Paleotempestology

Last updated

Paleotempestology is the study of past tropical cyclone activity by means of geological proxies as well as historical documentary records. The term was coined by American meteorologist Kerry Emanuel.

Contents

The usual approach in paleotempestology is the identification of deposits left by storms. Most commonly, these are overwash deposits in waterbodies close to the coast; other means are oxygen isotope ratio variations caused by tropical cyclone rainfall in trees or speleothems (cave deposits), and identifying beach ridges kicked up by storm waves. The occurrence rate of tropical cyclones can then be inferred from these deposits and sometimes also their intensity – typically the stronger events are the most easily recognizable ones –, by comparing them to deposits left by historical events.

Paleotempestological research has shown that in the Coast of the Gulf of Mexico and in Australia, the occurrence rate of intense tropical cyclones is about once every few centuries, and there are long-term variations in occurrence which are caused, for example, by shifts in their paths. Common problems in paleotempestology are confounding factors such as tsunami-generated deposits, and the fact that only some parts of the world have been investigated.

Definition and rationale

Paleotempestology is the estimation of tropical cyclone activity with the help of proxy data. The name was coined by Kerry Emanuel of the Massachusetts Institute of Technology; [1] the field has seen increased activity since the 1990s [2] and studies were first carried out in the United States of America [3] on the East Coast. [4]

The realisation that one cannot rely solely on historical records to infer past storm activity was a major driving force for the development of paleotempestology. [5] The historical record in many places is too short (one century at most) to properly determine the hazard produced by tropical cyclones, especially the rare very intense ones [1] which at times are undersampled by historical records; [6] in the United States, for example, only about 150 years of record are available, and only a small number of hurricanes classified as category 4 or 5 – the most destructive ones on the Saffir-Simpson scale – have come ashore, making it difficult to estimate the hazard level. [7] Such records may also not be representative for future weather patterns. [8] [9]

Information about past tropical cyclone occurrences can be used to constrain how their occurrences may change in the future, or about how they respond to large-scale climate modes, such as sea surface temperature changes. [1] In general, the origin and behaviour of tropical cyclone systems is poorly understood, [10] and there is concern that human-caused global warming will increase the intensity of tropical cyclones and the frequency of strong events by increasing sea surface temperatures. [11] [8]

Techniques

In general, paleotempestology is a complex field of science that overlaps with other disciplines like climatology and coastal geomorphology. [12] A number of techniques have been used to estimate the past hazards from tropical cyclones. [7] Many of these techniques have also been applied to studying extratropical storms, although research on this field is less advanced than on tropical cyclones. [4]

Overwash deposits

Overwash deposits in atolls, coastal lakes, marshes or reef flats are the most important paleoclimatological evidence of tropical cyclone strikes. When storms hit these areas, currents and waves can overtop barriers, erode these and other beach structures, and lay down deposits in the water bodies behind barriers. [13] [2] [14] Isolated breaches and especially widespread overtopping of coastal barriers during storms can generate fan-like, layered deposits behind the barrier. Individual layers can be correlated to particular storms in favourable circumstances; in addition they are often separated by a clear boundary from earlier sediments. [11] Such deposits have been observed in North Carolina after Hurricane Isabel in 2003, for example. [15] The intensity [3] and impacts of the tropical cyclone can also be inferred from overwash deposits [16] by comparing the deposits to these formed by known storms [3] and analyzing their lithology (their physical characteristics). [17] Additionally, thicker sediment layers usually correspond to stronger storm systems. [3] This procedure is not always clear-cut however. [18]

Several techniques have been applied to separate out storm overwash deposits from other sediments:

Generally, sites suitable for obtaining paleotempestology records are not found along the entire length of the coastline, [19] and depending on the properties of the site such as vegetation cover, [28] they might only track storms approaching from a certain direction. [17] Prerequisites for successful correlation of overwash deposits to tropical cyclones are: [29]

Dating and intensity determination

Various dating techniques can then be used to produce a chronology of tropical cyclone strikes at a given location and thus a recurrence rate; [2] [14] for example, at Lake Shelby in Alabama a return period of once every 318 years was determined. The storms in the Lake Shelby record have windspeeds of over 190 kilometres per hour (120 mph) [31] as Hurricane Ivan which in 2004 made landfall in the region at that intensity did not leave a deposit. [32] Based on geological considerations the minimum windspeed of storms recorded there might be 230 kilometres per hour (143 mph). [31]

For dating purposes radiometric dating procedures involving carbon-14, cesium-137, and lead-210 are most commonly used, often in combination. [26] Uranium series dating, [33] optically stimulated luminescence, [34] and correlations to human land use can also be used in some places. [20]

Beach ridges

Beach ridges and cheniers [2] form when storm surges, storm waves or tides deposit debris in ridges, with one ridge typically corresponding to one storm. [35] Ridges can be formed by coral rubble where coral reefs lie at the coast, [36] and can contain complicated layer structures, [37] shells, [38] pumice, [39] and gravel. [40] A known example is the ridge that Cyclone Bebe generated on Funafuti atoll in 1971. [41]

Beach ridges are common on the deltaic shores of China, and are indicative of increased typhoon activity. [3] They have also been found on the Australian coast facing the Great Barrier Reef and are formed from reworked corals. The height of each ridge appears to correlate with the intensity of the storm that produced it, and thus the intensity of the forming storm can be inferred by numerical modelling and comparison to known storms [42] and known storm surges. [43] Ridges tend to be older the farther inland they are; [44] they can also be dated through optically stimulated luminescence [45] and radiocarbon dating. [39] In addition, no tsunami-generated beach ridges have been observed, and tsunamis are important confounding factors in paleotempestology. [46]

