Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-mp689 Total loading time: 0 Render date: 2024-04-16T04:19:07.180Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  13 August 2009

Graham P. Weedon
Affiliation:
University of Luton
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
Time-Series Analysis and Cyclostratigraphy
Examining Stratigraphic Records of Environmental Cycles
, pp. 221 - 251
Publisher: Cambridge University Press
Print publication year: 2003

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

Algeo, T. J. (1993). Quantifying stratigraphic completeness: a probabilistic approach using paleomagnetic data. J. Geol. 101: 421–33CrossRefGoogle Scholar
Algeo, T. J. and Woods, A. D. (1994). Microstratigraphy of the Lower Mississippian Sunbury Shale: a record of solar-modulated climatic cyclicity. Geology 22: 795–82.3.CO;2>CrossRefGoogle Scholar
Allen, J. R. L. (1981). Lower Cretaceous tides revealed by cross-bedding with mud drapes. Nature 289: 579–81CrossRefGoogle Scholar
Allen, M. R. and Smith, L. A. (1996). Monte Carlo SSA: detecting irregular oscillations in the presence of colored noise. J. Clim. 9: 373–4042.0.CO;2>CrossRefGoogle Scholar
Allen, P. A. and Homewood, P. (1984). Evolution and mechanics of a Miocene tidal sandwave. Sedimentology 31: 63–81CrossRefGoogle Scholar
Alley, R. B. and MacAyeal, D. R. (1994). Ice-rafted debris associated with binge/purge oscillations of the Laurentide Ice Sheet. Paleoceanography 9: 503–11CrossRefGoogle Scholar
Alley, R. B., Shuman, C. A., Meese, D. A., Gow, A. J., Taylor, K. C., Cuffey, K. M., Fitzpatrick, J. J., Grootes, P. M., Zielinski, G. A., Ram, M., Spinelli, G. and Elser, B. (1997). Visual-stratigraphic dating of the GISP2 ice core: basis, reproducibility, and application. J. Geophys. Res. 102: 26,367–81CrossRefGoogle Scholar
Alley, R. B., Anandakrishnan, S. and Jung, P. (2001). Stochastic resonance in the North Atlantic. Paleoceanography 16: 190–8CrossRefGoogle Scholar
Anders, M. H., Krueger, S. W. and Sadler, P. M. (1987). A new look at sedimentation rates and the completeness of the stratigraphic record. J. Geol. 95: 1–14CrossRefGoogle Scholar
Anderson, D. M. (2001). Attenuation of millennial-scale events by bioturbation in marine sediments. Paleoceanography 16: 352–7CrossRefGoogle Scholar
Anderson, R. Y. (1961). Solar-terrestrial climatic patterns in varved sediments. Ann. NY Acad. Sci. USA 95: 424–35CrossRefGoogle Scholar
Anderson, R. Y. (1982). A long geoclimatic record from the Permian. J. Geophys. Res. 87: 7285–94CrossRefGoogle Scholar
Anderson, R. Y. (1984). Orbital forcing of evaporite sedimentation. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 147–62
Anderson, R. Y. (1986). The varve microcosm: propagator of cyclic bedding. Paleoceanography 1: 373–82CrossRefGoogle Scholar
Anderson, R. Y. (1992a). Long term changes in the frequency of occurrence of El Niño events. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Eds: H. F. Diaz and V. Markgraf. Cambridge University Press, Cambridge, pp. 193–200
Anderson, R. Y. (1992b). Possible connection between surface winds, solar activity and the Earth's magnetic field. Nature 358: 51–3CrossRefGoogle Scholar
Anderson, R. Y. (1996). Seasonal sedimentation: a framework for reconstructing climatic and environmental change. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication, No. 116. The Geological Society, London, pp. 1–15CrossRef
Anderson, R. Y. and Dean, W. E. (1988). Lacustrine varve formation through time. Palaeogeog. Palaeoclim. Palaeoecol. 62: 215–35CrossRefGoogle Scholar
Anderson, R. Y. and Koopmans, L. H. (1963). Harmonic analysis of varve time series. J. Geophys. Res. 68: 877–93CrossRefGoogle Scholar
Andrews, J. T., Jennings, A. E., Kerwin, M., Kirby, M., Manley, W., Miller, G. H., Bond, G. and MacLean, B. (1995). A Heinrich-like event, H-0 (DC-0): source(s) for detrital carbonate in the North Atlantic during the Younger Dryas chronozone. Paleoceanography 10: 943–52CrossRefGoogle Scholar
Andrews, J. T., Barber, D. C. and Jennings, A. E. (1999). Errors in generating time series and in dating events at Late Quaternary millennial (radiocarbon) time-scales: examples from Baffin Bay, NW Labrador Sea, and East Greenland. In: Mechanisms of Global Climate Change at Millennial Time Scales, Eds: P. U. Clark, R. S. Webb and L. D. Keigwin. Geophysical Monograph 112. American Geophysical Union, Washington, pp. 23–33CrossRef
Appenzeller, C., Stocker, T. F. and Anklin, M. (1998). North Atlantic Oscillation dynamics recorded in Greenland ice cores. Science 282: 446–9CrossRefGoogle ScholarPubMed
Archer, A. W. (1996). Reliability of lunar orbital periods extracted from ancient cyclic tidal rhythmites. Earth Planet. Sci. Lett. 141: 1–10CrossRefGoogle Scholar
Archer, A. W. and Johnson, T. W. (1997). Modelling of cyclic tidal rhythmites (Carboniferous of Indiana and Kansas, Precambrian of Utah, USA) as a basis for reconstruction of intertidal positioning and palaeotidal regimes. Sedimentology 44: 991–1010CrossRefGoogle Scholar
Archer, A. W., Kuecher, G. J. and Kvale, E. P. (1995). The role of tidal-velocity asymmetries in the deposition of silty tidal rhythmites (Carboniferous, eastern Interior Coal Basin, U.S.A.). J. Sed. Res. A65: 408–16Google Scholar
Arz, H. W., Pätzold, J. and Wefer, G. (1998). Correlated millennial-scale changes in surface hydrography and terrigenous sediment yield inferred from last-glacial marine deposits off Northeastern Brazil. Quat. Res. 50: 157–66CrossRefGoogle Scholar
Baker, P. A., Rigsby, C. A., Sltzer, G. O., Fritz, S. C., Lowenstein, T. K., Bacher, N. P. and Veliz, C. (2001). Tropical climate changes at millennial and orbital timescales on the Bolivian Altiplano. Nature 409: 698–701CrossRefGoogle ScholarPubMed
Barahona, M. and Poon, C-S. (1996). Detection of non-linear dynamics in short, noisy time series. Nature 381: 215–17CrossRefGoogle Scholar
Bard, E. (2001). Paleoceanographic implications of the difference in deep-sea sediment mixing between large and fine particles. Paleoceanography 16: 235–9CrossRefGoogle Scholar
Barlow, L. K., White, J. W. C., Barry, R. G., Rogers, J. C. and Grootes, P. M. (1993). The North Atlantic Oscillation signature in deuterium and deuterium excess signals in the Greenland Ice Sheet Project 2 ice core, 1840–1970. Geophys. Res. Lett. 20: 2901–4CrossRefGoogle Scholar
Barrell, J. (1917). Rhythms and the measurement of geological time. Bull. Geol. Soc. Am. 28: 745–904CrossRefGoogle Scholar
Barry, R. G. and Chorley, R. J. (1998). Atmosphere Weather and Climate, seventh edition. Rutledge, London, pp. 1–409
Beattie, P. D. and Dade, W. D. (1996). Is scaling in turbidite deposition consistent with forcing by earthquakes?J. Sed. Res. 66: 909–15Google Scholar
Beauchamp, K. G. (1975). Walsh Functions and their Applications. Academic Press, London, pp. 1–236
Beauchamp, K. G. (1984). Applications of Walsh and Related Functions, with an Introduction to Sequency Theory. Academic Press, New York, pp. 1–308
Beaufort, L. (1994). Climatic importance of the modulation of the 100 kyr cycle inferred from 16 m.y. long Miocene records. Paleoceanography 9: 821–34CrossRefGoogle Scholar
Beer, J., Blinov, A., Bonani, G., Finkel, R. C., Hofmann, H. J., Lehmann, B., Oeschger, H., Sigg, A., Schwander, J., Staffelbach, T., Stauffer, B., Suter, M. and Wölfli, W. (1990). Use of 10Be in polar ice to trace the 11-year cycle of solar activity. Nature 347: 164–6CrossRefGoogle Scholar
Beer, J., Mende, W. and Stellmacher, R. (2000). The role of the sun in climate forcing. Quat. Sci. Rev. 19: 403–15CrossRefGoogle Scholar
Beerbower, J. R. (1964). Cyclothems and cyclic depositional mechanisms in alluvial plain sedimentation. In: Symposium on Cyclic Sedimentation, Ed: D. F. Merriam. Kansas Geol. Surv. Bull. 169: 31–42Google Scholar
Behl, R. J., and Kennett, J. P. (1996). Brief interstadial events in the Santa Barbara basin, NE Pacific, during the past 60 kyr. Nature 379: 243–6CrossRefGoogle Scholar
Bender, M., Sowers, T., Dickson, M-L., Orchardo, J., Grootes, P., Mayewski, P. A. and Meese, D. A. (1994). Climate correlations between Greenland and Antarctica during the past 100,000 years. Nature 372: 663–6CrossRefGoogle Scholar
Benzi, R., Parisi, G., Sutera, A. and Vulpiani, A. (1982). Stochastic resonance in climatic change. Tellus 34: 10–16CrossRefGoogle Scholar
Berger, A. (1984). Accuracy and frequency stability of the Earth's orbital elements during the Quaternary. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 3–39
Berger, A. (1988). Milankovitch Theory and climate. Rev. Geophys. 26: 624–57CrossRefGoogle Scholar
Berger, A. (1989). The spectral characteristics of pre-Quaternary climate records, an example of the relationship between Astronomical Theory and Geosciences. In: Climate and Geosciences, Eds: A. Berger, S. Schneider, and J. C. Duplessy. Kluwer, Dordrecht, pp. 47–76
Berger, A. and Loutre, M. F. (1994). Astronomical forcing through geological time. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 15–24CrossRef
Berger, A. and Loutre, M. F. (1997). Intertropical latitudes and precessional and half-precessional cycles. Science 278: 1476–8CrossRefGoogle Scholar
Berger, A., Imbrie, J., Hays, J. D., Kukla, G. and Saltzman, B. (Eds) (1984). Milankovitch and Climate, Understanding the Response to Astronomical Forcing, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, 2 volumes, pp. 1–895
Berger, A., Loutre, M. F. and Dehant, V. (1989). Influence of the changing lunar orbit on the astronomical frequencies of pre-Quaternary insolation patterns. Paleoceanography 4: 555–64CrossRefGoogle Scholar
Berger, A., Mélice, J. L. and Mersch, I. (1990). Evolutive spectral analysis of sunspot data over the past 300 years. Philos. Trans. R. Soc. Lond. 330A: 529–41CrossRefGoogle Scholar
Berger, A., Loutre, M. F. and Mélice, J. L. (1997). Instability of the astronomical periods from 1.5 Myr BP to 0.5 Myr BP. Palaeoclimates 4: 1–42Google Scholar
Berggren, W. A., Kent, D. V., Swisher, C. C. and Aubry, M. (1995). A revised Cenozoic geochronology and chronostratigraphy. In: Geochronology Time Scales and Global Stratigraphic Correlation, Eds: W. A. Berggren, D. V. Kent, M. Aubry and J. Hardenbol. Society of Economic Paleontologists and Mineralogists (SEPM) Special Publication No. 54, pp. 129–212CrossRef
Bezrukov, S. M. and Vodyanoy, I. (1997). Stochastic resonance in non-dynamical systems without response thresholds. Nature 385: 319–21CrossRefGoogle ScholarPubMed
Bianchi, G. G. and McCave, I. N. (1999). Holocene periodicity in North Atlantic climate and deep-ocean flow south of Iceland. Nature 397: 515–17CrossRefGoogle Scholar
Bigg, G. R. (1996). The Oceans and Climate. Cambridge University Press, Cambridge, pp. 1–266
Biondi, F., Lange, C. B., Hughes, M. K. and Berger, W. H. (1997). Inter-decadal signals during the last millennium (AD 1117–1992) in the varve record of Santa Barbara basin, California. Geophys. Res. Lett. 24: 193–6CrossRefGoogle Scholar
Black, D. E., Peterson, L. C., Overpeck, J. T., Kaplan, A., Evans, M. N. and Kashgarian, M. (1999). Eight centuries of North Atlantic ocean atmosphere variability. Science 286: 1709–13CrossRefGoogle ScholarPubMed
Bloomfield, P. (1976). Fourier Analysis of Time Series: an Introduction. Wiley, London
Bond, G. and Lotti, R. (1995). Iceberg discharges into the North Atlantic on millennial time scales during the last glaciation. Science 267: 1005–10CrossRefGoogle ScholarPubMed
Bond, G. C., Heinrich, H., Broecker, W. S., Labeyrie, L., McManus, J. F., Andrews, J. T., Huon, S., Jantschik, R., Clasen, S., Simet, C., Tedesco, K., Klas, M., Bonani, G. and Ivy, S. (1992). Evidence for massive discharges of icebergs into the Northern Atlantic Ocean during the last glacial period. Nature 360: 245–9CrossRefGoogle Scholar
Bond, G., Broecker, W., Johnsen, S., McManus, J., Labeyrie, L., Jouzel, J. and Bonani, G. (1993). Correlations between climate records from North Atlantic sediments and Greenland ice. Nature 365: 143–7CrossRefGoogle Scholar
Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., deMenocal, P., Priore, P., Cullen, H., Hajdas, I. and Bonani, G. (1997). A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science 278: 1257–66CrossRefGoogle Scholar
Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M. N., Showers, W., Hoffmann, S., Lotti-Bond, R., Hajdas, I. and Bonani, G. (2001). Persistent solar influence on North Atlantic climate during the Holocene. Science 294: 2130–6CrossRefGoogle ScholarPubMed
Boss, S. K. and Rasmussen, K. A. (1995). Misuse of Fischer plots as sea-level curves. Geology 23: 221–42.3.CO;2>CrossRefGoogle Scholar
Boygle, J. (1993). The Swedish varve chronology — a review. Prog. Phys. Geog. 17: 1–19CrossRefGoogle Scholar
Bradley, W. H. (1929). The varves and climate of the Green River epoch. US Geol. Surv. Prof. Pap. 158: 87–110Google Scholar
Briffa, K. R. (2000). Annual climate variability in the Holocene: interpreting the message of ancient trees. Quat. Sci. Rev. 19: 87–105CrossRefGoogle Scholar
Briffa, K. R. and Osborn, T. J. (1999). Seeing the wood from the trees. Science 284: 926–7CrossRefGoogle Scholar
Briffa, K. R., Jones, P. D., Schweingruber, F. H. and Osborn, T. J. (1998a). Influence of volcanic eruptions on Northern hemisphere summer temperature over the past 600 years. Nature 393: 450–5CrossRefGoogle Scholar
Briffa, K. R., Schweingruber, F. H., Jones, P. D., Osborn, T. J., Harris, I. C., Shiyatov, S. G., Vaganov, E. A. and Grudd, H. (1998b). Trees tell of past climates: but are they speaking less clearly today?Philos. Trans. R. Soc. Lond. 353B: 65–73CrossRefGoogle Scholar
Briffa, K. R., Schweingruber, F. H., Jones, P. D., Osborn, T. J., Shiyatov, S. G. and Vaganov, E. A. (1998c). Reduced sensitivity of recent tree-growth to temperature at high northern latitudes. Nature 391: 678–82CrossRefGoogle Scholar
Briffa, K. R., Bartholin, T. S., Eckstein, D., Jones, P. D., Karlén, W., Schweingruber, F. H. and Zetterberg, P. (1990). A 1,400-year tree-ring record of summer temperatures in Fennoscandia. Nature 346: 434–9CrossRefGoogle Scholar
Broecker, W. S. (1997). Thermohaline circulation, the Achilles heel of our climate system: will man-made CO2 upset the current balance?Science 278: 1582–8CrossRefGoogle ScholarPubMed
Broecker, W. S. and Denton, G. H. (1990). The role of ocean-atmosphere reorganizations in glacial cycles. Geochim. Cosmochim. Acta 53: 2464–501Google Scholar
Broecker, W. S., Bond, G., Klas, M., Bonani, G. and Wolfli, W. (1990). A salt oscillator in the glacial Atlantic? 1, the concept. Paleoceanography 5: 469–77CrossRefGoogle Scholar
Broecker, W. S., Bond, G., Klas, M., Clark, E. and McManus, J. F. (1992). Origin of the northern Atlantic's Heinrich events. Clim. Dynam. 6: 265–73CrossRefGoogle Scholar
Brook, E. J., Sowers, T. and Orchardo, J. (1994). Rapid variations in methane concentration during the past 110,000 years. Science 273: 1087–91CrossRefGoogle Scholar
Buchanan, M. (2000). Ubiquity. The Science of History, …, or Why the World is Simpler than We Think. Weidenfield and Nicholson, London, pp. 1–230
Bull, D., Kemp, A. E. S. and Weedon, G. P. (2000). A 160,000 year old record of El Niño-Southern Oscillation in marine production and coastal run-off from Santa Barbara Basin, California. Geology 28: 1007–102.0.CO;2>CrossRefGoogle Scholar
Burroughs, W. J. (1992). Weather Cycles. Real or Imaginary? Cambridge University Press, Cambridge, pp. 1–207
Cande, S. C. and Kent, D. V. (1995). Revised calibration of the geomagnetic polarity time scale for the Late Cretaceous and Cenozoic. J. Geophys. Res. 100: 6093–5CrossRefGoogle Scholar
Cane, M. A. (1992). Tropical Pacific ENSO models: ENSO as a mode of the coupled system. In: Climate System Modelling, Ed: K. E. Trenberth. Cambridge University Press, Cambridge, pp. 583–614
Cannariato, K. G., Kennett, J. P. and Behl, R. J. (1999). Biotic response to late Quaternary rapid climate switches in Santa Barbara Basin: ecological and evolutionary implications. Geology 27: 63–62.3.CO;2>CrossRefGoogle Scholar
Carrs, B. W. and Neidell, N. S. (1966). A geological cyclicity detected by means of polarity coincidence correlation. Nature 212: 136–7CrossRefGoogle Scholar
Chan, M. J., Kvale, E., Archer, A. W. and Sonett, C. P. (1994). Oldest direct evidence of lunar-solar tidal forcing encoded in sedimentary rhythmites, Proterozoic Big Cottonwood Formation, central Utah. Geology 22: 791–42.3.CO;2>CrossRefGoogle Scholar
Chang, P., Ji, L. and Li, H. (1997). A decadal climate variation in the tropical Atlantic Ocean from thermodynamic air-sea interactions. Nature 385: 516–18CrossRefGoogle Scholar
Christensen, C. J., Gorsline, D. S., Hammond, D. E. and Lund, S. P. (1994). Non-annual laminations and explanation of anoxic basin-floor conditions in Santa Monica Basin, California Borderland, over the past four centuries. Mar. Geol. 116: 399–418CrossRefGoogle Scholar
Cisne, J. L. (1986). Earthquakes recorded stratigraphically on carbonate platforms. Nature 323: 320–2CrossRefGoogle Scholar
Clark, P. U., Webb, R. S. and Keigwin, L. D. (Eds) (1999). Mechanisms of Global Climate Change at Millennial Time Scales. Geophysical Monograph 112. American Geophysical Union, Washington, pp. 1–394CrossRef
Clark, P. U., Pisias, N. G., Stocker, T. F. and Weaver, A. J. (2002). The role of the thermohaline circulation in abrupt climate change. Nature 415: 863–9CrossRefGoogle ScholarPubMed
Cleaveland, M. K., Cook, E. R. and Stahle, D. W. (1992). Secular variability of the Southern Oscillation detected in tree-ring data from Mexico and the southern United States. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Eds: H. F. Diaz and V Markgraf. Cambridge University Press, Cambridge, pp. 271–91
Clemens, S. C. (1999). An astronomical tuning strategy for Pliocene sections: implications for global-scale correlation and phase relationships. Philos. Trans. R. Soc. Lond. 357: 1949–73CrossRefGoogle Scholar
Clemens, S. C. and Prell, W. L. (1990). Late Pleistocene variability of Arabian Sea summer monsoon winds and continental aridity: eolian records from the lithogenic component of deep-sea sediments. Paleoceanography 5: 109–45CrossRefGoogle Scholar
Clemens, S. C., Murray, D. W. and Prell, W. L. (1996). Nonstationary phase of the Plio-Pleistocene Asian Monsoon. Science 274: 943–8CrossRefGoogle ScholarPubMed
Clement, A. C., Seager, R. and Cane, M. A. (1999). Orbital controls on the El Niño/Southern Oscillation and the tropical climate. Paleoceanography 14: 441–56CrossRefGoogle Scholar
Cohen, A. L., Layne, G. D., Hart, S. R. and Lobel, P. S. (2001). Kinetic control of skeletal Sr/Ca in a symbiotic coral: implications for the paleotemperature proxy. Paleoceanography 16: 20–6CrossRefGoogle Scholar
Cole, J. E., Dunbar, R. B., McClanahan, T. R. and Muthiga, N. A. (2000). Tropical Pacific forcing of decadal SST variability in the western Indian Ocean over the past two centuries. Science 287: 617–19CrossRefGoogle ScholarPubMed
Cook, E. R., D'Arrigo, R. D. and Briffa, K. R. (1998). A reconstruction of the North Atlantic Oscillation using tree-ring chronologies from North America and Europe. The Holocene 8: 9–17CrossRefGoogle Scholar
Corrège, T., Delcroix, T., Récy, J., Beck, W., Caboich, G. and Cornec, F. L. (2000). Evidence for stronger El Niño-Southern Oscillation (ENSO) events in a mid-Holocene massive coral. Paleoceanography 15: 465–70CrossRefGoogle Scholar
Croll, J. (1864). On the physical cause of the change of climate during geological epochs. Phil. Mag. 28: 121–37CrossRefGoogle Scholar
Crowley, K. D., Duchon, C. E. and Rhi, J. (1986). Climate record in varved sediments of the Eocene Green River Formation. J. Geophys. Res. 91: 8637–47CrossRefGoogle Scholar
Crowley, T. J., Kim, K-Y., Mengel, J. G. and Short, D. A. (1992). Modeling 100,000-year climate fluctuations in pre-Pleistocene time series. Science 255: 705–7CrossRefGoogle ScholarPubMed
Crusius, J. and Anderson, R. F. (1992). Inconsistencies in accumulation rates of Black Sea sediments inferred from records of laminae and 210Pb. Paleoceanography 7: 215–27CrossRefGoogle Scholar
Cullen, H. M., D'Arrigo, R. D. and Cook, E. R. (2001). Multiproxy reconstructions of the North Atlantic. Paleoceanography 16: 27–39CrossRefGoogle Scholar
Currie, R. G. (1987). Examples and implications of 18.6- and 11-yr terms in world weather records. In: Climate History, Periodicity and Predictability, Eds: M. Rampino, J. E. Sanders, W. S. Newman and L. K. Kingsson. Van Nostrand Reinhold, New York, pp. 378–403
Curry, R. G., McCartney, M. S. and Joyce, T. M. (1998). Oceanic transport of subpolar climate signals to mid-depth subtropical waters. Nature 391: 575–7CrossRefGoogle Scholar
Curry, W. B. and Oppo, D. W. (1997). Synchronous high-frequency oscillations in tropical sea surface temperatures and North Atlantic deep water production during the last glacial cycle. Paleoceanography 12: 1–14CrossRefGoogle Scholar
Dalfes, H. N., Schneider, S. H. and Thompson, S. L. (1984). Effects of bioturbation on climatic spectra inferred from deep sea cores. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kula, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 481–92
Dalrymple, R. W. and Makino, Y. (1989). Description and genesis of tidal bedding in the Cobequid Bay — Salmon River estuary, Bay of Fundy, Canada. In: Sedimentary Facies of Active Plate Margins, Eds: A. Taira and F. Masuda. Terra, Tokyo, pp. 151–77
Dansgaard, W., Johnsen, S. J., Clausen, H. B., Dahl-Jensen, D., Gundestrup, N. S., Hammer, C. U., Hvidberg, C. S., Steffensen, J. P., Sveinbjörnsdottir, A. E., Jouzel, J. and Bond, G. (1993). Evidence for general instability of past climate from a 250-kyr ice-core record. Nature 364: 218–20CrossRefGoogle Scholar
D'Arrigo, R. D., Cook, E. R., Jacby, G. C. and Briffa, K. R. (1993). NAO and sea surface temperature signatures in tree-ring records from the North Atlantic sector. Quat. Sci. Rev. 12: 431–40CrossRefGoogle Scholar
Davis, J. C. (1973). Statistics and Data Analysis in Geology, first edition. Wiley, London, pp. 1–550
Davis, J. C. (1986). Statistics and Data Analysis in Geology, second edition. Wiley, Chichester, pp. 1–646
de Boer, P. L. and Smith, D. G. (Eds) (1994). Orbital Forcing and Cyclic Sequences. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 1–559
de Boer, P. L. and Wonders, A. A. H. (1984). Astronomically induced rhythmic bedding in Cretaceous pelagic sediments near Moria, Italy. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 177–90
Boer, P. L., Oost, A. P. and Visser, M. J. (1989). The diurnal inequality of the tide as a parameter for recognizing tidal influences. J. Sed. Petrol. 59: 912–21Google Scholar
Dehant, V., Loutre, M-F. and Berger, A. (1990). Potential impact of the Northern Hemisphere Quaternary ice sheets on the frequencies of the astroclimatic orbital parameters. J. Geophys. Res. 95: 7573–8CrossRefGoogle Scholar
Dessai, S. and Walter, M. E. (2000). Self-organised criticality and the atmospheric sciences: selected review, new findings and future directions. Contributed paper, NSF Workshop on Extreme Events (http://www.esig.ucar.edu/extremes/papers.html)
Dettinger, M. D., Ghil, M., Strong, C. M., Weibel, W. and Yiou, P. (1995). Software expedites singular-spectrum analysis of noisy time series. EOS Trans. AGU 76: 12–21Google Scholar
Diaz, H. E. and Markgraf, V. (Eds) (1992). El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation. Cambridge University Press, Cambridge
Dicke, R. H. (1978). Is there a chronometer hidden deep in the Sun?Nature 276: 676–80CrossRefGoogle Scholar
Dicke, R. H. (1979). Solar luminosity and the sunspot cycle. Nature 280: 24–7CrossRefGoogle Scholar
Dimitrov, B. D., Shangova-Gigoriadi, S. and Grigoriadis, E. D. (1998). Cyclicity in variations of incidence rates for breast cancer in different countries. Folia Med. (Plovdiv) 40: 66–71Google ScholarPubMed
Dokken, T. M. and Jansen, E. (1999). Rapid changes in the mechanism of ocean convection during the last glacial period. Nature 401: 458–61CrossRefGoogle Scholar
Droxler, A. W. and Schlager, W. (1985). Glacial versus interglacial sedimentation rates and oxygen-isotope record in the Bahamas. Geology 13: 799–8022.0.CO;2>CrossRefGoogle Scholar
Duff, P. M. D. and Walton, E. K. (1962). Statistical basis for cyclothems: a quantitative study of the sedimentary succession in the east Pennine coalfield. Sedimentology 1: 235–55CrossRefGoogle Scholar
Duff, P. M. D., Hallam, A. and Walton, E. K. (1967). Cyclic Sedimentation. Developments in Stratigraphy. Elsevier, Amsterdam, pp. 1–280
Dunbar, R. B., Wellington, G. M., Colgan, M. W. and Glynn, P. W. (1994). Eastern Pacific sea surface temperature since 1600AD: the δ18O record of climatic variability in Galápagos corals. Paleoceanography 9: 291–315CrossRefGoogle Scholar
Dunn, C. E. (1974). Identification of sedimentary cycles through Fourier analysis of geochemical data. Chem. Geol. 13: 217–32CrossRefGoogle Scholar
Duvall, T. L., D'Silva, S., Jefferies, S. M., Harvey, J. W. and Schou, J. (1996). Downflows under sunspots detected by helioseismic tomography. Nature 379: 235–7CrossRefGoogle Scholar
Eddy, J. A. (1976). The Maunder minimum. Science 192: 1189–202CrossRefGoogle ScholarPubMed
Egbert, G. D. and Ray, R. D. (2000). Significant dissipation of tidal energy in the deep ocean inferred from satellite altimeter data. Nature 405: 775–8CrossRefGoogle ScholarPubMed
Einsele, G. and Seilacher, A. (1991). Distinction of tempestites and turbidites. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, Berlin, pp. 377–82
