You are on page 1of 19

10.

11 Rock Coasts
WJ Stephenson, University of Otago, Dunedin, New Zealand
ME Dickson, University of Auckland, Private Bag, Auckland, New Zealand
AS Trenhaile, University of Windsor, Ontario, Canada
r 2013 Elsevier Inc. All rights reserved.

10.11.1 Introduction 289


10.11.2 Processes 290
10.11.2.1 Sea Level 291
10.11.2.2 The Role of Tides 291
10.11.2.3 Waves 291
10.11.2.3.1 Wave quarrying 291
10.11.2.3.2 Abrasion 292
10.11.2.4 Weathering 292
10.11.2.4.1 Salt weathering 293
10.11.2.4.2 Wetting and drying 293
10.11.2.5 Biology 293
10.11.2.6 Mass Movement 294
10.11.3 Rocky Coast Landforms 295
10.11.3.1 Relict, Composite, and Plunging Sea-Cliffs 296
10.11.3.2 Shore Platforms 297
10.11.3.2.1 Near-horizontal platforms 298
10.11.3.2.2 Sloping platforms 299
10.11.3.3 Inheritance 300
10.11.4 Rock Coast Modeling 300
10.11.4.1 Coastal Profiles 300
10.11.4.2 Plan Shape 302
10.11.5 Conclusions 303
References 303

Abstract

Rock coasts constitute the majority of the worlds shorelines and result from erosive processes. Rates of erosion vary from
almost nil to tens of meters per year. Cliff and shore platforms are the most common landform types but even these display
significant variations in morphology depending on sea-level history, tidal range, the relative contributions of marine and
subaerial processes, and geological properties of the rock. Smaller scale features such as stacks and arches are important
elements and contribute significantly to landscape values and subsequently, tourism economies. The evolutionary history of
rock coasts is difficult to reconstruct because of their erosional origin but modeling has revealed much about how such
coastlines have developed over the latter part of the Pleistocene and Holocene, and it has also provided important insights
into the modeling of these coasts to modern sea-level and coastal processes.

10.11.1 Introduction rock strength occurs along a very wide continuum. Sea cliffs
are the most characteristic landforms along the rocky shore-
Rock coasts are ubiquitous along the worlds coast lines, oc- lines. According to Emery and Kuhn (1982), cliffs, including
curring at all latitudes and in all morphogenetic environ- those cut into hard and soft rocks, as well as cohesive clays and
ments. Rock coasts can be defined as those that are composed sands, occur along about 80% of the worlds coasts (Figure 1),
of consolidated material (rock) of varying strength (Suna- although little work has been done to substantiate this figure.
mura, 1992). The strength of material forming rock coasts can Rock coasts may also be composed of shore platforms that
vary enormously from very hard granite to soft materials such either slope (411) or lie close to horizontal. Shore platforms
as clay and consolidated gravels. The common distinction are commonly front cliffs but in some cases platforms are
between hard and soft rock coasts is thus misleading, because backed by beaches and sand dunes. Unlike coasts composed
of accumulations of sediment, rock coasts are almost entirely
erosional in origin and consequently reconstructing the evo-
Stephenson, W.J., Dickson, M.E., Trenhaile, A.S., 2013. Rock coasts. In:
lutionary history of rock coast landforms presents significant
Shroder, J. (Editor in Chief), Sherman, D.J. (Ed.), Treatise on
Geomorphology. Academic Press, San Diego, CA, vol. 10, Coastal challenges to earth scientists. Furthermore the erosional origin
Geomorphology, pp. 289307. of rock coasts also presents significant management challenges

Treatise on Geomorphology, Volume 10 http://dx.doi.org/10.1016/B978-0-12-374739-6.00284-0 289


290 Rock Coasts

Figure 1 Distribution of cliffed coasts around the world. Adapted from Emery, K.O., Kuhn, G.G., 1982. Sea cliffs: their processes, profiles, and
classification. Geological Society of America Bulletin 93, 644654, with permission from Geological Society of America.

as rock coasts lack the capacity to recover from erosive events determinant of where coast lines occur, so it is first necessary
in the way beaches or other sedimentary environments return to examine the role of changing sea levels in the evolution of
sediment to their form. Paradoxically the scenic value attached rock coasts. Closely associated with sea-level changes are tides,
to rock coasts owes its origin to that same suite of erosional because these determine the vertical range within which rock
processes, so management and conservation requires main- coasts develop, and, to a certain degree, rock coast morph-
taining erosive processes that are not attractive in most coastal ology. The traditional role of waves and weathering are ex-
management strategies. plored. In addition, the roles of biological processes are also
Rock coasts are also at the center of several key debates in considered. Section 10.11.3 reviews how geological and
coastal geomorphology. For example, Cruslock et al. (2010) lithological frameworks determine landform morphology.
considered shore platforms in different genetic settings to Different rock coast landforms are presented and explained
demonstrate equifinality. The environmental significance of within a geological framework. Reconstructing the evolution-
boulder deposits are beginning to be better understood by the ary history of rock coasts is problematic because of an ero-
examination of cliff top deposits (Hansom et al., 2008) and sional origin, so that modeling is currently the best approach
viewed more critically through the consideration of geological to understand how rock coasts have evolved and how they
settings (Stephenson and Naylor, 2011a, b; Knight and Bur- might evolve into the future. Consequently Section 10.11.4
ningham, 2011). Large boulders deposited on cliff tops or far reviews modeling efforts that bring together our current state
inland beyond the reach of normal sea conditions have previ- of understanding of how rock coasts develop.
ously been attributed to tsunami (Bryant et al., 1992, 1996;
Young et al., 1996). Work on rock coasts is illustrating how
boulders and even megaclasts can in fact be attributed to (the 10.11.2 Processes
work of) extreme storm events (e.g., Hall et al., 2006, 2008;
Etienne and Paris, 2010; Switzer and Burston, 2010). Thus de- Rock coasts are exposed to a wide range of processes that are
bates over the role of tsunami as an agent of change of coastal unique in terms of geomorphic process response systems, es-
landscapes are becoming better understood and the role of pecially when compared to other coastal systems. In addition
storm waves is better recognized. What is yet to be understood to marine processes such as wave impact, quarrying, and
is whether these high magnitudelow frequency events make a abrasion, subaerial processes may reduce rock strength and in
significant contribution to long-term rock coast development. some settings can be the dominant process set. Subaerial
Today there are still many challenges to understanding the re- processes include the full range of physical and chemical
sponse of rock coasts to contemporary processes across a variety weathering processes and the relative significance of any par-
of scales and to interpreting the long term response, or evo- ticular one (e.g., wetting and drying, salt weathering, heating
lution, of these coasts (Naylor et al., 2010). and cooling, solution) is variable in time and space. In add-
In this chapter, we review these debates in rock coast ition to marine and subaerial processes, biological activity is
geomorphology and provide an overview of the role of dif- also important. Included here are not only erosion processes
ferent process and geological frameworks that lead to rock but also bioprotection that can act to reduce the efficiency
coast landforms. The evolutionary pathways that rock coasts of physical processes and construction where organisms con-
take are explored through the use of models. In Sec- struct reefs or accumulate sediment. Where rock coasts are
tion 10.11.2 we discuss the processes that drive rock coast wholly or partly composed of cliffs, mass movement is an
landform development. Sea level is the fundamental important process that causes cliff erosion and retreat.
Rock Coasts 291

10.11.2.1 Sea Level 10.11.2.3 Waves


Sea level is the primary control of coast evolution as it is the Waves impacting on rock coasts impart a variety of forces, but
basic spatial and temporal frame of reference upon which understanding the erosional efficiency of these forces still re-
processes responsible for shoreline development occur. Rock quires a great deal of empirical research. Until recently there
coasts evolve slower than sedimentary coasts because of the have been relatively few field studies reported in the literature.
greater resistance of their material composition, but as a Undoubtedly waves erode rock coasts and transport the wea-
consequence there is a much greater potential for rock coasts thered and unconsolidated debris, but the precise quantifi-
to be preserved in the landscape after sea-level changes. This cation of the relationship between wave forces and rock
has two important implications. First, many rock coasts may resistance, as expressed by Sunamura (1992, 1994), remains
not change rapidly enough to record all sea-level variations, elusive. Sunamura proposed that rock coast erosion occurs
particularly short term Holocene variations. Second, some when the erosive force of waves exceeds the resisting force of
rock coasts may have undergone more than one episode of rock. The challenge that remains is to accurately quantify both
development when sea-level reoccupies a similar elevation, or forces. Waves can erode by (1) directly expending a force (water
tectonic processes lift submerged shorelines to coincide with a hammer) by hydraulic quarrying of jointed blocks, or (2) by
contemporary sea level. The challenge for scientists investi- abrasion when waves mobilize loose sediment against bedrock.
gating rock coasts is to recognize the impact of multiple sea Our understanding of wave processes on shore platforms
levels occupying rock coasts and identify the morphologies has been hampered by a lack of field investigations. In recent
that have been inherited from earlier sea-level histories. times the number of field investigations has grown and new
Patterns of Quaternary sea-level change (2.5 Ma present) data reveal that the hydrodynamics of waves across shore
are generally known although the details of sea-level durations platforms are as complex as the surf zones of beaches. Work by
and elevations at specific sites are not (Pirazzoli, 1991). Sea- Stephenson and Kirk (2000a) considered how energy was
level change since the last two glacial periods are of most sig- transformed across platforms, noting the highly dissipative
nificance to rock coast development although some rock coasts nature of the platforms they studied. Taylor (2003) was the
may be inherited from sea levels dating back to the Tertiary, first to record wave energy in the infragravity band on shore
652.5 Ma ago (Trenhaile, 1989). It is likely that highly resist- platforms. Farrell et al. (2009) demonstrated that the wave
ant rock coasts have retained some degree of inheritance from breaking criterion of 0.42 h (water depth) for inner surf zone
the last interglacial at approximately 125 ka (Trenhaile, 2002). root-mean-squared wave height (Hrms) proposed by Thornton
More recently Trenhaile (2010a) has endeavored to examine the and Guza (1982) applied to the platform they studied in
role of relative sea-level rise during the Holocene, and this is Portugal. Prior to Taylor (2003) no attention had been given
considered further in Section 10.11.4. to infragravity energy or consideration of the geomorphic
consequence of such energy for rock coasts. This has changed
recently with a number of published studies examining wave
10.11.2.2 The Role of Tides transformations and energy dissipation across shore platforms
(Ogawa et al., 2011; Ogawa et al., 2012; Beetham and Kench,
If sea level determines where and for how long coastal processes
2011; Marshall and Stephenson, 2011). Unsurprisingly, infra-
act on shorelines, tides determine the vertical range within
gravity energy is present on shore platforms during rising and
which those processes operate. Even the location of supratidal
high tides when the platform acts to modulate incident wave
processes are partially determined by tidal range. Globally, rock
energy in a similar way as the surf zones of beaches excite
coast morphologies develop in all tidal environments and
infragravity motions. It remains to be seen what, if any, con-
vertical ranges. In addition, rock coast landforms develop in
sequence there is for the geomorphic evolution of rock coasts
lakes where tides are absent (Matthews et al., 1986; Allan et al.,
as a result of infragravity energy.
2002; Trenhaile, 2004a; Aarseth and Fossen, 2004). Lake levels
can fluctuate both naturally and from management inter-
vention (as in the case of hydropower generation) but whether
these fluctuations mimic tides with a geomorphic consequence 10.11.2.3.1 Wave quarrying
for lacustrine rock coasts remains unclear. Wave quarrying is thought to be the most important erosive
Tidal range has been identified as a primary control of process on shore platforms in the storm wave environment of
gross morphology of shore platforms. Trenhaile (1999) re- the northern hemisphere (Trenhaile, 1980, 1987). Quarrying
viewed the published data on shore platform gradient and is the breaking free and removal of rock fragments by shock
tidal range for a variety of coasts around the world. He found pressures from breaking waves, water hammer, and the com-
that there is a strong positive relationship between these pression of air in rock joints. Although shock pressures and
variables (r2 0.85), with steeper sloping platforms occurring water hammer are generated by breaking waves, both breaking
in areas with the highest tidal ranges, including in the Bay of and broken waves can cause compression of air in joints. Air is
Fundy in eastern Canada and the Severn Estuary in south compressed in joints and bedding planes as water rushes in,
western Britain. In addition, tidal range and type (diurnal, followed by a sudden explosive release as the water recedes
semidiurnal, or mixed) determine how and where wave energy (Trenhaile, 1987). Robinson (1977a, b) proposed that
is distributed within the vertical range of the tide and the quarrying becomes more effective when sand is washed into
horizontal extent of the shore profile. Fluctuations in baro- cracks and keeps them open. He called this process wedging.
metric pressure, wind, and wave set up can also change the Wave quarrying is limited in extent over a shore platform,
vertical range within which wave forces are expended. since maximum pressures occur at or near the mean water
292 Rock Coasts

