You are on page 1of 34

This is a featured article. Click here for more information.

Page semi-protected
Tropical cyclone
From Wikipedia, the free encyclopedia
Jump to navigationJump to search
"Hurricane" redirects here. For other uses, see Hurricane (disambiguation).
For technical reasons, "Hurricane #1" redirects here. For the band, see Hurricane
No. 1.

Hurricane Isabel (2003) as seen from orbit during Expedition 7 of the International
Space Station. The eye, eyewall, and surrounding rainbands, characteristics of
tropical cyclones in the narrow sense, are clearly visible in this view from space.
Part of a series on
Weather
Global tropical cyclone tracks-edit2.jpg
Temperate and polar seasons
Tropical seasons
Storms
Precipitation
Topics
Glossaries
Cumulus clouds in fair weather.jpeg Weather portal
vte
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/or squalls.
Depending on its location and strength, a tropical cyclone is referred to by
different names, including hurricane (/ˈhʌrɪkən, -keɪn/),[1][2][3] typhoon (/taɪ
ˈfuːn/), tropical storm, cyclonic storm, tropical depression, or simply cyclone.[4]
A hurricane is a tropical cyclone that occurs in the Atlantic Ocean and
northeastern Pacific Ocean, and a typhoon occurs in the northwestern Pacific Ocean;
in the south Pacific or Indian Ocean, comparable storms are referred to simply as
"tropical cyclones" or "severe cyclonic storms".[4]

"Tropical" refers to the geographical origin of these systems, which form almost
exclusively over tropical seas. "Cyclone" refers to their winds moving in a circle,
[5] whirling round their central clear eye, with their winds blowing
counterclockwise in the Northern Hemisphere and clockwise in the Southern
Hemisphere. The opposite direction of circulation is due to the Coriolis effect.
Tropical cyclones typically form over large bodies of relatively warm water. They
derive their energy through the evaporation of water from the ocean surface, which
ultimately recondenses into clouds and rain when moist air rises and cools to
saturation. This energy source differs from that of mid-latitude cyclonic storms,
such as nor'easters and European windstorms, which are fueled primarily by
horizontal temperature contrasts. Tropical cyclones are typically between 100 and
2,000 km (62 and 1,243 mi) in diameter. Every year during the late summer months
(July–September in the Northern Hemisphere and January–March in the Southern
Hemisphere), cyclones strike regions as far apart as the Gulf Coast of North
America, northwestern Australia, and eastern India and Bangladesh.

The strong rotating winds of a tropical cyclone are a result of the conservation of
angular momentum imparted by the Earth's rotation as air flows inwards toward the
axis of rotation. As a result, they rarely form within 5° of the equator.[6]
Tropical cyclones are almost unknown in the South Atlantic due to a consistently
strong wind shear and a weak Intertropical Convergence Zone.[7] Conversely, the
African easterly jet and areas of atmospheric instability give rise to cyclones in
the Atlantic Ocean and Caribbean Sea, while cyclones near Australia owe their
genesis to the Asian monsoon and 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 weaken quite rapidly
over land. This causes coastal regions to be particularly vulnerable to tropical
cyclones, compared to inland regions. 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 precipitation over a much smaller
area. This replenishing of moisture-bearing air after rain may cause multi-hour or
multi-day extremely heavy rain up to 40 kilometers (25 mi) from the coastline, far
beyond the amount of water that the local atmosphere holds at any one time. This in
turn can lead to river flooding, overland flooding, and a general overwhelming of
local water control structures across a large area. Although their effects on human
populations can be devastating, tropical cyclones can relieve drought conditions.
They also carry heat and energy away from the tropics and transport it towards
temperate latitudes, which could play an important role in regulating global
climate.

Part of a series on
Tropical cyclones
Structure
Central dense overcastDevelopmentEye
Effects
Warnings and watchesStorm surgePreparednessResponseNotable storms
Climatology and tracking
BasinsRSMCsScalesObservationForecastingRainfall forecastingRainfall climatology
Tropical cyclone naming
HistoryList of historical namesLists of retired names: Atlantic, Pacific hurricane,
Pacific typhoon, Philippine, Australian, South Pacific
Outline of tropical cyclones
Cyclone Catarina from the ISS on March 26 2004.JPG Tropical cyclones portal
vte

Contents
1 What is a tropical cyclone?
1.1 Wind field
1.2 Eye and center
1.3 Formation
1.4 Factors
1.5 Locations
1.6 Physics and energetics
1.7 Secondary circulation: a Carnot heat engine
1.8 Primary circulation: rotating winds
1.9 Maximum potential intensity
1.10 Interaction with the upper ocean
2 Movement
2.1 Environmental steering
2.2 Beta drift
2.3 Multiple storm interaction
2.4 Interaction with the mid-latitude westerlies
2.5 Landfall
3 Classification
3.1 Nomenclature and intensity classifications
3.2 Identification codes
3.3 Naming
4 Major basins and related warning centers
5 Preparations
6 Impacts
7 Response
8 Tropical Cyclone Intensity
8.1 Size
8.2 Rapid intensification
8.3 Dissipation
9 Climatology
9.1 Times
10 Influence of climate
10.1 Paleoclimate
10.2 Climate variability
10.3 Climate change
11 Observation and forecasting
11.1 Observation
11.2 Forecasting
12 Related cyclone types
13 Notable tropical cyclones
14 Popular culture
15 See also
15.1 Forecasting and preparation
15.2 Tropical cyclone seasons
16 References
17 External links
17.1 Warning centers
What is a tropical cyclone?

Diagram of a Northern hemisphere hurricane


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.[8][9] The
systems generally have a well-defined center which is surrounded by deep
atmospheric convection and a closed wind circulation at the surface.[8]
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 around 6000 B.C.[citation needed] However, 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.[10] These days, on
average around 80 to 90 named tropical cyclones form each year around the world, of
which over half of which develop hurricane-force winds of 65 kn (120 km/h; 75 mph)
or more.[10]

Tropical cyclones on either side of the Equator have their origins in the
Intertropical Convergence Zone, where winds blow from either the northeast or
southeast.[11] 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.
[11] These showers dissipate quite quickly, however, they can group together into
large clusters of thunderstorms.[11] This creates a flow of warm, moist, rapidly
rising air, which starts to rotate cyclonically as it interacts with the rotation
of the earth and leads to the development of a depression.[11]

Wind field
The near-surface wind field of a tropical cyclone is characterized by air rotating
rapidly around a center of circulation while also flowing radially inwards. At the
outer edge of the storm, air may be nearly calm; however, due to the Earth's
rotation, the air has non-zero absolute angular momentum. As air flows radially
inward, it begins to rotate cyclonically (counter-clockwise in the Northern
Hemisphere, and clockwise in the Southern Hemisphere) to conserve angular momentum.
At an inner radius, air begins to ascend to the top of the troposphere. This radius
is typically coincident with the inner radius of the eyewall, and has the strongest
near-surface winds of the storm; consequently, it is known as the radius of maximum
winds.[8] Once aloft, air flows away from the storm's center, producing a shield of
cirrus clouds.[12]

The previously mentioned processes result in a nearly axisymmetric wind field: Wind
speeds are low at the center, increase rapidly moving outwards to the radius of
maximum winds, and then decay more gradually with radius to large radii. However,
the wind field often exhibits additional spatial and temporal variability due to
the effects of localized processes, such as thunderstorm activity and horizontal
flow instabilities. In the vertical direction, winds are strongest near the surface
and decay with height within the troposphere.[13]

Eye and center


Main article: Eye (cyclone)

Thunderstorm activity in the eyewall of Cyclone Bansi as seen from the


International Space Station, on January 12, 2015
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 "eye". Weather in the eye is normally calm and
free of convective clouds, although the sea may be extremely violent.[14] 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.[15]
[16]

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".[16] 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.[14]

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.[17]

The eyewall may vary over time in the form of eyewall replacement cycles,
particularly in intense tropical cyclones. 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.[18]

Formation
Main article: Tropical cyclogenesis

Map of the cumulative tracks of all tropical cyclones during the 1985–2005 time
period. The Pacific Ocean west of the International Date Line sees more tropical
cyclones than any other basin, while there is almost no activity in the southern
hemisphere between Africa and 160˚W.

Map of all tropical cyclone tracks from 1945 to 2006. Equal-area projection.
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.[19]

Factors
Waves in the trade winds in the Atlantic Ocean—areas of converging winds that move
along the same track as the prevailing wind—create instabilities in the atmosphere
that may lead to the formation of hurricanes.
The formation of tropical cyclones is the topic of extensive ongoing research and
is still not fully understood.[20][needs update] While six factors appear to be
generally necessary, tropical cyclones may occasionally form without meeting all of
the following conditions. In most situations, water temperatures of at least 26.5
°C (79.7 °F) are needed down to a depth of at least 50 m (160 ft);[21] waters of
this temperature cause the overlying atmosphere to be unstable enough to sustain
convection and thunderstorms.[22] For tropical transitioning cyclones (i.e.
Hurricane Ophelia (2017)) a water temperature of at least 22.5 °C (72.5 °F) has
been suggested.[23]

Another factor is rapid cooling with height, which allows the release of the heat
of condensation that powers a tropical cyclone.[21] High humidity is needed,
especially in the lower-to-mid troposphere; when there is a great deal of moisture
in the atmosphere, conditions are more favorable for disturbances to develop.[21]
Low amounts of wind shear are needed, as high shear is disruptive to the storm's
circulation.[21] Tropical cyclones generally need to form more than 555 km (345 mi)
or five degrees of latitude away from the equator, allowing the Coriolis effect to
deflect winds blowing towards the low pressure center and creating a circulation.
[21] Lastly, a formative tropical cyclone needs a preexisting system of disturbed
weather. Tropical cyclones will not form spontaneously.[21] Low-latitude and low-
level westerly wind bursts associated with the Madden–Julian oscillation can create
favorable conditions for tropical cyclogenesis by initiating tropical disturbances.
[24]

Locations
Most tropical cyclones form in a worldwide band of thunderstorm activity near the
equator, referred to as the Intertropical Front (ITF), the Intertropical
Convergence Zone (ITCZ), or the monsoon trough.[25][26][27] Another important
source of atmospheric instability is found in tropical waves, which contribute to
the development of about 85% of intense tropical cyclones in the Atlantic Ocean and
become most of the tropical cyclones in the Eastern Pacific.[28][29][30] The
majority forms between 10 and 30 degrees of latitude away of the equator,[31] and
87% forms no farther away than 20 degrees north or south.[32][33] Because the
Coriolis effect initiates and maintains their rotation, tropical cyclones rarely
form or move within 5 degrees of the equator, where the effect is weakest.[32]
However, it is still possible for tropical systems to form within this boundary as
Tropical Storm Vamei and Cyclone Agni did in 2001 and 2004, respectively.[34][35]

