Advertisement

Too close for comfort

Tropical cyclones are becoming stronger and occurring at higher latitudes than in the past. Wang and Toumi report that cyclones' points of maximum intensity also have been coming closer to land since 1982, the distance decreasing by about 30 kilometers per decade as their paths drift poleward and westward (see the Perspective by Camargo and Wing). This shift could increase the danger that tropical cyclones pose to coastal populations.
Science, this issue p. 514; see also p. 458

Abstract

Poleward migrations of tropical cyclones have been observed globally, but their impact on coastal areas remains unclear. We investigated the change in global tropical cyclone activity in coastal regions over the period 1982–2018. We found that the distance of tropical cyclone maximum intensity to land has decreased by about 30 kilometers per decade, and that the annual frequency of global tropical cyclones increases with proximity to land by about two additional cyclones per decade. Trend analysis reveals a robust migration of tropical cyclone activity toward coasts, concurrent with poleward migration of cyclone locations as well as a statistically significant westward shift. This zonal shift of tropical cyclone tracks may be mainly driven by global zonal changes in environmental steering flow.
Tropical cyclones are one of the most devastating natural disasters in terms of their average destructiveness and annual frequency affecting coastal regions. There are ~100-year records of tropical cyclone activity in some regions (1, 2), but only since 1982 have tropical cyclones been monitored globally by satellites. Poleward migration of the locations of tropical cyclone maximum intensity in the past 40 years has been reported in the global “best-track” data (3) and is projected to continue regionally in the 21st century (4). This poleward migration has been traced back to the poleward shift of tropical cyclone formation, which is speculated to be linked to an anthropogenically forced tropical circulation expansion (5). These trends could change the coastal tropical cyclone risk in the future (4). However, the poleward migration of the locations of tropical cyclone formation and peak intensity by themselves may not necessarily indicate a change in coastal tropical cyclone risk directly, because these locations are commonly too far away to generate an impact on the coasts. The changes in coastal tropical cyclone activity and landfall frequency are a central and perhaps ultimate concern.
A tropical cyclone landfall is conventionally defined as the intersection of the surface center of a tropical cyclone with a coastline (6). To date, there has been no firm evidence of global trends of the frequency of tropical cyclones with maximum wind speed above the hurricane-force wind (64 knots) at landfall (7). However, a “near miss” or “indirect-hit” tropical cyclone track can also cause damage; for example, Sandy in 2012 and Dorian in 2019 skirted along the U.S. coast for a considerable time before making landfall. Investigating the trend of tropical cyclones entering coastal regions is therefore essential to better understand tropical cyclone risk.
Here we consider tropical cyclones, defined by the lifetime maximum intensity (LMI) reaching at least 34 knots (see supplementary materials), during the period 1982–2018. This is the period for which we have the highest confidence in the quality of data and its completeness in global basins (8). The poleward migration of tropical cyclone activity was first found at LMI (3). The annual mean distance of the locations of LMI to the nearest land (Fig. 1A) also shows an evident and statistically significant decreasing trend of –32 ± 31 km per decade (table S1). The fraction of annual tropical cyclones entering coastal regions, defined as the offshore area with a distance to the nearest land less than 200 km (Fig. 1B), shows a robust increase of 2.2 ± 1.9% per decade (table S1). We also find positive trends of the annual mean lifetime fraction that tropical cyclones spend in coastal regions, at a rate of 2.1 ± 1.2% per decade globally, and in the Northern and Southern hemispheres of 2.1 ± 1.6% and 2.1 ± 1.9% per decade, respectively (table S1). The consistently positive contribution from both hemispheres (table S1) to the three global trends shown in Fig. 1 suggests that the increase of tropical cyclone activity in coastal regions is a global phenomenon, although regional differences between basins are evident.
