El Niño–Southern Oscillation

From Wikipedia, the free encyclopedia
(Redirected from El Niño)
Impacts of El Niño on climate
Impacts of La Niña on climate
Changes to temperature and precipitation during El Niño (left) and La Niña (right). The top two maps are for Northern hemisphere winter, the bottom two for summer.[1]

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.[2][3] 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.[4] El Niño events can be more intense but La Niña events may repeat and last longer.

A key mechanism of ENSO is the Bjerknes feedback (named after Jacob Bjerknes in 1969) in which the atmospheric changes alter the sea temperatures that in turn alter the atmospheric winds in a positive feedback. Relaxed easterly trade winds result in a surge of warm surface waters to the east and reduced ocean upwelling on the equator.  In turn that leads to warmer sea surface temperatures (El Niño), a weaker Walker circulation (an east-west overturning circulation in the atmosphere) and further relaxed trade winds. Ultimately the warm waters in the western tropical Pacific are depleted enough that conditions return to normal. The exact mechanisms that cause the oscillation remain under study.

Each country that monitors the ENSO has a different threshold for what constitutes an El Niño or La Niña event, which is tailored to their specific interests.[5]

El Niño and La Niña affect the global climate and disrupt normal weather patterns, which as a result can lead to intense storms in some places and droughts in others.[6][7] El Niño events cause short-term (approximately 1 year in length) spikes in global average surface temperature while La Niña events cause short term surface cooling.[8] Therefore, the relative frequency of El Niño compared to La Niña events can affect global temperature trends on decadal timescales.[9] Developing countries dependent upon agriculture and fishing, particularly those bordering the Pacific Ocean, are the most affected.

In climate change science, ENSO is known as one of the internal climate variability phenomena.[10]: 23  Future trends in ENSO due to climate change are uncertain,[11] although climate change exacerbates the effects of droughts and floods. The IPCC Sixth Assessment Report summarized the state of the art of research in 2021 into the future of ENSO as follows: "In the long term, it is very likely that the precipitation variance related to El Niño–Southern Oscillation will increase"[10]: 113  and "It is very likely that rainfall variability related to changes in the strength and spatial extent of ENSO teleconnections will lead to significant changes at regional scale".[10]: 114 

Definition and terminology[edit]

Southern Oscillation Index timeseries from 1876 to 2023. The Southern Oscillation is the atmospheric component of El Niño. This component is an oscillation in surface air pressure between the tropical eastern and the western Pacific Ocean waters.

The El Niño–Southern Oscillation is a single climate phenomenon that periodically fluctuates between three phases: Neutral, La Niña or El Niño.[12] La Niña and El Niño are opposite phases in the oscillation which are deemed to occur when specific ocean and atmospheric conditions are reached or exceeded.[12]

An early recorded mention of the term "El Niño" ("The Boy" in Spanish) to refer to climate occurred in 1892, when Captain Camilo Carrillo told the geographical society congress in Lima that Peruvian sailors named the warm south-flowing current "El Niño" because it was most noticeable around Christmas.[13]

The capitalized term El Niño refers to the Christ child, Jesus, because periodic warming in the Pacific near South America is usually noticed around Christmas.[14]

Originally, the term El Niño applied to an annual weak warm ocean current that ran southwards along the coast of Peru and Ecuador at about Christmas time.[15] However, over time the term has evolved and now refers to the warm and negative phase of the El Niño–Southern Oscillation (ENSO). The original phrase, El Niño de Navidad, arose centuries ago, when Peruvian fishermen named the weather phenomenon after the newborn Christ.[16][17]

La Niña ("The Girl" in Spanish) is the colder counterpart of El Niño, as part of the broader ENSO climate pattern. In the past, it was also called an anti-El Niño[18] and El Viejo, meaning "the old man."[19]

A negative phase exists when atmospheric pressure over Indonesia and the west Pacific is abnormally high and pressure over the east Pacific is abnormally low, during El Niño episodes, and a positive phase is when the opposite occurs during La Niña episodes, and pressure over Indonesia is low and over the west Pacific is high.[20]

Fundamentals[edit]

Diagram showing a cross-section of the Pacific and related phenomena
The West Pacific is typically warmer than the East Pacific. The warmer waters lead to more cloudiness, rainfall, and low air pressure over the West Pacific. The buildup of warm waters towards the west also leads to a thicker layer of warm ocean water that lowers the depth of the thermocline.

On average, the temperature of the ocean surface in the tropical East Pacific is roughly 8–10 °C (14–18 °F) cooler than in the tropical West Pacific. The sea surface temperature (SST) of the West Pacific northeast of Australia averages around 28–30 °C (82–86 °F). SSTs in the East Pacific off the western coast of South America are closer to 20 °C (68 °F). Strong trade winds near the equator push water away from the East Pacific and towards the West Pacific.[21][a] This water is slowly warmed by the Sun as it moves west along the equator.[22] The ocean surface near Indonesia is typically around 0.5 m (1.5 ft) higher than near Peru because of the buildup of water in the West Pacific.[23][clarification needed] The thermocline, or the transitional zone between the warmer waters near the ocean surface and the cooler waters of the deep ocean,[24] is pushed downwards in the West Pacific due to this water accumulation.[23][b] Consequently, the thermocline is tilted across the tropical Pacific, rising from an average depth of about 140 m (450 ft) in the West Pacific to a depth of about 30 m (90 ft) in the East Pacific.[23]

Cooler deep ocean water takes the place of the outgoing surface waters in the East Pacific, rising to the ocean surface in a process called upwelling.[21][22][a] This process cools the East Pacific because the thermocline is closer to the ocean surface, leaving relatively little separation between the deeper cold water and the ocean surface.[23] Additionally, the northward-flowing Humboldt Current carries colder water from the Southern Ocean to the tropics in the East Pacific.[21] The combination of the Humboldt Current and upwelling maintains an area of cooler ocean waters off the coast of Peru.[21][22] The West Pacific lacks a cold ocean current and has less upwelling as the trade winds are usually weaker than in the East Pacific, allowing the West Pacific to reach warmer temperatures. These warmer waters provide energy for the upward movement of air. As a result, the warm West Pacific has on average more cloudiness and rainfall than the cool East Pacific.[21]

ENSO describes a quasi-periodic change of both oceanic and atmospheric conditions over the tropical Pacific Ocean.[21] These changes affect weather patterns across much of the Earth.[22] The tropical Pacific is said to be in one of three states of ENSO (also called "phases") depending on the atmospheric and oceanic conditions.[27] When the tropical Pacific roughly reflects the average conditions, the state of ENSO is said to be in the neutral phase. However, the tropical Pacific experiences occasional shifts away from these average conditions. If trade winds are weaker than average, the effect of upwelling in the East Pacific and the flow of warmer ocean surface waters towards the West Pacific lessen. This results in a cooler West Pacific and a warmer East Pacific, leading to a shift of cloudiness and rainfall towards the East Pacific. This situation is called El Niño. The opposite occurs if trade winds are stronger than average, leading to a warmer West Pacific and an cooler East Pacific. This situation is called La Niña and is associated with increased cloudiness and rainfall over the West Pacific.[21]

Bjerknes feedback[edit]

The close relationship between ocean temperatures and the strength of the trade winds was first identified by Jacob Bjerknes in 1969. Bjerknes also hypothesized that ENSO was a positive feedback system where the associated changes in one component of the climate system (the ocean or atmosphere) tend to reinforce changes in the other.[28]: 86  For example, during El Niño, the reduced contrast in ocean temperatures across the Pacific results in weaker trade winds, further reinforcing the El Niño state. This process is known as Bjerknes feedback.[29] Although these associated changes in the ocean and atmosphere often occur together, the state of the atmosphere may resemble a different ENSO phase than the state of the ocean or vice versa.[27] Because their states are closely linked, the variations of ENSO may arise from changes in both the ocean and atmosphere and not necessarily from an initial change of exclusively one or the other.[30][29] Conceptual models explaining how ENSO operates generally accept the Bjerknes feedback hypothesis. However, ENSO would perpetually remain in one phase if Bjerknes feedback were the only process occurring.[28]: 88  Several theories have been proposed to explain how ENSO can change from one state to the next, despite the positive feedback.[31] These explanations broadly fall under two categories.[32] In one view, the Bjerknes feedback naturally triggers negative feedbacks[clarification needed] that end and reverse the abnormal state of the tropical Pacific. This perspective implies that the processes that lead to El Niño and La Niña also eventually bring about their end, making ENSO a self-sustaining[clarification needed] process.[28]: 88  Other theories view the state of ENSO as being changed by irregular and external phenomena such as the Madden–Julian oscillation, tropical instability waves, and westerly wind bursts.[28]: 90 

Walker circulation[edit]

The three phases of ENSO relate to the Walker circulation, which was named after Gilbert Walker who discovered the Southern Oscillation during the early twentieth century. The Walker circulation is an east-west overturning circulation in the vicinity of the equator in the Pacific. Upward air is associated with high sea temperatures, convection and rainfall, while the downward branch occurs over cooler sea surface temperatures in the east. During El Niño, as the sea surface temperatures change so does the Walker Circulation. Warming in the eastern tropical Pacific weakens or reverses the downward branch, while cooler conditions in the west lead to less rain and downward air, so the Walker Circulation first weakens and may reverse.[33]: 185   

Southern Oscillation[edit]

The regions where the air pressure are measured and compared to generate the Southern Oscillation Index
Southern Oscillation Index correlated with mean sea level pressure.

The Southern Oscillation is the atmospheric component of ENSO. This component is an oscillation in surface air pressure between the tropical eastern and the western Pacific Ocean waters. The strength of the Southern Oscillation is measured by the Southern Oscillation Index (SOI). The SOI is computed from fluctuations in the surface air pressure difference between Tahiti (in the Pacific) and Darwin, Australia (on the Indian Ocean).[34]

El Niño episodes have negative SOI, meaning there is lower pressure over Tahiti and higher pressure in Darwin. La Niña episodes on the other hand have positive SOI, meaning there is higher pressure in Tahiti and lower in Darwin.

Low atmospheric pressure tends to occur over warm water and high pressure occurs over cold water, in part because of deep convection over the warm water. El Niño episodes are defined as sustained warming of the central and eastern tropical Pacific Ocean, thus resulting in a decrease in the strength of the Pacific trade winds, and a reduction in rainfall over eastern and northern Australia. La Niña episodes are defined as sustained cooling of the central and eastern tropical Pacific Ocean, thus resulting in an increase in the strength of the Pacific trade winds, and the opposite effects in Australia when compared to El Niño.

Although the Southern Oscillation Index has a long station record going back to the 1800s, its reliability is limited due to the latitudes of both Darwin and Tahiti being well south of the Equator, so that the surface air pressure at both locations is less directly related to ENSO.[35] To overcome this effect, a new index was created, named the Equatorial Southern Oscillation Index (EQSOI).[35][36] To generate this index, two new regions, centered on the Equator, were defined. The western region is located over Indonesia and the eastern one over the equatorial Pacific, close to the South American coast.[35] However, data on EQSOI goes back only to 1949.[35]

Three phases of sea surface temperature[edit]

The El Niño–Southern Oscillation is a single climate phenomenon that quasi-periodically fluctuates between three phases: Neutral, La Niña or El Niño.[12] La Niña and El Niño are opposite phases which require certain changes to take place in both the ocean and the atmosphere before an event is declared.[12] The cool phase of ENSO is La Niña, with SST in the eastern Pacific below average, and air pressure high in the eastern Pacific and low in the western Pacific. The ENSO cycle, including both El Niño and La Niña, causes global changes in temperature and rainfall.[37][38]

Neutral phase: Equatorial winds gather warm water pool toward the west. Warm pool in the west drives deep atmospheric convection. In the east local winds cause nutrient-rich cold water to upwell at the Equator and along the South American coast.
El Niño phase: Warm water pool approaches the South American coast. The absence of cold upwelling increases warming. Warm water and atmospheric convection move eastwards. In strong El Niños the deeper thermocline off South America means upwelled water is warm and nutrient poor.
La Niña phase: Warm water is farther west than usual.

Neutral phase[edit]

If the temperature variation from climatology is within 0.5 °C (0.9 °F), ENSO conditions are described as neutral. Neutral conditions are the transition between warm and cold phases of ENSO. Sea surface temperatures (by definition), tropical precipitation, and wind patterns are near average conditions during this phase.[39] Close to half of all years are within neutral periods.[40] During the neutral ENSO phase, other climate anomalies/patterns such as the sign of the North Atlantic Oscillation or the Pacific–North American teleconnection pattern exert more influence.[41]

El Niño phase[edit]

Loop of the 1997–98 El Niño event showing extreme sea surface temperature (SST) anomalies in the east tropical Pacific

El Niño conditions are established when the Walker circulation weakens or reverses and the Hadley circulation strengthens,[citation needed][clarification needed] leading to the developement of a band of warm ocean water in the central and east-central equatorial Pacific (approximately between the International Date Line and 120°W), including the area off the west coast of South America,[42][43] as upwelling of cold water occurs less or not at all offshore.[3]

This warming causes a shift in the atmospheric circulation, leading to higher air pressure in the western Pacific and lower in the eastern Pacific,[44] with rainfall reducing over Indonesia, India and northern Australia, while rainfall and tropical cyclone formation increases over the tropical Pacific Ocean.[45] The low-level surface trade winds, which normally blow from east to west along the equator, either weaken or start blowing from the other direction.[43]

El Niño phases are known to happen at irregular intervals of two to seven years, and lasts nine months to two years.[46] The average period length is five years. When this warming occurs for seven to nine months, it is classified as El Niño "conditions"; when its duration is longer, it is classified as an El Niño "episode".[47]

Timeline of El Niño episodes between 1900 and 2023.[48][49]

It is thought that there have been at least 30 El Niño events since 1900, with the 1982–83, 1997–98 and 2014–16 events among the strongest on record.[50] Since 2000, El Niño events have been observed in 2002–03, 2004–05, 2006–07, 2009–10, 2014–16, 2018–19,[51][52][53] and 2023–24.[54][55]

Major ENSO events were recorded in the years 1790–93, 1828, 1876–78, 1891, 1925–26, 1972–73, 1982–83, 1997–98, 2014–16, and 2023–24.[56][57][58] During strong El Niño episodes, a secondary peak in sea surface temperature across the far eastern equatorial Pacific Ocean sometimes follows the initial peak.[59]

La Niña phase[edit]

Sea surface temperature anomalies in November 2007, showing La Niña conditions

An especially strong Walker circulation causes La Niña, which is considered to be the cold oceanic and positive atmospheric phase of the broader El Niño–Southern Oscillation (ENSO) weather phenomenon, as well as the opposite of El Niño weather pattern,[18] where sea surface temperature across the eastern equatorial part of the central Pacific Ocean will be lower than normal by 3–5 °C (5.4–9 °F). The phenomenon occurs as strong winds blow warm water at the ocean's surface away from South America, across the Pacific Ocean towards Indonesia.[18] As this warm water moves west, cold water from the deep sea rises to the surface near South America.[18]

The movement of so much heat across a quarter of the planet, and particularly in the form of temperature at the ocean surface, can have a significant effect on weather across the entire planet. Tropical instability waves visible on sea surface temperature maps, showing a tongue of colder water, are often present during neutral or La Niña conditions.[60]

La Niña is a complex weather pattern that occurs every few years,[18] often persisting for longer than five months. El Niño and La Niña can be indicators of weather changes across the globe. Atlantic and Pacific hurricanes can have different characteristics due to lower or higher wind shear and cooler or warmer sea surface temperatures.

A timeline of all La Niña episodes between 1900 and 2023.[61][62] Note that each forecast agency has a different criteria for what constitutes a La Niña event, which is tailored to their specific interests.