level whose position on a shore platform and in front of a cliff abrasion as the result of sweeping, rolling, or dragging of
is controlled by the tide and wave set up (Trenhaile, 1987). rocks and sand across gently sloping rock surfaces, or the
The effectiveness of quarrying has been assessed largely on throwing of coarse material against steep surfaces. Abrasion
the basis of morphological evidence such as fresh block scars relies on the presence of sediment and sufficient wave energy
and angular blocks on platforms and in front of cliffs. Few to move it, and it generally occurs in a narrow zone, the extent
studies have attempted to determine the effectiveness of of which is controlled by the character of the tide and wave
quarrying using field data. This is largely because of the environment, and the accumulated sediment. Conversely,
difficulty of working in a very dynamic environment and Sunamura (1976) noted that large or very thick accumulations
problems with the availability of suitable instrumentation. of sediment can prevent abrasion by acting in a protective
Trenhaile and Kanyaya (2007) measured waves on platforms manner. Robinson (1977a, b, c) found that abrasion occurred
in the macrotidal Bay of Fundy on the Atlantic coast of North on a shore platform in northeastern England at a height of
America and the micromesotidal coast of Gaspe, Quebec, and 10 cm above a beach at a cliff foot and extended 14.5 cm
concluded that calculated wave stresses are capable of re- below the beach. Waves moved sediment at a depth of 5 cm
moving large rock blocks from the platform in this area. In- and as deep as 13.5 cm under the largest waves (Robinson,
stantaneous recordings of block removal are required to 1977c). Robinson (1977a, b, c) found that abrasion was more
improve our knowledge of this process but since it most often effective during winter when storms are of greater magnitude
occurs under during storm conditions, it is difficult to obtain and frequency. Abrasion rates in this setting were as high as
real time observations. Cruslock et al. (2010) used a block scar 4.5 cm yr1. Blanco-Chao et al. (2007) demonstrated how
inventory to illustrate the removal of blocks on the Swedish abrasion was effective on shore platforms where the landward
and Faro Island coasts in the Baltic Sea. Systematic repeat margin of the platform was buried by deposits of fluvio-nival
surveys revealed the removal of blocks over a 4-year period. and periglacial slope deposits. Sediments, liberated from the
Naylor and Stephenson (2010) used repeated photography slope deposits by erosion were made available in the upper
following storms to demonstrate the removal of blocks from a part of the tidal range as abrasive tools. In this zone, abrasion
platform surface on the Glamorgan coast in South Wales. Hall rates were between 0.13 and 1.8 mm yr1. These rates are
et al. (2006), Hansom et al. (2008), and Hall et al. (2008) similar to those measured with microerosion meters, reported
used a combination of historical photography, mapping, from other shore platforms where abrasion is absent (Ste-
dating, and modeling to show the effectiveness of wave phenson and Finlayson, 2009).
quarrying on cliffs and cliff tops in the British Isles. Blocks The significance of abrasion to shore platform develop-
deposited at the 15 m elevation on The Grind of the Navir on ment is unclear. Given that abrasion is a spatially and tem-
the Shetland Islands, Scotland, have been quarried from the porally discrete process, and works most effectively only where
cliff face during annual-frequency storms. Atlantic storm and when abrasive sediments accumulate, typically at the
waves have also been demonstrated as effective generators of landward margin of shore platforms, then large parts of shore
cliff top storm deposits by Suanez et al. (2009). Along the platforms are apparently eroded by other processes. Never-
coast of Banneg Island located in the archipelago of Mole`ne, theless, abrasion is significant in many areas because of rela-
in Brittany, France, wave quarrying has liberated blocks that tively high erosion rates (where they occur) and the landward
were subsequently deposited on top of cliffs. migration of the cliff-foot zone as shore platforms develop.
Boulders and megaclasts strewn along rock coasts on shore Conversely, many shore platforms have developed in the
platforms and cliff tops are frequently cited as evidence for the complete absence of abrasive materials.
occurrence of tsunami (Noormets et al., 2002, 2004; Scheffers
et al., 2005; Scheffers and Scheffers, 2006; Morton et al.,
2006). However such evidence remains controversial (Felton
10.11.2.4 Weathering
and Crook, 2003; Switzer and Burston, 2010) and in light of
recent work illustrating the ability of storm waves to deposit Coasts are perhaps the ideal weathering environment because
boulders on cliff tops the role of tsunami is being questioned of a plentiful supply of salts, water, temperature extremes, and
(Hall et al., 2006; Hansom et al., 2008; Hall et al., 2008; Paris bare rock surfaces. Although both physical and chemical
et al., 2011). Nevertheless, there are examples of very large weathering processes and biological activity occur, consider-
boulders located on cliff tops that are incapable of being ation of weathering usually focuses on (1) wetting and drying,
transported by storm waves, and tsunami remain the only (2) solution, and (3) salt weathering. Many rock types are
plausible explanation (e.g., Goto et al., 2010). Thus, both susceptible to chemical dissolution as well as physical fatigue
storm waves and tsunami are responsible for wave quarrying, from repeated wetting and drying by tides and aided by salt
but their relative contribution to long-term rock coast evo- water. In warmer environments wetting and drying can be
lution remains unclear. important while on higher latitude coasts freeze-thaw may be
more significant. The relative contribution of each type of
10.11.2.3.2 Abrasion weathering depends upon the suite of environmental con-
Abrasion has been identified as an important erosive process, ditions, so that the importance of each will be different spa-
particularly on shore platforms (Sunamura, 1976; Robinson tially and temporally. Seasonal variations are likely to be
1977a, b, c; Kennedy and Beban, 2005; Blanco-Chao et al., important in temperate environments. The suite of weathering
2007). Abrasion operates where sediments that are as hard or processes is also responsible for a number of micromeso scale
harder than the underlying bedrock accumulate and are landforms such as tafoni (Figure 2), pits, and model/karren
moved by wave action. Trenhaile (1987, p.25) defined (on limestone).
Rock Coasts 293

Figure 2 Honeycomb (tafoni) weathering. Vernier calipers are Figure 3 Water Morphology of water layer weathering morphology
230 mm long. of shallow pans surrounded by raised rims. The location of the raised
rims is determined by the occurrence of joints.

10.11.2.4.1 Salt weathering


Cooke and Smalley (1968) identified three mechanisms by Sedimentary rocks such as shales and mudstones are particu-
which salt causes weathering: (1) pressures exerted by crystals as larly susceptible to this form of weathering because the clay
they grow from solution, (2) pressures exerted by expanding minerals found in them expand on wetting and shrink on
salt crystals due to heating, and (3) pressures from volume drying (Yatsu, 1988). The observation that rock surfaces on
changes induced by hydration. Furthermore, three variables shore platforms undergo swelling when measured with a
control the effectiveness of salt weathering: (1) the nature of the microerosion meter has been linked with wetting and drying
salts and their solutions, (2) the properties of the affected (Kirk, 1977; Stephenson and Kirk, 2001; Stephenson et al.,
materials, and (3) the nature of the environment in which salts 2004; Kanyaya and Trenhaile, 2005; Trenaile, 2006; Porter and
may cause the materials to disintegrate (Cooke, 1979). Cooke Trenhaile, 2007; Porter et al., 2010a, b, c) and is consequently
and Smalley (1968) noted that the degree of saturation of the thought to contribute to rock stress, weakening, and eventual
solution is important, as is the duration of exposure of the bed erosion.
materials to supersaturated conditions. The capacity to absorb Wetting and drying is well recognized on shore platforms
water, porosity, microporosity, the rate at which solutions where flat surfaces allow water to pool. In this particular situ-
penetrate rocks and tensile strength, are all properties of rock ation the process has been termed water layer weathering. It
that control the effectiveness of salt weathering (Cooke, 1979). can operate wherever sea water can accumulate and evaporate.
Using laboratory experiments, Goudie et al. (1970) found that Water layer weathering was first recognized by Bartrum and
igneous rocks were little affected by salts, whereas chalk, lime- Turner (1928). The physical process of water layer weathering
stone, and sandstones broke down rapidly after repeated im- remains to be fully explained but most researchers consider it
mersion in saline solutions. Cooke and Smalley (1968) important in addition to wetting and drying, salt weathering,
proposed that the growth of crystals from solution is important chemical weathering, and the movement of solutions through
in humid coastal deserts, where dissolved salts are abundant. rock capillaries (Trenhaile, 1987). Kanyaya and Trenhaile
Salt weathering operating on rock coasts in warmer temperate (2005) and Trenhaile et al. (2006) used tidal simulators to
environments is probably a significant weathering agent, but replicate real time wetting and drying conditions on more than
little work has been conducted to establish the relative contri- 1000 cores comprising different rock types in a series of la-
bution of individual weathering processes relative to others on boratory experiments. Their results showed that wetting and
coastal rock coasts. A number of laboratory or simulation drying worked more effectively on coarse grained rock (sand-
studies have examined the relative roles of salts and other stone) when salt water was used compared to wetting and
weathering processes. Jerwood et al. (1990a, b) in a series of drying with fresh water. However the opposite result occurred
experiments using groups of chalk cubes immersed in water in finer grained rocks (basalt and argillites), with more erosion
and different salt solutions showed how different salts impacted occurring during wetting and drying using fresh water than salt
differently on freeze thaw weathering, for example. Sodium water. The processes of water layer weathering produces a dis-
chloride enhanced freeze thaw while magnesium sulfate in- tinct morphology where shallow pools are surrounded by
hibited freezing and subsequent rock fatigue. How the results of raised rims associated with joints (Figure 3). The raised rims
laboratory experiments apply to field settings remains unclear around the pool occur along joints and retain water in the joint.
(Tingstad, 2008). The significance of salt weathering is difficult The rims undergo fewer wetting and drying cycles than the
to determine, particularly when wetting and drying is also oc- pools which consequently weather and lower more quickly.
curring simultaneously.
10.11.2.5 Biology
10.11.2.4.2 Wetting and drying
The intertidal location and proximity to sea spray of many rock A common definition of bioerosion is the removal of lithic
coast landforms facilitates repeated wetting and drying. substrate by the direct action of organisms (Neumann, 1966).
294 Rock Coasts

A bewildering array of organisms are capable of eroding the Seychelles in the Indian Ocean and found that rates varied
bedrock on rock coasts and these operate from above the from 0.9 cm yr1 to 4.5 cm yr1. Trenhaile (1987) presented a
spring high tide line that is regularly splashed, but not sub- summary of 42 investigations of bioerosion rates which range
merged by ocean water (the supratidal zone), to depths that from 30 to 50 mm in 3 to 4 weeks by algae on carbonates, to
are permanently submerged (the subtidal zone). On rock 1 m yr1 on limestone by boring sea urchin (Paracentrotus
shores microboring occurs as a result of cyanobacteria and lividus). Naylor et al. (2012) has been critical of geomorph-
chlorophytes, and other boring activity is carried out by in- ologists for not paying greater attention to the role of
vertebrates including sponges and mollusca such as chinton, bioerosion and when bioerosion has been considered it has
gastropods, and bivalves. been in an oversimplified manner. They argue that erosion
Rates of bioerosion vary enormously from microns to rates reported as millimeters per year of downwearing fail to
meters per year, depending on the type and population disclose the complexity of bioerosion and advocate erosion
densities of the organisms responsible. Erosion is achieved rates instead be reported as volume loss. Naylor et al. (2012)
through grazing, chemical reaction to secretions, and boring. also attempt to show how bioerosion at the scale of milli-
Grazing gastropods feed by scraping the rock surface with their meters can influence erosion rates at larger scale and con-
radula to ingest the rock and digest the endolithic algae con- cluded that bioerosion is cross-scalar, influencing erosion at
tained in the rock (Figure 4). Grazing on chalk platforms larger scales.
along the English Channel by limpets (Patella modelli) con- In contrast to agents causing erosion, biota on rock coasts
tributed between 12 and 35% of the annual lowering rate of can also provide protection from other erosive processes
2.3 mm yr1 (Andrews and Williams, 2000). Trudgill (1976) (Naylor and Viles, 2002; Naylor et al., 2002). Hills (1949)
measured rates of boring organisms on Aldabra Atoll, part of modeling this protective role on platforms in Victoria, south-
eastern Australia, where the extensive covering of Hormosira
banksii (Neptunes necklace, sea grapes, or bubbleweed) pre-
vents erosion by mechanical or subaerial processes. Stephenson
and Kirk (1998) measured erosion rates on shore platforms and
found that winter rates were lower compared with summer. The
main reason for this was that cooler winter temperatures
allowed extensive covering of algae over the platform which
prevented drying, trapped sediment, and provided protection
from wave erosion. Bioprotection has not been widely studied
and the relative importance of it compared to bioerosion and
other erosion processes is unknown. Macroalgae (e.g., Durvil-
laea modelling, a common bull kelp) can dissipate wave energy
through mechanical movement and protect rock surfaces
against wave impact forces. The removal of macroalgae by
strong waves, however, often removes the rock to which they
were attached, resulting in block erosion (Figure 5).
Even less well understood than bioprotection is biocon-
struction on rock shores (coral reefs excluded). Encrustation of
Figure 4 Grazing track of a limpet. The limpet has traversed from rock surfaces by a variety of organisms such as barnacle and
left to right of the images scrapping the rock surface to remove and vermetid worms provides additional protection to rock sur-
ingest the rock. Endolithic algae are digested and the rock excreted. faces by the construction of sessile (fixed in one place) shells.
Naylor and Viles (2000) documented meter-scale reef building
on the seaward edge of shore platforms in southern Wales
(UK) by honeycomb worms (Sabellaria alveolata) that build
tubes by cementing sand grains together. These reefs may play
a role in dissipating wave energy by reducing the water depth
and forcing wave breaking.