Einsele, G., Ricken, W. and Seilacher, A. (1991). Cycles and events in stratigraphy — basic concepts and terms. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 1–19
Elliot, M., Labeyrie, L., Dokken, T. and Manthe, S. (2000). Coherent patterns of ice rafted debris deposits in the Nordic regions during the last glacial (10–60 ka). Earth Planet. Sci. Lett. 194: 151–63CrossRefGoogle Scholar
Elrick, M. and Hinnov, L. A. (1996). Millennial-scale climate origins for stratification in Cambrian and Devonian deep-water rhythmites, western USA. Palaeogeog. Palaeoclim. Palaeoecol. 123: 353–72CrossRefGoogle Scholar
Elrick, M., Read, J. A. and Coruh, C. (1991). Short-term paleoclimatic fluctuations expressed in lower Mississippian ramp-slope deposits, southwestern Montana. Geology 19: 799–8022.3.CO;2>CrossRefGoogle Scholar
Emery, W. J. and Thomson, R. E. (1997). Data Analysis Methods in Physical Oceanography. Elsevier, Amsterdam, pp. 1–634
Eriksson, K. A. and Simpson, E. L. (2000). Quantifying the oldest tidal record: the 3.2 Ga Moodies Group, Barberton Greenstone Belt, South Africa. Geology 28: 831–42.0.CO;2>CrossRefGoogle Scholar
Esper, J., Cook, E. R. and Schweingruber, F. H. (2002). Low-frequency signals in long tree-ring chronologies for reconstructing past temperature variability. Science 295: 2250–3CrossRefGoogle ScholarPubMed
Evans, J. W. (1972). Tidal growth increments in the cockle Clinocardium nuttalli. Science 176: 416–17CrossRefGoogle ScholarPubMed
Farley, K. A. and Patterson, D. B. (1995). A 100-kyr periodicity in the flux of extraterrestrial 3He to the seafloor. Nature 378: 600–3Google Scholar
Faure, G. (1986). Principles of Isotope Geology, second edition. Wiley, New York, pp. 1–589
Fedorov, A. V. and Philander, S. G. (2000). Is El Niño changing?Science 288: 1997–2002CrossRefGoogle ScholarPubMed
Finkel, R. C. and Nishiizumi, K. (1997). Beryllium 10 concentrations in the Greenland Ice Sheet Project 2 ice core from 3–40 ka. J. Geophys. Res. 102: 26,699–706CrossRefGoogle Scholar
Fischer, A. G. (1980). Gilbert — bedding rhythms and geochronology. Spec. Pap. Geol. Soc. Am. 183: 93–104Google Scholar
Fischer, A. G. (1986). Climatic rhythms recorded in strata. Annu. Rev. Earth Planet. Sci. 14: 351–76CrossRefGoogle Scholar
Fischer, A. G. and Bottjer, D. J. (Eds) (1991). Orbital forcing and sedimentary sequences. J. Sed. Petrol. 61: 1063–252Google Scholar
Fischer, A. G. and Roberts, L. T. (1991). Cyclicity in the Green River Formation (Lacustrine Eocene) of Wyoming. J. Sed. Petrol. 61: 1146–54Google Scholar
Fischer, A. G. and Schwarzacher, W. (1984). Cretaceous bedding rhythms under orbital control? In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 163–75
Fischer, A. G., de Boer, P. L. and Premoli Silva, I. (1990). Cyclostratigraphy. In: Cretaceous, Resources, Events and Rhythms, Eds: R. N. Ginsburg and B. Beaudoin. Kluwer, Dordrecht, pp. 139–72CrossRef
Fisher, R. A. (1929). Tests of significance in harmonic analysis. Proc. R. Soc. Series A, 125: 54–9CrossRefGoogle Scholar
Forte, A. M. and Mitrovica, J. X. A. (1997). A resonance in the Earth's obliquity and precession over the past 20 Myr driven by mantle convection. Nature 390: 676–9CrossRefGoogle Scholar
Foucault, A., Powichrowski, L. and Prud'Homme, A. (1987). Le côntrole astronomique de la sédimentation turbiditique: exemple du Flysch à Helminthoides des Alpes Ligures (Italie). C. R. Acad. Sci. Paris 305: 1007–11Google Scholar
Foukal, P. (1990). Solar luminosity variations over timescales of days to the past few solar cycles. Philos. Trans. R. Soc. Lond. 330A: 591–9CrossRefGoogle Scholar
Francois, R. and Bacon, M. (1994). Heinrich events in the North Atlantic: radiochemical evidence. Deep-Sea Res. 41: 315–34CrossRefGoogle Scholar
Friis-Christensen, E. and Lassen, K. (1991). Length of the solar cycle: an indicator of solar activity closely associated with climate. Science 254: 698–700CrossRefGoogle ScholarPubMed
Fröhlich, C. and Lean, J. (1998). The Sun's total irradiance: cycles, trends and related climate change uncertainties since 1976. Geophys. Res. Lett. 25: 4377–80CrossRefGoogle Scholar
Fronval, T., Jansen, E., Bloemendal, J. and Johnsen, S. (1995). Oceanic evidence for coherent fluctuations in Fennoscandian and Laurentide ice sheets on millennium timescales. Nature 374: 443–6CrossRefGoogle Scholar
Gagan, M. K., Ayliffe, L. K., Beck, J. W., Cole, J. E., Druffel, E. R. M., Dunbar, R. B. and Schrag, D. P. (2000). New views of tropical paleoclimates from corals. Quat. Sci. Rev. 19: 45–64CrossRefGoogle Scholar
Gallois, R. W. (2000). The stratigraphy of the Kimmeridge Clay (Upper Jurassic) in the RGGE Project boreholes at Swanworth Quarry and Metherhills, south Dorset. Proc. Geol. Assoc. 111: 265–80CrossRefGoogle Scholar
Ganopolski, A. and Rahmstorf, S. (2001). Rapid changes of global climate simulated in a coupled climate model. Nature 409: 153–8CrossRefGoogle Scholar
Ganopolski, A. and Rahmstorf, S. (2002). Abrupt glacial climate changes due to stochastic resonance. Phys. Rev. Lett. 88: Article 038501CrossRefGoogle ScholarPubMed
Gershenfeld, N., Schoner, B. and Metois, E. (1999). Cluster-weighted modelling for time-series analysis. Nature 397: 329–32CrossRefGoogle Scholar
Gilbert, G. K. (1895). Sedimentary measurement of geologic time. J. Geol. 3: 351–76CrossRefGoogle Scholar
Gilliland, R. L. (1981). Solar radius variations over the past 265 years. Astrophys. J. 248: 1144–55CrossRefGoogle Scholar
Gipp, M. R. (2001). Interpretation of climate dynamics from phase space portraits: is the climate system strange or just different?Paleoceanography 16: 335–51CrossRefGoogle Scholar
Glatzmaier, G. A., Coe, R. S., Hongre, L. and Roberts, P. H. (1999). The role of the Earth's mantle in controlling the frequency of geomagnetic reversals. Nature 401: 885–90CrossRefGoogle Scholar
Goldhammer, R. K., Dunn, P. A. and Hardie, L. A. (1990). Depositional cycles, composite sea-level changes, cycle stacking patterns and the hierarchy of stratigraphic forcing: examples from Alpine Triassic platform carbonates. Geol. Soc. Am. Bull. 102: 535–622.3.CO;2>CrossRefGoogle Scholar
Gorsline, D. S., Nava-Sanchez, E. and Murillo de Nava, J. (1996). A survey of occurrences of Holocene laminated sediments in California borderland basins: products of a variety of depositional processes. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication No. 116. The Geological Society, London, pp. 93–110CrossRef
Gough, D. (2000). News from the solar interior. Science 287: 2434–5CrossRefGoogle Scholar
Gradstein, F. M., Agterberg, F. P., Ogg, J. G., Hardenbol, J., Veen, P., Thierry, J. and Huang, Z. (1994). A Mesozoic Time Scale. J Geophys. Res. 99: 24,051–74CrossRefGoogle Scholar
Grassberger, P. (1986). Do climatic attractors exist?Nature 323: 609–12CrossRefGoogle Scholar
Greenland Ice-core Project Members (1993). Climate instability during the last interglacial period recorded in the GRIP ice core. Nature 364: 203–7CrossRef
Grootes, P. M. and Stuiver, M. (1997). Oxygen 18/16 variability in Greenland snow and ice with 10-3 to 105-year time resolution. J. Geophys. Res. 102: 26,455–470CrossRefGoogle Scholar
Gu, D. and Philander, S. G. H. (1997). Interdecadal climate fluctuations that depend on exchanges between the tropics and extratropics. Science 275: 805–7CrossRefGoogle ScholarPubMed
Guyodo, Y. and Valet, J-P. (1999). Global changes in intensity of the Earth's magnetic field during the past 800 kyr. Nature 399: 249–52CrossRefGoogle Scholar
Gwiazda, R. H., Hemming, S. R. and Broecker, W. S. (1996). Provenance of icebergs during Heinrich event 3 and the contrast to their sources during other Heinrich episodes. Paleoceanography 11: 371–8CrossRefGoogle Scholar
Haak, A. B. and Schlager, W. (1989). Compositional variations in calciturbidites due to sea-level fluctuations, late Quaternary, Bahamas. Geol. Runds. 78: 477–86CrossRefGoogle Scholar
Hagadorn, J. W. (1996). Laminated sediments of Santa Monica Basin, California continental borderland. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication No. 116. The Geological Society, London, pp. 111–20CrossRef
Hagelberg, T. K. and Pisias, N. G. (1990). Nonlinear response of Pliocene climate to orbital forcing: evidence from the eastern equatorial Pacific. Paleoceanography 5: 595–617CrossRefGoogle Scholar
Hagelberg, T. K., Pisias, N. G. and Elgar, S. (1991). Linear and nonlinear couplings between orbital forcing and the marine δ18O record during the late Neogene. Paleoceanography 6: 729–46CrossRefGoogle Scholar
Hagelberg, T. K., Bond, G. and deMenocal, P. (1994). Milankovitch band forcing of sub-Milankovitch climate variability during the Pleistocene. Paleoceanography 9: 545–58CrossRefGoogle Scholar
Haigh, J. D. (1996). The impact of solar variability on climate. Science 272: 981–4CrossRefGoogle ScholarPubMed
Halfman, J. D. and Johnson, T. C. (1988). High resolution record of cyclic climatic change during the past 4 ka from Lake Turkana, Kenya. Geology 16: 496–5002.3.CO;2>CrossRefGoogle Scholar
Hall, I. R., McCave, I. N., Shackleton, N. J., Weedon, G. P. and Harris, S. E. (2001). Glacial intensification of deep Pacific inflow and ventilation. Nature 412: 809–12CrossRefGoogle Scholar
Hallam, A. (1964). Origin of the limestone-shale rhythms in the Blue Lias of England: a composite theory. J. Geol. 72: 157–68CrossRefGoogle Scholar
Hammer, C., Mayewski, P. A., Peel, D. and Stuiver, M. (Eds) (1997). GISP2 and GRIP results. J. Geophys. Res. 102 part C12Google Scholar
Hansen, D. V. and Bezdek, H. F. (1996). On the nature of decadal anomalies in North Atlantic sea surface temperature. J. Geophys. Res. 101: 9749–58CrossRefGoogle Scholar
Hartmann, W. M. (1999). How we localize sound. Phys. Today 52: 24–9CrossRefGoogle Scholar
Hays, J. D., Imbrie, I. and Shackleton, N. J. (1976). Variations in the Earth's orbit: pacemaker of the ice ages. Science 194: 1121–32CrossRefGoogle ScholarPubMed
Heinrich, H. (1988). Origin and consequences of cyclic ice rafting in the northeast Atlantic Ocean during the past 130,000 years. Quat. Res. 29: 143–52CrossRefGoogle Scholar
Hendy, I. L. and Kennett, J. P. (1999). Latest Quaternary North Pacific surface-water responses imply atmosphere-driven climate instability. Geology 27: 291–42.3.CO;2>CrossRefGoogle Scholar
Herbert, T. D. (1992). Paleomagnetic calibration of Milankovitch cyclicity in Lower Cretaceous sediments. Earth Planet. Sci. Lett. 112: 15–28CrossRefGoogle Scholar
Herbert, T. D. (1993). Differential compaction in lithified deep-sea sediments is not evidence for “diagenetic unmixing”. Sediment. Geol. 84: 115–22CrossRefGoogle Scholar
Herbert, T. D. (1994). Reading orbital signals distorted by sedimentation: models and examples. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 483–507CrossRef
Herbert, T. D. and Fischer, A. G. (1985). Milankovitch climatic origin of mid-Cretaceous black shale rhythms in central Italy. Nature 321: 739–43CrossRefGoogle Scholar
Herschel, J. F. W. (1832). On the astronomical causes which may influence geological phenomena. Trans. Geol. Soc. 2nd Ser. 3: 393–9Google Scholar
Hibler, W. D. and Johnsen, S. J. (1979). The 20-yr cycle in Greenland ice core records. Nature 280: 481–3CrossRefGoogle Scholar
Hilgen, F. J. (1991). Astronomical calibration of Gauss to Matuyama sapropels in the Mediterranean and implications for the geomagnetic polarity timescale. Earth Planet. Sci. Lett. 104: 226–44CrossRefGoogle Scholar
Hilgen, F. J. and Langereis, G. C. (1988). The age of the Miocene-Pliocene boundary in the Capo Rossello area (Sicily). Earth Planet. Sci. Lett. 91: 214–22CrossRefGoogle Scholar
Hilgen, F. J. and Langereis, C. G. (1989). Periodicities of CaCO3 cycles in the Pliocene of Sicily: discrepancies with the quasi-periods of the Earth's orbital cycles?Terra Nova, 1: 409–15CrossRefGoogle Scholar
Hilgen, F. J., Krijgsman, W., Langereis, C. G., Lourens, L. J., Santarelli, A. and Zachariasse, W. J. (1995). Extending the astronomical (polarity) time scale into the Miocene. Earth Planet. Sci. Lett. 136: 495–510CrossRefGoogle Scholar
Hilgen, F. J., Abdul Aziz, W., Krijgsman, W., Langereis, C. G., Lourens, L. J., Meulenkamp, J. E., Raffi, I., Steenbrink, J., Turco, E., Vught, N., Wijbrans, J. R. and Zachariasse, W. J. (1999). Present status of the astronomical (polarity) time-scale for the Mediterranean Late Neogene. Philos. Trans. R. Soc. 357: 1931–47CrossRefGoogle Scholar
Hilgen, F., Schwarzacher, W. and Strasser, A. (2001). Concepts and definitions in cyclostratigraphy — Second Report of the cyclostratigraphy working group. (International Subcommission on Stratigraphic Nomenclature, IUGS Commission of Stratigraphy). Unpublished, 8p
Hinnov, L. A. (2000). New perspectives on orbitally-forced stratigraphy. Annu. Rev. Earth Planet. Sci. 28: 419–75CrossRefGoogle Scholar
Hinnov, L. A. and Goldhammer, R. K. (1991). Spectral analysis of the Middle Triassic Latemar Limestone. J. Sed. Petrol. 61: 1173–93Google Scholar
Hinnov, L. A. and Park, J. (1998). Detection of astronomical cycles in the stratigraphic record by frequency modulation (FM) analysis. J. Sed. Res. 68: 524–39CrossRefGoogle Scholar
Hinnov, L. A. and Park, J. (1999). Strategies for assessing Early-Middle (Pliensbachian-Aalenian) Jurassic cyclochronologies. Philos. Trans. R. Soc. Lond. 357: 1831–59CrossRefGoogle Scholar
Holmgren, K., Karlén, W., Lauritzen, S. E., Lee-Thorp, J. A., Partridge, T. C., Piketh, S., Repinski, P., Stevenson, C., Svanered, O. and Tyson, P. D. (1999). A 3000-year high-resolution stalagmite-based record of palaeoclimate for northeastern South Africa. The Holocene 9: 295–309CrossRefGoogle Scholar
House, M. R. (1985). A new approach to an absolute timescale from measurements of orbital cycles and sedimentary microrhythms. Nature 316: 721–5CrossRefGoogle Scholar
House, M. R. and Farrow, G. E. (1968). Daily growth banding in the shell of the cockle, Cardium edule. Nature 219: 1384–6CrossRefGoogle ScholarPubMed
House, M. R. and Gale, A. S. (Eds) (1995). Orbital Forcing Timescales and Cyclostratigraphy. Geological Society Special Publication No. 85. The Geological Society, London, pp. 51–66CrossRef
Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R. W., Larsen, R. M., Schou, J., Thompson, M. J. and Toomre, J. (2000). Dynamic variations at the base of the solar convective zone. Science 287: 2456–60CrossRefGoogle ScholarPubMed
Hoyt, V. and Schatten, K. H. (1997). The Role of the Sun in Climate Change. Oxford University Press, Oxford, pp. 1–279
Hubbard, B. B. (1996). The World According to Wavelets. The Story of a Mathematical Technique in the Making. A. K. Peters, Wellesley, Massachusetts, pp. 1–264
Hulme, M. and Barrow, E. (Eds) (1997). Climates of the British Isles. Routledge, London, pp. 1–454
Hunt, A. G. and Malin, P. E. (1998). Possible triggering of Heinrich events by ice-load-induced earthquakes. Nature 393: 155–8Google Scholar
Hurrell, J. W. (1995). Decadal trends in the North Atlantic Oscillation: regional temperatures and precipitation. Science 269: 676–9CrossRefGoogle ScholarPubMed
Hydrographic Office (1996). Admiralty Tide Tables 1997, Volume. 1, United Kingdom and Ireland including European Channel Ports. Hydrographic Office (UK)
Ifeachor, E. C. and Jervis, B. W. (1993). Digital Signal Processing. A Practical Approach. Addison-Wesley, Harlow, pp. 1–760
Imbrie, J. and Imbrie, J. Z. (1980). Modelling the climate response to orbital variations. Science 207: 943–53CrossRefGoogle Scholar
Imbrie, J. and Imbrie, K. P. (1979). Ice Ages. Solving the Mystery. Harvard University Press, Cambridge, Massachusetts, pp. 1–224CrossRef
Imbrie, J., Hays, J. D., Martinson, D. G., McIntyre, A. C., Mix, A. C., Morley, J. J., Pisias, N. G., Prell, W. L. and Shackleton, N. J. (1984). The orbital theory of Pleistocene climate: support from a revised chronology of the marine δ18O record. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 269–305
Imbrie, J., Berger, A., Boyle, E. A., Clemens, S. C., Duffy, A., Howard, W. R., Kukla, G., Kutzbach, J., Martinson, D. G., McIntyre, A., Mix, A. C., Molfino, B., Morley, J. J., Peterson, L. C., Pisias, N. G., Prell, W. L., Raymo, M. E., Shackleton, N. J. and Toggweiler, J. R. (1992). On the structure and origin of major glaciation cycles, 1: Linear responses to Milankovitch forcing. Paleoceanography 7: 701–38CrossRefGoogle Scholar
Imbrie, J., Berger, A., Boyle, E. A., Clemens, S. C., Duffy, A., Howard, W. R., Kukla, G., Kutzbach, J., Martinson, D. G., McIntyre, A., Mix, A. C., Molfino, B., Morley, J. J., Peterson, L. C., Pisias, N. G., Prell, W. L., Raymo, M. E., Shackleton, N. J. and Toggweiler, J. R. (1993a). On the structure and origin of the major glaciation cycles, 2: The 100,000-year cycle. Paleoceanography 8: 699–735CrossRefGoogle Scholar
Imbrie, J., Berger, A. and Shackleton, N. J. (1993b). Role of orbital forcing: a two-million-year perspective. In: Global Changes in the Perspective of the Past, Eds: J. A. Eddy and H. Oeschger. Wiley, Chichester, pp. 263–77
Ito, M., Nishikawa, T. and Sugimoto, H. (1999). Tectonic control of high-frequency depositional sequences with durations shorter than Milankovitch cyclicity: an example from the Pleistocene paleo-Tokyo Bay, Japan. Geology 27: 763–62.3.CO;2>CrossRefGoogle Scholar
James, I. N. and James, P. M. (1989). Ultra-low-frequency variability in a simple atmospheric circulation model. Nature 342: 53–5CrossRefGoogle Scholar
Jenkins, G. M. and Watts, D. G. (1969). Spectral Analysis and its Applications. Holden-Day, London
Jin, F-F., Neelin, J. D. and Ghil, M. (1994). El Niño on the Devil's staircase: annual subharmonic steps to chaos. Science 264: 70–2CrossRefGoogle Scholar
Johnsen, S. J., Clausen, H. B., Dansgaard, W., Gundestrup, N. S., Hammer, C. U., Andersen, U., Andersen, K. K., Hvidberg, C. S., Dahl-Jensen, D., Steffensen, J. P., Shoji, H., Sveinbjörnsdottir, Á.E., White, J., Jouzel, J. and Fisher, D. (1997). The δ18O record along the Greenland Ice Core Project deep ice core and the problem of possible Eemian climatic instability. J. Geophys. Res. 102: 26,397–410CrossRefGoogle Scholar
Jones, C. E., Jenkyns, H. C. and Hesselbo, S. B. (1994). Strontium isotopes in early Jurassic seawater. Geochim. Cosmochim. Acta 58: 3061–74CrossRefGoogle Scholar
Kanasewich, E. R. (1981). Time Sequence Analysis in Geophysics. University of Alberta Press, Alberta
Kantz, H. and Schreiber, T. (1997). Nonlinear Time Series Analysis. Cambridge University Press, Cambridge, pp. 1–304
Keeling, C. D. and Whorf, T. P. (2000). The 1,800-oceanic tidal cycle: a possible cause of rapid climate change. Proc. Natl. Acad. Sci. USA 97: 3814–19CrossRefGoogle ScholarPubMed
Keigwin, L. D. and Lehmann, S. J. (1994). Deep circulation change linked to HEINRICH event 1 and Younger Dryas in a mid depth North Atlantic core. Paleoceanography 9: 185–94CrossRefGoogle Scholar
Kemp, A. E. S. (Ed.) (1996). Palaeoclimatology and Palaeoceanography from Laminated Sediments. Geological Society Publication No. 116. The Geological Society, London
Kemp, A. E. S. and Baldauf, J. G. (1993). Vast Neogene diatom mat deposits from the eastern equatorial Pacific Ocean. Nature 362: 141–4CrossRefGoogle Scholar
Kemp, A. E. S., Baldauf, J. G. and Pearce, R. B. (1995). Origins and paleoceanographic significance of laminated diatom ooze from the eastern equatorial Pacific. Proc. Ocean Drill. Prog. Sci. Res. 138: 641–5Google Scholar
Kennedy, J. A. and Brassell, S. C. (1992). Molecular records of twentieth-century El Niño events in laminated sediments from the Santa Barbara basin. Nature 357: 62–4CrossRefGoogle Scholar
Kennett, J. P. and Ingram, B. L. (1995). A 20,000-year record of ocean circulation and climate change from the Santa Barbara Basin. Nature 377: 510–14CrossRefGoogle Scholar
Kent, D. V. (1999). Orbital tuning of geomagnetic polarity time-scales. Philos. Trans. R. Soc. Lond. 357: 1995–2007CrossRefGoogle Scholar
Kerr, R. A. (1999). Link between sunspots, stratosphere buoyed [sic]. Science 284: 234–5CrossRefGoogle Scholar
King, T. (1996). Quantifying nonlinearity and geometry in time series of climate. Quat. Sci. Rev. 15: 247–66CrossRefGoogle Scholar
Kirchner, J. W. (2002). Evolutionary speed limits inferred from the fossil record. Nature 415: 65–8CrossRefGoogle ScholarPubMed
Kominz, M. A. (1996). Whither cyclostratigraphy? Testing the gamma method on upper Pleistocene deep-sea sediments, North Atlantic Deep Sea Drilling Project site 609. Paleoceanography 11: 481–504CrossRefGoogle Scholar
Kominz, M. A. and Bond, G. (1990). A new method of testing periodicity in cyclic sediments — application to the Newark Supergroup. Earth Planet. Sci. Lett. 98: 233–44CrossRefGoogle Scholar
Kominz, M. A., Beavan, J., Bond, G. C. and McManus, J. (1991). Are cyclic sediments periodic? Gamma analysis and spectral analysis of Newark Supergroup lacustrine strata. In: Sedimentary Modeling: Computer Simulations and Methods for Improved Parameter Definition, Eds: K. Franseen, W. L. Watney, C. G. St. C. Kendall and W. Ross. Bull. Kansas Geol. Surv. 233: 319–34
Kortenkamp, S. J. and Dermott, S. F. (1998). A 100,000-year periodicity in the accretion rate of interplanetary dust. Science 280: 874–6CrossRefGoogle ScholarPubMed
Kotilainen, A. T. and Shackleton, N. J. (1995). Rapid climate variability in the North Pacific Ocean during the past 95,000 years. Nature 377: 323–6CrossRefGoogle Scholar
Krijgsman, W., Hilgen, F. J., Langereis, C. G. and Zachariasse, W. J. (1994a). The age of the Tortonian-Messinian boundary. Earth Planet. Sci. Lett. 121: 533–47CrossRefGoogle Scholar
Krijgsman, W., Langereis, C. G., Daams, R. and Meulen, A. J. (1994b). Magnetostratigraphic dating of the Middle Miocene climate change in the continental deposits of the Aragonian type area in the Calatayud-Teruel basin (central Spain). Earth Planet. Sci. Lett. 128: 513–26CrossRefGoogle Scholar
Krijgsman, W., Hilgen, F. J., Langereis, C. G., Santarelli, A. and Zachariasse, W. J. (1995). Late Miocene magnetostratigraphy, biostratigraphy and cyclostratigraphy in the Mediterranean. Earth Planet. Sci. Lett. 136: 475–94CrossRefGoogle Scholar
Kumar, K. K., Rajagopalan, B. and Cane, M. A. (1999). On the weakening relationship between the Indian Monsoon and ENSO. Science 284: 2156–9CrossRefGoogle ScholarPubMed
Kumar, P. and Foufoula-Georgiou, E. (1994). Wavelet analysis in geophysics: an introduction. In: Wavelets in Geophysics. Academic Press, London, pp. 1–43
Kvale, E. P., Archer, A. W. and Johnson, H. R. (1989). Daily, monthly, and yearly tidal cycles within laminated siltstone of the Mansfield Formation (Pennsylvanian) of Indiana. Geology 17: 365–82.3.CO;2>CrossRefGoogle Scholar
Kvale, E. P., Johnson, H. W., Sonett, C. P., Archer, A. W. and Zawistoski, A. (1999). Calculating the lunar retreat rates using tidal rhythmites. J. Sed. Res. 69: 1154–68CrossRefGoogle Scholar
Labitzke, K. and Loon, H. (1990). Associations between 11-year solar cycle, the quasi-biennial oscillation and the atmosphere: a summary of recent work. Philos. Trans. R. Soc. Lond. 330A: 577–89CrossRefGoogle Scholar
Lang, W. D., Spath, L. F., Cox, L. R. and Muir-Wood, H. M. (1928). The Belemnite Marls of Charmouth, a series in the Lower Lias of the Dorset Coast. Quart. J. Geol. Soc. 84: 179–257CrossRefGoogle Scholar
Laskar, J. (1989). A numerical experiment on the chaotic behaviour of the solar system. Nature 338: 237–8CrossRefGoogle Scholar
Laskar, J. (1999). The limits of Earth orbital calculations for geological time-scale use. Philos. Trans. R. Soc. Lond. 357: 1735–59CrossRefGoogle Scholar
Laskar, J., Joutel, F. and Robutel, P. (1993a). Stabilization of the Earth's obliquity by the Moon. Nature 361: 615–17CrossRefGoogle Scholar
Laskar, J., Joutel, F. and Boudin, F. (1993b). Orbital, precessional, and insolation quantities for the Earth from -20Myr to +10Myr. Astron. Astrophys. 270: 522–33Google Scholar
Latif, M. and Barnett, T. P. (1994). Causes of decadal climate variability over the North Pacific and North America. Science 266: 634–7CrossRefGoogle ScholarPubMed
Lau, K-M. and Sheu, P. J. (1988). Annual cycle, Quasi-Biennial Oscillation, and Southern Oscillation in global precipitation. J. Geophys. Res. 93: 10,975–88CrossRefGoogle Scholar
Lees, J. M. and Park, J. (1995). Multi-taper spectral analysis: a stand-alone C-subroutine. Comput. Geosci. 21: 199–236CrossRefGoogle Scholar
Treut, H. and Ghil, M. (1983). Orbital forcing, climatic interactions, and glaciation cycles. J. Geophys. Res. 88: 5167–90CrossRefGoogle Scholar
Liu, P. C. (1994). Wavelet spectrum analysis and ocean wind waves. In: Wavelets in Geophysics. Academic Press, London, pp. 151–66CrossRef
Lockwood, M., Stamper, R. and Wild, M. N. (1999). A doubling of the Sun's coronal magnetic field during the past 100 years. Nature 399: 437–9CrossRefGoogle Scholar
Lourens, J. L., Antonarakou, A., Hilgen, F. J., Hoof, A. A. M., Vergnaud-Grazzini, C. and Zachariasse, W. J. (1996). Evaluation of the Plio-Pleistocene astronomical timescale. Paleoceanography 11: 391–413CrossRefGoogle Scholar
Lourens, L. J., Wehausen, R. and Brumsack, H. J. (2001). Geological constraints on tidal dissipation and dynamical ellipticity of the Earth over the past three million years. Nature 409: 1029–33CrossRefGoogle ScholarPubMed
Lowrie, W. (1997). Fundamentals of Geophysics. Cambridge University Press, Cambridge, pp. 1–354
Lund, D. C. and Mix, A. C. (1998). Millennial-scale deep water oscillations: reflections of the North Atlantic in the deep Pacific from 10 to 60 ka. Paleoceanography 13: 10–19CrossRefGoogle Scholar
