Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-25T14:44:32.071Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  26 February 2019

Jan Zalasiewicz
Affiliation:
University of Leicester
Colin N. Waters
Affiliation:
University of Leicester
Mark Williams
Affiliation:
University of Leicester
Colin P. Summerhayes
Affiliation:
Scott Polar Research Institute, Cambridge
Get access
Type
Chapter
Information
The Anthropocene as a Geological Time Unit
A Guide to the Scientific Evidence and Current Debate
, pp. 287 - 355
Publisher: Cambridge University Press
Print publication year: 2019

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

Aarkrog, A. (2003). Input of anthropogenic radionuclides into the World Ocean. Deep-Sea Research II, 50(17–21), 25972606.Google Scholar
Abram, N. J., McGregor, H. V., Tierney, J. E., et al. (2016). Early onset of Industrial-era warming across the oceans and continents. Nature, 536, 411418.Google Scholar
Abram, N. J., Mulvaney, R., Wolff, E. W., et al. (2013). Acceleration of snow melt in an Antarctic Peninsula ice core during the twentieth century. Nature Geoscience, 6, 404411.Google Scholar
Abram, N. J., Thomas, E. R., McConnell, J. R., et al. (2010). Ice core evidence for a 20th century decline of sea ice in the Bellingshausen Sea, Antarctica. Journal of Geophysical Research, 115, D23101. doi:10.1029/2010JD014644.CrossRefGoogle Scholar
ACIA (2005). Arctic Climate Impact Assessment. Cambridge University Press.Google Scholar
Aguirre, E., and Pasini, G. (1985). The Pliocene–Pleistocene boundary. Episodes, 8, 116120.Google Scholar
Ahmad, S. M., Padmakumari, V. M., Raza, W., et al. (2011). High-resolution carbon and oxygen isotope records from a scleractinian (Porites) coral of Lakshadweep Archipelago. Quaternary International, 238, 107114.Google Scholar
Ahn, J., and Brook, E. J. (2007). Atmospheric CO2 and climate from 65 to 30 ka BP. Geophysical Research Letters, 34, L10703. doi:10.1029/2007GL029551.Google Scholar
Aitken, A. R. A., Roberts, J. L., van Ommen, T. D., et al. (2016). Repeated large-scale retreat and advance of Totten Glacier indicated by inland bed erosion. Nature, 533, 385389.Google Scholar
Aldridge, D. C., Elliot, P., and Moggridge, G. D. (2004). The recent and rapid spread of the zebra mussel (Dreissena polymorpha) in Great Britain. Biological Conservation, 119, 253261.Google Scholar
Allan, M., Fagel, N., Van Rampelbergh, M., et al. (2015). Lead concentrations and isotope ratios in speleothems as proxies for atmospheric metal pollution since the industrial revolution. Chemical Geology, 401, 140150.Google Scholar
Alley, R. B., Anandakrishnan, S., Jung, P., and Clough, A. (2001). Stochastic resonance in the North Atlantic: Further insights. In Seidov, D., Haupt, B. J., and Maslin, M., eds., The Oceans and Rapid Climate Change: Past, Present and Future (Geophysical Monographs), 126. Washington, DC: American Geophysical Union, pp. 5768.Google Scholar
Allwood, A. C., Grotzinger, J. P., Knoll, A. H., et al. (2009). Controls on development and diversity of early Archean stromatolites. Proceedings of the National Academy of Sciences (USA), 106, 95489555.Google Scholar
Alonso-Hernández, C. M., Tolosa, I., Mesa-Albernas, M., et al. (2015). Historical trends of organochlorine pesticides in a sediment core from the Gulf of Batabanó, Cuba. Chemosphere, 137, 95100.CrossRefGoogle Scholar
Al-Rousan, S., Pätzold, J., Al-Moghrabi, S., and Wefer, G. (2004). Invasion of anthropogenic CO2 recorded in planktonic foraminifera from the northern Gulf of Aquaba. International Journal of Earth Sciences, 93, 10661076.Google Scholar
Alvarez, L. W., Alvarez, W., Asaro, F., and Michel, H. (1984). Extra-terrestrial cause for the Cretaceous-Tertiary extinction: Experimental results and theoretical interpretation. Science, 208, 10951108.Google Scholar
Amos, S. E., and Yalcin, B. (2015). Hollow Glass Microspheres for Plastics, Elastomers, and Adhesives Compounds. Oxford: Elsevier.Google Scholar
Amundson, R., and Jenny, H. (1991). The place of humans in the state factor theory of ecosystems and their soils. Soil Science, 151, 99109.Google Scholar
Anderegg, W. R. L., Kane, J. M., and Anderegg, L. D. L. (2012). Consequences of widespread tree mortality triggered by drought and temperature stress. Nature Climate Change, 3, 3036.CrossRefGoogle Scholar
Andrady, A. (2011). Microplastics in the marine environment. Marine Pollution Bulletin, 62, 15961605.Google Scholar
Angus, I. (2016). Facing the Anthropocene: Fossil Capitalism and the Crisis of the Earth System. New York: Monthly Review Press.Google Scholar
Araújo, D. F., Boaventura, G. R., Machado, W., et al. (2017). Tracing of anthropogenic zinc sources in coastal environments using stable isotope composition. Chemical Geology, 449, 226235.CrossRefGoogle Scholar
Archer, D., Buffett, B., and Brovkin, V. (2009). Ocean methane hydrates as a slow tipping point in the global carbon cycle. Proceedings of the National Academy of Sciences (USA), 106, 2059620601.Google Scholar
Archer, D., Eby, M., Brovkin, V., et al. (2009). Atmospheric lifetime of fossil fuel carbon dioxide. Annual Review of Earth and Planetary Science, 37, 117134.Google Scholar
Archer, D., and Ganopolski, A. (2005). A movable trigger: Fossil fuel CO2 and the onset of the next glacial inception. Geochemistry, Geophysics, Geosystems, 6, Q05003. doi:10.1029/2004GC000891.Google Scholar
Arduino, G. (1760). Due lettere del Sig. Giovanni Arduino sopra varie sue osservazioni naturali. Nuova Raccolta d’Opuscoli Scientifici e Filologici, 6, 97132, 133–180.Google Scholar
Arenillas, I., Alegret, L., Arz, J. A., et al. (2002). Cretaceous-Tertiary boundary planktic foraminiferal mass extinction and biochronology at Le Ceiba and Bochil, Mexico and El Kef, Tunisia. Geological Society of America, Special Paper, 356, 253263.Google Scholar
Arienzo, M. M., McConnell, J. R., Chellman, N., et al. (2016). A method for continuous 239Pu determinations in Arctic and Antarctic ice cores. Environmental Science and Technology, 50(13), 70667073.Google Scholar
Armour, K. C., Marshall, J., Scott, J. R., et al. (2016). Southern Ocean warming delayed by circumpolar upwelling and equatorward transport. Nature Geoscience, 9, 549554.Google Scholar
Arnold, R. W., Szabolcs, J., and Targulian, V. O. (1990). Global Soil Change. Laxenburg, Austria: International Institute for Applied Systems Analysis.Google Scholar
Arrhenius, S. (1896). On the influence of carbonic acid in the air upon the temperature of the ground. London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science, Series 5, 41(251), 237276.Google Scholar
Arrhenius, S. (1908). Worlds in the Making: The Evolution of the Universe. New York: Harper and Bros.Google Scholar
Ashbee, P., Smith, I. F., and Evans, J. G. (1979). Excavation of three long barrows near Avebury, Wiltshire. Proceedings of the Prehistoric Society, 45, 207300.Google Scholar
Aubert, M., Brumm, A., Ramil, M., et al. (2014). Pleistocene cave art from Sulawesi, Indonesia. Nature, 514, 223227.Google Scholar
Aubry, M.-P., Ouda, K., Depuis, C., et al. (2007). The Global Standard Stratotype-section and Point (GSSP) for the base of the Eocene Series in the Dababiya section (Egypt). Episodes, 30, 271286.CrossRefGoogle Scholar
Autin, W. J., and Holbrook, J. M. (2012). Is the Anthropocene an issue of stratigraphy or pop culture? GSA Today, 22(7), 6061.Google Scholar
Avio, C. M., Gorbi, S., and Regoli, F. (2017). Plastics and microplastics in the oceans: From emerging pollutants to emerged threat. Marine Environmental Research, 128, 211.CrossRefGoogle ScholarPubMed
Babcock, L. E., Peng, S., Zhu, M., Xiao, S., and Ahlberg, P. (2014). Proposed reassessment of the Cambrian GSSP. Journal of African Earth Sciences, 98, 310.Google Scholar
Bacaloni, A., Insogna, S., and Zoccolillo, L. (2011). Remote zones air quality. Persistent organic pollutants: Sources, sampling and analysis. In Mazzeo, N., ed., Air Quality Monitoring, Assessment and Management. Rijeka, Croatia: InTech, pp. 223240.Google Scholar
Bacon, A. R., Richter, D. deB., Bierman, P. R., and Rood, D. H. (2012). Coupling meteoric 10Be with pedogenic losses of 9Be to improve soil residence time estimates on an ancient North American interfluve. Geology, 40, 847850.CrossRefGoogle Scholar
Bader, J., Mesquita, D. S., Hodges, K. I., et al. (2011). A review on Northern Hemisphere sea-ice, storminess and the North Atlantic Oscillation: Observations and projected changes. Atmospheric Research, 101(4), 809834.Google Scholar
Baker, T. D., and Jikells, A. R. (2015). Biogeochemical cycles: Heavy metals. In North, G. R., Pyle, J. A., and Zhang, F., eds., Encyclopedia of Atmospheric Sciences, 2nd ed., vol. 1. Academic Press, pp. 201204.Google Scholar
Bakker, P., Clark, P., Golledge, N. R., et al. (2017). Centennial-scale Holocene climate variations amplified by Antarctic Ice Sheet discharge. Nature, 541, 7276.Google Scholar
Balter, M. (2013). Archaeologists say the ‘Anthropocene’ is here—but it began long ago. Science, 340, 261262.Google Scholar
Bamber, J. L., Riva, R. E. M., Vermeersen, B. L. A., and LeBrocq, A. M. (2009). Reassessment of the potential sea-level rise from a collapse of the West Antarctic Ice Sheet. Science, 324, 901903.Google Scholar
Barbante, C., Spolaor, A., Cairns, W. R., and Boutron, C. (2017). Man’s footprint on the Arctic environment as revealed by analysis of ice and snow. Earth-Science Reviews, 168, 218231.Google Scholar
Bard, E., Raisbeck, G., Yiou, F., and Jouzel, J. (2000). Solar irradiance during the last 1200 years based on cosmogenic nuclides. Tellus B, 52, 985992.Google Scholar
Barker, G. (2002). A tale of two deserts: Contrasting desertification histories in Rome’s desert frontiers. World Archaeology, 33, 488507.Google Scholar
Barnes, D. K. A. (2005). Remote islands reveal rapid rise of Southern Hemisphere sea debris. Directions in Science, 5, 915921.Google Scholar
Barnes, D. K. A., Galgani, F., Thompson, R. C., and Barlaz, M. (2009). Accumulation and fragmentation of plastic debris in global environments. Philosophical Transactions of the Royal Society, B364, 19851998.Google Scholar
Barnett, R. L., Newton, T. L., Charman, D. J., and Gehrels, W. R. (2017). Salt-marsh testate amoebae as precise and widespread indicators of sea-level change. Earth-Science Reviews, 164, 193207.Google Scholar
Barnett, V., Payne, R., and Steiner, R. (1995). Agricultural Sustainability. New York: J. Wiley.Google Scholar
Barnola, J. M., Anklin, M., Porcheron, J., et al. (1995). CO2 evolution during the last millennium as recorded by Antarctic and Greenland ice. Tellus B, 47, 264272.Google Scholar
Barnosky, A. D. (2005). Effects of Quaternary climatic change on speciation in mammals. Journal of Mammalian Evolution, 12, 247256.Google Scholar
Barnosky, A. D. (2008). Megafauna biomass tradeoff as a driver of Quaternary and future extinctions. Proceedings of the National Academy of Sciences (USA), 105, 1154311548.Google Scholar
Barnosky, A. D. (2014). Palaeontological evidence for defining the Anthropocene. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 149165.Google Scholar
Barnosky, A. D., Holmes, M., Kirchholtes, R., et al. (2014). Prelude to the Anthropocene: Two new North American Land Mammal Ages (NALMAs). Anthropocene Review, 1(3), 225242.Google Scholar
Barnosky, A. D., and Lindsey, E. L. (2010). Timing of Quaternary megafaunal extinction in South America in relation to human arrival and climate change. Quaternary International, 217, 1029.Google Scholar
Barnosky, A. D., Matzke, N., Tomiya, S., et al. (2011). Has the Earth’s sixth mass extinction already arrived? Nature, 471, 5157.Google Scholar
Baroni, M., Savarino, J., Cole-Dai, J., et al. (2008). Anomalous sulfur isotope compositions of volcanic sulfate over the last millennium in Antarctic ice cores. Journal of Geophysical Research, 113, D20112. doi:10.1029/2008JD010185.Google Scholar
Barrett, E. M. (1963). The California oyster industry. Fish Bulletin, 123, 2103.Google Scholar
Barshis, D. J., Ladner, J. T., Oliver, T. A., et al. (2013). Genomic basis for coral resilience to climate change. Proceedings of the National Academy of Sciences (USA), 110, 13871392.Google Scholar
Bar-Yosef, O. (2002). The Upper Paleolithic revolution. Annual Reviews of Anthropology, 31, 363393.CrossRefGoogle Scholar
Baskin, J. (2015). Paradigm dressed as epoch: The ideology of the Anthropocene. Environmental Values, 24, 929.Google Scholar
Bassis, J. N., Petersen, S. V., and MacCathles, L. (2017). Heinrich events triggered by ocean forcing and modulated by isostatic adjustment. Nature, 542, 332334.Google Scholar
Basson, (2009). Sedimentation and Sustainable Use of Reservoirs and River Systems. Draft ICOLD Bulletin. http://icold-cigb.org/userfiles/files/CIRCULAR/CL1793Annex.pdf.Google Scholar
Battarbee, R. W. (1990). The causes of lake acidification, with special reference to the role of acid deposition. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 327, 339347.Google Scholar
Battarbee, R. W., Jones, V. J., Flower, R. J., et al. (1996). Palaeolimnological evidence for the atmospheric contamination and acidification of high Cairngorm lochs with special reference to Lochnagar. Botanical Journal of Scotland, 48, 7987.Google Scholar
Bauer, A. M., and Ellis, E. C. (2018). The Anthropocene divide: Obscuring understanding of social-environmental change. Current Anthropology, 59(2), 209227.Google Scholar
Baustian, J. J., Mendelssohn, I. A., and Hester, M. W. (2012). Vegetation’s importance in regulating surface elevation in a coastal salt marsh facing elevated rates of sea level rise. Global Change Biology, 18, 33773382.Google Scholar
Bax, N., Williamson, A., Aguero, M., et al. (2003). Marine invasive alien species: A threat to global biodiversity. Marine Policy, 27, 313323.Google Scholar
Beasley, T. M., Carpenter, R., and Jennings, C. D. (1982). Plutonium, 241Am and 137Cs ratios, inventories and vertical profiles in Washington and Oregon continental shelf sediments. Geochimica et Cosmochimica Acta, 46, 19311946.Google Scholar
Beerling, D. J. (2007). The Emerald Planet. Oxford University Press.Google Scholar
Beerling, D. J., and Royer, D. L. (2011). Convergent Cenozoic CO2 history. Nature Geoscience, 4, 418420.Google Scholar
Bell, R. E., Chu, W., Kingslake, J., et al. (2017). Antarctic ice shelf potentially stabilized by export of meltwater in surface river. Nature, 544, 344348.Google Scholar
Bender, M. L. (2013). Paleoclimate. Princeton Primers in Climate. Princeton University Press.Google Scholar
Bennett, C. E., Thomas, R., Williams, M., et al. (in press). The broiler chicken as a signal of a human reconfigured biosphere. Royal Society Open Science.Google Scholar
Benninger, L. K., and Dodge, R. E. (1986). Fallout plutonium and natural radionuclides in annual bands of the coral Montastrea annularis, St. Croix, U.S. Virgin Islands. Geochimica Cosmochimica Acta, 50, 27852797.Google Scholar
Berger, A., and Loutre, M. F. (2002). An exceptionally long interglacial ahead? Science, 297, 12871288.Google Scholar
Berner, R. A. (2004). The Phanerozoic Carbon Cycle: CO2 and O2. Oxford University Press.Google Scholar
Berrojalbiz, N., Lacorte, S., Calbet, A., et al. (2009). Accumulation and cycling of polycyclic aromatic hydrocarbons in zooplankton. Environmental Science & Technology, 43(7), 22952301.Google Scholar
Berry, E. W. (1925). The term Psychozoic. Science, 44, 16.Google Scholar
Berry, K. L. E., Seemann, J., Dellwig, O., et al. (2013). Sources and spatial distribution of heavy metals in scleractinian coral tissues and sediments from the Bocas del Toro Archipelago, Panama. Environmental Monitoring and Assessment, 185(11), 90899099.Google Scholar
Bertler, N. A. M., and Barrett, P. J. (2010). Vanishing polar ice sheets. In Dodson, J., ed., Changing Climates, Earth Systems and Society, International Year of Planet Earth, 49. Springer Science+Business Media B. V., pp. 4984.Google Scholar
Besseling, E., Wang, B., Lürling, M., and Koelmans, A. A. (2014). Nanoplastic affects growth of S. obliquus and reproduction of D. magna. Environmental Science and Technology, 48, 1233612343.Google Scholar
Bhattacharya, D., Agrawa, S., Aranda, M., et al. (2016). Comparative genomics explains the evolutionary success of reef-forming corals. eLife, 5, e13288. doi:10.7554/eLife.13288.Google Scholar
Biber, E. (2016). Law in the Anthropocene Epoch. UC Berkeley Public Law Research Paper No. 2834037 (September). https://papers.ssrn.com/sol3/papers.cfm?abstract_id=2834037.Google Scholar
Bidwell, O. W., and Hole, F. D. (1965). Man as a factor of soil formation. Soil Science, 99, 6572.Google Scholar
Bigalke, M., Weyer, S., Kobza, J., and Wilcke, W. (2010). Stable Cu and Zn isotope ratios as tracers of sources and transport of Cu and Zn in contaminated soil. Geochimica et Cosmochimica Acta, 74(23), 68016813.Google Scholar
Bigalke, M., Weyer, S., and Wilcke, W. (2011). Stable Cu isotope fractionation in soils during oxic weathering and podzolization. Geochimica et Cosmochimica Acta, 75(11), 31193134.Google Scholar
Bigler, M., Wagenbach, D., Fischer, H., et al. (2002). Sulphate record from a northeast Greenland ice core over the last 1200 years based on continuous flow analysis. Annals of Glaciology, 35, 250256.Google Scholar
Bigus, P., Tobiszewski, M., and Namieśnik, J. (2014). Historical records of organic pollutants in sediment cores. Marine Pollution Bulletin, 78(1), 2642.Google Scholar
Bindler, R. (2006). Mired in the past—looking to the future: Geochemistry of peat and the analysis of past environmental changes. Global and Planetary Change, 53(4), 209221.Google Scholar
Bing, H., Wu, Y., Zhou, J., Li, R., and Wang, J. (2016). Historical trends of anthropogenic metals in Eastern Tibetan Plateau as reconstructed from alpine lake sediments over the last century. Chemosphere, 148, 211219.CrossRefGoogle ScholarPubMed
Bintanja, R., and Andry, O. (2017). Towards a rain-dominated Arctic. Nature Climate Change, 7, 263267.Google Scholar
Bintanja, R., van Oldenborgh, G. J., Drijfhout, S. S., et al. (2013). Important role for ocean warming and increased ice-shelf melt in Antarctic sea-ice expansion. Nature Geoscience, 6, 376379.Google Scholar
Birch, G. F., Gunns, T. J., and Olmos, M. (2015). Sediment-bound metals as indicators of anthropogenic change in estuarine environments. Marine Pollution Bulletin, 101(1), 243257.Google Scholar
Bird, E. C. F. (1993). Submerging Coasts: The Effects of a Rising Sea Level on Coastal Environments. Chichester: Wiley.Google Scholar
Blais, J. M., Macdonald, R. W., Mackay, D., et al. (2007). Biologically mediated transport of contaminants to aquatic systems. Environmental Science & Technology, 41(4), 10751084.Google Scholar
Blasing, T. J. (2016). Recent Greenhouse Gas Concentrations. Carbon Dioxide Information Analysis Center. doi:10.3334/CDIAC/atg.032. http://cdiac.ornl.gov/pns/current_ghg.html.Google Scholar
Blöschl, G., Hall, J., Parajka, J., et al. (2017). Changing climate shifts timing of European floods. Science, 357, 588590.Google Scholar
Blunier, T., Chappellez, J., Schwander, J., et al. (1995). Variations in atmospheric methane concentration during the Holocene epoch. Nature, 374, 4649.Google Scholar
Bobrov, V. A., Bogush, A. A., Lenova, G. A., et al. (2011). Anomalous concentrations of zinc and copper in Highmoor Peat Bog, southeast coast of Lake Baikal. Doklady Akademii Nauk, 439, 784788.Google Scholar
Boerger, C. M., Lattin, G. L., Moore, S. L., and Moore, C. J. (2010). Plastic ingestion by planktivorous fishes in the North Pacific Central Gyre. Marine Pollution Bulletin, 60, 22752278.Google Scholar
Bogdal, C., Schmid, P., Zennegg, M., et al. (2009). Blast from the past: Melting glaciers as a relevant source for persistent organic pollutants. Environmental Science & Technology, 43(21), 81738177.Google Scholar
Bollhöfer, A., and Rosman, K. J. R. (2001). Isotopic source signatures for atmospheric lead: The Northern Hemisphere. Geochimica et Cosmochimica Acta, 65(11), 17271740.Google Scholar
Bolton, C. T., Hernández-Sánchez, M. T., Fuertes, M.-Á., et al. (2016). Decrease in coccolithophore calcification and CO2 since the middle Miocene. Nature Communications, 7, 10284. doi:10.1038/ncomms10284.Google Scholar
Bombelli, P., Howe, C. J., and Bertocchini, F. (2017a). Polyethylene bio-degradation by caterpillars of the wax moth Galleria mellonella. Current Biology, 27(8), R292–293.Google Scholar
Bombelli, P., Howe, C. J., and Bertocchini, F. (2017b). Response to Weber et al. Current Biology, 27(15), R745.Google Scholar
Bond, G., Heinrich, H., Broecker, W. S., et al. (1992). Evidence for massive discharges of icebergs into the North Atlantic Ocean during the last glacial period. Nature, 360, 245249.Google Scholar
Bond, G., Kromer, B., Beer, J., et al. (2001). Persistent solar influence on North Atlantic climate during the Holocene. Science, 294, 21302136.Google Scholar
Bond, G., Showers, W., Cheseby, M., et al. (1997). A pervasive millennial-scale cycle in North Atlantic Holocene and glacial climates. Science, 278, 12571266.Google Scholar
Bond, T. C., Bhardwaj, E., Dong, R., et al. (2007). Historical emissions of black and organic carbon aerosol from energy-related combustion, 1850–2000. Global Biogeochemical Cycles, 21, GB2018. doi:10.1029/2006GB002840.Google Scholar
Bonhommet, N., and Zähringer, J. (1969). Paleomagnetism and potassium argon age determinations of the Laschamp geomagnetic polarity event. Earth and Planetary Science Letters, 6, 4346.Google Scholar
Bonneuil, C., and Fressoz, J.-B. (2016). The Shock of the Anthropocene: The Earth, History and Us. Translated from French by Fernbach, D.. London, New York: Verso Books.Google Scholar
Borel, B. (2017). When the pesticides run out. Nature, 543, 302304.Google Scholar
Borsato, A., Frisia, S., Fairchild, I. J., et al. (2007). Trace element distribution in annual stalagmite laminae mapped by micrometer-resolution X-ray fluorescence: Implications for incorporation of environmentally significant species. Geochimica et Cosmochimica Acta, 71, 14941512.Google Scholar
Boucher, J., and Friot, D. (2017). Primary Microplastics in the Oceans: A Global Evaluation of Sources. Gland, Switzerland: IUCN. doi:10.2305/IUCN.CH.2017.01.en.Google Scholar
Boucher, O., Randall, D., Artaxo, P., et al. (2013). Clouds and aerosols. In Stocker, T. F., Qin, D., Plattner, G. K., et al., eds., Climate Change 2013: The Physical Science Basis, Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press, pp. 571657.Google Scholar
Boutron, C. F., Candelone, J. P., and Hong, S. (1995). Greenland snow and ice cores: Unique archives of large-scale pollution of the troposphere of the Northern Hemisphere by lead and other heavy metals. Science of the Total Environment, 160, 233241.Google Scholar
Bowell, R. J., Williams, K. P., Connelly, R. J., et al. (1999). Chemical containment of mine waste. In Metcalfe, R., and Rochelle, C. A., eds., Chemical Containment of Waste in the Geosphere. Geological Society, London, Special Publications, 157, pp. 213240.Google Scholar
Bowen, G. J., Maibauer, B. J., Kraus, M. J., et al. (2015). Two massive, rapid releases of carbon during the Paleocene-Eocene thermal maximum. Nature Geoscience, 8, 4447.Google Scholar
Boyce, J. I., Reinhardt, E. G., and Goodman, B. N. (2009). Magnetic detection of ship ballast mounds and anchorages at Caesarea Maritima, Israel. Journal of Archaeological Science, 36, 15161526.Google Scholar
Bracegirdle, T. J., Connolley, W. M., and Turner, J. (2008). Antarctic climate change over the twenty first century. Journal of Geophysical Research, 113, D03103. doi:10.1029/2007JD008933.CrossRefGoogle Scholar
Brantley, S. L., McDowell, W. H., Dietrich, W. E., et al. (2017). Designing a network of critical zone observatories to explore the living skin of the terrestrial Earth. Earth Surface Dynamics. doi:10.5194/esurf-2017–36.Google Scholar
Braun, H., Christl, M., Rahmstorf, S., et al. (2005). Possible solar origin of the 1,470-year glacial climate cycle demonstrated in a coupled mode. Nature, 438, 208211.Google Scholar
Breitenlechner, E., Stöllner, T., Thomas, P., et al. (2014). An interdisciplinary study on the environmental reflection of prehistoric mining actvities at the Mitterberg main lode (Salzburg, Austria). Archaeometry, 56(1), 102128.Google Scholar
Bridgewater, A. V. (1986). Refuse composition projections and recycling technology. Resources and Conservation, 12, 159174.Google Scholar
Brigham-Grette, J., Melles, M., Minyuk, P., et al. (2013). Pliocene warmth, polar amplification, and stepped Pleistocene cooling recorded in NE Arctic Russia. Science, 340(6139), 14211427.Google Scholar
Brimblecombe, P. (2005). The global sulfur cycle. In Schlesinger, W. H., ed., Biogeochemistry: Treatise on Geochemistry, vol. 8. Elsevier Science, pp. 645682.Google Scholar
Brodie, J., Waterhouse, J., Schaffelke, B., et al. (2013). Scientific Consensus Statement: Land Use Impacts on Great Barrier Reef Water Quality and Ecosystem Condition. Reef Water Quality Protection Plan Secretariat, the State of Queensland. http://reefplan.qld.gov.au/about/assets/scientific-consensus-statement-2013.pdf.Google Scholar
Broecker, W. S. (1998). Paleocean circulation during the last deglaciation: A bipolar seesaw? Paleoceanography, 13, 119121.Google Scholar
Broecker, W. S., and Stocker, T. F. (2006) The Holocene CO2 rise: Anthropogenic or natural? EOS, 87, 27.Google Scholar
Brook, B. W., and Barnosky, A. D. (2012). Quaternary extinctions and their link to climate change. In Hannah, L., ed., Saving a Million Species: Extinction Risk from Climate Change. Washington, DC: Island Press, pp. 179198.Google Scholar
Brown, A., Toms, P., Carey, C., and Rhodes, E. (2013). Geomorphology of the Anthropocene: Time-transgressive discontinuities of human-induced alluviation. Anthropocene, 1, 313.Google Scholar
Browne, M. A., Crump, P., Niven, S. J., et al. (2011). Accumulation of microplastic on shorelines worldwide: Sources and sinks. Environmental Science and Technology, 45, 91759179.Google Scholar
Browne, M. A., Galloway, T. S., and Thompson, R. C. (2010). Spatial patterns of plastic debris along estuarine shorelines. Environmental Science and Technology, 44, 34043409.Google Scholar
Bryant, R. B., and Galbraith, J. M. (2002). Incorporating anthropogenic processes in soil classification. In Eswaran, H., Ahrens, R., Rice, T. J., and Stewart, B. A., eds., Soil Classification: A Global Desk Reference. Boca Raton, Florida: CRC Press, pp. 5766.Google Scholar
Buddemeier, R. W., Baker, A. C., Fautin, D. G., and Jacobs, J. R., eds., (2004). The adaptive hypothesis of bleaching. Coral Health and Disease. Springer, pp. 427444.Google Scholar
Buddemeier, R. W., and Fautin, D. (1993). Coral bleaching as an adaptive mechanism. A testable hypothesis. Bioscience, 43, 320326.Google Scholar
Buddemeier, R. W., and Smith, S. V. (1988). Coral reef growth in an era of rapidly rising sea level: Predictions and suggestions for long-term research. Coral Reefs, 7, 5156.Google Scholar
Budov, V. V. (1994). Hollow glass microspheres. Use, properties, and technology (Review). Glass and Ceramics, 51(7), 230235.Google Scholar
Budyko, M. I. (1977). On present-day climate changes. Tellus, 29, 193204.Google Scholar
Buesseler, K. O., Lamborg, C. H., Boyd, P. W., et al. (2007). Revisiting carbon flux through the ocean’s twilight zone. Science, 316, 567570.Google Scholar
Buffon, G-L. L. de. (2018). The Epochs of Nature, ed. and trans. Zalasiewicz, J., Milon, A.-S., and Zalasiewicz, M.. University of Chicago Press.Google Scholar
Buggisch, W. (1991). The global Frasnian-Famennian “Kellwasser Event.” Geologische Rundschau, 80, 4972.Google Scholar
Buizert, C., and Schmittner, A. (2016). Southern Ocean control of glacial AMOC stability and Dansgaard-Oeschger interstadial duration. Palaeoceanography, 30, 15951612.Google Scholar
Bukata, A. R., and Kyser, T. K. (2007). Carbon and nitrogen isotope variations in tree-rings as records of perturbations in regional carbon and nitrogen cycles. Environmental Science and Technology, 41(4), 13311338.Google Scholar
Buol, S. W., Southard, R. J., Graham, R. C., and McDaniel, P. A. (2011). Soil Genesis and Classification. New York: John Wiley & Sons.Google Scholar
Burghardt, T. E., Pashkevich, A., and Żakowska, L. (2016). Influence of volatile organic compounds emissions from road marking paints on ground-level ozone formation: Case study of Kraków, Poland. Transportation Research Procedia, 14, 714723.Google Scholar
Burke, L., Reytar, K., Spalding, M., et al. (2011). Reefs at Risk Revisited. World Resource Institute, p. 115. http://wri.org/publication/reefs-risk-revisited.Google Scholar
Burney, D. A., James, H. F., Pigott Burney, L., et al. (2001). Fossil evidence for a diverse biota from Kaua’i and its transformation since human arrival. Ecological Monographs, 7, 615641.Google Scholar
Burnley, S. J. (2007). A review of municipal waste composition in the United Kingdom. Waste Management, 27, 12741285.Google Scholar
Burns, A., Pickering, M. D., Green, K. A., et al. (2017). Micromorphological and chemical investigation of late-Viking age grave fills at Hofstaðir, Iceland. Geoderma, 306, 183194. http://dx.doi.org/10.1016/j.geoderma.2017.06.021.Google Scholar
Butler, B. E. (1959). Periodic Phenomena in Landscapes as a Basis for Soil Studies. Melbourne, Australia: CSIRO.Google Scholar
Butler, C. D. (2016). Sounding the alarm: Health in the Anthropocene. International Journal of Environmental Research and Public Health, 13(7), E665.Google Scholar
Butterfield, N. J. (2011). Animals and the invention of the Phanerozoic Earth system. Trends in Ecology & Evolution, 26(2), 8187.Google Scholar
Caiazzo, L., Baccolo, G., Barbante, C., et al. (2017). Prominent features in isotopic chemical and dust stratigraphies in a coastal East Antarctic ice sheet (Eastern Wilkes Land). Chemosphere, 176, 273287.Google Scholar
Callender, E. (2014). Heavy metals in the environment – historical trends. Reference Module in Earth Systems and Environmental Sciences. In Holland, H. D., and Turekian, K. K., eds., Treatise on Geochemistry, vol. 11, 2nd ed. Elsevier Ltd, pp. 5989.Google Scholar
Campbell, J. E., Berry, J. A., Seibt, U., et al. (2017). Large historical growth in global terrestrial gross primary production. Nature, 544, 8487.Google Scholar
Campbell, J. E., Whelan, M. E., Seibt, U., et al. (2015). Atmospheric carbonyl sulfide sources from anthropogenic activity: Implications for carbon cycle constraints. Geophysical Research Letters, 42, 30043010.Google Scholar
Canadell, J. G., Le Quéré, C., Raupach, M. R., et al. (2007). Contributions to accelerating atmospheric CO2 growth from economic activity, carbon intensity, and efficiency of natural sinks. Proceedings of the National Academy of Science (USA), 104, 18866–1870.Google Scholar
Candelone, J. P., Hong, S., Pellone, C., and Boutron, C. F. (1995). Post-Industrial Revolution changes in large-scale atmospheric pollution of the northern hemisphere by heavy metals as documented in central Greenland snow and ice. Journal of Geophysical Research: Atmospheres, 100(D8), 1660516616.Google Scholar
Canfield, D. E., Glazer, A. N., and Falkowski, P. G. (2010). The evolution and future of Earth’s nitrogen cycle. Science, 330, 192196.Google Scholar
Cantin, N. E., and Lough, J. (2014). Surviving coral bleaching events: Porites growth anomalies on the Great Barrier Reef. PLoS One, 9(2), e88720.Google Scholar
Cao, Z. H., Ding, J. L., Hu, Z. H., et al. (2006). Ancient paddy soils from the Neolithic Age in China’s Yangtze River Delta. Naturwissenschaften, 93, 232236.Google Scholar
Capelotti, P. J. (2010). The Human Archaeology of Space: Lunar, Planetary and Interstellar Relics of Exploration. Jefferson: Mcfarland & Co.Google Scholar
Cappelletti, N., Skorupka, C. N., Migoya, M. C., et al. (2014). Behavior of dioxin like PCBs and PBDEs during early diagenesis of organic matter in settling material and bottom sediments from the sewage impacted Buenos Aires’ coastal area, Argentina. Bulletin of Environmental Contamination and Toxicology, 93(4), 388392.Google Scholar
Capron, E., Govin, A., Stone, E. J., et al. (2014). Temporal and spatial structure of multi-millennial temperature changes at high latitudes during the Last Interglacial. Quaternary Science Reviews, 103, 116133.Google Scholar
Carlton, J. T. (1979). History, biogeography, and ecology of the introduced marine and estuarine invertebrates of the Pacific coast of North America. Unpublished PhD dissertation, University of California, Davis.Google Scholar
Carlton, J. T. (1992). Introduced marine and estuarine mollusks of North America: An end-of-the-20th-century perspective. Journal of Shellfish Research, 11, 489505.Google Scholar
Carlton, J. T., and Cohen, A. N. (1998). Periwinkle’s progress: The Atlantic snail Littorina saxatilis (Mollusca: Gastropoda) establishes a colony on a Pacific shore. Veliger, 41, 333338.Google Scholar
