Denudation

Last updated

Denudation is the geological process in which moving water, ice, wind, and waves erode the Earth's surface, leading to a reduction in elevation and in relief of landforms and landscapes. Although the terms erosion and denudation are used interchangeably, erosion is the transport of soil and rocks from one location to another, [1] and denudation is the sum of processes, including erosion, that result in the lowering of Earth's surface. [2] Endogenous processes such as volcanoes, earthquakes, and tectonic uplift can expose continental crust to the exogenous processes of weathering, erosion, and mass wasting. The effects of denudation have been recorded for millennia but the mechanics behind it have been debated for the past 200 years[ when? ] and have only begun to be understood in the past few decades. [3] [ when? ]

Contents

Description

Denudation incorporates the mechanical, biological, and chemical processes of erosion, weathering, and mass wasting. Denudation can involve the removal of both solid particles and dissolved material. These include sub-processes of cryofracture, insolation weathering, slaking, salt weathering, bioturbation, and anthropogenic impacts. [4]

Factors affecting denudation include:

Historical theories

Charles Lyell, author of Principles of Geology, who established within the scientific community the concept of denudation and that idea that the surface of the Earth is shaped by gradual processes. Charles Lyell00.jpg
Charles Lyell, author of Principles of Geology, who established within the scientific community the concept of denudation and that idea that the surface of the Earth is shaped by gradual processes.

The effects of denudation have been written about since antiquity, although the terms "denudation" and "erosion" have been used interchangeably throughout most of history. [3] In the Age of Enlightenment, scholars began trying to understand how denudation and erosion occurred without mythical or biblical explanations. Throughout the 18th century, scientists theorized valleys are formed by streams running through them, not from floods or other cataclysms. [9] In 1785, Scottish physician James Hutton proposed an Earth history based on observable processes over an unlimited amount of time, [10] which marked a shift from assumptions based on faith to reasoning based on logic and observation. In 1802, John Playfair, a friend of Hutton, published a paper clarifying Hutton's ideas, explaining the basic process of water wearing down the Earth's surface, and describing erosion and chemical weathering. [11] Between 1830 and 1833, Charles Lyell published three volumes of Principles of Geology , which describes the shaping of the surface of Earth by ongoing processes, and which endorsed and established gradual denudation in the wider scientific community. [12]

W.M. Davis, the man who proposed the peneplanation cycle. William Morris Davis.jpg
W.M. Davis, the man who proposed the peneplanation cycle.

As denudation came into the wider conscience, questions of how denudation occurs and what the result is began arising. Hutton and Playfair suggested over a period of time, a landscape would eventually be worn down to erosional planes at or near sea level, which gave the theory the name "planation". [9] Charles Lyell proposed marine planation, oceans, and ancient shallow seas were the primary driving force behind denudation. While surprising given the centuries of observation of fluvial and pluvial erosion, this is more understandable given early geomorphology was largely developed in Britain, where the effects of coastal erosion are more evident and play a larger role in geomorphic processes. [9] There was more evidence against marine planation than there was for it. By the 1860s, marine planation had largely fallen from favor, a move led by Andrew Ramsay, a former proponent of marine planation who recognized rain and rivers play a more important role in denudation. In North America during the mid-19th century, advancements in identifying fluvial, pluvial, and glacial erosion were made. The work being done in the Appalachians and American West that formed the basis for William Morris Davis to hypothesize peneplanation, despite the fact while peneplanation was compatible in the Appalachians, it did not work as well in the more active American West. Peneplanation was a cycle in which young landscapes are produced by uplift and denuded down to sea level, which is the base level. The process would be restarted when the old landscape was uplifted again or when the base level was lowered, producing a new, young landscape. [13]

