Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Microbial Communities in Sunken Wood Are Structured by Wood-Boring Bivalves and Location in a Submarine Canyon

  • Sonja K. Fagervold,

    Affiliations Sorbonne Universités, UPMC, Univ Paris 06, UMR8222, LECOB, Observatoire Océanologique, Banyuls-sur-Mer, France, CNRS, UMR 8222, Laboratoire d'Écogéochimie des Environnements Benthiques (LECOB), Banyuls-sur-Mer, France

  • Chiara Romano,

    Affiliation Centre d'Estudis Avançats de Blanes (CEAB-CSIC), Blanes (Girona), Catalunya, Spain

  • Dimitri Kalenitchenko,

    Affiliations Sorbonne Universités, UPMC, Univ Paris 06, UMR8222, LECOB, Observatoire Océanologique, Banyuls-sur-Mer, France, CNRS, UMR 8222, Laboratoire d'Écogéochimie des Environnements Benthiques (LECOB), Banyuls-sur-Mer, France

  • Christian Borowski,

    Affiliation Max Planck Institute for Marine Microbiology, Bremen, Germany

  • Amandine Nunes-Jorge,

    Affiliation Max Planck Institute for Marine Microbiology, Bremen, Germany

  • Daniel Martin,

    Affiliation Centre d'Estudis Avançats de Blanes (CEAB-CSIC), Blanes (Girona), Catalunya, Spain

  • Pierre E. Galand

    pierre.galand@obs-banyuls.fr

    Affiliations Sorbonne Universités, UPMC, Univ Paris 06, UMR8222, LECOB, Observatoire Océanologique, Banyuls-sur-Mer, France, CNRS, UMR 8222, Laboratoire d'Écogéochimie des Environnements Benthiques (LECOB), Banyuls-sur-Mer, France

Abstract

The cornerstones of sunken wood ecosystems are microorganisms involved in cellulose degradation. These can either be free-living microorganisms in the wood matrix or symbiotic bacteria associated with wood-boring bivalves such as emblematic species of Xylophaga, the most common deep-sea woodborer. Here we use experimentally submerged pine wood, placed in and outside the Mediterranean submarine Blanes Canyon, to compare the microbial communities on the wood, in fecal pellets of Xylophaga spp. and associated with the gills of these animals. Analyses based on tag pyrosequencing of the 16S rRNA bacterial gene showed that sunken wood contained three distinct microbial communities. Wood and pellet communities were different from each other suggesting that Xylophaga spp. create new microbial niches by excreting fecal pellets into their burrows. In turn, gills of Xylophaga spp. contain potential bacterial symbionts, as illustrated by the presence of sequences closely related to symbiotic bacteria found in other wood eating marine invertebrates. Finally, we found that sunken wood communities inside the canyon were different and more diverse than the ones outside the canyon. This finding extends to the microbial world the view that submarine canyons are sites of diverse marine life.

Introduction

Debris of terrestrial plants can be exported from land to sea by rivers and streams, especially during flooding events [1]. Once saturated with water, the debris sinks and brings a discrete load of organic carbon to the ocean floor. The importance of plant debris for the oceans organic carbon cycle has recently been acknowledged [2]. However, sunken wood has long been in the center of scientific interest because it can harbor distinct and specialized faunal communities [3] and because sunken wood is hypothesized to play a key role in the maintenance and dispersion of chemosynthetic species in the deep sea [4], [5]. Among the emblematic species found on sunken wood are wood boring marine invertebrates, which use the wood matrix as shelter and food. A significant amount of energy is stored in the wood as cellulose which is degraded by cellulolytic organisms, and their degradation products can be used by animals that host microbial symbiotic communities [6][8].

Xylophaga Turton, 1822 from the family Pholadidae, is the most common genus of xylotrophic (wood eating) bivalves found in sunken wood in the deep sea [9], 10. These animals bore into wood using their shell and ingest the wood particles produced. They use wood as an energy source, but the origin of the cellulase used to hydrolyze the cellulose is still uncertain because studies on the physiology of Xylophaga do not exist. Shallow-water wood-boring bivalves from the family Teredinidae (known as shipworms) have been more extensively studied. Shipworms have strongly modified and enlarged gills that harbor cellulolytic and nitrogen-fixing Gammaproteobacteria [11], [12]. Cellulolytic enzymes produced by the gill endosymbionts are assumed to contribute to lignocellulose breakdown in the digestive system, however, this has not been demonstrated and the transfer mechanism of cellulases from the gills to the digestive tract remains unclear. The gills of Xylophaga are much smaller and less modified [13]. They also contain dense populations of endobacteria, but it is not known if these help digesting cellulose [14]. Further, one of the characteristics of Xylophaga species is that they fill their burrows with pellets of feces mixed with mucus while shipworms expel their fecal pellets outside the burrow. Thus, it is possible that Xylophaga create new niches for microbial colonization within the wood.

