23
Far-field impacts of a super typhoon on upper ocean phytoplankton dynamics Guoqiang Qiu 1, Xiaogang Xing 2, Fei Chai 2,3 , Xiao-Hai Yan 1,4 , Zhiyu Liu 1 , and Haili Wang 1* 1 1 State Key Laboratory of Marine Environmental Science, and Department of Physical Oceanography, 2 College of Ocean and Earth Sciences, Xiamen University, Xiamen, China 3 2 State Key Laboratory of Satellite Ocean Environment Dynamics, Second Institute of Oceanography, 4 Ministry of Natural Resources, Hangzhou, China 5 3 School of Marine Sciences, University of Maine, Orono, ME, USA 6 4 College of Earth, Ocean, and Environment, University of Delaware, Newark, DE, USA 7 These authors have contributed equally to this work 8 9 10 This manuscript has a main body of 4397 words, 6 figures and 0 table. 11 12 13 14 15 * Correspondence: 16 Haili Wang 17 [email protected] 18 19 20 21 In review

Far-field impacts of a super typhoon on upper ocean

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Far-field impacts of a super typhoon on upper ocean

Far-field impacts of a super typhoon on upper ocean phytoplankton

dynamics

Guoqiang Qiu1†, Xiaogang Xing2†, Fei Chai2,3, Xiao-Hai Yan1,4, Zhiyu Liu1, and Haili Wang1* 1

1State Key Laboratory of Marine Environmental Science, and Department of Physical Oceanography, 2

College of Ocean and Earth Sciences, Xiamen University, Xiamen, China 3

2State Key Laboratory of Satellite Ocean Environment Dynamics, Second Institute of Oceanography, 4

Ministry of Natural Resources, Hangzhou, China 5

3School of Marine Sciences, University of Maine, Orono, ME, USA 6

4College of Earth, Ocean, and Environment, University of Delaware, Newark, DE, USA 7

†These authors have contributed equally to this work 8

9

10

This manuscript has a main body of 4397 words, 6 figures and 0 table. 11

12

13

14

15

* Correspondence: 16

Haili Wang 17

[email protected] 18

19

20

21

In review

Page 2: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

2

This is a provisional file, not the final typeset article

Abstract 22

Super typhoon Rammasun (2014) traveled across the South China Sea on 07/16–07/18. Its far-field 23

impacts on phytoplankton dynamics in the upper ocean were documented by a Biogeochemical-Argo 24

(BGC-Argo) float located 200 km to the left of its track. Both surface chlorophyll-a concentration 25

(Chla) and particulate backscattering coefficient (bbp) were observed to increase substantially within 26

two distinct stages. The initial increase occurred during the passage of the typhoon, and the 27

subsequent increase happened 5 days after the typhoon. In contrast, depth-integrated Chla and bbp in 28

the upper 150 m underwent negligible changes throughout the entire period. The key lies in the fact 29

that surface phytoplankton increases in the far-field region resulted from the physically driven 30

vertical redistribution of particles, rather than from biological alternations. The first increase was 31

attributed to the typhoon-induced strong turbulence which deepened the surface mixed layer, and 32

thus entrained subsurface particles to the surface; the second was due to the post-typhoon adiabatic 33

quasi-geostrophic adjustment of the upper ocean that gradually raised the isopycnals (and thus 34

subsurface particles). These results challenge the prevailing wisdom on typhoon impacts, and thus 35

shed new lights on the nature of the upper ocean responses to typhoons from both physical and 36

biological perspectives. 37

38

Keywords: 39

particle redistribution, Typhoon, phytoplankton dynamics, far-field impacts, diapycnal mixing, 40

adiabatic quasi-geostrophic adjustment 41

42

1 Introduction 43

Typhoons (or hurricanes) are extreme weather events that have significant impacts on physical and 44

biogeochemical processes in the underlying upper ocean. Typhoons can induce strong mixing, 45

upwelling, near-inertial oscillations, mixed layer deepening and surface cooling (Dickey et al., 1998; 46

Price, 1981; Sanford et al., 2011; Stramma et al., 1986), and the induced baroclinic oscillations can 47

propagate outward far away from the typhoon track, leading to large scale changes in the thermocline 48

(Geisler, 1970). These physical responses can further impact biogeochemical processes in the upper 49

ocean. There are many studies reporting that surface chlorophyll-a concentration (Chla) substantially 50

increased after the passage of a typhoon (Babin et al., 2004; Black and Dickey, 2008; Lin et al., 51

2003), and two primary mechanisms have been proposed to explain such a phenomenon. The first 52

one highlights a direct physical process that the typhoon-induced strong mixing and upwelling can 53

bring subsurface phytoplankton into the surface mixed layer (i.e., the so-called particle entrainment 54

mechanism) (Chai et al., 2021). The other one is mainly a biological process that typhoons can pump 55

more nutrients into the upper layers from depth (i.e., the nutrient entrainment mechanism), which in 56

turn promote the growth of surface phytoplankton (Lin et al., 2003). Moreover, a two-stage increase 57

of the surface Chla has been identified by some studies, that is, an initial rapid increase attributed to 58

direct physical process, and a subsequent one mainly due to the biological response to increased 59

nutrient supply (Shang et al., 2008; Pan et al., 2017). 60

Not surprisingly, the extent of the surface Chla increase depends on the forcing factors including the 61

typhoon intensity, its translation speed, and the pre-typhoon ocean state (Lin, 2012). In general, 62

strong and slow typhoons inject more energy into the ocean, thereby facilitating the surface 63

In review

Page 3: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

3

phytoplankton increase; whereas strong upper ocean stratification diminishes vertical mixing and 64

thus inhibits the surface phytoplankton increase. These responses are usually asymmetric with 65

respect to the typhoon track. A greater Chla increase is often observed under or to the right of the 66

typhoon track, and it is usually attributed to the rightward stronger mixing due to both larger wind 67

speed and the stronger wind-current resonance (Babin et al., 2004). To date, most studies have 68

concentrated on this strong phytoplankton increase immediately under or to the right of the typhoon 69

track; while little attention has been given to upper ocean responses in more distant regions of the 70

direct typhoon influence, especially to the left of the typhoon track. It remains unclear as to what 71

mechanisms are responsible for the observed increase of surface phytoplankton in these regions due 72

to a lack of vertical profiling measurements. 73

Owing to the rough weather (and sea) conditions, the typhoon-induced changes of phytoplankton in 74

the upper ocean are hardly observable through either shipboard sampling or satellite remote sensing. 75

