You are on page 1of 13

Raised beach

A raised beach, coastal terrace,[1] or perched


coastline is a relatively flat, horizontal or gently inclined
surface of marine origin,[2] mostly an old abrasion
platform which has been lifted out of the sphere of wave
activity (sometimes called "tread"). Thus, it lies above or
under the current sea level, depending on the time of its
formation.[3][4] It is bounded by a steeper ascending
slope on the landward side and a steeper descending
slope on the seaward side[2] (sometimes called "riser").
Due to its generally flat shape, it is often used for
anthropogenic structures such as settlements and
infrastructure.[3]
A set of raised beaches at Kincraig Point in the
East Neuk of Fife, Scotland, near Earlsferry. A raised beach is an emergent coastal landform. Raised
beaches and marine terraces are beaches or wave-cut
platforms raised above the shoreline by a relative fall in
the sea level.[5]

Around the world, a combination of tectonic coastal uplift and


Quaternary sea-level fluctuations has resulted in the formation of
marine terrace sequences, most of which were formed during
separate interglacial highstands that can be correlated to marine
isotope stages (MIS).[6]

A marine terrace commonly retains a shoreline angle or inner edge,


the slope inflection between the marine abrasion platform and the
associated paleo sea-cliff. The shoreline angle represents the A raised beach, now at 4 metres
maximum shoreline of a transgression and therefore a paleo-sea (13 ft) above high tide, formed King's
level. Cave, Arran, below an earlier raised
beach at around 30 metres (98 ft)
height.

Contents
Morphology
Formation
Causes
Processes
Land and sea level history
Mapping and surveying Relict sea-cliffs at King's Cave on
Correlation and dating Arran's south-west coast
Correlational dating
Direct dating
Relevance for other research areas
Prominent examples
Related coastal geography
See also
References
External links

Morphology
The platform of a marine terrace usually has a
gradient between 1°–5° depending on the former
tidal range with, commonly, a linear to concave
profile. The width is quite variable, reaching up to
1,000 metres (3,300  ft), and seems to differ
between the northern and southern
[9]
hemispheres. The cliff faces that delimit the
platform can vary in steepness depending on the
relative roles of marine and subaerial
processes.[10] At the intersection of the former
shore (wave-cut/abrasion-) platform and the rising
cliff face the platform commonly retains a
shoreline angle or inner edge (notch) that Typical sequence of erosional marine terraces. 1) low
indicates the location of the shoreline at the time tide cliff/ramp with deposition, 2) modern shore (wave-
of maximum sea ingression and therefore a paleo- cut/abrasion-) platform, 3) notch/inner edge, modern
sea level.[11] Sub-horizontal platforms usually shoreline angle, 4) modern sea cliff, 5) old shore (wave-
terminate in a low tide cliff, and it is believed that cut/abrasion-) platform, 6) paleo-shoreline angle,
the occurrence of these platforms depends on tidal 7) paleo-sea cliff, 8) terrace cover deposits/marine
activity.[10] Marine terraces can extend for several deposits, colluvium, 9) alluvial fan, 10) decayed and
tens of kilometers parallel to the coast.[3] covered sea cliff and shore platform, 11) paleo-sea
level I, 12) paleo-sea level II. – after various
Older terraces are covered by marine and/or authors[1][3][7][8]
alluvial or colluvial materials while the uppermost
terrace levels usually are less well preserved.[12]
While marine terraces in areas of relatively rapid uplift rates (>  1  mm/year) can often be correlated to
individual interglacial periods or stages, those in areas of slower uplift rates may have a polycyclic origin
with stages of returning sea levels following periods of exposure to weathering.[2]

Marine terraces can be covered by a wide variety of soils with complex histories and different ages. In
protected areas, allochtonous sandy parent materials from tsunami deposits may be found. Common soil
types found on marine terraces include planosols and solonetz.[13]

Formation
It is now widely thought that marine terraces are formed during the separated highstands of interglacial
stages correlated to marine isotope stages (MIS).[14][15][16][17][18]

Causes
The formation of marine terraces is controlled by
changes in environmental conditions and by tectonic
activity during recent geological times. Changes in
climatic conditions have led to eustatic sea-level
oscillations and isostatic movements of the Earth's
crust, especially with the changes between glacial and
interglacial periods.

Processes of eustasy lead to glacioeustatic sea level


fluctuations due to changes of the water volume in the
oceans, and hence to regressions and transgressions of
the shoreline. At times of maximum glacial extent
Comparison of two sea level reconstructions during
during the last glacial period, the sea level was about
the last 500 Ma. The scale of change during the
100 metres (330 ft) lower compared to today. Eustatic
last glacial/interglacial transition is indicated with a
sea level changes can also be caused by changes in the
black bar.
void volume of the oceans, either through sedimento-
eustasy or tectono-eustasy.[19]

Processes of isostasy involve the uplift of continental crusts along with their shorelines. Today, the process
of glacial isostatic adjustment mainly applies to Pleistocene glaciated areas.[19] In Scandinavia, for instance,
the present rate of uplift reaches up to 10 millimetres (0.39 in)/year.[20]

In general, eustatic marine terraces were formed during separate sea level highstands of interglacial
stages[19][21] and can be correlated to marine oxygen isotopic stages (MIS).[22][23] Glacioisostatic marine
terraces were mainly created during stillstands of the isostatic uplift.[19] When eustasy was the main factor
for the formation of marine terraces, derived sea level fluctuations can indicate former climate changes. This
conclusion has to be treated with care, as isostatic adjustments and tectonic activities can be extensively
overcompensated by a eustatic sea level rise. Thus, in areas of both eustatic and isostatic or tectonic
influences, the course of the relative sea level curve can be complicated.[24] Hence, most of today's marine
terrace sequences were formed by a combination of tectonic coastal uplift and Quaternary sea level
fluctuations.