Publication of the Davisian cycle of erosion caused many geologists to begin looking for evidence of planation around the world. Unsatisfied with Davis's cycle due to evidence from the Western United States, Grove Karl Gilbert suggested backwearing of slopes would shape landscapes into pediplains, [14] and W.J. McGee named these landscapes pediments. This later gave the concept the name pediplanation when L.C. King applied it on a global scale. [15] The dominance of the Davisian cycle gave rise to several theories to explain planation, such as eolation and glacial planation, although only etchplanation survived time and scrutiny because it was based on observations and measurements done in different climates around the world and it also explained irregularities in landscapes. [16] The majority of these concepts failed, partly because Joseph Jukes, a popular geologist and professor, separated denudation and uplift in an 1862 publication that had a lasting impact on geomorphology. [17] These concepts also failed because the cycles, Davis's in particular, were generalizations and based on broad observations of the landscape rather than detailed measurements; many of the concepts were developed based on local or specific processes, not regional processes, and they assumed long periods of continental stability. [9]

Some scientists opposed the Davisian cycle; one was Grove Karl Gilbert, who, based on measurements over time, realized denudation is nonlinear; he started developing theories based on fluid dynamics and equilibrium concepts. Another was Walther Penck, who devised a more complex theory that denudation and uplift occurred at the same time, and that landscape formation is based on the ratio between denudation and uplift rates. His theory proposed geomorphology is based on endogenous and exogenous processes. [18] Penck's theory, while ultimately being ignored, returned to denudation and uplift occurring simultaneously and relying on continental mobility, even though Penck rejected continental drift. The Davisian and Penckian models were heavily debated for a few decades until Penck's was ignored and support for Davis's waned after his death as more critiques were made. One critic was John Leighly, who stated geologists did not know how landforms were developed, so Davis's theory was built upon a shaky foundation. [19]

From 1945 to 1965, a change in geomorphology research saw a shift from mostly deductive work to detailed experimental designs that used improved technologies and techniques, although this led to research over details of established theories, rather than researching new theories. Through the 1950s and 1960s, as improvements were made in ocean geology and geophysics, it became clearer Wegener's theory on continental drift was correct and that there is constant movement of parts (the plates) of Earth's surface. Improvements were also made in geomorphology to quantify slope forms and drainage networks, and to find relationships between the form and process, and the magnitude and frequency of geomorphic processes. [9] The final blow to peneplanation came in 1964 when a team led by Luna Leopold published Fluvial Processes in Geomorphology, which links landforms with measurable precipitation-infiltration runoff processes and concluded no peneplains exist over large areas in modern times, and any historical peneplains would have to be proven to exist, rather than inferred from modern geology. They also stated pediments could form across all rock types and regions, although through different processes. [20] Through these findings and improvements in geophysics, the study of denudation shifted from planation to studying which relationships affect denudation–including uplift, isostasy, lithology, and vegetation–and measuring denudation rates around the world. [9]

Measurement

Denudation is measured in the wearing down of Earth's surface in inches or centimeters per 1000 years. [21] This rate is intended as an estimate and often assumes uniform erosion, among other things, to simplify calculations. Assumptions made are often only valid for the landscapes being studied. Measurements of denudation over large areas are performed by averaging the rates of subdivisions. Often, no adjustments are made for human impact, which causes the measurements to be inflated. [22] [ ambiguous ] Calculations have suggested soil loss of up to 0.5 metres (20 in) caused by human activity will change previously calculated denudation rates by less than 30%. [23]

Denudation rates are usually much lower than the rates of uplift and average orogeny rates can be eight times the maximum average denudation. [24] The only areas at which there could be equal rates of denudation and uplift are active plate margins with an extended period of continuous deformation. [25]

Denudation is measured in catchment-scale measurements and can use other erosion measurements, which are generally split into dating and survey methods. Techniques for measuring erosion and denudation include stream load measurement, cosmogenic exposure and burial dating, erosion tracking, topographic measurements, surveying the deposition in reservoirs, landslide mapping, chemical fingerprinting, thermochronology, and analysis of sedimentary records in deposition areas. [26] The most common way of measuring denudation is from stream load measurements taken at gauging stations. [21] The suspended load, bed load, and dissolved load are included in measurements. The weight of the load is converted to volumetric units and the load volume is divided by the area of the watershed above the gauging station. [21] An issue with this method of measurement is the high annual variation in fluvial erosion, which can be up to a factor of five between successive years. [27] An important equation for denudation is the stream power law: , where E is erosion rate, K is the erodibility constant, A is drainage area, S is channel gradient, and m and n are functions that are usually given beforehand or assumed based on the location. [8] Most denudation measurements are based on stream load measurements and analysis of the sediment or the water chemistry.

