STATE OF THE CLIMATE IN 2019
THE TROPICS
H. J. Diamond and C. J. Schreck, Eds.
Special Online Supplement to the Bulletin of the American Meteorological Society, Vol.101, No. 8, August, 2020
https://doi.org/10.1175/BAMS-D-20-0077.1
Corresponding author: Howard J. Diamond / howard.diamond@noaa.gov
©2020 American Meteorological Society
For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
S185
Unauthenticated | Downloaded 12/04/22 09:14
AM UTC
4. THE TROPICS
STATE OF THE CLIMATE IN 2019
The Tropics
Editors
Jessica Blunden
Derek S. Arndt
Chapter Editors
Peter Bissolli
Howard J. Diamond
Matthew L. Druckenmiller
Robert J. H. Dunn
Catherine Ganter
Nadine Gobron
Rick Lumpkin
Jacqueline A. Richter-Menge
Tim Li
Ademe Mekonnen
Ahira Sánchez-Lugo
Ted A. Scambos
Carl J. Schreck III
Sharon Stammerjohn
Diane M. Stanitski
Kate M. Willett
Technical Editor
Andrea Andersen
BAMS Special Editor for Climate
Richard Rosen
American Meteorological Society
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S186
AM UTC
4. THE TROPICS
Cover credit:
Catastrophic Hurricane Dorian slowed to a crawl over Grand Bahama Island overnight and into Labor Day. On
Monday, September 2, 2019, GOES East captured a view of the Category 5 storm over Grand Bahama.
This GeoColor-enhanced imagery was created by NOAA's partners at the Cooperative Institute for Research in the
Atmosphere. The GOES East geostationary satellite, also known as GOES-16, provides coverage of the Western
Hemisphere, including the United States, the Atlantic Ocean and the eastern Pacific. The satellite's high-resolution
imagery provides optimal viewing of severe weather events, including thunderstorms, tropical storms, and hurricanes.
© NOAA
The Tropics is one chapter from the State of the Climate in 2019 annual report and is available
from https://doi.org/10.1175/BAMS-D-20-0077.1. Compiled by NOAA’s National Centers for
Environmental Information, State of the Climate in 2019 is based on contributions from
scientists from around the world. It provides a detailed update on global climate indicators,
notable weather events, and other data collected by environmental monitoring stations and
instruments located on land, water, ice, and in space. The full report is available from
https://doi.org/10.1175/2020BAMSStateoftheClimate.1.
How to cite this document:
Citing the complete report:
Blunden, J. and D. S. Arndt, Eds., 2020: State of the Climate in 2019. Bull. Amer. Meteor. Soc.,
101 (8), Si–S429, https://doi.org/10.1175/2020BAMSStateoftheClimate.1.
Citing this chapter:
Diamond, H.J. and C. J. Schreck, Eds., 2020: The Tropics [in “State of the Climate in 2019”].
Bull. Amer. Meteor. Soc., 101 (8), S185–S238, https://doi.org/10.1175/BAMS-D-20-0077.1.
Citing a section (example):
Chen, L., J. -J Luo, and A. D. Magee, 2020: Indian Ocean dipole [in “State of the Climate in
2019"]. Bull. Amer. Meteor. Soc., 101 (8), S229–S232, https://doi.org/10.1175/BAMS-D-20-0077.1.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
4 . T H E T12/04/22
R O P I C S09:14S187
Unauthenticated | Downloaded
AM UTC
Editor and Author Affiliations (alphabetical by name)
Baxter, Stephen, NOAA/NWS Climate Prediction Center, College Park,
Maryland
Bell, Gerald D., NOAA/NWS Climate Prediction Center, College Park, Maryland
Blake, Eric S., NOAA/NWS National Hurricane Center, Miami, Florida
Bringas, Francis G., NOAA/OAR Atlantic Oceanographic and Meteorological
Laboratory, Miami, Florida
Camargo, Suzana J., Lamont-Doherty Earth Observatory, Columbia University,
Palisades, New York
Chen, Lin, Institute for Climate and Application Research (ICAR)/KLME/ILCEC/
CIC-FEMD, Nanjing University of Information Science and Technology,
Nanjing, China
Coelho, Caio A. S., CPTEC/INPE Center for Weather Forecasts and Climate
Studies, Cachoeira Paulista, Brazil
Diamond, Howard J., NOAA/OAR Air Resources Laboratory, College Park,
Maryland
Domingues, Ricardo, Cooperative Institute for Marine and Atmospheric
Studies, University of Miami, Miami, Florida
Goldenberg, Stanley B., NOAA/OAR/AOML Hurricane Research Division,
Miami, Florida
Goni, Gustavo, NOAA/OAR/AOML Hurricane Research Division, Miami, Florida
Fauchereau, Nicolas, National Institute of Water and Atmospheric Research,
Ltd., Auckland, New Zealand
Halpert, Michael S., NOAA/NWS Climate Prediction Center, College Park,
Maryland
He, Qiong, Earth System Modeling Center, Nanjing University of Information
Science and Technology, Nanjing, China
Klotzbach, Philip J., Department of Atmospheric Science, Colorado State
University, Fort Collins, Colorado
Knaff, John A., NOAA/NESDIS Center for Satellite Applications and Research,
Fort Collins, Colorado
L'Heureux, Michelle, NOAA/NWS Climate Prediction Center, College Park,
Maryland
Landsea, Chris W., NOAA/NWS National Hurricane Center, Miami, Florida
Lin, I.-I., National Taiwan University, Taipei, Taiwan
Lorrey, Andrew M., National Institute of Water and Atmospheric Research,
Ltd., Auckland, New Zealand
Luo, Jing-Jia, Institute for Climate and Application Research (ICAR)/KLME/
ILCEC/CIC-FEMD, Nanjing University of Information Science and Technology,
Nanjing, China
Magee, Andrew D., Centre for Water, Climate and Land, School of
Environmental and Life Sciences, University of Newcastle, Callaghan, NSW,
Austrailia
Pasch, Richard J., NOAA/NWS National Hurricane Center, Miami, Florida
Pearce, Petra R., National Institute of Water and Atmospheric Research, Ltd.,
Auckland, New Zealand
Pezza, Alexandre B., Greater Wellington Regional Council, Wellington, New
Zealand
Rosencrans, Matthew, NOAA/NWS Climate Prediction Center, College Park,
Maryland
Schreck III, Carl J., North Carolina State University, Cooperative Institute for
Climate and Satellites – North Carolina (CICS-NC), Asheville, North Carolina
Trewin, Blair C., Australian Bureau of Meteorology, Melbourne, Victoria,
Australia
Truchelut, Ryan E., “WeatherTiger,” Tallahassee, Florida
Wang, Bin, Department of Atmospheric Science and IPRC, University of Hawaii,
Honolulu, Hawaii
Wang, H., NOAA/NWS Climate Prediction Center, College Park, Maryland
Wood, Kimberly M., Department of Geosciences, Mississippi State University,
Starkville, Mississippi
Woolley, John-Mark, National Institute of Water and Atmospheric Research,
Ltd., Auckland, New Zealand
Editoral and Production Team
Andersen, Andrea, Technical Editor, Innovative Consulting and Management
Services, LLC, NOAA/NESDIS National Centers for Environmental
Information, Asheville, North Carolina
Griffin, Jessicca, Graphics Support, Cooperative Institute for Satellite Earth
System Studies, North Carolina State University, Asheville, North Carolina
Hammer, Gregory, Content Team Lead, Communications and Outreach, NOAA/
NESDIS National Centers for Environmental Information, Asheville, North
Carolina
Love-Brotak, S. Elizabeth, Lead Graphics Production, NOAA/NESDIS National
Centers for Environmental Information, Asheville,
North Carolina
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Misch, Deborah J., Graphics Support, Innovative Consulting and Management
Services, LLC, NOAA/NESDIS National Centers for Environmental
Information, Asheville, North Carolina
Riddle, Deborah B., Graphics Support, NOAA/NESDIS National Centers for
Environmental Information, Asheville, North Carolina
Veasey, Sara W., Visual Communications Team Lead, Communications and
Outreach, NOAA/NESDIS National Centers for Environmental Information,
Asheville, North Carolina
Unauthenticated | Downloaded 12/04/22 09:14S188
AM UTC
4. THE TROPICS
4. Table of Contents
List of authors and affiliations ..................................................................................................S188
a. Overview ................................................................................................................................S190
b. ENSO and the tropical Pacific ................................................................................................ S191
1. Oceanic conditions ..................................................................................................... S192
2. Atmospheric circulation, temperature, and precipitation
anomalies during December–February 2018/19....................................................... S193
3. Atmospheric circulation, temperature, and precipitation anomalies during
March–May through September–November 2019 ..................................................S194
c. Tropical intraseasonal activity ............................................................................................... S195
d. Intertropical convergence zones ...........................................................................................S198
1. Pacific ..........................................................................................................................S198
2. Atlantic ....................................................................................................................... S199
e. Global monsoon summary .....................................................................................................S200
f. Tropical cyclones......................................................................................................................S203
1. Overview ....................................................................................................................S203
2. Atlantic basin .............................................................................................................S204
Sidebar 4.1: Hurricane Dorian: A devastating hurricane
for the northwest Bahamas ...........................................................................................S210
3. Eastern North Pacific and Central North Pacific basins ........................................... S212
4. Western North Pacific basin ......................................................................................S214
5. North Indian Ocean basin .........................................................................................S219
6. South Indian Ocean basin .........................................................................................S221
7. Australian basin..........................................................................................................S223
8. Southwest Pacific basin .............................................................................................S225
g. Tropical cyclone heat potential .............................................................................................S227
h. Indian Ocean dipole ...............................................................................................................S229
Appendix: Acronym List... ..........................................................................................................S233
References... ................................................................................................................................S235
*Please refer to Chapter 8 (Relevant datasets and sources) for a list of all climate variables and
datasets used in this chapter for analyses, along with their websites for more information and
access to the data.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S189
AM UTC
4. THE TROPICS
4. THE TROPICS
H. J. Diamond and C. J. Schreck, Eds.
a. Overview— H. J. Diamond and C. J. Schreck
The tropics in 2019 featured a weak El Niño event that began in January and ended in July.
Neutral ENSO conditions prevailed for the remainder of the year, although sea surface temperatures (SSTs) remained above normal in the central Pacific. The Oceanic Niño Index (ONI) met the
+0.5°C threshold for El Niño during September–December 2018 and November–December 2019.
However, the ocean–atmosphere coupling, normally an intrinsic aspect of El Niño, was missing
during both periods.
For the global tropics, combined land and ocean surface temperatures (measured 20°N–20°S)
registered +0.47°C above the 1981–2010 average. This makes 2019 the third-warmest year for the
tropics since records began in 1880, and the warmest since 2016. Data from the Global Precipitation Climatology Project indicate a mean annual total precipitation value of 1317 mm across the
20°N–20°S latitude band over land. This is 11 mm above the 1981–2010 average and ranks in the
middle tercile of the 1979–2019 period of record.
Globally, 96 named tropical cyclones (TCs; ≥34 kt; or 17 m s−1) were observed during the 2019
Northern Hemisphere (NH) season (January–December 2019) and the 2018/19 Southern Hemisphere (SH) season (July–June 2018/19; Table 4.2), as documented in IBTrACSv4 (Knapp et al.
2010). Overall, this number was well above the 1981–2010 global average of 82 TCs and similar to
the 95 TCs reported during 2018 (Diamond and Schreck 2019).
In terms of Accumulated Cyclone Energy (ACE; Bell et al. 2000), each NH basin was above its
1981–2010 average. The North and South Indian Ocean basins were in the top 10% of ACE recorded
for those basins at 85 × 104 kt2 and 154 × 104 kt2, respectively; and in fact, the ACE value in the
North Indian Ocean was the highest on record. In the western North Pacific, seven storms (six of
Category 5 intensity) out of a total of 28 accounted for 71% of the above-average seasonal ACE of
341 × 104 kt2. The North Atlantic basin had an ACE of nearly 145% of its 1981–2010 median value
but was well below the 241% of median recorded in 2017 (Bell et al. 2018). Category 5 Hurricanes
Dorian and Lorenzo alone accounted for >60% of the 2019 total. The Australian and southwest
Pacific basins were fairly quiet; each had an ACE that was below normal but still within the
middle tercile. The global total was near normal for 1981–2010 with 795 × 104 kt2. Five TCs across
the globe reached Saffir–Simpson Hurricane Wind Scale (SSHWS) Category 5 intensity level—two
in the North Atlantic and three in the western North Pacific.
From a socio-economic standpoint, the five Category 5 storms were significant in their effects.
Hurricane Dorian caused unprecedented and tremendous devastation, with over 70 fatalities and
damages totaling $3.4 billion (U.S. dollars). Hurricane Lorenzo as a post-tropical/extratropical
cyclone was the second-deadliest storm of the 2019 North Atlantic season, causing 19 deaths.
However, major impacts are not relegated to Category 5 storms, and Super Typhoon Faxai demonstrated that with total damages estimated at $9.3 billion (U.S. dollars). Faxai was one of the
strongest typhoons on record to affect Tokyo, Japan, killing three people and injuring 147, causing
extensive blackouts, and damaging more than 40 000 homes
The Indian Ocean dipole (IOD), an inherent air–sea coupling mode in the tropical Indian
Ocean, exhibited its greatest magnitude recorded since 1997, which was under extremely strong
El Niño conditions. The unique feature of the 2019 IOD event was that it occurred during neutral
ENSO conditions.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S190
AM UTC
4. THE TROPICS
In addition, tropical intraseasonal variability was especially prominent, with three distinct periods of Madden-Julian Oscillation (MJO) activity spanning a total of approximately eight months.
The editors of this chapter would like to insert a personal note recognizing the passing of a
past author of the Tropics Chapter. Our colleague and good friend A. Brett Mullan died of cancer
on 22 April 2020. Brett was a mainstay of this chapter having stewarded the section on the Pacific
Intertropical Convergence Zone from 2006 to 2018. Brett worked for New Zealand’s National Institute of Water and Atmosphere and made significant contributions and authored seminal papers
in meteorology. These included the analysis of SH climate and circulation variability over interannual (El Niño–Southern Oscillation [ENSO]) to interdecadal (interdecadal Pacific Oscillation)
timescales. His work in documenting the relationships of climate variability to long-term global
teleconnections has been a basis for seasonal climate prediction for New Zealand commencing in
the 1990s. He carried out research into climate change and modeling, with particular emphasis
on SH and New Zealand regional effects (Southern Oscillation, greenhouse warming, ocean–atmosphere coupled models and decadal variability, and integrated climate impact models). Over
his 40-year career, Brett’s contributions to meteorology and climate science and beyond were
tremendous. His outstanding work and significant scientific contributions will be his legacy,
and he will be greatly missed.
b. ENSO and the tropical Pacific— M. L’Heureux, G. D. Bell, and M. S. Halpert
The El Niño–Southern Oscillation (ENSO) is a coupled ocean–atmosphere climate phenomenon
over the tropical Pacific Ocean, with opposite phases called El Niño and La Niña. For historical
purposes, NOAA’s Climate Prediction Center (CPC) classifies and assesses the strength and duration of El Niño and La Niña using the Oceanic Niño Index (ONI; shown for mid-2018 through 2019
in Fig. 4.1). The ONI is the 3-month (seasonal) running average of sea surface temperature (SST)
anomalies in the Niño-3.4 region (5°N–5°S, 170°–120°W), currently calculated as the departure
from the 1986–2015 base period mean. El Niño is classified when the ONI ≥ +0.5°C for at least five
consecutive, overlapping seasons. La Niña is similarly defined but for ONI ≤ −0.5°C.
Using the ONI, the minimum threshold for El Niño was reached in September–November (SON)
2018, but the CPC did not declare the onset of El Niño until ocean–atmosphere coupling became
evident in January 2019 (Bell et al. 2019). ONI
values peaked and remained near +0.8°C for
five overlapping seasons (October–December [OND] until March–May [MAM]), then
decreased before El Niño ended in May–July
(MJJ) 2019. This episode was categorized as
weak because the ONI remained between
+0.5°C and +0.9ºC.
The ONI remained positive throughout
2019, and the central Pacific remained
warmer than usual. However, the remainder
of the year was classified as ENSO-neutral
as ONI values decreased to a minimum of
+0.1°C during July–September (JAS) and
August–October (ASO). During the autumn
and early winter, the ONI increased to +0.5°C Fig. 4.1. Time series of the ONI (ºC) from mid-2018 through 2019.
Overlapping 3-month seasons are labeled on the x-axis, with
in OND and +0.6°C in November–January initials indicating the first letter of each month in the season. Red
(NDJ), but the ocean–atmosphere coupling, bars indicate positive values in excess of +0.5ºC. ONI values are
which is normally an intrinsic aspect of El derived from the ERSST-v5 dataset (Huang et al. 2017) and are
based on departures from the 1986–2015 period monthly means.
Niño, was not present during this season.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S191
AM UTC
4. THE TROPICS
1) Oceanic conditions
Seasonal sea surface temperatures (SSTs) and anomalies during December–February (DJF)
2018/19 through SON 2019 are shown in Fig. 4.2. The El Niño during DJF and MAM is indicated
by positive SST anomalies across the central and eastern equatorial Pacific Ocean (Figs. 4.2a–d).
Throughout the event, anomalies exceeding +1.0°C were seen in the central and east-central
equatorial Pacific. These conditions reflected a weaker-than-average equatorial cold tongue in
the eastern Pacific and an eastward expansion of the western Pacific warm pool (approximated
by SSTs above 29°C) to well east of the date line (near 160°W; Fig. 4.2d).
Following the demise of El Niño, equatorial SST anomalies in the central Pacific Ocean remained quite high (near or above +1.0°C) throughout the year, while the anomalies decreased
in the eastern equatorial Pacific, returning to near zero during June–August (JJA) and SON 2019
(Figs. 4.2f,h). A sizable region of 30°C temperatures covered the western equatorial Pacific Ocean,
extending to the date line (Figs. 4.2e,g). Correspondingly, SST anomalies increased to +1.5°C in
the western equatorial Pacific (~170°E) during SON (Fig. 4.2h).
Consistent with the SST evolution, subsurface temperatures during DJF 2018/19 and MAM 2019
were above average across most of the equatorial Pacific (Figs. 4.3a,b). This warming reflected
deepening of the oceanic thermocline and reduced upwelling that accompanies El Niño. Although
ENSO-neutral conditions returned by summer, temperature anomalies near the date line remained
greater than +1.0°C between the surface and 150-m depth (Figs. 4.3c,d).
In contrast, in the far eastern equatorial Pacific, the thermocline was shallower than average,
consistent with the below-average temperatures in this region during JJA (Fig. 4.3c). By SON,
the thermocline and subsurface temperatures were near average across most of the equatorial
Pacific Ocean.
Fig. 4.2. Seasonal SST (left) and anomaly (right) for (a),(b) DJF 2018/19, (c),(d) MAM 2019, (e),(f) JJA
2019, and (g),(h) SON 2019. Contour interval for SST is 1°C. For SST anomaly, contour interval is 0.5°C
for anomalies between ±1ºC, and 1ºC for anomalies > 1ºC and < −1ºC. Anomalies are departures from
the 1981–2010 seasonal adjusted OI climatology (Reynolds et al. 2002).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S192
AM UTC
4. THE TROPICS
Fig. 4.3. Equatorial depth–longitude section of Pacific Ocean temperature anomalies (°C) from the
1981–2010 mean averaged between 5°N and 5°S during (a) DJF 2018/19, (b) MAM 2019, (c) JJA 2019,
and (d) SON 2019. The 20°C isotherm (thick solid line) approximates the center of the thermocline.
The data are derived from an analysis system that assimilates oceanic observations into an oceanic
general circulation model (Behringer et al. 1998).
2) Atmospheric circulation, temperature, and
precipitation anomalies during December–
February 2018/19
The patterns of tropical convection and
winds during DJF 2018/19 generally reflected
El Niño (Figs. 4.4a, 4.5a). In particular, tropical
convection (measured by Outgoing Longwave
Radiation [OLR]) was enhanced near the date
line (green shading) and suppressed over Indonesia (brown shading). Low-level (850 hPa)
tropical wind anomalies were westerly over the
western Pacific Ocean during DJF (Fig. 4.4a),
reflecting a weakening of the trade winds, an
indicator of a weaker Pacific Walker circulation
(Bjerknes 1969).
In the upper atmosphere (200 hPa), tropical
wind anomalies were mostly cross-equatorial
during DJF 2018/19, with flow from the Northern
Hemisphere (NH) subtropics to the Southern
Hemisphere (SH) over the eastern Pacific (Fig.
4.5a). Upper-level wind anomalies reflected
anomalous divergence in association with the
enhanced convection near the date line. Adjacent to this region, two anomalous upper-level
anticyclones flanked the equator, consistent
with El Niño.
Over the Pacific–North American region,
anomalies of 500-hPa heights and upper-level
winds during DJF 2018/19 generally did not
match those conventionally associated with El
Niño. The strengthened and southern-shifted
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Fig. 4.4. Anomalous 850-hPa wind vectors and speed
(contour interval is 2 m s −1) and anomalous OLR (shaded,
W m −2) during (a) DJF 2018/19, (b) MAM 2019, (c) JJA 2019,
and (d) SON 2019. Reference wind vector is below right of
color bar. Anomalies are departures from the 1981–2010
period monthly means. (Source: NCEP–NCAR reanalysis
[Kalnay et al. 1996].)
Unauthenticated | Downloaded 12/04/22 09:14S193
AM UTC
4. THE TROPICS
jet stream was only evident over the far eastern North Pacific Ocean instead of across the
central North Pacific Ocean as expected with
El Niño (Fig. 4.5a). Despite the lack of a clear
El Niño footprint, the anomalous circulation
was linked to increased precipitation over
California, the southeastern United States,
and Florida. However, enhanced precipitation
was also widespread over the entire contiguous United States, with the exception of the
Pacific Northwest and most of Texas, where
near- to slightly-below-average precipitation
occurred. As with the 500-hPa height anomalies, the temperature anomalies over North
America were also not consistent with El Niño
with below-average temperatures over western
Canada and the north-central United States,
and above-average temperatures over the
southern tier of the United States (see sections
7b1 and 7b2).