Thus, most studies have been based on the analysis of satellite remote sensed surface Chla several 76

days after the typhoon’s passage (e.g., Zhao et al., 2013). Recently, a new type of profiling platform, 77

Biogeochemical-Argo (BGC-Argo) float, has been developed, making all-weather physical and 78

biogeochemical observations at both high vertical (<5 m) and temporal (<5 d) resolutions possible 79

(Johnson et al., 2009). To date, two cases of phytoplankton response to a typhoon based on the 80

observations using BGC-Argo floats have been reported. After the weak but slow-moving tropical 81

cyclone Hudhud travelled across the Bay of Bengal, both surface and subsurface Chla was observed 82

to greatly increase under the storm track (Chacko, 2017). This bloom was attributed to the combined 83

effects of the particle entrainment and nutrient entrainment mechanisms, via strong mixing and 84

upwelling favored by sufficient wind-ocean interaction and a pre-existing cyclonic eddy. By contrast, 85

the strong but fast-moving typhoon Trami passed over the western Pacific, causing an increase in 86

surface Chla while a decrease in subsurface Chla to the right of the typhoon track, but no net increase 87

in the depth-integrated Chla (Chai et al., 2021). These observations demonstrate that the 88

phytoplankton responses depend on both the characteristics of the typhoon and the state of the 89

underlying ocean, and illustrate the great value of BGC-Argo float profiling observations as well. 90

The super typhoon Rammasun (2014) crossed the South China Sea (SCS) on 07/16–07/18, and the 91

vertical physical and biological changes were well observed by a BGC-Argo float located 200 km to 92

the left of the typhoon track with high vertical (2 m) and temporal (1 d) resolutions (Figure 1). The 93

observations showed that surface phytoplankton increased during and shortly after the typhoon’s 94

passage, but the change of the depth-integrated Chla in the upper 150 m was negligible during the 95

entire period. These unique observations provide a rare opportunity to examine the upper ocean 96

phytoplankton dynamics response to typhoon in the far-field region of typhoon influence. This work 97

extends the previous studies limited to surface phytoplankton changes, and seeks to shed light on the 98

physical and biological responses of the entire upper ocean to typhoon. 99

2 Materials and Methods 100

2.1 Float Data 101

The BGC-Argo float (Sea-Bird Navis BGC-i, No. 0348) was deployed at 14°N, 116°E in the central 102

SCS basin, more than 400 km away from the nearest coast (Figure 1). The float was equipped with a 103

CTD (SBE 41CP), a dissolved oxygen sensor (SBE 63), and a three-channel optical sensor 104

(MCOMS) to measure Chla, fluorescent dissolved organic matter, and the particulate backscattering 105

coefficient at 700 nm (bbp). It profiled every day with 2 m sampling resolution and surfaced at 106

midnight. The Chla and bbp profiles were smoothed by a 5-point running median filter to remove 107

In review

Page 4: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

4

This is a provisional file, not the final typeset article

unexpected spikes. In order to identify overall temporal changes, high-frequency variations were 108

removed using a 3-d moving average. The BGC-Argo float data are provided by Xiamen University 109

(http://121.192.191.43/BioArgo/0348_databin.txt). 110

2.2 Satellite Data 111

The typhoon track data was obtained from the Joint Typhoon Warning Center (JTWC, 112

http://www.usno.navy.mil/JTWC); six-hourly wind speed and stress data with a spatial resolution of 113

0.25°×0.25° were extracted from the Global Blended Mean Wind Fields product provided by 114

Copernicus Marine Environment Monitoring Service (CMEMS, http://marine.copernicus.eu/); three-115

hourly precipitation data with a spatial resolution of 0.25°×0.25° were extracted from the 3B42 116

product provided by the Tropical Rain Measuring Mission (TRMM) project 117

(ftp://trmmopen.gsfc.nasa.gov/pub/merged); daily sea level anomaly (SLA) data with a resolution of 118

0.25°×0.25° were extracted from the Global Ocean Gridded L4 Sea Surface Height and Derived 119

Variables Reprocessed product provided by CMEMS (http://marine.copernicus.eu/); daily satellite 120

sea surface temperature (SST) data with a spatial resolution of 0.25°×0.25° were extracted from the 121

Microwave-only Optimally Interpolated SST product from Remote Sensing Systems (RSS) 122

(http://www.remss.com/measurements/sea-surface-temperature); daily and 8-d composited Chla and 123

daily surface photosynthetically available radiation (PAR) data with a spatial resolution of 4 km×4 124

km were obtained from the GlobColour project (https://www.globcolour.info/). 125

2.3 Derived Data 126

Potential density (σ0) was calculated from the CTD measured pressure, temperature and conductivity 127

(salinity) based on the thermodynamic equation of Seawater-2010 (TEOS-10; McDougall and 128

Barker, 2011). The surface mixed layer depth (MLD) was calculated using a density threshold of 129

0.125 kg m-3 from the near-surface value at 10 m (Levitus, 1982). The deep Chla and bbp maximums 130

(DCM and DBM) were defined as their vertical maximum in each profile. The daily anomaly for 131

each variable was defined as the difference between the float-observed value and the pre-typhoon 132

value on 07/14. 133

Ekman pumping velocity (wE) was calculated as: 134

𝒘𝑬 =𝟏

𝒇𝝆(𝝏𝝉𝒚

𝝏𝒙−𝝏𝝉𝒙𝝏𝒚

) 135

where τx and τy are the eastward and northward wind stresses, f is the Coriolis parameter, and ρ is the 136

density of seawater. 137

The displacement of isotherms (∆h) in the pycnocline was calculated from the change in SLA (∆η) 138

using the reduced gravity approximation (Shay et al., 2000): 139

∆𝒉 = −𝒈

𝒈′∆𝜼 140

where 𝑔′ =Δ𝜌

𝜌𝑔, ∆ρ is the density increase from the surface mixed layer to the pycnocline base, and 141

g is gravity acceleration. 142

In review

Page 5: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

5

According to the two-layer linear ocean model (Geisler, 1970), the first baroclinic wave speed (C1) 143

was calculated as: 144

𝑪𝟏 = √𝒈′𝑯𝟏 145

where H1 is the upper layer thickness, which is practically indicated by the thermocline depth (Jaimes 146

