To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Geothermal gradient

From Wikipedia, the free encyclopedia

Temperature profile of inner Earth, schematic view (estimated). The red dashed line shows the minimum temperature for the respective mantle rock to melt. The geothermal gradient remains below the melting temperature of the rock, except in the asthenosphere. Sharp rises occur in the uppermost mantle and at the core–mantle boundary.

Geothermal gradient is the rate of change in temperature with respect to increasing depth in Earth's interior. As a general rule, the crust temperature rises with depth due to the heat flow from the much hotter mantle; away from tectonic plate boundaries, temperature rises in about 25–30 °C/km (72–87 °F/mi) of depth near the surface in the continental crust.[1] However, in some cases the temperature may drop with increasing depth, especially near the surface, a phenomenon known as inverse or negative geothermal gradient. The effects of weather, the Sun, and season only reach a depth of roughly 10–20 m (33–66 ft).

Strictly speaking, geo-thermal necessarily refers to Earth, but the concept may be applied to other planets. In SI units, the geothermal gradient is expressed as °C/km,[1] K/km,[2] or mK/m.[3] These are all equivalent.

Earth's internal heat comes from a combination of residual heat from planetary accretion, heat produced through radioactive decay, latent heat from core crystallization, and possibly heat from other sources. The major heat-producing nuclides in Earth are potassium-40, uranium-238, uranium-235, and thorium-232.[4] The inner core is thought to have temperatures in the range of 4000 to 7000 K, and the pressure at the centre of the planet is thought to be about 360 GPa (3.6 million atm).[5] (The exact value depends on the density profile in Earth.) Because much of the heat is provided for by radioactive decay, scientists believe that early in Earth's history, before nuclides with short half-lives had been depleted, Earth's heat production would have been much higher. Heat production was twice that of present-day at approximately 3 billion years ago,[6] resulting in larger temperature gradients within Earth, larger rates of mantle convection and plate tectonics, allowing the production of igneous rocks such as komatiites that are no longer formed.[7]

The top of the geothermal gradient is influenced by atmospheric temperature. The uppermost layers of the solid planet are at the temperature produced by the local weather, decaying to approximately the annual mean-average ground temperature (MAGT) at a shallow depth of about 10-20 metres depending on the type of ground, rock etc.;[8][9] [10][11][12] it is this depth which is used for many ground-source heat pumps.[13] The top hundreds of meters reflect past climate change;[14] descending further, warmth increases steadily as interior heat sources begin to dominate.

YouTube Encyclopedic

  • 1/5
    Views:
    14 485
    5 713
    1 000
    1 742
    329
  • Calculating a Geothermal Gradient
  • Geothermic Gradient || Mining videos || gas testing examination
  • 30. temperature gradient
  • well logging how to read the log
  • Geothermal Dry Rock Heat extraction equations

Transcription

Heat sources

Earth cutaway from core to exosphere
Geothermal drill machine in Wisconsin, USA

Temperature within Earth increases with depth. Highly viscous or partially molten rock at temperatures between 650 and 1,200 °C (1,200 and 2,200 °F) are found at the margins of tectonic plates, increasing the geothermal gradient in the vicinity, but only the outer core is postulated to exist in a molten or fluid state, and the temperature at Earth's inner core/outer core boundary, around 3,500 kilometres (2,200 mi) deep, is estimated to be 5650 ± 600 Kelvin.[15][16] The heat content of Earth is 1031 joules.[1]

The radiogenic heat from the decay of 238U and 232Th are now the major contributors to Earth's internal heat budget.

In Earth's continental crust, the decay of natural radioactive nuclides makes a significant contribution to geothermal heat production. The continental crust is abundant in lower density minerals but also contains significant concentrations of heavier lithophilic elements such as uranium. Because of this, it holds the most concentrated global reservoir of radioactive elements found in Earth.[19] Naturally occurring radioactive elements are enriched in the granite and basaltic rocks, especially in layers closer to Earth's surface.[20] These high levels of radioactive elements are largely excluded from Earth's mantle due to their inability to substitute in mantle minerals and consequent enrichment in melts during mantle melting processes. The mantle is mostly made up of high density minerals with higher concentrations of elements that have relatively small atomic radii, such as magnesium (Mg), titanium (Ti), and calcium (Ca).[19]

