Raised beach

Last updated
Raised beach and marine terraces at Water Canyon beach Santa-Rosa-Island-Water-Canyon-Beach.jpg
Raised beach and marine terraces at Water Canyon beach
A raised beach, now at 4 metres (13 ft) above high tide, formed King's Cave, Arran, below an earlier raised beach at around 30 metres (98 ft) height. Path from The Doon to King's Cave, Arran.jpg
A raised beach, now at 4 metres (13 ft) above high tide, formed King's Cave, Arran, below an earlier raised beach at around 30 metres (98 ft) height.

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]

Contents

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]

Relict sea-cliffs at King's Cave on Arran's south-west coast King's Cave, Isle of Arran.jpg
Relict sea-cliffs at King's Cave on Arran's south-west coast

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 maximum shoreline of a transgression and therefore a paleo-sea level.

Morphology

Typical sequence of erosional marine terraces. 1) low tide cliff/ramp with deposition, 2) modern shore (wave-cut/abrasion-) platform, 3) notch/inner edge, modern shoreline angle, 4) modern sea cliff, 5) old shore (wave-cut/abrasion-) platform, 6) paleo-shoreline angle, 7) paleo-sea cliff, 8) terrace cover deposits/marine deposits, colluvium, 9) alluvial fan, 10) decayed and covered sea cliff and shore platform, 11) paleo-sea level I, 12) paleo-sea level II. - after various authors Marine Terrace diagram(plain svg).svg
Typical sequence of erosional marine terraces. 1) low tide cliff/ramp with deposition, 2) modern shore (wave-cut/abrasion-) platform, 3) notch/inner edge, modern shoreline angle, 4) modern sea cliff, 5) old  shore (wave-cut/abrasion-) platform, 6) paleo-shoreline angle, 7) paleo-sea cliff, 8) terrace cover deposits/marine deposits, colluvium, 9)  alluvial fan, 10) decayed and covered sea cliff and shore platform, 11) paleo-sea level I, 12) paleo-sea level II.  after various authors

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 hemispheres. [9] 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 indicates the location of the shoreline at the time of maximum sea ingression and therefore a paleo-sea level. [11] Sub-horizontal platforms usually terminate in a low tide cliff, and it is believed that the occurrence of these platforms depends on tidal activity. [10] Marine terraces can extend for several tens of kilometers parallel to the coast. [3]

Older terraces are covered by marine and/or 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, allochthonous 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

Comparison of two sea level reconstructions during the last 500 Ma. The scale of change during the last glacial/interglacial transition is indicated with a black bar. Phanerozoic Sea Level.png
Comparison of two sea level reconstructions during the last 500 Ma. The scale of change during the last glacial/interglacial transition is indicated with a black bar.

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 during the last glacial period, the sea level was about 100 metres (330 ft) lower compared to today. Eustatic sea level changes can also be caused by changes in the 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. A deeper transgression of cliffs into the shoreline may completely destroy previous terraces; but 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. Current chronology relies principally on relative dating based on geomorphologic criteria, but in all cases the shoreline angle of the marine terraces is associated 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

Aerial photograph of the lowest marine terrace at Tongue Point, New Zealand Coast south of Wellington by Phillip Capper Flickr.jpg
Aerial photograph of the lowest marine terrace at Tongue Point, New Zealand

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 terraces can be correlated based on their size and continuity. Also, paleo-soils as well as glacial, fluvial, eolian and periglacial 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 14C 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 14C isotope. [24] In some cases, however, dating based on the 230Th/234U 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 10Be and 26Al 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 south of Choapa River in Chile. These terraces have been studied among others by Roland Paskoff. Terraza marina choapa.jpg
Marine terraces south of Choapa River in Chile. These terraces have been studied among others by Roland Paskoff.