and Shay, 2009). In the SCS basin, the depth of the 22 °C isotherm can be used as a good proxy for 147

the thermocline (Liu et al., 2001). 148

The local eddies were identified based on the daily SLA data following the approach of Chaigneau et 149

al. (2009). Firstly, the SLA is plotted with a uniform interval of 1 cm; the geometric center within the 150

innermost enclosed contour is considered as the center of a possible eddy; all the enclosed contours 151

vary monotonously outward from the eddy center belong to this eddy, and the outmost contour is 152

taken as the eddy edge; finally, only the detected eddies with the amplitude difference between the 153

center and edge larger than 2 cm are reserved. 154

3 Results and Discussion 155

3.1 Surface Responses Revealed by Satellite Observations 156

The category 5 (on the Saffir-Simpson Hurricane Wind Scale) typhoon Rammasun (2014) crossed 157

the SCS at a translation speed of ~6.5 m s-1 on 07/16–07/18. Satellite-observed SST was >28 °C over 158

the basin before the typhoon (Figure 2A), but it decreased immediately afterward (Figure 2B). The 159

surface cooling was attributed to the typhoon-induced strong mixing, upwelling and air–sea heat 160

exchange (Price, 1981). The cooling was asymmetric across the typhoon track, with the largest 161

decrease of −5.6 °C to the right of the typhoon track and about a factor of 3 less at the same distance 162

to the left. The asymmetric response was due to more intense mixing by both larger winds and 163

stronger wind–current resonance to the right of the typhoon track (Price, 1981). Before the typhoon, 164

satellite-observed Chla was ~0.1 mg m-3 (Figure 2C), which is typical for an oligotrophic region 165

(Ning et al., 2004). Ocean color sensors were largely ineffective due to the heavy cloud cover during 166

the typhoon. However, the 8-d composited observation showed a basin-scale increase in surface Chla 167

after the passage of the typhoon (Figure 2D). The increase in surface Chla, which coincided with the 168

surface cooling, was more pronounced to the right of the typhoon track. 169

The BGC-Argo float was located approximately 200 km to the left of the typhoon track (Figure 1), 170

~6 times of the radius of maximum wind, thus its observations represented a far-field responses of 171

the upper ocean to the typhoon. At the float location, the wind speed was very low (<6 m s-1) before 172

the typhoon, and increased rapidly during the typhoon’s passage, with a maximum value of 17 m s-1 173

on 07/17, and then decreased to the pre-typhoon level on 07/19 (Figure 3A). Precipitation was heavy 174

(18.2 mm h-1) during the passage, and continuous during 07/20–07/23 after the typhoon with a mean 175

of 1.7 mm h-1. SLA decreased from 7 cm before the typhoon to 2.6 cm on 07/25. Surface PAR 176

reached 57 Einstein m-2 d-1 on 07/14, which suggests there was intense solar radiation before the 177

typhoon; however, solar radiation was unobservable during the passage of the typhoon due to the 178

heavy cloud cover. The PAR recovered to the pre-typhoon level on 07/18–07/19 just after the 179

typhoon, but it became very low (8.4–21.8 Einstein m-2 d-1) again during 07/20–07/23 due to the 180

continuous precipitation (Figure 3A). 181

3.2 Physical Responses of the Upper Ocean as Observed by the BGC-Argo Float 182

In review

Page 6: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

6

This is a provisional file, not the final typeset article

Before the super typhoon Rammasun, the surface mixed layer was quite shallow (~26 m) due to weak 183

wind stirring and strong solar radiation (Figures 3A, 4A). Below the MLD, both the temperature and 184

σ0 profiles had large vertical gradients, approximately 0.1 °C m-1 and 0.05 kg m-4 (Figures 4A,E), 185

which implies strong subsurface stratification. The seasonal thermocline depth, indicated by the 186

depth of the 22 °C isotherm (Liu et al., 2001), reached 92 m. These vertical structures agreed well 187

with the climatological observations (Qu et al., 2007). 188

During the passage of the typhoon, MLD increased abruptly to 34 m due to wind stirring (Figure 189

4A). Consequently, upper thermocline water was entrained into the surface mixed layer: SST 190

decreased by 0.7 °C, and sea surface salinity (SSS) increased by 0.031 psu despite heavy 191

precipitation (Figure 3B). Although MLD showed only a modest increase of 8 m, the temperature, 192

salinity and σ0 all showed remarkable changes below the MLD (Figures 4B,D,F). The temperature 193

anomaly was largely negative above 75 m but positive below that depth, and the salinity anomaly 194

showed an opposite trend. The lower colder but saltier water was brought into the upper layer, while 195

the upper warmer and fresher water was mixed into the lower layer. Mechanical energy injected by 196

the typhoon disperses both vertically and horizontally (Gill et al., 1984). Vertically, the energy can 197

penetrate to depths well below the surface mixed layer by the near-inertial internal waves, causing 198

strong turbulence at subsurface (Price et al., 1994; Liu et al., 2008). Consequently, during this period 199

the increased subsurface diapycnal mixing altered the vertical distribution of physical variables 200

(Figures 4A-F). 201

After the passage of the typhoon, the MLD shoaled and reached 28 m on 07/21, which was close to 202

the pre-typhoon state (Figure 4A). However, temperature did not begin to recover immediately, but 203

continued to decrease (Figure 4B). SST decreased by 0.46 °C on 07/23 from the value on 07/17. The 204

surface cooling lasted for 8 d; a similar phenomenon has been observed for some other typhoons 205

(e.g., Domingues et al., 2015). The salinity anomaly above the MLD shows two distinct stages 206

(Figure 4D). During 07/18–07/20 salinity increased by 0.01–0.05 psu, which was consistent with the 207

temperature decrease. However, after 07/20, salinity decreased with the SSS even lower than the pre-208

typhoon value by 0.03 psu due to continuous precipitation (Figure 3). 209

After the typhoon, both isotherm and isopycnal showed continuous upward displacement during 210

07/16–07/23 (Figures 4A,E). In most previous studies, these uplifts have been attributed to the 211

intense typhoon-driven upwelling (i.e., positive Ekman pumping). However, this was not the case in 212

this situation. On the one hand, the typhoon-induced upwelling mainly occurred in the typhoon 213

forced stage, which time scale was usually <1 d (Jaimes and Shay, 2015) and is known not to persist 214

for a period of 8 d as observed in this case. On the other hand, the intense upwelling was confined 215

within a narrow region along the typhoon track, and correspondingly, a compensatory downwelling 216

occurred over a broader area outside the region of upwelling (Jaimes and Shay, 2015; Jullien et al., 217