Carlton, J. T., Thompson, J. K., Schemel, L. E., and Nichols, F. H. (1990). Remarkable invasion of San Francisco Bay (California, USA) by the Asian clam Potamocorbula amurensis. Marine Ecology Progress Series, 66, 8194.Google Scholar
Carver, M. (1987). Underneath English Towns: Interpreting Urban Archaeology. London: Batsford.Google Scholar
Caseldine, C. J., Turney, C., and Long, A. J. (2010). IPCC and Palaeoclimate: An Evolving Story? Journal of Quaternary Science, 25, 14.Google Scholar
Ceballos, G., Ehrlich, P. R., Barnosky, A. D., et al. (2015). Accelerated modern human–induced species losses: Entering the sixth mass extinction. Scientific Advances, 1(5), e1400253. doi:10.1126/sciadv.1400253.Google Scholar
Certini, G., and Scalenghe, R. (2011). Anthropogenic soils are the golden spikes for the Anthropocene. Holocene, 21, 12691274.Google Scholar
Chakrabarty, D. (2009). The climate of history: Four theses. Critical Inquiry, 35(2), 197222.Google Scholar
Chakrabarty, D. (2014). Conjoined histories. Critical Inquiry, 41(Autumn), 123.Google Scholar
Chappell, J., and Shackleton, N. J. (1986). Oxygen isotopes and sea level. Nature, 6093, 137140.Google Scholar
Charlson, R. J., Anderson, T. L., and McDuff, R. E. (1992). The sulfur cycle. In Butcher, S. S., Charlson, R. J., Orians, G. H., and Wolfe, G. V., eds., Global Biogeochemical Cycles. London: Academic Press, pp. 285300.Google Scholar
Charman, D. J., Roe, H. M., and Gehrels, W. R. (1998). The use of testate amoebae in studies of sea-level change: A case study from the Taf Estuary, South Wales, UK. Holocene, 8, 209218.Google Scholar
Chavez, F. P., Ryan, J., Lluch-Cota, S. E., and Ñiquen, M. (2003). From anchovies to sardines and back: Multidecadal change in the Pacific Ocean. Science, 299, 217221.Google Scholar
Chawla, F., Steinmann, P., Loizeau, J.-L., et al. (2010). Binding of 239Pu and 90Sr to organic colloids in soil solutions: Evidence from a field experiment. Environmental Science & Technology, 44(22), 85098514.Google Scholar
Chemekov, Y. F. (1983). Technogenic deposits. In INQUA Congress, 11, Moscow, Abstracts 3, p. 62.Google Scholar
Chen, C. W., Chen, C. F., and Dong, C. D. (2012). Copper contamination in the sediments of Salt River Mouth, Taiwan. Energy Procedia, 16, 901906.Google Scholar
Church, J. A., Clark, P. U., Cazenave, A., et al. (2013). Sea level change. In Stocker, T. F., Qin, D., Plattner, G. K., et al., eds., Climate Change 2013: The Physical Science Basis, Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press, pp. 11371216.Google Scholar
Church, J. A., and White, N. J. (2011). Sea-level rise from the late 19th to the early 21st Century. Surveys in Geophysics, 32, 585602.Google Scholar
Church, J. A., White, N. J., Aarup, T., et al. (2008). Understanding global sea levels: Past, present and future. Sustainability Science, 3, 922.Google Scholar
Clark, J. D., Beyene, Y., WoldeGabriel, G., et al. (2003). Stratigraphic, chronological and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature, 423, 747752.Google Scholar
Clark, J. D., de Heinzelin, J., Schick, K. D., et al. (1994). African Homo erectus: Old radiometric ages and young Oldowan assemblages in the Middle Awash Valley, Ethiopia. Science, 264, 19071910.Google Scholar
Clark, N., and Yusoff, K. (2017). Geosocial formations and the Anthropocene. Theory, Culture & Society, 34(2–3), 1275.Google Scholar
Clark, P. U., Shakun, J. D., Marcott, S. A., et al. (2016). Consequences of twenty-first-century policy for multi-millennial climate and sea-level change. Nature Climate Change, 6, 360369.Google Scholar
Clark, W. C. (1986). Sustainable development of the biosphere: Themes for a research program. In Clark, W. C., and Munn, R. E., eds., Sustainable Development of the Biosphere. IIASA, Laxenburg: Cambridge University Press, pp. 548.Google Scholar
Clarkson, C., Jacobs, Z., Marwick, B., et al. (2017). Human occupation of northern Australia by 65,000 years ago. Nature, 547, 306310.Google Scholar
Clarkson, O., Kasemann, S. A., Wood, R. A., et al. (2015). Ocean acidification and the Permo-Triassic mass extinction. Science, 348(6231), 229232.Google Scholar
Clette, F., Svalgaard, L., Vaquero, J. M., and Cliver, E. W. (2014). Revisiting the sunspot number. Space Science Review, 186, 35103.Google Scholar
Cocroft, W., and Schofield, G. (2012). The secret hill: Cold War archaeology of the Teufelsberg. British Archaeology, 126, 3843.Google Scholar
Codling, G., Vogt, A., Jones, P. D., et al. (2014). Historical trends of inorganic and organic fluorine in sediments of Lake Michigan. Chemosphere, 114, 203209.Google Scholar
Coe, M. T., and Foley, J. A. (2001). Human and natural impacts on the water resources of the Lake Chad basin. Journal of Geophysical Research, 106, 33493356.Google Scholar
Cofaigh, C. O., Davies, B. J., Livingstone, S. J., et al. (2014). Reconstruction of ice-sheet changes in the Antarctic Peninsula since the Last Glacial Maximum. Quaternary Science Reviews, 100, 87110.Google Scholar
Cohen, A. N. (2004). Invasions in the sea. Park Science, 22, 3741.Google Scholar
Cohen, A. N. (2011). The Exotics Guide: Non-native Marine Species of the North American Pacific Coast. Oakland, CA: Center for Research on Aquatic Bioinvasions, Richmond, CA, and San Francisco Estuary Institute. http://exoticsguide.org.Google Scholar
Cohen, A. N., and Carlton, J. T. (1998). Accelerating invasion rate in a highly invaded estuary. Science, 279, 555557.Google Scholar
Cohen, D. J. (2013). The advent and spread of early pottery in East Asia: New dates and new considerations for the world’s earliest ceramic vessels. Journal of Austronesian Studies, 4(2), 5590.Google Scholar
Cohen, S., Kettner, A. J., Syvitski, J. P. M., and Fekete, B. M. (2013). WBMsed: A distributed global-scale daily riverine sediment flux model–model description and validation. Computers & Geosciences, 53, 8093.Google Scholar
Cohen, T. J., Jansen, J. D., Gliganic, L. A., et al. (2015). Hydrological transformation coincided with megafaunal extinction in central Australia. Geology, 43, 195198.Google Scholar
Cole, J. E. (1996). Coral records of climate change: Understanding past variability in the tropical ocean-stmosphere. In Jones, P. D., Bradley, R. S., and Jouzel, J., eds., Climatic Fluctuations and Forcing: Mechanisms of the Last 2000 Years. Berlin: Springer, pp. 333355.Google Scholar
Cole, M., Lindeque, P., Fileman, E., et al. (2013). Microplastic ingestion by zooplankton. Environmental Science and Technology, 47, 66466655.Google Scholar
Collins, M., Knutti, R., Arblaster, J. M., et al. (2013). Long-term climate change: Projections, commitments and irreversibility. In Stocker, T. F., Qin, D., Plattner, G. K., et al., eds., Climate Change 2013: The Physical Science Basis, Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press, pp. 10291136.Google Scholar
Committee on Nonnative Oysters in the Chesapeake Bay, National Research Council (2004). Non-native Oysters in the Chesapeake Bay.Google Scholar
Conard, N. J. (2003). Palaeolithic ivory sculptures from southwestern Germany and the origin of figurative art. Nature, 426, 830832.Google Scholar
Conard, N. J. (2009). A female figurine from the basal Aurignacian of Hohle Fels Cave in southwestern Germany. Nature, 459, 248252.Google Scholar
Conway, K. W., Krautter, M., Barrie, J. V., and Neuweiler, M. (2001). Hexactinellid sponge reefs on the Canadian continental shelf: A unique “living fossil.” Geoscience Canada, 28, 7178.Google Scholar
Cook, A. J., Fox, A. J., Vaughan, D. G., and Ferrigno, J. G. (2005). Retreating glacier fronts on the Antarctic Peninsula over the past half-century. Science, 308, 541544.Google Scholar
Cook, A. J., Holland, P. R., Meredith, M. P., et al. (2016). Ocean forcing of glacier retreat in the western Antarctic Peninsula. Science, 353(6296), 283286.Google Scholar
Cook, A. J., van der Flierdt, T., Williams, T., et al. (2013). Dynamic behaviour of the East Antarctic ice sheet during Pliocene warmth. Nature Geoscience, 6, 765769.Google Scholar
Cook, A. J., and Vaughan, D. G. (2010). Overview of areal changes of the ice shelves on the Antarctic Peninsula over the past 50 years. Cryosphere, 4(10), 7798.Google Scholar
Cook, D. E., and Gale, S. J. (2005). The curious case of the date of introduction of leaded fuel to Australia: Implications for the history of Southern Hemisphere atmospheric lead pollution. Atmospheric Environment, 39(14), 25532557.Google Scholar
Cook, P. J., and Shergold, J. H. (eds.) (1986). Phosphate Deposits of the World: Volume 3 -Neogene to Modern Phosphorites. Cambridge University Press.Google Scholar
Cooke, C. A., and Bindler, R. (2015). Lake sediment records of preindustrial metal pollution. In Blais, J. M., Rosen, M. R., and Smol, J. P., eds., Environmental Contaminants: Using Natural Archives to Track Sources and Long-Term Trends of Pollution. Dordrecht: Springer, pp. 101119.Google Scholar
Cooper, A. H., Brown, T. J., Price, S. J., et al. (2018, in press). Humans are the most significant global geological driving force of the 21st century. Anthropocene Review.Google Scholar
Copper, P. (2001). Evolution, radiations, and extinctions in Proterozoic to mid-Paleozoic reefs. In Stanley, G. D. Jr., ed., The History and Sedimentology of Ancient Reef Systems. Topics in Geobiology, 17, pp. 89119. New York: Academic/Plenum Publishers.Google Scholar
Corcoran, P. L. (2015). Benthic plastic debris in marine and fresh water environments. Environmental Science Processes & Impacts, 17, 13631369.Google Scholar
Corcoran, P. L., Moore, C. J., and Jazvac, K. (2014). An anthropogenic marker horizon in the rock record. Geological Society of America Today, 24(6), 48.Google Scholar
Corcoran, P. L., Norris, T., Ceccanese, T., et al. (2015). Hidden plastics of Lake Ontario, Canada and their potential preservation in the sediment record. Environmental Pollution, 204, 1725.Google Scholar
Cordell, D., Drangert, J.-O., and White, S. (2009). The story of phosphorus: Global food security and food for thought. Global Environmental Change, 19, 292305.Google Scholar
Correggiari, A., Cattaneo, A., and Trincardi, F. (2005). The modern Po Delta system: Lobe switching and asymmetric prodelta growth. Marine Geology, 222, 4974.Google Scholar
Craig, O. E., Saul, H., Lucquin, A., et al. (2013). Earliest evidence for the use of pottery. Nature, 496, 351354.Google Scholar
Crucifix, M. (2012). Oscillators and relaxation phenomena in Pleistocene climate theory. Philosophical Transactions of the Royal Society A, 370, 11401165.Google Scholar
Crutzen, P. J. (2002). Geology of Mankind. Nature, 415(January), 23.Google Scholar
Crutzen, P. J. (2006). Albedo enhancement by stratospheric sulfur injections: A contribution to resolve a policy dilemma? Climate Change, 77, 211219.Google Scholar
Crutzen, P. J., and Stoermer, E. F. (2000). The “Anthropocene.” IGBP Global Change Newsletter, 41, 1718.Google Scholar
Cuffey, K. M., Clow, G. D., Steig, E. J., et al. (2016). Deglacial temperature history of West Antarctica. Proceedings of the National Academy of Science (USA), 113(50), 1424914254.Google Scholar
Cui, Y., Kump, L. R., Ridgwell, A. J., et al. (2011). Slow release of fossil carbon during the Palaeocene-Eocene Thermal Maximum. Nature Geoscience, 4(7), 481485.Google Scholar
Dachs, J., and Méjanelle, L. (2010). Organic pollutants in coastal waters, sediments, and biota: A relevant driver for ecosystems during the Anthropocene? Estuaries and Coasts, 33(1), 114.Google Scholar
Dalby, S. (2009). Security and Environmental Change. Cambridge/Boston: Polity.Google Scholar
Darwin, C. R. (1881). The Formation of Vegetable Mould. London: John Murray.Google Scholar
Davies, J. (2016). The Birth of the Anthropocene. Oakland: University of California Press.Google Scholar
Davis, H., and Turpin, E. (2015). Art in the Anthropocene: Encounters among Aesthetics, Politics, Environments and Epistemologies. London: Open Humanities Press.Google Scholar
Davis, R. V. (2011). Inventing the present: Historical roots of the Anthropocene. Earth Sciences History, 30(1), 6384.Google Scholar
Dawson, A. G., and Stewart, I. (2007). Tsunami deposits in the geological record. Sedimentary Geology, 200(3–4), 166183.Google Scholar
Day, J. W. Jr., Gunn, J. D., Folan, W., and Yanez-Arancibia, A. (2007). Emergence of complex societies after sea level stabilized. EOS, Transactions of the American Geophysical Union, 88, 169176.Google Scholar
De Beer, J., Price, S. J., and Ford, J. R. (2012). 3D modelling of geological and anthropogenic deposits at the World Heritage Site of Bryggen in Bergen, Norway. Quaternary International, 251, 107116.Google Scholar
de Bruyn, A. M. H., and Gobas, F. A. P. C. (2004). Modelling the diagenetic fate of persistent organic pollutants in organically enriched sediments. Ecological Modelling, 179(3), 405416.Google Scholar
Deacon, T. W. (2012). Incomplete Nature: How Mind Emerged from Matter. New York: Norton.Google Scholar
Dean, J. R., Leng, M. J., and Mackay, A. W. (2014). Is there an isotopic signature of the Anthropocene? Anthropocene Review, 1(3), 276287.Google Scholar
Dearing, J. A., and Jones, R. T. (2003). Coupling temporal and spatial dimensions of global sediment flux through lake and marine sediment records. Global and Planetary Change, 39, 147168.Google Scholar
Dearing, J. A., Yang, X.-D., Dong, X.-H., et al. (2012). Extending the timescale and range of ecosystem services through paleoenvironmental analyses, exemplified in the lower Yangtze basin. Proceedings of the National Academy of Sciences (USA), 109, E1111E1120.Google Scholar
DeCarlo, T. M., and Cohen, A. L. (2017). Dissepiments, density bands and signatures of thermal stress in Porites skeletons. Coral Reefs, 6, 749761.Google Scholar
DeConto, R. M., and Pollard, D. (2016). Contribution of Antarctica to past and future sea-level rise. Nature, 531, 591597.Google Scholar
Dedkov, A. P., and Mozzherin, V. I. (1992). Erosion and sediment yield in mountain regions of the world. In Walling, D. E., Davies, T. R., and Hasholt, B., eds., Erosion, Debris Flows and Environment in Mountain Regions, 209. Wallingford, UK: International Association of Hydrological Sciences (IAHS) Publication, pp. 2936.Google Scholar
DeFelice, R. C., Eldredge, L. G., and Carlton, J. T. (2001). Nonindigenous invertebrates. In Eldredge, L. G., and Smith, C. M., eds., A Guidebook of Introduced Marine Species in Hawaii. Bishop Museum Technical Report, p. 21.Google Scholar
Dehaut, A., Cassone, A.-L., Frère, L., et al. (2016). Microplastics in seafood: Benchmark protocol for their extraction and characterization. Environmental Pollution, 215, 223233.Google Scholar
Dekiff, J. H., Remy, D., Kasmeier, J., and Fries, E. (2014). Occurrence and spatial distribution of microplastics in sediments from Norderney. Environmental Pollution, 186, 248256.Google Scholar
Della Torre, C., Bergami, E., Salvati, A., et al. (2014). Accumulation and embryotoxicity of polystyrene nanoparticles at early stage of development of sea urchin embryos Paracentrotus lividus. Environmental Science and Technology, 48, 1230212311.Google Scholar
DeLong, K. L., Maupin, C. R., Flannery, J. A., et al. (2016). Refining temperature reconstructions with the Atlantic coral Siderastrea siderea. Palaeogeography, Palaeoclimatology, Palaeoecology, 462, 115.Google Scholar
Department for Environment Food and Rural Affairs (2009). Municipal Waste Composition: A Review of Municipal Waste Component Analyses (Project WR0119).Google Scholar
D’Errico, F., Henshilwood, C., Vanhaeren, M., and van Niekerk, K. (2005). Nassarius kraussianus shell beads from Blombos Cave: Evidence for symbolic behaviour in the Middle Stone Age. Journal of Human Evolution, 48, 324.Google Scholar
Deschamps, P., Durand, N., Bard, E., et al. (2012). Ice-sheet collapse and sea-level rise at the Bølling warming 14,600 years ago. Nature, 483, 559564.Google Scholar
Desnoyers, J. (1829). Observations sur un ensemble de dépôts marins plus récents que les terrains tertiaires du Bassin de la Seine et constituant une formation géologique distincte: Précédés d’un apercu de la nonsimultanéité des bassins tertiaires. Annales Scientifiques Naturelles, 16(171–214), 402419.Google Scholar
Diamond, J., and Bellwood, P. (2003). Farmers and their languages: The first expansions. Science, 300, 597603.Google Scholar
Diaz, R. J., and Rosenberg, R. (2008). Spreading dead zones and consequences for marine ecosystems. Science, 321, 926929.Google Scholar
Ding, Q., Schweiger, A., L’Heureux, M., et al. (2017). Influence of high-latitude atmospheric circulation changes on summertime Arctic sea ice. Nature Climate Change, 7, 289295.Google Scholar
Dodd, M. S., Papineau, D., Grenne, T., et al. (2017). Evidence for early life in Earth’s oldest hydrothermal vent precipitates. Nature, 543, 6064.Google Scholar
Dokuchaev, V. V. (1883). Russian chernozem. In Selected Works of V. V. Dokuchaev, vol. 1, trans. Kaner, N.. Jerusalem: Israel Program for Scientific Translations Ltd (for USDA-NSF), pp. 14419.Google Scholar
Dolan, A. M., Haywood, A. M., Hill, D. J., et al. (2011). Sensitivity of Pliocene ice sheets to orbital forcing. Palaeogeography, Palaeoclimatology, Palaeoecology, 309, 98110.Google Scholar
D’Olivo, J. P., and McCulloch, M. T. (2017). Response of coral calcification and calcifying fluid composition to thermally induced bleaching stress. Scientific Reports, 7(2207).Google Scholar
Domack, E. W., Amblas, D., Gilbert, R., et al. (2006). Subglacial morphology and glacial evolution of the Palmer deep outlet system, Antarctic Peninsula. Geomorphology, 75, 125142.Google Scholar
Donner, S. D., Rickbeil, G. J. M., and Heron, S. F. (2017). A new, high-resolution global mass coral bleaching database. PLoS ONE 12(4), e0175490.Google Scholar
Döring, M. (2007). Wasser für Gadara: 94 km langer Tunnel antiker Tunnel im Norden Jordaniens entdeckt. Querschnitt, 21, 2435Google Scholar
Douglas, I. (1993). Sediment transfer and siltation. In Turner, B. L., Clark, W. C., Kates, R. W., et al., eds., The Earth as Transformed by Human Action. Cambridge University Press, pp. 215234.Google Scholar
Douglas, I., and Lawson, N. (2001). The human dimensions of geomorphological work in Britain. Journal of Industrial Ecology, 4(2), 933.Google Scholar
Dounas, C., Davies, I., Triantafyllou, G., et al. (2007). Large-scale impacts of bottom trawling on shelf primary productivity. Continental Shelf Research, 27, 21982210.Google Scholar
Drage, D., Mueller, J. F., Birch, G., et al. (2015). Historical trends of PBDEs and HBCDs in sediment cores from Sydney estuary, Australia. Science of the Total Environment, 512, 177184.Google Scholar
Drijfhout, S., Bathiany, S., Brovkin, V., et al. (2015). Catalogue of abrupt shifts in Intergovernmental Panel on Climate Change climate models. Proceedings of the National Academy of Sciences (USA), 112(43), E5777E5786.Google Scholar
Druffel, E. R. M. (1996). Post-bomb radiocarbon records of surface corals from the tropical Atlantic Ocean. Radiocarbon, 38(3), 563572.Google Scholar
Duan, K., Thompson, L. G., Yao, T., et al. (2007). A 1000 year history of atmospheric sulfate concentrations in southern Asia as recorded by a Himalayan ice core. Geophysical Research Letters, 34, L01810. doi:10.1029/2006GL027456.Google Scholar
Dudal, R. (2005). The sixth factor of soil formation. Eurasian Soil Science, 38, S60.Google Scholar
Dunham, A. C. (1992). Developments in industrial mineralogy: I. The mineralogy of brick-making. Proceedings of the Yorkshire Geological Society, 49(2), 95104.Google Scholar
Ebbesson, J. (2014). Social-ecological security and international law in the Anthropocene. In Ebbesson, J., et al., eds., International Law and Changing Perceptions of Security: Liber Amicorum Said Mahmoudi. Boston/Leiden: Brill/Martinus Nijhoff, pp. 7192.Google Scholar
Eby, N., Hermes, R., Charnley, N., et al. (2010). Trinitite–the atomic rock. Geology Today, 26(5), 180185.Google Scholar
Edgeworth, M. (2011). Fluid Pasts: Archaeology of Flow. London: Bloomsbury Academic.Google Scholar
Edgeworth, M. (2013). Scale. In Graves-Brown, P., Harrison, R., and Piccini, A., eds., The Oxford Handbook of the Archaeology of the Contemporary World. Oxford University Press.Google Scholar
Edgeworth, M. (2014). The relationship between archaeological stratigraphy and artificial ground and its significance to the Anthropocene. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 91108.Google Scholar
Edgeworth, M., Richter, D. deB., Waters, C. N., et al. (2015). Diachronous beginnings of the Anthropocene: The lower bounding surface of anthropogenic deposits. Anthropocene Review, 2(1), 3358.Google Scholar
Edinburgh, T., and Fay, J. J. (2016). Estimating the extent of Antarctic summer sea ice during the Heroic Age of Antarctic Exploration. Cryosphere, 10, 27212730.Google Scholar
Edwards, K. J., and Whittington, G. (2001). Lake sediments, erosion and landscape change during the Holocene in Britain and Ireland. Catena, 42, 143173.Google Scholar
Eerkes-Medrano, D., Thompson, R. C., and Aldridge, D. C. (2015). Microplastics in freshwater systems: A review of the emerging threats, identification of knowledge gaps and prioritisation of research needs. Water Resources, 75, 6382.Google Scholar
Eichler, A., Tobler, L., Eyrikh, S., et al. (2014). Ice-core based assessment of historical anthropogenic heavy metal (Cd, Cu, Sb, Zn) emissions in the Soviet Union. Environmental Science & Technology, 48(5), 26352642.Google Scholar
Eighmy, J. L., and Sternberg, R. S. (eds.) (1990). Archaeomagnetic Dating. Tucson: University of Arizona Press.Google Scholar
Ekaykin, A. E., Vladimirova, D. O., Lipenkov, V. Y., and Masson-Delmotte, V. (2016). Climatic variability in Princess Elizabeth Land (East Antarctica) over the last 350 years. Climate of the Past, 13, 6171.Google Scholar
Ekdahl, E. J., Teranes, J. L., Guilderson, T. P., et al. (2004). Prehistorical record of cultural eutrophication from Crawford Lake, Canada. Geology, 32, 745748.Google Scholar
Ellis, E. C. (2011). Anthropogenic transformation of the terrestrial biosphere. Philosophical Transactions of the Royal Society A, 369, 10101035.Google Scholar
Ellis, E. C. (2015). Ecology in an anthropogenic biosphere. Ecological Monographs, 85, 287331.Google Scholar
Ellis, E. C., Maslin, M., Boivin, N., and Bauer, A. (2016). Involve social scientists in defining the Anthropocene. Nature, 540, 192193.Google Scholar
Elsig, J., Schmitt, J., Leuenberger, D., et al. (2009). Stable isotope constraints on Holocene carbon cycle changes from an Antarctic ice core. Nature, 461(7263), 507510.Google Scholar
Emerson, T. E., Hedman, K. M., and Simon, M. L. (2005). Marginal horticulturalists or maize agriculturalists? Archaeobotanical, paleopathological, and isotopic evidence relating to Langford Tradition Maize Consumption. Midcontinent Journal of Archaeology, 30, 67118.Google Scholar
Emmett, R., and Lekan, T. (2016). Whose Anthropocene? Revisiting Dipesh Chakrabarty’s “Four Theses.” Transformations in Environment and Society, 2016/2, Munich: Rachel Carson Center. http://environmentandsociety.org/sites/default/files/2015_new_final.pdf.Google Scholar
EPICA Community Members (2006). One-to-one coupling of glacial climate variability in Greenland and Antarctica. Nature, 444, 195198.Google Scholar
Epstein, S., Krishnamurthy, R. V., Oeschger, H., et al. (1990) Environmental information in the isotopic record in trees [and discussion]. Philosphical Transactions of the Royal Society of London A, 330, 427439.Google Scholar
Ericson, J. P., Vorosmarty, C. J., Dingman, S. L., et al. (2005). Effective sea-level rise and deltas: Causes of change and human dimension implications. Global and Planetary Change, 50, 6382.Google Scholar
Eriksen, M., Masin, S., Wilson, S., et al. (2013). Microplastic pollution in surface waters of the Laurentian Great Lakes. Marine Pollution Bulletin, 77, 177182.Google Scholar
Erisman, J. W., Sutton, M. A., Galloway, J., et al. (2008). How a century of ammonia synthesis changed the world. Nature Geoscience, 1, 636639.Google Scholar
Erlandson, J. (2013). Shell middens and other anthropogenic soils as global stratigraphic signatures of the Anthropocene. Anthropocene, 4, 2432.Google Scholar
Erwin, D. H., Laflamme, M., Tweedt, S. M., et al. (2011). The Cambrian conundrum: Early divergence and later ecological success in the early history of animals. Science, 334, 10911097.Google 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, 22502253.Google Scholar
Etheridge, D. M., Steele, L. P., Langenfelds, R. L., et al. (1996). Natural and anthropogenic changes in atmospheric CO2 over the last 1000 years from air in Antarctic ice and firn. Journal of Geophysical Research, 101(D2), 41154128.Google Scholar
Evans, D., Stephenson, M., and Shaw, R. (2009). The present and future use of “land” below ground. Land Use Policy, 26S, S302S316.Google Scholar
Evans, M. E., and Heller, F. (2003). Environmental Magnetism: Principles and Applications of Enviromagnetics. Amsterdam, Boston: Academic Press.Google Scholar
Fairbanks, R. G., and Matthews, R. K. (1978). The marine oxygen isotope record in Pleistocene coral, Barbados, West Indies. Quaternary Research, 10, 181196Google Scholar
Fairbridge, R. W. (1961). Eustatic changes in sea level. In Ahrens, L. H., Press, F., Rankama, K., and Runcorn, S. K., eds., Physics and Chemistry of the Earth, 4. New York: Pergamon, pp. 99185.Google Scholar
Fairchild, I. J. (2018). Geochemical records in speleothems. In DellaSala, D., and Goldstein, M. I., eds., Encyclopedia of the Anthropocene, vol. 1. Oxford: Elsevier. doi:10.1016/B978-0-12-809665-9.09775 -5.Google Scholar
Fairchild, I. J., and Frisia, S. (2014). Definition of the Anthropocene: A view from the underworld. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society Special Publication, 395, pp. 239254.Google Scholar
Fairchild, I. J., Loader, N. J., Wynn, P. M., et al. (2009). Sulfur fixation in wood mapped by synchrotron X-ray studies: Implications for environmental archives. Environmental Science and Technology, 43, 13101315.Google Scholar
Fairchild, I. J., Quest, M., Tucker, M. E., and Hendry, G. L. (1988). Chemical analysis of sedimentary rocks. In Tucker, M. E., ed., Techniques in Sedimentology. Oxford: Blackwells, pp. 274354.Google Scholar
Falcon, A. (2015). Aristotle on causality. In Zalta, E. N., ed., The Stanford Encyclopedia of Philosophy (Spring 2015 edition). https://plato.standord.edu/archives/spr2015/entries/aristotle-causality/.Google Scholar
Farbstein, R., Radić, D., Brajković, D., and Miracle, P. T. (2012). First Epigravettian ceramic figurines from Europe (Vela Spila, Croatia). PLoS ONE, 7, e41437.Google Scholar
Fekiacova, Z., Cornu, S., and Pichat, S. (2015). Tracing contamination sources in soils with Cu and Zn isotopic ratios. Science of the Total Environment, 517, 96105.Google Scholar
Feldpausch, T. R., Phillips, O. M., Brienen, T. J. W., et al. (2016). Amazon forest response to repeated droughts. Global Biogeochemical Cycles, 30, 964982.Google Scholar
Felis, T., Merkel, U., Asami, R., et al. (2012). Pronounced interannual variability in tropical South Pacific temperatures during Heinrich Stadial 1. Nature Communications, 3, 965.Google Scholar
Ferré, B., Durrieu de Madron, X., Estournel, C., et al. (2008). Impact of natural (waves and currents) and anthropogenic (trawl) resuspension on the export of particulate matter to the open ocean: Application to the Gulf of Lion (NW Mediterranean). Continental Shelf Research, 28, 20712091.Google Scholar
Ferretti, D. F., Miller, J. B., White, J. W. C., et al. (2005). Unexpected changes to the global methane budget over the last 2,000 years. Science, 309, 17141717.Google Scholar
Ferse, S. (2008). Artificial Reefs and Coral Transplantation: Fish Community Responses and Effects on Coral Recruitment in North Sulawesi/Indonesia. Dissertation. University of Bremen.Google Scholar
Fey, F., Dankers, N., Steenbergen, J., and Goudswaard, K. (2010). Development and distribution of the non-indigenous Pacific oyster (Crassostrea gigas) in the Dutch Wadden Sea. Aquaculture International, 18, 4559.Google Scholar
Field, D. B., Baumgartner, T. R., Charles, C. D., Ferreira-Bartrina, V., and Ohman, M. D. (2006). Planktonic foraminifera of the California Current reflect 20th-century warming. Science, 311, 6366.Google Scholar
Field, L. P., Milodowski, A. E., Shaw, R. P., et al. (2017). Unusual morphologies and the occurrence of ikaite (CaCO3・6H2O) pseudomorphs in rapid growth, hyperalkaline speleothem. Mineralogical Magazine, 81(3), 565589.Google Scholar
Fielding, C. R., Browne, G. H., Field, B., et al. (2011). Sequence stratigraphy of the ANDRILL AND-2A drillcore, Antarctica: A long-term, ice-proximal record of Early to Mid-Miocene climate, sea-level and glacial dynamism. Palaeogeography, Palaeoclimatology, Palaeoecology, 305, 337351.Google Scholar
Figueres, C., Schellnhuber, H. J., Whiteman, G., et al. (2017). Three years to safeguard our climate. Nature, 546, 593595.Google Scholar
Filippelli, G. (2002). The global phosphorus cycle. Reviews in Mineralogy and Geochemistry, 48, 391425.Google Scholar
Finlayson, B., Mithen, S. J., Najjar, M., et al. (2011). Architecture, sedentism, and social complexity at Pre-Pottery Neolithic A WF16, Southern Jordan. Proceedings of the National Academy of Sciences (USA), 108(20), 81838188.Google Scholar
Finney, S. C. (2014). The “Anthropocene” as a ratified unit in the ICS International Chronostratigraphic Chart: Fundamental issues that must be addressed by the Task Group. In Waters, C. N., Zalasiewicz, J. A., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 2328.Google Scholar
Finney, S. C., and Edwards, L. E. (2016). The “Anthropocene” epoch: Scientific decision or political statement? GSA Today, 26(2–3), 410.Google Scholar
Firestone, R. B., West, A., and Kennett, J. P., et al. (2007). Evidence for an extraterrestrial impact 12,900 years ago that contributed to the megafaunal extinctions and the younger Dryas cooling. Proceedings of the National Academy of Sciences (USA), 104, 1601616021.Google Scholar
Fischer, H., Wagenbach, D., and Kipfstuhl, J. (1998). Sulfate and nitrate firn concentrations on the Greenland ice sheet: 2. Temporal anthropogenic deposition changes. Journal of Geophysical Research, 103, 2193521942.Google Scholar
Fischer, V., Elsner, N. O., Brenke, N., et al. (2015). Plastic pollution of the Kuril–Kamchatka Trench area (NW Pacific). Deep Sea Research, Part II, 111, 399405.Google Scholar
Fischer-Kowalski, M., Krausmann, F., and Pallua, I., 2014. A sociometabolic reading of the Anthropocene: Modes of subsistence, population size and human impact on Earth. Anthropocene Review, 1, 833.Google Scholar
Fisher, A. T., Mankoff, K. D., Tulaczyk, S. M., et al. (2015). High geothermal heat flux measured below the West Antarctic Ice Sheet. Science Advances, 1(6), e1500093.Google Scholar
FitzGerald, D. M., Fenster, M. S., Argow, B. A., and Buynevich, I. V. (2008). Coastal impacts due to sea-level rise. Annual Review of Earth and Planetary Sciences, 36, 601647.Google Scholar
Flügel, E., and Senowbari-Daryan, B. (2001). Triassic reefs of the Tethys. In Stanley, G. D. Jr., ed., The History and Sedimentology of Ancient Reef Systems. Topics in Geobiology, 17. New York: Academic/Plenum Publishers, pp. 217249.Google Scholar
Fofonoff, P. W., Ruiz, G. M., Steves, B., et al. (2017). National Exotic Marine and Estuarine Species Information System. http://invasions.si.edu/nemesis/ (accessed 2 September 2017).Google Scholar
Fohlmeister, J., Kromer, B., and Mangini, A. (2011). The influence of soil organic matter age spectrum on the reconstruction of atmospheric 14C levels via stalagmites. Radiocarbon, 53, 99115.Google Scholar
Foley, R. A., and Lahr, M. M. (2015). Lithic landscapes: Early human impact from stone tool production on the central Saharan environment. PLo SONE, 10(3), e0116482. doi:10.1371/journal. pone.0116482.Google Scholar
Foley, S. F., Gronenborn, D., Andreae, M. O., et al. (2013). The Palaeoanthropocene: The beginnings of anthropogenic environmental change. Anthropocene, 3, 8388.Google Scholar
Food and Agriculture Organization (FAO) of the United Nations, World Food Situation, monthly update. http://fao.org/worldfoodsituation/csdb/en/ (accessed 22 June 2018).Google Scholar
Food and Agriculture Organization of the United Nations (FAOSTAT) (2017). FAOSTAT database. Available from FAO, Rome. http://faostat.fao.org/.Google Scholar
Forbes, D., and Syvitski, J. P. M. (1995). Paraglacial coasts. In Woodruffe, C., and Carter, R. W. G., eds., Coastal Evolution. Cambridge University Press, pp. 373424.Google Scholar
Forbes, R. J. (1958). Man the Maker: A History of Technology and Engineering. New York: Abelard-Schuman.Google Scholar
Ford, J. R., Kessler, H., Cooper, A. H., et al. (2010). An enhanced classification of artificial ground. British Geological Survey, Open Report OR/10/036.Google Scholar
