<<

CHAPTER GLACIAL LITHOFACIES AND 11 J. Lee British Geological Survey, Nottingham, United Kingdom

11.1 INTRODUCTION Reconstructing the environments, dynamics, and record of past glaciation requires a detailed knowl- edge of both the products of glaciation—effectively the , landforms, and glacitectonic structures that we see in the geological record, and the genetic processes that formed them (Fig. 11.1). Such an understanding of glacial deposits, landforms, and processes underpins our understanding of behaviour, the links between ice masses, climate change, and other feed- back mechanisms, and the applied significance of glaciated terrains from the perspective of resources and geohazards (Fig. 11.1). Building all of this knowledge into a robust geological model requires the employment of a systematic methodology for describing, recording, and interpreting geological evidence. For glacial sediments the principal method, and one routinely employed elsewhere in sedimentology, is the hierarchical lithofacies approach. In turn, understanding how these lithofacies and other glacial features (e.g., landforms and glacitectonic structures) fit together and correlate in both time and space is called stratigraphy. Stratigraphy is a key concept within . It enables the development of a framework of events and features that describe both the evolution of a geological succession and how that succession fits into the wider palaeoenvironmental picture or Earth . Within the context of glacial geology, e.g., a succession of glacigenic sediments and landforms in the Great Lakes region of Canada might document the repeated glaciation of the area. Studying the sediments (lithofacies) and landforms plus their temporal and spatial relationships (stratigraphy), would enable the number of ice advances, their flow directions, and associated geological processes to be reconstructed. These could then be linked-in, perhaps with , to the larger- scale behaviour of different lobes of the last Laurentide and ultimately an ice sheet model. Such an approach is equally valid fromanappliedapproachwithlevelsofgenetic interpretation replaced by applied interpretation for the purpose of developing applied thematic models. This chapter will review and illustrate the principles behind lithofacies analyses and strati- graphic approaches within glacial successions. It will outline the appropriate techniques and meth- ods that can be employed as well as the relevant complexities and challenges that may be encountered when applying them.

Past Glacial Environments. DOI: http://dx.doi.org/10.1016/B978-0-08-100524-8.00011-7 © 2018 Elsevier Ltd. All rights reserved. 377 378 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Geological Lithofacies model Geological Stratigraphy interpretation Applied Landforms model

Description Interpretation Application

FIGURE 11.1 The hierarchical approach to glacial geology encompassing description, interpretation, and application.

11.2 GEOLOGICAL COMPLEXITIES IN GLACIAL SEQUENCES The lithofacies approach (Section 11.3), together with many of the stratigraphic techniques outlined (Section 11.4) are commonly used within a range of geological contexts. However, their successful application to glacial sequences requires due care and consideration of a number of factors that control both preservation potential and the spatial and temporal complexity encountered within the geological record (Rose and Menzies, 1996). The preservation potential for continental (as opposed to marine) glacial sediments and land- forms is generally much lower than many other sedimentary environments. This is due to their susceptibility to reworking or removal on land during subsequent glacial events and by the vast range of glacial (e.g., mass-wasting, outburst floods) and nonglacial (e.g., periglacial and slope) processes that can operate within or adjacent to glaciated environments. Preservation is particularly low in upland areas but, by contrast, areas that lie to the periphery of major ice masses tend to act as sinks with improved preservation of both sediments and landforms. Despite this, even in these peripheral areas, glacial sequences are highly fragmented and disparate in appearance, reflecting the range of active processes that form and degrade them. The concepts of both time and space within geology are probably the most important yet chal- lenging aspects of geology to rationalize and are also highly relevant to glacial geology. Episodes of glaciation during the Earth’s history have occurred back as far as the Neproterozoic some 750 million ago. However, the geological time units (e.g., thousands of years to hundreds of millions of years) employed for much of the geological timescale (see Section 11.3.2)areincompre- hensible to us as humans because often we have no standard point of reference or a limited context. Understanding the relative speed and duration of geological processes or specific events that are pre- served within the geological record, provides geologists with that crucial reference point or context. This can in-part be achieved using geochronology, but also through the study of contemporary geo- logical (e.g., glacial) environments which can inform us of the rates and scales at which processes operate at. Some processes occur at very slow rates but may persist over a relatively long period of time. A prime example is the slow but steady accumulation of silts and clays in a glaciolacustrine basin. At the other end of the spectrum are catastrophic processes that occur rapidly, have a 11.2 GEOLOGICAL COMPLEXITIES IN GLACIAL SEQUENCES 379

3 2 8 5 1 27–26 ka 19 ka 7 2 16 ka 17 ka 1 8 17 ka 9 19 ka 16 ka 19 ka 7 6

16 ka 5 19 ka 3 16 ka 16 ka 1 2 17 ka 5 23–21 ka 2 4 19 ka 3 27 ka ? 5 4 16 ka 17 ka Be

ka 1 19 ka Be Be 19 ka 19 21 17 ka 16–15 ka 17–16 ka Be – to 17 ka 25 Outer limit of FIS and BIIS Be Retreat margins of FIS and BIIS

23

– 20 ka These numbers refer to marinal positions of the 21 ka Ages on ice limits in calendar years 1 southeastern margin of the FIS and are referred to in Kalm (2012) Be Beryllium 10 ages

FIGURE 11.2 Age estimates on the maximum extent and retreat positions of British Irish Ice Sheet and Fennoscandian Ice Sheet during the Last Glaciation (MIS 2) illustrating the time-transgressive and disparate nature of glacial systems. Note that the maximum spatial extent of these ice sheets varies temporally from between 21 and 26À27 ka. From Bo¨se et al., 2012. glaciations of northern . Quat. Sci. Rev. 44, 1À25.

significant geomorphic impact, but only last a short period of time—e.g., a dam failure and resulting outburst flood (Clague and Evans, 2000). Equally significant are phases of nondeposition, which are called a hiatus, which whilst not unique to glacial environments are nevertheless common. The occurrence of hiatuses reflects localized changes in sediment supply (e.g., cessation of sediment input into a small lake basin) or geological process (e.g., ice- decoupling); or much broader changes in palaeogeography (e.g., exposure of a subglacial surface during ). A spatial understanding of glacial processes is also important. This is well-illustrated by the concept of the so-called ‘’ (LGM)—the time during the last () when ice sheets reached their greatest extent (see chapter: Quaternary Glaciations and Chronology). Global ice volume is traditionally defined as peaking approximately 21,000 years ago (Clark and Mix, 2002). However, Northern Hemisphere ice sheets and sectors of individual ice sheets reached their maximum extent at different times (Fig. 11.2). For example, the British-Irish Ice Sheet reached its maximum extent at 27 ka BP, although the Irish and North Sea Ice Streams reached their maximum limits much later (Clark et al., 2012). Similar patterns can also be recog- nized around the southern sectors of the Fennoscandian Ice Sheet (Houmark-Nielsen and Kjær, 2003; Bo¨se et al., 2012; Kalm, 2012; Marks, 2012; Rinterknecht et al., 2012) and the margins of the Laurentide Ice Sheet (Dyke et al., 2002). Whilst highlighting the caution required when using the LGM term (Bo¨se et al., 2012), this example also draws attention to the asynchronous nature of many glacial systems and their coupling to the ‘global system’. Thus, conventional stratigraphic 380 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

correlation methods based upon ‘counting-up’/‘lithological similarity’ (lithostratigraphy) or ‘land- form preservation’/‘maximum extent’ (morphostratigraphy) are inherently weak and overlook independence (both within and between glacial systems) and the role of local processes, influences, and responses as controls within the glacial system (Rose and Menzies, 1996; Benn and Evans, 2010; Bo¨se et al., 2012). From the above outline, it can be seen that there are many factors that influence both the preser- vation and nature of glacial sequences. Within Section 11.3 the concept of ‘glacial lithofacies’, a key technique in describing and interpreting glacial sequences and one of the main the building blocks of stratigraphy, is introduced.

11.3 GLACIAL LITHOFACIES 11.3.1 INTRODUCTION Glacial sediments can be deposited (or tectonically accreted) in a range of different settings that may be defined by their geomorphology (e.g., glaciolacustrine, glaciofluvial) and position relative to a body of ice (e.g., subglacial, englacial, supraglacial, ice-marginal, and proglacial). Reconstructing these settings or depositional environments is the ultimate goal of glacial geology because it under- pins glacial stratigraphy, palaeoenvironmental interpretation and our understanding of the evolution and dynamics of glacial systems. However, reconstructing depositional environments from the sedi- mentary record can, in glacial systems, be a complicated task. Accordingly, glacial geologists fre- quently adopt a or lithofacies approach to describing and interpreting the sedimentary record. The term facies refers simply to the sum of the characteristics that make up a rock or sediment. This can encompass a variety of descriptive properties including the geometry of the rock or sediment body, its composition, the range of sedimentary structure and distribution of particle size, colour, and fos- sil content. The facies name given to a sediment or rock typically corresponds to either a dominant or distinctive property that allows a facies to be differentiated from an adjacent facies. Therefore, not all of the properties of a facies are given in its facies name, but should be detailed systematically within the facies description. Facies can also be subdivided into different components. Lithofacies,a term commonly used by glacial geologists, refers to the physical and chemical properties of a sedi- ment or rock. Biofacies, by contrast, relates to the floral and faunal content of a sediment or rock, and is unlikely to be used widely within glacial sediments with the possible exception of some glaciolacustrine or glaciomarine sequences. Ichnofacies are a type of biofacies that describe the presence of trace and may also be present is some basinal sequences. Building a robust lithofacies model based upon descriptive sedimentological and lithological data underpins all higher-level stratigraphic, process, and palaeoenvironmental interpretations of glacial deposits. Within Section 11.3.2 the range of methods and techniques that can be used to describe and characterize glacial lithofacies are outlined.

11.3.2 DESCRIBING AND CHARACTERIZING GLACIAL SEDIMENTS Lithofacies are defined and delineated by establishing: (1) their and structure (both sedimen- tary and glaciotectonic); and (2) their geometry and geometric relationship to adjacent lithofacies and 11.3 GLACIAL LITHOFACIES 381

the land surface. Key descriptive properties include a wide range of lithological, sedimentological, and structural proxies that can be described either in the field or within the laboratory (Table 11.1). Basic field descriptions may include texture, colour (Munsell colour value), density and consolidation, bedding and facies thickness, nature of lower and upper contacts, structures present, and directional measurements (e.g., palaeocurrents, clast fabric analysis, microstructures and fabrics, tectonic struc- tures). Supplementary data obtained via laboratory-based analyses can also provide valuable

Table 11.1 A Checklist Showing Basic Health and Safety and Field Equipment Requirements for Fieldwork—Specific Details Will Vary According to the Field Area and Scope of the Project Health and Safety • Hi-visibility clothing, hard hat, suitable footwear, GPS, and communication (e.g., mobile phone) • Food, drink, and appropriate field clothing for weather conditions and terrain • Awareness of local hazards (e.g., tides, topography, quarry machinery, etc.) Field Equipment • GPS and map: for location • Field notebook, pens, and pencils: for recording observations • Tape measure, scale bar, or surveying equipment: for measuring thickness and geometry • Compass clinometer: for orientation and structural measurements • Geological hammer: for breaking rocks/clasts • Spade and trowel: for cleaning unconsolidated sections • Munsell Colour Chart: for describing colour of rocks and unlithified sediment • Dilute hydrochloric acid: testing for carbonate content within rocks and unlithified sediment • Sample bags and marker pen: for collecting samples Describing a Field Section • Lithology • Sediment texture • Sediment colour (weathered and unweathered) • Thickness and geometry of facies • Sediment and clast composition • Sedimentary structure including directional measurements • Tectonic structure including style of deformation and geometry, way-up, etc. • and trace-fossil content • Clast fabric analysis Laboratory Analytical Techniques • Particle size analysis • Clast lithological analysis • Trace element geochemistry • Thin-section micromorphology • Clast morphology and roundness • Heavy mineral analysis

A further checklist is also shown detailing the types of data that should be systematically collected when describing a field section plus some of the main laboratory analytical techniques that can employed. 382 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

quantitative or qualitative detail. Commonly employed techniques include clast morphology, particle size analysis, and lithological/provenance analyses using clast content, heavy minerals, and geochem- istry. Thin-section micromorphology is another technique that can provide valuable descriptive infor- mation and aid genetic interpretation. It is particularly useful applied alongside macroscale studies for examining diamicton facies and deformed sediments that occur widely within glacial systems and may be generated by a variety of sedimentary, gravitational, and tectonic processes (Menzies and Maltman, 1992; Meer, 1993; Carr, 2001b; Lachniet et al., 2001; Phillips, 2006; Meer and Menzies, 2011; Phillips et al., 2013). Ultimately, the range of techniques deployed will depend on the nature of the glacial sediments plus the scope and resources of the research project. Further information on these methods is beyond the scope of this chapter and readers are directed to a comprehensive range of textbooks that provide relevant technical guidance (Goudie, 1990; Jones et al., 1999; Evans and Benn, 2004; Hubbard and Glasser, 2005; Phillips et al., 2011b; Gale and Hoare, 2012). Describing sediments in a systematic and objective way is central to constructing a robust litho- facies framework. However, the volume and variety of data employed can make visualizing and communicating these data in an effective manner challenging. To overcome this, sedimentologists have developed a shorthand scheme (commonly referred to as lithofacies codes), for the basic description of lithofacies. The lithofacies concept was originally introduced for describing braided river deposits (Miall, 1977, 1978) and subsequently extended to include diamicton facies (Eyles et al., 1983) and other facies found in glacial environments (Maizels, 1993; Benn and Evans, 1998; Evans and Benn, 2004)(Table 11.2). Lithofacies codes typically contain two descriptive qualifiers. Firstly, an upper-case qualifier describing modal lithology, namely: D, diamicton; B, boulders; G, gravel; GR, granules; S, sands; F, fines (silt and clay). The second part of a lithofacies code employs a lower-case qualifier that describes the internal structure. For example, the lithofacies code ‘Gm’ corresponds to ‘...gravel, massive...’, whilst Dms refers to ‘...diamicton, matrix- supported, stratified...’. The widespread application of lithofacies code schemes to glacial sedi- mentology demonstrates the value and relative simplicity of the approach. However, lithofacies nomenclature has yet to be standardized and many personal variants of coding schemes exist (Bridge, 1993). Equally, a suitable coding scheme that portrays the complex range of diamicton facies and associated processes has proven elusive (Benn and Evans, 1998). Displaying lithofacies data is commonly done by using either a sedimentary log/vertical profile or a section diagram. Sedimentary logs or vertical profiles (measured or composite) are a quick and relatively simple visual method for summarizing the vertical distribution of lithofa- cies and are particularly suitable for compilation in the field and describing cores and boreholes (Figs. 11.3 and 11.4). A sedimentary log commonly shows lithofacies thickness, modal grain size (scaled to the width of the polygon) lithology, the geometry and type of upper and lower lithofacies contacts, and finally, a lithofacies code (Fig. 11.4). Supplementary data, including directional measurements and quantitative lithological data, can also be shown in adjacent panels (Fig. 11.5). Whilst popular, the principal weakness of sedimentary logs is that little information on lateral bedding and facies geometry, sediment heterogeneity, and facies architec- ture can be depicted. To overcome this, an alternative approach is to use a section diagram. A section diagram comprises either a field sketch or a measured drawing of a field section and can similarly incorporate lithofacies codes together with various lithological and directional data (Fig. 11.6). 11.3 GLACIAL LITHOFACIES 383

Table 11.2 Lithofacies Short-Hand Codes Used for Recording the Lithology and Structure of Sediments With Field Sections or Boreholes Code Lithofacies Description Code Lithofacies Description

Poorly Sorted Sediment Particles of 0.063À2mm Diamictons (D) Admixture Sands (S) Diameter Dmm Matrix-supported, massive St Trough cross-bedding Dms Matrix-supported, stratified Sp Planar cross-bedding Dml Matrix-supported, laminated Sr (A) Type-A ripples Dcm Clast-supported, massive Sr (B) Type-B ripples Dcs Clast-supported, stratified Sr (S) Type-S ripples (c) Evidence of current reworking Scr Climbing ripples (r) Evidence of re-sedimentation Sh Horizontally-laminated (s) Evidence of shearing Sl Horizontal lamination with mud drapes (l) Evidence of loading Sfo Deltaic foresets (p) Includes clast pavements(s) Sfl Flaser bedding (i) Includes soft-sediment Se Erosional scours with intraclasts inclusions (ra) Includes glacitectonic rafts Sm Massive Sfu Fining-upwards Boulders (B) Particles .256 mm diameter Scu Coarsening-upwards Bms Matrix-supported, massive (d) With dropstones Bmg Matrix-supported, graded (l) With loading Bcm Clast-supported, massive (e) With syn-depositional extensional faults Bcg Clast-supported, graded (s) With shears Bfo Deltaic foresets BL Boulder lag or pavement Silts and clays (F) Particles less than 0.063 mm diameter Fl Finely laminated Gravels (G) Particles of 8À256 mm Fm Massive diameter Gms Matrix-supported, massive Fp Lenticular bedding Gm Clast-supported, massive Flv Fine lamination with rhythmites or varves Gsi Clast-supported, imbricated (d) With dropstones Gmi Clast-supported, massive, (l) With loading imbricated Gfo Deltaic foresets (e) With syn-depositional extensional faults Gh Horizontally bedded Gt Trough cross-bedding Gp Planar cross-bedding Gz Large-scale sinusoidal bedding (Continued) 384 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Table 11.2 Lithofacies Short-Hand Codes Used for Recording the Lithology and Structure of Sediments With Field Sections or Boreholes Continued Code Lithofacies Description Code Lithofacies Description

Poorly Sorted Sediment Particles of 0.063À2mm Diamictons (D) Admixture Sands (S) Diameter Gfu Fining-upwards Gcu Coarsening-upwards Go Open framework Gd Deformed bedding Glg Bedload lag Granules (GR) Particles of 2À8mm diameter GRcl Massive with clay laminae GRch Massive and infilling channels GRh Horizontally bedded GRm Massive GRmb Massive and pseudobedding GRmc Massive with isolated outsize clasts GRmi Massive with isolated, imbricate clasts GRmp Massive with pebble stringers GRo Open framework GRz Large-scale sinusoidal bedding GRfu Fining-upwards GRcu Coarsening-upwards GRt Trough cross-bedding GRp Planar cross-bedding GRfo Deltaic foresets

Modified from Benn, D.I., Evans, D.J.A., 2010. and Glaciation, second ed. Hodder Education, Abingdon.