Microorganisms are also present in the wood itself where they degrade wood cells. As submerged woods become anaerobic after only a few days [15], it is expected that pathways associated to fermentation, with electron acceptors other than oxygen, play a significant role. Indeed, studies on sunken wood in the deep Mediterranean revealed that this substrate can harbor rich bacterial communities [16][18] including fermenting bacteria, microorganisms involved in sulfur cycling and methane production, and new clades of Bacteria and Archaea with unknown physiologies [18]. As the application of molecular techniques has revealed a large diversity of microbes associated to sunken woods, they have also allowed a first understanding of the ecology of sunken wood microbial communities, showing that wood type, immersion time and the environmental conditions surrounding submerged wood may promote contrasted bacterial communities [19], [16][18]. For instance, bacterial communities associated with oak wood that had been artificially submerged in the Blanes Canyon (western Mediterranean) were dominated by Alphaproteobacteria of the family Rhodobacteriales, Gammaproteobacteria and Bacteroidetes [17]. In contrast, bacterial communities recovered with wood deployments from the Eastern Mediterranean differed on higher taxon levels and they were characterized by the presence of Flavobacteria [16]. Moreover, factors controlling community assembly remain poorly understood and the possible effect of wood-boring bivalves on bacterial community composition has never been explored.

This study aims at investigating the composition and diversity of microbial communities in wood, and in particular if external factors exhibit structuring influence. A major focus is on the effects of wood-boring bivalves on the wood associated microbial communities, in particular with respect to the abundant fecal pellets deposited by the bivalves in their burrows. We hypothesize that fecal pellets increase habitat diversity and that pellet communities are distinct from those in the wood matrix. Our second goal was to test if the special environmental conditions inside submarine canyons provide structuring influence on the microbial wood communities, as canyons are known to be hotspots for benthic biomass and productivity [20][22]. We used pine wood that was experimentally deployed in the deep Blanes Canyon off the Mediterranean Spanish coast and in similar water depths on an adjacent open slope area. We characterized the microbial communities in the wood matrix and fecal pellets by analyzing the bacterial 16S rRNA gene targeted by 454 pyrosequencing. In addition, we analyzed bacteria in Xylophaga gills to test if they relate to wood communities.

Materials and Methods

Experimental set up

Traps with triplicate cubes (8×8×8 cm) of pine and triplicate cubes of oak wood were deployed along the axis of Blanes Canyon at 900, 1200 and 1500 m water depth. Additional traps of pine and oak cubes were deployed next to Blanes Canyon on the western outer slope at 1200, 1500 and 1800 m depth (Figure 1 and Table 1). The oak wood samples have been presented previously [17] but are included in Table 1 for a comprehensive overview. The traps were suspended 20 m above the seafloor. Cubes from 1200 m depth were collected in November 2009 after 9 months of immersion and samples from 900, 1500 and 1800 m depths were collected after 12 months of immersion. Pine cubes from Blanes Canyon and the outer slope were used for the analysis of microbial communities in the wood and in the fecal pellets of wood boring Xylophaga bivalves. Microbial wood communities in the oak cubes from Blanes Canyon have been analyzed previously but they were used here for comparative analyses of microbial fecal pellet communities in pine and oak, as the pellets in these oak tubes were not previously analyzed.

thumbnail
Figure 1. Map of sampling sites and wood with Xylophaga burrows.

Map showing the position of the experimental moorings in relation to the Blanes Canyon (BC) and its outer slope (OS), drawn using GeoMapApp (http://www.geomapapp.org). Insert, cube of pine wood (8 X 8 X 8 cm) colonized by Xylophaga after 12 months of immersion in Blanes Canyon.

https://doi.org/10.1371/journal.pone.0096248.g001

thumbnail
Table 1. Overview of the different samples in the whole experiment.

https://doi.org/10.1371/journal.pone.0096248.t001

Immediately upon recovery of the cubes, wood chips to be used for microbial analysis were cut using sterilized tools, flash frozen in liquid nitrogen and kept at −20°C until further processing. Fecal pellets were collected from Xylophaga burrows in pine and oak cubes. Fecal pellets from individual wood triplicates were separately frozen at −20°C. Further, to estimate the % wood consumed, the rest of wood the cubes were carefully dissected by hand and all wood-boring bivalves were extracted and stored in 70% ethanol for taxonomic identification. To ensure that recently settled individuals were included, extraction was performed with the aid of a magnifier (2X) or a dissecting microscope. Shell length (SL) of each specimen of Xylophaga spp. was measured to the nearest 0.1 mm with digital calipers. The volume of each Xylophaga was calculated as the volume of a sphere with a radius equal to SL/2. Considering that each specimen created a burrow in the wood, the volume consumed by each Xylophaga was estimated to three time its volume. The % of wood consumed was estimated as the ratio between the volume occupied by Xylophaga spp. and the total volume of the wood cube.

No specific permissions were required for deploying submerged moorings in Blanes Canyon as it is not a protected area, moreover this study did not involve endangered or protected species.

Wood and pellet DNA extraction, PCR and pyrosequencing

Procedures for DNA extraction, PCR and pyrosequencing were performed as described earlier [17]. Briefly, representative pieces from each wood cube used for analyses of microbial wood communities (Table 1) were powdered by bead beating (RETSCH Mixer Mill. Retsch, Inc. MM 301) using 25-ml grinder jars (Retsch, Inc. MM 400 Stainless steel) and 20-mm diameter stainless steel balls. The grinder jars were dipped into liquid nitrogen to keep the wood brittle. The fecal pellet material was already in powder form after drying. Approximately 100 mg of the powders from woods or fecal pellets were used to extract genomic DNA with the Mobio PowerPlant kit (Ozyme, Saint-Quentin-en-Yvelines, France).