In other parts of the world, El Niño during
DJF is historically associated with positive
temperature anomalies over the northern half
of South America, Australia, Indonesia, southeast Asia, and southern Africa (Halpert and
Ropelewski 1992). All of these were apparent
during DJF 2018/19 (see relevant temperature Fig. 4.5. Anomalous 200-hPa wind vectors and speed
interval is 4 m s −1), and anomalous OLR (shaded,
sections in Chapter 7 for details), though un- (contour
−2
doubtedly with a partial contribution from the W m ), during (a) DJF 2018/19, (b) MAM 2019, (c) JJA 2019,
and (d) SON 2019. Reference wind vector is below right of
long-term climate change warming signal as color bar. Anomalies are departures from the 1981–2010
well (see section 2b1). El Niño was also likely period monthly means. (Source: NCEP–NCAR reanalysis
associated with above-average precipitation [Kalnay et al. 1996].)
across most of the southern tier of the United
States, Uruguay, and southeastern China during DJF 2018/19 (see relevant precipitation sections in Chapter 7 for details; Ropelewski and Halpert 1989). Likewise, El Niño likely played
some role in below-average precipitation over parts of southern Chile, northern South America,
South Africa, Indonesia, and Australia.
3) Atmospheric circulation, temperature, and precipitation anomalies during March–May
through September–November 2019
The pattern of wind anomalies over the equatorial Pacific Ocean changed from DJF to MAM
2019, with mostly near-average low-level winds (Fig. 4.4b) and anomalous upper-level easterlies
over the western Pacific Ocean during MAM (Fig. 4.5b). By this season, the El Niño was weakening from its boreal winter maximum. However, the East Asia–North Pacific jet stream was
stronger than average across most of the extratropical oceans (Fig. 4.5b), which is typical of El
Niño. Likewise, enhanced precipitation continued over California and much of the contiguous
United States (see section 7b2). Temperatures over the United States, however, were largely a
continuation of the DJF anomalies and not consistent with El Niño. Later in the year, the lowerlevel and upper-level winds were mostly near average over the equatorial Pacific (Figs. 4.4c,d
and 4.5c,d). During SON, convection was suppressed over the Maritime Continent, mostly in
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S194
AM UTC
4. THE TROPICS
association with the strengthening of the Indian Ocean dipole (IOD; section 4h). While SST
anomalies were positive over the western and central equatorial Pacific Ocean, there was no
corresponding increase in convection. In fact, OLR was weakly suppressed near the date line
(Figs. 4.4d, 4.5d).
c. Tropical intraseasonal activity—S. Baxter, C. Schreck, and G. D. Bell
Tropical intraseasonal variability was especially prominent during 2019. Two leading aspects
of this variability were the Madden-Julian Oscillation (MJO; Madden and Julian 1971, 1972, 1994;
Zhang 2005) and convectively coupled equatorial waves (Wheeler and Kiladis 1999; Kiladis et
al. 2009), which include equatorial Rossby waves and atmospheric Kelvin waves. There were
three distinct periods of MJO activity in 2019 spanning a total of approximately eight months
(Fig. 4.6), which were interspersed with the convectively coupled waves (Fig. 4.7). Between the
MJO periods, the tropical convective anomalies were dominated by lower frequency variability
and convectively coupled waves.
Fig. 4.6. Time–longitude section for 2019 of 5-day running
anomalous 200-hPa velocity potential (× 10 6 m2 s−1) averaged
between 5°N–5°S. For each day, the period mean is removed
prior to plotting. Green (brown) shading highlights likely
areas of anomalous divergence and rising motion (convergence and sinking motion). Red lines and labels highlight
the main MJO episodes. Anomalies are departures from the
1981–2010 base period daily means. (Source: NCEP–NCAR
reanalysis [Kalnay et al. 1996].)
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Fig. 4.7. Time–longitude section for 2019 of anomalous OLR
(W m−2) averaged between 10°N–10°S. Negative anomalies
indicate enhanced convection, positive anomalies indicate
suppressed convection. Contours identify anomalies filtered
for the MJO (black) and atmospheric Kelvin waves (red),
and equatorial Rossby waves (blue). Red labels highlight
the main MJO episodes. Contours are drawn at ±10 W m−2,
with the enhanced (suppressed) convective phase of these
phenomena indicated by solid (dashed) contours. Anomalies
are departures from the 1981–2010 base period daily means.
Unauthenticated | Downloaded 12/04/22 09:14S195
AM UTC
4. THE TROPICS
The MJO is a leading intraseasonal climate mode of tropical convective variability. Its convective
anomalies often have a similar spatial scale to El Niño–Southern Oscillation (ENSO) but differ in
that they exhibit a distinct eastward propagation and generally traverse the globe in 30–60 days.
The MJO affects weather patterns around the globe (Zhang 2013), including monsoons (Krishnamurti and Subrahmanyam 1982; Lau and Waliser 2012), tropical cyclones (TCs; Mo 2000; Frank
and Roundy 2006; Camargo et al. 2007; Schreck et al. 2012; Diamond and Renwick 2015), and
extratropical circulations (Knutson and Weickmann 1987; Kiladis and Weickmann 1992; Mo and
Kousky 1993; Kousky and Kayano 1994; Kayano and Kousky 1999; Cassou 2008; Lin et al. 2009;
Riddle et al. 2012; Schreck et al. 2013; Baxter et al. 2014). The MJO is often episodic, with periods
of moderate-to-strong activity followed by little or no activity. The MJO tends to be most active
during ENSO-neutral and weak ENSO periods and is often absent during strong El Niño events
(Hendon et al. 1999; Zhang and Gottschalck 2002; Zhang 2005). Common metrics for identifying
the MJO include time–longitude plots of anomalous 200-hPa velocity potential (Fig. 4.6) and
Outgoing Longwave Radiation (OLR; Fig. 4.7), as well as the Wheeler–Hendon (2004) Real-time
Multivariate MJO (RMM) index (Fig. 4.8). In the time–longitude plots, the MJO exhibits eastward
propagation from upper-left to lower-right. In the RMM, the MJO propagation and intensity are
seen as large, counter-clockwise circles around the origin. When considered together, these diagnostics point to three prolonged MJO episodes during 2019. MJO #1 was a strong and long-lasting
episode that continued from late
2018 (Baxter et al. 2019) through
mid-March 2019. MJO #2 began
in mid-April and persisted into
early July, while MJO #3 began in
mid-August and lasted through
late December. All three MJO periods were associated with either
westerly wind bursts (WWBs) or
trade wind surges (TWS) over the
central Pacific (Fig. 4.9a).
MJO #1 featured a zonal wave1 pattern of strong convective
anomalies. Its periodicity was
approximately 30 days during
January, slowing to about 45 days
during February and March (Figs.
4.6, 4.8a). The plot of anomalous
velocity potential (Fig. 4.6) shows
that MJO #1 circumnavigated the
globe nearly two times during
January–March. The RMM index
indicates the event was strongest
in late February and early March
(Fig. 4.8a). During late March,
Fig. 4.8. Wheeler–Hendon (2004) Real-time Multivariate MJO (RMM) index
coherent eastward propagation
for (a) Jan–Mar, (b) Apr–Jun, (c) Jul–Sep, and (d) Oct–Dec 2019. Each point
gave way to a more stationary
represents the MJO amplitude and location on a given day, and the conconvective pattern with uppernecting lines illustrate its propagation. Amplitude is indicated by distance
level divergence (convergence)
from the origin, with points inside the circle representing weak or no
centered over the west-central
MJO. The eight phases around the origin identify the region experiencing
Pacific Ocean (eastern Indian
enhanced convection, and counter-clockwise movement is consistent with
eastward propagation.
Ocean).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S196
AM UTC
4. THE TROPICS
Impacts from MJO #1 included notable WWB activity over the equatorial Pacific during January
and February (Fig. 4.9a). These WWBs initiated and reinforced the strongest downwelling oceanic Kelvin wave observed in 2019 (dashed line, Fig. 4.9b), which resulted in positive heat content
anomalies exceeding 2°C in early March. This downwelling wave reached the west coast of South
America during April. Prominent TWS were notably absent during early 2019.
MJO #2 occurred from mid-April to early July. Its periodicity was about 45 days, with nearly
canonical eastward propagation throughout its duration. The RMM index showed peak amplitude
during mid- to late May (Fig. 4.8b). Eastward propagation broke down during July, giving way to
less coherent convective anomalies punctuated by westward-moving equatorial Rossby waves.
MJO #2 resulted in alternating low-level zonal wind anomalies over the western and central
Pacific (Fig. 4.9a) that gave rise to both upwelling and downwelling oceanic Kelvin waves. TWS
events in April and June, respectively, resulted in upwelling oceanic Kelvin waves seen as local
minima in heat content anomalies (dotted lines, Fig. 4.9b). A WWB in May resulted in a downwelling oceanic Kelvin wave observed between the aforementioned upwelling periods.
The third and final MJO period of 2019 was associated with the emergence of a wave-1 convective pattern in late August. Both the RMM index and velocity potential anomalies reveal relatively slow propagation during mid-September through mid-October, when a westward-moving
equatorial Rossby wave (Figs. 4.7, 4.9a) interfered with the overall MJO signal. This interference
is seen as a distinct split in the MJO-suppressed phase during late September and early October
(Fig. 4.6). A similar split is visible but less prominent in the enhanced MJO phase at the same time.
Eastward propagation with a periodicity of nearly 40 days resumed in mid- to late October. MJO
#3 reached peak amplitude
in November (Fig. 4.8) as a
very strong suppressed phase
propagated across the Indian
Ocean (Fig. 4.7). Canonical
eastward propagation gave
way to a fast-moving atmospheric Kelvin wave in late
December.
MJO #3 resulted in t wo
prominent WWB events and
associated downwelling oceanic Kelvin waves in September and November, respectively. The first downwelling wave
reached the South American
coast in early December. A
Fig. 4.9. (a) Time–longitude section for 2019 of anomalous 850-hPa zonal
modest TWS in late October
wind (m s −1) averaged between 10°N–10°S. Contours identify anomalies
and the resulting upwelling
filtered for the MJO (black), atmospheric Kelvin waves (red), and equatoseparated t he t wo dow nrial Rossby waves (blue). Significant WWB and TWS over the equatorial
welling waves. MJO #3 also
Pacific that resulted in notable downwelling and upwelling oceanic Kelvin
appears to have played a role
waves are labeled. (b) Time–longitude section for 2019 of the anomalous
equatorial Pacific Ocean heat content, calculated as the mean temperature
in modulating Atlantic hurrianomaly between 0–300 m depth. Yellow/red (blue) shading indicates abovecane activity. During 4–14 Sep(below-) average heat content. Relative warming (dashed lines) and cooling
tember, no new named storm
(dotted lines) due to downwelling and upwelling equatorial oceanic Kelvin
formations occurred when the
waves are indicated. Anomalies are departures from the 1981–2010 base
MJO was producing enhanced
period pentad means. Data in (b) are derived from an analysis system that
upper-level divergence over
assimilates oceanic observations into an oceanic general circulation model
(Behringer et al. 1998).
the central and eastern Pacific
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S197
AM UTC
4. THE TROPICS
(Fig. 4.6), a pattern known to increase vertical wind shear over the tropical Atlantic and be unfavorable for tropical cyclogenesis (Klotzbach 2010). In contrast, the MJO likely contributed to
enhanced Atlantic TC activity during 14 September–2 October (five Atlantic named storms) and
18–30 October (four Atlantic named storms). In both periods, the suppressed phase of the MJO
produced anomalous upper-level convergence
over the central equatorial Pacific, a pattern
that acts to decrease the vertical wind shear
and increase activity over the tropical Atlantic.
d. Intertropical convergence zones
1) Pacific— N. Fauchereau
Tropical Pacific rainfall patterns are dominated by two convergence zones, the Intertropical Convergence Zone (ITCZ; Schneider
et al. 2014) north of the equator and the South
Pacific Convergence Zone (SPCZ; Vincent
1994). Figure 4.10 summarizes their combined
behavior during 2019 using rainfall estimated
from satellite microwave and infrared data in a
product known as CMORPH (Joyce et al. 2004).
Rainfall transects over 20°N–30°S are presented for each quarter of the year, averaged
across successive 30°-longitude bands, starting in the western Pacific at 150°E–180°. The
2019 seasonal variation is compared against
the 1998–2018 climatology.
From January through March, the positive
sea surface temperature (SST) anomalies in the
central Pacific were associated with considerable increases in precipitation around the date
line. During this time, large departures from
normal rainfall were recorded in February
just south of the equator (Figs. 4.10a, 4.11a)
within the SPCZ. A strongly intensified ITCZ
developed in March (Fig. 4.11b). Conversely,
well-below-normal rainfall was recorded in
the western Pacific and the Maritime Continent in February. Persistent dryness affected
many islands within Micronesia during the
first quarter of 2019 (PEAC 2019, Pacific ENSO
update).
Figure 4.12 shows a more detailed comparison of the western Pacific CMORPH rainfall
transect during January–March (JFM) 2019
relative to all other years in the satellite dataset. During JFM, the ITCZ was quite strong,
with the most exceptional rainfall anomalies—
approaching and exceeding the largest values
in the CMORPH dataset—recorded within the
ITCZ in the northern Pacific between 150°E
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Fig. 4.10. Rainfall rate (mm day−1) from CMORPH analysis
for the cross-section between 20°N and 30°S, for (a)
Jan–Mar, (b) Apr–Jun, (c) Jul–Sep, and (d) Oct–Dec 2019.
Each quarter’s panels show the 2019 rainfall (solid line),
and the 1998–2018 climatology (dotted line), for four 30°
sectors from 150°E–180° to 120°–90°W. (Source: CMORPH
[Joyce et al. 2004].)
Fig. 4.11. Rainfall anomalies (mm day−1) for (a) Feb and (b)
Mar 2019. The anomalies are calculated with respect to the
1998–2018 climatology. (Source: CMORPH [Joyce et al. 2004].)
Unauthenticated | Downloaded 12/04/22 09:14S198
AM UTC
4. THE TROPICS
and 180°. This pattern is atypical of the composite anomalies
associated with more canonical
El Niño conditions. However,
it is consistent with an atmospheric response to positive SST
anomalies centered around and
west of the date line, noting the
amplitude of the rainfall anomalies observed are still somewhat
unprecedented.
Rainfall anomalies broadly
Fig. 4.12. Rainfall rate (mm day−1) for Jan–Mar, for each year 1998 to 2018,
consistent with weak El Niño con- averaged over the longitude sector 150°E–180°. The cross-sections are
ditions persisted until about July, color-coded according to NOAA’s ONI, except 2019, which is shown in
after which most El Niño–South- black. Dotted lines are individual years, and solid lines are the average
ern Oscillation (ENSO) indicators over all years in each ENSO phase. Inset legend indicates how many years
dipped below El Niño thresholds went into each composite. (Source: CMORPH [Joyce et al. 2004].)
and ENSO-neutral conditions took
hold. However, the continued development of a positive Indian Ocean dipole (IOD; one of the
strongest on record) influenced rainfall patterns from September through the end of the year,
especially in the western Pacific (section 4h). During this period, dry conditions developed and
impacted some areas of the western Pacific and Maritime Continent again. At the same time, the
ITCZ shifted north of its climatological position in the central and eastern equatorial Pacific east
of the date line.
In November 2019, SST anomalies increased in the central and western Pacific. The continuation of positive IOD conditions well into December led to dry conditions forming across parts
of the western Pacific and the Maritime Continent. The SPCZ was clearly shifted northeast of its
climatological position in the southwest Pacific during December, leading to dry conditions across
Papua New Guinea, the Solomon Islands, Vanuatu, and New Caledonia.
2) Atlantic— A. B. Pezza and C. A. S. Coelho
The Atlantic ITCZ is a well-organized convective band that oscillates between approximately
5°–12°N during July–November and 5°N–5°S during January–May (Waliser and Gautier 1993;
Nobre and Shukla 1996). Equatorial atmospheric Kelvin waves can modulate ITCZ intraseasonal
variability (Guo et al. 2014). ENSO and the Southern Annular Mode (SAM) can also influence
the ITCZ on interannual time scales (Münnich and Neelin 2005). The SAM, also known as the
Antarctic Oscillation, describes the north–south movement of the westerly wind belt that circles
Antarctica, dominating the middle to higher latitudes of the Southern Hemisphere (SH). The
changing position of the westerly wind belt influences the strength and position of cold fronts
and midlatitude storm systems. During a positive SAM event, the belt of strong westerly winds
contracts toward Antarctica. Conversely, a negative SAM event reflects an expansion of the belt of
strong westerly winds towards the equator. The SAM, which was mostly positive in recent years,
started to oscillate between predominantly neutral and negative phases in 2019, with negative
values developing late in the year (see section 6b). This was consistent with an El Niño-like state
in the Pacific, with weak coupling between equatorial Pacific oceanic and atmospheric conditions.
This transition state was associated with an Atlantic ITCZ oscillating around its climatological
position. Occasional southern excursions during March and April contributed to positive rainfall
anomalies offshore and in some small areas of northeastern Brazil during the first half of the year
(Fig. 4.13). These bursts were associated with an anomalously warm Atlantic Ocean south of the
equator and a cool North Atlantic during the first half of the year. This SST pattern reverted to a
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S199
AM UTC
4. THE TROPICS
Fig. 4.13. Observed precipitation anomaly for tropical and subtropical South America (mm day−1) during (a) Jan–Jun and
(b) Jul–Dec 2019. Anomalies are calculated based on a 1998–2018 climatology. (Source: CMORPH [Joyce et al. 2004].)
more neutral set up from June to October, and then
re-intensified toward the end of the year.
The Atlantic Index (Pezza and Coelho 2019),
as defined by the SST south of the equator minus
the SST north of the equator over key areas of
influence for the ITCZ, reflects well the role of the
north–south gradient mechanism highlighted
above for 2019, with the ITCZ tending to shift toward the warmer side of this gradient (Fig. 4.14).
A weaker subtropical South Atlantic anticyclone
associated with a negative SAM also contributed
to re-establish a positive SST anomaly pattern
south of the equator toward the end of the year.
This pattern resulted in an abrupt increase in the
Atlantic Index (Fig. 4.14). This increase is also
consistent with possible atmospheric Kelvin wave
propagation, although the ITCZ was too far north
to be impacted.
e. Global monsoon summary— B. Wang and Q. He
The global monsoon is the dominant mode of
annual tropical–subtropical precipitation and circulation variability and thus a critical part of Earth’s
climate system (Wang and Ding 2008). Figure 4.15
shows global precipitation anomalies, focusing on
monsoon rainfall anomalies, especially over the
land monsoon region, for the monsoon seasons
in the (a) Southern Hemisphere (SH; November
2018–April 2019) and (b) Northern Hemisphere (NH;
May–October 2019), which constitute the global
monsoon year of 2018/19. Figure 4.16 shows the time
series of monsoon precipitation and low-level circulation indices (Yim et al. 2014) for each of the eight
regional monsoons. Note that these precipitation
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Fig. 4.14. (a) Atlantic ITCZ position inferred from OLR
(Liebmann and Smith 1996) during Mar 2019. The colored
thin lines indicate the approximate position for the six
pentads of the month. The black thick line indicates the
climatological position for Mar. SST anomalies for Mar
2019 based on the 1982–2018 climatology are shaded
(°C). Boxes indicate areas used to calculate the Atlantic
index. (b) Atlantic index for 2015–19, based on monthly
OISST (Smith et al. 2008) anomaly time series averaged
over the South Atlantic sector (SA box, 10°–50°W, 5°N–
5°S) minus the same averaged over the North Atlantic
sector (NA box, 20°–50°W, 5°–25°N). A positive index
indicates favorable conditions for enhanced Atlantic
ITCZ activity south of the equator.
Unauthenticated | Downloaded 12/04/22 09:14S200
AM UTC
4. THE TROPICS
Fig. 4.15. Precipitation anomalies (mm day−1) averaged for (a) the SH monsoon season: Nov 2018–Apr 2019 and (b) the NH
monsoon season: May–Oct 2019. Red lines outline the global monsoon precipitation domain defined by two climatological
conditions: first, the local monsoon season precipitation minus that of the cool season exceeds 300 mm and second, the
monsoon season precipitation constitutes at least 55% of the total annual amount (Wang and Ding 2008). Precipitation
indices for each regional monsoon are defined by the areal mean precipitation in the corresponding rectangular regions
(dashed blue), which are highly correlated with that of the corresponding real regional monsoon domains (Table 4.1).
(Source: GPCP [Huffman et al. 2009].)
Fig. 4.16. Summer mean precipitation (green) and circulation (red) indices for each of eight regional monsoons defined in
Table 4.1, normalized by their corresponding std. dev. In each panel, R denotes the correlation coefficient between the
seasonal mean precipitation and circulation indices (sample size: 40). Dashed lines indicate ±0.5 std. dev. The monsoon
seasons are May–Oct for the NH and Nov–Apr for the SH. The normalization method is discussed in Yim et al. (2014).
(Source: GPCP for precipitation; ERA-5 for circulation.)
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S201
AM UTC
4. THE TROPICS
indices represent the average precipitation amount over both land and ocean areas in the boxed
regions shown in Fig. 4.16. The definitions of the circulation indices for each monsoon region are
provided in Table 4.1. In most regions, the precipitation and circulation indices are well correlated,
with correlation coefficients ranging from 0.68 to 0.86, except for the southern African monsoon.
The correlation coefficients in Table 4.1 were computed using monthly mean data from 1979 to 2018
(sample size is 160). The precipitation and circulation indices together provide consistent measurements of the strength of each regional monsoon system.
Global land monsoon precipitation is strongly influenced by tropical sea surface temperature
(SST) anomalies, especially those associated with the El Niño–Southern Oscillation (ENSO; Wang
et al. 2012). As shown in Fig. 4.15a, during the SH monsoon season, precipitation increased over
the central-western Pacific and was suppressed over the Maritime Continent–Australian monsoon
region (Fig. 4.15a). This pattern was consistent with the SST anomalies associated with the weak
El Niño that occurred from January to July 2019. The South American monsoon was characterized
by below-normal precipitation and circulation intensity, especially a significant weakening of
the South American monsoon circulation (Fig. 4.16g). The Australian summer monsoon region
also received markedly less precipitation than normal, but the strength of the corresponding
circulation was near normal (Fig. 4.16h). The southern African summer monsoon precipitation
was near normal, but the circulation intensity was below normal (Fig. 4.16f). Overall, the SH summer monsoon was generally below normal with reduced precipitation and monsoon circulation,
although the degree of weakening varied in the three SH regional monsoons.
During the NH monsoon season, precipitation over the Maritime Continent was significantly
below normal with a prominent reduction of precipitation to the west of Sumatra over the tropical eastern Indian Ocean (Fig. 4.15b). On a regional scale, the northern African monsoon was
characterized by above-normal precipitation and circulation intensity, both of which reached
~1.5 std. dev. (Fig. 4.16e) above normal, indicating a strong monsoon year over northern Africa.