11.3.3 FACIES ANALYSIS, ASSOCIATIONS, AND SEQUENCES The lithofacies concept is a robust technique that can be used to describe, define, and delineate sediments leading to their genetic interpretation (Fig. 11.7). This interpretative component is called facies analysis and is a three-part process encompassing: (1) a process interpretation of individual lithofacies identifying modes of transportation and ; (2) identifying facies associations that enable adjacent lithofacies to be genetically linked and a palaeoenvironmental interpretation to be made; (3) recognizing patterns or cycles of facies associations enabling a facies sequence to be established. Identification of facies sequences can be achieved by either visual interpretation of a sedimentary log or by employing a statistical technique such as Markov chain analysis that can recognize cyclical patterns in data (Mack and James, 1986; Brierley, 1989). 11.3 GLACIAL LITHOFACIES 385

Locality: Log Number:

Date: Operator:

PARTICLE SIZE & STRUCTURE

SCALE NOTES silt f.sand m.sand diamicton c.sand boulders clay granules gravel LITHOLOGY DIRECTIONAL DATA LITHOFACIES CODE

FIGURE 11.3 Blank template for a sedimentary log/vertical profile. 386 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Lithology

Sand & Diamicton Boulders Gravel Gravel

Out-sized Sand Silt Clay clasts

Sedimentary and tectonic structures

Horizontal Trough Planar Massive bedding/ cross-bedding cross-bedding lamination

Current Wave Flaser Lenticular ripples ripples bedding bedding

Load Dewatering Imbrication structures structures Dropstone

r Reverse n Extensional Fold nose Folding fault faulting

b hf Slump Boudinage Intraclasts Hydrofracture

Basal contacts

Sharp and Gradational Loaded Intercalated conformable

Faulted Erosional

FIGURE 11.4 Descriptive symbols employed to summarize lithology, sedimentary and tectonic structure, and basal contacts within a sedimentary log/vertical profile. (A) Log A Log B Log C Log D Log E

(B)

FIGURE 11.5 (A) Vertical profile logs from southwest Ireland show the distribution of diamicton and sorted lithofacies. (B) Sedimentary logs and palaeocurrent data from Jokulhlaup Sandur deposits, south Iceland. (C) An example of logged borehole core from Norfolk, UK, showing Middle Pleistocene glacial sediments (blue) overlying preglacial shallow marine and fluvial (beige) sediments with quantitative particle size, calcium carbonate, and stone count data. (A) From O´ Cofaigh et al. (2011) Formation of a stratified subglacial ‘till’ assemblage by ice-marginal thrusting and glacier overriding. Boreas, 40, 1À14. (B) From Maizels, J., 1993. Lithofacies variation within sandur deposits: the role of runoff regime, flow dynamics and sediment supply characteristics. Sediment. Geol. 85, 299À325. (C) From Rose et al., 2002. Early and early Middle Pleistocene river, coastal and neotectonic processes, southeast Norfolk, England. Proc. Geol. Assoc. 113, 47À67. 388 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

(C)

THE GRANGE, LANGLEY

%Sand %Gravel 5

%Carb chert 0 0

100 50 %Rhax chert Coloured: colourless Qzt+VQ +Sch ratio

50 100 5 %Greensand chert 0 0 %Qzt+VQ +Sch 0 %Silt +Clay %CaCO3 Qzt:VQ ratio Flint: Qzt+VQ +Sch ratio CZSGP Samp No. 0050 0 50 0 2050 4050 0 5 0 5 10 5 1 2 26 Litho-facies Units 25

24 GL 23 F 22 21 2 0.00 2a 0.00 20 3x 0.00 4x 0.00 19 5x 0.00 18 17 GL 16 E 7 0.00 15 8 14 8a 9 13

Meters above O.D. Meters above 12 11 11 12a 13 0.00 0.00 0.00 10 GL D 14 0.00 0.00 0.00 9 8 17 7

6 GL 0.00 0.00 18x 0.00 5 B 19 0.00 0.00 0.00 4 20x 0.00 0.00 3 0.00 0.00 0.00 21 0.00 0.00 2 22 1 x GL 23 0.00 0 A 24 -1 25 -2

FIGURE 11.5 (Continued) 11.3 GLACIAL LITHOFACIES 389

(A)

6

4

2

0

0 10 20 30 40 50 60 70

80 90 100 110 120 130 140 150

160 170 180 190 200 210 220 230

Upper diamicton Lithofacies C (sand and gravel) Lithofacies B (sand and gravel)

Major bounding surfaces Palaeocurrent sample points (see Figure 6) FIGURE 11.6 (A) An example of a section diagram from Lleiniog, north Wales, UK, showing the distribution and geometry of diamicton and outwash facies. (B) A more detailed section diagram from 70 to 80 m showing the architecture of outwash sands and gravels and the development of a structure. From Lee et al., 2015. Sedimentary and structural evolution of a relict subglacial to subaerial drainage system and its hydrogeological implications: an example from Anglesey, north Wales, UK. Quat. Sci. Rev., 109, 88À110.

However, the genetic interpretation of lithofacies is seldom straightforward because many litho- facies can be deposited by similar mechanisms that occur across a range of different sedimentary environments. For example, a lithofacies comprising ‘planar cross-bedded sandstone’ can be depos- ited by either water-driven or aeolian mechanisms in a range of palaeoenvironments (e.g., desert, deltaic, fluvial, tidal flats, shallow marine, etc.). Thus, genetic qualifiers (e.g., ‘fluvial’ sandstone) should only be used in a facies name providing its palaeoenvironmental affinity is known with rea- sonable certainty. As a default, or where a palaeoenvironment interpretation is subjective, facies names should simply be descriptive (e.g., cross-bedded sandstone). Glacial environments perhaps embody the challenge of genetic interpretation more than many other palaeoenvironment. As previously discussed, this is because the range of factors that can influence both sedimentation and preservation are considerable. Extreme care and descriptive data from a range of different techniques are therefore required to interpret glacial sediments in a robust and objective way. Potential interpretive bias can be introduced by the range of analytical 390 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

(B)

Irish Sea till CF1 to CF3 - fracture event

Lithofacies C Obscured Fracture

Lithofacies B Downthrow

Obscured

Basin 2 m Basin 1 4 Synclinal folding

CF2 Flame-like Disharmonic CF2 Fold axis 3 & flame-like contacts CF3 CF2 bedding Normal faulting

2 CF1 CF3 CF1 Reverse faulting 1 CF3

0 CF3 Sinusoidal bedform Major bounding surface Stratified infill Planar & trough cross-bedding, Sand-filled fracture South including back-sets North

FIGURE 11.6 (Continued) techniques employed, the relative skills and experience of the researcher(s) and evolving conceptual trends in glacial geology (Benn and Evans, 2010). A classic example corresponds to debates sur- rounding the genetic interpretation of stratified diamicton lithofacies at Scarborough Bluffs, Lake Ontario, Canada (Evenson et al., 1977; Gibbard, 1980; May et al., 1980; Dreimanis, 1982; Hicock, 1992; Boyce and Eyles, 2000; Dreimanis and Gibbard, 2005); northern East Anglia in the United Kingdom (Eyles et al., 1989; Hart and Boulton, 1991b; Hart and Roberts, 1994; Roberts and Hart, 2005; Phillips et al., 2008); and the margins of the Irish Sea in the United Kingdom and Ireland (McCabe, 1987; Eyles and McCabe, 1989; McCarroll and Harris, 1992; O´ Cofaigh and Evans, 2001). In each of these localities, a combination of different palaeoenvironmental models (e.g., subglacial or subaqueous) and depositional mechanisms (e.g., debris flows, subaqueous rain-out, subglacial lodgement, and deformation) have been applied—each having major implications for the wider glaciological and palaeogeographical context. Despite the challenges of genetic interpreta- tion, the lithofacies approach provides a highly accessible and applicable approach to reconstructing past glacial environments and the range of different glacial lithofacies that may be encountered are reviewed in Chapters 4À10.

11.3.4 LANDSYSTEMS The value of a lithofacies approach for reconstructing glacial environments has been outlined above and is demonstrated by the vast range of published studies that utilize the technique. However, used 11.3 GLACIAL LITHOFACIES 391

e.g., herring-bone cross-bedding— Facies foresets dipping in opposite description directions

e.g., high-energy sedimentation, Facies process cyclic switches in palaeocurrent Interpretation direction

e.g., association to tidal rhythmites, Facies association flaser and lenticular bedding–tidal sedimentation within an estuary

e.g., cycles of tidal sedimentation punctuated by beach shore-face Facies sequence deposits (transgressive–regressive sequences)

FIGURE 11.7 A facies analysis flow chart showing the various descriptive and interpretative stages of investigation.

in isolation, lithofacies provide a somewhat restricted angle on reinterpreting past episodes of glacia- tion. The surface-form of lithofacies, where evident, provides an additional geomorphological context which can add significantly to a genetic and palaeoglaciological understanding. Associations of litho- facies and landforms are called sediment-landform associations and underpin the concept of glacial landsystems (Eyles, 1983a; Evans, 2003b). The glacial landsystems approach was initially conceived to characterize and subdivide glaciated terrains for engineering purposes based upon surface morphol- ogy and geological processes (Fookes et al., 1978) with subsequent incorporation of geology (Eyles, 1983a,b; Eyles and Menzies, 1983; Brodzikowski and van Loon, 1987), glacier morphology, and gla- cier dynamics into the approach (Benn and Evans, 1998, 2010; Evans, 2003a). ‘Landsystems’ are defined ‘... as areas of common terrain attributes, different from those of adjacent areas, in which recurring patterns of topography, soils and vegetation reflect the underlying geology, past erosional and depositional processes and climate...’(Benn and Evans, 2010, p. 584). They serve a similar role to the ‘facies sequence’ described above and can be further subdivided into land elements (cf. facies) and in turn land facets (cf. facies association). A land element, the smaller subdivision, comprises individual glacial landforms such as an or a . By contrast, a land facet represents a of land elements (landforms) that form a uni- form and continuous topographic expression—for instance an or drumlin field. The principal advantages of glacial landsystems are that they promote an integrated, multidisci- plinary, and hierarchical approach to glacial geology (Fig. 11.8). This approach links lithofacies and landforms to glacial processes and ultimately to palaeoglaciology at various temporal and spa- tial scales (Benn and Evans, 2010). The development of these so-called ‘process-form’ models has, over the past few decades, greatly enhanced our understanding of the dynamics of glaciation in 392 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Crevasse fill ridges Drumlin Flutes Concertina Dead ice esker Subglacial Overriden end moraine Blow-out till Pitted structure outwash Hummocky moraine Glacitectonic end moraine

FIGURE 11.8 An example of a glacial landsystem model from Eyjabakkajo¨kull, Iceland. From Schomacker et al., 2014. The Eyjabakkajo¨kull glacial landsystem, Iceland: geomorphic impact of multiple surges. Geomorphology 218, 98À107. both recently deglaciated areas (Sharp, 1985; Kjær and Kru¨ger, 2001; Evans and Twigg, 2002; Spedding and Evans, 2002; Schomacker and Kjær, 2008; Bennett and Evans, 2012) and relict glacial systems (Clayton and Moran, 1974; Boulton and Clark, 1990; Evans et al., 1999; Benn and Lukas, 2006; Kehew et al., 2012; Lee et al., 2013; Evans et al., 2014).

11.4 STRATIGRAPHY IN GLACIATED ENVIRONMENTS 11.4.1 INTRODUCTION Within Section 11.3, the concept of lithofacies, its practical application and role within the concept of glacial landsystems were explored. At a basic level, lithofacies provide a geological framework by which the properties of a rock or sediment can be characterized and a process-based interpretation derived. Understanding how geology and geomorphology evolve in time and space is called stratigraphy. It describes the sequential order that different rocks, sediments, or landforms were formed, enabling wider insight into the processes that formed them, the evolution of palaeoenvironments, relative timing of events, and age. Glacial stratigraphy encompasses this broad remit providing ‘...the methods and patterns by which we can reconstruct the history and patterns of past glaciations and environments...’ (Rose and Menzies, 1996, p. 253). Of equal significance, is the important applied role that stratigraphy plays in understanding the distribution of resources (e.g., minerals, hydrocarbons, and groundwater) and hazards (e.g., ground stability, running sand) in formerly glaciated terrains. 11.4 STRATIGRAPHY IN GLACIATED ENVIRONMENTS 393

INTERNATIONAL CHRONOSTRATIGRAPHIC CHART

www.stratigraphy.org International Commission on Stratigraphy v 2015/01

P

P

P

P

A

S

S

S S

numerical numerical numerical S numerical

S

S

S S

Series / / Age age (Ma) Series / Epoch Stage / Age Series / Epoch Stage / Age Erathem / Era System / Period S

G

G

G G EonothemErathem / Eon System / Era / Period EonothemErathem / Eon System/ Era / Period age (Ma) EonothemErathem / Eon System/ Era / Period age (Ma) / Eon G age (Ma) present ~ 145.0 358.9 ± 0.4 ~ 541.0 ±1.0 0.0117 Upper 152.1 ±0.9 ~ 635 0.126 Upper Neo- Middle 157.3 ±1.0 Upper ~ 720 Pleistocene 0.781 372.2 ±1.6 1.80 163.5 ±1.0 1000 166.1 ±1.2 Quaternary 2.58 382.7 ±1.6 168.3 ±1.3 Middle 1200 3.600 170.3 ±1.4 Middle 387.7 ±0.8 Meso- proterozoic 5.333 174.1 ±1.0 1400

Messinian 393.3 ±1.2 7.246 182.7 ±0.7 1600 11.63 Lower 190.8 ±1.0 Lower 407.6 ±2.6 13.82 1800 410.8 ±2.8 Proterozoic 15.97 199.3 ±0.3 Palaeo- 2050 201.3 ±0.2 419.2 ±3.2 proterozoic

20.44 Pridoli Rhyacian 423.0 ±2.3 23.03 ~ 208.5 2300 Ludlow 425.6 ±0.9 28.1 Upper 427.4 ±0.5 2500 Wenlock 430.5 ±0.7 Neo- 33.9 ~ 227 433.4 ±0.8 2800 37.8 ~ 237 Llandovery 438.5 ±1.1 Meso-

Bartonian 41.2 440.8 ±1.2 archean Middle ~ 242 443.8 ±1.5 3200 47.8 247.2 445.2 ±1.4 Palaeo- 251.2 Lower Archean archean 252.17 ±0.06 Upper Palaeogene 56.0 453.0 ±0.7 3600 254.14 ±0.07 59.2 Eo-

Wuchiapingian Palaeozoic Palaeocene 259.8 ±0.4 458.4 ±0.9 archean Phanerozoic 61.6 Phanerozoic 4000 265.1 ±0.4 Middle 66.0 467.3 ±1.1 268.8 ±0.5 470.0 ±1.4

Roadian ~ 4600 72.1 ±0.2 272.3 ±0.5 Lower 477.7 ±1.4 Units of all ranks are in the process of being defined by Global 283.5 ±0.6 Boundary Stratotype Section and Points (GSSP) for their lower 83.6 ±0.2 boundaries, including those of the Archean and Proterozoic, long 485.4 ±1.9 Upper 86.3 ±0.5 defined by Global Standard Stratigraphic Ages (GSSA). Charts and 290.1 ±0.26 Stage 10 detailed information on ratified GSSPs are available at the website ~ 489.5 89.8 ±0.3 http://www.stratigraphy.org. The URL to this chart is found below. 295.0 ±0.18 ~ 494 93.9 298.9 ±0.15 Numerical ages are subject to revision and do not define units in ~ 497 the Phanerozoic and the Ediacaran; only GSSPs do. For boundaries in the Phanerozoic without ratified GSSPs or without constrained 100.5 Upper 303.7 ±0.1 ~ 500.5 307.0 ±0.1 numerical ages, an approximate numerical age (~) is provided. Palaeozoic Series 3 Middle Moscovian ~ 504.5 Numerical ages for all systems except Lower Pleistocene, ~ 113.0 315.2 ±0.2 Stage 5

Mesozoic Permian,Triassic, and Precambrian are taken from

Cretaceous ~ 509 ‘A 2012’ by Gradstein et al. (2012); Lower Stage 4 323.2 ±0.4 those for the Lower Pleistocene, Permian, Triassic and Cretaceous ~ 125.0 Series 2 ~ 514 were provided by the relevant ICS subcommissions. Upper Stage 3 Barremian 330.9 ±0.2 Lower ~ 129.4 ~ 521 Colouring follows the Commission for the Geological Map of the World (http://www.ccgm.org) ~ 132.9 Middle Visean Stage 2 346.7 ±0.4 ~ 529 Chart drafted by K.M. Cohen, S.C. Finney, P.L. Gibbard (c) International Commission on Stratigraphy, January 2015 ~ 139.8 Lower To cite: Cohen, K.M., Finney, S.C., Gibbard, P.L. & Fan, J.-X. (2013; updated) ~ 145.0 358.9 ±0.4 541.0 ±1.0 The ICS International Chronostratigraphic Chart. Episodes 36: 199-204. URL: http://www.stratigraphy.org/ICSchart/ChronostratChart2015-01.pdf FIGURE 11.9 The 2015 International Commission on Stratigraphy chronostratigraphic chart. Cohen, K.M., Finney, S.C., Gibaard, P.L., Fan, J.-X., 2013. The ICS International Chronostratigraphic Chart. Episodes 36, 199À204.

