Abstract
Bivalves from the family Lucinidae host sulfur-oxidizing bacterial symbionts, which are housed inside specialized gill epithelial cells and are assumed to be acquired from the environment. However, little is known about the Lucinidae life cycle and symbiont acquisition in the wild. Some lucinid species broadcast their gametes into the surrounding water column, however, a few have been found to externally brood their offspring by the forming gelatinous egg masses. So far, symbiont transmission has only been investigated in one species that reproduces via broadcast spawning. Here, we show that the lucinid Loripes orbiculatus from the West African coast forms egg masses and these are dominated by diverse members of the Alphaproteobacteria, Clostridia, and Gammaproteobacteria. The microbial communities of the egg masses were distinct from those in the environments surrounding lucinids, indicating that larvae may shape their associated microbiomes. The gill symbiont of the adults was undetectable in the developmental stages, supporting horizontal transmission of the symbiont with environmental symbiont acquisition after hatching from the egg masses. These results demonstrate that L. orbiculatus acquires symbionts from the environment independent of the host’s reproductive strategy (brooding or broadcast spawning) and reveal previously unknown associations with microbes during lucinid early development.
Similar content being viewed by others
Introduction
Marine invertebrates have a remarkable range of reproductive modes, but in general there are two major strategies: 1) broadcast spawning, where gametes are released to the water column where fertilization occurs, and 2) brooding, where early development occurs in “masses” within or close to the parent [1]. These two modes have major consequences for dispersal capacity, as broadcast larvae are usually numerous, long-lived, and planktotrophic, whereas brooding species either produce viviparous juveniles or lecithotrophic larvae which usually settle close to their parental populations [2].
Gelatinous egg mass formation as a form of external brooding is a common reproductive strategy in marine invertebrates. It confers many advantages either through chemical defense and the production of antibacterial compounds [3, 4], or through protection from solar radiation and desiccation [5]. Egg mass brooding has been observed in fish [6], crustaceans [7], echinoids [8], polychaetes [9], and trochid gastropods [10] but this strategy is rare in bivalves. External egg masses have so far only been reported for six species of marine bivalves: the protobranchs Nucula delphinodonta and Turtonia minuta, the cardiid Parvicardium exiguum, the semelid Abra tenuis, and the chemosymbiotic lucinids Phacoides pectinatus and Loripes orbiculatus [11,12,13].
The bivalve family Lucinidae is one of the most species-rich families in the ocean today [14]. Their characteristic feature is a conspicuous symbiosis with sulfur-oxidizing gammaproteobacteria, which live inside host cells called bacteriocytes in the gill epithelia [14]. In this intimate mutualistic association, the host provides the symbiont with reduced compounds such as hydrogen sulfide and oxygen, driving the symbiont primary production that supports a substantial fraction of the animal host’s nutrition [15, 16]. Lucinid bivalves are found worldwide and can be enormously abundant. Some seagrass sediments can have >4000 individuals per m2 [17] and, in seagrass meadows these bivalves clearly play a key role in ecosystem functioning [18,19,20]. Despite this, little is known about their life cycle in the wild, although they are all assumed to lack symbionts in their early life stages and acquire them from the environment during development [21]. Further, “broadcast” spawning, the release of gametes into the water column, has been induced in the laboratory for the lucinid Codakia orbicularis [22].
In this study, we investigated reproduction and symbiont transmission in Loripes orbiculatus, which is distributed along the English coast, throughout the Mediterranean, Black Sea, and Atlantic coasts, to West Africa [11, 12, 18, 23]. In Mauritania, this species spawns seasonally, with a major spawning event in winter (December–January) and a minor one in summer (July), by releasing gelatinous egg masses that get attached to seagrass blades on intertidal mudflats [24]. Our goal was to investigate the microbial communities associated with the early life stages of a lucinid species that reproduces via brooding. In particular, we asked whether the gill endosymbiont that associates with the adults can be detected in these early life stages of a brooding host species, which would contrast with the horizontal transmission mode demonstrated for other lucinid species.
Material and methods
Sample collection and fixation
Egg masses and adult bivalves were collected in two consecutive seasons (December 2015 and January 2017) from intertidal seagrass mudflats around the village of Iwik located in the Banc d’Arguin National Park, Mauritania, West Africa (19°53.82′N, 16°18.63′W; Fig. 1). Developmental stages and adult specimens for molecular analyses were rinsed with 0.2 µm sterile, filtered seawater, and immediately fixed in RNAlater (Cat. No. AM7020; Life Technologies, Carlsbad, CA, USA) at 4 °C overnight and stored at −20 °C until extraction. Three adult Loripes orbiculatus individuals were dissected and gills were fixed in 4% paraformaldehyde (PFA), dehydrated into 70% ethanol, and stored at 4 °C until further analysis (Supplementary Methods). Sediment and seagrass rhizome debris was sampled in December 2015 for microbial community analysis. Samples (n = 5) were collected at the site using aseptic techniques from 0 to 8 cm depth, fixed in 96% ethanol and stored at −20 °C until further processing.
A Map showing the sample locations including seagrass densities in the study area based on the median Normalized Difference Vegetation index (NDVI) of the Sentinel-2 Images (Map created in GoogleEngine, filter date; “2020-01-01”, “2021-03-03”). B Adult bivalves on seagrass debris; scale bar = 1.2 cm. C Adult L. orbiculatus visualized under a stereo microscope, the asterisk marks one of the symbiont-housing organs, the gills; scale bar = 0.3 cm. D Egg mass (approximately 4 cm in size) of L. orbiculatus attached to seagrass leaves. Larvae are embedded in a gelatinous, collagen-like fiber structure; scale bare = 200 µm. E Veliger larvae inside the egg mass; scale bar = 20 µm.
Live egg masses were brought back to the laboratory for hatching and kept individually in petri dishes (100 mm diameter) provided with native, sterile sand and 0.2 µm sterile, filtered seawater (salinity 40‰) without any additives. The water was exchanged daily, and the development of larvae was monitored over a course of 54 days. Three larvae from each of the four sampling points (prior to hatching, 10, 12, and 54 days post hatching) were fixed for histological analysis in 4% PFA (in 0.01 M PBS, 10% sucrose), washed, dehydrated, and stored at 4 °C until embedding.
Histological sample processing
Adult gill tissue was embedded according to [25] with modifications (Supplementary Methods). A Leica RM2235 microtome (Leica, Nussloch, Germany) was used to cut the embedded gills into 5 µm sections. Dewaxing was performed with 90%, 80%, 70%, and 50% ethanol for 5 min each. Larvae were embedded in LR-White low viscosity resin (London Resin Company, London, UK) as described in ref. [26] and cut into 1-µm semithin sections using a Leica EM UC7 Ultramicrotome. For histological evaluation toluidine blue was applied to all sections, which were evaluated on an AxioImager A1 microscope (Zeiss, Oberkochen, Germany).