La Niña events have been observed for hundreds of years, and occurred on a regular basis during the early parts of both the 17th and 19th centuries.[63] Since the start of the 20th century, La Niña events have occurred during the following years:[64]

  1. 1903–04
  2. 1906–07
  3. 1909–11
  4. 1916–18
  5. 1924–25
  6. 1928–30
  7. 1938–39
  8. 1942–43
  9. 1949–51
  10. 1954–57
  11. 1964–65
  12. 1970–72
  13. 1973–76
  14. 1983–85
  15. 1988–89
  16. 1995–96
  17. 1998–2001
  18. 2005–06
  19. 2007–08
  20. 2008–09
  21. 2010–12
  22. 2016
  23. 2017–18
  24. 2020–23

Transitional phases[edit]

Transitional phases at the onset or departure of El Niño or La Niña can also be important factors on global weather by affecting teleconnections. Significant episodes, known as Trans-Niño, are measured by the Trans-Niño index (TNI).[65] Examples of affected short-time climate in North America include precipitation in the Northwest US[66] and intense tornado activity in the contiguous US.[67]

Variations[edit]

ENSO Modoki[edit]

Map showing Niño/Niña 1 to 4 regions, 3 and 4 being west and far west and much larger than 1 and 2 a coastal Peruvian/Ecuadorian zone differing subtly north–south

The first ENSO pattern to be recognised, called Eastern Pacific (EP) ENSO, to distinguish if from others,[68] involves temperature anomalies in the eastern Pacific. However, in the 1990s and 2000s, variations of ENSO conditions were observed, in which the usual place of the temperature anomaly (Niño 1 and 2) is not affected, but an anomaly also arises in the central Pacific (Niño 3.4).[69] The phenomenon is called Central Pacific (CP) ENSO,[68] "dateline" ENSO (because the anomaly arises near the dateline), or ENSO "Modoki" (Modoki is Japanese for "similar, but different").[70][71] There are variations of ENSO additional to the EP and CP types, and some scientists argue that ENSO exists as a continuum, often with hybrid types.[72]

The effects of the CP ENSO are different from those of the EP ENSO. The El Niño Modoki is associated with more hurricanes more frequently making landfall in the Atlantic.[73] La Niña Modoki leads to a rainfall increase over northwestern Australia and northern Murray–Darling basin, rather than over the eastern portion of the country as in a conventional EP La Niña.[74] Also, La Niña Modoki increases the frequency of cyclonic storms over Bay of Bengal, but decreases the occurrence of severe storms in the Indian Ocean overall.[75]

The first recorded El Niño that originated in the central Pacific and moved toward the east was in 1986.[76] Recent Central Pacific El Niños happened in 1986–87, 1991–92, 1994–95, 2002–03, 2004–05 and 2009–10.[77] Furthermore, there were "Modoki" events in 1957–59,[78] 1963–64, 1965–66, 1968–70, 1977–78 and 1979–80.[79][80] Some sources say that the El Niños of 2006-07 and 2014-16 were also Central Pacific El Niños.[81][82] Recent years when La Niña Modoki events occurred include 1973–1974, 1975–1976, 1983–1984, 1988–1989, 1998–1999, 2000–2001, 2008–2009, 2010–2011, and 2016–2017.[83][84][85]

The recent discovery of ENSO Modoki has some scientists believing it to be linked to global warming.[86] However, comprehensive satellite data go back only to 1979. More research must be done to find the correlation and study past El Niño episodes. More generally, there is no scientific consensus on how/if climate change might affect ENSO.[11]

There is also a scientific debate on the very existence of this "new" ENSO. A number of studies dispute the reality of this statistical distinction or its increasing occurrence, or both, either arguing the reliable record is too short to detect such a distinction,[87][88] finding no distinction or trend using other statistical approaches,[89][90][91][92][93] or that other types should be distinguished, such as standard and extreme ENSO.[94][95]

Likewise, following the asymmetric nature of the warm and cold phases of ENSO, some studies could not identify similar variations for La Niña, both in observations and in the climate models,[96] but some sources could identify variations on La Niña with cooler waters on central Pacific and average or warmer water temperatures on both eastern and western Pacific, also showing eastern Pacific Ocean currents going to the opposite direction compared to the currents in traditional La Niñas.[70][71][97]

ENSO Costero[edit]

Coined by the Peruvian Comité Multisectorial Encargado del Estudio Nacional del Fenómeno El Niño (ENFEN),[98] ENSO Costero, or ENSO Oriental, is the name given to the phenomenon where the sea-surface temperature anomalies are mostly focused on the South American coastline, especially from Peru and Ecuador.[99] Studies point many factors that can lead to its occurrence,[100] sometimes accompanying, or being accompanied, by a larger EP ENSO occurrence,[99] or even displaying opposite conditions from the observed ones in the other Niño regions when accompanied by Modoki variations.[101]

ENSO Costero events usually present more localized effects, with warm phases leading to increased rainfall over the coast of Ecuador, northern Peru and the Amazon rainforest, and increased temperatures over the northern Chilean coast,[98][102] and cold phases leading to droughts on the peruvian coast, and increased rainfall and decreased temperatures on its mountainous and jungle regions.[103]

Because they don't influence the global climate as much as the other types, these events present lesser and weaker correlations to other significant ENSO features, neither always being triggered by Kelvin waves,[98] nor always being accompanied by proportional Southern Oscillation responses.[104] According to the Coastal Niño Index (ICEN), strong El Niño Costero events include 1957, 1982–83, 1997–98 and 2015–16, and La Niña Costera ones include 1950, 1954–56, 1962, 1964, 1966, 1967–68, 1970–71, 1975–76 and 2013.[105]

Monitoring and declaration of conditions[edit]

The various "Niño regions" where sea surface temperatures are monitored to determine the current ENSO phase (warm or cold)

Currently, each country has a different threshold for what constitutes an El Niño event, which is tailored to their specific interests, for example:[5]

  • In the United States, its Climate Prediction Center and the International Research Institute for Climate and Society monitors the sea surface temperatures in the Niño 3.4 region, the tropical Pacific atmosphere and forecasts that NOAA's Oceanic Niño Index will equal or exceed .5 °C (0.90 °F) for several seasons in a row.[106] The Niño 3.4 region stretches from the 120th to 170th meridians west longitude astride the equator five degrees of latitude on either side, are monitored. It is approximately 3,000 kilometres (1,900 mi) to the southeast of Hawaii. The most recent three-month average for the area is computed, and if the region is more than 0.5 °C (0.9 °F) above (or below) normal for that period, then an El Niño (or La Niña) is considered in progress.[107]
  • The Australian Bureau of Meteorology looks at the trade winds, Southern Oscillation Index, weather models and sea surface temperatures in the Niño 3 and 3.4 regions, before declaring an ENSO event.[108]
  • The Japan Meteorological Agency declares that an ENSO event has started when the average five month sea surface temperature deviation for the Niño 3 region is over 0.5 °C (0.90 °F) for six consecutive months or longer.[109]
  • The Peruvian government declares that a ENSO Costero is under way if the sea surface temperature deviation in the Niño 1+2 regions equal or exceed 0.4 °C (0.72 °F) for at least three months.[105]
  • The United Kingdom's Met Office also uses a several month period to determine ENSO state.[110] When this warming or cooling occurs for only seven to nine months, it is classified as El Niño/La Niña "conditions"; when it occurs for more than that period, it is classified as El Niño/La Niña "episodes".[111]

Effects of ENSO on global climate[edit]

Refer to caption
This image shows three examples of internal climate variability measured between 1950 and 2012: the El Niño–Southern oscillation, the Arctic oscillation, and the North Atlantic oscillation.[112]

In climate change science, ENSO is known as one of the internal[clarification needed] climate variability phenomena. The other two main ones[clarification needed] are Pacific decadal oscillation and Atlantic multidecadal oscillation.[10]: 23 

La Niña impacts the global climate and disrupts normal weather patterns, which can lead to intense storms in some places and droughts in others.[113] El Niño events cause short-term (approximately 1 year in length) spikes in global average surface temperature while La Niña events cause short term cooling.[8] Therefore, the relative frequency of El Niño compared to La Niña events can affect global temperature trends on decadal timescales.[9]

Climate change[edit]

There is no sign that there are actual changes in the ENSO physical phenomenon due to climate change. Climate models do not simulate ENSO well enough to make reliable predictions. Future trends in ENSO are uncertain[11] as different models make different predictions.[114][115] It may be that the observed phenomenon of more frequent and stronger El Niño events occurs only in the initial phase of the global warming, and then (e.g., after the lower layers of the ocean get warmer, as well), El Niño will become weaker.[116] It may also be that the stabilizing and destabilizing forces influencing the phenomenon[clarification needed] will eventually compensate for each other.[117]

The consequences of ENSO in terms of the temperature anomalies and precipitation and weather extremes around the world are clearly increasing and associated with climate change. For example, recent scholarship (since about 2019) has found that climate change is increasing the frequency of extreme El Niño events.[118][119][120] Previously there was no consensus on whether climate change will have any influence on the strength or duration of El Niño events, as research alternately supported El Niño events becoming stronger and weaker, longer and shorter.[121][122]

Over the last several decades, the number of El Niño events increased, and the number of La Niña events decreased,[123] although observation of ENSO for much longer is needed to detect robust changes.[124]

Studies of historical data show the recent El Niño variation is most likely linked to global warming. For example, some results, even after subtracting the positive influence of decadal variation, are shown to be possibly present in the ENSO trend,[125] the amplitude of the ENSO variability in the observed data still increases, by as much as 60% in the last 50 years.[126] A study published in 2023 by CSIRO researchers found that climate change may have increased by two times the likelihood of strong El Niño events and nine times the likelihood of strong La Niña events.[127][128] The study stated it found a consensus between different models and experiments.[129]

The IPCC Sixth Assessment Report summarized the state of the art of research in 2021 into the future of ENSO as follows:

  • "In the long term, it is very likely that the precipitation variance related to El Niño–Southern Oscillation will increase"[10]: 113  and
  • "It is very likely that rainfall variability related to changes in the strength and spatial extent of ENSO teleconnections will lead to significant changes at regional scale".[10]: 114  and
  • "There is medium confidence that both ENSO amplitude and the frequency of high-magnitude events since 1950 are higher than over the period from 1850 and possibly as far back as 1400".[10]: 373 

Investigations regarding tipping points[edit]

The ENSO is considered to be a potential tipping element in Earth's climate[130] and, under the global warming, can enhance or alternate regional climate extreme events through a strengthened teleconnection.[131] For example, an increase in the frequency and magnitude of El Niño events have triggered warmer than usual temperatures over the Indian Ocean, by modulating the Walker circulation.[132] This has resulted in a rapid warming of the Indian Ocean, and consequently a weakening of the Asian Monsoon.[133]

Earlier (2008) list of tipping elements in the climate system.[134] When compared to later lists, the major differences are that in 2008 ENSO, Indian summer monsoon, Arctic ozone hole and all of Arctic sea ice were all listed as tipping points. Labrador-Irminger circulation, mountain glaciers and East Antarctic ice however were not included. This 2008 list also includes Antarctic bottom water (part of the Southern Ocean overturning circulation), which was left out of the 2022 list, but included in some subsequent ones.
The possibility that the El Niño–Southern Oscillation (ENSO) is a tipping element had attracted attention in the past.[135] Normally strong winds blow west across the South Pacific Ocean from South America to Australia. Every two to seven years, the winds weaken due to pressure changes and the air and water in the middle of the Pacific warms up, causing changes in wind movement patterns around the globe. This is known as El Niño and typically leads to droughts in India, Indonesia and Brazil, and increased flooding in Peru. In 2015/2016, this caused food shortages affecting over 60 million people.[136] El Niño-induced droughts may increase the likelihood of forest fires in the Amazon.[137] The threshold for tipping was estimated to be between 3.5 °C (6.3 °F) and 7 °C (13 °F) of global warming in 2016.[138] After tipping, the system would be in a more permanent El Niño state, rather than oscillating between different states. This has happened in Earth's past, in the Pliocene, but the layout of the ocean was significantly different from now.[135] So far, there is no definitive evidence indicating changes in ENSO behaviour,[137] and the IPCC Sixth Assessment Report concluded that it is "virtually certain that the ENSO will remain the dominant mode of interannual variability in a warmer world."[139] Consequently, the 2022 assessment no longer includes it in the list of likely tipping elements.[140]

Effects of ENSO on weather patterns[edit]

Colored bars show how El Niño years (red, regional warming) and La Niña years (blue, regional cooling) relate to overall global warming. The El Niño–Southern Oscillation has been linked to variability in longer-term global average temperature increase, with El Niño years usually corresponding to annual global temperature increases.
2023's June-July-August season was the warmest on record globally by a large margin, as El Niño conditions continued to develop.[141] 1998—a very strong El Niño year—also experienced a global temperature spike.

El Niño affects the global climate and disrupts normal weather patterns, which can lead to intense storms in some places and droughts in others.[6][7]

Tropical cyclones[edit]

Most tropical cyclones form on the side of the subtropical ridge closer to the equator, then move poleward past the ridge axis before recurving into the main belt of the Westerlies.[142] Areas west of Japan and Korea tend to experience many fewer September–November tropical cyclone impacts during El Niño and neutral years. During El Niño years, the break[clarification needed] in the subtropical ridge tends to lie near 130°E, which would favor the Japanese archipelago.[143]

Based on modeled and observed accumulated cyclone energy (ACE), El Niño years usually result in less active hurricane seasons in the Atlantic Ocean, but instead favor a shift to tropical cyclone activity in the Pacific Ocean, compared to La Niña years favoring above average hurricane development in the Atlantic and less so in the Pacific basin.[144]

Over the Atlantic Ocean, vertical wind shear is increased, which inhibits tropical cyclone genesis and intensification, by causing the westerly winds to be stronger.[145] The atmosphere over the Atlantic Ocean can also be drier and more stable during El Niño events, which can inhibit tropical cyclone genesis and intensification.[145] Within the Eastern Pacific basin: El Niño events contribute to decreased easterly vertical wind shear and favor above-normal hurricane activity.[146] However, the impacts of the ENSO state in this region can vary and are strongly influenced by background climate patterns.[146] The Western Pacific basin experiences a change in the location of where tropical cyclones form during El Niño events, with tropical cyclone formation shifting eastward, without a major change in how many develop each year.[145] As a result of this change, Micronesia is more likely, and China less likely, to be affected by tropical cyclones.[143] A change in the location of where tropical cyclones form also occurs within the Southern Pacific Ocean between 135°E and 120°W, with tropical cyclones more likely to occur within the Southern Pacific basin than the Australian region.[147][145] As a result of this change tropical cyclones are 50% less likely to make landfall on Queensland, while the risk of a tropical cyclone is elevated for island nations like Niue, French Polynesia, Tonga, Tuvalu, and the Cook Islands.[147][148][149]

Remote influence on tropical Atlantic Ocean[edit]

A study of climate records has shown that El Niño events in the equatorial Pacific are generally associated with a warm tropical North Atlantic in the following spring and summer.[150] About half of El Niño events persist sufficiently into the spring months for the Western Hemisphere Warm Pool to become unusually large in summer.[151] Occasionally, El Niño's effect on the Atlantic Walker circulation over South America strengthens the easterly trade winds in the western equatorial Atlantic region. As a result, an unusual cooling may occur in the eastern equatorial Atlantic in spring and summer following El Niño peaks in winter.[152] Cases of El Niño-type events in both oceans simultaneously have been linked to severe famines related to the extended failure of monsoon rains.[153]

Impacts on humans and ecosystems[edit]

Economic impacts[edit]

El Niño has the most direct impacts on life in the equatorial Pacific, its effects propagate north and south along the coast of the Americas, affecting marine life all around the Pacific. Changes in chlorophyll-a concentrations are visible in this animation, which compares phytoplankton in January and July 1998. Since then, scientists have improved both the collection and presentation of chlorophyll data.[clarification needed]

When El Niño conditions last for many months, extensive ocean warming and the reduction in easterly trade winds limits upwelling of cold nutrient-rich deep water, and its economic effect on local fishing for an international market can be serious.[154] Developing countries that depend on their own agriculture and fishing, particularly those bordering the Pacific Ocean, are usually most affected by El Niño conditions. In this phase of the Oscillation, the pool of warm water in the Pacific near South America is often at its warmest in late December.[155]

More generally, El Niño can affect commodity prices and the macroeconomy of different countries. It can constrain the supply of rain-driven agricultural commodities; reduce agricultural output, construction, and services activities; increase food prices; and may trigger social unrest in commodity-dependent poor countries that primarily rely on imported food.[156] A University of Cambridge Working Paper shows that while Australia, Chile, Indonesia, India, Japan, New Zealand and South Africa face a short-lived fall in economic activity in response to an El Niño shock, other countries may actually benefit from an El Niño weather shock (either directly or indirectly through positive spillovers from major trading partners), for instance, Argentina, Canada, Mexico and the United States. Furthermore, most countries experience short-run inflationary pressures following an El Niño shock, while global energy and non-fuel commodity prices increase.[157] The IMF estimates a significant El Niño can boost the GDP of the United States by about 0.5% (due largely to lower heating bills) and reduce the GDP of Indonesia by about 1.0%.[158]

Health and social impacts[edit]

Extreme weather conditions related to the El Niño cycle correlate with changes in the incidence of epidemic diseases. For example, the El Niño cycle is associated with increased risks of some of the diseases transmitted by mosquitoes, such as malaria, dengue fever, and Rift Valley fever.[159] Cycles of malaria in India, Venezuela, Brazil, and Colombia have now been linked to El Niño. Outbreaks of another mosquito-transmitted disease, Australian encephalitis (Murray Valley encephalitis—MVE), occur in temperate south-east Australia after heavy rainfall and flooding, which are associated with La Niña events. A severe outbreak of Rift Valley fever occurred after extreme rainfall in north-eastern Kenya and southern Somalia during the 1997–98 El Niño.[160]

ENSO conditions have also been related to Kawasaki disease incidence in Japan and the west coast of the United States,[161] via the linkage to tropospheric winds across the north Pacific Ocean.[162]

ENSO may be linked to civil conflicts. Scientists at The Earth Institute of Columbia University, having analyzed data from 1950 to 2004, suggest ENSO may have had a role in 21% of all civil conflicts since 1950, with the risk of annual civil conflict doubling from 3% to 6% in countries affected by ENSO during El Niño years relative to La Niña years.[163][164]

Ecological consequences[edit]

During the 1982–83, 1997–98 and 2015–16 ENSO events, large extensions of tropical forests experienced a prolonged dry period that resulted in widespread fires, and drastic changes in forest structure and tree species composition in Amazonian and Bornean forests. Their impacts do not restrict only vegetation, since declines in insect populations were observed after extreme drought and terrible fires during El Niño 2015–16.[165] Declines in habitat-specialist and disturbance-sensitive bird species and in large-frugivorous mammals were also observed in Amazonian burned forests, while temporary extirpation of more than 100 lowland butterfly species occurred at a burned forest site in Borneo.

