Ocean dynamical thermostat

Last updated

Ocean dynamical thermostat is a physical mechanism through which changes in the mean radiative forcing influence the gradients of sea surface temperatures in the Pacific Ocean and the strength of the Walker circulation. Increased radiative forcing (warming) is more effective in the western Pacific than in the eastern where the upwelling of cold water masses damps the temperature change. This increases the east-west temperature gradient and strengthens the Walker circulation. Decreased radiative forcing (cooling) has the opposite effect.

Contents

The process has been invoked to explain variations in the Pacific Ocean temperature gradients that correlate to insolation and climate variations. It may also be responsible for the hypothesized correlation between El Niño events and volcanic eruptions, and for changes in the temperature gradients that occurred during the 20th century. Whether the ocean dynamical thermostat controls the response of the Pacific Ocean to anthropogenic global warming is unclear, as there are competing processes at play; potentially, it could drive a La Niña-like climate tendency during initial warming before it is overridden by other processes.

Background

The equatorial Pacific is a key region of Earth in terms of its relative influence on the worldwide atmospheric circulation. A characteristic east-west temperature gradient is coupled to an atmospheric circulation, the Walker circulation, [1] and further controlled by atmospheric and oceanic dynamics. [2] The western Pacific features the so-called "warm pool", where the warmest sea surface temperatures (SSTs) of Earth are found. In the eastern Pacific conversely an area called the "cold tongue" is always colder than the warm pool even though they lie at the same latitude, as cold water is upwelled there. The temperature gradient between the two in turn induces an atmospheric circulation, the Walker circulation, [3] which responds strongly to the SST gradient. [4]

One important component of the climate is the El Niño-Southern Oscillation (ENSO), a mode of climate variability. During its positive/El Niño phase, waters in the central and eastern Pacific are warmer than normal while during its cold/La Niña they are colder than normal. Coupled to these SST changes the atmospheric pressure difference between the eastern and western Pacific changes. ENSO and Walker circulation variations have worldwide effects on weather, including natural disasters such as bushfires, droughts, floods and tropical cyclone activity. [5] The atmospheric circulation modulates the heat uptake by the ocean, the strength and position of the Intertropical Convergence Zone (ITCZ), tropical precipitation and the strength of the Indian monsoon. [6]

Original hypothesis by Clement et al.(1996) and Sun and Liu's (1996) precedent

Already in May 1996 Sun and Liu published a hypothesis that coupled interactions between ocean winds, the ocean surface and ocean currents can limit water temperatures in the western Pacific. [7] As part of that study, they found that increased equilibrium temperatures drive an increased temperature gradient between the eastern and western Pacific. [8]

The ocean dynamical thermostat mechanism was described in a dedicated publication by Clement et al. 1996 in a coupled ocean-atmosphere model of the equatorial ocean. Since in the western Pacific SSTs are only governed by stored heat and heat fluxes, while in the eastern Pacific the horizontal and vertical advection also play a role. Thus an imposed source of heating primarily warms the western Pacific, inducing stronger easterly winds that facilitate upwelling in the eastern Pacific and cool its temperature - a pattern opposite that expected from the heating. Cold water upwelled along the equator then spreads away from it, reducing the total warming of the basin. [9] The temperature gradient between the western and eastern Pacific thus increases, strengthening the trade winds and further increasing upwelling; this eventually results in a climate state resembling La Niña. [2] The mechanism is seasonal as upwelling is least effective in boreal spring and most effective in boreal autumn; thus it is mainly operative in autumn. [10] Due to the vertical temperature structure, ENSO variability becomes more regular during cooling by the thermostat mechanism, but is damped during warming. [11]

The model of Clement et al. 1996 only considers temperature anomalies and does not account for the entire energy budget. After some time, warming would spread to the source regions of the upwelled water and in the thermocline, eventually damping the thermostat. [9] The principal flaw in the model is that it assumes that the temperature of the upwelled water does not change over time. [2]

Later research

Later studies have verified the ocean dynamical thermostat mechanism for a number of climate models with different structures of warming [12] and also the occurrence of the opposite response - a decline in the SST gradient - in response to climate cooling. [13] In fully coupled models a tendency of the atmospheric circulation to intensify with decreasing insolation sometimes negates the thermostat response to decreased solar activity. [14] Liu, Lu and Xie 2015 proposed that an ocean dynamical thermostat can also operate in the Indian Ocean, [15] and the concept has been extended to cover the Indo-Pacific as a whole rather than just the equatorial Pacific. [16]

