Tropical cyclone

Page semi-protected
Source: Wikipedia, the free encyclopedia.
(Redirected from
Hurricanes
)

View of a tropical cyclone from space
Example of tropical cyclone Hurricane Florence in 2018 as viewed from space: The eye, eyewall, and surrounding rainbands are characteristics of tropical cyclones.

A tropical cyclone is a rapidly rotating

hurricane-force winds of 65 kn (120 km/h; 75 mph) or more.[1] 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
.

Tropical cyclones

horizontal temperature contrasts
. Tropical cyclones are typically between 100 and 2,000 km (62 and 1,243 mi) in diameter.

The strong rotating winds of a tropical cyclone are a result of the

Western Pacific Warm Pool
.

The primary energy source for these storms is warm ocean waters. These storms are therefore typically strongest when over or near water, and they weaken quite rapidly over land. This causes inland regions to be much less vulnerable to cyclones than coastal regions, with residents of tropical islands facing the greatest threat of all, although tidal flooding is often worse on continental coasts than on islands. Coastal damage may be caused by strong winds and rain, high waves (due to winds), storm surges (due to wind and severe pressure changes), and the potential of spawning tornadoes.

Tropical cyclones draw in air from a large area and concentrate the water content of that air (from atmospheric moisture and moisture evaporated from water) into

flooding, overland flooding, and a general overwhelming of local water control structures across a large area. The effects of tropical cyclones on human populations can be devastating, Every year tropical cyclones affect various regions of the globe including the Gulf Coast of North America, Australia, India, and Bangladesh
.

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.[2]

Definition and terminology

A tropical cyclone is the generic term for a warm-cored, non-frontal synoptic-scale low-pressure system over tropical or subtropical waters around the world.[3][4] The systems generally have a well-defined center which is surrounded by deep atmospheric convection and a closed wind circulation at the surface.[3] A tropical cyclone is generally deemed to have formed once mean surface winds in excess of 35 kn (65 km/h; 40 mph) are observed.[1] It is assumed at this stage that a tropical cyclone has become self-sustaining and can continue to intensify without any help from its environment.[1]

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", and such storms in the Indian Ocean can also be called "severe cyclonic storms".

Tropical refers to the geographical origin of these systems, which form almost exclusively over

Coriolis effect
.

Formation

A schematic diagram of a tropical cyclone
Diagram of a tropical cyclone in the Northern Hemisphere

Tropical cyclones tend to develop during the summer, but have been noted in nearly every month in most tropical cyclone basins. Tropical cyclones on either side of the Equator generally have their origins in the Intertropical Convergence Zone, where winds blow from either the northeast or southeast.[5] Within this broad area of low-pressure, air is heated over the warm tropical ocean and rises in discrete parcels, which causes thundery showers to form.[5] These showers dissipate quite quickly; however, they can group together into large clusters of thunderstorms.[5] This creates a flow of warm, moist, rapidly rising air, which starts to rotate cyclonically as it interacts with the rotation of the earth.[5]

Several factors are required for these thunderstorms to develop further, including

sea surface temperatures of around 27 °C (81 °F) and low vertical wind shear surrounding the system,[5][6] atmospheric instability, high humidity in the lower to middle levels of the troposphere, enough Coriolis force to develop a low-pressure center, and a pre-existing low-level focus or disturbance.[6]
There is a limit on tropical cyclone intensity which is strongly related to the water temperatures along its path.[7] and upper-level divergence.[8] An average of 86 tropical cyclones of tropical storm intensity form annually worldwide. Of those, 47 reach strength higher than 119 km/h (74 mph), and 20 become intense tropical cyclones (at least Category 3 intensity on the Saffir–Simpson scale).[9]

Climate oscillations such as El Niño–Southern Oscillation (ENSO) and the Madden–Julian oscillation modulate the timing and frequency of tropical cyclone development.[10][11][12][13] Rossby waves can aid in the formation of a new tropical cyclone by disseminating the energy of an existing, mature storm.[14][15] Kelvin waves can contribute to tropical cyclone formation by regulating the development of the westerlies.[16] Cyclone formation is usually reduced 3 days prior to the wave's crest and increased during the 3 days after.[17]

Formation regions and warning centers

Tropical cyclone basins and official warning centers
Basin Warning center Area of responsibility Notes
Northern Hemisphere
North Atlantic United States National Hurricane Center (Miami) Equator northward, African Coast – 140°W [18]
Eastern Pacific United States Central Pacific Hurricane Center (Honolulu) Equator northward, 140–180°W [18]
Western Pacific Japan Meteorological Agency Equator – 60°N, 180–100°E [19]
North Indian Ocean India Meteorological Department Equator northwards, 100–40°E [20]
Southern Hemisphere
South-West
Indian Ocean
Météo-France Reunion Equator – 40°S, African Coast – 90°E [21]
Australian region Indonesian Meteorology, Climatology,
and Geophysical Agency
(BMKG)
Equator – 10°S, 90–141°E [22]
Papua New Guinea National Weather Service Equator – 10°S, 141–160°E [22]
Australian Bureau of Meteorology 10–40°S, 90–160°E [22]
Southern Pacific Fiji Meteorological Service Equator – 25°S, 160°E – 120°W [22]
Meteorological Service of New Zealand 25–40°S, 160°E – 120°W [22]

The majority of tropical cyclones each year form in one of seven tropical cyclone basins, which are monitored by a variety of meteorological services and warning centres.

United States Government.[1] The Brazilian Navy Hydrographic Center names South Atlantic tropical cyclones, however the South Atlantic is not a major basin, and not an official basin according to the WMO.[23]

Intensity

Tropical cyclone intensity is based on wind speeds and pressure; relationships between winds and pressure are often used in determining the intensity of a storm.[24] Tropical cyclone scales such as the Saffir-Simpson Hurricane Wind Scale and Australia's scale (Bureau of Meteorology) only use wind speed for determining the category of a storm.[25][26] The most intense storm on record is Typhoon Tip in the northwestern Pacific Ocean in 1979, which reached a minimum pressure of 870 hPa (26 inHg) and maximum sustained wind speeds of 165 kn (85 m/s; 305 km/h; 190 mph).[27] The highest maximum sustained wind speed ever recorded was 185 kn (95 m/s; 345 km/h; 215 mph) in Hurricane Patricia in 2015—the most intense cyclone ever recorded in the Western Hemisphere.[28]

Factors that influence intensity

Warm sea surface temperatures are required in order for tropical cyclones to form and strengthen. The commonly-accepted minimum temperature range for this to occur is 26–27 °C (79–81 °F), however, multiple studies have proposed a lower minimum of 25.5 °C (77.9 °F).[29][30] Higher sea surface temperatures result in faster intensification rates and sometimes even rapid intensification.[31] High ocean heat content, also known as Tropical Cyclone Heat Potential, allows storms to achieve a higher intensity.[32] Most tropical cyclones that experience rapid intensification are traversing regions of high ocean heat content rather than lower values.[33] High ocean heat content values can help to offset the oceanic cooling caused by the passage of a tropical cyclone, limiting the effect this cooling has on the storm.[34] Faster-moving systems are able to intensify to higher intensities with lower ocean heat content values. Slower-moving systems require higher values of ocean heat content to achieve the same intensity.[33]

The passage of a tropical cyclone over the ocean causes the upper layers of the ocean to cool substantially, a process known as upwelling,[35] which can negatively influence subsequent cyclone development. This cooling is primarily caused by wind-driven mixing of cold water from deeper in the ocean with the warm surface waters. This effect results in a negative feedback process that can inhibit further development or lead to weakening. Additional cooling may come in the form of cold water from falling raindrops (this is because the atmosphere is cooler at higher altitudes). Cloud cover may also play a role in cooling the ocean, by shielding the ocean surface from direct sunlight before and slightly after the storm passage. All these effects can combine to produce a dramatic drop in sea surface temperature over a large area in just a few days.[36] Conversely, the mixing of the sea can result in heat being inserted in deeper waters, with potential effects on global climate.[37]

Vertical wind shear decreases tropical cyclone predicability, with storms exhibiting wide range of responses in the presence of shear.[38] Wind shear often negatively affects tropical cyclone intensification by displacing moisture and heat from a system's center.[39] Low levels of vertical wind shear are most optimal for strengthening, while stronger wind shear induces weakening.[40][41] Dry air entraining into a tropical cyclone's core has a negative effect on its development and intensity by diminishing atmospheric convection and introducing asymmetries in the storm's structure.[42][43][44] Symmetric, strong outflow leads to a faster rate of intensification than observed in other systems by mitigating local wind shear.[45][46][47] Weakening outflow is associated with the weakening of rainbands within a tropical cyclone.[48] Tropical cyclones may still intensify, even rapidly, in the presence of moderate or strong wind shear depending on the evolution and structure of the storm's convection.[49][50]

The size of tropical cyclones plays a role in how quickly they intensify. Smaller tropical cyclones are more prone to rapid intensification than larger ones.[51] The Fujiwhara effect, which involves interaction between two tropical cyclones, can weaken and ultimately result in the dissipation of the weaker of two tropical cyclones by reducing the organization of the system's convection and imparting horizontal wind shear.[52] Tropical cyclones typically weaken while situated over a landmass because conditions are often unfavorable as a result of the lack of oceanic forcing.[53] The Brown ocean effect can allow a tropical cyclone to maintain or increase its intensity following landfall, in cases where there has been copious rainfall, through the release of latent heat from the saturated soil.[54] Orographic lift can cause an significant increase in the intensity of the convection of a tropical cyclone when its eye moves over a mountain, breaking the capped boundary layer that had been restraining it.[55] Jet streams can both enhance and inhibit tropical cyclone intensity by influencing the storm's outflow as well as vertical wind shear.[56][57]

Rapid intensification

On occasion, tropical cyclones may undergo a process known as rapid intensification, a period in which the maximum sustained winds of a tropical cyclone increase by 30 kn (56 km/h; 35 mph) or more within 24 hours.[58] Similarly, rapid deepening in tropical cyclones is defined as a minimum sea surface pressure decrease of 1.75 hPa (0.052 inHg) per hour or 42 hPa (1.2 inHg) within a 24-hour period; explosive deepening occurs when the surface pressure decreases by 2.5 hPa (0.074 inHg) per hour for at least 12 hours or 5 hPa (0.15 inHg) per hour for at least 6 hours.[59] For rapid intensification to occur, several conditions must be in place. Water temperatures must be extremely high (near or above 30 °C (86 °F)), and water of this temperature must be sufficiently deep such that waves do not upwell cooler waters to the surface. On the other hand, Tropical Cyclone Heat Potential is one of such non-conventional subsurface oceanographic parameters influencing the cyclone intensity. Wind shear must be low; when wind shear is high, the convection and circulation in the cyclone will be disrupted. Usually, an anticyclone in the upper layers of the troposphere above the storm must be present as well—for extremely low surface pressures to develop, air must be rising very rapidly in the eyewall of the storm, and an upper-level anticyclone helps channel this air away from the cyclone efficiently.[60] However, some cyclones such as Hurricane Epsilon have rapidly intensified despite relatively unfavorable conditions.[61][62]

Dissipation

sheared tropical cyclone, with deep convection
slightly removed from the center of the system.

