ArticlePDF Available

Abstract and Figures

Phytoplankton blooms in coastal oceans can be beneficial to coastal fisheries production and ecosystem function, but can also cause major environmental problems1,2—yet detailed characterizations of bloom incidence and distribution are not available worldwide. Here we map daily marine coastal algal blooms between 2003 and 2020 using global satellite observations at 1-km spatial resolution. We found that algal blooms occurred in 126 out of the 153 coastal countries examined. Globally, the spatial extent (+13.2%) and frequency (+59.2%) of blooms increased significantly (P < 0.05) over the study period, whereas blooms weakened in tropical and subtropical areas of the Northern Hemisphere. We documented the relationship between the bloom trends and ocean circulation, and identified the stimulatory effects of recent increases in sea surface temperature. Our compilation of daily mapped coastal phytoplankton blooms provides the basis for global assessments of bloom risks and benefits, and for the formulation or evaluation of management or policy actions.
Content may be subject to copyright.
280 | Nature | Vol 615 | 9 March 2023
Article
Coastal phytoplankton blooms expand and
intensify in the 21st century
Yanhui Dai1,9, Shangbo Yang1,9, Dan Zhao1, Chuanmin Hu2, Wang Xu3, Donald M. Anderson4,
Yun Li5, Xiao-Peng Song6, Daniel G. Boyce7, Luke Gibson1, Chunmiao Zheng1,8 & Lian Feng1 ✉
Phytoplankton blooms in coastal oceans can be benecial to coastal sheries
production and ecosystem function, but can also cause major environmental
problems1,2—yet detailed characterizations of bloom incidence and distribution are
not available worldwide. Here we map daily marine coastal algal blooms between
2003 and 2020 using global satellite observations at 1-km spatial resolution. We found
that algal blooms occurred in 126 out of the 153 coastal countries examined. Globally,
the spatial extent (+13.2%) and frequency (+59.2%) of blooms increased signicantly
(P < 0.05) over the study period, whereas blooms weakened in tropical and
subtropical areas of the Northern Hemisphere. We documented the relationship
between the bloom trends and ocean circulation, and identied the stimulatory
eects of recent increases in sea surface temperature. Our compilation of daily
mapped coastal phytoplankton blooms provides the basis for global assessments
of bloom risks and benets, and for the formulation or evaluation of management
or policy actions.
Phytoplankton blooms are accumulations of microscopic algae in the
surface layer of fresh and marine water bodies. Although many blooms
can occur naturally, nutrients linked to anthropogenic eutrophication
are expected to intensify their frequency globally
2–4
. Many algal blooms
are beneficial, fixing carbon at the base of the food chain and support-
ing fisheries and ecosystems worldwide. However, proliferations of
algae that cause harm (termed harmful algal blooms (HABs)) have
become a major environmental problem worldwide5–7. For instance,
the toxins produced by some algal species can accumulate in the food
web, causing closures of fisheries as well as illness or mortality of marine
species and humans
8–10
. In other cases, the decay of a dense algal bloom
can deplete oxygen in bottom waters, forming anoxic ‘dead zones’ that
can cause fish and invertebrate die-offs and ecosystem restructuring,
with serious consequences for the well-being of coastal communities
1,11
.
Unfortunately, algal bloom frequency and distribution are projected
to increase with future climate change12,13, with some changes causing
adverse effects on aquatic ecosystems, fisheries and coastal resources.
Owing to substantial heterogeneity in space and time, algal blooms
are challenging to characterize on a large scale
5,14
, and thus present
knowledge does not allow us to answer one of the most fundamen-
tal questions: whether algal blooms have changed in recent decades
on a global basis
6,15,16
. For example, although HAB events have been
compiled into the UNESCO (United Nations Educational, Scientific,
and Cultural Organization) Intergovernmental Oceanographic Com-
mission Harmful Algae Event Database (HAEDAT) globally since 1985,
bloom trends are difficult to resolve, owing to inconsistent sampling
efforts and the diversity of the eco-environmental or socio-economic
effects
6
. Alternatively, satellite data have been used to monitor the
ocean surface continuously since 1997 and have enabled bloom detec-
tion in many coastal regions1719. However, the technical difficulties in
dealing with complex optical features across different types of coastal
waters have so far prohibited their application globally
20
. To fill this
knowledge gap, we developed a method to map global coastal algal
blooms and used this tool to examine satellite images between 2003
and 2020, addressing three fundamental questions: (1) where and how
frequently global coastal oceans have been affected by phytoplankton
blooms; (2) whether the blooms have expanded or intensified over the
past two decades, both globally and regionally; and (3) the identity of
the potential drivers.
Mapping global coastal phytoplankton blooms
We generated a satellite-based dataset of phytoplankton bloom occur-
rence to characterize the spatial and temporal patterns of algal blooms
in coastal oceans globally. The dataset was derived using global, 1-km
resolution daily observations from the ModerateResolutionImag-
ingSpectroradiometer (MODIS) onboard NASA’s Aqua satellite, and all
0.76 million images acquired by this satellite mission between 2003 and
2020 were used. We developed an automated method to detect phyto-
plankton blooms using MODIS images (Extended Data Fig.1) (Methods).
In this study, we define a phytoplankton bloom as the accumulation of
microscopic algae at the ocean surface that exhibits satellite-detectable
fluorescence signals
21
. However, whether a bloom produces toxins or is
harmful to humans or the marine environment is not distinguishable
from satellite data. We delineated bloom-affected areas (that is, the areas
where algal blooms were detected), and enumerated the bloom count
https://doi.org/10.1038/s41586-023-05760-y
Received: 17 May 2022
Accepted: 25 January 2023
Published online: 1 March 2023
Open access
Check for updates
1School of Environmental Science and Engineering, Southern University of Science and Technology, Shenzhen, China. 2College of Marine Science, University of South Florida, St. Petersburg,
FL, USA. 3Shenzhen Ecological and Environmental Monitoring Center of Guangdong Province, Shenzhen, China. 4Woods Hole Oceanographic Institution, Woods Hole, MA, USA. 5School of
Marine Science and Policy, College of Earth, Ocean, and Environment, University of Delaware, Lewes, DE, USA. 6Department of Geographical Sciences, University of Maryland, College Park,
MD, USA. 7Bedford Institute of Oceanography, Fisheries and Oceans Canada, Dartmouth, Nova Scotia, Canada. 8EIT Institute for Advanced Study, Ningbo, China. 9These authors contributed
equally: Yanhui Dai, Shangbo Yang. e-mail: fengl@sustech.edu.cn
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Nature | Vol 615 | 9 March 2023 | 281
at the 1-km pixel level (that is, the number of detected blooms per pixel)
(Fig.1). We further estimated the bloom frequency (dimensionless) by
integrating the bloom count and affected areas within 1° × 1° grid cells
(seeMethods), and this metric was used to examine temporal dynam
-
ics in bloom intensity. Validation with independent satellite samples
selected via several visual inspection techniques showed an overall accu-
racy level of more than 95% for our method, and comparisons using dis-
crete events in HAEDAT
6
indicated that we successfully identified bloom
counts for 79.3% of the historical HAB events in that database (Extended
Data Figs.2–6). We examined phytoplankton blooms in the exclusive
economic zones(EEZs) of 153 coastal countries and in 54 large marine
ecosystems (LMEs) (Extended Data Fig.7). Our study area encompasses
global continental shelves and outer margins of coastal currents, which
offer the majority of marine resources available for human use (seeMeth-
ods). Out of the 153 coastal countries examined, 126 were observed to
have phytoplankton blooms (Fig.1). The total bloom-affected area was
31.47 million km2, equivalent to approximately 24.2% of the global land
area and 8.6% of the global ocean area, with a median bloom count of
4.3 per year during the past 2 decades (Fig.1b). Europe (9.52 million
km2—30.3% of the total affected area) and North America (6.78 million
km
2
—21.5% of the total affected area) contributed the largest bloom
areas. By contrast, the most frequent blooms were found around Africa
and South America (median bloom counts of more than 6.3 per year).
Australia experienced the lowest frequency (2.4 per year) and affected
area (2.84 million km2—9.0% of the total affected area) of blooms.
Phytoplankton blooms occurred frequently in the eastern boundary
current systems (that is, California, Benguela, Humboldt and Canary),
northeastern USA, Latin America, the Baltic Sea, Northern Black Sea and
the Arabian Sea (Fig.1a). Five LMEs were found with the most frequent
blooms (annual median bloom count over 15), including Patagonian
Shelf, Northeast US continental Shelf, the Baltic Sea, Gulf of California
and Benguela Current (Extended Data Fig.7). These hotspots are often
reported as having a high incidence of algal blooms, some of which are
HABs, driven by either coastal upwelling or pronounced anthropogenic
nutrient enrichment
9,2226
. European LMEs showed mostly large propor-
tions of bloom-affected areas, and some also showed frequent bloom
occurrences. By contrast, Asian LMEs exhibited mainly infrequent
blooms, given their large affected areas. We identified more bloom
events in estuarine regions than along coasts in regions without major
river discharge (P < 0.05; Extended Data Fig.8), highlighting the critical
role of terrestrial nutrient sources in coastal algal blooms3.
Long-term trends
The total global bloom-affected area has expanded by 3.97 million km
2
(13.2%) between 2003 and 2020, equivalent to 0.14 million km2 yr−1
(P < 0.05; Fig.2). Furthermore, the number of countries with significant
bloom expansion was about 1.6 times those with a decreasing trend.
The global median bloom frequency showed an increasing rate of 59.2%
(+2.19% yr
−1
, P < 0.05) over the observed period. Spatially, areas showing
significant increasing trends (P < 0.05) in bloom frequency were 77.6%
larger than those with the opposite trends (Fig.2). Globally, our analysis
revealed overall consistent fluctuations between the bloom-affected
area and bloom frequency between 2003 and 2020 (Fig.2b). However,
there was no significant relationship between bloom extent and fre-
quency in 23 countries and 10 LMEs over the past two decades, under-
scoring the spatial and temporal variability of algal blooms and the
importance of continuous satellite monitoring.
The entire Southern Hemisphere was primarily characterized by
increased bloom frequency, although weakened blooms were also
sometimes found. In the Northern Hemisphere, the low latitude
(<30° N) coasts were mainly featured with strong bloom weakening
(Fig.2a), primarily in the California Current System and the Arabian
Sea. Bloom strengthening was found in the northern Gulf of Mexico and
the East and South China Seas, albeit at smaller magnitudes. At higher
latitudes, weakening blooms were detected mainly in the northeastern
North Atlantic and the Okhotsk Sea in the northwestern North Pacific.
Globally, the largest increases in bloom frequency were observed in six
a
1
10
30
50
SA AF EU NA AS AU Global
0
20
40
60
Bloom counts
b
EU NA AS SA AF AU
0
5
10
Affected area (×106 km2)
c
30.3%
21.5% 18.2%
11.8%
9.2%
9.0%
Global total affected area:
31.47 × 106 km2
Bloom counts
Fig. 1 | Globa l patterns o f coastal phy toplankton bl ooms betwe en 2003 and
2020. a, The spatial dist ribution of annu al mean bloom co unt based on dail y
satellite detections. b, Contine ntal and global s tatistics for a nnual mean
bloom count (South America (SA), n = 3,846, 125; Afric a (AF), n = 2,516, 225;
Europe (EU), n = 17,703,949; North Ameri ca (NA), n = 10,034,286; A sia (AS),
n = 5,371, 158; Australia (AU), n = 2, 781,998 pixel obser vations). The cen tre line
represen ts the median val ue, bottom and to p bounds of boxes are f irst and
third quart iles, and the whis kers show a maximum o f 1.5 times t he interquart ile
range. c, Conti nental stat istics for the lon g-term annual mean o f bloom-
affecte d areas (n = 18 years). The p ercentages s how the corresp onding
contribu tions to the glo bal total. The ba rs represent s .d. Open circle s are
the affec ted areas duri ngdifferentyears. M ap created usin g Python 3. 8.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
282 | Nature | Vol 615 | 9 March 2023
Article
major coastal current systems, including Oyashio (+6.31% yr−1), Alaska
(+5.22% yr−1), Canary (+4.28% yr−1), Malvinas (+3.02% yr−1), Gulf Stream
(+2.42% yr−1) and Benguela (+2.30% yr−1) (Figs.2a and3).