Water flows from the western Pacific into the Indian Ocean through straits between Australia and Asia, a phenomenon known as the Indonesian Throughflow. [17] Rodgers et al. 1999 postulated that stronger trade winds associated with the ocean dynamical thermostat may increase the sea level difference between the Indian and Pacific oceans, increasing the throughflow and cooling the Pacific further. [18] An et al. 2022 postulated a similar effect in the Indian Ocean could force changes to the Indian Ocean Dipole after carbon dioxide removal. [19]

Role in climate variability

The ocean dynamical thermostat has been used to explain:

Volcanic and solar influences

The ocean dynamical thermostat mechanism has been invoked to link volcanic eruptions to ENSO changes. [23] Volcanic eruptions can cool the Earth by injecting aerosols and sulfur dioxide into the stratosphere, which reflect incoming solar radiation. It has been suggested that in paleoclimate records volcanic eruptions are often followed by El Niño events, but it is questionable whether this applies to known historical eruptions [5] and results from climate modelling are equivocal. [13] In some climate models an ocean dynamical thermostat process causes the onset of El Niño events after volcanic eruptions, in others additional atmospheric processes override the effect of the ocean dynamical thermostat on Pacific SST gradients. [24]

The ocean dynamical thermostat process may explain variations in Pacific SSTs in the eastern Pacific that correlate to insolation changes [25] such as the Dalton Minimum [26] and to the solar cycle. [27] During the early and middle Holocene when autumn and summer insolation was increased, but also during the Medieval Climate Anomaly between 900-1300 AD, SSTs off Baja California in the eastern Pacific were colder than usual. Southwestern North America underwent severe megadroughts during this time, which could also relate to a La Niña-like tendency in Pacific SSTs. Conversely, during periods of low insolation [28] [29] and during the Little Ice Age SSTs increased. This region lies within the California Current which is influenced by the eastern Pacific [30] that controls the temperature of upwelled water. [31] This was further corroborated by analyses with additional foraminifera species. [32] Increased productivity in the ocean waters off Peru during the Medieval Climate Anomaly and the Roman Warm Period [33] between 50-400 AD, when the worldwide climate was warmer, [34] may occur through a thermostat-driven shallowing of the thermocline and increased upwelling of nutrient-rich waters. [33] Additional mechanisms connecting the equatorial Pacific climate to insolation changes have been proposed, however. [35]

Role in recent climate change

Changes in equatorial Pacific SSTs caused by anthropogenic global warming are an important problem in climate forecasts, as they influence local and global climate patterns. [36] The ocean dynamical thermostat mechanism is expected to reduce the anthropogenic warming of the eastern Pacific relative to the western Pacific, thus strengthening the SST gradient and the Walker circulation. This is opposed by a weakening of the Walker circulation [1] and the more effective evaporative cooling of the western Pacific under global warming. This compensation between different effects makes it difficult to estimate the eventual outcome of the Walker circulation and SST gradient. [37] In CMIP5 models it is usually not the dominating effect. [38]

The ocean dynamical thermostat has been invoked to explain contradictory changes in the Pacific Ocean in the 20th century. Specifically, there appears to be a simultaneous increase of the SST gradient, but also a weakening of the Walker circulation especially during boreal summer. All these observations are uncertain, owing to the particular choices of metrics used to describe SST gradients and Walker circulation strength, as well as measurement issues and biases. [37] However, the ocean dynamical thermostat mechanism could explain why the SST gradient has increased during global warming [12] and also why Walker circulation becomes stronger in autumn and winter, as these are the seasons when upwelling is strongest. [37] On the other hand, warming in the Atlantic Ocean and more generally changes in between-ocean temperature gradients may play a role. [39]

Projected future changes

Climate models usually depict an El Niño-like change, that is a decrease in the SST gradient. In numerous models, there is a time-dependent pattern with an initial increase in the SST gradient ("fast response") followed by a weakening of the gradient ("slow response") [2] especially but not only in the case of abrupt increases of greenhouse gas concentrations. [40] This may reflect a decreasing strength of the ocean dynamical thermostat with increasing warming [12] and the warming of the upwelled water, which occurs with a delay of a few decades after the surface warming [4] and is known as the "oceanic tunnel". [41] On the other hand, climate models might underestimate the strength of the thermostat effect. [42]