There are a number of ways a tropical cyclone can weaken, dissipate, or lose its tropical characteristics. These include making landfall, moving over cooler water, encountering dry air, or interacting with other weather systems; however, once a system has dissipated or lost its tropical characteristics, its remnants could regenerate a tropical cyclone if environmental conditions become favorable.[63][64]

A tropical cyclone can dissipate when it moves over waters significantly cooler than 26.5 °C (79.7 °F). This will deprive the storm of such tropical characteristics as a warm core with thunderstorms near the center, so that it becomes a remnant

extratropical cyclones. This transition can take 1–3 days.[66]

Should a tropical cyclone make landfall or pass over an island, its circulation could start to break down, especially if it encounters mountainous terrain.[67] When a system makes landfall on a large landmass, it is cut off from its supply of warm moist maritime air and starts to draw in dry continental air.[67] This, combined with the increased friction over land areas, leads to the weakening and dissipation of the tropical cyclone.[67] Over a mountainous terrain, a system can quickly weaken; however, over flat areas, it may endure for two to three days before circulation breaks down and dissipates.[67]

Over the years, there have been a number of techniques considered to try to artificially modify tropical cyclones.[68] These techniques have included using nuclear weapons, cooling the ocean with icebergs, blowing the storm away from land with giant fans, and seeding selected storms with dry ice or silver iodide.[68] These techniques, however, fail to appreciate the duration, intensity, power or size of tropical cyclones.[68]

Methods for assessing intensity

A variety of methods or techniques, including surface, satellite, and aerial, are used to assess the intensity of a tropical cyclone. Reconnaissance aircraft fly around and through tropical cyclones, outfitted with specialized instruments, to collect information that can be used to ascertain the winds and pressure of a system.[1] Tropical cyclones possess winds of different speeds at different heights. Winds recorded at flight level can be converted to find the wind speeds at the surface.[69] Surface observations, such as ship reports, land stations, mesonets, coastal stations, and buoys, can provide information on a tropical cyclone's intensity or the direction it is traveling.[1] Wind-pressure relationships (WPRs) are used as a way to determine the pressure of a storm based on its wind speed. Several different methods and equations have been proposed to calculate WPRs.[70][71] Tropical cyclones agencies each use their own, fixed WPR, which can result in inaccuracies between agencies that are issuing estimates on the same system.[71] The ASCAT is a scatterometer used by the MetOp satellites to map the wind field vectors of tropical cyclones.[1] The SMAP uses an L-band radiometer channel to determine the wind speeds of tropical cyclones at the ocean surface, and has been shown to be reliable at higher intensities and under heavy rainfall conditions, unlike scatterometer-based and other radiometer-based instruments.[72]

The Dvorak technique plays a large role in both the classification of a tropical cyclone and the determination of its intensity. Used in warning centers, the method was developed by Vernon Dvorak in the 1970s, and uses both visible and infrared satellite imagery in the assessment of tropical cyclone intensity. The Dvorak technique uses a scale of "T-numbers", scaling in increments of 0.5 from T1.0 to T8.0. Each T-number has an intensity assigned to it, with larger T-numbers indicating a stronger system. Tropical cyclones are assessed by forecasters according to an array of patterns, including curved banding features, shear, central dense overcast, and eye, in order to determine the T-number and thus assess the intensity of the storm.[73] The Cooperative Institute for Meteorological Satellite Studies works to develop and improve automated satellite methods, such as the Advanced Dvorak Technique (ADT) and SATCON. The ADT, used by a large number of forecasting centers, uses infrared geostationary satellite imagery and an algorithm based upon the Dvorak technique to assess the intensity of tropical cyclones. The ADT has a number of differences from the conventional Dvorak technique, including changes to intensity constraint rules and the usage of microwave imagery to base a system's intensity upon its internal structure, which prevents the intensity from leveling off before an eye emerges in infrared imagery.[74] The SATCON weights estimates from various satellite-based systems and microwave sounders, accounting for the strengths and flaws in each individual estimate, to produce a consensus estimate of a tropical cyclone's intensity which can be more reliable than the Dvorak technique at times.[75][76]

Intensity metrics

Multiple intensity metrics are used, including accumulated cyclone energy (ACE), the Hurricane Surge Index, the Hurricane Severity Index, the Power Dissipation Index (PDI), and integrated kinetic energy (IKE). ACE is a metric of the total energy a system has exerted over its lifespan. ACE is calculated by summing the squares of a cyclone's sustained wind speed, every six hours as long as the system is at or above tropical storm intensity and either tropical or subtropical.[77] The calculation of the PDI is similar in nature to ACE, with the major difference being that wind speeds are cubed rather than squared.[78] The Hurricane Surge Index is a metric of the potential damage a storm may inflict via storm surge. It is calculated by squaring the dividend of the storm's wind speed and a climatological value (33 m/s or 74 mph), and then multiplying that quantity by the dividend of the radius of hurricane-force winds and its climatological value (96.6 km or 60.0 mi). This can be represented in equation form as:

where v is the storm's wind speed and r is the radius of hurricane-force winds.[79] The Hurricane Severity Index is a scale that can assign up to 50 points to a system; up to 25 points come from intensity, while the other 25 come from the size of the storm's wind field.[80] The IKE model measures the destructive capability of a tropical cyclone via winds, waves, and surge. It is calculated as:

where p is the density of air, u is a sustained surface wind speed value, and dv is the volume element.[80][81]

Classification and naming

Classification

spiral banding and increased centralization, while the strongest (lower right) has developed an eye
.

Around the world, tropical cyclones are classified in different ways, based on the location (

International Dateline in the Northern Hemisphere, it becomes known as a typhoon. This happened in 2014 for Hurricane Genevieve, which became Typhoon Genevieve.[82] Within the Southern Hemisphere, it is either called a hurricane, tropical cyclone or a severe tropical cyclone, depending on if it is located within the South Atlantic, South-West Indian Ocean, Australian region or the South Pacific Ocean.[21][22] The descriptors for tropical cyclones with wind speeds below 65 kn (120 km/h; 75 mph) also vary by tropical cyclone basin and may be further subdivided into categories such as "tropical storm", "cyclonic storm", "tropical depression", or "deep depression".[19][20][18]

Naming

The practice of using

Clement Wragge who named systems between 1887 and 1907.[83][84] This system of naming weather systems subsequently fell into disuse for several years after Wragge retired, until it was revived in the latter part of World War II for the Western Pacific.[83][84] Formal naming schemes have subsequently been introduced for the North and South Atlantic, Eastern, Central, Western and Southern Pacific basins as well as the Australian region and Indian Ocean.[84]

At present, tropical cyclones are officially named by one of twelve

lists with one, three, or ten-minute sustained wind speeds of more than 65 km/h (40 mph) depending on which basin it originates.[18][20][21] However, standards vary from basin to basin with some tropical depressions named in the Western Pacific, while tropical cyclones have to have a significant amount of gale-force winds occurring around the center before they are named within the Southern Hemisphere.[21][22] The names of significant tropical cyclones in the North Atlantic Ocean, Pacific Ocean, and Australian region are retired from the naming lists and replaced with another name.[18][19][22] Tropical cyclones that develop around the world are assigned an identification code consisting of a two-digit number and suffix letter by the warning centers that monitor them.[22][85]

Related cyclone types

In addition to tropical cyclones, there are two other classes of

comma-shaped" cloud pattern.[88] Extratropical cyclones can also be dangerous when their low-pressure centers cause powerful winds and high seas.[89]

A subtropical cyclone is a weather system that has some characteristics of a tropical cyclone and some characteristics of an extratropical cyclone. They can form in a wide band of latitudes, from the equator to 50°. Although subtropical storms rarely have hurricane-force winds, they may become tropical in nature as their cores warm.[90]

Structure

Eye and center

The eye and surrounding clouds of 2018 Hurricane Florence as seen from the International Space Station

At the center of a mature tropical cyclone, air sinks rather than rises. For a sufficiently strong storm, air may sink over a layer deep enough to suppress cloud formation, thereby creating a clear "

convective clouds, although the sea may be extremely violent.[91] The eye is normally circular and is typically 30–65 km (19–40 mi) in diameter, though eyes as small as 3 km (1.9 mi) and as large as 370 km (230 mi) have been observed.[92][93]

The cloudy outer edge of the eye is called the "eyewall". The eyewall typically expands outward with height, resembling an arena football stadium; this phenomenon is sometimes referred to as the "stadium effect".[93] The eyewall is where the greatest wind speeds are found, air rises most rapidly, clouds reach their highest altitude, and precipitation is the heaviest. The heaviest wind damage occurs where a tropical cyclone's eyewall passes over land.[91]

In a weaker storm, the eye may be obscured by the central dense overcast, which is the upper-level cirrus shield that is associated with a concentrated area of strong thunderstorm activity near the center of a tropical cyclone.[94]

