May, V.J. & Hansom, J.D. 2003. Coastal Geomorphology of Great Britain, Geological Conservation Review Series No. 28. JNCC, Peterborough, ISBN 1 86107 4840.

The original source material for these web pages has been made available by the JNCC under the Open Government Licence 3.0. Full details in the JNCC Open Data Policy

Figures

(Figure 1.1) Geographical distribution of UK coastal research, based on a comprehensive computerized bibliography of books and papers on the geomorphology of the British Isles, containing some 9000 entries, compiled by K.M. Clayton. Of the 9000 entries, some 1400 are classified as dealing with coasts. These in turn are indexed under the 100 km squares of the National Grid and the number of published articles is shown (encircled number) for each relevant National Grid square, or combination of National Grid squares. As the map shows, they are strongly biased to the southern half of Britain. Because some articles cover the coast in more than one grid square, the total number of entries on this map is 1671.

(Figure 1.2) Geological map of Great Britain, also showing the locations of the Coastal Geomorphology GCR Sites. The map shows sedimentary rocks classified according to their age of deposition and igneous rocks according to their mode of origin. The numbers in the key indicate age in millions of years (Ma). (Permit number IPR/26–45C British Geological Survey. (NERC. All rights reserved.)

(Figure 1.3) Relative rock resistance for 71 different outcrops (divided by lithology and age) was established through computer analysis of data on altitude, dissection, and geology for a grid of kilometre squares covering Great Britain and the surrounding continental shelf. Six consistent classes were established using up to 19 variables in various combinations. White areas are unclassified. (From Clayton and Shamoon, 1998, fig. 1).

(Figure 1.4) Spring tidal amplitude (measured in metres) around the British coast. Elevations should be doubled to give spring tidal range. (UKDMAP 1998, NERC.)

(Figure 1.5) Significant wave height exceeded for 10% of the year (significant wave height is the mean value of the highest 1/3 of all waves). Wave height is one of the manifestations of the quantity of wave energy. (UKDMAP 1998 NERC.)

(Figure 1.6) Littoral sediment cells and subcells and direction of littoral drift. After Motyka and Brampton, 1993 and HR Wallingford, 1997. Cells are numbered 1 to 7 anticlockwise from St. Abb's Head for Scotland and there are three subcells within the Orcadian cell and two within the Shetland cell (shown in the inset); clockwise from St Abb's Head, cells are numbered 1–11 for England and Wales.

(Figure 1.7) Holocene sea-level history: (a) global view; (b) sea-level history in the area around The Wash (Norfolk/Lincolnshire). Three views of the global change in sea level over the last 10 000 years are shown in (a); the smooth curves combine data from different areas, a reconstruction based on a smaller region will show an irregular pattern over time (Shepard, 1963; Bloom, 1978; Monier, 1972). Because of the local effects of uplift and subsidence, it is increasingly recognized that such global sea-level curves have the potential to mislead and that local relative sea-level curves are generally more secure. The sea-level curve in (b) for the area around The Wash is based on an accreting sedimentary sequence preserved in an area that is subsiding at an average rate of 0.9 m ka−1. If this subsidence has been at a steady rate, then the local relative sea-level curve (the pecked line) can be converted to a eustatic curve (the solid line) by subtracting the effect of subsidence. (Based on Chorley et al., 1984, and Tooley and Shennan, 1987.)

(Figure 1.8) a. Uplift and subsidence, based on trends in sea level around the coast of Britain. These trends are established by comparing local sea-level curves with the eustatic trend shown in Figure 1.7b. Open circles indicate that the data were obtained by the author by extrapolation. The vertical scale on the graphs is in metres OD, the horizontal scale is in 103 sidereal years before present. (After Tooley and Shennan, 1987, p.136, fig. 4.9.) b. Map of isobases of uplift (positive values) and subsidence (negative values) following the Lateglacial and Holocene deglaciation of the British Isles. The rates shown (in mm a−1) are of crustal movement in Britain. Isobases cannot be drawn for much of southern England; point estimates are shown for guidance. (After Sherman, 1989, fig. 9.)

(Figure 1.9) Coastal engineering structures in Britain ('coastal protection' and 'sea defences'), shown as coastline with heavier line weight. Note the concentration on south-eastern England, Bristol channel, and north Wales and north-west England. (Data from Halcrow Group Ltd for England and Wales, published with permission from Department for Environment, Food and Rural Affairs; Scottish data after Ramsey and Brampton, 2000a-f.)

(Figure 2.1) (a) Clò Mór cliff (193m) to the east of Cape Wrath, Sutherland is a good example of a plunging chit with no shore platform development, which has been inherited from former sea levels. (b) Recession of the Chalk cliff at Sewerby, west of Flamborough Head, Yorkshire, has produced a steep lower cliff with a sloping shore platform whose upper junction is obscured by a gravel beach composed of chalk gravels together with glacial gravels derived from bevelling of the cliff-top till. (c) Rapid erosion of the soft and unconsolidated glacial till cliff at Atwick, Holdemess, Yorkshire, progresses by undercutting and rotational failure that is accentuated when the cliff-foot beach is thin or absent. This view looking north shows a very thin upper beach veneer over an area of exposed till shore platform (locally called an 'ord') whose surface is strewn with till blocks eroded from the cliff. (Photos: J.D. Hansom.)

(Figure 2.2) Coastal cliffs and their related shore platforms. A, cliff with intertidal ramp platform; B, cliff with shore platform at about high tide level; C, cliff coast with shore platform at about low tide level; D, plunging cliff with no shore platform; E, relict cliff with platform marked by emerged beach; F, typical inland cliff with talus at foot. (After Bird, 1984.)

(Figure 2.3) Processes of cliff retreat. SA = subaerial erosion of material, symbolized by the large arrow; M = marine erosion, symbolized by the fine arrows eroded material is removed offshore and alongshore by marine process. (a) SA<M; here steep cliffs, undercut by marine processes, develop. (b) SA=M; here a balance between the two sets of processes allows small beaches to develop at the toes of sloping cliffs; (c) SA>M; here subaerial mass movements by sliding produce a low stepped profile and marine transport of plentiful debris. On most coastal slopes, the rate of erosion of material falls far short of the ability of waves and tides to remove it, so that the slope angles are maintained (a,b). However, on weaker rocks (c) material is delivered at a rate controlled in large part by the ability of the sea to maintain removal and thus the rate of basal erosion, in which case slope angle will decline until sediment input matches the rate of removal. (After Hansom, 1988.)

(Figure 2.4) Toppling in the Torridonian sandstone cliffs south of Sandwood Bay, Sutherland. Dipping beds of well-jointed sandstones are subject to subaerial weathering and failure. Strong surf prohibits the debris from accumulating at the cliff foot. The stack in the distance is Am Buachaille (Gaelic for 'herdsman). (Photo: J.D. Hansom.)

(Figure 2.5) Cartoon depicting erosion of a vertical cliff under breaking-, broken- and unbroken-wave attack. Breaking waves cause the greatest amount of erosion. (Based on Sunamura, 1983, 1992 and Hansom, 1988.)

(Figure 2.6) Temporal variations in abrasion rate (AV/At), and beach elevation, h, expressed by the thickness of sand deposited at the cliff base, using data from wave-tank experiments. AV/At = volume of eroded material per unit time. (After Sunamura, 1976.)

(Figure 2.7) A classification of active sea-cliff forms according to comparative rates of subaerial erosion and marine erosion (SA = subaerial erosion; M = marine erosion). Type 'A' profiles are for cliffs of uniform resistance to erosion; type 'B', where a more resistant rock layer is present at the top; and type 'C', where there is a layer of more resistant rock at the base. (Based on Hansom, 1988, after Emery and Kuhn, 1982.)

(Figure 2.8) The development of caves, arches and stacks. Wave erosion is more effective along faults and joints where the rock is weaker, and so caves become excavated along these lines of weakness.

(Figure 2.9) Cliff height, and to some extent cliff form, is a function of the height of land cut by the cliffline. The photograph shows the cliff form of the Seven Sisters, Sussex, an almost straight cliffline truncates a series of dry valleys, the seven intervening ridges forming the Seven Sisters. (Photo: V.J. May)

(Figure 2.10) Plot of platform gradient against tidal range. Each point is a regional average of many surveyed profiles and suggests a direct relationship between gradient of platforms and spring tidal range. (After Trenhaile, 1987, p. 207.)

(Figure 2.11) Tidal duration curves from three locations as plotted in varying detail by two sources. Tidal duration is the length of time that still-water level occupies each elevation within the tidal range. (a) After Trenhaile and Layzell (1981); (b) after Carr and Graff (1982).

(Figure 2.12) Flow-diagram model of coastal cliff development over two glacial/interglacial cycles, starting from a vertical, unbevelled cliff profile, (a). During low sea levels, periglacial activity results in talus accumulation at the bases of cliffs. During high sea levels, the talus is removed and the cliff trimmed and stepped, and bevelled profiles (b) develop where the talus reached the cliff top during the last glacial stage, whereas multi-storied profiles (c) develop where the talus extended only part of the way up the cliff face. Both (b) and (c) cliff forms can be affected by a subsequent interglacial–glacial cycle, leading to the numerous possible complex stepped profiles (d) that depend on the resultant level of talus development between cycles. (After Griggs and Trenhaile, 1994.)

(Figure 2.13) The relationship between the occurrence of bevelled cliffs and ice limits in the British Isles. (After Griggs and Trenhaile, 1994.)

(Figure 3.1) High-cliffed coast of Great Britain, showing the location of the sites selected for the GCR specifically for coastal geomorphology features of hard-rock cliffs. Other coastal geomorphology GCR sites that include hard-rock cliffs in the assemblage are also indicated.

(Figure 3.2) Geological sketch map of the St Kilda archipelago showing the dominantly volcanic nature of the bedrock geology and the controlling effect of the granophyre sheets of the west in producing an approximately linear coastline. For relative geographical positions of the component islands of the archipelago, see Figures 3.1 and 3.4. (After Nature Conservancy Council, 1987.)

(Figure 3.3) a The west coast of Hirta, the main island of St Kilda, looking towards Dun, is characterized by stepped cliffs that steepen downwards to plunge steeply to well below sea level. (Photo J.D. Hansom.) Figure 3.3b Stac an Armin, seen here with the vertical plunging cliffs of Boreray (the second-largest island of St Kilda) in the foreground. This is the highest sea stack in Great Britain. (Photo J.D. Hansom.)

(Figure 3.4) Bathymetric map of the St Kilda archipelago. Depths are given in metres. Note the prominent break of slope where the roughly circular igneous complex stands proud of the otherwise low-gradient seabed. The seabed lies at c. 120 m and abruptly gives way to steep submarine cliffs that rise to a c. 60 m surface. In turn this surface abruptly gives way to submarine cliffs that may rise above sea level. Bathymetry is in metres below OD. (After Sutherland, 1984.)

(Figure 3.5) Geomorphological sketch map of the Villians of Hamnavoe showing extensive surfaces affected by both low-level wave-stripping in the south, and high-level wave-stripping in the north. For general location see Figure 3.1. (Modified from unpublished work by W Ritchie.)

(Figure 3.6) The Villians of Hamnavoe looking north towards South Head. The scoured surface is littered with both eroded boulders and debris thrown up by waves. Since some of this debris is of modern human origin (plastic fishing floats etc.) the waves that sweep the surface and emplace the debris and boulders are likely to be recent. (Photo J.D. Hansom.)

(Figure 3.7) The largest of three wave-emplaced boulder ridges that occur on top of a 15 m-high cliff some 50 m inland of the cliff edge at the Grind of the Navir, to the south of the Villians of Hamnavoe. Note 1.8 m-high figure for scale. (Photo J.D. Hansom.)

(Figure 3.8) Geomorphological sketch map of Papa Stour showing extensive wave-scoured cliff-top surfaces, together with stacks, caves, arches and geos. For general location, see Figure 3.1. (Modified from unpublished work by W Ritchie.)

(Figure 3.9) Fault-controlled stacks at Lamba Ness, Papa Stour. (Photo J.D. Hansom.)

(Figure 3.10) Coastal geomorphology of Foula. Sections 1–7 refer to descriptions in the text. The highest and most spectacular of these are Section 3 and Section 5 where the cliffs rise to 248 m at the Noup, and 376 m at the Kame respectively. See Figure 3.1 for general location. (After Pirkis, 1963.)

