller, P.J., Grootes, P.M., and Rullkatter, J., 2003, Asynchronous alkenone and foraminifera records from the Benguela upwelling system: Geochimica et Cosmochimica Acta, v. 67, p. 21572171, doi:10.1016/S0016 -7037(03)00168-6. Mollenhauer, G., McManus, J.F., Wagner, T., McCave, I.N., and Eglinton, T.I., 2011, Radiocarbon and 230Th data reveal rapid redistribution and temporal changes in sediment focussing at a North Atlantic drift: Earth and Planetary Science Letters, v. 301, p. 373381, doi:10.1016/j .epsl.2010.11.022. Muehlenbachs, K., and Clayton, R.N., 1976, Oxygen isotope composition of the oceanic crust and its bearing on seawater: Journal of Geophysical Research, v. 81, p. 43654369, doi:10.1029/JB081i023p04365. Mundil, R., Plfy, J., Renne, P.R., and Brack, P., 2010, The Triassic timescale: New constraints and a review of geochronological data, in Lucas, S.G., ed., The Triassic Timescale: Geological Society of London Special Publication 334, p. 4160, doi:10.1144/SP334.3. Murray, A.P., Edwards, D., Hope, J.M., Boreham, C.J., Booth, W.E., Alexander, R.A., and Summons, R.E., 1998, Carbon isotope biogeochemistry of plant resins and derived hydrocarbons: Organic Geochemistry, v. 29, p. 11991214, doi:10.1016/S0146-6380(98)00126-0. Nesbitt, H.W., 2003, Petrogenesis of siliciclastic sediments and sedimentary rocks, in Lentz, D.R., ed., Geochemistry of Sedimentary Rocks: Evolutionary Considerations to Mineral Deposit-Forming Environments, Volume 4: Saint Johns, Canada, Geological Association of Canada GeoText, p. 3951. Nesbitt, H.W., and Wilson, R.E., 1992, Recent chemical weathering of basalts: American Journal of Science, v. 292, p. 740777, doi:10.2475/ ajs.292.10.740. Nesbitt, H.W., and Young, G.M., 1984, Prediction of some weathering trends of plutonic and volcanic rocks based on thermodynamic and kinetic considerations: Geochimica et Cosmochimica Acta, v. 48, p. 15231534, doi:10.1016/0016-7037(84)90408-3. Nier, A.O., 1940, A mass spectrometer for routine isotope abundance measurements: The Review of Scientic Instruments, v. 11, p. 212216, doi:10.1063/1.1751688. Nittler, L.R., 2003, Presolar stardust in meteorites: Recent advances and scientic frontiers: Earth and Planetary Science Letters, v. 209, p. 259273, doi:10.1016/S0012-821X(02)01153-6. Noble, R.A., Alexander, R., Kagi, R.I., and Knox, J., 1985, Tetracyclic diterpenoid hydrocarbons in some Australian coals, sediments and crude oils: Geochimica et Cosmochimica Acta, v. 49, p. 21412147, doi:10.1016/0016-7037(85)90072-9.

spe500-08 1st pgs 40 Johnson et al.

page 40

Norton, D., and Knight, J.E., 1977, Transport phenomena in hydrothermal systems: Cooling plutons: American Journal of Science, v. 277, p. 937981, doi:10.2475/ajs.277.8.937. Norton, D., and Taylor, H.P.J., 1979, Quantitative simulation of the hydrothermal systems of crystallizing magma on the basis of transport theory and oxygen isotope data: An analysis of the Skaergaard intrusion: Journal of Petrology, v. 20, p. 421486, doi:10.1093/petrology/20.3.421. Nyquist, L.E., Kleine, T., Shih, C.-Y., and Reese, Y.D., 2009, The distribution of short-lived radioisotopes in the early solar system and the chronology of asteroid accretion, differentiation, and secondary mineralization: Geochimica et Cosmochimica Acta, v. 73, p. 51155136, doi:10.1016/j .gca.2008.12.031. Oehler, D.Z., Schopf, J.W., and Kvenvolden, K.A., 1972, Carbon isotopic studies of organic matter in Precambrian rocks: Science, v. 175, p. 12461248, doi:10.1126/science.175.4027.1246. Ohkouchi, N., Kawamura, K., Kawahata, H., and Okada, H., 1999, Depth ranges of alkenone production in the central Pacic Ocean: Global Biogeochemical Cycles, v. 13, p. 695704, doi:10.1029/1998GB900024. ONeil, J., Shaw, S.E., and Flood, R.H., 1977, Oxygen and hydrogen isotope compositions as indicators of granite genesis in the New England batholith, Australia: Contributions to Mineralogy and Petrology, v. 62, p. 313 328, doi:10.1007/BF00371018. ONeil, J., Carlson, R.W., Francis, D., and Stevens, R.K., 2008, Neodymium-142 evidence for Hadean mac crust: Science, v. 321, p. 18281831, doi:10.1126/science.1161925. ONions, R.K., Hamilton, P.J., and Evensen, N.M., 1977, Variations in 143Nd/144Nd and 87Sr/86Sr ratios in oceanic basalts: Earth and Planetary Science Letters, v. 34, p. 1322, doi:10.1016/0012-821X(77)90100-5. ONions, R.K., Hamilton, P.J., and Hooker, P.J., 1983, A Nd isotope investigation of sediments related to crustal development in the British Isles: Earth and Planetary Science Letters, v. 63, p. 229240, doi:10.1016/0012 -821X(83)90039-0. Orphan, V.J., House, C.H., Hinrichs, K.-U., McKeegan, K.D., and DeLong, E.F., 2002, Multiple archaeal groups mediate methane oxidation in anoxic cold seep sediments: Proceedings of the National Academy of Sciences of the United States of America, v. 99, p. 76637668, doi:10.1073/ pnas.072210299. Ourisson, G., and Albrecht, P., 1992, GeohopanoidsThe most abundant natural-products on Earth: Accounts of Chemical Research, v. 25, p. 398 402, doi:10.1021/ar00021a003. Ourisson, G., Albrecht, P., and Rohmer, M., 1979, The Hopanoids: Palaeochemistry and biochemistry of a group of natural products: Pure and Applied Chemistry, v. 51, p. 709729. Pancost, R.D., Crawford, N., Magness, S., Turner, A., Jenkyns, H.C., and Maxwell, J.R., 2004, Further evidence for the development of photic-zone euxinic conditions during Mesozoic oceanic anoxic events: Journal of the Geological Society of London, v. 161, p. 353364, doi:10.1144/0016764903-059. Park, R., and Epstein, S., 1960, Carbon isotope fractionation during photosynthesis: Geochimica et Cosmochimica Acta, v. 21, p. 110126, doi:10.1016/S0016-7037(60)80006-3. Patchett, P.J., and Tatsumoto, M., 1980, Hafnium isotope variations in oceanic basalts: Geophysical Research Letters, v. 7, p. 10771080, doi:10.1029/ GL007i012p01077. Patchett, P.J., Kouvo, O., Hedge, C.E., and Tatsumoto, M., 1981, Evolution of continental crust and mantle heterogeneity: Evidence from Hf isotopes: Contributions to Mineralogy and Petrology, v. 78, p. 279297, doi:10.1007/BF00398923. Patchett, P.J., White, W.M., Feldmann, H., Kielinczuk, S., and Hofmann, A.W., 1984, Hafnium/rare earth element fractionation in the sedimentary system and crustal recycling into the Earths mantle: Earth and Planetary Science Letters, v. 69, p. 365378, doi:10.1016/0012-821X(84)90195-X. Patterson, C., 1956, Age of meteorites and the Earth: Geochimica et Cosmochimica Acta, v. 10, p. 230237, doi:10.1016/0016-7037(56)90036-9. Patterson, C., and Tatsumoto, M., 1964, The signicance of lead isotopes in detrital feldspar with respect to chemical differentiation within the Earths mantle: Geochimica et Cosmochimica Acta, v. 28, p. 122, doi:10.1016/0016-7037(64)90052-3. Paul, B., Wood, J.A., Hergt, J., Danyushevsky, L., Kunihiro, T., and Nakamura, E., 2011, Melt inclusion Pb-isotopic analysis by LA-MC-ICPMS: Assessment of analytical performance and application to OIB genesis: Chemical Geology, v. 289, p. 210223, doi:10.1016/j.chemgeo.2011.08.005.