Fluorescence in situ hybridization (FISH)
Twelve larvae (three individuals from each of the four-time points) and gill tissues from three adult bivalves were hybridized with a symbiont-specific and general bacterial probes (Table S2; refs. [27,28,29]) in a 35% formamide buffer. The nonsense probe NON-388EUB was used to detect nonspecific binding. Slides with reaction mixture were incubated at 46 °C for 3 h in dark conditions (Supplementary Methods, Table S3). After hybridization, the slides were washed with pre-warmed (48 °C) washing buffer for 15 min and dried with compressed air. The sections were stained with 10 μg/ml DAPI and incubated in the dark for 5 min. After staining, slides were washed in ice-cold MilliQ water, mounted with CitifluorTM antifade mounting medium (Citifluor Products, Canterbury, UK) and stored at 4 °C. Sections were evaluated on a TCS SP8 X confocal laser scanning microscope (Leica, Wetzlar, Germany).
DNA extraction and marker gene amplification
DNA from gill tissues of four adult bivalves and from small pieces of 26 egg masses containing larvae was extracted using the DNAeasy Blood & Tissue Micro Kit (Qiagen, Hilden, Germany) following manufacturer’s instructions. Seven additional egg mass samples, from which larvae had been manually removed under aseptic conditions prior to DNA extraction served as controls. DNA from sediment and seagrass rhizome debris was extracted with the PowerSoil® DNA Isolation Kit (MoBio Laboratories, Inc., Carlsbad, CA, USA) following manufacturer’s instructions. The purified DNA samples were quantified using the Qubit 2.0 broad range assay (Life Technologies).
For molecular characterization of the host, we used the mitochondrial cytochrome b gene for PCR-based screenings. The following primers were used to amplify a 355 bp fragment of the lucinid cytB gene: cytb-F [30] and cytbR_new [31]. For DNA barcoding of the symbiont, the universal bacterial primers 616V and 1492R [32, 33] were used to amplify an ~1400 bp fragment of the 16S rRNA gene. The primer sequences, amplification protocols, and cycling conditions used for cytB and 16S rRNA gene PCRs are in Tables S2 and S4. Purified DNA (ZR96 DNA Cleanup Kit; Zymo Research, Orange, CA, USA) was sent for Sanger sequencing at Microsynth Austria.
16S rRNA amplicon sequencing
DNA from the egg masses, sediment, and seagrass rhizome debris was amplified in a two-step barcoding approach as described in [34] using the primer pair 341F/785R [35]. The first step PCR was done in triplicates for all samples to minimize stochastic PCR biases. Primers and detailed PCR protocols including cycling conditions are given in the Supplementary Material (Tables S1, S3). The triplicate PCR products were pooled and the ZR96 DNA Cleanup Kit (Zymo Research) was used, according to the manufacturer’s protocol, to purify the pooled PCR products to avoid carryover of primers and primer dimers to the second-step PCR.
Second step PCR reactions consisted of the same reagents as in the first step PCRs with one amendment—each reaction contained 50 µM barcode primers per reaction (Table S4). The second amplification step for egg mass PCR products was performed under the same conditions as described for the first step, but in 10 cycles, to avoid amplification biases due to high cycle numbers. The sediment and seagrass rhizome debris cycling conditions are provided in Table S4. The second step PCR products were purified using the same procedure as the first step PCR products. Purified PCR products were quantified using the Quant-iTTM PicoGreen® dsDNA Assay Kit (Invitrogen Life Technologies, Gaithersburg, MD, USA). All samples were pooled together in equimolar amounts and sent for Illumina MiSeq (2 × 250 bp) sequencing at Microsynth Austria.
Sequence processing and statistical analysis
Raw sequences were demultiplexed following the methods described in ref. [34] and further processed using dada2 v. 1.14.1 [36] in R v. 3.6.1 [37]. The demultiplexed reads were quality filtered, after which a total of 38 samples remained (26 from egg masses, seven from egg masses without larvae, and five from sediment and rhizome debris).
The reads from these samples were joined and an amplicon sequence variant (ASV) table was constructed. The ASV table was subsequently filtered to remove chimeras as well as any sequences resulting from mitochondrial or chloroplast DNA and taxonomy was assigned using the SILVA database v. 138.1 [38,39,40]. Singletons and doubletons were removed. For the detection of potential cross-contamination, we identified taxa that had higher total counts in any other sample library run on the same plate excluding the sample library analyzed in this study [34, 41]. These flagged taxa were then removed from the egg mass and sediment data when observed only in the contaminated samples from one sequencing run.
Additional 16S rRNA amplicon sequence datasets were obtained from NCBI’s SRA database (BioProject PRJNA282077 and PRJNA41903; from refs. [42, 43], respectively) to compare egg mass microbial community composition with the microbiota commonly found in the surrounding seagrass habitat. As 16S rRNA amplicon data from living seagrass at the sampling site was not available, we used amplicon sequences from seagrasses belonging to the genera Zostera and Cymodocea, which also occur in Mauritania. Specifically, external datasets contained 16S rRNA amplicon sequences from Zostera japonica (Netarts Bay, Oregon, United States), Zostera marina (Netarts Bay, Oregon, United States and Culatra Island, Portugal), Zostera noltei (Culatra Island, Portugal), and Cymodocea nodosa (Culatra Island, Portugal), as well as amplicon sequences from seagrass associated sediments and overlying seawater (Culatra Island, Portugal; Table S1). These sequence data were processed separately following the same protocol as previously described. The resulting ASV tables from all datasets were then merged using taxonomic assignments and the resulting table was normalized to account for compositional bias and differences in sequencing depth among samples with the Wrench package [44] using the W2 estimator.
All statistical analyses were conducted on the normalized ASV table using R v. 3.6.1 (ref. [37]). Diversity metrics, including Shannon–Wiener Diversity (H’) and Pielou’s evenness index (J), were calculated using the vegan package [45]. The distribution and homogeneity of variance in the data were determined using the Shapiro-Wilk Test and Levene’s Test to select the appropriate parametric (Student’s two-sided t-test) and nonparametric tests (Wilcoxon and Kruskal-Wallis Rank Sum). Differences in microbial community composition (normalized taxonomic group abundance and identity) were examined using non-metric multidimensional scaling ordination with Bray–Curtis distance in two dimensions. Community compositional differences due to sample type were assessed using permutational multivariate ANOVA (PERMANOVA) with the vegan package [45]. To identify taxonomic groups that had a significant association with egg masses, indicator species analysis was conducted using a point biserial correlation coefficient with the indicspecies package [46, 47] and resulting p values were adjusted using Benjamini–Hochberg’s False Discovery Rate (FDR correction).
