Skip to main content

ORIGINAL RESEARCH article

Front. Microbiol., 14 June 2018
Sec. Aquatic Microbiology

Microbial Community Structure–Function Relationships in Yaquina Bay Estuary Reveal Spatially Distinct Carbon and Nitrogen Cycling Capacities

  • 1Department of Microbiology, Oregon State University, Corvallis, OR, United States
  • 2Oak Ridge National Laboratory, Oak Ridge, TN, United States
  • 3Graduate School of Genome Science and Technology, The University of Tennessee, Knoxville, Knoxville, TN, United States
  • 4Department of Civil and Environmental Engineering, The University of Washington, Seattle, WA, United States
  • 5College of Earth, Ocean, and Atmospheric Sciences, Oregon State University, Corvallis, OR, United States
  • 6Lawrence Livermore National Laboratory, U.S. Department of Energy, Livermore, CA, United States

Linking microbial community structure to ecological processes requires understanding of the functional roles among individual populations and the factors that influence their distributions. These structure–function relationships are particularly difficult to disentangle in estuaries, due to highly variable physico-chemical conditions. Yet, examining microbe-mediated turnover of resources in these “bioreactor” ecosystems is critical for understanding estuarine ecology. In this study, a combined metagenomics and metaproteomics approach was used to show that the unequal distribution of microbial populations across the Yaquina Bay estuary led to a habitat-specific taxonomic and functional structure and a clear spatial distribution in microbe-mediated capacities for cycling of carbon and nitrogen. For example, size-fractionation revealed that communities inhabiting suspended particulate material encoded more diverse types of metabolisms (e.g., fermentation and denitrification) than those with a planktonic lifestyle, suggesting that the metabolic reactions can differ between size fractions of the same parcel of an estuarine water column. Similarly, communities inhabiting oligotrophic conditions in the lower estuary were enriched in genes involved in central carbon metabolism (e.g., TCA cycle), while communities in the upper estuary were enriched in genes typical of copiotrophic populations (e.g., cell growth, cell division). Integrating gene and protein data revealed that abundant populations of Flavobacteriales and Rhodobacterales encoded similar genomic functions, yet differed significantly in protein expression, dedicating a large proportion of their respective proteomes to rapid growth and division versus metabolic versatility and resource acquisition. This suggested potentially distinct life-strategies between these two co-occurring lineages and was concomitant with differing patterns of positive evolutionary selection on their encoded genes. Microbial communities and their functions across Yaquina Bay appear to be structured by population-level habitat preferences, resulting in spatially distinct elemental cycling, while within each community, forces such as competitive exclusion and evolutionary selection influence species life-strategies and may help maintain microbial diversity.

Introduction

Estuaries host communities of microorganisms that influence the exchange of nutrients between terrestrial, freshwater, and marine biomes (Cole and Caraco, 2001; Bauer et al., 2013). These boundary ecosystems often exhibit strong spatiotemporal gradients in salinity, turbidity, and resource availability, resulting in microbial communities with highly variable characteristics across these conditions (Crump et al., 2004; Jeffries et al., 2016). Biotic and abiotic dynamism is particularly evident in estuaries during winter along the Oregon coast when heavy overland precipitation and steep nearshore topography cause coastal river flooding events that transport significant amounts of suspended particulate matter into estuarine water columns (Hickey and Banas, 2003; Hastings et al., 2012; Goñi et al., 2013).

Microbes colonizing such particles represent an important component of aquatic biogeochemical cycles by acting to liberate particulate carbon and nutrients to the planktonic microbial loop through decomposition processes, and by supplementing higher trophic levels of the food web through resulting biomass production (Goulder, 1977; Wainright, 1990; Brown and Ozretich, 2009; Stocker, 2012). Studies examining microbial communities in Oregon estuaries have observed significant differences in the phylogenetic composition and respiratory activity between particle-associated and free-living microbial communities and across marine–estuarine–riverine gradients (Crump et al., 1998, 1999, 2017; Smith et al., 2013). Beyond phylogenetic characterization, however, the functional roles of microbial communities and their relative influence on ecological processes occurring in different habitats within heterogeneous estuarine systems have seldom been examined (Simon et al., 2014).

Applying a combination of high-throughput DNA sequencing and metaproteomics techniques, we investigated the phylogenetic and functional structure of microbial communities with two contrasting, operationally defined lifestyles (particle-associated and free-living) in two spatially separated locations (upper and lower) of Yaquina Bay, Oregon, United States. Using these data, we estimated how the mechanisms of microbe-mediated carbon and nitrogen turnover differed between the four habitats. This was achieved using previously identified metabolic marker genes and proteins, whose relative abundances provide a proxy for different components of aquatic microbe-mediated biogeochemical cycles (Lauro et al., 2011; Llorens-Marès et al., 2015). We hypothesized that habitat preference (i.e., a distribution biased toward a specific habitat) at the population level would lead to communities with distinct taxonomic and functional structures, corresponding with habitat-specific carbon and nitrogen cycling on both large (kilometers, between locations) and small (microns, between particle-attached and free-living) spatial scales (Hyndes et al., 2014) in the Yaquina Bay estuary.

To understand the ecology of successful microorganisms in this ecosystem, we also reconstructed the encoded genomes and expressed proteomes at the time of sampling from 15 populations in two dominant lineages (Flavobacteriales and Rhodobacterales) to determine whether their life-strategy – i.e., the set of traits used for survival, growth, and reproduction (Barnard et al., 2013) – contributed to their pervasiveness across habitats. To determine if life-strategy of the two taxa was consequential for their coexistence and codominance (i.e., whether they displayed overlapping or divergent life-strategies and, thus, occupied a similar or different niche space), we compared the relative abundances of their expressed proteins and their protein-coding genes. Similarly, we examined positive evolutionary selection in these two lineages using classical analyses of synonymous and non-synonymous mutation rates on protein-coding genes to determine whether evolutionary pressure on each lineage was related to its life-strategy.

Materials and Methods

Sample Collection and Processing

Surface water samples were collected on November 22, 2014 from two sites in Yaquina Bay (OR, United States). Samples were taken from the top 1 m of the water column at a tidal height of ∼1.8 m near slack current (-0.1 m/s) in both the riverine end-member of the estuary (hereafter, “upper”: 44.58°N, -123.99°W) and the coastal estuary mouth (hereafter, “lower”: 44.62°N, -124.04°W), which are separated by 8 km (Supplementary Figure S1). Along with tidal influence, Yaquina River flow was ∼17 m3/s (Oregon Water Resources Department, Station ID 14306030), resulting in upper and lower estuary sampling sites that represent differing salinity and nutrient zones during our sampling period (Brown and Ozretich, 2009; Shafer et al., 2016). Triplicate 12 L water samples were collected in acid-washed carboys and filtered through in-line 3-μm and 0.22-μm PES membranes (Pall Corporation, Ann Arbor, MI, United States) to recover operationally defined particle-associated and free-living microbial communities, respectively. Collected biomass was stored at -80°C until processing.

Community DNA was extracted using the CTAB protocol described in (Doherty et al., 2017). Total protein was extracted using SDS lysis method detailed in (Bryson et al., 2016, 2017). Extractions from 1 and 11 L of water yielded sufficient quantities of analytes for further analyses (1–5 μg DNA and 150–300 μg of protein, respectively).

16S Amplicon Library Construction and Analysis

The v4 locus of the 16S rRNA gene was amplified and sequenced on the Illumina MiSeq (Illumina, Inc., San Diego, CA, United States) using the two-step PCR protocol of the Nextera XT index kit. First-step PCRs were performed with 30 cycles using ∼50 ng of DNA, AccuStart II PCR ToughMix Polymerase (QuantaBio, Beverly, MA, United States) following manufacturer instructions, and primers encoding universal prokaryotic v4-complement sequences (515F and 806R) (Takahashi et al., 2014). AMPure XP beads were used for product purification. Second-step PCRs were performed per Nextera XT kit instructions.

Libraries were normalized by concentration, pooled, and sequenced using an Illumina MiSeq v2 kit (251 bp, paired-end reads), generating >17,000 reads per library. Reads were trimmed and quality filtered using FastQC and Sickle (Joshi and Fass, 2011). Low quality (<25 phred scores across 15 base sliding windows) sequences within reads were trimmed and >200 bp reads were retained. The “mothur” program was used to remove adapter regions and for read-pair assembly, unsupervised OTU clustering (97% identity), taxonomic assignment with the RDP classifier, rarefaction, relative abundance calculations, and OTU table generation (Schloss et al., 2009).

Metagenome Sequencing, Assembly, and Annotation

Metagenome libraries were prepared using Nextera XT (Illumina, Inc., San Diego, CA, United States) and Wafergen (Wafergen Bio-systems, Fremont, CA, United States) kits. Libraries were quantified with Bioanalyzer HS-DNA Chips, normalized by concentration, and pooled. Metagenome sequencing was performed on Illumina HiSeq 3000 using 151 bp long, paired-end reads and yielded >16 Gb per library (>32 M reads). Metagenome reads were filtered as defined for 16S amplicon reads, with the exception that the length threshold was 100 bases. IDBA-UD (Peng et al., 2013) was used for de novo metagenome assembly of the combined read sets from all three replicate community DNA samples in each of the four habitats (i.e., upper and lower estuary communities with both particle-associated and free-living lifestyles). All 16S-amplicon and metagenome sequences are available from NCBI (BioSample SAMN04917373, BioProject PRJNA320136, short read archive SRS1422236 with accession numbers SRX1738750-90 for 16S-amplicon reads, and SRX1738728-32 for metagenome reads).