Natural and anthropogenic effects
Increases in sea surface temperature (SST) can stimulate bloom occur-
rence. We found significant positive correlations (P < 0.05) between the
annual mean bloom frequency and the coincident SST (SST data were
averaged over the growth window of algal blooms within a year (Meth-
ods and Extended Data Fig.9)) in several high-latitude regions (>40° N),
such as the Alaska Current (r = 0.44), the Oyashio Current (r = 0.48) and
the Baltic Sea (r = 0.41) (Fig.3). These findings agree with previous stud-
ies, in which the bloom-favourable seasons in these temperate seas
have been extended under warmer temperatures2729. However, this
temperature-based mechanism did not apply to regions with inconsistent
trends between SST and bloom frequency, particularly for the substantial
bloom weakening in the tropical and subtropical areas (Figs.2a and 3b).
a
–1.0 –0.5 0 0.5 1.0
Slope (104 yr−1)
100 0 100
Fraction (%)
50° S
50° N
Latitude
Positive
(P < 0.05)
(P ≥ 0.05)
Negative
(P < 0.05)
(P ≥ 0.05)
2003 2006 2009 2012
Year
2015 2018
1
2
3
Bloom frequency (×103)
+2.19% yr−1 (P = 0.006)
b
25
30
35
Affected area (×106 km2)
+0.14 × 106 km2 yr−1 (P = 0.001)
Fig. 2 | Trends of glo bal coastal p hytoplankto n blooms bet ween 2003 and
2020. a, Spatial patter ns of the trends in b loom frequen cy at a 1° × 1° grid scal e.
The latit udinal profile s show the frac tions of grids w ith signif icant and
insignif icant tren ds (positive or ne gative) along the ea st–west direction.
b, Interannual v ariability an d trends in annua l median bloom f requency and t otal
global bloo m-affected ar ea. The linea r slopes and P-value (two -sided t-test) are
indicate d. The shading a ssociated w ith the bloom fr equency dat a represents
an uncert ainty level of 5% in b loom detect ion. Map creat ed using Py thon 3.8.
a
–2 –1 012
10−6 °C m−1 decade−1
b
–1.0 –0.5 00.5 1.0
°C decade−1
0
6
12
6
8
10
r = 0.81*
S = –4.26
r = –0.83*
1 California Current
26
28
2
5
8
14
19
24
r = 0.61*
S = 2.42
r = –0.02
2 Gulf Stream
12
14
16
0
3
6
8
10
r = 0.84*
S = 4.28
r = –0.63*
3 Canary Current
20
22
24
8
14
20
4 Malvinas Current
7
9
11
r = 0.83*
S = 3.02
r = 0.12
11
13
15
2003 2011 2019
10
25
40
10
13
16
r = 0.73*
S = 2.30
r = 0.16
5 Benguela Current
19
21
2003 2011 2019
0
5
10
12
14
16
6 Oyashio Current
r = 0.58*
S = 6.31 r = 0.48* 10
12
14
2003 2011
Year
2019
0
2
4
6
8
r = 0.09
S = 5.22
r = 0.44*
7 Alaska Current
7
9
11
2003 2011 2019
3
6
9
3
7
11
r = 0.30
S = 2.74 r = 0.41*
8 Baltic Sea
9
11
Bloom frequency (104)
SST (10–6 per m)
SST (
)
c
7
8
6
5
4
3
2
1
2003 2011 2019 2003 2011 2019 2003 2011 2019 2003 2011 2019
Fig. 3 | Effe cts of clim ate change on phy toplankton b looms. a,b, Global
pattern s of trends in SST g radient (a) and SST (b) fro m 2003 to 2020. c, L ong-
term change s in bloom frequ ency in the reg ions labelled i n a and b, and their
relations hip to the SST and SS T gradient. Lin ear slope (S) of bloo m frequency
and the correlation coefficient (r) between b loom frequenc y and the SST and
the SST gra dient (SST) are show n. Asteris ks indicate st atisticall y signific ant
(P<0.05) correlation s. Maps create d using ArcMap 10.4 a nd Python 3. 8.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Nature | Vol 615 | 9 March 2023 | 283
Changes in climate can also affect ocean circulation, altering ocean
mixing and the transport of nutrients that drive the growth of marine
phytoplankton and bloom formation
3032
. We used the spatial SST gradi-
ent (in °C m
−1
) as a proxy for the magnitude of oceanic mesoscale cur-
rents (the time-varying velocity of kinetic energy (also known as the eddy
kinetic energy (EKE))) by following the methods of a previous study33,
and examined its effects on algal blooms (Methods). The trend in the
SST gradient appeared more spatially aligned to bloom frequency than
SST. We found significant positive correlations (P < 0.05) between the
SST gradient and bloom frequency for various coastal current systems,
including the Canary (r = 0.84), Malvinas (r = 0.83), California (r = 0.81),
Benguela (r = 0.73), Gulf Stream (r = 0.61) and Oyashio (r = 0.58) currents.
Trends in bloom frequency in subtropical eastern boundary upwelling
regions (the California, Benguela and Canary currents) followed the
changes in mesoscale currents (Fig.3a,c). In the California Current Sys-
tem, the decrease in bloom frequency was probably due to the weakened
upwelling (represented by a reduced SST gradient and increased SST)
and thus lower nutrient supply
25
. Conversely, the Canary and Benguela
currents were characterized by strengthened upwelling and increased
bloom frequency. The two western boundary current systems at high
latitudes (Malvinas and Oyashio)—although characterized by less pro-
nounced upwelling
34
—exhibited a similar mechanism to the subtropical
eastern boundary regions. For the subtropical western boundary Gulf
Stream current, the strengthened current jets (manifested as a larger SST
gradient) brought more nutrients from the continental shelf
35
, trigger-
ing more algal blooms. Nevertheless, whether these changes in oceanic
mesoscale activities were responses to wind, stratification, the shear of
ocean currents or other factors
33
requires region-based investigations.
Global climate events, represented as the multivariate El Niño–
Southern Oscillation index36 (MEI), also showed connections with coastal
bloom frequency. The minimum MEI in 2010 (a strong LaNiña year)
was followed by a low bloom frequency in the following year, and the
largest MEI in 2015 (a strong El Niño year) was followed by the strongest
bloom frequency in 2016 (Fig.2b and Extended Data Fig.10a).
Changes in anthropogenic nutrient enrichment may have also con-
tributed to the trends in blooms
37
. For example, the decline in bloom
frequency in the Arabian Sea, without clear links to SST or SST gradient
changes, could result from decreased fertilizer use in the surround-
ing countries (such as Iran) (Extended Data Fig.10). By contrast, the
bloom strengthening in some Asian countries could be attributed to
surges in fertilizer use
38
. We examined trends in fertilizer usage (either
nitrogen or phosphorus) and bloom frequency and found high positive
correlations in China, Iran, Vietnam and the Philippines. Paradoxi-
cally, decreased fertilizer uses and increased bloom frequency were
identifiedin some countries, suggesting that nutrient control efforts
might have been counterbalanced by the stimulatory effects of climate
warming or other factors. Furthermore, the intensified aquaculture
industry in Finland, China, Algeria, Guinea, Vietnam, Argentina, Russia
and Uruguay may also be linked to their increased bloom incidence,
as suggested by the significant positive correlations (r > 0.5, P < 0.05)
between their aquaculture production and bloom frequency. A similar
relationship between aquaculture expansion and positive trends in HAB
incidence was reported from an analysis of HAEDAT data
6
. However,
analogous positive feedbacks for fertilizer or aquaculture were not
found in many other countries. Thus, an ecosystem model incorporat-
ing terrestrial and oceanic nutrient transport and nutrient–plankton
relationships of different species39 is required to quantify the contribu-
tions of natural and anthropogenic factors to algal blooms14.
Future implications
We acknowledge that our criteria for a detectable bloom event is
operationally defined by sensor sensitivities and other factors, and
that the bloom count metric used here may underestimate algal
bloom incidence, particularly compared to harmful events entered in
HAEDAT. For example, in a recent global analysis of the HAEDAT events,
Hallegraeff etal.
6
report a dozen or more events per year for each of nine
regions over a 33-year study period, compared to the global median
bloom count of 4.3 in this study. There are several possible explana-
tions for this discrepancy, such as the many low-cell-concentration
HABs that are not detectable from space but that can still cause harm,
as well as sensor sensitivities and algorithm thresholds. Furthermore,
our bloom count was averaged over all 1-km pixels within the EEZs,
whereas HAEDAT entries are based on discrete sampling regions. This
underestimation does not, however, alter the trends and other conclu
-
sions of this study, as the metrics used here were constant across time
and space. Underestimates would have been uniform across regions
globally. In this regard, it is of note that the study of Hallegraeff etal.6
found a significant link between the number of HAEDAT events over
time and the global expansion of aquaculture production, similar to
findings in our study.
The major contribution of our study is to provide a spatially and
temporally consistent characterization of global coastal algal blooms
between 2003 and 2020. Globally, increasing trends in algal bloom
area and frequency are apparent. Regionally, however, trends were
non-uniform owing to the compounded effects of changes in climate
(such as changes in SST and SST gradient and climate extremes),
anthropogenic eutrophication and aquaculture development. Our
daily mapping of bloom events offers valuable baseline information
to understand the mechanisms underlying the formation, mainte-
nance, and dissipation of algal blooms
5,40
. This could aid in developing
forecasting models (on either global or regional scales) that can help
minimize the consequences of harmful blooms, and can also help in
policy decisions relating to the control of nutrient discharges and
other HAB-stimulatory factors. Noting again that many blooms are
beneficial, particularly in terms of their positive effects on ecosystems
as well as on wild and farmed fisheries, the results here can also con-
tribute toward policies and management actions that sustain those
beneficial blooms.
Online content
Any methods, additional references, Nature Portfolio reporting summa-
ries, source data, extended data, supplementary information, acknowl-
edgements, peer review information; details of author contributions
and competing interests; and statements of data and code availability
are available at https://doi.org/10.1038/s41586-023-05760-y.
1. Breitburg, D. etal. Declining oxygen in the global ocean and coastal waters. Science 359,
eaam7240 (2018).
2. Anderson, D. M. Turning back the harmful red tide. Nature 388, 513–514 (1997).
3. Beman, J. M., Arrigo, K. R. & Matson, P. A. Agricultural runoff fuels large phytoplankton
blooms in vulnerable areas of the ocean. Nature 434, 211–214 (2005).
4. Heisler, J. etal. Eutrophication and harmful algal blooms: a scientiic consensus. Harmful
Algae 8, 3–13 (2008).
5. Anderson, D. M., Cembella, A. D. & Hallegraeff, G. M. Progress in understanding harmful
algal blooms: paradigm shifts and new technologies for research, monitoring, and
management. Annu. Rev. Mar. Sci. 4, 143–176 (2012).
6. Hallegraeff, G. M. etal. Perceived global increase in algal blooms is attributable to
intensiied monitoring and emerging bloom impacts. Commun. Earth Environ. 2, 117 (2021).
7. Smith, V. H. Eutrophication of freshwater and coastal marine ecosystems a global
problem. Environ. Sci. Pollut. Res. 10, 126–139 (2003).
8. Fleming, L . E. etal. Review of Florida red tide and human health effects. Harmful Algae
10, 224–233 (2011).
9. Richlen, M. L ., Morton, S. L., Jamali, E. A., Rajan, A. & Anderson, D. M. The catastrophic
2008–2009 red tide in the Arabian Gulf region, with observations on the identiication
and phylogeny of the ish-killing dinolagellate Cochlodinium polykrikoides. Harmful
Algae 9, 163–172 (2010).
10. Hallegraeff, G. & Bolch, C. Unprecedented toxic algal blooms impact on Tasmanian
seafood industry. Microbiol. Aust. 37, 143–144 (2016).
11. Diaz, R. J. & Rosenberg, R. Spreading dead zones and consequences for marine ecosystems.
Science 321, 926–929 (2008).
12. Barton, A. D., Irwin, A. J., Finkel, Z. V. & Stock, C. A. Anthropogenic climate change drives
shift and shufle in North Atlantic phytoplankton communities. Proc. Natl Acad. Sci. USA
113, 2964–2969 (2016).
13. Gobler, C. J. Climate change and harmful algal blooms: insights and perspective. Harmful
Algae 91, 101731 (2020).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
284 | Nature | Vol 615 | 9 March 2023
Article
14. Zohdi, E. & Abbaspour, M. Harmful algal blooms (red tide): a review of causes, impacts
and approaches to monitoring and prediction. Int. J. Environ. Sci. Technol. 16, 1789–1806
(2019).
15. Wells, M. L. etal. Future HAB science: directions and challenges in a changing climate.
Harmful Algae 91, 101632 (2020).
16. Rabalais, N. N., Turner, R. E., Díaz, R. J. & Justić, D. Global change and eutrophication of
coastal waters. ICES J. Mar. Sci. 66, 1528–1537 (2009).