Table 4.1. (Modified from Yim et al. 2014). Definition of the regional summer monsoon circulation indices
and their correlation coefficients (CCs) with the corresponding regional summer monsoon precipitation
indices for the period 1979–2018. All circulation indices are defined by the meridional shear of the zonal
wind at 850 hPa, which measures the intensity (relative vorticity) of the monsoon troughs at 850 hPa
except for northern African (NAF) and East Asian (EA). The NAF monsoon circulation index is defined by
the westerly monsoon strength: U850 (0°–15°N, 60°–10°W), and the EASM circulation index is defined by
the meridional wind strength: V850 (20°–40°N, 120°–140°E), which reflects the east–west thermal contrast
between the Asian continent and the western North Pacific. The precipitation indices are defined by the
areal mean precipitation over the blue box regions shown in Fig. 4.15. The correlation coefficients were
computed using monthly time series (160 summer months) (Jun–Sep [JJAS] in NH [1979–2018] and Dec–
Mar [DJFM] in SH [1979/80–2018/19]). Bolded numbers represent significance at the 99% confidence level.
Region
Definition of the circulation index
CC
Indian (ISM)
U850 (5° –15°N, 40°–80°E) minus
U850 (25°–35°N, 70°–90°E)
0.69
Western North Pacific (WNPSM)
U850 (5°–15°N, 100°–130°E) minus
U850 (20°–35°N, 110°–140°E)
0.80
East Asian (EASM)
V850 (20°–40°N, 120°–140°E)
0.70
North American (NASM)
U850 (5°–15°N, 130°–100°W) minus
U850 (20°–30°N, 110°–80°W)
0.83
Northern African (NAFSM)
U850 (0°–15°N, 60°–10°W)
0.68
South American (SASM)
U850 (20°–5°S, 70°–40°W) minus
U850 (35°–20°S, 70°–40°W)
0.82
Southern African (SAFSM)
U850 (15°S-0°, 10°–40°E) minus
U850 (25°–10°S, 40°–70°E)
0.53
Australian (AUSSM)
U850 (15°S–0°, 90°–130°E) minus
U850 (30°–20°S, 100°–140°E)
0.86
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S202
AM UTC
4. THE TROPICS
Boreal summer precipitation over India was significantly above normal, but precipitation over
Bangladesh and the Indo-China peninsula was below normal. The western North Pacific monsoon
precipitation was ~1 std. dev. above normal (Fig. 4.16b). Both the East Asian summer monsoon
rainfall and its circulation were near normal (Fig. 4.16c) with a dipolar pattern: above-normal
precipitation over the East China Sea to western Japan and below-normal precipitation along the
subtropical frontal zone extending from the middle Yangtze River Valley to the Korean peninsula
(Fig. 4.15b). The North American monsoon was characterized by both below-normal precipitation
and circulation intensity (Fig. 4.16d). Overall, total monsoon precipitation was above normal in the
NH and below normal in the SH. There was a notable contrast between the Eastern and Western
Hemispheres, with increased rainfall over the Eastern Hemisphere tropical monsoon regions and
reduced rainfall over those of the Western Hemisphere (e.g., the American monsoon regions).
f. Tropical cyclones
1) Overview— H. J. Diamond and C. J. Schreck
The IBTrACS dataset comprises historical tropical cyclone (TC) best-track data from numerous
sources around the globe, including all of the World Meteorological Organization (WMO) Regional
Specialized Meteorological Centers (RSMCs; Knapp et al. 2010). This dataset represents the most
complete compilation of global TC data. From
these data, Schreck et al. (2014) compiled
1981–2010 climatological values of TC activity
for each basin using statistics from both the
WMO RSMCs and the Joint Typhoon Warning
Center (JTWC). These values are referenced in
each subsection.
Tallying the global TC numbers is challenging and involves more than simply adding up
basin totals, because some storms cross TC
basin boundaries, some TC basins overlap,
and multiple agencies track and categorize
TCs. Compiling the activity using preliminary
IBTrACS data over all seven TC basins from
NOAA’s National Hurricane Center and the
JTWC (Fig. 4.17), the 2019 season (2018/19 in
the Southern Hemisphere [SH]) had 96 named
storms (sustained wind speeds ≥ 34 kt or 17
m s −1), which is one more than last season
(Diamond and Schreck 2019) and above the
1981–2010 average of 82 (Schreck et al. 2014).
The 2019 season also featured 53 hurricanes/
typhoons/cyclones (HTC; sustained wind
speeds ≥ 64 kt or 33 m s−1), which is above
the climatological average of 46 (Schreck et
al. 2014). During the 2019 season, 31 storms
reached major HTC status (sustained wind
speeds ≥ 96 kt or 49 m s−1), which is also above
the long-term average of 21 and five more than Fig. 4.17. (a) Global summary of TC tracks overlaid on associated
the 2018 season (Diamond and Schreck 2019). OISST anomalies (°C; Reynolds et al. 2002) for the 2019 season
All of these metrics were in the top 10% rela- relative to 1982–2010; (b) global TC counts; and (c) global ACE
values. Horizontal lines on (b) and (c) are 1981–2010 normals.
tive to 1981–2010 (Table 4.2).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S203
AM UTC
4. THE TROPICS
In sections 4f2–4f8, 2018/19
Table 4.2. Global counts of TC activity by basin for 2019. “+” means top
and 2019 seasonal TC activity
tercile; “++” is top 10%; “−” is bottom tercile; “−−” is bottom 10% (all
relative to 1981–2010). “+++” denotes record values for the entire IBTrACS
is described and compared to
period of record. Please note that some inconsistencies between Table
the historical record for each of
4.2 and the text of the various basin write-ups in section f exist and are
the seven WMO-defined hurriunavoidable, as tallying global TC numbers is challenging and involves
more than simply adding up basin totals, because some storms cross TC
cane basins. For simplicity, all
basin boundaries, some TC basins overlap, and multiple agencies are incounts are broken down by the
volved in tracking and categorizing TCs.
U.S. Saffir–Simpson HurriMajor
SS Cat 5
ACE (× 104 kt2)
Basin
TCs
HTCs
cane Wind Scale (SSHWS). The
HTCs
overall picture of global TCs
18
3
2
North Atlantic
6
130
during 2019 is shown in Fig.
+
+
++
4.17; actual counts by category
Eastern
19
7
4
0
97
North Pacific
are documented in Table 4.2.
Globally, five storms dur3
Western
263
27
16
10
North
Pacific
+
ing the year reached SSHWS
Category 5 strength (sustained
8
6
3
85
North Indian
0
+++
++
+++
+++
wind speeds ≥ 137 kt or 70.5
−1
m s ). This was one fewer than
11
10
8
154
South Indian
0
+
++
+++
++
recorded in 2016 (Diamond
and Schreck 2017), equal to
7
3
Australian Region
4
0
68
−
+
the number recorded in 2017
(Diamond and Schreck 2018),
4
0
0
25
Southwest Pacific
6
+
−
and six fewer than the total of
96
53
31
5
11 recorded in 2018 (Diamond
Global Totals
795
++
+
++
+
and Schreck 2019). The alltime record of 12 Category 5
global TCs was set in 1997 (Schreck et al. 2014).1
The five Category 5 storms were: Super Typhoons Wutip, Hagibis, and Halong in the western
North Pacific and Hurricanes Dorian and Lorenzo in the North Atlantic. Dorian caused unprecedented and tremendous devastation, with approximately 70 fatalities reported in the northwest
Bahamas and over $3.4 billion (U.S. dollars) in damages generated there. Dorian was responsible
for six fatalities in Florida and three in North Carolina and caused over $1 billion (U.S. dollars) in
damages in the United States. As a post-tropical cyclone, Dorian also caused considerable damages in Nova Scotia, Canada, with over $100 million (U.S. dollars) in damages reported. While
Lorenzo was a Category 5 storm for a short period of time, it was more deadly as a post-tropical/
extratropical cyclone. Lorenzo produced tropical storm force winds across portions of Ireland,
and was the second deadliest storm of the 2019 North Atlantic season, causing 19 deaths both
at sea and along the U.S. coast as a result of high-surf conditions. Sidebar 4.1 details the recordsetting and devastating local impacts of Hurricane Dorian.
2) Atlantic basin— G. D. Bell, E. S. Blake, C. W. Landsea, M. Rosencrans, H. Wang, S. B. Goldenberg, and R. J. Pasch
(I) 2019 SEASONAL ACTIVITY
The 2019 Atlantic hurricane season produced 18 named storms, of which six became hurricanes
and three achieved major hurricane status (Fig. 4.18a). The HURDAT2 1981–2010 seasonal averages
(included in IBTrACS) are 11.8 named storms, 6.4 hurricanes, and 2.7 major hurricanes (Landsea
and Franklin 2013). The 2019 seasonal Accumulated Cyclone Energy (ACE) value (Bell et al. 2000)
was 134% of the 1981–2010 median (which is 92.4 × 104 kt2; Fig. 4.18b), above NOAA’s threshold
1
SSHWS is based on 1-minute averaged winds, and the categories are defined at: https://www.weather.gov/mfl/saffirsimpson; the
Australian category scale is based on 10-minute averaged winds, and those categories are defined at: https://australiasevereweather
.com/cyclones/tropical_cyclone_intensity_scale.htm
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S204
AM UTC
4. THE TROPICS
(120%) for an above-normal season. The
numbers of named storms and major hurricanes were also both above average.
Therefore, the 2019 season was designated
as above normal by NOAA. This makes
2019 the fourth consecutive above-normal
season, tying the record set in 1998–2001.
This also marks the 17th above-normal
season of the 25 since the current Atlantic
high-activity era began in 1995 (Goldenberg
et al. 2001; Bell et al. 2019).
The previous high-activity era for which
fairly reliable data on TC counts and overall
hurricane strengths exist is 1950–70. That
period also featured numerous abovenormal seasons (10 out of 21), while the
intervening low-activity era of 1971–94 had
only 2 out of 24 (Bell et al. 2018). Note that Fig. 4.18. Seasonal Atlantic hurricane activity during 1950–2019.
the hurricane record is considered far less (a) Numbers of named storms (green), hurricanes (red), and
reliable before 1950, with exact season-to- major hurricanes (blue); 1981–2010 seasonal means shown by
season comparisons for ACE considered less solid colored lines. (b) ACE (Bell et al. 2000) index expressed
reliable before the mid-1970s and the start as percent of the 1981–2010 median value. Red, yellow, and
blue shadings correspond to NOAA classifications for above-,
of the geostationary satellite era (Landsea near-, and below-normal seasons, respectively (http://www.cpc
et al. 2006). Given these caveats, the best .ncep.noaa.gov/products/outlooks/ background_information
estimates suggest that the previous high- .shtml). Thick red horizontal line at 165% ACE value denotes
activity era actually spanned the period the threshold for an extremely active season. Vertical brown
lines separate high- and low-activity eras. Note: There is a low
from 1926–70 (Goldenberg et al. 2001).
The 18 named storms during 2019 are the bias in activity during the 1950s to the early 1970s due to the
lack of satellite imagery and technique (Dvorak) to interpret TC
sixth highest on record since 1950, while
intensity for systems over the open ocean. (Source: HURDAT2
the 2019 ACE value is only the 24th highest [Landsea and Franklin 2013] for TC counts.)
in that 69-year record. This disparity is in
part because two storms (Category 5 Hurricanes Dorian and Lorenzo) produced about 60% of
the season’s ACE. Meanwhile, eight of the named storms were very short-lived (<2 days). There
has been a large artificial increase in these “shorties” since 2000, with seasons averaging about
five per year since that time (Landsea et al. 2010). The increased ability to record these storms
primarily reflects new observational capabilities such as scatterometers, Advanced Microwave
Sounding Units, and the Advanced Dvorak Technique. Villarini et al. (2011) confirmed the lack
of association of the shorties’ time series with any known climate variability.
(II) STORM FORMATION REGIONS AND LANDFALLS
The vast majority of Atlantic TCs typically form during the peak months (August–October,
ASO) of the hurricane season. During 2019, 15 of the 18 named storms, five of the six hurricanes,
and all three major hurricanes formed during ASO.
Historically, the primary cause for an above-normal season is a sharp increase in activity
associated with storms that form within the Main Development Region (MDR), which spans the
tropical Atlantic Ocean and Caribbean Sea between 9.5°N and 21.5°N (Goldenberg and Shapiro
1996; Goldenberg et al. 2001; Bell and Chelliah 2006; Bell et al. 2017, 2018, 2019). For above-normal
seasons during 1981–2010, the ACE value associated with storms first named in the MDR averaged
155% of the median (Fig. 4.19a), compared to only 15.8% during below-normal seasons. During
2019, the MDR-related ACE value was 101% of the median.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S205
AM UTC
4. THE TROPICS
The nearly tenfold increase in ACE that
occurs on average during above-normal seasons reflects the fact that far more MDR-initiated storms eventually become hurricanes
(6.4 compared to 1.0) and major hurricanes
(4.4 compared to 0.4). These differences not
only reflect a nearly four-fold increase in the
number of named storms that form within
the MDR during above-normal seasons (9.3
compared to 2.5), but also a significantly
higher percentage of those storms that become hurricanes (72% compared to 39%) and
major hurricanes (44% compared to 17%;
Fig. 4.19b). These results are consistent with
those of Goldenberg et al. (2001), who noted
a five-fold increase in the number of Caribbean hurricanes for high- versus low-activity
eras. During 2019, six named storms formed
within the MDR, with three (50%) eventually becoming hurricanes and two (33%)
eventually becoming major hurricanes.
Thus, the MDR-related activity during 2019
Fig. 4.19. (a) Seasonal averages of specified storm metrics durwas relatively modest for an above-normal ing 1981–2010 associated with named storms initiated within
season in the entire basin, and no Caribbean the MDR. (b) Percentage of MDR-initiated named storms
during 1981–2010 that eventually became hurricanes (left)
hurricanes were recorded.
Two-thirds (67%) of the named storms dur- and major hurricanes (right). Red (blue) bars show results for
ing 2019 formed outside of the MDR, which is above-normal (below-normal) seasons. (Source: HURDAT2
[Landsea and Franklin 2013].)
a far higher percentage than the 1981–2010
average of 42% for above-normal seasons.
Five of those storms during 2019 formed over the Gulf of Mexico, tying a record with 2003 and
1957 for the most storms to form in that region. The other seven named storms (including one
hurricane) during 2019 formed over the North Atlantic north of the MDR, with all but one tropical
storm forming over the western North Atlantic (west of 55°W and north of 21.5°N). A relatively
high level of TC formation (six named storms including two hurricanes) also occurred over the
western North Atlantic in 2018 (Bell et al. 2019).
Regarding landfalls, the most significant landfalling storm of the 2019 Atlantic hurricane
season was Major Hurricane Dorian, which stalled over Abaco Island and Grand Bahama Island
in the northwest Bahamas during 1–2 September. Dorian spent much of this period at Category 5
intensity, resulting in widespread destruction and death. Dorian tied the Labor Day 1935 hurricane
for the strongest on record to make landfall (based on maximum wind speed) anywhere in the
Atlantic basin. While the intensity of Dorian was continually observed via satellite and extensively
measured by numerous NOAA and Air Force Reserve aircraft reconnaissance flights, the intensity
of the 1935 Labor Day storm was only approximated based on a reading from a single land-based
barometer, and the estimated maximum surface wind speed was derived using pressure-wind
relationships from that one observation.
By 6 September, Dorian weakened to a Category 1 hurricane and made landfall in North Carolina. Two other storms also made landfall in the United States during 2019. These storms were
Barry, which made landfall as a Category 1 hurricane in Louisiana on 13 July, and Tropical Storm
Imelda, which made landfall in Texas on 17 September.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S206
AM UTC
4. THE TROPICS
(III) SEA SURFACE TEMPERATURES
The MDR sea surface temperatures (SSTs) were
above average with an area-averaged SST anomaly
of +0.40°C (Fig. 4.20b). Most locations had departures between +0.25°C and +0.50°C. However, this
anomaly was only slightly higher (by 0.1°C) than
the remainder of the global tropics (Fig. 4.20c).
On multi-decadal time scales, the presence of
higher SST anomalies in the MDR compared to
the global tropics typifies the warm phase of the
Atlantic Multidecadal Oscillation (AMO; Enfield
and Mestas-Nuñez 1999; Bell and Chelliah 2006)
and is characteristic of Atlantic high-activity eras
such as 1950–70 and 1995–present (Goldenberg et
al. 2001; Vecchi and Soden 2007; Bell et al. 2018).
On interannual time scales, large fluctuations in
the relative anomalous warmth of the MDR can
also be seen. This variability can have nothing to
do with the AMO itself and instead reflect factors
such as fluctuations in the wind patterns across
the MDR, El Niño–Southern Oscillation (ENSO), the
Pacific-Decadal Oscillation, and Indian Ocean SST
variability. During ASO 2019, area-averaged SSTs in
both the tropical Indian and tropical Pacific Oceans
were the second highest (anomalies were +0.73°C
and +0.50°C, respectively) in the 1950–2019 record.
The reduction in the relative anomalous MDR
warmth, especially when compared to most years
since 1995, reflected these conditions and should
not be interpreted as an indicator that the warm
AMO phase has ended.
Another important SST signal during ASO reflected above-average SSTs in the western North
Atlantic (red box, Fig. 4.20a), where six TCs formed.
The area-averaged SST anomaly in this region
(+0.60°C) indicates a continuation of exceptional
warmth in that area that began in 2014 (Fig. 4.20d).
Fig. 4.20. (a) Aug–Oct (ASO) 2019 SST anomalies (°C).
(b)–(d) Time series of ASO area-averaged SST anomalies (black) and 5-point running mean of the time
series (red): (b) in the MDR (green box in (a), spanning
20°–87.5°W and 9.5°–21.5°N), (c) difference between
the MDR and the global Tropics (20°N–20°S), and (d)
in the western North Atlantic (red box in (a), spanning 55°–77.5°W and 21.5°–37.5°N). Anomalies are
departures from the 1981–2010 period means. (Source:
ERSST-v5 [Huang et al. 2017].)
(IV) ATMOSPHERIC CONDITIONS
Consistent with the ongoing high-activity era for
Atlantic hurricanes, an interrelated set of conditions
during ASO 2019 favored increased TC activity in the
MDR even if that region was relatively quiet in 2019. These included upper tropospheric anticyclonic
streamfunction anomalies across the subtropical North Atlantic, in association with an enhanced
subtropical ridge (Fig. 4.21a). A similar anomaly pattern was present across the subtropical South
Atlantic Ocean. This pronounced inter-hemispheric symmetry of the anticyclonic anomalies is
typical of an enhanced West African monsoon system (Bell and Chelliah 2006), which is the June–
September portion of the North African monsoon.
During 2019, these conditions were associated with upper-level easterly wind anomalies across
the MDR and lower-level westerly wind anomalies over the eastern half of the MDR (Fig. 4.21b). This
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S207
AM UTC
4. THE TROPICS
overall pattern resulted in anomalously
weak vertical wind shear across the MDR
(Figs. 4.22a,b). The area-averaged magnitude of the vertical wind shear within
the MDR was 7.0 m s−1 (Fig. 4.22c), which
is below the 8 m s−1 threshold considered
conducive to hurricane formation on a
monthly time scale (Bell et al. 2017).
Over the eastern half of the MDR, the
lower-level westerly wind anomalies reflected weaker easterly trade winds (Fig.
4.21b). These anomalies extended upward
to at least the 700-hPa level (Fig. 4.21c), the
approximate level of the African Easterly
Jet (AEJ). This contributed to a deep layer of
anomalous cyclonic relative vorticity (i.e.,
increased horizontal cyclonic shear) along
the equatorward flank of the AEJ. These
conditions are known to favor increased TC
activity within the MDR by helping African
easterly waves to be better maintained and
by providing an inherent cyclonic rotation
to their embedded convective cells (Bell et
al. 2004, 2006, 2017, 2018).
All of the above conditions are typical of
an enhanced West African monsoon system (Gray 1990; Hastenrath 1990; Landsea
et al. 1992; Bell and Chelliah 2006; Bell et
al. 2018). The strength of that monsoon is
a major factor contributing to observed
multidecadal fluctuations in Atlantic hurricane activity because it directly impacts
atmospheric conditions and TC formation
Fig. 4.21. Aug–Oct 2019: (a) 200-hPa streamfunction (contours,
and intensification within the MDR. Dur- interval is 5 × 106 m2 s−1) and anomalies (shaded), and anomalous
ing August–September, one indicator of vector winds (m s−1); (b) anomalous 1000-hPa heights (shaded, m)
the enhanced monsoon was an extensive and vector winds; and (c) anomalous 700-hPa cyclonic relative
−6 −1
area of anomalous 200-hPa divergence vorticity (shaded, × 10 s ) and vector winds. In (a), the upperacross western Africa, with an associ- level ridge discussed in the text is labeled and denoted by the
thick black line. In (c), the thick solid line indicates the axis of the
ated core of negative velocity potential
mean African Easterly Jet, hand-drawn based on total seasonal
anomalies (Fig. 4.23a). Another indicator wind speeds (not shown). Vector scales differ for each panel, and
was enhanced convection (shown by nega- are below right of the color bar. The green box denotes the MDR.
tive Outgoing Longwave Radiation [OLR] Anomalies are departures from the 1981–2010 means. (Source:
anomalies) in the African Sahel region (red NCEP–NCAR reanalysis [Kalnay et al. 1996].)
box, Fig. 4.23b). During August–September, OLR values in this region averaged 237 W m−2 (Fig. 4.23c). Values below 240 W m−2 indicate
deep tropical convection. These values are typical of the current high-activity era, whereas OLR
values generally above 240 W m−2 (indicating a weaker monsoon) were typical of the low-activity
period of the 1980s and early 1990s. These multidecadal fluctuations in monsoon strength coincide with opposing phases (warm and cold, respectively) of the AMO.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S208
AM UTC
4. THE TROPICS
Fig. 4.22. Aug–Oct (ASO) magnitude of the 200–850-hPa vertical wind shear (m s−1): (a) 2019 total magnitude and vector and
(b) 2019 anomalous magnitude and vector versus 1981–2010 means. (c)–(e) Time series of ASO vertical shear magnitude (black)
and 5-point running mean of the time series (red) averaged over (c) the MDR (green box in (a), (b) spanning 87.5°–20°W and
9.5°–21.5°N); (d) the western North Atlantic (red box in (a), (b) spanning 77.5°–55°W and 21.5°–37.5°N); and (e) the Gulf of
Mexico (blue box in (a), (b) spanning 97.5°–80°W and 21.5°–30°N). In (a) and (b), 2019 TC tracks (yellow lines) are shown and
vector scale (m s−1) is below right of color bar. (Source: NCEP–NCAR reanalysis [Kalnay et al. 1996].)
