Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-17T21:03:49.567Z Has data issue: false hasContentIssue false

From climate models to planetary habitability: temperature constraints for complex life

Published online by Cambridge University Press:  25 July 2016

Laura Silva*
Affiliation:
National Institute for Astrophysics, INAF-OATs, Trieste, Italy
Giovanni Vladilo
Affiliation:
National Institute for Astrophysics, INAF-OATs, Trieste, Italy
Patricia M. Schulte
Affiliation:
Department of Zoology, University of British Columbia, Vancouver, BC, Canada
Giuseppe Murante
Affiliation:
National Institute for Astrophysics, INAF-OATs, Trieste, Italy
Antonello Provenzale
Affiliation:
Institute of Geosciences and Earth Resources, CNR, Pisa, Italy

Abstract

In an effort to derive temperature-based criteria of habitability for multicellular life, we investigated the thermal limits of terrestrial poikilotherms, i.e. organisms whose body temperature and the functioning of all vital processes is directly affected by the ambient temperature. Multicellular poikilotherms are the most common and evolutionarily ancient form of complex life on earth. The thermal limits for the active metabolism and reproduction of multicellular poikilotherms on earth are approximately bracketed by the temperature interval 0°C ≤ T ≤ 50°C. The same interval applies to the photosynthetic production of oxygen, an essential ingredient of complex life, and for the generation of atmospheric biosignatures observable in exoplanets. Analysis of the main mechanisms responsible for the thermal thresholds of terrestrial life suggests that the same mechanisms would apply to other forms of chemical life. We therefore propose a habitability index for complex life, h050, representing the mean orbital fraction of planetary surface that satisfies the temperature limits 0°C ≤ T ≤ 50°C. With the aid of a climate model tailored for the calculation of the surface temperature of Earth-like planets, we calculated h050 as a function of planet insolation, S, and atmospheric columnar mass, Natm, for a few earth-like atmospheric compositions with trace levels of CO2. By displaying h050 as a function of S and Natm, we built up an atmospheric mass habitable zone (AMHZ) for complex life. At variance with the classic habitable zone, the inner edge of the complex life habitable zone is not affected by the uncertainties inherent to the calculation of the runaway greenhouse limit. The complex life habitable zone is significantly narrower than the habitable zone of dry planets. Our calculations illustrate how changes in ambient conditions dependent on S and Natm, such as temperature excursions and surface dose of secondary particles of cosmic rays, may influence the type of life potentially present at different epochs of planetary evolution inside the AMHZ.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2016 

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

Abe, Y., Abe-Ouchi, A., Sleep, N.H. & Zahnle, K.J. (2011). Habitable zone limits for dry planets. Astrobiology 11, 443460.Google Scholar
Addo-Bediako, A., Chown, S. L. & Gaston, K. J. (2000). Thermal tolerance, climatic variability and latitude. Proc Biol Sci. 267(1445), 739745.CrossRefGoogle ScholarPubMed
Arrhenius, S. (1915). Quantitative Laws in Biological Chemistry. Bell, London.Google Scholar
Atri, D., Hariharan, B. & Grießmeier, J.-M. (2013). Galactic cosmic ray-induced radiation dose on terrestrial exoplanets. Astrobiology 13, 910919.Google Scholar