At this point, it is relevant to introduce the two primary forms of stratigraphy. Firstly, relative stra- tigraphy involves the reconstruction of the sequential order in which rocks, sediments, and landforms were generated. For instance, which sediment was deposited first, which was deposited second, which was third and so-on. The second branch of stratigraphy is called absolute stratigraphy and this involves determining the age of the rocks, sediments, and landforms relative to geological time. In glacial geol- ogy, this can be achieved by applying a range of radiometric dating techniques to glacigenic materials, or through correlation with sequences or marker horizons of known age. Radiometric dating is beyond the scope of this chapter but readers are directed to Chapter 19, Geochronology Applied to Glacial Environments, Geochronology for further information on the subject. This section of the chapter will review the principles behind glacial stratigraphy. It outlines the range of relative stratigraphic methods that can be employed in glacial stratigraphy, discussing the relative pros and cons of each approach and providing case studies of where specific approaches have been applied successfully.

11.4.2 GEOLOGICAL TIME Geological time is an integral component of stratigraphy because it provides a universal standard— 4.54 billion years’ worth of Earth history—to which events of specific ages can be correlated 394 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

(Fig. 11.9). This geological timeline is subdivided into a range of geochronological units which in turn are subdivided in decreasing order of scale from Eons (the largest subdivision of geological time), Eras, Periods, Epochs, and finally to the smallest subdivision of time called Ages. An impor- tant distinction should be made here with chronostratigraphic units which correspond to units of geological materials such as rocks, sediments, landforms, or fossils. Chronostratigraphic subdivi- sions include Systems, Series, and Stages—the latter corresponds to the smallest chronostrati- graphic subdivision and is equivalent to Ages (geochronological subdivision). Within the Quaternary, further chronological subdivision has been made possible by the devel- opment of the Oxygen Isotope Record (Fig. 11.10). This record provides a global chronological framework based upon cyclical changes in the oxygen isotope composition of ice cores and deep marine sediments. Isotopic fractionation of oxygen into heavier and lighter isotopes is temperature- dependent so the proxy records cyclical changes in global temperature and a relative indicator of global ice volume and sea-level. Other more regional chronological records that span all or parts of the Quaternary also occur, most notably the loess-palaeosol sequences from China and central Eurasia, speleotherms in arid and low-latitude areas, and continental lake records. For further infor- mation about these records and the factors that drive climate change during the Quaternary, readers are directed to Lowe and Walker (2014) which provides an excellent overview of the topic.

11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 11.5.1 LITHOSTRATIGRAPHY 11.5.1.1 Summary Lithostratigraphy is a core stratigraphic technique and has been utilized in many former glacial environments. Lithostratigraphy is defined as ‘...the description and systematic organisation of...rocks... into distinctive named units based upon the lithological character of the rocks and their stratigraphic relations’ (Salvador, 1994, p. 31). Underpinning lithological data can include a variety of quantitative and qualitative data obtained from boreholes, shallow augering and geotech- nical site investigations as well as descriptions from field sections. When used effectively, lithostra- tigraphy offers the potential to build a succession in a more spatially inclusive three-dimensional manner rather than simply following a linear route (e.g., a series of coastal cliff sections). This lends itself to geological mapping and the approach is routinely employed by many Geological Survey organizations for surveying partly buried glaciated sequences or those that possess a degraded or subtle surface expression (McMillan and Merritt, 2012). Despite the common application of lithostratigraphy, successful use of the technique can be lim- ited by the lithological complexity of glacial sequences which are often fragmented, disparate, and heterogeneous in nature (Rose and Menzies, 1996). These characteristics are inherited from the vast range of geological processes (e.g., sedimentary, tectonic, and gravitational) and external factors (e.g., bedrock lithology, climate, groundwater availability) that operate in glacial environ- ments and occur over a variety of temporal and spatial scales. Common problems encountered include facies repetition and the often limited spatial extent of individual units. Tectonic processes, such as the overriding or pushing of preexisting strata by ice, provide a particular challenge because the products of glaciotectonism don’t often obey standard lithostratigraphic rules. Common 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 395

ODP 677 Deepwater 18O

Age (Ma) Palaeomag. 5 4 3 Holocene 0 2

BLA 5e 0.12 6 0.2 7 8

10 9 11 0.4 12 BRU 14

0.6 16

18 0.8 0.78 20 19 22

1.0 0.99 JAR 30 1.07 34 COB 1.2 1.19 36

1.4 QUATERNARY PLEISTOCENE MAT

54 1.6 GIL 1.65

62 1.78 1.8 64 OLD 68 1.95 2.0

REU 2.13

2.2 Gelasian Calabrian Ionian Taran.

2.4

2.6 2.58 GAU Piacen. NEOGENE PLIOCENE

FIGURE 11.10 The marine oxygen isotope record for the Quaternary showing the major (odd numbers) and glacial (even numbers) stages. Oxygen isotope data from Lisiecki, L.E., Raymo, M.E., 2007. Plio-Pleistocene climate evolution: trends and transitions in glacial cycle dynamics. Quat. Sci. Rev. 26, 5669. 396 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

products of glaciotectonism include the widespread development of angular bounding surfaces and the structural reordering of units either en bloc or by soft-sediment mixing (Banham, 1977; Aber et al., 1993; Larson et al., 2003; Burke et al., 2009).

11.5.1.2 Characterizing and building a lithostratigraphic framework The primary building blocks of a lithostratigraphic succession are called lithostratigraphic units. A lithostratigraphic unit is defined as a volume of rock or sediment of identifiable genetic origin that possesses an observable set of lithological, structural (sedimentary), geometric, petrographic, or palaeontological characteristics. Lithostratigraphy assumes that strata are the correct ‘way-up’ and that from bottom to top units within a succession will get progressively younger (Fig. 11.11A). This rule is called the and is a core principle of lithostratigraphy. Although widely applicable in areas that possess a relative simple layer-cake stratigraphy care should be taken where the contacts between units are unconformable (Fig. 11.11B and C) or where complex glaciotectonic deformation has occurred (Slater, 1926; Banham, 1975; Berthelsen, 1978; Aber and Ber, 2007; Lee et al., 2017). The lithology and structure of an individual rock or sediment (i.e., a lithofacies) can be deter- mined either quantitatively or qualitatively using a range of field and laboratory methods (see Section 11.3.2). Lithostratigraphy employs a hierarchical framework of nomenclature that defines lithostratigraphic units and these range in scale from ‘Supergroup’ (largest), ‘Group’, ‘Formation’, ‘Member’, to ‘Bed’ (smallest). The basic lithostratigraphic unit is the ‘Formation’ and this should be clearly defined based upon its lithological properties, geometry, and stratigraphic position. It should also be laterally traceable, being mappable at surface or at depth within the sub- surface. Division of a ‘Formation’ into several ‘Members’ enables more localized lithological detail to be classified. For example, a ‘Formation’ may comprise several different (superpositionally or lithologically) units of diamictons, or subdivision of an outwash sequence into sand-dominant members and gravel-dominant members. A ‘bed’ is the smallest lithostratigraphic subdivision but tends only to be named formally if they have specific stratigraphic, genetic, or applied significance.

(A) (B) (C) 3

2 Way-up 1

FIGURE 11.11 (A) Schematic diagram illustrating the ‘Law of Superposition’ with successive deposition (1À3) of units to form a ‘layer-cake’ sequence. (B) Geometry of an angular showing folding and erosion prior to deposition. (C) A disconformity showing several breaks in deposition and erosion within a sequence. 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 397

11.5.1.3 Application Lithostratigraphy is a widely employed stratigraphic technique within sedimentary successions but its sole application within glacial sequences is more limited as workers have increasingly recog- nized some of the problems that can arise. Within the Great Lakes region of North America, e.g., lithostratigraphy has traditionally under- pinned stratigraphic understanding of the succession deposited by several ice lobes along the south- ern sector of the Laurentide Ice Sheet (Mickelson et al., 1983; Karrow, 1984). The scheme by Karrow (1984) depicts a relatively straightforward, laterally extensive, layer-cake stratigraphy with diamicton units traced spatially on the basis of stratigraphic ordering, composition, and sediment dispersal patterns (Fig. 11.12). In reality, subsequent work has painted a more complex and dynamic glaciological picture for the southern sector of the Laurentide Ice Sheet with different ice

TIME HURON - GEORGIAN BAY LOBE L. ERIE TORONTO ST LAWRENCE VALLEY

TWO CREEKS 12 ka INT.

PT. HURON ST. St Joseph Halton Till Till MACKINAW 13 ka ST. Rannock Till Wentworth Till

Elma Till PORT Wartburg Till BRUCE Stratford Till ST. LATE WISCONSIN LATE Mornington Till Port Stanley Till Tavistock Till

Stirton Till Maryhill Till Gentilly Till ERIE INT.

NISSOURI Catfish Creek Till Catfish Creek Till ST.

PLUM PT. 25 ka INT. THORNCLIFFE CHERRYTREE FORMATION Meadowcliffe Till ST. ? Seminary Till PORT DUNWICH TALBOT 42–54 ka INT. T.

GUILDWOOD ST. ? Sunnybrook Till

ST PIERRE INT. 80 ka POTTERY RD FORMATION ST PIERRE FORMATION

NICOLET ST. Becancour Till

EARLY WISCONSINEARLY MID WISCONSIN SCARBOROUGH FORMATION

SANGAMONIAN 125 ka DON FORMATION

FIGURE 11.12 The glacial lithostratigraphy of the region around Lake Huron and the St Lawrence Valley along the southern flank of the Laurentide Ice Sheet. From Karrow, P., 1984. Quaternary stratigraphy and history, Great Lakes-St. Lawrence region. Quaternary stratigraphy of Canada—a Canadian contribution to IGCP Project 24, 84-10. 398 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

lobes acting independently of one another in response to both internal and external mechanisms (Karrow et al., 2000; Cutler et al., 2001; Dyke et al., 2002; Colgan et al., 2005). In the United Kingdom, lithostratigraphy has been widely applied to Middle and age glacial deposits that crop-out over 30% of the UK land area. The first formal lithos- tratigraphic framework (encompassing all of the UK Quaternary) was published by Bowen (1999) but superseded by the scheme of McMillan and Merritt (2012). This latter scheme is a mapping- based lithostratigraphic framework with nomenclature used on all new published British Geological Survey maps and models and is also used to underpin many applied thematic datasets (e.g., geoha- zards, hydrogeology, etc.). Within the scheme, glaciogenic deposits are subdivided based upon age (Group) depending on Anglian (MIS 12) or Devensian (MIS 4-2) age; ice-flow trajectory and provenance (Sub-Group); and finally, relative superposition and lithology (Formation) (Fig. 11.13). The resulting stratigraphic framework is both systematic and thorough but, by necessity, caries a vast array of stratigraphical terms and nomenclature. At a more regional scale, the Middle Pleistocene glacial deposits of East Anglia, UK, offer a classic example of attempts to apply lithostratigraphy to a glacial succession. Glacial deposits in East Anglia form arguably some of the most extensively studied and controversial sequences in with ongoing debates surrounding chronology, provenance, and the relative arrangement of the geological units (Preece et al., 2009; Rose, 2009; Lee et al., 2011). The region is widely believed to have been glaciated by oscillating lobes of North Sea and Central England ice resulting in the accretion of an extensive and thick sequence of diamicton and outwash units. Historically, the stratigraphic framework in East Anglia has been underpinned by lithostratigraphy with several schemes and palaeogeographic reconstructions published (Reid, 1882; Banham, 1968; Ehlers and Gibbard, 1991; Hart and Boulton, 1991a; Lunkka, 1994; Hamblin et al., 2005). A consensus view has, however, proven elusive (Rose, 2009). The reasons for this are not unique to eastern England but generic issues that relate to the application of lithostratigraphy to glacial ter- rains. Key factors include: (1) the reliance upon ‘linear route’ (e.g., coastal cliff sections) and point observations (e.g., quarries) rather than a fully integrated 3D geological framework; (2) the occur- rence of lithologically identical till units at multiple stratigraphic levels; (3) limited regional-scale stratigraphic continuity; and (4) limited genetic understanding of major till units and glaciotectonic complexes. Attempts to reconcile the stratigraphic framework have incorporated detailed sediment provenancing and geological mapping (Fig. 11.14)(Lee et al., 2004; Hamblin et al., 2005; Pawley, 2006; Rose, 2009). This has resolved some, but not all, of the stratigraphic issues within the region (Lee et al., 2017). More recent stratigraphic refinements have involved the application of a process-based kinetostratigraphic approach and this has revealed several additional ice advances and smaller phases of dynamic ice-marginal behaviour that have no or only a limited lithostrati- graphic signal (Lee and Phillips, 2008; Phillips et al., 2008; Waller et al., 2011; Fleming et al., 2013; Lee et al., 2013, 2017; Phillips and Lee, 2013).

