Skip to main content
PLOS ONE logoLink to PLOS ONE
. 2021 Apr 26;16(4):e0250604. doi: 10.1371/journal.pone.0250604

Micro-size plankton abundance and assemblages in the western North Pacific Subtropical Gyre under microscopic observation

Taketoshi Kodama 1,*, Tsuyoshi Watanabe 1, Yukiko Taniuchi 1, Akira Kuwata 1, Daisuke Hasegawa 1
Editor: Arga Chandrashekar Anil2
PMCID: PMC8075241  PMID: 33901250

Abstract

While primary productivity in the oligotrophic North Pacific Subtropical Gyre (NPSG) is changing, the micro-size plankton community has not been evaluated in the last 4 decades, prompting a re-evaluation. We collected samples over three years (2016–2018) from depths of 10 to 200 m (n = 127), and the micro-size plankton were identified and counted to understand the heterogeneity of micro-size plankton community structure. The assemblages were consistent to the those of 4 decades ago. Dinophyceae (dinoflagellates) were the most numerically abundant, followed by Cryptophyceae and Bacillariophyceae (diatoms). The other micro-size plankton classes (Cyanophyceae, Haptophyceae, Dictyochophyceae, Euglenophyceae, and Prasinophyceae) were not always detected, whereas only Trichodesmium spp. was counted in the Cyanophyceae. Other unidentified autotrophic and heterotrophic flagellates were also significantly present, and their numeric abundance was higher than or at the same level as was that of the Dinophyceae. In the Dinophyceae, Gymnodiniaceae and Peridiniales were abundant. The chlorophyll a concentration and these class-level assemblages suggested micro-size plankton is not a major primary producer in this area. We applied generalized additive models (GAMs) and principal coordination analyses (PCoAs) to evaluate the habitats of every plankton group and the heterogeneity of the assemblages. The GAMs suggested that every classified plankton abundance showed a similar response to salinity, and we observed differences in habitats in terms of temperature and nitrate concentrations. Based on the PCoAs, we observed unique communities at the 200 m depth layer compared with those at the other sampling layers. The site scores of PCoAs indicated that the micro-size plankton assemblages are most heterogeneous at the 10 m depth layer. At such depth, diazotrophic Cyanophyceae (Trichodesmium spp.) are abundant, particularly in less-saline water. Therefore, nitrogen fixation may contribute to the heterogeneity in the abundance and assemblages in the western NPSG.

Introduction

The size of primary producers is one of the keys to controlling the productivity of marine ecosystems [1]. In waters where pico-size phytoplankton are dominant, more trophic levels are required to convert primary production to useful forms; therefore, biomass transfer efficiency is low [1]. In other words, energy fixed by primary production is more efficiently transferred to higher trophic organisms in micro-plankton dominated waters [1]. In addition, over 100% of daily primary production is sometimes consumed by micro-size heterotrophic and mixotrophic plankton [2, 3]. Consequently, the abundance and composition of micro-size plankton comprise necessary information for evaluating the biological productivity of the ocean.

The subtropical open ocean occupies approximately 40% of the surface of the Earth. The stratification of the surface water column develops throughout the year, and nutrient supplies from the deeper layer are extremely limited; therefore, phytoplankton abundance is low and limited in this layer. It has been considered that the ecosystem of the subtropical open ocean is stable and is a ’climax-type’ community [4]. However, over the recent three decades, many studies have indicated the ecosystem of the subtropical open ocean as unstable and heterogeneous in space and time [5]. In the subtropical open ocean, pico- and nano-size phytoplankton is the dominant primary producers; however, micro-size phytoplankton is also present in significant numbers [57]. In particular, diatoms (Bacillariophyceae) are sometimes abundant (> 108 cells m-2) in the surface mixed layer [5, 8, 9], with most being nitrogen-fixing diatom symbioses [10].

The North Pacific Subtropical Gyre (NPSG), the widest gyre in the ocean, is sometimes divided into two provinces, namely west and east [11]. In the western NPSG, both nitrate and phosphate are depleted at nanomolar levels [12], and nitrogen fixation is significant, but not quite active [13]. The western NPSG is oligotrophic, but is known as the spawning and nursery grounds of commercially important migratory fish, such as Albacore (Thunnus alalunga), Skipjack tuna (Katsuwonus pelamis), Blue marlin (Makaira mazara), and Japanese eel (Anguilla japonica) [14, 15]. In the western NPSG, the micro-size plankton community was studied widely in the 1970s [6, 16]. The findings indicated that heterotrophic protists such as flagellates and dinoflagellates are dominant, with autotrophic coccolithophore and diatoms also significantly present. These communities are similar to those of the eastern NPSG [7]. However, with the recent effects of global warming, primary productivity has been changing, even in the nutrient-depleted oligotrophic ocean [17, 18]. Therefore, the micro-size plankton community in the western NPSG could have changed recently compared with the findings of the studies in the 1970s. In addition, descriptions of the classical morphologically identified plankton community are necessary for further studies using state-of-the-art technology, such as the metabarcoding technique.

In view of the above, we investigated the morphologically identified micro-size plankton community structure in the western NPSG. We evaluated the habitats of plankton groups by using empirical statistical models (generalized additive models, GAMs) and multivariate analyses (principal coordination analysis, PCoA). Our aim was to add to the basic knowledge on the micro-size plankton community structure in the oligotrophic western NPSG to understand the trophic structure and heterogeneity of this community.

Materials and methods

Sampling

We collected samples during three cruises on the R/V Kaiyo-maru (Japan Fisheries Agency) conducted in September and October 2016 (named the KY1604 cruise), September and October 2017 (KY1704 cruise), and June 2018 (KY1801 cruise). The cruises were conducted in the western NPSG between 12°N and 25°20´N and between 126°E and 143°E (Table 1, Fig 1). The primary aim was investigating the distribution of Japanese eel larvae; therefore, the observations for collections of seawater were limited to night-time collections.

Table 1. List of the sampling layers and numbers in three cruises.

Cruise Date Sampling layers Numbers of samples
KY1604 28th Sep. –31st Oct., 2016 10 m 9
50 m 9
100 m 9
200 m 9
SCM (100–130 m) 9
KY1704 29th Sep.–25th Oct., 2017 10 m 5
50 m 5
100 m 32
200 m 5
SCM (85–105 m) 5
KY1801 2nd Jun. –5th Jun., 2018 10 m 6
200 m 6
SCM-High (85–120 m) 6
SCM (125–155 m) 6
SCM-Low (150–175 m) 6

The SCM, SCM-High and SCM-Low denote subsurface chlorophyll maximum (the peak of in vivo chlorophyll fluorescence), the shallower and deeper SCM edges (approximately half of fluorescence values compared with the SCM), respectively.

Fig 1. Sampling locations of this study.

Fig 1

(a) Small-scale map of the western North Pacific, and (b) enlarged map of the sampling sites during the three cruises. The gray arrows in (a) denote schematic major ocean currents in the observation areas, but the Kuroshio current path is that of October 2017 based on the Quick Bulletin of Ocean Conditions issued by the Japan Coast Guard. (https://www1.kaiho.mlit.go.jp/KANKYO/KAIYO/qboc/index_E.html).

All observation and samplings during the cruises were permitted by Japan Fisheries Agency (SUISUI NO. 27–1253, NO. 29–28, and NO. 30–238). Our samplings were conducted in high seas or Japanese EEZ area during KY1604 and KY1704 cruises, and those during KY1801 cruise were permitted by the coastal States (the Federated States of Micronesia and United States of America) as the marine scientific research (MSR) in accordance with the United Nations Convention on the Law of the Sea (UNCLOS). The permission during KY1801 cruise was granted by Japan Fisheries Agency (No. U2018-001).

The samples for identifying micro-sized plankton abundance and assemblages were collected using a conductivity-temperature-depth profiler with carousel multiple sampling (CTD-CMS), which consists of a CTD sensor (SBE911plus, Sea-Bird Scientific, USA) and a Sea-Bird carousel water sampler (SBE32 with 10-litter Niskin-X sampling bottles). Chlorophyll fluorescence was monitored using a fluorometer (Seapoint Sensors, USA) attached to the CTD sensor. Water samples were collected at 10, 50, 100, and 200 m depths, and the subsurface chlorophyll maximum (SCM) layer at nine stations during the KY1604 cruise (Table 1). During the KY1704 cruise, the same vertical profiles were conducted at five stations in the northwestern part of the study area. At the other 28 stations, the samples were collected only at 100 m depth during the KY1704 cruise (Table 1). We did not collect a SCM sample when it was 100 m at one station during the KY1704 cruise. At six stations during the KY1801 cruise, the water samples were collected at depths of 10 and 200 m, and two edges and peak of the SCM to understand the variations in the SCM layers (Table 1). The shallower and deeper SCM edges (approximately half of fluorescence values compared with the SCM) were named SCM-High and SCM-Low, respectively. In total, 127 samples were collected during three cruises for analyses of the phytoplankton community structure.

For plankton assemblages and abundance analyses, 40 ml of acidified Lugol’s solution was added immediately to 1 L seawater (4% final concentration). The sample treatments were similar to those of Watanabe et al. [19]. The samples were kept in the dark at room temperature (~20°C) until the onshore treatments. Samples were concentrated by reverse filtration through 2 μm membrane filters [20]. Concentrated algal cells were settled in a chamber and counted using an inverted microscope [21]. Because we used acidified Lugol’s solution as the fixing reagent, some plankton groups with calcium carbonate plates, such as coccolithophores, were not detected in our samples. We counted plankton classified into eight classes, namely Bacillariophyceae (diatoms), Cyanophyceae, Cryptophyceae, Dinophyceae (dinoflagellates), Dictyochophyceae, Euglenophyceae, Haptophyceae, and Prasinophyceae. Bacillariophyceae, Cyanophyceae, Dinophyceae, Dictyochophyceae, Haptophyceae, and Prasinophyceae were classified mostly into genus level when possible. The identification of Cryptophyceae and Euglenophyceae remained at class level. The other autotrophic and heterotrophic flagellates, except eight classes, were not identified, but were counted (hereafter, described as flagellates). The phytoplankton species were identified following Tomas et al. [22]. Trichodesimum spp. is the only counted group in Cyanophyceae, with the other micro-size Cyanobacteria, such as Richelia intracellularis, not counted from these samples. In the case of Trichodesimum spp., the number of filaments were counted, but not that of cells. During the KY1704 and KY1801 cruises, diatom–diazotroph (R. intracellularis) associations (DDAs) were counted from the specimens collected on 10 μm filtered 2.3L of seawaters through a 10 μm filter following Chen et al. [23].