A more recent technique is cosmogenic isotope analysis, which is used in conjunction with stream load measurements and sediment analysis. This technique measures chemical weathering intensity by calculating chemical alteration in molecular proportions. [23] Preliminary research into using cosmogenic isotopes to measure weathering was done by studying the weathering of feldspar and volcanic glass, which contain most of the material found in the Earth's upper crust. The most common isotopes used are 26Al and 10Be; however, 10Be is used more often in these analyses. 10Be is used due to its abundance and, while it is not stable, its half-life of 1.39 million years is relatively stable compared to the thousand or million-year scale in which denudation is measured. 26Al is used because of the low presence of Al in quartz, making it easy to separate, and because there is no risk of contamination of atmospheric 10Be. [28] This technique was developed because previous denudation-rate studies assumed steady rates of erosion even though such uniformity is difficult to verify in the field and may be invalid for many landscapes; its use to help measure denudation and geologically date events was important. [29] On average, the concentration of undisturbed cosmogenic isotopes in sediment leaving a particular basin is inversely related to the rate at which that basin is eroding. In a rapidly-eroding basin, most rock will be exposed to only a small number of cosmic rays before erosion and transport out of the basin; as a result, isotope concentration will be low. In a slowly-eroding basin, integrated cosmic ray exposure is much greater and isotope concentration will be much higher. [23] Measuring isotopic reservoirs in most areas is difficult with this technique so uniform erosion is assumed. There is also variation in year-to-year measurements, which can be as high as a factor of three. [30]

Problems in measuring denudation include both the technology used and the environment. [26] Landslides can interfere with denudation measurements in mountainous regions, especially the Himalayas. [31] The tow main problems with dating methods are uncertainties in the measurements, both with equipment used and with assumptions made during measurement; and the relationship between the measured ages and histories of the markers. [26] This relates to the problem of making assumptions based on the measurements being made and the area being measured. Environmental factors such as temperature, atmospheric pressure, humidity, elevation, wind, the speed of light at higher elevations if using lasers or time of flight measurements, instrument drift, [26] chemical erosion, and for cosmogenic isotopes, climate and snow or glacier coverage. [31] When studying denudation, the Stadler effect, which states measurements over short time periods show higher accumulation rates and than measurements over longer time periods, should be considered. [32] In a study by James Gilully, the presented data suggested the denudation rate has stayed roughly the same throughout the Cenozoic era based on geological evidence; [33] however, given estimates of denudation rates at the time of Gilully's study and the United States' elevation, it would take 11-12 million years to erode North America; [27] well before the 66 million years of the Cenozoic. [34]

The research on denudation is primarily done in river basins and in mountainous regions like the Himalayas because these are very geologically active regions, [35] which allows for research between uplift and denudation. There is also research on the effects of denudation on karst because only about 30% of chemical weathering from water occurs on the surface. [36] Denudation has a large impact on karst and landscape evolution because the most-rapid changes to landscapes occur when there are changes to subterranean structures. [36] Other research includes effects on denudation rates; this research is mostly studying how climate [37] and vegetation [38] impact denudation. Research is also being done to find the relationship between denudation and isostasy; the more denudation occurs, the lighter the crust becomes in an area, which allows for uplift. [39] The work is primarily trying to determine a ratio between denudation and uplift so better estimates can be made on changes in the landscape. In 2016 and 2019, research that attempted to apply denudation rates to improve the stream power law so it can be used more effectively was conducted. [40] [41]

Examples

A) Villarrica Volcano, Chile, a volcano without effects of erosion and denudation
B) Chachahen Volcano, Mendoza Province, Argentina, a volcano with strong effects of erosion but no denudation
C) Cardiel Lake, Santa Cruz Province, Argentina, a volcanic area with strong effects of denudation, exposing subvolcanic rock body. Amotoki Denudation 1.jpg
A) Villarrica Volcano, Chile, a volcano without effects of erosion and denudation
B) Chachahén Volcano, Mendoza Province, Argentina, a volcano with strong effects of erosion but no denudation
C) Cardiel Lake, Santa Cruz Province, Argentina, a volcanic area with strong effects of denudation, exposing subvolcanic rock body.