10.11.2.6 Mass Movement


In addition to marine and weathering processes, sea cliffs are
subject to mass movement and other processes that drive
slope evolution. The type of mass movement occurring on
rock coasts (e.g., topple, fall, slide, avalanche, and flow) de-
pends on the climate, groundwater hydrology, and geological
and lithological framework of the coast (Trenhaile, 1987;
Sunamura, 1992; Hampton and Griggs, 2004). Understanding
Figure 5 Durvillaea antarctica washed up on a beach with bedrock the relative contributions of the subaerial, marine, and mass
attached. The block detached from a rock mass is 50 cm in the movement processes that control sea cliff erosion is complex
long axis. as these processes vary regionally and commonly work
Rock Coasts 295

Active
M >> SA

Boulders

M > SA

Inactive

Talus
M = SA

Figure 6 Cliff failure with associated rock fall in Blue Lias


limestone, Glamorgan, South Wales coast (UK).
Former
M < SA
together. Cliff erosion is commonly viewed as a result of Fan
marine processes where waves undercut cliffs and reduce cliff
support until collapse occurs, and the fallen debris is sub- (a) (b)
sequently removed by waves to expose the cliff thereby re-
Figure 7 Idealized cliff profiles cut in homogenous rocks, showing
peating the cycle (Figure 6). However, sea cliffs can continue
(a) geological history, and (b) relative efficacy of marine (M) and
to erode even when abandoned by the sea or in the absence of
subaerial (SA) processes. Adapted from Emery, K.O., Kuhn, G.G.,
undercutting because of mass wasting processes (Nott, 1990; 1982. Sea cliffs: their processes, profiles, and classification.
Lee and Clark, 2002). Consequently, coastal cliff profiles de- Geological Society of America Bulletin 93, 644654, with permission
cline in gradient. In many areas, cliff erosion is also a function from Geological Society of America.
of the amount, mobility, and morphology of the sediment or
mass movement debris at its base, which determine the ac- but in some locations profiles have been partially rejuvenated
cessibility of the cliff face to wave attack. In addition, surface by undercutting. In some cases there is little evidence of a
and ground water also play parts in producing conditions for formerly degraded profile.
cliff failure. Surface water running down cliff faces washes Shore platforms generally occur where cliff erosion has
material down slope. Groundwater causes loading and lubri- been active, although they may be partially or completely
cation in discontinuities, thus reducing the strength of rock covered by beaches. In contrast, where steep cliffs plunge into
materials. Seepage of ground water at the cliff face can erode deep water, wave energy is reflected from the cliff-face
material and weaken discontinuities. Water can also cause and marine erosion processes are ineffectual. These different
weathering and dissolution, thereby further reducing rock evolutionary pathways are heavily dependent on the local
strength. sea-level history, sediment supply, lithology, antecedent
topography, tectonics, and the local process regime.
Stacks and arches (Figure 8) are smaller scale features of
10.11.3 Rocky Coast Landforms many rock coasts and are commonly valued components of
coastal landscapes, important to tourism and sightseeing. The
Worldwide, the morphology of sea cliffs is highly varied, re- geomorphic significance of such features is seldom investi-
sulting from the complex interplay between different erosion gated. An exception is the work of Trenhaile et al. (1998) who
processes and the bedrock in which the cliffs have formed. assessed the role of rock hardness and joints in the develop-
Moreover, many rocky coasts have evolved slowly and may ment of stacks at Cape Hopewell Rocks, New Brunswick,
have partly inherited their morphology from processes oper- Canada. Stacks in rock with the same strength as the sur-
ating under different climatic conditions. Figure 7 provides a rounding shore platforms resulted from the patterns, spacing,
simple conceptual framework that considers the varied de- and location of jointing and faults. In other cases, stacks occur
velopment of sea-cliff morphology. During the Last Glacial, because they are formed from more resistant rock and become
global eustatic sea-level was as much as 120 m below present isolated as the surrounding shore line retreats. Earlier work by
(Chappell and Shackleton, 1986). Hence, for many thousands Shepard and Kuhn (1983) also noted the importance of
of years, many contemporary sea cliffs were subject only preferential erosion along joints and channels cut by surges in
to subaerial processes that degraded and rounded cliff the development of sea arches and caves. Johnson (1925)
profiles. Subsequent climatic warming was accompanied by a documented the development of a sea arch from a cave which
marine transgression that drove sediments landward, flooded broke through the other side of a promontory. In both ex-
embayments, and reexposed cliffed coasts to wave attack. At amples researchers utilized historical photographs to docu-
present, some degraded (former) profiles remain at the coast, ment the evolutionary sequence of arch development.
296 Rock Coasts

slope over wall, hogs back, and beveled cliffs (see Trenhaile,
1987). There has been much discussion of the origin of
composite profiles. Fisher (1866) originally suggested that
alternating periods of marine and subaerial erosion might be
responsible, whereas Challinor (1949) thought that the upper,
gently sloping surface and the lower, steep cliff face could form
contemporaneously. Improved recognition of the nature of
repeated Quaternary sea-level fluctuations led later researchers
toward the view that the degraded, rounded profile is likely
formed by subaerial processes during periods of lower sea
level, whereas with warmer climates, rising sea levels renew
wave attack and steepen the base of the cliff (e.g., Cotton,
1951; Guilcher, 1958). This view has been supported by
Kirkbys (1984) modeling that supported Savigears (1952)
view of the very gradual nature of degraded slope develop-
ment. The simplest development of composite cliffs involves
one glacialinterglacial cycle in which all previous evidence of
development have been destroyed. Flemming (1965) and
Cotton (1967) discussed such a possibility in the context of
cliffed profiles around the Auckland Islands in the Southern
Ocean. However, in other instances, cliff profiles may have
been produced by several phases of erosion across multiple
glacial/interglacial cycles (Wood, 1959; Griggs and Trenhaile,
1994).
The form of composite cliffs results from the interplay
between several factors. Arber (1949) discussed the interplay
between geology and the process environment while con-
Figure 8 Stack on Flower Pot Island, Bruce Peninsula, between sidering the development of cliff profiles in Devon and
Georgian Bay and Lake Huron. Cornwall along the southeastern coast of England (UK). She
noted that landward dipping strata generally give rise to hogs
back-type composite cliffs, but in cliffs without structural
10.11.3.1 Relict, Composite, and Plunging Sea-Cliffs
planes the resulting profile tends to be very steep with a flat
Former sea-cliffs (Figure 7) are cliffs that were exposed top, provided that marine processes are dominant. If there is a
to marine undercutting in the distant past, perhaps during balance between marine and subaerial processes, then the
the Last Interglacial, high sea level (124119 years BP), or upper portion of the profile may be rounded whereas the
even prior to that. Cliff profiles were subsequently degraded lower portion may be steep because of wave undercutting (as
during the Last Glacial, and for a number of local reasons depicted in Figure 7). In highly resistant rocks, vertical slopes
the cliffs were not reexcavated during the Holocene. can persist even in the absence of basal undercutting by wave
Good examples of such profiles occur on coasts exposed to action. Hence, where precipitous hard-rock cliffs occur in
tectonic uplift, such as the central Californian (USA) coastline, coastal environments, the profile that develops during glacial
where a suite of marine terraces exhibit paleocliff profiles that periods is not a degraded, rounded hillslope, but is instead a
become more rounded with age (Rosenbloom and Anderson, near-vertical cliff face mantled at its toe by talus. If the cliff is
1994). On tectonically stable coasts, most rocky shorelines high and the lithology resistant, there will be insufficient time
have been reactivated to some extent by Holocene raised for the talus slope to accumulate and build to the cliff top. At
sea level. Lord Howe Island (located 600 km east of Australia in the
The importance of different local controls was well Tasman Sea), for instance, very large talus slopes occur at the
demonstrated in Savigears (1952) classic study of slope- base of near-vertical basalt cliffs 800 m high (Dickson, 2004;
development in southern Wales (UK), where he noted that a Dickson and Woodroffe, 2005). Holocene sea-level rise has
laterally extending beach system in front of a cliffed hinterland been unable to remove these very large talus deposits, and the
had resulted in the cliffs progressively losing contact with the cliffs are relict in the sense that marine erosion processes are
sea, resulting in a gradual conversion of the cliffs to debris- currently ineffectual. Where lower cliffs occur around Lord
mantled slopes. The sequence of cliff profiles was immediately Howe Island, smaller talus slopes accumulated during the Last
identified as showing ergodic evidence of spacetime substi- Glacial, and wave action has since been sufficient to remove
tution (Burt, 2003). the talus. In some locations, the rocks are sufficiently hard that
These examples are typical of many observations that cliff upon removing talus, the cliff face was drowned, rather than
profiles are generally steepened under wave attack and de- eroded by sea-level rise. These are referred to as plunging cliffs.
graded under subaerial erosion. Some cliffs, known as com- An important feature of plunging cliffs is that deep water oc-
posite cliffs, consist of a number of distinct slope elements. A curs at the cliff toe, preventing waves breaking against the cliff
variety of terms have been used to describe composite cliffs face, and preventing an accumulation of debris within the
that have fairly gentle slope above a steep cliff face, including water column that might be used in abrasion (Cotton, 1949).
Rock Coasts 297

As a result, plunging cliffs are generally considered to be re- twentieth century made it clear that repeated sea-level move-
sistant to erosion (Figure 9). ments provided the mechanism necessary for continued cliff-
Davis provided one of the first specific descriptions of ing, rather than gradual subsidence.
plunging cliffs in basaltic rocks around the Pacific Islands and
also on Banks Peninsula, New Zealand (Davis, 1928). At a
10.11.3.2 Shore Platforms
similar time, Daly (1927) described plunging cliffs around the
island of Saint Helena in the South Atlantic Ocean. Darwin Plunging cliffs are resistant to marine erosion, but where cliffs
(1844, 1846) also considered the development of the cliffed do yield to erosive forces a range of morphologies can be
coast of Saint Helena, and pondered how marine processes produced. Shore platforms are one of the most distinctive fea-
could have continuously cut the cliffs around the island under tures. The critical conditions necessary for the erosion of rock
stable levels of land and sea. Darwin (1846, p. 26) concluded coasts were considered in detail by Tsujimoto (1987). He se-
that the island must have slowly subsided in order that waves lected 25 sites around the Japanese coast and quantified rock
could continue their attack with fresh and unimpaired vigor. resistance using measurements of compressive strength as well
Darwins theory was overlooked for many decades as re- as a discontinuity index based on longitudinal, wave velocity
searchers believed that waves could cut rock to a great depth measurements. Wave force at each site was calculated on the
below sea level, thereby enabling the continued delivery basis of nearshore bathymetry and incident wave conditions. As
of high energy waves to the cliff toe. Dietz and Menard Figure 10 illustrates, this work showed that when a threshold is
(1951) and Dietz (1963) eventually dispelled this notion, exceeded between rock resistance and wave force, rocky shore
and a greater understanding of Quaternary glaciations in the morphology changes from plunging cliffs to shore platforms.
A large proportion of the rocky coast literature is focused
on the horizontal or gently sloping surfaces that develop at the
base of cliffs as they retreat. Various terms have been used to
describe these features, but the most appropriate is shore
platform, which usefully has no genetic connotation, for there
remains some debate in respect to the relative importance of
different controls on shore platform evolution. At a broad
scale, a process-based distinction between sloping and near-
horizontal platforms is possible on the basis of tidal range. In
most areas where the tidal range is from the high end of the
mesotidal to the macrotidal, shore platforms are sloping
(typically 1.51 to more than 41), and they extend from the
cliff-platform junction to below the low tide level without a
clear break in slope. Platform gradient increases as tidal range
increases, implying that erosion at different tidal stages exerts
an important influence on platform development (e.g., see
Trenhaile, 1974, 1987, 1999; Trenhaile and Layzell, 1981). In
microtidal areas, shore platforms are usually near-horizontal,
Figure 9 Plunging cliff at the entrance to Sydney Harbor, Australia. with a gradient of about 11 or less, and often but not always

Shore platform (type A and B)


No platform (plunging cliff)

102
p (tonne m2)

Shore platform formation


No platform formation

101

101 102 103 104


S (tonne m2)
c

Figure 10 Demarcation between shore platform and plunging cliff. Reproduced from Tsujimoto, H., 1987. Dynamic conditions for shore
platform initiation. Science Report of the Institute of Geoscience, University of Tsukuba A8, 4593, with permission from Dspace.
298 Rock Coasts

they have a distinctive scarp at the seaward edge. Sunamura


(1992) referred to these morphologies as Type-A (sloping)
and Type-B (near-horizontal) and viewed the presence of a
seaward scarp as an essential distinction. One difficulty arises
in classification because some microtidal locations have plat-
forms with very low gradients and no clear seaward scarp (e.g.,
Edwards, 1951; Hills, 1972; Kennedy and Dickson, 2006). The
gradient of these platforms may be minimal, commonly 11 or
less, but the presence or absence of the scarp is generally used
in classifying the morphology as Type-A or Type-B. Potential
confusion exists, therefore, because microtidal platforms of
very low gradient without a clear seaward edge might be
considered Type-A alongside platforms that have developed in
quite different conditions in macrotidal environments. For the
purposes of this section we discuss near-horizontal platforms
(gradients of less than about 11), which generally occur in
microtidal areas, and sloping platforms which have gradients
of greater than about 11 and generally occur in high mesotidal
and macrotidal areas, separately. In the case of near-horizontal
platforms we specifically discuss the presence or absence of a
clearly marked seaward scarp.