Jerky tectonic uplifts can also lead to marked terrace steps while smooth relative sea level changes may not
result in obvious terraces, and their formations are often not referred to as marine terraces.[11]

Processes

Marine terraces often result from marine erosion along rocky coastlines[2] in temperate regions due to wave
attack and sediment carried in the waves. Erosion also takes place in connection with weathering and
cavitation. The speed of erosion is highly dependent on the shoreline material (hardness of rock[10]), the
bathymetry, and the bedrock properties and can be between only a few millimeters per year for granitic
rocks and more than 10 metres (33  ft) per year for volcanic ejecta.[10][25] The retreat of the sea cliff
generates a shore (wave-cut/abrasion-) platform through the process of abrasion. A relative change of the
sea level leads to regressions or transgressions and eventually forms another terrace (marine-cut terrace) at a
different altitude, while notches in the cliff face indicate short stillstands.[25]

It is believed that the terrace gradient increases with tidal range and decreases with rock resistance. In
addition, the relationship between terrace width and the strength of the rock is inverse, and higher rates of
uplift and subsidence as well as a higher slope of the hinterland increases the number of terraces formed
during a certain time.[26]
Furthermore, shore platforms are formed by denudation and marine-built terraces arise from accumulations
of materials removed by shore erosion.[2] Thus, a marine terrace can be formed by both erosion and
accumulation. However, there is an ongoing debate about the roles of wave erosion and weathering in the
formation of shore platforms.[10]

Reef flats or uplifted coral reefs are another kind of marine terrace found in intertropical regions. They are a
result of biological activity, shoreline advance and accumulation of reef materials.[2]

While a terrace sequence can date back hundreds of thousands of years, its degradation is a rather fast
process. On the one hand a deeper transgression of cliffs into the shoreline may completely destroy
previous terraces; on the other hand older terraces might be decayed[25] or covered by deposits, colluvia or
alluvial fans.[3] Erosion and backwearing of slopes caused by incisive streams play another important role
in this degradation process.[25]

Land and sea level history

The total displacement of the shoreline relative to the age of the associated interglacial stage allows
calculation of a mean uplift rate or the calculation of eustatic level at a particular time if the uplift is known.

In order to estimate vertical uplift, the eustatic position of the considered paleo sea levels relative to the
present one must be known as precisely as possible. Our chronology relies principally on relative dating
based on geomorphologic criteria, but in all cases we associated the shoreline angle of the marine terraces
with numerical ages. The best-represented terrace worldwide is the one correlated to the last interglacial
maximum (MIS 5e).[27][28][29] Age of MISS 5e is arbitrarily fixed to range from 130 to 116 ka[30] but is
demonstrated to range from 134 to 113 ka in Hawaii and Barbados with a peak from 128 to 116 ka on
tectonically stable coastlines. Older marine terraces well represented in worldwide sequences are those
related to MIS 9 (~303–339 ka) and 11 (~362–423 ka).[31] Compilations show that sea level was 3 ± 3
meters higher during MIS 5e, MIS 9 and 11 than during the present one and −1 ± 1 m to the present one
during MIS 7.[32][33] Consequently, MIS 7 (~180-240 ka) marine terraces are less pronounced and
sometimes absent. When the elevations of these terraces are higher than the uncertainties in paleo-eustatic
sea level mentioned for the Holocene and Late Pleistocene, these uncertainties have no effect on overall
interpretation.

Sequence can also occur where the accumulation of ice sheets have depressed the land so that when the ice
sheets melts the land readjusts with time thus raising the height of the beaches (glacio-isostatic rebound) and
in places where co-seismic uplift occur. In the latter case, the terrace are not correlated with sea level
highstand even if co-seismic terrace are known only for the Holocene.

Mapping and surveying


For exact interpretations of the morphology, extensive datings, surveying and mapping of marine terraces is
applied. This includes stereoscopic aerial photographic interpretation (ca. 1 : 10,000 – 25,000[11]), on-site
inspections with topographic maps (ca. 1  : 10,000) and analysis of eroded and accumulated material.
Moreover, the exact altitude can be determined with an aneroid barometer or preferably with a levelling
instrument mounted on a tripod. It should be measured with the accuracy of 1 cm (0.39 in) and at about
every 50–100 metres (160–330  ft), depending on the topography. In remote areas, the techniques of
photogrammetry and tacheometry can be applied.[24]

Correlation and dating


Different methods for dating and correlation of marine terraces can
be used and combined.

