Skip to main content Accessibility help
×
Hostname: page-component-84b7d79bbc-5lx2p Total loading time: 0 Render date: 2024-07-26T05:41:52.946Z Has data issue: false hasContentIssue false

Chapter 14 - Cosmogenic Nuclides

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

Abreu, J. A., Beer, J., Steinhilber, F., Tobias, S. M. and Weiss, N. O. (2008). For how long will the current grand maximum of solar activity persist? Geophys. Res. Lett. 35, L20109 14.Google Scholar
Abreu, J. A., Beer, J., Ferriz-Mas, A., McCracken, K. G. and Steinhilber, F. (2012). Is there a planetary influence on solar activity? Astron. & Astrophys. 548, A88 19.Google Scholar
Adkins, J. F. and Boyle, E. A. (1997). Changing atmospheric D14C and the record of deep water paleoventilation ages. Paleoceanography 12, 337–44.CrossRefGoogle Scholar
Anderson, E. C. and Libby, W. F. (1951). World-wide distribution of natural radiocarbon. Phys. Rev. 81, 64–9.Google Scholar
Andree, M., Oeschger, H., Broecker, W. et al. (1986). Limits on ventilation rates for the deep ocean over the last 12,000 years. Climate Dynamics 1, 5362.CrossRefGoogle Scholar
Andrews, J. N., Davis, S. N., Fabryka-Martin, J. et al. (1989). The in situ production of radioisotopes in rock matrices with particular reference to the Stripa granite. Geochim. Cosmochim. Acta 53, 1803–15.CrossRefGoogle Scholar
Arnold, J. R. (1956). Beryllium-10 produced by cosmic rays. Science 124, 584–5.CrossRefGoogle ScholarPubMed
Arnold, J. R. (1958). Trace elements and transport rates in the ocean. 2nd UN Conf. on Peaceful Uses of Atomic Energy 18, 344–6. IAEA.Google Scholar
Arnold, J. R. and Libby, W. F. (1949). Age determinations by radiocarbon content: Checks with samples of known age. Science 110, 678–80.CrossRefGoogle ScholarPubMed
Balco, G. and Rovey, C. W. (2008). An isochron method for cosmogenic-nuclide dating of buried soils and sediments. Amer. J. Sci. 308, 1083–114.CrossRefGoogle Scholar
Balco, G. and Rovey, C. W. (2010). Absolute chronology for major Pleistocene advances of the Laurentide Ice Sheet. Geology 38, 795–8.Google Scholar
Balco, G. and Shuster, D. L. (2009). 26Al–10Be–21Ne burial dating. Earth Planet. Sci. Lett. 286, 570–5.CrossRefGoogle Scholar
Bard, E., Arnold, M., Fairbanks, R. G. and Hamelin, B. (1993). 230Th–234U and 14C ages obtained by mass spectrometry on corals. Radiocarbon 35, 191–9.Google Scholar
Bard, E., Arnold, M., Hamelin, B., Tisnerat-Laborde, N. and Cabioch, G. (1998). Radiocarbon calibration by means of mass spectrometric 230Th/234U and 14C ages of corals: an updated database including samples from Barbados, Mururoa and Tahiti. Radiocarbon 40, 1085–92.CrossRefGoogle Scholar
Bard, E., Arnold, M., Mangerud, J. et al. (1994). The North Atlantic atmospheresea surface 14C gradient during the Younger Dryas climatic event. Earth Planet. Sci. Lett. 126, 275–87.CrossRefGoogle Scholar
Bard, E. and Frank, M. (2006). Climate change and solar variability: What's new under the sun?. Earth Planet. Sci. Lett. 248, 114.Google Scholar
Bard, E., Hamelin, B., Fairbanks, R. G. and Zindler, A. (1990a). Calibration of the 14C timescale over the past 30,000 years using mass spectrometric UTh ages from Barbados corals. Nature 345, 405–10.CrossRefGoogle Scholar
Bard, E., Hamelin, B., Fairbanks, R. G. et al. (1990b). U/Th and 14C ages of corals from Barbados and their use for calibrating the 14C timescale beyond 9000 years B.P. Nucl. Instr. Meth. in Phys. Res. B 52, 461–8.Google Scholar
Bard, E., Raisbeck, G. M., Yiou, F. and Jouzel, J. (1997). Solar modulation of cosmogenic nuclide production over the last millennium: comparison between 14C and 10Be records. Earth Planet. Sci. Lett. 150, 453–62.CrossRefGoogle Scholar
Baumgartner, S., Beer, J., Masarik, J. et al. (1998). Geomagnetic modulation of the 36Cl flux in the GRIP ice core, Greenland. Science 279, 1330–2.Google Scholar
Baxter, M. S. and Farmer, J. G. (1973). Radiocarbon: short-term variations. Earth Planet. Sci. Lett. 20, 295–9.Google Scholar
Bayes, T. and Price, R. (1763). An essay towards solving a problem in the doctrine of chances. Phil. Trans. 53, 370418.Google Scholar
Beck, J. W., Richards, D. A., Edwards, R. L. et al. (2001). Extremely large variations of atmospheric 14C concentration during the last glacial period. Science 292, 2453–8.CrossRefGoogle ScholarPubMed
Beer, J., McCracken, K. G., Abreu, J., Heikkilä, U. and Steinhilber, F. (2013). Cosmogenic radionuclides as an extension of the neutron monitor era into the past: potential and limitations. Space Sci. Rev. 176, 89100.Google Scholar
Beer, J., Andree, M., Oeschger, H. et al. (1984). The Camp Century 10Be record: implications for long-term variations of the geomagnetic dipole moment. Nucl. Instr. Meth. in Phys. Res. B 5, 380–4.Google Scholar
Beer, J., Oeschger, H., Finkel, R. C. et al. (1985). Accelerator measurements of 10Be: the 11 year solar cycle. Nucl. Instr. Meth. in Phys. Res. B 10, 415–18.Google Scholar
Beer, J., Siegenthaler, U., Bonani, G. et al. (1988). Information on past solar activity and geomagnetism from 10Be in the Camp Century ice core. Nature 331, 675–9.CrossRefGoogle Scholar
Bentley, H. W., Phillips, F. M., Davis, S. N. et al. (1982). Thermonuclear 36Cl pulse in natural water. Nature 300, 737–40.CrossRefGoogle Scholar
