Supplementary Information (doc 161K)

advertisement
1
Supporting Online Material
2
Supplementary Methods
3
Sampling
4
Seawater samples were collected on 72 occasions from January 2003-December 2008 from the
5
L4 sampling site (50° 15.00' N, 4° 13.02') of the Western Channel Observatory
6
(http://www.westernchannelobservatory.org.uk; Table S1, S2, S3). Nineteen environmental
7
variables were obtained from the WCO database, available on the website; these included surface
8
temperature, salinity, water density, silicate concentration, nitrate + nitrite concentration,
9
chlorophyll
concentration,
total
organic
nitrogen
(TON)
concentration,
ammonium
10
concentration, soluble reactive phosphorus (SRP) concentration, total organic carbon (TOC)
11
concentration, photo-active radiation (PAR), phytoplankton and zooplankton taxa abundances,
12
etc. Other variables measured were flow cytometric determination of bacterial abundance,
13
microscope counts of cryptophyte, picoeukaryote, nanoeukaryote, coccolithophore and small
14
dinoflagellate abundance. All methods for determining these variables are available on the WCO
15
website (http://www.westernchannelobservatory.org.uk/all_parameters.html) and are detailed in
16
(Widdicombe, Eloire et al. 2010).
17
18
DNA extraction, 16S rDNA V6 amplification and Pyrosequencing
19
Nucleic acid was extracted from 5 L seawater, collected from the surface and filtered
20
immediately through a 0.22 µm Sterivex cartridge (Millipore), which was stored at -80 °C. DNA
21
was isolated from each sample (Neufeld, Schafer et al. 2007) and then stored at -20 °C. DNA
22
was used as a template for V6-region 16S rRNA amplification using the method of Huber et al.
23
(Huber, Mark Welch et al. 2007) in sets of twelve. Each amplicon in set of twelve was labeled
1
24
with a unique multiplex identifier (MID) sequence (Table S8). Subsequently, amplicon product
25
pools for each year were pyrosequenced on ½ of a 454 pico-titre plate using the GS-flx platform.
26
All sequences have been submitted to the NCBI short reads archive under SRA009436, and
27
registered with the GOLD database (Gm00104). All data submitted are MIMARKS compliant
28
(Yilmaz, Kottmann et al. 2010).
29
30
Denoising or preclustering of pyrosequencing data
31
From the complete raw dataset of 747,496 sequences, there were 65,059 unique
32
sequences (100% nucleotide identity clustering (NIC)). To compensate for potential
33
overestimation of diversity using pyrosequencing (Quince, Lanzen et al. 2009), the data set was
34
further treated using a sequence clustering approach called Single Linkage Preclustering (SLP
35
(Huse, Welch et al. 2010)), and an algorithm based sff clustering approach called Denoiser V
36
0.851 (Reeder and Knight 2010). Denoiser clusters OTUs at 3% sequence identity following
37
alignment and clustering of the flowgrams. SLP uses a highly conservative 2% single-linkage
38
pre-clustering approach, followed by an average-linkage clustering based on pair-wise
39
alignments. From the combined 72 datasets, Denoiser produced 21,129 OTUs, while the
40
considerably more conservative SLP resulted in 8,794 OTUs. In this study, the smallest dataset
41
comprised 4101 sequences, whereas the largest dataset comprised 8 times the number of
42
sequences (32,826). This means that the larger dataset had eight times the number of
43
observations and hence could identify more novel OTUs purely on the basis of sample size.
44
Therefore, to enable accurate analysis of species richness and diversity estimates between
45
samples each dataset was randomly-resampled to the smallest sequencing effort using
46
daisychopper (Gilbert, Field et al. 2009). Following random re-sampling and reanalysis using the
2
47
denoising strategies, SLP generated 4,204 OTUs and Denoiser generated 10,215 OTUs. In both
48
datasets, ~50% of the OTUs were represented by only a single sequence (singletons).
49
50
Operational Taxonomic Unit picking and annotation
51
Annotation of data was performed using the Visualization and Analysis of Microbial
52
Population Structure project (VAMPS) website (http://vamps.mbl.edu/index.php). The VAMPS
53
workflow was used to create a profile of the unique sequences, their taxonomic assignment and
54
their abundance in each sample. OTUs in the each dataset were picked and annotated with the
55
GAST process (Sogin, Morrison et al. 2006), which is available through the VAMPS website
56
(http://vamps.mbl.edu/index.php). This process was used for all analyses in the study. However,
57
OTU prediction was also performed on the Denoiser treated data to provide a comparison (Fig.