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]

Furthermore, with the knowledge of eustatic sea level fluctuations, the speed of isostatic uplift can be estimated [49] and eventually the change of relative sea levels for certain regions can be 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

Quaternary marine terraces at Tongue Point, New Zealand Tongue-Point-by-John-Steedman-Flickr-edited.jpg
Quaternary marine terraces at Tongue Point, New Zealand

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, Gower Peninsula, Pembrokeshire and Cardigan Bay in Wales, Jura 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]

Air photograph of the marine terraced coastline north of Santa Cruz, California, note Highway 1 running along the coast along the lower terraces Marine terraces California.jpg
Air photograph of the marine terraced coastline north of Santa Cruz, California, note Highway 1 running along the coast along the lower terraces

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 marine terraces at Huon Peninsula, New Guinea, which extend over 80 km (50 mi) and rise over 600 m (2,000 ft) above present sea level [56] are currently on UNESCO's tentative list for world 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]

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

Related Research Articles

<span class="mw-page-title-main">Wave-cut platform</span> Narrow flat area created by erosion

A wave-cut platform, shore platform, coastal bench, or wave-cut cliff is the narrow flat area often found at the base of a sea cliff or along the shoreline of a lake, bay, or sea that was created by erosion. Wave-cut platforms are often most obvious at low tide when they become visible as huge areas of flat rock. Sometimes the landward side of the platform is covered by sand, forming the beach, and then the platform can only be identified at low tides or when storms move the sand.

<span class="mw-page-title-main">Last Glacial Period</span> Period of major glaciations of the Northern Hemisphere (115,000–12,000 years ago)

The Last Glacial Period (LGP), also known colloquially as the Last Ice Age or simply Ice Age, occurred from the end of the Last Interglacial to the end of the Younger Dryas, encompassing the period c. 115,000 – c. 11,700 years ago.

<span class="mw-page-title-main">Post-glacial rebound</span> Rise of land masses after glacial period

Post-glacial rebound is the rise of land masses after the removal of the huge weight of ice sheets during the last glacial period, which had caused isostatic depression. Post-glacial rebound and isostatic depression are phases of glacial isostasy, the deformation of the Earth's crust in response to changes in ice mass distribution. The direct raising effects of post-glacial rebound are readily apparent in parts of Northern Eurasia, Northern America, Patagonia, and Antarctica. However, through the processes of ocean siphoning and continental levering, the effects of post-glacial rebound on sea level are felt globally far from the locations of current and former ice sheets.

<span class="mw-page-title-main">Base level</span> Lowest limit for erosion processes

In geology and geomorphology a base level is the lower limit for an erosion process. The modern term was introduced by John Wesley Powell in 1875. The term was subsequently appropriated by William Morris Davis who used it in his cycle of erosion theory. The "ultimate base level" is the surface that results from projection of the sea level under landmasses. It is to this base level that topography tends to approach due to erosion, eventually forming a peneplain close to the end of a cycle of erosion.

<span class="mw-page-title-main">Los Angeles Basin</span> Sedimentary basin located along the coast of southern California

The Los Angeles Basin is a sedimentary basin located in Southern California, in a region known as the Peninsular Ranges. The basin is also connected to an anomalous group of east-west trending chains of mountains collectively known as the Transverse Ranges. The present basin is a coastal lowland area, whose floor is marked by elongate low ridges and groups of hills that is located on the edge of the Pacific Plate. The Los Angeles Basin, along with the Santa Barbara Channel, the Ventura Basin, the San Fernando Valley, and the San Gabriel Basin, lies within the greater Southern California region. The majority of the jurisdictional land area of the city of Los Angeles physically lies within this basin.

<span class="mw-page-title-main">Timeline of glaciation</span> Chronology of the major ice ages of the Earth

There have been five or six major ice ages in the history of Earth over the past 3 billion years. The Late Cenozoic Ice Age began 34 million years ago, its latest phase being the Quaternary glaciation, in progress since 2.58 million years ago.

<span class="mw-page-title-main">Kaikōura Peninsula</span>

The Kaikōura Peninsula is located in the northeast of New Zealand's South Island. It protrudes 5 kilometres (3.1 mi) into the Pacific Ocean. The town of Kaikōura is located on the north shore of the peninsula. The peninsula has been settled by Maori for approximately 1000 years, and by Europeans since the 1800s, when whaling operations began off the Kaikōura Coast. Since the end of whaling in 1922 whales have been allowed to thrive and the region is now a popular whale watching destination.

