small-scale variability of protistan planktonic

15
MARINE ECOLOGY PROGRESS SERIES Mar Ecol Prog Ser Vol. 548: 61–75, 2016 doi: 10.3354/meps11647 Published April 21 INTRODUCTION A central question in marine ecology is how envi- ronmental variation affects microbial community dy- namics. Because of the short generation times and rapid metabolic responses of aquatic microbes, pe- lagic food webs are considered to have maximum re- silience to environmental change (Laws 2003). Next- generation sequencing (NGS) methods provide a comprehensive assessment of marine microbial di- versity, including the detection of rare and ‘invisible’ taxa (e.g. Medinger et al. 2010, Behnke et al. 2011). Recent NGS studies have examined the statistical re- lationships among the dynamics of marine microbes, and the influence of environmental parameters meas- ured over various spatial and temporal scales. In doing so, they have determined the environmental conditions that influence individual microorganisms, and highlighted the significance of their biotic inter- actions. For example, existing studies on bacterial communities have indicated that environmental para- meters (such as physical mixing, day length, temper- © The authors 2016. Open Access under Creative Commons by Attribution Licence. Use, distribution and reproduction are un- restricted. Authors and original publication must be credited. Publisher: Inter-Research · www.int-res.com *Corresponding author: [email protected] Small-scale variability of protistan planktonic communities relative to environmental pressures and biotic interactions at two adjacent coastal stations Savvas Genitsaris, Sébastien Monchy, Elsa Breton, Eric Lecuyer, Urania Christaki* Laboratoire d’Océanologie et Géosciences (LOG), UMR CNRS 8187, Université du Littoral Côte d’Opale (ULCO), 32 av. Foch, 62930 Wimereux, France ABSTRACT: The aim of this study was to analyze planktonic protistan assemblages under differ- ent environmental pressures at 2 adjacent coastal stations (inshore and offshore; ~6.4 km apart) in the eastern English Channel (EEC). Samples were collected between March 2012 and June 2013, and analyzed using bioinformatic analysis and tag pyrosequencing of the V2-V3 hypervariable region of the 18S rRNA gene. In addition to the taxonomic composition of the protistan communi- ties, the detected operational taxonomic units (OTUs) were sorted into 6 major functional groups based on their trophic roles in marine systems. Comparisons of 13 environmental factors, includ- ing physical and chemical variables, indicated significant differences between the 2 stations. However, the only significant relationship between environmental pressures and protistan com- munity structure was detected when the trophic role of the OTUs was considered at the inshore station. In order to interpret differences in community structure, the effect of biotic interactions was investigated through the examination of the co-occurrence networks of the protistan commu- nities at both stations. In terms of the number of edges and connectivity of nodes, the analysis showed that the inshore station had more complex associations between OTUs than the offshore station. This strongly suggests that due to the higher and more variable environmental pressures the inshore station comparatively receives, its protistan community has developed a greater com- plexity of biotic connections. This in turn reflects the rapid responses of trophic interactions within the entire microbial community. KEY WORDS: Unicellular eukaryotes · Community organization · Environmental parameters · Biotic interactions · 18S rRNA · Pyrosequencing · Bioinformatics · Ecological tools OPEN PEN ACCESS CCESS

Upload: others

Post on 13-Jan-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Small-scale variability of protistan planktonic

MARINE ECOLOGY PROGRESS SERIESMar Ecol Prog Ser

Vol. 548: 61–75, 2016doi: 10.3354/meps11647

Published April 21

INTRODUCTION

A central question in marine ecology is how envi-ronmental variation affects microbial community dy-namics. Because of the short generation times andrapid metabolic responses of aquatic microbes, pe -lagic food webs are considered to have maximum re-silience to environmental change (Laws 2003). Next-generation sequencing (NGS) methods provide acomprehensive assessment of marine microbial di-versity, including the detection of rare and ‘invisible’

taxa (e.g. Medinger et al. 2010, Behnke et al. 2011).Recent NGS studies have examined the statistical re-lationships among the dynamics of marine microbes,and the influence of environmental parameters meas-ured over various spatial and temporal scales. In doing so, they have determined the environmentalconditions that influence individual microorganisms,and highlighted the significance of their biotic inter -actions. For example, existing studies on bacterialcommunities have indicated that environmental para -meters (such as physical mixing, day length, temper-

© The authors 2016. Open Access under Creative Commons byAttribution Licence. Use, distribution and reproduction are un -restricted. Authors and original publication must be credited.

Publisher: Inter-Research · www.int-res.com

*Corresponding author: [email protected]

Small-scale variability of protistan planktonic communities relative to environmental pressures and

biotic interactions at two adjacent coastal stations

Savvas Genitsaris, Sébastien Monchy, Elsa Breton, Eric Lecuyer, Urania Christaki*

Laboratoire d’Océanologie et Géosciences (LOG), UMR CNRS 8187, Université du Littoral Côte d’Opale (ULCO), 32 av. Foch, 62930 Wimereux, France

ABSTRACT: The aim of this study was to analyze planktonic protistan assemblages under differ-ent environmental pressures at 2 adjacent coastal stations (inshore and offshore; ~6.4 km apart) inthe eastern English Channel (EEC). Samples were collected between March 2012 and June 2013,and analyzed using bioinformatic analysis and tag pyrosequencing of the V2−V3 hypervariableregion of the 18S rRNA gene. In addition to the taxonomic composition of the protistan communi-ties, the detected operational taxonomic units (OTUs) were sorted into 6 major functional groupsbased on their trophic roles in marine systems. Comparisons of 13 environmental factors, includ-ing physical and chemical variables, indicated significant differences between the 2 stations.However, the only significant relationship between environmental pressures and protistan com-munity structure was detected when the trophic role of the OTUs was considered at the inshorestation. In order to interpret differences in community structure, the effect of biotic inter actionswas investigated through the examination of the co-occurrence networks of the protistan commu-nities at both stations. In terms of the number of edges and connectivity of nodes, the analysisshowed that the inshore station had more complex associations between OTUs than the offshorestation. This strongly suggests that due to the higher and more variable environmental pressuresthe inshore station comparatively receives, its protistan community has developed a greater com-plexity of biotic connections. This in turn reflects the rapid responses of trophic interactions withinthe entire microbial community.

KEY WORDS: Unicellular eukaryotes · Community organization · Environmental parameters ·Biotic interactions · 18S rRNA · Pyrosequencing · Bioinformatics · Ecological tools

OPENPEN ACCESSCCESS

Page 2: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 201662

ature, and nutrients) were more important in shapingcommunity structure than taxa interrelationships(e.g. Gilbert et al. 2012, Salter et al. 2015). Other stud-ies have suggested that the biotic interactions amongco-occurring taxa have been the main structural drivers of spatial and temporal assemblages (e.g.Fuhrman & Steele 2008, Shade et al. 2014, Székely &Langenheder 2014). In bacterial communities, func-tional similarity among taxonomically distinct groupsis common (e.g. Burke et al. 2011). Alternatively, ineukaryotes, morphological and phylogenetic related-ness does not necessarily correspond to ecological re-latedness (e.g. reviewed in Sherr et al. 2007, Caron& Countway 2009, Caron et al. 2012). For in stance,within the common high-level taxonomic group Alveolata, Dinophyceae are diatom grazers (e.g.Sherr & Sherr 1994, Grattepanche et al. 2011a, b),and marine alveolates (MALV) are most likely intra-cellular symbionts or parasites (e.g. Skovgaard etal. 2005, Harada et al. 2007, Massana & Pedrós-Alió2008); within the marine stra meno piles (MASTs),Bacillariophyceae (diatoms) are known autotrophs,and MASTs have been identified as free-living bac-terivorous heterotrophs (e.g. Massana et al. 2006);fungi are involved in the degradation of organic matter (e.g. Raghukumar 2004); and cercozoans aresuggested to exhibit mainly parasitic behavior (e.g.Tillmann et al. 1999, Schnepf & Kühn 2000).

The novelty of the present study was in consideringthe trophic roles of planktonic eukaryotes detectedby NGS in addition to their taxonomic position. Themain questions we investigated were: (1) Is there anincreased significance of inter-taxa relationships incounterbalancing environmental variability andshaping protistan community structure, and (2) Areeukaryotic functional roles (such as trophic traits)more informative than taxonomic affiliations in theinterpretation of biotic/abiotic interactions? Thesequestions were tested with a small-scale approach ina seasonal study (15 samples over a 15 mo period) atclosely located stations in the eastern English Chan-nel (EEC). In this area, physical and chemical vari-ables (winds, currents, tides, and nutrient concentra-tions) exhibit strong contrasts. The 2 coastal stations,hereafter referred to as ‘inshore’ and ‘offshore’, wereexpected to show differences in the prevailing envi-ronmental forcings, providing a framework withwhich to test plankton communities’ responses toenvironmental changes. The detected taxa (deter-mined by tag pyrosequencing) and environmentalparameters were analyzed by examining the co-occurrence networks of the protistan communities ateach station.