58
S1a). Following denoising with Denoiser v0.851, OTUs were picked using uclust (Edgar 2010)
59
version q1.2.15 at 97% sequence identity (maxaccepts=8, maxrejects=32). Representative
60
sequences for each OTU were aligned with PyNAST 1.1 (Caporaso, Bittinger et al. 2010) (with
61
min_length=50 to support the short read lengths), and a tree was built using FastTree (Price,
62
Dehal et al. 2009). Unweighted UniFrac distances were calculated between samples using the
63
FastTree tree and 5000 randomly selected sequences were used per sample to control for
64
differences in UniFrac distances which might arise due to differences in sampling (sequencing)
65
depth. Principle Coordinates Analysis (PCoA) was applied to the UniFrac distance matrices, and
66
plotted with an explicit time axis in units of days. Data analysis was performed using the
67
Quantitative Insights Into Microbial Ecology (QIIME) toolkit (Caporaso, Kuczynski et al. 2010).
68
69
Analysis of OTU richness dynamics
3
70
Changes in community structure through time were analysed using nonparametric multivariate
71
methods (Clarke 1993) implemented in Primer v 6 (Clarke and Gorley 2006). Resampled
72
abundances of OTUs (however defined) and subsets of OTUs (Proteobacteria, Bacterioidetes,
73
Cyanobacteria, all other phyla combined) were Log(N+1)-transformed, to downweight
74
contributions to intersample resemblances from the few most numerically abundant OTUs.
75
Intersample resemblances (Clarke and Gorley 2006) were calculated using the Bray-Curtis
76
similarity measure (Somerfield 2008) and the resemblance matrices ordinated using nonmetric
77
multidimensional scaling (MDS). To examine the relative effects of seasonal and interannual
78
variability samples were grouped into seasons (Winter: December, January, February; Spring:
79
March, April, May; Summer: June, July, August; Autumn: September, October, November) and
80
changes in community structure were analysed with 2-way permutation-based analysis of
81
variance (McArdle and Anderson 2001; Anderson, Gorley et al. 2008) using Season and Year as
82
factors, type III (partial) sums of squares and 999 permutations of residuals under a reduced
83
model.
84
Two univariate measures of community structure, the number of different OTUs (S) and
85
Simpson’s evenness index 1-’, were calculated for each sample. Variation in these indices, and
86
environmental variables, were analysed using multiple linear regression (MLR) to determine
87
temporal patterns.
88
regression on Serial Day (D), the number of days from 01/01/2003, a term representing a
89
seasonal cycle mirroring day-length, peaking in midwinter (DX1 = cos(2(d/365)), where d is
90
the number of days from the 20th December, the shortest day in the previous year) and a further
91
orthogonal term to determine the nature of cycles with peaks at other times of the year (DX2 =
92
sin(2(d/365))).
Temporal trends and the form of seasonal cycles were determined by
Univariate measures were further regressed on a range of (sometimes
4
93
transformed) explanatory variables using stepwise MLR.
The relative goodness of fit of
94
different regression models was assessed using the Akaike Information Criterion (AIC). When
95
necessary, single variables were transformed to normalise residuals. Changes in community
96
structure were also analysed using stepwise distance-based linear modelling of variation among
97
Bray-Curtis similarities. Analyses were performed using MINITAB v. 13 and the distance-based
98
linear modelling (DistLM) module within PERMANOVA+ v. 1 add-in for PRIMER v. 6
99
(McArdle and Anderson 2001; Anderson, Gorley et al. 2008). Inter-seasonal and inter-annual
100
variability among environmental variables were further analysed using correlation-based
101
principal components analysis (PCA).
