Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-m42fx Total loading time: 0 Render date: 2024-07-17T05:00:11.388Z Has data issue: false hasContentIssue false

Chapter 10 - K–Ar, Ar–Ar and U–He Dating

Published online by Cambridge University Press:  01 February 2018

Alan P. Dickin
Affiliation:
McMaster University, Ontario
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: 2018

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

Anders, E. (1964). Meteorite ages. Rev. Mod. Phys. 34, 287325.Google Scholar
Arnaud, N. O. and Kelley, S. P. (1997). Argon behaviour in gem-quality orthoclase from Madagascar: experiments and some consequences for 40Ar/39Ar geochronology. Geochim. Cosmochim. Acta 61, 3227–55.Google Scholar
Bachmann, O., Oberli, F., Dungan, M. A. et al. (2007). 40Ar/39Ar and U–Pb dating of the Fish Canyon magmatic system, San Juan Volcanic field, Colorado: Evidence for an extended crystallization history. Chem. Geol. 236, 134–66.Google Scholar
Baksi, A. K., Archibald, D. A. and Farrar, E. (1996). Intercalibration of 40Ar/39Ar dating standards. Chem. Geol. 129, 307–24.Google Scholar
Baksi, A. K., Hsu, V., McWilliams, M. O. and Farrar, E. (1992). 40Ar/39Ar dating of the Brunhes–Matuyama geomagnetic field reversal. Science 256, 356–7.Google Scholar
Beckinsale, R. D. and Gale, N. H. (1969). A reappraisal of the decay constants and branching ratio of 40K. Earth Planet. Sci. Lett. 6, 289–94.Google Scholar
Berger, G. W. (1975). 40Ar/39Ar step heating of thermally overprinted biotite, hornblende and potassium feldspar from Eldora, Colorado. Earth Planet. Sci. Lett. 26, 387408.CrossRefGoogle Scholar
Berger, G. W. and York, D. (1981a). Geothermometry from 40Ar/39Ar dating experiments. Geochim. Cosmochim. Acta 45, 795811.Google Scholar
Berger, G. W. and York, D. (1981b). 40Ar/39Ar dating of the Thanet gabbro, Ontario: looking through the metamorphic veil and implications for paleomagnetism. Can. J. Earth Sci. 18, 266–73.Google Scholar
Brereton, N. R. (1970). Corrections for interfering isotopes in the 40Ar/39Ar dating method. Earth Planet. Sci. Lett. 8, 427–33.Google Scholar
Buchan, K. L., Berger, G. W., McWilliams, M. O., York, D. and Dunlop, D. J. (1977). Thermal overprinting of natural remanent magnetization and K/Ar ages in metamorphic rocks. J. Geomag. Geoelectr. 29, 401–10.Google Scholar
Channell, J. E. T., Hodell, D. A., Singer, B. S. and Xuan, C. (2010). Reconciling astrochronological and 40Ar/39Ar ages for the Matuyama–Brunhes boundary and late Matuyama Chron. Geochem. Geophys. Geosys. 11 (12), 121.Google Scholar
Coe, R. S., Singer, B. S., Pringle, M. S. and Zhao, X. (2004). Matuyama–Brunhes reversal and Kamikatsura event on Maui: paleomagnetic directions, 40Ar/39Ar ages and implications. Earth Planet. Sci. Lett. 222, 667–84.Google Scholar
Cox, A. and Dalrymple, G. B. (1967). Statistical analysis of geomagnetic reversal data and the precision of potassium–argon dating. J. Geophys. Res. 72, 2603–14.Google Scholar
Cox, A., Doell, R. R. and Dalrymple, G. B. (1963). Geomagnetic polarity epochs and Pleistocene geochronology. Nature 198, 1049–51.Google Scholar
Dalrymple, G. B. and Lanphere, M. A. (1969). Potassium–Argon Dating. Freeman, 258 pp.Google Scholar
Dalrymple, G. B. and Lanphere, M. A. (1971). 40Ar/39Ar technique of K–Ar dating: a comparison with the conventional technique. Earth Planet. Sci. Lett. 12, 300–8.Google Scholar