MATERIALS AND METHODS

Sample collection

Sampling was carried out at 2 stations located6.4 km apart in the EEC SOMLIT (French Networkof Coastal Observatories). The inshore station (50°40.75’ N, 1° 31.17’ E, 27 m max. depth) was 1.6 kmfrom the coast, and the offshore station (50° 40.75’ N,1° 24.60’ E, 56 m max. depth) was 8 km from the coast(Fig. 1). A total of 30 sub-surface samples (2 to 3 mwater depth) were collected (15 from each site) in 2.5 lsterile polyethylene bottles twice a month during hightide, from 20 March 2012 to 25 June 2013 (16 and 14for each year, respectively), when local weather con-ditions permitted. After collection, samples were keptin the dark at in situ temperature, and filtered within2 h. Before filtering, samples were screened through a150 µm mesh to retain larger particles and most meta-zoa. Next, sequential filtration through 10, 3, and0.6 µm nucleopore filters (47 mm diameter) was per-formed using a low filtration pressure peristaltic pump

Fig. 1. Study area in the eastern English Channel, indicatingmajor estuaries (Somme, Authie and Canche) and the location

of the sampling stations (*)

Page 3: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities

(15 rpm) in order to avoid filter clumping and mini-mize organism disruption. The filters were immedi-ately stored at −80°C until DNA extraction.

Environmental variables

Seawater temperature (T, °C), salinity (S), andphoto synthetically active radiation (PAR) were meas-ured in situ with a CTD (Seabird SBE 19) equippedwith a biospherical PAR light sensor (QSP 2300, Bios-pherical Instruments). The diffuse attenuation coeffi-cient for down-welling irradiance (kd, m−1) was as -sessed from instantaneous vertical CTD profiles. Theaverage sub-surface daily light intensity (I, E m−2 d−1)experienced by phytoplankton (over the last 6 d priorto sampling) was estimated using the formula basedon Riley (1957):

(1)

where z is the depth at which samples were collected(2 m) and I0 is the daily incident light estimated fromglobal solar radiation (GSR, W m−2) measured contin-uously every 5 min with a solar radiation sensor(Vantage Pro, Davis) mounted on the laboratory roof bordering the seashore, in close proximity to thesampling area. Prior to calculating the coefficient,GSR was converted to PAR under the assumptionthat PAR = 50% of GSR, and 1 W m−2 = 0.36 E m−2 d−1

(Morel & Smith 1974). The level of dissolved oxygen(O2, mg l−1) was analyzed in triplicate by Winklermicro-titration with a Titrino848 (Methrom) accord-ing to the methodologies outlined in Aminot &Kérouel (2004). pH was measured with a pH1970i(WTW) pH meter. Nitrate (NO3

–), nitrite (NO2–), phos-

phate (PO43–), and silicate (SiO4

4–) concentrations(µM) were determined from 100 ml samples with anAlliance Integral Futura Autoanalyzer II based onAminot & Kérouel (2004). Ammonium (NH4

+, µM)was determined manually by fluorometry in accor-dance with Holmes et al. (1999), modified by Tayloret al. (2007). Chlorophyll a concentrations (chl a, µgl−1) were measured using a 10-AU Turner Designs®

fluorometer in accordance with Lorenzen (1967),after extraction of pigment retained on WhatmanGF/F glass fiber filters in 90% acetone for 12 h at4°C. Particulate organic carbon (POC, mg l−1) wasassessed with a NA2100 Frisons CHN analyzer afterthe GF/F filters were dried at 60°C for 24 h andexposed to vapors of 1 N HCl for 5 h. Suspended par-ticulate matter (SPM, mg l−1) was estimated byweighing the particulate matter collected by filtra -

tion on a GF/F filter. Additional details on environ-mental data acquisition and sample ana lysis can befound at http://somlit. epoc. u-bordeaux1. fr/fr/.

DNA extraction

After collectively pooling the 10, 3, and 0.6 µm fil-ters, the DNA of the planktonic organisms wasextracted and purified using a PowerWater® DNAisolation kit (Mobio Laboratories), following the man-ufacturer’s protocol. The samples contained between0.5 and 6 ng µl−1 of DNA as measured by the Qubit®2.0 fluorometer (Thermo Fischer Scientific).

PCR and tag pyrosequencing

DNA samples were amplified using the 2 eukary-otic primers 18S-82F (5’-GAA ACT GCG AAT GGCTC-3’) (López-García et al. 2003) and Euk-516r (5’-ACC AGA CTT GCC CTC C-3’) (Amann et al. 1990).These primers have been used successfully in previ-ous studies of the protistan community at this site(Monchy et al. 2012, Christaki et al. 2014, Genitsariset al. 2015); thus, they were selected in order to construct a dataset comparable with those studies.Polymerase chain reaction (PCR) was carried outaccording to standard conditions for Platinum TaqHigh-Fidelity DNA Polymerase (Invitrogen), with5 ng of environmental DNA as a template, using theGeneAmp PCR System Apparatus (Applied Biosys-tems). Tag pyrosequencing was carried out by Geno-Screen. The library was prepared following the pro-cedures described by Roche Diagnostics, with all 30samples sequenced on 1 plate run (15 samples per ½plate run) on a 454 GS FLX Titanium Sequencer, withthe long-read chemistry of Roche Applied Sciencessupplied for the GS-FLX system. Pyrosequenceswere submitted to GenBank-SRA under accessionnumber SRX768577.

Tag pyrosequencing quality filtering

Pyroreads were subjected to quality filtering usingmothur v.1.28.0, following standard operating proce-dures (Schloss et al. 2009, 2011). Flowgrams fromeach sample were extracted, separated according totheir tag, and de-noised using the mothur implemen-tation of ‘PyroNoise’ (Quince et al. 2009). The datasetwas de-replicated to the unique reads and alignedagainst the SILVA 108 database. Reads suspected

10

d

d

II e

k z

k z( )= − −

63

Page 4: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 2016

of being chimeras were removed using UCHIMEv.4.2.40 (Edgar 2010). Finally, the dataset was nor-malized to the sample with the lowest number ofreads, so that all samples contained 12 581 reads. Thereads were clustered into operational taxonomicunits (OTUs) at 97% similarity threshold, using theaverage neighbor method in mothur. Unique ampli-cons that occurred only once in the dataset wereremoved, as these were most likely erroneous se -quencing products (Kunin et al. 2010, Behnke et al.2011). For more details of PCR and tag pyrosequenc-ing procedures, see Genitsaris et al. (2015).

After tag pyrosequencing filtering and normaliza-tion, 1303 OTUs were produced. Taxonomic classifi-cation was assigned using BLASTN (Nucleotide−Nucleotide Basic Local Alignment Search Tool; Alt -schul et al. 1990) based on the Protist Ribosomal Ref-erence (PR2) curated database (built on GenBank203; October 2014) containing 23 003 sequences(Guillou et al. 2013). All reads affiliated with metazoawere removed from the dataset; accordingly, theremaining 1174 OTUs were protists.

Data analysis

The 1174 OTUs belonging to 9 taxonomic super-groups were sorted into 6 groups according to theirtrophic status (Table 1, Table S1 in the Supplement atwww. int-res. com/ articles/ suppl/ m548 p061_ supp.pdf). In eukaryotes, morphologically and phylogene -

tically similar taxa can exhibit many different typesof meta bolism. For example, dinoflagellates includegrazers, autotrophs, mixotrophs, and parasites.There fore, the strategy applied here was to individu-ally examine the 1174 OTUs and annotate them to atrophic group using the highest level of informationcurrently available. For many OTUs affiliated withmicroplankton and nanoplankton organisms, prima-rily dinoflagellates (e.g. Gyrodinium, Gymnodinium,and Karlo di ni um), ciliates (e.g. Strombidium), dia -toms (e.g. Lepto cyli d rus, Rhizosolenia), cryptophytes,coccolithophorids, chlorophytes, and the bloomingPhaeocystis globosa, confidence about their validaffiliation in the dataset was high. This was becausethey had been previously observed by microscopyand/ or flow cytometry in the same area (e.g. Grattep-anche et al. 2011a,b, Monchy et al. 2012, Bonato et al.2015, 2016). Conversely, for OTUs that were affili-ated with taxonomic groups impossible to detect withmicroscopy, consideration of their annotation tohigher taxonomic groups (e.g. family level) wastaken into account, which is considered reliable atthe OTU level (Bachy et al. 2013, Santoferrara et al.2014). For example, most taxa belonging to thegroups MALV and MAST are considered symbiontsand nano-grazers, respectively, according to theexisting literature (Skovgaard 2014, Massana et al.2006, respectively for MALV and MAST) (seeTable S1 in the Supplement)