Correlational dating

The morphostratigraphic approach focuses especially in regions of


marine regression on the altitude as the most important criterion to
distinguish coastlines of different ages. Moreover, individual marine
Aerial photograph of the lowest
terraces can be correlated based on their size and continuity. Also,
marine terrace at Tongue Point, New
paleo-soils as well as glacial, fluvial, eolian and periglacial Zealand
landforms and sediments may be used to find correlations between
terraces.[24] On New Zealand's North Island, for instance, tephra
and loess were used to date and correlate marine terraces.[34] At the terminus advance of former glaciers
marine terraces can be correlated by their size, as their width decreases with age due to the slowly thawing
glaciers along the coastline.[24]

The lithostratigraphic approach uses typical sequences of sediment and rock strata to prove sea level
fluctuations on the basis of an alternation of terrestrial and marine sediments or littoral and shallow marine
sediments. Those strata show typical layers of transgressive and regressive patterns.[24] However, an
unconformity in the sediment sequence might make this analysis difficult.[35]

The biostratigraphic approach uses remains of organisms which can indicate the age of a marine terrace.
For that, often mollusc shells, foraminifera or pollen are used. Especially Mollusca can show specific
properties depending on their depth of sedimentation. Thus, they can be used to estimate former water
depths.[24]

Marine terraces are often correlated to marine oxygen isotopic stages (MIS)[22] and can also be roughly
dated using their stratigraphic position.[24]

Direct dating

There are various methods for the direct dating of marine terraces and their related materials. The most
common method is 14 C radiocarbon dating,[36] which has been used, for example, on the North Island of
New Zealand to date several marine terraces.[37] It utilizes terrestrial biogenic materials in coastal
sediments, such as mollusc shells, by analyzing the 14 C isotope.[24] In some cases, however, dating based
on the 230 Th/234 U ratio was applied, in case detrital contamination or low uranium concentrations made
finding a high resolution dating difficult.[38] In a study in southern Italy paleomagnetism was used to carry
out paleomagnetic datings[39] and luminescence dating (OSL) was used in different studies on the San
Andreas Fault[40] and on the Quaternary Eupcheon Fault in South Korea.[41] In the last decade, the dating
of marine terraces has been enhanced since the arrival of terrestrial cosmogenic nuclides method, and
particularly through the use of 10 Be and 26 Al cosmogenic isotopes produced on site.[42][43][44] These
isotopes record the duration of surface exposure to cosmic rays.[45] This exposure age reflects the age of
abandonment of a marine terrace by the sea.

In order to calculate the eustatic sea level for each dated terrace, it is assumed that the eustatic sea-level
position corresponding to at least one marine terrace is known and that the uplift rate has remained
essentially constant in each section.[2]

Relevance for other research areas


Marine terraces play an important role in the research on tectonics
and earthquakes. They may show patterns and rates of tectonic
uplift[40][44][46] and thus may be used to estimate the tectonic
activity in a certain region.[41] In some cases the exposed
secondary landforms can be correlated with known seismic events
such as the 1855 Wairarapa earthquake on the Wairarapa Fault near
Wellington, New Zealand which produced a 2.7-metre (8 ft 10 in)
uplift.[47] This figure can be estimated from the vertical offset
between raised shorelines in the area.[48]
Marine terraces south of Choapa
Furthermore, with the knowledge of eustatic sea level fluctuations, River in Chile. These terraces have
the speed of isostatic uplift can be estimated[49] and eventually the been studied among others by
change of relative sea levels for certain regions can be Roland Paskoff.
reconstructed. Thus, marine terraces also provide information for
the research on climate change and trends in future sea level
changes.[10][50]

When analyzing the morphology of marine terraces, it must be considered, that both eustasy and isostasy
can have an influence on the formation process. This way can be assessed, whether there were changes in
sea level or whether tectonic activities took place.

Prominent examples
Raised beaches are found in a wide variety of coast and
geodynamical background such as subduction on the Pacific coasts
of South and North America, passive margin of the Atlantic coast
of South America,[51] collision context on the Pacific coast of
Kamchatka, Papua New Guinea, New Zealand, Japan, passive
margin of the South China Sea coast, on west-facing Atlantic
coasts, such as Donegal Bay, County Cork and County Kerry in
Ireland; Bude, Widemouth Bay, Crackington Haven, Tintagel,
Perranporth and St Ives in Cornwall, the Vale of Glamorgan,
Quaternary marine terraces at
Gower Peninsula, Pembrokeshire and Cardigan Bay in Wales, Jura
Tongue Point, New Zealand
and the Isle of Arran in Scotland, Finistère in Brittany and Galicia
in Northern Spain and at Squally Point in Eatonville, Nova Scotia
within the Cape Chignecto Provincial Park.

Other important sites include various coasts of New Zealand, e.g. Turakirae Head near Wellington being
one of the world's best and most thoroughly studied examples.[47][48][52] Also along the Cook Strait in
New Zealand, there is a well-defined sequence of uplifted marine terraces from the late Quaternary at
Tongue Point. It features a well preserved lower terrace from the last interglacial, a widely eroded higher
terrace from the penultimate interglacial and another still higher terrace, which is nearly completely
decayed.[47] Furthermore, on New Zealand's North Island at the eastern Bay of Plenty, a sequence of seven
marine terraces has been studied.[12][37]