Ford, J. R., Price, S. J., Cooper, A. H., and Waters, C. N. (2014). An assessment of lithostratigraphy for anthropogenic deposits. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 5589.Google Scholar
Foster, G. L., Lear, C. H., and Rae, J. W. B. (2012). The evolution of pCO2, ice volume and climate during the middle Miocene. Earth and Planetary Science Letters, 341 –344, 243254.Google Scholar
Foster, G. L., Royer, D. L., and Lunt, D. J. (2017). Future climate forcing potentially without precedent in the last 420 million years. Nature Communications, 8, 14845. doi:10.1038/ncomms14845.Google Scholar
Francey, R. J., Allison, C. E., Etheridge, D. M., et al. (1999). A 1000-year high precision record of δ13C in atmospheric CO2. Tellus B, 51, 170193.Google Scholar
Free, C. M., Jensen, O. P., Mason, S. A., et al. (2014). High-levels of microplastic pollution in a large, remote, mountain lake. Marine Pollution Bulletin, 85, 156163.Google Scholar
Freestone, D., Vidas, D., and Torres Camprubi, A. (2017). Sea level rise and impacts on maritime zones and limits: The work of the ILA Committee on International Law and Sea Level Rise. Korean Journal of International and Comparative Law, 5(1), 535.Google Scholar
Friedman, G. M. (1950). Benthos of Lake Sevan. Works of the Sevan Hydrobiological Station, 11, 793.Google Scholar
Friedman, T. (2016). Thank You for Being Late. Allen Lane/Penguin Books.Google Scholar
Frisia, S., Borsato, A., Fairchild, I. J., and Susini, J. (2005). Variations in atmospheric sulphate recorded in stalagmites by synchrotron micro-XRF and XANES analyses. Earth and Planetary Science Letters, 235, 729740.Google Scholar
Frisia, S., Borsato, A., Preto, N., and McDermott, F. (2003). Late Holocene annual growth in three Alpine stalagmites records the influence of solar activity and the North Atlantic Oscillation on winter climate. Earth and Planetary Science Letters, 216, 411424.Google Scholar
Fudge, T. J., Steig, E. J., Markle, B. R., et al. (2013). Onset of deglacial warming in West Antarctica driven by local orbital forcing. Nature, 500, 440446.Google Scholar
Fujii, T., Moynier, F., Abe, M., et al. (2013). Copper isotope fractionation between aqueous compounds relevant to low temperature geochemistry and biology. Geochimica et Cosmochimica Acta, 110, 2944.Google Scholar
Gabrieli, J., Cozzi, G., Vallelonga, P., et al. (2011). Contamination of Alpine snow and ice at Colle Gnifetti, Swiss/Italian Alps, from nuclear weapons tests. Atmospheric Environment, 45, 587593.Google Scholar
Gallardo, B., and Aldridge, D. C. (2013). Evaluating the combined threat of climate change and biological invasions on endangered species. Biological Conservation, 160, 225233.Google Scholar
Gallet, Y., Genevey, A., and Courtillot, V. (2003). On the possible occurrence of archeomagnetic jerks in the geomagnetic field over the past three millennia. Earth and Planetary Science Letters, 214, 237242.Google Scholar
Galloway, T. S., Cole, M., and Lewis, C. (2017). Interactions of microplastic debris throughout the marine ecosystem. Nature Ecology and Evolution, 1. doi:10.1038/s41559-017-0116.Google Scholar
Gałuszka, A., and Migaszewski, Z. M. (2017). Glass microspheres as a potential indicator of the Anthropocene: A first study in an urban environment. Holocene, 28(2), 323329.Google Scholar
Gałuszka, A., and Migaszewski, Z. M. (2018). Chemical signals of the Anthropocene. In DellaSala, D., and Goldstein, M. I., eds., Encyclopedia of the Anthropocene, vol. 1. Oxford: Elsevier.Google Scholar
Gałuszka, A., Migaszewski, Z. M., and Zalasiewicz, J. (2014). Assessing the Anthropocene with geochemical methods. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 221238.Google Scholar
Gambaryan, M. G. (1979). Temperature regime of Lake Sevan. Trudy Sevanskoi Gidrobiologicheskoi Stantsii, 17, 123129 [in Russian].Google Scholar
Ganopolski, A., Winkelmann, R., and Schellnhuber, H. J. (2016) Critical insolation–CO2 relation for diagnosing past and future glacial inception. Nature, 529, 200203.Google Scholar
Garcia, T., Féraud, G., Falguères, C., et al. (2010). Earliest human remains in Eurasia: New 40Ar/39Ar dating of the Dmanisi hominid-bearing levels, Georgia. Quaternary Geochronology, 5, 443451.Google Scholar
García-Alix, A., Jimenez-Espejo, F. J., Lozano, J. A., et al. (2013). Anthropogenic impact and lead pollution throughout the Holocene in Southern Iberia. Science of the Total Environment, 449, 451460.Google Scholar
García-Artola, A., Cearreta, A., and Leorri, E. (2015). Relative sea-level changes in the Basque coast (northern Spain, Bay of Biscay) during the Holocene and Anthropocene: The Urdaibai estuary case. Quaternary International, 364, 172180.Google Scholar
García-Ruiz, J. M., Beguería, S., Nadal-Romero, E., González-Hidalgo, J. C., Lana-Renault, N., and Sanjuán, Y. (2015). A meta-analysis of soil erosion rates across the world. Geomorphology, 239, 160173.Google Scholar
Garrett, R. G. (2000). Natural sources of metals to the environment. Human and Ecological Risk Assessment, 6, 945963.Google Scholar
Garrett, T. L. (2014). Long-run evolution of the global economy: 1. Physical basis. Earth’s Future, 2, 127151.Google Scholar
Gasperi, J., Dris, R., Mirande-Bret, C., et al. (2015). First overview of microplastics in indoor and outdoor air. In 15th EuCheMS International Conference on Chemistry and the Environment. https://hal-enpc.archives-ouvertes.fr/hal-01195546/.Google Scholar
Gattuso, J.-P., Magnan, A., Billé, R., et al. (2015). Contrasting futures for ocean and society from different anthropogenic CO2 emissions scenarios. Science, 349(6243).Google Scholar
Gauthier-Lafaye, F, Holliger, P., and Blanc, P. L. (1996). Natural fission reactors in the Franceville basin, Gabon: A review of the conditions and results of a “critical event” in a geologic system. Geochimica et Cosmochimica Acta, 60, 48314852.Google Scholar
Gehrels, R. (2010a). Sea-level changes since the Last Glacial Maximum: An appraisal of the IPCC Fourth Assessment Report. Journal of Quaternary Science, 25, 2638.Google Scholar
Gehrels, W. R. (2010b). Late Holocene land- and sea-level changes in the British Isles: Implications for future sea-level predictions. Quaternary Science Reviews, 29, 16481660.Google Scholar
Gehrels, W. R., Roe, H. M., and Charman, D. J. (2001). Foraminifera, testate amoebae and diatoms as sea-level indicators in UK saltmarshes: A quantitative multiproxy approach. Journal of Quaternary Science, 16, 201220.Google Scholar
Gehrels, W. R., and Shennan, I. (2015). Sea level in time and space: Revolutions and inconvenient truths. Journal of Quaternary Science, 30, 131143.Google Scholar
Gehrels, W. R., and Woodworth, P. L. (2013). When did modern rates of sea-level rise start? Global and Planetary Change, 100, 263277.Google Scholar
Gemmell, N. J., Schwartz, M. K., and Robertson, B. C. (2004). Moa were many. Proceedings of the Royal Society London B, 271, S430S432.Google Scholar
Genty, D., and Massault, M. (1999). Carbon transfer dynamics from bomb-14C and δ13C time series of a laminated stalagmite from SW France: Modelling and comparison with other stalagmite records. Geochimica et Cosmochimica Acta, 63, 15371548.Google Scholar
Georgescu-Roegen, N. (1975). Energy and economic myths. Southern Economic Journal, 41(3), 347381. (reprinted as chapter 1 [1972] in Energy and Economic Myths: Institutional and Analytical Economic Essays. New York: Pergamon.)Google Scholar
Gerasimov, I. P. (1979). Anthropogene and its major problem. Boreas, 8, 2330.Google Scholar
Gerig, B. S., Chaloner, D. T., Janetski, D. J., et al. (2015). Congener patterns of persistent organic pollutants establish the extent of contaminant biotransport by pacific salmon in the Great Lakes. Environmental Science & Technology, 50(2), 554563.Google Scholar
GESAMP (2016). Sources, Fate and Effects of Microplastics in the Marine Environment: Part Two of a Global Assessment, ed. Kershaw, P. J. and Rochman, C. M.. (IMO/FAO/ UNESCO-IOC/UNIDO/WMO/IAEA/UN/UNEP/UNDP Joint Group). Rep. Stud. GESAMP No. 93. http://gesamp.org/data/gesamp/files/file_element/0c50c023936f7ffd16506be330b43c56/rs93e.pdf.Google Scholar
Gevao, B., Boyle, E. A., Carrasco, G. G., et al. (2016). Spatial and temporal distributions of polycyclic aromatic hydrocarbons in the Northern Arabian Gulf sediments. Marine Pollution Bulletin, 112(1), 218224.Google Scholar
Geyer, R., Jambeck, J. R., Lavender Law, K. (2017). Production, use, and fate of all plastics ever made. Science Advances, 3, e1700782. doi:10.1126/sciadv.1700782.Google Scholar
GFDL (Geophysical Fluid Dynamics Laboratory) (2017). Global warming and hurricanes: An overview of current research results. https://gfdl.noaa.gov/global-warming-and-hurricanes/ (accessed June 2017).Google Scholar
Gherardi, F., Audigane, P., and Gaucher, E. C. (2012). Predicting long-term geochemical alteration of wellbore cement in a generic geological CO2 confinement site: Tackling a difficult reactive transport modelling challenge. Journal of Hydrology, 420 –421, 340359.Google Scholar
Gibbard, P. L., and Head, M. J. (2009). IUGS ratification of the Quaternary System/Period and the Pleistocene Series/Epoch with a base at 2.58 Ma. Quaternaire, 20(4), 411412.Google Scholar
Gibbard, P. L., and Head, M. J. (2010). The newly-ratified definition of the Quaternary System/Period and redefinition of the Pleistocene Series/Epoch, and comparison of proposals advanced prior to formal ratification. Episodes, 33, 152158.Google Scholar
Gibbard, P. L., Head, M. J., and Walker, M. J. C. (2010). Formal ratification of the Quaternary system/period and the Pleistocene series/Epoch with a base at 2.58 Ma. Journal of Quaternary Science, 25, 96102.Google Scholar
Gibbard, P. L., Smith, A. G., Zalasiewicz, J. A., et al. (2005). What status for the Quaternary? Boreas, 34, 16.Google Scholar
Gibbard, P. L., and Walker, M. J. C. (2014). The term “Anthropocene” in the context of formal geological classification. In Waters, C. N., Zalasiewicz, J. A., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 2937.Google Scholar
Gibbs, S. J., Bown, P. R., Sessa, J. A., et al. (2006). Nannoplankton extinction and origination across the Paleocene-Eocene Thermal Maximum. Science, 314(5806), 17701773.Google Scholar
Gilbert, A., Flowers, G. E., Miller, G. H., et al. (2017). Projected demise of Barnes Ice Cap: Evidence of an unusually warm 21st century Arctic. Geophysical Research Letters, 44(6), 28102816.Google Scholar
Gilbert, G. K. (1877). The Geology of the Henry Mountains. Washington, DC: US Government Printing Office.Google Scholar
Gillings, M. R., and Paulsen, I. T. (2014). Microbiology of the Anthropocene. Anthropocene, 5, 18.Google Scholar
Gingerich, P. D. (2006). Environment and evolution through the Paleocene–Eocene thermal maximum. Trends in Ecology & Evolution, 21(5), 246253.Google Scholar
Giosan, L., Syvitski, J., Constantinescu, S., and Day, J. (2014). Climate change: Protect the world’s deltas. Nature, 516, 3133.Google Scholar
Giuliani, S., Bellucci, L. G., Çağatay, M. N., et al. (2017). The impact of the 1999 Mw 7.4 event in the İzmit Bay (Turkey) on anthropogenic contaminant (PCBs, PAHs and PBDEs) concentrations recorded in a deep sediment core. Science of the Total Environment, 590 –591, 799808.Google Scholar
Glacken, C. J. (1956). Changing ideas of the habitable world. In Thomas, W. L. Jr., ed., Man’s Role in Changing the Face of the Earth. University of Chicago Press, pp. 7992.Google Scholar
Glasser, N. F., Harrison, S., Jansson, K. N., et al. (2011). Global sea-level contribution from the Patagonian Icefields since the Little Ice Age maximum. Nature Geoscience, 4.5(May 2011), 303307.Google Scholar
Glikson, A. (2013). Fire and human evolution: The deep-time blueprints of the Anthropocene. Anthropocene, 3, 8992.Google Scholar
Goldberg, E. D. (1997). Plasticizing the seafloor: An overview. Environmental Science and Technology, 18, 195201.Google Scholar
Golledge, N. R., Levy, R. H., McKay, R. M., and Naish, T. R. (2017). East Antarctic ice sheet most vulnerable to Weddell Sea warming. Geophysical Research Letters, 44, 23432351.Google Scholar
Gomez, B., Cui, Y., Kettner, A. J., et al. (2009). Simulating changes to the sediment transport regime of the Waipaoa River driven by climate change in the twenty-first century. Global and Planetary Change, 67, 153166.Google Scholar
Gonzalez, R. O., Strekopytov, S., Amato, F., et al. (2016). New insights from zinc and copper isotopic compositions into the sources of atmospheric particulate matter from two major European cities. Environmental Science & Technology, 50(18), 98169824.Google Scholar
Goodman, D., and Chant, C. (eds.) (1999). European Cities & Technology: Industrial to Post-Industrial City. London: Routledge.Google Scholar
Gori, G. (1965). La Forestal: La tragedia del quebracho Colorado. Buenos Aires: Platina/Stilcograf.Google Scholar
Goto-Azuma, K., and Koerner, R. M. (2001). Ice core studies of anthropogenic sulfate and nitrate trends in the Arctic. Journal of Geophysical Research, 106 (D5), 49594969.Google Scholar
Gowlett, J. A. J. (2016). The discovery of fire by humans: A long and convoluted process. Philosophical Transactions of the Royal Society B, 371, 20150164.Google Scholar
Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G. (eds.) (2012). The Geological Time Scale 2012. Elsevier.Google Scholar
Gradstein, F. M., Ogg, J. G., Smith, A. G., et al. (2004). A new geologic time scale with special reference to the Precambrian and Neogene. Episodes, 27, 83100.Google Scholar
Gradstein, F. M., Ogg, J. G., Smith, A. G. (eds.) (2005). A Geologic Time Scale 2004. Cambridge University Press.Google Scholar
Grasso, D. N. (2000). Geological effects of underground nuclear testing. United States Geological Survey Open-File Report 00–176.Google Scholar
Greb, L., Saric, B., Seyfried, H., and Leinfelder, R. R. (1996). Ökologie und Sedimentologie eines rezenten Rampensystems an der Karibikküste von Panamá. Profil, 10.Google Scholar
Greenbaum, J. S., Blankenship, D. D., Young, D. A., et al. (2015). Ocean access to a cavity beneath Totten Glacier in East Antarctica. Nature Geoscience, 8, 294298.Google Scholar
Greenop, R., Foster, G. L., Wilson, P. A., and Lear, C. H. (2014). Middle Miocene climate instability associated with high-amplitude CO2 variability. Paleoceanography, 29, 845853.Google Scholar
Gregory, M. R. (2009). Environmental implications of plastic debris in marine settings—entanglement, ingestion, smothering, hangers-on, hitch-hiking and alien invasions. Philosophical Transactions of the Royal Society B, 364, 20132025.Google Scholar
Gregory, M. R., and Andrady, A. L. (2003). Plastics in the marine environment. In Andrady, A. L., ed., Plastics and the Environment. New Jersey: Wiley & Sons, pp. 379401.Google Scholar
Griffin, E. (2010). A Short History of the British Industrial Revolution. London: Palgrave.Google Scholar
Grimalt, J. O., Fernandez, P., Berdie, L., et al. (2001). Selective trapping of organochlorine compounds in mountain lakes of temperate areas. Environmental Science & Technology, 35, 26902697.Google Scholar
Grinevald, J. (1987). On a holistic concept for deep and global ecology: The Biosphere. Fundamenta Scientiae, 8(2), 197226.Google Scholar
Grinevald, J. (1988). Sketch for a history of the idea of the Biosphere. In Bunyard, P., and Goldsmith, E., eds., GAIA, the Thesis, the Mechanisms, and the Implications. Camelford, Cornwall, UK: Wadebridge Ecological Centre, pp. 134. (reprinted in Bunyard, P., ed., Gaia in Action: Science of the Living Earth. Edinburgh: Floris Books, 1996, pp. 34–53.)Google Scholar
Grinevald, J. (2007). La Biosphère de l’Anthropocène: Climat et pétrole, la double menace. Repères transdisciplinaires (1824–2007). Geneva, Switzerland: Georg/Editions Médecine & Hygiène.Google Scholar
Grossman, E. L. (2012). Oxygen isotope stratigraphy. In Gradstein, F. M., Ogg, J. G, Schmitz, M., and Ogg, G., eds., The Geological Time Scale 2012. Elsevier, pp. 181206.Google Scholar
Grunwald, S., Thompson, J. A., and Boettinger, J. L. (2011). Digital soil mapping and modeling at continental scales: Finding solutions for global issues. (SSSA 75th Anniversary Special Paper). Soil Science Society of America Journal, 75(4), 12011213.Google Scholar
Gutjahr, M., Ridgwell, A., Sexton, P. F., et al. (2017). Very large release of mostly volcanic carbon during the Palaeocene-Eocene Thermal Maximum. Nature, 548(7669), 573777.Google Scholar
Haff, P. K. (2012). Technology and human purpose: The problem of solids transport on the Earth’s surface. Earth System Dynamics, 3, 149156.Google Scholar
Haff, P. K. (2014a). Technology as a geological phenomenon: Implications for human well-being. In Waters, C. N., Zalasiewicz, J., Williams, M., eds., A Stratigraphical Basis for the Anthropocene. Geological Society London, Special Publication, 395, pp. 301309.Google Scholar
Haff, P. K. (2014b). Humans and technology in the Anthropocene: Six rules. Anthropocene Review, 1, 126136.Google Scholar
Haff, P. K. (2016). Purpose in the Anthropocene: Dynamical role and physical basis. Anthropocene, 16, 5460.Google Scholar
Haigh, J. D., and Cargill, P. (2015). The Sun’s Influence on Climate. Princeton Primers in Climate. New Jersey: Princeton University Press.Google Scholar
Hall, N. M., Berry, K. L. E., Rintoul, L., and Hoogenboom, M. O. (2015). Microplastic ingestion in scleractinian corals. Marine Biology, 162, 725732.Google Scholar
Halpern, B. S., Walbridge, S., Selkoe, K. A., et al. (2008). A global map of human impact on marine ecosystems. Science, 319, 948952.Google Scholar
Hamilton, C. (2017). Defiant Earth: The Fate of Humans in the Anthropocene. Cambridge, UK: Polity Books.Google Scholar
Hamilton, C., and Grinevald, J. (2015). Was the Anthropocene anticipated? Anthropocene Review, 2(1), 5972.Google Scholar
Hammarlund, E. U., Dahl, T. W., Harper, D. A. T., et al. (2012). A sulfidic driver for the end-Ordovician mass extinction. Earth and Planetary Science Letters, 331 –332, 128139.Google Scholar
Han, Y. M., An, Z. S., and Cao, J. J. (2017). The Anthropocene: A potential stratigraphic definition based on black carbon, char, and soot records. In DellaSala, D., and Goldstein, M. I., eds., Encyclopedia of the Anthropocene, vol. 1. Oxford: Elsevier. doi:10.1016/B978–0–12–409548–9.10001–6.Google Scholar
Han, Y. M., Wei, C., Huang, R. J., et al. (2016). Reconstruction of atmospheric soot history in inland regions from lake sediments over the past 150 years. Scientific Reports, 6, 19151.Google Scholar
Hancock, G. J., Leslie, C., Everett, S. E., Tims, S. G., Brunskill, G. J., and Haese, R. (2011). Plutonium as a chronomarker in Australian and New Zealand sediments: A comparison with 137Cs. Journal of Environmental Radioactivity, 102, 919929.Google Scholar
Hancock, G. J., Tims, S. G., Fifield, L. K., and Webster, I. T. (2014). The release and persistence of radioactive anthropogenic nuclides. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 265281.Google Scholar
Hancock, T. (2016). Healthcare in the Anthropocene: Challenges and opportunities. Health Care Quarterly, 19(3), 1722.Google Scholar
Hancock, T. (2017). Population health promotion in the Anthropocene. In Rootman, I., et al., eds., Health Promotion in Canada, 4th ed. Toronto: Canadian Scholars Press.Google Scholar
Hancock, T., Capon, A., Dietrich, U., and Patrick, R. (2016). Governance for health in the Anthropocene. International Journal of Health Governance, 21(4), 120.Google Scholar
Hancock, T., Spady, D. W., and Soskolne, C. (eds.) (2015). Global change and public health: Addressing the ecological determinants of health. http://cpha.ca/uploads/policy/edh-brief.pdf.Google Scholar
Hanebuth, T. J. J., Lantzsch, H., and Nizou, J. (2015). Mud depocenters on continental shelves—appearance, initiation times, and growth dynamics. Geo-Marine Letters, 35(6), 487503.Google Scholar
Hanna, E., Navarro, F. J., Pattyn, F., et al. (2013). Ice-sheet mass balance and climate change. Nature, 498, 5159.Google Scholar
Hansen, J., Sato, M., Hearty, P., et al. (2016). Ice melt, sea level rise and superstorms: Evidence from paleoclimate data, climate modeling, and modern observations that 2°C global warming could be dangerous. Atmospheric Chemistry and Physics, 16, 37613812.Google Scholar
Hansen, P. H. (2013). The Summits of Modern Man: Mountaineering after the Enlightenment. Cambridge, MA: Harvard University Press.Google Scholar
Hanvey, J. S., Lewis, P. J., Lavers, J. L., et al. (2017). A review of analytical techniques for quantifying microplastics in sediments. Analytical Methods, 9, 13691383.Google Scholar
Haraway, D. (2015). Anthropocene, Capitalocene, Plantationocene, Chthulucene: Making kin. Environmental Humanities, 6, 159165.Google Scholar
Harmand, S., Lewis, J. E., Feibel, C. S., et al. (2015). 3.3-million-year-old stone tools from Lomekwi 3, west Turkana, Kenya. Nature, 521, 310315.Google Scholar
Harris, E. C. (2014). Archaeological stratigraphy as a paradigm for the Anthropocene. Journal of Contemporary Archaeology, 1(1), 105109.Google Scholar
Harris, T. D., and Smith, V. H. (2016). Do persistent organic pollutants stimulate cyanobacterial blooms? Inland Waters, 6(2), 124130.Google Scholar
Harshvardhan, K., and Jha, B. (2013). Biodegradation of low-density polyethylene by marine bacteria from pelagic waters, Arabian Sea, India. Marine Pollution Bulletin, 77, 100106.Google Scholar
Hart, A. B., and Lawn, C. J. (1977). Combustion of coal and oil power station boilers. CEGB Research, 5, 417.Google Scholar
Harvey, M. C., Brassel, S. C., Belcher, C. M., and Montanari, A. (2008) Combustion of fossil organic matter at the Cretaceous-Paleogene (K-P) boundary. Geology, 36, 355358.Google Scholar
Harwood, T. D., Tomlinson, I., Potter, C. A., and Knight, J. D. (2011). Dutch elm disease revisited: Past, present and future management in Great Britain. Plant Pathology, 60, 545555.Google Scholar
Hastings, M. G., Jarvis, J. C., and Steig, E. J. (2009). Anthropogenic impacts on nitrogen isotopes of ice-core nitrate. Science, 324, 1288.Google Scholar
Haughton, S. (1865). Manual of Geology. Dublin, London: Longman & Co.Google Scholar
Hautmann, M. (2012). Extinction: End-Triassic Mass Extinction. In Encyclopedia of Life Sciences. Chichester: John Wiley & Sons Ltd. doi:10.1002/9780470015902.a0001655.pub3.Google Scholar
Hawkins, E., Ortega, P., Suckling, E., et al. (2017) Estimating changes in global temperature since the pre-industrial period. Bulletin of the American Meteorological Society. doi:10.1175/BAMS-D-16-0007.1.Google Scholar
Hawkins, W., and Wohletz, K. (1996). Visual inspection for CTBT verification. Los Alamos National Laboratory, OAC Project Number: ST484A.Google Scholar
Hay, W. W. (1994). Pleistocene-Holocene fluxes are not the earth’s norm. In Hay, W., ed., Global Sedimentary Geofluxes, 599. Washington: National Academy of Sciences Press, pp. 1527.Google Scholar
Hay, W. W. (2013). Experimenting on a Small Planet: A Scholarly Entertainment. London: Springer.Google Scholar
Haygarth, P. M., and Jones, K. C. (1992). Atmospheric deposition of metals to agricultural surfaces. In Biogeochemistry of Trace Elements. Boca Raton: Lewis Publishers, pp. 249276.Google Scholar
Hazen, R. M., Papineau, D., Bleeker, W., et al. (2008). Mineral evolution. American Mineralogist, 93, 16391720.Google Scholar
Hazen, R. M., and Ferry, J. M. (2010). Mineral evolution: Mineralogy in the fourth dimension. Elements, 6, 912.Google Scholar
Hazen, R. M., Grew, E. S., Origlieri, M. J., and Downs, R. T. (2017). On the mineralogy of the “Anthropocene Epoch.” American Mineralogist, 102, 595611.Google Scholar
Head, M. J., Aubry, M.-P., Walker, M., et al. (2017). A case for formalizing subseries (subepochs) of the Cenozoic Era. Episodes, 40(1), 2227.Google Scholar
Head, M. J., and Gibbard, P. L. (2015). Formal subdivision of the Quaternary System/Period: Past, present, and future. Quaternary International, 383, 435.Google Scholar
Heaney, P. J. (2017). Defining minerals in the age of humans. American Mineralogist, 102, 925926.Google Scholar
Heard, B. P., Brook, B. W., Wigley, T. M. L., and Bradshaw, C. J. A. (2017). Burden of proof: A comprehensive review of the feasibility of 100% renewable-electricity systems. Renewable and Sustainable Energy Reviews, 76, 11221133.Google Scholar
Heim, S., and Schwarzbauer, J. (2013). Pollution history revealed by sedimentary records: A review. Environmental Chemistry Letters, 11(3), 255270.Google Scholar
Heim, S., Schwarzbauer, J., Littke, R., et al. (2004). Geochronology of anthropogenic pollutants in riparian wetland sediments of the Lippe River (Germany). Organic Geochemistry, 35(11–12), 14091425.Google Scholar
Heinrich, H. (1988). Origin and consequences of cyclic ice rafting in the northeast Atlantic Ocean during the past 130,000 years. Quaternary Research, 29, 142142.Google Scholar
Heiss, G., and Leinfelder, R. R. (2008). “Fünf vor Zwölf”– Verschwinden die Riffe? In Leinfelder, R. R., Heiss, G., and Moldrzyk, U., eds., “Abgetaucht”: Begleitbuch zur Sonderausstellung zum Internationalen Jahr des Riffes. Leinfelden-Echterdingen: Konradin-Verlag, pp. 182197.Google Scholar
Hemming, N. G., Guilderson, T. P., and Fairbanks, R. G. (1998). Seasonal variations in the boron isotopic composition of coral: A productivity signal? Global Biogeochemical Cycles, 12, 581586.Google Scholar
Hemming, N. G., and Hanson, G. N. (1992). Boron isotope composition and concentration in modern marine carbonates. Geochimica Cosmochimica Acta, 56, 537543.Google Scholar
Hempelmann, A., and Weber, W. (2012). Correlation between the sunspot number, the total solar irradiance, and the terrestrial insolation. Solar Physics, 277, 417430.Google Scholar
Hendy, I. L., Napier, T. J., and Schimmelmann, A. (2015). From extreme rainfall to drought: 250 years of annually resolved sediment deposition in Santa Barbara Basin, California. Quaternary International, 387, 312.Google Scholar
Hennissen, J. A. I., Head, M. J., De Schepper, S., and Groeneveld, J. (2014). Palynological evidence for a southward shift of the North Atlantic Current at ~2.6 Ma during the intensification of late Cenozoic Northern Hemisphere glaciation. Paleoceanography, 28. doi:10.1002/ 2013PA002543.Google Scholar
Hennissen, J. A. I., Head, M. J., De Schepper, S., and Groeneveld, J. (2015). Increased seasonality during the intensification of Northern Hemisphere glaciation at the Pliocene–Pleistocene boundary ~2.6 Ma. Quaternary Science Reviews, 129, 321332.Google Scholar
Hennissen, J. A. I., Head, M. J., De Schepper, S., and Groeneveld, J. (2017). Dinoflagellate cyst paleoecology during the Pliocene–Pleistocene climatic transition in the North Atlantic. Palaeogeography, Palaeoclimatology, Palaeoecology, 470, 81108.Google Scholar
Henshilwood, C. S., d’Errico, F., and Watts, I. (2009). Engraved ochres from the Middle Stone Age levels at Blombos Cave, South Africa. Journal of Human Evolution, 57, 2747.Google Scholar
Henshilwood, C. S., d’Errico, F., Yates, R., et al. (2002). Emergence of modern human behaviour: Middle Stone Age engravings from South Africa. Science, 295, 12781280.Google Scholar
Heringman, N. (2015). Deep time and the dawn of the Anthropocene. Representations, 129, 5685.Google Scholar
Herz-Fischler, R. (2009). The Shape of the Great Pyramid. Waterloo, Canada: Wilfrid Laurier University Press.Google Scholar
Herzschuh, U., Birks, J. B., Laepple, T., et al. (2016). Glacial legacies on interglacial vegetation at the Pliocene-Pleistocene transition in NE Asia. Nature Communications, 7, 11967. doi:10.1038/ncomms11967.Google Scholar
Hetzinger, S., Pfeiffer, M., Dullo, W.-C., et al. (2010). Rapid 20th century warming in the Caribbean and impact of remote forcing on climate in the northern tropical Atlantic as recorded in a Guadeloupe coral. Palaeogeography, Palaeoclimatology, Palaeoecology, 296, 111124.Google Scholar
Hey, E. (2016). International law and the Anthropocene. European Society of International Law (ESIL) Reflections, 5(10), 17.Google Scholar
Hibbard, K. A., Crutzen, P. J., Lambin, E. F., et al. (2006). Decadal interactions of humans and the environment. In Costanza, R., Graumlich, L., and Steffen, W., eds., Integrated History and Future of People on Earth. Dahlem Workshop Report, 96, pp. 341375.Google Scholar
Hicks, S., and Isaksson, E. (2006). Assessing source areas of pollutants from studies of fly ash, charcoal, and pollen from Svalbard snow and ice. Journal of Geophysical Research, 111(D02113). doi:10.1029/2005JD006167.Google Scholar
Higgins, S. (2016). Review: Advances in delta-subsidence research using satellite methods. Hydrogeology Journal, 24, 587600.Google Scholar
Higgins, S., Overeem, I., Tanaka, A., and Syvitski, J. P. M. (2013). Land subsidence at aquaculture facilities in the Yellow River delta, China. Geophysical Research Letters, 40, 38983902.Google Scholar
Hilgard, E. W. (1860). Report of the Geology and Agriculture of the State of Mississippi. Jackson, MS: E Barksdale, State Printer.Google Scholar
Hilgen, F. J., Lourens, L. J., and Van Dam, J. A. (2012). The Neogene Period. In Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G., eds., The Geological Time Scale 2012. Elsevier, pp. 923978.Google Scholar
Hillenbrand, C.-D., Smith, J. A., Hodell, D. A., et al. (2017). West Antarctic Ice Sheet retreat driven by Holocene warm water incursions. Nature, 547, 4348.Google Scholar
Hobbs, W. R., Massom, R. A., Stammerjohn, S., et al. (2016). A review of recent changes in Southern Ocean sea ice, their drivers and forcings. Global Planetary Change, 143, 228250.Google Scholar
Hoegh-Guldberg, O. (1999). Climate change, coral bleaching and the future of the world’s coral reefs. Marine and Freshwater Research, 50, 839–66.Google Scholar
Hoegh-Guldberg, O., and Bruno, J. F. (2010). The impact of climate change on the world’s marine ecosystems. Science, 328, 15231528.Google Scholar
Hoegh-Guldberg, O., Hoegh-Guldberg, H., Veron, J. E. N., et al. (2009). The Coral Triangle and Climate Change: Ecosystems, People and Societies at Risk. Brisbane: WWF Australia. http://wwf.de/fileadmin/fm-wwf/Publikationen-PDF/climate_change_coral_triangle_summary_report.pdf.Google Scholar
Hoegh-Guldberg, O., Mumby, P. J., Hooten, A. J., et al. (2007). Coral reefs under rapid climate change and ocean acidification. Science, 318, 17371742.Google Scholar
Hoesly, R. M., Smith, S. J., Feng, L., et al. (2018). Historical (1750–2014) anthropogenic emissions of reactive gases and aerosols from the Community Emissions Data System (CEDS). Geosicence Model Development Discussions, 11(1), 369408.Google Scholar
Hoffmann, G., and Reicherter, K. (2014). Reconstructing Anthropocene extreme flood events by using litter deposits. Global Planet Change, 123, 2228.Google Scholar
Hoffman, J., Clark, P. U., Parnell, A. C., and He, F. (2017). Regional and global sea-surface temperatures during the last interglaciation. Science, 335(6322), 276279.Google Scholar
Hofmann, A. (1981). The ecostratigraphic paradigm. Lethaia, 14, 17.Google Scholar
Holland, P. R., and Kwok, R. (2012). Wind-driven trends in Antarctic sea-ice drift. Nature Geoscience, 5, 872875.Google Scholar
Holliday, V. T. (2004). Soils in Archaeological Research. Oxford University Press.Google Scholar
Holmes, A. (1913). The Age of the Earth, 1st ed. Harper & Brothers.Google Scholar
Holtgrieve, G. W., Schindler, D. E., Hobbs, W. O., et al. (2011). A coherent signature of anthropogenic nitrogen deposition to remote watersheds of the northern hemisphere. Science, 334, 15451548.Google Scholar
Hom, W., Risebrough, R. W., Soutar, A., and Young, D. R. (1974). Deposition of DDE and polychlorinated biphenyls in dated sediments of the Santa Barbara Basin. Science, 184, 11971199.Google Scholar
Hong, S., Candelone, J. P., Patterson, C. C., and Boutron, C. F. (1994). Greenland ice evidences of hemispheric pollution for lead two millennia ago by Greek and Roman civilizations. Science, 265, 18411843.Google Scholar
Hong, S., Candelone, J. P., Patterson, C. C., and Boutron, C. F. (1996). History of ancient copper smelting pollution during Roman and medieval times recording in Greenland ice. Science, 272(5259), 246.Google Scholar
Hong, S., Lee, J., Kang, D., et al. (2014). Quantities, composition and sources of beach debris in Korea from the results of nationwide monitoring. Marine Pollution Bulletin, 84, 2734.Google Scholar
Hönisch, B., Hemming, N. G., Archer, D., et al. (2009). Atmospheric carbon dioxide concentration across the Mid-Pleistocene transition. Science, 324, 15511554.Google Scholar
Hönisch, B., Ridgwell, A., Schmidt, D. N., et al. (2012). The geological record of ocean acidification. Science, 335, 10581063.Google Scholar
Hooke, R. L. (2000). On the history of humans as geomorphic agents. Geology, 28(9), 843846.Google Scholar