We also collected samples for measurements of nutrients and chlorophyll a concentration. Approximately 10 mL of water was collected for nutrients, and kept frozen at -20°C until onshore analysis. The samples were thawed >24 h before analysis for recovering from polysilicate to monosilicate, after which the nutrient (nitrate, nitrite, phosphate, and silicate) concentrations were investigated using an autoanalyzer (QuAAtro, BLTEC, Japan). The measurement protocols depended mainly on information from Becker et al. [24]. For each chlorophyll (Chl) a sample, 300 mL of water was collected. The particles in the water were collected on four types of filter, namely a glass fiber filter (GF/F, Whatman, 0.7 μm mesh, Merck, Germany), 0.2 μm membrane filter (Nuclepore, Whatman, Merck, Germany), 3.0 μm membrane filter (Nuclepore, Whatman, Merck, Germany), and 10 μm membrane filter (Nuclepore, Whatman, Merck, Germany). The filters were immersed in N,N-dimethylformamide [25] and stored below -20°C in the dark until onshore measurements. Chl a concentrations were measured with a Turner Designs 10-AU fluorometer (USA) [26]. In this study, we used the results of the Whatman GF/F filter (as total Chl a concentration) and the Whatman Nuclepore 10 μm filter (10 μm Chl a concentration, hereafter). During the KY1604 cruise, the 10 μm Chl a concentration was lacking, except at the SCM layer.

As several previous studies have indicated that phytoplankton abundance and assemblages differ with the eddy structure (cyclonic or anticyclonic) in the open ocean [2729], we used relative vorticity (Ro) as the index of the eddy structure. Ro is calculated from the following equations

Ro=ζ/f (1)

where ζ is absolute vorticity, and f is the Coriolis parameter. The parameters for each sampling point were calculated from absolute geostrophic velocities, and geostrophic velocity anomalies from the sea level daily gridded data product (Product identifier: SEALEVEL_GLO_PHY_L4_REP_OBSERVATIONS_008_047) of the European Union Copernicus Marine Environment Monitoring Service. The positive and negative Ro indicated the position in the cyclonic and anticyclonic eddy, respectively.

Statistical analyses

All the statistical analyses were performed with R software [30]. We applied two statistical analyses for evaluating the relationship between the micro-size plankton community and environmental parameters. One is the generalized additive model (GAM) analysis, which is one of the empirical models for understanding the ecological habitats of target biota [31], and the other is PCoA, which is one of the classical multidimensional scaling techniques [32, 33]. We chose PCoA because our data sets did not meet the prerequisites for principal component analysis (PCA).

GAM analysis was applied to the cell density of every plankton group classified at class level, except flagellates, which were not identified into class level. We applied GAM using the gam function in the mgcv package [34] to examine the influence of oceanographic variables (i.e., temperature, salinity, nitrate, and silicate concentration) and sampling variables (i.e., latitude). We used a gamma distribution and a natural log link function to model plankton density. The full model is:

Pis(T)+s(S)+s(Lat)+s(logNO3)+s(Ro)+β (2)

where Pi represents the cell density of target plankton groups (i). T, S, Lat, logNO3, Ro, and β represent temperature, salinity, latitude, common log-transformed nitrate concentration, relative vorticity, and intercept, respectively. s is a spline function. The upper limits of the degrees of freedom for all smoothing terms (k) were set to 4 to avoid biologically impossible responses [35]. Multicollinearity in the environmental factors was checked using the variance inflation factor (VIF), and all VIF were <5 [36]. As regards the results, we did not include depth, which transformed as the categorical values (10 m, 50 m, 100 m, SCM, and 200 m), and the silicate and phosphate concentration in the model. We chose the optimal GAMs based on a corrected Akaike information criterion (AICc) value.

When the target taxon was absent in >30% of samples, we applied the delta-GAM technique. First, the target cell density data were transformed as present (1) or absent (0), and we applied GAMs with a binomial distribution and a logit link function (hereafter, present–absent GAM). Subsequently, after removing the absent sample data, GAM was re-performed with the gamma distribution and the natural log link function to model the target plankton cell density (hereafter, abundance GAM). The explanatory variables in the full model descriptions of the delta-GAM were the same with Eq 2. When the target taxon was absent in >80% of samples, we did not apply any GAMs to them.

The PCoA was used to evaluate the spatial heterogeneity of plankton assemblages. We applied PCoA to three types of communities, namely 1) classified into class level including the flagellates, 2) genus level data of the Dinophyceae, and 3) genus level data of the Bacillariophyceae. Our data contained significant null data; therefore, the distance among the samples were calculated with Bray–Curtis dissimilarity [33]. PCoA was conducted using the cmdscale function in the VEGAN package [37]. PCoA outputs scores for samples. The scores for target communities were calculated by the weighted average values. The relationships between environmental parameters and PCoA output scores for stations were calculated with the envfit function in the VEGAN package [37]. The environmental parameters were limited to those used in the GAMs.

Results

Environmental conditions

The SCM layer was formed at 115 ± 9 m (mean ± SD) in the KY1604, 94 ± 10 m in the KY1704, and 138 ± 10 m in the KY1801 cruises. SCM-High and SCM-Low in the KY1801 cruise were set at 102 ± 13 and 166 ± 8 m, respectively. Temperature decreased with depth; however, the temperature at 10 m was indicated the same as that at 50 m depth in the KY1704 cruise (Fig 2). Less-saline water was observed at 10 m depth during every cruise and, in particular, it was <34.5 at 10 m depth in the KY1801 cruise (Fig 2). The nitrate concentration was largely depleted from 10 m to the SCM, and remained at several μM at 200 m depth (Fig 2). Nitrate concentration was over 1 μM of in water where the temperature was below 20–25°C. The vertical distribution of silicate and phosphate concentrations was quite similar to the nitrate concentration, while the silicate concentration was not depleted <1 μM, even at a depth of 10 m. The total Chl a concentration was highest at the SCM or SCM-High layers, suggesting the fluorometer values were overestimating the chlorophyll a concentration in the deeper layers as indicated by Falkowski and Kiefer [38]. The total Chl a concentration was lowest at 200 m depth (Fig 2). The median values of 10 μm Chl a concentration were highest at SCM (Fig 2), but the contribution to the total Chl a concentration was limited to <5.0% in the median. The contribution to the total Chl a concentration increased at a depth of 200 m (Fig 2). In addition, the contribution of 10 μm Chl a was high at 10 m depth in the KY1801 cruise (Fig 2).

Fig 2.

Fig 2

Vertical distribution of (a) temperature, (b) salinity, (c) nitrate concentration, (d) total Chl a concentration, (e) 10 μm Chl a concentration, and (f) proportion of 10 μm Chl a concentration to total Chl a concentration for KY1604 (top), KY1704 (middle), and KY1801 (bottom). The thick vertical line in each box represents the median values. Boxes indicate the lower and upper quartiles. Horizontal lines extending from each box represent the minimum and maximum values. The open circles are outliers. On the nitrate concentration and proportion of 10 μm Chl a concentration, the x-axes are transformed as common logarithms.

Micro-size plankton abundance and assemblages

The layer of micro-sized plankton cell density varied widely from 1.1 ×103 to 7.3 ×104 cells L-1 (mean ± SD: 1.4 ± 1.3 ×104 cells L-1), and the vertical profiles differed among the cruises (Fig 3). In KY1604, the cell density was highest at the SCM (9.8 ± 3.4 ×103 cells L-1) and lowest at 200 m depth (3.3 ± 1.9 ×103 cells L-1). In the KY1704 cruise, the mean cell density was highest at 100 m depth (2.4 ± 1.6 ×104 cells L-1) and the next highest at 50 m depth (1.3 ± 0.6 ×104 cells L-1, Fig 3). In the KY1801 cruise, the total cell densities decreased with depth, with the mean cell abundance being highest at 10 m depth (3.7 ± 1.0 ×104 cells L-1).

Fig 3.

Fig 3

Vertical distribution of (a) micro-size plankton density (cells L-1) and (b) their mean community structure for KY1604 (top), KY1704 (middle), and KY1801 (bottom). Haptophyceae and Euglenophyceae are not present in the community structure panel. In the boxplots, the thick vertical line in each box represents the median values. Boxes indicate the lower and upper quartiles. Horizontal lines extending from each box represent the minimum and maximum values. The closed circles are the outliers. The x-axes are transformed square root values.

At the class level including flagellates, Dinophyceae (total mean ± SD: 6.2 ± 5.0 ×104 cells L-1) and flagellates (6.2 ± 6.6 ×104 cells L-1) were the dominant groups in all the samples. They were dominant in at least 62.6% of total cells (mean ± SD: 88.6 ± 7.4%). The third and fourth most dominant groups were the Cryptophyceae (1.0 ± 2.0 × 103 cells L-1) and Bacillariophyceae (0.6 ± 0.5 × 103 cells L-1). These findings were similar among the cruises and depths, but the contributions of Cryptophyceae and Bacillariophyceae were slightly higher below depths of 100 m (Fig 3). The other five groups were rare (mean abundance was <1% of the total cell density). Prasinophyceae was the fifth most dominant class on average. Pyramimonas spp. was the only identified in the Prasinophyceae and detected from 69 of 127 samples. Dictyochophyceae, only Dictyocha spp. identified and detected, was observed in 81 of 127 samples, but the numeric contribution was <2.7%. Trichodesmium spp. was the only counted in the Cyanophyceae class, which contributed <0.5% to the numerical base and was detected in 33 of 127 samples. In the glutaraldehyde-fixed samples, but limited to 10 m depth and SCM layers, DDAs (Richelia-Hemiaulus, Richelia-Rhizosolenia, and Richelia-Chaetoceros symbioses) were always <70 host cells L-1 at 10 m depth, and <9 host cells L-1 at the SCM layer. Haptophyceae (only detected Phaeocystis spp.) were observed only in two samples, but were dominant at 19.8% and 28.8% of total plankton cells in the samples. Euglenophyceae were observed only in three samples, with negligible contribution (<1.1% of total cells). Detailed vertical distributions of plankton density at class level showed that the KY1801 cruise differed from the other two cruises (Fig 4). In the KY1604 and KY1704 cruises, the median values of densities of planktons, except Cyanophyceae (Trichodesmium spp.), were highest at SCM or 100 m depth (Fig 4). The abundance of Trichodesmium spp. was highest at 10 m and 50 m depth in the KY1604 and KY1704 cruises, respectively. In contrast with the KY1604 and KY1704 cruises, the abundances, except Dictyochophyceae (Dictyocha spp.) and Cryptophyceae, were highest at 10 m depth in the KY1801 cruise.

Fig 4.

Fig 4

Vertical distributions of densities of the seven major plankton groups, i.e., (a) flagellates, (b) Dinophyceae, (c) Bacillariophyceae, (d) Cryptophyceae, (e) Prasinophyceae, (f) Dictyochophyceae, and (g) Cyanophyceae (Trichodesmium spp.) for KY1604 (top), KY1704 (middle), and KY1801 (bottom) cruises. The thick vertical line in each box represents the median values. Boxes indicate the lower and upper quartiles. Horizontal lines extending from each box represent the minimum and maximum values. The open circles are outliers. The x-axes are transformed logarithm values.

Dinophyceae were detected in 21 genera, one family (Gymnodiniaceae), and one order (Peridiniales). The contributions of Gymnodiniaceae and Peridiniales were large (Fig 5), with the cell numbers of Gymnodiniaceae dominant at 26.2–94.4% of Dinophyceae (mean 70.8%), and Peridiniales was dominant at 0–58.4% (mean 14.9%). When the mean contribution was calculated with every cruise and layer, only six genera were recorded >1% mean contribution of Dinophyceae (Gyrodinium, Oxytoxum, Pronoctiluca, Prorocentrum, Scrippsiella, and Heterocapsa, Fig 5). In the KY1604 cruise, Oxytoxum was abundant at genus level from 10 m to the SCM layer, and in the other two cruises, Heterocapsa was dominant or at the same level as Oxytoxum from 10 m to the SCM layer (Fig 5). At 200 m depth, Pronoctiluca was abundant for all three cruises (Fig 5).