Dalrymple, G. B. and Lanphere, M. A. (1974). 40Ar/39Ar age spectra of some undisturbed terrestrial samples. Geochim. Cosmochim. Acta 38, 715–38.Google Scholar
Dalrymple, G. B. and Moore, J. G. (1968). Argon 40: excess in submarine pillow basalts from Kilauea Volcano, Hawaii. Science 161, 1132–5.Google Scholar
Damon, P. E. and Kulp, L. (1958). Excess helium and argon in beryl and other minerals. Amer. Miner. 43, 433–59.Google Scholar
Deino, A. and Potts, R. (1990). Single-crystal 40Ar/39Ar dating of the Olorgesailie Formation, Southern Kenya Rift. J. Geophys. Res.B 95, 8453–70.Google Scholar
Dodson, M. H. (1973). Closure temperature in cooling geochronological and petrological systems. Contrib. Mineral. Petrol. 40, 259–74.Google Scholar
Doell, R. R. and Dalrymple, G. B. (1966). Geomagnetic polarity epochs: A new polarity event and the age of the Brunhes–Matuyama boundary. Science 152, 1060–1.Google Scholar
Dreyfus, G. B., Raisbeck, G. M., Parrenin, F. et al. (2008). An ice core perspective on the age of the Matuyama–Brunhes boundary. Earth Planet. Sci. Lett. 274, 151–6.Google Scholar
Endt, P. M. and Van der Leun, C. (1973). Energy levels of A = 21−44 nuclei (V). Nucl. Phys. A, 214, 1625.Google Scholar
Esser, R. P., McIntosh, W. C., Heizler, M. T. and Kyle, P. R. (1997). Excess argon in melt inclusions in zero-age anorthoclase feldspar from Mt Erebus, Antarctica, as revealed by the 40Ar/39Ar method. Geochim. Cosmochim. Acta 61, 3789–801.Google Scholar
Farley, K. A. (2000). Helium diffusion from apatite: General behavior as illustrated by Durango fluorapatite. J. Geophys. Res. B 105, 2903–14.Google Scholar
Farley, K. A. (2002). (U–Th)/He dating: Techniques, calibrations, and applications. Reviews in Mineralogy and Geochemistry 47, 819–44.Google Scholar
FitzGerald, J. D. and Harrison, T. M. (1993). Argon diffusion domains in K-feldspar I: microstructures in MH–10. Contrib. Mineral. Petrol. 113, 367–80.Google Scholar
Foland, K. A. (1974). Ar-40 diffusion in homogeneous orthoclase and an interpretation of Ar diffusion in K-feldspar. Geochim. Cosmochim. Acta 38, 151–66.Google Scholar
Foland, K. A., Linder, J. S., Laskowski, T. E. and Grant, K. (1984). 40Ar–39Ar dating of glauconies: measured 39Ar recoil loss from well-crystallized specimens. Chem. Geol. (Isot. Geosci. Sect.) 46, 241–64.Google Scholar
Gaber, L. J., Foland, K. A. and Corbato, C. E. (1988). On the significance of argon release from biotite and amphibole during 40Ar/39Ar vacuum heating. Geochim. Cosmochim. Acta 52, 2457–65.Google Scholar
Garner, E. L., Machlan, L. A. and Barnes, I. L. (1976). The isotopic composition of lithium, potassium, and rubidium in some Apollo 11, 12, 14, 15, and 16 samples. Proc. 6th Lunar Sci. Conf. Pergamon, pp. 1845–55.Google Scholar
Giletti, B. J. (1974). Diffusion related to geochronology. In: Hofmann, A. W., Giletti, B. J., Yoder, H. S. and Yund, R. A. (Eds) Geochemical Transport and Kinetics. Carnegie Inst. Wash., pp. 6176.Google Scholar
Hale, C. J. (1987). The intensity of the geomagnetic field at 3. 5 Ga: paleointensity results from the Komati Formation, Barberton Mountain Land, South Africa. Earth Planet. Sci. Lett. 86, 354–64.Google Scholar
Hanes, J. A., Clark, S. J. and Archibald, D. A. (1988). An 40Ar/39Ar geochronological study of the Elzevir batholith and its bearing on the tectonothermal history of the southwestern Grenville Province, Canada. Can. J. Earth Sci. 25, 1834–45.Google Scholar