(Figure 3.11) Coastal features of the West Coast of Orkney. Erosion of the Hoy Sandstone and Stromness flags (inset) has produced an impressive coast of steep cliffs, caves and stacks. (Modified from unpublished work by W Ritchie.)

(Figure 3.12) The Old Man of Hoy, West Coast of Orkney, showing incipient failure cracks. London's 'Big Ben' is shown for scale in the inset. (After Hansom and Evans, 1995.)

(Figure 3.13) The spectacular arch at Qui Ayre, Yesnaby, West Coast of Orkney, is one of several arches and columnar stacks in the area in various stages of development. (Photo J.D. Hansom.)

(Figure 3.14) Coastal geomorphology of north-east Caithness, Duncansby to Skirza Head GCR site. Descriptions of sections 1–4 and of representative profiles A-C are in the text. The geology of the area is predominantly composed of horizontally bedded Old Red Sandstones (ORS), which have been eroded into steep cliffs. (Modified from unpublished work by W Ritchie.)

(Figure 3.15) The three large stacks of Duncansby stand in stark contrast to the otherwise bleak and smooth landscape of the north-east coast of Caithness. Looking north towards Duncansby Head and South Ronaldsay, Orkney, in the background. (Photo: courtesy of Ken Crossan.)

(Figure 3.16) Geomorphological map and geological sketch map of Tarbat Ness, Ross and Cromarry, north-east Scotland. The eastern Moray Firth shore is fault-controlled and rocky with a prominent emerged cliffline. The northern Dornoch Firth shore has well-developed emerged gravel beach-ridges. At the Ness itself, the low rock shore platform is characterized by a range of well-developed weathering pits and tafoni that are rare on Scottish coasts. At 'S' an emerged till-plugged stack occurs in front of the relict cliff. (Modified from unpublished work by W Ritchie.)

(Figure 3.17) The submerged landscape of North Uist looking north-west over Lochmaddy. Submergence of a low undulating rock surface has resulted in a landscape of low rock basins, platforms and skerries with a range of tidal and salinity conditions. (Photo: P. & A. Macdonald/SNH.)

(Figure 3.18) Coastal geomorphology of the Loch Maddy–Sound of Harris area, North Uist, showing the extensive areas of intertidal rock platform, small islets and skerries produced by submergence of a pre-existing low-lying rocky surface. The eastern coast is fault-controlled. (Modified from unpublished work by W Ritchie.)

(Figure 3.19) Geomorphological map of the coast of northern Islay between Mala Bholsa and Rubha a'Mhail, northern Islay, showing a fine series of emerged rock platforms and beaches some of which have been capped by glacial moraines whose age informs the chronology for the platforms and beaches. MHWS = Mean High-Water Springs. For general location see Figure 3.1. (After Dawson, 1991.)

(Figure 3.20) The coast of northern Islay, south of Rubha a'Mhàil showing the High Rock Platform and its backing cliff. In the foreground the Main Rock Platform and its backing cliff is also well developed. Lateglacial and Postglacial emerged gravels also adorn parts of the coastline. (Photo: J.E.Gordon.)

(Figure 3.21) Geomorphological map of the Bullers of Buchan, north-east Scotland. The inset on the right shows the typical cliff profile relative to high-water mark (HWM). Much of the cliff tops are veneered by glacial till. (Modified from unpublished work by W Ritchie.)

(Figure 3.22) The Pot, Bullers of Buchan is a 60 m-deep enlarged blowhole connected to the sea by a 15 m-wide tunnel-like arch. (Photo J.D. Hansom.)

(Figure 3.23) Geomorphological map and geological sketch map of the Dunbar GCR site showing a series of emerged rock shore platforms that have been eroded across a varied geology and are backed by cliffs of varying heights. Platform A lies at c. 20 m OD and is ice-moulded; Platform B is intertidal; Platform C underlies emerged Holocene beach deposits and Platform D is subtidal. (After Gordon and Sutherland, 1993.)

(Figure 3.24) Geomorphological map and geological sketch map of St Abb's Head showing the heavily indented nature of the coast resulting from a strong structural control. (Modified from unpublished work by W. Ritchie.)

(Figure 3.25) Main features of the Tintagel coast (i) Start Point to Dennis Point: vertical and slope-over-wall cliffs; (ii) Trebarwith Strand: sand beach backed by cliffs over 90 m high; (iii) Hole Beach: caves developed on line of faults and thrust planes; (iv) Penhallic Point to West Cove: slope-over-wall; (v) West Cove to Bossiney Haven: complex coast with peninsulas at different stages of separation from mainland; (vi) Bossiney Haven: geo and arch. The inset shows characteristic slope-over-wall forms between Trebarwith Strand and Tintagel Island.

(Figure 3.26) Elephant Rock, Bossiney, showing the relationship of cliff features to vertical jointing. (Photo: V.J. May.)

(Figure 3.27) Major fault and thrust at Tintagel as the focus for marine erosion, cave and ultimately stack development. (Photo: V.J. May.)

(Figure 3.28) Examples of coastline development controlled by major faults, Penhallic Point and Barras Nose. See Figure 3.25 for general location. (After Wilson, 1952.)

(Figure 3.29) Erosional features of the south Pembrokeshire coast. (After John, 1978.)

(Figure 3.30) Cliff profiles, South Pembroke Cliffs GCR site. Cliffs are steep, near-vertical and occasionally overhang where the dip is to landward. (Photo: S. Campbell.)

(Figure 3.31) Arch and stack development. (A) Form of the arch and stack at The Green Bridge of Wales. (B) Interpretation of development of the feature. An initial arch develops on the line of a discontinuity, and extends up-dip by spalling and collapse of up-dip rock surfaces. The arch roof collapses and a new stack is isolated.

(Figure 3.32) Hartland Quay GCR site — showing the pattern of truncated valleys. The profiles A–A', B–B', C–C' are shown at the bottom of the figure. Section I lies to the north of Section II. (After Arber, 1911.)

(Figure 3.33) Cliffs, platform, beach and truncated valleys south of Hartland Quay. (Photo: Lou Johnson, www.walldngbritain.co.uk.)

(Figure 3.34) Cross profiles of Solfach and the Gwada Valley, showing the contrast between the ria of Solfach and the infilled former ria at Gwada.

(Figure 4.1) Location of significant soft-cliffed coasts and platforms in Great Britain, indicating the sites selected for the GCR specifically for soft-rock cliff geomorphology. Other coastal geomorphology sites that include soft-rock cliffs and sites selected for the Mass Movements GCR 'Block' that occur on the coast are also shown.

(Figure 4.2) Rates of retreat along the North Sea Coast of England from Bridlington to Clacton-on-Sea. Rates are shown as averages for each length of cliff; where the length of cliff exceeds 5 km, values are every 5 km along the coast. Values are totals (metres) for 100 years to 1980. See also Table 2.1 and Table 4.2 (Compiled by K.M. Clayton)

(Figure 4.3) Retreat of the coastal cliff at Dunwich, Suffolk, plotted on the 1589 map of Agas; the 1977 cliff top as surveyed by A.H.W Robinson. (After Robinson, 1980a, p.141)

(Figure 4.4) (a-c) Undercutting of the cliffs at Ladram Bay. (a) General view looking north showing the stacks associated with headlands; small pocket beaches occupy the bays (b) Two natural arches as they appeared at the beginning of the 20th century in a picture postcard, and (c) the present-day equivalent, view looking SSW The strata are dipping seawards. (Photos (a,c): V.J. May.)

(Figure 4.5) The cliffline, platforms and stacks at Ladram Bay. Characteristic profiles are shown (A-A' and B-B'). Of particular note are the absence of stacks below the high cliffs, the presence of strata with fewer discontinuities in the lower stacks, and the tendency for stacks to be associated with headlands.

(Figure 4.6) Pattern of seaward-facing micro-cliffs on the landward-dipping strata (the strike of the strata is indicated) on the low-gradient intertidal platform in Robin Hood's Bay.

(Figure 4.7) Shore platform at Robin Hood's Bay looking east from Mill Beck (see Figure 4.6 for location). (Photo: J.D. Hansom.)

(Figure 4.8) (a) Cross-sections, showing characteristic forms of the platform east of Watchet, where the dip of strata to landward or seaward strongly affects the pattern of micro-cliffs, (b) three characteristic platform profiles at Nash Point, Vale of Glamorgan (see GCR site report in the present chapter) where dip of strata is more uniform than at Watchet. Mean high- and low-water spring tide levels (MHWS and MLWS) and mean high- and low-water neap tide levels (MHWN and MLWN) are shown. (Part (b) is after Trenhaile, 1972.)

(Figure 4.9) Cliffs and shore platform at Kilve, Somerset (Photo: V.J. May)

(Figure 4.10) Nash Point, this view from directly above the site demonstrates the near-vertical nature of the cliffs and the width of platforms at low water. The micro-relief of the shore platforms is controlled largely by the  relative strengths of alternating beds of limestone and argillaceous rocks and jointing patterns, on this photograph particularly noticable in the vicinity of Nash Point itself (see also Figure 4.8b). (Photo: CCUCAP, © the Countryside Council for Wales.)

(Figure 4.11) The sediment budget of beaches between Lyme Regis (to the westmost part of the map) and Seatown. (After Bray, 1990a.)

(Figure 4.12) View looking south-east from Golden Cap, showing the depleted shingle beach at Seatown, platforms that are cut across folded strata, and the residual boulders at the west end of the beach (foreground). (Photo: V.J. May.)

(Figure 4.13) (a) Cross-section of the Black Ven system (Lyme Regis to Golden Cap GCR site) and sediment supply to its beach. See also Figure 4.11. In (b) the volumes of sediment (in m3 a−1) moving through the Black Ven beach are given. (Based on Brunsden, 1973 and Bray, 1990a.)

(Figure 4.14) The Spittles, east of Lyme Regis. (A) Main landslide scar — sand and chert cliff; (B) landslide storage and throughput system; (C) sea cliff and mud flows; (D) beach; (E) dissected shore platform. (Photo: V.J. May.)

(Figure 4.15) Variations in the rates of cliff retreat from Blackgang to the Needles (to the west of Scratchell's Bay), Isle of Wight. Cliff profiles for sections A to E are shown. (After Hutchinson, 1984.)

(Figure 4.16) The Needles and Scratchell's Bay, Isle of Wight, with narrow flint and chalk beach fed by contemporary rockfalls. (Photo: J.E. Gordon.)

(Figure 4.17) Sediment inputs from cliff retreat (m3 a−1) annual longshore potential sediment transport and variations with wind direction. See text for explanation. Total sediment input = 392 908 m3 a−1 (After Davies, 1997.)

(Figure 4.18) Differences of failure in the cliffs of south-west Isle of Wight, ranging from large rotational slides to shallow failures. (After Hutchinson, 1984.)

(Figure 4.19) Characteristic slope failures at Compton Down, looking west, showing shallow slides in chalk rock. (Photo: V.J. May.)

(Figure 4.20) View looking east from Compton Down where chalk pebbles typically survive for little more than 1 km owing to their erosion during longshore drift. Well-developed cusps commonly characterize this beach. (Photo: V.J. May.)

(Figure 4.21) Cliff-face failures west of Freshwater Bay. (Based on British Gas aerial survey, February 1996.)

(Figure 4.22) Sketch map of boulder ridges, South Foreland to St Margaret's Bay within the Kingsdown to Dover GCR site. Characteristic cliff-platform profiles through A-A1 and B-B1 are shown in the lower part of the diagram.

(Figure 4.23) View looking north of St Margaret's Bay, Kingsdown to Dover GCR site. (A) Small, fringing beach of flint, mostly derived from recent cliff falls; movement alongshore is restricted by fall debris; (B) large toe of a slide extending beyond low-water mark; (C) cliff being eroded where previous rock fall has been completely removed; (D) vegetated slope that developed behind a former slide toe and debris; these features then protected cliff-foot bedrock from erosion; (E) typical upper cliff profile above debis slopes. (Photo: V.J. May.)

(Figure 4.24) Langdon Bay. (a) Boulder rampart residue from earlier debris tongue; (b) in the foreground, talus from a clif failure is seen; in the background, residual boulder fields from flow-type failures are present; (c) parallel ridges bounding a large flow-failure that left the platform comparatively clear of large debris. (Photos: V.J. May.)

