Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-27T16:01:37.563Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  20 December 2019

Roger LeB. Hooke
Affiliation:
University of Maine, Orono
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 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

Aber, J. S., and Ber, A. (2007). Glaciotectonism. Developments in Quaternary Science, 6, Amsterdam: Elsevier, 246pp.Google Scholar
Adhikari, S., Ivins, E. R., and Larour, E. (2017). Mass transport waves amplified by intense Greenland melt and detected in solid earth deformation. Geophysical Research Letters, 44, 49654975.CrossRefGoogle Scholar
Adhikari, S., Nakawo, M., Seko, K., and Shakya, B. (2000). Dust influence on the melting process of glacier ice: experimental results from Lirung Glacier, Nepal Himalayas. In Nakawo, M., Raymond, C. F., and Fountain, A. (eds.) Debris-Covered Glaciers. IAHS Publication no. 264, 43–52.Google Scholar
Ahn, J., Headly, M., Wahlen, M., Brook, E. J., Mayewski, P. A., and Taylor, K. C. (2008). CO2 diffusion in polar ice: observations from naturally formed CO2 spikes in the Siple Dome (Antarctica) ice core. Journal of Glaciology, 54(187), 685695.Google Scholar
Allen, C. R., Kamb, W. B., Meier, M. F., and Sharp, R. P. (1960). Structure of lower Blue Glacier, Washington. Journal of Geology, 68(6), 601625.Google Scholar
Alley, R. B. (1988). Fabrics in polar ice sheets: development and prediction. Science, 240, 493495.Google Scholar
Alley, R. B. (1989a). Water pressure coupling of sliding and bed deformation: I. Water system. Journal of Glaciology, 35(119), 108118.Google Scholar
Alley, R. B. (1989b). Water-pressure coupling of sliding and bed deformation: II. Velocity-depth profiles. Journal of Glaciology, 35(119), 119129.CrossRefGoogle Scholar
Alley, R. B. (1991). Deforming bed origin for the southern Laurentide till sheets? Journal of Glaciology, 37(125), 6776.Google Scholar
Alley, R. B. (1992). Flow-law hypotheses for ice-sheet modeling. Journal of Glaciology, 38(129), 245256.CrossRefGoogle Scholar
Alley, R. B. (1993). In search of ice-stream sticky spots. Journal of Glaciology, 39(133), 447454.Google Scholar
Alley, R. B., and Whillans, I. M. (1991). Changes in the West Antarctic ice sheet. Science, 254(5034), 259263.CrossRefGoogle ScholarPubMed
Alley, R. B., et al. (1993). Abrupt increase in Greenland snow accumulation at the end of the Younger Dryas event. Nature, 362(6420), 527529.Google Scholar
Alley, R. B., Gow, A. J., and Meese, D. A. (1995). Mapping c-axis fabrics to study physical processes in ice. Journal of Glaciology, 41(137), 197203.CrossRefGoogle Scholar
Alley, R. B., Gow, A. J., Meese, D. A., Fitzpatrick, J. J., Waddington, E. D., and Bolzan, J. F. (1997). Grain-scale processes, folding, and stratigraphic disturbance in the GISP2 ice core. Journal of Geophysical Research, 102(C12), 2681926830.Google Scholar
Amundsen, R. (1912). The South Pole: an account of the Norwegian Antarctic expedition in the “Fram”, 1910–1912, Volumes 1 and 2, translated by A. G. Chater.Google Scholar
Anandakrishnan, S., Catania, G. A., Alley, R. B., and Horgan, H. J. (2007). Discovery of till deposition at the grounding line of Whillans ice stream. Science, 315, 18351838.Google Scholar
Andrews, L. C. et al. (2014). Direct observations of evolving subglacial drainage beneath the Greenland Ice Sheet. Nature, 514, 8083.Google Scholar
Arendt, A., Luthcke, S., Gardner, A., O’Neel, S., Hill, D., Moholdt, G., and Abdalati, W. (2013). Analysis of a GRACE global mascon solution for Gulf of Alaska glaciers. Journal of Glaciology, 59(217), 913924, doi: 10.3189/2013JoG12J197.Google Scholar
Ashley, G. M., Boothroyd, J. C., and Borns, H. W. Jr. (1991). Sedimentology of late Pleistocene (Laurentide) deglacial-phase deposits, eastern Maine; an example of a temperate marine grounded ice-sheet margin. Geological Society of America Special Paper, 261, 107125.Google Scholar
Atkinson, B. K. (1984). Subcritical crack growth in geological materials. Journal of Geophysical Research, 89(B6), 40774144.CrossRefGoogle Scholar
Atkinson, B. K., and Rawlings, R. D. (1981). Acoustic emission during stress corrosion cracking in rocks. In Simpson, D. W., and Richards, P. G. (eds.) Earthquake Prediction. An International Review. Ewing Series, 4. Washington, DC: American Geophysical Union, pp. 605619.Google Scholar
Attig, J. W., Mickelson, D. M., and Clayton, L. (1989). Late Wisconsin landform distribution and glacier bed conditions in Wisconsin. Sedimentary Geology, 62(3–4), 399405.Google Scholar
Azuma, N., Miyakoshi, T., Yokoyama, S., and Takata, M. (2012). Impeding effect of air bubbles on normal grain growth of ice. Journal of Structural Geology, 42, 184193.Google Scholar
Bahr, D. B., Dyurgerov, M., and Meier, M. F. (2009). Sea-level rise from glaciers and ice caps: a lower bound. Geophysical Research Letters, 36, L03501, doi:10.1029/2008GL036309.Google Scholar
Bahr, D. B., Meier, M. F., and Peckham, S. D. (1997). The physical basis of glacier volume-area scaling. Journal of Geophysical Research, 102(B9), 2035520362.CrossRefGoogle Scholar
Bahr, D. B., Pfeffer, W. T., and Kaser, G. (2015). A review of volume-area scaling of glaciers. Reviews of Geophysics, 53, 95140.Google Scholar
Baker, R. W. (1981). Textural and crystal-fabric anisotropies and the flow of ice masses. Science, 211(4486), 10431044.Google Scholar
Baker, R. W. (1982). A flow equation for anisotropic ice. Cold Regions Science and Technology, 6(3), 141148.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(5929), 901903.CrossRefGoogle ScholarPubMed
Banerjee, I., and McDonald, B. C. (1975). Nature of esker sedimentation. The Society of Economic Paleontologists and Mineralogists Special Publication 23, 132154.Google Scholar
Banwell, A. F., MacAyeal, D. R., and Sergienko, O. V. (2013). Breakup of Larsen B ice shelf triggered by chain reaction drainage of supraglacial lakes. Journal of Geophysical Research, 40, 15, doi: 10.1002/2013GL057694.Google Scholar
Barnes, P., Tabor, D., and Walker, J. C. F. (1971). Friction and creep of polycrystalline ice. Proceedings of the Royal Society of London, Series A, 324(1557), 127155.Google Scholar
Begeman, C. B., et al. (2018). Ocean stratification and low melt rates at the Ross ice shelf grounding zone. Journal of Geophysical Research, Oceans, 123(10), 74387452, doi: 10.1029/2018JC013987.CrossRefGoogle Scholar
Bell, J. F. (1941). Morphology of mechanical twinning in crystals. American Mineralogist, 26, 247261.Google Scholar
Beltaos, S. (2002). Collapse of floating ice covers under vertical loads: test data vs. theory. Cold Regions Science and Technology, 34, 191207.Google Scholar
Bender, M. L. (2002). Orbital tuning chronology for the Vostok climate record supported by trapped gas composition. Earth and Planetary Science Letters, 204, 275289, doi: 10.1016/S0012-821X(02)00980-9.CrossRefGoogle Scholar
Bender, M. L., Barnett, B., Dreyfus, G., Jouzel, J., and Porcelli, D. (2008). The contemporary degassing rate of 40Ar from the solid Earth. Proceedings of the National Academy of Science, USA, 105(24), 82328237.CrossRefGoogle ScholarPubMed
Benn, D. I. (1994). Fluted moraine formation and till genesis below a temperate valley glacier, Slettmarkbreen, Jotunheimen, southern Norway. Sedimentology, 41, 279292.CrossRefGoogle Scholar
Benn, D. I., Hulton, N. R. J., and Mottram, R. H. (2007). ‘Calving laws’, ‘sliding laws’ and the stability of tidewater glaciers. Annals of Glaciology, 46, 123130.Google Scholar
Benoist, J.-P. (1979). The spectral power density and shadowing function of a glacial microrelief at the decimetre scale. Journal of Glaciology, 23(89), 5766.Google Scholar
Benson, C. S. (1961). Stratigraphic studies in the snow and firn of the Greenland Ice Sheet. Folia Geographica Danica, 9, 1337.Google Scholar
Benson, C. S. (1962). Stratigraphic studies in the snow and firn of the Greenland Ice Sheet. U.S. Snow, Ice, and Permafrost Research Establishment. Research Report 70.Google Scholar
Berner, W., Oeschger, H., and Stauffer, B. (1980). Information on the CO2 cycle from ice core studies. Radiocarbon, 22(2), 227235.CrossRefGoogle Scholar
Bhattacharji, S. (1967). Mechanics of flow differentiation in ultramafic and mafic sills. The Journal of Geology, 75(1), 101112.Google Scholar
Biegel, R. L., Sammis, C. G., and Dieterich, J. H. (1989). The frictional properties of simulated gouge having a fractal particle distribution. Journal of Structural Geology, 11(7), 827846.Google Scholar
Bindschadler, R. A., and Scambos, T. A. (1991). Satellite-image-derived velocity field of an Antarctic ice stream. Science, 252(5003), 242246.Google Scholar
Bindschadler, R., and Vornberger, P. (1998). Changes in the West Antarctic Ice Sheet since 1963 from declassified satellite photography. Science, 279, 689692.Google Scholar
Bindschadler, R. A., King, M. A., Alley, R. B., Anandakrishnan, S., and Padman, L. (2003a). Tidally controlled stick-slip discharge of a West Antarctic ice stream. Science, 301, 10871089.Google Scholar
Bindschadler, R. A., Vornberger, P. L., King, M. A., and Padman, L. (2003b). Tidally-driven stick-slip motion in the mouth of Whillans Ice Stream, Antarctica. Annals of Glaciology, 36, 263272.CrossRefGoogle Scholar
Bishop, B. C. (1957). Shear moraines in the Thule area, northwest Greenland. U.S. Army Snow, Ice, and Permafrost Research Establishment, Research Report 17, 47 p.Google Scholar
Björnsson, H. (1992). Jökulhlaups in Iceland: prediction, characteristics, and simulation. Annals of Glaciology, 16, 95106.Google Scholar
Blankenship, D. D., Bentley, C. R., Rooney, S. T., and Alley, R. B. (1986). Seismic measurements reveal a saturated, porous layer beneath an active Antarctic ice stream. Nature, 322(6074), 5457.Google Scholar
Blackburn, T., Siman-Tov, S., Coble, M. A., Stock, G. M., Brodsky, E. E., and Hallet, B. (2019). Composition and formation age of the amorphous silica coating glacially polished surfaces. Geology, 47(4), 347350.Google Scholar
Bluemle, J. P., Lord, M. L., and Hunke, N. T. (1993). Exceptionally long, narrow drumlins formed in subglacial cavities, North Dakota. Boreas, 22, 1524.Google Scholar
Böðvarsson, G. (1955). On the flow of ice sheets and glaciers. Jökull, 5, 18.Google Scholar
Bolduc, A. M. (1992). The formation of eskers based on their morphology, stratigraphy and lithologic composition, Labrador, Canada. PhD thesis, Lehigh University.Google Scholar
Bond, G., et al. (1997). A pervasive millennial-scale cycle in north Atlantic holocene and glacial climates. Science, 278(5341), 12571266.Google Scholar
Boon, S., and Sharp, M. (2003). The role of hydrologically-driven ice fracture in drainage system evolution on an Arctic glacier. Geophysical Research Letters, 30(18), doi: 10.1029/2003GL018034Google Scholar
Borstad, C., Khazendar, A., Scheuchl, B., Morlighem, M., Larour, E., and Rignot, E. (2016). A constitutive framework for predicting weakening and reduced buttressing of ice shelves based on observations of progressive deterioration of the remnant Larsen B Ice Shelf. Geophysical Research Letters, 43, 20272035.Google Scholar
Boulton, G. S. (1987). A theory of drumlin formation by subglacial sediment deformation. In Menzies, J., and Rose, J. (eds.) Drumlin Symposium. Rotterdam: Balkema Publishers, pp. 2580.Google Scholar
Boulton, G. S., and Hindmarsh, R. C. A. (1987). Sediment deformation beneath glaciers: rheology and geological consequences. Journal of Geophysical Research, 92(B9), 90599082.Google Scholar
Bourne, A. J., et al. (2015). A tephra lattice for Greenland and a reconstruction of volcanic events spanning 25–45 Ka b2k. Quaternary Science Reviews, 118, 122141.CrossRefGoogle Scholar
Brand, G., Pohjola, V., and Hooke, R. LeB. (1987). Existence of a layer of deformable till at the base of Storglaciären, Sweden, revealed by electrical resistivity measurements. Journal of Glaciology, 33(115), 311314.Google Scholar
Broecker, W. S. (1998). Paleocean circulation during the last glaciation: a bipolar seesaw. Paleoceanography, 13(2), 119121.Google Scholar
Brown, C. S., Meier, M. F., and Post, A. (1982). Calving speed of Alaska tidewater glaciers, with application to Columbia Glacier. U.S. Geological Survey Professional Paper, 1258-C, C1C13.Google Scholar
Brown, N. L., Hallet, B., and Booth, D. B. (1987). Rapid soft-bed sliding of the Puget glacial lobe. Journal of Geophysical Research, 92(B9), 89858997.Google Scholar
Brugger, K. A. (2007). The non-synchronous response of Rabots Glaciär and Storglaciären, northern Sweden, to recent climatic change: a comparative study. Annals of Glaciology, 46, 275282.CrossRefGoogle Scholar
Brunt, K. M., and MacAyeal, D. R. (2014). Tidal modulation of ice=shelf flow: a viscous model of the Ross Ice Shelf. Journal of Glaciology, 60(221), 500508.CrossRefGoogle Scholar
Buckingham, E. (1914). On physically similar systems; illustrations of the use of dimensional equations. Physical Review, 4(4), 345376.Google Scholar
Budd, W. F. (1968). The longitudinal velocity profile of large ice masses. Union de Geodesie et Geophysique International. Association Internatiollale d ‘Hydrologie Scienti/ique. Assemblee generale de Bern, 25 Sept.–7 Oct. 1967. [Commission de Neiges et Glaces.] Rapports et discussions, pp. 58–77. (Publication No. 79 de I’ Association Internationale d’Hydrologie Scientifique.)Google Scholar
Budd, W. F. (1969). The dynamics of ice masses. Australian National Antarctic Expeditions Scientific Reports, Series A (IV) Glaciology. Publication No. 108.Google Scholar
Budd, W. F. (1970). The longitudinal stress and strain-rate gradients in ice masses. Journal of Glaciology, 9(55), 1927.Google Scholar
Budd, W. F., and Jacka, T. H. (1989). A review of ice rheology for ice sheet modelling. Cold Regions Science and Technology, 16(2), 107144.Google Scholar
Budd, W. F., Jensen, D., and Radok, U. (1971). Derived physical characteristics of the Antarctic Ice Sheet. Australian National Antarctic Expeditions Interim Reports, Series A (IV) Glaciology. Publication No. 120.Google Scholar
Butkovitch, T. R. (1954). The ultimate strength of ice. Snow, Ice, and Permafrost Research Establishment Research Report, 11, 12.Google Scholar
Canals, M., Urgeles, R., and Calafat, A. M. (2000). Deep sea-floor evidence of past ice streams off the Antarctic Peninsula. Geology, 28(1), 3134.2.0.CO;2>CrossRefGoogle Scholar
Carnahan, B., Luther, H. A., and Wilkes, J. O. (1969). Applied Numerical Methods. New York: John Wiley and Sons, Inc.Google Scholar
Carslaw, H. S., and Jaeger, J. C. (1959). Conduction of Heat in Solids. Oxford: Clarendon Press, 510 p.Google Scholar
Catania, G. A., Conway, H., Raymond, C. F., and Scambos, T. A. (2006). Evidence for floatation or near floatation in the mouth of Kamb Ice Stream, West Antarctica, prior to stagnation. Journal of Geophysical Research, 111, F01005, doi: 10.1029/2005JF000355Google Scholar
Catania, G. A., Hulbe, C., and Conway, H. (2010). Grounding-line basal melt rates determined using radar-derived internal stratigraphy. Journal of Glaciology, 56(197), 545554.Google Scholar
Chandler, D., Hubbard, A., Hubbard, B., and Nienow, P. (2006). A Monte Carlo error analysis for basal sliding velocity calculations, Journal of Geophysical Research, 111, F04005, doi: 10.1029/2006JF000476.Google Scholar