17. Blondeau-Patissier, D., Gower, J. F., Dekker, A. G., Phinn, S. R. & Brando, V. E. A review of
ocean color remote sensing methods and statistical techniques for the detection,
mapping and analysis of phytoplankton blooms in coastal and open oceans. Prog.
Oceanogr. 123, 123–144 (2014).
18. Wolny, J. L. etal. Current and future remote sensing of harmful algal blooms in the
Chesapeake Bay to support the shellish industry. Front. Mar. Sci. 7, 337 (2020).
19. Stumpf, R. P. etal. Monitoring Karenia brevis blooms in the Gulf of Mexico using satellite
ocean color imagery and other data. Harmful Algae 2, 147–160 (2003).
20. Bernard, S., Kudela, R. M., Robertson Lain, L. & Pitcher, G. Observation of Harmful Algal
Blooms with Ocean Colour Radiometry http://dx.doi.org/10.25607/OBP-1042 (IOCCG, 2021).
21. Hu, C. etal. Red tide detection and tracing using MODIS luorescence data: a regional
example in SW Florida coastal waters. Remote Sens. Environ. 97, 311–321 (2005).
22. Andersen, J. H. etal. Long-term temporal and spatial trends in eutrophication status of
the Baltic Sea. Biol. Rev. 92, 135–149 (2017).
23. Gómez, F. & Boicenco, L. An annotated checklist of dinolagellates in the Black Sea.
Hydrobiologia 517, 43–59 (2004).
24. Townsend, D. W., Pettigrew, N. R. & Thomas, A. C. Offshore blooms of the red tide
dinolagellate, Alexandrium sp., in the Gulf of Maine. Cont. Shelf Res. 21, 347–369 (2001).
25. Pitcher, G. C., Figueiras, F. G., Hickey, B. M. & Moita, M. T. The physical oceanography of
upwelling systems and the development of harmful algal blooms. Prog. Oceanogr. 85,
5–32 (2010).
26. López-Cortés, D. J. etal. The state of knowledge of harmful algal blooms of Margaleidinium
polykrikoides (a.k.a. Cochlodinium polykrikoides) in Latin America. Front. Mar. Sci. 6, 463
(2019).
27. Anderson, D. M. etal. Evidence for massive and recurrent toxic blooms of Alexandrium
catenella in the Alaskan Arctic. Proc. Natl Acad. Sci. USA 118, e2107387118 (2021).
28. Grifith, A. W., Doherty, O. M. & Gobler, C. J. Ocean warming along temperate western
boundaries of the Northern Hemisphere promotes an expansion of Cochlodinium
polykrikoides blooms. Proc. R. Soc. B 286, 20190340 (2019).
29. Conley, D. J. Save the Baltic Sea. Nature 486, 463–464 (2012).
30. Mahadevan, A., D’Asaro, E., Lee, C. & Perry, M. J. Eddy-driven stratiication initiates North
Atlantic spring phytoplankton blooms. Science 337, 54–58 (2012).
31. Chelton, D. B., Gaube, P., Schlax, M. G., Early, J. J. & Samelson, R. M. The inluence of
nonlinear mesoscale eddies on near-surface oceanic chlorophyll. Science 334, 328–332
(2011).
32. Boyce, D. G., Petrie, B., Frank, K. T., Worm, B. & Leggett, W. C. Environmental structuring of
marine plankton phenology. Nat. Ecol. Evol. 1, 1484–1494 (2017).
33. Martínez-Moreno, J. etal. Global changes in oceanic mesoscale currents over the satellite
altimetry record. Nat. Clim. Change 11, 397–403 (2021).
34. Kämpf, J. & Chapman, P. in Upwelling Systems of the World 31–65 (Springer, 2016).
35. Lee, T. N., Yoder, J. A. & Atkinson, L. P. Gulf Stream frontal eddy inluence on productivity
of the southeast US continental shelf. J. Geophys. Res. 96, 22191–22205 (1991).
36. Wolter, K. & Timlin, M. S. El Niño/Southern Oscillation behaviour since 1871 as diagnosed
in an extended multivariate ENSO index (MEI. ext). Int. J. Climatol. 31, 1074–1087 (2011).
37. Glibert, P. M. & Burford, M. A. Globally changing nutrient loads and harmful algal blooms:
recent advances, new paradigms, and continuing challenges. Oceanography 30, 58–69
(2017).
38. Lu, C. & Tian, H. Global nitrogen and phosphorus fertilizer use for agriculture production
in the past half century: shifted hot spots and nutrient imbalance. Earth Syst. Sci. Data 9,
181–192 (2017).
39. Falkowski, P. G., Barber, R. T. & Smetacek, V. Biogeochemical controls and feedbacks on
ocean primary production. Science 281, 200–206 (1998).
40. Wells, M. L. etal. Harmful algal blooms and climate change: Learning from the past and
present to forecast the future. Harmful Algae 49, 68–93 (2015).
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afiliations.
Open Access This article is licensed under a Creative Commons Attribution
4.0 International License, which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons licence,
and indicate if changes were made. The images or other third party material in this article are
included in the article’s Creative Commons licence, unless indicated otherwise in a credit line
to the material. If material is not included in the article’s Creative Commons licence and your
intended use is not permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a copy of this licence,
visit http://creativecommons.org/licenses/by/4.0/.
© The Author(s) 2023
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Methods
Data sources
MODIS on the Aqua satellite provides a global coverage within 1–2
days. All images acquired by this satellite mission from January 2003
to December 2020 were used in our study to detect global coastal
phytoplankton blooms, with a total of 0.76 million images. MODIS
Level-1A images were downloaded from the Ocean Biology Distributed
Active Archive Center (OB.DAAC) at NASA Goddard Space Flight Center
(GSFC), and were subsequently processed with SeaDAS software (ver-
sion 7.5) to obtain Rayleigh-corrected reflectance (R
rc
(dimensionless),
which was converted using the rhos (in sr
−1
) product (rhos × π) from
SeaDAS)
41
, remote sensing reflectance (R
rs
(sr
−1
)) and quality control
flags (l2_flags). If a pixel was flagged by any of the following, it was then
removed from phytoplankton bloom detection: straylight, cloud, land,
high sunglint, high solar zenith angle and high sensor zenith angle
(https://oceancolor.gsfc.nasa.gov/atbd/ocl2flags/). MODIS level-3
product for aerosol optical thicknesses (AOT) at 869 nm was also
obtained from OB.DAAC NASA GSFC (version R2018.0), which was
used to examine the impacts of aerosols on bloom trends.
We examined the algal blooms in the EEZs of 153 ocean-bordering
countries (excluding the EEZs in the Caspian Sea or around the
Antarctic), 126 of which were found with at least one bloom in the past
two decades. The EEZ dataset is available at https://www.marinere-
gions.org/download_file.php?name=World_EEZ_v11_20191118.zip.
The EEZs are up to 200 nautical miles (or 370 km) away from coast-
lines, which include all continental shelf areas and offer the majority of
marine resources available for human use. Regional statistics of algal
blooms were also performed for LMEs. LMEs encompass global coastal
oceans and outer edges of coastal currents areas, which are defined
by various distinct features of the oceans, including hydrology, pro-
ductivity, bathymetry and trophically dependent populations42.
Of the 66 LMEs identified globally, we excluded the Arctic and Ant-
arctic regions and examined 54 LMEs. The boundaries of LMEs were
obtained from https://www.sciencebase.gov/catalog/item/55c7772
2e4b08400b1fd8244.
We used HAEDAT to validate our satellite-detected phytoplankton
blooms in terms of presence or absence. The HAEDAT dataset (http://
haedat.iode.org) is a collection of records of HAB events, maintained
under the UNESCO Intergovernmental Oceanographic Commission
and with data archives since 1985. For each HAB event, the HAEDAT
records its bloom period (ranging from days to months) and geoloca-
tion. We merged duplicate entries when both the recorded locations
and times of the HAEDAT events were very similar to one another, and a
total number of 2,609 HAEDAT events were ultimately selected between
2003 and 2020.
We used the ¼° resolution National Oceanic and Atmospheric
Administration Optimum Interpolated SST (v. 2.1) data to examine the
potential simulating effects of warming on the global phytoplankton
trends. We also estimated the SST gradients following the method of
Martínez-Moreno33. As detailedin ref.33, the SST gradient can be used
as a proxy for the magnitude of oceanic mesoscale currents (EKE).
We used the SST gradient to explore the effects of ocean circulation
dynamics on algal blooms.
Fertilizer uses and aquaculture production for different countries
was used to examine the potential effects of nutrient enrichment
from humans on global phytoplankton bloom trends. Annual data
between 2003 and 2019 on synthetic fertilizer use, including nitrogen
and phosphorus, are available from https://ourworldindata.org/fer-
tilizers. Annual aquaculture production includes cultivated fish and
crustaceans in marine and inland waters, and sea tanks, and the data
between 2003 and 2018 are available from https://ourworldindata.
org/grapher/aquaculture-farmed-fish-production.
The MEI, which combines various oceanic and atmospheric variables36,
was used to examine the connections between El Niño–Southern
Oscillation activities and marine phytoplankton blooms. The dataset
is available from https://psl.noaa.gov/enso/mei/.
Development of an automated bloom detection method
A recent study by the UNESCO Intergovernmental Oceanographic
Commission revealed that globally reported HAB events have
increased
6
. However, such an overall increasing trend was found to
be highly correlated with recently intensified sampling efforts
6
. Once
this potential bias was accounted for by examining the ratio between
HAB events to the number of samplings5, there was no significant
global trend in HAB incidence, though there were increases in certain
regions. With synoptic, frequent, and large-scale observations, satel-
lite remote sensing has been extensively used to monitor algal blooms
in oceanic environments
1719
. For example, chlorophyll a (Chla) con-
centrations, a proxy for phytoplankton biomass, has been provided as
a standard product by NASA since the proof-of-concept Coastal Zone
Color Scanner (1978–1986) era
43,44
. The current default algorithm used
to retrieve Chla products is based on the high absorption of Chla at
the blue band45,46, which often shows high accuracy in the clear open
oceans but high uncertainties in coastal waters. This is because, in
productive and dynamic coastal oceans, the absorption of Chla in the
blue band can be obscured by the presence of suspended sediments
and/or coloured dissolved organic matter (CDOM)47. To address this
problem, various regionalized Chla algorithms have been developed48.
Unfortunately, the concentrations of the water constituents (CDOM,
sediment and Chla) can vary substantially across different coastal
oceans. As a result, a universal Chla algorithm that can accurately
estimate Chla concentrations in global coastal oceans is not currently
available.
Alternatively, many spectral indices have been developed to iden-
tify phytoplankton blooms instead of quantifying their bloom bio-
mass, including the normalized fluorescence line height21 (nFLH),
red tide index
49
(RI), algal bloom index
47
(ABI), red–blue difference
(RBD)50, Karenia brevis bloom index50 (KBBI) and red tide detec-
tion index51 (RDI). In practice, the most important task for these
index-based algorithms is to determine their optimal thresholds
for bloom classification. However, such optimal thresholds can be
regional-or image-specific
20
, due to the complexity of optical fea-
tures in coastal waters and/or the contamination of unfavourable
observational conditions (such as thick aerosols, thin clouds, and
so on), making it difficult to apply spectral-index-based algorithms
at a global scale.
To circumvent the difficulty in determining unified thresholds for
various spectral indices across global coastal oceans, an approach
from a recent study to classify algal blooms in freshwater lakes
52
was
adopted and modified here. In that study, the remotely sensed reflec-
tance data in three visible bands (red, green and blue) were converted
into two-dimensional colour space created by the Commission Interna-
tionale del’éclairage (CIE), in which the position on the CIE chromaticity
diagram represented the colour perceived by human eyes (Extended
Data Fig.1a). As the algal blooms in freshwater lakes were manifested
as greenish colours, the reflectance of bloom-containing pixels was
expected to be distributed in the green gamut of the CIE chromaticity
diagram; the stronger the bloom, the closer the distance to the upper
border of the diagram (the greener the water).
Here, the colour of phytoplankton blooms in the coastal oceans
can be greenish, yellowish, brownish, or even reddish53, owing to
the compositions of bloom species (diatoms or dinoflagellates) and
the concentrations of different water constituents. Furthermore, the
Chla concentrations of the coastal blooms are typically lower than
those in inland waters, thus demanding more accurate classification
algorithms. Thus, the algorithm proposed by Hou etal.