(Figure 4.25) Sketch map of the Beachy Head to Seaford Head GCR site, showing the five subdivisions of the site as described in the text.

(Figure 4.26) (a) Beachy Head, cliff top view looking east, the cliffs are characterized by slab failures in the lower cliff that gradually undermine the upper cliff. (b) Cliff collapse at Beachy Head, early 1999; the failure affected the whole cliff face and produced a very large debris area at the cliff foot. (Photos: V.J. May.)

(Figure 4.27) Relationships between joints, cliff morphology and retreat near Birling Gap. (Photo: V.J. May.)

(Figure 4.28) Detail of the chalk and flint platforms east of Birling Gap. (Photo: V.J. May.)

(Figure 4.29) The cave-arch-stack sequence at Handfast Point, looking north-east, with Old Harry Rocks to the right. (Photo: V.J. May.)

(Figure 4.30) Ballard Down. Views looking east from [SZ 038 810] (a) taken on 12 January 2001 and (b) on 16 January 2001, showing the development of the landslip over four days. In (a) note the chalk scar formed by the failure of the slope. In (b), note the rectangular scar of the shallow rockslide that followed removal of bedrock and weathered slope materials at the back of the earlier failure. (Photo: V.J. May.)

(Figure 4.31) Cave–arch–stack development at Handfast Point 1887–1996. (Sources: 1887 Ordnance Survey and May and Heeps, 1985)

(Figure 4.32) Multi-faceted northern cliffline west of Handfast Point towards Studland. 1. Vertical upper cliff; 2. vegetated debris slope; 3. lower vertical cliff; 4. smooth cliff-platform junction; 5. notch; 6. flint and chalk pocket beach; 7. chalk platform. resistant to erosion as their foot is formed of harder Chalk. May (1971b) outlined the relationship of the erosional forms to the jointing pattern.

(Figure 4.33) Wave refraction at Flamborough Head, showing variations in wave direction crossing the platform owing to wave refraction. See Figure 4.34 for location. (Based on aerial photographs in Pethick, 1984.)

(Figure 4.34) Sketch map of the Flamborough Head coastal geomorphology GCR site, showing the three main divisions of the locality.

(Figure 4.35) Flamborough Head, (a) looking east from Thornwick Nab. The upper cliff is in Devensian tills, the lower cliff in chalk with numerous caves, arches and platforms. (b) Looking WNW at Bempton Cliffs; steep cliffs with a short upper vegetated facet in tills. Pipe-like forms extend down the whole height of chalk cliff; the cliffs have a narrow platform with a cobble and boulder beach. (Photos: V.J. May.)

(Figure 4.36) Cave and blowhole development at Flamborough Head, shown schematically in plan view. There are several stages in the development of blowholes here. Stage 1: caves develop along major joints or faults. Stage 2: caves extend upwards into the overlying till, which begins to collapse allowing hollows to appear in the till. Stage 3: caves merge and blowholes coalesce. Stage 4: Further merging of caves, cave moves collapse, arches and/or geos develop. Subsequently, isolated blocks or stacks may develop.

(Figure 4.37) Sketch map of the Joss Bay coastal geomorphology GCR site.

(Figure 4.38) One of two stacks in Botany Bay. This stack was joined to the mainland in 1842 and became separated during the 19th century. (Photo: V.J. May.)

(Figure 4.39) Wave refraction and reflection in Porth Neigwl. Wave orthogonals show the direction of travel of waves and are drawn at right angles to the wave crest. Open arrows are also orthogonals for reflected waves.

(Figure 4.40) Lost villages of the Holderness coast. As the till has been easily eroded for hundreds of years at rates of 2 m a−1, there has been substantial loss of agricultural land and villages. (After Hansom, 1988)

(Figure 4.41) The relationship between cliff height and erosion along the Holderness coast. (After Valentin, 1971, in Steers, 1971a). For the cliff height profile, the vertical exaggeration is x 30.

(Figure 5.1) Beach morphology. Synonyms: The term 'ridge-and-runnel' is sometimes used for 'bar and trough'; 'ball and low' is the old name for 'bar and trough'; 'bar', 'offshore bar' etc., are old names for barrier islands, not to be confused with longshore bar; 'swash bar' is the old name for 'berm'; 'high-tide beach' is used for 'beach face'; low-tide beach' is used for the seaward edge of low-tide terrace. See also Figure 5.2. (After Pethick, 1984, p. 93.)

(Figure 5.2) (A) Beach terminology: (1) beach, (2) shore, (3) upper beach (cordon littoral), (4) foreshore, (5) break of slope between upper beach and foreshore, (6) inner side of beach ridge, (7) lagoon, (8) marsh, (9) berms, (10) storm beach, (11) coastline, (12) ridges and runnels on the foreshore, (13) channel on foreshore, (14) pool in runnel of foreshore, (15) beach cusp, (16) apex of cusp, (17) bay of cusp, (18) horn of cusp, (19) ripple marks. (B) Formation of rhomboidal ripple marks. (After Fairbridge, 1968, p. 67.)

(Figure 5.3) The relationship between mean grain size of sand and beach slope, (beach slope is given as a ratio, from 1:5 to 1:100). (After King, 1972a, p. 325.)

(Figure 5.4) Beach steepness in East Anglia as a function of the proportion of shingle. The scatter of points is largely a function of variations in exposure to higher wave energies. (After Clayton, 1992, p. 64.)

(Figure 5.5) Sources and sinks of coastal sediment can be quantified to produce a sediment budget. Note the human element in the coastal sediment budget. (After Davies, 1980.)

(Figure 5.6) The run-up (oblique swash) and longshore current contributions to longshore or littoral drift. The amount of sediment moved alongshore depends on the wave energy component oblique to the shore. (After Fairbridge, 1968 and Komar, 1976.)

(Figure 5.7) Some examples of English spits: (A) Spurn Head; (B) Orfordness; (C) Hurst Castle; and (I)) Dawlish Warren. While the plan form of spits varies greatly, they all require an updrift sediment feed to form. In most cases, especially shingle spits, the sediment supply has now greatly decreased. (After Pethick, 1984, p. 108.)

(Figure 5.8) The formation of beach cusps. Cusps vary in size, but typical separation is in the range 2–10 m. (After Pethick, 1984, p. 112.)

(Figure 5.9) Sweep zones are a means of establishing long-term trends in beach form and sediment volume. These examples show the recovery of the south Lincolnshire beaches after the storm surge of early 1953. The sweep zones mark the vertical range of successive profiles (usually surveyed every few months), and by establishing two or more time periods, longer-term trends can be separated from short-term changes linked to changing wave climate. MHST: mean high spring tide; MHNT: mean high neap tide. (After King, 1972a, p. 359.)

(Figure 5.10) Profile of ridge-and-runnel beach (Blackpool). Ridge-and-runnel beaches occur where wide sandy beaches are dominated by local waves and swell is excluded, as here within the limited fetch of the Irish Sea. The short-period waves require a steep beach slope for equilibrium, and this is achieved through the formation of a series of ridges separated by runnels. (After King, 1972a, p. 342.)

(Figure 5.11) Coastal barriers enclosing coastal lagoons, Slapton, Devon and Chesil Beach, Dorset. Each of these barriers show gradation in pebble size along the barrier; coarse material is at the southern end of the Slapton barrier at Hallsands, and at the eastern end of Chesil Beach. (After Bird, 1984, p. 144.)

(Figure 5.12) The shaping of a recurved spit, based on the outline of Hurst Castle Spit (see GCR site report in Chapter 6). Waves from A, arriving at an angle to the shore, set up longshore drifting which supplies sediment to the spit; waves from B and C determine the orientation of its seaward margin and recurved laterals respectively. (After King and McCullagh, 1971 and Bird, 1984, p. 148.)

(Figure 5.13) Coastal barriers backed by saltmarsh, North Norfolk Coast GCR site (see GCR site report in Chapter 11). The barriers and recurves carry sand dunes; behind are sheltered tidal inlets and extensive areas of salt-marsh, part of which has been reclaimed for grazing. (After Bird, 1984, p. 149.)

(Figure 5.14) The formation of coastal sand dunes on a prograding coast (A–B). The older dunes farthest inland develop relatively mature soils and vegetation, and often the sand differs in colour from the younger dunes nearest the beach. The pecked line on E shows the effect that rising sea level and reduction in sediment supply has on sand dunes. Most of the dunes of Britain show such frontal erosion to a greater or lesser degree. Most dunes on the Scottish western and northern coastas are erosional. (Based on Hansom, 2001; Hansom and Angus, 2001, after Bird, 1984, p. 180.)

(Figure 5.15) Sequential development of the dune ridges in a dune system. The blowthroughs of the second and third dune ridges eventually form into parabolic dunes in the older ridge. (After Pethick, 1984.)

(Figure 5.16) Ways in which cliff-top dunes may develop. (A) A transgressive dune truncated by cliff recession; (B) a dune formed at higher sea level stranded by cliff recession after a fall in relative sea level; (C) a dune formed during a lower sea level truncated by cliff recession as sea level rises; (D) a dune that has advanced from a neighbouring beach across a headland. (After Bird, 1984, p. 190.)

(Figure 6.1) Forms and typology of gravel and shingle structures and the GCR sites that represent them. The schematic diagrams show the plan form of the structure concerned. Italic type indicates presence of relict features at a site. In some cases gravel forms the core of the feature, and is now covered in sand.

(Figure 6.2) Coastal shingle and gravel structures around Britain, showing the location of the sites selected for the GCR specifically for gravel/shingle coast features, and some of the other larger gravel structures.

(Figure 6.3) Historical recession of position of beach crests at Westward Ho! (Based on Campbell and Bowen, 1989 and Keene, 1996.)

(Figure 6.4) Comparison of geomorphological form between Slapton Sands and Loe Bar. Slapton Sands encloses a large lagoon, part of which has been infilled by sediment and become a brackish wetland. At the Bar, a cliff-foot beach confined between headlands has blocked off a narrow estuary

(Figure 6.5) Loe Bar, looking approximately south-east, showing the bar and its washover features. (Photo: V.J. May.)

(Figure 6.6) Hallsands and Slapton Sands represent parts of a once-continuous gravel beach. Offshore, there is evidence of buried shorelines and a possible former barrier beach. The present-day shingle beach is separated by rock headlands. (After Hails, 1975a.)

(Figure 6.7) Wave refraction along the coast between Slapton Ley and Hallsands. The Skerries Bank affects waves entering Start Bay from the south-west. Wave energy is concentrated in locations such as Hallsands and Beesands during north-easterly winds. (After Hails, 1975a–c)

(Figure 6.8) View, looking north-west, of the shingle barrier beach of Slapton Sands, enclosing the freshwater lagoon, Slapton Ley. Artifical sea-walls protect Torcross in the foreground. (Photo: V J. May.)

(Figure 6.9) Cross-section of beach at Hallsands, showing the historic beach levels prior to dredging. (After Mottershead, 1986.)

(Figure 6.10) The ruins of the landward row of the houses of the former village of Hallsands. The seaward row of houses has completely disappeared. Compare with Figure 6.9. (Photo: V.J. May.)

(Figure 6.11) Sketch map of the Budleigh Salterton Beach GCR site.

(Figure 6.12) Budleigh Salterton beach, looking west, also showing the cliffs that provide the source material for the gravel beach. (Photo: V.J. May.)

(Figure 6.13) Chesil Beach. View looking north-west, from Portland, with Chesilton in the foreground. The beach reaches 14 m OD and over 150m wide at its eastern end, where limited washover still occurs in spite of artifical modifications. (Photo: J.D. Hansom.)

(Figure 6.14) Map and sections of Chesil Beach. For general location see Figure 6.2. (Based on borehole information in Carr and Blackleg, 1969, 1973; and Carr and Seaward, 1990.)

(Figure 6.15) Sediment profiles of Chesil Beach and The Fleet. Sample cores are shown in sequence along the beach and The Fleet. Some peat layers have been dated in cores from the bed of The Fleet (dates are given in years BP). (Based on Carr and Blackley, 1973; Coombe, 1996 and Whittaker, pers. comm.)

(Figure 6.16) West Bay, Chesil Beach, showing the retreat of the shoreline and lack of sediment at the western end of the modern Chesil Beach. (Photo: V.J. May.)

