<<

CHAPTER GLACIOAEOLIAN PROCESSES, , AND 8 E. Derbyshire1 and L.A. Owen2 1Royal Holloway, University of London, Surrey, United Kingdom, 2University of Cincinnati, Cincinnati, OH, United States

8.1 INTRODUCTION The frequently strong association of with present and former glaciation has been recognized for over 80 years from field observations (Hogbom, 1923). Bullard and Austin (2011) point out that the interaction between glacial dynamics, glaciofluvial, and aeolian transport in proglacial landscapes plays an important role, not only in local environmental systems, but also in the global context by affecting the amount of generated and transported. Moreover, glacial outwash plains have been cited as a significant source of dust in the Southern Hemisphere (Sugden et al., 2009) and must also have been important dust sources in the northern hemisphere (Bullard and Austin, 2011). Cold aeolian processes and landforms have been widely acknowledged in proglacial and paraglacial (e.g., Ballantyne, 2002; Seppa¨la¨, 2004). However, relatively little work has been undertaken on glacioaeolian processes, sediments, and landforms compared to other glacial systems. The ISI Web of Science does not even provide one reference to the term ‘glacioaeolian’ and Google Scholar provides a mere 61 references. Variations on the spell- ing of glacioaeolian, including glacioeolian, glacio-aeolian, glacio aeolian, and glacio-eolian yield less than 40 citations. Even the international journal Aeolian Research provides only one reference to the term glacioaeolian. A search of glacial aeolian yields 924 and 35,300 citations in the ISI Web of Science and Google Scholar, respectively. However, this includes reference to aeolian sedi- ments that are not of glacial origin, but were deposited during a glacial event. The classic textbook on , Glaciers and Glaciation,byBenn and Evans (2010) does not even list glacioaeolian in its index. Despite this, the glacioaeolian sediments and landforms are omnipresent in glacial and proglacial evironments, and many of the great belts owe their origin to glacioaeolian pro- cesses and the that is produced in the glacioaeolian environment. The importance of glacioaeolian sediments and landforms for palaeoenvironmental reconstruc- tions has a long history. The distribution of coversands, sand , and loess mantles in Western and North America, e.g., was used to develop the hypothesis of a glacial anticyclone associ- ated with the European (Hobbs, 1942, 1943a,b). This became a framework used by a num- ber of authors to characterize the extraglacial environment of the last glacial maximum in North America and Europe (e.g., Schultz and Frye, 1968; Dylik 1969; Demeck and Kukla, 1969; Galon, 1959; Velichko and Morozova, 1969; Tricart, 1970; Maarleveld, 1960; Dylikowa, 1969; Krajewski,

Past Glacial Environments. DOI: http://dx.doi.org/10.1016/B978-0-08-100524-8.00008-7 © 2018 Elsevier Ltd. All rights reserved. 273 274 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

1977; Poser, 1932, 1948, 1950; Reiter, 1961; Lill and Smalley, 1978; Williams, 1975, Arbogast et al., 2015; Sebe et al., 2015). Since glacioaeolian silt can be transported vast distances it has become an important proxy for climate change ice cores and lake records ( et al., 2008; Sudarchikova et al., 2015; Dorfman et al., 2015). For over a century, glaciers have been regarded as the greatest single source of aeolian silt, and the nature and distribution of much of the world’s loess (-deposited predominantly in the grain size range 0.01À0.05 mm, unstratified, relatively porous, dry, and yellow buff to brown yellow in colour—Lill and Smalley, 1978) have been explained in terms of the former presence of large ice sheets on the northern continents at various times during the Quaternary (Smalley, 1966a,b; Vita-Finzi and Smalley, 1970; Boulton, 1978; Smalley and Derbyshire, 1990). A number of authors have regarded the glacial grinding of rock as the -surface process capable of producing the huge volumes of silt making up the world’s loess lands (Smalley et al., 2014, and references therein). It is fair to say that the prevailing view still favours an origin in cold-climate conditions with increased aridity in which periglacial and montane environments are added to the strictly glacial. However, much work has recently emphasized the complex sources and pathways of the that comprises the great loess deposits and other associated aeolian deposits (e.g., Nie et al., 2015; Nottebaum et al., 2015; Vanneste et al., 2015). On a more local scale, the links between glacial deposition and aeolian reworking of sediments has been illustrated from a number of glacial forelands, an early example being the observations in Greenland by Poser (1932) and, more recently, by Bullard and Austin (2011). Information on sediment generation and sources, the operation of the aeolian processes, discrimination, and palaeoenvironmental reconstruction is, not surprisingly, widely scattered through the literature on , glacial and periglacial geomorphology, and Quaternary . In order to help to address the need to highlight the importance of glacioaeolian processes, landforms, and sediments in the study of glacial and geomorphology, we build on our chapter on glacioaeolian processes, sediments, and landforms that was published in the first edition of Past Glacial Environments: Sediments, Forms and Techniques (Derbyshire and Owen, 1996; Menzies, 1996). In particular, we draw on a selection of literature that emphasizes the relation between glaciers and aeolian sediments and landforms.

8.2 SEDIMENT PRODUCTION AND SOURCES Comminution of rock to produce fragments with a particle size susceptible to movement by aeolian suspension, , and traction is effected on a large scale by several mechanisms including glacial grinding, (notably by crystal growth), and fluvial and aeolian of particles in transport (Smalley et al., 2014; Woronko and Pisarska-Jamrozy, 2015; Fig. 8.1). More localized and less well documented processes include hydration effects, chemical and biological weathering, and mass slope failure in rock and soil including those associated with earthquake shock (Hewitt, 1988; Goudie, 2005; Graly et al., 2014; Hindshaw et al., 2016; Fig. 8.2). Most of the quartz in sedimentary rocks is derived initially from igneous and metamorphic types and is predominantly of sand size. Most of the sand-size particles in the world’s sedimentary system consist of quartz derived from massive plutonic rocks and gneiss with a mean grain size of 720 μm, 8.2 SEDIMENT PRODUCTION AND SOURCES 275

FIGURE 8.1 Views of supraglacial debris on the Khumbu in Nepal. Note the abundant fine sediments generated by rock fall processes that supply the debris, and abrasion and crushing of the debris as it is transported down glacier.

FIGURE 8.2 Rock debris on snow cones in the Himalaya of northern India illustrating the potential of such processes to generate abundant fine sediment that may be transported by aeolian processes into the glacioaeolian system. and sources such as schist with a mean quartz grain size of 440 μm, bringing the initial average particle size down to about 600 μm(Blatt, 1987; Livingstone and Warren, 1996). The ancient granitic rocks such as those of the heavily glaciated Laurentian and Baltic shields contain quartz in a stressed state, which facilitates comminution (Smalley, 1966a,b; Smalley et al., 2014). In contrast, the quartz in the Quaternary sands and silts of High Asia derives less from highly stressed granite and more from thick, recycled sedimentary units in the KarakoramÀHimalayanÀTibetan zone of crustal thickening. Break up and release of quartz particles results from the considerable energy 276 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

derived from tectonic and crystal-growth processes in this rapidly rising, cold, and rather dry region. The observation by Blatt (1987) that the δ18O values of quartz from plutonic igneous rocks are consistently lower than those found in quartz from many sedimentary and metamorphic sources is used by Smalley (1990) as a further distinction between loess derived from the Pleistocene ice sheet glaciations (glacial loess) and that produced in High Asia by tectonic and weathering processes ( loess). The production of large amounts of silt requires comminution of sand-grade material, as shown by considerable discussion of the relative importance of the glacial and nonglacial processes involved. Attrition in blowing and saltating sand grains produces some silt-size material, but the experimental results of Kuenen (1960) suggested that the volume is small; this led Smalley and Vita-Finzi (1968) to reject this origin for the great loess sheets. Instead, production of silt by the grinding action at the bed of glaciers and ice sheets has led some writers to express the view that this is the only mechanism capable of producing such large volumes of silt (e.g., Smalley and Cabrera, 1970; Boulton, 1978; Collinson, 1979). Other evidence, notably particle shape and surface texture, and the mineralogy of the grade (,2 μm) has been invoked to strengthen this view. Smalley and Cabrera (1970) interpreted the predominantly angular shape of loess grains as consis- tent with glacial breakage, and fracture surfaces were attributed to glacial processing of silt grains in some European loess (Smalley et al., 1973), as well as those in the Mississippi and New Zealand (Smalley, 1966a,b, 1971). Langroudietal.(2014)present empirical work that examined the micromechanics of sand-to-silt size reduction in the quartz material by a series of grinding experiments using a high-energy agate disc mill that provides analogous conditions similar to glacial abrasion. Their results suggested that the grain breakdown is not necessarily an energy-dependent process and that noncrystallographically pure (amorphous) quartz sand and silt are inherently breakable mate- rials through a fractal breakdown process. They also revealed that the internal defects in quartz are independent of size and energy input. In addition, Owen et al. (2002) show the importance of supraglacial processes in the production of silt by examining the shape, size, and thickness of supraglacial debris on Rakhiot and Chungphar glaciers in Northen Pakistan and Glacier de Cheilon in the Swiss Alps. There is accumulating evidence that weathering processes are important producers of silt (Assallay et al., 1998; , 2007). Mechanical breakage exploits the inherent planar microfrac- tures (, 1966; Riezebos and van der Waals, 1974; Moss and , 1975; Assallay et al., 1998) generated in sand-sized quartz grains by stress gradients during and as a result of contraction or cooling from the molten state. This can result from freezeÀthaw stressing (Zeuner, 1959; Minervin, 1984; Wright, 2000; Schwamborn et al., 2012; Woronko and Pisarska-Jamrozy, 2015; Fig. 8.3), and impact during fluvial transport (Moss et al., 1973; Whalley, 1979; Palmer, 1982; Attal and Lave,´ 2009), as well as impact-breakage during aeolian transport (Smalley and Vita- Finzi, 1968; Whalley et al., 1982, 1987; et al., 2002; Smalley et al., 2014). The products of such splitting are predominantly blade-shaped fragments (Smalley, 1966a,b), a view consistent with the observed dominance of face-to-face fabrics in the younger, little-modified of and Western Europe (Derbyshire and Mellors, 1988; Derbyshire et al., 1988). The laboratory and field evidence of silt production involves the combined processes of hydration and salt crystal growth (Goudie et al., 1970, 1979; Goudie, 1974, 2005; Sperling and Cooke, 1985; Rodriguez-Navarro and Doehne, 1999; Fig. 8.4). Salt weathering is a significant (A) B (B) Ice Grain Thin premelting film C

Ice

Grain Ice Grain Ice

(C) Ice (D) Ice Thin premelting Thin premelting film film D Breakage block

Grain Grain

Direction of ice expansion during In-grain weakness Thin premelting water film freezing σ n τ (E)

Unloading (dilation) microcracks Direction of grains movement σ n Normal stress τ Shear stress