Lyell, C. (1830). Principles of Geology, Volume 1. John Murray, London
Lyon, J. G. (2000). The solar wind-magnetosphere-ionosphere system. Science 288: 1987–91CrossRefGoogle ScholarPubMed
MacAyeal, D. R. (1993a). A low-order model of the Heinrich event cycle. Paleoceanography 8: 767–73CrossRefGoogle Scholar
MacAyeal, D. R. (1993b). Binge/Purge oscillations of the Laurentide Ice Sheet as a cause of the North Atlantic's Heinrich events. Paleoceanography 8: 775–84CrossRefGoogle Scholar
Mallat, S. (1998). A Wavelet Tour of Signal Processing. Academic Press, London, pp. 1–577
Mann, M. E. and Bradley, R. S. (1999). Northern hemisphere temperatures during the past millennium: inferences, uncertainties and limitations. Geophys. Res. Lett. 26: 759–62CrossRefGoogle Scholar
Mann, M. E. and Lees, J. (1996). Robust estimation of background noise and signal detection in climatic time series. Clim. Change 33: 409–45CrossRefGoogle Scholar
Mann, M. E., Bradley, R. S. and Hughes, M. K. (1998). Global-scale temperature patterns and climate forcing over the past six centuries. Nature 392: 779–87CrossRefGoogle Scholar
Markson, R. and Muir, M. (1980). Solar wind control of the Earth's electric field. Science 208: 979–90CrossRefGoogle ScholarPubMed
Martin, E. E., Shackleton, N. J., Zachos, N. J. and Flower, B. P. (1999). Orbitally-tuned Sr isotope chemostratigraphy for the late middle to late Miocene. Paleoceanography 14: 74–83CrossRefGoogle Scholar
Martinson, D. G., Menke, W. and Stoffa, P. (1982). An inverse approach to signal correlation. J. Geophys. Res. 87: 4807–18CrossRefGoogle Scholar
Matsuoka, J., Kano, A., Oba, T., Watanabe, T., Sakai, S. and Seto, K. (2001). Seasonal variation of stable isotopic composition recorded in a laminated tufa, SW Japan. Earth Planet. Sci. Lett. 192: 31–44CrossRefGoogle Scholar
Mayewski, P. A., Meeker, L. D., Twickler, M. S., Whitlow, S., Yang, Q., Lyons, W. B. and Prentice, M. (1997). Major features and forcing of high-latitude northern hemispheric atmospheric circulation using a 110,000-year-long glaciochemical series. J. Geophys. Res. 102: 26,345–66CrossRefGoogle Scholar
McClellan, J. H., Parks, T. W. and Rabiner, L. R. (1973). A computer program for designing optimum FIR linear phase digital filters. IEEE Trans. Audio Electroacoustics 21: 506–26CrossRefGoogle Scholar
McIntyre, A. and Molfino, B. (1996). Forcing of Atlantic equatorial and subpolar millennial cycles by precession. Science 274: 1867–70CrossRefGoogle ScholarPubMed
McManus, J. F., Bond, G. C., Broecker, W. S., Johnsen, S., Labeyrie, L. and Higgins, S. (1994). High-resolution climate records from the North Atlantic during the last interglacial. Nature 371: 326–9CrossRefGoogle Scholar
McManus, J. F., Bond, G. C., Broecker, W. S., Fleisher, M. Q. and Higgins, S. M. (1998). Radiometrically determined fluxes in the sub-polar North Atlantic during the last 140,000 years. Earth Planet. Sci. Lett. 135: 29–43CrossRefGoogle Scholar
McManus, J. F., Oppo, D. W. and Cullen, J. L. (1999). A 0.5 million-year record of millennial-scale climate variability in the North Atlantic. Science 283: 971–5CrossRefGoogle ScholarPubMed
McPhaden, M. J. and Yu, X. (1999). Equatorial waves and the 1997–98 El Niño. Geophys. Res. Lett. 26: 2961–4CrossRefGoogle Scholar
Medio, A. (1992). Chaotic Dynamics. Theory and Applications to Economics. Cambridge University Press, Cambridge, pp. 1–344
Meeker, L. D., Mayewski, P. A., Grootes, P. M., Alley, R. B. and Bond, G. C. (2001). Comment: “On sharp spectral lines in the climate record and the millennial peak” by C. Wunsch. Paleoceanography 16: 544–7CrossRefGoogle Scholar
Meko, D. M. (1992). Spectral properties of tree-ring data in the United States Southwest as related to El Niño/Southern Oscillation. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Ed: H. F. Diaz and V. Markgraf. Cambridge University Press, Cambridge, pp. 227–41
Melnyk, D. H., Smith, D. G. and Amiri-Garroussi, K. (1994). Filtering and frequency mapping as tools in subsurface cyclostratigraphy, with examples from the Wessex Basin, UK. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 35–46CrossRef
Meyers, S. R., Sageman, B. R. and Hinnov, L. A. (2001). Integrated quantitative stratigraphy of the Cenomanian-Turonian Bridge Creek Limestone Member using evolutive harmonic analysis and stratigraphic modeling. J. Sed. Res. 71: 628–44CrossRefGoogle Scholar
Middleton, G. V., Plotnick, R. E. and Rubin, D. M. (1995). Nonlinear Dynamics and Fractals. New Numerical Techniques for Sedimentary Data. Society of Economic Paleontologists and Mineralogists (SEPM) Short Course No. 36. SEPM, Tulsa, Oklahoma, pp. 1–174CrossRef
Milankovitch, M. (1941). Canon of insolation in the Ice-Age problem. [English translation by Israel Program for scientific translation, Jerusalem 1969.]R. Serbian Acad. Spec. Publ.132Google Scholar
Miller, D. J. and Eriksson, K. A. (1997). Late Mississippian prodeltaic rhythmites in the Appalachian Basin: a hierarchical record of tidal and climatic periodicities. J. Sed. Res. 67: 653–60Google Scholar
Miller, K. G., Feigenson, M. D., Wright, J. D. and Clement, B. M. (1991). Miocene isotope reference section, Deep Sea Drilling Project Site 608: an evaluation of isotope and biostratigraphic resolution. Paleoceanography 6: 33–52CrossRefGoogle Scholar
Mitchell, J. M. (1976). An overview of climatic variability and its causal mechanisms. Quat. Res. 6: 481–93CrossRefGoogle Scholar
Mitrovica, J. X., Forte, A. M. and Pan, R. (1997). Glaciation-induced variations in the Earth's precession frequency, obliquity and insolation over the last 2.6 Ma. Geophys. J. Int. 128: 270–84CrossRefGoogle Scholar
Molinie, A. J., Ogg, J. G. and Ocean Drilling Program Leg 129 Scientific Party (1990). Sedimentation rate curves and discontinuities from sliding-window spectral analysis of logs. Log Analyst, November–December, pp. 370–4
Morgans-Bell, H. S., Coe, A. L., Hesselbo, S. P., Jenkyns, H. C., Weedon, G. P., Marshall, J. E. A. and Williams, C. J. (2001). Integrated stratigraphy of the Kimmeridge Clay Formation (Upper Jurassic) based on exposures and boreholes in South Dorset, UK. Geol. Mag. 138: 511–39CrossRefGoogle Scholar
Morley, C. K., Vanhauwaert, P. and Batist, M. (2000). Evidence for high-frequency cyclic fault activity from high-resolution seismic reflection survey, Rukwa Rift, Tanzania. J. Geol. Soc. Lond. 157: 983–94CrossRefGoogle Scholar
Mudelsee, M. and Stattegger, K. (1994). Plio/Pleistocene climate modeling based on oxygen isotope time series from deep-sea sediment cores: the Grassberger-Procaccia algorithm and chaotic systems. Math. Geol. 26: 799–815CrossRefGoogle Scholar
Muller, R. A. and Macdonald, G. J. (1995). Glacial cycles and orbital inclination. Nature 377: 107–8CrossRefGoogle Scholar
Muller, R. A. and Macdonald, G. J. (1997a). Simultaneous presence of orbital inclination and eccentricity in proxy climate records from Ocean Drilling Program Site 806. Geology 25: 3–62.3.CO;2>CrossRefGoogle Scholar
Muller, R. A. and Macdonald, G. J. (1997b). Reply to comment on: Simultaneous presence of orbital inclination and eccentricity in proxy climate records from Ocean Drilling Program Site 806, by Schulz, M. and Mudelsee, M. Geology 25: 861–22.3.CO;2>CrossRefGoogle Scholar
Munk, W., Dzieciuck, M. and Jayne, S. (2002). Millennial climate variability: is there a tidal connection?J. Clim. 15: 370–852.0.CO;2>CrossRefGoogle Scholar
Munnecke, A., Westphal, H., Elrick, M. and Reijmer, J. J. G. (2001). The mineralogical composition of precursor sediments of calcareous rhythmites — a new approach. Int. J. Earth Sci. (Geol. Rundsch.) 90: 795–812Google Scholar
Murakoshi, N., Nakayama, N. and Masuda, F. (1995). Diurnal inequality pattern of tide in the upper Pleistocene Palaeo-Tokyo Bay: reconstruction from tidal deposits and growth-lines of fossil bivalves. Spec. Pub. Int. Assoc. Sed. 24: 289–300Google Scholar
Murray, B. C., Ward, W. R. and Yeung, S. C. (1973). Periodic insolation variation on Mars. Science 180: 638–40CrossRefGoogle ScholarPubMed
Naish, T. R. and Kamp, P. J. J. (1997). Sequence stratigraphy of sixth order (41 k.y.) Pliocene-Pleistocene cyclothems, Wanganui Basin, New Zealand: a case for the regressive systems tract. Geol. Soc. Am. Bull. 109: 978–992.3.CO;2>CrossRefGoogle Scholar
Naish, T. R., Abbott, S. T., Alloway, B. V., Beu, A. G., Carter, R. M., Edwards, A. R., Journeaux, T. J., Kamp, P. J. J., Pillans, B. J., Saul, G. S. and Woolfe, K. J. (1998). Astronomical calibration of a southern hemisphere Plio-Pleistocene reference section, Wanganui Basin, New Zealand. Quat. Sci. Rev. 17: 695–710CrossRefGoogle Scholar
National Research Council (1994). Solar Influences on Global Change. National Academy Press, Washington, pp. 1–163
Neff, U., Burns, S. J., Mangini, A., Mudelsee, M., Fleitmann, D. and Matter, A. (2001). Strong coherence between solar variability and the monsoon in Oman between 9 and 6 kyr ago. Nature 411: 290–3CrossRefGoogle ScholarPubMed
Nicolis, C. and Nicolis, G. (1984). Is there a climatic attractor?Nature 311: 529–32CrossRefGoogle Scholar
Ninnemann, U., Charles, C. D. and Hodell, D. A. (1999). Origin of global millennial scale climate events: constraints from the Southern Ocean deep sea sedimentary record. In: Mechanisms of Global Climate Change at Millennial Time Scales. Geophysical Monograph 112. American Geophysical Union, Washington, pp. 99–112CrossRef
Nowroozi, A. A. (1967). Table for Fisher's test of significance in harmonic analysis. Geophys. J. R. Astron. Soc. 12: 512–20CrossRefGoogle Scholar
O'Brien, N. R. (1996). Shale lamination and sedimentary processes. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Publication 116. The Geological Society, London, pp. 23–36CrossRef
Oeschger, H. and Beer, J. (1990). The past 5000 years history of solar modulation of cosmic radiation from 10Be and 14C studies. Philos. Trans. R. Soc. Lond. 330A: 471–80CrossRefGoogle Scholar
Olsen, G. H. (1977). Modern Electronics Made Simple. Allen, London, pp. 1–306
Olsen, P. E. (1986). A 40 million year lake record of early Mesozoic orbital climatic forcing. Science 234: 842–8CrossRefGoogle ScholarPubMed
Olsen, P. E. and Kent, D. V. (1996). Milankovitch climate forcing in the tropics of Pangaea during the Late Triassic. Palaeogeog. Palaeoclimat. Palaeoecol. 122: 1–26CrossRefGoogle Scholar
Olsen, P. E. and Kent, D. V. (1999). Long-period Milankovitch cycles from the Late Triassic and Early Jurassic of eastern North America and their implications for the calibration of the Early Mesozoic time-scale and the long-term behaviour of the planets. Philos. Trans. R. Soc. Lond. 357: 1761–86CrossRefGoogle Scholar
Oost, A. P., Dehaas, H., Ijnsen, F., Vandenboogert, J. M. and Boer, P. L. (1993). The 18.6 yr nodal cycle and its impact on tidal sedimentation. Sed. Geol. 87: 1–11CrossRefGoogle Scholar
Oppo, D. W. and Lehman, S. J. (1995). Suborbital timescale variability of North Atlantic deep water during the past 200,000 years. Paleoceanography 10: 901–10CrossRefGoogle Scholar
Otnes, R. K. and Enochson, L. (1978). Applied Time Series Analysis, Volume 1. Basic Techniques. Wiley, Chichester, pp. 1–449
Paillard, D., Labeyrie, L. and Yiou, P. (1996). Macintosh program performs time-series analysis. EOS Trans. AGU 77: 379CrossRefGoogle Scholar
Pälike, H. and Shackleton, N. J. (2000). Constraints on astronomical parameters from the geological record for the past 25 Myr. Earth Planet. Sci. Lett. 182: 1–14CrossRefGoogle Scholar
Pälike, H., Shackleton, N. J. and Röhl, U. (2001). Astronomical forcing in late Eocene marine sediments. Earth Planet. Sci. Lett. 193: 589–602CrossRefGoogle Scholar
Pantev, C., Oostenveld, R., Engelien, A., Ross, B., Roberts, L. E. and Hoke, M. (1998). Increased auditory cortical representation in musicians. Nature 392: 811–14CrossRefGoogle ScholarPubMed
Pardo-Igúzquiza, E. and Rodríguez-Tovar, F. J. (2000). The permutation test as a non-parametric method for statistical significance of power spectrum estimation in cyclostratigraphic research. Earth Planet. Sci. Lett. 181: 175–89CrossRefGoogle Scholar
Pardo-Igúzquiza, E., Chica-Olmo, M. and Rodríguez-Tovar, F. J. (1994). CYSTRATI: a computer program for spectral analysis of stratigraphic successions. Comput. Geosci. 20: 511–84CrossRefGoogle Scholar
Pardo-Igúzquiza, E., Schwarzacher, W. and Rodríguez-Tovar, F. J. (2000). A library of computer programs for assisting teaching and research in cyclostratigraphic analysis. Comput. Geosci. 26: 723–40CrossRefGoogle Scholar
Park, J. and Herbert, T. D. (1987). Hunting for periodicities in a mid-Cretaceous sedimentary series. J. Geophys. Res. 92: 14,027–40CrossRefGoogle Scholar
Paul, H. A., Zachos, J. C., Flower, B. P. and Tripati, A. (2000). Orbitally induced climate and geochemical variability across the Oligocene/Miocene boundary. Paleoceanography 15: 471–85CrossRefGoogle Scholar
Pearce, R. B., Kemp, A. E. S., Baldauf, J. G. and King, S. C. (1995). High-resolution sedimentology and micropaleontology of laminated diatomaceous sediments from the eastern equatorial Pacific Ocean. Proc. Ocean Drill. Prog. Sci. Res. 138: 647–63Google Scholar
Pearn, W. C. (1964). Finding the ideal cyclothem. Kansas Geol. Surv. Bull. 169: 399–413Google Scholar
Pelletier, J. D. (1997). Analysis and modeling of the natural variability of climate. J. Clim. 10: 1331–422.0.CO;2>CrossRefGoogle Scholar
Peper, T. and Cloetingh, S. (1995). Autocyclic perturbations of orbitally forced signals in the sedimentary record. Geology 23: 937–402.3.CO;2>CrossRefGoogle Scholar
Percival, D. B. and Walden, A. T. (1993). Spectral Analysis for Physical Applications. Multitaper and Conventional Univariate Techniques. Cambridge University Press, Cambridge, pp. 1–583CrossRef
Pestiaux, P. and Berger, A. (1984a). An optimal approach to the spectral characteristics of deep sea climatic records. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 417–45
Pestiaux, P. and Berger, A. (1984b). Impacts of deep-sea processes on paleoclimatic spectra. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 493–510
Pestiaux, P., Duplessy, J. C. and Berger, A. (1987). Paleoclimatic variability at frequencies ranging from 10-4 cycle per year to 10-3 cycle per year — evidence for nonlinear behavior of the climate system. In: Climate History, Periodicity and Predictability, Eds: M. Rampino, J. E. Sanders, W. S. Newman and L. K., Kingsson. Van Nostrand Reinhold, New York, pp. 285–98
Peters, S. E. and Foote, M. (2002). Determinants of extinction in the fossil record. Nature 416: 420–4CrossRefGoogle ScholarPubMed
Petit, J. R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J.-M., Basile, I., Bender, M., Chappellez, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V. M., Legrand, M., Lipenkov, V. Y., Lorius, C., Pépin, L., Ritz, C., Saltzman, E. and Stievenard, M. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399: 429–36CrossRefGoogle Scholar
Petterson, G. (1996). Varved sediments in Sweden: a brief review. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Publication 116. The Geological Society, London, pp. 73–7CrossRef
Philander, S. G. (1990). El Niño, La Niña and the Southern Oscillation. Academic Press, London, pp. 1–293
Pilskaln, C. H. and Pike, J. (2001). Formation of Holocene sedimentary laminae in the Black Sea and the role of the benthic flocculent layer. Paleoceanography 16: 1–19CrossRefGoogle Scholar
Pisias, N. G. (1983). Geologic time series from deep-sea sediments: time scales and distortion by bioturbation. Mar. Geol. 51: 99–113CrossRefGoogle Scholar
Pisias, N. G. and Mix, A. C. (1988). Aliasing of the geologic record and the search for long-period Milankovitch cycles. Paleoceanography 3: 613–19CrossRefGoogle Scholar
Pisias, N. G. and Moore, T. C. (1981). The evolution of Pleistocene climate: a time series approach. Earth Planet. Sci. Lett. 52: 450–8CrossRefGoogle Scholar
Pisias, N. G., Mix, A. C. and Zahn, R. (1990). Nonlinear response in the global climate system: evidence from benthic oxygen isotope record in core RC13–110. Paleoceanography 5: 147–60CrossRefGoogle Scholar
Plaut, G., Ghil, M. and Vautard, R. (1995). Interannual and interdecadal variability from a long temperature time series. Science 268: 710–13CrossRefGoogle Scholar
Plotnick, R. E. (1986). A fractal model for the distribution of stratigraphic hiatuses. J. Geol. 94: 885–90CrossRefGoogle Scholar
Prell, W. L., Imbrie, J., Martinson, D. G., Morley, J. J., Pisias, N. G., Shackleton, N. J. and Streeter, H. F. (1986). Graphic correlation of oxygen isotope stratigraphy: application to the Late Quaternary. Paleoceanography 1: 137–62CrossRefGoogle Scholar
Press, W. H., Teukolsky, S. A., Vetterling, W. T. and Flannery, B. P. (1992). Numerical Recipes, the Art of Scientific Computing. Cambridge University Press, Cambridge, pp. 1–963
Preston, F. W. and Henderson, J. H. (1964). Fourier series characterization of cyclic sediments for stratigraphic correlation. In: Symposium on Cyclic Sedimentation, Ed: D. F. Merriam. Bull Kansas Geol. Surv. 169: 415–25
Priestley, M. B. (1981). Spectral Analysis and Time Series. Academic Press, London, pp. 1–890
Priestley, M. B. (1988). Non-linear and Non-stationary Time Series Analysis. Academic Press, London, pp. 1–237
Proakis, J. G. and Menolakis, D. G. (1996). Digital Signal Processing, Principles, Algorithms and Applications. Prentice Hall, London, pp. 1–968
Prokoph, A. and Agterberg, F. P. (1999). Detection of sedimentary cyclicity and stratigraphic completeness by wavelet analysis: an application of Late Albian cyclostratigraphy of the western Canada sedimentary basin. J. Sed. Res. 69: 862–75CrossRefGoogle Scholar
Prokoph, A. and Barthelmes, F. (1996). Detection of nonstationarities in geological time series: wavelet transform of chaotic and cyclic sequences. Comp. Geosci. 22: 1097–108CrossRefGoogle Scholar
Prokoph, A., Fowler, A. D. and Patterson, R. T. (2000). Evidence for periodicity and nonlinearity in a high-resolution fossil record of long-term evolution. Geology 28: 867–702.0.CO;2>CrossRefGoogle Scholar
Pugh, D. T. (1987). Tides, Surges and Mean Sea-level. Wiley, Chichester, pp. 1–472
Qin, X., Tan, M., Liu, T., Wang, X., Li, T. and Lu, J. (1999). Spectral analysis of 1000-year stalagmite lamina-thickness record from Shihua Cavern, Beijing, China, and its climatic significance. The Holocene 9: 689–94CrossRefGoogle Scholar
Quinn, T. M., Taylor, F. W. and Crowley, T. J. (1993). A 173 year stable isotope record from a tropical south Pacific coral. Quat. Sci. Rev. 12: 407–12CrossRefGoogle Scholar
Quinn, T. M., Crowley, T. J., Taylor, F. W., Henin, C., Joannot, P. and Join, Y. (1998). A multicentury stable isotope record from a New Caledonia coral: interannual and decadal sea surface temperature variability in the southwest Pacific since 1657 A.D. Paleoceanography 13: 412–26CrossRefGoogle Scholar
Quinn, W. H. (1992). A study of Southern Oscillation-related climatic activity for AD622–1990 incorporating Nile River flood data. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Eds: H. F. Diaz and V. Markgraf. Cambridge University Press, Cambridge, pp. 119–49
Ramsay, A. T. S., Sykes, T. J. S. and Kidd, R. B. (1994a). Sedimentary hiatuses as indicators of fluctuating oceanic water masses: a new model. J. Geol. Soc. Lond. 151: 737–40CrossRefGoogle Scholar
Ramsay, A. T. S., Sykes, T. J. S. and Kidd, R. B. (1994b). Waxing (and waning) lyrical on hiatuses: Eocene-Quaternary Indian Ocean hiatuses as proxy indicators of water mass production. Paleoceanography 9: 857–77CrossRefGoogle Scholar
Räsänen, M. E., Linna, A. M., Santos, J. C. R. and Negri, F. R. (1995). Late Miocene tidal deposits in the Amazonian Foreland Basin. Science 269: 386–9CrossRefGoogle ScholarPubMed
Rast, M. P., Fox, P. A., Lin, H., Lites, B. W., Meisner, R. W. and White, O. R. (1999). Bright rings around sunspots. Nature 401: 678–9CrossRefGoogle Scholar
Raup, D. M. and Sepkoski, J. J. (1988). Testing for periodicity of extinction. Science 241: 94–6CrossRefGoogle ScholarPubMed
Raymo, M. E., Ruddiman, W. F., Backman, J., Clement, B. M. and Martinson, D. G. (1989). Late Pliocene variations in northern hemisphere ice sheets and North Atlantic deep-water circulation. Paleoceanography 4: 413–46CrossRefGoogle Scholar
Raymo, M. E., Ganley, K., Carter, S., Oppo, D. W. and McManus, J. (1998). Millennial-scale climate instability during the early Pleistocene epoch. Nature 392: 699–702CrossRefGoogle Scholar
Reijmer, J. J. G., ten Kate, W. G. H. Z., Sprenger, A. and Schlager, W. (1991). Calciturbidite composition related to exposure and flooding of a carbonate platform (Triassic, Eastern Alps). Sedimentology 38: 1059–74CrossRefGoogle Scholar
Rempel, A. W., Waddington, E. D., Wettlaufer, J. S. and Worster, M. G. (2001). Possible displacement of the climate signal in ancient ice by premelting and anomalous diffusion. Nature 411: 568–71CrossRefGoogle ScholarPubMed
Rial, J. A. (1999). Pacemaking the ice ages by frequency modulation of Earth's orbital eccentricity. Science 285: 564–8CrossRefGoogle ScholarPubMed
Richards, G. R. (1994). Orbital forcing and endogenous interactions: non-linearity, persistence and convergence in late Pleistocene climate. Quat. Sci. Rev. 13: 709–25CrossRefGoogle Scholar
Ricken, W. (1986). Diagenetic Bedding: A Model for Marl-Limestone Alternations. Lecture Notes Earth Science 6. Springer, Berlin, pp. 1–210
Ricken, W. (1991a). Time span assessment — an overview. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 773–94
Ricken, W. (1991b). Variation of sedimentation rates in rhythmically bedded sediments. Distinction between deposition types. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 167–87
Ricken, W. (1993). Sedimentation as a Three-Component System. Springer, London
Ricken, W. and Eder, W. (1991). Diagenetic modification of calcareous beds — an overview. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 430–49
Ridgwell, A. J., Watson, A. J. and Raymo, M. E. (1999). Is the spectral signature of the 100 kyr glacial cycle consistent with a Milankovitch origin?Paleoceanography 14: 437–40CrossRefGoogle Scholar
Riegel, W. (1991). Coal cyclothems and some models for their origin. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 733–50
Ripepe, M. and Fischer, A. G. (1991). Stratigraphic rhythms synthesized from orbital variations. In: Sedimentary Modeling: Computer Simulations and Methods for Improved Parameter Definition, Eds: K. Franseen, W. L. Watney, C. G. St. C. Kendall and W. Ross. Kansas State Geol. Surv. Bull. 233: 335–44
Ripepe, M., Roberts, L. T. and Fischer, A. G. (1991). ENSO and sunspot cycles in varved Eocene oil shales from image analysis. J. Sed. Petrol. 61: 1155–63Google Scholar
Rittenour, T. M., Brigham-Grette, J. and Mann, M. E. (2000). El Niño-like climate teleconnections in New England during the Late Pleistocene. Science 288: 1039–42CrossRefGoogle ScholarPubMed
Robock, A. (1996). Stratigraphic control of climate. Science 272: 972–3CrossRefGoogle Scholar
Rodbell, D. T., Seltzer, G. O., Anderson, D. M., Abbott, M. B., Enfield, D. B. and Newman, J. H. (1999). An ∼15,000-year record of El Niño-driven alluviation in southwestern Ecuador. Science 283: 516–20CrossRefGoogle ScholarPubMed
Rodwell, M. J., Rowell, D. P. and Folland, C. K. (1999). Oceanic forcing of the wintertime North Atlantic Oscillation and European climate. Nature 398: 320–3CrossRefGoogle Scholar
Roulier, L. M. and Quinn, T. M. (1995). Seasonal- and decadal-scale climatic variability in southwest Florida during the middle Pliocene: inferences from a coralline stable isotope record. Paleoceanography 10: 429–43CrossRefGoogle Scholar
Rubincam, D. P. (1995). Has climate changed the Earth's tilt?Paleoceanography 10: 365–72CrossRefGoogle Scholar
Ruddiman, W. F. (1977). Late Quaternary deposition of ice-rafted sand in the subpolar North Atlantic (latitude 40°N to 65°N). Geol. Soc. Am. Bull. 88: 1813–272.0.CO;2>CrossRefGoogle Scholar
Ruddiman, W. F. (1985). Climate studies in ocean cores. In: Paleoclimate Analysis and Modelling, Ed: A. D. Hecht. Kluwer, The Netherlands, pp. 197–257
Ruddiman, W. F. and McIntyre, A. (1984). Ice-age thermal response and climatic role of the surface Atlantic Ocean, 40°N to 63°N. Geol. Soc. Am. Bull. 95: 381–962.0.CO;2>CrossRefGoogle Scholar
Ruddiman, W. F., Raymo, M. and McIntyre, A. (1986). Matayama 41,000-year cycles, North Atlantic Ocean and northern hemisphere ice sheets. Earth Planet. Sci. Lett. 80: 117–29CrossRefGoogle Scholar
Ruddiman, W. F., Raymo, M. E., Martinson, D. G., Clement, B. M. and Backman, J. (1989). Pleistocene evolution: northern hemisphere ice sheets and North Atlantic Ocean. Paleoceanography 4: 353–412CrossRefGoogle Scholar
Rutherford, S. and D'Hondt, S. (2000). Early onset and tropical forcing of 100,000-year Pleistocene glacial cycles. Nature 408: 72–5CrossRefGoogle ScholarPubMed
Sadler, P. M. (1981). Sediment accumulation rates and the completeness of stratigraphic sections. J. Geol. 89: 569–84CrossRefGoogle Scholar
Sadler, P. M. and Strauss, D. J. (1990). Estimation of completeness of stratigraphical sections using empirical data and theoretical models. J. Geol. Soc. Lond. 147: 471–85CrossRefGoogle Scholar
Saji, N. H., Goswami, B. N., Vinayachandran, P. N. and Yamagata, T. (1999). A dipole mode in the tropical Indian Ocean. Nature 401: 360–3CrossRefGoogle ScholarPubMed
Saltzman, B. and Verbitsky, M. (1994). Late Pleistocene climatic trajectory in the phase space of global ice, ocean state, and CO2: observations and theory. Paleoceanography 9: 767–79CrossRefGoogle Scholar