Pavlov, A.A., and Kasting, J.F., 2002, Mass-independent fractionation of sulfur isotopes in Archean sediments: Strong evidence for an anoxic Archean atmosphere: Astrobiology, v. 2, p. 2741, doi:10.1089/153110702753621321. Payne, J.L., and Clapham, M.E., 2012, End-Permian mass extinction in the oceans: An ancient analog for the twenty-rst century?: Annual Review of Earth and Planetary Sciences, v. 40, p. 89111, doi:10.1146/annurev -earth-042711-105329. Pearson, A., McNichol, A.P., Benitez-Nelson, B.C., Hayes, J.M., and Eglinton, T.I., 2001, Origins of lipid biomarkers in Santa Monica Basin surface sediment: A case study using compound-specic 14C analysis: Geochimica et Cosmochimica Acta, v. 65, p. 31233137, doi:10.1016/S0016 -7037(01)00657-3. Pearson, A., Leavitt, W.D., Saenz, J.P., Summons, R.E., Tam, M.C.M., and Close, H.G., 2009, Diversity of hopanoids and squalene-hopene cyclases across a tropical land-sea gradient: Environmental Microbiology, v. 11, no. 5, p. 12081223, doi:10.1111/j.1462-2920.2008.01817.x. Pearson, D.G., and Wittig, N., 2008, Formation of Archean continental lithosphere and its diamonds: The root of the problem: Journal of the Geological Society of London, v. 165, p. 895914, doi:10.1144/0016-76492008-003. Peck, W.H., Valley, J.W., Wilde, S.A., and Graham, C.M., 2001, Oxygen isotope ratios and rare earth elements in 3.3 to 4.4 Ga zircons: Ion microprobe evidence for high d18O continental crust and oceans in the Early Archean: Geochimica et Cosmochimica Acta, v. 65, p. 42154229, doi:10.1016/ S0016-7037(01)00711-6. Peltzer, E.T., Bada, J.L., Schlesinger, G., and Miller, S.L., 1984, The chemical conditions on the parent body of the Murchison meteorite: Some conclusions based on amino, hydroxy and dicarboxylic acids: Advances in Space Research, v. 4, p. 6974, doi:10.1016/0273-1177(84)90546-5. Perry, E.C., Jr., 1967, The oxygen isotope chemistry of ancient cherts: Earth and Planetary Science Letters, v. 3, p. 6266, doi:10.1016/0012 -821X(67)90012-X. Perry, E.C., Jr., Monster, J., and Reimer, T., 1971, Sulfur isotopes in Swaziland system barites and the evolution of the Earths atmosphere: Science, v. 171, p. 10151016, doi:10.1126/science.171.3975.1015. Petaev, M.I., and Wood, J.A., 1998, The condensation with partial isolation (CWPI) model of condensation in the solar nebula: Meteoritics & Planetary Science, v. 33, p. 11231137, doi:10.1111/j.1945-5100.1998. tb01717.x. Peterman, Z.E., Hedge, C.E., and Tourtelot, H.A., 1970, Isotopic composition of strontium in sea water throughout Phanerozoic time: Geochimica et Cosmochimica Acta, v. 34, p. 105120, doi:10.1016/0016-7037(70)90154-7. Peters, K.E., Walters, C.C., and Moldowan, J.M., 2005, The Biomarker Guide: Biomarkers and Isotopes in Petroleum Exploration and Earth History: Cambridge, UK, Cambridge University Press, 700 p. Peterson, K.J., Summons, R.E., and Donoghue, P.C.J., 2007, Molecular palaeobiology: Palaeontology, v. 50, p. 775809, doi:10.1111/j.1475 -4983.2007.00692.x. Pettijohn, F.J., 1975, Sedimentary Rocks: New York, Harper and Row, 628 p. Philippot, P., van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J., and Van Kranendonk, M.J., 2007, Early Archaean microorganisms preferred elemental sulfur, not sulfate: Science, v. 317, p. 15341537, doi:10.1126/ science.1145861. Piepgras, D.J., and Wasserburg, G.J., 1980, Neodymium isotopic variations in seawater: Earth and Planetary Science Letters, v. 50, p. 128138, doi:10.1016/0012-821X(80)90124-7. Piepgras, D.J., Wasserburg, G.J., and Dasch, E.J., 1979, The isotopic composition of Nd in different ocean masses: Earth and Planetary Science Letters, v. 45, p. 223236, doi:10.1016/0012-821X(79)90125-0. Planavsky, N., Rouxel, O., Bekker, A., Shapiro, R., Fralick, P., and Knudsen, A., 2009, Iron-oxidizing microbial ecosystems thrived in late Paleoproterozoic redox-stratied oceans: Earth and Planetary Science Letters, v. 286, p. 230242, doi:10.1016/j.epsl.2009.06.033. Poitrasson, F., Mao, X.-L., Mao, S.S., Freydier, R., and Russo, R.E., 2003, Comparison of ultraviolet femtosecond and nanosecond laser ablation inductively coupled plasma mass spectrometry analysis in glass, monazite, and zircon: Analytical Chemistry, v. 75, p. 61846190, doi:10.1021/ ac034680a. Polyakov, V.B., and Mineev, S.D., 2000, The use of Mossbauer spectroscopy in stable isotope geochemistry: Geochimica et Cosmochimica Acta, v. 64, p. 849865, doi:10.1016/S0016-7037(99)00329-4. Popp, B.N., Laws, E.A., Bidigare, R.R., Dore, J.E., Hanson, K.L., and Wakeham, S.G., 1998, Effect of phytoplankton cell geometry on carbon isotopic