Fig. 4.23. (a) Aug–Sep 2019 anomalous 200-hPa velocity potential (× 10 6 m2 s −1) and divergent wind vectors (m s −1). (b)
Aug–Sep 2019 anomalous OLR (W m−2), with negative (positive) values indicating enhanced (suppressed) convection. (c)
Time series of Aug–Sep total OLR (black) and 5-point running mean of the time series (red) averaged over the African Sahel region (red box in (a) and (b) spanning 20°W–0° and 12.5°–17.5°N). In (a) the upper-level ridge discussed in the text is
labeled and denoted by the thick black line. In (b), contours show total OLR values of 220 W m−2 and 240 W m−2 . In (a) and
(b), the green box denotes the MDR. Anomalies are departures from the 1981–2010 means. (Source: NCEP–NCAR reanalysis
[Kalnay et al. 1996] for velocity potential and wind.)
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S209
AM UTC
4. THE TROPICS
SIDEBAR 4.1: Hurricane
Dorian: A devastating hurricane for the northwest Bahamas—
P. J. KLOTZBACH AND R. E. TRUCHELUT
The 2019 Atlantic hurricane season ended up slightly above
From 0600 UTC on 30 August to 1800 UTC on 31 August, Dorian
normal for most tropical cyclone (TC) parameters, with a total
underwent rapid intensification from 90 kt (46 m s−1) to 130 kt
of 18 named storms, six hurricanes, and three major hurricanes
(67 m s−1) with 24-hour intensification rates ranging between
occurring. By far, the most significant and devastating hurricane
30 kt (15 m s−1) and 35 kt (18 m s−1). Dorian slowed as it apof the 2019 season was Hurricane Dorian. Dorian will be most
proached the northwest Bahamas, then underwent another
remembered for the devastation that it caused in the northwest
burst of rapid intensification, becoming a Category 5 hurricane
Bahamas, especially on the Abaco Islands and on Grand Baas it approached Great Abaco Island.
hama Island. It was also the longest-lived (14 days as a named
Soon thereafter, Dorian reached its maximum intensity of
storm and 10 days as a hurricane) and most intense (1-minute
160 kt (82 m s−1) as it made landfall on Great Abaco Island
maximum sustained winds of 160 kt (82 ms−1) hurricane of the
on 1 September. In doing so, Dorian became the strongest
hurricane on record to make landfall in the Bahamas and
2019 season (Avila et al. 2020). Dorian also generated the most
tied with the Labor Day Hurricane of 1935 for the strongest
Accumulated Cyclone Energy (ACE) of any Atlantic hurricane,
landfalling hurricane on record anywhere in the Atlantic basin.
accounting for ~40% of basinwide ACE accrued in 2019. This
sidebar summarizes the meteorological history of Dorian along
The 160 kt (82 m s−1) intensity achieved by Dorian was also
with the notable records that the hurricane achieved during its
the strongest on record by any Atlantic hurricane outside of
track across the Atlantic. Historical landfall records from 1851–
the tropics (>23.5°N) in the satellite era (since 1966). Dorian
present are taken from the National Hurricane Center/Atlantic
tracked slowly over Great Abaco as the steering currents colOceanographic and Meteorological Laboratory archive located
lapsed, and the system effectively stalled after making landfall
at: http://www.aoml.noaa.gov/hrd/hurdat/All_U.S._Hurricanes.
on Grand Bahama Island with maximum sustained winds of
html, and Dorian’s observed values are taken from Avila et al
155 kt (80 m s−1) (Fig. SB4.2). Dorian was the first Category
(2020).
5 hurricane on record to make landfall on Grand Bahama
Dorian became a tropical depression (TD) on 24 August in the
Island. Its extremely slow forward movement caused devastatcentral tropical Atlantic and was upgraded to a tropical storm
ing wind, rain, and storm surge impacts over these islands.
(TS) shortly thereafter (Fig. SB4.1). Despite moving through an
During its first 24 hours over Grand Bahama Island, Dorian
environment of relatively low wind shear
and a warm sea surface (~28°–29°C),
considerable mid-level dry air inhibited
Dorian’s intensification early in its lifetime.
Dorian passed through the Windward
Islands on 27 August as a TS. Dorian’s
center reformed farther north after interacting with Saint Lucia, and its center
also reformed downshear (i.e., to the
east) due to moderate westerly shear. This
northeastward shift in track from where
the models were originally forecasting the
storm allowed it to avoid the elevated terrain of Hispaniola and Puerto Rico, which
would have likely weakened the storm. It
then turned northwestward and intensified as it moved into a more moisture-rich
environment. Dorian became a hurricane
as it tracked over Saint Croix on 28 August
and reached major hurricane intensity on Fig. SB4.1. NOAA’s National Hurricane Center Best Track Plot for Hurricane Dorian
30 August as it approached the Bahamas. (Avila et al. 2020).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
S210
Unauthenticated | Downloaded
12/04/22 09:14 AM UTC
4. THE TROPICS
weakened extremely slowly—from a 155 kt (80 m s−1) Category
minimum central pressure of 910 hPa was tied with Ivan for
the ninth-lowest lifetime minimum central pressure since 1980.
5 hurricane to a 115 kt (59 m s−1) Category 4 hurricane. Because
Dorian generated 49 × 104 kt2 ACE during its lifetime—the fifth
of this stalling motion and maintenance of strong hurricane
intensity, Dorian generated the most ACE in a 1° × 1° grid box
most for an August TC in the satellite era. It also generated 14
in the Atlantic basin in the satellite era (i.e., since 1966; Wood
named storm days, tying it with Felix (1995) for third place for
et al. 2020).
most named storm days by a storm forming in August in the
Land interaction, an increase in vertical wind shear, and
satellite era.
cold water upwelling continued to slowly reduce Dorian’s wind
Given its extreme intensity and slow forward speed over both
strength, and it weakened below major hurricane strength
Great Abaco Island and Grand Bahama Island, Dorian caused
late on 3 September. Dorian tracked northward offshore of the
tremendous devastation, with over 70 fatalities reported by
southeast United States and briefly regained major hurricane
the Bahamian Health Minister and $3.4 billion (U.S. dollars) in
strength on 5 September before weakening as it encountered
damage generated (Avila et al. 2020). Dorian was responsible
lower sea surface temperatures (SSTs) and stronger vertical
for four indirect fatalities in the United States and caused $1.6
wind shear. It brushed the South Carolina and North Carolina
billion (U.S. dollars) in damage. Dorian as a post-tropical cyclone
coasts, and Dorian made landfall on Cape Hatteras at 1230
also caused considerable damage in Nova Scotia, with over $100
UTC on 6 September as a Category 2 hurricane, with winds
million (U.S. dollars) in damage being reported.
−1
estimated at 85 kt (44 m s ), although most
of the strongest winds remained over water
to the east of the center (Avila et al. 2020). At
the time of its North Carolina landfall, Dorian’s
central pressure was 956 hPa, tying it with
Floyd (1999) and Florence (2018) for the sixth
lowest central pressure for a landfalling North
Carolina hurricane since 1950. Dorian became
extratropical as it accelerated northeastward,
but it also strengthened slightly during this
time. It made a final landfall as a post-tropical
cyclone in Nova Scotia on 7 September, bringing hurricane-force winds to portions of
Atlantic Canada. Dorian made a final landfall
as a post-tropical storm in Newfoundland on
8 September.
Dorian was an extremely long-lived storm
and set several records due to both its intensity and longevity. Its 160 kt (82 m s−1) winds
were tied with Gilbert (1988) and Wilma
(2005) for the second strongest on record
for an Atlantic hurricane in the satellite era
(since 1966), trailing only the 165 kt (85 m s−1) Fig. SB4.2. NOAA-18 infrared satellite image of Hurricane Dorian making
winds recorded by Allen (1980). Its lifetime landfall on Grand Bahama Island at 154 UTC on 2 Sep 2019.
Despite the above conditions, the 2019 TC activity for the MDR as a whole was relatively modest. This signal partly
reflected the limited activity (two tropical storms) over the Caribbean Sea due in part to anomalously strong upperlevel convergence (Fig. 4.23a) and sinking motion there. The modest activity was also associated with the synoptic
scale sinking motion typically found downstream of the mean ridge axis, which in this case extended across the Gulf
of Mexico and western subtropical North Atlantic (indicated by thick black line in Figs. 4.21a, 4.23a).
Two other aspects of the interannual variability during ASO 2019 include the relatively high number of six TC formations over the western subtropical North Atlantic and five over the Gulf of Mexico (yellow lines, Fig. 4.22). These
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
S211
Unauthenticated | Downloaded
12/04/22 09:14 AM UTC
4. THE TROPICS
are roughly double the 1981–2010 averages seen during above-normal seasons. In addition to
anomalously warm SSTs during ASO (Fig. 4.20a), both regions experienced below-average vertical
wind shear (Fig. 4.22b) with area-averaged shear values at or below 8 m s−1 (Figs. 4.22d,e). For the
Gulf of Mexico, the area-averaged shear was less than 6 m s−1 (Fig. 4.22e), which is comparable to
some of the lowest values in the record. These conditions were linked to the persistent, anomalous
upper-level ridge that extended across both regions (Fig. 4.21a).
3) Eastern North Pacific and Central North Pacific basins— K. M. Wood and C. J. Schreck
(I) SEASONAL ACTIVITY
Two agencies are responsible for issuing advisories and warnings in the eastern North Pacific
(ENP) basin: NOAA’s National Hurricane Center in Miami, Florida, covers the region from the
Pacific coast of North America to 140°W, and NOAA’s Central Pacific Hurricane Center in Honolulu, Hawaii, covers the central North Pacific (CNP) region between 140°W and the date line. This
section combines statistics from both regions.
A total of 19 named storms formed in the combined ENP/CNP basin, seven of which became
hurricanes and four became major hurricanes. The 1981–2010 IBTrACS seasonal averages for the
basin are 16.5 named storms, 8.5 hurricanes, and 4.0 major hurricanes (Schreck et al. 2014). Thus,
2019 storm counts were near normal (Fig. 4.24a). These storms occurred between the official start
date of the ENP season of 15 May and end date of 30 November. Hurricane Alvin first reached
tropical storm strength on 29 June—the latest first storm formation since 2016’s Tropical Storm
Agatha was named on 2 July. The final named storm, Raymond, dissipated on 17 November. Four
Fig. 4.24. (a) Annual storm counts for the eastern North Pacific by category during 1970–2019, with 1981–2010 average
denoted as dashed lines. (b) Annual ACE during 1970–2019, with 2019 in orange and the 1981–2010 average denoted by the
dashed line. (c) Daily ACE during 1981–2010 (solid black) and 2019 (solid green); accumulated daily ACE during 1981–2010
(dashed blue) and 2019 (dashed orange).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S212
AM UTC
4. THE TROPICS
of the 19 storms either formed within or entered the CNP basin from the east, placing 2019 slightly
below the 1981–2010 IBTrACS seasonal average of 4.7 named storms for the CNP.
Unlike 2018, which set a new record for basin-wide ACE (318 × 104 kt2; Wood et al. 2019), the
2019 seasonal ACE index was 98 × 104 kt2, or 74% of the 1981–2010 mean of 132 × 104 kt2 (Fig. 4.24b;
Bell et al. 2000; Schreck et al. 2014). The bulk of TC activity was confined to late June through
late September (Fig. 4.24c); no hurricanes developed in October or November.
Three TCs contributed more than half of 2019’s total ACE and reached Category 4 intensity
on the SSHWS. Each underwent rapid intensification (≥ 30 kt or 15.4 m s−1 in 24 hours) prior to
reaching peak intensity and then rapidly weakened (≤ −30 kt or −15.4 m s−1 in 24 hours; Wood and
Ritchie 2015). The strongest storm of the season, Hurricane Barbara (30 June–6 July) peaked at
135 kt (69 m s−1), just shy of Category 5 strength. Hurricane Erick (27 July–3 August) intensified by
50 kt (26 m s−1) in 24 hours, and Hurricane Kiko (12–24 September) reached Category 4 intensity
(115 kt; 59 m s−1) after similarly explosive intensification from 60 kt to 115 kt in 24 hours. All three
TCs maintained peak intensity for only 12 hours before weakening rapidly.
(II) ENVIRONMENTAL INFLUENCES ON THE 2019 SEASON
The El Niño of 2018/19 transitioned to a neutral state in mid-2019, and seasonal SSTs were
about average near most TC formation locations (Fig. 4.25a). Anomalous warmth dominated
the western part of the basin, particularly north of 10°N, but few storms crossed this region.
Below-average OLR was largely colocated with TC tracks (Fig. 4.25b), and vertical wind shear
was slightly weaker than normal where
most TCs formed (Fig. 4.25c). As in 2018,
the strongest easterly wind shear anomalies occurred in the central Pacific, but
few 2019 TCs reached that region. Again,
enhanced low-level westerlies dominated
west of 140°W, and the enhanced 850-hPa
easterly flow west of Central America (Fig.
4.25d) resembles the pattern in 2018 that
was attributed to anomalously strong gap
winds inf luenced by the Sierra Madre
mountain range (Kruk and Schreck 2019).
Both the Madden Julian Oscillation
(MJO) and convectively-coupled Kelvin
waves are known to affect ENP TC activity,
particularly cyclogenesis (e.g., Maloney
and Hartmann 2001; Aiyyer and Molinari
2008; Schreck and Molinari 2011; Ventrice et al. 2012a,b; Schreck 2015, 2016).
To examine convective variability during
the 2019 ENP hurricane season, Fig. 4.26
shows unfiltered, MJO-filtered, and Kelvin Fig. 4.25. 15 May–30 Nov 2019 anomaly maps of (a) SST (°C;
wave-filtered OLR anomalies computed Banzon and Reynolds 2013); (b) OLR (W m −2; Schreck et al.
using the methodology of Kiladis et al. 2018); (c) 200–850-hPa vertical wind shear (m s −1) vector (ar(2005, 2009). In general, the MJO remained rows) and scalar (shading) anomalies; and (d) 850-hPa wind
−1
weak in the ENP for much of the hurricane (m s ; arrows) and zonal wind (shading) anomalies. Anomalies
are relative to the annual cycle from 1981–2010, except for
season. However, a strong convectively
SST, which is relative to 1982–2010. Letters denote where
suppressed MJO phase in June likely con- each ENP TC attained tropical storm intensity. Wind data are
tributed to the late start to the season. The obtained from CFSR (Saha et al. 2014). The more westward A
subsequent weaker convective envelope represents “Akoni” and the more westward E “Ema.”
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S213
AM UTC
4. THE TROPICS
Fig. 4.26. Longitude–time Hovmöller diagram of 5°–15°N
average OLR (W m −2 ; Schreck et al. 2018). Unfiltered
anomalies from a daily climatology are shaded. Negative
anomalies (green) indicate enhanced convection. Anomalies filtered for Kelvin waves are contoured in blue at −10
W m −2 and MJO-filtered anomalies in black at ±10 W m −2 .
Letters denote the longitude and time when each ENP TC
attained tropical storm intensity. The second A represents
“Akoni” and the second E “Ema.”
may have contributed to the formations of Alvin, Barbara, and Cosme. Multiple Kelvin waves
crossed the ENP, potentially influencing the development of Dalila, Gil, Juliette, Narda, Octave, Priscilla, and Raymond. Beyond these influences, easterly waves—shown in Fig. 4.26
as westward-moving negative anomalies—likely supported the genesis of multiple storms
including Ivo, Priscilla, and Raymond.
(III) TROPICAL CYCLONE IMPACTS
Three TCs directly impacted land in 2019. Hurricane Lorena (17–22 September) made two landfalls in Mexico as a Category 1 hurricane, once in Jalisco and once in Baja California Sur, with one
reported death (Avila 2019). In addition, Lorena’s remnant moisture reached the southwestern
United States where it likely contributed to thunderstorms and heavy rain. Tropical Storm Narda
(29 September–1 October) also made landfall twice in Mexico, both times as a tropical storm, causing six fatalities largely attributed to freshwater flooding (Blake 2019). Short-lived Tropical Storm
Priscilla (20 October) made landfall hours after being named and dumped more than 150 mm of
rain in Nayarit. Its remnant moisture may have contributed to severe weather in the south-central
United States (Stewart 2019). Though none produced significant damage, three landfalling storms
in Mexico is well above the long-term average of 1.8 each year (Raga et al. 2013). Beyond landfalls,
remnants of Tropical Storm Raymond (15–17 November) may have influenced the development of
a low-pressure system that subsequently produced wind, rain, and flooding in the southwestern
United States (NOAA 2019).
4) Western North Pacific basin— S. J. Camargo
(I) OVERVIEW
The 2019 TC season in the western North Pacific (WNP) was slightly above normal by most
measures of TC activity. The data used here are primarily from JTWC best-track data for 1945–2018
and preliminary operational data for 2019. All statistics are based on the 1981–2010 climatological
period with the exception that landfall statistics use 1951–2010.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S214
AM UTC
4. THE TROPICS
A total of 28 TCs (climatological median = 26) reached tropical storm (TS) intensity in the WNP
during 2019, including Pabuk, which formed in December 2018. Of these, 17 reached typhoon
intensity (median = 16), with 4 reaching super typhoon status (≥ 130 kt, median = 3.5). There
were also three tropical depressions (TDs; median 3.5). While 61% of the tropical storms became
typhoons (median 64%), 23% of the typhoons intensified to super typhoons (median 24%). In
Fig. 4.27a, the number of storms in each category is shown for the period 1945–2019.
The Japan Meteorological Agency (JMA) total for 2019 was 29 storms (top tercile ≥ 29), also
including 2018 Tropical Storm Pabuk. While the JMA and JTWC totals are very close, there were
some differences between the two agencies.2 Kajiki was considered a TS by JMA but a depression by JTWC. Matmo was considered a severe TS by JMA and a typhoon by JTWC. Tapah was
classified as a TS for JTWC and a typhoon for JMA. Tropical Storm Sepat was not included as a
TC by JTWC, and Tropical Storm Four was not classified as a TC by JMA. Of the 29 TCs recorded
by JMA, nine were tropical storms (top quartile ≥ 7); three were severe tropical storms (bottom
quartile ≤ 4); and 17 were typhoons (top quartile ≥ 17). Fifty-nine percent of the storms reached
typhoon intensity (median 58%). The number of all TCs (1951–76) and TSs, severe TSs, and typhoons (1977–2019) according to the JMA are shown in Fig. 4.27b. The Philippine Atmospheric,
Geophysical and Astronomical Services Administration (PAGASA) named all 20 TCs that entered
its area of responsibility, including Tropical Depressions Amang, Chedeng, Goring, and Marilyn,
which were not named by JMA.
(II) SEASONAL ACTIVITY
The season started with Tropical Storm Pabuk, which formed on 30 December 2018 and lasted
until 7 January 2019, reaching TS status on 1 January. Super Typhoon Wutip was the season’s next
named storm and the second TC to reach super typhoon intensity in February in the historical record,
following Super Typhoon Higos (2005). Typhoon Mitag (2002) also formed in February, but reached
super typhoon intensity in March. No named storms formed during March–May (bottom quartile =
0 for each month). Only Tropical Storm Four was active in the month of June according to the JTWC
(bottom quartile ≤ 1), while JMA named Tropical Storm Sepat that month. Three TSs were active
during July: Mun, Danas, and Nari (bottom quartile ≤ 2). Tropical Storm Wipha formed at the end
of July, but was mostly active in August and therefore was considered as an August storm. Besides
Wipha, five other TCs occurred in August: Tropical Storms Bailu and Podul, Typhoons Francisco and
Krosa, and Super Typhoon Lekima. August had a total of six named storms (top quartile ≥ 6), three
typhoons (median = 3), and one super typhoon (top quartile ≥ 1). Five more named storms formed
in September (median = 5): Tropical Storms Peipah and Tapah and typhoons Faxai, Lingling, and
Mitag (median = 3). October was an active month with four typhoons: Hagibis, Neoguri, Bualoi,
and Matmo (top quartile ≥ 4), with Hagibis reaching super typhoon intensity. The basin continued
to be very active in November with six named storms (top quartile ≥ 3): Tropical Storm Fung-Wong
and Typhoons Halong (super typhoon), Nakri, Fengshen, Kalmaegi, and Kammuri (top quartile ≥
2). The six named storms and five typhoons matched the historical record for November, set in 1964
and 1968, respectively. The season ended with Typhoon Phanphone in December (median = 1).
As shown in Figs. 4.27c–f, the early season (January–June) was relatively quiet, with only three
tropical storms (bottom quartile ≤ 2.5) and one typhoon (bottom quartile ≤ 1) which reached super
typhoon intensity. The peak season (July–October) had near-normal activity with 18 named storms
(median = 17), 10 typhoons (bottom quartile ≤ 9), and two super typhoons (median = 2). In contrast,
the late season (November–December) was quite active, with seven named storms (top quartile ≥
4.5, maximum = 7) and six typhoons (top quartile ≥ 3) including one super typhoon (top quartile
= 1). The occurrence of six typhoons during November and December is a historical record. This
2
It is well known that there are systematic differences between the JMA and the JTWC datasets, which have been extensively documented in the literature (e.g. Wu et al. 2006; Nakazawa and Hoshino 2009; Song et al. 2010; Ying et al. 2011; Yu et al. 2012; Knapp
et al. 2013; Schreck et al. 2014).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S215
AM UTC
4. THE TROPICS
Fig. 4.27. (a) Number of tropical storms (TS) typhoons (TY) and super typhoons (STY) per year in the WNP for 1945–2019
based on JTWC. (b) Number of tropical cyclones (TC, all storms which reach TS intensity or higher) for 1951–1976; number
of TSs, severe tropical storms (STS) and TY for 1977–2019 based on JMA. Panel (c) shows the cumulative number of TCs
per month in the WNP in 2019 (black line) and climatology (1981–2010) as box plots (interquartile range: box, median: red
line, mean: blue asterisk, values in the top or bottom quartile: blue crosses, high [low] records in the 1945–2018 period:
red diamonds [circles]). Panel (e) is similar to panel (c) but for the number of TYs. Panels (d) and (f) show the number of
TCs and TYs per month in 2019 (black line) and the climatological mean (blue line); blue “+”signs denote the maximum and
minimum monthly historical records and the red error bars show the climatological interquartile range for each month. In
the case of no error bars, the upper and/or lower percentiles coincide with the median. (Sources: 1945–2018 JTWC best-track
dataset, 2019 JTWC preliminary operational track data, except for panel [b], which is 1951–2019 JMA best-track dataset.)
active late season compensated for the quiet early season, leading to a slightly above-average
typhoon season in terms of the JTWC numbers of named storms (28, median = 26), typhoons (17,
median = 16) and super typhoons (4, median = 3.5).