Initially, DNA extracts from individual triplicates of two selected treatments were amplified and sequenced: 1200 m pine wood chips and 1200 m pine pellet. Bacterial communities were then compared in a cluster analysis (see below), together with communities from oak wood obtained from a previous study from the Blanes Canyon [17]. The results revealed that community composition of individual triplicates were always more similar within than across treatments (Figure S1). We therefore pooled the triplicate DNA extracts from each treatment for further processing. This resulted in 6 pooled DNA extracts for wood (3 pine from the Blanes Canyon and 3 from the outer slope) and 5 pooled DNA extracts from fecal pellets (3 from pine and 2 from oak inside the Blanes Canyon) (Table 1).

A portion of the 16S rRNA gene was amplified by PCR using modified versions of universal bacterial 16S rRNA primers 27F (5'-AGRGTTTGATCMTGGCTCAG-3') [23] and 519R (5'-GTVTTACCGCGGCTGCTG-3') [24] as described in previously [17]. Amplicons obtained with the 27F primer were modified at the 5′ end by addition of the Roche 454 A-adaptor sequence and a 10-nucleotide identifier barcode (multiplex identifier, MID). Emulsion PCR and Roche 454 pyrosequencing (Genome Sequencer, FLX Titanium chemistry) were performed at the Genotoul platform of INRA, Toulouse (France).

Dissection, DNA extraction, PCR and pyrosequencing of Xylophaga gills

Bacterial 16S rRNA gene sequences from Xylophaga gill bacteria originated from a separate study using a different sequencing approach. Four individuals of Xylophaga sp. A, which was the most abundant wood-boring bivalve in the Blanes Canyon pine and oak deployments [25] were dissected. Three individuals originated from pine (Bla1.2, Bla10, Bla11), one from oak (Bla1.1; Table 1). DNA was extracted from their symbiont-containing gills following the protocol of Zhou et al. [26]. Amplification and 454 pyrosequencing of ∼480 bp long DNA fragments covering the V3 region of bacterial 16S rRNA genes was performed using primers bac339F (5′-CTCCTACGGGAGGCAGCAG-3′) and bac815R (5′-TTGTGCGGGCCCCCGTCAATT-3′) in a commercial laboratory (MR DNA, Shallowater, TX, USA). Bacterial DNA was amplified in a single-step PCR in which adaptors and barcodes were linked to the 5' region of the amplicons using the HotStarTaq Plus Master Mix Kit (Qiagen, Valencia, CA, USA). Amplification conditions were as follows: 94°C for 3 min, 28 cycles at 94°C for 30 s, 53°C for 40 s, 72°C for 1 min, and a final elongation step at 72°C. Equal concentrations of all PCR products were combined and purified with Agencourt Ampure beads (Agencourt Bioscience Corporation, MA, USA). Combined samples were sequenced with a Roche 454 FLX titanium instrument and reagents, following the manufacturer's procedures.

Attempts to amplify bacterial 16S rRNA genes from dissected gut tissue with the general bacterial primers 8F and 1492R [27] were unsuccessful and analysis of microbial gut communities was not further followed.

Sequence data analyses

All reads that had mismatches to the 16S rRNA primers, contained ambiguous nucleotides (N) or were shorter than 270 nucleotides (excl. the forward primer) were removed. The remaining sequences were subjected to stringent quality trimming to remove reads containing ≥3% bases with Phred values <27 (0.2% per-base error probability). This minimizes the influence of erroneous reads when clustering at 97% for OTU definition [28], [29]. Sequences were then de-replicated and clustered at a 97% threshold using Uclust [30]. Sequences from each OTU were classified by comparison to the Greengenes database [31]. Read quality filtering and length trimming, dereplication, clustering at 97% sequence identity, taxonomic classification and dataset partitioning based on barcodes were conducted with Pyrotagger [32]. The taxonomic affiliations of the most abundant OTUs (>1% of the sequences) were further verified against sequences from the NCBI databases using BLAST [33]. To compare bacterial communities for diversity analysis, all samples were randomly resampled to the size of the sample containing the fewest sequences (n = 798) using Daisy Chopper [34]. Calculation of the Shannon diversity index (H′) and cluster analysis were performed using the software PAST [35]. A similarity percentage analysis, SIMPER [36], was conducted to identify the phylotypes contributing the most to the dissimilarity between different samples. Sequences have been submitted to MG-RAST (http://metagenomics.anl.gov/linkin.cgi?project=5773) for public availability.

Because the 16S rRNA genes of free-living bacteria and gill bacteria from Xylophaga were amplified with different primer pairs, the obtained sequences did not cover identical gene fragments. However, the overlap included the entire hypervariable V3 region (E. coli positions 433–497) that is widely used in phylogenetic studies using next generation sequencing methods, and this region yields sufficient information for a direct comparison between wood, pellet and gill communities of Xylophaga. All sequences were therefore realigned and an OTU table based upon 100% sequence identity built from the common overlapping region of 80 bp. The analysis was done in mothur [37] using the Silva SEED database provided as a reference alignment.

Network association

A network analysis was conducted to characterize the relationships among bacterial OTUs as described earlier [38]. Maximal information-based nonparametric (MINE) statistics were applied by computing the maximal information coefficient (MIC) between each pair of OTUs [39]. MIC captures associations between data and provides a score that represents the strength of a relationship between data pairs. A matrix of MIC values >0.5 and corresponding to positive linear correlations was used with Cytoscape 2.8.3 to visualize the network of associations [40]. In these visualizations, bacterial OTUs are represented as nodes and are connected by lines that are proportional in length to the MIC value. The force-directed layout based on the Fruchterman-Rheingold Algorithm [41] was edge-weighted by the MIC value.