Wind-driven erosion or accumulation can alter the elevation of such ridges, and, in addition, the same ridge can be formed by more than one storm event [47] as has been observed in Australia. [48] Beach ridges can also shift around through non-storm processes after their formation [44] and can form through non-tropical cyclone processes. [49] Sedimentary texture can be used to infer the origin of a ridge from storm surges. [50]

Isotope ratios

Precipitation in tropical cyclones has a characteristic isotope composition with a depletion of heavy oxygen isotopes; carbon and nitrogen isotope data have also been used to infer tropical cyclone activity. [51] Corals can store oxygen isotope ratios which in turn reflect water temperatures, precipitation and evaporation; [52] these in turn can be related to tropical cyclone activity. [53] Fish otoliths and bivalves can also store such records, [54] as can trees where the oxygen isotope ratios of precipitation are reflected in the cellulose of trees, and can be inferred with the help of tree rings. [51] However, confounding factors like natural variation and soil properties also influence oxygen isotope ratios of tree cellulose. For these reasons, only the frequency of storms can be reliably estimated from tree ring isotopic records, not their intensity. [23]

Speleothems, deposits formed in caves through the dissolution and redeposition of dolomite and limestone, can store isotope signatures associated with tropical cyclones, especially in fast growing speleothems, areas with thin soils and speleothems which have undergone little alteration. Such deposits have a high temporal resolution, and are also protected from many confounding factors [23] although the extraction of annual layers has become possible only recently, with a two-week resolution (two separate layers correlated to two hurricanes that struck two weeks apart) achieved in one case. [55] However, the suitability of speleothems depends on the characteristics of the cave they are found in; caves that flood frequently may have their speleothems eroded or otherwise damaged, for example, making them less suitable for paleotempestology research. [56] Caves where speleothems form mainly during the offseason are also likely to miss tropical cyclones. [57] Very old records can be obtained from oxygen isotope ratios in rocks. [58]

Other techniques

Historical documents such as county gazettes in China, diaries, logbooks of travellers, official histories and old newspapers can contain information on tropical cyclones. [59] In China such records go back over a millennium, [3] while elsewhere it is usually confined to the last 130 years. [60] Such historical records however are often ambiguous or unclear, [1] they only record landfalling storms and sometimes confuse non-tropical systems or intense convective storms for tropical cyclones. [61] The frequency of shipwrecks has been used to infer past tropical cyclone occurrence, [17] such as has been done with a database of shipwrecks that the Spaniards suffered in the Caribbean. [62]

Aside from oxygen isotope ratios, [51] tree rings can also record information on storm-caused plant damage or vegetation changes, [63] such as thin tree rings due to storm-induced damage to a tree canopy, and saltwater intrusion and the resulting slowdown in tree growth. The term "dendrotempestology" is used in this context. [64] [62] [65] Speleothems can also store trace elements which can signal tropical cyclone activity [66] and mud layers formed by storm-induced cave flooding. [56] Droughts on the other hand can cause groundwater levels to drop enough that subsequent storms cannot induce flooding and thus fail to leave a record, as has been noted in Yucatan. [67]

Other techniques:

Timespans

A database of tropical cyclones going back to 6,000 BC has been compiled for the western North Atlantic Ocean. [84] In the Gulf of Mexico, records go back five millennia [14] but only a few typhoon [lower-alpha 1] records go back 5,000–6,000 years. [33] In general, tropical cyclone records do not go farther back than 5,000–6,000 years ago when the Holocene sea level rise levelled off; tropical cyclone deposits formed during sea level lowstands likely were reworked during sea level rise. Only tentative evidence exists of deposits from the last interglacial. [86] Tempestite deposits [87] and oxygen isotope ratios in much older rocks have also been used to infer the existence of tropical cyclone activity [58] as far back as the Jurassic. [87]

Results

Paleotempestological information has been used by the insurance industry in risk analysis [88] in order to set insurance rates. [65] The industry has also funded paleotempestological research. [89] Paleotempestology information is further of interest to archeologists, ecologists, and forest and water resource managers. [90]

Recurrence rates

The recurrence rate, the time gap between storms, is an important metric used to estimate tropical cyclone risk, and it can be determined by paleotempestological research. In the Gulf of Mexico, catastrophic hurricane strikes at given locations occur once about every 350 years in the last 3,800 years [14] or about 0.48%–0.39% annual frequency at any given site, [91] with a recurrence rate of 300 years or 0.33% annual probability at sites in the Caribbean and Gulf of Mexico; [92] category 3 or more storms occur at a rate of 3.9–0.1 category 3 or more storms per century in the northern Gulf of Mexico. [93] Elsewhere, tropical cyclones with intensities of category 4 or more occur about every 350 years in the Pearl River Delta (China), [94] one storm every 100–150 years at Funafuti and a similar rate in French Polynesia, [76] one category 3 or stronger every 471 years in St. Catherines Island (Georgia), [95] 0.3% each year for an intense storm in eastern Hainan, [96] one storm every 140–180 years in Nicaragua, [97] one intense storm every 200–300 years in the Great Barrier Reef [42] – formerly their recurrence rate was estimated to be one strong event every few millennia [98] – and one storm of category 2–4 intensity [99] every 190–270 years at Shark Bay in Western Australia. [100] Steady rates have been found for the Gulf of Mexico and the Coral Sea [101] for timespans of several millennia. [91]