Along many coasts of mainland and islands around the Pacific, marine terraces are typical coastal features.
An especially prominent marine terraced coastline can be found north of Santa Cruz, near Davenport,
California, where terraces probably have been raised by repeated slip earthquakes on the San Andreas
Fault.[40][53] Hans Jenny famously researched the pygmy forests of the Mendocino and Sonoma county
marine terraces. The marine terrace's "ecological staircase" of Salt Point State Park is also bound by the San
Andreas Fault.
Along the coasts of South America marine terraces are
present,[44][54] where the highest ones are situated where plate
margins lie above subducted oceanic ridges and the highest and
most rapid rates of uplift occur.[7][46] At Cape Laundi, Sumba
Island, Indonesia an ancient patch reef can be found at 475  m
(1,558 ft) above sea level as part of a sequence of coral reef terraces
with eleven terraces being wider than 100 m (330 ft).[55] The coral
Air photograph of the marine terraced
coastline north of Santa Cruz,
marine terraces at Huon Peninsula, New Guinea, which extend
California, note Highway 1 running over 80  km (50  mi) and rise over 600  m (2,000  ft) above present
along the coast along the lower sea level[56] are currently on UNESCO’s tentative list for world
terraces heritage sites under the name Houn Terraces - Stairway to the
Past.[57]

Other considerable examples include marine terraces rising up to 360  m (1,180  ft) on some Philippine
Islands[58] and along the Mediterranean Coast of North Africa, especially in Tunisia, rising up to 400  m
(1,300 ft).[59]

Related coastal geography


Uplift can also be registered through tidal notch sequences. Notches are often portrayed as lying at sea
level; however notch types actually form a continuum from wave notches formed in quiet conditions at sea
level to surf notches formed in more turbulent conditions and as much as 2 m (6.6 ft) above sea level.[60]
As stated above, there was at least one higher sea level during the Holocene, so that some notches may not
contain a tectonic component in their formation.

See also
Similar features
Bench (geology)
Fluvial terrace
Strandflat
Terrace (geology)
Beach erosion and accretion
Beach evolution
Beach morphodynamics
Beach nourishment
Modern recession of beaches
Paleoshoreline
Coastal management, to prevent coastal erosion and creation of beach
Coastal and oceanic landforms
Coastal development hazards
Coastal erosion
Coastal geography
Coastal engineering
Coastal and Estuarine Research Federation (CERF)
Erosion
Bioerosion
Blowhole
Natural arch
Wave-cut platform
Longshore drift
Deposition (sediment)
Coastal sediment supply
Sand dune stabilization
Submersion

