Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-27T09:14:51.871Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  04 May 2017

David C. Catling
Affiliation:
University of Washington
James F. Kasting
Affiliation:
Pennsylvania State University
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2017

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abbot, D. S., et al. (2012). Indication of insensitivity of planetary weathering behavior and habitable zone to surface land fraction. Astrophys. J. 756, 178.Google Scholar
Abbot, D. S. and Pierrehumbert, R. T. (2010). Mudball: Surface dust and Snowball Earth deglaciation. J. Geophys. Res. 115, D03104, doi: 10.1029/2009JD012007.Google Scholar
Abbot, D. S., et al. (2011). The Jormungand global climate state and implications for Neoproterozoic glaciations. J. Geophys. Res. 116, D18103, doi:10.1029/2011JD015927.Google Scholar
Abe, Y. (2011). Protoatmospheres and surface environment of protoplanets. Earth Moon Planets 108, 914.CrossRefGoogle Scholar
Abe, Y., et al. (2011). Habitable zone limits for dry planets. Astrobiology 11, 443460.Google Scholar
Abe, Y. and Matsui, T. (1988). Evolution of an impact-generated H2O-CO2 atmosphere and formation of a hot proto-ocean on Earth. J. Atmos. Sci. 45, 30813101.2.0.CO;2>CrossRefGoogle Scholar
Abelson, P. H. (1966). Chemical events on the primitive Earth. Proc. Nat. Acad. Sci. 55, 1365.CrossRefGoogle ScholarPubMed
Achterberg, R. K., et al. (2008). Titan’s middle-atmospheric temperatures and dynamics observed by the Cassini Composite Infrared Spectrometer. Icarus 194, 263277.Google Scholar
Achterberg, R. K., et al. (2011). Temporal variations of Titan’s middle-atmospheric temperatures from 2004 to 2009 observed by Cassini/CIRS. Icarus 211, 686698.Google Scholar
Ackiss, S. E. and Wray, J. (2014). Occurrences of possible hydrated sulfates in the southern high latitudes of Mars. Icarus 243, 311324.CrossRefGoogle Scholar
Acuna, M. H., et al. (1999). Global distribution of crustal magnetization discovered by the Mars Global Surveyor MAG/ER experiment. Science 284, 790793.CrossRefGoogle ScholarPubMed
Adams, E. Y. (2006). Titan’s thermal structure and the formation of a nitrogen atmosphere. University of Michigan, Ph.D. thesis, Ann Arbor, MI.Google Scholar
Agee, C. B., et al. (2013). Unique meteorite from early Amazonian Mars: Water-rich basaltic breccia Northwest Africa 7034. Science 339, 780785.Google Scholar
Agnor, C. and Asphaug, E. (2004). Accretion efficiency during planetary collisions. Astrophys. J. 613, L157L160.Google Scholar
Agnor, C. B. and Hamilton, D. P. (2006). Neptune’s capture of its moon Triton in a binary-planet gravitational encounter. Nature 441, 192194.CrossRefGoogle Scholar
Agol, E., et al. (2010). The climate of HD 189733b from fourteen transits and eclipses measured by Spitzer. Ap. J. 721, 18611877.CrossRefGoogle Scholar
Agol, E., et al. (2005). On detecting terrestrial planets with timing of giant planet transits. Mon. Not. R. Astron. Soc. 359, 567579.Google Scholar
Aharonson, O., et al. (2009). An asymmetric distribution of lakes on Titan as a possible consequence of orbital forcing. Nat. Geosci. 2, 851854.CrossRefGoogle Scholar
Aharonson, O., et al. (2002). Drainage basins and channel incision on Mars. P. Natl. Acad. Sci. USA 99, 17801783.CrossRefGoogle ScholarPubMed
Ahrens, T. J. (1993). Impact erosion of terrestrial planetary atmospheres. Annu. Rev. Earth Planet. Sci. 21, 525555.CrossRefGoogle Scholar
Alexander, B., et al. (2003). East Antarctic ice core sulfur isotope measurements over a complete glacial–interglacial cycle. J. Geophys. Res. 108, 4786, doi:10.1029/2003JD003513.Google Scholar
Alexander, R. D., et al. (2006). Photoevaporation of protoplanetary discs - II. Evolutionary models and observable properties. Mon. Not. R. Astron. Soc. 369, 229239.Google Scholar
Alfimova, N. A., et al. (2011). Mobility of cerium in the 2.8–2.1 Ga exogenous environments of the Baltic Shield: data on weathering profiles and sedimentary carbonates. Lithol. Miner. Resour. 46, 397408.CrossRefGoogle Scholar
Alibert, Y. and Mousis, O. (2007). Formation of Titan in Saturn’s subnebula: constraints from Huygens probe measurements. Astron. Astrophys. 465, 10511060.CrossRefGoogle Scholar
Allard, P. (1997). Endogenous magma degassing and storage at Mount Etna. Geophys. Res. Lett. 24, 22192222.Google Scholar
Allegre, C. J., et al. (1987). Rare gas systematics: Formation of the atmosphere, evolution and structure of the Earths mantle. Earth Planet. Sci. Lett. 81, 127150.Google Scholar
Allen, M. and Frederick, J. E. (1982). Effective photo-dissociation cross sections for molecular oxygen and nitric oxide in the Schumann–Runge bands. J. Atmos. Sci. 39, 20662075.2.0.CO;2>CrossRefGoogle Scholar
Allen, P. A. and Etienne, J. L. (2008). Sedimentary challenge to Snowball Earth. Nature Geosc. 1, 817825.Google Scholar
Allwood, A. C., et al. (2009). Controls on development and diversity of Early Archean stromatolites. P. Natl. Acad. Sci. USA 106, 95489555.Google Scholar
Allwood, A. C., et al. (2006). Stromatolite reef from the Early Archaean era of Australia. Nature 441, 714718.Google Scholar
ALMA-Partnership, et al. (2015). The 2014 ALMA Long Baseline Campaign: First Results from High Angular Resolution Observations toward the HL Tau Region. Astrophys. J. Lett. 808, L3, doi:10.1088/2041-8205/808/1/L3.Google Scholar
Alt, J. C. (1995). Sulfur isotopic profile through the oceanic crust: Sulfur mobility and seawater-crustal sulfur exchange during hydrothermal alteration. Geology 23, 585588.Google Scholar
Altabet, M. A. and Francois, R. (1994). Sedimentary nitrogen isotopic ratio as a recorder for surface ccean nitrate utilization. Global Biogeochemical Cycles 8, 103116.Google Scholar
Altermann, W. and Schopf, J. W. (1995). Microfossils from the Neoarchean Campbell Group, Griqualand West Sequence of the Transvaal Supergroup, and their paleoenvironmental and evolutionary Implications. Precambrian Res. 75, 6590.CrossRefGoogle ScholarPubMed
Altwegg, K. et al. (2015). 67P/Churyumov–Gerasimenko, a Jupiter family comet with a high D/H ratio. Science 347, doi: 10.1126/science.1261952.CrossRefGoogle ScholarPubMed
Amelin, Y., et al. (2010). U–Pb chronology of the Solar System’s oldest solids with variable 238U/235U. Earth Planet. Sci. Lett. 300, 343350.Google Scholar
Anbar, A. D., et al. (2007). A whiff of oxygen before the Great Oxidation Event? Science 317, 19031906.Google Scholar
Anbar, A. D. and Knoll, A. H. (2002). Proterozoic ocean chemistry and evolution: a bioinorganic bridge? Science 297, 11371142.Google Scholar
Anbar, A. D. and Rouxel, O. (2007). Metal stable isotopes in paleoceanography. Annu. Rev. Earth Planet. Sci. 35, 717746.Google Scholar
Anders, E. and Grevesse, N. (1989). Abundances of the elements – meteoritic and solar. Geochim. Cosmochim. Acta 53, 197214.CrossRefGoogle Scholar
Anderson, D. E. (1974). Mariner 6, 7, and 9 ultraviolet spectrometer experiment: Analysis of hydrogen Lyman alpha data. J. Geophys. Res. 79, 15131518.Google Scholar
Anderson, D. E. and Hord, C. W. (1971). Mariner 6 and Mariner 7 ultraviolet spectrometer experiment: analysis of hydrogen Lyman-alpha data. J. Geophys. Res. 76, 66666673.Google Scholar
Anderson, F. S., et al. (1999). Assessing the Martian surface distribution of aeolian sand using a Mars general circulation model. J. Geophys. Res. 104, 18 99119 002.Google Scholar
Anderson, G. M. (2005). Thermodynamics of Natural Systems. New York: Cambridge University Press.Google Scholar
Anderson, G. M. and Crerar, D. A. (1993). Thermodynamics in Geochemistry: The Equilibrium Model. New York: Oxford University Press.Google Scholar
Andre, M. J. (2011). Modelling 18O2 and 16O2 unidirectional fluxes in plants: I. Regulation of pre-industrial atmosphere. Biosystems 103, 239251.CrossRefGoogle ScholarPubMed
Andrews, D. G. (2010). An Introduction to Atmospheric Physics. New York: Cambridge University Press.CrossRefGoogle Scholar
Andrews, D. G., et al. (1987). Middle Atmosphere Dynamics. Orlando: Academic Press.Google Scholar
Andrews-Hanna, J. C., et al. (2007). Meridiani Planum and the global hydrology of Mars. Nature 446, 163166.Google Scholar
Andrews-Hanna, J. C., et al. (2010). Early Mars hydrology: Meridiani playa deposits and the sedimentary record of Arabia Terra. Journal of Geophysical Research-Planets 115.Google Scholar
Anglada-Escude, G., et al. (2016). A terrestrial planet candidate in a temperate orbit around Proxima Centauri. Nature 536, 437440.CrossRefGoogle Scholar
Ansan, V., et al. (2011). Stratigraphy, mineralogy, and origin of layered deposits inside Terby crater, Mars. Icarus 211, 273304.CrossRefGoogle Scholar
Archer, C. and Vance, D. (2006). Coupled Fe and S isotope evidence for Archean microbial Fe(III) and sulfate reduction. Geology 34, 153156.Google Scholar
Archer, D. (2005). Fate of fossil fuel CO2 in geologic time. J. Geophys. Res. 110.Google Scholar
Armstrong, B. H. (1968). Theory of diffusivity factor for atmospheric radiation. J. Quant. Spectros. Radiat. Transfer 8, 15771599.CrossRefGoogle Scholar
Armstrong, J. C. and Leovy, C. B. (2005). Long term wind erosion on Mars. Icarus 176, 5774.Google Scholar
Armstrong, J. C., et al. (2004). A 1 Gyr climate model for Mars: new orbital statistics and the importance of seasonally resolved polar processes. Icarus 171, 255271.Google Scholar
Arnold, G. L., et al. (2004). Molybdenum isotope evidence for widespread anoxia in mid-Proterozoic oceans. Science 304, 8790.Google Scholar
Arnold, L., et al. (2002). A test for the search for life on extrasolar planets-Looking for the terrestrial vegetation signature in the Earthshine spectrum. Astron. Astrophys. 392, 231237.CrossRefGoogle Scholar
Artemieva, N. and Lunine, J. I. (2005). Impact cratering on Titan - II. Global melt, escaping ejecta, and aqueous alteration of surface organics. Icarus 175, 522533.CrossRefGoogle Scholar
Arthur, M. A. (2000). Volcanic contributions to the carbon and sulfur geochemical cycles and global change. In: Encyclopedia of Volcanoes, ed. Sigurdsson, H., Academic Press, pp. 10451056.Google Scholar
Arthur, M. A., et al. (1994). Varve-calibrated records of carbonate and organic carbon accumulation over the last 2000 years in the Black Sea. Glob. Biogeochem. Cycles 8, 195217.Google Scholar
Arvidson, R. E., et al. (2014). Ancient aqueous environments at Endeavour Crater. Mars. Science 343, doi:10.1126/science.1248097.Google Scholar
Asael, D., et al. (2013). Coupled molybdenum, iron and uranium stable isotopes as oceanic paleoredox proxies during the Paleoproterozoic Shunga Event. Chem. Geol. 362, 193210.Google Scholar
Aston, F. W. (1924). The rarity of the inert gases on Earth. Nature 114, 786.CrossRefGoogle Scholar
Atkins, P. W. and Friedman, R. S. (2005). Molecular Quantum Mechanics. New York: Oxford University Press.Google Scholar
Atkinson, D. H., et al. (1997). Deep winds on Jupiter as measured by the Galileo probe. Nature 388, 649650.Google Scholar
Atreya, S. K. (1986). Atmospheres and Ionospheres of the Outer Planets and their Satellites. Heidelberg: Springer-Verlag.Google Scholar
Atreya, S. K. (2010). Atmospheric moons Galileo would have loved. Proc. IAU 6, 130140. doi:10.1017/S1743921310007349.Google Scholar
Atreya, S. K., et al. (2006). Titan’s methane cycle. Planet. Space Sci. 54, 11771187.Google Scholar
Atreya, S. K., et al. (1978). Evolution of a nitrogen atmosphere on Titan. Science 201, 611613.Google Scholar
Atreya, S. K., et al. (2009). Volatile origin and cycles: Nitrogen and methane. In: Titan from Cassini-Huygens, ed. Brown, R. H., et al., New York: Springer, pp. 7799.Google Scholar
Atreya, S. K., et al. (2003). Composition and origin of the atmosphere of Jupiter - an update, and implications for the extrasolar giant planets. Planet. Space Sci. 51, 105112.Google Scholar
Atreya, S. K., et al. (1991). Photochemistry and vertical mixing. In: Uranus, ed. Bergstralh, J. T., et al., Tucson, AZ: Univ. of Arizona Press, pp. 110146.Google Scholar
Atreya, S. K., et al. (2013). Primordial argon isotope fractionation in the atmosphere of Mars measured by the SAM instrument on Curiosity and implications for atmospheric loss. Geophys. Res. Lett. 40, 56055609.CrossRefGoogle Scholar
Atreya, S. K., et al. (1999). A comparison of the atmospheres of Jupiter and Saturn: deep atmospheric composition, cloud structure, vertical mixing, and origin. Planet. Space Sci. 47, 12431262.CrossRefGoogle ScholarPubMed
Aulbach, S. and Stagno, V. (2016). Evidence for a reducing Archean ambient mantle and its effects on the carbon cycle. Geology 44, 751754.Google Scholar
Awramik, S. M. and Barghoorn, E. S. (1977). Gunflint Microbiota. Precambrian Res. 5, 121142.Google Scholar
Axford, W. I. (1968). The polar wind and the terrestrial helium budget. J. Geophys. Res. 73, 68556859.Google Scholar
Bahcall, J. N., et al. (2001). Solar models: Current epoch and time dependences, neutrinos, and helioseismological properties. Ap. J. 555, 9901012.Google Scholar
Bailey, J. (2009). A comparison of water vapor line parameters for modeling the Venus deep atmosphere. Icarus 201, 444453.Google Scholar
Baker, N. L. and Leovy, C. B. (1987). Zonal winds near Venus cloud top level: a model study of the interaction between the zonal mean circulation and the semidiurnal tide. Icarus 69, 202220.Google Scholar
Baker, V. R. (1990). Spring sapping and valley network development, with case studies by R. C. Kochel, J. E. Laity, and A. D. Howard. In: Higgins, C. G., Coates, D. R., ed., Groundwater Geomorphology: The Role of Subsurface Water in Earth-Surface Processes and Landforms, Geol. Soc. Am. Spec. Pap., Vol. 252, pp. 235265.Google Scholar
Baker, V. R. (2001). Water and the martian landscape. Nature 412, 228236.Google Scholar
Baker, V. R. (2006). Geomorphological evidence for water on Mars. Elements 2, 139143.CrossRefGoogle Scholar
Baker, V. R., et al. (2015). Fluvial geomorphology on Earth-like planetary surfaces: A review. Geomorphology 245, 149182.Google Scholar
Baland, R.-M., et al. (2014). Titan’s internal structure inferred from its gravity field, shape, and rotation state. Icarus 237, 2941.Google Scholar
Balme, M. and Greeley, R. (2006). Dust devils on Earth and Mars. Rev. Geophys. 44, RG3003 doi:10.1029/2005RG000188.Google Scholar
Bandfield, J. L. (2002). Global mineral distributions on Mars. J. Geophys. Res. 107.Google Scholar
Bandfield, J. L., et al. (2003). Spectroscopic identification of carbonate minerals in the martian dust. Science 301, 10841087.CrossRefGoogle ScholarPubMed
Banks, P. M. and Kockarts, G. (1973). Aeronomy: Part B. New York: Academic Press.Google Scholar
Bannon, P. R., et al. (1997). Does the surface pressure equal the weight per unit area of a hydrostatic atmosphere? Bull. Am. Met. Soc. 78, 26372642.2.0.CO;2>CrossRefGoogle Scholar
Bao, H. M., et al. (2009). Stretching the envelope of past surface environments: Neoproterozoic glacial lakes from Svalbard. Science 323, 119122.Google Scholar
Bao, H. M., et al. (2008). Triple oxygen isotope evidence for elevated CO2 levels after a Neoproterozoic glaciation. Nature 453, 504506.Google Scholar
Barabash, S., et al. (2007). The loss of ions from Venus through the plasma wake. Nature 450, 650653.Google Scholar
Baraffe, I., et al. (1998). Evolutionary models for solar metallicity low-mass stars: Mass-magnitude relationships and color-magnitude diagrams. Astron. Astrophys. 337, 403412.Google Scholar
Baraffe, O., et al. (2002). Evolutionary models for low-mass stars and brown dwarfs: Uncertainties and limits at very young ages. Astron. Astrophys. 382, 563572.Google Scholar
Barakat, A. R. and Lemaire, J. (1990). Monte Carlo study of the escape of a minor species. Phys. Rev. A 42, 32913302.CrossRefGoogle ScholarPubMed
Barnes, R., et al. (2009). Tidal limits to planetary habitability. Astrophys. J. Lett. 700, L30L33.Google Scholar
Barnes, R., et al. (2013). Tidal Venuses: Triggering a climate catastrophe via tidal heating. Astrobiology 13, 225250.Google Scholar
Barnes, R., et al. (2010). CoRoT-7b: Super-Earth or Super-Io? Astrophys. J. Lett. 709, L95L98.Google Scholar
Barnhart, C. J., et al. (2009). Long-term precipitation and late-stage valley network formation: Landform simulations of Parana Basin, Mars. J. Geophys. Res. 114.Google Scholar
Barr, A. C. and Canup, R. M. (2010). Origin of the Ganymede-Callisto dichotomy by impacts during the late heavy bombardment. Nat. Geosci. 3, 164167.Google Scholar
Barr, A. C. and Citron, R. I. (2011). Scaling of melt production in hypervelocity impacts from high-resolution numerical simulations. Icarus 211, 913916.Google Scholar
Barth, C. A. (1974). Atmosphere of Mars. Annu. Rev. Earth Planet. Sci. 2, 333367.Google Scholar
Barth, C. A., et al. (1973). Mariner 9 ultraviolet spectrometer experiment: Seasonal variation of ozone on Mars. Science 179, 795796.Google Scholar
Barth, C. A., et al. (1972). Mariner 9 ultraviolet spectrometer experiment: Mars airglow spectroscopy and variations in Lyman alpha. Icarus 17, 457468.Google Scholar
Bartlett, D. H. (2002). Pressure effects on in vivo microbial processes. Biochim. Biophys. Acta 1595, 367381.Google Scholar
Batalha, N., et al. (2015). Testing the early Mars H2–CO2 greenhouse hypothesis with a 1-D photochemical model. Icarus 258, 337349.Google Scholar
Batalha, N. M. (2014). Exploring exoplanet populations with NASA’s Kepler Mission. P. Natl. Acad. Sci. USA 111, 1264712654.Google Scholar
Batalha, N. M., et al. (2013). Planetary candidates observed by Kepler. Iii. Analysis of the first 16 months of data. Astrophys. J. Supp. S. 204.Google Scholar
Bau, M., et al. (1997). Sources of rare-earth elements and iron in Paleoproterozoic iron-formations from the Transvaal Supergroup, South Africa: Evidence from neodymium isotopes. J. Geol. 105, 121129.Google Scholar
Baum, S. K. and Crowley, T. J. (2001). GCM response to late precambrian (similar to 590 Ma) ice-covered continents. Geophysical Research Letters 28, 583586.Google Scholar
Bean, J. L., et al. (2010). A ground-based transmission spectrum of the super-Earth exoplanet GJ 1214b. Nature 468, 669672.Google Scholar
Beaugé, C., et al. (2008). Planetary masses and orbital parameters from radial velocity measurements. In: Extrasolar Planets: Formation, Detection and Dynamics, Weinheim, Germany: Wiley-VCH Verlag GmbH & Co. KGaA, pp. 125.Google Scholar
Beaumont, V. and Robert, F. (1999). Nitrogen isotope ratios of kerogens in Precambrian cherts: A record of the evolution of atmosphere chemistry? Precambrian Res. 96, 6382.Google Scholar
Beck, P., et al. (2015). A Noachian source region for the “Black Beauty” meteorite, and a source lithology for Mars surface hydrated dust? Earth Planet. Sc. Lett. 427, 104111.CrossRefGoogle Scholar
Becker, R. H. and Clayton, R. N. (1972). Carbon isotopic evidence for origin of a Banded Iron Formation in Western Australia. Geochim. Cosmochim. Acta 36, 577.Google Scholar
Becker, R. H. and Pepin, R. O. (1984). The case for a martian origin of the Shergottites: Nitrogen and noble gases in EETA79001. Earth Planet. Sc. Lett. 69, 225242.Google Scholar
Beer, J. (2000). Long-term indirect indices of solar variability. Space Science Reviews 94, 5366.Google Scholar
Beerling, D. J. and Royer, D. L. (2011). Convergent Cenozoic CO2 history. Nat. Geosci. 4, 418420.Google Scholar
Béghin, C., et al. (2010). Titan’s native ocean revealed beneath some 45 km of ice by a Schumann-like resonance. C. R. Geosci. 342, 425433.Google Scholar
Bekker, A. and Holland, H. D. (2012). Oxygen overshoot and recovery during the early Paleoproterozoic. Earth Planet. Sc. Lett. 317, 295304.Google Scholar
Bekker, A., et al. (2004). Dating the rise of atmospheric oxygen. Nature 427, 117120.Google Scholar
Bekker, A., et al. (2008). Fractionation between inorganic and organic carbon during the Lomagundi (2.22–2.1 Ga) carbon isotope excursion. Earth Planet. Sci. Lett. 271, 278291.CrossRefGoogle Scholar
Bekker, A., et al. (2006). Carbon isotope record for the onset of the Lomagundi carbon isotope excursion in the Great Lakes area, North America. Precam. Res. 148, 145180.Google Scholar
Bekker, A., et al. (2010). Iron Formation: The sedimentary product of a complex interplay among mantle, tectonic, oceanic, and biospheric processes. Econ. Geol. 105, 467508.Google Scholar
Belcher, C. M. and McElwain, J. C. (2008). Limits for combustion in low O2 redefine paleoatmospheric predictions for the Mesozoic. Science 321, 11971200.CrossRefGoogle ScholarPubMed
Belcher, C. M., et al. (2010). Baseline intrinsic flammability of Earth’s ecosystems estimated from paleoatmospheric oxygen over the past 350 million years. P. Natl. Acad. Sci. USA 107, 22 44822 453.Google Scholar
Bell, D. R. and Anbar, A. D. (2013). Oxygen titration of continental lithosphere and the rise of atmospheric O2. American Geophysic Union Fall Mtg., p. V41E-05.Google Scholar
Bell, E. A., et al. (2015). Potentially biogenic carbon preserved in a 4.1 billion-year-old zircon. P. Natl. Acad. Sci. USA, doi: 10.1073/pnas.1517557112.CrossRefGoogle Scholar
Bell, J. M., et al. (2010a). Simulating the one-dimensional structure of Titan’s upper atmosphere: 1. Formulation of the Titan Global Ionosphere-Thermosphere Model and benchmark simulations. J. Geophys. Res. 115, E12002, doi:10.1029/2010JE003636.Google Scholar
Bell, J. M., et al. (2011). Simulating the one-dimensional structure of Titan’s upper atmosphere: 3. Mechanisms determining methane escape. J. Geophys. Res. 116.Google Scholar
Bell, J. M., et al. (2010b). Simulating the one-dimensional structure of Titan’s upper atmosphere: 2. Alternative scenarios for methane escape. J. Geophys. Res. 115.Google Scholar
Bell, J. M., et al. (2014). Developing a self-consistent description of Titan’s upper atmosphere without hydrodynamic escape. J. Geophys. Res. 119, 49574972.Google Scholar
Belton, M. J. S., et al. (1991). Images from Galileo of the Venus cloud deck. Science 253, 15311536.Google Scholar
Belton, M. J. S., et al. (1992). The Galileo Solid-State Imaging experiment. Space Sci. Rev. 60, 413455.Google Scholar
Belyaev, D. A., et al. (2012). Vertical profiling of SO2 and SO above Venus’ clouds by SPICAV/SOIR solar occultations. Icarus 217, 740751.Google Scholar
Ben-Jaffel, L. (2007). Exoplanet HD 209458b: Inflated hydrogen atmosphere but no sign of evaporation. Astrophysical Journal Letters 671, L61L64.Google Scholar
Ben-Jaffel, L. (2008). Spectral, spatial, and time properties of the hydrogen nebula around exoplanet HD 209458b. Astrophys. J. 688, 13521360.Google Scholar
Benedict, G. F., et al. (1999). Interferometric astrometry of Proxima Centauri and Barnard’s star using Hubble Space Telescope Fine Guidance Sensor 3: Detection limits for substellar companions. Astron. J. 118, 10861100.Google Scholar
Bengtson, S. (1994). The advent of animal skeletons. In: Early Life on Earth, ed. Bengtson, S., New York: Columbia University Press, pp. 412425.Google Scholar
Benner, S. A. (2009). Life, the Universe…and the Scientific Method. Gainesville, FL: FfAME Press.Google Scholar
Benner, S. A. (2010). Defining life. Astrobiology 10, 10211030.Google Scholar
Benner, S. A., et al. (2004). Is there a common chemical model for life in the universe? Curr. Opin. Chem. Biol. 8, 672689.Google Scholar
Bent, H. A. (1965). The Second Law: An Introduction to Classical and Statistical Thermodynamics. New York: Oxford University Press.Google Scholar
Benton, M. J. and Ayala, F. J. (2003). Dating the tree of life. Science 300, 16981700.Google Scholar
Bergin, E. A., et al. (2015). Tracing the ingredients for a habitable Earth from interstellar space through planet formation. P. Natl. Acad. Sci. USA 112, 89658970.Google Scholar
Bergman, N. M., et al. (2004). COPSE: A new model of biogeochemical cycling over Phanerozoic time. Am. J. Sci. 304, 397437.Google Scholar
Berkner, L. V. and Marshall, L. C. (1965). On the origin and rise of oxygen concentration in the Earth’s atmosphere. J. Atmos. Sci. 22, 225261.Google Scholar
Berkner, L. V. and Marshall, L. L. (1964). The history of oxygenic concentration in the Earth’s atmosphere. Disc. Faraday Soc. 34, 122141.CrossRefGoogle Scholar
Berkner, L. V. and Marshall, L. L. (1966). Limitation on oxygen concentration in a primitive planetary atmosphere. J. Atmos. Sci. 23, 133143.Google Scholar
Berkner, L. V. and Marshall, L. L. (1967). The rise of oxygen in the Earth’s atmosphere with notes on the martian atmosphere. Adv. Geophys. 12, 309331.Google Scholar
Bernal, J. D. (1951). The Physical Basis of Life. London: Routledge and Paul.Google Scholar
Berndt, M. E., et al. (1996). Reduction of CO2 during serpentinization of olivine at 300 ºC and 500 bar. Geology 24, 351354.Google Scholar
Berner, R. A. (1982). Burial of organic-carbon and pyrite sulfur in the modern ocean - Its geochemical and environmental significance. Am. J. Sci. 282, 451473.Google Scholar
Berner, R. A. (2004). The Phanerozoic Carbon Cycle: CO2 and O2. Oxford: Oxford University Press.Google Scholar
Berner, R. A. (2006a). GEOCARBSULF: A combined model for Phanerozoic atmospheric O2 and CO2. Geochim. Cosmochim. Acta 70, 56535664.Google Scholar
Berner, R. A. (2006b). Geological nitrogen cycle and atmospheric N2 over Phanerozoic time. Geology 34, 413415.Google Scholar
Berner, R. A. (2009). Phanerozoic atmospheric oxygen: new results using the Geocarbsulf Model. Am. J. Sci. 309, 603606.Google Scholar
Berner, R. A., et al. (2003). Phanerozoic atmospheric oxygen. Ann. Rev. Earth Planet. Sci. 31, 105134.Google Scholar
Berner, R. A. and Canfield, D. E. (1989). A new model for atmospheric oxygen over Phanerozoic time. Amer. J. Sci. 289, 333361.Google Scholar
Berner, R. A., et al. (1983). The carbonate-silicate geochemical cycle and its effect on atmospheric carbon dioxide over the past 100 million years. Amer. J. Sci. 283, 641683.Google Scholar
Berner, R. A. and Maasch, K. A. (1996). Chemical weathering and controls on atmospheric O2 and CO2: Fundamental principles were enunciated by J. J. Ebelmen in 1845. Geochim. Cosmochim. Acta 60, 16331637.Google Scholar
Berner, R. A. and Raiswell, R. (1983). Burial of organic-carbon and pyrite sulfur in sediments over Phanerozoic time – A new theory. Geochim. Cosmochim. Acta 47, 855862.Google Scholar
Berresheim, H. and Jaeschke, W. (1983). The contribution of volcanoes to the global atmospheric sulfur budget. J. Geophys. Res. 88, 37323740.Google Scholar
Berta, Z. K., et al. (2012). Transit detection in the MEarth survey of nearby M dwarfs: Bridging the clean-first, search-later divide. Astronom. J. 144, 145.Google Scholar
Berta-Thompson, Z. K., et al. (2015). A rocky planet transiting a nearby low-mass star. Nature 527, 204207.Google Scholar
Bertaux, J. L. (1975). Observed variations of the exospheric hydrogen density with the exospheric temperature. J. Geophys. Res. 80, 639642.Google Scholar
Bertaux, J. L., et al. (1978). Lyman-alpha observations of Venera-9 and Venera-10 .1. Nonthermal hydrogen population in exosphere of Venus. Planet. Space Sci. 26, 817831.Google Scholar
Bertaux, J. L., et al. (1982). Altitude profile of H in the atmosphere of Venus from Lyman alpha observations of Venera 11 and Venera 12 and origin of the hot exospheric component. Icarus 52, 221244.Google Scholar
Bertaux, J. L., et al. (1996). VEGA 1 and VEGA 2 entry probes: An investigation of local UV absorption (220–400 nm) in the atmosphere of Venus (SO2, aerosols, cloud structure). J. Geophys. Res. 101, 12 70912 745.Google Scholar
Bessell, M. S. (2005). Standard photometric systems. Annu. Rev. Astron. Astrophys. 43, 293336.Google Scholar
Bethe, H. A. (1939). Energy production in stars. Phys. Rev. 55, 01030103.Google Scholar
Betts, J. N. and Holland, H. D. (1991). The oxygen-content of ocean bottom waters, the burial efficiency of organic-carbon, and the regulation of atmospheric oxygen. Global Planet. Change 97, 518.Google Scholar
Bézard, B., et al. (1990). The deep atmosphere of Venus revealed by high-resolution nightside spectra. Nature 345, 508511.Google Scholar
Bezard, B., et al. (2009). Water vapor abundance near the surface of Venus from Venus Express/VIRTIS observations. J. Geophys. Res. 114.Google Scholar
Bezard, B., et al. (2014). The composition of Titan’s atmosphere. In: Titan: Surface, Atmosphere and Magnetosphere, ed. Muller-Wodarg, I., et al., New York: Cambridge University Press, pp. 158189.Google Scholar
Bibring, J. P., et al. (2007). Coupled ferric oxides and sulfates on the Martian surface. Science 317, 12061210.Google Scholar
Bibring, J. P., et al. (2005). Mars surface diversity as revealed by the OMEGA/Mars Express observations. Science 307, 15761581.Google Scholar
Bibring, J. P., et al. (2006). Global mineralogical and aqueous mars history derived from OMEGA/Mars express data. Science 312, 400404.Google Scholar
Bickle, M. J. (1986). Implications of melting for stabilization of the lithosphere and heat loss in the Archean. Earth Planet. Sci. Lett. 80, 314324.Google Scholar
Biemann, K. and Bada, J. (2011). Comment on “Reanalysis of the Viking results suggests perchlorate and organics at midlatitudes on Mars” by Rafael Navarro-Gonzalez et al. J. Geophys. Res. 116, E12001.Google Scholar
Birch, F. (1964). Density and composition of mantle and core. J. Geophys. Res. 69, 43774388.Google Scholar
Bird, G. A. (1994). Molecular gas dynamics and the direct simulation of gas flows. New York: Oxford University Press.Google Scholar
Bird, M. K., et al. (2005). The vertical profile of winds on Titan. Nature 438, 800802.Google Scholar
Bishop, J. L., et al. (2009). Mineralogy of Juventae Chasma: Sulfates in the light-toned mounds, mafic minerals in the bedrock, and hydrated silica and hydroxylated ferric sulfate on the plateau. J. Geophys. Res. 114, E00D09, doi:10.1029/2009JE003352.Google Scholar
Biver, N., et al. (2012). Ammonia and other parent molecules in comet 10P/Tempel 2 from Herschel/HIFI and ground-based radio observations. Astron. Astrophys, 539.Google Scholar
Bjerrum, C. J. and Canfield, D. E. (2002). Ocean productivity before about 1.9 Gyr ago limited by phosphorus adsorption onto iron oxides. Nature 417, 159162.Google Scholar
Bjerrum, C. J. and Canfield, D. E. (2004). New insights into the burial history of organic carbon on the early Earth. Geochem. Geophys. Geosys. 5, Q08001, doi:10.1029/2004GC000713.Google Scholar
Bjerrum, C. J. and Canfield, D. E. (2011). Towards a quantitative understanding of the late Neoproterozoic carbon cycle. P. Natl. Acad. Sci. USA 108, 55425547.Google Scholar
Blake, G. A. and Bergin, E. A. (2015). Prebiotic chemistry on the rocks. Nature 520, 161162.Google Scholar
Blake, R. E., et al. (2010). Phosphate oxygen isotopic evidence for a temperate and biologically active Archaean ocean. Nature 464, 1029–U89.Google Scholar
Blank, C. E. and Sanchez-Baracaldo, P. (2010). Timing of morphological and ecological innovations in the cyanobacteria – a key to understanding the rise in atmospheric oxygen. Geobiology 8, 123.Google Scholar
Bockelee-Morvan, D., et al. (2008). Large excess of heavy nitrogen in both hydrogen cyanide and cyanogen from comet 17P/Holmes. Astrophys. J. Lett. 679, L49L52.Google Scholar
Bogard, D. D. (1995). Impact ages of meteorites: A synthesis. Meteoritics 30, 244268.Google Scholar
Bogard, D. D. (1997). A reappraisal of the Martian Ar-36/Ar-38 ratio. J. Geophys. Res. 102, 16531661.Google Scholar
Bogard, D. D., et al. (2001). Martian volatiles: Isotopic composition, origin, and evolution. Space Sci. Rev. 96, 425458.Google Scholar
Bogard, D. D. and Garrison, D. H. (1998). Relative abundances of argon, krypton, and xenon in the Martian atmosphere as measured in Martian meteorites. Geochim. Cosmochim. Acta 62, 18291835.Google Scholar
Bogard, D. D. and Johnson, P. (1983). Martian gases in an Antarctic meteorite. Science 221, 651654.Google Scholar
Bogard, D. D. and Park, J. (2008). Ar-39–Ar-40 dating of the Zagami Martian shergottite and implications for magma origin of excess Ar-40. Meteorit. Planet. Sci. 43, 11131126.Google Scholar
Bohren, C. F. and Albrecht, B. A. (1998). Atmospheric Thermodynamics. New York: Oxford University Press.Google Scholar
Bohren, C. F. and Fraser, A. B. (1985). Colors of the sky. Phys. Teach. 23, 267272.Google Scholar
Bondi, H. (1952). On spherically symmetrical accretion. Mon. Not. R. Astron. Soc. 112, 195204.Google Scholar
Bonfils, X., et al. (2013). The HARPS search for southern extra-solar planets XXXI. The M-dwarf sample. Astron. Astrophys. 549, A109, doi: 10.1051/0004-6361/201014704.Google Scholar
Bonnefoy, M., et al. (2011). High angular resolution detection of β Pictoris b at 2.18 μm. Astron. Astrophys, 528.Google Scholar
Boone, R. H., et al. (2006). Metallicity in the solar neighborhood out to 60 pc. New Astronomy Reviews 50, 526529.Google Scholar
Boothroyd, A. I., et al. (1991). Our Sun II. Early mass loss of 0.1 Mo and the case of the missing lithium. Ap. J. 377, 318329.Google Scholar
Bordoni, S. and Schneider, T. (2008). Monsoons as eddy-mediated regime transitions of the tropical overturning circulation. Nat. Geosci. 1, 515519.Google Scholar
Bordoni, S. and Schneider, T. (2010). Regime transitions of steady and time-dependent Hadley circulations: Comparison of axisymmetric and eddy-permitting simulations. J. Atmos. Sci. 67, 16431654.Google Scholar
Borg, L. and Drake, M. J. (2005). A review of meteorite evidence for the timing of magmatism and of surface or near-surface liquid water on Mars. J. Geophys. Res. 110, E12S03, doi: 10.1029/2005JE002402.Google Scholar
Borucki, W. J. and Chameides, W. L. (1984). Lightning: Estimates of the rates of energy dissipation and nitrogen fixation. Rev. Geophys. 22, 363372.Google Scholar
Borucki, W. J., et al. (2011). Characteristics of planetary candidates observed by Kepler. I. Analysis of the first four months of data. Astrophys. J. 736, 19, doi: 10.1088/0004-637X/736/1/19.Google Scholar
Bosak, T., et al. (2009). Morphological record of oxygenic photosynthesis in conical stromatolites. P. Natl. Acad. Sci. USA 106, 10 93910 943.Google Scholar
Boss, A. P. (2005). Evolution of the solar nebula. VII. Formation and survival of protoplanets formed by disk instability. Astrophys. J. 629, 535548.Google Scholar
Boss, A. P. (2006). Gas giant protoplanets formed by disk instability in binary star systems. Astrophys. J. 641, 11481161.Google Scholar
Boss, A. P. (2008). Flux-limited diffusion approximation models of giant planet formation by disk instability. Astrophys. J. 677, 607615.Google Scholar
Boss, A. P. (2012). Giant planet formation by disc instability: flux-limited radiative diffusion and protostellar wobbles. Mon. Not. R. Astron. Soc. 419, 19301936.Google Scholar
Boss, A. P. and Ciesla, F. J. (2014). The solar nebula. In: Treatise on Geochemistry, ed. Holland, H. D., Turekian, K. K., New York: Elsevier, pp. 123.Google Scholar
Bottke, W. F., et al. (1995). Collisional lifetimes and impact statistics of near-Earth asteroids. In: Hazards Due to Comets and Asteroids, ed. Gehrels, T., Tucson: University of Arizona Press, pp. 337357.Google Scholar
Bottke, W. F., et al. (2015). Dating the Moon-forming impact event with asteroidal meteorites. Science 348, 321323.Google Scholar
Bottke, W. F., et al. (2012). An Archaean heavy bombardment from a destabilized extension of the asteroid belt. Nature 485, 7881.Google Scholar
Bougher, S. W., et al. (2000). Comparative terrestrial planet thermospheres 3. Solar cycle variation of global structure and winds at solstices. J. Geophys. Res. 105, 17 66917 692.Google Scholar
Bougher, S. W., et al. (2009). Solar cycle variability of Mars dayside exospheric temperatures: Model evaluation of underlying thermal balances. Geophysical Research Letters 36.Google Scholar
Boussau, B., et al. (2008). Parallel adaptations to high temperatures in the Archaean eon. Nature 456, 942–U74.Google Scholar
Bouvier, A., et al. (2009). Martian meteorite chronology and the evolution of the interior of Mars. Earth Planet. Sc. Lett. 280, 285295.Google Scholar
Bowring, S. A., et al. (2007). Geochronologic constraints on the chronostratigraphic framework of the neoproterozoic Huqf Supergroup, Sultanate of Oman. American Journal of Science 307, 10971145.Google Scholar
Boyd, E. S., et al. (2011). A late methanogen origin for molybdenum-dependent nitrogenase. Geobiology 9, 221232.Google Scholar
Boyd, E. S. and Peters, J. W. (2013). New insights into the evolutionary history of biological nitrogen fixation. Front. Microbiol. 4.Google Scholar
Boyle, R. A., et al. (2007). Neoproterozoic ‘snowball Earth’ glaciations and the evolution of altruism. Geobiology 5, 337349.Google Scholar
Boynton, W. V., et al. (2009). Evidence for calcium carbonate at the Mars Phoenix landing site. Science 325, 6164.Google Scholar
Brain, D. A. and Jakosky, B. M. (1998). Atmospheric loss since the onset of the Martian geologic record: Combined role of impact erosion and sputtering. J. Geophys. Res. 103, 22 68922 694.Google Scholar
Brantley, S. L. and Koepenick, K. W. (1995). Measured carbon dioxide emissions from Oldoinyo-Lengai and the skewed distribution of passive volcanic fluxes. Geology 23, 933936.Google Scholar
Brasier, M., et al. (2004). Earth’s oldest (similar to 3.5 Ga) fossils and the ‘Early Eden hypothesis’: Questioning the evidence. Orig. Life Evol. Biosph. 34, 257269.Google Scholar
Brasier, M. D., et al. (2015). Changing the picture of Earth’s earliest fossils (3.5–1.9 Ga) with new approaches and new discoveries. P. Natl. Acad. Sci. USA 112, 48594864.Google Scholar
Brasier, M. D., et al. (2002). Questioning the evidence for the Earth’s oldest fossils. Nature 416, 7681.Google Scholar
Brasier, M. D., et al. (2005). Critical testing of Earth’s oldest putative fossil assemblage from the similar to 3.5 Ga Apex Chert, Chinaman Creek, western Australia. Precam. Res. 140, 55102.Google Scholar
Brasier, M. D. and Lindsay, J. F. (1998). A billion years of environmental stability and the emergence of eukaryotes: New data from northern Australia. Geology 26, 555558.Google Scholar
Brasier, M. D., et al. (2011). Pumice as a remarkable substrate for the origin of life. Astrobiology 11, 725735.Google Scholar
Brass, G. W. (1980). Stability of brines on Mars. Icarus 42, 2028.Google Scholar
Brasseur, G. and Solomon, S. (2005). Aeronomy of the Middle Atmosphere: Chemistry and Physics of the Stratosphere and Mesosphere. Dordrecht: Springer.Google Scholar
Braterman, P. S., et al. (1983). Photooxidation of hydrated Fe+2 – significance for banded iron formations. Nature 303, 163164.Google Scholar
Breuer, D. and Spohn, T. (2003). Early plate tectonics versus single-plate tectonics on Mars: Evidence from magnetic field history and crust evolution. J. Geophys. Res. 108.Google Scholar
Brewer, A. W. (1949). Evidence for a world circulation provided by the measurements of helium and water vapour distribution in the stratosphere. Q. J. Roy. Meteor. Soc. 75, 351363.Google Scholar
Bridges, J. C., et al. (2001). Alteration assemblages in martian meteorites: Implications for near-surface processes. Space Science Reviews 96, 365392.Google Scholar
Bridges, J. C. and Warren, P. H. (2006). The SNC meteorites: basaltic igneous processes on Mars. J. Geol. Soc. London 163, 229251.Google Scholar
Bridges, N. T., et al. (2004). Insights on rock abrasion and ventifact formation from laboratory and field analog studies with applications to Mars. Planet. Space Sci. 52, 199213.Google Scholar
Brinkman, R. T. (1969). Dissociation of water vapor and evolution of oxygen in the terrestrial atmosphere. J. Geophys. Res. 74, 53555368.Google Scholar
Brinkmann, R. T. (1970). Departure from Jeans escape rate for H and He in the Earth’s atmosphere. Planet. Space Sci. 18, 449478.Google Scholar
Brinkmann, R. T. (1971). More comments on the validity of Jeans escape rate. Planet. Space Sci. 19, 791794.Google Scholar
Bristow, T. F., et al. (2011). A hydrothermal origin for isotopically anomalous cap dolostone cements from south China. Nature (advance online publication).Google Scholar
Bristow, T. F. and Kennedy, M. J. (2008). Carbon isotope excursions and the oxidant budget of the Ediacaran atmosphere and ocean. Geology 36, 863866.Google Scholar
Brocks, J. J., et al. (2003). A reconstruction of Archean biological diversity based on molecular fossils from the 2.78 to 2.45 billion-year-old Mount Bruce Supergroup, Hamersley Basin, Western Australia. Geochim. Cosmochim. Acta 67, 43214335.Google Scholar
Brocks, J. J., et al. (1999). Archean molecular fossils and the early rise of eukaryotes. Science 285, 10331036.Google Scholar
Broda, E. (1975). Beginning of photosynthesis. Origins Life Evol. Biosph. 6, 247251.Google Scholar
Broda, E. (1977). Evolution of photosynthesis. Precambrian Res. 4, 117132.Google Scholar
Broecker, W. S. (2015). The collision that changed the world. Elem. Sci. Anth. 3, 000061, doi: 10.12952/journal.elementa.000061.Google Scholar
Broecker, W. S. and Peng, T. H. (1982). Tracers in the Sea. Palisades, New York: Lamont Doherty Geol. Obs.Google Scholar
Brown, G. N. and Ziegler, W. T. (1979). Vapor pressure and heats of sublimation of liquids and solids of interest in cryogenics below 1-atm pressure. In: Advances in Cryogenic Engineering, ed. Timmerhaus, K., Snyder, H. A., New York: Plenum Press, pp. 662670.Google Scholar
Brown, M. E. (1997). A search for a sodium atmosphere around Ganymede. Icarus 126, 236238.Google Scholar
Brown, M. E. (2001). Potassium in Europa’s atmosphere. Icarus 151, 190195.Google Scholar
Brown, M. E. (2012). The compositions of Kuiper Belt Objects. Ann. Rev. Earth Planet. Sci. 40, 467495.Google Scholar
Brown, M. E. and Calvin, W. M. (2000). Evidence for crystalline water and ammonia ices on Pluto’s satellite Charon. Science 287, 107109.Google Scholar
Brown, M. E. and Hill, R. E. (1996). Discovery of an extended sodium atmosphere around Europa. Nature 380, 229231.Google Scholar
Brown, M. E. and Schaller, E. L. (2007). The mass of dwarf planet Eris. Science 316, 15851585.Google Scholar
Brown, R. H., et al. (1991). Triton’s global heat budget. Science 251, 14651467.Google Scholar
Brown, R. H., et al. (1990). Energy-Sources for Tritons Geyser-Like Plumes. Science 250, 431435.Google Scholar
Brown, R. H., et al. (2009). Titan from Cassini-Huygens. New York: Springer.Google Scholar
Brown, T. M., et al. (2001). Hubble Space Telescope time-series photometry of the transiting planet of HD 209458. Astrophysical Journal 552, 699709.Google Scholar
Brunner, B., et al. (2013). Nitrogen isotope effects induced by anammox bacteria. P. Natl. Acad. Sci. USA 110, 18 99418 999.Google Scholar
Buchhave, L. A., et al. (2014). Three regimes of extrasolar planet radius inferred from host star metallicities. Nature 509, 593595.Google Scholar
Buchhave, L. A., et al. (2012). An abundance of small exoplanets around stars with a wide range of metallicities. Nature 486, 375377.Google Scholar
Budd, G. E. (2008). The earliest fossil record of the animals and its significance. Phil. Trans. R. Soc. Lond. B 363, 14251434.Google Scholar
Budyko, M. I. (1969). The effect of solar radiation variations on the climate of the Earth. Tellus 21, 611619.Google Scholar
Buick, R. (1992). The antiquity of oxygenic photosynthesis: Evidence from stromatolites in sulphate-deficient Archaean lakes. Science 255, 7477.Google Scholar
Buick, R. (2007a). Did the Proterozoic ‘Canfield Ocean’ cause a laughing gas greenhouse? Geobiology 5, 97100.Google Scholar
Buick, R. (2007b). The earliest records of life on Earth. In: Planets and Life: The Emerging Science of Astrobiology, ed. Sullivan, W. T., Baross, J., Cambridge: Cambridge University Press, pp. 237264.Google Scholar
Buick, R. (2008). When did oxygenic photosynthesis evolve? Phil. Trans. R. Soc. Lond. B 363, 27312743.Google Scholar
Buick, R., et al. (1981). Stromatolite recognition in ancient rocks: An appraisal of irregularly laminated structures in an early Archean chert-barite unit from North Pole, Western Australia. Alcheringa 5, 161181.Google Scholar
Buick, R., et al. (1995a). Abiological origin of described stromatolites older than 3.2 Ga – Comment. Geology 23, 191191.Google Scholar
Buick, R., et al. (1995b). Stable isotopic compositions of carbonates from the Mesoproterozoic Bangemall Group, Northwestern Australia. Chem. Geol. 123, 153171.Google Scholar
Buie, M. W. and Grundy, W. M. (2000). The distribution and physical state of H2O on Charon. Icarus 148, 324339.Google Scholar
Bullock, M. A. and Grinspoon, D. H. (1996). The stability of climate on Venus. J. Geophys. Res. 101, 75217529.Google Scholar
Bullock, M. A. and Grinspoon, D. H. (2001). The recent evolution of climate on Venus. Icarus 150, 1937.Google Scholar
Bullock, M. A. and Moore, J. M. (2007). Atmospheric conditions on early Mars and the missing layered carbonates. Geophys. Res. Lett. 34.Google Scholar
Bunch, T. E. and Chang, S. (1980). Carbonaceous chondrites .2. Carbonaceous chondrite phyllosilicates and light-element geochemistry as indicators of parent body processes and surface conditions. Geochim. Cosmochim. Acta 44, 15431577.Google Scholar
Burgasser, A. J., et al. (2002). The spectra of T dwarfs. I. Near-infrared data and spectral classification. Astrophys. J. 564, 421451.Google Scholar
Burgisser, A. and Scaillet, B. (2007). Redox evolution of a degassing magma rising to the surface. Nature 445, 194197.Google Scholar
Burke, K., et al. (1976). In: Dominance Of Horizontal Movements, Arc, And Microcontinental Collisions During The Late Permobile Regime, New York: Wiley, pp. 113130.Google Scholar
Burns, S. J. and Matter, A. (1993). Carbon isotopic record of the latest Proterozoic from Oman. Eclogae Geol. Helv. 86, 595607.Google Scholar
Burr, D. M., et al. (2010). Inverted fluvial features in the Aeolis/Zephyria Plana region, Mars: Formation mechanism and initial paleodischarge estimates. J. Geophys. Res. 115.Google Scholar
Burrows, A. (2005). A theoretical look at the direct detection of giant planets outside the Solar System. Nature 433, 261268.Google Scholar
Burrows, A., et al. (2001). The theory of brown dwarfs and extrasolar giant planets. Rev. Mod. Phys. 73, 719765.Google Scholar
Burton, M. R., et al. (2013). Deep carbon emissions from volcanoes. Rev. Mineral. Geochem. 75, 323354.Google Scholar
Busigny, V., et al. (2013). Nitrogen cycle in the Late Archean ferruginous ocean. Chem. Geol. 362, 115130.Google Scholar
Busse, F. H. (1976). Simple model of convection in Jovian atmosphere. Icarus 29, 255260.Google Scholar
Busse, F. H. (1983). A model of mean zonal flows in the major planets. Geophys. Astro. Fluid. 23, 153174.Google Scholar
Butler, S. and Peltier, W. (2002). Thermal evolution of Earth: Models with time-dependent layering of mantle convection which satisfy the Urey ratio constraint. J. Geophys. Res. 107, 3-13-15.Google Scholar
Butterfield, N. J. (2000). Bangiomorpha pubescens n. gen., n. sp.: Implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterozoic radiation of eukaryotes. Paleobiology 26, 386404.Google Scholar
Butterfield, N. J. (2009). Oxygen, animals and oceanic ventilation: an alternative view. Geobiology 7, 17.Google Scholar
Butterfield, N. J. (2011). Animals and the invention of the Phanerozoic Earth system. Trends Ecol. Evol. 26, 8187.Google Scholar
Butterfield, N. J. (2015). Proterozoic photosynthesis – a critical review. Palaeontology 58, 953972.Google Scholar
Byrne, S., et al. (2009). Distribution of mid-latitude ground ice on Mars from new impact craters. Science 325, 16741676.Google Scholar
Caballero, R. and Huber, M. (2013). State-dependent climate sensitivity in past warm climates and its implications for future climate projections. P. Natl. Acad. Sci. USA 110, 14 16214 167.Google Scholar
Cabane, M., et al. (1993). Fractal aggregates in Titan atmosphere. Planet. Space Sci. 41, 257267.Google Scholar
Cairns-Smith, A. G. (1978). Precambrian solution photochemistry, inverse segregation, and banded iron formations. Nature 276, 807808.Google Scholar
Caldeira, K. (1995). Long-term control of atmospheric carbon dioxide: Low temperature seafloor alteration or terrestrial silicate rock weathering. Am. J. Sci. 295, 10771114.Google Scholar
Caldeira, K., et al. (1993). Cooling in the Late Cenozoic. Nature 361, 123124.Google Scholar
Caldeira, K. and Kasting, J. F. (1992a). The life span of the biosphere revisited. Nature 360, 721723.Google Scholar
Caldeira, K. and Kasting, J. F. (1992b). Susceptibility of the early Earth to irreversible glaciation caused by carbon dioxide clouds. Nature 359, 226228.Google Scholar
Came, R. E., et al. (2007). Coupling of surface temperatures and atmospheric CO2 concentrations during the Palaeozoic era. Nature 449, 198202.Google Scholar
Cameron, A. C. (2016). Extrasolar planetary transists. In: Methods of Detecting Exoplanets: 1st Advanced School on Exoplanetary Science, ed. Bozza, V. et al. Springer, Cham, pp. 89131.Google Scholar
Cameron, V., et al. (2009). A biomarker based on the stable isotopes of nickel. P. Natl. Acad. Sci. USA 106, 10 94410 948.Google Scholar
Campbell, A. J., et al. (2011). Refugium for surface life on Snowball Earth in a nearly-enclosed sea? A first simple model for sea-glacier invasion. Geophys. Res. Lett. 38.Google Scholar
Campbell, B., et al. (2008). SHARAD radar sounding of the Vastitas Borealis Formation in Amazonis Planitia. J. Geophys. Res. 113, 110.Google Scholar
Campbell, I. H. and Allen, C. M. (2008). Formation of supercontinents linked to increases in atmospheric oxygen. Nature Geosci. 1, 554558.Google Scholar
Campbell, I. H. and Squire, R. J. (2010). The mountains that triggered the Late Neoproterozoic increase in oxygen: The Second Great Oxidation Event. Geochim. Cosmochim. Acta 74, 41874206.Google Scholar
Campbell, I. H. and Taylor, S. R. (1993). No water, no granites no continents, no oceans. Geophys. Res. Lett. 10, 10611064.Google Scholar
Canfield, D. E. (1998). A new model for Proterozoic ocean chemistry. Nature 396, 450453.Google Scholar
Canfield, D. E. (2004). The evolution of the Earth surface sulfur reservoir. Am. J. Sci. 304, 839861.Google Scholar
Canfield, D. E. (2005). The early history of atmospheric oxygen: homage to Robert M. Garrels. Ann. Rev. Earth Planet. Sci. 33, 136.Google Scholar
Canfield, D. E. (2014). Proterozoic atmospheric oxygen. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 197216.Google Scholar
Canfield, D. E. and Farquhar, J. (2009). Animal evolution, bioturbation, and the sulfate concentration of the oceans. P. Natl. Acad. Sci. USA 106, 81238127.Google Scholar
Canfield, D. E., et al. (2010). The evolution and future of Earth’s nitrogen cycle. Science 330, 192196.Google Scholar
Canfield, D. E., et al. (2005). Aquatic Geomicrobiology. San Diego, CA: Elsevier Academic Press.Google Scholar
Canfield, D. E., et al. (2013). Oxygen dynamics in the aftermath of the Great Oxidation of Earth’s atmosphere. P. Natl. Acad. Sci. USA 110, 16 73616 741.Google Scholar
Canfield, D. E., et al. (2007). Late-Neoproterozoic deep-ocean oxygenation and the rise of animal life. Science 315, 9295.Google Scholar
Canfield, D. E. and Raiswell, R. (1999). The evolution of the sulfur cycle. Amer. J. Sci. 299, 697723.Google Scholar
Canfield, D. E., et al. (2006). Early anaerobic metabolisms. Phil. Trans. R. Soc. Lond. B 361, 18191834.Google Scholar
Canfield, D. E. and Teske, A. (1996). Late-Proterozoic rise in atmospheric oxygen concentration inferred from phylogenetic and sulfur isotope studies. Nature 382, 127132.Google Scholar
Canil, D. (1997). Vanadium partitioning and the oxidation state of Archaean komatiite magmas. Nature 389, 842845.Google Scholar
Canil, D. (2002). Vanadium in peridotites, mantle redox and tectonic environments: Archean to present. Earth Planet. Sci. Lett. 195, 7590.Google Scholar
Cannat, M., et al. (2010). Serpentinization and associated hydrogen and methane fluxes at slow spreading ridges. Geophysical Monograph Series, Vol. 188, pp. 241264.Google Scholar