In seasonally dry tropical forests, which are more drought tolerant, researchers found that El Niño induced drought increased seedling mortality. In a research published in October 2022, researchers studied seasonally dry tropical forests in a national park in Chiang Mai in Thailand for 7 years and observed that El Niño increased seedling mortality even in seasonally dry tropical forests and may impact entire forests in long run.[166]

Coral bleaching[edit]

Following the El Nino event in 1997 – 1998, the Pacific Marine Environmental Laboratory attributes the first large-scale coral bleaching event to the warming waters.[167]

Most critically, global mass bleaching events were recorded in 1997-98 and 2015–16, when around 75-99% losses of live coral were registered across the world. Considerable attention was also given to the collapse of Peruvian and Chilean anchovy populations that led to a severe fishery crisis following the ENSO events in 1972–73, 1982–83, 1997–98 and, more recently, in 2015–16. In particular, increased surface seawater temperatures in 1982-83 also lead to the probable extinction of two hydrocoral species in Panamá, and to a massive mortality of kelp beds along 600 km of coastline in Chile, from which kelps and associated biodiversity slowly recovered in the most affected areas even after 20 years. All these findings enlarge the role of ENSO events as a strong climatic force driving ecological changes all around the world – particularly in tropical forests and coral reefs.[168]

Impacts by region[edit]

Observations of ENSO events since 1950 show that impacts associated with such events depend on the time of year.[169] While certain events and impacts are expected to occur, it is not certain that they will happen.[169] The impacts that generally do occur during most El Niño events include below-average rainfall over Indonesia and northern South America, and above average rainfall in southeastern South America, eastern equatorial Africa, and the southern United States.[169]

Africa[edit]

Between 50,000 and 100,000 people died during the 2011 East Africa drought.[170]

La Niña results in wetter-than-normal conditions in southern Africa from December to February, and drier-than-normal conditions over equatorial east Africa over the same period.[171]

The effects of El Niño on rainfall in southern Africa differ between the summer and winter rainfall areas. Winter rainfall areas tend to get higher rainfall than normal and summer rainfall areas tend to get less rain. The effect on the summer rainfall areas is stronger and has led to severe drought in strong El Niño events.[172][173]

Sea surface temperatures off the west and south coasts of South Africa are affected by ENSO via changes in surface wind strength.[174] During El Niño the south-easterly winds driving upwelling are weaker which results in warmer coastal waters than normal, while during La Niña the same winds are stronger and cause colder coastal waters. These effects on the winds are part of large scale influences on the tropical Atlantic and the South Atlantic High-pressure system, and changes to the pattern of westerly winds further south. There are other influences not known to be related to ENSO of similar importance. Some ENSO events do not lead to the expected changes.[174]

Antarctica[edit]

Many ENSO linkages exist in the high southern latitudes around Antarctica.[175] Specifically, El Niño conditions result in high-pressure anomalies over the Amundsen and Bellingshausen Seas, causing reduced sea ice and increased poleward heat fluxes in these sectors, as well as the Ross Sea. The Weddell Sea, conversely, tends to become colder with more sea ice during El Niño. The exact opposite heating and atmospheric pressure anomalies occur during La Niña.[176] This pattern of variability is known as the Antarctic dipole mode, although the Antarctic response to ENSO forcing is not ubiquitous.[176]

Asia[edit]

In Western Asia, during the region's November–April rainy season, there is increased precipitation in the El Niño phase and reduced precipitation in the La Niña phase on average.[177][178]

During El Niño years: As warm water spreads from the west Pacific and the Indian Ocean to the east Pacific, it takes the rain with it, causing extensive drought in the western Pacific and rainfall in the normally dry eastern Pacific. Singapore experienced the driest February in 2010 since records began in 1869, with only 6.3 mm of rain falling in the month. The years 1968 and 2005 had the next driest Februaries, when 8.4 mm of rain fell.[179]

During La Niña years, the formation of tropical cyclones, along with the subtropical ridge position, shifts westward across the western Pacific Ocean, which increases the landfall threat in China.[180] In March 2008, La Niña caused a drop in sea surface temperatures over Southeast Asia by 2 °C (3.6 °F). It also caused heavy rains over the Philippines, Indonesia, and Malaysia.[181]

Australia[edit]

Across most of the continent, El Niño and La Niña have more impact on climate variability than any other factor. There is a strong correlation between the strength of La Niña and rainfall: the greater the sea surface temperature and Southern Oscillation difference from normal, the larger the rainfall change.[182]

During El Niño events, the shift in rainfall away from the Western Pacific may mean that rainfall across Australia is reduced.[183] Over the southern part of the continent, warmer than average temperatures can be recorded as weather systems are more mobile and fewer blocking areas of high pressure occur.[183] The onset of the Indo-Australian Monsoon in tropical Australia is delayed by two to six weeks, which as a consequence means that rainfall is reduced over the northern tropics.[183] The risk of a significant bushfire season in south-eastern Australia is higher following an El Niño event, especially when it is combined with a positive Indian Ocean Dipole event.[183]

Effects of the El Niño–Southern Oscillation in Australia are present across most of Australia, particularly the north and the east, and are one of the main climate drivers of the country. Associated with seasonal abnormality in many areas in the world, Australia is one of the continents most affected and experiences extensive droughts alongside considerable wet periods that cause major floods. There exist three phases — El Niño, La Niña, and Neutral, which help to account for the different states of ENSO.[184] Since 1900, there have been 28 El Niño and 19 La Niña events in Australia including the current 2023 El Niño event, which was declared on 17th of September in 2023.[185][186][187][188] The events usually last for 9 to 12 months, but some can persist for two years, though the ENSO cycle generally operates over a time period from one to eight years.[189]

Through La Niña years the eastern seaboard of Australia records above-average rainfall usually creating damaging floods due to stronger easterly trade winds from the Pacific towards Australia, thus increasing moisture in the country. Conversely, El Niño events will be associated with a weakening, or even a setback, of the prevailing trade winds, and this, results in reduced atmospheric moisture in the country.[190] Many of the worst bushfires in Australia accompany ENSO events, and can be exacerbated by a positive Indian Ocean Dipole, where they would tend to cause a warm, dry and windy climate.[191]

Europe[edit]

El Niño's effects on Europe are controversial, complex and difficult to analyze, as it is one of several factors that influence the weather over the continent and other factors can overwhelm the signal.[192][193]

North America[edit]

La Niña causes mostly the opposite effects of El Niño: above-average precipitation across the northern Midwest, the northern Rockies, Northern California, and the Pacific Northwest's southern and eastern regions.[194] Meanwhile, precipitation in the southwestern and southeastern states, as well as southern California, is below average.[195] This also allows[clarification needed] for the development of many stronger-than-average hurricanes in the Atlantic and fewer in the Pacific.

ENSO is linked to rainfall over Puerto Rico.[clarification needed][196] During an El Niño, snowfall is greater than average across the southern Rockies and Sierra Nevada mountain range, and is well-below normal across the Upper Midwest and Great Lakes states. During a La Niña, snowfall is above normal across the Pacific Northwest and western Great Lakes.[197]

In Canada, La Niña will, in general, cause a cooler, snowier winter, such as the near-record-breaking amounts of snow recorded in the La Niña winter of 2007–2008 in eastern Canada.[198][199]

In the spring of 2022, La Niña caused above-average precipitation and below-average temperatures in the state of Oregon. April was one of the wettest months on record, and La Niña effects, while less severe, were expected to continue into the summer.[200]

Over North America, the main temperature and precipitation impacts of El Niño generally occur in the six months between October and March.[201][202] In particular, the majority of Canada generally has milder than normal winters and springs, with the exception of eastern Canada where no significant impacts occur.[203] Within the United States, the impacts generally observed during the six-month period include wetter-than-average conditions along the Gulf Coast between Texas and Florida, while drier conditions are observed in Hawaii, the Ohio Valley, Pacific Northwest and the Rocky Mountains.[201]

Study of more recent weather events over California and the southwestern United States indicate that there is a variable relationship between El Niño and above-average precipitation, as it strongly depends on the strength of the El Niño event and other factors.[201] Though it has been historically associated with high rainfall in California, the effects of El Niño depend more strongly on the "flavor"[clarification needed] of El Niño than its presence or absence, as only "persistent El Niño" events lead to consistently high rainfall.[204][205]

To the north across Alaska, La Niña events lead to drier than normal conditions, while El Niño events do not have a correlation towards dry or wet conditions. During El Niño events, increased precipitation is expected in California due to a more southerly, zonal, storm track.[206] During La Niña, increased precipitation is diverted into the Pacific Northwest due to a more northerly storm track.[207] During La Niña events, the storm track shifts far enough northward to bring wetter than normal winter conditions (in the form of increased snowfall) to the Midwestern states, as well as hot and dry summers.[208] During the El Niño portion of ENSO, increased precipitation falls along the Gulf coast and Southeast due to a stronger than normal, and more southerly, polar jet stream.[209]

Isthmus of Tehuantepec[edit]

The synoptic condition for the Tehuantepecer, a violent mountain-gap wind in between the mountains of Mexico and Guatemala, is associated with high-pressure system forming in Sierra Madre of Mexico in the wake of an advancing cold front, which causes winds to accelerate through the Isthmus of Tehuantepec. Tehuantepecers primarily occur during the cold season months for the region in the wake of cold fronts, between October and February, with a summer maximum in July caused by the westward extension of the Azores-Bermuda high pressure system. Wind magnitude is greater during El Niño years than during La Niña years, due to the more frequent cold frontal incursions during El Niño winters.[210] Tehuantepec winds reach 20 knots (40 km/h) to 45 knots (80 km/h), and on rare occasions 100 knots (190 km/h). The wind's direction is from the north to north-northeast.[211] It leads to a localized acceleration of the trade winds in the region, and can enhance thunderstorm activity when it interacts with the Intertropical Convergence Zone.[212] The effects can last from a few hours to six days.[213] Between 1942 and 1957, La Niña had an impact that caused isotope changes in the plants of Baja California, and that had helped scientists to study his impact.[214]

Pacific islands[edit]

During an El Niño event, New Zealand tends to experience stronger or more frequent westerly winds during their summer, which leads to an elevated risk of drier than normal conditions along the east coast.[215] There is more rain than usual though on New Zealand's West Coast, because of the barrier effect of the North Island mountain ranges and the Southern Alps.[215]

Fiji generally experiences drier than normal conditions during an El Niño, which can lead to drought becoming established over the Islands.[216] However, the main impacts on the island nation is felt about a year after the event becomes established.[216] Within the Samoan Islands, below average rainfall and higher than normal temperatures are recorded during El Niño events, which can lead to droughts and forest fires on the islands.[217] Other impacts include a decrease in the sea level, possibility of coral bleaching in the marine environment and an increased risk of a tropical cyclone affecting Samoa.[217]

In the late winter and spring during El Niño events, drier than average conditions can be expected in Hawaii.[218] On Guam during El Niño years, dry season precipitation averages below normal, but the probability of a tropical cyclone is more than triple what is normal, so extreme short duration rainfall events are possible.[219] On American Samoa during El Niño events, precipitation averages about 10 percent above normal, while La Niña events are associated with precipitation averaging about 10 percent below normal.[220]

South America[edit]

The effects of El Niño in South America are direct and strong. An El Niño is associated with warm and very wet weather months in April–October along the coasts of northern Peru and Ecuador, causing major flooding whenever the event is strong or extreme.[221]

Because El Niño's warm pool feeds thunderstorms above, it creates increased rainfall across the east-central and eastern Pacific Ocean, including several portions of the South American west coast. The effects of El Niño in South America are direct and stronger than in North America. An El Niño is associated with warm and very wet weather months in April–October along the coasts of northern Peru and Ecuador, causing major flooding whenever the event is strong or extreme.[222] The effects during the months of February, March, and April may become critical along the west coast of South America, El Niño reduces the upwelling of cold, nutrient-rich water that sustains large fish populations, which in turn sustain abundant sea birds, whose droppings support the fertilizer industry. The reduction in upwelling leads to fish kills off the shore of Peru.[223]

The local fishing industry along the affected coastline can suffer during long-lasting El Niño events. Peruvian fisheries collapsed during the 1970s due to overfishing following the 1972 El Niño Peruvian anchoveta reduction.[224] The fisheries were previously the world's largest, however, this collapse led to the decline of these fisheries. During the 1982–83 event, jack mackerel and anchoveta populations were reduced, scallops increased in warmer water, but hake followed cooler water down the continental slope, while shrimp and sardines moved southward, so some catches decreased while others increased.[225] Horse mackerel have increased in the region during warm events. Shifting locations and types of fish due to changing conditions create challenges for the fishing industry. Peruvian sardines have moved during El Niño events to Chilean areas. Other conditions provide further complications, such as the government of Chile in 1991 creating restrictions on the fishing areas for self-employed fishermen and industrial fleets.

Southern Brazil and northern Argentina also experience wetter than normal conditions during El Niño years, but mainly during the spring and early summer. Central Chile receives a mild winter with large rainfall, and the Peruvian-Bolivian Altiplano is sometimes exposed to unusual winter snowfall events. Drier and hotter weather occurs in parts of the Amazon River Basin, Colombia, and Central America.[226]

During a time of La Niña, drought affects the coastal regions of Peru and Chile.[227] From December to February, northern Brazil is wetter than normal.[227] La Niña causes higher than normal rainfall in the central Andes, which in turn causes catastrophic flooding on the Llanos de Mojos of Beni Department, Bolivia. Such flooding is documented from 1853, 1865, 1872, 1873, 1886, 1895, 1896, 1907, 1921, 1928, 1929 and 1931.[228]

Galápagos Islands[edit]

The Galápagos Islands are a chain of volcanic islands, nearly 600 miles west of Ecuador, South America.[229] in the Eastern Pacific Ocean. These islands support a wide diversity of terrestrial and marine species.[230] The ecosystem is based on the normal trade winds which influence upwelling of cold, nutrient rich waters to the islands.[231] During an El Niño event the trade winds weaken and sometimes blow from west to east, which causes the Equatorial current to weaken, raising surface water temperatures and decreasing nutrients in waters surrounding the Galápagos. El Niño causes a trophic cascade which impacts entire ecosystems starting with primary producers and ending with critical animals such as sharks, penguins, and seals.[232] The effects of El Niño can become detrimental to populations that often starve and die back during these years. Rapid evolutionary adaptations are displayed amongst animal groups during El Niño years to mitigate El Niño conditions.[233]

History[edit]

Average equatorial Pacific temperatures, published in 2009.