Results

Wood degradation and bacterial diversity

Pine wood showed different levels of degradation depending of the location, wood loss being higher in BC than in OS (18.3 vs 2.8% in average, respectively) (Figure S2). Xylophaga spp. had degraded large inner parts of the cubes, leaving their burrows filled with wood pellets (Figure 1, insert). Sequencing of the bacterial 16S rRNA gene originating from wood and pellet yielded a total of 55 874 quality checked sequences. The amount of sequences varied among samples but rarefaction analysis (Figure S3) shows that the sequencing effort was not exhaustive. Bacterial community diversity was significantly higher in the wood immerged in Blanes Canyon compared to wood deployed on the open slope (Figure S3), as estimated with the number of OTUs and the Shannon index (t-test, p = 0.01, Table 2). On the other hand, the bacterial diversity found in the pellets filling the burrows of Xylophaga spp. from the canyon did not differ significantly from that in the wood matrix (p = 0.09).

Bacterial community composition in the wood

Cluster analysis at the OTU level based on a Bray Curtis distance matrix showed that bacterial communities could be separated in two main groups: pellet and wood communities (Figure 2). Within the wood samples, communities grouped according to location. All samples from Blanes Canyon grouped together and were separate from the open slope samples. Further, a comparison of sequences from this study with oak wood sequences from Fagervold et al. [17] showed that pine wood communities were different from oak wood communities (Figure S1). This was also true when comparing the pellet samples only, in that they grouped according to wood type. Depth was not a structuring factor for the community composition in pellets (Figure 2).

thumbnail
Figure 2. Sample clustering.

Dendrogram based on Bray-Curtis distance representing the similarity between bacterial communities sequenced from the wood matrix and from burrow pellets obtained from wood immerged at various depths in Blanes Canyon (BC) and its adjacent open slope (OS).

https://doi.org/10.1371/journal.pone.0096248.g002

The microbial composition differed at high taxonomic level (phylum/class level) (Figure 3). Pine wood samples contained more Alphaproteobacteria, Planctomycetes and Deltaproteobacteria while pine pellet communities contained more Gammaproteobacteria and Bacteroidetes (Figure 3). Regarding wood samples, canyon communities harbored less Alphaproteobacteria and more Deltaproteobacteria than the open slope. Concerning pellet samples, oak contained more Alphaproteobacteria than pine. In turn, gill communities were dominated by Gammaproteobacteria sequences with the exception of the sample collected from oak, Bla11, which contained more Alphaproteobacteria (Figure 3).

thumbnail
Figure 3. Distribution of bacterial taxa.

Relative proportion of the most abundant bacterial phylum or class (>1% of the sequences) found in wood, pellet and gill (Bla) samples from Blanes Canyon (BC) and its adjacent open slope (OS). *  =  contained significantly less sequences than the rest of the samples (n = 263).

https://doi.org/10.1371/journal.pone.0096248.g003

OTU co-occurrence and taxonomy in the wood

Microbial communities in pine pellets from Blanes Canyon shared very few OTUs with wood samples from the same wood cubes, and the few OTUs in common were not abundant (Figure 4). Among the typical pine pellet OTUs, the most abundant were OTU 31, a Gammaproteobacterium distantly related to unpublished sequences from a marine biofilm, followed by OTU 10, a Bacteroidetes distantly related (90%) to algae associated communities, a Gammaproteobacterium (OTU 12) distantly related (93%) to sequence from sunken wood [42], and an Epsilonproteobacterium (OTU 15) previously detected as abundant in sequences cloned from pine pellets (OTU6, [19]). Among less abundant pellet OTUs, one Gammaproteobacterium (OTU 83) exhibited 99% similarity to a symbiotic bacterium from the Teredinidae Lyrodus pedicellatus [43].

thumbnail
Figure 4. OTU network.

Network showing associations between OTUs found in wood and pellets from wood immerged at various depths in Blanes Canyon (BC) and its adjacent open slope (OS). OTUs are represented as nodes (white circles) in the network and the size of each node is proportional to the number of sequences contained in each OTU. Lines connect the nodes that are the most correlated (MIC values>0.5). The identifying numbers of the most abundant OTUs are written in the nodes. Colors highlight associations between the OTUs that most explain the differences between groups of samples (Oak pellet, Pine pellet, BC Pine wood and OS Pine wood). OTUs that best explain differences between samples were identified by a SIMPER analysis.

https://doi.org/10.1371/journal.pone.0096248.g004

Pine pellet communities had more in common with those from oak pellets than from the pine wood. Nevertheless, the oak pellets contained more Alphaproteobacteria closely related to coral tissue clones [44], [45] (OTU 21) (97%). Among the OTU shared between pellet samples, many belonged to Bacteroidetes often distantly related to sponge bacteria (OTU 62 and 77, 94–95% similarity, [46] or deep sediments (OTU 45, 95% similarity, [47]).

On the other hand, communities from wood in the canyon shared many OTUs with open slope communities, as illustrated by the short distance separating these samples in the network (Figure 4). In pine wood from the canyon, an abundant Planctomycetes OTU (OTU 14) was 100% similar to sequences found in the digestive tract of a chiton, a wood associated mollusk [42], and it was 99% identical to digestive tract bacteria found in a sea urchin from a wood fall [48]. Further, an abundant Deltaproteobacterium OTU (OTU 51) was closely related (99%) to clones found in sunken wood [42]. A less abundant Gammaproteobacterium was distantly related to Teredinibacter turnerae, an intracellular endosymbiont of shallow water marine wood-boring Teredinidae [49]. In wood from the open slope, the most abundant OTU (OTU 5) belonged to the Rhodobacteriales order of Alphaproteobacteria and was 100% similar to a sequence found earlier in oak wood from Blanes Canyon (Blanes 1043, [17]) and distantly related (91%) to bacteria found in guts of the marine wood-feeding gastropod Pectinodonta sp. (Patellogastropoda, Mollusca) [50] and the chiton Nierstraszella lineata [42]. Another abundant OTU (OTU 7) was identified as a Gammaproteobacterium related to clones from gorgon or seafloor lava (96%). This OTU was not detected in previous oak wood samples from Blanes Canyon.