The total ACE in 2019 (Fig. 4.28a) was slightly below normal. As seen in Fig. 4.28b, the value
for February, however, was the highest in the historical record. From March until July, the
monthly ACE was in the bottom quartile of the monthly climatologies, with zero ACE values for
March, April, and May. The August ACE was in the below-average quartile (25%–50%), and the
September ACE was also in the bottom quartile of the monthly climatological distribution. The
October and December ACE values were in the above-average quartile (50%–75%) of the climatological distributions, while November ACE was in the top quartile. The five months of October,
November, August, September, and February contributed 26%, 22%, 16%, 14%, and 15% of the
ACE respectively, summing to 93% of the seasonal ACE. In descending order of storm ACE, Super
Typhoons Hagibis (top 5%), Wutip (top decile), and Halong, Typhoons Bualoi and Kammuri, and
Super Typhoon Lekima placed in the top quartile of historical ACE per storm. Together, these
six storms contributed 57% of the seasonal ACE, with Hagibis and Wutip contributing 14% and
12%, respectively.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S216
AM UTC
4. THE TROPICS
Fig. 4.28. (a) ACE per year in the WNP for 1945–2019. The solid green line indicates the median for 1981–2010; dashed lines
show the climatological 25th and 75th percentiles. (b) ACE per month in 2019 (black line) and the 1981–2010 median (blue
line); red error bars indicate the 25th and 75th percentiles. In case of no error bars, the upper and/or lower percentiles
coincide with the median. The blue “+” signs denote the maximum and minimum values during the 1945–2018. (Source:
1945–2018 JTWC best-track dataset; 2019 JTWC preliminary operational track data.)
The mean genesis location in 2019 was 14.4°N, 138.4°E, slightly northwest of the climatological mean of 13.2°N, 142.8°E (std. dev. of 1.9° latitude and 5.6° longitude). The mean track position
in 2019 was 19.0°N, 133.9°E, similarly northwest of the climatology mean of 17.3°N, 136.6°E (std.
dev. of 1.4° latitude and 4.7° longitude). There is a well-known connection between genesis and
track shifts in the WNP basin and ENSO phase (Chia and Ropelewski 2002; Camargo et al. 2007).
However, the 2019 northwest shift in TC genesis and track cannot be attributed to El Niño, as
there were neutral ENSO conditions during the peak typhoon season.
There were 110.25 named storm days in the WNP in 2019 (median = 113 days). The WNP had
50.25 typhoon days (bottom quartile ≤ 49.5 days) and 21.5 major typhoon days (SSHWS Categories
3–5; median = 21). The percentage of days with typhoons and major typhoons was 32% (bottom
quartile ≤ 33%) and 14% (median = 13.9%) respectively. The median lifetime for TCs reaching TS
intensity was 6.25 days (bottom percentile ≤ 6.25 days) and for those reaching typhoon intensity
was eight days (bottom quartile ≤ 7.75 days). The longest-lived named storm in 2019 was Typhoon
Matmo (12.25 days), followed by Major Typhoons Krosa (11.25 days), Wutip (11 days), and Kammuri
(10.5 days)—all of which were in the top quartile (≥ 10.5 days). Tropical depression One was very
long-lived as well (18 days). Of the 28 tropical storms and typhoons, 17 had a lifetime at or below
the median (7.75 days), with 12 in the bottom quartile (≤ 5.25 days). The maximum number of TCs
(and typhoons) active simultaneously in 2019 was three and occurred 7–9 November (Super Typhoon Halong and Typhoons Matmo and Nakri). The historical record is six (14–15 August 1996).
Including TDs, 23 storms made landfall in 2019, above the 90th percentile compared with
the 1951–2010 climatology. Landfall here is defined when the storm track is over land, and the
previous location was over the ocean. In order to include landfall over small islands, tracks were
interpolated from 6-hourly to 15-minute intervals, using a high-resolution land mask. Seven storms
made landfall as TDs (above the 95th percentile ≥ 7) and eight as tropical storms (top quartile ≥
8). Six TCs made landfall as Category 1–2 typhoons on the SSHWS scale (median = 5): Francisco,
Lekima, Faxai, Hagibis, Matmo, and Phanphone. Super Typhoons Lingling and Kammuri made
landfall as major typhoons (Category ≥ 3; median = 2). Lingling affected both South and North
Korea—the latter of which is not hit frequently by typhoons. Kammuri made landfall in the Bicol region of the Philippines on 2 December, followed by Typhoon Phanphone’s landfall in the
country’s eastern Samar region on 24 December. Five storms made landfall in Japan in 2019 (top
quartile ≥ 5), with the strongest being Typhoons Faxai and Hagibis.
(III) ENVIRONMENTAL CONDITIONS
Figures 4.29a–e show the July–October (JASO) environmental conditions associated with typhoon
activity in 2019. The 2018/19 El Niño transitioned to near- to below-normal SSTs in the eastern Pacific
during the beginning of the peak typhoon season (July to mid-September). From mid-September
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S217
AM UTC
4. THE TROPICS
Fig. 4.29. (a) SST anomalies (°C) for Jul–Oct (JASO) 2019. (b) Potential intensity anomalies in JASO 2019 (kt). (c) Relative
humidity (%) 600-hPa anomalies in JASO 2019. (d) Genesis potential index (unitless) anomalies in JASO 2019. First position
of storms in JASO 2019 are marked with an asterisk. (e) 850-hPa zonal winds (m s −1) in JASO 2019. (Source: atmospheric
variables: NCEP–NCAR reanalysis data [Kalnay et al. 1996]; SST [Smith et al. 2008].)
until the end of the calendar year, above-normal SSTs expanded from the date line into the eastern
Pacific. Below-normal eastern Pacific SSTs are clearly seen in the JASO SST anomalies (Fig. 4.29a),
with above-average SST anomalies extending northeastward from the equatorial central Pacific
around these cold anomalies. The above-average SST anomalies in the central Pacific are reflected
in other environmental variables, such as positive potential intensity anomalies (Fig. 4.29b) in the
eastern part of the basin near the date line. There was also a positive band of 600-hPa relative humidity anomalies between 130°–160°E extending from the equatorial region to ~40°N (Fig. 4.29c).
For the genesis potential index (GPI; Fig. 4.29d; Emanuel and Nolan 2004; Camargo et al. 2007),
anomalies are observed in a recurving narrow band between 10°–20°N. Many TC genesis locations
occurred close to or just southwest of this region. The extent of the monsoon trough, defined by 850hPa zonal winds (Fig. 4.29e), extended to 150°E, despite below-normal SSTs in the eastern Pacific.
Several cases of TC genesis occurred just north of these westerly anomalies.
(IV) TROPICAL CYCLONE IMPACTS
Many storms had social and economic impacts in Asia in 2019, particularly Typhoons Lekima,
Faxai, and Hagibis. Typhoon Lekima made landfall in China as the fifth-strongest landfalling
typhoon to affect the country since 1949, according to the China Meteorological Administration.
Lekima’s heavy rainfall and long duration over China led to many historical daily precipitation
records across the country, particularly in Zhejiang Province, where the typhoon made its first
landfall. Lekima then passed over Shanghai and Jiangsu Province, before making a second
landfall in Shangdong Province. Lekima left 48 dead and 21 missing, and displaced 1.7 million
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S218
AM UTC
4. THE TROPICS
people. Damages were estimated to be $9.3 billion (U.S. dollars). Typhoon Faxai impacted Japan
as one of the strongest typhoons on record to affect Tokyo, killing three people and injuring 147,
causing extensive blackouts, and damaging more than 40 000 homes. Japan’s economic losses
across several sectors due to Faxai are estimated at $7 billion (U.S. dollars). In October, Super
Typhoon Hagibis affected the Tokyo region. The storm's record-breaking rainfall led to extensive
flooding of rivers and dams and multiple landslides. At least 95 people were killed, 460 injured,
and economic losses exceeded $10 billion (U.S. dollars).
5) North Indian Ocean basin— A. D. Magee and C. J. Schreck
(I) SEASONAL ACTIVITY
The North Indian Ocean (NIO) TC season typically occurs between April and December, with
two peaks of activity: May–June and October–December, due to the presence of the monsoon
trough over tropical waters of the NIO during these periods. Tropical cyclone genesis typically
occurs in the Arabian Sea and Bay of Bengal between 8°–15°N. The Bay of Bengal, on average,
experiences four times more TCs than the Arabian Sea (Dube et al. 1997).
The 2019 NIO TC season was a particularly active and record-breaking TC season with eight
named storms, six cyclones, and three major cyclones (tied 1999), compared to the IBTrACS–JTWC
1981–2010 climatology of 4.9, 1.5, and 0.7, respectively (Fig 4.30a). One event, Cyclone Kyarr, was
the second-most intense cyclone ever observed in the Arabian Sea. The 2019 NIO TC season was
also the second-costliest on record with losses exceeding $11 billion (U.S. dollars).
Record-breaking ACE index values and a strongly positive Indian Ocean dipole (IOD) event
characterized the 2019 NIO TC season (refer to the legend of Fig. 4.38 for the definition of IOD and
its polarity). The 2019 seasonal ACE index was 85 × 104 kt2. It nearly doubled the previous record
holders (2007 and 2013 each had about 45 × 104 kt2) and was over four times the 1981–2010 ACE
climatology (19 × 104 kt2; Fig 4.30b). The strong positive IOD event that marked the latter half of the
2019 season is clearly seen in Fig 4.31a, where anomalously warm SSTs occurred in the western
tropical Indian Ocean (10°N–10°S, 50°–70°E).
In addition, enhanced convection (Fig 4.31b)
and negative vertical wind shear anomalies
(Fig 4.31c) provided favorable conditions in
the Arabian Sea, contributing to the high
number of events there. Although positive
IOD events historically result in fewer TCs in
the NIO (Yuan and Cao 2012), this was not the
case for the 2019 TC season.
(II) NOTEWORTHY TROPICAL CYLONES
AND IMPACTS
The first severe cyclonic storm of the 2019
NIO cyclone season, Cyclone Fani (27 April–3
May), developed unusually close to the equator, at 2.7°N, just west of Sumatra. Strong
vertical wind shear impeded further development until 29 April when Fani intensified into
a severe cyclonic storm. On 30 April, favorable conditions aided further intensification
before Fani recurved north-northeastward
toward India. It then underwent additional
intensification, reaching its peak intensity
of 135 kt (69 m s−1) and a minimum central
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Fig. 4.30. Annual TC statistics for the NIO for 1980–2019:
(a) number of named storms, cyclones, and major cyclones
and (b) estimated annual ACE index (× 10 4 kt 2) for all TCs
during which they were at least tropical storm strength or
greater intensity (Bell et al. 2000). Horizontal lines represent
1981–2010 climatology.
Unauthenticated | Downloaded 12/04/22 09:14S219
AM UTC
4. THE TROPICS
Fig. 4.31. 15 Sep 2019–15 Dec 2019 NIO anomaly maps of (a) SST (°C; Banzon and Reynolds 2013), (b) OLR (W m −2 ;
Schreck et al. 2018), (c) 200–850-hPa vertical wind shear (m s −1) vector (arrows) and scalar (shading) anomalies, and
(d) 850-hPa winds (m s −1, arrows) and zonal wind (shading) anomalies. Anomalies are relative to the annual cycle
from 1981–2010, except for SST, which is relative to 1982–2010 due to data availability. Letter symbols denote where
each NIO TC attained its initial tropical storm intensity. Wind data are obtained from CFSR (Saha et al. 2014).
pressure of 917 hPa, equivalent to a strong Category 4 system on the SSHWS. Fani made landfall
near Odisha, India, on 3 May with 1-minute maximum sustained wind speeds of 120 kt (62 m s−1).
Fani eventually weakened and moved into Bangladesh on 4 May before dissipating the following
day. In total, 89 people were killed with estimated damages of approximately $8.1 billion (U.S.
dollars). A storm surge of approximately 1.5 m and heavy rainfall resulted in extensive damage,
including to agricultural land, where more than 30% of crops were damaged. In Bangladesh, 17
people were killed, many by lightning. Around 160 000 acres of farmland was destroyed, with
agricultural losses in Bangladesh totaling $4.6 million (U.S. dollars).
Cyclone Vayu (10–17 June) formed in the southeastern Arabian Sea, aided by a strong convective
pulse of the MJO. Vayu reached peak intensity on 13 June, with 1-minute maximum sustained winds
of 95 kt (48 m s−1), and a minimum central pressure of 950 hPa, a Category 2 SSHWS equivalent
system. During its lifetime, Vayu’s track recurved several times before weakening to a tropical low
off the coast of Gujarat, India, and passed over the coast on 18 June. Vayu’s proximity to Gujarat
and surrounding regions resulted in eight deaths and an estimated $140 000 (U.S. dollars) in
damages. Cyclone-generated waves and swells resulted in flooding of low-lying areas in Oman
and Pakistan. Vayu contributed to an approximate one-week delay in the northward migration of
the Indian monsoon, which was already delayed by weakening El Niño conditions that persisted
during the early part of the 2019 monsoon season.
The fourth cyclone of the season, Hikaa (22–25 September), developed in the Arabian Sea and
intensified into a severe cyclonic storm, reaching peak intensity of 85 kt (43 m s−1) and a minimum
central pressure of 972 hPa, a Category 2 SSHWS equivalent system. Hikaa tracked toward the
west before making landfall near Duqm, Oman, on 24 September. As a result of Hikaa, a boat
carrying 11 Indian fishermen reportedly sank, while another boat sank off the coast of Duqm.
Super Cyclone Kyarr (24 October–1 November) was the second-most intense cyclone ever
observed in the Arabian Sea with a peak intensity of 135 kt (69 m s−1) and a minimum central
pressure of 923 hPa. After forming in the southeastern Arabian Sea, high SSTs and low vertical
wind shear favored rapid intensification. Kyarr reached Super Cyclonic Storm strength (India
Meteorological Department 2020) on 27 October—the first in the Arabian Sea since Cyclone Gonu
in 2007. Unfavorable conditions resulted in a weakening of Kyarr, and it later dissipated on 1
November. No fatalities were recorded as a result of Kyarr; however, strong winds and intense
rainfall caused flash flooding in Goa, India. High tide and extreme sea levels associated with
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S220
AM UTC
4. THE TROPICS
Kyarr caused coastal flooding in Oman, with extensive damage to the Muttrah Corniche as well
as a portion of the coastline of the United Arab Emirates.
Cyclone Maha (30 October–6 November), the season’s fourth cyclonic storm to originate in the
Arabian Sea (compared to an average of one), intensified in a similar fashion to, and occurred concurrently with, Cyclone Kyarr. Maha underwent rapid intensification from depression to cyclonic
storm within a 12-hour period. The peak intensity of Maha on 4 November was 105 kt (54 m s−1) and
a minimum central pressure of 959 hPa, a Category 3 SSHWS equivalent system. Maha generally
tracked toward the northwest throughout its lifetime, parallel to the west coast of India, and generated storm surge up to 0.5 m (at Asarsa and Tankaria) on 2 November. Upwelling of cooler SSTs
weakened Maha, and it made landfall as a depression near Gujarat and dissipated shortly thereafter.
Cyclone Bulbul (7–10 November) originated in the Bay of Bengal from a previous disturbance,
Severe Tropical Storm Matmo, and emerged into the north Andaman Sea. After tracking westnorthwestward toward the central Bay of Bengal, Bulbul moved to the north, intensifying to a
very severe cyclonic storm on 8 November, with 1-minute maximum wind speeds of 85 kt (43 m s−1)
and a minimum central pressure of 971 hPa, a Category 2 SSHWS equivalent system. Bulbul made
landfall near the Sagar Islands of West Bengal on 9 November. It brought significant rainfall, with
reports of 24-hour accumulations of up to 202 mm in Canning, West Bengal. In total, 41 people
died, with estimated damage totaling $3.38 billion (U.S. dollars). In the state of Odisha, rainfall
caused agricultural damage, including an estimated 200 000 ha of damaged crops. In Bangladesh, more than two million people fled to shelters, 25 people died, and approximately 14% of
Bangladesh’s total farmland was damaged, equating to agricultural losses of approximately $31
million (U.S. dollars).
6) South Indian Ocean basin— A. D. Magee and C. J. Schreck
(I) SEASONAL ACTIVITY
The South Indian Ocean (SIO) TC basin extends south of the equator and from the African
coastline to 90°E. In the SIO, TC genesis typically occurs south of 10°S. While the SIO TC season
extends year-round, from July to June, the majority of activity occurs between November and April
when the ITCZ is located in the SH. The 2018/19 TC season includes TCs that occurred from July
2018–June 2019. Landfalling TCs typically impact Madagascar, Mozambique, and the Mascarene
Islands, including Mauritius and Réunion Island. The Regional Specialized Meteorological Centre (RSMC) on La Réunion is the official monitoring agency for TC activity within the SIO basin.
The 2018/19 SIO season had 11 named storms, 10 cyclones, and eight major cyclones (Fig 4.32a),
compared to the IBTrACS–JTWC 1981–2010 mean of 9.1, 5.5, and 2.9, respectively (Schreck et al.,
2014). The eight major cyclones broke the previous record of seven in 1993/94. The 2018/19 SIO season also had a record-breaking number of cyclone days, 39 days in total, overtaking the previous
records of 1993/94 (36 days) and 2001/02 (35 days). Unfortunately, the season also set records for
deaths and economic losses with over 1300 fatalities and total damage exceeding $2.3 billion (U.S.
dollars). Cyclone Idai caused the majority of deaths and damage and was one of the worst natural
disasters on record to impact southern Africa.
The 2018/19 seasonal ACE index was 154 × 104 kt2, above the 1981–2010 SIO average of 92 × 104
2
kt (Fig. 4.32b). Cyclone-favorable environmental conditions, including anomalously warm SSTs
(Fig. 4.33a), enhanced convection (Fig. 4.33b), and anomalously weak shear (Fig. 4.33c) were
present where the majority of TCs developed. The presence of low-level westerly anomalies along
10°S enhanced cyclonic vorticity for many systems, excluding TCs east of 70°E, where easterly
anomalies predominated.
(II) NOTEWORTHY TROPICAL CYCLONES AND IMPACTS
The first named cyclone of the season intensified to a Category 3 SSHWS equivalent system,
with maximum 1-minute sustained winds of 100 kt (51 m s−1) and a minimum central pressure of
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S221
AM UTC
4. THE TROPICS
Fig. 4.32. Annual TC statistics for the SIO for 1980–2019:
(a) number of named storms, cyclones, and major cyclones
and (b) estimated annual ACE index (× 10 4 kt 2) for all TCs
during which they were at least tropical storm strength or
greater intensity (Bell et al. 2000). Horizontal lines represent
1981–2010 climatology.
965 hPa. After tracking in a west-southwesterly
direction toward Madagascar, Cyclone Alcide
(6–13 November 2018) quickly weakened due to
less favorable conditions and did not make landfall, although it did cause minor damage on the
Mauritian island of Agaléga.
Cyclone Desmond (20–22 January 2019) formed
as a TD off the east coast of Mozambique and reFig. 4.33. Nov 2018–Apr 2019 SIO anomaly maps of (a) SST
curved several times before tracking toward the
(°C; Banzon and Reynolds 2013); (b) OLR (W m−2; Schreck
northeast. Desmond intensified into a moderate
et al. 2018); (c) 200–850-hPa vertical wind shear (m s −1)
−1
TS with a peak intensity of 45 kt (23 m s ) and
vector (arrows) and scalar anomalies (shading), and (d)
850-hPa winds (m s −1 arrows) and zonal wind anomalies
minimum central pressure of 993 hPa. Desmond
(shading). Anomalies are relative to the annual cycle from
made landfall in Mozambique approximately
1981–2010, except for SST, which is relative to 1982–2010.
200 km north of Beira, bringing 277 mm of rainLetter symbols denote where each SIO TC attained its
fall over a 24-hour period. Significant flooding
initial tropical storm intensity. (Source: Wind data from
resulted in deaths of over 1000 livestock and
CFSR [Saha et al. 2014].)
affected approximately 60 000 ha of crops.
Cyclone Galena (6–15 February) intensified northeast of Mauritius and reached a peak intensity
of 120 kt (61 m s−1) with a minimum central pressure of 933 hPa, a Category 4 SSHWS equivalent
system. It passed within 35 km of Rodrigues where wind gusts of 90 kt (46 m s−1) were recorded.
Winds associated with Galena devastated the agricultural sector on Rodrigues and damaged approximately 90% of the island’s electricity grid.
Cyclone Idai (4–16 March), a Category 3 SSHWS equivalent system, was the deadliest TC ever
recorded in the SIO basin. Over 1300 people lost their lives, and 3 million people were affected or
displaced across Mozambique, Zimbabwe, Malawi, and Madagascar. Idai made two landfalls over
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S222
AM UTC
4. THE TROPICS
Mozambique. It remained over Mozambique for five days after its first landfall (4 March) before moving offshore. Just before Idai’s second landfall, the system intensified, reaching peak intensity with
maximum 1-minute sustained winds of 105 kt (54 m s−1) and minimum central pressure of 944 mb.
Its second landfall was near Beira, Mozambique, on 15 March, and it remained over Mozambique
for three days. Multiple landfalls amplified the impacts associated with Idai, which are described
in Sidebar 7.3.
Cyclone Joaninha (22−31 March), a Category 4 SSHWS equivalent system, formed to the east of
Madagascar. On 24 March, Joaninha achieved peak intensity, with maximum 1-minute sustained
winds of 115 kt (59 m s−1) and a minimum central pressure of 937 hPa. Cyclone Joaninha was a
slow-moving storm and passed within ~80 km of Rodrigues, Mauritius, with cyclonic conditions
persisting there for more than 34 hours. Wind gusts up to 100 kt (51 m s−1) and rainfall accumulations of 200 mm were recorded, resulting in widespread power cuts and flooding.
Cyclone Kenneth (23–26 April) was the most intense landfalling TC in Mozambique’s observational record and also resulted in significant damage to the Comoro Islands, Tanzania, and
Mozambique. At its peak, Kenneth reached a maximum intensity of 125 kt (64 m s−1), a category
4 SSHWS equivalent system. It passed ~60 km north of Grande Comore Island and resulted in
significant impacts there, which are described in Sidebar 7.3. Kenneth made landfall on 25 April,
north of Pemba, Mozambique, with 1-minute sustained winds of 120 kt (61 m s−1). Kenneth’s
widespread destruction in Mozambique came as the nation was still coming to terms with the
substantial impacts of TC Idai, just six weeks before.