52
was modified
when using the CIE chromaticity space for bloom detection in marine
environments. Specifically, we used the following coordinate conver-
sion formulas to obtain the xy coordinate values in the CIE colour space:
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
xXXYZ
yYXYZ
XRGB
YRGB
ZRGB
=/(++)
=/(++)
=2.7689+1.7517 +1.1302
=1.0000 +4.5907+0.0601
=0.0000 +0.0565+5.5943
(1
)
where R, G and B represent the R
rc
at 748 nm, 678 nm (fluorescence
band) and 667 nm in the MODIS Aqua data, respectively. By contrast,
the R, G and B channels used in Hou etal.52 were the red, green and blue
bands. We used the fluorescence band for the G channel because, for a
given region, the 678 nm signal increases monotonically with the Chla
concentration for blooms of moderate intensity
21
, which is similar to the
response of greenness to freshwater algal blooms. Thus, the converted
y value in the CIE coordinate system represents the strength of the
fluorescence. In practice, for pixels with phytoplankton blooms, the
converted colours in the chromaticity diagram will be located within
the green, yellow or orange–red gamut (see Extended Data Fig.1a); the
stronger the fluorescence signal is, the closer the distance to the upper
border of the CIE diagram (larger y value). By contrast, for bloom-free
pixels without a fluorescence signal, their converted xy coordinates will
be located in the blue or purple gamut. Therefore, we can determine
a lower boundary in the CIE two-dimensional coordinate system to
separate bloom and non-bloom pixels, similar to the method proposed
by Hou etal.52.
We selected 53,820 bloom-containing pixels from the MODIS R
rc
data
as training samples to determine the boundary of the CIE colour space.
These sample points were selected from nearshore waters worldwide
where frequent phytoplankton blooms have been reported (Extended
Data Fig.2); the algal species included various speciesof dinoflagellates
and diatoms
20
. A total of 80 images was used, which were acquired from
different seasons and across various bloom magnitudes, to ensure that
the samples used could almost exhaustively represent the different
bloom conditions in the coastal oceans.
We combined the MODIS FLHRrc (fluorescence line height based on
Rrc) and enhanced red–green–blue composite (ERGB) to delineate
bloom pixels manually. The FLHRrc image was calculated as:
RFRFRF
RF
FLH −[ ×+
−×(678 667)/(748 667)] (2
)
Rrcrc678 678rc667 667 rc748748
rc667 667
where R
rc667
, R
rc678
and R
rc74 8
are the R
rc
at 667, 678 and 748 nm, respec-
tively, and F
667
, F
678
and F
748
are the corresponding extraterrestrial solar
irradiance. ERGB composite images were generated using Rrc of three
bands at 555 (R), 488 (G) and 443 nm (B). Although phytoplankton-rich
andsediment-rich waters have high FLHRrc values, they appear as dark-
ish and bright features in the ERGB images (Extended Data Fig.3),
respectively
21
. In fact, visual examination with fluorescence signals
and ERGB has been widely accepted as a practical way to deline-
ate coastal algal blooms on a limited number of images21,54,55. Note
that the FLHRrc here was slightly different from the NASA standard
nFLH product
56
, as the latter is generated using R
rs
(corrected for both
Rayleigh and aerosol scattering) instead of R
rc
(with residual effects of
aerosols). However, when using the NASA standard algorithm to further
perform aerosol scattering correction over R
rc
, 20.7% of our selected
bloom-containing pixels failed to obtain valid R
rs
(without retrievals
or flagged as low quality), especially for those with strong blooms (see
examples in Extended Data Fig.4). Likewise, we also found various
nearshore regions with invalid Rrs retrievals. By contrast, Rrc had valid
data for all selected samples and showed more coverage in nearshore
coastal waters. The differences between R
rs
and R
rc
were because the
assumptions for the standard atmospheric correction algorithm do
not hold for bloom pixels or nearshore waters with complex optical
properties57. In fact, Rrc has been used as an alternative to Rrs in various
applications in complex waters58,59.
We converted the Rrc data of 53,820 selected sample pixels into
the xy coordinates in the CIE colour space (Extended Data Fig.1a).
As expected, these samples of bloom-containing pixels were located
in the upper half of the chromaticity diagram (the green, yellow and
orange–red gamut) (Extended Data Fig.1a). We determined the lower
boundary of these sample points in the chromaticity diagram, which
represents the lightest colour and thus the weakest phytoplankton
blooms; any point that falls above this boundary represents stronger
blooms. The method to determine the boundary was similar to Hou
etal.
52
: we first binned the sample points according to the x value in
the chromaticity diagram and estimated the 1st percentile (Q1%) of
the corresponding Y for each bin; then, we fit the Q1% using two-order
polynomial regression. Sensitivity analysis with Q
0.3
% (the three-sigma
value) resulted in minor changes (<1%) in the resulting bloom areas for
single images. Notably, sample points were rarely located near white
points (x = 1/3 and y = 1/3, represent equal reflection from three RGB
bands) in the diagram, and we used two polynomial regressions to
determine the boundaries for x values greater and less than 1/3, which
can be expressed as:
yxxx=4.8093−3.0958 +0.8357<
1
3
(3)
12
yxxx=4.9040−3.5759 +0.9862>
1
3
(4)
22
Based on the above, if a pixel’s xy coordinate (converted from R
rc
spectrum) satisfies the conditions of (x < 1/3 AND y > y
1
) or (x > 1/3 AND
y > y2), it is classified as a ‘bloom’ pixel.
Depending on the local region and application purpose, the mean-
ing of ‘phytoplankton bloom’ may differ. Here, for a global applica-
tion, the pixelwise bloom classification is based on the relationship
(represented using the CIE colour space) between Rrc in the 667-, 678-
and 754-nm bands derived from visual interpretation of the 80 pairs
of FLH
Rrc
and ERGB imagery. Instead of a simple threshold, we used a
lower boundary of the sample points in the chromaticity diagram to
define a bloom. In simple words, a pixel is classified as a bloom if its
fluorescence signal is detectable (the associated xy coordinate in the
CIE colour space located above the lower boundary). Histogram of the
nFLH values from the 53,820 training pixels demonstrated the minimum
value of ~0.02 mWcm
−2
μm
−1
(Extended Data Fig.1a), which is in line
with the lower-bound signal of K. brevis blooms on the West Florida
shelf21,47. Note that, such a minimum nFLH is determined from the global
training pixels, and it does not necessarily represent a unified lower
bound for phytoplankton blooms across the entire globe, especially
considering that fluorescence efficiency may be a large variable across
different regions. Different regions may have different lower bounds
of nFLH to define a bloom, and such variability is represented by the
predefined boundary in the CIE chromaticity diagram in our study. Cor-
respondingly, although the accuracy of Chla retrievals may have large
uncertainties in coastal waters, the histogram of the 53,820 training
pixels shows a lower bound of ~1 mg m−3 (Extended Data Fig.1a). Simi-
larly to nFLH, such a lower bound may not be applicable to all coastal
regions, as different regions may have different lower bounds of Chla
for bloom definition.
Although the MODIS cloud (generated by SeaDAS with R
rc869
 < 0.027)
and associated straylight flags can be used to exclude most clouds, we
found that residual errors from thin clouds or cloud shadows could
affect the spectral shape and cause misclassification for bloom detec-
tions. Thus, we designed two spectral indices to remove such effects:
RR RRIndex1 =n −n −(n−n)×0.5 (5)
rc488rc443 rc555 rc443
RR RRIndex2 =n −n −(n−n)×0.5 (6)
rc555 rc469rc645 rc469
Content courtesy of Springer Nature, terms of use apply. Rights reserved
where Index1 and Index2 were used to remove cloud shadows and
clouds, respectively. The nRrc443, nRrc488 and nRrc555 in index1 are the
normalized R
rc
, obtained by normalizing R
rc488
. Similar calculations
were performed for index2. The purpose of normalizations is to elimi-
nate the effect of the absolute magnitude of the reflectance, so that
the thresholds of these two indices are influenced by only the relative
magnitude (spectral shape). We determined thresholds for Index1
(>0.12) and Index2 (<0.012) through trial-and-error and ensured that
the misclassifications caused by residual errors from clouds and cloud
shadows could be effectively removed. After applying the cloud/cloud
shadow and various other masks that are associated with l2_flags, we
obtained an annual mean valid pixel observation (N
vobs
) of ~2.0 × 10
5
for global 1° × 1° grid cells, and the fluctuation patterns and trends of
N
vobs
, either annually or seasonally, are different from that of the global
bloom frequency and affected areas (see Supplementary Fig.1).
Assessments of the algorithm performance
In addition to phytoplankton blooms, macroalgal blooms (Sargassum
and Ulva) frequently occur in many coastal oceans6063. To verify
whether our CIE-fluorescence algorithm could eliminate such impacts,
we compared the spectra between micro-and macroalgal blooms (see
Extended Data Fig.1b). We found that the spectral shapes of Sargassum
and Ulva are substantially different from those of microalgae, particu-
larly for the three bands used for CIE coordinate conversion. The con-
verted xy coordinates for macroalgae were located in the purple–red
gamut of the CIE diagram, which was far below the predefined bound-
ary (Extended Data Fig.1). Moreover, our algorithm is not affected by
highly turbid waters for the following two reasons: first, extremely high
turbidity tends to saturate the MODIS ocean bands
64
, which can be eas-
ily excluded; second, without a fluorescence peak, the reflectance of
unsaturated turbid waters, after conversion to CIE coordinates, will be
located below the predefined boundary (see example in Extended Data
Fig.1b). We also confirmed that the spectral shapes of coccolithophore
blooms are different from dinoflagellates and diatoms (see example in
Extended Data Fig.1b), and thus they are excluded from our algorithm.
Three different types of validation methods were adopted to dem-
onstrate the reliability of the proposed CIE-fluorescence algorithm for
phytoplankton bloom detection in global coastal oceans, including
visual inspections of the RGB, ERGB and FLH
Rrc
images, verifications
using independent manually delineated algal blooms, and comparisons
with the reported HAB events from the HAEDAT dataset.
First, we selected MODIS Aqua images from different locations where
coastal phytoplankton blooms have been recorded in the published
literature. We visually compared the RGB, ERGB, and FLH
Rrc
images, and
our algorithm detected bloom patterns (see examples in Extended Data
Fig.3). Comparisons from various images worldwide showed that our
algorithm could successfully identify regions with high FLHRrc values
and brownish-to-darkish features on the ERGB images, which can be
considered phytoplankton blooms.
Second, we delineated additional 15,466 bloom-containing pixels
from 35 images covering different coastal areas, using the same visual
inspection and manual delineation method as for the training sample
pixels. Moreover, we also selected 14,149 bloom-free pixels (bright or
turquoise green colours on ERGB images or low FLHRrc values) within
the same images as bloom-containing images. We applied our algo-
rithm to all these pixels, and compared the algorithm-identified and
manually delineated results. Our CIE-fluorescence algorithm showed
high values in both producer and user accuracies (92.04% and 98.63%)
(Supplementary Table1), and appeared effective at identifying bloom
pixels and excluding false negatives (blooms classified as non-blooms)
and false positives (non-blooms classified as blooms).
Third, we validated the satellite-detected phytoplankton blooms
using insitu reported HAB events from the HAEDAT dataset. For each
HAB event in the HAEDAT dataset, we obtained all MODIS images over
the reported bloom period (from days to months). Within each year, we
estimated the ratio between the number of satellite images with ‘bloom
detected’ against the number of valid images (see definition above)
during the bloom periods across the entire globe (Supplementary
Table1). Moreover, we calculated the number of events with at least
one successful satellite bloom detection (Ns), and then estimated the
ratio between N
s
and the total HAB events for each year. Results showed
that substantial amounts (averaged at 51.2%) of satellite observations
acquired during the HAB event periods were found with phytoplank-
ton blooms by our algorithm. Overall, 79.3% of the global HAB events
were successfully identified by satellite. The discrepancies between
satellite and insitu observations could be explained by the following
reasons: first, our study focused only on the phytoplankton blooms
that are resolvable by satellite fluorescence signals; other types of HAB
events in the HAEDAT dataset may not have been detectable by satellite
observations, such as events with lower phytoplankton biomass but
high toxicity, occurrences at the subsurface layers, or fluorescence
signals overwhelmed by suspended sediments6567. Second, although
the HAEDAT recorded HAB events could be sustained for long periods,
high biomass of surface algae may not have occurred every day within
this period due to the changes in stratification, precipitation, wind, ver-
tical migration of cells, and many other factors
68
. Third, the spatial scale
of certain HAB events may have been too small to be identified using
the 1-km resolution MODIS observations. Fourth, a reduced MODIS
satellite observation frequency by the contaminations of clouds and
land adjacency effects69. Therefore, we believe the underestimations
of satellite-detected blooms compared to the insitu reported HAB
events were mainly due to inconsistencies between the two observa-
tions rather than uncertainties in our algorithm.