Harper, C. T. (1967). On the interpretation of potassium–argon ages from Precambrian shields and Phanerozoic orogens. Earth Planet. Sci. Lett. 3, 128–32.Google Scholar
Harrison, T. M. (1990). Some observations on the interpretation of feldspar 40Ar/39Ar results. Chem. Geol. (Isot. Geosci. Sect.) 80, 219–29.Google Scholar
Harrison, T. M., Duncan, I. and McDougall, I. (1985). Diffusion of 40Ar in biotite: Temperature, pressure and compositional effects. Geochim. Cosmochim. Acta 49, 2461–8.Google Scholar
Harrison, T. M. and McDougall, I. (1981). Excess 40Ar in metamorphic rocks from Broken Hill, New South Wales: implications for 40Ar/39Ar age spectra and the thermal history of the region. Earth Planet. Sci. Lett. 55, 123–49.Google Scholar
Hart, S. R. (1964). The petrology and isotopic–mineral age relations of a contact zone in the Front Range, Colorado. J. Geol. 72, 493525.Google Scholar
Hart, S. R. and Dodd, R. T. (1962). Excess radiogenic argon in pyroxenes. J. Geophys. Res. 67, 2998–9.CrossRefGoogle Scholar
Heizler, M. T., Lux, D. R. and Decker, E. R. (1988). The age and cooling history of the Chain of Ponds and Big Island Pond plutons and the Spider Lake granite, west-central Maine and Quebec. Amer. J. Sci. 288, 925–52.Google Scholar
Hodges, K. V., Hames, W. E. and Bowring, S. A. (1994). 40Ar/39Ar age gradients in micas from a high-temperature – low-pressure metamorphic terrane: evidence for very slow cooling and implications for the interpretation of age spectra. Geology 22, 55–8.2.3.CO;2>CrossRefGoogle Scholar
House, M. A., Farley, K. A. and Kohn, B. P. (1999). An empirical test of helium diffusion in apatite: borehole data from the Otway basin, Australia. Earth Planet. Sci. Lett. 170, 463–74.Google Scholar
Hurley, P. M. (1954). The helium age method and the distribution and migration of helium in rocks. In: Faul, H. (Ed.) Nuclear Geology. Wiley, pp. 301–29.Google Scholar
Johnson, R. G. (1982). Brunhes–Matuyama magnetic reversal dated at 790,000 yr B.P. by marine–astronomical correlations. Quaternary Res. 17, 135–47.Google Scholar
Jourdan, F. and Renne, P. R. (2007). Age calibration of the Fish Canyon sanidine 40Ar/39Ar dating standard using primary K–Ar standards. Geochim. Cosmochim. Acta 71, 387402.Google Scholar
Kaneoka, I. (1974). Investigation of excess argon in ultramafic rocks from the Kola peninsula by the 40Ar/39Ar method. Earth Planet. Sci. Lett. 22, 145–56.CrossRefGoogle Scholar
Kapusta, Y., Steinitz, G., Akkerman, A. et al. (1997). Monitoring the deficit of 39Ar in irradiated clay fractions and glauconites: modelling and analytical procedure. Geochim. Cosmochim. Acta 61, 4671–8.Google Scholar
Kelley, S. P. (2002). Excess argon in K–Ar and Ar–Ar geochronology. Chem. Geol. 188, 122.Google Scholar
Kelley, S. P. and Turner, G. (1991). Laser probe 40Ar–39Ar measurements of loss profiles within individual hornblende grains from the Giants Range granite, northern Minnesota, USA. Earth Planet. Sci. Lett. 107, 634–48.Google Scholar
Kelley, S. P., Arnaud, N. O. and Turner, S. P. (1994). High spatial resolution 40Ar/39Ar investigations using an ultra-violet laser probe extraction technique. Geochim. Cosmochim. Acta 58, 3519–25.Google Scholar
Kuiper, K. F., Deino, A., Hilgen, F. J. et al. (2008). Synchronizing rock clocks of Earth history. Science 320, 500–4.Google Scholar
Lanphere, M. A. and Dalrymple, G. B. (1966). Simplified bulb tracer system for argon analysis. Nature 209, 902–3.Google Scholar