FIGURE 8.3 Schematic of the development of of quartz grains by freezing and expansion of water (AÀD) and (E) frictional sliding between two rough grains. (A) Frozen sandy-sized deposit. (B) Fragment of quartz grain with ingrain weaknesses and thin unfrozen water (premelting film) adjacent to the surface of the grain within frozen sand sediment. (C) The growth of the crystal ice causes the increase in the pressure on the unfrozen water film and disintegration of the quartz grain along the ingrain weaknesses. It occurs when thickness of the unfrozen water film becomes less than irregularities on the surface of particles. At the beginning small conchoidal fractures are formed. (D) Multiple cycles of the freezeÀthaw cause the overlapping of conchoidal features and forming breakage block microstructures. (A–D) After Woronko, B., Pisarska-Jamrozy, M., 2015. Micro-scale frost weathering of sand-sized quartz grains microscale frost weathering of quartz grains. Periglacial Process (1045–6740), doi:10.1002/ppp.1855; (E) after Blenkinsop, T., 2000. Deformation microstructures and mechanisms in minerals and rocks. Kluwer Academic Publishers, New York, 150 pp. 278 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.4 Cavernous weathering and removal of rock varnish by salt crystal growth and aeolian abrasion in Zanskar, northern India. Please note handheld GPS for scale. cause of rock tafoni and silt production in a range of dry environments including intertropical plains and piedmonts which are of particular interest (Cooke, 1981; Goudie and Day, 1981; Bradley et al., 1978), polar (Prebble, 1967; Waragai 1999; Matsuoka et al., 2006), and dry mountain slopes (Goudie, 1983). Quartz sand grains treated with saturated sulphate solutions in a hot desert environment as simulated in a climatic cabinet, showed abundant fracturing and commi- nution that produced angular silt-sized particles after only 40 cycles of temperature and humidity oscillation (Goudie et al., 1979). Scanning electron microscopy of the fragments resulting from the experiments of Sperling and Cooke (1985), e.g., revealed conchoidal and stepped fracture surfaces which are indistinguishable from those attributed by some authors specifically to glacial crushing. Activity of this kind may be high where salts become concentrated as in saturated sites such as arroyos, alluvial fans, and playa basins, locations in which the finer-grained sediments also tend to coat the surface. A combination of salt enrichment and contraction cracking on desiccation in wetting and drying cycles yields silt-sized bundles of clay-grade particles (Yaalon and Ganor, 1974; Dare-Edwards, 1984), which are removed by the wind at threshold velocities significantly lower than those required to lift the grains making up such aggregates (Gillette et al., 1980). It thus appears that the cold but semiarid to arid environments that cover so much of central Asia are capable of producing abundant silt-grade materials which, following concentration in wadis, fans and plains, are readily susceptible to deflation (Yaalon, 1969; Bruins and Yaalon, 1979; Smalley and Smalley, 1983). Van Kamp (2010) emphasized that humid tropical yield quartz-rich sands and kaolinite-rich muds, along with Si, Mg, Ca, Na, and K in solution, whereas is commonly more abundant than quartz in sands in arid- and semi-arid climates. The intimate association of till matrices of glacially comminuted , periglacial river sediments, and the European loess sheets has sustained a strong body of opinion favouring the 8.2 SEDIMENT PRODUCTION AND SOURCES 279

FIGURE 8.5 Map of North America showing extent of the Laurentide ice sheet at the last glacial maximum and wind patterns derived from a modelled glacial anticyclone (COHMAP Members, 1988) compared to those derived from loess distributions. After Muhs, D.R., Bettis III, E.A., 2000. Geochemical variations in Peoria Loess of western Iowa indicate Last Glacial paleowinds of midcontinental North America. Quat. Res. 53, 49À61 and Bettis III, E.A., Muhs, D.R., Roberts, H.M., Wintle, A.G., 2003. Last Glacial loess in the conterminous USA. Quat. Sci. Rev., 22, 1907À1946.

glacioaeolian origin of loess (e.g., Hobbs, 1933; Smalley, 1966a,b). Similarly, the loess in North America has been attributed to a glacial origin, with glacioaeolian sediment derived by deflation of glacial deposit in the forelands as the North American ice sheets retreated (Muhs et al., 2000, 2013; Bettis et al., 2003; Fig. 8.5). However, the extension of this explanation to the Loess of China (Smalley, 1968) was dependent on derivation of glacial silt from the of subtropi- cal China to the southeast, which, according to prevailing opinion at that time, as summarized in the map of Sun and Yang (1961), underwent at least three glaciations during the Pleistocene. Reevaluation has concluded that there is no unequivocal evidence of Pleistocene glaciation of the Yangtze River (Derbyshire, 1983, 1992). A further attempt to explain the main body of the Chinese loess in glacioaeolian terms postulates derivation from the Qinghai-Xizang (Tibet) Plateau 280 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

as a result of a glacial anticyclone induced by widespread ice sheets (Smalley and Krinsley, 1978). Nie et al. (2015) used a multitechnique sedimentary provenance dataset from the Yellow River to show that substantial amounts of sediment eroded from northeast Tibet and carried by the river’s upper reach are stored in the Chinese Loess Plateau and the western . Their finding revises the understanding of the origin of the Chinese Loess Plateau suggesting that the Loess Plateau is a major terrestrial sink for Yellow River sediment eroded from the glaciated northeastern margin of the .

8.3 WIND ACTION AROUND GLACIERS Wind action in and adjacent to glacially covered is expressed directly in the of bedrock surfaces and unlithified deposits, and transportation of the detached particles to be deposited as a variety of aeolian sediments and landforms (Figs. 8.6 and 8.7). Wind also acts in a number of ways that influence landforms indirectly, most notably in its effect upon snow distribution on glacial forelands, but consideration of aeolian processes will be limited in this section to direct effects. Seppa¨la¨ (2004) provides a comprehensive review of these aeolian processes in cold environments.

FIGURE 8.6 Wind-eroded rock surfaces, boulders, and pebbles in glacial environments. (A) Tor at B5000 m above sea level in Zanskar, northern India. (B) Rock outcrop, Gurla Mandata, southern Tibet. (C) Wind-polished boulder on , Gurla Mandata, southern Tibet. (D) Ventifacts on in Tashkurgan valley, southeastern Tibet. 8.3 WIND ACTION AROUND GLACIERS 281

FIGURE 8.7 View of dry deflated lake bed, a source of fine sediment, along Baltoro Glacier in the Karakoram of northern Pakistan.

Sources of the abrasive particles essential to wind erosion of bedrock surfaces have been outlined above. To these should be added the hard granular snow of the high polar regions, where grain hardness reaches a Moh value of six in snow at a temperature of 250C(Bird, 1967). Wind abrasion produces a suite of distinctive landforms including asymmetrically smoothed pebbles and cobbles (ventifacts), streamlined ridges () and grooves, polished pavements, and a variety of cavernous forms, under the general term of tafoni, in which both weathering and aeolian abrasion appear to play a part (Samuelson, 1926; Fristrup, 1952; Pissart, 1966; Cailleux, 1968; Bogacki, 1970; Akerman, 1980; McKenna-Neuman and , 1986; Koster and Dijkmans, 1988; Andre´ and Hall, 2005; Mackay and Burn, 2005; Matsuoka et al., 2006; Strini et al., 2008; Bockheim, 2010; Figs. 8.4, 8.6, and 8.8). The distribution of wind-eroded surfaces is uneven, however, and there are extensive areas in, e.g., northern Canada that are almost devoid of such features (Bird, 1967), localized effects being observed only in mechanically weak (Saint-Onge, 1965). The dry forelands of polar and continental glaciers and ice sheets, on the other hand, display abundant and well-developed forms. Early reports from northeast Greenland (Fristrup, 1952) added many obser- vations from the ice-free enclaves (Sekyra, 1969) and the dry valleys of Antarctica (Nichols, 1966). Ventifacts have been observed in many lithologies including coarse-grained granites, but the com- monest abrasion-polished surfaces on both pebbles and exposed bedrock occur in the finer-grained basic intrusives. Frost-shattered coarse-grained granite contrasts with wind-polished dykes and veins of microgranite in Taylor Valley, Antarctica. Ventifact orientation leaves little doubt that gla- cially induced katabatic move very dense, very cold air at high velocities in winter from the 282 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.8 Yardangs in the Qaidam Basin in northwest Tibet. The height of each at this location is about 20 m.

Antarctic ice sheet and its outlet (valley) glaciers along the deep troughs known as ‘dry valleys’. The wind pattern is not always so simple, however, as can be seen in the aeolian abrasion of both and gneissic bedrock around the containing Sandy Glacier, which lies almost 2000 m above the floor of Wright dry valley. That the very strong winds blow across as well as along the axis of this dry valley is attested by the composition of the englacial sediments of the cir- que glacier (Dort et al., 1969). The cavernous weathering forms seen in the Antarctic dry valleys have also been explained, in part, by wind erosion (Strini et al., 2008). There appears to be rather general agreement that weath- ering processes weaken rock surfaces, which are then etched by wind abrasion. Periodic localized melting by direct isolation of thin accumulations of snow around rock stacks and boulders provides a brief wetting phase used by Cailleux (1968) to explain the mobilization of salts (notably mirabil- ite) to form veins and efflorescences which effect substantial localized granular disaggregation of the coarser-grained surfaces and old glacial boulders. Such features are also well known from Greenland (Washburn, 1969) and the glaciated desert valley of the Karakoram in northern Pakistan and China (Fig. 8.4). Thus weakened, rock surfaces are much more susceptible to wind abrasion than in the unweathered state. However, the observed asymmetry of the tafoni does not simply reflect present-day dominant wind directions. Case hardening, by which the surface of a boulder becomes relatively stronger than the tafoni by evaporation of solutions leaving a thin cemented surface layer, certainly plays a role in both Antarctic and high mountain deserts (Andre´ and Hall, 2005; Strini et al., 2008). Such surfaces may develop through a number of stages until it, too, is removed in part by aeolian abrasion (Fig. 8.9)(Derbyshire et al., 1984). Although a preferred orientation has been recorded in some groups of tafoni, the orientations do not appear to match current dominant wind directions in Antarctica (Cailleux and Calkin, 1963) or in the 8.3 WIND ACTION AROUND GLACIERS 283

FIGURE 8.9 Conceptual model of landscape evolution in southeastern Nebraska. The model starts after initial fluvial dissection of glaciated landscape and slow accumulation of oldest loess unit (Kennard Formation). Landscape diverges between large areas of net upbuilding through loess accumulation and steep slopes that experience net erosion and retreat. After Mason, J.A., 2015. Up in the refrigerator: geomorphic response to periglacial environments in the Upper Mississippi River Basin, USA. Geomorphology 248, 363À381.