2012). The observed negative Ekman pumping at the BGC-Argo float location suggests that the float 218

measurements were conducted in such a downwelling region (Figure 5A). In this case, we attribute 219

the long-lasting upward displacement to the quasi-geostrophic response of the upper ocean to 220

typhoon forcing. Both barotropic and baroclinic oscillations can be induced in the ocean by the 221

passage of a typhoon (Geisler, 1970). According to the two-layer ocean model, when typhoon 222

translation speed exceeds the first baroclinic wave speed, a geostrophically balanced baroclinic ridge 223

is generated along the track, which can also propagate as Rossby waves. In the present case, the 224

typhoon translation speed was around 6.5 m s-1 and the first baroclinic wave speed was calculated to 225

be 1.5 m s-1; thus, the formation of a baroclinic ridge was to be expected. A primary feature of the 226

first baroclinic mode is that the sea surface height changes in the opposite direction to the vertical 227

displacement of the pycnocline or thermocline. This effect was clearly shown by our observations. 228

In review

Page 7: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

7

After the passage of the typhoon, SLA decreased continuously (Figure 3A), which was coincident 229

with the upward displacement of the isotherms and isopycnals (Figures 4A,E). The estimated 230

isotherm displacement from the pre-typhoon condition (Δh) was basically consistent with the 231

subsurface observations (Figure 5C). The westward propagation speed of first mode Rossby wave 232

can be practically estimated from the time-series of altimeter observations (Chelton and Schlax, 233

1996). According to their estimation, this speed at the latitude of the float location (14°N) is 234

approximately 0.15 m s-1, and thus the upward displacement forced by the adiabatic quasi-235

geostrophic adjustment could last for 15 d in theory. However, in a real case, the ocean is always 236

forced by various physical dynamics, and this phenomenon would be obscured by new forcing. 237

Nevertheless, our observation of 8 d uplifting revealed the great impact of a super typhoon on the far-238

filed region. In principle, sea surface depression is also influenced by the barotropic process and local 239

eddy development. However, the SCS basin is deeper than 4000 m, and the depth-averaged 240

horizontal velocity calculated from the neighboring float locations was <0.01 m s-1, and thus the 241

barotropic sea surface depression was negligible (Ginis and Sutyrin, 1995). Local eddy influence was 242

also excluded as the float was located far away from eddies before and after the typhoon (Figures 243

5B,D). 244

3.3 Biological Responses of the Upper Ocean as Observed by the BGC-Argo Float 245

In stratified waters, Chla usually exhibits a bell-shaped vertical distribution, and DCM appears near 246

the top of the nutricline where phytoplankton growth is optimal (Cullen, 2015). Before the typhoon 247

Rammasun, a pronounced DCM of 1 mg m-3 occurred at 82 m (Figures 4G, 6A). Previous studies 248

have shown that the nutricline in the SCS basin is deep (~60 m) in summer (Tseng et al., 2005). To 249

clearly show the phytoplankton redistribution, we divided the upper 150 m into 3 layers: upper (0–50 250

m), intermediate (50–100 m) and lower (100–150 m). DCM was located in the intermediate layer 251

before the typhoon, and the depth-integrated Chla in this layer was 28 mg m-2, about 5 and 2 times 252

higher than the values in the upper and lower layers, respectively (Figure 6B). A similar vertical 253

distribution (Figures 4I, 6C) can be seen for bbp, the proxy for particulate organic carbon or 254

phytoplankton biomass (Cetinić et al., 2012; Graff et al., 2015). 255

Coinciding with the physical responses during the passage of the typhoon, the increased mixing also 256

altered the vertical distribution of phytoplankton, and brought subsurface phytoplankton into the 257

surface mixed layer. As a result, the surface Chla increased from 0.06 to 0.22 mg m-3 (Figure 3B). 258

Below the MLD, Chla anomaly displayed a two-layer structure: increasing by 0.06–0.2 mg m-3 above 259

75 m, and decreasing by 0–0.4 mg m-3 below (Figure 4H). Depth-integrated Chla in the upper layer 260

(0–50 m) increased from 5.2 to 13.4 mg m-2, but the depth-integrated values in the intermediate and 261

lower layers decreased by 3.5 and 2.4 mg m-2, respectively (Figure 6B). The depth-integrated Chla in 262

the upper 150 m increased slightly by 2.2 mg m-2, implying that biological alternations (i.e., net 263

phytoplankton growth) was very limited and that the typhoon-induced physical processes were more 264

significant in determining the vertical phytoplankton re-distribution and the consequent increase in 265

surface Chla. The bbp anomaly showed a two-layer structure as well (Figures 4J, 6C). Depth-266

integrated bbp in the upper layer increased by 0.0034, while the depth-integrated values in the 267

intermediate and lower layers decreased by 0.0091 and 0.0016; the total bbp within 150 m showed a 268

slight decrease of 0.007 (Figure 6D). It should be noted that, phytoplankton can adjust their 269

Chla:Carbon ratio in response to the changes of light, temperature, and nutrients (e.g., Laws and 270

Bannister, 1980; Siegel et al., 2013). When the light decreases, phytoplankton can increase their Chla 271

content to promote the photosynthesis, resulting in a higher Chla:Carbon ratio (Laws and Bannister, 272

1980). This effect is the so-called phytoplankton photoacclimation. During the passage of the 273

typhoon, the surface PAR was rather low due to the heavy cloud (Figure 3A), and the 274

In review

Page 8: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

8

This is a provisional file, not the final typeset article

photoacclimation adjustment can give rise to a higher increase extent of the upper layer Chla than 275

that of bbp (Figures 6B,D). However, the depth-integrated Chla in the total column (0–150 m) 276

increased negligibly, suggesting that photoacclimation process plays a minor role in the total Chla 277

increase. The changes in Chla and bbp suggests that there was almost no net phytoplankton biomass 278

increase in the euphotic layer during the typhoon, and that surface increases were primarily due to the 279

vertical redistribution of particles (and thus phytoplankton) driven by typhoon-induced vertical 280

mixing. 281

After the passage of the typhoon, both DCM and DBM showed continuous uplifting (Figures 4G,I), 282

accompanied by the upward displacement of the isotherms and isopycnals (Figures 4A,E). DCM 283

depth rose to 49 m on 07/23, which was 33 m shallower than the pre-typhoon depth and was similar 284

to the isopycnal displacement of 23 kg m-3 (Figure 4G); DBM depth showed a similar vertical 285

displacement (Figure 4I). These shifts brought more subsurface phytoplankton into the surface 286

mixed layer; 5 d after the typhoon, both surface Chla and bbp increased again (Figure 3B). Both the 287