Predicted coding sequences (CDS) on contigs were determined using Prodigal (Hyatt et al., 2012). Taxonomic assignment of CDS was based on best hits to the RefSeq database (downloaded June 14, 2014) using the Diamond algorithm in “blastp” mode (Buchfink et al., 2015). Diamond blastp was chosen for annotation to the Class taxonomic level based on speed and sensitivity tests comparing the algorithm with traditional blastp and Diamond blastp in sensitive mode (Supplementary Figure S2). The GhostKoala server was used to functionally annotate each CDS by assigning a KEGG Orthology (KO) number (Kanehisa et al., 2016). Additional functional assignments of CDS were based on best rpsBLAST hits to protein families from the Pfam (Finn et al., 2016) and cluster of orthologous genes (COG) (Huerta-Cepas et al., 2016) databases.

Read depths of contigs/scaffolds and of CDS were defined by separately aligning reads from all libraries against each of the four resulting assembles using the Bowtie2 program (Langmead and Salzberg, 2012). The relative abundances of CDS were approximated using the reads per kilobase of genome equivalents (RPKG) metric, which was calculated with “MicrobeCensus.” This metric uses CDS read-depth and a genome equivalent metric based on metagenome library size and single-copy marker genes in order to normalize relative abundances of CDS across metagenomes (Nayfach and Pollard, 2015).

Contigs from each metagenome were automatically binned into organism-specific sets using the “maxbin2.1” program, which uses contig read-depths and tetranucleotide (TN) frequency data for bin assignment (Wu et al., 2016). More than 600 total bins were auto-generated. The lineage-specific workflow of “CheckM” (Parks et al., 2015) was used to assess initial bin quality by calculating completeness and contamination metrics. Bins with >85% completeness (83 bins; bolded typeface in Supplementary Table S1) were manually curated using “Vizbin” (Laczny et al., 2015) and “mmgenome” (Karst et al., 2016). Outlying contigs were identified in plots of GC content, TN frequency, and read-depths and removed. Resulting bins were reassessed in “CheckM,” and those with >85% completeness and <10% contamination were analyzed further (40 bins; bolded and italicized typeface in Supplementary Table S1). Taxonomy of manually curated bins was determined by BLAST-based annotation of binned CDS, by comparison to the “amphora2” marker gene set (Wu and Scott, 2012), and with “CheckM” rerun with the taxon-specific workflow. Thirty-nine bins with >85% completeness and <10% contamination were retained for further population-level analyses.

Peptide Mass Spectrometry, Database Construction, and Protein Identification

Extracted proteins were digested with trypsin at room temperature (enzyme:substrate ratio of 1:100 w:w). Twenty-five micrograms of peptides were analyzed for each of the 12 metaproteomes using 2D-Liquid Chromatography Tandem Mass Spectrometry (LC-MS/MS) with previously described conditions (Bryson et al., 2016). LC-MS/MS measurements were performed on an LTQ Orbitrap Elite mass spectrometer (Thermo Scientific, Waltham, MA, United States) (Washburn et al., 2001). MS and MS/MS scans were acquired with a resolution of 30,000 and 15,000, respectively, and the 10 most abundant precursor ions were selected for MS/MS analysis by higher-energy collisional dissociation.

Sipros Ensemble (Guo et al., 2017) was used to search all MS spectra against a peptide database constructed from all metagenome CDS clustered at 100% identity using cd-hit (Li and Godzik, 2006), and with reverse sequences of all peptides added as decoys to calculate false discovery rate (FDR) (Bryson et al., 2016). Searches considered 7–60 residue peptides with a maximum of two missed tryptic cuts. The two-peptide rule – one unique plus one shared peptide, or two unique peptides – was used to determine confident protein identifications, and a 1% FDR at the peptide level was implemented as in (Bryson et al., 2016). Normalized balanced spectral counts (NBSC) of peptides for identified proteins were used as measures of protein relative abundances across all samples (Supplementary Table S2). The proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE (Vizcaíno et al., 2016) partner repository with the dataset identifier PXD008093.

DNA Sequence Analyses

The relative abundances of KO, Pfam, and COG protein families within each metagenome were approximated based on the cumulative RPKG of all CDS assigned to a given protein family. The relative abundance of COG categories was calculated by summing RPKG for all CDS assigned to COG numbers in each COG category.

To infer carbon and nitrogen cycling potential of each microbial community, relative abundances of metabolic marker genes (KO numbers) defined in previous studies (Lauro et al., 2011; Llorens-Marès et al., 2015; Vavourakis et al., 2016) were calculated using marker KO RPKG ratios. A total of 40 marker genes were used in the analysis, representing 10 microbe-mediated elemental cycling reactions (Supplementary Table S3). Metabolic potential for each metagenome was calculated as in (Lauro et al., 2011) and averaged across replicates. The relative abundance of proteins (in NBSC) encoded by marker KO genes were similarly analyzed. Statistically significant differences in the relative abundances of element cycling processes between habitats were assessed using ANOVA with the Tukey–Kramer post hoc test and Benjamin-Hochberg FDR controlled at 5% using the “STAMP” software package (Parks et al., 2014).

OTU relative abundances based on 16S amplicon sequencing data were summed at Order-level taxonomic groupings for community diversity analyses. Biplots of principal component analyses (PCA) examining the beta-diversity of communities were produced using the “pca3d” library in R (Weiner, 2017). Weighted Unifrac calculations, PCA, Mantel tests, Procrustes analyses, and permutational ANOVA tests were performed in R using the “vegan” package (Oksanen et al., 2017). Faith’s (1992) phylogenetic alpha-diversity was calculated using the “picante” package and significant differences between samples were defined using pairwise Welch’s t-tests. R scripts and input data for statistical calculations can be found in Supplementary Table S2.

Significantly discriminating taxonomic and functional features between communities were identified using the linear discriminant analysis (LDA) method implemented in “LEfSe” (Segata et al., 2011). “Enrichment” of a feature refers to one with an LDA score >2.0 (log 10) using the strict all-against-all multi-class analysis and a p < 0.05 cutoff for the among- and between-class LEfSe tests. LEfSe, by default, performs significance testing on normalized relative abundances. Thus, a correction variable was added to input files when the feature abundance was in RPKG, which is based on a unique genome equivalent metric calculated for each sample (Nayfach, personal communication).

Positive Selection Detection on Groups of Orthologous Proteins

ProteinOrtho was used to predict orthologous protein groups (OGs) shared between populations (Lechner et al., 2011). OGs were scanned for evidence of horizontal transfer using GENECONV, and those with significant hits were not considered in further analyses (Sawyer, 1989; Posada and Crandall, 2001). The “ETE3” toolkit (Huerta-Cepas et al., 2016) was used to discover evidence of positive selection on OGs using Codeml and PAML. Clustal Omega with trimAL and RaxML were implemented for multiple codon sequence alignments and tree-building, respectively (Sievers et al., 2011; Stamatakis, 2014). Trees were visualized with the Interactive Tree of Life server (Letunic and Bork, 2016).

Models M1 vs. M2, M7 vs. M8, and M8 vs. M8a were used as sets of null and alternative models, respectively, for defining a threshold of significance for positive selection (Yang et al., 2000). A significant difference in dN/dS ratio (omega) based on likelihood ratio tests (LRT) of model outcomes with the BEB posterior probability method was used to infer the best model fit for a given OG alignment (Yang et al., 2005). LRT p-values from all tests were corrected by controlling at a FDR of 5%. If an OG had a corrected q-value < 0.05 under any of the three models, this gene group was considered to be under putative positive selection (e.g., as in Tang and Zhang, 2007).

Results

Community Taxonomic Structure and Diversity

Amplicon sequencing of 16S v4 loci and clustering at 97% identity yielded 6,792 OTUs from ∼240,000 total paired-end reads (>17,000 paired-end reads per replicate) across all four sampled habitat communities (Supplementary Table S4). The OTU-level Unifrac distance matrix and Order-level Euclidean distance matrix calculated from OTU counts yielded highly similar community beta-diversity patterns (Mantel’s test, Pearson r = 0.962, p < 0.001), and PCA likewise resulted in similar data shape between these two metrics (Procrustes symmetric correlation = 0.982, p < 0.001). Therefore, community structure analysis was conducted using Euclidean PCA and biplotting of Order-level taxa relative abundances across samples to determine which groups most strongly drove differences in community structure (Figure 1). PCA indicated that 23.4% of community structure variance was explained by sampling location and 20.1% by lifestyle (Figure 1). However, only location significantly differentiated overall sample grouping, while lifestyle did not (Permutational ANOVA, perms = 999; df = 11; location: p = 0.002, R2 = 0.54; lifestyle: p = 0.14, R2 = 0.85). In contrast, community alpha-diversity was not different between the upper and lower estuary communities (p = 0.45). Rather, communities in the particle-associated lifestyle were significantly more diverse than their free-living counterparts both overall (t-test, p < 0.0001), and within each location (each p < 0.01; Supplementary Figure S3).

FIGURE 1
www.frontiersin.org

FIGURE 1. Euclidean distance PCA with feature biplots of taxonomic community structure. Plotted data are relative abundance of 16S OTUs grouped at Order-level taxa. Sample triplicates per habitat (points) are plotted alongside a set of the most important Orders driving sample separation (vectors). Arrow direction and magnitude represent both the covariance between features (Orders) and their effect on PC loadings. Ellipses represent 95% confidence intervals for centroid positions. FL, free-living; PA, particle-attached.

Population-Level Habitat Preferences

Of the 6,792 OTUs recovered, 157 (2.3%) were overrepresented in one of the four estuary habitats based on significant relative abundance changes across samples using linear discriminant analysis (LDA score >2). Despite being a minor subset of all OTUs, these populations represented 75.5% of all sampled 16S reads, indicating that relatively high-abundance populations exhibited habitat preference (Supplementary Table S5).