Cantor, B., et al. (2002). Multiyear Mars Orbiter Camera (MOC) observations of repeated Martian weather phenomena during the northern summer season. J. Geophys. Res. 107, doi:10.1029/2001JE001588.Google Scholar
Cantor, B. A., et al. (2001). Martian dust storms: 1999 Mars Orbiter Camera observations. J. Geophys. Res. 106, 2365323687.Google Scholar
Canup, R. M. (2005). A giant impact origin of Pluto–Charon. Science 307, 546550.Google Scholar
Canuto, V. M., et al. (1982). UV radiation from the young Sun and oxygen levels in the pre-biological paleoatmosphere. Nature 296, 816820.Google Scholar
Canuto, V. M., et al. (1983). The young Sun and the atmosphere and photochemistry of the early Earth. Nature 305, 281286.Google Scholar
Carlson, R. W. (1999). A tenuous carbon dioxide atmosphere on Jupiter’s moon Callisto. Science 283, 820821.Google Scholar
Carmichael, I. S. E. (2002). The andesite aqueduct: perspectives on the evolution of intermediate magmatism in west-central (105-99 degrees W) Mexico. Contrib. Mineral. Petr. 143. 641663.Google Scholar
Carr, M. H. (1986). Mars: a water-rich planet. Icarus 68, 187216.Google Scholar
Carr, M. H. (1990). D/H on Mars: Effects of floods, volcanism, impacts and polar processes. Icarus 87, 210–27.Google Scholar
Carr, M. H. (1996). Water on Mars. New York: Oxford University Press.Google Scholar
Carr, M. H. (2006). The Surface of Mars. Cambridge: Cambridge University Press.Google Scholar
Carr, M. H. and Clow, G. D. (1981). Martian channels and valleys – Their characteristics, distribution, and age. Icarus 48, 91117.Google Scholar
Carr, M. H. and Head, J. W. (2003). Oceans on Mars: An assessment of the observational evidence and possible fate. J. Geophys. Res. 108, 5042.Google Scholar
Carr, M. H. and Head, J. W. (2010). Geologic history of Mars. Earth Planet. Sci. Lett. 294, 185203.Google Scholar
Carr, M. H. and Head, J. W. (2015). Martian surface/near-surface water inventory: Sources, sinks, and changes with time. Geophys. Res. Lett. 42, 726732.Google Scholar
Carr, M. H. and Malin, M. C. (2000). Meter-scale characteristics of martian channels and valleys. Icarus 146, 366386.Google Scholar
Carr, R. H., et al. (1985). Martian atmospheric carbon-dioxide and weathering products in SNC meteorites. Nature 314, 248250.Google Scholar
Carrier, B. L. and Kounaves, S. P. (2015). The origins of perchlorate in the Martian soil. Geophys. Res. Lett. 42, 37393745.Google Scholar
Carroll, B. W. and Ostlie, D. A. (2007). An Introduction to Modern Astrophysics. San Francisco: Pearson Addison-Wesley.Google Scholar
Carter, J., et al. (2013). Hydrous minerals on Mars as seen by the CRISM and OMEGA imaging spectrometers: Updated global view. Journal of Geophysical Research-Planets 118, 831858.Google Scholar
Casciotti, K. L. (2009). Inverse kinetic isotope fractionation during bacterial nitrite oxidation. Geochim. Cosmochim. Acta 73, 20612076.Google Scholar
Cassan, A., et al. (2012). One or more bound planets per Milky Way star from microlensing observations. Nature 481, 167169.Google Scholar
Cassanelli, J. P., et al. (2015). Sources of water for the outflow channels on Mars: Implications of the Late Noachian “icy highlands” model for melting and groundwater recharge on the Tharsis rise. Planet. Space Sci. 108, 5465.Google Scholar
Cassata, W. S., et al. (2012). Trapped Ar isotopes in meteorite ALH 84001 indicate Mars did not have a thick ancient atmosphere. Icarus 221, 461465.Google Scholar
Catling, D. and Kasting, J. F. (2007). Planetary atmospheres and life. In: Planets and Life: The Emerging Science of Astrobiology, ed. Sullivan, W. T. and Baross, J. A., Cambridge: Cambridge University Press, pp. 91116.Google Scholar
Catling, D. C. (2004). Planetary science: On Earth, as it is on Mars? Nature 429, 707708.Google Scholar
Catling, D. C. (2013). Astrobiology: A Very Short Introduction. Oxford: Oxford University Press.Google Scholar
Catling, D. C. (2014). The Great Oxidation Event Transition. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., 2nd edn. New York: Elsevier, pp. 191233.Google Scholar
Catling, D. C. (2015). Planetary Atmospheres. In: Treatise on Geophysics, ed. Schubert, G., 2nd edn. Oxford: Elsevier, pp. 429472.Google Scholar
Catling, D. C. and Claire, M. W. (2005). How Earth’s atmosphere evolved to an oxic state: A status report. Earth Planet. Sci. Lett. 237, 120.Google Scholar
Catling, D. C., et al. (2004). Understanding the evolution of atmospheric redox state from the Archaean to the Proterozoic. In: Field Forum on Processes on the Early Earth, ed. Reimold, W. U. and Hofmann, A., Kaapvaal Craton, S. Africa: University of Witwatersrand, pp. 1719.Google Scholar
Catling, D. C., et al. (2007). Anaerobic methanotrophy and the rise of atmospheric oxygen. Phil. Trans. R. Soc. Lond. A 365, 18671888.Google Scholar
Catling, D. C., et al. (2010). Atmospheric origins of perchlorate on Mars and in the Atacama. J. Geophys. Res. 115, E00E11.Google Scholar
Catling, D. C., et al. (2005). Why O2 is required by complex life on habitable planets and the concept of planetary “oxygenation time”. Astrobiology 5, 415438.Google Scholar
Catling, D. C., et al. (2012). Does the Vastitas Borealis Formation contain oceanic or volcanic deposits? Third Int. Conf. on Early Mars, 7031.Google Scholar
Catling, D. C., et al. (2006). Light-toned layered deposits in Juventae Chasma, Mars. Icarus 181, 2651.Google Scholar
Catling, D. C. and Zahnle, K. J. (2013). An impact erosion stability limit controlling the existence of atmospheres on exoplanets and solar system bodies. 44th Lunar Planet. Sci. Conf. 2665.Google Scholar
Catling, D. C., et al. (2001). Biogenic methane, hydrogen escape, and the irreversible oxidation of early Earth. Science 293, 839843.Google Scholar
Catling, D. C., et al. (2002). What caused the second rise of O2 in the late Proterozoic? Methane, sulfate, and irreversible oxidation. Astrobiology 2, 569 (Abstract).Google Scholar
Catling, D. C. and Zahnle, K. L. (2009). The Planetary Air Leak. Sci. Am. 300, 3643.Google Scholar
Cavosie, A. J., et al. (2005). Magmatic delta O-18 in 4400–3900 Ma detrital zircons: A record of the alteration and recycling of crust in the Early Archean. Earth Planet. Sci. Lett. 235, 663681.Google Scholar
Chaffin, M. S., et al. (2014). Unexpected variability of Martian hydrogen escape. Geophys. Res. Lett. 41, 314320.Google Scholar
Chaloner, W. G. (1989). Fossil charcoal as an indicator of paleoatmospheric oxygen level. J. Geol. Soc. London 146, 171174.Google Scholar
Chamberlain, J. W. (1963). Planetary coronae and atmospheric evaporation. Planet. Space Sci. 11, 901960.Google Scholar
Chamberlain, J. W. and Campbell, F. J. (1967). Rate of evaporation of a non-Maxwellian atmosphere. Ap. J. 149, 687705.Google Scholar
Chamberlain, J. W. and Hunten, D. M. (1987). Theory of Planetary Atmospheres. Orlando: Academic Press.Google Scholar
Chamberlain, J. W. and Smith, G. R. (1971). Rate of evaporation of a non-Maxwellian atmosphere. Planet. Space Sci. 19, 675684.Google Scholar
Chambers, J. E. (2007). On the stability of a planet between Mars and the asteroid belt: Implications for the Planet V hypothesis. Icarus 189, 386400.Google Scholar
Chambers, J. E. (2009). Planetary migration: What does it mean for planet formation? Annu. Rev. Earth. Pl. Sc. 37, 321344.Google Scholar
Chambers, J. E. (2014). Planet Formation. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 5572.Google Scholar
Chambers, J. E., et al. (1996). The stability of multi-planet systems. Icarus 119, 261268.Google Scholar
Chambers, L. H., et al. (2002). Examination of new CERES data for evidence of tropical Iris feedback. Journal of Climate 15, 37193726.Google Scholar
Chameides, W. and Walker, J. C. G. (1975). Possible variation of ozone in troposphere during course of geologic time. Am. J. Sci. 275, 737752.Google Scholar
Chameides, W. L., et al. (1977). NOx production in lightning. J. Atmos. Sci. 34, 143149.Google Scholar
Chameides, W. L. and Walker, J. C. G. (1981). Rates of fixation by lightning of carbon and nitrogen in possible primitive terrestrial atmospheres. Origins of Life 11, 291302.Google Scholar
Chandrasekhar, S. (1960). Radiative Transfer. New York: Dover Publications.Google Scholar
Chang, S., et al. (1983). Prebiotic organic syntheses and the origin of life. In: Earth’s Earliest Biosphere: Its Origin and Evolution, ed. Schopf, J. W., Princeton, New Jersey: Princeton University Press, pp. 5392.Google Scholar
Chapman, G. A., et al. (2012). Comparison of TSI from SORCE TIM with SFO Ground-Based Photometry. Sol. Phys. 276, 3541.Google Scholar
Chapman, M. G., et al. (2003). Possible Juventae Chasma subice volcanic eruptions and Maja Valles ice outburst floods on Mars: Implications of Mars Global Surveyor crater densities, geomorphology, and topography. J. Geophys. Res. 108, 5113.Google Scholar
Chapman, M. G., et al. (2010a). Noachian–Hesperian geologic history of the Echus Chasma and Kasei Valles system on Mars: New data and interpretations. Earth Planet. Sc. Lett. 294, 256271.Google Scholar
Chapman, M. G., et al. (2010b). Amazonian geologic history of the Echus Chasma and Kasei Valles system on Mars: New data and interpretations. Earth Planet. Sc. Lett. 294, 238255.Google Scholar
Chapman, S. and Cowling, T. G. (1939). The Mathematical Theory Of Non-Uniform Gases: An Account of the Kinetic Theory of Viscosity, Thermal Conduction, and Diffusion In Gases. Cambridge: Cambridge University Press.Google Scholar
Chapman, S., et al. (1990). The Mathematical Theory of Non-Uniform Gases: An Account of the Kinetic Theory of Viscosity, Thermal Conduction and Diffusion In Gases (3rd edition). Cambridge: Cambridge University Press.Google Scholar
Chapman, S. and Lindzen, R. S. (1970). Atmospheric Tides: Thermal and Gravitational. New York: Gordon and Breach.Google Scholar
Charbonneau, D., et al. (2000). Detection of planetary transits across a sun-like star. Ap. J. 529, L45L49.Google Scholar
Charbonneau, D., et al. (2002). Detection of an extrasolar planet atmosphere. Astrophys. J. 568, 377384.Google Scholar
Charlson, R. J., et al. (1987). Oceanic phytoplankton, atmospheric sulfur, cloud albedo and climate. Nature 326, 655661.Google Scholar
Charnay, B., et al. (2015). Methane storms as a driver of Titan’s dune orientation. Nat. Geosci. 8, 362366.Google Scholar
Charnay, B., et al. (2014). Titan’s past and future: 3D modeling of a pure nitrogen atmosphere and geological implications. Icarus 241, 269279.Google Scholar
Charnay, B., et al. (2013). Exploring the faint young Sun problem and the possible climates of the Archean Earth with a 3-D GCM. J. Geophys. Res. 118, 10 41410 431.Google Scholar
Chatterjee, S. and Tan, J. C. (2014). Inside-out Planet Formation. Astrophys. J. 780.Google Scholar
Chaufray, J., et al. (2008). Observation of the hydrogen corona with SPICAM on Mars Express. Icarus 195, 598613.Google Scholar
Chaufray, J. Y., et al. (2012). Hydrogen density in the dayside venusian exosphere derived from Lyman-alpha observations by SPICAV on Venus Express. Icarus 217, 767778.Google Scholar
Chen, G. Q. and Ahrens, T. J. (1997). Erosion of terrestrial planet atmosphere by surface motion after a large impact. Physics of the Earth and Planetary Interiors 100, 2126.Google Scholar
Chen, X., et al. (2015). Rise to modern levels of ocean oxygenation coincided with the Cambrian radiation of animals. Nat. Commun. 6.Google Scholar
Cheng, M., et al. (2015). Mo marine geochemistry and reconstruction of ancient ocean redox states. Sci. China Earth Sci., 1–11.Google Scholar
Cho, J. Y. K. and Polvani, L. M. (1996). The morphogenesis of bands and zonal winds in the atmospheres on the giant outer planets. Science 273, 335337.Google Scholar
Choi, J., et al. (2013). Precise Doppler monitoring of Barnard’s Star. Astrophys. J. 764, 131, doi: 10.1088/0004-637X/764/2/131.Google Scholar
Christensen, P. R. (2003). Formation of recent martian gullies through melting of extensive water-rich snow deposits. Nature 422, 4548.Google Scholar
Christensen, P. R. (2006). Water at the poles and in permafrost regions of Mars. Elements 2, 151155.Google Scholar
Christensen, P. R., et al. (2001). Global mapping of Martian hematite mineral deposits: Remnants of water-driven processes on early Mars. J. Geophys. Res. 106, 23 87323 885.Google Scholar
Christensen, P. R., et al. (2004). Initial results from the Mini-TES experiment in Gusev crater from the Spirit rover. Science 305, 837842.Google Scholar
Christopher, J. B., et al. (2015). Terrestrial planet occurrence rates for the Kepler GK dwarf sample. Astrophys. J. 809, 8, doi: 10.1088/0004-637X/809/1/8.Google Scholar
Chumakov, N. M. (2008). A problem of total glaciations on the Earth in the Late Precambrian. Stratigr. Geo. Correl. 16, 107119.Google Scholar
Chyba, C. and Sagan, C. (1992). Endogenous production, exogenous delivery and impact-shock synthesis of organic molecules: an inventory for the origins of life. Nature 355, 125132.Google Scholar
Chyba, C. F. (1987). The cometary contribution to the oceans of primitive Earth. Nature 330, 632635.Google Scholar
Chyba, C. F. (1990). Extraterrestrial amino acids and terrestrial life. Nature 348, 113114.Google Scholar
Chyba, C. F. and Phillips, C. B. (2007). Europa. In: Planets and Life: The Emerging Science of Astrobiology, ed. Sullivan, W. T. and Baross, J. A., Cambridge: Cambridge University Press.Google Scholar
Chyba, C. F., et al. (1990). Cometary delivery of organic molecules to the early Earth. Science 249, 366373.Google Scholar
Cicerone, R. (1989). Analysis of sources and sinks of atmospheric nitrous oxide (N2O). J. Geophys. Res. 94, 18 26518 271.Google Scholar
Clack, J. A. (2000). The origin of the tetrapods. In: Amphibian Biology, ed. Heatwole, H. and Carroll, R. L., Chipping Norton, Australia: Surrey Beatty, pp. 9791029.Google Scholar
Claire, M., et al. (2006). Biogeochemical modelling of the rise in atmospheric oxygen. Geobiology 4, 239269.Google Scholar
Claire, M. W. (2008). Quantitative modeling of the rise in atmospheric oxygen. University of Washington, Ph.D., Seattle.Google Scholar
Claire, M. W., et al. (2014). Modeling the signature of sulfur mass-independent fractionation produced in the Archean atmosphere. Geochim. Cosmochim. Acta 141, 365380.Google Scholar
Claire, M. W., et al. (2012). The evolution of solar flux from 2 nm to 160 microns: Quantitative estimates for planetary studies. Astrophys. J. 757, 95 doi:10.1088/0004-637X/757/1/95.Google Scholar
Clancy, R., et al. (2004). A measurement of the 362 GHz absorption line of Mars atmospheric H2O2. Icarus 168, 116121.Google Scholar
Clark, B. C. (1993). Geochemical components in Martian soil. Geochim. Cosmochim. Acta 57, 45754581.Google Scholar
Clark, B. C., et al. (2005). Chemistry and mineralogy of outcrops at Meridiani Planum. Earth Planet. Sc. Lett. 240, 7394.Google Scholar
Clark, I. D. (1971). Chemical kinetics of CO2 atmospheres. J. Atmos. Sci. 28, 847858.Google Scholar
Clarke, J. T., et al. (2009). HST observations of the extended hydrogen corona of Mars. Bull. Am. Astron. Soc. 41, 49.11 (abstract).Google Scholar
Clarke, J. T., et al. (2014). A rapid decrease of the hydrogen corona of Mars. Geophys. Res. Lett. 41, 80138020.Google Scholar
Clayton, R. N. (2002). Solar System : Self-shielding in the solar nebula. Nature 415, 860861.Google Scholar
Cleaves, H. J., et al. (2008). A reassessment of prebiotic organic synthesis in neutral planetary atmospheres. Origins Life Evol. Biosph. 38, 105115.Google Scholar
Clifford, S. M. and Parker, T. J. (2001). The evolution of the martian hydrosphere: Implications for the fate of a primordial ocean and the current state of the northern plains. Icarus 154, 4079.Google Scholar
Cloud, P. (1976). Beginnings of biospheric evolution and their biogeochemical consequences. Paleobiol. 2, 351387.Google Scholar
Cloud, P. E. (1968). Atmospheric and hydrospheric evolution on the primitive earth. Both secular accretion and biological and geochemical processes have affected earth’s volatile envelope. Science 160, 729736.Google Scholar
Clough, S. A. and Iacono, M. J. (1995). Line-by-line calculation of atmospheric fluxes and cooling rates. 1. Application to carbon dioxide, ozone, methane, nitrous oxide, and the halocarbons. J. Geophys. Res. 100, 16 51916 535.Google Scholar
Clough, S. A., et al. (1992). Line-by-line calculations of atmospheric fluxes and cooling rates: Application to water vapor. J. Geophys. Res. 97, 15 76115 785.Google Scholar
Coates, A. J., et al. (2007). Discovery of heavy negative ions in Titan’s ionosphere. Geophys. Res. Lett. 34.Google Scholar
Coates, A. J., et al. (2009). Heavy negative ions in Titan’s ionosphere: Altitude and latitude dependence. Planet. Space Sci. 57, 18661871.Google Scholar
Cohen, B. A., et al. (2000). Support for the lunar cataclysm hypothesis from lunar meteorite impact melt ages. Science 290, 17541756.Google Scholar
Cohen, P. A., et al. (2009). Large spinose microfossils in Ediacaran rocks as resting stages of early animals. P. Natl. Acad. Sci. U.S.A. 106, 65196524.Google Scholar
Colaprete, A., et al. (2005). Albedo of the south pole on Mars determined by topographic forcing of atmosphere dynamics. Nature 435, 184188.Google Scholar
Colaprete, A. and Toon, O. B. (2003). Carbon dioxide clouds in an early dense Martian atmosphere. J. Geophys. Res. 108, 5025, doi:10.1029/2002JE001967.Google Scholar
Colburn, D., et al. (1989). Diurnal variations in optical depth at Mars. Icarus 79, 159–89.Google Scholar
Colman, A. S. and Holland, H. D. (2000). The global diagenetic flux of phosphorus from marine sediments to the oceans: Redox sensitivity and the control of atmospheric oxygen levels. In: Marine Authigenesis: From Global to Microbial, ed. C. R. Glenn, et al.: SEPM Special Pub. No. 66.Google Scholar
Comas-Solá, J. (1908). Observationes des satellites principauz de Jupiter et de Titan. Astron. Nachr. 179.Google Scholar
Compston, W. and Pidgeon, R. T. (1986). Jack Hills, Evidence of more very old detrital zircons in Western Australia. Nature 321, 766769.Google Scholar
Condon, D., et al. (2005). U–Pb ages from the neoproterozoic Doushantuo Formation, China. Science 308, 9598.Google Scholar
Condon, D. J., et al. (2002). Neoproterozoic glacial-rainout intervals: Observations and implications. Geology 30, 3538.Google Scholar
Connelly, J. N., et al., Pb–Pb chronometry and the early Solar System. Geochim. Cosmochim. Acta, in press, doi:10.1016/j.gca.2016.10.044.Google Scholar
Conrad, R. (1996). Soil microorganisms as controllers of atmospheric trace gases (H2, CO, CH4, OCS, N2O, and NO). Microbiol. Rev. 60, 609640.Google Scholar
Coogan, L. A. and Dosso, S. E. (2015). Alteration of ocean crust provides a strong temperature dependent feedback on the geological carbon cycle and is a primary driver of the Sr-isotopic composition of seawater. Earth Planet. Sc. Lett. 415, 3846.Google Scholar
Coogan, L. A. and Gillis, K. M. (2013). Evidence that low-temperature oceanic hydrothermal systems play an important role in the silicate–carbonate weathering cycle and long-term climate regulation. Geochem. Geophys. Geosys. 14, 17711786.Google Scholar
Cook, J. C., et al. (2007). Near-infrared spectroscopy of Charon: Possible evidence for cryovolcanism on Kuiper Belt objects. Astrophys. J. 663, 14061419.Google Scholar
Cook, K. H. (2004). Hadley circulation dynamics. In: The Hadley Circulation: Past, Present and Future, ed. Diaz, H. F. and Bradley, R. S., Dordrecht: Kluwer, pp. 6183.Google Scholar
Cooney, C. L. (1975). Thermophilic anaerobic digestion of solid waste for fuel gas production. J. Biotech. Bioengineer. 17, 11191135.Google Scholar
Cooper, J. F., et al. (2001). Energetic ion and electron irradiation of the icy Galilean satellites. Icarus 149, 133159.Google Scholar
Cope, M. J. and Chaloner, W. G. (1980). Fossil charcoal as evidence of past atmospheric composition. Nature 283, 647649.Google Scholar
Cordier, D., et al. (2009). An estimate of the chemical composition of Titan’s lakes. Ap. J. 707, L128L131.Google Scholar
Cordier, D., et al. (2012). Titan’s lakes chemical composition: Sources of uncertainties and variability. Planet. Space Sci. 61, 99107.Google Scholar
Corliss, J. B., et al. (1981). An hypothesis concerning the relationship between submarine hot springs and the origin of life on Earth. Ocean Acta, 59–69.Google Scholar
Cornell, R. M. and Schwertmann, U. (1996). The iron oxides : structure, properties, reactions, occurrences and uses. Weinheim ; Cambridge: VCH.Google Scholar
Correia, A. C. M. and Laskar, J. (2001). The four final rotation states of Venus. Nature 411, 767770.Google Scholar
Correia, A. C. M. and Laskar, J. (2003). Long-term evolution of the spin of Venus II. Numerical simulations. Icarus 163, 2445.Google Scholar
Correia, A. C. M., et al. (2003). Long-term evolution of the spin of Venus I. Theory. Icarus 163, 123.Google Scholar
Costard, F., et al. (2002). Formation of recent Martian debris flows by melting of near-surface ground ice at high obliquity. Science 295, 110113.Google Scholar
Coughenour, C. L., et al. (2009). Tides, tidalites, and secular changes in the Earth-Moon system. Earth Sci. Rev. 97, 5979.Google Scholar
Cowan, N. B., et al. (2009). Alien maps of an ocean-bearing world. Astrophys. J. 700, 915923.Google Scholar
Craddock, R. A. and Howard, A. D. (2002). The case for rainfall on a warm, wet early Mars. J. Geophys. Res. 107, 5111, doi:10.1029/2001JE001505.Google Scholar
Crary, F., et al. (2010). Upper limits on carbon group ions near the orbit of Titan: Implications for methane escape from Titan. 38th COSPAR Scientific Assembly, 5.Google Scholar
Cravens, T. E., et al. (1997). Photochemical sources of non-thermal neutrals for the exosphere of Titan. Planetary and Space Science 45, 889896.Google Scholar
Crespin, A., et al. (2008). Diagnostics of Titan’s stratospheric dynamics using Cassini/CIRS data and the 2-dimensional IPSL circulation model. Icarus 197, 556571.Google Scholar
Crisp, D. (1997). Absorption of sunlight by water vapor in cloudy conditions: A partial explanation for the cloud absorption anomaly. Geophys. Res. Lett. 24, 571574.Google Scholar
Croft, S. K., et al. (1995). The geology of Triton. In: Neptune and Triton, ed. Cruikshank, D. P., Tucson: University of Arizona Press, pp. 879947.Google Scholar
Crossfield, I. J. M. (2015). Observations of exoplanet atmospheres. Publ. Astron. Soc. Pac. 127, 941960.Google Scholar
Crow, C. A., et al. (2011). Views from EPOXI: colors in our solar system as an analog for extrasolar planets. Astrophys. J. 729, 130.Google Scholar
Crowe, S. A., et al. (2013). Atmospheric oxygenation three billion years ago. Nature 501, 535.Google Scholar
Crowell, J. C. (1999). Pre-Mesozoic Ice Ages: Their Bearing on Understanding the Climate System. Vol. GSA Memoir 192. Geological Society of America.Google Scholar
Crowley, T. J. (2000). CLIMAP SSTs re-revisited. Climate Dyn. 16, 241255.Google Scholar
Crowley, T. J., et al. (2001). CO2 levels required for deglaciation of a “Near-Snowball” Earth. Geophys. Res. Lett. 28, 283286.Google Scholar
Crowley, T. J. and North, G. R. (1991). Paleoclimatology. New York: Oxford University Press.Google Scholar
Cruikshank, D. P., et al. (2000). Water ice on Triton. Icarus 147, 309316.Google Scholar
Crutzen, P. J. (1976). Possible importance of CSO for sulfate layer of stratosphere. Geophys. Res. Lett. 3, 7376.Google Scholar
Crutzen, P. J. (1979). Role of NO and NO2 in the chemistry of the troposphere and stratosphere. Annu. Rev. Earth Planet. Sci. 7, 443472.Google Scholar
Crutzen, P. J. and Zimmermann, P. H. (1991). The changing photochemistry of the troposphere. Tellus A 43, 136151.Google Scholar
Cui, J., et al. (2011). The implications of the H2 variability in Titan’s exosphere. J. Geophys. Res. 116.Google Scholar
Cui, J., et al. (2008). Distribution and escape of molecular hydrogen in Titan’s thermosphere and exosphere. J. Geophys. Res. 113, E10004, doi:10.1029/2007JE003032.Google Scholar
Cull, S. C., et al. (2010). Concentrated perchlorate at the Mars Phoenix landing site: Evidence for thin film liquid water on Mars. Geophys. Res. Lett. 37, L22203.Google Scholar
Cunha, D., et al. (2015). Spin evolution of Earth-sized exoplanets, including atmospheric tides and core–mantle friction. Int. J. Astrobiol. 14, 233254.Google Scholar
Cunningham, N. J., et al. (2015). Detection of Callisto’s oxygen atmosphere with the Hubble Space Telescope. Icarus 254, 178189.Google Scholar
Curtis, A. R. and Goody, R. M. (1956). Thermal radiation in the upper atmosphere. Proc. R. Soc. Lond. A 236, 193206.Google Scholar
Cushing, M. C., et al. (2011). The discovery of Y Dwarfs using data from the Wide-Field Infrared Survey Explorer (WISE). Astrophys. J. 743.Google Scholar
Cutter, G. A. and Bruland, K. W. (1984). The marine biogeochemistry of selenium: a re-evaluation. Limnol. Oceanogr. 29, 11791192.Google Scholar
Cuzzi, J. N., et al. (2008). Toward planetesimals: dense chondrule clumps in the protoplanetary nebula. Astrophys. J. 687, 14321447.Google Scholar
Czaja, A. D., et al. (2010). Iron and carbon isotope evidence for ecosystem and environmental diversity in the similar to 2.7 to 2.5 Ga Hamersley Province, Western Australia. Earth Planet. Sci. Lett. 292, 170180.Google Scholar
Czaja, A. D., et al. (2013). Biological Fe oxidation controlled deposition of banded iron formation in the ca. 3770 Ma Isua Supracrustal Belt (West Greenland). Earth Planet. Sc. Lett. 363, 192203.Google Scholar
Czaja, A. D., et al. (2012). Evidence for free oxygen in the Neoarchean ocean based on coupled iron-molybdenum isotope fractionation. Geochim. Cosmochim. Acta 86, 118137.Google Scholar
D’Hondt, S., et al. (2002). Metabolic activity of subsurface life in deep-sea sediments. Science 295, 20672070.Google Scholar
Dahl, T. W., et al. (2010). Devonian rise in atmospheric oxygen correlated to the radiations of terrestrial plants and large predatory fish. P. Natl. Acad. Sci. U.S.A. 107, 17 91117 915.Google Scholar
Dai, A. and Wang, J. (1999). Diurnal and semidiurnal tides in global surface pressure fields. J. Atmos. Sci. 56, 38743891.Google Scholar
Daines, S. J. and Lenton, T. M. (2016). The effect of widespread early aerobic marine ecosystems on methane cycling and the Great Oxidation. Earth Planet. Sci. Lett. 434, 4251.Google Scholar
Dalcanton, J., et al. (2015). From Cosmic Birth to Living Earths: The Future of UVOIR Space Astronomy. arXiv preprint arXiv:1507.04779.Google Scholar
Danielache, S. O., et al. (2008). High-precision spectroscopy of 32S, 33S, and 34S sulfur dioxide: Ultraviolet absorption cross sections and isotope effects. J. Geophys. Res. 113, D17314.Google Scholar
Darling, A. and Whipple, K. (2015). Geomorphic constraints on the age of the western Grand Canyon. Geosphere 11, 958976.Google Scholar
Dauphas, N. (2003). The dual origin of the terrestrial atmosphere. Icarus 165, 326339.Google Scholar
Dauphas, N., et al. (2014). Geochemical arguments for an Earth-like Moon-forming impactor. Phil. Trans R. Soc. Lond. A 372.Google Scholar
Dauphas, N. and Kasting, J. F. (2011). Low pCO2 in the pore water, not in the Archean atmosphere. Nature 474, E1, doi:10.1038/nature09960.Google Scholar
Dauphas, N. and Pourmand, A. (2011). Hf-W-Th evidence for rapid growth of Mars and its status as a planetary embryo. Nature 473, 489–U227.Google Scholar
Dauphas, N., et al. (2000). The late asteroidal and cometary bombardment of Earth as recorded in water deuterium to protium ratio. Icarus 148, 508512.Google Scholar
David, L. A. and Alm, E. J. (2011). Rapid evolutionary innovation during an Archaean genetic expansion. Nature 469, 9396.Google Scholar
Davies, G. F. (1980). Thermal histories of convective Earth models and constraints on radiogenic heat production in the Earth. J. Geophys. Res. 85, 25172530.Google Scholar
Davies, G. F. (2002). Stirring geochemistry in mantle convection models with stiff plates and slabs. Geochim. Cosmochim. Acta 66, 31253142.Google Scholar
Davies, G. F. (2006). Gravitational instability of the early Earth’s upper mantle and the viability of early plate tectonics. Earth Planet. Sci. Lett. 243, 376382.Google Scholar
de Kok, R. J., et al. (2014). HCN ice in Titan’s high-altitude southern polar cloud. Nature 514, 65.Google Scholar
de Wit, M. J. and Furnes, H. (2016). 3,5-Ga hydrothermal fields and diamictites in the Barberton Greenstone Belt – Paleoarchean crust in cold environments. Sci. Adv. 2, e1500368.Google Scholar
DeBergh, C., et al. (1991). Deuterium on Venus – Observations from Earth. Science 251, 547549.Google Scholar
Degens, E. T. and Epstein, S. (1962). Relationship between O18/O16 ratios in coexisting carbonates, cherts and diatomites. Amer. Assoc. Petrol. Geol. Bull. 46, 534542.Google Scholar
Del Genio, A. D., et al. (2009). Saturn atmospheric structure and dynamics. In: Saturn from Cassini-Huygens, ed. Dougherty, M., et al., New York: Springer, pp. 113160.Google Scholar
Del Genio, A. D. and Suozzo, R. J. (1987). A comparative study of rapidly and slowly rotating dynamic regimes in a terrestrial General Circulation Model. J. Atmos. Sci. 44, 973986.Google Scholar
Del Genio, A. D. and Zhou, W. (1996). Simulations of superrotation on slowly rotating planets: Sensitivity to rotation and initial condition. Icarus 120, 332343.Google Scholar
Delano, J. W. (2001). Redox history of the Earth’s interior: implications for the origin of life. Orig. Life Evol. Biosph. 31, 311341.Google Scholar
Delorme, P., et al. (2013). Direct-imaging discovery of a 12–14 Jupiter-mass object orbiting a young binary system of very low-mass stars. Astronomy & Astrophysics 553, L5.Google Scholar
Delsemme, A. H. (2001). An argument for the cometary origin of the biosphere. Am. Sci. 89, 432442.Google Scholar
Derry, L. A., et al. (1992). Sedimentary cycling and environmental change in the Late Proterozoic: Evidence from stable and radiogenic isotopes. Geochim. Cosmochim. Acta 56, 13171329.Google Scholar
Derry, L. A. (2014). Organic carbon cycling and the lithosphere. In: Treatise on Geochemistry (Second Edition), ed. Holland, H. D. and Turekian, K. K.. Oxford: Elsevier, pp. 239249.Google Scholar
Des Marais, D. J., et al. (1992). Carbon isotope evidence for the stepwise oxidation of the Proterozoic environment. Nature 359, 605609.Google Scholar
DesMarais, D. J., et al. (1992). Carbon isotope evidence for the stepwise oxidation of the Proterozoic environment. Nature 359, 605609.Google Scholar
Dessler, A. E., et al. (2013). Stratospheric water vapor feedback. P. Natl. Acad. Sci. USA 110, 1808718091.Google Scholar
Dessler, A. E., et al. (2008). Water-vapor climate feedback inferred from climate fluctuations, 2003–2008. Geophys. Res. Lett. 35, L20704, doi:10.1029/2008GL035333.Google Scholar
Dhuime, B., et al. (2012). A Change in the Geodynamics of Continental Growth 3 Billion Years Ago. Science 335, 13341336.Google Scholar
Di Achille, G. and Hynek, B. M. (2010). Ancient ocean on Mars supported by global distribution of deltas and valleys. Nature Geoscience 3, 459463.Google Scholar
Diakonov, I., et al. (1994). Thermodynamic properties of iron-oxides and hydroxides .1. Surface and bulk thermodynamic properties of goethite (alpha-FeOOH) up to 500 K. Eur. J. Mineral. 6, 967983.Google Scholar
Dickens, A. F., et al. (2004). Reburial of fossil organic carbon in marine sediments. Nature 427, 336339.Google Scholar
Dickey, J. O., et al. (1994). Lunar laser ranging – a continuing legacy of the Apollo Program. Science 265, 482490.Google Scholar
Dickson, J. L., et al. (2015). Recent climate cycles on Mars: Stratigraphic relationships between multiple generations of gullies and the latitude dependent mantle. Icarus 252, 8394.Google Scholar
Dickson, J. L., et al. (2007). Martian gullies in the southern mid-latitudes of Mars: Evidence for climate-controlled formation of young fluvial features based upon local and global topography. Icarus 188, 315323.Google Scholar
Dieterich, S., B., et al. (2012). The Solar Neighborhood. XXVIII. The multiplicity fraction of nearby stars from 5 to 70 AU and the brown dwarf desert around M Dwarfs. Astron. J. 144, 64.Google Scholar
Dima, I. M. and Wallace, J. M. (2003). On the seasonality of the Hadley cell. J. Atmos. Sci. 60, 15221527.Google Scholar
Diniega, S., et al. (2013). A new dry hypothesis for the formation of martian linear gullies. Icarus 225, 526537.Google Scholar
Dobrijevic, M., et al. (2014). Coupling of oxygen, nitrogen, and hydrocarbon species in the photochemistry of Titan’s atmosphere. Icarus 228, 324346.Google Scholar
Dobrovolskis, A. R. (1980). Atmospheric tides and the rotation of Venus .2. Spin evolution. Icarus 41, 1835.Google Scholar
Dobrovolskis, A. R. (2009). Insolation patterns on synchronous exoplanets with obliquity. Icarus 204, 110.Google Scholar
Dobrovolskis, A. R. and Ingersoll, A. P. (1980). Atmospheric tides and the totation of Venus .1. Tidal theory and the balance of torques. Icarus 41, 117.Google Scholar
Dobson, G. M. B. (1956). Origin and distribution of the polyatomic molecules in the atmosphere. Proc. R. Soc. Lond. A 236, 187193.Google Scholar
Dohnanyi, J. S. (1972). Interplanetary objects in review: Statistics of their masses and dynamics. Icarus 17, 1.Google Scholar
Dole, S. H. (1964). Habitable Planets for Man. New York: Blaisdell Publishing.Google Scholar
Domagal-Goldman, S. D., et al. (2011). Using biogenic sulfur gases as remotely detectable biosignatures on anoxic planets. Astrobiology 11, 419441.Google Scholar
Domagal-Goldman, S. D., et al. (2014). Abiotic ozone and oxygen in atmospheres similar to prebiotic Earth. Astrophys. J. 792, doi: 10.1088/0004-637X/792/2/90.Google Scholar
Donahue, T. M. (1995). Evolution of water reservoirs on Mars from D/H ratios in the atmosphere and crust. Nature 374, 432434.Google Scholar
Donahue, T. M., et al. (1982). Venus was wet: A measurement of the ratio of deuterium to hydrogen. Science 216, 630633.Google Scholar
Donahue, T. M. and Pollack, J. B. (1983). Origin and evolution of the atmosphere of Venus. In: Venus, ed. Hunten, D. M., et al., Tucson: University of Arizona Press, pp. 10031036.Google Scholar
Donnadieu, Y., et al. (2004). A ‘snowball Earth’ climate triggered by continental break-up through changes in runoff. Nature 428, 303306.Google Scholar
Doolittle, W. F. (2009). The practice of classification and the theory of evolution, and what thedemise of Charles Darwin’s tree of life hypothesis means for both of them. Phil. Trans. R. Soc. Lond. B 364, 22212228.Google Scholar
Dos Santos, P. C., et al. (2012). Distribution of nitrogen fixation and nitrogenase-like sequences amongst microbial genomes. BMC Genomics 13, 162.Google Scholar
Dowling, T. E. (1995). Dynamics of Jovian atmospheres. Annu. Rev. Fluid Mech. 27, 293334.Google Scholar
Drake, M. J. and Righter, K. (2002). Determining the composition of the Earth. Nature 416, 3944.Google Scholar
Dreibus, G. and Wänke, H. (1987). Volatiles on Earth and Mars, A comparison. Icarus 71, 225–40.Google Scholar
Dressing, C. D. and Charbonneau, D. (2015). The occurrence of potentially habitable planets orbiting M dwarfs estimated from the full Kepler dataset and an empirical measurement of the detection sensitivity. Astrophys. J. 807, 45.Google Scholar
Driese, S. G., et al. (2011). Neoarchean paleoweathering of tonalite and metabasalt: Implications for reconstructions of 2.69 Ga early terrestrial ecosystems and paleoatmospheric chemistry. Precambrian Res. 189, 117.Google Scholar
Du, S. Y., et al. (2011). The kinetics study of the S + S-2 -> S-3 reaction by the chaperone mechanism. Journal of Chemical Physics 134.Google Scholar
Dudley, R. (1998). Atmospheric oxygen, giant Paleozoic insects and the evolution of aerial locomotor performance. J. Exp. Biol. 201, 10431050.Google Scholar
Dundas, C. M., et al. (2014). HiRISE observations of new impact craters exposing Martian ground ice. J. Geophys. Res. 119, 109127.Google Scholar
Dundas, C. M., et al. (2012). Seasonal activity and morphological changes in martian gulllies. Icarus, doi:10.1016/j.icarus.2012.04.005.Google Scholar
Dundas, C. M., et al. (2015). Long-term monitoring of martian gully formation and evolution with MRO/HiRISE. Icarus 251, 244263.Google Scholar
Dundas, C. M. and Keszthelyi, L. P. (2014). Emplacement and erosive effects of lava in south Kasei Valles, Mars. J. Volcanol. Geoth. Res. 282, 92102.Google Scholar
Dutkiewicz, A., et al. (2006). Biomarkers from Huronian oil-bearing fluid inclusions: An uncontaminated record of life before the Great Oxidation Event. Geology 34, 437440.Google Scholar
Duxbury, N. S. and Brown, R. H. (1997). The role of an internal heat source for the eruptive plumes on Triton. Icarus 125, 8393.Google Scholar
Dworkin, L. P., et al. (2001). Self-assembling amphiphilic molecules: Synthesis in simulated interstellar/precometary ices. P. Natl. Acad. Sci. USA 98, 815819.Google Scholar
Dyudina, U. A., et al. (2010). Detection of visible lightning on Saturn. Geophys. Res. Lett. 37, doi:10.1029/2010GL043188.Google Scholar
Edgett, K. S. (2005). The sedimentary rocks of Sinus Meridiani: Five key observations from data acquired by the Mars Global Surveyor and Mars Odyssey orbiters. Mars 1, 558.Google Scholar
Edgett, K. S. and Malin, M. C. (2002). Martian sedimentary rock stratigraphy: Outcrops and interbedded craters of northwest Sinus Meridiani and southwest Arabia Terra. Geophys. Res. Lett. 29.Google Scholar
Edson, A., et al. (2011). Atmospheric circulations of terrestrial planets orbiting low mass stars,. Icarus 212, 113.Google Scholar
Edwards, C. S. and Ehlmann, B. L. (2015). Carbon sequestration on Mars. Geology, doi:10.1130/G36983.1.Google Scholar
Egami, F. (1974). Inorganic types of fermentation and anaerobic respirations in evolution of energy-yielding metabolism. Origins Life Evol. Biosph. 5, 405413.Google Scholar
Egami, F. (1976). Comment on position of nitrate respiration in metabolic evolution. Origins Life Evol. Biosph. 7, 7172.Google Scholar
Ehlmann, B. (2010). Diverse aqueous environments during Mars’ first billion years: The emerging view from orbital visible-near infrared spectroscopy. Geochem. News 142.Google Scholar
Ehlmann, B., et al. (2011). Subsurface water and clay mineral formation during the early history of Mars. Nature 479, 5360.Google Scholar
Ehlmann, B. L., et al. (2013). Geochemical consequences of widespread clay mineral formation in Mars’ ancient crust. Space Sci. Rev. 174, 329364.Google Scholar
Ehlmann, B. L. and Edwards, C. S. (2014). Mineralogy of the martian surface. Annu. Rev. Earth Pl. Sc. 42, 291315.Google Scholar
Ehlmann, B. L., et al. (2008a). Clay minerals in delta deposits and organic preservation potential on Mars. Nature Geoscience 1, 355358.Google Scholar
Ehlmann, B. L., et al. (2010). Geologic setting of serpentine deposits on Mars. Geophysical Research Letters 37, L06201, doi:10.1029/2010GL042596.Google Scholar
Ehlmann, B. L., et al. (2008b). Orbital identification of carbonate-bearing rocks on Mars. Science 322, 18281832.Google Scholar
Ehlmann, B. L., et al. (2009). Identification of hydrated silicate minerals on Mars using MRO-CRISM: Geologic context near Nili Fossae and implications for aqueous alteration. J. Geophys. Res. 114.Google Scholar
Ehrenreich, A. and Widdel, F. (1994). Anaerobic oxidation of ferrous iron by purple bacteria, a new type of phototrophic metabolism. Appl. Environ. Microbiol. 60, 45174526.Google Scholar
Eigenbrode, J. L. and Freeman, K. H. (2006). Late Archean rise of aerobic microbial ecosystems. P. Natl. Acad. Sci. USA 103, 15 75915 764.Google Scholar
Eigenbrode, J. L., et al. (2008). Methylhopane biomarker hydrocarbons in Hamersley Province sediments provide evidence for Neoarchean aerobiosis. Earth Planet. Sci. Lett. 273, 323331.Google Scholar
Eiler, J. M. (2007). “Clumped-isotope” geochemistry–The study of naturally-occurring, multiply-substituted isotopologues. Earth Planet. Sci. Lett. 262, 309327.Google Scholar
Eiler, J. M. (2011). Paleoclimate reconstruction using carbonate clumped isotope thermometry. Quaternary Sci. Rev. 30, 35753588.Google Scholar
El Albani, A., et al. (2010). Large colonial organisms with coordinated growth in oxygenated environments 2.1 Gyr ago. Nature 466, 100104.Google Scholar
El Albani, A., et al. (2014). The 2.1 Ga old Francevillian biota: Biogenicity, taphonomy and biodiversity. PLoS ONE 9, e99438 doi:10.1371/journal.pone.0099438.Google Scholar
Elachi, C. and Van Zyl, J. (2006). Introduction to the Physics and Techniques of Remote Sensing. Hoboken, N.J.: Wiley.Google Scholar
Elkins-Tanton, L. T. (2008). Linked magma ocean solidification and atmospheric growth for Earth and Mars. Earth Planet. Sc. Lett. 271, 181191.Google Scholar
Elkins-Tanton, L. T. (2012). Magma oceans in the inner Solar System. Ann. Rev. Earth Planet. Sci. 40, 113139.Google Scholar
Elliot, J. L., et al. (1998). Global warming on Triton. Nature 393, 765767.Google Scholar
Ellis, A. M., et al. (2005). Electronic and Photoelectron Spectroscopy: Fundamentals and Case Studies. New York: Cambridge University Press.Google Scholar
Ellis, R. J. (1979). Most abundant protein in the world. Trends Biochem. Sci. 4, 241244.Google Scholar
Elwood-Madden, M. E., et al. (2009). How long was Meridiani Planum wet? Applying a jarosite stopwatch to determine the duration of aqueous diagenesis. Geology 37, 635638.Google Scholar
Embleton, B. J. and Williams, G. E. (1986). Low paleolatitude of deposition for the late Precambrian periglacial varvitesin South Australia: implications for paleoclimatology. Earth Planet. Sci. Lett. 79, 419430.Google Scholar
Embley, T. M. and Williams, T. A. (2015). Steps on the road to eukaryotes. Nature 521, 169170.Google Scholar
Encrenaz, T., et al. (2004). Hydrogen peroxide on Mars: Evidence for spatial and seasonal variations. Icarus 170, 424429.Google Scholar
Encrenaz, T., et al. (2008). Simultaneous mapping of H2O and H2O2 on Mars from infrared high-resolution imaging spectroscopy. Icarus 195, 547556.Google Scholar
Endo, Y., et al. (2015). Photoabsorption cross-section measurements of S-32, S-33, S-34, and S-36 sulfur dioxide from 190 to 220 nm. J. Geophys. Res. 120, 25462557.Google Scholar
England, G. L., et al. (2002). Paleoenvironmental significance of rounded pyrite in siliclastic sequences of the Late Archean Witwatersrand Basin: Oxygen-deficient atmosphere or hydrothermal alteration? Sedimentol. 49, 11331156.Google Scholar
Eriksson, P. G. and Cheney, E. S. (1992). Evidence for the transition to an oxygen-rich atmosphere during the evolution of red beds in the lower Proterozoic sequences of southern Africa. Precamb. Res. 54, 257269.Google Scholar
Erkaev, N. V., et al. (2007). Roche lobe effects on the atmospheric loss from “Hot Jupiters”. Astron. Astrophys. 472, 329334.Google Scholar
Erwin, D. H., et al. (2011). The Cambrian Conundrum: Early divergence and later ecological success in the early history of animals. Science 334, 10911097.Google Scholar
Erwin, J., et al. (2013). Hybrid fluid/kinetic modeling of Pluto’s escaping atmosphere. Icarus 226, 375384.Google Scholar
Esposito, L. W. (1984). Sulfur dioxide: Episodic injection shows evidence for active Venus volcanism. Science 223, 10721074.Google Scholar
Etiope, G., et al. (2009). Terrestrial methane seeps and mud volcanoes: A global perspective of gas origin. Mar. Petrol. Geol. 26, 333344.Google Scholar
Etiope, G. and Klusman, R. W. (2002). Geologic emissions of methane to the atmosphere. Chemosphere 49, 777789.Google Scholar
Etiope, G., et al. (2008). Reappraisal of the fossil methane budget and related emission from geologic sources. Geophys. Res. Lett. 35.Google Scholar
Etiope, G. and Lollar, B. S. (2013). Abiotic methane on Earth. Rev. Geophys. 51, 276299.Google Scholar
Eugster, H. P. and Wones, D. R. (1962). Stability relations of the ferruginous biotite, annite. J. Petrol. 3, 82125.Google Scholar
Evans, D. A., et al. (1997). Low-latitude glaciation in the Proterozoic era. Nature 386, 262266.Google Scholar
Evans, K. A., et al. (2012). Oxidation state of subarc mantle. Geology 40, 783786.Google Scholar
Fahr, H. J. (1976). Reduced hydrogen temperatures in the transition region between thermosphere and exosphere. Ann. Geophys. 32, 277282.Google Scholar
Fahr, H. J. and Weidner, B. (1977). Gas evaporation from collision determined from planetary exospheres. Mon. Not. R. Astron. Soc. 180, 593612.Google Scholar
Fairen, A. G. (2010). A cold and wet Mars. Icarus 208, 165175.Google Scholar
Fairen, A. G., et al. (2009). Stability against freezing of aqueous solutions on early Mars. Nature 459, 401404.Google Scholar
Fairen, A. G., et al. (2003). Episodic flood inundations of the northern plains of Mars. Icarus 165, 5367.Google Scholar
Fairen, A. G., et al. (2004). Inhibition of carbonate synthesis in acidic oceans on early Mars. Nature 431, 423426.Google Scholar
Fairen, A. G., et al. (2002). An origin for the linear magnetic anomalies on Mars through accretion of terranes: Implications for dynamo timing. Icarus 160, 220223.Google Scholar
Falkowski, P. G. (1997). Evolution of the nitrogen cycle and its influence on the biological CO2 pump in the oceans. Nature 387, 272275.Google Scholar
Falkowski, P. G., et al. (2005). The rise of oxygen over the past 205 million years and the evolution of large placental mammals. Science 309, 22022204.Google Scholar
Falkowski, P. G., et al. (1992). Natural versus anthropogenic factors affecting low-level cloud albedo over the North Atlantic. Science 256, 13111313.Google Scholar
Fallick, A. E., et al. (2008). The ancient anoxic biosphere was not as we know it. In: Biosphere Origin and Evolution, ed. Dobretsov, N., et al., New York: Springer, pp. 169188.Google Scholar
Fani, R., et al. (2000). Molecular evolution of nitrogen fixation: the evolutionary history of the nifD, nifK, nifE, and nifN genes. J Mol. Evol. 51, 111.Google Scholar
Fardeau, M. L. and Belaich, J. P. (1986). Energetics of the growth of Methanococcus-Thermolithotrophicus. Arch. Microbiol. 144, 381385.Google Scholar
Farley, K. A. and Neroda, E. (1998). Noble gases in the Earth’s mantle. Ann. Rev. Earth Planet. Sci. 26, 189218.Google Scholar
Farley, K. A. and Poreda, R. J. (1993). Mantle neon and atmospheric contamination. Earth Planet. Sc. Lett. 114, 325339.Google Scholar
Farman, J. C., et al. (1985). Large losses of total ozone in Antarctica reveal seasonal ClOx/NOx interaction. Nature 315, 207210.Google Scholar
Farmer, C. B. and Houghton, J. T. (1966). Collision-induced absorption in Earth’s atmosphere. Nature 209, 1341.Google Scholar
Farquhar, J., et al. (2000a). Atmospheric influence of Earth’s earliest sulfur cycle. Science 289, 756758.Google Scholar
Farquhar, J., et al. (2008). Sulfur and oxygen isotope study of sulfate reduction in experiments with natural populations from Faellestrand, Denmark. Geochim. Cosmochim. Acta 72, 28052821.Google Scholar
Farquhar, J., et al. (2007). Implications from sulfur isotopes of the Nakhla meteorite for the origin of sulfate on Mars. Earth Planet. Sci. Lett. 264, 18.Google Scholar
Farquhar, J., et al. (2001). Observation of wavelength-sensitive mass-independent sulfur isotope effects during SO2 photolysis: application to the early atmosphere. J. Geophys. Res. 106, 111.Google Scholar
Farquhar, J., et al. (2000b). Evidence of atmospheric sulfur in the martian regolith from sulphur isotopes in meteorites. Nature 404, 5052.Google Scholar
Farquhar, J. and Wing, B. A. (2003). Multiple sulfur isotopes and the evolution of the atmosphere. Earth Planet. Sci. Lett. 213, 113.Google Scholar
Farquhar, J., et al. (2010). Connections between sulfur cycle evolution, sulfur isotopes, sediments, and base metal sulfide deposits. Econ. Geol. 105, 509533.Google Scholar
Farquhar, J., et al. (2011). Geological constraints on the origin of oxygenic photosynthesis. Photosyn. Res. 107, 1136.Google Scholar
Farquhar, J., et al. (2014). Geologic and geochemical constraints on Earth’s early atmosphere. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 91138.Google Scholar
Fassett, C. I. and Head, J. W. (2005). Fluvial sedimentary deposits on Mars: Ancient deltas in a crater lake in the Nili Fossae region. Geophys. Res. Lett. 32.Google Scholar
Fassett, C. I. and Head, J. W. (2008a). The timing of martian valley network activity: Constraints from buffered crater counting. Icarus 195, 6189.Google Scholar
Fassett, C. I. and Head, J. W. (2008b). Valley network-fed, open-basin lakes on Mars: Distribution and implications for Noachian surface and subsurface hydrology. Icarus 198, 3756.Google Scholar
Fassett, C. I. and Minton, D. A. (2013). Impact bombardment of the terrestrial planets and the early history of the Solar System. Nat. Geosci. 6, 520524.Google Scholar
Fedo, C. M. and Whitehouse, M. J. (2002). Metasomatic origin of quartz-pyroxene rock, Akilia, Greenland, and implications for Earth’s earliest life. Science 296, 14481452.Google Scholar
Fegley, B. (2012). Practical Chemical Thermodynamics for Geoscientists. Academic Press.Google Scholar
Fegley, B. (2014). Venus. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 127148.Google Scholar
Fegley, B. and Osborne, R. (2013). Practical Chemical Thermodynamics for Geoscientists. London: Academic Press.Google Scholar
Fegley, B. and Schaefer, L. (2010). Cosmochemistry of the Biogenic Elements C, H, N, O, and S. In: Astrobiology: Emergence, Search and Detection of Life, ed. Basuik, V. A., Stevenson Ranch, CA: Am. Sci. Publishers, pp. 2349.Google Scholar
Fegley, B. and Zolotov, M. Y. (2000). Chemistry of sodium, potassium, and chlorine in volcanic gases on Io. Icarus 148, 193210.Google Scholar
Feldman, P., et al. (2011). Rosetta-Alice observations of exospheric hydrogen and oxygen on Mars. Icarus 214, 394399.Google Scholar
Feldman, W. C., et al. (2004). Global distribution of near-surface hydrogen on Mars. J. Geophys. Res. 109, E09006.Google Scholar
Ferraz-Mello, S., et al. (2008). Tidal friction in close-in satellites and exoplanets: The Darwin theory revisited. Celestial Mech. Dyn. Ast. 101, 171201.Google Scholar
Ferreira, D., et al. (2011). Climate determinism revisited: Multiple equilibria in a complex climate model. J. Climate 24, 9921012.Google Scholar
Feulner, G. (2012). The faint young Sun problem. Rev. Geophys. 50.Google Scholar
Feynman, R. P., et al. (1963). The Feynman Lectures On Physics. Reading, Mass: Addison-Wesley.Google Scholar
Fiebig, J., et al. (2009). Excess methane in continental hydrothermal emissions is abiogenic. Geology 37, 495498.Google Scholar
Fike, D. A., et al. (2015). Rethinking the ancient sulfur cycle. Annu. Rev. Earth Pl. Sc. 43, 593622.Google Scholar
Fike, D. A., et al. (2006). Oxidation of the Ediacaran Ocean. Nature 444, 744747.Google Scholar
Fink, U., et al. (1972). Water vapor in the atmosphere of Venus. Icarus 17, 617631.Google Scholar
Finlayson-Pitts, J., B. J. and Pitts, J. N. (2000). Chemistry of the Upper and Lower Atmosphere : Theory, Experiments, and Applications. San Diego: Academic Press.Google Scholar
Fischer, D. A. and Valenti, J. (2005). The planet-metallicity correlation. Astrophys. J. 622, 11021117.Google Scholar
Fischer, T. P. (2008). Fluxes of volatiles (H2O, CO2, N2, Cl, F) from arc volcanoes. Geochem. J. 42, 2138.Google Scholar
Fischer, W. W., et al. (2014). Archean “whiffs of oxygen” tied to post-depositional processes. Mineral. Mag. 78.Google Scholar
Fishbaugh, K. E., et al. (2007). On the origin of gypsum in the Mars north polar region. J. Geophys. Res. 112.Google Scholar
Fisher, J. A., et al. (2005). A survey of martian dust devil activity using Mars Global Surveyor Mars Orbiter Camera images. J. Geophys. Res. 110.Google Scholar
Fishman, J. and Crutzen, P. J. (1977). Numerical study of tropospheric photochemistry using a one-dimensional model. J. Geophys. Res. 82, 58975906.Google Scholar