7) Australian basin— B.C. Trewin
(I) SEASONAL ACTIVITY
The 2018/19 TC season was near normal in the broader Australian basin (areas south of the
equator and between 90°E and 160°E,3 which includes Australian, Papua New Guinea, and Indonesian areas of responsibility). The season produced 11 TCs, which is near the 1983/84–2010/11
average4 of 10.8, and is consistent with neutral ENSO conditions. The 1981–2010 IBTrACS seasonal
averages for the basin are 9.9 named storms, 7.5 TCs, and 4.0 major TCs, which compares with
the 2018/19 counts of 10, six, and two, respectively (Fig 4.34).
There were six TCs in the western sector5 of the broader Australian region during 2018/19, four
in the northern sector, and five in the eastern sector.6 Three systems made landfall in Australia
as TCs (two on multiple occasions), affecting Queensland and the Northern Territory, while a
fourth approached the coast closely enough to have major impacts on land on the Pilbara coast
in Western Australia. All cyclone categories referred to in this section are based on the Australian
cyclone categorization scale.
(II) LANDFALLING AND OTHER SIGNIFICANT TROPICAL CYCLONES
The strongest cyclone of the season was Veronica, which approached the Pilbara coast in late
March. Veronica was named at 1800 UTC on 19 March near 15°S, 120°E. It intensified rapidly over
the following 24 hours while moving generally west-southwest, and it reached Australian Category
5 intensity at 0000 UTC on 21 March, near 17°S, 118°E, with maximum sustained 10-minute wind
speeds of 115 kt (59 m s−1) and a central pressure of 928 hPa. It weakened slightly as it moved toward
3
4
5
6
The Australian Bureau of Meteorology’s warning area overlaps both the southern Indian Ocean and southwest Pacific.
Averages are taken from 1983/84 onward as that is the start of consistent satellite coverage of the region.
The western sector covers areas between 90°E and 125°E. The eastern sector covers areas east of the eastern Australian coast to
160°E, as well as the eastern half of the Gulf of Carpentaria. The northern sector covers areas from 125°E east to the western half of
the Gulf of Carpentaria. The western sector incorporates the Indonesian area of responsibility, while the Papua New Guinea area
of responsibility is incorporated in the eastern sector.
Trevor and Owen passed through both the northern and eastern sectors, Wallace through both the northern and western sectors,
and Lili through both the northern sector and the Indonesian warning area of responsibility, which is included with the western
sector.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S223
AM UTC
4. THE TROPICS
the coast but was still at Category 4 intensity
near the coast, about 100 km northeast of
Karratha, at 0000 UTC on 24 March. Veronica
then remained near-stationary, moving less
than 50 km in 24 hours, while slowly weakening, before resuming its movement west, parallel to the Pilbara coast. It weakened below
TC intensity by 0000 UTC on 26 March. The
remnant low crossed North West Cape later
that day before dissipating to the west.
While Veronica did not make landfall as a
TC, its prolonged presence caused extended
shutdowns of mining, oil, and gas industries
in the region, with economic losses from
lost production estimated at over $1.4 billion
(U.S. dollars). There was also heavy rain in
the Pilbara region, with 72-hour totals for
24–26 March of 634 mm at Indee, 572 mm
at Sherlock, 539 mm at Mallina, 470 mm at Fig. 4.34. Annual TC statistics for the Australian basin for
Upper North Pole (near Marble Bar), and 356 1980–2019: (a) number of named storms, cyclones, and major
4
2
mm at Port Hedland. Local and river flooding cyclones and (b) the estimated annual ACE (× 10 kt ) for all
caused traffic disruptions and some livestock TCs at least tropical storm strength or greater intensity (Bell et
al. 2000). The 1981–2010 means (horizontal lines) are included
losses.
in both (a) and (b).
On 17 March, TC Trevor formed in the Coral
Sea at 1800 UTC. It moved west while intensifying and made its initial landfall as a Category 3 system near Lockhart River, on the Cape York Peninsula, around 0900 UTC on 19 March. It weakened
to a Category 1 system while crossing the Peninsula, before re-emerging south of Weipa on the
morning of 21 March. Once over the Gulf of Carpentaria, Trevor reintensified rapidly while moving southwest, reaching Category 4 intensity early on 23 March with maximum sustained winds
of 95 kt (49 m s−1) while off the coast west of the Northern Territory–Queensland border. It made
landfall around 0000 UTC at slightly below-peak intensity, east of Port McArthur on the Northern
Territory coast. The system weakened below TC intensity that evening as it moved inland, but it
remained as a remnant low for several days, initially moving south through the eastern Northern
Territory and then east through Queensland, finally dissipating near Richmond on 28 March.
There was substantial tree and some building damage near the point of Trevor’s initial landfall at Lockhart River. The second landfall was in a sparsely populated area, and few impacts
were reported. Storm surge heights east of that landfall reached 1.8 m at Burketown and 1.7 m at
Mornington Island. Precautionary evacuations were carried out in a number of communities on
the island of Groote Eylandt and parts of the Northern Territory coast. The heaviest rainfalls from
Trevor were near the point of its Cape York Peninsula landfall, with 302 mm (and a two-day total
of 421 mm) at Lockhart River on 20 March and 211 mm at Aurukun on 21 March. East of Trevor’s
second landfall, Westmoreland Station received 282 mm on 24 March. Following landfall, numerous daily rainfalls exceeding 100 mm were recorded in the eastern Northern Territory and far
western Queensland, including 178 mm on 27 March at Trepell Airport, north of Boulia. The postlandfall rains caused widespread flooding on both sides of the Northern Territory–Queensland
border, with significant cattle losses in some areas. Floodwaters moved south and eventually
contributed to a partial filling of Lake Eyre.
TC Owen initially formed in the Coral Sea on 2 December, but soon weakened and moved
west before making landfall near Port Douglas as a tropical low early on 10 December. The system crossed Cape York Peninsula and emerged over the Gulf of Carpentaria, moving west and
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S224
AM UTC
4. THE TROPICS
intensifying to Category 3 intensity with maximum sustained winds of 80 kt (41 m s−1). It briefly
touched the Northern Territory coast north of Port McArthur early on 13 December, at peak intensity, before beginning to move east and almost retracing its path across the Gulf. It made landfall
again on the east coast of the Gulf, near the mouth of the Gilbert River, at slightly below-peak
intensity, at about 1900 UTC on 14 December. Owen weakened to a tropical low that crossed Cape
York Peninsula a second time and re-emerged into the Coral Sea. The two cyclone-intensity landfalls of Owen were both in remote, sparsely populated areas, and few impacts were reported. The
major impacts were from flooding on the east coast during its tropical low phases. On 10 December,
Kirrama Range (west of Cardwell) received 349 mm, but the most extreme rainfall occurred where
the low moved offshore late in its lifetime. On 16 December, 681 mm fell at Halifax, the heaviest daily total recorded in Australia in December, and a number of other sites on the east coast
exceeded 500 mm. There was substantial local flooding and some crop damage was reported.
The season’s other landfall was Penny, which peaked at Category 2 intensity in the Coral Sea
after crossing Cape York Peninsula from the west and reforming. It made landfall near Weipa as a
Category 1 system on the afternoon of 1 January. Savannah remained well off the coast of western
Australia while peaking at Category 4 intensity in mid-March, but the precursor low brought heavy
rain to the Indonesian island of Java, with substantial flooding and some loss of life.
8) Southwest Pacific basin— J.-M. Woolley, P. R. Pearce, A. M. Lorrey, and H. J. Diamond
(I) SEASONAL ACTIVITY
The 2018/19 TC season in the southwest Pacific officially began in November 2018 and ended
in April 2019; however there was both early and late activity in this region with “out of season”
storms. Storm track data for 2018/19 were gathered from the Fiji Meteorological Service, Australian
Bureau of Meteorology, and New Zealand MetService, Ltd. The southwest Pacific basin (defined by
Diamond et al. 2012 as 135°E–120°W) had nine TCs, including four severe TCs (based on the Australian TC intensity scale). As noted in section
4e1, Fig. 4.35 shows the standardized TC
distribution based on the basin spanning
the area from 160°E–120°W to avoid overlaps with the Australian basin that could
result in double counting of storms. However, it is important to use the definition of
the southwest Pacific basin of Diamond et
al. (2012) as that is how annual TC outlooks
are produced and disseminated.
The 1981–2010 Southwest Pacific Enhanced Archive of Tropical Cyclones
(SPEArTC) indicates a seasonal average
of 10.4 named TCs and 4.3 severe TCs. The
2018/19 TC season therefore had near-normal activity with nine named TCs, of which
four were severe. The first and last storm
formed outside of the formally defined TC
season, with TC Liua occurring in the Solomon Sea in late September 2018 and TC Ann Fig. 4.35. Annual TC statistics for the southwest Pacific for
1980–2019: (a) number of named storms, cyclones, and major
developing in May 2019. The ratio of severe
cyclones and (b) the estimated annual ACE index (× 10 4 kt 2) for
TCs relative to the total number of named all TCs at least tropical storm strength or greater intensity (Bell
TCs in 2018/19 was 44%, which is 6% lower et al. 2000). The 1981–2010 means (horizontal lines) are included
than the previous season.
in both (a) and (b).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S225
AM UTC
4. THE TROPICS
(II) LANDFALLING AND OTHER SIGNIFICANT TROPICAL CYCLONES
The first named TC of the 2018/19 season was reported as a tropical disturbance on 24 September 2018 to the east-northeast of Port Moresby in Papua New Guinea. On 26 September, the
system moved southeast and intensified into a Category 1 TC named TC Liua. After intensifying
to Category 1, TC Liua turned west and began to track toward Port Moresby, weakening into a TD
on 28 September and further dissipating over the northern Coral Sea over the following days. TC
Liua’s peak 10-minute sustained winds were 40 kt (21 m s–1) and its minimum central pressure
was 994 hPa.
Severe TC Owen began as a low-pressure system over the Solomon Islands that developed into
a tropical low on 29 November. The system became more organized the following day as it tracked
southwest toward Tagula Island, then strengthened further as it tracked over the Coral Sea in
favorable conditions. On 2 December, the system was classified as a Category 1 TC, but Owen
weakened rapidly on 4 December and was downgraded to a tropical low. The degradation into a
tropical low was temporary, as this system made landfall north of Cardwell, Queensland, on 10
December and re-attained Category 1 intensity on 11 December over the Gulf of Carpentaria. TC
Owen looped and tracked back east, peaking as a Category 3 severe TC with maximum 10-minute
sustained winds of 81 kt (42 m s−1). On 15 December, TC Owen made landfall near Kowanyama
as a low-end Category 3 severe TC. TC Owen’s passage over northern Queensland brought heavy
rainfall to the region.
Penny was the third TC of the season, which began as a tropical low located near the eastern coastline of Cape York Peninsula, Queensland, in late December 2018. The system tracked
westward, emerging in the Gulf of Carpentaria on 31 December before turning eastward and
strengthening into a Category 1 storm on the same day. On 1 January, TC Penny made landfall on
the western Cape York Peninsula coastline, south of Weipa and was downgraded to a gale-force
tropical low as it weakened over land. On 2 January, TC Penny achieved Category 1 status again
after reorganizing over the Coral Sea. TC Penny’s peak 10-minute sustained winds were 51 kt (26
m s−1), and its minimum central pressure was 987 hPa.
TC Mona began as a tropical low near the southern Solomon Islands in a trough stretching
across the northern Coral Sea. On 3 January, TC Mona achieved Category 1 status, north of Fiji.
Mona intensified to Category 2 status the following day. It then tracked south toward Fiji and
dissipated on 7 January. Approximately 2000 people took shelter in evacuation centers, and 30
roads were closed, mostly due to floods and some landslides. TC Mona’s peak 10-minute sustained
winds were 51 kt (26 m s−1), and its minimum central pressure was 985 hPa.
Severe TC Oma began as a tropical low which had developed within an active monsoon trough
along the coast of Vanuatu on 7 February. On 11 February, Oma intensified into a TC, quickly
reaching Category 2 TC intensity. Oma achieved Category 3 TC status on 16 February, and again
on 19 February following a brief weakening. Oma’s peak 10-minute sustained winds were 70 kt
(36 m s−1), and its minimum central pressure was 974 hPa. Oma weakened to a Category 2 TC as
it tracked southwest toward the Australian coast. On 22 February, TC Oma transitioned into a
subtropical cyclone while turning to the northeast and continued to weaken further over the following days as it tracked farther in this direction. On 27 February, Oma turned eastward, while
situated over Vanuatu, and dissipated on 28 February.
During early February, TC Oma pushed a bulk carrier freighter aground on a coral reef in the
Solomon Islands, resulting in an oil spill, with an estimated cleanup cost of $50 million (U.S.
dollars). Vanuatu was affected for several days by persistent heavy rain, damaging surf, and
strong winds, particularly in the northern provinces of Malampa, Sanma, and Torba. Storm surge
reportedly extended up to 50 m inland in some locations, impacting houses along the coast, particularly those constructed using traditional methods. In Torba, communications and transport
links to the north were disrupted while flooding cut off road access to main services such as the
hospital. New Caledonia was also impacted by heavy rain and damaging winds from TC Oma.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S226
AM UTC
4. THE TROPICS
Thousands of people there were left without power while flooding made some roads impassable.
Agriculture in New Caledonia was significantly affected, and the French government released
$1.43 million (U.S. dollars) for recovery. Queensland was hit by large swells for about one week,
causing significant beach erosion. More than 30 people required rescue, with some hospitalized,
due to turbulent waters. One person drowned just off North Stradbroke Island. Heavy winds also
damaged Cavendish banana plantations in Cudgen, New South Wales.
Severe TC Pola began as a tropical disturbance that formed northeast of Tonga on 23 February.
Pola intensified into a TD while moving slowly southward. Pola became a Category 1 TC on 26
February and intensified into a Category 2 TC later that day. On 27 February, the system became
a severe TC. On 28 February, Pola reached its peak intensity as a Category 4 TC with 10-minute
sustained winds of 89 kt (46 m s−1) and a minimum central pressure of 950 hPa.
Severe TC Trevor originated as a tropical low which formed off of the east coast of Papua New
Guinea on 15 March. The system tracked southeast, crossing Papua New Guinea south of Port
Moresby on 16 March. On 19 March, Trevor made landfall on the far northeast of the Queensland
coast as a Category 3 severe TC and crossed Cape York Peninsula, downgrading to a Category 1
storm as it did so. As TC Trevor tracked southwest across the Gulf of Carpentaria, it intensified
rapidly to a Category 4 system and then made landfall on the Northern Territory’s Gulf coastline
east of Borroloola on 23 March. The storm weakened as it moved inland. TC Trevor’s peak 10-minute
sustained winds were 94 kt (49 m s−1), and its minimum central pressure was 950 hPa. Flooding
in Queensland associated with the cyclone caused a farm to suffer loss of cattle and damage to
equipment estimated to cost at least $710 000 (U.S. dollars). There was little reported in terms of
major damage or injuries in the Northern Territory.
TC Ann originated from a tropical low that formed on 7 May, east of Honiara in the Solomon
Islands. The low tracked slowly toward the southwest in a favorable environment, passing close
to Honiara on 8 May and then moved southward, passing between the Australian cyclone region
and South Pacific cyclone region three times over several days. On 11 May, the system intensified
into a Category 1 TC before turning west-northwest and further strengthening over the Coral Sea.
On 12 May, Ann reached peak intensity as a Category 2 TC with 10-minute sustained winds of 51
kt (26 m s−1) and a central barometric pressure of 993 hPa. TC Ann weakened to a gale-force tropical low on 14 May and made landfall near Lockhart River on Cape York Peninsula on 15 May. The
system continued to track west-northwest for several days and dissipated as a tropical low near
East Timor on 18 May. Impacts associated with TC Ann were relatively minor, with heavy rainfall
and gusts experienced in many areas south of where the system made landfall as a tropical low.
g. Tropical cyclone heat potential— R. Domingues, G. J. Goni, J. A. Knaff, I-I Lin, and F. Bringas
Upper-ocean thermal conditions observed during 2019 within the seven tropical cyclone (TC)
basins are described here with respect to the long-term mean (1993–2018) and to conditions observed in 2018. The analysis focuses on vertically integrated temperature conditions based on
the Tropical Cyclone Heat Potential (TCHP; e.g., Goni et al. 2009, 2017) which is calculated as the
integrated heat content between the sea surface and the depth of the 26°C isotherm (the minimum
temperature required for genesis and intensification, Leipper and Volgenau 1972; Dare and McBride
2011). The TCHP is an indicator of the amount of heat stored in the upper ocean and available
to fuel TC intensification and modulates TC-induced sea surface temperature (SST) cooling and
ocean−hurricane enthalpy fluxes (e.g., Lin et al. 2013). Areas in the ocean with TCHP values above
50 kJ cm−2 have been associated with TC intensification and rapid intensification (e.g., Shay et al.
2000; Mainelli et al. 2008; Lin et al. 2014; Knaff et al. 2018), provided that atmospheric conditions
are also favorable. Salinity in the upper layers also modulates upper-ocean turbulent mixing and,
thus, can also impact the depth of the 26°C isotherm and the corresponding TCHP values (e.g.,
Balaguru et al. 2015; Domingues et al. 2015).
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S227
AM UTC
4. THE TROPICS
The analysis developed here focuses
primarily on seasonal TCHP anomalies
(Fig. 4.36) calculated as departures
from the long-term mean (1993–2019)
for the primary months of TC activity
in each hemisphere: June–November
2019 in the Northern Hemisphere (NH)
and November 2018–April 2019 in the
Southern Hemisphere (SH). Differences
between the 2019 and 2018 seasons are
also analyzed (Fig. 4.37). In any given
TC basin, TCHP anomalies can exhibit
large spatial and temporal variability
Fig. 4.36. Global anomalies of TCHP during 2019 computed as
described in the text. Boxes indicate the seven regions where TCs
linked with large mesoscale ocean feaoccur: from left to right, Southwest Indian, North Indian, West
tures, and short-term, interannual (e.g.,
North Pacific, Southeast Indian, South Pacific, East Pacific, and
El Niño-Southern Oscillation [ENSO]),
North Atlantic (shown as Gulf of Mexico and tropical Atlantic
and longer-term ocean variability, such
separately). The green lines indicate the trajectories of all TCs
as the Pacific Decadal Variability.
reaching at least Category-1 (1-min average wind ≥ 64 kts, 34
The 2019 TC season exhibited abovem s −1) and above during Nov 2018–Apr 2019 in the SH and Jun–
Nov 2019 in the NH. The numbers above each box correspond to
normal TCHP anomalies, which are
the number of Category-1 and above cyclones that travel within
favorable for TC development and
each box. The Gulf of Mexico conditions are shown in the inset
intensification, in most TC basins (Fig.
in the lower right corner.
4.36). TCHP values also increased in
most basins from 2018 to 2019 (Fig. 4.37), with notable warming of 20 kJ cm−2 with respect to 2018
observed at: (1) portions of the Gulf of Mexico associated with Loop Current dynamics; (2) large
areas in the South and North Indian Ocean basins; and (3) the western North Pacific basin Main
Development Region (MDR; Lin et al. 2014), i.e., east of the Philippines between 5°N and 20°N,
and 100°–170°E. Negative TCHP anomalies with respect to long-term conditions (Fig. 4.36) and
the 2018 season (Fig. 4.37) were only observed in the southeast Indian basin and near the eastern
portion of the South Pacific basin.
Both the North and southwest Indian Ocean basins exhibited considerably large TCHP values
in 2019 (Fig. 4.36), with anomalies as large as ~30 kJ cm−2 larger than the long-term average in
most of the North Indian basin, including the Bay of Bengal and Arabian Sea; and ~20 kJ cm−2 in
the southeast Indian basin. In particular, TCHP values were consistently larger than 90 kJ cm−2 in
the North Indian basin and 70 kJ cm−2 in the southeast basin (not shown). Consistent with these
substantially warmer conditions, both
the North and southwest Indian basins
were characterized by above-normal TC
activity. In the North Indian basin, the
2019 TC season was one of the most active on record (see section 4f5; Fig. 4.36).
In the southwest Indian basin, the 2019
TC season was the most active, costliest,
and deadliest on record (see section 4f6).
In the North Pacific, upper-ocean
thermal conditions are largely modulated by the state of ENSO (e.g., Lin et al.
2014, 2020; Zheng et al. 2015), which can
Fig. 4.37. TCHP difference between the 2019 and 2018 tropical
impact conditions both in the western
cyclone seasons (Jun–Nov in the NH and Nov–Apr in the SH).
and eastern North Pacific basins. During
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S228
AM UTC
4. THE TROPICS
the 2019 TC season, ENSO conditions switched from neutral in late 2018 to a weak El Niño in
early 2019 and back to neutral conditions by mid-2019. Associated with the neutral ENSO state,
the MDR within the western North Pacific basin exhibited TCHP values approximately 10–20 kJ
cm−2 larger than the long-term mean (Fig. 4.36) and ~20 kJ cm−2 larger than 2018 conditions (Fig.
4.37). These anomalies led to absolute TCHP values of 120 kJ cm−2 or larger over the MDR and of at
least 70 kJ cm−2 over most of this basin. Among the TCs that formed in this basin, Super Typhoon
Hagibis was a notable TC that experienced rapid intensification while traveling over areas with
TCHP of 100 kJ cm−2 or larger, where it became Category 5 (not shown). Another notable case is
Super Typhoon Halong, which also rapidly intensified over the MDR in areas with large TCHP
values (~100 kJ cm−2) in November, reaching a maximum wind speed of 155 kts (80 m s−1). Halong
was the most intense TC globally in 2019, but fortunately did not make landfall.
In the eastern North Pacific basin, TCHP values were consistently larger than long-term average conditions by 10–30 kJ cm−2 (Fig. 4.36). Compared to 2018 conditions, TCHP values were ~20
kJ cm−2 larger in 2019 over the central part of the basin between 180°W and 120°W and slightly
cooler by less than 10 kJ cm−2 closer to Central America. Of note, Major Hurricane Erick’s rapid
intensification west of 140°E was aided by the higher TCHP in this region.