Phylogenetic analysis of host and symbiont marker genes
For phylogenetic reconstruction of the host mitochondrial cytochrome B gene, 17 full-length sequences were extracted from lucinid metagenomes from the NCBI SRA database (BioProject PRJNA679177 from [48]). We assembled mitochondrial genomes from the metagenome read libraries to obtain peptide sequences of the cytochrome B gene, which has been established as a marker gene in lucinid molecular taxonomic studies [49]. Adapter-trimmed reads were used as input for mitochondrial genome assembly with Novoplasty on default parameters and using a publicly available L. orbiculatus mitochondrial genome (EF043341.1) as the seed sequence [50]. The mitochondrial genome assemblies were then functionally annotated using the MITOS2 webserver [51].
In addition to the metagenome-obtained sequences, we used two full-length cytB gene sequences from GenBank to generate a matrix (accession no. EF043341.1 and EF043342.1). In total, sequences of 15 Loripes individuals together with sequences of four individuals from three different lucinid genera: Lucinella, Clathrolucina, and Divalinga were used to generate a phylogenetic tree. The cytB sequence of Divalinga quadrisulcata was used as an outgroup. All metagenome-derived cytB sequences were translated into protein sequences in the software Geneious v. 11.0.3 [52] using the invertebrate mitochondrial translation table 5.
The symbiont 16S rRNA phylogeny was reconstructed using 26 near full-length sequences (>1300 bp), two of which originated from the aforementioned Bioproject PRJNA679177 while the remaining 24 sequences were obtained from GenBank (see results, Fig. S2). Allochromatium vinosum was used as the outgroup.
Symbiont sequences were aligned individually using MAFFT v7 [53] with the Q-INS-I algorithm [54] in Geneious. Host sequences were translated into AA sequences and aligned in MAFFT v7 with the E-INS-I algorithm. Host and symbiont phylogenies were reconstructed with the maximum likelihood-based tool IQTree [55] using the Ultrafast Bootstrap Approximation UFBoot [56] with 1000 bootstrap runs. In addition to maximum likelihood, support values were generated using approximate Bayes and SH-aLRT analyses [57, 58]. The TIM + F + I (for Maximum likelihood) and the TN + F + G4 (for Bayesian) substitution models were the best fit for the 16S alignment and the mtZOA+I (for Maximum likelihood) and mtZOA (for Bayesian) for the cytB alignment. Phylogenetic trees were visualized in iTol [59].
Results
Lucinids from Mauritania reproduce by releasing gelatinous egg masses
The early developmental stages of L. orbiculatus, encapsulated within gelatinous egg masses, were abundant and easily obtained during the low tide on the intertidal mudflats in Banc d’Arguin during the entire research stay (2 weeks in December 2015 and January 2017; Fig. 1). We observed that adult bivalves deposited transparent gelatinous egg masses, which were 3–4 cm in diameter and still connected to the parent’s mucus tube, by attaching them to the leaves of the seagrasses Zostera sp. and Cymodocea sp. (Fig. 1 and other personal observations). The matrix of the egg masses was characterized by a fibrous structure, which was made more conspicuous following fixation and likely comprises collagenous structural proteins such as was reported in the egg capsules of elasmobranchs [60] or the antimicrobial mucus of some heterobranch gastropods [61, 62]. The live egg masses were densely packed with trochophore and/or veliger larvae (Fig. 1). The larvae actively rotated and extended their velum within the egg masses, activities that became even more apparent after hatching.
“Hatched” pediveliger larvae that had left the egg mass were observed over a period of 54 days, during which they grew from ~200 µm to ~360 µm. Fully formed gill filaments were only seen in “hatched” individuals and never in larvae while still inside the egg mass (Fig. 2). No additional gill filaments were formed, nor did the filaments appear to grow over the course of 54 days in the laboratory (Fig. 2). The absence of gill filaments in “unhatched” larvae suggest that gill development is initiated only upon hatching.
Scale bars = 20 µm. A Individual larva prior to hatching removed from inside the egg mass and fixed. No gills are visible. B Larva fixed 10 days after hatching. Individual, small gill filaments are already detectable (highlighted in white box). C Larva 21 days after hatching, gills visible. D Larva 54 days after hatching from the egg mass. Gills are shown in white boxes for C and D.
The cytB sequences amplified from adult gills and egg masses were 98% identical to NCBI database entries for Loripes lacteus, a synonym for Loripes orbiculatus Poli [63] from Croatia. Phylogenetic reconstruction placed the metagenome-derived full-length cytB gene sequences from Mauritanian lucinids into a monophyletic, well-supported clade (95% bootstrap support) within the genus Loripes, although they were distinct from others sampled in Europe (Fig. 3). The closest relatives of the Mauritanian clams are those from Slovenia, Croatia, Montenegro, Spain and France. These findings suggest that there are genetic differences amongst different Mediterranean and Atlantic populations of Loripes orbiculatus, but further taxonomic analyses would be required to describe these differences in detail.
Support values are given in Bayesian Posterior Probability, Ultrafast Bootstrap and Maximum Likelihood SH-like aLRT; values ≥85%/0.9/95% are marked with black circle segments. Bivalves from Mauritania are highlighted in dark purple. The publicly available sequence of Divalinga quadrisulcata (SRS7805870) was used as the outgroup and the tree was visualized using IQ Tree.
Mauritanian lucinid symbionts belong to the species group Ca. Thiodiazotropha taylori
All of the 16S rRNA gene sequences amplified from the gills of adult Loripes from Mauritania were virtually identical (with single nucleotide changes) and belonged to the recently described lucinid symbiont species Candidatus Thiodiazotropha taylori [48]. The Mauritanian clam symbionts formed a highly supported monophyletic clade together with other described Ca. Thiodiazotropha in phylogenetic analysis of partial 16S rRNA genes (bootstrap support value = 94%). Like the hosts, and consistent with previous phylogenomic analyses, the closest relatives of this symbiont group were symbionts associated with the Mediterranean L. orbiculatus (ref. [48]; Fig. S1).
Diverse microbes were associated with the egg masses but the intracellular symbiont was not detected
A total of 479 ASVs belonging to 174 unique taxonomic groups were common to all egg mass samples. A similar number of taxonomic groups were found in both the microbial communities from egg mass pieces including larvae (161 groups) and egg mass pieces with the larvae removed (155 groups), indicating that these organisms possibly reside in the gelatinous mass rather than on the surface of or within the larvae. Microbial community composition, including taxonomic identity and relative abundance, did not differ significantly between egg mass pieces with and without larvae (PERMANOVA, R2 = 0.02, p > 0.05). Additionally, median diversity (H’) and evenness (J) were not significantly different (p = 0.35 and p = 0.50, respectively) between communities from egg masses with larvae present (H’ = 2.94; J = 0.90) and without (H’ = 3.21; J = 0.90). As no significant differences between the two sample types were observed, the egg mass samples with and without larvae were combined for further analyses.