Bentley, H. W., Phillips, F. M., Davis, S. N. et al. (1986). Chlorine 36 dating of very old groundwaters, 1, The Great Artesian Basin, Australia. Water Resources Res. 22, 19912002.CrossRefGoogle Scholar
Bierman, P., Gillespie, A., Caffee, M. and Elmore, D. (1995). Estimating erosion rates and exposure ages with 36Cl produced by neutron activation. Geochim. Cosmochim. Acta 59, 3779–98.Google Scholar
Broecker, W., Barker, S., Clark, E. et al. (2004). Ventilation of the glacial deep Pacific Ocean. Science 306, 1169–72.Google Scholar
Broecker, W., Clark, E. and Barker, S. (2008). Near constancy of the Pacific Ocean surface to mid-depth radiocarbon-age difference over the last 20 kyr. Earth Planet. Sci. Lett. 274, 322–6.CrossRefGoogle Scholar
Broecker, W. S. and Denton, G. H. (1989). The role of oceanatmosphere reorganizations in glacial cycles. Geochim. Cosmochim. Acta 53, 2465–501.Google Scholar
Broecker, W. S., Gerard, R., Ewing, M. and Heezen, B. C. (1960). Natural radiocarbon in the Atlantic Ocean. J. Geophys. Res. 65, 2903–31.Google Scholar
Broecker, W. S. and Peng, T. H. (1982). Tracers in the Sea. Lamont-Doherty Geol. Obs. 690 pp.Google Scholar
Brown, E. T., Edmond, J. M., Raisbeck, G. M. et al. (1992). Beryllium isotope geochemistry in tropical basins. Geochim. Cosmochim. Acta 56, 1607–24.Google Scholar
Brown, L., Klein, J. and Middleton, R. (1985). Anomalous isotopic concentrations in the sea off Southern California. Geochim. Cosmochim. Acta 49, 153–7.Google Scholar
Brown, L., Klein, J., Middleton, R., Sacks, I. S. and Tera, F. (1982). 10Be in island-arc volcanoes and implications for subduction. Nature 299, 718–20.Google Scholar
Brown, L., Stensland, G. J., Klein, J. and Middleton, R. (1989). Atmospheric deposition of 7Be and 10Be. Geochim. Cosmochim. Acta 53, 135–42.CrossRefGoogle Scholar
Bruns, M., Rhein, M., Linick, T. W. and Suess, H. E. (1983). The atmospheric 14C level in the 7th millennium BC. P.A.C.T. (Physical And Chemical Techniques in Archaeology) 8, 511–16.Google Scholar
Bucha, V. and Neustupny, E. (1967). Changes in the Earth's magnetic field and radiocarbon dating. Nature 215, 261–3.CrossRefGoogle Scholar
Buck, C. E., Litton, C. D. and Smith, A. F. (1992). Calibration of radiocarbon results pertaining to related archaeological events. J. Archaeological Sci. 19, 497512.Google Scholar
Burke, A., and Robinson, L. F. (2012). The Southern Ocean's role in carbon exchange during the last deglaciation. Science 335, 557–61.Google Scholar
Campin, J.-M., Fichefet, T. and Duplessy, J.-C. (1999). Problems with using radiocarbon to infer ocean ventilation rates for past and present climates. Earth Planet. Sci. Lett. 165, 1724.Google Scholar
Chengde, S., Beer, J., Tungsheng, L. et al. (1992). 10Be in Chinese loess. Earth Planet. Sci. Lett. 109, 169–77.Google Scholar
Chmeleff, J., von Blanckenburg, F., Kossert, K. and Jakob, D. (2010). Determination of the 10Be half-life by multicollector ICP–MS and liquid scintillation counting. Nucl. Instr. Meth. in Phys. Res. B 268, 192–9.Google Scholar
Cohen, T. J. and Lintz, P. R. (1974). Long term periodicities in the sunspot cycle. Nature 250, 398400.Google Scholar
Craig, H. (1954). Carbon-13 in plants and the relationships between carbon-13 and carbon-14 variations in nature. J. Geol. 62, 115–49.CrossRefGoogle Scholar
Craig, H. (1957). Isotopic standards for carbon and oxygen and correction factors for mass-spectrometric analysis of carbon dioxide. Geochim. Cosmochim. Acta 12, 133–49.Google Scholar
Davis, R. and Schaeffer, O. A. (1955). Chlorine-36 in Nature. Ann. N. Y. Acad. Sci. 62, 105–22.Google Scholar
De Jong, A. F. M., Mook, W. G. and Becker, B. (1979). Confirmation of the Suess wiggles: 32003700 BC. Nature 280, 48–9.Google Scholar
De Vries, H. (1958). Variation in concentration of radiocarbon with time and location on Earth. Proc. Konikl. Ned. Akad. Wetenschap B 61, 94102.Google Scholar
De Vries, H. and Barendsen, G. W. (1953). Radiocarbon dating by a proportional counter filled with carbon dioxide. Physica 19, 9871003.CrossRefGoogle Scholar
DeVries, T. and Primeau, F. (2010). An improved method for estimating water-mass ventilation age from radiocarbon data. Earth and Planetary Science Letters 295, 367–78.CrossRefGoogle Scholar
Druffel, E. M. (1996). Post-bomb radiocarbon records of surface corals from the tropical Atlantic Ocean. Radiocarbon 38, 563–72.Google Scholar
Eddy, J. A. (1976). The Maunder minimum. Science 192, 1189–202.Google Scholar
Eddy, J. A. (1977). Climate and the changing sun. Climate Change 1, 173–90.Google Scholar
Edwards, R. L., Beck, J. W., Burr, G. S. et al. (1993a). A large drop in atmospheric 14C/12C and reduced melting in the Younger Dryas, documented with 230Th ages of corals. Science 260, 962–8.Google Scholar
Edwards, C. M. H., Morris, J. D. and Thirlwall, M. F. (1993b). Separating mantle from slab signatures in arc lavas using B/Be and radiogenic isotope systematics. Nature 362, 530–3.Google Scholar
Elmore, D., Gove, H. E., Ferraro, R. et al. (1980). Determination of 129I using tandem accelerator mass spectrometry. Nature 286, 138–40.Google Scholar
Elmore, D., Tubbs, L. E., Newman, D. et al. (1982). 36Cl bomb pulse measured in a shallow ice core from Dye 3, Greenland. Nature 300, 735–7.CrossRefGoogle Scholar
Elsasser, W., Ney, E. P. and Winckler, J. R. (1956). Cosmic-ray intensity and geomagnetism. Nature 178, 1226–7.Google Scholar