Finally, in the North Atlantic basin, TCHP values were ~10 kJ cm−2 above the long-term average
(Fig. 4.36) in most parts of the basin, and warmer than 2018 in the central part of the basin between
60°W and 30°W and in the Gulf of Mexico, where the Loop Current extended northward and shed a
warm core ring. Associated with these conditions, the North Atlantic basin exhibited above-normal
hurricane activity for the fourth consecutive year. Higher TCHP values over the central portion
of the basin likely contributed to the rapid intensification of five of the total six hurricanes that
developed in that region of the North Atlantic in 2019 (Fig. 4.36). Hurricane Dorian, now regarded
as the most powerful hurricane on record for the Atlantic outside of the tropics (>23.5°N) in the
satellite era (since 1966), reached its peak intensity while traveling over areas with TCHP values
consistently above 70 kJ cm−2 and as large as 90 kJ cm−2 (not shown). These conditions are well
above the 50 kJ cm−2 minimum threshold required to support Atlantic hurricane intensification
(Mainelli et al. 2008). In addition to high TCHP values, Dorian traveled and intensified over areas
with low surface salinity values associated with the Amazon and Orinoco riverine plumes (not
shown). Areas with this type of low surface salinity are known for favoring TC intensification by
creating barrier layer conditions that suppress upper-ocean mixing, maintaining enthalpy fluxes
from the ocean into the hurricane (e.g., Balaguru et al. 2015; Domingues et al. 2015).
In summary, upper-ocean conditions conducive for TC development and intensification observed in 2019 were associated with higher-than-normal values of TCHP in most TC basins in
2019. Notable warming with respect to 2018 was also recorded in most basins, especially in the
Gulf of Mexico, the west North Pacific, and the Indian Ocean, particularly the Arabian Sea. These
warmer-than-usual conditions contributed to the more intense and above-normal TC activity in
most of these basins.
h. Indian Ocean dipole— L. Chen, J.-J. Luo, and A.D. Magee
The Indian Ocean dipole (IOD) is an inherent air–sea coupling mode in the tropical Indian
Ocean. It originates from local air–sea interaction in the Indian Ocean and/or the forcing associated with the El Niño–Southern Oscillation (ENSO) in the tropical Pacific (Saji et al. 1999; Luo et
al. 2010). Typically, IOD events develop in boreal summer, peak in boreal autumn, and terminate
rapidly in early boreal winter. During the late boreal spring to autumn 2019, a positive IOD (pIOD)
with extreme intensity occurred for the first time since 1997. Prior to the pIOD event in 1997, the
previous extreme pIOD event occurred in 1994 (Luo et al. 2007, 2008).
In the tropical Pacific, a weak El Niño occurred in the boreal winter of 2018/19 and returned to
neutral conditions by the boreal summer of 2019, but the sea surface anomalously warmed there
during the autumn of 2019 (Fig. 4.38c). In the tropical Indian Ocean, a weak pIOD occurred during
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S229
AM UTC
4. THE TROPICS
the autumn of 2018 but rapidly deteriorated
early in the winter of 2018/19 (Figs. 4.38a,b;
Chen and Luo 2019). For the first four
months of 2019 (Figs. 4.38 a,b), IOD-related
sea surface temperature (SST) anomalies
were near zero. Meanwhile, weak surface
easterly wind anomalies prevailed over
the central equatorial Indian Ocean during
the boreal winter of 2018/19 (partly due to
the remote influence of the weak El Niño).
These anomalies weakened to near zero in
March–April 2019 (Fig. 4.38b). Both pIODrelated SST anomalies (SSTA) and easterly
wind anomalies started to grow sharply
beginning in May 2019 (Fig. 4.38b). The
initial SSTA in the southeastern Indian
Ocean exhibited cooling along the south
coast of Java in May 2019, and then the
cooling signal gradually strengthened and
expanded toward the west coast of Sumatra and eastern equatorial Indian Ocean
(Figs. 4.39b–d). The positive SSTA in the
western equatorial Indian Ocean can be
traced back to the persistent warming SSTA Fig. 4.38. (a) Monthly anomalies of SST (°C; solid lines) and precipi−1
associated with the Indian Ocean basin tation (mm day ; dashed lines) in the eastern pole (IODE; 0°−10°S,
mode throughout the late 2018/19 boreal 90°−110°E; blue lines) and the western pole (IODW; 10°N−10°S,
50°−70°E; red lines) of the IOD. (b) As in (a), but for the IOD index
winter and early 2019 spring (Figs. 4.39a,b).
(measured by the SST difference between IODW and IODE, green
Then the anomalously warm SSTA in the line) and surface zonal wind anomaly (m s−1) in the central equatowestern Indian Ocean maintained its rial IO (Ucio; 5°N−5°S, 70°−90°E; black line). (c) As in (a), but for
intensity throughout June−October 2019 the SST anomalies in the Niño-3.4 region (5°N−5°S, 170°−120°W;
(Figs. 4.38a, 4.39c,d). The negative SSTA in black line) and the tropical IO (IOB; 10°N−20°S, 40°−120°E; red
the eastern pole started to grow from May line). Anomalies are relative to 1982−2019. (Sources: NOAA OISST
and continued to increase quickly until [Reynolds et al. 2002]; GPCP precipitation [Huffman et al. 2009];
and JRA-55 atmospheric reanalysis [Ebita et al. 2011].)
October (Figs. 4.38a, 4.39b–d).
Since the pIOD started to grow in May,
positive precipitation anomalies developed near the western pole with dry anomalies near the
eastern pole (Fig. 4.38a). This pattern indicates that the precipitation anomalies in the equatorial Indian Ocean were well coupled with the easterly wind anomalies in the central equatorial
Indian Ocean and SSTA throughout the development of this pIOD event. Before the development
of the IOD-related SSTA, a positive precipitation anomaly occurred near the eastern pole of the
IOD in April (Fig. 4.38a), which might be associated with atmospheric high-frequency “noise.”
This positive precipitation near the eastern pole may have played a role in inducing the initial
southeasterly wind anomaly along the south coast of Java and southwest coast of Sumatra in
April, which caused the positive Bjerknes feedback (Bjerknes 1969) over the following months,
ultimately leading to the pIOD event.
The pIOD in 2019, whose Dipole Mode Index (DMI) attained ~2.1°C in October 2019, exhibited
the greatest magnitude in the observational record since 1997 (Fig. 4.40c). The surface zonal
wind anomaly in the central equatorial Indian Ocean related to the pIOD in 2019 ranked only
second to the extraordinary pIOD event in 1997 (Fig. 4.40d). In contrast to the extreme pIOD in
1997 that occurred with an extremely strong El Niño, the 2019 pIOD event was accompanied by a
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S230
AM UTC
4. THE TROPICS
neutral ENSO state in the tropical Pacific (Fig.
4.40e). There is no clear evidence supporting
that remote processes in the tropical Pacific
played an essential role in generating the
pIOD event in 2019. Rather, it appears that
the development of this extreme pIOD event
was largely generated by local processes in
the Indian Ocean. This is different from the
majority of pIOD events, which have often
co-occurred with El Niño events (e.g., 6 out
of 10 past pIOD events since 1980 co-occurred
with El Niño, as shown in Fig. 4.40). It is
also worth noting that the positive SSTA in
the western pole reached ~0.8°C and the
negative SSTA in the eastern pole reached ~
−1.3°C in late autumn of 2019 (Figs. 4.40a,b).
The former ranked first among all historical
pIOD events, which may be traced back to the
continuous enhancement of tropical Indian
Ocean warming during recent decades (Luo
et al. 2012).
Impacts associated with this strong pIOD
event were widespread and preconditioned a
number of events across the globe. In Australia, the austral spring of 2019 was the driest
on record, and along with a particularly dry
austral winter, fueled an unusually early
start to the bushfire season (see section 7h4
and Sidebar 7.6 for details). Fires continued to
(contours:
burn into early 2020. This strong pIOD event Fig. 4.39. SST (°C; colored scale) and precipitation
−8, −6, −4, −2, −1, 0, 1, 2, 4, 6 mm day−1; solid/dashed/bold
resulted in significant flooding in eastern Afcurves denote positive/negative/zero values) anomalies
rica, with some regions in the Horn of Africa during (a) Dec 2018−Feb 2019, (b) Mar−May 2019, (c) Jun−
seeing up to 300% above-average rainfall be- Aug 2019, and (d) Sep−Nov 2019. Anomalies are relative to
tween October and mid-November, ranking 1982–2019. (Sources: NOAA OISST [Reynolds et al. 2002];
among the wettest rainfall seasons in east GPCP precipitation analysis [Huffman et al. 2009].)
Africa in at least 40 years. Approximately
300 people died, and a further 3.4 million people were affected across the region (Famine Early
Warning Systems Network 2020). The strong pIOD has also been associated with the ongoing
drought and smoke haze in Indonesia.
In summary, the strongest pIOD event since 1997 occurred in October 2019. During the course
of the growth of this pIOD event, equatorial zonal wind, precipitation, and SST anomalies in the
equatorial Indian Ocean all coupled well with each other. As shown in Fig. 4.40f, in April–May,
low-level southeasterly anomalies prevailed near the south coasts of Java and Sumatra, and the
negative SSTA near the eastern pole started to grow rapidly. Concurrently, weak, warm SSTA
persisted near the western pole (which may be associated with the prolonged Indian Ocean basin warming during early 2019). As a result, low-level easterly wind anomalies started to grow in
the central equatorial Indian Ocean in May. Through the positive Bjerknes feedback, the pIOD
event was generated, and the corresponding anomaly signal peaked during the autumn of 2019.
In December, the IOD-related SST, precipitation, and wind anomalies quickly deteriorated. The
extreme pIOD event in 2019 seems to have originated from air–sea feedback processes in the
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S231
AM UTC
4. THE TROPICS
Indian Ocean itself, rather than being induced by the remote influence of El Niño. Interestingly,
such a unique development feature of the pIOD in 2019 differs from many of the past pIOD events
that co-occurred with El Niño events.
Fig. 4.40. Monthly SST anomalies in the (a) IODW, (b) IODE, and (c) the Dipole Mode Index (DMI, the SST anomaly difference
between the IODW and the IODE) during 11 pIOD events since the 1980s. (d) As in (c) but for the surface zonal wind anomaly
(m s−1) in the central equatorial Indian Ocean (70°−90°E; 5°N−5°S). (e)−(f) As in (a)−(b), but for the monthly SST anomalies
in the Niño-3.4 region (170°−120°W; 5°N−5°S) and the tropical Indian Ocean basin (40°−120°E; 20°N−20°S). (Sources: NOAA
OISST [Reynolds et al. 2002]; GPCP precipitation [Huffman et al. 2009]; and JRA-55 atmospheric reanalysis [Ebita et al. 2011].)
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S232
AM UTC
4. THE TROPICS
APPENDIX: Acronym List
ACE
AEJ
AMO
ASO
CMORPH
CNP
CPC
DJF
DMI
ENP
ENSO
GPI
HTC
HURDAT2
IBTrACS
IOD
ITCZ
JAS
JASO
JFM
JJA
JMA
JTWC
MAM
MDR
MJJ
MJO
NDJ
NH
NIO
OLR
OND
ONI
PAGASA
pIOD
RMM
RMSCs
SAM
SH
SIO
SON
SPCZ
SPEArTC
SSHWS
SSHWS
SST
Accumulated Cyclone Energy
African Easterly Jet
Atlantic Multidecadal Oscillation
August-October
Climate Prediction Center morphing method
central North Pacific
Climate Prediction Center
December-February
Dipole Mode Index
eastern North Pacific
El Niño-Southern Oscillation
genesis potential index
hurricanes/typhoons/cyclones
(National Hurricane Center’s) Hurricane Database
International Best Track Archive for Climate Stewardship
Indian Ocean dipole
Intertropical Convergence Zone
July-September
July-October
January-March
June-August
Japan Meteorological Agency
Joint Typhoon Warning Center
March-May
Main Development Region
May-July
Madden Julian Oscillation
November-January
Northern Hemisphere
North Indian Ocean
Outgoing Longwave Radiation
October-December
Oceanic Niño Index
Philippine Atmospheric, Geophysical and Astronomical Services
Administration
positive Indian Ocean dipole
Real-time Multivariate MJO
Regional Specialized Meteorological Centers
Southern Annular Mode
Southern Hemisphere
South Indian Ocean
September-November
South Pacific Convergence Zone
Southwest Pacific Enhanced Archive of Tropical Cyclones
Saffir-Simpson
Saffir-Simpson Hurricane Wind Scale
sea surface temperature
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S233
AM UTC
4. THE TROPICS
TC
TCHP
TD
TS
TWS
WMO
WNP
WWBs
tropical cyclone
Tropical Cyclone Heat Potential
tropical depression
tropical storm
trade wind surges
World Meteorological Organization
western North Pacific
westerly wind bursts
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Unauthenticated | Downloaded 12/04/22 09:14S234
AM UTC
4. THE TROPICS
References
Aiyyer, A., and J. Molinari, 2008: MJO and tropical cyclogenesis in the Gulf of
Mexico and eastern Pacific: Case study and idealized numerical modeling. J.
Atmos. Sci., 65, 2691–2704, https://doi.org/10.1175/2007JAS2348.1.
Avila, L. A., 2019: Hurricane Lorena. National Hurricane Center Tropical Cyclone
Rep., 20 pp., www.nhc.noaa.gov/data/tcr/EP152019_Lorena.pdf.
—, S. A. Stewart, R. Berg, and A. B. Hagen, 2020: Hurricane Dorian. National
Hurricane Center Tropical Cyclone Rep., 74 pp., www.nhc.noaa.gov/data/
tcr/AL052019_Dorian.pdf.
Balaguru, K., G. R. Foltz, L. R. Leung, E. A. D’Asaro, K. A. Emanuel, H. Liu, and
S. E. Zedle, 2015: Dynamic potential intensity: An improved representation
of the ocean’s impact on tropical cyclones. Geophys. Res. Lett., 42, 6739–
6746, https://doi.org/10.1002/2015GL064822.
Banzon, V. F., and R. W. Reynolds, 2013: Use of WindSat to extend a microwavebased daily optimum interpolation sea surface temperature time series. J.
Climate, 26, 2557–2562, https://doi.org/10.1175/JCLI-D-12-00628.1.
Baxter, S., S. Weaver, J. Gottschalck, and Y. Xue, 2014: Pentad evolution of
wintertime impacts of the Madden–Julian Oscillation over the contiguous
United States. J. Climate, 27, 7356–7367, https://doi.org/10.1175/JCLID-14-00105.1.
—, C. Schreck, and G. D. Bell, 2019: ttTropical intraseasonal activity [in
“State of the Climate in 2018”]. Bull. Amer. Meteor. Soc., 100 (9), S105–
S107, https://doi.org/10.1175/2019BAMSStateoftheClimate.1.
Behringer, D. W., M. Ji, and A. Leetmaa, 1998: An improved coupled model
for ENSO prediction and implications for ocean initialization. Part I: The
ocean data assimilation system. Mon. Wea. Rev., 126, 1013–1021, https://
doi.org/10.1175/1520-0493(1998)126<1013:AICMFE>2.0.CO;2.
Bell, G. D., and M. Chelliah, 2006: Leading tropical modes associated with interannual and multi-decadal fluctuations in North Atlantic hurricane activity. J.
Climate, 19, 590–612, https://doi.org/10.1175/JCLI3659.1.
,Bull. Amer. Meteor. Soc.100—, and Coauthors, 2000: The 1999 North
Atlantic hurricane season: A climate perspective [in “State of the Climate in 1999”]. Bull. Amer. Meteor. Soc., 81 (6), S1–S68, https://doi.
org/10.1175/1520-0477(2000)81[s1:CAF]2.0.CO;2.
—, and Coauthors, 2004: The 2003 North Atlantic Hurricane Season: A climate perspective [in “State of the Climate in 2003”]. Bull. Amer. Meteor.
Soc., 85 (6), S1–S68, https://doi.org/10.1175/1520-0477-85.6s.S1.
—, and Coauthors, 2006: The 2005 North Atlantic Hurricane Season: A climate perspective [in “State of the Climate in 2005”]. Bull. Amer. Meteor.
Soc., 86 (6), S1–S68, https://doi.org/10.1175/1520-0477-87.6.S1.
—, and Coauthors, 2017: [The tropics] the 2016 North Atlantic hurricane
season: A climate perspective [in “State of the Climate in 2016”]. Bull.
Amer. Meteor. Soc., 98 (8), S109–S112, https://doi.org/10.1175/2017BAM
SStateoftheClimate.1.
—, and Coauthors, 2018: [The tropics] the 2017 North Atlantic hurricane
season: A climate perspective [in “State of the Climate in 2017”]. Bull.
Amer. Meteor. Soc., 99 (8), S114–S118, https://doi.org/10.1175/2018BAM
SStateoftheClimate.1.
—, M. S. Halpert, and M. L’Heureux, 2019: ENSO and the tropical Pacific [in
“State of the Climate in 2018”]. Bull. Amer. Meteor. Soc., 100 (9), S101–
S104, https://doi.org/10.1175/2019BAMSStateoftheClimate.1.
Bjerknes, J., 1969: Atmospheric teleconnections from the equatorial Pacific. Mon. Wea. Rev., 97, 163–172, https://doi.org/10.1175/15200493(1969)097<0163:ATFTEP>2.3.CO;2.
Blake, E. S., 2019: Tropical Storm Narda. National Hurricane Center Tropical
Cyclone Rep., 14 pp., www.nhc.noaa.gov/data/tcr/EP162019_Narda.pdf.
Camargo, S. J., K. A. Emanuel, and A. H. Sobel, 2007: Use of a genesis potential
index to diagnose ENSO effects on tropical cyclone genesis. J. Climate, 20,
4819–4834, https://doi.org/10.1175/JCLI4282.1.
Cassou, C., 2008: Intraseasonal interaction between the Madden Julian oscillation and the North Atlantic oscillation. Nature, 455, 523–527. Letter,
https://doi.org/10.1038/nature07286.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Chen, L., and J.-J. Luo, 2019: Indian Ocean dipole [in “State of the Climate in
2018”]. Bull. Amer. Meteor. Soc., 100 (9), S138–S140, https://doi.org/10.11
75/2019BAMSStateoftheClimate.1.
Chia, H. H., and C. F. Ropelewski, 2002: The interannual variability in the genesis location of tropical cyclones in the northwest Pacific. J. Climate, 15,
2934–2944, https://doi.org/10.1175/1520-0442(2002)015<2934:TIVITG>2
.0.CO;2.
CDare, R. A., and J. L. McBride, 2011: Sea surface temperature response to tropical cyclones. Mon. Wea. Rev., 139, 3798–3808, https://doi.org/10.1175/
MWR-D-10-05019.1.
Diamond, H. J., and J. A. Renwick, 2015: The climatological relationship between tropical cyclones in the southwest Pacific and the Madden–Julian
Oscillation. Int. J. Climatol., 35, 676–686, https://doi.org/10.1002/joc.4012.
—, and C. J. Schreck III, Eds, 2017: The tropics [in “State of the Climate in
2016”]. Bull. Amer. Meteor. Soc., 98 (8), S93–S128, https://doi.org/10.1175/
2017BAMSStateoftheClimate.1..
—, and —, Eds, 2018: The tropics [in “State of the Climate in 2017”]. Bull.
Amer. Meteor. Soc., 99 (8), S101–S141, https://doi.org/10.1175/2018BAMS
StateoftheClimate.1..
—, and —, Eds, 2019: The tropics [in “State of the Climate in 2018”]. Bull.
Amer. Meteor. Soc., 100 (9), S101–S140, https://doi.org/10.1175/2019BAM
SStateoftheClimate.1..
—, A. M. Lorrey, K. R. Knapp, and D. H. Levinson, 2012: Development of an
enhanced tropical cyclone tracks database for the southwest Pacific from
1840 to 2011. Int. J. Climatol., 32, 2240–2250, https://doi.org/10.1002/
joc.2412.
Domingues, R., G. Goni, F. Bringas, G. Halliwell, S.-K. Lee, H.-S. Kim, J. Morell,
and J. Dong, 2015: Upper ocean response to Hurricane Gonzalo (2014): Results from unique observations by underwater gliders. Geophys. Res. Lett.,
42, 7131–7138, https://doi.org/10.1002/2015GL065378.
Dube, S. K., D. Rao, P. C. Sinha, T. S. Murty, and N. Bahuluyan, 1997: Storm
surge in the Bay of Bengal and Arabian Sea: The problem and its prediction.
Mausam, 48, 288–304.
Ebita, A., and Coauthors, 2011: The Japanese 55-year reanalysis “JRA-55”: An
interim report. SOLA, 7, 149–152, https://doi.org/10.2151/SOLA.2011-038.
Emanuel, K. A., and D. S. Nolan, 2004: Tropical cyclone activity and the
global climate system. Proc. 26th Conf. on Hurricanes and Tropical Meteorology, Miami, FL, Amer. Meteor. Soc., 10A.2, https://ams.confex.com/
ams/26HURR/techprogram/paper_75463.htm.
Enfield, D. B., and A. M. Mestas-Nuñez, 1999: Multiscale variabilities in global
sea surface temperatures and their relationships with tropospheric climate patterns. J. Climate, 12, 2719–2733, https://doi.org/10.1175/15200442(1999)012<2719:MVIGSS>2.0.CO;2.
Famine Early Warning Systems Network, 2020: 2019 short rains in East Africa
among the wettest on historical record,. FEWS NET, , https://fews.net/eastafrica/special-report/january-29-2020.
Frank, W. M., and P. E. Roundy, 2006: The role of tropical waves in tropical cyclogenesis. Mon. Wea. Rev., 134, 2397–2417, https://doi.org/10.1175/
MWR3204.1.
Goldenberg, S. B., and L. J. Shapiro, 1996: Physical mechanisms for the association
of El Niño and West African rainfall with Atlantic major hurricane activity. J. Climate, 9, 1169–1187, https://doi.org/10.1175/1520-0442(1996)009<1169:PM
FTAO>2.0.CO;2.
—, C. W. Landsea, A. M. Mestas-Nuñez, and W. M. Gray, 2001: The recent
increase in Atlantic hurricane activity: Causes and implications. Science,
293, 474–479, https://doi.org/10.1126/science.1060040.
Goni, G. J., and Coauthors, 2009: Applications of satellite-derived ocean measurements to tropical cyclone intensity forecasting. Oceanography, 22,
190–197, https://doi.org/10.5670/oceanog.2009.78.
Unauthenticated | Downloaded 12/04/22 09:14S235
AM UTC
4. THE TROPICS
—, and Coauthors, 2017: Autonomous and Lagrangian Ocean observations
for Atlantic tropical cyclone studies and forecasts. Oceanography, 30, 92–
103, https://doi.org/10.5670/oceanog.2017.227.
Gray, W. M., 1990: Strong association between West African rainfall and
U.S. landfall of intense hurricanes. Science, 249, 1251–1256, https://doi.
org/10.1126/science.249.4974.1251.
Guo, Y., X. Jiang, and D. E. Waliser, 2014: Modulation of the convectively coupled Kelvin waves over South America and the tropical Atlantic Ocean in association with the Madden–Julian oscillation. J. Atmos. Sci., 71, 1371–1388,
https://doi.org/10.1175/JAS-D-13-0215.1.