Karakoram where residual snowpatch localization (a combined product of isolation and an indirect effect of wind action) appears to be the essential cause. The erosive action of the wind in the form of deflation from the surfaces of unlithified deposits such as till, glaciofluvial, and glaciolacustrine facies is probably far more significant in terms of volume of material moved per unit of time, than is the case of abrasion of even severely weathered bedrock. The winnowing of fines (sands and silts) from freshly deposited proglacial sediments may be sufficiently rapid to produce a stone pavement within just a few years. Blowing sand and silt has been observed on many glacial outwash plains including, e.g., those in the Yukon (Nickling, 1978), Baffin Island (Church, 1972; McKenna-Neuman and Gilbert, 1986), Iceland (Bogacki, 1970; Mountney and , 2006; Dagsson-Waldhauserova et al., 2015), west Spitsbergen (Riezebos et al., 1986), and in the Karakoram glacial valleys, the Pamir, and Tibet (Fig. 8.10). Blown silt and sand may cap landforms such as moraines and form sand ramps (Figs. 8.11 and 8.12). The relation between particle size, fluid, and impact threshold velocities is summarized in Fig. 8.13 (Mabbutt, 1977). Particles ,0.1 mm in diameter are moved by suspension and particles .0.1 mm in diameter are moved by saltation, while those .0.6 mm are moved by creep. Threshold velocities for entrainment of particles, in traction, saltation, and suspension loads, are reached with higher frequency in winter on snow-free surfaces in high-latitude glacial forelands. Wind velocities and air densities may be sufficiently high in the Antarctic winter to move particles of gravel grade by trac- tion, saltation, and even suspension (Selby et al., 1974; J. J. M. van der Meer, Personal communica- tion). Moreover, since the density of air varies inversely with temperature and because the drag force 284 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.10 in the Shigar Valley in the Karakoram of northern Pakistan. Glacial winds move vast amounts of sediment through this valley during the summer.

on a particle moving in air is proportional to air density, it is generally speculated that cold winds are more effective in aeolian transport than warm winds (McKenna Neuman, 1993). As McKenna Neuman (1993) highlights, for temperatures typical of current seasonal variation in the Canadian Arctic, the flux of sediment moved by winter winds at 230C could exceed that moved in summer at 10C by a factor of at least 1.16, assuming transport-limited conditions. Particle size has a strong influence not only on the mechanism of transport but also on the sedi- mentary characteristics and landforms (Fig. 8.13). Aeolian sand is deposited as dunes and sheets within short distances of its source in glacial outwash plains, lake basins, and proglacial coastal flats. In contrast, aeolian silts are transported predominantly in suspension and may be carried hun- dreds of kilometres or more from their proglacial source locations, contributing a fertile and readily worked fraction to the soils of extensive regions remote from the glaciers. Glacial winds also vary over time and glaciers oscillate due to climate change. Ballantyne (2002) highlights the importance of reworking of sediment during , aeolian processes being particularly important in redistributing sediment. This is a view supported in numerous other studies (e.g., Kadlec et al., 2015; Mason, 2015; Martignier et al., 2015). Moreover, the loess records highlight a strong control of climate in its sedimentology, notably enhancing deposition of silt dur- ing glacial times, while palaeosols dominate during interglacials. Aeolian archives in mountain areas are particularly sensitive to climate change, providing important data on the nature of palaeoenviornmental change (Stauch, 2015; Fig. 8.14). 8.3 WIND ACTION AROUND GLACIERS 285

Aeolian sands 0.01 Sum Fluvio-aeolian sands n = 85 0.005 density Probability

0 0.004 Great Britain Sum 0.004 n = 43 density

Probability 0.002

0 0.003

SW France Sum n = 26 0.002

0.001

0 0.0015 Portugal 0.001 Sum n = 11 density0.0005 Probability density Probability

0 11.4 ka event Balling Allerad 8.2 ka event 9.3 ka event Younger Dryas Younger

12.5 25 50

(ppb) 100

2+ 200

Ca 400 800 GI-1 GI-2 GI-3 GI-4 GS-2 GS-3 GS-1 0 10 20 30 Age-4 (ka)

FIGURE 8.11 Summed probability densities of ages for the OSL-dated west European aeolian records in the Netherlands, Great Britain, Portugal, and southwest France compared with the INTIMATE event stratigraphy of Rasmussen et al. (2014). After Sitzia, L., Bertran, P., Bahain, J.-J., Bateman, M.D., Hernandez, M., Garon, H., et al., 2015. The Quaternary coversands of southwest France. Quat. Sci. Rev. 124, 84À105. 286 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.12 Glacioaeolian sands and silts capping hummocky moraines in the Waqia Valley in the Pamir.

8.4 GLACIOAEOLIAN SEDIMENTS AND LANDFORMS The wind acts as an agent of deposition on glacier surfaces, on the morainic and outwash areas adjacent to glaciers, and in more distant extraglacial terrain (Seppa¨la¨, 2004). Given that the wastage of glaciers tends to destroy supraglacial aeolian landforms and to rework aeolian deposits, it is not surprising that much aeolian on glacier surfaces has gone relatively unreported. Dry conditions tend to restrict aeolian entrainment and transport to the winter season. Nevertheless, wind-winnowing of the supraglacial load of some Alpine glaciers also takes place locally in summer when the debris dries out and eventually collapses as the subjacent glacier ice melts down, the sand and coarser grades being redeposited by sliding and rolling while the silts are blown off. This is an important process of grain size fractionation in the terminal zone of glaciers with hot, dry summers as in the Karakoram, Himalaya, and some valleys on the eastern slope of the Southern Alps of New Zealand. Aeolian deposition of sediments on glacier surfaces appears to be substantial only in the case of hyperarid polar and continental glaciers such as the Qilian Shan of northern Tibet and the Alpine glaciers in the dry valley region of eastern Antarctica. Most glaciers in the latter region show a finely dispersed silt component along the englacial foliation derived from aeolian deposition of dust and snow in the accumulation zone. On some Antarctic glaciers, the supraglacial load is predominantly of wind-rippled sand deposited by winds blowing across the glacier from adjacent debris-strewn desert surfaces. The intercalation of such sand with snow in the glacier’s accumulation zone is sometimes preserved in the englacial foliation. In the case of Sandy Glacier in Wright Valley, Antarctica, the consistency of the alternation of thin layers of sand and clean ice in the snout area has been used to support the view that movement in this cirque 8.4 GLACIOAEOLIAN SEDIMENTS AND LANDFORMS 287

Dust Sand grain diameter (mm)

0.010.04 0.1 0.2 0.4 0.6 0.8 1.8 1.5 2.0 2.5 3.0 8

60 Approximate threshold wind velocity 1 cm above surface (m s ) –1 40 Fluid threshold 6

Impact threshold 20 Drag velocity (cm s velocity Drag 4 –1 )

0 Creep Saltation Normal Suspension (local) Extreme Suspension (export)

Loess

Dunes Swales, sand sheets, and ripples

Regs

421 01 Grain diameter (∅)

FIGURE 8.13 The relation between particle size, fluid, and impact thresholds, dominant mode of aeolian transport and aeolian landforms. Mabbutt, J.A., Desert Landforms,figure, r 1977 Massachusetts Institute of Technology, by permission of The MIT Press.

glacier has been by simple rotational sliding with little deformation of the original sedimentary structure derived from aeolian deposition above the equilibrium line (Dort, 1967; Dort et al., 1969). The sands and silts blown across the proglacial plains and beyond are deposited as dunes and a number of sheet-like or planar bodies. Sand dunes derived from glacial outwash plains are known from many parts of the world including Antarctica (Webb and McKelvey, 1959; Bourke et al., 2009; Bristow et al., 2011), Alaska (Koster and Dijkmans, 1988; Busacca et al., 2003), the Himalaya, and Tibet (Owen, 2014; Figs. 8.15 and 8.16). Dunes of barchan, transverse, and whaleback form have been (A) Qilan Shan

Qinghai Lake

Gonghe Basin

Qaidam Basin Loess Sand Palaeosol Donggi Cona

Southern Tibet 2000 4000 6000 8000 10,00012,000 14,000 16,00018,000 20,000 (B) 520 510 )

–3 500 490 (W m 480 Insolation 30°N (C) 470 –10 –9 –8 –7

O record –6 18 δ

Dongge Cave –5 (D) –4 –29 –28 –27 –26 cellulose

O C. Mullieensis O C. –25 18

δ –24 (E) High Low Effective moisture Effective 2000 4000 6000 8000 10,00012,000 14,000 16,00018,000 20,000 AGE (years)

FIGURE 8.14 Palaeoclimatic comparison of the aeolian phases and selected records. (A) Phases of enhanced sediment accumulation and soil formation on the Tibetan Plateau. (B) June summer insolation at 30 degrees North of Berger and Loutre (1991). (C) δ18O record from Dongge Cave of Dykoski et al. (2005). (D) δ13C record from Hongyuan peat record of Hong et al. (2003) and (E) effective moisture for monsoonal central Asia of Herzschuh (2006). After Stauch, G., 2015. Geomorphological and palaeoclimate dynamics recorded by the formation of aeolian archives on the Tibetan Plateau. Earth-Sci. Rev. 150, 393À408. 8.4 GLACIOAEOLIAN SEDIMENTS AND LANDFORMS 289

FIGURE 8.15 field at the confluence of the Shigar and Indus valleys in the Karakoram of northern Pakistan. Glacial winds converge at this confluence reworking fluvial and glacial deposits to form a complex assemblage of pyramidal and barchan dunes.

FIGURE 8.16 Barchan dunes in southern Tibet. 290 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

described from the Antarctic dry valleys where they characteristically consist of interstratified sand and snow-derived ice which acts as a cement (Calkin and Rutford, 1974; Bristow et al., 2010a,b, 2011). Deposition of sand with snow in this way is predominantly a winter process. Although, outside the polar regions, such niveoaeolian deposits (Rochette and Cailleux, 1971; Christiansen, 1998)tendtobe destroyed by seasonal melting, some structures have been recognized in dunes (Steidmann, 1973; Ahlbrandt and Andrews, 1978; Ballantyne and Whittington, 1987; Koster and Dijkmans, 1988; Belanger´ and Filion, 1991; Bateman, 2013). For example, Ahlbrandt and Andrews (1978) recognize cold-climate deposits with distinctive tensional, compressional, and dissipation in Colorado. Small-scale sand accumulation in the lee of vegetation tussocks (sand shadows), and in the lee of moraines, gravel braids in outwash plains and beach ridges, are common on glacial forelands. Processes of deflation may also contribute to dune morphologies such as the classic blow-out dunes in Iceland (Bogacki, 1970). Parabolic dunes are particularly common in cold-climate regions, often form- ing extensive complexes (Ahlbrandt and Andrews, 1978; Koster, 1988; Belanger´ and Filion, 1991; Sitzia et al., 2015; Yan and Baas, 2015). The formation and movement of dunes has been observed with good examples being described in northwest Alaska (Koster and Dijkmans, 1988) and on the sub-Arctic eastern coast of the Hudson Bay (Belanger´ and Filion, 1991). Reconstructions of formation and move- ments of dunes include work in Colorado (Ahlbrandt and Andrews, 1978), western Nebraska (Smith, 1965) and the Netherlands (Koster, 1988). Sands with varying amounts of silt occur as sheets in down-wind locations adjacent to outwash plains and temporary, dried-out lakes or ponds. Sand sheets, often with distinctive subhorizontal planar bedding are best known from the Pleistocene record of the Netherlands where they have been termed ‘coversands’ (Maarleveld, 1960). Fig. 8.17 shows the distribution of the coversand belts in Europe with notable studies being undertaken in Belgium (Vandenberghe and Gullentops, 1977; Vandenberghe, 1983), Denmark (Kolstrup et al., 2007), France (Sitzia et al., 2015), Germany (Koster, 1988; Schwan, 1988), the Netherlands (Vandenberghe, 1991, 1993; Kasse, 1999, 2002; Vandenberghe and Kasse, 2008), Poland (Manikowska, 1991), and Scandinavia (Seppa¨la¨,1995;Ka¨yhko¨ et al., 1999). This, along with evidence such as the distribution of ventifacts and loess, the percentage of frosted coarse sand grains in aeolian sediments (Fig. 8.17b), sedimentary structures indicating dominant palaeowind directions, as well as evidence from other glacial and paraglacial sediments, are important in reconstructing Pleistocene palaeoenvironments. This has facilitated reconstruction of ice margins (Fig. 8.17a) and the configuration of the past atmospheric pressure system during the last glaciation. Cailleux (1942), e.g., believed that areas with higher percentages of frosted grains were regions that had experienced the most severe periglacial activity and strongest winds. Increases in aeolian activity and the formation of coversands are thought to have coincided with increased aridity during periods of colder climate. In the European Lowlands the deposition of coversands during late glacial times occurred during the Early Dryas Stadial and the Late Dryas Stadial, with older coversands deposited in the Earliest Dryas Stadial and the Pleniglacial (Kukla, 1975; Koster 1988). Sarntheim’s (1978) computer simulation of the distribution of moisture during the last glacial and present interglacial, and evidence for the distribution and extent of past deserts show that increased aridity during the last glacial maximum enhanced aeolian activity, the formation of dunes, and the extension of deserts, while during the climatic optimum deserts and dune formation were at a mini- mum. Using stratigraphic analysis and numerical dating of Pleistocene coversands in southwest 20° 10° W 0° 10° E 20° 30° 40° 50° 60° 70° N (A) ? Max. extent of glaciation in Europe ? A Anglian limit Ural S Saale limit ? Extent of last glaciation D Deversian limit ? W Weichselian limit Loess and Loessic deposits ? (based on Bordes 1969, Hobbs 1943) ? Extent of sand belts 60° ? based on Koster 1988 Vetifacts 0 500 km