  • According to An and Im 2014, in an oceanic dynamical model a doubling of carbon dioxide concentrations initially cools the eastern Pacific cold tongue, but a further increase in carbon dioxide concentrations eventually causes the cooling to stop and the cold tongue to shrink. [43] [44] Their model does not consider changes in the thermocline temperature, which would tend to occur after over a decade of global warming. [45]
  • According to Luo et al. 2017, the ocean dynamical thermostat eventually is overwhelmed first by a weakening of the trade winds and increased ocean stratification which decrease the supply of cold water to the upwelling zones, [46] and second by the arrival of warmer subtropical waters there. [47] In their model, the transition takes about a decade. [48]
  • According to Heede, Fedorov and Burls 2020, the greater climate warming outside of the tropics than inside of them eventually causes the water arriving to the upwelling regions to warm [49] and the oceanic currents that transport it to weaken. [50] This negates the thermostat effect after about two decades in the case of an abrupt increase of greenhouse gas concentrations, and after about half to one century when greenhouse gas concentrations are increasing more slowly. [51]
  • With further warming of the subsurface ocean, the strength of the ocean dynamical thermostat is expected to decline, because the decreasing stratification means that momentum is less concentrated in the surface layer and thus upwelling decreases. [12]
  • According to Heede and Fedorov 2021, in some climate models the thermostat mechanism initially prevails over other mechanisms and causes a cooling of the subtropical and central Pacific. [52] Eventually most models converge to an equatorial warming pattern. [53]
  • Zhou et al. 2022 found that in carbon dioxide removal scenarios, the thermostat amplifies precipitation changes. [54]

Other contexts

The term "ocean dynamical thermostat" has also been used in slightly different contexts:

Notes

  1. The Equatorial Undercurrent is a strong ocean current under the surface of the equatorial Pacific, [37] which is powered by westward surface winds. These transport water westwards, in turn inducing an eastward pressure force that powers the Equatorial Undercurrent while the wind-driven pressure acts to retard it. [55] The Equatorial Undercurrent reaches the surface in the eastern Pacific and is the main source of upwelled water there. [56]

Related Research Articles

<span class="mw-page-title-main">El Niño–Southern Oscillation</span> Climate phenomenon that periodically fluctuates between three phases

El Niño–Southern Oscillation (ENSO) is a climate phenomenon that exhibits irregular quasi-periodic variation in winds and sea surface temperatures over the tropical Pacific Ocean. It affects the climate of much of the tropics and subtropics, and has links (teleconnections) to higher latitude regions of the world. The warming phase of the sea surface temperature is known as El Niño and the cooling phase as La Niña. The Southern Oscillation is the accompanying atmospheric component, which is coupled with the sea temperature change. El Niño is associated with higher than normal air sea level pressure over Indonesia, Australia and across the Indian Ocean to the Atlantic. La Niña has roughly the reverse pattern: high pressure over the central and eastern Pacific and lower pressure through much of the rest of the tropics and subtropics. The two phenomena last a year or so each and typically occur every two to seven years with varying intensity, with neutral periods of lower intensity interspersed. El Niño events can be more intense but La Niña events may repeat and last longer.

<span class="mw-page-title-main">Ocean current</span> Directional mass flow of oceanic water generated by external or internal forces

An ocean current is a continuous, directed movement of seawater generated by a number of forces acting upon the water, including wind, the Coriolis effect, breaking waves, cabbeling, and temperature and salinity differences. Depth contours, shoreline configurations, and interactions with other currents influence a current's direction and strength. Ocean currents are primarily horizontal water movements.

<span class="mw-page-title-main">Intertropical Convergence Zone</span> Meteorological phenomenon

The Intertropical Convergence Zone, known by sailors as the doldrums or the calms because of its monotonous windless weather, is the area where the northeast and the southeast trade winds converge. It encircles Earth near the thermal equator though its specific position varies seasonally. When it lies near the geographic Equator, it is called the near-equatorial trough. Where the ITCZ is drawn into and merges with a monsoonal circulation, it is sometimes referred to as a monsoon trough.