Atri, D. & Melott, A. L. (2014). Cosmic Rays and Terrestrial Life: a Brief Review. Astropart. Phys. 53, 186190.Google Scholar
Bartik, K., Bruylants, G., Locci, E. & Reisse, J. (2011). Liquid water: a necessary condition for all forms of life? In Origins and Evolution of Life: an Astrobiological Perspective, ed. Gargaud, M., López-García, P. & Martin, H., pp. 205217. Cambridge Univ. Press, Cambridge, UK.CrossRefGoogle Scholar
Batalha, N.M. et al. (2013). Planetary candidates observed by Kepler. III. Analysis of the first 16 months of data. Astrophys. J. Suppl. 204, 24 (21 pp).Google Scholar
Berg, J., Tymoczko, J.L. & Stryer, L. (2007). Biochemistry, 6th edn. W. H. Freeman and Co., New York.Google Scholar
Bressan, A., Marigo, P., Girardi, L., Salasnich, B., Dal Cero, C., Rubele, S. & Nanni, A. (2012). PARSEC: stellar tracks and isochrones with the PAdova and TRieste stellar evolution code. M.N.R.A.S. 427, 127.Google Scholar
Budisa, N. & Schulze-Makuch, D. (2014). Supercritical carbon dioxide and its potential as a life-sustaining solvent in a planetary environment. Life 4, 331340.Google Scholar
Bustamante, C., Keller, D. & Oster, G. (2001). The physics of molecular motors. Acc. Chem. Res. 34, 412420.Google Scholar
Caldeira, K. & Kasting, J.F. (1992). The life span of the biosphere revisited. Nature 360, 721723.CrossRefGoogle ScholarPubMed
Carigi, L., García-Rojas, J. & Meneses-Goytia, S. (2013). Chemical evolution and the galactic habitable zone of M31. Rev. Mex. Astron. Astrofís. 49, 253273.Google Scholar
Carr, A.G., Mammucarib, R. & Fosterb, N.R. (2011). A review of subcritical water as a solvent and its utilisation for the processing of hydrophobic organic compounds. Chem. Eng. J. 172, 117.Google Scholar
Catling, D.C., Glein, C.R., Zahnle, K.J. & McKay, C.P. (2005). Why O2 is required by complex life on habitable planets and the concept of planetary “Oxygenation Time”. Astrobiology 5(3), 415438.Google Scholar
Cavicchioli, R. & Thomas, T. (2003). Extremophiles. In The Desk Encyclopedia of Microbiology, ed. Schaechter, M., pp. 436453. Elsevier, London, UK.Google Scholar
Chyba, C.F. & Hand, K.P. (2005). Astrobiology: the study of the living universe. Ann. Rev. Astron. Astrophys. 43, 3174.Google Scholar
Clarke, A. (2014). The thermal limits to life on Earth. Int. J. Astrobiology 13(2), 141154.CrossRefGoogle Scholar
Cleaves, H.J. & Chalmers, J.H. (2004). Extremophiles may be irrelevant to the origin of life. Astrobiology 4, 19.Google Scholar
Cleland, C.E. & Chyba, C.F. (2002). Defining ‘Life’. Orig. Life Evol. Biosph. 32, 387393.CrossRefGoogle ScholarPubMed
Cresto Aleina, F., Baudena, M., D'Andrea, F. & Provenzale, A. (2013). Multiple equilibria on planet Dune: climate-vegetation dynamics on a sandy planet. Tellus B 65, 17662. http://dx.doi.org/10.3402/tellusb.v65i0.17662.CrossRefGoogle Scholar
Daniel, R.M. & Danson, M.J. (2010). A new understanding of how temperature affects the catalytic activity of enzymes. Trends Biochem. Sci. 35(10), 584591.Google Scholar
Daniel, R.M., Danson, M.J., Eisenthal, R., Lee, C.K. & Peterson, M.E. (2007). New parameters controlling the effect of temperature on enzyme activity. Biochem. Soc. Trans. 35, 15431546.Google Scholar
Danovaro, R., Dell'Anno, A., Pusceddu, A., Gambi, C., Heiner, I. & Kristensen, R.M. (2010). The first metazoa living in permanently anoxic conditions. BMC Biol. 8, 30.Google Scholar