Because R
rc
depends not only on water colour but also on aerosols
(type and concentration) and solar and viewing geometry, a sensitivity
analysis was used to determine whether such variables could impact
bloom detection. Aerosol reflectance (ρ
a
) with different AOTs at 869
nm was simulated using the NASA-recommended maritime aerosol
model (r75f02, with a relative humidity of 75% and a fine-mode frac-
tion of 2%). Then, ρ
a
of each MODIS band was added to R
rc
images, and
the resulting bloom areas with and without added ρ
a
were compared.
Results showed that even with a change of 0.02 in AOT at 869 nm, the
bloom areas showed minor changes (<2%) in the tested images; minor
changes were also found when we used different aerosol models to
conduct ρ
a
simulations
70
. Note that 0.02 represents the high end of
the AOT intra-annual variability in coastal oceans (see Extended Data
Fig.5), and the associated interannual changes are much smaller.
Thus, the use of R
rc
instead of the fully atmospherically corrected
reflectance R
rs
could have limited impacts on our detected global
bloom trend.
We also tried various index-based algorithms developed in previ-
ous studies. However, results showed that all these methods require
image-specific thresholds to accurately determine algal bloom bounda-
ries for different coastal regions (see Extended Data Fig.6). By con-
trast, although our CIE-fluorescence algorithm may lead to different
bloom thresholds for different regions, it can identify bloom pixels
without adjusting the coefficients and, therefore, is more suitable for
global-scale bloom assessment efforts.
We acknowledge that our satellite-detected algal blooms represent
only high amounts of phytoplankton biomass on the ocean surfaces
without distinguishing whether such blooms produce toxins or are
harmful to marine environments. Furthermore, with only limited
spectral information from MODIS, it is difficult to discriminate the
phytoplankton species of algal blooms; such information could help
to improve our understanding of the impacts of these phytoplankton
blooms. However, we expect these challenges to be addressed soon with
the scheduled launch ofthe Plankton,Aerosol,Cloud, oceanEcosystem
(PACE)mission by NASA in 2024, where the hyperspectral measure-
ments over a broad spectrum (350–885 nm) will make species-level
classifications possible71.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
Exploring the patterns and trends of global coastal
phytoplankton blooms
We applied the CIE-fluorescence algorithm to all MODIS Aqua level-2 R
rc
images, and a total number of 0.76 million images between 2003 and
2020 were processed. We mapped all detected blooms into 1-km daily
scale level-3 composites. The number of bloom counts within a year
for each location can be easily enumerated, and the long-term annual
mean values were then estimated (Fig.1a). We further calculated the
total global bloom-affected area (the areas where algal blooms were
detected at least once) for each year and examined their changes over
time (Fig.2b).
We defined bloom frequency (dimensionless) to represent the den-
sity of phytoplankton blooms for a year by integrating the bloom count
and bloom-affected areas within 1°×1° grid cells within that year, which
is expressed as:
n
NMBloom frequency= (7
)
i
n
i
=1
where M
i
is the enumerated bloom count for each 1-km resolution pixel
in a year within one 1° × 1° grid cell, and n represents the associated
number of bloom-affected pixels in the same cell (the number of pixels
with M
i
 > 0), and N is the total number of 1-km MODIS pixels in this grid
cell. We estimated the bloom frequency for each year between 2003 and
2020, and determined the long-term trend over global EEZs through a
linear least-squares regression (see Fig.2a).
Continental and country-level statistics were performed for bloom
count, bloom-affected areas, and bloom frequency (Fig.1b,c and
Supplementary Table2), using boundaries for the EEZs of different
ocean-bordering countries (see above). Similar statistics were also con-
ducted for 54 LMEs (Extended Data Fig.7 and Supplementary Table3).
Correlations with SST and SST gradient
To assess the impacts of climate change on long-term trends in coastal
phytoplankton blooms, we correlated the annual mean bloom fre-
quency and the associated SST and SST gradient in various coastal
current systems for grid cells with significant changes in bloom fre-
quency (Fig.3c). The SST and SST gradient were averaged over the
growth window within a year, assuming that the changes within the
growth window, either in water temperatures or ocean circulations,
play more important roles in the bloom trends compared to other
seasons32.
We determined the growth window of phytoplankton blooms for
each 1° × 1° grid cell (Extended Data Fig.9a) using the following method:
first, we estimated the proportion of cumulative bloom-affected pixels
within the grid cells for a year. Second, a generalized additive model72
was used to determine the shape of the phenological curves (Extended
Data Fig.9b), where a log link function and a cubic cyclic regression
spline smoother were applied73,74. Third, the timing of maximum
bloom-affected areas (TMBAA) was then determined by identifying
the inflection point on the bloom growth curve (Extended Data Fig.9c).
To facilitate comparisons across Northern and Southern Hemispheres,
the year in the Southern Hemisphere was shifted forward by 183 days
(Extended Data Fig.9c). We characterized the similarity of the bloom
growth curve between different grid cells and grouped them into three
distinct clusters using a fuzzy c-means cluster analysis method
75,76
.
We found uniform distributions of the clusters over large geographic
areas. Cluster I is mainly distributed in mid-low latitudes (<45° N and
<30° S), where the maximum bloom-affected areas were expected in the
early period of the year. Cluster II was mostly found in higher latitudes,
with bloom developments (quasi-) synchronized with increases in SST.
Cluster III was detected along the coastlines, where the bloom-affected
areas increase throughout the entire year. In practice, the growth win-
dow for clusters I and III was set as the entire year, and that for cluster II
was set from day 150 to day 270 within the year. We further found that
the TMBAA for cluster II showed small changes over the entire period
(Extended Data Fig.9d), indicating relatively stable phenological cycles
for those phytoplankton blooms32,77.
Reporting summary
Further information on research design is available in theNature Port-
folio Reporting Summary linked to this article.
Data availability
The satellite-based dataset of global coastal algal bloom at 1-km resolu-
tion and the associated code are available at https://doi.org/10.5281/
zenodo.7359262.Source data are provided with this paper.
41. Hu, C. A novel ocean color index to detect loating algae in the global oceans. Remote
Sens. Environ. 113, 2118–2129 (2009).
42. Sherman, K. Adaptive management institutions at the regional level: the case of large
marine ecosystems. Ocean Coast. Manag. 90, 38–49 (2014).
43. Gordon, H. R., Clark, D. K., Mueller, J. L. & Hovis, W. A. Phytoplankton pigments from the
Nimbus-7 coastal zone color scanner: comparisons with surface measurements. Science
210, 63–66 (1980).
44. Moore, J. K. & Abbott, M. R. Phytoplankton chlorophyll distributions and primary
production in the Southern Ocean. J. Geophys. Res. 105, 28709–28722 (2000).
45. Hu, C. etal. Improving satellite global chlorophyll a data products through algorithm
reinement and data recovery. J. Geophys. Res. 124, 1524–1543 (2019).
46. Hu, C., Lee, Z. & Franz, B. Chlorophyll a algorithms for oligotrophic oceans: a novel
approach based on three-band relectance difference. J. Geophys. Res. 117, C01011
(2012).
47. Hu, C. & Feng, L. Modiied MODIS luorescence line height data product to improve
image interpretation for red tide monitoring in the eastern Gulf of Mexico. J. Appl. Remote
Sens. 11, 012003 (2016).
48. Siswanto, E. etal. Empirical ocean-color algorithms to retrieve chlorophyll-a, total
suspended matter, and colored dissolved organic matter absorption coeficient in the
Yellow and East China Seas. J. Oceanogr. 67, 627–650 (2011).
49. Ahn, Y.-H. & Shanmugam, P. Detecting the red tide algal blooms from satellite ocean
color observations in optically complex Northeast-Asia coastal waters. Remote Sens.
Environ. 103, 419–437 (2006).
50. Amin, R. etal. Novel optical techniques for detecting and classifying toxic dinolagellate
Karenia brevis blooms using satellite imagery. Opt. Express 17, 9126–9144 (2009).
51. Shen, F., Tang, R., Sun, X. & Liu, D. Simple methods for satellite identiication of algal
blooms and species using 10-year time series data from the East China Sea. Remote Sens.
Environ. 235, 111484 (2019).
52. Hou, X. etal. Global mapping reveals increase in lacustrine algal blooms over the past
decade. Nat. Geosci. 15, 130–134 (2022).
53. Dierssen, H. M., Kudela, R. M., Ryan, J. P. & Zimmerman, R. C. Red and black tides:
quantitative analysis of water-leaving radiance and perceived color for phytoplankton,
colored dissolved organic matter, and suspended sediments. Limnol. Oceanogr. 51,
2646–2659 (2006).
54. Zhao, J., Temimi, M. & Ghedira, H. Characterization of harmful algal blooms (HABs) in the
Arabian Gulf and the Sea of Oman using MERIS luorescence data. ISPRS J. Photogramm.
Remote Sens. 101, 125–136 (2015).
55. Qi, L. etal. Noctiluca blooms in the East China Sea bounded by ocean fronts. Harmful
Algae 112, 102172 (2022).
56. Behrenfeld, M. J. etal. Satellite-detected luorescence reveals global physiology of
ocean phytoplankton. Biogeosciences 6, 779–794 (2009).
57. Gordon, H. R. Atmospheric correction of ocean color imagery in the Earth Observing
System era. J. Geophys. Res. 102, 17081–17106 (1997).
58. Feng, L., Hou, X., Li, J. & Zheng, Y. Exploring the potential of Rayleigh-corrected relectance
in coastal and inland water applications: a simple aerosol correction method and its merits.
ISPRS J. Photogramm. Remote Sens. 146, 52–64 (2018).
59. Feng, L., Hu, C., Han, X., Chen, X. & Qi, L. Long-term distribution patterns of chlorophyll-a
concentration in China’s largest freshwater lake: MERIS full-resolution observations with a
practical approach. Remote Sens. 7, 275–299 (2015).
60. Xiao, J. etal. An anomalous bi-macroalgal bloom caused by Ulva and Sargassum
seaweeds during spring to summer of 2017 in the western Yellow Sea, China. Harmful
Algae 93, 101760 (2020).
61. Teichberg, M. etal. Eutrophication and macroalgal blooms in temperate and tropical
coastal waters: nutrient enrichment experiments with Ulva spp. Glob. Change Biol. 16,
2624–2637 (2010).
62. Viaroli, P. etal. Nutrient and iron limitation to Ulva blooms in a eutrophic coastal lagoon
(Sacca di Goro, Italy). Hydrobiologia 550, 57–71 (2005).
63. Dierssen, H. M., Chlus, A. & Russell, B. Hyperspectral discrimination of loating mats of
seagrass wrack and the macroalgae Sargassum in coastal waters of Greater Florida Bay
using airborne remote sensing. Remote Sens. Environ. 167, 247–258 (2015).
64. Hu, C. etal. Dynamic range and sensitivity requirements of satellite ocean color sensors:
learning from the past. Appl. Opt. 51, 6045–6062 (2012).
65. Trainer, V. L. etal. Pelagic harmful algal blooms and climate change: lessons from
nature’s experiments with extremes. Harmful Algae 91, 101591 (2020).
66. Mardones, J. I. etal. Disentangling the environmental processes responsible for the
world’s largest farmed ish-killing harmful algal bloom: Chile, 2016. Sci. Total Environ.
766, 144383 (2021).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
67. Gilerson, A. etal. Fluorescence component in the relectance spectra from coastal
waters. II. Performance of retrieval algorithms. Opt. Express 16, 2446–2460 (2008).
68. Lee, J. H., Harrison, P. J., Kuang, C. & Yin, K. in The Environment in Asia Paciic Harbours
(ed. Wolanski, E.) 187–206 (Springer, 2006).
69. Feng, L. & Hu, C. Cloud adjacency effects on top-of-atmosphere radiance and ocean
color data products: a statistical assessment. Remote Sens. Environ. 174, 301–313 (2016).
70. Ahmad, Z. etal. New aerosol models for the retrieval of aerosol optical thickness and
normalized water-leaving radiances from the SeaWiFS and MODIS sensors over coastal
regions and open oceans. Appl. Opt. 49, 5545–5560 (2010).
71. Werdell, P. J. etal. The Plankton, Aerosol, Cloud, Ocean Ecosystem mission: status,
science, advances. Bull. Am. Meteorol. Soc. 100, 1775–1794 (2019).
72. Hastie, T. J. Generalized Additive Models (Routledge, 2017).
73. Wood, S. N. Generalized Additive Models: An Introduction with R (CRC press, 2017).
74. Macgregor, C. J. etal. Climate-induced phenology shifts linked to range expansions in
species with multiple reproductive cycles per year. Nat. Commun. 10, 4455 (2019).