The three most abundant Orders in all samples were Pelagibacterales, Flavobacteriales, and Rhodobacterales, collectively representing 39–54% of total 16S reads (Supplementary Figure S4). The relative habitat preferences of several populations in these three lineages influenced divergent habitat taxonomic structures (Figure 1). For example, the SAR11 clade member HTCC1062 (Pelagibacterales) was cumulatively the most abundant Yaquina Bay organism (OTUs 1, 14, 55) and exhibited clear habitat preference for the free-living fraction of the marine-influenced lower estuary (each OTU LDA score >3.2; Supplementary Table S5). In the Flavobacteriales and Rhodobacterales lineages, notable populations with significant habitat preference included a Cellulophaga (OTU 2; Flavobacteriales) in the free-living lifestyle of the upper estuary (LDA = 4.8) and a Thalassobius (OTU 4; Rhodobacterales) in the particulate-associated lifestyle of the upper estuary (LDA = 4.5).

An archaeal population of Nitrosopumilales (OTU 30), which represented >98% of estuary Thaumarchaeota, was strongly associated with the free-living fraction of the lower estuary (LDA score = 3.6), causing this group to be as influential as Pelagibacterales in driving unique taxonomic structure of the habitat despite its lower relative abundance (Figure 1). Likewise, OTUs belonging to lineages that typically inhabit anaerobic or microaerophilic environments (e.g., Desulfovibrionales) were relatively rare community members (cumulatively 0.5–2.5%; Supplementary Figure S5), but their enrichment in the particle-associated habitats drove distinct alpha-diversity (Supplementary Figure S4) and beta-diversity structure (Figure 1) of this lifestyle in each estuary location.

Community Functional Structure

Assembly of ∼62 Gbp shotgun metagenome DNA sequence yielded ∼1 M contigs (1.8 Gbp) and ∼2.5 M predicted CDS across all four sampled habitats (Supplementary Table S4). Between 71 and 78% of CDS were assigned to a taxonomic group and between 32 and 58% were annotated to COG, KEGG, or Pfam protein families. Functional structure based on KEGG-annotated CDS relative abundance was significantly correlated to taxonomic structure inferred from 16S analyses (Mantel’s test, Pearson r = 0.832; p < 0.0001), leading to the similar ordination pattern (Figures 1, 2).

FIGURE 2
www.frontiersin.org

FIGURE 2. Euclidean distance PCA of functional community structure. Plotted data are KO RPKG relative abundance across samples based on metagenome CDS annotation. Ellipses represent 95% confidence intervals for centroid positions. FL, free-living; PA, particle-attached.

Of the 4,764 KEGG protein families identified across all metagenomes, 432 (9.1%) were biased in their distribution between samples (Supplementary Table S6). Several functional pathways showed clear patterns of habitat enrichment based on the distribution of KEGG families (Figure 2 and Supplementary Figure S6), including photosynthesis in the particulate-associated fractions (Supplementary Figure S6A), ABC transport in the free-living fractions, flagellar motility in the upper estuary (Supplementary Figure S6C), TCA cycle and glycolysis in the lower estuary (Supplementary Figures S6D,E), and ribosomal and transcription in the upper estuary (Supplementary Figures S6F,G).

Community Carbon and Nitrogen Cycling Capacities

Out of 40 marker genes for 13 biogeochemical flux processes, 31 genes and 10 processes were identified in our metagenome assemblies (Supplementary Table S3). Most missing genes were involved in processes not expected to occur at detectable levels in aerobic estuarine surface water (anammox and methanogenesis). Of the 31 metagenome-encoded marker genes, 21 proteins were identified in metaproteomes (Supplementary Table S3). The relative abundance of marker genes and marker proteins within each habitat were correlated, indicating that the gene abundance for C- and N-cycling processes was generally a useful indicator of protein abundance within our samples (Figure 3).

FIGURE 3
www.frontiersin.org

FIGURE 3. Major processes within the (A) C- or (B) N-cycles and relative abundance of biological marker genes that catalyze fluxes. Arrow sizes show relative abundance of total marker gene RPKG for each process. Bar plots show marker genes (left-axis) and lines plots show marker proteins (right-axis). Dotted arrows represent a process with no recovered marker genes or proteins. Axis labels and scales are equivalent for all plots and only drawn on one plot for clarity. Significantly non-equal potential across habitats was determined with a FDR-corrected ANOVA: q < 0.05). Error bars represent standard error of the mean for each set of replicate samples. MP, metaproteome; MG, metagenome in ANOVA results.

As expected, the encoded genetic capacity (metagenome) for aerobic C respiration is likely the dominant pathway of estuary C flux and was equal across all habitats (FDR-corrected ANOVA, q = 0.18; Figure 3A). Aerobic C fixation potential, on the other hand, was significantly higher in the particle-associated fractions of both locations, and marker gene relative abundance for anaerobic C fixation and fermentation were overrepresented in particle-attached community of the lower estuary (Figure 3A). Notably, carbon monoxide (CO) oxidation genes were highly enriched in upper estuary samples, and over 68% of CO oxidation capacity was annotated to the Roseobacter clade (e.g., Roseovarius, Ruegeria).

With regard to N-cycling processes, the encoded capacity for N assimilation is likely the dominant pathway of estuary N flux and was equal across all habitats (FDR-corrected ANOVA, q = 0.13; Figure 3B). Importantly, though, encoded community capacity for N mineralization appeared to be highest in the particle-attached lifestyles in each location. Marker gene and marker protein relative abundances suggested enrichment in ammonification and denitrification potential in the marine-influenced lower estuary, particularly in the particle-associated communities (Figure 3B). On the other hand, marker gene abundance for N fixation was significantly enriched in the upper estuary, though detected proteins did not show this trend.

A comparison of the taxonomic distribution of 16S rRNA reads, metagenome CDS, and the N or C marker genes showed that many taxa appear to contribute disproportionately to community biogeochemical cycling capacity compared to their relative abundance in the community (Supplementary Figure S7). These taxa and processes included ammonification in Epsilon-proteobacteria (ammonification: 85%, 16S: 0.4%), fermentation in Aquificae and Planctomycetia (fermentation: 63%, 16S: 0.02%; fermentation: 9%, 16S: 0.7%), N fixation in Alpha-proteobacteria (N fixation: 81.3%, 16S: 15.3%), and aerobic C fixation in Eukaryota (C fixation: 10%, 16S: 1.4%).

Population-Level Life-Strategies

Flavobacteriales and Rhodobacterales were ubiquitous, though unevenly distributed (Figure 1), across all samples; their combined pangenome (>450,000 CDS) and panproteome (>4,500 proteins) represented 24.7 and 31.4% of total sampled genes and proteins, respectively (Figure 4). Fifteen well-curated population Flavobacteriales and Rhodobacterales genomes were defined from the metagenome contig binning and curation process, while bins from the other relatively abundant taxon, Pelagibacterales, could not be easily resolved (data not shown). This was reflected in 16S v4 diversity; of OTUs over 0.1% relative abundance, three were in Pelagibacterales, while 18 were Flavobacteriales and 10 were Rhodobacterales. The robust dataset of genes, proteins, and curated genomes made Flavobacteriales and Rhodobacterales lineages good candidates to test whether co-dominance across the estuary was corollary to divergent strategies for growth and resource acquisition, which is a hypothesis that has been explored for co-dominant microbial populations in other ecosystems (e.g., Violle et al., 2011).

FIGURE 4
www.frontiersin.org

FIGURE 4. Comparison of (A) pangenome and (B) panproteome content of Rhodobacterales (Rb) and Flavobacteriales (Fb). Plots show cumulative relative abundance of lineage-specific (A) CDS and (B) proteins annotated to COG numbers and grouped at COG category level (represented by letters as points on each plot). Input and results of correlation analysis from genomic and proteomic data are shown above each plot. The inset in (B) is an expansion of the origin region (0–0.02 on x-axis and 0–0.015 on the y-axis). COG categories are colored according to their classification: blue, information storage and processing; green, metabolism; red, cellular processes and signaling; black, unclassified. Dotted lines indicate a 1:1 relationship. COG categories shown but not labeled are: D, cell cycle control, cell division, chromosome partitioning; M, cell wall/membrane/envelope biogenesis; N, cell motility; O, post-translational modification, protein turnover, and chaperones; T, signal transduction mechanisms; U, intracellular trafficking, secretion, and vesicular transport; V, defense mechanisms; B, chromatin structure and dynamics; L, replication, recombination and repair; C, energy production and conversion; F, nucleotide transport and metabolism; I, lipid transport and metabolism; R, general function prediction only; S, function unknown.

While the Flavobacteriales and Rhodobacterales pangenomes, as determined by the relative abundance of all metagenome CDS annotated to each group, were correlated at the conservative functional classification level of COG category (Pearson’s ρ = 0.81), their panproteomes, as determined by the relative abundance of all proteins annotated to each group, were weakly correlated at this broad resolution (ρ = -0.07), indicating that proteins of different functions were being expressed by each group during sampling (Figure 4).

The panproteome of Rhodobacterales populations was relatively enriched for metabolism functions, with over 50% of proteins annotated to amino acid (E), carbohydrate (G), nucleotide (F), and secondary metabolite (Q) transport and metabolism COG categories (Figure 4). In contrast, Flavobacteriales populations devoted just 6.8% of proteins to these four COG categories and were instead enriched in translation (J, 27.4%) and transcription (K, 16.7%) functions relative to Rhodobacterales (7.4 and 2.1%, respectively). Flavobacteriales inorganic ion transport and metabolism proteins was also highly disproportionate (P, 27.0%) relative to Rhodobacterales (3.1%), though the majority of these proteins (>93%, predominately SusC and SusD) are involved in the Bacteroidetes starch utilization system (Reintjes et al., 2017).