Fig. 1 Landward migration of global tropical cyclone activity.
(A) Time series of annual-mean distance to land of the locations of lifetime maximum intensity (LMI). (B) As in (A), but for the fraction of annual tropical cyclones entering coastal regions. (C) As in (A), but for the time fraction of annual-mean life span spent in coastal regions. Dashed lines are historical data; solid lines are linear trends, with shaded areas denoting 95% confidence intervals.
The trend of the frequency of coastal tropical cyclones increases very significantly with proximity to land by one cyclone per decade per 1000 km (Fig. 2 and table S2). There are ~2 ± 2 additional tropical cyclones per decade in the 200-km coastal region globally (Fig. 2). Consistent with this, the fraction of annual tropical cyclones entering coastal regions (fig. S1A) and the annual-mean lifetime fraction that tropical cyclones spend in coastal regions (fig. S1B) increase with proximity to land. The statistical significance remains by excluding short-lived storms (9) and using an alternative set of tropical cyclone records (table S3). These clear increases in trend with reduced distance-to-land threshold suggest a global-scale landward migration of tropical cyclone activity.
Fig. 2 Increasing trend of global tropical cyclone activity in coastal regions when reducing the distance-to-land threshold for coastal region definition.
Changes in trends of the frequency of global annual tropical cyclones entering coastal regions are shown as a consequence of reducing the distance-to-land threshold from 2000 km to 200 km at 100-km intervals. The dashed line, solid line, and shading show the historical data, linear trend, and 95% confidence interval of the trend, respectively. On average there are 45 cyclones per year reaching coastal regions within 200 km of land.
Considering the geometry of global land-sea distribution, zonal and/or meridional shifts of tropical cyclone locations may lead to a change of tropical cyclone activity in coastal regions. Our analysis shows that from 1982–1999 to 2000–2018, the epochal mean tropical cyclone location in the global basins migrated not only poleward but also westward (Table 1). The globally epochal poleward and westward shifts are both statistically significant. If we measure the track shift in terms of degrees, the zonal shift is considerably larger than the meridional change. Tropical cyclone activity is being shifted westward in the West Pacific, East Pacific, and North and South Indian Ocean. These four basins account for ~75% of global tropical cyclones for 1982–2018, which accounts for the global mean westward track shifts. There are eastward track shifts in the South Pacific and no significant zonal shifts in the North Atlantic. These basin-wise changes in tropical cyclone locations can also be seen in the density change of tropical cyclone locations between the two epochs (fig. S2).
Global NHEM SHEM WPAC EPAC NATL NIO SIO SPAC
Westward (degrees longitude) +0.8 ± 0.2 +1.0 ± 0.3 +0.4 ± 0.3 +1.6 ± 0.4 +1.3 ± 0.7 –0.1 ± 0.5 +2.7 ± 0.4 +1.4 ± 0.5 –1.6 ± 0.6
Poleward (degrees latitude) +0.3 ± 0.1 +0.2 ± 0.1 +0.6 ± 0.2 +0.5 ± 0.2 +0.4 ± 0.2 –0.3 ± 0.3 –0.8 ± 0.3 +0.6 ± 0.2 +0.5 ± 0.3
Zonal steering (m s–1) +0.3 ± 0.3 +0.2 ± 0.3 +0.4 ± 0.6 +0.4 ± 0.4 +0.4 ± 0.3 +0.1 ± 0.3 –0.1 ± 0.3 +0.3 ± 0.5 +0.5 ± 0.7
Table 1 Changes in mean tropical cyclone location and deep-layer steering flow between the periods 1982–1999 and 2000–2018.
The seasonal mean steering flow is calculated from the reanalysis product ERA5. The meridional steering flow is not shown because the absolute mean epochal change is less than 0.1 m s–1 in all regions. Statistical significance is indicated in bold at the 95% confidence interval. The differences are calculated globally, in the hemispheres (NHEM, Northern Hemisphere; SHEM, Southern Hemisphere), and in the individual basins (WPAC, West Pacific; EPAC, East Pacific; NATL, North Atlantic; NIO, North Indian Ocean; SIO, South Indian Ocean; SPAC, South Pacific).