(Figure 6.17) Chesil Beach (a) relationships between modern beach and dated peats and water levels (mean high-water springs and mean low-water springs, MHWS and MLWS, are shown). By c. 5000 years BP, the supply of flint was able to create a barrier beach atop an earlier sand ridge and estuarine peats. (b) Seabed features of eastern Lyme Bay and their relationship to Chesil Beach. Note the relation of bedrock exposures and seabed contours to the present shore, which probably affected the development of the earlier beach form. 'First attack' indicates the bathymetric contour representing the shoreline first attacked by the sea at the date shown.

(Figure 6.18) (A) Porlock barrier and back-barrier; (B) barrier crest and back-barrier changes before and after the 1996 barrier breaching; (C) barrier profile changes due to the 1996 storm.

(Figure 6.19) Overview of Porlock barrier (October, 1997) looking east. The 1996 barrier breach is identifiable, as are the storm generated washover fans at point NW (New Works sluice gate); HP is Hurlstone Point. (Photo: W B. Whalley.)

(Figure 6.20) Distal recurves at Hurst Castle Spit — the history of geomorphological development. (After Nicholls and Webber, 1987a.)

(Figure 6.21) (a) Changes in the profile of Hurst Castle Spit. (After Nicholls and Webber, 1987a.) (b) 1996 coast protection works at Hurst Castle Spit. The pecked line in (b) delimits the saltmarsh edge.

(Figure 6.22) Aerial photo of Hurst Castle spit. 1. Distal end of modern beach; 2. Groynes protecting Henrician (16th century) castle; 3. and 4. earlier recurves; 5. saltmarsh — the seaward edge of saltmarsh is undergoing retreat; 6. Spartina anglica-dominated saltmarsh, declining in area; 7. coastal defences at Keyhaven; 8. most commonly overtopped and artificially rebuilt section of beach ridge; 9. waves approaching from south-west. For discussion of the saltmarsh features, see GCR site report in Chapter 10 for Keyhaven Marsh. (Photo: courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 6.23) Historical changes at Pagham Harbour 1785–1961. (After Robinson, 1955.)

(Figure 6.24) Sediment pathways at Pagham Harbour. Arrows show sediment pathways with estimated annual volumes. (Based on Lewis and Duvivier, 1976; Hooke et al., 1996; and Harlow, 1979.)

(Figure 6.25) The Ayres of Swinister: a triple gravel barrier. Only the southern barrier is a true tombolo, the others are spits that enclose The Houb, a tidal basin. For general location, see Figure 6.2.

(Figure 6.26) South Ayre and North Ayre at high tide looking north-east towards Swinister Voe, showing the very sheltered nature of the site. Fish farms can be seen at the North Ayre (upper left of the photograph) (Photo: J.D. Hansom.)

(Figure 6.27) Washover lobes of gravel on the tombolo of South Ayre (to the right), with Fora Ness in the distance. Intertidal peats, which extend subtidally, are exposed at low tide in the lagoon between South Ayre and the unnamed barrier to the north-east (on the left). (Photo: Lorne Gill/SNH.)

(Figure 6.28) Graphs of modelled relative sea level against time over the last 16 000 years, along a south-north transect from Shetland to the Firth of Forth. (After Lambeck, 1993; Hansom, 2001.)

(Figure 6.29) Sketch map of the Whiteness Head area. The GCR site lies entirely within the eastern site area; the arrow indicates direction of net longshore sediment movement.

(Figure 6.30) Map of historical changes in the Whiteness Head Spit between 1946 and 1973. (After Smith, 1974.)

(Figure 6.31) Historical evolution of the Whiteness Head Spit between 1880 and 1991. In 111 years the spit has lengthened considerably and the creek morphology has changed. Note the pronounced change between 1958 and 1991 when the McDermott construction yard was built and a prominent channel was dredged on the south side of the spit. (After Stapleton and Pethick, 1996.)

(Figure 6.32) The extensive gravel ridges and emerged coastal and fluvial terraces of the Spey mouth in 1963. At this time, the river was diverted west by over 1 km, threatening the village of Kingston in the right centre of the view. (Photo from Gemmell et al., 2001b)

(Figure 6.33) Spey Bay showing coastal gravel strandplain backed by emerged marine (and fluvial) terraces. Land over 16 m is mainly glaciofluvial sands and gravels. MoD is a Ministry of Defence weapons testing range. (After Ritchie, 1983.)

(Figure 6.34) The emerged gravel ridges of Spey Bay descend in a 'staircase' from 9–10 m OD to the present-day beach. The greatest extent of the unvegetated gravel occurs to the west of Kingston (see Figure 6.35), where this picture was taken. (Photo: J.D. Hansom.)

(Figure 6.35) Movement of the River Spey mouth between 1870 and 1960. (After Grove, 1955 and Gemmell et al., 2001a.)

(Figure 6.36) Diagrammatic representation of the Spey Bay sediment budget. Scale approximate. (After Gemmell et al., 2001a.)

(Figure 6.37) Geomorphology of western Jura in the area of South Shian Bay, showing the 'staircase' of emerged gravel ridges. (After Dawson, 1993.)

(Figure 6.38) An unbroken 'staircase' of unvegetated emerged gravel beaches falls from c. 30 m OD to sea level on the West Coast of Jura. Looking eastwards towards Glenbatrick. (Photo: J.D. Hansom.)

(Figure 6.39) Cliff erosion and ness migration at Benacre Ness. The ness moves at 25 m a−1 to the north. The early accounts interpret the movement of the ness northwards as a result of accretion on the updrift side of the ness. The alternative view is that transport is towards the north (see Figure 6.40) and that accretion occurs on the lee (northern) side of the ness. Hardy (1966) suggests a reversal of movement of both the spit and direction of transport. (After Williams, 1956.)

(Figure 6.40) Longshore transport data for the East Anglian coast, showing estimated volumes and transport directions related to major shingle features. (After Cambers, 1975).

(Figure 6.41) Sketch map of the Orfordness–Shingle Street area.

(Figure 6.42) Schematic diagrams of shingle ridge groups representing developmental phases of Orfordness (A–B) spit development; (C–D) ness development; (E) extended spit with distal recurves; (F) additions to ness; (G–H) storm beach additions to spit. Each diagram portrays the north-south position accurately; the east-west position is arbitrary. (Based on Carr 1969b, 1972, 1973.)

(Figure 6.43) Variations in ridge patterns of Orfordness, in the southern part of the ness, the northern part of the spit with an earlier recurved spit fronted by individual shingle ridges, now largely destroyed, and also at the distal end of the spit, showing recurves. (Based on Carr, 1973; Green and McGregor, 1988.)

(Figure 6.44) Historical changes in the position of distal features at Orfordness. (After various authors, mainly Carr, 1965; and Green and McGregor, 1988.)

(Figure 6.45) Historical distal changes at Orfordness. showing development of major ridge crests.

(Figure 6.46) The cuspate foreland, Dungeness, Kent. The pecked lines 1 to 3 indicate former positions of the original spit over time, showing the downdrift extension of the spit across the bay. Saltmarsh has formed behind the outer shingle barrier. Over time, updrift erosion and downdrift deposition led to rotation of the feature from position 1 to 3. Land-claim of the marsh occurred in two phases — in the north it was drained in the Roman period, and in the 13th century diversion of the River Rother from its course north of Lydd to its new exit at Camber Castle led to the draining of the southern marshes. (After Bird, 1984, p. 159.)

(Figure 6.47) The historical evolution of Rye Bay. Dates indicate shoreline and beach area from contemporary maps and charts. (After Lovegrove, 1953; and Eddison, 1998.)

(Figure 6.48) Major zones of shingle at Dungeness.

(Figure 6.49) Schematic representation of the characteristic shingle ridge patterns and profiles at Dungeness. The vertical variation in ridge altitude is typically about 3m.

(Figure 6.50) Historical sediment pathways and development at Dungeness. Each schematic map shows the probable sediment movements associated with the erosional and accretional trends in the shoreline.

(Figure 6.51) Eastward development of Dungeness. The orientation of the beach ridges change and the ness forms is preserved in ridges dated between 600 AD and 1000 AD. The natural 'Open Pits' are areas of naturally lower and enclosed land that is seasonally or in some cases permanently freshwater. (Based on Steers, 1946a and Eddison, 1983b.)

(Figure 7.1) Great Britain sandy beaches and coastal dunes, also indicating the location of GCR machair–dune sites (see chapter 9) and other coastal geomorphology GCR sites that contain dunes in the assemblage.

(Figure 7.2) Key geomorphological features of Marsden Bay, Marsden Lea to Lizard Point.

(Figure 7.3) Marsden Bay — view looking towards the north-west showing the Magnesian Limestone cliffs and stacks and stumps. (Photo: V.J. May.)

(Figure 7.4) Historical dune development at South Haven. The 'Training Bank' extends south-eastwards from point X. (After Diver, 1933.)

(Figure 7.5) View looking north-eastwards towards Sandbanks and Poole Harbour, with Shell Bay (SB) to the right foreground, to the east of South Haven Point (SHP). Gravel Point (GP) lies in the forground to the left (see Figure 7.4 for sketch map). Brownsea Island, in the centre of Poole Harbour (see Figure 10.2, Chapter 10, for map), lies to the WNW of Sandbanks, just out of view. (Photo: ukaerialphotography.co.uk.)

(Figure 7.6) Cave relationships at Redend Point, South Haven Peninsula GCR site (see Figure 7.4 for location). (a) Cave height, h; width, w; length, 1. (b) Relationships between cave height (h), and w and I.

(Figure 7.7) Upton and Gwithian Towans GCR site. Both on the mainland and on the stack the sequence a–d is as follows: (a) dune grasses on blown sand; (b) thin sandy soil on weathered clay and angular intermittent gravel-sized clasts; (c) weathered bedrock; (d) bedrock. (Photo: V.J. May)

(Figure 7.8) Relationships between dunes and cliffs at Peter's Point. Profiles through section A, B and C are shown.

(Figure 7.9) Aerial photograph of dunes and Crow Point. 1, Westward Ho! cobble beach; 2, Taw–Torridge estuary; 3, Crow Point; 4, Airy Point; 5, Braunton Burrows showing main dune ridges and blowthroughs; 6, ridge-and-runnel beach. (Photo: courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 7.10) Braunton Burrows and Westward Ho! GCR sites, showing locations of emerged beaches and generalized geomorphology. See also Figure 7.11 for photograph of the area around Crow Point.

(Figure 7.11) Emerged beach profile and dune features at Braunton Burrows GCR site. (a) Section through emerged beach and possible former dunes at Saunton Down; (b) section through the central slack within the main dunes, showing that the dunes lie on both marine clay and gravels and sand resting on the underlying CuIm Measures bedrock.; (c) cross-section of the dunes showing the relationship of the slacks to the water table. (Based on Keene 1996; Willis 1985; and Willis et al., 1959a.)

(Figure 7.12) Key geomorphological features of Oxwich Bay, together with a typical profile.

(Figure 7.13) Oxwich Burrows. Linear dune ridges are evident, with alignment close to right angles to the shore at the western end of the dunes. Towards the east, the ridges retain this orientation close to the shore, but have a more east-west alignment inland. Similar ridges are absent from the dunes east of the stream mouth. (Photo: courtesy Cambridge University Collection of Aerial Photographs © Countryside Council for Wales.)

(Figure 7.14) Key geomorphological features and profile of the Tywyn Aberffraw GCR site. (After Robinson and Milward, 1983.)

(Figure 7.15) Aerial photograph of Aberffraw, Anglesey, for comparison with Figure 7.14. (Photo: Cambridge University Collection of Aerial Photographs © Countryside Council for Wales.)

(Figure 7.16) Ainsdale National Nature Reserve, view looking towards the west, North of Fisherman's Path. The site important for geomorphology (it is one of the three largest dune systems of the west coast of England and Wales) as well as for wildlife. In the middle distance a 'toadscrape' has been created to encourage natterjack toads. (Photo: copyright English Nature.)

(Figure 7.17) (a) Modern cross-section and zonation (eight zones) of active dune shore and nearshore zone. (After Parker, 1975.) (b) Historical schematic summary of dated peats. (After Tooley, 1978.)

(Figure 7.18) Dune-front processes at Ainsdale.

(Figure 7.19) Luce Sands is located at the head of a long linear embayment that is floored by extensive areas of sands and gravels. The result of unidirectional wave activity is that sediment is transported northwards on to the beach at Luce Sands. (After Single and Hansom, 1994.)