Denudation exposes deep subvolcanic structures on the present surface of the area where volcanic activity once occurred. Subvolcanic structures such as volcanic plugs and dikes are exposed by denudation.

Other examples include:

Related Research Articles

<span class="mw-page-title-main">Erosion</span> Natural processes that remove soil and rock

Erosion is the action of surface processes that removes soil, rock, or dissolved material from one location on the Earth's crust and then transports it to another location where it is deposited. Erosion is distinct from weathering which involves no movement. Removal of rock or soil as clastic sediment is referred to as physical or mechanical erosion; this contrasts with chemical erosion, where soil or rock material is removed from an area by dissolution. Eroded sediment or solutes may be transported just a few millimetres, or for thousands of kilometres.

<span class="mw-page-title-main">Sediment</span> Particulate solid matter that is deposited on the surface of land

Sediment is a naturally occurring material that is broken down by processes of weathering and erosion, and is subsequently transported by the action of wind, water, or ice or by the force of gravity acting on the particles. For example, sand and silt can be carried in suspension in river water and on reaching the sea bed deposited by sedimentation; if buried, they may eventually become sandstone and siltstone through lithification.

<span class="mw-page-title-main">Geomorphology</span> Scientific study of landforms

Geomorphology is the scientific study of the origin and evolution of topographic and bathymetric features generated by physical, chemical or biological processes operating at or near Earth's surface. Geomorphologists seek to understand why landscapes look the way they do, to understand landform and terrain history and dynamics and to predict changes through a combination of field observations, physical experiments and numerical modeling. Geomorphologists work within disciplines such as physical geography, geology, geodesy, engineering geology, archaeology, climatology, and geotechnical engineering. This broad base of interests contributes to many research styles and interests within the field.

<span class="mw-page-title-main">Peneplain</span> Low-relief plain formed by protracted erosion

In geomorphology and geology, a peneplain is a low-relief plain formed by protracted erosion. This is the definition in the broadest of terms, albeit with frequency the usage of peneplain is meant to imply the representation of a near-final stage of fluvial erosion during times of extended tectonic stability. Peneplains are sometimes associated with the cycle of erosion theory of William Morris Davis, but Davis and other workers have also used the term in a purely descriptive manner without any theory or particular genesis attached.

<span class="mw-page-title-main">Baltic Shield</span> Ancient segment of Earths crust

The Baltic Shield is a segment of the Earth's crust belonging to the East European Craton, representing a large part of Fennoscandia, northwestern Russia and the northern Baltic Sea. It is composed mostly of Archean and Proterozoic gneisses and greenstone which have undergone numerous deformations through tectonic activity. It contains the oldest rocks of the European continent with a thickness of 250–300 km.

<span class="mw-page-title-main">Raised beach</span> Emergent coastal landform

A raised beach, coastal terrace, or perched coastline is a relatively flat, horizontal or gently inclined surface of marine origin, mostly an old abrasion platform which has been lifted out of the sphere of wave activity. Thus, it lies above or under the current sea level, depending on the time of its formation. It is bounded by a steeper ascending slope on the landward side and a steeper descending slope on the seaward side. Due to its generally flat shape, it is often used for anthropogenic structures such as settlements and infrastructure.

<span class="mw-page-title-main">Walther Penck</span> German geologist (1888–1923)

Walther Penck was a geologist and geomorphologist known for his theories on landscape evolution. Penck is noted for criticizing key elements of the Davisian cycle of erosion, concluding that the process of uplift and denudation occur simultaneously, at gradual and continuous rates. Penck's idea of parallel slope retreat led to revisions of Davis's cycle of erosion.

The geographic cycle, or cycle of erosion, is an idealized model that explains the development of relief in landscapes. The model starts with the erosion that follows uplift of land above a base level and ends, if conditions allow, in the formation of a peneplain. Landscapes that show evidence of more than one cycle of erosion are termed "polycyclical". The cycle of erosion and some of its associated concepts have, despite their popularity, been a subject of much criticism.