Lanphere, M. A. and Dalrymple, G. B. (1971). A test of the 40Ar/39Ar age spectrum technique on some terrestrial materials. Earth Planet. Sci. Lett. 12, 359–72.Google Scholar
Lanphere, M. A. and Dalrymple, G. B. (1976). Identification of excess 40Ar by the 40Ar/39Ar age spectrum technique. Earth Planet. Sci. Lett. 32, 141–8.Google Scholar
Layer, P. W., Hall, C. M. and York, D. (1987). The derivation of 40Ar/39Ar age spectra of single grains of hornblende and biotite by laser step-heating. Geophys. Res. Lett. 14, 757–60.CrossRefGoogle Scholar
Lee, J. (1995a). Rapid uplift and rotation of mylonitic rocks from beneath a detachment fault: insights from potassium feldspar 40Ar/39Ar thermochronology, northern Snake range, Nevada. Tectonics 14, 5477.Google Scholar
Lee, J. K. W. (1995b). Multipath diffusion in geochronology. Contrib. Mineral. Petrol. 120, 6082.Google Scholar
Lee, J. K. W. and Aldama, A. A. (1992). Multipath diffusion: a general numerical model. Comput. Geosci. 18, 531–55.Google Scholar
Lee, J. K. W., Onstott, T. C., Cashman, K. V., Cumbest, R. J. and Johnson, D. (1991). Incremental heating of hornblende in vacuo: implications for 40Ar/39Ar geochronology and the interpretation of thermal histories. Geology 19, 872–6.Google Scholar
Lee, J. K. W., Onstott, T. C. and Hanes, J. A. (1990). An 40Ar/39Ar investigation of the contact effects of a dyke intrusion, Kapuskasing Structural Zone, Ontario. Contrib. Mineral. Petrol. 105, 87105.Google Scholar
Lo, C.-H., Lee, J. K. W. and Onstott,. C. (2000). Argon release mechanisms of biotite in vacuo and the role of short-circuit diffusion and recoil. Chem. Geol. 165, 135–66.Google Scholar
Lopez Martinez, M., York, D., Hall, C. M. and Hanes, J. A. (1984). Oldest reliable 40Ar/39Ar ages for terrestrial rocks: Barberton Mountain komatiites. Nature 307, 352–4.Google Scholar
Lovera, O. M. (1992). Computer programs to model 40Ar/39Ar diffusion data from multidomain samples. Comput. Geosci. 18, 789813.Google Scholar
Lovera, O. M., Grove, M. and Harrison, T. M. (2002). Systematic analysis of K-feldspar 40Ar/39Ar step heating results II: relevance of laboratory argon diffusion properties to nature. Geochim. Cosmochim. Acta 66, 1237–55.CrossRefGoogle Scholar
Lovera, O. M., Grove, M., Harrison, T. M. and Mahon, K. I. (1997). Systematic analysis of K-feldspar 40Ar/39Ar step heating results: I. Significance of activation energy determinations. Geochim. Cosmochim. Acta 61, 3171–92.Google Scholar
Lovera, O. M., Heizler, M. T. and Harrison, T. M. (1993). Argon diffusion domains in K-feldspar II: kinetic properties of MH–10. Contrib. Mineral. Petrol. 113, 381–93.Google Scholar
Lovera, O. M., Richter, F. M. and Harrison, T. M. (1989). The 40Ar/39Ar thermochronometry for slowly-cooled samples having a distribution of domain sizes. J. Geophys. Res. 94, 17 917–35.Google Scholar
Lovera, O. M., Richter, F. M. and Harrison, T. M. (1991). Diffusion domains determined by 39Ar released during step heating. J. Geophys. Res. 96, 2057–69.Google Scholar
McDowell, F. W. (1983). K−Ar dating: Incomplete extraction of radiogenic argon from alkali feldspar. Chem. Geol. 41, 119–26.CrossRefGoogle Scholar
McDowell, F. W., McIntosh, W. C. and Farley, K. A. (2005). A precise 40Ar–39Ar reference age for the Durango apatite (U–Th)/He and fission-track dating standard. Chem. Geol. 214, 249–63.Google Scholar