The Hoxnian Stage was a middle Pleistocene stage of the geological history of the British Isles. It was an interglacial which preceded the Wolstonian Stage and followed the Anglian Stage. It is equivalent to Marine Isotope Stage 11. Marine Isotope Stage 11 started 424,000 years ago and ended 374,000 years ago. The Hoxnian is divided into sub-stages Ho I to Ho IV.

Sequence stratigraphy is a branch of geology, specifically a branch of stratigraphy, that attempts to discern and understand historic geology through time by subdividing and linking sedimentary deposits into unconformity bounded units on a variety of scales. The essence of the method is mapping of strata based on identification of surfaces which are assumed to represent time lines, thereby placing stratigraphy in chronostratigraphic framework allowing understanding of the evolution of the Earth's surface in a particular region through time. Sequence stratigraphy is a useful alternative to a purely lithostratigraphic approach, which emphasizes solely based on the compositional similarity of the lithology of rock units rather than time significance. Unconformities are particularly important in understanding geologic history because they represent erosional surfaces where there is a clear gap in the record. Conversely within a sequence the geologic record should be relatively continuous and complete record that is genetically related.

<span class="mw-page-title-main">Terrace (geology)</span> A step-like landform

In geology, a terrace is a step-like landform. A terrace consists of a flat or gently sloping geomorphic surface, called a tread, that is typically bounded on one side by a steeper ascending slope, which is called a "riser" or "scarp". The tread and the steeper descending slope together constitute the terrace. Terraces can also consist of a tread bounded on all sides by a descending riser or scarp. A narrow terrace is often called a bench.

<span class="mw-page-title-main">Quaternary glaciation</span> Series of alternating glacial and interglacial periods

The Quaternary glaciation, also known as the Pleistocene glaciation, is an alternating series of glacial and interglacial periods during the Quaternary period that began 2.58 Ma and is ongoing. Although geologists describe this entire period up to the present as an "ice age", in popular culture this term usually refers to the most recent glacial period, or to the Pleistocene epoch in general. Since Earth still has polar ice sheets, geologists consider the Quaternary glaciation to be ongoing, though currently in an interglacial period.

<span class="mw-page-title-main">Marine Isotope Stage 11</span> Marine isotope stage between 424,000 and 374,000 years ago

Marine Isotope Stage 11 or MIS 11 is a Marine Isotope Stage in the geologic temperature record, covering the interglacial period between 424,000 and 374,000 years ago. It corresponds to the Hoxnian Stage in Britain.

<span class="mw-page-title-main">Fossil beach</span>

A fossil beach, also known as a paleo-beach, fossil strandline or raised beach, is an ancient oceanic or lacustrine beach preserved in fossil form due to a change in water level or sea level, or because of a shift in terrain elevation. It is often present as a sediment layer or terrace, with beach-related fossils and features, above the present shoreline.

<span class="mw-page-title-main">Erosion and tectonics</span> Interactions between erosion and tectonics and their implications

The interaction between erosion and tectonics has been a topic of debate since the early 1990s. While the tectonic effects on surface processes such as erosion have long been recognized, the opposite has only recently been addressed. The primary questions surrounding this topic are what types of interactions exist between erosion and tectonics and what are the implications of these interactions. While this is still a matter of debate, one thing is clear, Earth's landscape is a product of two factors: tectonics, which can create topography and maintain relief through surface and rock uplift, and climate, which mediates the erosional processes that wear away upland areas over time. The interaction of these processes can form, modify, or destroy geomorphic features on Earth's surface.