Evans, J. C., Rancitelli, L. A. and Reeves, J. H. (1979). 26Al content of Antarctic meteorites: implications for terrestrial ages and bombardment history. Proc. 10th Lunar Planet. Sci. Conf., 1061–72.Google Scholar
Evans, J. C. and Reeves, J. H. (1987). 26Al survey of Antarctic meteorites. Earth Planet. Sci. Lett. 82, 223–30.Google Scholar
Fabryka-Martin, J., Bentley, H., Elmore, D. and Airey, P. L. (1985). Natural iodine-129 as an environmental tracer. Geochim. Cosmochim. Acta 49, 337–47.Google Scholar
Fabryka-Martin, J., Davis, S. N. and Elmore, D. (1987). Applications of 129I and 36Cl in hydrology. Nucl. Instr. Meth. in Phys. Res. B 29, 361–71.Google Scholar
Fabryka-Martin, J., Davis, S. N., Elmore, D. and Kubik, P. W. (1989). In situ production and migration of 129I in the Stripa granite, Sweden. Geochim. Cosmochim. Acta 53, 1817–23.CrossRefGoogle Scholar
Fan, C. Y., Chen, T. M., Yun, S. X. and Dai, K. M. (1986). Radiocarbon activity variation in dated tree rings grown in Mackenzie delta. Radiocarbon 28, 300–5.CrossRefGoogle Scholar
Fehn, U. (2012). Tracing crustal fluids: Applications of natural 129I and 36Cl. Ann. Rev. Earth Planet. Sci. 40, 45.Google Scholar
Fehn, U., Holdren, G. R., Elmore, D. et al. (1986). Determination of natural and anthropogenic 129I in marine sediments. Geophys. Res. Lett. 13, 137–9.CrossRefGoogle Scholar
Fehn, U., Moran, J. E., Snyder, G. T. and Muramatsu, Y. (2007a). The initial 129I/I ratio and the presence of ‘old’ iodine in continental margins. Nucl. Instr. Meth. in Phys. Res. B 259, 496502.Google Scholar
Fehn, U., Snyder, G. T. and Muramatsu, Y. (2007b). Iodine as a tracer of organic material: 129I results from gas hydrate systems and fore arc fluids. J. Geochem. Explor. 95, 6680.Google Scholar
Ferguson, C. W. (1970). Dendrochronology of Bristlecone pine, Pinus aristata. Establishment of a 7484-year chronology in the White Mountains of eastern-central California, USA. In: Olsson, I. U. (Ed.) Radiocarbon Variations and Absolute Chronology, Proc. 12th Nobel Symp. Wiley, pp. 571–93.Google Scholar
Ferguson, C. W. and Graybill, D. A. (1983). Dendrochronology of Bristlecone pine: a progress report. Radiocarbon 25, 287–8.Google Scholar
Feulner, G. and Rahmstorf, S. (2010). On the effect of a new grand minimum of solar activity on the future climate on Earth. Geophys. Res. Lett. 37, L05707, 15.Google Scholar
Fink, D., Middleton, R., Klein, J. and Sharma, P. (1990). 41Ca measurement by accelerator mass spectrometry and applications. Nucl. Instr. Meth. in Phys. Res. B 47, 7996.Google Scholar
Fleitmann, D., Burns, S. J., Mudelsee, M. et al. (2003). Holocene forcing of the Indian monsoon recorded in a stalagmite from southern Oman. Science 300, 1737–9.Google Scholar
Forbush, S. E. (1954). Worldwide cosmic-ray variations, 1937–1952. J. Geophys. Res. 59, 525–42.Google Scholar
Frank, M., Schwarz, B., Baumann, S. et al. (1997). A 200 ka record of cosmogenic radionuclide production rate and geomagnetic field intensity from 10Be in globally stacked deep-sea sediments. Earth Planet. Sci. Lett. 149, 121–9.Google Scholar
Galbraith, E. D., Jaccard, S. L., Pedersen, T. F. et al. (2007). Carbon dioxide release from the North Pacific abyss during the last deglaciation. Nature 449, 890–3.Google Scholar
Godwin, H. (1962). Half-life of radiocarbon. Nature 195, 984.Google Scholar
Goldberg, E. D. and Arrhenius, G. O. S. (1958). Chemistry of Pacific pelagic sediments. Geochim. Cosmochim. Acta 13, 153212.Google Scholar
Goldstein, S. J., Lea, D. W., Chakraborty, S., Kashgarian, M. and Murrell, M. T. (2001). Uranium-series and radiocarbon geochronology of deep-sea corals: implications for Southern Ocean ventilation rates and the ocean carbon cycle. Earth Planet. Sci. Lett. 193, 167–82.CrossRefGoogle Scholar
Goslar, T., Arnold, M., Bard, E. et al. (1995). High concentration of atmospheric 14C during the Younger Dryas cold episode. Nature 377, 414–17.Google Scholar
Gove, H. E. (1987). Tandem-accelerator mass-spectrometry measurements of 36Cl, 129I and osmium isotopes in diverse natural samples. Phil. Trans. Roy. Soc. Lond. A 323, 103–19.Google Scholar
Graf, T., Kohl, C. P., Marti, K. and Nishiizumi, K. (1991). Cosmic-ray produced neon in Antarctic rocks. Geophys. Res. Lett. 18, 203–6.Google Scholar
Granger, D. E., Fabel, D. and Palmer, A. N. (2001). PliocenePleistocene incision of the Green river, Kentucky, determined from radioactive decay of cosmogenic 26Al and 10Be in Mammoth Cave sediments. GSA Bulletin 113, 825–36.Google Scholar
Granger, D. E., Gibbon, R. J., Kuman, K. et al. (2015). New cosmogenic burial ages for Sterkfontein Member 2 Australopithecus and Member 5 Oldowan. Nature 522, 85–8.Google Scholar
Granger, D. E. and Muzikar, P. F. (2001). Dating sediment burial with in situ-produced cosmogenic nuclides: theory, techniques, and limitations. Earth Planet. Sci. Lett. 188, 269–81.Google Scholar
Gray, L. J., Beer, J., Geller, M. et al. (2010). Solar influences on climate. Rev. Geophys. 48, RG4001, 153.Google Scholar
Guyodo, Y. and Valet, J. P. (1996). Relative variations in geomagnetic intensity from sedimentary records: the past 200,000 years. Earth Planet. Sci. Lett. 143, 2336.Google Scholar
Hammer, C. U., Clausen, H. B. and Tauber, H. (1986). Ice-core dating of the Pleistocene/Holocene boundary applied to a calibration of the 14C time scale. Radiocarbon 28, 284–91.CrossRefGoogle Scholar