Xylophaga gill community

A total of 8260 sequences were obtained after quality check from the four individuals of Xylophaga (Table 1). The number of sequences varied between samples as did the diversity, but the gill community from Xylophaga was always less diverse than wood or pellet communities (Table 1). Bacteria from gills of Xylophaga were different from those found in wood and pellets (Figure 5). Only 17 Xylophaga OTUs (8%) were found in the pellets or in wood while pellets and wood shared 55 of their OTUs (21 to 26%) (Figure 5).

thumbnail
Figure 5. Venn diagram representing bacterial OTUs unique and common to the wood samples, the pellet and the Xylophaga gills.

https://doi.org/10.1371/journal.pone.0096248.g005

Many sequences from the gills of Xylophaga were closely related to those of symbiotic bacteria from shipworms. In specimen Bla1.2, at least 96% of the sequences were from possible symbionts. The most abundant OTU, which represented 84% of the sequences, was 99% similar to a sequence of a Gammaproteobacterium from the teredinid shipworm L. pedicellatus (Clone RT14, [43]). In Bla10, at least 75% of the sequences were from possible symbionts. The most abundant OTU (73%) was 95% similar to the 16S rRNA gene of the T7901 strain of the Gammaproteobacterium T. turnerae [49]. In Bla11, more than 45% of the sequences were possible symbionts. The most abundant OTUs were less dominant than in the other Xylophaga specimens with only 24% of the sequences distantly affiliated to a symbiont of L. pedicellatus (Clone RT20, [43]). For specimen Bla1.1, we probably failed to amplify specific gill symbionts as the most abundant OTU related to symbionts from Teredinidae represented only 5% of the sequences.

Discussion

The incubation of wood pieces in Blanes Canyon and its adjacent open slope showed that the boring activity of Xylophaga sp. A transformed the wood environment by creating distinct niches for bacterial communities. The communities associated with the gills of Xylophaga were different and probably composed of symbiotic bacteria, while the niche created by the pellets promoted the development of bacterial communities that were very different from those in the wood matrix. The large differences observed at phylum and OTU levels suggest that the pellet populations represent a distinct community rather than a subset of the wood communities.

The accumulation of pellets consolidated with mucus is typical for Xylophaga [13]. While Teredinidae expulse the products of their wood boring activity from their burrows, Xylophaga species most often form chimneys of compacted fecal pellets around their siphons, thus lining the bored tunnels. These pellets have a very different chemical composition compared to wood, as up to 80% of the wood cellulose may be lost during digestion by Xylophaga spp. [51]. Pellets may thus form a new substrate for microorganisms that is depleted in cellulose but enriched in mucus, in comparison to wood. Alternatively, the community found in pellets might be similar to that found inside the gut of Xylophaga. These animals ingest the wood shavings produced by their shell, which they use as rasp to dig burrows. Wood particles are then stored in a large caecum before passing through the stomach and intestine. However, previous studies found few microbes in the caecum of Xylophaga [14] and Teredinidae [52] while large numbers of bacteria occurred in the fecal pellets in the intestines of Teredinidae [52]. It is not known if bacterial communities follow the same spatial distribution in the digestive tracts of Xylophaga, but similar life strategies and common evolutionary history of the two wood-boring bivalve groups [53] may suggest so. Bacteria colonizing the pellets may thus exit the digestive tract with the excreted feces and survive in the burrows. However this hypothesis could not be verified in this study, as the gut microbiota was not analyzed. Further, communities from pine pellets were different from those in oak pellets. This difference could be due to the presence of a second species, Xylophaga sp. B, that was predominantly colonizing pine while it was rare in oak [25]. It is possible that this species has different gut microflora than the ubiquitously abundant Xylophaga sp. A [25] and this could influence the composition of pellet communities in pine and oak. However, it is also possible that the composition of the wood itself is controlling the composition of pellet communities. Remarkably, we could not identify known cellulolytic strains among the bacteria detected in the fecal pellets. However, since cellulotyic pathways are widespread over many bacterial phyla, it cannot be excluded that the uncultured bacteria we detected may be able to degrade cellulose. One metabolic pathway that might be inferred from the identified OTUs is fermentation. OTU 17 shares 98% sequence identity to the fermentative Alphaproteobacterium Polymorphum gilvum [54], and this close relationship may be an indication that fermentative bacteria can take advantage of the wood remains.

We detected several abundant OTUs in the gills of Xylophaga sp. A closely related to sequences from bacteria associated to wood-boring Teredinidae. The very close similarity (99%) of these OTUs to symbionts from the shipworm Lyrodus pedicellatus is a strong indication that we were able to detect gill symbionts from Xylophaga sp. A. Bacteria had been observed earlier in gills of Xylophaga [14] but they have never been isolated or taxonomically characterized. Further, the most abundant OTU in Xylophaga sp. A matched the symbionts of the most abundant phylotype (clade P3) in the Teredinidae L. pedicellatus [43]. We also detected sequences related to the cultivated shipworm endosymbiont Teredinibacter turnerae, a cellulolytic Gammaproteobacterium that has been isolated from many teredinid host species [12]. The presence of similar sequences in deep-sea and shallow environments indicates that wood-boring endosymbionts may be adapted to a wide range of environments and hosts.