Evolutionary Pressure on Life-Strategies

The 10 Flavobacteriales and five Rhodobacterales population genomes recovered by metagenome contig binning had an average completeness of 93.84% and an average contamination of 2.68% (Supplementary Table S7). In total, 73 and 138 single-copy, orthologous gene groups (OGs) were shared in the Flavobacteriales and Rhodobacterales populations, respectively. A concatenated amino acid alignment of these OGs was used to estimate the phylogenetic relationship between populations in each lineage (Supplementary Figure S8), and positive selection on OGs was determined based on the classic dN/dS substitution ratio (Yang et al., 2000). Positive selection was observed in seven (9.6%) Flavobacteriales and 10 (7.2%) Rhodobacterales OGs (Table 1). Model statistics are reported in Supplementary Tables S8, S9. Similar rates have been reported in other non-pathogenic, or host-dependent, microbial populations (e.g., 4.5% in Nandi et al., 2010; 9% in Bulgarelli et al., 2015).

TABLE 1
www.frontiersin.org

TABLE 1. Genes in the Rhodobacterales and Flavobacteriales lineages that are under putative positive selection.

In Rhodobacterales, over half of genes under selection were related to resource transport and metabolism (Table 1). The positively selected genes glnG, mgtE, phoU, and aapM code for proteins that regulate or directly facilitate the import and utilization of metabolic substrates. In addition, gene sequences of the metabolic enzymes chlN and argF, as well as a predicted sulfurtransferase, thioesterase, and amidase, each were under positive selection.

In contrast, positively selected genes in the Flavobacteriales were typically assigned to cell growth, repair, and replication functions (Table 1). A histidyl-tRNA synthetase gene (hisS), a DNA replication and repair gene (recR), and two outer-membrane biogenesis and protein folding chaperones (hlpA and ftsI) were each under positive selection. The positively selected gene purB encodes the enzyme adenylosuccinate synthetase, which is involved in purine nucleotide biosynthesis during cellular replication and division.

Discussion

The taxonomic structure of estuarine bacterioplankton communities is known to vary across salinity gradients (e.g., Crump et al., 1999; Smith et al., 2013) and between particle-attached and free-living fractions (e.g., D’Ambrosio et al., 2014; Zhang et al., 2016), but little is known about the physiological capabilities that results from this variability. Our study expands on this previous work by quantifying the functional capabilities and activities of taxa present under four distinct spatially separated and size-fractioned habitat conditions in the Yaquina Bay estuary.

Using amplicon sequencing of the v4 region of the 16S gene, we determined that microbial community-level taxonomic structure in Yaquina Bay was distinct between the upper and lower estuary margins and the free-living and particle-attached lifestyles in each location. Ordinations of these patterns in bacterial diversity based on 16S amplicon sequencing changed little using OTU-level or Order-level annotations and were each similar to taxonomic structure as determined by metagenome CDS sequence annotation (Supplementary Figure S9). Overall, we found patterns in the distributions of microbial populations that were similar to those in other temperate estuaries (e.g., Bouvier and del Giorgio, 2002; Ortmann and Santos, 2016).

The Relative Abundance of Key Taxonomic Groups Distinguished the Four Sampled Habitats

Two highly abundant lineages, the Flavobacteriales and Rhodobacterales, contained 11 and 19 OTU-level populations that exhibited distinct habitat preferences. Both lineages had multiple OTUs enriched in each of the four habitats, resulting in no clear consensus of Order-level habitat preference. However, their two most abundant populations, a Cellulophaga (Flavobacteriales) and a Thalassobius (Rhodobacterales), were each relatively enriched in the upper estuary, leading to the significant influence of Flavobacteriales and Rhodobacterales in driving upper estuary communities apart from those inhabiting the lower estuary.

Pelagibacterales were also highly abundant in the estuary, but unlike Flavobacteriales and Rhodobacterales, the three Pelagibacterales OTUs exhibited strong habitat preferences, each being highly enriched in the planktonic fraction of the marine-influenced lower estuary margin and being most depleted in the particle-associated lifestyle of the upper estuary. These populations, which were all classified as SAR11 clade members, were most closely related to coastal (Ortmann and Santos, 2016), rather than brackish ecotypes (Smith et al., 2013), suggesting that they may have entered the estuary by tidal intrusion. This distribution of Pelagibacterales is consistent with the well-characterized planktonic lifestyle of SAR11 Subclade I (Giovannoni et al., 1990) and with its observed decrease in its relative abundance across another marine-estuary-river continuum (Ortmann and Santos, 2016).

Nitrosopumilales enrichment in the marine-influenced lower estuary (Hewson et al., 2014; Hugerth et al., 2015) and Burkholderiales in the riverine-influenced upper estuary (Bouvier and del Giorgio, 2002; Silveira et al., 2011) also corroborate previous findings of microbial population habitat range. Similarly, the high relative abundance of anaerobic or microaerophilic populations on suspended particulates is a common observation in estuarine environments (Crump et al., 1999; Waidner and Kirchman, 2007; Zhang et al., 2016), and may be the result of oxygen-depleted niches in particle biofilms (Dang et al., 2011), or suspension of colonized grains from the anoxic sediment by physical mixing (Crump and Baross, 1996; Baker et al., 2015). The specialist populations in particle-associated communities contributed to a significantly higher phylogenetic and metabolic diversity in the community associated with these samples, which is consistent with studies of suspended particulate material in other aquatic ecosystems (Smith et al., 2013; Dang and Lovell, 2015; Yung et al., 2016).

Overall, the results from community taxonomic structure analyses substantiate those from previous studies showing that both lateral gradients (Crump et al., 2004; Henriques et al., 2004) and filtration fraction of the water-column (Crump et al., 1999; Zhang et al., 2016) are significant predictors of microbial community structure in estuaries.

Distinct Taxonomic Structure of Communities in Each Habitat Was Correlated With Distinct Functional Structure

We hypothesized that the distinct community taxonomic structure of each habitat would correspond to distinct functional structure, leading to differences in the capacity of each community to perform ecological processes such as carbon and nitrogen turnover. If supported, this habitat-specific ecological function may indicate how biogeochemical cycling in the estuarine water-column is spatially organized and provide a more complete understanding of how resources move through this ecosystem. If no significant functional differences were evident between habitats, it would suggest that communities encode functional redundancy despite their distinct taxonomic structure, leading to a uniform spatial distribution of biogeochemical cycling capacity and a dissociation between microbial population diversity and functional diversity (Allison and Martiny, 2008; Delgado-Baquerizo et al., 2016).

To determine community functional structure, we sequenced metagenomes of all samples and measured relative abundances of protein-coding gene families. Community functional structure was significantly correlated with taxonomic structure, leading to similar ordination shapes of OTU and KEGG relative abundance data. Lower estuary communities were enriched in central carbon metabolism functions (i.e., glycolysis and TCA cycle), while upper estuary communities were enriched in cell growth, transcription, and translation genes (Supplementary Figure S6). This result tracked with the distribution of microbial taxa with different trophic strategies: genomes of copiotrophic populations, such as those in the Rhodobacterales and Flavobacteriales lineages that were enriched in the upper estuary, typically encode relatively more transcriptional and ribosomal genes in their larger genomes, whereas oligotrophs, such as Pelagibacterales that were enriched in the lower estuary, often encode streamlined genomes with higher proportions of genes involved in resource transport and metabolism (Cottrell and Kirchman, 2016). Similar observations have been made across physico-chemical gradients in the brackish Baltic Sea, where key pathways and core metabolic processes were organized spatially by salinity gradients (Dupont et al., 2014).

Similarly, the relative enrichment of photosynthetic genes in particle-associated communities was related to the increased relative abundance of Cyanobacteria and chloroplast-like 16S v4 rRNA sequences in this size-fraction. In contrast, ABC transporters were overrepresented in free-living communities, which may be related to the enrichment of planktonic prokaryotes (e.g., SAR11, whose genome has a high density of transporter functions) in this water-column filtration fraction.

Enrichment for flagellar motility genes in the upper estuary differentiated the functional structure of communities in this margin from those in the lower estuary, which again may reflect the biased distribution of copiotrophic and oligotrophic lineages. The single flagellar motility gene that was overrepresented in the lower rather than upper estuary was an archaeal flagella subunit (K07332). This was expected considering that all 7 archaeal OTUs with significant habitat preferences were enriched in the marine-influenced lower estuary margin (OTUs 6, 30, 35, 42, 78, 161, 256).

Communities Differed in Capacities for Carbon and Nitrogen Transformation

Based on these significant biases in the distributions of microbial community function across habitats, we hypothesized that microbe-mediated biogeochemical flux would be habitat-specific, leading to the spatial distribution of carbon- (C) and nitrogen- (N) cycling processes in the estuary. To test this hypothesis, we used metagenome and metaproteome data to estimate the relative abundance of marker genes and proteins involved in central C and N utilization pathways within each habitat community. Relative abundance ratios of these 31 functional markers were used to infer their capacity for catalyzing 10 major C- and N-cycling processes.

Out of the 31 metagenome marker genes, we detected 21 marker proteins in our samples, reflecting the difficulty in capturing total protein diversity from complex communities and the low recovery of protein fragments such as membrane-spanning domains (Williams and Cavicchioli, 2014). Regardless, all 10 C- and N-cycling processes we examined were represented by at least one marker protein, and the relative abundances of marker proteins generally tracked those of the marker genes that encoded them.