Heisinger, B., Lal, D., Jull, A. T. et al. (2002). Production of selected cosmogenic radionuclides by muons: 1. Fast muons. Earth Planet. Sci. Lett. 200, 345–55.Google Scholar
Henning, W. (1987). Accelerator mass spectrometry of heavy elements: 36Cl to 205Pb. Phil. Trans. Roy. Soc. Lond. A 323, 8799.Google Scholar
Hillam, J., Groves, C. M., Brown, D. M. et al. (1990). Dendrochronology of the English Neolithic. Antiquity 64, 210–20.Google Scholar
Hoffmann, D. L., Beck, J. W., Richards, D. A. et al. (2010). Towards radiocarbon calibration beyond 28 ka using speleothems from the Bahamas. Earth Planet. Sci. Lett. 289, 110.Google Scholar
Hofmann, H. J., Beer, J., Bonani, G. et al. (1987). 10Be half-life and AMS-standards. Nucl. Instr. Meth. in Phys. Res. B 29, 32–6.Google Scholar
Hua, Q., Barbetti, M., Fink, D. et al. (2009). Atmospheric 14C variations derived from tree rings during the early Younger Dryas. Quaternary Sci. Rev. 28, 2982–90.CrossRefGoogle Scholar
Huber, B. (1970). Dendrochronology of central Europe. In: Olsson, I. U. (Ed.) Radiocarbon Variations and Absolute Chronology, Proc. 12th Nobel Symp. Wiley, pp. 233–5.Google Scholar
Hughen, K. A., Overpeck, J. T., Lehman, S. J. et al. (1998). Deglacial changes in ocean circulation from an extended radiocarbon calibration. Nature 391, 65–8.Google Scholar
Jull, A. T. and Burr, G. S. (2006). Accelerator mass spectrometry: Is the future bigger or smaller? Earth Planet. Sci. Lett. 243, 305–25.Google Scholar
Kamen, M. D. (1963). Early history of carbon-14. Science 140, 584–90.Google Scholar
Keigwin, L. D. and Lehman, S. J. (2015). Radiocarbon evidence for a possible abyssal front near 3.1 km in the glacial equatorial Pacific Ocean. Earth Planet. Sci. Lett. 425, 93104.CrossRefGoogle Scholar
Kelly, P. M. and Wigley, T. M. L. (1992). Solar cycle length, greenhouse forcing and global climate. Nature 360, 328–30.Google Scholar
Key, R., Quay, P. D., Jones, G. A. et al. (1996). WOCE AMS radiocarbon I: Pacific Ocean results (P6, P16, P17). Radiocarbon 38, 425518.Google Scholar
Kieser, W. E., Beukens, R. P., Kilius, L. R., Lee, H. W. and Litherland, A. E. (1986). Isotrace radiocarbon analysis – equipment and procedures. Nucl. Instr. Meth. in Phys. Res. B 15, 718–21.Google Scholar
Kitagawa, H. and van der Plicht, J. (1998). Atmospheric radiocarbon calibration to 45,000 yr B.P.: Late glacial fluctuations and cosmogenic isotope production. Science 279, 1187–90.Google Scholar
Klein, J., Fink, D., Middleton, R., Nishiizumi, K. and Arnold, J. (1991). Determination of the half-life of 41Ca from measurements of Antarctic meteorites. Earth Planet. Sci. Lett. 103, 7983.Google Scholar
Kober, F., Ivy-Ochs, S., Schlunegger, F. et al. (2007). Denudation rates and a topography-driven rainfall threshold in northern Chile: multiple cosmogenic nuclide data and sediment yield budgets. Geomorphology 83, 97120.CrossRefGoogle Scholar
Kok, Y. S. (1999). Climatic influence in NRM and 10Be-derived geomagnetic paleointensity data. Earth Planet. Sci. Lett. 166, 105–19.Google Scholar
Korschinek, G., Bergmaier, A., Faestermann, T. et al. (2010). A new value for the half-life of 10Be by heavy-ion elastic recoil detection and liquid scintillation counting. Nucl. Instr. Meth. in Phys. Res. B 268, 187–91.Google Scholar
Krivova, N. A., Vieira, L. E. A. and Solanki, S. K. (2010). Reconstruction of solar spectral irradiance since the Maunder minimum. J. Geophys. Res.: Space Phys. 115, A12112, 111.Google Scholar
Kromer, B., Friedrich, M., Hughen, K. A. et al. (2004). Late glacial 14C ages from a floating, 1382-ring pine chronology. Radiocarbon 46, 1203–9.Google Scholar
Ku, T. L., Wang, L., Luo, S. and Southon, J. R. (1995). 26Al in seawater and 26Al/10Be as paleo-flux tracer. Geophys. Res. Lett. 22, 2163–6.CrossRefGoogle Scholar
Ku, T. L., Kusakabe, M., Measures, C. I. et al. (1990). Beryllium isotope distribution in the western North Atlantic: a comparison to the Pacific. Deep-Sea Res. 37, 795808.Google Scholar
Kusakabe, M., Ku, T. L., Southon, J. R. et al. (1987). The distribution of 10Be and 9Be in ocean water. Nucl. Instr. Meth. in Phys. Res. B 29, 306–10.Google Scholar
Labrie, D. and Reid, J. (1981). Radiocarbon dating by infrared laser spectroscopy. Appl. Phys. 24, 3816.CrossRefGoogle Scholar
Laj, C., Kissel, C., Mazaud, A. et al.. (2002). Geomagnetic field intensity, North Atlantic Deep Water circulation and atmospheric Δ 14 C during the last 50 kyr. Earth Planet. Sci. Lett. 200, 177–90.Google Scholar
Laken, B. A., Pallé, E., Čalogović, J. and Dunne, E. M. (2012). A cosmic ray–climate link and cloud observations. J. Space Weather Space Climate 2, A18.Google Scholar
Lal, D. (1988). In situ-produced cosmogenic isotopes in terrestrial rocks. Ann. Rev. Earth Planet. Sci. 16, 355–88.Google Scholar
Lal, D. (1991). Cosmic ray labeling of erosion surfaces: in situ nuclide production rates and erosion models. Earth Planet. Sci. Lett. 104, 424–39.Google Scholar
Lal, D. (2007). Recycling of cosmogenic nuclides after their removal from the atmosphere; special case of appreciable transport of 10Be to polar regions by aeolian dust. Earth Planet. Sci. Lett. 264, 177–87.Google Scholar
Lal, D. and Peters, B. (1967). Cosmic-ray produced radioactivity on the Earth. In: Handbook of Physics 46/2. Springer, pp. 551612.Google Scholar
Lao, Y., Anderson, R. F., Broecker, W. S. et al. (1992). Increased production of cosmogenic 10Be during the Last Glacial Maximum. Nature 357, 576–8.Google Scholar