Hooke, R. L., Martín-Ducque, J. F., and Pedraza, J. (2012). Land transformation by humans: A review. GSA Today, 22(12), 410.Google Scholar
Hori, K., Tanabe, S., Saito, Y., et al. (2004). Delta initiation and Holocene sea-level change: Example from the Song Hong (Red River) delta, Vietnam. Sedimentary Geology, 164, 237249.Google Scholar
Horng, C.-S., Huh, C.-A., Chen, K.-H., et al. (2009). Air pollution history elucidated from anthropogenic spherules and their magnetic signatures in marine sediments offshore of Southwestern Taiwan. Journal of Marine Systems, 76, 468478.Google Scholar
Horton, R. (2013). Offline: Planetary health—a new vision for the post-2015 era. The Lancet, 382, 1012.Google Scholar
Hou, X.-G., Siveter, D. J., Siveter, D. J., et al. (2017). The Cambrian Fossils of Chengjiang, China: The Flowering of Early Animal Life, 2nd ed. Oxford: Wiley Blackwell.Google Scholar
Hounslow, M. W. (2018). Magnetic particulates as markers of fossil fuel burning. In DellaSala, D., and Goldstein, M. I., eds., Encyclopedia of the Anthropocene, vol. 1. Oxford: Elsevier.Google Scholar
Hovhannissian, R. H. (1994). Lake Sevan, Yesterday, Today. Erevan: Armenia National Academy of Sciences [in Russian, with extended summaries in English and Armenian].Google Scholar
Hovhannissian, R. H. (1996). Evolution of eutrophication processes of Lake Sevan and how to control them. In Lake Sevan: Problems and Strategies of Action, pp. 81–85. Proceeding of the international conference, Sevan, Armenia, 13–16 October 1996, Yeravan.Google Scholar
Howard, J. L. (2014). Proposal to add anthrostratigraphic and technostratigraphic units to the stratigraphic code for classification of anthropogenic Holocene deposits. Holocene, 24(12), 18561861.Google Scholar
Hu, B., Yang, Z., Wang, H., et al. (2009). Sedimentation in the Three Gorges Dam and the future trend of Changjiang (Yangtze River) sediment flux to the sea. Hydrology of Earth System Science, 13, 22532264.Google Scholar
Hu, Z., and Gao, S. (2008). Upper crustal abundances of trace elements: A revision and update. Chemical Geology, 253(3), 205221.Google Scholar
Hua, Q., Barbetti, M., and Rakowski, A. Z. (2013). Atmospheric radiocarbon for the period 1950–2010. Radiocarbon, 55(4), 20592072.Google Scholar
Hublin, J.-J., Ben-Ncer, A., Bailey, S. E., et al. (2017). New fossils from Jebel Irhoud, Morocco and the pan-African origin of Homo sapiens. Nature, 546, 289292.Google Scholar
Hughes, T. P., Baird, A. H., and Bellwood, D. R. (2003). Climate change, human impacts, and the resilience of coral reefs. Science, 301(5635), 929933.Google Scholar
Hughes, T. P., Barner, M. K., Bellwood, D. R., et al. (2017a). Coral reefs in the Anthropocene. Nature, 546, 8290.Google Scholar
Hughes, T. P., Kerry, J. T., Alvarez-Noriega, M., et al. (2017b). Global warming and recurrent mass bleaching of corals. Nature, 543(7645), 373377.Google Scholar
Hume, B. C. C., D’Angelo, C., Smith, E. G., et al. (2015). Symbiodinium thermophilum sp. nov., a thermotolerant symbiotic alga prevalent in corals of the world’s hottest sea, the Persian/Arabian Gulf. Scientific Reports, 5(8562). doi:10.1038/srep08562.Google Scholar
Humphreys, G. S., and Wilkinson, M. T. (2007). The soil production function: A brief history and its rediscovery. Geoderma, 139, 7378.Google Scholar
Hupy, J. P., and Schaetzl, R. J. (2006). Introducing “bombturbation,” a singular type of soil disturbance and mixing. Soil Science, 171(11), 823836.Google Scholar
Hussain, I., and Hamid, H. (2003). Plastics in agriculture. In Andrady, A. L., ed., Plastics and the Environment. Hoboken: Wiley, pp. 185209.Google Scholar
Hutton, J. (1795). Theory of the Earth with Proofs and Illustrations (in Four Parts), Edinburgh vol. 1, vol. 2, vol. 3. London: Geological Society, 1899.Google Scholar
Huybers, P., and Denton, G. (2008). Antarctic temperature at orbital timescales controlled by local summer duration. Nature Geoscience, 1, 787792.Google Scholar
Huybers, P., and Langmuir, C. (2009). Feedback between deglaciation, volcanism, and atmospheric CO2. Earth and Planetary Science Letters, 286(3–4), 479491.Google Scholar
ICOLD (2009). Sedimentation and sustainable use of reservoirs and river systems. ICOLD bulletin. CIGB-ICOLD.Google Scholar
Imhof, H. K., Ivleva, N. P., Schmid, J., et al. (2013). Contamination of beach sediments of a subalpine lake with microplastic particles. Current Biology, 23, R867–868.Google Scholar
Intergovernmental Panel on Climate Change (IPCC) (2013). Climate Change 2013: The Physical Science Basis, Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press.Google Scholar
International Council on Mining and Metals (ICMM) (2012). Trends in the mining and metals industry: Mining’s contribution to sustainable development, October 2012. London.Google Scholar
International Energy Agency (IEA) (2017). Key World Energy Statistics. http:// iea.org/statistics/.Google Scholar
International Law Association (ILA) (2016). Committee on International Law and Sea Level Rise: Interim Report. London: International Law Association.Google Scholar
International Union for the Conservation of Nature and Natural Resources (IUCN) (2014). The IUCN Red List of Threatened Species. Available at iucnredlist.org/.Google Scholar
Iozza, S., Müller, C. E., Schmid, P., et al. (2008). Historical profiles of chlorinated paraffins and polychlorinated biphenyls in a dated sediment core from Lake Thun (Switzerland). Environmental Science & Technology, 42(4), 10451050.Google Scholar
Irabien, M. J., García-Artola, A., Cearreta, A., and Leorri, E. (2015). Chemostratigraphic and lithostratigraphic signatures of the Anthropocene in estuarine areas from the eastern Cantabrian coast (N. Spain). Quaternary International, 364, 196205.Google Scholar
Ivar do Sul, J. A., and Costa, M. F. (2014). The present and future of microplastic pollution in the marine environment. Environmental Pollution, 185, 352364.Google Scholar
Ivar do Sul, J. A., Costa, M. F., Silva-Cavalcanti, J., and Araújo, M. C. B. (2014). Plastic debris retention and exportation by a mangrove forest patch. Marine Pollution Bulletin, 78, 252257.Google Scholar
Jackson, J. B. C. (1997). Reefs since Columbus. Corals Reefs, 16, 2332.Google Scholar
Jackson, M. D., Landis, E. N., Brune, P. F., et al. (2014). Mechanical resilience and cementitious processes in Imperial Roman architectural mortar. Proceedings of the National Academy of Sciences (USA), 111(52), 1848418489.Google Scholar
Jahan, K., Axe, L. B., Sandhu, N. K., et al. (2011). Heavy Metal Contamination in Highway Marking Glass Beads. New Jersey Department of Transportation.Google Scholar
Jambeck, J. R., Geyer, R., Wilcox, C., et al. (2015). Plastic waste inputs from land into the ocean. Science, 347, 768771.Google Scholar
James, L. A. (2013). Legacy sediment: Definitions and processes of episodically produced anthropogenic sediment. Anthropocene, 2, 1626.Google Scholar
Jamieson, A. J., Malkocs, T., Piertney, S. B., et al. (2017). Bioaccumulation of persistent organic pollutants in the deepest ocean fauna. Nature Ecology & Evolution, 1(0051).Google Scholar
Jansen, E., Overpeck, J., Briffa, K. R., et al. (2007). Palaeoclimate. In Solomon, S., Qin, D., Manning, M., et al., eds., Climate Change 2007: The Physical Science Basis; Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press, pp. 433497.Google Scholar
Jaspers, V. L. B., Megson, D., and O’Sullivan, G. (2014). POPs in the terrestrial environment. In O’Sullivan, G., and Sandau, C. D., eds., Environmental Forensics for Persistent Organic Pollutants. Amsterdam: Elsevier, pp. 291356.Google Scholar
Jeandel, C. (1981). Comportement du Plutonium dans les mileux naturels (Lacustre, Fluvial et Estuarien). Thèse 3° Cycle, Université de Paris VII.Google Scholar
Jeandel, C., and Oelkers, E. H. (2015). The influence of terrigenous particulate material dissolution on ocean chemistry and global element cycles. Chemical Geology, 395, 5066.Google Scholar
Jeker, P., and Krahenbühl, U. (2001). Sulfur profiles of the twentieth century in peat bogs of the Swiss Midlands measured by ICP-OES and by IC. Chimia, 55, 10291032.Google Scholar
Jenkyn, T. W. (1854a). Lessons in geology XLVI, chapter IV: On the effects of organic agents on the Earth’s crust. Popular Educator, 4, 139141.Google Scholar
Jenkyn, T. W. (1854b). Lessons in geology XLIX, chapter V: On the classification of rocks (section IV: On the tertiaries). Popular Educator, 4, 312316.Google Scholar
Jennings, N. S. (2011). Mining and quarrying. In Armstrong, J. R., and Menon, R., eds., Encyclopedia of Occupational Health and Safety. Geneva: International Labor Organization.Google Scholar
Jeong, S., Howat, I. M., and Bassis, J. N. (2016). Accelerated ice shelf rifting and retreat at Pine Island Glacier, West Antarctica. Geophysical Research Letters, 43, 1172011725.Google Scholar
Jevrejeva, S., Grinsted, A., Moore, J. C., and Holgate, S. J. (2006). Nonlinear trends and multiyear cycles in sea level records. Journal of Geophysical Research, 111, C09012.Google Scholar
Jevrejeva, S., Moore, J. C., Grinsted, A., and Woodworth, P. L. (2008). Recent global sea-level acceleration started over 200 years ago? Geophysical Research Letters, 35, L08715.Google Scholar
Jimenez, H., and Ruiz, G. M. (2016). Contribution of non-native species to soft-sediment marine community structure of San Francisco Bay, California. Biological Invasions, 18, 20072016.Google Scholar
Jinhui, L., Yuan, C., and Wenjing, X. (2017). Polybrominated diphenyl ethers in articles: A review of its applications and legislation. Environmental Science and Pollution Research, 24, 43124321.Google Scholar
Johannsson, O. E., Mills, E. L., and O’Gorman, R. (1991). Changes in the nearshore and offshore zooplankton communities in Lake Ontario: 1981–88. Canadian Journal of Fisheries and Aquatic Sciences, 48, 15461557.Google Scholar
Johnsen, S. I., and Taugbøl, T. (2010). NOBANIS – invasive alien species fact sheet: Pacifastacus leniusculus. Online Database of the European Network on Invasive Alien Species – NOBANIS. http://nobanis.org (accessed 23 June 2017).Google Scholar
Johnson, B. (2018). The reputed plague pits of London. The History Magazine. https://historic-uk.com/HistoryMagazine/DestinationsUK/LondonPlaguePits/ (accessed June 2018).Google Scholar
Johnson, L. (2017). Fisheries minister to announce protection for ancient glass sponge reefs: 9,000 year old reefs expected to be protected by new Marine Protected Area on BC Coast. CBCNews, 15 February 2017. http://cbc.ca/news/canada/british-columbia/leblanc-sponge-announcement-1.3984590Google Scholar
Johnson, L. C. (1995). China’s Pompeii: Twelfth century Dongjing. Historian, 58, 4968.Google Scholar
Jomelli, V., Favier, V., Vuille, M., et al. (2014). A major advance of tropical Andean glaciers during the Antarctic cold reversal. Nature, 513, 224228.Google Scholar
Joos, F., Prentice, I. C., Sitch, S., et al. (2001). Global warming feedbacks on terrestrial carbon uptake under the Intergovernmental Panel on Climate Change (IPCC) emission scenarios. Global Biogeochemical Cycles, 15, 891908.Google Scholar
Juarrero, A. (1999). Dynamics in Action: Intentional Behavior as a Complex System. Cambridge: MIT Press.Google Scholar
Juniper, T. (2013). What Has Nature Ever Done for Us? London: Profile Books.Google Scholar
Kabata-Pendias, A. (2010). Trace Elements in Soils and Plants, 4th ed. Boca Raton: CRC Press.Google Scholar
Kakonyi, G., and Ahmed, I. (2013). Cadmium and lead: Contamination. In Jørgensen, S. E., ed., Encyclopedia of Environmental Management. Boca Raton: CRC Press.Google Scholar
Kallenborn, R., Halsall, C., Dellong, M., and Carlsson, P. (2012). The influence of climate change on the global distribution and fate processes of anthropogenic persistent organic pollutants. Journal of Environmental Monitoring, 14(11), 28542869.Google Scholar
Kannan, K., Johnson-Restrepo, B., Yohn, S. S., et al. (2005). Spatial and temporal distribution of polycyclic aromatic hydrocarbons in sediments from Michigan inland lakes. Environmental Science & Technology, 39(13), 47004706.Google Scholar
Kaplan, M. R., Schaefer, J. M., Denton, G. H., et al. (2010). Glacier retreat in New Zealand during the Younger Dryas Stadial. Nature, 467, 194197.Google Scholar
Karkanas, P., Shahack-Gross, R., Ayalon, A., et al. (2007). Evidence for habitual use of fire at the end of the Lower Palaeolithic: Site-formation processes at Qesem Cave, Israel. Journal of Human Evolution, 53, 197212.Google Scholar
Kaser, G., Cogley, J. G., Dyurgerov, M. B., et al. (2006). Mass balance of glaciers and ice caps: Consensus estimates for 1961–2004. Geophysical Research Letters, 33, L19501. doi:10.1029/2006GL027511.Google Scholar
Kasirajan, S., and Ngouajio, M. (2012). Polyethylene and biodegradable mulches for agricultural applications: A review. Agronomy for Sustainable Development, 32, 501529.Google Scholar
Kaspari, S. D., Schwikowski, M., Gysel, M., et al. (2011). Recent increase in black carbon concentrations from a Mt. Everest ice core spanning 1860–2000 AD. Geophysical Research Letters, 38, L04703.Google Scholar
Kawamura, H., Matusoka, N., Momoshima, N., et al. (2006). Isotopic evidence in tree rings for historical changes in atmospheric sulfur sources. Environmental Science and Technology, 40, 57505754.Google Scholar
Keeling, C. D. (1960). The concentration and isotopic abundance of carbon dioxide in the atmosphere. Tellus, 12(2), 200203.Google Scholar
Keenan, T. F., Prentice, C., Canadell, J. G., et al. (2016). Recent pause in the growth rate of atmospheric CO2 due to enhanced terrestrial carbon uptake. Nature Communications, 7, 13428. doi:10.1038/ncomms13428.Google Scholar
Kelleher, J. (2016). Studying Anthropocene sedimentation behind a 19th century dam in western Connecticut. Keck Geology Consortium, Short Contributions, 29th Annual Symposium Volume.Google Scholar
Keller, G., Adatte, T., Stinnesbeck, W., et al. (2004). Chicxulub impact predates the K–T boundary mass extinction. Proceedings of the National Academy of Sciences (USA), 101, 37533758.Google Scholar
Kelly, A. E., Reuer, M. K., Goodkin, N. F., and Boyle, E. A. (2009). Lead concentrations and isotopes in corals and water near Bermuda, 1780–2000. Earth and Planetary Science Letters, 283, 93100.Google Scholar
Kelly, J. M., Scarpino, P., Berry, H., et al. (eds.) (2017). Rivers of the Anthropocene. University of California Press.Google Scholar
Kemp, A. C., Hawkes, A. D., Donnelly, J. P., et al. (2015). Relative sea-level change in Connecticut (USA) during the last 2200 yrs. Earth and Planetary Science Letters, 428, 217229.Google Scholar
Kennett, J. P. (1982). Marine Geology. New York: Prentice-Hall.Google Scholar
Kettner, A. J., Gomez, B., and Syvitski, J. P. M. (2007). Modeling suspended sediment discharge from the Waipaoa River system, New Zealand: The last 3000 years. Water Resources Research, 43, W07411. doi:10.1029/2006WR005570.Google Scholar
Kettner, A. J., and Syvitski, J. P. M. (2009). Fluvial responses to environmental perturbations in the Northern Mediterranean since the Last Glacial Maximum. Quaternary Science Reviews, 28, 23862397.Google Scholar
Key, R. M., Kozyr, A., Sabine, C. L., et al. (2004). A global ocean carbon climatology: Results from Global Data Analysis Project (GLODAP). Global Biogeochemical Cycles, 18, GB4031. doi:10.1029/2004GB002247.Google Scholar
Khan, A., and Lemmen, C. (2014). Bricks and urbanism in the Indus Valley rise and decline. Cornell University Library. arXiv:1303.1426v2 [physics.hist-ph]. Last revised 24 July 2014.Google Scholar
Khan, S. A., Sasgen, I., Bevis, M., et al. (2016). Geodetic measurements reveal similarities between post–Last Glacial Maximum and present-day mass loss from the Greenland ice sheet. Science Advances, 2, e1600931.Google Scholar
Kim, R. E., and Bosselmann, K. (2013). International environmental law in the Anthropocene: Towards a purposive system of multilateral environmental agreements. Transnational Environmental Law, 2, 285309.Google Scholar
Kinnard, C., Zdanowicz, C. M., Fisher, D. A., et al. (2011). Reconstructed changes in Arctic sea ice over the past 1450 years. Nature, 479, 509512.Google Scholar
Kinnard, C., Zdanowicz, C. M., Koerner, R. M., and Fisher, D. A. (2008). A changing Arctic seasonal ice zone: Observations from 1870–2003 and possible oceanographic consequences. Geophysical Research Letters, 35, L02507.Google Scholar
Kintisch, E. (2013). Can coastal marshes rise above it all? Science, 341, 480481.Google Scholar
Klee, R. J., and Graedel, T. E. (2004). Elemental cycles: A status report on human or natural dominance. Annual Review of Environment and Resources, 29, 69107.Google Scholar
Klein, G. D. (2015). The “ANTHROPOCENE”: What is its geological utility? (Answer: It has none!). Episodes, September 2015, 218.Google Scholar
Knoll, A. H., Walter, M. R., Narbonne, G. M., and Christie-Blick, N. (2006). The Ediacaran Period: A new addition to the geologic time scale. Lethaia, 39, 1330.Google Scholar
Knowlton, N. (2001). Sea urchin recovery from mass mortality: New hope for Caribbean coral reefs? Proceedings of the National Academy of Sciences (USA), 98(9), 48224824.Google Scholar
Knudsen, M. F., Seidenkrantz, M.-S., Jacoben, B. H., and Kuijpers, A. (2011). Tracking the Atlantic Multidecadal Oscillation through the last 8,000 years. Nature Communications, 2, 178.Google Scholar
Kobashi, T., Shindell, D. T., Kodera, K., et al. (2013). On the origin of multidecadal to centennial Greenland temperature anomalies over the past 800 yr. Climate of the Past, 9, 583596.Google Scholar
Koch, P. L., and Barnosky, A. D. (2006). Late Quaternary extinctions: State of the debate. Annual Review of Ecology, Evolution, and Systematics, 37, 215250.Google Scholar
Koch, P. L., Zachos, J. C., and Gingerich, P. D. (1992). Correlation between isotope records in marine and continental carbon reservoirs near the Paleocene/Eocene boundary. Nature, 358, 319322.Google Scholar
Koide, M., Griffin, J. J., and Goldberg, E. D. (1975). Records of plutonium fallout in marine and terrestrial samples. Journal of Geophysical Research, 80, 41534162.Google Scholar
Kominz, M. A., Browning, J. V., Miller, K. G., et al. (2008). Late Cretaceous to Miocene sea-level estimates from the New Jersey and Delaware coastal plain coreholes: An error analysis. Basin Research, 20(2), 211226.Google Scholar
Kondolf, G. M., Gao, Y., Annandale, G. W., et al. (2014). Sustainable sediment management in reservoirs and regulated rivers: Experiences from five continents. Earth’s Future, 2, 256280.Google Scholar
Konrad, H., Gilbert, L., Cornford, S. L., et al. (2017). Uneven onset and pace of ice dynamic imbalance in the Amundsen Sea Embayment, West Antarctica. Geophysical Research Letters. doi:10.1002/2016GL070733.Google Scholar
Kopp, R. E., Horton, B. P., Kemp, A. C., and Tebaldi, C. (2015). Past and future sea-level rise along the coast of North Carolina, USA. Climatic Change, 132, 693707.Google Scholar
Kopp, R. E., Kemp, A. C., Bittermann, K., et al. (2016). Temperature-driven global sea-level variability in the Common Era. Proceedings of the National Academy of Sciences (USA), 113, E1434E1441.Google Scholar
Kopp, R. E., Simons, F. J., Mitrovica, J. X., et al. (2009). Probabilistic assessment of sea level during the last interglacial stage. Nature, 462, 863868.Google Scholar
Kotzé, L. J. (2014). Rethinking global environmental law and governance in the Anthropocene. Journal of Energy & Natural Resources Law, 32, 121156.Google Scholar
Krachler, M., Zheng, J., Fisher, D., and Shotyk, W. (2009). Global atmospheric As and Bi contamination preserved in 3000 year old Arctic ice. Global Biogeochemical Cycles, 23, GB3011.Google Scholar
Kramer, N., Wohl, E. E., and Harry, D. L. (2011). Using ground penetrating radar to “unearth” buried beaver dams. Geology, 40, 4346.Google Scholar
Krause, J. C., Diesing, M., and Arlt, G. (2010). The physical and biological impact of sand extraction: A case study of the western Baltic Sea. Journal of Coastal Research, 51, 215226.Google Scholar
Krotkov, N. A., McLinden, C. A., Li, C., et al. (2016). Aura OMI observations of regional SO2 and NO2 pollution changes from 2005 to 2015. Atmospheric Chemistry and Physics, 16, 46054629.Google Scholar
Kubo, Y., Syvitski, J. P. M., and Tanabe, S. (2006). An application of the hydrological model HYDROTREND to the paleo-Tonegawa: Numerical estimates of sediment discharge for the last 13,000 years. Journal of the Geological Society of Japan, 112, 719729.Google Scholar
Kubota, K., Yokoyama, Y., Ishikawa, T., and Suzuki, A. (2015). A new method for calibrating a boron isotope paleo-pH proxy using massive Porites corals. Geochemistry, Geophysics, Geosystems, 16(9), 33333342.Google Scholar
Kummu, M., and Varis, O. (2011). The world by latitudes: A global analysis of human population, development level and environment across the north-south axis over the past half century. Applied Geography, 31, 495507.Google Scholar
Kunasek, S. A, Alexander, B., Steig, E. J., et al. (2010). Sulfate sources and oxidation chemistry over the past 230 years from sulfur and oxygen isotopes of sulfate in a West Antarctic ice core. Journal of Geophysical Research, 115, D18313. doi:10.1029/2010JD013846.Google Scholar
Kuoppamaa, M., Goslar, T., and Hicks, S. (2009). Pollen accumulation rates as a tool for detecting land-use changes in a sparsely settled boreal forest. Vegetation History and Archaeobotany, 18, 205217.Google Scholar
Kurashov, E. A., and Belyakov, V. P. (1987). Role of meiofauna in benthic communities of various types in the Lattgalian lakes. Hydrobiological Journal, 23, 4650.Google Scholar
Kurz, W. A., and Apps, M. J. (1999). A 70-year retrospective analysis of carbon fluxes in the Canadian forest sector. Ecological Applications, 9, 526547.Google Scholar
Lal, R., Reicosky, D. C., and Hanson, J. D. (2007). Evolution of the plow over 10,000 years and the rationale for no-till farming. Soil and Tillage Research, 93(1), 112.Google Scholar
Lambeck, K., Esat, T., and Potter, E. (2002). Links between climate and sea levels for the past three million years. Nature, 419, 199206.Google Scholar
Landing, E. (1994). Precambrian–Cambrian global stratotype ratified and a new perspective of Cambrian time. Geology, 22, 179182.Google Scholar
Landing, E., Geyer, G., Brasier, M. D., and Bowring, S. A. (2013). Cambrian evolutionary radiation: Context, correlation, and chronostratigraphy-overcoming deficiencies of the first appearance datum (FAD) concept. Earth Science Reviews, 123, 133172.Google Scholar
Landrigan, P. J., Fuller, R., Acosta, N. J. R., et al. (2017). The Lancet Commission on pollution and health. The Lancet, 391, 407408.Google Scholar
Larson, G., Karlsson, E. K., Perri, A., et al. (2012). Rethinking dog domestication by integrating genetics, archeology, and biogeography. Proceedings of the National Academy of Sciences (USA), 109, 88788883.Google Scholar
Latour, B. (2015). Face à Gaïa. Les Empêcheurs de Penser en Rond/La Découverte, Paris.Google Scholar
Latti, S. J. (1932). The lake bed of Sevan: Materials on the investigation of Lake Sevan and its basin. Tiflis.Google Scholar
Lavanchy, V. M. H., Gäggler, H. W., Schotterer, U., Schwikowski, M., and Baltensperger, U. (1999). Historical record of carbonaceous particle concentrations from a European high-alpine glacier (Colle Gnifetti, Switzerland). Journal of Geophysical Research, 104(D17), 2122721236.Google Scholar
Lavers, J. L., and Bond, A. L. (2017). Exceptional and rapid accumulation of anthropogenic debris on one of the world’s most remote and pristine islands. Proceedings of the National Academy of Sciences (USA), 114(23), 60526055.Google Scholar
Law, K. L., Morét-Ferguson, S. E., Goodwin, D. S., et al. (2014). Distribution of surface plastic debris in the Eastern Pacific Ocean from an 11-year data set. Environmental Science and Technology, 48, 47324738.Google Scholar
Lay, M. G. (1992). Ways of the World: A History of the World’s Roads and of the Vehicles That Used Them. New Bunswick, New Jersey: Rutgers University Press.Google Scholar
Le Quéré, C., Andrew, R. M., Canadell, J. P., et al. (2016). Global carbon budget 2016. Earth System Science Data, 8, 605649.Google Scholar
Le Quéré, C., Andrew, R. M., Friendlingstein, P., et al. (2018). Global carbon budget 2017. Earth System Science Data, 10, 405448.Google Scholar
Le Quéré, C., Moriarty, R., Andrew, R. M., et al. (2015). Global carbon budget 2014. Earth Systems Science Data, 7, 4785.Google Scholar
Leão, Z. M. A. N. (1982). Morphology, Geology and Developmental History of the Southernmost Coral Reefs of Western Atlantic, Abrolhos Bank, Brazil. PhD Dissertation, Rosenstiel School of Marine and Atmospheric Sciences. Florida: University of Miami.Google Scholar
Leão, Z. M. A. N., and Kikuchi, R. (2001). The Abrolhos reefs of Brazil. Ecological Studies, 144, 8392.Google Scholar
Lechner, A., Keckeis, H., Lumesberger-Loisl, F., and Zens, B. (2014). The Danube so colourful: A potpourri of plastic litter outnumbers fish larvae in Europe’s second largest river. Environmental Pollution, 188, 177181.Google Scholar
Lee, J.-M., Eltgroth, S. F., Boyle, E. A., and Adkins, J. F. (2017). The transfer of bomb radiocarbon and anthropogenic lead to the deep North Atlantic Ocean observed from a deep sea coral. Earth and Planetary Science Letters, 458, 223232.Google Scholar
Lee, S.-H., Povinec, P. P., Wyse, E., et al. (2005). Distribution and inventories of 90Sr, 137Cs, 241Am and Pu isotopes in sediments of the Northwest Pacific Ocean. Marine Geology, 216(4), 249263.Google Scholar
Legovich, N. A. (1979). “Blooms” in the water of Lake Sevan (Observations 1964–1972). Materials of Sevan Hydrobiological Station, 17, 5174 [in Russian].Google Scholar
Leinfelder, R. R. (1997). Coral reefs and carbonate platforms within a siliciclastic setting: General aspects and examples from the Late Jurassic of Portugal. Proceedings of the 8th International Coral Reef Symposium, 2, 17371742.Google Scholar
Leinfelder, R. R. (2001). Jurassic reef ecosystems. In Stanley, G. D. Jr., ed., The History and Sedimentology of Ancient Reef Systems. Topics in Geobiology Series, 17, pp. 251309.Google Scholar
Leinfelder, R. R. (2017). Das Zeitalter des Anthropozäns und die Notwendigkeit der grossen Transformation: Welche Rollen spielen Umweltpolitik und Umweltrecht? Zeitschrift für Umweltrecht (ZUR), 28(5), 259266, Nomos. http://zur.nomos.de/fileadmin/zur/doc/Aufsatz_ZUR_17_05.pdf.Google Scholar
Leinfelder, R. R., and Haum, R. (2015). Ozeane. In Kersten, J., ed., Inwastement: Abfall in Umwelt und Gesellschaft. Bielefeld: Transcript-Verlag, pp. 153179.Google Scholar
Leinfelder, R. R., and Leão, Z. M. A. N. (2000). Increasing Reef Complexity-Decreasing Reef Flexibility Through Time, and a Unique Exception – The Evolutionary Relic Reefs of Brazil. Rio de Janeiro, Brazil: 31st International Geological Congress, abstract vol.Google Scholar
Leinfelder, R. R., and Nose, M. (1999). Increasing complexity - decreasing flexibility: A different perspective of reef evolution through time. Profil, 17, 135147 (Univ. Stuttgart).Google Scholar
Leinfelder, R. R., and Schmid, D. U. (2000). Mesozoic reefal thrombolites and other microbolites. In Riding, R., ed., Microbial Sediments. Berlin: Springer, pp. 289294.Google Scholar
Leinfelder, R. R., Schmid, D. U., Nose, M., and Werner, W. (2002). Jurassic reef patterns: The expression of a changing globe. In Flügel, E., Kiessling, W., and Golonka, J., eds., Phanerozoic Reef Patterns. Tulsa: SEPM Special Publication, 72, pp. 465520.Google Scholar
Leinfelder, R. R., Seemann, J., Heiss, G. A., and Struck, U. (2012). Could “ecosystem atavisms” help reefs to adapt to the Anthropocene? Proceedings of the 12th International Coral Reef Symposium, Cairns, Australia, 9–13 July 2012. 2B Coral reefs: Is the past the key to the future? http://icrs2012.com/proceedings/manuscripts/ICRS2012_2B_2.pdf.Google Scholar
Leinfelder, R. R., and Wilson, R. C. L. (1998). Third order sequences in an Upper Jurassic rift-related second order sequence, Central Lusitanian Basin, Portugal. In Graciansky, P.-C. de, Hardenbol, J., Jacquin, T., and Vail, P., eds., Mesozioc-Cenozoic Sequence Stratigraphy of European Basins. Tulsa: SEPM Special Publication, 60, pp. 507525.Google Scholar
Lenton, T. (2016). Earth System Science: A Very Short Introduction. Oxford University Press.Google Scholar
Lenton, T., and Watson, A. (2011). Revolutions That Made the Earth. Oxford University Press.Google Scholar
Letcher, T. M. (2016). Climate Change: Observed Impacts on Planet Earth, 2nd ed. Elsevier.Google Scholar
Levy, R., Harwood, D., Florindo, F., et al. (2016). Antarctic ice sheet sensitivity to atmospheric CO2 variations in the early to mid-Miocene. Proceedings of the National Academy of Sciences (USA), 113(13), 34533458.Google Scholar
Lewin, J. (2012). Enlightenment and the GM floodplain. Earth Surface Processes and Landforms, 38, 1729.Google Scholar
Lewis, S. L., and Maslin, M. A. (2015). Defining the Anthropocene. Nature, 519, 171180.Google Scholar
Lindahl, P., Asami, R., Iryu, Y., et al. (2011). Sources of plutonium to the tropical Northwest Pacific Ocean (1943–1999) identified using a natural coral archive. Geochemica et Cosmochimica Acta, 75, 13461356.Google Scholar
Lirman, D., Schopmeyer, S., Galvan, V., et al. (2014). Growth dynamics of the threatened Caribbean staghorn coral Acropora cervicornis: Influence of host genotype, symbiont identity, colony size, and environmental setting. PLoS ONE, 9(9), e107253. http://doi.org/10.1371/journal.pone.0107253.Google Scholar
Lisiecki, L., and Raymo, M. E. (2005). A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography, 20, PA1003. doi:10.1029/2004PA001071.Google Scholar
Lithner, D., Larsson, A., Dave, G. (2011). Environmental and health hazard ranking and assessment of plastic polymers based on chemical composition. Science of the Total Environment, 409, 33093324.Google Scholar
Liu, W., Xie, S.-P., Liu, A., and Zhu, J. (2017). Overlooked possibility of a collapsed Atlantic Meridional Overturning Circulation in warming climate. Science Advances, 3(1), e1601666. doi:10.1126/sciadv.1601666.Google Scholar
Livingston, H. D., and Povinec, P. P. (2000). Anthropogenic marine radioactivity. Ocean & Coastal Management, 43(8–9), 689712.Google Scholar
Livingston, H. D., Povinec, P. P., Ito, T., et al. (2001). The behaviour of plutonium in the Pacific Ocean. In Kudo, A., ed., Radioactivity in the Environment, vol. 1, Plutonium in the Environment. Amsterdam: Elsevier, pp. 267292.Google Scholar
Loader, N. J., Young, G. H. F., Grudd, H., and McCarroll, D. (2013). Stable carbon isotopes from Torneträsk, northern Sweden provide a millennial length reconstruction of summer sunshine and its relationship to Arctic circulation. Quaternary Science Reviews, 62, 97113.Google Scholar
Lobelle, D., and Cunliffe, M. (2011). Early microbial biofilm formation on marine plastic debris. Marine Pollution Bulletin, 62, 197200.Google Scholar
Loch, K., Loch, W., and Anlauf, H. (2007). Der Zustand der Steinkorallen in maledivischen Riffen und die Regeneration nach dem 1998er Korallenbleichen. Bufus, 37. http://bufus.sbg.ac.at/Info/Info37/Info37–2.htm.Google Scholar
Loch, K., Loch, W., Schuhmacher, H., and See, W. R. (2002). Coral recruitment and regeneration on a Maldivian reef 21 months after the coral bleaching event of 1998. Marine Ecology, 23, 219236Google Scholar
Long, A. J., Barlow, N. L. M., Gehrels, W. R., et al. (2014). Contrasting records of sea-level change in the eastern and western North Atlantic during the last 300 years. Earth and Planetary Science Letters, 388, 110122.Google Scholar
Lopes-Lima, M., Sousa, R., Geist, J., et al. (2016). Conservation status of freshwater mussels in Europe: State of the art and future challenges. Biological Reviews, 92, 572607.Google Scholar
Lorgeoux, C., Moilleron, R., Gasperi, J., et al. (2016). Temporal trends of persistent organic pollutants in dated sediment cores: Chemical fingerprinting of the anthropogenic impacts in the Seine River basin, Paris. Science of the Total Environment, 541, 13551363.Google Scholar
Lough, J. M., and Barnes, D. J. (2000). Environmental controls on growth of the massive coral Porites. Journal of Experimental Marine Biology and Ecology, 245(2), 225243.Google Scholar
Loulergue, L., Schilt, A., Spahni, R., et al. (2008). Orbital and millennial-scale features of atmospheric CH4 over the past 800,000 years. Nature, 453, 383386.Google Scholar
Lovejoy, S. (2014a). Scaling fluctuation analysis and statistical hypothesis testing of anthropogenic warming, Climate Dynamics, 42, 23392351.Google Scholar
Lovejoy, S. (2014b). Return periods of global climate fluctuations and the pause. Geophysical Research Letters, 41, 47044710.Google Scholar
Lovejoy, S. (2015). Climate Closure. EOS, 96. doi:10.1029/2015EO037499.Google Scholar
Lovelock, J. E., and Margulis, L. (1974). Atmospheric homeostasis by and for the biosphere: The Gaia hypothesis. Tellus, 26, 210.Google Scholar