References
1. Pinter, N (2010): 'Coastal Terraces, Sealevel, and Active Tectonics' (educational exercise),
from "Archived copy" (https://web.archive.org/web/20101010230028/http://www.geology.siu.
edu/people/pinter/pdf/CoastalExercise.pdf) (PDF). Archived from the original (http://www.geo
logy.siu.edu/people/pinter/pdf/CoastalExercise.pdf) (PDF) on 2010-10-10. Retrieved
2011-04-21. [02/04/2011]
2. Pirazzoli, PA (2005a): 'Marine Terraces', in Schwartz, ML (ed) Encyclopedia of Coastal
Science. Springer, Dordrecht, pp. 632–633
3. Strahler AH; Strahler AN (2005): Physische Geographie. Ulmer, Stuttgart, 686 p.
4. Leser, H (ed)(2005): ‚Wörterbuch Allgemeine Geographie. Westermann&Deutscher
Taschenbuch Verlag, Braunschweig, 1119 p.
5. "The Nat -" (http://www.sdnhm.org/research/paleontology/sdshoreline.html).
www.sdnhm.org.
6. Johnson, ME; Libbey, LK (1997). "Global review of Upper Pleistocene (Substage 5e) Rocky
Shores: tectonic segregation, substrate variation and biological diversity". Journal of Coastal
Research.
7. Goy, JL; Macharé, J; Ortlieb, L; Zazo, C (1992). "Quaternary shorelines in Southern Peru: a
Record of Global Sea-level Fluctuations and Tectonic Uplift in Chala Bay". Quaternary
International. 15–16: 9–112. Bibcode:1992QuInt..15...99G (https://ui.adsabs.harvard.edu/ab
s/1992QuInt..15...99G). doi:10.1016/1040-6182(92)90039-5 (https://doi.org/10.1016%2F104
0-6182%2892%2990039-5).
8. Rosenbloom, NA; Anderson, RS (1994). "Hillslope and channel evolution in a marine
terraced landscape, Santa Cruz, California". Journal of Geophysical Research. 99 (B7):
14013–14029. Bibcode:1994JGR....9914013R (https://ui.adsabs.harvard.edu/abs/1994JG
R....9914013R). doi:10.1029/94jb00048 (https://doi.org/10.1029%2F94jb00048).
9. Pethick, J (1984): An Introduction to Coastal Geomorphology. Arnold&Chapman&Hall, New
York, 260p.
10. Masselink, G; Hughes, MG (2003): Introduction to Coastal Processes & Geomorphology.
Arnold&Oxford University Press Inc., London, 354p.
11. Cantalamessa, G; Di Celma, C (2003). "Origin and chronology of Pleistocene marine
terraces of Isla de la Plata and of flat, gently dipping surfaces of the southern coast of Cabo
San Lorenzo (Manabí, Ecuador)". Journal of South American Earth Sciences. 16 (8): 633–
648. Bibcode:2004JSAES..16..633C (https://ui.adsabs.harvard.edu/abs/2004JSAES..16..63
3C). doi:10.1016/j.jsames.2003.12.007 (https://doi.org/10.1016%2Fj.jsames.2003.12.007).
12. Ota, Y; Hull, AG; Berryman, KR (1991). "Coseismic Uplift of Holocene Marine Terraces in the
Pakarae River Area, Eastern North Island, New Zealand". Quaternary Research. 35 (3):
331–346. Bibcode:1991QuRes..35..331O (https://ui.adsabs.harvard.edu/abs/1991QuRes..3
5..331O). doi:10.1016/0033-5894(91)90049-B (https://doi.org/10.1016%2F0033-5894%289
1%2990049-B).
13. Finkl, CW (2005): 'Coastal Soils' in Schwartz, ML (ed) Encyclopedia of Coastal Science.
Springer, Dordrecht, pp. 278–302
14. James, N.P.; Mountjoy, E.W.; Omura, A. (1971). "An early Wisconsin reef Terrace at
Barbados, West Indies, and its climatic implications". Geological Society of America
Bulletin. 82 (7): 2011–2018. Bibcode:1971GSAB...82.2011J (https://ui.adsabs.harvard.edu/a
bs/1971GSAB...82.2011J). doi:10.1130/0016-7606(1971)82[2011:aewrta]2.0.co;2 (https://do
i.org/10.1130%2F0016-7606%281971%2982%5B2011%3Aaewrta%5D2.0.co%3B2).
15. Chappell, J (1974). "Geology of coral terraces, Huon Peninsula, New Guinea: a study of
Quaternary tectonic movements and sea Level changes". Geological Society of America
Bulletin. 85 (4): 553–570. Bibcode:1974GSAB...85..553C (https://ui.adsabs.harvard.edu/abs/
1974GSAB...85..553C). doi:10.1130/0016-7606(1974)85<553:gocthp>2.0.co;2 (https://doi.or
g/10.1130%2F0016-7606%281974%2985%3C553%3Agocthp%3E2.0.co%3B2).
16. Bull, W.B., 1985. Correlation of flights of global marine terraces. In: Morisawa M. & Hack J.
(Editor), 15th Annual Geomorphology Symposium. Hemel Hempstead, State University of
New York at Binghamton, pp. 129–152.
17. Ota, Y (1986). "Marine terraces as reference surfaces in late Quaternary tectonics studies:
examples from the Pacific Rim". Royal Society of New Zealand Bulletin. 24: 357–375.
18. Muhs, D.R.; et al. (1990). "Age Estimates and Uplift Rates for Late Pleistocene Marine
Terraces: Southern Oregon Portion of the Cascadia Forearc" (https://digitalcommons.unl.ed
u/usgsstaffpub/164). Journal of Geophysical Research. 95 (B5): 6685–6688.
Bibcode:1990JGR....95.6685M (https://ui.adsabs.harvard.edu/abs/1990JGR....95.6685M).
doi:10.1029/jb095ib05p06685 (https://doi.org/10.1029%2Fjb095ib05p06685).
19. Ahnert, F (1996) – Einführung in die Geomorphologie. Ulmer, Stuttgart, 440 p.