Flament, N., et al. (2008). A case for late-Archaean continental emergence from thermal evolution models and hypsometry. Earth Planet. Sc. Lett. 275, 326336.Google Scholar
Flannery, D. T. and Walter, M. R. (2012). Archean tufted microbial mats and the Great Oxidation Event: new insights into an ancient problem. Australian Journal of Earth Sciences 59, 111.Google Scholar
Flasar, F. M. (1983). Oceans on Titan. Science 221, 5557.Google Scholar
Flasar, F. M. (1998). The dynamic meteorology of Titan. Planet. Space Sci. 46, 11251147.Google Scholar
Flasar, F. M., et al. (2009). Atmosperic dynamics and meteorology. In: Titan from Cassini–Huygens, ed. Brown, R. H., et al., New York: Springer.Google Scholar
Flasar, F. M., et al. (1981). Titan’s atmosphere: Temperature and dynamics. Nature 292, 693698.Google Scholar
Fleagle, R. G. and Businger, J. A. (1980). An Introduction to Atmospheric Physics. New York: Academic Press.Google Scholar
Flowers, R. M. and Farley, K. A. (2012). Apatite He-4/He-3 and (U-Th)/He Evidence for an Ancient Grand Canyon. Science 338, 16161619.Google Scholar
Forbes, J., et al. (2004). Tides in the middle and upper atmospheres of Mars and Venus. Adv. Space Res. 33, 125131.Google Scholar
Forbes, J. M. (2002). Wave coupling in terrestrial planetary atmospheres. In: Atmospheres in the Solar System: Comparative Aeronomy ed. Mendillo, M., et al., Washington, D. C.: AGU, pp. 171190.Google Scholar
Forget, F. and Pierrehumbert, R. T. (1997). Warming early Mars with carbon dioxide clouds that scatter infrared radiation. Science 278, 12731276.Google Scholar
Forget, F., et al. (2013). 3D modelling of the early martian climate under a denser CO2 atmosphere: Temperatures and CO2 ice clouds. Icarus 222, 8199.Google Scholar
Formisano, V., et al. (2004). Detection of methane in the atmosphere of Mars. Science 306, 17581761.Google Scholar
Forterre, P. (2015). The universal tree of life: An update. Front. Microbiol. 6, Article 717, doi: 10.3389/fmicb.2015.00717.Google Scholar
Fortes, A. D. (2000). Exobiological implications of a possible ammonia-water ocean inside Titan. Icarus 146, 444452.Google Scholar
Fossati, L., et al. (2013). Absorbing gas around the WASP-12 planetary system. Astrophys. J. Lett. 766, L20 doi:10.1088/2041-8205/766/2/L20.Google Scholar
Fossati, L., et al. (2010). Metals in the exosphere of the highly irradiated planet WASP-12b. Astrophys. J. Lett. 714, L222L227.Google Scholar
Fowler, D., et al. (2013). The global nitrogen cycle in the twenty-first century. Phil. Trans. R. Soc. Lond. B 368.Google Scholar
Fowler, D., et al. (2009). Atmospheric composition change: Ecosystems-atmosphere interactions. Atmos. Env. 43, 51935267.Google Scholar
Fox, J. and Hac, A. (2009). Photochemical escape of oxygen from Mars: A comparison of the exobase approximation to a Monte Carlo method. Icarus 204, 527544.Google Scholar
Fox, J. L. (1993). The production and escape of nitrogen atoms on Mars. J. Geophys. Res. 98, 32973310.Google Scholar
Fox, J. L. (2007). Comment on the papers “Production of hot nitrogen atoms in the martian thermosphere” by F. Bakalian and “Monte Carlo computations of the escape of atomic nitrogen from Mars” by F. Bakalian and R.E. Hartle. Icarus 192, 296301.Google Scholar
Fox, J. L. and Bakalian, F. M. (2001). Photochemical escape of atomic carbon from Mars. J. Geophys. Res. 106, 28 78528 795.Google Scholar
Fox, J. L. and Hac, A. (1997). The 15N/14N isotope fractionation in dissociative recombination of N2+. J. Geophys. Res. 102, 91919204.Google Scholar
Frakes, L. A. (1979). Climates Throughout Geologic Time. New York: Elsevier.Google Scholar
Fralick, P. and Riding, R. (2015). Steep Rock Lake: Sedimentology and geochemistry of an Archean carbonate platform. Earth Sci. Rev. 151, 132175.Google Scholar
Franck, S., et al. (2000a). Habitable zone for Earth-like planets in the solar system. Planet. Space Sci. 48, 10991105.Google Scholar
Franck, S., et al. (2000b). Reduction of biosphere life span as a consequence of geodynamics. Tellus B 52, 94107.Google Scholar
Franck, S., et al. (1999). Modelling the global carbon cycle for the past and future evolution of the earth system. Chem. Geol. 159, 305317.Google Scholar
Francois, L. M. and Walker, J. C. G. (1992). Modelling the Phanerozoic carbon cycle and climate: Constraints from the 87Sr/86Sr isotopic ratio of seawater. Amer. J. Sci. 292, 81135.Google Scholar
Franz, H., et al. (2015). Reevaluated martian atmospheric mixing ratios from the mass spectrometer on the Curiosity rover. Planet. Space Sci. 109110, 154158.Google Scholar
Franz, H. B., et al. (2014). Isotopic links between atmospheric chemistry and the deep sulphur cycle on Mars. Nature 508, 364368.Google Scholar
Free, A. and Barton, N. H. (2007). Do evolution and ecology need the Gaia hypothesis? Trends Ecol. Evol. 22, 611619.Google Scholar
Frei, R., et al. (2009). Fluctuations in Precambrian atmospheric oxygenation recorded by chromium isotopes. Nature 461, 250–U125.Google Scholar
Freissinet, C., et al. (2015). Organic molecules in the Sheepbed Mudstone, Gale Crater. Mars. J. Geophys. Res. 120, 495514.Google Scholar
French, K. L., et al. (2015). Reappraisal of hydrocarbon biomarkers in Archean rocks. P. Natl. Acad. Sci. USA 112, 59155920.Google Scholar
Frey, H. V. (2006a). Impact constraints on the age and origin of the lowlands of Mars. Geophys. Res. Lett. 33, 14.Google Scholar
Frey, H. V. (2006b). Impact constraints on, and a chronology for, major events in early Mars history. J. Geophys. Res. 111, 111.Google Scholar
Frey, H. V., et al. (2002). Ancient lowlands on Mars. Geophys. Res. Lett. 29.Google Scholar
Friedson, A. J., et al. (2009). A global climate model of Titan’s atmosphere and surface. Planet. Space Sci. 57, 19311949.Google Scholar
Frierson, D. M. W., et al. (2006). A gray-radiation aquaplanet moist GCM. Part I: Static stability and eddy scale. J. Atmos. Sci. 63, 25482566.Google Scholar
Frierson, D. M. W., et al. (2007). Width of the Hadley cell in simple and comprehensive general circulation models. Geophys. Res. Lett. 34, doi:10.1029/2007GL031115.Google Scholar
Frost, B. R. (1991). Introduction to oxygen fugacity and its petrologic importance. In: Reviews In Mineralogy, ed. Lindsley, D. H., Washington, D.C.: Mineralogical Society of America, pp. 19.Google Scholar
Frost, D. J., et al. (2004). Experimental evidence for the existence of iron-rich metal in the Earth’s lower mantle. Nature 428, 409412.Google Scholar
Frost, D. J. and McCammon, C. A. (2008). The redox state of Earth’s mantle. Ann. Rev. Earth Planet. Sci. 36, 389420.Google Scholar
Fu, Q., et al. (2002). Tropical cirrus and water vapor: an effective Earth infrared iris feedback? Atmos. Chem. Phys. 2, 3137.Google Scholar
Fujii, Y., et al. (2010). Colors of a second Earth: Estimating the fractional areas of ocean, land, and vegetation of Earth-Like exoplanets. Astrophys. J. 715, 866880.Google Scholar
Fulchignoni, M., et al. (2005). In situ measurements of the physical characteristics of Titan’s environment. Nature 438, 785791.Google Scholar
Gaeman, J., et al. (2012). Sustainability of a subsurface ocean within Triton’s interior. Icarus 220, 339347.Google Scholar
Gaidos, E. (2013). Candidate planets in the habitable zones of Kepler stars. Astrophys. J. 770, 90, doi: 10.1088/0004-637X/770/2/90.Google Scholar
Gaidos, E. and Marion, G. (2003). Geological and geochemical legacy of a cold early Mars. J. Geophys. Res. 108.Google Scholar
Gaillard, F. and Scaillet, B. (2009). The sulfur content of volcanic gases on Mars. Earth Planet. Sci. Lett. 279, 3443.Google Scholar
Gaillard, F., et al. (2011). Atmospheric oxygenation caused by a change in volcanic degassing pressure. Nature 478, 229233.Google Scholar
Gaines, S. M., et al. (2009). Echoes of Life: What Fossil Molecules Reveal about Earth History. New York: Oxford University Press.Google Scholar
Galli, A., et al. (2006a). Energetic hydrogen and oxygen atoms observed on the nightside of Mars. Space Sci. Rev. 126, 267296.Google Scholar
Galli, A., et al. (2006b). The hydrogen exospheric density profile measured with ASPERA-3/NPD. Space Sci. Rev. 126, 447467.Google Scholar
Galtier, N., et al. (1999). A non-hyperthermophilic common ancestor to extant life forms. Science, 283, 220221.Google Scholar
Ganguli, S. B. (1996). The polar wind. Rev. Geophys. 34, 311348.Google Scholar
Garcia, R. R. and Solomon, S. (1985). The effect of breaking gravity waves on the dynamics and chemical composition of the mesosphere and lower thermosphere. J. Geophys. Res. 92, 38503868.Google Scholar
Garcia-Munoz, A. (2007). Physical and chemical aeronomy of HD 209458b. Planet. Space Sci. 55, 14261455.Google Scholar
Garnier, P., et al. (2008). The lower exosphere of Titan: Energetic neutral atoms absorption and imaging. J. Geophys. Res. 113, A10216, doi:10.1029/2008JA013029.Google Scholar
Garrels, R. M., et al. (1973). Genesis of Precambrian iron formations and development of atmospheric oxygen. Econ. Geol. 68, 11731179.Google Scholar
Garvin, J., et al. (2009). Isotopic evidence for an aerobic nitrogen cycle in the latest Archean. Science 323, 10451048.Google Scholar
Gaucher, E. A., et al. (2008). Palaeotemperature trend for Precambrian life inferred from resurrected proteins. Nature 451, 704707.Google Scholar
Gaucher, E. A., et al. (2003). Inferring the palaeoenvironment of ancient bacteria on the basis of resurrected proteins. Nature 425, 285288.Google Scholar
Gear, C. W. (1971). Numerical Initial Value Problems In Ordinary Differential Equations. Englewood Cliffs, N.J.: Prentice-Hall.Google Scholar
Geboy, N. J., et al. (2013). Re-Os age constraints and new observations of Proterozoic glacial deposits in the Vazante Group, Brazil. Precambrian Research 238, 199213.Google Scholar
Geminale, A., et al. (2011). Mapping methane in Martian atmosphere with PFS-MEX data. Planet. Space Sci. 59, 137148.Google Scholar
Genda, H. and Abe, Y. (2003). Survival of a proto-atmosphere through the stage of giant impacts: The mechanical aspects. Icarus 164, 149162.Google Scholar
Genda, H. and Abe, Y. (2005). Enhanced atmospheric loss on protoplanets at the giant impact phase in the presence of oceans. Nature 433, 842844.Google Scholar
Gendrin, A., et al. (2005). Sulfates in martian layered terrains: the OMEGA/Mars Express view. Science 307, 15871591.Google Scholar
Gensel, P. G. (2008). The Earliest Land Plants. Ann. Rev. Ecol. Evol. Sys. 39, 459477.Google Scholar
Genthner, B. R. S. and Bryant, M. P. (1982). Growth of eubacterium: Limosum with carbon monoxide as the energy source. Appl. Environ. Microbiol. 43, 7074.Google Scholar
George, S. C., et al. (2008). Preservation of hydrocarbons and biomarkers in oil trapped inside fluid inclusions for > 2 billion years. Geochim. Cosmochim. Acta 72, 844870.Google Scholar
Gerasimov, M. V. and Mukhin, L. M. (1984). Studies of the chemical composition of gaseous phase released from laser pulse evaporated rocks and meteorite materials. Proc. Lunar Planet. Sci. Conf. XV, 298–299.Google Scholar
Gerlach, T. M. (2011). Volcanic versus anthropogenic carbon dioxide. EOS Trans. AGU 92, 201202.Google Scholar
German, C. R. and Seyfried, W. E. (2014). Hydrothermal processes. In: Treatise on Geochemistry, 2nd Edn, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 181222.Google Scholar
Ghatan, G. J. and Zimbelman, J. R. (2006). Paucity of candidate coastal constructional landforms along proposed shorelines on Mars: Implications for a northern lowlands-filling ocean. Icarus 185, 171196.Google Scholar
Gierasch, P. J. (1975). Meridional circulation and maintenance of Venus atmospheric rotation. J. Atmos. Sci. 32, 10381044.Google Scholar
Gierasch, P. J., et al. (1997). The general circulation of the Venus atmosphere: An assessment. In: Venus II, ed. Bougher, S. W., et al., Tucson: University of Arizona Press, pp. 459500.Google Scholar
Giggenbach, W. (1997). Relative importance of thermodynamic and kinetic processes in governing the chemical and isotopic composition of carbon gases in high-heatflow sedimentary basins. Geochim. Cosmochim. Acta 61, 37633785.Google Scholar
Giggenbach, W. F. and Matsuo, S. (1991). Evaluation of results from Second and Third IAVCEI Field Workshops on Volcanic Gases. Appl. Geochem. 6, 125141.Google Scholar
Giggenbach, W. H. (1996). Chemical composition of volcanic gases. In: Monitoring and Mitigation of Volcano Hazards, ed. Scarpa, R. and Tilling, R. I., Berlin: Springer-Verlag, pp. 221256.Google Scholar
Gillessen, S., et al. (2009). Monitoring stellar orbits around the massive black hole in the galactic center. Astrophys. J. 692, 10751109.Google Scholar
Giorgi, F. and Chameides, W. L. (1985). The rainout parameterization in a photochemical model. J. Geophys. Res. 90, 78727880.Google Scholar
Gladman, B. (1993). Dynamics of systems of 2 close planets. Icarus 106, 247263.Google Scholar
Gladstone, G. R., et al. (2016). The atmosphere of Pluto as observed by New Horizons. Science 351, 1280, doi:10.1126/science.aad8866.Google Scholar
Glasspool, I. J. and Scott, A. C. (2010). Phanerozoic concentrations of atmospheric oxygen reconstructed from sedimentary charcoal. Nat. Geosci. 3, 627630.Google Scholar
Glavin, D. P., et al. (2013). Evidence for perchlorates and the origin of chlorinated hydrocarbons detected by SAM at the Rocknest aeolian deposit in Gale Crater. J. Geophys. Res. 118, 19551973.Google Scholar
Glein, C. R., et al. (2015). The pH of Enceladus’ ocean. Geochim. Cosmochim. Acta 162, 202219.Google Scholar
Glein, C. R., et al. (2009). The absence of endogenic methane on Titan and its implications for the origin of atmospheric nitrogen. Icarus 204, 637644.Google Scholar
Glein, C. R. and Shock, E. L. (2013). A geochemical model of non-ideal solutions in the methane-ethane-propane-nitrogen-acetylene system on Titan. Geochim. Cosmochim. Acta 115, 217240.Google Scholar
Glotch, T. D., et al. (2010). Distribution and formation of chlorides and phyllosilicates in Terra Sirenum, Mars. Geophysical Research Letters 37,.Google Scholar
Glotch, T. D. and Rogers, A. D. (2007). Evidence for aqueous deposition of hematite- and sulfate-rich light-toned layered deposits in Aureum and Iani Chaos, Mars. J. Geophys. Res. 112, E06001, doi:10.1029/2006JE002863.Google Scholar
Godderis, Y., et al. (2007). Coupled modeling of global carbon cycle and climate in the Neoproterozoic: links between Rodinia breakup and major glaciations. Comptes Rendus Geoscience 339, 212222.Google Scholar
Godderis, Y., et al. (2003). The Sturtian ‘snowball’ glaciation: fire and ice. Earth and Planetary Science Letters 211, 112.Google Scholar
Godderis, Y. and Veizer, J. (2000). Tectonic control of chemical and isotopic composition of ancient oceans: The impact of continental growth. Am. J. Sci. 300, 434461.Google Scholar
Godfrey, L. V. and Falkowski, P. G. (2009). The cycling and redox state of nitrogen in the Archaean ocean. Nature Geosci., doi: 10.1038/NGEO633.Google Scholar
Gogarten-Boekels, M., et al. (1995). The effects of heavy meteorite bombardment on the early evolution: The emergence of the three domains of life. Orig. Life and Evol. Biosph. 25, 251264.Google Scholar
Golabek, G. J., et al. (2011). Origin of the martian dichotomy and Tharsis from a giant impact causing massive magmatism. Icarus 215, 346357.Google Scholar
Gold, T. (1964). Outgassing processes on the Moon and Venus. In: The Origin and Evolution of Atmospheres and Oceans, ed. Brancazio, P. J. and Cameron, A. G. W., New York: Wiley, pp. 249256.Google Scholar
Gold, T. and Soter, S. (1969). Atmospheric tides and resonant rotation of Venus. Icarus 11, 356366.Google Scholar
Gold, T. and Soter, S. (1971). Atmospheric tides and 4-day circulation on Venus. Icarus 14, 1620.Google Scholar
Goldblatt, C., et al. (2009). Nitrogen-enhanced greenhouse warming on early Earth. Nat. Geosci. 2, 891896.Google Scholar
Goldblatt, C., et al. (2006). Bistability of atmospheric oxygen and the Great Oxidation. Nature 443, 683686.Google Scholar
Goldblatt, C., et al. (2013). Low simulated radiation limit for runaway greenhouse climates. Nat. Geosci. 6, 661667.Google Scholar
Goldblatt, C. and Watson, A. J. (2012). The runaway greenhouse: implications for future climate change, geoengineering and planetary atmospheres. Phil. Trans. R. Soc. Lond. A 370, 41974216.Google Scholar
Goldblatt, C. and Zahnle, K. J. (2011). Clouds and the faint young Sun paradox. Clim. Past 7, 203220.Google Scholar
Golden, D. C., et al. (2008). Hydrothermal synthesis of hematite spherules and jarosite: Implications for diagenesis and hematite spherule formation in sulfate outcrops at Meridiani Planum, Mars. Am. Mineral. 93, 12011214.Google Scholar
Goldreich, P. (1966). History of the lunar orbit. Rev. Geophys. 4, 411439.Google Scholar
Goldreich, P. and Soter, S. (1966). Q in Solar System. Icarus 5, 375389.Google Scholar
Goldspiel, J. M. and Squyres, S. W. (2000). Groundwater sapping and valley formation on Mars. Icarus 148, 176192.Google Scholar
Golombek, M. P., et al. (2006). Erosion rates at the Mars Exploration Rover landing sites and long-term climate change on Mars. J. Geophys. Res. 111.Google Scholar
Gomes, R., et al. (2005). Origin of the cataclysmic Late Heavy Bombardment period of the terrestrial planets. Nature 435, 466469.Google Scholar
Gonzalez, G. (1999). Is the Sun anomalous? Astron. Geophys. 40, 2529.Google Scholar
Gonzalez, G., et al. (2001). The Galactic Habitable Zone: Galactic chemical evolution. Icarus 152, 185200.Google Scholar
Goodman, J. C. (2006). Through thick and thin: Marine and meteoric ice in a “Snowball Earth” climate. Geophys. Res. Lett. 33.Google Scholar
Goodman, J. C. and Pierrehumbert, R. T. (2003). Glacial flow of floating marine ice in ‘Snowball Earth’. J. Geophys. Res. 108, 3308.Google Scholar
Goody, R. M. (1964). Atmospheric Radiation. Oxford: Clarendon Press.Google Scholar
Goody, R. M. and Belton, M. J. S. (1967). Radiative relaxation times for Mars: a discussion of Martian atmospheric dynamics. Planet. Space Sci. 15, 247–56.Google Scholar
Goody, R. M. and Walker, J. C. G. (1972). Atmospheres. Englewood Cliffs, NJ: Prentice Hall.Google Scholar
Goody, R. M. and Yung, Y. L. (1989). Atmospheric Radiation: Theoretical Basis. New York: Oxford University Press.Google Scholar
Gough, D. O. (1981). Solar interior structure and luminosity variations. Solar Phys. 74, 2134.Google Scholar
Grad, H. (1949). On the kinetic theory of rarefied gases. Comm. Pure Appl. Math. 2, 331407.Google Scholar
Gradie, J. and Tedesco, E. (1982). Compositional structure of the asteroid belt. Science 216, 14051407.Google Scholar
Gradstein, F. M. (2012). The Geologic Time Scale 2012. Boston: Elsevier.Google Scholar
Graedel, T. E. and Crutzen, P. J. (1993). Atmospheric Change: An Earth System Perspective. New York: W.H. Freeman.Google Scholar
Graedel, T. E., et al. (1991). Early solar mass loss: A potential solution to the weak sun paradox. Geophys. Res. Lett. 18, 18811884.Google Scholar
Graham, D. W. (2002). Noble gas isotope geochemistry of mid-ocean ridge and ocean island basalts: Characterization of mantle source reservoirs. Rev. Mineral. Geochem. 47, 247317.Google Scholar
Grandstaff, D. E. (1980). Origin of uraniferous conglomerates at Elliot Lake, Canada and Witwatersrand, South-Africa – Implications for oxygen in the Precambrian atmosphere. Precam. Res. 13, 126.Google Scholar
Grant, J. A., et al. (2008). HiRISE imaging of impact megabreccia and sub-meter aqueous strata in Holden Crater, Mars. Geology 36, 195198.Google Scholar
Grant, J. A. and Wilson, S. A. (2011). Late alluvial fan formation in southern Margaritifer Terra, Mars. Geophy. Res. Lett. 38.Google Scholar
Greeley, R., et al. (1997). Aeolian processes and features on Venus. In: Venus II, ed. Bougher, S. W., et al., Tucson: University of Arizona Press, pp. 547589.Google Scholar
Greeley, R., et al. (2005). Fluid lava flows in Gusev crater. Mars. J. Geophys. Res. 110, E05008.Google Scholar
Greeley, R. and Iversen, J. D. (1985). Wind as a Geological Process on Earth, Mars, Venus, and Titan. New York: Cambridge University Press.Google Scholar
Greenwood, J. P., et al. (2008). Hydrogen isotope evidence for loss of water from Mars through time. Geophys. Res. Lett 35.Google Scholar
Grenfell, T. C. and Warren, S. G. (1999). Representation of a nonspherical ice particle by a collection of independent spheres for scattering and absorption of radiation. J. Geophys. Res. 104, 31 69731 709.Google Scholar
Grevesse, N., et al. (2007). The solar chemical composition. Space Sci. Rev. 130, 105114.Google Scholar
Griessmeier, J. M., et al. (2005). Cosmic ray impact on extrasolar earth-like planets in close-in habitable zones. Astrobiology 5, 587603.Google Scholar
Griffith, C. A. (2009). Storms, polar deposits and the methane cycle in Titan’s atmosphere. Phil. Trans. R. Soc. Lond. A 367, 713728.Google Scholar
Griffith, C. A., et al. (2012). Possible tropical lakes on Titan from observations of dark terrain. Nature 486, 237239.Google Scholar
Griffith, C. A., et al. (2008). Titan’s tropical storms in an evolving atmosphere. Astrophys. J. Lett. 687, L41L44.Google Scholar
Griffith, C. A., et al. (2006). Evidence for a polar ethane cloud on Titan. Science 313, 16201622.Google Scholar
Griffith, C. A. and Zahnle, K. (1995). Influx of cometary volatiles to planetary Moons: The atmospheres of 1000 possible Titans. J. Geophys. Res. 100, 1690716922.Google Scholar
Grima, C., et al. (2009). North polar deposits of Mars: Extreme purity of the water ice. Geophysical Research Letters 36.Google Scholar
Grimm, R. E., et al. (2014). Water budgets of martian recurring slope lineae. Icarus 233, 316327.Google Scholar
Grinspoon, D. (1987). Was Venus wet? Deuterium reconsidered. Science 238, 17021704.Google Scholar
Grinspoon, D. H. (1993). Implications of the high D/H ratio for the sources of water in Venus’ atmosphere. Nature 363, 428431.Google Scholar
Groller, H., et al. (2014). Hot oxygen and carbon escape from the martian atmosphere. Planet. Space Sci. 98, 93105.Google Scholar
Gross, J., et al. (2013). Petrography, mineral chemistry, and crystallization history of olivine-phyric shergottite NWA 6234: A new melt composition. Meteoritics & Planetary Science 48, 854871.Google Scholar
Gross, S. H. (1972). Exospheric temperature of hydrogen-dominated planetary atmospheres. J. Atmos. Sci. 29, 214218.Google Scholar
Gross, S. H. (1974). Atmospheres of Titan and Galilean satellites. J. Atmos. Sci. 31, 14131420.Google Scholar
Grotzinger, J. P., et al. (2005). Stratigraphy and sedimentology of a dry to wet eolian depositional system, Burns formation, Meridiani Planum, Mars. Earth and Planetary Science Letters 240, 1172.Google Scholar
Grotzinger, J. P., et al. (2015a). Curiosity’s mission of exploration at Gale Crater, Mars. Elements 11, 1926.Google Scholar
Grotzinger, J. P., et al. (2011). Enigmatic origin of the largest-known carbon isotope excursion in Earth’s history. Nature Geoscience 4, 285292.Google Scholar
Grotzinger, J. P., et al. (2015b). Deposition, exhumation, and paleoclimate of an ancient lake deposit, Gale crater. Mars. Science 350, doi: 10.1126/science.aac7575.Google Scholar
Grotzinger, J. P. and Knoll, A. H. (1999). Stromatolites in Precambrian carbonates: Evolutionary mileposts or environmental dipsticks? Annu. Rev. Earth Planet. Sci. 27, 313358.Google Scholar
Grotzinger, J. P. and Rothman, D. H. (1996). An abiotic model for stromatolite morphogenesis. Nature 383, 423425.Google Scholar
Grotzinger, J. P., et al. (2014). A habitable fluvio-lacustrine environment at Yellowknife Bay, Gale Crater, Mars. Science 343, 1242777, doi: 10.1126/science.1242777.Google Scholar
Gruber, N. and Sarmiento, J. L. (1997). Global patterns of marine nitrogen fixation and denitrification. Global Biogeochem. Cyc. 11, 235266.Google Scholar
Grundy, W. M. and Buie, M. W. (2001). Distribution and evolution of CH4, N2, and CO ices on Pluto’s surface: 1995 to 1998. Icarus 153, 248263.Google Scholar
Grundy, W. M., et al. (2007). New horizons mapping of Europa and Ganymede. Science 318, 234237.Google Scholar
Guillot, T. (1999). Interiors of giant planets inside end outside the solar system. Science 286, 7277.Google Scholar
Guillot, T. and Gautier, D. (2015). Giant planets. In: Treatise on Geophysics (Second Edition), ed. Schubert, G., Oxford: Elsevier, pp. 529557.Google Scholar
Guillot, T., et al. (2004). The interior of Jupiter. In: Jupiter: The Planet, Satellites, and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press, pp. 3557.Google Scholar
Gulbis, A. A. S., et al. (2006). Charon’s radius and atmospheric constraints from observations of a stellar occultation. Nature 439, 4851.Google Scholar
Gulick, V. C. (1998). Magmatic intrusions and a hydrothermal origin for fluvial valleys on Mars. J. Geophys. Res. 103, 1936519387.Google Scholar
Gulick, V. C. (2001). Origin of the valley networks on Mars: a hydrological perspective. Geomorphology 37, 241268.Google Scholar
Gunning, H. E. and Strausz, O. P. (1966). The reactions of sulfur atoms. In: Advan. Photochem., ed. Noyes, W. A., et al., Hoboken, NJ: Wiley, pp. 143194.Google Scholar
Gutowsky, H. S. (1976). Halocarbons: Effects on Stratospheric Ozone. National Academy of Sciences.Google Scholar
Guzik, J. A., et al. (1987). A comparison between mass-losing and standard solar models. Ap. J. 319, 957965.Google Scholar
Haberle, R. M. (2013). Estimating the power of Mars’ greenhouse effect. Icarus 223, 619620.Google Scholar
Haberle, R. M., et al. (2017). Early Mars. In: The Atmosphere and Climate of Mars, ed. Haberle, R. M. et al., New York: Cambridge University Press.Google Scholar
Haberle, R. M., et al. (2008). The effect of ground ice on the Martian seasonal CO2 cycle. Planet. Space Sci. 56, 251255.Google Scholar
Haberle, R. M., et al. (1993a). Atmospheric effects on the utility of solar power on Mars In: Resources of Near-Earth Space, ed. Lewis, J., Matthews, M. S., Tucson: University of Arizona Press, pp. 845885.Google Scholar
Haberle, R. M., et al. (2001). On the possibility of liquid water on present-day Mars. J. Geophys. Res. 106, 23 31723 326.Google Scholar
Haberle, R. M., et al. (2003). Orbital change experiments with a Mars general circulation model. Icarus 161, 6689.Google Scholar
Haberle, R. M., et al. (1993b). Mars atmospheric dynamics as simulated by the NASA Ames General Circulation Model. 1. The zonal-mean circulation. J. Geophys. Res. 98, 30933123.Google Scholar
Habicht, K. S., et al. (2002). Calibration of sulfate levels in the Archean ocean. Science 298, 23722374.Google Scholar
Hadley, G. (1735). Concerning the cause of the general trade-winds. Phil. Trans. R. Soc. 39, 5862.Google Scholar
Haisch, K. E., et al. (2001). Disk frequencies and lifetimes in young clusters. Astrophys. J. 553, L153L156.Google Scholar
Haldane, J. B. S. (1928). Possible Worlds and Other Papers. London: Harper & Brothers Publishers.Google Scholar
Haldane, J. B. S. (1929). The origin of life. Rationalist Annual 148, 310.Google Scholar
Halevy, I. (2013). Production, preservation, and biological processing of mass-independent sulfur isotope fractionation in the Archean surface environment. P. Natl. Acad. Sci. USA 110, 17 64417 649.Google Scholar
Halevy, I. and Head, J. W. (2014). Episodic warming of early Mars by punctuated volcanism. Nat. Geosci. 7, 865868.Google Scholar
Halevy, I., et al. (2010). Explaining the structure of the Archean mass-independent sulfur isotope record. Science 329, 204207.Google Scholar
Halevy, I., et al. (2007). A sulfur dioxide climate feedback on early Mars. Science 318, 19031907.Google Scholar
Hall, D. T., et al. (1998). The far-ultraviolet oxygen airglow of Europa and Ganymede. Astrophys. J. 499, 475481.Google Scholar
Halliday, A. N. (2013). The origins of volatiles in the terrestrial planets. Geochim. Cosmochim. Acta 105, 146171.Google Scholar
Hallis, L. J., et al. (2012). Magmatic water in the martian meteorite Nakhla. Earth Planet. Sc. Lett. 359, 8492.Google Scholar
Hallmann, C. and Summons, R. E. (2014). Paleobiological clues to early atmospheric evolution. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., 2nd edn, New York: Elsevier, pp. 139155.Google Scholar
Halmer, M. M., et al. (2002). The annual volcanic gas input into the atmosphere, in particular into the stratosphere: a global data set for the past 100 years. J. Volcanol. Geotherm. Res. 115, 511528.Google Scholar
Halverson, G. P. (2006). A Neoproterozoic chronology. In: Neoproterozoic Geobiology and Paleobiology, ed. Xiao, S. and Kaufman, A. J., New York: Kluwer, pp. 231271.Google Scholar
Halverson, G. P., et al. (2005). Toward a Neoproterozoic composite carbon-isotope record. GSA Bull. 117, 11811207.Google Scholar
Halverson, G. P. and Hurtgen, M. T. (2007). Ediacaran growth of the marine sulfate reservoir. Earth Planet. Sci. Lett. 263, 3244.Google Scholar
Halverson, G. P., et al. (2010). Neoproterozoic chemostratigraphy. Precambrian Research 182, 337350.Google Scholar
Hamano, K., et al. (2013). Emergence of two types of terrestrial planet on solidification of magma ocean. Nature 497, 607610.Google Scholar
Han, T. M. (1988). Origin of magnetite in Precambrian iron-formations of low metamorphic grade. In: Proc. Seventh IAGOD Symposium, ed. Zachrisson, E., Stuttgart: E. Schweizerbart, pp. 641656.Google Scholar
Han, T. M. and Runnegar, B. (1992). Megascopic eukaryotic algae from the 2.1 billion-year-old Negaunee iron-formation, Michigan. Science 257, 232235.Google Scholar
Han, T. T., et al. (2011). Global features and trends of the tropopause derived from GPS/CHAMP RO data. Sci. China: Phys. Mech. Astron. 54, 365374.Google Scholar
Hanel, R. A. (1981). Fourier spectroscopy on planetary missions including Voyager. Proc. Soc. Photo-Opt. Instrum. Eng. 289, 331344.Google Scholar
Hanel, R. A., et al. (1981). Albedo, internal heat, and energy balance of Jupiter: Preliminary results of the Voyager infrared investigation. J. Geophys. Res. 86, 87058712.Google Scholar
Hanel, R. A., et al. (1972). Nimbus 4 infrared spectroscopy experiment. 1. Calibrated thermal emission spectra. J. Geophys. Res. 77, 26292641.Google Scholar
Hansen, B. M. S. (2009). Formation of the terrestrial planets from a narrow annulus. Astrophys. J. 703, 11311140.Google Scholar
Hansen, C. J. and Paige, D. A. (1996). Seasonal nitrogen cycles on Pluto. Icarus 120, 247265.Google Scholar
Hansen, C. J., et al. (2011). The composition and structure of the Enceladus plume. Geophys. Res. Lett. 38.Google Scholar
Hansen, J., et al. (2008). Target atmospheric CO2: Where should humanity aim? Open Access Atmos. Sci. J. 2, 217231.Google Scholar
Hapke, B. (2012). Theory of Reflectance and Emittance Spectroscopy. New York: Cambridge University Press.Google Scholar
Hapke, B. and Nelson, R. (1975). Evidence for an elemental sulfur component of clouds from Venus spectrophotometry. J. Atmos. Sci. 32, 12121218.Google Scholar
Haqq-Misra, J. D., et al. (2008). A revised, hazy methane greenhouse for the Archean Earth. Astrobiol. 8, 11271137.Google Scholar
Haqq-Misra, J., et al. (2016). Limit cycles can reduce the width of the habitable zone. Astrophys. J. 827, doi:10.3847/0004–637x/827/2/120.Google Scholar
Hardisty, D. S., et al. (2014). An iodine record of Paleoproterozoic surface ocean oxygenation. Geology 42, 619622.Google Scholar
Harland, W. B. and Rudwick, M. J. S. (1964). The great infra-cambrian ice age. Sci. Am. 211, 2836.Google Scholar
Harman, C. E., et al. (2015). Abiotic O2 levels on planets aroudn F, G, K, and M stars: Possible false positives for life? Astrophys. J., in press.Google Scholar
Harnmeijer, J. P. (2009). Squeezing blood from a stone: Inferences into the life and depositional environments of the Early Archaean. University of Washington, PhD thesis, Seattle.Google Scholar
Harper, C. L. and Jacobsen, S. B. (1996). Noble gases and Earth’s accretion. Science 273, 18141818.Google Scholar
Harries, J., et al. (1996). On the distribution of mesospheric molecular hydrogen inferred from HALOE measurements of H2O and CH4. Geophy. Res. Lett. 23, 297300.Google Scholar
Harrison, J. F., et al. (2010). Atmospheric oxygen level and the evolution of insect body size. Proc. R. Soc. Lond. B 277, 19371946.Google Scholar
Harrison, T. M. (2009). The Hadean crust: Evidence from > 4 Ga zircons. Ann. Rev. Earth Planet. Sci. 37, 479505.Google Scholar
Hart, M. H. (1975). Explanation for absence of extraterrestrials on Earth. Q. J. Roy. Astron. Soc. 16, 128135.Google Scholar
Hart, M. H. (1978). The evolution of the atmosphere of the Earth. Icarus 33, 2339.Google Scholar
Hart, M. H. (1979). Habitable zones around main sequence stars. Icarus 37, 351357.Google Scholar
Hart, M. H. (1982). Atmospheric evolution, the Drake Equation, and DNA: Sparse life in an infinite Universe. In: In Extraterrestrials: Where Are They?, ed. Zuckerman, M. H. H. a. B., New York: Pergamon Press, pp. 154164.Google Scholar
Harteck, P. and Jensen, J. H. D. (1948). Uber Den Sauerstoffgehalt Der Atmosphare (On the oxygen content of the atmosphere). Z. Naturforsch. A 3, 591595.Google Scholar
Hartle, R. E. and Grebowsky, J. M. (1990). Upward ion flow in ionospheric holes on Venus. J. Geophys. Res. 95, 3137.Google Scholar
Hartle, R. E., et al. (2006). Initial interpretation of Titan plasma interaction as observed by the Cassini plasma spectrometer: Comparisons with Voyager 1. Planet. Space Sci. 54, 12111224.Google Scholar
Hartman, H. and McKay, C. P. (1995). Oxygenic photosynthesis and the oxidation state of Mars. Planet. Space Sci. 43, 123128.Google Scholar
Hartmann, D. L. (1994). Global Physical Climatology. San Diego: Academic Press.Google Scholar
Hartmann, L., et al. (2005). IRAC observations of Taurus pre-main-sequence stars. Astrophys. J. 629, 881896.Google Scholar
Hartmann, W. K. and Davis, D. R. (1975). Satellite-sized planetesimals and lunar origin. Icarus 24, 504515.Google Scholar
Hartmann, W. K., et al. (2000). The time-dependent intense bombardment of the primordial Earth/Moon system. In: Origin of the Moon and Earth, ed. Canup, R. M. and Righter, K., Tucson, AZ: University of Arizona Press, pp. 493512.Google Scholar
Hartmann, W. K. and Neukum, G. (2001). Cratering chronology and the evolution of Mars. Space Sci. Rev. 96, 165194.Google Scholar
Hartogh, P., et al. (2010). Herschel/HIFI observations of Mars: First detection of O2 at submillimetre wavelengths and upper limits on HCl and H2O2. Astron. Astrophys. 521.Google Scholar
Hartogh, P., et al. (2011a). Direct detection of the Enceladus water torus with Herschel. Astron. Astrophys. 532, article ID L2.Google Scholar
Hartogh, P., et al. (2011b). Ocean-like water in the Jupiter-family comet 103P/Hartley 2. Nature 478, 218220.Google Scholar
Hashimoto, G. L. and Abe, Y. (2005). Climate control on Venus: Comparison of the carbonate and pyrite models. Planet. Space Sci. 53, 839848.Google Scholar
Hashimoto, G. L., et al. (2007). The chemical composition of the early terrestrial atmosphere: Formation of a reducing atmosphere from CI-like material. J. Geophys. Res. 112, E05010.Google Scholar
Haskin, L. A. (1998). The Imbrium impact event and the thorium distribution at the lunar highlands surface. J. Geophys. Res. 103, 16791689.Google Scholar
Haswell, C. A. (2010). Transiting Exoplanets. Cambridge: Cambridge University Press.Google Scholar
Haswell, C. A., et al. (2012). Near-ultraviolet absorption, chromospheric activity, and star-planet interactions in the WASP-12 system. Astrophys. J. 760, 79: doi:10.1088/0004-637X/760/1/79.Google Scholar
Hattori, S., et al. (2011). Ultraviolet absorption cross sections of carbonyl sulfide isotopologues (OCS)-S-32, (OCS)-S-33, (OCS)-S-34 and (OCS)-C-13: isotopic fractionation in photolysis and atmospheric implications. Atmos. Chem. Phys. 11, 10 29310 303.Google Scholar
Hauber, E., et al. (2013). Asynchronous formation of Hesperian and Amazonian-aged deltas on Mars and implications for climate. J. Geophys. Res. 118, 15291544.Google Scholar
Hauck, S. A. and Phillips, R. J. (2002). Thermal and crustal evolution of Mars. J. Geophys. Res. 107.Google Scholar
Hawkesworth, C., et al. (2009). A matter of preservation. Science 323, 4950.Google Scholar
Hawkesworth, C. J., et al. (2010). The generation and evolution of the continental crust. J. Geol. Soc. 167, 229248.Google Scholar
Hayashi, C., et al. (1979). Earths melting due to the blanketing effect of the primordial dense atmosphere. Earth. Planet. Sc. Lett. 43, 2228.Google Scholar
Hayashi, C., et al. (1985). Formation of the Solar System. In: Protostars and Protoplanets II, ed. Black, D. C. and Matthews, M. S., Tucson: University of Arizona Press, pp. 11001153.Google Scholar
Hayes, A. G. (2016). The lakes and seas of Titan. Annu. Rev. Earth Planet. Sci. 44, 5783.Google Scholar
Hayes, J. M. (1983). Geochemical evidence bearing on the origin of aerobiosis, a speculative hypothesis. In: Earth’s Earliest Biosphere: Its Origin and Evolution, ed. Schopf, J. W., Princeton, New Jersey: Princeton University Press, pp. 291301.Google Scholar
Hayes, J. M. (1994). Global methanotrophy at the Archean-Preoterozoic transition. In: Early Life on Earth, ed. Bengtson, S., New York: Columbia University Press, pp. 220236.Google Scholar
Hayes, J. M. and Waldbauer, J. R. (2006). The carbon cycle and associated redox processes through time. Phil. Trans. R. Soc. Lond. B 361, 931950.Google Scholar
Hays, P. B. and Liu, V. C. (1965). On loss of gases from a planetary atmosphere. Planet. Space Sci. 13, 11851211.Google Scholar
Head, J. W., et al. (2002). Northern lowlands of Mars: Evidence for widespread volcanic flooding and tectonic deformation in the Hesperian Period. J. Geophys. Res. 107, 5003.Google Scholar
Head, J. W., et al. (2003). Recent ice ages on Mars. Nature 426, 797802.Google Scholar
Hebrard, E. and Marty, B. (2014). Coupled noble gas-hydrocarbon evolution of the early Earth atmosphere upon solar UV irradiation. Earth Planet. Sc. Lett. 385, 4048.Google Scholar
Hecht, M. H. (2002). Metastability of liquid water on Mars. Icarus 156, 373386.Google Scholar
Hecht, M. H., et al. (2009). Detection of perchlorate and the soluble chemistry of martian soil at the Phoenix Lander site. Science 325, 6467.Google Scholar
Hedelt, P., et al. (2010). Titan’s atomic hydrogen corona. Icarus 210, 424435.Google Scholar
Heimann, A., et al. (2010). Fe, C, and O isotope compositions of banded iron formation carbonates demonstrate a major role for dissimilatory iron reduction in similar to 2.5 Ga marine environments. Earth Planet. Sci. Lett. 294, 818.Google Scholar
Heimpel, M. and Aurnou, J. (2007). Turbulent convection in rapidly rotating spherical shells: A model for equatorial and high latitude jets on Jupiter and Saturn. Icarus 187, 540557.Google Scholar
Heimpel, M., et al. (2005). Simulation of equatorial and high-latitude jets on Jupiter in a deep convection model. Nature 438, 193196.Google Scholar
Heising, S., et al. (1999). Chlorobium ferrooxidans sp. nov., a phototrophic green sulfur bacterium that oxidizes ferrous iron in coculture with a “Geospirillum” sp. strain. Arch. Microbiol. 172, 116124.Google Scholar
Heising, S. and Schink, B. (1998). Phototrophic oxidation of ferrous iron by a Rhodomicrobium vannielii strain. Microbiol. 144, 22632269.Google Scholar
Held, I. M. (1975). Momentum transport by quasi-geostrophic eddies. J. Atmos. Sci. 32, 14941497.Google Scholar
Held, I. M. (2000). General circulation of the atmosphere. In: 2000 Program in Geophysical Fluid Dynamics, ed. Thiffeault, J.-L., Woods Hole, MA: Woods Hole Oceanographic Institution, pp. 154 (http://www.whoi.edu/page.do?pid=13076).Google Scholar
Held, I. M. and Hou, A. Y. (1980). Non-linear axially-symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci. 37, 515533.Google Scholar
Held, I. M. and Phillips, P. J. (1990). A barotropic model of the Interaction between the Hadley-Cell and a Rossby wave. J. Atmos. Sci. 47, 856869.Google Scholar
Heldmann, J. L., et al. (2007). Observations of martian gullies and constraints on potential formation mechanisms II. The northern hemisphere. Icarus 188, 324344.Google Scholar
Heldmann, J. L. and Mellon, M. T. (2004). Observations of martian gullies and constraints on potential formation mechanisms. Icarus 168, 285304.Google Scholar
Helled, R., et al. (2011). Jupiter’s moment of inertia: A possible determination by Juno. Icarus 216, 440448.Google Scholar
Helled, R., et al. (2014). Giant planet formation, evolution, and internal structure. In: Protostars and Planets VI, ed. Beuther, H., et al., Tucson, AZ: Univ. of Arizona Press, pp. 643665.Google Scholar
Helled, R., et al. (2008). Grain sedimentation in a giant gaseous protoplanet. Icarus 195, 863870.Google Scholar
Helling, C. and Casewell, S. (2014). Atmospheres of brown dwarfs. Astron. Astrophys. Rev. 22, 80.Google Scholar
Hemming, N. G. and Honisch, B. (2007). Boron isotopes in marine carbonate sediments adn the pH of the ocean. In: Proxies in Late Cenozoic Paleoceanography, ed. de Vernal, A., Hillaire-Marcel, C., Amsterdam: Elsevier, pp. 717734.Google Scholar
Henderson, P. and Henderson, G. M. (2009). The Cambridge Handbook of Earth Science Data. New York: Cambridge University Press.Google Scholar
Heng, K., et al. (2014). Analytical models of exoplanetary atmospheres. II. Radiative transfer via the two-stream approximation. Astrophys. J. Suppl. S. 215, 4, doi:10.1088/0067-0049/215/1/4.Google Scholar
Henkes, G. A., et al. (2014). Temperature limits for preservation of primary calcite clumped isotope paleotemperatures. Geochim. Cosmochim. Acta 139, 362382.Google Scholar
Herd, C. D. K. (2003). The oxygen fugacity of olivine-phyric martian basalts and the components within the mantle and crust of Mars. Meteorit. Planet. Sci. 38, 17931805.Google Scholar
Herd, C. D. K. (2006). Insights into the redox history of the NWA 1068/1110 martian basalt from mineral equilibria and vanadium oxybarometry. Am. Min. 91, 16161627.Google Scholar
Herrick, R. R., et al. (1997). Morphology and morphometry of impact craters. In: Venus II: Geology, Geophysics, Atmospghere, and Solar Wind Environment, Tucson: University of Arizona Press, pp. 10151046.Google Scholar
Herwartz, D., et al. (2014). Identification of the giant impactor Theia in lunar rocks. Science 344, 11461150.Google Scholar
Herzberg, C., et al. (2010). Thermal history of the Earth and its petrological expression. Earth Planet. Sc. Lett. 292, 7988.Google Scholar
Hide, R. (1969). Dynamics of atmospheres of major planets with an appendix on viscous boundary layer at rigid bounding surface of an electrically-conducting rotating fluid in presence of a magnetic field. J. Atmos. Sci. 26, 841853.Google Scholar
Hide, R. (1970). Equatorial jets in planetary atmospheres. Nature 225, 254255.Google Scholar
Higgs, P. G. and Lehman, N. (2015). The RNA world: molecular cooperation at the origins of life. Nat. Rev. Genet. 16, 717.Google Scholar
Hillier, J. K., et al. (2007). The composition of Saturn’s E ring. Mon. Not. R. Astron. Soc. 377, 15881596.Google Scholar
Hinrichs, K. U. (2002). Microbial fixation of methane carbon at 2.7 Ga: Was an anaerobic mechanism possible? Geochem. Geophys. Geosys. 3, doi:10.1029/2001GC000286.Google Scholar
Hinrichs, K. U. and Boetius, A. (2002). The anaerobic oxidation of methane: New insights in microbial ecology and biogeochemistry. In: Ocean Margin Systems, ed. Wefer, G., et al., Berlin: Springer-Verlag, pp. 457477.Google Scholar
Hinrichs, K. U., et al. (1999). Methane-consuming archaebacteria in marine sediments. Nature 398, 802805.Google Scholar
Hitchcock, D. R. and Lovelock, J. E. (1967). Life detection by atmospheric analysis. Icarus 7, 149.Google Scholar
Hitzman, M. W., et al. (2010). Formation of sedimentary rock-hosted stratiform copper deposits through Earth history. Econ. Geol. 105, 627639.Google Scholar
Hoashi, M., et al. (2009). Primary haematite formation in an oxygenated sea 3.46 billion years ago. Nat. Geosci. 2, 301306.Google Scholar
Hobbs, P. V. (2000). Introduction to Atmospheric Chemistry. New York: Cambridge University Press.Google Scholar
Hodges, R. R. (1999). An exospheric perspective of isotopic fractionation of hydrogen on Venus. J. Geophys. Res. 104, 84638471.Google Scholar
Hodges, R. R. and Tinsley, B. A. (1981). Charge exchange in the Venus ionosphere as the source of exospheric hydrogen. J. Geophys. Res. 86, 76497656.Google Scholar
Hoeppe, G. (2007). Why the Sky is Blue? Discovering the Color of Life. Princeton, N.J.: Princeton University Press.Google Scholar
Hoffert, M. I. and Covey, C. (1992). Deriving global climate sensitivity from paleoclimate reconstructions. Nature 360, 573576.Google Scholar
Hoffman, P. F. (2013). The Great Oxidation and a Siderian snowball Earth: MIF-S based correlation of Paleoproterozoic glacial epochs. Chemical Geology 362, 143156.Google Scholar
Hoffman, P. F., et al. (1998). A Neoproterozoic Snowball Earth. Science 281, 13421346.Google Scholar
Hoffman, P. F. and Li, Z. X. (2009). A palaeogeographic context for Neoproterozoic glaciation. Palaeogeogr. Palaeocl. 277, 158172.Google Scholar
Hoffman, P. F. and Schrag, D. P. (2002). The Snowball Earth hypothesis: Testing the limits of global change. Terra Nova 14, 129155.Google Scholar
Hoffmann, H. J. (1976). Precambrian microflora, Belcher Islands, Canada: Significance and systematics. J. Paleontol. 50, 10501073.Google Scholar
Hofgartner, J. D., et al. (2014). Transient features in a Titan sea. Nature Geoscience 7, 493496.Google Scholar
Hofgartner, J. D. and Lunine, J. I. (2013). Does ice float in Titan’s lakes and seas? Icarus 223, 628631.Google Scholar
Hofmann, A., et al. (2009). Multiple sulphur and iron isotope composition of detrital pyrite in Archaean sedimentary rocks: A new tool for provenance analysis. Earth Planet. Sci. Lett. 286, 436445.Google Scholar
Hofmann, H. J., et al. (1999). Origin of 3.45 Ga coniform stromatolites in Warrawoona Group, Western Australia. Geol. Soc. Am. Bull. 111, 12561262.Google Scholar
Hoinka, K. P. (1998). Statistics of the global tropopause pressure. Mon. Weather Rev. 126, 33033325.Google Scholar
Hoke, M. R. T., et al. (2011). Formation timescales of large Martian valley networks. Earth Planet. Sc. Lett. 312, 112.Google Scholar
Holland, G., et al. (2009). Meteorite Kr in Earth’s mantle suggests a late accretionary source for the atmosphere. Science 326, 15221525.Google Scholar
Holland, H. D. (1962). Model for the evolution of the Earth’s atmosphere. In: Petrologic Studies: A Volume to Honor A.F. Buddington, ed. Engel, A. E. J., et al., New York: Geol. Soc. Am., pp. 447477.Google Scholar
Holland, H. D. (1973a). Ocean water, nutrients, and atmospheric oxygen through geologic time. In: Proceedings of Symposium of Hydrogeochemistry and Biogeochemistry, ed. Ingerson, E., Washington, DC: The Clark Co., pp. 6881.Google Scholar
Holland, H. D. (1973b). Oceans: Possible source of iron in iron-formations. Econ. Geol. 68, 11691172.Google Scholar
Holland, H. D. (1978). The Chemistry of the Atmosphere and Oceans. New York: Wiley.Google Scholar
Holland, H. D. (1979). Metals in black shales: Reassessment. Econ. Geol. 74, 16761680.Google Scholar
Holland, H. D. (1984). The Chemical Evolution of the Atmosphere and Oceans. Princeton: Princeton University Press.Google Scholar
Holland, H. D. (2002). Volcanic gases, black smokers, and the Great Oxidation Event. Geochim. Cosmochim. Acta 66, 38113826.Google Scholar
Holland, H. D. (2006). The oxygenation of the atmosphere and oceans. Phil. Trans. R. Soc. Lond. B 361, 903915.Google Scholar
Holland, H. D. (2009). Why the atmosphere became oxygenated: A proposal. Geochim. Cosmochim. Acta 73, 52415255.Google Scholar
Holland, H. D. (2011). Discovering the history of atmospheric oxygen. In: Frontiers in Geochemistry: Contribution of Geochemistry to the Study of the Earth, ed. Harmon, R. S. and Parker, A., Oxford: Blackwell Publishing Ltd., pp. 4360.Google Scholar
Holland, H. D. and Zbinden, E. A. (1988). Paleosols and the evolution of the atmosphere: part I. In: Physical and Chemical Weathering in Geochemical Cycles, ed. Lerman, A. and Meybeck, M., Dordrecht: Reidel, pp. 6182.Google Scholar
Hollingsworth, J. L. and Barnes, J. R. (1996). Forced stationary planetary waves in Mars’s winter atmosphere. J. Atmos. Sci. 53, 428448.Google Scholar
Hollingsworth, J. L., et al. (2007). A simple-physics global circulation model for Venus: Sensitivity assessments of atmospheric superrotation. Geophys. Res. Lett. 34.Google Scholar
Holloway, J. R. (2004). Redox reactions in seafloor basalts: possible insights into silicic hydrothermal systems. Chem. Geol. 210, 225230.Google Scholar
Holman, M. J. and Murray, N. W. (2005). The use of transit timing to detect terrestrial-mass extrasolar planets. Science 307, 12881291.Google Scholar
Holser, W. T., et al. (1988). Geochemical cycles of carbon and sulfur. In: Chemical Cycles in the Evolution of the Earth, ed. Gregor, C. B., et al., New York: Wiley, pp. 105173.Google Scholar
Holton, J. R. (2004). An Introduction to Dynamic Meteorology. New York: Elsevier Academic Press.Google Scholar
Holton, J. R. and Hakim, G. J. (2013). An Introduction to Dynamic Meteorology (5th Edition). Amsterdam: Elsevier, Academic Press.Google Scholar
Holzer, T. E., et al. (1971). Comparison of kinetic and hydrodynamic models of an expanding ion exosphere. J. Geophys. Res. 76, 2453.Google Scholar
Hood, A. V. S. and Wallace, M. W. (2014). Marine cements reveal the structure of an anoxic, ferruginous Neoproterozoic ocean. J. Geol. Soc. London 171, 741744.Google Scholar
Hopkins, M., et al. (2008). Low heat flow inferred from > 4 Gyr zircons suggests Hadean plate boundary interactions. Nature 456, 493496.Google Scholar
Hopper, J. P. and Leverington, D. W. (2014). Formation of Hrad Vallis (Mars) by low viscosity lava flows. Geomorphology 207, 96113.Google Scholar
Horandl, E. and Hadacek, F. (2013). The oxidative damage initiation hypothesis for meiosis. Plant Reprod. 26, 351367.Google Scholar
Horita, J. (2005). Some perspectives on isotope biosignatures for early life. Chem. Geol. 218, 171186.Google Scholar
Horita, J. and Berndt, M. E. (1999). Abiogenic methane formation and isotopic fractionation under hydrothermal conditions. Science 285, 10551057.Google Scholar
Horner, J., et al. (2009). Differences between the impact regimes of the terrestrial planets: Implications for primordial D:H ratios. Planet. Space Sci. 57, 13381345.Google Scholar
Horst, S. M., et al. (2008). Origin of oxygen species in Titan’s atmosphere. J. Geophys. Res. 113.Google Scholar
Houghton, J. T. (2002). The Physics of Atmospheres. New York: Cambridge University Press.Google Scholar
Houghton, J. T., et al. (1995). Climate Change, 1994: Radiative Forcing Of Climate Change And An Evaluation Of The IPCC IS92 Emission Scenarios. New York: Cambridge University. Press.Google Scholar
House, C. H., et al. (2003a). Geobiological analysis using whole genome-based tree building applied to the Bacteria, Archea, and Eukarya. Geobiology 1, 1526.Google Scholar
House, C. H., et al. (2003b). Carbon isotopic fractionation by Archaeans and other thermophilic prokaryotes. Org. Geochem. 34, 345356.Google Scholar
Howard, A. D. (2007). Simulating the development of Martian highland landscapes through the interaction of impact cratering, fluvial erosion, and variable hydrologic forcing. Geomorphology 91, 332363.Google Scholar
Howard, A. D., et al. (2005). An intense terminal epoch of widespread fluvial activity on early Mars: 1. Valley network incision and associated deposits. J. Geophys. Res. 110.Google Scholar
Howard, A. W. (2013). Observed properties of extrasolar planets. Science 340, 572576.Google Scholar
Howett, C. J. A., et al. (2010). Thermal inertia and bolometric Bond albedo values for Mimas, Enceladus, Tethys, Dione, Rhea and Iapetus as derived from Cassini/CIRS measurements. Icarus 206, 573593.Google Scholar