Sequence reads from all egg mass samples were assigned to 21 taxonomic classes belonging to 12 phyla. The bacterial class Alphaproteobacteria (39.4%) comprised the largest portion of relative read abundance on average across all samples, followed by Clostridia (33.0%), Gammaproteobacteria (7.6%), Bacilli (3.7%), Bacteroidia (3.5%), Cyanobacteriia (4.1%), Saccharimonadia (2.4%), and Actinobacteria (1.5%; Fig. 4).
The remaining bacterial classes comprised less than 1% of relative read abundance. No ASVs belonging to the archaeal domain were detected in the egg mass microbial communities. The alphaproteobacterial ASVs assigned to the families Rhodobacteraceae and Rhizobiaceae comprised the largest proportions of the reads. Whereas the most abundant representatives from the Clostridia class belonged to the families Ruminococcaceae and Lachnospiraceae. None of the reads from egg mass samples were assigned to the genus Ca. Thiodiazotropha, which is consistent with the horizontal transmission of symbionts occurring at a later host life-stage (Fig. 5).
Egg masses are deposited by attaching them to seagrass blades while they are still connected to the mother’s mucus tube. The larvae inside the egg masses are aposymbiotic and undergo development via a trochophore (ref. [11]; ~8 days after egg mass release) and a veliger larval stage (ref. [11]; ~12 days after egg mass formation) before they hatch as mature pediveliger larvae. Gill development is initiated upon hatching and the sulfur-oxidizing endosymbionts colonize the settled juvenile. Environmental bacteria are shown in pink, blue, yellow, and brown. The specific sulfur-oxidizing symbiont is depicted in red. Broken lines indicate that we could not verify the presence of a membrane that separates the larvae inside the egg mass although such a barrier has been reported from other lucinid larvae [13]. Not shown to scale.
Microbial communities in the egg masses are distinct from those in other seagrass microhabitats
As seawater, seagrass leaves and rhizome samples from Mauritania were not available, we used 16S rRNA gene amplicon sequence data from other published studies to compare the microbial communities found in L. orbiculatus egg masses to microbial communities typically found in the different microhabitats in a seagrass meadow (seagrass leaves and rhizomes, sediment, and seawater; Table S1). There were 838 taxonomic groups identified in all samples from the seagrass microhabitats and egg masses (n = 114). Microbial community composition, including both taxonomic identity and abundance, was significantly different depending on microhabitat type (PERMANOVA, R2 = 0.25, p = 0.001; Fig. 6; Table S5). Alpha diversity was also significantly different between the habitat types; the sediment and seagrass root microbial communities had significantly higher diversity (H’) than the egg mass communities (Kruskal-Wallis, Χ2 = 62.19, p < 0.001; Dunn’s test, adjusted p < 0.05; Table S6). Sediment communities had the highest diversity of any habitat type. Sediment also had the highest number of taxonomic groups, and both sediment and seagrass root microbial communities contained more taxonomic groups than the egg mass communities (Kruskal-Wallis, Χ2 = 61.477, p < 0.001; Dunn’s test, adjusted p < 0.05; Table S6). However, the egg mass microbial communities had significantly higher evenness (J’) than both the seagrass root, seagrass leaf and sediment microbial communities (Kruskal-Wallis, Χ2 = 25.999, p = 0.001; Dunn’s test, adjusted p < 0.05; Table S6).
Sediment samples include material from three different depths: surface layer = 0–2 cm; intermediate layer = 2–5 cm; bottom layer = 5–8 cm. Seagrass rhizome debris sampled from the bottom layer was sieved and treated as a sediment sample. Analysis is based on data from Illumina sequencing and Bray–Curtis dissimilarities. Different study locations are given in different shapes whereas sample types are indicated by color (stress = 0.1796; k = 2).
While Ca. Thiodiazotropha was not detected in the egg mass microbial communities, sequences assigned to this genus were observed in the sediment microbial communities from Banc d’Arguin, Mauritania (n = 4) as well as in the seagrass rhizome (n = 19) and seawater communities (n = 1) from other locations. ASVs belonging to this genus comprised 0.88% of relative read abundance on average, ranging from 0.03 to 9.60% relative abundance in a single sample.
Indicator species analysis identified 51 taxonomic groups that showed a significant association with egg masses (point biserial correlation, adjusted p < 0.05) when compared with the seagrass environment. These 51 indicator taxa belonged to six bacterial phyla (Actinobacteriota, Bacteroidota, Cyanobacteria, Firmicutes, Patescibacteria, and Proteobacteria). The alphaproteobacterial genera Oceanicaulis and Tropicimonas had the strongest association with the egg mass environment. Many of the other indicator taxa also belonged to the class Alphaproteobacteria including eight that belonged to the Rhodobacterales order. The bacterial order Lachnospirales within the Firmicutes phylum was also frequently represented with 12 indicator taxa assigned to this taxonomic designation. Many of the indicator taxa we identified comprised the largest portions of relative abundance on average, including the genera Tuzzerella (3.29%) and Subdoligranulum (3.00%), which both belong to the Firmicutes phylum.
Discussion
In Banc d’Arguin, Mauritania, L. orbiculatus produces gelatinous egg masses, an intriguing phenomenon considering that bivalves rarely reproduce via “external” brooding. This provides further evidence of the diversity of reproductive strategies within the Lucinidae family [13, 22]. Previously, one spawning and one brooding species were reported, we now confirm earlier morphological reports that the species Loripes orbiculatus is capable of brooding via external gelatinous egg masses [11, 12, 22, 13]. Ovipositioning in sand-dwelling molluscs such as bivalves usually requires a suitable substrate to which the animals can attach their egg capsules or masses. In soft-bottom habitats such as intertidal mudflats however, the animals need to find alternatives to rocky substrates. Some animals have overcome these limitations by traveling long distances to find suitable places for oviposition while others deposit their egg capsules onto living organisms such as algae or conspecifics [64,65,66]. While reports of lucinid ovipositioning in nature are scarce [67], we observed that L. orbiculatus egg masses were mostly attached to seagrass leaves or the leaf stem. This suggests that in a habitat lacking adequate substrates for deposition it can be advantageous to attach the egg masses to alternative substrates such as seagrass leaves to enhance offspring survival. Mutualistic interactions between bivalves and seagrasses have been hypothesized based on metabolic interactions, but a reproductive benefit for lucinids of associating with seagrasses has not been previously considered [18, 68]. This discovery may explain the recent increase in the L. orbiculatus population as a result of a tremendous increase in seagrass cover in the study area [69].