? ? D ? A ? W 50° S S

Ice C line ap Ap Pyrenees Ice Cap Caucasus Ice Cap

20° 10° W 0° 10° E 20° 30° 40° 50° 60° 70° N (B) 0 500

>80% km Ural Ice Cap 60-80%

40-60% 20-40% n 60° o <20% ti ia c la G

f

o

f n t o o n t ti n e

a t te i x ac x

E l e G

t x s a a

L M

50°

Ice Ca Caucasus Ice line p Pyrenees Ap Cap Ice Cap

20° 10° W 0° 10° E 20° 30° 40° 50° 60° FIGURE 8.17 Maps of northern Europe comparing: (A) the distribution of loess and coversand, and reconstruction of former ice extents (B) the percentage frosted coarse sand grains. (A) After Bordes F., 1969. Le loess en France. Bull. Assoc. Franc¸. Et. Quaternaire (A.F.E.Q.) Suppl. VII. Congre’s Int. I.N.Q.U.A., 69 pp.; Hobbs, W.H., 1943a. The glacial anticyclones and the continental glaciers of North America. Am. Philos. Soc. Proc., 86, 368– 402; Hobbs, W.H., 1943b. The glacial anticyclone and the European continental glacier. Am. J. Sci., 241, 333–336; and Koster, E. A., 1988. Ancient and modem cold-climate aeolian sand deposition: a review. J. Quat. Sci. 3, 69–83. (B) after Cailleux, A., 1942. Les actions eoliennes periglaciaires en Europe. Soc. Ge´ol. France Me´m 46, 1À176. 292 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

France, Sitzia et al. (2015) constructed a renewed chronostratigraphic framework for sand deposi- tion, showing the development of transgressive dune fields since at least the Middle Pleistocene (MIS 10). Three main phases of accumulation occurred during the last glacial. The oldest (64À42 ka) is associated with wet sandsheet facies, histic horizons, and zibar-type dune fields and the Late Pleniglacial (24À14 ka) corresponds to the main phase of coversand extension in a drier context. These compare favourably with chronologies elsewhere in Europe (Fig. 8.11). By examining the Loveland loess in Iowa, Muhs et al. (2013) tested whether Earth was dustier during the last glacial period and that the increased gustiness (stronger, more frequent winds) would have enhanced dustiness and that this, in turn, would be recorded in the loess record. The Loveland loess is one of the thickest deposits of last-glacial-age (Peoria) loess in the world. Based on the geochemistry of the loess, Muhs et al. (2013) were able to show that the loess was derived not only from glaciogenic sources of the Missouri River, but from distal loess of nonglacial sources in Nebraska. Optically stimulated luminescence (OSL) dating of the loess succession showed that deposition began after B27 ka continuing to B17 ka (Fig. 8.18). The OSL ages also indicate that

FIGURE 8.18 Relation to insolation, Laurentide ice sheet oscillations, and Loveland loess deposition in Iowa. (A) insolation from 31 to 17 ka at the top of the at 65 degrees North in July; (B) timeÀdistance diagram showing the southerly extent of the Laurentide ice sheet in the midcontinent of North America from B31 to B17 ka; (C) stratigraphy, OSL ages (shown on part B as squares with error bars), and coarse/fine silt ratios at Loveland, Iowa; and (D) particle size data; and loess mass accumulation rates. After Muhs, D.R., Bettis, E.A., Roberts, H.M., Harlan, S.S., Paces, J.B., , R.L., 2013. Chronology and provenance of lastglacial (Peoria) loess in western Iowa and paleoclimatic implications. Quat. Res. 80, 468À481. 8.4 GLACIOAEOLIAN SEDIMENTS AND LANDFORMS 293

mass accumulation rates of loess were not constant, but highest and grain size coarsest at B23 ka. During this time B10 m of loess accumulated in # 2 ka. Muhs et al. (2013) argued that the timing of the coarsest grain size and highest mass accumulation rates, indicating strongest winds, coincides with a summer-insolation minimum at high latitudes in North America and the maximum south- ward extent of the Laurentide ice sheet. Their observations help to confirm the view that increased dustiness during the last glacial was driven largely by enhanced gustiness, forced by a steepened meridional temperature gradient. The down-wind fining of aeolian deposits from sand to sandy silt and to silt has been demon- strated in a number of currently glaciated regions including Spitsbergen (Riezebos et al., 1986; Van Vliet-Lano¨e and Hequette, 1987). In the case of the Holmstro¨mbreen area of Spitsbergen, commi- nution of mica also increases rapidly down-wind from the source sites, a characteristic noted in the Pleistocene coversands of the Netherlands (Riezebos et al., 1986). Periodic wetting and drying are important, even in very dry polar regions where vertical cracks arising from desiccation, thermal contraction, or both, may become filled with aeolian sands to form sand wedges (Washburn, 1980; Fig. 8.19). Other indicators of periodic moistening of the surface laminae of wind-deposited sedi- ments are adhesion structures (the so-called adhesion ripples and warts: Kocurek and Fielder, 1982) known from both modern and Pleistocene proglacial aeolian sand sheets (Ruegg, 1983). The deposition of loess, coversands, and other aeolian deposits may be complex, with sediments capping, infilling, and burying landforms (Figs. 8.12 and 8.20). Mason (2015) provides an interest- ing model for aeolian deposition and landscape evolution in southeastern Nebraska that illustrates some of the complexity (Fig. 8.9). The particle size of loess is dominantly medium silt (Fig. 8.21), while the coversands have much wider particle size variability and are often loamy. The mineralogy of the aeolian sediments is strongly influenced by source area although carbonates, frequently occurring as cements, are common in many loess deposits. Loess typically includes quartz, plagioclase, K-feldspar, mica, calcite (sometimes with dolomite), and phyllosilicate clay minerals (smectite, chlorite, mica, and kaolinite), with bulk geochemical dominated by SiO2 that ranges from B45% to 75%, but is

FIGURE 8.19 Sand-wedge near Tso Mor in southern Tibet. These formed as drifting glacioaeolian sands filled thermal contraction cracks in polygonal ground. 294 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.20 Dunes and sand ramps south of Kongur Shan in the Pamir (upper) and north of Gurla Mandata in southern Tibet (lower), derived from the glaciers and associated outwash plains.

typically 55À65%. Muhs (2013) provides ternary plots to show that SiO2/10, CaO 1 MgO, and Na2O 1 K2O can be used to show the relative abundance of quartz, carbonates, and , respectively, and how compositions compare with the most common rock types (Fig. 8.22). Most loess has a composition that is very close to average shale, but this varies between regions, as shown in Fig. 8.22.

8.5 FACIES The facies progression coversandsÀsandy loessÀloess is well documented from the proglacial fore- lands around the Northern Hemisphere Pleistocene ice sheets (Ruegg, 1981, 1983; Schwan, 1986, 1987; Koster, 1988; Figs. 8.23 and 8.24). Fig. 8.23 is a schematic diagram showing the possible and sediment associations in a glacioaeolian-dominated region with lithofacies variations shown in a section. The scale has been omitted because such facies variations can change over short distances (several kilometres), as is the case in the European Alps or Iceland, as well as over 8.5 FACIES 295

FIGURE 8.21 Plots of mean particle size and degree of sorting (standard deviation of the mean particle size) of loess from various regions. After Muhs, D.R., 2013. Loess and its geomorphic, stratigraphic, and paleoclimatic significance in the Quaternary. In: Shroder, J., Lancaster, N., Sherman, D.J., Baas, A.C.W. (Eds.), Treatise on Geomorphology. Academic Press, San Diego, CA, vol. 11, Aeolian Geomorphology, pp. 149À183.

many hundreds or thousands of kilometres as in the case in Europe, North America, and Asia. Facies may vary seasonally as shown in Fig. 8.24 (Schwan, 1986). Over large areas particular litho- facies associations are geographically governed by their proximity to the ice margins, that is, the dominant sediment source area. Much, although by no means all, loess is derived from the silt pro- duced by subglacial grinding. At the present time, loess is generally not accumulating in great quan- tities even adjacent to the polar deserts, although thick deposits of Holocene age are known from the Yukon and central Alaska (Pew´ e,´ 1955), and continuous, but thin (B1 m) loess mantles in extensive areas such as Søndre Strømfjord in west Greenland (Meer, J.J.M. van der, Personal communication). The proglacial loess of the last ice-sheet maximum in Europe contains sedimentary indicators of a complex environmental history. Ruegg (1981, 1983), Schwan (1986, 1987),andKoster (1988) all recognize a broad succession of facies associations that vary through time from fluvial to aeolian-lacustrine to aeolian in northwest Europe, representing climatic amelioration (Fig. 8.25). There is also abundant evidence of syndepositional permafrost development in the form of contrac- tion cracks, sand wedges, and pseudomorphs of ice veins and ice wedges (Harry and Gozdzik, 1988; 296 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

Average rocks 0 Noncalcerous loess 1 1 0

SiO 0.8 0.2 0.2 SiO 2 /10 0.8 O 0.4 2 0.6 2 /10 Granite 0.6 0.4 O +0.4 K 2 Shale O 0.6 Na 2 New Zealand 0.2 Basalt 0.8 loess Sandstone

0.6 1 O + K Alaskan 0 0 2 0.4 1 0.8 0.6 0.4 0.2 0 loess Na Limestone CaO + MgO 0.8 0.2 Icelandic (A) loess 0 1 1 0.8 0.6 0.4 0.2 0 Calcerous loess 0 0 Calcerous loess 1 1 North America CaO + MgO SiO 0.2 0.2 SiO 0.8 0.8 2 /10 2 /10 0.4 0.4 O 0.6 O 2 0.6 2 Nebraska Siberian 0.6 O + K 0.6 O + K loess 2 loess 0.4 2 0.4 Na Na Illinois 0.8 loess 0.8 0.2 0.2 Argentine Chinese loess Iowa loess loess 1 0 0 1 1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0

(B) (C) CaO + MgO CaO + MgO

FIGURE 8.22 Ternary plots of major element chemistry of bulk loess from various regions. (A) Noncalcareous loess; (B) calcareous loess from Asia and South America; (C) calcareous loess from North America. Also shown as an inset is average compositions of common rocks. After Muhs, D.R., 2013. Loess and its geomorphic, stratigraphic, and paleoclimatic significance in the Quaternary. In: Shroder, J., Lancaster, N., Sherman, D.J., Baas, A.C.W. (Eds.), Treatise on Geomorphology. Academic Press, San Diego, CA, vol. 11, Aeolian Geomorphology, pp. 149À183.