11.5.2 KINETOSTRATIGRAPHY 11.5.2.1 Summary Kinetostratigraphy (also referred to as ) is a stratigraphic technique developed principally by the pioneering work of Berthelsen (1978) centred upon the intensely glaciotectonized 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 399

FIGURE 11.13 Surface distribution of Middle and Late Pleistocene glacigenic groups and subgroups within the United Kingdom. From McMillan, A.A., Merritt, J.W. 2012. A new Quaternary and Neogene Lithostratigraphical framework for and the Isle of Man. Proc. Geol. Assoc. 123, 679À691. 400 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Bowen et al. (1986) Hamblin et al. (2005) Lee et al. (2015)

EASTERN EASTERN EASTERN WAVENEY NE WAVENEY NE WAVENEY NE WESTERN FEN WESTERN FEN WESTERN FEN VALLEY NORFOLK VALLEY NORFOLK VALLEY NORFOLK FENS MARGIN FENS MARGIN FENS MARGIN Tectonics Tottenhill Tottenhill Sands & Briton’s5 Gravels5 Lane S&G Briton’s 5 Oadby Oadby Till Lane S&G Till Weybourne Weybourne Marly 4 A5 A6 Drift 4 TownTill Town Till 4

Oadby Third Till Bacton

3 A4 Cromer Till Green Bacton Lowestoft Till Till3 Green Bozeat Lowestoft Till 3 Till Till

2 Second Walcott 2 Walcott Cromer Till Till 2 Lowestoft Till Till A2 A3 Corton Corton First Norwich Brickearth 1 H’bro 1 1 Till Till Cromer Till Till 1 H’bro Till1 A1

Scandinavian ice BGF BGF - Briton’s Lane Glac. Fm Glaciotectonite WF WGF BGF SCF - Sheringham Cliffs Pennine ice Glacigenic Fm Pennine ice WGF SGF LF WGF - Wolston Glac. Fm LGF SGF North Sea ice LGF - Lowestoft Glac. Fm British North Sea ice NSDF WF - Wolston Fm LGF HGF - Happisburgh Glac. Fm LF - Lowestoft Fm HGF HGF NSDF - North Sea Drift Fm FIGURE 11.14 Various stratigraphic models for the Middle Pleistocene age glacial deposits of northern East Anglia. (A) Conventional lithostratigraphic approach (after Bowen, D.Q., Rose, J., McCabe, A.M., Sutherland, D.G., 1986. Correlation of Quaternary glaciations in England, Ireland, Scotland and Wales. Quat. Sci. Rev. 5, 299À340). (B) Modified lithostratigraphic approach encompassing geological mapping and sediment provenancing (after Hamblin, R.J.O., Moorlock, B.S.P., Rose, J., Lee, J.R., Riding, J.B., Booth, S.J., et al., 2005. Revised Pre-Devensian glacial stratigraphy in Norfolk, England, based on mapping and till provenance. Geol. Mijnbouw 84, 77À85). (C) Hybrid stratigraphic approach encompassing kinetostratigraphy and geomorphology (after Lee, J.R., Phillips, E.R., Rose, J., Vaughan-Hirsch, D.P. The Middle Pleistocene glacial evolution of northern East Anglia, UK: a dynamic tectonostratigraphicÀparasequence approach. J. Quat. Sci., in press). sequences of Denmark (Berthelsen, 1978; Aber, 1979; Houmark-Nielsen and Berthelsen, 1981). The technique differentiates packages of deformed sediment and rock based upon their deformation histories and the recognition of key kinematic (direction of movement) indicators that can differenti- ate phases of ice movement. Application of kinetostratigraphy has focused on glacial sequences where glaciotectonic deformation, remobilization, and translation of preexisting strata (bedrock and/ or superficial) dominate over sedimentary processes. In these glaciotectonic terrains core lithostrati- graphic principles such as ‘way-up’ and the Law of Superposition cannot readily be applied (Twiss and Moores, 1992). Instead, strata within glaciotectonized terrains may be partitioned by laterally extensive tectonic detachments (rather than sedimentary contacts), and have undergone soft-sediment mixing, meso- to large scale deformation including folding, thrusting, and fracturing 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 401

(Bluemle and Clayton, 1984; Croot, 1987; Aber et al., 1995; Van der Wateren, 1995; Pedersen, 1996; Rattas and Kalm, 2004; Thomas et al., 2004; Le Heron et al., 2005; Aber and Ber, 2007; Phillips et al., 2011b; Lee et al., 2017). The study of glaciotectonic structures provides a powerful insight into the controls of glacier-induced sediment deformation, glacier behaviour, and a range of applied issues including ground stability, land sliding, and reservoirs for water, hydrocarbons, and gas hydrates (Lee and Phillips, 2013). The approach employs a similar methodology to those used by geologists for reconstructing the tectonic evolution of orogenic belts (Berthelsen, 1978). Successful application requires a sound knowledge of structural geology and this undoubtedly has historically restricted its routine deploy- ment in glacial stratigraphy (Rose and Menzies, 1996). Nevertheless, the approach overcomes many of the issues of utilizing lithostratigraphy on glacial sequences (Aber and Ber, 2007) includ- ing recognizing ice advances and changes in glacial processes that may have no lithostratigraphic signature (Lee et al., 2013). The role the stratigraphic technique provides in understanding how materials are deformed by glacier-related stresses and in reconstructing glacial processes (an inte- gral part of glacial stratigraphy) cannot be overstated (Rose and Menzies, 1996).

11.5.2.2 Characterizing and building kinetostratigraphic units Stratigraphic subdivisions following the kinetostratigraphy approach are called kinetostratigraphic units. Individual kinetostratigraphic units are defined as all of the lithofacies and glaciotectonic deformation that possess a uniform pattern and direction of ice movement (Berthelsen, 1978). An important feature of any is its base which in kinetostratigraphy refers to the limit of penetrative deformation. A good example of the limit of penetrative deformation is the classic subglacial deforming bed profile including a subglacial till and deformed underlying materials called a glaciotectonite (see chapter: Subglacial Processes and Sediments) (Banham, 1977; Berthelsen, 1978; Benn and Evans, 1996). Glaciotectonic deformation produced by individual ice advances can be described as domainal where deformation of primary strata occurs during the same ice advance. Or, by contrast, extradomainal where deformation occurs within older or underlying strata formed during an earlier or unrelated event. Unlike lithostratigraphy that employs a rigid hierarchy for nomenclature, no formal nomencla- ture beyond an individual ‘kinetostratigraphic unit’ has been formalized. However, when building kinetostratigraphic units, certain rules must be obeyed (Berthelsen, 1978; Aber and Ber, 2007). Firstly, the attribution of unit status only occurs when consecutive ice advances come from differ- ent directions; secondly, kinematic indicators must clearly correspond to the direction of ice movement; and thirdly, each kinetostratigraphic unit must display a consistent regional pattern of ice movement. These rules provide strong guidance for the application of kinetostratigraphy. Essentially, that in a given scenario where several successive ice advances occur from a single direction (e.g., the northeast), all lithofacies and glaciotectonic deformation relating to these advances will collectively form only one kinetostratigraphic unit. This strict definition is somewhat limiting because local and regional ice flow directions can vary, reflecting: (1) irregular topography of the glacier bed; (2) variations in the glacier’s surface profile; (3) changes in stress field and rheological response within a material; and (4) within a lobate model of ice movement the direction of ice flow could vary by up to 120 degrees (Dreimanis, 1999; Aber and Ber, 2007). To overcome these issues, a modified approach has been developed which bases stratigraphic subdivision upon structural 402 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

architecture much like architectural analysis approach of bounding surfaces applied to sediments and sedimentary rocks (Pedersen, 2014). Alternatively, zones of glacitectonism can be subdivided into structural domains based upon associations of deformation styles and/or kinematic indicators (Lee and Phillips, 2008; Phillips et al., 2008; Lee et al., 2013; Vaughan-Hirsch et al., 2013). Building a kinetostratigraphic sequence requires the thorough and systematic description and measurement of a variety of glaciotectonic structures not only to reconstruct the relative ordering of deformation but also to establish the ice flow direction. Glaciotectonic structures include fractures, faults and joints, folds, foliations, and lineations with those structures that record a sense of displace- ment/shear the most significant (Van der Wateren et al., 2000; McCarroll and Rijsdijk, 2003; Phillips et al., 2011b). Clast (orientation) fabric data can also provide valuable directional and genetic information (Dowdeswell et al., 1985) but requires careful interpretation (Kjær and Kru¨ger, 1998; Bennett et al., 1999; Carr and Rose, 2003). Further guidance on relevant structural techniques can be found in Chapter 14, Geographic Information Systems and Glacial Environments (Glaciotectonics).

11.5.2.3 Applications Several good examples exist within the literature for the deployment of kinetostratigraphy to glacial deposits. A particular benefit of the approach is that it can be applied at a variety of spatial scales from regional scale, to field section scale, and even to a microscopic scale (Berthelsen, 1978). The paper by Berthelsen (1978) is arguably the classic work on kinetostratigraphy and presents the stratigraphic concept, methodology, and several practical examples of its application. Fig. 11.15 shows the application of kinetostratigraphy to three localities (AÀC) with kinematic indicators

SW NE

III

II

I A B C A B C

FIGURE 11.15 The kinetostratigraphic approach to reconstructing glacial stratigraphy showing a hypothetical example. (From Berthelsen, A., 1978.The methodology of kineto-stratigraphy as applied to glacial geology. Bull. Geol. Soc. Den. 27, 25À38). Vertical profile logs from three localities (AÀC) are shown on the left together with the direction of ice advance determined from glaciotectonic structures. On the right, the resolved stratigraphy showing three (IÀIII) distinctive ice advances with corresponding sediment packages. 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 403

shown relative to a vertical profile (left-hand figure). Segments of these vertical profiles are then grouped together into kinetostratigraphic units (IÀIII) based upon directional properties. Several further papers based upon the classic Danish sequences show the practical application of the tech- nique (Aber, 1979; Houmark-Nielsen and Berthelsen, 1981) whilst the technique has also been applied in the (Van der Wateren, 1995) and (Houmark-Nielsen et al., 2001). In the United Kingdom, spectacular glacitectonized sequences have been recognized on the Isle of Man (Slater, 1931; Thomas, 1984) with the island forming a temporary and dynamic ice- marginal still-stand location during the collapse of the Late Pleistocene Irish Sea (Roberts et al., 2007; Thomas and Chiverrell, 2011; Chiverrell et al., 2013). The Bride Moraine is the most distinctive morainic feature on the island (Slater, 1931; Thomas, 1984) with glacitectonic structures and kinematic data recording the progressive evolution and construction of the landform and syntectonic sedimentation during an ice push from the north-northeast (Fig. 11.16). Kinetostratigraphy can also be applied at the microscale, such as the novel paper by Phillips et al. (2011a) present a new ‘microstructural mapping’ technique for recording polyphase deformation within which glacial deposits are introduced. The technique, which is based upon a similar approach employed by metamorphic petrologists, develops a series of ‘structural domains’ comprising similarly oriented detrital grains (microfabric). These are then analyzed relative to other significant microstructures (e.g., plasmic fabrics, faults, folds, turbate structures, etc.) to develop a kinetostratigraphic framework that describes the genetic evolution of the host sediment (Fig. 11.17).

11.5.3 MORPHOSTRATIGRAPHY 11.5.3.1 Summary Morphostratigraphy, as traditionally defined, is a stratigraphical method that subdivides topography based upon its surface form (Frye and Willman, 1962). The technique establishes a relative order of landform development and is a powerful approach where geological processes produce distinc- tive landforms (Gripp, 1924). Modern or relatively fresh glacial sequences naturally lend them- selves to morphostratigraphy because of the range of distinctive landforms that glacial processes can produce (Sissons, 1967). These include morainic landforms that relate to previous ice-marginal positions, through to smaller landforms or associations of landforms that record a sequence of specific events within the glacial sequence (Sissons, 1967; Rose and Letzer, 1977; Boulton et al., 1985, 2001; Boulton and Clark, 1990; Livingstone et al., 2012). The development of new high-resolution bathymetric and 3D seismic datasets means that the technique is increasingly becoming employed in former or recently deglaciated terrains in offshore settings (Wellner et al., 2006; Winsborrow et al., 2010; O´ Cofaigh et al., 2012b). Whilst widely employed, several historical problems persist with morphostratigraphy. First, the resolution of data capture relative to the scale of the study area and the landforms can raise signifi- cant issues relating to detail being overlooked, captured inaccurately, or misinterpreted (Smith et al., 2006; Lu¨thgens and Bo¨se, 2012). Secondly, it is assumed that both the absolute and relative ages of the landforms (e.g., terminal ) are clearly known and this is not always the case (Lu¨thgens et al., 2010). Thirdly, morphostratigraphy assumes that the rate of landscape preservation is proportional to time—thus, the older the landforms, the lower the degree of preservation. In FIGURE 11.16 The structure of the Bride Moraine on the Isle of Man, UK (see inset showing map on figure). (A) Section reproduced from Slater, G., 1931. The structure of the Bride Moraine, Isle of Man. Proc. Liverpool Geol. Soc. 14, 184À196. (B) Section measured by Thomas, G.S.P., 1984. The origin of the glacio-dynamic structure of the Bride Moraine, Isle of Man. Boreas 13, 355À364. Note similarity in sections described 50 years apart. (C) Structural map across the foreshore in front of the moraine. (D) Summary section showing stratigraphy and extent of kineto- stratigraphic zone. (E) Contour plot (Wulff, lower hemisphere) of poles to foliation. Contours at 1% on 432 observations. (F) Plunge of fold axes. Arrow shows inferred direction of readvance. (G) Faults: solid circles, poles to low-angled overthrusts, open circles, poles to high-angled reversed faults. (H) Stages in the development of the moraine. Stratigraphic units: 1, Bride Mb; 2, Cronk-ny-Arrey Laa Mb; 3, Ballvarkish Mb; 4, Kionlough Mb. From Thomas, G.S.P., Chiverrell, R.C., 2011. Styles of structural deformation and syn-tectonic sedimentation around the margins of the Late Devensian Irish Sea Ice-stream: the Isle of Man, Llyn Peninsula and County Wexford. In: Phillips, E.R., Lee, J.R., Evans, H. (Eds.), Glacitectonic-Field Guide. Quaternary Research Association, London, pp. 59À78. FIGURE 11.17 An example of the micromorphological approach employed by Phillips et al. (2011a) to reconstruct the kinetostratigraphic framework of a poly- deformed glacial deposit. From Phillips, E., Meer, J.J.M. van der, Ferguson, A., 2011a. A new ‘microstructural mapping’ methodology for the identification, analysis and interpretation of polyphase deformation within subglacial sediments. Quat. Sci. Rev. 30, 2570À2596. 406 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

practice, this assumption can be difficult to apply. This is partly, because the scale and preservation of landforms is also influenced by mode of formation and lithology; but also because once formed, glacial landforms are highly susceptible to modification and erosion by a range of subaerial processes (Rose and Menzies, 1996; Ballantyne, 2002b), mass wastage (Everest and Bradwell, 2003), or burial by more recent deposits (Hughes, 2010). Once any of these factors come into force, morphostratigraphic precision drops rapidly. Many workers now employ supporting techniques to help subdivide glacial sequences including lithostratigraphy (Bowen, 1978; Pitka¨ranta, 2009) and sedimentology (Owen et al., 1998; Hughes et al., 2005). Such a sedimentÀlandform assemblage approach—also encompassing a genetic process-link—underpins the modern ‘landsystems’ approach to glacial geology (Evans, 2003a). Furthermore, a much greater emphasis is now placed on utilizing geochronology to aid morphostra- tigraphic subdivision. Not only does this help the development of more robust morphostratigraphic frameworks, but can also aid palaeoenvironmental reconstruction and ice-marginal dynamics (Larsen et al., 2006; Wysota et al., 2009; Clark et al., 2012; Lu¨thgens and Bo¨se, 2012; Marks, 2012; Rinterknecht et al., 2012).

11.5.3.2 Characterizing and building morphostratigraphic units Stratigraphic subdivisions that follow a morphostratigraphic approach are called morphostrati- graphic units. A morphostratigraphic unit was originally defined as ‘...comprising a body of rock that is identified primarily from the surface form its deploys; it may or may not be distinctive litho- logically from contiguous units; it may or may not transgress time through-out its extent’ (Frye and Willman, 1962, p. 7). Strictly speaking, this definition incorporates geology (lithology and subsur- face geometry) into stratigraphic subdivision and should be considered as a ‘hybrid’ morphostrati- graphic approach. Such an approach has been widely deployed in glacial terrains alongside other stratigraphic techniques such as lithostratigraphy and allostratigraphy. Collectively, these approaches can provide a powerful technique that incorporates lithology and surface/subsurface geometry. A purely morphostratigraphic unit, by contrast, can be defined as ‘... a body of rock or unconsolidated sediment that can be characterized by its surface form ...’. Many studies adopt such an approach which is suitable for characterizing and subdividing glacial landforms that occur in the landscape based upon their surface form. The identification of morphostratigraphic units requires highly detailed morphological or geomorphological mapping undertaken in the field and/or remote mapping using high-resolution remote datasets such as aerial photographs and digital surface models. Data collected during morphological mapping should include quantitative measurements of the surface topography such as the position of slope breaks, slope angles, and surface texture. Geomorphological mapping adds an interpretative element to surveying with individual genetic landforms or groups of mapped landforms. Supportive geometric data (e.g., size, orientation, slope angles) helps not only to quantify the scale of the landforms but can also be useful for recognizing the superimposition of different ‘events’ or processes, and up-scaling for inclusion in glaciological models. The reso- lution and quality of morphological or geomorphological mapping will be governed by a range of factors. These include: (1) the scientific or applied objectives and drivers of the survey; (2) the scale of the survey relative to that of the landforms; (3) the resolution of available remote datasets-if used; and (4) land-use history and the level of anthropogenic modification of the land surface. 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 407

11.5.3.3 Application Many examples exist within the literature that demonstrate the practical application of morphostra- tigraphy to glacial sequences. In upland areas, for instance, the technique has been routinely deployed to record glacier extent, interpret glacier behaviour and dynamics, and examine the inter- action of glacial processes with other nonglacial landforms and processes (e.g., periglacial, slope, etc.). This has led to the building of elaborate glacier reconstructions that depict the shape, surface profile, and equilibrium line altitudes of glaciers (Fig. 11.18)(Ballantyne, 1989, 2002a; Carr, 2001a; Hubbard et al., 2005; Finlayson et al., 2011; Bendle and Glasser, 2012). Of particular applied importance has been the utilization of morphostratigraphic techniques for indicating areas of geohazard risk (e.g., and glacier outburst floods) in high mountainous areas (Zimmermann et al., 1986; Huggel et al., 2004; Ka¨a¨b et al., 2005; Quincey et al., 2005, 2007). Morphostratigraphy has also been employed at a wider regional scale to map the spatial arrange- ments of glacial landforms in order to reconstruct the extent and dynamics of glaciation in both relict and recently deglaciated landscapes (Boulton et al., 1985, 2001; Stokes and Clark, 2001; Evans and Twigg, 2002; Phillips et al., 2010; Clark et al., 2012). An excellent example is provided by Evans et al. (2014) from southern Alberta where the authors have used morphostratigraphy to reconstruct the retreat of the Central Alberta Ice Stream, part of the Laurentide Ice Sheet (Fig. 11.19). The authors were able to recognize four individual landform assemblage zones: (1) minor morainic trans- verse ridges; (2) hummocky terrain; (3) fluted terrain; and (4) meltwater spillways. Collectively, the association of these landforms was interpreted to indicate a highly dynamic ice stream comprising several lobate snouts which underwent phases of cold, temperate, and polythermal basal behaviour punctuated by surge activity. Recent technological developments in capturing sea bed bathymetry data have enabled geomorphologists to recognize similar glacial landforms superimposed upon the sea bed and these have proved particularly helpful for developing onshoreÀoffshore stratigraphic correlations and studying submarine glacial landscapes adjacent to contemporary ice sheets (Fig. 11.20)(Wellner et al., 2006; Bradwell et al., 2008; van Landeghem et al., 2009; Winkelmann et al., 2010; O´ Cofaigh et al., 2012a; Graham et al., 2013; Dove et al., 2015). One of the biggest challenges within morphostratigraphy is building a robust framework of the morphostratigraphic units that demonstrates the relative order with which they were generated (Bo¨se et al., 2012; Lu¨thgens and Bo¨se, 2012). Now many geomorphologists employ geochronologi- cal techniques alongside morphostratigraphy to assist with this and numerous examples exist within the literature that adopt this approach (Rinterknecht et al., 2006; Lu¨thgens et al., 2010; Kalm, 2012; Bo¨se et al., 2012; Clark et al., 2012). The benefit of using this approach is that it also provides a detailed insight into rates and scales of ice-marginal retreat and this is fundamental in constraining glaciological models (Clark et al., 2012).