<span class="mw-page-title-main">Weichselian glaciation</span> Last glacial period and its associated glaciation in northern parts of Europe

The Weichselian glaciation was the last glacial period and its associated glaciation in northern parts of Europe. In the Alpine region it corresponds to the Würm glaciation. It was characterized by a large ice sheet that spread out from the Scandinavian Mountains and extended as far as the east coast of Schleswig-Holstein, northern Poland and Northwest Russia. This glaciation is also known as the Weichselian ice age, Vistulian glaciation, Weichsel or, less commonly, the Weichsel glaciation, Weichselian cold period (Weichsel-Kaltzeit), Weichselian glacial (Weichsel-Glazial), Weichselian Stage or, rarely, the Weichselian complex (Weichsel-Komplex).

A raised shoreline is an ancient shoreline exposed above current water level. These landforms are formed by a relative change in sea level due to global sea level rise, isostatic rebound, and/or tectonic uplift. These surfaces are usually exposed above modern sea level when a heavily glaciated area experiences a glacial retreat, causing water levels to rise. This area will then experience post-glacial rebound, effectively raising the shoreline surface.

<span class="mw-page-title-main">Maureen Raymo</span> American paleoclimatologist and marine geologist

Maureen E. "Mo" Raymo is an American paleoclimatologist and marine geologist. She is the Co-Founding Dean of the Columbia Climate School, Director of the Lamont–Doherty Earth Observatory of Columbia University, the G. Unger Vetlesen Professor of Earth & Environmental Sciences, and Director of the Lamont–Doherty Core Repository at the Lamont–Doherty Earth Observatory of Columbia University. She is the first female climate scientist and first female scientist to head the institution.

<span class="mw-page-title-main">River terraces (tectonic–climatic interaction)</span>

Terraces can be formed in many ways and in several geologic and environmental settings. By studying the size, shape, and age of terraces, one can determine the geologic processes that formed them. When terraces have the same age and/or shape over a region, it is often indicative that a large-scale geologic or environmental mechanism is responsible. Tectonic uplift and climate change are viewed as dominant mechanisms that can shape the earth’s surface through erosion. River terraces can be influenced by one or both of these forcing mechanisms and therefore can be used to study variation in tectonics, climate, and erosion, and how these processes interact.

<span class="mw-page-title-main">Marine Isotope Stage 5</span> Marine isotope stage in the geologic temperature record, between 130,000 and 80,000 years ago

Marine Isotope Stage 5 or MIS 5 is a marine isotope stage in the geologic temperature record, between 130,000 and 80,000 years ago. Sub-stage MIS 5e corresponds to the Last Interglacial, also called the Eemian or Sangamonian, the last major interglacial period before the Holocene, which extends to the present day. Interglacial periods which occurred during the Pleistocene are investigated to better understand present and future climate variability. Thus, the present interglacial, the Holocene, is compared with MIS 5 or the interglacials of Marine Isotope Stage 11.

The geology of the Baltic Sea is characterized by having areas located both at the Baltic Shield of the East European Craton and in the Danish-North German-Polish Caledonides. Historical geologists make a distinction between the current Baltic Sea depression, formed in the Cenozoic era, and the much older sedimentary basins whose sediments are preserved in the zone. Although glacial erosion has contributed to shape the present depression, the Baltic trough is largely a depression of tectonic origin that existed long before the Quaternary glaciation.