10.11.3.2.1 Near-horizontal platforms


Contemporary shore platforms, as opposed to those inherited
from previous high stands of sea level, were probably initiated at
the beginning of the Holocene still-stand (i.e., 60007000 years
BP). Sunamura (1975, 1976, 1991) conducted laboratory ex-
periments on model cliffs composed of Portland cement, quartz
sand, and water. When subjected to wave attack, notches de- Figure 11 Seaward edge of a horizontal platform on Kaikoura
veloped in the cliff face and deepened over the experiments, Peninsula, South Island, New Zealand, showing abrupt seaward edge
which generally ran for around 100 h (Sunamura, 1992). In the sometimes referred to as a low tide cliff.
field a continuation of this process would lead to a stress state in
the cliff face that would result in cliff collapse. Subsequent debris
removal would expose a cliff and near-horizontal platform with
an abrupt seaward scarp (Figure 11). Basal erosion of cliffs is an
important process, and the simple cantilever-collapse mech-
anism plays an important role in the gradual evolution of many
rocky coasts. However, it is interesting to note that recent
measurements on cliffs in North Yorkshire (UK) show that
complex rock-fall patterns occur on the faces of sea-cliffs without
a high degree of connectivity to basal erosion processes (Lim
et al., 2005, 2010a, b).
Near-horizontal platforms with a characteristic seaward
scarp have been described in many areas, particularly around
Australasia (e.g., Bartrum, 1924, 1926; Jutson, 1939; Edwards,
1941; Hills, 1949; Gill, 1967; Cotton, 1963; Bird and Dent,
1966; Kirk, 1977; Stephenson and Kirk, 2000a, b; Dickson,
2006; Kennedy and Dickson, 2006; Kennedy, 2010) and Japan
(e.g., Takahashi, 1977; Tsujimoto, 1987; Sunamura, 1992).
The width of these platforms is measured from the cliff-toe to Figure 12 A near-horizontal platform, 100130 m wide, cut in
the seaward edge, and varies between narrow platforms only a sandstone at Pakiri, 65 km northeast of Auckland, North Island, New
few meters wide to platforms hundreds of meters in width Zealand. The platform is elevated just above high tide level and
(Figure 12). Platform elevation also varies considerably, with plunges into 1013 m water depth.
some platforms formed around low-tide level, others at a
range of inter-tidal elevations, and yet others at a range of when salt weathering processes operate on structural weak-
supratidal elevations up to several meters above high tide. nesses in the rock mass (e.g., Ongley, 1940). These structural
Multiple Holocene shore platform elevations are preserved on benches differ from shore platforms, which develop at an
some coasts due to tectonic uplift (e.g., Kennedy and Beban, elevation that is not predetermined by rock structure.
2005). On stable coasts it is also possible to form benches in The laboratory results of Sunamura (1991) imply that
cliff faces at multiple elevations tens of meters above sea level shore platform elevation increases as rock resistance increases.
Rock Coasts 299

This observation is consistent with many field observations platforms. He described the results of laboratory experiments
(e.g., Gill, 1967; Trenhaile, 1969, 1987; Hills, 1971; Kirk, (Sunamura, 1991) that show that the submerged part of a
1977; Reffell, 1978; Gill and Lang, 1983; Thornton and Ste- homogenous cliff is free from erosion, because the assailing
phenson, 2006). Sunamura (1991) also deduced that platform force of breaking waves decreases exponentially with water
elevation should be higher where nearshore water depth is depth. Sunamura (1992) further pointed out that following
greater. At Lord Howe Island (Tasman Sea) the elevation of storms, field observations show that seaward scarps com-
basalt shore platforms positively correlated both with rock monly remain covered in marine flora/fauna. He concluded
resistance and nearshore water depth (Dickson, 2006). Some that the seaward scarp does not recede under present sea level
researchers have also noted that higher platforms tend to form provided that platforms are cut in insoluble rocks without
where wave exposure is greater (e.g., Kirk, 1977; Reffell, 1978). significant lithological structures. This latter proviso is par-
Multiple controls on platform elevation mean that their ticularly important, for in many situations shore platform
potential use as paleosea-level markers is fraught with dif- development occurs in areas where there are significant
ficulty. One of the earliest descriptions of near-horizontal lithological structures. Despite this, most recent studies of
platforms was made by Dana (1849), who described surfaces shore platforms in microtidal settings assume that the seaward
with a clear seaward scarp around the Bay of Islands, New edge has remained stationary over the course of platform de-
Zealand. Bartrum (1916, 1926) noted that the elevation of velopment. This is clearly not the case in the Bay of Islands
these platforms is a little below HWL, and subsequently de- (North Island, New Zealand), where joints generally run
veloped a theory that rock saturation level might control parallel to the cliff face and have been heavily eroded, in some
platform elevation. This formed an early focus for debate instances causing a platform to be completely dissected,
around the relative roles of weathering and wave erosion causing a decrease in platform width (Kennedy et al., 2010).
processes. Trenhaile and Mercan (1984) subsequently showed Significant erosion of the seaward edge of platforms affords
that rock saturation does not control elevation, and Kennedy one possible explanation for anomalous relationships be-
et al. (2010) recently surveyed the platforms in the Bay of tween platform width and exposure to wave energy, although
Islands, showing that they occur at a range of elevations be- many other factors could be important from site to site.
tween a little above mean high spring tide and closer to mean A second important matter in respect to the seaward scarp
sea level. They noted that platform elevation tends to increase concerns the issue of preservation. It is possible along short
both with exposure to wave action and with the depth of water sections of coast in which the tidal range is micro or low
immediately in front of the platform. mesotidal, to observe platforms with very gentle slopes alter-
A level of agreement has emerged with respect to controls nating between areas that have a clearly marked seaward scarp
on the elevation of near-horizontal platforms, but no such and those that descend beneath sea level without a clear edge
consistency is apparent with respect to platform width. Several (e.g., Healy, 1968; Stephenson and Kirk, 2001). Gill (1972)
studies have noted that wider platforms occur where rock re- noted that the seaward edge is generally preserved where rock
sistance is lowest (e.g., Agar, 1960; Everard et al., 1964; resistance is higher, and Bird and Dent (1966) commented
Takahashi, 1977). Similarly, the narrowest platforms commonly that local availability of abrasive material seems to be im-
occur where more resistant rocks limit the extent of platform portant. In addition to distinguishing a critical condition be-
widening. However, in some areas resistant rocks are associ- tween plunging cliffs and shore platforms, the data collected
ated with wider platforms, whereas less resistant rocks have by Tsujimoto (1987) also provides for discrimination between
narrower platforms (e.g., Edwards, 1941; So, 1965). Similar sloping (Type-A) and near-horizontal (Type-B) platforms. In
discordance exists with respect to exposure to wave energy, rocks of lesser resistance the seaward edge is destroyed by wave
because width is commonly greater where wave exposure is attack leading to the sloping form.
greater (e.g., Swan, 1971; Trenhaile, 1972), but in some in-
stances wider platforms actually occur in sheltered embay- 10.11.3.2.2 Sloping platforms
ments (e.g., Edwards, 1941; Hills, 1949; Bird and Dent, 1966). A strong positive correlation occurs between tidal range and
A number of factors obfuscate platformwidth relationships. the gradient of sloping platforms (Trenhaile, 1974, 1987).
For instance, some shore platforms may have been formed Seaward scarps generally do not form in high mesotidal set-
over more than one sea-level cycle (e.g., Brooke et al., 1994). tings (tidal range 3 to 4 m) and are absent from macrotidal
Another issue concerns the assumption that the seaward edge environments regardless of rock type. Sloping shore platforms
of shore platforms marks the origin of platform cutting. Many have been the subject of many conceptual models. A widely
early researchers considered the possibility that the seaward held view of their development follows the theoretical work of
scarp of near-horizontal platforms might retreat (e.g., Bar- Challinor (1949) who proposed that such platforms might
trum, 1926; Jutson, 1939; Edwards, 1951; Hills, 1949). On retreat self-parallel, with constant geometry, provided that
aeolianite coasts the edge of shore platforms can be seen to be rates of downcutting are sufficiently rapid to allow continued
undercut with many dislodged blocks in front of the platform wave attack at the cliff toe. Several mathematical models
(e.g., Gill, 1972; Dickson and Woodroffe, 2005). Blocks in subsequently expanded on this work. Flemming (1965) sim-
front of the seaward scarp can be seen in other lithologies also, plified the important processes involved in self-parallel retreat
but as Stephenson (2000) pointed out, because of the slow by neglecting the possible effects of rock weathering, tidal
rate of change, it is not always clear whether these blocks have movement, and beach development. The model implied that
been sourced from the seaward scarp or the landward cliff. cliff recession decreases with time but never ceases. By con-
Sunamura (1992) noted that there has been little con- trast, the equations developed by Sunamura (1978a, b, 1992)
sideration of wave dynamics on the seaward edge of shore imply that erosion rates decline over time and eventually
300 Rock Coasts

become zero when the wave assailing force is below the re- (Woodroffe et al., 1995; Dickson and Woodroffe, 2005), and
sistance of the cliff rock. This suggests that platform geometry the possible effects of Holocene higher sea level have been
becomes wider and flatter as rock resistance decreases, al- noted in Sydney, Australia (Kennedy, 2010). An analysis of
though the model does not account for weathering, or the high-resolution, laser scan data from Shag Point, South Island,
accumulation of beach materials, or the effect of tide range. In New Zealand, which revealed the extent of two distinct sur-
contrast Trenhaile (1983) specifically modeled the effects of faces, raises the possibility that Holocene sea-level movements
variable rates of erosion at different tidal levels. Again the initiated platforms at more than one elevation (Palamara
model showed that initial rates of platform development are et al., 2007).
rapid, but eventually that rates of erosion at high and low tidal Trenhaile (2010a) considered the issue of a higher Holo-
levels become similar, resulting in an equilibrium morph- cene sea level using a numerical model. Interestingly, although
ology. Walkden and Hall (2005) and Walkden and Dickson the model results supported previous work that showed a link
(2008) recently explored the development of sloping plat- between tidal range and platform slope, and between rock
forms in soft rocks using a numerical model, and reached a resistance and platform type, the model runs imply that
conclusion similar to that of early conceptual models: erosion Holocene sea-level history provides a further reason for the
rates decrease asymptotically from some initial condition (e.g., difference between sloping platforms and subhorizontal
a vertical cliff), and eventually the sloping rock shore profile platforms. This example highlights the importance of model-
dynamically adjusts to wave forcing but over time maintains a ing as a valuable method for understanding longer term rock
similar profile, provided that boundary conditions are con- coast development.
stant. Similar equilibrium profiles develop with stable sea level
in Trenhailes (2010b) cohesive-clay coast model.
Field studies have shown that the size and amount of debris
on sloping platforms leads to changes in the platform profile 10.11.4 Rock Coast Modeling
(Robinson, 1977a, b, c). On chalk cliffs, more elevated plat-
forms were noted in embayments near Kent, England (UK) Rock coasts develop very slowly and attempts to determine
because sand in embayments resulted in wave runup accessing modes and rates of long-term development are constrained by
higher parts of the cliff than on headlands (Wood, 1968). the general absence of datable deposits and suitable dating
However, the reverse situation has also been noted on the Isle techniques. Many attempts have been made to model the
of Thanet in Kent (So, 1965), where lower platform surfaces evolution of rock coasts (e.g., Trenhaile, 2003), although their
have been attributed to the abrasive role of sediment in low- inherent complexity the product of myriad geological and
ering platforms. Wright (1970) noted that alongshore variation morphogenic factors makes it difficult to represent all the
in the cliffplatform junction on the southern coast of England variables that play a role in their development. Although
generally correlates with changes in tidal range, but the lo- Sanders (1968) and Sunamura (1973, 1975) studied the
cations of more resistant strata are also important (Wright, erosion of plaster and cement blocks in wave tanks and
1967). Similarly, Trenhaile (1972) showed that in the Vale of flumes, most rock coast models have been mathematical and
Glamorgan, Wales, the elevation of the cliffplatform junction focused on the two-dimensional development of coastal pro-
increases as the magnitude of tidal range increases, but con- files, and especially with intertidal shore platforms and sub-
siderable variability is owing to differences in local lithology. marine continental shelves.