McDougall, I. and Harrison, T. M. (1988). Geochronology and Thermochronology by the 40Ar/39Ar Method. Oxford University Press, 212 pp.Google Scholar
McDougall, I. and Harrison, T. M. (1999). Geochronology and Thermochronology by the 40Ar/39Ar Method. 2nd Edn. Oxford University Press, 269 pp.Google Scholar
McDougall, I., Polach, H. A. and Stipp, J. J. (1969). Excess radiogenic argon in young subaerial basalts from the Auckland volcanic field, New Zealand. Geochim. Cosmochim. Acta 33, 1485520.Google Scholar
McDougall, I. and Tarling, D. H. (1964). Dating geomagnetic polarity zones. Nature 202, 171–2.Google Scholar
Mankinen, E. A. and Dalrymple, G. B. (1979). Revised geomagnetic polarity time scale for the interval 0 to 5 m.y. B.P. J. Geophys. Res. 84, 615–26.Google Scholar
Megrue, G. H. (1967). Isotopic analysis of rare gases with a laser microprobe. Science 157, 1555–6.Google Scholar
Megrue, G. H. (1973). Spatial distribution of 40Ar/39Ar ages in lunar breccia 14301. J. Geophys. Res. 78, 3216–21.Google Scholar
Merrihue, C. and Turner, G. (1966). Potassium–argon dating by activation with fast neutrons. J. Geophys. Res. 71, 2852–7.Google Scholar
Min, K., Mundil, R., Renne, P. R. and Ludwig, K. R. (2000). A test for systematic errors in 40Ar/39Ar geochronology through comparison with U/Pb analysis of a 1.1-Ga rhyolite. Geochim. Cosmochim. Acta 64, 7398.CrossRefGoogle Scholar
Mitchell, J. G. (1968). The argon-40/argon-39 method for potassium–argon age determination. Geochim. Cosmochim. Acta 32, 781–90.CrossRefGoogle Scholar
Mussett, A. E. and Dalrymple, G. B. (1968). An investigation of the source of air Ar contamination in K–Ar dating. Earth Planet. Sci. Lett. 4, 422–6.Google Scholar
Naumenko, M. O., Mezger, K., Nägler, T. F. and Villa, I. M. (2013). High precision determination of the terrestrial 40K abundance. Geochim. Cosmochim. Acta 122, 353–62.Google Scholar
Onstott, T. C., Hall, C. M. and York, D. (1989). 40Ar/39Ar thermo-chronometry of the Imataca complex, Venezuela. Precamb. Res. 42, 255–91.Google Scholar
Onstott, T. C., Miller, M. L., Ewing, R. C., Arnold, G. W. and Walsh, D. S. (1995). Recoil refinements: implications for the 40Ar/39Ar dating technique. Geochim. Cosmochim. Acta 59, 1821–34.Google Scholar
Onstott, T. C., Phillips, D. and Pringle-Goodell, L. (1991). Laser microprobe measurement of chlorine and argon zonation in biotite. Chem. Geol. 90, 145–68.Google Scholar
Parsons, I., Brown, W. L. and Smith, J. V. (1999). 40Ar/39Ar thermochronology using alkali feldspars: real thermal history or mathematical mirage of microtexture? Contrib. Mineral. Petrol. 136, 92110.CrossRefGoogle Scholar
Phillips, D. and Onstott, T. C. (1988). Argon isotopic zoning in mantle phlogopite. Geology 16, 542–6.Google Scholar
Pickles, C. S., Kelley, S. P., Reddy, S. M. and Wheeler, J. (1997). Determination of high spatial resolution argon isotope variations in metamorphic biotites. Geochim. Cosmochim. Acta 61, 3809–33.Google Scholar
Reiners, P. W. and Farley, K. A. (2001). Influence of crystal size on apatite (U–Th)/He thermochronology: an example from the Bighorn Mountains, Wyoming. Earth Planet. Sci. Lett. 188, 413–20.Google Scholar
Renne, P. R., Deino, A. L., Hilgen, F. J. et al. (2013). Time scales of critical events around the Cretaceous–Paleogene boundary. Science 339, 684–7.Google Scholar