However, it has also been found that the occurrence rates of tropical cyclone measured with instrumental data over historical time can be significantly different from the actual occurrent rate. In the past, tropical cyclones were far more frequent in the Great Barrier Reef [42] and the northern Gulf of Mexico than today; [102] in Apalachee Bay, strong storms occur every 40 years, not every 400 years as documented historically. [103] Serious storms in New York occurred twice in 300 years [104] not once every millennium or less. [105] In general, the area of Australia appears to be unusually inactive in recent times by the standards of the past 550–1500 years, [106] and the historical record underestimates the incidence of strong storms in Northeastern Australia. [107]

Long term fluctuations

Long-term variations of tropical cyclone activity have also been found. The Gulf of Mexico saw increased activity between 3,800 and 1,000 years ago with a fivefold increase of category 4–5 hurricane activity, [108] and activity at St. Catherines Island and Wassaw Island was also higher between 2,000 and 1,100 years ago. [109] This appears to be a stage of increased tropical cyclone activity spanning the region from New York to Puerto Rico, [110] while the last 1,000 years have been inactive both there and in the Gulf Coast. [111] Before 1400 AD, the Caribbean and the Gulf of Mexico were active while the East Coast of the United States was inactive, followed by a reversal that lasted until 1675 AD; [112] in an alternative interpretation, the US Atlantic coast and the Caribbean saw low activity between 950 AD and 1700 with a sudden increase around 1700. [33] It is unclear whether in the Atlantic hurricane activity is more regionally modulated or basin-wide. [113] Such fluctuations appear to mainly concern strong tropical cyclone systems, at least in the Atlantic; weaker systems have a more steady pattern of activity. [114] Rapid fluctuations over short timespans have also been observed. [90]

In the Atlantic Ocean, the so-called "Bermuda High" hypothesis stipulates that changes in the position of this anticyclone can cause storm paths to alternate between landfalls on the East Coast and the Gulf Coast [11] [115] but also Nicaragua. [116] Paleotempestological data support this theory [117] although additional findings on Long Island and Puerto Rico have demonstrated that storm frequency is more complex [111] as active periods appear to correlate between the three sites. [118] A southward shift of the High has been inferred to have occurred 3,000 [119] –1,000 years ago, [120] and has been linked with the "hurricane hyperactivity" period in the Gulf of Mexico between 3,400 and 1,000 years ago. [121] Conversely a decrease in hurricane activity is recorded after the mid-millennium period [122] and after 1,100 the Atlantic changes from a pattern of widespread activity to a more geographically confined one. [123] Between 1,100-1,450 the Bahamas and the Florida Gulf Coast were frequently struck while between 1,450-1,650 activity was higher in New England. [124] Furthermore, a tendency to a more northerly storm track may be associated with a strong North Atlantic Oscillation [125] while the Neoglacial cooling is associated with a southward shift. [121] In West Asia, high activity in the South China Sea coincides with low activity in Japan and vice versa. [126] [127]

Role of climate modes

The influence of natural trends on tropical cyclone activity has been recognized in paleotempestology records, such as a correlation between Atlantic hurricane tracks [128] and activity with the status of the ITCZ; [129] [130] [131] position of the Loop Current (for Gulf of Mexico hurricanes); [91] North Atlantic Oscillation; sea surface temperatures [132] and the strength of the West African Monsoon; [133] and Australian cyclone activity and the Pacific Decadal Oscillation. [134] Increased insolation – either from solar activity [135] or from orbital variations – have been found to be detrimental to tropical cyclone activity in some regions. [136] In the first millennium AD, warmer sea surface temperatures in the Atlantic as well as more restricted anomalies may be responsible for stronger regional hurricane activity. [137] The climate mode dependency of tropical cyclone activity may be more pronounced in temperate regions where tropical cyclones find less favourable conditions. [138]

Among the known climate modes that influence tropical cyclone activity in paleotempestological records are ENSO phase variations, which influence tropical cyclone activity in Australia and the Atlantic, [139] but also their path as has been noted for typhoons. [140] [141] [142] [143] More general global correlations have been found, such as a negative correlation between tropical cyclone activity in Japan and the North Atlantic [136] and correlation between the Atlantic and Australia on the one hand [144] and between Australia and French Polynesia on the other hand. [145]

Influence of long-term temperature variations

The effect of general climate variations have also been found. Hurricane [146] and typhoon tracks tend to shift north (e.g. Amur Bay) during warm periods and south (e.g. South China) during cold periods, [61] [147] patterns that might be mediated by shifts in the subtropical anticyclones. [111] These patterns (northward shift with warming) has been observed as a consequence of human-induced global warming and the end of the Little Ice Age [146] but also after volcanic eruptions (southward shift with cooling); [148] some volcanic eruptions have been linked to decreased hurricane activity, although this observation is not universal. [149]

The Dark Ages Cold Period has been linked to decreased activity off Belize. [150] Initially the Medieval Climate Anomaly featured increased activity across the Atlantic, but later activity decreased along the US East Coast. [151] During the 1350 to present interval in the Little Ice Age, there were more but weaker storms in the Gulf of Mexico [152] while hurricane activity did not decrease in western Long Island. [118] Colder waters may have impeded tropical cyclone activity in the Gulf of Mexico during the Little Ice Age. [153] Increased hurricane activity during the last 300 years in the Caribbean may also correlate to the Little Ice Age. [154] The Little Ice Age may have been accompanied by more but weaker storms in the South China Sea relative to preceding or following periods. [155] [156]