During human history[edit]

ENSO conditions have occurred at two- to seven-year intervals for at least the past 300 years, but most of them have been weak.[234]

El Niño may have led to the demise of the Moche and other pre-Columbian Peruvian cultures.[235] A recent study suggests a strong El Niño effect between 1789 and 1793 caused poor crop yields in Europe, which in turn helped touch off the French Revolution.[236] The extreme weather produced by El Niño in 1876–77 gave rise to the most deadly famines of the 19th century.[237] The 1876 famine alone in northern China killed up to 13 million people.[238]

The phenomenon had long been of interest because of its effects on the guano industry and other enterprises that depend on biological productivity of the sea. It is recorded that as early as 1822, cartographer Joseph Lartigue, of the French frigate La Clorinde under Baron Mackau, noted the "counter-current" and its usefulness for traveling southward along the Peruvian coast.[239][240][241]

Charles Todd, in 1888, suggested droughts in India and Australia tended to occur at the same time;[242] Norman Lockyer noted the same in 1904.[243] An El Niño connection with flooding was reported in 1894 by Victor Eguiguren (1852–1919) and in 1895 by Federico Alfonso Pezet (1859–1929).[244][240][245] In 1924, Gilbert Walker (for whom the Walker circulation is named) coined the term "Southern Oscillation".[246] He and others (including Norwegian-American meteorologist Jacob Bjerknes) are generally credited with identifying the El Niño effect.[247]

The major 1982–83 El Niño led to an upsurge of interest from the scientific community. The period 1990–95 was unusual in that El Niños have rarely occurred in such rapid succession.[248][249][unreliable source?][250] An especially intense El Niño event in 1998 caused an estimated 16% of the world's reef systems to die. The event temporarily warmed air temperature by 1.5 °C, compared to the usual increase of 0.25 °C associated with El Niño events.[251] Since then, mass coral bleaching has become common worldwide, with all regions having suffered "severe bleaching".[252]

Around 1525, when Francisco Pizarro made landfall in Peru, he noted rainfall in the deserts, the first written record of the impacts of El Niño.[253]

In geologic timescales[edit]

Evidence is also strong for El Niño events during the early Holocene epoch 10,000 years ago.[234] Different modes of ENSO-like events have been registered in paleoclimatic archives, showing different triggering methods, feedbacks and environmental responses to the geological, atmospheric and oceanographic characteristics of the time. These paleorecords can be used to provide a qualitative basis for conservation practices.[254]

Scientists have also found chemical signatures of warmer sea surface temperatures and increased rainfall caused by El Niño in coral specimens that are around 13,000 years old.[253]

Series/ epoch Age of archive / Location / Type of archive or proxy Description and references
Mid Holocene 4150 ya / Vanuatu Islands / Coral core Coral bleaching in Vanuatu coral records, indication of shoaling of thermocline, is analyzed for Sr/Ca and U/Ca content, from which temperature is regressed. The temperature variability shows that during the mid-Holocene, changes in the position of the anticyclonic gyre produced average to cold (La Niña) conditions, which were probably interrupted by strong warm events (El Niño), which might have produced the bleaching, associated to decadal variability.[255]
Holocene 12000ya / Bay of Guayaquil, Ecuador / Pollen content of marine core Pollen records show changes in precipitation, possibly related to variability of the position of the ITCZ, as well as the latitudinal maxima of the Humboldt Current, which both depend on ENSO frequency and amplitude variability. Three different regimes of ENSO influence are found in the marine core.[256]
Holocene 12000ya /

Pallcacocha Lake, Ecuador / Sediment core

Core shows warm events with periodicities of 2–8 years, which become more frequent over the Holocene until about 1,200 years ago, and then decline, on top of which there are periods of low and high ENSO-related events, possibly due to changes in insolation.[257][258]
LGM 45000ya / Australia / Peat core Moisture variability in the Australian core shows dry periods related to frequent warm events (El Niño), correlated to DO events. Although no strong correlation was found with the Atlantic Ocean, it is suggested that the insolation influence probably affected both oceans, although the Pacific Ocean seems to have the most influence on teleconnection in annual, millennial and semi-precessional timescales.[259]
Pleistocene 240 Kya / Indian and Pacific oceans / Coccolithophore in 9 deep sea cores 9 deep cores in the equatorial Indian and Pacific show variations in primary productivity, related to glacial-interglacial variability and precessional periods (23 ky) related to changes in the thermocline. There is also indication that the equatorial areas can be early responders to insolation forcing.[260]
Pliocene 2.8 Mya / Spain / Lacustrine laminated sediments core The basin core shows light and dark layers, related to summer/autumn transition where more/less productivity is expected. The core shows thicker or thinner layers, with periodicities of 12, 6–7 and 2–3 years, related to ENSO, North Atlantic Oscillation (NAO) and Quasi-biennial Oscillation (QBO), and possibly also insolation variability (sunspots).[261]
Pliocene 5.3 Mya / Equatorial Pacific / Foraminifera in deep sea cores Deep sea cores at ODP site 847 and 806 show that the Pliocene warm period presented permanent El Niño-like conditions, possibly related to changes in the mean state of extratropical regions[262] or changes in ocean heat transport resulting from increased tropical cyclone activity.[263]
Miocene 5.92-5.32 Mya / Italy / Evaporite varve thickness The varve close to the Mediterranean shows 2–7 year variability, closely related to ENSO periodicity. Model simulations show that there is more correlation with ENSO than NAO, and that there is a strong teleconnection with the Mediterranean due to lower gradients of temperature.[264]

Related patterns[edit]

Madden–Julian oscillation[edit]

A Hovmöller diagram of the 5-day running mean of outgoing longwave radiation showing the MJO. Time increases from top to bottom in the figure, so contours that are oriented from upper-left to lower-right represent movement from west to east.
The Madden–Julian oscillation (MJO) is the largest element of the intraseasonal (30- to 90-day) variability in the tropical atmosphere. It was discovered in 1971 by Roland Madden and Paul Julian of the American National Center for Atmospheric Research (NCAR).[265] It is a large-scale coupling between atmospheric circulation and tropical deep atmospheric convection.[266][267] Unlike a standing pattern like the El Niño–Southern Oscillation (ENSO), the Madden–Julian oscillation is a traveling pattern that propagates eastward, at approximately 4 to 8 m/s (14 to 29 km/h; 9 to 18 mph), through the atmosphere above the warm parts of the Indian and Pacific oceans. This overall circulation pattern manifests itself most clearly as anomalous rainfall.

Link to the El Niño-Southern oscillation[edit]

There is strong year-to-year (interannual) variability in Madden–Julian oscillation activity, with long periods of strong activity followed by periods in which the oscillation is weak or absent. This interannual variability of the MJO is partly linked to the El Niño–Southern Oscillation (ENSO) cycle. In the Pacific, strong MJO activity is often observed 6 to 12 months prior to the onset of an El Niño episode, but is virtually absent during the maxima of some El Niño episodes, while MJO activity is typically greater during a La Niña episode. Strong events in the Madden–Julian oscillation over a series of months in the western Pacific can speed the development of an El Niño or La Niña but usually do not in themselves lead to the onset of a warm or cold ENSO event.[268] However, observations suggest that the 1982-1983 El Niño developed rapidly during July 1982 in direct response to a Kelvin wave triggered by an MJO event during late May.[269] Further, changes in the structure of the MJO with the seasonal cycle and ENSO might facilitate more substantial impacts of the MJO on ENSO. For example, the surface westerly winds associated with active MJO convection are stronger during advancement toward El Niño and the surface easterly winds associated with the suppressed convective phase are stronger during advancement toward La Niña.[270] Globally, the interannual variability of the MJO is most determined by atmospheric internal dynamics, rather than surface conditions.[clarification needed]

Pacific decadal oscillation[edit]

PDO positive phase global pattern
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 (meaning time periods of a few years to as much as time periods of multiple decades). There is evidence of reversals in the prevailing polarity (meaning changes in cool surface waters versus warm surface waters within the region) 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.

Mechanisms[edit]

ENSO can influence the global circulation pattern thousands of kilometers away from the equatorial Pacific through the "atmospheric bridge". During El Niño events, deep convection and heat transfer to the troposphere is enhanced over the anomalously warm sea surface temperature, this ENSO-related tropical forcing generates Rossby waves that propagate poleward and eastward and are subsequently refracted back from the pole to the tropics. The planetary waves form at preferred locations both in the North and South Pacific Ocean, and the teleconnection pattern is established within 2–6 weeks.[271] ENSO driven patterns modify surface temperature, humidity, wind, and the distribution of clouds over the North Pacific that alter surface heat, momentum, and freshwater fluxes and thus induce sea surface temperature, salinity, and mixed layer depth (MLD) anomalies.

Pacific Meridional Mode[edit]

The SST and wind anomalies of the PMM positive phase
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 PMM is not the same thing as the El Niño-Southern Oscillation (ENSO), but there is evidence that PMM events can trigger ENSO events, especially Central Pacific El Niño events. The PMM state can also modulate hurricane activity in the East Pacific and typhoon activity in the West Pacific oceans and alter precipitation on the continents surrounding the Pacific Ocean. The South Pacific Ocean has a PMM-like mode known as the "South Pacific Meridional Mode" (SPMM) that also influences the ENSO cycle.

In the early 21st century, the intensity of the 2014–16 El Niño event and the highly active 2018 Pacific hurricane and typhoon seasons have been attributed to positive PMM events. With anthropogenic global warming, PMM activity is likely to increase, and some scientists have proposed that a loss of Antarctic and especially Arctic sea ice will induce future positive PMM events.

South Pacific Meridional Mode[edit]

The "South Pacific Meridional Mode" (SPMM) is an analogous climate mode in the south Pacific;[272] Zhang, Clement and Di Nezio proposed its existence in 2014[273] and it operates in a nearly identical manner to the northern hemisphere PMM[274] albeit according to You and Furtado (2018) with SST anomalies peaking during (austral) summer and wind anomalies during (austral) winter.[275] According to Middlemas et al. (2019), cloud radiative feedbacks counteract the persistence of SPMM.[276] The SPMM has been further related to a different climate mode known as the "South Pacific Quadrupole"[277] and the "South Pacific subtropical dipole mode".[278]

Unlike the PMM, the South Pacific Meridional Mode has a more extensive influence on the Pacific Ocean than the northern PMM, by impacting the equator instead of remaining within the southern hemisphere,[279] for example, and favouring the onset of "canonical" East Pacific El Niño events instead of Central Pacific El Niño events like PMM.[280] This is because the southern trade winds in the east Pacific cross the equator into the northern hemisphere and can thus "transport" the effects of the South Pacific Meridional Mode northward. Ocean dynamics in the cold tongue region may also play a role.[281][282] The exact relation between SPMM and ENSO onset is still unclear.[280] The failure of an expected El Niño event to develop in 2014 has been explained by an unfavourable state of the SPMM in that year.[283] Apart from ENSO development, the SPMM has impacts on the Chilean Desventuradas Islands and Juan Fernandez Island according to Dewitte et al. (2021).[284] Kim et al. (2022) proposed that cooling in the Southern Ocean can force a negative SPMM state.[285]

See also[edit]

  • Ocean dynamical thermostat – Physical mechanism affecting sea surface temperatures in the Pacific Ocean
  • Recharge oscillator – Theory to explain the periodical variation of the sea surface temperature and thermocline depth

For La Niña:

For El Niño:

Notes[edit]

  1. ^ a b Along the western coast of South America, water near the ocean surface is pushed westward due to the combination of the trade winds and the Coriolis effect. This process is known as Ekman transport. Colder water from deeper in the ocean rises along the continental margin to replace the near-surface water.[26]
  2. ^ The total weight of a column of ocean water is almost the same in the western and east Pacific. Because the warmer waters of the upper ocean are slightly less dense than the cooler deep ocean, the thicker layer of warmer water in the western Pacific means the thermocline there must be deeper. The difference in weight must be enough to drive any deep water return flow.[25]: 12 

References[edit]