Physics and energetics

Tropical cyclones exhibit an overturning circulation where air inflows at low


levels near the surface, rises in thunderstorm clouds, and outflows at high levels
near the tropopause.[36]
The three-dimensional wind field in a tropical cyclone can be separated into two
components: a "primary circulation" and a "secondary circulation". The primary
circulation is the rotational part of the flow; it is purely circular. The
secondary circulation is the overturning (in-up-out-down) part of the flow; it is
in the radial and vertical directions. The primary circulation is larger in
magnitude, dominating the surface wind field, and is responsible for the majority
of the damage a storm causes, while the secondary circulation is slower but governs
the energetics of the storm.[citation needed]

Secondary circulation: a Carnot heat engine


A tropical cyclone's primary energy source is heat from the evaporation of water
from the surface of a warm ocean, previously heated by sunshine. The energetics of
the system may be idealized as an atmospheric Carnot heat engine.[37] First,
inflowing air near the surface acquires heat primarily via evaporation of water
(i.e. latent heat) at the temperature of the warm ocean surface (during
evaporation, the ocean cools and the air warms). Second, the warmed air rises and
cools within the eyewall while conserving total heat content (latent heat is simply
converted to sensible heat during condensation). Third, air outflows and loses heat
via infrared radiation to space at the temperature of the cold tropopause. Finally,
air subsides and warms at the outer edge of the storm while conserving total heat
content. The first and third legs are nearly isothermal, while the second and
fourth legs are nearly isentropic. This in-up-out-down overturning flow is known as
the secondary circulation. The Carnot perspective provides an upper bound on the
maximum wind speed that a storm can attain.[citation needed]

Scientists estimate that a tropical cyclone releases heat energy at the rate of 50
to 200 exajoules (1018 J) per day,[38] equivalent to about 1 PW (1015 Watt). This
rate of energy release is equivalent to 70 times the world energy consumption of
humans and 200 times the worldwide electrical generating capacity, or to exploding
a 10-megaton nuclear bomb every 20 minutes.[38][39][needs update]

Primary circulation: rotating winds


The primary rotating flow in a tropical cyclone results from the conservation of
angular momentum by the secondary circulation. Absolute angular momentum on a
rotating planet {\displaystyle M}M is given by

{\displaystyle M={\frac {1}{2}}fr^{2}+vr}M = \frac{1}{2}fr^2 + vr


where {\displaystyle f}f is the Coriolis parameter, {\displaystyle v}v is the
azimuthal (i.e. rotating) wind speed, and {\displaystyle r}r is the radius to the
axis of rotation. The first term on the right hand side is the component of
planetary angular momentum that projects onto the local vertical (i.e. the axis of
rotation). The second term on the right hand side is the relative angular momentum
of the circulation itself with respect to the axis of rotation. Because the
planetary angular momentum term vanishes at the equator (where {\displaystyle
f=0}f=0 ), tropical cyclones rarely form within 5° of the equator.[6][40]

As air flows radially inward at low levels, it begins to rotate cyclonically in


order to conserve angular momentum. Similarly, as rapidly rotating air flows
radially outward near the tropopause, its cyclonic rotation decreases and
ultimately changes sign at large enough radius, resulting in an upper-level
anticyclone. The result is a vertical structure characterized by a strong cyclone
at low levels and a strong anticyclone near the tropopause; from thermal wind
balance, this corresponds to a system that is warmer at its center than in the
surrounding environment at all altitudes (i.e. "warm-core"). From hydrostatic
balance, the warm core translates to lower pressure at the center at all altitudes,
with the maximum pressure drop located at the surface.[13]

Maximum potential intensity


Main article: Maximum potential intensity
Due to surface friction, the inflow only partially conserves angular momentum.
Thus, the sea surface lower boundary acts as both a source (evaporation) and sink
(friction) of energy for the system. This fact leads to the existence of a
theoretical upper bound on the strongest wind speed that a tropical cyclone can
attain. Because evaporation increases linearly with wind speed (just as climbing
out of a pool feels much colder on a windy day), there is a positive feedback on
energy input into the system known as the Wind-Induced Surface Heat Exchange
(WISHE) feedback.[37] This feedback is offset when frictional dissipation, which
increases with the cube of the wind speed, becomes sufficiently large. This upper
bound is called the "maximum potential intensity", {\displaystyle v_{p}}v_p, and is
given by

{\displaystyle v_{p}^{2}={\frac {C_{k}}{C_{d}}}{\frac {T_{s}-T_{o}}{T_{o}}}\Delta


k}v_p^2 = \frac{C_k}{C_d}\frac{T_s - T_o}{T_o}\Delta k
where {\displaystyle T_{s}}T_{s} is the temperature of the sea surface,
{\displaystyle T_{o}}T_{o} is the temperature of the outflow ([K]),
{\displaystyle \Delta k}\Delta k is the enthalpy difference between the surface and
the overlying air ([J/kg]), and {\displaystyle C_{k}}C_{k} and {\displaystyle
C_{d}}C_{d} are the surface exchange coefficients (dimensionless) of enthalpy and
momentum, respectively.[41] The surface-air enthalpy difference is taken as
{\displaystyle \Delta k=k_{s}^{*}-k}\Delta k = k^*_s-k, where {\displaystyle
k_{s}^{*}}k^*_s is the saturation enthalpy of air at sea surface temperature and
sea-level pressure and {\displaystyle k}k is the enthalpy of boundary layer air
overlying the surface.

The maximum potential intensity is predominantly a function of the background


environment alone (i.e. without a tropical cyclone), and thus this quantity can be
used to determine which regions on Earth can support tropical cyclones of a given
intensity, and how these regions may evolve in time.[42][43] Specifically, the
maximum potential intensity has three components, but its variability in space and
time is due predominantly to the variability in the surface-air enthalpy difference
component {\displaystyle \Delta k}\Delta k.

Interaction with the upper ocean

Chart displaying the drop in surface temperature in the Gulf of Mexico as


Hurricanes Katrina and Rita passed over
The passage of a tropical cyclone over the ocean causes the upper layers of the
ocean to cool substantially, which can 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.[44]
Conversely, the mixing of the sea can result in heat being inserted in deeper
waters, with potential effects on global climate.[45]

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".[46]

Environmental steering
Environmental steering is the dominant term. Conceptually, it represents the
movement of the storm due to prevailing winds and other wider environmental
conditions, similar to "leaves carried along by a stream".[47] 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 in which the storm takes place. In this way,
tropical cyclone motion may be represented to first-order simply as advection of
the storm by the local environmental flow. This environmental flow is termed the
"steering flow".[citation needed]

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


west trade winds on the equatorial side of the subtropical ridge—a persistent high-
pressure area over the world's subtropical oceans.[47] 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.[29] These
waves are the precursors to many tropical cyclones within this region.[28] 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 Inter-tropical Convergence Zone and the monsoon trough.[48]
Additionally, tropical cyclone motion can be influenced by transient weather
systems, such as extratropical cyclones.[citation needed]

Beta drift
In addition to environmental steering, a tropical cyclone will tend to drift slowly
poleward and westward, a motion known as "beta drift". 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. It is
induced indirectly by the storm itself, the result of a feedback between the
cyclonic flow of the storm and its environment.[citation needed]

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.
[citation needed]

Multiple storm interaction


Main article: Fujiwhara effect
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.[49]

Interaction with the mid-latitude westerlies


See also: Westerlies

Storm track of Typhoon Ioke, showing recurvature off the 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 jet
stream or an extratropical cyclone. This motion, termed "recurvature", commonly
occurs near the western edge of the major ocean basins, where the jet stream
typically has a poleward component and extratropical cyclones are common.[50] An
example of tropical cyclone recurvature was Typhoon Ioke in 2006.[51]

Landfall
See also: Tropical cyclogenesis § Unusual areas of formation
The landfall of a tropical cyclone occurs when a storm's surface center (eye if a
stronger cyclone) moves over a coastline.[52] Storm conditions may be experienced
on the coast and inland hours before landfall; in fact, a tropical cyclone can
launch its strongest winds over land, yet not make landfall. NOAA uses the term
"direct hit" to describe when a location (on the left side of the eye) falls within
the radius of maximum winds (or twice that radius if on the right side), whether or
not the hurricane's eye made landfall.[52]

Classification
Nomenclature and intensity classifications
Main article: Tropical cyclone scales
Three tropical cyclones of the 2006 Pacific typhoon season at different stages of
development. The weakest (left) demonstrates only the most basic circular shape. A
stronger storm (top right) demonstrates 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, the structure of the system and its intensity. For example, within the
Northern Atlantic and Eastern Pacific basins, a tropical cyclone with wind speeds
of over 65 kn (75 mph; 120 km/h) is called a hurricane, while it is called a
typhoon or a severe cyclonic storm within the Western Pacific or North Indian
Oceans.[53][54][55] 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.[56][57]