We do not find any statistically significant global change in cyclone zonal translation velocity, duration, or zonal shifts of cyclone genesis (table S4). The lack of regionally significant zonal genesis location change is different from a reported poleward shift of genesis in the Pacific Ocean basins (10). The relative changes in meridional cyclone translation velocity in the second epoch agree with the meridional track shifts in all the basins.
Tropical cyclone tracks are primarily determined by the environmental steering flow and genesis location (11, 12). We find significant enhancement of westward steering at a global scale in all basins except the North Indian Ocean (Table 1). Consistent general circulation patterns are confirmed in all three reanalysis products (fig. S3, A to C). The global mean change in zonal steering (+0.3 m s–1; Table 1) is about 11% of the global mean zonal steering speed (2.6 m s–1 in ERA5 reanalysis; fig. S3) and global mean cyclone zonal translation speed (2.7 m s–1). The West Pacific, East Pacific, and South Indian Ocean show both westward track shift and enhanced westward environmental steering between the epochs. The West Pacific is the only basin that shows a consistent link among seasonal westward steering enhancement, larger westward cyclone velocity, and westward track shift. This robust linkage to seasonal mean environmental conditions may be due to the considerably larger fraction of storm days in the season in the West Pacific relative to the other basins (13).
An environment with reduced vertical wind shear and/or increased potential intensity is favorable for tropical cyclone development (14, 15) and may also contribute to geographical shifts of tropical cyclone locations (3). We find a significant relative decrease in vertical wind shear from the east to west globally (Fig. 3), which is confirmed in individual basins with reanalysis products (fig. S4, A to F). The South Pacific is the only basin where neither steering flow nor vertical wind shear fully explains the regional eastward track shift. However, we do find a relative increase in potential intensity to the east and a decrease to the west in that basin (fig. S4L). In the North Atlantic, the shear weakens only in the central part of the basin (fig. S4C) with no zonal change in either steering (Table 1) or potential intensity (fig. S4I), which is consistent with the absence of a significant zonal track shift. Further analysis reveals that the zonal potential intensity change is primarily modulated by the variation of the atmospheric convective available potential energy (fig. S5).
Fig. 3 Epochal change in the global-mean basin-wise climate condition as a function of longitude.
The climate condition shown here is calculated as the relative change in vertical wind shear in the second epoch, 2000–2018, with respect to that in the first epoch, 1982–1999. The vertical wind shear is extracted from the reanalysis product ERA5. The mean vertical wind shear is calculated in seven of the 10° longitude bins, from west to east, in all the basins. The black dashed line shows the mean relative change in the seven longitude bins. The solid line shows the linear fit with ordinary least-squares regression. The shading denotes the 95% confidence interval of the fit.
Given the dominant role of steering flow in tropical cyclone tracks (12) and a larger epochal relative change of steering (Table 1) than wind shear (Fig. 3) in our analysis, it is very likely that the westward shift of cyclone tracks is mainly due to the enhanced westward steering. Anomalous westward steering may increase tropical cyclone frequency in the west of the basins, and relatively reduced vertical wind shear provides a further favorable environment for cyclones there. The quantitative contribution of the three environmental factors needs further exploration.