Renne, P. R., Deino, A. L., Walter, R. C. et al. (1994). Intercalibration of astronomical and radioisotopic time. Geology 22, 783–6.Google Scholar
Renne, P. R., Mundil, R., Balco, G., Min, K. and Ludwig, K. R. (2010). Joint determination of 40K decay constants and 40Ar/40K for the Fish Canyon sanidine standard, and improved accuracy for 40Ar/39Ar geochronology. Geochim. Cosmochim. Acta 74, 5349–67.CrossRefGoogle Scholar
Renne, P. R., Swisher, C. C., Deino, A. L. et al. (1998). Intercalibration of standards, absolute ages and uncertainties in 40Ar/39Ar dating. Chem. Geol. 145, 117–52.Google Scholar
Rex, D. C., Guise, P. G. and Wartho, J.-A. (1993). Disturbed 40Ar/39Ar spectra from hornblendes: thermal loss or contamination? Chem. Geol. (Isot. Geosci. Sect.) 103, 271–81.Google Scholar
Richter, F. M., Lovera, O. M., Harrison, T. M. and Copeland, P. (1991). Tibetan tectonics from 40Ar/39Ar analysis of a single K-feldspar sample. Earth Planet. Sci. Lett. 105, 266–78.Google Scholar
Rivera, T. A., Storey, M., Zeeden, C., Hilgen, F. J. and Kuiper, K. (2011). A refined astronomically calibrated 40Ar/39Ar age for Fish Canyon sanidine. Earth Planet. Sci. Lett. 311, 420–6.Google Scholar
Roberts, H. J., Kelley, S. P. and Dahl, P. S. (2001). Obtaining geologically meaningful 40Ar–39Ar ages from altered biotite. Chem. Geol. 172, 277–90.Google Scholar
Roddick, J. C. and Farrar, E. (1971). High initial argon ratios in hornblendes. Earth Planet. Sci. Lett. 12, 208–14.Google Scholar
Schmitz, M. D. and Bowring, S. A. (2001). U–Pb zircon and titanite systematics of the Fish Canyon Tuff: an assessment of high-precision U–Pb geochronology and its application to young volcanic rocks. Geochim. Cosmochim. Acta 65, 2571–87.Google Scholar
Schmitz, M. D. and Kuiper, K. F. (2013). High-precision geochronology. Elements 9 (1), 2530.Google Scholar
Shackleton, N. J., Berger, A. and Peltier, W. R. (1990). An alternative astronomical calibration of the lower Pleistocene timescale based on ODP Site 677. Trans. Roy. Soc. Edinburgh: Earth Sci. 81, 251–61.Google Scholar
Simon, J. I., Renne, P. R. and Mundil, R. (2008). Implications of pre-eruptive magmatic histories of zircons for U–Pb geochronology of silicic extrusions. Earth Planet. Sci. Lett. 266, 182–94.Google Scholar
Smith, P. E., Evensen, N. M. and York, D. (1993). First successful 40Ar/39Ar dating of glauconies: argon recoil in single grains of cryptocrystalline material. Geology 21, 41–4.Google Scholar
Spell, T. L. and McDougall, I. (1992). Revisions to the age of the Brunhes–Matuyama boundary and the Pleistocene geomagnetic polarity timescale. Geophys. Res. Lett. 19, 1181–4.Google Scholar
Steiger, R. H. and Jager, E. (1977). IUGS Subcommission on Geochronology: convention on the use of decay constants in geo- and cosmochronology. Earth Planet. Sci. Lett. 36, 359–62.Google Scholar
Stone, J. O., Balco, G. A., Sugden, D. E. et al. (2003). Holocene deglaciation of Marie Byrd land, west Antarctica. Science 299, 99102.Google Scholar
Strutt, R. J. (1905). On the radio-active minerals. Proc. Roy. Soc. Lond. A 76, 88101.Google Scholar
Suganuma, Y., Okada, M., Horie, K. et al. (2015). Age of Matuyama–Brunhes boundary constrained by U–Pb zircon dating of a widespread tephra. Geology 43, 491–4.Google Scholar
Tauxe, L., Deino, A. D., Behrensmeyer, A. K. and Potts, R. (1992). Pinning down the Brunhes/Matuyama and upper Jaramillo boundaries: a reconciliation of orbital and isotopic time scales. Earth Planet. Sci. Lett. 109, 561–72.Google Scholar
