Skip to main content
  • Original Article
  • Published:

Antimicrobial activity and diversity of bacteria associated with Taiwanese marine sponge Theonella swinhoei

Abstract

Marine sponges often rely on other epiphytes for protection from harmful predators. To understand the diversity and antimicrobial activity present among epiphytic bacteria isolated from marine sponge. We used both the 16S rRNA tag pyrosequencing method and the culture-based method to investigate the bacterial communities of Theonella swinhoei collected off the shore of southern Taiwan. Eight-hundred and eighteen operational taxonomic units (OTUs; 97% sequence similarity) were identified from 23,700 sponge-derived sequence tags. The bacteria associated with T. swinhoei were found to be highly diverse—as many as 12 different phyla of bacteria were identified. However, in terms of population evenness, the community was dominated by two phyla—Acidobacteria (71.54%) and Chloroflexi (19.60%). A total of 700 bacterial strains were isolated and cultured from samples of the sponge T. swinhoei. Within these culturable strains, only 12% were Actinomycetes. Despite the low percentage of Actinobacteria from the samples, among the 51 strains of culturable bacteria that showed high antimicrobial activity, a great majority (62%) were Actinomycetes (30 strains of Streptomyces and 1 strain each of Micromonospora and Brevibacterium). The remaining isolates that produced antimicrobial compounds were Gammaproteobacteria (10 strains of Pseudoalteromonas) and Firmicutes (8 and 1 strains of Bacillus and Paenibacillus, respectively). We speculated that many more Actinomycetes are yet to be isolated from T. swinhoei microbiota. Advanced techniques, such as high-throughput culture and culturome, should allow the isolation and purification of these medically important groups of bacteria from sponge.

Introduction

The ocean carries many microbes, most of which are attached to other marine organisms (Burgess et al. 1999; Nithyanand et al. 2011). Sponges host a large community of microorganisms, including various bacteria, fungi, unicellular algae, and archaea (Taylor et al. 2007; Webster and Taylor 2012). The biomass of microbes in some sponges can constitute up to 40–60% of the host’s total biomass (Hill et al. 2006). Many of these microbial symbionts are photosynthetic, whereas others are nitrogen fixers and nitrifiers. They form a complex community to support the many nutritional, health, and defense needs of the sponge (Bayer et al. 2008; Flatt et al. 2005; Hoffmann et al. 2009; Osinga et al. 2001; Thomas et al. 2010; Wilkinson and Fay 1979; Wilkinson et al. 1981). Although most of these microbes are yet to be cultured, recent massive parallel DNA sequencing studies have provided new insight into bacterial communities in sponges. Sponges are associated with more than 47 bacterial phyla (Reveillaud et al. 2014). Very often, genetically related sponges collected from very different geographical locations harbor a similar microbiota, suggesting the presence of specific associations between sponges and bacterial symbionts (Reveillaud et al. 2014; Thoms et al. 2003).

Marine organisms may produce chemicals of great value (Leal et al. 2012; Mehbub et al. 2014). Accumulated evidence has shown that some natural products formerly attributed to being produced by the sponge host are actually produced by symbiotic bacteria (Unson and Faulkner 1993; Unson et al. 1994). Moreover, it has also been demonstrated that more than 80% of FDA-approved drugs or pharmaceutical agents in clinical trials isolated from marine organisms are predicted to originate from associated microbes (Gerwick and Moore 2012). Sponges are a primitive group of ocean animals belonging to the phylum Porifera. Many antimicrobial compounds have been isolated from sponge-associated microbes (Graca et al. 2013; Hentschel et al. 2001; Thomas et al. 2010; Unson and Faulkner 1993; Unson et al. 1994). Over the years, researchers have noticed that new species of different genera of Actinobacteria, commonly present in marine sponges, can produce novel antimicrobial compounds (Abdelmohsen et al. 2014). Graca et al. (2013) studied the activities of bacterial symbionts from the sponge Erylus discophorus, and reported that as many as 31% of the isolates could produce antimicrobial metabolites. Among the antimicrobial-producing bacteria, Pseudovibrio, Vibrio, and Bacillus are the three most common genera. On the other hand, Pseudoalteromonas, a member of the γ-Proteobacteria family, and many α-Proteobacteria are the most abundant bacteria that produce antimicrobial compounds from sponges Aplysina aerophoba and Aplysina cavernicola (Hentschel et al. 2001).