  1. ^ Wald, Lucien (2021). "Definitions of time: from year to second". Fundamentals of solar radiation. Boca Raton: CRC Press. ISBN 978-0-367-72588-4.
  2. ^ Climate Prediction Center (2005-12-19). "Frequently Asked Questions about El Niño and La Niña". National Centers for Environmental Prediction. Archived from the original on 2009-08-27. Retrieved 2009-07-17.
  3. ^ a b Trenberth, K.E.; P.D. Jones; P. Ambenje; R. Bojariu; D. Easterling; A. Klein Tank; D. Parker; F. Rahimzadeh; J.A. Renwick; M. Rusticucci; B. Soden; P. Zhai. "Observations: Surface and Atmospheric Climate Change". In Solomon, S.; D. Qin; M. Manning; et al. (eds.). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press. pp. 235–336. Archived from the original on 2017-09-24. Retrieved 2014-06-30.
  4. ^ "El Niño, La Niña and the Southern Oscillation". MetOffice. Archived from the original on 2023-10-27. Retrieved 2015-08-18.
  5. ^ a b Becker, Emily (4 December 2014). "December's ENSO Update: Close, but no cigar". ENSO Blog. Archived from the original on 22 March 2016.
  6. ^ a b "El Niño and La Niña". New Zealand's National Institute of Water and Atmospheric Research. 27 February 2007. Archived from the original on 19 March 2016. Retrieved 11 April 2016.
  7. ^ a b Emily Becker (2016). "How Much Do El Niño and La Niña Affect Our Weather? This fickle and influential climate pattern often gets blamed for extreme weather. A closer look at the most recent cycle shows that the truth is more subtle". Scientific American. 315 (4): 68–75. doi:10.1038/scientificamerican1016-68. PMID 27798565.
  8. ^ a b Brown, Patrick T.; Li, Wenhong; Xie, Shang-Ping (27 January 2015). "Regions of significant influence on unforced global mean surface air temperature variability in climate models: Origin of global temperature variability". Journal of Geophysical Research: Atmospheres. 120 (2): 480–494. doi:10.1002/2014JD022576. hdl:10161/9564.
  9. ^ a b Trenberth, Kevin E.; Fasullo, John T. (December 2013). "An apparent hiatus in global warming?". Earth's Future. 1 (1): 19–32. Bibcode:2013EaFut...1...19T. doi:10.1002/2013EF000165.
  10. ^ a b c d e f g IPCC, 2021: Climate Change 2021: The Physical Science Basis Archived 2023-12-08 at the Wayback Machine. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change Archived 2023-05-26 at the Wayback Machine [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 2391 pp. doi:10.1017/9781009157896.
  11. ^ a b c Collins, M.; An, S-I; Cai, W.; Ganachaud, A.; Guilyardi, E.; Jin, F-F; Jochum, M.; Lengaigne, M.; Power, S.; Timmermann, A.; Vecchi, G.; Wittenberg, A. (2010). "The impact of global warming on the tropical Pacific Ocean and El Niño". Nature Geoscience. 3 (6): 391–7. Bibcode:2010NatGe...3..391C. doi:10.1038/ngeo868. Archived from the original on 2019-09-14. Retrieved 2019-01-10.
  12. ^ a b c d L'Heureux, Michelle (5 May 2014). "What is the El Niño–Southern Oscillation (ENSO) in a nutshell?". ENSO Blog. Archived from the original on 9 April 2016.
  13. ^ Carrillo, Camilo N. (1892) "Disertación sobre las corrientes oceánicas y estudios de la correinte Peruana ó de Humboldt" Archived 2023-10-30 at the Wayback Machine (Dissertation on the ocean currents and studies of the Peruvian, or Humboldt's, current), Boletín de la Sociedad Geográfica de Lima, 2 : 72–110. [in Spanish] From p. 84: Archived 2023-10-30 at the Wayback Machine "Los marinos paiteños que navegan frecuentemente cerca de la costa y en embarcaciones pequeñas, ya al norte ó al sur de Paita, conocen esta corriente y la denomination Corriente del Niño, sin duda porque ella se hace mas visible y palpable después de la Pascua de Navidad." (The sailors [from the city of] Paita who sail often near the coast and in small boats, to the north or the south of Paita, know this current and call it "the current of the Boy [el Niño]", undoubtedly because it becomes more visible and palpable after the Christmas season.)
  14. ^ "El Niño Information". California Department of Fish and Game, Marine Region. Archived from the original on 2019-10-27. Retrieved 2014-06-30.
  15. ^ Trenberth, Kevin E (December 1997). "The Definition of El Niño". Bulletin of the American Meteorological Society. 78 (12): 2771–2777. Bibcode:1997BAMS...78.2771T. doi:10.1175/1520-0477(1997)078<2771:TDOENO>2.0.CO;2.
  16. ^ "The Strongest El Nino in Decades Is Going to Mess With Everything". Bloomberg.com. 21 October 2015. Archived from the original on 11 February 2022. Retrieved 18 February 2017.
  17. ^ "How the Pacific Ocean changes weather around the world". Popular Science. Archived from the original on 3 January 2022. Retrieved 19 February 2017.
  18. ^ a b c d e "What are "El Niño" and "La Niña"?". National Ocean Service. oceanservice.noaa.gov. National Oceanic and Atmospheric Administration. February 10, 2020. Archived from the original on 11 January 2023. Retrieved 11 September 2020.
  19. ^ "What is "La Niña"?". Tropical Atmosphere Ocean project / Pacific Marine Environmental Laboratory. National Oceanic and Atmospheric Administration. 24 March 2008. Archived from the original on 16 December 2008. Retrieved 17 July 2009.
  20. ^ "The Southern Oscillation and its Links to the ENSO Cycle". www.cpc.ncep.noaa.gov. NOAA National Weather Service Climate Prediction Centre. Archived from the original on 19 January 2024. Retrieved 19 January 2024.
  21. ^ a b c d e f g "El Niño Southern Oscillation (ENSO)". About Australian climate. Bureau of Meteorology. Archived from the original on 22 January 2024. Retrieved 22 January 2024.
  22. ^ a b c d "El Niño, La Niña and Australia's Climate" (PDF). Bureau of Meteorology. February 2005. Archived (PDF) from the original on 22 January 2024. Retrieved 22 January 2024.
  23. ^ a b c d "Effects of ENSO in the Pacific". National Weather Service. Retrieved 22 January 2024.
  24. ^ "What is ENSO?". IRI/LDEO Climate Data Library. International Research Institute for Climate and Society. Retrieved 22 January 2024.
  25. ^ Sarachik, Edward S.; Cane, Mark A. (2010). The El Niño-Southern Oscillation Phenomenon. Cambridge: Cambridge University Press. ISBN 978-0-521-84786-5.
  26. ^ "Wind Driven Surface Currents: Upwelling and Downwelling Background". Ocean Motion and surface currents. NASA. Retrieved 22 January 2024.
  27. ^ a b L'Heureux, Michelle (5 May 2014). "What is the El Niño–Southern Oscillation (ENSO) in a nutshell?". ENSO Blog. Climate.gov. Retrieved 22 January 2024.
  28. ^ a b c d Wang, Chunzai; Deser, Clara; Yu, Jin-Yi; DiNezio, Pedro; Clement, Amy (2017). "El Niño and Southern Oscillation (ENSO): A Review" (PDF). In Glynn, Peter W.; Manzello, Derek P.; Enochs, Ian C. (eds.). Coral Reefs of the Eastern Tropical Pacific. Coral Reefs of the World. Vol. 8. Springer. pp. 85–106. doi:10.1007/978-94-017-7499-4_4. ISBN 978-94-017-7498-7. Retrieved 22 January 2024. {{cite book}}: |journal= ignored (help)
  29. ^ a b L'Heureux, Michelle (23 October 2020). "The Rise of El Niño and La Niña". ENSO Blog. Climate.gov. Retrieved 22 January 2024.
  30. ^ Fox, Alex (5 October 2023). "What is El Niño?". Scripps Institution of Oceanography. San Diego, California: University of California–San Diego. Retrieved 22 January 2024.
  31. ^ Wang, Chunzai (1 November 2018). "A review of ENSO theories". National Science Review. 5 (6): 813–825. doi:10.1093/nsr/nwy104.
  32. ^ Yang, Song; Li, Zhenning; Yu, Jin-Yi; Hu, Xiaoming; Dong, Wenjie; He, Shan (1 November 2018). "El Niño–Southern Oscillation and its impact in the changing climate". National Science Review. 5 (6): 840–857. doi:10.1093/nsr/nwy046.
  33. ^ Trenberth, Kevin (2022). Chapter 12: El Niño. In: The changing flow of energy through the climate system. Cambridge New York, NY Port Melbourne: Cambridge University Press. ISBN 978-1-108-97903-0.
  34. ^ "Climate glossary — Southern Oscilliation Index (SOI)". Bureau of Meteorology (Australia). 2002-04-03. Archived from the original on 2017-12-26. Retrieved 2009-12-31.
  35. ^ a b c d Barnston, Anthony (2015-01-29). "Why are there so many ENSO indexes, instead of just one?". NOAA. Archived from the original on 2015-09-05. Retrieved 2015-08-14.
  36. ^ International Research Institute for Climate and Society. "Southern Oscillation Index (SOI) and Equatorial SOI". Columbia University. Archived from the original on 2015-11-17. Retrieved 2015-08-14.
  37. ^ Climate Prediction Center (19 December 2005). "Frequently Asked Questions about El Niño and La Niña". National Centers for Environmental Prediction. Archived from the original on 27 August 2009. Retrieved 17 July 2009.
  38. ^ Sergey K. Gulev; Peter W. Thorne; Jinho Ahn; Frank J. Dentener; Catia M. Domingues; Sebastian Gerland; Daoyi Gong; Darrell S. Kaufman; Hyacinth C. Nnamchi; Johannes Quaas; Juan Antonio Rivera; Shubha Sathyendranath; Sharon L. Smith; Blair Trewin; Karina von Shuckmann; Russell S. Vose. "Changing state of the climate system" (PDF). In Valérie Masson-Delmotte; Panmao Zhai; Anna Pirani; Sarah L. Connors; C. Péan; Sophie Berger; Nada Caud; Y. Chen; Leah Goldfarb; Melissa I. Gomis; Mengtian Huang; Katherine Leitzell; Elisabeth Lonnoy; J. B. Robin Matthews; Thomas K. Maycock; Tim Waterfield; Özge Yelekçi; R. Yu; Botao Zhou (eds.). Climate Change 2021: The Physical Science Basis. The contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK: Cambridge University Press. Archived (PDF) from the original on 2022-03-02. Retrieved 2024-01-18.
  39. ^ Climate Prediction Center Internet Team (2012-04-26). "Frequently Asked Questions about El Niño and La Niña". National Oceanic and Atmospheric Administration. Archived from the original on 2020-05-02. Retrieved 2014-06-30.
  40. ^ International Research Institute for Climate and Society (February 2002). "More Technical ENSO Comment". Columbia University. Archived from the original on 2014-07-14. Retrieved 2014-06-30.
  41. ^ State Climate Office of North Carolina. "Global Patterns – El Niño-Southern Oscillation (ENSO)". North Carolina State University. Archived from the original on 2014-06-27. Retrieved 2014-06-30.
  42. ^ "Australian Climate Influences: El Niño". Australian Bureau of Meteorology. Archived from the original on 24 March 2016. Retrieved 4 April 2016.
  43. ^ a b L'Heureux, Michelle (5 May 2014). "What is the El Niño–Southern Oscillation (ENSO) in a nutshell?". ENSO Blog. Archived from the original on 9 April 2016. Retrieved 7 April 2016.
  44. ^ Intergovernmental Panel on Climate Change (2007). "Climate Change 2007: Working Group I: The Physical Science Basis: 3.7 Changes in the Tropics and Subtropics, and the Monsoons". World Meteorological Organization. Archived from the original on 2014-07-14. Retrieved 2014-07-01.
  45. ^ "What is El Niño and what might it mean for Australia?". Australian Bureau of Meteorology. Archived from the original on 18 March 2016. Retrieved 10 April 2016.
  46. ^ Climate Prediction Center (19 December 2005). "ENSO FAQ: How often do El Niño and La Niña typically occur?". National Centers for Environmental Prediction. Archived from the original on 27 August 2009. Retrieved 26 July 2009.
  47. ^ National Climatic Data Center (June 2009). "El Niño / Southern Oscillation (ENSO) June 2009". National Oceanic and Atmospheric Administration. Retrieved 26 July 2009.
  48. ^ "Historical El Niño/La Niña episodes (1950–present)". United States Climate Prediction Center. 1 February 2019. Archived from the original on 29 November 2014. Retrieved 15 March 2019.
  49. ^ "El Niño - Detailed Australian Analysis". Australian Bureau of Meteorology. Archived from the original on 3 May 2021. Retrieved 3 April 2016.
  50. ^ "El Niño in Australia" (PDF). Bom.gov.au. Archived (PDF) from the original on 7 March 2022. Retrieved 1 March 2022.
  51. ^ Brian Donegan (14 March 2019). "El Niño Conditions Strengthen, Could Last Through Summer". The Weather Company. Archived from the original on 15 March 2019. Retrieved 15 March 2019.
  52. ^ "El Nino is over, NOAA says". Al.com. 8 August 2019. Archived from the original on 5 September 2019. Retrieved 5 September 2019.
  53. ^ "Here comes El Nino: It's early, likely to be big, sloppy and add even more heat to a warming world". The Independent. 2023-06-08. Archived from the original on 2023-06-10. Retrieved 2023-06-23.
  54. ^ Henson, Bob (9 June 2023). "NOAA makes it official: El Niño is here". Yale Climate Connections. Archived from the original on 10 June 2023. Retrieved 11 June 2023.
  55. ^ "El Niño Outlook ( June 2023 - December 2023 )". Climate Prediction Division. Japan Meteorological Agency. 9 June 2023. Archived from the original on 2 May 2023. Retrieved 12 June 2023. El Niño conditions are considered to be present in the equatorial Pacific.
  56. ^ Davis, Mike (2001). Late Victorian Holocausts: El Niño Famines and the Making of the Third World. London: Verso. p. 271. ISBN 978-1-85984-739-8.
  57. ^ "Very strong 1997-98 Pacific warm episode (El Niño)". Archived from the original on 3 May 2021. Retrieved 28 July 2015.
  58. ^ Sutherland, Scott (16 February 2017). "La Niña calls it quits. Is El Niño paying us a return visit?". The Weather Network. Archived from the original on 18 February 2017. Retrieved 17 February 2017.
  59. ^ Kim, WonMoo; Wenju Cai (2013). "Second peak in the far eastern Pacific sea surface temperature anomaly following strong El Niño events". Geophys. Res. Lett. 40 (17): 4751–4755. Bibcode:2013GeoRL..40.4751K. doi:10.1002/grl.50697. S2CID 129885922.
  60. ^ "August 2016 ENSO update;Wavy Gravy". Climate.gov.uk. Archived from the original on 11 December 2022. Retrieved 16 October 2021.
  61. ^ Cold and warm episodes by season. Climate Prediction Center (Report). National Oceanic and Atmospheric Administration. Archived from the original on September 26, 2023. Retrieved September 11, 2020.
  62. ^ La Niña – Detailed Australian analysis (Report). Australian Bureau of Meteorology. Archived from the original on 28 December 2017. Retrieved 3 April 2016.
  63. ^ Druffel, Ellen R. M.; Griffin, Sheila; Vetter, Desiree; Dunbar, Robert B.; Mucciarone, David M. (16 March 2015). "Identification of frequent La Niña events during the early 1800s in the east equatorial Pacific". Geophysical Research Letters. 42 (5): 1512–1519. Bibcode:2015GeoRL..42.1512D. doi:10.1002/2014GL062997. S2CID 129644802. Archived from the original on 15 January 2023. Retrieved 26 February 2022.
  64. ^ The following sources identified the listed "La Niña years":
  65. ^ Trenberth, Kevin E.; Stepaniak, David P. (15 April 2001). "Indices of El Niño Evolution". Journal of Climate. 14 (8): 1697–1701. Bibcode:2001JCli...14.1697T. doi:10.1175/1520-0442(2001)014<1697:LIOENO>2.0.CO;2. Archived from the original on 23 December 2019. Retrieved 27 August 2019.
  66. ^ Kennedy, Adam M.; D. C. Garen; R. W. Koch (2009). "The association between climate teleconnection indices and Upper Klamath seasonal streamflow: Trans-Niño Index". Hydrol. Process. 23 (7): 973–84. Bibcode:2009HyPr...23..973K. CiteSeerX 10.1.1.177.2614. doi:10.1002/hyp.7200. S2CID 16514830.
  67. ^ Lee, Sang-Ki; R. Atlas; D. Enfield; C. Wang; H. Liu (2013). "Is there an optimal ENSO pattern that enhances large-scale atmospheric processes conducive to tornado outbreaks in the U.S?". J. Climate. 26 (5): 1626–1642. Bibcode:2013JCli...26.1626L. doi:10.1175/JCLI-D-12-00128.1.
  68. ^ a b Kao, Hsun-Ying; Jin-Yi Yu (2009). "Contrasting Eastern-Pacific and Central-Pacific Types of ENSO". J. Climate. 22 (3): 615–632. Bibcode:2009JCli...22..615K. CiteSeerX 10.1.1.467.457. doi:10.1175/2008JCLI2309.1.
  69. ^ Larkin, N. K.; Harrison, D. E. (2005). "On the definition of El Niño and associated seasonal average U.S. Weather anomalies". Geophysical Research Letters. 32 (13): L13705. Bibcode:2005GeoRL..3213705L. doi:10.1029/2005GL022738.