Lowenthal, D. (2000). George Perkins Marsh: Prophet of Conservation. Seattle & London: University of Washington Press.Google Scholar
Lu, Z., Letcher, R. J., Chu, S., et al. (2015). Spatial distributions of polychlorinated biphenyls, polybrominated diphenyl ethers, tetrabromobisphenol A and bisphenol A in Lake Erie sediment. Journal of Great Lakes Research, 41(3), 808817.Google Scholar
Lüdecke, C., and Summerhayes, C. P. (2012). The Third Reich in Antarctica: The Third German Antarctic Expedition 1938–39. Bluntisham Books & Erskine Press.Google Scholar
Lundstrom, M. (2003). Moore’s law forever? Science, 299, 210211.Google Scholar
Lunt, D. J., Haywood, A. M., Schmidt, G. A., et al. (2012). On the causes of mid-Pliocene warmth and polar amplification. Earth and Planetary Science Letters, 321 –322, 128138.Google Scholar
Lüthi, D., Le Floch, M., Bereiter, B., et al. (2008). High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature, 453, 379382.Google Scholar
Lyon, T. L., and Buckman, H. O. (1946). The Nature and Properties of Soils. New York: Macmillan Co.Google Scholar
MA (Millennium Ecosystem Assessment) (2005) Ecosystems and Human Well-Being: Synthesis. Washington, DC: Island Press.Google Scholar
Ma, Z., Melville, D. S., Liu, J., et al. (2014). Rethinking China’s new great wall. Science, 346(6212), 912914.Google Scholar
MacAyeal, D. R. (1993). Binge/purge oscillations of the Laurentide ice sheet as a cause of the North Atlantic’s Heinrich Events. Paleoceanography, 8(6), 775784.Google Scholar
MacFarling Meure, C., Etheridge, D. E., Trudinger, C., et al. (2006). Law Dome CO2, CH4 and N2O ice core records extended to 2000 years BP. Geophysical Research Letters, 33(14), L14810. doi:10.1029/2006GL026152.Google Scholar
Mackintosh, A. N., Anderson, B. M., Lorrey, A. M., et al. (2017). Regional cooling caused recent New Zealand glacier advances in a period of global warming. Nature Communications. doi:10.1038/ncomms14202.Google Scholar
Maher, B. A. (1986). Characterisation of soils by mineral magnetic measurements. Physics of the Earth and Planetary Interiors, 42, 7692.Google Scholar
Maher, B. A., and Thompson, R. (1992). Paleoclimatic significance of the mineral magnetic record of the Chinese loess and paleosols. Quaternary Research, 37(2), 155170.Google Scholar
Mahiques, M. M., Hanebuth, T. J. J., Martins, C. C., et al. (2016). Mud depocentres on the continental shelf: A neglected sink for anthropogenic contaminants from the coastal zone. Environmental Earth Sciences, 75(44). doi:10.1007/s12665–015–4782-z.Google Scholar
Mahmood, K. (1987). Reservoir sedimentation: Impact, extent and mitigation. World Bank Technical Paper, 71.Google Scholar
Malakoff, D. (2002). Trawling’s a drag for marine life, say studies. Science, 298, 2123.Google Scholar
Malm, A., and Hornborg, A. (2014). The geology of mankind? A critique of the Anthropocene narrative. Anthropocene Review, 1, 6269.Google Scholar
Mann, M. E., Miller, S. K., Rahmstorf, S., et al. (2017). Record temperature streak bears anthropogenic fingerprint. Geophysical Research Letters, 44. doi:10.1002/2017GL074056.Google Scholar
Mao, J., Ribes, A., Yan, B., et al. (2016). Human-induced greening of the northern extratropical land surface. Nature Climate Change, 6, 959964.Google Scholar
Marchant, J. (2017). Biologists rush to study creatures living beneath Larsen C ice shelf before they disappear. Nature, 549, 443.Google Scholar
Marcott, S. A., Bauska, T. K., Buizert, C., et al. (2014). Centennial-scale changes in the global carbon cycle during the last deglaciation. Nature, 514, 616619.Google Scholar
Marcott, S. A., Shakun, J. D., Clark, P. U., and Mix, A. (2013). A reconstruction of regional and global temperature for the past 11,300 years. Science, 339(6124), 11981201.Google Scholar
Mari, M., and Domingo, J. L. (2010). Toxic emissions from crematories: A review. Environment International, 36(1), 131137.Google Scholar
Marie, B., Zanella-Cléon, I., Guichard, N., et al. (2011). Novel proteins from the calcifying shell matrix of the Pacific oyster Crassostrea gigas. Marine Biotechnology, 13, 11591168.Google Scholar
Marini, F. (2003). Natural microtektites versus industrial glass beads: An appraisal of contamination problems. Journal of Non-Crystalline Solids, 323(1), 104110.Google Scholar
Markosyan, A. G. (1959). Benthos production in Lake Sevan. In Trudy VI Soveshchaniya po Problemam Biologii Vnutrennikh Vod. Moscow, Leningrad, pp. 139145 [in Russian].Google Scholar
Marsh, G. P. (1864). Man and Nature; Or, Physical Geography as Modified by Human Action. New York: Charles Scribner (reprinted by Lowenthal, D., ed., Cambridge, MA: Belknap Press/Harvard University Press, 1965).Google Scholar
Marsh, G. P. (1874). The Earth as Modified by Human Action: A New Edition of “Man and Nature” Charles Scribner. New York: Armstrong & Co.Google Scholar
Martín, J., Puig, P., Palanques, A., et al. (2015). Commercial bottom trawling as a driver of sediment dynamics and deep seascape evolution in the Anthropocene. Anthropocene, 7, 115.Google Scholar
Martin, K., and Sauerborn, J. (2013). Agroecology. Netherlands: Springer.Google Scholar
Martinez, A., Hadnott, B. N., Awad, A. M., et al. (2017). Release of airborne polychlorinated biphenyls from New Bedford Harbor results in elevated concentrations in the surrounding air. Environmental Science & Technology Letters, 4(4), 127131.Google Scholar
Martinez, J. M., Guyot, J. L., Filizola, N., and Sondag, F. (2009). Increase in suspended sediment discharge of the Amazon River assessed by monitoring network and satellite data. Catena, 79, 257264.Google Scholar
Martínez-García, B., Pascual, A., Baceta, J. I., and Murelaga, X. (2013). Estudio de los foraminíferos bentónicos del “beach-rock” de Azkorri (Getxo, Bizkaia). Geogaceta, 53, 2932.Google Scholar
Martinez-Porchas, M., and Martinez-Cordova, L. R., 2012. World aquaculture: Environmental impacts and troubleshooting alternatives. Scientific World Journal, 2012, 389623.Google Scholar
Marx, S. K., Rashid, S., and Stromsoe, N. (2016). Global-scale patterns in anthropogenic Pb contamination reconstructed from natural archives. Environmental Pollution, 213, 283298.Google Scholar
Maselli, V., and Trincardi, F. (2013). Man made deltas. Scientific Reports, 3, 1926.Google Scholar
Masiokas, M. H., Rivera, A., Espizua, L. E., et al. (2009). Glacier fluctuations in extratropical South America during the past 1000 years. Palaeogeography, Palaeoclimatology, Palaeoecology, 281, 242268.Google Scholar
Masson-Delmotte, V., Schulz, M., Abe-Ouchi, A., et al. (2013). Information from paleoclimate archives. In Stocker, T. F., Qin, D., Plattner, G. K., et al., eds., Climate Change 2013: The Physical Science Basis, Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press, pp. 383464.Google Scholar
Masson-Delmotte, V., Swingedouw, D., Landais, A., et al. (2012). Greenland climate change: From the past to the future. Wiley Interscience Reviews: Climate Change, 3(5), 427449.Google Scholar
Massos, A., and Turner, A. (2017). Cadmium, lead and bromine in beached microplastics. Environmental Pollution, 227, 139145.Google Scholar
Mathew, W. M. (1970). Peru and the British guano market, 1840–1870. Economic History Review 23(1), 112128.Google Scholar
Matisoo-Smith, E., Roberts, R. M., Irwin, G. J., et al. (1998). Patterns of prehistoric human mobility in Polynesia indicated by mtDNA from the Pacific rat. Proceedings of the National Academy of Sciences (USA), 95, 1514515150.Google Scholar
Mattey, D. M., Lowry, D., Duffet, J., et al. (2008). A 53 year seasonally resolved oxygen and carbon isotope record from a modern Gibraltar speleothem: Reconstructed drip water and relationship to local precipitation. Earth and Planetary Science Letters, 269, 8095.Google Scholar
Mayewski, P. A., Carleton, A. M., Birkel, S. D., et al. (2017). Ice core and climate reanalysis analogs to predict Antarctic and southern hemisphere climate changes. Quaternary Science Reviews, 155, 5066.Google Scholar
McCauley, D. J., Pinsky, M. L., Palumbi, S. R., et al. (2015). Marine defaunation: Animal loss in the global ocean. Science, 347, 1255641.Google Scholar
McConnell, J. R., and Edwards, R. (2008). Coal burning leaves toxic heavy metal legacy in the Arctic. Proceedings of the National Academy of Sciences (USA), 105(34), 1214012144.Google Scholar
McConnell, J. R., Edwards, R., Kok, G. L., et al. (2007). 20th-century industrial black carbon emissions altered Arctic climate forcing. Science, 317, 13811384.Google Scholar
McConnell, J. R., Kipfstuhl, S., and Fischer, H. (2006). The NGT and PARCA shallow ice core arrays in Greenland: A brief overview. PAGES News, 14(1), 1314.Google Scholar
McCormick, A., Hoellin, T. J., Mason, S. A., et al. (2014). Microplastic is an abundant and distinct microbial habitat in an urban river. Environmental Science and Technology, 48, 1186311871.Google Scholar
McDougall, I., Brown, F. H., and Fleagle, J. G. (2005). Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature, 433, 733736.Google Scholar
McFarlane, D. A., Lundberg, J., and Neff, H. (2014). A speleothem record of early British and Roman mining at Charterhouse, Mendip, England. Archaeometry, 56(3), 431443.Google Scholar
McGann, M., Grossman, E. E., Takesue, R. K., et al. (2012). Arrival and expansion of the invasive foraminifera Trochammina hadai Uchio in Padilla Bay, Washington. Northwest Science, 86, 926.Google Scholar
McGann, M., Sloan, D., and Cohen, A. N. (2000). Invasion by a Japanese marine microorganism in western North America. Hydrobiologia, 421, 2530.Google Scholar
McGann, M., Sloan, D., and Wan, E. (2002). Biostratigraphy beneath central San Francisco Bay along the San Francisco-Oakland Bay Bridge transect. In Crustal Structure of the Coastal and Marine San Francisco Bay Region, California, pp. 11–28.Google Scholar
McInerney, F. A., and Wing, S. L. (2011). The Paleocene-Eocene Thermal Maximum: A perturbation of carbon cycle, climate, and biosphere with implications for the future. Annual Review of Earth and Planetary Sciences, 39, 489516.Google Scholar
McKee, K., Rogers, K., and Saintilan, H. (2012). Response of salt marsh and mangrove wetlands to changes in atmospheric CO2, climate, and sea level. In Middleton, B. A., ed., Global Change and the Function and Distribution of Wetlands, Dordrecht: Springer, pp. 6396.Google Scholar
McLean, D. (1991). Magnetic spherules in recent lake sediments. Hydrobiologica, 214, 9197.Google Scholar
McLinden, C. A., Fioletov, V., Shephard, M. W., et al. (2016). Space-based detection of missing sulfur dioxide sources of global air pollution. Nature Geoscience, 9, 496500.Google Scholar
McMichael, A. (1993). Planetary Overload. Cambridge University Press.Google Scholar
McMillan, A. A., and Powell, J. H. (1993). BGS Rock Classification Scheme: The classification of artificial (man-made) ground and natural superficial deposits. Version 2. British Geological Survey Technical Report, WG/93/46/R.Google Scholar
McMillan, A. A., and Powell, J. H. (1999). BGS Rock Classification Scheme: Classification of artificial (man-made) ground and natural superficial deposits; Applications to geological maps and datasets in the UK, vol. 4. British Geological Survey Research Report, RR/99/004.Google Scholar
McMinn, A., Hallegraeff, G. M., Thomson, P., et al. (1997). Cyst and radionucleotide evidence for the recent introduction of the toxic dinoflagellate Gymnodinium catenatum into Tasmanian waters. Marine Ecology Progress Series, 161, 165172.Google Scholar
McNeely, J. (2001). Invasive species: A costly catastrophe for native biodiversity. Land Use and Water Resources Research, 1, 110.Google Scholar
McNeill, J., Barrie, F. R., Buck, W. R., et al. (2012). International Code of Nomenclature for algae, fungi, and plants (Melbourne Code). Regnum Vegetabile 154. Koeltz Scientific Books. ISBN 978-3-87429-425-6.Google Scholar
McNeill, J. R. (2001). Something New under the Sun: An Environmental History of the Twentieth-Century World. New York: W. W. Norton & Company.Google Scholar
McNeill, J. R., and Engelke, P. (2016). The Great Acceleration: An Environmental History of the Anthropocene since 1945. Cambridge, MA: Harvard University Press.Google Scholar
McPherron, S., Alemseged, Z., Marean, C. W., et al. (2010). Evidence for stone-tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature, 466, 857860.Google Scholar
Mee, L. (2012). Between the devil and the deep blue sea: The coastal zone in an era of globalisation. Estuarine, Coastal and Shelf Science, 96, 18.Google Scholar
Meehl, G. A., Arblaster, J. M., Bitz, C. M., et al. (2016). Antarctic sea-ice expansion between 2000 and 2014 driven by tropical Pacific decadal climate variability. Nature Geoscience, 9, 590595.Google Scholar
Meharg, A. A., and Killham, K. (2003). A pre-industrial source of dioxins and furans. Nature, 421, 909910.Google Scholar
Mehta, P. K., and Monteiro, P. J. M. (2006). Concrete: Microstructure, properties, and materials, 3rd ed. New York: McGraw-Hill.Google Scholar
Meier, K. J. S., Beaufort, L., Heussner, S., and Ziveri, P. (2014). The role of ocean acidification in Emiliania huxleyi coccolith thinning in the Mediterranean Sea. Biogeosciences, 11, 28572869.Google Scholar
Melchin, M. J., Sadler, P. M., and Cramer, B. D. (2012). The Silurian Period. In Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G., eds., The Geological Time Scale 2012. Elsevier, pp. 526558.Google Scholar
Mernild, S. H., Lipscomb, W. H., Bahr, D. B., et al. (2013). Global glacier retreat: A revised assessment of committed mass losses and sampling uncertainties. The Cryosphere Discussions, 7, 19872005.Google Scholar
Merrill, G. P. (1897). A Treatise on Rocks: Rock-Weathering and Soils. New York: Macmillan Co.Google Scholar
Merritts, D., Walter, R., Rahnis, M., et al. (2011). Anthropocene streams and base-level controls from historic dams in the unglaciated mid-Atlantic region, USA. Philosophical Transactions of the Royal Society A, 369, 9761009.Google Scholar
Meshkova, T. M. (1976). Eutrophication of Lake Sevan. Biological Journal of Armenia, 29, 1422 [in Russian].Google Scholar
Messager, E., Lordkipanidze, D., Kvavadze, E., et al. (2010). Palaeoenvironmental reconstruction of Dmanisi site (Georgia) based on palaeobotanical data. Quaternary International, 223 –224, 2027.Google Scholar
Meybeck, M. (2003). Global analysis of river systems: From Earth System controls to Anthropocene syndromes. Philosophical Transactions of the Royal Society B, 358, 19351955.Google Scholar
Meybeck, M., Laroche, L., Darr, H. H., and Syvitski, J. P. M. (2003). Global variability of total suspended solids and their fluxes in rivers. Global and Planetary Change, 39(1/2), 6593.Google Scholar
Michelutti, N., Simonetti, A., Briner, J. P., et al. (2009). Temporal trends of pollution Pb and other metals in east-central Baffin Island inferred from lake sediment geochemistry. Science of the Total Environment, 407(21), 56535662.Google Scholar
Mighall, T. M., Timberlake, S., Foster, I. D., et al. (2009). Ancient copper and lead pollution records from a raised bog complex in Central Wales, UK. Journal of Archaeological Science, 36(7), 15041515.Google Scholar
Miles, B. W. J., Stokes, C. R., Veili, A., and Cox, N. J. (2013). Rapid, climate-driven changes in outlet glaciers on the Pacific coast of East Antarctica. Nature, 500, 563566.Google Scholar
Miller, G., Magee, J., Smith, M., et al. (2016). Human predation contributed to the extinction of the Australian megafunal bird Genyornis newtoni ~ 47 ka. Nature Communications, 7, 10496. doi:10.1038/ncomms10496.Google Scholar
Miller, G. H., Brigham-Grette, J., Alley, R. B., et al. (2010a). Temperature and precipitation history of the Arctic. Quaternary Science Reviews, 29, 16791715.Google Scholar
Miller, G. H., Brigham-Grette, J., Alley, R. B., et al. (2010b). Arctic amplification: Can the past constrain the future? Quaternary Science Reviews, 29(15–16), 17791790.Google Scholar
Miller, G. H., Geirsdottir, A., Zhong, Y., et al. (2012). Abrupt onset of the Little Ice Age triggered by volcanism and sustained by sea-ice/ocean feedbacks. Geophysical Research Letters, 39, L02708. doi:10.1029/2011GL050168.Google Scholar
Miller, G. H., Lehman, S. J., Refsnider, K. A., et al. (2013). Unprecedented recent summer warmth in Arctic Canada. Geophysical Research Letters, 40, 57455751.Google Scholar
Miller, K. G., Kopp, R. E., Horton, B. P., et al. (2013). A geological perspective on sea-level rise and its impacts along the U.S. mid-Atlantic coast. Earth’s Future, 1, 318.Google Scholar
Miller, K. G., and Wright, J. D. (2017). Success and failure in Cenozoic global correlations using golden spikes: A geochemical and magnetostratigraphic perspective. Episodes. 10.18814/epiiugs/2017/v40i1/017003.Google Scholar
Miller, K. G., Wright, J. D., Browning, J. V., et al. (2012). High tide of the warm Pliocene: Implications of global sea level for Antarctic deglaciation. Geology, 49, 407421.Google Scholar
Milliman, J. D., and Syvitski, J. P. M. (1992). Geomorphic/tectonic control of sediment discharge to the ocean: The importance of small mountainous rivers. Journal of Geology, 100, 525544.Google Scholar
Milne, G. A., Gehrels, W. R., Hughes, C. W., and Tamisiea, M. E. (2009). Identifying the causes of sea-level change. Nature Geoscience, 2, 471478.Google Scholar
Molina, E., Alegret, L., Arenillas, I., et al. (2006). The Global Boundary Stratotype Section and Point for the base of the Danian Stage (Paleocene, Paleogene, “Tertiary”, Cenozoic) at El Kef, Tunisia: Original definition and revision. Episodes, 29, 263273.Google Scholar
Monge, G., Jimenez-Espejo, F. J., García-Alix, A., et al. (2015). Earliest evidence of pollution by heavy metals in archaeological sites. Scientific Reports, 5, 14252.Google Scholar
Monnin, E., Indermühle, A., Dällenbach, A., et al. (2001). Atmospheric CO2 concentrations over the last glacial termination. Science, 291, 112114.Google Scholar
Monteith, D. T., Evans, C. D., Henrys, P. A., et al. (2014). Trends in the hydrochemistry of acid-sensitive surface waters in the UK 1988–2008. Ecological Indicators, 37, 287303.Google Scholar
Montzka, S. A., Aydin, M., Battle, M., et al. (2004). A 350-year atmospheric history for carbonyl sulfide inferred from Antarctic firn air and air trapped in ice. Journal of Geophysical Research, 109, D22302. doi:10.1029/2004JD004686.Google Scholar
Moore, A. M. T., Hillman, G. C., and Legge, A. J. (2000). Village on the Euphrates: From Foraging to Farming at Abu Hureyra. Oxford University Press.Google Scholar
Moore, C. J., Moore, S. L., Leecaster, M. K., and Weisberg, S. B. (2001). A comparison of plastic and plankton in the North Pacific Central Gyre. Marine Pollution Bulletin, 42, 12971300.Google Scholar
Mora, C., Dousset, B., Caldwell, I. R., et al. (2017). Global risk of deadly heat. Nature Climate Change, 7, 501506.Google Scholar
Morehead, M. D., Syvitski, J. P. M., Hutton, E. W. H., and Peckham, S. D. (2003). Modeling the inter-annual and intra-annual variability in the flux of sediment in ungauged river basins. Global and Planetary Change, 39(1/2), 95110.Google Scholar
Morf, L. S., Tremp, J., Gloor, R., et al. (2005). Brominated flame retardants in waste electrical and electronic equipment: Substance flows in a recycling plant. Environmental Science & Technology, 39(22), 86918699.Google Scholar
Morritt, D., Stefanoudis, P. V., Pearce, D., et al. (2014). Plastic in the Thames: A river runs through it. Marine Pollution Bulletin, 78, 196200.Google Scholar
Morton, O. (2015). The Planet Remade: How Geoengineering Could Change the World. London: Granta Publications.Google Scholar
Moura, R. L., Amado-Filho, G. M., Moraes, F. C., et al. (2016). An extensive reef system at the Amazon River mouth. Science Advances, 2(4), e1501252. doi:10.1126/sciadv.1501252.Google Scholar
Muir, D. C. G., and de Wit, C. A. (2010). Trends of legacy and new persistent organic pollutants in the circumpolar arctic: Overview, conclusions and recommendations. Science of the Total Environment, 408, 30443051.Google Scholar
Muir, D. C. G., and Howard, P. H. (2006). Are there other persistent organic pollutants? A challenge for environmental chemists. Environmental Science & Technology, 40(23), 71577166.Google Scholar
Muir, D. C. G., and Rose, N. L. (2007). Persistent organic pollutants in the sediments of Lochnagar. In Rose, N. L., ed., Lochnagar: The Natural History of a Mountain Lake, Developments in Paleoenvironmental Research, 12. Dordrecht: Springer, pp. 375402.Google Scholar
Mulder, T., and Syvitski, J. P. M. (1996). Climatic and morphologic relationships of rivers: Implications of sea level fluctuations on river loads. Journal of Geology, 104, 509523.Google Scholar
Muri, G., Wakeham, S., and Rose, N. L. (2006). Records of atmospheric delivery of pyrolysis-derived pollutants in recent mountain lake sediments of the Julian Alps (NW Slovenia). Environmental Pollution, 139, 461468.Google Scholar
Murphy, B. H., Farley, K. A., and Zachos, J. C. (2010). An extraterrestrial 3He-based timescale for the Paleocene-Eocene thermal maximum (PETM) from Walvis Ridge, IODP Site 1266. Geochimica et Cosmochimca Acta, 74, 50985108.Google Scholar
Muscheler, R., Beer, J., and Kubik, P. W. (2004). Long-term solar variability and climate change based on radionuclide data from ice cores. In Pap, J. M., and Fox, P., eds., Solar Variability and Its Effects on Climate. Geophysical Monograph, 114. American Geophysical Union, pp. 221235.Google Scholar
Myers, R. A., and Worm, B. (2003). Rapid worldwide depletion of predatory fish communities. Nature, 423, 280283.Google Scholar
Naish, T. R., Powell, R., Levy, R., et al. (2009). Obliquity-paced Pliocene West Antarctic oscillations. Nature, 458, 322328.Google Scholar
Naish, T. R., and Wilson, G. S. (2009). Constraints on the amplitude of mid-Pliocene (3.6–2.4 Ma) eustatic sea-level fluctuations from the New Zealand shallow-marine sediment record. Philosophical Transactions of the Royal Society of London A, 367, 169187.Google Scholar
Naredo, J. M., and Gutiérrez, L. (eds.) (2005). La Incidencia de la Especie Humana sobre la Faz de la Tierra (1955–2005). Fundacion César Manrique, Universidad de Granada.Google Scholar
National Academies of Sciences, Engineering, and Medicine (NASEM) (2015). A Strategic Vision for NSF Investment in Antarctic and Southern Ocean Research. Washington, DC: The National Academies Press.Google Scholar
National Academies of Sciences, Engineering, and Medicine (NASEM) (2017). Antarctic Sea Ice Variability in the Southern Ocean-Climate System. Washington, DC: The National Academies Press. doi:10.17226/24696.Google Scholar
National Funeral Directors Association (NFDA) (2016). The 2016 NFDA Cremation and Burial Report: Research, Statistics and Projections.Google Scholar
National Snow and Ice Data Center (NSIDC) (2017). http://nsidc.org/.Google Scholar
NEEM Community Members (2013). Eemian interglacial reconstructed from a Greenland folded ice core. Nature, 493, 489494.Google Scholar
Négrel, P., Kloppmann, W., Garcin, M., and Giot, D. (2004). Lead isotope signatures of Holocene fluvial sediments from the Loire River valley. Applied Geochemistry, 19, 957972.Google Scholar
Netz, R., and Noel, W. (2007). The Archimedes Codex. New York: DaCapo Press.Google Scholar
Neukom, R., Gergis, J., Karoly, D. J., et al. (2014). Interhemispheric temperature variability over the past millennium. Nature Climate Change, 4, 362367.Google Scholar
Newman, L., Wanner, H., and Kiefer, T. (2009). Towards a global synthesis of the climate of the last two millennia: Workshop of the pages 2k regional network, Corvallis, USA, 7 July 2009. PAGES News, 17(3), 130131.Google Scholar
Nickel, E. H. (1995a). Definition of a mineral. Canadian Mineralogist, 33, 689690.Google Scholar
Nickel, E. H. (1995b). Mineral names applied to synthetic substances. Canadian Mineralogist, 33, 1335.Google Scholar
Nickel, E. H., and Grice, J. D. (1998). The IMA Commission on New Minerals and Mineral Names: Procedures and guidelines on mineral nomenclature. Canadian Mineralogist, 36, 913926.Google Scholar
Nir, D. (1983). Man, a Geomorphological Agent: An Introduction to Anthropic Geomorphology. Keter Publication Jerusalem.Google Scholar
Nirei, H., Furuno, K., Osamu, K., et al. (2012). Classification of man made strata for assessment of geopollution. Episodes, 35(2), 333336.Google Scholar
Nisbet, E. G., Dlugokencky, E. J., Manning, M. R., et al. (2016). Rising atmospheric methane: 2007–2014 growth and isotopic shift. Global Biogeochemical Cycles, 30, 13561370.Google Scholar
NOAA (National Oceanic and Atmospheric Administration) (2017). Global analysis: Annual 2016. https://ncdc.noaa.gov/sotc/global/201613.Google Scholar
Nobre, C. A., and de Simone Borma, L. (2009). “Tipping points” for the Amazon forest. Current Opinion in Environmental Sustainability, 1, 2836.Google Scholar
Nogués-Bravo, D., Rodríguez, J., Hortal, J., et al. (2008). Climate change, humans, and the extinction of the woolly mammoth. PLoS Biology, 6(4), e79. doi:10.1371/journal.pbio.0060079.Google Scholar
Nordhaus, W. D. (1996). Do real-output and real-wage measures capture reality? The history of lighting suggests not. In Bresnahan, T. F., and Gordon, R. J., eds., The Economics of New Goods. http://nber.org/chapters/c6064.Google Scholar
North American Commission on Stratigraphic Nomenclature (NACSN) (1983). North American Stratigraphic Code. American Association of Petroleum Geologists B, 67, 841875.Google Scholar
North American Commission on Stratigraphic Nomenclature (NACSN) (2005). North American Stratigraphic Code. American Association of Petroleum Geologists B, 89, 15471591.Google Scholar
Novak, M., Adamová, M., Wieder, R. K., and Bottrell, S. H. (2005). Sulfur mobility in peat. Applied Geochemistry, 20, 673681.Google Scholar
Novak, M., Sipkova, A., Chrastny, V., et al. (2016). Cu-Zn isotope constraints on the provenance of air pollution in Central Europe: Using soluble and insoluble particles in snow and rime. Environmental Pollution, 218, 11351146.Google Scholar
Novakov, T., Ramanathan, V., Hansen, J. E., et al. (2003). Large historical changes of fossil-fuel black carbon aerosols. Geophysical Research Letters, 30(6), 1324, 57–1–57–4.Google Scholar
Nuelle, M.-T., Dekiff, J. H., Remy, D., and Fries, E. (2014). A new analytical approach for monitoring microplastics in marine sediments. Environmental Pollution, 184, 161169.Google Scholar
Obbard, R. W., Sadri, S., Wong, Y. Q., et al. (2014). Global warming releases microplastic legacy frozen in Arctic Sea ice. Earth’s Future, 2, 315320.Google Scholar
Oberle, F. K. J., Storlazzi, C. D., and Hanebuth, T. J. J. (2016). What a drag: Quantifying the global impact of chronic bottom trawling on continental shelf sediment. Journal of Marine Systems, 159, 109119.Google Scholar
Oberle, F. K. J., Swarzenski, P. W., Reddy, C. M., et al. (2015). Deciphering the lithological consequences of bottom trawling to sedimentary habitats on the shelf. Journal of Marine Systems, 159, 120131.Google Scholar
O’Connor, B. (2001). The origins and development of the British coprolite industry. Mining History: The Bulletin of the Peak District Mines Historical Society, 14(5), 4657.Google Scholar
Officer, C. B., Hallam, A., Drake, C. L., and Devine, J. D. (1987). Late Cretaceous and paroxysmal Cretaceous/Tertiary extinctions. Nature, 326, 143149.Google Scholar
Ogg, J. G., and Hinnov, L. A. (2014). Cretaceous. In Gradstein, F. M., Ogg, J. G., Schmitz, M. D., and Ogg, G., eds., The Geologic Time Scale. Elsevier, pp. 793853.Google Scholar
Oldfield, F. (2015). Can the magnetic signatures from inorganic fly ash be used to mark the onset of the Anthropocene? Anthropocene Review, 2(1), 313.Google Scholar
Oldfield, F., and Richardson, N. (1990). Lake sediment magnetism and atmospheric deposition. Philosophical Transactions of the Royal Society of London B, 327, 325330.Google Scholar
Oldfield, F., Thompson, R., and Barber, K. E. (1978). Changing atmospheric fallout of magnetic particles recorded in recent ombrotrophic peat. Science, 199, 679680.Google Scholar
Olóriz, F., Caracuel, J. E., and Rodríguez-Tovar, F. J. (1995). Using ecostratigraphic trends in sequence stratigraphy. In Haq, B. U., ed., Sequence Stratigraphy and Depositional Response to Eustatic, Tectonic and Climatic. Dordrecht: Springer-Publ., pp. 5985.Google Scholar
Oppenheimer, M., and Alley, R. B. (2016). How high will the seas rise? Science, 354, 13751377.Google Scholar
Oreskes, N. (1999). The Rejection of Continental Drift: Theory and Method in American Earth Science. New York: Oxford University Press.Google Scholar
Oreskes, N. (2003). Plate Tectonics: An Insider’s History of the Modern Theory of the Earth. Boulder, Westview Press.Google Scholar
Orio, A. A., and Botkin, D. B. (eds.) (1986). Man’s role in changing the global environment (papers presented at an international conference, Venice, Italy, 21–26 October 1985). Science of Total Environment, 55, 10399; 56, 1–415.Google Scholar
Orsi, A. J., Kawamura, K., Masson-Delmotte, V., et al. (2017). The recent warming trend in North Greenland. Geophysical Research Letters. doi:10.1002/2016GL072212.Google Scholar
Osborn, F. (1948). Our Plundered Planet. New York: Little, Brown and Company.Google Scholar
Osterman, L. E., Poore, R. Z., and Swarzenski, P. W. (2008). The last 1000 years of natural and anthropogenic low-oxygen bottom-water on the Louisiana shelf, Gulf of Mexico. Marine Micropaleontology, 66(3–4), 291303.Google Scholar
Ostrovsky, I. S. (1983). Productivity of the common zoobenthos species and their role in Lake Sevan ecosystem. Moscow [in Russian].Google Scholar
O’Sullivan, G., and Megson, D. (2014). Brief overview: Discovery, regulation, properties, and fate of POPs. In O’Sullivan, G., and Sandau, C. D., eds., Environmental Forensics for Persistent Organic Pollutants. Amsterdam: Elsevier, pp. 120.Google Scholar
Our World in Data (OWID) (2017). University of Oxford, OurWorldInData.org.Google Scholar
Overeem, I., Syvitski, J. P. M., Hutton, E. W. H., and Kettner, A. J. (2005). Stratigraphic variability due to uncertainty in model boundary conditions: A case study of the New Jersey Shelf over the last 21,000 years. Marine Geology, 224, 2341.Google Scholar
Overland, J. E., Wood, K. R., and Wang, M. (2011). Warm Arctic — cold continents: Climate impacts of the newly open Arctic Sea. Polar Research, 30. doi:10.3402/polar.v30i0.15787.Google Scholar
Pacyna, J. M., and Pacyna, E. G. (2001). An assessment of global and regional emissions of trace metals to the atmosphere from anthropogenic sources worldwide. Environmental Reviews, 9(4), 269298.Google Scholar
Pacyna, J. M., and Pacyna, E. G. (2016). Environmental Determinants of Human Health. Molecular and Integrative Toxicology. Switzerland: Springer International Publishing.Google Scholar
Padilla, D. K. (2010). Context-dependant impacts of a non-native ecosystem engineer, the Pacific oyster Crassostrea gigas. Integrative and Comparative Biology, 50, 213225.Google Scholar
Pagani, M., Lemarchand, D., Spivack, A., and Gaillardet, J. A. (2005). Critical evaluation of the boron isotope pH-proxy: The accuracy of ancient ocean pH estimates. Geochimica Cosmochimica Acta, 69, 953961.Google Scholar
Palanques, A., Guillén, J., and Puig, P. (2001). Impact of bottom trawling on water turbidity and muddy sediment of an unfished continental shelf. Limnology and Oceanography, 46, 11001110.Google Scholar
Palmieri, A., Shah, F., Annandale, G. W., and Dinar, A. (2003). Reservoir Conservation Vol. 1: The RESCON Approach: Economic and Engineering Evaluation of Alternative Strategies for Managing Sedimentation in Storage Reservoirs. International Bank for Reconstruction and Development, the World Bank.Google Scholar
Pandolfi, J. M. (2015). Incorporating uncertainty in predicting the future response of coral reefs to climate change. Annual Review of Ecology, Evolution, and Systematics, 46, 281303.Google Scholar
Paolo, F. S., Fricker, H. A., and Padman, L. (2015). Volume loss from Antarctic ice shelves is accelerating. Science, 348(6232), 327331.Google Scholar
Parrenin, F., Masson-Delmotte, V., Köhler, P., et al. (2013). Synchronous change of atmospheric CO2 and Antarctic temperature during the last deglacial warming. Science, 339(6123), 10601063.Google Scholar
Patris, N., Delmas, R. J., Legrand, M., et al. (2002). First sulfur isotope measurements in central Greenland ice cores along the preindustsrial and industrial periods. Journal of Geophysical Research, 107(D11), 4115. doi:10.1029/2001JD000672.Google Scholar
Patton, H., Hubbard, A., Andreassen, K., et al. (2017). Deglaciation of the Eurasian ice sheet complex. Quaternary Science Reviews, 169, 148172.Google Scholar
Paull, C. K., Ussler, W. III, Mitts, P. J., et al. (2006). Discordant 14C-stratigraphies in upper Monterey Canyon: A signal of anthropogenic disturbance. Marine Geology, 233, 2136.Google Scholar