10.11.3.3 Inheritance 10.11.4.1 Coastal Profiles


Understanding the role of inheritance in rock coast evolution is The first models of coastal profiles were qualitative, structured
one of the most significant problems in rock coast research within a cycle of erosion and based on the premise, which was
(Trenhaile, 2002). Several studies have shown that because of prevalent in the eighteenth and early part of the nineteenth
the slow rate of change of rock coasts, shore platform morph- century, that submarine erosion is effective down to depths of
ology observed today may have been partly formed during the about 100 m or more, and thus able to produce extensive
Last Interglacial or before (Bird and Dent, 1966; Woodroffe marine surfaces while sea level is stationary (Ramsey, 1846;
et al., 1992; Young and Bryant, 1993; Brooke et al., 1994; Stone Johnson, 1919). Davis (1896) inferred, in diagrammatic form,
et al., 1996; Trenhaile et al., 1999; Blanco Chao et al., 2003; that wave-eroded submarine surfaces become wider through
Gomez-Pujol et al., 2006). Although based on measurements of time while maintaining a constant gradient, whereas Johnson
relatively rapid erosion rates, Kirk (1977), and later Stephenson (1919) proposed that surface gradients decrease, albeit at a
and Kirk (1998) argued that the shore platforms on Kaikoura slow rate in the early stages of development, as surface width
Peninsula, South Island, New Zealand are contemporary fea- increases. Conversely, Challinor (1949) thought that sub-
tures developed in the last 6000 to 7000 years. marine surfaces undergo parallel slope retreat, without any
It is important to consider if Holocene sea-level fluctu- change in width or gradient, as they migrate landward. Several
ations (where they occurred) influenced platform develop- workers also suggested that shore platforms experience paral-
ment to any great extent. There have been few detailed studies lel slope retreat (Edwards, 1941; Bird, 1968; Trenhaile, 1972).
of the effect of Holocene sea-level change on rock coasts, al- These conceptual models were based on the assumption that
though the effect of a Holocene highstand on calcarenite the negative feedback relationship between bottom gradient,
shore platform morphology is clear at Lord Howe Island rates of wave attenuation, and the energy reaching the cliff
Rock Coasts 301

foot produces a constant rate of erosion at different elevations The present morphology of slowly changing rock coasts is
within the intertidal zone. also the result of Quaternary changes in sea level, and in many
Most of the early mathematical models were concerned areas changes in the elevation of the land. Scheidegger (1962,
with erosion with constant sea level in tideless seas (Flem- 1970) and King (1963) modeling the effect of steadily rising
ming, 1965; Scheidegger, 1970; Sunamura, 1976, 1978a, b). and falling sea level on rock coasts, and Sunamura (1978b)
Trenhailes (1983) model calculated erosion rates at the high modeling the development of erosional continental shelves,
and low tidal levels but did not consider rates at intermediate albeit in a tideless sea, during the Holocene transgression. The
elevations. Trenhaile and Layzells (1981) model was the first effect of different Holocene relative sea-level curves, which are
to consider erosion within the intertidal zone, based on the characteristic of different geographic locations, was examined
percentage frequency or total time each year that the water by Trenhaile and Byrne (1986), based on Trenhailes (1983)
surface, and consequently wave action, is at each intertidal wave erosional model, and the same model was used to in-
elevation (the tidal duration value). vestigate the formation of continental shelves and wide coastal
Mathematical models developed in Japan related the rate of terraces with changing sea level over five glacialinterglacial
cliff erosion (dx/dt) to the height of the waves at the cliff foot cycles (Trenhaile, 1989). Cinque et al. (1995) and Anderson
and the compressive and impact strength of the rocks (Hor- et al. (1999) included, in addition to Quaternary changes in
ikawa and Sunamura, 1967). Sunamura (1977) proposed that: sea level, the effect of tectonic uplift on marine terrace for-
mation. Trenhailes (2000) model, which considered the way
dx=dt p 1nfw =fr 1 that wave energy is expended within the intertidal zone, was
used to study: (1) the effect of high sea level during the last
where dx/dt is the mean cliff erosion rate, fw is the assailing two interglacial sea stages on contemporary intertidal shore
force of the waves, and fr is the resisting force of the rocks. The platforms (Trenhaile, 2001a); (2) the evolution of shore
term fr can be approximated by (rgH/Sc) C, where r is the platforms and erosional continental and volcanic island
density of sea water, g is the acceleration due to gravity, H is shelves during the Quaternary (Trenhaile, 2001b; Quartau
the wave height at the cliff base, Sc is the compressive strength et al., 2010); (3) the formation of subaerial and submarine
of the rock, and C is a dimensionless constant. Tsujimoto terraces during the Quaternary on tectonically mobile coasts
(1987) provided expressions for fw and fr based on wave (Trenhaile, 2002); and (4) the effect of Holocene changes in
pressure and the longitudinal sound wave velocity in the rock relative sea level on subaerial to submarine profiles (Trenhaile,
body, as well as nondimensional constants representing 2010a). The main conclusions of these investigations were
abrasion and the effect of weathering. Although these in- that:
vestigations attempted to identify and quantify factors that
1. Shore platforms are over-steepened and truncated by fall-
determine rates of cliff erosion, it is questionable whether
ing and rising sea level during the onset and end of glacial
these simple variables can adequately represent the driving
stages, respectively, and then modified and restored to a
and resisting forces. There are also many other variables to
state of quasiequilibrium during the ensuing interglacials,
consider, including the effect of tidal variations, which deter-
largely by erosion at the high tidal level (Trenhaile, 2001b).
mine the frequency of wave attack at the cliff foot, the dip and
2. Marine terraces are formed during interglacial stages on
strike of the rock, and the presence, mobility and quantity of
rising landmasses and during glacial stages on subsiding
cliff-foot deposits, and whether they function as protective or
landmasses. The number of terraces increases with the rate
erosive agents (Trenhaile, 1999). It is unlikely that math-
of uplift and subsidence and with the slope of the land
ematical equations can adequately represent the effect of
(Trenhaile, 2002).
many of these factors, and rock coast models may therefore
3. Although tidal range is also important, falling relative
have to continue to rely, at least in part, upon empirical data.
sea level from its mid-Holocene maximum, 1 to 2 m above
Field evidence indicates that most mechanical wave ero-
its present level, promoted subhorizontal platform devel-
sion occurs through processes that are closely associated with
opment in Australasia and over much of the Southern
the water surface, including water hammer and air com-
Hemisphere, whereas the asymptotic rise in sea level over
pression in joints, bedding planes, and other rock crevices
much of the Northern Hemisphere promoted the devel-
(Sanders, 1968; Robinson, 1977a; Trenhaile, 1987). Trenhaile
opment of sloping platforms (Trenhaile, 2010a).
(2000) developed a wave-erosion model that used basic wave
equations to explore the interaction between wave dynamics, Most rock coast modeling has been concerned with
tidal regimes (tidal range, tidal duration distribution), rock mechanical wave erosion on sloping platforms in mega to
resistance, coastal morphology, and erosion at the shoreline. mesotidal environments, and the additional effect of wea-
The model predicted that, under stable sea-level conditions, thering has only been considered recently. Downwearing data
shore platform width increases and gradient decreases until a from transverse arrays of microerosion meters was incorpor-
state of static equilibrium is reached, when the attenuated ated into a model that considered the role of weathering in
waves at each tidal elevation are unable to erode the rock. conjunction with wave quarrying, abrasion, and beach devel-
Because the breaker depth is related to the wave height, high opment (Trenhaile, 2005), and in a wave erosional model
waves expend more of their energy in crossing wide, turbulent to investigate the development of sloping and horizontal
surf zones than lower waves crossing narrow surf zones; con- shore platforms (Trenhaile, 2008a, b). Other models have
sequently, there was no consistent relationship in the model considered the effect of weathering, acting alone, in shore
runs between the width of the shore platforms and deep water platform development (Trenhaile, 2001c, 2004b). These
wave height. models suggested that weathering is unable to produce wide
302 Rock Coasts

shore platforms by itself, although it can play an important Kamphuiss (1986) erosional equation and described a broad
direct role in lowering and widening wave-cut surfaces, as well coastal system that could be perturbed to represent manage-
as facilitating wave quarrying by acting along joints and other ment interventions and climate change (Dickson et al., 2007).
rock discontinuities, and by lowering platform surfaces, re- Walkden and Dickson (2008) used the model to explore the
ducing rates of wave attenuation. effects of sea-level rise on soft-rock shores with little beach
Many coasts are composed of rocks that break down into sediment. Ashton et al. (2011) reproduced the basic finding of
clays and other fine-grained particles that are carried offshore, Walkden and Dickson (2008) and used a simpler model to
or of rocks that break down too slowly to produce extensive discuss the potential responses of different types of cliffed
sandy or pebble beaches. Nevertheless, there is commonly a coasts to changes in the rate of sea-level rise. Trenhailes
thin strip of sand or pebble at the cliff foot, or in small pockets (2009) model combined elements of previous hard rock
trapped against scarps or in topographic depressions, and models (Trenhaile, 2000, 2001a) with additional expressions
some rocky intertidal zones are covered entirely in beach representing the effect of wave-generated bottom shear stres-
sediment. The occurrence and morphodynamics of platform- ses, and depending on beach thickness, determined using
beaches has received little attention in the literature, although Trenhailes (2004c) model, the abrasional or protective role of
many beaches have a rock foundation and will therefore re- beach material. This combined model has also been used to
spond in a different way to storms and rising sea level than study the effect of rising sea level and increased storminess on
beaches on entirely depositional coasts. Trenhaile (2004c) hard and soft rock coasts (Trenhaile, 2010a, b, 2011).
modeled beaches on shore platforms to determine the con-
ditions under which deposition takes place, and how plat-
form-beaches adjust to changing wave conditions. The model
10.11.4.2 Plan Shape
was based on the assumptions that: sediment accumulation
occurs where the local gradient of the platform is lower than Much less work has been conducted on the plan shape than
the gradient of the beach face (which depends on grain size, on the sectional shape of rock coasts, or in integrating changes
breaker height, and wave period); and deposition takes place in the vertical and horizontal planes. Many rocky coasts have
on the most landward of the suitable sites. Consequently, if an irregular plan shape consisting of protruding headlands
the surface is linear and has a gradient that is lower than the and intervening fiord or ria inlets, or shallower bays that are
beach face for the prevailing wave conditions, sediment ac- commonly occupied by bay-head beaches. Crenulated estu-
cumulates at the cliff foot, and then, depending on sediment arine coasts were created by the drowning of river valleys by
availability, extends increasingly seaward. On concave and rising sea level during the Holocene, and their morphology is
convex platform profiles, accumulation begins on the gently not directly related to differences in the resistance of the rocks
sloping, seaward portions and on the more gently sloping to erosion, although river valleys frequently develop along
landward portions, respectively, as long as these local gradi- faultlines or other zones of structural or lithological weakness.
ents are lower than the beach face gradient. During storms, In southeastern Australia, for example, the length of estuarine-
platform-beaches become thinner and more gently sloping bays is related to the area of the associated drainage basins,
and, depending on such factors as sediment grain size, plat- and their size depends upon the stream order at the coast and
form gradient, wave characteristics, and storm duration, may the degree to which the sea has penetrated into the drainage
disappear completely. The model suggested that because network (Bishop and Cowell, 1997). Conversely, bays that are
platform and sandy beach face gradients are generally similar, not at the mouth of river valleys are generally produced by
only a small proportion of the material eroded from the cliff differential erosion of material that is weaker than in the ad-
and platform can be stored in the intertidal zone. The beach- jacent headlands. Variations in erosional efficacy along a coast
platform model was also used as a component of a wave may reflect the occurrence of different rock strata striking at
erosional-weathering model to predict spatial and temporal high angles to the coast, the presence of folded rocks, faulting,
variations in the protection afforded to the underlying rock or more subtle, and harder to distinguish, variations in rock
surface by thick beach deposits and the abrasive effects of structure. The amount of debris that accumulates at the foot of
thinner accumulations (Trenhaile, 2005). a cliff, protecting it from erosion, is partly determined by the
The term soft rock is commonly used to refer to cohesive height of the cliff. This factor is of little importance in clays
clays, shales, and other weak or weathered materials. Cohesive and other fine-grained materials that are quickly broken up
clay coasts represent a soft rock subset of a broad spectrum and carried offshore in suspension, but it can play an im-
of rock coast processes and forms, with cliffs and shore plat- portant role in rocks, that because of joint density and bed
forms that are similar to those in quasihomogeneous shales. thickness, break down into large rock fragments that remain at
Soft rock coasts experience fairly rapid coastal erosion the cliff foot for years. Therefore, variations in cliff height can
and the consequent threat to human life, and property has influence cliff recession rates, and because of low cliffs around
generated a need for predictive bluff recession and other ero- the mouths of rivers, cause bays to develop without any
sional models with constant and rising lake and sea levels marked alongshore variation in the resistance of the rock to
(Bray and Hooke, 1997; Kamphuis, 1986; Nairn et al., 1986; erosion.
Sallenger et al., 2005; Lee, 2005; Collins and Sitar, 2008; The long-term development of crenulated coasts is gov-
Hapke et al., 2009). erned by several important feedback mechanisms. As bays
Several attempts have been made to model the effect of develop along outcrops of weaker material, offshore gradients
changing sea level on soft rock coasts (Dar and Dar, 2009). become much lower than off the adjacent headlands and
Walkden and Halls (2005) model was partly based on the waves are consequently more attenuated in the bays.
Rock Coasts 303

