Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-24T23:33:21.379Z Has data issue: false hasContentIssue false

Maintenance of permeable habitable subsurface environments by earthquakes and tidal stresses

Published online by Cambridge University Press:  16 April 2012

Norman H. Sleep*
Affiliation:
Department of Geophysics, Stanford University, Stanford, CA 94025, USA e-mail: norm@stanford.edu

Abstract

Life inhabits the subsurface of the Earth down to depths where temperature precludes it. Similar conditions are likely to exist within the traditional habitable zone for objects between 0.1 Earth mass (Mars) and 10 Earth masses (superearth). Long-term cooling and internal radioactivity maintain surface heat flow on the Earth. These heat sources are comparable and likely to be comparable in general within old rocky planets. Surface heat flow scales with mass divided by surface area and hence with surface gravity. The average absolute habitable subsurface thickness scales inversely with heat flow and gravity. Surface gravity varies by only 0.4 g for Mars to 3.15 g for a superearth. This range is less than the regional variation of heat flow on the Earth. Still ocean-boiling asteroid impacts (if they occur) are more likely to sterilize the thin habitable subsurface of large objects than thick habitable subsurface of small ones. Tectonics self-organizes to maintain subsurface permeability and habitability within both stable and active regions on the Earth. Small earthquakes within stable regions allow sudden mixing of water masses. Large earthquakes at plate boundaries allow surface water to descend to great habitable depths. Seismic shaking near major faults cracks shallow rock forming permeable regolith. Strong tidal strains form a similar porous regolith on small bodies such as Enceladus with weak stellar heating. This regolith may be water-saturated within rocky bodies and thus habitable.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2012