Hren, M. T., et al. (2009). Oxygen and hydrogen isotope evidence for a temperate climate 3.42 billion years ago. Nature 462, 205208.Google Scholar
Hsu, H. W., et al. (2015). Ongoing hydrothermal activities within Enceladus. Nature 519, 207210.Google Scholar
Hu, R. Y., et al. (2012). Photochemistry in terrestrial exoplanet atmospheres. I. Photochemistry model and benchmark cases. Astrophys. J. 761.Google Scholar
Huang, S. S. (1959). Occurrence of life in the universe. Amer. Scientist 47, 397402.Google Scholar
Huang, S. S. (1960). Life outside the solar system. Scientific American 202, 5563.Google Scholar
Hubbard, W. B., et al. (1995). The interior of Neptune. In: Neptune and Triton, ed. Cruikshank, D. P., Tucson: Univ. of Arizona Press, pp. 109140.Google Scholar
Hubbard, W. B., et al. (1993). The occultation of 28 Sgr by Titan. Astronomy & Astrophysics 269, 541563.Google Scholar
Huber, C. and Wachtershauser, G. (1997). Activated acetic acid by carbon fixation on (Fe,Ni)S under primordial conditions. Science 276, 245247.Google Scholar
Huber, C. and Wachtershauser, G. (1998). Peptides by activation of amino acids with CO on (Ni,Fe) surfaces: implications for the origin of life. Science 281, 670672.Google Scholar
Hulburt, E. O. (1953). Explanation of the brightness and color of the sky, particularly the twilight sky. J. Opt. Soc. Am. 43, 113118.Google Scholar
Hulme, M. (1995). Estimating global change in precipitation. Weather 50, 3442.Google Scholar
Humayun, M., et al. (2013). Origin and age of the earliest Martian crust from meteorite NWA 7533. Nature 503, 513516.Google Scholar
Hunt, B. G. (1979). Effects of past variations of the Earth’s rotation rate on climate. Nature 281, 188191.Google Scholar
Hunten, D. M. (1973). The escape of light gases from planetary atmospheres. J. Atmos. Sci. 30, 14811494.Google Scholar
Hunten, D. M. (1975). Estimates of stratospheric pollution by an analytic model. Proc. Nat. Acad. Sci. 72, 4711.Google Scholar
Hunten, D. M. (1979a). Capture of Phobos and Deimos by protatmospheric drag. Icarus 37, 113123.Google Scholar
Hunten, D. M. (1979b). Possible oxidant sources in the atmosphere and surface of Mars. J Mol. Evol. 14, 7178.Google Scholar
Hunten, D. M. (1979c). Possible oxidant sources in the atmosphere and surface of Mars. J. Mol. Evol. 14, 7178.Google Scholar
Hunten, D. M. (1990). Kuiper Prize Lecture: Escape of atmospheres, ancient and modern. Icarus 85, 120.Google Scholar
Hunten, D. M. (2002). Exospheres and planetary escape. In: Aeronomic systems on planets, moons, and comets, ed. Mendillo, M., et al., Washington, D. C.: AGU, pp. 191202.Google Scholar
Hunten, D. M. (2006). The sequestration of ethane on Titan in smog particles. Nature 443, 669670.Google Scholar
Hunten, D. M. and Donahue, T. M. (1976). Hydrogen loss from the terrestrial planets. Ann. Rev. Earth Planet Sci. 4, 265292.Google Scholar
Hunten, D. M., et al. (1989). Escape of atmospheres and loss of water. In: Origin And Evolution Of Planetary And Satellite Atmospheres, ed. Atreya, S. K., et al., Tucson, AZ: University of Arizona.Google Scholar
Hunten, D. M., et al. (1987). Mass fractionation in hydrodynamic escape. Icarus 69, 532549.Google Scholar
Hunten, D. M. and Strobel, D. F. (1974). Production and escape of terrestrial hydrogen. J. Atmos. Sci. 31, 305317.Google Scholar
Hunten, D. M. and Watson, A. J. (1982). Stability of Pluto’s atmosphere. Icarus 51, 665667.Google Scholar
Hurtgen, M. T., et al. (2005). Neoproterozoic sulfur isotopes, the evolution of microbial sulfur species, and the burial efficiency of sulfide as sedimentary pyrite. Geology 33, 4144.Google Scholar
Hurtgen, M. T., et al. (2002). The sulfur isotopic composition of Neoproterozoic seawater sulfate: implications for a snowball Earth? Earth Planet. Sci. Lett. 203, 413429.Google Scholar
Hussmann, H., et al. (2006). Subsurface oceans and deep interiors of medium-sized outer planet satellites and large trans-neptunian objects. Icarus 185, 258273.Google Scholar
Huston, D. L. and Logan, G. A. (2004). Barite, BIFs and bugs: evidence for the evolution of the Earth’s early hydrosphere. Earth Planet. Sci. Lett. 220, 4155.Google Scholar
Hutchins, K. S. and Jakosky, B. M. (1996). Evolution of Martian atmospheric argon: Implications for sources of volatiles. J. Geophys. Res. 101, 1493314949.Google Scholar
Hutchins, K. S., et al. (1997). Impact of a paleomagnetic field on sputtering loss of Martian atmospheric argon and neon. J. Geophys. Res. 102, 91839189.Google Scholar
Hyde, W. T., et al. (2000). Neoproterozoic ‘snowball Earth’ simulations with a coupled climate/ice-sheet model. Nature 405, 425429.Google Scholar
Hynek, B. M., et al. (2010). Updated global map of martian valley networks and implications for climate and hydrologic processes. J. Geophys. Res. 115, E09008, doi:10.1029/2009JE003548.Google Scholar
Hynek, B. M. and Phillips, R. J. (2008). The stratigraphy of Meridiani Planum, Mars, and implications for the layered deposits’ origin. Earth Planet. Sci. Lett. 274, 214220.Google Scholar
Hynek, B. M., et al. (2003). Explosive volcanism in the Tharsis region: Global evidence in the Martian geologic record. J. Geophys. Res. 108, 5111, doi:10.1029/2003JE002062.Google Scholar
Iess, L., et al. (2012). The tides of Titan. Science 337, 457459.Google Scholar
Iess, L., et al. (2010). Gravity field, shape, and moment of inertia of Titan. Science 327, 13671369.Google Scholar
Ignatiev, N. I., et al. (1997). Water vapour in the lower atmosphere of Venus: A new analysis of optical spectra measured by entry probes. Adv. Space Res. 19, 11591168.Google Scholar
Iizuka, T., et al. (2015). Meteorite zircon constraints on the bulk Lu-Hf isotope composition and early differentiation of the Earth. P. Natl. Acad. Sci. USA 112, 53315336.Google Scholar
Imbrie, J. and Imbrie, K. P. (1979). Ice Ages: Solving the Mystery. Cambridge, MA: Harvard University Press.Google Scholar
Inaba, S. and Ikoma, M. (2003). Enhanced collisional growth of a protoplanet that has an atmosphere. Astron. Astrophys. 410, 711723.Google Scholar
Ingersoll, A. B. and Cuzzi, J. N. (1969). Dynamics of Jupiter’s cloud bands. J. Atmos. Sci. 26, 981985.Google Scholar
Ingersoll, A. P. (1969). The runaway greenhouse: A history of water on Venus. J. Atmos. Sci. 26, 11911198.Google Scholar
Ingersoll, A. P. (1970). Mars: Occurrence of liquid water. Science 168, 972973.Google Scholar
Ingersoll, A. P. (2013). Planetary Climates. Princeton, N.J.: Princeton Uniersity Press.Google Scholar
Ingersoll, A. P. and Dobrovolskis, A. R. (1978). Venus rotation and atmospheric tides. Nature 275, 3738.Google Scholar
Ingersoll, A. P., et al. (2004). Dynamics of Jupiter’s atmosphere. In: Jupiter: The Planet, Satellites, and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press.Google Scholar
Ingersoll, A. P. and Ewald, S. P. (2011). Total particulate mass in Enceladus plumes and mass of Saturn’s E ring inferred from Cassini ISS images. Icarus 216, 492506.Google Scholar
Irwin, P. (2009). Giant Planets of Our Solar System: Atmospheres, Composition, and Structure. Chichester, UK: Springer.Google Scholar
Irwin, R. P., et al. (2005). An intense terminal epoch of widespread fluvial activity on early Mars: 2. Increased runoff and paleolake development. J. Geophys. Res. 110.Google Scholar
Irwin, R. P., et al. (2015). Paleohydrology of Eberswalde crater, Mars. Geomorphology 240, 83101.Google Scholar
Ishimaru, R., et al. (2011). Oxidizing proto-atmosphere on Titan: Constraint from N-2 formation by impact shock. Astrophys. J. Lett. 741.Google Scholar
Isley, A. E. (1995). Hydrothermal plumes and the delivery of iron to banded iron-formation. J. Geol. 103, 169185.Google Scholar
Israel, G., et al. (2005). Complex organic matter in Titan’s atmospheric aerosols from in situ pyrolysis and analysis. Nature 438, 796799.Google Scholar
Jackson, B., et al. (2008a). Tidal heating of terrestrial extrasolar planets and implications for their habitability. Mon. Not. R. Astron. Soc. 391, 237245.Google Scholar
Jackson, B., et al. (2008b). Tidal heating of extrasolar planets. Astrophys. J. 681, 16311638.Google Scholar
Jackson, B., et al. (2010). The roles of tidal evolution and evaporative mass loss in the origin of CoRoT-7 b. Mon. Not. R. Astron. Soc. 407, 910922.Google Scholar
Jacob, D. J. (1999). Introduction to Atmospheric Chemistry. Princeton, N.J.: Princeton University Press.Google Scholar
Jacobsen, S. B., et al. (2008). Isotopes as clues to the origin and earliest differentiation history of the Earth. Phil. Trans. R. Soc. Lond. A 366, 41294162.Google Scholar
Jacobson, M. Z. (2005). Fundamentals of Atmospheric Modeling (2nd Edn). New York: Cambridge University Press.Google Scholar
Jacovi, R. and Bar-Nun, A. (2008). Removal of Titan’s noble gases by their trapping in its haze. Icarus 196, 302304.Google Scholar
Jaeger, W. L., et al. (2007). Athabasca Valles, Mars: A lava-draped channel system. Science 317, 17091711.Google Scholar
Jaeger, W. L., et al. (2010). Emplacement of the youngest flood lava on Mars: A short, turbulent story. Icarus 205, 230243.Google Scholar
Jakosky, B. M. and Farmer, C. B. (1982). The seasonal and global behaviour of water vapour in the Mars atmosphere: Complete global results of the Viking atmospheric water detector experiment. J. Geophys. Res. 87, 29993019.Google Scholar
Jakosky, B. M., et al. (2015). MAVEN observations of the response of Mars to an interplanetary coronal mass ejection. Science 350, doi: 10.1126/science.aad0210.Google Scholar
Jakosky, B. M. and Jones, J. H. (1997). The history of Martian volatiles. Rev. Geophys. 35, 116.Google Scholar
Jakosky, B. M. and Mellon, M. T. (2004). Water on Mars. Phys. Today 57, 7176.Google Scholar
Jakosky, B. M., et al. (1994). Mars atmospheric loss and isotopic fractionation by solar-wind-induced sputtering and photochemical escape. Icarus 111, 271288.Google Scholar
Jakosky, B. M. and Phillips, R. J. (2001). Mars’ volatile and climate history. Nature 412, 237244.Google Scholar
Janssen, M. A. (1993). Atmospheric Remote Sensing by Microwave Radiometry. New York: Wiley.Google Scholar
Jansson, K. W. and Johansen, A. (2014). Formation of pebble-pile planetesimals. Astron. Astrophys. 570.Google Scholar
Jarrard, R. D. (2003). Subduction fluxes of water, carbon dioxide, chlorine, and potassium. Geochem. Geophys. Geosys. 4, doi: 10.1029/2002gc000392.Google Scholar
Jaumann, R., et al. (2009). Geology and surface processes on Titan. In: Titan from Cassini-Huygens, ed. Brown, R. H., et al., New York: Springer, pp. 75140.Google Scholar
Javaux, E. J., et al. (2001). Morphological and ecological complexity in early eukaryotic ecosystems. Nature 412, 6669.Google Scholar
Javaux, E. J., et al. (2004). TEM evidence for eukaryotic diversity in mid-Proterozoic oceans. Geobiology 2, 121132.Google Scholar
Javoy, M., et al. (2010). The chemical composition of the Earth: Enstatite chondrite models. Earth Planet. Sci. Lett. 293, 259268.Google Scholar
Jeans, J. H. (1954). The Dynamical Theory of Gases. New York: Dover.Google Scholar
Jenkins, G. S. (1993). A general circulation model study of the effects of faster rotation, enhanced CO2 concentrations, and solar forcing: implications for the Faint Young Sun Paradox. J. Geophys. Res. 98, 20 80320 811.Google Scholar
Jenkins, G. S. (1996). A sensitivity study of changes in Earth’s rotation rate with an atmospheric general circulation model. Global and Planet. Change 11, 141154.Google Scholar
Jenkins, G. S., et al. (1993). Precambrian climate: the effects of land area and Earth’s rotation rate. J. Geophys. Res. 98, 87858791.Google Scholar
Jennings, D. E., et al. (2009). Titan’s surface brightness temperatures. Astrophys. J. Lett. 691, L103L105.Google Scholar
Jensen, S., et al. (2000). Complex trace fossils from the terminal Proterozoic of Namibia. Geology 28, 143146.Google Scholar
Jerolmack, D. J., et al. (2004). A minimum time for the formation of Holden Northeast fan, Mars. Geophys. Res. Lett. 31.Google Scholar
Ji, Q., et al. (2002). The earliest known eutherian mammal. Nature 416, 816822.Google Scholar
Jia, Y. F. and Kerrich, R. (2004). Nitrogen 15-enriched Precambrian kerogen and hydrothermal systems. Geochem. Geophys. Geosys. 5, doi:10.1029/2004GC000716.Google Scholar
Johansen, A., et al. (2014). The multifaceted planetesimal formation process. In: Protostars and Planets VI, ed. Beuther, H., et al., Tucson, AZ: University of Arizona Press, pp. 571594.Google Scholar
Johansen, A., et al. (2015). Growth of asteroids, planetary embryos, and Kuiper belt objects by chondrule accretion. Sci. Adv. 1, e1500109, doi:10.1126/sciadv.1500109.Google Scholar
Johnson, A. P., et al. (2008a). The Miller volcanic spark discharge experiment. Science 322, 404404.Google Scholar
Johnson, B. and Goldblatt, C. (2015). The nitrogen budget of Earth. Earth Sci. Rev. 148, 150173.Google Scholar
Johnson, B. C. and Melosh, H. J. (2012). Impact spherules as a record of an ancient heavy bombardment of Earth. Nature 485, 7577.Google Scholar
Johnson, C. M., et al. (2008b). The iron isotope fingerprints of redox and biogeochemical cycling in the modern and ancient Earth. Ann. Rev. Earth Planet. Sci. 36, 457493.Google Scholar
Johnson, C. M., et al. (2013a). Iron formation carbonates: Paleoceanographic proxy or recorder of microbial diagenesis? Geology 41, 11471150.Google Scholar
Johnson, H. E. and Axford, W. I. (1969). Production and loss of He3 in Earth’s atmosphere. J. Geophys. Res. 74, 2433.Google Scholar
Johnson, J. A., et al. (2010). Giant planet cccurrence in the stellar mass-metallicity plane. Publ. Astron. Soc. Pac. 122, 905915.Google Scholar
Johnson, J. E., et al. (2014). O2 constraints from Paleoproterozoic detrital pyrite and uraninite. Bull. Geol. Soc. Am. 126, 813830.Google Scholar
Johnson, J. E., et al. (2013b). Manganese-oxidizing photosynthesis before the rise of cyanobacteria. P. Natl. Acad. Sci. USA 110, 11 23811 243.Google Scholar
Johnson, J. E., et al. (2013c). Reply to Jones and Crowe: Correcting mistaken views of sedimentary geology, Mn-oxidation rates, and molecular clocks. P. Natl. Acad. Sci. USA 110, E4119E4120.Google Scholar
Johnson, R. E., et al. (2009). Composition and detection of Europa’s sputter-induced atmosphere. In: Europa, ed. Pappalardo, R. T., et al., Tucson: University of Arizona Press, pp. 507527.Google Scholar
Johnson, R. E., et al. (2004). Radiation effects on the surfafces of the Galilean satellites. In: Jupiter: The Planet, Satellites, and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press.Google Scholar
Johnson, R. E., et al. (2008c). Exospheres and atmospheric escape. Space Sci. Rev. 139, 355397.Google Scholar
Johnson, R. E., et al. (2002). Energy distributions for desorption of sodium and potassium from ice: The Na/K ratio at Europa. Icarus 156, 136142.Google Scholar
Johnson, R. E., et al. (2013d). Molecular-kinetic simulations of escape from the ex-planet and exoplanets: Criterion for transonic flow. Astrophys. J. Lett. 768.Google Scholar
Johnson, T. M. and Bullen, T. (2004). Mass-dependent fractionation of selenium and chromium isotopes in low-temperature environments. Rev. Mineral. Geochem. 55, 289317.Google Scholar
Johnston, D. T., et al. (2010). An emerging picture of Neoproterozoic ocean chemistry: Insights from the Chuar Group, Grand Canyon, USA. Earth Planet. Sci. Lett. 290, 6473.Google Scholar
Johnston, D. T., et al. (2005). Active microbial sulfur disproportionation in the Mesoproterozoic. Science 310, 14771479.Google Scholar
Jones, C. and Crowe, S. A. (2013). No evidence for manganese-oxidizing photosynthesis. P. Natl. Acad. Sci. USA 110, E4118.Google Scholar
Jones, T. D. and Lewis, J. S. (1987). Estimated impact shock production of N2 and organic compounds on early Titan. Icarus 72, 381393.Google Scholar
Jöns, H.-P. (1985). Late sedimentation and late sediments in the northern lowlands on Mars. Lunar Planet. Sci. Conf. XVI, 414–415.Google Scholar
Jorgensen, U. G., et al. (2009). The Earth-Moon system during the late heavy bombardment period - Geochemical support for impacts dominated by comets. Icarus 204, 368380.Google Scholar
Joseph, J. H., et al. (1976). Delta-Eddington approximation for radiative flux transfer. J. Atmos. Sci. 33, 24522459.Google Scholar
Joshi, M. (2003). Climate model studies of synchronously rotating planets. Astrobiology 3, 415427.Google Scholar
Joshi, M. M. and Haberle, R. M. (2012). Suppression of the water ice and snow albedo feedback on planets orbiting red dwarf stars and the subsequent widening of the habitable zone. Astrobiology 12, 38.Google Scholar
Joshi, M. M., et al. (1997). Simulations of the atmospheres of synchronously rotating terrestrial planets orbiting M dwarfs: Conditions for atmospheric collapse and the implications for habitability. Icarus 129, 450465.Google Scholar
Jouannic, G., et al. (2015). Laboratory simulation of debris flows over sand dunes: Insights into gully-formation (Mars). Geomorphology 231, 101115.Google Scholar
Joyce, G. F. (1989). RNA evolution and the origins of life. Nature 338, 217224.Google Scholar
Joyce, G. F. (1994). Foreward. In: Origins of life : the central concepts, ed. Deamer, D. W., Fleischaker, G. R., Boston: Jones and Bartlett Publishers, pp. xixii.Google Scholar
Junge, C. E., et al. (1975). Model calculations for the terrestrial carbon cycle: Carbon isotope geochemistry and evolution of photosynthetic oxygen. J. Geophys. Res. 80, 45424552.Google Scholar
Kadoya, S. and Tajika, E. (2014). Conditions for oceans on Earth-Like planets orbiting within the habitable zone: Importance of volcanic CO2 degassing. Astrophys. J. 790, 107.Google Scholar
Kah, L. C., et al. (2004). Low marine sulphate and protracted oxygenation of the Proterozoic biosphere. Nature 431, 834838.Google Scholar
Kah, L. C. and Riding, R. (2007). Mesoproterozoic carbon dioxide levels inferred from calcified cyanobacteria. Geology 35, 799802.Google Scholar
Kahn, R. A., et al. (1992). The martian dust cycle. In: Mars, ed. Kieffer, H. H., et al., Tucson: University of Arizona Press, pp. 10171053.Google Scholar
Kaib, N. A. and Chambers, J. E. (2015). The fragility of the terrestrial planets during a giant planet instability. Mon. Not. R. Astron. Soc., in press.Google Scholar
Kanzaki, Y. and Murakami, T. (2015). Estimates of atmospheric CO2 in the Neoarchean-Paleoproterozoic from paleosols. Geochim. Cosmochim. Acta 159, 190219.Google Scholar
Kappler, A. and Newman, D. K. (2004). Formation of Fe(III)-minerals by Fe(II)-oxidizing photoautotrophic bacteria. Geochim. Cosmochim. Acta 68, 12171226.Google Scholar
Kargel, J. S. and Lewis, J. S. (1993). The composition and early evolution of Earth. Icarus 105, 125.Google Scholar
Karhu, J. A. and Holland, H. D. (1996). Carbon isotopes and the rise of atmospheric oxygen. Geology 24, 867870.Google Scholar
Karkoschka, E. (1994). Spectrophotometry of the Jovian planets and Titan at 300 nm to 1000-nm wavelength: The Methane Spectrum. Icarus 111, 174192.Google Scholar
Karkoschka, E. (1998). Methane, ammonia, and temperature measurements of the Jovian planets and Titan from CCD-spectrophotometry. Icarus 133, 134146.Google Scholar
Karlsson, N. B., et al. (2015). Volume of Martian midlatitude glaciers from radar observations and ice flow modeling. Geophys. Res. Lett. 42, 26272633.Google Scholar
Karlstrom, K. E., et al. (2014). Formation of the Grand Canyon 5 to 6 million years ago through integration of older palaeocanyons. Nat. Geosci. 7, 239244.Google Scholar
Karttunen, H. (2007). Fundamental astronomy. Berlin: Springer-Verlag.Google Scholar
Kasemann, S. A., et al. (2005). Boron and calcium isotope composition in Neoproterozoic carbonate rocks from Namibia: evidence for extreme environmental change. Earth Planet. Sc. Lett. 231, 7386.Google Scholar
Kasemann, S. A., et al. (2010). Neoproterozoic ice ages, boron isotopes, and ocean acidification: Implications for a snowball Earth. Geology 38, 775778.Google Scholar
Kashefi, K. and Lovley, D. R. (2003). Extending the upper temperature limit for life. Science 301, 934934.Google Scholar
Kasper, M., et al. (2010). EPICS: direct imaging of exoplanets with the E-ELT. SPIE Astronomical Telescopes+ Instrumentation. International Society for Optics and Photonics, pp. 77352E–77352E-9.Google Scholar
Kaspi, Y. and Showman, A. P. (2015). Atmospheric dynamics of terrestrial exoplanets over a wide range of orbital and atmospheric parameters. Astrophys. J. 804, 60, doi:10.1088/0004-637x/804/1/60.Google Scholar
Kasting, J. F. (1982). Stability of ammonia in the primitive terrestrial atmosphere. J. Geophys. Res. 87, 30913098.Google Scholar
Kasting, J. F. (1985). Photochemical consequences of enhanced CO2 levels in Earth’s early atmosphere. In: The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present, ed. Sundquist, E. T. and Broecker, W. S., Washington D.C.: American Geophysical Union, pp. 612622.Google Scholar
Kasting, J. F. (1987). Theoretical constraints on oxygen and carbon dioxide concentrations in the Precambrian atmosphere. Precambrian Res. 34, 205229.Google Scholar
Kasting, J. F. (1988). Runaway and moist greenhouse atmospheres and the evolution of Earth and Venus. Icarus 74, 472494.Google Scholar
Kasting, J. F. (1990). Bolide impacts and the oxidation state of carbon in the Earth’s early atmosphere. Origins of Life 20, 199231.Google Scholar
Kasting, J. F. (1991). CO2 condensation and the climate of early Mars. Icarus 94, 113.Google Scholar
Kasting, J. F. (1992). Models relating to Proterozoic atmospheric and oceanic chemistry. In: The Proterozoic Biosphere: A Multidisciplinary Study, ed. Schopf, J. W. and Klein, C., Cambridge: Cambridge University Press, pp. 11851187.Google Scholar
Kasting, J. F. (1993). Earth’s early atmosphere. Science 259, 920926.Google Scholar
Kasting, J. F. (1997). Habitable zones around low mass stars and the search for extraterrestrial life. Origins of Life 27, 291307.Google Scholar
Kasting, J. F. (2001). The rise of atmospheric oxygen (Perspective). Science 293, 819820.Google Scholar
Kasting, J. F. (2010). How to Find a Habitable Planet. Princeton, N.J.: Princeton University Press.Google Scholar
Kasting, J. F. (2013). What caused the rise of atmospheric O2? Chem. Geol. 362, 1325.Google Scholar
Kasting, J. F. and Ackerman, T. P. (1986). Climatic consequences of very high CO2 levels in the Earth’s early atmosphere. Science 234, 13831385.Google Scholar
Kasting, J. F. and Brown, L. L. (1998). Setting the stage: the early atmosphere as a source of biogenic compounds. In: The Molecular Origins of Life: Assembling the Pieces of the Puzzle, ed. Brack, A., New York: Cambridge University. Press, pp. 3556.Google Scholar
Kasting, J. F. and Canfield, D. E. (2012). The global oxygen cycle. In: Fundamentals of Geobiology, ed. Konhauser, K. O., et al., Oxford: Wiley-Blackwell, pp. 93104.Google Scholar
Kasting, J. F. and Catling, D. (2003). Evolution of a habitable planet. Ann. Rev. Astron. Astrophys. 41, 429463.Google Scholar
Kasting, J. F., et al. (2012). Atmospheric oxygenation and volcanism. Nature 487, E1.Google Scholar
Kasting, J. F., et al. (2015). Stratospheric temperatures and water loss from moist greenhouse atmospheres of Earth-like planets. Astrophys. J. Lett. 813, L3, doi: 10.1088/2041-8205/813/1/13.Google Scholar
Kasting, J. F. and Donahue, T. M. (1980). The evolution of atmospheric ozone. J. Geophys. Res. 85, 32553263.Google Scholar
Kasting, J. F., et al. (1993a). Mantle redox evolution and the oxidation state of the Archean atmosphere. J. Geol. 101, 245257.Google Scholar
Kasting, J. F., et al. (1985). Oxidant abundances in rainwater and the evolution of atmospheric oxygen. J. Geophys. Res. 90, 10 49710 510.Google Scholar
Kasting, J. F., et al. (2006). Paleoclimates, ocean depth, and the oxygen isotopic composition of seawater. Earth Planet. Sci. Lett. 252, 8293.Google Scholar
Kasting, J. F., et al. (2014). Remote life-detection criteria, habitable zone boundaries, and the frequency of Earth-like planets around M and late K stars. P. Natl. Acad. Sci. USA 111, 12 64112 646.Google Scholar
Kasting, J. F., et al. (2001). A coupled ecosystem-climate model for predicting the methane concentration in the Archean atmosphere. Origins Life Evol. Biosph. 31, 271285.Google Scholar
Kasting, J. F. and Pollack, J. B. (1983). Loss of water from Venus. I. Hydrodynamic escape of hydrogen. Icarus 53, 479508.Google Scholar
Kasting, J. F., et al. (1984a). Response of Earth’s atmosphere to increases in solar flux and implications for loss of water from Venus. Icarus 57, 335355.Google Scholar
Kasting, J. F., et al. (1984b). Effects of high CO2 levels on surface temperature and atmospheric oxidation state of the early earth. J. Atmos. Chem. 1, 403428.Google Scholar
Kasting, J. F., et al. (1989a). Climate evolution on the terrestrial planets. In: Origin and Evolution of Planetary and Satellite Atmospheres, ed. Matthews, M. S., Tucson, Arizona: University. of Arizona Press, pp. 423449.Google Scholar
Kasting, J. F., et al. (1988). How climate evolved on the terrestrial planets. Scientific Am. 256, 9097.Google Scholar
Kasting, J. F. and Walker, J. C. G. (1981). Limits on oxygen concentration in the prebiological atmosphere and the rate of abiotic fixation of nitrogen. J. Geophys. Res. 86, 11471158.Google Scholar
Kasting, J. F., et al. (1993b). Habitable zones around main sequence stars. Icarus 101, 108128.Google Scholar
Kasting, J. F., et al. (1989b). Sulfur, ultraviolet radiation, and the early evolution of life. Origins Life Evol. Biosph. 19, 95108.Google Scholar
Kasting, J. F., et al. (1983). Photochemistry of methane in the Earth’s early atmosphere. Precambrian Res. 20, 121148.Google Scholar
Kato, S., et al. (1999). The k-distribution method and correlated-k approximation for a shortwave radiative transfer model. J. Quant. Spectrosc. Radiat. Transf. 62, 109121.Google Scholar
Kato, Y., et al. (2009). Hematite formation by oxygenated groundwater more than 2.76 billion years ago. Earth Planet. Sc. Lett. 278, 4049.Google Scholar
Kaufman, A. J. (1997). Palaeoclimatology – An ice age in the tropics. Nature 386, 227228.Google Scholar
Kaufman, A. J., et al. (2007). Late Archean biospheric oxygenation and atmospheric evolution. Science 317, 19001903.Google Scholar
Kawahara, H. and Fujii, Y. (2010). Global mapping of Earth-Like exoplanets from scattered light curves. Astrophys. J. 720, 13331350.Google Scholar
Keir, R. S. (2010). A note on the fluxes of abiogenic methane and hydrogen from mid-ocean ridges. Geophys. Res. Lett. 37.Google Scholar
Keller, C. B. and Schoene, B. (2012). Statistical geochemistry reveals disruption in secular lithospheric evolution about 2.5 Gyr ago. Nature 485, 490495.Google Scholar
Kelley, D. S., et al. (2001). An off-axis hydrothermal vent field near the Mid-Atlantic Ridge at 30o N. Nature 412, 145149.Google Scholar
Kelley, D. S., et al. (2005). A serpentinite-hosted ecosystem: the Lost City hydrothermal vent field. Science 307, 14281434.Google Scholar
Kelley, K. A. and Cottrell, E. (2009). Water and the oxidation state of subduction zone magmas. Science 325, 605607.Google Scholar
Kemp, A. I. S. and Hawkesworth, C. J. (2014). Growth and differentiation of the continental crust from isotope studies of accessory minerals. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 379421.Google Scholar
Kendall, B., et al. (2009). Re-Os and Mo isotope systematics of black shales from the Middle Proterozoic Velkerri and Wollogorang Formations, McArthur Basin, northern Australia. Geochim. Cosmochim. Ac. 73, 25342558.Google Scholar
Kendall, B., et al. (2010). Pervasive oxygenation along late Archaean ocean margins. Nature Geosci. 3, 647652.Google Scholar
Kennedy, M., et al. (2006). Late Precambrian oxygenation; inception of the clay mineral factory. Science 311, 14461449.Google Scholar
Kennedy, M., et al. (2008). Snowball Earth termination by destabilization of equatorial permafrost methane clathrate. Nature 453, 642645.Google Scholar
Kenyon, S. J. and Bromley, B. C. (2006). Terrestrial planet formation. I. The transition from oligarchic growth to chaotic growth. Astron. J. 131, 18371850.Google Scholar
Kerby, R., et al. (1983). Single-carbon catabolism in acetogens: Analysis of carbon flow in Acetobacterium-Woodii and Butyribacterium-Methylotrophicum by fermentation and C-13 nuclear magnetic-resonance measurement. J. Bacteriol. 155, 12081218.Google Scholar
Kerby, R. and Zeikus, J. G. (1983). Growth of Clostridium-Thermoaceticum on H2/CO2 or CO as energy source. Curr. Microbiol. 8, 2730.Google Scholar
Kerr, G. B., et al. (2015). The Palaeoproterozoic global carbon cycle: insights from the Loch Maree Group, NW Scotland. J. Geol. Soc., doi:10.1144/jgs2014-042.Google Scholar
Kerrich, R., et al. (2006). Secular variations in N-isotopes in terrestrial reservoirs and ore deposits. Geol. Soc. Am. Mem. 198, 81104.Google Scholar
Kerrick, D. (2001). Present and past nonanthropogenic CO2 degassing from the solid Earth. Rev. Geophys. 39, 565585.Google Scholar
Kerrick, D. M., et al. (1995). Convective hydrothermal CO2 emission from high heat-flow regions. Chem. Geol. 121, 285293.Google Scholar
Kerridge, J. F. (1985). Carbon, hydrogen and nitrogen in carbonaceous chondrites – Abundances and isotopic compositions in bulk samples. Geochim. Cosmochim. Acta 49, 17071714.Google Scholar
Khare, B. N. and Sagan, C. (1975). Cyclic octatomic sulfur: Possible infrared and visible chromophore in clouds of Jupiter. Science 189, 722723.Google Scholar
Khare, B. N., et al. (1984). Optical constants of organic tholins produced in a simulated Titanian atmosphere: From soft X-ray to microwave frequencies. Icarus 60, 127137.Google Scholar
Kharecha, P., et al. (2005). A coupled atmosphere-ecosystem model of the early Archean Earth. Geobiology 3, 5376.Google Scholar
Khurana, K. K., et al. (2009). Electromagnetic induction from Europa’s ocean and deep interior. In: Europa, ed. Pappalardo, R. T., et al., Tucson: University of Arizona Press.Google Scholar
Kiang, N. Y., et al. (2007a). Spectral signatures of photosynthesis. II. Coevolution with other stars and the atmosphere on extrasolar worlds. Astrobiology 7, 252274.Google Scholar
Kiang, N. Y., et al. (2007b). Spectral signatures of photosynthesis. I. Review of Earth organisms. Astrobiology 7, 222251.Google Scholar
Kieffer, H. H., et al. (1992). The planet Mars from antiquity to present. In: Mars, ed. Kieffer, H. H., et al., Tucson: University of Arizona Press, pp. 133.Google Scholar
Kiehl, J. T. and Dickinson, R. E. (1987). A study of the radiative effects of enhanced atmospheric CO2 and CH4 on early earth surface temperatures. J. Geophys. Res. 92, 29912998.Google Scholar
Kiehl, J. T. and Trenberth, K. E. (1997). Earth’s annual global mean energy budget. Bull. Am. Meteorol. Soc. 78, 197208.Google Scholar
Kienert, H., et al. (2012). Faint young Sun problem more severe due to ice-albedo feedback and higher rotation rate of the early Earth. Geophys. Res. Lett. 39, L23710.Google Scholar
Kilner, B., et al. (2005). Low-latitude glaciation in the Neoproterozoic of Oman. Geology 33, 413416.Google Scholar
Kim, K. M., et al. (2012). Protein domain structure uncovers the origin of aerobic metabolism and the rise of planetary oxygen. Structure 20, 6776.Google Scholar
King, P. L. and McSween, H. Y. (2005). Effects of H2O, pH, and oxidation state on the stability of Fe minerals on Mars. J. Geophys. Res. 110, 115.Google Scholar
Kirchner, J. W. (1989). The Gaia Hypothesis – can it be tested? Reviews of Geophysics 27, 223235.Google Scholar
Kirchner, J. W. (2002). The Gaia hypothesis: Fact, theory, and wishful thinking. Climatic Change 52, 391408.Google Scholar
Kirchner, J. W. (2003). The Gaia hypothesis: Conjectures and refutations. Climatic Change 58, 2145.Google Scholar
Kirk, R. L., et al. (1995). Triton’s plume: Discovery, characteristics and models. In: Neptune and Triton, ed. Cruikshank, D. P., Tucson: University of Arizona Press, pp. 807877.Google Scholar
Kirkpatrick, J. D., et al. (1999). Dwarfs cooler than “M”: The definition of spectral type “L” using discoveries from the 2-Micron All-Sky Survey (2MASS). Astrophysical Journal 519, 802833.Google Scholar
Kirschke, S., et al. (2013). Three decades of global methane sources and sinks. Nat. Geosci. 6, 813823.Google Scholar
Kirschvink, J. L. (1992). In: Late Proterozoic low-latitude global glaciation: the snowball Earth, ed. Schopf, J. W. and Klein, C., Cambridge: Cambridge University Press, pp. 5152.Google Scholar
Kirschvink, J. L., et al. (2000). Paleoproterozoic snowball Earth: Extreme climatic and geochemical global change and its biological consequences. P. Natl. Acad. Sci. USA 97, 14001405.Google Scholar
Kirschvink, J. L. and Kopp, R. E. (2008). Palaeoproterozoic ice houses and the evolution of oxygen-mediating enzymes: the case for a late origin of photosystem II. Phil. Trans. R. Soc Lond.. 363, 27552765.Google Scholar
Kite, E. S., et al. (2013). Seasonal melting and the formation of sedimentary rocks on Mars, with predictions for the Gale Crater mound. Icarus 223, 181210.Google Scholar
Kite, E. S., et al. (2013). Seasonal melting and the formation of sedimentary rocks on Mars, with predictions for the Gale Crater mound. Icarus 223, 181210.Google Scholar
Kite, E. S., et al. (2014). Low palaeopressure of the martian atmosphere estimated from the size distribution of ancient craters. Nature Geoscience 7, 335339.Google Scholar
Kivelson, M. G., et al. (2002). The permanent and inductive magnetic moments of Ganymede. Icarus 157, 507522.Google Scholar
Kivelson, M. G. and Russell, C. T. (1995). Introduction to Space Physics. New York: Cambridge University Press.Google Scholar
Klein, C. (2005). Some Precambrian banded iron-formations (BIFs) from around the world: Their age, geologic setting, mineralogy, metamorphism, geochemistry, and origin. Am. Mineral. 90, 14731499.Google Scholar
Klein, C. and Beukes, N. J. (1992). Time distribution, stratigraphy, and sedimentologic setting, and geochemistry of Precambrian iron formations. In: The Proterozoic Biosphere: A Multidisciplinary Study, ed. Schopf, J. W. and Klein, C., Cambridge: Cambridge University Press, pp. 139146.Google Scholar
Klein, C. and Beukes, N. J. (1993). Sedimentology and geochemistry of the glaciogenic Late Proterozoic Rapitan iron-formation in Canada. Econ. Geol. Bull. Soc. 88, 542565.Google Scholar
Klein, C., et al. (1987). Filamentous microfossils in the Early Proterozoic Transvaal Supergroup: Their morphology, significance, and paleoenvironmental setting. Precambrian Res. 36, 8194.Google Scholar
Klein, H. P. (1979). Viking Mission and the Search for Life on Mars. Rev. Geophys. 17, 16551662.Google Scholar
Klein, H. P. (1998). The search for life on Mars: What we learned from Viking. J. Geophys. Res. 103, 28 46328 466.Google Scholar
Kleine, T. and Rudge, J. F. (2011). Chronometry of meteorites and the formation of the Earth and Moon. Elements 7, 4146.Google Scholar
Kleinhans, M. G. (2005). Flow discharge and sediment transport models for estimating a minimum timescale of hydrological activity and channel and delta formation on Mars. J. Geophys. Res. 110.Google Scholar
Kleinhans, M. G. (2010). A tale of two planets: geomorphology applied to Mars’ surface, fluvio-deltaic processes and landforms. Earth Surf. Proc. Land. 35, 102117.Google Scholar
Kleinhans, M. G., et al. (2010). Palaeoflow reconstruction from fan delta morphology on Mars. Earth Planet. Sc. Lett. 294, 378392.Google Scholar
Klingelhofer, G., et al. (2006). Two earth years of Mossbauer studies of the surface of Mars with MIMOS II. Hyperfine Interactions 170, 169177.Google Scholar
Kliore, A., et al. (1965). Occultation experiment: Results of first direct measurement of Mars atmosphere and ionosphere. Science 149, 1243.Google Scholar
Kliore, A. J., et al. (2002). Ionosphere of Callisto from Galileo radio occultation observations. J. Geophys. Res. 107.Google Scholar
Klose, K. B., et al. (1992). Mineral equilibria and the high radar reflectivity of Venus mountaintops. J. Geophys. Res. 97, 16 35316 369.Google Scholar
Knauth, L. P. (2005). Temperature and salinity history of the Precambrian ocean: implications for the course of microbial evolution. Palaeogeog., Palaeoclimat., Palaeoecol. 219, 5369.Google Scholar
Knauth, L. P. and Epstein, S. (1976). Hydrogen and oxygen isotope ratios in nodular and bedded cherts. Geochim. Cosmochim. Acta 40, 10951108.Google Scholar
Knauth, L. P. and Kennedy, M. J. (2009). The late Precambrian greening of the Earth. Nature 460, 728732.Google Scholar
Knauth, P. and Lowe, D. R. (2003). High Archean climatic temperature inferred from oxygen isotope geochemistry of cherts in the 3.5 Ga Swaziland Supergroup, South Africa. GSA Bull. 115, 566580.Google Scholar
Kneissl, T., et al. (2010). Distribution and orientation of northern-hemisphere gullies on Mars from the evaluation of HRSC and MOC-NA data. Earth Planet. Sc. Lett. 294, 357367.Google Scholar
Knoll, A. H. (1979). Archean photoautotrophy – some alternatives and limits. Orig. Life Evol. Biosph. 9, 313327.Google Scholar
Knoll, A. H. (1992). The early evolution of the eukaryotes: a geological perspective. Science 256, 622627.Google Scholar
Knoll, A. H. (2003). Life on a Young Planet: The First Three Billion Years of Evolution on Earth. Princeton, N.J.; Oxford: Princeton University Press.Google Scholar
Knoll, A. H. and Carroll, S. B. (1999). Early animal evolution: Emerging views from comparative biology and geology. Science 284, 21292137.Google Scholar
Knoll, A. H., et al. (2006a). Eukaryotic organisms in Proterozoic oceans. Phil. Tran. R. Soc. Lond. B 361, 10231038.Google Scholar
Knoll, A. H., et al. (2004). A new period for the geologic time scale. Science 305, 621622.Google Scholar
Knoll, A. H., et al. (2006b). The Ediacaran Period: a new addition to the geologic time scale. Lethaia 39, 1330.Google Scholar
Knollenberg, R. G. and Hunten, D. M. (1979). Clouds of Venus: Particle-size distribution measurements. Science 203, 792795.Google Scholar
Knutson, H. A., et al. (2014). A featureless transmission spectrum for the Neptune-mass exoplanet GJ 436b. Nature 505, 6668.Google Scholar
Knutson, H. A., et al. (2007). A map of the day–night contrast of the extrasolar planet HD 189733b. Nature 447, 183186.Google Scholar
Koeberl, C. (2006). The record of impact processes on the early Earth: A review of the first 2.5 billion years. GSA Special Papers 405, 122.Google Scholar
Kohler, I., et al. (2013). Biological carbon precursor to diagenetic siderite with spherical structures in iron formations. Nat. Commun. 4, 1741, doi :10.1038/ncomms2770.Google Scholar
Kohn, J. P., et al. (1976). 3-Phase solid-liquid-vapor equilibria of binary-n-alkane systems (ethane-n-octane, ethane-n-decane, ethane-n-dodecane). J. Chem. Eng. Data 21, 360362.Google Scholar
Kok, J. F. and Renno, N. O. (2009). Electrification of wind-blown sand on Mars and its implications for atmospheric chemistry. Geophys. Res. Lett. 36.Google Scholar
Komabayashi, M. (1967). Discrete equilibrium temperatures of a hypothetical planet with the atmosphere and the hydrosphere of one component-two phase system under constant solar radiation. J. Meteor. Soc. Japan 45, 137139.Google Scholar
Komabayashi, M. (1968). Conditions for the existence of the atmosphere and oceans. Shizen 23, 2431 (in Japanese).Google Scholar
Komatsu, G., et al. (1993). Stratigraphy and erosional landforms of layered deposits in Valles Marineris, Mars. J. Geophys. Res. 98, 11 10511 121.Google Scholar
Komatsu, G., et al. (2004). Interior layered deposits of Valles Marineris, Mars: Analogous subice volcanism related to Baikal rifting, southern Siberia. Planet. Space Sci. 52, 167187.Google Scholar
Komiya, T., et al. (1999). Plate tectonics at 3.8-3.7 Ga: Field evidence from the Isua Accretionary Complex, southern West Greenland. J. Geol. 107, 515554.Google Scholar
Konhauser, K. (2007). Introduction to Geomicrobiology. Malden, MA: Blackwell Publishing.Google Scholar
Konhauser, K. O., et al. (2007a). Decoupling photochemical Fe(II) oxidation from shallow-water BIF deposition. Earth Planet. Sci. Lett. 258, 87100.Google Scholar
Konhauser, K. O., et al. (2002). Could bacteria have formed the Precambrian banded iron formations? Geology 30, 10791082.Google Scholar
Konhauser, K. O., et al. (2007b). Was there really an Archean phosphate crisis? Science 315, 1234.Google Scholar
Konhauser, K. O., et al. (2011). Aerobic bacterial pyrite oxidation and acid rock drainage during the Great Oxidation Event. Nature 478, 369.Google Scholar
Konhauser, K. O., et al. (2009). Oceanic nickel depletion and a methanogen famine before the Great Oxidation Event. Nature 458, 750754.Google Scholar
Konhauser, K. O., et al. (2015). The Archean nickel famine revisited. Astrobiology 15, 804815.Google Scholar
Konn, C., et al. (2015). The production of methane, hydrogen, and organic compounds in ultramafic-hosted hydrothermal vents of the mid-Atlantic Ridge. Astrobiology 15, 381399.Google Scholar
Kopp, G. and Lean, J. L. (2011). A new, lower value of total solar irradiance: Evidence and climate significance. Geophys. Res. Lett. 38, L01706, doi:10.1029/2010GL045777.Google Scholar
Kopp, R. E., et al. (2005). The Paleoproterozoic Snowball Earth: A climate disaster triggered by the evolution of oxygenic photosynthesis. P. Natl. Acad. Sci. USA 102, 11 13111 136.Google Scholar
Kopparapu, R. K., et al. (2013). Habitable zones around main-sequence stars: New estimates. Astrophys. J. 765, doi: 10.1088/0004-637X/765/2/131.Google Scholar
Kopparapu, R. K., et al. (2014). Habitable zones around main-sequence stars: Dependence on planetary mass. Astrophys. J. Lett. 787.Google Scholar
Korenaga, J. (2006). Archean geodynamics and the thermal evolution of Earth. In: Archean Geodynamics and Environments, ed. Benn, K., Washington DC: American Geophys. Union, pp. 732.Google Scholar
Korenaga, J. (2007). Eustasy, supercontinental insulation, and the temporal variability of terrestrial heat flux. Earth Planet. Sci. Lett. 257, 350358.Google Scholar
Korenaga, J. (2008a). Plate tectonics, flood basalts and the evolution of Earth’s oceans. Terra Nova 20, 419439.Google Scholar
Korenaga, J. (2008b). Urey ratio and the structure and evolution of Earth’s mantle. Reviews of Geophysics 46.Google Scholar
Korycansky, D. G. and Zahnle, K. J. (2005). Modeling crater populations on Venus and Titan. Planet. Space Sci. 53, 695710.Google Scholar
Korycansky, D. G. and Zahnle, K. J. (2011). Titan impacts and escape. Icarus 211, 707721.Google Scholar
Koskinen, T. T., et al. (2014). Thermal escape from extrasolar giant planets. Phil. Trans. R. Soc. A 372, 20130089.Google Scholar
Kostiuk, T., et al. (2006). Stratospheric global winds on Titan at the time of Huygens descent. J. Geophys. Res. 111, E07S03.Google Scholar
Kouchinsky, A., et al. (2012). Chronology of early Cambrian biomineralization. Geol. Mag. 149, 221251.Google Scholar
Kounaves, S. P., et al. (2010). Soluble sulfate in the martian soil at the Phoenix landing site. Geophy. Res. Lett. 37, L09201, doi:10.1029/2010GL042613.Google Scholar
Kouvaris, L. C. and Flasar, F. M. (1991). Phase-equilibrium of methane and nitrogen at low-temperatures – application to Titan. Icarus 91, 112124.Google Scholar
Kraal, E. R., et al. (2008). Catalogue of large alluvial fans in martian impact craters. Icarus 194, 101110.Google Scholar
Kral, T. A., et al. (1998). Hydrogen consumption by methanogens on the early Earth. Origins of Life and Evol. of the Biosph. 28, 311319.Google Scholar
Krasnopolsky, V. (2000). On the deuterium abundance on Mars and some related problems. Icarus 148, 597602.Google Scholar
Krasnopolsky, V. A. (1993). Photochemistry of the martian atmosphere (mean conditions). Icarus 101, 313332.Google Scholar
Krasnopolsky, V. A. (1999). Hydrodynamic flow of N2 from Pluto. J. Geophys. Res. 104, 59555962.Google Scholar
Krasnopolsky, V. A. (2002). Mars’ upper atmosphere and ionosphere at low, medium, and high solar activities: Implications for evolution of water. J. Geophys. Res. 107.Google Scholar
Krasnopolsky, V. A. (2006). Photochemistry of the martian atmosphere: Seasonal, latitudinal, and diurnal variations. Icarus 185, 153170.Google Scholar
Krasnopolsky, V. A. (2010). The photochemical model of Titan’s atmosphere and ionosphere: A version without hydrodynamic escape. Planet. Space Sci. 58, 15071515.Google Scholar
Krasnopolsky, V. A. (2011). Atmospheric chemistry on Venus, Earth, and Mars: Main features and comparison. Planet. Space Sci. 59, 952964.Google Scholar
Krasnopolsky, V. A. (2012). Search for methane and upper limits to ethane and SO2 on Mars. Icarus 217, 144152.Google Scholar
Krasnopolsky, V. A. and Feldman, P. D. (2001). Detection of molecular hydrogen in the atmosphere of Mars. Science 294, 19141917.Google Scholar
Krasnopolsky, V. A. and Lefevre, F. (2013). Chemistry of the atmospheres of Mars, Venus, and Titan. In: Comparative Climatology of Terrestrial Planets, ed. Mackwell, S. J., et al., Tucson: University of Arizona Press, pp. 231275.Google Scholar
Krasnopolsky, V. A. and Pollack, J. B. (1994). H2O–H2SO4 system in Venus’ clouds and OCS, CO, and H2SO4 profiles in Venus’ troposphere. Icarus 109, 5878.Google Scholar
Kraus, R. G., et al. (2011). Impacts onto H2O ice: Scaling laws for melting, vaporization, excavation, and final crater size. Icarus 214, 724738.Google Scholar
Kreidberg, L., et al. (2014). Clouds in the atmosphere of the super-Earth exoplanet GJ 1214b. Nature 505, 6972.Google Scholar
Kreslavsky, M. A. and Head, J. W. (2002). Fate of outflow channel effluents in the northern lowlands of Mars: The Vastitas Borealis Formation as a sublimation residue from frozen ponded bodies of water. J. Geophys. Res. 107, 5121.Google Scholar
Kreslavsky, M. A., et al. (2015). The resurfacing history of Venus: Constraints from buffered crater densities. Icarus 250, 438450.Google Scholar
Kress, M. E. and McKay, C. P. (2004). Formation of methane in comet impacts: implications for Earth, Mars, and Titan. Icarus 168, 475483.Google Scholar
Krimigis, S. M., et al. (2013). Search for the exit: Voyager 1 at heliosphere’s border with the galaxy. Science 341, 144147.Google Scholar
Krissansen-Totton, J., et al. (2015). A statistical analysis of the carbon isotope record from the Archean to Phanerozoic and implications for the rise of oxygen. Amer. J. Sci. 315, 275316.Google Scholar
Krissansen-Totton, J., et al. (2016a). On detecting biospheres from chemical disequilibrium in planetary atmospheres. Astrobiology, 16, 3967.Google Scholar
Krissansen-Totton, J., et al. (2016b). Is the Pale Blue Dot unique? Optimized photometric bands for identifying Earth-like exoplanets. Astrophys. J. 871, 31, doi: 10.3847/0004-637X/817/1/31.Google Scholar
Kruijer, T. S., et al. (2014). Protracted core formation and rapid accretion of protoplanets. Science 344, 11501154.Google Scholar
Krull-Davatzes, A. E., et al. (2010). Evidence for a low-O2 Archean atmosphere from nickel-rich chrome spinels in 3.24 Ga impact spherules, Barberton greenstone belt, South Africa. Earth Planet. Sci. Lett. 296, 319328.Google Scholar
Krupp, R., et al. (1994). The Early Precambrian atmosphere and hydrosphere: Thermodynamic constraints from mineral deposits. Econ. Geol. 89, 15811598.Google Scholar
Kuhn, W. R. and Atreya, S. K. (1979). Ammonia photolysis and the greenhouse effect in the primordial atmosphere of the Earth. Icarus 37, 207213.Google Scholar
Kuiper, G. P. (1944). Titan: A satellite with an atmosphere. Astrophys. J. 100, 378383.Google Scholar
Kuipers, G., et al. (2013). Periglacial evidence for a 1.91–1.89 Ga old glacial period at low latitude, Central Sweden. Geol. Today 29, 218221.Google Scholar
Kumar, S., et al. (1983). Nonthermal escape of hydrogen and deuterium from Venus and implications for loss of water. Icarus 55, 369389.Google Scholar
Kump, L. R. and Barley, M. E. (2007). Increased subaerial volcanism and the rise of atmospheric oxygen 2.5 billion years ago. Nature 448, 10331036.Google Scholar
Kump, L. R., et al. (2011). Isotopic evidence for massive oxidation of organic matter following the Great Oxidation Event. Science 334, 16941696.Google Scholar
Kump, L. R., et al. (2001). The rise of atmospheric oxygen and the “upside-down” Archean mantle. Geol. Geochem. Geophys. (online) 2.Google Scholar
Kump, L. R., et al. (2010). The Earth System. Upper Saddle River, NJ: Pearson.Google Scholar
Kump, L. R. and Seyfried, W. E. (2005). Hydrothermal Fe fluxes during the Precambrian: Effect of low oceanic sulfate concentrations and low hydrostatic pressure on the composition of black smokers. Earth Planet. Sc. Lett. 235, 654662.Google Scholar
Kunzmann, M., et al. (2013). Zn isotope evidence for immediate resumption of primary productivity after snowball Earth. Geology 41, 2730.Google Scholar
Kuramoto, K. and Matsui, T. (1994). Formation of a hot proto-atmosphere on the accreting giant icy satellite: Implications for the origin and evolution of Titan, Ganymede, and Callisto. J. Geophys. Res. 99, 21 18321 200.Google Scholar
Kuramoto, K. and Matsui, T. (1996). Partitioning of H and C between the mantle and the core during the core formation in the Earth: its implications for the atmospheric evolution and redox state of the early mantle. J. Geophys. Res. 101, 14 90914 932.Google Scholar
Kuramoto, K., et al. (2013). Effective hydrodynamic hydrogen escape from an early Earth atmosphere inferred from high-accuracy numerical simulation. Earth Planet. Sci. Lett. 375, 312318.Google Scholar
Kurata, F. (1975). Solubility of heavier hydrocarbons in liquid methane. Research Report, RR-14. Gas Processors Association.Google Scholar
Kurokawa, H. and Kaltenegger, L. (2013). Atmospheric mass-loss and evolution of short-period exoplanets: the examples of CoRoT-7b and Kepler-10b. Mon. Not. R. Astron. Soc. 433, 32393245.Google Scholar
Kurokawa, H. and Nakamoto, T. (2014). Mass-loss evolution of close-in exoplanets: evaporation of hot Jupiters and the effect on population. Astrophysical Journal 783.Google Scholar
Kurokawa, H., et al. (2014). Evolution of water reservoirs on Mars: Constraints from hydrogen isotopes in martian meteorites. Earth Planet. Sc. Lett. 394, 179185.Google Scholar
Kurosawa, K. (2015). Impact-driven planetary desiccation: The origin of the dry Venus. Earth Planet. Sci. Lett. 429, 181190.Google Scholar
Kurster, M., et al. (2003). The low-level radial velocity variability in Barnard’s star (= GJ 699): Secular acceleration, indications for convective redshift, and planet mass limits. Astron. Astrophys. 403, 10771087.Google Scholar
Kurzweil, F., et al. (2013). Atmospheric sulfur rearrangement 2.7 billion years ago: Evidence for oxygenic photosynthesis. Earth Planet. Sci. Lett. 366, 1726.Google Scholar
Kurzweil, F., et al. (2015a). Coupled sulfur, iron and molybdenum isotope data from black shales of the Teplá-Barrandian unit argue against deep ocean oxygenation during the Ediacaran. Geochim. Cosmochim. Acta 171, 121142.Google Scholar
Kurzweil, F., et al. (2015b). Continuously increasing δ98Mo values in Neoarchean black shales and iron formations from the Hamersley Basin. Geochim. Cosmochim. Acta 164, 523542.Google Scholar
Kuzuhara, M., et al. (2013). Direct imaging of a cold Jovian exoplanet in orbit around the Sun-like star GJ 504. The Astrophysical Journal 774, 11.Google Scholar
Laakso, T. A. and Schrag, D. P. (2014). Regulation of atmospheric oxygen during the Proterozoic. Earth Planet. Sci. Lett. 388, 8191.Google Scholar
Lacis, A. A., et al. (2013). The role of long-lived greenhouse gases as principal LW control knob that governs the global surface temperature for past and future climate change. Tellus B 65, 19734, http://dx.doi.org/10.3402/tellusb.v65i0.19734.Google Scholar
Lacis, A. A., et al. (2010). Atmospheric CO2: Principal control knob governing Earth’s temperature. Science 330, 356359.Google Scholar