Fig 5.

Fig 5

Vertical distribution of assemblages of (a) Dinophyceae and (b) Bacillariophyceae for KY1604 (top), KY1704 (middle), and KY1801 (bottom) cruises.

Bacillariophyceae were detected in 34 genera, all identified into genus level. The vertical distribution of the composition of Bacillariophyceae in the three cruises are shown in Fig 5. The genus of which the contributions were always <5% of the total Bacillariophyceae in every cruise and layer is treated as the other Bacillariophyceae. The contribution to the total Bacillariophyceae abundance differed among the layers (Fig 5), at depths of 10 and 50 m, Navicula and Nitzschia were dominant on average and, below the SCM, Fragilariopsis was dominant on average. Pseudo-nitzschia was significantly present below 100 m depth. Cylindrotheca was also significantly present at 100 m and the SCM layer in the KY1604 and KY1704 cruises and significantly present at 10 m depth in the KY1801 cruise. Thalassiosira was sometimes abundant >20% of Bacillariophyceae (e.g., at 200 m in the KY1704 cruise).

Response to environmental parameters based on GAMs

Because the Haptophyceae and Euglenophyceae were rarely detected in the samples, we did not perform GAMs on these two classes. Flagellates, Dinophyceae, Bacillariophyceae, and Cryptophyceae were detected and observed from >70% of samples, and we performed the GAMs only on their abundance. The model descriptions and deviance explained values are presented in Table 2. The full model was the least-AICc model in the case of Dinophyceae and Cryptophyceae and, in the other two groups, Ro was removed from the optimal model (Table 2). In other words, temperature, salinity, latitude, and nitrate concentration were selected as the explanatory variables in all the optimal models of these four groups (Table 2). While Ro remained in the optimal models in the cases of Dinophyceae and Cryptophyceae, the response was not significant (p > 0.05).

Table 2. Corrected Akaike information criterion (AICc)-selected best GAM descriptions and deviance explained for every micro-size plankton abundance.

Class Least-AICc models Deviance explained (%)
Flagellates
(abundance) s(T) + s(S) + s(Lat) + s(logNO3) + β 43.3
Dinophyceae
(abundance) s(T) + s(S) + s(Lat) + s(logNO3) + s(Ro) + β 47.1
Bacillariophyceae
(abundance) s(T) + s(S) + s(Lat) + s(logNO3) + β 33.4
Cryptophyceae
(abundance) s(T) + s(S) + s(Lat) + s(logNO3) + s(Ro) + β 33.9
Prasinophyceae
(present–absent) s(S) + s(Lat) + s(logNO3) + s(Ro) + β 27.4
(abundance) s(S) + s(logNO3) + β 34.3
Dictyochophyceae
(present–absent) s(T) + s(S) + s(Ro) + β 28.2
(abundance) s(T) + s(Lat) + s(logNO3) + s(Ro) + β 21.8
Cyanophyceae
(present–absent) s(T) +s(Lat) + β 34.0
(abundance) s(S) + s(Lat) + β 33.6

The s denotes the smoothing function, and T, S, Lat, logNO3, Ro, and β denote water temperature, sea salinity, latitude, common logarithm transformed nitrate concentration, relative vorticity, and intercept, respectively.

Temperature showed unimodal effects (Dinophyceae and Bacillariophyceae) or monotonical negative effects with warming (Dinophyceae and Cryptophyceae) to the abundance of every group (Fig 6). Here, the spline function indicated the relationship between the target group abundance and the target parameter after the other parameters were fixed, and the additive effect 0 was set as the mean abundance of targets. In other words, the negative additive effects indicated that the target abundances are expected lower than the mean abundance, and the positive effects indicated the target abundances are expected higher than the mean abundance. When the additive effect of one parameter was observed wider range than the other parameter, the effects of that parameter was larger than the other parameters.

Fig 6.

Fig 6

Partial effects of GAMs on (a) flagellates, (b) Dinophyceae, (c) Bacillariophyceae, (d) Cryptophyceae. The black lines are the smoothing terms, and shadows denote the 95% confidence intervals. The blank panels denote the parameters that were not selected in using the AIC selection.

The effects of temperature showed a peak at 23–24°C to the abundances of Bacillariophyceae, and at 20–21°C to those of Dinophyceae (Fig 6). The response to salinity was consistent in all four groups, and the abundance decreased linearly with salinity (Fig 6). The response to latitude showed unimodal effects (Flagellates and Cryptophyceae) or monotonical negative effects with high latitude (Dinophyceae and Bacillariophyceae, Fig 6). The response to the nitrate concentration was similar among four groups, except Bacillariophyceae (Fig 6). The abundance of Bacillariophyceae decreased under a nitrate-depleted condition and showed a peak at sub-micromolar condition (Fig 6).

The delta GAMs were applied to Dictyochophyceae, Prasinophyceae, and Cyanophyceae (Fig 7, Table 2). Different environmental parameters were selected (Table 2), and different environmental responses were sometimes shown between the present–absent GAMs and the abundance GAMs (Fig 7).

Fig 7.

Fig 7

Partial effects of delta GAMs on (a) Prasinophyceae, (b) Dictyochophyceae, and (c) Cyanophyceae (Trichodesmium spp.). The blue line with blue shadow indicates the result of the present–absent GAM, and that of the red the result of the abundance GAM. The shadows denote the 95% confidential intervals.

Temperature was not selected in the GAMs on Prasinophyceae presence and abundance (Fig 7A). Temperature was selected as explanatory variables in the cases of Dictyochophyceae and Cyanophyceae (only their presence, Fig 7). The presence of Dictyochophyceae showed a peak between 24 and 26°C (Fig 7), and the temperature effect on their abundance was selected but was not significant (p > 0.05, Fig 7B). The presence probability of Cyanophyceae increased in warm waters (Fig 7C). The response to salinity was ambiguous to the presence probability when selected as explanatory variable (Fig 7A and 7B), but showed significantly negative effects to the abundances of Prasinophyceae and Cyanophyceae (Fig 7A and 7C). The latitude showed that the presence probabilities of Prasinophyceae and Cyanophyceae increased in the high-latitude area. On the other hand, the abundance of Dictyochophyceae was high in the low-latitude area, and the abundance of Cyanophyceae was ambiguous. Both the presence probability and abundance were high in nitrate-depleted water in the case of the Prasinophyceae (Fig 7A). The presence probability of Dictyochophyceae was also high in low-nitrate water, less than 1 μM. The nitrate concentration was not selected as the explanatory variable in the case of Cyanophyceae. The Ro was selected in the case of Dictyochophyceae, but the effects on their presence probability and abundance were not significant.

Variations in assemblages based on PCoA

The PCoA, applied to class-level plankton assemblages, summarized approximately three quarters of variation in assemblages with first and second axes (Axis 1 and 2, respectively) based on the eigenvalues (Fig 8). Temperature, common logarithm nitrate concentration, varied significantly with the assemblages summarized in Axis 1 and 2 of class-level PCoA (p < 0.05); however, the other parameters (i.e., salinity, latitude, and Ro) were not significant. Axis 1 explained 57.6% of variations and Axis 2 explained 18.7% of variations, denoting that positive values were warm and low nitrate concentrations (Fig 8). The weighting average score of every class denoted that two rare classes, Haptophyceae (Hap) and Euglenophyceae (Eug), differed from the other classes (Fig 8). The other classes were plotted in similar positions and their score was positive along Axis 1. Along Axis 2, Cyanophyceae (Cya) and Dictyochophyceae (Dic) were scored positive, but the other classes, i.e., Bacillariophyceae (Bac), Dinophyceae (Din), Prasinophyceae (Pra), Cryptophyceae (Cry) and flagellates (Fla), were scored negative. The 95% ellipse of every sampling layer based on site scores of class-level PCoA denoted that plots of clusters of sites scored at the 200 m depth layer differed slightly from the that of other layers. In addition, the site scores at 10 m depth layer varied most widely (Fig 8).

Fig 8.

Fig 8

Triplots of the results on the PCoA of (a) class level, (b) those of the Dinophyceae, and (c) those of the Bacillariophyceae. The gray and blue plots are sample scores and target group scores, and the arrows are environmental parameters. The ellipses divided by color and line type denote plotted ranges of sample scores at every sampling layer regardless of the cruises. The abbreviations TEMP, SAL, and NO3 denote temperature, salinity, and nitrate concentrations, respectively. The abbreviations of target group (dark blue fonts) are shown in the main text.

The sample scores of the PCoA of the Dinophyceae showed the Dinophyceae community at 200 m depth differed from others, as well as the other PCoA (Fig 8). On the other hand, the scores of every genus, including Peridiniales (Peri) and Gymnodiniaceae (Gymn), plotted to quite similar positions along Axis 1. Along Axis 2, Heterocapsa (Hete), Gyrodinium (Gyro), Pronoctiluca (Pron), and Gymnodiniaceae plotted in the negative, and Oxytoxum (Oxyt), Prorocentrum (Proc), Scrippsiella (Scri), and Peridiniales plotted in the positive. Temperature and nitrate concentration were correlated significantly with the site scores.

The PCoA applied to abundance of the genus of Bacillariophyceae summarized 39.4% of variations with Axis 1 and 2. Temperature, salinity, and nitrate concentration significantly explained the variations. Axis 1 explained 24.8% of variation and Axis 2 explained 14.6% of variation, denoting that negative values plotted in the warm, less-saline, and nitrate-poor conditions (Fig 8). The sample scores denoted that the Bacillariophyceae community at 200 m depth differed from those at the other depths, as well as the class level and Dinophyceae community structure. The scores of every genus showed two clusters, one of which was consistent with Rhizosolenia (Rhi), Navicula (Nav), Hemiaulus (Hem), and Nitzschia (Nit), and the other was consistent with Chaetoceros (Cha), Cylindrotheca (Cyl), Fragilariopsis (Fra), Nanoneis (Nan), Pseudo-nitzschia (Pse), and Thalassiosira (Tha). The scores of samples at depths of 10 and 50 m plotted near the first groups.

Discussion

Our study focused on micro-size plankton abundance and assemblages in the western NPSG. Our numerical abundance-based micro-size plankton community structure showed results similar to carbon-based results in the western NPSG [6] and numerical abundance-based results [16] conducted over three decades. Interestingly, comparing the results of this study and the observations of half a century ago [6, 16] indicated that the micro-size plankton community structure could be considered conservative in the western NPSG, whereas primary production in the western NPSG depicted slight increase [18]. Our investigation, however, could not determine the density of coccolithophores that make a significant contribution as the primary producer of this area [6, 16], with ~10% of carbon explained by coccolithophores [6]. The abundance of coccolithophores was between that of Bacillariophyceae and Dinophyceae [16]. Therefore, further investigation is necessary to improve understanding of the entire micro-size plankton community of this area.