Marine sponge Theonella swinhoei (class Demospongiae, order Tetractinellida, family Theonellidae) is distributed throughout the Indo-West Pacific. Forty percent of the biomass of the sponge is contributed by microbes (Jin et al. 2014; Keren et al. 2015). Natural T. swinhoei sponges produce products such as bioactive polyketides (Bewley et al. 1996; Piel et al. 2004) and glycopeptides (Schmidt et al. 2000). The bacterial community inhabiting T. swinhoei has been studied using the clone library technique (Hentschel et al. 2002) and culture-based methods (Keren et al. 2015; Lavy et al. 2014). While these methods can identify the dominant members of the community, many of the low-abundance taxa in the community are likely to have been missed. Pyrosequencing DNA technology is an ultra-deep DNA sequencing method that provides a large number of DNA sequence reads in a single run. When this deep sequencing method is applied to detect 16S DNA in environmental samples, many low-abundance taxa can be identified. We analyzed the diversity of bacteria inhabiting T. swinhoei using the tagged 16S rRNA pyrosequencing DNA sequencing method, and isolated bacteria, with a particular focus on actinomycetes with antimicrobial activities from T. swinhoei.

Materials and methods

Sponge collection and processing

Specimens of sponge T. swinhoei were collected during February 2013 at the inlet of a nuclear power plant within Kenting National Park, southern Taiwan (21°94′42.60″N, 120°79′31.97″E) by hand at a depth of about 3–5 m below sea level. The samples were placed in sterile plastic bags in a cold water environment for transportation to the laboratory immediately after collection. Three pieces of sponge tissue (each approximately 0.5 g wet weight) from different parts of one sponge were cut with a sterile scalpel, thoroughly washed twice with filtered (0.22 μm) seawater, pooled together and homogenized with 5 ml filtered seawater in a sterile mortar, then transferred to a 50-ml centrifuge tube. The sponge homogenate prepared was used for bacterial community DNA extraction and sponge-associated bacteria cultivation.

Bacterial community DNA extraction and 16S rRNA gene tag sequencing

Total genomic DNA for two T. swinhoei samples, A and B, was extracted and purified using a GeneMark genomic DNA purification kit (GeneMark, Taipei, Taiwan) according to the manufacturer’s protocol. PCR, amplicon purification, and quality assessment, as well as tag sequencing, were performed at Welgene Biotech (Taipei, Taiwan) following the protocols described previously (Hentschel et al. 2002; Lavy et al. 2014). Samples A and B are collected at the same place and the location distance between them is less than 1 m. Primer sets 27F (5′-AGAGTTTGATCCTGGCTCAG-3′) and 533R (5′-TTACCG CGGCTGCTGGCAC-3′) were used to amplify the V1–V3 hypervariable regions of bacterial 16S rRNA genes. 16S rRNA gene tag sequencing was performed on a Roche 454 Life Science FLX pyrosequencer (454 Life Sciences, Barnford, CT, USA).

Pyrosequencing data handling and bioinformatics analysis

All trimming, clustering, and classification were performed using the Mothur software package (Reveillaud et al. 2014; Schloss et al. 2011, 2009). Prior to cluster analysis, sequences that had the low-quality reads (reads shorter than 200 bp and with an average quality score above 25) (Simister et al. 2012), had homopolymer longer than 8 bp, and contained chimeras identified using the chimer.uchime command in Mothur were trimmed from analysis (Schloss et al. 2011, 2009). The clean reads were then aligned against the SILVA database (release 119) (Pruesse et al. 2007) and clustered into operational taxonomic units (OTUs) at a 3% dissimilarity cutoff point based on the furthest neighbor-clustering algorithm. The taxonomic position of each OTU was assigned from the SILVA database (Pruesse et al. 2007). The RDP-classifier was used to assign a phylotype for representative sequence from each OTU, with an 80% confidence threshold (Cole et al. 2005). Alpha and beta diversity metrics, including the Shannon-Weaver diversity index, the Chao1 richness estimator, and the abundance-based coverage estimator (ACE) as well as the Bray-Curtis dissimilarity, the Morisita-Horn dissimilarity, and the Jaccard dissimilarity, were also calculated using Mothur software. The coverage was estimated using program Nonpareil version 3.3, using alignment algorithm (Rodriguez-R and Konstantinidis 2014).