20. Lehmkuhl, F; Römer, W (2007): 'Formenbildung durch endogene Prozesse: Neotektonik', in
Gebhardt, H; Glaser, R; Radtke, U; Reuber, P (ed) Geographie, Physische Geographie und
Humangeographie. Elsevier, München, pp. 316–320
21. James, NP; Mountjoy, EW; Omura, A (1971). "An Early Wisconsin Reef Terrace at Barbados,
West Indies, and ist Climatic Implications". Geological Society of America Bulletin. 82 (7):
2011–2018. doi:10.1130/0016-7606(1971)82[2011:AEWRTA]2.0.CO;2 (https://doi.org/10.11
30%2F0016-7606%281971%2982%5B2011%3AAEWRTA%5D2.0.CO%3B2).
22. Johnson, ME; Libbey, LK (1997). "Global Review of Upper Pleistocene (Substage 5e) Rocky
Shores: Tectonic Segregation, Substrate Variation, and Biological Diversity". Journal of
Coastal Research. 13 (2): 297–307.
23. Muhs, D; Kelsey, H; Miller, G; Kennedy, G; Whelan, J; McInelly, G (1990). " 'Age Estimates
and Uplift Rates for Late Pleistocene Marine Terraces' Southern Oregon Portion of the
Cascadia Forearc' " (https://digitalcommons.unl.edu/usgsstaffpub/164). Journal of
Geophysical Research. 95 (B5): 6685–6698. Bibcode:1990JGR....95.6685M (https://ui.adsa
bs.harvard.edu/abs/1990JGR....95.6685M). doi:10.1029/jb095ib05p06685 (https://doi.org/10.
1029%2Fjb095ib05p06685).
24. Worsley, P (1998): 'Altersbestimmung – Küstenterrassen', in Goudie, AS (ed)
Geomorphologie, Ein Methodenhandbuch für Studium und Praxis. Springer, Heidelberg, pp.
528–550
25. Anderson, RS; Densmore, AL; Ellis, MA (1999). "The Generation and degradation of Marine
Terraces". Basin Research. 11 (1): 7–19. Bibcode:1999BasR...11....7A (https://ui.adsabs.har
vard.edu/abs/1999BasR...11....7A). doi:10.1046/j.1365-2117.1999.00085.x (https://doi.org/1
0.1046%2Fj.1365-2117.1999.00085.x).
26. Trenhaile, AS (2002). "Modeling the development of marine terraces on tectonically mobile
rock coasts". Marine Geology. 185 (3–4): 341–361. Bibcode:2002MGeol.185..341T (https://u
i.adsabs.harvard.edu/abs/2002MGeol.185..341T). doi:10.1016/S0025-3227(02)00187-1 (htt
ps://doi.org/10.1016%2FS0025-3227%2802%2900187-1).
27. Pedoja, K.; Bourgeois, J.; Pinegina, T.; Higman, B. (2006). "Does Kamchatka belong to
North America? An extruding Okhotsk block suggested by coastal neotectonics of the
Ozernoi Peninsula, Kamchatka, Russia" (http://ir.scsio.ac.cn/handle/344004/4418). Geology.
34 (5): 353–356. Bibcode:2006Geo....34..353P (https://ui.adsabs.harvard.edu/abs/2006Ge
o....34..353P). doi:10.1130/g22062.1 (https://doi.org/10.1130%2Fg22062.1).
28. Pedoja, K.; Dumont, J-F.; Lamothe, M.; Ortlieb, L.; Collot, J-Y.; Ghaleb, B.; Auclair, M.;
Alvarez, V.; Labrousse, B. (2006). "Quaternary uplift of the Manta Peninsula and La Plata
Island and the subduction of the Carnegie Ridge, central coast of Ecuador". South American
Journal of Earth Sciences. 22 (1–2): 1–21. Bibcode:2006JSAES..22....1P (https://ui.adsabs.h
arvard.edu/abs/2006JSAES..22....1P). doi:10.1016/j.jsames.2006.08.003 (https://doi.org/10.
1016%2Fj.jsames.2006.08.003).
29. Pedoja, K.; Ortlieb, L.; Dumont, J-F.; Lamothe, J-F.; Ghaleb, B.; Auclair, M.; Labrousse, B.
(2006). "Quaternary coastal uplift along the Talara Arc (Ecuador, Northern Peru) from new
marine terrace data" (http://ir.scsio.ac.cn/handle/344004/4442). Marine Geology. 228 (1–4):
73–91. Bibcode:2006MGeol.228...73P (https://ui.adsabs.harvard.edu/abs/2006MGeol.228...
73P). doi:10.1016/j.margeo.2006.01.004 (https://doi.org/10.1016%2Fj.margeo.2006.01.004).
30. Kukla, G.J.; et al. (2002). "Last Interglacial Climates" (https://digitalcommons.unl.edu/cgi/vie
wcontent.cgi?article=1173&context=usgsstaffpub). Quaternary Research. 58 (1): 2–13.
Bibcode:2002QuRes..58....2K (https://ui.adsabs.harvard.edu/abs/2002QuRes..58....2K).
doi:10.1006/qres.2001.2316 (https://doi.org/10.1006%2Fqres.2001.2316). S2CID 55262041
(https://api.semanticscholar.org/CorpusID:55262041).
31. Imbrie, J. et al., 1984. The orbital theory of Pleistocene climate: support from revised
chronology of the marine 18O record. In: A. Berger, J. Imbrie, J.D. Hays, G. Kukla and B.
Saltzman (Editors), Milankovitch and Climate. Reidel, Dordrecht, pp. 269–305.
32. Hearty, P.J.; Kindler, P. (1995). "Sea-Level Highstand Chronology from Stable Carbonate
Platforms (Bermuda and the Bahamas)". Journal of Coastal Research. 11 (3): 675–689.
33. Zazo, C (1999). "Interglacial sea levels". Quaternary International. 55 (1): 101–113.
Bibcode:1999QuInt..55..101Z (https://ui.adsabs.harvard.edu/abs/1999QuInt..55..101Z).
doi:10.1016/s1040-6182(98)00031-7 (https://doi.org/10.1016%2Fs1040-6182%2898%2900
031-7).
34. Berryman, K (1992). "A stratigraphic age of Rotoehu Ash and late Pleistocene climate