Pauly, D. (2010). 5 Easy Pieces: The Impact of Fisheries on Marine Systems. Island Press.Google Scholar
Pauly, D., Christensen, V., Guenette, S., et al. (2002). Towards sustainability in world fisheries. Nature, 418, 689695.Google Scholar
Pearce, F. (2017). A year on thin ice. New Scientist, 234(3120), 3337.Google Scholar
Pearson, A. J., Pizzuto, J. E., and Vargas, R. (2015). Influence of run of river dams on floodplain sediments and carbon dynamics. Geoderma, 272, 5163.Google Scholar
Pearson, P. N., and Palmer, M. R. (2000). Atmospheric carbon dioxide concentrations over the past 60 million years. Nature, 406, 695699.Google Scholar
Pedro, J. B., Rasmussen, S. O., and Van Ommen, T. D (2012). Tightened constraints on the time-lag between Antarctic temperature and CO2 during the last deglaciation. Climate of the Past, 8, 12131221.Google Scholar
Peloggia, A. U. G., Oliveira, A. M. S., Oliveira, A. A., et al. (2014). Technogenic geodiversity: A proposal on the classification of artificial ground. Quaternary and Environmental Geosciences, 5(1), 2840.Google Scholar
Penman, D. E., Hönisch, B., Zeebe, R. E., et al. (2014). Rapid and sustained ocean acidification during the Paleocene-Eocene Thermal Maximum. Paleoceanography, 29, 357369.Google Scholar
Pereira, N. S., Sial, A. N., Kikuchi, R. K. P., et al. (2015). Coral-based climate records from tropical South Atlantic: 2009/2010 ENSO event in C and O isotopes from Porites corals (Rocas Atoll, Brazil). Anais da Academia Brasileira de Ciência, 87(4). doi:10.1590/0001–3765201520150072.Google Scholar
Periman, R. D. (2006). Visualizing the Anthropocene: Human land use history and environmental management. In Aguirre-Bravo, C., Pellicane, P. J., Burns, D. P., and Draggan, S., eds., Monitoring Science and Technology Symposium: Unifying Knowledge for Sustainability in the Western Hemisphere. Fort Collins: US Department of Agriculture, pp. 558564.Google Scholar
Perry, W. E. (1992). The utilization of by-products and waste products in the production of commercial fertilizers. Fertilizer Research, 32, 111114.Google Scholar
Petit, J. R., Jouzel, J., Raynaud, D., et al. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399, 429436.Google Scholar
Petoukhov, V., and Semenov, V. A. (2010). A link between reduced Barents-Kara sea ice and cold winter extremes over northern continents. Journal of Geophysical Research, 115, D21111. doi:10.1029/2009JD013568.Google Scholar
Pfeffer, W. T., Arendt, A. A., Bliss, A., et al. (2014). The Randolph Glacier Inventory: A globally complete inventory of glaciers. Journal of Glaciology, 60(221), 537552.Google Scholar
Pham, C. K., Ramirez-Llodra, E., Alt, C. H. S., et al. (2014). Marine litter density and distribution in European seas, from shelves to deep basins. PLoS ONE, 9, e95839.Google Scholar
Pillans, B., and Gibbard, P. (2012). The Quaternary Period. In Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G., eds., The Geological Time Scale 2012. Elsevier, pp. 9791010.Google Scholar
Pillans, B., and Naish, T. (2004). Defining the Quaternary. Quaternary Science Reviews, 23, 22712282.Google Scholar
Piperno, D. R. (2011). The origins of plant cultivation and domestication in the New World tropics: Patterns, process, and new developments. Current Anthropology, 52, S453S470.Google Scholar
Piruzyan, L. A., Malenkov, A. G., Barenboim, G. M., and Precoda, N. (1980). Chemistry and the Biosphere. Environment, 22(10), 2530.Google Scholar
Pla, S., Monteith, D., Flower, R., and Rose, N. (2009). The recent palaeolimnology of a remote Scottish loch with special reference to the relative impacts of regional warming and atmospheric contamination. Freshwater Biology, 54, 505523.Google Scholar
Planchon, F. A., Boutron, C. F., Barbante, C., et al. (2002). Changes in heavy metals in Antarctic snow from Coats Land since the mid-19th to the late-20th century. Earth and Planetary Science Letters, 200(1), 207222.Google Scholar
Plass, G. N. (1961). The influence of infrared absorptive molecules on the climate. Annals of the New York Academy of Sciences, 95, 6171.Google Scholar
PlasticsEurope (2016). Plastics – the facts 2016: An analysis of European latest plastics production, demand and waste data. http://plasticseurope.org/Document/plastics---the-facts-2016-15787.aspx?FolID=2.Google Scholar
Plater, A. J., Ridgway, J., Appleby, P. J., et al. (1998). Historical contaminant fluxes in the Tees Estuary, UK: Geochemical, magnetic and radionuclide evidence. Marine Pollution Bulletin, 37, 343360.Google Scholar
Plumb, K. A. (1991). New Precambrian time scale. Episodes, 14(2), 139140.Google Scholar
Pohl, M. E. D., Piperno, D. R., Pope, K. O., and Jones, J. G. (2007). Microfossil evidence for pre-Columbian maize dispersals in the neotropics from San Andrés, Tabasco, Mexico. Proceedings of the National Academy of Sciences (USA), 104, 68706875.Google Scholar
Pohl, T., Al-Muqdadi, S. W., Ali, M. H., et al. (2013). Discovery of a living coral reef in the coastal waters of Iraq. Scientific Reports, 4, 4250. doi:10.1038/srep04250.Google Scholar
Polato, N. R., Voolstra, C. R., and Schnetzer, J. (2010). Location-specific responses to thermal stress in larvae of the reef-building coral Montastraea faveolata. PLoS ONE, 5(6), e11221. doi:10.1371/journal.pone.0011221.Google Scholar
Pollard, D., and DeConto, R. (2005). Hysteresis in Cenozoic Antarctic ice-sheet variations. Global and Planetary Change, 45, 921.Google Scholar
Pollard, D., and DeConto, R. (2009). Modelling West Antarctic ice sheet growth and collapse through the past five million years. Nature, 458, 329333.Google Scholar
Pollard, D., DeConto, R. M., and Alley, R. B. (2015). Potential Antarctic Ice Sheet retreat driven by hydrofracturing and ice cliff failure. Earth and Planetary Science Letters, 412, 112121.Google Scholar
Poloczanska, E. S., Burrows, M. T., Brown, C. J., et al. (2016). Responses of marine organisms to climate change across oceans. Frontiers in Marine Science, 3(62). doi:10.3389/fmars.2016.00062.Google Scholar
Pompeani, D. P., Abbott, M. B., Bain, D. J., et al. (2015). Copper mining on Isle Royale 6500–5400 years ago identified using sediment geochemistry from McCargoe Cove, Lake Superior. Holocene, 25, 253262.Google Scholar
Pompeani, D. P., Abbott, M. B., Steinman, B. A., and Bain, D. J. (2013). Lake sediments record prehistoric lead pollution related to early copper production in North America. Environmental Science & Technology, 47(11), 55455552.Google Scholar
Pongratz, R., and Heumann, K. G. (1999). Production of methylated mercury, lead, and cadmium by marine bacteria as a significant natural source for atmospheric heavy metals in polar regions. Chemosphere, 39(1), 89102.Google Scholar
Porter, S. (2011). The rise of predators. Geology, 39, 607608.Google Scholar
Prado, J., Martinez-Maza, C., and Alberdi, M. T. (2015). Megafauna extinction in South America: A new chronology for the Argentine Pampas. Palaeogeography, Palaeoclimatology, Palaeoecology, 425, 4149.Google Scholar
Preunkert, S., and Legrand, M. (2013). Towards a quasi-complete reconstruction of past atmospheric aerosol load and composition (organic and inorganic) over Europe since 1920 inferred from Alpine ice cores. Climates of the Past, 9, 14031416.Google Scholar
Price, S. J., Ford, J. R., Cooper, A. H., and Neal, C. (2011). Humans as major geological and geomorphological agents in the Anthropocene: The significance of artificial ground in Great Britain. In Zalasiewicz, J. A., Williams, M., Haywood, A., and Ellis, M., eds., The Anthropocene: A New Epoch of Geological Time. Philosophical Transactions of the Royal Society (Series A), 369, pp. 10561084.Google Scholar
Pritchard, H. D., Ligtenberg, S. R. M., Fricker, H. A., et al. (2012). Antarctic ice-sheet loss driven by basal melting of ice shelves. Nature, 484, 502505.Google Scholar
Pugh, D., and Woodworth, P. (2014). Sea-Level Science: Understanding Tides, Surges, Tsunamis and Mean Sea-Level Changes. Cambridge University Press.Google Scholar
Puig, P., Canals, M., Company, J. B., et al. (2012). Ploughing the deep sea floor. Nature, 489, 286289.Google Scholar
Punmia, B. C., Jain, A. K., and Jain, A. K. (2004). Basic Civil Engineering. New Delhi, India: Laxmi Publications (P) Ltd.Google Scholar
Qiu, Y. W., Zhang, G., Liu, G. Q., et al. (2009). Polycyclic aromatic hydrocarbons (PAHs) in the water column and sediment core of Deep Bay, South China. Estuarine, Coastal and Shelf Science, 83(1), 6066.Google Scholar
Quinto, F., Hrnecek, E., Krachler, M., et al. (2013). Determination of 239Pu, 240Pu, 241Pu and 242Pu at femtogram and attogram levels—evidence for the migration of fallout plutonium in an ombrotrophic peat bog profile. Environmental Science: Processes & Impacts, 15(4), 839847.Google Scholar
Rackow, T., Wesche, C., Timmermann, R., et al. (2016). A simulation of small to giant Antarctic iceberg evolution: Differential impact on climatology estimates. Journal of Geophysical Research – Oceans, 122(4), 31703190.Google Scholar
Radivojević, M., Rehren, T., Pernicka, E., et al. (2010). On the origins of extractive metallurgy: New evidence from Europe. Journal of Archaeological Science, 37, 27752787.Google Scholar
Raes, F., Van Dingenen, R., Vignati, E., et al. (2000). Formation and cycling of aerosols in the global troposphere. Atmospheric Environment, 34(25), 42154240.Google Scholar
Rahmstorf, S., Foster, G., and Cahill, N. (2017). Global temperature evolution: Recent trends and some pitfalls. Environmental Research Letters, 12(5).Google Scholar
Rakowski, A. Z., Nadeau, M.-J., Nakamura, T., et al. (2013). Radiocarbon method in environmental monitoring of CO2 emission. Nuclear Instruments and Methods in Physics Research B, 294, 503507.Google Scholar
Ramirez-Llodra, E., Tyler, P. A., Baker, M. C., et al. (2011). Man and the last great wilderness: Human impact on the deep sea. PLoS ONE, 6, e22588. doi.org/10.1371/journal.pone.0022588.Google Scholar
Rasmussen, B., and Buick, R. (1999). Redox state of the Archaean atmosphere: Evidence from detrital heavy minerals in ca 3250–2750 Ma sandstones from the Pilbara Craton, Australia. Geology, 27, 115118.Google Scholar
Rathje, W. L., and Murphy, C. (2001). Rubbish!: The Archaeology of Garbage. Tucson: University of Arizona Press.Google Scholar
Rauch, J. N. (2010). Global spatial indexing of the human impact on Al, Cu, Fe and Zn mobilization. Environmental Science and Technology, 44, 57285734.Google Scholar
Rauch, J. N. (2012). The present understanding of Earth’s global anthrobiogeochemical metal cycles. Mineral Economics, 25(1), 715.Google Scholar
Rauch, J. N., and Graedel, T. E. (2007). Earth’s anthrobiogeochemical copper cycle. Global Biogeochemical Cycles, 21(2). doi:10.1029/2006GB002850.Google Scholar
Rauch, J. N., and Pacyna, J. M. (2009). Earth’s global Ag, Al, Cr, Cu, Fe, Ni, Pb, and Zn cycles. Global Biogeochemical Cycles, 23(2). doi:10.1029/2008GB003376.Google Scholar
Rausch, N., Nieminen, T., Ukonmaanaho, L., et al. (2005). Comparison of atmospheric deposition of copper, nickel, cobalt, zinc, and cadmium recorded by Finnish peat cores with monitoring data and emission records. Environmental Science & Technology, 39(16), 59895998.Google Scholar
Razali, M. N., Effendi, M. L. H. M, Musa, M., and Yunus, R. M. (2016). Formulation of bitumen from industrial waste. ARPN Journal of Engineering and Applied Sciences, 11(8), 52445250.Google Scholar
Reading, H. G. (ed.) (1996). Sedimentary Environments: Processes, Facies and Stratigraphy, 3rd ed. Malden, MA: Blackwell Publishing.Google Scholar
Reager, J. T., Gardner, A. S., Famiglietti, J. S., et al. (2016). A decade of sea level rise slowed by climate-driven hydrology. Science, 351, 699703.Google Scholar
Reed, A. J., Mann, M. E., Emanuel, K. A., et al. (2015). Increased threat of tropical cyclones and coastal flooding to New York City during anthropogenic era. Proceedings of the National Academy of Sciences (USA), 112(41), 1261012615.Google Scholar
Reed, C. (2015). Dawn of the Plasticene age. New Scientist, 225 (3006, 31 January 2015), 2832.Google Scholar
Reis, L. (2014). Occurrence of macro- and microplastics in sediments from Fehmarm. BSc. Unpublished thesis, Freie Universität, Berlin.Google Scholar
Remane, J. (1997). Foreword: Chronostratigraphic standards: How are they defined and when should they be changed? Quaternary International, 40, 34.Google Scholar
Remane, J. (2003). Chronostratigraphic correlations: Their importance for the definition of geochronologic units. Palaeogeography, Palaeoclimatology, Palaeoecology, 196, 718.Google Scholar
Remane, J., Bassett, M. G., Cowie, J. W., et al. (1996). Revised guidelines for the establishment of global chronostratigraphic standards by the International Commission on Stratigraphy (ICS). Episodes, 19, 7781.Google Scholar
Renaud, F., Syvitski, J. P. M., Sebesvari, Z., et al. (2013). Tipping from the Holocene to the Anthropocene: How threatened are major world deltas? Current Opinion in Environmental Sustainability, 5, 644654.Google Scholar
Renberg, I., and Battarbee, R. W. (1990). The SWAP Palaeolimnology Programme: A synthesis. In Mason, B. J (ed). The Surface Waters Acidification Programme. Cambridge University Press, pp. 281300.Google Scholar
Renberg, I., Brännvall, M.-L., Bindler, R., and Emteryd, O. (2000). Atmospheric lead pollution history during four millennia (2000 BC to 2000 AD) in Sweden. Ambio, 29(3), 150156.Google Scholar
Renberg, I., Persson, M., and Emteryd, O. (1994). Pre-industrial atmospheric lead contamination detected in Swedish lake sediments. Nature, 368, 323326.Google Scholar
Restrepo, J. D., and Syvitski, J. P. M. (2006). Assessing the effect of natural controls and land use change on sediment yield in a major Andean river: The Magdalena drainage basin, Colombia. Ambio, 35, 6574.Google Scholar
Revkin, A. C. (1992). Global Warming: Understanding the Forecast (American Museum of Natural History, Environmental Defense Fund). New York: Abbeville Press.Google Scholar
Reynolds, P., Planke, S., Millett, J. M., et al. (2017). Hydrothermal vent complexes offshore Northeast Greenland: A potential role in driving the PETM. Earth and Planetary Science Letters, 467, 7278.Google Scholar
Reznikov, S. A. (1984). Biogenic elements of sediments in Lake Sevan. Trudy Sevanskoi Gidrobiologicheskoi Stantsii, 19, 517 [in Russian].Google Scholar
Rhoads, B. L., Quinn, W., and Lewis, W. A. (2016). Historical changes in channel network extent and channel planform in an intensively managed landscape: Natural versus human-induced effects. Geomorphology, 252, 1731.Google Scholar
Ribeiro, S., Amorim, A., Abrantes, F., and Ellegaard, M. (2016). Environmental change in the Western Iberia Upwelling Ecosystem since the preindustrial period revealed by dinoflagellate cyst records. Holocene, 26, 874889.Google Scholar
Ribeiro, S., Amorim, A., Andersen, T. J., et al. (2012). Reconstructing the history of an invasion: The toxic phytoplankton species Gymnodinium catenatum in the Northeast Atlantic. Biological Invasions, 14, 969985.Google Scholar
Richards, Z. T., and Beger, M. (2011). A quantification of the standing stock of macro-debris in Majuro lagoon and its effect on hard coral communities. Marine Pollution Bulletin, 62, 16931701.Google Scholar
Richter, D. deB. (2007). Humanity’s transformation of Earth’s soil: Pedology’s new frontier. Soil Science, 172, 957967.Google Scholar
Richter, D. deB., Bacon, A. R., Megan, L. M., et al. (2011). Human–soil relations are changing rapidly: Proposals from SSSA’s cross-divisional Soil Change Working Group. Soil Science Society of America Journal, 75(6), 20792084.Google Scholar
Richter, D. deB., Grün, R., Joannes-Boyau, R., et al. (2017). The age of the hominin fossils from Jbel Irhoud, Morocco, and the origins of the Middle Stone Age. Nature, 546, 293296.Google Scholar
Richter, D. deB., and Markewitz, D. (1995). How deep is soil? BioScience, 45, 600609.Google Scholar
Richter, D. deB., and Yaalon, D. H. (2012). “The changing model of soil” revisited. Soil Science Society of America Journal, 76, 766778.Google Scholar
Richter-Menge, J., Overland, J. E., and Mathis, J. T. (eds.) (2016). Arctic Report Card 2016. http://arctic.noaa.gov/Report-Card.Google Scholar
Rietbroek, R., Brunnabend, S.-E., Kusche, J., et al. (2016). Revisiting the contemporary sea-level budget on global and regional scales. Proceedings of the National Academy of Sciences (USA), 113, 15041509.Google Scholar
Rigaud, J.-P., Texier, P.-J., Parkington, J., and Poggenpoel, C. (2006). South African Middle Stone Age chronology: New excavations at Diepkloof Rock Shelter; Preliminary results. Comptes Rendus Palevol, 5, 839849.Google Scholar
Rignot, E., Casassa, G., Gogineni, P., et al. (2004). Accelerated ice discharge from the Antarctic Peninsula following the collapse of Larsen B ice shelf. Geophysical Research Letters, 31, L18401. doi:10.1029/2004GL020697.Google Scholar
Rignot, E., Velicogna, I., van den Broeke, M. R., et al. (2011). Acceleration of the contribution of the Greenland and Antarctic ice sheets to sea level rise. Geophysical Research Letters, 38, L05503. doi:10.1029/2011GL046583.Google Scholar
Rijsdijk, K. F, Hume, J. P., Bunnik, F., et al. (2009). Mid-Holocene vertebrate bone concentration-Lagerstätte on oceanic island Mauritius provides a window into the ecosystem of the dodo (Taphus cucullatus). Quaternary Science Reviews, 28, 1424.Google Scholar
Rillig, M. C. (2012). Microplastic in terrestrial ecosystems and the soil? Environmental Science and Technology, 46, 64536454.Google Scholar
Rintoul, S. R., Silvano, A., Pena-Molino, B., et al. (2016). Ocean heat drives rapid basal melt of the Totten Ice Shelf. Science Advances, 2, e1601610.Google Scholar
Rio, D., Sprovieri, R., Castradori, D., and Di Stefano, E. (1998). The Gelasian Stage (Upper Pliocene): A new unit of the global standard chronostratigraphic scale. Episodes, 21(2), 8287.Google Scholar
Rivas, V., Cendero, A., Hurtado, M., et al. (2006). Geomorphic consequences of urban development and mining activities: An analysis of study areas in Spain and Argentina. Geomorphology, 73, 185206.Google Scholar
Roberts, C. (2007). The Unnatural History of the Sea: The Past and Future of Humanity and Fishing (Gaia Thinking). Washington, DC: Island Press.Google Scholar
Roberts, C. (2013). Ocean of Life. London: Penguin.Google Scholar
Roberts, J., Moy, A., Plummer, C., et al. (2017). A revised Law Dome age model (LD2017) and implications for last glacial climate. Climate of the Past Discussions. https://doi.org/10.5194/cp-2017-96.Google Scholar
Roberts, R. G., Flannery, T. F., Ayliffe, L. K., et al. (2001). New ages for the last Australian megafauna: Continent-wide extinction about 46,000 years ago. Science, 292, 18881892.Google Scholar
Robertson, J. (2017). Runoff pollution from Cyclone Debbie flooding sweeps into Great Barrier Reef. The Guardian, 11 April 2017. https://theguardian.com/australia-news/2017/apr/11/run-off-pollution-from-cyclone-debbie-flooding-sweeps-into-great-barrier-reef (accessed June 2017).Google Scholar
Robinson, N. (2012). Beyond sustainability: Environmental management for the Anthropocene epoch. Journal of Public Affairs, 12, 181194.Google Scholar
Robison, B. H., Reisenbichler, K. R., and Sherlock, R. E. (2005). Giant larvacean houses: Rapid carbon transport to the deep sea floor. Science, 308, 16091611.Google Scholar
Rochman, C., Browne, M. A., Halpern, B., et al. (2013). Classify plastic waste as hazardous. Nature, 494, 169171.Google Scholar
Rockström, J., Gaffney, O., Rogelj, J., et al. (2017). A roadmap for rapid decarbonization. Science, 355, 12691271.Google Scholar
Rockström, J., Steffen, W., Noone, K., et al. (2009). A safe operating space for humanity. Nature, 461, 472475.Google Scholar
Roebroeks, W., and Villa, P. (2011). On the earliest evidence for habitual use of fire in Europe. Proceedings of the National Academy of Sciences (USA), 108, 52095214.Google Scholar
Roger, J. (ed.) (1962). Buffon: Les Époques de la Nature. Edition critique, Mémoires du Muséum National d’Histoire Naturelle. Nouvelle Série, Série C, Sciences de la Terre, Tome X, Paris.Google Scholar
Röhl, U., Westerhold, T., Bralower, T. J., and Zachos, J. C. (2007). On the duration of the Paleocene-Eocene thermal maximum (PETM). Geochemistry, Geophysics, Geosystems, 8 (12), Q12002. doi:10.1029/2007GC001784.Google Scholar
Rohling, E. J., Haigh, I. D., Foster, G. L., et al. (2013). A geological perspective on potential future sea-level rise. Scientific Reports, 3, 3461.Google Scholar
Rohwer, F., and Youle, M. (2010). Coral reefs in the microbial seas. San Francisco, CA: Plaid Press.Google Scholar
Rolison, J. M., Landing, W. M., Luke, W., et al. (2013). Isotopic composition of species-specific atmospheric Hg in a coastal environment. Chemical Geology, 336, 3749.Google Scholar
Roosevelt, A. C. (2013). The Amazon and the Anthropocene: 13,000 years of human influence in a tropical rainforest. Anthropocene, 4, 6987.Google Scholar
Rose, N. L. (1996). Inorganic ash spheres as pollution tracers. Environmental Pollution, 91, 245252.Google Scholar
Rose, N. L. (2001). Fly-ash particles. In Last, W. M., and Smol, J. P., eds., Tracking Environmental Change Using Lake Sediments: Volume 2. Physical and Chemical Techniques. Dordrecht, Netherlands: Kluwer Academic Publishers, pp. 319349.Google Scholar
Rose, N. L. (2015). Spheroidal carbonaceous fly-ash particles provide a globally synchronous stratigraphic marker for the Anthropocene. Environmental Science & Technology, 49, 41554162.Google Scholar
Rose, N. L. (2018). Spheroidal carbonaceous fly ash particles in the Anthropocene. In DellaSala, D. A., and Goldstein, M. I., eds., Encyclopaedia of the Anthropocene, vol. 1. Oxford: Elsevier.Google Scholar
Rose, N. L., and Appleby, P. G. (2005). Regional applications of lake sediment dating by spheroidal carbonaceous particle analysis I. Journal of Paleolimnology, 34, 349361.Google Scholar
Rose, N. L., Backus, S., Karlsson, H., and Muir, D. C. G. (2001). An historical record of toxaphene and its congeners in a remote lake in western Europe. Environmental Science & Technology, 35, 13121319.Google Scholar
Rose, N. L., Jones, V. J., Noon, P. E., et al. (2012). Long-range transport of pollutants to the Falkland Islands and Antarctica: Evidence from lake sediment fly-ash particle records. Environmental Science & Technology, 46, 98819889.Google Scholar
Rose, N. L., and Ruppel, M. (2015). Environmental archives of contaminant particles. In Blais, J. M., Rosen, M. R., and Smol, J. P., eds., Environmental Contaminants: Using Natural Archives to Track Sources and Long-Term Trends of Pollution. Developments in Paleoenvironmental Research, 18. Dordecht, Netherlands: Springer, pp. 187221.Google Scholar
Rosenbaum, M. S., McMillan, A. A., Powell, J. H., et al. (2003). Classification of artificial (man-made) ground. Engineering Geology, 69, 399409.Google Scholar
Rosman, K. J., Chisholm, W., Hong, S., et al. (1997). Lead from Carthaginian and Roman Spanish mines isotopically identified in Greenland ice dated from 600 BC to 300 AD. Environmental Science & Technology, 31(12), 34133416.Google Scholar
Rosswall, T., Liss, P., Rapley, C., et al. (2015). Reflections on Earth-System Science. Global Change, 84, 813.Google Scholar
Rothman, A. (2007). Slave Country: American Expansion and the Origins of the New South. Cambridge, MA: Harvard University Press.Google Scholar
Royal Society (2009). Geoengineering the Climate: Science, Governance and Uncertainty. Report 10/9 RS1636. London: Royal Society.Google Scholar
Royer, D. L. (2006). CO2-forced climate thresholds during the Phanerozoic. Geochimica Cosmochimica Acta, 70, 56655675.Google Scholar
Royer, D. L., Berner, R. A., Montañez, I. P., et al. (2004). CO2 as a primary driver of Phanerozoic climate. GSA Today, 14(3), 410.Google Scholar
Rubino, M., Etheridge, D. M., Trudinger, C. M., et al. (2013). A revised 1000 year atmospheric δ13C-CO2 record from Law Dome and South Pole, Antarctica. Journal of Geophysical Research, 118, 84828499.Google Scholar
Rubino, M., Etheridge, D. M., Trudinger, C. M., et al. (2016). Low atmospheric CO2 levels during the Little Ice Age due to cooling-induced terrestrial uptake. Nature Geoscience, 9, 691694.Google Scholar
Ruddiman, W. F. (1977). Late Quaternary deposition of ice-rafted sand in the subpolar North Atlantic (Lat 40° to 65°N). Bulletin of the Geological Society of America, 88, 18131827.Google Scholar
Ruddiman, W. F. (2003). The anthropogenic Greenhouse Era began thousands of years ago. Climatic Change, 61, 261293.Google Scholar
Ruddiman, W. F. (2005). Plows, Plagues and Petroleum. Princeton University Press.Google Scholar
Ruddiman, W. F. (2013). Anthropocene. Annual Review of Earth and Planetary Sciences, 41, 4568.Google Scholar
Ruddiman, W. F. (2014). Earth’s Climate: Past and Future, 3rd ed. New York: W. H. Freeman.Google Scholar
Ruddiman, W. F., Ellis, E. C., Kaplan, J. O., and Fuller, D. Q. (2015a). Defining the epoch we live in. Science, 348, 3839.Google Scholar
Ruddiman, W. F., Fuller, D. Q., Kutzbach, J. E., et al. (2015b). Late Holocene climate: Natural or anthropogenic? Reviews of Geophysics, 54(1), 93118.Google Scholar
Rudwick, M. J. S. (2005). Bursting the Limits of Time: The Reconstruction of Geohistory in the Age of Revolution. Chicago University Press.Google Scholar
Rudwick, M. J. S. (2008). Worlds before Adam: The Reconstruction of Geohistory in the Age of Reform. Chicago University Press.Google Scholar
Ruhe, R. V., and Scholtes, W. H. (1956). Age and development of soil landscapes in relation to climate and vegetational changes in Iowa. Soil Science Society of America Proceedings, 20, 264273.Google Scholar
Ruppel, C. D., and Kessler, J. D. (2017). The interaction of climate change and methane hydrates. Reviews of Geophysics, 55(1), 126168.Google Scholar
Rushton, A. W. A., Brück, P. M., Molyneux, S. G., et al. (2011). A revised correlation of the Cambrian rocks in the British Isles. Geological Society of London, Special Report, 25.Google Scholar
Ryan, P. G., Moore, C. J., van Franeker, J. A., and Moloney, C. L. (2009). Monitoring the abundance of plastic debris in the marine environment. Philosophical Transactions of the Royal Society B, 364, 19992012.Google Scholar
Sabine, C. L., Feely, R. A., Gruber, N., et al. (2004). The oceanic sink for anthropogenic CO2. Science, 305, 367371.Google Scholar
Sadri, S. S., and Thompson, R. C. (2014). On the quantity and composition of floating plastic debris entering and leaving the Tamar Estuary, Southwest England. Marine Pollution Bulletin, 81, 5560.Google Scholar
Sagan, C. (1994). Pale Blue Dot: A Vision of the Human Future in Space, 1st ed. New York: Random House.Google Scholar
Sagnotti, L., Scardia, G., Giaccio, B., et al. (2014). Extremely rapid directional change during Matuyama-Brunhes geomagnetic polarity reversal. Geophysical Journal International, 199, 11101124.Google Scholar
Saito, Y., Chaimanee, N., Jarupongsakul, T., and Syvitski, J. P. M. (2007). Shrinking megadeltas in Asia: Sea-level rise and sediment reduction impacts from case study of the Chao Phraya delta. Imprint Newsletter of the IGBP/IHDP Land Ocean Interaction in the Coastal Zone, 2007(2), 39.Google Scholar
Sakurai, T., Ohno, H., Motoyama, H., and Uchida, T. (2016). Micro-droplets containing sulfate in the Dome Fuji deep ice core, Antarctica: Findings using micro-Raman spectroscopy. Journal of Raman Spectroscopy, 48, 448452.Google Scholar
Salomons, W., and Förstner, U. (2012). Metals in the Hydrocycle. Berlin: Springer Science & Business Media.Google Scholar
Salvador, A. (ed.) (1994). International Stratigraphic Guide: A Guide to Stratigraphic Classification, Terminology, and Procedure, 2nd ed. Boulder, Colorado: International Subcommission on Stratigraphic Classification of IUGS International Commission on Stratigraphy and the Geological Society of America.Google Scholar
Salzmann, U., Williams, M., Haywood, A. M., et al. (2011) Climate and environment of a Pliocene warm world. Palaeogeography, Palaeoclimatology, Palaeoecology, 309, 18.Google Scholar
Samways, M. (1999). Translocating fauna to foreign lands: Here comes the Homogenocene. Journal of Insect Conservation, 3, 6566.Google Scholar
Sandom, C., Faurby, S., Sandel, B., and Svenning, J.-C. (2013). Global late Quaternary megafauna extinctions linked to humans, not climate change. Proceedings of the Royal Society B, 281. doi:10.1098/rspb.2013.3254.Google Scholar
Sarg, J. F. (1988). Carbonate sequence stratigraphy. In Wilgus, C. K., Hastings, B. S., Posamentier, H., et al., eds., Sea-Level Changes: An Integrated Approach. Tulsa: SEPM Special Publication, 42, pp. 155181.Google Scholar
Saunois, M., Jackson, R. B., Bousquet, P., et al. (2016). The growing role of methane in anthropogenic climate change. Environmental Research Letters, 11. doi:10.1088/1748–9326/11/12/120207.Google Scholar
Schaefer, J. M., Finkel, R. C., Balco, G., et al. (2016). Greenland was nearly ice-free for extended periods during the Pleistocene. Nature, 540, 252255.Google Scholar
Schaetzl, R. J., and Thompson, M. L. (2015). Soils: Genesis and Geomorphology. Cambridge University Press.Google Scholar
Scheffers, B. R., De Meester, L., Bridge, T. C. L., et al. (2016). The broad footprint of climate change from genes to biomes to people. Science, 354. doi:10.1126/science.aaf7671.Google Scholar
Schellnhuber, H. J., Rahmstorf, S., and Winkelmann, R. (2016). Why the right climate target was agreed in Paris. Nature Climate Change, 6, 649653.Google Scholar
Scherer, R. P., DeConto, R. M., Pollard, D., and Alley, R. B. (2016). Windblown Pliocene diatoms and East Antarctic Ice Sheet retreat. Nature Communications, 7. doi:10.1038/ncomms12957.Google Scholar
Schimmelmann, A., Hendy, I. L., Dunn, L., et al. (2013). Revised ~2000-year chronostratigraphy of partially varved marine sediment in Santa Barbara Basin, California. GFF, 135(3–4), 258264.Google Scholar
Schimmelmann, A., Lange, C. B., Schieber, J., et al. (2016). Varves in marine sediments: A review. Earth-Science Reviews, 159, 215246.Google Scholar
Schimper, W. P. (1874). Traité de Paléontologie Végétale. Paris: J. B. Ballière, 3.Google Scholar
Schlager, W. (1992). Sedimentology and sequence stratigraphy of reefs and carbonate platforms. American Association of Petroleum Geologists Contin. Educ. Course Note Series, 34, Tulsa.Google Scholar
Schlining, K., von Thun, S., Kuhnz, L., et al. (2013). Debris in the deep: Using a 22-year video annotation database to survey marine litter in Monterey Canyon, central California, USA. Deep-Sea Research Part I: Ocean Research Paper, 79, 96105.Google Scholar
Schmid, P., Bogdal, C., Blüthgen, N., et al. (2010). The missing piece: Sediment records in remote mountain lakes confirm glaciers being secondary sources of persistent organic pollutants. Environmental Science & Technology, 45(1), 203208.Google Scholar
Schmidt, H., and Reimers, C. E. (1991). The recent history of trace metal accumulation in the Santa Barbara Basin, southern California borderland. Estuarine, Coastal and Shelf Science, 33, 485500.Google Scholar
Schmitt, J., Schneider, R., Elsig, J., et al. (2012). Carbon isotope constraints on the deglacial CO2 rise from ice cores. Science, 336, 711714.Google Scholar
Schoepf, V., Stat, M., Falter, J. L., and McCulloch, M. T. (2015). Limits to the thermal tolerance of corals adapted to a highly fluctuating, naturally extreme temperature environment. Scientific Reports, 5. doi:10.1038/srep17639.Google Scholar
Schofield, P. J. (2009). Geographic extent and chronology of the invasion of non-native lionfish (Pterois volitans [Linnaeus 1758] and P. miles [Bennett 1828]) in the Western North Atlantic and Caribbean Sea. Aquatic Invasions, 4(3), 473479.Google Scholar
Scholz, D., Frisia, S., Borsato, A., et al. (2012). Holocene climate variability in north-eastern Italy: Potential influence of the NAO and solar activity recorded by speleothem data. Climate of the Past, 8, 13671383.Google Scholar
Schulte, P., Alegret, L., Arenillas, I., et al. (2010). The Chicxulub asteroid impact and mass extinction at the Cretaceous-Paleogene boundary. Science, 327, 12141218.Google Scholar
Scott, D. B., and Medioli, F. S. (1980). Quantitative studies of marsh foraminiferal distributions in Nova Scotia: Implications for sea level studies. Cushman Foundation for Foraminiferal Research, Special Publication 17, 158.Google Scholar
Scott, D. B., Medioli, F. S., and Schafer, C. T. (2001). Monitoring in Coastal Environments Using Foraminifera and Thecamoebian Indicators. Cambridge University Press.Google Scholar
Scott, K. (2013). International law in the Anthropocene: Responding to the geoengineering challenge. Michigan Journal of International Law, 34, 309358.Google Scholar