Lagrange, A. M., et al. (2010). A giant planet imaged in the disk of the young star β pictoris. Science 329, 5759.Google Scholar
Lainey, V., et al. (2009). Strong tidal dissipation in Io and Jupiter from astrometric observations. Nature 459, 957959.Google Scholar
Lake, J. A., et al. (1984). Eocytes: A new ribosome structure indicates a kingdom with a close relationship to eukaryotes. P. Natl. Acad. Sci. USA 81, 37863790.Google Scholar
Lalonde, S. V. and Konhauser, K. O. (2015). Benthic perspective on Earth’s oldest evidence for oxygenic photosynthesis. P. Natl. Acad. Sci. USA 112, 9951000.Google Scholar
Lamb, D. M., et al. (2009). Evidence for eukaryotic diversification in the similar to 1800 million-year-old Changzhougou Formation, North China. Precam. Res. 173, 93104.Google Scholar
Lamb, M. P., et al. (2008). Formation of Box Canyon, Idaho, by megaflood: Implications for seepage erosion on Earth and Mars. Science 320, 10671070.Google Scholar
Lamb, M. P., et al. (2006). Can springs cut canyons into rock? J. Geophys. Res. 111.Google Scholar
Lambrechts, M. and Johansen, A. (2012). Rapid growth of gas-giant cores by pebble accretion. Astron. Astrophys. 544.Google Scholar
Lammer, H. (2013). Origin and Evolution of Planetary Atmospheres: Implications for Habitability. New York: Springer.Google Scholar
Lammer, H. and Bauer, S. J. (1991). Nonthermal atmospheric escape from Mars and Titan. J. Geophys. Res. 96, 18191825.Google Scholar
Lammer, H., et al. (2013). Outgassing history and escape of the Martian atmosphere and water inventory. Space Sci. Rev. 174, 113154.Google Scholar
Lammer, H., et al. (2008). Atmospheric escape and evolution of terrestrial planets and satellites. Space Sci. Rev. 139, 399436.Google Scholar
Lammer, H., et al. (2006). Loss of hydrogen and oxygen from the upper atmosphere of Venus. Planet. Space Sci. 54, 14451456.Google Scholar
Lammer, H., et al. (2003a). Loss of water from Mars: Implications for the oxidation of the soil. Icarus 165, 925.Google Scholar
Lammer, H., et al. (2007). Coronal Mass Ejection (CME) activity of low mass M stars as an important factor for the habitability of terrestrial exoplanets. II. CME-induced ion pick up of Earth-like exoplanets in close-in habitable zones. Astrobiology 7, 185207.Google Scholar
Lammer, H., et al. (2009). Determining the mass loss limit for close-in exoplanets: what can we learn from transit observations? Astron. Astrophys. 506, 399410.Google Scholar
Lammer, H., et al. (2003b). Atmospheric loss of exoplanets resulting from stellar X-ray and extreme-ultraviolet heating. Astrophys. J. 598, L121L124.Google Scholar
Land, L. S. (1995). Oxygen and carbon isotopic composition of Ordovician brachiopods: Implications for Coeval seawater – Comment. Geochim. Cosmochim. Acta 59, 28432844.Google Scholar
Lange, M. A. and Ahrens, T. J. (1982). The evolution of an impact generated atmosphere. Icarus 51, 96120.Google Scholar
Lange, M. A. and Ahrens, T. J. (1986). Shock-induced CO2 loss from CaCO3 – Implications for early planetary atmospheres. Earth Planet. Sci. Lett. 77, 409418.Google Scholar
Langevin, Y., et al. (2005). Sulfates in the north polar region of Mars detected by OMEGA/Mars express. Science 307, 15841586.Google Scholar
Lapen, T. J., et al. (2010). A younger age for ALH84001 and its geochemical link to shergottite sources in Mars. Science 328, 347351.Google Scholar
Lara, L. M., et al. (2014). A time-dependent photochemical model for Titan’s atmosphere and the origin of H2O. Astron. Astrophys. 566.Google Scholar
Lasaga, A. C., et al. (1971). Primordial oil slick. Science 174, 5355.Google Scholar
Lasaga, A. C., et al. (1985). An improved geochemical model of atmospheric CO2 fluctuations over the past 100 million years. In: The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present, ed. Sundquist, E. T. and Broecker, W. S., Washington, DC: American Geophysical Union, pp. 397411.Google Scholar
Laskar, J. (2000). On the spacing of planetary systems. Phys. Rev. Lett. 84, 32403243.Google Scholar
Laskar, J. and Correia, A. C. M. (2004). The rotation of extra-solar planets. In: Extrasolar Planets: Today and Tomorrow, ed. Beaulieu, J.-P., et al.: Astronom. Soc. of the Pacific, pp. 401409.Google Scholar
Laskar, J., et al. (2004). Long term evolution and chaotic diffusion of the insolation quantities of Mars. Icarus 170, 343364.Google Scholar
Lasue, J., et al. (2013). Quantitative assessments of the martian hydrosphere. Space Sci. Rev. 174, 155212.Google Scholar
Lavvas, P. P., et al. (2008a). Coupling photochemistry with haze formation in Titan’s atmosphere, part I: Model description. Planet. Space Sci. 56, 2766.Google Scholar
Lavvas, P. P., et al. (2008b). Coupling photochemistry with haze formation in Titan’s atmosphere. Part II: Results and validation with Cassini/Huygens data. Planet. Space Sci. 56, 6799.Google Scholar
Le Deit, L., et al. (2010). Morphology, stratigraphy, and mineralogical composition of a layered formation covering the plateaus around Valles Marineris, Mars: Implications for its geological history. Icarus 208, 684703.Google Scholar
Le Heron, P., et al. (2010). Sea ice−free conditions during the Sturtian glaciation (early Cryogenian), South Australia. Geology 39, 134.Google Scholar
Le Hir, G., et al. (2010). Toward the snowball Earth deglaciation. Climate Dynamics 35, 285297.Google Scholar
Lean, J. and Rind, D. (1998). Climate forcing by changing solar radiation. Journal of Climate 11, 30693094.Google Scholar
Leather, J., et al. (2002). Neoproterozoic snowball Earth under scrutiny: Evidence from the Fiq glaciation of Oman. Geology 30, 891894.Google Scholar
LeBlanc, F. (2010). An Introduction to Stellar Astrophysics. Chichester, UK: Wiley.Google Scholar
Lebonnois, S., et al. (2003). Atomic and molecular hydrogen budget in Titan’s atmosphere. Icarus 161, 474485.Google Scholar
Lebonnois, S., et al. (2012). Angular momentum budget in General Circulation Models of superrotating atmospheres: A critical diagnostic. J. Geophys. Res. 117.Google Scholar
Lebonnois, S., et al. (2010). Superrotation of Venus’ atmosphere analyzed with a full general circulation model. J. Geophys. Res. 115.Google Scholar
Lecavalier des Etangs, A., et al. (2004). Atmospheric escape from hot Jupiters. Astron. Astrophys. 418, L1L4.Google Scholar
Lecavalier Etangs, L. D., et al. (2008). Rayleigh scattering in the transit spectrum of HD189733b. Astron. Astrophys. 481, L83L86.Google Scholar
Leconte, J., et al. (2013). Increased insolation threshold for runaway greenhouse processes on Earth-like planets. Nature 504, 268271.Google Scholar
Leconte, J., et al. (2015). Asynchronous rotation of Earth-mass planets in the habitable zone of lower-mass stars. Science 347, 632635.Google Scholar
Lecuyer, C., et al. (1998). The hydrogen isotope composition of seawater and the global water cycle. Chem. Geol. 145, 249261.Google Scholar
Lecuyer, C. and Ricard, Y. (1999). Long-term fluxes and budget of ferric iron: implication for the redox states of the Earth’s mantle and atmosphere. Earth Planet. Sci. Lett. 165, 197211.Google Scholar
Lecuyer, C., et al. (2000). Comparison of carbon, nitrogen and water budgets on Venus and the Earth. Earth Planet. Sc. Lett. 181, 3340.Google Scholar
Lederberg, J. (1965). Signs of life: criterion-system of exobiology. Nature 207, 913.Google Scholar
Lee, C. T. A., et al. (2005). Similar V/Sc systematics in MORB and arc basalts: Implications for the oxygen fugacities of their mantle source regions. J. Petrol. 46, 23132336.Google Scholar
Lee, C. T. A., et al. (2010). The redox state of arc mantle using Zn/Fe systematics. Nature 468, 681685.Google Scholar
Lee, H., et al. (2016). Massive and prolonged deep carbon emissions associated with continental rifting. Nat. Geosci. 9, 145149.Google Scholar
Lee, J. Y., et al. (2006). A redetermination of the isotopic abundances of atmospheric Ar. Geochim. Cosmochim. Acta 70, 45074512.Google Scholar
Lee, S. and Kim, H. K. (2003). The dynamical relationship between subtropical and eddy-driven jets. J. Atmos. Sci. 60, 14901503.Google Scholar
Lee, Y. N. and Schwartz, S. E. (1981). Evaluation of the rate of uptake of nitrogen dioxide by atmospheric and surface liquid water. J. Geophys. Res. 86, 11 97111 983.Google Scholar
Lefevre, F., et al. (2008). Heterogeneous chemistry in the atmosphere of Mars. Nature 454, 971975.Google Scholar
Lefevre, F. and Forget, F. (2009). Observed variations of methane on Mars unexplained by known atmospheric chemistry and physics. Nature 460, 720723.Google Scholar
Leger, A., et al. (1993). Search for primitive life on a distant planet: Relevance of O2 and O3 dectections. Astron. Astrophys. 277, 309313.Google Scholar
Leighton, R. B. and Murray, B. C. (1966). Behaviour of carbon dioxide and other volatiles on Mars. Science 153, 136144.Google Scholar
Leitzinger, M., et al. (2011). Could CoRoT-7b and Kepler-10b be remnants of evaporated gas or ice giants? Planet. Space Sci. 59, 14721481.Google Scholar
Lellouch, E., et al. (1997). Monitoring of mesospheric structure and dynamics. In: Venus II: Geology, Geophysics, Atmosphere, and Solar Wind Environment, ed. Bougher, S. W., et al., Tucson, AZ: University of Arizona Press, pp. 295324.Google Scholar
Lellouch, E., et al. (2011a). High resolution spectroscopy of Pluto’s atmosphere: detection of the 2.3 μm CH4 bands and evidence for carbon monoxide. Astron. Astrophys. 530.Google Scholar
Lellouch, E., et al. (2011b). The tenuous atmospheres of Pluto and Triton explored by CRIRES on the VLT. ESO Messenger 145, 2023.Google Scholar
Lellouch, E., et al. (2003). Volcanically emitted sodium chloride as a source for Io’s neutral clouds and plasma torus. Nature 421, 4547.Google Scholar
Lemaire, J. F., et al. (2007). History of kinetic polar wind models and early observations. J. Atmos. Sol-Terr. Phys. 69, 19011935.Google Scholar
Lenardic, A., et al. (2004). Growth of the hemispheric dichotomy and the cessation of plate tectonics on Mars. J. Geophys. Res. 109, E02003, doi:10.1029/2003JE002172.Google Scholar
Lenton, T. M. (1998). Gaia and natural selection. Nature 394, 439447.Google Scholar
Lenton, T. M., et al. (2014). Co-evolution of eukaryotes and ocean oxygenation in the Neoproterozoic era. Nature Geoscience 7, 257265.Google Scholar
Lenton, T. M. and Watson, A. J. (2000). Redfield revisited: II. What regulates the oxygen content of the atmosphere? Global Biogeochem. Cyc. 14, 249268.Google Scholar
Leone, G. (2014). A network of lava tubes as the origin of Labyrinthus Noctis and Valles Marineris on Mars. J. Volcanol. Geoth. Res. 277, 18.Google Scholar
Leone, G., et al. (2014). Three-dimensional simulations of the southern polar giant impact hypothesis for the origin of the Martian dichotomy. Geophys. Res. Lett. 41, 2014GL062261.Google Scholar
Leovy, C. (1982a). Martian meteorological variability. Adv. Space Res. 2, 1944.Google Scholar
Leovy, C. (2001). Weather and climate on Mars. Nature 412, 245249.Google Scholar
Leovy, C. B. (1964). Simple models of thermally driven mesospheric circulation. J. Atmos. Sci. 21, 327341.Google Scholar
Leovy, C. B. (1973). Rotation of upper-atmosphere of Venus. J. Atmos. Sci. 30, 12181220.Google Scholar
Leovy, C. B. (1977). The atmosphere of Mars. Sci. Am. 237, 3443.Google Scholar
Leovy, C. B. (1982b). Control of the homopause level. Icarus 50, 311321.Google Scholar
Leovy, C. B. (1987). Zonal winds near Venus cloud top level: An analytic model of the equatorial wind speed. Icarus 69, 193201.Google Scholar
Leovy, C. B. and Mintz, Y. (1969). Numerical simulation of the weather and climate of Mars. Journal of Atmospheric Sciences 26, 1167–90.Google Scholar
Leovy, C. B., et al. (1973). Mechanisms for Mars dust storms. J. Atmos. Sci. 30, 749762.Google Scholar
Lepland, A., et al. (2013). The earliest phosphorites: Radical change in the phosphorus cycle ruring the Palaeoproterozoic. In: Reading the Archive of Earth’s Oxygenation, ed. Melezhik, V. A. e. a., Berlin: Springer, pp. 12751296.Google Scholar
Lepland, A., et al. (2005). Questioning the evidence for Earth’s earliest life - Akilia revisited. Geology 33, 7779.Google Scholar
Lepland, A., et al. (2011). Fluid-deposited graphite and its geobiological implications in early Archean gneiss from Akilia, Greenland. Geobiology 9, 29.Google Scholar
Lesniak, M. V. and Desch, S. J. (2011). Temperature structure of protoplanetary disks undergoing layered accretion. Astrophys. J. 740, 118.Google Scholar
Levenson, B. P. (2015). Why Hart found narrow ccospheres: A minor science mystery solved. Astrobiology 15, 327330.Google Scholar
LeVeque, R. J. (2002). Finite Volume Methods for Hyperbolic Problems. Cambridge: Cambridge University Press.Google Scholar
Leverington, D. W. (2004). Volcanic rilles, streamlined islands, and the origin of outflow channels on Mars. J. Geophys. Res. 109, 114.Google Scholar
Leverington, D. W. (2007). Was the Mangala Valles system incised by volcanic flows? J. Geophys. Res. 112, 122.Google Scholar
Leverington, D. W. (2011). A volcanic origin for the outflow channels of Mars: Key evidence and major implications. Geomorphology 132, 5175.Google Scholar
Levin, L. A. (2002). Deep-ocean life where oxygen is scarce. Am. Sci. 90, 436444.Google Scholar
Levin, L. A. (2003). Oxygen minimum zone benthos: Adaptation and community response to hypoxia. Oceanogr. Mar. Biol. 41, 145.Google Scholar
Levine, J. S., et al. (1979). The evolution and variability of atmospheric ozone over geologic time. Icarus 39, 295309.Google Scholar
Levine, X. J. and Schneider, T. (2011). Response of the Hadley Circulation to climate change in an aquaplanet GCM coupled to a simple representation of ocean heat transport. J. Atmos. Sci. 68, 769783.Google Scholar
Levison, H. F. and Dones, L. (2014). Comet populations and cometary dynamics. In: Encyclopedia of the Solar System, ed. Spohn, T., et al., 3rd edn. Boston: Elsevier, pp. 705719.Google Scholar
Levison, H. F., et al. (2001). Could the lunar “Late heavy bombardment” have been triggered by the formation of Uranus and Neptune? Icarus 151, 286306.Google Scholar
Levison, H. F., et al. (2015). Growing the gas-giant planets by the gradual accumulation of pebbles. Nature 524, 322324.Google Scholar
Levy, H. (1971). Normal atmosphere – large radical and formaldehyde concentrations predicted. Science 173, 141143.Google Scholar
Levy, J. S., et al. (2014). Sequestered glacial ice contribution to the global Martian water budget: Geometric constraints on the volume of remnant, midlatitude debris-covered glaciers. J. Geophys. Res. 119, 21882196.Google Scholar
Lew, S. K. (1967). The problem of hydrogen escape in the Earth’s upper atmosphere – a reappraisal. University of California, Los Angeles, PhD, Los Angeles, CA.Google Scholar
Lewis, G. N. and Randall, M. (1923). Thermodynamics and the Free Energy of Chemical Substances. New York: McGraw-Hill.Google Scholar
Lewis, J. S. (2004). Physics and Chemistry of the Solar System (2nd Edn). Boston: Elsevier.Google Scholar
Lewis, J. S. and Prinn, R. G. (1984). Planets and Their Atmospheres: Origin and Evolution. Orlando, Florida: Academic Press.Google Scholar
Lewis, S. R., et al. (2007). Assimilation of thermal emission spectrometer atmospheric data during the Mars Global Surveyor aerobraking period. Icarus 192, 327347.Google Scholar
Li, C., et al. (2010). A stratified redox model for the Ediacaran ocean. Science 328, 8083.Google Scholar
Li, D. W. and Pierrehumbert, R. T. (2011). Sea glacier flow and dust transport on Snowball Earth. Geophys. Res. Lett. 38.Google Scholar
Li, L. M., et al. (2011). The global energy balance of Titan. Geophy. Res. Lett. 38.Google Scholar
Li, W. Q., et al. (2013). An anoxic, Fe(II)-rich, U-poor ocean 3.46 billion years ago. Geochim. Cosmochim. Acta 120, 6579.Google Scholar
Li, W. Q., et al. (2012). U-Th-Pb isotope data indicate phanerozoic age for oxidation of the 3.4 Ga Apex Basalt. Earth Planet. Sci. Lett. 319, 197206.Google Scholar
Li, Z. X. A. and Lee, C. T. A. (2004). The constancy of upper mantle fO(2) through time inferred from V/Sc ratios in basalts. Earth Planet. Sci. Lett. 228, 483493.Google Scholar
Lian, Y. and Showman, A. P. (2010). Generation of equatorial jets by large-scale latent heating on the giant planets. Icarus 207, 373393.Google Scholar
Liang, M. C., et al. (2006). Production of hydrogen peroxide in the atmosphere of a Snowball Earth and the origin of oxygenic photosynthesis. P. Natl. Acad. Sci. USA 103, 18 89618 899.Google Scholar
Liang, M. C., et al. (2005). Atmosphere of Callisto. Journal of Geophysical Research-Planets 110.Google Scholar
Lichtenberg, K. A., et al. (2010). Stratigraphy of hydrated sulfates in the sedimentary deposits of Aram Chaos, Mars. J. Geophys. Res. 115, E00D17, doi:10.1029/2009JE003353.Google Scholar
Lide, D. R. (2011). Handbook of Chemistry and Physics. Boca Raton, FL: CRC Press.Google Scholar
Limaye, S. S. and Rengel, M. (2013). Atmospheric circulation and dynamics. In: Towards Understanding the Climate of Venus, ed. Bengtsson, L., et al., New York: Springer, pp. 5570.Google Scholar
Lin, B., et al. (2002). The iris hypothesis: A negative or positive cloud feedback? Journal of Climate 15, 37.Google Scholar
Lin, Y., et al. (2011). Multiple-sulfur isotope effects during photolysis of carbonyl sulfide. Atmos. Chem. Phys. 11, 10 28310 292.Google Scholar
Lindemann, T. A. and Dobson, G. M. B. (1923). A theory of meteors, and the density and temperature of the outer atmosphere to which it leads. Proc. R. Soc. Lond. A 102, 411437.Google Scholar
Lindsay, J. F. and Brasier, M. D. (2002). Did global tectonics drive early biosphere evolution? Carbon isotope record from 2.6 To 1.9 Ga carbonates of Western Australian basins. Precam. Res. 114, 134.Google Scholar
Lindzen, R. S. (1971). Atmospheric tides. In: Mathematical Problems in the Geophysical Sciences, No. 2 : Inverse Problems, Dynamo Theory and Tides, ed. Reid, W. H., Providance, RI: Amer. Math. Soc., pp. 293362.Google Scholar
Lindzen, R. S., et al. (2001). Does the earth have an adaptive infrared iris? Bull. Am. Met. Soc. 82, 417432.Google Scholar
Lindzen, R. S. and Hou, A. Y. (1988). Hadley circulations for zonally averaged heating centred off the equator. J. Atmos. Sci. 45, 24162447.Google Scholar
Line, M. R., et al. (2010). High temperature photochemistry in the atmosphere of HD 189733b. Astrophys. J. 717, 496502.Google Scholar
Lineweaver, C. H., et al. (2004). The Galactic habitable zone and the age distribution of complex life in the Milky Way. Science 303, 5962.Google Scholar
Linsky, J. L., et al. (2010). Observations of mass loss from the transiting exoplanet HD 209458b*. Astrophys. J. 717, 12911299.Google Scholar
Liou, K.-N. (2002). An Introduction to Atmospheric Radiation. Amsterdam; London: Academic Press.Google Scholar
Liss, P. S., et al. (1997). Marine sulphur emissions. Phil. Trans. R. Soc. Lond. B 352, 159168.Google Scholar
Liss, P. S. and Slater, P. G. (1974). Flux of gases across air-sea interface. Nature 247, 181184.Google Scholar
Lissauer, J. J. (2007). Planets formed in habitable zones of M dwarf stars probably are deficient in volatiles. Astrophys. J. 660, L149L152.Google Scholar
Little, B., et al. (1999). Galileo images of lightning on Jupiter. Icarus 142, 306323.Google Scholar
Liu, A. G., et al. (2015). Remarkable insights into the paleoecology of the Avalonian Ediacaran macrobiota. Gondwana Res. 27, 13551380.Google Scholar
Liu, A. G., et al. (2010). First evidence for locomotion in the Ediacara biota from the 565 Ma Mistaken Point Formation, Newfoundland. Geology 38, 123126.Google Scholar
Liu, S. C. and Donahue, T. M. (1974). The aeronomy of hydrogen in the atmosphere of the earth. J. Atmos. Sci. 31, 11181136.Google Scholar
Lodders, K. (2003). Solar system abundances and condensation temperatures of the elements. Astrophys. J. 591, 12201247.Google Scholar
Lodders, K. (2010a). Atmospheric chemistry of the gas giant planets. Geochem. News 142, 111.Google Scholar
Lodders, K. (2010b). Solar system abundances of the elements. In: Principles and Perspectives in Cosmochemistry, ed. Goswami, A. and Reddy, B. E., Berlin: Springer, pp. 379417.Google Scholar
Lodders, K. and Fegley, B. (1998). The Planetary Scientist’s Companion. New York: Oxford University Press.Google Scholar
Loeb, N. G., et al. (2009). Toward optimal closure of the Earth’s top-of-atmosphere radiation budget. J. Climate 22, 748766.Google Scholar
Logan, G. A., et al. (1995). Terminal Proterozoic reorganization of biogeochemical cycles. Nature 376, 5356.Google Scholar
Lopez, E. D. and Fortney, J. J. (2013). The role of core mass in controlling evaporation: The Kepler radius distribution and the Kepler-36 density dichotomy. Astrophysical Journal 776.Google Scholar
Lopez, E. D. and Fortney, J. J. (2014). Understanding the mass-radius relation for sub-Neptunes: Radius as a proxy for composition. Astrophys. J. 792, 1, doi: 10.1088/0004-637X/792/1/1.Google Scholar
López-Puertas, M. and Taylor, F. W. (2001). Non-LTE Radiative Transfer in the Atmosphere. London: World Scientific.Google Scholar
Lora, J. M., et al. (2015). GCM simulations of Titan’s middle and lower atmosphere and comparison to observations. Icarus 250, 516528.Google Scholar
Lorenz, R. D. (2000). The weather on Titan. Science 290, 467468.Google Scholar
Lorenz, R. D., et al. (2009). Seasonal change on Titan. In: Titan from Cassini-Huygens, ed. Brown, R. H., et al., New York: Springer.Google Scholar
Lorenz, R. D., et al. (1997). Photochemically driven collapse of Titan’s atmosphere. Science 275, 642644.Google Scholar
Lorenz, R. D., et al. (2008). Titan’s inventory of organic surface materials. Geophys. Res. Lett. 35.Google Scholar
Lorenz, R. D., et al. (2006). Titan’s damp ground: Constraints on Titan surface thermal properties from the temperature evolution of the Huygens GCMS inlet. Meteorit. Planet. Sci. 41, 17051714.Google Scholar
Lorenz, R. D., et al. (2011). Hypsometry of Titan. Icarus 211, 699706.Google Scholar
Lorius, C., et al. (1990). The ice-core record: Climate sensitivity and future greenhouse warming. Nature 347, 139145.Google Scholar
Love, G. D., et al. (2009). Fossil steroids record the appearance of Demospongiae during the Cryogenian period. Nature 457, 718721.Google Scholar
Lovelock, J. (1988). The Ages of Gaia. New York: W.W. Norton.Google Scholar
Lovelock, J. E. (1965). A physical basis for life detection experiments. Nature 207, 568570.Google Scholar
Lovelock, J. E. (1972). Gaia as seen through atmosphere. Atmos. Environ. 6, 579–&.Google Scholar
Lovelock, J. E. (1975). Thermodynamics and the recognition of alien biospheres. Proc. R. Soc. Lond.B 189, 167181.Google Scholar
Lovelock, J. E. (1979). Gaia: A New Look at Life on Earth. Oxford: Oxford University Press.Google Scholar
Lovelock, J. E. (1989). Geophysiology, the science of Gaia. Rev. Geophys. 27, 215222.Google Scholar
Lovelock, J. E. (1991). Gaia: The Practical Science of Planetary Medicine. London: Gaia Books.Google Scholar
Lovelock, J. E. and Watson, A. J. (1982). The regulation of carbon dioxide and climate: Gaia or Geochemistry. Planet. Space Sci. 30, 795802.Google Scholar
Lovelock, J. E. and Whitfield, M. (1982). Life span of the biosphere. Nature 296, 561563.Google Scholar
Lowe, D. R. and Byerly, G. R. (2015). Geologic record of partial ocean evaporation triggered by giant asteroid impacts, 3.29–3.23 billion years ago. Geology 43, 535538.Google Scholar
Lowe, D. R., et al. (2014). Recently discovered 3.42-3.23 Ga impact layers, Barberton Belt, South Africa: 3.8 Ga detrital zircons, Archean impact history, and tectonic implications. Geology 42, 747750.Google Scholar
Lu, J., et al. (2009). Cause of the widening of the tropical belt since 1958. Geophys. Res. Lett. 36, L03803, doi:10.1029/2008GL036076.Google Scholar
Lucchitta, B. K., et al. (1994). Topography of Valles-Marineris - Implications for erosional and structural history. J. Geophys. Res. 99, 37833798.Google Scholar
Luger, R. and Barnes, R. (2015). Extreme water loss and abiotic O2 buildup on planets throughout the habitable zones of M dwarfs. Astrobiology 15, 119143.Google Scholar
Luger, R., et al. (2015). Habitable evaporated cores: Transforming mini-Neptunes into Super-Earths in the habitable zones of M dwarfs. Astrobiology 15, 5788.Google Scholar
Lunine, J., et al. (2009). The origin and evolution of Titan. In: Titan from Cassini-Huygens, ed. Brown, R. H., et al., New York: Springer, pp. 3559.Google Scholar
Lunine, J. I. (2010). Titan and habitable planets around M-dwarfs. Faraday Discuss. 147, 405418.Google Scholar
Lunine, J. I., et al. (2003). The origin of water on Mars. Icarus 165, 18.Google Scholar
Lunine, J. I. and Lorenz, R. D. (2009). Rivers, lakes, dunes, and rain: crustal processes in Titan’s methane cycle. Ann. Rev. Earth Planet. Sci. 37, 299320.Google Scholar
Lunine, J. I., et al. (1998). Some speculations on Titan’s past, present and future. Planetary and Space Science 46, 10991107.Google Scholar
Lunine, J. I. and Nolan, M. C. (1992). A massive early atmosphere on Triton. Icarus 100, 221234.Google Scholar
Lunine, J. I., et al. (2011). Dynamical models of terrestrial planet formation. Adv. Sci. Lett. 4, 325338.Google Scholar
Lunine, J. I., et al. (1983). Ethane ocean on Titan. Science 222, 12291230.Google Scholar
Lupu, R. E., et al. (2014). The atmospheres of Earth-like planets after giant impact events. Astrophys. J. 784.Google Scholar
Luth, R. W. and Canil, D. (1993). Ferric iron in mantle-derived pyroxenes and a new oxybarometer for the mantle. Contrib. Mineral. Petrol. 113, 236248.Google Scholar
Lynd, L., et al. (1982). Carbon monoxide metabolism of the methylotrophic acidogen Butyribacterium Methylotrophicum. J. Bacteriol. 149, 255263.Google Scholar
Lyons, J. R. (2009). Atmospherically-derived mass-independent sulfur isotope signatures, and incorporation into sediments. Chem. Geol. 267, 164174.Google Scholar
Lyons, J. R. and Young, E. D. (2005). CO self-shielding as the origin of oxygen isotope anomalies in the early solar nebula. Nature 435, 317320.Google Scholar
Lyons, T. W., et al. (2009). Tracking euxinia in the ancient ocean: A multiproxy perspective and Proterozoic case study. Ann. Rev. Earth Planet. Sci. 37, 507534.Google Scholar
Lyons, T. W., et al. (2014). The rise of oxygen in Earth’s early ocean and atmosphere. Nature 506, 307315.Google Scholar
Lyons, T. W. and Severmann, S. (2006). A critical look at iron paleoredox proxies: New insights from modern euxinic marine basins. Geochim. Cosmochim. Acta 70, 56985722.Google Scholar
Macdonald, F. A., et al. (2010). Calibrating the Cryogenian. Science 327, 12411243.Google Scholar
Macdonald, F. A., et al. (2013). The stratigraphic relationship between the Shuram carbon isotope excursion, the oxygenation of Neoproterozoic oceans, and the first appearance of the Ediacara biota and bilaterian trace fossils in northwestern Canada. Chem. Geol. 362, 250272.Google Scholar
MacGregor, A. M. (1927). The problem of the Precambrian atmosphere. S. African J. Sci. 24, 155172.Google Scholar
Machado, A. D. (1987). On the origin and age of the Steep Rock Buckshot, Ontario, Canada. Chem. Geol. 60, 337349.Google Scholar
Macintosh, B., et al. (2015). Discovery and spectroscopy of the young jovian planet 51 Eri b with the Gemini Planet Imager. Science 350, 6467.Google Scholar
Macouin, M., et al. (2015). Is the Neoproterozoic oxygen burst a supercontinent legacy? Front. Earth Sci. 3, 44, doi:10.3389/feart.2015.00044.Google Scholar
Madeleine, J. B., et al. (2011). Revisiting the radiative impact of dust on Mars using the LMD Global Climate Model. J. Geophys. Res. 116.Google Scholar
Magalhaes, J. A., et al. (2002). The stratification of Jupiter’s troposphere at the Galileo probe entry site. Icarus 158, 410433.Google Scholar
Magee, B. A., et al. (2009). INMS-derived composition of Titan’s upper atmosphere: Analysis methods and model comparison. Planet. Space Sci. 57, 18951916.Google Scholar
Mahaffy, P. R., et al. (2015a). Volatile and Isotopic Imprints of Ancient Mars. Elements 11, 5156.Google Scholar
Mahaffy, P. R., et al. (2015b). The imprint of atmospheric evolution in the D/H of Hesperian clay minerals on Mars. Science 347, 412414.Google Scholar
Maher, L. J. and Tinsley, B. A. (1977). Atomic hydrogen escape rate due to charge-exchange with hot plasmaspheric ions. J. Geophys. Res. 82, 689695.Google Scholar
Maheshwari, A., et al. (2010). Global nature of the Paleoproterozoic Lomagundi carbon isotope excursion A review of occurrences in Brazil, India, and Uruguay. Precam. Res. 182, 274299.Google Scholar
Maindl, T. I., et al. (2015). Impact induced surface heating by planetesimals on early Mars. Astron. Astrophys. 574, A22, doi: 10.1051/0004-6361/201424256.Google Scholar
Mak, M. (2011). Atmospheric Dynamics. Cambridge ; New York: Cambridge University Press.Google Scholar
Malin, M. C., et al. (2008). Climate, weather, and north polar observations from the Mars Reconnaissance Orbiter Mars Color Imager. Icarus 194, 501512.Google Scholar
Malin, M. C. and Edgett, K. S. (1999). Oceans or seas in the Martian northern lowlands: High resolution imaging tests of proposed coastlines. Geophys. Res. Lett. 26, 30493052.Google Scholar
Malin, M. C. and Edgett, K. S. (2000a). Evidence for recent groundwater seepage and surface runoff on Mars. Science 288, 23302335.Google Scholar
Malin, M. C. and Edgett, K. S. (2000b). Sedimentary rocks of early Mars. Science 290, 19271937.Google Scholar
Malin, M. C. and Edgett, K. S. (2001). Mars Global Surveyor Mars Orbiter Camera: Interplanetary cruise through primary mission. J. Geophys. Res. 106, 2342923570.Google Scholar
Malin, M. C. and Edgett, K. S. (2003). Evidence for persistent flow and aqueous sedimentation on early Mars. Science 302, 19311934.Google Scholar
Malin, M. C., et al. (2010). An overview of the 1985–2006 Mars Orbiter Camera science investigation. Mars 5, 160.Google Scholar
Malin, M. C., et al. (2006). Present-day impact cratering rate and contemporary gully activity on Mars. Science 314, 15731577.Google Scholar
Mallmann, G. and O’Neill, H. S. C. (2009). The crystal/melt partitioning of V during mantle melting as a function of oxygen fugacity compared with some other elements (Al, P, Ca, Sc, Ti, Cr, Fe, Ga, Y, Zr and Nb). J. Petrol. 50, 17651794.Google Scholar
Maloof, A. C., et al. (2010). Possible animal-body fossils in pre-Marinoan limestones from South Australia. Nature Geosci. 3 (9), 653659.Google Scholar
Manabe, S. and Strickler, R. F. (1964). Thermal equilibrium of the atmosphere with a convective adjustment. J. Atmos. Sci. 21, 361385.Google Scholar
Manabe, S. and Wetherald, R. T. (1967). Thermal equilibrium of the atmosphere with a given distribution of relative humidity. J. Atmos. Sci. 24, 241259.Google Scholar
Mancinelli, R. L. and McKay, C. P. (1988). The evolution of nitrogen cycling. Origins of Life 18, 311325.Google Scholar
Mandt, K. E., et al. (2009). Isotopic evolution of the major constituents of Titan’s atmosphere based on Cassini data. Planetary and Space Science 57, 19171930.Google Scholar
Manga, M., et al. (2012). Wet surface and dense atmosphere on early Mars suggested by the bomb sag at Home Plate, Mars. Geophys. Res. Lett. 39.Google Scholar
Mangold, N. and Howard, A. D. (2013). Outflow channels with deltaic deposits in Ismenius Lacus, Mars. Icarus 226, 385401.Google Scholar
Mangold, N., et al. (2012). The origin and timing of fluvial activity at Eberswalde crater, Mars. Icarus 220, 530551.Google Scholar
Manning, C. E., et al. (2006a). Geology, age and origin of supracrustal rocks at Akilia, West Greenland. American Journal of Science 306, 303366.Google Scholar
Manning, C. V., et al. (2006b). Thick and thin models of the evolution of carbon dioxide on Mars. Icarus 180, 3859.Google Scholar
Marchi, S., et al. (2013). High-velocity collisions from the lunar cataclysm recorded in asteroidal meteorites. Nat. Geosci. 6, 303307.Google Scholar
Marchi, S., et al. (2014). Widespread mixing and burial of Earth’s Hadean crust by asteroid impacts. Nature 511, 578582.Google Scholar
Marcq, E., et al. (2008). A latitudinal survey of CO, OCS, H2O, and SO2 in the lower atmosphere of Venus: Spectroscopic studies using VIRTIS-H. J. Geophys. Res. 113, doi: 10.1029/2008JE003074.Google Scholar
Marcq, E., et al. (2006). Remote sensing of Venus’ lower atmosphere from ground-based IR spectroscopy: Latitudinal and vertical distribution of minor species. Planet. Space Sci. 54, 13601370.Google Scholar
Margulis, L. and Lovelock, J. E. (1974). Biological modulation of the Earth’s atmosphere. Icarus 21, 471489.Google Scholar
Margulis, L., et al. (1976). Reassessment of roles of oxygen and ultraviolet light in Precambrian evolution. Nature 264, 620624.Google Scholar
Marinova, M. M., et al. (2008). Mega-impact formation of the Mars hemispheric dichotomy. Nature 453, 12161219.Google Scholar
Markiewicz, W. J., et al. (2014). Glory on Venus cloud tops and the unknown UV absorber. Icarus 234, 200203.Google Scholar
Marley, M. S. (2010). The atmospheres of extrasolar planets. EAS Publications Series 41, 411428.Google Scholar
Marley, M. S., et al. (2013). Clouds and hazes in exoplanets. In: Comparative Climatology of Terrestrial Planets, ed. Mackwell, S. J., et al., Tucson: University of Arizona Press, pp. 367391.Google Scholar
Marounina, N., et al. (2015). Evolution of Titan’s atmosphere during the Late Heavy Bombardment. Icarus 257, 324335.Google Scholar
Marrero, T. R. and Mason, E. A. (1972). Gaseous diffusion coefficients. J. Phys. Chem. Ref. Data 1, 3118.Google Scholar
Marshall, A. O., et al. (2014). Multiple generations of carbonaceous material deposited in Apex chert by basin-scale pervasive hydrothermal fluid flow. Gondwana Res. 25, 284289.Google Scholar
Marshall, A. O. and Marshall, C. P. (2013). Comment on “Biogenicity of Earth’s earliest fossils: A resolution of the controversy” by J. W. Schopf and A. B. Kudryavtsev, Gondwana Research, Volume 22, Issue 3–4, Pages 761–771. Gondwana Res. 23, 16541655.Google Scholar
Marshall, C. P., et al. (2011). Haematite pseudomicrofossils present in the 3.5-billion-year-old Apex Chert. Nat. Geosci. 4, 240243.Google Scholar
Marshall, H. G., et al. (1988). Long-term climate change and the geochemical cycle of carbon. J. Geophys. Res. 93, 791802.Google Scholar
Marshall, J. and Plumb, R. A. (2008). Atmosphere, Ocean, and Climate Dynamics: An Introductory Text. Boston: Academic Press.Google Scholar
Martin, R. V., et al. (2002). An improved retrieval of tropospheric nitrogen dioxide from GOME. J. Geophys. Res. 107, 4437, doi: 10.1029/2001JD001027.Google Scholar
Martin, W., et al. (2008). Hydrothermal vents and the origin of life. Nat. Rev. Microbiol. 6, 805814.Google Scholar
Marty, B. (2012). The origins and concentrations of water, carbon, nitrogen and noble gases on Earth. Earth Planet. Sc. Lett. 313314, 5666.Google Scholar
Marty, B., et al. (2011). A 15N-poor isotopic composition for the Solar System as shown by Genesis solar wind samples. Science 332, 15331536.Google Scholar
Marty, B. and Dauphas, N. (2003). The nitrogen record of crust-mantle interaction and mantle convection from Archean to present. Earth Planet. Sci. Lett. 206, 397410.Google Scholar
Marty, B. and Yokochi, R. (2006). Water in the early Earth. Rev. Mineral. Geochem. 62, 421450.Google Scholar
Marty, B. and Zimmermann, L. (1999). Volatiles (He, C, N, Ar) in mid-ocean ridge basalts: Assesment of shallow-level fractionation and characterization of source composition. Geochim. Cosmochim. Acta 63, 36193633.Google Scholar
Marty, B., et al. (2013). Nitrogen isotopic composition and density of the Archean atmosphere. Science 342, 101104.Google Scholar
Martyn, D. F. and Pulley, O. O. (1936). The temperature and constituents of the upper atmosphere. Proc. R. Soc. Lond. A 154, 455486.Google Scholar
Mason, R. (1990). Petrology of the Metamorphic Rocks. Boston: Unwin Hyman.Google Scholar
Masse, M., et al. (2008). Mineralogical composition, structure, morphology, and geological history of Aram Chaos crater fill on Mars derived from OMEGA Mars Express data. J. Geophys. Res. 113, E12006.Google Scholar
Massie, S. T. and Hunten, D. M. (1981). Stratospheric eddy diffusion coefficients from tracer data. J. Geophys. Res. 86, 98599868.Google Scholar
Mastrogiuseppe, M., et al. (2014). The bathymetry of a Titan sea. Geophys. Res. Lett. 41, 14321437.Google Scholar
Mather, T. A. (2015). Volcanoes and the environment: Lessons for understanding Earth’s past and future from stydies of present-day volcanic emissions. J. Volcanol. Geoth. Res., doi:10.1016/j.jvolgeores.2015.08.016.Google Scholar
Matsubara, Y., et al. (2013). Hydrology of early Mars: Valley network incision. J. Geophys. Res. 118, 13651387.Google Scholar
Matsui, T. and Abe, Y. (1986a). Evolution of an impact-induced atmosphere and magma ocean on the accreting Earth. Nature 319, 303305.Google Scholar
Matsui, T. and Abe, Y. (1986b). Impact-induced atmospheres and oceans on Earth and Venus. Nature 322, 526528.Google Scholar
Matsuo, T. and Tamura, M. (2010). Second-earth imager for TMT (SEIT): a proposal and concept Description. SPIE Astronomical Telescopes + Instrumentation. International Society for Optics and Photonics, pp. 773584–773584-9.Google Scholar
Mattey, D. P. (1987). Carbon isotopes in the mantle. Terra Cognita 7, 3137.Google Scholar
Mattioli, G. S. and Wood, B. J. (1986). Upper mantle oxygen fugacity recorded by spinel lherzolites. Nature 322, 626628.Google Scholar
Maynard, J. B. (2010). The chemistry of manganese ores through time: A signal of increasing diversity of Earth-surface environments. Econ. Geol. 105, 535552.Google Scholar
Mayor, M., et al. (2014). Doppler spectroscopy as a path to the detection of Earth-like planets. Nature 513, 328335.Google Scholar
Mayor, M. and Queloz, D. (1995). A Jupiter-mass companion to a solar-type star. Nature 378, 355359.Google Scholar
Mazzullo, S. J. (2000). Organogenic dolomitization in peritidal to deep-sea sediments. J. Sediment. Res. 70, 1023.Google Scholar
McBride, N. and Gilmour, I. (2004). An Introduction to the Solar System. New York: Cambridge University Press.Google Scholar
McCammon, C. (2005). The paradox of mantle redox. Science 308, 807808.Google Scholar
McCauley, J. F. (1978). Geologic Map of the Coprates Quadrangle of Mars, scale 1:5,000,000. U.S. Geol. Surv. Misc. Inv. Series Map I-897.Google Scholar
McCleese, D. J., et al. (2010). Structure and dynamics of the Martian lower and middle atmosphere as observed by the Mars Climate Sounder: Seasonal variations in zonal mean temperature, dust, and water ice aerosols. J. Geophys. Res. 115.Google Scholar
McCollom, T. M., et al. (2007). Could erosion of Meridiani Planum represent a significant contributer to global sulfate-rich martian soils. 38th Lunar Planet. Sci. Conf., no. 2151 (abstract).Google Scholar
McCollom, T. M. and Seewald, J. S. (2007). Abiotic synthesis of organic compounds in deep-sea hydrothermal environments. Chem. Rev. 107, 382401.Google Scholar
McCord, T. B., et al. (2008). Titan’s surface: Search for spectral diversity and composition using the Cassini VIMS investigation. Icarus 194, 212242.Google Scholar
McDermott, J. M., et al. (2015). Pathways for abiotic organic synthesis at submarine hydrothermal fields. P. Natl. Acad. Sci. USA 112, 76687672.Google Scholar
McDonough, W. F. (2003). Compositional model for the Earth’s core. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., Oxford: Pergamon, pp. 547568.Google Scholar
McElroy, M. B. (1972). Mars: An evolving atmosphere. Science 175, 443445.Google Scholar
McElroy, M. B. and Donahue, T. M. (1972). Stability of the Martian atmosphere. Science 177, 986988.Google Scholar
McElroy, M. B., et al. (1982). Escape of hydrogen from Venus. Science 215, 16141615.Google Scholar
McEwen, A. S. (2013). Mars in Motion: The surface of Mars changes all the time. Is flowing water one of the causes? Sci. Am. 308, 5865.Google Scholar
McEwen, A. S., et al. (2014). Recurring slope lineae in equatorial regions of Mars. Nature Geoscience 7, 5358.Google Scholar
McEwen, A. S., et al. (2004). The lithosphere and surface of Io. In: Jupiter: The planet, satellites, and magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press, pp. 307328.Google Scholar
McEwen, A. S., et al. (2011). Seasonal flows on warm martian slopes. Science 333, 740743.Google Scholar
McFadden, K. A., et al. (2008). Pulsed oxidation and bioloical evolution in the Ediacaran Doushantuo Formation. P. Natl. Acad. Sci. USA 105, 31973202.Google Scholar
McGouldrick, K. and Toon, O. B. (2007). An investigation of possible causes of the holes in the condensational Venus cloud using a microphysical cloud model with a radiative-dynamical feedback. Icarus 191, 124.Google Scholar
McGouldrick, K., et al. (2011). Sulfuric acid aerosols in the atmospheres of the terrestrial planets. Planet. Space Sci. 59, 934941.Google Scholar
McGovern, W. E. (1973). Potential atmospheric composition of smaller bodies in Solar System and some aspects of planetary evolution. J. Geophys. Res. 78, 274280.Google Scholar
McGrath, M. A., et al. (2004). Satellite atmospheres. In: Jupiter: The planet, satellites, and magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press.Google Scholar
McIlveen, J. F. R. (1992). Fundamentals of Weather and Climate. London: Chapman & Hall.Google Scholar
McKay, C. P. (2000). Thickness of tropical ice and photosynthesis on a snowball Earth. Geophys. Res. Lett. 27, 21532156.Google Scholar
McKay, C. P. (2004). Wet and cold thick atmosphere on early Mars. J. Phys. IV France 121, 283288.Google Scholar
McKay, C. P., et al. (1997). Temperature lapse rate and methane in Titan’s troposphere. Icarus 129, 498505.Google Scholar
McKay, C. P., et al. (1989). The thermal structure of Titan’s atmosphere. Icarus 80, 2353.Google Scholar
McKay, C. P., et al. (1991). The greenhouse and antigreenhouse effects on Titan. Science 253, 11181121.Google Scholar
McKay, C. P., et al. (1993). Coupled atmosphere ocean models of Titan’s past. Icarus 102, 8898.Google Scholar
McKay, C. P., et al. (1988). High temperature shock formation of N2 and organics on primordial Titan. Nature 332, 520522.Google Scholar
McKay, C. P. and Smith, H. D. (2005). Possibilities for methanogenic life in liquid methane on the surface of Titan. Icarus 178, 274276.Google Scholar
McKeegan, K. D., et al. (2006). Isotopic compositions of cometary matter returned by Stardust. Science 314, 17241728.Google Scholar
McKenzie, D. and Nimmo, F. (1999). The generation of Martian floods by the melting of ground ice above dykes. Nature 397, 231233.Google Scholar
McKinnon, W. B. (2002). Planetary science – Out on the edge. Nature 418, 135–+.Google Scholar
McKinnon, W. B. (2010). Argon-40 outgassing from Titan and Enceladus: A tale of two satellites. 41st Lunar Planet. Sci. Conf., 2718 (abstract).Google Scholar
McKinnon, W. B., et al. (1995). Origin and evolution of Triton. In: Neptune and Triton, ed. Cruikshank, D. P., Tucson: University of Arizona Press, pp. 807877.Google Scholar
McKinnon, W. B., et al. (2008). Structure and evolution of Kuiper Belt Objects and dwarf planets. In: The Solar System Beyond Neptune, ed. Barucci, M. A., et al., Tucson: University of Arizona Press, pp. 213241.Google Scholar
McLennan, S. M. and Grotzinger, J. P. (2008). The sedimentary rock cycle on Mars. In: The Martian Surface: Composition, Mineralogy and Physical Properties, ed. Bell, J., New York: Cambridge University Press, pp. 541577.Google Scholar
McLoughlin, N., et al. (2008). Growth of synthetic stromatolites and wrinkle structures in the absence of microbes: Implications for the early fossil record. Geobiology 6, 95105.Google Scholar
McNutt, R. L. (1989). Models of Pluto upper atmosphere. Geophys. Res. Lett. 16, 12251228.Google Scholar
McQuarrie, D. A. and Simon, J. D. (1997). Physical Chemistry: A Molecular Approach. Sausalito: University Science Books.Google Scholar
McSween, H. Y. and McLennan, S. M. (2014). Mars. In: Treatise on Geochemistry (2nd Edn), ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 251300.Google Scholar
Meadows, V. S. and Crisp, D. (1996). Ground-based near-infrared observations of the Venus nightside: The thermal structure and water abundance near the surface. J. Geophys. Res. 101, 45954622.Google Scholar
Meech, K. J., et al. (2011). Epoxi: Comet 103P/Hartley 2 observations from a worldwide campaign. Astrophys. J. Lett. 734.Google Scholar
Melezhik, V. A. and Fallick, A. E. (2010). On the Lomagundi-Jatuli carbon isotopic event: The evidence from the Kalix Greenstone Belt, Sweden. Precam. Res. 179, 165190.Google Scholar
Melezhik, V. A., et al. (2013a). The Palaeoproterozoic perturbation of the global carbon cycle: The Lomagundi-Jatuli isotopic event. In: Reading the Archive of Earth’s Oxygenation, ed. Melezhik, V. A. e. a., Berlin: Springer, pp. 11111150.Google Scholar
Melezhik, V. A., et al. (1999). Extreme 13Ccarb enrichment in ca. 2.0 Ga magnesite-stromatolite-dolomite- ‘red beds’ association in a global context: A case for the world-wide signal enhanced by a local environment. Earth Sci. Rev. 48, 71120.Google Scholar
Melezhik, V. A., et al. (2007). Temporal constraints on the Paleoproterozoic Lomagundi-Jatuli carbon isotopic event. Geology 35, 655658.Google Scholar
Melezhik, V. A., et al. (2013b). Huronian-age glaciation. In: Reading the Archive of Earth’s Oxygenation, ed. Melezhik, V. A. e. a., Berlin: Springer, pp. 10591109.Google Scholar
Melnik, Y. P. (1982). Precambrian Banded Iron-Formations: Physicochemical Conditions Of Formation. Amsterdam: Elsevier.Google Scholar
Melosh, H. J. and Vickery, A. M. (1989). Impact erosion of the primordial atmosphere of Mars. Nature 338, 487489.Google Scholar
Menou, K. (2015). Climate stability of habitable Earth-like planets. Earth Planet. Sc. Lett. 429, 2024.Google Scholar
Menou, K., et al. (2003). “Weather” variability of close-in extrasolar giant planets. Astrophys. J. 587, L113L116.Google Scholar
Menou, K. and Rauscher, E. (2009). Atmospheric circulation of hot Jupiters: A shallow three-dimensional model. Astrophys. J. 700, 887897.Google Scholar
Mercer, C. M., et al. (2015). Refining lunar impact chronology through high spatial resolution 40Ar/39Ar dating of impact melts. Sci. Adv. 1, e1400050 doi:10.1126/sciadv.1400050.Google Scholar
Merlis, T. M. and Schneider, T. (2011). Changes in zonal surface temperature gradients and Walker Circulations in a wide range of climates. J. Climate 24, 47574768.Google Scholar
Merryfield, W. J. and Shizgal, B. D. (1994). Discrete velocity model for an escaping single-component atmosphere. Planet. Space Sci. 42, 409419.Google Scholar
Michael, M., et al. (2005). Ejection of nitrogen from Titan’s atmosphere by magnetospheric ions and pick-up ions. Icarus 175, 263267.Google Scholar
Michalski, G., et al. (2004). Long term atmospheric deposition as the source of nitrate and other salts in the Atacama Desert, Chile: New evidence from mass-independent oxygen isotopic compositions. Geochim. Cosmochim. Acta 68, 40234038.Google Scholar
Michalski, J. and Niles, P. B. (2010). Deep crustal carbonate rocks exposed by meteor impact on Mars. Nature Geosci. 3, 751755.Google Scholar
Michalski, J. and Niles, P. B. (2012). Atmospheric origin of martian interior layered deposits: Links to climate change and global sulfur cycle. Geology 40, 419422.Google Scholar
Migliorini, A., et al. (2012). Investigation of air temperature on the nightside of Venus derived from VIRTIS-H on board Venus-Express. Icarus 217, 640647.Google Scholar
Mikucki, J. A., et al. (2009). A contemporary microbially maintained subglacial ferrous “ocean”. Science 324, 397400.Google Scholar
Miller, C. A., et al. (2011). Re-assessing the surface cycling of molybdenum and rhenium. Geochim. Cosmochim. Acta 75, 71467179.Google Scholar
Miller, S. L. (1953). A production of amino acids under possible primitive Earth conditions. Science 117, 528529.Google Scholar
Miller, S. L. (1955). Production of some organic compounds under possible primitive Earth conditions. J. Am. Chem. Soc. 77, 23512361.Google Scholar
Miller, S. L. and Bada, J. L. (1988). Submarine hot springs and the origin of life. Nature 334, 609611.Google Scholar
Miller, S. L. and Schlesinger, G. (1984). Carbon energy yields in atmospheres containing CH4, CO, and CO2. Origins of Life 14, 8390.Google Scholar
Milliken, R. E. and Bish, D. L. (2010). Sources and sinks of clay minerals on Mars. Philos. Mag. 90, 22932308.Google Scholar
Milliken, R. E., et al. (2008). Opaline silica in young deposits on Mars. Geology 36, 847850.Google Scholar
Mills, B., et al. (2011). Timing of Neoproterozoic glaciations linked to transport-limited global weathering. Nature Geosci. 4, 861864.Google Scholar
Mills, D. B. and Canfield, D. E. (2014). Oxygen and animal evolution: Did a rise of atmospheric oxygen trigger the origin of animals? Bioessays 36, doi:10.1002/bies.201400101.Google Scholar
Mills, F. P. and Allen, M. (2007). A review of selected issues concerning the chemistry in Venus’ middle atmosphere. Planet. Space Sci. 55, 17291740.Google Scholar
Mills, F. P., et al. (2007). Atmospheric composition, chemistry, and clouds. In: Exploring Venus as a Terrestrial Planet, American Geophysical Union, pp. 73100.Google Scholar
Ming, D. W., et al. (2014). Volatile and organic compositions of sedimentary rocks in Yellowknife Bay, Gale Crater, Mars. Science 343, 1245267, doi: 10.1126/science.1245267.Google Scholar
Misra, A., et al. (2014). Using dimers to measure biosignatures and atmospheric pressure for terrestrial exoplanets. Astrobiology 14, 6786.Google Scholar
Mitchell, J. L. and Vallis, G. K. (2010). The transition to superrotation in terrestrial atmospheres. J. Geophys. Res. 115, E12008, doi:10.1029/2010JE003587.Google Scholar
Mlawer, E. J., et al. (1997). Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model of the longwave. J. Geophys. Res. 102, 16 66316 682.Google Scholar
Mojzsis, S. J., et al. (1996). Evidence for life on Earth before 3,800 million years ago. Nature 384, 5559.Google Scholar
Mojzsis, S. J., et al. (2001). Oxygen isotope evidence from ancient zircons for liquid water at the Earth’s surface 4,300 Myr ago. Nature 409, 178181.Google Scholar
Monga, N. and Desch, S. (2015). External photoevaporation of the solar nebula: Jupiter’s noble gas enrichments. Astrophys. J. 798.Google Scholar
Monteith, J. L. (2012). Principles of Environmental Physics. Oxford: Academic.Google Scholar
Montmessin, F., et al. (2011). A layer of ozone detected in the nightside upper atmosphere of Venus. Icarus 216, 8285.Google Scholar
Montmessin, F., et al. (2005). Modeling the annual cycle of HDO in the Martian atmosphere. J. Geophys. Res. 110, E03006, doi: 10.1029/2004JE002357.Google Scholar
Montzka, S. A., et al. (2007). On the global distribution, seasonality, and budget of atmospheric carbonyl sulfide (COS) and some similarities to CO2. J. Geophys. Res. 112, D09302, doi:10.1029/2006JD007665.Google Scholar
Moorbath, S., et al. (1973). Early Archaean age for Isua iron formation, West Greenland. Nature 245, 138139.Google Scholar
Moore, J. M. and Howard, A. D. (2005). Large alluvial fans on Mars. J. Geophys. Res. 110.Google Scholar
Moore, J. M., et al. (2003). Martian layered fluvial deposits: implications for Noachian climate scenarios. Geophys. Res. Lett. 30.Google Scholar
Moore, J. M. and Pappalardo, R. T. (2011). Titan: An exogenic world? Icarus 212, 790806.Google Scholar
Moore, W. B. and Webb, A. A. G. (2013). Heat-pipe Earth. Nature 501, 501505.Google Scholar
Moores, E. M. (1986). The Proterozoic ophiolite problem, continental emergence, and the Venus connection. Science 234, 6568.Google Scholar
Moores, E. M. (1993). Neoproterozoic oceanic crustal thinning, emergence of continents, and origin of the Phanerozoic ecosystem; a model. Geol. 21, 58.Google Scholar