Sander, B. (1936). Beiträge zur Kenntnis der Anlargerungsgefüge. Mineral. Petrogr. Mitt. 48: 27–139Google Scholar
Schaaf, M. and Thurow, J. (1997). Tracing short cycles in long records: the study of inter-annual to inter-centennial climate change from long sediment records, examples from the Santa Barbara Basin. J. Geol. Soc. Lond. 154: 613–22CrossRefGoogle Scholar
Schiffelbein, P. (1984). Effect of benthic mixing on the information content of deep sea stratigraphical signals. Nature 311: 651–3CrossRefGoogle Scholar
Schiffelbein, P. and Dorman, L. (1986). Spectral effects of time-depth nonlinearities in deep sea sediment records: a demodulation technique for realigning time and depth scales. J. Geophys. Res. 91: 3821–35CrossRefGoogle Scholar
Schlesinger, M. E. and Ramankutty, N. (1994). An oscillation in the global climate system of period 65–70 years. Nature 367: 723–6CrossRefGoogle Scholar
Schmitz, W. J. and McCarthey, M. S. (1993). On the North Atlantic circulation. Rev. Geophys. 31: 29–49CrossRefGoogle Scholar
Scholz, C. H. (1998). Earthquakes and friction laws. Nature 391: 37–42CrossRefGoogle Scholar
Schulz, H., Rad, U. and Erlenkeuser, H. (1998). Correlation between Arabian Sea and Greenland climate oscillations of the past 110,000 years. Nature 393: 54–7CrossRefGoogle Scholar
Schulz, M. and Stattegger, K. (1997). Spectrum: spectral analysis of unevenly spaced paleoclimatic time series. Comput. Geosci. 23: 929–45CrossRefGoogle Scholar
Schuster, A. (1898). On the investigation of hidden periodicities with application to a supposed 26-day period of meteorological phenomenon. Terr. Mag. Atmos. Elect. 3: 13–41CrossRefGoogle Scholar
Schwarzacher, W. (1964). An application of statistical time-series analysis of a limestone-shale sequence. J. Geol. 72: 195–213CrossRefGoogle Scholar
Schwarzacher, W. (1975). Sedimentation Models and Quantitative Stratigraphy. Elsevier, Amsterdam, pp. 1–382
Schwarzacher, W. (1991). Milankovitch cycles and the measurement of time. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken, and A. Seilacher. Springer, London, pp. 855–63
Schwarzacher, W. (1993). Cyclostratigraphy and the Milankovitch Theory. Elsevier, Amsterdam, pp. 1–225
Schwarzacher, W. (1998). Stratigraphic resolution, cycles and sequences. In: Sequence Stratigraphy — Concepts and Applications, Eds: F. M. Gradstein, K. O. Sandvik and N. J. Milton. Norwegian Petroleum Society Special Publication No. 8. Elsevier, Amsterdam, pp. 1–8
Schwarzacher, W. and Fischer, A. G. (1982). Limestone-shale bedding and perturbations of the Earth's orbit. In: Cyclic and Event Stratification, Eds: G. Einsele and A. Seilacher. Springer, London, pp. 72–95CrossRef
Schweingruber, F. H., Echstein, D., Serre-Bachet, F. and Braeker, O. U. (1990). Identification, presentation and interpretation of event years and pointer years in dendrochronology. Dendrochronologia 8: 9–38Google Scholar
Seidov, D. and Maslin, M. (1999). North Atlantic deep water circulation collapse during Heinrich events. Geology 27: 23–62.3.CO;2>CrossRefGoogle Scholar
Sellwood, B. W. (1970). The relation of trace fossils to small scale sedimentary cycles in the British Lias. In: Trace Fossils, Eds: T. P. Crimes and J. C. Harper. Seel House Press, Liverpool, pp. 489–504
Sepkoski, J. J. (1989). Periodicity in extinction and the problem of catastrophism in the history of life. J. Geol. Soc. 146: 7–19CrossRefGoogle ScholarPubMed
Sepkoski, J. J. and Raup, D. M. (1986). Periodicity in marine extinction events. In: Dynamics of Extinction, Ed: D. K. Elliott. Wiley, Chichester, pp. 3–36
Shackleton, N. J. (2000). The 100,000-year ice-age cycle identified and found to lag temperature, carbon dioxide, and orbital eccentricity. Science 289: 1897–902CrossRefGoogle ScholarPubMed
Shackleton, N. J. and Crowhurst, S. (1997). Sediment fluxes based on an orbitally tuned time scale 5 Ma to 14 Ma, Site 926. Proc. Ocean Drill. Prog. 154: 69–82Google Scholar
Shackleton, N. J. and Imbrie, J. (1990). The δ18O spectrum of oceanic deep water over a five-decade band. Clim. Change 16: 217–30CrossRefGoogle Scholar
Shackleton, N. J. and Opdyke, N. D. (1973). Oxygen isotope and palaeomagnetic stratigraphy of equatorial Pacific core V28-238: oxygen isotope temperatures and ice volumes on a 105 year and 106 year scale. Quat. Res. 3: 39–55CrossRefGoogle Scholar
Shackleton, N. J. and Pisias, N. G. (1985). Atmospheric carbon dioxide, orbital forcing and climate. In: The Carbon Cycle and Atmospheric CO2: Natural Variations Archaean to Present, Eds: E. T. Sundquist and W. S. Broecker. Geophysical Monograph 32. American Geophysical Union, Washington, pp. 303–17CrossRef
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. R. Soc. Edinburgh, Earth Sci. 81: 251–61CrossRefGoogle Scholar
Shackleton, N. J., Crowhurst, S. J., Hagelberg, T., Pisias, N. G. and Schneider, D. A. (1995a). A new late Neogene timescale: application to Leg 138 sites. Proc. Ocean Drill. Prog. Sci. Res. 138: 73–101Google Scholar
Shackleton, N. J., Hagelberg, T. K. and Crowhurst, S. J. (1995b). Evaluating the success of astronomical tuning: pitfalls of using coherence as a criterion for assessing pre-Pleistocene timescales. Paleoceanography 10: 693–7CrossRefGoogle Scholar
Shackleton, N. J., Hall, M. A. and Pate, D. (1995c). Pliocene isotope stratigraphy of Site 846. Proc. Ocean Drill. Prog. Sci. Res. 138: 337–55Google Scholar
Shackleton, N. J., Crowhurst, S. J., Weedon, G. P. and Laskar, J. (1999a). Astronomical calibration of Oligocene-Miocene time. Philos. Trans. R. Soc. Lond. 357: 1907–29CrossRefGoogle Scholar
Shackleton, N. J., McCave, I. N. and Weedon, G. P. (Eds) (1999b). Astronomical (Milankovitch) calibration of the geological time-scale. Philos. Trans. R. Soc. Lond. 357: 1733–2007CrossRefGoogle Scholar
Shen, G. T., Cole, J. E., Lea, D. W., Linn, L. J., McConnaughey, E. A. and Fairbanks, R. G. (1992). Surface ocean variability at Galapagos from 1936–1982: calibration of geochemical tracers in corals. Paleoceanography 7: 563–88CrossRefGoogle Scholar
Shindell, D., Rind, D., Balachandran, N., Lean, J. and Lonergan, P. (1999). Solar cycle variability, ozone and climate. Science 284: 305–8CrossRefGoogle ScholarPubMed
Shindell, D. T., Schmidt, G. A., Mann, M. E., Rind, D. and Waple, A. (2001). Solar forcing of regional climate during the Maunder Minimum. Science 294: 2149–52CrossRefGoogle ScholarPubMed
Short, D. A., Mengel, J. G., Crowley, T. J., Hyde, W. T. and North, G. R. (1991). Filtering of Milankovitch cycles by Earth's geography. Quat. Res. 35: 157–73CrossRefGoogle Scholar
Sigman, D. M. and Boyle, E. A. (2000). Glacial/interglacial variations in atmospheric carbon dioxide. Nature 407: 859–69CrossRefGoogle ScholarPubMed
Sloan, L. C. and Huber, M. (2001). Eocene oceanic responses to orbital forcing on precessional time scales. Paleoceanography 16: 101–11CrossRefGoogle Scholar
Smith, D. G. (1994). Cyclicity or chaos? Orbital forcing versus non-linear dynamics. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 531–44CrossRef
Smith, N. D., Phillips, A. C. and Powell, R. D. (1990). Tidal drawdown: a mechanism for producing cyclic sediment laminations in glaciomarine deltas. Geology 18: 10–132.3.CO;2>CrossRefGoogle Scholar
Solanki, S. K., Schüssler, M. and Fligge, M. (2000). Evolution of the Sun's large-scale magnetic field since the Maunder minimum. Nature 408: 445–7CrossRefGoogle ScholarPubMed
Sonett, C. P. and Chan, M. A. (1998). Neoproterozoic Earth-Moon dynamics: rework of the 900Ma Big Cottonwood Canyon tidal laminae. Geophys. Res. Lett. 25: 539–42CrossRefGoogle Scholar
Sonett, C. P. and Finney, S. A. (1990). The spectrum of radiocarbon. Philos. Trans. R. Soc. Lond. 330A: 413–26CrossRefGoogle Scholar
Sonett, C. P. and Williams, G. E. (1985). Solar periodicities expressed in varves from glacial Skilak Lake, southern Alaska. J. Geophys. Res. 90: 12,019–26CrossRefGoogle Scholar
Sonett, C. P., Finney, S. A. and Williams, C. P. (1988). The lunar orbit in the late Precambrian and the Elatina sandstone laminae. Nature 335: 806–8CrossRefGoogle Scholar
Sonett, C. P., Kvale, E. P., Zakharian, A., Chan, M. J. and Demko, T. M. (1996). Late Proterozoic and Paleozoic tides, retreat of the Moon, and rotation of the Earth. Science 273: 100–4CrossRefGoogle Scholar
Spencer-Cervato, C. (1998). Changing depth distribution of hiatuses during the Cenozoic. Paleoceanography 13: 178–82CrossRefGoogle Scholar
Stauffer, B., Blunier, T., Dällenbach, A., Indermühle, A., Scwander, J., Stocker, T. F., Tschumi, J., Chappellaz, J., Raynaud, D., Hammer, C. U. and Clausen, H. B. (1998). Atmospheric CO2 concentration and millennial-scale climate change during the last glacial period. Nature 392: 59–62CrossRefGoogle Scholar
Stewart, I. (1990). Does God Play Dice? The New Mathematics of Chaos. Penguin, London, pp. 1–317
Stigler, S. M. and Wagner, M. J. (1987). A substantial bias in nonparametric tests for periodicity in geophysical data. Science 238: 940–5CrossRefGoogle ScholarPubMed
Stigler, S. M. and Wagner, M. J. (1988). Testing for periodicity of extinction. Science 241: 96–9CrossRefGoogle ScholarPubMed
Stockton, C. W., Boggess, W. R. and Meko, D. M. (1985). Climate and tree rings. In: Paleoclimate Analysis and Modelling, Ed: A. D. Hecht. Wiley, Chichester, pp. 71–161
Strauss, D. and Sadler, P. M. (1989). Stochastic models for the completeness of stratigraphic sections. J. Int. Assoc. Math. Geol. 21: 37–59CrossRefGoogle Scholar
Street-Perrott, F. and Perrott, R. A. (1990). Abrupt fluctuations in the tropics: the influence of Atlantic Ocean circulation. Nature 343: 607–12CrossRefGoogle Scholar
Stuiver, M. and Braziunas, T. F. (1989). Atmospheric 14C and century-scale solar oscillations. Nature 338: 405–7CrossRefGoogle Scholar
Stuiver, M. and Quay, P. D. (1980). Changes in atmospheric carbon-14 attributed to a variable sun. Science 207: 11–19CrossRefGoogle ScholarPubMed
Stuiver, M., Grootes, P. M. and Braziunas, T. F. (1995). The GISP2 δ18O climate record of the past 16,500 years and the role of the Sun, oceans and volcanoes. Quat. Res. 44: 341–54CrossRefGoogle Scholar
Stuiver, M., Braziunas, T. F., Grootes, P. M. and Zielinski, G. A. (1997). Is there evidence for solar forcing of climate in the GISP2 oxygen isotope record?Quat. Res. 48: 259–66CrossRefGoogle Scholar
Sugihara, G. and May, R. M. (1990). Nonlinear forecasting as a way of distinguishing chaos from measurement error in time series. Nature 344: 734–41CrossRefGoogle ScholarPubMed
Sutton, R. T. and Allen, M. R. (1997). Decadal predictability of North Atlantic sea surface temperature and climate. Nature 388: 563–7CrossRefGoogle Scholar
Sztanó, O. and Boer, P. L. (1995). Basin dimensions and morphology as controls on amplification of tidal motions (the Early Miocene North Hungarian Bay). Sedimentology 42: 665–82CrossRefGoogle Scholar
Taner, M. T., Koehler, F. and Sheriff, R. E. (1979). Complex seismic trace analysis. Geophysics 44: 1041–63CrossRefGoogle Scholar
Taylor, A. H., Jordan, M. B. and Stephens, J. A. (1998). Gulf Stream shifts following ENSO events. Nature 393: 638CrossRefGoogle Scholar
Taylor, C. A. (1965). Physics of Musical Sounds. English University Press, Aylesbury, pp. 1–196
Taylor, C. A. (1976). Sounds of Music. BBC, London, pp. 1–183
Taylor, K. C., Lamorey, G. W., Doyle, G. A., Alley, R. B., Grootes, P. M., Mayewski, P. A., White, J. W. C. and Barlow, L. K. (1993). The “flickering switch” of late Pleistocene climate change. Nature 361: 432–6CrossRefGoogle Scholar
Tessier, B. and Gigot, P. (1989). A vertical record of different tidal cyclicities: an example from the Miocene marine molasse of Digne (Haute Provence, France). Sedimentology 36: 767–76CrossRefGoogle Scholar
Thomson, D. J. (1982). Spectrum estimation and harmonic analysis. Proc. IEEE 70: 1055–96CrossRefGoogle Scholar
Thomson, D. J. (1990). Quadratic-inverse spectrum estimates; applications to paleoclimatology. Philos. Trans. R. Soc. Lond. 332A: 539–97CrossRefGoogle Scholar
Thomson, D. J. (1995). The seasons, global temperature, and precession. Science 268: 59–68CrossRefGoogle ScholarPubMed
Thomson, J., Higgs, N. C. and Clayton, T. (1995). A geochemical criterion for the recognition of Heinrich events and estimation of their depositional fluxes by the 230Thexcess profiling method. Earth Planet. Sci. Lett. 135: 29–43CrossRefGoogle Scholar
Thunell, R., Pride, C., Tappa, E. and Muller-Karger, F. (1993). Varve formation in the Gulf of California: insights from time series sediment trap sampling and remote sensing. Quat. Sci. Rev. 12: 451–64CrossRefGoogle Scholar
Thunell, R. C., Tappa, E. and Anderson, D. M. (1995). Sediment fluxes and varve formation in Santa Barbara Basin, offshore California. Geology 23: 1083–62.3.CO;2>CrossRefGoogle Scholar
Tiedemann, R. and Franz, S. O. (1997). Deep-water circulation, chemistry, and terrigenous sediment supply in the equatorial Atlantic during the Pliocene, 3.3–2.6 Ma and 5–4.5 Ma. Proc. Ocean Drill. Prog. Sci. Res. 154: 299–318Google Scholar
Tinsley, B. A. (1996). Correlations of atmospheric dynamics with solar wind induced changes of air-Earth current density into cloud tops. J. Geophys. Res. 101: 29,701–14CrossRefGoogle Scholar
Tinsley, B. A. (1997). Do effects of global atmospheric electricity on clouds cause climate changes?EOS Trans. Am. Geophys. Union 78: 341–9CrossRefGoogle Scholar
Tipper, J. C. (1983). Rates of sedimentation, and stratigraphical completeness. Nature 302: 696–8CrossRefGoogle Scholar
Tong, H. (1990). Non-linear Time Series. Oxford University Press, Oxford, pp. 1–564
Toon, O. B., Pollack, J. B., Ward, W., Burns, J. A. and Bilski, K. (1980). The astronomical theory of climate change on Mars. Icarus 44: 552–607CrossRefGoogle Scholar
Torrence, C. and Compo, G. P. (1998). A practical guide to wavelet analysis. Bull. Am. Meteorol. Soc. 79: 61–782.0.CO;2>CrossRefGoogle Scholar
Trauth, M. H. (1998). TURBO: a dynamic-probabilistic simulation to study the effects of bioturbation on paleoceanographic time series. Comput. Geosci. 24: 433–41CrossRefGoogle Scholar
Trauth, M. H., Sarnthein, M. and Arnold, M. (1997). Bioturbational mixing depth and carbon flux at the seafloor. Paleoceanography 12: 517–26CrossRefGoogle Scholar
Tsonis, A. A. and Elsner, J. B. (1992). Nonlinear prediction as a way of distinguishing chaos from random fractal sequences. Nature 358: 217–20CrossRefGoogle Scholar
Tudhope, A. W., Shimmield, G. B., Chilcott, C. P., Jebb, M., Fallick, A. E. and Dalgleish, A. N. (1995). Recent changes in climate in the far western equatorial Pacific and their relationship to the Southern Oscillation; oxygen isotope records from massive corals, Papua New Guinea. Earth Planet. Sci. Lett. 136: 575–90CrossRefGoogle Scholar
Turcotte, D. L. (1997). Fractals and Chaos in Geology and Geophysics, 2nd Edition. Cambridge University Press, Cambridge, pp. 1–398
Tziperman, E., Stone, L., Cane, M. A. and Jarosh, H. (1994). El Niño chaos: overlapping of resonances between the seasonal cycle and the Pacific ocean-atmosphere oscillator. Science 264: 72–4CrossRefGoogle ScholarPubMed
Ulrych, T. J. and Bishop, T. N. (1975). Maximum entropy spectral analysis and autoregressive decomposition. Rev. Geophys. Space Phys. 13: 183–200CrossRefGoogle Scholar
Urban, F. E., Cole, J. E. and Overpeck, J. T. (2000). Influence of mean climate change on climate variability from a 155-year tropical Pacific coral record. Nature 407: 989–93Google ScholarPubMed
Vail, P. R., Mitchum, R. M., Todd, R. G., Widmer, J. W., Thompson, S., Sangree, J. B., Bubb, J. N. and Hatlelid, W. G. (1977). Seismic stratigraphy and global changes of sealevel. In: Seismic Stratigraphy — Application to Hydrocarbon Exploration, Ed: C. E. Payton. Am. Assoc. Petrol. Geol. Mem. 26: 46–212
Vail, P. R., Audemard, F., Bowman, S. A., Eisner, P. N. and Perez-Cruz, C. (1991). The stratigraphic signatures of tectonics, eustasy and sedimentology — an overview. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 617–59
Van Echelpoel, E. (1994). Identification of regular sedimentary cycles using Walsh spectral analysis with results from the Boom Clay Formation, Belgium. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer, and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 63–74CrossRef
Geel, B., Raspopov, O. M., Renssen, H., Plicht, J., Dergachev, V. A. and Meijer, H. A. J. (1999). The role of solar forcing upon climate change. Quat. Sci. Rev. 18: 331–8CrossRefGoogle Scholar
Vautard, R. and Ghil., M. (1989). Singular spectrum analysis in non-linear dynamics, with applications to palaeoclimatic time series. Physica D 35: 395–424CrossRefGoogle Scholar
Visser, M. J. (1980). Neap-Spring cycles reflected in Holocene subtidal large-scale bedform deposits: a preliminary note. Geology 8: 543–62.0.CO;2>CrossRefGoogle Scholar
Wales, D. J. (1991). Calculating the rate of loss of information from chaotic time series by forecasting. Nature 350: 485–8CrossRefGoogle Scholar
Walther, G. (1997). Absence of correlation between the solar neutrino flux and the sunspot number. Phys. Rev. Lett. 79: 4522–4CrossRefGoogle Scholar
Wanless, H. R. and Weller, J. M. (1932). Correlation and extent of Pennsylvanian cyclothems. Geol. Soc. Am. Bull. 43: 1003–16CrossRefGoogle Scholar
Webster, P. J. and Palmer, T. N. (1997). The past and the future of El Niño. Nature 390: 562–4CrossRefGoogle Scholar
Webster, P. J., Moore, A. M., Loschnigg, J. P. and Leben, R. R. (1999). Coupled ocean-atmosphere dynamics in the Indian Ocean during 1997–98. Nature 401: 356–60CrossRefGoogle ScholarPubMed
Weedon, G. P. (1989). The detection and illustration of regular sedimentary cycles using Walsh power spectra and filtering, with examples from the Lias of Switzerland. J. Geol. Soc. Lond. 146: 133–44CrossRefGoogle Scholar
Weedon, G. P. (1993). The recognition and stratigraphic implications of orbital forcing of climate and sedimentary cycles. In: Sedimentology Review, Ed: V. P. Wright. Blackwell, Oxford, pp. 31–50CrossRef
Weedon, G. P. and Jenkyns, H. C. (1990). Regular and irregular climatic cycles and the Belemnite Marls (Pliensbachian, Lower Jurassic, Wessex Basin). J. Geol. Soc. Lond. 147: 915–18CrossRefGoogle Scholar
Weedon, G. P. and Jenkyns, H. C. (1999). Cyclostratigraphy and the Early Jurassic timescale: data from the Belemnite Marls, Dorset, southern England. Geol. Soc. Am. Bull. 111: 1823–402.3.CO;2>CrossRefGoogle Scholar
Weedon, G. P. and Read, W. A. (1995). Orbital-climatic forcing of Namurian cyclic sedimentation from spectral analysis of the Limestone Coal Group, Central Scotland. In: Orbital Forcing Timescales and Cyclostratigraphy, Eds: M. R. House and A. S. Gale. Geological Society Special Publication No. 85. The Geological Society, London, pp. 51–66CrossRef
Weedon, G. P. and Shimmield, G. B. (1991). Late Pleistocene upwelling and productivity variations in the northwest Indian Ocean deduced from spectral analyses of geochemical data from sites 722 and 724. Proc. Ocean Drill. Program Sci. Res. 117: 431–43Google Scholar
Weedon, G. P., Shackleton, N. J. and Pearson, P. N. (1997). The Oligocene timescale and cyclostratigraphy on the Ceara Rise, western equatorial Atlantic. Proc. Ocean Drill. Prog. Sci. Res. 154: 101–14Google Scholar
Weedon, G. P., Jenkyns, H. C., Coe, A. L. and Hesselbo, S. P. (1999). Astronomical calibration of the Jurassic time-scale from cyclostratigraphy in British mudrock formations. Philos. Trans. R. Soc. Lond. 357: 1787–813CrossRefGoogle Scholar
Weiss, N. O. (1990). Periodicity and aperiodicity in solar magnetic activity. Philos. Trans. R. Soc. Lond. 330A: 617–25CrossRefGoogle Scholar
Wells, J. W. (1963). Coral growth and geochronometry. Nature 197: 948–50CrossRefGoogle Scholar
Weltje, G. and Boer, P. L. (1993). Astronomically induced paleoclimatic oscillations reflected in Pliocene turbidite deposits on Corfu (Greece): implications for the interpretation of higher order cyclicity in ancient turbidite systems. Geology 21: 307–102.3.CO;2>CrossRefGoogle Scholar
Whitcombe, L. J. (1996). A FORTRAN program to calculate tidal heights using the simplified harmonic method of tidal prediction. Comp. Geosci. 22: 817–21CrossRefGoogle Scholar
White, J. C. W., Barlow, L. K., Fisher, D., Grootes, P., Jouzel, J., Johnsen, S. J., Stuiver, M. and Clausen, H. (1997). The climate signal in the stable isotopes of snow from Summit, Greenland: results of comparisons with modern climate observations. J. Geophys. Res. 102: 26,425–39CrossRefGoogle Scholar
Wiesenfield, K. and Moss, F. (1995). Stochastic resonance and the benefits of noise: from ice ages to crayfish and SQUIDS. Nature 373: 33–6CrossRefGoogle Scholar
Wilkinson, B. H., Drummond, C. N., Rothman, E. D. and Diedrich, N. W. (1997). Stratal order in peritidal carbonate sequences. J. Sed. Res. 67: 1068–82Google Scholar
Wilkinson, B. H., Drummond, C. N., Diedrich, N. W. and Rothman, E. D. (1999). Poisson processes of carbonate accumulation on Paleozoic and Holocene platforms. J. Sed. Res. 69: 338–50CrossRefGoogle Scholar
Williams, G. E. (1981). Sunspot periods in the late Precambrian glacial climate and solar-planetary relations. Nature 291: 624–8CrossRefGoogle Scholar
Williams, G. E. (1986). The solar cycle in Precambrian time. Sci. Am. 255: 88–95CrossRefGoogle Scholar
Williams, G. E. (1988). Cyclicity in the Late Precambrian Elatina Formation, South Australia: solar or tidal signature?Clim. Change 13: 117–28CrossRefGoogle Scholar
Williams, G. E. (1989). Late Precambrian tidal rhythmites in South Australia and the history of the Earth's rotation. J. Geol. Soc. Lond. 146: 97–111CrossRefGoogle Scholar
Williams, G. E. (1991). Milankovitch-band cyclicity in bedded halite deposits contemporaneous with Late Ordovician-Early Silurian glaciation, Canning Basin, western Australia. Earth Planet. Sci. Lett. 103: 143–55CrossRefGoogle Scholar
Williams, G. E. and Sonett, C. P. (1985). Solar signature in sedimentary cycles from the late Precambrian Elatina Formation, Australia. Nature 318: 523–7CrossRefGoogle Scholar
Williams, G. P. (1997). Chaos Theory Tamed. National Academy Press, Washington, pp. 1–499
Williams, R. B. G. (1984). Introduction to Statistics for Geographers and Earth Scientists. Macmillan, London, pp. 1–349CrossRef
Willson, R. C. and Hudson, H. S. (1988). Solar luminosity variations in solar cycle 21. Nature 332: 810–12CrossRefGoogle Scholar
Willson, R. C. and Hudson, H. S. (1991). The Sun's luminosity over a complete solar cycle. Nature 351: 42–4CrossRefGoogle Scholar
Wilson, D. S. (1993). Confirmation of the astronomical calibration of the magnetic polarity timescale from sea-floor spreading rates. Nature 364: 788–90CrossRefGoogle Scholar
Worthington, P. F. (1990). Sediment cyclicity from well logs. In: Geological Applications of Wireline Logs, Eds: A. Hurst, M. A. Lovell and A. C. Morton. Geological Society Special Publication No. 48. The Geological Society, London, pp. 123–32CrossRef
Wunsch, C. (1999). The interpretation of short climate records, with comments on the North Atlantic and Southern Oscillations. Bull. Am. Meteorol. Soc. 80: 245–552.0.CO;2>CrossRefGoogle Scholar
Wunsch, C. (2000a). On sharp spectral lines in the climate record and the millennial peak. Paleoceanography 15: 417–24CrossRefGoogle Scholar
Wunsch, C. (2000b). Moon, tides and climate. Nature 405: 743–4CrossRefGoogle Scholar
Yang, C-S. and Baumfalk, Y. A. (1994). Milankovitch cyclicity in the Upper Rotliegend Group of the Netherlands offshore. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication No. 19. Blackwell, Oxford, pp. 47–61CrossRef
Yang, C-S. and Nio, S-D. (1985). The estimation of palaeohydrodynamic processes from subtidal deposits using time series analysis methods. Sedimentology 32: 41–57CrossRefGoogle Scholar
Yiou, P., Baert, E. and Loutre, M. F. (1996). Spectral analysis of climate data. Surve. Geophys. 17: 619–63CrossRefGoogle Scholar
Yiou, P., Sornette, D. and Ghil, M. (2000). Data-adaptive wavelets and multi-scale singular-spectrum analysis. Physica D 142: 254–90CrossRefGoogle Scholar
Young, P. C. (1999). Nonstationary time series analysis. Prog. Environ. Sci. 1: 3–48Google Scholar
Yu, Z. and Ito, E. (1999). Possible forcing of century-scale drought frequency in the northern Great Plains. Geology 27: 263–62.3.CO;2>CrossRefGoogle Scholar
Zachos, J. C., Flower, B. P. and Paul, H. (1997). Orbitally paced climate oscillations across the Oligocene/Miocene boundary. Nature 388: 567–70CrossRefGoogle Scholar
Zachos, J. C., Shackleton, N. J., Revenaugh, J. S., Palike, H. and Flower, B. P. (2001). Climate response to orbital forcing across the Oligocene-Miocene boundary. Science 292: 274–8CrossRefGoogle Scholar
Zhang, R-H., Rothstein, L. M. and Busalacchi, A. J. (1998). Origin of upper-ocean warming and El Niño change on decadal scales in the tropical Pacific Ocean. Nature 391: 879–83CrossRefGoogle Scholar
Zolitschka, B. (1996). Image analysis and microscopic investigation of annually laminated lake sediments from Fatetteville Green Lake (NY, USA) Lake C2 (NWT, Canada) and Holzmaar (Germany): a comparison. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication No. 116. The Geological Society, London, pp. 49–55
Algeo, T. J. (1993). Quantifying stratigraphic completeness: a probabilistic approach using paleomagnetic data. J. Geol. 101: 421–33CrossRefGoogle Scholar
Algeo, T. J. and Woods, A. D. (1994). Microstratigraphy of the Lower Mississippian Sunbury Shale: a record of solar-modulated climatic cyclicity. Geology 22: 795–82.3.CO;2>CrossRefGoogle Scholar