Observing animal reproduction in nature remains challenging and we cannot exclude the possibility that different individuals belonging to the same bivalve species may be capable of both broadcast spawning and brooding, depending on external factors such as environmental conditions. In fact, the coexistence of different modes of reproduction within closely related, co-occurring marine invertebrates is common and well-documented from asteroids [70, 71], bivalves [72], pteropods [73], ophiuroids [74, 75], chitons [76], limpets [77, 78], and nereids [79]. It was even reported that both brooding and broadcasting exist within the same female individual of the sea star species Pteraster militaris [80]. Although the selection pressures driving the evolution of broadcast spawning versus brooding in Lucinidae are unclear, we would expect the mode of reproduction to have numerous consequences for the host’s biology. For example, broadcast spawners presumably disperse greater distances than brooders [1]. This may be detectable in the genetic population structure of the animal hosts, as brooders would be expected to show lower levels of population connectivity and genetic diversity compared to broadcast spawning species as was observed in closely related species of the brittle star Ophioderma longicauda [75].
Another consequence of egg masses is that they provide a new habitat for other microbes. We show that diverse microorganisms were associated with the egg masses, which form a microhabitat that was distinct from other seagrass environments. There were no significant differences in egg mass bacterial community compositions over time (December 2015 and January 2017), indicating that the population structure remained stable over a period of at least 1 year. This further suggests that certain conditions or factors intrinsic to the egg masses shape community composition.
Similar bacterial communities were found before and after larvae were removed from the egg masses which implies that microbes most likely associate with the gelatinous mass rather than the larvae. We observed an additional membrane or barrier around fertilized eggs similar to the capsules surrounding larvae in Phacoides pectinatus [13], but individual membranes were not visible at any later developmental stage (data not shown). These observations raise the possibility that larvae were in direct contact with bacteria in the gelatinous mass from an early developmental stage. Imaging studies such as FISH in the egg masses would be needed to pinpoint the exact location of the bacteria.
Phytoplankton blooms that preceed L. orbiculatus spawning events in May and between November and December [81, 24] might explain the abundance of cyanobacteria (predominantly from the order Phormidesmiales and Synechococcales) and diatoms in the egg masses. Some members of these orders supply oxygen to larvae in egg masses of various aquatic invertebrates and amphibians and are therefore thought to contribute to larval fitness [82,83,84]. Moreover, diatoms associate ubiquitously with alphaproteobacteria in many aquatic habitats [85, 86] and thus could play an important role for L. orbiculatus in the larval stage dominated by alphaproteobacterial groups. In addition to oxygen production, phototrophic microbes in the egg masses could provide an additional source of nutrition to the developing larvae [83, 87]. Reads belonging to the bacterial class Clostridia were surprisingly among the most abundant in the egg masses. Most of these representatives (e.g., Ruminococcaceae, Lachnospiraceae) are among the most abundant taxa in the gut microbiome of humans and other animals including birds [88]. Several migratory shorebird species (e.g., Calidris pusilla, C. alpina, C. canutus, Arenaria interpres and others) harbor Clostridia in their gastrointestinal tract from an early age on, which are thought to be acquired from the environment [89, 90]. The exact role of Clostridia in wild birds is still unknown but their occurrence is positively correlated with weight gain [91] and might influence the survival rates of migratory birds. The occurrence of Clostridia in egg masses could possibly be related to the presence of high numbers of these foraging shorebirds on the intertidal mudflat [92]. Understanding the functions of the egg mass-associated microbes will require additional studies; however, the known metabolic capabilities of microbes related to members of the egg mass community could shed light upon their successful colonization of the egg masses or help generate hypotheses about their potential functions (for more details see the Supplementary Discussion).
Although the egg masses host such diverse microbial communities, one particular microbe was conspicuous by its absence—the chemosynthetic symbiont that reaches extremely high abundance in all adult L. orbiculatus. This supports previous reports of horizontal transmission in another lucinid species, Codakia orbicularis, with environmental symbiont acquisition in a later life-stage of the host [22, 93]. Bacteria affiliated with the genus Ca. Thiodiazotropha are widespread in seagrass root environments; however, very few studies have directly compared the diversity of host-associated and free-living Ca. Thiodiazotropha at the same site [42, 43, 29]. We found exact matches between host-associated and free-living stages in sediment samples from Banc d’Arguin. These observations are all consistent with the acquisition of symbionts from the environment during development, independent of reproductive strategy (broadcast or brooding).
Conclusions
This study revealed differences in the reproductive biology and molecular identity of lucinid species that were previously assigned to a single species. The genetic diversity within the genus Loripes is therefore probably still underestimated. Moreover, we show that differences in the life cycle of lucinids and their mode of reproduction are not necessarily linked to differences in symbiont transmission mode, which is similar to patterns observed in corals where there is no strong link between reproductive mode of the host, and transmission mode of the symbionts [94]. Future studies will reveal the ecological drivers of these different reproductive modes, and their consequences for fundamental aspects of host biology such as dispersal and population structure. In addition, if the communities of specific microbes associated with the egg masses play beneficial roles for the host, this would indicate shifts in reliance on different microbes during ontogeny, as has been observed for animals ranging from insects and echinoderms to humans (e.g., [95,96,97]). Finally, our observation that Loripes egg masses attach to seagrasses adds important components in understanding the functioning of intertidal flats and the interplay between lucinids and seagrass beds.
Data availability
The 16S rRNA gene amplicon datasets generated in this study are available in the NCBI Sequence Read Archive under BioProject accession number PRJNA783529.
References
Barnes RSK, Calow PP, Olive PJ, Golding DW, Spicer JI. The invertebrates: a synthesis. 3rd ed. Wiley-Blackwell, Hoboken, NJ; 2009.
Thorson G. Reproductive and larval ecology of marine bottom invertebrates. Biol Rev. 1950;25:1–45.
Ebel R, Marin A, Proksch P. Organ-specific distribution of dietary alkaloids in the marine opisthobranch Tylodina perversa. Biochem Syst Ecol. 1999;27:769–77.
Benkendorff K, Davis AR, Bremner JB. Chemical defense in the egg masses of benthic invertebrates: an assessment of antibacterial activity in 39 mollusks and 4 polychaetes. J Invertebr Pathol. 2001;78:109–18.
Przeslawski R. A review of the effects of environmental stress on embryonic development within intertidal gastropod egg masses. Molluscan Res. 2004;24:43–63.
Breder CM, Rosen DE. Modes of reproduction in fishes. Natural History Press, Garden City, NY; 1966.