Climate modes such as the El Niño–Southern Oscillation (ENSO) and Pacific Decadal Oscillation (PDO), considered as measures of mainly internal unforced variability, can modulate tropical cyclone activities (1619) and global atmospheric circulation in the tropics (20, 21). We find that ENSO has a very limited impact on annual trends. However, the PDO may substantially contribute to the trends in the Northern Hemisphere (table S1). The epochal steering flow change is similar to the steering difference between PDO warm and cold phases (fig. S3, D to F). The PDO index had a phase change in 1998 (5), which also separates the first and second epochs (pre- and post-2000) in our analysis into a warm and cold PDO phase, respectively.
Given no strong trend of global annual frequency of tropical cyclones (22), an increasing trend of annual coastal cyclone fraction agrees with a rising frequency of coastal tropical cyclones shown here. The lifetime fraction trend in Fig. 1C seems to agree with a recently reported global slowdown of tropical cyclone translation speed (23) because the annual mean tropical cyclone duration has not changed significantly during the period 1982–2018 (table S4). However, this causal hypothesis needs to be confirmed, as changes in cyclone moving speed are disputed (24).
We find statistically significant landward migration of tropical cyclone activity. The lack of a statistically significant trend of actual landfall (+1 ± 2 cyclones per decade globally) may be due to the rapid intensity reduction before and during landfall. Major tropical cyclones have decayed more rapidly after LMI (25). This could stabilize the annual frequency of landfall defined by a fixed intensity threshold. Thus, the lack of a global trend of annual mean landfall frequency, as has been previously reported (7) and confirmed in our analysis, may not be contradictory to the increase in tropical cyclone activity in coastal regions reported here.
Long-term zonal migration of tropical cyclones, as found in our analysis, has been previously shown directly or indirectly in some individual basins for different metrics. In the West Pacific, westward migration of the tropical cyclone maximum intensity locations was reported (26), which was related to the strengthened tropical Pacific Walker circulation driven by the zonally enhanced sea surface temperature gradient. No detectable trend of U.S. landfall hurricane frequency has emerged (22). There has been a decreasing trend of landfall frequency by severe tropical cyclones in eastern Australia (2). In the North Indian Ocean, the reduction of local wind shear over the Arabian Sea has made tropical cyclone development more favorable, and this may continue in the future (27).
The relatively limited length of global tropical cyclone observations, limitations of climate modeling, and possible association with the PDO limit our ability to attribute tropical cyclone migration toward coasts to anthropogenic forcing. However, we note that global epochal westward shifts are still present when considering only PDO-neutral years (table S5). The West Pacific LMI distance to land, fraction of coastal cyclones, and coastal time fraction trends since 1950 are still significant after the PDO is removed (table S1). This suggests that factors other than the PDO are also important. It is argued that greenhouse gas emissions have contributed to the observed changes in regional distribution of tropical cyclones since 1980, and yet these trends may not persist in the 21st century (28).
The enhanced westward tropical cyclone steering may also be consistent with Hadley circulation expansion. The PDO phase change and Hadley circulation expansion are related (20, 21). The poleward migration of cyclone locations has also been related to the anomalous sinking and rising motions due to the meridional expansion of the Hadley circulation (5). Long-term zonal shifts of tropical precipitation emerge in atmospheric reanalysis data (29). The Walker circulation dominates the large-scale zonal motion in the tropics. An attempt has been made to establish a link between the covariability of tropical cyclone genesis locations and both the Hadley and Walker circulations in the West Pacific (30). The combined Walker and Hadley circulation variation is a plausible avenue for future study to understand the global tropical cyclone changes presented here. The considerable impacts of coastal tropical cyclones warrant further study of the changes in coastal tropical cyclone activity, as identified here, and its future projections and the consequent change in risk.