Delaye, L. & Lazcano, A. (2005). Prebiological evolution and the physics of the origin of life. Phys. Life Rev. 2, 4764.CrossRefGoogle ScholarPubMed
Des Marais, D.J., Harwit, M.O., Jucks, K.W., Kasting, J.F., Lin, D.N.C., Lunine, J.I., Schneider, J., Seager, S., Traub, W.A. & Woolf, N.J. (2002). Remote sensing of planetary properties and biosignatures on extrasolar terrestrial planets. Astrobiology 2, 153181.Google Scholar
Des Marais, D.J. et al. (2003). The NASA astrobiology roadmap. Astrobiology 3, 219235.CrossRefGoogle ScholarPubMed
Dole, S.H. (1964). Habitable Planets for Man. Blaisdell Pub. Co., New York.Google Scholar
Dressing, C.D., Spiegel, D.S., Scharf, C.A., Menou, K. & Raymond, S.N. (2010). Habitable climates: the influence of eccentricity. Astrophys. J. 721, 12951307.Google Scholar
Ferrari, F. & Szuszkiewicz, E. (2009). Cosmic rays: a review for astrobiologists. Astrobiology 9(4), 413436.Google Scholar
Fields, B.D., Athanassiadou, T. & Johnson, S.R. (2008). Supernova collisions with the heliosphere. Astrophys. J. 678, 549.Google Scholar
Forgan, D. (2014) Assessing circumbinary habitable zones using latitudinal energy balance modelling. MNRAS 437, 1352.Google Scholar
Foreman-Mackey, D., Hogg, D.W. & Morton, T.D. (2014). Exoplanet population inference and the abundance of Earth analogs from noisy, incomplete catalogs. Astrophys. J. 795, 64.Google Scholar
Fox, R.F. & Choi, M.H. (2001). Rectified Brownian motion and kinesin motion along microtubules. Phys. Rev. E 63, 051901051913.CrossRefGoogle ScholarPubMed
Frank, E.A., Meyer, B.S. & Mojzsis, S.J. (2014). A radiogenic heating evolution model for cosmochemically Earth-like exoplanets. Icarus 243, 274286.CrossRefGoogle Scholar
Gates, D.M. (1980). Biophysical Ecology. Springer-Verlag, New York.Google Scholar
Goldblatt, C. & Watson, A.J. (2012). The runaway greenhouse: implications for future climate change, geoengineering and planetary atmospheres. Phil. Trans. R. Soc. A 370, 41974216.Google Scholar
Gowanlock, M.G., Patton, D.R. & McConnell, S.M. (2011). A model of habitability within the Milky Way galaxy. Astrobiology 11(9), 855873.Google Scholar
Grießmeier, J.-M.A., Stadelmann, A., Motschmann, U., Belisheva, N.K., Lammer, H. & Biernat, H.K. (2005). Cosmic ray impact on extrasolar earth-like planets in close-in habitable zones. Astrobiology 5(5), 587603.Google Scholar
Grosberg, R.K. & Strathmann, R.R. (2007). The evolution of multicellularity: a minor major transition? Ann. Rev. of Ecol. Evol. Syst. 38, 621654.CrossRefGoogle Scholar
Harman, C.E., Schwieterman, E.W., Schottelkotte, J.C. & Kasting, J.F. (2015). Abiotic O2 levels on planets around F, G, K, and M stars: possible false positives for life?. Astrophys. J. 812, 137.Google Scholar
Hart, M.H. (1978). The evolution of the atmosphere of the earth. Icarus 33, 23.CrossRefGoogle Scholar
Hart, M.H. (1979). Habitable zones about main sequence stars. Icarus 37, 35.Google Scholar
Hedelt, P., von Paris, P., Godolt, M., Gebauer, S., Grenfell, J.L., Rauer, H., Schreier, F., Selsis, F. & Trautmann, T. (2013). Spectral features of Earth-like planets and their detectability at different orbital distances around F, G, and K-type stars. Astronom. Astrophys. 553, 9.Google Scholar
Hegde, S. & Kaltenegger, L. (2013). Colors of extreme exo-earth environments. Astrobiology 13, 4756.Google Scholar