The recent results of eukaryotic 18S rRNA gene sequence compositions based on metabarcoding technique also showed Dinophyceae are dominant >75% of eukaryotic community in the western NPSG [39]. The metabarcoding results [39] were consistent with our microscopic observations, with the exception of Dinophyceae which comprised a larger proportion of the community in metabarcoding results. This is likely due to Dinophyceae having larger nuclear genome sizes compared to other microplankton [40]. In addition, the contribution of other flagellates was remained very low in the metabarcoding technique results [39]. This observation suggests that while microscopic observation is a time consuming technique, it is still a necessary tool to understand the micro-size plankton community in the oligotrophic ocean; furthermore, this indicates that the metabarcoding technique should best be coupled with more traditional methods such as microscopic observations.

The contribution of micro-size plankton to primary production is low, as the contributions of >10 μm-Chl a concentration to the total Chl a concentration in the discrete samples were almost <10%. In addition, at every depth, flagellates and Dinophyceae were abundant. The dominant Dinophyceae in our study area (Peridiniales and Gymnodiniaceae) are considered heterotrophs [4143]. In addition, most of the pico- and nano-size phytoplankton of this area are considered mixotrophs [44]. The proportion of autotrophic flagellates was not evaluated in this study. However, we did consider the micro-size plankton community as not comprising mainly primary producers, but were ranked as secondary or more higher producers. These predacious flagellates and Dinophyceae prey on pico-size plankton, which is abundant in this region. The Dinophyceae were detected from larval eel guts based on the metabarcoding technique [45]. Therefore, the microbial loop is suggested important for the biological production of this area, and micro-size plankton is the link between pico-size plankton and higher trophic animals.

The elevations of abundances in the less-saline water are the common feature, except Dictyochophyceae, shown in GAMs. The mechanism of high micro-size plankton abundance in the less-saline water was not explained in previous studies, but Shiozaki et al. [46] reported that active nitrogen fixation and primary production are observed in the less-saline (<34.2) water mass of the North Equatorial Current. In the less-saline water observed by Shiozaki et al. [46], Trichodesmium (and R. intracellularis) increased, similar to the findings of the present study. Therefore, we considered that the elevation of micro-size plankton in the less-saline water could be attributed to active nitrogen fixation. The importance of nitrogen fixation in this area is also pointed out in the results of the nitrogen isotope ratios of pelagic fish and squid muscle [47].

We could not detect the effects of meso-scale eddies on the micro-size plankton abundance (and also assemblages), whereas other studies [2729] have reported that the phytoplankton abundance vary with the eddies. Several studies focused on the variations in chlorophyll a concentration, but micro-size plankton abundance and assemblages vary with meso-scale eddies, as reported by Davis and McGillicuddy [48]. Two reasons were considered for our results on the meso-scale eddies, namely 1) the difference in observation of oceanic basins, and 2) the limitations of our approach (observations and statistical models). First, previous studies had been conducted in the central and eastern NPSG [2729], and the heterogeneity of plankton abundance and assemblages with the meso-scale eddies are observed rarely in the western NPSG. Second, our data sets were a mixture of observations of three cruises, of which the primary aims were investigation of the distribution of Japanese eel larvae. Well-organized observations would be necessary for evaluation of the effects of meso-scale eddies on micro-size plankton abundance and assemblages.

The heterogeneity of the plankton assemblages was another topic of our study. Among the dominant four groups at the class level, similar responses to salinity and latitude were observed in the GAMs. Habitat segregation was caused mainly by temperature and nitrate concentration among the four classes, based on the GAMs. Our results in this regard are consistent with the results of the PCoA. Comparing the response to temperature among four groups, Bacillariophyceae increased in abundance and proportion in the micro-zooplankton in warm water, and Cryptophyceae and flagellates increased in cold water. The response to the nitrate concentration indicated that Bacillariophyceae abundance increased at the nitracline, and the other three groups were increased in the nitrate-depleted waters. Collinearity was not high, but the relationship between temperature and nitrate concentration showed negative correlations; therefore, we rarely found warm and nitrate-rich condition nor cold and nitrate-poor condition at our study sites. We considered that these conflicted optimal conditions observed in these four classes could be attributed to the mixing of several genera, as shown in the PCoA of Bacillariophyceae.

Wei et al. [49] reported that Bacillariophyceae are more abundant than are the Dinophyceae in the western NPSG. Compared with micro-size plankton assemblages in the eastern NPSG [7], our results were similar at class level, but Bacillariophyceae abundance in the eastern NPSG was higher than that indicated in our study [7]. In the eastern NPSG, nitrogen-fixing diatom symbioses blooms are sometimes observed [10], and vertical migration of Rhizosolenia is observed [50]. During the KY1704 and KY1801 cruises, DDAs abundance was <70 host cells L-1, and such a low abundance of DDAs could be the reason for low Bacillariophyceae abundance in our observations compared with the results from Wei et al. [49] and Venrick [7]. Therefore, we considered that Bacillariophyceae-dominant waters could be present in the western NPSG, but at a limited scale. Our results on the abundance of Bacillariophyceae are consistent with the findings of Hashihama et al. [51], who reported that the abundance of Bacillariophyceae was several tens of cells L-1 at the surface of the western NPSG, except a diatom bloom in a cyclonic eddy at the northern edge of NPSG.

The results of PCoAs indicated that community structure at 200 m depth differed from those at other depths. At 200 m depth, nutrient concentration was at micromolar level, and the depth at which light intensity is 1% of the surface irradiance is reported as <~150 m in the western NPSG [46, 52]. The mean attenuation coefficient of photosensitive available radiation (PAR) was -0.0457 m-1 during the other western NPSG cruise in 2019, which was conducted under the same project of this study, and it suggested that a compensation depth of 143 m for a light level of 0.1%, and a compensation depth of 95 m for a light level of 1%. Therefore, light is the potential limiting factor in the growth of micro-size phytoplankton at 200 m depth. Similar environmental conditions were observed at the SCM-Low layer (at 166 ± 8 m), but the community structure differed from that at 200 m depth. Therefore, a unique micro-size plankton community was formed at 200 m depth compared with the other sunlit layers.

Latasa et al. [53] investigated the fine-scale vertical distribution of the phytoplankton community structure in the subtropical North Atlantic and concluded that nutrients and light are the most evident environmental variables for the difference in community structure. Some studies in the western NPSG indicated that the phytoplankton community in the deep ocean depended mainly on sinking from the sunlit surface [54, 55]. In this study, however, sinking from the shallower layers were not considered as the main factor of plankton community at 200 m depths because PCoA showed the micro-size plankton community at 200 m depth was unique. Some of the phytoplankton community at 200 m depth, such as Fragilariopsis could adapt to the low-light condition.

Comparing the heterogeneity of plankton assemblages among the sampling layers showed the plankton communities at 10 m depth were most heterogeneous based on PCoA, except those of Bacillariophyceae. At 10 m depth, the nitrate concentration was almost depleted, and the heterogeneous plankton assembly of this layer was not based on the heterogeneous nitrate supply from the deep water. We considered that nitrogen fixation could make the community heterogeneous in this area. Biotic nitrogen fixation occurs everywhere [56], but not homogenously in the western NPSG [13, 46, 52]. The micro-size plankton abundances were elevated in the less-saline and Trichodesmium-rich surface water. In addition, a part of unicellular nano-size Cyanobacteria has nitrogen fixation ability, and is distributed patchily in our observation area [57]. The difference in the diazotroph community links to the difference in primary production [58]. Therefore, we considered that the micro-size plankton community would be heterogeneous because both the quality and quantity of the nitrogen input via biological nitrogen fixation are heterogeneous at a depth of 10 m.

The micro-size plankton is not the major primary producers of the western NPSG, but they are one of the key groups in the ecosystems linked to the higher trophic organisms. Watanabe et al. [39] reported the direct importance of micro-size plankton for the larval eels growing up in the western NPSG. The other important role of the micro-size plankton of this area is the size-up of organic matter. The western equatorial Pacific is the nursery ground of the Skipjack tuna K. pelamis, and the primary production of that area are positively correlated to the stock (biomass) of K. pelamis in that area [59]. However, K. pelamis larvae are carnivores as well as the other tuna larvae [60]: they prey meso- and macro-size plankton. The pico- and nano size plankton is the main primary producers in the western NPSG, and thus the size-up processes of organic matter by micro-size plankton are essential for the higher trophic level organisms in the western NPSG. The surface micro-plankton abundance was considered to be controlled with nitrogen fixation in our study, the primary production in the less-saline and Trichodesmium-rich water is doubled comparing to the saline and Trichodesmium-poor water in the western NPSG [46], and the nitrogen isotope ratios of pelagic fish and squid muscle also indicates the importance of nitrogen fixation as the nitrogen source in the marine ecosystem of this area [47]; these facts indicated micro-size plankton links primary production to the higher trophic level organisms including commercially viable migratory fish in the ecosystems of this area.

Conclusion

We investigated micro-size plankton communities and their habitats in the western NPSG based on microscopic observation. Heterotrophic plankton was indicated as abundant in the micro-size plankton community; therefore, microbial loops are considered important for biological production. The assemblages did not differ markedly in the euphotic layer. A unique community was observed only at a depth of 200 m. Most heterogeneous communities of micro-size plankton were recorded at 10 m depth, suggesting that a heterogeneous nitrogen fixation amount and Cyanobacteria groups made the plankton community heterogeneous. In addition, abundant micro-size plankton was observed in surface waters, where diazotrophic micro-size Cyanophyceae (Trichodesmium spp.) are present in rich abundance. Accordingly, understanding the microbial loops and the abundance and community of the diazotrophs is a prerequisite to further evaluation of biological productivity in the western NPSG. We consider genetical approaches, including environmental DNA, as promising methods for understanding the productivity of the western NPSG, but the genetical approaches are partly compatible approaches of microscopic observations at this time.

Acknowledgments

We thank the captains, crews, and scientists of RV Kaiyo-Maru, the Fisheries Agency of Japan, for their invaluable support with sampling. We also thank Kyoko Kawanobe for counting of the large phytoplankton.

Data Availability

All data are available from Mendeley Data (http://dx.doi.org/10.17632/hmy5jzccyx.2).

Funding Statement

This work was supported by grants from the Project of the Bio-oriented Technology Research Advancement Institution, NARO (the special scheme project on advanced research and development for next-generation technology) to all. The funder did not play any role in neither the study design, data collection and analysis, decision to publish, nor preparation of the manuscript.