102
103
Analysis of community composition dynamics
104
Bacterial operational taxonomic units (OTUs) were formed from the sequence tags by single-
105
linkage preclustering using pairwise alignments, followed by an average linkage clustering (SLP
106
/ PW-AL (Huse, Welch et al. 2010)). Taxonomic identities for bacteria are based off of 454 16S
107
V6 sequence tags run through the Global Alignment for Sequence Taxonomy (GAST) process,
108
and represent the best agreed upon match for each OTU cluster within the SILVA98 ICOMM
109
database. Taxonomic IDs and representative sequences for the OTUs used in this study are
110
included as supplementary tables (Tables SX). OTU data and sequence data used in this study is
111
archived
112
(http://vampsarchive.mbl.edu/diversity/diversity_old.php). Eukaryotic counts and environmental
113
data were gathered from the PML L4 dataset.
under
ICM_PML_Bv6-Western
English
Channel
114
For bacterial OTUs, the 300 most abundant (i.e. greatest average abundance), 300 most
115
common (i.e. OTUs which occurred most often during the 6 years), and the 300 most variable
5
116
(i.e. OTUs whose abundance changed the most over 6 years) were selected for downstream
117
analysis. Only those bacterial OTUs that occurred a minimum of least 7 months for bacteria were
118
included. Also included in the analysis were 76 eukaryotic variables that were present for a
119
minimum of one third of the dates sampled.
120
121
Discriminant Function Analysis
122
Discriminant Function Analysis (DFA (Manly 2005)) and was used to identify variables (e.g.
123
bacterial OTUs) that best predict the months (Fuhrman, Hewson et al. 2006). Bacterial OTUs
124
and Eukaryotic variables were log(x+1) transformed and taken as percentages of the sample
125
total. Environmental variables were treated as described above. To avoid false patterns, the
126
sample dates were randomized to disrupt the temporal relationships and the DFA was performed
127
on the randomized data, resulting in no similar patterns to those observed with the temporal
128
relationships intact. Time series analysis (Philippi 1993) was used to identify cyclical patterns
129
and significant autocorrelations among months as evidence of a repeating pattern, based upon
130
bacterial OTUs only. Multiple regression analyses (Rasmussen, Heissey et al. 2001) were used to
131
test whether the temporally repeatable patterns in distribution and abundance are predictable by
132
one or more environmental and eukaryotic factors. We used SYSTAT ® (v.11, Systat Software
133
Inc, Chicago, IL, USA) and R 2.10.1 (Team 2010) to perform these analyses.
134
135
Association networks
136
To examine associations between the microbial populations and their environment, we analyzed
137
the correlations, of the bacterial OTUs with each other, with eukaryotic counts, and with biotic
138
and abiotic parameters over time using local similarity analysis (LSA, (Ruan, Dutta et al. 2006;
6
139
Fuhrman and Steele 2008; Fuhrman 2009), Steele, J.A., Countway, P.D., Xia, L., Vigil, P.D.,
140
Beman, J.M., et al. Marine bacterial, archaeal, and protistan association networks reveal
141
ecological linkages. ISME J, In Press). LSA calculates contemporaneous and time-lagged
142
correlations based on normalized ranked data and produces correlation coefficients that are
143
analogous to a Spearman’s ranked correlation (Ruan, Dutta et al. 2006). Any parameters (OTUs
144
or physico-chemical parameters) that occurred in fewer than 7 months were excluded from the
145
analysis. We included 392 variables in each analysis: 300 bacterial OTUs, 74 Eukaryotic
146
parameters, and 18 environmental parameters (Table S1,2,3). To sort through, condense, and
147
visualize the correlations generated by the analysis, we used Cytoscape (Shannon, Markiel et al.
148
2003) to create subnetworks (Figs. 6,7 S2-4) showing highly correlated subsets of the 398
149
variables with significant local similarity correlations (all shown were p ≤ 0.005) with a false
150
discovery rate (i.e. q-value) less than 0.003 (Storey 2002). False discovery rates are calculated
151
from the observed p-value distribution and vary with the p-value (e.g. for the connections with p
152
= 0.005, 0.001, q = 0.003, 0.0009, respectively). These correlation networks include nodes that
153
consisted of bacterial OTUs, Eukaryotic OTUs, and the environmental parameters. The local
154
similarity scores (rank correlations that may include a 1-3 month time lag) among the OTUs and
155
the environmental parameters constituted the edges, i.e. connections between nodes.