Chandler, D. M., et al. (2013). Evolution of the subglacial drainage system beneath the Greenland Ice Sheet revealed by tracers. Nature Geoscience, 6, 195198.Google Scholar
Chen, J., and Ohmura, A. (1990). Estimation of alpine water resources and their change since the 1870s. Hydrology in Mountainous regions: I. Hydrological measurements; the water cycle (Proceedings of two Lausanne Symposia, August 1990). IAHS Publication No. 193, 127–135.Google Scholar
Church, J. A., et al., (2001). Changes in sea level. In Houghton, J. T., et al. (eds.) Climate Change 2001: The Science Basis, Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press, pp. 641693.Google Scholar
Church, J. A., et al. (2013). Sea level change. In Stocker, T. F., 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. Cambridge: Cambridge University Press, pp. 11371216.Google Scholar
Clark, C. D., and Stokes, C. R. (2001). Extent and basal characteristics of the M’Clintock Channel Ice Stream. Quaternary International, 86, 81101.Google Scholar
Clark, P. U., and Hansel, A. R. (1989). Clast ploughing, lodgement and glacier sliding over a soft glacier bed. Boreas, 18(3), 201207.Google Scholar
Clarke, G. K. C. (1982). Glacier outburst floods from “Hazard Lake”, Yukon Territory, and the problem of flood magnitude prediction. Journal of Glaciology, 28(98), 321.Google Scholar
Clarke, G. K. C. (1976). Thermal regulation of glacier surging. Journal of Glaciology, 16(74), 231250.CrossRefGoogle Scholar
Clarke, T. S., and Echelmeyer, K. (1996). Seismic-reflection evidence for a deep subglacial trough beneath Jakobshavn Isbræ, West Greenland. Journal of Glaciology, 43(141), 219232.Google Scholar
Clayton, L., and Cherry, J. A. (1967). Pleistocene superglacial and ice-walled lakes of west-central North America. In Clayton, L., and Freers, T. F. (eds.) Glacial Geology of the Missouri Coteau and adjacent areas. Grand Forks, ND: N. Dakota Geological Survey Miscellaneous Series, 30, pp. 4752.Google Scholar
Clayton, L. (1967). Stagnant-glacier features of the Missouri Coteau in North Dakota. In Clayton, L., and Freers, T. F. (eds.) Glacial Geology of the Missouri Coteau and adjacent areas. Grand Forks, ND: N. Dakota Geological Survey Miscellaneous Series, 30, pp. 2545.Google Scholar
Cohen, D. (2000). Rheology of ice at the bed of Engabreen, Norway. Journal of Glaciology, 46(155), 611621.Google Scholar
Cole-Dai, J., Budner, D. M., and Ferris, D. G. (2006). High-speed, high resolution, and continuous chemical analysis of ice core using a melter and ion chromatography. Environmental Science and Technology, 40, 67646769.CrossRefGoogle ScholarPubMed
Colgan, W., Pfeffer, W. T., Rajaram, H., Abdalati, W., and Balog, J. (2012). Monte Carlo ice flow modeling projects a new stable configuration for Columbia Glacier, Alaska, c. 2020. The Cryosphere, 6, 13951409.Google Scholar
Collins, I. F. (1968). On the use of the equilibrium equations and flow law in relating the surface and bed topography of glaciers and ice sheets. Journal of Glaciology, 7(50), 199204.Google Scholar
Conway, H., Hall, B. L., Denton, G. H., Gades, A. M., and Waddington, E. D. (1999). Past and future grounding-line retreat of the West Antarctic Ice Sheet. Science, 286, 280283.Google Scholar
Cook, E. R., D’Arrigo, R. D., and Briffa, K. R. (1998). A reconstruction of the North Atlantic Oscillation using tree-ring chronologies from North America and Europe. The Holocene, 8, 917.Google Scholar
Costa, J. B. (1983). Paleohydraulic reconstruction of flash-flood peaks from boulder deposits in the Colorado Front Range. Geological Society of America Bulletin, 94, 9861004.Google Scholar
Craig, H. (1961). Isotopic variations in meteoric waters. Science, 133, 17021703.Google Scholar
Cuffey, K. M., and Clow, G. D. (1997). Temperature, accumulation, and ice sheet elevation in central Greenland through the last glacial transition. Journal of Geophysical Research, 102(C12), 2638326396.Google Scholar
Cuffey, K. M., and Paterson, W. S. B. (2010). The Physics of Glaciers. Burlington, MA: Butterworth-Heinemann/Elsevier, 693pp.Google Scholar
Cullather, R. I., Bromwich, D. H., and Van Woert, M. L. (1996). Interannual variations in Antarctic precipitation related to El Niño-Southern Oscillation. Journal of Geophysical Research, 101(D14), 1910919118.Google Scholar
Cutler, P. M., Colgan, P. M., and Mickelson, D. M. (2002). Sedimentologic evidence for outburst floods from the Laurentide Ice Sheet margin in Wisconsin, USA: implications for tunnel-channel formation. Quaternary International, 90, 2340.Google Scholar
Cutler, P. M., MacAyeal, D. R., Mickelson, D. M., Parizek, B. R., and Colgan, P. M. (2000). A numerical investigation of ice-lobe – permafrost interaction around the southern Laurentide ice sheet. Journal of Glaciology, 46(153), 311325.Google Scholar
Dahl-Jensen, D., and Gundestrup, N. S. (1987). Constitutive properties of ice at Dye 3, Greenland. International Association of Hydrological Sciences Publication 170 (Symposium at Vancouver 1987 - The physical basis for ice sheet modelling), 31–43.Google Scholar
Dahl-Jensen, D., and Johnsen, S. J. (1986). Palaeotemperatures still exist in the Greenland ice sheet. Nature, 320, 250252.Google Scholar
Dahl-Jensen, D., Mosegaard, K., Gundestrup, N., Clow, G. D., Johnsen, S. J., Hansen, A. W., and Balling, N. (1998). Past temperatures directly from the Greenland Ice Sheet. Science, 282, 268271.Google Scholar
Damsgaard, A., Egholm, D. L., Beem, L. H., Tulaczyk, S., Larsen, N. K., Piotrowski, J. A., and Siegfried, M. R. (2016). Ice flow dynamics forced by water pressure variation in subglacial granular beds. Geophysical Research Letters, 43, 1216512173.Google Scholar
Dansgaard, W., et al. (1993). Evidence for general instability of past climate from a 250-kyr ice-core record. Nature, 364, 218220.Google Scholar
Das, S. B., Joughlin, I., Behn, M. D., Howat, I. M., King, M. A., Lizarralde, D., and Bhatia, M. P. (2008). Fracture propagation to the base of the Greenland Ice Sheet during supraglacial lake drainage. Science, 320(5877), 778781.Google Scholar
Dash, J. G. (1995). The premelting of ice and its environmental consequences. Reports on Progress in Physics, 58, 115167.Google Scholar
de La Chapelle, S., Duval, P., and Baudet, B. (1995). Compressive creep of polycrystalline ice containing a liquid phase. Scripta Metallurgica et Materialia, 33(3), 447450.Google Scholar
Deeley, R. M., and Parr, P. H. (1914). The Hintereis Glacier. Philosophical Magazine, 6, 153176.Google Scholar
Delmas, R. J., Ascencio, J.-M., and Legrand, M. (1980). Polar ice evidence that atmospheric CO2 20,000 yr BP was 50% of present. Nature, 284(5752), 155157.Google Scholar
Demorest, M. (1941). Glacier flow and its bearing on the classification of glaciers. Geological Society of America Bulletin, 52(12), 20242025.Google Scholar
Demorest, M. (1942). Glacier regimens and ice movement within glaciers. American Journal of Science, 240(1), 3166.Google Scholar
Desai, C. S. (2000). Evaluation of liquefaction using disturbed state and energy approaches. Journal of Geotechnical and Geoenvironmental Engineering, 126(7), 618631.Google Scholar
Ditlevsen, P. D., Andersen, K. K., and Svensson, A. (2007). The DO-events are probably noise induced: statistical investigation of the claimed 1470 years cycle. Climate of the Past, 3, 129134.Google Scholar
Donner, J. J. (1965). The quaternary of Finland. In Rankama, K. (ed.) The Quaternary. New York, Interscience, pp. 199272.Google Scholar
Dowdeswell, J. A., and Fugelli, E. M. G. (2012). The seismic architecture and geometry of grounding-zone wedges formed at the marine margins of past ice sheets. Geological Society of America Bulletin, 124(11/12), 17501761.Google Scholar
Dunlop, P., and Clark, C. D. (2006). The morphological characteristics of ribbed moraine. Quaternary Science Reviews, 25, 16681691.Google Scholar
Dunse, T., Schellenberger, T., Hagen, J. O., Kääb, A., Schuler, T. V., and Reijmer, C. H. (2015). Glacier-surge mechanisms promoted by a hydro-thermodynamic feedback to summer melt. The Cryosphere, 9, 197215.Google Scholar
Duval, P. (1977). The role of water content on the creep rate of polycrystalline ice. In Isotopes and impurities in snow and ice. Proceedings of the Grenoble Symposium, August–September 1975, International Association of Scientific Hydrology, Publication, 118, 29–33.Google Scholar
Duval, P., Ashby, M. F., and Anderman, I. (1983). Rate-controlling processes in the creep of polycrystalline ice. Journal of Physical Chemistry, 87(21), 40664074.Google Scholar
Duval, P., and Castelnau, O. (1995). Dynamic recrystallization of ice in polar ice sheets. Journal de Physique IV, Colloque C3, supplement to Journal de Physique III, 5, C3-197C3-205.Google Scholar
Dykes, R. C., and Brook, M. S. (2010). Terminus recession, proglacial lake expansion and 21st century calving retreat of Tasman Glacier, New Zealand. New Zealand Geographer, 66, 203217.Google Scholar
Echelmeyer, K., and Harrison, W. D. (1999). Ongoing margin migration of Ice Stream B, Antarctica. Journal of Glaciology, 45(150), 361369.Google Scholar
Echelmeyer, K., and Wang, Z. (1987). Direct observation of basal sliding and deformation of basal drift at sub-freezing temperatures. Journal of Glaciology, 33(113), 8398.Google Scholar
Echelmeyer, K. A., Harrison, W. D., Larsen, C., and Mitchell, J. E. (1994). The role of the margins in the dynamics of an active ice stream. Journal of Glaciology, 40(136), 527538.Google Scholar
Ehlers, J., and Stephan, H.-J. (1979). Forms at the base of till strata as indicators of ice movement. Journal of Glaciology, 22(87), 345355.Google Scholar
Elsberg, D. H., Harrison, W. D., Echelmeyer, K. A., and Krimmel, R. M. (2001). Quantifying the effects of climate and surface change on glacier mass balance. Journal of Glaciology, 47(159), 649658.Google Scholar
Engelhardt, H., and Kamb, B. (1997). Basal hydraulic system of a West Antarctic ice stream: constraints from borehole observations. Journal of Glaciology, 43(144), 207230.Google Scholar
Engelhardt, H., and Kamb, B. (1998). Basal sliding of ice stream B, West Antarctica. Journal of Glaciology, 44(147), 223230.Google Scholar
Engelhardt, H. F., Humphrey, N., Kamb, B., and Fahnestock, M. (1990). Physical conditions at the base of a fast moving Antarctic ice stream. Science, 248(4951), 5759.Google Scholar
EPICA community members. (2004). Eight glacial cycles from an Antarctic ice core. Nature, 429, 623628.Google Scholar
Etchecopar, A. (1977). A plane kinematic model of progressive deformation in a polycrystalline aggregate. Tectonophysics, 39, 121139.Google Scholar
Eyles, N., Salden, J. A., and Gilroy, S. (1982). A depositional model for stratigraphic complexes and facies superimposition in lodgement till. Boreas, 11(4), 317333.Google Scholar
Fagan, B. (2008). The Great Warming. New York: Bloomsbury Press. 282pp.Google Scholar
Fairbanks, R. (1989). A 17,000-year glacio-eustatic sea level record: influence of glacial melting rates on the Younger Dryas event and deep ocean circulation. Nature, 232, 637742.Google Scholar
Fastook, J. L., and Chapman, J. E. (1989). A map-plane finite-element model: three modeling experiments. Journal of Glaciology, 35(119), 4852.Google Scholar
Fastook, J. L., and Holmlund, P. (1994). A glaciological model of the Younger Dryas event in Scandinavia. Journal of Glaciology, 40(134), 125131.Google Scholar
Feldmann, J., and Levermann, A. (2015). Collapse of the West Antarctic Ice Sheet after local destabilization of the Admundsen Basin. Proceedings of the National Academy of Sciences, 112(46), 1419114196.Google Scholar
Fischer, H., Siggaard-Andersen, M.-L., Ruth, U., Röthlisberger, R., and Wolff, E. (2007). Glacial/interglacial changes in mineral dust and sea-salt records in polar ice cores: sources, transport, and deposition. Reviews of Geophysics, 45, RG1002, doi: 10.1029/2005RG000192.Google Scholar
Fisher, D. A. (1987). Enhanced flow of Wisconsin ice related to solid conductivity through strain history and recrystallization. International Association of Scientific Hydrology Publication, 170, 4551.Google Scholar
Fisher, D. A., and Koerner, R. M. (1986). On the special rheological properties of ancient microparticle-laden Northern Hemisphere ice as derived from bore-hole and core measurements. Journal of Glaciology, 32(112), 501510.Google Scholar
Fisher, D. A., Reeh, N., and Langley, K. (1985). Objective reconstructions of the late Wisconsinan Laurentide ice sheet and the significance of deformable beds. Gèographie Physique et Quaternaire, 39, 229238.Google Scholar
Fountain, A. G. (1989). The storage of water in, and hydraulic characteristics of, the firn of South Cascade Glacier, Washington State, U.S.A. Annals of Glaciology, 13, 6975.Google Scholar
Fountain, A. G., Jacobel, R. W., Schlichting, R., and Jansson, P. (2005). Fractures as the main pathways of water flow in temperate glaciers. Nature, 433, 618621.Google Scholar
Fountain, A. G., and Walder, J. S. (1998). Water flow through temperate glaciers. Reviews of Geophysics, 36(3) 299328.Google Scholar
Fowler, A. C. (2000). An instability mechanism for drumlin formation. In Maltman, A., Hambrey, M. J., and Hubbard, B. (eds.) Deformation of Glacial Materials. London: Geological Society, 176, pp. 307319.Google Scholar
Fowler, A. C., and Larson, D. A. (1980). On the flow of polythermal glaciers: II. Surface wave analysis. Proceedings of the Royal Society, London, A370, 155171.Google Scholar
Fowler, A. C., Murray, T., and Ng, F. S. L. (2001). Thermally controlled glacier sliding. Journal of Glaciology, 47(159), 527538.CrossRefGoogle Scholar
Fricker, H. A., Coleman, R., Padman, L., Scambos, T. A., Bohlander, J., and Brunt, K. M. (2009). Mapping the grounding zone of Amery Ice Shelf, East Antarctica using InSAR, MODIS, and ICESat. Antarctic Science, 21(5), 515532.Google Scholar
Fricker, H. A., and Padman, L. (2006). Ice shelf grounding zone structure from ICESat laser altimetry. Geophysical Research Letters, 33, L15502, doi: 10.1029/2006GL026907.Google Scholar
Fried, M. J., et al. (2015). Distributed subglacial discharge drives significant submarine melt at a Greenland tidewater glacier. Geophysical Research Letters, 42, 93289336.Google Scholar
Giovinetto, M. B., and Zwally, H. J. (2000). Spatial distribution of net surface accumulation on the Antarctic ice sheet. Annals of Glaciology, 31, 171178.Google Scholar
Glasser, N. F., and Scambos, T. A. (2008). A structural glaciological analysis of the 2002 Larsen B ice-shelf collapse. Journal of Glaciology, 54(184), 315.Google Scholar
Glasstone, S., Laidler, K. J., and Eyring, H. (1941). The Theory of Rate Processes. New York: McGraw-Hill.Google Scholar
Glen, J. W. (1955). The creep of polycrystalline ice. Proceedings of the Royal Society, Series A228(1175), 519538.Google Scholar
Glen, J. W. (1958). The flow law of ice. A discussion of the assumptions made in glacier theory, their experimental foundations and consequences. International Association of Scientific Hydrology, 47, 171183.Google Scholar
Glen, J. W. (1963). Contribution to the discussion. International Association of Scientific Hydrology Bulletin, 8(2), 68.Google Scholar
Gogineni, S., Chuah, T., Allen, C., Jezek, K., and Moore, R. K. (1998). An improved coherent radar depth sounder. Journal of Glaciology, 44(148), 659669.Google Scholar