Sequences similarity search between 16S rRNA genes of isolates and OTUs as well as OTUs and SILVA database were conducted using stand-alone BLAST+ program (version 2.2.30) from NCBI (Camacho et al. 2009).

Sponge-associated bacteria cultivation

Seven different selective media, soil agar (SA), Gause modified agar I (GIA), actinomycete isolation agar (AIA), M1A agar (M1A), glucose-peptone-yeast extract agar (GPYA), Gause mineral agar (GHA), and peptone-yeast extract agar (PYA), were used to culture and isolate bacteria from the sponge homogenate. The compositions of these media were as follows: soil agar—1 ml soil extract (200 g soil mixed in 1 l 50% seawater, followed by boiling for 30 min), 1 ml vitamin complex solution (0.5 mg/ml calcium pantothenate, 0.5 mg/ml nicotinic acid, 0.05 mg/ml thiamin chloride, 0.05 mg/ml biotin), 0.05 g/l nalidixic acid, 0.025 g/l nystatin, and 20 g/l agar (Bredholdt et al. 2007); Gause modified agar I—20 g/l soluble starch, 0.5 g/l MgSO4·7H2O, 0.5 g/l K2HPO4, 1 g/l KNO3, 0.5 g/l NaCl, 0.01 g/l FeSO4·7H2O, 0.05 g/l nalidixic acid, 0.025 g/l nystatin, and 20 g/l agar (Terekhova et al. 1991); actinomycete isolation agar—22 g/l actinomycete isolation broth, 0.025 g/l nystatin, 0.05 g/l cycloheximide, and 20 g/l agar; M1A agar—10 g/l starch, 4 g/l yeast extract, 2 g/l peptone, 0.05 g/l nystatin, 0.05 g/l cycloheximide, and 20 g/l agar (Jensen et al. 2005); glucose-peptone-yeast extract agar—20 g/l glucose, 5 g/l peptone, 5 g/l yeast extract, 0.05 g/l streptomycin, and 20 g/l agar; Gause mineral agar—20 g/l soluble starch, 1 g/l KNO3, 0.5 g/l MgSO4·7H2O, 0.5 g/l K2HPO4, 0.01 g/l FeSO4·7H2O, 0.025 g/l nystatin, 0.05 g/l cycloheximide, and 20 g/l agar (Ivanitskaia et al. 1978); and peptone-yeast extract agar—5 g/l peptone, 5 g/l yeast extract, 0.025 g/l nystatin, 0.05 g/l cycloheximide, and 20 g/l agar. All the above media were supplemented with 50% sea water to increase the chance of isolating bacteria adapted to the marine environment. After plating of the sponge homogenate, the plates were incubated at 25 °C for 5 days to 2 weeks. Individual colonies were chosen and streaked onto fresh M1 agar (M1A without nystatin and cycloheximide) until pure cultures were obtained. For long-term preservation, cultures were grown at 25 °C in M1 broth and stored at − 80 °C in 20% glycerol solution.

Antimicrobial activity screening

Two screening steps, primary and secondary screening, were used to screen sponge isolates for antimicrobial activity. Primary screening was performed to select preliminary antimicrobial isolates using the agar block method (Chen et al. 2012; Stern et al. 2006) against a Gram-negative strain (Vibrio parahaemolyticus), a Gram-positive strain (Staphylococcus aureus), and a yeast (Candida albicans). Isolates that exhibited an inhibition zone > 1 mm in diameter against at least one test strain in primary screening were selected for secondary screening using the agar block method (Stern et al. 2006) against five test bacteria, namely Escherichia coli, V. parahaemolyticus, Pseudomonas aeruginosa, S. aureus, and C. albicans.

Test microorganisms were cultured as follows: V. parahaemolyticus (BCRC 10806) and P. aeruginosa (BCRC 10303) were cultured in marine broth at 25 °C for 24 h and maintained on marine agar; C. albicans (BCRC 22903) was cultured in YM broth at 30 °C for 48 h and maintained on YM agar; and S. aureus (ATCC 12600) and E. coli (ATCC 11775) were cultured in LB broth at 37 °C for 18 h and maintained on LB agar.