156
157
Supplementary results
158
Comparing OTU richness across 6-years using SLP, Denoiser and no-treament
159
The robust seasonal pattern was very similar whether the raw data, Denoiser-treated data or SLP-
160
treated data was used for analysis (correlations based on observed richness in all 72 time points;
161
raw-denoised 0.74; raw-SLP 0.76; denoised-SLP 0.78; Figure S1), which indicates that the
7
162
while these denoising methods have a significant impact on the estimation of richness, they have
163
little impact on the observed seasonal patterns in richness (S – OTU abundance).
164
165
Alpha and Beta-diversity analysis based on Denoiser results
166
Alpha diversity, measured in observed OTUs (S) and Phylogenetic Diversity (PD) (Faith 1992),
167
was relatively constant across the time-series within a defined range, but showed distinct cyclical
168
patterns with maxima in winter and minima in summer (Figure S5a). These trends were
169
consistent irrespective of the applied denoising algorithm. The mean S per time point was 694,
170
with an average minimum of 282 in summer and maximum of 1157 in winter. A cyclic pattern
171
was observed in beta diversity, measured in UniFrac distances (Lozupone and Knight 2005)
172
between the samples (Figure S5b). UniFrac distances, or the fraction of branch length in a
173
phylogenetic tree representing the reads observed in a pair of samples that is unique to either
174
sample, were smallest between time points from the same seasons and larger when comparing
175
different seasons. For example, January of 2003 is more similar to January of 2008 in terms of
176
phylogenetic composition of the samples than either is to July of 2003.
177
178
Community Composition Dynamics
179
Diatoms were not correlated with any bacterial OTUs (Fig. S2), yet diatoms correctly predicted
180
the bacterial community pattern in all but the 300 most common OTUs, which was significantly
181
predicted by phytoplankton counts, although the variance explained by eukaryotes (adjusted r2 =
182
0.18-0.25) was lower for eukaryotic counts compared to environmental factors (Table S6). The
183
correlation networks reveal that diatoms and phytoplankton are correlated to primary production,
184
PAR, and day-length. The diatoms, and other phytoplankton, may also be reacting to a seasonal
8
185
shift in nutrients or energy and providing a change in community to which the bacteria are
186
responding. It is also likely that the broad view that these counts are providing misses
187
interactions between individual eukaryotic and bacterial taxa, but that comparing community-
188
wide indicators (e.g. DFA, dbLM) is allowing the signals to be detected.
189
190
Supplementary References
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
Anderson, M., R. Gorley, et al. (2008). PERMANOVA+ for PRIMER: Guide to Software and Statistical
Methods. Plymouth, PRIMER-E.
Caporaso, J. G., K. Bittinger, et al. (2010). "PyNAST: a flexible tool for aligning sequences to a template
alignment." Bioinformatics 26(2): 266-267.
Caporaso, J. G., J. Kuczynski, et al. (2010). "QIIME allows analysis of high-throughput community
sequencing data." Nature Methods 7(5): 335-336.
Clarke, K. and R. Gorley (2006). PRIMER: User Manual/Tutorial. Plymouth, PRIMER-E.
Clarke, K. R. (1993). "Non-parameteric multivariate analyses of changes in community structure."
Austral Ecology 18(1): 117-143.
Edgar, R. C. (2010). USARCH, http://www.drive5.com/usearch.
Faith, D. P. (1992). "Conservation Evaluation and Phylogenetic Diversity." Biological Conservation
61(1): 1-10.
Fuhrman, J. A. (2009). "Microbial community structure and its functional implications." Nature
459(7244): 193-199.
Fuhrman, J. A., I. Hewson, et al. (2006). "Annually reoccurring bacterial communities are predictable
from ocean conditions." Proc Natl Acad Sci U S A 103(35): 13104-13109.
Fuhrman, J. A. and J. A. Steele (2008). "Community Structure of marine bacterioplankton: patterns,
networks, and relationships to function." Aquatic Microbial Ecology 53(1): 69-81.
Gilbert, J. A., D. Field, et al. (2009). "The seasonal structure of microbial communities in the Western
English Channel." Environ Microbiol 11(12): 3132-3139.
Huber, J. A., D. Mark Welch, et al. (2007). "Microbial population structures in the deep marine
biosphere." Science 318(5847): 97-100.
Huse, S. M., D. M. Welch, et al. (2010). "Ironing out the wrinkles in the rare biosphere through improved
OTU clustering." Environmental Microbiology 12(7): 1889-1898.