The eyewall may vary over time in the form of

Outer rainbands can organize into an outer ring of thunderstorms that slowly moves inward, which is believed to rob the primary eyewall of moisture and angular momentum. When the primary eyewall weakens, the tropical cyclone weakens temporarily. The outer eyewall eventually replaces the primary one at the end of the cycle, at which time the storm may return to its original intensity.[95]

Size

There are a variety of metrics commonly used to measure storm size. The most common metrics include the radius of maximum wind, the radius of 34-knot (17 m/s; 63 km/h; 39 mph) wind (i.e.

isobar (ROCI), and the radius of vanishing wind.[96][97] An additional metric is the radius at which the cyclone's relative vorticity field decreases to 1×10−5 s−1.[93]

Size descriptions of tropical cyclones
ROCI (Diameter) Type
Less than 2 degrees latitude Very small/minor
2 to 3 degrees of latitude Small
3 to 6 degrees of latitude Medium/Average/Normal
6 to 8 degrees of latitude Large
Over 8 degrees of latitude Very large[98]

On Earth, tropical cyclones span a large range of sizes, from 100–2,000 km (62–1,243 mi) as measured by the radius of vanishing wind. They are largest on average in the northwest Pacific Ocean basin and smallest in the northeastern Pacific Ocean basin.[99] If the radius of outermost closed isobar is less than two degrees of latitude (222 km (138 mi)), then the cyclone is "very small" or a "midget". A radius of 3–6 latitude degrees (333–670 km (207–416 mi)) is considered "average sized". "Very large" tropical cyclones have a radius of greater than 8 degrees (888 km (552 mi)).[98] Observations indicate that size is only weakly correlated to variables such as storm intensity (i.e. maximum wind speed), radius of maximum wind, latitude, and maximum potential intensity.[97][99] Typhoon Tip is the largest cyclone on record, with tropical storm-force winds 2,170 km (1,350 mi) in diameter. The smallest storm on record is Tropical Storm Marco of 2008, which generated tropical storm-force winds only 37 km (23 mi) in diameter.[100]

Movement

The movement of a tropical cyclone (i.e. its "track") is typically approximated as the sum of two terms: "steering" by the background environmental wind and "beta drift".[101] Some tropical cyclones can move across large distances, such as Hurricane John, the second longest-lasting tropical cyclone on record, which traveled 13,280 km (8,250 mi), the longest track of any Northern Hemisphere tropical cyclone, over its 31-day lifespan in 1994.[102][103][104]

Environmental steering

Environmental steering is the primary influence on the motion of tropical cyclones.[105] It represents the movement of the storm due to prevailing winds and other wider environmental conditions, similar to "leaves carried along by a stream".[106]

Physically, the winds, or flow field, in the vicinity of a tropical cyclone may be treated as having two parts: the flow associated with the storm itself, and the large-scale background flow of the environment.[105] Tropical cyclones can be treated as local maxima of vorticity suspended within the large-scale background flow of the environment.[107] In this way, tropical cyclone motion may be represented to first-order as advection of the storm by the local environmental flow.[108] This environmental flow is termed the "steering flow" and is the dominant influence on tropical cyclone motion.[105] The strength and direction of the steering flow can be approximated as a vertical integration of the winds blowing horizontally in the cyclone's vicinity, weighted by the altitude at which those winds are occurring. Because winds can vary with height, determining the steering flow precisely can be difficult.

The pressure altitude at which the background winds are most correlated with a tropical cyclone's motion is known as the "steering level".[107] The motion of stronger tropical cyclones is more correlated with the background flow averaged across a thicker portion of troposphere compared to weaker tropical cyclones whose motion is more correlated with the background flow averaged across a narrower extent of the lower troposphere.[109] When wind shear and latent heat release is present, tropical cyclones tend to move towards regions where potential vorticity is increasing most quickly.[110]

Climatologically, tropical cyclones are steered primarily westward by the east-to-west

subtropical ridge—a persistent high-pressure area over the world's subtropical oceans.[106] In the tropical North Atlantic and Northeast Pacific oceans, the trade winds steer tropical easterly waves westward from the African coast toward the Caribbean Sea, North America, and ultimately into the central Pacific Ocean before the waves dampen out.[111] These waves are the precursors to many tropical cyclones within this region.[112] In contrast, in the Indian Ocean and Western Pacific in both hemispheres, tropical cyclogenesis is influenced less by tropical easterly waves and more by the seasonal movement of the Intertropical Convergence Zone and the monsoon trough.[113] Other weather systems such as mid-latitude troughs and broad monsoon gyres can also influence tropical cyclone motion by modifying the steering flow.[109][114]

Beta drift

In addition to environmental steering, a tropical cyclone will tend to drift poleward and westward, a motion known as "beta drift".[115] This motion is due to the superposition of a vortex, such as a tropical cyclone, onto an environment in which the Coriolis force varies with latitude, such as on a sphere or beta plane.[116] The magnitude of the component of tropical cyclone motion associated with the beta drift ranges between 1–3 m/s (3.6–10.8 km/h; 2.2–6.7 mph) and tends to be larger for more intense tropical cyclones and at higher latitudes. It is induced indirectly by the storm itself as a result of feedback between the cyclonic flow of the storm and its environment.[117][115]

Physically, the cyclonic circulation of the storm advects environmental air poleward east of center and equatorial west of center. Because air must conserve its angular momentum, this flow configuration induces a cyclonic gyre equatorward and westward of the storm center and an anticyclonic gyre poleward and eastward of the storm center. The combined flow of these gyres acts to advect the storm slowly poleward and westward. This effect occurs even if there is zero environmental flow.[118][119] Due to a direct dependence of the beta drift on angular momentum, the size of a tropical cyclone can affect the influence of beta drift on its motion; beta drift imparts a greater influence on the movement of larger tropical cyclones than that of smaller ones.[120][121]

Multiple storm interaction

A third component of motion that occurs relatively infrequently involves the interaction of multiple tropical cyclones. When two cyclones approach one another, their centers will begin orbiting cyclonically about a point between the two systems. Depending on their separation distance and strength, the two vortices may simply orbit around one another, or else may spiral into the center point and merge. When the two vortices are of unequal size, the larger vortex will tend to dominate the interaction, and the smaller vortex will orbit around it. This phenomenon is called the Fujiwhara effect, after Sakuhei Fujiwhara.[122]

Interaction with the mid-latitude westerlies

Japanese coast in 2006

Though a tropical cyclone typically moves from east to west in the tropics, its track may shift poleward and eastward either as it moves west of the subtropical ridge axis or else if it interacts with the mid-latitude flow, such as the

Effects

Natural phenomena caused or worsened by tropical cyclones

Tropical cyclones out at sea cause large waves,

heavy rain, floods and high winds, disrupting international shipping and, at times, causing shipwrecks.[125] Tropical cyclones stir up water, leaving a cool wake behind them, which causes the region to be less favorable for subsequent tropical cyclones.[36] On land, strong winds can damage or destroy vehicles, buildings, bridges, and other outside objects, turning loose debris into deadly flying projectiles. The storm surge, or the increase in sea level due to the cyclone, is typically the worst effect from landfalling tropical cyclones, historically resulting in 90% of tropical cyclone deaths.[126] Cyclone Mahina produced the highest storm surge on record, 13 m (43 ft), at Bathurst Bay, Queensland, Australia, in March 1899.[127] Other ocean-based hazards that tropical cyclones produce are rip currents and undertow. These hazards can occur hundreds of kilometers (hundreds of miles) away from the center of a cyclone, even if other weather conditions are favorable.[128][129]
The broad rotation of a landfalling tropical cyclone, and vertical wind shear at its periphery, spawns
eyewall mesovortices, which persist until landfall.[130] Hurricane Ivan produced 120 tornadoes, more than any other tropical cyclone.[131] Lightning activity is produced within tropical cyclones; this activity is more intense within stronger storms and closer to and within the storm's eyewall.[132][133] Tropical cyclones can increase the amount of snowfall a region experiences by delivering additional moisture.[134] Wildfires can be worsened when a nearby storm fans their flames with its strong winds.[135][136]

Effect on property and human life

total collapse of houses, cars and facilities
Aftermath of Hurricane Ike in Bolivar Peninsula, Texas
The number of $1 billion Atlantic hurricanes almost doubled from the 1980s to the 2010s, and inflation-adjusted costs have increased more than elevenfold.[137] The increases have been attributed to climate change and to greater numbers of people moving to coastal areas.[137]

Tropical cyclones regularly affect the coastlines of most of

Category 3 intensity).[146]

In

Sahara Desert,[147] or otherwise strike the Horn of Africa and Southern Africa.[148][149] Cyclone Idai in March 2019 hit central Mozambique, becoming the deadliest tropical cyclone on record in Africa, with 1,302 fatalities, and damage estimated at US$2.2 billion.[150][151] Réunion island, located east of Southern Africa, experiences some of the wettest tropical cyclones on record. In January 1980, Cyclone Hyacinthe produced 6,083 mm (239.5 in) of rain over 15 days, which was the largest rain total recorded from a tropical cyclone on record.[152][153][154] In Asia, tropical cyclones from the Indian and Pacific oceans regularly affect some of the most populated countries on Earth. In 1970, a cyclone struck Bangladesh, then known as East Pakistan, producing a 6.1 m (20 ft) storm surge that killed at least 300,000 people; this made it the deadliest tropical cyclone on record.[155] In October 2019, Typhoon Hagibis struck the Japanese island of Honshu and inflicted US$15 billion in damage, making it the costliest storm on record in Japan.[156] The islands that comprise Oceania, from Australia to French Polynesia, are routinely affected by tropical cyclones.[157][158][159] In Indonesia, a cyclone struck the island of Flores in April 1973, killing 1,653 people, making it the deadliest tropical cyclone recorded in the Southern Hemisphere.[160][161]

South Atlantic Ocean is generally inhospitable to the formation of a tropical storm.[169] However, in March 2004, Hurricane Catarina struck southeastern Brazil as the first hurricane on record in the South Atlantic Ocean.[170]

Environmental effects

Although cyclones take an enormous toll in lives and personal property, they may be important factors in the

saltmarshes and Mangrove forests, can be severely damaged or destroyed by tropical cyclones, which erode land and destroy vegetation.[180][181] Tropical cyclones can cause harmful algae blooms to form in bodies of water by increasing the amount of nutrients available.[182][183][184] Insect populations can decrease in both quantity and diversity after the passage of storms.[185] Strong winds associated with tropical cyclones and their remnants are capable of felling thousands of trees, causing damage to forests.[186]

When hurricanes surge upon shore from the ocean, salt is introduced to many freshwater areas and raises the

acids onshore when they make landfall. The floodwater can pick up the toxins from different spills and contaminate the land that it passes over. These toxins are harmful to the people and animals in the area, as well as the environment around them.[188] Tropical cyclones can cause oil spills by damaging or destroying pipelines and storage facilities.[189][182][190] Similarly, chemical spills have been reported when chemical and processing facilities were damaged.[190][191][192] Waterways have become contaminated with toxic levels of metals such as nickel, chromium, and mercury during tropical cyclones.[193][194]

Tropical cyclones can have an extensive effect on geography, such as creating or destroying land.