Halpert, M. S., and C. F. Ropelewski, 1992: Surface temperature patterns associated with the southern oscillation. J. Climate, 5, 577–593, https://doi.
org/10.1175/1520-0442(1992)005<0577:STPAWT>2.0.CO;2.
Hastenrath, S., 1990: Decadal-scale changes of the circulation in the tropical
Atlantic sector associated with Sahel drought. Int. J. Climatol., 10, 459–
472, https://doi.org/10.1002/joc.3370100504.
Hendon, H., C. Zhang, and J. Glick, 1999: Interannual variation of the Madden–
Julian oscillation during austral summer. J. Climate, 12, 2538–2550, https://
doi.org/10.1175/1520-0442(1999)012<2538:IVOTMJ>2.0.CO;2.
Huang, B., and Coauthors, 2017: Extended Reconstructed Sea Surface Temperature, version 5 (ERSSTv5): Upgrades, validations, and intercomparisons. J.
Climate, 30, 8179–8205, https://doi.org/10.1175/JCLI-D-16-0836.1.
Huffman, G. J., R. F. Adler, D. T. Bolvin, and G. Gu, 2009: Improving the global
precipitation record: GPCP version 2.1. Geophys. Res. Lett., 36, L17808, ,
https://doi.org/10.1029/2009GL040000.
India Meteorological Department, 2020: Statement on climate of India during
2019. India Meteorological Department Climate Research and Services.A,
accessed 14 January 2020, https://mausam.imd.gov.in/backend/assets/
press_release_pdf/Statement_on_Climate_of_India_during_2019.pdf.
Joyce, R. J., J. E. Janowiak, P. A. Arkin, and P. Xie, 2004: CMORPH: A method
that produces global precipitation estimates from passive microwave and
infrared data at high spatial and temporal resolution. J. Hydrometeor., 5,
487–503, https://doi.org/10.1175/1525-7541(2004)005<0487:CAMTPG>2.
0.CO;2.
Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437–471, https://doi.org/10.1175/15200477(1996)077<0437:TNYRP>2.0.CO;2.
Kayano, M., and V. Kousky, 1999: Intraseasonal (30-60 day) variability in the
global tropics: Principal modes and their evolution. Tellus, 51A, 373–386,
https://doi.org/10.3402/tellusa.v51i3.13459.
Kiladis, G. N., and K. M. Weickmann, 1992: Circulation anomalies associated
with tropical convection during northern winter. Mon. Wea. Rev., 120,
1900–1923, https://doi.org/10.1175/1520-0493(1992)120<1900:CAAWTC
>2.0.CO;2.
—, K. H. Straub, and P. T. Haertel, 2005: Zonal and vertical structure of
the Madden–Julian oscillation. J. Atmos. Sci., 62, 2790–2809, https://doi.
org/10.1175/JAS3520.1.
Kiladis, G. N., M. C. Wheeler, P. T. Haertel, K. H. Straub, and P. E. Roundy, 2009:
Convectively coupled equatorial waves. Rev. Geophys., 47, RG2003, https://
doi.org/10.1029/2008RG000266.
Klotzbach, P. J., 2010: On the Madden–Julian oscillation–Atlantic hurricane relationship. J. Climate, 23, 282–293, https://doi.org/10.1175/2009JCLI2978.1.
Knaff, J. A., C. R. Sampson, and K. D. Musgrave, 2018: An operational rapid intensification prediction aid for the western North Pacific. Wea. Forecasting, 33,
799–811, https://doi.org/10.1175/WAF-D-18-0012.1.
Knapp, K. R., M. C. Kruk, D. H. Levinson, H. J. Diamond, and C. J. Neumann,
2010: The International Best Track Archive for Climate Stewardship (IBTrACS): Unifying tropical cyclone data. Bull. Amer. Meteor. Soc., 91, 363–
376, https://doi.org/10.1175/2009BAMS2755.1.
—, J. A. Knaff, C. R. Sampson, G. M. Riggio, and A. D. Schnapp, 2013: A pressure-based analysis of the historical western North Pacific tropical cyclone
intensity record. Mon. Wea. Rev., 141, 2611–2631, https://doi.org/10.1175/
MWR-D-12-00323.1.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Knutson, T. R., and K. M. Weickmann, 1987: 30–60 day atmospheric oscillations: Composite life cycles of convection and circulation anomalies. Mon. Wea. Rev., 115, 1407–1436, https://doi.org/10.1175/15200493(1987)115<1407:DAOCLC>2.0.CO;2.
Kousky, V. E., and M. T. Kayano, 1994: Principal modes of outgoing longwave
radiation and 250-mb circulation for the South American sector. J. Climate,
7, 1131–1143, https://doi.org/10.1175/1520-0442(1994)007<1131:PMOOLR
>2.0.CO;2.
Krishnamurti, T. N., and D. Subrahmanyam, 1982: The 30–50 day mode
at 850 mb during MONEX. J. Atmos. Sci., 39, 2088–2095, https://doi.
org/10.1175/1520-0469(1982)039<2088:TDMAMD>2.0.CO;2.
Kruk, M. C., and C. J. Schreck, 2019: tEastern North Pacific basin [in “State of
the Climate in 2018”]. Bull. Amer. Meteor. Soc., 100 (9), S121–S124, https://
doi.org/10.1175/2019BAMSStateoftheClimate.1.
Landsea, C. W., and J. L. Franklin, 2013: Atlantic hurricane database uncertainty and presentation of a New database format. Mon. Wea. Rev., 141,
3576–3592, https://doi.org/10.1175/MWR-D-12-00254.1.
—, W. M. Gray, P. W. Mielke, and K. J. Berry, 1992: Long-term variations of western Sahelian monsoon rainfall and intense U.S. landfalling hurricanes. J. Climate, 5, 1528–1534, https://doi.org/10.1175/15200442(1992)005<1528:LTVOWS>2.0.CO;2.
—, B. A. Harper, K. Hoarau, and J. A. Knaff, 2006: Can we detect trends in
extreme tropical cyclones? Science, 313, 452–454, https://doi.org/10.1126/
science.1128448.
—, G. A. Vecchi, L. Bengtsson, and T. R. Knutson, 2010: Impact of duration thresholds on Atlantic tropical cyclone counts. J. Climate, 23, 2508–2519, https://
doi.org/10.1175/2009JCLI3034.1.
Lau, W. K.-M., and D. E. Waliser, 2012: Intraseasonal Variability in the Atmosphere-Ocean Climate System. Springer, 642 pp.
Leipper, D. F., and D. Volgenau, 1972: Hurricane heat potential of the Gulf of
Mexico. J. Phys. Oceanogr., 2, 218–224, https://doi.org/10.1175/15200485(1972)002<0218:HHPOTG>2.0.CO;2.
Liebmann, B., and C. A. Smith, 1996: Description of a complete (interpolated)
outgoing longwave radiation dataset. Bull. Amer. Meteor. Soc., 77, 1275–
1277, https://doi.org/10.1175/1520-0477-77.6.1274.
Lin, H., G. Brunet, and J. Derome, 2009: An observed connection between the
North Atlantic oscillation and the Madden–Julian oscillation. J. Climate, 22,
364–380, https://doi.org/10.1175/2008JCLI2515.1.
Lin, I.-I., and Coauthors, 2013: An ocean coupling potential intensity index
for tropical cyclones. Geophys. Res. Lett., 40, 1878–1882, https://doi.
org/10.1002/grl.50091.
—, I.-F. Pun, and C.-C. Lien, 2014: “Category-6” super-typhoon Haiyan in
global warming hiatus: Contribution from subsurface ocean warming. Geophys. Res. Lett., 41, 8547–8553, https://doi.org/10.1002/2014GL061281.
—, and Coauthors, 2020: ENSO and tropical cyclones. El Niño Southern Oscillation in a Changing Climate, Geophys. Monogr., Vol. 254, Amer. Geophys.
Union, 377–408.
Luo, J.-J., S. Masson, S. Behera, and T. Yamagata, 2007: Experimental forecasts of the Indian Ocean dipole using a coupled OAGCM. J. Climate, 20,
2178–2190, https://doi.org/10.1175/JCLI4132.1.
—, S. Behera, Y. Masumoto, H. Sakuma, and T. Yamagata, 2008: Successful
prediction of the consecutive IOD in 2006 and 2007. Geophys. Res. Lett., 35,
L14S02, https://doi.org/10.1029/2007GL032793.
—, R. Zhang, S. K. Behera, Y. Masumoto, F.-F. Jin, R. Lukas, and T. Yamagata,
2010: Interaction between El Niño and extreme Indian Ocean dipole. J.
Climate, 23, 726–742, https://doi.org/10.1175/2009JCLI3104.1.
—, W. Sasaki, and Y. Masumoto, 2012: Indian Ocean warming modulates
Pacific climate change. Proc. Natl. Acad. Sci. USA, 109, 18 701–18 706, ,
https://doi.org/10.1073/pnas.1210239109.
Madden, R., and P. Julian, 1971: Detection of a 40–50 day oscillation in the
zonal wind in the tropical Pacific. J. Atmos. Sci., 28, 702–708, https://doi.
org/10.1175/1520-0469(1971)028<0702:DOADOI>2.0.CO;2.
Unauthenticated | Downloaded 12/04/22 09:14S236
AM UTC
4. THE TROPICS
—, and —, 1972: Description of global-scale circulation cells in the tropics with a 40–50 day period. J. Atmos. Sci., 29, 1109–1123, https://doi.
org/10.1175/1520-0469(1972)029<1109:DOGSCC>2.0.CO;2.
—, and —, 1994: Observations of the 40–50-day tropical oscillation:
A review. Mon. Wea. Rev., 122, 814–837, https://doi.org/10.1175/15200493(1994)122<0814:OOTDTO>2.0.CO;2.
Mainelli, M., M. DeMaria, L. Shay, and G. Goni, 2008: Application of oceanic heat content estimation to operational forecasting of recent Atlantic category 5 hurricanes. Wea. Forecasting, 23, 3–16, https://doi.
org/10.1175/2007WAF2006111.1.
Maloney, E. D., and D. L. Hartmann, 2001: The Madden–Julian oscillation,
barotropic dynamics, and North Pacific tropical cyclone formation. Part I:
Observations. J. Atmos. Sci., 58, 2545–2558, https://doi.org/10.1175/15200469(2001)058<2545:TMJOBD>2.0.CO;2.
Masters, J., 2019: Africa’s Hurricane Katrina: Tropical Cyclone Idai causes an
extreme catastrophe. Weather Underground, accessed 14 January 2020,
www.wunderground.com/cat6/Africas-Hurricane-Katrina-Tropical-Cyclone-Idai-Causes-Extreme-Catastrophe.
Mo, K. C., 2000: The association between intraseasonal oscillations and tropical storms in the Atlantic basin. Mon. Wea. Rev., 128, 4097–4107, https://
doi.org/10.1175/1520-0493(2000)129<4097:TABIOA>2.0.CO;2.
—, and V. E. Kousky, 1993: Further analysis of the relationship between circulation anomaly patterns and tropical convection. J. Geophys. Res., 98,
5103–5113, https://doi.org/10.1029/92JD02952.
Münnich, M., and J. D. Neelin, 2005: Seasonal influence of ENSO on the Atlantic ITCZ and equatorial South America. Geophys. Res. Lett., 32, L21709,
https://doi.org/10.1029/2005GL023900.
Nakazawa, T., and S. Hoshino, 2009: Intercomparison of Dvorak parameters in
the tropical cyclone datasets over the western North Pacific. SOLA, 5, 33–36,
https://doi.org/10.2151/SOLA.2009-009.
NOAA, 2019: Synoptic discussion, November 2019., NOAA/NCEI, www.ncdc.
noaa.gov/sotc/synoptic/201911#November-16-22.
Nobre, P., and J. Shukla, 1996: Variations of sea surface temperature, wind
stress and rainfall over the tropical Atlantic and South America. J. Climate,
9, 2464–2479, https://doi.org/10.1175/1520-0442(1996)009<2464:VOSST
W>2.0.CO;2.
PEAC, 2019: Pacific ENSO update. Pacific ENSO Applications Center, www.
weather.gov/peac/update.
Pezza, A. B., and C. A. S. Coelho, 2019: The Atlantic Intertropical convergence
zone [in “State of the Climate in 2018”]. Bull. Amer. Meteor. Soc., 100 (9),
S109–S110, https://doi.org/10.1175/2019BAMSStateoftheClimate.1.
Raga, G. B., B. Bracamontes-Ceballos, L. Farfán, and R. Romero-Centeno, 2013:
Landfalling tropical cyclones on the Pacific coast of Mexico: 1850–2010.
Atmósfera, 26, 209–220, https://doi.org/10.1016/S0187-6236(13)71072-5.
Reynolds, R. W., N. A. Rayner, T. M. Smith, D. C. Stokes, and W. Wang, 2002:
An improved in situ and satellite SST analysis for climate. J. Climate, 15,
1609–1625, https://doi.org/10.1175/1520-0442(2002)015<1609:AIISAS>2.
0.CO;2.
Riddle, E., M. Stoner, N. Johnson, M. L’Heureux, D. Collins, and S. Feldstein, 2012:
The impact of the MJO on clusters of wintertime circulation anomalies
over the North American region. Climate Dyn., 40, 1749–1766, https://doi.
org/10.1007/S00382-012-1493-Y.
Ropelewski, C. F., and M. S. Halpert, 1989: Precipitation patterns associated
with the high index phase of the Southern Oscillation. J. Climate, 2, 268–
284,
https://doi.org/10.1175/1520-0442(1989)002<0268:PPAWTH>2.0.
CO;2.
Saha, S., and Coauthors, 2014: The NCEP Climate Forecast System version 2. J.
Climate, 27, 2185–2208, https://doi.org/10.1175/JCLI-D-12-00823.1.
Saji, N. H., B. N. Goswami, P. N. Vinayachandran, and T. Yamagata, 1999: A
dipole mode in the tropical Indian ocean. Nature, 401, 360–363, https://
doi.org/10.1038/43854.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Schneider, T., T. Bischoff, and G. H. Haug, 2014: Migrations and dynamics
of the intertropical convergence zone. Nature, 513, 45–53, https://doi.
org/10.1038/nature13636.
Schreck, C. J., 2015: Kelvin waves and tropical cyclogenesis: A global survey. Mon. Wea. Rev., 143, 3996–4011, https://doi.org/10.1175/MWRD-15-0111.1.
—, 2016: Convectively coupled Kelvin waves and tropical cyclogenesis in a
semi-Lagrangian framework. Mon. Wea. Rev., 144, 4131–4139, https://doi.
org/10.1175/MWR-D-16-0237.1.
—, and J. Molinari, 2011: Tropical cyclogenesis associated with Kelvin waves
and the Madden–Julian oscillation. Mon. Wea. Rev., 139, 2723–2734,
https://doi.org/10.1175/MWR-D-10-05060.1.
—, —, and A. Aiyyer, 2012: A global view of equatorial waves and tropical cyclogenesis. Mon. Wea. Rev., 140, 774–788, https://doi.org/10.1175/
MWR-D-11-00110.1.
—, J. M. Cordeira, and D. Margolin, 2013: Which MJO events affect North
American temperatures? Mon. Wea. Rev., 141, 3840–3850, https://doi.
org/10.1175/MWR-D-13-00118.1.
—, K. R. Knapp, and J. P. Kossin, 2014: The impact of best track discrepancies on global tropical cyclone climatologies using IBTrACS. Mon. Wea. Rev.,
142, 3881–3899, https://doi.org/10.1175/MWR-D-14-00021.1.
—, H.-T. Lee, and K. R. Knapp, 2018: HIRS outgoing longwave radiation—daily climate data record: Application toward identifying tropical
subseasonal variability. Remote Sens., 10, 1325, https://doi.org/10.3390/
rs10091325.
Shay, L. K., G. J. Goni, and P. G. Black, 2000: Effects of a warm oceanic feature on Hurricane Opal. Mon. Wea. Rev., 128, 1366–1383, https://doi.
org/10.1175/1520-0493(2000)128<1366:EOAWOF>2.0.CO;2.
Smith, T. M., R. W. Reynolds, T. C. Peterson, and J. Lawrimore, 2008: Improvements to NOAA’s historical merged land–ocean surface temperature analysis (1880–2006). J. Climate, 21, 2283–2296, https://doi.
org/10.1175/2007JCLI2100.1.
Song, J.-J., Y. Wang, and L. Wu, 2010: Trend discrepancies among three best
track data sets of western North Pacific tropical cyclones. J. Geophys. Res.,
115, D12128, https://doi.org/10.1029/2009JD013058.
Stewart, S. R., 2019: Tropical Storm Priscilla. National Hurricane Center Tropical
Cyclone Rep., 9 pp., www.nhc.noaa.gov/data/tcr/EP192019_Priscilla.pdf.
Vecchi, G. A., and B. J. Soden, 2007: Effect of remote sea surface temperature
change on tropical cyclone potential intensity. Nature, 450, 1066–1070,
https://doi.org/10.1038/NATURE06423.
Ventrice, M. J., C. D. Thorncroft, and M. A. Janiga, 2012a: Atlantic tropical
cyclogenesis: A three-way interaction between an African easterly wave,
diurnally varying convection, and a convectively coupled atmospheric Kelvin wave. Mon. Wea. Rev., 140, 1108–1124, https://doi.org/10.1175/MWRD-11-00122.1.
—, —, and C. J. Schreck, 2012b: Impacts of convectively coupled Kelvin
waves on environmental conditions for Atlantic tropical cyclogenesis. Mon.
Wea. Rev., 140, 2198–2214, https://doi.org/10.1175/MWR-D-11-00305.1.
Villarini, G., G. A. Vecchi, T. R. Knutson, and J. A. Smith, 2011: Is the recorded
increase in short duration North Atlantic tropical storms spurious? J. Geophys. Res., 116, D10114, https://doi.org/10.1029/2010JD015493.
Vincent, D. G., 1994: The South Pacific Convergence Zone (SPCZ): A review. Mon. Wea. Rev., 122, 1949–1970, https://doi.org/10.1175/15200493(1994)122<1949:TSPCZA>2.0.CO;2.
Waliser, D. E., and C. Gautier, 1993: A satellite-derived climatology of
the ITCZ. J. Climate, 6, 2162–2174, https://doi.org/10.1175/15200442(1993)006<2162:ASDCOT>2.0.CO;2.
Wang, B., and Q. Ding, 2008: Global monsoon: Dominant mode of annual variation in the tropics. Dyn. Atmos. Ocean, 44, 165–183, https://doi.org/10.1016/J.
DYNATMOCE.2007.05.002.
—, J. Liu, H. J. Kim, P. J. Webster, and S. Y. Yim, 2012: Recent change of the
global monsoon precipitation (1979–2008). Climate Dyn., 39, 1123–1135,
https://doi.org/10.1007/s00382-011-1266-z.
Unauthenticated | Downloaded 12/04/22 09:14S237
AM UTC
4. THE TROPICS
Wheeler, M., and G. N. Kiladis, 1999: Convectively coupled equatorial
waves: Analysis of clouds and temperature in the wavenumber-frequency domain. J. Atmos. Sci., 56, 374–399, https://doi.org/10.1175/15200469(1999)056<0374:CCEWAO>2.0.CO;2.
—, and H. H. Hendon, 2004: An all-season real-time multivariate MJO index:
Development of an index for monitoring and prediction. Mon. Wea. Rev.,
132, 1917–1932, https://doi.org/10.1175/1520-0493(2004)132<1917:AAR
MMI>2.0.CO;2.
Wood, K. M., and E. A. Ritchie, 2015: A definition for rapid weakening in the
North Atlantic and eastern North Pacific. Geophys. Res. Lett., 42, 10 091–
10 097, https://doi.org/10.1002/2015GL066697.
—, P. J. Klotzbach, J. M. Collins, and C. J. Schreck, 2019: The record‐setting 2018 eastern North Pacific hurricane season. Geophys. Res. Lett., 46,
10 072–10 081, https://doi.org/10.1029/2019GL083657.
—, —, —, L.-P. Caron, R. E. Truchelut, and C. J. Schreck III, 2020: Factors affecting the 2019 Atlantic hurricane season and the role of the Indian Ocean Dipole. Geophys. Res. Lett., 47, e2020GL087781, https://doi.
org/10.1029/2020GL087781.
Wu, M.-C., K.-H. Yeung, and W.-L. Chang, 2006: Trends in western North Pacific
tropical cyclone intensity. Eos, Trans. Amer. Geophys. Union, 87, 537–548,
https://doi.org/10.1029/2006EO480001.
Yim, S. Y., B. Wang, J. Liu, and Z. W. Wu, 2014: A comparison of regional
monsoon variability using monsoon indices. Climate Dyn., 43, 1423–1437,
https://doi.org/10.1007/s00382-013-1956-9.
AU G U S T 2 0 2 0 | S t a t e o f t h e C l i m a t e i n 2 0 1 9
Ying, M., E.-J. Cha, and H. J. Kwon, 2011: Comparison of three western North
Pacific tropical best track datasets in a seasonal context. J. Meteor. Soc.
Japan, 89, 211–224, https://doi.org/10.2151/jmsj.2011-303.
Yu, H., Y. Lu, P.-Y. Chen, and W. C. Zhou, 2012: Intensity change characteristics
of tropical cyclones in the western North Pacific as revealed by three different datasets. J. Trop. Meteor., 18, 119–126.
Yuan, J., and J. Cao, 2012: North Indian Ocean tropical cyclone activities influenced by the Indian Ocean Dipole mode. Sci. China Earth Sci., 56, 855–865,
https://doi.org/10.1007/s11430-012-4559-0.
Zhang, C., 2005: Madden–Julian oscillation. Rev. Geophys., 43, RG2003,
https://doi.org/10.1029/2004RG000158.
—, 2013: Madden–Julian oscillation: Bridging weather and climate. Bull.
Amer. Meteor. Soc., 94, 1849–1870, https://doi.org/10.1175/BAMSD-12-00026.1.
—, and J. Gottschalck, 2002: SST anomalies of ENSO and the Madden–Julian
oscillation in the equatorial Pacific. J. Climate, 15, 2429–2445, https://doi.
org/10.1175/1520-0442(2002)015<2429:SAOEAT>2.0.CO;2.
Zheng, Z.-W., I.-I. Lin, B. Wang, H.-C. Huang, and C.-H. Chen, 2015: A long neglected damper in the El Nino–Typhoon relationship: A ‘Gaia-like’ process.
Sci. Rep., 5, 11103, https://doi.org/10.1038/srep11103.
Unauthenticated | Downloaded 12/04/22 09:14S238
AM UTC
4. THE TROPICS