Lozupone, C. and R. Knight (2005). "UniFrac: a new phylogenetic method for comparing microbial
communities." Appl Environ Microbiol 71(12): 8228-8235.
Manly, B. F. J. (2005). Multivariate Statistical Methods, A Primer. Boca Raton, Chapman & Hall CRC.
McArdle, B. and M. Anderson (2001). "Fitting multivariate models to community data: a comment on
distance-based redundancy analysis." Ecology 82: 290-297.
Neufeld, J. D., H. Schafer, et al. (2007). "Stable-isotope probing implicates Methylophaga spp and novel
Gammaproteobacteria in marine methanol and methylamine metabolism." Isme Journal 1(6): 480-491.
Philippi, T. E. (1993). Multiple Regression Herbivory. Design and Analysis of Ecological Experiments.
S. M. Scheiner and J. Gurevitch. New York, Chapman & Hall CRC: 183-210.
Price, M. N., P. S. Dehal, et al. (2009). "FastTree: computing large minimum evolution trees with profiles
instead of a distance matrix." Mol Biol Evol 26(7): 1641-1650.
Quince, C., A. Lanzen, et al. (2009). "Accurate determination of microbial diversity from 454
pyrosequencing data." Nature Methods 6(9): 639-U627.
9
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
Rasmussen, P. W., D. M. Heissey, et al. (2001). Time Series Intervention Analysis: Unreplicated Large
Scale Experiments. Design and Analysis of Ecological Experiments. S. M. Scheiner and J. Gurevitch.
Oxford, Oxford University Press: 158-177.
Reeder, J. and R. Knight (2010). "Rapidly denoising pyrosequencing amplicon reads by exploiting rankabundance distributions." Nature Methods 7(9): 668-669.
Ruan, Q. S., D. Dutta, et al. (2006). "Local similarity analysis reveals unique associations among marine
bacterioplankton species and environmental factors." Bioinformatics 22(20): 2532-2538.
Shannon, P., A. Markiel, et al. (2003). "Cytoscape: A software environment for integrated models of
biomolecular interaction networks." Genome Research 13(11): 2498-2504.
Sogin, M. L., H. G. Morrison, et al. (2006). "Microbial diversity in the deep sea and the underexplored
"rare biosphere"." Proceedings of the National Academy of Sciences of the United States of America
103(32): 12115-12120.
Somerfield, P. J. (2008). "Identification of the Bray-Curtis similarity index: Comment on Yoshioka
(2008)." Marine Ecology Progress Series 372: 303-306.
Storey, J. D. (2002). "A direct approach to false discovery rates." Journal of the Royal Statistical Society
Series B-Statistical Methodology 64: 479-498.
Team, R. D. C. (2010). R: A language and environment for statistical computing. Vienna, R Foundation
for Statistical Computing.
Widdicombe, C. E., D. Eloire, et al. (2010). "Long-term phytoplankton community dynamics in the
Western English Channel." Journal of Plankton Research 32(5): 643-655.
Yilmaz, P., R. Kottmann, et al. (2010). "The Minimum Information about a Marker Gene Sequence
(MIMARKS) specification." Nature Precedings.
252
10
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
Table S1 – Sampling dates, and environmental conditions for each date. Data are MIMARKS
compliant (Yilmaz, Kottmann et al. 2010). DX1 and DX2 were calculated using the equations
outlined in the methodology. Na – data not used in analysis. Am – Ammonia (µmol L-1); Chla Chlorophyll a (µg/L); NOx – NO2+NO3 (µmol L-1); Sa – Salinity (Practical Salinity Units); Si –
Silicate (µmol L-1); SRP – Soluble Reactive Phosphate (µmol L-1); Tmp – Temperature (°C);
TOC – Total Organic Carbon (µmol L-1); TON – Total Organic Nitrogen (µmol L-1); SD – Serial
Day; DL – Day Length (hours); PAR – Photosynthetically Active Radiation (µmol
photons/m2/s); NAO – North Atlantic Oscilation; PP – Primary Productivity; dPP – Daily
Primary Productivity; MLD – Mixed Layer Depth (m).