Moores, E. M. (2002). Pre-1 Ga (pre-Rodinian) ophiolites: Their tectonic and environmental implications. GSA Bull. 114, 8095.Google Scholar
Morbidelli, A., et al. (2000). Source regions and timescales for the delivery of water to the Earth. Meteoritics and Planet. Sci. 35, 13091320.Google Scholar
Moreira, M. (2013). Noble gas constraints on the origin and evolution of Earth’s volatiles. Geochem. Perspectives 2, 229403.Google Scholar
Moreno, R., et al. (2012). The abundance, vertical distribution and origin of H2O in Titan’s atmosphere: Herschel observations and photochemical modelling. Icarus 221, 753767.Google Scholar
Moresi, L. and Solomatov, V. (1998). Mantle convection with a brittle lithosphere: thoughts on the global tectonic styles of the Earth and Venus. Geophys. J. Int. 133, 669682.Google Scholar
Morii, H., et al. (1987). Energetic analysis of the growth of Methanobrevibacter-Arboriphilus A2 in hydrogen-limited continuous cultures. Biotechnology and Bioengineering 29, 310315.Google Scholar
Morner, N. A. and Etiope, G. (2002). Carbon degassing from the lithosphere. Global And Planetary Change 33, 185203.Google Scholar
Moroz, V. I., et al. (1985). Solar and thermal radiation in the Venus atmosphere. Adv. Space Res. 5, 197232.Google Scholar
Morris, R. V., et al. (2006). Mössbauer mineralogy of rock, soil, and dust at Gusev crater, Mars: Spirit’s journey through weakly altered olivine basalt on the plains and pervasively altered basalt in the Columbia Hills. Journal of Geophysical Research-Planets 111.Google Scholar
Morris, R. V., et al. (2010). Identification of carbonate-rich outcrops on Mars by the Spirit Rover. Science 329, 421424.Google Scholar
Moses, J. I., et al. (2004). The stratosphere of Jupiter. In: Jupiter: The Planet, Satellites, and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press.Google Scholar
Moses, J. I., et al. (2011). Disequilibrium carbon, oxygen, and nitrogen chemistry in the atmospheres of HD 189733b and HD 209458b. Astrophys. J. 737.Google Scholar
Moses, J. I., et al. (2002). Photochemistry of a volcanically driven atmosphere on Io: Sulfur and oxygen species from a Pele-type eruption. Icarus 156, 76106.Google Scholar
Moskalenko, N. I., et al. (1979). Pressure-induced infrared radiation absorption in atmospheres. Bull. Acad. Sci. USSR, Atmospheric and Oceanic Physics 15, 632637.Google Scholar
Mottl, M. J. and Wheat, C. G. (1994). Hydrothermal circulation through mid-ocean ridge flanks: fluxes of heat and magnesium. Geochim. Cosmochim. Acta 58, 22252237.Google Scholar
Mouginot, J., et al. (2010). The 3–5 MHz global reflectivity map of Mars by MARSIS/Mars Express: Implications for the current inventory of subsurface H2O. Icarus 210, 612625.Google Scholar
Mousis, O., et al. (2014). Equilibrium composition between liquid and clathrate reservoirs on Titan. Icarus 239, 3945.Google Scholar
Mousis, O., et al. (2009a). A primordial origin for the atmospheric methane of Saturn’s moon Titan. Icarus 204, 749751.Google Scholar
Mousis, O., et al. (2011). Removal of Titan’s atmospheric noble gases by their sequestration in surface clathrates. Astrophys. J. Lett. 740.Google Scholar
Mousis, O., et al. (2009b). Clathration of volatiles in the solar nebula and implications for the origin of Titan’s atmosphere. Astrophys. J. 691, 17801786.Google Scholar
Mueller, R. F. and Saxena, S. K. (1977). Chemical Petrology. New York: Springer-Verlag.Google Scholar
Mulders, G., D., et al. (2015). The snow line in viscous disks around low-mass stars: Implications for water delivery to terrestrial planets in the habitable zone. Astrophys. J. 807, 9.Google Scholar
Mumma, M. J., et al. (2009). Strong release of methane on Mars in northern summer 2003. Science 323, 10411045.Google Scholar
Murakami, T., et al. (2011). Quantification of atmospheric oxygen levels during the Paleoproterozoic using paleosol compositions and iron oxidation kinetics. Geochim. Cosmochim. Acta 75, 39824004.Google Scholar
Murakami, T., et al. (2001). Direct evidence of Late Archean to Early Proterozoic anoxic atmosphere from a product of 2.5 Ga old weathering. Earth Planet. Sci. Lett. 184, 523.Google Scholar
Murchie, S., et al. (2009a). Evidence for the origin of layered deposits in Candor Chasma, Mars, from mineral composition and hydrologic modeling. J. Geophys. Res. 114, E00D05, doi:10.1029/2009JE003343.Google Scholar
Murchie, S. L., et al. (2009b). A synthesis of Martian aqueous mineralogy after 1 Mars year of observations from the Mars Reconnaissance Orbiter. J. Geophys. Res. 114, E00D06, doi:10.1029/2009JE003342.Google Scholar
Murray, C. D. and Dermott, S. F. (2001). Solar System Dynamics. New York: Cambridge Univ. Press.Google Scholar
Murray-Clay, R. A., et al. (2009). Atmospheric escape from hot Jupiters. Astrophys. J. 693, 2342.Google Scholar
Mushkin, A., et al. (2010). Spectral constraints on the composition of low-albedo slope streaks in the Olympus Mons Aureole. Geophys. Res. Lett. 37, L22201, doi: 10.1029/2010GL044535.Google Scholar
Mustard, J. F., et al. (2001). Evidence for recent climate change on Mars from the identification of youthful near-surface ground ice. Nature 412, 411414.Google Scholar
Mustard, J. F., et al. (2012). Sequestration of volatiles in the martian crust through hydrated minerals: Implications for planetary evolution. Third Int. Conf. on Early Mars, 7075.Google Scholar
Mutch, T. A., et al. (1976). The Geology of Mars. Princeton: Princeton University Press.Google Scholar
Nagy, A. F., et al. (2001). Hot carbon densities in the exosphere of Mars. J. Geophys. Res. 106, 2156521568.Google Scholar
Nair, H., et al. (1994). A photochemical model of the Martian atmosphere. Icarus 111, 124150.Google Scholar
Nakajima, S., et al. (1992). A study on the “runaway greenhouse effect” with a one-dimensional radiative-convective equilibrium model. J. Atmos. Sci. 49, 22562266.Google Scholar
Nakamura, K. and Kato, Y. (2004). Carbonatization of oceanic crust by the seafloor hydrothermal activity and its significance as a CO2 sink in the Early Archean. Geochim. Cosmochim. Acta 68, 45954618.Google Scholar
Nappo, C. J. (2012). An Introduction to Atmospheric Gravity Waves (2nd Edn). Waltham, MA: Academic Press.Google Scholar
Nappo, C. J. (2013). An Introduction to Atmospheric Gravity Waves (2nd Edn). Waltham, MA: Elsevier.Google Scholar
Narbonne, G. M. (2005). The Ediacara biota: Neoproterozoic origin of animals and their ecosystems. Ann. Rev. Earth Planet. Sci. 33, 421442.Google Scholar
Narbonne, G. M. and Gehling, J. G. (2003). Life after the Snowball: The oldest complex Ediacaran fossils. Geology 31, 2730.Google Scholar
National Research Council. (1979). Stratospheric Ozone Depletion by Halocarbons: Chemistry and Transport. National Academy of Sciences.Google Scholar
Navarra, A. and Boccaletti, G. (2002). Numerical general circulation experiments of sensitivity to Earth rotation rate. Clim. Dynam. 19, 467483.Google Scholar
Navarro-Gonzalez, R., et al. (2001). A possible nitrogen crisis for Archean life due to reduced nitrogen fixation by lightning. Nature 412, 6164.Google Scholar
Navarro-Gonzalez, R., et al. (2010). Reanalysis of the Viking results suggests perchlorate and organics at midlatitudes on Mars. J. Geophys. Res. 115.Google Scholar
Navarro-Gonzalez, R., et al. (2011). Correction to “Reanalysis of the Viking results suggests perchlorate and organics at midlatitudes on Mars”. J. Geophys. Res. 116, E08011.Google Scholar
Nedelcu, A. M., et al. (2004). Sex as a response to oxidative stress: a twofold increase in cellular reactive oxygen species activates sex genes. Proc. R. Soc. Lond. B–Biol. Sci. 271, 15911596.Google Scholar
Nedell, S. S., et al. (1987). Origin and evolution of the layered deposits in the Valles Marineris, Mars. Icarus 70, 409441.Google Scholar
Neumann, G. A., et al. (2004). Crustal structure of Mars from gravity and topography. J. Geophys. Res. 109.Google Scholar
Newman, C. E., et al. (2011). Stratospheric superrotation in the TitanWRF model. Icarus 213, 636654.Google Scholar
Newsom, H. E. and Sims, K. W. W. (1991). Core formation during early accretion of the Earth. Science 252, 926933.Google Scholar
Nicklas, R. W., et al. (2017). The redox history of the Archean Mantle: Evidence from komatiites, in preparation.Google Scholar
Niemann, H. B., et al. (2010). Composition of Titan’s lower atmosphere and simple surface volatiles as measured by the Cassini-Huygens probe gas chromatograph mass spectrometer experiment. J. Geophys. Res. 115.Google Scholar
Nier, A. O. and McElroy, M. B. (1977). Composition and structure of Mars’ upper atmosphere: Results from the neutral mass spectrometers on Viking 1 and 2. J. Geophys. Res. 82, 4341–49.Google Scholar
Niles, P. B., et al. (2013). Geochemistry of carbonates on Mars: Implications for climate history and nature of aqueous environments. Space Sci. Rev. 174, 301328.Google Scholar
Niles, P. B. and Michalski, J. (2009). Meridiani Planum sediments on Mars formed through weathering in massive ice deposits. Nature Geoscience 2, 215220.Google Scholar
Nimmo, F. (2000). Dike intrusion as a possible cause of linear Martian magnetic anomalies. Geology 28, 391394.Google Scholar
Nimmo, F., et al. (2008). Implications of an impact origin for the martian hemispheric dichotomy. Nature 453, 1220–U32.Google Scholar
Nimmo, F. and Stevenson, D. J. (2000). Influence of early plate tectonics on the thermal evolution and magnetic field of Mars. J. Geophys. Res. 105, 1196911979.Google Scholar
Nimmo, F. and Tanaka, K. (2005). Early crustal evolution of mars. Annu. Rev. Earth Pl. Sc. 33, 133161.Google Scholar
Nisbet, E., et al. (2007a). Creating habitable zones, at all scales, from planets to mud micro-habitats, on earth and on mars. Space Sci. Rev. 129, 79121.Google Scholar
Nisbet, E. G. (2002). Fermor lecture: The influence of life on the face of the Earth: garnets and moving continents. In: The Early Earth: Physical, Chemical and Biological Development, ed. Fowler, C. M. R., et al., London: Geological Soc. of London.Google Scholar
Nisbet, E. G. and Fowler, C. M. R. (2014). The early history of life. In: Treatise on Geochemistry (2nd Edn), ed. Holland, H. D. and Turekian, K. K., Oxford: Elsevier, pp. 142.Google Scholar
Nisbet, E. G., et al. (2012). The regulation of the air: a hypothesis. Solid Earth 3, 8796.Google Scholar
Nisbet, E. G., et al. (2007b). The age of Rubisco: The evolution of oxygenic photosynthesis. Geobiology 5, 311335.Google Scholar
Nisbet, E. G. and Nisbet, R. E. R. (2008). Methane, oxygen, photosynthesis, rubisco and the regulation of the air through time. Phil. Trans. R. Soc. Lond. B 363, 27452754.Google Scholar
Nisbet, E. G. and Sleep, N. H. (2001). The habitat and nature of early life. Nature 409, 10831091.Google Scholar
Nishiizumi, K., et al. (1991). Cosmic ray produced 10Be and 26Al in Antarctic rocks: Exposure and erosion history. Earth Planet. Sci. Lett. 104, 440454.Google Scholar
Nishizawa, M., et al. (2007). Speciation and isotope ratios of nitrogen in fluid inclusions from seafloor hydrothermal deposits at similar to 3.5 Ga. Earth Planet. Sci. Lett. 254, 332344.Google Scholar
Nixon, C. A., et al. (2008). Isotopic ratios in Titan’s atmosphere from Cassini CIRS limb sounding: CO2 at low and midlatitudes. Astrophys. J. Lett. 681, L101L103.Google Scholar
Noel, A., et al. (2015). Mineralogy, morphology and stratigraphy of the light-toned interior layered deposits at Juventae Chasma. Icarus 251, 315331.Google Scholar
Nogueira, E., et al. (2011). Reassessing the origin of Triton. Icarus 214, 113130.Google Scholar
Nordstrom, D. K. and Munoz, J. L. (1994). Geochemical Thermodynamics. Boston: Blackwell Scientific Publications.Google Scholar
Norman, M. D., et al. (2003). Chronology, geochemistry, and petrology of a ferroan noritic anorthosite clast from Descartes breccia 67215: Clues to the age, origin, structure, and impact history of the lunar crust. Meteorit. Planet. Sci. 38, 645661.Google Scholar
Norman, M. D., et al. (2010). Imbrium provenance for the Apollo 16 Descartes terrain: Argon ages and geochemistry of lunar breccias 67016 and 67455. Geochim. Cosmochim. Acta 74, 763783.Google Scholar
Norman, M. D. and Nemchin, A. A. (2014). A 4.2 billion year old impact basin on the Moon: U-Pb dating of zirconolite and apatite in lunar melt rock 67955. Earth Planet. Sci. Lett. 388, 387398.Google Scholar
North, G. R. (1975). Theory of energy-balance climate models. J. Atmos. Sci. 32, 20332043.Google Scholar
Nursall, J. R. (1959). Oxygen as a prerequisite to the origin of the metazoa. Nature 183, 11701172.Google Scholar
Nutman, A. P. (2006). Antiquity of the oceans and continents. Elements 2, 223227.Google Scholar
Nutman, A. P., et al. (1997). Recognition of >=3850 Ma water-lain sediments in West Greenland and their significance for the early Archaean Earth. Geochim. Cosmochim. Acta 61, 24752484.Google Scholar
Nutman, A. P., et al. (2016). Rapid emergence of life shown by discovery of 3,700-million-year-old microbial structures. Nature 537, 535538.Google Scholar
O’Brien, D. P., et al. (2014). Water delivery and giant impacts in the ‘Grand Tack’ scenario. Icarus 239, 7484.Google Scholar
Oakley, P. H. H. and Cash, W. (2009). Construction of an Earth model: Analysis of exoplanet light curves and mapping the next Earth with the New Worlds Observer. Astrophys. J. 700, 14281439.Google Scholar
Och, L. M. and Shields-Zhou, G. A. (2012). The Neoproterozoic oxygenation event: Environmental perturbations and biogeochemical cycling. Earth Sci. Rev. 110, 2657.Google Scholar
Oduro, H., et al. (2011). Evidence of magnetic isotope effects during thermochemical sulfate reduction. P. Natl. Acad. Sci. USA 108, 17 63517 638.Google Scholar
Ody, A., et al. (2015). Candidates source regions of martian meteorites as identified by OMEGA/MEx. Icarus 258, 366383.Google Scholar
Ohtomo, Y., et al. (2014). Evidence for biogenic graphite in early Archaean Isua metasedimentary rocks. Nat. Geosci. 7, 2528.Google Scholar
Ojakangas, R. W., et al. (2014). The Tayla conglomerate: an Archean (~2.7 Ga) glaciomarine formation, Western Dharwar Craton, Southern India. Curr. Sci. 106, 287396.Google Scholar
Ojha, L., et al. (2015). Spectral evidence for hydrated salts in recurring slope lineae on Mars. Nature Geosci., doi: 10.1038/ngeo2546.Google Scholar
Ojima, M. and Podosek, F. A. (2002). Noble gas geochemistry. New York: Cambridge University Press.Google Scholar
Olson, S. L., et al. (2013). Quantifying the areal extent and dissolved oxygen concentrations of Archean oxygen oases. Chem. Geol. 362, 3543.Google Scholar
Olson, S. L., et al. (2016). Limited role for methane in the mid-Proterozoic greenhouse. P. Natl Acad. Sci. USA 113, 11 44711 452.Google Scholar
Ono, S. (2001). Detrital uraninite and the early Earth’s atmosphere: SIMS analyses of uraninite in the Elliot Lake District and the dissolution kinetics of natural uraninite. Penn State University, PhD, College Station, PA.Google Scholar
Ono, S., et al. (2003). New insights into Archean sulfur cycle from mass-independent sulfur isotope records from the Hamersley Basin, Australia. Earth Planet. Sci. Lett. 213, 1530.Google Scholar
Ono, S., et al. (2007). S-33 constraints on the seawater sulfate contribution in modern seafloor hydrothermal vent sulfides. Geochim. Cosmochim. Acta 71, 11701182.Google Scholar
Ono, S., et al. (2013). Contribution of isotopologue self-shielding to sulfur mass-independent fractionation during sulfur dioxide photolysis. J. Geophys. Res. 118, 24442454.Google Scholar
Oparin, A. I. (1924). Proiskhozhdenie Zhizni (The Origin of Life). Moscow: Izd. Moskovskii Rabochii (in Russian).Google Scholar
Oparin, A. I. (1938). The Origin of Life. New York: MacMillan.Google Scholar
Oparin, A. I. (1957). The Origin of Life. New York: Academic Press.Google Scholar
Oparin, A. I. (1968). Genesis and Evolutionary Development of Life. New York: Academic Press.Google Scholar
Opik, E. J. (1963). Selective escape of gases. Geophys. J. R. Astron. Soc. 7, 490509.Google Scholar
Oppenheimer, C., et al. (2014). Volcanic degassing: Process and impact. In: Treatise on Geochemistry (2nd Edn), ed. Holland, H. D. and Turekian, K. K., New York: Elsevier.Google Scholar
Orgel, L. E. (1986). RNA catalysis and the origins of life. Journal of Theoretical Biology 123, 127149.Google Scholar
Oro, J. (1961). Comets and the formation of biochemical compounds on the primitive Earth. Nature 190, 389390.Google Scholar
Osborn, H. F. (1917). The Origin and Evolution of Life: On the Theory of Action, Reaction and Interaction of Energy. New York: Charles Scribner’s Sons.Google Scholar
Osterloo, M. M., et al. (2010). Geologic context of proposed chloride-bearing materials on Mars. J. Geophys. Res. 115.Google Scholar
Owen, J. E. and Jackson, A. P. (2012). Planetary evaporation by UV & X-ray radiation: basic hydrodynamics. Mon. Not. R. Astron. Soc. 425, 29312947.Google Scholar
Owen, J. E. and Wu, Y. Q. (2013). Kepler planets: A tale of evaporation. Astrophys. J. 775, 105, doi: 10.1088/0004-637X/775/2/105.Google Scholar
Owen, T. (1980). The search for early forms of life in other planetary systems: future possibilities afforded by spectroscopic techniques. In: Strategies for Search for Life in the Universe, ed. Papagiannis, M. D., Dordrecht: Reidel, pp. 177185.Google Scholar
Owen, T. (1992). The composition and early history of the atmosphere of Mars. In: Mars, ed. Kieffer, H. H., et al., Tucson: University of Arizona Press, pp. 818834.Google Scholar
Owen, T. and Bar-Nun, A. (1995). Comets, impacts, and atmospheres. Icarus 116, 215226.Google Scholar
Owen, T. and Bar-Nun, A. (1998). From the interstellar medium to planetary atmospheres via comets. Faraday Discuss. 109, 453462.Google Scholar
Owen, T., et al. (1992). Possible cometary origin of heavy noble gases in the atmospheres of Venus, Earth, and Mars. Nature 358, 4346.Google Scholar
Owen, T., et al. (1977). The composition of the atmosphere at the surface of Mars. J. Geophys. Res. 82, 46354639.Google Scholar
Owen, T., et al. (1999). Saturn VI (Titan). IAU Circ. 7306.Google Scholar
Owen, T., et al. (1979). Early Earth: An enhanced carbon dioxide greenhouse to compensate for reduced solar luminosity. Nature 277, 640642.Google Scholar
Owen, T. and Encrenaz, T. (2006). Compositional constraints on giant planet formation. Planet. Space Sci. 54, 11881196.Google Scholar
Owen, T., et al. (1988). Deuterium on Mars: The abundance of HDO and the value of D/H. Science 240, 1767–70.Google Scholar
Owen, T. and Niemann, H. B. (2009). The origin of Titan’s atmosphere: some recent advances. Phil. Trans. R. Soc. Lond. A 367, 607615.Google Scholar
Owen, T. C., et al. (1993). Surface ices and the atmospheric composition of Pluto. Science 261, 745748.Google Scholar
Oyama, V. I. and Berdahl, B. J. (1979). Model of Martian surface chemistry. J. Mol. Evol. 14, 199210.Google Scholar
Pahlevan, K. and Stevenson, D. J. (2007). Equilibration in the aftermath of the lunar-forming giant impact. Earth Planet. Sc. Lett. 262, 438449.Google Scholar
Palle, E., et al. (2008). Identifying the rotation rate and the presence of dynamic weather on extrasolar earth-like planets from photometric observations. Astrophys. J. 676, 13191329.Google Scholar
Palle, E., et al. (2003). Earthshine and the Earth’s albedo: 2. Observations and simulations over 3 years. J. Geophys. Res. 108, 4710, doi:10.1029/2003jd003611.Google Scholar
Palomba, E., et al. (2009). Evidence for Mg-rich carbonates on Mars from a 3.9 μm absorption feature. Icarus 203, 5865.Google Scholar
Papineau, D. (2010). Global biogeochemical changes at both ends of the Proterozoic: Insights from phosphorites. Astrobiology 10, 165181.Google Scholar
Papineau, D., et al. (2009). High primary productivity and nitrogen cycling after the Paleoproterozoic phosphogenic event in the Aravalli Supergroup, India. Precam. Res. 171, 3756.Google Scholar
Parker, E. N. (1963). Interplanetary Dynamical Processes. New York: Interscience Publishers.Google Scholar
Parker, E. T., et al. (2011). Primordial synthesis of amines and amino acids in a 1958 Miller H2S-rich spark discharge experiment. P. Natl. Acad. Sci. USA 108, 55265531.Google Scholar
Parker, T. J., et al. (1993). Coastal geomorphology of the martian Northern Plains. J. Geophys. Res. 98, 11 06111 078.Google Scholar
Parker, T. J., et al. (2010). The northern plains: A martian oceanic basin? In: Lakes on Mars, eds. Cabrol, N. A., Grin, E. A.. New York: Elsevier.Google Scholar
Parkinson, T. D. and Hunten, D. M. (1972). Spectroscopy and aeronomy of O2 on Mars. J. Atmos. Sci. 29, 13801390.Google Scholar
Parmentier, V. and Guillot, T. (2014). A non-grey analytical model for irradiated atmospheres. I. Derivation. Astron. Astrophys. 562, A133, doi:10.1051/0004-6361/201322342.Google Scholar
Parmentier, V., et al. (2015). A non-grey analytical model for irradiated atmospheres II. Analytical vs. numerical solutions. Astron. Astrophys. 574, A35, doi:10.1051/0004-6361/201323127.Google Scholar
Parnell, J., et al. (2010). Early oxygenation of the terrestrial environment during the Mesoproterozoic. Nature 468, 290293.Google Scholar
Partin, C. A., et al. (2013). Large-scale fluctuations in Precambrian atmospheric and oceanic oxygen levels from the record of U in shales. Earth Planet. Sci. Lett. 369, 284293.Google Scholar
Pasek, M. and Lauretta, D. (2008). Extraterrestrial flux of potentially prebiotic C, N, and P to the early Earth. Origins Life Evol. B. 38, 521.Google Scholar
Pavlov, A. A., et al. (2003). Methane-rich Proterozoic atmosphere? Geology 31, 8790.Google Scholar
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 2, 2741.Google Scholar
Pavlov, A. A., et al. (2001). UV-shielding of NH3 and O2 by organic hazes in the Archean atmosphere. J. Geophys. Res. 106, 23 26723 287.Google Scholar
Pavlov, A. A., et al. (2000). Greenhouse warming by CH4 in the atmosphere of early Earth. J. Geophys. Res. 105, 11,98111,990.Google Scholar
Pavlov, A. A., et al. (2012). Degradation of the organic molecules in the shallow subsurface of Mars due to irradiation by cosmic rays. Geophys. Res. Lett. 39.Google Scholar
Payne, J. L., et al. (2009). Two-phase increase in the maximum size of life over 3.5 billion years reflects biological innovation and environmental opportunity. P. Natl. Acad. Sci. USA 106, 2427.Google Scholar
Payne, J. L., et al. (2011). The evolutionary consequences of oxygenic photosynthesis: a body size perspective. Photosynth. Res. 107, 3757.Google Scholar
Peale, S. J. (1977). Rotational histories of the natural satellites. In: Planetary Satellites, ed. Burns, J. A., Tucson, AZ: University of Arizona Press.Google Scholar
Pearl, J. C. and Conrath, B. J. (1991). The albedo, effective temperature, and energy balance of Neptune, as determined from Voyager data. J. Geophys. Res. 96, 18 92118 930.Google Scholar
Pearl, J. C., et al. (1990). The albedo, effective temperature, and energy balance of Uranus, as determined from Voyager IRIS data. Icarus 84, 1228.Google Scholar
Pearson, D. G., et al. (2003). Mantle samples included in volcanic rocks: Xenoliths and diamonds. In: Treatise on Geochemistry - Volume 2: The Mantle and Core, ed. Carlson, R. W., Amsterdam: Elsevier, pp. 171276.Google Scholar
Pechmann, J. B. and Ingersoll, A. P. (1984). Thermal tides in the atmosphere of Venus: Comparison of model results with observations. J. Atmos. Sci. 41, 32903313.Google Scholar
Peck, W. H., et al. (2001). Oxygen isotope ratios and rare earth elements in 3.3 to 4.4 Ga zircons: Ion microprobe evidence for high delta O-18 continental crust and oceans in the Early Archean. Geochim. Cosmochim. Acta 65, 42154229.Google Scholar
Pecoits, E., et al. (2015). Atmospheric hydrogen peroxide and Eoarchean iron formations. Geobiology 13, 114.Google Scholar
Pedone, M., et al. (2014). Tunable diode laser measurements of hydrothermal/volcanic CO2 and implications for the global CO2 budget. Solid Earth 5, 12091221.Google Scholar
Peixoto, J. P. and Oort, A. H. (1984). Physics of climate. Rev. Mod. Phys. 56, 365429. © 1984 by The American Physical Society.Google Scholar
Peltier, W. R., et al. (2007). Snowball Earth prevention by dissolved organic carbon remineralization. Nature 450, 813–U1.Google Scholar
Pepe, F., et al. (2014). ESPRESSO: The next European exoplanet hunter. Astronomische Nachrichten 335, 820.Google Scholar
Pepin, R. O. (1989). Atmospheric compositions: Key similarities and differences. In: Origin and Evolution of Planetary and Satellite Atmospheres, ed. Atreya, S. K., et al., Tucson, AZ: University of Arizona Press, pp. 291305.Google Scholar
Pepin, R. O. (1991). On the origin and evolution of terrestrial planet atmospheres and meteoritic volatiles. Icarus 92, 279.Google Scholar
Pepin, R. O. (2000). On the isotopic composition of primordial xenon in terrestrial planet atmospheres. Space Sci. Rev. 92, 371395.Google Scholar
Pepin, R. O. (2006). Atmospheres on the terrestrial planets: Clues to origin and evolution. Earth Planet. Sci. Lett. 252, 114.Google Scholar
Pepin, R. O. (2013). Comment on “Chondritic-like xenon trapped in Archean rocks: A possible signature of the ancient atmosphere” by M. Pujol, B. Marty, R. Burgess [Earth Planet. Sci. Lett. 308 (2011) 298–306]. Earth Planet. Sci. Lett. 371, 294295.Google Scholar
Pepin, R. O. and Porcelli, D. (2006). Xenon isotope systematics, giant impacts, and mantle degassing on the early Earth. Earth Planet. Sci. Lett. 250, 470485.Google Scholar
Pepin, R. O., et al. (2012). Helium, neon, and argon composition of the solar wind as recorded in gold and other Genesis collector materials. Geochim. Cosmochim. Acta 89, 6280.Google Scholar
Peralta, J., et al. (2007). A reanalysis of Venus winds at two cloud levels from Galileo SSI images. Icarus 190, 469477.Google Scholar
Peralta, J., et al. (2012). Solar migrating atmospheric tides in the winds of the polar region of Venus. Icarus 220, 958970.Google Scholar
Perez-de-Tejada, H. (1987). Plasma flow in the Mars magnetosphere. J. Geophys. Res. 92, 47134718.Google Scholar
Perkins, J. P., et al. (2015). Amplification of bedrock canyon incision by wind. Nat. Geosci. 8, 305310.Google Scholar
Pernice, H., et al. (2004). Laboratory evidence for a key intermediate in the Venus atmosphere: Peroxychloroformyl radical. P. Natl. Acad. Sci. USA 101, 14 00714 010.Google Scholar
Perrier, S., et al. (2006). Global distribution of total ozone on Mars from SPICAM/MEX UV measurements. J. Geophys. Res. 111, E09S06.Google Scholar
Perry, E. C., et al. (1978). The oxygen isotope composition of 3800 m.y. old metamorphosed chert and iron formation from Isukasia, West Greenland. J. Geol. 86, 223239.Google Scholar
Perryman, M. A. C. (2014). The Exoplanet Handbook. New York: Cambridge University Press.Google Scholar
Pestova, O. N., et al. (2005). Polythermal study of the systems M(ClO4)(2)-H2O(M2+ = Mg2+, Ca2+, Sr2+, Ba2+). Russian J. Appl. Chem. 78, 409413.Google Scholar
Peters, K. E., et al. (1978). Correlation of carbon and nitrogen stable isotope ratios in sedimentary organic matter. Limnol. Oceanogr. 23, 598604.Google Scholar
Peterson, C. (1981). A secondary origin for the central plateau of Hebes Chasma. Proc. Lunar Planet. Sci. Conf. 11th, 1459–1471.Google Scholar
Peterson, K. J., et al. (2008). The Ediacaran emergence of bilaterians: congruence between the genetic and the geological fossil records. Phil. Trans. R. Soc. Lond. B 363, 14351443.Google Scholar
Petigura, E. A., et al. (2013). Prevalence of Earth-size planets orbiting Sun-like stars. P. Natl. Acad. Sci. USA 110, 1927319278.Google Scholar
Petty, G. W. (2006). A First Course in Atmospheric Radiation. Madison, Wis.: Sundog Pub.Google Scholar
Pfalzner, S., et al. (2015). The formation of the solar system. Phy. Scr. 90, 068001 doi:10.1088/0031-8949/90/6/068001.Google Scholar
Pham, L. B. S., et al. (2011). Effects of impacts on the atmospheric evolution: Comparison between Mars, Earth, and Venus. Planet. Space Sci. 59, 10871092.Google Scholar
Phillips, R. J., et al. (2011). Massive CO(2) ice deposits sequestered in the South Polar layered deposits of Mars. Science 332, 838841.Google Scholar
Phillips, R. J., et al. (1992). Impact craters and Venus resurfacing history. J. Geophys. Res. 97, 15 923.Google Scholar
Phillips, R. J., et al. (2001). Ancient geodynamics and global-scale hydrology on Mars. Science 291, 25872591.Google Scholar
Pierazzo, E., et al. (2008). Validation of numerical codes for impact and explosion cratering: Impacts on strengthless and metal targets. Meteorit. Planet. Sci. 43, 19171938.Google Scholar
Pierazzo, E. and Chyba, C. F. (1999). Amino acid survival in large cometary impacts. Meteorit. Planet. Sci. 34, 909918.Google Scholar
Pieri, D. C. (1980). Geomorphology of valleys on Mars: A summary of morphology, distribution, age and origin. Science 210, 895897.Google Scholar
Pierrard, V. (2003). Evaporation of hydrogen and helium atoms from the atmospheres of Earth and Mars. Planet. Space Sci. 51, 319327.Google Scholar
Pierrehumbert, R. and Gaidos, E. (2011). Hydrogen greenhouse planets beyond the habitable zone. Astrophys. J. Lett. 734.Google Scholar
Pierrehumbert, R. T. (2004). High levels of atmospheric carbon dioxide necessary for the termination of global glaciation. Nature 429, 646649.Google Scholar
Pierrehumbert, R. T. (2010). Principles of Planetary Climate. Cambridge: Cambridge University Press.Google Scholar
Pierrehumbert, R. T. (2011). A pallette of climates for Gliese 581g. Astrophys. J. Lett. 726.Google Scholar
Pierrehumbert, R. T., et al. (2011). Climate of the Neoproterozoic. Annu. Rev. Earth Planet. Sci. 39, 417460.Google Scholar
Pierrehumbert, R. T. and Swanson, K. L. (1995). Baroclinic instability. Annu. Rev. Fluid Mech. 27, 419467.Google Scholar
Pinti, D. L., et al. (2001). Nitrogen and argon signatures in 3.8 To 2.8 Ga metasediments: Clues on the chemical state of the Archean ocean and the deep biosphere. Geochim. Cosmochim. Acta 65, 23012315.Google Scholar
Pinti, D. L., et al. (2007). Biogenic nitrogen and carbon in Fe-Mn-oxyhydroxides from an Archean chert, Marble Bar, Western Australia. Geochem. Geophys. Geosys. 8, doi:10.1029/2006GC001394.Google Scholar
Pinti, D. L., et al. (2009). Isotopic fractionation of nitrogen and carbon in Paleoarchean cherts from Pilbara craton, Western Australia: Origin of N-15-depleted nitrogen. Geochim. Cosmochim. Acta 73, 38193848.Google Scholar
Pinti, D. L., et al. (2013). Comment on “Biogenicity of Earth’s earliest fossils: a resolution of the controversy” by J. William Schopf and Anatoliy B. Kudryavtsev, Gondwana Research 22 (2012), 761–771. Gondwana Res. 23, 16521653.Google Scholar
Pinto, J. P., et al. (1980). Photochemical production of formaldehyde in the earth’s primitive atmosphere. Science 210, 183185.Google Scholar
Planavsky, N., et al. (2010). Rare Earth Element and yttrium compositions of Archean and Paleoproterozoic Fe formations revisited: New perspectives on the significance and mechanisms of deposition. Geochim. Cosmochim. Acta 74, 63876405.Google Scholar
Planavsky, N. J., et al. (2014a). Evidence for oxygenic photosynthesis half a billiion years before the Great Oxidation Event. Nat. Geosci. 7, 283286.Google Scholar
Planavsky, N. J., et al. (2012). Sulfur record of rising and falling marine oxygen and sulfate levels during the Lomagundi event. P. Natl. Acad. Sci. USA 109, 18 30018 305.Google Scholar
Planavsky, N. J., et al. (2011). Widespread iron-rich conditions in the mid-Proterozoic ocean. Nature 477, 448451.Google Scholar
Planavsky, N. J., et al. (2014b). Low Mid-Proterozoic atmospheric oxygen levels and the delayed rise of animals. Science 346, 635638.Google Scholar
Plankensteiner, K., et al. (2006). Amino acids on the rampant primordial Earth: Electric discharges and the hot salty ocean. Mol. Divers. 10, 37.Google Scholar
Plaut, J. J., et al. (2007). Subsurface radar sounding of the south polar layered deposits of Mars. Science 316, 9295.Google Scholar
Pleskot, L. K. and Miner, E. D. (1981). Time variability of martian bolometric albedo. Icarus 45, 179201.Google Scholar
Pogge von Strandmann, P. A. E., et al. (2015). Selenium isotope evidence for post-glacial oxygenation trends in the Ediacaran ocean. Nat. Commun., submitted.Google Scholar
Pohl, W. (1989). Comparative geology of magnesite and ocurrences. Monograph Series on Mineral Deposits 28, 113.Google Scholar
Poirier, J. P. (1994). Light elements in the Earths outer core: a critical review. Phys. Earth Planet. In. 85, 319337.Google Scholar
Pollack, J. B. (1971). A nongrey calculation of the runaway greenhouse: implications for Venus’ past and present. Icarus 14, 295306.Google Scholar
Pollack, J. B. (1979). Climatic change on the terrestrial planets. Icarus 37, 479553.Google Scholar
Pollack, J. B. (1991). Kuiper Prize Lecture: Present and past climates of the terrestrial planets. Icarus 91, 173198.Google Scholar
Pollack, J. B., et al. (1979). Properties and effects of dust particles suspended in the Martian atmosphere. J. Geophys. Res. 84, 29292945.Google Scholar
Pollack, J. B., et al. (1993). Near-infrared light from Venus nightside: A spectroscopic analysis. Icarus 103, 142.Google Scholar
Pollack, J. B., et al. (1977). Calculation of Saturn’s gravitational contraction history. Icarus 30,111128.Google Scholar
Pollack, J. B., et al. (1990). Simulations of the general circulation of the Martian atmosphere 1. Polar Processes. J. Geophys. Res. 95, 14471473.Google Scholar
Pollack, J. B., et al. (1996). Formation of the giant planets by concurrent accretion of solids and gas. Icarus 124, 6285.Google Scholar
Pollack, J. B., et al. (1987). The case for a wet, warm climate on early Mars. Icarus 71, 203224.Google Scholar
Pollard, D. and Kasting, J. F. (2005). Snowball Earth: A thin-ice model with flowing sea glaciers. J. Geophys. Res. 110, C07010, doi: 10.1029/2004JC002525.Google Scholar
Pollard, D. and Kasting, J. F. (2006). Reply to comment by Stephen G. Warren and Richard E. Brandt on “Snowball earth: A thin-ice solution with flowing sea glaciers”. J. Geophys. Res. 111, C09017, doi: 10.1029/2006JC003488.Google Scholar
Polyak, V., et al. (2008). Age and evolution of the Grand Canyon revealed by U-Pb dating of water table-type speleothems. Science 319, 13771380.Google Scholar
Pondrelli, M., et al. (2011). Geological, geomorphological, facies and allostratigraphic maps of the Eberswalde fan delta. Planet. Space Sci. 59, 11661178.Google Scholar
Pope, E. C., et al. (2012). Isotope composition and volume of Earth’s early oceans. P. Natl. Acad. Sci. USA 109, 43714376.Google Scholar
Popp, M., et al. (2016). Transition to a moist greenhouse with CO2 and solar forcing. Nat. Commun. 7, 10627, doi:10.1038/ncomms10627.Google Scholar
Porcelli, D. and Elliott, T. (2008). The evolution of He isotopes in the convecting mantle and the preservation of high He-3/He-4 ratios. Earth Planet. Sci. Lett. 269, 175185.Google Scholar
Porcelli, D. and Wasserburg, G. J. (1995). Mass transfer of helium, neon, argon, and xenon through a steady-state upper mantle. Geochim. Cosmochim. Acta 59, 49214937.Google Scholar
Porco, C. C., et al. (2006). Cassini observes the active South Pole of Enceladus. Science 311, 13931401.Google Scholar
Porco, C. C., et al. (2003). Cassini imaging of Jupiter’s atmosphere, satellites, and rings. Science 299, 15411547.Google Scholar
Postawko, S. E. and Kuhn, W. R. (1986). Effect of the greenhouse gases (CO2, H2O, SO2) on Martian paleoclimate. J. Geophys. Res. (Proc. Lunar Planet. Sci. Conf. 16th) 91, D431D438.Google Scholar
Postberg, F., et al. (2011). A salt-water reservoir as the source of a compositionally stratified plume on Enceladus. Nature 474, 620622.Google Scholar
Postgate, J. R. (1987). Nitrogen Fixation. London: Edward Arnold.Google Scholar
Posth, N. R., et al. (2014). Biogenic Fe(III) minerals: From formation to diagenesis and preservation in the rock record. Earth Sci. Rev. 135, 103121.Google Scholar
Posth, N. R., et al. (2013). Simulating Precambrian banded iron formation diagenesis. Chem. Geol. 362, 6673.Google Scholar
Poulton, S. W. and Canfield, D. E. (2011). Ferruginous conditions: A dominant feature of the ocean through Earth’s history. Elements 7, 107112.Google Scholar
Poulton, S. W., et al. (2004). The transition to a sulphidic ocean ~1.84 billion years ago. Nature 431, 173177.Google Scholar
Poulton, S. W., et al. (2010). Spatial variability in oceanic redox structure 1.8 billion years ago. Nature Geosci. 3, 486490.Google Scholar
Poulton, S. W. and Raiswell, R. (2002). The low-temperature geochemical cycle of iron: From continental fluxes to marine sediment deposition. Am. J. Sci. 302, 774805.Google Scholar
Prasad, N. and Roscoe, S. M. (1996). Evidence of anoxic to oxic atmosphere change during 2.45–2.22 Ga from lower and upper sub-Huronian paleosols, Canada. Catena 27, 105121.Google Scholar
Prather, M. (1996). Time scales in atmospheric chemistry: Theory, GWPs for CH4 and CO, and runaway growth. Geophys. Res. Lett. 23, 25972600.Google Scholar
Prather, M., et al. (2001). Atmospheric chemistry and greenhouse gases. In: Climate Change 2001, The Scientific Basis, ed. Houghton, J. T., et al., Cambridge; New York: Cambridge University Press, pp. 239287.Google Scholar
Prausnitz, J. M., et al. (1999). Molecular Thermodynamics Of Fluid-Phase Equilibria. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Prave, A. R. (2002). Life on land in the Proterozoic: Evidence from the Torridonian rocks of northwest Scotland. Geology 30, 811814.Google Scholar
Prentice, I. C., et al. (2001). The carbon cycle and atmospheric carbon dioxide. In: Climate Change 2001: the Scientific Basis ed. Houghton, J. T., et al., Cambridge; New York: Cambridge University Press, pp. 183238.Google Scholar
Press, W. H. (2007). Numerical Recipes: The Art Of Scientific Computing. New York: Cambridge University Press.Google Scholar
Preston, G. T. and Prausnit, J. M. (1970). Thermodynamics of solid solubility in cryogenic solvents. Ind. Eng. Chem .Proc. Dd. 9, 264271.Google Scholar
Prinn, R. G. (1971). Photochemistry of HCl and other minor constituents in atmosphere of Venus. J. Atmos. Sci. 28, 10581068.Google Scholar
Prinn, R. G. (2014). Ozone, hydroxyl radical, and oxidative capacity. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier.Google Scholar
Prinn, R. G. and Fegley, B. (1987). The atmospheres of Venus, Earth, and Mars: A critical comparison. Annu. Rev. Earth Planet. Sci. 15, 171212.Google Scholar
Prinn, R. G. and Fegley, B. (1989). Solar nebula chemistry: origin of planetary, satellite and cometary volatiles. In: Origin and Evolution of Planetary and Satellite Atmospheres, ed. Atreya, S. K., et al., Tucson, AZ: University of Arizona Press, pp. 78136.Google Scholar
Prinn, R. G., et al. (1995). Atmospheric trends and ilfetime of CH3CCl3 and global OH concentrations. Science 269, 187192.Google Scholar
Proskurowski, G., et al. (2008). Abiogenic hydrocarbon production at Lost City hydrothermal field. Science 319, 604607.Google Scholar
Pujol, M., et al. (2011). Chondritic-like xenon trapped in Archean rocks: A possible signature of the ancient atmosphere. Earth Planet. Sc. Lett. 308, 298306.Google Scholar
Pujol, M., et al. (2013). Argon isotopic composition of Archaean atmosphere probes early Earth geodynamics. Nature 498, 8790.Google Scholar
Putzig, N. E., et al. (2014). SHARAD soundings and surface roughness at past, present, and proposed landing sites on Mars: Reflections at Phoenix may be attributable to deep ground ice. J. Geophys. Res. 119, 19361949.Google Scholar
Quinn, R. C., et al. (2013). Perchlorate radiolysis on Mars and the origin of martian soil reactivity. Astrobiology 13, 515520.Google Scholar
Raack, J., et al. (2015). Present-day seasonal gully activity in a south polar pit (Sisyphi Cavi) on Mars. Icarus 251, 226243.Google Scholar
Radebaugh, J., et al. (2008). Dunes on Titan observed by Cassini Radar. Icarus 194, 690703.Google Scholar
Raff, R. A. and Raff, E. C. (1970). Respiratory mechanisms and the metazoan fossil record. Nature 228, 10031005.Google Scholar
Raghoebarsing, A. A., et al. (2006). A microbial consortium couples anaerobic methane oxidation to denitrification. Nature 440, 918921.Google Scholar
Rairden, R. L., et al. (1986). Geocoronal imaging with Dynamics Explorer. J. Geophys. Res. 91, 36133630.Google Scholar
Ramanathan, V. and Carmichael, G. (2008). Global and regional climate changes due to black carbon. Nat. Geosci. 1, 221227.Google Scholar
Ramanathan, V. and Inamdar, A. (2006). The radiative forcing due to clouds and water vapor. In: Frontiers of Climate Modeling, ed. Kiehl, J. T., Ramanathan, V., New York: Cambridge University Press, pp. 119151.Google Scholar
Ramaswamy, V., et al. (2001). Radiative forcing of climate change. In: Climate Change 2001: Working Group I: The Scientific Basis, ed. Houghton, J. T., et al., New York: Cambridge University Press, pp. 349416.Google Scholar
Rambler, M. and Margulis, L. (1980). Bacterial resistance to ultraviolet irradiation under anaerobiosis: Implications for pre-Phanerozoic evolution. Science 210, 638640.Google Scholar
Rameau, J., et al. (2013). Discovery of a probable 4–5 Jupiter-mass exoplanet to HD 95086 by direct imaging. Astrophys. J. Lett. 772, L15.Google Scholar
Ramirez, R. M. and Kaltenegger, L. (2014). The habitable zones of pre-main-sequence stars. Astrophys. J. Lett. 797.Google Scholar
Ramirez, R. M., et al. (2014a). Warming early Mars with CO2 and H2. Nat. Geosci. 7, 5963.Google Scholar
Ramirez, R. M., et al. (2014b). Can increased atmospheric CO2 levels trigger a runaway greenhouse? Astrobiology 14, 714731.Google Scholar
Randel, W. J. and Held, I. M. (1991). Phase speed spectra of transient eddy fluxes and critical layer absorption. J. Atmos. Sci. 48, 688697.Google Scholar
Rannou, P., et al. (1997). A new interpretation of scattered light measurements at Titan’s limb. J. Geophys. Res. 102, 10 99711 013.Google Scholar
Rannou, P., et al. (1995). Titan’s geometric albedo: Role of the fractal structure of the aerosols. Icarus 118, 355372.Google Scholar
Rannou, P., et al. (2004). A coupled dynamics-microphysics model of Titan’s atmosphere. Icarus 170, 443462.Google Scholar
Rannou, P., et al. (2003). A model of Titan’s haze of fractal aerosols constrained by multiple observations. Planet. Space Sci. 51, 963976.Google Scholar
Rannou, P., et al. (2006). The latitudinal distribution of clouds on Titan. Science 311, 201205.Google Scholar
Rashby, S. E., et al. (2007). Biosynthesis of 2-methylbacteriohopanepolyols by an anoxygenic phototroph. P. Natl. Acad. Sci. USA 104, 15 09915 104.Google Scholar
Rasmussen, B. and Buick, R. (1999). Redox state of the Archean atmosphere: Evidence from detrital heavy minerals in ca. 3250–2750 Ma sandstones from the Pilbara Craton, Australia. Geology 27, 115118.Google Scholar
Rasmussen, B., et al. (2008). Reassessing the first appearance of eukaryotes and cyanobacteria. Nature 455, 11011104.Google Scholar
Rasmussen, B., et al. (2014). Hematite replacement of iron-bearing precursor sediments in the 3.46-b.y.-old Marble Bar Chert, Pilbara craton, Australia. Bull. Geol. Soc. Am., doi:10.1130/B31049.1.Google Scholar
Rasmussen, B., et al. (2016). Dust to dust: Evidence for the formation of “primary” hematite dust in banded iron formations via oxidation of iron silicate nanoparticles. Precam. Res. 284, 4963.Google Scholar
Rasool, S. I. and DeBergh, C. (1970). The runaway greenhouse and the accumulation of CO2 in the Venus atmosphere. Nature 226, 10371039.Google Scholar
Raub, T. D., et al. (2007). Siliciclastic prelude to Elatina–Nuccaleena deglaciation: Lithostratigraphy and rock magnetism of the base of the Ediacaran system. Spec. Publ. Geol. Soc. London 286, 5375.Google Scholar
Raulin, F., et al. (2012). Prebiotic-like chemistry on Titan. Chem. Soc. Rev. 41, 53805393.Google Scholar
Raymann, K., et al. (2015). The two-domain tree of life is linked to a new root for the Archaea. P. Natl Acad. Sci. USA 112, 66706675.Google Scholar
Raymo, M. E. and Ruddiman, W. F. (1992). Tectonic forcing of Late Cenozoic climate. Nature 359, 117122.Google Scholar
Raymond, J., et al. (2004). The natural history of nitrogen fixation. Molec. Biol. Evol. 21, 541554.Google Scholar
Raymond, S. N. (2014). Terrestrial planet formation at home and abroad. In: Protostars and Planets VI, ed. Beuther, H., et al., Tucson, AZ: University of Arizona Press, pp. 595618.Google Scholar
Raymond, S. N., et al. (2006). High-resolution simulations of the final assembly of Earth-like planets I: Terrestrial accretion and dynamics. Icarus 183, 265282.Google Scholar
Raymond, S. N., et al. (2010). Formation of Terrestrial Planets. In: Formation and Evolution of Exoplanets, ed. Barnes, Rony, Wiley - VCH Verlag GmbH & Co. KGaA, pp. 123143.Google Scholar
Read, P. L. (2013). The dynamics and circulation of Venus atmosphere. In: Towards Understanding the Climate of Venus, ed. Bengtsson, L., et al., New York: Springer, pp. 73110.Google Scholar
Read, P. L. and Lewis, S. R. (2003). The Martian Climate Revisited: Atmosphere and Environment of a Desert Planet. London: Praxis.Google Scholar
Redfield, A. C. (1958). The biological control of chemical factors in the environment. Am. Sci. 46, 205221.Google Scholar
Reese, C. C., et al. (2010). Impact origin for the Martian crustal dichotomy: Half emptied or half filled? J. Geophys. Res. 115.Google Scholar
Reinhard, C. T. and Planavsky, N. J. (2011). Mineralogical constraints on Precambiran pCO2. Nature 474, E1, doi:10.1038/nature09959.Google Scholar
Reinhard, C. T., et al. (2013a). Long-term sedimentary recycling of rare sulphur isotope anomalies. Nature 497, 100104.Google Scholar
Reinhard, C. T., et al. (2013b). Proterozoic ocean redox and biogeochemical stasis. P. Natl. Acad. Sci. USA 110, 53575362.Google Scholar
Reinhard, C. T., et al. (2009). A late Archean sulfidic sea stimulated by early oxidative weathering of the continents. Science 326, 713716.Google Scholar
Reuschel, M., et al. (2012). Isotopic evidence for a sizeable seawater sulfate reservoir at 2.1 Ga. Precambrian Res. 19295, 7888.Google Scholar
Rhines, P. B. (1975). Waves and turbulence on a beta plane. J. Fluid Mech. 69, 417443.Google Scholar
Ribas, I., et al. (2005). Evolution of the solar activity over time and effects on planetary atmospheres. I. High-energy irradiances (1–1700 ångström). Astrophys. J. 622, 680694.Google Scholar
Ricardo, A., et al. (2004). Borate minerals stabilize ribose. Science 303, 196196.Google Scholar
Richardson, M. I. and Wilson, R. J. (2002). A topographically forced asymmetry in the martian circulation and climate. Nature 416, 298301.Google Scholar
Ridgwell, A. J., et al. (2003). Carbonate deposition, climate stability, and neoproterozoic ice ages. Science 302, 859862.Google Scholar
Riding, R., et al. (2014). Idenfitication of an Archean marine oxygen oasis. Precam. Res. 251, 232237.Google Scholar
Riedel, K. and Lassey, K. (2008). Detergent of the atmosphere. Water Atmos. 16, 2223.Google Scholar
Roach, L. H., et al. (2010a). Diagenetic haematite and sulfate assemblages in Valles Marineris. Icarus 207, 659674.Google Scholar
Roach, L. H., et al. (2010b). Hydrated mineral stratigraphy of Ius Chasma, Valles Marineris. Icarus 206, 253268.Google Scholar
Robbins, S. J. and Hynek, B. M. (2012). A new global database of Mars impact craters >= 1 km: 1. Database creation, properties, and parameters. J. Geophys. Res. 117.Google Scholar
Roberson, A. L., et al. (2011). Greenhouse warming by nitrous oxide and methane in the Proterozoic Eon. Geobiology 9, 313320.Google Scholar
Robert, F. (2001). The origin of water on Earth. Science 293, 10561058.Google Scholar
Robert, F. and Chaussidon, M. (2006). A palaeotemperature curve for the Precambrian oceans based on silicon isotopes in cherts. Nature 443, 969972.Google Scholar
Robertson, M. P. and Joyce, G. F. (2012). The origins of the RNA World. CSH Perspect. Biol. 4.Google Scholar
Robin, C. M. I. and Bailey, R. C. (2009). Simultaneous generation of Archean crust and subcratonic roots by vertical tectonics. Geology 37, 523526.Google Scholar
Robinson, J. M. (1990). Lignin, land plants, and fungi – Biological evolution affecting Phanerozoic oxygen balance. Geology 18, 607610.Google Scholar
Robinson, T. D. and Catling, D. C. (2012). An analytic radiative-convective model for planetary atmospheres. Astrophys. J. 757, 104 doi:10.1088/0004-637X/757/1/104.Google Scholar
Robinson, T. D. and Catling, D. C. (2014). Common 0.1 bar tropopause in thick atmospheres set by pressure-dependent infrared transparency. Nat. Geosci. 7, 1215.Google Scholar
Robinson, T. D., et al. (2014a). Detection of ocean glint and ozone absorption using LCROSS Earth observations. Astrophys. J. 787, 171.Google Scholar
Robinson, T. D., et al. (2014b). Titan solar occultation observations reveal transit spectra of a hazy world. P. Natl. Acad. Sci. USA 111, 90429047.Google Scholar
Robinson, T. D., et al. (2010). Detecting oceans on extrasolar planets using the glint effect. Astrophys. J. Lett. 721, L67L71.Google Scholar
Robinson, T. D., et al. (2011). Earth as an extrasolar planet: Earth model validation using EPOXI Earth observations. Astrobiology 11, 393408.Google Scholar
Robock, A. (2000). Volcanic eruptions and climate. Rev. Geophys. 38, 191219.Google Scholar
Rodehacke, C. B., et al. (2013). An open ocean region in Neoproterozoic glaciations would have to be narrow to allow equatorial icesheets. Geophys. Res. Lett. 40, 55035507.Google Scholar
Rodgers, C. D. and Walshaw, C. D. (1966). Computation of infra-red cooling rate in planetary atmospheres. Q. J. R. Met. Soc. 92, 6792.Google Scholar
Rodler, F. and Lopez-Morales, M. (2014). Feasibility studies for the detection of O2 in an Earth-like exoplanet. Astrophys. J. 781, 54.Google Scholar
Rodriguez, S., et al. (2011). Titan’s cloud seasonal activity from winter to spring with Cassini/VIMS. Icarus 216, 89110.Google Scholar
Roe, H. G. and Grundy, W. M. (2012). Buoyancy of ice in the CH4–N2 system. Icarus 219, 733736.Google Scholar
Rogers, L. A. (2015). Most 1.6-Earth-radii planets are not rocky. Ap. J. 801, 41.Google Scholar
Rondanelli, R. and Lindzen, R. S. (2010). Can thin cirrus clouds in the tropics provide a solution to the faint young Sun paradox? J. Geophys. Res. 115.Google Scholar
Rooney, A. D., et al. (2014). Re–Os geochronology and coupled Os–Sr isotope constraints on the Sturtian snowball Earth. P. Natl. Acad. Sci. USA 111, 5156.Google Scholar
Rooney, A. D., et al. (2015). A Cryogenian chronology: Two long-lasting synchronous Neoproterozoic glaciations. Geology 43, 459462.Google Scholar
Roscoe, S. M. (1969). Huronian rocks and uraniferous conglomerates in the Canadian Shield. Geol. Surv. Can. Pap. 68–40, 205 pp.Google Scholar
Rose, B. E. J. and Marshall, J. (2009). Ocean heat transport, sea ice, and multiple climate states: Insights from energy balance models. J. Atmos. Sci. 66, 28282843.Google Scholar
Rosing, M. T. (1999). 13C-depleted carbon microparticles in >3700-Ma sea-floor sedimentary rocks from West Greenland. Science 283, 674676.Google Scholar
Rosing, M. T. and Bird, D. K. (2007). Constraints on atmospheric H-2 from banded iron formations. Geochim. Cosmochim. Acta 71, A852.Google Scholar
Rosing, M. T., et al. (2010). No climate paradox under the faint early Sun. Nature 464, 744-U117.Google Scholar
Rosing, M. T. and Frei, R. (2004). U-Rich Archaean seafloor sediments from Greenland – Indications of > 3700 Ma oxygenic photosynthesis. Earth Planet. Sci. Lett. 217, 237244.Google Scholar
Roskar, R., et al. (2008). Riding the spiral waves: Implications of stellar migration for the properties of galactic disks. Astrophys. J. Lett. 684, L79L82.Google Scholar
Rossow, W. B., et al. (1982). Cloud feedback: A stabilizing effect for the early Earth? Science 217, 12451247.Google Scholar
Roth, L., et al. (2014). Transient water vapor at Europa’s south pole. Science 343, 171174.Google Scholar