Alpha-diversity estimators (the richness estimatorSChao1, and the Simpson, Equitability, and Berger-

64

Functional groups Trophic role Example of representative taxonomic groups

Autotrophs (auto) Primary producers: fix CO2 to produce Bacillariophyta (diatoms) organic material using light energy Chlorophyta

Mixotrophs (mixo) Either autotroph or heterotroph that can Cryptophyta, Haptophyta use mix of sources of carbon and energya Ciliates, dinoflagellates

Picoplankton grazers (picoG)b Graze on heterotrophic bacteria and other Apusozoa, Discoba picoplankton MAST, Telonemia, Choanoflagellida

Nano- and microplankton grazers Graze on nano- and microplankton Ciliates, dinoflagellates(nanoG, microG)b (e.g. on nanoflagellates, diatoms)

Symbionts/parasites/decomposers Infect hosts, decompose and Fungi, MALV, Labyrinthulea, (symbdec) remineralize organic matter Perkinsea, Omycetes

aMixotrophs can be autotrophs, which supplement photosynthesis by bacterivory and/or dissolved organic matterassimilation, or heterotrophs that use the chloroplasts of their ingested prey for photosynthesis, or have symbiotic algae(e.g. mixotrophic ciliates). A perfect mixotroph would use organic and inorganic carbon equally well. For more details,see Table S1 in the Supplement

bPicoG, nanoG, and microG are heterotrophic grazers

Table 1. Functional groups presented in this study based on their trophic role and prey grazing size. For more details on repre-sentative taxonomic groups, see Table S1 in the Supplement at www.int-res.com/articles/suppl/m548p061_supp.pdf

Page 5: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities 65

Parker indices) were calculated for all samples.These diversity estimators were compared for eachdate with a bootstrapping randomization procedure,and the p-value was computed based on 1000 ran-dom permutation pairs. The coefficients of variation(CVs) of the 13 environmental parameters measuredwere tested for each date with the Fligner-Killeentest statistic. The mean ranks from both stations werealso compared using the Wilcoxon test. The abovetests were calculated using PAST v.3c software(Hammer et al. 2001).

We used co-inertia analysis (COIA) to investigatethe coupling between the 2 stations in terms of envi-ronmental variables and plankton assemblages, andalso between the environmental variables and plank-ton assemblages at each station (Dolédec & Chessel1994, Dray et al. 2003). The principle of COIA con-sists of finding co-inertia axes by maximizing thecovariance between the row coordinates of 2 matri-ces. It defines axes that simultaneously explain thehighest possible variance in each of the 2 matricesand describes their closest possible common struc-ture. A PCA (principal component analysis) was per-formed on each matrix, and then the COIA analysiswas applied (Dray et al. 2003, see their Fig. 1). Toestimate the strength of the coupling between the 2matrices, a multidimensional correlation coefficient(RV) was calculated, and statistical significance wastested using a Monte Carlo permutation procedurewith 1000 permutations. COIA was carried out withthe ‘ade4’ package in R (Dray & Dufour 2007).Finally, in order to define the variables that were themost important in structuring the COIA scatterplot,Spearman’s correlation coefficients were calculatedbetween all variables and COIA coordinates. Envi-ronmental and abundance data were log(x + 1) trans-formed to ensure normal distribution (Legendre &Legendre 1998), and environmental data were stan-dardized to render them dimensionless (z-scores;Kenkel 2006).

For the network analysis, 2 matrices correspondingto offshore (881 OTUs) and inshore (919 OTUs) readnumbers were integrated into a ‘total’ matrix con-taining 1800 rows (the sum of the offshore andinshore OTUs). In total, 626 OTUs were commonbetween the 2 stations; thus all OTUs received a dis-tinct coding for each station (e.g. oOTU0001 for off-shore, and iOTU0001 for inshore). Consequently, n ×(n − 1) / 2 pairings were analyzed; where n is the totalnumber of OTUs in the analysis matrix (i.e. 1800OTUs). Markedly, 626 pairings corresponded to thecommon OTUs. The relationship between OTUs wascharacterized through maximal information-based

nonparametric exploration (MINE) statistics by com-puting the maximal information coefficient (MIC)between each pair (Reshef et al. 2011). MIC capturesassociations between data and provides a score thatrepresents the strength of the relationship betweendata pairs. OTU pairings where MIC > 0.8 were usedto visualize the co-occurrence relationships betweenOTUs within and between the different stations. The0.8 MIC threshold was selected to depict the strongestOTU pairings (see Reshef et al. 2011); this was con-sidered as the best compromise in order to clearlyvisualize the most representative co-occurrence pat-terns at different trophic levels during the studyperiod. The network of MIC associations of OTUswas constructed and respective topological parame-ters were calculated using Cytoscape v.3.0 (Smoot etal. 2011).

RESULTS

Community composition and seasonal succession

A total of 898296 raw reads were produced fromthe sequencing of all samples. After quality check-ing, de-noising, and chimera removal, 591 307 readsremained. Sub-sampling and removal of metazoanreads resulted in a total of 162 705 reads at the off-shore station and 160 912 at the inshore station.Results in all samples identified 1174 unique OTUsaffiliated with protists at the inshore (919 OTUs) andoffshore (881 OTUs) stations. Among these, 626 OTUswere observed at both stations, while 293 and 255were only found at the inshore or offshore station,respectively. The ratio of observed to expected OTUs(SChao1) was >75% in all cases, and >90% in themajority of samples (Table 2). The 1174 OTUs wereaffiliated with 9 supergroups (Amoebozoa, Apuso -zoa, Alve olata, Archeaplastida, Excavata, Hacro bia,Opis tho konta, Rhizaria, and Strameno piles) and mis-cellaneous other protists (Fig. 2), which were furtherdivided to 34 taxonomic groups (Table S1 in the Supplement at www. int-res. com/ articles/ suppl/ m548p061_ supp. pdf). The relative OTU richness of thesupergroups was similar at both stations over thesampling period (Fig. 2). The supergroup Alveolatahad the highest OTU richness at both sites (341 OTUsat the inshore and 325 OTUs at the offshore station)followed by Stramenopiles (228 and 219 OTUs,respectively) (Fig. 2).

The OTUs were further sorted into 6 functional/trophic groups (Table 1, for more details see Table S1in the Supplement): autotrophs (auto), mixotrophs

Page 6: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 2016

(mixo), picoplankton grazers (picoG),nanoplankton grazers (nanoG), micro-plankton grazers (microG), and a groupmainly represented by parasites and/orsymbionts. However, since parasitismdefines a very specific mechanism ofsymbiosis, parasitic groups such asMALV, Apicomplexa, and Perkinsea canbe also considered as symbionts (TableS1). With the exception of 1 OTU belong-ing Hypho chy trio my ceta, the other OTU-rich group was the fungi, which includedparasites and decomposers (56 and 76OTUs offshore and inshore, respectively;Table S1). Historically, marine fungi havebeen poorly investigated, and the OTUaffiliation here contained a low percent-age, indicating no definite phylogeneticdistance be tween the different fungigroups. Accordingly, although there wasno binding connection between parasiticand de composer fungi, they were allo-cated here into a symbiont/decomposergroup, hereafter referred to as ‘symbdec’(see Table 1).

In general, symbdec was the predomi-nant group in most samples in terms ofOTU richness; reaching 124 OTUs in thein shore sample collected on 3 Oct 2012(Fig. 3a) and contributing >20% of thetotal number of OTUs in 28 of the 30 sam-ples (Fig. 3b). This was followed bypicoG, which also had the highest OTUrichness (64 OTUs) in the 3 Oct 2012 sam-ple (Fig. 3a). The least dominant group

66

Sampling date No. SChao1 Simpson Equitability Berger-Parker (station) OTUs (D) (H/Hmax)