Hillenius, W.J. & Ruben, J.A. (2004). The evolution of endothermy in terrestrial vertebrates: Who? When? Why? Physiol. Biochem. Zool. 77, 10191042.CrossRefGoogle Scholar
Hobbs, J.K., Jiao, W., Easter, A.D., Parker, E.J., Schipper, L.A. & Arcus, V.L. (2013). Change in heat capacity for enzyme catalysis determines temperature dependence of enzyme catalyzed rates. ACS Chem. Biol. 8(11), 23882393.Google Scholar
Jakosky, B.M. et al. (2015). MAVEN observations of the response of Mars to an interplanetary coronal mass ejection. Science 350, 0210.CrossRefGoogle Scholar
Jeffrey, G.A. & Saenger, W. (1991). Hydrogen Bonding in Biological Structures. Springer-Verlag, Berlin.Google Scholar
Kaltenegger, L., Fridlund, M. & Kasting, J. (2002). Review on habitability and biomarkers. Earth-like Planets Moons 514, 277282.Google Scholar
Kaltenegger, L., Traub, W.A. & Jucks, K.W. (2007). Spectral evolution of an earth-like planet. Astrophys. J 658, 598616.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. (1993). Earth's early atmosphere. Science 259, 920926.Google Scholar
Kasting, J.F. & Catling, D. (2003). Evolution of a habitable planet. Annu. Rev. Astron. Astrophys. 41, 429463.Google Scholar
Kasting, J.F., Whitmore, D.P. & Reynolds, R.T. (1993). Habitable zones around main sequence stars. Icarus 101, 108128 (K93).Google Scholar
Kolb, V.M. (2007). On the applicability of the Aristotelian principles to the definition of life. Int. J. of Astrobiol. 6, 5157.Google Scholar
Koonin, E.V. (2015). Origin of eukaryotes from within archea, archeal eukaryome and bursts of gene gain: eukaryogenesis just made easier? Phil. Trans. R. Soc. B 370, 20140333.Google Scholar
Kopparapu, R.K., Ramirez, R., Kasting, J.F., Eymet, V., Robinson, T.D., Mahadevan, S., Terrien, R.C., Domagal-Goldman, S., Meadows, V. & Deshpande, R. (2013). Habitable zones around main-sequence stars: new estimates. Astroph. J. 765, 131.Google Scholar
Kopparapu, R.K., Ramirez, R.M., SchottelKotte, J., Kasting, J.F., Domagal-Goldman, S. & Eymet, V. (2014). Habitable zones around main-sequence stars: dependence on planetary mass. Astroph. J. 787, L29.CrossRefGoogle Scholar
Lane, N. (2014). Bioenergetic constraints on the evolution of complex life. Cold Spring Harb. Perspect. Biol. 6, a015982.Google Scholar
Lazcano, A. (2008). Towards a definition of life: the impossible quest? Space Sci. Rev. 135, 510.Google Scholar
Leconte, J., Forget, F., Charnay, B., Wordsworth, R., Selsis, F., Millour, E. & Spiga, A. (2013a). 3D climate modeling of close-in land planets: Circulation patterns, climate moist bistability, and habitability. Astron. Astrophys. 554, A69.Google Scholar
Leconte, J., Forget, F., Charnay, B., Wordsworth, R. & Pottier, A. (2013b). Increased insolation threshold for runaway greenhouse processes on Earth-like planets. Nature 504, 268271.CrossRefGoogle ScholarPubMed
Lee, C.K., Daniel, R.M., Shepherd, C., Saul, D., Cary, S.C., Danson, M.J., Eisenthal, R. & Peterson, M.E. (2007). Eurythermalism and the temperature dependence of enzyme activity. FASEB J. 21(8), 19341941.Google Scholar
Li, K., Pahlevan, K., Kirschvink, J.L. & Yung, Y.L. (2009). Atmospheric pressure as a natural climate regulator for a terrestrial planet with a biosphere. PNAS 106(24), 95769579.Google Scholar
Lineweaver, C.H., Fenner, Y. & Gibson, B.K. (2004). The galactic habitable zone and the age distribution of complex life in the Milky Way. Science 303, 5962.Google Scholar
Lovegrove, B.G. (2012). The evolution of endothermy in Cenozoic mammals: a plesiomorphic apomorphic continuum. Biol. Rev. 87, 128162.Google Scholar