Scrivener, K. L., and Kirkpatrick, R. J. (2008). Innovation in use and research on cementitious material. Cement and Concrete Research, 38, 128136.Google Scholar
Seemann, J. (2013). The use of 13C and 15N isotope labeling techniques to assess heterotrophy of corals. Journal of Experimental Marine Biology and Ecology, 442(88). doi:10.1016/j.jembe.2013.01.004.Google Scholar
Seemann, J., Carballo-Bolaños, R., Berry, K. L., et al. (2012). Importance of heterotrophic adaptations of corals to maintain energy reserves. Proceedings of the 12th International Coral Reef Symposium, Cairns, Australia, 9–13 July 2012. 19A Human impacts on coral reef. http://icrs2012.com/proceedings/manuscripts/ICRS2012_19A_4.pdf.Google Scholar
Sen, I. S., and Peuckner-Ehrenbrink, B. (2012). Anthropogenic disturbance of element cycles at the Earth’s surface. Environmental Science and Technology, 46, 86018609.Google Scholar
Setälä, O., Fleming-Lehtinen, V., and Lehtiniemi, M. (2014). Ingestion and transfer of microplastics in the planktonic food web. Environmental Pollution, 185, 7783.Google Scholar
Shah, A. A., Hasan, F., Hameed, A., and Ahmed, S. (2008). Biological degradation of plastics: A comprehensive review. Biotechnology Advances, 26, 246265.Google Scholar
Shakhova, N., Semiletov, I., Leifer, I., et al. (2013). Ebullition and storm induced methane release from the East Siberian Arctic shelf. Nature Geoscience, 7(1), 6470.Google Scholar
Shakhova, N., Semiletov, I., Salyk, A., and Yusupov, V. (2010). Extensive methane venting to the atmosphere from sediments of the East Siberian Arctic shelf. Science, 327, 1246.Google Scholar
Shakun, J. D., Clark, P. U., He, F., et al. (2012). Global warming preceded by increasing carbon dioxide concentrations during the last deglaciation. Nature, 383, 4954.Google Scholar
Shaler, N. S. (1905). Man and the Earth. New York: Fox, Duffield & Co.Google Scholar
Shamberger, K. E. F., Cohen, A. L., Golbuu, Y., et al. (2014). Diverse coral communities in naturally acidified waters of a Western Pacific reef. Geophysical Research Letters, 41(2), 499504.Google Scholar
Shaviv, N. J. (2002). The spiral structure of the Milky Way, cosmic rays, and ice age epochs on Earth. New Astronomy, 8, 3977.Google Scholar
Shaviv, N. J., and Veizer, J. (2003). Celestial driver of Phanerozoic climate? GSA Today, 13(7), 410.Google Scholar
Shell International BV (2013). New Lens Scenarios. A Shift in Perspective for a World in Transition. http://shell.com/content/dam/royaldutchshell/documents/corporate/scenarios-newdoc.pdf.Google Scholar
Shen, B., Wu, J., and Zhao, Z. (2017). A ~150-year record of human impact in the Lake Wuliangsu (China) watershed: Evidence from polycyclic aromatic hydrocarbon and organochlorine pesticide distributions in sediments. Journal of Limnology, 76(1), 129136.Google Scholar
Shepherd, A., Ivins, E. R., Geruo, A., et al. (2012). A reconciled estimate of ice-sheet mass balance. Science, 338, 11831189.Google Scholar
Sheppard, C. (2006). Trawling the sea bed. Marine Pollution Bulletin, 52, 831835.Google Scholar
Sherlock, R. L. (1922). Man as a Geological Agent: An Account of His Action on Inanimate Nature. London: H. F. & G. Witherby.Google Scholar
Sherwood, O. A., Heikoop, J. M., Scott, D. B., et al. (2005a). Stable isotopic composition of deep-sea gorgonian corals Primnoa spp.: A new archive of surface processes. Marine Ecology Progress Series, 301, 135148.Google Scholar
Sherwood, O. A., Lehmann, M. F., Schubert, C. J., et al. (2011). Nutrient regime shift in the western North Atlantic indicated by compound-specific δ15N of deep-sea gorgonian corals. Proceedings of the National Academy of Sciences (USA), 108(3), 10111015.Google Scholar
Sherwood, O. A., Scott, D. B., Risk, M. J., and Guilderson, T. P. (2005b). Radiocarbon evidence for annual growth rings in a deep sea octocoral (Primnoa resedaeformis). Marine Ecology Progress Series, 301, 129134.Google Scholar
Shevenell, A. E. (2016). Drilling and modeling studies expose Antarctica’s Miocene secrets. Proceedings of the National Academy of Sciences (USA), 113(13), 34193421.Google Scholar
Shields, W. J., Ahn, S., Pietari, J., et al. (2014). Atmospheric fate and behavior of POPs. In O’Sullivan, G., and Sandau, C. D., eds., Environmental Forensics for Persistent Organic Pollutants. Amsterdam: Elsevier, pp. 199289.Google Scholar
Shields-Zhou, G. A., Porter, S., and Halverson, G. P. (2016). A new rock-based definition for the Cryogenian Period. Episodes, 39, 38.Google Scholar
Shimada, I., and Cavallaro, R. (1985). Monumental adobe architecture of the late prehispanic northern north coast of Peru. Journal de la Société des Américanistes, 71, 4178.Google Scholar
Sholkovitz, E. R., and Mann, D. R. (1984). The pore water chemistry of 239,240Pu and 137Cs in sediments of Buzzards Bay, Massachusetts. Geochimica et Cosmochimica Acta, 48, 11071114.Google Scholar
Shotyk, W., Appleby, P. G., Bicalho, B., et al. (2016). Peat bogs in northern Alberta, Canada reveal decades of declining atmospheric Pb contamination. Geophysical Research Letters, 43(18), 99649974.Google Scholar
Shotyk, W., Krachler, M., Martinez-Cortizas, A., et al. (2002). A peat bog record of natural, pre-anthropogenic enrichments of trace elements in atmospheric aerosols since 12370 14C yr BP, and their variation with Holocene climate change. Earth and Planetary Science Letters, 199(1), 2137.Google Scholar
Shotyk, W., Weiss, D., Appleby, P. G., et al. (1998). History of atmospheric lead deposition since 12,370 14C yr BP from a peat bog, Jura Mountains, Switzerland. Science, 281, 16351640.Google Scholar
Sigl, M., Winstrup, M., McConnell, J. R., et al. (2015). Timing and climate forcing of volcanic eruptions for the past 2,500 years. Nature, 523, 543549.Google Scholar
Simonyan, A. A. (1988). Zooplankton in changing conditions of a water body (case study: Lake Sevan). Unpublished PhD Thesis, Leningrad [in Russian].Google Scholar
Simonyan, A. A. (1991). Zooplankton of the Lake Sevan. Yerevan [in Russian].Google Scholar
Simpson, I. A. (1997). Relict properties of anthropogenic deep top soils as indicators of infield management in Marwick, West Mainland, Orkney. Journal of Archaeological Science, 24(4), 365380.Google Scholar
Sinkkonen, S., and Paasivirta, J. (2000). Degradation half-life times of PCDDs, PCDFs and PCBs for environmental fate modeling. Chemosphere, 40(9–11), 943949.Google Scholar
Skinner, L. C., Fallon, S., Waelbroeck, C., et al. (2010). Ventilation of the deep Southern Ocean and deglacial CO2 rise. Science, 328, 11471151.Google Scholar
Sloan, D. (1992). The Yerba Buena mud: Record of the last-interglacial predecessor of San Francisco Bay, California. GSA Bulletin, 104, 716727.Google Scholar
Slowikowski, A. (1995). “The greatest depository of archaeological material”: The role of pottery in ploughzone archaeology. In Shepherd, E., ed., Interpreting Stratigraphy. Hunstanton: Witley Press, 5, pp. 1520.Google Scholar
Smil, V. (2010). Energy Transitions: History, Requirements, Prospects. Cambridge: MIT Press.Google Scholar
Smil, V. (2015). It’s too soon to call this the Anthropocene. IEEE Spectrum, 52(6), 28.Google Scholar
Smith, A. G., Barry, T., Bown, P., et al. (2014). GSSPs, global stratigraphy and correlation. In Smith, D. G., Bailey, R. J., Burgess, P. M., and Fraser, A. J., eds., Strata and Time: Probing the Gaps in Our Understanding. Geological Society, London, Special Publications, 404, pp. 3767.Google Scholar
Smith, B. D., and Zeder, M. A. (2013). The onset of the Anthropocene. Anthropocene, 4, 813.Google Scholar
Smith, C. L., Fairchild, I. J., Spötl, C., et al. (2009). Chronology-building using objective identification of annual signals in trace element profiles of stalagmites. Quaternary Geochronology, 4, 1121.Google Scholar
Smith, D. M., Zalasiewicz, J. A., Williams, M., et al. (2010). Holocene drainage of the English Fenland: Roddons and their environmental significance. Proceedings of the Geologists’ Association, 121, 256269.Google Scholar
Smith, J. A., Andersen, T. J., Shortt, M., et al. (2016). Sub-ice-shelf sediments record twentieth century retreat of Pine Island Glacier. Nature, 541, 7780.Google Scholar
Smith, J. N., and Levy, E. M. (1990). Geochronology for polycyclic aromatic hydrocarbon contamination in sediments of the Saguenay Fjord. Environmental Science & Technology, 24, 874879.Google Scholar
Smith, S. D. A., and Markic, A. (2013). Estimates of marine debris accumulation on beaches are strongly affected by the temporal scale of sampling. PloS ONE, 8, 16.Google Scholar
Smith, S. J., Van Aardenne, J., Klimont, Z., et al. (2011). Anthropogenic sulfur dioxide emissions: 1850–2005. Atmospheric Chemistry and Physics, 11, 11011116.Google Scholar
Smol, J. P. (2008). Pollution of Lakes and Rivers: An Environmental Perspective, 2nd ed. Wiley-Blackwell.Google Scholar
Snir, A., Nadel, D., Groman-Yaroslavski, I., et al. (2015). The origin of cultivation and proto-weeds, long before Neolithic farming. PLoS ONE, 10(7), e0131422. doi:10.1371/journal.pone.0131422.Google Scholar
Snowball, I., Hounslow, M. W., and Nilsson, A. (2014). Geomagnetic and mineral magnetic characterisation of the Anthropocene. In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 119141.Google Scholar
Snyder, N. P., Rubin, D. M., Alpers, C. N., et al. (2004). Estimating accumulation rates and physical properties of sediment behind a dam: Englebright Lake, Yuba River, northern California. Water Resources Research, 40, W11301. doi:10.1029/2004WR003279.Google Scholar
Sokolov, B. S. (1986). Ekostratigrafiya, yeye mesto i rol’ ν sovremennoy stratigrafii. In Kaljo, D. L., and Klaamann, E. R., eds., Teoriya i opyt ekostratigrafii (The Theory and Practice of Ecostratigraphy), pp. 9–18; Inst. Geol. AN Eston. SSR, Valgus Press, Tallinn, 1986. (Using translated version: Ecostratigraphy, its place and role in modern stratigraphy, International Geology Review, 30[1], 3–10, 1988. doi:10.1080/00206818809465980.)Google Scholar
Solomon, S., Plattner, G.-K. Knutti, R., and Friedlingstein, P. (2009). Irreversible climate change due to carbon dioxide emissions. Proceedings of the National Academy of Sciences (USA), 106, 17041709.Google Scholar
Sousa, R., Pilotto, F., Aldridge, D. C. (2011). Fouling of European freshwater bivalves (Unionidae) by the invasive zebra mussel (Dreissena polymorpha). Freshwater Biology, 56, 867876.Google Scholar
Spencer, K., and O’Shea, F. T. (2014). The hidden threat of historical landfills on eroding and low-lying coasts. ECSA Bulletin, Summer 2014, 16–17.Google Scholar
Stanley, D. J., and Warne, A. G. (1997). Holocene sea-level change and early human utilization of deltas. GSA Today, 7, 17.Google Scholar
Stanley, G. D. Jr. (2001a). Introduction to reef ecosystems and their evolution. In Stanley, G. D. Jr., ed., The History and Sedimentology of Ancient Reef Systems. Topics in Geobiology, 17, pp. 139.Google Scholar
Stanley, G. D. Jr. (ed.) (2001b). The History and Sedimentology of Ancient Reef Systems. Topics in Geobiology, 17.Google Scholar
Stefani, M., and Vincenzi, M. (2005). The interplay of eustasy, climate and human activity in the late Quaternary depositional evolution and sedimentary architecture of the Po Delta system. Marine Geology, 222 –223, 1948.Google Scholar
Steffen, W., Broadgate, W., Deutsch, L., et al. (2015). The trajectory of the Anthropocene: The Great Acceleration. Anthropocene Review, 2(1), 8198.Google Scholar
Steffen, W., Crutzen, P. J., and McNeill, J. R. (2007). The Anthropocene: Are humans now overwhelming the great forces of Nature? Ambio, 36, 614621.Google Scholar
Steffen, W., Grinevald, J., Crutzen, P., and McNeill, J. (2011). The Anthropocene: Conceptual and historical perspectives. Philosophical Transactions of the Royal Society A, 369, 842867.Google Scholar
Steffen, W., Leinfelder, R., Zalasiewicz, J., et al. (2016). Stratigraphic and Earth System approaches in defining the Anthropocene. Earth’s Future, 8, 324345.Google Scholar
Steffen, W., Rockström, J., Richardson, K., et al. (2018). Trajectories of the Earth System in the Anthropocene. Proceedings of the National Academy of Sciences (USA).Google Scholar
Steffen, W., Sanderson, A., Tyson, P. D., et al. (2004). Global Change and the Earth System: A Planet under Pressure. The IGBP Book Series. Berlin, Heidelberg, New York: Springer-Verlag.Google Scholar
Steffensen, J. P., Andersen, K. K., Bigler, M., et al. (2008). High-resolution Greenland ice core data show abrupt climate change happens in a few years. Science, 321, 680683.Google Scholar
Steig, E. J., Ding, Q., White, J. C., et al. (2013). Recent climate and ice-sheet changes in West Antarctica compared with the past 2,000 years. Nature Geoscience, 6, 372375.Google Scholar
Stein, R., Fahl, K., Schade, I., et al. (2017). Holocene variability in sea ice cover, primary production, and Pacific-Water inflow and climate change in the Chukchi and East Siberian Seas (Arctic Ocean). Journal of Quaternary Science, 32(3), 362379.Google Scholar
Steinhilber, F., Abreu, J. A., Beer, J., et al. (2012). 9,400 years of cosmic radiation and solar activity from ice cores and tree rings. Proceedings of the National Academy of Sciences, 109(16), 59675971.Google Scholar
Stenni, B., Curran, M. A. J., Abram, N. J., et al. (2017). Antarctic climate variability on regional and continental scales over the last 2000 years. Climate of the Past, 13, 16091634.Google Scholar
Stickley, C. E. (2014). The sea ice thickens. Nature Geoscience, 7, 165166.Google Scholar
Stocker, T. F. (1998). The seesaw effect. Science, 282, 6162.Google Scholar
Stoppani, A. (1873). Corso di Geologia, vol. 2. Geologia Stratigrafica. Milan: G. Bernardoni e G. Brigola.Google Scholar
Stordeur, D., and Khawam, R. (2007). Les crânes surmodelés de Tell Aswad (PPNB, Syrie): Premier regard sur l’ensemble, premières réflexions. Syria, 84, 532.Google Scholar
Stow, D. A. V. (2001). Deep sea sediment drifts. Steele. Steele, J. H., Thorpe, S. A., and Turekian, K. K. (eds.), Encyclopaedia of Ocean Sciences. London: Academic Press.Google Scholar
Strasser, T. F., Panagopoulou, E., Runnels, C. N., et al. (2010). Stone Age seafaring in the Mediterranean: Evidence from the Plakias Region for Lower Palaeolithic and Mesolithic habitation of Crete. Hesperia, 79, 145190.Google Scholar
Strecker, A. L., and Arnott, S. E. (2010). Complex interactions between regional dispersal of native taxa and an invasive species. Ecology, 91, 10351047.Google Scholar
Stuart, A. J., Kosintsev, P. A., Higham, T. F. G., and Lister, A. M. (2004). Pleistocene to Holocene extinction dynamics in giant deer and woolly mammoth. Nature, 431, 684689.Google Scholar
Stuart, A. J., Sulerzhitsky, L. D., Orlova, L. A., et al. (2002). The latest woolly mammoths (Mammuthus primigenius Blumenbach) in Europe and Asia: A review of the current evidence. Quaternary Science Reviews, 21, 15591569.Google Scholar
Sucharovà, J., Suchara, I., Hola, M., et al. (2012). Top-/bottom-soil ratios and enrichment factors: What do they really show? Applied Geochemistry, 27(1), 138145.Google Scholar
Summerhayes, C. P. (2010). Climate change: A creeping catastrophe. Bulletin of the World Health Organization, 88(6). http://dx.doi.org/10.1590/S0042–96862010000600007.Google Scholar
Summerhayes, C. P. (2015). Earth’s Climate Evolution. Chichester: Wiley.Google Scholar
Summerhayes, C. P., Ellis, J. P., and Stoffers, P. (1985). Estuaries as sinks for sediment and industrial waste: A case history from the Massachusetts coast. Contributions to Sedimentology, 14, 147.Google Scholar
Surovell, T. A., Holliday, V. T., Gingerich, J. A. M., et al. (2009). An independent evaluation of the Younger Dryas extraterrestrial impact hypothesis. Proceedings of the National Academy of Sciences (USA), 106, 1815518158.Google Scholar
Svendsen, J. I., Alexanderson, H., Astakhov, V. I., et al. (2004). Late Quaternary ice sheet history of northern Eurasia. Quaternary Science Reviews, 23, 12291271.Google Scholar
Swart, N. (2017). Natural causes of Arctic sea-ice loss. Nature Climate Change, 7, 239241.Google Scholar
Swart, P. K., Greer, L., Rosenheim, B. E., et al. (2010). The 13C Suess effect in scleractinian corals mirror changes in the anthropogenic CO2 inventory of the surface oceans. Geophysical Research Letters, 37, L05604.Google Scholar
Swindles, G. T., Watson, E., Turner, T. E., et al. (2015). Spheroidal carbonaceous particles are a defining stratigraphic marker for the Anthropocene. Scientific Reports, 5, 10264. doi:10210.11038/srep10264.Google Scholar
Syvitski, J. P. M. (1993). Glaciomarine environments in Canada: An overview. Canadian Journal of Earth Sciences, 30, 354371.Google Scholar
Syvitski, J. P. M. (Editor) (2003a). The supply and flux of sediment along hydrological pathways: Anthropogenic influences at the global scale. Global and Planetary Change, 39(1/2), 1199.Google Scholar
Syvitski, J. P. M. (2003b). Sediment fluxes and rates of sedimentation. In Middleton, G. V., ed., Encyclopedia of Sediments and Sedimentary Rocks. Dordrecht, Netherlands: Kluwer Academic Publishers, pp. 600606.Google Scholar
Syvitski, J. P. M. (2008). Deltas at Risk. Sustainability Science, 3, 2332.Google Scholar
Syvitski, J. P. M., Burrell, D. C., and Skei, J. M. (1987). Fjords: Processes & Products. New York: Springer-Verlag.Google Scholar
Syvitski, J. P. M., Harvey, N., Wollanski, E., et al. (2005a). Dynamics of the coastal zone. In Crossland, C. J., Kremer, H. H., Lindeboom, H. J., et al., eds., Global Fluxes in the Anthropocene. Berlin: Springer, pp. 3994.Google Scholar
Syvitski, J. P. M., and Kettner, A. (2011). Sediment flux and the Anthropocene. Philosophical Transactions of the Royal Society A, 369(1938), 957975.Google Scholar
Syvitski, J. P. M., Kettner, A. J., Correggiari, A., and Nelson, B. W. (2005b). Distributary channels and their impact on sediment dispersal. Marine Geology, 222 –223, 7594.Google Scholar
Syvitski, J. P. M., Kettner, A. J., Overeem, I., et al. (2009). Sinking deltas due to human activities. Nature Geoscience, 2, 681689.Google Scholar
Syvitski, J. P., Kettner, A. J., Overeem, I., et al. (2017). Latitudinal controls on siliciclastic sediment production and transport. SEPM Special Issue No. 108, Latitudinal Controls on Stratigraphic Models and Sedimentary Concepts, 1–15, Tulsa OK.Google Scholar
Syvitski, J. P. M., Kettner, A., Peckham, S. D., and Kao, S. J. (2005c). Predicting the flux of sediment to the coastal zone: Application to the Lanyang watershed, northern Taiwan. Journal of Coastal Research, 21, 580587.Google Scholar
Syvitski, J. P. M., and Milliman, J. D. (2007). Geology, geography and humans battle for dominance over the delivery of sediment to the coastal ocean. Journal of Geology, 115, 119.Google Scholar
Syvitski, J. P. M., and Saito, Y. (2007). Morphodynamics of deltas under the influence of Humans. Global and Planetary Changes, 57, 261182.Google Scholar
Syvitski, J. P. M., Vörösmarty, C., Kettner, A. J., and Green, P. (2005d). Impact of humans on the flux of terrestrial sediment to the global coastal ocean. Science, 308, 376380.Google Scholar
Szabó, J. (2010). Anthropogenic geomorphology: Subject and system. In Szabó, J., Dávid, L., and Lóczy, D., eds., Anthropogenic Geomorphology: A guide to Man-Made Landforms. Dordecht, Heidelberg, London, New York: Springer, pp. 310.Google Scholar
Ta, T. K. O., Nguyen, V. L., Tateishi, M., et al. (2002). Sediment facies and late Holocene progradation of the Mekong River delta in Bentre Province, southern Vietnam: An example of evolution from a tide-dominated to a tide- and wave- dominated delta. Sedimentary Geology, 152, 313325.Google Scholar
Tagliabue, A., Aumont, O., and Bopp, L. (2014). The impact of different external sources of iron on the global carbon cycle. Geophysical Research Letters. doi:10.1002/2013GL059059.Google Scholar
Tanabe, S., Saito, Y., Vu, Q. L., et al. (2006). Holocene evolution of the Song Hong (Red River) delta system, northern Vietnam. Sedimentary Geology, 187, 2961.Google Scholar
Tansel, B., and Yildiz, B. S. (2011). Goal-based waste management strategy to reduce persistence of contaminants in leachate at municipal solid waste landfills. Environment, Development and Sustainability, 13, 821831.Google Scholar
Targulian, V. O., and Goryachkin, S. V. (2008). Soil memory: Types of records, carriers, hierarchy and diversity. Revista Mexicana Ciencias Geologica, 21, 18.Google Scholar
Tarolli, P., Preti, F., and Romano, N. (2014). Terraced landscapes: From an old best practice to a potential hazard for soil degradation due to land abandonment. Anthropocene, 6, 1025.Google Scholar
Tarolli, P., and Sofia, G. (2016). Human topographic signatures and derived geomorphic processes. Geomorphology, 255, 140161.Google Scholar
Terrington, R. L., Silva, É. C. N., Waters, C. N., et al. (2018). Quantifying anthropogenic modification of the shallow geosphere in central London, UK. Geomorphology, 319, 1534.Google Scholar
Tessler, Z., Vörösmarty, C., Grossberg, M., et al. (2015). Profiling risk and sustainability in coastal deltas of the world. Science, 349(6248), 638643.Google Scholar
Tesson, M., Labaune, C., and Gensous, B. (2005). Small rivers contribution to the Quaternary evolution of a Mediterranean littoral system: The western gulf of Lion, France. Marine Geology, 222 –223, 313334.Google Scholar
Thevenon, F., Guédron, S., Chiaradia, M., et al. (2011). (Pre-) historic changes in natural and anthropogenic heavy metals deposition inferred from two contrasting Swiss Alpine lakes. Quaternary Science Reviews, 30(1), 224233.Google Scholar
Thiede, J., Jessen, C., Knutz, P., et al. (2010). Millions of years of Greenland Ice Sheet history recorded in ocean sediments. Polarforschung, 80(3), 141159.Google Scholar
Thiengo, S. C., Faraco, F. A., Salgado, N. C., et al. (2007). Rapid spread of an invasive snail in South America: The giant African snail, Achatina fulica in Brasil. Biological Invasions, 9, 693702.Google Scholar
Thomas, E. (2003). Benthic foraminiferal record across the Initial Eocene Thermal Maximum, Southern Ocean Site 690: Causes and consequences of globally warm climates in the Early Paleogene. Geological Society of America Special Paper, 369, 319331.Google Scholar
Thomas, E. R., Hosking, J. S., Tuckwell, R. R., et al. (2015). Twentieth century increase in snowfall in coastal West Antarctica. Geophysical Research Letters, 42, 93879393.Google Scholar
Thomas, E. R., Van Wessem, J. M., Roberts, J., et al. (2017). Regional Antarctic show accumulation over the past 1000 years. Climate of the Past, 13, 14911513.Google Scholar
Thomas, W. L. Jr. (ed.) (1956). Man’s Role in Changing the Face of the Earth. Wenner-Gren Foundation for Anthropological Research Symposium, Princeton, June 1955. University of Chicago Press.Google Scholar
Thompson, L. G. (2010). Climate change: The evidence and our options. Behavior Analyst, 33(2), 153170.Google Scholar
Thompson, L. G., Mosley-Thompson, E., Brecher, H., et al. (2006). Abrupt tropical climate change: Past and present. Proceedings of the National Academy of Sciences (USA), 103(28), 1053610543.Google Scholar
Thompson, R., Battarbee, R. W., O’Sullivan, P. E., and Oldfield, F. (1975). Magnetic susceptibility of lake sediments. Limnology and Oceanography, 20, 687698.Google Scholar
Thompson, R. C., Moore, C., vom Saal, F. S., and Swan, S. H. (2009). Plastics, the environment and human health: Current consensus and future trends. Philosophical Transactions of the Royal Society B, 364, 21532166.Google Scholar
Thompson, R., and Morton, D. J. (1978). Magnetic susceptibility and particle-size distribution in recent sediments of the Loch Lomond drainage basin, Scotland. Journal of Sedimentary Petrology, 49, 801811.Google Scholar
Thompson, R., and Oldfield, F. (1986). Environmental Magnetism. London: Allen & Unwin.Google Scholar
Thomson, J., Brown, L., Nixon, S., et al. (2000). Bioturbation and Holocene sediment accumulation fluxes in the north-east Atlantic Ocean (Benthic Boundary Layer experiment sites). Marine Geology, 169, 2139.Google Scholar
Tickell, C. (2011). Societal responses to the Anthropocene. Philosophical Transactions of the Royal Society A, 369, 926932.Google Scholar
Tierney, J. E., Abram, N. J., Anchukaitis, K. J., et al. (2015). Tropical sea surface temperatures for the past four centuries reconstructed from coral archives. Paleoceanography, 30, 226252.Google Scholar
Tobiszewski, M., and Namieśnik, J. (2012). PAH diagnostic ratios for the identification of pollution emission sources. Environmental Pollution, 162, 110119.Google Scholar
Toggweiler, R. (2008). Origin of the 100,000-year timescale in Antarctic temperatures and atmospheric CO2. Paleoceanography, 23, PA2211.Google Scholar
Tollefson, J. (2017). Larsen C’s big divide: Collapse of nearby Antarctic ice shelves offers a glimpse of the future. Nature, 542, 402403.Google Scholar
Torres Camprubi, A. (2016). Statehood under Water: Challenges of Sea-Level Rise to the Continuity of Pacific Island States. Boston/Leiden: Brill/Martinus Nijhoff.Google Scholar
Toynbee, J. (1996). Death and Burial in the Roman World. Baltimore, MD: John Hopkins University Press, reprint.Google Scholar
Tubau, X., Canals, M., Lastras, G., and Rayo, X. (2015). Marine litter on the floor of deep submarine canyons of the Northwestern Mediterranean Sea: The role of hydrodynamic processes. Progress in Oceanography, 134, 379403.Google Scholar
Tully, J. (2009). A Victorian ecological disaster: Imperialism, the telegraph, and Gutta-Percha. Journal of World History, 20, 559579.Google Scholar
Turner, B. L. II, Clark, W. C., Kates, R. W., et al., (eds.) (1990). The Earth as Transformed by Human Action: Global and Regional Changes in the Biosphere over the Past 300 Years. Cambridge University Press.Google Scholar
Turner, J., Binschadler, R., Convey, P., et al. (eds.) (2009a). Antarctic Climate Change and the Environment. Cambridge, UK: Scientific Committee on Antarctic Research, pp. 389393.Google Scholar
Turner, J., Comiso, J. C., Marshall, G. J., et al. (2009b). Non-annular atmospheric circulation change induced by stratospheric ozone depletion and its role in the recent increase of Antarctic sea ice extent. Geophysical Research Letters, 36, L08502. doi:10.1029/2009GL037524.Google Scholar
Turner, J., Lu, H., White, I., et al. (2016). Absence of 21st century warming on Antarctic Peninsula consistent with natural variability. Nature, 535, 411415.Google Scholar
Turner, J., Summerhayes, C. P., Sparrow, M. D., et al. (2017). Antarctic climate change and the environment – 2017 update. Information Paper 80, Antarctic Treaty Consultative Meeting 40, Beijing, China.Google Scholar
Turra, A., Manzano, A. B., Dias, R. J. S., et al. (2014). Three-dimensional distribution of plastic pellets in sandy beaches: Shifting paradigms. Scientific Reports, 4, 4435.Google Scholar
Turvey, S., Pitman, R. L., Taylor, B. L., et al. (2007). First human-caused extinction of a cetacean species? Biology Letters, 3, 537540.Google Scholar
Tyndall, J. (1868). On Radiation, the 1865 Rede Lecture. New York: D. Appleton and Co. Also included in his 1871 book Fragments of Science, available as a Project Gutenberg e-Book.Google Scholar
Tyrrell, T. (2011). Anthropogenic modification of the oceans. Philosophical Transactions of the Royal Society A, 369, 887908.Google Scholar
Tyrrell, T., and Zeebe, R. E. (2004). History of carbonate ion concentration over the last 100 million years. Geochimica Cosmochimica Acta, 68, 35213530.Google Scholar
Tzedakis, P. C., Crucifix, M., Mitsui, T., and Wolff, E. W. (2017). A simple rule to determine which insolation cycles lead to interglacials. Nature, 452, 427432.Google Scholar
Uhrqvist, O., and Linnér, B.-O. (2015). Narratives of the past for Future Earth: The historiography of global environmental change research. Anthropocene Review, 2(2), 159173.Google Scholar
Ulanowicz, R. E. (1997). Ecology, the Ascendent Perspective. New York: Columbia University Press.Google Scholar
Underwood, J. R. (2001). Anthropic rocks as a fourth basic class. Environmental & Engineering Geoscience, 7(1), 104110.Google Scholar
UNEP (United Nations Environment Programme) (2013). Environmental Risks and Challenges of Anthropogenic Metals Flows and Cycles: A Report of the Working Group on the Global Metal Flows to the International Resource Panel.Google Scholar
UNESCO (2011). Sediment issues and sediment management in large river basins: Interim case study synthesis report. International Sediment Initiative Technical Documents in Hydrology, IRTCES 2011.Google Scholar
United Nations Convention on the Law of the Sea (1982). United Nations Treaty Series, 1833, 3.Google Scholar
United Nations, Department of Economic and Social Affairs, Population Division (2014). World Urbanization Prospects: The 2014 Revision, Highlights (ST/ESA/SER. A/352).Google Scholar
United Nations, Department of Economic and Social Affairs, Population Division (2015). World Urbanization Prospects: The 2015 Revision, Key Findings and Advance Tables. Working Paper No. ESA/P/WP.241.Google Scholar
United Nations Scientific Committee on the Effects of Atomic Radiation (UNSCEAR) (2000). Sources and Effects of Ionizing Radiation, 2000 Report, vol. 1. New York: United Nations.Google Scholar
United States Geological Survey (USGS) (2010). Aluminium statistics. In Kelly, T. D., and Matos, G. R., eds., Historical Statistics for Mineral and Material Commodities in the United States. US Geological Survey Data Series, 140. http://pubs.usgs.gov/ds/2005/140/ (accessed 16 December 2012).Google Scholar
United States Geological Survey (USGS) (2016). http://minerals.usgs.gov/minerals/pubs/commodity/cement/.Google Scholar
US CLIVAR Project Office (2012). Understanding the Dynamic Response of Greenland’s Marine Terminating Glaciers to Oceanic and Atmospheric Forcing. A whitepaper by the US CLIVAR Working Group on Greenland Ice Sheet-Ocean Interactions (GRISO), Report 2012–2, US CLIVAR Project Office, Washington, DC, 20006.Google Scholar
Vai, G. B. (2007). A history of chronostratigraphy. Stratigraphy, 4, 8397.Google Scholar
Van Cappellen, P., and Wang, Y. (1995). Metal cycling in surface sediments: Modeling the interplay of transport and reaction. In Allen, H. E., ed., Metal Contaminated Aquatic Sediments. Chelsea, MI: Ann Arbor Press, pp. 2164.Google Scholar
Van Cauwenberghe, L., Vanreusel, A., Mees, J., and Janssen, C. R. (2013). Microplastic pollution in deep-sea sediments. Environmental Pollution, 182, 495499.Google Scholar
van der Gon, H. D., van het Bolscher, M., Visschedijk, A., and Zandveld, P. (2007). Emissions of persistent organic pollutants and eight candidate POPs from UNECE–Europe in 2000, 2010 and 2020 and the emission reduction resulting from the implementation of the UNECE POP protocol. Atmospheric Environment, 41(40), 92459261.Google Scholar
Van der Velde, G., and Rajagopal, S. (2010). Zebra Mussels in Europe. Leiden: Backhuys Publishers.Google Scholar
van der Voet, E., Salminen, R., Eckelman, M., et al. (1995). Metal cycling in surface sediments: Modeling the interplay of transport and reaction. In Allen, H. E., ed., Metal Contaminated Aquatic Sediments. Chelsea: Ann Arbor Press, pp. 2164.Google Scholar
Van Kranendonk, M. J., Altermann, W., Beard, B. L., et al. (2012). A Chronostratigraphic Division of the Precambrian. In Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G., eds., The Geologic Time Scale 2012, vol. 1. Elsevier, pp. 299392.Google Scholar
Van Oppen, M. J. H., Oliver, J. K., Putnam, H. M., and Gates, R. D. (2015). Building coral reef resilience through assisted evolution. Proceedings of the National Academy of Sciences (USA), 112(8), 23072313.Google Scholar
Vandenberghe, N., Hilgen, F. J., and Speijer, R. P. (2012). The Paleogene Period. In Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G., eds., The Geological Time Scale 2012. Elsevier, pp. 853922.Google Scholar
Vandiver, P. B., Soffer, O., Klima, B., and Svoboda, J. (1989). The origins of ceramic technology at Dolni Věstonice, Czechoslovakia. Science, 246(4933), 10021008.Google Scholar
Vane, C. H., Chenery, S. R., Harrison, I., et al. (2011). Chemical signatures of the Anthropocene in the Clyde estuary, UK: Sediment-hosted Pb, 207/206Pb, total petroleum hydrocarbon, polyaromatic hydrocarbon and polychlorinated biphenyl pollution records. Philosophical Transactions of the Royal Society A, 369, 10851111.Google Scholar
Velicogna, I., Sutterley, T. S., and van den Broeke, M. R. (2014). Regional acceleration in ice mass loss from Greenland and Antarctica using GRACE time-variable gravity data. Journal of Geophysical Research Space Physics, 41, 81308137.Google Scholar
Vella, C., Fleury, T.-J., Raccasi, G., et al. (2005). Evolution of the Rhône delta plain in the Holocene. Marine Geology, 222 –223, 235265.Google Scholar
Velzboer, I., Kwadijk, C. J. A. F., and Koelmans, A. A. (2014). Strong sorption of PCBs to nanoplastics, microplastics, carbon tubules, and fullerenes. Environmental Science and Technology, 48, 48694876.Google Scholar