Svensson, 1988), as well as indications of widespread reworking by a number of processes including frost creep, , rainbeat, rainwash, and snow meltwater flow (Maarleveld, 1964; Mu¨cher and Vreeken, 1981; Vreeken and Mu¨cher 1981). In the glaciated valleys of some of the world’s greatest mountain ranges, the considerable volumes of silt produced by subglacial abrasion occur in a number of ‘storages’, which persist for varying lengths of time depending on the climatic conditions affecting, in particular, the den- sity of the vegetation cover, and also on the dynamism of the proglacial landscape as influenced by climate, topography, and tectonics. Silt, initially stored as the matrix in the till of large mor- aines, is reworked as mudflows, is incorporated in outwash rivers as , and is held 8.5 FACIES 297

Production of siltsand &

Transportsilt of sand &

Depositionof silt & fine sand

Erosion & deflation of silt & fine sand Cycles of aeolian sedimentation

Transport

General fining of

Grain size in aeolian sediments Deposition of silt & fine sand Loeses Sand Silty

Erosion & Silts deflation Nonglacial diamictons Glacial diamictons Deflation of fine sand & silt Ventifacts

FIGURE 8.23 LandformÀsediment associations in the glacioaeolian environment. (1) Silt and fine sand formation by weathering on high steep valley slopes; (2) silts and sands produced and supplied to the glacier by rock fall and other mass movement processes; (3) silts deposited in a supraglacial lake; (4) silts, sands, and other rock debris falling into and incorporated into the ice; (5) silts and sands washed into a small lake; (6) terraces composed of lacustrine silts and sands produced as lake dries or drains; (7) fine sediment produced by overland flow and deposited at base of slope; (8) meltwater stream feeding a proglacial lake; (9) proglacial lake into which lacustrine silts and sands are deposited; (10) parabolic dune; (11) ice-contact lake; (12) deflation of outwash surface; (13) barchan dune; (14) longitudinal dunes; (15) deflation hollow; (16) deflation of lacustrine silts and sands in a dried- up proglacial lake; (17) rock-strewn surface, reg-like; (18) hummocky moraine; (19) aeolian sand infilling depressions within till ridges; (20) end moraines; (21) meltwater stream dissecting the end moraine; (22) coversands; (23) floodplain sands and gravels; (24) alluvial fan; (25) river terraces capped by loess; (26) vegetated surface with formation of palaeosols; (27) fines being deflated from floodplain sediment; (28) loess hills. After Derbyshire, E., Owen, L.A., 1996. Glacioaeolian processes, sediments and landforms. In: Menzies, J. (Ed.), Past Glacial Environments: Sediments, Forms and Techniques, Wiley, Chichester, pp. 213À237. 298 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.24 Schematic representation of seasonal aeolian deposition in the periglacial environment. After Schwan, J., 1986. The origin of horizontal alternating bedding in Weichselian aeolian sands in Northwestern Europe. Sediment. Geol. 49, 73À108.

in enclosed depressions as lake beds (Owen, 1996). The volumes of such stored silts are consid- erable in the Karakoram and Himalaya (Li et al., 1984; Fort et al., 1989; Owen 1996; Benn and Owen, 2002), but the occurrence within such mountain regions of silts as aeolian accumulations tends to be limited to topographically favourable locations where thicknesses rarely exceed a metre or two (Owen et al., 1992; Lehmkuhl et al., 2000; Lee et al., 2014; Fig. 8.26). As with the Pleistocene loess of Western Europe, there is much evidence of colluvial reworking. Most silt-grade material is blown considerable distances from the immediate proglacial source sites in the high mountains, and the thick loess deposits of regions such as the Potwar Plateau in north- ern Pakistan are probably derived mainly from glacial silts that have been concentrated into thick alluvial sequences by the major rivers along the KarakoramÀHimalayan piedmont. Smalley et al. (1973) suggested that much of the silt that occurs in the North European Plain and hill lands, including the Munich region originated from glaciers, and then was fluvially transported, deposited as floodplain sediments, and subsequently deflated and deposited by aeo- lian processes to form loess as mentioned previously with the Chinese loess. These examples help to demonstrate the often polygenic origin and complex dynamics of the glacioaeolian system. 8.5 FACIES 299

Depositional Facies Structures environment

6 Dune sand facies (older inland dunes Dry aeolian Dune-foreset cross-bedding (sub) horizontal and river dunes) lamination 5 Sand sheet facies A (younger cover sands) Deflation surface, desert Evenly lamination/even horizontal or slightly pavement inclined parallel lamination, rarely cross- bedded/low-angle cross-lamination, granule Dry aeolian (seasonal frost?) and adhesion ripples/occasional strings of small pebbles, deflation levels, small frost cracks and cryogenic deformations 4 Sand sheet facies B (older cover sands) Moist aeolian (permafrost?) Evenly lamination(’layer-cake’)/horizontal alternating bedding, silty laminate or silt Wet aeolian (permafrost?) layers, adhesion ripples, frost wedges and major cryogenic deformations, vertical- platy cracks

3 (Aeolian-) lacustrine facies Shallow pools, aeolian Evenly or wave-ripple laminated, silt and supplied material gytja layers, adhesion lamination 2 (Local) flowing water Fining-upward sequence of 1 Or fluviatile facies Water current velocity low A or B Low energy (ripple phase) Climbing-ripple cross-lamination, scour fast running water troughs, horizontal lamintion,adhesion lamination Low energy (dune phase) On fast running water large-scale trough cross-bedding, sand with scattered granules, no cryogenic deformations High energy very fast running water

FIGURE 8.25 Facies associations of Weisheselian dune sands and sand sheets in northwestern Europe. The facies are in sequential order; however, facies type 6 and 5 may be synchronous and facies 4 and 3 can be in reverse order. After Koster, E.A., 1988. Ancient and modem cold-climate aeolian sand deposition: a review. J. Quat. Sci. 3, 69À83 according to Ruegg, G.H.J., 1981. Sedimentary features and grain size of glaciofluvial and periglacial Pleistocene deposits in The Netherlands and adjacent parts of Western Germany. Verh. Naturwiss. Verein Hamburg 24, 133À154; Ruegg, G.H.J., 1983. Periglacial aeolian evenly laminated sandy deposits in the Late Pleistocene of NW Europe, a facies unrecorded in modem sedimentology handbooks. In: Brookfield, M.S., Ahlbrandt, T.S. (Eds.), Aeolian Sediments and Processes, Elsevier Scientific Publications, Amsterdam, pp. 455À482; Schwan, J., 1986. The origin of horizontal alternating bedding in Weichselian aeolian sands in Northwestern Europe. Sediment. Geol., 49, 73À108; Schwan, J., 1988. The structure and genesis of Weichselian to early Hologene aeolian sand sheets in western Europe. Sediment. Geol. 55, 197À232. 300 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

FIGURE 8.26 Terrace section at Pangboche in the Khumbu Himal comprising, from bottom up: glacial diamict, aeolian silts, debris flow diamict, and colluviated loess.

8.6 CONCLUSION Studies of contemporary glacioaeolian processes and sedimentation are still few compared to the research undertaken on other aspects of glacial geology. However, studies of ancient glacioaeolian sys- tems, if loess is included, are numerous. To interpret fully the vast thickness of Quaternary glacioaeolian sediments present throughout the world requires an understanding of the production, transport, and deposition of sediment by aeolian processes within glacial and proglacial environments. Glacioaeolian sediments have great potential for use as palaeoenvironmental indicators for continental regions, especially with regard to reconstructing palaeoclimatic change. This is particularly true of the thick loess sequences in northwestern Europe, Midwestern USA, and central China that contain palaeo- sols. There is still much to be learned about the process and timing of palaeosol formation and associ- ated aeolian sediments within such sequences. Moreover, the study, interpretation, and mapping of glacioaeolian sand and silt have important economic aspects. Some of the richest agricultural regions of the world are located on glacioaeolian sediments; they also yield some of the best sand deposits for industrial use, such as aggregates, concrete, and manufacturing, which includes heat-resistant bricks, ceramics, and glass. Clearly, an understanding of the sediment geometries and facies variability is important for the exploration, exploitation and sustainable development of these resources.

REFERENCES

Ahlbrandt, T.S., Andrews, S., 1978. Distinctive sedimentary features of cold-climate eolian deposits, North Park, Colorado. Palaeogeogr. Palaeoclimatol. 25, 327À351. REFERENCES 301

Akerman, J., 1980. Studies on periglacial geomorphology in West Spitsbergen, Meddelanden Fran Lunds Universities Geografiska Institution Avhanglingrar, vol. 89. pp. 1À297. Andre,´ M.-F., Hall, K., 2005. Honeycomb development on Alexander Island, glacial history of George VI Sound and palaeoclimatic implications (Two Step Cliffs/ Oasis, W Antarctica). Geomorphology 65, 117À138. Arbogast, A.F., Luehmann, M.D., Miller, B.A., Wernette, P.A., , K.M., Waha, J.D., et al., 2015. Late-Pleistocene paleowinds and aeolian sand mobilization in north-central Lower Michigan. Aeolian Res. 16, 109À116. Assallay, A.M., Rogers, C.D.F., Smalley, I.J., Jefferson, I.F., 1998. Silt 2À62 μm, 9À4 phi. Earth-Sci. Rev. 45, 61À88. Attal, M., Lave,´ J., 2009. Pebble abrasion during fluvial transport: experimental results and implications for the evolution of the sediment load along rivers. J. Geophys. Res. Earth Surf. 114, F04023. Ballantyne, C.K., 2002. Paraglacial geomorphology. Quat. Sci. Rev. 21, 1935À2017. Ballantyne, C.K., Whittington, G., 1987. Niveoaeolian sand deposits on An Teallach, Wester , Scotland. Trans. R. Soc. Edinburgh Earth Sci. 78, 51À63. Bateman, M.D., 2013. Aeolian processes in periglacial environments. In: Shroder, J. (Ed.), Treatise on Geomorphology, vol. 8. Academic Press, San Diego, pp. 416À429. Belanger,´ S., Filion, L., 1991. Niveo-aeolian sand deposition in subarctic dunes, eastern coast of Hudson Bay, Quebec,´ Canada. J. Quat. Sci. 6, 27À37. Blenkinsop, T., 2000. Deformation Microstructures and Mechanisms in Minerals and Rocks. Kluwer Academic Publishers, New York, 150 pp. Benn, D.I., Evans, D.J.A., 2010. Glaciers and Glaciation. second ed. Routledge, London, 816 pp. Benn, D.I., Owen, L.A., 2002. Himalayan glacial sedimentary environments: a framework for reconstructing and dating former glacial extents in high mountain regions. Quat. Intl. 97À98, 3À26. Berger, A., Loutre, M.F., 1991. Insolation values for the climate of the last 10 million years. Quat. Sci. Rev. 10, 297À317. Bettis III, E.A., Muhs, D.R., Roberts, H.M., Wintle, A.G., 2003. Last Glacial loess in the conterminous USA. Quat. Sci. Rev. 22, 1907À1946. Bird, J.B., 1967. The Physiography of Arctic Canada. John Hopkins Press, Baltimore, 336 pp. Blatt, H., 1987. Oxygen isotopes and the origin of quartz. J. Sediment. Petrol. 57, 373À377. Bockheim, J.G., 2010. Soil preservation and ventifact recycling from dry-based glaciers in Antarctica. Antarct. Sci. 22, 409À417. Bogacki, M., 1970. Eolian processes on the forefield of the Skeidararjo´ ¨kull (Iceland). Acad. Pol. Sci. Bull. Ser.´ Sci. Geol.´ Geogr.´ 18, 279À287. Bordes, F., 1969. Le loess en France. Bull. Assoc. Franc¸. Et. Quaternaire (A.F.E.Q.) (Suppl. VII), Congre`s Int. I.N.Q.U.A., 69 pp. Boulton, G.S., 1978. Boulder shapes and grain-size distributions of debris as indicators of transport paths through a glacier and till genesis. Sedimentology 25, 773À799. Bourke, M.C., Ewing, R.C., Finnegan, D., McGowan, H.A., 2009. Sand dune movement in the Valley. Antarct. Geomorphol. 109, 148À160. Bradley, W.C., , J.T., Twidale, C.R., 1978. Role of salts in development of granitic tafoni, South Australia. J. Geol. 86, 647À654. Bristow, C.S., Jol, H.M., Augustinus, P., Wallis, I., 2010a. Slipfaceless whaleback dunes in a polar desert, Victoria Valley, Antarctica: insights from ground penetrating radar. Geomorphology 114, 361À372. Bristow, C.S., Augustinus, P.C., Wallis, I.C., Jol, H.M., Rhodes, E.J., 2010b. Investigation of the age and migration of reversing dunes in Antarctica using GPR and OSL, with implications for GPR on Mars. Earth . Sci. Lett. 289, 30À42. 302 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