The location of the sunken wood, a factor that is linked to different levels of wood degradation, clearly shaped the community composition in the wood matrix. Wood deposited in the canyon, which was largely and more rapidly degraded and colonized by the wood-boring bivalves, harbored different and more diverse communities than wood in the open slope. Canyons channel the transport of organic matter from the continental shelf to the deep sea, which is specially enhanced during dense shelf water cascading events [55]. This process in combination with frequent storms from easterly directions in the western Mediterranean area and offshore convection appear to be the main drivers for the transfer of organic matter to the deep Mediterranean Sea [56], [57] and they may in particular contribute to high benthic biomass and productivity that make the hot spot character of submarine canyon ecosystems [21]. Our results thus extend to the microbial world the view that submarine canyons are sites of enhanced marine diversity. The difference in bacterial community composition between woods immerged inside and outside the canyon may also be due to different sources of microbes colonizing that wood. Allochthonous bacteria could be transported to the canyon as a result of ecosystem forcing events, which may lead to the establishment of different communities in the canyon water column or sediments compared to the open slope. However, we cannot discard the possibility that the difference in bacterial communities between locations are linked to the different phases of the wood decomposition process. The pine wood communities where characterized by a large proportion of Alphaproteobacteria followed by Planctomycetes and Gammaproteobacteria. These classes of bacteria were also observed previously in oak wood [17], emphasizing their role in wood degradation. However, there were differences between pine and oak at the OTU level. The most abundant pine OTUs were absent in oak, like the Gammaproteobacterium OTU 7, or less abundant, as in the case of the Alphaproteobacterium OTU 5. Some pine OTUs were also 100% similar to sequences found on sunken wood, inside a chiton gut [42], and in a sea urchin collected around the Vanuatu Island [48]. This suggests that some bacteria associated with sunken wood are distributed worldwide. The presence of these bacteria in the guts of different wood-ingesting animal species suggests that they may either be associated to a large variety of macroorganisms or that they may represent globally distributed sunken-wood microorganisms that are ingested by the animals together with the wood.

Supporting Information

Figure S1.

Clustering of all samples. Hierarchical clustering using the Bray Curtis index showing the similarity of the microbial communities between the different samples.

https://doi.org/10.1371/journal.pone.0096248.s001

(PDF)

Figure S2.

Wood consumtion. Percentage of consumed pine wood after immersion at various depths in Blanes Canyon (BC) and its adjacent open slope (OS), expressed as percentage of the initial volume of the respective wood cubes.

https://doi.org/10.1371/journal.pone.0096248.s002

(PDF)

Figure S3.

Rarefaction curve. Number of sequences versus OTUs formed for all sequenced samples, except the gill samples. Broadly, blue lines represents BC Pine Wood, redish lines OS Pine Wood, greenish lines Pine Pellet sampels and purple lines oak pellets.

https://doi.org/10.1371/journal.pone.0096248.s003

(PDF)

Acknowledgments

We thank all members of the projects PROMETEO and DOSMARES and, especially the team of Prof. A. Calafat and the crew of R/V Garcia del Cid for helping in deploying and collecting the wood samples.

Author Contributions

Conceived and designed the experiments: PEG SKF CR DM. Performed the experiments: DM CR. Analyzed the data: SKF PEG DK CR DM CB ANJ. Wrote the paper: SKF PEG DK CR DM CB ANJ.