Crude extract preparation

Sponge isolates that exhibited antimicrobial activity were cultured in 500-ml flasks containing 100 ml M1 medium with 50% seawater. Flasks were incubated at 25 °C on a rotatory shaker at 150 rpm. After 5 days of incubation, the fermented broths were extracted twice with ethyl acetate (100 ml × 2). The solvent extracts were combined and evaporated to dryness under vacuum. The extracts obtained were weighed and stored at − 20 °C. The extracts were then used for cytotoxic activity assays.

Detection of cytotoxic activity

The cytotoxic activity of the crude extract was studied by MTT assays (Lu et al. 2009). The cell lines used were cancer cell lines MCF-7, MDA-MB-231, and T47-D from breast tumors, MOLT-4 and K-562 from leukemia, and DLD-1 and HCT-116 from colorectal cancer. All cell lines were purchased from the ATCC collection. Cells were maintained in RPMI 1640 medium supplemented with 10% fetal calf serum, 2 mM glutamine, and antibiotics (100 units/ml penicillin and 100 μg/ml streptomycin) at 37 °C in a humidified atmosphere of 5% CO2.

Gene sequencing and phylogenetic analysis

Isolates that exhibited antimicrobial activity were identified by their 16S rRNA gene sequence. A 16S rRNA fragment of each strain was amplified using the universal 16S rRNA bacterial primer pair 27F (5′-GAGAGTTTGATCCTGGCTCAG-3′) and 1492R (5′-CTACGGCTACCTTGTTACGA-3′). Sequencing was performed by Tri-I Biotech (Taipei, Taiwan).

A neighbor-joining tree was constructed with Kimura’s two-parameter model using MEGA software ver. 5.0 (Tamura et al. 2011). The statistical significance of the resultant tree topology was evaluated by bootstrap analysis, involving the construction of 1000 trees from randomly resampled data.

Nucleotide sequence accession numbers

The nucleotide sequences of isolates obtained in this study have been deposited in the GenBank database under accession numbers from MH311883 to MH311933. The raw pyrosequencing reads obtained in this study were deposited in the NCBI Sequence Read Archive (SRA) with the accession number SRP145405.

Results

Diversity of sponge-associated bacteria with pyrosequencing data

To investigate the diversity of bacteria associated with T. swinhoei at a deeper level, total DNA obtained from sponge extracts was used as a template for 16S rDNA tag pyrosequencing. A total of 27,548 raw gene sequence reads (approximately 13,179,125 bp) spanning the V1–V3 variable regions were obtained from two T. swinhoei samples, samples A and B. Table 1 presents a summary of the OTUs, Coverage, Shannon-Weaver and Simpson diversity indexes, and ACE and Chao1 indexes for richness. The high differences between OTUs and ACE or Chao1 suggested that further sequencing of the samples may reveal greater OTU richness. We found that although we obtained more reads and coverage from sample B, the diversity and richness in sample B were slightly higher than those of sample A.

Table 1 Summary of numbers of reads and OTUs, in addition to coverage and diversity estimators, of the 16S rRNA gene sequences of the bacterial community of marine sponge Theonella swinhoei (samples A and B)

After removing low-quality reads, we obtained 23,700 clean reads, 10,573 from sample A and 13,127 from sample B, with an average length of 485.24 and 488.70 bp, respectively. The 23,700 reads of bacterial communities of the sponge T. swinhoei could be classified into 12 phyla (Fig. 1a). The bacterial community structure of samples A and B varied little at the phylum level. Acidobacteria (71.54% of total reads), Chloroflexi (19.60%), Proteobacteria (2.62%), Planctomycetes (2.58%), Actinobacteria (1.70%), Cyanobacteria (0.56%), and Firmicutes (0.25%) were the seven major phyla. Thus, on the phylum level, more than 70% of the bacterial reads from T. swinhoei were affiliated with phylum Acidobacteria. This result is in line with the median value of the Shannon-Weaver diversity index. In addition to the phyla mentioned above, eight less-dominant bacteria, belonging to the phyla Bacteroidetes, Verrucomicrobia, Deferribacteres, and candidate phylum TM6 and TM7, were also identified from the sponge samples.