interpretation based on marine terrace chronology, Mahia Peninsula, North Island, New
Zealand". New Zealand Journal of Geology and Geophysics. 35: 1–7.
doi:10.1080/00288306.1992.9514494
(https://doi.org/10.1080%2F00288306.1992.9514494).
35. Bhattacharya, JP; Sheriff, RE (2011). "Practical problems in the application of the sequence
stratigraphic method and key surfaces: integrating observations from ancient fluvial–deltaic
wedges with Quaternary and modelling studies". Sedimentology. 58 (1): 120–169.
Bibcode:2011Sedim..58..120B (https://ui.adsabs.harvard.edu/abs/2011Sedim..58..120B).
doi:10.1111/j.1365-3091.2010.01205.x (https://doi.org/10.1111%2Fj.1365-3091.2010.01205.
x).
36. Schellmann, G; Brückner, H (2005): 'Geochronology', in Schwartz, ML (ed) Encyclopedia of
Coastal Science. Springer, Dordrecht, pp. 467–472
37. Ota, Y (1992). "Holocene marine terraces on the northeast coast of North Island, New
Zealand, and their tectonic significance". New Zealand Journal of Geology and Geophysics.
35 (3): 273–288. doi:10.1080/00288306.1992.9514521 (https://doi.org/10.1080%2F0028830
6.1992.9514521).
38. Garnett, ER; Gilmour, MA; Rowe, PJ; Andrews, JE; Preece, RC (2003). "230Th/234U dating
of Holocene tufas: possibilities and problems". Quaternary Science Reviews. 23 (7–8): 947–
958. Bibcode:2004QSRv...23..947G (https://ui.adsabs.harvard.edu/abs/2004QSRv...23..947
G). doi:10.1016/j.quascirev.2003.06.018 (https://doi.org/10.1016%2Fj.quascirev.2003.06.01
8).
39. Brückner, H (1980): 'Marine Terrassen in Süditalien. Eine quartärmorphologische Studie
über das Küstentiefland von Metapont', Düsseldorfer Geographische Schriften, 14,
Düsseldorf, Germany: Düsseldorf University
40. Grove, K; Sklar, LS; Scherer, AM; Lee, G; Davis, J (2010). "Accelerating and spatially
varying crustal uplift and ist geomorphic expression, San Andreas Fault zone north of San
Francisco, California". Tectonophysics. 495 (3): 256–268. Bibcode:2010Tectp.495..256G (htt
ps://ui.adsabs.harvard.edu/abs/2010Tectp.495..256G). doi:10.1016/j.tecto.2010.09.034 (http
s://doi.org/10.1016%2Fj.tecto.2010.09.034).
41. Kim, Y; Kihm, J; Jin, K (2011). "Interpretation of the rupture history of a low slip-rate active
fault by analysis of progressive displacement accumulation: an example from the Quaternary
Eupcheon Fault, SE Korea". Journal of the Geological Society, London. 168 (1): 273–288.
Bibcode:2011JGSoc.168..273K (https://ui.adsabs.harvard.edu/abs/2011JGSoc.168..273K).
doi:10.1144/0016-76492010-088 (https://doi.org/10.1144%2F0016-76492010-088).
S2CID 129506275 (https://api.semanticscholar.org/CorpusID:129506275).
42. Perg, LA; Anderson, RS; Finkel, RC (2001). "Use of a new 10Be and 26Al inventory method
to date marine terraces, Santa Cruz, California, USA". Geology. 29 (10): 879–882.
Bibcode:2001Geo....29..879P (https://ui.adsabs.harvard.edu/abs/2001Geo....29..879P).
doi:10.1130/0091-7613(2001)029<0879:uoanba>2.0.co;2 (https://doi.org/10.1130%2F0091-
7613%282001%29029%3C0879%3Auoanba%3E2.0.co%3B2).
43. Kim, KJ; Sutherland, R (2004). "Uplift rate and landscape development in southwest
Fiordland, New Zealand, determined using 10Be and 26Al exposure dating of marine
terraces". Geochimica et Cosmochimica Acta. 68 (10): 2313–2319.
Bibcode:2004GeCoA..68.2313K (https://ui.adsabs.harvard.edu/abs/2004GeCoA..68.2313K).
doi:10.1016/j.gca.2003.11.005 (https://doi.org/10.1016%2Fj.gca.2003.11.005).
44. Saillard, M; Hall, SR; Audin, L; Farber, DL; Hérail, G; Martinod, J; Regard, V; Finkel, RC;
Bondoux, F (2009). "Non-steady long-term uplift rates and Pleistocene marine terrace
development along the Andean margin of Chile (31°S) inferred from 10Be dating". Earth and
Planetary Science Letters. 277 (1–2): 50–63. Bibcode:2009E&PSL.277...50S (https://ui.adsa
bs.harvard.edu/abs/2009E&PSL.277...50S). doi:10.1016/j.epsl.2008.09.039 (https://doi.org/1
0.1016%2Fj.epsl.2008.09.039).
45. Gosse, JC; Phillips, FM (2001). "Terrestrial in situ cosmogenic nuclides: theory and
application". Quaternary Science Reviews. 20 (14): 1475–1560.
Bibcode:2001QSRv...20.1475G (https://ui.adsabs.harvard.edu/abs/2001QSRv...20.1475G).
CiteSeerX 10.1.1.298.3324 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.298.
3324). doi:10.1016/s0277-3791(00)00171-2 (https://doi.org/10.1016%2Fs0277-3791%280
0%2900171-2).
46. Saillard, M; Hall, SR; Audin, L; Farber, DL; Regard, V; Hérail, G (2011). "Andean coastal
uplift and active tectonics in southern Peru: 10Be surface exposure dating of differentially
uplifted marine terrace sequences (San Juan de Marcona, ~15.4°S)". Geomorphology. 128