Magill, J. & Galy, J. (2005). Radioactivity Radionuclides Radiation. Springer-Verlag, Berlin, Heidelberg and European Communities.Google Scholar
Márquez, L.M., Redman, R.S., Rodriguez, R.J. & Roossinck, M.J. (2007). A virus in a fungus in a plant: three-way symbiosis required for thermal tolerance. Science 315, 513.CrossRefGoogle Scholar
Mayor, M. et al. (2011). The HARPS search for southern extra-solar planets. XXXIV. Occurrence, mass distribution and orbital properties of super-Earths and Neptune-mass planets. ArXiv e-prints, arXiv:1109.2497.Google Scholar
Mayor, M., Lovis, C. & Santos, N.C. (2014). Doppler spectroscopy as a path to the detection of Earth-like planets. Nature 513, 328335.Google Scholar
McKay, C.P. (2014). Requirements and limits for life in the context of exoplanets. PNAS 111, 1262812633.Google Scholar
Minorsky, P.V. (2003). The hot and the classic. Plant Physiol. 132, 2526.Google Scholar
Misra, A., Meadows, V., Claire, M. & Crisp, D. (2014). Using dimers to measure biosignatures and atmospheric pressure for terrestrial exoplanets. Astrobiology 14(2), 6786.Google Scholar
Narita, N., Enomoto, T., Masaoka, S. & Kusakabe, N. (2015). Titania may produce abiotic oxygen atmospheres on habitable exoplanets. Scientific Rep. 5, 13977.Google Scholar
North, G.R., Cahalan, R.F. & Coackley, J.A. (1981). Energy balance climate models. Rev. Geophys. Space phys. 19, 91121.Google Scholar
O'Malley-James, J.T., Greaves, J.S., Raven, J.A. & Cockell, C.S. (2013). Swansong biospheres: refuges for life and novel microbial biospheres on terrestrial planets near the end of their habitable lifetimes. Int. J. Astrobiol. 12, 99112.Google Scholar
O'Malley-James, J.T., Cockell, C.S., Greaves, J.S. & Raven, J.A. (2014). Swansong biospheres II: the final signs of life on terrestrial planets near the end of their habitable lifetimes. Int. J. Astrobiol. 13, 229243.Google Scholar
Oster, G. & Wang, H. (2003). How protein motors convert chemical energy into mechanical work. In Molecular Motors, ed. Schliwa, M., pp. 207227. Wiley-VHC, Weinheim.Google Scholar
Pallé, E., Zapatero Osorio, M.R., Barrena, R., Montañés-Rodríguez, P. & Martín, E.L. (2009). Earth's transmission spectrum from lunar eclipse observations. Nature 459, 814816.CrossRefGoogle ScholarPubMed
Parnell, J., Boyce, A.J. & Blamey, N.J.F. (2010). Follow the methane: the search for a deep biosphere, and the case for sampling serpentinites, on Mars. Int. J. Astrobiol. 9, 193200.Google Scholar
Peterson, M.E., Daniel, R.M., Danson, M.J. & Eisenthal, R. (2007). The dependence of enzyme activity on temperature: determination and validation of parameters. Biochem. J. 402(2), 331337.CrossRefGoogle ScholarPubMed
Pierrehumbert, R. & Gaidos, E. (2011). Hydrogen greenhouse planets beyond the habitable zone. ApJ 734, L13.Google Scholar
Pierrehumbert, R.T. (2010). Principles of Planetary Climate. Cambridge Univ. Press, Cambridge, UK.Google Scholar
Podio, L. et al. (2013). Water vapor in the protoplanetary disk of DG Tau. Astrophys. J. 766, L5.Google Scholar
Pörtner, H.O. (2002). Climate variations and the physiological basis of temperature dependent biogeography: systemic to molecular hierarchy of thermal tolerance in animals. Comp. Biochem. Physiol. Part A 132, 739.CrossRefGoogle ScholarPubMed
Prantzos, N. (2008). On the “Galactic Habitable Zone”. Space Sci. Rev. 135, 313322.Google Scholar