<span class="mw-page-title-main">Sea surface temperature</span> Water temperature close to the oceans surface

Sea surface temperature (SST), or ocean surface temperature, is the ocean temperature close to the surface. The exact meaning of surface varies in the literature and in practice. It is usually between 1 millimetre (0.04 in) and 20 metres (70 ft) below the sea surface. Sea surface temperatures greatly modify air masses in the Earth's atmosphere within a short distance of the shore. Local areas of heavy snow can form in bands downwind of warm water bodies within an otherwise cold air mass. Warm sea surface temperatures can develop and strengthen cyclones over the Ocean. Experts call this process tropical cyclogenesis. Tropical cyclones can also cause a cool wake. This is due to turbulent mixing of the upper 30 metres (100 ft) of the ocean. Sea surface temperature changes during the day. This is like the air above it, but to a lesser degree. There is less variation in sea surface temperature on breezy days than on calm days. Ocean currents, such as the Atlantic Multidecadal Oscillation, can affect sea surface temperatures over several decades. Thermohaline circulation has a major impact on average sea surface temperature throughout most of the world's oceans.

<span class="mw-page-title-main">Pacific decadal oscillation</span> Recurring pattern of climate variability

The Pacific decadal oscillation (PDO) is a robust, recurring pattern of ocean-atmosphere climate variability centered over the mid-latitude Pacific basin. The PDO is detected as warm or cool surface waters in the Pacific Ocean, north of 20°N. Over the past century, the amplitude of this climate pattern has varied irregularly at interannual-to-interdecadal time scales. There is evidence of reversals in the prevailing polarity of the oscillation occurring around 1925, 1947, and 1977; the last two reversals corresponded with dramatic shifts in salmon production regimes in the North Pacific Ocean. This climate pattern also affects coastal sea and continental surface air temperatures from Alaska to California.

<span class="mw-page-title-main">Walker circulation</span> Theorerical model of air flow in the atmosphere

The Walker circulation, also known as the Walker cell, is a conceptual model of the air flow in the tropics in the lower atmosphere (troposphere). According to this model, parcels of air follow a closed circulation in the zonal and vertical directions. This circulation, which is roughly consistent with observations, is caused by differences in heat distribution between ocean and land. In addition to motions in the zonal and vertical direction the tropical atmosphere also has considerable motion in the meridional direction as part of, for example, the Hadley Circulation.

<span class="mw-page-title-main">Mixed layer</span> Layer in which active turbulence has homogenized some range of depths

The oceanic or limnological mixed layer is a layer in which active turbulence has homogenized some range of depths. The surface mixed layer is a layer where this turbulence is generated by winds, surface heat fluxes, or processes such as evaporation or sea ice formation which result in an increase in salinity. The atmospheric mixed layer is a zone having nearly constant potential temperature and specific humidity with height. The depth of the atmospheric mixed layer is known as the mixing height. Turbulence typically plays a role in the formation of fluid mixed layers.

<span class="mw-page-title-main">Atlantic meridional overturning circulation</span> System of surface and deep currents in the Atlantic Ocean

The Atlantic meridional overturning circulation (AMOC) is the "main current system in the South and North Atlantic Oceans". As such, it is a component of Earth's oceanic circulation system and plays an important role in the climate system. If the strength of the AMOC changes this could have impacts on some elements of the climate system. The AMOC includes currents at the surface as well as at great depths in the Atlantic Ocean. These currents are driven by changes in the atmospheric weather as well as by changes in temperature and salinity. They collectively make up one half of the global thermohaline circulation that encompasses the flow of major ocean currents. The other half is the Southern Ocean overturning circulation.

<span class="mw-page-title-main">Hadley cell</span> Tropical atmospheric circulation feature

The Hadley cell, also known as the Hadley circulation, is a global-scale tropical atmospheric circulation that features air rising near the equator, flowing poleward near the tropopause at a height of 12–15 km (7.5–9.3 mi) above the Earth's surface, cooling and descending in the subtropics at around 25 degrees latitude, and then returning equatorward near the surface. It is a thermally direct circulation within the troposphere that emerges due to differences in insolation and heating between the tropics and the subtropics. On a yearly average, the circulation is characterized by a circulation cell on each side of the equator. The Southern Hemisphere Hadley cell is slightly stronger on average than its northern counterpart, extending slightly beyond the equator into the Northern Hemisphere. During the summer and winter months, the Hadley circulation is dominated by a single, cross-equatorial cell with air rising in the summer hemisphere and sinking in the winter hemisphere. Analogous circulations may occur in extraterrestrial atmospheres, such as on Venus and Mars.