Palumbi SR, Johnson BA. A note on the influence of life-history stage on metabolic adaptation: the responses of Limulus eggs and larvae to hypoxia. Prog Clin Biol Res. 1982;81:115–24.
Whitaker DM. On the rate of oxygen consumption by fertilized and unfertilized eggs: IV. Chaetopterus and Arbacia punctulata. J Gen Physiol. 1933;16:475.
Strathmann MF. Reproduction and development of marine invertebrates of the northern Pacific coast: data and methods for the study of eggs, embryos, and larvae. University of Washington Press, Seattle; 1987.
Fretter V, Graham A. British prosobranch molluscs. Their functional anatomy and ecology. Ray Society, London; 1962.
Pelseneer P. Notes d’embryologie Malacologiques. Bull Biol Fr Belg. 1926; 88–112.
Zakhvatkina KA. Larvae of bivalve mollusks of the Sevastopol region of the Black Sea. AN SSSR. Trudy Sevastoplskoi Biologicheskoi Stantsii. 1959;11:108–152. [In Russian, Transl. by Evelyn Wells, Virginia Inst. Mar. Sci. Transl. Series 1966;15:1–41]
Collin R, Giribet G. Report of a cohesive gelatinous egg mass produced by a tropical marine bivalve. Invertebr Biol. 2010;129:165–71.
Taylor JD, Glover EA. Lucinidae (Bivalvia)–the most diverse group of chemosymbiotic molluscs. Zool J Linn Soc. 2006;148:421–38.
Cary SC, Vetter RD, Felbeck H. Habitat characterization and nutritional strategies of the endosymbiont-bearing bivalve Lucinoma. Mar Ecol Prog Ser. 1989;55:31–45.
Le Pennec M, Beninger PG, Herry A. Feeding and digestive adaptations of bivalve molluscs to sulphide-rich habitats. Comp Biochem Phys A. 1995;111:183–9.
van der Geest M, van Gils JA, van der Meer J, Olff H, Piersma T. Suitability of calcein as an in-situ growth marker in burrowing bivalves. J Exp Mar Biol Ecol. 2011;399:1–7.
van der Heide T, Govers LL, de Fouw J, Olff H, van der Geest M, van Katwijk MM, et al. A three-stage symbiosis forms the foundation of seagrass ecosystems. Science. 2012;336:1432–4.
van Gils JA, van der Geest M, Leyrer J, Oudman T, Lok T, Onrust J, et al. Toxin constraint explains diet choice, survival and population dynamics in a molluscivore shorebird. Proc R Soc B. 2013;280:20130861.
Higgs ND, Newton J, Attrill MJ. Caribbean spiny lobster fishery is underpinned by trophic subsidies from chemosynthetic primary production. Curr Biol. 2016;26:3393–8.
Fisher CR. Chemoautotrophic and methanotrophic symbioses in marine invertebrates. Rev Aquat Sci. 1990;2:399–436.
Gros O, Frenkiel L, Mouëza M. Embryonic, larval, and post-larval development in the symbiotic clam Codakia orbicularis (Bivalvia: Lucinidae). Invertebr Biol. 1997;116:86–101.
Allen JA. On the basic form and adaptations to habitat in the Lucinacea (Eulamellibranchia). Philos Trans R Soc B. 1958;241:421–84.
van der Geest M, Sall AA, Ely SO, Nauta RW, van Gils JA, Piersma T. Nutritional and reproductive strategies in a chemosymbiotic bivalve living in a tropical intertidal seagrass bed. Mar Ecol Prog Ser. 2014;501:113–26.
Steedman H. Polyester Wax: a new ribboning embedding medium for histology. Nature. 1957;179:1345.
Klose J, Polz MF, Wagner M, Schimak MP, Gollner S, Bright M. Endosymbionts escape dead hydrothermal vent tubeworms to enrich the free-living population. Proc Natl Acad Sci USA. 2015;112:11300–5.
Amann RI, Binder BJ, Olson RJ, Chisholm SW, Devereux R, Stahl DA. Combination of 16S rRNA-targeted oligonucleotide probes with flow cytometry for analyzing mixed microbial populations. Appl Environ Microbiol. 1990;56:1919–25.
Daims H, Brühl A, Amann R, Schleifer K-H, Wagner M. The domain-specific probe EUB338 is insufficient for the detection of all Bacteria: Development and evaluation of a more comprehensive probe set. Syst Appl Microbiol. 1999;22:434–44.
Martin BC, Middleton JA, Fraser MW, Marshall IP, Scholz VV, Hausl B, et al. Cutting out the middle clam: lucinid endosymbiotic bacteria are also associated with seagrass roots worldwide. ISME J. 2020;14:2901–5.
Taylor JD, Glover EA, Williams ST. Ancient chemosynthetic bivalves: systematics of Solemyidae from eastern and southern Australia (Mollusca: Bivalvia). Mem Queensl Mus. 2008;54.
Taylor JD, Glover EA, Smith L, Dyal P, Williams ST. Molecular phylogeny and classification of the chemosymbiotic bivalve family Lucinidae (Mollusca: Bivalvia). Zool J Linn Soc. 2011;163:15–49.
Juretschko S, Timmermann G, Schmid M, Schleifer KH, Pommerening-Röser A, Koops HP, et al. Combined molecular and conventional analyses of nitrifying bacterium diversity in activated sludge: Nitrosococcus mobilis and Nitrospira-like bacteria as dominant populations. Appl Environ Microb. 1998;64:3042–51.
Lane DJ. 16S/23S rRNA sequencing. In: Stackebrandt E, Goodfellow M (eds). Nucleic acid techniques in bacterial systematics. John Wiley and Sons, New York; 1991. p. 115–75.
Herbold C, Pelikan C, Kuzyk O, Hausmann B, Angel R, Berry D, et al. A flexible and economical barcoding approach for highly multiplexed amplicon sequencing of diverse target genes. Front Microbiol. 2015;6:731.
Klindworth A, Pruesse E, Schweer T, Peplies J, Quast C, Horn M, et al. Evaluation of general 16S ribosomal RNA gene PCR primers for classical and next-generation sequencing-based diversity studies. Nucleic Acids Res. 2013;41:e1–e1.
Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJA, Holmes SP. DADA2: High-resolution sample inference from Illumina amplicon data. Nat Methods. 2016;13:581–3.
R Core Team. R: a language and environment for statistical computing. R Foundation for Statistical Computing, https://www.r-project.org. Vienna, Austria; 2019.
Quast C, Pruesse E, Yilmaz P, Gerken J, Scweer T, Yarza P, et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucl Acids Res. 2013;41:590–6.
Yilmaz P, Parfrey LW, Yarza P, Gerken J, Pruesse E, Quast C, et al. The SILVA and “All-species Living Tree Project (LTP)” taxonomic frameworks. Nucleic Acids Res. 2014;42:643–8.