75. Bezdek, J. C. Pattern Recognition with Fuzzy Objective Function Algorithms (Springer,
2013).
76. Bi, S. etal. Optical classiication of inland waters based on an improved fuzzy C-means
method. Opt. Express 27, 34838–34856 (2019).
77. Kheireddine, M., Mayot, N., Ouhssain, M. & Jones, B. H. Regionalization of the Red Sea
based on phytoplankton phenology: a satellite analysis. J. Geophys. Res. 126,
e2021JC017486 (2021).
Acknowledgements We thank NASA for providing global MODIS satellite images, and the
Intergovernmental Oceanographic Commission (IOC) of UNESCO for providing the HAEDAT
dataset. L.F. was supportedby the National Natural Science Foundation of China (no.
41890852, 42271322 and 41971304). D.M.A. was supportedby the Woods Hole Center for
Oceans and Human Health (National Science Foundation grant OCE-1840381 and National
Institutes of Health grants NIEHS-1P01-ES028938-01).
Author contributions Y.D. and S.Y.: methodology, data processing and analyses, and writing.
L.F.: conceptualization, methodology, funding acquisition, supervision and writing. D.Z.: data
processing and analysis. C.H., W.X., D.M.A., Y.L., X.-P.S., D.G.B., L.G. and C.Z. participated in
interpreting the results and reining the manuscript.
Competing interests The authors declare no competing interests.
Additional information
Supplementary information The online version contains supplementary material available at
https://doi.org/10.1038/s41586-023-05760-y.
Correspondence and requests for materials should be addressed to Lian Feng.
Peer review information Nature thanks Bryan Franz, Bingkun Luo and the other, anonymous,
reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are
available.
Reprints and permissions information is available at http://www.nature.com/reprints.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
Extende d Data Fig. 1 | Devel opment of the C IE-fluor escence al gorithm to
detect phytoplankton blooms using MODIS satellite imagery. (a). A1: The
density p lot of manually deli neated bloom -containing p ixels in the CIE
coordinate system (n = 53,820), and the ir distributi on in the CIE color sp ace
(box in A2). A3 : Histograms o f nFLH and Chla for the del ineated pixels , obtained
using NAS A standard algori thms47,57. (b) MODIS true color c omposites an d
selected spect ra for phytoplankton blooms, macroalgal blooms (Ulva and
Sargassum), coccolitho phore blooms, an d sediment-rich turb id waters. The x-y
numbers in dicate their co rresponding p ositions in th e CIE coordinate s ystem.
The black re ctangular b oxes in the three lower p anels highligh t different
spectr al shapes bet ween phytopl ankton blooms an d other features n ear the
fluore scence band . Maps created u sing ArcMap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Extende d Data Fig. 2 | MOD IS-det ected bloo m count withi n certain yea rs
for several coas tal region s with frequ ently repor ted blooms . The MODIS
obser vational year is ann otated with in each panel, an d overlaid points ind icate
insitu re corded harmf ul algal bloom event s from the Harmf ul Algae Event
Databas e (HAEDAT) within the s ame year. The lower rig ht panel shows the
locatio ns of all the HAEDAT record s that were used for algo rithm validatio ns in
this study (Sup plementar y Table1), which also d emonstrate s the increas e in
sampling ef fort in the mos t recent years. Cr eated using Ar cMap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
Extende d Data Fig. 3 | Per formance of t he CIE-fl uorescen ce algorithm for
phytoplankton bloom detection in 12 selected coastal oceans. From left
to right are th e RGB-true color co mposite, ERGB c omposite, FLH Rrc, and the
bloom area ( green pixels) det ected by the CI E-fluoresce nce algorithm . Created
using ArcM ap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Extende d Data Fig. 4 | Exa mples showi ng disadvant ages of usi ng NASA
standard Rrs (i.e., with t he removal of both Rayl eigh and aer osol scatt ering)
in algal bloom detection. From left to righ t are the RGB compos ites, ERGB,
nFLH, and th e bloom areas ( green pixels) dete cted by the CIE-f luorescen ce
algorithm ( based on Rrc, w ithout the removal o f aerosol scat tering). Sub stantial
amounts of i nvalid Rrs retrieval s can be obser ved in the red-en circled areas in
which severe blo oms can be found . Additionally, nFLH show s high values at
cloud edge s (yellow-encircl ed areas), making it chal lenging to use a s imple
threshold to c lassify bloo ms. However, such problem s can be circumvent ed in
our CIE-flu orescence al gorithm. Creat ed using ArcMap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
Extende d Data Fig. 5 | Se nsitivit y analysis of th e impacts of a erosols on
bloom detection. (a) Response s of bloom area (BA) to ch anges in aeroso l
optical thi ckness (AOT). Aero sol reflec tance (ρa) w ith AOTs of 0.01 and 0.02 at
869-nm is simula ted and added to th e MODIS image s, and the resulti ng bloom
areas (g reen pixels) with and w ithout adde d ρa are compared. T he left colum ns
show the RGB com posites, and t he right three c olumns show the blo om areas
under diffe rent AOTs. The percent ages of BA changes a re annotated in t he
panels. (b) The stand ard deviation be tween the 12 m onthly mean valu es of AOT
in global coa stal waters (i.e ., 66.7% of the int ra-annual variab ility), and the
histogra m is shown in (c). Maps create d using ArcMap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Extende d Data Fig. 6 | Com parison of d ifferent i ndex-based algo rithms in
algal bloom detection in various coastal regions. Image-specific thresholds
(annotated w ithin the panel s) are required (label ed within the pa nels) for RI50,
ABI (estima ted with FLHRrc)48, R BD51, KBBI51, and RDI52 to delineate accurate
bloom area s (i.e., high nFLH valu es, which appe ar as bright and d arkish feature s
on the ERGB ima ges). The left panel s are the bloom area s (green pi xels)
extract ed using our CIE-f luorescen ce algorithm. T he RGB-true colo r and ERGB
composit es are shown in Ex tended Data Fi g.3. Created using A rcMap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
Extende d Data Fig. 7 | An nual median b loom count an d the proport ion of
bloom-af fected ar eas for large ma rine ecosys tems (LMEs). (a) Annu al
median bloom count, (b) proportion of blo om-affected a reas. The dat a are
ordered from t he largest to the sm allest. The L MEs are groupe d by continent,
and their nam es, numbers, a nd location s are shown in (a) and (b). Map create d
using Py thon 3.8.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Extende d Data Fig. 8 | Com parison of b loom counts i n the estuari ne and
non-estuarine regions. Boxplot s for long-term mean b loom count in the
estuarin e (n = 13,622 pixel ob servation s) and non-estuari ne (n = 361,604 pixel
obser vations) regions . Compariso n analysis was per formed by two sid ed
Welch’s t-test (P < 0.001).Upper an d lower bounds are f irst and third qua rtiles,
the bar in the mi ddle represent s the median valu e, and the whiskers sh ow the
minimum and ma ximum values . Sixty-two es tuarine zones f rom large rivers
were selec ted, and the bou ndary of each zo ne was manually de lineated
according to high-resolution satellite images.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Article
Extende d Data Fig. 9 | Clu sters of dif ferent bloo m growth paths . (a) The
spatial distribution of different clusters. The fractions of different clusters
across dif ferent latitude s are summarize d. (b) The developm ent of the
maximum blo om-affecte d areas within a ye ar within 1° × 1° grid c ells, where
all global gr id cells are group ed into three dis tinct cluste rs according to the
similarit y of the bloom grow th curve. Th e colored bond cur ves represe nt the
mean value s of all the grid cell s, and their mean S ST and associ ated standard
deviation s are shown with da shed lines and g ray shading. T he proportion s of
different c lusters in the glo bal bloom-affe cted areas are a nnotated. (c) and (f)
The mean t iming of the maxi mum bloom-affec ted areas (T MBAA) and the
associa ted standard de viations bet ween 2003 an d 2019. The whole year in the
Souther n Hemisphere is s hifted forwar d by 183 days in (c). Maps created u sing
Python 3.8.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Extende d Data Fig. 10 | Cha nges in clim ate extremes , global ferti lizer uses ,
and fi shery prod uction over the pa st two decad es. (a) Changes in t he
bi-monthly Mu ltivariate El Niñ o–Southern O scillation (ENS O) index (MEI)
betwee n 2002 and 2020. Po sitive and negat ive MEI values repre sent EI Niño
and La Niña even ts, respec tively. The dots show a nnual mean value s.
(b–c) Trends of nitrogen an d phosphorus f rom 2003 to 2019 for differe nt
countries. (d) Trends of fisher y production f rom 2003 to 2018. G ray indicates
no data. Ma ps created usin g ArcMap 10.4.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1
nature portfolio | reporting summary March 2021
Corresponding author(s): Lian Feng
Last updated by author(s): Nov 23, 2022
Reporting Summary
Nature Portfolio wishes to improve the reproducibility of the work that we publish. This form provides structure for consistency and transparency
in reporting. For further information on Nature Portfolio policies, see our Editorial Policies and the Editorial Policy Checklist.
Statistics
For all statistical analyses, confirm that the following items are present in the figure legend, table legend, main text, or Methods section.
n/a Confirmed
The exact sample size (n) for each experimental group/condition, given as a discrete number and unit of measurement
A statement on whether measurements were taken from distinct samples or whether the same sample was measured repeatedly
The statistical test(s) used AND whether they are one- or two-sided
Only common tests should be described solely by name; describe more complex techniques in the Methods section.
A description of all covariates tested
A description of any assumptions or corrections, such as tests of normality and adjustment for multiple comparisons
A full description of the statistical parameters including central tendency (e.g. means) or other basic estimates (e.g. regression coefficient)
AND variation (e.g. standard deviation) or associated estimates of uncertainty (e.g. confidence intervals)
For null hypothesis testing, the test statistic (e.g. F, t, r) with confidence intervals, effect sizes, degrees of freedom and P value noted
Give P values as exact values whenever suitable.
For Bayesian analysis, information on the choice of priors and Markov chain Monte Carlo settings
For hierarchical and complex designs, identification of the appropriate level for tests and full reporting of outcomes
Estimates of effect sizes (e.g. Cohen's d, Pearson's r), indicating how they were calculated
Our web collection on statistics for biologists contains articles on many of the points above.
Software and code
Policy information about availability of computer code
Data collection The satellite data were obtained from the U.S. National Aeronautics and Space Administration (NASA) Goddard Space Flight Center (GSFC).
Data analysis SeaDAS (Version 7.5) were used to analyze the satellite images.
For manuscripts utilizing custom algorithms or software that are central to the research but not yet described in published literature, software must be made available to editors and
reviewers. We strongly encourage code deposition in a community repository (e.g. GitHub). See the Nature Portfolio guidelines for submitting code & software for further information.
Data
Policy information about availability of data
All manuscripts must include a data availability statement. This statement should provide the following information, where applicable:
- Accession codes, unique identifiers, or web links for publicly available datasets
- A description of any restrictions on data availability
- For clinical datasets or third party data, please ensure that the statement adheres to our policy
The MODIS Aqua data can be obtained from the U.S. National Aeronautics and Space Administration (NASA) Goddard Space Flight Center (GSFC).
The in situ reported HAB data are available from events from http://haedat.iode.org.
The Exclusive economic zones (EEZs) dataset is available at https://www.marineregions.org/download_file.php?name=World_EEZ_v11_20191118.zip.
The boundaries of large marine ecosystems (LMEs) were obtained from https://www.sciencebase.gov/catalog/item/55c77722e4b08400b1fd8244.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
nature portfolio | reporting summary March 2021
Annual data between 2003 and 2019 on synthetic fertilizer use, including nitrogen and phosphorus, are available from https://ourworldindata.org/fertilizers.
Annual aquaculture production includes cultivated fish and crustaceans in marine and inland waters, and sea tanks, and the data between 2003 and 2018 are
available from https://ourworldindata.org/grapher/aquaculture-farmed-fish-production.
The dataset is available from https://psl.noaa.gov/enso/mei/.
Human research participants
Policy information about studies involving human research participants and Sex and Gender in Research.
Reporting on sex and gender Use
the
terms
sex
(biological
attribute)
and
gender
(shaped
by
social
and
cultural
circumstances)
carefully
in
order
to
avoid
confusing both terms. Indicate if findings apply to only one sex or gender; describe whether sex and gender were considered in
study design whether sex and/or gender was determined based on self-reporting or assigned and methods used. Provide in the
source data disaggregated sex and gender data where this information has been collected, and consent has been obtained for
sharing of individual-level data; provide overall numbers in this Reporting Summary. Please state if this information has not
been collected. Report sex- and gender-based analyses where performed, justify reasons for lack of sex- and gender-based
analysis.