(Figure 7.20) The generalized coastal geomorphology of Luce Sands and Torrs Warren showing the wide intertidal area backed by extensive, largely stabilized sand dune. In the central section of the bay, two large areas of dune have been levelled for military use, and access to these areas and to the adjacent intertidal area is restricted. (After Single and Hansom, 1994.)

(Figure 7.21) The extensive and well-vegetated dune system of Torrs Warren has developed atop a series of emerged gravel ridges. Sections of these ridges are found in swales within the dune system and on the floors of healed blowthroughs. (Photo: J.D. Hansom.)

(Figure 7.22) Sandwood Bay, Sutherland, is dominated by a large and highly dynamic area of blown sand and machair that lies between the sea and the freshwater Sandwood Loch. Arrows show slope direction. (After Ritchie and Mather, 1969.)

(Figure 7.23) This view of the broad sweep of Sandwood Bay from the south shows the large areas of bare sand that indicate a high degree of dynamism at the beach–dune edge and within the dune-complex. Note the development of low tombolos linking the skerries to the beach crest (arrowed). Depending on the state of the tide these can be quite prominent features. (Photo: J.D. Hansom.)

(Figure 7.24) Looking south from the dune-capped gravel bar of Sandwood Bay towards the stack of Am Buachaille ('the Herdsman') in the distance. The low embryo dunes in the foreground lie adjacent to dune pillars, he eroded remnants of a more extensive dune cordon. (Photo: Lorne Gill, SNH.)

(Figure 7.25) The geomorphology of Torrisdale and Invernaver is bisected by a glacially scoured rock ridge that is flanked on either side by glaciofluvial terraces that are capped by windblown sands. The unvegetated upper beach is wide and backed by low dunes. Areas of saltmarsh occur along the exit of the River Borgie. (After Ritchie and Mather, 1969.)

(Figure 7.26) The large glaciofluvial terrace at Invernaver viewed from the east is flanked and capped by blown-sand deposits that also climb the ridge behind. The surface of the terrace also supports a wealth of archaeological remains including hut circles and cist burials. (Photo: J.D. Hansom.)

(Figure 7.27) The intertidal saltmarsh and sandflats of the River Borgie exit looking north-west over the low dune area and beach of Torrisdale Bay in the middle distance. (Photo: J.D. Hansom.)

(Figure 7.28) The coastal landforms of Dunnet Bay and dunes showing a coastal dune edge that is both undercut by frontal erosion and punctuated in several places by large, linear, blowthrough corridors. (Based on Ritchie and Mather, 1970a and Hansom and Rennie, 2003.)

(Figure 7.29) The wide expanse of Dunnet Bay looking west over the indented exit of the Burn of Midsand. Much of the coastal edge comprises mature dunes whose edge is now steep and undercut and whose surfaces now support re-invigorated marram growth. (Photo: J.D. Hansom.)

(Figure 7.30) Balta Island, Unst, Shetland, is low in the west and high in the east. It is mainly rocky except where sand is blown up-slope from the beach at South Links. (After MacTaggart, 1999.)

(Figure 7.31) The geomorphology of Balta, Unst. There are no dunes but instead the site supports a wide expanse of climbing dune grassland some of which has been eroded into low escarpments. In places the dune surface has been eroded down to a base level of calcarenite by both wind deflation and rill erosion. (After MacTaggart, 1999.)

(Figure 7.32) Generalized coastal features of Strathbeg, showing enclosure of the Loch by gravel ridges and a series of old dune ridges fronted by lower foredunes. Heights are in metres OD. The detailed sections a-c are shown in Figure 7.33a-c. (After Walton, 1956.)

(Figure 7.33) a–c Detailed coastal geomorphology of the (a) south, (b) central and (c) north sections of Strathbeg, showing the extensive series of shore-parallel dune ridges punctuated by the outlet from the loch. Representative sections through x–x are also shown. Arrows indicate direction of slope. The figure is continued overleaf (After Ritchie et al., 1978.) Figure 7.33b.Coastal geomorphology of Strathbeg.

(Figure 7.34) Possible evolution of the Strathbeg area during the Holocene Epoch showing the southward extension of gravel ridges and progressive closure of the former embayment. (After Walton, 1956.)

(Figure 7.35) The Loch of Strathbeg, as mapped in 1755 by the military surveys of William Roy. Note the loch exit is located in the south, but also that an artificial channel across the north end of the beach was in existence to allow the loch to drain northwards. (Photograph © The British Library from the British Library Special Collections, Maps C9b 31.)

(Figure 7.36) A large coalesced blowthrough in the southern part of Strathbeg. The loch is visible in the top right. The figure provides the scale. (Photo: J.D. Hansom.)

(Figure 7.37) Oblique cartoon of the Sands of Forvie and the northern part of Foveran. (After Ritchie et al., 1978.)

(Figure 7.38) Extensive areas of bare sand are visible in this oblique aerial view of Forvie and Foveran looking north along the axis of the Ythan estuary. In the middle right of the image, the large unvegetated dune of South Forvie allows sand to traverse the peninsula from the North Sea intertidal area in the east to the inner Ythan estuary in the west. In addition, sand still moves northwards from South Forvie but at a much smaller scale than in the past. These former sand movements contributed to the migration and development of several large parabolic dune systems that now rest on the higher ground of north Forvie. (Photo: P and A. Macdonald).

(Figure 7.39) The coastal landforms of the Sands of Forvie showing the bare sand and dune-arc dominated southern part, and the largely stabilized and vegetated northern part, which also hosts the nine groups of parabolic dunes. (After Ritchie et al., 1978.)

(Figure 7.40) The great dome of bare sand that dominates south Forvie is subject to active aeolian activity and sand movement. (Photo: J.D. Hansom.)

(Figure 7.41) The postulated phases of sand movement over Forvie. The lines relate to sand limits as follows: 1 = the northern limit of sand before about 0 BC; 2 = the sand limit a few hundred years after 0 BC (there was little further northward encroachment until at least the 8th century AD); 3 = the limit of the area inundated by sand early in the 15th century; 4 = position of the sand front by the end of the 15th century; 5 = line reached by 1688. Further small advances are shown by dated boundaries. (After Ritchie, 1992.)

(Figure 7.42) Location of Tentsmuir and Barry Links in St Andrews Bay. Tentsmuir and Barry Links have built out eastward of the main Postglacial (Holocene) shoreline at the mouth of the Tay estuary. Extensive intertidal and subtidal sand banks have also accreted at Abertay and Gaa Sands in the zone where river discharge interacts with open coast tides and waves. (After Ferentinos and McManus, 1981.)

(Figure 7.43) Generalized coastal geomorphology of Barry Links in 1981 showing the erosion of the narrow cordon of recent dunes and the linear nature of the series of older dune ridges, some of which are associated with parabolic forms downwind. As a result of concerns over erosion, a boulder revetment was built in 1992/1993 from the town of Carnoustie to extend along c. 3.5 km of the eastern shore. The section through A-B is shown in Figure 7.45. (After Wright, 1981.)

(Figure 7.44) Barry Links looking north showing the high dune edge at Buddon Ness itself and dune ridges of the south (estuarine) side. Clearly visible are the long linear dune ridges, some with parabolic forms, that have been truncated by erosion on the eastern (North Sea) shore. (Photo: P and A. Macdonald/SNH.)

(Figure 7.45) Stylized 3.5 km cross-section (along the line A–B on Figure 7.43) of Barry Links and Buddon Ness as reconstructed from borehole data. Barry Links sits atop substantial thicknesses of marine and shoreface deposits and suggests that this estuary-mouth site has undergone continued deposition over much of the Holocene Epoch. (After Paterson, 1981.)

(Figure 7.46) The relationship between mean parabolic dune orientation and the resultant vector of the wind polygon using dune orientations and locations in 1956. Note the eastern limits of the parabolic dunes in 1956 in comparison with their positions in 1981 as plotted in Figure 7.43. (After Landsberg, 1956, from Hansom, 1988.)

(Figure 7.47) Mid-flood and mid-ebb tidal stream patterns in St Andrew's Bay based on a combination of direct measurement and hydraulic modelling. The open coast at Tentsmuir is affected by northward movement on the flood and south-eastward movement on the ebb, whereas the open coast at Barry Links is affected by southward movement on both the flood and the ebb. (After Ferentinos and McManus, 1981.)

(Figure 7.48) The coastal landforms of Tentsmuir showing the extensive areas of sandflat, foredunes and intertidal sandbanks that extend out to Abertay Sands. Erosional edges are found in the south of Tentsmuir and along parts of the Tay estuary coast. (Based on Ritchie, 1979b, and McManus and Wal, 1996.)

(Figure 7.49) A spectacular oblique aerial photograph looking east towards the exit of the Tay at low tide with Tentsmuir and Abertay Sands extending into the distance on the south side and on the north Barry Links with Gaa Sands extending beyond. The recent sand accretions of Tentsmuir Point can be seen in the foreground. (Photo: P and A. Macdonald/ SNH.)

(Figure 7.50) Long-term changes in the position of south and north Tentsmuir showing a general trend of erosion in the south and accretion in the north. (Compiled from McManus and Wal, 1996.)

(Figure 8.1) Sand spits and their associated structures, indicating some key representative GCR sites.

(Figure 8.2) The location of sand spits in Great Britain, also indicating other coastal geomorphology GCR sites that contain sand spits in the assemblage. (Modified after Pethick, 1984).

(Figure 8.3) Pwll-ddu Bay. See Figure 8.4 for explanation. (Photo: Cambridge University Collection of Aerial Photographs © Countryside Council for Wales.)

(Figure 8.4) The succession of beaches at Pwll-ddu Bay: (1) oldest, (5) youngest (as yet undated). 1, 3 and 5 are the higher ridges that dominate the site.

(Figure 8.5) A east–west beach section at Ynyslas. The large arrow indicates the position of the submerged forest beds. (After Campbell and Bowen, 1989.)

(Figure 8.6) Historical changes at East Head. (After May, 1975.)

(Figure 8.7) Schematic diagram showing the key features including sediment transfers at East Head. (Sa = sand; Sh = shingle). Sand and shingle transported out of Chichester Harbour may be added by wave and aeolian action to longshore transport from the south-east. This may account for the excess of sediment reaching the spit over longshore transport. (After Harlow, 1982.)

(Figure 8.8) (a) The cyclical evolution of Spurn Head as envisaged by de Boer (1964). Over a period of about 240 years, the spit extends, beginning to develop a spatulate distal end after about 150 years. The neck of the spit is breached and a new cycle of spit growth begins. (b) The key features and 19th century development of Spurn Head. The log shows the sediments underlying Old Den. (After IECS, 1992.)

(Figure 8.9) Active processes at Spurn Head. Wave-refraction models indicate areas upon which wave energy concentrates. With waves from different directions (shown by the wave orthogonals), the zones of erosion and accretion change, giving rise to possible breaching from both North Sea and Humber sides of the spit. The wave-refraction models show wave convergence and divergence for waves from south-east and north-east, in both cases for waves 1 m-high, of period 8 seconds and at high tide. (After Halcrow, 1988; and IECS, 1992.)

(Figure 8.10) Key geomorphological features of Dawlish Warren, showing differences in slope on dunes and the upper beach, and differences in sediment sizes. n = number of observations of slope angle; R = mean slope angle; a = standard deviation. (I) = −log, (grain diameter in mm ); the grain-size profile for estuarine material and for Seaton, Devon, are shown for comparison.

(Figure 8.11) Aerial photograph of Dawlish Warren with the main geomorphological features numbered. 1 = Exe estuary, main channel; 2 = active recurved distal end (Warren Point); 3 = saltmarsh; 4 = inner spit (largely modified); 5 = outer spit; 6 = proximal end coastal protection works; 7 = intertidal sandbanks; 8 = prevailing and dominant wave direction (from the south-east). (Photo: courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 8.12) The key features of Gibraltar Point. Tidal currents based on Dugdale (1977). The West Dunes developed during the 19th century and the sandy ridge of Gibraltar Point since the 1920s.

(Figure 8.13) Location of scars along the coast of Walney Island, Lancashire.

(Figure 8.14) Former positions of the ness at Winterton, indicating a rapid southwards change in position between the 1880s and 1900s, but a subsequent movement northwards to the 1980s, and then a return southwards in the 1990s.