Gold, L. W. (1958). Some observations on the dependence of strain on stress for ice. Canadian Journal of Physics, 36(10), 12651275.Google Scholar
Goldsby, D. (2009). Superplastic flow of ice relevant to glacier and ice-sheet mechanics. In Knight, P. (ed.) Glacier Science and Environmental Change. Oxford: Wiley-Blackwell, 527pp.Google Scholar
Goldsby, D. L., and Kohlstedt, D. L. (1997). Grain boundary sliding in fine-grained Ice I. Scripta Materialia, 37(9), 13991406.Google Scholar
Goldsby, D. L., and Kohlstedt, D. L. (2001). Superplastic deformation of ice: experimental observations. Journal of Geophysical Research, 106(B6), 1101711030.Google Scholar
Goldsby, D. L., and Kohlstedt, D. L. (2002). Reply to comment by P. Duval and M. Montagnat on “Superplastic deformation of ice: experimental observations”. Journal of Geophysical Research, 107(B11), 1101711030, doi: 10.1029/2002JB001842.Google Scholar
Goldsmith, H. L., and Mason, S. G. (1961). Axial migration of particles in Poiseuille flow. Nature, 190(4781), 10951096.CrossRefGoogle Scholar
Goldstein, R. M., Engelhardt, H., Kamb, B., and Frolich, R. M. (1993). Satellite radar interferometry for monitoring ice sheet motion: application to an Antarctic ice stream. Science, 262(5139), 15251530.Google Scholar
Goldthwait, R. P. (1951). Development of end moraines in east-central Baffin Island. Journal of Geology, 59(6), 567577.Google Scholar
Goujon, C., Barnola, J.-M., and Ritz, C. (2003). Modeling the densification of polar firn including heat diffusion: application to close off characteristics and gas isotopic fractionation for Antarctica and Greenland sites. Journal of Geophysical Research-Atmospheres, 108(D24), 4792, doi: 10.1029/2002JD003319.Google Scholar
Gow, A. J. (1969). On the rates of growth of grains and crystals in south polar firn. Journal of Glaciology, 8(53), 241252.Google Scholar
Gow, A. J., and Williamson, T. (1976). Rheological implications of the internal structure and crystal fabrics of the West Antarctic ice sheet as revealed by deep core drilling at Byrd Station. Geological Society of America Bulletin, 87, 16651677.Google Scholar
Gravenor, C. P. (1955). The origin and significance of prairie mounds. American Journal of Science, 253, 475481.Google Scholar
Gravenor, C. P., and Kupsch, W. O. (1959). Ice disintegration features in western Canada. Journal of Geology, 67, 4867.Google Scholar
Gravenor, C. P., and Meneley, W. A. (1958). Glacial flutings in central and northern Alberta. American Journal of Science, 256, 715728.Google Scholar
Greve, R., Zwinger, T., and Gong, Y. (2014). On the pressure dependence of the rate factor in Glen’s flow law. Journal of Glaciology, 60(220), 397398.Google Scholar
Griffith, A. A. (1921). The phenomena of rupture and flow in solids. Philosophical Transactions of the Royal Society of London, A221, 163197.Google Scholar
Griffith, A. A. (1924). Theory of rupture. Proceedings of the First International Congress on Applied Mechanics, Delft, 55–63.Google Scholar
Grootes, P. M., Stuiver, M., White, J. W. C., Johnsen, S., and Jouzel, J. (1993). Comparison of oxygen isotope records from the GISP 2 and GRIP Greenland ice cores. Nature, 366, 552554.Google Scholar
Grove, J. M. (1988). The Little Ice Age. London: Methuen.Google Scholar
Groves, G. W., and Kelly, A. (1969). Change of shape due to dislocation climb. Philosophical Magazine, 19, 977986.Google Scholar
Gudmundsson, G. H. (2007). Tides and the flow of Rutford Ice Stream, West Antarctica. Journal of Geophysical Research, 112, F04007, doi: 10.1029/2006JF000731.Google Scholar
Gudmundsson, G. H. (2011). Ice stream response to ocean tides and the form of the basal sliding law. The Cryosphere, 5, 259270.Google Scholar
Gudmundsson, G. H., Raymond, C. F., and Bindschadler, R. (1998). The origin and longevity of flow stripes on Antarctic ice streams. Annals of Glaciology, 27, 145152.Google Scholar
Gupta, P., Noone, D., Galewsky, J., Sweeney, C., and Vaughn, B. H. (2009). Demonstration of high-precision continuous measurements of water vapor isotopologues in laboratory and remote field deployments using wavelength-scanned cavity ring-down spectroscopy (WS-CRDS) technology. Rapid Communications in Mass Spectroscopy, 23, 25342542.Google Scholar
Haefeli, R. (1962). The ablation gradient and the retreat of a glacier tongue. In Symposium of Obergurgl, International Association of Scientific Hydrology, Publication 58, 49–59.Google Scholar
Hallet, B. (1976a). Deposits formed by subglacial precipitation of CaCO3. Geological Society of America Bulletin, 87(7), 10031015.Google Scholar
Hallet, B. (1976b). The effect of subglacial chemical processes on sliding. Journal of Glaciology, 17(76), 209221.Google Scholar
Hallet, B. (1979a). A theoretical model of glacial abrasion. Journal of Glaciology, 23(89), 3950.Google Scholar
Hallet, B. (1979b). Subglacial regelation water film. Journal of Glaciology, 23(89), 321334.Google Scholar
Hallet, B. (1981). Glacial abrasion and sliding: their dependence on the debris concentration in basal ice. Annals of Glaciology, 2, 2328.Google Scholar
Hallet, B. (1996). Glacial quarrying: a simple theoretical model. Annals of Glaciology, 22, 18.Google Scholar
Hallet, B., and Anderson, R. S. (1980). Detailed glacial geomorphology of a proglacial bedrock area at Castleguard Glacier, Alberta, Canada. Zeitschrift fur Gletscherkunde und Glazialgeologie, 16, 171184.Google Scholar
Hallet, B., Lorrain, R. D., and Souchez, R. A. (1978). The composition of basal ice from a glacier sliding over limestones. Geological Society of America Bulletin, 89(2), 314320.Google Scholar
Hamilton, G. S., Whillans, I., and Morgan, P. (1998). First point measurements of ice-sheet thickness change in Antarctica. Annals of Glaciology, 27, 125129.Google Scholar
Hamilton, W. C., and Ibers, J. A. (1968). Hydrogen Bonding in Solids; Methods of Molecular Structure Determination. New York: W. A. Benjamin.Google Scholar
Hammer, C. U. (1980). Acidity of polar ice cores in relation to absolute dating, past volcanism, and radio-echoes. Journal of Glaciology, 25(93), 359372.Google Scholar
Hanna, E., Cappelen, J., Fettweis, X., Huybrechts, P., Luckman, A., and Ribergaard, M. H. (2009). Hydrologic response of the Greenland ice sheet: the role of oceanographic warming. Hydrological Processes, 23, 730.Google Scholar
Hanson, B. (1995). A fully three-dimensional finite-element model applied to velocities on Storglaciären, Sweden. Journal of Glaciology, 41(137), 91102.Google Scholar
Hanson, B., and Hooke, R. LeB. (2000). A model study of the forces involved in glacier calving. Journal of Glaciology, 46(153), 188194.Google Scholar
Hanson, B., and Hooke, R. LeB. (2003). Buckling rate and overhang development at a calving face. Journal of Glaciology, 49(167), 577586.Google Scholar
Hanson, B., Hooke, R. LeB., and Grace, E. M. Jr. (1998). Short-term velocity and water-pressure measurements down-glacier from a riegel, Storglaciären, Sweden. Journal of Glaciology, 44(147), 359367.Google Scholar
Harrison, W. (1958). Marginal zones of vanished glaciers reconstructed from preconsolidation-pressure values of overridden silts. The Journal of Geology, 66(1), 7295.Google Scholar
Harrison, W. D. (1972). Temperature of a temperate glacier. Journal of Glaciology, 11(61), 1529.Google Scholar
Harrison, W. D., Echelmeyer, K. A., and Larsen, C. F. (1998). Measurement of temperature in a margin of Ice Stream B, Antarctica: implications for margin migration and lateral drag. Journal of Glaciology, 44(148), 615624.Google Scholar
Harrison, W. D., Raymond, C. F., Echelmeyer, K. A., and Krimmel, R. M. (2003). A macroscopic approach to glacier dynamics. Journal of Glaciology, 49(164), 1321.Google Scholar
Hättestrand, C., and Kleman, J. (1999). Ribbed moraine formation. Quaternary Science Reviews, 18, 4361.Google Scholar
Hausmann, M. R. (1990). Engineering Principles of Ground Modification. New York: McGraw-Hill.Google Scholar
Hays, J. D., Imbrie, J., and Shackleton, N. S. (1976). Variations in the Earth’s orbit: pacemaker of the ice ages. Science, 194(4270), 11211132.Google Scholar
Hebrand, M., and Åmark, M. (1989). Esker formation and glacier dynamics in eastern Skåne and adjacent areas, southern Sweden. Boreas, 18(1), 6781.Google Scholar
Higgins, J. A., et al. (2014). Atmospheric composition 1 million years ago from blue ice in the Allan Hills, Antarctica. Proceedings of the National Academy of Science, USA, 112(22), 68876891.Google Scholar
Hindmarsh, R. C. A. (1998). Drumlinization and drumlin-forming instabilities: viscous till mechanisms. Journal of Glaciology, 44(147), 293314.Google Scholar
Hobbs, P. V. (1974). Ice Physics. New York: Oxford Clarendon Press.Google Scholar
Hock, R., and Hooke, R. LeB. (1993). Further tracer studies of internal drainage in the lower part of the ablation area of Storglaciären, Sweden. Geological Society of America Bulletin, 105(4), 537546.Google Scholar
Hodge, S. M., Trabant, D. C., Krimmel, R. M., Heinrichs, T. A., March, R. S., and Joshberger, E. G. (1998). Climate variations and changes in mass balance of three glaciers in western North America. Journal of Climate, 11, 21612179.Google Scholar
Hoffman, M. J., et al. (2016). Greenland subglacial drainage evolution regulated by weakly connected regions of the bed. Nature Communications, 7(13903), doi: 10.1038/ncomms13903.Google Scholar
Hogg, A. E., and Gudmundsson, G. H. (2017). Impacts of the Larsen-C ice shelf calving event. Nature Climate Change, 7, 540542.Google Scholar
Holland, P. R., Corr, H. F. J., Vaughan, D. G., Jenkins, A., and Skvarca, P. (2009). Marine ice in Larsen ice shelf. Geophysical Research Letters, 36, L11604, doi: 10.1029/2009GL038162.Google Scholar
Holmlund, P. (1987). Mass balance of Storglaciären during the 20th century. Geografiska Annaler, 69A(3–4), 439447.Google Scholar
Holmlund, P. (1988). Internal geometry and evolution of moulins, Storglaciären, Sweden. Journal of Glaciology, 34(117), 242248.Google Scholar
Hong, S., Candelone, J.-P., Patterson, C. C., and Boutron, C. F. (1994). Greenland ice evidence of hemispheric lead pollution two millennia ago by Greek and Roman civilizations. Science, 265, 18411843.Google Scholar
Hooke, R. LeB. (1970). Morphology of the ice-sheet margin near Thule, Greenland. Journal of Glaciology, 9(57), 303324.Google Scholar
Hooke, R. LeB. (1973a). Flow near the margin of the Barnes Ice Cap and the development of ice-cored moraines. Geological Society of America Bulletin, 84(12), 39293948.Google Scholar
Hooke, R. LeB. (1973b). Structure and flow in the margin of Barnes Ice Cap, Baffin Island, N.W.T., Canada. Journal of Glaciology, 12(66), 423438.Google Scholar
Hooke, R. LeB. (1976). Pleistocene ice at the base of the Barnes Ice Cap, Baffin Island, N.W.T., Canada. Journal of Glaciology, 17(75), 4960.Google Scholar
Hooke, R. LeB. (1981). Flow law for polycrystalline ice in glaciers: comparison of theoretical predictions, laboratory data, and field measurements. Reviews of Geophysics and Space Physics, 19(4), 664672.Google Scholar
Hooke, R. LeB. (1984). On the role of mechanical energy in maintaining subglacial conduits at atmospheric pressure. Journal of Glaciology, 30(105), 180187.Google Scholar
Hooke, R. LeB. (1989). Englacial and subglacial hydrology: a qualitative review. Arctic and Alpine Research, 21(3), 221233.Google Scholar
Hooke, R. LeB. (1991). Positive feedbacks associated with the erosion of glacial cirques and overdeepenings. Geological Society of America Bulletin, 103(8), 11041108.Google Scholar
Hooke, R. LeB., Alexander, E. C. Jr., and Gustafson, R. J. (1980). Temperature profiles in Barnes Ice Cap, Baffin Island, Canada, and heat flux from the subglacial terrane. Canadian Journal of Earth Sciences, 17(9), 11741188.Google Scholar
Hooke, R. LeB., Calla, P., Holmlund, P., Nilsson, M., and Stroeven, A. (1989). A three‑year record of seasonal variations in surface velocity, Storglaciären, Sweden. Journal of Glaciology, 35(120), 235247.CrossRefGoogle Scholar
Hooke, R. LeB., and Clausen, H. B. (1982). Wisconsin and Holocene δ18O variations, Barnes Ice Cap, Canada. Geological Society of America Bulletin, 93(8), 784789.Google Scholar
Hooke, R. LeB., Cummings, D. I., Lesemann, J.-E., and Sharpe, D. R. (2013). Genesis of dispersal plumes in till. Canadian Journal of Earth Sciences, 50(8), 847855.Google Scholar
Hooke, R. LeB., Dahlin, B. B., and Kauper, M. T. (1972). Creep of ice containing dispersed fine sand. Journal of Glaciology, 11(63), 327336.Google Scholar
Hooke, R. LeB., and Elverhøi, A. (1996). Sediment flux from a fjord during glacial periods, Isfjorden, Spitsbergen. Global and Planetary Change, 12, 237249.Google Scholar
Hooke, R. LeB., and Fastook, J. (2007). Thermal conditions at the bed of the Laurentide Ice Sheet in Maine during deglaciation: implications for esker formation. Journal of Glaciology, 53(183) 646658.Google Scholar
Hooke, R. LeB., Gould, J. E., and Brzozowski, J. (1983). Near-surface temperatures near and below the equilibrium line on polar and subpolar glaciers. Zeitschrift für Gletscherkunde und Glazialgeologie, 19(1), 125.Google Scholar
Hooke, R. LeB., and Hanson, B. H. (1986). Borehole deformation experiments, Barnes Ice Cap, Canada. Cold Regions Science and Technology, 12(3), 261276.Google Scholar
Hooke, R. LeB., Hanson, B., Iverson, N. R., Jansson, P., and Fischer, U. H. (1997). Rheology of till beneath Storglaciären, Sweden. Journal of Glaciology, 43(143), 172179.Google Scholar
Hooke, R. LeB., and Hudleston, P. J. (1978). Origin of foliation in glaciers. Journal of Glaciology, 20(83), 285299.Google Scholar
Hooke, R. LeB., and Hudleston, P. J. (1980). Ice fabrics in a vertical flowplane, Barnes Ice Cap, Canada. Journal of Glaciology, 25(92), 195214.Google Scholar
Hooke, R. LeB., and Hudleston, P. J. (1981). Ice fabrics from a borehole at the top of the South Dome, Barnes Ice Cap, Baffin Island. Geological Society of America Bulletin, 92(5), 274281.Google Scholar
Hooke, R. LeB., and Iverson, N. R. (1995). Grain size distribution in deforming subglacial tills: role of grain fracture. Geology, 23(1), 5760.Google Scholar
Hooke, R. LeB., and Jennings, C. (2006). On the formation of tunnel valleys. Quaternary Science Reviews, 25(11–12), 13641372.Google Scholar
Hooke, R. LeB., Johnson, G. W., Brugger, K. A., Hanson, B., and Holdsworth, G. (1987). Changes in mass balance, velocity, and surface profile along a flow line on Barnes Ice Cap. (1970–1984). Canadian Journal of Earth Sciences, 24(8), 15501561.Google Scholar
Hooke, R. LeB., Laumann, T., and Kohler, J. (1990). Subglacial water pressures and the shape of subglacial conduits. Journal of Glaciology, 36(122), 6771.Google Scholar
Hooke, R. LeB., and Medford, A. (2013). Are drumlins a result of a thermo-mechanical instability? Quaternary Research, 79(3), 458464.Google Scholar
Hooke, R. LeB., and Pohjola, A. (1994). Hydrology of a segment of a glacier situated in an overdeepening, Storglaciären, Sweden. Journal of Glaciology, 40(134), 140148.Google Scholar
Hooke, R. LeB., Pohjola, V., Jansson, P., and Kohler, J. (1992). Intra-seasonal changes in deformation profiles revealed by borehole studies, Storglaciären, Sweden. Journal of Glaciology, 38(130), 348358.Google Scholar
Hooke, R. LeB., Wold, B., and Hagen, J. O. (1985). Subglacial hydrology and sediment transport at Bondhusbreen, southwest Norway. Geological Society of America Bulletin, 96(3), 388397.Google Scholar