Tropical Cyclone Classifications


Beaufort
scale 1-minute sustained winds
(NHC/CPHC/JTWC) 10-minute sustained winds
(WMO/JMA/MF/BOM/FMS) NE Pacific &
N Atlantic
NHC/CPHC[53] NW Pacific
JTWC NW Pacific
JMA N Indian Ocean
IMD[55] SW Indian Ocean
MF Australia & S Pacific
BOM/FMS[57]
0–7 <32 knots (37 mph; 59 km/h) <28 knots (32 mph; 52 km/h) Tropical
Depression Tropical Depression Tropical Depression Depression Zone of
Disturbed Weather Tropical Disturbance
7 33 knots (38 mph; 61 km/h) 28–29 knots (32–33 mph; 52–54 km/h) Deep
Depression Tropical Disturbance Tropical Depression
8 34–37 knots (39–43 mph; 63–69 km/h) 30–33 knots (35–38 mph; 56–61 km/h)
Tropical Storm Tropical Storm Tropical Depression Tropical Low
9–10 38–54 knots (44–62 mph; 70–100 km/h) 34–47 knots (39–54 mph; 63–87 km/h)
Tropical Storm Cyclonic Storm Moderate
Tropical Storm Category 1
Tropical Cyclone
11 55–63 knots (63–72 mph; 102–117 km/h) 48–55 knots (55–63 mph; 89–102
km/h) Severe
Tropical Storm Severe
Cyclonic Storm Severe
Tropical Storm Category 2
Tropical Cyclone
12+ 64–71 knots (74–82 mph; 119–131 km/h) 56–63 knots (64–72 mph; 104–117
km/h) Category 1
Hurricane Typhoon
72–82 knots (83–94 mph; 133–152 km/h) 64–72 knots (74–83 mph; 119–133 km/h)
Typhoon Very Severe
Cyclonic Storm Tropical Cyclone Category 3
Severe Tropical Cyclone
83–95 knots (96–109 mph; 154–176 km/h) 73–83 knots (84–96 mph; 135–154 km/h)
Category 2
Hurricane
96–97 knots (110–112 mph; 178–180 km/h) 84–85 knots (97–98 mph; 156–157 km/h)
Category 3
Major Hurricane Very Strong Typhoon
98–112 knots (113–129 mph; 181–207 km/h) 86–98 knots (99–113 mph; 159–181 km/h)
Extremely Severe
Cyclonic Storm Intense
Tropical Cyclone Category 4
Severe Tropical Cyclone
113–122 knots (130–140 mph; 209–226 km/h) 99–107 knots (114–123 mph; 183–198 km/h)
Category 4
Major Hurricane
123–129 knots (142–148 mph; 228–239 km/h) 108–113 knots (124–130 mph; 200–209 km/h)
Violent Typhoon Category 5
Severe Tropical Cyclone
130–136 knots (150–157 mph; 241–252 km/h) 114–119 knots (131–137 mph; 211–220 km/h)
Super
Typhoon Super
Cyclonic Storm Very Intense
Tropical Cyclone
>136 knots (157 mph; 252 km/h) >120 knots (138 mph; 222 km/h) Category 5
Major Hurricane
Identification codes
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. These codes start at 01 every year and are assigned in order to
systems, which have the potential to develop further, cause significant impact to
life and property or when the warning centers start to write advisories on the
system.[57][58]

Tropical Cyclone Numbering[58][59][60]


Naming
Main articles: Tropical cyclone naming and History of tropical cyclone naming
The practice of using names to identify tropical cyclones goes back many years,
with systems named after places or things they hit before the formal start of
naming.[67][68] The system currently used provides positive identification of
severe weather systems in a brief form, that is readily understood and recognized
by the public.[67][68] The credit for the first usage of personal names for weather
systems is generally given to the Queensland Government Meteorologist Clement
Wragge who named systems between 1887 and 1907.[67][68] 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.[67][68] 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.[68]

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


services and retain their names throughout their lifetimes to provide ease of
communication between forecasters and the general public regarding forecasts,
watches, and warnings.[67] Since the systems can last a week or longer and more
than one can be occurring in the same basin at the same time, the names are thought
to reduce the confusion about what storm is being described.[67] Names are assigned
in order from predetermined lists with one, three, or ten-minute sustained wind
speeds of more than 65 km/h (40 mph) depending on which basin it originates.[53]
[55][56] 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.[56][57] 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.[53][54][57]

Major basins and related warning centers


Main articles: Tropical cyclone basins and Regional Specialized Meteorological
Centre
Tropical cyclone basins and official warning centers
Basin Warning center Area of responsibility Notes
Northern Hemisphere
North Atlantic United States National Hurricane Center Equator northward,
African Coast – 140°W [53]
Eastern Pacific United States Central Pacific Hurricane Center Equator northward,
140–180°W [53]
Western Pacific Japan Meteorological Agency Equator – 60°N, 180–100°E [54]
North Indian Ocean India Meteorological Department Equator northwards, 100–
40°E [55]
Southern Hemisphere
South-West
Indian Ocean Météo-France Reunion Equator – 40°S, African Coast – 90°E
[56]
Australian region Indonesian Meteorology, Climatology,
and Geophysical Agency (BMKG) Equator – 10°S, 90–141°E [57]
Papua New Guinea National Weather Service Equator – 10°S, 141–160°E [57]
Australian Bureau of Meteorology 10–40°S, 90–160°E [57]
Southern Pacific Fiji Meteorological Service Equator – 25°S, 160°E – 120°W [57]
Meteorological Service of New Zealand 25–40°S, 160°E – 120°W [57]
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.[10] Ten of these warning centres worldwide are designated as either a
Regional Specialized Meteorological Centre or a Tropical Cyclone Warning Centre by
the World Meteorological Organisation's tropical cyclone programme.[10] These
warning centres issue advisories which provide basic information and cover a
systems present, forecast position, movement and intensity, in their designated
areas of responsibility.[10] Meteorological services around the world are generally
responsible for issuing warnings for their own country, however, there are
exceptions, as the United States National Hurricane Center and Fiji Meteorological
Service issue alerts, watches and warnings for various island nations in their
areas of responsibility.[10][57] The United States Joint Typhoon Warning Center
(JTWC) and Fleet Weather Center (FWC) also publicly issue warnings, about tropical
cyclones on behalf of the United States Government.[10] 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.[69]

Preparations
Main article: Tropical cyclone preparedness
Ahead of the formal season starting, people are urged to prepare for the effects of
a tropical cyclone by politicians and weather forecasters amongst others. They
prepare by determining their risk to the different types of weather, tropical
cyclones cause, checking their insurance coverage and emergency supplies, as well
as determining where to evacuate to if needed. When a tropical cyclone develops and
is forecast to impact land each member nation of the World Meteorological
Organization issues various watches and warnings to cover the expected impacts.
However, there are some exceptions with the United States National Hurricane Center
and Fiji Meteorological Service responsible for issuing warnings for other nations
in their area of responsibility.[citation needed]

Impacts
Main article: Effects of tropical cyclones
Tropical cyclones out at sea cause large waves, heavy rain, flood and high winds,
disrupting international shipping and, at times, causing shipwrecks.[70] Tropical
cyclones stir up water, leaving a cool wake behind them, which causes the region to
be less favorable for subsequent tropical cyclones.[44] 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.[71]
The broad rotation of a landfalling tropical cyclone, and vertical wind shear at
its periphery, spawns tornadoes. Tornadoes can also be spawned as a result of
eyewall mesovortices, which persist until landfall.[72]

Over the past two centuries, tropical cyclones have been responsible for the deaths
of about 1.9 million people worldwide. Large areas of standing water caused by
flooding lead to infection, as well as contributing to mosquito-borne illnesses.
Crowded evacuees in shelters increase the risk of disease propagation.[71] Tropical
cyclones significantly interrupt infrastructure, leading to power outages, bridge
destruction, and the hampering of reconstruction efforts.[71][73] On average, the
Gulf and east coasts of the United States suffer approximately US$5 billion (1995
US $) in cyclone damage every year. The majority (83%) of tropical cyclone damage
is caused by severe hurricanes, category 3 or greater. However, category 3 or
greater hurricanes only account for about one-fifth of cyclones that make landfall
every year.[74]

Although cyclones take an enormous toll in lives and personal property, they may be
important factors in the precipitation regimes of places they impact, as they may
bring much-needed precipitation to otherwise dry regions.[75] Tropical cyclones
also help maintain the global heat balance by moving warm, moist tropical air to
the middle latitudes and polar regions,[76] and by regulating the thermohaline
circulation through upwelling.[77] The storm surge and winds of hurricanes may be
destructive to human-made structures, but they also stir up the waters of coastal
estuaries, which are typically important fish breeding locales. Tropical cyclone
destruction spurs redevelopment, greatly increasing local property values.[78]

When hurricanes surge upon shore from the ocean, salt is introduced to many
freshwater areas and raises the salinity levels too high for some habitats to
withstand. Some are able to cope with the salt and recycle it back into the ocean,
but others can not release the extra surface water quickly enough or do not have a
large enough freshwater source to replace it. Because of this, some species of
plants and vegetation die due to the excess salt.[79] In addition, hurricanes can
carry toxins and acids onto shore when they make landfall. The flood water can pick
up the toxins from different spills and contaminate the land that it passes over.
The toxins are very harmful to the people and animals in the area, as well as the
environment around them. The flooding water can also spark many dangerous oil
spills.[80]

Response
Main article: Tropical cyclone response

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.
[81] Hurricane response requires coordination between federal, tribal, state,
local, and private entities.[82] 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.[83]

Hurricane responders face many hazards. Hurricane responders may be exposed to


chemical and biological contaminants including stored chemicals, sewage, human
remains, and mold growth encouraged by flooding,[84][85][86] as well as asbestos
and lead that may be present in older buildings.[85][87] Common injuries arise from
falls from heights, such as from a ladder or from level surfaces; from
electrocution in flooded areas, including from backfeed from portable generators;
or from motor vehicle accidents.[84][87][88] Long and irregular shifts may lead to
sleep deprivation and fatigue, increasing the risk of injuries, and workers may
experience mental stress associated with a traumatic incident. Additionally, 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.[84]
[87]

Tropical Cyclone Intensity


Size
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[89]
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 wind (i.e. gale
force), the radius of outermost closed isobar (ROCI), and the radius of vanishing
wind.[90][91] An additional metric is the radius at which the cyclone's relative
vorticity field decreases to 1×10−5 s−1.[16]

On Earth, tropical cyclones span a large range of sizes, from 100–2,000 kilometres
(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.[92] 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)).[89] 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.[91][92]

Rapid intensification
Main article: 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 knots or more within 24 hours.[93] 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.[94]
However, some cyclones such as Hurricane Epsilon have rapidly intensified despite
relatively unfavorable conditions.[95][96]

Dissipation
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 been declared dissipated or lost its tropical characteristics, there is
a chance that a tropical cyclone could regenerate should it move back over open
waters.[citation needed]

Landfall
Should a tropical cyclone make landfall or pass over an island, then its
circulation could start to break down, especially if it is a mountainous island.
[97] Should a system make landfall on a continent such as Africa, Australia, India
or the United States, then it would be cut off from its supply of warm moist
maritime air and would start to draw in dry continental air.[97] This combined with
the increased friction over land areas, leads to the weakening and dissipation of
the tropical cyclone.[97] Over a mountainous terrain, a system can quickly weaken,
however, over flat areas, it may take two to three days for the circulation to
break down and dissipate.[97]

Factors

Tropical Storm Kyle, in 2020, is an example of a sheared tropical cyclone, with


deep convection slightly removed from the center of the system.
A tropical cyclone can dissipate when it moves over waters significantly below 26.5
°C (79.7 °F). This will cause the storm to lose its tropical characteristics, such
as a warm core with thunderstorms near the center, and become a remnant low-
pressure area. These remnant systems may persist for up to several days before
losing their identity. This dissipation mechanism is most common in the eastern
North Pacific. Weakening or dissipation can occur if it experiences vertical wind
shear, causing the convection and heat engine to move away from the center; this
normally ceases development of a tropical cyclone.[98] In addition, its interaction
with the main belt of the Westerlies, by means of merging with a nearby frontal
zone, can cause tropical cyclones to evolve into extratropical cyclones. This
transition can take 1–3 days.[99]