McLaren MR, Callahan BJ. Silva 138.1 prokaryotic SSU taxonomic training data formatted for DADA2. 2021. https://doi.org/10.5281/zenodo.4587955. Accessed 22 Nov 2021.
Lesaulnier CC, Herbold CW, Pelikan C, Berry D, Gérard D, Le Coz X, et al. Bottled aqua incognita: microbiota assembly and dissolved organic matter diversity in natural mineral waters. Microbiome. 2017;5:1–17.
Cucio C, Engelen A, Costa R, Muyzer G. Rhizosphere microbiomes of european seagrasses are selected by the plant but are not species specific. Front Microbiol. 2016;7:440.
Crump BC, Wojahn JM, Tomas F, Mueller RS. Metatranscriptomics and amplicon sequencing reveal mutualisms in seagrass microbiomes. Front Microbiol. 2018;9:388.
Kumar R, Jain A, Khatait JP. Optimal reaction wrench measuring platform. In: Proceedings of ASME International Mechanical Engineering Congress and Exposition. Volume 13: Design, reliability, safety, and risk. Pittsburgh, Pennsylvania, USA; 2018. https://doi.org/10.1115/IMECE2018-87057.
Oksanen J, Blanchet FG, Friendly M, Kindt R, Legendre P, McGlinn D, et al. Vegan: Community Ecology Package. R package Version 2.5-6. 2019. https://CRAN.R-project.org/package=vegan. Accessed 15 Aug 2021.
De Cáceres M, Legendre P. Associations between species and groups of sites: indices and statistical inference. Ecol. 2009;90:3566–74.
De Cáceres M, Jansen F, Dell N. Indicspecies Package: Relationship between species and groups of sites. Version 1.7.9. 2020. https://vegmod.github.io/software/indicspecies.
Osvatic JT, Wilkins LG, Leibrecht L, Leray M, Zauner S, Polzin J, et al. Global biogeography of chemosynthetic symbionts reveals both localized and globally distributed symbiont groups. Proc Natl Acad Sci USA. 2021; https://doi.org/10.1073/pnas.2104378118.
Taylor JD, Glover EA. Unloved, paraphyletic or misplaced: new genera and species of small to minute lucinid bivalves and their relationships (Bivalvia, Lucinidae). ZooKeys. 2019;899:109.
Dierckxsens N, Mardulyn P, Smits G. NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017;45:e18–e18.
Donath A, Jühling F, Al-Arab M, Bernhart SH, Reinhardt F, Stadler PF, et al. Improved annotation of protein-coding genes boundaries in metazoan mitochondrial genomes. Nucleic Acids Res. 2019;47:10543–52.
Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, et al. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28:1647–9.
Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30:772–80.
Katoh K, Toh H. Recent developments in the MAFFT multiple sequence alignment program. Brief Bioinform. 2008;9:286–98.
Nguyen LT, Schmidt HA, von Haeseler A, Minh BQ. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol. 2015;32:268–74.
Minh BQ, Nguyen MAT, von Haeseler A. Ultrafast approximation for phylogenetic bootstrap. Mol Biol Evol. 2013;30:1188–95.
Anisimova M, Gil M, Dufayard JF, Dessimoz C, Gascuel O. Survey of branch support methods demonstrates accuracy, power, and robustness of fast likelihood-based approximation schemes. Syst Biol. 2011;60:685–99.
Guindon S, Dufayard JF, Lefort V, Anisimova M, Hordijk W, Gascuel O. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol. 2010;59:307–21.
Letunic I, Bork P. Interactive Tree Of Life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021;49:293–6.
Wourms JP. Reproduction and development in chondrichthyan fishes. Am Zool. 1977;17:379–410.
Abdillahi SM, Maaß T, Kasetty G, Strömstedt AA, Baumgarten M, Tati R, et al. Collagen VI contains multiple host defense peptides with potent in vivo activity. J Immunol. 2018;201:1007–20.
Pitt SJ, Hawthorne JA, Garcia-Maya M, Alexandrovich A, Symonds RC, Gunn A. Identification and characterisation of anti-Pseudomonas aeruginosa proteins in mucus of the brown garden snail, Cornu aspersum. Br J Biomed Sci. 2019;76:129–36.
Poli JX. Testacea utriusque siciliae eorumque historia et anatome tabulis aeneis illustrata. Parma, Regio Typographeio. Vol. 2, p. ixlix [1-2], 75-264, pls 19-39. 1795. https://www.biodiversitylibrary.org/page/44019644.
Barnett PRO, Hardy BLS, Watson J. Substratum selection and egg-capsule deposition in Nassarius reticulatus (L.). J Exp Mar Biol Ecol. 1980;45:95–103.
Rittschof D, Sawardecker P, Petry C. Chemical mediation of egg capsule deposition by mud snails. J Chem Ecol. 2002;28:2257–69.
von Dassow YJ, Strathmann RR. Full of eggs and no place to lay them: hidden cost of benthic development. Mar Ecol Prog Ser. 2005;294:23–34.
van der Geest M. Multi-trophic interactions within the seagrass beds of Banc d’Arguin, Mauritania. A chemosynthesis-based intertidal ecosystem. PhD [dissertation]. Groningen: University of Groningen; 2013. https://pure.rug.nl/ws/portalfiles/portal/13681090.
Cardini U, Bartoli M, Lücker S, Mooshammer M, Polzin J, Lee RW, et al. Chemosymbiotic bivalves contribute to the nitrogen budget of seagrass ecosystems. ISME J. 2019;13:3131–34.
El-Hacen EM, Sidi Cheikh MA, Bouma TJ, Olff H, Piersma T. Long-term changes in seagrass and benthos at Banc d’Arguin, Mauritania, the premier intertidal system along the East Atlantic Flyway. Glob. Ecol Conserv. 2020;24:e01364.
Menge BA. Brood or broadcast? The adaptive significance of different reproductive strategies in the two intertidal sea stars Leptasterias hexactis and Pisaster ochraceus. Mar Biol. 1975;31:87–100.
Emson RH, Crump RG. Description of a new species of Asterina (Asteroidea), with an account of its ecology. J Mar Biol Assoc UK. 1979;59:77–94.
Sastry AN. Pelecypoda (Excluding Ostreidae). In: Giese AC, Pearse JS (eds). Reproduction of marine invertebrates. Academic Press, New York; 1979. p. 113–292.
Lalli CM, Wells FE. Brood protection in an epipelagic thecosomatous pteropod, Spiratella (“Limacina”) Inflata (D'Orbigny). Bull Mar Sci. 1973;23:933–41.