11.5.4 ALLOSTRATIGRAPHY AND ARCHITECTURAL ELEMENTS ANALYSIS 11.5.4.1 Summary Allostratigraphy is a stratigraphical method that subdivides geological sequences based upon a hier- archical framework of bounding surfaces or discontinuities that serve to compartmentalize discrete packages of sediment or rock (NACSN, 1983). Therefore allostratigraphy adopts a quite different conceptual approach to lithostratigraphy by focusing on the geometric arrangement of units 408 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

(A)

40 25 35 30

N

R

O L

F

O

H

T

R

I F 2000 m 5 5 7 7 Fluted drift Till sheet boulders Mountain summits Glacially-transported Peat Striae Dissected drift Ice-moulded bedrock Ice-moulded bedrock, shattered bedrock Frost bodies Water Contours at 100 metre intervals Roches moutonnées direction showing in metres n

o 0 0

0

D 0

3

3

h 867

Loc 0

0

00 00 0 1 1 1

0

0

2 200

I

I

I I

I

I I

I I I

I I

I

I

I

I I I

I I I I

I

I

I I

I I

I I

0

I 0

100 I 1

0

0 200 2 200

0

0

e

1 100 0 100 0

Meltwater channel Meltwater v Outwash, alluvium Outwash, Mountain-top detritus Moraines Drift limit or trimline or raised beaches or raised glacier limit Inferred 7 7

I

l

I

I e

I

I I I p

I

I

I 0 0 0

I S

0 0

I 0

I

2 2

I 2

I

e I

100 1 100

I

0 h

I

0 t I

i I 8

I

2 200 I

I 0 4 400 400 3

o

0

0

0 3 338

0 a 0

300 3 300 h

h

5 500

500 I

0

c 3 300 0 I

G

0

I

I

0 I a o

I I

d I I L

I I

I

I

I I

n I

I I

I

I I

ù

0

6 I

I I 0

400 4 400 I

6 I I

I

I I I I D Dùn da Ghaoithe I I

I I I I I 766 7 I

I

100 1 100

I 2 200 I 200

L 0 I I I 0

I I I 0

I

I I 0

I I 2 20

L 1 I I I I 20

0 I

I 4 0 0 I 0 I U 741 7 I

I I

g

I r I 3 300 I r 300

M r I I 0

a u I 0 1100

I e 0

I g 4004 I I 0 I I I 0

Sgurr S Dearg D 0

I

F I

I

7 I

I

3 I I

I

I O I

637 6 I

I I

I I

2 200 I

I 0 I

0

I I 5 0 5

00 00 0

I I

6 6 1 1 I 1

D I

I I

I 200200

I I I

I I I I I

N I I 8

0

0

400 4 400

I I

I

M

I I 9

I

I I

h E

I 0 U R 698 6

I 0

R c

n 200 2 200 I I O

I a O n

GORM G O GORM

I i

e

C

COIRE I COIRE r e 3003 0

S

I 0

0

I Creach Beinn B 2200 0 C

300 3 I 0

0 g I 0

I I 50050 s I I

I I i I

I 0

I I I

0

I 0

I I

I I

0 U

579 0

I 4 400 I 400 I I 200

0 I I 2

4004 0 I I I

0 0 I I I 3 300 I I

5500 I I I I I

0 I I I I I I I h

I I

I I

I I

I I I c

I

I

I I I I

300300 I o I

I I

I

I

I

I I L Loch Uisg I

I I

I I

I

I I n

I I

I

I I

I I

3003 I I

I 3 I

h n I

I I I

2002 0 I i

I I 6 0 d I 0

e i

763 7 I I I

0

I

1 100

I

a I I

B Beinn

0 l I

I

I I

I

I

I

0 I

I I

I

I I

I I

a I

I

I

I I I I T Talaidh Talaidh

I I I

I I I

I I I I I

I I I I I I

I

I

I I I I I I I 0 I 0 I

202 I I I I

I I I I I I I

I I I I I I I I I I I I I I I

I I I I

I I I I I I

I I 0

I I 0

500 5 500

I I I 0

0

I I

0

0

I

0

I I

0 I 0 I

300 3 300

I

4 400

I 400

A I

2 200 0 I

I 0 S I

I I

1 100 I I R I 0

I 0

I e

I

O 0

I i 300

F I 3 I 2 200 I I I 200 I I u

I I I I 0 I I

6 I

I I 0 I I

N I I I I B 1 100

I 5 I 100

I I Loch Buie 0

717 0

E I I I I I

6 I

I 456 4 6

L I I n I

I I

I e

I

G I

I I 300 3 300 0 0

I B Ben Buie I I I

I I I I I

I

I

E E I

I 500500 I 2

R R

I 0 0 Loch Fuaran

I 602 6 O O I I 0 2200

I I M I I I

I I

I I

2 200 I

N 0 n

I I

0 I 40400 I

2002 n E

0 I I i

0 I L

I I e 00

3003 I

I I

I I

h G I

I

I I 4

I I I 1 100

0 b 100

I I 0 - a 704 7 I 0 I

I a I r 0

B I I

r I

0 I

3 300 4 400 300

I o 0 I 496 0 Corra-bheinn h C I I

c 4 0

I I 0

I I 0 1 100

o I I 100

I I

I I 704 7 2 20

I 20 I I 0 0 0

L I I 0

I I

3 300

11000 300

0 2002 0 0 0 0 I I I I I 4400

I I

0 0

503

I

0 0

4 400 0 0

100 I 1

I I

0 5 5

0 5 5

2200

0

0 2 100

1 7 3300 5 572

I 0

0

I 2 0

I 0

I I 3 300 0 300 I I I 702 7 I I I 1 I I

I I 9

I 7 591 5 6 I 0 00 202 I 8 867 4004 0 000 0

I 0 5

I I 0

I I

6 600 I 600 I

I I

I I I I

l

I n

0 e

0

a e I

200

2 r

I e o

B Ben

I K 1100 More M More 00 6

00 6 966 9 a 3003

n 200200

2 200

I

0 0 I h 1001 0

0 5500 0 I

c 0

o

0 0

0 0 3 300

300 4400 Loch L 33000 0 Scridain © Crown Copyright. Ordnance Survey 40 35 30 25 An EDINA Digimap/JISC supplied service

FIGURE 11.18 (A) A geomorphological map of central and eastern Mull, Scotland. (B) A reconstruction of the Mull Icefield and other former glaciers in central and eastern Mull, Scotland. From Ballantyne, C.K., 2002. The Loch Lomond Readvance of the Isle of Mull, Scotland: glacier reconstruction and paleoclimate implications. J. Quat. Sci. 17, 759À771. 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 409

FIGURE 11.18 (Continued) (A) (B)

CAIS (n(normalormal fflowlow phphase)ase)

?

((normalnormal HPIS flow East East pphase)hase) Lobe HPIS Lobe ((streamingstreaming ? phphase)ase)

Hfs_5Hfs_5 CAIS (normal flow phase)

(C) (D) HPIS (s(streamingtreaming pphase)hase)

Hfs_4

CAIS (streaming(streaming pphase)hase) EEastast (normal HPIS LLobeobe flow phase) EastEast CCA_fs1A_fs1 LobeLobe

CAIS (normal flow phase)

(E) N

FIGURE 11.19 (AÀD) Reconstructed palaeoglaciology of the southern Alberta ice streams/lobes based upon published dates and geomorphological evidence. (E) Digital elevation covering SE Alberta showing the development of minor transverse ridges (MTR type 1) produced by glaciotectonic thrusting, overriding, and streamlining. From Evans, D.J., Young, N.J., O´ Cofaigh, C. 2014. Glacial geomorphology of terrestrial-terminating fast flow lobes/ice stream margins in the southwest Laurentide Ice Sheet. Geomorphology 204, 86À113. 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 411

N

FIGURE 11.20 Submarine bathymetry data (A) from western Scotland, UK showing glacially streamlined bedforms and recessional moraines superimposed upon bedrock strata. Summarized in (B). From Dove, D., Arosio, R., Finlayson, A., Bradwell, T., Howe, J.A., 2015. Submarine glacial landforms record Late Pleistocene ice- sheet dynamics, Inner Hebrides, Scotland. Quat. Sci. Rev. 123, 76À90. themselves rather than their lithological properties (Salvador, 1994; Rawson et al., 2002). In other words, whereas the ‘bricks’ represent the key stratigraphic building blocks in lithostratigraphy, in allostratigraphy, it is the ‘mortar’ between the ‘bricks’ that forms the stratigraphic framework. Allostratigraphy also shares many of the ‘architectural’ principles that underpin sequence stratigra- phy (Emery and Myers, 1996; Miall, 1997). However, unlike the latter, allostratigraphy is not driven by cyclical changes in sea-level or sedimentation (Slomka and Eyles, 2015). From a practi- cal perspective, allostratigraphy is essentially the same as the ‘unconformity-bound’ approach advocated by Chang (1975) and Salvador (1994). Allostratigraphy is, in principle, well-suited to deployment within glaciogenic sequences due to the often widespread occurrence of bounding surfaces including erosion surfaces, hiatuses, and tec- tonic detachments (Hughes, 2010). However, the stratigraphic technique remains somewhat underutilized in glacial geology (Ra¨sa¨nen et al., 2009). Used on its own, allostratigraphy does not enable either the characterization of deposits or the subdivision of sequences that possess complex 412 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

genetic histories—key challenges when working in glacial terrains (Slomka and Eyles, 2015). To overcome these issues, many researchers advocate adoption of a hybrid allostratigraphic approach encompassing other techniques such as lithofacies analysis, lithostratigraphy, or morphostratigraphy (Hughes, 2010). Architectural element analysis (AEA) is one such approach that utilizes both allos- tratigraphy and lithofacies analyses (Miall, 1985; NACSN, 2005). It enables the multiscaled charac- terization and delineation of sediment facies and geometry relative to a hierarchical framework of bounding surfaces that collectively define the sediment architecture (Miall, 1985, 1988, 1996; Fielding, 2006; Slomka and Eyles, 2013, 2015). The approach has been successfully applied to a range of glaciogenic settings enabling improved spatial extrapolating of sediment geometry through the subsurface (Eyles et al., 1998; Boyce and Eyles, 2000; Slomka and Eyles, 2013). The technique has obvious value for improving applied understanding of the properties of complex glacial succes- sions and is increasingly being utilized with groundwater modelling and hydrocarbon exploration (Clark and Pickering, 1996; Slomka and Eyles, 2013). 11.5.4.2 Characterizing and building allostratigraphic units/AEA subdivision Allostratigraphic subdivisions of sediment or rock are called allostratigraphic units. They are defined by bounding surfaces or discontinuities that separate units or rock and sediment regardless of lithology (NACSN, 1983). Typical bounding surfaces or discontinuities in glacial sequences include disconformities and constructional or erosional produced by sedimentary (Ra¨sa¨nen et al., 2009; Slomka and Eyles, 2015) or tectonic processes (Lønne, 2001; Lee et al., 2017). Bounding surfaces may also include (where preserved) the surfaces of a range of ice-contact and meltwater landforms (Passchier et al., 2010; Slomka and Eyles, 2015). Allostratigraphic units can therefore be used to define: (1) contiguous, discontinuity-bound (sedimentary or tectonic), homogeneous or heterogeneous deposits; (2) geographically separated deposits; and (3) packages of sediment that contain preserved palaeo surfaces including palaeosols. In a similar manner to lithos- tratigraphy, allostratigraphy also employs a relatively simple hierarchical nomenclature that defines different allostratigraphic units. In increasing scale and rank, formal allostratigraphic units follow a hierarchy of allomember, alloformation, and allogroup, with the alloformation representing the fundamental mappable unit and by inference, being laterally extensive (NACSN, 1983). By contrast, the AEA approach is a more informal approach without the stratigraphic nomen- clature that allostratigraphy uses (NACSN, 2005). When employed, stratigraphic subdivisions are often simply referred to as allostratigraphic units. However, correct hierarchical interpretation of the bounding surfaces or lithofacies contacts is critical to the approach. A sevenfold hierarchical framework of bounding surfaces was originally conceived by Miall (1985) and subsequently adapted by Slomka and Eyles (2015) for ice-marginal environments (Fig. 11.21). Lower-order bounding surfaces (first- to third-order surfaces) record self-generated surfaces developed within a depositional system, for instance, bedform migration and the development of sets, cosets, and reactivation surfaces. Fourth-order surfaces are larger-scale surfaces that delineate genetically related lithofacies associations to form individual bedforms or landforms. Higher-level surfaces record larger spatial changes in the depositional system. Fifth-order surfaces define groups of genetically related lithofacies that form landform tracts (e.g., terraces). Sixth-order surfaces separate genetically related lithofacies and landforms that form individual landsystem tracts (e.g., glaciofluvial from subglacial). Finally, seventh-order surfaces form the base of major depo- sitional complexes, separating deposits that correspond to different landsystems (e.g., glacial, marine, fluvial, etc.). FIGURE 11.21 A bounding surface hierarchy (1À7) for classifying bounding surfaces using the architectural element analysis approach. From Slomka, J.M., Eyles, C.H. 2015. Architectural-landsystem analysis of a modern glacial landscape, So´lheimajo¨kull, southern Iceland. Geomorphology 230, 75À97. 414 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

11.5.4.3 Application Pure allostratigraphy has received little attention from glacial geologists, but several excellent case studies demonstrate the value of architectural elements analysis to a range of different glacial settings including subaqueous (Eyles et al., 1998; Lønne, 2001), terrestrial (subglacial) (Boyce and Eyles, 2000), and glaciofluvial (Slomka and Eyles, 2013, 2015). The papers by Eyles et al. (1998) and Lønne (2001) apply AEA to subaqueous glacial settings. Eyles et al. (1998) employ the approach in their study of Permian-age marine and glaciomarine strata within New South Wales, Australia, which at the time lay adjacent to part of the glaciated Gondwana Supercontinent. Six separate lithofacies associations record sedimentation and back- ground ice-rafted input into various marine settings including storm-dominated shallow inner shelf, deeper-water mid-shelf, and turbidite-dominated upper continental slope. Two different bounding discontinuities were recognized (erosional surfaces and flooding surfaces) that were interpreted to relate to glacio-eustatic change, sediment supply, and basin tectonics. By contrast, Lønne (2001) employed AEA to reconstruct the dynamic evolution of tidewater morainic systems relative to ice-front dynamics, sediment input, and background marine depositional processes (Fig. 11.22). Five allostratigraphic units were defined (Units AÀE) and record the principal morphogenic and sedimentological changes that occur within an in-front system during ice advance (Unit A), ice- front still-stand (Unit B), deltaic sedimentation (Unit C), ice-front retreat and moraine abandonment (Unit D), and finally emerge (Unit E). Two distinctive assemblages of allostratigraphic units were also identified within tidewater moraines, reflecting formation during either single or multiple ice advances. The AEA approach has also been applied to studying the internal architecture of till sequences. In an example from Ontario, Canada, Boyce and Eyles (2000) used the technique to demonstrate the three-dimensional form and heterogeneity of a Late Pleistocene till sheet based upon outcrop, borehole, and geophysical data. Theywereabletoidentifyasuccessionofdiscrete diamicton units bounded by discontinuities and overlain by subglacial meltwater deposits and boulder pavements. Bounding discontinuities between the diamicton units were interpreted to reflect major changes in the dynamics of the Laurentide Ice Sheet. AEA has also been employed by Slomka and Eyles (2013) to reconstruct the evolution of a relict glaciofluvial sequence adjacent to Lake Ontario, Canada. The authors recognize six distinctive sediment package-types, so-called ‘element associations’, based on lithofacies and hierarchical bounding surface associa- tions (Fig. 11.23). These record the progressive change from braided river (EA1), to delta-front (EA2), to braided river and delta-top (EA3), to delta-front and lacustrine (EA4), braided river and deltaic (EA5), and finally, fluvial (EA6) styles of sedimentation (Slomka and Eyles, 2013). Critically, the authors also highlight the important applied value the technique has in reconstruct- ing the subsurface distribution of alluvial deposits which is important from a hydrocarbon and water resource perspective.