The two most abundant metabolic pathways, aerobic C respiration and N assimilation, had equal capacity across all habitat metagenomes and metaproteomes, suggesting that differences in taxonomic composition did not affect the capacity of the community to perform these functions. On the other hand, the relative abundance of markers for processes requiring specific environmental conditions was not conserved across habitats. For example, denitrification and fermentation genes and proteins were relatively more abundant in the particle-associated communities, especially in the lower estuary margin. The marker genes for these processes were predominately annotated to facultative or strictly anaerobic OTUs, which correspondingly were enriched in the particle-attached communities of both the upper and lower estuary, indicating that these metabolic specialists could be biogeochemically important members of the particle-associated communities (Jørgensen, 1977; Etcheber et al., 2007; Garnier et al., 2010; Zhang et al., 2016).

The relative abundance of encoded marker genes and expressed marker proteins for CO oxidation suggested that this process has the potential to contribute significantly to the C cycle of Yaquina Bay. CO oxidation capacity was highest in the upper estuary communities, and the majority (68%) of marker genes for this process were annotated to Roseobacter populations. Many OTUs in this clade were significantly enriched in the upper estuary (e.g., OTU 15, Litoreibacter; OTU 4, Thalassobius), and indeed Roseobacters have been proposed to be the major drivers of biotic CO oxidation in coastal ecosystems (Moran et al., 2004; Tolli et al., 2006). Estuarine CO is primarily produced through the photolysis or oxidation of particulate matter (Miller et al., 2002; Song et al., 2015), and the Yaquina Bay estuary and Yaquina river are known to be supersaturated in CO throughout the year (Butler et al., 1987). Thus, we hypothesize that a steady supply of CO from decaying particulate organic matter entering Yaquina Bay by riverine transport is consumed by Roseobacters in the upper estuary. Given the relative magnitude of this biogeochemical process based on marker gene and protein abundance, CO oxidation may be a key characteristic of upper estuary ecology, and changes in Roseobacter abundance or CO production may significantly change the flux of C from Yaquina Bay.

Unexpectedly, we did not detect KO marker genes for nitrification (K10944, K10945, or K10946), despite finding that the Nitrosopumilales lineage represented 0.6% of all 16S amplicons. When considering sample metagenomes, however, the relative abundance of genes annotated to these nitrifiers was an order of magnitude lower than 16S relative abundance (0.06%) and most recovered Nitrosopumilales genes were of housekeeping or unknown function (data not shown). This result indicates that the absence of nitrification markers in our samples may have been due to under-sampling of the genetic diversity in the system, rather true biological absence.

Importantly, our results highlight potential keystone organisms in the microbial food webs of the estuary surface water. For example, populations of Epsilon-proteobacteria (predominately of Sulfurospirillum) represented an ammonification marker gene abundance that was highly disproportional to their relative community abundance. Significant contribution to biogeochemical cycling by such populations in low-abundance lineages has been supported by studies examining rare biosphere activity (Campbell et al., 2011).

Microbial Populations Appear to Use Varying Strategies to Succeed in the Estuarine Environment

To understand what characteristics may be allowing microorganisms to coexist across these four distinct habitats, we examined the pangenome and panproteome of the Flavobacteriales and Rhodobacterales lineages. The co-dominance of these lineages is a common observation in estuarine ecosystems (Dong et al., 2014; Colatriano et al., 2015), and they have been shown in experimental and observational studies to grow and acquire resources using different strategies (Alonso-Sáez and Gasol, 2007; Poretsky et al., 2010; Bryson et al., 2016, 2017). Thus, we hypothesized that their ability to coexist across the estuary may be due to the use of different strategies for growth and reproduction (Hibbing et al., 2010). This hypothesis is based on the phylogenetic limiting theory, which states that when populations minimize their overlap in life-strategy, deleterious interactions can be reduced between them, curtailing forces such as competitive exclusion that drive species with similar life-strategies out of a single habitat (Violle et al., 2011).

Our results supported the divergence in life-strategies of Flavobacteriales and Rhodobacterales, which we inferred based on the significantly different types of proteins expressed by each lineage at the time of sampling. Rhodobacterales populations dedicated a relatively larger part of their proteome to metabolic versatility and resource acquisition, while Flavobacteriales were enriched in transcription/translation and specialized starch transport functions. Similar patterns of protein-level expression and de-novo protein production have been recorded previously in these lineages using stable-isotope-probing experiments in Monterey Bay, in which Rhodobacterales accounted for the highest total amount of substrate assimilation (glucose, amino acids, peptides), while Flavobacteriales bloomed under all substrates but were significantly isotope-enriched only in a starch treatment (Bryson et al., 2017).

Interestingly, the types of functions under positive evolutionary selection in each lineage, which we determined using high-quality population genomes acquired from metagenome binning, appeared to be those that were related their respective life-strategies. This suggests that evolutionary selection may be a mechanism that maintains lineage life-strategies over time (Violle et al., 2011), allowing for the stable coexistence of Flavobacteriales and Rhodobacterales across the habitats we sampled. Indeed, the relationship between life-strategy and evolutionary selection has been proposed to contribute significantly to present-day observations of microbial function and community assembly (Luo and Moran, 2015). However, because our sampling was designed to focus on spatial rather than temporal investigation, it will be necessary to verify our interpretations over time (e.g., seasonal) or in other estuarine or coastal ecosystems.

Taken together, results from the investigation of Order-level pangenomes and panproteomes showed that the highly abundant and co-occurring Flavobacteriales and Rhodobacterales lineages were differentiated from one another not by broad functional groups that were encoded at the genome level, but by functions they expressed at the protein level. Results from the examination of evolutionary selection on genes from highly resolved, population-level genomes in each lineage showed that positive selection on OGs encoded in Flavobacteriales and Rhodobacterales appeared to be acting on gene sequences with the types of functions that confer the life-strategy of both groups (i.e., genes in Table 1 were biased toward growth and replication functions for Flavobacteriales and metabolism-related functions in Rhodobacterales). This is the first such example of a connection between life-strategy and evolutionary selection in estuarine microorganisms, though a link between functions under positive selection and functions that are important in defining the life-strategy of microbial populations has been observed previously, including studies of deep sea bacteria (Campanaro et al., 2008) and extremophilic archaea (Gunbin et al., 2009).

Data Availability

All 16S-amplicon and metagenome sequences are available from NCBI (BioSample SAMN04917373, BioProject PRJNA320136, short read archive SRS1422236 with accession numbers SRX1738750-90 for 16S-amplicon reads and SRX1738728-32 for metagenome reads). The metaproteome protein data are available from the PRIDE archive under ID PXD008093.

Author Contributions

BK, SB, XM, RH, CP, and RM designed the experiments. BK, SB, XM, and RM coordinated and carried out the sampling procedures. BK, ZL, RH, CP, and RM performed the analytical procedures and acquired data. All authors contributed to discussion of results. ZL, RH, CP, and RM contributed reagents, materials, and technology. BK and RM designed and prepared figures and tables and drafted the original manuscript. All authors contributed critical revisions during the editing process, and all authors approved the final manuscript draft for submission.

Funding

This work was funded by the Gordon and Betty Moore Foundation Marine Microbiology Initiative (Grant Id: GBMF3302). Part of this work was performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344. Oak Ridge National Laboratory resources, including the Oak Ridge Leadership Computing Facility, were used in the research and are supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC05-00OR22725. BC was supported by National Science Foundation grant OCE-0424602.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank the Oregon State University Center for Genomic Research and Biocomputing and for their support in data collection and analysis. Hatfield Marine Science Center staff and facilities were instrumental in sample collection and processing. We also thank the reviewers of the manuscript, whose comments and suggestions significantly improved the presentation of this research.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fmicb.2018.01282/full#supplementary-material

References

Allison, S. D., and Martiny, J. B. H. (2008). Resistance, resilience, and redundancy in microbial communities. Proc. Natl. Acad. Sci. U.S.A. 105, 11512–11519. doi: 10.1073/pnas.0801925105

PubMed Abstract | CrossRef Full Text | Google Scholar

Alonso-Sáez, L., and Gasol, J. M. (2007). Seasonal variations in the contributions of different bacterial groups to the uptake of low-molecular-weight compounds in northwestern mediterranean coastal waters. Appl. Environ. Microbiol. 73, 3528–3535. doi: 10.1128/AEM.02627-06

PubMed Abstract | CrossRef Full Text | Google Scholar

Baker, B. J., Lazar, C. S., Teske, A. P., and Dick, G. J. (2015). Genomic resolution of linkages in carbon, nitrogen, and sulfur cycling among widespread estuary sediment bacteria. Microbiome 3:14. doi: 10.1186/s40168-015-0077-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Barnard, R. L., Osborne, C. A., and Firestone, M. K. (2013). Responses of soil bacterial and fungal communities to extreme desiccation and rewetting. ISME J. 7, 2229–2241. doi: 10.1038/ismej.2013.104

PubMed Abstract | CrossRef Full Text | Google Scholar

Bauer, J. E., Cai, W.-J., Raymond, P. A., Bianchi, T. S., Hopkinson, C. S., and Regnier, P. A. G. (2013). The changing carbon cycle of the coastal ocean. Nature 504, 61–70. doi: 10.1038/nature12857

PubMed Abstract | CrossRef Full Text | Google Scholar

Bouvier, T. C., and del Giorgio, P. A. (2002). Compositional changes in free-living bacterial communities along a salinity gradient in two temperate estuaries. Limnol. Oceanogr. 47, 453–470. doi: 10.4319/lo.2002.47.2.0453

CrossRef Full Text | Google Scholar

Brown, C. A., and Ozretich, R. J. (2009). Coupling between the coastal ocean and yaquina bay: Oregon: importance of oceanic inputs relative to other nitrogen sources. Estuaries Coasts 32, 219–237. doi: 10.1007/s12237-008-9128-6