Funafuti Atoll, by nearly 20%.[195][197][198] Hurricane Walaka destroyed the small East Island in 2018,[196][199] which destroyed the habitat for the endangered Hawaiian monk seal, as well as, threatened sea turtles and seabirds.[200] Landslides frequently occur during tropical cyclones and can vastly alter landscapes; some storms are capable of causing hundreds to tens of thousands of landslides.[201][202][203][204] Storms can erode coastlines over an extensive area and transport the sediment to other locations.[194][205][206]

Climatology

Tropical cyclones have occurred around the world for millennia. Reanalyses and research are being undertaken to extend the historical record, through the usage of proxy data such as overwash deposits, beach ridges and historical documents such as diaries.[207] Major tropical cyclones leave traces in overwash records and shell layers in some coastal areas, which have been used to gain insight into hurricane activity over the past thousands of years.[208] Sediment records in Western Australia suggest an intense tropical cyclone in the 4th millennium BC.[207] Proxy records based on paleotempestological research have revealed that major hurricane activity along the Gulf of Mexico coast varies on timescales of centuries to millennia.[209][210] In the year 957, a powerful typhoon struck southern China, killing around 10,000 people due to flooding.[211] The Spanish colonization of Mexico described "tempestades" in 1730,[212] although the official record for Pacific hurricanes only dates to 1949.[213] In the south-west Indian Ocean, the tropical cyclone record goes back to 1848.[214] In 2003, the Atlantic hurricane reanalysis project examined and analyzed the historical record of tropical cyclones in the Atlantic back to 1851, extending the existing database from 1886.[215]

Before satellite imagery became available during the 20th century, many of these systems went undetected unless it impacted land or a ship encountered it by chance.

National Aeronautics and Space Administration in 1960 but were not declared operational until 1965.[1] However, it took several years for some of the warning centres to take advantage of this new viewing platform and develop the expertise to associate satellite signatures with storm position and intensity.[1]

Each year on average, around 80 to 90 named tropical cyclones form around the world, of which over half develop hurricane-force winds of 65 kn (120 km/h; 75 mph) or more.[1] Worldwide, tropical cyclone activity peaks in late summer, when the difference between temperatures aloft and sea surface temperatures is the greatest. However, each particular basin has its own seasonal patterns. On a worldwide scale, May is the least active month, while September is the most active month. November is the only month in which all the tropical cyclone basins are in season.[218] In the Northern Atlantic Ocean, a distinct cyclone season occurs from June 1 to November 30, sharply peaking from late August through September.[218] The statistical peak of the Atlantic hurricane season is September 10. The Northeast Pacific Ocean has a broader period of activity, but in a similar time frame to the Atlantic.[219] The Northwest Pacific sees tropical cyclones year-round, with a minimum in February and March and a peak in early September.[218] In the North Indian basin, storms are most common from April to December, with peaks in May and November.[218] In the Southern Hemisphere, the tropical cyclone year begins on July 1 and runs all year-round encompassing the tropical cyclone seasons, which run from November 1 until the end of April, with peaks in mid-February to early March.[218][22]

Of various

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 to China and much greater intensity in the Philippines.[222] The Atlantic Ocean experiences depressed activity due to increased vertical wind shear across the region during El Niño years.[223] Tropical cyclones are further influenced by the Atlantic Meridional Mode, the Quasi-biennial oscillation and the Madden–Julian oscillation.[220][224]

Season lengths and averages
Basin Season
start
Season
end
Tropical
cyclones
Refs
North Atlantic June 1 November 30 14.4 [225]
Eastern Pacific May 15 November 30 16.6 [225]
Western Pacific January 1 December 31 26.0 [225]
North Indian January 1 December 31 12 [226]
South-West Indian July 1 June 30 9.3 [225][21]
Australian region November 1 April 30 11.0 [227]
Southern Pacific November 1 April 30 7.1 [228]
Total: 96.4

Influence of climate change

The 20-year average of the number of annual Category 4 and 5 hurricanes in the Atlantic region has approximately doubled since the year 2000.[229]
Perceptions in the United States differ along political lines, on whether climate change was a "major factor" contributing to various extreme weather events experienced by respondents.[230] "Severe storms" includes hurricanes.

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.[2] Tropical cyclones use warm, moist air as their fuel. As climate change is warming ocean temperatures, there is potentially more of this fuel available.[231] Between 1979 and 2017, there was a global increase in the proportion of tropical cyclones of Category 3 and higher on the Saffir–Simpson scale. The trend was most clear in the North Atlantic and in the Southern Indian Ocean. In the North Pacific, tropical cyclones have been moving poleward into colder waters and there was no increase in intensity over this period.[232] With 2 °C (3.6 °F) warming, a greater percentage (+13%) of tropical cyclones are expected to reach Category 4 and 5 strength.[2] A 2019 study indicates that climate change has been driving the observed trend of rapid intensification of tropical cyclones in the Atlantic basin. Rapidly intensifying cyclones are hard to forecast and therefore pose additional risk to coastal communities.[233]

Warmer air can hold more water vapor: the theoretical maximum water vapor content is given by the

due to global warming.[237]

There is currently no consensus on how climate change will affect the overall frequency of tropical cyclones.[2] A majority of climate models show a decreased frequency in future projections.[238] For instance, a 2020 paper comparing nine high-resolution climate models found robust decreases in frequency in the Southern Indian Ocean and the Southern Hemisphere more generally, while finding mixed signals for Northern Hemisphere tropical cyclones.[239] Observations have shown little change in the overall frequency of tropical cyclones worldwide,[240] with increased frequency in the North Atlantic and central Pacific, and significant decreases in the southern Indian Ocean and western North Pacific.[241] There has been a poleward expansion of the latitude at which the maximum intensity of tropical cyclones occurs, which may be associated with climate change.[242] In the North Pacific, there may also have been an eastward expansion.[236] Between 1949 and 2016, there was a slowdown in tropical cyclone translation speeds. It is unclear still to what extent this can be attributed to climate change: climate models do not all show this feature.[238]

A study review article published in 2021 concluded that the geographic range of tropical cyclones will probably expand poleward in response to climate warming of the

Hadley circulation.[243]

Observation and forecasting

Observation

Aerial view of storm clouds
Sunset view of Hurricane Isidore's rainbands photographed at 2,100 m (7,000 ft)
WP-3D Orion is used to go into the eye of a hurricane
for data collection and measurements purposes.

Historically, tropical cyclones have occurred around the world for thousands of years, with one of the earliest tropical cyclones on record estimated to have occurred in Western Australia in around 4000 BC.[207] However, before satellite imagery became available during the 20th century, there was no way to detect a tropical cyclone unless it impacted land or a ship encountered it by chance.[1]

Intense tropical cyclones pose a particular observation challenge, as they are a dangerous oceanic phenomenon, and weather stations, being relatively sparse, are rarely available on the site of the storm itself. In general, surface observations are available only if the storm is passing over an island or a coastal area, or if there is a nearby ship. Real-time measurements are usually taken in the periphery of the cyclone, where conditions are less catastrophic and its true strength cannot be evaluated. For this reason, there are teams of meteorologists that move into the path of tropical cyclones to help evaluate their strength at the point of landfall.[244]

Tropical cyclones are tracked by

visible and infrared images from space, usually at half-hour to quarter-hour intervals. As a storm approaches land, it can be observed by land-based Doppler weather radar. Radar plays a crucial role around landfall by showing a storm's location and intensity every several minutes.[245] Other satellites provide information from the perturbations of GPS signals, providing thousands of snapshots per day and capturing atmospheric temperature, pressure, and moisture content.[246]

2005 hurricane season. A similar mission was also completed successfully in the western Pacific Ocean.[248]

Track errors plotted over time
A general decrease in error trends in tropical cyclone path prediction is evident since the 1970s

Forecasting

High-speed computers and sophisticated simulation software allow forecasters to produce

satellites and other sensors, scientists have increased the accuracy of track forecasts over recent decades.[249] However, scientists are not as skillful at predicting the intensity of tropical cyclones.[250] The lack of improvement in intensity forecasting is attributed to the complexity of tropical systems and an incomplete understanding of factors that affect their development. New tropical cyclone position and forecast information is available at least every six hours from the various warning centers.[251][252][253][254][255]

Geopotential height

In meteorology, geopotential heights are used when creating forecasts and analyzing pressure systems. Geopotential heights represent the estimate of the real height of a pressure system above the average sea level.[256] Geopotential heights for weather are divided up into several levels. The lowest geopotential height level is 850 hPa (25.10 inHg), which represents the lowest 1,500 m (5,000 ft) of the atmosphere. The moisture content, gained by using either the relative humidity or the precipitable water value, is used in creating forecasts for precipitation.[257] The next level, 700 hPa (20.67 inHg), is at a height of 2,300–3,200 m (7,700–10,500 ft); 700 hPa is regarded as the highest point in the lower atmosphere. At this layer, both vertical movement and moisture levels are used to locate and create forecasts for precipitation.[258] The middle level of the atmosphere is at 500 hPa (14.76 inHg) or a height of 4,900–6,100 m (16,000–20,000 ft). The 500 hPa level is used for measuring atmospheric vorticity, commonly known as the spin of air. The relative humidity is also analyzed at this height in order to establish where precipitation is likely to materialize.[259] The next level occurs at 300 hPa (8.859 inHg) or a height of 8,200–9,800 m (27,000–32,000 ft).[260] The top-most level is located at 200 hPa (5.906 inHg), which corresponds to a height of 11,000–12,000 m (35,000–41,000 ft). Both the 200 and 300 hPa levels are mainly used to locate the jet stream.[261]