References

  • 1.Ryther JH. Photosynthesis and fish production in the sea. Science. 1969;166(3901):72–6. 10.1126/science.166.3901.72 [DOI] [PubMed] [Google Scholar]
  • 2.Sherr EB, Sherr BF. Bacterivory and herbivory: Key roles of phagotrophic protists in pelagic food webs. Microb Ecol. 1994;28(2):223–35. 10.1007/BF00166812 [DOI] [PubMed] [Google Scholar]
  • 3.Sherr EB, Sherr BF. Significance of predation by protists in aquatic microbial food webs. Antonie Van Leeuwenhoek. 2002;81(1–4):293–308. 10.1023/a:1020591307260 [DOI] [PubMed] [Google Scholar]
  • 4.Karl DM. A sea of change: Biogeochemical variability in the North Pacific Subtropical Gyre. Ecosystems. 1999;2(3):181–214. 10.1007/s100219900068 [DOI] [Google Scholar]
  • 5.Karl DM, Letelier RM. Nitrogen fixation-enhanced carbon sequestration in low nitrate, low chlorophyll seascapes. Mar Ecol Prog Ser. 2008;364:257–68. 10.3354/meps07547 [DOI] [Google Scholar]
  • 6.Furuya K. Subsurface chlorophyll maximum in the tropical and subtropical western Pacific Ocean: Vertical profiles of phytoplankton biomass and its relationship with chlorophylla and particulate organic carbon. Mar Biol. 1990;107(3):529–39. 10.1007/bf01313438 [DOI] [Google Scholar]
  • 7.Venrick EL. Phytoplankton in an oligotrophic ocean: Species structure and interannual variability. Ecology. 1990;71(4):1547–63. 10.2307/1938291 [DOI] [Google Scholar]
  • 8.White AE, Spitz YH, Letelier RM. What factors are driving summer phytoplankton blooms in the North Pacific Subtropical Gyre? J Geophys Res. 2007;112(C12). 10.1029/2007JC004129 [DOI] [Google Scholar]
  • 9.Scharek R, Latasa M, Karl DM, Bidigare RR. Temporal variations in diatom abundance and downward vertical flux in the oligotrophic North Pacific gyre. Deep Sea Research I. 1999;46(6):1051–75. 10.1016/S0967-0637(98)00102-2 [DOI] [Google Scholar]
  • 10.Villareal TA, Brown CG, Brzezinski MA, Krause JW, Wilson C. Summer diatom blooms in the North Pacific Subtropical Gyre: 2008–2009. PLoS One. 2012;7(4):e33109. 10.1371/journal.pone.0033109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Longhurst R. Ecological geography of the sea Academic Press; 2007. [Google Scholar]
  • 12.Hashihama F, Furuya K, Kitajima S, Takeda S, Takemura T, Kanda J. Macro-scale exhaustion of surface phosphate by dinitrogen fixation in the western North Pacific. Geophys Res Lett. 2009;36(3):L03610. 10.1029/2008gl036866 [DOI] [Google Scholar]
  • 13.Shiozaki T, Furuya K, Kodama T, Kitajima S, Takeda S, Takemura T, et al. New estimation of N2 fixation in the western and central Pacific Ocean and its marginal seas. Global Biogeochem Cycles. 2010;24(1):GB1015. 10.1029/2009gb003620 [DOI] [Google Scholar]
  • 14.Nishikawa Y, Honma M, Ueyanagi S, Kikawa S. Average distribution of larvae of oceanic species of scombroid fishes, 1956–1981. Shimizu 1985. 99 p. [Google Scholar]
  • 15.Tsukamoto K. Oceanic biology: spawning of eels near a seamount. Nature. 2006;439(7079):929. 10.1038/439929a [DOI] [PubMed] [Google Scholar]
  • 16.Taniguchi A. Distribution of microzooplankton in the Philippine Sea and the Celebes Sea in summer, 1972. J Oceanogr Soc Japan. 1977;33(2):82–9. 10.1007/bf02110013 [DOI] [Google Scholar]
  • 17.Behrenfeld MJ, O’Malley RT, Boss ES, Westberry TK, Graff JR, Halsey KH, et al. Revaluating ocean warming impacts on global phytoplankton. Nat Clim Change. 2015;6(3):323–30. 10.1038/nclimate2838 [DOI] [Google Scholar]
  • 18.Behrenfeld MJ, O’Malley RT, Siegel DA, McClain CR, Sarmiento JL, Feldman GC, et al. Climate-driven trends in contemporary ocean productivity. Nature. 2006;444(7120):752–5. 10.1038/nature05317 [DOI] [PubMed] [Google Scholar]
  • 19.Watanabe T, Taniuchi Y, Kakehi S, Sakami T, Kuwata A. Seasonal succession in the diatom community of Sendai Bay, northern Japan, following the 2011 off the Pacific coast of Tohoku earthquake. J Oceanogr. 2017;73(1):133–44. 10.1007/s10872-016-0387-8 [DOI] [Google Scholar]
  • 20.Dodson AN, Thomas WH. Concentrating plankton in a centle fashion. Limnol Oceanogr. 1964;9(3):455–6. 10.4319/lo.1964.9.3.0455 [DOI] [Google Scholar]
  • 21.Utermöhl H. Zur vervollkommung der quantitativen phytolankton methodik. Mitt Int Ver Limnol. 1958;9:1–39. [Google Scholar]
  • 22.Tomas CR. Identifying marine phytoplankton: Elsevier; 1997. [Google Scholar]
  • 23.Chen YL, Tuo S, Chen HY. Co-occurrence and transfer of fixed nitrogen from Trichodesmium spp. to diatoms in the low-latitude Kuroshio Current in the NW Pacific. Mar Ecol Prog Ser. 2011;421:25–38. [Google Scholar]
  • 24.Becker S, Aoyama M, Woodward EMS, Bakker K, Coverly S, Mahaffey C, et al. GO-SHIP repeat hydrography nutrient manual: The precise and accurate determination of dissolved Iinorganic nutrients in seawater, using continuous flow analysis methods. Front Mar Sci. 2020;7(908). 10.3389/fmars.2020.581790 [DOI] [Google Scholar]
  • 25.Suzuki R, Ishimaru T. An improved method for the determination of phytoplankton chlorophyll using N, N-dimethylformamide. J Oceanogr Soc Japan. 1990;46(4):190–4. 10.1007/bf02125580 [DOI] [Google Scholar]
  • 26.Welschmeyer NA. Fluorometric analysis of chlorophyll a in the presence of chlorophyll b and pheopigments. Limnol Oceanogr. 1994;39(8):1985–92. 10.4319/lo.1994.39.8.1985 [DOI] [Google Scholar]
  • 27.Chow CH, Cheah W, Tai J-H, Liu S-F. Anomalous wind triggered the largest phytoplankton bloom in the oligotrophic North Pacific Subtropical Gyre. Sci Rep. 2019;9(1):15550. 10.1038/s41598-019-51989-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Guidi L, Calil PHR, Duhamel S, Björkman KM, Doney SC, Jackson GA, et al. Does eddy-eddy interaction control surface phytoplankton distribution and carbon export in the North Pacific Subtropical Gyre? J Geophys Res Biogeosci. 2012;117(G2). 10.1029/2012jg001984 [DOI] [Google Scholar]
  • 29.Xiu P, Chai F. Eddies affect subsurface phytoplankton and oxygen distributions in the North Pacific Subtropical Gyre. Geophys Res Lett. 2020;47(15):e2020GL087037. 10.1029/2020gl087037 [DOI] [Google Scholar]
  • 30.R Core Team. R: A language and environment for statistical computing. In: R Foundation for Statistical Computing, editor. Vienna, Austria: https://www.R-project.org/. 2020.
  • 31.Guisan A, Zimmermann NE. Predictive habitat distribution models in ecology. Ecol Model. 2000;135(2–3):147–86. 10.1016/s0304-3800(00)00354-9 [DOI] [Google Scholar]
  • 32.Legendre P, Anderson MJ. Distance‐based redundancy analysis: testing multispecies responses in multifactorial ecological experiments. Ecol Monogr. 1999;69(1):1–24. [Google Scholar]
  • 33.Legendre P, Legendre L. Numerical ecology. Amsterdam: Elsevier; 2012. [Google Scholar]
  • 34.Wood SN. Generalized additive models: an introduction with R: CRC press; 2017. [Google Scholar]
  • 35.Guisan A, Thuiller W, Zimmermann NE. Habitat suitability and distribution models: with applications in R: Cambridge University Press; 2017. [Google Scholar]
  • 36.Borcard D, Gillet Fo, Legendre P. Numerical ecology with R. New York: Springer; 2011. xi, 306 p. p. [Google Scholar]
  • 37.Oksanen J, Kindt R, Legendre P, O’Hara B, Stevens MHH, Oksanen MJ, et al. The vegan package. Community ecology package. 2007;10:631–7. [Google Scholar]
  • 38.Falkowski P, Kiefer DA. Chlorophyll a fluorescence in phytoplankton: relationship to photosynthesis and biomass. J Plankton Res. 1985;7(5):715–31. 10.1093/plankt/7.5.715 [DOI] [Google Scholar]
  • 39.Watanabe T, Nagai S, Kawakami Y, Asakura T, Kikuchi J, Inaba N, et al. 18S rRNA gene sequences of leptocephalus gut contents, particulate organic matter, and biological oceanographic conditions in the western North Pacific. Sci Rep. 2021;11(1):5488. 10.1038/s41598-021-84532-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Shoguchi E, Shinzato C, Kawashima T, Gyoja F, Mungpakdee S, Koyanagi R, et al. Draft Assembly of the Symbiodinium minutum Nuclear Genome Reveals Dinoflagellate Gene Structure. Curr Biol. 2013;23(15):1399–408. 10.1016/j.cub.2013.05.062 [DOI] [PubMed] [Google Scholar]
  • 41.Skovgaard A. Role of chloroplast retention in a marine dinoflagellate. Aquat Microb Ecol. 1998;15(3):293–301. [Google Scholar]
  • 42.Jeong HJ, Yoo YD, Kim JS, Seong KA, Kang NS, Kim TH. Growth, feeding and ecological roles of the mixotrophic and heterotrophic dinoflagellates in marine planktonic food webs. Ocean Sci J. 2010;45(2):65–91. 10.1007/s12601-010-0007-2 [DOI] [Google Scholar]
  • 43.Stoecker DK. Mixotrophy among Dinoflagellates. The Journal of Eukaryotic Microbiology. 1999;46(4):397–401. 10.1111/j.1550-7408.1999.tb04619.x [DOI] [Google Scholar]
  • 44.Sato M, Hashihama F. Assessment of potential phagotrophy by pico- and nanophytoplankton in the North Pacific Ocean using flow cytometry. Aquat Microb Ecol. 2019;82(3):275–88. [Google Scholar]
  • 45.Chow S, Inaba N, Nagai S, Kurogi H, Nakamura Y, Yanagimoto T, et al. Molecular diet analysis of Anguilliformes leptocephalus larvae collected in the western North Pacific. PLoS One. 2019;14(11):e0225610. 10.1371/journal.pone.0225610 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Shiozaki T, Kodama T, Kitajima S, Sato M, Furuya K. Advective transport of diazotrophs and importance of their nitrogen fixation on new and primary production in the western Pacific warm pool. Limnol Oceanogr. 2013;58(1):49–60. 10.4319/lo.2013.58.1.0049 [DOI] [Google Scholar]
  • 47.Ohshimo S, Madigan DJ, Kodama T, Tanaka H, Komoto K, Suyama S, et al. Isoscapes reveal patterns of δ13C and δ15N of pelagic forage fish and squid in the Northwest Pacific Ocean. Prog Oceanogr. 2019;175:124–38. 10.1016/j.pocean.2019.04.003 [DOI] [Google Scholar]
  • 48.Davis CS, McGillicuddy DJ. Transatlantic abundance of the N2-fixing colonial cyanobacterium Trichodesmium. Science. 2006;312(5779):1517–20. 10.1126/science.1123570 [DOI] [PubMed] [Google Scholar]
  • 49.Wei Y, Chen Z, Guo C, Zhong Q, Wu C, Sun J. Physiological and ecological responses of photosynthetic processes to oceanic properties and phytoplankton communities in the oligotrophic western Pacific Ocean. Front Microbiol. 2020;11(1774). 10.3389/fmicb.2020.01774 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Villareal TA, Altabet MA, Culver-Rymsza K. Nitrogen transport by vertically migrating diatom mats in the North Pacific Ocean. Nature. 1993;363(6431):709–12. 10.1038/363709a0 [DOI] [Google Scholar]
  • 51.Hashihama F, Kanda J, Maeda Y, Ogawa H, Furuya K. Selective depressions of surface silicic acid within cyclonic mesoscale eddies in the oligotrophic western North Pacific. Deep-Sea Res I. 2014;90:115–24. 10.1016/j.dsr.2014.05.004 [DOI] [Google Scholar]
  • 52.Shiozaki T, Furuya K, Kodama T, Takeda S. Contribution of N2 fixation to new production in the western North Pacific Ocean along 155°E. Mar Ecol Prog Ser. 2009;377:19–32. 10.3354/meps07837 [DOI] [Google Scholar]
  • 53.Latasa M, Cabello AM, Morán XAG, Massana R, Scharek R. Distribution of phytoplankton groups within the deep chlorophyll maximum. Limnol Oceanogr. 2017;62(2):665–85. 10.1002/lno.10452 [DOI] [Google Scholar]
  • 54.Shiozaki T, Hirose Y, Hamasaki K, Kaneko R, Ishikawa K, Harada N. Eukaryotic phytoplankton contributing to a seasonal bloom and carbon export revealed by tracking sequence variants in the western North Pacific. Front Microbiol. 2019;10(2722):2722. 10.3389/fmicb.2019.02722 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sohrin R, Isaji M, Obara Y, Agostini S, Suzuki Y, Hiroe Y, et al. Distribution of Synechococcus in the dark ocean. Aquat Microb Ecol. 2011;64(1):1–14. [Google Scholar]
  • 56.Shiozaki T, Kondo Y, Yuasa D, Takeda S. Distribution of major diazotrophs in the surface water of the Kuroshio from northeastern Taiwan to south of mainland Japan. J Plankton Res. 2018;40(4):407–19. 10.1093/plankt/fby027 [DOI] [Google Scholar]
  • 57.Sato M, Hashihama F, Kitajima S, Takeda S, Furuya K. Distribution of nano-sized cyanobacteria in the western and central Pacific Ocean. Aquat Microb Ecol. 2010;59(3):273–82. 10.3354/ame01397 [DOI] [Google Scholar]
  • 58.Shiozaki T, Bombar D, Riemann L, Sato M, Hashihama F, Kodama T, et al. Linkage between dinitrogen fixation and primary production in the oligotrophic South Pacific Ocean. Global Biogeochem Cycles. 2018;32(7):1028–44. 10.1029/2017gb005869 [DOI] [Google Scholar]
  • 59.Yen KW, Lu HJ. Spatial-temporal variations in primary productivity and population dynamics of skipjack tuna Katsuwonus pelamis in the western and central Pacific Ocean. Fish Sci. 2016;82(4):563–71. 10.1007/s12562-016-0992-x [DOI] [Google Scholar]
  • 60.Young JW, Davis TLO. Feeding ecology of larvae of southern bluefin, albacore and Skipjack tunas (Pisces, Scombridae) in the eastern Indian ocean. Mar Ecol Prog Ser. 1990;61(1–2):17–29. 10.3354/meps061017 [DOI] [Google Scholar]