Fig. 1
figure 1

Taxonomic diversity of the bacterial community associated with marine sponge Theonella swinhoei at the phylum level (samples A and B)

The clean reads were assigned to 818 unique OTUs (based on a 97% sequence identity), among which sample A contained 529 OTUs, sample B contained 524 OTUs. Only 235 OTUs (28.73% of total OTUs) were shared between samples A and B, but these OTUs include 95.97% total reads. To know the difference between two samples, the Bray-Curtis, Morisita-Horn, and Jaccard dissimilarity were calculated to describing the dissimilarity between two communities and their values were 0.114, 0.004, and 0.712714, respectively. The inconsistence between Jaccard index and two other indices is Jaccard index does not consider the abundance of species. Therefore, we can conclude that the difference in the microbial communities between two samples is small, but rare phylotypes between them are different.

Four-hundred and fifty-five OTUs (55.50% of the total OTUs) or 5529 reads (22.85% of the total reads) could be classified into 50 separate families. The inability to resolve the taxonomic ranking of these bacteria below the family level suggested that numerous bacteria associated with T. swinhoei are still unclassified in the literature or unknown. Caldilineaceae (19.19%) was the most abundant family detected from T. swinhoei, followed by Planctomycetaceae (2.49%), Coxiellaceae (0.26%), Anaerolineaceae (0.18%), Sinobacteraceae (0.18%), Lachnospiraceae (0.12%), and Rhodospirillaceae (0.12%) (Fig. 1b; Supplementary File 1).

The top 20 most abundant OTUs, which were responsible for 81% of the total sequences of the two sponge samples in this study, were compared with the GenBank nt database using blastn, and the results are listed in Table 2. Among the 20 most abundant OTUs, 10 were affiliated to Acidobacteria, 8 to Chloroflexi, and 1 to Actinobacteria and Planctomycetes.

Table 2 BLAST analysis of the 20 most abundant OTUs in Theonella swinhoei samples A and B

Antimicrobial activity analysis

A total of 700 isolates (including 606 bacteria and 94 actinomycetes) were cultured from homogenates of the marine sponge T. swinhoei, collected at Nanwan Bay in southern Taiwan, using seven types of culture media. They were subsequently evaluated in terms of their antimicrobial activity using the agar block method in primary screening. Among the isolates, 15% (105), including 65 bacteria and 40 actinomycetes, were found to possess antimicrobial activity against at least one of the three test strains, V. parahaemolyticus, S. aureus, and C. albicans (Table 3). Isolates belonging to actinomycetes showed a greater percentage of antimicrobial activity (42.55%) than bacteria (10.73%).

Table 3 Number of strains isolated, number (percentage) of isolates exhibiting antimicrobial activity, and number of antimicrobial isolates with seawater requirements from marine sponge Theonella swinhoei using the agar block method in primary screening

Fifty-one isolates (including 19 bacteria and 32 actinomycetes) for which an inhibition zone > 1 mm in diameter was observed against at least one test strain were selected for secondary antimicrobial screening against five pathogens, S. aureus, E. coli, V. parahaemolyticus, P. aeruginosa, and C. albicans, and the results are presented in Fig. 2a, b. Among the 51 isolates, 84.31% inhibited a Gram-positive bacterium (S. aureus), 54.90% inhibited Gram-negative bacteria (E. coli, V. parahaemolyticus, and/or P. aeruginosa), 39.22% inhibited a fungus (C. albicans), and 49.02% inhibited both Gram-positive and Gram-negative bacteria. Five isolates (9.80%), including four bacteria, PYAN1-2, PYAN10-2, PYAN10-21, and PYAN1-3, and one actinomycete, SAU1-231, inhibited more than three test strains. The percentage of susceptibility of S. aureus, E. coli, V. parahaemolyticus, P. aeruginosa, and C. albicans to the 51 sponge isolates was 84.31, 23.53, 47.06, 7.84, and 39.22% (Fig. 2b), respectively, and the percentage of susceptibility to the 51 sponge isolates with inhibition zone diameters > 5 mm was 56.86, 11.76, 27.45, 3.92, and 25.49%, respectively.