  70. ^ a b Yuan Yuan; HongMing Yan (2012). "Different types of La Niña events and different responses of the tropical atmosphere". Chinese Science Bulletin. 58 (3): 406–415. Bibcode:2013ChSBu..58..406Y. doi:10.1007/s11434-012-5423-5.
  71. ^ a b Cai, W.; Cowan, T. (17 June 2009). "La Niña Modoki impacts Australia autumn rainfall variability". Geophysical Research Letters. 36 (12): L12805. Bibcode:2009GeoRL..3612805C. doi:10.1029/2009GL037885.
  72. ^ Johnson, Nathaniel C. (1 July 2013). "How Many ENSO Flavors Can We Distinguish?". Journal of Climate. 26 (13): 4816–4827. Bibcode:2013JCli...26.4816J. doi:10.1175/JCLI-D-12-00649.1. S2CID 55416945.
  73. ^ Kim, Hye-Mi; Webster, Peter J.; Curry, Judith A. (3 July 2009). "Impact of Shifting Patterns of Pacific Ocean Warming on North Atlantic Tropical Cyclones". Science. 325 (5936): 77–80. Bibcode:2009Sci...325...77K. doi:10.1126/science.1174062. PMID 19574388. S2CID 13250045.
  74. ^ Cai, W.; Cowan, T. (2009). "La Niña Modoki impacts Australia autumn rainfall variability". Geophysical Research Letters. 36 (12): L12805. Bibcode:2009GeoRL..3612805C. doi:10.1029/2009GL037885. ISSN 0094-8276.
  75. ^ M R Ramesh Kumar (2014-04-23). "El Nino, La Nina and the Indian sub-continent". Society for Environmental Communications. Archived from the original on 2014-07-21. Retrieved 2014-07-25.
  76. ^ S. George Philander (2004). Our Affair with El Niño: How We Transformed an Enchanting Peruvian Current Into a Global Climate Hazard. Princeton University Press. ISBN 978-0-691-11335-7.
  77. ^ "Study Finds El Niños are Growing Stronger". NASA. Archived from the original on 17 November 2022. Retrieved 3 August 2014.
  78. ^ Takahashi, K.; Montecinos, A.; Goubanova, K.; Dewitte, B. (2011). "Reinterpreting the Canonical and Modoki El Nino" (PDF). Geophysical Research Letters. 38 (10): n/a. Bibcode:2011GeoRL..3810704T. doi:10.1029/2011GL047364. hdl:10533/132105. S2CID 55675672. Archived (PDF) from the original on 2019-05-03. Retrieved 2019-08-12.
  79. ^ Different Impacts of Various El Niño Events (PDF) (Report). NOAA. Archived (PDF) from the original on 2023-07-25. Retrieved 2024-01-18.
  80. ^ The Enhanced Drying Effect of Central Pacific El Niño on US Winters (Report). IOP Science. Archived from the original on 3 September 2015. Retrieved 5 February 2023..
  81. ^ Monitoring the Pendulum (Report). IOP Science. doi:10.1088/1748-9326/aac53f.
  82. ^ "El Nino's Bark is Worse than its Bite". The Western Producer. Archived from the original on 14 January 2019. Retrieved 11 January 2019.
  83. ^ Yuan, Yuan; Yan, HongMing (2012). "Different types of La Niña events and different responses of the tropical atmosphere". Chinese Science Bulletin. 58 (3): 406–415. Bibcode:2013ChSBu..58..406Y. doi:10.1007/s11434-012-5423-5.
  84. ^ Tedeschi, Renata G.; Cavalcanti, Iracema F. A. (23 April 2014). "Influência dos ENOS Canônico e Modoki na precipitação da América do Sul" (PDF) (in Portuguese). Instituto Nacional de Pesquisas Espaciais/Centro de Previsão de Tempo e Estudos Climáticos. Archived from the original (PDF) on 23 October 2014. Retrieved 27 September 2014.
  85. ^ For evidence of La Niña Modoki, and identification of La Niña Modoki year:
  86. ^ Yeh, Sang-Wook; Kug, Jong-Seong; Dewitte, Boris; Kwon, Min-Ho; Kirtman, Ben P.; Jin, Fei-Fei (September 2009). "El Niño in a changing climate". Nature. 461 (7263): 511–4. Bibcode:2009Natur.461..511Y. doi:10.1038/nature08316. PMID 19779449. S2CID 4423723.
  87. ^ Nicholls, N. (2008). "Recent trends in the seasonal and temporal behaviour of the El Niño Southern Oscillation". Geophys. Res. Lett. 35 (19): L19703. Bibcode:2008GeoRL..3519703N. doi:10.1029/2008GL034499. S2CID 129372366.
  88. ^ McPhaden, M.J.; Lee, T.; McClurg, D. (2011). "El Niño and its relationship to changing background conditions in the tropical Pacific Ocean". Geophys. Res. Lett. 38 (15): L15709. Bibcode:2011GeoRL..3815709M. doi:10.1029/2011GL048275. S2CID 9168925.
  89. ^ Giese, B.S.; Ray, S. (2011). "El Niño variability in simple ocean data assimilation (SODA), 1871–2008". J. Geophys. Res. 116 (C2): C02024. Bibcode:2011JGRC..116.2024G. doi:10.1029/2010JC006695. S2CID 85504316.
  90. ^ Newman, M.; Shin, S.-I.; Alexander, M.A. (2011). "Natural variation in ENSO flavors" (PDF). Geophys. Res. Lett. 38 (14): L14705. Bibcode:2011GeoRL..3814705N. doi:10.1029/2011GL047658. Archived (PDF) from the original on 2020-01-24. Retrieved 2019-08-27.
  91. ^ Yeh, S.-W.; Kirtman, B.P.; Kug, J.-S.; Park, W.; Latif, M. (2011). "Natural variability of the central Pacific El Niño event on multi-centennial timescales" (PDF). Geophys. Res. Lett. 38 (2): L02704. Bibcode:2011GeoRL..38.2704Y. doi:10.1029/2010GL045886. Archived (PDF) from the original on 2019-12-03. Retrieved 2019-08-27.
  92. ^ Hanna Na; Bong-Geun Jang; Won-Moon Choi; Kwang-Yul Kim (2011). "Statistical simulations of the future 50-year statistics of cold-tongue El Niño and warm-pool El Niño". Asia-Pacific J. Atmos. Sci. 47 (3): 223–233. Bibcode:2011APJAS..47..223N. doi:10.1007/s13143-011-0011-1. S2CID 120649138.
  93. ^ L'Heureux, M.; Collins, D.; Hu, Z.-Z. (2012). "Linear trends in sea surface temperature of the tropical Pacific Ocean and implications for the El Niño-Southern Oscillation". Climate Dynamics. 40 (5–6): 1–14. Bibcode:2013ClDy...40.1223L. doi:10.1007/s00382-012-1331-2.
  94. ^ Lengaigne, M.; Vecchi, G. (2010). "Contrasting the termination of moderate and extreme El Niño events in coupled general circulation models". Climate Dynamics. 35 (2–3): 299–313. Bibcode:2010ClDy...35..299L. doi:10.1007/s00382-009-0562-3. S2CID 14423113. Archived from the original on 2019-12-03. Retrieved 2019-01-10.
  95. ^ Takahashi, K.; Montecinos, A.; Goubanova, K.; Dewitte, B. (2011). "ENSO regimes: Reinterpreting the canonical and Modoki El Niño" (PDF). Geophys. Res. Lett. 38 (10): L10704. Bibcode:2011GeoRL..3810704T. doi:10.1029/2011GL047364. hdl:10533/132105. S2CID 55675672. Archived (PDF) from the original on 2019-05-03. Retrieved 2019-08-12.
  96. ^ Kug, J.-S.; Jin, F.-F.; An, S.-I. (2009). "Two types of El Niño events: Cold Tongue El Niño and Warm Pool El Niño". J. Climate. 22 (6): 1499–1515. Bibcode:2009JCli...22.1499K. doi:10.1175/2008JCLI2624.1. S2CID 6708133.
  97. ^ Shinoda, Toshiaki; Hurlburt, Harley E.; Metzger, E. Joseph (2011). "Anomalous tropical ocean circulation associated with La Niña Modoki". Journal of Geophysical Research: Oceans. 115 (12): C12001. Bibcode:2011JGRC..11612001S. doi:10.1029/2011JC007304.
  98. ^ a b c "El Niño, La Niña, ENSO, ENOS, El Niño Modoki, El Niño Canónico, El Niño Extraordinario, El Niño Godzilla, El Niño Costero, El Niño Oriental ¿En qué consisten realmente y cómo afectan al Ecuador?". Instituto Oceanográfico de la Armada del Ecuador (in Spanish). Retrieved 11 February 2024.
  99. ^ a b "ANTECEDENTES DE "EL NIÑO COSTERO"". INSTITUTO DEL MAR DEL PERU (in Spanish). Retrieved 11 February 2024.
  100. ^ Hu, Zeng-Zhen; Huang, Bohua; Zhu, Jieshun; Kumar, Arun; McPhaden, Michael J. (6 June 2018). "On the variety of coastal El Niño events". Climate Dynamics. 52 (12): 7537–7552. doi:10.1007/s00382-018-4290-4. S2CID 135045763. Retrieved 11 February 2024.
  101. ^ ZENTENO, HERMOGENES EDGARD GONZALES (2022). PREDICCIÓN DEL FENÓMENO EL NIÑO MEDIANTE ÍNDICES OCEÁNICOS E INFLUENCIA DE LA ZONA DE CONVERGENCIA INTERTROPICAL EN EL NORTE PERUANO (PDF) (PhD) (in Spanish). National Agrarian University. Retrieved 11 February 2024.
  102. ^ Aste, Fiorella (17 March 2017). "Cómo afecta El Niño costero a Chile, el fenómeno que ha dejado a más de 60 mil damnificados en Perú" (in Spanish). La Tercera. Retrieved 11 February 2024.
  103. ^ Blume, Daniela Valdivia (19 January 2024). "¿Se viene La Niña en Perú? Enfen explica lo que podría suceder en los siguientes meses" (in Spanish). Infobae. Retrieved 11 February 2024.
  104. ^ Takahashi, Ken; Martínez, Alejandra G. (2019-06-01). "The very strong coastal El Niño in 1925 in the far-eastern Pacific". Climate Dynamics. 52 (12): 7389–7415. Bibcode:2019ClDy...52.7389T. doi:10.1007/s00382-017-3702-1. hdl:20.500.12816/738. ISSN 1432-0894. S2CID 134011107.
  105. ^ a b "Eventos El Niño y La Niña Costeros" (in Spanish). Comité Multisectorial Encargado del Estudio Nacional del Fenómeno El Niño. Retrieved 11 February 2024.
  106. ^ Becker, Emily (27 May 2014). "How will we know when an El Niño has arrived?". ENSO Blog. Archived from the original on 22 March 2016.
  107. ^ Climate Prediction Center (2014-06-30). "ENSO: Recent Evolution, Current Status and Predictions" (PDF). National Oceanic and Atmospheric Administration. pp. 5, 19–20. Archived (PDF) from the original on 2005-03-05. Retrieved 2014-06-30.
  108. ^ "ENSO Tracker: About ENSO and the Tracker". Australian Bureau of Meteorology. Archived from the original on 15 January 2023. Retrieved 4 April 2016.
  109. ^ "Historical El Niño and La Niña Events". Japan Meteorological Agency. Archived from the original on 14 July 2022. Retrieved 4 April 2016.
  110. ^ Met Office (2012-10-11). "El Niño, La Niña and the Southern Oscillation". United Kingdom. Archived from the original on 2023-10-27. Retrieved 2014-06-30.
  111. ^ National Climatic Data Center (June 2009). "El Niño / Southern Oscillation (ENSO) June 2009". National Oceanic and Atmospheric Administration. Retrieved 2009-07-26.
  112. ^ "Climate.gov". NOAA. Global Climate Dashboard > Climate Variability. Archived from the original on 3 July 2011. Retrieved 22 December 2017.
  113. ^ "El Niño and La Niña". New Zealand: National Institute of Water and Atmospheric Research. 2007-02-27. Archived from the original on 19 March 2016. Retrieved 11 April 2016.
  114. ^ Merryfield, William J. (2006). "Changes to ENSO under CO2 Doubling in a Multimodel Ensemble". Journal of Climate. 19 (16): 4009–27. Bibcode:2006JCli...19.4009M. CiteSeerX 10.1.1.403.9784. doi:10.1175/JCLI3834.1.
  115. ^ Guilyardi, E.; Wittenberg, Andrew; Fedorov, Alexey; Collins, Mat; Wang, Chunzai; Capotondi, Antonietta; Van Oldenborgh, Geert Jan; Stockdale, Tim (2009). "Understanding El Nino in Ocean-Atmosphere General Circulation Models: Progress and Challenges" (PDF). Bulletin of the American Meteorological Society. 90 (3): 325–340. Bibcode:2009BAMS...90..325G. doi:10.1175/2008BAMS2387.1. hdl:10871/9288. S2CID 14866973. Archived (PDF) from the original on 2021-04-29. Retrieved 2021-01-21.
  116. ^ Meehl, G. A.; Teng, H.; Branstator, G. (2006). "Future changes of El Niño in two global coupled climate models". Climate Dynamics. 26 (6): 549–566. Bibcode:2006ClDy...26..549M. doi:10.1007/s00382-005-0098-0. S2CID 130825304. Archived from the original on 2019-12-28. Retrieved 2019-08-12.
  117. ^ Philip, Sjoukje; van Oldenborgh, Geert Jan (June 2006). "Shifts in ENSO coupling processes under global warming". Geophysical Research Letters. 33 (11): L11704. Bibcode:2006GeoRL..3311704P. doi:10.1029/2006GL026196.
  118. ^ "Climate Change is Making El Niños More Intense, Study Finds". Yale E360. Archived from the original on 2022-04-25. Retrieved 2022-04-19.
  119. ^ Wang, Bin; Luo, Xiao; Yang, Young-Min; Sun, Weiyi; Cane, Mark A.; Cai, Wenju; Yeh, Sang-Wook; Liu, Jian (2019-11-05). "Historical change of El Niño properties sheds light on future changes of extreme El Niño". Proceedings of the National Academy of Sciences. 116 (45): 22512–22517. Bibcode:2019PNAS..11622512W. doi:10.1073/pnas.1911130116. ISSN 0027-8424. PMC 6842589. PMID 31636177.
  120. ^ Jiu,Liping; Song,Mirong; Zhu,Zhu; Horton, Radley M; Hu,Yongyun; Xie,Shang-Ping (23 Aug 2022). "Arctic sea-ice loss is projected to lead to more frequent strong El Niño events". Nature Communications. 13 (1): 4952. Bibcode:2022NatCo..13.4952L. doi:10.1038/s41467-022-32705-2. PMC 9399112. PMID 35999238.
  121. ^ Di Liberto, Tom (11 September 2014). "ENSO + Climate Change = Headache". ENSO Blog. Archived from the original on 18 April 2016.
  122. ^ Collins, Mat; An, Soon-Il; Cai, Wenju; Ganachaud, Alexandre; Guilyardi, Eric; Jin, Fei-Fei; Jochum, Markus; Lengaigne, Matthieu; Power, Scott; Timmermann, Axel; Vecchi, Gabe; Wittenberg, Andrew (23 May 2010). "The impact of global warming on the tropical Pacific Ocean and El Niño". Nature Geoscience. 3 (6): 391–397. Bibcode:2010NatGe...3..391C. doi:10.1038/ngeo868. Archived from the original on 14 September 2019. Retrieved 10 January 2019.
  123. ^ Trenberth, Kevin E.; Hoar, Timothy J. (January 1996). "The 1990–1995 El Niño–Southern Oscillation event: Longest on record". Geophysical Research Letters. 23 (1): 57–60. Bibcode:1996GeoRL..23...57T. CiteSeerX 10.1.1.54.3115. doi:10.1029/95GL03602.
  124. ^ Wittenberg, A.T. (2009). "Are historical records sufficient to constrain ENSO simulations?". Geophys. Res. Lett. 36 (12): L12702. Bibcode:2009GeoRL..3612702W. doi:10.1029/2009GL038710. S2CID 16619392.
  125. ^ Fedorov, Alexey V.; Philander, S. George (16 June 2000). "Is El Niño Changing?". Science. 288 (5473): 1997–2002. Bibcode:2000Sci...288.1997F. doi:10.1126/science.288.5473.1997. PMID 10856205. S2CID 5909976.
  126. ^ Zhang, Qiong; Guan, Yue; Yang, Haijun (2008). "ENSO Amplitude Change in Observation and Coupled Models". Advances in Atmospheric Sciences. 25 (3): 331–6. Bibcode:2008AdAtS..25..361Z. CiteSeerX 10.1.1.606.9579. doi:10.1007/s00376-008-0361-5. S2CID 55670859.
  127. ^ Logan, Tyne (18 May 2023). "El Niño and La Niña have become more extreme and frequent because of climate change, study finds". ABC. Archived from the original on 16 July 2023. Retrieved 17 July 2023.
  128. ^ Readfearn, Graham (18 May 2023). "Global heating has likely made El Niños and La Niñas more 'frequent and extreme', new study shows". The Guardian. Archived from the original on 16 July 2023. Retrieved 17 July 2023.
  129. ^ Cai, Wenju; Ng, Benjamin; Geng, Tao; Jia, Fan; Wu, Lixin; Wang, Guojian; Liu, Yu; Gan, Bolan; Yang, Kai; Santoso, Agus; Lin, Xiaopei; Li, Ziguang; Liu, Yi; Yang, Yun; Jin, Fei-Fei; Collins, Mat; McPhaden, Michael J. (June 2023). "Antropogenic impacts on twentieth - century ENSO variability changes". Nature Reviews Earth & Environment. 4 (6): 407–418. Bibcode:2023NRvEE...4..407C. doi:10.1038/s43017-023-00427-8. S2CID 258793531. Archived from the original on 17 July 2023. Retrieved 17 July 2023.
  130. ^ Lenton, T. M.; Held, H.; Kriegler, E.; Hall, J. W.; Lucht, W.; Rahmstorf, S.; Schellnhuber, H. J. (12 February 2008). "Tipping elements in the Earth's climate system". Proceedings of the National Academy of Sciences. 105 (6): 1786–1793. doi:10.1073/pnas.0705414105. PMC 2538841. PMID 18258748.
  131. ^ Simon Wang, S.-Y.; Huang, Wan-Ru; Hsu, Huang-Hsiung; Gillies, Robert R. (16 October 2015). "Role of the strengthened El Niño teleconnection in the May 2015 floods over the southern Great Plains". Geophysical Research Letters. 42 (19): 8140–8146. Bibcode:2015GeoRL..42.8140S. doi:10.1002/2015GL065211.
  132. ^ Roxy, Mathew Koll; Ritika, Kapoor; Terray, Pascal; Masson, Sébastien (15 November 2014). "The Curious Case of Indian Ocean Warming*,+" (PDF). Journal of Climate. 27 (22): 8501–8509. Bibcode:2014JCli...27.8501R. doi:10.1175/JCLI-D-14-00471.1. S2CID 42480067. Archived (PDF) from the original on 3 September 2019. Retrieved 10 January 2019.
  133. ^ Roxy, Mathew Koll; Ritika, Kapoor; Terray, Pascal; Murtugudde, Raghu; Ashok, Karumuri; Goswami, B. N. (November 2015). "Drying of Indian subcontinent by rapid Indian Ocean warming and a weakening land-sea thermal gradient". Nature Communications. 6 (1): 7423. Bibcode:2015NatCo...6.7423R. doi:10.1038/ncomms8423. PMID 26077934.