Hooyer, T. S., and Iverson, N. R. (2002). Flow mechanism of the Des Moines Lobe of the Laurentide ice sheet. Journal of Glaciology, 48(163), 575586.Google Scholar
Houghton, J. T., et al. (2001). Climate Change 2001: The Scientific Basis. Report of the Intergovernmental Panel on Climate Change (IPCC). Cambridge: Cambridge University Press, 881pp.Google Scholar
Howell, D., Behringer, R. P., and Veje, C. (1999). Stress fluctuations in a 2D granular Couette experiment: a continuous transition. Physical Review Letters, 82(96), 52415244.Google Scholar
Hudleston, P. J. (1976). Recumbent folding in the base of the Barnes Ice Cap, Baffin Island, Northwest Territories, Canada. Geological Society of America Bulletin, 87(12), 16781683.Google Scholar
Hudleston, P. J. (1989). The association of folds and veins in shear zones. Journal of Structural Geology, 11(8), 949957.Google Scholar
Hudleston, P. J. (2015). Structures and fabric in glacial ice: a review. Journal of Structural Geology, 81, 127.Google Scholar
Hudleston, P. J., and Hooke, R. LeB. (1980). Cumulative deformation in the Barnes Ice Cap and implications for the development of foliation. Tectonophysics, 66, 127146.Google Scholar
Hughes, T. (1992). Abrupt climate change related to unstable ice-sheet dynamics: toward a new paradigm. Palaeogeography, Palaeoclimatology, Palaeoecology, 97, 203234.Google Scholar
Hulbe, C., Joughin, I., Morse, D., and Bindschadler, R. A. (2000). Tributaries to West Antarctic ice streams: characteristics deduced from numerical modelling of ice flow. Annals of Glaciology, 31, 184190.Google Scholar
Hull, D. (1969). Introduction to Dislocations. New York: Pergamon Press.Google Scholar
Humphrey, N. F., and Raymond, C. F. (1994). Hydrology, erosion and sediment production in a surging glacier: Variegated Glacier, Alaska, 1982–83. Journal of Glaciology, 40(136), 539552.Google Scholar
Husband, D. M., Mondy, L. A., Ganani, E., and Graham, A. L. (1994). Direct measurements of shear-induced particle migration in suspensions of bimodal spheres. Rheologica Acta, 33(3), 185192.Google Scholar
Hutchinson, J. W. (1976). Bounds and self-consistent estimates for creep of polycrystalline materials. Proceedings of the Royal Society, Series A., 348, 101127.Google Scholar
Hutter, K. (1981). The effect of longitudinal strain on the shear stress of an ice sheet. In defense of using stretched coordinates. Journal of Glaciology, 27(95), 3956.Google Scholar
Hutter, K. (1983). Theoretical glaciology. Tokyo, Japan: D. Reidel Publishing, 510pp.Google Scholar
Huybrechts, Ph. (1990). A 3-D model for the Antarctic ice sheet: a sensitivity study on the glacial-interglacial contrast. Climate Dynamics, 5, 7992.CrossRefGoogle Scholar
Huybrechts, Ph. (2002). Sea-level changes at the LGM from ice-dynamic reconstructions of the Greenland and Antarctic ice sheets during the glacial cycles. Quaternary Science Reviews, 21(1-3), 203231.Google Scholar
Huybrechts, P., de Nooze, P., and Decleir, H. (1989). Numerical modeling of Glacier d’Argentière and its historical front variations. In Oerlemans, J. (ed.) Glacier Fluctuations and Climatic Change. The Netherlands: Kluwer Academic, pp. 373389.Google Scholar
Huybrechts, P., Payne, T., and The EISMINT Intercomparison Group. (1996). The EISMINT benchmarks for testing ice-sheet models. Annals of Glaciology, 23, 112.Google Scholar
Huybrechts, P., Steinhage, D., Wilhelms, F., and Bamber, J. (2000). Balance velocities and measured properties of the Antarctic ice sheet from a new compilation of gridded data for modelling. Annals of Glaciology, 30, 5260.Google Scholar
Huybrechts, P., and T’Siobbel, S. (1995). Thermomechanical modeling of northern hemisphere ice sheets with a two-level mass balance parameterization. Annals of Glaciology, 21, 111117.Google Scholar
Iken, A. (1981). The effect of subglacial water pressure on the sliding velocity of a glacier in an idealized numerical model. Journal of Glaciology, 27(97), 407421.Google Scholar
Iken, A., and Bindschadler, R. A. (1986). Combined measurements of subglacial water pressure and surface velocity of Findelengletscher, Switzerland: conclusions about the drainage system and sliding mechanism. Journal of Glaciology, 32(110), 101119.Google Scholar
Iken, A., and Truffer, M. (1997). The relationship between subglacial water pressure and velocity of Findelengletscher during its advance and retreat. Journal of Glaciology, 43(144), 328338.Google Scholar
Irons, B. M., and Shrive, N. G. (1987). Numerical Methods in Engineering and Applied Science: Numbers are Fun. New York: John Wiley & Sons, 248pp.Google Scholar
Iverson, N. A., Kalteyer, D., Dunbar, N. W., Kurbatov, A., and Yates, M. (2017). Advancements and best practices for analysis and correlation of tephra and cryptotephra in ice. Quaternary Geochronology, 40, 4555.Google Scholar
Iverson, N. R. (1989). Theoretical and experimental analyses of glacial abrasion and quarrying. PhD thesis, University of Minnesota, Minneapolis, 233pp.Google Scholar
Iverson, N. R. (1991a). Morphology of glacial striae: implications for abrasion of glacier beds and fault surfaces. Geological Society of America Bulletin, 103, 13081316.Google Scholar
Iverson, N. R. (1991b). Potential effects of subglacial water-pressure fluctuations on quarrying. Journal of Glaciology, 37(125), 2736.Google Scholar
Iverson, N. R. (1993). Regelation of ice through debris at glacier beds: implications for sediment transport. Geology, 21(6), 559562.Google Scholar
Iverson, N. R. (1999). Coupling between a glacier and a soft bed: II. Model results. Journal of Glaciology, 45(149), 4153.Google Scholar
Iverson, N. R. (2000). Sediment entrainment by a soft-bedded glacier: a model based on regelation into the bed. Earth Surface Processes and Landforms, 25, 881893.Google Scholar
Iverson, N. R. (2010). Shear resistance and continuity of subglacial till: hydrology rules. Journal of Glaciology, 56(200), 11041114.Google Scholar
Iverson, N. R. (2012). A theory of glacial quarrying for landscape evolution models. Geology, 40(8), 679682.Google Scholar
Iverson, N. R., Baker, R. W., and Hooyer, T. S. (1997). A ring-shear device for the study of till deformation: tests on tills with contrasting clay contents. Quaternary Science Reviews, 16, 10571066.Google Scholar
Iverson, N. R., et al. (2003). Effects of basal debris on glacier flow. Science, 301, 8184.Google Scholar
Iverson, N. R., and Hooyer, T. S. (2004). Estimating the sliding velocity of a Pleistocene ice sheet from plowing structures in the geologic record. Journal of Geophysical Research, 109, F04006, doi: 10.1029/2004JF000132.Google Scholar
Iverson, N. R., Hooyer, T. S., and Baker, R. W. (1998). Ring-shear studies of till deformation: coulomb-plastic behavior and distributed strain in glacier beds. Journal of Glaciology, 44(148), 634642.Google Scholar
Iverson, N. R., et al. (2007). Soft-bed experiments beneath Engabreen, Norway: regelation infiltration, basal slip and bed deformation. Journal of Glaciology, 53(182), 323340.Google Scholar
Iverson, N. R., Hooyer, T. S., and Hooke, R. LeB. (1996). A laboratory study of sediment deformation: stress heterogeneity and grain-size evolution. Annals of Glaciology, 22, 167175.Google Scholar
Iverson, N. R., and Iverson, R. M. (2001). Distributed shear of subglacial till due to Coulomb slip. Journal of Glaciology, 47(158), 481488.Google Scholar
Iverson, N. R., and Petersen, B. B. (2011). A new laboratory device for study of subglacial processes: first results on ice-bed separation during sliding. Journal of Glaciology, 57(206), 11351146, doi: 10.3189/002214311798843458.Google Scholar
Iverson, N. R., and Slemmons, D. J. (1995). Intrusion of ice into porous media by regelation: a mechanism of sediment entrainment by glaciers. Journal of Geophysical Research, 100(B7), 1021910230.Google Scholar
Iverson, N. R., and Souchez, R. (1996). Isotopic signature of debris-rich ice formed by regelation into a subglacial sediment bed. Geophysical Research Letters, 23(10), 11511154.Google Scholar
Jacka, T. H., and Budd, W. F. (1989). Isotropic and anisotropic flow relations for ice dynamics. Annals of Glaciology, 12(1), 8184, doi: 10.3198/1989AoG12-1-81-84.Google Scholar
Jacka, T. H., and Maccagnan, M. (1984). Ice crystallographic and strain rate changes with strain in compression and extension. Cold Regions Science and Technology, 8(3), 269286.Google Scholar
Jackson, M., and Kamb, B. (1997). The marginal shear stress of Ice Stream B, West Antarctica. Journal of Glaciology, 43(145), 415426.Google Scholar
Jacobel, R. W., Scambos, T. A., Nereson, N. A., and Raymond, C. F. (2000). Changes in the margin of Ice Stream C, Antarctica. Journal of Glaciology, 46(152), 102110.Google Scholar
Jacobel, R. W., Scambos, T. A., Raymond, C. F., and Gades, A. M. (1996). Changes in the configuration of ice stream flow from the West Antarctic Ice Sheet. Journal of Geophysical Research, 101(B3), 54995504.Google Scholar
Jacobson, H. P., and Raymond, C. F. (1998). Thermal effects on the location of ice stream margins. Journal of Geophysical Research, 103(B9), 1211112122.Google Scholar
Jansson, E. P. (1995). Water pressure and basal sliding on Storglaciären, northern Sweden. Journal of Glaciology, 41(138), 232240.Google Scholar
Jenkins, A. (2011). Convection-driven melting near the grounding lines of ice shelves and tidewater glaciers. Journal of Physical Oceanography, 41, 22792294.Google Scholar
Jóhannesson, T., Raymond, C. F., and Waddington, E. (1989). Time-scale for adjustment of glaciers to changes in mass balance. Journal of Glaciology, 35(121), 355369.Google Scholar
Johnsen, S. J., Dansgaard, W., Clausen, H. B., and Langway, C. C. Jr. (1972). Oxygen isotope profiles through the Antarctic and Greenland ice sheets. Nature, 235(5339), 429434.Google Scholar
Johnson, A. (1960). Variation in surface elevation of the Nisqually glacier Mt. Rainier, Washington. International Association of Scientific Hydrology Bulletin, 19, 5460.Google Scholar
Johnson, W., and Mellor, P. B. (1962). Plasticity for Mechanical Engineers. Princeton, NJ: Van Nostrand, Ltd., 412pp. (There is also a 1973 edition, in which the relevant pages are 44–49).Google Scholar
Jones, N. (1982). The formation of glacial flutings in east-central Alberta. Proceedings - Guelph Symposium on Geomorphology, 6, 4970.Google Scholar
Jones, S. (1982). The confined compressive strength of polycrystalline ice. Journal of Glaciology, 28(98), 171177.Google Scholar
Joughin, I., et al. (1999). Tributaries of West Antarctic ice streams revealed by RADARSAT interferometry. Science, 286, 283286.Google Scholar
Joughin, I., et al. (2005). Continued deceleration of Whillans Ice Stream. Geophysical Research Letters, 32, L22501, doi: 10.1029/2005GL024319.Google Scholar
Joughin, I., MacAyeal, D. R., and Tulaczyk, S. (2004). Basal shear stress of the Ross ice streams from control method inversions. Journal of Geophysical Research, 109, B09405, doi: 10.1029/2003JB002960.Google Scholar
Jouzel, J. (2013). A brief history of ice core science over the last 50 yr. Climate of the Past, 9, 25252547.Google Scholar
Kamb, B. (1959). Ice petrofabric observations from Blue Glacier, Washington, in relation to theory and experiment. Journal of Geophysical Research, 64(11), 18911909.Google Scholar
Kamb, B. (1965). Structure of ice VI. Science, 150(3693), 205209.Google Scholar
Kamb, B. (1970). Sliding motion of glaciers: theory and observation. Reviews of Geophysics and Space Physics, 8(4), 673728.Google Scholar
Kamb, B. (1972). Experimental recrystallization of ice under stress. In Heard, H. C., Borg, I. Y., Carter, N. L., and Raleigh, C. B. (eds.) Flow and Fracture of Rocks. Washington, DC: American Geophysical Union, 16, 211241.Google Scholar
Kamb, B. (1987). Glacier surge mechanism based on linked cavity configuration of the basal water conduit system. Journal of Geophysical Research, 92(B9), 90839100.Google Scholar
Kamb, B. (2001). Basal zone of the West Antarctic ice streams and its role in lubrication of their rapid motion. In Alley, R. B., and Bindschadler, R. A. (eds.) The West Antarctic Ice Sheet: Behavior and Environment. Washington, DC: American Geophysical Union, Antarctic Research Series, 77, 157201.Google Scholar
Kamb, B., and Echelmeyer, K. A. (1986). Stress-gradient coupling in glacier flow: 1. Longitudinal averaging of the influence of ice thickness and surface slope. Journal of Glaciology, 32(111), 267284.Google Scholar
Kamb, B., et al. (1985). Glacier surge mechanism: 1982–1983 surge of Variegated Glacier, Alaska. Science, 227(4686), 469479.Google Scholar
Kamb, B., and LaChapelle, E. (1964). Direct observation of the mechanism of glacier sliding over bedrock. Journal of Glaciology, 5(38), 159172.Google Scholar
Kanninen, M. F., and Popelar, C. H. (1985). Advanced Fracture Mechanics. New York: Oxford University Press, 563pp.Google Scholar
Kaspari, S., Hooke, R. LeB., Mayewski, P. A., Kang, S., Hou, S., and Qin, D. (2008). Snow accumulation rate on Qomolangma (Mt. Everest), Himalaya: synchroneity with sites across the Tibetan Plateau on 50–100 year timescales. Journal of Glaciology, 54(185), 343352.Google Scholar
Kaspari, S., Mayewski, P. A., Dixon, D. A., Spikes, V. B., Sneed, S. B., Handley, M. J., and Hamilton, G. S. (2004). Climate variability in West Antarctica derived from annual accumulation rate records from ITASE Firn/ice cores. Annals of Glaciology, 39, 585594.Google Scholar
Kaufman, B. G., et al. (2014). Centennial-scale variability of the Southern Hemisphere westerly wind belt in the eastern Pacific over the past two millennia. Climate of the Past, 10, 11251144.Google Scholar
Kawamura, K., et al. (2007). Northern Hemisphere forcing of climatic cycles in Antarctica over the past 360 000 years. Nature, 448, 912916.Google Scholar
Keegan, K. M., Albert, M. R., McConnell, J. R., and Baker, I. (2014). Climate change and forest fires synergistically drive widespread melt events of the Greenland Ice Sheet, Proceedings to the National Academy of Sciences, May 19, 2014, www.pnas.org/cgi/doi/10.1073/ pnas.1405397111.Google Scholar
Kendall, K. (1978). The impossibility of comminuting small particles by compression. Nature, 272, 710711.Google Scholar
Kenneally, J. (2003). Crevassing and calving of glacial ice. PhD thesis, University of Maine, Orono, 145pp.Google Scholar
Kenneally, J. P., and Hughes, T. (2006). Calving giant icebergs: old principles, new applications. Antarctic Science, 18(3), 409419.Google Scholar
Ketcham, W. M., and Hobbs, P. V. (1969). An experimental determination of the surface energies of ice. Philosophical Magazine, 19(162), 11611173.Google Scholar
Kimura, S., Nichols, K. W., and Venables, E. (2015). Estimation of shelf melt rates in the presence of a thermohaline staircase. Journal of Physical Oceanography, 45(1), 133148.Google Scholar
Kinosita, S. (1962). Transformation of snow into ice by plastic compression. Low Temperature Science, A, 20, 131157.Google Scholar
Kirby, S. H., Durham, W. B., Beeman, M. L., Heard, H. C., and Daley, M. A. (1987). Inelastic properties of ice Ih at low temperatures and high pressures. Journal de Physique Colloques, 48(C1), C1-227C1-232.Google Scholar
Kirkbride, M. P. (2011). Debris-covered glaciers. In Singh, V. P., Singh, P., and Haritashya, U. K. (eds.) Encyclopedia of Snow, Ice and Glaciers. Dordrecht: Springer, pp. 190192, doi: 10.1007/978-90-481-2642-2.Google Scholar