(3): 178–190. Bibcode:2011Geomo.128..178S (https://ui.adsabs.harvard.edu/abs/2011Geo
mo.128..178S). doi:10.1016/j.geomorph.2011.01.004 (https://doi.org/10.1016%2Fj.geomorp
h.2011.01.004).
47. Crozier, MJ; Preston NJ (2010): 'Wellington's Tectonic Landscape: Astride a Plate Boundary'
in Migoń, P. (ed) Geomorphological Landscapes of the World. Springer, New York, pp. 341–
348
48. McSaveney; et al. (2006). "Late Holocene uplift of beach ridges at Turakirae Head, south
Wellington coast, New Zealand". New Zealand Journal of Geology & Geophysics. 49 (3):
337–358. doi:10.1080/00288306.2006.9515172 (https://doi.org/10.1080%2F00288306.200
6.9515172). S2CID 129074978 (https://api.semanticscholar.org/CorpusID:129074978).
49. Press, F; Siever, R (2008): Allgemeine Geologie. Spektrum&Springer, Heidelberg, 735 p.
50. Schellmann, G; Radtke, U (2007). "Neue Befunde zur Verbreitung und
chronostratigraphischen Gliederung holozäner Küstenterrassen an der mittel- und
südpatagonischen Atlantikküste (Argentinien) – Zeugnisse holozäner
Meeresspiegelveränderungen". Bamberger Geographische Schriften. 22: 1–91.
51. Rostami, K.; Peltier, W.R.; Mangini, A. (2000). "Quaternary marine terraces, sea-level
changes and uplift history of Patagonia, Argentina: comparisons with predictions of the ICE-
4G (VM2) model for the global process of glacial isostatic adjustment". Quaternary Science
Reviews. 19 (14–15): 1495–1525. Bibcode:2000QSRv...19.1495R (https://ui.adsabs.harvar
d.edu/abs/2000QSRv...19.1495R). doi:10.1016/s0277-3791(00)00075-5 (https://doi.org/10.1
016%2Fs0277-3791%2800%2900075-5).
52. Wellman, HW (1969). "Tilted Marine Beach Ridges at Cape Turakirae, N.Z.". Tuatara. 17 (2):
82–86.
53. Pirazzoli, PA (2005b.): 'Tectonics and Neotectonics', Schwartz, ML (ed) Encyclopedia of
Coastal Science. Springer, Dordrecht, pp. 941–948
54. Saillard, M; Riotte, J; Regard, V; Violette, A; Hérail, G; Audin, A; Riquelme, R (2012). "Beach
ridges U-Th dating in Tongoy bay and tectonic implications for a peninsula-bay system,
Chile". Journal of South American Earth Sciences. 40: 77–84.
Bibcode:2012JSAES..40...77S (https://ui.adsabs.harvard.edu/abs/2012JSAES..40...77S).
doi:10.1016/j.jsames.2012.09.001 (https://doi.org/10.1016%2Fj.jsames.2012.09.001).
55. Pirazzoli, PA; Radtke, U; Hantoro, WS; Jouannic, C; Hoang, CT; Causse, C; Borel Best, M
(1991). "Quaternary Raised Coral-Reef Terraces on Sumba Island, Indonesia". Science. 252
(5014): 1834–1836. Bibcode:1991Sci...252.1834P (https://ui.adsabs.harvard.edu/abs/1991S
ci...252.1834P). doi:10.1126/science.252.5014.1834 (https://doi.org/10.1126%2Fscience.25
2.5014.1834). PMID 17753260 (https://pubmed.ncbi.nlm.nih.gov/17753260).
S2CID 36558992 (https://api.semanticscholar.org/CorpusID:36558992).
56. Chappell, J (1974). "Geology of Coral Terraces, Huon Peninsula, New Guinea: A Study of
Quaternary Tectonic Movements and Se-Level Changes". Geological Society of America
Bulletin. 85 (4): 553–570. Bibcode:1974GSAB...85..553C (https://ui.adsabs.harvard.edu/abs/
1974GSAB...85..553C). doi:10.1130/0016-7606(1974)85<553:gocthp>2.0.co;2 (https://doi.or
g/10.1130%2F0016-7606%281974%2985%3C553%3Agocthp%3E2.0.co%3B2).
57. UNESCO (2006): Huon Terraces – Stairway to the Past. from
https://whc.unesco.org/en/tentativelists/5066/ [13/04/2011]
58. Eisma, D (2005): 'Asia, eastern, Coastal Geomorphology', in Schwartz, ML (ed)
Encyclopedia of Coastal Science. Springer, Dordrecht, pp. 67–71
59. Orme, AR (2005): 'Africa, Coastal Geomorphology', in Schwartz, ML (ed) Encyclopedia of
Coastal Science. Springer, Dordrecht, pp. 9–21
60. Rust, D.; Kershaw, S. (2000). "Holocene tectonic uplift patternes in northeastern Sicily:
evidence from marine notches in coastal outcrops". Marine Geology. 167 (1–2): 105–126.
Bibcode:2000MGeol.167..105R (https://ui.adsabs.harvard.edu/abs/2000MGeol.167..105R).
doi:10.1016/s0025-3227(00)00019-0 (https://doi.org/10.1016%2Fs0025-3227%2800%2900
019-0).

External links
Notes (https://web.archive.org/web/20030320144646/http://www.nahste.ac.uk/subj/r/1973/)
at NAHSTE
US Geological Survey Marine Terrace Fact Sheet - Wikimedia link (https://commons.wikime
dia.org/wiki/File:USGS_Marine_Terraces_of_California_Fact_Sheet.pdf), USGS link (https://
pubs.usgs.gov/fs/2018/3002/fs20183002.pdf)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Raised_beach&oldid=1099160501"

This page was last edited on 19 July 2022, at 08:53 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License 3.0;


additional terms may apply. By
using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the
Wikimedia Foundation, Inc., a non-profit organization.

You might also like