Chla and bbp vertical anomalies were positive above 75 m and negative below that depth (Figures 288

4H,J). Depth-integrated Chla and bbp in the upper layer increased, while those integrated values in 289

the intermediate and lower layers decreased; However, both the depth-integrated Chla and bbp above 290

150 m showed an insignificant decrease of 1.5 mg m-2 and 0.009 (Figures 6B,D). Similarly, 291

phytoplankton photoacclimation effect also contributed to the upper layer Chla increase, but the 292

impact to the total Chla within the whole water column was negligible. These observations suggest 293

that there was no phytoplankton bloom after the typhoon and that the surface increases in 294

phytoplankton were mainly due to the uplift of subsurface particles. 295

Other studies have also recognized two stages of surface phytoplankton increase (Pan et al., 2017). 296

Their first stage was characterized by a remarkable deepening of MLD, to the extent of exceeding 297

DCM, entraining subsurface phytoplankton into the surface mixed layer. In our case, the float was 298

located far away from the typhoon track, and wind speed at this location was not large enough so that 299

the observed MLD increased weakly; however, the mechanical energy transferred into the ocean was 300

transported into the ocean depth by near-inertial internal waves, which causes intense subsurface 301

turbulence (Price et al., 1994). This wave-induced mixing redistributed the particles, matching an 302

increase in surface phytoplankton along with a decrease in subsurface. 303

Studies have attributed the delayed increase in surface phytoplankton to biological alternations, 304

arguing that the phytoplankton blooms lagged the nutrient supplement by 3–6 days (Walker et al., 305

2005; Zheng and Tang., 2007). However, we observed that the upper ocean was highly stratified 306

before the typhoon, and the nutricline was far from MLD. It is therefore unlikely that substantial 307

quantities of nutrients were entrained into the mixed layer from depth. Moreover, in some occasions, 308

plentiful nutrients are horizontally transported to the offshore regions by the reinforced river 309

discharge (Zhao et al., 2013) or jet flow (Davis and Yan, 2004). In summer, a strong coastal 310

upwelling has often been observed off the Vietnam coast, mainly attributed to the strong wind-driven 311

offshore Ekman transport and Ekman pumping (e.g., Lu et al., 2018; Tang et al., 2004a; Xiao et al., 312

2020). This upwelling water characterized with low SST, high Chla, and high nutrients, can be 313

transported eastward by the jet flow, resulting in a surface phytoplankton bloom in the open region 314

(e.g., Tang et al., 2004a; 2004b; Liu et al., 2020; Liu and Tang, 2012). The bloom induced by the jet 315

flow is usually found in the region within 400 km off the Vietnam coast (e.g., Tang et al., 2004a), 316

and cannot extend to the BGC-Argo float location (116°E, 14°N) even in the larger summer monsoon 317

year (e.g., Liu and Tang, 2012; Xiao et al., 2020). The satellite-observed SST showed that the 318

Vietnam coastal upwelling water was indeed transported to the open area, but the surface cooling 319

induced by typhoon and coastal upwelling can still be distinguished (Figure 2B). Therefore, the 320

In review

Page 9: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

9

likelihood of horizontal advection can be excluded. Furthermore, if nutrients were supplied into the 321

euphotic layer by vertical entrainment or horizontal advection, the total phytoplankton biomass in the 322

water column is expected to increase. Therefore, we conclude that the second stage increase in 323

surface phytoplankton was also due to vertical redistribution of the particles. The typhoon-induced 324

adiabatic quasi-geostrophic adjustment can produce large scale upward displacement of the 325

thermocline. Although the BGC-Argo float was located in a region of negative Ekman pumping, 326

subsurface particles can be uplifted for long periods, in which case some subsurface particles would 327

be brought into the surface mixed layer. 328

Therefore, the super typhoon Rammasun did not increase the total Chla and bbp in the distant region 329

either during or after its passage, suggesting that the total phytoplankton biomass and the associated 330

primary production were not promoted. In the far-field regions, the typhoon-induced surface Chla 331

increase resulted from particle vertical redistribution, and was not necessarily connected with the 332

nutrient supply. Only the surface (satellite) observations may overestimate the phytoplankton 333

biological responses. 334

Our observations suggest that: (1) spatially, although the typhoon-induced upwelling (positive 335

Ekman pumping) region was narrow and local, the response of isopycnal uplifting may occur on a 336

much larger scale; (2) temporally, although the typhoon-induced upwelling was usually a transient 337

process, the uplifting of the thermocline in the distant regions can last for a long period (>1 week) 338

due to its quasi-geostrophic nature; (3) the spatial extent and duration of these physical processes 339

support the increase in surface Chla several days after the typhoon without necessitating an increase 340

in net phytoplankton growth; and (4) it is likely that previous studies based only on satellite data 341

overestimated the extent of typhoon-induced phytoplankton increase and the associated carbon 342

fixation, especially in the far-field regions, because they neglected to consider important physical 343

processes that redistribute particles between the surface and subsurface layers. 344

4 Conclusions 345

Physical and biological responses to the super typhoon Rammasun (2014) were investigated using a 346

BGC-Argo float located some 200 km to the left of its track. The float observations showed that there 347

were two stages in the increase of surface phytoplankton, but the negligible change in the depth-348

integrated values of phytoplankton in the upper 150 m. The initial stage surface increase was 349

attributed to the typhoon-induced diapycnal mixing, which entrained subsurface particles directly 350

into the upper layer; the second stage increase was due to adiabatic quasi-geostrophic adjustment of 351

the upper ocean to the typhoon forcing, which uplifted subsurface particles to the surface. In the far-352

field region of the typhoon influence, the surface bloom was due primarily to particle redistribution 353

rather than biological alternations. This study provides some new insight into the physical and 354

biological responses of the upper ocean to a typhoon. The impacts of typhoons on ocean are 355

extremely complex, and our observations cannot represent the whole picture of the biological 356

responses to a typhoon. In the future, more vertical profiling observations across a broad region are 357

necessary to comprehensively quantify oceanic biological responses to a typhoon. 358