11.5.5 11.5.5.1 Summary Sequence stratigraphy is a stratigraphic approach that uses genetic cycles (commonly regressive and transgressive sea-level cycles) to drive predicted changes in sediment supply and sediment storage space, which in turn control the geometry and character of the sedimentary record FIGURE 11.22 (A) An allostratigraphic model for submarine moraine generation showing five allostratigraphic units (AÀE) and bounding discontinuities produced by different ice-contact and sedimentary processes. (BÀM) The range of possible moraine architectures developed during single or multiple ice advances. From Lønne, I., 2001. Dynamics of marine glacier termini read from moraine architecture. Geology 29, 199À202. FIGURE 11.23 Architectural element analysis applied to glaciofluvial sequences in Ontario, Canada. The photomosaics and sketches demonstrate the vertical and lateral lithofacies variability plus the wider lithofacies geometry defined by clearly developed fourth- and fifth-order bounding surfaces (see Fig. 11.19 for hierarchical significance of bounding surfaces). From Slomka, J.M., Eyles, C.H., 2013. Characterising heterogeneity in a glaciofluvial deposit using architectural elements, Limehouse, Ontario, Canada. Canad. J. Earth Sci. 50, 911À929. 11.5 STRATIGRAPHIC APPROACHES WITHIN GLACIATED ENVIRONMENTS 417

(Van Wagoner et al., 1988; Catuneanu et al., 2009; Emery and Myers, 2009). The technique is in part an evolution of the allostratigraphic approach described above (see Section 11.5.4) and was pioneered during the 1970s by exploration geologists searching for hydrocarbon reservoirs. Application of sequence stratigraphy has focused largely, but not exclusively, upon marine sedi- mentary sequences (Van Wagoner et al., 1988) and their extension into strata deposited in marginal and terrestrial environments such as fluvial and lacustrine (Wright and Marriott, 1993; Shanley and McCabe, 1994). However, despite widespread usage in classic areas of hydrocarbon exploration (Van Wagoner et al., 1988; Sharland et al., 2013), application of the technique elsewhere has proven more problematic (Catuneanu, 2006) and some have questioned its widespread applicability (Miall and Miall, 2001). Particularly controversial has proven the adoption of the ‘global-eustasy paradigm’ by the sequence stratigraphy community, with others highlighting the inherent spatial complexity (the ‘complexity paradigm’) of sea-level relative to local and regional controls (Miall and Miall, 2001). Nevertheless, sequence stratigraphy has been applied to glacial-marine sequences (Eyles and Eyles, 1992; Pedersen, 2012) with a small body of literature promoting its use in both pre-Quaternary (Proust and Deynoux, 1994; Etienne et al., 2007; Ghienne et al., 2007) and Quaternary deposits (Boulton, 1990; Brookfield and Martini, 1999). Issues surrounding local and regional influences on sea-level (and in turn sediment supply and sediment storage) are especially relevant in glacial environments (Brookfield and Martini, 1999; Benn and Evans, 2010; Pedersen, 2012). Significant ‘complexities’ may include: (1) differential glacio-isostatic loading and unload- ing of the crust across the margins of a basin which can generate a localized sedimentary response; (2) the production of erosion surfaces by subglacial erosion and meltwater incision beneath sea-level; and (3) the supply of sediment and accommodation space being controlled by local water level, the depositional surface, and point of sediment input relative to the glacier margin.

11.5.5.2 Application A missing element within sequence stratigraphy is that of a formally agreed nomenclature for defining stratigraphic units and, for this reason, unlike other stratigraphic techniques within this chapter, such a section has purposefully been omitted. Readers are instead directed to the wide range of literature that describes various types of sequence stratigraphic nomenclature that may be deployed (Emery and Myers, 1996; Catuneanu, 2006; Emery and Myers, 2009). However, several examples highlight the ways and challenges associated with employing sequence stratigraphy in glacial settings. The paper by Boulton (1990) adopts a sequence stratigraphic approach to distinguishing between sediments derived from different glacial environments relative to changing sea-levels within a glacial cycle. The resulting model (Fig. 11.24) demonstrates the temporal and spatial development of sediment architecture within a glaciomarine setting during transgressive and regres- sive phases. Brookfield and Martini (1999) provide an informative and thought-provoking perspec- tive on the application of genetic sequence stratigraphy to glacial environments. They argue that glacial sequences are deposited independently of sea-level due to the level at which sediment enters a water body and the fact that in lacustrine environments, water-level may be higher than sea-level and subjected to marked temporal variability. Brookfield and Martini (1999) suggest that genetic sequence stratigraphy should be abandoned in favour of a nongenetic allostratigraphic approach that simply describes the geometric arrangement of the stratigraphic units (Fig. 11.25). Upper Lower Foreshore Offshore shoreface shoreface

Highstand systems tract

Transgressive surface Pelagice / Hemipelagic

Stage 1: High sea-level

Erosion Sequence boundary

Submarine fan

Stage 2: Sea-level fall (regression)

Fluvial Estuarine Lowstand systems tract

Slope deposits

Stage 3: Low sea-level

Condensed facies Maximum flooding surface Transgressive systems tract

Pelagic / Hemipelagic

Stage 4: Rising sea-level (transgressive)

Upper Lower Foreshore shoreface shoreface Offshore Highstand systems tract

Stage 5: High sea-level (aggradation / progradation)

FIGURE 11.24 The conceptual sequence stratigraphic model showing depositional sequences, sequence boundaries, and sequence tracts relative to a sea-level rise and fall cycle. Modified from Nichols, G., 2001. Sedimentology and Stratigraphy. Blackwell Science, Oxford. 11.6 SUMMARY 419

FIGURE 11.25 A sequence stratigraphic model showing glaciomarine architecture driven by sea-level changes during a glacialÀinterglacial cycle. From Boulton, G.S., 1990. Sedimentary and sea-level changes during glacial cycles and their controls on glacimarine facies architecture. In: Dowdeswell, J.A., Scourse, J.D. (Eds.), Glacimarine Environments: Processes and Products. Geological Society, London, Special Publication, pp. 15À52.

11.6 SUMMARY This section of the chapter has provided an overview of the range of stratigraphically techniques that are typically deployed by geologists and geomorphologists working in glaciated terrains. Principal techniques utilized include lithostratigraphy, morphostratigraphy, allostratigraphy, AEA, kinetostratigraphy (tectonostratigraphy), and sequence stratigraphy. As can be seen within the over- view, each technique has its relative strengths and weaknesses and these together with the purpose of the stratigraphic model have to be borne in mind when employing and applying a particular approach. Many workers now consider that a hybrid stratigraphic approach encompassing relevant components of two or more techniques (including geochronology if available) is required to unravel the true stratigraphic complexity of glacial successions (Hughes, 2010). 420 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

11.7 CONCLUSIONS Within this chapter the importance of lithofacies analysis and several different stratigraphic approaches have been demonstrated and numerous examples have been used to illustrate their practical application and the practical and theoretical problems that may be encountered. The importance of time and space within glacial systems and how these fit into the wider geological timescale have also been discussed. Glacial geology thus offers a unique challenge to geologists and geomorphologists alike, because the highly heterogeneous and fragmented nature of the processes and products that occur in glacial environments are seldom replicated in other environ- ments. Nevertheless, with careful and methodological use of these techniques and a hierarchical approach to describing and interpreting field and remote data, it is possible to build strong and robust geological models of glaciated terrains. The methods deployed will depend upon the scien- tific and applied objectives of the research including temporal and spatial scale and the resources and skills that are available. However, the ultimate aim of any geological model is to be multi- scaled and applicable at various scales from site scale to continental, hemispheric, and global scale. This remains a substantial challenge in glacial geology and geomorphology but is essential to develop numerical models of glacial systems and fully understand their applied significance.

REFERENCES

Aber, J., Bluemle, J., Brigham-Grette, J., Dredge, L., Sauchyn, D., Ackerman, D., 1995. Glaciotectonic map of North America, 1: 6,500,000. Geological Society of America, Maps and Charts Series. Aber, J.S., 1979. Kineto-stratigraphy at Hvideklint, Møn, Denmark and its regional significance. Bull. Geol. Soc. Den. 28, 81À93. Aber, J.S., Ber, A., 2007. Glaciotectonism. Elsevier, Amsterdam. Aber, J.S., Bluemle, J.P., Brigham-Grette, J., Dredge, L.A., Sauchyn, D.J., Ackerman, D.L., 1993. Glaciotectonic data base and mapping of North America. Can. Plains Proc. 25, 177À200. Ballantyne, C.K., 1989. The Loch Lomond Readvance on the Isle of Skye, Scotland: glacier reconstruction and palaeoclimatic implications. J. Quat. Sci. 4, 95À108. Ballantyne, C.K., 2002a. The Loch Lomond Readvance on the Isle of Mull, Scotland: glacier reconstruction and palaeoclimatic implications. J. Quat. Sci. 17, 759À771. Ballantyne, C.K., 2002b. Paraglacial geomorphology. Quat. Sci. Rev. 21, 1935À2017. Banham, P.H., 1968. A preliminary note on the Pleistocene stratigraphy of northeast Norfolk. Proc. Geol. Assoc. 79, 507À512. Banham, P.H., 1975. Glaciotectonic structures: a general discussion with particular reference to the Contorted Drift of Norfolk. In: Wright, A.E., Moseley, F. (Eds.), Ice Ages: Ancient and Modern. Seel House Press, Liverpool, pp. 69À84. Banham, P.H., 1977. Glaciotectonites in till stratigraphy. Boreas 6, 101À105. Bendle, J.M., Glasser, N.F., 2012. Palaeoclimatic reconstruction from Lateglacial ( ) glaciers in Snowdonia, North Wales. Proc. Geol. Assoc. 123, 130À145. Benn, D.I., Evans, D.J.A., 1996. The interpretation and classification of subglacially-deformed materials. Quat. Sci. Rev. 15, 23À52. Benn, D.I., Evans, D.J.A., 1998. Glaciers and Glaciation. first ed. Arnold, London. Benn, D.I., Evans, D.J.A., 2010. Glaciers and Glaciation. second ed. Hodder Education, Abingdon. REFERENCES 421

Benn, D.I., Lukas, S., 2006. Younger Dryas glacial landsystems in North West Scotland: an assessment of modern analogues and palaeoclimatic implications. Quat. Sci. Rev. 25, 2390À2408. Bennett, G.L., Evans, D.J., 2012. Glacier retreat and landform production on an overdeepened glacier foreland: the debris-charged glacial landsystem at Kv´ıarjo´ ¨kull, Iceland. Earth Surf. Process. Landforms 37, 1584À1602. Bennett, M.R., Waller, R.I., Glasser, N.F., Hambrey, M.J., Huddart, D., 1999. Glacigenic clast fabrics: genetic fingerprint or wishful thinking? J. Quat. Sci. 14, 125À135. Berthelsen, A., 1978. The methodology of kineto-stratigraphy as applied to glacial geology. Bull. Geol. Soc. Den 27, 25À38. Bluemle, J.P., Clayton, L., 1984. Large-scale glacial thrusting and related processes in North Dakota. Boreas 13, 279À299. Bo¨se, M., Lu¨thgens, C., Lee, J.R., Rose, J., 2012. Quaternary glaciations of northern Europe. Quat. Sci. Rev. 44, 1À25. Boulton, G.S., 1990. Sedimentary and sea-level changes during glacial cycles and their controls on glacimarine facies architecture. In: Dowdeswell, J.A., Scourse, J.D. (Eds.), Glacimarine Environments: Processes and Products. Geological Society, London, pp. 15À52, Special Publication. Boulton, G.S., Clark, C., 1990. The Laurentide ice sheet through the last glacial cycle: the topology of drift lineations as a key to the dynamic behaviour of former ice sheets. Trans. R. Soc. Edinburgh Earth Sci. 81, 327À347. Boulton, G.S., Smith, G.D., Jones, A.S., Newsome, J., 1985. Glacial geology and of the last mid-latitude ice sheets. J. Geol. Soc. 142, 447À474. Boulton, G.S., Dongelmans, P., Punkari, M., Broadgate, M., 2001. Palaeoglaciology of an ice sheet through a glacial cycle: the European ice sheet through the Weichselian. Quat. Sci. Rev. 20, 591À625. Bowen, D.Q., 1978. Quaternary Geology: A Stratigraphic Framework for Multidisciplinary Work. Pergamon Press, Oxford, NY. Bowen, D.Q., 1999. A Revised Correlation of Quaternary Deposits in the British Isles. Geological Society of London, Oxford, London. Bowen, D.Q., Rose, J., McCabe, A.M., Sutherland, D.G., 1986. Correlation of Quaternary glaciations in England, Ireland, Scotland and Wales. Quat. Sci. Rev. 5, 299À340. Boyce, J.I., Eyles, N., 2000. Architectural element analysis applied to glacial deposits: internal geometry of a late Pleistocene till sheet, Ontario, Canada. Geol. Soc. Am. Bull. 112, 98À118. Bradwell, T., Stoker, M.S., Golledge, N.R., Wilson, C.K., Merritt, J.W., Long, D., et al., 2008. The northern sector of the last British Ice Sheet: maximum extent and demise. Earth-Sci. Rev. 88, 207À226. Bridge, J.S., 1993. Description and interpretation of fluvial deposits: a critical perspective. Sedimentology 40, 801À810. Brierley, G.J., 1989. River planform facies models: the sedimentology of braided, wandering and meandering reaches of the Squamish River, British Columbia. Sediment. Geol. 61, 17À35. Brodzikowski, K., van Loon, A.J., 1987. A systematic classification of glacial and periglacial environments, facies and deposits. Earth-Sci. Rev. 24, 297À381. Brookfield, M.E., Martini, I.P., 1999. Facies architecture and sequence stratigraphy in glacially influenced basins: basic problems and water-level/glacier input-point controls (with an example from the Quaternary of Ontario, Canada). Sediment. Geol. 123, 183À197. Burke, H., Phillips, E., Lee, J.R., Wilkinson, I.P., 2009. Imbricate thrust stack model for the formation of glaciotectonic rafts: an example from the Middle Pleistocene of north Norfolk, UK. Boreas 38, 620À637. Carr, S., 2001a. A glaciological approach for the discrimination of Loch Lomond glacial landforms in the Brecon Beacons, South Wales. Proc. Geol. Assoc. 112, 253À262. Carr, S.J., 2001b. Micromorphological criteria for discriminating subglacial and glacimarine sediments: evidence from a contemporary tidewater glacier, Spitsbergen. Quat. Intl. 86, 71À79. 422 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Carr, S.J., Rose, J., 2003. Till fabric patterns and significance: particle response to subglacial stress. Quat. Sci. Rev. 22, 1415À1426. Catuneanu, O., 2006. Principles of Sequence Stratigraphy. Elsevier, Amsterdam. Catuneanu, O., Abreu, V., Bhattacharya, J.P., Blum, M.D., Dalrymple, R.W., Eriksson, P.G., et al., 2009. Towards the standardization of sequence stratigraphy. Earth-Sci. Rev. 92, 1À33. Chang, K.H., 1975. Unconformity-bounded stratigraphic units. Geol. Soc. Am. Bull. 86, 1544À1552. Chiverrell, R.C., Thrasher, I.M., Thomas, G.S.P., Lang, A., Scourse, J.D., van Landeghem, K.J.J., et al., 2013. Bayesian modelling the retreat of the Irish Sea Ice Stream. J. Quat. Sci. 28, 200À209. Clague, J.J., Evans, S.G., 2000. A review of catastrophic drainage of moraine-dammed lakes in British Columbia. Quat. Sci. Rev. 19, 1763À1783. Clark, C.D., Hughes, A.L., Greenwood, S.L., Jordan, C., Sejrup, H.P., 2012. Pattern and timing of retreat of the last British-Irish Ice Sheet. Quat. Sci. Rev. 44, 112À146. Clark, J.D., Pickering, K.T., 1996. Architectural elements and growth patterns of submarine channels: applica- tion to hydrocarbon exploration. Am. Assoc. Pet. Geol. Bull. 80, 194À220. Clark, P.U., Mix, A.C., 2002. Ice sheets and sea level of the Last Glacial Maximum. Quat. Sci. Rev. 21, 1À7. Clayton, L., Moran, S.R., 1974. A glacial process-form model. In: Coates, D.R. (Ed.), Glacial Geomorphology. State Univeristy of New York, Binghamton, NY, pp. 89À119. Colgan, P.M., Mickelson, D.M., Cutler, P.M., 2005. Ice-marginal terrestrial landsystems: Southern Laurentide Ice Sheet margin. In: Evans, D.J.A. (Ed.), Glacial Landsystems. Hodder Arnold, London. Croot, D.G., 1987. Glacio-tectonic structures-a mesoscale model of thin-skinned thrust sheets. J. Struct. Geol. 9, 797À808. Cutler, P.M., Mickelson, D.M., Colgan, P.M., MacAyeal, D.R., Parizek, B.R., 2001. Influence of the Great Lakes on the dynamics of the southern Laurentide Ice Sheet: numerical experiments. Geology 29, 1039À1042. Dove, D., Arosio, R., Finlayson, A., Bradwell, T., Howe, J.A., 2015. Submarine glacial landforms record Late Pleistocene ice-sheet dynamics, Inner Hebrides, Scotland. Quat. Sci. Rev. 123, 76À90. Dowdeswell, J.A., Hambrey, M.J., Wu, R., 1985. A comparison of clast fabric and shape in Late Precambrian and modern glacigenic sediments. J. Sediment. Res. 55, 691À704. Dreimanis, A., 1982. Two origins of the stratified Catfish Creek Till at Plum Point, Ontario, Canada. Boreas 11, 173À180. Dreimanis, A., 1999. Need for three-dimensional analysis of structural elements in glacial deposits for determi- nation of direction of glacier movement. Geol. Soc. Am. Special Papers 337, 59À67. Dreimanis, A., Gibbard, P., 2005. Stratigraphy and sedimentation of the stratotype sections of the Catfish Creek Drift Formation between Bradtville and Plum Point, north shore, Lake Erie, southwestern Ontario, Canada. Boreas 34, 101À122. Dyke, A.S., Andrews, J.T., Clark, P.U., England, J.H., Miller, G.H., Shaw, J., et al., 2002. The Laurentide and Innuitian ice sheets during the Last Glacial Maximum. Quat. Sci. Rev. 21, 9À31. Ehlers, J., Gibbard, P.L., 1991. Anglian glaciation and glacial deposits in Britain and the adjoining offshore regions. In: Ehlers, J., Gibbard, P.L., Rose, J. (Eds.), Glacial Deposits in Great Britain and Ireland. Balkema, Rotterdam, pp. 17À24. Emery, D., Myers, K., 1996. Sequence Stratigraphy. Blackwell, Oxford. Emery, D., Myers, K., 2009. Sequence Stratigraphy. John Wiley & Sons, Chichester. Etienne, J.L., Allen, P.A., Rieu, R., Le Guerroue,´ E., 2007. glaciated basins: a critical review of the hypothesis by comparison with phanerozoic glaciations. In: Hambrey, M.J., Christoffersen, P., Glasser, N.F., Hubbard, B. (Eds.), Glacial Sedimentary Processes and Products. Blackwell Publishing Ltd, Oxford, pp. 343À399. Evans, D.J., Lemmen, D.S., Rea, B.R., 1999. Glacial landsystems of the southwest Laurentide Ice Sheet: mod- ern Icelandic analogues. J. Quat. Sci. 14, 673À691. Evans, D.J., Young, N.J., Cofaigh, C.O´ ., 2014. Glacial geomorphology of terrestrial-terminating fast flow lobes/ice stream margins in the southwest Laurentide Ice Sheet. Geomorphology 204, 86À113. REFERENCES 423