Decision Letter 0

Arga Chandrashekar Anil

23 Feb 2021

PONE-D-20-37689

Micro-size plankton abundance and assemblies in the western North Pacific Subtropical Gyre under microscopic observation

PLOS ONE

Dear Dr. Kodama,

Thank you for submitting your manuscript to PLOS ONE. After careful consideration, we feel that it has merit but does not fully meet PLOS ONE’s publication criteria as it currently stands. Therefore, we invite you to submit a revised version of the manuscript that addresses the points raised during the review process.

The review that I have received indicates the usefulness of this manuscript to the area of research and has made several suggestions. A thorough revision is sought for further consideration.

Please submit your revised manuscript by Apr 09 2021 11:59PM. If you will need more time than this to complete your revisions, please reply to this message or contact the journal office at plosone@plos.org. When you're ready to submit your revision, log on to https://www.editorialmanager.com/pone/ and select the 'Submissions Needing Revision' folder to locate your manuscript file.

Please include the following items when submitting your revised manuscript:

  • A rebuttal letter that responds to each point raised by the academic editor and reviewer(s). You should upload this letter as a separate file labeled 'Response to Reviewers'.

  • A marked-up copy of your manuscript that highlights changes made to the original version. You should upload this as a separate file labeled 'Revised Manuscript with Track Changes'.

  • An unmarked version of your revised paper without tracked changes. You should upload this as a separate file labeled 'Manuscript'.

If you would like to make changes to your financial disclosure, please include your updated statement in your cover letter. Guidelines for resubmitting your figure files are available below the reviewer comments at the end of this letter.

If applicable, we recommend that you deposit your laboratory protocols in protocols.io to enhance the reproducibility of your results. Protocols.io assigns your protocol its own identifier (DOI) so that it can be cited independently in the future. For instructions see: http://journals.plos.org/plosone/s/submission-guidelines#loc-laboratory-protocols

We look forward to receiving your revised manuscript.

Kind regards,

Arga Chandrashekar Anil, Ph. D., D. Agr.,

Academic Editor

PLOS ONE

Journal Requirements:

When submitting your revision, we need you to address these additional requirements.

1. Please ensure that your manuscript meets PLOS ONE's style requirements, including those for file naming. The PLOS ONE style templates can be found at

https://journals.plos.org/plosone/s/file?id=wjVg/PLOSOne_formatting_sample_main_body.pdf and

https://journals.plos.org/plosone/s/file?id=ba62/PLOSOne_formatting_sample_title_authors_affiliations.pdf

2. We note that you have stated that you will provide repository information for your data at acceptance. Should your manuscript be accepted for publication, we will hold it until you provide the relevant accession numbers or DOIs necessary to access your data. If you wish to make changes to your Data Availability statement, please describe these changes in your cover letter and we will update your Data Availability statement to reflect the information you provide.

3. We note that you have included the phrase “data not shown” in your manuscript. Unfortunately, this does not meet our data sharing requirements. PLOS does not permit references to inaccessible data. We require that authors provide all relevant data within the paper, Supporting Information files, or in an acceptable, public repository. Please add a citation to support this phrase or upload the data that corresponds with these findings to a stable repository (such as Figshare or Dryad) and provide and URLs, DOIs, or accession numbers that may be used to access these data. Or, if the data are not a core part of the research being presented in your study, we ask that you remove the phrase that refers to these data.

[Note: HTML markup is below. Please do not edit.]

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #1: Yes

**********

2. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #1: Yes

**********

3. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #1: No

**********

4. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #1: No

**********

5. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #1: Summary: Authors called for a re-evaluation of the micro-size plankton community in the western NPSG to understand the hetero/homogeneity of their community structure. They collected plankton samples over 3 years (2016-2018) in the euphotic zone (<200 m) and analyzed via microscopy. It is commendable the amount of work that was conducted for this manuscript, using traditional methodology such as microscopy. The work is interesting in that the authors utilized several modeling approaches to elucidate drivers of specific microplankton populations in the NPSG. This work is comprehensive and thorough, despite its claiming that it was a “side project” during expeditions that had an alternative main focus. I believe that it will be a strong contribution to the field, with several key edits that I recommend below. Also, most of the manuscript is written in intelligible fashion and in standard English, except for a few choices of words (e.g. heterogenous) that are not correct, and also the use of parentheses to convey patterns of the opposite trend is confusing. In regards to these examples, the manuscript must be revised to be made more clear.

Major comments:

-This paper provides an overview of microplankton diversity and community structure in the western NPSG through microscopy. The method used is not necessarily novel, and I would have liked to see a combination or groundtruthing of more cutting edge techniques (e.g. DNA metabarcoding) with classic methodologies such as microscopy. I know that is not possible, so perhaps there could be a comparison with studies done with DNA in the area or similar areas to provide this context.

-Production measurements of microplankton would also provide greater understanding of the microplankton contribution and importance to ecosystem functioning. There are several references to this in the Discussion, but there must be historical measurements of carbon fixation in the area? This would also lean into the fisheries discussion that will set this study in a broader context.

-Many of the figures need more explanation in the Results. A few lines to explain the GAMs (Fig. 6-7), in terms of the basics (additive effect, the scales for each panel and how they differ), would be beneficial for the reader.

-I think that a Table showing the depths of the SCMs during each cruise (which is specified in L 226-228) would be more clear, or label the depths in the figures. It is also confusing that the SCM was 94 +/- 10 m in KY1704, but it is plotted beneath 100 m in the plots. I see you need to bin the samples this way because the samples weren’t always taken at the same depth, but it is a bit misleading to see a “depth profile” without the same range between the values on the y-axis. You could also try depth-integrating all the parameters to a specific light bin – 0-100m, 100 m to SCM, SCM to 200 m to break up the parameters around the Chlorophyll a profile. Where is your 1% surface light level? You could alternatively break it up according to light, but this may not be important for the heterotrophic folks – though they would likely be driven by their photosynthetic prey.

-Perhaps the SCM values in the Table could be combined with how many samples were taken at each depth and each cruise? You state that 127 total samples were taken – what is the breakdown (of the box plots in Figs. 3 and 4)? Also, are the environmental parameter sample count the same as the plankton sample count (in Fig. 2)?

Minor Comments:

Title: do they mean “assemblages” versus “assemblies”?

Abstract:

why is a re-evaluation necessary? Maybe one line expanding on what you describe in Introduction:

L31: Peridiniales?

L40: heterogeneous?

L40-43: major conclusions are lacking – the focus of the study is about reevaluating the microplankton, and the major conclusion from the abstract indicate nitrogen fixation is a large contributor.

-Introduction:

L47-49: Instead of using the parentheses, break up the pico vs micro comparison into another sentence.

L60: heterogeneous

Methods:

L96: were these measurements taken on an expedition that had a primary aim to investigate Japanese eel larvae? What does it mean that “other observations were not always organized well in terms of space and time”?

L115: fluorescence

L128: assemblages?

L168-L177: are there satellite altimetry showing sea level anomaly in this study region?

Results:

L234: nitracline is not the depth where there is 1 uM of nitrate concentration, itʻs the region of rapid change in nitrate concentration relative to depth. So for your cruises it looks like between 100-150 m.

Figure 1: the y-axis should be plotted on a linear scale, relative to the SCM. What are the SCM-High, SCM, and SCM-Low depths for each cruise? – applies to all the figures. Or have a table that tells the readers what the SCM depths are for each cruise?

L239-240: why do you think fluorometer values were overestimating the chl a concentration in the deeper layers?

Figure 4: It’s difficult to compare abundances, do they go from highest to less from left to right?

Figure 5: (A) Dinophyceae spelled with a y

L393 and 396 both start with “On the other hand…”

Figure 8b: Dinophyceae speeled with a y

L428: TEM should be TEMP

Discussion:

L466-468: Is this contribution of >10 um chl a concentration integrated from all depths sampled, or per depth? I would think that chl a concentrations and contributions would differ in varying light conditions (e.g. photoadaptation) so this must be evaluated accordingly with respect to light fields.