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

Abbott, D.L., Burgess, L., Longhi, J. & Smith, W.H.F. (1994). An empirical thermal history of the Earth's upper mantle. J. Geophys. Res. 99, 1383513850.CrossRefGoogle Scholar
Abramov, O. & Spencer, J.R. (2009). Endogenic heat from Enceladus’ south polar fractures: new observations, and models of conductive surface heating. Icarus 199, 189196.Google Scholar
Aoi, S., Kunugi, T. & Fujiwara, H. (2008). Trampoline effect in extreme ground motions. Science 322, 727730, doi:10.1126/science.1163113.CrossRefGoogle Scholar
Araki, T. et al. (2005). Experimental investigation of geologically produced antineutrinos with KamLAND. Nature 436, 499503.CrossRefGoogle ScholarPubMed
Bakun, W.H. et al. (2005). Implications for prediction and hazard assessment from the 2004 Parkfield earthquake. Nature 437, 969974, doi:10.1038/nature04067.Google Scholar
Blythe, A.E., d'Alessio, M.A. & Bürgmann, R. (2004). Constraining the exhumation and burial history of the SAFOD pilot hole with apatite fission track and (U-Th)/He thermochronometry. Geophys. Res. Lett. 31, L15S16, doi:10.1029/2003GL019407.CrossRefGoogle Scholar
Bradbury, K.K., Evans, J.P., Chester, J.S., Chester, F.M. & Kirschner, D.L. (2011). Lithology and internal structure of the San Andreas fault at depth based on characterization of Phase 3 whole-rock core in the San Andreas Fault Observatory at Depth (SAFOD) borehole. Earth Planet. Sci. Lett. 310, 131144.CrossRefGoogle Scholar
Brune, J.N. (2001). Shattered rock and precarious rock evidence for strong asymmetry in ground motions during thrust faulting. Bull. Seismol. Soc. Am. 91(3), 441447, doi:10.1785/0120000118.Google Scholar
Bryson, C.E., Cazcarra, V. & Levenson, L.L. (1974). Sublimation rates and vapor pressures of H2O, CO2, N2O, and Xe. J. Chem. Eng. 19, 107110.Google Scholar
Chao, K. & Peng, Z.G. (2009). Temporal changes of seismic velocity and anisotropy in the shallow crust induced by the 10/22/1999 M6.4 Chia-Yi, Taiwan earth- quake. Geophys. J. Int. 179, 18001816, doi:10.1111/j.1365-246X.2009.04384.x.Google Scholar
Collins, G.C., McKinnon, W.B., Moore, J.M., Nimmo, F., Pappalardo, R.T., Procktor, L.M. & Schenk, P.M. (2009). Tectonics of outer planet satellites. In Planetary Tectonics, ed. Watters, T.R. & Schultz, R.A., pp. 264350. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
d'Alessio, M.A., Williams, C.F. & Bürgmann, R. (2006). Frictional strength heterogeneity and surface heat flow: implications for the strength of the creeping San Andreas fault. J. Geophys. Res. 111, B05410, doi:10.1029/2005JB003780.Google Scholar
Delbo, M., Dell'oro, A., Harris, A.W., Mottola, S. & Mueller, M. (2007). Thermal inertia of near-Earth asteroids and implications for the magnitude of the Yarkovsky effect. Icarus 190, 236249.CrossRefGoogle Scholar
Dor, O., Yildirim, C., Rockwell, T.K., Ben-Zion, Y., Emre, O., Sisk, M. & Duman, T.Y. (2008). Geological and geomorphologic asymmetry across the rupture zones of the 1943 and 1944 earthquakes on the North Anatolian Fault: possible signals for preferred earthquake propagation direction. Geophys. J. Int. 173, 483504, doi:10.1111/j.1365-246X.2008.03709.x.Google Scholar
Elkhoury, J.E., Brodsky, E.E. & Agnew, D.C. (2006). Seismic waves increase permeability. Nature 441, 11351138, doi:10.1038/nature04798.Google Scholar
England, P. & Bickle, M. (1984). Continental thermal and tectonic regimes during the Archean. J. Geol. 92, 353367.CrossRefGoogle Scholar
Fischer, A.D. & Sammis, C.G. (2009). Dynamic driving of small shallow events during strong motion. Bull. Seismol. Soc. Am. 99, 17201729, doi:10.1785/0120080293.CrossRefGoogle Scholar
Fischer, A.D., Sammis, C.G., Chen, Y.L. & Teng, T.L. (2008a). Dynamic triggering by strong motion P and S waves: evidence from 1999 Chi-Chi, Taiwan earthquake. Bull. Seismol. Soc. Am. 98, 580592, doi:10.1785/0120070155.CrossRefGoogle Scholar
Fischer, A.D., Peng, Z.G. & Sammis, C.G. (2008b). Dynamic triggering of high-frequency bursts by strong motions during the 2004 Parkfield earthquake sequence. Geophys. Res. Lett. 35, L12305, doi:10.1029/2008GL033905.Google Scholar
Fontaine, F.J., Olive, J.A., Cannat, M., Escartin, J. & Perol, T. (2011). Hydrothermally-induced melt lens cooling and segmentation along the axis of fast- and intermediate-spreading centers. Geophys. Res. Lett. 38, L14307, doi:10.1029/2011GL047798.Google Scholar
Fortney, J.J., Marley, M.S. & Barnes, J.W. (2007). Planetary radii across five orders of magnitude in mass and stellar insolation: application to transits. Astrophys. J. 659, 16611672.CrossRefGoogle Scholar
Fulton, P.M. & Saffer, D.M. (2009). Effect of thermal refraction on heat flow near the San Andreas Fault, Parkfield, California. J. Geophys. Res. 114, B06408, doi:10.1029/2008JB005796.Google Scholar