Population characteristics Describe
the
covariate-relevant
population
characteristics
of
the
human
research
participants
(e.g.
age,
genotypic
information, past and current diagnosis and treatment categories). If you filled out the behavioural & social sciences study
design questions and have nothing to add here, write "See above."
Recruitment Describe
how
participants
were
recruited.
Outline
any
potential
self-selection
bias
or
other
biases
that
may
be
present
and
how these are likely to impact results.
Ethics oversight Identify
the
organization(s)
that
approved
the
study
protocol.
Note that full information on the approval of the study protocol must also be provided in the manuscript.
Field-specific reporting
Please select the one below that is the best fit for your research. If you are not sure, read the appropriate sections before making your selection.
Life sciences Behavioural & social sciences Ecological, evolutionary & environmental sciences
For a reference copy of the document with all sections, see nature.com/documents/nr-reporting-summary-flat.pdf
Ecological, evolutionary & environmental sciences study design
All studies must disclose on these points even when the disclosure is negative.
Study description This study developed a novel method to map global coastal algal blooms and used this tool to examine satellite images between
2003 and 2020, addressing three fundamental questions: 1) where and how frequently have global coastal oceans been affected by
phytoplankton blooms? 2) have the blooms expanded or intensified over the past two decades, both globally and regionally? and 3)
what are the potential drivers?
Research sample Three separate samples were selected. 1) MODIS Aqua images were used to develop the phytoplankton bloom extraction algorithm,
2) MODIS Aqua images and were used to verify the reliability of the algorithm and the accuracy of the phytoplankton bloom
extraction results, and 3) in situ reported HAB events from the HAEDAT dataset were used to validate the accuracy of the
phytoplankton bloom extraction results.
Sampling strategy A total of 115 MODIS Aqua images were selected from the different locations where coastal phytoplankton blooms have been
recorded in the published literature, of which 80 were used for algorithm development and 35 were used for algorithm validation. A
total number of 2609 HAB events that occurred in the coastal area were selected from the HAEDAT dataset.
Data collection The HAEDAT dataset is a collection of records of harmful algal bloom (HAB) events , maintained under the UNESCO
Intergovernmental Oceanographic Commission and with data archives since 1985.
Timing and spatial scale The satellite data were acquired from different seasons and across various phytoplankton bloom magnitudes between 2003 and
2020, and HAB data from 2003 to 2020 in the HAEDAT dataset were used.
Data exclusions No data were excluded from analysis.
Reproducibility Our results could easily be reproduced with existing datasets.
Randomization Excluding data affected by clouds, a total of 0.76 million MODIS Aqua images from 2003 to 2020 were used to extract phytoplankton
blooms in global coastal area.
Blinding Not applicable in our study.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
nature portfolio | reporting summary March 2021
Did the study involve field work? Yes No
Reporting for specific materials, systems and methods
We require information from authors about some types of materials, experimental systems and methods used in many studies. Here, indicate whether each material,
system or method listed is relevant to your study. If you are not sure if a list item applies to your research, read the appropriate section before selecting a response.
Materials & experimental systems
n/a Involved in the study
Antibodies
Eukaryotic cell lines
Palaeontology and archaeology
Animals and other organisms
Clinical data
Dual use research of concern
Methods
n/a Involved in the study
ChIP-seq
Flow cytometry
MRI-based neuroimaging
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... In recent years, global climate change and intensified anthropogenic activities have contributed to increasing HAB frequency worldwide. Between 2003 and 2020, algal bloom events were documented in over 80% of coastal nations worldwide, with their geographical coverage expanding by 13.2% alongside a 59.1% increase in occurrence frequency [21]. While the coexistence of algae and seagrass is natural in healthy marine ecosystems, excessive algal growth disrupts the ecological balance of seagrass habitats, resulting in cascading negative impacts [19] (Fig. 1). ...
... Elevated surface temperatures intensify water column stratification, creating favorable conditions for algal proliferation by reducing vertical mixing and maintaining algal populations within the photic zone; this thermally induced stratification leads to earlier bloom initiation and extended bloom duration [104]. Furthermore, temperaturedriven alterations in oceanic circulation patterns enhance upwelling intensity, particularly along eastern boundary currents, thereby elevating nutrient flux to surface waters and optimizing stoichiometric conditions (N:P > 16) for algal biomass accumulation [21]. ...
... Concurrently, temperature elevation within the 5-25 °C range enhances enzymatic catalytic efficiency in algal cells, accelerating photosynthetic processes and shortening cellular division cycles [21]. For instance, diatom proliferation rates increase by 33% when temperatures rise from 15 °C to 20 °C, significantly compressing their reproductive timelines [21]. ...
Article
Full-text available
Purpose of Review Harmful algal blooms (HABs) present a growing threat to seagrass ecosystems, significantly impacting their ecological functions and blue carbon potential. Understanding the complex interactions between HABs and seagrasses is crucial for developing adaptive management strategies to protect seagrass ecosystems. Recent Findings Recent studies reveal that global HAB events have significantly expanded both geographically and in frequency over the past two decades. The geomorphological processes and depositional environments of seagrass meadows, along with the effects of climate change, act as contemporary drivers that influence algal invasion, presence, and retention within seagrass ecosystems. Emerging research demonstrates that macroalgal blooms can significantly accelerate seagrass carbon loss by enhancing decomposition rates and increasing greenhouse gas emissions from blue carbon stocks. Seagrass allelopathy and associated algicidal bacteria play crucial roles in natural HAB control. Advanced monitoring techniques combining artificial intelligence with remote sensing have achieved significant improvements in detecting and tracking HAB events and seagrass ecosystems. Summary This review provides a comprehensive analysis of HAB-seagrass interactions, documenting diverse types of HABs affecting seagrass beds, including macroalgal and microalgal blooms. We examine key environmental factors contributing to HABs in seagrass ecosystems, particularly eutrophication, global warming, and ocean acidification, and analyze their complex impact mechanisms, including light limitation, resource competition, biogeochemical alterations, and toxin effects. Natural defense mechanisms of seagrasses, particularly allelopathy and algicidal bacteria, offer potential solutions for HAB control. Effective protection of these valuable blue carbon resources requires integrated adaptive management strategies, combining advanced monitoring technologies, water quality improvement measures, and community-based conservation approaches.
... Hue angle is widely used in color analysis and serves as an intermediate variable for estimating the Forel-Ule index (a traditional index of water color assessment), enabling the assessment of hue variations (23,24). The CIE system provides a standard for digitally expressing water color changes of satellite images and is now used to conduct related studies (22,(24)(25)(26). ...
... The fluctuation amplitude of α* anomaly ( a * a ) exhibits more pronounced fluctuations in coastal (SD > 10°) and highlatitude waters (SD, ~5°), whereas mid-low latitude oceans have relatively lower variability (SD < 3°) (Fig. 2C). Coastal waters, influenced by terrestrial inputs, algal blooms, and human activity (26,31,32), experience pronounced seasonal fluctuations, with a * a variations > 20° in certain regions (e.g., coast of the Bering Sea, the Argentine coast, and the marginal seas of China). ...
... The reported distribution of algal blooms through satellite analyses is similar to identified coastal areas with high hue fluctuations (25), underscoring the potential impact of phytoplankton blooms on water color anomalies. Natural triggers, aquaculture, and fertilizer use in coastal regions seem to contribute to algal blooms (25,26,(68)(69)(70). ...
Article
Full-text available
Ocean change leaves a potentially important imprint on ocean colorimetry. Here, we present an overview and current evaluation of the global ocean color variability from 1998 to 2022, and satellites observe that 36% of oceans (~122 million square kilometers, derived from valid observations) have experienced changes ( P < 0.1). In this context, 25% of the area (formerly blue hue) is turning light blue or green, while the remaining 11% becomes bluer, mainly concentrating in the low-latitude oceans. This study further identifies a “direct” notable impact of both sea surface temperature (SST) and climate on ocean colorimetry tendency and anomaly, especially in the low-latitude oceans. Extreme SST events cause “distinct” ocean colorimetry anomalies, although 94% of cases involve relatively small SST fluctuations. Causal analysis reveals important impacts of climate change on equatorial ocean dynamics, particularly ENSO events. Our findings prove the low-latitude oceans as one of the core changing regions that respond to climate change in the early 21st century.
... Additionally, some HAB-forming species release harmful toxins that accumulate through the food web posing severe health risks to humans and other organisms (Yan et al., 2022). Over the past few decades, the frequency and intensity of HABs have increased globally (Dai et al., 2023), driven by factors such as eutrophication, global warming, and anthropogenic activities (Xiao et al., 2019). To address these challenges, it is essential to develop effective and environmentally friendly strategies to control algal outbreaks and mitigate their associated risks. ...
Article
Full-text available
Harmful algal blooms (HABs) are increasing in frequency and intensity worldwide, posing significant threats to aquatic ecosystems, fisheries, and human health. While chemical algicides are widely used for HABs control due to their rapid efficacy, the lack of systematic data integration and concerns over environmental toxicity limit their broader application. To address these challenges, we developed AlgicideDB, a manually curated database containing 1,672 algicidal records on 542 algicides targeting 110 algal species. Using this database, we analyzed the physicochemical properties of algicides and proposed an algicide-likeness scoring function to facilitate the exploration of compounds with antialgal properties. Additionally, we evaluated the acute toxicity of algicidal compounds to non-target aquatic organisms of different trophic levels to assess their ecological risks. The platform also incorporates a large language model (LLM) enhanced by retrieval-augmented generation (RAG) to address HAB-related queries, supporting decision-making and facilitating knowledge dissemination. AlgicideDB, available at http://algicidedb.ocean-meta.com/#/ , serves as an innovative and comprehensive platform to explore algicidal compounds and facilitate the development of safe and effective HAB control strategies.
... VER 40% of the global population resides in coastal regions, where anthropogenic activities have led to a deterioration in water quality [1]. Water quality is a determining factor in the suitability of water for various applications, including recreation, drinking, fishing, agriculture, and industry [2,3]. ...
Article
Accurate and high spatiotemporal resolution water quality data are critical for the effective management of marine and coastal ecosystems. However, accurate atmospheric correction under high solar zenith angles (SZA) remains a challenge, introducing substantial uncertainties in satellite-derived water quality indicators (WQI) under high SZA. With an attempt to fill the gap, this study evaluated three types of strategies for satellite retrieval of suspended particulate matter (SPM) and chlorophyll-a (Chl-a) concentrations from top-of-atmosphere reflectance (), Rayleigh-corrected reflectance () and remote sensing reflectance (Rrs), respectively. The models, named XGBWQI, based on three types of remote sensing data were tested with in-situ data and compared with the Geostationary Ocean Color Imager (GOCI) standard algorithms. Results showed that: (i)-based XGBWQI had the best accuracy (R 2 = 0.90 and MAPD = 14.65% for SPM, R 2 = 0.85 and MAPD = 5.34% for Chl-a); (ii) model testing results with in-situ data also confirmed the advantage of-based XGBWQI over other models (R 2 = 0.88, MAPD = 26.9% and MRPD =11.8% for SPM, R 2 = 0.78, MAPD = 43.3% and MRPD =-15.5% for Chl-a); and (iii) the XGBWQI models obtained more valid WQI values for GOCI images under high SZA and successfully revealed the diurnal variations of a red tide event in the Yellow Sea and the SPM dynamics in the East China Sea. Therefore,-based XGBWQI models were recommended as the best strategy for satellite retrievals of WQI under high SZA. The methods can serve as an effective tool in retrieving WQI in coastal waters under high SZA, and thus contribute to better and high-frequency water quality monitoring.
Article
Marine mammal skin, in contact with seawater containing diverse chemicals, reflects species health and environmental quality. The contributions of natural toxins and anthropogenic contaminants to the effects of such chemical mixtures remain poorly quantified. Using skin fibroblast cells from the Indo-Pacific finless porpoise and humpback dolphin, we assessed the toxic potential of seawater extracts, focusing on cytotoxicity and intracellular reactive oxygen species (ROS) formation. Among the 38 studied chemicals prevalent in seawater, four algal toxins were 1–6 orders of magnitude more potent than 30 anthropogenic chemicals, including antibiotics, ultraviolet filters, per- and polyfluoroalkyl substances (PFASs), and polyaromatic hydrocarbons. Pectenotoxin-2 accounted for 92% of the cytotoxicity triggered by the mixture of all studied chemicals, which collectively explained 34% of seawater-induced cytotoxicity in porpoise cells. For ROS induction, although all studied chemicals collectively explained a small fraction (<1%) of the effects elicited by seawater extracts in both cell lines, okadaic acid and gymnodimine accounted for ∼80% of the mixture effects of all chemicals, with additional contributions from PFASs. Extending the approach to other coastal habitats where concentration data are available revealed algal toxins as dominant contributors among the known contaminant mixtures eliciting dermal toxic potential. This study provides novel insights to guide the identification of toxicity contributors across dermal health end points, with a balanced perspective on natural toxins and anthropogenic contaminants in addressing their mixture effects on sentinel species health.