Kleman, J., and Borgström, I. (1994). Glacial landforms indicative of a partly frozen bed. Journal of Glaciology, 40(135), 255264.Google Scholar
Kleman, J., and Hättestrand, C. (1999). Frozen-bed Fennoscandian and Laurentide ice sheets during the Last Glacial Maximum. Nature, 402, 6366.Google Scholar
Kranenborg, E. J., and Dijkstra, H. A. (1998). On the evolution of double-diffusive intrusions into a stably stratified liquid: a study of the layer merging process. International Journal of Heat and Mass Transfer, 41(18), 27432756.Google Scholar
Kruk, N. S., Vendrame, I. F., Rocha, H. R, Chou, S. C., and Cabral, O. (2010). Downward longwave radiation estimates for clear and all-sky conditions in the Sertãozinho region of São Paulo, Brazil. Theoretical and Applied Climatology, 99, 115123.Google Scholar
Kuhn, M. (1981). Climate and glaciers. International Association of Scientific Hydrology, 131, 320.Google Scholar
Kuhn, M. (1989). The response of the equilibrium line altitude to climate fluctuations: theory and observations. In Oerlemans, J. (ed.) Glacier Fluctuations and Climatic Change. Dordrecht: Kluwer Academic Publishers, pp. 407417.Google Scholar
Lambe, T. W., and Whitman, R. V. (1969). Soil Mechanics. New York: John Wiley, 553pp.Google Scholar
Landais, A., et al. (2010). What drives millennial and orbital variations of δ18Oatm? Quaternary Science Reviews, 29, 235246.Google Scholar
Langdon, T. G. (1991). The physics of superplastic deformation. Materials Science and Engineering, A137, 111.Google Scholar
Lawn, B. (1993). Fracture of Brittle Solids, 2nd ed. Cambridge: Cambridge University Press, 378pp.Google Scholar
Lawson, D. E., Strasser, J. C., Evenson, E. B., Alley, R. B., Larson, G. J., and Arcone, S. A. (1998). Glaciohydraulic supercooling: a freeze-on mechanism to create stratified, debris-rich basal ice: I. Field evidence. Journal of Glaciology, 44(148), 547562.Google Scholar
Lazzara, M. A., Jezek, K. C., Scambos, T. A., MacAyeal, D. R., and van der Veen, C. J. (1999). On the recent calving of icebergs from the Ross Ice Shelf. Polar Geography, 23(3), 201212.Google Scholar
Lazzara, M. A., Jezek, K. C., Scambos, T. A., MacAyeal, D. R., and van der Veen, C. J. (2008). On the recent calving of icebergs from Ross Ice Shelf. Polar Geography, 31(1–2), 1526.Google Scholar
Legrand, M., and Mayewski, P. A. (1997). Glaciochemistry of polar ice cores: a review. Reviews of Geophysics, 35(3), 219243.Google Scholar
Lemieux-Dudon, B., et al. (2010). Consistent dating for Antarctic and Greenland ice cores. Quaternary Science Reviews, 29, 820.Google Scholar
Leonard, K. C., Bell, R. E., and Studinger, M. (2003). The influence of subglacial topography on accumulation rates at Lake Vostok. American Geophysical Union Annual Meeting, December 7–12, 2003.Google Scholar
Leonard, K. C., and Fountain, A. G. (2003). Map-based methods for estimating glacier equilibrium line altitudes. Journal of Glaciology, 49(166), 329336.Google Scholar
Levermann, A., Albrecht, T., Winkelmann, R., Martin, M. A., Haseloff, M., and Joughin, I. (2012). Kinematic first-order calving law implies potential for abrupt ice-shelf retreat. The Cryosphere, 6, 273286.Google Scholar
Lewis, E. L., and Perkin, R. G. (1983). Supercooling and energy exchange near the Arctic Ocean surface. Journal of Geophysical Research, 88(C12) 76817685.Google Scholar
Lewis, E. L., and Perkin, R. G. (1986). Ice pumps and their rates. Journal of Geophysical Research, 91(C10) 1175611762.Google Scholar
Li, J., Jacka, T. H., and Budd, W. F. (1996). Deformation rates in combined compression and shear for ice which is initially isotropic and after the development of strong anisotropy. Annals of Glaciology, 23, 247252.Google Scholar
Lighthill, M. J., and Whitham, G. B. (1955). On kinematic waves, I. Flood movement in long rivers. Proceedings of the Royal Society, Series A, 229(1178), 281316.Google Scholar
Lile, R. C. (1978). The effect of anisotropy on the creep of polycrystalline ice. Journal of Glaciology, 21(85), 475483.Google Scholar
Lipenkov, V. Ya., Barkov, N. I., Duval, P., and Pimienta, P. (1989). Crystalline texture of the 2083 m ice core at Vostok station, Antarctica. Journal of Glaciology, 35(121), 392398.Google Scholar
Lisiecki, L. E., and Raymo, M. E. (2005). A Pleistocene-Pliocene stack of 57 globally distributed benthic δ18O records. Paleoceanography, 20, PA1003, doi: 10.1029/2004PA001071.Google Scholar
Liu, C.-H., Nagel, S. R., Schecter, D. A., Coppersmith, S. N., Majumdar, S., Narayan, O., and Witten, T. A. (1995). Force fluctuations in bead packs. Science, 269(5223), 513515.Google Scholar
Liu, H., Jezek, K. C., and Li, B. (1999). Development of an Antarctic DEM database by integrating cartographic and remotely sensed data: a GIS approach. Journal of Geophysical Research, 104(B10), 2319923213.Google Scholar
Lliboutry, L. (1964). Traité de Glaciologie, Vol 1. Paris: Masson and Co.Google Scholar
Lliboutry, L. (1968). General theory of subglacial cavitation and sliding of temperate glaciers. Journal of Glaciology, 7(49), 2158.Google Scholar
Lliboutry, L. (1970). Ice flow law from ice sheet dynamics. Proceedings of the International Symposium on Antarctic Glaciological Exploration, Hanover, NH, 3-7 September, 1968; International Association of Scientific Hydrology, Publication no. 86, 216228.Google Scholar
Lliboutry, L. (1976). Physical processes in temperate glaciers. Journal of Glaciology, 16(74), 151158.Google Scholar
Lliboutry, L. (1979). Local friction laws for glaciers: a critical review and new openings. Journal of Glaciology, 23(89), 6795.Google Scholar
Lliboutry, L. (1983). Modifications to the theory of intraglacial waterways for the case of subglacial ones. Journal of Glaciology, 29(102), 216226.Google Scholar
Lliboutry, L. (1996). Temperate ice permeability, stability of water veins and percolation of internal meltwater. Journal of Glaciology, 42(141), 201211.Google Scholar
Loewe, F. (1970). Screen temperatures and 10 m temperatures. Journal of Glaciology, 9(56), 263268.Google Scholar
Loulergue, L., et al. (2007). New constraints on the gas age-ice age difference along the EPICA ice cores, 0-50 kyr. Climate of the Past, 3, 527540.Google Scholar
Luckman, A., Benn, D. I., Cottier, F., Bevan, S., Nilsen, F., and Inall, M. (2015). Calving rates at tidewater glaciers vary strongly with ocean temperature. Nature Communications, 6(8566), doi: 10.1038/ncomms9566.Google Scholar
Lűthi, D., et al. (2008). High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature, 453, 379382.Google Scholar
MacAyeal, D. R. (1993a). A low-order model of the Heinrich event cycle. Paleoceanography, 8(6), 767773.Google Scholar
MacAyeal, D. R. (1993b). Binge/purge oscillations of the Laurentide Ice Sheet as a cause of the North Atlantic’s Heinrich events. Paleoceanography, 8(6), 775784.Google Scholar
MacAyeal, D. R., Bindschadler, R. A., and Scambos, T. A. (1995). Basal friction of Ice Stream E, West Antarctica. Journal of Glaciology, 41(138), 247262.Google Scholar
MacAyeal, D. R., Scambos, T. A., Hulbe, C. L., and Fahnestock, M. A. (2003). Catastrophic ice-shelf break-up by an ice-shelf-fragment-capsize mechanism. Journal of Glaciology, 49(164), 2236.Google Scholar
Macheret, Y. Y., Cherkasov, P. A., and Bobrova, L. I. (1988). Tolschina i ob’em lednikov Djungarskogo Alatau po danniy aeroradiozondirovaniya. (Thickness and volume of glaciers of the Dzhungar Mountains according to results of aero-radio-sounding). Materialy Glyatsiologicheskikh Issled. Khronika Obsuzhdeniya, 62, 5971.Google Scholar
Machguth, H., Purves, R. S., Oerlemans, J., Hoelzle, M., and Paul, F. (2008). Exploring uncertainty in glacier mass balance modelling with Monte Carlo simulation. The Cryosphere, 2, 191204, doi: 10.5194/tc-2-191-2008.Google Scholar
Makinson, K., Holland, P. R., Jenkins, A., Nicholls, K. W., and Holland, D. M. (2011). Influence of tides on melting and freezing beneath Filchner-Ronne Ice Shelf, Antarctica. Geophysical Research Letters, 38, L06601, doi: 10.1029/2010GL046462.Google Scholar
Makinson, K., King, M. A., Nicholls, K. W., and Gudmundsson, G. H. (2012). Diurnal and semidiurnal tide-induced lateral movement of Ronne Ice Shelf, Antarctica. Geophysical Research Letters, 39, L10501, doi: 10.1029/2012GL051636.Google Scholar
Mandl, G., de Jong, L. N. J., and Maltha, A. (1977). Shear zones in granular material – an experimental study of their structure and mechanical genesis. Rock Mechanics, 9(2-3), 95144.Google Scholar
Mantua, N. J., and Hare, S. R. (2002). The Pacific decadal oscillation. Journal of Oceanography, 58, 3544.Google Scholar
Marshall, E. W. (1959). Stratigraphic use of particulates in polar ice caps. Geological Society of America Bulletin, 70, 1643.Google Scholar
Marshall, S. J., Tarasov, L., Clarke, G. K. C., and Peltier, W. R. (2000). Glaciological reconstruction of the Laurentide Ice Sheet: physical processes and modeling challenges. Canadian Journal of Earth Sciences, 37, 769793.Google Scholar
Martín, C., Gudmundsson, G. H., Pritchard, H. D., and Gagliardini, O. (2009). On the effects of anisotropic rheology on ice flow, internal structure, and the age-depth relationship at ice divides. Journal of Geophysical Research, 114(F4), F04001, doi: 10.1029/2008JF001204.Google Scholar
Massonnet, D., and Feigl, K. L. (1998). Radar interferometry and its application to changes in the Earth’s surface. Reviews of Geophysics, 36(4), 441500.Google Scholar
Matsuda, M., and Wakahama, G. (1978). Crystallographic structure of polycrystalline ice. Journal of Glaciology, 21(85), 607620.Google Scholar
Matthews, J. B. (1981). The seasonal circulation of Glacier Bay, Alaska fjord system. Estuarine, Coastal, and Shelf Science, 12, 679700.Google Scholar
Matthews, W. H. (1974). Surface profiles of the Laurentide ice sheet in its marginal areas. Journal of Glaciology, 13(67), 3743.Google Scholar
Maxwell, K. D. (2002). Pacific decadal oscillation and Arizona precipitation. Available online from: http://www.wrh.noaa.gov/wrhq/02TAs/0208/.Google Scholar
Mayewski, P., et al. (1994). Changes in atmospheric circulation and ocean ice cover over the North Atlantic during the last 41,000 years. Science, 263, 17471751.Google Scholar
Mayewski, P. A., Sneed, S. B., Birkel, S. D., Kurbatov, A. V., and Maasch, K. A. (2014). Holocene warming marked by abrupt onset of longer summers and reduced storm frequency around Greenland. Journal of Quaternary Science, 29(1), 99104, doi: 10.1002/jqs.2684.Google Scholar
McConnell, J. R., Lamorey, G. W., Lambert, S. W., and Taylor, K. C. (2002). Continuous ice-core chemical analyses using inductively coupled plasma mass spectrometry. Environmental Science and Technology, 36, 711.Google Scholar
Meese, D. A., et al. (1997). The Greenland Ice Sheet Project 2 depth-age scale: methods and results. Journal of Geophysical Research, 102(C12), 2641126423.Google Scholar
Meier, M. F. (1961). Mass budget of South Cascade Glacier. 1957–1960. U.S. Geological Survey Professional Paper, 424-B, 206211.Google Scholar
Meier, M. F. (1965). Glaciers and climate. In Wright, H. E. Jr., and Frey, D. G. (eds.) The Quaternary of the United States. Princeton, NJ: Princeton University Press, 795805.Google Scholar
Meier, M. F., et al. (1994). Mechanical and hydrologic basis for the rapid motion of a large tidewater glacier: 1. Observations. Journal of Geophysical Research, 99(B8), 1521915229.Google Scholar
Meier, M., et al. (2007). Glaciers dominate eustatic sea‐level rise in the 21st century. Science, 317, 10641067.Google Scholar
Meier, M. F., and Post, A. (1987). Fast tidewater glaciers. Journal of Geophysical Research, 92(B9), 90519058.Google Scholar
Meier, M. F., Rasmussen, L. A., Krimmel, R. M., Olsen, R. W., and Frank, D. (1985). Photogrammetric determination of surface altitude, terminus position, and ice velocity of Columbia Glacier, Alaska. U.S. Geological Survey Professional Paper, 1258-F, F1F40.Google Scholar
Meierbachtol, T., Harper, J., and Humphrey, N. (2013). Basal drainage system response to increasing surface melt on the Greenland Ice Sheet. Science, 341, 777779.Google Scholar
Mellor, M., and Testa, R. (1969). Effect of temperature on the creep of ice. Journal of Glaciology, 8(52), 131145.Google Scholar
Menzies, J., and Shilts, W. W. (1996). Subglacial environments. In Menzies, J. (ed.) Past Glacial Environments–Sediments, Forms, and Techniques. Glacial Environments, Vol. 2. Oxford: Butterworth-Heinemann, pp. 15136.Google Scholar
Meyer, C. R., Fernandes, M. C., Creyts, T. T., and Rice, J. R. (2016). Effects of ice deformation on Röthlisberger channels and implications for transitions in subglacial hydrology. Journal of Glaciology, 62(234), 750762.Google Scholar
Mickelson, D. M. (1987). Central Lowlands. In Graf, W. L. (ed.) Geomorphic Systems of North America. Boulder, CO: Geological Society of America, Centennial Special Volume 2, pp. 111118.Google Scholar
Mikesell, T. D., van Wijk, K., Otheim, L. T., Marshall, H.-P., and Kurbatov, A. (2017). Laser ultrasound observations of mechanical property variations in ice cores. Geosciences, 7(3), 47, doi: 10.3390/geosciences7030047.Google Scholar
Miller, J. K. (1984). Model for Clastic Indicator Trains in Till: Prospecting in Areas of Glaciated Terrain. London: The Institution of Mining and Metallurgy, pp. 6977.Google Scholar
Minchew, B. M., Meyer, C. R., Robel, A. A., Gudmundsson, G. H., and Simons, M. (2018). Processes controlling the downstream evolution of ice rheology in glacier shear margins: case study on Rutford Ice Stream, West Antarctica. Journal of Glaciology, 64(246), 583594.Google Scholar
Minchew, B. M., Simons, M., Riel, B., and Milillo, P. (2017). Tidally induced variations in vertical and horizontal motion on Rutford Ice Stream, West Antarctica, inferred from remotely sensed observations. Journal of Geophysical Research: Earth Surface, 122(1), 167190.Google Scholar
Mingle, J. (2015). Fire and Ice: Soot, Solidarity, and Survival on the Roof of the World. New York: St. Martin’s Press, 450pp.Google Scholar
Minzoni, R. T., Anderson, J. B., Wellner, J. S., and Fernandez, R. (2017). Antarctic Peninsula glacier synchroneity increased during the Holocene: implications for modern recession. Geological Society of America Annual Meeting Abstracts with Programs, Paper 51-14.Google Scholar
Mitchell, J. K. (1993). Fundamentals of Soil Behavior, 2nd ed. New York: John Wiley.Google Scholar
Mitchell, J. K., Campanella, R. G., and Singh, A. (1968). Soil creep as a rate process. Journal of the Soil Mechanics and Foundations Division, American Society of Civil Engineers, 94(SM1), 231253.Google Scholar
Mitchell, J. K., and Soga, K. (2005). Fundamentals of Soil Behavior, 3rd ed. Hoboken, NJ: John Wiley, 577pp.Google Scholar
Montagnat, M., and Duval, P. (2000). Rate controlling processes in the creep of polar ice, influence of grain boundary migration associated with recrystallization. Earth and Planetary Science Letters, 183, 179186.Google Scholar
Montagnat, M., and Duval, P. (2004). Dislocations in ice and deformation mechanisms: from single crystals to polar ice. Defect and Diffusion Forum, 229, 4354.Google Scholar
Mooers, H. D. (1989). On the formation of tunnel valleys of the Superior Lobe, Central Minnesota. Quaternary Research, 32, 2435.Google Scholar