<span class="mw-page-title-main">Abrasion (geology)</span> Process of erosion

Abrasion is a process of erosion that occurs when material being transported wears away at a surface over time, commonly happens in ice and glaciers. The primary process of abrasion is physical weathering. Its the process of friction caused by scuffing, scratching, wearing down, marring, and rubbing away of materials. The intensity of abrasion depends on the hardness, concentration, velocity and mass of the moving particles. Abrasion generally occurs in four ways: glaciation slowly grinds rocks picked up by ice against rock surfaces; solid objects transported in river channels make abrasive surface contact with the bed and walls; objects transported in waves breaking on coastlines; and by wind transporting sand or small stones against surface rocks. Abrasion is the natural scratching of bedrock by a continuous movement of snow or glacier downhill. This is caused by a force, friction, vibration, or internal deformation of the ice, and by sliding over the rocks and sediments at the base that causes the glacier to move.

Attrition is the process of erosion that occurs during rock collision and transportation. The transportation of sediment chips and smooths the surfaces of bedrock; this can be through water or wind. Rocks undergoing attrition erosion are often found on or near the bed of a stream. Attrition is also partially responsible for turning boulders into smaller rocks and eventually to sand.

<span class="mw-page-title-main">Pediment (geology)</span> Very gently sloping inclined bedrock surface

A pediment, also known as a concave slope or waning slope, is a very gently sloping (0.5°–7°) inclined bedrock surface. It is typically a concave surface sloping down from the base of a steeper retreating desert cliff, escarpment, or surrounding a monadnock or inselberg, but may persist after the higher terrain has eroded away.

<span class="mw-page-title-main">Backswamp</span> Environment on a floodplain where deposits settle after a flood

In geology, a backswamp is a type of depositional environment commonly found in a floodplain. It is where deposits of fine silts and clays settle after a flood. These deposits create a marsh-like landscape that is often poorly drained and usually lower than the rest of the floodplain.

<span class="mw-page-title-main">Dissolved load</span>

Dissolved load is the portion of a stream's total sediment load that is carried in solution, especially ions from chemical weathering. It is a major contributor to the total amount of material removed from a river's drainage basin, along with suspended load and bed load. The amount of material carried as dissolved load is typically much smaller than the suspended load, though this is not always the case, particularly when the available river flow is mostly harnessed for purposes such as irrigation or industrial uses. Dissolved load comprises a significant portion of the total material flux out of a landscape, and its composition is important in regulating the chemistry and biology of the stream water.

River channel migration is the geomorphological process that involves the lateral migration of an alluvial river channel across its floodplain. This process is mainly driven by the combination of bank erosion of and point bar deposition over time. When referring to river channel migration, it is typically in reference to meandering streams. In braided streams, channel change is driven by sediment transport.

Surface exposure dating is a collection of geochronological techniques for estimating the length of time that a rock has been exposed at or near Earth's surface. Surface exposure dating is used to date glacial advances and retreats, erosion history, lava flows, meteorite impacts, rock slides, fault scarps, cave development, and other geological events. It is most useful for rocks which have been exposed for between 103 and 106 years.

<span class="mw-page-title-main">Planation surface</span> Large-scale surface that is almost flat

In geology and geomorphology a planation surface is a large-scale surface that is almost flat with the possible exception of some residual hills. The processes that form planation surfaces are labelled collectively planation and are exogenic. Planation surfaces are planated regardless of bedrock structures. On Earth, they constitute some of the most common landscapes. Geological maps indicate that planation surfaces may comprise 65% of the landscapes on Saturn's largest moon, Titan, which hosts a hydrological cycle of liquid methane. Peneplains and pediplains are types of planation surfaces planated respectively by "peneplanation" and "pediplanation". In addition to these there are planation surfaces proposed to be formed by cryoplanation, marine processes, areal glacial erosion and salt weathering. The term planation surface is often preferred over others because some more specific planation surface types and processes remain controversial. Etchplains are weathered planation surfaces.