CrossRef Full Text | Google Scholar

Bryson, S., Li, Z., Chavez, F., Weber, P. K., Pett-Ridge, J., Hettich, R. L., et al. (2017). Phylogenetically conserved resource partitioning in the coastal microbial loop. ISME J. 11, 2781–2792. doi: 10.1038/ismej.2017.128

PubMed Abstract | CrossRef Full Text | Google Scholar

Bryson, S., Li, Z., Pett-Ridge, J., Hettich, R. L., Mayali, X., Pan, C., et al. (2016). Proteomic stable isotope probing reveals taxonomically distinct patterns in amino acid assimilation by coastal marine bacterioplankton. mSystems 1:e00027-15. doi: 10.1128/mSystems.00027-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Buchfink, B., Xie, C., and Huson, D. H. (2015). Fast and sensitive protein alignment using DIAMOND. Nat. Methods 12, 59–60. doi: 10.1038/nmeth.3176

PubMed Abstract | CrossRef Full Text | Google Scholar

Bulgarelli, D., Garrido-Oter, R., Münch, P. C., Weiman, A., Dröge, J., Pan, Y., et al. (2015). Structure and function of the bacterial root microbiota in wild and domesticated barley. Cell Host Microbe 17, 392–403. doi: 10.1016/j.chom.2015.01.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Butler, J. H., Jones, R. D., Garber, J. H., and Gordon, L. I. (1987). Seasonal distributions and turnover of reduced trace gases and hydroxylamine in Yaquina Bay, Oregon. Geochim. Cosmochim. Acta 51, 697–706. doi: 10.1016/0016-7037(87)90080-9

CrossRef Full Text | Google Scholar

Campanaro, S., Treu, L., and Valle, G. (2008). Protein evolution in deep sea bacteria: an analysis of amino acids substitution rates. BMC Evol. Biol. 8:313. doi: 10.1186/1471-2148-8-313

PubMed Abstract | CrossRef Full Text | Google Scholar

Campbell, B. J., Yu, L., Heidelberg, J. F., and Kirchman, D. L. (2011). Activity of abundant and rare bacteria in a coastal ocean. Proc. Natl. Acad. Sci. U.S.A. 108, 12776–12781. doi: 10.1073/pnas.1101405108

PubMed Abstract | CrossRef Full Text | Google Scholar

Colatriano, D., Ramachandran, A., Yergeau, E., Maranger, R., Gélinas, Y., and Walsh, D. A. (2015). Metaproteomics of aquatic microbial communities in a deep and stratified estuary. Proteomics 15, 3566–3579. doi: 10.1002/pmic.201500079

PubMed Abstract | CrossRef Full Text | Google Scholar

Cole, J. J., and Caraco, N. F. (2001). Carbon in catchments: connecting terrestrial carbon losses with aquatic metabolism. Mar. Freshw. Res. 52, 101–110. doi: 10.1071/MF00084

CrossRef Full Text | Google Scholar

Cottrell, M. T., and Kirchman, D. L. (2016). Transcriptional control in marine copiotrophic and oligotrophic bacteria with streamlined genomes. Appl. Environ. Microbiol. 82, 6010–6018. doi: 10.1128/AEM.01299-16

PubMed Abstract | CrossRef Full Text | Google Scholar

Crump, B. C., Armbrust, E. V., and Baross, J. A. (1999). Phylogenetic analysis of particle-attached and free-living bacterial communities in the Columbia river, its estuary, and the adjacent coastal ocean. Appl. Environ. Microbiol. 65, 3192–3204.

PubMed Abstract | Google Scholar

Crump, B. C., and Baross, J. A. (1996). Particle-attached bacteria and heterotrophic plankton associated with the Columbia River estuarine turbidity maxima. Mar. Ecol. Prog. Ser. 138, 265–273. doi: 10.3354/meps138265

CrossRef Full Text | Google Scholar

Crump, B. C., Baross, J. A., and Simenstad, C. A. (1998). Dominance of particle-attached bacteria in the Columbia River estuary, USA. Aquat. Microb. Ecol. 14, 7–18. doi: 10.3354/ame014007

CrossRef Full Text | Google Scholar

Crump, B. C., Fine, L. M., Fortunato, C. S., Herfort, L., Needoba, J. A., Murdock, S., et al. (2017). Quantity and quality of particulate organic matter controls bacterial production in the Columbia River estuary. Limnol. Oceanogr. 62, 2713–2731. doi: 10.1002/lno.10601

CrossRef Full Text | Google Scholar

Crump, B. C., Hopkinson, C. S., Sogin, M. L., and Hobbie, J. E. (2004). Microbial biogeography along an estuarine salinity gradient: combined influences of bacterial growth and residence time. Appl. Environ. Microbiol. 70, 1494–1505. doi: 10.1128/AEM.70.3.1494-1505.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

D’Ambrosio, L., Ziervogel, K., MacGregor, B., Teske, A., and Arnosti, C. (2014). Composition and enzymatic function of particle-associated and free-living bacteria: a coastal/offshore comparison. ISME J. 8, 2167–2179. doi: 10.1038/ismej.2014.67

PubMed Abstract | CrossRef Full Text | Google Scholar

Dang, H., Chen, R., Wang, L., Shao, S., Dai, L., Ye, Y., et al. (2011). Molecular characterization of putative biocorroding microbiota with a novel niche detection of Epsilon- and Zetaproteobacteria in Pacific Ocean coastal seawaters. Environ. Microbiol. 13, 3059–3074. doi: 10.1111/j.1462-2920.2011.02583.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Dang, H., and Lovell, C. R. (2015). Microbial surface colonization and biofilm development in marine environments. Microbiol. Mol. Biol. Rev. 80, 91–138. doi: 10.1128/MMBR.00037-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Delgado-Baquerizo, M., Giaramida, L., Reich, P. B., Khachane, A. N., Hamonts, K., Edwards, C., et al. (2016). Lack of functional redundancy in the relationship between microbial diversity and ecosystem functioning. J. Ecol. 104, 936–946. doi: 10.1111/1365-2745.12585

CrossRef Full Text | Google Scholar

Doherty, M., Yager, P. L., Moran, M. A., Coles, V. J., Fortunato, C. S., Krusche, A. V., et al. (2017). Bacterial biogeography across the Amazon river-ocean continuum. Front. Microbiol. 8:882. doi: 10.3389/fmicb.2017.00882

PubMed Abstract | CrossRef Full Text | Google Scholar

Dong, H.-P., Hong, Y.-G., Lu, S., and Xie, L.-Y. (2014). Metaproteomics reveals the major microbial players and their biogeochemical functions in a productive coastal system in the northern South China Sea. Environ. Microbiol. Rep. 6, 683–695. doi: 10.1111/1758-2229.12188

PubMed Abstract | CrossRef Full Text | Google Scholar

Dupont, C. L., Larsson, J., Yooseph, S., Ininbergs, K., Goll, J., Asplund-Samuelsson, J., et al. (2014). Functional tradeoffs underpin salinity-driven divergence in microbial community composition. PLoS One 9:e89549. doi: 10.1371/journal.pone.0089549

PubMed Abstract | CrossRef Full Text | Google Scholar

Etcheber, H., Taillez, A., Abril, G., Garnier, J., Servais, P., Moatar, F., et al. (2007). Particulate organic carbon in the estuarine turbidity maxima of the Gironde, Loire and Seine estuaries: origin and lability. Hydrobiologia 588, 245–259. doi: 10.1007/s10750-007-0667-9

CrossRef Full Text | Google Scholar

Faith, D. P. (1992). Conservation evaluation and phylogenetic diversity. Biol. Conserv. 61, 1–10. doi: 10.1016/0006-3207(92)91201-3

CrossRef Full Text | Google Scholar

Finn, R. D., Coggill, P., Eberhardt, R. Y., Eddy, S. R., Mistry, J., Mitchell, A. L., et al. (2016). The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 44, D279–D285. doi: 10.1093/nar/gkv1344

PubMed Abstract | CrossRef Full Text | Google Scholar

Garnier, J., Billen, G., Némery, J., and Sebilo, M. (2010). Transformations of nutrients (N, P, Si) in the turbidity maximum zone of the Seine estuary and export to the sea. Estuarine Coast. Shelf Sci. 90, 129–141. doi: 10.1016/j.ecss.2010.07.012

CrossRef Full Text | Google Scholar

Giovannoni, S. J., Britschgi, T. B., Moyer, C. L., and Field, K. G. (1990). Genetic diversity in Sargasso Sea bacterioplankton. Nature 345, 60–63. doi: 10.1038/345060a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Goñi, M. A., Hatten, J. A., Wheatcroft, R. A., and Borgeld, J. C. (2013). Particulate organic matter export by two contrasting small mountainous rivers from the Pacific Northwest, U.S.A. J. Geophys. Res. Biogeosci. 118, 112–134. doi: 10.1002/jgrg.20024

CrossRef Full Text | Google Scholar

Goulder, R. (1977). Attached and free bacteria in an estuary with abundant suspended solids. J. Appl. Bacteriol. 43, 399–405. doi: 10.1111/j.1365-2672.1977.tb00766.x

CrossRef Full Text | Google Scholar

Gunbin, K. V., Afonnikov, D. A., and Kolchanov, N. A. (2009). Molecular evolution of the hyperthermophilic archaea of the Pyrococcus genus: analysis of adaptation to different environmental conditions. BMC Genomics 10:639. doi: 10.1186/1471-2164-10-639

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, X., Li, Z., Yao, Q., Mueller, R. S., Eng, J. K., Tabb, D. L., et al. (2017). Sipros ensemble improves database searching and filtering for complex metaproteomics. Bioinformatics 34, 795–802. doi: 10.1093/bioinformatics/btx601