The response of tropical cyclones to future global warming is of great interest. The Holocene Climatic Optimum did not induce increased tropical cyclone strikes in Queensland and phases of higher hurricane activity on the Gulf Coast are not associated with global warming; [33] however warming has been correlated with typhoon activity in the Gulf of Thailand [157] and marine warming with typhoon activity in the South China Sea, [158] increased hurricane activity in Belize (which increased during the Medieval Warm Period) [159] and during the Mesozoic when carbon dioxide caused warming episodes [87] such as the Toarcian anoxic event. [160]

After-effects of tropical cyclones

A correlation between hurricane strikes and subsequent wildfire activity [161] and vegetation changes has been noted in the Alabamian [162] and Cuban paleotempestological record. [163] In St. Catherines Island, cultural activity ceased at the time of increased storm activity, [164] and both Taino settlement of the Bahamas [92] and Polynesian expansion across the Pacific may have been correlated to decreased tropical cyclone activity. [145] Tropical cyclone induced alteration in oxygen isotope ratios may mask isotope ratio variations caused by other climate phenomena, which may thus be misinterpreted. [165]

On the other hand, the Classic Maya collapse may or may not coincide with, and have been caused by, a decrease in tropical cyclone activity. [166] [167] Tropical cyclones are important for preventing droughts in the southeastern US. [168] Paleotempestology has found evidence that the Kamikaze typhoons that impeded the Mongol invasions of Japan did, in fact, exist. [169]

Other patterns

Sites in the Bahamas show more strong storms in the northern Bahamas than the southern ones, presumably because storms approaching the southern Bahamas have passed over the Greater Antilles before and have lost much of their intensity there. [170] Atmospheric conditions favourable for tropical cyclone activity in the "main development region" [lower-alpha 2] of the Atlantic are correlated to unfavourable conditions along the East Coast. [172] The anti-correlation between Gulf of Mexico and Bahamas activity with the US East Coast activity may be due to active hurricane seasons - which tend to increase storm activity in the former - being accompanied by unfavourable climatological conditions along the East Coast. [173]

Problems

Paleotempestological reconstructions are subject to a number of limitations, [24] including the presence of sites suited for the obtainment of paleotempestological records, [19] changes in the hydrological properties of the site due to e.g. sea level rise [24] which increases the sensitivity to weaker storms [174] and "false positives" caused by for example non-tropical cyclone-related floods, sediment winnowing, wind-driven transport, tides, tsunamis, [24] bioturbation [17] and non-tropical storms such as nor'easters [175] or winter storm, the latter of which however usually result in lower surges. [176] In particular, tsunamis are a problem for paleotempestological studies in the Indian and Pacific Ocean; [177] one technique that has been used to differentiate the two is the identification of traces of runoff which occurs during storms but not during tsunamis. [178] Coastal paleotempestology records are based on storm surge, and do not always reflect wind speeds, [179] e.g in large and slow-moving storms. [180]

Not all of the world has been investigated with paleotempestological methods; among the places thus researched are Belize, the Carolinas of North America, northern coasts of the Gulf of Mexico, the northeastern United States, [19] (in a lesser measure) the South Pacific islands and tropical Australia. [60] Conversely China, [181] Cuba, Florida, Hispaniola, Honduras, the Lesser Antilles and North America north of Canada are poorly researched. The presence of research institutions active in paleotempestology and suitable sites for paleotempestological research and tropical cyclone landfalls may influence whether a given location is researched or not. [19] In the Atlantic Ocean, research has been concentrated on regions where hurricanes are common rather than more marginal areas. [182]

Paleotempestology records mostly record activity during the Holocene [181] and tend to record mainly catastrophic storms as these are the ones most likely to leave evidence. [6] In addition, as of 2017 there has been little effort in making comprehensive databases of paleotempestological data or in attempting regional reconstructions from local results. [182] Different sites have different intensity thresholds and thus capture different storm populations, [151] and the same layer can be caused by a landfall of a weaker storm closer to the site or a landfall at a larger distance of a stronger storm. [183]

Also, paleotempestological records, especially overwash records in marshes, are often grossly incomplete with questionable geochronology. Deposition mechanism are poorly documented, and it is often not clear how to identify storm deposits. [184] The magnitude of overwash deposits is fundamentally a function of storm surge height, which, however, is not a function of storm intensity. [74] Overwash deposits are regulated by the height of the overwashed barrier and there is no expectation that it will remain stable over time; [185] tropical cyclones themselves have been observed eroding such barriers [186] and such barrier height decreases (e.g. through storm erosion or sea level rise) may induce a spurious increase of tropical cyclone deposits over time. [187] Successive overwash deposits can be difficult to distinguish, and they are easily eroded by subsequent storms. [188] Storm deposits can vary strongly even a short distance from the landfall point, [189] even over few tens of metres, [190] and changes in tropical cyclone activity recorded at one site might simply reflect the stochastic nature of tropical cyclone landfalls. [172] In particular, in core tropical cyclone activity regions weather variations rather than large-scale modes may control tropical cyclone activity. [191]