Article
Seawater transparency provides critical insight into marine ecological dynamics and serves as a foundational indicator for fisheries management, environmental monitoring, and coastal resource development. Among various indicators, the Secchi disk depth (SDD) is widely used to quantify seawater transparency in marine environmental monitoring. This study develops a remote sensing inversion model for estimating the SDD in the coastal waters of Qinhuangdao, utilizing Sentinel-3 OLCI satellite imagery and in situ measurements. The model is based on the CIE hue angle and demonstrates high accuracy (R2 = 0.93, MAPE = 7.88%, RMSE = 0.25 m), outperforming traditional single-band, band-ratio, and multi-band approaches. Using the proposed model, we analyzed the monthly and interannual variations of SDD in Qinhuangdao’s coastal waters from 2018 to 2024. The results reveal a clear seasonal pattern, with SDD values generally increasing and then decreasing throughout the year, primarily driven by the East Asian monsoon and other natural factors. Notably, the average annual SDD in 2018 was significantly lower than in subsequent years (2019–2024), which is closely associated with comprehensive water management and pollution reduction initiatives in the Bohai Sea region. These findings highlight marked improvements in the coastal marine environment and underscore the benefits of China’s ecological civilization strategy, particularly the principle that “lucid waters and lush mountains are invaluable assets.”
Article
Full-text available
Algal blooms constitute an emerging threat to global inland water quality, yet their spatial and temporal distribution at the global scale remains largely unknown. Here we establish a global bloom database, using 2.91 million Landsat satellite images from 1982 to 2019 to characterize algal blooms in 248,243 freshwater lakes, representing 57.1% of the global lake area. We show that 21,878 lakes (8.8%) spread across six continents have experienced algal blooms. The median bloom occurrence of affected lakes was 4.6%, but this frequency is increasing; we found increased bloom risks in the 2010s, globally (except for Oceania). The most pronounced increases were found in Asia and Africa, mostly in developing countries that remain reliant on agricultural fertilizer. As algal blooms continue to expand in scale and magnitude, this baseline census will be vital towards future risk assessments and mitigation efforts.
Article
Full-text available
The current average state of Red Sea phytoplankton phenology needs to be resolved in order to study future variations that could be induced by climate change. Moreover, a regionalization of the Red Sea could help to identify areas of interest and guide in situ sampling strategies. Here, a clustering method used 21 years of satellite surface chlorophyll‐a concentration observations to characterize similar regions of the Red Sea. Four relevant phytoplankton spatiotemporal patterns (i.e., bio‐regions) were found and linked to biophysical interactions occurring in their respective areas. Two of them, located in the northern part the Red Sea, were characterized by a distinct winter‐time phytoplankton bloom induced by mixing events or associated with a convergence zone. The other two, located in the southern regions, were characterized by phytoplankton blooms in summer and winter which might be under the influence of water advected into the Red Sea from the Gulf of Aden in response to the seasonal monsoon winds. Some observed inter‐annual variabilities in these bio‐regions suggested that physical mechanisms could be highly variable in response to variations in air‐sea heat fluxes and ENSO phases in the northern and southern half of the Red Sea, respectively. This study reveals the importance of sustaining in situ measurements in the Red Sea to build a full understanding about the physical processes that contribute to phytoplankton production in this basin.
Article
Full-text available
Significance The neurotoxin-producing dinoflagellate Alexandrium catenella is shown to be distributed widely and at high concentrations in bottom sediments and surface waters of the Alaskan Arctic. Future blooms are likely to be large and frequent given hydrographic and bathymetric features that support high cell and cyst accumulations, and warming temperatures that promote bloom initiation from cysts in bottom sediments and cell division in surface waters. As the region undergoes an unprecedented regime shift, the exceptionally widespread and dense cyst and cell distributions represent a significant threat to Arctic communities that are heavily dependent upon subsistence harvesting of marine resources. These observations also highlight how warming can facilitate range expansions of harmful algal bloom species into waters where temperatures were formerly unfavorable.
Article
Full-text available
Global trends in the occurrence, toxicity and risk posed by harmful algal blooms to natural systems, human health and coastal economies are poorly constrained, but are widely thought to be increasing due to climate change and nutrient pollution. Here, we conduct a statistical analysis on a global dataset extracted from the Harmful Algae Event Database and Ocean Biodiversity Information System for the period 1985–2018 to investigate temporal trends in the frequency and distribution of marine harmful algal blooms. We find no uniform global trend in the number of harmful algal events and their distribution over time, once data were adjusted for regional variations in monitoring effort. Varying and contrasting regional trends were driven by differences in bloom species, type and emergent impacts. Our findings suggest that intensified monitoring efforts associated with increased aquaculture production are responsible for the perceived increase in harmful algae events and that there is no empirical support for broad statements regarding increasing global trends. Instead, trends need to be considered regionally and at the species level.
Article
Full-text available
Oceanic mesoscale eddies play a profound role in mixing tracers such as heat, carbon and nutrients, thereby regulating regional and global climate. Yet, it remains unclear how the eddy field has varied over the past few decades. Furthermore, climate model predictions generally do not resolve mesoscale eddies, which could limit their accuracy in simulating future climate change. Here we show a global statistically significant increase of ocean eddy activity using two independent observational datasets of surface mesoscale eddy variability (one estimates surface currents, and the other is derived from sea surface temperature). Maps of mesoscale variability trends show heterogeneous patterns, with eddy-rich regions showing a significant increase in mesoscale variability of 2–5% per decade, while the tropical oceans show a decrease in mesoscale variability. This readjustment of the surface mesoscale ocean circulation has important implications for the exchange of heat and carbon between the ocean and atmosphere.
Article
Full-text available
Harmful algal bloom (HAB) species in the Chesapeake Bay can negatively impact fish, shellfish, and human health via the production of toxins and the degradation of water quality. Due to the deleterious effects of HAB species on economically and environmentally important resources, such as oyster reef systems, Bay area resource managers are seeking ways to monitor HABs and water quality at large spatial and fine temporal scales. The use of satellite ocean color imagery has proven to be a beneficial tool for resource management in other locations around the world where high-biomass, nearly monospecific HABs occur. However, remotely monitoring HABs in the Chesapeake Bay is complicated by the presence of multiple, often co-occurring, species and optically complex waters. Here we present a summary of common marine and estuarine HAB species found in the Chesapeake Bay, Alexandrium monilatum, Karlodinium veneficum, Margalefidinium polykrikoides, and Prorocentrum minimum, that have been detected from space using multispectral data products from the Ocean and Land Colour Imager (OLCI) sensor on the Sentinel-3 satellites and identified based on in situ phytoplankton data and ecological associations. We review how future hyperspectral instruments will improve discrimination of potentially harmful species from other phytoplankton communities and present a framework in which satellite data products could aid Chesapeake Bay resource managers with monitoring water quality and protecting shellfish resources.
Article
Full-text available
Climate change is transforming aquatic ecosystems. Coastal waters have experienced progressive warming, acidification, and deoxygenation that will intensify this century. At the same time, there is a scientific consensus that the public health, recreation, tourism, fishery, aquaculture, and ecosystem impacts from harmful algal blooms (HABs) have all increased over the past several decades. The extent to which climate change is intensifying these HABs is not fully clear, but there has been a wealth of research on this topic this century alone. Indeed, the United Nations' Intergovernmental Panel on Climate Change's (IPCC) Special Report on the Ocean and Cryosphere in a Changing Climate (SROCC) approved in September 2019 was the first IPCC report to directly link HABs to climate change. In the Summary for Policy Makers, the report made the following declarations with "high confidence": In addition, the report specifically outlines a series of linkages between heat waves and HABs. These statements about HABs and climate change and the high levels of confidence ascribed to them provides clear evidence that the field of HABs and climate change has matured and has, perhaps, reached a first plateau of certainty. While there are well-documented global trends in HABs being promoted by human activity, including climate change, individual events are driven by local, regional, and global drivers, making it critical to carefully evaluate the conditions and responses at appropriate scales. It is within this context that the first Special Issue on Climate Change and Harmful Algal Blooms is published in Harmful Algae.
Article
A recent study demonstrated the possibility of satellite-based detection of surface blooms of the heterotrophic dinoflagellate, red Noctiluca scintillans (RNS). The study further documented RNS bloom patterns in the East China Sea (ECS) between 2000 and 2017. Here, complemented by more recent satellite observations between 2018 and 2020, the 21-year bloom record shows that while bloom distributions vary in different years and annual cumulative bloom footprint shows an increasing trend, the 21-year cumulative bloom footprint is bounded by major ocean fronts such as the Kuroshio Front. Of all observations, 2020 is a critical year to “complete” the footprint as extra discharge from the 2020 Yangtze River flood event, combined with ocean currents, transports the bloom to the most northeast location of the footprint although the riverine influence reaches at least 128°E, well beyond the RNS footprint.
Article
The dictyochophyte microalga Pseudochattonella verruculosa was responsible for the largest farmed fish mortality ever recorded in the world, with losses for the Chilean salmon industry amounting to US$ 800M in austral summer 2016. Super-scale climatic anomalies resulted in strong vertical water column stratification that stimulated development of a dynamic P. verruculosa thin layer (up to 38 μg chl a L⁻¹) for several weeks in Reloncaví Sound. Hydrodynamic modeling (MIKE 3D) indicated that the Sound had extremely low flushing rates (between 121 and 200 days) in summer 2016. Reported algal cell densities of 7,000 - 20,000 cells mL⁻¹ generated respiratory distress in fish that was unlikely due to low dissolved oxygen (permanently >4 mg L⁻¹). Histological examination of salmon showed that gills were the most affected organ with significant tissue damage and circulatory disorders. It is possible that some of this damage was due to a diatom bloom that preceded the Pseudochattonella event, thereby rendering the fish more susceptible to Pseudochattonella. No correlation between magnitude of fish mortality and algal cell abundance nor fish age was evident. Algal cultures revealed rapid growth rates and high cell densities (up to 600,000 cells mL⁻¹), as well as highly complex life cycle stages that can be easily overlooked in monitoring programs. In cell-based bioassays, Chilean P. verruculosa was only toxic to the RTgill-W1 cell line following exposures to high cell densities of lysed cells (>100,000 cells mL⁻¹). Fatty acid profiles of a cultured strain showed elevated concentrations of potentially ichthyotoxic, long-chain polyunsaturated fatty acids (PUFAs) (69.7% ± 1.8%)- stearidonic (SDA, 18:4ω3 – 28.9%), and docosahexaenoic acid (DHA, 22:6ω3 - 22.3%), suggesting that lipid peroxidation may help to explain the mortalities, though superoxide production by Pseudochattonella was low (< 0.21 ± 0.19 pmol O2⁻ cell⁻¹ h⁻¹). It therefore remains unknown what the mechanisms of salmon mortality were during the Pseudochattonella bloom. Multiple mitigation strategies were used by salmon farmers during the event, with only delayed seeding of juvenile fish into the cages and towing of cages to sanctuary sites being effective. Airlift pumping, used effectively against other fish-killing HABs in the US and Canada was not effective, perhaps because it brought subsurface layers of Pseudochattonella to the surface, or and it also may have lysed the fragile cells, rendering them more lethal. The present study highlights knowledge gaps and inefficiency of contingency plans by the fish farming industry to overcome future fish-killing algal blooms under future climate change scenarios. The use of new technologies based on molecular methods for species detection, good farm practices by fish farms, and possible mitigation strategies are discussed.
Article
An unprecedented bi-macroalgal bloom caused by Ulva prolifera and Sargassum horneri occurred from spring to summer of 2017 in the western Yellow Sea (YS) of China, where annual large-scale green tides have prevailed for a decade. The distinct genesis and blooming dynamics of the two seaweed species were detected and described. Unlike the consistent raft-origin of the floating Ulva biomass, the massive pelagic S. horneri was derived from multiple sources (residual seaweeds from the previous winter bloom and those drifting from offshore water in the south). The scale of the green tide in 2017 was found smaller than the previous four years. We then discussed a number of hypotheses attributing to this reduction, including reduced epiphytic green algae from aquaculture rafts and the influences of the massive pelagic S. horneri. However, further research is needed to identify the origin of the pelagic S. horneri in the western YS and any affiliations with the benthic populations, and to elucidate the interactions of this species with the annual green tides and the ensuing consequences.