Additionally, as the bays become deeper, wave refraction in- terrestrial laser scanning, cosmogenic dating, and wave re-
creasingly concentrates wave energy on the headlands and cording, is improving our understanding of these common
reduces it in the bays; bay erosion may be further inhibited by coastal landforms. However, significant challenges remain
the accumulation of eroded material in bay-head beaches. particularly in terms of representing rock resistance to erosion
Therefore, theory suggests that crenulated coasts eventually processes across a range of scales, unraveling inheritance, and
attain a state of equilibrium, whereby stronger, less attenuated the long term evolution of cliffs and shore platforms. Further
and refracted waves erode largely bare cliff foots on headlands field-based investigations of surf zone hydrodynamics are re-
at the same rate as weaker, more attenuated and refracted quired to better understand the erosional effectiveness of wave
waves erode beach-protected cliffs in bays (Johnson, 1919; processes and the significance, if any, of infragravity energy on
Muir-Wood, 1971; Komar, 1976; Philpott, 1984). It might be rock coasts. The role of weathering and biological processes in
expected that the depth of the bays would increase with the reducing rock resistance still has not been fully explained. For
difference in the erosional resistance of the rocks in the bays example, the relative contributions of bioerosion, protection,
and headlands, and decrease with the distance between the and construction at any one site should be assessed. Further
headlands. modeling will aid in predicting how rock coasts respond to
The development of crenulated coasts has important im- climate change and sea-level rise.
plications for the prediction and mitigation of cliff erosion,
and particularly for anticipating the effect of rising sea level.
We lack reliable long-term records of episodic cliff recession
References
on crenulated coasts, however, and it is not known whether
Aarseth, I., Fossen, H., 2004. Holocene lacustrine rock platform around Storavatnet,
these coasts trend toward equilibrium or whether there has Ostery, western Norway. The Holocene 14, 589596.
been enough time for such states to have developed. Fur- Agar, R., 1960. Postglacial erosion of the north Yorkshire coast from the Tees
thermore, it cannot be assumed that coasts become increas- estuary to Ravenscar. Proceedings of the Yorkshire Geological Society 32,
ingly crenulated, albeit at a decreasing rate, until an 408425.
equilibrium is attained. Profile adjustment to changing sea Allan, J.C., Stephenson, W.J., Kirk, R.M., Taylor, A., 2002. Lacustrine shore
platforms at Lake Waikaremoana, North Island, New Zealand. Earth Surface
level during the Quaternary would have disturbed any simple Processes and Landforms 27, 207220.
trend toward an equilibrium state (Trenhaile, 2002) and cre- Anderson, R.S., Densmore, A.L., Ellis, M.A., 1999. The generation and degradation
nulated coasts are also affected in many areas by tectonic or of marine terraces. Basin Research 11, 719.
isostatic adjustments in the elevation of the land. Andrews, C., Williams, R.B.G., 2000. Limpet erosion of chalk shore platforms in
southeast England. Earth Surface Processes and Landforms 25, 13711381.
Small amounts of abrasive beach material can increase
Arber, M.A., 1949. Cliff profiles of Devon and Cornwall. Geographical Journal 114,
rates of cliff and platform erosion in bays, but large amounts 191197.
may protect the cliff and allow erosion of the exposed head- Ashton, A., Walkden, M.J.A., Dickson, M.E., 2011. Equilibrium responses
lands to reduce the degree of crenulation; this could eventu- of rocky coasts to changes in the rate of sea level rise. Marine Geology 284,
ally permit bay-head sediment to escape alongshore. 217229.
Bartrum, J.A., 1916. High-water rock-platforms: a phase of shore-line erosion.
Therefore, beaches may play an important role in the devel-
Transactions and Proceedings of the New Zealand Institute 48, 132134.
opment of rocky coasts where there is a copious supply of Bartrum, J.A., 1924. The shore-platform of the west coast near Auckland: its storm
beach material from rivers or from rapidly eroding cliffs, and wave origin. Australian and New Zealand Association for the Advancement of
particularly where, as on plate collision coasts, the rocks are Science 16, 493495.
fairly homogeneous alongshore, and the plan shape presents Bartrum, J.A., 1926. Abnormal shore platforms. Journal of Geology 34, 793807.
Bartrum, J.A., Turner, F.J., 1928. Pillow Lavas, Peridotites, and Associated Rocks
few obstacles to longshore transport. It has been proposed From Northernmost New Zealand. Transactions New Zealand Institute 59,
that longshore variations in beach width, and consequently in 98138.
cliff and shore platform abrasion and protection, may provide Beetham, E.P., Kench, P.S., 2011. Field observations of infragravity waves and their
an additional explanation for the development of crenulated behavior on rock shore platforms. Earth Surface Processes and Landforms
coasts which does not depend upon differences in the resist- 36(4), 18721888.
Bird, E.C.F., Dent, O.F., 1966. Shore platforms on the south coast of New South
ance of the rocks to erosion. Preliminary numerical modeling Wales. Australian Geographer 10, 7180.
suggests that variations in the alongshore distribution of Bird, E.C.F., 1968. Coasts. Australian National University Press, Canberra, 246 p.
beach sediment may result in self-organization of headlands, Bishop, P., Cowell, P., 1997. Lithological and drainage network determinants of the
and that the headlands and intervening stretches of cliff- character of drowned, embayed coastlines. The Journal of Geology 105, 685699.
Blanco-Chao, R., Costa-Casais, M., Martnez-Cortizas, A., Perez-Alberti, A.,
backed pocket beaches all retreat at the same rate (Limber
Trenhaile, A.S., 2003. Evolution and inheritance of a rock coast: western Galicia,
et al., 2008, 2009). northwestern Spain. Earth Surface Processes and Landforms 28, 757775.
Blanco-Chao, R., Perez-Alberti, A., Trenhaile, A.S., Costa-Casais, M., Valcarcel-Diaz,
M., 2007. Shore platform abrasion in a para-periglacial environment, Galicia,
10.11.5 Conclusions northwestern Spain. Geomorphology 83, 136151.
Bray, M.J., Hooke, J.M., 1997. Prediction of soft-cliff retreat with accelerating sea-
level rise. Journal of Coastal Research 13, 453467.
Rock coasts dominate the worlds shorelines and will face Brooke, B.P., Young, R.W., Bryant, E.A., Murray-Wallace, C.V., Price, D.M., 1994. A
significant pressure from increasing development and occu- Pleistocene origin for shore platforms along the northern Illawarra coast, New
pation, and sea-level rise predicted by climate warming South Wales. Australian Geographer 25, 178185.
models. A resurgence in research activity in the last 15 years Bryant, E.A., Young, R.W., Price, D.M., 1992. Evidence of tsunami sedimentation on
the southeastern coast of Australia. Journal of Geology 100, 753765.
has resulted in significant advances in our understanding of Bryant, E.A., Young, R.W., Price, D.M., 1996. Tsunami as a major control on
the role of different processes and rates of development. The coastal evolution, southeastern Australia. Journal of Coastal Research 12,
application of new technologies and methods such as 831840.
304 Rock Coasts

Burt, T.P., 2003. Some observations on slope development in South Wales: Views on Some Aspects of the Geology of Cornwall and Devon. Royal
Savigear and Kirkby revisited. Progress in Physical Geography 27, 581595. Geological Society of Cornwall, Penzance, pp. 283310.
Challinor, J., 1949. A principle in coastal geomorphology. Geography 34, Farrell, E.J., Granja, H., Cappietti, L., Ellis, J.T., Li, B., Sherman, D.J., 2009. Wave
212215. Transformation across a rock platform, Belinho, Portugal. Journal of Coastal
Chappell, J., Shackleton, N.J., 1986. Oxygen isotopes and sea level. Nature 324, Research (Special Issue) 56, 4448.
137140. Felton, E.A., Crook, K.A.W., 2003. Evaluating the impacts of huge waves on rocky
Cinque, A., De Pippo, T., Romano, P., 1995. Coastal slope terracing and relative shorelines: an essay review of the book Tsunami The Underrated Hazard.
sea-level changes: deductions based on computer simulations. Earth Surface Marine Geology 197, 112.
Processes and Landforms 20, 87103. Fisher, O., 1866. On the disintegration of a chalk cliff. Geological Magazine 3,
Collins, B.D., Sitar, N., 2008. Processes of coastal bluff erosion in weakly lithified 354356.
sands, Pacifica, California, USA. Geomorphology 97, 483501. Fleming, C.A., 1965. Two-storied cliffs at the Auckland Islands. Transactions of the
Cooke, R.U., 1979. Laboratory Simulation of Salt Weathering Processes in Arid Royal Society of New Zealand 3, 171174.
Environments. Earth Surface Processes and Landforms 4, 347359. Flemming, N.C., 1965. Form and relation to present sea level of Pleistocene marine
Cooke, R.U., Smalley, I.J., 1968. Salt Weathering in Deserts. Nature 220, erosion features. Journal of Geology 73, 799811.
12261227. Gill, E.D., 1967. The dynamics of the shore platform process, and its relation to
Cotton, C.A., 1949. Plunging cliffs, Lyttleton Harbor. In: Collins, B.W. (Ed.), Bold changes in sea-level. Proceedings of the Royal Society of Victoria 80, 183192.
Coasts. A. H. & A. W.. Reed, Wellington, pp. 4651. Gill, E.D., 1972. The relationship of present shore platforms to past sea levels.
Cotton, C.A., 1951. Atlantic gulfs, estuaries and cliffs. In: Collins, B.W. (Ed.), 1974. Boreas 1, 125.
Bold Coasts. A. H. & A. W. Reed, Wellington, pp. 6983. Gill, E.D., Lang, J.G., 1983. Micro-erosion meter measurements of rock wear on the
Cotton, C.A., 1963. Levels of planation of marine benches. Zeitschrift fur Otway coast of southeast Australia. Marine Geology 52, 141156.
Geomorphologie 7, 97111. Gomez-Pujol, L., Cruslock, E.M., Fornos, J.J., Swantesson, J.O.H., 2006.
Cotton, C.A., 1967. Plunging cliffs and Pleistocene coastal cliffing in the Southern Unravelling factors that control shore platforms and cliffs in microtidal coasts:
Hemisphere. In: Collins, B.W. (Ed.), 1974. Bold Coasts. A. H. & A. W. Reed, the case of Mallorcan, Catalonian and Swedish coasts. Zeitschrift fur
Wellington, pp. 232245. Geomorphologie, Supplementband 144, 117135.
Cruslock, E.M., Naylor, L.A., Foote, Y.L., Swantesson, J.O.H., 2010. Goto, K., Miyagi, K., Kawamata, H., Imamura, F., 2010. Discrimination of boulders
Geomorphologic equifinality: a comparison between shore platforms in Hoga deposited by tsunamis and storm waves at Ishigaki Island, Japan. Marine
Kusten and Faro, Sweden and the Vale of Glamorgan, South Wales, UK. Geology 269, 3445.
Geomorphology 114, 7888. Goudie, A., Cooke, R.U., Evans, I., 1970. Experimental Investigation of rock
Daly, R.A., 1927. The geology of Saint Helena Island. Proceedings of the American weathering by salts. Area 4, 4248.
Academy of Arts and Sciences 62, 3192. Griggs, G.B., Trenhaile, A.S., 1994. Coastal cliffs and platforms. In: Carter, R.W.G.,
Dana, J.D., 1849. Geology, Report of the U.S. Exploration Expedition during the Woodroffe, C.D. (Eds.), Coastal Evolution. Cambridge University Press,
Years 1838 to 1842. Putnam. New York. Cambridge, pp. 425450.
Dar, I.A., Dar, M.A., 2009. Prediction of shoreline recession using geospatial Guilcher, A., 1958. Coastal and Submarine Morphology. Methuen, London.
technology: a case study of Chennai Coast, Tamil Nadu, India. Journal of Hall, A.M., Hansom, J.D., Jarvis, J., 2008. Patterns and rates of erosion produced
Coastal Research 25, 12761286. by high energy wave processes on hard rock headlands: the Grind of the Navir,
Darwin, C.R., 1844. Geological observations on the volcanic islands visited during Shetland, Scotland. Marine Geology 248, 2846.
the voyage of H.M.S. Beagle, together with some brief notices of the geology of Hall, A.M., Hansom, J.D., Williams, D.M., Jarvis, J., 2006. Distribution,
Australia and the Cape of Good Hope. Being the second part of the geology of geomorphology and lithofacies of cliff-top storm deposits, Examples
the voyage of the Beagle, under the command of Capt. Fitzroy, R.N. during the from the high-energy coasts of Scotland and Ireland. Marine Geology 232,
years 1832 to 1836. Smith Elder and Co, London. 131155.
Darwin, C.R., 1846. Geological observations on South America. Being the Third Hampton, M.A., Griggs, G.B. (Eds.), 2004. Formation, Evolution, and Stability of
Part of the Teology of the Voyage of the Beagle, under the Command of Capt. Coastal Cliffs Status and trends. United States Geological Survey. Professional
Fitzroy, R.N. during the Years 1832 to 1836. Smith Elder and Co, London. paper 1693.
Davis, W.M., 1896. Plains of marine and subaerial denudation. Geological Society Hansom, J.D., Barltrop, N.D.P., Hall, A.M., 2008. Modeling the processes of cliff-
of America Bulletin 7, 377398. top erosion and deposition under extreme storm waves. Marine Geology 253,
Davis, W.M., 1928. The Coral Reef Problem. American Geographical Society, New 3650.
York, Special Publication. 9. Hapke, C., Reid, D., Richmond, B., 2009. Rates and trends of coastal change in
Dickson, M.E., 2004. The development of talus slopes around Lord Howe Island California and the regional behavior of the beach and cliff system. Journal of
and implications for the history of island plantation. Australian Geographer 35, Coastal Research 25, 603615.
223238. Healy, T.R. 1968. Shore platform morphology on the Whangaparaoa Peninsula,
Dickson, M.E., 2006. Shore platform development around Lord Howe Island, Auckland. Conference Series, New Zealand Geographical Society. 5, 163-168.
southwest Pacific. Geomorphology 76, 295315. Hills, E.S., 1949. Shore platforms. Geological Magazine 86, 137152.
Dickson, M.E., Woodroffe, C.D., 2005. Rock coast morphology in relation to Hills, E.S., 1971. A study of cliffy coast profiles in Victoria, Australia. Zeitschrift fur
lithology and wave exposure, Lord Howe Island, southwest Pacific. Zeitschrift fur Geomorphologie 15, 137180.
Geomorphologie 49, 239251. Hills, E.S., 1972. Shore platforms and wave ramps. Geological Magazine. 109,
Dickson, M.E., Walkden, M.J.A., Hall, J.W., 2007. Systemic impacts of climate 8188.
change on an eroding coastal region over the twenty-first century. Climatic Horikawa, K., Sunamura, T., 1967. A study on erosion of coastal cliffs by using
Change 84, 141166. aerial photographs. Coastal Engineering in Japan 10, 6783.
Dietz, R.S., Menard, H.W., 1951. Origin of abrupt changes in slope at the Jerwood, L.C., Robinson, D.A., Williams, R.B.G., 1990a. Experimental frost and salt
continental shelf margin. Bulletin of the American Association of Petrolium weathering of chalk .1. Earth Surface Processes and Landforms 15, 611624.
Geologists 35, 19942016. Jerwood, L.C., Robinson, D.A., Williams, R.B.G., 1990b. Experimental frost and salt
Dietz, R.S., 1963. Wave-base, marine profile equilibrium, and wave-built terraces. weathering of chalk 2. Earth Surface Processes and Landforms 15, 699708.
Geological Society of America Bulletin 74, 971990. Johnson, D.W., 1919. Shore Processes and Shoreline Development. Wiley, New
Edwards, A.B., 1941. Storm wave platforms. Journal of Geomorphology 4, York, 584 pp.
223236. Johnson, D.W., 1925. The New England-Acadian Shoreline. Wiley, New York, N.Y,
Edwards, A.B., 1951. Wave action in shore platform formation. Geological 608 pp.
Magazine. 88, 4149. Jutson, J.T., 1939. Shore platforms near Sydney, New South Wales. Journal of
Emery, K.O., Kuhn, G.G., 1982. Sea cliffs: their processes, profiles, and Geomorphology 2, 236250.
classification. Geological Society of America Bulletin 93, 644654. Kamphuis, J.W., 1986. Erosion of cohesive bluffs, a model and a formula.
Etienne, S., Paris, R., 2010. Boulder accumulations related to storms on the south In: Skafel, M.G. (Ed.), Proceedings of the Symposium on Cohesive Shores.
coast of the Reykjanes Peninsula (Iceland). Geomorphology 114, 5570. National Research Council of Canada (ACROSES), Ottawa, pp. 226245.
Everard, C.E., Lawrence, R.H., Witherick, M.E., Wright, L.W., 1964. Raised beaches Kanyaya, J.I., Trenhaile, A.S., 2005. Tidal wetting and drying on shore platforms: an
and marine geomorphology. In: Hosking, K.F.G., Shrimpton, G.J. (Eds.), Present experimental assessment. Geomorphology 70, 129146.
Rock Coasts 305