Allen, J. R. L. (1981). Lower Cretaceous tides revealed by cross-bedding with mud drapes. Nature 289: 579–81CrossRefGoogle Scholar
Allen, M. R. and Smith, L. A. (1996). Monte Carlo SSA: detecting irregular oscillations in the presence of colored noise. J. Clim. 9: 373–4042.0.CO;2>CrossRefGoogle Scholar
Allen, P. A. and Homewood, P. (1984). Evolution and mechanics of a Miocene tidal sandwave. Sedimentology 31: 63–81CrossRefGoogle Scholar
Alley, R. B. and MacAyeal, D. R. (1994). Ice-rafted debris associated with binge/purge oscillations of the Laurentide Ice Sheet. Paleoceanography 9: 503–11CrossRefGoogle Scholar
Alley, R. B., Shuman, C. A., Meese, D. A., Gow, A. J., Taylor, K. C., Cuffey, K. M., Fitzpatrick, J. J., Grootes, P. M., Zielinski, G. A., Ram, M., Spinelli, G. and Elser, B. (1997). Visual-stratigraphic dating of the GISP2 ice core: basis, reproducibility, and application. J. Geophys. Res. 102: 26,367–81CrossRefGoogle Scholar
Alley, R. B., Anandakrishnan, S. and Jung, P. (2001). Stochastic resonance in the North Atlantic. Paleoceanography 16: 190–8CrossRefGoogle Scholar
Anders, M. H., Krueger, S. W. and Sadler, P. M. (1987). A new look at sedimentation rates and the completeness of the stratigraphic record. J. Geol. 95: 1–14CrossRefGoogle Scholar
Anderson, D. M. (2001). Attenuation of millennial-scale events by bioturbation in marine sediments. Paleoceanography 16: 352–7CrossRefGoogle Scholar
Anderson, R. Y. (1961). Solar-terrestrial climatic patterns in varved sediments. Ann. NY Acad. Sci. USA 95: 424–35CrossRefGoogle Scholar
Anderson, R. Y. (1982). A long geoclimatic record from the Permian. J. Geophys. Res. 87: 7285–94CrossRefGoogle Scholar
Anderson, R. Y. (1984). Orbital forcing of evaporite sedimentation. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 147–62
Anderson, R. Y. (1986). The varve microcosm: propagator of cyclic bedding. Paleoceanography 1: 373–82CrossRefGoogle Scholar
Anderson, R. Y. (1992a). Long term changes in the frequency of occurrence of El Niño events. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Eds: H. F. Diaz and V. Markgraf. Cambridge University Press, Cambridge, pp. 193–200
Anderson, R. Y. (1992b). Possible connection between surface winds, solar activity and the Earth's magnetic field. Nature 358: 51–3CrossRefGoogle Scholar
Anderson, R. Y. (1996). Seasonal sedimentation: a framework for reconstructing climatic and environmental change. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication, No. 116. The Geological Society, London, pp. 1–15CrossRef
Anderson, R. Y. and Dean, W. E. (1988). Lacustrine varve formation through time. Palaeogeog. Palaeoclim. Palaeoecol. 62: 215–35CrossRefGoogle Scholar
Anderson, R. Y. and Koopmans, L. H. (1963). Harmonic analysis of varve time series. J. Geophys. Res. 68: 877–93CrossRefGoogle Scholar
Andrews, J. T., Jennings, A. E., Kerwin, M., Kirby, M., Manley, W., Miller, G. H., Bond, G. and MacLean, B. (1995). A Heinrich-like event, H-0 (DC-0): source(s) for detrital carbonate in the North Atlantic during the Younger Dryas chronozone. Paleoceanography 10: 943–52CrossRefGoogle Scholar
Andrews, J. T., Barber, D. C. and Jennings, A. E. (1999). Errors in generating time series and in dating events at Late Quaternary millennial (radiocarbon) time-scales: examples from Baffin Bay, NW Labrador Sea, and East Greenland. In: Mechanisms of Global Climate Change at Millennial Time Scales, Eds: P. U. Clark, R. S. Webb and L. D. Keigwin. Geophysical Monograph 112. American Geophysical Union, Washington, pp. 23–33CrossRef
Appenzeller, C., Stocker, T. F. and Anklin, M. (1998). North Atlantic Oscillation dynamics recorded in Greenland ice cores. Science 282: 446–9CrossRefGoogle ScholarPubMed
Archer, A. W. (1996). Reliability of lunar orbital periods extracted from ancient cyclic tidal rhythmites. Earth Planet. Sci. Lett. 141: 1–10CrossRefGoogle Scholar
Archer, A. W. and Johnson, T. W. (1997). Modelling of cyclic tidal rhythmites (Carboniferous of Indiana and Kansas, Precambrian of Utah, USA) as a basis for reconstruction of intertidal positioning and palaeotidal regimes. Sedimentology 44: 991–1010CrossRefGoogle Scholar
Archer, A. W., Kuecher, G. J. and Kvale, E. P. (1995). The role of tidal-velocity asymmetries in the deposition of silty tidal rhythmites (Carboniferous, eastern Interior Coal Basin, U.S.A.). J. Sed. Res. A65: 408–16Google Scholar
Arz, H. W., Pätzold, J. and Wefer, G. (1998). Correlated millennial-scale changes in surface hydrography and terrigenous sediment yield inferred from last-glacial marine deposits off Northeastern Brazil. Quat. Res. 50: 157–66CrossRefGoogle Scholar
Baker, P. A., Rigsby, C. A., Sltzer, G. O., Fritz, S. C., Lowenstein, T. K., Bacher, N. P. and Veliz, C. (2001). Tropical climate changes at millennial and orbital timescales on the Bolivian Altiplano. Nature 409: 698–701CrossRefGoogle ScholarPubMed
Barahona, M. and Poon, C-S. (1996). Detection of non-linear dynamics in short, noisy time series. Nature 381: 215–17CrossRefGoogle Scholar
Bard, E. (2001). Paleoceanographic implications of the difference in deep-sea sediment mixing between large and fine particles. Paleoceanography 16: 235–9CrossRefGoogle Scholar
Barlow, L. K., White, J. W. C., Barry, R. G., Rogers, J. C. and Grootes, P. M. (1993). The North Atlantic Oscillation signature in deuterium and deuterium excess signals in the Greenland Ice Sheet Project 2 ice core, 1840–1970. Geophys. Res. Lett. 20: 2901–4CrossRefGoogle Scholar
Barrell, J. (1917). Rhythms and the measurement of geological time. Bull. Geol. Soc. Am. 28: 745–904CrossRefGoogle Scholar
Barry, R. G. and Chorley, R. J. (1998). Atmosphere Weather and Climate, seventh edition. Rutledge, London, pp. 1–409
Beattie, P. D. and Dade, W. D. (1996). Is scaling in turbidite deposition consistent with forcing by earthquakes?J. Sed. Res. 66: 909–15Google Scholar
Beauchamp, K. G. (1975). Walsh Functions and their Applications. Academic Press, London, pp. 1–236
Beauchamp, K. G. (1984). Applications of Walsh and Related Functions, with an Introduction to Sequency Theory. Academic Press, New York, pp. 1–308
Beaufort, L. (1994). Climatic importance of the modulation of the 100 kyr cycle inferred from 16 m.y. long Miocene records. Paleoceanography 9: 821–34CrossRefGoogle Scholar
Beer, J., Blinov, A., Bonani, G., Finkel, R. C., Hofmann, H. J., Lehmann, B., Oeschger, H., Sigg, A., Schwander, J., Staffelbach, T., Stauffer, B., Suter, M. and Wölfli, W. (1990). Use of 10Be in polar ice to trace the 11-year cycle of solar activity. Nature 347: 164–6CrossRefGoogle Scholar
Beer, J., Mende, W. and Stellmacher, R. (2000). The role of the sun in climate forcing. Quat. Sci. Rev. 19: 403–15CrossRefGoogle Scholar
Beerbower, J. R. (1964). Cyclothems and cyclic depositional mechanisms in alluvial plain sedimentation. In: Symposium on Cyclic Sedimentation, Ed: D. F. Merriam. Kansas Geol. Surv. Bull. 169: 31–42Google Scholar
Behl, R. J., and Kennett, J. P. (1996). Brief interstadial events in the Santa Barbara basin, NE Pacific, during the past 60 kyr. Nature 379: 243–6CrossRefGoogle Scholar
Bender, M., Sowers, T., Dickson, M-L., Orchardo, J., Grootes, P., Mayewski, P. A. and Meese, D. A. (1994). Climate correlations between Greenland and Antarctica during the past 100,000 years. Nature 372: 663–6CrossRefGoogle Scholar
Benzi, R., Parisi, G., Sutera, A. and Vulpiani, A. (1982). Stochastic resonance in climatic change. Tellus 34: 10–16CrossRefGoogle Scholar
Berger, A. (1984). Accuracy and frequency stability of the Earth's orbital elements during the Quaternary. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 3–39
Berger, A. (1988). Milankovitch Theory and climate. Rev. Geophys. 26: 624–57CrossRefGoogle Scholar
Berger, A. (1989). The spectral characteristics of pre-Quaternary climate records, an example of the relationship between Astronomical Theory and Geosciences. In: Climate and Geosciences, Eds: A. Berger, S. Schneider, and J. C. Duplessy. Kluwer, Dordrecht, pp. 47–76
Berger, A. and Loutre, M. F. (1994). Astronomical forcing through geological time. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 15–24CrossRef
Berger, A. and Loutre, M. F. (1997). Intertropical latitudes and precessional and half-precessional cycles. Science 278: 1476–8CrossRefGoogle Scholar
Berger, A., Imbrie, J., Hays, J. D., Kukla, G. and Saltzman, B. (Eds) (1984). Milankovitch and Climate, Understanding the Response to Astronomical Forcing, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, 2 volumes, pp. 1–895
Berger, A., Loutre, M. F. and Dehant, V. (1989). Influence of the changing lunar orbit on the astronomical frequencies of pre-Quaternary insolation patterns. Paleoceanography 4: 555–64CrossRefGoogle Scholar
Berger, A., Mélice, J. L. and Mersch, I. (1990). Evolutive spectral analysis of sunspot data over the past 300 years. Philos. Trans. R. Soc. Lond. 330A: 529–41CrossRefGoogle Scholar
Berger, A., Loutre, M. F. and Mélice, J. L. (1997). Instability of the astronomical periods from 1.5 Myr BP to 0.5 Myr BP. Palaeoclimates 4: 1–42Google Scholar
Berggren, W. A., Kent, D. V., Swisher, C. C. and Aubry, M. (1995). A revised Cenozoic geochronology and chronostratigraphy. In: Geochronology Time Scales and Global Stratigraphic Correlation, Eds: W. A. Berggren, D. V. Kent, M. Aubry and J. Hardenbol. Society of Economic Paleontologists and Mineralogists (SEPM) Special Publication No. 54, pp. 129–212CrossRef
Bezrukov, S. M. and Vodyanoy, I. (1997). Stochastic resonance in non-dynamical systems without response thresholds. Nature 385: 319–21CrossRefGoogle ScholarPubMed
Bianchi, G. G. and McCave, I. N. (1999). Holocene periodicity in North Atlantic climate and deep-ocean flow south of Iceland. Nature 397: 515–17CrossRefGoogle Scholar
Bigg, G. R. (1996). The Oceans and Climate. Cambridge University Press, Cambridge, pp. 1–266
Biondi, F., Lange, C. B., Hughes, M. K. and Berger, W. H. (1997). Inter-decadal signals during the last millennium (AD 1117–1992) in the varve record of Santa Barbara basin, California. Geophys. Res. Lett. 24: 193–6CrossRefGoogle Scholar
Black, D. E., Peterson, L. C., Overpeck, J. T., Kaplan, A., Evans, M. N. and Kashgarian, M. (1999). Eight centuries of North Atlantic ocean atmosphere variability. Science 286: 1709–13CrossRefGoogle ScholarPubMed
Bloomfield, P. (1976). Fourier Analysis of Time Series: an Introduction. Wiley, London
Bond, G. and Lotti, R. (1995). Iceberg discharges into the North Atlantic on millennial time scales during the last glaciation. Science 267: 1005–10CrossRefGoogle ScholarPubMed
Bond, G. C., Heinrich, H., Broecker, W. S., Labeyrie, L., McManus, J. F., Andrews, J. T., Huon, S., Jantschik, R., Clasen, S., Simet, C., Tedesco, K., Klas, M., Bonani, G. and Ivy, S. (1992). Evidence for massive discharges of icebergs into the Northern Atlantic Ocean during the last glacial period. Nature 360: 245–9CrossRefGoogle Scholar
Bond, G., Broecker, W., Johnsen, S., McManus, J., Labeyrie, L., Jouzel, J. and Bonani, G. (1993). Correlations between climate records from North Atlantic sediments and Greenland ice. Nature 365: 143–7CrossRefGoogle Scholar
Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., deMenocal, P., Priore, P., Cullen, H., Hajdas, I. and Bonani, G. (1997). A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science 278: 1257–66CrossRefGoogle Scholar
Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M. N., Showers, W., Hoffmann, S., Lotti-Bond, R., Hajdas, I. and Bonani, G. (2001). Persistent solar influence on North Atlantic climate during the Holocene. Science 294: 2130–6CrossRefGoogle ScholarPubMed
Boss, S. K. and Rasmussen, K. A. (1995). Misuse of Fischer plots as sea-level curves. Geology 23: 221–42.3.CO;2>CrossRefGoogle Scholar
Boygle, J. (1993). The Swedish varve chronology — a review. Prog. Phys. Geog. 17: 1–19CrossRefGoogle Scholar
Bradley, W. H. (1929). The varves and climate of the Green River epoch. US Geol. Surv. Prof. Pap. 158: 87–110Google Scholar
Briffa, K. R. (2000). Annual climate variability in the Holocene: interpreting the message of ancient trees. Quat. Sci. Rev. 19: 87–105CrossRefGoogle Scholar
Briffa, K. R. and Osborn, T. J. (1999). Seeing the wood from the trees. Science 284: 926–7CrossRefGoogle Scholar
Briffa, K. R., Jones, P. D., Schweingruber, F. H. and Osborn, T. J. (1998a). Influence of volcanic eruptions on Northern hemisphere summer temperature over the past 600 years. Nature 393: 450–5CrossRefGoogle Scholar
Briffa, K. R., Schweingruber, F. H., Jones, P. D., Osborn, T. J., Harris, I. C., Shiyatov, S. G., Vaganov, E. A. and Grudd, H. (1998b). Trees tell of past climates: but are they speaking less clearly today?Philos. Trans. R. Soc. Lond. 353B: 65–73CrossRefGoogle Scholar
Briffa, K. R., Schweingruber, F. H., Jones, P. D., Osborn, T. J., Shiyatov, S. G. and Vaganov, E. A. (1998c). Reduced sensitivity of recent tree-growth to temperature at high northern latitudes. Nature 391: 678–82CrossRefGoogle Scholar
Briffa, K. R., Bartholin, T. S., Eckstein, D., Jones, P. D., Karlén, W., Schweingruber, F. H. and Zetterberg, P. (1990). A 1,400-year tree-ring record of summer temperatures in Fennoscandia. Nature 346: 434–9CrossRefGoogle Scholar
Broecker, W. S. (1997). Thermohaline circulation, the Achilles heel of our climate system: will man-made CO2 upset the current balance?Science 278: 1582–8CrossRefGoogle ScholarPubMed
Broecker, W. S. and Denton, G. H. (1990). The role of ocean-atmosphere reorganizations in glacial cycles. Geochim. Cosmochim. Acta 53: 2464–501Google Scholar
Broecker, W. S., Bond, G., Klas, M., Bonani, G. and Wolfli, W. (1990). A salt oscillator in the glacial Atlantic? 1, the concept. Paleoceanography 5: 469–77CrossRefGoogle Scholar
Broecker, W. S., Bond, G., Klas, M., Clark, E. and McManus, J. F. (1992). Origin of the northern Atlantic's Heinrich events. Clim. Dynam. 6: 265–73CrossRefGoogle Scholar
Brook, E. J., Sowers, T. and Orchardo, J. (1994). Rapid variations in methane concentration during the past 110,000 years. Science 273: 1087–91CrossRefGoogle Scholar
Buchanan, M. (2000). Ubiquity. The Science of History, …, or Why the World is Simpler than We Think. Weidenfield and Nicholson, London, pp. 1–230
Bull, D., Kemp, A. E. S. and Weedon, G. P. (2000). A 160,000 year old record of El Niño-Southern Oscillation in marine production and coastal run-off from Santa Barbara Basin, California. Geology 28: 1007–102.0.CO;2>CrossRefGoogle Scholar
Burroughs, W. J. (1992). Weather Cycles. Real or Imaginary? Cambridge University Press, Cambridge, pp. 1–207
Cande, S. C. and Kent, D. V. (1995). Revised calibration of the geomagnetic polarity time scale for the Late Cretaceous and Cenozoic. J. Geophys. Res. 100: 6093–5CrossRefGoogle Scholar
Cane, M. A. (1992). Tropical Pacific ENSO models: ENSO as a mode of the coupled system. In: Climate System Modelling, Ed: K. E. Trenberth. Cambridge University Press, Cambridge, pp. 583–614
Cannariato, K. G., Kennett, J. P. and Behl, R. J. (1999). Biotic response to late Quaternary rapid climate switches in Santa Barbara Basin: ecological and evolutionary implications. Geology 27: 63–62.3.CO;2>CrossRefGoogle Scholar
Carrs, B. W. and Neidell, N. S. (1966). A geological cyclicity detected by means of polarity coincidence correlation. Nature 212: 136–7CrossRefGoogle Scholar
Chan, M. J., Kvale, E., Archer, A. W. and Sonett, C. P. (1994). Oldest direct evidence of lunar-solar tidal forcing encoded in sedimentary rhythmites, Proterozoic Big Cottonwood Formation, central Utah. Geology 22: 791–42.3.CO;2>CrossRefGoogle Scholar
Chang, P., Ji, L. and Li, H. (1997). A decadal climate variation in the tropical Atlantic Ocean from thermodynamic air-sea interactions. Nature 385: 516–18CrossRefGoogle Scholar
Christensen, C. J., Gorsline, D. S., Hammond, D. E. and Lund, S. P. (1994). Non-annual laminations and explanation of anoxic basin-floor conditions in Santa Monica Basin, California Borderland, over the past four centuries. Mar. Geol. 116: 399–418CrossRefGoogle Scholar
Cisne, J. L. (1986). Earthquakes recorded stratigraphically on carbonate platforms. Nature 323: 320–2CrossRefGoogle Scholar
Clark, P. U., Webb, R. S. and Keigwin, L. D. (Eds) (1999). Mechanisms of Global Climate Change at Millennial Time Scales. Geophysical Monograph 112. American Geophysical Union, Washington, pp. 1–394CrossRef
Clark, P. U., Pisias, N. G., Stocker, T. F. and Weaver, A. J. (2002). The role of the thermohaline circulation in abrupt climate change. Nature 415: 863–9CrossRefGoogle ScholarPubMed
Cleaveland, M. K., Cook, E. R. and Stahle, D. W. (1992). Secular variability of the Southern Oscillation detected in tree-ring data from Mexico and the southern United States. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Eds: H. F. Diaz and V Markgraf. Cambridge University Press, Cambridge, pp. 271–91
Clemens, S. C. (1999). An astronomical tuning strategy for Pliocene sections: implications for global-scale correlation and phase relationships. Philos. Trans. R. Soc. Lond. 357: 1949–73CrossRefGoogle Scholar
Clemens, S. C. and Prell, W. L. (1990). Late Pleistocene variability of Arabian Sea summer monsoon winds and continental aridity: eolian records from the lithogenic component of deep-sea sediments. Paleoceanography 5: 109–45CrossRefGoogle Scholar
Clemens, S. C., Murray, D. W. and Prell, W. L. (1996). Nonstationary phase of the Plio-Pleistocene Asian Monsoon. Science 274: 943–8CrossRefGoogle ScholarPubMed
Clement, A. C., Seager, R. and Cane, M. A. (1999). Orbital controls on the El Niño/Southern Oscillation and the tropical climate. Paleoceanography 14: 441–56CrossRefGoogle Scholar
Cohen, A. L., Layne, G. D., Hart, S. R. and Lobel, P. S. (2001). Kinetic control of skeletal Sr/Ca in a symbiotic coral: implications for the paleotemperature proxy. Paleoceanography 16: 20–6CrossRefGoogle Scholar
Cole, J. E., Dunbar, R. B., McClanahan, T. R. and Muthiga, N. A. (2000). Tropical Pacific forcing of decadal SST variability in the western Indian Ocean over the past two centuries. Science 287: 617–19CrossRefGoogle ScholarPubMed
Cook, E. R., D'Arrigo, R. D. and Briffa, K. R. (1998). A reconstruction of the North Atlantic Oscillation using tree-ring chronologies from North America and Europe. The Holocene 8: 9–17CrossRefGoogle Scholar
Corrège, T., Delcroix, T., Récy, J., Beck, W., Caboich, G. and Cornec, F. L. (2000). Evidence for stronger El Niño-Southern Oscillation (ENSO) events in a mid-Holocene massive coral. Paleoceanography 15: 465–70CrossRefGoogle Scholar
Croll, J. (1864). On the physical cause of the change of climate during geological epochs. Phil. Mag. 28: 121–37CrossRefGoogle Scholar
Crowley, K. D., Duchon, C. E. and Rhi, J. (1986). Climate record in varved sediments of the Eocene Green River Formation. J. Geophys. Res. 91: 8637–47CrossRefGoogle Scholar
Crowley, T. J., Kim, K-Y., Mengel, J. G. and Short, D. A. (1992). Modeling 100,000-year climate fluctuations in pre-Pleistocene time series. Science 255: 705–7CrossRefGoogle ScholarPubMed
Crusius, J. and Anderson, R. F. (1992). Inconsistencies in accumulation rates of Black Sea sediments inferred from records of laminae and 210Pb. Paleoceanography 7: 215–27CrossRefGoogle Scholar
Cullen, H. M., D'Arrigo, R. D. and Cook, E. R. (2001). Multiproxy reconstructions of the North Atlantic. Paleoceanography 16: 27–39CrossRefGoogle Scholar
Currie, R. G. (1987). Examples and implications of 18.6- and 11-yr terms in world weather records. In: Climate History, Periodicity and Predictability, Eds: M. Rampino, J. E. Sanders, W. S. Newman and L. K. Kingsson. Van Nostrand Reinhold, New York, pp. 378–403
Curry, R. G., McCartney, M. S. and Joyce, T. M. (1998). Oceanic transport of subpolar climate signals to mid-depth subtropical waters. Nature 391: 575–7CrossRefGoogle Scholar
Curry, W. B. and Oppo, D. W. (1997). Synchronous high-frequency oscillations in tropical sea surface temperatures and North Atlantic deep water production during the last glacial cycle. Paleoceanography 12: 1–14CrossRefGoogle Scholar
Dalfes, H. N., Schneider, S. H. and Thompson, S. L. (1984). Effects of bioturbation on climatic spectra inferred from deep sea cores. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kula, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 481–92
Dalrymple, R. W. and Makino, Y. (1989). Description and genesis of tidal bedding in the Cobequid Bay — Salmon River estuary, Bay of Fundy, Canada. In: Sedimentary Facies of Active Plate Margins, Eds: A. Taira and F. Masuda. Terra, Tokyo, pp. 151–77
Dansgaard, W., Johnsen, S. J., Clausen, H. B., Dahl-Jensen, D., Gundestrup, N. S., Hammer, C. U., Hvidberg, C. S., Steffensen, J. P., Sveinbjörnsdottir, A. E., Jouzel, J. and Bond, G. (1993). Evidence for general instability of past climate from a 250-kyr ice-core record. Nature 364: 218–20CrossRefGoogle Scholar
D'Arrigo, R. D., Cook, E. R., Jacby, G. C. and Briffa, K. R. (1993). NAO and sea surface temperature signatures in tree-ring records from the North Atlantic sector. Quat. Sci. Rev. 12: 431–40CrossRefGoogle Scholar
Davis, J. C. (1973). Statistics and Data Analysis in Geology, first edition. Wiley, London, pp. 1–550
Davis, J. C. (1986). Statistics and Data Analysis in Geology, second edition. Wiley, Chichester, pp. 1–646
de Boer, P. L. and Smith, D. G. (Eds) (1994). Orbital Forcing and Cyclic Sequences. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 1–559
de Boer, P. L. and Wonders, A. A. H. (1984). Astronomically induced rhythmic bedding in Cretaceous pelagic sediments near Moria, Italy. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 177–90
Boer, P. L., Oost, A. P. and Visser, M. J. (1989). The diurnal inequality of the tide as a parameter for recognizing tidal influences. J. Sed. Petrol. 59: 912–21Google Scholar
Dehant, V., Loutre, M-F. and Berger, A. (1990). Potential impact of the Northern Hemisphere Quaternary ice sheets on the frequencies of the astroclimatic orbital parameters. J. Geophys. Res. 95: 7573–8CrossRefGoogle Scholar
Dessai, S. and Walter, M. E. (2000). Self-organised criticality and the atmospheric sciences: selected review, new findings and future directions. Contributed paper, NSF Workshop on Extreme Events (http://www.esig.ucar.edu/extremes/papers.html)
Dettinger, M. D., Ghil, M., Strong, C. M., Weibel, W. and Yiou, P. (1995). Software expedites singular-spectrum analysis of noisy time series. EOS Trans. AGU 76: 12–21Google Scholar
Diaz, H. E. and Markgraf, V. (Eds) (1992). El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation. Cambridge University Press, Cambridge
Dicke, R. H. (1978). Is there a chronometer hidden deep in the Sun?Nature 276: 676–80CrossRefGoogle Scholar
Dicke, R. H. (1979). Solar luminosity and the sunspot cycle. Nature 280: 24–7CrossRefGoogle Scholar
Dimitrov, B. D., Shangova-Gigoriadi, S. and Grigoriadis, E. D. (1998). Cyclicity in variations of incidence rates for breast cancer in different countries. Folia Med. (Plovdiv) 40: 66–71Google ScholarPubMed
Dokken, T. M. and Jansen, E. (1999). Rapid changes in the mechanism of ocean convection during the last glacial period. Nature 401: 458–61CrossRefGoogle Scholar
Droxler, A. W. and Schlager, W. (1985). Glacial versus interglacial sedimentation rates and oxygen-isotope record in the Bahamas. Geology 13: 799–8022.0.CO;2>CrossRefGoogle Scholar
Duff, P. M. D. and Walton, E. K. (1962). Statistical basis for cyclothems: a quantitative study of the sedimentary succession in the east Pennine coalfield. Sedimentology 1: 235–55CrossRefGoogle Scholar
Duff, P. M. D., Hallam, A. and Walton, E. K. (1967). Cyclic Sedimentation. Developments in Stratigraphy. Elsevier, Amsterdam, pp. 1–280
Dunbar, R. B., Wellington, G. M., Colgan, M. W. and Glynn, P. W. (1994). Eastern Pacific sea surface temperature since 1600AD: the δ18O record of climatic variability in Galápagos corals. Paleoceanography 9: 291–315CrossRefGoogle Scholar
Dunn, C. E. (1974). Identification of sedimentary cycles through Fourier analysis of geochemical data. Chem. Geol. 13: 217–32CrossRefGoogle Scholar
Duvall, T. L., D'Silva, S., Jefferies, S. M., Harvey, J. W. and Schou, J. (1996). Downflows under sunspots detected by helioseismic tomography. Nature 379: 235–7CrossRefGoogle Scholar
Eddy, J. A. (1976). The Maunder minimum. Science 192: 1189–202CrossRefGoogle ScholarPubMed
Egbert, G. D. and Ray, R. D. (2000). Significant dissipation of tidal energy in the deep ocean inferred from satellite altimeter data. Nature 405: 775–8CrossRefGoogle ScholarPubMed
Einsele, G. and Seilacher, A. (1991). Distinction of tempestites and turbidites. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, Berlin, pp. 377–82
Einsele, G., Ricken, W. and Seilacher, A. (1991). Cycles and events in stratigraphy — basic concepts and terms. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 1–19
Elliot, M., Labeyrie, L., Dokken, T. and Manthe, S. (2000). Coherent patterns of ice rafted debris deposits in the Nordic regions during the last glacial (10–60 ka). Earth Planet. Sci. Lett. 194: 151–63CrossRefGoogle Scholar
Elrick, M. and Hinnov, L. A. (1996). Millennial-scale climate origins for stratification in Cambrian and Devonian deep-water rhythmites, western USA. Palaeogeog. Palaeoclim. Palaeoecol. 123: 353–72CrossRefGoogle Scholar
Elrick, M., Read, J. A. and Coruh, C. (1991). Short-term paleoclimatic fluctuations expressed in lower Mississippian ramp-slope deposits, southwestern Montana. Geology 19: 799–8022.3.CO;2>CrossRefGoogle Scholar