Application to non-tropical storms

Paleotempestological research has been mostly carried out in low-latitude regions [192] but research in past storm activity has been conducted in the British Isles, France and the Mediterranean. [193] Increases in storm activity on the European Atlantic coast have been noted AD 1350–1650, AD 250–850, AD 950–550, 1550–1350 BC, 3550–3150 BC, and 5750–5150 BC. [194] In southern France, a recurrence rate of 0.2% per year of catastrophic storms has been inferred for the last 2,000 years. [195]

Storm records indicate increased storm activity during colder periods such as the Little Ice Age, Medieval Dark Age and Iron Age Cold Epoch. [196] During cold periods, increased temperature gradients between the polar and low-latitude regions increase baroclinic storm activity. Changes in the North Atlantic Oscillation may also play a role. [195]

Examples

See also

Notes

  1. Typhoons are tropical cyclones in the West Pacific. [85]
  2. The "main development region" is an area between 10° and 20° northern latitude and between 20° and 60° western longitude in the Atlantic where numerous hurricanes form. [171]

Related Research Articles

<span class="mw-page-title-main">Subtropical cyclone</span> Cyclonic storm with both tropical and extratropical characteristics

A subtropical cyclone is a weather system that has some characteristics of both tropical and extratropical cyclones.

This is a list of all known or suspected Atlantic hurricanes up to 1599. Although most storms likely went unrecorded, and many records have been lost, recollections of hurricane occurrences survive from some sufficiently populated coastal areas, and rarely, ships at sea that survived the tempests.

<span class="mw-page-title-main">Typhoon</span> Types of tropical cyclone that develops in the Northern Hemisphere going to the eastern side

A typhoon is a tropical cyclone that develops between 180° and 100°E in the Northern Hemisphere. This region is referred to as the Northwestern Pacific Basin, accounting for almost one-third of the world's annual tropical cyclones. The term hurricane refers to a tropical cyclone in the northcentral and northeast Pacific, and the north Atlantic. For organizational purposes, the northern Pacific Ocean is divided into three regions: the eastern, central, and western. The Regional Specialized Meteorological Center (RSMC) for tropical cyclone forecasts is in Japan, with other tropical cyclone warning centres for the northwest Pacific in Hawaii, the Philippines, and Hong Kong. Although the RSMC names each system, the main name list itself is coordinated among 18 countries that have territories threatened by typhoons each year.

<span class="mw-page-title-main">Atlantic hurricane</span> Tropical cyclone that forms in the Atlantic Ocean

An Atlantic hurricane is a type of tropical cyclone that forms in the Atlantic Ocean primarily between June and November. The terms "hurricane", "typhoon", and "cyclone" can be used interchangeably to describe this weather phenomenon. These storms are rotating, organized systems of clouds and thunderstorms that originate over tropical or subtropical waters and have closed low-level circulation, not to be confused with tornadoes. They form over low pressure systems. In the North Atlantic, central North Pacific, and eastern North Pacific, the term "hurricane" is mainly used, whereas "typhoon" is more commonly used for storms originating in the western North Pacific. The term "cyclone" is used in the South Pacific and Indian Ocean.

<span class="mw-page-title-main">Eye (cyclone)</span> Central area of calm weather in a tropical cyclone

The eye is a region of mostly calm weather at the center of a tropical cyclone. The eye of a storm is a roughly circular area, typically 30–65 kilometers in diameter. It is surrounded by the eyewall, a ring of towering thunderstorms where the most severe weather and highest winds of the cyclone occur. The cyclone's lowest barometric pressure occurs in the eye and can be as much as 15 percent lower than the pressure outside the storm.

<span class="mw-page-title-main">Extratropical cyclone</span> Type of cyclone

Extratropical cyclones, sometimes called mid-latitude cyclones or wave cyclones, are low-pressure areas which, along with the anticyclones of high-pressure areas, drive the weather over much of the Earth. Extratropical cyclones are capable of producing anything from cloudiness and mild showers to severe gales, thunderstorms, blizzards, and tornadoes. These types of cyclones are defined as large scale (synoptic) low pressure weather systems that occur in the middle latitudes of the Earth. In contrast with tropical cyclones, extratropical cyclones produce rapid changes in temperature and dew point along broad lines, called weather fronts, about the center of the cyclone.

<span class="mw-page-title-main">Tropical cyclone</span> Rapidly rotating storm system

A tropical cyclone is a rapidly rotating storm system characterized by a low-pressure center, a closed low-level atmospheric circulation, strong winds, and a spiral arrangement of thunderstorms that produce heavy rain and squalls. Depending on its location and strength, a tropical cyclone is referred to by different names, including hurricane, typhoon, tropical storm, cyclonic storm, tropical depression, or simply cyclone. A hurricane is a strong tropical cyclone that occurs in the Atlantic Ocean or northeastern Pacific Ocean, and a typhoon occurs in the northwestern Pacific Ocean. In the Indian Ocean and South Pacific, comparable storms are referred to as "tropical cyclones". In modern times, on average around 80 to 90 named tropical cyclones form each year around the world, over half of which develop hurricane-force winds of 65 kn or more. Tropical cyclones carry heat and energy away from the tropics and transport it towards temperate latitudes, which plays an important role in regulating global climate.