Tremblay, M. M., Shuster, D. L. and Balco, G. (2014). Cosmogenic noble gas paleothermometry. Earth Planet. Sci. Lett. 400, 195205.Google Scholar
Trieloff, M., Jessberger, E. K. and Fieni, C. (2001). Comment on “40Ar/39Ar age of plagioclase from Acapulco meteorite and the problem of systematic errors in cosmochronology” by Paul. R. Renne. Earth Planet. Sci. Lett. 190, 267–9.Google Scholar
Turner, G. (1968). The distribution of potassium and argon in chondrites. In: Ahrens, L. H. (Ed.) Origin and Distribution of the Elements. Pergamon, pp. 387–97.Google Scholar
Turner, G. (1969). Thermal histories of meteorites by the 39Ar–40Ar method. In: Millman, P. M. (Ed.) Meteorite Research. Reidel, pp. 407–17.Google Scholar
Turner, G. (1971a). Argon 40–argon 39 dating: the optimisation of irradiation parameters. Earth Planet. Sci. Lett. 10, 227–34.Google Scholar
Turner, G. (1971b). 40Ar/39Ar ages from the lunar maria. Earth Planet. Sci. Lett. 11, 169–91.Google Scholar
Turner, G. (1972). 40Ar–39Ar age and cosmic ray irradiation history of the Apollo 15 anorthosite, 15415. Earth Planet. Sci. Lett. 14, 169–75.Google Scholar
Turner, G. and Cadogan, P. H. (1974). Possible effects of 39Ar recoil in 40Ar/39Ar dating. Proc. 5th Lunar Sci. Conf., pp. 1601–15.Google Scholar
Turner, G., Miller, J. A. and Grasty, R. L. (1966). Thermal history of the Bruderheim meteorite. Earth Planet. Sci. Lett. 1, 155–7.Google Scholar
Villa, I. M. (1997). Direct determination of 39Ar recoil distance. Geochim. Cosmochim. Acta 61, 689–91.Google Scholar
Villa, I. M. (1998). Reply to the comment by T. M. Harrison, M. Grove, and O. M. Lovera on “Direct determination of 39Ar recoil distance”. Geochim. Cosmochim. Acta 62, 349.Google Scholar
Vincent, E. A. (1960). Analysis by gravimetric and volumetric methods, flame photometry, colorimetry and related techniques. In: Smales, A. A. and Wager, L. R. (Eds) Methods in Geochemistry. Interscience. pp. 3380.Google Scholar
Wartho, J.-A., Kelley, S. P., Brooker, R. A. et al. (1999). Direct measurement of Ar diffusion profiles in a gem-quality Madagascar K-feldspar using the ultra-violet laser ablation microprobe (UVLAMP). Earth Planet. Sci. Lett. 170, 141–53.Google Scholar
Wolf, R. A., Farley, K. A. and Silver, L. T. (1996). Helium diffusion and low-temperature thermochronometry of apatite. Geochim. Cosmochim. Acta 60, 4231–40.Google Scholar
Wright, N., Layer, P. W. and York, D. (1991). New insights into thermal history from single grain 40Ar/39Ar analysis of biotite. Earth Planet. Sci. Lett. 104, 70–9.Google Scholar
York, D. (1978). A formula describing both magnetic and isotopic blocking temperatures. Earth Planet. Sci. Lett. 39, 8993.Google Scholar
York, D. (1984). Cooling histories from 40Ar/39Ar age spectra: implications for Precambrian plate tectonics. Ann. Rev. Earth Planet. Sci. 12, 383409.Google Scholar
York, D., Hall, C. M., Yanase, Y., Hanes, J. A. and Kenyon, W. J. (1981). 40Ar/39Ar dating of terrestrial minerals with a continuous laser. Geophys. Res. Lett. 8, 1136–8.Google Scholar
Zeitler, P. K., Herczeg, A. L., McDougall, I. and Honda, M. (1987). U–Th–He dating of apatite: A potential thermochronometer. Geochim. Cosmochim. Acta 51, 2865–8.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.

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.

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.

Available formats
×