Emery, W. J. and Thomson, R. E. (1997). Data Analysis Methods in Physical Oceanography. Elsevier, Amsterdam, pp. 1–634
Eriksson, K. A. and Simpson, E. L. (2000). Quantifying the oldest tidal record: the 3.2 Ga Moodies Group, Barberton Greenstone Belt, South Africa. Geology 28: 831–42.0.CO;2>CrossRefGoogle Scholar
Esper, J., Cook, E. R. and Schweingruber, F. H. (2002). Low-frequency signals in long tree-ring chronologies for reconstructing past temperature variability. Science 295: 2250–3CrossRefGoogle ScholarPubMed
Evans, J. W. (1972). Tidal growth increments in the cockle Clinocardium nuttalli. Science 176: 416–17CrossRefGoogle ScholarPubMed
Farley, K. A. and Patterson, D. B. (1995). A 100-kyr periodicity in the flux of extraterrestrial 3He to the seafloor. Nature 378: 600–3Google Scholar
Faure, G. (1986). Principles of Isotope Geology, second edition. Wiley, New York, pp. 1–589
Fedorov, A. V. and Philander, S. G. (2000). Is El Niño changing?Science 288: 1997–2002CrossRefGoogle ScholarPubMed
Finkel, R. C. and Nishiizumi, K. (1997). Beryllium 10 concentrations in the Greenland Ice Sheet Project 2 ice core from 3–40 ka. J. Geophys. Res. 102: 26,699–706CrossRefGoogle Scholar
Fischer, A. G. (1980). Gilbert — bedding rhythms and geochronology. Spec. Pap. Geol. Soc. Am. 183: 93–104Google Scholar
Fischer, A. G. (1986). Climatic rhythms recorded in strata. Annu. Rev. Earth Planet. Sci. 14: 351–76CrossRefGoogle Scholar
Fischer, A. G. and Bottjer, D. J. (Eds) (1991). Orbital forcing and sedimentary sequences. J. Sed. Petrol. 61: 1063–252Google Scholar
Fischer, A. G. and Roberts, L. T. (1991). Cyclicity in the Green River Formation (Lacustrine Eocene) of Wyoming. J. Sed. Petrol. 61: 1146–54Google Scholar
Fischer, A. G. and Schwarzacher, W. (1984). Cretaceous bedding rhythms under orbital control? In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 163–75
Fischer, A. G., de Boer, P. L. and Premoli Silva, I. (1990). Cyclostratigraphy. In: Cretaceous, Resources, Events and Rhythms, Eds: R. N. Ginsburg and B. Beaudoin. Kluwer, Dordrecht, pp. 139–72CrossRef
Fisher, R. A. (1929). Tests of significance in harmonic analysis. Proc. R. Soc. Series A, 125: 54–9CrossRefGoogle Scholar
Forte, A. M. and Mitrovica, J. X. A. (1997). A resonance in the Earth's obliquity and precession over the past 20 Myr driven by mantle convection. Nature 390: 676–9CrossRefGoogle Scholar
Foucault, A., Powichrowski, L. and Prud'Homme, A. (1987). Le côntrole astronomique de la sédimentation turbiditique: exemple du Flysch à Helminthoides des Alpes Ligures (Italie). C. R. Acad. Sci. Paris 305: 1007–11Google Scholar
Foukal, P. (1990). Solar luminosity variations over timescales of days to the past few solar cycles. Philos. Trans. R. Soc. Lond. 330A: 591–9CrossRefGoogle Scholar
Francois, R. and Bacon, M. (1994). Heinrich events in the North Atlantic: radiochemical evidence. Deep-Sea Res. 41: 315–34CrossRefGoogle Scholar
Friis-Christensen, E. and Lassen, K. (1991). Length of the solar cycle: an indicator of solar activity closely associated with climate. Science 254: 698–700CrossRefGoogle ScholarPubMed
Fröhlich, C. and Lean, J. (1998). The Sun's total irradiance: cycles, trends and related climate change uncertainties since 1976. Geophys. Res. Lett. 25: 4377–80CrossRefGoogle Scholar
Fronval, T., Jansen, E., Bloemendal, J. and Johnsen, S. (1995). Oceanic evidence for coherent fluctuations in Fennoscandian and Laurentide ice sheets on millennium timescales. Nature 374: 443–6CrossRefGoogle Scholar
Gagan, M. K., Ayliffe, L. K., Beck, J. W., Cole, J. E., Druffel, E. R. M., Dunbar, R. B. and Schrag, D. P. (2000). New views of tropical paleoclimates from corals. Quat. Sci. Rev. 19: 45–64CrossRefGoogle Scholar
Gallois, R. W. (2000). The stratigraphy of the Kimmeridge Clay (Upper Jurassic) in the RGGE Project boreholes at Swanworth Quarry and Metherhills, south Dorset. Proc. Geol. Assoc. 111: 265–80CrossRefGoogle Scholar
Ganopolski, A. and Rahmstorf, S. (2001). Rapid changes of global climate simulated in a coupled climate model. Nature 409: 153–8CrossRefGoogle Scholar
Ganopolski, A. and Rahmstorf, S. (2002). Abrupt glacial climate changes due to stochastic resonance. Phys. Rev. Lett. 88: Article 038501CrossRefGoogle ScholarPubMed
Gershenfeld, N., Schoner, B. and Metois, E. (1999). Cluster-weighted modelling for time-series analysis. Nature 397: 329–32CrossRefGoogle Scholar
Gilbert, G. K. (1895). Sedimentary measurement of geologic time. J. Geol. 3: 351–76CrossRefGoogle Scholar
Gilliland, R. L. (1981). Solar radius variations over the past 265 years. Astrophys. J. 248: 1144–55CrossRefGoogle Scholar
Gipp, M. R. (2001). Interpretation of climate dynamics from phase space portraits: is the climate system strange or just different?Paleoceanography 16: 335–51CrossRefGoogle Scholar
Glatzmaier, G. A., Coe, R. S., Hongre, L. and Roberts, P. H. (1999). The role of the Earth's mantle in controlling the frequency of geomagnetic reversals. Nature 401: 885–90CrossRefGoogle Scholar
Goldhammer, R. K., Dunn, P. A. and Hardie, L. A. (1990). Depositional cycles, composite sea-level changes, cycle stacking patterns and the hierarchy of stratigraphic forcing: examples from Alpine Triassic platform carbonates. Geol. Soc. Am. Bull. 102: 535–622.3.CO;2>CrossRefGoogle Scholar
Gorsline, D. S., Nava-Sanchez, E. and Murillo de Nava, J. (1996). A survey of occurrences of Holocene laminated sediments in California borderland basins: products of a variety of depositional processes. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication No. 116. The Geological Society, London, pp. 93–110CrossRef
Gough, D. (2000). News from the solar interior. Science 287: 2434–5CrossRefGoogle Scholar
Gradstein, F. M., Agterberg, F. P., Ogg, J. G., Hardenbol, J., Veen, P., Thierry, J. and Huang, Z. (1994). A Mesozoic Time Scale. J Geophys. Res. 99: 24,051–74CrossRefGoogle Scholar
Grassberger, P. (1986). Do climatic attractors exist?Nature 323: 609–12CrossRefGoogle Scholar
Greenland Ice-core Project Members (1993). Climate instability during the last interglacial period recorded in the GRIP ice core. Nature 364: 203–7CrossRef
Grootes, P. M. and Stuiver, M. (1997). Oxygen 18/16 variability in Greenland snow and ice with 10-3 to 105-year time resolution. J. Geophys. Res. 102: 26,455–470CrossRefGoogle Scholar
Gu, D. and Philander, S. G. H. (1997). Interdecadal climate fluctuations that depend on exchanges between the tropics and extratropics. Science 275: 805–7CrossRefGoogle ScholarPubMed
Guyodo, Y. and Valet, J-P. (1999). Global changes in intensity of the Earth's magnetic field during the past 800 kyr. Nature 399: 249–52CrossRefGoogle Scholar
Gwiazda, R. H., Hemming, S. R. and Broecker, W. S. (1996). Provenance of icebergs during Heinrich event 3 and the contrast to their sources during other Heinrich episodes. Paleoceanography 11: 371–8CrossRefGoogle Scholar
Haak, A. B. and Schlager, W. (1989). Compositional variations in calciturbidites due to sea-level fluctuations, late Quaternary, Bahamas. Geol. Runds. 78: 477–86CrossRefGoogle Scholar
Hagadorn, J. W. (1996). Laminated sediments of Santa Monica Basin, California continental borderland. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication No. 116. The Geological Society, London, pp. 111–20CrossRef
Hagelberg, T. K. and Pisias, N. G. (1990). Nonlinear response of Pliocene climate to orbital forcing: evidence from the eastern equatorial Pacific. Paleoceanography 5: 595–617CrossRefGoogle Scholar
Hagelberg, T. K., Pisias, N. G. and Elgar, S. (1991). Linear and nonlinear couplings between orbital forcing and the marine δ18O record during the late Neogene. Paleoceanography 6: 729–46CrossRefGoogle Scholar
Hagelberg, T. K., Bond, G. and deMenocal, P. (1994). Milankovitch band forcing of sub-Milankovitch climate variability during the Pleistocene. Paleoceanography 9: 545–58CrossRefGoogle Scholar
Haigh, J. D. (1996). The impact of solar variability on climate. Science 272: 981–4CrossRefGoogle ScholarPubMed
Halfman, J. D. and Johnson, T. C. (1988). High resolution record of cyclic climatic change during the past 4 ka from Lake Turkana, Kenya. Geology 16: 496–5002.3.CO;2>CrossRefGoogle Scholar
Hall, I. R., McCave, I. N., Shackleton, N. J., Weedon, G. P. and Harris, S. E. (2001). Glacial intensification of deep Pacific inflow and ventilation. Nature 412: 809–12CrossRefGoogle Scholar
Hallam, A. (1964). Origin of the limestone-shale rhythms in the Blue Lias of England: a composite theory. J. Geol. 72: 157–68CrossRefGoogle Scholar
Hammer, C., Mayewski, P. A., Peel, D. and Stuiver, M. (Eds) (1997). GISP2 and GRIP results. J. Geophys. Res. 102 part C12Google Scholar
Hansen, D. V. and Bezdek, H. F. (1996). On the nature of decadal anomalies in North Atlantic sea surface temperature. J. Geophys. Res. 101: 9749–58CrossRefGoogle Scholar
Hartmann, W. M. (1999). How we localize sound. Phys. Today 52: 24–9CrossRefGoogle Scholar
Hays, J. D., Imbrie, I. and Shackleton, N. J. (1976). Variations in the Earth's orbit: pacemaker of the ice ages. Science 194: 1121–32CrossRefGoogle ScholarPubMed
Heinrich, H. (1988). Origin and consequences of cyclic ice rafting in the northeast Atlantic Ocean during the past 130,000 years. Quat. Res. 29: 143–52CrossRefGoogle Scholar
Hendy, I. L. and Kennett, J. P. (1999). Latest Quaternary North Pacific surface-water responses imply atmosphere-driven climate instability. Geology 27: 291–42.3.CO;2>CrossRefGoogle Scholar
Herbert, T. D. (1992). Paleomagnetic calibration of Milankovitch cyclicity in Lower Cretaceous sediments. Earth Planet. Sci. Lett. 112: 15–28CrossRefGoogle Scholar
Herbert, T. D. (1993). Differential compaction in lithified deep-sea sediments is not evidence for “diagenetic unmixing”. Sediment. Geol. 84: 115–22CrossRefGoogle Scholar
Herbert, T. D. (1994). Reading orbital signals distorted by sedimentation: models and examples. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 483–507CrossRef
Herbert, T. D. and Fischer, A. G. (1985). Milankovitch climatic origin of mid-Cretaceous black shale rhythms in central Italy. Nature 321: 739–43CrossRefGoogle Scholar
Herschel, J. F. W. (1832). On the astronomical causes which may influence geological phenomena. Trans. Geol. Soc. 2nd Ser. 3: 393–9Google Scholar
Hibler, W. D. and Johnsen, S. J. (1979). The 20-yr cycle in Greenland ice core records. Nature 280: 481–3CrossRefGoogle Scholar
Hilgen, F. J. (1991). Astronomical calibration of Gauss to Matuyama sapropels in the Mediterranean and implications for the geomagnetic polarity timescale. Earth Planet. Sci. Lett. 104: 226–44CrossRefGoogle Scholar
Hilgen, F. J. and Langereis, G. C. (1988). The age of the Miocene-Pliocene boundary in the Capo Rossello area (Sicily). Earth Planet. Sci. Lett. 91: 214–22CrossRefGoogle Scholar
Hilgen, F. J. and Langereis, C. G. (1989). Periodicities of CaCO3 cycles in the Pliocene of Sicily: discrepancies with the quasi-periods of the Earth's orbital cycles?Terra Nova, 1: 409–15CrossRefGoogle Scholar
Hilgen, F. J., Krijgsman, W., Langereis, C. G., Lourens, L. J., Santarelli, A. and Zachariasse, W. J. (1995). Extending the astronomical (polarity) time scale into the Miocene. Earth Planet. Sci. Lett. 136: 495–510CrossRefGoogle Scholar
Hilgen, F. J., Abdul Aziz, W., Krijgsman, W., Langereis, C. G., Lourens, L. J., Meulenkamp, J. E., Raffi, I., Steenbrink, J., Turco, E., Vught, N., Wijbrans, J. R. and Zachariasse, W. J. (1999). Present status of the astronomical (polarity) time-scale for the Mediterranean Late Neogene. Philos. Trans. R. Soc. 357: 1931–47CrossRefGoogle Scholar
Hilgen, F., Schwarzacher, W. and Strasser, A. (2001). Concepts and definitions in cyclostratigraphy — Second Report of the cyclostratigraphy working group. (International Subcommission on Stratigraphic Nomenclature, IUGS Commission of Stratigraphy). Unpublished, 8p
Hinnov, L. A. (2000). New perspectives on orbitally-forced stratigraphy. Annu. Rev. Earth Planet. Sci. 28: 419–75CrossRefGoogle Scholar
Hinnov, L. A. and Goldhammer, R. K. (1991). Spectral analysis of the Middle Triassic Latemar Limestone. J. Sed. Petrol. 61: 1173–93Google Scholar
Hinnov, L. A. and Park, J. (1998). Detection of astronomical cycles in the stratigraphic record by frequency modulation (FM) analysis. J. Sed. Res. 68: 524–39CrossRefGoogle Scholar
Hinnov, L. A. and Park, J. (1999). Strategies for assessing Early-Middle (Pliensbachian-Aalenian) Jurassic cyclochronologies. Philos. Trans. R. Soc. Lond. 357: 1831–59CrossRefGoogle Scholar
Holmgren, K., Karlén, W., Lauritzen, S. E., Lee-Thorp, J. A., Partridge, T. C., Piketh, S., Repinski, P., Stevenson, C., Svanered, O. and Tyson, P. D. (1999). A 3000-year high-resolution stalagmite-based record of palaeoclimate for northeastern South Africa. The Holocene 9: 295–309CrossRefGoogle Scholar
House, M. R. (1985). A new approach to an absolute timescale from measurements of orbital cycles and sedimentary microrhythms. Nature 316: 721–5CrossRefGoogle Scholar
House, M. R. and Farrow, G. E. (1968). Daily growth banding in the shell of the cockle, Cardium edule. Nature 219: 1384–6CrossRefGoogle ScholarPubMed
House, M. R. and Gale, A. S. (Eds) (1995). Orbital Forcing Timescales and Cyclostratigraphy. Geological Society Special Publication No. 85. The Geological Society, London, pp. 51–66CrossRef
Howe, R., Christensen-Dalsgaard, J., Hill, F., Komm, R. W., Larsen, R. M., Schou, J., Thompson, M. J. and Toomre, J. (2000). Dynamic variations at the base of the solar convective zone. Science 287: 2456–60CrossRefGoogle ScholarPubMed
Hoyt, V. and Schatten, K. H. (1997). The Role of the Sun in Climate Change. Oxford University Press, Oxford, pp. 1–279
Hubbard, B. B. (1996). The World According to Wavelets. The Story of a Mathematical Technique in the Making. A. K. Peters, Wellesley, Massachusetts, pp. 1–264
Hulme, M. and Barrow, E. (Eds) (1997). Climates of the British Isles. Routledge, London, pp. 1–454
Hunt, A. G. and Malin, P. E. (1998). Possible triggering of Heinrich events by ice-load-induced earthquakes. Nature 393: 155–8Google Scholar
Hurrell, J. W. (1995). Decadal trends in the North Atlantic Oscillation: regional temperatures and precipitation. Science 269: 676–9CrossRefGoogle ScholarPubMed
Hydrographic Office (1996). Admiralty Tide Tables 1997, Volume. 1, United Kingdom and Ireland including European Channel Ports. Hydrographic Office (UK)
Ifeachor, E. C. and Jervis, B. W. (1993). Digital Signal Processing. A Practical Approach. Addison-Wesley, Harlow, pp. 1–760
Imbrie, J. and Imbrie, J. Z. (1980). Modelling the climate response to orbital variations. Science 207: 943–53CrossRefGoogle Scholar
Imbrie, J. and Imbrie, K. P. (1979). Ice Ages. Solving the Mystery. Harvard University Press, Cambridge, Massachusetts, pp. 1–224CrossRef
Imbrie, J., Hays, J. D., Martinson, D. G., McIntyre, A. C., Mix, A. C., Morley, J. J., Pisias, N. G., Prell, W. L. and Shackleton, N. J. (1984). The orbital theory of Pleistocene climate: support from a revised chronology of the marine δ18O record. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 269–305
Imbrie, J., Berger, A., Boyle, E. A., Clemens, S. C., Duffy, A., Howard, W. R., Kukla, G., Kutzbach, J., Martinson, D. G., McIntyre, A., Mix, A. C., Molfino, B., Morley, J. J., Peterson, L. C., Pisias, N. G., Prell, W. L., Raymo, M. E., Shackleton, N. J. and Toggweiler, J. R. (1992). On the structure and origin of major glaciation cycles, 1: Linear responses to Milankovitch forcing. Paleoceanography 7: 701–38CrossRefGoogle Scholar
Imbrie, J., Berger, A., Boyle, E. A., Clemens, S. C., Duffy, A., Howard, W. R., Kukla, G., Kutzbach, J., Martinson, D. G., McIntyre, A., Mix, A. C., Molfino, B., Morley, J. J., Peterson, L. C., Pisias, N. G., Prell, W. L., Raymo, M. E., Shackleton, N. J. and Toggweiler, J. R. (1993a). On the structure and origin of the major glaciation cycles, 2: The 100,000-year cycle. Paleoceanography 8: 699–735CrossRefGoogle Scholar
Imbrie, J., Berger, A. and Shackleton, N. J. (1993b). Role of orbital forcing: a two-million-year perspective. In: Global Changes in the Perspective of the Past, Eds: J. A. Eddy and H. Oeschger. Wiley, Chichester, pp. 263–77
Ito, M., Nishikawa, T. and Sugimoto, H. (1999). Tectonic control of high-frequency depositional sequences with durations shorter than Milankovitch cyclicity: an example from the Pleistocene paleo-Tokyo Bay, Japan. Geology 27: 763–62.3.CO;2>CrossRefGoogle Scholar
James, I. N. and James, P. M. (1989). Ultra-low-frequency variability in a simple atmospheric circulation model. Nature 342: 53–5CrossRefGoogle Scholar
Jenkins, G. M. and Watts, D. G. (1969). Spectral Analysis and its Applications. Holden-Day, London
Jin, F-F., Neelin, J. D. and Ghil, M. (1994). El Niño on the Devil's staircase: annual subharmonic steps to chaos. Science 264: 70–2CrossRefGoogle Scholar
Johnsen, S. J., Clausen, H. B., Dansgaard, W., Gundestrup, N. S., Hammer, C. U., Andersen, U., Andersen, K. K., Hvidberg, C. S., Dahl-Jensen, D., Steffensen, J. P., Shoji, H., Sveinbjörnsdottir, Á.E., White, J., Jouzel, J. and Fisher, D. (1997). The δ18O record along the Greenland Ice Core Project deep ice core and the problem of possible Eemian climatic instability. J. Geophys. Res. 102: 26,397–410CrossRefGoogle Scholar
Jones, C. E., Jenkyns, H. C. and Hesselbo, S. B. (1994). Strontium isotopes in early Jurassic seawater. Geochim. Cosmochim. Acta 58: 3061–74CrossRefGoogle Scholar
Kanasewich, E. R. (1981). Time Sequence Analysis in Geophysics. University of Alberta Press, Alberta
Kantz, H. and Schreiber, T. (1997). Nonlinear Time Series Analysis. Cambridge University Press, Cambridge, pp. 1–304
Keeling, C. D. and Whorf, T. P. (2000). The 1,800-oceanic tidal cycle: a possible cause of rapid climate change. Proc. Natl. Acad. Sci. USA 97: 3814–19CrossRefGoogle ScholarPubMed
Keigwin, L. D. and Lehmann, S. J. (1994). Deep circulation change linked to HEINRICH event 1 and Younger Dryas in a mid depth North Atlantic core. Paleoceanography 9: 185–94CrossRefGoogle Scholar
Kemp, A. E. S. (Ed.) (1996). Palaeoclimatology and Palaeoceanography from Laminated Sediments. Geological Society Publication No. 116. The Geological Society, London
Kemp, A. E. S. and Baldauf, J. G. (1993). Vast Neogene diatom mat deposits from the eastern equatorial Pacific Ocean. Nature 362: 141–4CrossRefGoogle Scholar
Kemp, A. E. S., Baldauf, J. G. and Pearce, R. B. (1995). Origins and paleoceanographic significance of laminated diatom ooze from the eastern equatorial Pacific. Proc. Ocean Drill. Prog. Sci. Res. 138: 641–5Google Scholar
Kennedy, J. A. and Brassell, S. C. (1992). Molecular records of twentieth-century El Niño events in laminated sediments from the Santa Barbara basin. Nature 357: 62–4CrossRefGoogle Scholar
Kennett, J. P. and Ingram, B. L. (1995). A 20,000-year record of ocean circulation and climate change from the Santa Barbara Basin. Nature 377: 510–14CrossRefGoogle Scholar
Kent, D. V. (1999). Orbital tuning of geomagnetic polarity time-scales. Philos. Trans. R. Soc. Lond. 357: 1995–2007CrossRefGoogle Scholar
Kerr, R. A. (1999). Link between sunspots, stratosphere buoyed [sic]. Science 284: 234–5CrossRefGoogle Scholar
King, T. (1996). Quantifying nonlinearity and geometry in time series of climate. Quat. Sci. Rev. 15: 247–66CrossRefGoogle Scholar
Kirchner, J. W. (2002). Evolutionary speed limits inferred from the fossil record. Nature 415: 65–8CrossRefGoogle ScholarPubMed
Kominz, M. A. (1996). Whither cyclostratigraphy? Testing the gamma method on upper Pleistocene deep-sea sediments, North Atlantic Deep Sea Drilling Project site 609. Paleoceanography 11: 481–504CrossRefGoogle Scholar
Kominz, M. A. and Bond, G. (1990). A new method of testing periodicity in cyclic sediments — application to the Newark Supergroup. Earth Planet. Sci. Lett. 98: 233–44CrossRefGoogle Scholar
Kominz, M. A., Beavan, J., Bond, G. C. and McManus, J. (1991). Are cyclic sediments periodic? Gamma analysis and spectral analysis of Newark Supergroup lacustrine strata. In: Sedimentary Modeling: Computer Simulations and Methods for Improved Parameter Definition, Eds: K. Franseen, W. L. Watney, C. G. St. C. Kendall and W. Ross. Bull. Kansas Geol. Surv. 233: 319–34
Kortenkamp, S. J. and Dermott, S. F. (1998). A 100,000-year periodicity in the accretion rate of interplanetary dust. Science 280: 874–6CrossRefGoogle ScholarPubMed
Kotilainen, A. T. and Shackleton, N. J. (1995). Rapid climate variability in the North Pacific Ocean during the past 95,000 years. Nature 377: 323–6CrossRefGoogle Scholar
Krijgsman, W., Hilgen, F. J., Langereis, C. G. and Zachariasse, W. J. (1994a). The age of the Tortonian-Messinian boundary. Earth Planet. Sci. Lett. 121: 533–47CrossRefGoogle Scholar
Krijgsman, W., Langereis, C. G., Daams, R. and Meulen, A. J. (1994b). Magnetostratigraphic dating of the Middle Miocene climate change in the continental deposits of the Aragonian type area in the Calatayud-Teruel basin (central Spain). Earth Planet. Sci. Lett. 128: 513–26CrossRefGoogle Scholar
Krijgsman, W., Hilgen, F. J., Langereis, C. G., Santarelli, A. and Zachariasse, W. J. (1995). Late Miocene magnetostratigraphy, biostratigraphy and cyclostratigraphy in the Mediterranean. Earth Planet. Sci. Lett. 136: 475–94CrossRefGoogle Scholar
Kumar, K. K., Rajagopalan, B. and Cane, M. A. (1999). On the weakening relationship between the Indian Monsoon and ENSO. Science 284: 2156–9CrossRefGoogle ScholarPubMed
Kumar, P. and Foufoula-Georgiou, E. (1994). Wavelet analysis in geophysics: an introduction. In: Wavelets in Geophysics. Academic Press, London, pp. 1–43
Kvale, E. P., Archer, A. W. and Johnson, H. R. (1989). Daily, monthly, and yearly tidal cycles within laminated siltstone of the Mansfield Formation (Pennsylvanian) of Indiana. Geology 17: 365–82.3.CO;2>CrossRefGoogle Scholar
Kvale, E. P., Johnson, H. W., Sonett, C. P., Archer, A. W. and Zawistoski, A. (1999). Calculating the lunar retreat rates using tidal rhythmites. J. Sed. Res. 69: 1154–68CrossRefGoogle Scholar
Labitzke, K. and Loon, H. (1990). Associations between 11-year solar cycle, the quasi-biennial oscillation and the atmosphere: a summary of recent work. Philos. Trans. R. Soc. Lond. 330A: 577–89CrossRefGoogle Scholar
Lang, W. D., Spath, L. F., Cox, L. R. and Muir-Wood, H. M. (1928). The Belemnite Marls of Charmouth, a series in the Lower Lias of the Dorset Coast. Quart. J. Geol. Soc. 84: 179–257CrossRefGoogle Scholar
Laskar, J. (1989). A numerical experiment on the chaotic behaviour of the solar system. Nature 338: 237–8CrossRefGoogle Scholar
Laskar, J. (1999). The limits of Earth orbital calculations for geological time-scale use. Philos. Trans. R. Soc. Lond. 357: 1735–59CrossRefGoogle Scholar
Laskar, J., Joutel, F. and Robutel, P. (1993a). Stabilization of the Earth's obliquity by the Moon. Nature 361: 615–17CrossRefGoogle Scholar
Laskar, J., Joutel, F. and Boudin, F. (1993b). Orbital, precessional, and insolation quantities for the Earth from -20Myr to +10Myr. Astron. Astrophys. 270: 522–33Google Scholar
Latif, M. and Barnett, T. P. (1994). Causes of decadal climate variability over the North Pacific and North America. Science 266: 634–7CrossRefGoogle ScholarPubMed
Lau, K-M. and Sheu, P. J. (1988). Annual cycle, Quasi-Biennial Oscillation, and Southern Oscillation in global precipitation. J. Geophys. Res. 93: 10,975–88CrossRefGoogle Scholar
Lees, J. M. and Park, J. (1995). Multi-taper spectral analysis: a stand-alone C-subroutine. Comput. Geosci. 21: 199–236CrossRefGoogle Scholar
Treut, H. and Ghil, M. (1983). Orbital forcing, climatic interactions, and glaciation cycles. J. Geophys. Res. 88: 5167–90CrossRefGoogle Scholar
Liu, P. C. (1994). Wavelet spectrum analysis and ocean wind waves. In: Wavelets in Geophysics. Academic Press, London, pp. 151–66CrossRef
Lockwood, M., Stamper, R. and Wild, M. N. (1999). A doubling of the Sun's coronal magnetic field during the past 100 years. Nature 399: 437–9CrossRefGoogle Scholar
Lourens, J. L., Antonarakou, A., Hilgen, F. J., Hoof, A. A. M., Vergnaud-Grazzini, C. and Zachariasse, W. J. (1996). Evaluation of the Plio-Pleistocene astronomical timescale. Paleoceanography 11: 391–413CrossRefGoogle Scholar
Lourens, L. J., Wehausen, R. and Brumsack, H. J. (2001). Geological constraints on tidal dissipation and dynamical ellipticity of the Earth over the past three million years. Nature 409: 1029–33CrossRefGoogle ScholarPubMed
Lowrie, W. (1997). Fundamentals of Geophysics. Cambridge University Press, Cambridge, pp. 1–354
Lund, D. C. and Mix, A. C. (1998). Millennial-scale deep water oscillations: reflections of the North Atlantic in the deep Pacific from 10 to 60 ka. Paleoceanography 13: 10–19CrossRefGoogle Scholar
Lyell, C. (1830). Principles of Geology, Volume 1. John Murray, London
Lyon, J. G. (2000). The solar wind-magnetosphere-ionosphere system. Science 288: 1987–91CrossRefGoogle ScholarPubMed
MacAyeal, D. R. (1993a). A low-order model of the Heinrich event cycle. Paleoceanography 8: 767–73CrossRefGoogle Scholar
MacAyeal, D. R. (1993b). Binge/Purge oscillations of the Laurentide Ice Sheet as a cause of the North Atlantic's Heinrich events. Paleoceanography 8: 775–84CrossRefGoogle Scholar
Mallat, S. (1998). A Wavelet Tour of Signal Processing. Academic Press, London, pp. 1–577
Mann, M. E. and Bradley, R. S. (1999). Northern hemisphere temperatures during the past millennium: inferences, uncertainties and limitations. Geophys. Res. Lett. 26: 759–62CrossRefGoogle Scholar