Lean, J., Skumanich, A. and White, O. (1992). Estimating the Sun's radiative output during the Maunder Minimum. Geophys. Res. Lett. 19, 1591–4.Google Scholar
Lee, H. W., Galindo-Uribarri, A., Chang, K. H., Kilius, L. R. and Litherland, A. E. (1984). The 12CH2+2 molecule and radiocarbon dating by accelerator mass spectrometry. Nucl. Instrum. Meth. in Phys. Res. B 5, 208–10.Google Scholar
Libby, W. F. (1952). Radiocarbon Dating. University of Chicago Press, 124 pp.Google Scholar
Libby, W. F. (1970). Ruminations on radiocarbon dating. In: Olsson, I. U. (Ed.) Radiocarbon Variations and Absolute Chronology, Proc. 12th Nobel Symp. Wiley, pp. 629–40.Google Scholar
Litherland, A. E. (1987). Fundamentals of accelerator mass spectrometry. Phil. Trans. Roy. Soc. Lond. A 323, 521.Google Scholar
Litherland, A. E., Zhao, X. L. and Kieser, W. E. (2011). Mass spectrometry with accelerators. Mass Spec. Rev. 30, 1037–72.Google Scholar
Liu, B., Phillips, F. M., Fabryka-Martin, J. T., Fowler, M. M. and Stone, W. D. (1994). Cosmogenic 36Cl accumulation in unstable landforms 1. effects of the thermal neutron distribution. Water Resour. Res. 30, 3115–25.Google Scholar
Lockwood, M. (2010). Solar change and climate: an update in the light of the current exceptional solar minimum. Proc. Roy. Soc. A 466, 303–29.Google Scholar
Mahara, Y., Habermehl, M. A., Hasegawa, T. et al. (2009). Groundwater dating by estimation of groundwater flow velocity and dissolved 4He accumulation rate calibrated by 36Cl in the Great Artesian Basin, Australia. Earth Planet. Sci. Lett. 287, 4356.Google Scholar
Mangini, A., Segl, M., Bonani, G. et al. (1984). Mass-spectrometric 10Be dating of deep-sea sediments applying the Zurich tandem accelerator. Nucl. Instr. Meth. in Phys. Res. B 5, 353–8.Google Scholar
Mann, M. E. and Park, J. (1994). Global-scale modes of surface temperature variability on interannual to century timescales. J. Geophys. Res.: Atm. 99, 25819–33.Google Scholar
Marchitto, T. M., Lehman, S. J., Ortiz, J. D., Flückiger, J. and van Geen, A. (2007). Marine radiocarbon evidence for the mechanism of deglacial atmospheric CO2 rise. Science, 316, 1456–9.Google Scholar
Marrero, S. M., Phillips, F. M., Caffee, M. W. and Gosse, J. C. (2016). CRONUS-Earth cosmogenic 36Cl calibration. Quaternary Geochron. 31, 199219.Google Scholar
Maunder, E. W. (1922). The sun and sun-spots, 18201920. Monthly Notices Roy. Astron. Soc. 82, 534–43.Google Scholar
Maycock, A. C., Ineson, S., Gray, L. J. et al. (2015). Possible impacts of a future grand solar minimum on climate: Stratospheric and global circulation changes. J. Geophys. Res.: Atm. 120, 9043–58.Google Scholar
Mazaud, A., Laj, C., Bard, E., Arnold, M. and Tric, E. (1991). Geomagnetic field control of 14C production over the last 80 ka: implications for the radiocarbon time-scale. Geophys Res. Lett. 18, 1885–8.Google Scholar
McCorkell, R., Fireman, E. L. and Langway, C. C. (1967). Aluminium-26 and Beryllium-10 in Greenland Ice. Science 158, 1690–2.Google Scholar
McCracken, K. G., Beer, J. and Steinhilber, F. (2014). Evidence for planetary forcing of the cosmic ray intensity and solar activity throughout the past 9400 years. Solar Phys. 289, 3207–29.Google Scholar
Measures, C. I. and Edmond, J. M. (1982). Beryllium in the water column of the central North Pacific. Nature 297, 51–3.Google Scholar
Merrill, J. R., Lyden, E. F. X., Honda, M. and Arnold, J. R. (1960). The sedimentary geochemistry of the beryllium isotopes. Geochim. Cosmochim. Acta 18, 108–29.Google Scholar
Middleton, R. and Klein, J. (1987). 26Al: measurement and applications. Phil. Trans. Roy. Soc. Lond. A 323, 121–43.Google Scholar
Mitchell, S. G., Matmon, A., Bierman, P. R. et al. (2001). Displacement history of a limestone normal fault scarp, northern Israel, from cosmogenic 36Cl. J. Geophys. Res. 106, 4247–64.Google Scholar
Monaghan, M. C., Klein, J. and Measures, C. I. (1988). The origin of 10Be in island-arc volcanic rocks. Earth Planet. Sci. Lett. 89, 288–98.Google Scholar
Monnin, E., Indermühle, A., Dällenbach, A. et al. (2001). Atmospheric CO2 concentrations over the last glacial termination. Science 291, 112–14.Google Scholar
Mook, W. G. and Streurman, H. J. (1983). Physical and chemical aspects of radiocarbon dating. P.A.C.T. (Physical And Chemical Techniques in Archaeology) 8, 3155.Google Scholar
Moran, J. E., Fehn, U. and Teng, R. T. (1998). Variations in 129I/127I ratios in recent marine sediments: evidence for a fossil organic component. Chem. Geol. 152, 193203.Google Scholar
Morris, J. D., Leeman, W. P. and Tera, F. (1990). The subducted component in island arc lavas: constraints from Be isotopes and B–Be systematics. Nature 344, 31–6.CrossRefGoogle ScholarPubMed
Morris, J. D. and Tera, F. (1989). 10Be and 9Be in mineral separates and whole-rocks from island arcs: implications for sediment subduction. Geochim. Cosmochim. Acta 53, 3197–206.Google Scholar
Naylor, J. C. and Smith, A. F. M. (1988). An archaeological inference problem. J. Amer. Stat. Assoc. 83, 588–95.Google Scholar
Neff, U., Burns, S. J., Mangini, A. et al. (2001). Strong coherence between solar variability and the monsoon in Oman between 9 and 6 kyr ago. Nature 411, 290–3.Google Scholar
Nishiizumi, K., Arnold, J. R., Elmore, D et al. (1979). Measurements of 36Cl in Antarctic meteorites and Antarctic ice using a van de Graaff accelerator. Earth Planet. Sci. Lett. 45, 285–92.Google Scholar