(Figure 8.15) Different interpretations of the sediment transfers at Winterton Ness. In the 1880s, according to Steers (1964a) and the Shoreline Management Plan (North Norfolk District Council et al., 1996), net sediment transport was southwards and the ness moved in the same direction. Others have suggested that transport is from the south, and Cambers (1975) and Halcrow (1988) agree on transport from both south and north with a transfer offshore and the ness extending seawards.

(Figure 8.16) Context of Morfa Harlech and Morfa Dyffryn — key geomorphological features.

(Figure 8.17) Key features of Morfa Harlech. (After Steers, 1946a.)

(Figure 8.18) Aerial photograph of Moth Harlech with the main geomorphological features numbered. 1 = former mainland; 2 = linear stable dunes ('grey dunes'); 3 = active 'yellow' dunes; 4 = zone of active blowthroughs; 5 = relict blowthroughs with SW—NE-aligned linear dunes; 6 = dune and slack topography; 7 = recurved linear dunes; 8 = former distal spits; 9 = 19th century distal features; 10 = modern distal dunes; 11 = intertidal sandflats. (Photo: courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 8.19) The main processes and sediment transfers at Morfa Harlech.

(Figure 8.20) The historical development of Morfa Dyffryn. During the eighteenth century the sand beach was separated from the morainic hill of Mochras by the channel of the Afon Artro and formed a spit with recurves at its northern end. In 1829, the Afon Artro was diverted to the east of Mochras. About the same time, the southern beach was transgressing inland. By 1830, the spit had closed the former river mouth and had joined Mochras. In the south, a new spit was developing northwards across the mouth of the Afon Ysgethin.

(Figure 8.21) Aerial photograph of part of the northern sector of Moth Dyffryn with sand transfers and the main geomorphological features numbered. 1 = till boulder and cobble beach derived from erosion of Mochras; 2 = main active zone of dunes and spit distal link with former island; 3 = major blowthrough; 4 = bar merging with beach — maintains sand supply to dunes; 5 = intertidal ridge and runnel; 6 = prevailing and dominant wave direction. (Photo: courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 8.22) Active dunes of Morfa Dyffryn migrating eastwards (in the foreground) are affected by a large blowthrough to the centre right. (Photo: V.J. May)

(Figure 8.23) St Ninian's tombolo, looking south-west towards St Ninian's Isle. Dunes flank either extremity of the sandy tombolo. During the highest spring tides the central part may be completely covered in water. (Photo: G. Satterley/SNH.)

(Figure 8.24) St Ninian's tombolo connects the Shetland mainland to an offshore island, and represents deposition from waves travelling south onto the north side, and vice versa in the south, in a very sheltered environment.

(Figure 8.25) Change in tombolo position. In the early 20th century the tombolo was wider than at present, but with its axis in a similar position. Aerial photographs taken in March 1993 showed that the tombolo had migrated northwards by about 30 to 40 m. Subsequent topographical surveys later in the same year showed that the tombolo was migrating back southwards, suggesting that the northward shift was a temporary feature caused by southerly winds. MLWS, MHWS represent the position of mean low- and high-water springs, respectively. (Source: J. Swale, SNH.)

(Figure 8.26) Key features of the coast of the Isles of Scilly.

(Figure 8.27) Geomorphological map of central Sanday showing the two tombolos that enclose Cata Sand and Little Sea. Note the orientation of the gravel ridges in Cata Sand. MHWS: Mean High-Water Springs; MLWS: Mean Low-Water Springs. (After Rennie and Hansom, 2001.)

(Figure 8.28) Looking north-east along the dune-capped tombolo in the Bay of Newark. Older intertidal gravel ridges can be seen extending inland towards the north in Cata Sand. (Photo: J.D. Hansom.)

(Figure 8.29) Coastline of Sanday in 1822 (from Thomson, 1832). Note the modem marine inlet at Cata Sand is mapped as a low area of land, possibly machair; the modem marine inlet of Little Sea is mapped as a freshwater lake and Start Island is mapped as a promontory with a lighthouse at the end.

(Figure 9.1) Distribution of machair in Scotland. Other than Sandwood, Torrisdale and Balta (see Chapter 7), all the sites included in the GCR fulfil both the geomorphological and vegetational definition of machair. Small vegetational differences in the above sites have resulted in the label 'probable machair'. Ongoing work that interprets the geomorphology and botany of machair aims to provide a definitive machair diagnostic test in the future and so the above classification will be subject to slight modification (Angus, 2003, pers. comm.). (After Hansom and Angus, 2001.)

(Figure 9.2) A diagrammatic representation of the beach–dune–machair system showing the general landward transport of sand broken by seaward returns via wind and streams. (After Mather and Ritchie, 1977.)

(Figure 9.3) The Holocene development of machair from approximately 6500 thousand years ago to present, showing the switch from conditions of accretion of the dunes to erosion and recycling of dune sands into machair. (a) early–mid Holocene; (b) late Holocene; (c) present day. (After Hansom and Angus, 2001.)

(Figure 9.4) A typical machair landscape of partly-drowned rock basins connected on the seaward side by wide sandy beaches and on the landward side by dune cordons backed by expanses of windblown machair sand. Looking north-east from North Uist over Vallay Strand in the foreground to Hornish and Lingay in the distance. (Photo: P and A. Macdonald/SNH.)

(Figure 9.5) Geomorphology of Machir Bay, Islay, showing a mix of machair types including substantial terraces at the rear of the system covered by high machair. (After Ritchie and Crofts, 1974.)

(Figure 9.6) Geomorphology of the Eoligarry isthmus. Note the narrow Atlantic beach of Tràigh Eais and the extensive flat beach of Tràigh Mhór between which lie a cordon of high dunes punctuated deeply by blowthroughs. The otherwise extensive machair surfaces are extremely narrow at the southern end of the isthmus. The position of the west-east cross-section of Figure 9.7 is indicated. (After Hansom and Comber, 1996.)

(Figure 9.7) Representative cross-section levelled west to east over Eoligarry (see Figure 9.6 for line of section). Note the expanse of Tràigh Mhór and the relatively narrow cordon of coastal dunes that are currently undergoing severe wind erosion. (After Hansom and Comber, 1996.)

(Figure 9.8) Looking south over the large blowthroughs on the west side of Eoligarry. Some remedial work has been undertaken but deflation is now so extensive that several dunes have been reduced to isolated 'buttes'. Tr-Ai& Eais is to the right. (Photo: J.D. Hansom.)

(Figure 9.9) The extensive machair lands of South Uist, looking north, with the exit of the Howmore River in the foreground, Ardivachar, Loch Bee, Gualan and North Uist in the distance at the top of the photo. (Photo: P and A. Macdonald/SNH.)

(Figure 9.10) (overleaf) The geomorphology of the area around Ardivachar, South Uist. A narrow cordon of active dunes separates the intertidal sandy beach from machair surfaces that are punctuated by erosional terraces. The west side of Loch Bee is subject to gradual infilling by machair sands. (After Ritchie, 1971.)

(Figure 9.11) A representative cross-section over Dremisdale Machair, South Uist, showing the relationship between landforms, vegetation and soil characteristics. Note the landward decline of calcium carbonate content of the sand from high values close to the beach. Dremisdale is sited just north of the Howmore River exit (see Figure 9.9). (After Mather and Ritchie, 1977.)

(Figure 9.12) The geomorphology of Hornish and Lingay Strands including Machairs Robach and Newton. Well-developed beaches, dunes and machair have benefited from the relative protection from westerly waves offered by the headland at Sollas and the island of Lingay (see Figure 9.4). (After Ritchie, 1971.)

(Figure 9.13) The geomorphology of Pabbay, Sound of Harris, showing the extensive area of climbing machair and the low-lying area east of Haltosh Point that has been infilled by beach and dune accretion since 1857, when it was a marine inlet. (After Ritchie, 1980.)

(Figure 9.14) The geomorphology of Luskentyre and Corran Seilebost. Extensive intertidal sands front the dune and machair landforms of Luskentyre where several large blowthroughs occur. The dunes reach 35 m OD and are the highest free-standing dunes in Harris. Numbered locations on Luskentyre are referred to in the text. The complex of beach–dune–machair features may be the remnant of a once-larger machair that has been fragmented by submergence. (After Ritchie and Mather, 1970b.)

(Figure 9.15) The rocky peninsula of Crago lies in the foreground of this aerial oblique looking north-west over Tràigh Luskentyre to the dunes and machair beyond. The western tip is highly active with clear evidence of extensive wind-blow. The free-standing spit of the tip of Corran Seilebost is also in view in the centre left. The island of Taransay (at the top) provides shelter from westerly waves. (Photo: P and A. Macdonald/SNH.)

(Figure 9.16) The geomorphology of the small embayment of Mangersta, Lewis. A narrow beach separates an area of low-lying sand from the sea. Mangersta is an example of a machair complex that has been eroded and deflated down to the water table. (After Ritchie and Mather, 1970b.)

(Figure 9.17) The central deflated surface of Mangersta viewed from the north-east. Climbing machair veneers the flanking slopes and occasional washover of the fronting beach occurs during storms. (Photo: J.D. Hansom.)

(Figure 9.18) The geomorphology of Tràigh na Berie showing a mainly eroding coastal edge (see Figure 9.20). Several small lochs that lie behind the machair surfaces are subject to infill by blown sand. (Modified from Ritchie and Mather, 1970b.)

(Figure 9.19) An aerial view of Tràigh na Berie from the south-west taken in the mid-1980s shows a dynamic coastal edge subject to pressure from tourist caravans and resulting in substantial wind-blow and destabilization. Caravans are now restricted to the central section of the dune and machair area. (Photo: S. Angus.)

(Figure 9.20) This view of Tràigh na Berie from the east (October, 2001) shows a wide intertidal zone backed by an undercut dune cordon and machair to landward. (Photo: J.D. Hansom.)

(Figure 9.21) The geomorphology of Balnakeil–An Fharaid is dominated by the easterly transport of blown sand from the beaches of Balnakeil Bay into dune and machair surfaces that climb the slopes of the An Fharaid peninsula. Some of the larger linear blowthroughs channel sand to cascade over the cliff on the eastern edge of the peninsula. (After Hansom, 1998.)

(Figure 9.22) Looking north over the large linear blowthrough at Balnakeil, a highly dynamic feature that channels wind flow to the east to allow sand to be blown onto the adjacent climbing machair surfaces as well as to cascade over the cliff edge at Flirum, at the right of the scene. (Photo: J.D. Hansom.)

(Figure 10.1) The generalized distribution of active saltmarshes in Great Britain. Key to GCR sites described in the present chapter or Chapter 11 (coastal assemblage GCR sites): 1. Morrich More; 2. Culbin; 3. North Norfolk Coast; 4. St Osyth Marsh; 5. Dengie Marsh; 6. Keyhaven Marsh, Hurst Castle; 7. Burly Inlet, Carmarthen Bay; 8. Solway Firth, North and South shores; 9. Solway Firth, Cree Estuary; 10. Loch Gruinart, Islay, 11. Holy Island. (After Pye and French, 1993.)

(Figure 10.2) Marshes in Poole Harbour, Dorset. Common cord-grass Spartina anglica saltmarsh has developed here since 1899, and this is backed in the upper reaches by Phragmites australis reedswamp, where salinity is reduced by freshwater inflow. Saltmarsh has been reclaimed by embanking, especially near the northern urbanized fringes. (After Bird, 1984, p. 214; based on original map by V.J. May, updated to 2000)

(Figure 10.3) Typical saltmarsh vegetation zonation: the dominant species found in England and Wales at each level are named in the boxes. In Scotland, the sandy saltmarshes are dominated by common saltmarsh-grass Puccinellia. (After Carter, 1988, p. 344.)

(Figure 10.4) Map of the saltmarsh at Gibraltar Point, Lincolnshire, recording the position in 1951. The marsh was growing on the landward side of the spit; the area was re-surveyed in 1959, by which time 15–30 cm of sediment had built up the marsh surface over most of the area, and the low-lying mud and sand of 1951 had been colonized by common cord-grass Spartina. (After King, 1972a, p. 428.)

(Figure 10.5) The distribution of Spartina anglica in England and Wales. (After Hubbard and Stebbings, 1967.)

(Figure 10.6) The velocities of a single water particle during a tidal cycle as it moves from a creek channel onto a mudflat surface. (After Pethick, 1984.)