Acknowledgments

We thank C. Landsea, two anonymous reviewers, and K. Emanuel for helpful comments. Funding: Supported by the UK-China Research and Innovation Partnership Fund through the Met Office Climate Science for Service Partnership (CSSP) China as part of the Newton Fund. Author contributions: S.W. and R.T. conceived the study. S.W. performed the analysis. Both authors discussed the results and jointly contributed to writing the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: The tropical cyclone best-track data can be downloaded from the National Centers for Environmental Information website (www.ncdc.noaa.gov/ibtracs/index.php). The ERA5, MERRA, and NCEP/NCAR reanalysis data are available at the European Centre for Medium-Range Weather Forecasts (www.ecmwf.int/en/forecasts/datasets/reanalysis-datasets/era5), the NASA Modeling and Assimilation Data and Information Services Center (https://gmao.gsfc.nasa.gov/reanalysis/MERRA-2/), and the NOAA Physical Sciences Laboratory (https://psl.noaa.gov/data/reanalysis/reanalysis.shtml), respectively.

Supplementary Material

Summary

Materials and Methods
Figs. S1 to S6
Tables S1 to S5
References (3140)

Resources

File (abb9038_wang_sm.pdf)

References and Notes

1
J. C. L. Chan, K.-S. Liu, M. Xu, Q. Yang, Variations of frequency of landfalling typhoons in East China, 1450-1949. Int. J. Climatol. 32, 1946–1950 (2012).
2
J. Callaghan, S. B. Power, Variability and decline in the number of severe tropical cyclones making land-fall over eastern Australia since the late nineteenth century. Clim. Dyn. 37, 647–662 (2011).
3
J. P. Kossin, K. A. Emanuel, G. A. Vecchi, The poleward migration of the location of tropical cyclone maximum intensity. Nature 509, 349–352 (2014).
4
J. P. Kossin, K. A. Emanuel, S. J. Camargo, Past and projected changes in western North Pacific tropical cyclone exposure. J. Clim. 29, 5725–5739 (2016).
5
S. Sharmila, K. J. E. Walsh, Recent poleward shift of tropical cyclone formation linked to Hadley cell expansion. Nat. Clim. Chang. 8, 730–736 (2018).
6
NOAA, Glossary of National Hurricane Center Terms; www.nhc.noaa.gov/aboutgloss.shtml.
7
J. Weinkle, R. Maue, R. Pielke Jr., ., Historical global tropical cyclone landfalls. J. Clim. 25, 4729–4735 (2012).
8
K. R. Knapp, M. C. Kruk, Quantifying interagency differences in tropical cyclone best-track wind speed estimates. Mon. Weather Rev. 138, 1459–1473 (2010).
9
C. W. Landsea, G. A. Vecchi, L. Bengtsson, T. R. Knutson, Impact of duration thresholds on Atlantic tropical cyclone counts. J. Clim. 23, 2508–2519 (2010).
10
A. S. Daloz, S. J. Camargo, Is the poleward migration of tropical cyclone maximum intensity associated with a poleward migration of tropical cyclone genesis? Clim. Dyn. 50, 705–715 (2018).
11
L. Wu, B. Wang, Assessing Impacts of Global Warming on Tropical Cyclone Tracks. J. Clim. 17, 1686–1698 (2004).
12
J. C. L. Chan, the Physics of Tropical Cyclone Motion. Annu. Rev. Fluid Mech. 37, 99–128 (2005).
13
B. Wang, Y. Yang, Q. H. Ding, H. Murakami, F. Huang, Climate control of the global tropical storm days (1965–2008). Geophys. Res. Lett. 37, L07704 (2010).
14
K. A. Emanuel, Thermodynamic control of hurricane intensity. Nature 401, 665–669 (1999).
15
M. Ting, J. P. Kossin, S. J. Camargo, C. Li, Past and Future Hurricane Intensity Change along the U.S. East Coast. Sci. Rep. 9, 7795 (2019).
16
S. J. Camargo, A. W. Robertson, A. G. Barnston, M. Ghil, Clustering of eastern North Pacific tropical cyclone tracks: ENSO and MJO effects. Geochem. Geophys. Geosyst. 9, Q06V05 (2008).
17
J. P. Kossin, S. J. Camargo, M. Sitkowski, Climate modulation of North Atlantic hurricane tracks. J. Clim. 23, 3057–3076 (2010).
18
S. J. Camargo, A. W. Robertson, S. J. Gaffney, P. Smyth, M. Ghil, Cluster analysis of typhoon tracks. Part II: Large-scale circulation and ENSO. J. Clim. 20, 3654–3676 (2007).
19
H. A. Ramsay, S. J. Camargo, D. Kim, Cluster analysis of tropical cyclone tracks in the Southern Hemisphere. Clim. Dyn. 39, 897–917 (2012).