L515 and other places: When there is the polar opposite condition in parentheses, itʻs confusing. Please clarify.

L554-556, 562, 564, 574: heterogenous means “of foreign origin” and not the same as heterogeneous, which is the word I think you are trying to use.

L573: heterogenetic is not a word, I believe.

L579 to end: This ending sentence about eDNA for migratory fish comes out of nowhere – a few lines linking the study of microphytoplankton to migratory fish (like you do in Intro) in the Discussion would be beneficial.

Acknowledgments:

Any funding to acknowledge?

**********

6. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: No

[NOTE: If reviewer comments were submitted as an attachment file, they will be attached to this email and accessible via the submission site. Please log into your account, locate the manuscript record, and check for the action link "View Attachments". If this link does not appear, there are no attachment files.]

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com/. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Registration is free. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email PLOS at figures@plos.org. Please note that Supporting Information files do not need this step.

Attachment

Submitted filename: PONE_D_20_37689_Review.docx

PLoS One. 2021 Apr 26;16(4):e0250604. doi: 10.1371/journal.pone.0250604.r002

Author response to Decision Letter 0


16 Mar 2021

We copied the response letter which attached after the manuscript file.

Response to Reviewer #1

Summary: Authors called for a re-evaluation of the micro-size plankton community in the western NPSG to understand the hetero/homogeneity of their community structure. They collected plankton samples over 3 years (2016-2018) in the euphotic zone (<200 m) and analyzed via microscopy. It is commendable the amount of work that was conducted for this manuscript, using traditional methodology such as microscopy. The work is interesting in that the authors utilized several modeling approaches to elucidate drivers of specific microplankton populations in the NPSG. This work is comprehensive and thorough, despite its claiming that it was a “side project” during expeditions that had an alternative main focus. I believe that it will be a strong contribution to the field, with several key edits that I recommend below. Also, most of the manuscript is written in intelligible fashion and in standard English, except for a few choices of words (e.g. heterogenous) that are not correct, and also the use of parentheses to convey patterns of the opposite trend is confusing. In regards to these examples, the manuscript must be revised to be made more clear.

We appreciated the reviewer’s constructive comments. As indicated in the reviewer’s comment, this manuscript is the side project, and the data which we can use in this study were limited. For example, the observations were mostly conducted in night-time, and thus we cannot evaluate the light conditions. In this revision, we referred the results on the DNA metabarcoding approach collected in the same cruises but limited. We believe the further discussion using the metabarcoding results make our manuscript better in this revision.

Major comments:

-This paper provides an overview of microplankton diversity and community structure in the western NPSG through microscopy. The method used is not necessarily novel, and I would have liked to see a combination or groundtruthing of more cutting edge techniques (e.g. DNA metabarcoding) with classic methodologies such as microscopy. I know that is not possible, so perhaps there could be a comparison with studies done with DNA in the area or similar areas to provide this context.

The DNA metabarcoding results in the limited cruise and stations were published in very recent (Watanabe et al. 2021 Scientific Reports); we can refer their results in the revised MS. The results of the DNA metabarcoding showed that Dinophytes are dominant group (over 90% in >3–10 µm) in the western NPSG. The Radiolaria and Copepods were following groups in >10 µm. The DNA metabarcoding approach also cannot clear the community structure of flagellates. We considered the results of metabarcoding are interesting, but we questioned its quantitative capability. Therefore, the morphological analyses were still effective for understanding the plankton community in the ocean. We added the comparison between metagenetic approaches and morphological approaches in the discussion (L. 478-488), and revised the conclusion (L625-628).

-Production measurements of microplankton would also provide greater understanding of the microplankton contribution and importance to ecosystem functioning. There are several references to this in the Discussion, but there must be historical measurements of carbon fixation in the area? This would also lean into the fisheries discussion that will set this study in a broader context.

We accepted this comment. We referred some studies on the primary production of this area (Shiozaki et al. 2013 and Yen & Lu 2016), and added the discussion the relationships between fisheries and micro-sized plankton (L593-611).

-Many of the figures need more explanation in the Results. A few lines to explain the GAMs (Fig. 6-7), in terms of the basics (additive effect, the scales for each panel and how they differ), would be beneficial for the reader.

We accepted this comment. We added (L366-373).

-I think that a Table showing the depths of the SCMs during each cruise (which is specified in L 226-228) would be more clear, or label the depths in the figures. It is also confusing that the SCM was 94 +/- 10 m in KY1704, but it is plotted beneath 100 m in the plots. I see you need to bin the samples this way because the samples weren’t always taken at the same depth, but it is a bit misleading to see a “depth profile” without the same range between the values on the y-axis. You could also try depth-integrating all the parameters to a specific light bin – 0-100m, 100 m to SCM, SCM to 200 m to break up the parameters around the Chlorophyll a profile. Where is your 1% surface light level? You could alternatively break it up according to light, but this may not be important for the heterotrophic folks – though they would likely be driven by their photosynthetic prey.

-Perhaps the SCM values in the Table could be combined with how many samples were taken at each depth and each cruise? You state that 127 total samples were taken – what is the breakdown (of the box plots in Figs. 3 and 4)? Also, are the environmental parameter sample count the same as the plankton sample count (in Fig. 2)?

On these two comments, we added information on the depth of SCM in Table 1 (L101), and revised the figures as the order of the sampling depth. We are sorry, but we cannot measure light condition in these cruises because most of the observations were conducted in night-time. We added the information of light-conditions in the other cruises conducted in the same areas and in day-time (L560-564) for supporting that 200 m depth was not a sun-lit layer.

Minor Comments:

Title: do they mean “assemblages” versus “assemblies”?

We revised as suggested. We revised in the text as well (L1).

Abstract:

why is a re-evaluation necessary? Maybe one line expanding on what you describe in Introduction:

We added (L17-20).

L31: Peridiniales?

L40: heterogeneous?

Sorry for careless mistakes. We revised (L31, and L40)

L40-43: major conclusions are lacking – the focus of the study is about reevaluating the microplankton, and the major conclusion from the abstract indicate nitrogen fixation is a large contributor.

We added as suggested (L23-24)

-Introduction:

L47-49: Instead of using the parentheses, break up the pico vs micro comparison into another sentence.

We added as suggested (L47-50).

L60: heterogeneous

Sorry for typo; we revised (L61 and others).

Methods:

L96: were these measurements taken on an expedition that had a primary aim to investigate Japanese eel larvae? What does it mean that “other observations were not always organized well in terms of space and time”?

We revised the sentence for clarifying the meaning (L97-99).

L115: fluorescence

L128: assemblages?

We revised (L124 and L137).

L168-L177: are there satellite altimetry showing sea level anomaly in this study region?

Yes, but the sea level anomaly with satellites was 8 days composited data; therefore we chose the reanalysis data. Sea surface height anomaly was reflected in the geostrophic velocity anomalies in our study. We described the production ID (L186).

Results:

L234: nitracline is not the depth where there is 1 uM of nitrate concentration, itʻs the region of rapid change in nitrate concentration relative to depth. So for your cruises it looks like between 100-150 m.

We revised the sentence (L244-245).

Figure 1: the y-axis should be plotted on a linear scale, relative to the SCM. What are the SCM-High, SCM, and SCM-Low depths for each cruise? – applies to all the figures. Or have a table that tells the readers what the SCM depths are for each cruise?

This is the comment on Figure 2. The SCM layer was different among the stations. When the average depth of SCM was put in the y-axis, the balance of the figure was bad. Therefore, we added the new table (Table 1), and describe the depth of SCM as well as depth of SCM-High and SCM-Low. The relationship between SCM and 100 m depth during the KY1704 cruise were revised.

L239-240: why do you think fluorometer values were overestimating the chl a concentration in the deeper layers?

The reason was not cleared in our study, but the Falkowski and Keifer (1985) reviewed some reasons that in vivo fluorometer often does not adequately represent in situ chlorophyll a concentration. We added (L250).

Figure 4: It’s difficult to compare abundances, do they go from highest to less from left to right?

We arranged the x-axis at the same scales (from 0 to 120) among the taxon but different orders (Figure 4).

Figure 5: (A) Dinophyceae spelled with a y

Sorry for typo. We revised (Figure 5).

L393 and 396 both start with “On the other hand…”

We revised (L405).

Figure 8b: Dinophyceae speeled with a y

Sorry for typo. We revised. (Figure 8)

L428: TEM should be TEMP

We revised as suggested (L440).

Discussion:

L466-468: Is this contribution of >10 um chl a concentration integrated from all depths sampled, or per depth? I would think that chl a concentrations and contributions would differ in varying light conditions (e.g. photoadaptation) so this must be evaluated accordingly with respect to light fields.

This is the results of discrete samples and extracted chlorophyll a concentration. We consider the photoadaptation can be ignored in the analyses of extracted chlorophyll a concentration. We revised the sentence and clarify the sentence (L489-491).

L515 and other places: When there is the polar opposite condition in parentheses, itʻs confusing. Please clarify.

We revised (L539-540).

L554-556, 562, 564, 574: heterogenous means “of foreign origin” and not the same as heterogeneous, which is the word I think you are trying to use.

L573: heterogeneous is not a word, I believe.

Sorry for typo. We revised (L580 and others).

L579 to end: This ending sentence about eDNA for migratory fish comes out of nowhere – a few lines linking the study of microphytoplankton to migratory fish (like you do in Intro) in the Discussion would be beneficial.

We added the discussion on the relationships between micro size plankton and migratory fish (L593-611).

Acknowledgments:

Any funding to acknowledge?

In the PLOS submission guideline, we found these sentences in the acknowledgements: Do not include funding sources in the Acknowledgments or anywhere else in the manuscript file. Funding information should only be entered in the financial disclosure section of the submission system.

Attachment

Submitted filename: PONE_D_20_37689_response_letter.docx

Decision Letter 1

Arga Chandrashekar Anil

7 Apr 2021

PONE-D-20-37689R1

Micro-size plankton abundance and assemblages in the western North Pacific Subtropical Gyre under microscopic observation

PLOS ONE

Dear Dr. Kodama,

Thank you for submitting your manuscript to PLOS ONE. After careful consideration, we feel that it has merit but does not fully meet PLOS ONE’s publication criteria as it currently stands. Therefore, we invite you to submit a revised version of the manuscript that addresses the points raised during the review process.

The reviewer has recommended corrections to the text. Based on the opinion of the reviewer and my own assessment the manuscript can be considered for publication. Before the acceptance of the manuscript, I would like to see the corrected version.

Please submit your revised manuscript by May 22 2021 11:59PM. If you will need more time than this to complete your revisions, please reply to this message or contact the journal office at plosone@plos.org. When you're ready to submit your revision, log on to https://www.editorialmanager.com/pone/ and select the 'Submissions Needing Revision' folder to locate your manuscript file.

Please include the following items when submitting your revised manuscript:

  • A rebuttal letter that responds to each point raised by the academic editor and reviewer(s). You should upload this letter as a separate file labeled 'Response to Reviewers'.

  • A marked-up copy of your manuscript that highlights changes made to the original version. You should upload this as a separate file labeled 'Revised Manuscript with Track Changes'.

  • An unmarked version of your revised paper without tracked changes. You should upload this as a separate file labeled 'Manuscript'.