References

  1. 1 2 Pinter, N (2010): 'Coastal Terraces, Sealevel, and Active Tectonics' (educational exercise), from "Archived copy" (PDF). Archived from the original (PDF) on 2010-10-10. Retrieved 2011-04-21.{{cite web}}: CS1 maint: archived copy as title (link) [02/04/2011]
  2. 1 2 3 4 5 6 7 Pirazzoli, PA (2005a): 'Marine Terraces', in Schwartz, ML (ed) Encyclopedia of Coastal Science. Springer, Dordrecht, pp. 632–633
  3. 1 2 3 4 5 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 -". 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. 1 2 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. doi:10.1016/1040-6182(92)90039-5.
  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. doi:10.1029/94jb00048.
  9. Pethick, J (1984): An Introduction to Coastal Geomorphology. Arnold&Chapman&Hall, New York, 260p.
  10. 1 2 3 4 5 6 Masselink, G; Hughes, MG (2003): Introduction to Coastal Processes & Geomorphology. Arnold&Oxford University Press Inc., London, 354p.
  11. 1 2 3 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. doi:10.1016/j.jsames.2003.12.007.
  12. 1 2 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. doi:10.1016/0033-5894(91)90049-B. S2CID   129630764.
  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. doi:10.1130/0016-7606(1971)82[2011:aewrta]2.0.co;2.
  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. doi:10.1130/0016-7606(1974)85<553:gocthp>2.0.co;2.
  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". Journal of Geophysical Research. 95 (B5): 6685–6688. Bibcode:1990JGR....95.6685M. doi:10.1029/jb095ib05p06685.
  19. 1 2 3 4 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. Bibcode:1971GSAB...82.2011J. doi:10.1130/0016-7606(1971)82[2011:AEWRTA]2.0.CO;2.
  22. 1 2 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'". Journal of Geophysical Research. 95 (B5): 6685–6698. Bibcode:1990JGR....95.6685M. doi:10.1029/jb095ib05p06685.
  24. 1 2 3 4 5 6 7 8 Worsley, P (1998): 'Altersbestimmung – Küstenterrassen', in Goudie, AS (ed) Geomorphologie, Ein Methodenhandbuch für Studium und Praxis. Springer, Heidelberg, pp. 528–550
  25. 1 2 3 4 Anderson, RS; Densmore, AL; Ellis, MA (1999). "The Generation and degradation of Marine Terraces". Basin Research. 11 (1): 7–19. Bibcode:1999BasR...11....7A. doi:10.1046/j.1365-2117.1999.00085.x. S2CID   19075109.
  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. doi:10.1016/S0025-3227(02)00187-1.
  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". Geology. 34 (5): 353–356. Bibcode:2006Geo....34..353P. doi:10.1130/g22062.1.
  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. doi:10.1016/j.jsames.2006.08.003. S2CID   59487926.
  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". Marine Geology. 228 (1–4): 73–91. Bibcode:2006MGeol.228...73P. doi:10.1016/j.margeo.2006.01.004. S2CID   129024575.
  30. Kukla, G.J.; et al. (2002). "Last Interglacial Climates". Quaternary Research. 58 (1): 2–13. Bibcode:2002QuRes..58....2K. doi:10.1006/qres.2001.2316. S2CID   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. doi:10.1016/s1040-6182(98)00031-7.
  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): 1–7. Bibcode:1992NZJGG..35....1B. doi: 10.1080/00288306.1992.9514494 .
  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. doi:10.1111/j.1365-3091.2010.01205.x. S2CID   128395986.
  36. Schellmann, G; Brückner, H (2005): 'Geochronology', in Schwartz, ML (ed) Encyclopedia of Coastal Science. Springer, Dordrecht, pp. 467–472
  37. 1 2 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. Bibcode:1992NZJGG..35..273O. doi: 10.1080/00288306.1992.9514521 .
  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. doi:10.1016/j.quascirev.2003.06.018.
  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. 1 2 3 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. doi:10.1016/j.tecto.2010.09.034.
  41. 1 2 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. doi:10.1144/0016-76492010-088. S2CID   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. doi:10.1130/0091-7613(2001)029<0879:uoanba>2.0.co;2.
  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. doi:10.1016/j.gca.2003.11.005.
  44. 1 2 3 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. doi:10.1016/j.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. CiteSeerX   10.1.1.298.3324 . doi:10.1016/s0277-3791(00)00171-2.
  46. 1 2 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. doi:10.1016/j.geomorph.2011.01.004.
  47. 1 2 3 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. 1 2 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. Bibcode:2006NZJGG..49..337M. doi: 10.1080/00288306.2006.9515172 . S2CID   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. doi:10.1016/s0277-3791(00)00075-5.
  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. doi:10.1016/j.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. doi:10.1126/science.252.5014.1834. PMID   17753260. S2CID   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. doi:10.1130/0016-7606(1974)85<553:gocthp>2.0.co;2.
  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. doi:10.1016/s0025-3227(00)00019-0.