<span class="mw-page-title-main">Tropical cyclone observation</span>

Tropical cyclone observation has been carried out over the past couple of centuries in various ways. The passage of typhoons, hurricanes, as well as other tropical cyclones have been detected by word of mouth from sailors recently coming to port or by radio transmissions from ships at sea, from sediment deposits in near shore estuaries, to the wiping out of cities near the coastline. Since World War II, advances in technology have included using planes to survey the ocean basins, satellites to monitor the world's oceans from outer space using a variety of methods, radars to monitor their progress near the coastline, and recently the introduction of unmanned aerial vehicles to penetrate storms. Recent studies have concentrated on studying hurricane impacts lying within rocks or near shore lake sediments, which are branches of a new field known as paleotempestology. This article details the various methods employed in the creation of the hurricane database, as well as reconstructions necessary for reanalysis of past storms used in projects such as the Atlantic hurricane reanalysis.

<span class="mw-page-title-main">Tropical cyclone basins</span> Areas of tropical cyclone formation

Traditionally, areas of tropical cyclone formation are divided into seven basins. These include the north Atlantic Ocean, the eastern and western parts of the northern Pacific Ocean, the southwestern Pacific, the southwestern and southeastern Indian Oceans, and the northern Indian Ocean. The western Pacific is the most active and the north Indian the least active. An average of 86 tropical cyclones of tropical storm intensity form annually worldwide, with 47 reaching hurricane/typhoon strength, and 20 becoming intense tropical cyclones, super typhoons, or major hurricanes.

<span class="mw-page-title-main">Effects of tropical cyclones</span> Events including rain, wind, storm surge, and tornadoes

The effects of tropical cyclones include heavy rain, strong wind, large storm surges near landfall, and tornadoes. The destruction from a tropical cyclone, such as a hurricane or tropical storm, depends mainly on its intensity, its size, and its location. Tropical cyclones remove forest canopy as well as change the landscape near coastal areas, by moving and reshaping sand dunes and causing extensive erosion along the coast. Even well inland, heavy rainfall can lead to landslides in mountainous areas. Their effects can be sensed over time by studying the concentration of the Oxygen-18 isotope within caves.

<span class="mw-page-title-main">Hurricane Alley</span> Area of warm water in the Atlantic Ocean

Hurricane Alley is an area of warm water in the Atlantic Ocean stretching from the west coast of northern Africa to the east coast of Central America and Gulf Coast of the Southern United States. Many hurricanes form within this area. The sea surface temperature of the Atlantic in Hurricane Alley has grown slightly warmer over the past decades. A particularly warm summer in 2005 led climate scientists to begin studying whether this trend would lead to an increase in hurricane activity.

<span class="mw-page-title-main">Eyewall replacement cycle</span> Meteorological process around and within the eye of intense tropical cyclones

In meteorology, eyewall replacement cycles, also called concentric eyewall cycles, naturally occur in intense tropical cyclones, generally with winds greater than 185 km/h (115 mph), or major hurricanes. When tropical cyclones reach this intensity, and the eyewall contracts or is already small, some of the outer rainbands may strengthen and organize into a ring of thunderstorms—a new, outer eyewall—that slowly moves inward and robs the original, inner eyewall of its needed moisture and angular momentum. Since the strongest winds are in a tropical cyclone's eyewall, the storm usually weakens during this phase, as the inner wall is "choked" by the outer wall. Eventually the outer eyewall replaces the inner one completely, and the storm may re-intensify.

<span class="mw-page-title-main">Tropical cyclones and climate change</span> Impact of climate change on tropical cyclones

Climate change can affect tropical cyclones in a variety of ways: an intensification of rainfall and wind speed, a decrease in overall frequency, an increase in the frequency of very intense storms and a poleward extension of where the cyclones reach maximum intensity are among the possible consequences of human-induced climate change. Tropical cyclones use warm, moist air as their source of energy or "fuel". As climate change is warming ocean temperatures, there is potentially more of this fuel available.

<span class="mw-page-title-main">Automated Tropical Cyclone Forecasting System</span> Software used to predict and forecast tropical cyclogenesis and to track tropical cyclones

The Automated Tropical Cyclone Forecasting System (ATCF) is a piece of software originally developed to run on a personal computer for the Joint Typhoon Warning Center (JTWC) in 1988, and the National Hurricane Center (NHC) in 1990. ATCF remains the main piece of forecasting software used for the United States Government, including the JTWC, NHC, and Central Pacific Hurricane Center. Other tropical cyclone centers in Australia and Canada developed similar software in the 1990s. The data files with ATCF lie within three decks, known as the a-, b-, and f-decks. The a-decks include forecast information, the b-decks contain a history of center fixes at synoptic hours, and the f-decks include the various fixes made by various analysis center at various times. In the years since its introduction, it has been adapted to Unix and Linux platforms.

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

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

<span class="mw-page-title-main">Tropical cyclone effects by region</span> Tropical cyclone effects and impacts

Tropical cyclones regularly affect the coastlines of most of Earth's major bodies of water along the Atlantic, Pacific, and Indian oceans. Also known as hurricanes, typhoons, or other names, tropical cyclones have caused significant destruction and loss of human life, resulting in about 2 million deaths since the 19th century. Powerful cyclones that make landfall – moving from the ocean to over land – are some of the most impactful, although that is not always the case. An average of 86 tropical cyclones of tropical storm intensity form annually worldwide, with 47 reaching hurricane/typhoon strength, and 20 becoming intense tropical cyclones, super typhoons, or major hurricanes.