Table S2 – Eukaryotic organism abundance measurements used in this study. H – Heterotrophic;
A – Autotrophic; M – Mixotrophic; H – Herbivores; O – Omnivores; C(B)
Carnivores(Bacteriovores); C – Carnivores; na – data not used in analysis. Ac - Acartia clausii;
Apc – Appendicularia; Bra - Brachyura; Bry - Bryozoa (Cyphonautes larvae); Cal - Calanoida;
Ch - Calanus helgolandicus; Ch(f) - Calanus helgolandicus (female); Ch(m) - Calanus
helgolandicus (male); Cet - Centropages spp.; Cty - Centropages typicus; Cha - Chaetognatha;
Cho - Chordata; Cil - CILIATES; Cin - Cirripede nauplii; Cir - Cirripedia; Cla - Cladocera; Cni Cnidaria; Coc - Coccolithophorids; Cod - Colorless Dinophycaea; CGy - Colourless Gyrodinium
sp. Med.; CGy(s) - Colourless Gyrodinium sp. Small; Cgn - Copepod eggs and nauplii; Con Copepod nauplii; Cop – Copepoda; Ca - Corycaeus anglicus; Cr - Crustacean; Crp Cryptomonad; Cyc - Cyclopoida (Oithona spp.); Del - Decapod larvae; De - Decapoda; Dia DIATOMS; Ech - Echinodermata; Echl - Echinodermata larvae; Ehux - Emiliania huxleyi; Eua Euterpina acutifrons; Eva - Evadne spp.; FE - Fish eggs; Fla2 - Flagellate 2µm; Fla FLAGELLATES; Fla5 - Flagellates 5µm; Gud - Guinardia delicatula; Ha - Harpacticoida; Hy Hydromedusae; Ka - Katodinium sp. Small; Ll - Lamellibranch larvae; Ma - Malacostraca; MiZ
- MICROZOOPLANKTON; Mo – Mollusca; Myr - Myrionecta rubra; Nic - Nitzschia
closterium; On - Oncaea spp.; Ppc - Para-Pseudo-Clauso-Ctenocalanus; Pap - Paracalanus
parvus; Pas - Paralia sulcata; Pyt - PHYTOPLANKTON; Po - Podon spp.; Poe Poecilostomatoida; Pol - Polychaeta; Pll - Polychaete larvae; Pnd - Pseudo-nitzschia
delicatissima; Pce - Pseudocalanus elongates; Sar - SARCODINA; Sct - Scripsiella trochoidea;
Sip - Siphonophorae; Stm - Stauroneis membranacea; StL - Strombidium sp. Large; StM Strombidium sp. Medium; StS - Strombidium sp. Small; Tl - Temora longicornis; Tr Torodinium cf. robustum; Uc - Unidentified Chaetognath; Uo - Unidentified Oncaea; Ur Urochordata; Zf - ZOO-FLAGELLATES.
TABLE S3 – Minimal Information about a MARKer gene Sequence (MIMARKS) compliant
data for sample information and sequencing protocol information.
Table S4 Permutation-based analysis of variance tests of differences among seasons and years
for different taxonomic groupings, using Bray-Curtis similarities calculated from Log(N+1)transformed abundances.
Table S5 Discriminant function analysis results from the first discriminant axis (DFA1) for the
bacterioplankton community sampled monthly from the English Channel. The OTU
identification given is the best consensus identification for the OTU cluster based on the ICoMM
SILVA98 taxonomy (http://vampsarchive.mbl.edu/diversity/diversity_old.php).
11
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
Table S6 Multiple Regression Analysis results for DFA1 from the bacterioplankton community
predicted by environmental and eukaryotic parameters.
Table S7. Summary of results from stepwise distance-based linear modeling of inter-sample
similarities (or distances for single variables) variables on temporal variables (serial day = D,
solstice term = DX1, equinox term = DX2) and measured or modeled variables. S – Species
richness or number of operational taxonomic units. 1-λ' is the Simpsons dominance coefficient,
1-(1- λ a') refers to the indicative evenness of the community.