Vernadsky, V. I. (1924). La Géochimie. Paris: Librairie Félix Acan. (Lectures at the Sorbonne in 1922–1923).Google Scholar
Vernadsky, V. I. (1945). The Biosphere and the Noosphere. American Scientist, 33(1), 112.Google Scholar
Vernadsky, V. I. (1997). Scientific Thought as a Planetary Phenomenon. Translated from the Russian [1938, 1977, 1991] by Starostin, B. A.. Moscow: Nongovernmental Ecological V. I. Foundation.Google Scholar
Vernadsky, V. I. [1926](1998). The Biosphere. Translated from the Russian by Langmuir, D. B., revised and annotated by McMenamin, M. A. S.. New York: Copernicus (Springer-Verlag).Google Scholar
Veron, J. E. N. (1995). Corals in Space and Time: The Biogeography & Evolution of the Scleractinia. Ithaca, NY: Cornell University Press.Google Scholar
Veron, J. E. N. (2008). Mass extinctions and ocean acidification: Biological constraints on geological dilemmas. Coral Reefs, 27, 459472.Google Scholar
Verpoorter, C., Kutser, T., Seekell, D. A., and Tranvik, L. J. (2014). A global inventory of lakes based on high-resolution satellite imagery. Geophysical Research Letters, 41(18), 63966402.Google Scholar
Vidas, D. (2010). Responsibility for the seas. In Vidas, D., ed., Law, Technology and Science for Oceans in Globalization. Boston/Leiden: Brill/Martinus Nijhoff, pp. 340.Google Scholar
Vidas, D. (2011). The Anthropocene and the international law of the sea. Philosophical Transactions of the Royal Society A, 369, 909925.Google Scholar
Vidas, D. (2015). The Earth in the Anthropocene – and the world in the Holocene? European Society of International Law (ESIL) Reflections, 4(6), 17.Google Scholar
Vidas, D., Fauchald, O. K., Jensen, Ø., and Tvedt, M. W. (2015a). International law for the Anthropocene? Shifting perspectives in regulation of the oceans, environment and genetic resources. Anthropocene, 9, 113.Google Scholar
Vidas, D., Zalasiewicz, J., and Williams, M. (2015b). What is the Anthropocene: And why is it relevant for international law? Yearbook of International Environmental Law, 25, 323.Google Scholar
Vienna Convention on the Law of Treaties (1969). United Nations Treaty Series, 1155, 331.Google Scholar
Viers, J., Dupré, B., and Gaillardet, J. (2009). Chemical composition of suspended sediments in world rivers: New insights from a new database. Science of the Total Environment, 407(2), 853868.Google Scholar
Vigney, J.-D. (2011). The origins of animal domestication and husbandry: A major change in the history of humanity and the biosphere. Comptes Rendus Biologies, 334, 171181.Google Scholar
Villavicencio, N. A., Lindsey, E. L., Martin, F. M., et al. (2015). Combination of humans, climate, and vegetation change triggered Late Quaternary megafauna extinction in the Última Esperanza region, southern Patagonia, Chile. Ecography, 38, 116.Google Scholar
Villmoare, B., Kimbel, W. H., Seyoum, C., et al. (2015). Early Homo at 2.8 Ma from Ledi-Geraru, Afar, Ethiopia. Science, 347, 13521355.Google Scholar
Vinther, B. M., Buchardt, S. L., Clausen, H. B., et al. (2009). Holocene thinning of the Greenland ice sheet. Nature, 461, 385388.Google Scholar
Vinuales, J. E. (2016). Law and the Anthropocene. C-EENRG Working Papers, 2016–4 (September), University of Cambridge, 1–72.Google Scholar
Vogel, J. C. (1970). Groningen radiocarbon dates IX. Radiocarbon, 12(2), 444471.Google Scholar
von Glasow, R., Bobrowski, N., and Kern, C. (2009). The effects of volcanic eruptions on atmospheric chemistry. Chemical Geology, 263, 131142.Google Scholar
Vörösmarty, C., Meybeck, M., Fekete, B., et al. (2003). Anthropogenic sediment retention: Major global-scale impact from the population of registered impoundments. Global and Planetary Change, 39, 169190.Google Scholar
Wacey, D., Kilburn, M. R., Saunders, M., et al. (2011). Microfossils of sulphur-metabolizing cells in 3.4-billion-year-old rocks of Western Australia. Nature Geoscience, 4, 698702.Google Scholar
Wadhams, P. (2016). A Farewell to Ice: A Report from the Arctic. London: Allen Lane.Google Scholar
Wagreich, M., and Draganits, E. (2018). Early mining and smelting lead anomalies in geological archives as potential stratigraphic markers for the base of an early Anthropocene. The Anthropocene Review. doi:10.1177/2053019618756682.Google Scholar
Walker, M. J. C., Berkelhammer, M., Björck, S., et al. (2012). Formal subdivision of the Holocene Series/Epoch: A discussion paper by a working group of INTIMATE (integration of ice-core, marine and terrestrial records) and the Subcommission on Quaternary Stratigraphy (International Commission on Stratigraphy). Journal of Quaternary Science, 27, 649659.Google Scholar
Walker, M. J. C., Berkelhammer, M., Björck, S., et al. (2016). Formal subdivision of the Holocene Series/Epoch: Three proposals by a working group of members of INTIMATE (integration of ice-core, marine and terrestrial records) and the Subcommission on Quaternary Stratigraphy. Unpublished proposal submitted to the ICS Subcommission on Quaternary Stratigraphy.Google Scholar
Walker, M. J. C., Johnsen, S., Rasmussen, O. S., et al. (2009). Formal definition and dating of the GSSP (Global Stratotype Section and Point) for the base of the Holocene using the Greenland NGRIP ice core, and selected auxiliary records. Journal of Quaternary Science, 24, 317.Google Scholar
Walling, D. E., and Fang, D. (2003). Recent trends in the suspended sediment loads of the world’s rivers. Global and Planetary Change, 39, 111126.Google Scholar
Walsh, J. E. (2009). A comparison of Arctic and Antarctic climate change, present and future. Antarctic Science, 21(3), 179188.Google Scholar
Walsh, J. E., and Chapman, W. L. (2001). 20th-century sea-ice variations from observational data. Annals of Glaciology, 33, 444448.Google Scholar
Walter, R. C., and Merritts, D. J. (2008). Natural streams and the legacy of water-powered mills. Science, 319, 299304.Google Scholar
Wang, H., Saito, Y., Zhang, Y., et al. (2011). Recent changes of sediment flux to the western Pacific Ocean from major rivers in East and Southeast Asia. Earth Science Reviews, 108, 80100.Google Scholar
Wang, H., Yang, Z., Saito, Y., et al. (2007). Stepwise decreases of the Huanghe (Yellow River) sediment load (1950–2004): Impacts from climate changes and human activities. Global Planetary Change, 57, 331354.Google Scholar
Wang, M., and Overland, J. E. (2009). A sea ice free summer Arctic within 30 years? Geophysical Research Letters, 36(7). doi:10.1029/2009GL037820.Google Scholar
Wania, F. (2003). Assessing the potential of persistent organic chemicals for long-range transport and accumulation in polar regions. Environmental Science & Technology, 37(7), 13441351.Google Scholar
Wanner, H., Beer, J., Butikofer, J., et al. (2008). Mid- to Late Holocene climate change: An overview. Quaternary Science Reviews, 27, 17911828.Google Scholar
Ward, C. V., Tocheri, M. W., Plavcan, J. M., et al. 2014. Early Pleistocene third metacarpal from Kenya and the evolution of modern human-like hand morphology. Proceedings of the National Academy of Sciences (USA), 111, 121124.Google Scholar
Ward, L. (2015). The London County Council Bomb Damage Maps. Thames & Hudson.Google Scholar
Waste Online (2004). History of waste and recycling information sheet. http://dl.dropbox.com/u/21130258/resources/informationsheets/historyofwaste.htm.Google Scholar
Waters, C. N., Graham, C., Tapete, D., et al. (2018, in press). Recognising anthropogenic modification of the subsurface in the geological record. Quarterly Journal of Engineering Geology and Hydrogeology.Google Scholar
Waters, C. N., Northmore, K., Prince, G., et al. (1996). Volume 2: A technical guide to ground conditions. In Waters, C. N., Northmore, K., Prince, G., and Marker, B. R., eds., A Geological Background for Planning and Development in the City of Bradford Metropolitan District. British Geological Survey Technical Report WA/96/1.Google Scholar
Waters, C. N., Syvitski, J. P. M., Gałuszka, A., et al. (2015). Can nuclear weapons fallout mark the beginning of the Anthropocene Epoch? Bulletin of the Atomic Scientists, 71(3), 4657.Google Scholar
Waters, C. N., and Zalasiewicz, J. (2018). Concrete: The most abundant novel rock type of the Anthropocene. In DellaSala, D., and Goldstein, M. I., eds., Encyclopedia of the Anthropocene, vol. 1. Oxford: Elsevier. https://doi.org/10.1016/B978-0-12-809665-9.09775 -5.Google Scholar
Waters, C. N., Zalasiewicz, J., Summerhayes, C., et al. (2016). The Anthropocene is functionally and stratigraphically distinct from the Holocene. Science, 351(6269), 137.Google Scholar
Waters, C. N., Zalasiewicz, J., Summerhayes, C., et al. (2018). Global Boundary Stratotype Section and Point (GSSP) for the Anthropocene Series: Where and how to look for a potential candidate. Earth-Science Reviews, 178, 379429.Google Scholar
Waters, C. N., Zalasiewicz, J., Williams, M., et al. (eds.) (2014). A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395.Google Scholar
Watling, L., and Norse, E. A. (1998). Disturbance of the seabed by mobile fishing gear: A comparison to forest clearcutting. Conservation Biology, 12, 11801197.Google Scholar
Watmough, S. A. (1999). Monitoring historical changes in soil and atmospheric trace metal levels by dendrochemical analysis. Environmental Pollution, 106, 391403.Google Scholar
Watters, D. L., Yoklavich, M. M., Love, M. S., and Schroeder, D. M. (2010). Assessing marine debris in deep seafloor habitats off California. Marine Pollution Bulletin, 60, 131138.Google Scholar
Weaver, P. P. E., Wynn, R. B., Kenyon, N. H., and Evans, J. (2000). Continental margin sedimentation, with special reference to the north-east Atlantic margin. Sedimentology, 47(1), 239256.Google Scholar
Weaver, T. D., and Roseman, C. C. (2008). New developments in the genetic evidence for modern human origins. Evolutionary Anthropology, 17, 6980.Google Scholar
Webby, B. D. (1998). Steps toward a global standard for Ordovician stratigraphy. Newsletters on Stratigraphy, 36(1), 133.Google Scholar
Weber, C., Pusch, S., and Opatz, T. (2017). Correspondence: Polyethylene bio-degradation by caterpillars? Current Biology, 27(15), R744R745.Google Scholar
Wei, G., McCulloch, M. T., Mortimer, G., et al. (2009). Evidence for ocean acidification in the Great Barrier Reef of Australia. Geochimica et Cosmochimica Acta, 73, 23322346.Google Scholar
Wei, S., Wang, Y., Lam, J. C., et al. (2008). Historical trends of organic pollutants in sediment cores from Hong Kong. Marine Pollution Bulletin, 57(6), 758766.Google Scholar
Weidman, C. R., and Jones, G. A. (1993). A shell-derived time history of bomb 14C on Georges Bank and its Labrador Sea implications. Journal of Geophysical Research, 98(C8), 1457714588.Google Scholar
Weightman, G. (2007). The Industrial Revolutionaries: The Creation of the Modern World, 1776–1914. New York: Grove Press.Google Scholar
Wellman, C., Osterloff, P. L., and Mohiuddin, U. (2003). Fragments of the earliest land plants. Nature, 425, 282285.Google Scholar
Wells, S., and Hanna, R. (1992). The Greenpeace Book of Coral Reefs. London: Cameron.Google Scholar
West, G. (2017). Scale. New York: Penguin Press.Google Scholar
Westbrook, G. K., Thatcher, K. E., Rohling, E. J., et al. (2009). Escape of methane gas from the seabed along the West Spitzbergen continental margin. Geophysical Research Letters, 36(15). doi:10.1029/2009GL039191.Google Scholar
Wetzel, R. G. (2001). Limnology: Lake and River Systems. San Diego: Academic Press.Google Scholar
White, S. J. O., and Hemond, H. F. (2012). The anthrobiogeochemical cycle of indium: A review of the natural and anthropogenic cycling of indium in the environment. Critical Reviews in Environmental Science and Technology, 42(2), 155186.Google Scholar
Whiteman, G., Hope, C., and Wadhams, P. (2013). Vast costs of Arctic change. Nature, 499, 401403.Google Scholar
Whitmee, S., Haines, A., Beyrer, C., et al. (2015). Safeguarding human health in the Anthropocene epoch: Report of The Rockefeller Foundation–Lancet Commission on planetary health. The Lancet, 386, 19732028.Google Scholar
Wikipedia (2017). World population estimates. https://en.wikipedia.org/wiki/World_population_estimates.Google Scholar
Wilkinson, B. H., and McElroy, B. J. (2007). The impact of humans on continental erosion and sedimentation. Geological Society of America Bulletin, 119, 140156.Google Scholar
Wilkinson, I. P., and Gulakyan, S. Z. (2010). Holocene to recent Ostracoda of Lake Sevan, Armenia: Biodiversity and ecological controls. Stratigraphy, 7, 301315.Google Scholar
Wilkinson, I. P., Poirier, C., Head, M. J., et al. (2014). Microbiotic signatures of the Anthropocene in marginal marine and freshwater palaeoenvironments. In Waters, C. N., Zalasiewicz, J. A., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 185219.Google Scholar
Wilkinson, T. J. (2003). Archaeological Landscapes of the Near East. Tucson: University of Arizona Press.Google Scholar
Williams, A. T., and Simmons, S. L. (1996). The degradation of plastic litter in rivers: Implications for beaches. Journal of Coastal Conservation, 2, 6372.Google Scholar
Williams, M., Edgeworth, M., Zalasiewicz, J., et al. (2019, in press). Underground metro systems: A durable geological proxy of rapid urban population growth and energy consumption during the Anthropocene. Anthropocene. In Benjamin, C., Quaedakers, E., and Baker, D., eds., The Routledge Handbook of Big History (Routledge Companions).Google Scholar
Williams, M., Zalasiewicz, J., Davies, N., et al. (2015a). Humans as the third evolutionary stage of biosphere engineering of rivers. Anthropocene, 7, 5763.Google Scholar
Williams, M., Zalasiewicz, J., Haff, P. K., et al. (2015b). The Anthropocene biosphere. Anthropocene Review, 2, 196219.Google Scholar
Williams, M., Zalasiewicz, J., Haywood, A., and Ellis, M. (eds.) (2011). The Anthropocene: A new epoch of geological time? Philosophical Transactions of the Royal Society A, 369, 8331112.Google Scholar
Williams, M., Zalasiewicz, J., and Waters, C. N. (2017). The Anthropocene: A geological perspective. In Heikkurinen, P., ed., Sustainability and Peaceful Coexistence for the Anthropocene. Routledge Series on Transnational Law and Governance. Oxon.: Taylor & Francis, pp. 1630.Google Scholar
Williams, M., Zalasiewicz, J., Waters, C. N., and Landing, E. (2014). Is the fossil record of complex animal behaviour a stratigraphical analogue for the Anthropocene? In Waters, C. N., Zalasiewicz, J., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 143148.Google Scholar
Williams, M., Zalasiewicz, J., Waters, C. N., et al. (2016). The Anthropocene: A conspicuous stratigraphical signal of anthropogenic changes in production and consumption across the biosphere. Earth’s Future, 4, 3453.Google Scholar
Willmore, P. L. (1949). Seismic experiments on the North German explosions, 1946 to 1947. Philosophical Transactions of the Royal Society A, 242(843), 123151.Google Scholar
Wilmshurst, J., Hunt, T. L., Lipo, C. P., and Anderson, A. J. (2011). High-precision radiocarbon dating shows recent and rapid initial colonization of East Polynesia. Proceedings of the National Academy of Sciences (USA), 108, 18151820.Google Scholar
Wilson, M. J., and Bell, N. (1996). Acid deposition and heavy metal mobilization. Applied Geochemistry, 11(1), 133137.Google Scholar
Wing, S. L., and Currano, E. D. (2013). Plant response to a global greenhouse 56 million years ago. American Journal of Botany, 100, 12341254.Google Scholar
Wolfe, A. P., Hobbs, W. O., Birks, H. H., et al. (2013). Stratigraphic expressions of the Holocene–Anthropocene transition revealed in sediments from remote lakes. Earth-Science Reviews, 116, 1734.Google Scholar
Wolff, E. W. (2011). Greenhouse gases in the Earth System: A palaeoclimate perspective. Philosophical Transactions of the Royal Society A, 369, 21332147.Google Scholar
Wolff, E. W. (2013). Ice sheets and nitrogen. Philosophical Transactions of the Royal Society B, 368, 20130127. http://dx.doi.org/10.1098/rstb.2013.0127.Google Scholar
Wolff, E. W. (2014). Ice sheets and the Anthropocene. In Waters, C. N., Zalasiewicz, J. A., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society, London, Special Publications, 395, pp. 255263.Google Scholar
Wolff, E. W., Fischer, H., and Rõthlisberger, R. (2009). Glacial terminations as southern warmings without northern control. Nature Geoscience, 2, 206209.Google Scholar
Wolff, E. W., and Suttie, E. D. (1994). Antarctic snow record of southern hemisphere lead pollution. Geophysical Research Letters, 21, 781784.Google Scholar
Wolff, E. W., Suttie, E. D., and Peel, D. A. (1999). Antarctic snow record of cadmium, copper, and zinc content during the twentieth century. Atmospheric Environment, 33(10), 15351541.Google Scholar
Wood, K. (2008). Microspheres: Fillers filled with possibilities. Composites World, 14(2), 2832.Google Scholar
Wood, K. R., and Overland, J. E. (2010). Early 20th century Arctic warming in retrospect. International Journal of Climatology, 30, 12691279.Google Scholar
Wood, R. (1999). Reef Evolution. Oxford University Press.Google Scholar
Woodall, L. C., Sanchez-Vidal, A., Canals, M., et al. (2014). The deep sea is a major sink for microplastic debris. Royal Society Open Science, 1, 140317.Google Scholar
Woodroffe, S. A., and Horton, B. P. (2005). Holocene sea-level changes in the Indo-Pacific. Journal of Asian Earth Sciences, 25, 2943.Google Scholar
Woods, W. I. (2008). Amazonian Dark Earths: Wim Sombroek’s Vision. New York: Springer.Google Scholar
Woodworth, P. L., Gehrels, W. R., and Nerem, R. S. (2011). Nineteenth and twentieth century changes in sea level. Oceanography, 24, 8093.Google Scholar
Wright, G. R. H. (1985). Ancient Building in South Syria and Palestine, vol. 1. Leiden-Köln: E. J. Brill.Google Scholar
Wright, J. D., and Schaller, M. F. (2013). Evidence for a rapid release of carbon at the Paleocene-Eocene thermal maximum. Proceedings of the National Academy of Sciences (USA), 110(40), 1590815913.Google Scholar
Wright, S. L., Thompson, R. C., and Galloway, T. S. (2013). The physical impacts of microplastics on marine organisms: A review. Environmental Pollution, 178, 483492.Google Scholar
Wrigley, E. A. (2010). Energy and the English Industrial Revolution. Cambridge University Press.Google Scholar
Wroe, S., and Field, J. (2006). A review of the evidence for a human role in the extinction of Australian megafauna and an alternative interpretation. Quaternary Science Reviews, 25, 26922703.Google Scholar
Wroe, S., Field, J. H., Archer, M., et al. (2013). Climate change frames debate over the extinction of megafauna in Sahul (Pleistocene Australia-New Guinea). Proceedings of the National Academy of Sciences (USA), 110, 87778781.Google Scholar
Wynn, P. M., Borsato, A., Baker, A., et al. (2013). Biogeochemical cycling of sulphur in karst and transfer into speleothem archives at Grotta di Ernesto, Italy. Biogeochemistry, 114, 255267.Google Scholar
Wynn, P. M., Fairchild, I. J., Baker, A., et al. (2008). Isotopic archives of sulphate in speleothems. Geochimica Cosmochimica Acta, 72, 24652477.Google Scholar
Wynn, P. M., Fairchild, I. J., Frisia, C., et al. (2010). High-resolution sulphur isotope analysis of speleothem carbonate by secondary ionisation mass spectrometry. Chemical Geology, 271, 101107.Google Scholar
Wynn, P. M., Loader, N. J., and Fairchild, I. J. (2014). Interrogating trees for isotopic archives of atmospheric sulphur deposition and comparison to speleothem records. Environmental Pollution, 187, 98105.Google Scholar
Wypych, G. (2016). Fillers – origin, chemical composition, properties, and morphology. In Wypych, G., ed., Handbook of Fillers. Amsterdam: Elsevier, pp. 13266.Google Scholar
Wu, X.-H., Zhang, C., Goldberg, P., Cohen, D., et al. (2013). Early pottery at 20,000 years ago in Xianrendong Cave, China. Science, 336, 16961700.Google Scholar
Xu, B. Q., Cao, J. J., Hansen, J., Yao, T. D., et al. (2009). Black soot and the survival of Tibetan glaciers. Proceedings of the National Academy of Sciences (USA), 106, 2211422118.Google Scholar
Xu, J. (2003). Naturally and anthropogenically accelerated sedimentation in the Lower Yellow River, China, over the past 13.000 years. Geografiska Annaler, Series A, 80(1), 6778.Google Scholar
Yaalon, D. H., and Yaron, B. (1966). Framework for man-made soil changes – an outline of metapedogenesis. Soil Science, 102, 272277.Google Scholar
Yan, H., Sun, L., Wang, Y., et al. (2010). A 2000-year record of copper pollution in South China Sea derived from seabird excrements: A potential indicator for copper production and civilization of China. Journal of Paleolimnology, 44(2), 431442.Google Scholar
Yan, X.-H., Boyer, T., Trenberth, K., Karl, T. R., et al. (2016). The global warming hiatus: Slowdown or redistribution? Earth’s Future, 4. doi:10.1002/2016EF000417.Google Scholar
Yang, C., Rose, N. L., Turner, S. D., et al. (2016). Hexabromocyclododecanes, polybrominated diphenyl ethers, and polychlorinated biphenyls in radiometrically dated sediment cores from English lakes, ~ 1950–present. Science of the Total Environment, 541, 721728.Google Scholar
Yang, H., Engstrom, D. R., and Rose, N. L. (2010). Recent changes in atmospheric mercury deposition recorded in the sediments of remote equatorial lakes in the Rwenzori Mountains, Uganda. Environmental Science & Technology, 44, 65706575.Google Scholar
Yang, H., Lohmann, G., Wei, W., et al. (2016). Intensification and poleward shift of subtropical western boundary currents in a warming climate. Journal of Geophysical Research Oceans, 121(7), 49284945.Google Scholar
Yang, H., Rose, N. L., and Battarbee, R. W. (2001). Dating of recent catchment peats using spheroidal carbonaceous particle (SCP) concentration profiles with particular reference to Lochnagar, Scotland. Holocene, 11, 593597.Google Scholar
Yang, H., Rose, N. L., Battarbee, R. W., and Boyle, J. F. (2002). Mercury and lead budgets for Lochnagar, a Scottish mountain lake and its catchment. Environmental Science & Technology, 36, 13831388.Google Scholar
Yang, J., Yang, Y., Wu, W. M., Zhao, J., and Jiang, L. (2014). Evidence of polyethylene biodegradation by bacterial strains from the guts of plastic-eating waxworms. Environmental Science and Technology, 48, 1377613784.Google Scholar
Yang, Y., Dong, G., Zhang, S., et al. (2017). Copper content in anthropogenic sediments as a tracer for detecting smelting activities and its impact on environment during prehistoric period in Hexi Corridor, Northwest China. Holocene, 27(2), 282291.Google Scholar
Yang, Z., Wang, H., Saito, Y., et al. (2006). Dam impacts on the Changjiang (Yangtze River) sediment discharge to the sea: The past 55 years and after the Three Gorges Dam. Water Resources Research, 42, W04407. doi:10.1029/2005WR003970.Google Scholar
Yasuhara, M., Hunt, G., Breitburg, D., et al. (2012). Human-induced marine ecological degradation: Micropaleontological perspectives. Ecology and Evolution, 2, 32423268.Google Scholar
Yin, Z., Maoyan, Z., Davidson, E. H., et al. (2015). Sponge grade body fossil with cellular resolution dating 60 Myr before the Cambrian. Proceedings of the National Academy of Sciences (USA), 112, E1453–1460.Google Scholar
Yoshida, S., Hiraga, K., Takehana, T., et al. (2016). A bacterium that degrades and assimilates poly(ethylene terephthalate). Science, 351(6278), 11961199.Google Scholar
Yu, Z., Loisel, J., Brosseau, D. P., and Beilman, D. W. (2010). Global peatland dynamics since the Last Glacial Maximum. Geophysical Research Letters, 37, L13402. doi:10.1029/2010GL043584.Google Scholar
Zachos, J. C., Bohaty, S. M., John, C. M., et al. (2007). The Palaeocene-Eocene carbon isotope excursion: Constraints from individual shell planktonic foraminifer records. Philosophical Transactions of the Royal Society A, 365, 18291842.Google Scholar
Zachos, J. C., Dickens, G. R., and Zeebe, R. E. (2008). An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature, 451, 279283.Google Scholar
Zachos, J. C., Pagani, M., Sloan, M., et al. (2001). Trends, rhythms and aberrations in global climate 65 Ma to present. Science, 292, 686693.Google Scholar
Zachos, J. C., Röhl, U., Schellenberg, S. A., et al. (2005). Rapid acidification of the ocean during the Paleocene–Eocene thermal maximum. Science, 308, 1611–161.Google Scholar
Zachos, J. C., Wara, M. W., Bohaty, S., et al. (2003). A transient rise in tropical sea surface temperature during the Paleocene-Eocene thermal maximum. Science, 302(5650), 15511554.Google Scholar
Zalasiewicz, J., Cita, M. B., Hilgen, F., et al. (2013). Chronostratigraphy and geochronology: A proposed realignment. GSA Today, 23(3), 48.Google Scholar
Zalasiewicz, J., Kryza, R., and Williams, M. (2014a). The mineral signature of the Anthropocene. In Waters, C. N., Zalasiewicz, J. A., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society of London Special Publication, 395, pp. 109117.Google Scholar
Zalasiewicz, J., Steffen, W., Leinfelder, R., et al. (2017a). Petrifying Earth process: The stratigraphic imprint of key Earth System parameters in the Anthropocene. In Clark, N., and Yusoff, K., eds., Theory Culture & Society, Special Issue: Geosocial Formations and the Anthropocene; Theory, Culture & Society, 34, pp. 83104.Google Scholar
Zalasiewicz, J., Waters, C. N., Barnosky, A. D., et al. (2015a). Colonization of the Americas, “Little Ice Age” climate, and bomb-produced carbon: Their role in defining the Anthropocene. Anthropocene Review, 2(2), 117127.Google Scholar
Zalasiewicz, J., Waters, C. N., and Head, M. J. (2017b). Anthropocene: Its stratigraphic basis. Nature, 541(7637), 289.Google Scholar
Zalasiewicz, J., Waters, C. N., Head, M. J., et al. (2018). The geological and Earth System reality of the Anthropocene: Reply to Bauer, A. M. and Ellis, E. C. The Anthropocene Divide: Obscuring understanding of social-environmental change. Current Anthropology, 59(2), 220223.Google Scholar
Zalasiewicz, J., Waters, C. N., Ivar do Sul, J., et al. (2016a). The geological cycle of plastics and their use as a stratigraphic indicator of the Anthropocene. Anthropocene, 13, 417.Google Scholar
Zalasiewicz, J., Waters, C. N., Summerhayes, C. P., et al. (2017c). The Working Group on the Anthropocene: Summary of evidence and interim recommendations. Anthropocene, 19, 5560.Google Scholar
Zalasiewicz, J., Waters, C. N., and Williams, M. (2014b). Human bioturbation, and the subterranean landscape of the Anthropocene. Anthropocene, 6, 39.Google Scholar
Zalasiewicz, J., Waters, C. N., Williams, M., et al. (2015b). When did the Anthropocene begin? A mid-twentieth century boundary level is stratigraphically optimal. Quaternary International, 383, 196203.Google Scholar
Zalasiewicz, J., Waters, C. N., Wolfe, A. P., et al. (2017d). Making the case for a formal Anthropocene: An analysis of ongoing critiques. Newsletters on Stratigraphy, 50, 205226.Google Scholar
Zalasiewicz, J., and Williams, M. (2012). The Goldilocks Planet: An Earth History of Climate Change. Oxford University Press.Google Scholar
Zalasiewicz, J., and Williams, M. (2014). The Anthropocene: A comparison with the Ordovician-Silurian boundary. Rendiconti Lincei – Scienze Fisiche e Naturali, 25(1), 512.Google Scholar
Zalasiewicz, J., Williams, M., Fortey, R., et al. (2011a). Stratigraphy of the Anthropocene. In Zalasiewicz, J. A., Williams, M., Haywood, A., and Ellis, M., eds., The Anthropocene: A New Epoch of Geological Time. Philosophical Transactions of the Royal Society A, 369, pp. 10361055.Google Scholar
Zalasiewicz, J., Williams, M., Haywood, A., and Ellis, M. (2011b). The Anthropocene: A new epoch of geological time? Philosophical Transactions of the Royal Society, A369, 835841.Google Scholar
Zalasiewicz, J., Williams, M., Smith, A., et al. (2008). Are we now living in the Anthropocene? GSA Today, 18(2), 48.Google Scholar
Zalasiewicz, J., Williams, M., and Waters, C. N. (2014c). Can an Anthropocene Series be defined and recognized? In Waters, C. N., Zalasiewicz, J. A., Williams, M., et al., eds., A Stratigraphical Basis for the Anthropocene. Geological Society of London Special Publication, 395, pp. 3954.Google Scholar
Zalasiewicz, J., Williams, M., Waters, C. N., et al. (2014d). The technofossil record of humans. Anthropocene Review, 1, 3443.Google Scholar
Zalasiewicz, J., Williams, M., Waters, C. N., et al. (2016b). Scale and diversity of the physical technosphere: A geological perspective. Anthropocene Review. 4(1), 922.Google Scholar
Zalasiewicz, J., and Zalasiewicz, M. (2015). Battle scars. New Scientist. 36–39.Google Scholar
Zannoni, D., Valotto, G., Visin, F., and Rampazzo, G. (2016). Sources and distribution of tracer elements in road dust: The Venice mainland case of study. Journal of Geochemical Exploration, 166, 6472.Google Scholar
Zareitalabad, P., Siemens, J., Hamer, M., and Amelung, W. (2013). Perfluorooctanoic acid (PFOA) and perfluorooctanesulfonic acid (PFOS) in surface waters, sediments, soils and wastewater: A review on concentrations and distribution coefficients. Chemosphere, 91, 725732.Google Scholar
Zarfl, C., and Matthies, M. (2010). Are marine plastic particles transport vectors for organic pollutants to the Arctic? Marine Pollution Bulletin, 60(10), 18101814.Google Scholar
Zayasu, T. Y., and Shinzato, C. (2016). Hope for coral reef rehabilitation: Massive synchronous spawning by outplanted corals in Okinawa, Japan. Coral Reefs, 35, 12951295.Google Scholar
Zbyszewski, M., Corcoran, P. L., and Hockin, A. (2014). Comparison of the distribution and degradation of plastic debris along the shorelines of the Great Lakes, North America. Journal of Great Lakes Research, 40, 288299.Google Scholar
Zeder, M. A. (2011). The origins of agriculture in the Near East. Current Anthropology, 52, 221235.Google Scholar
Zeebe, R. E., Dickens, G. R., Ridgwell, A., et al. (2014). Onset of carbon isotope excursion at the Paleocene-Eocene thermal maximum took millennia, not 13 years. Proceedings of the National Academy of Sciences, 111(12), E1062E1063.Google Scholar
Zeebe, R. E., Ridgwell, A., and Zachos, J. C. (2016). Anthropogenic carbon release rate unprecedented during the past 66 million years. Nature Geoscience, 9, 325329.Google Scholar
Zeebe, R. E., Zachos, J. C., and Dickens, G. R. (2009). Carbon dioxide forcing alone insufficient to explain Palaeocene–Eocene Thermal Maximum warming. Nature Geoscience, 2(8), 576580.Google Scholar
Zemp, D. C., Schleussner, C. F., Barbosa, H. M. J., and Rammig, A. (2017). Deforestation effects on Amazon forest resilience. Geophysical Research Letters, 44 (12), 61826190.Google Scholar
Zemp, M., Frey, H., Gärtner-Roer, I., et al. (2015). Historically unprecedented global glacier decline in the early 21st century. Journal of Glaciology, 61(228), 745762.Google Scholar
Zettler, E. R., Mincer, T. J., and Amaral-Zettler, L. A. (2013). Life in the “Plastisphere”: Microbial communities on plastic marine debris. Environmental Science and Technology, 47, 71377146.Google Scholar
Zhao, S., Zhu, L., Wang, T., and Li, D. (2014). Suspended microplastics in the surface water of the Yangtze estuary system, China: First observations on occurrence, distribution. Marine Pollution Bulletin, 86, 562568.Google Scholar
Zheng, J., Shotyk, W., Krachler, M., and Fisher, D. A. (2007). A 15,800-year record of atmospheric lead deposition on the Devon Island ice cap, Nunavut, Canada: Natural and anthropogenic enrichments, isotopic composition and predominant sources. Global Biogeochemical Cycles, 21, GB2027.Google Scholar
Zhu, R. X., Potts, R., Xie, F., et al. (2004). New evidence on the earliest human presence at high northern latitudes in northeast Asia. Nature, 431, 559562.Google Scholar
Zhu, Z., Piao, S., Myneni, R. B., et al. (2016). Greening of the Earth and its drivers. Nature Climate Change, 6, 791795.Google Scholar
Zielinski, G. A., Mayewski, P. A., Meeker, L. D., et al. (1994). Record of volcanism since 7000 B.C. from the GISP2 Greenland ice core and implications for the volcano-climate system. Science, 264, 948952.Google Scholar
Zinke, J., Loveday, B. R., Reason, C. J. C., et al. (2014). Madagascar corals track sea surface temperature variability in the Agulhas Current core region over the past 334 years. Scientific Reports, 4, 4393. doi:10.1038/srep04393.Google Scholar
Zolitschka, B., Francus, P., Ojala, A. E. K., and Schimmelmann, A. (2015). Varves in lake sediments: A review. Quaternary Science Reviews, 117, 141.Google Scholar
Zong, Y., and Horton, B. P. (1999). Diatom-based tidal-level transfer functions as an aid in reconstructing Quaternary history of sea-level movements in the UK. Journal of Quaternary Science, 14, 153167.Google Scholar
Zushi, Y., Tamada, M., Kanai, Y., and Masunaga, S. (2010). Time trends of perfluorinated compounds from the sediment core of Tokyo Bay, Japan (1950s–2004). Environmental Pollution, 158, 756763.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×