20/03/2012 (O) 124 136 0.24 0.54 0.4720/03/2012 (I) 58 63 0.08 0.77 0.1609/05/2012 (O) 122 126 0.09 0.67 0.2109/05/2012 (I) 172 203 0.07 0.65 0.1905/06/2012 (O) 148 150 0.11 0.62 0.2705/06/2012 (I) 85 88 0.13 0.59 0.2521/06/2012 (O) 230 246 0.06 0.69 0.1421/06/2012 (I) 196 203 0.07 0.65 0.1904/07/2012 (O) 96 102 0.25 0.57 0.4904/07/2012 (I) 101 119 0.13 0.71 0.3403/09/2012 (Ο) 184 192 0.12 0.60 0.2503/09/2012 (I) 300 313 0.02 0.80 0.0603/10/2012 (O) 298 311 0.03 0.78 0.1103/10/2012 (I) 58 63 0.08 0.77 0.1613/11/2012 (O) 146 148 0.05 0.77 0.1813/11/2012 (I) 317 326 0.03 0.77 0.1011/02/2013 (O) 199 207 0.05 0.72 0.1611/02/2013 (I) 205 211 0.04 0.73 0.1326/02/2013 (O) 101 119 0.13 0.71 0.3426/02/2013 (I) 80 86 0.19 0.60 0.4126/03/2013 (O) 164 167 0.14 0.58 0.3326/03/2013 (I) 165 175 0.40 0.39 0.6208/04/2013 (O) 158 175 0.09 0.63 0.2108/04/2013 (I) 59 60 0.37 0.46 0.5927/05/2013 (O) 296 335 0.05 0.71 0.2027/05/2013 (I) 183 190 0.13 0.57 0.2310/06/2013 (O) 230 260 0.06 0.66 0.1510/06/2013 (I) 138 146 0.27 0.46 0.4925/06/2013 (O) 128 169 0.74 0.18 0.8625/06/2013 (I) 205 231 0.36 0.43 0.59

Table 2. Number of operational taxonomic units (OTUs), richness estima-tor (SChao1) and the heterogeneity of the diversity indexes (Simpson’s D,Equit ability [H: Shannon diversity index] and Berger-Parker) at the SOM-LIT offshore (O) and inshore (I) stations in the eastern English Channel

from March 2012 to June 2013 (dates are given as dd/mm/yyyy)

Amoebozoa (3) Apusozoa (5)

Alveolata (325)

Archeaplastida (41)

Excavata (14) Hacrobia (75)

Opisthokonta (97)

Rhizaria (74)

Stramenopiles(219)

Other Protists (26)

Amoebozoa (5) Apusozoa (12)

Alveolata (341)

Archeaplastida (42)Excavata (8)

Hacrobia (75)

Opisthokonta (113)

Rhizaria (67)

Stramenopiles(228)

Other Protists (32)

Offshore Inshore

Fig. 2. Number of operational taxonomic units (OTUs) in the 9 protistan supergroups and other protists detected at the in-shore and offshore sampling stations based on comparisons against the Protist Ribosomal Reference (PR2) database using the

‘blastn’ function (see also Table 1, Table S1 in the Supplement at www. int-res. com/ articles/suppl/ m548p061_ supp. pdf)

Page 7: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities

was nanoG, which contributed ~10% of the totalnumber of OTUs in most samples (Fig. 3b), and in anextreme case (4 July 2012) were represented by only1 OTU (Fig. 3a).

The highest number of OTUs in the offshore stationwas observed on 3 Oct 2012 (298 OTUs) and at theinshore station on 3 Sep 2012 (300 OTUs). The lowestOTUs richness was observed at the offshore stationon 4 July 2012 (96 OTUs) and at the inshore stationon 20 Mar 2012 and 3 Oct 2012 (both 58 OTUs)(Table 2). In certain cases, a large variation in OTUrichness was observed between the 2 stations (themost ex treme example occurred on 3 Oct 2012, whenthe difference in OTU richness was 240 OTUs). TheSimpson index (D) was highest in the samples with

the lowest richness, and vice versa. Fluctuationsranged from 0.02 on 3 Sep 2012 at the inshore station(the sample with the highest species richness in theentire dataset) to 0.74 on 25 Jun 2013 at the offshorestation (Table 2). This was when the protistan com-munity was dominated by a single OTU affiliatedwith Phaeocystis globosa (OTU0005), which com-prised 86% of the total number of reads (Berger-Parker index = 0.86.). The Berger-Parker indexshowed large variations, ranging from 0.06 on 3 Sep2012 at the in shore station to 0.86 on 25 Jun 2013 atthe offshore station. These values were the same asthose of the Simpson index, while the Shannon Equi-tability index fluctuated from 0.18 to 0.80, withhigher values reflecting low variation between spe-cies abundances within the community. A compari-son of these diversity estimators indicated significantstatistical differences among all indices for all datesbetween the 2 stations (p < 0.05), except for the Equi-tability index on 13 Nov 2012 and 11 Dec 2012 (p =0.17 and 0.11, respectively).

Environmental parameters and community structure

Overall, higher mean values of kd, pH, NO3–+NO2

–,POC, SPM, and chl a, and lower values of S, SiO4

4–

and PO43– were found at the inshore relative to the

offshore station (Fig. 4, Table S2). These higher val-ues were reflected in significant differences of themean ranks (Wilcoxon test) for kd, S, pH, NO3

–+NO2–,

POC, and SPM. The comparison of CVs of the13 environmental variables showed significant dif-ferences for kd, S, and SiO4

4– (Table 3). When COIAwas applied to test the seasonal evolution of the envi-ronmental variables of the 2 stations, it showed a sig-nificant environmental coupling between offshoreand inshore stations over the study period (RV = 0.80,p = 0.001).

COIA analysis indicated that neither taxonomic(RV = 0.33, p = 0.36), nor trophic groups (RV = 0.27,p = 0.22), displayed any significant coupling betweenthe 2 stations over the sampling period (Table 4). Fur-thermore, no significant relationship between theenvironmental variables and the 9 taxonomic super-groups was found at either station (p = 0.067 and0.075 for offshore and inshore, respectively; Table 4).The only significant coupling between environmen-tal variables and trophic groups was found for theinshore station (RV = 0.41, p = 0.033; Fig. 5). Accord-ingly, the first 2 axes explained 89% of the total vari-ance (Table 4). The environmental variables kd, PAR,

67

Fig. 3. Absolute and relative number of operational taxo-nomic units (OTUs) in the trophic groups detected on eachsampling date at the 2 sampling stations (inshore [I] andoffshore [O]). The trophic groups were formed based onthe trophic status of the detected OTUs in marine systemsas inferred from the literature (see Table 1, Table S1 in the

Supplement). Dates are dd/mm

Page 8: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 2016

T, O2, NH4+, NO3

–+NO2–, PO4

3–, and SPM showed thestrongest correlations with the 2 COIA axes (rS; p <0.005, and p < 0.001 for SPM; Table S3 in the Supple-ment). Among the trophic groups, microG andnanoG showed the strongest correlations with theCOIA co-ordinates (rS; p < 0.001 and 0.005 respec-tively; Table S3), followed by mixo, auto, and picoG(rS; p < 0.01 and 0.05 respectively; Table S3). The par-asites were not correlated with any of the axes. Thescatterplot of the COIA analysis also indicated howfar apart the stations were relative to their explana-

tory and response variables, and showed a relativelyclear seasonal pattern with the autumn−winter andsummer−spring stations grouped together (Fig. 5).The trophic group microG was associated withwaters exhibiting relatively low nutrients but highPAR and high salinity, while the inverse was true forthe auto group. The trophic groups mixo, picoG, andnanoG were all associated with relatively warm (highT values) and clear (high PAR but low kd values)waters, but that were relatively poor in O2, chl a, andSPM (Fig. 5).

68

In Off

Chl

a (

μg l–1

) Si

O44–

(μM

) N

H4+ (

μM)

S k d

(m–1

) 0.7

0.6 0.5

0.4 0.3

0.2 0.1

35.5

35

34.5

34

33.5

33

5

4

3

2

1

0

8

6

4

2

0

POC

(mg

l–1)

NO

3– + N

O2–

(μM

) O

2 (m

g l–1

) PA

R (E

m–2

d–1

)

50

40

30

20

10 5

8.5

7.5

6.5

5.5 5

20

16

12

8

4

0

1.4

1

0.6

0.2 0

SPM

(mg

l–1)

PO43–

(μM

) pH

T

(C°)

20

16

12

8

4

8.5

8.4

8.3

8.2

8.18.05

1.2

1 0.8 0.6

0.4

0.2 0

8

6

4

2

0

In Off In Off 16

12

8

4

0

Fig. 4. Environmental parameters examined in this study to determineprotistan community composition at 2 sampling stations (in: inshore;off: offshore). kd: light attenuation; PAR: photosynthetically active radi-ation; T: temperature; S: salinity; POC: particulate organic carbon;SPM: suspended particulate matter. Red cross: mean; mid-line: median;box limits: first and third quartiles; whiskers: 1.5× the interquartilerange; open circles: outliers; blue points: minimum and maximum for

each parameter

Page 9: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities

Biotic interactions

The next step was to investigate the degree ofbiotic interactions based on OTU co-occurrence pat-terns by network analysis according to the MIC val-ues. The network of OTU associations for both sta-tions based on MIC > 0.8 is shown in Fig. 6. TheOTUs were presented according to their trophic role(see Table 1, Table S1 in the Supplement). The net-work indicated that the inshore station had morecomplex associations between OTUs, in terms ofnumber of edges and connectivity of nodes (size ofthe node), than the offshore station. Additionally, alow number of connections between the 2 stationswere detected since only 5% of the possible connec-tions exhibited MIC > 0.8, and were therefore pre-sented in the network (Fig. 6). MicroG and nanoGprovided a modest contribution to the number of

nodes, since a low number of OTUs belonging tothese trophic groups showed connections with MIC >0.8 (27 and 20 OTUs, respectively). In addition, thesegroups showed low connectivity between themselvesand other groups included in the network, as indi-cated by the relatively small size of the nodes (Fig. 6).Furthermore, symbdec, auto, mixo, and picoG ap -peared to be more important at the inshore station,where they were represented by a higher number ofOTUs (nodes), which also had more connections(edges) between themselves and other OTUs. TheOTUs belonging to these groups comprised >75% ofthe total number of nodes and >80% of the totalnumber of edges.