<span class="mw-page-title-main">Atlantic multidecadal oscillation</span> Climate cycle that affects the surface temperature of the North Atlantic

The Atlantic Multidecadal Oscillation (AMO), also known as Atlantic Multidecadal Variability (AMV), is the theorized variability of the sea surface temperature (SST) of the North Atlantic Ocean on the timescale of several decades.

<span class="mw-page-title-main">Indian Ocean Dipole</span> Climatic and oceanographic cycle affecting Southeast Asia, Australia and Africa

The Indian Ocean Dipole (IOD), also known as the Indian Niño, is an irregular oscillation of sea surface temperatures in which the western Indian Ocean becomes alternately warmer and then colder than the eastern part of the ocean.

<span class="mw-page-title-main">Tropical instability waves</span> Ocean waves generated near the equator

Tropical instability waves, often abbreviated TIW, are a phenomenon in which the interface between areas of warm and cold sea surface temperatures near the equator form a regular pattern of westward-propagating waves. These waves are often present in the Atlantic Ocean, extending westward from the African coast, but are more easily recognizable in the Pacific, extending westward from South America. They have an average period of about 30 days and wavelength of about 1100 kilometers, and are largest in amplitude between June and November. They are also largest during La Niña conditions, and may disappear when strong El Niño conditions are present.

The Atlantic Equatorial Mode or Atlantic Niño is a quasiperiodic interannual climate pattern of the equatorial Atlantic Ocean. It is the dominant mode of year-to-year variability that results in alternating warming and cooling episodes of sea surface temperatures accompanied by changes in atmospheric circulation. The term Atlantic Niño comes from its close similarity with the El Niño-Southern Oscillation (ENSO) that dominates the tropical Pacific basin. For this reason, the Atlantic Niño is often called the little brother of El Niño. The Atlantic Niño usually appears in northern summer, and is not the same as the Atlantic Meridional (Interhemispheric) Mode that consists of a north-south dipole across the equator and operates more during northern spring. The equatorial warming and cooling events associated with the Atlantic Niño are known to be strongly related to rainfall variability over the surrounding continents, especially in West African countries bordering the Gulf of Guinea. Therefore, understanding of the Atlantic Niño has important implications for climate prediction in those regions. Although the Atlantic Niño is an intrinsic mode to the equatorial Atlantic, there may be a tenuous causal relationship between ENSO and the Atlantic Niño in some circumstances.

The Tropical Atlantic SST Dipole refers to a cross-equatorial sea surface temperature (SST) pattern that appears dominant on decadal timescales. It has a period of about 12 years, with the SST anomalies manifesting their most pronounced features around 10–15 degrees of latitude off of the Equator. It is also referred to as the interhemispheric SST gradient or the Meridional Atlantic mode.

There are a number of explanations of the asymmetry of the Intertropical Convergence Zone (ITCZ), known by sailors as the Doldrums.

The Tropical Atlantic Variability (TAV) is influenced by internal interaction and external effects. TAV can be discussed in different time scales: seasonal and interannual. tav can be discussed in different time scales:seasonal and interannual.and external effects.

Pacific Centennial Oscillation is a climate oscillation predicted by some climate models.

<span class="mw-page-title-main">Cyclonic Niño</span> Climatological phenomenon

Cyclonic Niño is a climatological phenomenon that has been observed in climate models where tropical cyclone activity is increased. Increased tropical cyclone activity mixes ocean waters, introducing cooling in the upper layer of the ocean that quickly dissipates and warming in deeper layers that lasts considerably more, resulting in a net warming of the ocean.

<span class="mw-page-title-main">Pacific Meridional Mode</span> Climate mode in the North Pacific

Pacific Meridional Mode (PMM) is a climate mode in the North Pacific. In its positive state, it is characterized by the coupling of weaker trade winds in the northeast Pacific Ocean between Hawaii and Baja California with decreased evaporation over the ocean, thus increasing sea surface temperatures (SST); and the reverse during its negative state. This coupling develops during the winter months and spreads southwestward towards the equator and the central and western Pacific during spring, until it reaches the Intertropical Convergence Zone (ITCZ), which tends to shift north in response to a positive PMM.