20
R. J. Allen, M. Kovilakam, The Role of Natural Climate Variability in Recent Tropical Expansion. J. Clim. 30, 6329–6350 (2017).
21
P. W. Staten, J. Lu, K. M. Grise, S. M. Davis, T. Birner, Re-examining tropical expansion. Nat. Clim. Chang. 8, 768–775 (2018).
22
T. Knutson, S. J. Camargo, J. C. L. Chan, K. Emanuel, C.-H. Ho, J. Kossin, M. Mohapatra, M. Satoh, M. Sugi, K. Walsh, L. Wu, Tropical Cyclones and Climate Change Assessment: Part I: Detection and Attribution. Bull. Am. Meteorol. Soc. 100, 1987–2007 (2019).
23
J. P. Kossin, A global slowdown of tropical-cyclone translation speed. Nature 558, 104–107 (2018).
24
M. Yamaguchi, J. C. L. Chan, I. J. Moon, K. Yoshida, R. Mizuta, Global warming changes tropical cyclone translation speed. Nat. Commun. 11, 47 (2020).
25
S. Wang, T. Rashid, H. Throp, R. Toumi, A Shortening of the Life Cycle of Major Tropical Cyclones. Geophys. Res. Lett. 47, (2020).
26
D.-S. R. Park, C.-H. Ho, J.-H. Kim, Growing threat of intense tropical cyclones to East Asia over the period 1977–2010. Environ. Res. Lett. 9, 014008 (2014).
27
H. Murakami, G. A. Vecchi, S. Underwood, Increasing frequency of extremely severe cyclonic storms over the Arabian Sea. Nat. Clim. Chang. 7, 885–889 (2017).
28
H. Murakami, T. L. Delworth, W. F. Cooke, M. Zhao, B. Xiang, P. C. Hsu, Detected climatic change in global distribution of tropical cyclones. Proc. Natl. Acad. Sci. U.S.A. 117, 10706–10714 (2020).
29
M. P. Byrne, A. G. Pendergrass, A. D. Rapp, K. R. Wodzicki, Response of the Intertropical Convergence Zone to Climate Change: Location, Width, and Strength. Curr. Clim. Change Rep. 4, 355–370 (2018).
30
H. Zhao, J. Zhang, P. J. Klotzbach, S. Chen, Recent Increased Covariability of Tropical Cyclogenesis Latitude and Longitude over the Western North Pacific during the Extended Boreal Summer. J. Clim. 32, 8167–8179 (2019).
31
K. R. Knapp, M. C. Kruk, D. H. Levinson, H. J. Diamond, C. J. Neumann, The International Best Track Archive for Climate Stewardship, IBTrACS. Bull. Am. Meteorol. Soc. 91, 363–376 (2010).
32
Copernicus Climate Change Service (C3S), “ERA5: Fifth generation of ECMWF atmospheric reanalyses of the global climate.” Copernicus Climate Change Service Climate Data Store (2017); https://cds.climate.copernicus.eu/cdsapp#!/home.
33
M. M. Rienecker, M. J. Suarez, R. Gelaro, R. Todling, J. Bacmeister, E. Liu, M. G. Bosilovich, S. D. Schubert, L. Takacs, G.-K. Kim, S. Bloom, J. Chen, D. Collins, A. Conaty, A. da Silva, W. Gu, J. Joiner, R. D. Koster, R. Lucchesi, A. Molod, T. Owens, S. Pawson, P. Pegion, C. R. Redder, R. Reichle, F. R. Robertson, A. G. Ruddick, M. Sienkiewicz, J. Woollen, MERRA: NASA’s Modern-Era Retrospective Analysis for Research and Applications. J. Clim. 24, 3624–3648 (2011).
34
E. Kalnay, M. Kanamitsu, R. Kistler, W. Collins, D. Deaven, L. Gandin, M. Iredell, S. Saha, G. White, J. Woollen, Y. Zhu, A. Leetmaa, R. Reynolds, M. Chelliah, W. Ebisuzaki, W. Higgins, J. Janowiak, K. C. Mo, C. Ropelewski, J. Wang, R. Jenne, D. Joseph, The NCEP/NCAR 40-Year Reanalysis Project. Bull. Am. Meteorol. Soc. 77, 437–471 (1996).
35
D. S. Wilks, Statistical Methods in the Atmospheric Sciences (Academic Press, 2011).
36
K. Emanuel, Sensitivity of Tropical Cyclones to Surface Exchange Coefficients and a Revised Steady-State Model Incorporating Eye Dynamics. J. Atmos. Sci. 52, 3969–3976 (1995).
37
A. G. Bamston, M. Chelliah, S. B. Goldenberg, Documentation of a highly ENSO‐related sst region in the equatorial pacific: Research note. Atmos.-Ocean 35, 367–383 (1997).
38
K. Ashok, S. K. Behera, S. A. Rao, H. Weng, T. Yamagata, El Niño Modoki and its possible teleconnection. J. Geophys. Res. Oceans 112, 1–27 (2007).
39
Y. Zhang, J. M. Wallace, D. S. Battisti, ENSO-like interdecadal variability: 1900-93. J. Clim. 10, 1004–1020 (1997).
40
N. J. Mantua, S. R. Hare, Y. Zhang, J. M. Wallace, R. C. Francis, A Pacific Interdecadal Climate Oscillation with Impacts on Salmon Production. Bull. Am. Meteorol. Soc. 78, 1069–1079 (1997).