spe500-08 1st pgs Five decades of advances in geochemistry


fractionation: Geochimica et Cosmochimica Acta, v. 62, p. 6977, doi:10.1016/S0016-7037(97)00333-5. Prahl, F., Wolf, G. V., and Sparrow, M. A., 2003, Physiological impacts on alkenone paleothermometry: Paleoceanography, v. 18, 1025, doi:10.1029/2002PA000803, 2. Prettyman, T.H., Hagerty, J.J., Elphic, R.C., Feldman, W.C., Lawrence, D.J., McKinney, G.W., and Vaniman, D.T., 2006, Elemental composition of the lunar surface: Analysis of gamma ray spectroscopy data from Lunar Prospector: Journal of Geophysical Research, v. 111, no. E12, E12007, doi:10.1029/2005JE002656. Puchtel, I.S., Brandon, A.D., Humayun, M., and Walker, R.J., 2005, Evidence for the early differentiation of the core from Pt-Re-Os isotope systematics of 2.8-Ga komatiites: Earth and Planetary Science Letters, v. 237, p. 118134, doi:10.1016/j.epsl.2005.04.023. Raiswell, R., and Caneld, D.E., 2012, The iron biogeochemical cycle past and present: Geochemical Perspectives, v. 1, p. 1220, doi:10.7185/ geochempersp.1.1. Rampen, S.W., Schouten, S., Abbas, B., Elda Panoto, F., Muyzer, G., Campbell, C.N., Fehling, J., and Sinninghe Damst, J.S., 2007, On the origin of 24-norcholestanes and their use as age-diagnostic biomarkers: Geology, v. 35, p. 419422, doi:10.1130/G23358A.1. Rasmussen, B., Fletcher, I.R., Brocks, J.J., and Kilburn, M.R., 2008, Reassessing the rst appearance of eukaryotes and cyanobacteria: Nature, v. 455, p. 11011104, doi:10.1038/nature07381. Raymo, M.E., Ruddiman, W.F., and Froelich, P.N., 1988, Inuence of late Cenozoic mountain building on ocean geochemical cycles: Geology, v. 16, p. 649653, doi:10.1130/0091-7613(1988)016<0649:IOLCMB >2.3.CO;2. Reed, S.J.B., 1995, Electron probe microanalysis, in Potts, P.J., Bowles, J.F.W., Reed, S.J.B., and Cave, M.R., eds., Microprobe Techniques in the Earth Sciences: London, Chapman & Hall, p. 4989. Richard, P., Shimizu, N., and Allgre, C.J., 1976, 143Nd/146Nd, a natural tracer: An application to oceanic basalts: Earth and Planetary Science Letters, v. 31, p. 269278, doi:10.1016/0012-821X(76)90219-3. Righter, K., 2003, Metal-silicate partitioning of siderophile elements and core formation in the early Earth: Annual Review of Earth and Planetary Sciences, v. 31, p. 135174, doi:10.1146/annurev.earth.31.100901.145451. Righter, K., and Drake, M.J., 1996, Core formation in Earths Moon, Mars and Vesta: Icarus, v. 124, p. 513529, doi:10.1006/icar.1996.0227. Rino, S., Komiya, T., Windley, B.F., Katayama, I., Motoki, A., and Hirata, T., 2004, Major episodic increases of continental crustal growth determined from zircon ages of river sands; implications for mantle overturns in the early Precambrian: Physics of the Earth and Planetary Interiors, v. 146, p. 369394, doi:10.1016/j.pepi.2003.09.024. Rohmer, M., 2003, Mevalonate-independent methylerythritol phosphate pathway for isoprenoid biosynthesis. Elucidation and distribution: Pure and Applied Chemistry, v. 75, p. 375387, doi:10.1351/pac200375020375. Rohmer, M., 2010, Hopanoids, in Timmis, K.N., ed., Handbook of Hydrocarbon and Lipid Microbiology: Berlin, Springer, p. 133142. Rohmer, M., Dastillung, M., and Ourisson, G., 1980, Hopanoids from C30 to C35 in recent muds: Naturwissenschaften, v. 67, p. 456458, doi:10.1007/ BF00405643. Rohmer, M., Bouvier-Nave, P., and Ourisson, G., 1984, Distribution of hopanoid triterpanes in prokaryotes: Journal of General Microbiology, v. 130, p. 11371150. Rohmer, M., Knani, M., Simonin, P., Sutter, B., and Sahm, H., 1993, Isoprenoid biosynthesis in bacteria: A novel pathway for the early steps leading to isopentenyl diphosphate: Biochemical Journal, v. 295, part 2, p. 517524. Rosell-Mel, A., Eglinton, G., Paumann, U., and Sarnthein, M., 1995, Atlantic core-top calibration of the U37K index as a sea-surface palaeotemperature indicator: Geochimica et Cosmochimica Acta, v. 59, p. 30993107, doi:10.1016/0016-7037(95)00199-A. Rouxel, O.J., Bekker, A., and Edwards, K.J., 2005, Iron isotope constraints on the Archean and Paleoproterozoic ocean redox state: Science, v. 307, p. 10881091, doi:10.1126/science.1105692. Rudnick, R.L., and Walker, R.J., 2009, Interpreting age from Re-Os isotopes in peridotites: Lithos, v. 112S, p. 10831095, doi:10.1016/j .lithos.2009.04.042. Russell, W.A., Papanastassiou, D.A., and Tombrello, T.A., 1978, Ca isotope fractionation on the Earth and other solar system materials: Geochimica et Cosmochimica Acta, v. 42, p. 10751090, doi:10.1016/0016 -7037(78)90105-9.