PubMed Abstract | CrossRef Full Text | Google Scholar

Hastings, R. H., Goñi, M. A., Wheatcroft, R. A., and Borgeld, J. C. (2012). A terrestrial organic matter depocenter on a high-energy margin: the Umpqua River system, Oregon. Cont. Shelf Res. 3, 78–91. doi: 10.1016/j.csr.2012.04.002

CrossRef Full Text | Google Scholar

Henriques, I. S., Almeida, A., Cunha,Â., and Correia, A. (2004). Molecular sequence analysis of prokaryotic diversity in the middle and outer sections of the Portuguese estuary Ria de Aveiro. FEMS Microbiol. Ecol. 49, 269–279. doi: 10.1016/j.femsec.2004.04.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Hewson, I., Eggleston, E. M., Doherty, M., Lee, D. Y., Owens, M., Shapleigh, J. P., et al. (2014). Metatranscriptomic analyses of plankton communities inhabiting surface and subpycnocline waters of the chesapeake bay during oxic-anoxic-oxic transitions. Appl. Environ. Microbiol. 80, 328–338. doi: 10.1128/AEM.02680-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Hibbing, M. E., Fuqua, C., Parsek, M. R., and Peterson, S. B. (2010). Bacterial competition: surviving and thriving in the microbial jungle. Nat. Rev. Microbiol. 8, 15–25. doi: 10.1038/nrmicro2259

PubMed Abstract | CrossRef Full Text | Google Scholar

Hickey, B. M., and Banas, N. S. (2003). Oceanography of the U. S. Pacific Northwest coastal ocean and estuaries with application to coastal ecology. Estuaries 26, 1010–1031. doi: 10.1007/BF02803360

CrossRef Full Text | Google Scholar

Huerta-Cepas, J., Szklarczyk, D., Forslund, K., Cook, H., Heller, D., Walter, M. C., et al. (2016). eggNOG 4.5: a hierarchical orthology framework with improved functional annotations for eukaryotic, prokaryotic and viral sequences. Nucleic Acids Res. 44, D286–D293. doi: 10.1093/nar/gkv1248

PubMed Abstract | CrossRef Full Text | Google Scholar

Hugerth, L. W., Larsson, J., Alneberg, J., Lindh, M. V., Legrand, C., Pinhassi, J., et al. (2015). Metagenome-assembled genomes uncover a global brackish microbiome. Genome Biol. 16:279. doi: 10.1186/s13059-015-0834-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Hyatt, D., LoCascio, P. F., Hauser, L. J., and Uberbacher, E. C. (2012). Gene and translation initiation site prediction in metagenomic sequences. Bioinformatics 28, 2223–2230. doi: 10.1093/bioinformatics/bts429

PubMed Abstract | CrossRef Full Text | Google Scholar

Hyndes, G. A., Nagelkerken, I., McLeod, R. J., Connolly, R. M., Lavery, P. S., and Vanderklift, M. A. (2014). Mechanisms and ecological role of carbon transfer within coastal seascapes. Biol. Rev. 89, 232–254. doi: 10.1111/brv.12055

PubMed Abstract | CrossRef Full Text | Google Scholar

Jeffries, T. C., Schmitz Fontes, M. L., Harrison, D. P., Van-Dongen-Vogels, V., Eyre, B. D., Ralph, P. J., et al. (2016). Bacterioplankton dynamics within a large anthropogenically impacted Urban estuary. Front. Microbiol. 6:1438. doi: 10.3389/fmicb.2015.01438

PubMed Abstract | CrossRef Full Text | Google Scholar

Jørgensen, B. B. (1977). Bacterial sulfate reduction within reduced microniches of oxidized marine sediments. Mar. Biol. 41, 7–17. doi: 10.1007/BF00390576

CrossRef Full Text | Google Scholar

Joshi, N., and Fass, J. (2011). Sickle: A Sliding-Window, Adaptive, Quality-Based Trimming Tool for FastQ Files (Version 1.33) [Software]. Available at: https://github.com/najoshi/sickle

Google Scholar

Kanehisa, M., Sato, Y., and Morishima, K. (2016). BlastKOALA and GhostKOALA: KEGG tools for functional characterization of genome and metagenome sequences. J. Mol. Biol. 428, 726–731. doi: 10.1016/j.jmb.2015.11.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Karst, S. M., Kirkegaard, R. H., and Albertsen, M. (2016). mmgenome: a toolbox for reproducible genome extraction from metagenomes. bioRxiv [Preprint]. doi: 10.1101/059121

CrossRef Full Text

Laczny, C. C., Sternal, T., Plugaru, V., Gawron, P., Atashpendar, A., Margossian, H. H., et al. (2015). VizBin - an application for reference-independent visualization and human-augmented binning of metagenomic data. Microbiome 3:1. doi: 10.1186/s40168-014-0066-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Langmead, B., and Salzberg, S. L. (2012). Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359. doi: 10.1038/nmeth.1923

PubMed Abstract | CrossRef Full Text | Google Scholar

Lauro, F. M., DeMaere, M. Z., Yau, S., Brown, M. V., Ng, C., Wilkins, D., et al. (2011). An integrative study of a meromictic lake ecosystem in Antarctica. ISME J. 5, 879–895. doi: 10.1038/ismej.2010.185

PubMed Abstract | CrossRef Full Text | Google Scholar

Lechner, M., Findeiß, S., Steiner, L., Marz, M., Stadler, P. F., and Prohaska, S. J. (2011). Proteinortho: detection of (Co-)orthologs in large-scale analysis. BMC Bioinformatics 12:124. doi: 10.1186/1471-2105-12-124

PubMed Abstract | CrossRef Full Text | Google Scholar

Letunic, I., and Bork, P. (2016). Interactive tree of life (iTOL) v3: an online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 44, W242–W245. doi: 10.1093/nar/gkw290

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, W., and Godzik, A. (2006). Cd-hit: a fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics 22, 1658–1659. doi: 10.1093/bioinformatics/btl158

PubMed Abstract | CrossRef Full Text | Google Scholar

Llorens-Marès, T., Yooseph, S., Goll, J., Hoffman, J., Vila-Costa, M., Borrego, C. M., et al. (2015). Connecting biodiversity and potential functional role in modern euxinic environments by microbial metagenomics. ISME J. 9, 1648–1661. doi: 10.1038/ismej.2014.254

PubMed Abstract | CrossRef Full Text | Google Scholar

Luo, H., and Moran, M. A. (2015). How do divergent ecological strategies emerge among marine bacterioplankton lineages? Trends Microbiol. 23, 577–584. doi: 10.1016/j.tim.2015.05.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, W. L., Moran, M., Sheldon, W. M., Zepp, R. G., and Opsahl, S. (2002). Determination of apparent quantum yield spectra for the formation of biologically labile photoproducts. Limnol. Oceanogr. 47, 343–352. doi: 10.4319/lo.2002.47.2.0343

CrossRef Full Text | Google Scholar

Moran, M. A., Buchan, A., González, J. M., Heidelberg, J. F., Whitman, W. B., Kiene, R. P., et al. (2004). Genome sequence of Silicibacter pomeroyi reveals adaptations to the marine environment. Nature 432, 910–913. doi: 10.1038/nature03170

PubMed Abstract | CrossRef Full Text | Google Scholar

Nandi, T., Ong, C., Singh, A. P., Boddey, J., Atkins, T., Sarkar-Tyson, M., et al. (2010). A genomic survey of positive selection in Burkholderia pseudomallei provides insights into the evolution of accidental virulence. PLoS Pathog. 6:e1000845. doi: 10.1371/journal.ppat.1000845

PubMed Abstract | CrossRef Full Text | Google Scholar

Nayfach, S., and Pollard, K. S. (2015). Average genome size estimation improves comparative metagenomics and sheds light on the functional ecology of the human microbiome. Genome Biol. 16:51. doi: 10.1186/s13059-015-0611-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., McGlinn, D., et al. (2017). vegan: Community Ecology Package. Available at: https://cran.r-project.org/web/packages/vegan/index.html

Google Scholar

Ortmann, A. C., and Santos, T. T. L. (2016). Spatial and temporal patterns in the Pelagibacteraceae across an estuarine gradient. FEMS Microbiol. Ecol. 92:fiw133. doi: 10.1093/femsec/fiw133

PubMed Abstract | CrossRef Full Text | Google Scholar

Parks, D. H., Imelfort, M., Skennerton, C. T., Hugenholtz, P., and Tyson, G. W. (2015). CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 25, 1043–1055. doi: 10.1101/gr.186072.114

PubMed Abstract | CrossRef Full Text | Google Scholar

Parks, D. H., Tyson, G. W., Hugenholtz, P., and Beiko, R. G. (2014). STAMP: statistical analysis of taxonomic and functional profiles. Bioinformatics 30, 3123–3124. doi: 10.1093/bioinformatics/btu494

PubMed Abstract | CrossRef Full Text | Google Scholar

Peng, Y., Leung, H. C. M., Yiu, S.-M., Lv, M.-J., Zhu, X.-G., and Chin, F. Y. L. (2013). IDBA-tran: a more robust de novo de Bruijn graph assembler for transcriptomes with uneven expression levels. Bioinformatics 29, i326–i334. doi: 10.1093/bioinformatics/btt219

PubMed Abstract | CrossRef Full Text | Google Scholar

Poretsky, R. S., Sun, S., Mou, X., and Moran, M. A. (2010). Transporter genes expressed by coastal bacterioplankton in response to dissolved organic carbon. Environ. Microbiol. 12, 616–627. doi: 10.1111/j.1462-2920.2009.02102.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Posada, D., and Crandall, K. A. (2001). Evaluation of methods for detecting recombination from DNA sequences: computer simulations. Proc. Natl. Acad. Sci. U.S.A. 98, 13757–13762. doi: 10.1073/pnas.241370698