Mooers, H. D. (1990). Ice marginal thrusting of drift and bedrock: thermal regime subglacial aquifers, and glacial surges. Canadian Journal of Earth Sciences, 27(6), 849862.Google Scholar
Moor, E., and Stauffer, B. (1984). A new dry extraction system for gasses in ice. Journal of Glaciology, 30(106), 358361.Google Scholar
Moore, P. L. (2014). Deformation of debris-ice mixtures. Reviews of Geophysics, 52, 435467.Google Scholar
Moore, P. L., Iverson, N. R., and Cohen, D. (2010). Conditions for thrust faulting in a glacier. Journal of Geophysical Research, 115, 15, doi: 10.1029/2009JF001307. F02005.Google Scholar
More, A. F., et al. (2017). Next-generation ice core technology reveals true minimum natural levels of lead (Pb) in the atmosphere: insights from the Black Death. GeoHealth, 1(4), 211219.Google Scholar
Morgan, V. I. (1972). Oxygen isotope evidence for bottom freezing on the Amery Ice Shelf. Journal of Glaciology, 24(90), 393394.Google Scholar
Motyka, R. J., Dryer, W. P., Amundson, J., Truffer, M., and Fahnestock, M. (2013). Rapid submarine melting driven by subglacial discharge, LeConte Glacier, Alaska, U.S.A. Geophysical Research Letters, 40, 51535158, doi: 10.1002/grl.51011.Google Scholar
Motyka, R. J., Hunter, L., Echelmeyer, K. A., and Connor, K. (2003). Submarine melting at the terminus of a temperate tidewater glacier, LeConte Glacier, Alaska, U.S.A. Annals of Glaciology, 36, 5765.Google Scholar
Moran, S. R., Clayton, L., Hooke, R. LeB., Fenton, M. M., and Andriashek, L. D. (1980). Glacier bed landforms of the prairie region of North America. Journal of Glaciology, 25(93), 457476.Google Scholar
Morse, D. L., Waddington, E. D., and Steig, E. J. (1998). Ice age storm trajectories from radar stratigraphy at Taylor Dome, Antarctica. Geophysical Research Letters, 25(17), 33833386.Google Scholar
Müller, F. (1962). Zonation in the accumulation area of the glaciers of Axel Heiberg Island, N.W.T., Canada. Journal of Glaciology, 4(33), 302318.Google Scholar
Murozumi, M., Chow, T. J., and Patterson, C. (1969). Chemical concentrations of pollutant lead aerosols, terrestrial dusts and sea salts in Greenland and Antarctic snow strata. Geochimica et Cosmochimica Acta, 33, 12471294.Google Scholar
Nadai, A. (1950). Theory of Flow and Fracture of Solids, volume 1, 2nd ed. New York: McGraw Hill, 572pp.Google Scholar
Nakase, A., and Kamei, T. (1986). Influence of strain rate on undrained shear strength characteristics of Ko-consolidated cohesive soils. Soils and Foundations, 26, 8595.Google Scholar
Nakawo, M., and Rana, B. (1999). Estimation of ablation rate of glacier ice under a supraglacial debris layer. Geografiska Annaler, 81A(4), 695701.Google Scholar
Neftel, A., Oeschger, H., and Stauffer, B. (1988). CO2 record in the Byrd ice core 50,000-5,000 years BP. Nature, 331, 609611.Google Scholar
Nereson, N. A., Raymond, C. F., Waddington, E. D., and Jacobel, R. W. (1998). Migration of the Siple Dome ice divide, West Antarctica. Journal of Glaciology, 44(148), 643652.Google Scholar
Ng, F. S. L. (1999). A mathematical model of wide subglacial water drainage channels. In Wettlaufer, J. S., Dash, J. G., and Untersteiner, N. (eds.) Ice Physics and the Natural Environment. Berlin: Springer-Verlag, pp. 325327 (NATO ASI Series I: Global Environmental Change 56).Google Scholar
Ng, F. S. L. (2000a). Canals under sediment-based ice sheets. Annals of Glaciology, 30, 146152.Google Scholar
Ng, F. S. L. (2000b). Coupled ice-till deformation near subglacial channels and cavities. Journal of Glaciology, 46(155), 580598.Google Scholar
Ng, F., and Hallet, B. (2002). Patterning mechanisms in subglacial carbonate dissolution and deposition. Journal of Glaciology, 48(162), 386400.Google Scholar
NGRIP members. (2004). High-resolution record of Northern Hemisphere climate extending into the last interglacial period. Nature, 431, 147.Google Scholar
Nguyen, V. B., Darnige, T., Bruand, A., and Clement, E. (2011). Creep and fluidity of a real granular packing near jamming. Physical Review Letters, 107, 5.Google Scholar
Nicholls, K. W., Corr, H. F. J., Stewart, C. L., Lok, L. B., Brennan, P. V., and Vaughan, D. G. (2015). A ground-based radar for measuring vertical strain rates and time-varying basal melt rates in ice sheets and shelves. Journal of Glaciology, 61(230), 10791087, doi: 10.3189/2015JoG15J073.Google Scholar
Nick, F. M., van der Veen, C. J., Vieli, A., and Benn, D. I. (2010). A physically based calving model applied to marine outlet glaciers and implications for the glacier dynamics. Journal of Glaciology, 56(199), 781794.Google Scholar
NSIDC. (2017). National Snow and Ice Data Center. Available online from: https://nsidc.org/cryosphere/seaice/processes/albedo.html [Accessed 15 August 2017].Google Scholar
Nygård, A., Sejrup, H. P., Haflidason, H., Lekens, W. A. H., Clark, C. D., and Bigg, G. R. (2007). Extreme sediment and ice discharge from marine-based ice streams: new evidence from the North Sea. Geology, 35(5), 395398.Google Scholar
Nye, J. F. (1951). The flow of glaciers and ice sheets as a problem in plasticity. Proceedings of the Royal Society of London, Series A, 207(1091), 554572.Google Scholar
Nye, J. F. (1952a). Reply to Mr. Joel E. Fisher’s comments. Journal of Glaciology, 2(11), 5253.Google Scholar
Nye, J. F. (1952b). Mechanics of glacier flow. Journal of Glaciology, 2(12), 8293.Google Scholar
Nye, J. F. (1953). The flow law of ice from measurements in glacier tunnels, laboratory experiments, and the Jungfraufirn borehole experiment. Proceedings of the Royal Society of London, Series A, 219(1139), 477489.Google Scholar
Nye, J. F. (1955). Comments on Dr. Lowe’s letter and notes on crevasses. Journal of Glaciology, 2(17), 512514.Google Scholar
Nye, J. F. (1957). The distribution of stress and velocity in glaciers and ice sheets. Proceedings of the Royal Society of London, Series A, 239(1216), 113133.Google Scholar
Nye, J. F. (1960). The response of glaciers and ice sheets to seasonal and climatic changes. Proceedings of the Royal Society, Series A, 256(1287), 559584.Google Scholar
Nye, J. F. (1963). On the theory of the advance and retreat of glaciers. Geophysical Journal of the Royal Astronomical Society, 7(4), 432456.Google Scholar
Nye, J. F. (1965). The flow of a glacier in a channel of rectangular, elliptic, or parabolic cross section. Journal of Glaciology, 5(41), 661690.Google Scholar
Nye, J. F. (1969a). The calculation of sliding of ice over a wavy surface using a Newtonian viscous approximation. Proceedings of the Royal Society of London, Series A, 311(1506), 445467.Google Scholar
Nye, J. F. (1969b). The effect of longitudinal stress on the shear stress at the base of an ice sheet. Journal of Glaciology, 8(53), 207213.Google Scholar
Nye, J. F. (1970). Glacier sliding without cavitation in a linear viscous approximation. Proceedings of the Royal Society of London, Series A, 315(1522), 381403.Google Scholar
Nye, J. F. (1973a). The motion of ice past obstacles. In Whalley, E., Jones, S. J., and Gold, L. W. (eds.) The Physics and Chemistry of Ice. Ottawa: Royal Society of Canada, pp. 387394.Google Scholar
Nye, J. F. (1973b). Water at the bed of a glacier. IUGG-AIHS Symposium on the Hydrology of Glaciers, Cambridge, September 7–13, 1969. International Association of Scientific Hydrology Publication 95, 189–194.Google Scholar
Nye, J. F. (1976). Water flow in glaciers: Jökulhlaups, tunnels and veins. Journal of Glaciology, 17(76), 181207.Google Scholar
Nye, J. F., and Frank, F. C. (1973). Hydrology of intergranular veins in a temperate glacier. IUGG-AIHS Symposium on the Hydrology of Glaciers, Cambridge, 7–13 September 1969. International Association of Scientific Hydrology Publication, 95, 157161.Google Scholar
Nye, J. F., and Mae, S. (1972). The effect of non-hydrostatic stress on intergranular water veins and lenses in ice. Journal of Glaciology, 11(61), 81101.Google Scholar
O’Leary, M. O., and Christoffersen, P. (2013). Calving on tidewater glaciers amplified by submarine frontal melting. The Cryosphere, 7, 119128.Google Scholar
O’Leary, M. O., and Luckman, A. (2017). Larsen C calves trillion ton iceberg. Available online from: http://www.projectmidas.org/blog/calving/.Google Scholar
Osterberg, E. C., et al. (2017). The 1200 year composite ice core record of Aleutian Low intensification. Geophysical Research Letters, 44, 74477454.Google Scholar
Osterberg, E. C., Handley, M. J., Sneed, S. B., Mayewski, P. A., and Kreutz, K. J. (2006). Continuous sampling ice core melter system with discrete sampling for major ion, trace element, and stable isotope analyses. Environmental Science and Technology, 40, 33553361.Google Scholar
Östrem, G. (1959). Ice melting under a thin layer of moraine, and the existence of ice cores in moraine ridges. Geografiska Annaler, 41(4), 228230.Google Scholar
Padman, L., Siegfried, M. R., and Fricker, H. A. (2018). Ocean tide influences on the Antarctic and Greenland ice sheets. Reviews of Geophysics, 56, 142184, doi.org/10.1002/2016RG000546.Google Scholar
Painter, T. H., Flanner, M. D., Kaser, G., Marzeion, B., VanCuren, R. A., and Abdalati, W. (2013). End of the Little Ice Age in the Alps forced by industrial black carbon. PNAS, 110(38), 1521615221.Google Scholar
Park, J. W., Gourmelen, N., Shepherd, A., Kim, S. W., Vaughan, D. G., and Wingham, D. J. (2013). Sustained retreat of the Pine Island Glacier. Geophysical Research Letters, 40, 21372142.Google Scholar
Parker, G. (1979). Hydraulic geometry of active gravel rivers. Journal of the Hydraulics Division, American Society of Civil Engineers, 105(HY9), 11851201.Google Scholar
Paterson, W. S. B. (1971). Temperature measurements in Athabasca Glacier, Alberta, Canada. Journal of Glaciology, 10(60), 339349.Google Scholar
Paterson, W. S. B. (1977). Secondary and tertiary creep of glacier ice as measured by borehole closure rates. Reviews of Geophysics and Space Physics, 15(1), 4755.Google Scholar
Paterson, W. S. B. (1991). Why ice-age ice is sometimes “soft”. Cold Regions Science and Technology, 20, 7598.Google Scholar
Paterson, W. S. B., et al. (1977). An oxygen isotope climatic record from the Devon Island ice cap. Arctic Canada. Nature, 266(5602), 508511.Google Scholar
Patterson, C. J. (1994). Tunnel-valley fans of the St. Croix moraine, east-central Minnesota, USA. In Warren, W., and Croot, C. (eds.) Formation and Deformation of Glacial Deposits. Rotterdam: Balkema, pp. 6987.Google Scholar
Patterson, C. J. (1997). Southern Laurentide ice lobes were created by ice streams: Des Moines lobe in Minnesota, USA. Sedimentary Geology, 111, 247261.Google Scholar
Patterson, C. J., and Hooke, R. LeB. (1995). Physical environment of drumlin formation. Journal of Glaciology, 41(137), 3038.Google Scholar
Pattyn, F., et al. (2008). Benchmark experiments for higher-order and full-Stokes ice sheet models (ISMIP–HOM). The Cryosphere, 2, 95108.Google Scholar
Payne, A. J., et al. (2000). Results from the EISMINT model intercomparison: the effects of thermomechanical coupling. Journal of Glaciology, 46(153), 227238.Google Scholar
Peiró, J., and Sherwin, S. (2005). Finite difference, finite element, and finite volume methods for partial differential equations. In Yip, S. (ed.) Handbook of Materials Modeling, Volume 1: Methods and Models. The Netherlands: Springer, pp. 132.Google Scholar
Pelto, M. S., and Warren, C. R. (1991). Relationship between tidewater glacier calving velocity and water depth at the calving front. Annals of Glaciology, 15, 115118.Google Scholar
Perol, T., Rice, J. R., Platt, J. D., and Suckale, J. (2015). Subglacial hydrology and ice stream margin locations. Journal of Geophysical Research Earth Surface, 120, 13521368, doi: 10.1002/2015JF003542.Google Scholar
Petit, J. R., et al. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399, 429436.Google Scholar
Pfeffer, W. T. (1992). Stress-induced foliation in the terminus of Variegated Glacier, Alaska, U.S.A., formed during the 1982–83 surge. Journal of Glaciology, 38(129), 213222.Google Scholar
Philberth, K., and Federer, B. (1971). On the temperature profile and age profile in the central part of cold ice sheets. Journal of Glaciology, 10(58), 314.Google Scholar
Philip, J. R. (1980). Thermal fields during regelation. Cold Regions Science and Technology, 3, 193203.Google Scholar
Pimienta, P., and Duval, P. (1987). Rate controlling processes in the creep of polar glacier ice. Journal de Physique, 48, Colloque C1, Supplement to no. 3, C1-243C1-248.Google Scholar
Pimienta, P., Duval, P., and Lipenkov, V. Ya. (1988). Mechanical behavior of ice along the 2040 m Vostok core, Antarctica. Annals of Glaciology, 10, 137140.Google Scholar
Piotrowski, J. A., Mickelson, D. M., Tulaczyk, S., Krzyszkowski, D., and Junge, F. W. (2001). Were deforming subglacial beds beneath past ice sheets really widespread? Quaternary International, 86, 139150.Google Scholar
Pohjola, V. A., and Rogers, J. C. (1997). Atmospheric circulation and variations in Scandinavian glacier mass balance. Quaternary Research, 47, 2936, doi: 10.1006/qres.1996.1859.Google Scholar
Pralong, A., and Funk, M. (2005). Dynamic damage model of crevasse opening and application to glacier calving. Journal of Geophysical Research, 110, B01309, doi: 10.1029/2004JB003104.Google Scholar
Price, S. F., Conway, H., and Waddington, E. D. (2007a). Evidence for late Pleistocene thinning of Siple Dome, West Antarctica. Journal of Geophysical Research, 112, F03021, doi: 10.1029/2006JF000725.Google Scholar
Price, S. F., Waddington, E. D., and Conway, H. (2007b). A full-stress, thermomechanical flow band model using the finite volume method. Journal of Geophysical Research, 112, F03020, doi: 10.1029/2006JF000724.Google Scholar
Prior, D. J., et al. (2015). Making EBSD on water ice routine. Journal of Microscopy, 259(3), 237256.Google Scholar
Radko, T. (2013). Double-Diffusive Convention. Cambridge: Cambridge University Press, 342pp.Google Scholar
Raisbeck, G. M., et al. (2017). An improved north-south synchronization of ice core records around the 41 kyr 10Be peak. Climate of the Past, 13, 217229.Google Scholar
Raisbeck, G. M., Yiou, F., Jouzel, J., and Stocker, T. F. (2007). Direct north-south synchronization of abrupt climate change record in ice cores using Beryllium 10. Climate of the Past, 3, 541547.Google Scholar
Ram, M., Illing, M., Weber, P., Koenig, G., and Kaplan, M. (1995). Polar ice stratigraphy from laser-light scattering: scattering from ice. Geophysical Research Letters, 22(24), 35253527.Google Scholar
Ramanathan, V., and Carmichael, G. (2008). Global and regional climate change due to black carbon. Nature Geoscience, 1, 221227.Google Scholar
Ramsay, J. G., and Graham, R. H. (1970). Strain variation in shear belts. Canadian Journal of Earth Sciences, 7, 786813.Google Scholar
Ramseier, R. O. (1967). Self-diffusion of tritium in natural and synthetic ice monocrystals. Journal of Applied Physics, 36(6), 25532556.Google Scholar
Rasmussen, E. M. (1984). El Niño: the ocean/atmosphere connection. Oceanus, 27(2), 512.Google Scholar
Rasmussen, L. A., and Meier, M. F. (1982). Continuity equation model of the predicted drastic retreat of Columbia Glacier, Alaska. U.S. Geological Survey Professional Paper, 1258-F, A1A23.Google Scholar
Ratcliffe, E. H. (1962). Thermal conductivity of ice. Philosophical Magazine, 7, 11971203.Google Scholar