Table S8 – Bacterial 16S rRNA V6 specific primers used for amplification of the V6 region of
16S rRNA gene. Lowercase base pairs indicate the 454-GS-FLX A (for forward primers) or B
(for reverse primers) adapter. Large uppercase base pairs indicate the multiplex identifier (MID)
label used to differentiate the 9 primers into 12 groups for pooling. Primers originate from Huber
et al (2007)
12
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
Figure S1 – A - Comparison of OTU abundance of raw (blue) and error-corrected sequence data
(Red – denoiser, Green – SRP) from all 72 time points in the L4 data series. Each time point was
resampled to an equal sequencing effort of 4505 sequences. The red line demonstrates OTU
abundance following the Denoiser software, while the green line demonstrates the OTU
abundance following the more conservative Single Linkage Preclustering approach. B –
Comparison of S, and diversity estimators, caho1 and ACE.
Figure S2 - Most connected cluster from a highly correlated (r > 0.7, p < 0.001, q<0.0015) subset
of the 300 most variable bacterial OTUs. The correlation network shows a loosely connected
cluster (A) and a highly connected cluster (B). The loosely connected cluster consists of OTUs
from the Bacteroidetes (light orange) and Proteobacteria phyla (blue) along with eukaryotes
(red), phytoplankton (green), clustered around environmental variables (grey). The highly
connected cluster consists entirely of bacterial OTUs including bacteroidetes, proteobacteria
(blue shades), deferribacteres (dark orange), cyanobacteria (light green). Solid lines represent
postive correlations, dashed lines represent negative correlations. Black lines show no time
delay, red arrows are delayed one month. Figure S7 contains a key for color ids.
Figure S3A - Subnetwork of correlated (r > 0.5, p < 0.005, q<0.0048) variables built around
mixotrophic eukaryotes and the variables directly connected to mixotrophic eukaryotes from the
300 most abundant bacterial OTUs. The zooplankton are red, phytoplankton are green,
environmental variables are grey, proteobacteria are blue, bacteroidetes are orange,
verrucomicrobia are purple, cyanobacteria are light green. Black lines show no time delay and
red arrows are delayed one month. Solid lines represent positive correlations; red lines represent
negative correlations. S3B - Subnetwork of correlated (r > 0.5, p < 0.005, q<0.0048) variables
built around autotrophic eukaryotes and the variables directly connected to autotrophic
eukaryotes from the 300 most abundant bacterial OTUs. The zooplankton are red, phytoplankton
are green, environmental variables are grey, proteobacteria are blue, bacteroidetes are orange,
verrucomicrobia are purple, cyanobacteria are light green. Black lines show no time delay and
red arrows are delayed one month. Solid lines represent positive correlations; red lines represent
negative correlations.
Figure S4A - Subnetwork of highly correlated (r > 0.7, p < 0.001, q<0.0008) variables built
around environmental parameters and the variables directly connected to the environmental
parameters from the 300 most variable bacterial OTUs. The zooplankton are red, phytoplankton
are green, environmental variables are grey, proteobacteria are blue, bacteroidetes are orange.
Black lines show no time delay. Red arrows are delayed one month, green arrows are delayed
two months, blue arrows are delayed 3 months. Solid lines represent positive correlations; red
lines represent negative correlations. S4B - Subnetwork of highly correlated (r > 0.7, p < 0.001,
q<0.0015) variables built around environmental parameters from the 300 most common bacterial
OTUs. The zooplankton are red, phytoplankton are green, environmental variables are grey,
proteobacteria are blue, bacteroidetes are orange. Black lines show no time delay. Red arrows are
delayed one month, green arrows are delayed two months, blue arrows are delayed 3 months.
Solid lines represent positive correlations; red lines represent negative correlations.
13
359
360
361
362
363
364
365
366
367
368
369
Appendum to Figures: Color key showing identifications for the variables used in the correlation
networks. Bacterial OTUs are defined by class, eukaryotes are divided into zooplankton and
phytoplankton in Figures 4, 5, S2-S4.
Figure S5 - (a) Alpha diversity (observed OTUs: solid line, left axis; phylogenetic diversity:
dashed line, right axis) and (b) beta diversity (unweighted UniFrac) by month spanning six years
of marine water sampling at the L4 site in the Western English Channel. A cyclic pattern is
observed in both alpha diversity, with species richness peaking in the winter months; and beta
diversity with, for example, samples from the winter (red/blue) of 2003 being more similar in
phylogenetic composition to the winter of 2008 than either is to the summer (green) of 2003.
14
Download