Conflict of Interest 359

The authors declare that the research was conducted in the absence of any commercial or financial 360

relationships that could be construed as a potential conflict of interest. 361

Author Contributions 362

In review

Page 10: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

10

This is a provisional file, not the final typeset article

HW performed the observations and initiated this work. FC designed the concept. ZL contributed to 363

the physical analysis. XY provided suggestions on concept design, data analysis, and manuscript 364

writing. GQ and XX analyzed the data. GQ prepared the figures and wrote the manuscript. All 365

authors revised the manuscript. 366

Funding 367

This work was supported by the National Key Research & Development Program of China 368

(2019YFA0606702), the National Natural Science Foundation of China (41876032, 41721005, 369

41576100, 91858202, 41630963, 41776003), the Scientific Research Fund of the Second Institute of 370

Oceanography, MNR (14283), and the National Basic Research Program of China (2015CB954002). 371

Acknowledgments 372

The authors thank Dr. Huijie Xue, Dr. Zhumin Lu, Dr. Yunwei Yan and Dr. Yi Yu for their insightful 373

suggestions. 374

Contribution to the Field Statement 375

Surface phytoplankton biomass usually increases after a typhoon, which has commonly been 376

attributed to particle entrainment and/or nutrient entrainment through the enhanced mixing and 377

upwelling. Most previous studies have focused on the remarkable increase under or to the right of the 378

track, and less attention has been given to phytoplankton responses in the distant region, especially to 379

the left of the track. The mechanisms responsible for such often-observed far-field surface blooms 380

remain unknown for lack of vertical measurements. The super typhoon Rammasun (2014) traveled 381

across the South China Sea, and its induced vertical changes were recorded by a Biogeochemical-382

Argo float located 200 km to the left of its track. Surface phytoplankton biomass increased during 383

and shortly after the passage of the typhoon, but depth-integrated values in the total water column 384

were almost unchanged over the entire period. We suggest that the surface phytoplankton increases 385

were due to vertical redistribution of the phytoplankton driven by typhoon-induced vertical mixing 386

and adiabatic quasi-geostrophic adjustment. The surface blooms in the more distant regions 387

influenced by the typhoon were therefore primarily driven by particle redistribution rather than 388

biological alternations. These findings differ from the results of previous studies, and so provide new 389

insights into physical and biological responses to typhoons. 390

391

References 392

Babin, S.M., Carton, J.A., Dickey, T.D., and Wiggert, J.D. (2004). Satellite evidence of hurricane-393

induced phytoplankton blooms in an oceanic desert. J. Geophys. Res. Oceans 109(C3), C03043. 394

doi: 10.1029/2003jc001938. 395

Black, W.J., and Dickey, T.D. (2008). Observations and analyses of upper ocean responses to tropical 396

storms and hurricanes in the vicinity of Bermuda. J. Geophys. Res. Oceans 113(C8), C08009. 397

doi: 10.1029/2007jc004358. 398

Cetinić, I., Perry, M.J., Briggs, N.T., Kallin, E., D'Asaro, E.A., and Lee, C.M. (2012). Particulate 399

organic carbon and inherent optical properties during 2008 North Atlantic Bloom Experiment. 400

J. Geophys. Res. Oceans 117(C6), C06028. doi: 10.1029/2011jc007771. 401

Chacko, N. (2017). Chlorophyll bloom in response to tropical cyclone Hudhud in the Bay of Bengal: 402

Bio-Argo subsurface observations. Deep Sea Res. I 124, 66-72. doi: 10.1016/j.dsr.2017.04.010. 403

In review

Page 11: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

11

404

Chai, F., Wang, Y., Xing, X., Yan, Y., Xue, H., Wells, M., et al. (2021). A Limited Effect of Sub-Tropical 405

Typhoons on Phytoplankton Dynamics. Biogeosciences 18(3), 849-859. doi: 10.5194/bg-18-406

849-2021. 407

Chaigneau, A., Eldin, G., and Dewitte, B. (2009). Eddy activity in the four major upwrelling systems 408

from satellite altimetry (1992-2007). Prog. Oceanogr. 83(1-4), 117-123. doi: 409

10.1016/j.pocean.2009.07.012. 410

Chelton, D.B., and Schlax, M.G. (1996). Global observations of oceanic Rossby waves. Science 411

272(5259), 234-238. doi: 10.1126/science.272.5259.234. 412

Cullen, J.J. (2015). Subsurface Chlorophyll Maximum Layers: Enduring Enigma or Mystery Solved? 413

Annu. Rev. Mar. Sci. 7(1), 207-239. doi: 10.1146/annurev-marine-010213-135111. 414

Davis, A., and Yan, X.H. (2004). Hurricane forcing on chlorophyll-a concentration off the northeast 415

coast of the US. Geophys. Res. Lett. 31(17), L17304. doi: 10.1029/2004gl020668. 416

Dickey, T., Frye, D., McNeil, J., Manov, D., Nelson, N., Sigurdson, D., et al. (1998). Upper-ocean 417

temperature response to Hurricane Felix as measured by the Bermuda testbed mooring. Mon. 418

Wea. Rev. 126(5), 1195-1201. doi: 10.1175/1520-0493(1998)126<1195:uotrth>2.0.co;2. 419

Domingues, R., Goni, G., Bringas, F., Lee, S.-K., Kim, H.-S., Halliwell, G., et al. (2015). Upper ocean 420

response to Hurricane Gonzalo (2014): Salinity effects revealed by targeted and sustained 421

underwater glider observations. Geophys. Res. Lett. 42(17), 2015GL065378. doi: 422

10.1002/2015GL065378. 423

Geisler, J.E. (1970). Linear theory of the response of a two layer ocean to a moving hurricane. Geophys. 424

Fluid Dyn. 1(1-2), 249-272. doi: 10.1080/03091927009365774. 425

Gill, A.E. (1984). On the Behavior of Internal Waves in the Wakes of Storms. J. Phys. Oceanogr. 14(7), 426

1129-1151. doi: 10.1175/1520-0485(1984)014<1129:OTBOIW>2.0.CO;2. 427

Ginis, I., and Sutyrin, G. (1995). Hurricane-Generated Depth-Averaged Currents and Sea Surface 428