Society and culture

Preparations

Evacuation route sign on Tulane Avenue in New Orleans shows lines from long standing floodwaters after Hurricane Katrina

Ahead of the formal season starting, people are urged to

watches and warnings to cover the expected effects.[265] However, there are some exceptions with the United States National Hurricane Center and Fiji Meteorological Service responsible for issuing or recommending warnings for other nations in their area of responsibility.[266][267][268]
: 2–4 

An important decision in individual preparedness is determining if and when to evacuate an area that will be affected by a tropical cyclone.[269] Tropical cyclone tracking charts allow people to track ongoing systems to form their own opinions regarding where the storms are going and whether or not they need to prepare for the system being tracked, including possible evacuation. This continues to be encouraged by the National Oceanic and Atmospheric Administration and National Hurricane Center.[270]

Response

View of tropical cyclone damage from a helicopter
Relief efforts for Hurricane Dorian in the Bahamas

Hurricane response is the disaster response after a hurricane. Activities performed by hurricane responders include assessment, restoration, and demolition of buildings; removal of debris and waste; repairs to land-based and maritime infrastructure; and public health services including search and rescue operations.[271] Hurricane response requires coordination between federal, tribal, state, local, and private entities.[272] According to the National Voluntary Organizations Active in Disaster, potential response volunteers should affiliate with established organizations and should not self-deploy, so that proper training and support can be provided to mitigate the danger and stress of response work.[273]

Hurricane responders face many hazards. Hurricane responders may be exposed to chemical and biological contaminants including stored chemicals,

heat stress is a concern as workers are often exposed to hot and humid temperatures, wear protective clothing and equipment, and have physically difficult tasks.[274][277]