page 41 41

Rtters, H., Sass, H., Cypionka, H., and Rullkotter, J., 2002, Phospholipid analysis as a tool to study complex microbial communities in marine sediments: Journal of Microbiological Methods, v. 48, p. 149160, doi:10.1016/S0167-7012(01)00319-0. Saal, A.E., Hart, S.R., Shimizu, N., Hauri, E.H., and Layne, G.D., 1998, Pb isotopic variability in melt inclusions from oceanic island basalts, Polynesia: Science, v. 282, p. 14811484, doi:10.1126/science.282.5393.1481. Salters, V.J.M., 1996, The generation of mid-ocean ridge basalts from the Hf and Nd isotope perspective: Earth and Planetary Science Letters, v. 141, p. 109123, doi:10.1016/0012-821X(96)00070-2. Salters, V.J.M., and Hart, S.R., 1991, The mantle sources of ocean ridges, islands and arcs: The Hf-isotope connection: Earth and Planetary Science Letters, v. 104, p. 364380, doi:10.1016/0012-821X(91)90216-5. Savin, S.M., and Epstein, S., 1970, The oxygen and hydrogen isotope geochemistry of clay minerals: Geochimica et Cosmochimica Acta, v. 34, p. 2542, doi:10.1016/0016-7037(70)90149-3. Schidlowski, M., 1987, Application of stable carbon isotopes to early biochemical evolution on Earth: Annual Review of Earth and Planetary Sciences, v. 15, p. 4772, doi:10.1146/annurev.ea.15.050187.000403. Schidlowski, M., Eichman, R., and Junge, C.E., 1975, Precambrian sedimentary carbonates: Carbon and oxygen isotope geochemistry and implication for the terrestrial oxygen budget: Precambrian Research, v. 2, p. 169, doi:10.1016/0301-9268(75)90018-2. Schidlowski, M., Appel, P.W.U., Eichmann, R., and Junge, C.E., 1979, Carbon isotope geochemistry of the 3.7109-yr-old Isua sediments, West Greenland: Implications for the Archaean carbon and oxygen cycles: Geochimica et Cosmochimica Acta, v. 43, p. 189199, doi:10.1016/0016 -7037(79)90238-2. Schouten, S., Hopmans, E.C., Pancost, R.D., and Sinninghe Damst, J.S.,2000a, Widespread occurrence of structurally diverse tetraether membrane lipids: Evidence for the ubiquitous presence of low-temperature relatives of hyperthermophiles: Proceedings of the National Academy of Sciences of the United States of America, v. 97, p. 14,42114,426, doi:10.1073/pnas.97.26.14421. Schouten, S., Van Kaam-Peters, H.M.E., Rijpstra, W.I.C., Schoell, M., and Sinninghe Damst, J.S., 2000b, Effects of an oceanic anoxic event on the stable carbon isotopic composition of early Toarcian carbon: American Journal of Science, v. 300, p. 122, doi:10.2475/ajs.300.1.1. Schouten, S., Hopmans, E.C., Schefu, E., and Sinninghe Damst, J.S., 2002, Distributional variations in marine crenarchaeotal membrane lipids: A new tool for reconstructing ancient sea water temperatures?: Earth and Planetary Science Letters, v. 204, p. 265274, doi:10.1016/S0012 -821X(02)00979-2. Schouten, S., Hopmans, E.C., Forster, A., van Breugel, Y., Kuypers, M.M.M., and Sinninghe Damste, J.S., 2003, Extremely high sea-surface temperatures at low latitudes during the Middle Cretaceous as revealed by archaeal membrane lipids: Geology, v. 31, p. 10691072, doi:10.1130/G19876.1. Schouten, S., Hopmans, E.C., Baas, M., Boumann, H., Standfest, S., Konneke, M., Stahl, D.A., and Sinninghe Damst, J.S., 2008, Intact membrane lipids of Candidatus Nitrosopumilus maritimus, a cultivated representative of the cosmopolitan mesophilic group I Crenarchaeota: Applied and Environmental Microbiology, v. 74, p. 24332440, doi:10.1128/AEM.01709-07. Seifert, W.K., and Moldowan, J.M., 1978, Applications of steranes, terpanes and monoaromatics to the maturation, migration and source of crude oils: Geochimica et Cosmochimica Acta, v. 42, p. 7795, doi:10.1016/0016 -7037(78)90219-3. Seifert, W.K., and Moldowan, J.M., 1980, The effect of thermal stress on source-rock quality as measured by hopane stereochemistry: Physics and Chemistry of the Earth, v. 12, no. 229237. Sephton, M.A., Looy, C.V., Brinkhuis, H., Wignall, P.B., de Leeuw, J.W., and Visscher, H., 2005, Catastrophic soil erosion during the end-Permian biotic crisis: Geology, v. 33, p. 941944, doi:10.1130/G21784.1. Sephton, M.A., Love, G.D., Watson, J.S., Verchovsky, A.B., Wright, I.P., Snape, C.E., and Gilmour, I., 2004, Hydropyrolysis of insoluble carbonaceous matter in the Murchison meteorite: New insights into its macromolecular structure: Geochimica et Cosmochimica Acta, v. 68, p. 13851393, doi:10.1016/j.gca.2003.08.019. Sepulveda, J., Wendler, J.E., Summons, R.E., and Hinrichs, K.-U., 2009, Rapid resurgence of marine productivity after the Cretaceous-Paleogene mass extinction: Science, v. 326, p. 129132, doi:10.1126/science.1176233. Sessions, A.L., Burgoyne, T.W., Schimmelmann, A., and Hayes, J.M., 1999, Fractionation of hydrogen isotopes in lipid biosynthesis: Organic Geochemistry, v. 30, p. 11931200, doi:10.1016/S0146-6380(99)00094-7.

spe500-08 1st pgs 42 Johnson et al.