Present-day major heat-producing nuclides[21]
Nuclide Heat release

[W/kg nuclide]

Half-life

[years]

Mean mantle concentration

[kg nuclide/kg mantle]

Heat release

[W/kg mantle]

238U 9.46 × 10−5 4.47 × 109 30.8 × 10−9 2.91 × 10−12
235U 56.9 × 10−5 0.704 × 109 0.22 × 10−9 0.125 × 10−12
232Th 2.64 × 10−5 14.0 × 109 124 × 10−9 3.27 × 10−12
40K 2.92 × 10−5 1.25 × 109 36.9 × 10−9 1.08 × 10−12

The geothermal gradient is steeper in the lithosphere than in the mantle because the mantle transports heat primarily by convection, leading to a geothermal gradient that is determined by the mantle adiabat, rather than by the conductive heat transfer processes that predominate in the lithosphere, which acts as a thermal boundary layer of the convecting mantle.[citation needed]

Heat flow

Heat flows constantly from its sources within Earth to the surface. Total heat loss from Earth is estimated at 44.2 TW (4.42 × 1013 Watts).[22] Mean heat flow is 65 mW/m2 over continental crust and 101 mW/m2 over oceanic crust.[22] This is 0.087 watt/square metre on average (0.03 percent of solar power absorbed by Earth[23]), but is much more concentrated in areas where the lithosphere is thin, such as along mid-ocean ridges (where new oceanic lithosphere is created) and near mantle plumes.[24] Earth's crust effectively acts as a thick insulating blanket which must be pierced by fluid conduits (of magma, water or other) in order to release the heat underneath. More of the heat in Earth is lost through plate tectonics, by mantle upwelling associated with mid-ocean ridges. Another major mode of heat loss is by conduction through the lithosphere, the majority of which occurs in the oceans due to the crust there being much thinner and younger than under the continents.[22][25]

The heat of Earth is replenished by radioactive decay at a rate of 30 TW.[26] The global geothermal flow rates are more than twice the rate of human energy consumption from all primary sources. Global data on heat-flow density are collected and compiled by the International Heat Flow Commission (IHFC) of the IASPEI/IUGG.[27]

Direct application

Heat from Earth's interior can be used as an energy source, known as geothermal energy. The geothermal gradient has been used for space heating and bathing since ancient Roman times, and more recently for generating electricity. As the human population continues to grow, so does energy use and the correlating environmental impacts that are consistent with global primary sources of energy. This has caused a growing interest in finding sources of energy that are renewable and have reduced greenhouse gas emissions. In areas of high geothermal energy density, current technology allows for the generation of electrical power because of the corresponding high temperatures. Generating electrical power from geothermal resources requires no fuel while providing true baseload energy at a reliability rate that constantly exceeds 90%.[19] In order to extract geothermal energy, it is necessary to efficiently transfer heat from a geothermal reservoir to a power plant, where electrical energy is converted from heat by passing steam through a turbine connected to a generator.[19] The efficiency of converting the geothermal heat into electricity depends on the temperature difference between the heated fluid (water or steam) and the environmental temperature, so it is advantageous to use deep, high-temperature heat sources. On a worldwide scale, the heat stored in Earth's interior provides an energy that is still seen as an exotic source. About 10 GW of geothermal electric capacity is installed around the world as of 2007, generating 0.3% of global electricity demand. An additional 28 GW of direct geothermal heating capacity is installed for district heating, space heating, spas, industrial processes, desalination and agricultural applications.[1]

Variations

The geothermal gradient varies with location and is typically measured by determining the bottom open-hole temperature after borehole drilling. Temperature logs obtained immediately after drilling are however affected due to drilling fluid circulation. To obtain accurate bottom hole temperature estimates, it is necessary for the well to reach stable temperature. This is not always achievable for practical reasons.

In stable tectonic areas in the tropics, a temperature-depth plot will converge to the annual average surface temperature. However, in areas where deep permafrost developed during the Pleistocene, a low temperature anomaly can be observed that persists down to several hundred metres.[28] The Suwałki cold anomaly in Poland has led to the recognition that similar thermal disturbances related to Pleistocene-Holocene climatic changes are recorded in boreholes throughout Poland, as well as in Alaska, northern Canada, and Siberia.