Rothman, D. H., et al. (2003). Dynamics of the Neoproterozoic carbon cycle. P. Natl. Acad. Sci. USA 100, 81248129.Google Scholar
Rouxel, O. J., et al. (2005). Iron isotope constraints on the Archean and Paleoproterozoic ocean redox state. Science 307, 10881091.Google Scholar
Royer, D. L., et al. (2007). Climate sensitivity constrained by CO2 concentrations over the past 420 million years. Nature 446, 530532.Google Scholar
Rubey, W. W. (1951). Geological history of seawater. An attempt to state the problem. Geol. Soc. Am. Bull. 62, 11111148.Google Scholar
Rubey, W. W. (1955). Development of the hydrosphere and atmosphere, with special reference to probable composition of the early atmosphere. In: Crust of the Earth, ed. Poldervaart, A., New York: Geol. Soc. Am., pp. 631650.Google Scholar
Rubie, D. C., et al. (2011). Heterogeneous accretion, composition and core-mantle differentiation of the Earth. Earth Planet. Sc. Lett. 301, 3142.Google Scholar
Rubie, D. C. and Jacobson, S. A. (2015). Mechanisms and geochemical models of core formation. In: Deep Earth: Physics and Chemistry of the Lower Mantle and Core. AGU Monograph, ed. Fischer, R. and Terasaki, H., Washington DC: Amer. Geophys. Union.Google Scholar
Rubie, D. C., et al. (2015a). Accretion and differentiation of the terrestrial planets with implications for the compositions of early-formed Solar System bodies and accretion of water. Icarus 248, 89108.Google Scholar
Rubie, D. C., et al. (2015b). Formation of the Earth’s core. In: Treatise on Geophysics, ed. Schubert, G., New York: Elsevier, pp. 4379.Google Scholar
Rubinstein, C. V., et al. (2010). Early Middle Ordovician evidence for land plants in Argentina (eastern Gondwana). New Phytol. 188, 365369.Google Scholar
Ruff, S. W., et al. (2014). Evidence for a Noachian-aged ephemeral lake in Gusev crater, Mars. Geology 42, 359362.Google Scholar
Ruiz, J., et al. (2011). The thermal evolution of Mars as constrained by paleo-heat flows. Icarus 215, 508517.Google Scholar
Ruiz-Mirazo, K., et al. (2014). Prebiotic systems chemistry: new perspectives for the origins of life. Chem. Rev. 114, 285366.Google Scholar
Runnegar, B. (1982). Oxygen requirements, biology and phylogenetic significance of the late Precambrian worm Dickinsonia, and the evolution of the burrowing habit. Alcheringa 6, 223239.Google Scholar
Runnegar, B. (1991). Oxygen and the early evolution of the Metazoa. In: Metazoan Life without Oxygen, ed. Bryant, C., New York: Chapman and Hall, pp. 6587.Google Scholar
Runnegar, B., et al. (2002). Mass-independent and mass-dependent sulfur processing throughout the Archean. Geochim. Cosmochim. Acta 66, A655–A655.Google Scholar
Russell, G. L., et al. (2013). Fast atmosphere–ocean model runs with large changes in CO2. Geophys. Res. Lett. 40, 57875792.Google Scholar
Russell, H. N. (1916). On the albedo of the Moon and planets. Ap. J. 43.Google Scholar
Russell, J. M., et al. (1996). Satellite confirmation of the dominance of chlorofluorocarbons in the global stratospheric chlorine budget. Nature 379, 526529.Google Scholar
Russell, M. J. (1996). The generation at hot springs of sedimentary ore deposits, microbialites and life. Ore Geology Reviews 10, 199214.Google Scholar
Russell, S. S., et al. (2006). Timescales of the protoplanetary disk. In: Meteorites and the Early Solar System II, ed. Lauretta, D. S. and McSween, H. Y., Tuscon, AZ: University of Arizona Press, pp. 233251.Google Scholar
Ryan, S., et al. (2006). Mauna Loa volcano is not a methane source: Implications for Mars. Geophysical Research Letters 33.Google Scholar
Ryder, G., et al. (2000). Heavy bombardment on the Earth at ~3.85 Ga: The search for petrographic and geochemical evidence. In: Origin of the Moon and Earth, ed. Canup, R. M. and Righter, K., Tucson, AZ: University of Arizona Press, pp. 475492.Google Scholar
Rye, R. and Holland, H. D. (1998). Paleosols and the evolution of atmospheric oxygen: A critical review. Amer. J. Sci. 298, 621672.Google Scholar
Rye, R. and Holland, H. D. (2000). Life associated with a 2.76 Ga ephemeral pond?: Evidence from Mount Roe #2 paleosol. Geology 28, 483486.Google Scholar
Rye, R., et al. (1995). Atmospheric carbon dioxide concentrations before 2.2 billion years ago. Nature 378, 603605.Google Scholar
Sackmann, I. J. and Boothroyd, A. I. (2003). Our Sun. V. A bright young Sun consistent with helioseismology and warm temperatures on ancient Earth and Mars. Astrophys. J. 583, 10241039.Google Scholar
Safronov, V. S. (1972). Evolution of the Protoplanetary Cloud and Formation of the Earth and Planets. NASA.Google Scholar
Sagan, C. (1973). Ultraviolet selection pressure on earliest organisms. J. Theor. Biol. 39, 195200.Google Scholar
Sagan, C. (1977). Reducing greenhouses and temperature history of Earth and Mars. Nature 269, 224226.Google Scholar
Sagan, C. (1994). Pale Blue Dot: A Vision of the Human Future in Space. New York: Random House.Google Scholar
Sagan, C. and Chyba, C. (1997). The early faint Sun paradox: Organic shielding of ultraviolet-labile greenhouse gases. Science 276, 12171221.Google Scholar
Sagan, C. and Khare, B. N. (1979). Tholins: Organic chemistry of interstellar grains and gas. Nature 277, 102107.Google Scholar
Sagan, C., et al. (1993a). Polycyclic aromatic hydrocarbons in the atmospheres of Titan and Jupiter. Astrophys. J. 414, 399405.Google Scholar
Sagan, C. and Mullen, G. (1972). Earth and Mars: Evolution of atmospheres and surface temperatures. Science 177, 5256.Google Scholar
Sagan, C., et al. (1993b). A search for life on Earth from the Galileo spacecraft. Nature 365, 715721.Google Scholar
Sahagian, D. L. and Maus, J. E. (1994). Basalt vesicularity as a measure of atmospheric pressure and palaeoelevation. Nature 372, 449451.Google Scholar
Sahoo, S. K., et al. (2012). Ocean oxygenation in the wake of the Marinoan glaciation. Nature 489, 546549.Google Scholar
Salaris, M. and Cassisi, S. (2005). Evolution of Stars and Stellar Populations. Hoboken, NJ: Wiley.Google Scholar
Salvatore, M. R., et al. (2010). Definitive evidence of Hesperian basalt in Acidalia and Chryse planitiae. J. Geophys. Res. 115, E07005.Google Scholar
Sam, H. and Yoram, L. (2014). Densities and eccentricities of 139 Kepler planets from transit time variations. Astrophys. J. 787, 80.Google Scholar
Sanchez-Lavega, A. (2010). An Introduction to Planetary Atmospheres. Boca Raton, FL: CRC Press/Taylor & Francis.Google Scholar
Sanchez-Lavega, A., et al. (2004). Clouds in planetary atmospheres: A useful application of the Clausius–Clapeyron equation. Am. J. Phys. 72, 767774.Google Scholar
Sandor, B. J., et al. (2010). Sulfur chemistry in the Venus mesosphere from SO2 and SO microwave spectra. Icarus 208, 4960.Google Scholar
Sano, Y. and Pillinger, C. T. (1990). Nitrogen isotopes and N2/Ar ratios in cherts: An attempt to measure time evolution of atmospheric delta-N-15 value. Geochem. J. 24, 315325.Google Scholar
Sanudo, J., et al. (1997). Bounds for the mass of the atmosphere. J. Geophys. Res. – Atmospheres 102, 70077009.Google Scholar
Sasaki, S. and Nakazawa, K. (1988). Origin of isotopic fractionation of terrestrial Xe: Hydrodynamic fractionation during escape of the primordial H2–He atmosphere. Earth Planet. Sc. Lett. 89, 323334.Google Scholar
Satkoski, A. M., et al. (2015). A redox-stratified ocean 3.2 billion years ago. Earth Planet. Sci. Lett. 430, 4353.Google Scholar
Satoh, M. (2004). Atmospheric circulation dynamics and general circulation models. Chichester, UK: Praxis Pub.Google Scholar
Saunders, A. and Reichow, M. (2009). The Siberian Traps and the End-Permian mass extinction: a critical review. Chinese Sci. Bull. 54, 2037.Google Scholar
Sausen, R. and Santer, B. D. (2003). Use of changes in tropopause height to detect human influences on climate. Meteorol. Z. 12, 131136.Google Scholar
Schaber, G. G. et al. (1992). Geology and distribution of impact craters on Venus: what are they telling us? J. Geophys. Res. 97, 13 25713 301.Google Scholar
Schaefer, L. and Fegley, B. (2010). Chemistry of atmospheres formed during accretion of the Earth and other terrestrial planets. Icarus 208, 438448.Google Scholar
Schaller, E. L. and Brown, M. E. (2007). Volatile loss and retention on Kuiper Belt objects. Astrophys. J. 659, L61L64.Google Scholar
Scharf, C. A. (2009). Extrasolar Planets and Astrobiology. Sausalito, CA: University Science Books.Google Scholar
Schenk, P. M., et al. (2004). Ages and interiors: The cratering record of the Galilean satellites. In: Jupiter: The Planet, Satellites, and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press.Google Scholar
Schenk, P. M. and Zahnle, K. (2007). On the negligible surface age of Triton. Icarus 192, 135149.Google Scholar
Schidlowski, M. (1988). A 3,800-million-year isotopic record of life from carbon in sedimentary rocks. Nature 333, 313318.Google Scholar
Schidlowski, M., et al. (1976). Carbon isotope geochemistry of Precambrian Lomagundi carbonate province, Rhodesia. Geochim. Cosmochim. Acta 40, 449455.Google Scholar
Schidlowski, M., et al. (1983). Isotopic inferences of ancient biochemistries: carbon, sulfur, hydrogen, and nitrogen. In: Earth’s Earliest Biosphere: Its Origin and Evolution, ed. Schopf, J. W., Princeton, New Jersey: Princeton University Press, pp. 149186.Google Scholar
Schinder, P. J., et al. (2011). The structure of Titan’s atmosphere from Cassini radio occultations. Icarus 215, 460474.Google Scholar
Schirrmeister, B. E., et al. (2013). Evolution of multicellularity coincided with increased diversification of cyanobacteria and the Great Oxidation Event. P. Natl. Acad. Sci. USA 110, 17911796.Google Scholar
Schlichting, H. E., et al. (2015). Atmospheric mass loss during planet formation: The importance of planetesimal impacts. Icarus 247, 8194.Google Scholar
Schmidt, G. A., et al. (2010). Attribution of the present-day total greenhouse effect. J. Geophys. Res. 115, D20106, doi:10.1029/2010JD014287.Google Scholar
Schmidt, P. W. and Williams, G. E. (1995). The Neoproterozoic climatic paradox - equatorial paleolatitude for marinoan glaciation near sea-level in south Australia. Earth and Planetary Science Letters 134, 107124.Google Scholar
Schmidt, P. W. and Williams, G. E. (2008). Palaeomagnetism of red beds from the Kimberley Group, Western Australia: Implications for the palaeogeography of the 1.8 Ga King Leopold glaciation. Precambrian Res. 167, 267280.Google Scholar
Schmidt, P. W., et al. (1991). Low paleolatitude of Late Proterozoic glaciation - early timing of remanence in hematite of the Elatina Formation, South-Australia. Earth Planet. Sci. Lett. 105, 355367.Google Scholar
Schmitt, B., et al. (1994). Identification of 3 absorption bands in the 2 micron spectrum of Io. Icarus 111, 79105.Google Scholar
Schneider, D. A., et al. (2002). Age of volcanic rocks and syndepositional iron formations, Marquette Range Supergroup: Implications for the tectonic setting of Paleoproterozoic, iron formations of the Lake Superior region. Can. J. Earth Sci. 39, 9991012.Google Scholar
Schneider, P. (2014). Extragalactic Astronomy and Cosmology: An Introduction. Berlin: Springer.Google Scholar
Schneider, T. (2006). The general circulation of the atmosphere. Annu. Rev. Earth Planet. Sci. 34, 655688.Google Scholar
Schneider, T. and Bordoni, S. (2008). Eddy-mediated regime transitions in the seasonal cycle of a Hadley circulation and implications for monsoon dynamics. J. Atmos. Sci. 65, 915934.Google Scholar
Schneider, T., et al. (2012). Polar methane accumulation and rainstorms on Titan from simulations of the methane cycle. Nature 481, 5861.Google Scholar
Schneider, T. and Liu, J. J. (2009). Formation of jets and equatorial superrotation on Jupiter. J. Atmos. Sci. 66, 579601.Google Scholar
Schoell, M. and Wellmer, F. W. (1981). Anomalous C-13 depletion in Early Precambrian graphites from Superior-Province, Canada. Nature 290, 696699.Google Scholar
Schofield, J. T. and Taylor, F. W. (1983). Measurements of the mean, solar-fixed temperature and cloud structure of the middle atmosphere of Venus. Q. J. R. Meteor. Soc. 109, 5780.Google Scholar
Scholz, A., et al. (2012). Substellar objects in nearby young clusters (SONYC) VI: The planetary-mass domain of NGC1333. Astrophys. J. 756, 24.Google Scholar
Schonheit, P., et al. (1980). Growth: Parameters (Ks, Mu-Max, Ys) of methanobacterium-thermoautotrophicum. Arch. Microbiol. 127, 5965.Google Scholar
Schopf, J. W. (1993). Microfossils of the Early Archean Apex chert: new evidence for the antiquity of life. Science 260, 640646.Google Scholar
Schopf, J. W. (2000). Solution to Darwin’s dilemma: Discovery of the missing Precambrian record of life. P. Natl. Acad. Sci. USA 97, 69476953.Google Scholar
Schopf, J. W., et al. (1983). Evolution of Earth’s earliest ecosystems: recent progress and unsolved problems. In: Earth’s Earliest Biosphere: Its Origin and Evolution, ed. Schopf, J. W., Princeton, NJ: Princeton University Press, pp. 361384.Google Scholar
Schopf, J. W. and Kudryavtsev, A. B. (2012). Biogenicity of Earth’s earliest fossils: A resolution of the controversy. Gondwana Res. 22, 761771.Google Scholar
Schopf, J. W., et al. (2002). Laser-Raman imagery of Earth’s earliest fossils. Nature 416, 7376.Google Scholar
Schopf, J. W. and Kudryavtsev, A. B. (2013). Reply to the comments of DL Pinti, R. Mineau and V. Clement, and of AO Marshall and CP Marshall on “Biogenicity of Earth’s earliest fossils: A resolution of the controversy” by J. William Schopf and Anatoliy B. Kudryavtsev, Gondwana Research 22 (2012), 761–771. Gondwana Res. 23, 16561658.Google Scholar
Schopf, J. W. and Packer, B. M. (1987). Early Archean (3.3 billion to 3.5 billion-year-old) microfossils from Warrawoona Group, Australia. Science 237, 7073.Google Scholar
Schrag, D. P., et al. (2002). On the initiationof a snowball Earth. Geochem. Geophys. Geosystems 3.Google Scholar
Schrag, D. P., et al. (2013). Authigenic carbonate and the history of the global carbon cycle. Science 339, 540543.Google Scholar
Schrag, D. P. and Hoffman, P. F. (2001). Geophysics - Life, geology and snowball Earth. Nature 409, 306.Google Scholar
Schrauzer, G. N. and Guth, T. D. (1977). Photolysis of water and photoreduction of nitrogen on titanium dioxide. J. Am. Chem. Soc. 99, 71897193.Google Scholar
Schroder, S., et al. (2008). Rise in seawater sulphate concentration associated with the Paleoproterozoic positive carbon isotope excursion: evidence from sulphate evaporites in the similar to 2.2-2.1 Gyr shallow-marine Lucknow Formation, South Africa. Terra Nova 20, 252.Google Scholar
Schubert, G. (1983). General circulation and the dynamical state of the Venus circulation. In: Venus, ed. Hunten, D. M., et al., Tucson: University of Arizona Press, pp. 681765.Google Scholar
Schubert, G., et al. (2010). Evolution of icy satellites. Space Sci. Rev. 153, 447484.Google Scholar
Schubert, G., et al. (2000). Geophysics: Timing of the Martian dynamo. Nature 408, 666667.Google Scholar
Schubert, G. and Soderlund, K. M. (2011). Planetary magnetic fields: Observations and models. Phys. Earth Planet. In. 187, 92108.Google Scholar
Schubert, G., et al. (2001). Mantle Convection in the Earth and Planets. Cambridge: Cambridge University Press.Google Scholar
Schunk, R. W. (1977). Mathematical structure of transport equations for multispecies flows. Rev. Geophys. 15, 429445.Google Scholar
Schunk, R. W. (1988). The polar wind. In: Modeling Magnetospheric Plasma, ed. Moore, T. E., Waite, J. H., Washington, D. C.: AGU, pp. 219228.Google Scholar
Schunk, R. W. and Nagy, A. (2009). Ionospheres: Physics, Plasma Physics, and Chemistry. New York: Cambridge University Press.Google Scholar
Schwarzschild, M. (1958). Structure and Evolution of the Stars. Princeton, NJ: Princeton University Press.Google Scholar
Schwieterman, E. W., et al. (2015). Detecting and constraining N2 abundances in planetary atmospheres using collisional pairs. Astrophys. J. 810, 57.Google Scholar
Scott, A. C. (2000). The Pre-Quaternary history of fire. Palaeogeogr. Palaeoclim. Palaeoecol. 164, 281329.Google Scholar
Scott, C., et al. (2008). Tracing the stepwise oxygenation of the Proterozoic ocean. Nature 452, 456459.Google Scholar
Scott, C., et al. (2014). Pyrite multiple-sulfur isotope evidence for rapid expansion and contraction of the early Paleoproterozoic seawater sulfate reservoir. Earth Planet. Sci. Lett. 389, 95104.Google Scholar
Seager, S. (2010). Exoplanet Atmospheres: Physical Processes. Princeton, N.J.: Princeton University Press.Google Scholar
Seager, S. (2013). Exoplanet habitability. Science 340, 577581.Google Scholar
Seager, S. and Bains, W. (2015). The search for signs of life on exoplanets at the interface of chemistry and planetary science. Science Advances 1, http://dx.doi.org/10.1126/sciadv.1500047.Google Scholar
Seager, S., et al. (2013a). A biomass-based model to estimate the plausibility of exoplanet biosignature gases. Astrophys. J. 775, 104.Google Scholar
Seager, S., et al. (2013b). Biosignature gases in H2-dominated atmospheres on rocky exoplanets. Astrophys. J. 777, 95.Google Scholar
Seager, S., et al. (2015). Exo-S: Starshade Probe-Class Exoplanet Direct Imaging Mission Concept Final Report. Pasadena, California.Google Scholar
Seager, S. and Sasselov, D. D. (2000). Theoretical transmission spectra during extrasolar giant planet transits. Astrophys. J. 537, 916921.Google Scholar
Seager, S., et al. (2012). An astrophysical view of Earth-based metabolic biosignature gases. Astrobiology 12, 6182.Google Scholar
Seager, S., et al. (2005). Vegetation’s red edge: a possible spectroscopic biosignature of extraterrestrial plants. Astrobiology 5, 372390.Google Scholar
Sefton-Nash, E. and Catling, D. C. (2008). Hematitic concretions at Meridiani Planum, Mars: Their growth timescale and possible relationship with iron sulfates. Earth Planet. Sci. Lett. 269, 365375.Google Scholar
Segatz, M., et al. (1988). Tidal dissipation, surface heat flow, and figure of viscoelastic models of Io. Icarus 75, 187206.Google Scholar
Segura, A., et al. (2005). Biosignatures from earth-like planets around M dwarfs. Astrobiology 5, 706725.Google Scholar
Segura, A., et al. (2003). Ozone concentrations and ultraviolet fluxes on Earth-like planets around other stars. Astrobiology 3, 689708.Google Scholar
Segura, A., et al. (2007). Abiotic production of O2 and O3 in high-CO2 terrestrial atmospheres. Astrobiology 7, 494495.Google Scholar
Segura, T. L., et al. (2008). Modeling the environmental effects of moderate-sized impacts on Mars. J. Geophys. Res. 113.Google Scholar
Segura, T. L., et al. (2002). Environmental effects of large impacts on Mars. Science 298, 19771980.Google Scholar
Segura, T. L., et al. (2013). The effects of impacts on the climates of terrestrial planets. In: Comparative Climatology of Terrestrial Planets, ed. Mackwell, S. J., et al., Tucson, AZ: University of Arizona Press.Google Scholar
Seiff, A. and Kirk, D. B. (1977). Structure of the atmosphere of Mars in summer at mid-latitudes. J. Geophys. Res. 82, 43644378.Google Scholar
Seiff, A., et al. (1980). Measurements of thermal structure and thermal constrasts in the atmosphere of Venus and related dynamical observations: Results from the four Pioneer Venus probes. J. Geophys. Res. 85, 79037933.Google Scholar
Seinfeld, J. H. and Pandis, S. N. (1998). Atmospheric Chemistry and Physics. New York: Wiley.Google Scholar
Seinfeld, J. H. and Pandis, S. N. (2006). Atmospheric Chemistry and Physics: From Air Pollution to Climate Change. Hoboken, N.J.: Wiley.Google Scholar
Sekine, Y., et al. (2011a). Replacement and late formation of atmospheric N2 on undifferentiated Titan by impacts. Nat. Geosci. 4, 359362.Google Scholar
Sekine, Y., et al. (2005). The role of Fischer–Tropsch catalysis in the origin of methane-rich Titan. Icarus 178, 154164.Google Scholar
Sekine, Y., et al. (2011b). Osmium evidence for synchronicity between a rise in atmospheric oxygen and Palaeoproterozoic deglaciation. Nat. Commun. 2.Google Scholar
Sekine, Y., et al. (2011c). Manganese enrichment in the Gowganda Formation of the Huronian Supergroup: A highly oxidizing shallow-marine environment after the last Huronian glaciation. Earth Planet. Sci. Lett. 307, 201210.Google Scholar
Sekiya, M., et al. (1981). Dissipation of the primordial terrestrial atmosphere due to irradiation of the solar far-UV during T-Tauri stage. Prog. Theor. Phys. 66, 13011316.Google Scholar
Sekiya, M., et al. (1980a). Dissipation of the primordial terrestrial atmosphere due to irradiation of the solar EUV. Prog. Theor. Phys. 64, 19681985.Google Scholar
Sekiya, M., et al. (1980b). Dissipation of the rare-gases contained in the primordial Earth’s atmosphere. Earth Planet. Sci. Lett. 50, 197201.Google Scholar
Sellers, W. D. (1969). A climate model based on the energy balance of the earth-atmosphere system. J. Appl. Meteor. 8, 392400.Google Scholar
Selsis, F., et al. (2002). Signature of life on exoplanets: Can Darwin produce false positive detections? Astron. Astrophys. 388, 9851003.Google Scholar
Senft, L. E. and Stewart, S. T. (2007). Modeling impact cratering in layered surfaces. J. Geophys. Res. 112.Google Scholar
Senft, L. E. and Stewart, S. T. (2008). Impact crater formation in icy layered terrains on Mars. Meteorit. Planet. Sci. 43, 19932013.Google Scholar
Senft, L. E. and Stewart, S. T. (2011). Modeling the morphological diversity of impact craters on icy satellites. Icarus 214, 6781.Google Scholar
Sessions, A. L., et al. (2009). The continuing puzzle of the Great Oxidation Event. Curr. Biol. 19, R567R574.Google Scholar
Settle, M. (1979). Formation and deposition of volcanic sulfate aerosols on Mars. J. Geophys. Res. 84, 83438354.Google Scholar
Setzmann, U. and Wagner, W. (1991). A new equation of state and tables of thermodynamic properties for methane covering the range from the meling line to 625 K at pressures up to 1000 MPa. J. Phys. Chem. Ref. Data 20, 10611155.Google Scholar
Sforna, M. C., et al. (2014). Structural characterization by Raman hyperspectral mapping of organic carbon in the 3.46 billion-year-old Apex chert, Western Australia. Geochim. Cosmochim. Acta 124, 1833.Google Scholar
Shaheen, R., et al. (2015). Carbonate formation events in ALH 84001 trace the evolution of the Martian atmosphere. P. Natl. Acad. Sci. USA 112, 336341.Google Scholar
Shalygin, E. V., et al. (2015). Active volcanism on Venus in the Ganiki Chasma rift zone. Geophys. Res. Lett. 42, 47624769.Google Scholar
Shanks, W. C. and Seyfried, W. E. (1987). Stable isotope studies of vent fluids and chimney minerals, Southern Juan-De-Fuca Ridge: Sodium metasomatism and seawater sulfate reduction. J. Geophys. Res. 92, 11 38711 399.Google Scholar
Shapiro, R. (1995). The prebiotic role of adenine: A critical analysis. Origins Life Evol. Biosph. 25, 8398.Google Scholar
Shapley, H. (1953). Climatic Change: Evidence, Causes, and Effects. Cambridge: Harvard University Press.Google Scholar
Sharma, A., et al. (2002). Microbial activity at gigapascal pressures. Science 295, 15141516.Google Scholar
Sharp, C. M. and Burrows, A. (2007). Atomic and molecular opacities for brown dwarf and giant planet atmospheres. Astrophys. J. Suppl. S. 168, 140166.Google Scholar
Sharp, R. P. (1973). Mars: Fretted and chaotic terrain. J. Geophys. Res. 78, 40734083.Google Scholar
Shea, M. A. and Smart, D. F. (2000). Fifty years of cosmic radiation data. Space Sci. Rev. 93, 229262.Google Scholar
Sheldon, N. D. (2006). Precambrian paleosols and atmospheric CO2 levels. Precambrian Res. 147, 148155.Google Scholar
Sheldon, N. D. and Tabor, N. J. (2009). Quantitative paleoenvironmental and paleoclimatic reconstruction using paleosols. Earth Sci. Rev. 95, 152.Google Scholar
Shematovich, V. I., et al. (2003). Nitrogen loss from Titan. J. Geophys. Res. – Planets 108.Google Scholar
Shen, Y. and Buick, R. (2004). The antiquity of microbial sulfate reduction. Earth Sci. Rev. 64, 243272.Google Scholar
Shen, Y., et al. (2001). Isotopic evidence for microbial sulfate reduction in the early Archean era. Nature 410, 7781.Google Scholar
Shen, Y., et al. (2003). Evidence for low sulphate and anoxia in a mid-Proterozoic marine basin. Nature 423, 632635.Google Scholar
Shen, Y. N., et al. (2002). Middle proterozoic ocean chemistry: Evidence from the McArthur Basin, northern Australia. Am. J. Sci. 302, 81109.Google Scholar
Shen, Y. N., et al. (2008). On the coevolution of Ediacaran oceans and animals. P. Natl. Acad. Sci. USA 105, 73767381.Google Scholar
Shetty, S. and Marcus, P. S. (2010). Changes in Jupiter’s Great Red Spot (1979-2006) and Oval BA (2000-2006). Icarus 210, 182201.Google Scholar
Shibuya, T., et al. (2012). Depth variation of carbon and oxygen isotopes of calcites in Archean altered upper oceanic crust: Implications for the CO2 flux from ocean to oceanic crust in the Archean. Earth Planet. Sci. Lett. 321, 6473.Google Scholar
Shields, A. L., et al. (2013). The effect of host star spectral energy distribution and ice-albedo feedback on the climate of extrasolar planets. Astrobiology 13, 715739.Google Scholar
Shields, G. and Veizer, J. (2002). Precambrian marine carbon isotope database: version 1.1. Geol. Geochem. Geophys. 3, 10.1029/2001GC000266.Google Scholar
Shields, G. A. (2007). A normalized seawater strontium isotope curve: Possible implications for Neoproterozoic Cambrian weathering rates and further oxygenation of the Earth. Earth 2, 3542.Google Scholar
Shields-Zhou, G. and Och, L. (2011). The case for a Neoproterozoic Oxygenation Event: Geochemical evidence and biological consequences. GSA Today 21, 411.Google Scholar
Shine, K. P., et al. (2012). The water vapour continuum: Brief history and recent cevelopments. Surv. Geophys. 33, 535555.Google Scholar
Shirey, S. B., et al. (2013). Diamonds and the geology of mantle carbon. Rev. Mineral. Geochem. 75, 355421.Google Scholar
Shizgal, B. and Blackmore, R. (1986). A collisional kinetic theory of a plane parallel evaporating planetary atmosphere. Planet. Space Sci. 34, 279291.Google Scholar
Shizgal, B. and Lindenfeld, M. J. (1980). Further studies of non-Maxwellian effects associated with the thermal escape of a planetary atmosphere. Planet. Space Sci. 28, 159163.Google Scholar
Shizgal, B. D. and Arkos, G. G. (1996). Nonthermal escape of the atmospheres of Venus, Earth and Mars. Rev. Geophys. 34, 483505.Google Scholar
Shklovskii, I. S. (1951). On the possibility of explaining the difference in chemical compostion of the Earth and Sun by thermal dissipation of light gases. Astron. Zh. 28, 234243.Google Scholar
Shklovskii, I. S. and Sagan, C. (1966). Intelligent Life in the Universe. San Francisco, CA: Holden-Day.Google Scholar
Shock, E. L. (1990). Geochemical constraints on the origin of organic compounds in hydrothermal systems. Origins Life Evol. Biosph. 20, 331367.Google Scholar
Showman, A. P., et al. (2010). Atmospheric circulation of extrasolar planets. In: Exoplanets, ed. Seager, S., Tucson: University of Arizona Press, pp. 471516.Google Scholar
Showman, A. P., et al. (2013a). Doppler signatures of the atmospheric circulation on Hot Jupiters. Astrophys. J. 762, doi:10.1088/0004-637X/762/1/24.Google Scholar
Showman, A. P. and Guillot, T. (2002). Atmospheric circulation and tides of “51 Pegasus b-like” planets. Astron. Astrophys. 385, 166180.Google Scholar
Showman, A. P. and Polvani, L. M. (2011). Equatorial superrotation on tidally locked exoplanets. Astrophys. J. 738, 71, doi:10.1088/0004-637X/738/1/71.Google Scholar
Showman, A. P., et al. (2013b). Atmospheric circulation of terrestrial exoplanets. In: Comparative Climatology of Terrestrial Planets, ed. Mackwell, S. J., et al., Tucson: University of Arizona Press, pp. 277326.Google Scholar
Shu, F. H. (1991). The Physics of Astrophysics: Volume II, Gas Dynamics. Mill Valley, Calif.: University Science Books.Google Scholar
Shuvalov, V. (2009). Atmospheric erosion induced by oblique impacts. Meteorit. Planet. Sci. 44, 10951105.Google Scholar
Shuvalov, V., et al. (2014). Impact induced erosion of hot and dense atmospheres. Planet. Space Sci. 98, 120127.Google Scholar
Sicardy, B., et al. (2006). Charon’s size and an upper limit on its atmosphere from a stellar occultation. Nature 439, 5254.Google Scholar
Sicardy, B., et al. (2011). A Pluto-like radius and a high albedo for the dwarf planet Eris from an occultation. Nature 478, 493496.Google Scholar
Sicardy, B., et al. (2003). Large changes in Pluto’s atmosphere as revealed by recent stellar occultations. Nature 424, 168170.Google Scholar
Sigman, D. M., et al. (2008). Nitrogen isotopes in the ocean. In: Encyclopedia of Ocean Sciences (2nd Edn), ed. Steele, J. H., et al., San Diego: Academic Press, pp. 4054.Google Scholar
Silburt, A., et al. (2015). A statistical reconstruction of the planet population around Kepler solar-type stars. Astrophys. J. 799, 180.Google Scholar
Sillanpaa, I., et al. (2011). Cassini Plasma Spectrometer and hybrid model study on Titan’s interaction: Effect of oxygen ions. J. Geophys. Res. 116.Google Scholar
Silva-Tamayo, J. C., et al. (2010). Global perturbation of the marine Ca isotopic composition in the aftermath of the Marinoan global glaciation. Precambrian Res. 182, 373381.Google Scholar
Simoncini, E., et al. (2013). Quantifying drivers of chemical disequilibrium: theory and application to methane in the Earth’s atmosphere. Earth Syst. Dynam. 4, 317331.Google Scholar
Simpson, G. C. (1927). Some studies in terrestrial radiation. Mem. R. Met. Soc. 11, 6995.Google Scholar
Sittler, E. C., et al. (2006). Energetic nitrogen ions within the inner magnetosphere of Saturn. J. Geophys. Res. 111.Google Scholar
Slack, J. F. and Cannon, W. F. (2009). Extraterrestrial demise of banded iron formations 1.85 billion years ago. Geology 37, 10111014.Google Scholar
Slack, J. F., et al. (2007). Suboxic deep seawater in the late Paleoproterozoic: Evidence from hematitic chert and iron formation related to seafloor-hydrothermal sulfide deposits, central Arizona, USA. Earth Planet. Sci. Lett. 255, 243256.Google Scholar
Sleep, N. H. (1994). Martian plate tectonics. J. Geophys. Res. 99, 56395655.Google Scholar
Sleep, N. H. (2005). Dioxygen over geologic time. In: Metal Ions in Biological Systems, Vol. 43, Biogeochemical Cycles of Elements, ed. Sigel, A., et al., Boca Raton, FL: Taylor & Francis, pp. 4973.Google Scholar
Sleep, N. H. (2007). Plate tectonics through time. In: Treatise on Geophysics, Amsterdam: Elsevier, pp. 145169.Google Scholar
Sleep, N. H. and Windley, B. F. (1982). Archean plate tectonics: Constraints and inferences. J. Geol. 90, 363379.Google Scholar
Sleep, N. H. and Zahnle, K. (1998). Refugia from asteroid impacts on early Mars and the early Earth. J. Geophys. Res. 103, 28 52928 544.Google Scholar
Sleep, N. H. and Zahnle, K. (2001). Carbon dioxide cycling and implications for climate on ancient Earth. J. Geophys. Res. 106, 13731399.Google Scholar
Sleep, N. H., et al. (1989). Annihilation of ecosystems by large asteroid impacts on the early Earth. Nature 342, 139142.Google Scholar
Slinn, W. G. N., et al. (1978). Some aspects of the transfer of atmospheric trace constituents past the air–sea interface. Atmos. Environ. 12, 20552087.Google Scholar
Smith, D. E., et al. (2009a). Time variations of Mars’ gravitational field and seasonal changes in the masses of the polar ice caps. J. Geophys. Res. 114.Google Scholar
Smith, G. R. and Hunten, D. M. (1990). Study of planetary atmospheres by absorptive occultations. Rev. Geophys. 28, 117143.Google Scholar
Smith, G. S. (2005). Human color vision and the unsaturated blue color of the daytime sky. Am. J. Phys. 73, 590597.Google Scholar
Smith, M. D., et al. (2001). Thermal Emission Spectrometer results: Mars atmospheric thermal structure and aerosol distribution. J. Geophys. Res. 106, 23 92923 945.Google Scholar
Smith, M. L., et al. (2014). The formation of sulfate, nitrate and perchlorate salts in the martian atmosphere. Icarus 231, 5164.Google Scholar
Smith, P. H., et al. (2009b). H2O at the Phoenix Landing Site. Science 325, 5861.Google Scholar
Smrekar, S. E., et al. (2010). Recent hotspot volcanism on Venus from VIRTIS emissivity data. Science 328, 605608.Google Scholar
Smyth, W. H. and Marconi, M. L. (2006). Europa’s atmosphere, gas tori, and magnetospheric implications. Icarus 181, 510526.Google Scholar
Snellen, I. A. G., et al. (2010). The orbital motion, absolute mass and high-altitude winds of exoplanet HD 209458b. Nature 465, 10491051.Google Scholar
Snellen, I. A. G., et al. (2013). Finding extraterrestrial life using ground-based high-dispersion spectroscopy. Astrophys. J. 764, 182.Google Scholar
Snyder, C. W. (1979). Planet Mars as seen at the end of the Viking Mission. J. Geophys. Res. 84, 84878519.Google Scholar
Sobolev, V. V. (1975). Light Scattering in Planetary Atmospheres. Oxford: Pergamon.Google Scholar
Soderblom, J. M., et al. (2012). Modeling specular reflections from hydrocarbon lakes on Titan. Icarus 220, 744751.Google Scholar
Soderblom, L. A., et al. (1990). Tritons geyser-like plumes: Discovery and basic characterization. Science 250, 410415.Google Scholar
Sohl, L. E., et al. (1999). Paleomagnetic polarity reversals in Marinoan (ca. 600 Ma) glacial deposits of Australia: Implications for the duration of low-latitude glaciation in neoproterozoic time. Geol. Soc. Am. Bull. 111, 11201139.Google Scholar
Solanki, S. K. and Krivova, N. A. (2003). Can solar variability explain global warming since 1970? J. Geophys. Res. 108, 1200, doi: 10.1029/2002JA009753.Google Scholar
Solanki, S. K., et al. (2004). Unusual activity of the Sun during recent decades compared to the previous 11,000 years. Nature 431, 10841087.Google Scholar
Solomon, S. (1999). Stratospheric ozone depletion: A review of concepts and history. Rev. Geophys. 37, 275316.Google Scholar
Som, S. M., et al. (2012). Air density 2.7 billion years ago limited to less than twice modern levels by fossil raindrop imprints. Nature 484, 359362.Google Scholar
Som, S. M., et al. (2016). Air pressure limited to less than half modern levels 2.7 billion years ago. Nat. Geosci. 9, 448451.Google Scholar
Spang, A., et al. (2015). Complex archaea that bridge the gap between prokaryotes and eukaryotes. Nature 521, 173179.Google Scholar
Sparks, W. B., et al. (2016). Probing for evidence of plumes on Europa with HST/STIS. Astrophys. J. 829, 121.Google Scholar
Spencer, J. R. (1990). Nitrogen frost migration on Triton: A historical model. Geophys. Res. Lett. 17, 17691772.Google Scholar
Spencer, J. R., et al. (2000). Discovery of gaseous S2 in Io’s Pele plume. Science 288, 12081210.Google Scholar
Spencer, J. R. and Nimmo, F. (2013). Enceladus: An active ice world in the Saturn system. Annu. Rev. Earth Planet. Sci. 41, 693717.Google Scholar
Spencer, J. R., et al. (2006). Cassini encounters Enceladus: Background and the discovery of a south polar hot spot. Science 311, 14011405.Google Scholar
Spergel, D., et al. (2015). Wide-Field InfrarRed Survey Telescope-Astrophysics Focused Telescope Assets WFIRST-AFTA 2015 Report, arXiv preprint arXiv:1503.03757.Google Scholar
Sperling, E. A., et al. (2013). Oxygen, ecology, and the Cambrian radiation of animals. P. Natl. Acad. Sci. USA 110, 13 44613 451.Google Scholar
Sperling, E. A., et al. (2014). Redox heterogeneity of subsurface waters in the Mesoproterozoic ocean. Geobiology 12, 373386, doi: 10.1111/gbi.12091.Google Scholar
Spitzer, L. J. (1952). In: The Terrestrial Atmosphere Above 300 km, ed. Kuiper, G. P., Chicago: University of Chicago Press, pp. 211247.Google Scholar
Sprague, A. L., et al. (2004). Mars’ south polar Ar enhancement: A tracer for south polar seasonal meridional mixing. Science 306, 13641367.Google Scholar
Sprague, A. L., et al. (2007). Mars’ atmospheric argon: Tracer for understanding Martian atmospheric circulation and dynamics. J. Geophys. Res. 112, E03S02, doi:10.1029/2005je002597.Google Scholar
Squyres, S. W. and Knoll, A. H. (2005). Sedimentary rocks at Meridiani Planum: Origin, diagetiesis, and implications for life on Mars. Earth Planet. Sci. Lett. 240, 110.Google Scholar
Squyres, S. W., et al. (2009). Exploration of Victoria Crater by the Mars Rover Opportunity. Science 324, 10581061.Google Scholar
Squyres, S. W., et al. (2006). Two years at Meridiani Planum: Results from the Opportunity Rover. Science 313, 14031407.Google Scholar
Stamper, C. C., et al. (2014). Oxidised phase relations of a primitive basalt from Grenada, Lesser Antilles. Contrib. Mineral. Petr. 167, Art. 954.Google Scholar
Stanley, B. D., et al. (2011). CO2 solubility in Martian basalts and Martian atmospheric evolution. Geochim. Cosmochim. Acta 75, 59876003.Google Scholar
Stapelfeldt, K., et al. (2015). Exo-C Imaging Nearby Worlds: Exoplanet Direct Imaging: coronograph probe mission study final report. Pasadena, California.Google Scholar
Steele, A., et al. (2012). A reduced organic carbon component in martian basalts. Science 337, 212215.Google Scholar
Steemans, P., et al. (2009). Origin and radiation of the earliest vascular land plants. Science 324, 353353.Google Scholar
Stephan, K., et al. (2010). Specular reflection on Titan: Liquids in Kraken Mare. Geophysical Research Letters 37.Google Scholar
Stephens, G. L. and Tjemkes, S. A. (1993). Water vapour and its role in the Earth’s greenhouse. Aust. J. Phys. 46, 149166.Google Scholar
Stephens, S. K. (1995a). Carbonate formation on Mars: Experiments and Models. California Institute of Technology, PhD, Pasadena, CA.Google Scholar
Stephens, S. K. (1995b). Carbonates on Mars: Experimental results. Proc. Lunar Planet. Sci. Conf. XXVI, 13551356.Google Scholar
Stepinski, T. F., et al. (2004). Martian geomorphology from fractal analysis of drainage networks. J. Geophys. Res. 109.Google Scholar
Stern, J. C., et al. (2015a). Evidence for indigenous nitrogen in sedimentary and aeolian deposits from the Curiosity rover investigations at Gale crater, Mars. P. Natl. Acad. Sci. USA 112, 42454250.Google Scholar
Stem, R. J. (2007). When and how did plate tectonics begin? Theoretical and empirical considerations. Chinese Sci. Bull. 52, 578591.Google Scholar
Stern, S. A. (2008). On the atmospheres of objects in the Kuiper Belt. In: The Solar System Beyond Neptune, ed. Barucci, M. A., et al., Tucson: University of Arizona Press, pp. 365380.Google Scholar
Stern, S. A., et al. (2015b). The Pluto system: Initial results from its exploration by New Horizons. Science 350, doi:10.1126/science.aad1815.Google Scholar
Stevens, B. and Bony, S. (2013). Water in the atmosphere. Phys. Today 66, 2934.Google Scholar
Stevenson, D. J. (1975). Thermodynamics and phase separation of dense fully ionized hydrogen–helium fluid mixtures. Phys. Rev. B 12, 39994007.Google Scholar
Stevenson, D. J. (1983). The nature of the Earth prior to the oldest known rock record: the Hadean Earth. In: Earth’s Earliest Biosphere: Its Origin and Evolution, ed. Schopf, J. W., Princeton, New Jersey: Princeton University Press, pp. 3240.Google Scholar
Stevenson, D. J. (1987). Origin of the Moon – the collision hypothesis. Ann. Rev. Earth Planet. Sci. 15, 271315.Google Scholar
Stevenson, D. J. (1999). Life-sustaining planets in interstellar space? Nature 400, 3232.Google Scholar
Stevenson, D. J. (2001). Mars’ core and magnetism. Nature 412, 214219.Google Scholar
Stevenson, D. J. and Salpeter, E. E. (1977a). Dynamics and helium distribution in hydrogen-helium fluid planets. Astrophys. J. Suppl. S. 35, 239261.Google Scholar
Stevenson, D. J. and Salpeter, E. E. (1977b). Phase diagram and transport properties for hydrogen–helium fluid planets. Astrophys. J. Suppl. S. 35, 221237.Google Scholar
Stevenson, K. B., et al. (2014). Thermal structure of an exoplanet atmosphere from phase-resolved emission spectroscopy. Science 346, 838841.Google Scholar
Stiles, B. W., et al. (2008). Determining Titan’s spin state from Cassini RADAR images. Astron. J. 135, 16691680.Google Scholar
Stillman, D. E., et al. (2014). New observations of martian southern mid-latitude recurring slope lineae (RSL) imply formation by freshwater subsurface flows. Icarus 233, 328341.Google Scholar
Stocker, T. F., et al. (2013). IPCC, 2013: Climate Change: The Physical Science Basis. New York: Cambridge University Press.Google Scholar
Stofan, E. R., et al. (2007). The lakes of Titan. Nature 445, 6164.Google Scholar
Stone, E. C., et al. (2013). Voyager 1 observes low-energy galactic cosmic rays in a region depleted of heliospheric ions. Science 341, 150153.Google Scholar
Stoney, G. J. (1898). Of atmospheres on planets and satellites. Astrophys. J. 7, 2555.Google Scholar
Stoney, G. J. (1900a). Note on inquiries as to the escape of gases from atmospheres. Astrophys. J. 12, 201207.Google Scholar
Stoney, G. J. (1900b). On the escape of gases from planetary atmospheres according to the kinetic theory. I. Astrophys. J. 11, 251258.Google Scholar
Stoney, G. J. (1900c). On the escape of gases from planetary atmospheres according to the kinetic theory. II. Astrophys. J. 11, 357372.Google Scholar
Stoney, G. J. (1904). Escape of gases from atmospheres. Astrophys. J. 20, 6978.Google Scholar
Strauss, H., et al. (2001). The sulfur isotopic composition of Neoproterozoic to early Cambrian seawater – evidence from the cyclic Hanseran evaporites, NW India. Chem. Geol. 175, 1728.Google Scholar
Stribling, R. and Miller, S. L. (1987). Energy yields for hydrogen cyanide and formaldehyde syntheses: the HCN and amino acid concentrations in the primitive ocean. Origins Life Evol. Biosph. 17, 261273.Google Scholar
Strobel, D. F. (2002). Aeronomic systems on planets, moons, and comets. In: Atmospheres in the Solar System: Comparative Aeronomy ed. Mendillo, M., et al., Washington, D. C.: AGU, pp. 722.Google Scholar
Strobel, D. F. (2008a). N2 escape rates from Pluto’s atmosphere. Icarus 193, 612619.Google Scholar
Strobel, D. F. (2008b). Titan’s hydrodynamically escaping atmosphere. Icarus 193, 588594.Google Scholar
Strobel, D. F. (2009). Titan’s hydrodynamically escaping atmosphere: Escape rates and the structure of the exobase region. Icarus 202, 632641.Google Scholar
Strobel, D. F. (2010). Molecular hydrogen in Titan’s atmosphere: Implications of the measured tropospheric and thermospheric mole fractions. Icarus 208, 878886.Google Scholar
Strobel, D. F. (2012). Hydrogen and methane in Titan’s atmosphere: chemistry, diffusion, escape, and the Hunten limiting flux principle. Can. J. Phys. 90, 795805.Google Scholar
Strobel, D. F., et al. (2009). Atmospheric structure and composition. In: Titan from Cassini-Huygens, ed. Brown, R. H., et al., New York: Springer, pp. 235257.Google Scholar
Strughold, H. (1953). The Green and Red Planet. Albuquerque: University of New Mexico Press.Google Scholar
Strughold, H. (1955). The ecosphere of the Sun. Avia. Med. 26, 323328.Google Scholar
Stubenrauch, C. J., et al. (2013). Assessment of global cloud datasets from satellites: Project and database initiated by the GEWEX radiation panel. Bull. Am. Met. Soc. 94, 10311049.Google Scholar
Stüeken, E. E. (2016). Nitrogen in ancient mud: A biosignature? Astrobiology 16, 730735.Google Scholar
Stüeken, E. E., et al. (2015a). Selenium isotopes support free O2 in the latest Archean. Geology 43, 259262.Google Scholar
Stüeken, E. E., et al. (2015b). The evolution of the global selenium cycle: Secular trends in Se isotopes and abundances. Geochim. Cosmochim. Acta 162, 109125.Google Scholar
Stüeken, E. E., et al. (2015c). Isotopic evidence for biological nitrogen fixation by molybdenum-nitrogenase from 3.2 Gyr. Nature 520, 666670.Google Scholar
Stüeken, E. E., et al. (2015d). Nitrogen isotope evidence for alkaline lakes on late Archean continents. Earth Planet. Sc. Lett. 411, 110.Google Scholar
Stüeken, E. E., et al. (2012). Contributions to late Archaean sulphur cycling by life on land. Nat. Geosci. 5, 722725.Google Scholar
Su, H., et al. (2008). Variations of tropical upper tropospheric clouds with sea surface temperature and implications for radiative effects. J. Geophys. Res. 113,.Google Scholar
Sumi, T., et al. (2011). Unbound or distant planetary mass population detected by gravitational microlensing. Nature 473, 349352.Google Scholar
Summons, R. E., et al. (2006). Steroids, triterpenoids and molecular oxygen. Phil. Trans. R. Soc. B 361, 951968.Google Scholar
Summons, R. E. and Brocks, J. J. (2004). Sedimentary hydrocarbons, biomarkers for early life. In: Treatise on Geochemistry – Volume 8: Biogeochemistry, ed. Schlesinger, W. H., Amsterdam: Elsevier, pp. 63115.Google Scholar
Sun, S., et al. (2015). Primary hematitte in Neoarchean to Paleoproterozoic oceans. GSA Bull. 127, 850861.Google Scholar
Sutherland, J. D. (2016). The origin of life: Out of the blue. Angew. Chem. Int. Edit. 55, 104121.Google Scholar
Sutter, B., et al. (2012). The detection of carbonate in the martian soil at the Phoenix Landing site: A laboratory investigation and comparison with the Thermal and Evolved Gas Analyzer (TEGA) data. Icarus 218.Google Scholar
Svensen, H. and Jamtveit, B. (2010). Metamorphic fluids and global environmental changes. Elements 6, 179182.Google Scholar
Svensmark, H. (2007). Cosmoclimatology: a new theory emerges. Astron. Geophys. 48, 1824.Google Scholar
Svetsov, V. V. (2007). Atmospheric erosion and replenishment induced by impacts of cosmic bodies upon the Earth and Mars. Solar Syst. Res. 41, 2841.Google Scholar
Swanson-Hysell, N. L., et al. (2010). Cryogenian glaciation and the onset of carbon-isotope decoupling. Science 328, 608611.Google Scholar
Swindle, T. D., et al. (1986). Xenon and other noble gases in Shergottites. Geochim. Cosmochim. Acta 50, 10011015.Google Scholar
Symonds, R. B., et al. (1994). Volcanic-gas studies: Methods, results, and applications. In: Volatiles in Magmas, ed. Carroll, M. R. and Holloway, J. R.: Mineralogical Society of America.Google Scholar
Szczepanieccieciak, E., et al. (1978). Estimation of solubility of solidified substances in liquid methane by Preston-Prausnitz Method. Cryogenics 18, 591600.Google Scholar
Szymanski, A., et al. (2010). High oxidation state during formation of Martian nakhlites. Meteorit. Planet. Sci. 45, 2131.Google Scholar
Tajika, E. (1998). Mantle degassing of major and minor volatile elements during the Earth’s history. Geophys. Res. Lett. 25, 39913994.Google Scholar
Takahashi, H., et al. (1963). In: Comparative Biogeochemistry, ed. Florkin, M. and Mason, H. S., New York: Academic Press, pp. 91.Google Scholar
Takai, K., et al. (2008). Cell proliferation at 122 degrees C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high-pressure cultivation. P. Natl. Acad. Sci. USA 105, 10 94910 954.Google Scholar
Tam, S. W. Y., et al. (2007). Kinetic modeling of the polar wind. J. Atmos. Sol-Terr. Phys. 69, 19842027.Google Scholar
Tanaka, K. L. (1986). The stratigraphy of Mars. J. Geophys. Res. 91, E139E158.Google Scholar
Tanaka, K. L., et al. (2014). The digital global geologic map of Mars: Chronostratigraphic ages, topographic and crater morphologic characteristics, and updated resurfacing history. Planet. Space Sci. 95, 1124.Google Scholar
Tappert, R., et al. (2013). Stable carbon isotopes of C3 plant resins and ambers record changes in atmospheric oxygen since the Triassic. Geochim. Cosmochim. Acta 121, 240262.Google Scholar
Taran, V. Y. A. and Giggenbach, W. F. (2003). Geochemistry of light hydrocarbons in subduction-related volcanic and hydrothermal fluids. In: Special Publication 10: Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of Processes within the Earth, ed. Simmons, S. F. and Graham, I., Littleton, CO: Soc. of Econ. Geol., pp. 6174.Google Scholar
Taylor, F. and Grinspoon, D. (2009). Climate evolution of Venus. J. Geophys. Res. 114, E00B40, doi: 10.1029/2008JE003316.Google Scholar
Taylor, F. W. (2010). Planetary Atmospheres. Oxford; New York: Oxford University Press.Google Scholar
Taylor, F. W., et al. (1997). Near-infrared sounding of the lower atmosphere of Venus. In: Venus II–Geology, Geophysics, Atmosphere, and Solar Wind Environment, ed. Bougher, S. W., et al., Tucson, AZ: University of Arizona Press.Google Scholar
Taylor, F. W., et al. (1983). The thermal balance of the middle and upper atmosphere of Venus. In: Venus, ed. Hunten, D. M., et al., Tucson: University of Arizona Press, pp. 650680.Google Scholar
Taylor, G. J. (2013). The bulk composition of Mars. Chemie der Erde - Geochemistry 73, 401420.Google Scholar
Taylor, J. R. (2005). Classical Mechanics. Sausalito, Calif.: University Science Books.Google Scholar
Taylor, S. R. (2001). Solar System Evolution. New York: Cambridge University Press.Google Scholar
Teanby, N. A., et al. (2008). Titan’s winter polar vortex structure revealed by chemical tracers. J. Geophys. Res. 113, E12003, doi:10.1029/2008JE003218.Google Scholar
Teanby, N. A., et al. (2012). Active upper-atmosphere chemistry and dynamics from polar circulation reversal on Titan. Nature 491, 732735.Google Scholar
Tellmann, S., et al. (2009). Structure of the Venus neutral atmosphere as observed by the Radio Science experiment VeRa on Venus Express. J. Geophys. Res. 114.Google Scholar
Tenishev, V., et al. (2010). An approach to numerical simulation of the gas distribution in the atmosphere of Enceladus. J. Geophys. Res. 115.Google Scholar
Tenishev, V., et al. (2014). Effect of the Tiger Stripes on the water vapor distribution in Enceladus’ exosphere. J. Geophys. Res. 119, 26582667.Google Scholar
Teolis, B. D., et al. (2010). Cassini finds an oxygen–carbon dioxide atmosphere at Saturn’s icy moon Rhea. Science 330, 18131815.Google Scholar
Tera, F., et al. (1974). Isotopic evidence for a terminal lunar cataclysm. Earth Planet. Sci. Lett. 22, 121.Google Scholar
Thiemens, M. H. (2006). History and applications of mass-independent isotope effects. Annu. Rev. Earth Pl. Sc. 34, 217262.Google Scholar
Thiemens, M. H. and Heidenreich, J. E. (1983). The mass-independent fractionation of oxygen: A novel isotope effect and its possible cosmochemical implications. Science 219, 10731075.Google Scholar
Thomas, C., et al. (2007). Clathrate hydrates as a sink of noble gases in Titan’s atmosphere. Astron. Astrophys. 474, L17L20.Google Scholar
Thomas, G. E. and Stamnes, K. (1999). Radiative Transfer in the Atmosphere and Ocean. Cambridge; New York: Cambridge University Press.Google Scholar
Thomas, P. C., et al. (2009). Residual south polar cap of Mars: Stratigraphy, history, and implications of recent changes. Icarus 203, 352375.Google Scholar
Thomas, P. C., et al. (2016). Enceladus’s measured physical libration requires a global subsurface ocean. Icarus 264, 3747.Google Scholar
Thomazo, C., et al. (2011). Extreme 15N-enrichments in 2.72-Gyr-old sediments: evidence for a turning point in the nitrogen cycle. Geobiology 9, 107120.Google Scholar
Thomazo, C. and Papineau, D. (2013). Biogeochemical cycling of nitrogen on the early Earth. Elements 9, 345351.Google Scholar
Thomazo, C., et al. (2009). Biological activity and the Earth’s surface evolution: Insights from carbon, sulfur, nitrogen and iron stable isotopes in the rock record. Comptes Rendus Palevol 8, 665678.Google Scholar
Thommes, E. W., et al. (2002). The formation of Uranus and Neptune among Jupiter and Saturn. Astron. J. 123, 28622883.Google Scholar
Thompson, A. M. and Cicerone, R. J. (1986). Possible perturbations to atmospheric CO, CH4, and OH. J. Geophys. Res. 91, 853864.Google Scholar
Thompson, R. O. R. Y. (1971). Why there is an intense eastward current in the North Atlantic but not in the South Atlantic. J. Phys. Ocean. 1, 235237.Google Scholar
Tian, F., et al. (2010). Photochemical and climate consequences of sulfur outgassing on early Mars. Earth Planet. Sci. Lett. 295, 412418.Google Scholar
Tian, F., et al. (2014). High stellar FUV/NUV ratio and oxygen contents in the atmospheres of potentially habitable planets. Earth Planet. Sci. Lett. 385, 2227.Google Scholar