The calculated topological parameters indicatedthat the inshore station had more complex OTU asso-ciations than the offshore station (Fig. 6). In particu-lar, the clustering coefficient (representing the ratioof existing links connecting a node’s neighbors toeach other divided by the maximum possible numberof such links), the shortest path (lowest number ofedges needed to pass through all nodes), the averagepath length (average number of edges along theshortest paths for all possible pairs of networknodes), and the average number of neighbors were,in all cases, about 2 times higher at the inshore sta-tion. Furthermore, the inshore station had a highercentralization value (an expression of how ‘tightly’the graph organized around its most central point;0.11 vs. 0.03), indicating the prevalence of focalpoints at the inshore station around which the rest ofthe nodes were organized. Finally, the inshore sta-tion showed a higher number of nodes (108 vs. 98),higher values of network density (which describesthe portion of the potential connections in a networkthat are actual connections; 0.04 vs. 0.02), networkdiameter (the longest of all the calculated shortestpaths between 2 nodes in a network; 10 vs. 9), andhigher heterogeneity (1.08 vs. 0.64) (Fig. 6).

69

Offshore/ Offshore/ Offshore/ Offshore Offshore Inshore Inshore inshore vs. inshore vs. inshore vs. vs. vs. vs. vs. Environmental Taxonomic Trophic Environmental/ Environmental/ Environmental/ Environmental/ taxonomic trophic taxonomic trophic

p-value 0.001 0.36 0.27 0.067 0.189 0.075 0.033RV 0.8 0.33 0.22 0.40 0.34 0.37 0.41% Axis 1 67.3 60.4 72.4 69.5 47.4 40.4 63.8% Axis 2 91.0 77.8 96.5 14.6 27.5 18.8 25.4

Table 4. Co-inertia analysis (PCA−PCA COIA) between the environmental variables, taxonomic, and trophic groups at the 2 sam-pled stations (offshore and inshore) in the eastern English Channel. The p-value (randomization test with 1000 permutations)shows the significance (in bold) of the relationship between the 2 matrices in each case. The correlation between the matrices isgiven by the RV-coefficient. Percentage of covariation is explained by the first 2 COIA axes (% Axis 1 and Axis 2, respectively)

Parameters Filinger-Killeen test Wilcoxon test

kd (m−1) 0.0008 0.0344PAR (E m−2 d−1) 0.4170 nsT (°C) 0.3117 nsSalinity 0.0225 0.0007O2 (ml l−1) 0.3920 nspH 0.1497 0.01707NH4 (µM) 0.3644 nsNO3

–+NO2– (µM) 0.2464 ns

PO43– (µM) 0.1368 0.0006

SiO44– (µM) 0.0118 ns

POC (mg l−1) 0.3737 0.0012SPM (mg l−1) 0.1957 0.0105Chl a (µg l−1) 0.4658 ns

Table 3. p-values of the Flinger-Killeen test for equal coeffi-cients of variation (CV) and the Wilcoxon test between the en-vironmental variables of the inshore versus offshore stations.Significant p-values (p < 0.05) are shown in bold; ns: not sig-nificant. kd: diffuse attenuation coefficient for down-wellingirradiance; PAR: photosynthetically active radiation; POC: par-ticulate organic carbon; SPM: suspended particulate matter

Page 10: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 2016

DISCUSSION

In this era of substantial global change, a growinginterest in ecosystem function and response (e.g.Duffy & Stachowicz 2006) has stimulated the devel-opment of methods for assessing change in commu-nities and measuring their response to disturbance(Magurran 2011). Here, NGS was used to determinethe composition of protistan communities in 2 closelylocated coastal stations and to understand the mainfactors (both biotic and abiotic) controlling their com-munity structure.

The EEC has a megatidal regime in which strongtidal currents alternate and remain essentially paral-lel to the coast; the coastal water drifts nearshore,separated from the open sea. The lowest salinity val-ues measured in all inshore station samples con-firmed that the continental inputs from the bays ofthe Somme, Authie, and Canche (which consist ofless saline and more turbid coastal waters) are re -stricted to the coastal area and are separated fromthe offshore waters by an unstable tidal front (Brylin-ski & Lagadeuc 1990, Lagadeuc et al. 1997). Thishydrological regime is responsible for the different

70

Ax1

Ax2

Ax3

x-axes

Ax1

Ax2

Ax3

y-axes

Eigenvalues

d = 1

1

2 3

4

5

6

7 8 9

10

11

12

13

14

15

d = 0.2

auto

microG

mixo picoG nanoG

parasites

d = 0.2

kd

PAR

T

S

O2

pH NH4

NO3+NO2

PO4 SiO4

POC

SPM CHLa

Inshore

y Canonical weights x Canonical weights

p-value = 0.033RV = 0.41Ax1 (%) = 63.9Ax2 (%) = 25.5

Sampling Dates1. 20/03/20122. 09/05/20123. 05/06/20124. 21/06/20125. 04/07/20126. 03/09/20127. 03/10/20128. 13/11/20129. 11/02/201310. 26/02/201311. 26/03/201312. 08/04/201313. 27/05/201314. 10/06/201315. 25/06/2013

Ax1 = 41.6 %Ax2 = 22.9 %Ax3 = 16.1 %

Ax1 = 50.0 %Ax2 = 27.5 %Ax3 = 10.8 %

Fig. 5. Co-inertia analysis (PCA−PCA COIA) between environmental variables and operational taxonomic unit (OTU) abun-dance sorted into trophic groups. The x-axes show projections of the first 3 PCA components from the explanatory variables (en-vironmental variables); y-axes show those of the response variables (trophic groups). The circles represent a view of the rotationneeded to associate the 2 datasets. p-values were calculated using the Monte Carlo permutation test (1000 permutations). Thesample scatterplot shows how far apart the samples were relative to their explanatory and response variables. The beginning ofthe arrow shows the position of the sample described by the explanatory variables, and the end by the response variables. Vari-ables were log(x + 1) transformed, and environmental variables were additionally standardized. RV: correlation coefficient

between the 2 tables (‘R’ for correlation and ‘V’ for vectorial)

Page 11: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities

environmental forcing that prevails at the inshoreand offshore stations. Indeed, the measurements andanalyses of 13 different parameters (see Table S2 inthe Supplement at www. int-res. com/ articles/ suppl/m548p061_ supp. pdf) showed that although the envi-ronmental variables presented a quasi-identical sea-sonal evolution over the duration of the study at the 2stations (COIA analysis; Table 4), statistical differ-ences relative to their variability and mean rankswere detected. Several parameters reflecting thequantity of the suspended material in the water (live,detritus, or inorganic material) such as chl a, POC,SPM, and kd were significantly higher at the inshorestation (Fig. 4), while the concentrations of nutrientssuch as SiO4

4– and NO3–+NO2

–, were significantly different between the 2 stations (Table 3).

In a previous study dealing with mesozooplanktoncommunities at the same site, Brylinski & Aelbrecht

(1993) reported that distinct zooplank-ton populations were maintaineddespite the close proximity of the sta-tions sampled. This was attributed tothe different environmental pressuresexhibited at the 2 stations caused bythe hydrochemical characteristics inthe area (Brylinski & Aelbrecht 1993).Indeed, various studies have suggestedthat the spatiotemporal dy namics inmicrobial planktonic composition areoften related to variability in environ-mental parameters (Hullar et al. 2006,Stomp et al. 2011, Read et al. 2015,Wang et al. 2015), although coherentpatterns are often not apparent.