In areas of Holocene uplift and erosion (Fig. 1) the shallow gradient will be high until it reaches a point (labeled "Inflection point" in the figure) where it reaches the stabilized heat-flow regime. If the gradient of the stabilized regime is projected above this point to its intersection with present-day annual average temperature, the height of this intersection above present-day surface level gives a measure of the extent of Holocene uplift and erosion. In areas of Holocene subsidence and deposition (Fig. 2) the initial gradient will be lower than the average until it reaches a point where it joins the stabilized heat-flow regime.

Variations in surface temperature, whether daily, seasonal, or induced by climate changes and the Milankovitch cycle, penetrate below Earth's surface and produce an oscillation in the geothermal gradient with periods varying from a day to tens of thousands of years, and an amplitude which decreases with depth. The longest-period variations have a scale depth of several kilometers.[29][30] Melt water from the polar ice caps flowing along ocean bottoms tends to maintain a constant geothermal gradient throughout Earth's surface.[29][dubious ][verification needed]

If the rate of temperature increase with depth observed in shallow boreholes were to persist at greater depths, temperatures deep within Earth would soon reach the point where rocks would melt. We know, however, that Earth's mantle is solid because of the transmission of S-waves. The temperature gradient dramatically decreases with depth for two reasons. First, the mechanism of thermal transport changes from conduction, as within the rigid tectonic plates, to convection, in the portion of Earth's mantle that convects. Despite its solidity, most of Earth's mantle behaves over long time-scales as a fluid, and heat is transported by advection, or material transport. Second, radioactive heat production is concentrated within the crust of Earth, and particularly within the upper part of the crust, as concentrations of uranium, thorium, and potassium are highest there: these three elements are the main producers of radioactive heat within Earth. Thus, the geothermal gradient within the bulk of Earth's mantle is of the order of 0.5 kelvin per kilometer, and is determined by the adiabatic gradient associated with mantle material (peridotite in the upper mantle).[31]

Negative geothermal gradient

Negative geothermal gradients occur where temperature decreases with depth. This occurs in the upper few hundreds of meters near the surface. Because of the low thermal diffusivity of rocks, deep underground temperatures are hardly affected by diurnal or even annual surface temperature variations. At depths of a few meters, underground temperatures are therefore similar to the annual average surface temperature. At greater depths, underground temperatures reflect a long-term average over past climate, so that temperatures at the depths of dozens to hundreds of meters contain information about the climate of the last hundreds to thousands of years. Depending on the location, these may be colder than current temperatures due to the colder weather close to the last ice age, or due to more recent climate change.[32][33][14]

Negative geothermal gradients may also occur due to deep aquifers, where heat transfer from deep water by convection and advection results in water at shallower levels heating adjacent rocks to a higher temperature than rocks at a somewhat deeper level.[34]

Negative geothermal gradients are also found at large scales in subduction zones.[35] A subduction zone is a tectonic plate boundary where oceanic crust sinks into the mantle due to the high density of the oceanic plate relative to the underlying mantle. Since the sinking plate enters the mantle at a rate of a few centimeters per year, heat conduction is unable to heat the plate as quickly as it sinks. Therefore, the sinking plate has a lower temperature than the surrounding mantle, resulting in a negative geothermal gradient.[35]