<span class="mw-page-title-main">River incision</span>

River incision is the narrow erosion caused by a river or stream that is far from its base level. River incision is common after tectonic uplift of the landscape. Incision by multiple rivers result in a dissected landscape, for example a dissected plateau. River incision is the natural process by which a river cuts downward into its bed, deepening the active channel. Though it is a natural process, it can be accelerated rapidly by human factors including land use changes such as timber harvest, mining, agriculture, and road and dam construction. The rate of incision is a function of basal shear-stress. Shear stress is increased by factors such as sediment in the water, which increase its density. Shear stress is proportional to water mass, gravity, and WSS:

The glacial buzzsaw is a hypothesis claiming erosion by warm-based glaciers is key to limit the height of mountains above certain threshold altitude. To this the hypothesis adds that great mountain massifs are leveled towards the equilibrium line altitude (ELA), which would act as a “climatic base level”. Starting from the hypothesis it has been predicted that local climate restricts the maximum height that mountain massifs can attain by effect of uplifting tectonic forces. It follows that as local climate is cooler at higher latitudes the highest mountains are lower there compared to the tropics where glaciation is and has been more limited. The mechanism behind the glacial buzzsaw effect would be the erosion of small glaciers that are mostly unable to erode much below the equilibrium line altitude since they do not reach these altitudes because of increased ablation. Instead, large valley glaciers may easily surpass the equilibrium line altitude and do therefore not contribute to a glacial buzzsaw effect. This is said to be the case of the Patagonian ice fields where lack of buzzsaw effect results in rapid tectonic uplift rates.

<span class="mw-page-title-main">Climatic geomorphology</span>

Climatic geomorphology is the study of the role of climate in shaping landforms and the earth-surface processes. An approach used in climatic geomorphology is to study relict landforms to infer ancient climates. Being often concerned about past climates climatic geomorphology considered sometimes to be an aspect of historical geology. Since landscape features in one region might have evolved under climates different from those of the present, studying climatically disparate regions might help understand present-day landscapes. For example, Julius Büdel studied both cold-climate processes in Svalbard and weathering processes in tropical India to understand the origin of the relief of Central Europe, which he argued was a palimpsest of landforms formed at different times and under different climates.

Hillslope evolution is the changes in the erosion rates, erosion styles and form of slopes of hills and mountains over time.