Artificial dissipation
Over the years, there have been a number of techniques considered to try and
artificially modify tropical cyclones.[100] 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.[100]
However, these techniques have failed to appreciate the duration, intensity, power
or size of tropical cyclones.[100]

Climatology
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 around 6000 B.C. However, 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. As a result, the ability of climatologists
to make long-term analysis of tropical cyclones is limited by the amount of
reliable historical data. However, reanalyses and research are being undertaken to
extend the historical record, through the usage of poxy data such as overwash
deposits, beach ridges and historical documents such as diaries. During the 1940s,
routine aircraft reconnaissance started in both the Atlantic and Western Pacific
basin during the mid-1940s, which provided ground truth data, however, early
flights were only made once or twice a day.[citation needed]

Times
In the Northern Atlantic Ocean, a distinct cyclone season occurs from June 1 to
November 30, sharply peaking from late August through September.[19] 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.[101] The Northwest Pacific sees tropical cyclones year-round, with a
minimum in February and March and a peak in early September.[19] In the North
Indian basin, storms are most common from April to December, with peaks in May and
November.[19] 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.[19]
[57]

Season lengths and averages


Basin Season
start Season
end Tropical
cyclones Refs
North Atlantic June 1 November 30 12.1 [102]
Eastern Pacific May 15 November 30 16.6 [102]
Western Pacific January 1 December 31 26.0 [102]
North Indian January 1 December 31 12 [103]
South-West Indian July 1 June 30 9.3 [102][56]
Australian region November 1 April 30 11.0 [104]
Southern Pacific November 1 April 30 7.1 [105]
Total: 94.1
Influence of climate
See also: Atlantic hurricane reanalysis project
Ambox current red Asia Australia.svg
This section needs to be updated. Please update this article to reflect recent
events or newly available information. (September 2020)

Atlantic Multidecadal Variability (1856–2013)


Often in part because of the threat of hurricanes, many coastal regions had sparse
population between major ports until the advent of automobile tourism; therefore,
the most severe portions of hurricanes striking the coast may have gone unmeasured
in some instances. The combined effects of ship destruction and remote landfall
severely limit the number of intense hurricanes in the official record before the
era of hurricane reconnaissance aircraft and satellite meteorology. Although the
record shows a distinct increase in the number and strength of intense hurricanes,
therefore, experts regard the early data as suspect.[106]

Paleoclimate
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.[107][108] Few major hurricanes struck the Gulf coast during 3000–1400
BC and again during the most recent millennium. These quiescent intervals were
separated by a hyperactive period during 1400 BC and 1000 AD, when the Gulf coast
was struck frequently by catastrophic hurricanes and their landfall probabilities
increased by 3–5 times. This millennial-scale variability has been attributed to
long-term shifts in the position of the Azores High,[108] which may also be linked
to changes in the strength of the North Atlantic oscillation.[109]

Climate variability
Of various modes of variability, El Niño-Southern Oscillation has the largest
impact on tropical cyclone activity.[110] Most tropical cyclones form on the side
of the subtropical ridge closer to the equator, then move poleward past the ridge
axis before recurving into the main belt of the Westerlies.[111] When the
subtropical ridge position shifts due to El Niño, so will the preferred tropical
cyclone tracks. Areas west of Japan and Korea tend to experience much fewer
September–November tropical cyclone impacts during El Niño and neutral years.
During El Niño years, the break in the subtropical ridge tends to lie near 130°E
which would favor the Japanese archipelago.[112] During El Niño years, Guam's
chance of a tropical cyclone impact is one-third more likely than of the long-term
average.[113] The tropical Atlantic Ocean experiences depressed activity due to
increased vertical wind shear across the region during El Niño years.[114] During
La Niña years, the formation of tropical cyclones, along with the subtropical ridge
position, shifts westward across the western Pacific Ocean, which increases the
landfall threat to China and much greater intensity in the Philippines.[112]
Although more common since 1995, few above-normal hurricane seasons occurred during
1970–94.[115] Destructive hurricanes struck frequently from 1926 to 1960, including
many major New England hurricane you so in browser s. Twenty-one Atlantic tropical
storms formed in 1933, a record only recently exceeded in 2005, which saw 28
storms. Tropical hurricanes occurred infrequently during the seasons of 1900–25;
however, many intense storms formed during 1870–99. During the 1887 season, 19
tropical storms formed, of which a record 4 occurred after November 1 and 11
strengthened into hurricanes. Few hurricanes occurred in the 1840s to 1860s;
however, many struck in the early 19th century, including an 1821 storm that made a
direct hit on New York City. Some historical weather experts say these storms may
have been as high as Category 4 in strength.[116]

According to the Azores High hypothesis, an anti-phase pattern is expected to exist


between the Gulf of Mexico coast and the Atlantic coast. During the quiescent
periods, a more northeasterly position of the Azores High would result in more
hurricanes being steered towards the Atlantic coast. During the hyperactive period,
more hurricanes were steered towards the Gulf coast as the Azores High was shifted
to a more southwesterly position near the Caribbean. Such a displacement of the
Azores High is consistent with paleoclimatic evidence that shows an abrupt onset of
a drier climate in Haiti around 3200 14C years BP,[117] and a change towards more
humid conditions in the Great Plains during the late-Holocene as more moisture was
pumped up the Mississippi Valley through the Gulf coast.

Climate change
Main article: Tropical cyclones and climate change
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 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.[118]

Tropical cyclones use warm, moist air as their fuel. As climate change is warming
ocean temperatures, there is potentially more of this fuel available.[119] 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.[120] With 2 °C warming, a greater percentage (+13%) of
tropical cyclones are expected to reach Category 4 and 5 strength.[118] 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.[121]

There is currently no consensus on how climate change will affect the overall
frequency of tropical cyclones.[118] A majority of climate models show a decreased
frequency in future projections.[122] 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.[123] Observations have shown
little change in the overall frequency of tropical cyclones worldwide,[124] with
increased frequency in the North Atlantic and central Pacific, and significant
decreases in the southern Indian Ocean and western North Pacific.[125]

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.[126] In
the North Pacific, there may also have been an eastward expansion.[127] 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.[122]

Warmer air can hold more water vapor: the theoretical maximum water vapor content
is given by the Clausius–Clapeyron relation, which yields ≈7% increase in water
vapor in the atmosphere per 1 °C warming.[128][129] All models that were assessed
in a 2019 review paper show a future increase of rainfall rates.[118] Additional
sea level rise will increase storm surge levels.[127][130] It is plausible that
extreme wind waves see an increase as a consequence of changes in tropical
cyclones, further exacerbating storm surge dangers to coastal communities.[122] The
compounding effects from floods, storm surge, and terrestrial flooding (rivers) are
projected to increase due to global warming.[130]

Observation and forecasting


Observation
Main article: Tropical cyclone observation

Sunset view of Hurricane Isidore's rainbands photographed at 7,000 feet (2,100 m)

"Hurricane Hunter" – WP-3D Orion is used to go into the eye of a hurricane for data
collection and measurements purposes.
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.
[131]

Tropical cyclones far from land are tracked by weather satellites capturing 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.[132]

In situ measurements, in real-time, can be taken by sending specially equipped


reconnaissance flights into the cyclone. In the Atlantic basin, these flights are
regularly flown by United States government hurricane hunters.[133] The aircraft
used are WC-130 Hercules and WP-3D Orions, both four-engine turboprop cargo
aircraft. These aircraft fly directly into the cyclone and take direct and remote-
sensing measurements. The aircraft also launch GPS dropsondes inside the cyclone.
These sondes measure temperature, humidity, pressure, and especially winds between
flight level and the ocean's surface. A new era in hurricane observation began when
a remotely piloted Aerosonde, a small drone aircraft, was flown through Tropical
Storm Ophelia as it passed Virginia's Eastern Shore during the 2005 hurricane
season. A similar mission was also completed successfully in the western Pacific
Ocean. This demonstrated a new way to probe the storms at low altitudes that human
pilots seldom dare.[134]

A general decrease in error trends in tropical cyclone path prediction is evident


since the 1970s
Forecasting
See also: Tropical cyclone track forecasting, Tropical cyclone prediction model,
and Tropical cyclone rainfall forecasting
Because of the forces that affect tropical cyclone tracks, accurate track
predictions depend on determining the position and strength of high- and low-
pressure areas, and predicting how those areas will change during the life of a
tropical system. The deep layer mean flow, or average wind through the depth of the
troposphere, is considered the best tool in determining track direction and speed.
If storms are significantly sheared, use of wind speed measurements at a lower
altitude, such as at the 70 kPa pressure surface (3,000 metres or 9,800 feet above
sea level) will produce better predictions. Tropical forecasters also consider
smoothing out short-term wobbles of the storm as it allows them to determine a more
accurate long-term trajectory.[135] High-speed computers and sophisticated
simulation software allow forecasters to produce computer models that predict
tropical cyclone tracks based on the future position and strength of high- and low-
pressure systems. Combining forecast models with increased understanding of the
forces that act on tropical cyclones, as well as with a wealth of data from Earth-
orbiting satellites and other sensors, scientists have increased the accuracy of
track forecasts over recent decades.[136] However, scientists are not as skillful
at predicting the intensity of tropical cyclones.[137] 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.[138][139][140][141][142]

Related cyclone types

Hurricane Gustav on September 9, 2002, the first system to be given a name as a


subtropical cyclone
See also: Cyclone, Extratropical cyclone, and Subtropical cyclone
In addition to tropical cyclones, there are two other classes of cyclones within
the spectrum of cyclone types. These kinds of cyclones, known as extratropical
cyclones and subtropical cyclones, can be stages a tropical cyclone passes through
during its formation or dissipation.[143] An extratropical cyclone is a storm that
derives energy from horizontal temperature differences, which are typical in higher
latitudes. A tropical cyclone can become extratropical as it moves toward higher
latitudes if its energy source changes from heat released by condensation to
differences in temperature between air masses; although not as frequently, an
extratropical cyclone can transform into a subtropical storm, and from there into a
tropical cyclone.[144] From space, extratropical storms have a characteristic
"comma-shaped" cloud pattern.[145] Extratropical cyclones can also be dangerous
when their low-pressure centers cause powerful winds and high seas.[146]

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.[147]

Notable tropical cyclones


Main articles: List of tropical cyclone records, Lists of Atlantic hurricanes, and
List of Pacific hurricanes
Tropical cyclones that cause extreme destruction are rare, although when they
occur, they can cause great amounts of damage or thousands of fatalities.