Kennedy, D.M., 2010. Geological control on the morphology of estuarine shore Noormets, R., Crook, K.A.W., Felton, E.A., 2004. Sedimentology of rocky shorelines:
platforms: Middle Harbour, Sydney, Australia. Geomorphology 114, 7177. 3.: hydrodynamics of megaclast emplacement and transport on a shore platform,
Kennedy, D.M., Beban, J.G., 2005. Shore platform morphology on a rapidly Oahu, Hawaii. Sedimentary Geology 172, 4165.
uplifting coast, Wellington, New Zealand. Earth Surface Processes and Noormets, R., Felton, E.A., Crook, K.A.W., 2002. Sedimentology of rocky shorelines:
Landforms 30, 823832. 2 Shoreline megaclasts on the north shore of Oahu, Hawaii origins and
Kennedy, D.M., Dickson, M.E., 2006. Lithological control on the elevation of shore history. Sedimentary Geology 150, 3145.
platforms in a microtidal setting. Earth Surface Processes and Landforms 31, Nott, J.F., 1990. The role of sub-aerial processes in sea cliff retreat a south east
15751584. Australian example. Zeitschrift fur Geomorphologie 34, 7585.
Kennedy, D.M., Paulik, R., Dickson, M.E., 2010. Subaerial weathering versus wave Ogawa, H., Dickson, M.E., Kench, P.S., 2011. Wave transformation on a sub-
processes in shore platform development: reappraising the Old Hat Island horizontal shore platform, Tatapouri, North Island, New Zealand. Continental
evidence. Earth Surface Processes and Landforms 36(5), 686694. Shelf Research 31, 14091419.
King, C.A.M., 1963. Some problems concerning marine planation and the formation Ogawa, H., Kench, P.S., Dickson, M.E., 2012. Field measurements of wave
of erosion surfaces. Transactions of the Institute of British Geographers 33, characteristics on a near-horizontal shore platform, Mahia Peninsula,
2943. North Island, New Zealand. 50(2), 179192.
Kirk, R.M., 1977. Rates and forms of erosion on intertidal platforms at Kaikoura Ongley, M., 1940. Note on coastal benches formed by spray weathering. New
Peninsula, South Island, New Zealand. New Zealand Journal of Geology and Zealand Journal of Science and Technology 22, 34b35b.
Geophysics 20, 571613. Palamara, D.R., Dickson, M.E., Kennedy, D.M., 2007. Defining shore platform
Kirkby, M.J., 1984. Modeling cliff development on south Wales: savigear reviewed. boundaries using airborne laser scan data: a preliminary investigation. Earth
Zeitschrift fur Geomorphologie 28, 405426. Surface Processes and Landforms 32, 945953.
Knight, J., Burningham, H., 2011. Boulder dynamics on an Atlantic-facing rock Paris, R., Naylor, L.A., Stephenson, W.J., 2011. Boulders as a signature of storms
coastline, northwest Ireland. Marine Geology 283, 5665. on rock coasts. Marine Geology 283, 111.
Komar, P.D., 1976. Beach processes and Sedimentation. Prentice-Hall, Englewood Philpott, K.L., 1984. Comparison of cohesive coasts and beach coasts.
Cliffs, NJ. In: Kamphuis, J.W. (Ed.), Proceedings, Coastal Engineering in Canada. Queens
Lee, E.M., 2005. Coastal cliff recession risk: a simple judgement-based model. University. Kingston, Ontario, pp. 227244.
Quarterly Journal of Engineering Geology and Hydrogeology 38, 89104. Pirazzoli, P.A., 1991. World Atlas of Holocene Sea-level Changes. Elsevier,
Lee, M., Clark, M., 2002. Investigation and Management of Soft Rock Cliffs. Amsterdam, 300p.
Thomas Telford, London, 382 pp. Porter, N.J., Trenhaile, A.S., 2007. Short-term rock surface expansion and
Lim, M., Petley, D.N., Rosser, N.J., Allison, R.J., Long, A.J., 2005. Combined contraction in the intertidal zone. Earth Surface Processes and Landforms 32,
digital photogrammetry and time-of-flight laser scanning for monitoring cliff 13791397.
evolution. The Photogrammetric Record 20, 109129. Porter, N.J., Trenhaile, A.S., Prestanski, K.J., Kanyaya, J.I., 2010a. Shore platform
Lim, M., Rosser, N.J., Allison, R.J., Petley, D.N., 2010a. Erosional processes in the downwearing in eastern Canada: the mega-tidal Bay of Fundy. Geomorphology
hard rock coastal cliffs at Staithes, North Yorkshire. Geomorphology 114, 1221. 118, 112.
Lim, M., Rosser, N.J., Petley, D.N., Norman, E.C., Barlow, J., Brain, M.J., 2010b. Porter, N.J., Trenhaile, A.S., Prestanski, K.J., Kanyaya, J.I., 2010b. Shore platform
Characterising coastal cliff erosional environments: microclimate-rock interactions. downwearing in eastern Canada: microtidal Gaspe, Quebec. Geomorphology 116,
In: Williams, A.L., Pinches, G.M., Chin, C.Y., McMorran, T.J., Massey, C.I. (Eds.), 7786.
Geologically Active. Taylor & Francis Group, London, pp. 393400. Porter, N.J., Trenhaile, A.S., Prestanski, K., Kanyaya, J.I. 2010c. Patterns of surface
Limber, P., Murray, A.B., Littlewood, R., Valvo, L., 2008. Numerical modeling of downwearing on shore platforms in eastern Canada. Earth Surface Processes
large scale rocky coastline evolution. Eos Trans. AGU 89(53), Fall Meet. Suppl., and Landforms 35, 17931810.
Abstract OS23A-1250. Quartau, R., Trenhaile, A.S., Mitchell, N.C., Tempera, F., 2010. Development of
Limber, P.W., Murray, A.B., Adams, P., 2009. The growth of headlands in a simple volcanic insular shelves: insights from observations and modelling of Faial
model of rocky coastline evolution. Eos Trans. AGU 90(52), Fall Meet. Suppl., Island in the Azores Archipelago. Marine Geology 275, 6683.
Abstract EP33C-02. Ramsey, A.C., 1846. On the denudation of south Wales and the adjacent counties
Marshall, R.J.E., Stephenson, W.J., 2011. The morphodynamics of shore platforms: of England. Memoirs, Geological. Survey. G.B. 1, 297335.
interactions between waves and morphology. Marine Geology 288, 1831. Reffell, G., 1978. Descriptive Analysis of the Subaqueous Extensions of Subaerial
Matthews, J.A., Dawson, A.G., Shakesby, R.A., 1986. Lake shoreline development, Rock Platforms. University of Sydney, Sydney, B.A. Thesis.
frost weathering and rock platform erosion in an alpine periglacial environment. Robinson, L.A., 1977a. Marine erosive processes at the Cliff Foot. Marine Geology
Boreas 15, 3350. 23, 257271.
Morton, R.A., Richmond, B.M., Jaffe, B.E., Gelfenbaum, G. 2006. Reconnaissance Robinson, L.A., 1977b. Erosive Processes on the Shore Platform of Northeast
investigation of Caribbean extreme wave deposits preliminary observations, Yorkshire Shore Platform. Marine Geology 23, 339361.
interpretations, and research directions. USGS Open File Report 20061293. Robinson, L.A., 1977c. The Morphology and Development of the Northeast
Muir-Wood, A.M., 1971. Engineering aspects of coastal landslides. Proceedings Yorkshire Shore Platform. Marine Geology 23, 237255.
Institute of Civil Engineers 50, 257276. Rosenbloom, N.A., Anderson, R.S., 1994. Hillslope and channel evolution in a
Nairn, R.B., Pinchin, B.M., Philpott, K.L., 1986. Cohesive profile development marine terraced landscape, Santa Cruz, California. Journal of Geophysical
model. In: Skafel, M.G. (Ed.), Proceedings of the Symposium on Cohesive Research 99(B7), 1401314029.
Shores. National Research Council of Canada (ACROSES), Ottawa, pp. 246261. Sallenger, A.H., Krabill, W., Brock, J., Swift, R., Manizade, S., Stockdon, H., 2005.
Naylor, L.A., Coombes, M.A., Viles, H.A., 2012. Reconceptualising the role Sea-cliff erosion as a function of beach changes and extreme wave runup during
of organisms in the erosion of rock coasts: A new model. Geomorphology the 19971998 El Nino. Marine Geology 187, 279297.
157158, 1730. Sanders, N.K., 1968. Wave tank experiments on the erosion of rocky coasts.
Naylor, L.A., Stephenson, W.J., 2010. On the role of discontinuities in mediating Proceedings of the Royal Society of Tasmania 102, 1116.
shore platform erosion. Geomorphology 114, 89100. Savigear, R.A.G., 1952. Some observations on slope development in South Wales.
Naylor, L.A., Stephenson, W.J., Trenhaile, A.S., 2010. Rock coast geomorphology: Institute of British Geographers, Transactions 18, 3152.
recent advances and future research directions. Geomorphology 114, 311. Scheffers, A., Scheffers, S., Kelletat, D., 2005. Paleotsunami relics on the
Naylor, L.A., Viles, H.A., 2000. A temperate reef builder: an evaluation of the southern and central Antillean Island Arc. Journal of Coastal Research 21,
growth, morphology and composition of Sabellaria alveolata (L.) colonies on 263273.
carbonate platforms in South Wales. Geological Society Special Publication 178, Scheffers, A., Scheffers, S., 2006. Documentation of the Impact of Hurricane Ivan
919. on the Coastline of Bonaire (Netherlands Antilles). Journal of Coastal Research
Naylor, L.A., Viles, H.A., 2002. A new technique for evaluating short-term rates of 22, 14371450.
coastal bioerosion and bioprotection. Geomorphology 47, 3144. Scheidegger, A.E., 1962. Marine terraces. Pure and Applied Geophysics 52, 6982.
Naylor, L.A., Viles, H.A., Carter, N.E.A., 2002. Biogeomorphology revisited: looking Scheidegger, A.E., 1970. Theoretical Geomorphology. Springer-Verlag, New York,
towards the future. Geomorphology 47, 314. 435 p.
Neumann, A.C., 1966. Observations on coastal erosion in Bermuda and Shepard, F.P., Kuhn, G.G., 1983. History of sea arches and remnant stacks of La
measurement of the boring rate of the Sponge Cliona lampa. Limnology Jolla, California, and their bearing on similar features elsewhere. Marine
Oceanography 11, 92108. Geology 51, 139161.
306 Rock Coasts