In addition, strong spatial variabilityin the taxonomic composition of protis-tan communities has been documentedat distant oceanic (Not et al. 2009, Per-nice et al. 2013) and coastal locations(Logares et al. 2012, 2014, Massana etal. 2015), reflecting different environ-mental pressures. Furthermore, a strik-ing consistency in the relative propor-tions of protistan assemblages amongclosely located sites (2 to 21 km apart)has been shown (Lie et al. 2013, Loga-res et al. 2014, Massana et al. 2015). Inthis study, the protistan communities ofthe 2 sampling sites shared >60% ofthe total number of OTUs over thewhole sampling period, and had simi-lar taxonomic composition (Fig. 2).However, comparison of several diver-

sity estimators (shown in Table 2) revealed that theindices differed significantly between the 2 sites forall dates. The α-diversity estimators showed consis-tent patterns with previous studies conducted in thearea with NGS tools (Genitsaris et al. 2015); forexample, the lowest equitability and highest domi-nance indices (Simpson and Berger-Parker indices)that were calculated during the Phaeocystis globosapeaks. However, these estimators have a relativeweight in comparisons among samples analyzedwithin the same pipeline, as they are based on thenumber of se quences, not the number of detectedindividuals. Among eukaryotes, the number of rRNAgene copies depends on cell size, and varies from oneto several thousand (Zhu et al. 2005), providingbiased estimates of taxa abundance.

Our results confirm differences in environmentalparameters and community structure between the 2

71

Fig. 6. Network diagram of the operational taxonomic units (OTUs; nodes) showingcorrelations (edges) with maximal information coefficient (MIC) values >0.8. Dif-ferent colors: different trophic groups, based on the trophic status of the detectedOTUs in marine systems as inferred from the literature (see Table 1, Table S1 in theSupplement). The size of the nodes is analogous to the clustering coefficient foreach OTU. Table shows the topological parameters characterizing the individualnetwork of each sampling station (inshore and offshore).Avg.:average;Nb.:number

Page 12: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 201672

stations. In a seasonal succession investigation of theprotistan communities in the EEC at a single site,Genitsaris et al. (2015) showed with canonical analysisthat the measured environmental parameters ex-plained only 30% of the protistan seasonal variation,and that biotic interactions among co-occurring taxaappeared to be the main structural drivers of the tem-poral assemblages detected. A similarly low percent-age of protistan variation explained by environmentalparameters was reported by Chow et al. (2014), in astudy that attempted to characterize ecological rela-tionships between viruses, bacteria, and protists inthe ocean. That study found that the predictable vari-ability within protists was approximately 10% accord-ing to environmental factors (especially day length).The advantage of the analysis used here (COIA) isthat it is based on partial least-squares regression;consequently, it places no restrictions on the numberof variables that can be analyzed (unlike classiccanonical models). The only significant coupling be-tween the 13 environmental parameters and the pro-tist community was found at the inshore station, afterthe protists were sorted into trophic groups (Table 4).

The trophic group symbdec included organisms in-volved in organic matter degradation, i.e. parasites,symbionts, and decomposers (Table 1). Most of thesequences belonged to MALV I, which have a widehost spectrum, whereas MALV II (which were alsoabundant) are considered dinoflagellate parasites(Massana et al. 2011). The considerable abundanceand diversity of MALV suggests interactions with var-ious hosts (Skovgaard et al. 2005, Massana & Pedrós-Alió 2008), such as dinoflagellates (Christaki et al.2015). However, direct evidence that most MALVclades are parasitic is lacking (Worden et al. 2015).The associations observed between MALV and theirhosts are equally consistent with symbiotic relation-ships (Bråte et al. 2012), and environmental factorscould potentially shift a relationship from commensalto pathogenic (Worden et al. 2015). In the presentstudy, the absence of a correlation between symbdecand the COIA co-ordinates indicates that the variabil-ity of this trophic group at the inshore station cannotbe explained by any of the variables chosen in thisstudy (Fig. 5). This could be because the organismsincluded in this trophic group are associated withother organisms, and thus, there is a lag-time in theirresponse to perturbations in the environment. Statis-tical analysis based on lagged as soci ations can beused to predict interactions in time series includingseveral seasonal cycles and with equidistant samplingpoints (e.g. Steele et al. 2011, Xia et al. 2011, 2013,Chow et al. 2014, re viewed in Faust et al. 2015).

The trophic group mikroG was associated withwaters exhibiting relatively low nutrients but highPAR and high salinity, while the inverse was true forthe auto group. This pattern is consistent with theplankton succession evidenced in the EEC (Gratte -panche et al. 2011a, Christaki et al. 2014, Genitsariset al. 2015). Within the autotroph-affiliated OTUs, aP. globosa-related OTU and various diatom-relatedOTUs were among the dominant groups. Everyspring, a P. globosa bloom is preceded and followedby communities of colonial diatoms. MicroG, consist-ing mainly of heterotrophic dinoflagellates, accountfor a large fraction of grazing on phytoplankton, andappear at the end of the diatom blooms when nutri-ents are depleted (Grattepanche et al. 2011b). Thetrophic groups mixo, picoG, and nanoG were allassociated with relatively warm (high T values) andclear (high PAR, but low kd values) waters, but whichwere relatively low in O2, chl a, and SPM (Fig. 5).Pico- and nanograzers consume heterotrophic bacte-ria and small phytoplankton cells, while the ‘mixo’group includes many of these small phytoplanktonand several small dinoflagellates. Ciliates, whichdominated the nanoG group, are known to increaseat the end of the Phaeocystis bloom (between Mayand July) each year, and graze on the free P. globosacells liberated by the senescent colonies (Gratte -panche et al. 2011a,b).

Overall, the environmental parameters explainonly a small part of the community structure, andonly at the inshore station. Therefore, in order to pro-vide additional information on the protistan commu-nities’ complexities, the next step was to investigatethe effect of biotic interactions by examining the co-occurrence networks of the protistan communities ateach site. These networks reveal elements of the nat-ural history of various microbes (Steele et al. 2011).Our data analysis revealed 2 sub-networks corre-sponding to each site, highlighting differences be -tween the inshore and offshore communities’ organi-zation (Fig. 6). In particular, the inshore stationexhibited more complex interactions among the con-nected OTUs. This was shown by the higher valuesof almost all topological parameters calculated foreach sub-network corresponding to each samplingsite. Furthermore, the inshore station had a highercentralization value, indicating the prevalence offocal points at this station, around which the rest ofthe nodes were organized. The higher centralizationvalue, and most importantly the higher clusteringcoefficient of the inshore station, suggests that it mayhave ‘small world’ properties, with a few highly con-nected nodes (e.g. the autotroph cluster) compared to

Page 13: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities 73

the offshore station (Watts & Strogatz 1998). Thisproperty would render the inshore protistan commu-nity more resilient to environmental pressures, andmore susceptible to removal of highly connected spe-cies/nodes (Montoya et al. 2006). Indeed, it has beenshown that in trophic webs, robustness increaseswith connectivity independent of taxa abundance(Dunne et al. 2002) in the same way that increasingspecies richness can stabilize community structure inthe face of perturbation (Tilman et al. 1997). Al -though classical theory of ecosystem resilience pre-dicts that the stability of food webs should decreasewith increasing complexity (May 1974, McCann et al.1998), Vallina & Le Quéré (2011) demonstrated thatwith an increase of food web complexity, the overallstability of the ecosystem in fact increases by increas-ing resistance to climatic perturbations. Undoubt-edly, the use of molecular data is foreseeable in thefuture exploration of ecological questions (e.g. Faust& Raes 2012, Faust et al. 2015, Fuhrman et al. 2015and references therein). However, the terminologyand metrics used in macroecology must be appliedwith greater caution because the methods availablewith which to characterize microbial communitiesremain at ‘intermediate resolution’. This is becausethere is still no means to ob tain accurate quantitativedata and to determine the accuracy of extraction, bio -informatic analysis, and detection methods (Øvreås &Curtis 2011).

In summary, our results suggest that due to thehigher and more variable environmental pressuresacting on the inshore station, its protistan commu-nity has differentiated from that of the offshore site.This has resulted in the development of more com-plex connections that are reflected the rapid re -sponse of trophic interactions within the wholemicrobial community. Network analysis demon-strated different de grees of complexity in commu-nity structure be tween the sites, and highlighted thepotential importance of microbial interactions incounterbalancing environmental variability at these2 closely located coastal stations. In addition, theanalysis based on functional/ trophic roles was foundto be more informative than that based solely ontaxonomic affiliation.

Acknowledgements. This study was supported by the‘Nord-Pas de Calais’ FRB-DEMO (FRB 2013) project and theSOMLIT network. The authors thank C. Georges for helpwith sampling, and J. M. Brylinski and A. Senchev for help-ful discussions. We thank www.englisheditor.webs.com forEnglish proofing. We are also thankful to the anonymousreferees for suggestions that helped improve the originalmanuscript.