See also

References

  1. ^ a b c d Fridleifsson, Ingvar B.; Bertani, Ruggero; Huenges, Ernst; Lund, John W.; Ragnarsson, Arni; Rybach, Ladislaus (2008-02-11). O. Hohmeyer and T. Trittin (ed.). "The possible role and contribution of geothermal energy to the mitigation of climate change" (PDF). IPCC Scoping Meeting on Renewable Energy Sources, Proceedings. Luebeck, Germany: 59–80. CiteSeerX 10.1.1.362.1202. Archived from the original on 2013-03-12. Retrieved 2013-11-03.
  2. ^ Jones, M. Q. W. (July 2018). "Virgin rock temperatures and geothermal gradients in the Bushveld Complex". Journal of the Southern African Institute of Mining and Metallurgy. 118 (7): 671–680. doi:10.17159/2411-9717/2018/v118n7a1. ISSN 2225-6253.
  3. ^ "Component parts of the World Heat Flow Data Collection". Pangaea. Global Heat Flow Compilation Group. 11 April 2013. doi:10.1594/PANGAEA.810104. Retrieved 2021-09-23.
  4. ^ Sanders, Robert (2003-12-10). "Radioactive potassium may be major heat source in Earth's core". UC Berkeley News. Retrieved 2007-02-28.
  5. ^ Alfè, D.; Gillan, M. J.; Vocadlo, L.; Brodholt, J.; Price, G. D. (2002). "The ab initio simulation of the Earth's core" (PDF). Philosophical Transactions of the Royal Society. 360 (1795): 1227–44. Bibcode:2002RSPTA.360.1227A. doi:10.1098/rsta.2002.0992. PMID 12804276. S2CID 21132433. Retrieved 2007-02-28.
  6. ^ a b Turcotte, DL; Schubert, G (2002). "4". Geodynamics (2nd ed.). Cambridge, England, UK: Cambridge University Press. pp. 136–7. ISBN 978-0-521-66624-4.
  7. ^ Vlaar, N; Vankeken, P; Vandenberg, A (1994). "Cooling of the earth in the Archaean: Consequences of pressure-release melting in a hotter mantle". Earth and Planetary Science Letters. 121 (1–2): 1–18. Bibcode:1994E&PSL.121....1V. doi:10.1016/0012-821X(94)90028-0.
  8. ^ Kalogirou, Soteris & Florides, Georgios. (2004). Measurements of Ground Temperature at Various Depths, conference paper 3rd International Conference on Sustainable Energy Technologies, Nottingham, UK, https://www.researchgate.net/publication/30500372_Measurements_of_Ground_Temperature_at_Various_Depths https://ktisis.cut.ac.cy/bitstream/10488/870/3/C55-PRT020-SET3.pdf
  9. ^ Williams G. and Gold L. Canadian Building Digest 180m 1976. National Research Council of Canada, Institute for Research in Construction. https://nrc-publications.canada.ca/eng/view/ft/?id=386ddf88-fe8d-45dd-aabb-0a55be826f3f,
  10. ^ "Groundwater temperature's measurement and significance - National Groundwater Association". National Groundwater Association. 23 August 2015. Archived from the original on 23 August 2015.
  11. ^ "Mean Annual Air Temperature - MATT". www.icax.co.uk.
  12. ^ "Ground Temperatures as a Function of Location, Season, and Depth". builditsolar.com.
  13. ^ Rafferty, Kevin (April 1997). "An Information Survival Kit for the Prospective Residential Geothermal Heat Pump Owner" (PDF). Geo-Heat Centre Quarterly Bulletin. Vol. 18, no. 2. Klmath Falls, Oregon: Oregon Institute of Technology. pp. 1–11. ISSN 0276-1084. Archived from the original (PDF) on 17 February 2012. Retrieved 2009-03-21. The author issued an updated version Archived 2013-02-17 at the Wayback Machine of this article in February 2001.
  14. ^ a b Huang, S., H. N. Pollack, and P. Y. Shen (2000), Temperature trends over the past five centuries reconstructed from borehole temperatures, Nature, 403, 756–758.
  15. ^ Alfe, D.; M. J. Gillan; G. D. Price (2003-02-01). "Thermodynamics from first principles: temperature and composition of Earth's core" (PDF). Mineralogical Magazine. 67 (1): 113–123. Bibcode:2003MinM...67..113A. doi:10.1180/0026461026610089. S2CID 98605003. Archived from the original (PDF) on 2007-03-16. Retrieved 2007-03-01.