Bristow, C.S., Augustinus, P., Rhodes, E.J., Wallis, I.C., Jol, H.M., 2011. Is climate change affecting rates of dune migration in Antarctica? Geology 39, 831À834. Bruins, H.J., Yaalon, D.H., 1979. Stratigraphy of the Netivot section in the desert loess of the Negev (Israel). Acta Geol. Acad. Sci. Hung. 22, 161À169. Bullard, J.E., Austin, M., 2011. Dust generation on a proglacial floodplain, West Greenland. J. Aeolian Res. 3, 43À54. Busacca, A.J., Beget,´ J.E., Markewich, H.W., Muhs, D.R., Lancaster, N., Sweeney, M.R., 2003. Eolian sedi- ments. Dev. Quat. Sci. 1, 275À309. Cailleux, A., 1942. Les actions eoliennes periglaciaires en Europe. Soc. Geol.´ France Mem.´ 46, 1À176. Cailleux, A., 1968. Periglacial of McMurdo Strait (Antarctica). 17, 57À90. Cailleux, A., Calkin, P.E., 1963. Orientation of hollows in cavernously weathered boulders in Antarctica. Biuletyn Peryglacialny 12, 147À150. Calkin, P.E., Rutford, R.H., 1974. The sand dunes of Victoria Valley, Antarctica. Geogr. Rev. 64, 189À216. Christiansen, H.H., 1998. Nivation forms and processes in unconsolidated sediments, NE Greenland. Earth Surf. Processes Landforms 23, 751À760. Church, M., 1972. Baffin Island Sandurs: a study of Arctic . Geol. Surv. Can. Bull. 216, 280 pp. COHMAP Members, 1988. Climatic changes of the last 18,000 years: observations and model simulations. Science 241, 1043À1052. Cooke, R.U., Doornkamp, J.C., 1990. Geomorphology in Environmental Management. A New Introduction. second ed Clarendon Press, Oxford, 515 pp. Collinson, J.D., 1979. In: Reading, H.G. (Ed.), Sedimentary Environments and Facies. Elsevier, New York. Cooke, R.U., 1981. Salt weathering in deserts. Proc. Geol. Assoc. 92, 1À16. Dagsson-Waldhauserova, P., Arnalds, O., Olafsson, H., Hladil, J., Skala, R., Navratil, T., et al., 2015. Snow- Dust Storm: unique case study from Iceland, March 6-7, 2013. Aeolian Res. 16, 69À74. Dare-Edwards, A.T., 1984. Loessic clays of south-east Australia. Loess Lett. Suppl. 2, 3À16. Demeck, J., Kukla, J., 1969. Periglazialzone, Loess and Paleolithikum der Tschekoslowakei. Tschechoslowakische Akademie der Wissenschaften, Geographisches Institut, Brno, 157 pp. Derbyshire, E., 1983. The Lushan dilemma: Pleistocene glaciation south of the Chang Jiang (Yangtze River). Z. Geomorphol. NF 27, 445À471. Derbyshire, E., 1992. Origine et remise en cause de la theorie´ des glaciations pleistoce´ `cenes de la Chine du sudest: le cas des massifs de Lushan et Huagnshan. Ann. Geogr.´ 566, 472À490. Derbyshire, E., Mellors, T.W., 1988. Geological and geotechnical characteristics of some loess and loessic soils from China and Britain: a comparison. Eng. Geol. 25, 135À175. Derbyshire, E., Owen, L.A., 1996. Glacioaeolian processes, sediments and landforms. In: Menzies, J. (Ed.), Past Glacial Environments: Sediments, Forms and Techniques. Wiley, Chichester, pp. 213À237. Derbyshire, E., Li, J.J., Perrott, F.A., Xu, S.Y., Waters, R.S., 1984. Quaternary glacial history of the Hunza valley, Karakoran Mountains, Pakistan. In: Miller, K.J. (Ed.), The International Karakoram Project 1980, vol. 2. Cambridge University Press, Cambridge, pp. 456À495. Derbyshire, E., Billard, A., Van Vliet-Lanoe¨, B., Lautridou, J.ÀP., Cremaschi, M., 1988. Loess and palaeoen- vironment: some results of a European joint programme of research. J. Quat. Sci. 3, 147À169. Dorfman, J.M., Stoner, J.S., Finkenbinder, M.S., Abbott, M.B., Xuan, C., St-Onge, G., 2015. A 37,000-year environ- mental magnetic record of aeolian dust deposition from Burial Lake, Arctic Alaska. Quat. Sci. Rev. 128, 81À97. Dort Jr., W., 1967. Internal structure of Sandy Glacier, Southern Victoria Land, Antarctica. J. Glaciol. 6, 529À540. Dort Jr., W., Roots, E.F., Derbyshire, E., 1969. -ice relationships, Sandy Glacier, Southern Victoria Land, Antarctica. Geogr. Ann. 51(A), 104À111. Dykoski, C.A., Edwards, R.L., Cheng, H., Yuan, D., Cai, Y., Zhang, M., et al., 2005. A high-resolution, absolute-dated Holocene and deglacial Asian record from Dongge cave, China. Earth Planet. Sci. Lett. 233, 71À86. REFERENCES 303

Dylik, J., 1969. L’action du vent pendant le dernier aˆge froid sur le territoire de la Pologne Centrale. Biuletyn Peryglacjalny 20, 29À44. Dylikowa, A., 1969. Le proble`me des dunes interieures´ en Pologne a˘ la lumie`re des etudes´ de structure. Biuletyn Peryglacjalny 20, 45À80. Fort, M., Burbank, D.W., Freytet, P., 1989. Lacustrine sedimentation in a semiarid alpine setting: an example from Ladakh, northwestern Himalaya. Quat. Res. 31, 332À350. Fristrup, B., 1952. Wind erosion within the Arctic deserts. Geogr. Tidsskr. 51À56. Galon, R., 1959. New investigations of inland dunes in Poland. Przegl. Geogr. 31, 93À110. Gillette, D.A., Adams, J., Endo, A., Smith, D., Kihl, R., 1980. Threshold velocities for imput of the soil parti- cles into the air by desert soils. J. Geophys. Res. 85, 5621À5630. Goudie, A.S., 1974. Further experimental investigation of rock weathering by salt and other mechanical pro- cesses. Z. Geomorphol., Suppl. 21, 1À12. Goudie, A.S., 1983. Dust storms in space and time. Prog. Phys. Geogr. 7, 502À530. Goudie, A.S., 2005. Rocks under pressure. Phys. World 18, 22. Goudie, A.S., Day, M.J., 1981. Disintegration of fan sediments in Death Valley, California, by salt weathering. Phys. Geogr. 1, 126À137. Goudie, A.S., Cooke, R.U., Evans, I., 1970. Experimental investigation of rock weathering by salts. Area 2, 42À48. Goudie, A.S., Cooke, R.U., Doornkamp, J.C., 1979. The formation of silt from quartz dune sand by salt- weathering processes in deserts. J. Arid Environ. 2, 105À112. Graly, J.A., Humphrey, N.F., Landowski, C.M., Harper, J.T., 2014. Chemical weathering under the Greenland ice sheet. Geology 42, 551À554. Harry, D.G., Gozdzik, J.S., 1988. Ice wedges: growth, thaw transformation, and palaeoenvironmental signifi- cance. J. Quat. Sci. 3, 39À550. Herzschuh, U., 2006. Palaeo-moisture evolution inmonsoonal Central Asia during the last 50,000 years. Quat. Sci. Rev. 25, 163À178. Hewitt, K., 1988. Catastrophic deposits in the Karakoram Himalaya. Science 242, 64À67. Hindshaw, R.S., Heaton, T.H.E., Boyd, E.S., Lindsay, M.R., Tipper, E.T., 2016. Influence of glaciation on mechanisms of mineral weathering in two high Arctic catchments. Chem. Geol. 420, 37À50. Hobbs, W.H., 1933. The origin of the loess associated with continental glaciation based upon studies in Greenland. International Geographical Congress, Paris 1931. C. R. Geosci. 2, 408À411. Hobbs, W.H., 1942. Wind À the dominant transportation agent within extra-marginal zones to continental gla- ciers. J. Geol. 50, 556À559. Hobbs, W.H., 1943a. The glacial anticyclones and the continental glaciers of North America. Am. Philos. Soc. Proc. 86, 368À402. Hobbs, W.H., 1943b. The glacial anticyclone and the European continental glacier. Am. J. Sci. 241, 333À336. Hogbom, I., 1923. Ancient inland dunes of northern and middle Europe. Geogr. Ann. 5, 113À243. Hong, Y.T., Hong, B., Lin, Q.H., Zhu, Y.X., Shibata, Y., Hirota, M., et al., 2003. Correlation between Indian Ocean summer monsoon and North Atlantic climate during the Holocene. Earth Planet. Sci. Lett. 211, 371À380. Kadlec, J., Kocurek, G., Mohrig, D., Shinde, D.P., Murari, M.K., Varma, V., et al., 2015. Response of fluvial, aeolian, and lacustrine systems to late Pleistocene to Holocene climate change, Lower Moravian Basin, Czech Republic. Geomorphology 232, 193À208. Kasse, C., 1999. Late Pleniglacial and Late Glacial aeolian phases in the Netherlands. GeoArchaeo-Rhein 3, 61À82. Kasse, C., 2002. Sandy aeolian deposits and environments and their relation to climate during the Last Glacial Maximum and Lateglacial in northwest and central Europe. Prog. Phys. Geogr. 26, 507À532. Ka¨yhko¨, J.A., Worsley, P., Pye, K., Clarke, M.L., 1999. A revised chronology for aeolian activity in subarctic Fennoscandia during the Holocene. Holocene 9, 195À205. 304 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