Rathbun, A. P., Marone, C., Alley, R. B., and Anandakrishnan, S. (2008). Laboratory study of the frictional rheology of sheared till. Journal of Geophysical Research, 113(F2), F02020, doi: 10.1029/2007JF000815.Google Scholar
Raymond, C. F. (1971). Flow in a transverse section of Athabasca Glacier, Alberta, Canada. Journal of Glaciology, 10(58), 5584.Google Scholar
Raymond, C. F. (1973). Inversion of flow measurements for stress and rheological parameters in a valley glacier. Journal of Glaciology, 12(64), 1944.Google Scholar
Raymond, C. F. (1983). Deformation in the vicinity of ice divides. Journal of Glaciology, 29(103), 357373.Google Scholar
Raymond, C. F. (2000). Energy balance of ice streams. Journal of Glaciology, 46(155), 665674.Google Scholar
Raymond, C. F., Echelmeyer, K. A., Whillans, I. M., and Doake, C. S. M. (2001). Ice stream shear margins in the West Antarctic Ice Sheet: behavior and environment. Antarctic Research Series, 77, 137155.Google Scholar
Raymond, C. F., and Harrison, W. D. (1975). Some observations on the behavior of liquid and gas phases in temperate glacier ice. Journal of Glaciology, 14(71), 213234.Google Scholar
Raymond, C. F., and Harrison, W. D. (1988). Evolution of Variegated Glacier, U.S.A., prior to its surge. Journal of Glaciology, 34(117), 154165.Google Scholar
Raynaud, D. (2012). The integrity of the ice record of greenhouse gases with special focus on atmospheric CO2. Лёд и Снег (Ice and Snow), 118(2), 514.Google Scholar
Raynaud, D., Jouzel, J., Barnola, J.-M., Chappellaz, J., Delmas, R. J., and Lorius, C. (1993). The ice core record of greenhouse gases. Science, 259(5097), 926934.Google Scholar
Raynaud, D., and Whillans, I. M. (1982). Air content of the Byrd core and past changes in the West Antarctic Ice Sheet. Annals of Glaciology, 3, 269273.Google Scholar
Reeh, N. (1968). On the calving of ice from floating glaciers and ice shelves. Journal of Glaciology, 7(50), 215232.Google Scholar
Rempel, A. W. (2008). A theory for ice–till interactions and sediment entrainment beneath glaciers. Journal of Geophysical Research, Earth Surface, 113(F1), doi: 10.1029/2007JF000870.Google Scholar
Retzlaff, R., and Bentley, C. R. (1993). Timing of stagnation of Ice Streams A, West Antarctica, from short-pulse radar studies of buried surface crevasses. Journal of Glaciology, 39(133), 553561.Google Scholar
Retzlaff, R., Lord, N., and Bentley, C. R. (1993). Airborne-radar studies: ice streams A, B and C, West Antarctica. Journal of Glaciology, 39(133), 495506.Google Scholar
Rignot, E. (1998). Fast recession of a West Antarctic glacier. Science, 281, 549551.Google Scholar
Rignot, E. (2001). Evidence for rapid retreat and mass loss of Thwaites Glacier, West Antarctica. Journal of Glaciology, 47(157), 213222.Google Scholar
Rignot, E. (2008). Changes in West Antarctic ice stream dynamics observed with ALOS PALSAR data. Geophysical Research Letters, 35, L12505, doi: 10.1029/2008GL033365.Google Scholar
Rignot, E., Fenty, I., Xu, Y., Cai, C., and Kemp, C. (2015). Undercutting of marine-terminating glaciers in West Greenland. Geophysical Research Letters, 42, 59095917.Google Scholar
Rignot, E., Mouginot, J., Morlighem, M., Seroussi, H., and Scheuchl, B. (2014). Widespread, rapid grounding line retreat of Pine Island, Thwaites, Smith, and Kohler glaciers, West Antarctica, from 1992 to 2011. Geophysical Research Letters, 41, 35023509.Google Scholar
Rignot, E., Velicogna, I., van den Broeke, M. R., Monaghan, A., and Lenaerts, J. T. M. (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
Rist, M. A., Sammonds, P. R., Murrell, S. A. F., Meredith, P. G., Doake, C. S. M., Oerter, H., and Matsuki, K. (1999). Experimental and theoretical fracture mechanics applied to Antarctic ice and surface crevassing. Journal of Geophysical Research, 104(B2), 29732987.Google Scholar
Robin, G. deQ. (1955). Ice movement and temperature distribution in glaciers and ice sheets. Journal of Glaciology, 2(18), 523532.Google Scholar
Robin, G. DeQ. (1976). Is the basal ice of a temperate glacier at the pressure melting point? Journal of Glaciology, 16(74), 183195.Google Scholar
Robin, G. DeQ. (1979). Formation, flow, and disintegration of ice shelves. Journal of Glaciology, 24(90), 259271.Google Scholar
Roe, G. H., and Baker, M. B. (2014). Glacier response to climate perturbations: an accurate linear geometric model. Journal of Glaciology, 60(222), 670684.Google Scholar
Roe, G. H., and O’Neal, M. A. (2009). The response of glaciers to intrinsic climate variability: observations and models of late-Holocene variations in the Pacific Northwest. Journal of Glaciology, 55(193), 839854.Google Scholar
Römkens, M. J. M., and Miller, R. D. (1973). Migration of mineral particles in ice with a temperature gradient. Journal of Colloidal and Interface Science, 42(1), 103111.Google Scholar
Rose, J. (1989). Glacier stress patterns and sediment transfer associated with the formation of superimposed flutes. Sedimentary Geology, 62, 151176.Google Scholar
Rosenberg, R. (2005). Why is ice slippery? Physics Today, 58(12), 5055.Google Scholar
Rosier, S. H. R., Gudmundsson, G. H., and Green, J. A. M. (2015). Temporal variations in the flow of a large Antarctic ice stream controlled by tidally induced changes in the subglacial water system. The Cryosphere, 9, 16491661.Google Scholar
Röthlisberger, H. (1972). Water pressure in intra- and subglacial channels. Journal of Glaciology, 11(62), 177204.Google Scholar
Röthlisberger, H. (1974). Möglichkeiten und grenzen der gletscherüberwachung (Possibilities and limits of glacier surveillance). Neuen Zürcher Zeitung, 196, 15p.Google Scholar
Röthlisberger, H., and Iken, A. (1981). Plucking as an effect of water-pressure variations at the glacier bed. Annals of Glaciology, 2, 57-62.Google Scholar
Rott, H., Rack, W., Skvarca, P., and De Angelis, H., 2002. Northern Larsen Ice Shelf, Antarctica: further retreat after collapse. Annals of Glaciology, 34, 277282.Google Scholar
Ruddiman, W. F. (2003). The anthropogenic greenhouse era began thousands of years ago. Climate Change, 61, 261293.Google Scholar
Ruddiman, W. F. (2007). The early anthropogenic hypothesis: challenges and responses. Reviews of Geophysics, 45, RG4001, doi: 10.1029/2006RG000207.Google Scholar
Russell-Head, D. S., and Budd, W. F. (1979). Ice-sheet flow properties derived from bore-hole shear measurements combined with ice-core studies. Journal of Glaciology, 24(90), 117130.Google Scholar
Sammis, C. G., King, G., and Biegel, R. (1987). The kinematics of gouge deformation. Pure and Applied Geophysics, 125(5), 777812.Google Scholar
Sane, S. M., Desai, C. S., Jenson, J. W., Contractor, D. N., Carlson, A. E., and Clark, P. U. (2008). Disturbed state constitutive modeling of two Pleistocene tills. Quaternary Science Reviews, 27, 267283.Google Scholar
Scambos, T. A., Bohlander, J. A., Shuman, C. A., and Skvarca, P. (2004). Glacier acceleration and thinning after ice shelf collapse in the Larsen B embayment, Antarctica. Geophysical Research Letters, 31, L18402.Google Scholar
Scambos, T. A., Hulbe, C., Fahnestock, M., and Bohlander, J. (2000). The link between climate warming and break-up of ice shelves in the Antarctic Peninsula. Journal of Glaciology, 46(154), 516530.Google Scholar
Scambos, T. A., Hulbe, C., Fahnestock, M. A., and Bohlander, J. (2003). The link between climate warming and break-up of ice shelves in the Antarctic Peninsula. Journal of Glaciology, 46(154), 516530.Google Scholar
Schellenberger, T., Dunse, T., Kääb, A., Schuler, T. V., Hagen, J. O., and Reijmer, C. H. (2017). Multi-year surface velocities and sea-level rise contribution of the Basin-3 and Basin-2 surges, Austfonna, Svalbard. The Cryosphere, doi: 10.5194/tc-2017-5, 2017.Google Scholar
Schmidt, R., and Housen, K. (1995). Problem solving with dimensional analysis. The Industrial Physicist, 1, 2124.Google Scholar
Schneider, N., and Cornuelle, B. D. (2005). The forcing of the Pacific Decadal Oscillation. Journal of Climate, 18, 43554373.Google Scholar
Schodlok, M. P., Menemenlis, D., Rignot, E., and Studinger, M. (2012). Sensitivity of the ice-shelf/ocean system to the sub-ice-shelf cavity shape measured by NASA IceBridge in Pine Island Glacier, West Antarctica. Annals of Glaciology, 53(60), 156162.Google Scholar
Schoof, C. G. (2012). Thermally-driven migration of ice stream margins. Journal of Fluid Mechanics, 712, 552578.Google Scholar
Schoof, C. G., and Clarke, G. K. C. (2008). A model for spiral flows in basal ice and the formation of subglacial flutes based on a Reiner–Rivlin rheology for glacial ice. Journal of Geophysical Research, Earth Surface, 113(B5), doi: 10.1029/2007JB004957.Google Scholar
Schulson, E. M. (1999). The structure and mechanical behavior of ice. Journal of the Minerals Metals & Materials Society, 52(1), 2127.Google Scholar
Scott, J. B. T., Gudmundsson, G. H., Smith, A. M., Bingham, R. G., Pritchard, H. D., and Vaughan, D. G. (2009). Increased rate of acceleration on Pine Island Glacier strongly coupled to changes in gravitational driving stress. The Cryosphere, 2, 125131.Google Scholar
Seaberg, S. Z., Seaberg, J. Z., Hooke, R. LeB., and Wiberg, D. W. (1988). Character of the englacial and subglacial drainage system in the lower part of the ablation area of Storglaciären, Sweden, as revealed by dye-trace studies. Journal of Glaciology, 34(117), 217227.Google Scholar
Segall, P. (1984). Rate‑dependent extensional deformation resulting from crack growth in rock. Journal of Geophysical Research, 89(B6), 41854195.Google Scholar
Shabtaie, S., and Bentley, C. R. (1988). Ice-thickness map of the West Antarctic ice streams by radar sounding. Annals of Glaciology, 11, 126136.Google Scholar
Sharp, M. (1982). Modification of clasts in lodgement tills by glacial erosion. Journal of Glaciology, 28(100), 475481.Google Scholar
Shaw, J., and Freschauf, R. C. (1973). A kinematic discussion of the formation of glacial flutings. Canadian Geographer, 17(1), 1935.Google Scholar
Shepherd, A., et al. (2012). A reconciled estimate of ice-sheet mass balance. Science, 338(6111), 11831189.Google Scholar
Shreve, R. L. (1967). Migration of air bubbles, vapor figures, and brine pockets in ice under a temperature gradient. Journal of Geophysical Research, 72(16), 40934100.Google Scholar
Shreve, R. L. (1972). Movement of water in glaciers. Journal of Glaciology, 11(62), 205214.Google Scholar
Shreve, R. L. (1984). Glacier sliding at subfreezing temperatures. Journal of Glaciology, 30(106), 341347.Google Scholar
Shreve, R. L. (1985a). Esker characteristics in terms of glacier physics, Katahdin Esker system, Maine. Geological Society of America Bulletin, 96(5), 639646.Google Scholar
Shreve, R. L. (1985b). Late Wisconsin ice-surface profile calculated from esker paths and types, Katahdin esker system, Maine. Quaternary Research, 23(1), 2737.Google Scholar
Shreve, R. L., and Sharp, R. P. (1970). Internal deformation and thermal anomalies in lower Blue Glacier, Mount Olympus, Washington, USA. Journal of Glaciology, 9(55), 6586.Google Scholar
Shumskii, P. A. (1964). Principles of Structural Glaciology. New York: Dover.Google Scholar
Shumway, J. R., and Iverson, N. R. (2009). Magnetic fabrics of the Douglas Till of the Superior lobe: exploring bed-deformation kinematic. Quaternary Science Reviews, 28, 107119.Google Scholar
Siegenthaler, U., and Oeschger, H. (1987). Biospheric CO2 emissions during the past 200 years reconstructed by deconvolution of ice core data. Tellus B: Chemical and Physical Meteorology, 39(1–2), 140154.Google Scholar
Sigl, M., Abram, N. J., Gabrieli, J., Jenk, T. M., Osmont, D., and Schwikowski, M. (2018). 19th century glacier retreat in the Alps preceded the emergence of industrial black carbon deposition on high-alpine glaciers. The Cryosphere, 12, 33113331.Google Scholar
Sih, G. C. (1973). Handbook of Stress-Intensity Factors; Stress-Intensity Factor Solutions and Formulas for Reference. Bethlehem, PA: Institute of Fracture and Solid Mechanics, Leigh University.Google Scholar
Siman-Tov, S., Stock, G. M., Brodsky, E. E., and White, J. C. (2017). The coating layer of glacial polish. Geology, 45(11), 987990.Google Scholar
Smith, A. M., Murray, T., Nicholls, K. W., Makinson, K., Aðalgeirsdóttir, G., Behar, A. E., and Vaughan, D. G. (2007). Rapid erosion, drumlin formation, and changing hydrology beneath an Antarctic ice stream. Geology, 35(2), 127130.Google Scholar
Sneed, S. B., et al. (2015). New LA-ICP-MS cryocell and calibration technique for sub-millimeter analysis of ice cores. Journal of Glaciology, 61(226), 233242.Google Scholar
Sneed, W. A., Hooke, R. LeB., and Hamilton, G. S. (2008). Thinning of the south dome of Barnes Ice Cap, arctic Canada, over the past two decades. Geology, 36(1), 7174.Google Scholar
Sokolnikoff, I. S., and Redheffer, R. M. (1958). Mathematics of Physics and Modern Engineering. New York: McGraw Hill, 810pp.Google Scholar
Sommerfeld, R., and LaChapelle, E. (1970). The classification of snow metamorphism. Journal of Glaciology, 9(55), 317.Google Scholar
Souchez, R. A., and Lorrain, R. D. (1978). Origin of the basal ice layer from Alpine glaciers indicated by its chemistry, Journal of Glaciology, 20(83), 319328.Google Scholar
Spencer, M. K. (2005). Paleoclimatic change inferred through firnification processes. PhD thesis, The Pennsylvania State University, 196pp.Google Scholar
Spencer, M. K., Alley, R. B., and Fitzpatrick, J. J. (2006). Developing a bubble-number density paleoclimatic indicator for glacier ice. Journal of Glaciology, 52(178), 358364.Google Scholar
Steffensen, J. P., et al. (2008). High-resolution Greenland ice core data show abrupt climate change happens in a few years. Science, 321(5889), 680684.Google Scholar
Stehle, N. S. (1967). Migration of air bubbles in ice under a temperature gradient. In Öura, H. (ed.) Physics of Snow and Ice: International Conference on Low Temperature Science 1966 Proceedings. Sapporo: Institute of Low Temperature Science, Hokkaido University, 1(1), 219232.Google Scholar
Stephenson, S. N., and Doake, C. S. M. (1982). Dynamic behavior of Rutford Ice Stream. Annals of Glaciology, 3, 295299.Google Scholar
Stevens, L. A., et al. (2015). Greenland supraglacial lake drainages triggered by hydrologically induced basal slip. Nature, 522, 7376.Google Scholar
Straneo, F., Heimbach, P. (2013). North Atlantic warming and the retreat of Greenland’s outlet glaciers. Nature, 504, 3643.Google Scholar
Strang, G., and Fix, G. J. (1973). An Analysis of the Finite-Element Method. New York: Prentice Hall, 306pp.Google Scholar
Stocker, T. F., and Johnsen, S. J. (2003). A minimum thermodynamic model for the bipolar seesaw. Paleoceanography, 18(4).Google Scholar
Stokes, C. R., Fowler, A. C., Clark, C. D., Hindmarsh, R. C. A., and Spagnolo, M. (2013). The instability theory of drumlin formation and its explanation of their varied composition and internal structure. Quaternary Science Reviews, 62, 7796.Google Scholar
Stokes, C. R., Lian, O. B., Tulaczyk, S., and Clark, C. D. (2008). Superimposition of ribbed moraine on a palaeo-ice-stream bed: implications for ice stream dynamics and shutdown. Earth Surface Processes and Landforms, 33, 593609.Google Scholar
Stone, G. H. (1899). The glacial gravels of Maine and their associated deposits. U.S. Geological Survey Monograph 34, 499pp.Google Scholar