Flooding after the 1991 Bangladesh cyclone, which killed around 140,000 people.
The 1970 Bhola cyclone is considered to be the deadliest tropical cyclone on
record, which killed around 300,000 people, after striking the densely populated
Ganges Delta region of Bangladesh on November 13, 1970.[148] Its powerful storm
surge was responsible for the high death toll.[149] The North Indian cyclone basin
has historically been the deadliest basin.[71][150] Elsewhere, Typhoon Nina killed
nearly 100,000 in China in 1975 due to a 100-year flood that caused 62 dams
including the Banqiao Dam to fail.[151] The Great Hurricane of 1780 is the
deadliest North Atlantic hurricane on record, killing about 22,000 people in the
Lesser Antilles.[152] A tropical cyclone does not need to be particularly strong to
cause memorable damage, primarily if the deaths are from rainfall or mudslides.
Tropical Storm Thelma in November 1991 killed thousands in the Philippines,[153]
although the strongest typhoon to ever make landfall on record was Typhoon Haiyan
in November 2013, causing widespread devastation in Eastern Visayas, and killing at
least 6,300 people in the Philippines alone. In 1982, the unnamed tropical
depression that eventually became Hurricane Paul killed around 1,000 people in
Central America.[154]

Hurricane Harvey and Hurricane Katrina are estimated to be the costliest tropical
cyclones to impact the United States mainland, each causing damage estimated at
$125 billion.[155] Harvey killed at least 90 people in August 2017 after making
landfall in Texas as a low-end Category 4 hurricane. Hurricane Katrina is estimated
as the second-costliest tropical cyclone worldwide,[156] causing $81.2 billion in
property damage (2008 USD) alone,[157] with overall damage estimates exceeding $100
billion (2005 USD).[156] Katrina killed at least 1,836 people after striking
Louisiana and Mississippi as a major hurricane in August 2005.[157] Hurricane Maria
is the third most destructive tropical cyclone in U.S. history, with damage
totaling $91.61 billion (2017 USD), and with damage costs at $68.7 billion (2012
USD), Hurricane Sandy is the fourth most destructive tropical cyclone in U.S
history. The Galveston Hurricane of 1900 is the deadliest natural disaster in the
United States, killing an estimated 6,000 to 12,000 people in Galveston, Texas.
[158] Hurricane Mitch caused more than 10,000 fatalities in Central America, making
it the second deadliest Atlantic hurricane in history. Hurricane Iniki in 1992 was
the most powerful storm to strike Hawaii in recorded history, hitting Kauai as a
Category 4 hurricane, killing six people, and causing U.S. $3 billion in damage.
[159] Other destructive Eastern Pacific hurricanes include Pauline and Kenna, both
causing severe damage after striking Mexico as major hurricanes.[160][161] In March
2004, Cyclone Gafilo struck northeastern Madagascar as a powerful cyclone, killing
74, affecting more than 200,000 and becoming the worst cyclone to affect the nation
for more than 20 years.[162]

The relative sizes of Typhoon Tip, Cyclone Tracy, and the Contiguous United States
The most intense storm on record is Typhoon Tip in the northwestern Pacific Ocean
in 1979, which reached a minimum pressure of 870 hectopascals (25.69 inHg) and
maximum sustained wind speeds of 165 knots (85 m/s) or 190 miles per hour (310
km/h).[163] The highest maximum sustained wind speed ever recorded was 185 knots
(95 m/s) or 215 miles per hour (346 km/h) in Hurricane Patricia in 2015—the most
intense cyclone ever recorded in the Western Hemisphere.[164] Typhoon Nancy in 1961
also had recorded wind speeds of 185 knots (95 m/s) or 215 miles per hour (346
km/h), but recent research indicates that wind speeds from the 1940s to the 1960s
were gauged too high, and thus it is no longer considered the storm with the
highest wind speed on record.[165] Likewise, a surface-level gust caused by Typhoon
Paka on Guam in late 1997 was recorded at 205 knots (105 m/s) or 235 miles per hour
(378 km/h). Had it been confirmed, it would be the strongest non-tornadic wind ever
recorded on the Earth's surface, but the reading had to be discarded since the
anemometer was damaged by the storm.[166] The World Meteorological Organization
established Barrow Island (Queensland) as the location of the highest non-tornado
related wind gust at 408 kilometres per hour (254 mph)[167] on April 10, 1996,
during Severe Tropical Cyclone Olivia.[168]

In addition to being the most intense tropical cyclone on record based on pressure,
Tip is the largest cyclone on record, with tropical storm-force winds 2,170
kilometres (1,350 mi) in diameter. The smallest storm on record, Tropical Storm
Marco, formed during October 2008 and made landfall in Veracruz. Marco generated
tropical storm-force winds only 37 kilometres (23 mi) in diameter. Marco broke the
record of 1974's Cyclone Tracy.[169]
Hurricane John is the longest-lasting tropical cyclone on record, lasting 31 days
in 1994. Before the advent of satellite imagery in 1961, however, many tropical
cyclones were underestimated in their durations.[170] John is also the longest-
tracked tropical cyclone in the Northern Hemisphere on record, with a path of 8,250
mi (13,280 km).[171] Cyclone Rewa of the 1993–94 South Pacific and Australian
region cyclone seasons had one of the longest tracks observed within the Southern
Hemisphere, traveling a distance of over 5,545 mi (8,920 km) during December 1993
and January 1994.[171]

Popular culture
Main article: Tropical cyclones in popular culture
In popular culture, tropical cyclones have made several appearances in different
types of media, including films, books, television, music, and electronic games.
[172] These media often portray tropical cyclones that are either entirely
fictional or based on real events.[172] For example, George Rippey Stewart's Storm,
a best-seller published in 1941, is thought to have influenced meteorologists on
their decision to assign female names to Pacific tropical cyclones.[68] Another
example is the hurricane in The Perfect Storm, which describes the sinking of the
Andrea Gail by the 1991 Perfect Storm.[173] Hurricanes have been featured in parts
of the plots of series such as The Simpsons, Invasion, Family Guy, Seinfeld,
Dawson's Creek, Burn Notice and CSI: Miami.[172][174][175][176][177] The 2004 film
The Day After Tomorrow includes several mentions of actual tropical cyclones and
features fantastical "hurricane-like", albeit non-tropical, Arctic storms.[178]
[179]