Evans, D.J.A., 2003a. Glacial Landsystems. Hodder Arnold, London. Evans, D.J.A., 2003b. Introduction to glacial landsystems. In: Evans, D.J.A. (Ed.), Glacial Landsystems. Hodder Arnold, London, p. 532. Evans, D.J.A., Benn, D.I., 2004. A Practical Guide to the Study of Glacial Sediments. Arnold, London, p. 266. Evans, D.J.A., Twigg, D.R., 2002. The active temperate glacial landsystem: a model based on Breiðamerkurjo¨kull and Fjallsjo¨kull, Iceland. Quat. Sci. Rev. 21, 2143À2177. Evenson, E.B., Dreimanis, A., Newsome, J.W., 1977. Subaquatic flow tills: a new interpretation for the genesis of some laminated till deposits. Boreas 6, 115À133. Everest, J., Bradwell, T., 2003. Buried glacier ice in southern Iceland and its wider significance. Geomorphology 52, 347À358. Eyles, C.H., Eyles, N., Gostin, V.A., 1998. Facies and allostratigraphy of high-latitude, glacially influenced marine strata of the Early Permian southern Sydney Basin, Australia. Sedimentology 45, 121À162. Eyles, N., 1983a. Glacial geology: a landsytems approach. In: Eyles, N. (Ed.), Glacial Geology. Pergamon, Oxford, pp. 1À8. Eyles, N., 1983b. The glaciated valley landsystem. In: Eyles, N. (Ed.), Glacial Geology. Pergamon, Oxford, pp. 91À110. Eyles, N., Eyles, C.H., 1992. Glacial depositional systems. In: Walker, R.G., James, N.P. (Eds.), Facies Models: Response to Sea-Level Change. Geological Association of Canada, Toronto, ON, pp. 73À100. Eyles, N., McCabe, A.M., 1989. The Late Devensian ,22,000 BP Irish Sea Basin: the sedimentary record of a collapsed ice sheet margin. Quat. Sci. Rev. 8, 307À351. Eyles, N., Menzies, J., 1983. The subglacial landsystem. In: Eyles, N. (Ed.), Glacial Geology. Pergamon, Oxford, pp. 19À70. Eyles, N., Eyles, C.H., Miall, A.D., 1983. Lithofacies types and vertical profile models; an alternative approach to the description and environmental interpretation of glacial diamict and diamictite sequences. Sedimentology 30, 393À410. Eyles, N., Eyles, C.H., McCabe, A.M., 1989. Sedimentation in an ice-contact subaqueous setting: the mid-Pleistocene ‘North Sea Drifts’ of Norfolk, U.K. Quat. Sci. Rev. 8, 57À74. Fielding, C.R., 2006. Upper flow regime sheets, lenses and scour fills: extending the range of architectural elements for fluvial sediment bodies. Sediment. Geol. 190, 227À240. Finlayson, A., Golledge, N., Bradwell, T., Fabel, D., 2011. Evolution of a Lateglacial mountain icecap in northern Scotland. Boreas 40, 536À554. Fleming, E.J., Stevenson, C.T.E., Petronis, M.S., 2013. New insights into the deformation of a Middle Pleistocene glaciotectonised sequence in Norfolk, England through magnetic and structural analysis. Proc. Geol. Assoc. 124, 834À854. Fookes, P., Gordon, D., Higginbottom, I., 1978. Glacial landforms, their deposits and engineering characteris- tics. In: The Engineering Behaviour of Glacial Materials: Proceedings of the Symposium on the Engineering Behaviour of Glacial Materials, Birmingham, UK, pp. 18À51. Frye, J.C., Willman, H.B., 1962. Morphostratigraphic units in Pleistocene stratigraphy. Am. Assoc. Pet. Geol. Bull. 46, 112À113. Gale, S.J., Hoare, P.G., 2012. Quaternary Sediments: Petrographic Methods for the Study of Unlithified Rocks. second ed. The Blackburn Press, New Jersey. Ghienne, J.-F., Le Heron, D.P., Moreau, J., Denis, M., Deynoux, M., 2007. The Late Ordovician glacial sedimentary system of the North Gondwana platform. In: Hambrey, M.J., Christoffersen, P., Glasser, N.F., Hubbard, B. (Eds.), Glacial Sedimentary Processes and Products. Blackwell Publishing Ltd, Oxford, pp. 295À320. Gibbard, P.L., 1980. The origin of stratified Catfish Creek Till by basal melting. Boreas 9, 71À85. Goudie, A.S., 1990. Geomorphological Techniques. second ed. Routledge, London. Graham, A.G., Dutrieux, P., Vaughan, D.G., Nitsche, F.O., Gyllencreutz, R., Greenwood, S.L., et al., 2013. Seabed corrugations beneath an Antarctic revealed by autonomous underwater vehicle survey: 424 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

origin and implications for the history of . J. Geophys. Res. Earth Surf. 118, 1356À1366. Gripp, K., 1924. U¨ ber die a¨ußerste Grenze der letzten Vereisung in Nordwest-Deutschland. Mitt. Geogr. Ges. Hamburg 36, 159À245. Hamblin, R.J.O., Moorlock, B.S.P., Rose, J., Lee, J.R., Riding, J.B., Booth, S.J., et al., 2005. Revised Pre-Devensian glacial stratigraphy in Norfolk, England, based on mapping and till provenance. Geol. Mijnbouw 84, 77À85. Hart, J.K., Boulton, G.S., 1991a. The glacial drifts of northeastern Norfolk. In: Ehlers, J., Gibbard, P.L., Rose, J. (Eds.), Glacial Deposits in Great Britain and Ireland. A. A. Balkema, Rotterdam, pp. 233À243. Hart, J.K., Boulton, G.S., 1991b. The interrelations of glaciotectonic and glaciodepositional processes within the glacial environment. Quat. Sci. Rev. 10, 335À350. Hart, J.K., Roberts, D.H., 1994. Criteria to distinguish between subglacial glaciotectonic and glaciomarine sedimentation, I. Deformation styles and sedimentolgy. Sediment. Geol. 91, 191À213. Hicock, S.R., 1992. Lobal interactions and rheologic superposition in subglacial till near Bradtville, Ontario, Canada. Boreas 21, 73À88. Houmark-Nielsen, M., Berthelsen, A., 1981. Kineto-stratigtraphic evaluation and presentation of glacial-strati- graphic data, with examples from northern Samsø, Denmark. Boreas 10, 411À422. Houmark-Nielsen, M., Kjær, K.H., 2003. Southwest Scandinavia, 40À15 kyr BP: palaeogeography and environmental change. J. Quat. Sci. 18, 769À786. Houmark-Nielsen, M., Demidov, I., Funder, S., Grøsfjeld, K., Kjær, K.H., Larsen, E., et al., 2001. Early and Middle Valdaian glaciations, ice-dammed lakes and periglacial interstadials in northwest Russia: new evidence from the Pyoza River area. Global Planet. Change 31, 215À237. Hubbard, B., Glasser, N.F., 2005. Field Tecniques in Glaciology and Glacial Geomorphology. Wiley, Chichester. Hubbard, A., Hein, A.S., Kaplan, M.R., Hulton, N.R.J., Glasser, N., 2005. A modelling reconstruction of the Last Glacial Maximum ice sheet and its deglaciation in the vicinity of the northern Patagonian icefield, South America. Geogr. Ann. A Phys. Geogr. 87, 375À391. Huggel, C., Haeberli, W., Ka¨a¨b, A., Bieri, D., Richardson, S., 2004. An assessment procedure for glacial hazards in the Swiss Alps. Can. Geotech. J. 41, 1068À1083. Hughes, P., Gibbard, P., Woodward, J., 2005. Quaternary glacial records in mountain regions: a formal strati- graphical approach. Episodes 28, 85À92. Hughes, P.D., 2010. Geomorphology and Quaternary stratigraphy: the roles of morpho-, litho-, and allostrati- graphy. Geomorphology 123, 189À199. Jones, A.P., Tucker, M.E., Hart, J.K., 1999. The Description and Analysis of Quaternary Stratigraphic Field Sections. Technical Guide 7. Quaternary Research Association, London, p. 295. Ka¨a¨b,A.,Huggel,C.,Fischer,L.,Guex,S.,Paul,F.,Roer,I.,Salzmann,N.,Schlaefli,S.,Schmutz,K., Schneider, D., Strozzi, T., 2005. Remote sensing of glacier-and permafrost-related hazards in high mountains: an overview. Nat. Hazards Earth Syst. Sci. 5, 527À554. Kalm, V., 2012. Ice-flow pattern and extent of the last Scandinavian Ice Sheet southeast of the . Quat. Sci. Rev. 44, 51À59. Karrow, P., 1984. Quaternary stratigraphy and history, Great lakes-St. Lawrence region. Quaternary stratigra- phy of Canada—a Canadian contribution to IGCP Project 24, 84-10. Karrow, P.F., Dreimanis, A., Barnett, P.J., 2000. A proposed diachronic revision of late Quaternary time- stratigraphic classification in the eastern and northern Great Lakes area. Quat. Res. 54, 1À12. Kehew, A.E., Esch, J.M., Kozlowski, A.L., Ewald, S.K., 2012. Glacial landsystems and dynamics of the Saginaw Lobe of the Laurentide Ice Sheet, Michigan, USA. Quat. Intl. 260, 21À31. Kjær, K.H., Kru¨ger, J., 1998. Does clast size influence fabric strength? J. Sediment. Res. 68, 746À749. REFERENCES 425