  134. ^ Lenton, Timothy M.; Held, Hermann; Kriegler, Elmar; Hall, Jim W; Lucht, Wolfgang; Rahmstorf, Stefan; Schellnhuber, Hans Joachim (2008-02-12). "Tipping elements in the Earth's climate system". PNAS. 105 (6): 1786–1793. Bibcode:2008PNAS..105.1786L. doi:10.1073/pnas.0705414105. PMC 2538841. PMID 18258748.
  135. ^ a b Wunderling, Nico; Donges, Jonathan F.; Kurths, Jürgen; Winkelmann, Ricarda (3 June 2021). "Interacting tipping elements increase risk of climate domino effects under global warming". Earth System Dynamics. 12 (2): 601–619. Bibcode:2021ESD....12..601W. doi:10.5194/esd-12-601-2021. ISSN 2190-4979. S2CID 236247596.
  136. ^ "Tipping Points: Why we might not be able to reverse climate change". ClimateScience. Retrieved 17 July 2022.
  137. ^ a b Duque-Villegas, Mateo; Salazar, Juan Fernando; Rendón, Angela Maria (2019). "Tipping the ENSO into a permanent El Niño can trigger state transitions in global terrestrial ecosystems". Earth System Dynamics. 10 (4): 631–650. Bibcode:2019ESD....10..631D. doi:10.5194/esd-10-631-2019. ISSN 2190-4979. S2CID 210348791.
  138. ^ Schellnhuber, Hans Joachim; Rahmstorf, Stefan; Winkelmann, Ricarda (2016). "Why the right climate target was agreed in Paris". Nature Climate Change. 6 (7): 649–653. Bibcode:2016NatCC...6..649S. doi:10.1038/nclimate3013. ISSN 1758-6798.
  139. ^ Arias, Paola A.; Bellouin, Nicolas; Coppola, Erika; Jones, Richard G.; et al. (2021). "Technical Summary" (PDF). IPCC AR6 WG1. p. 88.
  140. ^ Armstrong McKay, David (9 September 2022). "Exceeding 1.5°C global warming could trigger multiple climate tipping points – paper explainer". climatetippingpoints.info. Retrieved 2 October 2022.
  141. ^ "August Climate Bulletins / Summer 2023: the hottest on record". Copernicus Programme. 6 September 2023. Archived from the original on 8 September 2023.
  142. ^ Joint Typhoon Warning Center (2006). "3.3 JTWC Forecasting Philosophies" (PDF). Archived from the original (PDF) on 5 July 2012. Retrieved 11 February 2007.
  143. ^ a b Wu, M. C.; Chang, W. L.; Leung, W. M. (2004). "Impacts of El Niño–Southern Oscillation Events on Tropical Cyclone Landfalling Activity in the Western North Pacific". Journal of Climate. 17 (6): 1419–28. Bibcode:2004JCli...17.1419W. CiteSeerX 10.1.1.461.2391. doi:10.1175/1520-0442(2004)017<1419:ioenoe>2.0.co;2.
  144. ^ Patricola, Christina M.; Saravanan, R.; Chang, Ping (15 July 2014). "The Impact of the El Niño–Southern Oscillation and Atlantic Meridional Mode on Seasonal Atlantic Tropical Cyclone Activity". Journal of Climate. 27 (14): 5311–5328. Bibcode:2014JCli...27.5311P. doi:10.1175/JCLI-D-13-00687.1.
  145. ^ a b c d Landsea, Christopher W; Dorst, Neal M (1 June 2014). "Subject: G2) How does El Niño-Southern Oscillation affect tropical cyclone activity around the globe?". Tropical Cyclone Frequently Asked Question. United States National Oceanic and Atmospheric Administration's Hurricane Research Division. Archived from the original on 9 October 2014.
  146. ^ a b "Background Information: East Pacific Hurricane Outlook". United States Climate Prediction Center. 27 May 2015. Archived from the original on 9 May 2009. Retrieved 7 April 2016.
  147. ^ a b "What is El Niño and what might it mean for Australia?". Australian Bureau of Meteorology. Archived from the original on 18 March 2016. Retrieved 10 April 2016.
  148. ^ "Southwest Pacific Tropical Cyclone Outlook: El Niño expected to produce severe tropical storms in the Southwest Pacific" (Press release). New Zealand National Institute of Water and Atmospheric Research. 14 October 2015. Archived from the original on 12 December 2015. Retrieved 22 October 2014.
  149. ^ "El Nino is here!" (Press release). Tonga Ministry of Information and Communications. 11 November 2015. Archived from the original on 25 October 2017. Retrieved 8 May 2016.
  150. ^ Enfield, David B.; Mayer, Dennis A. (1997). "Tropical Atlantic sea surface temperature variability and its relation to El Niño–Southern Oscillation". Journal of Geophysical Research. 102 (C1): 929–945. Bibcode:1997JGR...102..929E. doi:10.1029/96JC03296.
  151. ^ Lee, Sang-Ki; Chunzai Wang (2008). "Why do some El Niños have no impact on tropical North Atlantic SST?". Geophysical Research Letters. 35 (L16705): L16705. Bibcode:2008GeoRL..3516705L. doi:10.1029/2008GL034734.
  152. ^ Latif, M.; Grötzner, A. (2000). "The equatorial Atlantic oscillation and its response to ENSO". Climate Dynamics. 16 (2–3): 213–218. Bibcode:2000ClDy...16..213L. doi:10.1007/s003820050014. S2CID 129356060.
  153. ^ Davis, Mike (2001). Late Victorian Holocausts: El Niño Famines and the Making of the Third World. London: Verso. p. 271. ISBN 978-1-85984-739-8.
  154. ^ WW2010 (28 April 1998). "El Niño". University of Illinois at Urbana-Champaign. Archived from the original on 19 September 2023. Retrieved 17 July 2009.{{cite web}}: CS1 maint: numeric names: authors list (link)
  155. ^ "El Niño Information". California Department of Fish and Game, Marine Region. Archived from the original on 2021-11-22. Retrieved 2024-01-18.
  156. ^ "Study reveals economic impact of El Niño". University of Cambridge. 11 July 2014. Archived from the original on 28 July 2014. Retrieved 25 July 2014.
  157. ^ Cashin, Paul; Mohaddes, Kamiar & Raissi, Mehdi (2014). "Fair Weather or Foul? The Macroeconomic Effects of El Niño" (PDF). Cambridge Working Papers in Economics. Archived from the original (PDF) on 28 July 2014.
  158. ^ "International Monetary Fund". Imf.org. Archived from the original on 1 March 2022. Retrieved 1 March 2022.
  159. ^ "El Niño and its health impact". allcountries.org. Archived from the original on 20 January 2011. Retrieved 10 October 2017.
  160. ^ "El Niño and its health impact". Health Topics A to Z. Archived from the original on 20 January 2011. Retrieved 1 January 2011.
  161. ^ Ballester, Joan; Jane C. Burns; Dan Cayan; Yosikazu Nakamura; Ritei Uehara; Xavier Rodó (2013). "Kawasaki disease and ENSO-driven wind circulation" (PDF). Geophysical Research Letters. 40 (10): 2284–2289. Bibcode:2013GeoRL..40.2284B. doi:10.1002/grl.50388. Archived (PDF) from the original on 2020-11-22. Retrieved 2024-01-18.
  162. ^ Rodó, Xavier; Joan Ballester; Dan Cayan; Marian E. Melish; Yoshikazu Nakamura; Ritei Uehara; Jane C. Burns (10 November 2011). "Association of Kawasaki disease with tropospheric wind patterns". Scientific Reports. 1: 152. Bibcode:2011NatSR...1E.152R. doi:10.1038/srep00152. ISSN 2045-2322. PMC 3240972. PMID 22355668.
  163. ^ Hsiang, S. M.; Meng, K. C.; Cane, M. A. (2011). "Civil conflicts are associated with the global climate". Nature. 476 (7361): 438–441. Bibcode:2011Natur.476..438H. doi:10.1038/nature10311. PMID 21866157. S2CID 4406478.
  164. ^ Quirin Schiermeier (2011). "Climate cycles drive civil war". Nature. 476: 406–407. doi:10.1038/news.2011.501.
  165. ^ França, Filipe; Ferreira, J; Vaz-de-Mello, FZ; Maia, LF; Berenguer, E; Palmeira, A; Fadini, R; Louzada, J; Braga, R; Oliveira, VH; Barlow, J (10 February 2020). "El Niño impacts on human-modified tropical forests: Consequences for dung beetle diversity and associated ecological processes". Biotropica. 52 (1): 252–262. Bibcode:2020Biotr..52..252F. doi:10.1111/btp.12756.
  166. ^ "El Niño increases seedling mortality even in drought-tolerant forests". ScienceDaily. Archived from the original on 2022-11-01. Retrieved 2022-11-01.
  167. ^ "FAQs | El Nino Theme Page – A comprehensive Resource". www.pmel.noaa.gov. Archived from the original on 13 November 2016. Retrieved 12 November 2016.
  168. ^ França, FM; Benkwitt, CE; Peralta, G; Robinson, JPW; Graham, NAJ; Tylianakis, JM; Berenguer, E; Lees, AC; Ferreira, J; Louzada, J; Barlow, J (2020). "Climatic and local stressor interactions threaten tropical forests and coral reefs". Philosophical Transactions of the Royal Society B. 375 (1794): 20190116. doi:10.1098/rstb.2019.0116. PMC 7017775. PMID 31983328.
  169. ^ a b c Barnston, Anthony (19 May 2014). "How ENSO leads to a cascade of global impacts". ENSO Blog. Archived from the original on 26 May 2016.
  170. ^ "Slow response to East Africa famine 'cost lives'". BBC News. 18 January 2012. Archived from the original on 4 April 2022. Retrieved 27 February 2022.
  171. ^ "La Niña weather likely to last for months". Scoop News (Scoop.co.nz). 12 October 2010. Archived from the original on 28 June 2011. Retrieved 27 February 2022.
  172. ^ "Southern Africa: El Niño, Positive Indian Ocean Dipole Forecast and Humanitarian Impact (October 2023)". reliefweb.int. OCHA. 16 October 2023. Retrieved 20 January 2024.
  173. ^ Brugnara, Yuri; Brönnimann, Stefan; Grab, Stefan; Steinkopf, Jessica; Burgdorf, Angela-Maria; Wilkinson, Clive; Allan, Rob (October 2023). "South African extreme weather during the 1877–1878 El Niño". Weather. 78 (10): 286–293. Bibcode:2023Wthr...78..286B. doi:10.1002/wea.4468.
  174. ^ a b Nhesvure, B. (2020). Impacts of ENSO on coastal South African sea surface temperatures. Faculty of Science, Department of Oceanography. Retrieved from http://hdl.handle.net/11427/32954/
  175. ^ Turner, John (2004). "The El Niño–Southern Oscillation and Antarctica". International Journal of Climatology. 24 (1): 1–31. Bibcode:2004IJCli..24....1T. doi:10.1002/joc.965. S2CID 129117190.
  176. ^ a b Yuan, Xiaojun (2004). "ENSO-related impacts on Antarctic sea ice: a synthesis of phenomenon and mechanisms". Antarctic Science. 16 (4): 415–425. Bibcode:2004AntSc..16..415Y. doi:10.1017/S0954102004002238. S2CID 128831185.
  177. ^ Barlow, M., H. Cullen, and B. Lyon, 2002: Drought in central and southwest Asia: La Niña, the warm pool, and Indian Ocean precipitation. J. Climate, 15, 697–700
  178. ^ Nazemosadat, M. J., and A. R. Ghasemi, 2004: Quantifying the ENSO-related shifts in the intensity and probability of drought and wet periods in Iran. J. Climate, 17, 4005–4018
  179. ^ "channelnewsasia.com - February 2010 is driest month for S'pore since records began in 1869". 3 March 2010. Archived from the original on 3 March 2010.
  180. ^ Wu, M. C.; Chang, W. L.; Leung, W. M. (2004). "Impacts of El Niño–Southern Oscillation events on tropical cyclone landfalling activity in the western north Pacific". Journal of Climate. 17 (6): 1419–1428. Bibcode:2004JCli...17.1419W. CiteSeerX 10.1.1.461.2391. doi:10.1175/1520-0442(2004)017<1419:ioenoe>2.0.co;2.
  181. ^ Hong, Lynda (13 March 2008). "Recent heavy rain not caused by global warming". Channel News Asia. Archived from the original on 14 May 2008. Retrieved 22 June 2008.
  182. ^ Power, Scott; Haylock, Malcolm; Colman, Rob; Wang, Xiangdong (1 October 2006). "The Predictability of Interdecadal Changes in ENSO Activity and ENSO Teleconnections". Journal of Climate. 19 (19): 4755–4771. Bibcode:2006JCli...19.4755P. doi:10.1175/JCLI3868.1. ISSN 0894-8755. S2CID 55572677.
  183. ^ a b c d "What is El Niño and what might it mean for Australia?". Australian Bureau of Meteorology. Archived from the original on 18 March 2016. Retrieved 10 April 2016.
  184. ^ The beasts to our east: What are El Ninos and La Ninas? By Peter Hannam from the Sydney Morning Herald, December 29, 2020.
  185. ^ La Niña In Australia Bureau of Meteorology. www.bom.gov.au
  186. ^ El Niño in Australia Bureau of Meteorology. www.bom.gov.au
  187. ^ "Climate Driver Update". Bureau of Meteorology. Bureau of Meteorology. 17 September 2023.
  188. ^ King, Andrew (13 September 2022). "La Niña, 3 years in a row: a climate scientist on what flood-weary Australians can expect this summer". The Conversation.
  189. ^ What is La Nina and what does it mean for your summer? By Peter Hannam and Laura Chung. The Sydney Morning Herald. November 25, 2021.
  190. ^ "Climate glossary — Southern Oscillation Index (SOI)". Bureau of Meteorology (Australia). 2002-04-03. Retrieved 2009-12-31.
  191. ^ Australian Climate Extremes – Fire, BOM. Retrieved 2 May 2007.
  192. ^ "What are the prospects for the weather in the coming winter?". Met Office News Blog. United Kingdom Met Office. 29 October 2015. Archived from the original on 20 April 2016.
  193. ^ Ineson, S.; Scaife, A. A. (7 December 2008). "The role of the stratosphere in the European climate response to El Niño". Nature Geoscience. 2 (1): 32–36. Bibcode:2009NatGe...2...32I. doi:10.1038/ngeo381.
  194. ^ "La Niña is coming. Here's what that means for winter weather in the U.S." NPR. 22 October 2021. Archived from the original on 20 December 2021. Retrieved 21 December 2021.
  195. ^ "ENSO Diagnostic Discussion". Climate Prediction Center. National Oceanic and Atmospheric Administration. 5 June 2014. Archived from the original on 26 June 2014.
  196. ^ San Juan, Puerto Rico Weather Forecast Office (2010-09-02). "The Local Impacts of ENSO across the Northeastern Caribbean". National Weather Service Southern Region Headquarters. Archived from the original on 2014-07-14. Retrieved 2014-07-01.
  197. ^ Climate Prediction Center. ENSO Impacts on United States Winter Precipitation and Temperature. Archived 2008-04-12 at the Wayback Machine Retrieved on 2008-04-16.
  198. ^ "A never-ending winter". Canada's top ten weather stories for 2008. Environment Canada. 2008-12-29. number 3. Archived from the original on 7 August 2011.
  199. ^ ENSO evolution, status, and forecasts (PDF). Climate Prediction Center (Report) (update ed.). National Oceanic and Atmospheric Administration. 2005-02-28. Archived from the original (PDF) on 2005-05-15.
  200. ^ "If la Niña continues, what does that mean for Oregon this summer?". 29 April 2022. Archived from the original on 26 April 2023. Retrieved 17 January 2024.
  201. ^ a b c Halpert, Mike (12 June 2014). "United States El Niño Impacts". ENSO Blog. Archived from the original on 26 May 2016.
  202. ^ Barnston, Anthony (12 June 2014). "With El Niño likely, what climate impacts are favored for this summer?". ENSO Blog. Archived from the original on 30 March 2016.
  203. ^ "El Niño: What are the El Niño impacts in Canada?". Environment and Climate Change Canada. 2 December 2015. Archived from the original on 22 March 2016.
  204. ^ Oetting, Jeremiah (11 May 2018). "El Nino "flavors" affect California rainfall". www.earthmagazine.org. Archived from the original on 2022-06-17. Retrieved 2022-04-18.
  205. ^ Lee, Sang-Ki; Lopez, Hosmay; Chung, Eui-Seok; DiNezio, Pedro; Yeh, Sang-Wook; Wittenberg, Andrew T. (2018-01-28). "On the Fragile Relationship Between El Niño and California Rainfall". Geophysical Research Letters. 45 (2): 907–915. Bibcode:2018GeoRL..45..907L. doi:10.1002/2017GL076197. ISSN 0094-8276. S2CID 35504261.
  206. ^ Monteverdi, John and Jan Null. WESTERN REGION TECHNICAL ATTACHMENT NO. 97-37 NOVEMBER 21, 1997: El Niño and California Precipitation. Archived December 27, 2009, at the Wayback Machine Retrieved on 2008-02-28.
  207. ^ Mantua, Nathan. La Niña Impacts in the Pacific Northwest. Archived 2007-10-22 at the Wayback Machine Retrieved on 2008-02-29.
  208. ^ Reuters. La Nina could mean dry summer in Midwest and Plains. Archived 2008-04-21 at the Wayback Machine Retrieved on 2008-02-29.
  209. ^ Climate Prediction Center. El Niño (ENSO) Related Rainfall Patterns Over the Tropical Pacific. Archived 2010-05-28 at the Wayback Machine Retrieved on 2008-02-28.
  210. ^ Romero-Centeno, Rosario; Zavala-Hidalgo, Jorge; Gallegos, Artemio; O'Brien, James J. (1 August 2003). "Isthmus of Tehuantepec Wind Climatology and ENSO Signal". Journal of Climate. 16 (15): 2628–2639. Bibcode:2003JCli...16.2628R. doi:10.1175/1520-0442(2003)016<2628:iotwca>2.0.co;2. S2CID 53654865.
  211. ^ American Meteorological Society (2012-01-26). "Tehuantepecer". Glossary of Meteorology. Archived from the original on 2014-01-11. Retrieved 2013-05-16.