See also
Tropical cyclones portal
Disaster preparedness
Extraterrestrial vortex
Space hurricane
Hurricane Alley
Hypercane
List of Atlantic hurricanes
List of the most intense tropical cyclones
List of tropical cyclone records
List of wettest tropical cyclones by country
Outline of tropical cyclones
Whirlwind
Forecasting and preparation
Catastrophe modeling
Tropical cyclone engineering
Hurricane-proof building
Tropical cyclone watches and warnings
Tropical cyclone seasons
Tropical cyclones in 2021
Atlantic hurricane (2021 Atlantic hurricane season)
Pacific hurricane season (2021 Pacific hurricane season)
Pacific typhoon season (2021 Pacific typhoon season)
North Indian Ocean tropical cyclone (2021 North Indian Ocean cyclone season
South-West Indian Ocean tropical cyclone (2020–21 South-West Indian Ocean cyclone
season)
Australian region tropical cyclone (2020–21 Australian region cyclone season)
South Pacific tropical cyclone (2020–21 South Pacific cyclone season)
Mediterranean tropical-like cyclone
South Atlantic tropical cyclone
References
"hurricane"
. Oxford dictionary. Archived
from the original on October 6, 2014. Retrieved October 1, 2014.
"Hurricane – Definition and More from the Free Merriam-Webster Dictionary"
. Archived
from the original on September 12, 2017. Retrieved October 1, 2014.
"Definition of "hurricane" – Collins English Dictionary"
. Archived
from the original on October 6, 2014. Retrieved October 1, 2014.
"What is the difference between a hurricane, a cyclone, and a typhoon?"
. OCEAN FACTS. National Ocean Service. Archived
from the original on December 25, 2016. Retrieved December 24, 2016.
"Oxford English Dictionary"
. OED Online. Oxford University Press. 2017. Retrieved September 10, 2017. a
general term for all storms or atmospheric disturbances in which the wind has a
circular or whirling course.
Henderson-Sellers, A.; Zhang, H.; Berz, G.; Emanuel, K.; Gray, W.; Landsea, C.;
Holland, G.; Lighthill, J.; Shieh, S.L.; Webster, P.; McGuffie, K. (1998).
"Tropical Cyclones and Global Climate Change: A Post-IPCC Assessment"
. Bulletin of the American Meteorological Society. 79 (1): 19–38.
Bibcode:1998BAMS...79...19H
. doi:10.1175/1520-0477(1998)079<0019:TCAGCC>2.0.CO;2
. S2CID 9935617
.
Landsea, Chris (July 13, 2005). "Why doesn't the South Atlantic Ocean experience
tropical cyclones?"
. Atlantic Oceanographic and Meteorological Laboratory. National Oceanographic and
Atmospheric Administration. Archived
from the original on June 23, 2012. Retrieved June 9, 2018.
"Glossary of NHC Terms"
. United States National Hurricane Center. Archived
from the original on February 16, 2021. Retrieved February 18, 2021.
"Tropical cyclone facts: What is a tropical cyclone?"
. United Kingdom Met Office. Archived
from the original on February 2, 2021. Retrieved February 25, 2021.
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.
"Tropical cyclone facts: How do tropical cyclones form?"
. United Kingdom Met Office. Archived
from the original on February 2, 2021. Retrieved March 1, 2021.
Marine Meteorology Division. "Cirrus Cloud Detection"
(PDF). Satellite Product Tutorials. Monterey, CA: United States Naval Research
Laboratory. p. 1. Archived
(PDF) from the original on February 16, 2013. Retrieved June 4, 2013.
Frank, W.M. (1977). "The structure and energetics of the tropical cyclone I. Storm
structure"
. Monthly Weather Review. 105 (9): 1119–1135. Bibcode:1977MWRv..105.1119F
. doi:10.1175/1520-0493(1977)105<1119:TSAEOT>2.0.CO;2
.
National Weather Service (October 19, 2005). "Tropical Cyclone Structure"
. JetStream – An Online School for Weather. National Oceanic & Atmospheric
Administration. Archived
from the original on December 7, 2013. Retrieved May 7, 2009.
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. Retrieved December 14, 2006.
Annamalai, H.; Slingo, J.M.; Sperber, K.R.; Hodges, K. (1999). "The Mean Evolution
and Variability of the Asian Summer Monsoon: Comparison of ECMWF and NCEP–NCAR
Reanalyses"
. Monthly Weather Review. 127 (6): 1157–1186. Bibcode:1999MWRv..127.1157A
. doi:10.1175/1520-0493(1999)127<1157:TMEAVO>2.0.CO;2
.
American Meteorological Society. "AMS Glossary: C"
. Glossary of Meteorology. Allen Press. Archived
from the original on January 26, 2011. Retrieved December 14, 2006.
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.
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 5, 2009. Retrieved July 25, 2006.
Ross., Simon (1998). Natural Hazards
(Illustrated ed.). Nelson Thornes. p. 96. ISBN 978-0-7487-3951-6. Retrieved May 7,
2009.
Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division.
"Frequently Asked Questions: How do tropical cyclones form?"
. National Oceanic and Atmospheric Administration. Archived from the original
on August 27, 2009. Retrieved July 26, 2006.
Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division.
"Frequently Asked Questions: Why do tropical cyclones require 80 °F (27 °C) ocean
temperatures to form?"
. National Oceanic and Atmospheric Administration. Archived from the original
on August 23, 2006. Retrieved July 25, 2006.
Ron McTaggart-Cowan; Canada Emily L.; Jonathan G. Fairman Jr.; Thomas J. Galarneau
Jr.; David M. Schultz (2015). "Revisiting the 26.5°C Sea Surface Temperature
Threshold for Tropical Cyclone Development"
. Bulletin of the American Meteorological Society. 96 (11): 1929–1943.
Bibcode:2015BAMS...96.1929M
. doi:10.1175/BAMS-D-13-00254.2
. Archived
from the original on June 13, 2017. Retrieved October 15, 2017.
Kikuchi, Kazuyoshi; Wang, Bin; Fudeyasu, Hironori (2009). "Genesis of tropical
cyclone Nargis revealed by multiple satellite observations"
(PDF). Geophysical Research Letters. 36 (6): L06811. Bibcode:2009GeoRL..36.6811K
. doi:10.1029/2009GL037296
. Archived
(PDF) from the original on December 29, 2010. Retrieved February 21, 2010.
Korek, Fritz (November 21, 2000). "Marine Meteorological Glossary"
. Marine Knowledge Centre. Archived from the original
on December 11, 2008. Retrieved May 6, 2009.
"Formation of Tropical Cyclones"
. Philippine Atmospheric, Geophysical and Astronomical Services Administration.
2008. Archived from the original on September 2, 2012. Retrieved May 6, 2009.
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.
Avila, L.A.; Pasch, R.J. (1995). "Atlantic Tropical Systems of 1993"
. Monthly Weather Review. 123 (3): 887–896. Bibcode:1995MWRv..123..887A
. doi:10.1175/1520-0493(1995)123<0887:ATSO>2.0.CO;2
.
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.
Landsea, C.W. (1993). "A Climatology of Intense (or Major) Atlantic Hurricanes"
. Monthly Weather Review. 121 (6): 1703–1713. Bibcode:1993MWRv..121.1703L
. doi:10.1175/1520-0493(1993)121<1703:ACOIMA>2.0.CO;2
.
Dowdy, A.J.; Qi, L.; Jones, D.; Ramsay, H.; Fawcett, R.; Kuleshov, Y. (2012).
"Tropical Cyclone Climatology of the South Pacific Ocean and Its Relationship to El
Niño–Southern Oscillation"
. Journal of Climate. 25 (18): 6108–6122. Bibcode:2012JCli...25.6108D
. doi:10.1175/JCLI-D-11-00647.1
.
Neumann, Charles J. "Worldwide Tropical Cyclone Tracks 1979–88"
. Global Guide to Tropical Cyclone Forecasting. Bureau of Meteorology. Archived
from the original
on June 2, 2011. Retrieved December 12, 2006.
Henderson-Sellers; et al. (October 8, 2002). "Tropical Cyclones and Global Climate
Change: A Post-IPCC Assessment"
. National Oceanic and Atmospheric Administration. Archived
from the original on June 22, 2012. Retrieved May 7, 2009.
"Monthly Global Tropical Cyclone Summary, December 2001"
. Gary Padgett. Australian Severe Weather Index. Archived
from the original on February 23, 2009. Retrieved May 6, 2009.
"Annual Tropical Cyclone Report 2004"
(PDF). Joint Typhoon Warning Center. 2006. Archived
(PDF) from the original on December 6, 2013. Retrieved May 6, 2009.
Emanuel, Kerry (February 8, 2006). "Anthropogenic Effects on Tropical Cyclone
Activity"
. Massachusetts Institute of Technology. Archived
from the original on March 30, 2009. Retrieved May 7, 2009.
Emanuel, K A. (1986). "An Air-Sea Interaction Theory for Tropical Cyclones. Part
I: Steady-State Maintenance"
. Journal of the Atmospheric Sciences. 43 (6): 585–605. Bibcode:1986JAtS...43..585E
. doi:10.1175/1520-0469(1986)043<0585:AASITF>2.0.CO;2
.
"NOAA FAQ: How much energy does a hurricane release?"
. National Oceanic & Atmospheric Administration. August 2001. Archived
from the original on June 22, 2012. Retrieved June 30, 2009.
"Hurricanes: Keeping an eye on weather's biggest bullies"
. University Corporation for Atmospheric Research. March 31, 2006. Archived from
the original
on April 25, 2009. Retrieved May 7, 2009.
Barnes, Gary. "Hurricanes and the equator"
. University of Hawaii. Archived
from the original on August 5, 2013. Retrieved August 30, 2013.
Bister, M.; Emanuel, K.A. (1998). "Dissipative heating and hurricane intensity".
Meteorology and Atmospheric Physics. 65 (3–4): 233–240. Bibcode:1998MAP....65..233B
. doi:10.1007/BF01030791
. S2CID 123337988
.
Emanuel, K. (2000). "A Statistical Analysis of Tropical Cyclone Intensity"
. Monthly Weather Review. 128 (4): 1139–1152. Bibcode:2000MWRv..128.1139E
. doi:10.1175/1520-0493(2000)128<1139:ASAOTC>2.0.CO;2
.
Knutson, T.R.; McBride, J.L.; Chan, J.; Emanuel, K.; Holland, G.; Landsea, C.;
Held, I.; Kossin, J.P.; Srivastava, A.K.; Sugi, M. (2010). "Tropical cyclones and
climate change". Nature Geoscience. 3 (3): 157–163. Bibcode:2010NatGe...3..157K
. doi:10.1038/ngeo779
. hdl:11343/192963
.
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.
Fedorov, Alexey V.; Brierley, Christopher M.; Emanuel, Kerry (February 2010).
"Tropical cyclones and permanent El Niño in the early Pliocene epoch". Nature. 463
(7284): 1066–1070. Bibcode:2010Natur.463.1066F
. doi:10.1038/nature08831
. hdl:1721.1/63099
. ISSN 0028-0836
. PMID 20182509
. S2CID 4330367
.
Holland, G.J. (1983). "Tropical Cyclone Motion: Environmental Interaction Plus a
Beta Effect"
. Journal of the Atmospheric Sciences. 40 (2): 328–342. Bibcode:1983JAtS...40..328H
. doi:10.1175/1520-0469(1983)040<0328:TCMEIP>2.0.CO;2
. S2CID 124178238
.
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 June 23, 2012. Retrieved July 25, 2006.
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.
"Fujiwhara effect describes a stormy waltz"
. USA Today. November 9, 2007. Archived
from the original on November 5, 2012. Retrieved February 21, 2008.
"Section 2: Tropical Cyclone Motion Terminology"
. United States Naval Research Laboratory. April 10, 2007. Archived
from the original on June 23, 2012. Retrieved May 7, 2009.
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.
National Hurricane Center (2016). "Glossary of NHC/TPC Terms"
. United States National Oceanic and Atmospheric Administration. Archived
from the original on June 1, 2014. Retrieved April 30, 2016.
RA IV Hurricane Committee. Regional Association IV Hurricane Operational Plan 2019