Nishiizumi, K., Elmore, D. and Kubik, P. W. (1989a). Update on terrestrial ages of Antarctic meteorites. Earth Planet. Sci. Lett. 93, 299313.Google Scholar
Nishiizumi, K., Kohl, C. P., Arnold, J. R. et al. (1991a). Cosmic-ray produced 10Be and 26Al in Antarctic rocks: exposure and erosion history. Earth Planet. Sci. Lett. 104, 440–54.Google Scholar
Nishiizumi, K., Kohl, C. P., Shoemaker, J. R. et al. (1991b). In situ 10Be and 26Al exposure ages at Meteor Crater, Arizona. Geochim. Cosmochim. Acta 55, 2699–703.Google Scholar
Nishiizumi, K., Lal, D., Klein, J., Middleton, R. and Arnold, J. R. (1986). Production of 10Be and 26Al by cosmic rays in terrestrial quartz in situ and implications for erosion rates. Nature 319, 134–6.Google Scholar
Nishiizumi, K., Winterer, E. L., Kohl, C. P. et al. (1989b). Cosmic ray production rates of 10Be and 26Al in quartz from glacially polished rocks. J. Geophys. Res. 94, 17 907–15.Google Scholar
Nydal, R. (2000). Radiocarbon in the ocean. Radiocarbon 42, 8198.Google Scholar
Oeschger, H., Houtermans, J., Loosli, H. and Wahlen, M. (1970). The constancy of cosmic radiation from isotope studies in meteorites and on the Earth. In: Olsson, I. U. (Ed.) Radiocarbon Variations and Absolute Chronology, Proc. 12th Nobel Symp. Wiley, pp. 471–98.Google Scholar
Oktay, S. D., Santschi, P. H., Moran, J. E. and Sharma, P. (2000). The 129I bomb pulse recorded in Mississippi River Delta sediments: results from isotopes of I, Pu, Cs, Pb, and C. Geochim. Cosmochim. Acta 64, 989–96.Google Scholar
Olsson, I. U., El-Daoushy, M. F. A. F., Abdel-Mageed, A. I. and Klasson, M. (1974). A comparison of different methods for pretreatment of bones. Geol. Foren. Stockh. Forhandl. 96, 171–81.Google Scholar
Olsson, I. U. (1980). 14C in extractives from wood. Radiocarbon 22, 515–24.CrossRefGoogle Scholar
Ostlund, H. G. and Rooth, C. G. H. (1990). The North Atlantic tritium and radiocarbon transients 19721983. J. Geophys. Res. 95, 20 147–65.Google Scholar
Pavich, M. J., Brown, L., Valette-Silver, J. N., Klein, J. and Middleton, R. (1985). 10Be analysis of a Quaternary weathering profile in the Virginia Piedmont. Geology 13, 3941.Google Scholar
Pearson, G. W., Pilcher, J. R., Baillie, M. G. L. and Hillam, J. (1977). Absolute radiocarbon dating using a low altitude European tree-ring calibration. Nature 270, 25–8.Google Scholar
Phillips, F. M., Argento, D. C., Balco, G. et al. (2016). The CRONUS-Earth project: a synthesis. Quaternary Geochron. 31, 119–54.Google Scholar
Phillips, F. M., Leavy, B. D., Jannik, N. O., Elmore, D. and Kubik, P. W. (1986). The accumulation of cosmogenic chlorine-36 in rocks: a method for surface exposure dating. Science 231, 41–3.Google Scholar
Phillips, F. M., Mattick, J. L. and Duval, T. A. (1988). Chlorine 36 and tritium from nuclear weapons fallout as tracers for long-term liquid and vapour movement in desert soils. Water Resour. Res. 24, 1877–91.Google Scholar
Phillips, F. M., Zreda, M. G., Gosse, J. C. et al. (1997). Cosmogenic 36Cl and 10Be ages of Quaternary glacial and fluvial deposits of the Wind River Range, Wyoming. GSA Bull. 109, 1453–63.Google Scholar
Plummer, M. A., Phillips, F. M., Fabryka-Martin, J. et al. (1997). Chlorine-36 in fossil rat urine: an archive of cosmogenic nuclide deposition during the past 40,000 years. Science 277, 538–41.Google Scholar
Purser, K. H., Liebert, R. B., Litherland, A. E. et al. (1977). An attempt to detect stable N-ions from a sputter ion source and some implications of the results for the design of tandems for ultra-sensitive carbon analysis. Rev. Phys. Appl. 12, 1487–92.Google Scholar
Raisbeck, G. M., Yiou, F., Fruneau, M., Lieuvin, M. and Loiseaux, J. M. (1978). Measurements of 10Be in 1,000- and 5,000-year-old Antarctic ice. Nature 275, 731–3.Google Scholar
Raisbeck, G. M., Yiou, F., Fruneau, M. et al. (1979). Deposition rate and seasonal variations in precipitation of cosmogenic 10Be. Nature 282, 279–80.CrossRefGoogle Scholar
Raisbeck, G. M., Yiou, F., Fruneau, M. et al. (1981a). Cosmogenic 10Be concentrations in Antarctic ice during the past 30,000 years. Nature 292, 825–6.Google Scholar
Raisbeck, G. M., Yiou, F., Fruneau, M. et al. (1980). 10Be concentration and residence time in the deep ocean. Earth Planet. Sci. Lett. 51, 275–8.Google Scholar
Raisbeck, G. M., Yiou, F. and Zhou, S. Z. (1994). Palaeointensity puzzle. Nature 371, 207–8.Google Scholar
Ralph, E. K. (1971). Carbon-14 dating. In: Michael, H. N. and Ralph, E. K. (Eds) Dating Techniques for the Archaeologist. M.I.T. Press, pp. 148.Google Scholar
Ralph, E. K. and Michael, H. N. (1974). Twenty-five years of radiocarbon dating. Amer. Scient. 62, 553–60.Google Scholar
Ramsey, C. B. (1995). Radiocarbon calibration and analysis of stratigraphy; the OxCal program. Radiocarbon 37, 425–30.Google Scholar
Reimer, P. J., Bard, E., Bayliss, A. et al. (2013). IntCal13 and Marine13 radiocarbon age calibration curves 0–50,000 years cal BP. Radiocarbon 55, 1869–87.Google Scholar
Reyss, J. L., Yokoyama, Y. and Guichard, F. (1981). Production cross sections of 26Al, 22Na, 7Be from argon and of 10Be, 7Be from nitrogen: implications for the production rates of 26Al and 10Be in the atmosphere. Earth Planet. Sci. Lett. 53, 203–10.Google Scholar