Suckale, J., Platt, J. D., Perol, T., and Rice, J. R. (2014). Deformation-induced melting in the margins of the West Antarctic ice streams. Journal of Geophysical Research: Earth Surface, 119, 10041025, doi: 10.1002/2013JF003008.Google Scholar
Sundal, A. V., Shepard, A., Nienow, P., Hanna, E., Palmer, S., and Huybrechts, P. (2011). Melt-induced speed-up of Greenland ice sheet offset by efficient subglacial drainage. Nature, 469, 521524.Google Scholar
Suwa, M., and Bender, M. L. (2008). Chronology of the Vostok ice core constrained by O2/N2 ratios of occluded air, and its implication for the Vostok climate records. Quaternary Science Reviews, 27, 10931106, doi: 10.1016/j.quascirev.2008.02.017.Google Scholar
Swithinbank, C. (1969). Giant icebergs in the Weddell Sea, 1967–68. Polar Record, 14(91), 477478.Google Scholar
Tarasov, L., and Peltier, W. R. (1999). Impact of thermomechanical ice sheet coupling on a model of the 100 kyr ice age cycle. Journal of Geophysical Research, 105(D4), 95179545.Google Scholar
Taylor, K. C., et al. (2004). Dating the Siple Dome (Antarctica) ice core by manual and computer interpretation of annual layering. Journal of Glaciology, 50(170), 453461.Google Scholar
Taylor, L. D. (1963). Structure and fabric on the Burroughs Glacier, south-east Alaska. Journal of Glaciology, 4(36), 731752.Google Scholar
Tedesco, M., Doherty, S., Fettweis, X., Alexander, P., Jeyaratnam, J., and Stroeve, J. (2016). The darkening of the Greenland ice sheet: trends, drivers, and projections (1981–2100). The Cryosphere, 10, 477496.Google Scholar
Terzaghi, K. (1950). Mechanics of landslides. In Application of Geology to Engineering Practice. New York: Geological Society of America, Berkey Volume, 83125.Google Scholar
Thomas, R. H. (1973). The creep of ice shelves: theory. Journal of Glaciology, 12(64), 4553.Google Scholar
Thomas, R. H., and Bentley, C. R. (1978). A model for Holocene retreat of the West Antarctic Ice Sheet. Quaternary Research, 10, 150170.Google Scholar
Thomason, J. F., and Iverson, N. R. (2008). A laboratory study of particle ploughing and pore-pressure feedback: a velocity-weakening mechanism for soft glacier beds. Journal of Glaciology, 54(184), 169181.Google Scholar
Thomason, J. F., and Iverson, N. R. (2009). Deformation of the Batestown till of the Lake Michigan lobe, Laurentide ice sheet. Journal of Glaciology, 55(189), 131146.Google Scholar
Thompson, L. G. (1977). Variations in microparticle concentration, size distribution and elemental composition found in Camp Century, Greenland, and Byrd station, Antarctica, deep ice cores. International Association of Hydrological Sciences Publication, 118, 351364.Google Scholar
Tika, T. E., Vaughan, P. R., and Lemos, L. J. (1996). Fast shearing of preexisting shear zones in soil. Géotechnique, 46(2), 197233.Google Scholar
Tresca, M. H. (1864). Mémoire sur l’écoulement des corps solides soumis à de fortes pressions. Comptes Rendus des Séances de l’Academie des Sciences, Paris, 59, 754758.Google Scholar
Treverrow, A., Budd, W. F., Jacka, T. H., and Warner, R. C. (2012). The tertiary creep of polycrystalline ice: experimental evidence for stress-dependent levels of strain-rate enhancement. Journal of Glaciology, 58(208), 301314.Google Scholar
Truffer, M., Echelmeyer, K. A., and Harrison, W. D. (2001). Implications of till deformation on glacier dynamics. Journal of Glaciology, 47(156), 123134.Google Scholar
Truffer, M., Harrison, W. D., and Echelmeyer, K. A. (2000). Glacier motion dominated by processes deep in underlying till. Journal of Glaciology, 46(153), 213221.Google Scholar
Tulaczyk, S. (1999). Ice sliding over weak, fine-grained tills: dependence of ice-till interactions on till granulometry. In Mickelson, D. M., and Attig, J. W. (eds.) Glacial Processes Past and Present. Boulder, CO: Geological Society of America Special Paper, 337, 159177.Google Scholar
Tulaczyk, S., Kamb, B., Scherer, R., and Engelhardt, H. (1998). Sedimentary processes at the base of a West Antarctic ice stream: constraints from textural and compositional properties of subglacial debris. Journal of Sedimentary Research, 68, 487496.Google Scholar
Tulaczyk, S., Kamb, W. B., and Engelhardt, H. F. (2000a). Basal mechanics of Ice Stream B, West Antarctica 1. Till mechanics. Journal of Geophysical Research, 105(B1), 463481.Google Scholar
Tulaczyk, S., Kamb, W. B., and Engelhardt, H. F. (2000b). Basal mechanics of Ice Stream B, West Antarctica 2. Undrained plastic bed model. Journal of Geophysical Research, 105(B1), 483494.Google Scholar
Tulaczyk, S., Kamb, W. B., and Engelhardt, H. F. (2001). Estimates of effective stress beneath a modern West Antarctic ice stream from till preconsolidation and void ratio. Boreas, 30, 101114.Google Scholar
Ussing, N. V. (1903). On Jyllands hedesletter og teorierne om deres dannelse. [On Jyllands meltwater outwash plains and theories of their origin]. Oversigt over det Kongelige Danske Videnskabernes Selskabs Forhandlinger, 2, 99165.Google Scholar
Vallon, M., Petit, J.-R., and Fabre, B. (1976). Study of an ice core to bedrock in the accumulation zone of an alpine glacier. Journal of Glaciology, 17(75), 113128.Google Scholar
Van Beaver, H. G. (1971). The significance of the distribution of clasts within the Great Pond esker and adjacent till. MS thesis, University of Maine, Orono, 61pp.Google Scholar
van de Wal, R. S. W., and Oerlemans, J. (1995). Response of valley glaciers to climatic change and kinematic waves: a study with a numerical ice flow model. Journal of Glaciology, 41(137), 142152.Google Scholar
van de Wal, R. S. W., et al. (2015). Self-regulation of ice flow varies across the ablation area in south-west Greenland. The Cryosphere, 9, 603611.Google Scholar
Van der Veen, C. J. (1998). Fracture mechanics approach to penetration of surface crevasses on glaciers. Cold Regions Science and Technology, 27, 3147.Google Scholar
Van der Veen, C. J., and Whillans, I. M. (1989). Force budget: I. Theory and numerical methods. Journal of Glaciology, 35(119), 5360.Google Scholar
Van der Veen, C. J., and Whillans, I. M. (1994). Development of fabric in ice. Cold Regions Science and Technology, 22, 171195.Google Scholar
Vaughan, D. G. (1993). Relating the occurrence of crevasses to surface strain rates. Journal of Glaciology, 39(132), 255266.Google Scholar
Vaughan, D. G. (1995). Tidal flexure at ice shelf margins. Journal of Geophysical Research, 100(B4), 62136224.Google Scholar
Vaughan, D. G., Corr, H. F. J., Doake, C. S. M., and Waddington, E. D. (1999). Distortion of isochronous layers in ice revealed by ground penetrating radar. Nature, 398(6725), 323326.Google Scholar
Vaughan, D. G., et al. (2013). Observations: cryosphere. In Stocker, T. F., 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. Cambridge: Cambridge University Press, pp. 317382.Google Scholar
Velicogna, I. (2009). Increasing rates of ice mass loss from the Greenland and Antarctic ice sheets revealed by GRACE. Geophysical Research Letters, 36, L19503, doi: 10.1029/2009GL040222.Google Scholar
Vivian, R., and Bocquet, G. (1973). Subglacial cavitation phenomena under the glacier d’Argentière, Mont Blanc, France. Journal of Glaciology, 12(66), 439451.Google Scholar
Vogel, S. W., et al. (2005). Subglacial conditions during and after stoppage of an Antarctic Ice Stream: is reactivation imminent? Geophysical Research Letters, 32, L14502, doi: 10.1029/2005GL022563.Google Scholar
von Mises, R. (1928). Mechanik der plastischen Formänderung von Kristallen. Zeitschrift fur Angewandte, Mathematik und Mechanik, 8, 161185.Google Scholar
Waite, A. H., and Schmidt, S. J. (1961). Gross errors in height indication from pulsed radar altimeters operating over thick ice or snow. Institute of Radio Engineers International Convention Record, 5, 3853.Google Scholar
Walder, J. S. (1982). Stability of sheet flow of water beneath temperate glaciers and implications for glacier surging. Journal of Glaciology, 28(99), 273293.Google Scholar
Walder, J. S., and Fowler, A. (1994). Channelized subglacial drainage over a deformable bed. Journal of Glaciology, 40(134), 315.Google Scholar
Walder, J. S., and Hallet, B. (1979). Geometry of former subglacial water channels and cavities. Journal of Glaciology, 23(89), 335346.Google Scholar
Walters, R., and Meier, M. F. (1989). Variability of glacier mass balances in Western North America. In Peterson, D. H. (ed.) Aspects of Climate Variability in the Pacific and Western Americas. Washington, DC: American Geophysical Union, Geophysical Monograph 55, pp. 365374.Google Scholar
Walters, R. A., Joshberger, E. G., and Driedger, C. L. (1988). Columbia Bay, Alaska: an ‘upside down’ estuary. Estuarine, Coastal, and Shelf Science, 26, 607617.Google Scholar
Wang, W., and Warner, R. (1999). Modelling of anisotropic ice flow in Law Dome, East Antarctica. Annals of Glaciology, 29, 184190.Google Scholar
Warren, C. R., Glasser, N. F., Harrison, S., Winchester, V., Kerr, A., and Rivera, A. (1995). Characteristics of tide-water calving at Glaciar San Rafael, Chile. Journal of Glaciology, 41(138), 273289.Google Scholar
Warren, S. G., Dadic, R., Mullen, P., Schneebeli, M., and Brandt, R. E. (2012). Albedo of bare ice near the Trans-Antarctic Mountains to represent sea-glaciers on the tropical ocean of Snowball Earth. American Geophysical Union, Fall Meeting 2012, Abstract #U11A-06.Google Scholar
Watson, E., and Luckman, B. H. (2004). Tree-ring-based mass-balance estimates for the past 300 years at Peyto Glacier, Alberta, Canada. Quaternary Research, 62, 918.Google Scholar
Weertman, J. (1957a). On sliding of glaciers. Journal of Glaciology, 3(21), 3338.Google Scholar
Weertman, J. (1957b). Deformation of floating ice shelves. Journal of Glaciology, 3(21), 3842.Google Scholar
Weertman, J. (1964). Glacier sliding. Journal of Glaciology, 5(39), 287303.Google Scholar
Weertman, J. (1972). General theory of water flow at the base of a glacier or ice sheet. Reviews of Geophysics and Space Physics, 10(1), 287333.Google Scholar
Weertman, J. (1973). Can a water-filled crevasse reach the bottom surface of a glacier? International Association of Scientific Hydrology Publication, 95, 139145.Google Scholar
Weertman, J. (1983). Creep deformation of ice. Annual Reviews of Earth and Planetary Science, 11, 215240.Google Scholar
Weertman, J., and Birchfield, G. E. (1983). Stability of sheet water flow under a glacier. Journal of Glaciology, 29(103), 374382.Google Scholar
Westgate, J. A. 1968. Linear sole markings in Pleistocene till. Geological Magazine, 105(6), 501505.Google Scholar
Whillans, I. M. (1977). The equation of continuity and its application to the ice sheet near “Byrd” Station, Antarctica. Journal of Glaciology, 18(80), 359371.Google Scholar
Whillans, I. M., Bolzan, J., and Shabtaie, S. (1987). Velocity of ice streams B and C, Antarctica. Journal of Geophysical Research, 92(B9), 88958902.Google Scholar
Whillans, I. M., and Johnsen, S. J. (1983). Longitudinal variations in glacial flow: theory and test using data from the Byrd Station strain network, Antarctica. Journal of Glaciology, 29(101), 7897.Google Scholar
Whillans, I. M., and Tseng, Y.-H. (1995). Automatic tracking of crevasses on satellite images. Cold Regions Science and Technology, 23(2), 201214.Google Scholar
Whillans, I. M., and van der Veen, C. J. (1993). New and improved determinations of velocity of Ice Streams B and C, West Antarctica. Journal of Glaciology, 39(133), 483490.Google Scholar
Whillans, I. M., and van der Veen, C. J. (1997). The role of lateral drag in the dynamics of Ice Stream B, Antarctica. Journal of Glaciology, 43(144), 231237.Google Scholar
Whitlow, S., Mayewski, P. A., and Dibb, J. E. (1992). A comparison of major chemical species input timing and accumulation at South Pole and Summit Greenland. Atmospheric Environment, 26, 20452054.Google Scholar
Wilen, L. A., Diprinzio, C. L., Alley, R. B., and Azuma, N. (2003). Development, principles, and applications of automated ice fabric analyzers. Microscopy Research and Technique, 62(1), 218.Google Scholar
Winberry, J. P., Anandakrishnan, S., Alley, R. B., Bindschadler, R. A., and King, M. A. (2009). Basal mechanics of ice streams: insights from the stick-slip motion of Whillans Ice Stream, Antarctica. Journal of Geophysical Research, 114, F01016, doi: 10.1029/2008JF001035.Google Scholar
Winberry, J. P., Anandakrishnan, S., Alley, R. B., Wiens, D. A., and Pratt, M. J. (2014). Tidal pacing, skipped slips and the slowdown of Whillans Ice Stream, Antarctica. Journal of Glaciology, 60(222), 795807.Google Scholar
Winsborrow, M. C. M., Clark, C. D., and Stokes, C. R. (2004). Ice streams and the Laurentide Ice Sheet. Géographie physique et Quaternaire, 58(2–3), 269280.Google Scholar
Winski, D. A., et al. (2017). Industrial-age doubling of snow accumulation in the Alaska Range linked to tropical ocean warming. Scientific Reports, 7, 17869, doi: 10.1038/s41598–017-18022-5.Google Scholar
Winstrup, M. (2016). A hidden Markov model approach to infer timescales for high-resolution climate archives. Proceedings of the Twenty-Eighth AAAI Conference on Innovative Applications (IAAI-16), July 27–31, 2014, Québec, Canada, 4053–4060.Google Scholar
Winstrup, M., et al. (2019). A 2700-year annual time scale and accumulation history for an ice core from Roosevelt Island, West Antarctica. Climate of the Past, 15, 751779.Google Scholar
Wiscombe, W. J., and Warren, S. G. (1981). A model for the spectral albedo of snow. I: pure snow. Journal of Atmospheric Sciences, 37, 27122733.Google Scholar
Wolovick, M. J., Creyts, T. T., Buck, W. R., and Bell, R. E. (2014). Traveling slippery patches produce thickness-scale folds in ice sheets. Geophysical Research Letters, 41, 88958901.Google Scholar
Yen, Y.-C. (1981). Review of the thermal properties of snow, ice and sea ice. Cold Regions Research and Engineering Laboratories Report 81-10, 27pp.Google Scholar
Zwally, H. J., Abdalati, W., Herring, T., Larson, K., Saba, J., and Steffen, K. (2002). Surface melt-induced acceleration of Greenland Ice-Sheet flow. Science, 297, 218222.Google Scholar
Zwally, H. J., and Giovinetto, M. B. (2000). Spatial distribution of net surface mass balance on Greenland. Annals of Glaciology, 31, 126132.Google Scholar
Zwally, H. J., Li, J., Robbins, J. W., Saba, J. L., Yi, D., and Brenner, A. C. (2015). Mass gains of the Antarctic ice sheet exceed losses. Journal of Glaciology, 62(230) 10191036.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Roger LeB. Hooke, University of Maine, Orono
  • Book: Principles of Glacier Mechanics
  • Online publication: 20 December 2019
  • Chapter DOI: https://doi.org/10.1017/9781108698207.021
Available formats
×

Save book to Dropbox

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

  • References
  • Roger LeB. Hooke, University of Maine, Orono
  • Book: Principles of Glacier Mechanics
  • Online publication: 20 December 2019
  • Chapter DOI: https://doi.org/10.1017/9781108698207.021
Available formats
×

Save book to Google Drive

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

  • References
  • Roger LeB. Hooke, University of Maine, Orono
  • Book: Principles of Glacier Mechanics
  • Online publication: 20 December 2019
  • Chapter DOI: https://doi.org/10.1017/9781108698207.021
Available formats
×