Fulton, P.M., Saffer, D.M., Harris, R.N. & Bekins, B.A. (2004). Re-evaluation of heat flow data near Parkfield, CA: evidence for a weak San Andreas Fault. Geophys. Res. Lett. 31, L15S15, doi:10.1029/2003GL019378.Google Scholar
Galer, S.J.G. & Mezger, K. (1998). Metamorphism denudation and sea level in the Archean and cooling of the Earth. Precambrian Res. 92, 387412.CrossRefGoogle Scholar
Gaucher, E.A., Govindarajan, S., Omjoy, K. & Ganesh, O.K. (2008). Palaeotemperature trend for Precambrian life inferred from resurrected proteins. Nature 451, 704708.Google Scholar
Girty, G.H., Biggs, M.A. & Berry, R.W. (2008). An unusual occurrence of probable Pleistocene corestone with a Cretaceous dioritic enclave, Peninsular Ranges, California. Catena 74, 4357, doi:10.1016/j.catena.2008.03.009.Google Scholar
Hartman, H. & McKay, C.P. (1995). Oxygenic photosynthesis and the oxidation sate of Mars. Planet. Space Phys. 43, 123128.Google Scholar
Hashin, Z. & Shtrikman, S. (1963). A variational approach to the elastic behaviour of multiphase materials. J. Mech. Phys. Solids 11, 127140.Google Scholar
Hoehler, T.M. (2005). Biochemistry of dihydrogen (H2). In Metal Ions in Biological Systems, Biogeochemical Cycles of Elements, ed. Sigel, A., Sigel, H. & Sigel, R.K.O., vol. 43, pp. 948. Taylor & Francis, Boca Raton, FL.Google Scholar
Hoehler, T.M., Alperin, M.J., Albert, D.B. & Martens, C.S. (1998). Thermodynamic control on H2 concentrations in an anoxic marine sediment. Geochim. Cosmochim. Acta 62, 17451756.Google Scholar
Holdsworth, R.E., van Diggelen, E.W.E., Spiers, C.J., de Bresser, J.H.P., Walker, R.J. & Bowen, L. (2011). Fault rocks from the SAFOD core samples: implications for weakening at shallow depths along the San Andreas Fault, California. J. Struct. Geol. 33, 132144.Google Scholar
Hopkins, M., Harrsion, M.T., Manning, C.E. (2008). Low heat flow inferred from >4 Gyr zircons suggests Hadean plate boundary interactions. Nature 456, 493496.CrossRefGoogle ScholarPubMed
Howett, C.J.A., Spencer, J.R., Pearl, J. & Segura, M. (2011). High heat flow from Enceladus’ south polar region measured using 10–600 cm−1 Cassini/CIRS data. J. Geophys. Res. 116, E03003, doi:10.1029/2010JE003718.Google Scholar
Korenaga, J. (2008). Urey ratio and the structure and evolution of Earth's mantle. Rev. Geophys. 46, RG2007, doi:10.1029/2007RG000241.Google Scholar
Langseth, M., Keihm, S.J. & Peters, K. (1976). Revised lunar heat-flow values. Proc. Lunar Sci. Conf. 7, 31433171.Google Scholar
Lin, L.-H.et al. (2005). Radiolytic H2 in continental crust: nuclear power for deep subsurface microbial communities. Geochem. Geophys. Geosyst. 6, Q07003, doi:10.1029/2004GC000907.CrossRefGoogle Scholar
Mat, W.-K., Xue, H. & Wong, T.-F. (2008). The genomics of LUCA. Front. Biosci 13, 56055613.CrossRefGoogle ScholarPubMed
McCalpin, J.P. & Hart, E.W. (2003). Ridge-Top Spreading Features and Relationship to Earthquakes, San Gabriel Mountains Region, Part B: Paleoseismic Investigations of Ridgetop Depressions, in Ridge-Top Spreading in California [CD-ROM], CGS CD 2003-05, Contrib. 6, 51 pp., California Geological Survey, Sacramento.Google Scholar
McDonough, W.F. & Sun, S.-S. (1995). The composition of the Earth. Chem. Geol. 120, 223253.Google Scholar
McKay, C.P., Porco, C.C., Athleide, T., Davis, W.L. & Kral, T.R. (2008). The possible origin and persistence of life on Enceladus and detection of biomarkers in the plume. Astrobiology 8, 909919.Google Scholar
McSween, H.Y. Jr., Ghosh, A., Grimm, R.E., Wilson, L. & Young, E.D. (2002). Thermal evolution models of asteroids. In Asteroids III, ed. Bottke, W.F. Jr., Cellino, A., Paolicchi, P. & Binzel, R.P., pp. 559571. University of Arizona Press, Tucson.CrossRefGoogle Scholar
Miller, S.L. & Lazcano, A. (1995). The origin of life – did it occur at high temperatures. J. Mol. Evol. 41, 689692.Google Scholar
Mittempergher, S., Di Toro, G., Gratier, J.P., Hadizadeh, J., Smith, S.A.F. & Spiess, R. (2011). Evidence of transient increases of fluid pressure in SAFOD phase III cores. Geophys. Res. Lett. 38, L03301, doi:10.1029/2010GL046129.Google Scholar
Noda, H., Dunham, E.M. & Rice, J.R. (2009). Earthquake ruptures with thermal weakening and the operation of major faults at low overall stress levels. J. Geophys. Res. 114, B07302, doi:10.1029/2008JB006143.Google Scholar
O'Neill, C. & Nimmo, F. (2010). The role of episodic overturn in generating the surface geology and heat flow on Enceladus. Nature Geosci. 3, 8891, Doi:10.1038/NGEO731.Google Scholar
Prialnik, D. & Merk, R. (2008). Growth and evolution of small porous icy bodies with an adaptive-grid thermal evolution code I. Application to Kuiper belt objects and Enceladus. Icarus 197, 211220.CrossRefGoogle Scholar