References

  1. "Erosion". Encyclopedia Britannica. Archived from the original on 2015-06-21.
  2. Dixon, John C.; Thorn, Colin E. (2005). "Chemical weathering and landscape development in mid-latitude alpine environments". Geomorphology. 67 (1–2): 127–145. Bibcode:2005Geomo..67..127D. doi:10.1016/j.geomorph.2004.07.009. ISSN   0169-555X.
  3. 1 2 Shukla, Dericks Praise (2017-10-18), "Introductory Chapter: Geomorphology", Hydro-Geomorphology - Models and Trends, InTech, doi: 10.5772/intechopen.70823 , ISBN   978-953-51-3573-9
  4. 1 2 Smithson, P et al (2008) Fundamentals of the Physical Environment (4th ed.)
  5. Kemp, David B.; Sadler, Peter M.; Vanacker, Veerle (2020-11-26). "The human impact on North American erosion, sediment transfer, and storage in a geologic context". Nature Communications. 11 (1): 6012. Bibcode:2020NatCo..11.6012K. doi: 10.1038/s41467-020-19744-3 . ISSN   2041-1723. PMC   7691505 . PMID   33243971.
  6. Blatt H, Middleton G, Murray R (1980). Origin of sedimentary rocks 2e. pp. 245–250.
  7. "Weathering and Landforms 5.1". www.radford.edu. Archived from the original on 2015-03-13. Retrieved 2021-04-16.
  8. 1 2 3 K., Lutgens, Frederick (2018). Essentials of geology. Pearson. ISBN   978-0-13-444662-2. OCLC   958585873.{{cite book}}: CS1 maint: multiple names: authors list (link)
  9. 1 2 3 4 5 6 Orme, A.R. (2013). "1.12 Denudation, Planation, and Cyclicity: Myths, Models, and Reality". Treatise on Geomorphology. 1: 205–232. doi:10.1016/B978-0-12-374739-6.00012-9. ISBN   9780080885223.
  10. Hutton, James Geologe (1997). Theory of the Earth with proofs and illustrations: in four parts. The Geological Society. OCLC   889722081.
  11. Playfair, John (2009). Illustrations of the Huttonian Theory of the Earth. Cambridge: Cambridge University Press. doi:10.1017/cbo9780511973086. ISBN   978-0-511-97308-6.
  12. Lyell, Charles (1830–1833). Principles of Geology: Being an Attempt to Explain the Former Changes of the Earth's Surface, By Reference to Causes Now in Operation. John Murray.
  13. Davis, William M. (November 1899). "The Geographical Cycle". The Geographical Journal. 14 (5): 481–504. doi:10.2307/1774538. ISSN   0016-7398. JSTOR   1774538.
  14. Gilbert, G.K. (1877). "Report on the geology of the Henry Mountains". Monograph. p. 212. doi:10.3133/70039916.
  15. Charles., King, Lester (1967). The morphology of the earth: a study and synthesis of world scenery. Oliver & Boyd. OCLC   421411.{{cite book}}: CS1 maint: multiple names: authors list (link)
  16. Wayland, E.J. (1934). "Rifts, Rivers, Rains and Early Man in Uganda". The Journal of the Royal Anthropological Institute of Great Britain and Ireland. 64: 333–352. doi:10.2307/2843813. JSTOR   2843813.
  17. Jukes, J. B. (1862-02-01). "On the Mode of Formation of some of the River-valleys in the South of Ireland". Quarterly Journal of the Geological Society. 18 (1–2): 378–403. doi:10.1144/gsl.jgs.1862.018.01-02.53. hdl: 2027/uiug.30112068306734 . ISSN   0370-291X. S2CID   130701405.
  18. Oldroyd, D.R. "1.5 Geomorphology in the First Half of the Twentieth Century". Treatise on Geomorphology.
  19. Leighly, John. "Symposium: Walther Penck's Contribution to Geomorphology: [Introduction]". Annals of the Association of American Geographers.
  20. Leopold, Luna B. (1964). Fluvial Processes in Geomorphology. Courier Dover Publications. ISBN   978-0-486-84552-4. OCLC   1137178795.
  21. 1 2 3 Ritter, D.F. 1967. Rates of denudation. Jour. Geol. Educ. 15, C.E.G.S. short rev. 6:154-59
  22. Judson, S. 1968. Erosion of the land. Am. Scientist 56:356-74
  23. 1 2 3 BIERMAN, PAUL; STEIG, ERIC J. (February 1996). <125::aid-esp511>3.0.co;2-8 "Estimating Rates of Denudation Using Cosmogenic Isotope Abundances in Sediment". Earth Surface Processes and Landforms. 21 (2): 125–139. Bibcode:1996ESPL...21..125B. doi:10.1002/(sici)1096-9837(199602)21:2<125::aid-esp511>3.0.co;2-8. ISSN   0197-9337.
  24. Schumm, Stanley Alfred (1963). "The disparity between present rates of denudation and orogeny". Shorter Contributions to General Geology. doi: 10.3133/pp454h . ISSN   2330-7102.
  25. Burbank, D.W., and Beck, R.A. 1991. Rapid long-term rates of denudation. Geology 19:1169-72