LITERATURE CITED

Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ (1990)Basic local alignment search tool. J Mol Biol 215: 403−410

Amann RI, Blinder BJ, Olson RJ, Chisholm SW, Devereux R,Stahl DA (1990) Combination of 16S rRNA-targeted oli-gonucleotide probes with flow cytometry for analyzingmixed microbial populations. Appl Environ Microbiol 56: 1919−1925

Aminot A, Kérouel R (2004) Dissolved organic carbon, nitrogen, and phosphorus in the N-E Atlantic and the N-W Mediterranean with particular reference to non-refractory fractions and degradation. Deep-Sea Res I 51: 1975−1999

Bachy C, Dolan JR, López-García P, Deschamps P, MoreiraD (2013) Accuracy of protist diversity assessments: mor-phology compared with cloning and direct pyrosequenc-ing of 18S rRNA genes and ITS regions using the con-spicuous tintinnid ciliates as a case study. ISME J 7: 244−255

Behnke A, Engel M, Christen R, Nebel M, Klein RR, StoeckT (2011) Depicting more accurate pictures of protistancommunity complexity using pyrosequencing of hyper-variable SSU rRNA gene regions. Environ Microbiol 13: 340−349

Bonato S, Christaki U, Lefebvre S, Lizon F, Thyssen M, Arti-gas LF (2015) High spatial variability of phytoplanktonassessed by flow cytometry, in a dynamic productivecoastal area, in spring: the eastern English Channel.Estuar Coast Shelf Sci 154: 214−223

Bonato S, Breton E, Didry M, Lizon F and others (2016) Spa-tio-temporal patterns in phytoplankton assemblages ininshore−offshore gradients using flow-cytometry: a casestudy in the eastern English Channel. J Mar Syst 156: 76−85

Bråte J, Krabberød AK, Dolven JK, Ose RF, Kristensen T,Bjørklund KR, Shalchian-Tabrizi K (2012) Radiolariaassociated with large diversity of marine alveolates. Pro-tist 163: 767−777

Brylinski JM, Aelbrecht D (1993) Plankton transfers andcoastal front in the Dover Strait. Oceanol Acta 16: 671−676

Brylinski JM, Lagadeuc Y (1990) L’interface eaux côtières/eaux du large dans le Pas-de-Calais (côte française): unezone frontale. C R Acad Sci Paris Ser II 311: 535−540

Burke C, Steinberg P, Rusch D, Kjelleberg S, Thomas T(2011) Bacterial community assembly based on func-tional genes rather than species. Proc Natl Acad Sci USA108: 14288−14293

Caron DA, Countway PD (2009) Hypotheses on the role ofthe protistan rare biosphere in a changing world. AquatMicrob Ecol 57: 227−238

Caron DA, Countway PD, Jones AC, Kim DY, Schnetzer A(2012) Marine protistan diversity. Annu Rev Mar Sci 4: 467−493

Chow CET, Kim DY, Sachdeva R, Caron DA, Fuhrman JA(2014) Top-down controls on bacterial community struc-ture: microbial network analysis of bacteria, T4-likeviruses and protists. ISME J 8: 816−829

Christaki U, Kormas KA, Genitsaris S, Georges C, Sime-Ngando T, Viscogliosi E, Monchy S (2014) Winter-sum-mer succession of unicellular eukaryotes in a meso-eutrophic coastal system. Microb Ecol 67: 13−23

Christaki U, Georges C, Genitsaris S, Monchy S (2015) Micro-zooplankton community associated with phytoplankton

Page 14: Small-scale variability of protistan planktonic

Mar Ecol Prog Ser 548: 61–75, 2016

blooms in the naturally iron-fertilized Kerguelen area(Southern Ocean). FEMS Microbiol Ecol 91: fiv068, doi: 10.1093/femsec/fiv068

Dolédec S, Chessel D (1994) Co-inertia analysis: an alterna-tive method for studying species-environment relation-ships. Freshw Biol 31: 277−294

Dray S, Dufour AB (2007) The ade4 package: implementingthe duality diagram for ecologists. J Stat Softw 22: 1−20

Dray S, Chessel D, Thioulouse J (2003) Co-inertia analysisand the linking of ecological data tables. Ecology 84: 3078−3089

Duffy JE, Stachowicz JJ (2006) Why biodiversity is impor-tant to oceanography: potential roles of genetic, species,and trophic diversity in pelagic ecosystem processes.Mar Ecol Prog Ser 311: 179−189

Dunne JA, Williams RJ, Martinez ND (2002) Network struc-ture and biodiversity loss in food webs: robustnessincreases with connectance. Ecol Lett 5: 558−567

Edgar RC (2010) Search and clustering orders of magnitudefaster than BLAST. Bioinformatics 26: 2460−2461

Faust K, Raes J (2012) Microbial interactions: from networksto models. Nat Rev Microbiol 10: 538−550

Faust K, Lahti L, Gonze D, de Vos WM, Raes J (2015) Meta -genomics meets time series analysis: unraveling micro-bial community dynamics. Curr Opin Microbiol 25: 56−66

Fuhrman JA, Steele JA (2008) Community structure of mar-ine bacterioplankton: patterns, networks, and relation-ships to function. Aquat Microb Ecol 53: 69−81

Fuhrman JA, Cram JA, Needham DM (2015) Marine micro-bial community dynamics and their ecological interpre-tation. Nat Rev Microbiol 13: 133−146

Genitsaris S, Monchy S, Viscogliosi E, Sime-Ngando T, Fer-reira S, Christaki U (2015) Seasonal variations of marineprotist community structure based on taxon-specifictraits using the eastern English Channel as a modelcoastal system. FEMS Microbiol Ecol 91: fiv034 doi: 10.1093/ femsec/ fiv034

Gilbert JA, Steele JA, Caporaso JG, Steinbrück L and others(2012) Defining seasonal marine microbial communitydynamics. ISME J 6: 298−308

Grattepanche JD, Breton E, Brylinksi JM, Lecuyer E, Chris-taki U (2011a) Succession of primary producers and micro -grazers in a coastal system dominated by Phaeocystsis glo-bosa blooms. J Plankton Res 33: 37−50

Grattepanche JD, Vincent D, Breton E, Christaki U (2011b)Phytoplankton growth and microzooplankton grazingduring a spring bloom in the eastern English Channel.J Exp Mar Biol Ecol 404: 87−97

Guillou L, Bachar D, Audic S, Bass D and others (2013) TheProtist Ribosomal Reference database (PR2): a catalog ofunicellular eukaryote small sub-unit rRNA sequenceswith curated taxonomy. Nucleic Acids Res 41: D597−D604

Hammer Ø, Harper DAT, Ryan PD (2001) PAST: paleonto-logical statistics software package for education anddata analysis. Paleontol Electron 4: art4 http://palaeo-electronica.org/2001_1/past/issue1_01.htm

Harada A, Ohtsuka S, Horiguchi T (2007) Species of the par-asitic genus Dubusquella are members of the enigmaticMarine Alveolate Group I. Protist 158: 337−347

Holmes RM, Aminot A, Kérouel R, Hooker BA, Peterson BJ(1999) A simple and precise method for measuringammonium in marine and freshwater ecosystems. Can JFish Aquat Sci 56: 1801−1808

Hullar MA, Kaplan LA, Stahl DA (2006) Recurring seasonaldynamics of microbial communities in stream habitats.

Appl Environ Microbiol 72: 713−722Kenkel NC (2006) On selecting an appropriate multivariate

analysis. Can J Plant Sci 86: 663−676Kunin V, Engelbrektson A, Ochman H, Hugenholtz P (2010)

Wrinkles in the rare biosphere: pyrosequencing errorscan lead to artificial inflation of diversity estimates. Env-iron Microbiol 12: 118−123

Lagadeuc Y, Brylinski JM, Aelbrecht D (1997) Temporalvariability of the vertical stratification of a front in a tidalRegion Of Freshwater Influence (ROFI) system. J MarSyst 12: 147−155

Laws EA (2003) Partitioning of microbial biomass in pelagicaquatic communities: maximum resiliency as a food weborganizing construct. Aquat Microb Ecol 32: 1−10

Legendre P, Legendre L (1998) Numerical ecology, 2nd edn.Elsevier, Amsterdam

Lie AAY, Kim DY, Schnetzer A, Caron DA (2013) Small-scaletemporal and spatial variations in protistan communitycomposition at the San Pedro Ocean time-series stationoff the coast of southern California. Aquat Microb Ecol70: 93−110

Logares R, Audic S, Santini S, Pernice MC, de Vargas C,Massana R (2012) Diversity patterns and activity ofuncultured marine heterotrophic flagellates unveiledwith pyrosequencing. ISME J 6: 1823−1833

Logares R, Audic S, Bass D, Bittner L and others (2014) Pat-terns of rare and abundant marine microbial eukaryotes.Curr Biol 24: 813−821

López-García P, Philippe H, Gail F, Moreira D (2003) Auto -chthonous eukaryotic diversity in hydrothermal sedi-ment and experimental microcolonizers at the Mid-Atlantic Ridge. Proc Natl Acad Sci USA 100: 697−702