Tian, F., et al. (2009). Thermal escape of carbon from the early Martian atmosphere. Geophysical Research Letters 36.Google Scholar
Tian, F., et al. (2011). Revisiting HCN formation in Earth’s early atmosphere. Earth Planet. Sci. Lett. 308, 417423.Google Scholar
Tian, F. and Toon, O. B. (2005). Hydrodynamic escape of nitrogen from Pluto. Geophys. Res. Lett. 32.Google Scholar
Tian, F., et al. (2005). A hydrogen-rich early Earth atmosphere. Science 308, 10141017.Google Scholar
Till, C. B., et al. (2012). The beginnings of hydrous mantle wedge melting. Contrib. Mineral. Petr. 163, 669688.Google Scholar
Tinetti, G., et al. (2006). Detectability of planetary characteristics in disk-averaged spectra II: Synthetic spectra and light-curves of earth. Astrobiology 6, 881900.Google Scholar
Tinsley, B. A. (1974). Hydrogen in the upper atmosphere. Fund. Cosmic Phy. 1, 201300.Google Scholar
Tirard, S., et al. (2010). The definition of life: A brief history of an elusive scientific endeavor. Astrobiology 10, 10031009.Google Scholar
Titov, D. V., et al. (2007). Radiation in the atmosphere of Venus. In: Exploring Venus as a Terrestrial Planet, ed. Esposito, L. W., et al., Washington, D.C.: AGU, pp. 121138.Google Scholar
Titov, D. V., et al. (2013). Radiative energy balance in the Venus atmosphere. In: Towards Understanding the Climate of Venus, ed. Bengtsson, L., et al., New York: Springer, pp. 2353.Google Scholar
Tobie, G., et al. (2006). Episodic outgassing as the origin of atmospheric methane on Titan. Nature 440, 6164.Google Scholar
Tokano, T. (2009). The dynamics of Titan’s troposphere. Phil. Trans. R. Soc. Lond. A 367, 633648.Google Scholar
Tokano, T. (2010). Relevance of fast westerlies at equinox for the eastward elongation of Titan’s dunes. Aeolian Res. 2, 113127.Google Scholar
Tokano, T. (2013). Wind-induced equatorial bulge in Venus and Titan general circulation models: Implication for the simulation of superrotation. Geophys. Res. Lett. 40, 45384543.Google Scholar
Tokar, R. L., et al. (2012). Detection of exospheric O2+ at Saturn’s moon Dione. Geophys. Res. Lett. 39.Google Scholar
Tolbert, N. E., et al. (1995). The oxygen and carbon-dioxide compensation points of C3 plants: Possible role in regulating atmospheric oxygen. P. Natl. Acad. Sci. USA 92, 11 23011 233.Google Scholar
Tolstikhin, I. N. and O’Nions, R. K. (1994). The Earths missing xenon: A combination of early degassing and of rare gas loss from the atmosphere. Chem. Geol. 115, 16.Google Scholar
Tomasko, M. G., et al. (2005). Rain, winds and haze during the Huygens probe’s descent to Titan’s surface. Nature 438, 765778.Google Scholar
Tomasko, M. G., et al. (2008). A model of Titan’s aerosols based on measurements made inside the atmosphere. Planet. Space Sci. 56, 669707.Google Scholar
Tomasko, M. G., et al. (2009). Limits on the size of aerosols from measurements of linear polarization in Titan’s atmosphere. Icarus 204, 271283.Google Scholar
Toner, J. D., et al. (2014a). The formation of supercooled brines, viscous liquids, and low-temperature perchlorate glasses in aqueous solutions relevant to Mars. Icarus 233, 3647.Google Scholar
Toner, J. D., et al. (2014b). Soluble salts at the Phoenix Lander site, Mars: A reanalysis of the Wet Chemistry Laboratory data. Geochim. Cosmochim. Acta 136, 142168.Google Scholar
Toner, J. D., et al. (2015). A revised Pitzer model for low-temperature soluble salt assemblages at the Phoenix site, Mars. Geochim. Cosmochim. Acta 166, 327343.Google Scholar
Tonks, W. B. and Melosh, H. J. (1993). Magma ocean formation due to giant impacts. J. Geophys. Res. 98, 53195333.Google Scholar
Toon, O. B. and Farlow, N. H. (1981). Particles above the tropopause: Measurements and models of stratospheric aerosols, meteoric debris, nacreous clouds, and noctilucent clouds. Annu. Rev. Earth Planet. Sci. 9, 1958.Google Scholar
Toon, O. B., et al. (1989). Rapid calculation of radiative heating rates and photodissociation rates in inhomogeneous multiple scattering atmospheres. J. Geophys. Res. 94, 16 28716 301.Google Scholar
Toon, O. B., et al. (2010). The formation of Martian river valleys by impacts. Annu. Rev. Earth. Pl. Sc. 38, 303322.Google Scholar
Toon, O. B., et al. (1982). The ultraviolet absorber on Venus: Amorphous sulfur. Icarus 51, 358373.Google Scholar
Toro, E. F. (1999). Riemann Solvers and Numerical Methods for Fluid Dynamics: A Practical Introduction. Berlin: Springer-Verlag.Google Scholar
Tosca, N. J. and Knoll, A. H. (2009). Juvenile chemical sediments and the long term persistence of water at the surface of Mars. Earth Planet. Sc. Lett. 286, 379386.Google Scholar
Touma, J. and Wisdom, J. (1993). The chaotic obliquity of Mars. Science 259, 12941297.Google Scholar
Touma, J. and Wisdom, J. (1998). Resonances in the early evolution of the Earth-Moon system. Astron. J. 115, 16531663.Google Scholar
Towe, K. M. (1981). Environmental conditions surrounding the origin and early Archean evolution of life: a hypothesis. Precambrian Res. 16, 110.Google Scholar
Trafton, L. (1980). Does Pluto have a substantial atmosphere. Icarus 44, 5361.Google Scholar
Trafton, L. M. (1966). Pressure-induced monochromatic translational absorption coefficients for homopolar and non-polar gases and gas mixtures with particular application to H2. Ap. J. 146, 558571.Google Scholar
Trafton, L. M. (1998). Planetary Atmospheres: The role of collision-induced absorption. In: Molecular Complexes in Earth’s Planetary, Cometary and Interstellar Atmospheres, ed. Vigasin, A. A. and Slanina, Z., London: World Scientific, pp. 177193.Google Scholar
Trafton, L. M., et al. (1997). Escape processes at Pluto and Charon. In: Pluto and Charon, ed. Stern, S. A. and Tholen, D. J., Tucson: University of Arizona Press, pp. 475522.Google Scholar
Trail, D., et al. (2007). Constraints on Hadean zircon protoliths from oxygen isotopes, Ti-thermometry, and rare earth elements. Geochem. Geophys. Geosys. 8, Q06014.Google Scholar
Trail, D., et al. (2011). The oxidation state of Hadean magmas and implications for early Earth’s atmosphere. Nature 480, 7983.Google Scholar
Trainer, M. G., et al. (2006). Organic haze on Titan and the early Earth. P. Natl. Acad. Sci. USA 103, 18 03518 042.Google Scholar
Trauger, J. T. and Lunine, J. I. (1983). Spectroscopy of molecular oxygen in the atmospheres of Venus and Mars. Icarus 55, 272281.Google Scholar
Treiman, A. H. and Irving, A. J. (2008). Petrology of Martian meteorite Northwest Africa 998. Meteorit. Planet. Sci. 43, 829854.Google Scholar
Trenberth, K. E., et al. (2009). Earth’s global energy budget. Bull. Am. Met. Soc. 90, 311323.Google Scholar
Trenberth, K. E. and Guillemot, C. J. (1994). The total mass of the atmosphere. J. Geophys. Res. 99, 23 07923 088.Google Scholar
Trenberth, K. E. and Stepaniak, D. P. (2003). Seamless poleward atmospheric energy transports and implications for the Hadley circulation. J. Climate 16, 37063722.Google Scholar
Trieloff, M., et al. (2003). The distribution of mantle and atmospheric argon in oceanic basalt glasses. Geochim. Cosmochim. Acta 67, 12291245.Google Scholar
Trindade, R. I. F., et al. (2003). Low-latitude and multiple geomagnetic reversals in the Neoproterozoic Puga cap carbonate, Amazon craton. Terra Nova 15, 441446.Google Scholar
Tsikos, H. and Moore, J. M. (1997). Petrography and geochemistry of the Paleoproterozoic Hotazel iron-formation, Kalahari manganese field, South Africa: Implications for Precambrian manganese metallogenesis. Econ. Geol. 92, 8797.Google Scholar
Tucker, O. J., et al. (2012). Thermally driven escape from Pluto’s atmosphere: A combined fluid/kinetic model. Icarus 217, 408415.Google Scholar
Tucker, O. J. and Johnson, R. E. (2009). Thermally driven atmospheric escape: Monte Carlo simulations for Titan’s atmosphere. Planet. Space Sci. 57, 18891894.Google Scholar
Tucker, O. J., et al. (2013). Diffusion and thermal escape of H2 from Titan’s atmosphere: Monte Carlo simulations. Icarus 222, 149158.Google Scholar
Turcotte, D. L. (1993). An episodic hypothesis for Venusian tectonics. J. Geophys. Res. 98, 17 06117 068.Google Scholar
Turner, E. C. and Bekker, A. (2015). Thick sulfate evaporite accumulations marking a mid-Neoproterozoic oxygenation event (Ten Stone Formation, Northwest Territories, Canada). GSA Bull., doi: 10.1130/B31268.1.Google Scholar
Turner, G. (1989). The outgassing history of the Earth’s atmosphere. J. Geol. Soc. London 146, 147154.Google Scholar
Turner, J., et al. (2009). Record low surface air temperature at Vostok station, Antarctica. J. Geophys. Res. 114, 114.Google Scholar
Turtle, E. P., et al. (2011a). Rapid and extensive surface changes near Titan’s equator: Evidence of April showers. Science 331, 14141417.Google Scholar
Turtle, E. P., et al. (2011b). Shoreline retreat at Titan’s Ontario Lacus and Arrakis Planitia from Cassini Imaging Science Subsystem observations. Icarus 212, 957959.Google Scholar
Turtle, E. P., et al. (2009). Cassini imaging of Titan’s high-latitude lakes, clouds, and south-polar surface changes. Geophys. Res. Lett. 36, doi: 10.1029/2008GL036186.Google Scholar
Tyburczy, J. A., et al. (1986). Shock-induced volatile loss from a carbonaceous chondrite – Implications for planetary accretion. Earth Planet. Sci. Lett. 80, 201207.Google Scholar
Tyler, S. A. and Barghoorn, E. S. (1954). Occurrence of structurally preserved plants in Pre-Cambrian rocks of the Canadian Shield. Science 119, 606608.Google Scholar
Tziperman, E., et al. (2011). Biologically induced initiation of Neoproterozoic snowball-Earth events. P. Natl. Acad. Sci. USA 108, 1509115096.Google Scholar
Ueno, Y., et al. (2009). Geological sulfur isotopes indicate elevated OCS in the Archean atmosphere, solving faint young sun paradox. P. Natl. Acad. Sci. USA 106, 14 78414 789.Google Scholar
Ueno, Y., et al. (2008). Quadruple sulfur isotope analysis of ca. 3.5 Ga Dresser Formation: New evidence for microbial sulfate reduction in the early Archean. Geochim. Cosmochim. Acta 72, 56755691.Google Scholar
Ueno, Y., et al. (2006). Evidence from fluid inclusions for microbial methanogenesis in the early Archaean era. Nature 440, 516519.Google Scholar
Unwin, S. C., et al. (2008). Taking the measure of the universe: Precision astrometry with SIM PlanetQuest. Publ. Astron. Soc. Pac. 120, 3888.Google Scholar
Usui, T., et al. (2012). Origin of water and mantle-crust interactions on Mars inferred from hydrogen isotopes and volatile element abundances of olivine-hosted melt inclusions of primitive shergottites. Earth Planet. Sci. Lett. 357, 119129.Google Scholar
Utsunomiya, S., et al. (2003). Iron oxidation state of a 2.45-Byr-old paleosol developed on mafic volcanics. Geochim. Cosmochim. Acta 67, 213221.Google Scholar
Vakoch, D. A. and Dowd, M. F. (2015). The Drake Equation: Estimating the Prevalence of Extraterrestrial Life Through the Ages. Cambridge: Cambridge University Press.Google Scholar
Valeille, A., et al. (2010). Water loss and evolution of the upper atmosphere and exosphere over martian history. Icarus 206, 2839.Google Scholar
Valentine, D. L. (2002). Biogeochemistry and microbial ecology of methane oxidation in anoxic environments: a review. Anton. Leeuw. Int. J. G. 81, 271282.Google Scholar
Vallis, G. K. (2006). Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. New York: Cambridge University Press.Google Scholar
van Berk, W. and Fu, Y. J. (2011). Reproducing hydrogeochemical conditions triggering the formation of carbonate and phyllosilicate alteration mineral assemblages on Mars (Nili Fossae region). J. Geophys. Res. 116.Google Scholar
van Berk, W., et al. (2012). Reproducing early Martian atmospheric carbon dioxide partial pressure by modeling the formation of Mg–Fe–Ca carbonate identified in the Comanche rock outcrops on Mars. J. Geophys. Res. 117.Google Scholar
Van Cappellen, P. and Ingall, E. D. (1996). Redox stabilization of the atmosphere and oceans by phosphorus-limited marine production. Science 271, 493496.Google Scholar
Van de Kamp, P. (1969). Parallax proper motion acceleration and orbital motion of Barnard’s Star. Astron. J. 74, 238240.Google Scholar
Van de Kamp, P. (1975). Astrometric study of Barnard’s star from plates taken with Sproul 61-cm refractor. Astron. J. 80, 658661.Google Scholar
van de Kamp, P. (1982). The planetary system of Barnard’s star. Vistas in Astronomy 26, 141157.Google Scholar
van den Boorn, S. H. J. M., et al. (2007). Dual role of seawater and hydrothermal fluids in Early Archean chert formation: Evidence from silicon isotopes. Geology 35, 939942.Google Scholar
Van Kranendonk, M. J. (2006). Volcanic degassing, hydrothermal circulation and the flourishing of early life on Earth: A review of the evidence from c. 3490–3240 Ma rocks of the Pilbara Supergroup, Pilbara Craton, Western Australia. Earth Sci. Rev. 74, 197240.Google Scholar
Van Kranendonk, M. J. (2011). Morphology as an indicator of biogenicity for 2.5-3.2 Ga fossil stromatolites from the Pilbara Craton, Western Australia. In: Advances in Stromatolite Geobiology, ed. Reitner, J., et al., Berlin: Springer-Verlag, pp. 537554.Google Scholar
Van Trump, J. E. and Miller, S. L. (1973). Carbon monoxide on the primitive Earth. Earth Planet. Sci. Lett. 20, 145150.Google Scholar
Van Valen, L. (1971). The history and stability of atmospheric oxygen. Science 171, 439443.Google Scholar
Van Zeggeren, F. and Storey, S. H. (1970). The Computation of Chemical Equilibria. New York: Cambridge University Press.Google Scholar
van Zuilen, M. A., et al. (2002). Reassessing the evidence for the earliest traces of life. Nature 418, 627630.Google Scholar
Vasavada, A. R., et al. (2012). Lunar equatorial surface temperatures and regolith properties from the Diviner Lunar Radiometer Experiment. J. Geophys. Res. 117, E00H18, doi:10.1029/2011JE003987.Google Scholar
Vasavada, A. R., et al. (1999). Near-surface temperatures on Mercury and the Moon and the stability of polar ice deposits. Icarus 141, 179193.Google Scholar
Vasavada, A. R. and Showman, A. P. (2005). Jovian atmospheric dynamics: an update after Galileo and Cassini. Rep. Prog. Phys. 68, 19351996.Google Scholar
Veeder, G. J., et al. (1994). Io’s heat flow from infrared radiometry: 1983–1993. J. Geophys. Res. 99, 17 09517 162.Google Scholar
Veizer, J. and Prokoph, A. (2015). Temperatures and oxygen isotopic composition of Phanerozoic oceans. Earth Sci. Rev. 146, 92104.Google Scholar
Vickery, A. M. and Melosh, H. J. (1990). Atmospheric erosion and impactor retention in large impacts, with application to mass extinctions. In: Global Catastrophes in Earth History. Geol. Soc. Sp. Paper 247, ed. Sharpton, V. L., Ward, P. D., Boulder, Colo.: Geol. Soc. Amer., pp. 289300.Google Scholar
Vidal-Madjar, A., et al. (2003). An extended upper atmosphere around the extrasolar planet HD209458b. Nature 422, 143146.Google Scholar
Vidal-Madjar, A., et al. (2004). Detection of oxygen and carbon in the hydrodynamically escaping atmosphere of the extrasolar planet HD 209458b. Astrophys. J. 604, L69L72.Google Scholar
Vidal-Madjar, A., et al. (2008). Exoplanet HD 209458b (Osiris(1)): Evaporation strengthened. Astrophys. J. Lett. 676, L57L60.Google Scholar
Viehmann, S., et al. (2014). Decoupled Hf-Nd isotopes in Neoarchean seawater reveal weathering of emerged continents. Geology 42, 115118.Google Scholar
Villanueva, G. L., et al. (2015). Strong water isotopic anomalies in the martian atmosphere: Probing current and ancient reservoirs. Science 348, 218221.Google Scholar
Vincendon, M. (2015). Identification of Mars gully activity types associated with ice composition. J. Geophys. Res. 120, doi:10.1002/2015JE004909.Google Scholar
Viola, D., et al. (2015). Expanded secondary craters in the Arcadia Planitia region, Mars: Evidence for tens of Myr-old shallow subsurface ice. Icarus 248, 190204.Google Scholar
Vogel, N., et al. (2011). Argon, krypton, and xenon in the bulk solar wind as collected by the Genesis mission. Geochim. Cosmochim. Acta 75, 30573071.Google Scholar
Voigt, A. and Abbot, D. S. (2012). Sea-ice dynamics strongly promote Snowball Earth initiation and destabilize tropical sea-ice margins. Climate of the Past 8, 20792092.Google Scholar
Voigt, A., et al. (2011). Initiation of a Marinoan Snowball Earth in a state-of-the-art atmosphere-ocean general circulation model. Climate of the Past 7, 249263.Google Scholar
Volkov, A. N. and Johnson, R. E. (2013). Thermal escape in the hydrodynamic regime: Reconsideration of Parker’s isentropic theory based on results of kinetic simulations. Astrophys. J. 765.Google Scholar
Volkov, A. N., et al. (2013). Expansion of monatomic and diatomic gases from a spherical source into a vacuum in a gravitational field. Fluid Dynam. 48, 239250.Google Scholar
Volkov, A. N., et al. (2011a). Thermally Driven Atmospheric Escape: Transition from Hydrodynamic to Jeans Escape. Astrophys. J. Lett. 729.Google Scholar
Volkov, A. N., et al. (2011b). Kinetic simulations of thermal escape from a single component atmosphere. Phys. Fluids 23, 066601.Google Scholar
Von Bloh, W., et al. (2003). Biogenic enhancement of weathering and the stability of the ecosphere. Geomicrobiol. J. 20, 501511.Google Scholar
Von Damm, K. L. (1990). Seafloor hydrothermal activity: black smoker chemistry and chimneys. Ann. Rev. Earth Planet. Sci. 18, 173204.Google Scholar
Von Damm, K. L. (1995). Controls on the chemistry and temporal variability of seafloor hydrothermal fluids. In: Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, ed. Humphris, S. E., et al., Washington DC: American Geophysical Union, pp. 222247.Google Scholar
Vuitton, V., et al. (2008). Formation and distribution of benzene on Titan. J. Geophys. Res. 113, E05007, doi:10.1029/2007JE002997.Google Scholar
Wacey, D. (2009). Early Life on Earth: A Practical Guide. New York: Springer.Google Scholar
Wacey, D., et al. (2011). Microfossils of sulphur-metabolizing cells in 3.4-billion-year-old rocks of Western Australia. Nat. Geosci. 4, 698702.Google Scholar
Wacey, D., et al. (2012). Taphonomy of very ancient microfossils from the similar to 3400 Ma Strelley Pool Formation and similar to 1900 Ma Gunflint Formation: New insights using a focused ion beam. Precambrian Res. 220, 234250.Google Scholar
Wachtershauser, G. (1988a). Before enzymes and templates: Theory of surface metabolism. Microbiol. Rev. 52, 452484.Google Scholar
Wachtershauser, G. (1988b). Pyrite formation, the first energy source for life - a hypothesis. Syst. Appl. Microbiol. 10, 207210.Google Scholar
Wachtershauser, G. (1990). Evolution of the first metabolic cycles. P. Natl. Acad. Sci. USA 87, 200204.Google Scholar
Wachtershauser, G. (1992). Groundworks for an evolutionary biochemistry - the iron sulfur world. Prog. Biophys. Mol. Biol. 58, 85201.Google Scholar
Wade, J. and Wood, B. J. (2005). Core formation and the oxidation state of the Earth. Earth Planet. Sc. Lett. 236, 7895.Google Scholar
Wadhwa, M. (2001). Redox state of Mars’ upper mantle and crust from Eu anomalies in Shergottite pyroxenes. Science 291, 15271530.Google Scholar
Wadhwa, M. (2008). Redox conditions on small bodies, the Moon and Mars. Rev. Mineral. Geochem. 68, 493510.Google Scholar
Waite, J. H., et al. (2009a). Liquid water on Enceladus from observations of ammonia and 40Ar in the plume. Nature 460, 487490.Google Scholar
Waite, J. H., et al. (2005). Ion Neutral Mass Spectrometer results from the first flyby of Titan. Science 308, 982986.Google Scholar
Waite, J. H., et al. (2007). The process of tholin formation in Titan’s upper atmosphere. Science 316, 870875.Google Scholar
Waite, J. H., et al. (2009b). High-altitude production of Titan’s aerosols. In: Titan from Cassini-Huygens, ed. Brown, R. H., et al., New York: Springer, pp. 201214.Google Scholar
Waldbauer, J. R., et al. (2009). Late Archean molecular fossils from the Transvaal Supergroup record the antiquity of microbial diversity and aerobiosis. Precam. Res. 169, 2847.Google Scholar
Walker, J. C. G. (1977). Evolution of the Atmosphere. New York: Macmillan.Google Scholar
Walker, J. C. G. (1978). Oxygen and hydrogen in the primitive atmosphere. Pure Appl. Geophys. 116, 222231.Google Scholar
Walker, J. C. G. (1980). The oxygen cycle. In: The Natural Environment and the Biogeochemical Cycles, ed. Hutzinger, O., Berlin: Springer-Verlag, pp. 87104.Google Scholar
Walker, J. C. G. (1982). The earliest atmosphere of the Earth. Precambrian Res. 17, 147171.Google Scholar
Walker, J. C. G. (1984). Suboxic diagenesis in banded iron formations. Nature 309, 340342.Google Scholar
Walker, J. C. G. (1985). Carbon dioxide on the early Earth. Orig. Life 16, 117127.Google Scholar
Walker, J. C. G. (1986). Impact erosion of planetary atmospheres. Icarus 68, 8798.Google Scholar
Walker, J. C. G. (1987). Was the Archaean biosphere upside down? Nature 329, 710712.Google Scholar
Walker, J. C. G. (1990). Precambrian evolution of the climate system. Palaeogeograph. Palaeoclimat. Palaeoecol. 82, 261289.Google Scholar
Walker, J. C. G. and Brimblecombe, P. (1985). Iron and sulfur in the pre-biologic ocean. Precambrian Res. 28, 205222.Google Scholar
Walker, J. C. G., et al. (1981). A negative feedback mechanism for the long-term stabilization of Earth’s surface temperature. J. Geophys. Res. 86, 97769782.Google Scholar
Walker, J. C. G. and Kasting, J. F. (1992). Effects of fuel and forest conservation on predicted levels of atmospheric carbon dioxide. Global Planet. Change 97, 151189.Google Scholar
Walker, J. C. G., et al. (1983). Environmental evolution of the Archean-Early Proterozoic Earth. In: Earth’s Earliest Biosphere: Its Origin and Evolution, ed. Schopf, J. W., Princeton, New Jersey: Princeton University Press, pp. 260290.Google Scholar
Walker, J. C. G. and Lohmann, K. C. (1989). Why the oxygen isotopic composition of seawater changes with time. Geophys. Res. Lett. 16, 323326.Google Scholar
Wall, S., et al. (2010). Active shoreline of Ontario Lacus, Titan: A morphological study of the lake and its surroundings. Geophys. Res. Lett. 37.Google Scholar
Wallace, J. M. and Hobbs, P. V. (2006). Atmospheric Science: An Introductory Survey. Burlington, MA: Elsevier Academic Press.Google Scholar
Wallace, M. W., et al. (2014). Enigmatic chambered structures in Cryogenian reefs: The oldest sponge-grade organisms? Precambrian Res. 255, 109123.Google Scholar
Wallmann, K. (2004). Impact of atmospheric CO2 and galactic cosmic radiation on Phanerozoic climate change and the marine delta 18-O record. Geol. Geochem. Geophys. 5.Google Scholar
Walsh, K. J., et al. (2011). A low mass for Mars from Jupiter’s early gas-driven migration. Nature 475, 206209.Google Scholar
Walsh, M. M. (1992). Microfossils and possible microfossils from the Early Archean Onverwacht Group, Barbeton Mountain Land, S. Africa. Precambrian Res. 52, 271293.Google Scholar
Walter, F. M. and Barry, D. C. (1991). Pre- and main-sequence evolution of solar activity. In: The Sun in Time, ed. Sonett, C. P., et al., Tucson, AZ: University of Arizona Press, pp. 633657.Google Scholar
Walter, M. (1995). Biogeochemistry – Fecal pellets in world events. Nature 376, 1617.Google Scholar
Wang, C. L., et al. (2014). Rare earth element and yttrium compositions of the Paleoproterozoic Yuanjiacun BIF in the Luliang area and their implications for the Great Oxidation Event (GOE). Sci. China Earth Sci. 57, 24692485.Google Scholar
Wang, J. and Fischer, D. A. (2015). Revealing a universal planet-metallicity correlation for planets of different solar-type stars. Astron. J. 149.Google Scholar
Wang, J. G., et al. (2012). Evolution from an anoxic to oxic deep ocean during the Ediacaran–Cambrian transition and implications for bioradiation. Chem. Geol. 306, 129138.Google Scholar
Wanke, H. (1981). Constitution of terrestrial planets. Phil. Trans. R. Soc. Lond. A 303, 287302.Google Scholar
Wänke, H. (2001). Geochemical evidence for a close genetic relationship of Earth and Moon. Earth Moon and Planets 856, 445452.Google Scholar
Wänke, H., et al. (2001). Chemical composition of rocks and soils at the Pathfinder site. Space Sci. Rev. 96, 317330.Google Scholar
Wänke, H. and Dreibus, G. (1994). Chemistry and accretion history of Mars. Phil. Trans. R. Soc. Lond. A 349, 285293.Google Scholar
Ward, P. D. B. D. (2000). Rare Earth: Why Complex Life is Uncommon in the Universe. New York: Copernicus.Google Scholar
Ward, W. R. (1974). Climatic variations on Mars. 1. Astronomical theory of insolation. J. Geophys. Res. 79, 33753386.Google Scholar
Ward, W. R. (1992). Long-term orbital and spin dynamics of Mars. In: Mars, ed. Kieffer, H. H., et al., Tucson: University of Arizona Press, pp. 298320.Google Scholar
Warneck, P. (2000). Chemistry of the Natural Atmosphere. New York: Academic Press.Google Scholar
Warren, S. G. and Brandt, R. E. (2006). Comment on “Snowball earth: A thin-ice solution with flowing sea glaciers” by David Pollard and James F. Kasting. J. Geophys. Res. 111, C09016, doi: 10.1029/2005JC003411.Google Scholar
Warren, S. G., et al. (2002). Snowball Earth: Ice thickness on the tropical ocean. Journal Of Geophysical Research-Oceans 107.Google Scholar
Watanabe, Y., et al. (2009). Anomalous fractionations of sulfur isotopes during thermochemical sulfate reduction. Science 324, 370373.Google Scholar
Watson, A. J., et al. (1984). Temperatures in a runaway greenhouse on the evolving Venus: Implications for water loss. Earth Planet. Sci. Lett. 68, 16.Google Scholar
Watson, A. J., et al. (1981). The dynamics of a rapidly escaping atmosphere: applications to the evolution of Earth and Venus. Icarus 48, 150166.Google Scholar
Wayne, R. P. (2000). Chemistry of Atmospheres: An Introduction to the Chemistry of the Atmospheres of Earth, the Planets, and their Satellites. Oxford: Clarendon Press.Google Scholar
Weaver, C. P. and Ramanathan, V. (1995). Deductions from a simple climate model: Factors governing surface temperature and atmospheric thermal structure. J. Geophys. Res. 100, 11 58511 591.Google Scholar
Weber, R. C., et al. (2011). Seismic detection of the lunar core. Science 331, 309312.Google Scholar
Webster, C. R., et al. (2015). Mars methane detection and variability at Gale crater. Science 347, 415417.Google Scholar
Webster, C. R., et al. (2013). Isotope ratios of H, C, and O in CO2 and H2O of the martian atmosphere. Science 341, 260263.Google Scholar
Wedepohl, K. H. (1995). The composition of the continental crust. Geochim. Cosmochim. Acta 59, 12171232.Google Scholar
Weidenschilling, S. J. (1977). Aerodynamics of solid bodies in solar nebula. Mon. Not. R. Astron. Soc. 180, 5770.Google Scholar
Weisberg, M. K., et al. (2006). Systematics and evaluation of meteorite classification. In: Meteorites and the Early Solar System II, ed. Lauretta, D. S. and McSween, H. Y., Tucson, AZ: University of Arizona Press, pp. 1952.Google Scholar
Weiss, B. P., et al. (2008). Paleointensity of the ancient Martian magnetic field. Geophys. Res. Lett. 35.Google Scholar
Weiss, J. W. (2004). Planetary Parameters. In: Jupiter: The Planet, Satellites and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press.Google Scholar
Weiss, L. M. and Marcy, G. W. (2014). The mass-radius relation for 65 exoplanets smaller than 4 Earth radii. Astrophys. J. Lett. 783, L6.Google Scholar
Weiss, M. C., et al. (2016). The physiology and habitat of the last universal common ancestor. Nature Microbiology 1, 16116 doi:10.1038/nmicrobiol.2016.116.Google Scholar
Weitz, C. M., et al. (2006). Soil grain analyses at Meridiani Planum, Mars. J. Geophys. Res. 111, E12S04, doi:10.1029/2005JE002541.Google Scholar
Weitz, C. M., et al. (2015). Mixtures of clays and sulfates within deposits in western Melas Chasma, Mars. Icarus 251, 291314.Google Scholar
Weitz, C. M., et al. (2010). Mars Reconnaissance Orbiter observations of light-toned layered deposits and associated fluvial landforms on the plateaus adjacent to Valles Marineris. Icarus 205, 73102.Google Scholar
Welhan, J. A. (1988). Origins of methane in hydrothermal systems. Chem. Geol. 71, 183198.Google Scholar
Wellman, C. H., et al. (2003). Fragments of the earliest land plants. Nature 425, 282285.Google Scholar
Wen, H. J., et al. (2015). Reconstruction of early Cambrian ocean chemistry from Mo isotopes. Geochim. Cosmochim. Acta 164, 116.Google Scholar
Werner, S. C. (2008). The early martian evolution – Constraints from basin formation ages. Icarus 195, 4560.Google Scholar
Werner, S. C. (2009). The global martian volcanic evolutionary history. Icarus 201, 4468.Google Scholar
Werner, S. C., et al. (2014). The source crater of martian shergottite meteorites. Science 343, 13431346.Google Scholar
Werner, S. C. and Tanaka, K. L. (2011). Redefinition of the crater-density and absolute-age boundaries for the chronostratigraphic system of Mars. Icarus 215, 603607.Google Scholar
West, R. A., et al. (2004). Jovian clouds and haze. In: Jupiter: The Planet, Satellites, and Magnetosphere, ed. Bagenal, F., et al., New York: Cambridge University Press, pp. 79104.Google Scholar
West, R. A., et al. (2009). Clouds and aerosols in Saturn’s atmosphere. In: Saturn from Cassini-Huygens, ed. Dougherty, M., et al., New York: Springer, pp. 161180.Google Scholar
West, R. A., et al. (1991). Clouds and aerosols in the Uranian atmosphere. In: Uranus, ed. Bergstralh, J. T., et al., Tucson: University of Arizona Press, pp. 296324.Google Scholar
West, R. A., et al. (2016). Cassini imaging science subsystem observations of Titan’s south polar cloud. Icarus 270, 399408.Google Scholar
Westall, F. (2005). Life on the early Earth: A sedimentary view. Science 308, 366367.Google Scholar
Westall, F., et al. (2001). Early Archean fossil bacteria and biofilms in hydrothermally-influenced sediments from the Barberton greenstone belt, South Africa. Precambrian Res. 106, 93116.Google Scholar
Whelley, P. L. and Greeley, R. (2008). The distribution of dust devil activity on Mars. J. Geophys. Res. 113.Google Scholar
Whewell, W. (1853). Of the Plurality of Worlds: An Essay. London: John W. Parker and Son.Google Scholar
Whipple, F. L. (1972). On certain aerodynamic processes for asteroids and comets. In: From Plasma to Planet; Proceedings, ed. Elvius, A., New York: Wiley, pp. 211232.Google Scholar
Whitehill, A. R., et al. (2015). SO2 photolysis as a source for sulfur mass-independent isotope signatures in stratospehric aerosols. Atmos. Chem. Phys. 15, 18431864.Google Scholar
Widdel, F., et al. (1993). Ferrous iron oxidation by anoxygenic phototrophic bacteria. Nature 362, 834836.Google Scholar
Wiens, R. C., et al. (1986). The case for a martian origin of the shergottites, 2. Trapped and Indigenous gas components in EETA 79001 glass. Earth Planet. Sc. Lett. 77, 149158.Google Scholar
Wiens, R. C. and Pepin, R. O. (1988). Laboratory shock emplacement of noble gases, nitrogen, and carbon dioxide into basalt, and implications for trapped gases in shergottite EETA-79001. Geochim. Cosmochim. Acta 52, 295307.Google Scholar
Wiktorowicz, S. J. and Ingersoll, A. P. (2007). Liquid water oceans in ice giants. Icarus 186, 436447.Google Scholar
Wilde, S. A., et al. (2001). Evidence from detrital zircons for the existence of continental crust and oceans on Earth 4.4 Gyr ago. Nature 409, 175178.Google Scholar
Wildman, R. A., et al. (2004). Burning of forest materials under late Paleozoic high atmospheric oxygen levels. Geology 32, 457460.Google Scholar
Wilhelms, D. E. and Squyres, S. W. (1984). The martian hemispheric dichotomy may be due to a giant Impact. Nature 309, 138140.Google Scholar
Wille, M., et al. (2007). Evidence for a gradual rise of oxygen between 2.6 and 2.5 Ga from Mo isotopes and Re-PGE signatures in shales. Geochim. Cosmochim. Acta 71, 24172435.Google Scholar
Williams, D. A., et al. (2005). Erosion by flowing Martian lava: New insights for Hecates Tholus from Mars Express and MER data. J. Geophys. Res. 110.Google Scholar
Williams, G. E. (1975). Late Precambrian glacial climate and the Earth’s obliquity. Geol. Mag. 112, 441465.Google Scholar
Williams, G. E. (1981). Sunspot Periods in the Late Precambrian Glacial Climate and solar-planetary relations. Nature 291, 624628.Google Scholar
Williams, G. E. (1989). Late Precambrian tidal rhythmites in South Australia and the history of the Earth’s rotation. J. Geol. Soc. 146, 97111.Google Scholar
Williams, G. E. (1997). Precambrian length of day and the validity of tidal rhythmite paleotidal values. Geophys. Res. Lett 24, 421424.Google Scholar
Williams, G. E. (2000). Geological constraints on the Precambrian history of earth’s rotation and the moon’s orbit. Rev. Geophys. 38, 3759.Google Scholar
Williams, G. E. (2005). Subglacial meltwater channels and glaciofluvial deposits in the Kimberley Basin, Western Australia: 1.8 Ga low-latitude glaciation coeval with continental assembly. Journal of the Geological Society 162, 111124.Google Scholar
Williams, G. P. (1979). Planetary circulations. 2. Jovian quasi-geostrophic regime. J. Atmos. Sci. 36, 932968.Google Scholar
Williams, G. P. (2003). Jovian dynamics. Part III: Multiple, migrating, and equatorial jets. J. Atmos. Sci. 60, 12701296.Google Scholar
Williams, K. E., et al. (2009). Ancient melting of mid-latitude snowpacks on Mars as a water source for gullies. Icarus 200, 418425.Google Scholar
Williams, R. G. and Follows, M. (2011). Ocean Dynamics and the Carbon Cycle: Principles and Mechanisms. New York: Cambridge University Press.Google Scholar
Williams, R. M., et al. (2000). Flow rates and duration within Kasei Valles, Mars: Implications for the formation of a martian ocean. Geophys. Res. Lett. 27, 10731076.Google Scholar
Wills, C. and Bada, J. (2000). The Spark of Life : Darwin and the Primeval Soup. Cambridge, Mass.: Perseus Pub.Google Scholar
Willson, L. A., et al. (1987). Mass loss on the main sequence. Comments Astrophys. 12, 1734.Google Scholar
Wilson, E. H. and Atreya, S. K. (2004). Current state of modeling the photochemistry of Titan’s mutually dependent atmosphere and ionosphere. J. Geophys. Res. 109, E06002, doi:10.1029/2003JE002181.Google Scholar
Wilson, E. H. and Atreya, S. K. (2009). Titan’s Carbon Budget and the Case of the Missing Ethane. J. Phys. Chem. A 113, 11 22111 226.Google Scholar
Wilson, H. F. and Militzer, B. (2010). Sequestration of noble gases in giant planet interiors. Phys. Rev. Lett. 104, 121101.Google Scholar
Wilson, R. J., et al. (2002). Traveling waves in the Northern Hemisphere of Mars. Geophys. Res. Lett. 29, 1684, doi:10.1029/2002GL014866.Google Scholar
Winn, J. N. and Fabrycky, D. C. (2015). The occurrence and architecture of exoplanetary systems. Annu. Rev. Astron. Astrophys. 53, 409447.Google Scholar
Withers, P. and Catling, D. (2010). Observations of atmospheric tides on Mars at the season and latitude of the Phoenix atmospheric entry. Geophys. Res. Lett 37, L24204.Google Scholar
Woese, C. R. (2005). Evolving biological organization. In: Microbial Phylogeny and Evolution: Concepts and Controversies, ed. Sapp, J., Oxford: Oxford University Press, pp. 99117.Google Scholar
Wolery, T. J. and Sleep, N. H. (1976). Hydrothermal circulation and geochemical flux at midocean ridges. J. Geol. 84, 249275.Google Scholar
Wolf, E. T. and Toon, O. B. (2010). Fractal organic hazes provided an ultraviolet shield for early Earth. Science 328, 12661268.Google Scholar
Wolf, E. T. and Toon, O. B. (2013). Hospitable Archean climates simulated by a General Circulation Model. Astrobiology 13, 656673.Google Scholar
Wolf, E. T. and Toon, O. B. (2014). Delayed onset of runaway and moist greenhouse climates for Earth. Geophys. Res. Lett. 41, 167172.Google Scholar
Wolf, E. T. and Toon, O. B. (2015). The evolution of habitable climates under the brightening Sun. J. Geophys. Res. 120, 57755794.Google Scholar
Wones, D. R. and Gilbert, M. C. (1969). The fayalite–magnetite–quartz assemblage between 600 º and 800 ºC. Amer. J. Sci. 267 -A, 480488.Google Scholar
Wood, B. E., et al. (2002). Measured mass loss rates of solar-like stars as a function of age and activity. Ap. J. 574, 412425.Google Scholar
Wood, B. E., et al. (2005). New mass-loss measurements from astrospheric Ly alpha absorption. Astrophys. J. 628, L143L146.Google Scholar
Wood, B. J. and Halliday, A. N. (2010). The lead isotopic age of the Earth can be explained by core formation alone. Nature 465, 767–U4.Google Scholar
Wood, B. J., et al. (1996). Water and carbon in Earth’s mantle. Phil. Trans. R. Soc. Lond. A. 354, 14951511.Google Scholar
Wood, B. J. and Virgo, D. (1989). Upper mantle oxidation state: ferric iron contents of lherzolite spinels by 57Fe Mossbauer spectroscopy and resultant oxygen fugacities. Geochim. Cosmochim. Acta 53, 12771291.Google Scholar
Wood, B. J., et al. (2006). Accretion of the Earth and segregation of its core. Nature 441, 825833.Google Scholar
Woodland, A. B. and Koch, M. (2003). Variation in oxygen fugacity with depth in the upper mantle beneath the Kaapvaal craton, Southern Africa. Earth Planet. Sci. Lett. 214, 295310.Google Scholar
Woolf, N. J., et al. (2002). The spectrum of earthshine: A pale blue dot observed from the ground. Astrophys. J. 574, 430433.Google Scholar
Wordsworth, R., et al. (2010). Infrared collision-induced and far-line absorption in dense CO2 atmospheres. Icarus 210, 992997.Google Scholar
Wordsworth, R., et al. (2013). Global modelling of the early martian climate under a denser CO2 atmosphere: Water cycle and ice evolution. Icarus 222, 119.Google Scholar
Wordsworth, R. and Pierrehumbert, R. (2013). Hydrogen–nitrogen greenhouse warming in Earth’s early atmosphere. Science 339, 6467.Google Scholar
Wordsworth, R. and Pierrehumbert, R. (2014). Abiotic oxygen-dominated atmospheres on terrestrial habitable zone planets. Astrophys. J. Lett. 785, L20.Google Scholar
Wordsworth, R., et al. Transient reducing greenhouse warming on early Mars. Geophys. Res. Lett.; doi:10.1002/2016GL071766, submitted.Google Scholar
Wray, J. J., et al. (2009). Diverse aqueous environments on ancient Mars revealed in the southern highlands. Geology 37, 10431046.Google Scholar
Wray, J. J., et al. (2016). Orbital evidence for more widespread carbonate-bearing rocks on Mars. J. Geophys. Res. 121, 652677.Google Scholar
Wright, J. T. and Gaudi, B. S. (2013). Exoplanet detection methods. In: Planets, Stars and Stellar Systems, ed. Oswalt, T., et al.: Springer Netherlands, pp. 489540.Google Scholar
Wuebbles, D. J. and Hayhoe, K. (2002). Atmospheric methane and global change. Earth Sci. Rev. 57, 177210.Google Scholar
Xiao, S. (2014). Oxygen and early animal evolution. In: Treatise on Geochemistry, ed. Holland, H. D. and Turekian, K. K., New York: Elsevier, pp. 231250.Google Scholar
Xiao, S. H. and Laflamme, M. (2009). On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota. Trends Ecol. Evol. 24, 3140.Google Scholar
Yamamoto, M. and Takahashi, M. (2009). Dynamical effects of solar heating below the cloud layer in a Venus-like atmosphere. J. Geophys. Res. 114.Google Scholar
Yan, F., et al. (2015). High-resolution transmission spectrum of the Earth’s atmosphere-seeing Earth as an exoplanet using a lunar eclipse. Int. J. Astrobiol. 14, 255266.Google Scholar
Yang, J., et al. (2014a). Strong dependence of the inner edge of the habitable zone on planetary rotation rate. Astrophys. J. Lett. 787, L2.Google Scholar
Yang, J., et al. (2013). Stabilizing cloud feedback dramatically expands the habitable zone of tidally locked planets. Astrophys. J. Lett. 771, L45, doi:10.1088/2041-8205/771/2/L45.Google Scholar
Yang, J., et al. (2012a). The initiation of modern “Soft Snowball” and “Hard Snowball” climates in CCSM3. Part I: The influences of solar luminosity, CO2 concentration, and the sea ice/snow albedo parameterization. J. Climate 25, 27112736.Google Scholar
Yang, J., et al. (2012b). The initiation of modern “Soft Snowball” and “Hard Snowball” climates in CCSM3. Part II: Climate dynamic feedbacks. J. Climate 25, 27372754.Google Scholar
Yang, W. and Holland, H. D. (2002). The redox sensitive trace elements Mo, U and Re in Precambrian carbonaceous shales: Indicators of the Great Oxidation Event (abstr.). Geol. Soc. Am. Ann. Mtng 34, 382.Google Scholar
Yang, X., et al. (2014b). A relatively reduced Hadean continental crust and implications for the early atmosphere and crustal rheology. Earth Planet. Sc. Lett. 393, 210219.Google Scholar
Yang, X. Z., et al. (2014c). A relatively reduced Hadean continental crust and implications for the early atmosphere and crustal rheology. Earth Planet. Sc. Lett. 393, 210219.Google Scholar
Yelle, R. V., et al. (2008). Methane escape from Titan’s atmosphere. J. Geophys. Res. 113.Google Scholar
Yelle, R. V., et al. (2001). Structure of the Jovian stratosphere at the Galileo probe entry site. Icarus 152, 331346.Google Scholar
Yelle, R. V., et al. (1995). Lower atmospheric structure and surface-atmosphere interactions on Triton. In: Neptune and Triton, ed. Cruikshank, D. P., Tucson: University of Arizona Press, pp. 10311105.Google Scholar
Yen, A. S., et al. (2005). An integrated view of the chemistry and mineralogy of martian soils. Nature 436, 4954.Google Scholar
Yen, A. S., et al. (2008). Hydrothermal processes at Gusev Crater: An evaluation of Paso Robles class soils. J. Geophys. Res. 113, E06S10, doi:10.1029/2007JE002978.Google Scholar
Yeo, G. M. (1981). The Late Proterozoic Rapitan glaciation in the northern Cordillera. In: Proterozoic Basins of Canada, ed. Campbell, F. H.: Geological Survey of Canada Paper 81–10, pp. 2546.Google Scholar
Yin, L. M., et al. (2007). Doushantuo embryos preserved inside diapause egg cysts. Nature 446, 661663.Google Scholar
Yoder, C. F. (1997). Venus spin dynamics. In: Venus II, ed. Bougher, S. W., et al., Tucson: University of Arizona Press, pp. 10871124.Google Scholar
Young, G. M. (1991). Stratigraphy, Sedimentology, and Tectonic Setting of the Huronian Supergroup. Toronto: G. A. Canada, 34 pp.Google Scholar
Young, G. M. (2002). Stratigraphic and tectonic settings of Proterozoic glaciogenic rocks and Banded Iron-Formations: Relevance to the Snowball Earth debate. J. African Earth Sci. 35, 451466.Google Scholar
Young, G. M. and Gostin, V. A. (1989). An exceptionally thick Upper Proterozoic (Sturtian) glacial succession in the Mount Painter Area, South Australia. Geol. Soc. Am. Bull. 101, 834845.Google Scholar
Young, G. M. and Nesbitt, H. W. (1999). Paleoclimatology and provenance of the glaciogenic Gowganda Formation (Paleoproterozoic), Ontario, Canada: A chemostratigraphic approach. Geol. Soc. Am. Bull. 111, 264274.Google Scholar
Young, G. M., et al. (1998). Earth’s oldest reported glaciation; physical and chemical evidence from the Archean Mozaan Group (~ 2.9 Ga) of South Africa. J. Geol. 106, 523538.Google Scholar
Young, L. A., et al. (1997). Detection of gaseous methane on Pluto. Icarus 127, 258262.Google Scholar
Young, L. D. G. (1971). Interpretation of high resolution spectra of Mars .2. Calculations of CO2 abundance, rotational temperature and surface pressure. J. Quant. Spectrosc. Ra. Trans. 11, 1075.Google Scholar
Young, R. A. and Crow, R. (2014). Paleogene Grand Canyon incompatible with Tertiary paleogeography and stratigraphy. Geosphere 10, 664679.Google Scholar
Yuan, X. L., et al. (2005). Lichen-like symbiosis 600 million years ago. Science 308, 10171020.Google Scholar
Yung, Y. L., et al. (1984). Photochemistry of the atmosphere of Titan: comparison between model and observations. Ap. J. Supp. 55, 465506.Google Scholar
Yung, Y. L. and DeMore, W. B. (1982). Photochemistry of the stratosphere of Venus: Implications for atmospheric evolution. Icarus 51, 199247.Google Scholar
Yung, Y. L. and DeMore, W. B. (1999). Photochemistry of Planetary Atmospheres. New York: Oxford University Press.Google Scholar
Yung, Y. L. and McElroy, M. B. (1977). Stability of an oxygen atmosphere on Ganymede. Icarus 30, 97103.Google Scholar
Yung, Y. L., et al. (1989). Hydrogen and deuterium loss from the terrestrial atmosphere: A quantitative assessment of nonthermal escape fluxes. J. Geophys. Res. 94, 14 97114 989.Google Scholar
Yung, Y. L., et al. (1988). HDO in the martian atmosphere: Implications for the abundance of crustal water. Icarus 76, 146159.Google Scholar
Zachos, J. C., et al. (2008). An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 451, 279283.Google Scholar
Zahnle, K. (1993a). Planetary noble gases. In: Protostars and Protoplanets II, ed. Black, D. C. and Matthews, M. S., Tucson, AZ: University of Arizona Press, pp. 13051338.Google Scholar
Zahnle, K. (2015). Play it again, SAM. Science 347, 370371.Google Scholar
Zahnle, K., et al. (2007). Emergence of a habitable planet. Space Sci. Rev. 129, 3578.Google Scholar
Zahnle, K., et al. (2006). The loss of mass-independent fractionation in sulfur due to a Palaeoproterozoic collapse of atmospheric methane. Geobiology 4, 271283.Google Scholar
Zahnle, K., et al. (1998). Cratering rates on the Galilean satellites. Icarus 136, 202222.Google Scholar
Zahnle, K., et al. (2011). Is there methane on Mars? Icarus 212, 493503.Google Scholar
Zahnle, K., et al. (2008). Photochemical instability of the ancient Martian atmosphere. J. Geophys. Res. 113, E11004.Google Scholar
Zahnle, K., et al. (1990). Mass fractionation of noble gases in diffusion-limited hydrodynamic hydrogen escape. Icarus 84, 502527.Google Scholar
Zahnle, K., et al. (1992). Impact-generated atmospheres over Titan, Ganymede, and Callisto. Icarus 95, 123.Google Scholar
Zahnle, K., et al. (2010). Earth’s earliest atmospheres. CSH Perspect. Biol. 2, doi: 10.1101/cshperspect.a004895.Google Scholar
Zahnle, K. J. (1986). Photochemistry of methane and the formation of hydrocyanic acid (HCN) in the Earth’s early atmosphere. J. Geophys. Res. 91, 28192834.Google Scholar
Zahnle, K. J. (1993b). Xenological constraints on the impact erosion of the early Martian atmosphere. J. Geophys. Res. 98, 10 89910 913.Google Scholar
Zahnle, K. J. (1998). Origins of Atmospheres. In: Origins, ed. Woodward, C. E., et al., San Francisco: Astron. Soc. Pacific, pp. 364391.Google Scholar
Zahnle, K. J. (2000). Hydrodynamic escape of ionized xenon from ancient atmospheres. Bull. Am. Astron. Soc. 32, 1044.Google Scholar
Zahnle, K. J., et al. (2013). The rise of oxygen and the hydrogen hourglass. Chem. Geol. 362, 2634.Google Scholar
Zahnle, K. J. and Kasting, J. F. (1986). Mass fractionation during transonic escape and implications for loss of water from Mars and Venus. Icarus 68, 462480.Google Scholar
Zahnle, K. J., et al. (1988). Evolution of a steam atmosphere during Earth’s accretion. Icarus 74, 6297.Google Scholar
Zahnle, K. J. and Sleep, N. (1997). Impacts and the early evolution of life. In: Comets and the Origin and Evolution of Life, ed. Thomas, P. J., et al., New York: Springer, pp. 175208.Google Scholar
Zahnle, K. J. and Sleep, N. H. (2002). Carbon dioxide cycling through the mantle and its implications for the climate of the ancient Earth. Geol. Soc. Lond. Sp. Publ. 199, 231257.Google Scholar
Zahnle, K. J. and Walker, J. C. G. (1982). The evolution of solar ultraviolet luminosity. Rev. Geophys. Space Phys. 20, 280292.Google Scholar
Zalucha, A. M., et al. (2011). An analysis of Pluto occultation light curves using an atmospheric radiative-conductive model. Icarus 211, 804818.Google Scholar
Zasova, L. V., et al. (1981). Vertical distribution of SO2 in upper cloud layer of Venus and origin of U.V.-absorption. Adv. Space Res. 1, 1316.Google Scholar
Zdunkowski, W., et al. (2007). Radiation in the Atmosphere: A Course in Theoretical Meteorology. Cambridge: Cambridge University Press.Google Scholar
Zebker, H., et al. (2014). Surface of Ligeia Mare, Titan, from Cassini altimeter and radiometer analysis. Geophy. Res. Lett. 41, 308313.Google Scholar
Zehr, J. P., et al. (1995). Diversity of heterotrophic nitrogen-fixation genes in a marine cyanobacterial mat. Appl. Environ. Microbiol. 61, 25272532.Google Scholar
Zelinka, M. D. and Hartmann, D. L. (2011). The observed sensitivity of high clouds to mean surface temperature anomalies in the tropics. J. Geophys. Res. 116.Google Scholar
Zellem, R. T., et al. (2014). The 4.5 μm full orbit phase curve of the hot Jupiter HD 209458b. Astrophys. J. 790, doi:10.1088/0004-637x/790/1/53.Google Scholar
Zeng, X. (2010). What is the atmosphere’s effect on Earth’s surface temperature. EOS Trans. AGU 91, 134135.Google Scholar
Zent, A. P. and Fanale, F. P. (1986). Possible Mars brines: Equilibrium and kinetic considerations. J. Geophys. Res. 91, D439D445.Google Scholar
Zent, A. P. and McKay, C. P. (1994). The chemical reactivity of the Martian soil and implications for future missions. Icarus 108, 146157.Google Scholar
Zent, A. P. and Quinn, R. C. (1995). Simultaneous adsorption of CO2 and H2O under Mars-like conditions and application to the evolution of the martian climate. J. Geophys. Res. 100, 53415349.Google Scholar
Zerkle, A. L., et al. (2012). A bistable organic-rich atmosphere on the Neoarchaean Earth. Nat. Geosci. 5, 359363.Google Scholar
Zhang, X., et al. (2012). Sulfur chemistry in the middle atmosphere of Venus. Icarus 217, 714739.Google Scholar
Zhang, X., et al. (2013). Radiative forcing of the stratosphere of Jupiter, Part I: Atmospheric cooling rates from Voyager to Cassini. Planet. Space Sci. 88, 325.Google Scholar
Zhang, X. L., et al. (2014). Triggers for the Cambrian explosion: Hypotheses and problems. Gondwana Res. 25, 896909.Google Scholar
Zhu, S. and Chen, H. (1995). Megascopic multicellular organisms from the 1700-million-year-old Tuanshanzi Formation in the Jixian area, North China. Science 270, 620622.Google Scholar
Ziering, S. and Hu, P. N. (1967). Thermal escape from planetary atmospheres in presence of a gravitational field. Astronaut. Acta 13, 327340.Google Scholar
Ziering, S., et al. (1968). Thermal escape problem. 2. Transition domain in spherical geometry. Phys. Fluids 11, 13271334.Google Scholar
Zimmer, C., et al. (2000). Subsurface oceans on Europa and Callisto: Constraints from Galileo magnetometer observations. Icarus 147, 329347.Google Scholar
Zinder, S. H. (1993). Physiological ecology of methanogens. In Methanogenesis: Ecology, Physiology, Biochemistry and Genetics, ed. Ferry, J. G., New York: Chapman and Hall, pp. 128206.Google Scholar
Zmolek, P., et al. (1999). Large mass independent sulfur isotope fractionations during the photopolymerization of (CS2)-C-12 and (CS2)-C-13. J. Phys. Chem. A 103, 24772480.Google Scholar
Zolotov, M. Y. and Shock, E. L. (2001). Composition and stability of salts on the surface of Europa and their oceanic origin. J. Geophys. Res. 106, 32 81532 827.Google Scholar
Zsom, A., et al. (2013). Toward the minimum inner edge distance of the habitable zone. Astrophys. J. 778, 109, doi: 10.1088/0004-637X/778/2/109.Google Scholar
Zuber, M. T., et al. (2007). Density of Mars’ south polar layered deposits. Science 317, 17181719.Google Scholar
Zugger, M. E., et al. (2010). Light scattering from exoplanet oceans and atmospheres. Astrophys. J. 723, 11681179.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Bibliography
  • David C. Catling, University of Washington, James F. Kasting, Pennsylvania State University
  • Book: Atmospheric Evolution on Inhabited and Lifeless Worlds
  • Online publication: 04 May 2017
  • Chapter DOI: https://doi.org/10.1017/9781139020558.020
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Bibliography
  • David C. Catling, University of Washington, James F. Kasting, Pennsylvania State University
  • Book: Atmospheric Evolution on Inhabited and Lifeless Worlds
  • Online publication: 04 May 2017
  • Chapter DOI: https://doi.org/10.1017/9781139020558.020
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Bibliography
  • David C. Catling, University of Washington, James F. Kasting, Pennsylvania State University
  • Book: Atmospheric Evolution on Inhabited and Lifeless Worlds
  • Online publication: 04 May 2017
  • Chapter DOI: https://doi.org/10.1017/9781139020558.020
Available formats
×