page 42

Sharp, Z.D., 1990, A laser-based microanalytical method for the in situ determination of oxygen isotope ratios of silicates and oxides: Geochimica et Cosmochimica Acta, v. 54, p. 13531357, doi:10.1016/0016-7037(90)90160-M. Shen, S.-z., Crowley, J.L., Wang, Y., Bowring, S.A., Erwin, D.H., Sadler, P.M., Cao, C.-q., Rothman, D.H., Henderson, C.M., Ramezani, J., Zhang, H., Shen, Y., Wang, X.-d., Wang, W., Mu, L., Li, W.-z., Tang, Y.-g., Liu, X.-l., Liu, L.-j., Zeng, Y., Jiang, Y.-f., and Jin, Y.-g., 2011, Calibrating the endPermian mass extinction: Science, v. 334, p. 13671372, doi:10.1126/ science.1213454. Sheppard, S.M.F., 1986, Characterization and isotopic variations in natural waters: Reviews in Mineralogy, v. 16, p. 165183. Shields, G.A., and Veizer, J., 2002, The Precambrian marine carbonate isotope database: Version 1.1: Geochemistry Geophysics Geosystems, v. 3, no. 6, 12 p., doi:10.1029/2001GC000266. Shirey, S.B., and Richardson, S.H., 2011, Start of the Wilson cycle at 3 Ga shown by diamonds from subcontinental mantle: Science, v. 333, p. 434 436, doi:10.1126/science.1206275. Shirey, S.B., and Walker, R.J., 1998, The Re-Os isotope system in cosmochemistry and high-temperature geochemistry: Annual Review of Earth and Planetary Sciences, v. 26, p. 423500, doi:10.1146/annurev .earth.26.1.423. Sikes, E.L., and Volkman, J.K., 1993, Calibration of alkenone unsaturation ratios (Uk37) for paleotemperature estimation in cold polar waters: Geochimica et Cosmochimica Acta, v. 57, p. 18831889, doi:10.1016/0016-7037(93)90120-L. Silln, L.G., 1961, The physical chemistry of sea water, in Sears, M., ed., Oceanography: Washington, D.C., American Association for the Advancement of Science, p. 549581. Simon, J.I., Reid, M.R., and Young, E.D., 2007, Lead isotopes by LA-MC-ICPMS: Tracking the emergence of mantle signatures in an evolving silicic magma system: Geochimica et Cosmochimica Acta, v. 71, p. 20142035, doi:10.1016/j.gca.2007.01.023. Simoneit, B.R.T., Grimalt, J.O., Wang, T.G., Cox, R.E., Hatcher, P.G., and Nissenbaum, A., 1986, Cyclic terpenoids of contemporary resinous plant detritus and of fossil woods, ambers and coals: Organic Geochemistry, v. 10, p. 877889, doi:10.1016/S0146-6380(86)80025-0. Sinninghe Damst, J.S., and Koopmans, M.P., 1997, The fate of carotenoids in sediments: An overview: Pure and Applied Chemistry, v. 69, p. 2067 2074, doi:10.1351/pac199769102067. Sinninghe Damst, J.S., Rijpstra, W.I.C., Hopmans, E.C., Prahl, F.G., Wakeham, S.G., and Schouten, S., 2002, Distribution of membrane lipids of planktonic Crenarchaeota in the Arabian Sea: Applied and Environmental Microbiology, v. 68, p. 29973002, doi:10.1128/AEM.68.6.2997-3002.2002. Sinninghe Damst, J.S., Muyzer, G., Abbas, B., Rampen, S.W., Masse, G., Allard, W.G., Belt, S.T., Robert, J.-M., Rowland, S.J., Moldowan, J.M., Barbanti, S.M., Fago, F.J., Denisevich, P., Dahl, J., Trindade, L.F., and Schouten, S., 2004, The rise of the rhizosolenid diatoms: Science, v. 304, p. 584587, doi:10.1126/science.1096806. Sinninghe Damst, J.S., Rijpstra, W.I.C., Geenevasen, J.A.J., Strous, M., and Jetten, M.S.M., 2005, Structural identication of ladderane and other membrane lipids of planctomycetes capable of anaerobic ammonium oxidation (anammox): The FEBS Journal, v. 272, p. 42704283, doi:10.1111/j.1742-4658.2005.04842.x. Smith, F.A., and Freeman, K.H., 2006, Inuence of physiology and climate on D of leaf wax n-alkanes from C3 and C4 grasses: Geochimica et Cosmochimica Acta, v. 70, p. 11721187, doi:10.1016/j.gca.2005.11.006. Solomon, G.C., and Taylor, H.P.J., 1989, Isotopic evidence for the origin of Mesozoic and Cenozoic granitic plutons in the northern Great Basin: Geology, v. 17, p. 591594, doi:10.1130/0091-7613(1989)017<0591 :IEFTOO>2.3.CO;2. Spooner, E.T.C., 1976, The strontium isotopic composition of seawater, and seawater-oceanic crust interaction: Earth and Planetary Science Letters, v. 31, p. 167174, doi:10.1016/0012-821X(76)90108-4. Squyres, S.W., Grotzinger, J.P., Arvidson, R.E., Bell, J.F., Calvin, W., Christensen, P.R., Clark, B.C., Crisp, J.A., Farrand, W.H., Herkenhoff, K.E., Johnson, J.R., Klingelhaufer, G., Knoll, A.H., McLennan, S.M., McSween, H.Y., Morris, R.V., Rice, J.W., Rieder, R., and Soderblom, L.A., 2004, In situ evidence for an ancient aqueous environment at Meridiani Planum, Mars: Science, v. 306, p. 17091714, doi:10.1126/science.1104559. Stacey, J.S., and Kramers, J.D., 1975, Approximation of terrestrial lead isotope evolution by a two-stage model: Earth and Planetary Science Letters, v. 26, p. 207221, doi:10.1016/0012-821X(75)90088-6.