See also

References

  1. ^ a b c d e f g h i j k l m n o p q Global Guide to Tropical Cyclone Forecasting: 2017 (PDF) (Report). World Meteorological Organization. April 17, 2018. Archived (PDF) from the original on July 14, 2019. Retrieved September 6, 2020.
  2. ^
    ISSN 0003-0007
    .
  3. ^ a b "Glossary of NHC Terms". United States National Hurricane Center. Archived from the original on February 16, 2021. Retrieved February 18, 2021.
  4. ^ "Tropical cyclone facts: What is a tropical cyclone?". United Kingdom Met Office. Archived from the original on February 2, 2021. Retrieved February 25, 2021.
  5. ^ a b c d e "Tropical cyclone facts: How do tropical cyclones form?". United Kingdom Met Office. Archived from the original on February 2, 2021. Retrieved March 1, 2021.
  6. ^
    Landsea, Chris. "How do tropical cyclones form?". Frequently Asked Questions. Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. Archived
    from the original on August 27, 2009. Retrieved October 9, 2017.
  7. ^ Berg, Robbie. "Tropical cyclone intensity in relation to SST and moisture variability" (PDF). Rosenstiel School of Marine, Atmospheric, and Earth Science (University of Miami). Archived (PDF) from the original on June 10, 2011. Retrieved September 23, 2010.
  8. S2CID 53341455
    . Retrieved October 4, 2022.
  9. from the original on October 2, 2012. Retrieved October 19, 2006.
  10. ^ Landsea, Christopher. "AOML Climate Variability of Tropical Cyclones paper". Atlantic Oceanographic and Meteorological Laboratory. Archived from the original on October 26, 2021. Retrieved September 23, 2010.
  11. S2CID 17409876
    .
  12. from the original on October 2, 2022. Retrieved October 5, 2022.
  13. .
  14. .
  15. .
  16. .
  17. .
  18. ^ a b c d e f RA IV Hurricane Committee. Regional Association IV Hurricane Operational Plan 2019 (PDF) (Report). World Meteorological Organization. Archived (PDF) from the original on July 2, 2019. Retrieved July 2, 2019.
  19. ^ a b c d WMO/ESCP Typhoon Committee (March 13, 2015). Typhoon Committee Operational Manual Meteorological Component 2015 (PDF) (Report No. TCP-23). World Meteorological Organization. pp. 40–41. Archived (PDF) from the original on October 1, 2015. Retrieved March 28, 2015.
  20. ^ a b c d WMO/ESCAP Panel on Tropical Cyclones (November 2, 2018). Tropical Cyclone Operational Plan for the Bay of Bengal and the Arabian Sea 2018 (PDF) (Report No. TCP-21). World Meteorological Organization. pp. 11–12. Archived (PDF) from the original on July 2, 2019. Retrieved July 2, 2019.
  21. ^ a b c d e RA I Tropical Cyclone Committee (November 9, 2012). Tropical Cyclone Operational Plan for the South-West Indian Ocean: 2012 (PDF) (Report No. TCP-12). World Meteorological Organization. pp. 11–14. Archived (PDF) from the original on March 29, 2015. Retrieved March 29, 2015.
  22. ^ a b c d e f g h i j k RA V Tropical Cyclone Committee (2023). Tropical Cyclone Operational Plan for the South-East Indian Ocean and the Southern Pacific Ocean 2023 (PDF) (Report). World Meteorological Organization. Retrieved October 23, 2023.
  23. ^ "Normas Da Autoridade Marítima Para As Atividades De Meteorologia Marítima" (PDF) (in Portuguese). Brazilian Navy. 2011. Archived from the original (PDF) on February 6, 2015. Retrieved October 5, 2018.
  24. S2CID 19031120
    .
  25. ^ "What is a Tropical Cyclone?". Bureau of Meteorology. Archived from the original on October 3, 2022. Retrieved October 7, 2022.
  26. ^ "Saffir-Simpson Hurricane Wind Scale". National Hurricane Center. Archived from the original on June 20, 2020. Retrieved October 7, 2022.
  27. .
  28. ^ Pasch, Richard (October 23, 2015). "Hurricane Patricia Discussion Number 14". National Hurricane Center. Archived from the original on October 25, 2015. Retrieved October 23, 2015. Data from three center fixes by the Hurricane Hunters indicate that the intensity, based on a blend of 700 mb-flight level and SFMR-observed surface winds, is near 175 kt. This makes Patricia the strongest hurricane on record in the National Hurricane Center's area of responsibility (AOR) which includes the Atlantic and the eastern North Pacific basins.
  29. from the original on April 28, 2021. Retrieved April 28, 2021.
  30. from the original on April 28, 2021. Retrieved April 28, 2021.
  31. from the original on April 28, 2021. Retrieved April 28, 2021.
  32. ^ Brown, Daniel (April 20, 2017). "Tropical Cyclone Intensity Forecasting: Still a Challenging Proposition" (PDF). National Hurricane Center. p. 7. Archived (PDF) from the original on April 27, 2021. Retrieved April 27, 2021.
  33. ^ from the original on April 27, 2021. Retrieved April 27, 2021.
  34. (PDF) from the original on April 27, 2021. Retrieved April 27, 2021.
  35. .
  36. ^ a b D'Asaro, Eric A. & Black, Peter G. (2006). "J8.4 Turbulence in the Ocean Boundary Layer Below Hurricane Dennis". University of Washington. Archived (PDF) from the original on March 30, 2012. Retrieved February 22, 2008.
  37. S2CID 4330367
    .
  38. .
  39. ^ Stovern, Diana; Ritchie, Elizabeth. "Modeling the Effect of Vertical Wind Shear on Tropical Cyclone Size and Structure" (PDF). American Meteorological Society. pp. 1–2. Archived (PDF) from the original on June 18, 2021. Retrieved April 28, 2021.
  40. S2CID 73622535
    .
  41. .
  42. . Retrieved May 14, 2022.
  43. .
  44. from the original on May 14, 2022. Retrieved May 14, 2022.
  45. .
  46. from the original on May 14, 2022. Retrieved May 15, 2022.
  47. .
  48. .
  49. .
  50. .
  51. .
  52. .
  53. . Retrieved October 7, 2022.
  54. from the original on May 15, 2022. Retrieved May 20, 2022.
  55. .
  56. from the original on October 7, 2022. Retrieved October 7, 2022.
  57. .
  58. ^ "Glossary of NHC Terms". United States National Oceanic and Atmospheric Administration's National Hurricane Center. Archived from the original on September 12, 2019. Retrieved June 2, 2019.
  59. .
  60. ^ Diana Engle. "Hurricane Structure and Energetics". Data Discovery Hurricane Science Center. Archived from the original on May 27, 2008. Retrieved October 26, 2008.
  61. Miami, Florida: National Hurricane Center. Archived
    from the original on March 21, 2021. Retrieved February 4, 2021.
  62. ^ Cappucci, Matthew (October 21, 2020). "Epsilon shatters records as it rapidly intensifies into major hurricane near Bermuda". The Washington Post. Archived from the original on December 10, 2020. Retrieved February 4, 2021.
  63. ^ Lam, Linda (September 4, 2019). "Why the Eastern Caribbean Sea Can Be a 'Hurricane Graveyard'". The Weather Channel. TWC Product and Technology. Archived from the original on July 4, 2021. Retrieved April 6, 2021.
  64. ^ Sadler, James C.; Kilonsky, Bernard J. (May 1977). The Regeneration of South China Sea Tropical Cyclones in the Bay of Bengal (PDF) (Report). Monterey, California: Naval Environmental Prediction Research Facility. Archived (PDF) from the original on June 22, 2021. Retrieved April 6, 2021 – via Defense Technical Information Center.
  65. from the original on August 14, 2021. Retrieved November 22, 2020.
  66. ^ United States Naval Research Laboratory (September 23, 1999). "Tropical Cyclone Intensity Terminology". Tropical Cyclone Forecasters' Reference Guide. Archived from the original on July 12, 2012. Retrieved November 30, 2006.
  67. ^ a b c d "Anatomy and Life Cycle of a Storm: What Is the Life Cycle of a Hurricane and How Do They Move?". United States Hurricane Research Division. 2020. Archived from the original on February 17, 2021. Retrieved February 17, 2021.
  68. ^ a b c "Attempts to Stop a Hurricane in its Track: What Else has been Considered to Stop a Hurricane?". United States Hurricane Research Division. 2020. Archived from the original on February 17, 2021. Retrieved February 17, 2021.
  69. from the original on April 24, 2021. Retrieved April 23, 2021.
  70. .
  71. ^ .
  72. ^ Meissner, Thomas; Ricciardulli, L.; Wentz, F.; Sampson, C. (April 18, 2018). "Intensity and Size of Strong Tropical Cyclones in 2017 from NASA's SMAP L-Band Radiometer". American Meteorological Society. Archived from the original on April 21, 2021. Retrieved April 21, 2021.
  73. ^ DeMaria, Mark; Knaff, John; Zehr, Raymond (2013). Satellite-based Applications on Climate Change (PDF). Springer. pp. 152–154. Archived (PDF) from the original on April 22, 2021. Retrieved April 21, 2021.
  74. from the original on April 21, 2021. Retrieved April 21, 2021.
  75. from the original on April 21, 2021. Retrieved April 21, 2021.
  76. from the original on April 21, 2021. Retrieved April 21, 2021.
  77. .
  78. .
  79. .
  80. ^ .
  81. .
  82. ^ "Learn the difference between hurricanes, cyclones and typhoons", ABC, Inc., KGO-TV San Francisco, Channel 7 News. Retrieved May 25, 2023.
  83. ^
    S2CID 201717866. Archived from the original
    (PDF) on November 29, 2014. Retrieved August 25, 2014.
  84. ^ a b c d e Dorst, Neal M (October 23, 2012). "They Called the Wind Mahina: The History of Naming Cyclones". Hurricane Research Division, Atlantic Oceanographic and Meteorological Laboratory. National Oceanic and Atmospheric Administration. p. Slides 8–72.
  85. ^ Office of the Federal Coordinator for Meteorological Services and Supporting Research (May 2017). National Hurricane Operations Plan (PDF) (Report). National Oceanic and Atmospheric Administration. pp. 26–28. Archived (PDF) from the original on October 15, 2018. Retrieved October 14, 2018.
  86. ^ Lander, Mark A.; et al. (August 3, 2003). "Fifth International Workshop on Tropical Cyclones". World Meteorological Organization. Archived from the original on May 9, 2009. Retrieved May 6, 2009.
  87. ^ Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. "Frequently Asked Questions: What is an extra-tropical cyclone?". National Oceanic and Atmospheric Administration. Archived from the original on February 9, 2007. Retrieved July 25, 2006.
  88. ^ "Lesson 14: Background: Synoptic Scale". University of Wisconsin–Madison. February 25, 2008. Archived from the original on February 20, 2009. Retrieved May 6, 2009.
  89. ^ "An Overview of Coastal Land Loss: With Emphasis on the Southeastern United States". United States Geological Survey. 2008. Archived from the original on February 12, 2009. Retrieved May 6, 2009.
  90. ^ Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. "Frequently Asked Questions: What is a sub-tropical cyclone?". National Oceanic and Atmospheric Administration. Archived from the original on October 11, 2011. Retrieved July 25, 2006.
  91. ^
    National Oceanic & Atmospheric Administration. Archived
    from the original on December 7, 2013. Retrieved May 7, 2009.
  92. ^ Pasch, Richard J.; Eric S. Blake; Hugh D. Cobb III; David P. Roberts (September 28, 2006). "Tropical Cyclone Report: Hurricane Wilma: 15–25 October 2005" (PDF). National Hurricane Center. Archived (PDF) from the original on March 4, 2016. Retrieved December 14, 2006.
  93. ^ from the original on August 1, 2020. Retrieved December 12, 2019.
  94. ^ American Meteorological Society. "AMS Glossary: C". Glossary of Meteorology. Allen Press. Archived from the original on January 26, 2011. Retrieved December 14, 2006.
  95. ^ Atlantic Oceanographic and Hurricane Research Division. "Frequently Asked Questions: What are "concentric eyewall cycles" (or "eyewall replacement cycles") and why do they cause a hurricane's maximum winds to weaken?". National Oceanic and Atmospheric Administration. Archived from the original on December 6, 2006. Retrieved December 14, 2006.
  96. Bureau of Meteorology. May 7, 2009. Archived from the original
    on June 1, 2011. Retrieved May 6, 2009.
  97. ^ .
  98. ^ a b "Q: What is the average size of a tropical cyclone?". Joint Typhoon Warning Center. 2009. Archived from the original on October 4, 2013. Retrieved May 7, 2009.
  99. ^
    S2CID 123276607
    .