Mann, M. E. and Lees, J. (1996). Robust estimation of background noise and signal detection in climatic time series. Clim. Change 33: 409–45CrossRefGoogle Scholar
Mann, M. E., Bradley, R. S. and Hughes, M. K. (1998). Global-scale temperature patterns and climate forcing over the past six centuries. Nature 392: 779–87CrossRefGoogle Scholar
Markson, R. and Muir, M. (1980). Solar wind control of the Earth's electric field. Science 208: 979–90CrossRefGoogle ScholarPubMed
Martin, E. E., Shackleton, N. J., Zachos, N. J. and Flower, B. P. (1999). Orbitally-tuned Sr isotope chemostratigraphy for the late middle to late Miocene. Paleoceanography 14: 74–83CrossRefGoogle Scholar
Martinson, D. G., Menke, W. and Stoffa, P. (1982). An inverse approach to signal correlation. J. Geophys. Res. 87: 4807–18CrossRefGoogle Scholar
Matsuoka, J., Kano, A., Oba, T., Watanabe, T., Sakai, S. and Seto, K. (2001). Seasonal variation of stable isotopic composition recorded in a laminated tufa, SW Japan. Earth Planet. Sci. Lett. 192: 31–44CrossRefGoogle Scholar
Mayewski, P. A., Meeker, L. D., Twickler, M. S., Whitlow, S., Yang, Q., Lyons, W. B. and Prentice, M. (1997). Major features and forcing of high-latitude northern hemispheric atmospheric circulation using a 110,000-year-long glaciochemical series. J. Geophys. Res. 102: 26,345–66CrossRefGoogle Scholar
McClellan, J. H., Parks, T. W. and Rabiner, L. R. (1973). A computer program for designing optimum FIR linear phase digital filters. IEEE Trans. Audio Electroacoustics 21: 506–26CrossRefGoogle Scholar
McIntyre, A. and Molfino, B. (1996). Forcing of Atlantic equatorial and subpolar millennial cycles by precession. Science 274: 1867–70CrossRefGoogle ScholarPubMed
McManus, J. F., Bond, G. C., Broecker, W. S., Johnsen, S., Labeyrie, L. and Higgins, S. (1994). High-resolution climate records from the North Atlantic during the last interglacial. Nature 371: 326–9CrossRefGoogle Scholar
McManus, J. F., Bond, G. C., Broecker, W. S., Fleisher, M. Q. and Higgins, S. M. (1998). Radiometrically determined fluxes in the sub-polar North Atlantic during the last 140,000 years. Earth Planet. Sci. Lett. 135: 29–43CrossRefGoogle Scholar
McManus, J. F., Oppo, D. W. and Cullen, J. L. (1999). A 0.5 million-year record of millennial-scale climate variability in the North Atlantic. Science 283: 971–5CrossRefGoogle ScholarPubMed
McPhaden, M. J. and Yu, X. (1999). Equatorial waves and the 1997–98 El Niño. Geophys. Res. Lett. 26: 2961–4CrossRefGoogle Scholar
Medio, A. (1992). Chaotic Dynamics. Theory and Applications to Economics. Cambridge University Press, Cambridge, pp. 1–344
Meeker, L. D., Mayewski, P. A., Grootes, P. M., Alley, R. B. and Bond, G. C. (2001). Comment: “On sharp spectral lines in the climate record and the millennial peak” by C. Wunsch. Paleoceanography 16: 544–7CrossRefGoogle Scholar
Meko, D. M. (1992). Spectral properties of tree-ring data in the United States Southwest as related to El Niño/Southern Oscillation. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Ed: H. F. Diaz and V. Markgraf. Cambridge University Press, Cambridge, pp. 227–41
Melnyk, D. H., Smith, D. G. and Amiri-Garroussi, K. (1994). Filtering and frequency mapping as tools in subsurface cyclostratigraphy, with examples from the Wessex Basin, UK. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 35–46CrossRef
Meyers, S. R., Sageman, B. R. and Hinnov, L. A. (2001). Integrated quantitative stratigraphy of the Cenomanian-Turonian Bridge Creek Limestone Member using evolutive harmonic analysis and stratigraphic modeling. J. Sed. Res. 71: 628–44CrossRefGoogle Scholar
Middleton, G. V., Plotnick, R. E. and Rubin, D. M. (1995). Nonlinear Dynamics and Fractals. New Numerical Techniques for Sedimentary Data. Society of Economic Paleontologists and Mineralogists (SEPM) Short Course No. 36. SEPM, Tulsa, Oklahoma, pp. 1–174CrossRef
Milankovitch, M. (1941). Canon of insolation in the Ice-Age problem. [English translation by Israel Program for scientific translation, Jerusalem 1969.]R. Serbian Acad. Spec. Publ.132Google Scholar
Miller, D. J. and Eriksson, K. A. (1997). Late Mississippian prodeltaic rhythmites in the Appalachian Basin: a hierarchical record of tidal and climatic periodicities. J. Sed. Res. 67: 653–60Google Scholar
Miller, K. G., Feigenson, M. D., Wright, J. D. and Clement, B. M. (1991). Miocene isotope reference section, Deep Sea Drilling Project Site 608: an evaluation of isotope and biostratigraphic resolution. Paleoceanography 6: 33–52CrossRefGoogle Scholar
Mitchell, J. M. (1976). An overview of climatic variability and its causal mechanisms. Quat. Res. 6: 481–93CrossRefGoogle Scholar
Mitrovica, J. X., Forte, A. M. and Pan, R. (1997). Glaciation-induced variations in the Earth's precession frequency, obliquity and insolation over the last 2.6 Ma. Geophys. J. Int. 128: 270–84CrossRefGoogle Scholar
Molinie, A. J., Ogg, J. G. and Ocean Drilling Program Leg 129 Scientific Party (1990). Sedimentation rate curves and discontinuities from sliding-window spectral analysis of logs. Log Analyst, November–December, pp. 370–4
Morgans-Bell, H. S., Coe, A. L., Hesselbo, S. P., Jenkyns, H. C., Weedon, G. P., Marshall, J. E. A. and Williams, C. J. (2001). Integrated stratigraphy of the Kimmeridge Clay Formation (Upper Jurassic) based on exposures and boreholes in South Dorset, UK. Geol. Mag. 138: 511–39CrossRefGoogle Scholar
Morley, C. K., Vanhauwaert, P. and Batist, M. (2000). Evidence for high-frequency cyclic fault activity from high-resolution seismic reflection survey, Rukwa Rift, Tanzania. J. Geol. Soc. Lond. 157: 983–94CrossRefGoogle Scholar
Mudelsee, M. and Stattegger, K. (1994). Plio/Pleistocene climate modeling based on oxygen isotope time series from deep-sea sediment cores: the Grassberger-Procaccia algorithm and chaotic systems. Math. Geol. 26: 799–815CrossRefGoogle Scholar
Muller, R. A. and Macdonald, G. J. (1995). Glacial cycles and orbital inclination. Nature 377: 107–8CrossRefGoogle Scholar
Muller, R. A. and Macdonald, G. J. (1997a). Simultaneous presence of orbital inclination and eccentricity in proxy climate records from Ocean Drilling Program Site 806. Geology 25: 3–62.3.CO;2>CrossRefGoogle Scholar
Muller, R. A. and Macdonald, G. J. (1997b). Reply to comment on: Simultaneous presence of orbital inclination and eccentricity in proxy climate records from Ocean Drilling Program Site 806, by Schulz, M. and Mudelsee, M. Geology 25: 861–22.3.CO;2>CrossRefGoogle Scholar
Munk, W., Dzieciuck, M. and Jayne, S. (2002). Millennial climate variability: is there a tidal connection?J. Clim. 15: 370–852.0.CO;2>CrossRefGoogle Scholar
Munnecke, A., Westphal, H., Elrick, M. and Reijmer, J. J. G. (2001). The mineralogical composition of precursor sediments of calcareous rhythmites — a new approach. Int. J. Earth Sci. (Geol. Rundsch.) 90: 795–812Google Scholar
Murakoshi, N., Nakayama, N. and Masuda, F. (1995). Diurnal inequality pattern of tide in the upper Pleistocene Palaeo-Tokyo Bay: reconstruction from tidal deposits and growth-lines of fossil bivalves. Spec. Pub. Int. Assoc. Sed. 24: 289–300Google Scholar
Murray, B. C., Ward, W. R. and Yeung, S. C. (1973). Periodic insolation variation on Mars. Science 180: 638–40CrossRefGoogle ScholarPubMed
Naish, T. R. and Kamp, P. J. J. (1997). Sequence stratigraphy of sixth order (41 k.y.) Pliocene-Pleistocene cyclothems, Wanganui Basin, New Zealand: a case for the regressive systems tract. Geol. Soc. Am. Bull. 109: 978–992.3.CO;2>CrossRefGoogle Scholar
Naish, T. R., Abbott, S. T., Alloway, B. V., Beu, A. G., Carter, R. M., Edwards, A. R., Journeaux, T. J., Kamp, P. J. J., Pillans, B. J., Saul, G. S. and Woolfe, K. J. (1998). Astronomical calibration of a southern hemisphere Plio-Pleistocene reference section, Wanganui Basin, New Zealand. Quat. Sci. Rev. 17: 695–710CrossRefGoogle Scholar
National Research Council (1994). Solar Influences on Global Change. National Academy Press, Washington, pp. 1–163
Neff, U., Burns, S. J., Mangini, A., Mudelsee, M., Fleitmann, D. and Matter, A. (2001). Strong coherence between solar variability and the monsoon in Oman between 9 and 6 kyr ago. Nature 411: 290–3CrossRefGoogle ScholarPubMed
Nicolis, C. and Nicolis, G. (1984). Is there a climatic attractor?Nature 311: 529–32CrossRefGoogle Scholar
Ninnemann, U., Charles, C. D. and Hodell, D. A. (1999). Origin of global millennial scale climate events: constraints from the Southern Ocean deep sea sedimentary record. In: Mechanisms of Global Climate Change at Millennial Time Scales. Geophysical Monograph 112. American Geophysical Union, Washington, pp. 99–112CrossRef
Nowroozi, A. A. (1967). Table for Fisher's test of significance in harmonic analysis. Geophys. J. R. Astron. Soc. 12: 512–20CrossRefGoogle Scholar
O'Brien, N. R. (1996). Shale lamination and sedimentary processes. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Publication 116. The Geological Society, London, pp. 23–36CrossRef
Oeschger, H. and Beer, J. (1990). The past 5000 years history of solar modulation of cosmic radiation from 10Be and 14C studies. Philos. Trans. R. Soc. Lond. 330A: 471–80CrossRefGoogle Scholar
Olsen, G. H. (1977). Modern Electronics Made Simple. Allen, London, pp. 1–306
Olsen, P. E. (1986). A 40 million year lake record of early Mesozoic orbital climatic forcing. Science 234: 842–8CrossRefGoogle ScholarPubMed
Olsen, P. E. and Kent, D. V. (1996). Milankovitch climate forcing in the tropics of Pangaea during the Late Triassic. Palaeogeog. Palaeoclimat. Palaeoecol. 122: 1–26CrossRefGoogle Scholar
Olsen, P. E. and Kent, D. V. (1999). Long-period Milankovitch cycles from the Late Triassic and Early Jurassic of eastern North America and their implications for the calibration of the Early Mesozoic time-scale and the long-term behaviour of the planets. Philos. Trans. R. Soc. Lond. 357: 1761–86CrossRefGoogle Scholar
Oost, A. P., Dehaas, H., Ijnsen, F., Vandenboogert, J. M. and Boer, P. L. (1993). The 18.6 yr nodal cycle and its impact on tidal sedimentation. Sed. Geol. 87: 1–11CrossRefGoogle Scholar
Oppo, D. W. and Lehman, S. J. (1995). Suborbital timescale variability of North Atlantic deep water during the past 200,000 years. Paleoceanography 10: 901–10CrossRefGoogle Scholar
Otnes, R. K. and Enochson, L. (1978). Applied Time Series Analysis, Volume 1. Basic Techniques. Wiley, Chichester, pp. 1–449
Paillard, D., Labeyrie, L. and Yiou, P. (1996). Macintosh program performs time-series analysis. EOS Trans. AGU 77: 379CrossRefGoogle Scholar
Pälike, H. and Shackleton, N. J. (2000). Constraints on astronomical parameters from the geological record for the past 25 Myr. Earth Planet. Sci. Lett. 182: 1–14CrossRefGoogle Scholar
Pälike, H., Shackleton, N. J. and Röhl, U. (2001). Astronomical forcing in late Eocene marine sediments. Earth Planet. Sci. Lett. 193: 589–602CrossRefGoogle Scholar
Pantev, C., Oostenveld, R., Engelien, A., Ross, B., Roberts, L. E. and Hoke, M. (1998). Increased auditory cortical representation in musicians. Nature 392: 811–14CrossRefGoogle ScholarPubMed
Pardo-Igúzquiza, E. and Rodríguez-Tovar, F. J. (2000). The permutation test as a non-parametric method for statistical significance of power spectrum estimation in cyclostratigraphic research. Earth Planet. Sci. Lett. 181: 175–89CrossRefGoogle Scholar
Pardo-Igúzquiza, E., Chica-Olmo, M. and Rodríguez-Tovar, F. J. (1994). CYSTRATI: a computer program for spectral analysis of stratigraphic successions. Comput. Geosci. 20: 511–84CrossRefGoogle Scholar
Pardo-Igúzquiza, E., Schwarzacher, W. and Rodríguez-Tovar, F. J. (2000). A library of computer programs for assisting teaching and research in cyclostratigraphic analysis. Comput. Geosci. 26: 723–40CrossRefGoogle Scholar
Park, J. and Herbert, T. D. (1987). Hunting for periodicities in a mid-Cretaceous sedimentary series. J. Geophys. Res. 92: 14,027–40CrossRefGoogle Scholar
Paul, H. A., Zachos, J. C., Flower, B. P. and Tripati, A. (2000). Orbitally induced climate and geochemical variability across the Oligocene/Miocene boundary. Paleoceanography 15: 471–85CrossRefGoogle Scholar
Pearce, R. B., Kemp, A. E. S., Baldauf, J. G. and King, S. C. (1995). High-resolution sedimentology and micropaleontology of laminated diatomaceous sediments from the eastern equatorial Pacific Ocean. Proc. Ocean Drill. Prog. Sci. Res. 138: 647–63Google Scholar
Pearn, W. C. (1964). Finding the ideal cyclothem. Kansas Geol. Surv. Bull. 169: 399–413Google Scholar
Pelletier, J. D. (1997). Analysis and modeling of the natural variability of climate. J. Clim. 10: 1331–422.0.CO;2>CrossRefGoogle Scholar
Peper, T. and Cloetingh, S. (1995). Autocyclic perturbations of orbitally forced signals in the sedimentary record. Geology 23: 937–402.3.CO;2>CrossRefGoogle Scholar
Percival, D. B. and Walden, A. T. (1993). Spectral Analysis for Physical Applications. Multitaper and Conventional Univariate Techniques. Cambridge University Press, Cambridge, pp. 1–583CrossRef
Pestiaux, P. and Berger, A. (1984a). An optimal approach to the spectral characteristics of deep sea climatic records. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 417–45
Pestiaux, P. and Berger, A. (1984b). Impacts of deep-sea processes on paleoclimatic spectra. In: Milankovitch and Climate, Eds: A. Berger, J. Imbrie, J. D. Hays, G. Kukla, B. Saltzman, NATO ASI Series C126. Reidel, Dordrecht, volume 1, pp. 493–510
Pestiaux, P., Duplessy, J. C. and Berger, A. (1987). Paleoclimatic variability at frequencies ranging from 10-4 cycle per year to 10-3 cycle per year — evidence for nonlinear behavior of the climate system. In: Climate History, Periodicity and Predictability, Eds: M. Rampino, J. E. Sanders, W. S. Newman and L. K., Kingsson. Van Nostrand Reinhold, New York, pp. 285–98
Peters, S. E. and Foote, M. (2002). Determinants of extinction in the fossil record. Nature 416: 420–4CrossRefGoogle ScholarPubMed
Petit, J. R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J.-M., Basile, I., Bender, M., Chappellez, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V. M., Legrand, M., Lipenkov, V. Y., Lorius, C., Pépin, L., Ritz, C., Saltzman, E. and Stievenard, M. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399: 429–36CrossRefGoogle Scholar
Petterson, G. (1996). Varved sediments in Sweden: a brief review. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Publication 116. The Geological Society, London, pp. 73–7CrossRef
Philander, S. G. (1990). El Niño, La Niña and the Southern Oscillation. Academic Press, London, pp. 1–293
Pilskaln, C. H. and Pike, J. (2001). Formation of Holocene sedimentary laminae in the Black Sea and the role of the benthic flocculent layer. Paleoceanography 16: 1–19CrossRefGoogle Scholar
Pisias, N. G. (1983). Geologic time series from deep-sea sediments: time scales and distortion by bioturbation. Mar. Geol. 51: 99–113CrossRefGoogle Scholar
Pisias, N. G. and Mix, A. C. (1988). Aliasing of the geologic record and the search for long-period Milankovitch cycles. Paleoceanography 3: 613–19CrossRefGoogle Scholar
Pisias, N. G. and Moore, T. C. (1981). The evolution of Pleistocene climate: a time series approach. Earth Planet. Sci. Lett. 52: 450–8CrossRefGoogle Scholar
Pisias, N. G., Mix, A. C. and Zahn, R. (1990). Nonlinear response in the global climate system: evidence from benthic oxygen isotope record in core RC13–110. Paleoceanography 5: 147–60CrossRefGoogle Scholar
Plaut, G., Ghil, M. and Vautard, R. (1995). Interannual and interdecadal variability from a long temperature time series. Science 268: 710–13CrossRefGoogle Scholar
Plotnick, R. E. (1986). A fractal model for the distribution of stratigraphic hiatuses. J. Geol. 94: 885–90CrossRefGoogle Scholar
Prell, W. L., Imbrie, J., Martinson, D. G., Morley, J. J., Pisias, N. G., Shackleton, N. J. and Streeter, H. F. (1986). Graphic correlation of oxygen isotope stratigraphy: application to the Late Quaternary. Paleoceanography 1: 137–62CrossRefGoogle Scholar
Press, W. H., Teukolsky, S. A., Vetterling, W. T. and Flannery, B. P. (1992). Numerical Recipes, the Art of Scientific Computing. Cambridge University Press, Cambridge, pp. 1–963
Preston, F. W. and Henderson, J. H. (1964). Fourier series characterization of cyclic sediments for stratigraphic correlation. In: Symposium on Cyclic Sedimentation, Ed: D. F. Merriam. Bull Kansas Geol. Surv. 169: 415–25
Priestley, M. B. (1981). Spectral Analysis and Time Series. Academic Press, London, pp. 1–890
Priestley, M. B. (1988). Non-linear and Non-stationary Time Series Analysis. Academic Press, London, pp. 1–237
Proakis, J. G. and Menolakis, D. G. (1996). Digital Signal Processing, Principles, Algorithms and Applications. Prentice Hall, London, pp. 1–968
Prokoph, A. and Agterberg, F. P. (1999). Detection of sedimentary cyclicity and stratigraphic completeness by wavelet analysis: an application of Late Albian cyclostratigraphy of the western Canada sedimentary basin. J. Sed. Res. 69: 862–75CrossRefGoogle Scholar
Prokoph, A. and Barthelmes, F. (1996). Detection of nonstationarities in geological time series: wavelet transform of chaotic and cyclic sequences. Comp. Geosci. 22: 1097–108CrossRefGoogle Scholar
Prokoph, A., Fowler, A. D. and Patterson, R. T. (2000). Evidence for periodicity and nonlinearity in a high-resolution fossil record of long-term evolution. Geology 28: 867–702.0.CO;2>CrossRefGoogle Scholar
Pugh, D. T. (1987). Tides, Surges and Mean Sea-level. Wiley, Chichester, pp. 1–472
Qin, X., Tan, M., Liu, T., Wang, X., Li, T. and Lu, J. (1999). Spectral analysis of 1000-year stalagmite lamina-thickness record from Shihua Cavern, Beijing, China, and its climatic significance. The Holocene 9: 689–94CrossRefGoogle Scholar
Quinn, T. M., Taylor, F. W. and Crowley, T. J. (1993). A 173 year stable isotope record from a tropical south Pacific coral. Quat. Sci. Rev. 12: 407–12CrossRefGoogle Scholar
Quinn, T. M., Crowley, T. J., Taylor, F. W., Henin, C., Joannot, P. and Join, Y. (1998). A multicentury stable isotope record from a New Caledonia coral: interannual and decadal sea surface temperature variability in the southwest Pacific since 1657 A.D. Paleoceanography 13: 412–26CrossRefGoogle Scholar
Quinn, W. H. (1992). A study of Southern Oscillation-related climatic activity for AD622–1990 incorporating Nile River flood data. In: El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation, Eds: H. F. Diaz and V. Markgraf. Cambridge University Press, Cambridge, pp. 119–49
Ramsay, A. T. S., Sykes, T. J. S. and Kidd, R. B. (1994a). Sedimentary hiatuses as indicators of fluctuating oceanic water masses: a new model. J. Geol. Soc. Lond. 151: 737–40CrossRefGoogle Scholar
Ramsay, A. T. S., Sykes, T. J. S. and Kidd, R. B. (1994b). Waxing (and waning) lyrical on hiatuses: Eocene-Quaternary Indian Ocean hiatuses as proxy indicators of water mass production. Paleoceanography 9: 857–77CrossRefGoogle Scholar
Räsänen, M. E., Linna, A. M., Santos, J. C. R. and Negri, F. R. (1995). Late Miocene tidal deposits in the Amazonian Foreland Basin. Science 269: 386–9CrossRefGoogle ScholarPubMed
Rast, M. P., Fox, P. A., Lin, H., Lites, B. W., Meisner, R. W. and White, O. R. (1999). Bright rings around sunspots. Nature 401: 678–9CrossRefGoogle Scholar
Raup, D. M. and Sepkoski, J. J. (1988). Testing for periodicity of extinction. Science 241: 94–6CrossRefGoogle ScholarPubMed
Raymo, M. E., Ruddiman, W. F., Backman, J., Clement, B. M. and Martinson, D. G. (1989). Late Pliocene variations in northern hemisphere ice sheets and North Atlantic deep-water circulation. Paleoceanography 4: 413–46CrossRefGoogle Scholar
Raymo, M. E., Ganley, K., Carter, S., Oppo, D. W. and McManus, J. (1998). Millennial-scale climate instability during the early Pleistocene epoch. Nature 392: 699–702CrossRefGoogle Scholar
Reijmer, J. J. G., ten Kate, W. G. H. Z., Sprenger, A. and Schlager, W. (1991). Calciturbidite composition related to exposure and flooding of a carbonate platform (Triassic, Eastern Alps). Sedimentology 38: 1059–74CrossRefGoogle Scholar
Rempel, A. W., Waddington, E. D., Wettlaufer, J. S. and Worster, M. G. (2001). Possible displacement of the climate signal in ancient ice by premelting and anomalous diffusion. Nature 411: 568–71CrossRefGoogle ScholarPubMed
Rial, J. A. (1999). Pacemaking the ice ages by frequency modulation of Earth's orbital eccentricity. Science 285: 564–8CrossRefGoogle ScholarPubMed
Richards, G. R. (1994). Orbital forcing and endogenous interactions: non-linearity, persistence and convergence in late Pleistocene climate. Quat. Sci. Rev. 13: 709–25CrossRefGoogle Scholar
Ricken, W. (1986). Diagenetic Bedding: A Model for Marl-Limestone Alternations. Lecture Notes Earth Science 6. Springer, Berlin, pp. 1–210
Ricken, W. (1991a). Time span assessment — an overview. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 773–94
Ricken, W. (1991b). Variation of sedimentation rates in rhythmically bedded sediments. Distinction between deposition types. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 167–87
Ricken, W. (1993). Sedimentation as a Three-Component System. Springer, London
Ricken, W. and Eder, W. (1991). Diagenetic modification of calcareous beds — an overview. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 430–49
Ridgwell, A. J., Watson, A. J. and Raymo, M. E. (1999). Is the spectral signature of the 100 kyr glacial cycle consistent with a Milankovitch origin?Paleoceanography 14: 437–40CrossRefGoogle Scholar
Riegel, W. (1991). Coal cyclothems and some models for their origin. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 733–50
Ripepe, M. and Fischer, A. G. (1991). Stratigraphic rhythms synthesized from orbital variations. In: Sedimentary Modeling: Computer Simulations and Methods for Improved Parameter Definition, Eds: K. Franseen, W. L. Watney, C. G. St. C. Kendall and W. Ross. Kansas State Geol. Surv. Bull. 233: 335–44
Ripepe, M., Roberts, L. T. and Fischer, A. G. (1991). ENSO and sunspot cycles in varved Eocene oil shales from image analysis. J. Sed. Petrol. 61: 1155–63Google Scholar
Rittenour, T. M., Brigham-Grette, J. and Mann, M. E. (2000). El Niño-like climate teleconnections in New England during the Late Pleistocene. Science 288: 1039–42CrossRefGoogle ScholarPubMed
Robock, A. (1996). Stratigraphic control of climate. Science 272: 972–3CrossRefGoogle Scholar
Rodbell, D. T., Seltzer, G. O., Anderson, D. M., Abbott, M. B., Enfield, D. B. and Newman, J. H. (1999). An ∼15,000-year record of El Niño-driven alluviation in southwestern Ecuador. Science 283: 516–20CrossRefGoogle ScholarPubMed
Rodwell, M. J., Rowell, D. P. and Folland, C. K. (1999). Oceanic forcing of the wintertime North Atlantic Oscillation and European climate. Nature 398: 320–3CrossRefGoogle Scholar
Roulier, L. M. and Quinn, T. M. (1995). Seasonal- and decadal-scale climatic variability in southwest Florida during the middle Pliocene: inferences from a coralline stable isotope record. Paleoceanography 10: 429–43CrossRefGoogle Scholar
Rubincam, D. P. (1995). Has climate changed the Earth's tilt?Paleoceanography 10: 365–72CrossRefGoogle Scholar
Ruddiman, W. F. (1977). Late Quaternary deposition of ice-rafted sand in the subpolar North Atlantic (latitude 40°N to 65°N). Geol. Soc. Am. Bull. 88: 1813–272.0.CO;2>CrossRefGoogle Scholar
Ruddiman, W. F. (1985). Climate studies in ocean cores. In: Paleoclimate Analysis and Modelling, Ed: A. D. Hecht. Kluwer, The Netherlands, pp. 197–257
Ruddiman, W. F. and McIntyre, A. (1984). Ice-age thermal response and climatic role of the surface Atlantic Ocean, 40°N to 63°N. Geol. Soc. Am. Bull. 95: 381–962.0.CO;2>CrossRefGoogle Scholar
Ruddiman, W. F., Raymo, M. and McIntyre, A. (1986). Matayama 41,000-year cycles, North Atlantic Ocean and northern hemisphere ice sheets. Earth Planet. Sci. Lett. 80: 117–29CrossRefGoogle Scholar
Ruddiman, W. F., Raymo, M. E., Martinson, D. G., Clement, B. M. and Backman, J. (1989). Pleistocene evolution: northern hemisphere ice sheets and North Atlantic Ocean. Paleoceanography 4: 353–412CrossRefGoogle Scholar
Rutherford, S. and D'Hondt, S. (2000). Early onset and tropical forcing of 100,000-year Pleistocene glacial cycles. Nature 408: 72–5CrossRefGoogle ScholarPubMed
Sadler, P. M. (1981). Sediment accumulation rates and the completeness of stratigraphic sections. J. Geol. 89: 569–84CrossRefGoogle Scholar
Sadler, P. M. and Strauss, D. J. (1990). Estimation of completeness of stratigraphical sections using empirical data and theoretical models. J. Geol. Soc. Lond. 147: 471–85CrossRefGoogle Scholar
Saji, N. H., Goswami, B. N., Vinayachandran, P. N. and Yamagata, T. (1999). A dipole mode in the tropical Indian Ocean. Nature 401: 360–3CrossRefGoogle ScholarPubMed
Saltzman, B. and Verbitsky, M. (1994). Late Pleistocene climatic trajectory in the phase space of global ice, ocean state, and CO2: observations and theory. Paleoceanography 9: 767–79CrossRefGoogle Scholar
Sander, B. (1936). Beiträge zur Kenntnis der Anlargerungsgefüge. Mineral. Petrogr. Mitt. 48: 27–139Google Scholar
Schaaf, M. and Thurow, J. (1997). Tracing short cycles in long records: the study of inter-annual to inter-centennial climate change from long sediment records, examples from the Santa Barbara Basin. J. Geol. Soc. Lond. 154: 613–22CrossRefGoogle Scholar
Schiffelbein, P. (1984). Effect of benthic mixing on the information content of deep sea stratigraphical signals. Nature 311: 651–3CrossRefGoogle Scholar
Schiffelbein, P. and Dorman, L. (1986). Spectral effects of time-depth nonlinearities in deep sea sediment records: a demodulation technique for realigning time and depth scales. J. Geophys. Res. 91: 3821–35CrossRefGoogle Scholar