Steinhoefel, G., Horn, I., and Von Blanckenburg, F., 2009, Micro-scale tracing of Fe and Si isotope signatures in banded iron formation using femtosecond laser ablation: Geochimica et Cosmochimica Acta, v. 73, p. 5343 5360, doi:10.1016/j.gca.2009.05.037. Stevenson, R.K., and Patchett, P.J., 1990, Implications for the evolution of continental crust from Hf isotope systematics of Archean detrital zircons: Geochimica et Cosmochimica Acta, v. 54, p. 16831697, doi:10.1016/0016-7037(90)90400-F. Sturt, H.F., Summons, R.E., Smith, K., Elvert, M., and Hinrichs, K.-U., 2004, Intact polar membrane lipids in prokaryotes and sediments deciphered by high-performance liquid chromatography/electrospray ionization multistage mass spectrometrynew biomarkers for biogeochemistry and microbial ecology: Rapid Communications in Mass Spectrometry, v. 18, p. 617628, doi:10.1002/rcm.1378. Summons, R.E., and Powell, T.G., 1987, Identication of aryl isoprenoids in source rocks and crude oils: Biological markers for the green sulphur bacteria: Geochimica et Cosmochimica Acta, v. 51, p. 557566, doi:10.1016/0016-7037(87)90069-X. Summons, R.E., Brassell, S.C., Eglinton, G., Evans, E., Horodyski, R.J., Robinson, N., and Ward, D.M., 1988, Distinctive hydrocarbon biomarkers from fossiliferous sediment of the Late Proterozoic Walcott Member, Chuar Group, Grand Canyon, Arizona: Geochimica et Cosmochimica Acta, v. 52, p. 26252637, doi:10.1016/0016-7037(88)90031-2. Summons, R.E., Bradley, A.S., Jahnke, L.L., and Waldbauer, J.R., 2006, Steroids, triterpenoids and molecular oxygen: Philosophical Transactions of the Royal Society of London, ser. B, Biological Sciences, v. 361, p. 951 968, doi:10.1098/rstb.2006.1837. Sundquist, E.T., and Visser, K., 2003, The geologic history of the carbon cycle, in Holland, H.D., and Turekian, K.K., eds., Treatise on Geochemistry, Volume 8, Biogeochemistry: Oxford, UK, Elsevier, p. 425472. Sutton, S.R., Caffee, M.W., and Dove, M.T., 2006, Synchrotron radiation, neutron, and mass spectrometry techniques at user facilities: Elements, v. 2, p. 1521, doi:10.2113/gselements.2.1.15. Swisher, C.C., III, Grajales-Nishimura, J.M., Montanari, A., Margolis, S.V., Claeys, P., Alvarez, W., Renne, P., Cedillo-Pardoa, E., Maurrasse, F.J., Curtis, G.H., Smit, J., and McWilliams, M.O., 1992, Coeval 40Ar/39Ar ages of 65.0 million years ago from Chicxulub crater melt rock and Cretaceous-Tertiary boundary tektites: Science, v. 257, p. 954958, doi:10.1126/science.257.5072.954. Talbot, H.M., Summons, R.E., Jahnke, L.L., Cockell, C.S., Rohmer, M., and Farrimond, P., 2008, Cyanobacterial bacteriohopanepolyol signatures from cultures and natural environmental settings: Organic Geochemistry, v. 39, p. 232263, doi:10.1016/j.orggeochem.2007.08.006. Tang, Ming, and Anders, E., 1988, Isotopic anomalies of Ne, Xe, and C in meteorites. II. Interstellar diamond and SiC: Carriers of exotic noble gases: Geochimica et Cosmochimica Acta, v. 52, p. 12351244, doi:10.1016/0016-7037(88)90277-3. Tatsumoto, M., 1966, Genetic relations of oceanic basalts as indicated by lead isotopes: Science, v. 153, p. 10941101, doi:10.1126/science .153.3740.1094. Taylor, H.P., Jr., 1980, The effects of assimilation of country rocks by magmas on 18O/16O and 87Sr/86Sr systematics in igneous rocks: Earth and Planetary Science Letters, v. 47, p. 243254, doi:10.1016/0012 -821X(80)90040-0. Taylor, H.P., Jr., and Sheppard, S.M.F., 1986, Igneous rocks: I. Processes of isotopic fractionation and isotopic systematics: Reviews in Mineralogy, v. 16, p. 227271. Taylor, L.A., Nazarov, M.A., Demidova, S.I., and Patchen, A., 2001, A new mare basalt from Oman: Meteoritics & Planetary Science, v. 36, p. A204. Taylor, S.R., and Jakes, P., 1974, The geochemical evolution of the Moon, in Proceedings of the Lunar and Planetary Science Conference, Volume 5: Houston, Texas, Lunar and Planetary Institute, p. 12871305. Taylor, S.R., and McLennan, S.M., 1985, The Continental Crust: Its Composition and Evolution: Oxford, UK, Blackwell Scientic Publications, 312 p. Taylor, S.R., and McLennan, S.M., 2009, Planetary Crusts: Their Composition, Origin and Evolution: Cambridge, UK, Cambridge University Press, 378 p. Taylor, S.R., Pieters, C.M., and MacPherson, G.J., 2006, Earth-Moon system, planetary science, and lessons learned: Reviews in Mineralogy and Geochemistry, v. 60, p. 657704, doi:10.2138/rmg.2006.60.7. Tera, F., Brown, L., Morris, J., Sacks, I.S., Klein, J., and Middleton, R., 1986, Sediment incorporation in island-arc magmas: Inferences from 10Be:

spe500-08 1st pgs Five decades of advances in geochemistry


Geochimica et Cosmochimica Acta, v. 50, p. 535550, doi:10.1016/0016 -7037(86)90103-1. Thiagarajan, N., Adkins, J., and Eiler, J., 2011, Carbonate clumped isotope thermometry of deep-sea corals and implications for vital effects: Geochimica et Cosmochimica Acta, v. 75, p. 44164425, doi:10.1016/j .gca.2011.05.004. Thirlwall, M.F., 1991, Long-term reproducibility of multicollector Sr and Nd isotope ratio analyses: Chemical GeologyIsotope Geoscience Section, v. 94, p. 85104, doi:10.1016/0168-9622(91)90002-E. Thode, H.G., Monster, J., and Dunford, H.B., 1961, Sulphur isotope geochemistry: Geochimica et Cosmochimica Acta, v. 25, p. 159174, doi:10.1016/0016-7037(61)90074-6. Tissot, B.P., and Welte, D.H., 1978, Petroleum Formation and Occurrence: A New Approach to Oil and Gas Exploration: New York, Springer-Verlag, 554 p. Tomascak, P. B., 2004, Developments in the understanding and application of lithium isotopes in the earth and planetary sciences: Reviews in Mineralogy and Geochemistry, v. 55, p. 153195. Tripati, A.K., Eagle, R.A., Thiagarajan, N., Gagnon, A.C., Bauch, H., Halloran, P.R., and Eiler, J.M., 2010, 13C-18O isotope signatures and clumped isotope thermometry in foraminifera and coccoliths: Geochimica et Cosmochimica Acta, v. 74, p. 56975717, doi:10.1016/j.gca.2010.07.006. Tsikos, H., Matthews, A., Erel, Y., and Moore, J.M., 2010, Iron isotopes constrain biogeochemical redox cycling of iron and manganese in a Palaeoproterozoic stratied basin: Earth and Planetary Science Letters, v. 298, no. 12, p. 125134, doi:10.1016/j.epsl.2010.07.032. Twitchett, R.J., Looy, C.V., Morante, R., Visscher, H., and Wignall, P.B., 2001, Rapid and synchronous collapse of marine and terrestrial ecosystems during the end-Permian biotic crisis: Geology, v. 29, p. 351354, doi:10.1130/0091-7613(2001)029<0351:RASCOM>2.0.CO;2. Tyrrell, S., Haughton, P.D.W., and Daly, J.S., 2007, Drainage reorganization during breakup of Pangea revealed by in-situ Pb isotopic analysis of detrital K-feldspar: Geology, v. 35, p. 971974, doi:10.1130/G4123A.1. Urey, H.C., 1947, The thermodynamic properties of isotopic substances: Journal of the Chemical Society, May 1947, p. 562581. Valaas-Hyslop, E., Valley, J.W., Johnson, C.M., and Beard, B.L., 2008, The effects of metamorphism on O and Fe isotope compositions in the Biwabik iron-formation, northern Minnesota: Contributions to Mineralogy and Petrology, v. 155, p. 313328, doi:10.1007/s00410-007-0244-2. Valley, J.W., 1986, Stable isotope geochemistry of metamorphic rocks: Reviews in Mineralogy, v. 16, p. 445489. Valley, J.W., and Kita, N.T., 2009, In situ oxygen isotope geochemistry by ion microprobe: Toronto, Mineralogical Associate of Canada, Short Course Volume 41, p. 1963. Valley, J.W., Peck, W.H., King, E.M., and Wilde, S.A., 2002, A cool early Earth: Geology, v. 30, p. 351354, doi:10.1130/0091-7613(2002)030<0351 :ACEE>2.0.CO;2. Valley, J.W., Cavosie, A.J., Fu, B., Peck, W.H., and Wilde, S.A., 2006, Comment on Heterogeneous Hadean hafnium: Evidence of continental crust at 4.4 to 4.5 Ga: Science, v. 312, p. 1139, doi:10.1126/science.1125301. Van Kaam-Peters, H.M.E., Rijpstra, W.I.C., De Leeuw, J.W., and Sinninghe Damst, J.S., 1998a, A high resolution biomarker study of different lithofacies of organic sulfur-rich carbonate rocks of a Kimmeridgian lagoon (French southern Jura): Organic Geochemistry, v. 28, p. 151177, doi:10.1016/S0146-6380(97)00132-0. Van Kaam-Peters, H.M.E., Schouten, S., Kester, J., and Sinninghe Damst, J.S., 1998b, Controls on the molecular and carbon isotopic composition of organic matter deposited in a Kimmeridgian euxinic shelf sea: Evidence for preservation of carbohydrates through sulfurisation: Geochimica et Cosmochimica Acta, v. 62, p. 32593283, doi:10.1016/S0016 -7037(98)00231-2. van Zuilen, M.A., Lepland, A., and Arrhenius, G., 2002, Reassessing the evidence for the earliest traces of life: Nature, v. 418, p. 627630, doi:10.1038/nature00934. Veizer, J., and Compston, W., 1976, 87Sr/86Sr in Precambrian carbonates as an index of crustal evolution: Geochimica et Cosmochimica Acta, v. 40, p. 905914, doi:10.1016/0016-7037(76)90139-3. Veizer, J., Compston, W., Clauer, N., and Schidlowski, M., 1983, 87Sr/86Sr in Late Proterozoic carbonates: Evidence for a mantle event at 900 Ma ago: Geochimica et Cosmochimica Acta, v. 47, p. 295302, doi:10.1016/0016 -7037(83)90142-4. Vervoort, J.D., and Blichert-Toft, J., 1999, Evolution of the depleted mantle: Hf isotope evidence from juvenile rocks through time: Geochi-