References

  1. 1. West AJ, Lin C-W, Lin T-C, Hilton RG, Liu S-H, et al. (2011) Mobilization and transport of coarse woody debris to the oceans triggered by an extreme tropical storm. Limnol Oceanogr 56: 77–85.
  2. 2. Bianchi TS (2011) The role of terrestrially derived organic carbon in the coastal ocean: A changing paradigm and the priming effect. Proc Natl Acad Sci 108: 19473–19481.
  3. 3. Wolff T (1979) Macrofaunal utilization of plant remains in the deep sea. Sarsia 64: 117–136.
  4. 4. Distel D, Baco A, Chuang E, Morrill W, Cavanaugh C, et al. (2000) Do mussels take wooden steps to deep-sea vents. Nature 403: 725–726.
  5. 5. Jones WJ, Won YJ, Maas PAY, Smith PJ, Lutz RA, et al. (2006) Evolution of habitat use by deep-sea mussels. Mar Biol 148: 841–851.
  6. 6. Gaudron SM, Pradillon F, Pailleret M, Duperron S, Le Bris N, et al. (2010) Colonization of organic substrates deployed in deep-sea reducing habitats by symbiotic species and associated fauna. Mar Environ Res 70: 1–12.
  7. 7. Laurent MCZ, Gros O, Brulport J-P, Gaill F, Le Bris N (2009) Sunken wood habitat for thiotrophic symbiosis in mangrove swamps. Mar Environ Res 67: 83–88.
  8. 8. Leschine SB (1995) Cellulose degradation in anaerobic environments. Ann Rev Microbiol 49: 399–426.
  9. 9. Turner RD (1973) Wood-boring bivalves, opportunistic species in the deep sea. Science 180: 1377–1379.
  10. 10. Turner RD (2000) On the subfamily Xylophagainae (Family Pholadidae, Bivalvia, Mollusca). Bull Mus Comp Zool 157: 223–308.
  11. 11. Distel D, DeLong E, Waterbury J (1991) Phylogenetic characterization and in situ localization of the bacterial symbiont of shipworms (Teredinidae: Bivalvia) by using 16S rRNA sequence analysis and oligodeoxynucleotide probe hybridization. Appl Environ Microbiol 57: 2376–2382.
  12. 12. Distel DL, Morrill W, MacLaren-Toussaint N, Franks D, Waterbury J (2002) Teredinibacter turnerae gen. nov., sp. nov., a dinitrogen-fixing, cellulolytic, endosymbiotic gamma-proteobacterium isolated from the gills of wood-boring molluscs (Bivalvia: Teredinidae). Int J Syst Evol Micr 52: 2261–2269.
  13. 13. Purchon DR (1941) On the biology and relationships of the lamellibranch Xylophaga dorsalis (Turton). J Mar Biol Ass U K 25: 1–39.
  14. 14. Distel DL, Roberts SJ (1997) Bacterial endosymbionts in the gills of the deep-sea wood-boring bivalves Xylophaga atlantica and Xylophaga washingtona. Biol Bull 192: 253–261.
  15. 15. Yücel M, Galand PE, Fagervold SK, Contreira-Pereira L, Le Bris N (2013) Sulfide production and consumption in degrading wood in the marine environment. Chemosphere 90: 403–409.
  16. 16. Bienhold C, Pop Ristova P, Wenzhöfer F, Dittmar T, Boetius A (2013) How deep-sea wood falls sustain chemosynthetic life. PLoS ONE 8: e53590.
  17. 17. Fagervold SK, Bessette S, Romano C, Martin D, Plyuscheva M, et al. (2013) Microbial communities associated with the degradation of oak wood in the Blanes submarine canyon and its adjacent open slope (NW Mediterranean). Prog Oceanogr 118: 137–143.
  18. 18. Fagervold SK, Galand PE, Zbinden M, Gaill F, Lebaron P, et al. (2012) Sunken woods on the ocean floor provide diverse specialized habitats for microorganisms. FEMS Microbiol Ecol 82: 616–628.
  19. 19. Bessette S, Fagervold SK, Romano C, Martin D, Le Bris N, et al. (2013) Diversity of bacterial communities on sunken woods in the Mediterranean Sea. J Mar Sci Tech 22: 60–66.
  20. 20. Cartes JE, Maynou F, Fanelli E, Romano C, Mamouridis V, et al. (2009) The distribution of megabenthic, invertebrate epifauna in the Balearic Basin (western Mediterranean) between 400 and 2300 m: Environmental gradients influencing assemblages composition and biomass trends. J Sea Res 61: 244–257.
  21. 21. De Leo FC, Smith CR, Rowden AA, Bowden DA, Clark MR (2010) Submarine canyons: hotspots of benthic biomass and productivity in the deep sea. Proc R Soc B 22: 2783–2792.
  22. 22. Vetter EW, Smith CR, De Leo FC (2010) Hawaiian hotspots: enhanced megafaunal abundance and diversity in submarine canyons on the oceanic islands of Hawaii. Mar Ecol 31: 183–199.
  23. 23. Vergin KL, Urbach E, Stein JL, DeLong EF, Lanoil BD, et al. (1998) Screening of a fosmid library of marine environmental genomic DNA fragments reveals four clones related to members of the order Planctomycetales. Appl Environ Microbiol 64: 3075–3078.
  24. 24. Teske A, Alm E, Regan JM, Toze S, Rittmann BE, et al. (1994) Evolutionary relationships among ammonia- and nitrite-oxidizing bacteria. J Bacteriol 176: 6623–6630.
  25. 25. Romano C, Voight JR (2013) Company JB, Plyuscheva M, Martin D (2013) Submarine canyons as the preferred habitat for wood-boring species of Xylophaga (Mollusca, Bivalvia). Prog Oceanogr 118: 175–187.
  26. 26. Zhou J, Bruns MA, Tiedje JM (1996) DNA recovery from soils of diverse composition. Appl Environ Microbiol 62: 316–322.
  27. 27. Muyzer G, Teske A, Wirsen CO, Jannasch HW (1995) Phylogenetic relationships of Thiomicrospira species and their identification in deep-sea hydrothermal vent samples by denaturing gradient gel electrophoresis of 16S rDNA fragments. Arch Microbiol 164: 165–172.
  28. 28. Huse SM, Welch DM, Morrison HG, Sogin ML (2010) Ironing out the wrinkles in the rare biosphere through improved OTU clustering. Environ Microbiol 12: 1889–1898.