(PDF) (Report). World Meteorological Organization. Retrieved July 2, 2019.
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 September 4, 2015. Retrieved March 28, 2015.
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. Retrieved
July 2, 2019.
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.
RA V Tropical Cyclone Committee (October 8, 2020). Tropical Cyclone Operational
Plan for the South-East Indian Ocean and the Southern Pacific Ocean 2020
(PDF) (Report). World Meteorological Organization. pp. I-4–II-9 (9–21). Retrieved
October 10, 2020.
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.
Retrieved October 14, 2018.
United States Naval Research Laboratory, Marine Meteorology Division (June 8,
2010). "Best Track/Objective Aid/Wind Radii Format"
. Monterey: United States Navy. Retrieved October 15, 2018.
"Tropical Cyclone Names"
. Met Office (United Kingdom Meteorological Office). Retrieved October 17, 2018.
"RSMC Tokyo - Typhoon Center"
. Japan Meteorological Agency. Retrieved October 19, 2018.
"過去の台風資料"
(in Japanese). Japan Meteorological Agency. Retrieved October 19, 2018.
"Saisons cycloniques archivées"
(in French). Météo-France La Réunion.
"Monthly Global Tropical Cyclone Summary March 2004"
. Australia Severe Weather.
"Rare Tropical Cyclone Forms Off Brazil"
. EarthWeek. Retrieved October 18, 2018.
"Observed and forecast tracks: southern hemisphere 2016-17"
. Met Office (United Kingdom Meteorological Office). Retrieved October 17, 2018.
Smith, Ray (1990). "What's in a Name?"
(PDF). Weather and Climate. The Meteorological Society of New Zealand. 10 (1): 24–
26. doi:10.2307/44279572
. JSTOR 44279572
. S2CID 201717866
. Archived from the original
(PDF) on November 29, 2014. Retrieved August 25, 2014.
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.
"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.
Roth, David & Cobb, Hugh (2001). "Eighteenth Century Virginia Hurricanes"
. NOAA. Archived
from the original on May 1, 2013. Retrieved February 24, 2007.
Shultz, J.M.; Russell, J.; Espinel, Z. (2005). "Epidemiology of Tropical Cyclones:
The Dynamics of Disaster, Disease, and Development"
. Epidemiologic Reviews. 27: 21–35. doi:10.1093/epirev/mxi011
. PMID 15958424
.
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.
Staff Writer (August 30, 2005). "Hurricane Katrina Situation Report #11"
(PDF). Office of Electricity Delivery and Energy Reliability (OE) United States
Department of Energy. Archived from the original
(PDF) on November 8, 2006. Retrieved February 24, 2007.
Burroughs, William James (2007). Climate change : a multidisciplinary approach
(2nd ed.). Cambridge: Cambridge University Press. ISBN 978-0-521-87015-3.
National Oceanic and Atmospheric Administration. 2005 Tropical Eastern North
Pacific Hurricane Outlook.
Archived
May 28, 2015, at WebCite. Retrieved May 2, 2006.
National Weather Service (October 19, 2005). "Tropical Cyclone Introduction"
. JetStream – An Online School for Weather. National Oceanic & Atmospheric
Administration. Archived
from the original on June 22, 2012. Retrieved September 7, 2010.
Emanuel, Kerry (July 2001). "Contribution of tropical cyclones to meridional heat
transport by the oceans"
. Journal of Geophysical Research. 106 (D14): 14771–14781.
Bibcode:2001JGR...10614771E
. doi:10.1029/2000JD900641
.
Christopherson, Robert W. (1992). Geosystems: An Introduction to Physical
Geography. New York: Macmillan Publishing Company. pp. 222–224. ISBN 978-0-02-
322443-0.
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.
Cappielo, Dina (2005). "Spills from hurricanes stain coast With gallery"
. Houston Chronicle. Archived
from the original on April 25, 2014. Retrieved February 12, 2014.
"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.
"Before You Begin – The Incident Command System (ICS)"
. American Industrial Hygiene Association. Archived from the original
on September 29, 2018. Retrieved September 26, 2018.
"Volunteer"
. National Voluntary Organizations Active in Disaster. Archived
from the original on September 29, 2018. Retrieved September 25, 2018.
"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.
"Hazardous Materials and Conditions"
. American Industrial Hygiene Association. Archived from the original
on September 29, 2018. Retrieved September 26, 2018.
"Mold and Other Microbial Growth"
. American Industrial Hygiene Association. Archived from the original
on September 29, 2018. Retrieved September 26, 2018.
"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.
"Electrical Hazards"
. American Industrial Hygiene Association. Archived from the original
on September 29, 2018. Retrieved September 26, 2018.
"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.
"Global Guide to Tropical Cyclone Forecasting: chapter 2: Tropical Cyclone
Structure"
. Bureau of Meteorology. May 7, 2009. Archived from the original
on June 1, 2011. Retrieved May 6, 2009.
Chavas, D.R.; Emanuel, K.A. (2010). "A QuikSCAT climatology of tropical cyclone
size". Geophysical Research Letters. 37 (18): n/a. Bibcode:2010GeoRL..3718816C
. doi:10.1029/2010GL044558
. hdl:1721.1/64407
.
Merrill, Robert T (1984). "A comparison of Large and Small Tropical cyclones"
. Monthly Weather Review. 112 (7): 1408–1418. Bibcode:1984MWRv..112.1408M
. doi:10.1175/1520-0493(1984)112<1408:ACOLAS>2.0.CO;2
. S2CID 123276607
.
"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.
Diana Engle. "Hurricane Structure and Energetics"
. Data Discovery Hurricane Science Center. Archived from the original
on May 27, 2008. Retrieved October 26, 2008.
Brad Reinhart; Daniel Brown (October 21, 2020). "Hurricane Epsilon Discussion
Number 12"
. nhc.noaa.gov. Miami, Florida: National Hurricane Center. Retrieved February 4,
2021.
Cappucci, Matthew (October 21, 2020). "Epsilon shatters records as it rapidly
intensifies into major hurricane near Bermuda"
. The Washington Post. Retrieved February 4, 2021.
"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.
Chang, Chih-Pei (2004). East Asian Monsoon
. World Scientific. ISBN 978-981-238-769-1. OCLC 61353183
.
United States Naval Research Laboratory (September 23, 1999). "Tropical Cyclone
Intensity Terminology"
. Tropical Cyclone Forecasters' Reference Guide. Archived
from the original on June 23, 2012. Retrieved November 30, 2006.
"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.
McAdie, Colin (May 10, 2007). "Tropical Cyclone Climatology"
. National Hurricane Center. Archived
from the original on March 21, 2015. Retrieved June 9, 2007.
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.
http://www.rsmcnewdelhi.imd.gov.in/images/pdf/publications/annual-rsmc-report/rsmc-
2018.pdf
"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.
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.
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.
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.
Liu, Kam-biu; Fearn, Miriam L. (2000). "Reconstruction of Prehistoric Landfall
Frequencies of Catastrophic Hurricanes in Northwestern Florida from Lake Sediment
Records". Quaternary Research. 54 (2): 238–245. Bibcode:2000QuRes..54..238L
. doi:10.1006/qres.2000.2166
.
Elsner, James B.; Liu, Kam-biu; Kocher, Bethany (2000). "Spatial Variations in
Major U.S. Hurricane Activity: Statistics and a Physical Mechanism"
. Journal of Climate. 13 (13): 2293–2305. Bibcode:2000JCli...13.2293E
. doi:10.1175/1520-0442(2000)013<2293:SVIMUS>2.0.CO;2
.
Ramsay, Hamish (2017). "The Global Climatology of Tropical Cyclones"
. Oxford Research Encyclopedia of Natural Hazard Science.
doi:10.1093/acrefore/9780199389407.001.0001/acrefore-9780199389407-e-79#acrefore-
9780199389407-e-79-div1-4
.
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.
Wu, M.C.; Chang, W.L.; Leung, W.M. (2004). "Impacts of El Niño–Southern
Oscillation Events on Tropical Cyclone Landfalling Activity in the Western North
Pacific". Journal of Climate. 17 (6): 1419–1428. Bibcode:2004JCli...17.1419W
. CiteSeerX 10.1.1.461.2391
. doi:10.1175/1520-0442(2004)017<1419:IOENOE>2.0.CO;2
.
Pacific ENSO Applications Climate Center. "Pacific ENSO Update: 4th Quarter, 2006.
Vol. 12 No. 4"
. Archived
from the original on June 24, 2012. Retrieved March 19, 2008.
Rappaport, Edward N. (1999). "Atlantic Hurricane Season of 1997"
(PDF). Monthly Weather Review. 127 (9): 2012–2026. Bibcode:1999MWRv..127.2012R
. doi:10.1175/1520-0493(1999)127<2012:AHSO>2.0.CO;2
. Archived
(PDF) from the original on December 7, 2013. Retrieved July 18, 2009.
Risk Management Solutions (March 2006). "U.S. and Caribbean Hurricane Activity
Rates"
(PDF). Archived from the original
(PDF) on June 14, 2007. Retrieved November 30, 2006.
Center for Climate Systems Research. "Hurricanes, Sea Level Rise, and New York
City"
. Columbia University. Archived from the original
on January 2, 2007. Retrieved November 29, 2006.
Higuera-Gundy, Antonia; Brenner, Mark; Hodell, David A.; Curtis, Jason H.; Leyden,
Barbara W.; Binford, Michael W. (1999). "A 10,300 14C yr Record of Climate and
Vegetation Change from Haiti". Quaternary Research. 52 (2): 159–170.
Bibcode:1999QuRes..52..159H
. doi:10.1006/qres.1999.2062
.
Knutson, Thomas; Camargo, Suzana J.; Chan, Johnny C. L.; Emanuel, Kerry; Ho,
Chang-Hoi; Kossin, James; Mohapatra, Mrutyunjay; Satoh, Masaki; Sugi, Masato;
Walsh, Kevin; Wu, Liguang (August 6, 2019). "Tropical Cyclones and Climate Change
Assessment: Part II. Projected Response to Anthropogenic Warming"
. Bulletin of the American Meteorological Society. 101 (3): BAMS–D–18–0194.1.
doi:10.1175/BAMS-D-18-0194.1
. ISSN 0003-0007
.
"Major tropical cyclones have become '15% more likely' over past 40 years"
. Carbon Brief. May 18, 2020. Retrieved August 31, 2020.
Kossin, James P.; Knapp, Kenneth R.; Olander, Timothy L.; Velden, Christopher S.
(May 18, 2020). "Global increase in major tropical cyclone exceedance probability
over the past four decades"
(PDF). Proceedings of the National Academy of Sciences. 117 (22): 11975–11980.
doi:10.1073/pnas.1920849117
. ISSN 0027-8424
. PMC 7275711
. PMID 32424081
.
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.
Walsh, K. J. E.; Camargo, S. J.; Knutson, T. R.; Kossin, J.; Lee, T. -C.;
Murakami, H.; Patricola, C. (December 1, 2019). "Tropical cyclones and climate
change"
. Tropical Cyclone Research and Review. 8 (4): 240–250.
doi:10.1016/j.tcrr.2020.01.004
. ISSN 2225-6032
.
Roberts, Malcolm John; Camp, Joanne; Seddon, Jon; Vidale, Pier Luigi; Hodges,
Kevin; Vannière, Benoît; Mecking, Jenny; Haarsma, Rein; Bellucci, Alessio;
Scoccimarro, Enrico; Caron, Louis-Philippe (2020). "Projected Future Changes in
Tropical Cyclones Using the CMIP6 HighResMIP Multimodel Ensemble"
. Geophysical Research Letters. 47 (14): e2020GL088662. Bibcode:2020GeoRL..4788662R
. doi:10.1029/2020GL088662
. ISSN 1944-8007
. PMC 7507130
. PMID 32999514
. S2CID 221972087
.
"Hurricanes and Climate Change"
. Union of Concerned Scientists. Retrieved September 29, 2019.