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

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

References

Citations

  1. 1 2 3 4 Oliva, Peros & Viau 2017, p. 172.
  2. 1 2 3 4 5 6 Fan & Liu 2008, p. 2908.
  3. 1 2 3 4 5 6 7 Fan & Liu 2008, p. 2910.
  4. 1 2 Goslin & Clemmensen 2017, p. 81.
  5. Oliva et al. 2018, p. 1664.
  6. 1 2 Frappier et al. 2007, p. 529.
  7. 1 2 Liu 2004, p. 444.
  8. 1 2 Donnelly et al. 2014, p. 2.
  9. Frappier et al. 2007, p. 530.
  10. Donnelly 2009, p. 763.
  11. 1 2 3 Donnelly 2009, p. 764.
  12. Liu 2004, p. 447.
  13. 1 2 Xiong et al. 2018, p. 150.
  14. 1 2 3 4 Liu 2004, p. 445.
  15. Liu 2010, p. 11.
  16. Fan & Liu 2008, p. 2909.
  17. 1 2 3 4 Bregy et al. 2018, p. 28.
  18. Oliva et al. 2018, p. 90.
  19. 1 2 3 4 5 6 Oliva, Peros & Viau 2017, p. 173.
  20. 1 2 3 Oliva, Peros & Viau 2017, p. 180.
  21. Oliva, Peros & Viau 2017, pp. 179–180.
  22. Oliva, Peros & Viau 2017, p. 177.
  23. 1 2 3 Oliva, Peros & Viau 2017, p. 182.
  24. 1 2 3 4 Oliva, Peros & Viau 2017, p. 183.
  25. Han et al. 2023, p. 3.
  26. 1 2 Oliva, Peros & Viau 2017, p. 178.
  27. Hippensteel & Garcia 2014, p. 1170.
  28. Xiong et al. 2018, p. 155.
  29. 1 2 3 4 5 Donnelly et al. 2014, p. 8.
  30. Harris, Martin & Hippensteel 2005, p. 1033.
  31. 1 2 Elsner, Jagger & Liu 2008, p. 368.
  32. Elsner, Jagger & Liu 2008, p. 369.
  33. 1 2 3 4 Fan & Liu 2008, p. 2917.
  34. Brill et al. 2017, p. 135.
  35. Hayne & Nott 2001, p. 509.
  36. Nott 2015, p. 130.
  37. Nott 2015, p. 133.
  38. Nott 2015, p. 139.
  39. 1 2 Nott 2015, p. 141.
  40. Nott 2015, p. 140.
  41. Nott 2004, p. 435.
  42. 1 2 3 Fan & Liu 2008, p. 2911.
  43. Nott 2015, p. 144.
  44. 1 2 Nott 2015, p. 134.
  45. Nott 2015, p. 136.
  46. Brückner et al. 2016, p. 2819.
  47. Goslin & Clemmensen 2017, p. 88,91.
  48. Nott 2015, p. 135.
  49. 1 2 Nott 2004, p. 437.
  50. Nott 2015, p. 138.
  51. 1 2 3 Oliva, Peros & Viau 2017, p. 181.
  52. Zinke et al. 2008, p. 11.
  53. Zinke et al. 2008, p. 13.
  54. Frappier et al. 2007, p. 533.
  55. Fan & Liu 2008, p. 2914.
  56. 1 2 Frappier et al. 2014, p. 5149.
  57. James, Banner & Hardt 2015.
  58. 1 2 Kolodny, Calvo & Rosenfeld 2009, p. 387.
  59. Liu 2004, pp. 444–445.
  60. 1 2 Nott 2004, p. 433.
  61. 1 2 Han et al. 2023, p. 4.
  62. 1 2 3 Domínguez-Delmás, Harley & Trouet 2016, p. 3169.
  63. Knapp, Maxwell & Soulé 2016, p. 312.
  64. Grissino-Mayer, Miller & Mora 2010, p. 291.
  65. 1 2 3 Travis 2000, p. 3.
  66. Frappier et al. 2007, p. 532.
  67. Frappier et al. 2014, p. 5152.
  68. Fan & Liu 2008, p. 2912.
  69. Frappier et al. 2007, p. 531.
  70. Nott 2004, p. 438.
  71. Liu 2010, p. 9.
  72. Xiong et al. 2018, p. 152.
  73. Woodruff, Donnelly & Okusu 2009, p. 1774.
  74. 1 2 Xiong et al. 2018, p. 157.
  75. 1 2 3 Donnelly et al. 2014, p. 6.
  76. 1 2 Ford et al. 2018, p. 918.
  77. Goslin & Clemmensen 2017, p. 91.
  78. Goslin & Clemmensen 2017, p. 93.
  79. Goslin & Clemmensen 2017, p. 95.
  80. Brandon et al. 2013, p. 2994.
  81. Astakhov et al. 2019, pp. 62–63.
  82. Harris, Martin & Hippensteel 2005, p. 1034.
  83. Han et al. 2023, p. 5.
  84. Oliva et al. 2018, p. 1665.
  85. Astakhov et al. 2015, p. 383.
  86. Nott 2004, p. 434.
  87. 1 2 3 Krencker et al. 2015, p. 129.
  88. Liu 2004, p. 446.
  89. Travis 2000, p. 2.
  90. 1 2 Frappier et al. 2007, p. 534.
  91. 1 2 3 Bregy et al. 2018, p. 39.
  92. 1 2 Park 2012, p. 900.
  93. Williams 2013, p. 181.
  94. Fan & Liu 2008, p. 2913.
  95. Braun et al. 2017, p. 370.