Kjær, K.H., Kru¨ger, J., 2001. The final phase of dead-ice moraine development: processes and sediment architecture, Ko¨tlujo¨kull, Iceland. Sedimentology 48, 935À952. Lachniet, M.S., Larson, G.J., Lawson, D.E., Evenson, E.B., Alley, R.B., 2001. Microstructures of sediment flow deposits and subglacial sediments: a comparison. Boreas 30, 254À264. Larsen, E., Kjær, K.H., Demidov, I.N., Funder, S., Grøsfjeld, K., Houmark-Nielsen, M., et al., 2006. Late Pleistocene glacial and lake history of northwestern Russia. Boreas 35, 394À424. Larson, G.J., Ehlers, J., Gibbard, P.L., 2003. Large-scale glaciotectonic deformation in the Great Lakes basin, USA and Canada. Boreas 32, 370À385. Le Heron, D.P., Sutcliffe, O.E., Whittington, R.J., Craig, J., 2005. The origins of glacially related soft-sediment deformation structures in Upper Ordovician glaciogenic rocks: implication for ice-sheet dynamics. Palaeogeogr. Palaeoclimatol. Palaeoecol. 218, 75À103. Lee, J.R., Phillips, E.R., 2008. Progressive soft sediment deformation within a subglacial shear zone-a hybrid mosaic-pervasive deformation model for Middle Pleistocene glaciotectonised sediments from eastern England. Quat. Sci. Rev. 27, 1350À1362. Lee, J.R., Phillips, E.R., 2013. Glacitectonics—a key approach to examining ice dynamics, substrate rheology and ice-bed coupling. Proc. Geol. Assoc. 124, 731À737. Lee, J.R., Booth, S.J., Hamblin, R.J.O., Jarrow, A.M., Kessler, H., Moorlock, B.S.P., et al., 2004. A new stratigraphy for the glacial deposits around Lowestoft, Great Yarmouth, North Walsham, and Cromer, East Anglia, UK. Bull. Geol. Soc. Norfolk 53, 3À60. Lee, J.R., Rose, J., Hamblin, R.J., Moorlock, B.S., Riding, J.B., Phillips, E., Barendregt, R.W., Candy, I., 2011. The Glacial History of the British Isles during the Early and Middle Pleistocene: implications for the long-term development of the British Ice Sheet. In: Ehlers, J., Gibbard, P.L., Hughes, P.D. (Eds.), Quaternary GlaciationsÀExtent and Chronology, A Closer look. Developments in Quaternary Science. Elsevier, Amsterdam, pp. 59À74. Lee, J.R., Phillips, E., Booth, S.J., Rose, J., Jordan, H.M., Pawley, S.M., Warren, M., Lawley, R.S., 2013. A polyphase glacitectonic model for ice-marginal retreat and terminal moraine development: the Middle Pleistocene British Ice Sheet, northern Norfolk, UK. Proc. Geol. Assoc. 124, 753À777. Lee, J.R., Phillips, E., Rose, J., Vaughan Hirsch, D., 2017. The Middle Pleistocene glacial evolution of northern East Anglia, UK: a dynamic tectonostratigraphicÀparasequence approach. J. Quat. Sci., 33, 231À260. Lisiecki, L.E., Raymo, M.E., 2007. Plio-Pleistocene climate evolution: trends and transitions in glacial cycle dynamics. Quat. Sci. Rev. 26, 56À69. Livingstone, S.J., Evans, D.J., O´ Cofaigh, C., Davies, B.J., Merritt, J.W., Huddart, D., Mitchell, W.A., Roberts, D.H., Yorke, L., 2012. Glaciodynamics of the central sector of the last BritishÀIrish Ice Sheet in Northern England. Earth-Sci. Rev. 111, 25À55. Lønne, I., 2001. Dynamics of marine glacier termini read from moraine architecture. Geology 29, 199À202. Lowe, J.J., Walker, M.J.C., 2014. Reconstructing Quaternary Environments. third ed. Routledge. Lunkka, J.P., 1994. Sedimentation and lithostratigraphy of the North Sea Drift and Lowestoft Till Formations in the coastal cliffs of northeast Norfolk. J. Quat. Sci. 9, 209À233. Lu¨thgens, C., Bo¨se, M., 2012. From morphostratigraphy to geochronology—on the dating of ice marginal positions. Quat. Sci. Rev. 44, 26À36. Lu¨thgens, C., Bo¨se, M., Krbetschek, M., 2010. On the age of the young morainic morphology in the area ascribed to the maximum extent of the Weichselian glaciation in north-eastern Germany. Quat. Intl. 222, 72À79. Mack, G.H., James, W., 1986. Cyclic sedimentation in the mixed siliciclastic-carbonate Abo-Hueco transitional zone (Lower Permian), southwestern New Mexico. J. Sediment. Res. 56. Maizels, J., 1993. Lithofacies variations within sandur deposits: the role of runoff regime, flow dynamics and sediment supply characteristics. Sediment. Geol. 85, 299À325. 426 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Marks, L., 2012. Timing of the Late Vistulian (Weichselian) glacial phases in Poland. Quat. Sci. Rev. 44, 81À88. May, R.W., Dreimanis, A., Stankowski, W., 1980. Quantitative evaluation of clast fabrics within the Catfish Creek Till, Bradtville, Ontario. Can. J. Earth Sci. 17, 1064À1074. McCabe, A.M., 1987. Quaternary deposits and glacial stratigraphy in Ireland. Quat. Sci. Rev. 6, 259À299. McCarroll, D., Harris, C., 1992. The glacigenic deposits of western Lleyn, north Wales: terrestrial or marine? J. Quat. Sci. 7, 19À29. McCarroll, D., Rijsdijk, K.F., 2003. Deformation styles as a key for interpreting glacial depositional environ- ments. J. Quat. Sci. 18, 473À489. McMillan, A.A., Merritt, J.W., 2012. A new Quaternary and Neogene Lithostratigraphical Framework for Great Britain and the Isle of Man. Proc. Geol. Assoc. 123, 679À691. Menzies, J., Maltman, A.J., 1992. Microstructures in diamictons—evidence of subglacial bed conditions. Geomorphology 6, 27À40. Meer, J.J.M. van der, 1993. Microscopic evidence of subglacial deformation. Quat. Sci. Rev. 12, 553À587. Meer, J.J.M. van der, Menzies, J., 2011. The micromorphology of unconsolidated sediments. Sediment. Geol. 238, 213À232. Miall, A.D., 1977. A review of the braided-river depositional environment. Earth-Sci. Rev. 13, 1À62. Miall, A.D., 1978. Lithofacies types and vertical profile models in braided river deposits: a summary. In: Miall, A.D. (Ed.), Fluvial Sedimentology. Canadian Society of Petroleum Geologists, Memoir 5, pp. 597À604. Miall, A.D., 1985. Architectural-element analysis: a new method of facies analysis applied to fluvial deposits. EarthÀSc. Rev. 22, 261À308. Miall, A.D., 1988. Architectural elements and bounding surfaces in fluvial deposits: anatomy of the (lower jurassic), Southwest Colorado. Sediment. Geol. 55, 233À262. Miall, A.D., 1996. The Geology of Fluvial Deposits. Springer Verlag, Berlin. Miall, A.D., 1997. The Geology of Stratigraphic Sequences. Springer-Verlag, Berlin. Miall, A.D., Miall, C.E., 2001. Sequence stratigraphy as a scientific enterprise: the evolution and persistence of conflicting paradigms. Earth-Sci. Rev. 54, 321À348. Mickelson, D.M., Clayton, L., Fullerton, D.S., Borns, H.W., 1983. The late Wisconsin glacial record of the Laurentide Ice Sheet in the United States. In: Porter, S.C. (Ed.), Late-Quaternary Environments of the United States, Vol.1, The Late Pleistocene. University of Minnesota Press, Minneapolis, MN, pp. 3À37. NACSN, 1983. North American stratigraphic code. Am. Assoc. Pet. Geol. Bull. 67, 841À875. NACSN, 2005. North American Stratigraphic Code. Am. Assoc. Pet. Geol. Bull. 89, 1547À1591. O´ Cofaigh, C., Evans, D.J.A., 2001. Sedimentary evidence for deforming bed conditions associated with a grounded Irish Sea glacier, southern Ireland. J. Quat. Sci. 16, 435À454. O´ Cofaigh, C., Dunlop, P., Benetti, S., 2012a. Marine geophysical evidence for Late Pleistocene ice sheet extent and recession off northwest Ireland. Quat. Sci. Rev. 44, 147À159. O´ Cofaigh, C., Telfer, M.W., Bailey, R.M., Evans, D.J.A., 2012b. Late Pleistocene and ice sheet limits, southern Ireland. Quat. Sci. Rev. 44, 160À179. Owen, L.A., Derbyshire, E., Fort, M., 1998. The Quaternary glacial history of the Himalaya. Quat. Proc. 6, 91À120. Passchier, S., Laban, C., Mesdag, C.S., Rijsdijk, K.F., 2010. Subglacial bed conditions during Late Pleistocene glaciations and their impact on ice dynamics in the southern North Sea. Boreas 39, 633À647. Pawley, S.M., 2006. Quaternary Glaciations of North and West Norfolk (Unpublished Ph.D. thesis). University of London, p. 453. Pedersen, S.A.S., 1996. Progressive glaciotectonic deformation in Weichselian and Palaeogene deposits at Feggeklit, northern Denmark. Bull. Geol. Soc. Den 42, 153À174. REFERENCES 427

Pedersen, S.A.S., 2012. Glaciodynamic sequence stratigraphy. In: Huuse, M., Redfern, J., Le Heron, D.P., Dixon, R., Moscariello, A. (Eds.), Glaciogenic Reservoirs and Hydrocarbon Systems. Geological Society, London, pp. 29À51, Special Publications. Pedersen, S.A.S., 2014. Architecture of glaciotectonic complexes. Geosciences 4, 26À296. Phillips, E., 2006. Micromorphology of a debris flow deposit: evidence of basal shearing, hydrofracturing, liquefaction and rotational deformation during emplacement. Quat. Sci. Rev. 25, 720À738. Phillips, E., Lee, J.R., 2013. Development of a subglacial drainage system and its effect on glacitectonism within the polydeformed Middle Pleistocene (Anglian) glacigenic sequence of north Norfolk, Eastern England. Proc. Geol. Assoc. 124, 855À875. Phillips, E., Lee, J.R., Burke, H., 2008. Progressive proglacial to subglacial deformation and syntectonic sedimentation at the margins of the Mid-Pleistocene British Ice Sheet: evidence from north Norfolk, UK. Quat. Sci. Rev. 27, 1848À1871. Phillips, E., Everest, J., Diaz-Doce, D., 2010. Bedrock controls on subglacial landform distribution and geo- morphological processes: evidence from the Late Devensian Irish Sea Ice Stream. Sediment. Geol. 232, 98À118. Phillips, E., Meer, J.J.M. van der, Ferguson, A., 2011a. A new ‘microstructural mapping’ methodology for the identification, analysis and interpretation of polyphase deformation within subglacial sediments. Quat. Sci. Rev. 30, 2570À2596. Phillips, E.R., Lee, J.R., Evans, H., 2011b. Glacitectonics-Field Guide. Quaternary Research Association, p. 263. Phillips, E., Lipka, E., Meer, J.J.M. van der, 2013. Micromorphological evidence of liquefaction, injection and sediment deposition during of glaciers. Quat. Sci. Rev. 81, 114À137. Pitka¨ranta, R., 2009. Lithostratigraphy and age estimations of the Pleistocene erosional remnants near the centre of the Scandinavian glaciations in western Finland. Quat. Sci. Rev. 28, 166À180. Preece, R.C., Parfitt, S.A., Coope, G.R., Penkman, K.E.H., Ponel, P., Whittaker, J.E., 2009. Biostratigraphic and aminostratigraphic constraints on the age of the Middle Pleistocene glacial succession in north Norfolk, UK. J. Quat. Sci. 24, 557À580. Proust, J.N., Deynoux, M., 1994. Marine to non-marine sequence architecture on a intracratonic glacially- related basin, Late Proterozoic of the West African platform in west Mali. In: Deynoux, M., Miller, J.M. G., Domack, E.W., Eyles, N., Fairchild, I.J., Young, G.M. (Eds.), The Earth’s Glacial Record: Facies Models and Geodynamic Evolution. Cambridge University Press, Cambridge, pp. 121À145. Quincey, D., Lucas, R., Richardson, S., Glasser, N., Hambrey, M., Reynolds, J., 2005. Optical remote sensing techniques in high-mountain environments: application to glacial hazards. Prog. Phys. Geogr. 29, 475À505. Quincey, D., Richardson, S.D., Luckman, A., Lucas, R.M., Reynolds, J.M., Hambrey, M.J., Glasser, N.F., 2007. Early recognition of hazards in the Himalaya using remote sensing datasets. Global Planet. Change 56, 137À152. Ra¨sa¨nen, M., Auri, J., Huitti, J., Klap, A., Virtasalo, J., 2009. A shift from lithostratigraphic to allostratigraphic classification of Quaternary glacial deposits. GSA Today 19, 4À11. Rattas, M., Kalm, V., 2004. Glaciotectonic deformation patterns in Estonia. Geol. Q. 48, 15À22. Rawson, P.F., Allen, P.M., Brenchley, P.J., Cope, J.C.W., Gale, A.S., Evans, J.A., Gibbard, P.L., Gregory, F. J., Hailwood, E.A., Hesselbo, S.P., Knox, R.W.O’B, Marshall, J.E.A., Oates, M., Riley, N.J., Smith, A.G., Trewin, N., Zalasiewicz, J.A., 2002. Stratigraphic Procedure. The Geological Society, London. Reid, C., 1882. The Geology of the Country Around Cromer HMSO, London. Rinterknecht, V.R., Clark, P.U., Raisbeck, G.M., Yiou, F., Bitinas, A., Brook, E.J., Marks, L., Zelcs,ˇ V., Lunkka, J.P., Pavlovskaya, I.E., Piotrowski, J.A., 2006. The last deglaciation of the southeastern sector of the Scandinavian Ice Sheet. Science 311, 1449À1452. 428 CHAPTER 11 GLACIAL LITHOFACIES AND STRATIGRAPHY

Rinterknecht, V.R., Braucher, R., Bo¨se, M., Bourle`s, D., Mercier, J.L., 2012. Late Quaternary ice sheet extents in northeastern Germany inferred from surface exposure dating. Quat. Sci. Rev. 44, 89À95. Roberts, D.H., Hart, J.K., 2005. The deforming bed characteristics of a stratified till assemblage in north East Anglia, UK: investigating controls on sediment rheology and strain signatures. Quat. Sci. Rev. 24, 123À140. Roberts, D.H., Dackombe, R.V., Thomas, G.S.P., 2007. Palaeo-ice streaming in the central sector of the British—Irish Ice Sheet during the Last Glacial Maximum: evidence from the northern Irish Sea Basin. Boreas 36, 115À129. Rose, J., 2009. Early and Middle Pleistocene landscapes of eastern England. Proc. Geol. Assoc. 120, 3À33. Rose, J., Letzer, J.M., 1977. Superimposed . J. Glaciol. 18, 471À480. Rose, J., Menzies, J., 1996. Glacial stratigraphy. In: Menzies, J. (Ed.), Past Glacial Environments-Sediments, Forms and Techniques. Butterworth-Heinemann, Oxford, pp. 253À284. Salvador, A., 1994. International Stratigraphic Guide. second ed. The International Union of Geological Sciences and The Geological Society of America, Boulder, CO. Schomacker, A., Kjær, K.H., 2008. Quantification of dead-ice melting in ice-cored moraines at the high-Arctic glacier Holmstro¨mbreen, Svalbard. Boreas 37, 211À225. Shanley, K.W., McCabe, P.J., 1994. Perspectives on the sequence stratigraphy of continental strata. Am. Assoc. Pet. Geol. Bull. 78, 544À568. Sharland, P.R., Archer, R., Casey, D., Davies, R., Hall, S.H., Heward, A.P., Horbury, A.D., Simmons, M.D., 2013. Arabian plate sequence stratigraphy. GeoArabia J. Middle East Pet. Geosci. 18, 371. Sharp, M., 1985. Sedimentation and stratigraphy at Eyjabakkajo¨kull—an Icelandic surging glacier. Quat. Res. 24, 268À284. Sissons, J.B., 1967. The Evolution of Scotland’s Scenery. Oliver & Boyd, Edinburgh. Slater, G., 1926. Glacial tectonics as reflected in disturbed drift deposits. Proc. Geol. Assoc. 37, 392À400. Slater, G., 1931. The structure of the Bride Moraine, Isle of Man. Proc. Liverpool Geol. Soc. 14, 184À196. Slomka, J.M., Eyles, C.H., 2013. Characterizing heterogeneity in a glaciofluvial deposit using architectural ele- ments, Limehouse, Ontario, Canada. Can. J. Earth Sci. 50, 911À929. Slomka, J.M., Eyles, C.H., 2015. ArchitecturalÀlandsystem analysis of a modern glacial landscape, So´lheimajo¨kull, southern Iceland. Geomorphology 230, 75À97. Smith, M.J., Rose, J., Booth, S.J., 2006. Geomorphological mapping of glacial landforms from remotely sensed data: an evaluation of the principal data sources and an assessment of their quality. Geomorphology 76, 148À165. Spedding, N., Evans, D.J.A., 2002. Sediments and landforms at Kvıarjo´ ¨kull, southeast Iceland: a reappraisal of the glaciated valley landsystem. Sediment. Geol. 149, 21À42. Stokes, C.R., Clark, C.D., 2001. Palaeo-ice streams. Quat. Sci. Rev. 20, 1437À1457. Thomas, G.S.P., 1984. The origin of the glacio-dynamic structure of the Bride Moraine, Isle of Man. Boreas 13, 355À364. Thomas, G.S.P., Chiverrell, R.C., 2011. Styles of structural deformation and syn-tectonic sedimentation around the margins of the Late Devensian Irish Sea Ice-stream: the Isle of Man, Llyn Peninsula and County Wexford. In: Phillips, E.R., Lee, J.R., Evans, H. (Eds.), Glacitectonic-Field Guide. Quaternary Research Association, London, pp. 59À78. Thomas, G.S.P., Chiverrell, R.C., Huddart, D., 2004. Ice-marginal depositional responses to readvance episodes in the Late Devensian deglaciation of the Isle of Man. Quat. Sci. Rev. 23, 85À106. Twiss, R.J., Moores, E.M., 1992. Structural Geology. W.H. Freeman and Company, New York, NY. Van der Wateren, F.M., 1995. Structural Geology and Sedimentology of Push Moraines. Mededelingen Rijks Geologische Dienst Nr 54, Haarlem. REFERENCES 429

Van der Wateren, F.M., Kluiving, S.J., Bartek, L.R., 2000. Kinematic indicators of subglacial shearing. In: Maltman, A.J., Hubbard, B., Hambrey, M.J. (Eds.), Deformation of Glacial Materials. Geological Society, London, pp. 259À278. Special Publication No. 176. van Landeghem, K.J.J., Wheeler, A.J., Mitchell, N.C., 2009. Seafloor evidence for palaeo-ice streaming and calving of the grounded Irish Sea Ice Stream: implications for the interpretation of its final deglaciation phase. Boreas 38, 119À131. Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M.J., Vail, P.R., Sard, J.F., Loutit, T.S., Hardenbol, J., 1988. An overview of the fundamentals of sequence stratigraphy and key definitions. In: Wilgus, C.K., Hastings, B.S., Kendall, C.G.S.C., Posamentier, H.W., Ross, C.A., Van Wagoner, J.C. (Eds.), Sea Level Changes: An Integrated Approach. Society of Economic Paleontologists and Mineralogists, pp. 39À45, Special Publication No. 42. Vaughan-Hirsch, D.P., Phillips, E., Lee, J.R., Hart, J.K., 2013. Micromorphological analysis of poly-phase deformation associated with the transport and emplacement of glaciotectonic rafts at West Runton, north Norfolk, UK. Boreas 42, 376À394. Waller, R., Murton, J., Phillips, E., Lee, J.R., Whiteman, C.A., 2011. Sand intraclast development and the deformation of glacially-overridden permafrost, West Runton. In: Phillips, E., Lee, J.R., Evans, H.M. (Eds.), Glacitectonics: Field Guide. Quaternary Research Association, pp. 162À181. Wellner, J.S., Heroy, D.C., Anderson, J.B., 2006. The death mask of the : comparison of glacial geomorphic features across the continental shelf. Geomorphology 75, 157À171. Winkelmann, D., Jokat, W., Jensen, L., Schenke, H.-W., 2010. Submarine end moraines on the continental shelf off NE GreenlandÀImplications for Lateglacial dynamics. Quat. Sci. Rev. 29, 1069À1077. Winsborrow, M.C., Andreassen, K., Corner, G.D., Laberg, J.S., 2010. Deglaciation of a marine-based ice sheet: Late Weichselian palaeo-ice dynamics and retreat in the southern reconstructed from onshore and offshore glacial geomorphology. Quat. Sci. Rev. 29, 424À442. Wright, P.V., Marriott, S.B., 1993. The sequence stratigraphy of fluvial depositional systems: the role of flood- plain sediment storage. Sediment. Geol. 86, 203À210. Wysota, W., Molewski, P., Sokołowski, R.J., 2009. Record of the ice lobe advances in the Late Weichselian glacial sequence in north-central Poland. Quat. Intl. 207, 26À41. Zimmermann, M., Bichsel, M., Kienholz, H., 1986. Mountain hazards mapping in the Khumbu Himal, Nepal. Mt. Res. Dev. 29À40.