Schlesinger, M. E. and Ramankutty, N. (1994). An oscillation in the global climate system of period 65–70 years. Nature 367: 723–6CrossRefGoogle Scholar
Schmitz, W. J. and McCarthey, M. S. (1993). On the North Atlantic circulation. Rev. Geophys. 31: 29–49CrossRefGoogle Scholar
Scholz, C. H. (1998). Earthquakes and friction laws. Nature 391: 37–42CrossRefGoogle Scholar
Schulz, H., Rad, U. and Erlenkeuser, H. (1998). Correlation between Arabian Sea and Greenland climate oscillations of the past 110,000 years. Nature 393: 54–7CrossRefGoogle Scholar
Schulz, M. and Stattegger, K. (1997). Spectrum: spectral analysis of unevenly spaced paleoclimatic time series. Comput. Geosci. 23: 929–45CrossRefGoogle Scholar
Schuster, A. (1898). On the investigation of hidden periodicities with application to a supposed 26-day period of meteorological phenomenon. Terr. Mag. Atmos. Elect. 3: 13–41CrossRefGoogle Scholar
Schwarzacher, W. (1964). An application of statistical time-series analysis of a limestone-shale sequence. J. Geol. 72: 195–213CrossRefGoogle Scholar
Schwarzacher, W. (1975). Sedimentation Models and Quantitative Stratigraphy. Elsevier, Amsterdam, pp. 1–382
Schwarzacher, W. (1991). Milankovitch cycles and the measurement of time. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken, and A. Seilacher. Springer, London, pp. 855–63
Schwarzacher, W. (1993). Cyclostratigraphy and the Milankovitch Theory. Elsevier, Amsterdam, pp. 1–225
Schwarzacher, W. (1998). Stratigraphic resolution, cycles and sequences. In: Sequence Stratigraphy — Concepts and Applications, Eds: F. M. Gradstein, K. O. Sandvik and N. J. Milton. Norwegian Petroleum Society Special Publication No. 8. Elsevier, Amsterdam, pp. 1–8
Schwarzacher, W. and Fischer, A. G. (1982). Limestone-shale bedding and perturbations of the Earth's orbit. In: Cyclic and Event Stratification, Eds: G. Einsele and A. Seilacher. Springer, London, pp. 72–95CrossRef
Schweingruber, F. H., Echstein, D., Serre-Bachet, F. and Braeker, O. U. (1990). Identification, presentation and interpretation of event years and pointer years in dendrochronology. Dendrochronologia 8: 9–38Google Scholar
Seidov, D. and Maslin, M. (1999). North Atlantic deep water circulation collapse during Heinrich events. Geology 27: 23–62.3.CO;2>CrossRefGoogle Scholar
Sellwood, B. W. (1970). The relation of trace fossils to small scale sedimentary cycles in the British Lias. In: Trace Fossils, Eds: T. P. Crimes and J. C. Harper. Seel House Press, Liverpool, pp. 489–504
Sepkoski, J. J. (1989). Periodicity in extinction and the problem of catastrophism in the history of life. J. Geol. Soc. 146: 7–19CrossRefGoogle ScholarPubMed
Sepkoski, J. J. and Raup, D. M. (1986). Periodicity in marine extinction events. In: Dynamics of Extinction, Ed: D. K. Elliott. Wiley, Chichester, pp. 3–36
Shackleton, N. J. (2000). The 100,000-year ice-age cycle identified and found to lag temperature, carbon dioxide, and orbital eccentricity. Science 289: 1897–902CrossRefGoogle ScholarPubMed
Shackleton, N. J. and Crowhurst, S. (1997). Sediment fluxes based on an orbitally tuned time scale 5 Ma to 14 Ma, Site 926. Proc. Ocean Drill. Prog. 154: 69–82Google Scholar
Shackleton, N. J. and Imbrie, J. (1990). The δ18O spectrum of oceanic deep water over a five-decade band. Clim. Change 16: 217–30CrossRefGoogle Scholar
Shackleton, N. J. and Opdyke, N. D. (1973). Oxygen isotope and palaeomagnetic stratigraphy of equatorial Pacific core V28-238: oxygen isotope temperatures and ice volumes on a 105 year and 106 year scale. Quat. Res. 3: 39–55CrossRefGoogle Scholar
Shackleton, N. J. and Pisias, N. G. (1985). Atmospheric carbon dioxide, orbital forcing and climate. In: The Carbon Cycle and Atmospheric CO2: Natural Variations Archaean to Present, Eds: E. T. Sundquist and W. S. Broecker. Geophysical Monograph 32. American Geophysical Union, Washington, pp. 303–17CrossRef
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. R. Soc. Edinburgh, Earth Sci. 81: 251–61CrossRefGoogle Scholar
Shackleton, N. J., Crowhurst, S. J., Hagelberg, T., Pisias, N. G. and Schneider, D. A. (1995a). A new late Neogene timescale: application to Leg 138 sites. Proc. Ocean Drill. Prog. Sci. Res. 138: 73–101Google Scholar
Shackleton, N. J., Hagelberg, T. K. and Crowhurst, S. J. (1995b). Evaluating the success of astronomical tuning: pitfalls of using coherence as a criterion for assessing pre-Pleistocene timescales. Paleoceanography 10: 693–7CrossRefGoogle Scholar
Shackleton, N. J., Hall, M. A. and Pate, D. (1995c). Pliocene isotope stratigraphy of Site 846. Proc. Ocean Drill. Prog. Sci. Res. 138: 337–55Google Scholar
Shackleton, N. J., Crowhurst, S. J., Weedon, G. P. and Laskar, J. (1999a). Astronomical calibration of Oligocene-Miocene time. Philos. Trans. R. Soc. Lond. 357: 1907–29CrossRefGoogle Scholar
Shackleton, N. J., McCave, I. N. and Weedon, G. P. (Eds) (1999b). Astronomical (Milankovitch) calibration of the geological time-scale. Philos. Trans. R. Soc. Lond. 357: 1733–2007CrossRefGoogle Scholar
Shen, G. T., Cole, J. E., Lea, D. W., Linn, L. J., McConnaughey, E. A. and Fairbanks, R. G. (1992). Surface ocean variability at Galapagos from 1936–1982: calibration of geochemical tracers in corals. Paleoceanography 7: 563–88CrossRefGoogle Scholar
Shindell, D., Rind, D., Balachandran, N., Lean, J. and Lonergan, P. (1999). Solar cycle variability, ozone and climate. Science 284: 305–8CrossRefGoogle ScholarPubMed
Shindell, D. T., Schmidt, G. A., Mann, M. E., Rind, D. and Waple, A. (2001). Solar forcing of regional climate during the Maunder Minimum. Science 294: 2149–52CrossRefGoogle ScholarPubMed
Short, D. A., Mengel, J. G., Crowley, T. J., Hyde, W. T. and North, G. R. (1991). Filtering of Milankovitch cycles by Earth's geography. Quat. Res. 35: 157–73CrossRefGoogle Scholar
Sigman, D. M. and Boyle, E. A. (2000). Glacial/interglacial variations in atmospheric carbon dioxide. Nature 407: 859–69CrossRefGoogle ScholarPubMed
Sloan, L. C. and Huber, M. (2001). Eocene oceanic responses to orbital forcing on precessional time scales. Paleoceanography 16: 101–11CrossRefGoogle Scholar
Smith, D. G. (1994). Cyclicity or chaos? Orbital forcing versus non-linear dynamics. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 531–44CrossRef
Smith, N. D., Phillips, A. C. and Powell, R. D. (1990). Tidal drawdown: a mechanism for producing cyclic sediment laminations in glaciomarine deltas. Geology 18: 10–132.3.CO;2>CrossRefGoogle Scholar
Solanki, S. K., Schüssler, M. and Fligge, M. (2000). Evolution of the Sun's large-scale magnetic field since the Maunder minimum. Nature 408: 445–7CrossRefGoogle ScholarPubMed
Sonett, C. P. and Chan, M. A. (1998). Neoproterozoic Earth-Moon dynamics: rework of the 900Ma Big Cottonwood Canyon tidal laminae. Geophys. Res. Lett. 25: 539–42CrossRefGoogle Scholar
Sonett, C. P. and Finney, S. A. (1990). The spectrum of radiocarbon. Philos. Trans. R. Soc. Lond. 330A: 413–26CrossRefGoogle Scholar
Sonett, C. P. and Williams, G. E. (1985). Solar periodicities expressed in varves from glacial Skilak Lake, southern Alaska. J. Geophys. Res. 90: 12,019–26CrossRefGoogle Scholar
Sonett, C. P., Finney, S. A. and Williams, C. P. (1988). The lunar orbit in the late Precambrian and the Elatina sandstone laminae. Nature 335: 806–8CrossRefGoogle Scholar
Sonett, C. P., Kvale, E. P., Zakharian, A., Chan, M. J. and Demko, T. M. (1996). Late Proterozoic and Paleozoic tides, retreat of the Moon, and rotation of the Earth. Science 273: 100–4CrossRefGoogle Scholar
Spencer-Cervato, C. (1998). Changing depth distribution of hiatuses during the Cenozoic. Paleoceanography 13: 178–82CrossRefGoogle Scholar
Stauffer, B., Blunier, T., Dällenbach, A., Indermühle, A., Scwander, J., Stocker, T. F., Tschumi, J., Chappellaz, J., Raynaud, D., Hammer, C. U. and Clausen, H. B. (1998). Atmospheric CO2 concentration and millennial-scale climate change during the last glacial period. Nature 392: 59–62CrossRefGoogle Scholar
Stewart, I. (1990). Does God Play Dice? The New Mathematics of Chaos. Penguin, London, pp. 1–317
Stigler, S. M. and Wagner, M. J. (1987). A substantial bias in nonparametric tests for periodicity in geophysical data. Science 238: 940–5CrossRefGoogle ScholarPubMed
Stigler, S. M. and Wagner, M. J. (1988). Testing for periodicity of extinction. Science 241: 96–9CrossRefGoogle ScholarPubMed
Stockton, C. W., Boggess, W. R. and Meko, D. M. (1985). Climate and tree rings. In: Paleoclimate Analysis and Modelling, Ed: A. D. Hecht. Wiley, Chichester, pp. 71–161
Strauss, D. and Sadler, P. M. (1989). Stochastic models for the completeness of stratigraphic sections. J. Int. Assoc. Math. Geol. 21: 37–59CrossRefGoogle Scholar
Street-Perrott, F. and Perrott, R. A. (1990). Abrupt fluctuations in the tropics: the influence of Atlantic Ocean circulation. Nature 343: 607–12CrossRefGoogle Scholar
Stuiver, M. and Braziunas, T. F. (1989). Atmospheric 14C and century-scale solar oscillations. Nature 338: 405–7CrossRefGoogle Scholar
Stuiver, M. and Quay, P. D. (1980). Changes in atmospheric carbon-14 attributed to a variable sun. Science 207: 11–19CrossRefGoogle ScholarPubMed
Stuiver, M., Grootes, P. M. and Braziunas, T. F. (1995). The GISP2 δ18O climate record of the past 16,500 years and the role of the Sun, oceans and volcanoes. Quat. Res. 44: 341–54CrossRefGoogle Scholar
Stuiver, M., Braziunas, T. F., Grootes, P. M. and Zielinski, G. A. (1997). Is there evidence for solar forcing of climate in the GISP2 oxygen isotope record?Quat. Res. 48: 259–66CrossRefGoogle Scholar
Sugihara, G. and May, R. M. (1990). Nonlinear forecasting as a way of distinguishing chaos from measurement error in time series. Nature 344: 734–41CrossRefGoogle ScholarPubMed
Sutton, R. T. and Allen, M. R. (1997). Decadal predictability of North Atlantic sea surface temperature and climate. Nature 388: 563–7CrossRefGoogle Scholar
Sztanó, O. and Boer, P. L. (1995). Basin dimensions and morphology as controls on amplification of tidal motions (the Early Miocene North Hungarian Bay). Sedimentology 42: 665–82CrossRefGoogle Scholar
Taner, M. T., Koehler, F. and Sheriff, R. E. (1979). Complex seismic trace analysis. Geophysics 44: 1041–63CrossRefGoogle Scholar
Taylor, A. H., Jordan, M. B. and Stephens, J. A. (1998). Gulf Stream shifts following ENSO events. Nature 393: 638CrossRefGoogle Scholar
Taylor, C. A. (1965). Physics of Musical Sounds. English University Press, Aylesbury, pp. 1–196
Taylor, C. A. (1976). Sounds of Music. BBC, London, pp. 1–183
Taylor, K. C., Lamorey, G. W., Doyle, G. A., Alley, R. B., Grootes, P. M., Mayewski, P. A., White, J. W. C. and Barlow, L. K. (1993). The “flickering switch” of late Pleistocene climate change. Nature 361: 432–6CrossRefGoogle Scholar
Tessier, B. and Gigot, P. (1989). A vertical record of different tidal cyclicities: an example from the Miocene marine molasse of Digne (Haute Provence, France). Sedimentology 36: 767–76CrossRefGoogle Scholar
Thomson, D. J. (1982). Spectrum estimation and harmonic analysis. Proc. IEEE 70: 1055–96CrossRefGoogle Scholar
Thomson, D. J. (1990). Quadratic-inverse spectrum estimates; applications to paleoclimatology. Philos. Trans. R. Soc. Lond. 332A: 539–97CrossRefGoogle Scholar
Thomson, D. J. (1995). The seasons, global temperature, and precession. Science 268: 59–68CrossRefGoogle ScholarPubMed
Thomson, J., Higgs, N. C. and Clayton, T. (1995). A geochemical criterion for the recognition of Heinrich events and estimation of their depositional fluxes by the 230Thexcess profiling method. Earth Planet. Sci. Lett. 135: 29–43CrossRefGoogle Scholar
Thunell, R., Pride, C., Tappa, E. and Muller-Karger, F. (1993). Varve formation in the Gulf of California: insights from time series sediment trap sampling and remote sensing. Quat. Sci. Rev. 12: 451–64CrossRefGoogle Scholar
Thunell, R. C., Tappa, E. and Anderson, D. M. (1995). Sediment fluxes and varve formation in Santa Barbara Basin, offshore California. Geology 23: 1083–62.3.CO;2>CrossRefGoogle Scholar
Tiedemann, R. and Franz, S. O. (1997). Deep-water circulation, chemistry, and terrigenous sediment supply in the equatorial Atlantic during the Pliocene, 3.3–2.6 Ma and 5–4.5 Ma. Proc. Ocean Drill. Prog. Sci. Res. 154: 299–318Google Scholar
Tinsley, B. A. (1996). Correlations of atmospheric dynamics with solar wind induced changes of air-Earth current density into cloud tops. J. Geophys. Res. 101: 29,701–14CrossRefGoogle Scholar
Tinsley, B. A. (1997). Do effects of global atmospheric electricity on clouds cause climate changes?EOS Trans. Am. Geophys. Union 78: 341–9CrossRefGoogle Scholar
Tipper, J. C. (1983). Rates of sedimentation, and stratigraphical completeness. Nature 302: 696–8CrossRefGoogle Scholar
Tong, H. (1990). Non-linear Time Series. Oxford University Press, Oxford, pp. 1–564
Toon, O. B., Pollack, J. B., Ward, W., Burns, J. A. and Bilski, K. (1980). The astronomical theory of climate change on Mars. Icarus 44: 552–607CrossRefGoogle Scholar
Torrence, C. and Compo, G. P. (1998). A practical guide to wavelet analysis. Bull. Am. Meteorol. Soc. 79: 61–782.0.CO;2>CrossRefGoogle Scholar
Trauth, M. H. (1998). TURBO: a dynamic-probabilistic simulation to study the effects of bioturbation on paleoceanographic time series. Comput. Geosci. 24: 433–41CrossRefGoogle Scholar
Trauth, M. H., Sarnthein, M. and Arnold, M. (1997). Bioturbational mixing depth and carbon flux at the seafloor. Paleoceanography 12: 517–26CrossRefGoogle Scholar
Tsonis, A. A. and Elsner, J. B. (1992). Nonlinear prediction as a way of distinguishing chaos from random fractal sequences. Nature 358: 217–20CrossRefGoogle Scholar
Tudhope, A. W., Shimmield, G. B., Chilcott, C. P., Jebb, M., Fallick, A. E. and Dalgleish, A. N. (1995). Recent changes in climate in the far western equatorial Pacific and their relationship to the Southern Oscillation; oxygen isotope records from massive corals, Papua New Guinea. Earth Planet. Sci. Lett. 136: 575–90CrossRefGoogle Scholar
Turcotte, D. L. (1997). Fractals and Chaos in Geology and Geophysics, 2nd Edition. Cambridge University Press, Cambridge, pp. 1–398
Tziperman, E., Stone, L., Cane, M. A. and Jarosh, H. (1994). El Niño chaos: overlapping of resonances between the seasonal cycle and the Pacific ocean-atmosphere oscillator. Science 264: 72–4CrossRefGoogle ScholarPubMed
Ulrych, T. J. and Bishop, T. N. (1975). Maximum entropy spectral analysis and autoregressive decomposition. Rev. Geophys. Space Phys. 13: 183–200CrossRefGoogle Scholar
Urban, F. E., Cole, J. E. and Overpeck, J. T. (2000). Influence of mean climate change on climate variability from a 155-year tropical Pacific coral record. Nature 407: 989–93Google ScholarPubMed
Vail, P. R., Mitchum, R. M., Todd, R. G., Widmer, J. W., Thompson, S., Sangree, J. B., Bubb, J. N. and Hatlelid, W. G. (1977). Seismic stratigraphy and global changes of sealevel. In: Seismic Stratigraphy — Application to Hydrocarbon Exploration, Ed: C. E. Payton. Am. Assoc. Petrol. Geol. Mem. 26: 46–212
Vail, P. R., Audemard, F., Bowman, S. A., Eisner, P. N. and Perez-Cruz, C. (1991). The stratigraphic signatures of tectonics, eustasy and sedimentology — an overview. In: Cycles and Events in Stratigraphy, Eds: G. Einsele, W. Ricken and A. Seilacher. Springer, London, pp. 617–59
Van Echelpoel, E. (1994). Identification of regular sedimentary cycles using Walsh spectral analysis with results from the Boom Clay Formation, Belgium. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer, and D. G. Smith. International Association of Sedimentologists Special Publication 19. Blackwell, Oxford, pp. 63–74CrossRef
Geel, B., Raspopov, O. M., Renssen, H., Plicht, J., Dergachev, V. A. and Meijer, H. A. J. (1999). The role of solar forcing upon climate change. Quat. Sci. Rev. 18: 331–8CrossRefGoogle Scholar
Vautard, R. and Ghil., M. (1989). Singular spectrum analysis in non-linear dynamics, with applications to palaeoclimatic time series. Physica D 35: 395–424CrossRefGoogle Scholar
Visser, M. J. (1980). Neap-Spring cycles reflected in Holocene subtidal large-scale bedform deposits: a preliminary note. Geology 8: 543–62.0.CO;2>CrossRefGoogle Scholar
Wales, D. J. (1991). Calculating the rate of loss of information from chaotic time series by forecasting. Nature 350: 485–8CrossRefGoogle Scholar
Walther, G. (1997). Absence of correlation between the solar neutrino flux and the sunspot number. Phys. Rev. Lett. 79: 4522–4CrossRefGoogle Scholar
Wanless, H. R. and Weller, J. M. (1932). Correlation and extent of Pennsylvanian cyclothems. Geol. Soc. Am. Bull. 43: 1003–16CrossRefGoogle Scholar
Webster, P. J. and Palmer, T. N. (1997). The past and the future of El Niño. Nature 390: 562–4CrossRefGoogle Scholar
Webster, P. J., Moore, A. M., Loschnigg, J. P. and Leben, R. R. (1999). Coupled ocean-atmosphere dynamics in the Indian Ocean during 1997–98. Nature 401: 356–60CrossRefGoogle ScholarPubMed
Weedon, G. P. (1989). The detection and illustration of regular sedimentary cycles using Walsh power spectra and filtering, with examples from the Lias of Switzerland. J. Geol. Soc. Lond. 146: 133–44CrossRefGoogle Scholar
Weedon, G. P. (1993). The recognition and stratigraphic implications of orbital forcing of climate and sedimentary cycles. In: Sedimentology Review, Ed: V. P. Wright. Blackwell, Oxford, pp. 31–50CrossRef
Weedon, G. P. and Jenkyns, H. C. (1990). Regular and irregular climatic cycles and the Belemnite Marls (Pliensbachian, Lower Jurassic, Wessex Basin). J. Geol. Soc. Lond. 147: 915–18CrossRefGoogle Scholar
Weedon, G. P. and Jenkyns, H. C. (1999). Cyclostratigraphy and the Early Jurassic timescale: data from the Belemnite Marls, Dorset, southern England. Geol. Soc. Am. Bull. 111: 1823–402.3.CO;2>CrossRefGoogle Scholar
Weedon, G. P. and Read, W. A. (1995). Orbital-climatic forcing of Namurian cyclic sedimentation from spectral analysis of the Limestone Coal Group, Central Scotland. In: Orbital Forcing Timescales and Cyclostratigraphy, Eds: M. R. House and A. S. Gale. Geological Society Special Publication No. 85. The Geological Society, London, pp. 51–66CrossRef
Weedon, G. P. and Shimmield, G. B. (1991). Late Pleistocene upwelling and productivity variations in the northwest Indian Ocean deduced from spectral analyses of geochemical data from sites 722 and 724. Proc. Ocean Drill. Program Sci. Res. 117: 431–43Google Scholar
Weedon, G. P., Shackleton, N. J. and Pearson, P. N. (1997). The Oligocene timescale and cyclostratigraphy on the Ceara Rise, western equatorial Atlantic. Proc. Ocean Drill. Prog. Sci. Res. 154: 101–14Google Scholar
Weedon, G. P., Jenkyns, H. C., Coe, A. L. and Hesselbo, S. P. (1999). Astronomical calibration of the Jurassic time-scale from cyclostratigraphy in British mudrock formations. Philos. Trans. R. Soc. Lond. 357: 1787–813CrossRefGoogle Scholar
Weiss, N. O. (1990). Periodicity and aperiodicity in solar magnetic activity. Philos. Trans. R. Soc. Lond. 330A: 617–25CrossRefGoogle Scholar
Wells, J. W. (1963). Coral growth and geochronometry. Nature 197: 948–50CrossRefGoogle Scholar
Weltje, G. and Boer, P. L. (1993). Astronomically induced paleoclimatic oscillations reflected in Pliocene turbidite deposits on Corfu (Greece): implications for the interpretation of higher order cyclicity in ancient turbidite systems. Geology 21: 307–102.3.CO;2>CrossRefGoogle Scholar
Whitcombe, L. J. (1996). A FORTRAN program to calculate tidal heights using the simplified harmonic method of tidal prediction. Comp. Geosci. 22: 817–21CrossRefGoogle Scholar
White, J. C. W., Barlow, L. K., Fisher, D., Grootes, P., Jouzel, J., Johnsen, S. J., Stuiver, M. and Clausen, H. (1997). The climate signal in the stable isotopes of snow from Summit, Greenland: results of comparisons with modern climate observations. J. Geophys. Res. 102: 26,425–39CrossRefGoogle Scholar
Wiesenfield, K. and Moss, F. (1995). Stochastic resonance and the benefits of noise: from ice ages to crayfish and SQUIDS. Nature 373: 33–6CrossRefGoogle Scholar
Wilkinson, B. H., Drummond, C. N., Rothman, E. D. and Diedrich, N. W. (1997). Stratal order in peritidal carbonate sequences. J. Sed. Res. 67: 1068–82Google Scholar
Wilkinson, B. H., Drummond, C. N., Diedrich, N. W. and Rothman, E. D. (1999). Poisson processes of carbonate accumulation on Paleozoic and Holocene platforms. J. Sed. Res. 69: 338–50CrossRefGoogle Scholar
Williams, G. E. (1981). Sunspot periods in the late Precambrian glacial climate and solar-planetary relations. Nature 291: 624–8CrossRefGoogle Scholar
Williams, G. E. (1986). The solar cycle in Precambrian time. Sci. Am. 255: 88–95CrossRefGoogle Scholar
Williams, G. E. (1988). Cyclicity in the Late Precambrian Elatina Formation, South Australia: solar or tidal signature?Clim. Change 13: 117–28CrossRefGoogle Scholar
Williams, G. E. (1989). Late Precambrian tidal rhythmites in South Australia and the history of the Earth's rotation. J. Geol. Soc. Lond. 146: 97–111CrossRefGoogle Scholar
Williams, G. E. (1991). Milankovitch-band cyclicity in bedded halite deposits contemporaneous with Late Ordovician-Early Silurian glaciation, Canning Basin, western Australia. Earth Planet. Sci. Lett. 103: 143–55CrossRefGoogle Scholar
Williams, G. E. and Sonett, C. P. (1985). Solar signature in sedimentary cycles from the late Precambrian Elatina Formation, Australia. Nature 318: 523–7CrossRefGoogle Scholar
Williams, G. P. (1997). Chaos Theory Tamed. National Academy Press, Washington, pp. 1–499
Williams, R. B. G. (1984). Introduction to Statistics for Geographers and Earth Scientists. Macmillan, London, pp. 1–349CrossRef
Willson, R. C. and Hudson, H. S. (1988). Solar luminosity variations in solar cycle 21. Nature 332: 810–12CrossRefGoogle Scholar
Willson, R. C. and Hudson, H. S. (1991). The Sun's luminosity over a complete solar cycle. Nature 351: 42–4CrossRefGoogle Scholar
Wilson, D. S. (1993). Confirmation of the astronomical calibration of the magnetic polarity timescale from sea-floor spreading rates. Nature 364: 788–90CrossRefGoogle Scholar
Worthington, P. F. (1990). Sediment cyclicity from well logs. In: Geological Applications of Wireline Logs, Eds: A. Hurst, M. A. Lovell and A. C. Morton. Geological Society Special Publication No. 48. The Geological Society, London, pp. 123–32CrossRef
Wunsch, C. (1999). The interpretation of short climate records, with comments on the North Atlantic and Southern Oscillations. Bull. Am. Meteorol. Soc. 80: 245–552.0.CO;2>CrossRefGoogle Scholar
Wunsch, C. (2000a). On sharp spectral lines in the climate record and the millennial peak. Paleoceanography 15: 417–24CrossRefGoogle Scholar
Wunsch, C. (2000b). Moon, tides and climate. Nature 405: 743–4CrossRefGoogle Scholar
Yang, C-S. and Baumfalk, Y. A. (1994). Milankovitch cyclicity in the Upper Rotliegend Group of the Netherlands offshore. In: Orbital Forcing and Cyclic Sequences, Eds: P. L. de Boer and D. G. Smith. International Association of Sedimentologists Special Publication No. 19. Blackwell, Oxford, pp. 47–61CrossRef
Yang, C-S. and Nio, S-D. (1985). The estimation of palaeohydrodynamic processes from subtidal deposits using time series analysis methods. Sedimentology 32: 41–57CrossRefGoogle Scholar
Yiou, P., Baert, E. and Loutre, M. F. (1996). Spectral analysis of climate data. Surve. Geophys. 17: 619–63CrossRefGoogle Scholar
Yiou, P., Sornette, D. and Ghil, M. (2000). Data-adaptive wavelets and multi-scale singular-spectrum analysis. Physica D 142: 254–90CrossRefGoogle Scholar
Young, P. C. (1999). Nonstationary time series analysis. Prog. Environ. Sci. 1: 3–48Google Scholar
Yu, Z. and Ito, E. (1999). Possible forcing of century-scale drought frequency in the northern Great Plains. Geology 27: 263–62.3.CO;2>CrossRefGoogle Scholar
Zachos, J. C., Flower, B. P. and Paul, H. (1997). Orbitally paced climate oscillations across the Oligocene/Miocene boundary. Nature 388: 567–70CrossRefGoogle Scholar
Zachos, J. C., Shackleton, N. J., Revenaugh, J. S., Palike, H. and Flower, B. P. (2001). Climate response to orbital forcing across the Oligocene-Miocene boundary. Science 292: 274–8CrossRefGoogle Scholar
Zhang, R-H., Rothstein, L. M. and Busalacchi, A. J. (1998). Origin of upper-ocean warming and El Niño change on decadal scales in the tropical Pacific Ocean. Nature 391: 879–83CrossRefGoogle Scholar
Zolitschka, B. (1996). Image analysis and microscopic investigation of annually laminated lake sediments from Fatetteville Green Lake (NY, USA) Lake C2 (NWT, Canada) and Holzmaar (Germany): a comparison. In: Palaeoclimatology and Palaeoceanography from Laminated Sediments, Ed: A. E. S. Kemp. Geological Society Special Publication No. 116. The Geological Society, London, pp. 49–55

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.

  • References
  • Graham P. Weedon, University of Luton
  • Book: Time-Series Analysis and Cyclostratigraphy
  • Online publication: 13 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511535482.009
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.

  • References
  • Graham P. Weedon, University of Luton
  • Book: Time-Series Analysis and Cyclostratigraphy
  • Online publication: 13 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511535482.009
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.

  • References
  • Graham P. Weedon, University of Luton
  • Book: Time-Series Analysis and Cyclostratigraphy
  • Online publication: 13 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511535482.009
Available formats
×