Lorenzen CJ (1967) Determination of chlorophyll andpheo-pigments: spectrophotometric equations. LimnolOceanogr 12: 343−346

Magurran AE (2011) Measuring biological diversity in time(and space). In: Magurran AE, McGill BJ (eds) Biologicaldiversity: frontiers in measurement and assessment.Oxford University Press, Oxford, p 85−96

Massana R, Pedrós-Alió C (2008) Unveiling new microbialeukaryotes in the surface ocean. Curr Opin Microbiol 11: 213−218

Massana R, Terrado R, Forn I, Lovejoy C, Pedrós-Alió C(2006) Distribution and abundance of uncultured hetero-trophic flagellates in the world oceans. Environ Micro-biol 8: 1515−1522

Massana R, Pernice M, Bunge JA, del Campo J (2011)Sequence diversity and novelty of natural assemblagesof picoeukaryotes from the Indian Ocean. ISME J 5: 184−195

Massana R, Gobet A, Audic S, Bass D and others (2015) Mar-ine protist diversity in European coastal waters and sedi-ments as revealed by high-throughput sequencing. Envi-ron Microbiol 17: 4035−4049

May RM (1974) Stability and complexity in model ecosys-tems. Princeton University Press, Princeton, NJ

McCann K, Hastings A, Huxel GR (1998) Weak trophic inter -actions and the balance of nature. Nature 395: 794−798

Medinger R, Nolte V, Pandey RV, Jost S, Ottenwälder B,Schlötterer C, Boenigk J (2010) Diversity in a hiddenworld: potential and limitation of next-generation sequen-cing for surveys of molecular diversity of eukaryoticmicroorganisms. Mol Ecol 19(Suppl 1): 32−40

Monchy S, Grattepanche JD, Breton E, Meloni D and others(2012) Microplanktonic community structure in a coastal

74

Page 15: Small-scale variability of protistan planktonic

Genitsaris et al.: Spatial variability in protistan communities

system relative to a Phaeocystis bloom inferred frommorphological and tag pyrosequencing methods. PLoSONE 7: e39924

Montoya JM, Pimm SL, Solé RV (2006) Ecological networksand their fragility. Nature 442: 259−264

Morel A, Smith RC (1974) Relation between total quanta andtotal energy for aquatic photosynthesis. Limnol Oceanogr19: 591−600

Not F, del Campo J, Balagué V, de Vargas C, Massana R(2009) New insights into the diversity of marinepicoeukaryotes. PLoS ONE 4: e7143

Øvreås L, Curtis TP (2011) Microbial diversity and ecology.In: Magurran AE, McGill BJ (eds) Biological diversity: frontiers in measurement and assessment. Oxford Uni-versity Press, Oxford, p 221−236

Pernice MC, Logares R, Guillou L, Massana R (2013) Gen-eral patterns of diversity in major marine microeukaryotelineages. PLoS ONE 8: e57170

Quince C, Lanzén A, Curtis TP, Davenport RJ and others(2009) Accurate determination of microbial diversityfrom 454 pyrosequencing data. Nat Methods 6: 639−641

Raghukumar S (2004) The role of fungi in marine detritalprocesses. In: Ramaiah NE (ed) Marine microbiology: facts, opportunities. National Institute of Oceanography,Goa, p 91−101

Read DS, Gweon HS, Bowes MJ, Newbold LK, Field D, BaileyMJ, Griffiths RI (2015) Catchment-scale biogeography ofriverine bacterioplankton. ISME J 9: 516−526

Reshef DN, Reshef YA, Finucane HK, Grossman SR and oth-ers (2011) Detecting novel associations in large data sets.Science 334: 1518−1524

Riley GA (1957) Phytoplankton of the North Central Sar-gasso Sea, 1950–52. Limnol Oceanogr 2:252–270

Salter I, Galand PE, Fagervold SK, Lebaron P and others(2015) Seasonal dynamics of active SAR11 ecotypes inthe oligotrophic Northwest Mediterranean Sea. ISME J9: 347−360

Santoferrara LF, Grattepanche JD, Katz LA, McManus GB(2014) Pyrosequencing for assessing diversity of eukary-otic microbes: analysis of data on marine planktonic cili-ates and comparison with traditional methods. EnvironMicrobiol 16: 2752−2763

Schloss PD, Westcott SL, Ryabin T, Hall JR and others (2009)Introducing mothur: open-source, platform-indepen-dent, community-supported software for describing andcomparing microbial communities. Appl Environ Micro-biol 75: 7537−7541

Schloss PD, Gevers D, Westcott SL (2011) Reducing theeffects of PCR amplification and sequencing artifacts on16S rRNA-based studies. PLoS ONE 6: e27310

Schnepf E, Kühn SF (2000) Food uptake and fine structure ofCryothecomonas longipes ps. nov., a marine nanoflagel-late incertae sedis feeding phagotrophically on largediatoms. Helgol Mar Res 54: 18−32

Shade A, Jones SE, Caporaso JG, Handelsman J, Knight R,Fierer N, Gilbert JA (2014) Conditionally rare taxa dispro-portionally contribute to temporal changes in microbialdiversity. MBio 5: e01371-14, doi: 10. 1128/ mBio. 01371-14

Sherr EB, Sherr BF (1994) Bacterivory and herbivory: key

roles of phagotrophic protists in pelagic food webs.Microb Ecol 28: 223−235

Sherr BF, Sherr EB, Caron DA, Vaulot D, Worden AZ (2007)Oceanic protists. Oceanography 20: 130−134

Skovgaard A (2014) Dirty tricks in the plankton: diversity androle of marine parasitic protists. Acta Protozool 53: 51−62

Skovgaard A, Massana R, Balague V, Saiz E (2005) Phyloge-netic position of the copepod-infesting parasite Syndini -um turbo (Dinoflagellata, Syndinea). Protist 156: 413−423

Smoot ME, Ono K, Ruscheinski J, Wang PL, Ideker T (2011)Cytoscape 2.8: new features for data integration and net-work visualization. Bioinformatics 27: 431−432

Steele JA, Countway PD, Xia L, Vigil PD and others (2011)Marine bacterial, archaeal and protistan association net-works reveal ecological linkages. ISME J 5: 1414−1425

Stomp M, Huisman J, Mittelbach GG, Litchman E, Klaus-meier CA (2011) Large-scale biodiversity patterns infreshwater phytoplankton. Ecology 92: 2096−2107

Székely AJ, Langenheder S (2014) The importance of spe-cies sorting differs between habitat generalists and spe-cialists in bacterial communities. FEMS Microbiol Ecol87: 102−112

Taylor BW, Keep CF, Hall RO Jr, Koch BJ, Tronstad LM,Flecker AS, Ulseth AJ (2007) Improving the fluorometricammonium method: matrix effects, background fluores-cence, and standard additions. J N Am Benthol Soc 26: 167−177

Tillmann U, Hesse KJ, Tillmann A (1999) Large-scale para-sitic infection of diatoms in the Northfrisian Wadden Sea.J Sea Res 42: 255−261

Tilman D, Lehman C, Thompson K (1997) Plant diversity andecosystem productivity: theoretical considerations. ProcNatl Acad Sci USA 94: 1857−1861

Vallina SM, Le Quéré C (2011) Stability of complex foodwebs: resilience, resistance and the average interactionstrength. J Theor Biol 272: 160−173

Wang Y, Liu L, Chen H, Yang J (2015) Spatiotemporaldynamics and determinants of bacterial and micro eukar -yote communities in a Chinese subtropical river. ApplMicrobiol Biotechnol 99: 9255−9266

Watts DJ, Strogatz SH (1998) Collective dynamics of ‘small-world’ networks. Nature 393: 440−442

Worden AZ, Follows MJ, Giovannoni SJ, Wilken S, Zimmer-man AE, Keeling PJ (2015) Rethinking the marine carboncycle: factoring in the multifarious lifestyles of microbes.Science 347: 1257594

Xia LC, Steele JA, Cram JA, Cardon ZG and others (2011)Extended local similarity analysis (eLSA) of microbialcommunity and other time series data with replicates.BMC Syst Biol 5(Suppl 2): S15, doi: 10.1186/1752-0509-5-S2-S15

Xia LC, Dongmei A, Cram J, Fuhrman JA, Sun F (2013) Effi-cient statistical significance approximation for local asso-ciation analysis of high-throughput time series data.Bioinformatics 29: 230−237

Zhu F, Massana R, Not F, Marie D, Vaulot D (2005) Mappingof picoeukaryotes in marine ecosystems with quantita-tive PCR of the 18S rRNA gene. FEMS Microbiol Ecol 52: 79−92

75

Editorial responsibility: Toshi Nagata, Kashiwanoha, Japan

Submitted: October 13, 2015; Accepted: February 2, 2016Proofs received from author(s): April 4, 2016

➤➤