Hendler G. Sex-reversal and viviparity in Ophiolepis kieri, n. sp., with notes on viviparous brittlestars from the Caribbean (Echinodermata: Ophiuroidea). Proc Biol Soc Wash. 1979;92:783–95.
Weber AT, Mérigot B, Valière S, Chenuil A. Influence of the larval phase on connectivity: strong differences in the genetic structure of brooders and broadcasters in the Ophioderma longicauda species complex. Mol Ecol. 2015;24:6080–94.
Pearse JS. Polyplacophora. In: Giese AC, Pearse JS (eds). Reproduction of marine invertebrates, Volume 5, Molluscs: pelecypods and lesser classes. Academic Press, New York;1979. p. 27–85.
Gallardo CS. Two modes of development in the morphospecies Crepidula dilatata (Gastropoda: Calyptraeidae) from Southern Chile. Mar Biol. 1977;39:241–51.
Collin R. The effects of mode of development on phylogeography and population structure of North Atlantic Crepidula (Gastropoda: Calyptraeidae). Mol Ecol. 2001;10:2249–62.
Lucey NM, Lombardi C, DeMarchi L, Schulze A, Gambi MC, Calosi P. To brood or not to brood: are marine invertebrates that protect their offspring more resilient to ocean acidification? Sci Rep. 2015;5:1–7.
McClary DJ, Mladenov PV. Reproductive pattern in the brooding and broadcasting sea star Pteraster militaris. Mar Biol. 1989;103:531–40.
Le Pennec M, Herry A, Johnson M, Beninger P. Nutrition-gametogenesis relationship in the endosymbiont host-bivalve Loripes lucinalis (Lucinidae) from reducing coastal habitats. In: Eleftheriou A, Ansell AD, Smith CJ (eds). Biology and ecology of shallow coastal waters. Olsen and Olsen, Fredensborg; 1995. p. 139–42.
Pinder A, Friet S. Oxygen transport in egg masses of the amphibians Rana sylvatica and Ambystoma maculatum: convection, diffusion and oxygen production by algae. J Exp Biol. 1994;197:17–30.
Strathmann RR. Form, function, and embryonic migration in large gelatinous egg masses of arenicolid worms. Invertebr Biol. 2000;119:319–28.
Peyton KA, Hanisak MD, Lin J. Marine algal symbionts benefit benthic invertebrate embryos deposited in gelatinous egg masses. J Exp Mar Biol Ecol. 2004;307:139–64.
Schäfer H, Abbas B, Witte H, Muyzer G. Genetic diversity of ‘satellite’ bacteria present in cultures of marine diatoms. FEMS Microbiol Ecol. 2002;42:25–35.
Grossart HP, Levold F, Allgaier M, Simon M, Brinkhoff T. Marine diatom species harbour distinct bacterial communities. Environ Microbiol. 2005;7:860–73.
Sivakumar N, Sundararaman M, Selvakumar G. Efficacy of micro algae and cyanobacteria as a live feed for juveniles of shrimp Penaeus monodon. Afr J Biotechnol. 2011;10:11594–9.
Stanley D, Hughes RJ, Geier MS, Moore RJ. Bacteria within the gastrointestinal tract microbiota correlated with improved growth and feed conversion: challenges presented for the identification of performance-enhancing probiotic bacteria. Front Microbiol. 2016;7:187.
Ryu H, Grond K, Verheijen B, Elk M, Buehler DM, Santo Domingo JW. Intestinal microbiota and species diversity of Campylobacter and Helicobacter spp. in migrating shorebirds in Delaware Bay. Appl Environ Microb. 2014;80:1838–47.
Grond K, Lanctot RB, Jumpponen A, Sandercock BK. Recruitment and establishment of the gut microbiome in arctic shorebirds. FEMS Microbiol Ecol. 2017;93:fix142.
Angelakis E, Raoult D. The increase of Lactobacillus species in the gut flora of newborn broiler chicks and ducks is associated with weight gain. PLoS One. 2010;5:e10463.
Oudman T, Schekkerman H, Kidee A, van Roomen M, Camara M, Smit C, et al. Changes in the waterbird community of the Parc National du Banc d’Arguin, Mauritania, 1980–2017. Bird Conserv Int. 2020;30:618–33. https://doi.org/10.1017/S0959270919000431.
Gros O, Duplessis MR, Felbeck H. Embryonic development and endosymbiont transmission mode in the symbiotic clam Lucinoma aequizonata (Bivalvia: Lucinidae). Int J Inver Rep Dev. 1999;36:93–103.
Hartmann AC, Baird AH, Knowlton N, Huang D. The paradox of environmental symbiont acquisition in obligate mutualisms. Curr Biol. 2017;27:3711–6.
Dominguez-Bello MG, Godoy-Vitorino F, Knight R, Blaser MJ. Role of the microbiome in human development. Gut. 2019;68:1108–14.
Nyholm SV. In the beginning: egg–microbe interactions and consequences for animal hosts. Philos Trans R Soc B. 2020;375:20190593 https://doi.org/10.1098/rstb.2019.0593.
Carrier TJ, Reitzel AM. Symbiotic life of echinoderm larvae. Front Ecol Evol. 2020;7:509.
Acknowledgements
We thank the entire team among Han Olff, Theunis Piersma, Anne Dekinga and Job ten Horn as well as the staff at the station in Parc National du Banc d’Arguin for logistically supporting us before, during and after sampling in Iwik (Mauritania). Thanks to Jan van Gils for involving us in research at Iwik, to Martin Kunert for support in the lab in Vienna and to Belinda Martin of Ooid Scientific for illustration. We thank our funding sources including the Austrian Academy of Sciences (ÖAW) for a DOC Fellowship to SZ, the Aqua Terra Zoo Vienna (Haus des Meeres) for awarding the Hans and Lotte Hass Prize for Marine Ecology to SZ, the Vienna Science and Technology Fund (WWTF) for a Vienna Research Group grant to JMP, the European Research Council (ERC) for the Starting Grant 802494 EvoLucin to JMP and the Austrian Science Fund (FWF) for the project MAINTAIN DOC 69 doc.fund (to JMP and SZ) and project 31010 to MM.
Author information
Authors and Affiliations
Contributions
SZ and JMP conceptualized this study. SZ, MM, JP, and EH-ME-H. did work in the field. SZ and JP performed work in the laboratory. SZ, MM and BY performed phylogenetic and 16S rRNA analyses. All authors wrote and edited the paper.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Zauner, S., Vogel, M., Polzin, J. et al. Microbial communities in developmental stages of lucinid bivalves. ISME COMMUN. 2, 56 (2022). https://doi.org/10.1038/s43705-022-00133-4
Received:
Revised:
Accepted:
Published:
DOI: https://doi.org/10.1038/s43705-022-00133-4