  26. 1 2 3 4 Turowski, Jens M.; Cook, Kristen L. (2016-08-31). "Field techniques for measuring bedrock erosion and denudation". Earth Surface Processes and Landforms. 42 (1): 109–127. doi: 10.1002/esp.4007 . ISSN   0197-9337.
  27. 1 2 Judson, Sheldon; Ritter, Dale F. (1964-08-15). "Rates of regional denudation in the United States". Journal of Geophysical Research. 69 (16): 3395–3401. Bibcode:1964JGR....69.3395J. doi:10.1029/jz069i016p03395. ISSN   0148-0227.
  28. Nishiizumi, K.; Lal, D.; Klein, J.; Middleton, R.; Arnold, J. R. (January 1986). "Production of 10Be and 26Al by cosmic rays in terrestrial quartz in situ and implications for erosion rates". Nature. 319 (6049): 134–136. Bibcode:1986Natur.319..134N. doi:10.1038/319134a0. ISSN   0028-0836. S2CID   4335625.
  29. Kohl, C.P; Nishiizumi, K (September 1992). "Chemical isolation of quartz for measurement of in-situ -produced cosmogenic nuclides". Geochimica et Cosmochimica Acta. 56 (9): 3583–3587. Bibcode:1992GeCoA..56.3583K. doi:10.1016/0016-7037(92)90401-4. ISSN   0016-7037.
  30. Lupker, Maarten; et al. (2012). "10Be-derived Himalayan denudation rates and sediment budgets in the Ganga basin". Earth and Planetary Science Letters. 334: 146. Bibcode:2012E&PSL.333..146L. doi:10.1016/j.epsl.2012.04.020.
  31. 1 2 Ojha, Lujendra; Ferrier, Ken L.; Ojha, Tank (2019-02-26). "Millennial-scale denudation rates in the Himalaya of Far Western Nepal". doi: 10.5194/esurf-2019-7 . hdl: 10150/635019 .{{cite journal}}: Cite journal requires |journal= (help)
  32. Sadler, Peter M. (September 1981). "Sediment Accumulation Rates and the Completeness of Stratigraphic Sections". The Journal of Geology. 89 (5): 569–584. Bibcode:1981JG.....89..569S. doi:10.1086/628623. ISSN   0022-1376. S2CID   140202963.
  33. Gilluly, James (1955-05-01). "Review". Pacific Historical Review. 24 (2): 187–189. doi:10.2307/3634584. ISSN   0030-8684. JSTOR   3634584.
  34. "Cenozoic". www.usgs.gov. Archived from the original on 2021-04-28. Retrieved 2021-04-19.
  35. "The Himalayas [This Dynamic Earth, USGS]". pubs.usgs.gov. Archived from the original on 2006-04-14. Retrieved 2021-04-26.
  36. 1 2 Gabrovšek, Franci (2008). "On concepts and methods for the estimation of dissolutional denudation rates in karst areas". Geomorphology. 106 (1–2): 9–14. doi:10.1016/j.geomorph.2008.09.008.
  37. Wasak-Sęk, Katarzyna (2021). "Buffering role of soil in chemical denudation in mountainous areas affected by windfall events – In light of experimental research". Geomorphology. 381: 107642. Bibcode:2021Geomo.38107642W. doi:10.1016/j.geomorph.2021.107642. S2CID   234056976.
  38. Torres Acosta, Veronica; Schildgen, Taylor; Clarke, Brian; Scherler, Dirk; Bookhagen, Bodo; Wittmann, Hella; von Blankenburg, Friedhelm; Strecker, Manfred (2014-05-01). "Effects of vegetation cover on landscape denudation rates". EGU General Assembly 2014, held 27 April - 2 May 2014 in Vienna, Austria. p. 8857. Bibcode:2014EGUGA..16.8857T. 8857.
  39. Gilchrist, A.R. (1990). "Differential denudation and flexural isostasy in formation of rifted-margin upwarps". Nature. 346 (6286): 739–742. Bibcode:1990Natur.346..739G. doi:10.1038/346739a0. S2CID   42743054.
  40. Harel, M.-A. (2016). "Global analysis of the stream power law parameters based on worldwide 10Be denudation rates" (PDF). Geomorphology. 268: 184. Bibcode:2016Geomo.268..184H. doi:10.1016/j.geomorph.2016.05.035. hdl: 20.500.11820/a17cd979-3d06-461c-9f14-4a6afa545c10 .
  41. Yuan, X. P.; Braun, J.; Guerit, L.; Rouby, D.; Cordonnier, G. (2019). "A New Efficient Method to Solve the Stream Power Law Model Taking Into Account Sediment Deposition". Journal of Geophysical Research: Earth Surface. 124 (6): 1346–1365. Bibcode:2019JGRF..124.1346Y. doi:10.1029/2018JF004867. ISSN   2169-9011. S2CID   146610951.
  42. Motoki, Akihisa; Sichel, Susanna. Avaliação de aspectos texturais e estruturais de corpos vulcânicos e sub vulcânicos e sua relação com oambiente de cristalização, com base em exemplos do Brasil, Argentina e Chile (PDF) (in Portuguese).
  43. "Betsiboka Estuary, Madagascar". earthobservatory.nasa.gov. 2004-04-12. Archived from the original on 2020-07-20. Retrieved 2021-04-22.