  96. Zhou et al. 2019, pp. 14–15.
  97. McCloskey & Liu 2012, p. 462.
  98. Hayne & Nott 2001, p. 510.
  99. Nott 2011b, p. 722.
  100. Nott 2011b, p. 713.
  101. Nott 2004, p. 441.
  102. Liu 2010, p. 59.
  103. Muller et al. 2017, p. 23.
  104. Sullivan et al. 2014, p. 7.
  105. Sullivan et al. 2014, p. 1.
  106. Muller et al. 2017, p. 5.
  107. Muller et al. 2017, p. 9.
  108. Williams 2013, p. 170.
  109. Braun et al. 2017, p. 366.
  110. Braun et al. 2017, p. 371.
  111. 1 2 3 Fan & Liu 2008, p. 2918.
  112. Wallace et al. 2019, p. 4.
  113. Yao et al. 2020, p. 15.
  114. McCloskey & Liu 2013, p. 279.
  115. Liu 2010, p. 36.
  116. McCloskey & Liu 2012, p. 463.
  117. Liu 2010, p. 39.
  118. 1 2 Scileppi & Donnelly 2007, p. 22.
  119. Volin et al. 2013, p. 17215.
  120. Peros et al. 2015, p. 1492.
  121. 1 2 Park 2012, p. 892.
  122. Rodysill et al. 2020, p. 7.
  123. Wallace et al. 2021, p. 19.
  124. Wallace et al. 2021, p. 2.
  125. Liu 2010, p. 37.
  126. Yue et al. 2019, p. 68.
  127. Zhou et al. 2019, p. 11.
  128. van Hengstum et al. 2014, p. 112.
  129. Wallace et al. 2019, p. 8.
  130. Muller et al. 2017, p. 36.
  131. van Hengstum et al. 2016, p. 7.
  132. Muller et al. 2017, p. 21.
  133. van Hengstum et al. 2014, p. 110-111.
  134. Liu et al. 2016, p. 66.
  135. Haig & Nott 2016, p. 2849.
  136. 1 2 Muller et al. 2017, p. 17.
  137. Donnelly et al. 2015, p. 50.
  138. Wallace et al. 2020, p. 14.
  139. Denniston et al. 2015, p. 4578-4579.
  140. Zhou et al. 2017, p. 7.
  141. Cook et al. 2015, pp. 3–4.
  142. Zhou et al. 2019, p. 2.
  143. Yang et al. 2020, p. 2248.
  144. Nott & Forsyth 2012, p. 4.
  145. 1 2 Toomey, Donnelly & Tierney 2016, p. 501.
  146. 1 2 Breitenbach et al. 2016, p. 6.
  147. Astakhov et al. 2019, p. 69.
  148. Breitenbach et al. 2016, p. 5.
  149. Muller et al. 2017, pp. 26–28.
  150. Schmitt et al. 2020, p. 11.
  151. 1 2 Wallace et al. 2021, p. 16.
  152. van Hengstum et al. 2015, p. 53.
  153. Rodysill et al. 2020, p. 9.
  154. LeBlanc et al. 2017, p. 147.
  155. Xiong et al. 2020, p. 1702.
  156. Tao et al. 2021, p. 3.
  157. Williams et al. 2016, p. 75.
  158. Yue et al. 2019, p. 69.
  159. Droxler, Bentley & Denommee 2014, p. 5.
  160. Krencker et al. 2015, p. 120.
  161. Liu 2010, p. 45.
  162. Liu 2010, p. 46.
  163. Peros et al. 2015, p. 1493.
  164. Braun et al. 2017, p. 367.
  165. Frappier 2013, p. 3642.
  166. Smith 2023, p. 36.
  167. Medina-Elizalde et al. 2016, p. 1.
  168. Knapp, Maxwell & Soulé 2016, pp. 319–320.
  169. Smith 2023, p. 41.
  170. Wallace et al. 2021, p. 14.
  171. Ercolani et al. 2015, p. 17.
  172. 1 2 Wallace et al. 2019, p. 5.
  173. Wallace et al. 2021, p. 17.
  174. Liu 2010, p. 14.
  175. Oliva, Peros & Viau 2017, p. 185.
  176. Liu 2010, p. 15.
  177. Astakhov et al. 2019, p. 62.
  178. Chagué-Goff et al. 2016, p. 346.
  179. Zhu et al. 2023, p. 2024.
  180. Zhu et al. 2023, p. 2026.
  181. 1 2 Du et al. 2016, p. 78.
  182. 1 2 Oliva, Peros & Viau 2017, p. 184.
  183. Smith 2023, p. 31.
  184. Hippensteel 2010, p. 52.
  185. Nott 2004, p. 439.
  186. Nott 2004, p. 440.
  187. Donnelly et al. 2014, p. 9.
  188. Chaumillon et al. 2017, p. 164.
  189. Harris, Martin & Hippensteel 2005, p. 1028.
  190. Hippensteel & Garcia 2014, p. 1169.
  191. Wallace et al. 2020, p. 13.
  192. Dezileau et al. 2011, p. 290.
  193. Pouzet et al. 2018, p. 432.
  194. Pouzet et al. 2018, p. 446.
  195. 1 2 Dezileau et al. 2011, p. 295.
  196. Pouzet et al. 2018, p. 445.

General sources

Further reading