Murakami, Hiroyuki; Delworth, Thomas L.; Cooke, William F.; Zhao, Ming; Xiang,
Baoqiang; Hsu, Pang-Chi (2020). "Detected climatic change in global distribution of
tropical cyclones"
. Proceedings of the National Academy of Sciences. 117 (20): 10706–10714.
doi:10.1073/pnas.1922500117
. ISSN 0027-8424
. PMID 32366651
.
James P. Kossin; Kerry A. Emanuel; Gabriel A. Vecchi (2014). "The poleward
migration of the location of tropical cyclone maximum intensity". Nature. 509
(7500): 349–352. Bibcode:2014Natur.509..349K
. doi:10.1038/nature13278
. hdl:1721.1/91576
. PMID 24828193
. S2CID 4463311
.
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.
Thomas R. Knutson; Joseph J. Sirutis; Ming Zhao (2015). "Global Projections of
Intense Tropical Cyclone Activity for the Late Twenty-First Century from Dynamical
Downscaling of CMIP5/RCP4.5 Scenarios"
. Journal of Climate. 28 (18): 7203–7224. Bibcode:2015JCli...28.7203K
. doi:10.1175/JCLI-D-15-0129.1
.
Knutson; et al. (2013). "Dynamical Downscaling Projections of Late 21st Century
Atlantic Hurricane Activity: CMIP3 and CMIP5 Model-based Scenarios"
. Journal of Climate. 26 (17): 6591–6617. Bibcode:2013JCli...26.6591K
. doi:10.1175/JCLI-D-12-00539.1
.
"Hurricane Harvey shows how we underestimate flooding risks in coastal cities,
scientists say"
. The Washington Post. August 29, 2017.
Florida Coastal Monitoring Program. "Project Overview"
. University of Florida. Archived from the original
on May 3, 2006. Retrieved March 30, 2006.
"Observations"
. Central Pacific Hurricane Center. December 9, 2006. Archived
from the original on June 24, 2012. Retrieved May 7, 2009.
403rd Wing. "The Hurricane Hunters"
. 53rd Weather Reconnaissance Squadron. Archived
from the original on June 24, 2012. Retrieved March 30, 2006.
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.
"Influences on Tropical Cyclone Motion"
. United States Navy. Archived
from the original on June 24, 2012. Retrieved April 10, 2007.
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 June 24, 2012. Retrieved November 30, 2006.
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 June 24, 2012. Retrieved November 30, 2006.
"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.
Fiji Meteorological Service (2017). "Services"
. Archived
from the original on June 18, 2017. Retrieved June 4, 2017.
Joint Typhoon Warning Center (2017). "Products and Service Notice"
. United States Navy. Archived
from the original on June 9, 2017. Retrieved June 4, 2017.
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.
Japan Meteorological Agency (2017). "Notes on RSMC Tropical Cyclone Information"
. Archived
from the original on March 19, 2017. Retrieved June 4, 2017.
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.
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.
"Lesson 14: Background: Synoptic Scale"
. University of Wisconsin–Madison. February 25, 2008. Archived
from the original on February 20, 2009. Retrieved May 6, 2009.
"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.
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.
"New Mortality Records Announced"
(Press release). World Meteorological Organization. 2017. Archived
from the original on June 26, 2018. Retrieved June 25, 2018.
Landsea, Chris (1993). "Which tropical cyclones have caused the most deaths and
most damage?"
. Hurricane Research Division. Archived
from the original on June 24, 2012. Retrieved February 23, 2007.
Frank, N.L.; Husain, S.A. (1971). "The Deadliest Tropical Cyclone in History"
. Bulletin of the American Meteorological Society. 52 (6): 438–445.
Bibcode:1971BAMS...52..438F
. doi:10.1175/1520-0477(1971)052<0438:TDTCIH>2.0.CO;2
.
Anderson-Berry, Linda J. Fifth International Workshop on Tropycal Cyclones: Topic
5.1: Societal Impacts of Tropical Cyclones.
Archived
September 22, 2012, at WebCite. Retrieved February 26, 2008.
National Hurricane Center (April 22, 1997). "The Deadliest Atlantic Tropical
Cyclones, 1492–1996"
. National Oceanic and Atmospheric Administration. Archived
from the original on January 28, 2008. Retrieved March 31, 2006.
Joint Typhoon Warning Center. "Typhoon Thelma (27W)"
(PDF). 1991 Annual Tropical Cyclone Report. Archived
(PDF) from the original on December 6, 2013. Retrieved March 31, 2006.
Gunther, E.B.; Cross, R.L.; Wagoner, R.A. (1983). "Eastern North Pacific Tropical
Cyclones of 1982"
. Monthly Weather Review. 111 (5): 1080–1102. Bibcode:1983MWRv..111.1080G
. doi:10.1175/1520-0493(1983)111<1080:ENPTCO>2.0.CO;2
.
Costliest U.S. tropical cyclones tables update
(PDF) (Report). United States National Hurricane Center. January 12, 2018.
Archived
(PDF) from the original on January 26, 2018. Retrieved January 12, 2018.
"Hurricane Damages Sour to New Levels"
. Earth Policy Institute. 2006. Archived from the original
on December 13, 2006. Retrieved February 23, 2007.
Knabb, Richard D.; Rhome, Jamie R.; Brown, Daniel P. (December 20, 2005).
"Tropical Cyclone Report: Hurricane Katrina: 23–30 August 2005"
(PDF). National Hurricane Center. Retrieved May 30, 2006.
National Hurricane Center. Galveston Hurricane 1900.
Archived
July 9, 2006, at the Wayback Machine. Retrieved February 24, 2008.
Central Pacific Hurricane Center. "Hurricane Iniki Natural Disaster Survey Report"
. National Oceanic and Atmospheric Administration. Archived
from the original on July 16, 2015. Retrieved March 31, 2006.
Lawrence, Miles B. (November 7, 1997). "Preliminary Report: Hurricane Pauline: 5–
10 October 1997"
. National Hurricane Center. Archived from the original
on September 25, 2005. Retrieved March 31, 2006.
Franklin, James L. (December 26, 2002). "Tropical Cyclone Report: Hurricane Kenna:
22–26 October 2002"
(PDF). National Hurricane Center. Archived
from the original on July 16, 2014. Retrieved March 31, 2006.
World Food Programme (2004). "WFP Assists Cyclone And Flood Victims in Madagascar"
. Archived from the original
on February 14, 2009. Retrieved February 24, 2007.
Dunnavan, G.M.; Diercks, J.W. (1980). "An Analysis of Super Typhoon Tip (October
1979)"
. Monthly Weather Review. 108 (11): 1915–1923. Bibcode:1980MWRv..108.1915D
. doi:10.1175/1520-0493(1980)108<1915:AAOSTT>2.0.CO;2
.
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.
Atlantic Oceanographic and Meteorological Laboratory, Hurricane Research Division.
"Frequently Asked Questions: Which is the most intense tropical cyclone on record?"
. NOAA. Archived from the original
on December 6, 2010. Retrieved July 25, 2006.
Houston, Sam; Greg Forbes; Arthur Chiu (August 17, 1998). "Super Typhoon Paka's
(1997) Surface Winds Over Guam"
. National Weather Service. Archived
from the original on November 5, 2015. Retrieved March 30, 2006.
World Record Wind Gust: 408 km/h
Archived
January 20, 2013, at the Wayback Machine. World Meteorological Organization.
Courtney et. al. 2012
, Documentation and verification of the world extreme wind gust record: 113.3 m/s
on Barrow Island, Australia, during passage of tropical cyclone Olivia, AMOJ 62,
p1-9.
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.
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 19, 2009. Retrieved June 12, 2013.
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 19, 2009. Retrieved June 12, 2013.
Dorst, Neal; Hurricane Research Division (June 1, 2013). "Subject: J4) What
fictional books, plays, poems, and movies have been written involving tropical
cyclones?"
. Tropical Cyclone Frequently Asked Questions. National Oceanic and Atmospheric
Administration. Retrieved March 30, 2013.
McCown, Sean (December 13, 2004). "Unnamed Hurricane 1991"
. Satellite Events Art Gallery: Hurricanes. National Climatic Data Center. Archived
from the original on December 7, 2013. Retrieved February 4, 2007.
"Hurricane Neddy – Episode Overview"
. Yahoo! TV. Archived from the original
on May 6, 2009. Retrieved February 26, 2008.
"Family Guy: One if by Clam, Two if by Sea – Summary"
. starpulse.com. Archived from the original
on May 6, 2009. Retrieved February 26, 2008.
"Dawson's Creek – Hurricane"
. Yahoo! TV. Archived from the original
on May 6, 2009. Retrieved February 25, 2008.
"CSI: Miami Episodes – Episode Detail: Hurricane Anthony"
. TV Guide. Archived from the original
on May 6, 2009. Retrieved February 25, 2008.
"The Day After Tomorrow Movie Synopsis"
. Tribute.ca. Archived from the original
on May 6, 2009. Retrieved February 26, 2008.
"The Day After Tomorrow (2004)"
. The New York Times. Archived
from the original on March 10, 2011. Retrieved February 26, 2008.
External links
Look up tropical cyclone in Wiktionary, the free dictionary.
Wikimedia Commons has media related to Tropical cyclones.
Wikisource has original text related to this article:
The Hurricane
Wikivoyage has a travel guide for Cyclones.
Warning centers
US National Hurricane Center
– North Atlantic, Eastern Pacific
Central Pacific Hurricane Center
– Central Pacific
Japan Meteorological Agency
– Western Pacific
India Meteorological Department
– Indian Ocean
Météo-France – La Reunion
– South Indian Ocean from 30°E to 90°E
Indonesian Meteorological Department
– South Indian Ocean from 90°E to 125°E, north of 10°S
Australian Bureau of Meteorology
– South Indian Ocean and South Pacific Ocean from 90°E to 160°E
Papua New Guinea National Weather Service
– South Pacific east of 160°E, north of 10°S
Fiji Meteorological Service
– South Pacific west of 160°E, north of 25° S
Meteorological Service of New Zealand Limited
– South Pacific west of 160°E, south of 25°S
vte
Cyclones and anticyclones of the world (Centers of action)
vte
Seasons
vte
Elements of nature
Authority control Edit this at Wikidata
GND: 4186319-7
LCCN: sh85138066
MA: 29141058
, 2780210877
NARA: 10638964
NDL: 00575300
Categories: Tropical cyclonesTropical cyclone meteorologyClimate change and
hurricanesMeteorological phenomenaTypes of cycloneVorticesWeather hazardsStorm
Navigation menu
Not logged in
Talk
Contributions
Create account
Log in
ArticleTalk
ReadView sourceView historySearch
Search Wikipedia
Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
Contribute
Help
Learn to edit
Community portal
Recent changes
Upload file
Tools
What links here
Related changes
Special pages
Permanent link
Page information
Cite this page
Wikidata item
Print/export
Download as PDF
Printable version
In other projects
Wikimedia Commons
Wikinews

Languages
Deutsch
Español
Français
한국어
Italiano
Русский
Tagalog
Tiếng Việt
中文
107 more
Edit links
This page was last edited on 9 March 2021, at 14:10 (UTC).
Text is available under the Creative Commons Attribution-ShareAlike License;
additional terms may apply. By using this site, you agree to the Terms of Use and
Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation,
Inc., a non-profit organization.
Privacy policyAbout WikipediaDisclaimersContact WikipediaMobile
viewDevelopersStatisticsCookie statementWikimedia FoundationPowered by MediaWiki

You might also like