page 43 43

mica et Cosmochimica Acta, v. 63, p. 533556, doi:10.1016/S0016 -7037(98)00274-9. Vervoort, J.D., Patchett, P.J., Blichert-Toft, J., and Albarde, F., 1999, Relationships between Lu-Hf and Sm-Nd isotopic systems in the global sedimentary system: Earth and Planetary Science Letters, v. 168, p. 7999, doi:10.1016/S0012-821X(99)00047-3. Vinogradov, A.P., 1943, Biogeochemical research in the U.S.S.R.: Nature, v. 151, p. 659661, doi:10.1038/151659a0. Voegelin, A.R., Nagler, T.F., Beukes, N.J., and Lacassie, J.P., 2010, Molybdenum isotopes in Late Archean carbonate rocks: Implications for early Earth oxygenation: Precambrian Research, v. 182, no. 12, p. 7082, doi:10.1016/j.precamres.2010.07.001. Voice, P.J., Kowalewski, M., and Eriksson, K.A., 2011, Quantifying the timing and rate of crustal evolution: Global compilation of radiometrically dated detrital zircon grains: The Journal of Geology, v. 119, p. 109126, doi:10.1086/658295. Volkman, J.K., 2003, Sterols in microorganisms: Applied Microbiology and Biotechnology, v. 60, p. 495506. Volkman, J.K., Eglinton, G., Corner, E.D.S., and Forsberg, T.E.V., 1980, Long-chain alkenes and alkenones in the marine coccolithophorid Emiliania huxleyi: Phytochemistry, v. 19, p. 26192622, doi:10.1016/S0031 -9422(00)83930-8. Volkman, J.K., Barrerr, S.M., Blackburn, S.I., and Sikes, E.L., 1995, Alkenones in Gephyrocapsa oceanica: Implications for studies of paleoclimate: Geochimica et Cosmochimica Acta, v. 59, p. 513520, doi:10.1016/0016-7037(95)00325-T. Von Blanckenburg, F., Mamberti, M., Schoenberg, R., Kamber, B.S., and Webb, G.E., 2008, The iron isotope composition of microbial carbonate: Chemical Geology, v. 249, p. 113128, doi:10.1016/j.chemgeo.2007.12.001. Wacey, D., Kilburn, M.R., Saunders, M., Cliff, J., and Brasier, M.D., 2011, Microfossils of sulphur-metabolizing cells in 3.4-billion-year-old rocks of Western Australia: Nature Geoscience, v. 4, p. 698702, doi:10.1038/ ngeo1238. Waldbauer, J.R., Sherman, L.S., Sumner, D.Y., and Summons, R.E., 2009, Late Archean molecular fossils from the Transvaal Supergroup record the antiquity of microbial diversity and aerobiosis: Precambrian Research, v. 169, p. 2847, doi:10.1016/j.precamres.2008.10.011. Walker, R.J., 2009, Highly siderophile elements in the Earth, Moon and Mars: Update and implications for planetary accretion and differentiation: Chemie der Erde, v. 69, p. 101125, doi:10.1016/j.chemer.2008.10.001. Walker, R.J., Carlson, R.W., Shirey, S.B., and Boyd, F.R., 1989, Os, Sr, Nd, and Pb isotope systematics of southern African peridotite xenoliths: Implications for the chemical evolution of subcontinental mantle: Geochimica et Cosmochimica Acta, v. 53, p. 15831595, doi:10.1016/0016 -7037(89)90240-8. Walker, R.J., Morgan, J.W., and Horan, M.F., 1995, Osmium-187 enrichment in some plumes: Evidence for core-mantle interaction?: Science, v. 269, p. 819822, doi:10.1126/science.269.5225.819. Wang, Z., Schauble, E.A., and Eiler, J.M., 2004, Equilibrium thermodynamics of multiply substituted isotopologues of molecular gases: Geochimica et Cosmochimica Acta, v. 68, p. 47794797, doi:10.1016/j.gca.2004.05.039. Watanabe, Y., Farquhar, J., and Ohmoto, H., 2009, Anomalous fractionations of sulfur isotopes during thermochemical sulfate reduction: Science, v. 324, p. 370373, doi:10.1126/science.1169289. Weijers, J.W.H., Schouten, S., Spaargaren, O.C., and Sinninghe Damst, J.S., 2006, Occurrence and distribution of tetraether membrane lipids in soils: Implications for the use of the TEX86 proxy and the BIT index: Organic Geochemistry, v. 37, p. 16801693, doi:10.1016/j .orggeochem.2006.07.018. Welander, P.V., Hunter, R.C., Zhang, L., Sessions, A.L., Summons, R.E., and Newman, D.K., 2009, Hopanoids play a role in membrane integrity and pH homeostasis in Rhodopseudomonas palustris TIE-1: Journal of Bacteriology, v. 191, p. 61456156, doi:10.1128/JB.00460-09. Welander, P.V., Coleman, M.L., Sessions, A.L., Summons, R.E., and Newman, D.K., 2010, Identication of a methylase required for 2-methylhopanoid production and implications for the interpretation of sedimentary hopanes: Proceedings of the National Academy of Sciences of the United States of America, v. 107, p. 85378542, doi:10.1073/pnas.0912949107. Welander, P.V., Doughty, D.M., Wu, C.H., Mehay, S., Summons, R.E., and Newman, D.K., 2012, Identication and characterization of Rhodopseudomonas palustris TIE-1 hopanoid biosynthesis mutants: Geobiology, v. 10, p. 163177, doi:10.1111/j.1472-4669.2011.00314.x.