Kocurek, G., Fielder, G., 1982. Adhesion structures. J. Sediment. Petrol. 52, 1229À1241. Kolstrup, E., Murray, A., Possnert, G., 2007. Luminescence and radiocarbon ages from laminated Lateglacial aeolian sediments in western Jutland, Denmark. Boreas 36, 314À325. Koster, E.A., 1988. Ancient and modem cold-climate aeolian sand deposition: a review. J. Quat. Sci. 3, 69À83. Koster, E.A., Dijkmans, J.W.A., 1988. Niveo-aeolian deposits and denivation forms, with special reference to the Greak Kobuk Sand Dunes, Northwestern Alaska. Earth Surf. Processes Landforms 13, 153À170. Krajewski, K., 1977. Po´znoplejstocenskie i Holocenskie procesy wydmotwo´rcze w Pradolinie Warszawsko- Berlinskiej w wid_ach Warty i Neru. Acta Geogr. Lodz. 39, 87 pp. Kuenen, P.H., 1960. Experimental abrasion 4: Eolian action. J. Geol. 68, 427À449. Kukla, G.J., 1975. Loess stratigraphy of central Europe. In: Butzer, K.W., Issac, G.L. (Eds.), After the Australopithecines: Stratigraphy, Ecology, and Culture in the Middle Pleistocene. Mouton Publishers, The Hague, pp. 99À188. Lambert, F., Delmonte, D., Petit, J.R., et al., 2008. Dust-climate couplings over the past 800 000 years from the EPICA Dome C . Nature 452, 616À619. Langroudi, A.A., Jefferson, I., O’hara-Dhand, K., Smalley, I., 2014. Micromechanics of quartz sand breakage in a fractal context. Geomorphology 211, 1À10. Lee, S.Y., Seong, Y.B., Owen, L.A., Murari, M.K., Lim, H.S., Yoon, H.I., et al., 2014. Late Quaternary glacia- tion in the Nun-Kun massif, northwestern India. Boreas 43, 67À89. Lehmkuhl, F., Klinge, M., Rees-, J., Rhodes, E.J., 2000. Late Quaternary eolian sedimentation in central and south-eastern Tibet. Quat. Intl. 68, 117À132. Li, J.J., Derbyshire, E., Xu, S.Y., 1984. Glacial and paraglacial sediments of the Hunza valley, northwest Karakoram, Pakistan: a preliminary analysis. In: Miller, K. (Ed.), The International Karakoram Project 1980, 2. Cambridge University Press, Cambridge, pp. 496À535. Lill, G.O., Smalley, I.J., 1978. Distribution of loess in Britain. Proc. Geol. Assoc. 88, 57À65. Livingstone, I., Warren, A., 1996. Aeolian Geomorphology. Longman, Harlow, 211 pp. Mabbutt, J.A., 1977. Desert Landforms. MIT Press, Cambridge. Mackay, J.R., Burn, C.R., 2005. A long-term field study (1951À2003) of ventifacts formed by katabatic winds at Paulatuk, western Arctic coast, Canada. Can. J. Earth Sci. 42, 1615À1635. Manikowska, B., 1991. Vistulian and Holocene aeolian activity, pedostratigraphy and relief evolution in Central Poland. Z. Geomorph. 90, 131À141. Maarleveld, G.C., 1960. Wind directions and cover sands in The Netherlands. Biuletyn Peryglacjalny 8, 49À58. Maarleveld, G.C., 1964. Periglacial phenomena in The Netherlands during different parts of the Wu¨rm time. Biuletyn Peryglacjalny 14, 251À256. Martignier, L., Nussbaumer, M., Adatte, T., Gobat, J.-M., Verrecchia, E.P., 2015. Assessment of a locally- sourced loess system in Europe: The Swiss Jura Mountains. Aeolian Res. 18, 11À21. Mason, J.A., 2015. Up in the refrigerator: geomorphic response to periglacial environments in the Upper Mississippi River Basin, USA. Geomorphology 248, 363À381. Matsuoka, N., Thomachot, C.E., Oguchi, C.T., Hatta, T., Abe, M., Matsuzaki, H., 2006. Quaternary bedrock erosion and landscape evolution in the Sør Rondane Mountains, East Antarctica: reevaluating rates and processes. Geomorphology 81, 408À420. McKenna Neuman, C., 1993. A review of aeolian transport processes in cold environments. Prog. Phys. Geogr. 17, 137À155. McKenna-Neuman, D.P., Gilbert, R., 1986. Aeolian processes landforms in glaciofluvial environments of southeastern Baffin Island, N.W.T. Canada. In: Nicklings, W.G. (Ed.), Aeolian Geomorphology. Allen & Unwin, Boston, MA, pp. 213À235. Menzies, J. (Ed.), 1996. Past Glacial Environments: Sediments, Forms and Techniques. Wiley, Chichester. REFERENCES 305

Minervin, A.V., 1984. Cryogenic processes in loess formation in central Asia. In: Velichko, A.A. (Ed.), Late Quaternary Environments of the Soviet Union. Longman, London, pp. 133À140. Moss, A.J., 1966. Origin, shaping and significance of quartz sand grains. J. Geol. Soc. Aust. 13, 97À136. Moss, A.J., Green, P., 1975. Sand and silt grains: predetermination of their formation and properties by microfractures in quartz. J. Geol. Soc. Aust. 22, 485À495. Moss, A.J., Walker, P.H., Hutka, J., 1973. Fragmentation of granitic quartz in water. Sedimentology 20, 489À511. Mountney, N.P., Russell, A.J., 2006. Coastal aeolian dune development, So´lheimasandur, southern Iceland. Sediment. Geol. 192, 167À181. Mu¨cher, H.J., Vreeken, W.J., 1981. (Re)deposition of loess in southern Limbourg, The Netherlands: 2. Micromorphology of the Lower Silt Loam complex and comparison with deposits produced under labo- ratory conditions. Earth Surf. Processes Landforms 6, 355À363. Muhs, D.R., 2013. Loess and its geomorphic, stratigraphic, and paleoclimatic significance in the Quaternary. In: Shroder, J., Lancaster, N., Sherman, D.J., Baas, A.C.W. (Eds.), Treatise on Geomorphology, vol. 11. Academic Press, San Diego, CA, pp. 149À183. , Aeolian Geomorphology. Muhs, D.R., Bettis III, E.A., 2000. Geochemical variations in Peoria Loess of western Iowa indicate Last Glacial paleowinds of midcontinental North America. Quat. Res. 53, 49À61. Muhs, D.R., Bettis, E.A., Roberts, H.M., Harlan, S.S., Paces, J.B., Reynolds, R.L., 2013. Chronology and provenance of last-glacial (Peoria) loess in western Iowa and paleoclimatic implications. Quat. Res. 80, 468À481. Nichols, R.L., 1966. Geomorphology of Antarctica. In: Tedrow, J.C.F. (Ed.), Antarctic Soils and Soil Forming Processes. American Geophysical Union, pp. 1À46. , Antarctic Research Series, 8. Nickling, W.G., 1978. Eolian during dust storms: Slims River Valley, Yukon Territory. Can. J. Earth Sci. 15, 1069À1084. Nie, J., Stevens, T., Rittner, M., Stockli, D., Garzanti, E., Limonta, M., et al., 2015. Loess Plateau storage of Northeastern Tibetan Plateau-derived Yellow River sediment. Nat. Commun. 6, 8511. Nottebaum, V., Lehmkuhl, F., Stauch, G., Lu, H., Yi, S., 2015. Late Quaternary aeolian sand deposition sus- tained by fluvial reworking and sediment supply in the Hexi Corridor—an example from northern Chinese drylands. Geomorphology 250, 113À127. Owen, L.A., 1996. Quaternary lacustrine deposits in a high-energy semi-arid mountain environment, Karakoram Mountains, northern Pakistan. J. Quat. Sci. 11, 461À483. Owen, L.A., 2014. Himalayan landscapes of India. In: Kale, V.S. (Ed.), Landscapes and Landforms of India. World Geomorphological Landscapes. Springer Science 1 Business Media, Dordrecht, pp. 41À52. Owen, L.A., White, B., Rendell, H., Derbyshire, E., 1992. Loessic silts in the western Himalayas: their sedi- mentology, genesis and age. Catena 19, 493À509. Owen, L.A., Derbyshire, E., Scott, C.H., 2002. Contemporary sediment production and transfer in high-altitude glaciers. Sediment. Geol. 155, 13À36. Palmer, A.S., 1982. Stratigraphy and Selected Properties of Loess in Wairarapa, New Zealand (Unpublished Ph.D. thesis). Victoria University, Wellington, New Zealand. Pew´ e,´ T.L., 1955. Origin of the upland silt newar Fairbanks, Alaska. Geol. Soc. Am. Bull. 66, 699À724. Pissart, A., 1966. Le roˆle geomorphologique´ du vent dans la region de Mould Bay (Ile Prince Patrick—N.W. T.—Canada). Z. Geomorphol. 10, 226À236. Poser, H., 1932. Eingie Untersuchungen zur Morphologie Ostgro¨nlands. Medd. Gronl. 94, 55 pp. Poser, H., 1948. Aolische Ablaerungen und Klima des Spa¨tglazials in Mittel-und Westeuropa. Die Naturwiss. 35, 269À276, 307À312. Poser, H., 1950. Zur Rekonstruktion der Spa¨tglazialen Luftdruckverhaltnisse in Mittel und Westeuropa auf Grund der Vorzeitlichen Binrendu¨nen. Erdkunde 4, 81À88. Prebble, M.M., 1967. Cavernous weathering in the Taylor dry valley, Victoria Land, Antarctica. Nature 216, 1194À1195. 306 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