  100. ^ Dorst, Neal; Hurricane Research Division (May 29, 2009). "Frequently Asked Questions: Subject: E5) Which are the largest and smallest tropical cyclones on record?". National Oceanic and Atmospheric Administration's Atlantic Oceanographic and Meteorological Laboratory. Archived from the original on December 22, 2008. Retrieved June 12, 2013.
  101. S2CID 124178238
    .
  102. ^ Dorst, Neal; Hurricane Research Division (January 26, 2010). "Subject: E6) Frequently Asked Questions: Which tropical cyclone lasted the longest?". National Oceanic and Atmospheric Administration's Atlantic Oceanographic and Meteorological Laboratory. Archived from the original on May 6, 2009. Retrieved June 12, 2013.
  103. ^ Dorst, Neal; Delgado, Sandy; Hurricane Research Division (May 20, 2011). "Frequently Asked Questions: Subject: E7) What is the farthest a tropical cyclone has travelled?". National Oceanic and Atmospheric Administration's Atlantic Oceanographic and Meteorological Laboratory. Archived from the original on May 6, 2009. Retrieved June 12, 2013.
  104. ^ "Deadly cyclone Freddy has become Earth's longest-lived tropical storm". the Washington Post. March 7, 2023. Retrieved September 27, 2023.
  105. ^
    S2CID 58921153
    .
  106. ^ a b Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. "Frequently Asked Questions: What determines the movement of tropical cyclones?". National Oceanic and Atmospheric Administration. Archived from the original on July 16, 2012. Retrieved July 25, 2006.
  107. ^ .
  108. .
  109. ^ .
  110. .
  111. ^ Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. "Frequently Asked Questions: What is an easterly wave?". National Oceanic and Atmospheric Administration. Archived from the original on July 18, 2006. Retrieved July 25, 2006.
  112. .
  113. ^ DeCaria, Alex (2005). "Lesson 5 – Tropical Cyclones: Climatology". ESCI 344 – Tropical Meteorology. Millersville University. Archived from the original on May 7, 2008. Retrieved February 22, 2008.
  114. .
  115. ^ a b Wang, Bin; Elsberry, Russell L.; Yuqing, Wang; Liguang, Wu (1998). "Dynamics in Tropical Cyclone Motion: A Review" (PDF). Chinese Journal of the Atmospheric Sciences. 22 (4). Allerton Press: 416–434. Archived (PDF) from the original on June 17, 2021. Retrieved April 6, 2021 – via University of Hawaii.
  116. .
  117. .
  118. .
  119. .
  120. .
  121. .
  122. ^ "Fujiwhara effect describes a stormy waltz". USA Today. November 9, 2007. Archived from the original on November 5, 2012. Retrieved February 21, 2008.
  123. ^ "Section 2: Tropical Cyclone Motion Terminology". United States Naval Research Laboratory. April 10, 2007. Archived from the original on February 12, 2012. Retrieved May 7, 2009.
  124. ^ Powell, Jeff; et al. (May 2007). "Hurricane Ioke: 20–27 August 2006". 2006 Tropical Cyclones Central North Pacific. Central Pacific Hurricane Center. Archived from the original on March 6, 2016. Retrieved June 9, 2007.
  125. ^ Roth, David & Cobb, Hugh (2001). "Eighteenth Century Virginia Hurricanes". NOAA. Archived from the original on May 1, 2013. Retrieved February 24, 2007.
  126. ^
    PMID 15958424
    .
  127. .
  128. from the original on May 26, 2022. Retrieved May 25, 2022.
  129. from the original on May 26, 2022. Retrieved May 25, 2022.
  130. ^ Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. "Frequently Asked Questions: Are TC tornadoes weaker than midlatitude tornadoes?". National Oceanic and Atmospheric Administration. Archived from the original on September 14, 2009. Retrieved July 25, 2006.
  131. ^ Grazulis, Thomas P.; Grazulis, Doris (February 27, 2018). "Top 25 Tornado-Generating Hurricanes". The Tornado Project. St. Johnsbury, Vermont: Environmental Films. Archived from the original on December 12, 2013. Retrieved November 8, 2021.
  132. S2CID 56304603
    .
  133. .
  134. . Retrieved May 25, 2022.
  135. from the original on January 24, 2022. Retrieved May 25, 2022.
  136. . Retrieved May 25, 2022.
  137. ^ a b Philbrick, Ian Pasad; Wu, Ashley (December 2, 2022). "Population Growth Is Making Hurricanes More Expensive". The New York Times. Archived from the original on December 6, 2022. Newspaper states data source: NOAA.
  138. ^ Haque, Ubydul; Hashizume, Masahiro; Kolivras, Korine N; Overgaard, Hans J; Das, Bivash; Yamamoto, Taro (March 16, 2011). "Reduced death rates from cyclones in Bangladesh: what more needs to be done?". Bulletin of the World Health Organization. Archived from the original on October 5, 2020. Retrieved October 12, 2020.
  139. ^ "Hurricane Katrina Situation Report #11" (PDF). Office of Electricity Delivery and Energy Reliability (OE) United States Department of Energy. August 30, 2005. Archived from the original (PDF) on November 8, 2006. Retrieved February 24, 2007.
  140. S2CID 224926212
    . Retrieved May 25, 2022.
  141. (PDF) from the original on May 26, 2022. Retrieved May 25, 2022.
  142. from the original on May 26, 2022. Retrieved May 25, 2022.
  143. from the original on May 26, 2022. Retrieved May 25, 2022.
  144. from the original on May 26, 2022. Retrieved May 25, 2022.
  145. .
  146. from the original on October 2, 2012. Retrieved October 19, 2006.
  147. ^ Belles, Jonathan (August 28, 2018). "Why Tropical Waves Are Important During Hurricane Season". Weather.com. Archived from the original on October 1, 2020. Retrieved October 2, 2020.
  148. ^ Schwartz, Matthew (November 22, 2020). "Somalia's Strongest Tropical Cyclone Ever Recorded Could Drop 2 Years' Rain In 2 Days". NPR. Archived from the original on November 23, 2020. Retrieved November 23, 2020.
  149. S2CID 54879038
    .
  150. ^ Masters, Jeff. "Africa's Hurricane Katrina: Tropical Cyclone Idai Causes an Extreme Catastrophe". Weather Underground. Archived from the original on March 22, 2019. Retrieved March 23, 2019.
  151. ^ "Global Catastrophe Recap: First Half of 2019" (PDF). Aon Benfield. Archived (PDF) from the original on August 12, 2019. Retrieved August 12, 2019.
  152. ^ Lyons, Steve (February 17, 2010). "La Reunion Island's Rainfall Dynasty!". The Weather Channel. Archived from the original on February 10, 2014. Retrieved February 4, 2014.
  153. ^ Précipitations extrêmes (Report). Meteo France. Archived from the original on February 21, 2014. Retrieved April 15, 2013.
  154. .
  155. .
  156. ^ Weather, Climate & Catastrophe Insight: 2019 Annual Report (PDF) (Report). AON Benfield. January 22, 2020. Archived (PDF) from the original on January 22, 2020. Retrieved January 23, 2020.
  157. ^ Sharp, Alan; Arthur, Craig; Bob Cechet; Mark Edwards (2007). Natural hazards in Australia: Identifying risk analysis requirements (PDF) (Report). Geoscience Australia. p. 45. Archived (PDF) from the original on October 31, 2020. Retrieved October 11, 2020.
  158. ^ The Climate of Fiji (PDF) (Information Sheet: 35). Fiji Meteorological Service. April 28, 2006. Archived (PDF) from the original on March 20, 2021. Retrieved April 29, 2021.
  159. ^ Republic of Fiji: Third National Communication Report to the United Nations Framework Convention on Climate Change (PDF) (Report). United Nations Framework Convention on Climate Change. April 27, 2020. p. 62. Archived (PDF) from the original on July 6, 2021. Retrieved August 23, 2021.
  160. ^ "Death toll". The Canberra Times. Australian Associated Press. June 18, 1973. Archived from the original on August 27, 2020. Retrieved April 22, 2020.
  161. ^ Masters, Jeff. "Africa's Hurricane Katrina: Tropical Cyclone Idai Causes an Extreme Catastrophe". Weather Underground. Archived from the original on August 4, 2019. Retrieved March 23, 2019.
  162. ^ "Billion-Dollar Weather and Climate Disasters". National Centers for Environmental Information. Archived from the original on August 11, 2021. Retrieved August 23, 2021.
  163. ^ a b Blake, Eric S.; Zelensky, David A. Tropical Cyclone Report: Hurricane Harvey (PDF) (Report). National Hurricane Center. Archived (PDF) from the original on January 26, 2018. Retrieved August 23, 2021.
  164. ^ Franklin, James L. (February 22, 2006). Tropical Cyclone Report: Hurricane Vince (PDF) (Report). National Hurricane Center. Archived (PDF) from the original on October 2, 2015. Retrieved August 14, 2011.
  165. ^ Blake, Eric (September 18, 2020). Subtropical Storm Alpha Discussion Number 2 (Report). National Hurricane Center. Archived from the original on October 9, 2020. Retrieved September 18, 2020.
  166. from the original on May 23, 2022. Retrieved May 23, 2022.
  167. ^ Pielke, R. A. Jr.; Rubiera, J; Landsea, C; Fernández, M. L.; Klein, R (2003). "Hurricane Vulnerability in Latin America & The Caribbean" (PDF). National Hazards Review. Archived (PDF) from the original on August 10, 2006. Retrieved July 20, 2006.
  168. ^ Rappaport, Ed (December 9, 1993). Tropical Storm Bret Preliminary Report (GIF) (Report). National Hurricane Center. p. 3. Archived from the original on March 3, 2016. Retrieved August 11, 2015.
  169. ^ Landsea, Christopher W. (July 13, 2005). "Subject: Tropical Cyclone Names: G6) Why doesn't the South Atlantic Ocean experience tropical cyclones?". Tropical Cyclone Frequently Asked Question. United States National Oceanic and Atmospheric Administration's Hurricane Research Division. Archived from the original on March 27, 2015. Retrieved February 7, 2015.
  170. (PDF) from the original on August 30, 2021. Retrieved May 23, 2022.
  171. ^ National Oceanic and Atmospheric Administration. 2005 Tropical Eastern North Pacific Hurricane Outlook. Archived June 12, 2015, at the Wayback Machine. Retrieved May 2, 2006.
  172. ^ "Summer tropical storms don't fix drought conditions". ScienceDaily. May 27, 2015. Archived from the original on October 9, 2021. Retrieved April 10, 2021.
  173. .
  174. .
  175. from the original on June 14, 2012. Retrieved September 7, 2010.
  176. .
  177. ^ Alex Fox. (20 June 2023). "New Measurements Suggest Tropical Cyclones May Influence Global Climate". UC San Diego. Scripps Institution of Oceanography website Retrieved 30 June 2023.
  178. PMID 37339203
    .
  179. .
  180. .
  181. from the original on May 17, 2022. Retrieved May 27, 2022.
  182. ^ .
  183. from the original on January 19, 2022. Retrieved May 19, 2022.
  184. . Retrieved May 19, 2022.
  185. .
  186. ^ "Tempestade Leslie provoca grande destruição nas Matas Nacionais" [Storm Leslie wreaks havoc in the National Forests]. Notícias de Coimbra (in Portuguese). October 17, 2018. Archived from the original on January 28, 2019. Retrieved May 27, 2022.
  187. ^ Doyle, Thomas (2005). "Wind damage and Salinity Effects of Hurricanes Katrina and Rita on Coastal Baldcypress Forests of Louisiana" (PDF). Archived (PDF) from the original on March 4, 2016. Retrieved February 13, 2014.
  188. ^ Cappielo, Dina (2005). "Spills from hurricanes stain coast With gallery". Houston Chronicle. Archived from the original on April 25, 2014. Retrieved February 12, 2014.
  189. (PDF) from the original on January 20, 2022. Retrieved May 19, 2022.
  190. ^ . Retrieved May 21, 2022.
  191. . Retrieved May 19, 2022.
  192. . Retrieved May 19, 2022.
  193. ^ Cañedo, Sibely (March 29, 2019). "Tras el Huracán Willa, suben niveles de metales en río Baluarte" [After Hurricane Willa, metal levels rise in the Baluarte River] (in Spanish). Noreste. Archived from the original on September 30, 2020. Retrieved May 19, 2022.
  194. ^
    PMID 32818899
    .
  195. ^ from the original on January 25, 2021. Retrieved May 21, 2022.
  196. ^ from the original on May 17, 2022. Retrieved May 21, 2022.
  197. from the original on May 16, 2022. Retrieved May 21, 2022.
  198. . Retrieved May 21, 2022.
  199. from the original on May 12, 2022. Retrieved May 20, 2022.
  200. from the original on May 17, 2022. Retrieved May 27, 2022.
  201. from the original on May 17, 2022. Retrieved May 21, 2022.
  202. from the original on May 17, 2022. Retrieved May 21, 2022.
  203. from the original on May 17, 2022. Retrieved May 21, 2022.
  204. .
  205. from the original on May 17, 2022. Retrieved May 21, 2022.
  206. .
  207. ^ from the original on December 21, 2020. Retrieved March 13, 2021.