spe500-08 1st pgs 44 Johnson et al.

page 44

Weyer, S., Anbar, A.D., Gerdes, A., Gordon, G.W., Algeo, T.J., and Boyle, E.A., 2008, Natural fractionation of 238U/235U: Geochimica et Cosmochimica Acta, v. 72, p. 345359, doi:10.1016/j.gca.2007.11.012. White, A.F., and Brantley, S.L., 1995, Chemical weathering rates of silicate minerals: Reviews in Mineralogy, v. 31, p. 583. White, W.M., 2012, Geochemistry: New York, Wiley-Blackwell, 656 p. Wignall, P., and Hallam, A., 1992, Anoxia as a cause of the Permian/Triassic mass extinction: Facies evidence from northern Italy and the western United States: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 93, p. 2146. Williford, K.H., Van Kranendonk, M.J., Ushikubo, T., Kozdon, R., and Valley, J.W., 2011, Constraining atmospheric oxygen and seawater sulfate concentrations during Paleoproterozoic glaciation: In situ sulfur three-isotope microanalysis of pyrite from the Turee Creek Group, Western Australia: Geochimica et Cosmochimica Acta, v. 75, p. 56865705, doi:10.1016/j .gca.2011.07.010. Williford, K.H., Ushikubo, T., Schopf, J.W., Lepot, K., Kitajima, K., and Valley, J.W., 2013, Preservation and detection of microstructural and taxonomic correlations in the carbon isotopic compositions of individual Precambrian microfossils: Geochimica et Cosmochimica Acta, v. 104, p. 165 182, doi:10.1016/j.gca.2012.11.005. Winter, B.L., and Knauth, L.P., 1992, Stable isotope geochemistry of cherts and carbonates from the 2.0 Ga Gunint Iron Formation: Implications for the depositional setting, and the effects of diagenesis and metamorphism: Precambrian Research, v. 59, p. 283313, doi:10.1016/0301-9268(92)90061-R. Wood, J.A., and Hashimoto, A., 1993, Mineral equilibrium in fractionated nebular systems: Geochimica et Cosmochimica Acta, v. 57, p. 23772388, doi:10.1016/0016-7037(93)90575-H. Wood, J.A., Dickey, J.S., Marvin, U.B., and Poweel, B.N., Lunar anorthosites and a geophysical model of the Moon, in Levinson, A.A., ed., Proceedings of the Apollo 11 Lunar Science Conference 1970; Volume 1: Mineralogy and Petrology: New York, Pergamon Press, p. 965988. Wuchter, C., Schouten, S., Wakeham, S.G., and Sinninghe Damst, J.S., 2005, Temporal and spatial variation in tetraether membrane lipids of marine Crenarchaeota in particulate organic matter: Implications for TEX86 paleothermometry: Paleoceanography, v. 20, PA3013, doi:10.1029/2004PA001110. Xie, S., Pancost, R.D., Huang, J., Wignall, P.B., Yu, J., Tang, X., Chen, L., Huang, X., and Lai, X., 2007, Changes in the global carbon cycle

occurred as two episodes during the Permian-Triassic crisis: Geology, v. 35, p. 10831086, doi:10.1130/G24224A.1. Yamaguchi, K.E., Johnson, C.M., Beard, B.L., and Ohmoto, H., 2005, Biogeochemical cycling of iron in the ArcheanPaleoproterozoic Earth: Constraints from iron isotope variations in sedimentary rocks from the Kaapvaal and Pilbara cratons: Chemical Geology, v. 218, p. 135169, doi:10.1016/j.chemgeo.2005.01.020. Yoneda, S., and Grossman, L., 1995, Condensation of CaO-MgO-Al2O3-SiO2 liquids from cosmic gases: Geochimica et Cosmochimica Acta, v. 59, p. 34133444, doi:10.1016/0016-7037(95)00214-K. Young, E.D., and Galy, A., 2004, The isotope geochemistry and cosmochemistry of magnesium: Reviews in Mineralogy and Geochemistry, v. 55, p. 197230, doi:10.2138/gsrmg.55.1.197. Yuen, G., Blair, N., Des Marais, D.J., and Chang, S., 1984, Carbon isotope composition of low molecular weight hydrocarbons and monocarboxylic acids from Murchison meteorite: Nature, v. 307, p. 252254, doi:10.1038/307252a0. Zachos, J.C., Schouten, S., Bohaty, S., Quattlebaum, T., Sluijs, A., Brinkhuis, H., Gibbs, S.J., and Bralower, T.J., 2006, Extreme warming of midlatitude coastal ocean during the Paleocene-Eocene thermal maximum: Inferences from TEX86 and isotope data: Geology, v. 34, p. 737740, doi:10.1130/G22522.1. Zahnle, K., Claire, M., and Catling, D., 2006, The loss of mass-independent fractionation in sulfur due to a Palaeoproterozoic collapse of atmospheric methane: Geobiology, v. 4, p. 271283, doi:10.1111/j.1472 -4669.2006.00085.x. Zartman, R.E., and Doe, B.R., 1981, PlumbotectonicsThe model: Tectonophysics, v. 75, p. 135162, doi:10.1016/0040-1951(81)90213-4. Zindler, A., and Hart, S.R., 1986, Chemical geodynamics: Annual Review of Earth and Planetary Sciences, v. 14, p. 493571, doi:10.1146/annurev .ea.14.050186.002425. Zindler, A., Jagoutz, E., and Goldstein, S., 1982, Nd, Sr and Pb isotopic systematics in a three-component mantle: A new perspective: Nature, v. 298, p. 519523, doi:10.1038/298519a0. Zinner, E., 2004, Presolar grains, in Davis, A.M., ed., Treatise on Geochemistry; Volume 1: Meteorites, Comets, and Planets: Oxford, UK, Elsevier, p. 1739. MANUSCRIPT ACCEPTED BY THE SOCIETY 11 FEBRUARY 2013

Printed in the USA

Das könnte Ihnen auch gefallen