Elevation. J. Phys. Oceanogr. 25(6), 1218-1242. doi: 10.1175/1520-429

0485(1995)025<1218:hgdaca>2.0.co;2. 430

Graff, J.R., Westberry, T.K., Milligan, A.J., Brown, M.B., Dall'Olmo, G., van Dongen-Vogels, V., et al. 431

(2015). Analytical phytoplankton carbon measurements spanning diverse ecosystems. Deep 432

Sea Res. I 102, 16-25. doi: 10.1016/j.dsr.2015.04.006. 433

Jaimes, B., and Shay, L.K. (2009). Mixed Layer Cooling in Mesoscale Oceanic Eddies during 434

Hurricanes Katrina and Rita. Mon. Wea. Rev. 137(12), 4188-4207. doi: 435

10.1175/2009mwr2849.1. 436

Jaimes, B., and Shay, L.K. (2015). Enhanced Wind-Driven Downwelling Flow in Warm Oceanic Eddy 437

Features during the Intensification of Tropical Cyclone Isaac (2012): Observations and Theory. 438

J. Phys. Oceanogr. 45(6), 1667-1689. doi: 10.1175/jpo-d-14-0176.1. 439

Johnson, K.S., Berelson, W.M., Boss, E., Chase, Z., Claustre, H., Emerson, S.R., et al. (2009). 440

Observing biogeochemical cycles at global scales with profiling floats and gliders: prospects 441

for a global array. Oceanography 22(3), 216-225. 442

Jullien, S., Menkes, C.E., Marchesiello, P., Jourdain, N.C., Lengaigne, M., Koch-Larrouy, A., et al. 443

(2012). Impact of Tropical Cyclones on the Heat Budget of the South Pacific Ocean. J. Phys. 444

Oceanogr. 42(11), 1882-1906. doi: 10.1175/jpo-d-11-0133.1. 445

Laws, E.A., and Bannister, T.T. (1980). Nutrient- and light-limited growth of Thalassiosira fluviatilis 446

in continuous culture with implications for phytoplankton growth in the ocean. Limnol. 447

Oceanogr. 25(3), 457-473. 448

Levitus, S. (1982). “Climatological atlas of the world ocean,” in Prof. Paper 13, National Oceanic and 449

Atmospheric Administration, Rockville, MD 450

Lin, I.I. (2012). Typhoon-induced phytoplankton blooms and primary productivity increase in the 451

western North Pacific subtropical ocean. J. Geophys. Res. Oceans 117, C03039. doi: 452

In review

Page 12: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

12

This is a provisional file, not the final typeset article

10.1029/2011jc007626. 453

Lin, I.I., Liu, W.T., Wu, C.C., Wong, G.T.F., Hu, C.M., Chen, Z.Q., et al. (2003). New evidence for 454

enhanced ocean primary production triggered by tropical cyclone. Geophys. Res. Lett. 30(13), 455

1718. doi: 10.1029/2003gl017141. 456

Liu, C., and Tang, D. (2012). Spatial and temporal variations in algal blooms in the coastal waters of 457

the western South China Sea. J. Hydro.-Environ. Res. 6(3), 239-247. doi: 458

https://doi.org/10.1016/j.jher.2012.02.002. 459

Liu, L.L., Wang, W., and Huang, R.X. (2008). The mechanical energy input to the ocean induced by 460

tropical cyclones. J. Phys. Oceanogr. 38(6), 1253-1266. doi: 10.1175/2007jpo3786.1. 461

Liu, Q., Jia, Y., Liu, P., Wang, Q., and Chu, P.C. (2001). Seasonal andJ. Phys. Oceanogr.intraseasonal 462

thermocline variability in the central south China Sea. Geophys. Res. Lett. 28(23), 4467-4470. 463

doi: 10.1029/2001GL013185. 464

Liu, S., Li, J., Sun, L., Wang, G., Tang, D., Huang, P., et al. (2020). Basin-wide responses of the South 465

China Sea environment to Super Typhoon Mangkhut (2018). Sci. Total Environ. 731, 139093. 466

doi: https://doi.org/10.1016/j.scitotenv.2020.139093. 467

Lu, W.F., Oey, L.Y., Liao, E.H., Zhuang, W., Yan, X.H., and Jiang, Y.W. (2018). Physical modulation 468

to the biological productivity in the summer Vietnam upwelling system. Ocean Sci. 14(5), 469

1303-1320. doi: 10.5194/os-14-1303-2018. 470

McDougall, T. J., and Barker, P. M. (2011). Getting started with TEOS-10 and the Gibbs Seawater 471

(GSW) oceanographic toolbox. London: SCOR/IAPSO WG 127, 1–28. 472

Ning, X., Chai, F., Xue, H., Cai, Y., Liu, C., and Shi, J. (2004). Physical-biological oceanographic 473

coupling influencing phytoplankton and primary production in the South China Sea. J. Geophys. 474

Res. Oceans 109, C10005. doi: 10.1029/2004jc002365. 475

Pan, S.S., Shi, J., Gao, H.W., Guo, X.Y., Yao, X.H., and Gong, X. (2017). Contributions of physical 476

and biogeochemical processes to phytoplankton biomass enhancement in the surface and 477

subsurface layers during the passage of Typhoon Damrey. J. Geophys. Res. Biogeosci. 122(1), 478

212-229. doi: 10.1002/2016jg003331. 479

Price, J.F. (1981). Upper Ocean Response to a Hurricane. J. Phys. Oceanogr. 11(2), 153-175. doi: 480

10.1175/1520-0485(1981)011<0153:uortah>2.0.co;2. 481

Price, J.F., Sanford, T.B., and Forristall, G.Z. (1994). Forced stage response to a moving hurricane. J. 482

Phys. Oceanogr. 24(2), 233-260. doi: 10.1175/1520-0485(1994)024<0233:fsrtam>2.0.co;2. 483

Qu, T., Du, Y., Gan, J., and Wang, D. (2007). Mean seasonal cycle of isothermal depth in the South 484

China Sea. J. Geophys. Res. Oceans 112(C2), C02020. doi: 10.1029/2006JC003583. 485

Sanford, T.B., Price, J.F., and Girton, J.B. (2011). Upper-Ocean Response to Hurricane Frances (2004) 486

Observed by Profiling EM-APEX Floats. J. Phys. Oceanogr. 41(6), 1041-1056. doi: 487