Roberts, M. L., Burton, J. R., Elder, K. L. et al. (2010). A high-performance 14C accelerator mass spectrometry system. Radiocarbon 52, 226–35.Google Scholar
Robinson, C., Raisbeck, G. M., Yiou, F., Lehman, B. and Laj, C. (1995). The relationship between 10Be and geomagnetic field strength records in central North Atlantic sediments during the last 80 ka. Earth Planet. Sci. Lett. 136, 551–7.Google Scholar
Scafetta, N. (2014). Discussion on the spectral coherence between planetary, solar and climate oscillations: a reply to some critiques. Astrophys. Space Sci. 354, 275–99.Google Scholar
Scafetta, N. (2016). High resolution coherence analysis between planetary and climate oscillations. Advances in Space Res. 57, 2121–35.CrossRefGoogle Scholar
Scafetta, N. and Willson, R. C. (2013). Planetary harmonics in the historical Hungarian aurora record (1523–1960). Planet. Space Sci. 78, 3844.Google Scholar
Schlesinger, M. E. and Ramankutty, N. (1992). Implications for global warming of intercycle solar irradiance variations. Nature 360, 330–3.Google Scholar
Shackleton, N. J., Duplessy, J.-C., Arnold, M. et al. (1988). Radiocarbon age of last glacial Pacific deep water. Nature 335, 708–11.Google Scholar
Sharma, M. (2002). Variations in solar magnetic activity during the last 200 000 years: is there a Sun–climate connection? Earth Planet. Sci. Lett. 199, 459–72.Google Scholar
Shepard, M. K., Arvidson, R. E., Caffee, M., Finkel, R. and Harris, L. (1995). Cosmogenic exposure ages of basalt flows: Lunar Crater volcanic field, Nevada. Geology 23, 21–4.Google Scholar
Siegenthaler, U. and Sarmiento, J. L. (1993). Atmospheric carbon dioxide and the ocean. Nature 365, 119–25.Google Scholar
Sigmarsson, O., Condomines, M., Morris, J. D. and Harmon, R. S. (1990). Uranium and 10Be enrichments by fluids in Andean arc magmas. Nature 346, 163–5.Google Scholar
Sikes, E. L., Samson, C. R., Guilderson, T. P. and Howard, W. R. (2000). Old radiocarbon ages in the southwest Pacific Ocean during the last glacial period and deglaciation. Nature 405, 555–9.Google Scholar
Simpson, J. A. (1951). Neutrons produced in the atmosphere by cosmic radiations. Phys. Rev. 83, 1175–88.Google Scholar
Skinner, L. C., Fallon, S., Waelbroeck, C., Michel, E. and Barker, S. (2010). Ventilation of the deep Southern Ocean and deglacial CO2 rise. Science 328, 1147–51.Google Scholar
Skinner, L., McCave, I. N., Carter, L. et al. (2015). Reduced ventilation and enhanced magnitude of the deep Pacific carbon pool during the last glacial period. Earth Planet. Sci. Lett. 411, 4552.Google Scholar
Snyder, G., Aldahan, A. and Possnert, G. (2010). Global distribution and long-term fate of anthropogenic 129I in marine and surface water reservoirs. Geochem. Geophys. Geosys. 11, Q04010, 119.Google Scholar
Snyder, G. T. and Fehn, U. (2002). Origin of iodine in volcanic fluids: 129I results from the Central American Volcanic Arc. Geochim. Cosmochim. Acta 66, 3827–38.Google Scholar
Snyder, G. T., Fehn, U. and Goff, F. (2002). Iodine isotope ratios and halide concentrations in fluids of the Satsuma–Iwojima volcano, Japan. Earth,Planets Space 54, 265–73.Google Scholar
Steinhilber, F., Abreu, J. A., Beer, J. et al. (2012). 9,400 years of cosmic radiation and solar activity from ice cores and tree rings. Proc. Nat. Acad. Sci. 109, 5967–71.Google Scholar
Steinhilber, F., Beer, J. and Fröhlich, C. (2009). Total solar irradiance during the Holocene. Geophys. Res. Lett. 36 L19704, 15.Google Scholar
Stensland, G. J., Brown, L., Klein, J. and Middleton, R. (1983). Beryllium-10 in rain. EOS 64, 283 (abs.).Google Scholar
Stocker, T. F., Qin, D., Plattner, G. K. et al. Eds.(2014). IPCC, Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press.Google Scholar
Stone, J. O., Allan, G. L., Fifield, L. K. and Cresswell, R. G. (1996). Cosmogenic chlorine-36 from calcium spallation. Geochim. Cosmochim. Acta 60, 679–92.Google Scholar
Stuiver, M. (1961). Variations in radiocarbon concentration and sunspot activity. J. Geophys. Res. 66, 273–6.Google Scholar
Stuiver, M. (1965). Carbon-14 content of 18th- and 19th-century wood: variations correlated with sunspot activity. Science 149, 533–4.Google Scholar
Stuiver, M., Braziunas, T. F., Becker, B. and Kromer, B. (1991). Climatic, solar, oceanic and geomagnetic influences on late-glacial and Holocene atmospheric 14C/12C change. Quat. Res. 35, 124.Google Scholar
Stuiver, M. and Pearson, G. W. (1986). High-precision radiocarbon time-scale calibration from the present to 500 BC. Radiocarbon 28, 805–38.Google Scholar
Stuiver, M. and Quay, P. D. (1981). Atmospheric 14C changes resulting from fossil fuel CO2 release and cosmic-ray flux variability. Earth Planet. Sci. Lett. 53, 349–62.Google Scholar
Stuiver, M., Quay, P. D. and Ostlund, H. G. (1983). Abyssal water carbon-14 distribution and the age of the world oceans. Science 219, 849–51.Google Scholar
Suess, H. E. (1955). Radiocarbon concentrations in modern wood. Science 122, 415–17.Google Scholar
Suess, H. E. (1965). Secular variations of the cosmic-ray-produced carbon 14 in the atmosphere and their interpretations. J. Geophys. Res. 70, 5937–52.Google Scholar
Suess, H. E. (1970). Bristlecone-pine calibration of the radiocarbon time-scale 5200 B.C. to the present. In: Olsson, I. U. (Ed.) Radiocarbon Variations and Absolute Chronology, Proc. 12th Nobel Symp. Wiley, pp. 303–11.Google Scholar
Suess, H. E. and Strahm, C. (1970). The Neolithic of Auvernier, Switzerland. Antiquity 44, 91–9.Google Scholar