(Figure 10.7) Stages in the evolution of a tidal creek as a saltmarsh encroachment takes place, forming terraces on either side of a deepening channel (B), the sides of which eventually become unstable (C). (After Bird, 1984, p. 213.)

(Figure 10.8) The intricate, dendritic creek network of a mature saltmarsh surface, Stacey, north Norfolk. (After Pethick, 1984, p. 159.)

(Figure 10.9) The distribution of saltpans on a saltmarsh at Blakeney Point, north Norfolk. (After Pethick, 1984, p. 164.)

(Figure 10.10) Cheniers and rates of change at (a) St Osyth Marsh (the arrows show the position of the northern end of the spit at different times) and (b) Dengie Marsh. Older cheniers occur landwards of the embankments at Dengie Marsh. Both the patterns of change in the marsh edge at Dengie Marsh and the spit at St Osyth show the tendency for these features to fluctuate in position, with erosion alternating with accretion. (Based in part on Greensmith and Tucker, 1975 and Burd, 1992.)

(Figure 10.11) Keyhaven Marshes. (1) Distal point of Hurst Castle spit; (2) saltpans; (3) major creek; (4) retreating saltmarsh edge and chenier formation; (5) dominant channel draining upper marsh. (Photo: courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 10.12) Location of the saltmarshes of the inner Solway Firth including the Upper Solway Flats and marshes on the south shore and the saltmarshes of the Solway Firth (north shore). The 2842 ha of saltmarsh found at these sites comprises 79% of all the saltmarsh in the Solway and 8% of all British saltmarshes. (After Pye and French, 1993.)

(Figure 10.13) A view looking south over Kirkconnell Merse on the west side of the River Nith towards Airds Point in the middle distance and Southerness beyond. The saltmarsh is grazed and is crossed by many well-developed creeks that drain to a prominent terrace along the Nith. Part of the saltmarsh of Caerlaverock can be seen on the east side of the river and to the south lie extensive sandflats. (Photo: P and A. Macdonald/SNH).

(Figure 10.14) Erosion and accretion of the Caerlaverock saltmarsh edge between 1946 and 1973. Eastward migration of the main channel of the Nith resulted in erosion of the western side of Caerlaverock. Between 1973 and 1999, the channel had migrated back to the west of the bay and approximately occupied its 1946 route. In spite of this, Harvey (2000) has shown that erosion continues to dominate the west side of the salt-marsh (see Figure 10.15). In the east, close to the exit of the Lochar Water, accretion is the long-term trend. (After Rowe, 1978.)

(Figure 10.15) Erosion on the west side of Caerlaverock saltmarsh. 1 m-high erosional terraces cut into the Puccinellia-dominated high-marsh surface are common on this shore but in places exposed to south-westerly wave activity, they are also subject to further surface-stripping of vegetation. (Photo: J.D. Hansom).

(Figure 10.16) Elevated concentrations of the radionuclide 137Cs occur at varying depths beneath the saltmarsh surfaces of all of the Scottish Solway saltmarshes, including at Caerlaverock. The time-integrated profile of Sellafield discharges shown here, comes from nearby Southwick saltmarsh, and shows peak concentrations at 0.30 m depth below the high-marsh surface, which represent input of sediments from the outer firth that peaked in 1978 and have declined since. These data can be used to calculate a sedimentation rate over time that can be compared with direct measurements using sedimentation plates or pins. (After Harvey, 2000.)

(Figure 10.17) Migration of the Rivers Esk and Eden over the period 1846–1973 has contributed to rapid accretion and westward migration of Rockcliffe Marsh that continues today. A healthy supply of sediment comes from the extensive nearshore and intertidal sandflats, augmented by fluvial sediment from the two rivers. (After Rowe, 1978.)

(Figure 10.18) Saltmarshes and emerged carse surfaces in the Cree estuary, Wigtown Bay. The estuary is shallow and extensive sandflats are exposed at MLWS. Muddier sediments are restricted to the tidal reaches of the Cree River itself in the north where small saltmarshes fringe its course, particularly on the inside of meander loops.

(Figure 10.19) A view looking south over the meanders of the Cree estuary towards Jultock Point in the distance on the right (west) side. Extensive sandflats are visible to the south of the forest in the middle distance. Emerged carse surfaces (old saltmarsh deposits) dominate the foreground, separated from the present salt-marsh by small cliffs that can be seen in places along the main river channel. (Photo: P. and A. Macdonald/SNH.)

(Figure 10.20) The linear saltmarshes at the Crook of Baldoon, on the western side of the Cree estuary appear to have undergone rapid accretion since the edge was mapped in 1847 but have mainly undergone edge retreat since 1973. Landward embankments along much of the Cree testify to land-claim for agricultural purposes in the past. Linear creeks at Baldoon have been artificially straightened and deepened.

(Figure 10.21) The coastal geomorphology of Loch Gruinart, Islay is dominated by a history of sea-level changes with emerged erosional and depositional features flanking the north-south axis of the loch. In the shelter provided by Ardnave and Killinallan Points, extensive linear and loch-head saltmarshes have developed, some of which are extending onto intertidal gravels. (After Ritchie and Crofts, 1974.)

(Figure 10.22) The changing coastline of the Loch Gruinart–Loch Indaal area, Islay, at 7000, 9000, 11 000 and 13 000 years BP showing phases of marine inundation and land emergence. Since 7000 years ago the relative sea level has shown a more or less constant falling trend towards the position of the present coastline. (After Dawson and Dawson, 1997.)

(Figure 11.1) The GCR site of Culbin is a large and complex gravel strandplain composed of suites of partially visible emerged ridges capped by large sand dunes. The pattern of the underlying gravels can be reconstructed into a series of westward-extending gravel spits; the updrift erosion of the earlier spits fed the downdrift extension of the more recent ones that have been dated using historical maps and aerial photographs. Extensive sandflats and saltmarshes have developed in the shelter of the westward-extending spits. See Figure 11.2 for the section X—Y. (After Hansom, 1999.)

(Figure 11.2) The gravel ridges over a 1000 m transect from x–y (see Figure 11.1) show two groups of emerged ridges, the most seaward of which decline rapidly in height towards the north-west. (After Comber, 1995.)

(Figure 11.3) Parabolic sand dunes undergoing erosion at the exit of the River Findhorn, Culbin. Erosion of this section of the Culbin foreland feeds sand to fuel accretion at the downdrift Buckie Loch spit. Harvesting of timber over a 20–30 m-wide dune edge zone is part of a management regime designed to reduce erosion caused by disruption of the dune surface by toppling, and to allow mechanical harvesting to be carried out in safety (Photo: J.D. Hansom.)

(Figure 11.4) The magnificent gravel spit of The Bar at Culbin is extending westwards towards the town of Nairn at approximately 15 m a−1. The sandy Buckle Loch spit can be seen in the upper middle distance and the narrow erosional neck at the eastern end of The Bar. This updrift erosion has truncated earlier ridges and is now encroaching into the area of saltmarsh that has developed behind the bar and may ultimately threaten the larger area of saltmarsh that lies in the right foreground. (Photo: P. and A. Macdonald/SNH.)

(Figure 11.5) Spectacular recurves extend from the active outer beach ridges at The Bar at Culbin into the sheltered area behind. The inner recurved parts of the gravel ridges support small areas of heather, gorse, broom and pine whereas the intertidal flats between the gravel ridges support small areas of saltmarsh (see Figure 11.4). (Photo: P and A. Macdonald/SNH.)

(Figure 11.6) An extensive area of saltpans and linear creeks characterize the area of saltmarsh that has developed at the heads of the two intertidal lagoons that lie either side of the area where The Bar at Culbin is attached to the mainland (see Figure 11.4). (Photo: J.D. Hansom.)

(Figure 11.7) Morrich More is a series of emerged sand beach ridges adorned with sand dunes, which together form a progradational strand-plain jutting out into the Dornoch Firth. Saltmarsh interdigitates between the sand ridges in the north and east of the structure. The numbered coastlines correspond to the approximate locations of reconstructed and actual coastal positions. Heights of the sand ridges are given in metres above OD. (After Hansom, 1999.)

(Figure 11.8) Aerial view of Morrich More from the north-west. The islands of Patterson and Innis Mhòr (the eastern end of which is seen to the right) are separated by a tidal channel, which connects with a large area of sand accretion in the shelter provided by Innis Mhòr. The inlet of Inver Bay in the centre middle distance supports large areas of saltmarsh, which interdigitates between sand-beach ridges capped with blown sand. (Photo: P and A. Macdonald/SNH.)

(Figure 11.9) A remotely sensed image of the narrow upper beach and eroding edge on the western flank of Morrich More showing the large fixed parabolic dunes and the area of low dunes downwind. Intertidal bars are well-developed on the 1 km-wide intertidal flats on the Dornoch Firth side of Morrich More, and may indicate the direction of sediment movement under the flood tide (onshore) and ebb tide (alongshore to the northeast). North is at the top of the image. The arrow indicates direction of migration of the parabolic dunes. (After Hansom and Leafe, 1990.)

(Figure 11.10) The large parabolic dunes on the western flank of Morrich More have migrated north-east (towards the camera) but have since stabilized, mainly by marram colonization. The low hummocky dunes in the foreground have been influenced in the past by sand blown from the parabolic dune field but are now also stable and covered mainly with marram and smaller areas of heather. (Photo: J.D. Hansom.)

(Figure 11.11) Former coastal positions of Morrich More, based mainly on historical sources and Ordnance Survey maps and aerial photography. The north-east coast has extended by about 1 km since 1730, but the western flank has eroded by varying amounts over the same period. Sediment eroded from the west is moved north-east by tidal streams and waves at high tide to be deposited in the area behind Innis Mhòr and Patterson Island. (After Hansom and Leafe, 1990.)

(Figure 11.12) Sketch map of the key geomorphological features and sediment transfers of Carmarthen Bay. See also Figure 11.17 for details of the Rhossili Bay area. (Offshore transfers derived in part from Bather and Thomas, 1989.)

(Figure 11.13) Variations in accretion and erosion since 1950, Carmarthen Bay.

(Figure 11.14) The Devensian geomorphology of Carmarthen Bay, with stylized sections through selected emerged beach sites. (After Campbell and Bowen, 1989.)

(Figure 11.15) Pembrey: older dunes can be seen to the right of the photograph — these are now conifer plantations. Post-19th century accretion is evident in the middle and left background. A blowthrough is present in the foreground. (Photo: V.J. May.)

(Figure 11.16) A deeply incised saltmarsh channel in the muds of Llanrhidian saltmarsh, Loughor Estuary. (Photo: J.D. Hansom.)

(Figure 11.17) Geomorphological features of Rhossili Bay and Whiteford Burrows.

(Figure 11.18) Rhossili Bay seen from the Carboniferous Limestone headland looking north towards Burry Holms and Llangennith Burrows. Rhosilli Down to the right of the photograph is fronted by low cliffs formed in periglacial head and fluvioglacial material. Erosion of these cliffs feeds the narrow fringing cobble beach. The wide intertidal beach is mainly sand. (Photo: J.D. Hansom)

(Figure 11.19) Limestone intertidal solution features near Worms Head, typically up to 0.3 m in height. (Photo: V.J. May.).

(Figure 11.20) a and b Sketch map of the key features of the area of Newborough Warren and Moth Dinlle. (b) Detail of the entrance to the Menai Strait (Photo: courtesy Cambridge University Collection of Aerial Photographs Countryside Council for Wales.)

(Figure 11.21) Key features of Newborough Warren. The western part of the site has a series of sand spits and dunes extending from a former rocky shore, of which Llanddwyn Island (Ynys Llanddwyn) is a seaward extension. The eastern shore changes from low, climbing dunes on a rock base to an extensive area of migrating dunes on a sandplain. The Abermenai spit owes its form in part to strengthening by an artifical breakwater.

(Figure 11.22) Aerial photograph of part of Newborough Warren (see Figure 11.21 for location), which shows the contrasts between the coastal active and migrating dunes, the linear sub-parallel ridges that extend inland and area of mostly stabilized parabolic dunes. Intertidal sand ridges may provide a pathway for sand transport feeding the dunes in the east, but gravels underlie parts of the intertidal area in the west. (Photo: courtesy Cambridge University Collection of Aerial Photographs © Countryside Council for Wales.)