Precht, H., Christophersen, J., Hensel, H. & Larcher, W. (1973). Temperature and Life. Springer-Verlag, Berlin, Heidelberg.Google Scholar
Provenzale, A. (2014). Climate models. Rend. Fis. Acc. Lincei 25, 4958.Google Scholar
Rasool, S.I. & de Bergh, C. (1970). The runaway greenhouse and the accumulation of CO2 in the Venus atmosphere. Nature 226, 10371039.Google Scholar
Ravaux, J., Hamel, G., Zbinden, M., Tasiemski, A.A., Boutet, I., Léger, N., Tanguy, A., Jollivet, D. & Shillito, B. (2013). Thermal limit for metazoan life in question: in vivo heat tolerance of the Pompeii Worm. PLoS ONE 8(5), e64074.Google Scholar
Ribas, I., Guinan, E.F., Güdel, M. & Audard, M. (2005). Evolution of the solar activity over time and effects on planetary atmospheres. I. High-energy irradiances (1–1700 Å). Astrophys. J. 622, 680694.Google Scholar
Ruben, J. (1995). The evolution of endothermy in mammals and birds: from physiology to fossils. Ann. Rev. of Physiology 57(1), 6995.Google Scholar
Rugheimer, S., Kaltenegger, L., Zsom, A., Segura, A. & Sasselov, D. (2013). Spectral fingerprints of earth-like planets around FGK stars. Astrobiology 13(3), 251269.Google Scholar
Salaris, M. & Cassisi, S. (2005). Evolution of Stars and Stellar Populations. John Wiley and Sons, West Sussex, England.Google Scholar
Schmidt-Nielsen, K. (1997). Animal Physiology: Adaptation and Environment. Cambridge University Press, Cambridge, UK.Google Scholar
Schulte, P.M. (2015). The effects of temperature on aerobic metabolism: towards a mechanistic understanding of the responses of ectotherms to a changing environment. J. Exp. Biol. 218(12), 18561866.Google Scholar
Schulze-Makuch, D. & Irwin, L.N. (2008). Life in the Universe. Expectations and Constraints, 2nd edn. Springer, Berlin.CrossRefGoogle Scholar
Seager, S. & Deming, D. (2010). Exoplanet atmospheres. Annu. Rev. Astron. Astrophys. 48, 631672.Google Scholar
Segura, A., Meadows, V.S., Kasting, J.F., Crisp, D. & Cohen, M. (2007). Abiotic formation of O2 and O3 in high-CO2 terrestrial atmospheres. Astron. Astrophys. 472, 665679.Google Scholar
Selsis, F., Kasting, J.F., Levrard, B., Paillet, J., Ribas, I. & Delfosse, X. (2007). Habitable planets around the star Gliese 581? A&A 476, 13731387.Google Scholar
Shock, E.L. & Holland, M.E. (2007). Quantitative habitability. Astrobiology 7, 839851.Google Scholar
Som, S.M., Catling, D.C., Harnmeijer, J.P., Polivka, P.M. & Buick, R. (2012). Air density 2.7 billion years ago limited to less than twice modern levels by fossil raindrop imprints. Nature 484, 359.Google Scholar
Spang, A., Saw, J.H., Jørgensen, S.L., Zaremba-Niedzwiedzka, K., Martijn, J., Lind, A.E., van Eijk, R., Schleper, C., Guy, L. & Ettema, T.J.G. (2015). Complex archea that bridge the gap between prokaryotes and eukaryotes. Nature 521, 173179.Google Scholar
Spiegel, D.S., Menou, K. & Scharf, C.A. (2008). Habitable climates. ApJ 681, 16091623.Google Scholar
Spiegel, D.S., Menou, K. & Scharf, C.A. (2009). Habitable climates: the influence of obliquity. ApJ 691, 596610.Google Scholar
Spiegel, D.S., Raymond, S.N., Dressing, C.D., Scharf, C.A. & Mitchell, J.L. (2010). Generalized Milankovitch cycles and long-term climatic habitability. ApJ 721, 13081318.Google Scholar
Spitoni, E., Matteucci, F. & Sozzetti, A. (2014). The galactic habitable zone of the Milky Way and M31 from chemical evolution models with gas radial flows. MNRAS 440, 25882598.Google Scholar