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Embedded figures cannot be submitted, and we discourage the use of figures within eLetters in general. If a figure is essential, please include a link to the figure within the text of the eLetter. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 371 | Issue 6528
29 January 2021

Submission history

Received: 26 March 2020
Accepted: 30 November 2020
Published in print: 29 January 2021

Permissions

Request permissions for this article.

Acknowledgments

We thank C. Landsea, two anonymous reviewers, and K. Emanuel for helpful comments. Funding: Supported by the UK-China Research and Innovation Partnership Fund through the Met Office Climate Science for Service Partnership (CSSP) China as part of the Newton Fund. Author contributions: S.W. and R.T. conceived the study. S.W. performed the analysis. Both authors discussed the results and jointly contributed to writing the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: The tropical cyclone best-track data can be downloaded from the National Centers for Environmental Information website (www.ncdc.noaa.gov/ibtracs/index.php). The ERA5, MERRA, and NCEP/NCAR reanalysis data are available at the European Centre for Medium-Range Weather Forecasts (www.ecmwf.int/en/forecasts/datasets/reanalysis-datasets/era5), the NASA Modeling and Assimilation Data and Information Services Center (https://gmao.gsfc.nasa.gov/reanalysis/MERRA-2/), and the NOAA Physical Sciences Laboratory (https://psl.noaa.gov/data/reanalysis/reanalysis.shtml), respectively.

Authors

Affiliations

Department of Physics, Imperial College London, London SW7 2BU, UK.
Ralf Toumi
Department of Physics, Imperial College London, London SW7 2BU, UK.

Funding Information

Notes

*
Corresponding author. Email: [email protected]

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. Estimation of the Spring Tide Bedload Transport at the Eastern Entrance of the Qiongzhou Strait, Water, 15, 4, (724), (2023).https://doi.org/10.3390/w15040724
    Crossref
  2. Relationship between South China Sea Summer Monsoon and Western North Pacific Tropical Cyclones Linkages with the Interaction of Indo-Pacific Pattern, Atmosphere, 14, 4, (645), (2023).https://doi.org/10.3390/atmos14040645
    Crossref
  3. Increased U.S. coastal hurricane risk under climate change, Science Advances, 9, 14, (2023)./doi/10.1126/sciadv.adf0259
    Abstract
  4. Has There Been a Recent Shallowing of Tropical Cyclones?, Geophysical Research Letters, 50, 5, (2023).https://doi.org/10.1029/2022GL102184
    Crossref
  5. Phase Shifts of the PDO and AMO Alter the Translation Distance of Global Tropical Cyclones, Earth's Future, 11, 3, (2023).https://doi.org/10.1029/2022EF003079
    Crossref
  6. Have atmospheric extremes changed in the past?, Science of Weather, Climate and Ocean Extremes, (81-126), (2023).https://doi.org/10.1016/B978-0-323-85541-9.00009-2
    Crossref
  7. Divergence of tropical cyclone hazard based on wind-weighted track distributions in the Coral Sea, over 50 years, Natural Hazards, (2023).https://doi.org/10.1007/s11069-022-05780-3
    Crossref
  8. Response Process of Coastal Hypoxia to a Passing Typhoon in the East China Sea, Frontiers in Marine Science, 9, (2022).https://doi.org/10.3389/fmars.2022.892797
    Crossref
  9. Dynamical Projections of the Mean and Extreme Wave Climate in the Bohai Sea, Yellow Sea and East China Sea, Frontiers in Marine Science, 9, (2022).https://doi.org/10.3389/fmars.2022.844113
    Crossref
  10. Enhanced understanding of changes in tropical cyclones’ landfall frequency over mainland China, Frontiers in Earth Science, 10, (2022).https://doi.org/10.3389/feart.2022.932843
    Crossref
  11. See more
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Check Access

Log in to view the full text

AAAS ID LOGIN

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

Log in via OpenAthens.
Log in via Shibboleth.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Media

Figures

Multimedia

Tables

Share

Share

Share article link

Share on social media