  208. ^ Liu, Kam-biu (1999). Millennial-scale variability in catastrophic hurricane landfalls along the Gulf of Mexico coast. 23rd Conference on Hurricanes and Tropical Meteorology. Dallas, TX: American Meteorological Society. pp. 374–377.
  209. S2CID 140723229
    .
  210. ^ G. Huang; W.W. S. Yim (January 2001). "Reconstruction of an 8,000-year record of typhoons in the Pearl River estuary, China" (PDF). University of Hong Kong. Archived (PDF) from the original on July 20, 2021. Retrieved April 2, 2021.
  211. ^ Arnold Court (1980). Tropical Cyclone Effects on California. NOAA technical memorandum NWS WR; 159. Northridge, California: California State University. pp. 2, 4, 6, 8, 34. Archived from the original on October 1, 2018. Retrieved February 2, 2012.
  212. ^ "Atlantic hurricane best track (HURDAT version 2)" (Database). United States National Hurricane Center. April 5, 2023. Retrieved April 12, 2024. Public Domain This article incorporates text from this source, which is in the public domain.
  213. ^ Philippe Caroff; et al. (June 2011). Operational procedures of TC satellite analysis at RSMC La Reunion (PDF) (Report). World Meteorological Organization. Archived from the original on April 27, 2021. Retrieved April 22, 2013.
  214. ^ Christopher W. Landsea; et al. "Documentation for 1851–1910 Alterations and Additions to the HURDAT Database". The Atlantic Hurricane Database Re-analysis Project. Hurricane Research Division. Archived from the original on June 15, 2021. Retrieved April 27, 2021.
  215. ^ Neumann, Charles J. "1.3: A Global Climatology". Global Guide to Tropical Cyclone Forecasting. Bureau of Meteorology. Archived from the original on June 1, 2011. Retrieved November 30, 2006.
  216. from the original on August 13, 2021. Retrieved April 17, 2021.
  217. ^ a b c d e Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division. "Frequently Asked Questions: When is hurricane season?". National Oceanic and Atmospheric Administration. Archived from the original on May 6, 2009. Retrieved July 25, 2006.
  218. ^ McAdie, Colin (May 10, 2007). "Tropical Cyclone Climatology". National Hurricane Center. Archived from the original on March 21, 2015. Retrieved June 9, 2007.
  219. ^ from the original on August 15, 2021.
  220. ^ Joint Typhoon Warning Center (2006). "3.3 JTWC Forecasting Philosophies" (PDF). United States Navy. Archived (PDF) from the original on November 29, 2007. Retrieved February 11, 2007.
  221. ^ .
  222. .
  223. from the original on August 15, 2021
  224. ^ a b c d Hurricane Research Division. "Frequently Asked Questions: What are the average, most, and least tropical cyclones occurring in each basin?". National Oceanic and Atmospheric Administration's Atlantic Oceanographic and Meteorological Laboratory. Retrieved December 5, 2012.
  225. ^ "Report on Cyclonic Disturbances Over North Indian Ocean During 2018" (PDF). Archived from the original (PDF) on May 11, 2020.
  226. ^ "Australian Tropical Cyclone Outlook for 2019 to 2020". Australian Bureau of Meteorology. October 11, 2019. Archived from the original on October 14, 2019. Retrieved October 14, 2019.
  227. ^ 2019–20 Tropical Cyclone Season Outlook [in the] Regional Specialised Meteorological Centre Nadi – Tropical Cyclone Centre (RSMC Nadi – TCC) Area of Responsibility (AOR) (PDF) (Report). Fiji Meteorological Service. October 11, 2019. Archived (PDF) from the original on October 11, 2019. Retrieved October 11, 2019.
  228. ^ Leonhardt, David; Moses, Claire; Philbrick, Ian Prasad (September 29, 2022). "Ian Moves North / Category 4 and 5 Atlantic hurricanes since 1980". The New York Times. Archived from the original on September 30, 2022. Source: NOAA - Graphic by Ashley Wu, The New York Times (cites for 2022— data)
  229. ^ Ajasa, Amudalat; Clement, Scott; Guskin, Emily (August 23, 2023). "Partisans remain split on climate change contributing to more disasters, and on their weather becoming more extreme". The Washington Post. Archived from the original on August 23, 2023.
  230. ^ "Major tropical cyclones have become '15% more likely' over past 40 years". Carbon Brief. May 18, 2020. Archived from the original on August 8, 2020. Retrieved August 31, 2020.
  231. (PDF) from the original on November 19, 2020. Retrieved October 6, 2020.
  232. ^ Collins, M.; Sutherland, M.; Bouwer, L.; Cheong, S.-M.; et al. (2019). "Chapter 6: Extremes, Abrupt Changes and Managing Risks" (PDF). IPCC Special Report on the Ocean and Cryosphere in a Changing Climate. p. 602. Archived (PDF) from the original on December 20, 2019. Retrieved October 6, 2020.
  233. from the original on January 5, 2020. Retrieved October 6, 2020.
  234. from the original on October 5, 2020. Retrieved October 6, 2020.
  235. ^ a b Collins, M.; Sutherland, M.; Bouwer, L.; Cheong, S.-M.; et al. (2019). "Chapter 6: Extremes, Abrupt Changes and Managing Risks" (PDF). IPCC Special Report on the Ocean and Cryosphere in a Changing Climate. p. 603. Archived (PDF) from the original on December 20, 2019. Retrieved October 6, 2020.
  236. ^ a b "Hurricane Harvey shows how we underestimate flooding risks in coastal cities, scientists say". The Washington Post. August 29, 2017. Archived from the original on August 30, 2017. Retrieved August 30, 2017.
  237. ^
    ISSN 2225-6032
    .
  238. .
  239. ^ "Hurricanes and Climate Change". Union of Concerned Scientists. Archived from the original on September 24, 2019. Retrieved September 29, 2019.
  240. PMID 32366651
    .
  241. .
  242. from the original on January 4, 2022. Retrieved January 4, 2022.
  243. ^ Florida Coastal Monitoring Program. "Project Overview". University of Florida. Archived from the original on May 3, 2006. Retrieved March 30, 2006.
  244. ^ "Observations". Central Pacific Hurricane Center. December 9, 2006. Archived from the original on February 12, 2012. Retrieved May 7, 2009.
  245. ^ "NOAA harnessing the power of new satellite data this hurricane season". National Oceanic and Atmospheric Administration. June 1, 2020. Archived from the original on March 18, 2021. Retrieved March 25, 2021.
  246. 53rd Weather Reconnaissance Squadron. Archived from the original on May 30, 2012. Retrieved March 30, 2006.{{cite web}}: CS1 maint: numeric names: authors list (link
    )
  247. ^ Lee, Christopher. "Drone, Sensors May Open Path Into Eye of Storm". The Washington Post. Archived from the original on November 11, 2012. Retrieved February 22, 2008.
  248. ^ National Hurricane Center (May 22, 2006). "Annual average model track errors for Atlantic basin tropical cyclones for the period 1994–2005, for a homogeneous selection of "early" models". National Hurricane Center Forecast Verification. National Oceanic and Atmospheric Administration. Archived from the original on May 10, 2012. Retrieved November 30, 2006.
  249. ^ National Hurricane Center (May 22, 2006). "Annual average official track errors for Atlantic basin tropical cyclones for the period 1989–2005, with least-squares trend lines superimposed". National Hurricane Center Forecast Verification. National Oceanic and Atmospheric Administration. Archived from the original on May 10, 2012. Retrieved November 30, 2006.
  250. ^ "Regional Specialized Meteorological Center". Tropical Cyclone Program (TCP). World Meteorological Organization. April 25, 2006. Archived from the original on August 14, 2010. Retrieved November 5, 2006.
  251. ^ Fiji Meteorological Service (2017). "Services". Archived from the original on June 18, 2017. Retrieved June 4, 2017.
  252. ^ Joint Typhoon Warning Center (2017). "Products and Service Notice". United States Navy. Archived from the original on June 9, 2017. Retrieved June 4, 2017.
  253. ^ National Hurricane Center (March 2016). "National Hurricane Center Product Description Document: A User's Guide to Hurricane Products" (PDF). National Oceanic and Atmospheric Administration. Archived (PDF) from the original on June 17, 2017. Retrieved June 3, 2017.
  254. ^ "Notes on RSMC Tropical Cyclone Information". Japan Meteorological Agency. 2017. Archived from the original on March 19, 2017. Retrieved June 4, 2017.
  255. ^ "Geopotential Height". National Weather Service. Archived from the original on March 24, 2022. Retrieved October 7, 2022.
  256. ^ "Constant Pressure Charts: 850 mb". National Weather Service. Archived from the original on May 4, 2022. Retrieved October 7, 2022.
  257. ^ "Constant Pressure Charts: 700 mb". National Weather Service. Archived from the original on June 29, 2022. Retrieved October 7, 2022.
  258. ^ "Constant Pressure Charts: 500 mb". National Weather Service. Archived from the original on May 21, 2022. Retrieved October 7, 2022.
  259. ^ "Constant Pressure Charts: 300 mb". National Weather Service. Archived from the original on October 7, 2022. Retrieved October 7, 2022.
  260. ^ "Constant Pressure Charts: 200 mb". National Weather Service. Archived from the original on August 5, 2022. Retrieved October 7, 2022.
  261. ^ "Hurricane Seasonal Preparedness Digital Toolkit". Ready.gov. February 18, 2021. Archived from the original on March 21, 2021. Retrieved April 6, 2021.
  262. .
  263. ^ Morrissey, Shirley A.; Reser, Joseph P. (May 1, 2003). "Evaluating the Effectiveness of Psychological Preparedness Advice in Community Cyclone Preparedness Materials". The Australian Journal of Emergency Management. 18 (2): 46–61. Archived from the original on May 23, 2022. Retrieved April 6, 2021.
  264. ^ "Tropical Cyclones". World Meteorological Organization. April 8, 2020. Archived from the original on December 18, 2023. Retrieved April 6, 2021.
  265. ^ "Fiji Meteorological Services". Ministry of Infrastructure & Meteorological Services. Ministry of Infrastructure & Transport. Archived from the original on August 14, 2021. Retrieved April 6, 2021.
  266. ^ "About the National Hurricane Center". Miami, Florida: National Hurricane Center. Archived from the original on October 12, 2020. Retrieved April 6, 2021.
  267. from the original on November 14, 2020. Retrieved April 6, 2021.
  268. ^ "National Hurricane Center - "Be Prepared"". Retrieved November 9, 2023.
  269. ^ National Ocean Service (September 7, 2016). "Follow That Hurricane!" (PDF). National Oceanic and Atmospheric Administration. Retrieved June 2, 2017.
  270. ^ "OSHA's Hazard Exposure and Risk Assessment Matrix for Hurricane Response and Recovery Work: List of Activity Sheets". U.S. Occupational Safety and Health Administration. 2005. Archived from the original on September 29, 2018. Retrieved September 25, 2018.
  271. ^ "Before You Begin – The Incident Command System (ICS)". American Industrial Hygiene Association. Archived from the original on September 29, 2018. Retrieved September 26, 2018.
  272. ^ "Volunteer". National Voluntary Organizations Active in Disaster. Archived from the original on September 29, 2018. Retrieved September 25, 2018.
  273. ^ a b c "Hurricane Key Messages for Employers, Workers and Volunteers". U.S. National Institute for Occupational Safety and Health. 2017. Archived from the original on November 24, 2018. Retrieved September 24, 2018.
  274. ^ a b "Hazardous Materials and Conditions". American Industrial Hygiene Association. Archived from the original on September 29, 2018. Retrieved September 26, 2018.
  275. ^ "Mold and Other Microbial Growth". American Industrial Hygiene Association. Archived from the original on September 29, 2018. Retrieved September 26, 2018.
  276. ^ a b c "OSHA's Hazard Exposure and Risk Assessment Matrix for Hurricane Response and Recovery Work: Recommendations for General Hazards Commonly Encountered during Hurricane Response and Recovery Operations". U.S. Occupational Safety and Health Administration. 2005. Archived from the original on September 29, 2018. Retrieved September 25, 2018.
  277. ^ "Electrical Hazards". American Industrial Hygiene Association. Archived from the original on September 29, 2018. Retrieved September 26, 2018.

External links