  29. 29. Kunin V, Engelbrektson A, Ochman H, Hugenholtz P (2010) Wrinkles in the rare biosphere: pyrosequencing errors can lead to artificial inflation of diversity estimates. Environ Microbiol 12: 118–123.
  30. 30. Edgar RC (2010) Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26: 2460–2461.
  31. 31. DeSantis TZ, Hugenholtz P, Larsen N, Rojas M, Brodie EL, et al. (2006) Greengenes, a chimera-checked 16S rRNA gene database and workbench compatible with ARB. Appl Environ Microbiol 72: 5069–5072.
  32. 32. Kunin V, Hugenholtz P (2010) PyroTagger: A fast, accurate pipeline for analysis of rRNA amplicon pyrosequence data. The Open Journal 1..
  33. 33. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, et al. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res 25: 3389–3402.
  34. 34. Gilbert JA, Field D, Swift P, Newbold L, Oliver A, et al. (2009) The seasonal structure of microbial communities in the Western English Channel. Environ Microbiol 11: 3132–3139.
  35. 35. Hammer Ø, Harper DAT, Ryan PD (2001) PAST: Paleontological statistics software package for education and data analysis. Paleontol Electro 4: 9.
  36. 36. Clarke KR (1993) Non-parametric multivariate analyses of changes in community structure. Aust J Ecol 18: 117–143.
  37. 37. Schloss PD, Westcott SL, Ryabin T, Hall JR, Hartmann M, et al. (2009) Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl Environ Microbiol 75: 7537–7541.
  38. 38. Blanquer A, Uriz MJ, Galand PE (2013) Removing environmental sources of variation to gain insight on symbionts vs. transient microbes in high and low microbial abundance sponges. Environ Microbiol 15: 3008–3019.
  39. 39. Reshef DN, Reshef YA, Finucane HK, Grossman SR, McVean G, et al. (2011) Detecting novel associations in large data sets. Science 334: 1518–1524.
  40. 40. Smoot ME, Ono K, Ruscheinski J, Wang PL, Ideker T (2011) Cytoscape 2.8: new features for data integration and network visualization. Bioinformatics 27: 431–432.
  41. 41. Fruchterman TMJ, Reingold EM (1991) Graph drawing by force-directed placement. Software Pract Exper 21: 1129–1164.
  42. 42. Duperron S, Pottier MA, Leger N, Gaudron SM, Puillandre N, et al. (2013) A tale of two chitons: is habitat specialisation linked to distinct associated bacterial communities? FEMS Microbiol Ecol 83: 552–567.
  43. 43. Luyten YA, Thompson JR, Morrill W, Polz MF, Distel DL (2006) Extensive variation in intracellular symbiont community composition among members of a single population of the wood-boring bivalve Lyrodus pedicellatus (Bivalvia: Teredinidae). Appl Environ Microbiol 72: 412–417.
  44. 44. Klaus JS, Janse I, Heikoop JM, Sanford RA, Fouke BW (2007) Coral microbial communities, zooxanthellae and mucus along gradients of seawater depth and coastal pollution. Environ Microbiol 9: 1291–1305.
  45. 45. Sunagawa S, DeSantis TZ, Piceno YM, Brodie EL, DeSalvo MK, et al. (2009) Bacterial diversity and white plague disease-associated community changes in the Caribbean coral Montastraea faveolata. ISME J 3: 512–521.
  46. 46. Sipkema D, Holmes B, Nichols SA, Blanch HW (2009) Biological characterisation of Haliclona (?gellius) sp.: sponge and associated microorganisms. Microb Ecol 58: 903–920.
  47. 47. Schauer R, Bienhold C, Ramette A, Harder J (2009) Bacterial diversity and biogeography in deep-sea surface sediments of the South Atlantic Ocean. ISME J 4: 159–170.
  48. 48. Becker PT, Samadi S, Zbinden M, Hoyoux C, Compère P, et al. (2009) First insights into the gut microflora associated with an echinoid from wood falls environments. Cah Biol Mar 50: 343–352.
  49. 49. Yang JC, Madupu R, Durkin AS, Ekborg NA, Pedamallu CS, et al. (2009) The complete genome of Teredinibacter turnerae T7901: An intracellular endosymbiont of marine wood-boring bivalves (shipworms). PLoS ONE 4: e6085.
  50. 50. Zbinden M, Pailleret M, Ravaux J, Gaudron SM, Hoyoux C, et al. (2010) Bacterial communities associated with the wood-feeding gastropod Pectinodonta sp. (Patellogastropoda, Mollusca). FEMS Microbiol Ecol 74: 450–463.
  51. 51. Boynton LC, Miller RC (1927) The occurence of a cellulase in the shipworm. J Biol Chem 75: 613–618.
  52. 52. Betcher MA, Fung JM, Han AW, O'Connor R, Seronay R, et al. (2012) Microbial distribution and abundance in the digestive system of five shipworm species (Bivalvia: Teredinidae). PLoS ONE 7: e45309.
  53. 53. Distel DL, Amin M, Burgoyne A, Linton E, Mamangkey G, et al. (2011) Molecular phylogeny of Pholadoidea Lamarck, 1809 supports a single origin for xylotrophy (wood feeding) and xylotrophic bacterial endosymbiosis in Bivalvia. Mol Phylogenet Evol 61: 245–254.
  54. 54. Cai M, Wang L, Cai H, Li Y, Tang Y-Q, et al. (2011) Rubrimonas shengliensis sp. nov. and Polymorphum gilvum gen. nov., sp. nov., novel members of Alphaproteobacteria from crude oil contaminated saline soil. Syst Appl Microbiol 34: 321–327.
  55. 55. Canals M, Puig P, de Madron XD, Heussner S, Palanques A, et al. (2006) Flushing submarine canyons. Nature 444: 354–357.
  56. 56. Canals M, Company J, Martín D, Sànchez-Vidal A, Ramírez-Llodrà E (2013) Integrated study of Mediterranean deep canyons: Novel results and future challenges. Prog Oceanogr 118: 1–27.
  57. 57. Canals M, Danovaro R, Heussner S, Lykousis V, Puig P, et al. (2009) Cascades in mediterranean submarine grand canyons. Oceanogr 22: 26–43.