Sun, B. and Bradley, R. S. (2002). Solar influences on cosmic rays and cloud formation: A reassessment. J. Geophys. Res.: Atm. 107 AAC5, 112.Google Scholar
Suter, M., Jacob, S. W. A. and Synal, H. A. (2000). Tandem AMS at sub-MeV energies–Status and prospects. Nucl. Instr. Meth. in Phys. Res. B 172, 144–51.Google Scholar
Svensmark, H. and Friis-Christensen, E. (1997). Variation of cosmic ray flux and global cloud coverage – a missing link in solar–climate relationships. J. Atm. Solar–Terrest. Phys. 59, 1225–32.Google Scholar
Tanaka, S. and Inoue, T. (1979). 10Be dating of North Pacific sediment cores up to 2.5 million years B.P. Earth Planet. Sci. Lett. 45, 181–7.Google Scholar
Tauber, H. (1970). The Scandinavian varve chronology and C-14 dating. In: Olsson, I. U. (Ed.) Radiocarbon Variations and Absolute Chronology, Proc. 12th Nobel Symp. Wiley, pp. 173–96.Google Scholar
Tera, F., Brown, L., Morris, J. et al. (1986). Sediment incorporation in island-arc magmas: inferences from 10Be. Geochim. Cosmochim. Acta 50, 535–50.Google Scholar
Thellier, E. O. (1941). Sur la verification d'une methode permettant de determiner l'intensite du champ magnetique terrestre dans le passe. Compte Rendu Acad. Sci. Paris 212, 281.Google Scholar
Tobias, S. M. (1996). Grand minimia in nonlinear dynamos. Astron. Astrophys 307, L2124.Google Scholar
Torgersen, T., Habermehl, M. A., Phillips, F. M. et al. (1991). Chlorine 36 dating of very old groundwater 3. Further studies in the Great Artesian Basin, Australia. Water. Resources. Res. 27, 3201–13.Google Scholar
Turekian, K. K., Cochran, J. K., Krishnaswami, S. et al. (1979). The measurement of 10Be in manganese nodules using a tandem van de Graaff accelerator. Geophys. Res. Lett. 6, 417–20.Google Scholar
Ullman, W. J. and Aller, R. C. (1980). Dissolved iodine flux from estuarine sediments and implications for the enrichment of iodine at the sediment water interface. Geochim. Cosmochim. Acta 44, 1177–84.Google Scholar
Usoskin, I. G., Solanki, S. K. and Kovaltsov, G. A. (2007). Grand minima and maxima of solar activity: new observational constraints. Astron. Astrophys. 471, 301–9.Google Scholar
Vieira, L. E. A., Solanki, S. K., Krivova, N. A. and Usoskin, I. (2011). Evolution of the solar irradiance during the Holocene. Astron. Astrophys. 531, A6 120.Google Scholar
von Blankenburg, F., O'Nions, R. K., Belshaw, N. S., Gibb, A. and Hein, J. R. (1996). Global distribution of beryllium isotopes in deep ocean water as derived from FeMn crusts. Earth Planet. Sci. Lett. 141, 213–26.Google Scholar
Wagner, G., Laj, C., Beer, J. et al. (2001). Reconstruction of the paleoaccumulation rate of central Greenland during the last 75 ka using the cosmogenic radionuclides 36Cl and 10Be and geomagnetic field intensity data. Earth Planet. Sci. Lett. 193, 515–21.Google Scholar
Wang, L., Ku, T. L., Luo, S., Southon, J. R. and Kusakabe, M. (1996). 26Al10Be systematics in deep-sea sediments. Geochim. Cosmochim. Acta 60, 109–19.Google Scholar
Wolfli, W. (1987). Advances in accelerator mass spectrometry. Nucl. Instr. Meth. in Phys. Res. B 29, 113.Google Scholar
Yamazaki, T. and Oda, H. (2002). Orbital influence on Earth's magnetic field: 100,000-year periodicity in inclination. Science 295, 2435–8.Google Scholar
Yiou, F., Raisbeck, G. M., Bourles, D., Lorius, C. and Barkov, N. I. (1985). 10Be in ice at Vostok Antarctica during the last climatic cycle. Nature 316, 616–17.Google Scholar
Yokoyama, Y., Guichard, F., Reyss, J. L., Van, N. H. (1978). Oceanic residence times of dissolved beryllium and aluminium deduced from cosmogenic tracers 10Be and 26Al. Science 201, 1016–17.Google Scholar
You, C. F., Lee, T. and Li, Y. H. (1989). The partition of Be between soil and water. Chem. Geol. 77, 105–18.Google Scholar
Zreda, M. and Noller, J. S. (1998). Ages of prehistoric earthquakes revealed by cosmogenic chlorine-36 in a bedrock fault scarp at Hebgen Lake. Science 282, 1097–9.Google Scholar
Zreda, M. G., Phillips, F. M., Elmore, D. et al. (1991). Cosmogenic chlorine-36 production rates in terrestrial rocks. Earth Planet. Sci. Lett. 105, 94109.Google Scholar
Zreda, M. G. and Phillips, F. M. (1995). Surface exposure dating by cosmogenic chlorine-36 accumulation. In: Beck., C. (Ed.) Dating in Exposed and Surface Contexts, University of New Mexico Press, pp. 161–83.Google Scholar
Zreda, M. G., Phillips, F. M. and Elmore, D. (1994). Cosmogenic 36Cl accumulation in unstable landforms 2. Simulations and measurements on eroding moraines. Water Resour. Res. 30, 3127–36.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.

  • Cosmogenic Nuclides
  • Alan P. Dickin, McMaster University, Ontario
  • Book: Radiogenic Isotope Geology
  • Online publication: 01 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781316163009.015
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.

  • Cosmogenic Nuclides
  • Alan P. Dickin, McMaster University, Ontario
  • Book: Radiogenic Isotope Geology
  • Online publication: 01 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781316163009.015
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.

  • Cosmogenic Nuclides
  • Alan P. Dickin, McMaster University, Ontario
  • Book: Radiogenic Isotope Geology
  • Online publication: 01 February 2018
  • Chapter DOI: https://doi.org/10.1017/9781316163009.015
Available formats
×