Replogle, C.T. (2011). Corestone and saprock development in a zone of precariously balanced rocks, Peninsular Ranges, southern California: speculations on the effects of ground shaking during earthquakes. MS Thesis, 67 pp., San Diego State University, San Diego, CA.Google Scholar
Ryder, G. (2002). Mass flux in the ancient Earth-Moon system and benign implications for the origin of life on Earth. J. Geophys. Res. 107(E4), 5022, DOI:10.1029/2001JE001583.Google Scholar
Sasselov, D.D., Valencia, D. & O'Connell, R.J. (2008). Massive terrestrial planets (super-Earths): detailed physics of their interiors. Physica Scripta T130, 014035, DOI:10.1088/0031-8949/2008/T130/014035.Google Scholar
Schleicher, A.M., Tourscher, S.N., van der Pluijm, B.A. & Warr, L.N. (2009). Constraints on mineralization, fluid-rock interaction, and mass transfer during faulting at 2–3 km depth from the SAFOD drill hole. J. Geophys. Res. 114, B04202, doi:10.1029/2008JB006092.Google Scholar
Seager, S., Kuchner, M., Hier-Majurnder, C.A. & Militzer, B. (2007). Mass-radius relationships for solid exoplanets. Astrophys. J. 669, 12791297.CrossRefGoogle Scholar
Sherwood Lollar, B., Lacrampe-Couloume, G., Slater, G.F., Ward, J., Moser, D.P., Gihring, T.M., Lin, L.H. & Onstott, T.C. (2006). Unraveling abiogenic and biogenic sources of methane in the Earth's deep subsurface. Chem. Geol. 226, 328339.Google Scholar
Sleep, N.H. (2007). Plate tectonics through time. In Treatise on Geophysics, ed. Schubert, G., vol. 9, pp. 101117. Elsevier, Oxford.Google Scholar
Sleep, N.H. (2011). Seismically damaged regolith as self-organized fragile geological feature. Geochem. Geophys. Geosyst. 12, Q12013, doi:10.1029/2011GC003837.Google Scholar
Sleep, N.H. & Ma, S. (2008). Production of brief extreme ground acceleration pulses by nonlinear mechanisms in the shallow subsurface. Geochem. Geophys. Geosyst. 9, Q03008, doi:10.1029/2007GC001863.Google Scholar
Sleep, N.H. & Zoback, M.D. (2007). Did earthquakes keep the early crust habitable? Astrobiology 7(6), 10231032.CrossRefGoogle ScholarPubMed
Sotin, C., Grasset, O. & Mocquet, A. (2007). Mass-radius curve for extrasolar Earth-like planets and ocean planets. Icarus 191, 337351.CrossRefGoogle Scholar
Takai, K. et al. (2008). Cell proliferation at 122 °C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high-pressure cultivation. Proc. Natl. Acad. Sci. U.S.A. 105, 1094910954.Google Scholar
Townend, J. & Zoback, M.D. (2000). How faulting keeps the crust strong. Geology 28, 399402.2.0.CO;2>CrossRefGoogle Scholar
Valencia, D., O'Connell, R.J. & Sasselov, D.D. (2006). Massive terrestrial planets. Icarus 181, 54554.Google Scholar
Valencia, D., O'Connell, R.J. & Sasselov, D.D. (2007a). Inevitability of plate tectonics on super-earths. Astrophys. J. 670, L45L48.Google Scholar
Valencia, D., Sasselov, D.D. & O'Connell, R.J. (2007b). Radius and structure models of the first super-Earth planet. Astrophys. J. 656, 545551.Google Scholar
Wang, C.-Y. (2011). High pore pressure, or its absence, in the San Andreas Fault. Geology 39(11), 10471050, doi:10.1130/G32294.1.Google Scholar
Wechsler, N., Rockwell, T.K. & Ben-Zion, Y. (2009). Application of high resolution DEM data to detect rock damage from geomorphic signals along the central San Jacinto Fault. Geomorphology 113, 8296, doi:10.1016/j.geomorph. 2009.06.007.Google Scholar
Wiersberg, T. & Erzinger, J. (2008). On the origin and spatial distribution of gas at seismogenic depths of the San Andreas Fault from drilling mud gas analysis. Appl. Geochem. 23, 16751690, doi:10.1016/j.apgeochem.2008.01.012.Google Scholar
Wiersberg, T. & Erzinger, J. (2011). Chemical and isotope compositions of drilling mud gas from the San Andreas Fault Observatory at Depth (SAFOD) boreholes: implications on gas migration and the permeability structure of the San Andreas Fault. Chem. Geol. 284, 148159.CrossRefGoogle Scholar
Yomogida, K. & Matsui, T. (1983). Physical properties of ordinary chondrites. J. Geophys. Res. 88(B11), 95139533, doi:10.1029/JB088iB11p09513.Google Scholar
Zahnle, K.J. & Sleep, N.H. (2006). Impacts and the early evolution of life. In Comets and the Origin and Evolution of Life, 2nd edn, ed. Thomas, P.J., Hicks, R.D., Chyba, C.F. & McKay, C.P., pp. 207251. Springer, New York.Google Scholar
Zahnle, K., Arndt, N., Cockell, C., Halliday, A., Nisbet, E., Selsis, F. & Sleep, N.H. (2007). Emergence of a habitable planet. Space Sci. Rev. 129, 3578.Google Scholar
Zoback, M.D., Hickman, S. & Ellsworth, W. (2010). Scientific drilling into the San Andreas fault zone. Eos Trans. Am. Geophys. Union 91, 197204, doi:10.1029/2010EO220001.Google Scholar
Zoback, M.D. & Townend, J. (2001). Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere. Tectonophysics 336, 1930.Google Scholar