PubMed Abstract | CrossRef Full Text | Google Scholar

Reintjes, G., Arnosti, C., Fuchs, B. M., and Amann, R. (2017). An alternative polysaccharide uptake mechanism of marine bacteria. ISME J. 11, 1640–1650. doi: 10.1038/ismej.2017.26

PubMed Abstract | CrossRef Full Text | Google Scholar

Sawyer, S. (1989). Statistical tests for detecting gene conversion. Mol. Biol. Evol. 6, 526–538. doi: 10.1093/oxfordjournals.molbev.a040567

PubMed Abstract | CrossRef Full Text | Google Scholar

Schloss, P. D., Westcott, S. L., Ryabin, T., Hall, J. R., Hartmann, M., Hollister, E. B., et al. (2009). Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl. Environ. Microbiol. 75, 7537–7541. doi: 10.1128/AEM.01541-09

PubMed Abstract | CrossRef Full Text | Google Scholar

Segata, N., Izard, J., Waldron, L., Gevers, D., Miropolsky, L., Garrett, W. S., et al. (2011). Metagenomic biomarker discovery and explanation. Genome Biol. 12:R60. doi: 10.1186/gb-2011-12-6-r60

PubMed Abstract | CrossRef Full Text | Google Scholar

Shafer, D., Swannack, T., Saltus, S., Kaldy, J., and Davis, A. (2016). Development and validation of a habitat suitability model for the non-indigenous seagrass Zostera japonica in North America. Manage. Biol. Invasions 7, 141–155. doi: 10.3391/mbi.2016.7.2.02

CrossRef Full Text | Google Scholar

Sievers, F., Wilm, A., Dineen, D., Gibson, T. J., Karplus, K., Li, W., et al. (2011). Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 7, 539. doi: 10.1038/msb.2011.75

PubMed Abstract | CrossRef Full Text | Google Scholar

Silveira, C. B., Vieira, R. P., Cardoso, A. M., Paranhos, R., Albano, R. M., and Martins, O. B. (2011). Influence of salinity on bacterioplankton communities from the Brazilian rain forest to the coastal Atlantic ocean. PLoS One 6:e17789. doi: 10.1371/journal.pone.0017789

PubMed Abstract | CrossRef Full Text | Google Scholar

Simon, H. M., Smith, M. W., and Herfort, L. (2014). Metagenomic insights into particles and their associated microbiota in a coastal margin ecosystem. Front. Microbiol. 5:466. doi: 10.3389/fmicb.2014.00466

PubMed Abstract | CrossRef Full Text | Google Scholar

Smith, M. W., Zeigler Allen, L., Allen, A. E., Herfort, L., and Simon, H. M. (2013). Contrasting genomic properties of free-living and particle-attached microbial assemblages within a coastal ecosystem. Front. Microbiol. 4:120. doi: 10.3389/fmicb.2013.00120

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, G., Richardson, J. D., Werner, J. P., Xie, H., and Kieber, D. J. (2015). Carbon monoxide photoproduction from particles and solutes in the delaware estuary under contrasting hydrological conditions. Environ. Sci. Technol. 49, 14048–14056. doi: 10.1021/acs.est.5b02630

PubMed Abstract | CrossRef Full Text | Google Scholar

Stamatakis, A. (2014). RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313. doi: 10.1093/bioinformatics/btu033

PubMed Abstract | CrossRef Full Text | Google Scholar

Stocker, R. (2012). Marine microbes see a Sea of gradients. Science 338, 628–633. doi: 10.1126/science.1208929

PubMed Abstract | CrossRef Full Text | Google Scholar

Takahashi, S., Tomita, J., Nishioka, K., Hisada, T., and Nishijima, M. (2014). Development of a prokaryotic universal primer for simultaneous analysis of bacteria and archaea using next-generation sequencing. PLoS One 9:e105592. doi: 10.1371/journal.pone.0105592

PubMed Abstract | CrossRef Full Text | Google Scholar

Tang, F., and Zhang, C. (2007). Evidence for positive selection on the E2 gene of bovine viral diarrhoea virus type 1. Virus Genes 35, 629–634. doi: 10.1007/s11262-007-0122-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Tolli, J. D., Sievert, S. M., and Taylor, C. D. (2006). Unexpected diversity of bacteria capable of carbon monoxide oxidation in a coastal marine environment, and contribution of the roseobacter-associated clade to total CO oxidation. Appl. Environ. Microbiol. 72, 1966–1973. doi: 10.1128/AEM.72.3.1966-1973.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Vavourakis, C. D., Ghai, R., Rodriguez-Valera, F., Sorokin, D. Y., Tringe, S. G., Hugenholtz, P., et al. (2016). Metagenomic insights into the uncultured diversity and physiology of microbes in four hypersaline Soda lake brines. Front. Microbiol. 7:211. doi: 10.3389/fmicb.2016.00211

PubMed Abstract | CrossRef Full Text | Google Scholar

Violle, C., Nemergut, D. R., Pu, Z., and Jiang, L. (2011). Phylogenetic limiting similarity and competitive exclusion: phylogenetic relatedness and competition. Ecol. Lett. 14, 782–787. doi: 10.1111/j.1461-0248.2011.01644.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Vizcaíno, J. A., Csordas, A., del-Toro, N., Dianes, J. A., Griss, J., Lavidas, I., et al. (2016). 2016 update of the PRIDE database and its related tools. Nucleic Acids Res. 44, D447–D456. doi: 10.1093/nar/gkv1145

PubMed Abstract | CrossRef Full Text | Google Scholar

Waidner, L. A., and Kirchman, D. L. (2007). Aerobic anoxygenic phototrophic bacteria attached to particles in turbid waters of the delaware and chesapeake estuaries. Appl. Environ. Microbiol. 73, 3936–3944. doi: 10.1128/AEM.00592-07

PubMed Abstract | CrossRef Full Text | Google Scholar

Wainright, S. C. (1990). Sediment-to-water fluxes of particulate material and microbes by resuspension and their contribution to the planktonic food web. Mar. Ecol. Prog. Ser. 62, 271–281. doi: 10.3354/meps062271

CrossRef Full Text | Google Scholar

Washburn, M. P., Wolters, D., and Yates, J. R. III (2001). Large-scale analysis of the yeast proteome by multidimensional protein identification technology. Nat. Biotechnol. 19, 242–247. doi: 10.1038/85686

PubMed Abstract | CrossRef Full Text | Google Scholar

Weiner, J. (2017). pca3d: Three Dimensional PCA Plots. R Package Version 0.8. Available at: https://cran.r-project.org/web/packages/pca3d/index.html

Google Scholar

Williams, T. J., and Cavicchioli, R. (2014). Marine metaproteomics: deciphering the microbial metabolic food web. Trends Microbiol. 22, 248–260. doi: 10.1016/j.tim.2014.03.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, M., and Scott, A. J. (2012). Phylogenomic analysis of bacterial and archaeal sequences with AMPHORA2. Bioinformatics 28, 1033–1034. doi: 10.1093/bioinformatics/bts079

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, Y. W., Simmons, B. A., and Singer, S. (2016). MaxBin 2.0: an automated binning algorithm to recover genomes from multiple metagenomic datasets. Bioinformatics 32, 605–607. doi: 10.1093/bioinformatics/btv638

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, Z., Nielsen, R., Goldman, N., and Pedersen, A.-M. (2000). Codon-substitution models for heterogeneous selection pressure at amino acid sites. Genetics 155, 431–449.

Google Scholar

Yang, Z., Wong, W. S. W., and Nielsen, R. (2005). Bayes Empirical Bayes inference of amino acid sites under positive selection. Mol. Biol. Evol. 22, 1107–1118. doi: 10.1093/molbev/msi097

PubMed Abstract | CrossRef Full Text | Google Scholar

Yung, C.-M., Ward, C. S., Davis, K. M., Johnson, Z. I., and Hunt, D. E. (2016). Diverse and temporally-variable particle-associated microbial communities are insensitive to bulk seawater environmental parameters. Appl. Environ. Microbiol. 82, 3431–3437. doi: 10.1128/AEM.00395-16

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Xiao, W., and Jiao, N. (2016). Linking biochemical properties of particles to particle-attached and free-living bacterial community structure along the particle density gradient from freshwater to open ocean. J. Geophys. Res. Biogeosci. 121, 2261–2274. doi: 10.1002/2016JG003390

CrossRef Full Text | Google Scholar

Keywords: metagenomics, metaproteomics, estuary, biogeochemical cycling, Yaquina Bay, free-living, particle-attached

Citation: Kieft B, Li Z, Bryson S, Crump BC, Hettich R, Pan C, Mayali X and Mueller RS (2018) Microbial Community Structure–Function Relationships in Yaquina Bay Estuary Reveal Spatially Distinct Carbon and Nitrogen Cycling Capacities. Front. Microbiol. 9:1282. doi: 10.3389/fmicb.2018.01282

Received: 09 March 2018; Accepted: 25 May 2018;
Published: 14 June 2018.

Edited by:

Lasse Riemann, University of Copenhagen, Denmark

Reviewed by:

Sara Beier, Leibniz Institute for Baltic Sea Research Warnemü (LG), Germany
Markus V. Lindh, Lund University, Sweden
Julie Dinasquet, University of California, San Diego, United States

Copyright © 2018 Kieft, Li, Bryson, Crump, Hettich, Pan, Mayali and Mueller. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Brandon Kieft, kieft1bp@gmail.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.