So, C.L., 1965. Coastal platforms of the Isle of Thanet, Kent. Transactions of the Trenhaile, A.S., 1980. Shore platforms: a neglected coastal feature. Progress in
Institute of British Geographers 37, 147156. Physical Geography 4, 123.
Stephenson, W.J., 2000. Shore platforms: remain a neglected coastal feature? Trenhaile, A.S., 1983. The width of shore platforms; a theoretical approach.
Progress in Physical Geography 24, 311327. Geografiska Annaler 65A, 147158.
Stephenson, W.J., Finlayson, B.L., 2009. Measuring erosion with the micro-erosion Trenhaile, A.S., 1987. The Geomorphology of Rock Coasts. Clarendon Press,
meter contributions to understanding landform evolution and building stone Oxford.
decay. Earth Science Reviews 95, 5362. Trenhaile, A.S., 1989. Sea level oscillations and the development of rock coasts.
Stephenson, W.J., Kirk, R.M., 1998. Rates and patterns of erosion on intertidal In: Lakhan, V.C., Trenhaile, A.S. (Eds.), Applications in Coastal Modeling.
shore platforms, Kaikoura Peninsula, South Island, New Zealand. Earth Surface Elsevier, Amsterdam, pp. 271295.
Processes and Landforms 23, 10711085. Trenhaile, A.S., 1999. The width of shore platforms in Britain, Canada, and Japan.
Stephenson, W.J., Kirk, R.M., 2000a. Development of shore platforms on Kaikoura Journal of Coastal Research 15, 355364.
Peninsula, South Island, New Zealand. Part One: the role of waves. Trenhaile, A.S., 2000. Modeling the development of wave-cut shore platforms.
Geomorphology 32, 2141. Marine Geology 166, 163178.
Stephenson, W.J., Kirk, R.M., 2000b. Development of shore platforms on Kaikoura Trenhaile, A.S., 2001a. Modeling the effect of late Quaternary interglacial sea levels
Peninsula, South Island, New Zealand. II: the role of subaerial weathering. on wave-cut shore platforms. Marine Geology 172, 205223.
Geomorphology 32, 4356. Trenhaile, A.S., 2001b. Modeling the Quaternary evolution of shore platforms and
Stephenson, W.J., Kirk, R.M., 2001. Surface swelling of coastal bedrock on inter- erosional continental shelves. Earth Surface Processes and Landforms 26,
tidal shore platforms, Kaikoura Peninsula, South Island, New Zealand. 11031128.
Geomorphology 41, 521. Trenhaile, A.S., 2001c. Modeling the effect of weathering on the evolution and
Stephenson, W.J., Naylor, L.A., 2011a. Geological controls on boulder production in morphology of shore platforms. Journal of Coastal Research 17, 398406.
a rock coast setting: insights from South Wales, UK. Marine Geology 283, Trenhaile, A.S., 2002. Modeling the development of sloping marine terraces on
1224. tectonically mobile rock coasts. Marine Geology 185, 341361.
Stephenson, W.J., Naylor, L.A., 2011b. Within site geological contingency and its Trenhaile, A.S., 2003. Modeling shore platforms : present status and future
effect on rock coast erosion. Journal of Coastal Research SI 64, 831835. developments. In: Lakhan, V.C. (Ed.), Advances in Coastal Modeling. Elsevier,
Stephenson, W.J., Taylor, A.J., Hemmingsen, M.A., Tsujimoto, H., Kirk, R.M., 2004. Amsterdam, pp. 393409.
Short-term microscale topographic changes of coastal bedrock on shore Trenhaile, A.S., 2004a. Lacustrine shore platforms in southwestern Ontario, Canada.
platforms. Earth Surface Processes and Landforms 29, 16631673. Zeitschrift fur Geomorphologie 48, 441459.
Stone, J., Lambeck, K., Fifield, L.K., Evans, J.M., Cresswell, R.G., 1996. A Trenhaile, A.S., 2004b. Modeling the effect of tidal wetting and drying on shore
lateglacial age for the Main Rock Platform, western Scotland. Geology 24, platform development. Journal of Coastal Research 20, 10491060.
707710. Trenhaile, A.S., 2004c. Modeling the accumulation and dynamics of beaches on
Suanez, S., Fichaut, B., Magne, R., 2009. Cliff-top storm deposits on Banneg shore platforms. Marine Geology 206, 5572.
Island, Brittany, France: Effects of giant waves in the Eastern Atlantic Ocean. Trenhaile, A.S., 2005. Modeling the effect of waves, weathering and beach
Sedimentary Geology 220, 1228. development on shore platform development. Earth Surface Processes and
Sunamura, T., 1973. Coastal cliff erosion due to waves field investigations and Landforms 30, 613634.
laboratory experiments. Journal Faculty Engineering University of Tokyo 32, Trenhaile, A., 2006. Tidal wetting and drying on shore platforms, an experimental
186. study of surface expansion and contraction. Geomorphology 76, 316331.
Sunamura, T., 1975. A laboratory study of wave-cut platform formation. Journal of Trenhaile, A.S., 2008a. The development of subhorizontal shore platforms in
Geology 83, 389397. microtidal environments. Zeitschrift fur Geomorphologie 52, 105124.
Sunamura, T., 1976. Feedback relationship in wave erosion of laboratory rocky Trenhaile, A.S., 2008b. Modeling the role of weathering on shore platform
coast. Journal of Geology 84, 427437. development. Geomorphology 94, 2439.
Sunamura, T., 1977. A relationship between wave-induced cliff erosion and erosive Trenhaile, A.S., 2009. Modeling the erosion of cohesive clay coasts. Coastal
force of wave. Journal of Geology 85, 613618. Engineering 56, 5972.
Sunamura, T., 1978a. A mathematical model of submarine platform development. Trenhaile, A.S., 2010a. The effect of Holocene changes in relative sea level on the
Mathematical Geology 10, 5358. morphology of rocky coasts. Geomorphology 114, 3041.
Sunamura, T., 1978b. A model of the development of continental shelves having Trenhaile, A.S., 2010b. Modeling cohesive clay coast evolution and response to
erosional origin. Geological Society of America Bulletin 89, 504510. climate change. Marine Geology 277, 1120.
Sunamura, T., 1991. The elevation of shore platforms: a laboratory approach to the Trenhaile, A.S. 2011. Predicting the response of hard and soft rock coasts to
unsolved problem. Journal of Geology 99, 761766. changes in sea level and wave height. Climatic Change 109, 599615.
Sunamura, T., 1992. Geomorphology of rocky coasts. Wiley, New York, 302p. Trenhaile, A.S., Byrne, M.L., 1986. A theoretical investigation of rock coasts, with
Sunamura, T., 1994. Rock control in coastal geomorphic processes. Transactions particular reference to shore platforms. Geografiska Annaler 68A, 114.
Japanese Geomorphological Union 15, 253272. Trenhaile, A.S., Kanyaya, J.I., 2007. The role of wave erosion on sloping and
Swan, S.B.S.T.C., 1971. Coastal geomorphology in a humid tropical low energy horizontal shore platforms in macro- and mesotidal environments. Journal of
environment: the islands of Singapore. Journal of Tropical Geography 33, Coastal Research 23, 298309.
4361. Trenhaile, A.S., Layzell, M.G.J., 1981. Shore platforms morphology and the tidal
Switzer, A.D., Burston, J.M., 2010. Competing mechanisms for boulder deposition duration factor. Transactions of the Institute of British Geomorphologists 6,
on the southeast Australian coast. Geomorphology 114, 4254. 82102.
Takahashi, T., 1977. Shore Platforms in Southwestern Japan Geomorphological Trenhaile, A.S., Mercan, D.W., 1984. Frost weathering and the saturation of coastal
Study. Coastal Landform Study Society, Southwestern Japan, Osaka. rocks. Earth Surface Processes and Landforms 9, 321331.
Taylor, A.J., 2003. Change and processes of change on shore platforms. University Trenhaile, A.S., Pepper, D.A., Trenhaile, R.W., Dalimonte, M., 1998. Stacks and
of Canterbury, Christchurch, 386 pp. PhD Thesis. notches at Hopewell Rocks, New Brunswick, Canada. Earth Surface Processes
Thornton, E.B., Guza, R.T., 1982. Energy saturation and phase speeds measured on and Landforms 23, 975988.
natural beaches. Journal of Geophysical Research 87, 94999508. Trenhaile, A.S., Perez Alberti, A., Martnez Cortizas, A., Costa Casais, M., Blanco
Thornton, L.E., Stephenson, W.J., 2006. Rock strength: a control of shore platform Chao, R., 1999. Rock Coast Inheritance: an Example from Galicia, Northwestern
elevation. Journal of Coastal Research 22, 224231. Spain. Earth Surface Processes and Landforms 24, 605621.
Tingstad, A., 2008. Simulation of salt weathering in a closely replicated coastal Trenhaile, A.S., Porter, N.J., Kanyaya, J.I., 2006. Shore platform processes in
environment. Geografiska Annaler 90A, 165171. eastern Canada. Geographie Physique et Quaternaire 60, 1930.
Trenhaile, A.S., 1969. A geomorphological investigation of shore platforms and Trudgill, S.T., 1976. The marine erosion of limestones on Aldabra Atoll, Indian
high-water rock ledges in the Vale of Glamorgan. PhD Thesis, University of Ocean. Zeitschrift fur Geomorphologie Supplementband 26, 164200.
Wales. Tsujimoto, H., 1987. Dynamic conditions for shore platform initiation. Science
Trenhaile, A.S., 1972. The shore platforms of the Vale of Glamorgan, Wales. Report of the Institute of Geoscience, University of Tsukuba A8, 4593.
Transactions of the Institute of British Geographers 56, 127144. Walkden, M., Dickson, M., 2008. Equilibrium erosion of soft rock shores with a
Trenhaile, A.S., 1974. The geometry of shore platforms in England and Wales. shallow or absent beach under increased sea level rise. Marine Geology 251,
Transactions of the Institute of British Geographers 62, 129142. 7584.
Rock Coasts 307

Walkden, M.J.A., Hall, J.W., 2005. A predictive mesoscale model of the erosion and Wright, L.W., 1967. Some characteristics of the shore platforms of the English
profile development of soft rock shores. Coastal Engineering 52, 535563. Channel Coast and the northern part of the North Island of New Zealand.
Wood, A., 1959. The erosional history of the cliffs around Aberystwyth. Liverpool Zeitschrift fur Geomorphologie 11, 3646.
and Manchester. Geology Journal 2, 271287. Wright, L.W., 1970. Variation in the level of the cliff/shore platform junction along
Wood, A., 1968. Beach platforms in the chalk of Kent, England. Zeitschrift fur the south coast of Great Britain. Marine Geology 9, 347353.
Geomorphologie 12, 107113. Yatsu, E. 1988. The nature of weathering an introduction. Sozosha, Tokyo, Japan.
Woodroffe, C.D., Bryant, E.A., Price, D.M., Short, S.A., 1992. Quaternary inheritance Young, R.W., Bryant, E.A., 1993. Coastal rock platforms and ramps of Pleistocene
of coastal landforms, Cobourg Peninsula, Northern Territory. Australian and Tertiary age in southern New South Wales, Australia. Zeitschrift fur
Geographer 23, 101115. Geomorphologie 37, 257272.
Woodroffe, C.D., Heijnis, H., Price, D.M., Murray-Wallace, C.V., Bryant, E.A., Young, R.W., Bryant, E.A., Price, D.M., 1996. Catastrophic wave (tsunami?)
Brooke, B., 1995. Late Quaternary sea-level highstands in the Tasman Sea: transport of boulders in southern New South Wales, Australia. Zeitschrift fur
evidence from Lord Howe Island. Marine Geology 125, 6172. Geomorphologie 40, 191207.

Biographical Sketch

Wayne Stephenson is currently a Senior Lecturer in Geography in the Department of Geography at the University of Otago in Dunedin, New Zealand. From
1999 to 2010 he was a member of the Geography Department at the University of Melbourne in Australia. He holds a PhD and MSc from the University of
Canterbury. His research is focussed on the geomorphology of rock coasts.

Dr. Mark Dickson is a Senior Lecturer in the School of Environment, University of Auckland. He studied Geography at Massey University before completing a
PhD at the University of Wollongong where he studied the geomorphic evolution of hard rock shorelines around Lord Howe Island. Mark worked as a
postdoctoral researcher at the University of Bristol where he modelled the effects of climate change on rapidly eroding soft-rock coasts. He then returned to
New Zealand on a FRST Postdoctoral Fellowship and worked at the National Institute of Water and Atmospheric Research modeling on chronically eroding
gravel shorelines. Mark has been at the University of Auckland since 2008. His current research interests include cliff erosion and shore platform devel-
opment, prograded barrier coasts and chenier plains, gravel coasts, and a range of coastal management issues associated with these landforms.

Alan Trenhaile (BSc, PhD, DSc) grew up in Abertillery, South Wales (UK). He was an undergraduate and graduate
student in the Geography Department of University College Swansea (now Swansea University), a constituent
college of the University of Wales. Upon completion of his doctorate, which was on the wide shore platforms
along the northern coast of the megatidal Severn Estuary, he accepted a position in the University of Windsor,
Ontario, Canada, where he now holds the title University Prof. in the Department of Earth and Environmental
Sciences. He has continued to work on various aspects of rock coasts. In addition to mathematical modeling of
hard and soft rock coastal evolution with changes in relative sea level and climate, his recent work, in the field and
through laboratory experiments, has been concerned with the processes operating on the sloping and horizontal
shore platforms of eastern Canada. Work is also being conducted in several other areas, including near La Paz,
Mexico, in the southern Baja Peninsula, Galicia in northwestern Spain, and on several island archipelagos in the
Atlantic Ocean. In addition to many research papers and several edited volumes, Dr. Trenhaile is the author of a
number of books, including: The Geomorphology of Rock Coasts (Oxford University Press, Oxford, UK);
Coastal Dynamics and Landforms (Oxford University Press, Oxford, UK); The Geomorphology of Canada
(Oxford University Press, Toronto); and Geomorphology: a Canadian Perspective (Oxford University Press,
Toronto), which is now in its fourth edition.

You might also like