If you would like to make changes to your financial disclosure, please include your updated statement in your cover letter. Guidelines for resubmitting your figure files are available below the reviewer comments at the end of this letter.

If applicable, we recommend that you deposit your laboratory protocols in protocols.io to enhance the reproducibility of your results. Protocols.io assigns your protocol its own identifier (DOI) so that it can be cited independently in the future. For instructions see: http://journals.plos.org/plosone/s/submission-guidelines#loc-laboratory-protocols. Additionally, PLOS ONE offers an option for publishing peer-reviewed Lab Protocol articles, which describe protocols hosted on protocols.io. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols.

We look forward to receiving your revised manuscript.

Kind regards,

Arga Chandrashekar Anil, Ph. D., D. Agr.,

Academic Editor

PLOS ONE

Journal Requirements:

Please review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the rebuttal letter that accompanies your revised manuscript. If you need to cite a retracted article, indicate the article’s retracted status in the References list and also include a citation and full reference for the retraction notice.

[Note: HTML markup is below. Please do not edit.]

Reviewers' comments:

Reviewer's Responses to Questions

Comments to the Author

1. If the authors have adequately addressed your comments raised in a previous round of review and you feel that this manuscript is now acceptable for publication, you may indicate that here to bypass the “Comments to the Author” section, enter your conflict of interest statement in the “Confidential to Editor” section, and submit your "Accept" recommendation.

Reviewer #1: All comments have been addressed

**********

2. Is the manuscript technically sound, and do the data support the conclusions?

The manuscript must describe a technically sound piece of scientific research with data that supports the conclusions. Experiments must have been conducted rigorously, with appropriate controls, replication, and sample sizes. The conclusions must be drawn appropriately based on the data presented.

Reviewer #1: Yes

**********

3. Has the statistical analysis been performed appropriately and rigorously?

Reviewer #1: Yes

**********

4. Have the authors made all data underlying the findings in their manuscript fully available?

The PLOS Data policy requires authors to make all data underlying the findings described in their manuscript fully available without restriction, with rare exception (please refer to the Data Availability Statement in the manuscript PDF file). The data should be provided as part of the manuscript or its supporting information, or deposited to a public repository. For example, in addition to summary statistics, the data points behind means, medians and variance measures should be available. If there are restrictions on publicly sharing data—e.g. participant privacy or use of data from a third party—those must be specified.

Reviewer #1: Yes

**********

5. Is the manuscript presented in an intelligible fashion and written in standard English?

PLOS ONE does not copyedit accepted manuscripts, so the language in submitted articles must be clear, correct, and unambiguous. Any typographical or grammatical errors should be corrected at revision, so please note any specific errors here.

Reviewer #1: Yes

**********

6. Review Comments to the Author

Please use the space provided to explain your answers to the questions above. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. (Please upload your review as an attachment if it exceeds 20,000 characters)

Reviewer #1: Review #2

Micro-size plankton abundance and assemblages in the western North Pacific Subtropical Gyre under microscopic observation

In this version, the authors have addressed all of the comments of the reviewers to satisfaction. I have made few, mostly grammatical suggestions here:

L5 typo in the short title

L18-20 Edit first sentence to: “While primary productivity in the oligotrophic North Pacific Subtropical Gyre (NPSG) is changing, the micro-size plankton community has not been evaluated in the last 3 decades, prompting a re-evaluation.”

L23 Don’t need s in community structures.

L24 “The assemblages were consistent with those identified 4 decades previously.” Also 4 decades? Thought the first sentence indicated 3 decades.

L25 Dinophyceae were “the most numerically abundant, followed by Cryptophyceae and Bacillariophyceae (diatoms)”.

L42 Therefore, nitrogen fixation “may contribute” to the heterogeneity …

L49-50 In other words, energy fixed by primary production is more efficiently transferred to higher trophic organisms in micro-plankton dominated waters.

L61 is instead of as

L62 phytoplankton are “the dominant primary producers”;

L64 abundant but is it significant? What % of cell counts or Chl a?

L69 western NPSG is oligotrophic (no ‘the’)

L70 migratory fish?

L88-89 To understand the trophic structure and heterogeneity of this community.

L93 three cruises on the R/V …

L98-99 therefore, the observations for collections of seawater were limited to night-time collections.

Table 1 is very informative and much improved, thank you!

L121 – collected using … (no “by”)

L244 Nitrate concentration was over 1 uM of in water “where” the temperature was below …

L250 …layers as indicated “by” Falokowski …

L472 …primary production in the wester NPSG depicted slight increase.

L480 The metabarcoding results were consistent “with” our microscopic observation, with the exception of Dinophyceae which comprised a larger proportion of the community in metabarcoding results. This is likely due to Dinophyceae having largernuclear genome sizes compared to other microplankton.

L484-490 This observation suggests that while microscopic observation is a time consuming technique, it is still a necessary tool to understand the micro…..; furthermore, this indicates that the metabarcoding technique should best be coupled with more traditional methods such as microscopic observations.

**********

7. PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: No

[NOTE: If reviewer comments were submitted as an attachment file, they will be attached to this email and accessible via the submission site. Please log into your account, locate the manuscript record, and check for the action link "View Attachments". If this link does not appear, there are no attachment files.]

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com/. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Registration is free. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email PLOS at figures@plos.org. Please note that Supporting Information files do not need this step.

PLoS One. 2021 Apr 26;16(4):e0250604. doi: 10.1371/journal.pone.0250604.r004

Author response to Decision Letter 1


8 Apr 2021

Response to Reviewer

In this version, the authors have addressed all of the comments of the reviewers to satisfaction. I have made few, mostly grammatical suggestions here:

We thank the reviewer’s comments. I revised the manuscript as suggested.

L5 typo in the short title

We revised from “Micro-sizes” to “Micro-size” (L5).

L18-20 Edit first sentence to: “While primary productivity in the oligotrophic North Pacific Subtropical Gyre (NPSG) is changing, the micro-size plankton community has not been evaluated in the last 3 decades, prompting a re-evaluation.”

We revised as suggested (L18-20).

L23 Don’t need s in community structures.

We revised from “community structures” to “community structure” (L22) as well as L556.

L24 “The assemblages were consistent with those identified 4 decades previously.” Also 4 decades? Thought the first sentence indicated 3 decades.

We revised from “3 decades” to “4 decades” of the first sentence (L19).

L25 Dinophyceae were “the most numerically abundant, followed by Cryptophyceae and Bacillariophyceae (diatoms)”.

We revised as suggested (L25).

L42 Therefore, nitrogen fixation “may contribute” to the heterogeneity …

We revised as suggested (L41).

L49-50 In other words, energy fixed by primary production is more efficiently transferred to higher trophic organisms in micro-plankton dominated waters.

We revised as suggested (L48-50).

L61 is instead of as

L62 phytoplankton are “the dominant primary producers”;

We revised “is the dominant primary producers” (L61-62).

L64 abundant but is it significant? What % of cell counts or Chl a?

We added “(> 108 cells m-2) in the surface mixed layer” with a new reference Scharek et al. 1999 DSRI (L64).

L69 western NPSG is oligotrophic (no ‘the’)

We revised as suggested (L69).

L70 migratory fish?

We revised as suggested (L70).

L88-89 To understand the trophic structure and heterogeneity of this community.

We revised as suggested (L88-89).

L93 three cruises on the R/V …

We revised as suggested (L93).

L98-99 therefore, the observations for collections of seawater were limited to night-time collections.

We revised as suggested (L98-99).

Table 1 is very informative and much improved, thank you!

We also thank the comments.

L121 – collected using … (no “by”)

We revised as suggested (L121).

L244 Nitrate concentration was over 1 uM of in water “where” the temperature was below

We revised as suggested (L245).

L250 …layers as indicated “by” Falokowski …

We revised as suggested (L250).

L472 …primary production in the wester NPSG depicted slight increase.

We revised as suggested (L472).

L480 The metabarcoding results were consistent “with” our microscopic observation, with the exception of Dinophyceae which comprised a larger proportion of the community in metabarcoding results. This is likely due to Dinophyceae having largernuclear genome sizes compared to other microplankton.

We revised as suggested (L481-483).

L484-490 This observation suggests that while microscopic observation is a time consuming technique, it is still a necessary tool to understand the micro…..; furthermore, this indicates that the metabarcoding technique should best be coupled with more traditional methods such as microscopic observations.

We revised as suggested (L485-489).

Attachment

Submitted filename: Response_letter_PONE-D-20-37689-R2.docx

Decision Letter 2

Arga Chandrashekar Anil

12 Apr 2021

Micro-size plankton abundance and assemblages in the western North Pacific Subtropical Gyre under microscopic observation

PONE-D-20-37689R2

Dear Dr. Kodama,

We’re pleased to inform you that your manuscript has been judged scientifically suitable for publication and will be formally accepted for publication once it meets all outstanding technical requirements.

Within one week, you’ll receive an e-mail detailing the required amendments. When these have been addressed, you’ll receive a formal acceptance letter and your manuscript will be scheduled for publication.

An invoice for payment will follow shortly after the formal acceptance. To ensure an efficient process, please log into Editorial Manager at http://www.editorialmanager.com/pone/, click the 'Update My Information' link at the top of the page, and double check that your user information is up-to-date. If you have any billing related questions, please contact our Author Billing department directly at authorbilling@plos.org.

If your institution or institutions have a press office, please notify them about your upcoming paper to help maximize its impact. If they’ll be preparing press materials, please inform our press team as soon as possible -- no later than 48 hours after receiving the formal acceptance. Your manuscript will remain under strict press embargo until 2 pm Eastern Time on the date of publication. For more information, please contact onepress@plos.org.

Kind regards,

Arga Chandrashekar Anil, Ph. D., D. Agr.,

Academic Editor

PLOS ONE

Additional Editor Comments (optional):

Reviewers' comments:

Acceptance letter

Arga Chandrashekar Anil

15 Apr 2021

PONE-D-20-37689R2

Micro-size plankton abundance and assemblages in the western North Pacific Subtropical Gyre under microscopic observation

Dear Dr. Kodama:

I'm pleased to inform you that your manuscript has been deemed suitable for publication in PLOS ONE. Congratulations! Your manuscript is now with our production department.

If your institution or institutions have a press office, please let them know about your upcoming paper now to help maximize its impact. If they'll be preparing press materials, please inform our press team within the next 48 hours. Your manuscript will remain under strict press embargo until 2 pm Eastern Time on the date of publication. For more information please contact onepress@plos.org.

If we can help with anything else, please email us at plosone@plos.org.

Thank you for submitting your work to PLOS ONE and supporting open access.

Kind regards,

PLOS ONE Editorial Office Staff

on behalf of

Professor Arga Chandrashekar Anil

Academic Editor

PLOS ONE

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    Attachment

    Submitted filename: PONE_D_20_37689_Review.docx

    Attachment

    Submitted filename: PONE_D_20_37689_response_letter.docx

    Attachment

    Submitted filename: Response_letter_PONE-D-20-37689-R2.docx

    Data Availability Statement

    All data are available from Mendeley Data (http://dx.doi.org/10.17632/hmy5jzccyx.2).


    Articles from PLoS ONE are provided here courtesy of PLOS

    RESOURCES