  212. ^ Fett, Bob (2002-12-09). "World Wind Regimes – Central America Gap Wind Tutorial". United States Naval Research Laboratory Monterey, Marine Meteorology Division. Archived from the original on 2013-02-18. Retrieved 2013-05-16.
  213. ^ Arnerich, Paul A. "Tehuantepecer Winds of the West Coast of Mexico". Mariners Weather Log. 15 (2): 63–67.
  214. ^ Martínez-Ballesté, Andrea; Ezcurra, Exequiel (2018). "Reconstruction of past climatic events using oxygen isotopes in Washingtonia robusta growing in three anthropic oases in Baja California". Boletín de la Sociedad Geológica Mexicana. 70 (1): 79–94. doi:10.18268/BSGM2018v70n1a5.
  215. ^ a b "El Niño's impacts on New Zealand's climate". New Zealand's National Institute of Water and Atmospheric Research. 19 October 2015. Archived from the original on 19 March 2016. Retrieved 11 April 2016.
  216. ^ a b "ENSO Update, Weak La Nina Conditions Favoured" (PDF). Fiji Meteorological Service. Archived from the original (PDF) on 7 November 2017.
  217. ^ a b "Climate Summary January 2016" (PDF). Samoa Meteorology Division, Ministry of Natural Resources and Environment. January 2016. Archived (PDF) from the original on 10 April 2017. Retrieved 2021-05-02.
  218. ^ Chu, Pao-Shin. Hawaii Rainfall Anomalies and El Niño. Retrieved on 2008-03-19.
  219. ^ Pacific ENSO Applications Climate Center. Pacific ENSO Update: 4th Quarter, 2006. Vol. 12 No. 4. Archived 2012-10-22 at the Wayback Machine Retrieved on 2008-03-19.
  220. ^ Pacific ENSO Applications Climate Center. RAINFALL VARIATIONS DURING ENSO. Archived 2008-04-21 at the Wayback Machine Retrieved on 2008-03-19.
  221. ^ "Atmospheric Consequences of El Niño". University of Illinois. Archived from the original on 2014-10-06. Retrieved 2010-05-31.
  222. ^ "Atmospheric Consequences of El Niño". University of Illinois. Archived from the original on 6 October 2014. Retrieved 31 May 2010.
  223. ^ WW2010 (28 April 1998). "El Niño". University of Illinois at Urbana-Champaign. Archived from the original on 19 September 2023. Retrieved 17 July 2009.{{cite web}}: CS1 maint: numeric names: authors list (link)
  224. ^ "An El Niño Fish Tale". scied.ucar.edu. Archived from the original on 2023-12-14. Retrieved 2023-11-26.
  225. ^ Pearcy, W. G.; Schoener, A. (1987). "Changes in the marine biota coincident with the 1982-83 El Niño in the northeastern subarctic Pacific Ocean". Journal of Geophysical Research. 92 (C13): 14417–28. Bibcode:1987JGR....9214417P. doi:10.1029/JC092iC13p14417. Archived from the original on 22 September 2012. Retrieved 22 June 2008.
  226. ^ Sharma, P. D.; P.D, Sharma (2012). Ecology And Environment. Rastogi Publications. ISBN 978-81-7133-905-1. Archived from the original on 2024-01-20. Retrieved 2024-01-18.
  227. ^ a b "La Niña follows El Niño, the GLOBE El Niño Experiment continues". Archived from the original on 15 October 2011. Retrieved 31 May 2010.
  228. ^ van Valen, Gary (2013). Indigenous Agency in the Amazon. Tucson, Arizona: University of Arizona Press. p. 10.
  229. ^ "Biodiversity". Galapagos Conservancy. Archived from the original on June 28, 2022. Retrieved June 24, 2022.
  230. ^ Karnauskas, Kris. "El Niño and the Galapagos". Climate.gov. Archived from the original on November 29, 2022. Retrieved Nov 17, 2022.
  231. ^ Vargas (2006). "Biological effects of El Niño on the Galápagos penguin". Biological Conservation. 127 (1): 107–114. Bibcode:2006BCons.127..107V. doi:10.1016/j.biocon.2005.08.001.
  232. ^ Edgar (2010). "El Niño, grazers and fisheries interact to greatly elevate extinction risk for Galapagos marine species". Global Change Biology. 16 (10): 2876–2890. Bibcode:2010GCBio..16.2876E. doi:10.1111/j.1365-2486.2009.02117.x. S2CID 83795836.
  233. ^ Holmgren (2001). "El Niño effects on the dynamics of terrestrial ecosystems". Trends in Ecology and Evolution. 16 (2): 89–94. doi:10.1016/S0169-5347(00)02052-8. PMID 11165707.
  234. ^ a b Carrè, Matthieu; et al. (2005). "Strong El Niño events during the early Holocene: stable isotope evidence from Peruvian sea shells". The Holocene. 15 (1): 42–7. Bibcode:2005Holoc..15...42C. doi:10.1191/0959683605h1782rp. S2CID 128967433.
  235. ^ Brian Fagan (1999). Floods, Famines and Emperors: El Niño and the Fate of Civilizations. Basic Books. pp. 119–138. ISBN 978-0-465-01120-9.
  236. ^ Grove, Richard H. (1998). "Global Impact of the 1789–93 El Niño". Nature. 393 (6683): 318–9. Bibcode:1998Natur.393..318G. doi:10.1038/30636. S2CID 205000683.
  237. ^ Ó Gráda, C. (2009). "Ch. 1: The Third Horseman". Famine: A Short History. Princeton University Press. ISBN 9780691147970. Archived from the original on 12 January 2016. Retrieved 3 March 2010.
  238. ^ "Dimensions of need - People and populations at risk". Fao.org. Archived from the original on 10 October 2017. Retrieved 28 July 2015.
  239. ^ Lartigue (1827). Description de la Côte Du Pérou, Entre 19° et 16° 20' de Latitude Sud, ... [Description of the Coast of Peru, Between 19° and 16° 20' South Latitude, ...] (in French). Paris, France: L'Imprimerie Royale. pp. 22–23. Archived from the original on 2024-01-20. Retrieved 2024-01-18. From pp. 22–23: "Il est néanmoins nécessaire, au sujet de cette règle générale, de faire part d'une exception ... dépassé le port de sa destination de plus de 2 ou 3 lieues; ... " (It is nevertheless necessary, with regard to this general rule, to announce an exception which, in some circumstances, might shorten the sailing. One said above that the breeze was sometimes quite fresh [i.e., strong], and that then the counter-current, which bore southward along the land, stretched some miles in length; it is obvious that one will have to tack in this counter-current, whenever the wind's force will permit it and whenever one will not have gone past the port of one's destination by more than 2 or 3 leagues; ...)
  240. ^ a b Pezet, Federico Alfonso (1896), "The Counter-Current "El Niño," on the Coast of Northern Peru", Report of the Sixth International Geographical Congress: Held in London, 1895, Volume 6, pp. 603–606
  241. ^ Findlay, Alexander G. (1851). A Directory for the Navigation of the Pacific Ocean -- Part II. The Islands, Etc., of the Pacific Ocean. London: R. H. Laurie. p. 1233. M. Lartigue is among the first who noticed a counter or southerly current.
  242. ^ "Droughts in Australia: Their causes, duration, and effect: The views of three government astronomers [R.L.J. Ellery, H.C. Russell, and C. Todd]," The Australasian (Melbourne, Victoria), 29 December 1888, pp. 1455–1456. From p. 1456: Archived 16 September 2017 at the Wayback Machine "Australian and Indian Weather" : "Comparing our records with those of India, I find a close correspondence or similarity of seasons with regard to the prevalence of drought, and there can be little or no doubt that severe droughts occur as a rule simultaneously over the two countries."
  243. ^ Lockyer, N. and Lockyer, W.J.S. (1904) "The behavior of the short-period atmospheric pressure variation over the Earth's surface," Archived 2023-04-03 at the Wayback Machine Proceedings of the Royal Society of London, 73 : 457–470.
  244. ^ Eguiguren, D. Victor (1894) "Las lluvias de Piura" Archived 2023-10-30 at the Wayback Machine (The rains of Piura), Boletín de la Sociedad Geográfica de Lima, 4 : 241–258. [in Spanish] From p. 257: Archived 2023-10-30 at the Wayback Machine "Finalmente, la época en que se presenta la corriente de Niño, es la misma de las lluvias en aquella región." (Finally, the period in which the El Niño current is present is the same as that of the rains in that region [i.e., the city of Piura, Peru].)
  245. ^ Pezet, Federico Alfonso (1896) "La contra-corriente "El Niño", en la costa norte de Perú" Archived 2023-10-30 at the Wayback Machine (The counter-current "El Niño", on the northern coast of Peru), Boletín de la Sociedad Geográfica de Lima, 5 : 457-461. [in Spanish]
  246. ^ Walker, G. T. (1924) "Correlation in seasonal variations of weather. IX. A further study of world weather," Memoirs of the Indian Meteorological Department, 24 : 275–332. From p. 283: "There is also a slight tendency two quarters later towards an increase of pressure in S. America and of Peninsula [i.e., Indian] rainfall, and a decrease of pressure in Australia : this is part of the main oscillation described in the previous paper* which will in future be called the 'southern' oscillation." Available at: Royal Meteorological Society Archived 18 March 2017 at the Wayback Machine
  247. ^ Cushman, Gregory T. "Who Discovered the El Niño-Southern Oscillation?". Presidential Symposium on the History of the Atmospheric Sciences: People, Discoveries, and Technologies. American Meteorological Society (AMS). Archived from the original on 1 December 2015. Retrieved 18 December 2015.
  248. ^ "The El Niño Phenomenon Returns". Wild Singapore. Archived from the original on April 3, 2023. Retrieved May 8, 2022.
  249. ^ Sinamaw Zeleke Wallie (January 2019). Economic Impact from El Niños (Thesis). Debark University. Archived from the original on April 3, 2023. Retrieved May 8, 2022 – via Academia.Edu.
  250. ^ Trenberth, Kevin E.; Hoar, Timothy J. (January 1996). "The 1990–95 El Niño–Southern Oscillation event: Longest on record". Geophysical Research Letters. 23 (1): 57–60. Bibcode:1996GeoRL..23...57T. CiteSeerX 10.1.1.54.3115. doi:10.1029/95GL03602.
  251. ^ Trenberth, K. E.; et al. (2002). "Evolution of El Niño – Southern Oscillation and global atmospheric surface temperatures". Journal of Geophysical Research. 107 (D8): 4065. Bibcode:2002JGRD..107.4065T. CiteSeerX 10.1.1.167.1208. doi:10.1029/2000JD000298.
  252. ^ Marshall, Paul; Schuttenberg, Heidi (2006). A reef manager's guide to coral bleaching. Townsville, Qld.: Great Barrier Reef Marine Park Authority. ISBN 978-1-876945-40-4. Archived from the original on 2023-07-30. Retrieved 2024-01-18.
  253. ^ a b "El Niño 2016". Atavist. 6 October 2015. Archived from the original on 26 February 2018. Retrieved 18 January 2024.
  254. ^ Willis, Katherine J; Araújo, Miguel B; Bennett, Keith D; Figueroa-Rangel, Blanca; Froyd, Cynthia A; Myers, Norman (28 February 2007). "How can a knowledge of the past help to conserve the future? Biodiversity conservation and the relevance of long-term ecological studies". Philosophical Transactions of the Royal Society B: Biological Sciences. 362 (1478): 175–187. doi:10.1098/rstb.2006.1977. PMC 2311423. PMID 17255027.
  255. ^ Corrège, Thierry; Delcroix, Thierry; Récy, Jacques; Beck, Warren; Cabioch, Guy; Le Cornec, Florence (August 2000). "Evidence for stronger El Niño-Southern Oscillation (ENSO) Events in a Mid-Holocene massive coral". Paleoceanography. 15 (4): 465–470. Bibcode:2000PalOc..15..465C. doi:10.1029/1999pa000409.
  256. ^ Seillès, Brice; Sánchez Goñi, Maria Fernanda; Ledru, Marie-Pierre; Urrego, Dunia H; Martinez, Philippe; Hanquiez, Vincent; Schneider, Ralph (April 2016). "Holocene land–sea climatic links on the equatorial Pacific coast (Bay of Guayaquil, Ecuador)". The Holocene. 26 (4): 567–577. Bibcode:2016Holoc..26..567S. doi:10.1177/0959683615612566. hdl:10871/18307. S2CID 130306658.
  257. ^ Rodbell, Donald T.; Seltzer, Geoffrey O.; Anderson, David M.; Abbott, Mark B.; Enfield, David B.; Newman, Jeremy H. (22 January 1999). "An ~15,000-Year Record of El Niño-Driven Alluviation in Southwestern Ecuador". Science. 283 (5401): 516–520. Bibcode:1999Sci...283..516R. doi:10.1126/science.283.5401.516. PMID 9915694. S2CID 13714632.
  258. ^ Moy, Christopher M.; Seltzer, Geoffrey O.; Rodbell, Donald T.; Anderson, David M. (2002). "Variability of El Niño/Southern Oscillation activity at millennial timescales during the Holocene epoch". Nature. 420 (6912): 162–165. Bibcode:2002Natur.420..162M. doi:10.1038/nature01194. PMID 12432388. S2CID 4395030.
  259. ^ Turney, Chris S. M.; Kershaw, A. Peter; Clemens, Steven C.; Branch, Nick; Moss, Patrick T.; Fifield, L. Keith (2004). "Millennial and orbital variations of El Niño/Southern Oscillation and high-latitude climate in the last glacial period". Nature. 428 (6980): 306–310. Bibcode:2004Natur.428..306T. doi:10.1038/nature02386. PMID 15029193. S2CID 4303100.
  260. ^ Beaufort, Luc; Garidel-Thoron, Thibault de; Mix, Alan C.; Pisias, Nicklas G. (28 September 2001). "ENSO-like Forcing on Oceanic Primary Production During the Late Pleistocene". Science. 293 (5539): 2440–2444. Bibcode:2001Sci...293.2440B. doi:10.1126/science.293.5539.2440. PMID 11577233.
  261. ^ Muñoz, Arsenio; Ojeda, Jorge; Sánchez-Valverde, Belén (2002). "Sunspot-like and ENSO/NAO-like periodicities in lacustrinelaminated sediments of the Pliocene Villarroya Basin (La Rioja,Spain)". Journal of Paleolimnology. 27 (4): 453–463. Bibcode:2002JPall..27..453M. doi:10.1023/a:1020319923164. S2CID 127610981.
  262. ^ Wara, Michael W.; Ravelo, Ana Christina; Delaney, Margaret L. (29 July 2005). "Permanent El Niño-Like Conditions During the Pliocene Warm Period". Science. 309 (5735): 758–761. Bibcode:2005Sci...309..758W. CiteSeerX 10.1.1.400.7297. doi:10.1126/science.1112596. PMID 15976271. S2CID 37042990.
  263. ^ Fedorov, Alexey V.; Brierley, Christopher M.; Emanuel, Kerry (February 2010). "Tropical cyclones and permanent El Niño in the early Pliocene epoch". Nature. 463 (7284): 1066–1070. Bibcode:2010Natur.463.1066F. doi:10.1038/nature08831. hdl:1721.1/63099. PMID 20182509. S2CID 4330367.
  264. ^ Galeotti, Simone; von der Heydt, Anna; Huber, Matthew; Bice, David; Dijkstra, Henk; Jilbert, Tom; Lanci, Luca; Reichart, Gert-Jan (May 2010). "Evidence for active El Niño Southern Oscillation variability in the Late Miocene greenhouse climate". Geology. 38 (5): 419–422. Bibcode:2010Geo....38..419G. doi:10.1130/g30629.1. S2CID 140682002.
  265. ^ Madden, Roland A.; Julian, Paul R. (1971-07-01). "Detection of a 40–50 Day Oscillation in the Zonal Wind in the Tropical Pacific". Journal of the Atmospheric Sciences. 28 (5): 702–708. Bibcode:1971JAtS...28..702M. doi:10.1175/1520-0469(1971)028<0702:DOADOI>2.0.CO;2. ISSN 0022-4928.
  266. ^ Zhang, Chidong (2005). "Madden-Julian Oscillation". Rev. Geophys. 43 (2): RG2003. Bibcode:2005RvGeo..43.2003Z. CiteSeerX 10.1.1.546.5531. doi:10.1029/2004RG000158. S2CID 33003839.
  267. ^ "Madden-Julian oscillation forecast research". University of East Anglia. Archived from the original on 9 March 2012. Retrieved 22 February 2012.
  268. ^ Jon Gottschalck & Wayne Higgins (2008-02-16). "Madden Julian Oscillation Impacts" (PDF). Climate Prediction Center. Retrieved 2009-07-17.
  269. ^ Roundy, P.E.; Kiladis, G.N. (2007). "Analysis of a Reconstructed Oceanic Kelvin Wave Dynamic Height Dataset for the Period 1974–2005". J. Climate. 20 (17): 4341–55. Bibcode:2007JCli...20.4341R. doi:10.1175/JCLI4249.1.
  270. ^ Roundy, P.E.; Kravitz, J.R. (2009). "The Association of the Evolution of Intraseasonal Oscillations to ENSO Phase". J. Climate. 22 (2): 381–395. Bibcode:2009JCli...22..381R. doi:10.1175/2008JCLI2389.1.
  271. ^ Liu, Zhengyu; Alexander Michael (2007). "Atmospheric bridge, oceanic tunnel, and global climate teleconnections". Reviews of Geophysics. 45 (2): 2. Bibcode:2007RvGeo..45.2005L. doi:10.1029/2005RG000172.
  272. ^ Stuecker 2018, p. 1.
  273. ^ Zhang, Clement & Di Nezio 2014, p. 772.
  274. ^ Amaya 2019, p. 302.
  275. ^ You & Furtado 2018, p. 10145.
  276. ^ Middlemas et al. 2019, p. 4670.
  277. ^ Ding et al. 2020, p. 2.
  278. ^ Zheng et al. 2018, p. 2198.
  279. ^ Zhang, Clement & Di Nezio 2014, p. 773.
  280. ^ a b Amaya 2019, p. 303.
  281. ^ Zhang, Clement & Di Nezio 2014, p. 781.
  282. ^ You & Furtado 2018, p. 10142.
  283. ^ Maeda et al. 2016, p. 17.
  284. ^ Dewitte et al. 2021, p. 239.
  285. ^ Kim et al. 2022, p. 4.

Sources[edit]

External links[edit]