Rasmussen, S.O., Bigler, M., Blockley, S.P., Blunier, T., Buchardt, S.L., Clausen, H.B., et al., 2014. A stratigraphic framework for abrupt climatic changes during the Last Glacial period based on three synchronized Greenland ice-core records: refining and extending the INTIMATE event stratigraphy. Quat. Sci. Rev. 106, 14À28. Reiter, E.R., 1961. Meterologie der Strahlstro¨me (Jet Streams). Springer Verlag, Wien, 473 pp. Riezebos, P.A., van der Waals, L., 1974. Silt-sized quartz particles: a proposed source. Sediment. Geol. 12, 279À285. Riezebos, P.A., Boulton, G.S., Meer, J.J.M., van der Ruegg, G.H.J., Beets, D.J., Castel, I.I.Y., et al., 1986. Products and effects of modern aeolian activity on a nineteenth-century glacier-pushed ridge in West Spitsbergen, Svalbard. Arct. Alp. Res. 18, 389À396. Rochette, J.-C., Cailleux, A., 1971. Depo´ ˆts niveo-´ eoliens´ annuels a` Poste-de-la-Baleine, Nouveau-Quebec.´ Rev. Geogr.´ Montreal´ 25, 35À41. Rodriguez-Navarro, C., Doehne, E., 1999. Salt weathering: influence of evaporation rate, supersaturation and crystallization pattern. Earth Surf. Processes Landforms 24, 191À209. Ruegg, G.H.J., 1981. Sedimentary features and grain size of glaciofluvial and periglacial Pleistocene deposits in The Netherlands and adjacent parts of Western Germany. Verh. Naturwiss. Verein Hamburg 24, 133À154. Ruegg, G.H.J., 1983. Periglacial aeolian evenly laminated sandy deposits in the Late Pleistocene of NW Europe, a facies unrecorded in modem sedimentology handbooks. In: Brookfield, M.S., Ahlbrandt, T.S. (Eds.), Aeolian Sediments and Processes. Elsevier Scientific Publications, Amsterdam, pp. 455À482. Saint-Onge, D.A., 1965. La Geomorphologie´ de I’Iˆlle Ellef Ringnes, Territories du NordOuest, Canada. Canada Department of Mines and Technical Surveys, Geographical Branch Geographical Paper, Ottawa, No. 38, 46 pp. Samuelson, C., 1926. Studien U¨ ber die Wirkungen des Windes in den Kalten und gema¨ssigten Erdtielen. Bull. Geol. Inst. Univ. Uppsala 20, 57À230. Sarntheim, M., 1978. Sand deserts during glacial maximum and climatic optimum. Nature 272, 43À46. Schultz, C.B., Frye, J.C., 1968. ). Loess and Related Eolian Deposits of the World. University of Nebraska Press, Lincoln, 369 pp. Schwamborn, G., Schirrmeister, L., Fru¨tsch, F., Diekmann, B., 2012. Quartz weathering in freeze-thaw cycles: experiment and application to the el’gygytgyn crater lake record for tracing Siberian permafrost history. Geogr. Ann. Series A Phys. Geogr. 94, 481À499. Schwan, J., 1986. The origin of horizontal alternating bedding in Weichselian aeolian sands in Northwestern Europe. Sediment. Geol. 49, 73À108. Schwan, J., 1988. The structure and genesis of Weichselian to early Hologene aeolian sand sheets in western Europe. Sediment. Geol. 55, 197À232. Schwan, K., 1987. Sedimentologie characteristics of a fluvial to aeolian succession in Weichselian talsand in the Emsland (FRG). Sediment. Geol. 52, 273À289. Sebe, K., Roetzel, R., Fiebig, M., Luthgens, C., 2015. Pleistocene wind system in eastern Austria and its impact on landscape evolution. Catena 134, 59À74. Sekyra, J., 1969. Periglacial phenomena in the oases and the mountains of the Enderby Land and the Dronning Maud Land (Easy Antarctica). Biuletyn Peryglacjalny 19, 277À289. Selby, M.J., Rains, R.B., Palmer, R.W.P., 1974. Eolian deposits of the ice-free Victoria Valley, southern Victoria Land, Antarctica. N. Z. J. Geol. Geophys. 17, 543À562. Seppa¨la¨, M., 1995. Deflation and redeposition of sand dunes in Finnish Lapland. Quat. Sci. Rev. 14, 799À809. Seppa¨la¨, M., 2004. Wind as a Geomorphic Agent in Cold Climates. Cambridge University Press, Cambridge, 378 pp. Sitzia, L., Bertran, P., Bahain, J.-J., Bateman, M.D., Hernandez, M., Garon, H., et al., 2015. The Quaternary coversands of southwest France. Quat. Sci. Rev. 124, 84À105. Smalley, I.J., 1966a. Formation of quartz sand. Nature 211, 476À479. REFERENCES 307

Smalley, I.J., 1966b. The properties of glacial loess and the formation of loess deposits. J. Sediment. Petrol. 36, 669À676. Smalley, I.J., 1968. The loess deposits and Neolithic culture of Northern China. Man 3, 224À241. Smalley, I.J., 1971. ‘In-situ’ theories of loess formation and the significance of the calcium carbonate content of loess. EarthÀSci. Rev. 7, 67À85. Smalley, I.J., Cabrera, J.G., 1970. The shape and surface texture of loess particles. Geol. Soc. Am. Bull. 81, 1591À1596. Smalley, I.J., 1990. Possible mechanisms for the model coarse-silt quartz particles in loess deposits. Quat. Int 7/8, 23À27. Smalley, I.J., Derbyshire, E., 1990. The definition of ice-sheet and mountain loess. Area 22, 300À301. Smalley, I.J., Krinsley, D.H., 1978. Loess deposits associated with deserts. Catena 5, 53À66. Smalley, I.J., Smalley, V., 1983. Loess material and loess deposits: formation, distribution and consequences. In: Brookfield, M.E., Ahlbrandt, T.S. (Eds.), Eolian Sediments and Processes. Elsevier, Amsterdam, pp. 51À68. Smalley, I.J., Vita-Finzi, C., 1968. The formation of fine particles in sandy deserts and the nature of ‘desert’ loess. J. Sediment. Petrol. 38, 766À774. Smalley, I.J., Krinsley, D.H., Vita-Finzi, C., 1973. Observations on the Kaisertu¨hl loess. Geol. Mag. 110, 29À36. Smalley, I., O’Hara-Dhand, K., Kwong, J., 2014. China: materials for a loess landscape. Catena 117, 100À107. Smith, H.T.U., 1965. Dune morphology and chronology in central and western Nebraska. J. Geol 73, 557À578. Smith, B.J., Wright, J.S., Whalley, W.B., 2002. Sources of non-glacial, loess-size quartz silt and the origins of “desert loess”. Earth-Sci. Rev. 59, 1À26. Sperling, C.H.B., Cooke, R.U., 1985. Laboratory simulation of rock weathering by salt crystallization and hydration processes in hot, arid environments. Earth Surf. Processes Landforms 10, 541À555. Stauch, G., 2015. Geomorphological and palaeoclimate dynamics recorded by the formation of aeolian archives on the Tibetan Plateau. Earth-Sci. Rev. 150, 393À408. Steidmann, J.R., 1973. Ice and snow in eolian sand dunes of southwestern Wyoming. Science 179, 796À798. Strini, A., Guglielmin, M., Hall, K., 2008. Tafoni development in a cryotic environment: an example from Northern Victoria Land, Antarctica. Earth Surf. Processes Landforms 33, 1502À1519. Sudarchikova, N., Mikolajewicz, U., Timmreck, C., O’Donnell, D., Schurgers, G., Sein, D., et al., 2015. Modelling of mineral dust for interglacial and glacial climate conditions with a focus on Antarctica. Clim. Past 11, 765À779. Sugden, D.E., McCulloch, R.D., Bory, A.J.-M., Hein, A.S., 2009. Influence of Patagonian glaciers on Antarctic dust deposition during the last glacial period. Nat. Geosci . Available from: http://dx.doi.org/ 10.1038/NGEO474. Sun, T.C., Yang, H.J., 1961. The great Ice Age glaciation in China. Acta Geol. Sin. 41, 234À244. Svensson, H., 1988. Ice-wedge casts and relict polygonal patterns in Scandinavia. J. Quat. Sci. 3, 57À67. Tricart, J., 1970. Geomorphology of Cold Environments. Macmillan, London, 320 pp. Van Kamp, P.C.D.E., 2010. Arkose, subarkose, quartz sand, and associated muds derived from felsic plutonic rocks in glacial to tropical humid climates. J. Sediment. Res. 80, 895À918. Van Vliet-Lano¨e, B., Hequette, A., 1987. Activite´ eolienne´ et sables limoneaux sur les versants exposes´ au nord-est de la Peninsule du Brogger, Spitzberg du Nord-Ouest (Svalbard). In: Pecsi, M., French, H.M. (Eds.), Loess and Periglacial Phenomena. Akademia Kiado, Budapest, pp. 103À123. Vandenberghe, J., 1983. Late Weichselian river dune formation Grote Nete Valley, Central Belgium. Z. Geomorphol. 45, 251À263. 308 CHAPTER 8 GLACIOAEOLIAN PROCESSES, SEDIMENTS, AND LANDFORMS

Vandenberghe, J., 1991. Changing conditions of aeolian sand deposition during the last deglaciation period. Z. Geomorphol. Suppl. 90, 193À207. Vandenberghe, J., 1993. Changing fluvial processes under changing periglacial conditions. Z. Geomorphol. Suppl. 88, 17À28. Vandenberghe, J., Gullentops, F., 1977. Contribution to the stratigraphy of the Weichselian Pleniglacial in the Belgian coversand area. Geol. Mijnb 56, 123À128. Vandenberghe, J., Kasse, C., 2008. Les formations sableuses en milieux periglaciaires: sables de couverture et sables dunaires. In: Dewolf, Y., Bourrie, G. (Eds.), Les Formations Superficielles. Ellipses, Paris, pp. 317À321. Vanneste, H., De Vleeschouwer, K., Mart´ınez-Cortizas, A., von Scheffer, C., Piotrowska, N., Coronato, A., et al., 2015. Late-glacial elevated dust deposition linked to westerly wind shifts in southern South America. Sci. Rep. 5, 11670. Velichko, A.A., Morozova, T.D., 1969. Structure of the loess mass of the Russian plain. Akad. Nauk SSSR Izv. Ser. Geogr. 4, 18À29 (in Russian). Vita-Finzi, C., Smalley, I.J., 1970. Origin of quartz silt: comments on a note by Ph.H.Kuenen. J. Sediment. Petrol. 40, 1367À1369. Vreeken, W.J., Mu¨cher, H.J., 1981. (Re)deposition of loess in southern Limbourg, The Netherlands. 1. Field evidence for conditions of deposition of the lower silt loam complex. Earth Surf. Processes Landforms 6, 337À354. Waragai, T., 1999. Weathering processes on rock surfaces in the Hunza Valley, Karakoram, North Pakistan. Z. Geomorphol. Suppl. 119, 119À136. Washburn, A.L., 1969. Weathering, frost action and in the Mesters Vig District, Northeast Greenland. Medd. Grønl. 176, 303 pp. Washburn, A.L., 1973. Periglacial Processes and Environments. Edward Arnold, London, 320 pp. Washburn, A.L., 1980. Geocryology: A Survey of Periglacial Processes and Environments. Edward Arnold Publishing, London, 406 pp. Webb, P.N., McKelvey, P.C., 1959. Geological investigations in South Victoria Land, Antarctica. N. Z. J. Geol. Geophys. 2, 120À136. Whalley, W.B., 1979. Quartz silt production and sand grain surface textures from fluvial and glacial environ- ments. Scanning Electron Microsc. 1979, 547À551. Whalley, W.B., Smith, B.J., McAllister, J.J., Edwards, A.J., 1987. Aeolian abrasion of quartz particles and the production of silt-size fragments: preliminary results. In: Frostick, L., Reid, I. (Eds.), Desert Sediments: Ancient and Modern. Geological Society of London, Special Publication 35, pp. 129À138. Williams, R.G.B., 1975. The British climate during the Last Glaciation: an interpretation based on periglacial phenomenon. In: Wright, A.E., Moseley, F. (Eds.), Ice Ages: Ancient and Modern. Seel House Press, Liverpool, pp. 95À120. , Geological Journal, Special Issue 6. Woronko, B., Pisarska-Jamrozy, M., 2015. Micro-scale frost weathering of sand-sized quartz grains micro- scale frost weathering of quartz grains. Permafrost Periglacial Process. (1045À6740), p. n doi:10.1002/ ppp.1855. Wright, J.S., 2000. The Spalling of overgrowths during experimental freeze-thaw of a quartz sandstone as a mechanism of quartz silt production. Micron 31, 631À638. Wright, J.S., 2007. An overview of the role of weathering in the production of quartz silt. Sediment. Geol. 202, 337À351. Yaalon, D.H., 1969. Origin of desert loess. Abstracts. In: 8th INQUA congress, Paris, 2, p. 755. Yaalon, D.H., Ganor, E., 1974. The influence of dust on soils during the Quaternary. Soil Sci. 116, 146À155. Yan, N., Baas, A.C.W., 2015. Parabolic dunes and their transformations under environmental and climatic changes: towards a conceptual framework for understanding and prediction. Glob. Planet. Change 124, 123À148. Zeuner, F.E., 1959. The Pleistocene Period. Hutchinson, London, 447 pp.