10.1175/2010jpo4313.1. 488

Shang, S.L., Li, L., Sun, F.Q., Wu, J.Y., Hu, C.M., Chen, D.W., et al. (2008). Changes of temperature 489

and bio-optical properties in the South China Sea in response to Typhoon Lingling, 2001. 490

Geophys. Res. Lett. 35(10), L10602. doi: 10.1029/2008gl033502. 491

Shay, L.K., Goni, G.J., and Black, P.G. (2000). Effects of a warm oceanic feature on Hurricane Opal. 492

Mon. Wea. Rev. 128(5), 1366-1383. doi: 10.1175/1520-0493(2000)128<1366:eoawof>2.0.co;2. 493

Siegel, D.A., Behrenfeld, M.J., Maritorena, S., McClain, C.R., Antoine, D., Bailey, S.W., et al. (2013). 494

Regional to global assessments of phytoplankton dynamics from the SeaWiFS mission. Remote 495

Sens. Environ. 135(0), 77-91. doi: 10.1016/j.rse.2013.03.025. 496

Stramma, L., Cornillon, P., and Price, J.F. (1986). Satellite observations of sea surface cooling by 497

hurricanes. J. Geophys. Res. Oceans 91(C4), 5031-5035. doi: 10.1029/JC091iC04p05031. 498

Tang, D.L., Kawamura, H., Van Dien, T., and Lee, M. (2004a). Offshore phytoplankton biomass 499

increase and its oceanographic causes in the South China Sea. Mar. Ecol. Prog. Ser. 268, 31-500

41. doi: 10.3354/meps268031. 501

In review

Page 13: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

13

Tang, D.L., Kawamura, H., Doan-Nhu, H., and Takahashi, W. (2004b). Remote sensing oceanography 502

of a harmful algal bloom off the coast of southeastern Vietnam. J. Geophys. Res. Oceans 503

109(C3). doi: https://doi.org/10.1029/2003JC002045. 504

Tseng, C.-M., Wong, G.T.F., Lin, I.I., Wu, C.R., and Liu, K.K. (2005). A unique seasonal pattern in 505

phytoplankton biomass in low-latitude waters in the South China Sea. Geophys. Res. Lett. 32(8), 506

L08608. doi: 10.1029/2004gl022111. 507

Walker, N.D., Leben, R.R., and Balasubramanian, S. (2005). Hurricane-forced upwelling and 508

chlorophyll a enhancement within cold-core cyclones in the Gulf of Mexico. Geophys. Res. 509

Lett. 32(18), L18610. doi: 10.1029/2005gl023716. 510

Xiao, F., Wu, Z.F., Lyu, Y.S., and Zhang, Y.C. (2020). Abnormal Strong Upwelling off the Coast of 511

Southeast Vietnam in the Late Summer of 2016: A Comparison with the Case in 1998. 512

Atmosphere 11(9). doi: 10.3390/atmos11090940. 513

Zhao, H., Han, G., Zhang, S., and Wang, D. (2013). Two phytoplankton blooms near Luzon Strait 514

generated by lingering Typhoon Parma. J. Geophys. Res. Biogeosci. 118(2), 412-421. doi: 515

10.1002/jgrg.20041. 516

Zheng, G.M., and Tang, D.L. (2007). Offshore and nearshore chlorophyll increases induced by typhoon 517

winds and subsequent terrestrial rainwater runoff. Mar. Ecol. Prog. Ser. 333, 61-74. doi: 518

10.3354/meps333061. 519

Captions 520

FIGURE 1. Wind speed at 0 h UTC on 07/17 in the SCS and the track of Rammasun (black line). The 521

black dots represent the 6-hourly typhoon centers, and the red pentagram shows the BGC-Argo float. 522

FIGURE 2. Satellite-observed SST on (A) 07/14 and (B) 07/18, and surface Chla on (C) 07/14 and 523

(D) composited for 07/20–07/27. 524

FIGURE 3. Time series of (A) satellite-observed 6-hourly wind speed, 3-hourly precipitation, daily 525

SLA, and daily PAR at the float location, and of (B) float-observed surface temperature, salinity, Chla 526

and bbp. The period when Rammasun crossed SCS is denoted by the two vertical dashed lines. 527

FIGURE 4. Time series of vertical distributions of float-observed (A) temperature, (C) salinity, (E) 528

σ0, (G) Chla and (I) bbp, alongside their corresponding anomalies (right panel). The solid black lines in 529

(A-J), white line in (G), blue line in (G), and green in (I) represent the MLD, depth of isopycnic of σ0 530

= 23 kg m-3, depth of DCM, and depth of DBM, respectively. 531

FIGURE 5. (A) Time series of Ekman pumping at the float location. (B, D) Daily SLA on 07/14 and 532

07/18, with solid (dashed) contours for the edges of anticyclonic (cyclonic) eddies. (C) Estimated 533

vertical displacement of the thermocline (Δh), the float-observed depth of the 23 (24) °C isotherm, and 534

depth of isopycnic of σ0 = 23 (23.5) kg m-3. 535

FIGURE 6. (A, C) Vertical distributions of the time-averaged Chla and bbp before (07/14–07/15), 536

during (07/16–07/17), and after (07/19–07/23) the typhoon’s passage. (B, D) Time series of the depth-537

integrated Chla and bbp in the upper (0–50 m), intermediate (50–100 m) and lower layers (100–150 m), 538

and the entire column (0–150 m), respectively. 539

540

541

In review

Page 14: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

14

This is a provisional file, not the final typeset article

Figure 1 542

543

544

545

546

Figure 2 547

548

549

550

In review

Page 15: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

15

Figure 3 551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

In review

Page 16: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

16

This is a provisional file, not the final typeset article

Figure 4 568

569

570

571

572

573

574

In review

Page 17: Far-field impacts of a super typhoon on upper ocean

Phytoplankton Response to Typhoon

17

Figure 5 575

576

577

578

579

580

Figure 6 581

582

583

In review

Page 18: Far-field impacts of a super typhoon on upper ocean

Figure 1.TIFF

In review

Page 19: Far-field impacts of a super typhoon on upper ocean

Figure 2.TIFF

In review

Page 20: Far-field impacts of a super typhoon on upper ocean

Figure 3.TIFF

In review

Page 21: Far-field impacts of a super typhoon on upper ocean

Figure 4.TIFF

In review

Page 22: Far-field impacts of a super typhoon on upper ocean

Figure 5.TIFF

In review

Page 23: Far-field impacts of a super typhoon on upper ocean

Figure 6.TIFF

In review