The recharge oscillator model for El Niño–Southern Oscillation (ENSO) is a theory described for the first time in 1997 by Jin., which explains the periodical variation of the sea surface temperature (SST) and thermocline depth that occurs in the central equatorial Pacific Ocean. The physical mechanisms at the basis of this oscillation are periodical recharges and discharges of the zonal mean equatorial heat content, due to ocean-atmosphere interaction. Other theories have been proposed to model ENSO, such as the delayed oscillator, the western Pacific oscillator and the advective reflective oscillator. A unified and consistent model has been proposed by Wang in 2001, in which the recharge oscillator model is included as a particular case.

References

  1. 1 2 3 Coats & Karnauskas 2018, p. 6245.
  2. 1 2 3 4 Luo et al. 2017, p. 2812.
  3. An et al. 2012, p. 1373.
  4. 1 2 An et al. 2012, p. 1374.
  5. 1 2 McGregor & Timmermann 2011, p. 2178.
  6. Heede, Fedorov & Burls 2020, p. 6101.
  7. Sun & Liu 1996, p. 1148.
  8. Sun & Liu 1996, p. 1149.
  9. 1 2 Clement et al. 1996, p. 2192.
  10. Clement et al. 1996, pp. 2192–2193.
  11. Clement et al. 1996, p. 2193.
  12. 1 2 3 4 An & Im 2014, p. 174.
  13. 1 2 McGregor & Timmermann 2011, p. 2179.
  14. Marchitto et al. 2010, p. 1378.
  15. Liu, Lu & Xie 2015, pp. 1044–1045.
  16. Heede, Fedorov & Burls 2021, p. 2520.
  17. Rodgers et al. 1999, p. 20551.
  18. Rodgers et al. 1999, p. 20567.
  19. An et al. 2022, p. 20.
  20. Hertzberg & Schmidt 2014.
  21. Giralt, Moreno & Bao 2007, p. 350.
  22. Zhao et al. 2016, p. 6820.
  23. Alfaro-Sánchez et al. 2018, p. 936.
  24. McGregor & Timmermann 2011, p. 2187.
  25. Emile-Geay et al. 2011.
  26. Trouet & Taylor 2010, p. 961.
  27. Sun et al. 2022, p. 8.
  28. Jiménez-Moreno, Anderson & Shinker 2021, p. 7.
  29. Marchitto et al. 2010, p. 1380.
  30. Kelly et al. 2016.
  31. Marchitto et al. 2010, pp. 1378–1379.
  32. Grist et al. 2013.
  33. 1 2 Salvatteci et al. 2014, p. 727.
  34. Salvatteci et al. 2014, p. 716.
  35. Metcalfe & Nash 2012, p. 51.
  36. Ying, Huang & Huang 2016, p. 433.
  37. 1 2 3 4 Coats & Karnauskas 2018, p. 6246.
  38. Heede, Fedorov & Burls 2020, p. 6105.
  39. Heede, Fedorov & Burls 2020, p. 6102.
  40. Heede, Fedorov & Burls 2020, p. 6114.
  41. Heede, Fedorov & Burls 2021, p. 2506.
  42. Luo, Wang & Dommenget 2018, p. 1343.
  43. Heede & Fedorov 2021, p. 696.
  44. An & Im 2014, p. 181.
  45. An & Im 2014, p. 182.
  46. Luo et al. 2017, p. 2822.
  47. Luo et al. 2017, p. 2824.
  48. Luo et al. 2017, p. 2826.
  49. Heede, Fedorov & Burls 2020, p. 6108.
  50. Heede, Fedorov & Burls 2020, p. 6109.
  51. Heede, Fedorov & Burls 2020, p. 6113.
  52. Heede & Fedorov 2021, p. 697.
  53. Heede & Fedorov 2021, p. 698.
  54. Zhou et al. 2022, p. 1.
  55. Coats & Karnauskas 2018, p. 6247.
  56. 1 2 Coats & Karnauskas 2018, p. 6248.
  57. Coats & Karnauskas 2018, p. 6258.
  58. Coats & Karnauskas 2018, p. 6259.
  59. Heede, Fedorov & Burls 2020, p. 6103.
  60. Heede, Fedorov & Burls 2020, p. 6106.
  61. Heede, Fedorov & Burls 2020, p. 6107.

Sources