(Figure 11.23) Sketch map of the key geomorphological and historical changes to Holy Island. Bold arrows show the dates of the main channels draining estuaries.

(Figure 11.24) Cross-section and profile of the main geomorphological features of Budle Bay and Ross Links. (based in part on interpretation of aerial photographs, see Figure 11.25.)

(Figure 11.25) Aerial photograph mosaic showing the main features of Ross Links and Budle Bay. 1, cliff-foot dunes; 2, sand-waves in Budle Bay; 3, intertidal sandflats and mudflats; 4, Budle spit; 5, prevailing wave direction; 6, saltmarsh and intertidal mudflats; 7, possible former beach ridges; 8, dunes of Ross Links; 9, linear shore-parallel dunes decreasing in altitude towards shoreline. (Photo courtesy Cambridge University Collection of Aerial Photographs, Crown Copyright, Great Scotland Yard.)

(Figure 11.26) Coves Haven, on the northern coast of Holy Island. The underlying Carboniferous Limestone is covered by till and emerged beach sediments, which are covered by dunes aligned from north-west to south-east. (Photo copyright English Nature.)

(Figure 11.27) Summary features and recent dynamics of the North Norfolk Coast GCR site from Hunstanton to Sheringham, and east to west distance of about 40 km. (After Halcrow, 1988.)

(Figure 11.28) (a) The distinctive tabular chalk cliffs of Hunstanton, looking north. (b) Cartoon of potential transport mechanisms of rockfall debris from the failure of the chalk cliffs. (Photo: V.J. May.)

(Figure 11.29) (a) Brancaster beach-dunes, sand and shingle beach with regular shore-normal cusps. (b) Eroded dune-face remnants of World War II defence structures and associated retreating foredune scarp. (Photos: V.J. May.)

(Figure 11.30) (a) Scolt Head Island geomorphological features. (Based mainly on Steers, 1946b, 1960; Bird, 1984, 1985; Halcrow 1988.) (b) Blakeney Point geomorphological features. (Based mainly on Steers, 1946b; Bird, 1984, 1985; Flalcrow, 1988.)

(Figure 11.31) Relationship of saltmarsh elevation to tides and age of the marsh surface at Missel Marsh, Lower Hut Marsh, upper Hut Marsh and Hut Marsh. H = 2.385 – 1.411 e – 0.014t is a best-fit line based on the relationship between marsh height (H), tides per year (e) and age of marsh surface (t). (After Pethick, 1980b; 1981.)

(Figure 11.32) Geological map of the Dorset coast from Lulworth Cove to Studland Bay.

(Figure 11.33) Summary geomorphological character of the coast between Furry Cliff and Poole Harbour.

(Figure 11.34) Cliff retreat at Furzy Cliff

(Figure 11.35) Cross-sections of the beaches of Ringstead Bay, and sketch map showing locations of sections.

(Figure 11.36) Geomorphology and cross profile of the White Nothe landslides in the Dorset Coast GCR site.

(Figure 11.37) Sonargraph of the seabed between White Nothe and Bat's Head, showing an area of tilted strata (dipping from north-east to south-west) and joint and block patterns. The upper right hand area shows blocks broken from the strata. (Image by permission of C. Heeps.)

(Figure 11.38) Profiles of the submerged rock ridges between Bat's Head and Worbarrow Tout. The ridges are formed by steeply dipping Purbeckian and Portlandian strata between the headlands along this coast. Typically the northern face of the ridges is formed by the structural surface of strata, which here dip northwards (inland). Opposite Lulworth Cove, the ridges are less than 1 m in height, whereas in Worbarrow Bay they attain 9 m. C = Cave; A = Arch; hanging dry valleys are shown with a bracketed arrowhead, whereas valleys draining to sea level are shown without a bracket. Locations of the profiles 1 to 6 are indicated in the plan view.

(Figure 11.39) Sonargraph of the Worbarrow Bay submerged rock reef. The echo-sound profile shows the depth profile of the seabed and the distinct, steep, backwall of the ridge. Either side of the echo-sound trace, the sonar image shows the seabed patterns of the individual ridges formed by the eroded, dipping strata. The alignment of the sand ripples indicates that sand movement is alongshore. (Photo reproduced by permission of C. Heeps.)

(Figure 11.40) Changes in shingle beach profiles in Worbarrow Bay, Ringstead Bay and Durdle Door. Erosion and recovery of beaches can take place over very short timescales. These profiles are based upon monthly re-surveys of sample transects. (After Heeps, 1986.) can be as steep as 30°. These steep profile areas are the most exposed and are composed of the coarsest and best-sorted sediments (Arkell, 1947; Heeps, 1986).

(Figure 11.41) Looking eastwards across Worbarrow Bay, showing the outcrop of Purbeck and Portland beds at Worbarrow Tout, the Wealden cliffs undergoing erosion and the shingle storm-beach with cusp development. (Photo: V.J. May.)

(Figure 11.42) Gad Cliff. The upper cliff is in Portland Stone; the debris and boulder field, well-vegetated by scrub, lies on Portland Sand and upper Kimmeridge Clay. The boulder beach has alternating ramparts and baylets related to differential erosion (associated with bedding in the Kimmeridge Clay) and debris toes. (Photo: V.J. May.)

(Figure 11.43) Broad Bench, a Kimmeridge shale platform, showing block removal at platform edge. (Photo: V.J. May.)

(Figure 11.44) The cliffs between Hounstout and St Aldhelm's Head are characterized by large landslides of as yet undetermined age. Large Portland Stone boulder fields provide protection against wave attack, but where they are absent shoreline retreat produces steep lower cliffs. (Photo: V.J. May)

(Figure 11.45) Emmet's Hill and St Aldhelm's Head cross-profiles. For locations of profiles A–C, see Figure 11.47.

(Figure 11.46) Landslide features between Hounstout and St Aldhelm's Head. (After May, 1997a.)

(Figure 11.47) Truncated valley infill at Seacombe Valley [SY 983 785]. The altitude of the valley floor is about 7 m. (Photo: V.J. May.)

(Figure 11.48) Development of Worbarrow Bay from initial flooding of former valley along line of Arish Mell gap. The breakthrough in the vicinity of Mupe Rocks develops in early stages in similar way to that at Stair Hole. Subsequently, both Worbarrow and Mupe bays merge with increasing asymmetry as the eastern shoreline of Worbarrow Bay becomes more exposed to waves from English Channel.

Tables

(Table 1.1) Number of items in the computerized bibliography of geomorphology of Britain that are classified as 'Coasts' (total 1400), by year of publication.

(Table 1.2) Number of items under selected keywords (some items appear more than once as several keywords are allocated to each).

(Table 1.3) Geographical analysis of the British coastal literature, using selected grid squares only.

(Table 1.4) General order of resistance to erosion of British rock types (from Clayton and Shamoon, 1998).

(Table 1.5) Morphosedimentological classification of the British coast (based on European Commission (1998 – the CORINE project érosion cotieré).

(Table 1.6) Main features of each GCR Site, broadly following the classification of King, 1978, to show where different features are represented.

(Table 1.7) Coastal Annex I habitats occurring in the UK (from McLeod et al., 2002.)

(Table 2.1) Likely recession rates in different materials (compiled by Carter, 1988, from data in Sunamura, 1983).

(Table 2.2) Primary, secondary and tertiary controls on cliff form (based on May, 1997a).

(Table 2.3) Candidate and possible Special Areas of Conservation in Great Britain supporting Habitats Directive Annex I habitat 'Vegetated sea cliffs of the Atlantic and Baltic coasts' and/or 'Submerged or partially submerged sea caves' as qualifying European features. Non-significant occurrences of these habitats on SACs selected for other features are not included. (Source: JNCC International Designations Database, November 2002.)

(Table 3.1) Hard-rock cliff GCR sites, including those sites described in other chapters of the present volume that include hard-rock cliffs in the assemblage.

(Table 3.2) Altitude and orientation of some cliff-top boulder deposits in Shetland (after Hansom et al., in press).

(Table 4.1) The main features of soft-rock cliff coastal geomorphology GCR sites, including coastal geomorphology GCR sites described in other chapters of the present volume that contain soft-rock cliffs in the assemblage. Sites described in the present chapter are in bold typeface.

(Table 4.2) Rates of cliff-top retreat of soft-cliffed coasts (from various sources).

(Table 4.3) North Yorkshire coast cliff retreat rates in m a−1 (based on Agar, 1960).

(Table 4.4) Land-loss by natural sections of the Holdemess coast, 1852–1952 (Valentin, 1954, 1971).

(Table 5.1) Classification of beach structures based on their plan form (after Pethick, 1984); outline definitions are provided in the glossary of the present volume.

(Table 6.1) Main features and sediment sources of gravel/shingle beach and ness GCR sites, including coastal geomorphology GCR sites described in other chapters of the present volume that contain shingle beach/ness structures in the assemblage.

(Table 6.2) Candidate and possible Special Areas of Conservation in Great Britain supporting Habitats Directive Annex I habitat 'Perennial vegetation of stony banks' and/or Annual vegetation of drift lines' as qualifying European features. Non-significant occurrences of these habitats on SACS selected for other features are not included. (Source: JNCC International Designations Database, July 2002.)

(Table 6.3) Westerly extension of the active gravel beach (West Spey Bay). (From Gemmell et al., 2001b.)

(Table 6.4) Development phases at Dungeness. Ridge height data are mainly from Lewis and Balchin (1940).

(Table 7.1) Main features and present-day sediment sources of dune types. Exemplar sites described in the present chapter are in bold typeface. See also Table 7.2. (Based on Ranwell, 1972.)

(Table 7.2) Main features, sediment sources, tidal ranges of sandy beach and dune GCR sites, including coastal geomorphology GCR sites described in other chapters of the present volume that contain dune features in the assemblage. It should be noted that all of the machair sites in Chapter 9 have dune features (see Table 9.1). Sites described in the present chapter are in bold typeface.

(Table 7.3) Calcium carbonate content of upper beach/foredune in selected coastal geomorphology GCR sites. Sites described in the present chapter are in bold typeface. (Based in part on Goudie, 1990, and various sources cited by Ritchie and Mather, 1984.)

(Table 7.4) Variations in calcium carbonate content and pH in foredunes and main dunes. (Based on Salisbury, 1952; and Willis, 1985)

(Table 7.5) Candidate and possible Special Areas of Conservation in Great Britain supporting Habitats Directive Annex I coastal dune habitat(s) (other than machair) as qualifying European features. Non-significant occurrences of these habitats on SACs selected for other features are not included. (Source: JNCC International Designations Database, July 2002.)

(Table 8.1) The main features of sediment sources and tidal ranges of sand spit GCR sites, including coastal geomorphology GCR sites described in other chapters of the present volume that contain important sand spit structures in the assemblage of features. Many machair sites have small sandspits — see Chapter 9. (Sites described in the present chapter are in bold typeface )

(Table 8.2) Area of East Head — historical data from 1846 to 1996

(Table 9.1) Machair GCR sites

(Table 9.2) Candidate Special Areas of Conservation supporting Habitats Directive Annex I habitat 'Machair' as a qualifying European feature. (Source: JNCC International Designations Database, July 2002.)

(Table 10.1) Candidate and possible Special Areas of Conservation in Great Britain supporting Habitats Directive Annex I coastal saltmarsh habitat(s) as qualifying European features. Non-significant occurrences of these habitats on SACs selected for other features are not included. (Source: JNCC International Designations Database, July 2002.)

(Table 10.2) Characteristic geomorphological features of some of the main Solway Firth saltmarshes.

(Table 10.3) Estimated areal accretion in hectares between 1864 and 1946, 1946 and 1973, 1973 and 1993 for selected inner Solway saltmarshes. (Based on data from Marshall, 1962; Rowe, 1978 and Pye and French, 1993.) All areas in ha. Caerlaverock Marsh is in the Solway Firth (north shore) GCR site.

(Table 11.1) Main geomorphological features of the 'Coastal Assemblage' GCR sites.

(Table 11.2) CORINE categories, data for the Carmarthen Bay, North Norfolk Coast, Purbeck (Dorset Coast) and Newborough Warren/Morfa Dinlle GCR sites; measurements are in km.

(Table 11.3) Summary of saltmarsh development in north Norfolk.

(Table 11.4) Rates of cliff-top retreat since c. 1900 on the Dorset Coast.