Stan-Lotter, H. (2007). Extremophiles, the physicochemical limits of life (growth and survival). In Complete Course in Astrobiology, ed. Horneck, G. & Rettberg, P., pp. 121150. Wiley-VCH, Weinheim, Germany.Google Scholar
Steel, H., Verdoodt, F., Čerevková, A., Couvreur, M., Fonderie, P., Moens, T. & Bert, W. (2013). Survival and colonization of nematodes in a composting process. Invertebr. Biol. 132(2), 108119.Google Scholar
Stein, C., Finnenkötter, A., Lowman, J.P. & Hansen, U. (2011). The pressure-weakening effect in super-Earths: consequences of a decrease in lower mantle viscosity on surface dynamics. Geophys. Res. Lett. 38, L21201.Google Scholar
Stevenson, A. et al. (2015). Multiplication of microbes below 0.690 water activity: implications for terrestrial and extraterrestrial life. Environ. Microbiol. 17(2), 257277.Google Scholar
Valencia, D., O'Connell, R.J. & Sasselov, D.D. (2007). Inevitability of plate tectonics on super-earths. ApJ 670, L45L48.Google Scholar
Vasseur, D.A., DeLong, J.P., Gilbert, B., Greig, H.S., Harley, C.D.G., McCann, K.S., Savage, V., Tunney, T.D. & O'Connor, M.I. (2015). Increased temperature variation poses a greater risk to species than climate warming. Proc. R. Soc. B 281, 20132612.Google Scholar
Vladilo, G., Murante, G., Silva, L., Provenzale, A., Ferri, G. & Ragazzini, G. (2013). The habitable zone of earth-like planets with different levels of atmospheric pressure. ApJ 767, 6587.Google Scholar
Vladilo, G., Silva, L., Murante, G., Filippi, L. & Provenzale, A. (2015). Modeling the surface temperature of earth-like planets. ApJ 804, 50 (20 pp).CrossRefGoogle Scholar
von Paris, P., Grenfell, J.L., Hedelt, P., Rauer, H., Selsis, F. & Stracke, B. (2013). Atmospheric constraints for the CO2 partial pressure on terrestrial planets near the outer edge of the habitable zone. Astron. Astrophys. 549, A94.Google Scholar
Walker, J.C.G., Hays, P.B. & Kasting, J.F. (1981). A negative feedback mechanism for the long-term stabilization of Earth's surface temperature. J. Geophys. Res. 86(C10), 97769782.Google Scholar
Ward, D.M. & Castenholz, R.W. (2000). Cyanobacteria in geothermal habitats. In The Ecology of Cyanobacteria, ed. Whitton, B.A. & Potts, M., pp. 3759. Kluwer Academic Publishers, Dordrecht, Netherlands.Google Scholar
Watson, A.J. & Lovelock, J.E. (1983). Biological homeostasis of the global environment: the parable of Daisyworld. Tellus B 35B, 284289.Google Scholar
Williams, D.M. & Kasting, J.F. (1997). Habitable planets with high obliquities. Icarus 129, 254267.Google Scholar
Williams, D.M. & Pollard, D. (2002). Earth-like worlds on eccentric orbits: excursions beyond the habitable zone. IJAsB 1, 6169.Google Scholar
Wood, A.J., Ackland, G.J., Dyke, J.G., Williams, H.T.P. & Lenton, T.M. (2008). Daisyworld: a review. Rev. Geophys. 46, RG1001. doi: 10.1029/2006RG000217.Google Scholar
Wordsworth, R. & Pierrehumbert, R. (2014). Abiotic oxygen-dominated atmospheres on terrestrial habitable zone planets. ApJ 785, L20.Google Scholar
Yang, J., Boué, G., Fabrycky, D.C. & Abbot, D.S. (2014). Strong dependence of the inner edge of the habitable zone on planetary rotation rate. ApJL 787, L2.Google Scholar
Zank, G.P. & Frisch, P.C. (1999). Consequences of a change in the galactic environment of the Sun. Astrophys. J. 518, 965973.Google Scholar
Zsom, A., Seager, S., de Wit, J. & Stamenkovic, V. (2013). Towards the minimum inner edge distance of the habitable zone. ApJ 778, 109.Google Scholar