  16. ^ Steinle-Neumann, Gerd; Lars Stixrude; Ronald Cohen (2001-09-05). "New Understanding of Earth's Inner Core". Carnegie Institution of Washington. Archived from the original on 2006-12-14. Retrieved 2007-03-01.
  17. ^ Anuta, Joe (2006-03-30). "Probing Question: What heats the earth's core?". physorg.com. Retrieved 2007-09-19.
  18. ^ Johnston, Hamish (19 July 2011). "Radioactive decay accounts for half of Earth's heat". PhysicsWorld.com. Institute of Physics. Retrieved 18 June 2013.
  19. ^ a b c d William, G. E. (2010). Geothermal Energy: Renewable Energy and the Environment (pp. 1-176). Boca Raton, FL: CRC Press.
  20. ^ Wengenmayr, R., & Buhrke, T. (Eds.). (2008). Renewable Energy: Sustainable Energy Concepts for the future (pp. 54-60). Weinheim, Germany: WILEY-VCH Verlag GmbH & Co. KGaA.
  21. ^ Turcotte, D. L.; Schubert, G. (2002). "4". Geodynamics (2nd ed.). Cambridge, England, UK: Cambridge University Press. p. 137. ISBN 978-0-521-66624-4.
  22. ^ a b c Pollack, Henry N., et.al.,Heat flow from Earth's interior: Analysis of the global data set, Reviews of Geophysics, 31, 3 / August 1993, p. 273  Archived 2011-08-11 at the Wayback Machine doi:10.1029/93RG01249
  23. ^ "Climate and Earth's Energy Budget". NASA. 2009-01-14.
  24. ^ Richards, M. A.; Duncan, R. A.; Courtillot, V. E. (1989). "Flood Basalts and Hot-Spot Tracks: Plume Heads and Tails". Science. 246 (4926): 103–107. Bibcode:1989Sci...246..103R. doi:10.1126/science.246.4926.103. PMID 17837768. S2CID 9147772.
  25. ^ Sclater, John G; Parsons, Barry; Jaupart, Claude (1981). "Oceans and Continents: Similarities and Differences in the Mechanisms of Heat Loss". Journal of Geophysical Research. 86 (B12): 11535. Bibcode:1981JGR....8611535S. doi:10.1029/JB086iB12p11535.
  26. ^ Rybach, Ladislaus (September 2007). "Geothermal Sustainability" (PDF). Geo-Heat Centre Quarterly Bulletin. Vol. 28, no. 3. Klamath Falls, Oregon: Oregon Institute of Technology. pp. 2–7. ISSN 0276-1084. Archived from the original (PDF) on 2018-03-08. Retrieved 2018-03-07.
  27. ^ www.ihfc-iugg.org IHFC: International Heat Flow Commission - Homepage. Retrieved 18/09/2019.
  28. ^ The Frozen Time, from the Polish Geological Institute Archived 2010-10-27 at the Wayback Machine
  29. ^ a b Stacey, Frank D. (1977). Physics of the Earth (2nd ed.). New York: John Wiley & Sons. ISBN 0-471-81956-5. pp. 183-4
  30. ^ Sleep, Norman H.; Kazuya Fujita (1997). Principles of Geophysics. Blackwell Science. ISBN 0-86542-076-9. pp. 187-9
  31. ^ Turcotte, D. L.; Schubert, G. (2002). "4". Geodynamics (2nd ed.). Cambridge, England, UK: Cambridge University Press. p. 187. ISBN 978-0-521-66624-4.
  32. ^ Lachenbruch, A. H., & Marshall, B. V. (1986). Changing climate: geothermal evidence from permafrost in the Alaskan Arctic. Science, 234(4777), 689-696.
  33. ^ Šafanda, J., Szewczyk, J., & Majorowicz, J. (2004). Geothermal evidence of very low glacial temperatures on a rim of the Fennoscandian ice sheet. Geophysical Research Letters, 31(7).
  34. ^ Ziagos, J. P., & Blackwell, D. D. (1986). A model for the transient temperature effects of horizontal fluid flow in geothermal systems. Journal of Volcanology and Geothermal Research, 27(3-4), 371-397.
  35. ^ a b Ernst, W.G., (1976) Petrologic Phase Equilibria, W.H. Freeman, San Francisco.

"Geothermal Resources". DOE/EIA-0603(95) Background Information and 1990 Baseline Data Initially Published in the Renewable Energy Annual 1995. Retrieved May 4, 2005.

This page was last edited on 29 February 2024, at 23:07
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.