figure 2

Cytotoxic activity analysis

Twenty-seven isolates (including 13 bacteria and 14 actinomycetes) for which an inhibition zone > 10 mm in diameter was observed against at least one test strain, or that were able to inhibit at least four test strains in secondary screening, were selected for evaluation of cytotoxic activity. Among these isolates, 29.6% (eight isolates, AIAC10-11, AIAC1-21, GIC10-1, GIC1-13, M1C1-37, PYAC20-24, PYAU1-1, PYAN10-21), including one bacterium and seven actinomycetes, were found to exert cytotoxic activity against at least five of the seven human cancer cell lines (Table 4). We consider that these eight isolates have potential for further study, in particular actinomycetes AIAC10-11 and GIC1-13, and bacterium PYAN10-21, which had an IC50 < 1 μg/ml for at least two cancer cell lines.

Table 4 IC50 values of crude extracts of bacterial strains isolated from marine sponge Theonella swinhoei against cancer cell lines MCF-7, MDA-MB-231, and T47-D from breast tumors, MOLT-4 and K-562 from leukemia, and DLD-1 and HCT-116 from colorectal cancer using the MTT method

Identification of bacteria producing antibiotics

16S rRNA gene sequences of the 51 selected isolates that exhibited antimicrobial activities were sequenced and subsequently aligned to construct phylogenetic trees (Fig. 3a, b). It should be noted that these 51 isolates were, therefore, neither a representative sample of culturable bacteria nor a representative sample of culturable bacteria with antimicrobial activities. Among the 51 isolates, 9 belonged to the phylum Firmicutes, 32 to the phylum Actinobacteria, and 10 to the phylum Proteobacteria. In addition, we used the Ribosomal Data Project (RDP) classifier tool (Cole et al. 2005) to classify bacteria based on their 16S rRNA sequences. The 51 isolates were classified into six genera: Streptomyces (30 isolates), Pseudoalteromonas (10), Bacillus (8), Brevibacterium (1), Micromonospora (1), and Paenibacillus (1). Streptomyces, Pseudoalteromonas, and Bacillus were the three major genera, accounting for 94% of the antibiotic-producing isolates.

Fig. 3
figure 3

Phylogenetic trees of selected bacterial strains isolated from sponge Theonella swinhoei. The trees were constructed by comparison of an approximate 650-bp region of the 16S rRNA gene sequence using neighbor-joining analysis of a distance matrix with Kimura’s two-parameter model. Bootstrap values (expressed as percentages of 100 replications) greater than 60% are shown at branch points. The 16S rRNA sequences of Methanobacterium paludis and Thermococcus radiophilus were used as outgroups. The scale bar represents 0.05 substitutions per nucleotide position

Overlap of antimicrobial isolates and 454 sequences

In the present study, we found that all of the antimicrobial isolates obtained from the sponge samples belonged to three bacterial phyla, namely Actinobacteria (62.8%), Proteobacteria (19.6%), and Firmicutes (17.4%), among which 70 (8.56% of the total OTUs), 187 (8.56%), and 29 OTUs (3.55%), respectively, were found in the 454 pyrosequencing data set (Supplementary File 1). At the genus level, only three genera, Pseudoalteromonas, Streptomyces, and Bacillus, were found to each have one OTU, and no Brevibacterium, Micromonospora, or Paenibacillus OTUs were found in the pyrosequencing data set (Supplementary File 1). BLAST comparison of the 16S rRNA gene sequences of the antimicrobial isolated bacteria in the pyrosequencing data indicated that only OTU692 and OTU618, found within the T. swinhoei metagenome, were closely related (98 and 96% identities, respectively) to 16S rRNA genes from the eight Pseudoalteromonas and one Streptomyces isolates, respectively. This lack of overlap between culture-based isolates and metagenomic sequences was consistent with studies of the microbiomes of sponges Rhopaloeides odorabile (Webster and Hill 2001) and Haliclona simulans (Kennedy et al. 2009).

Discussion

Many researchers believe that symbiotic microbes of marine invertebrates are an untapped source of bioactive compounds (Burgess et al. 1999; Nithyanand et al. 2011). However, just a small fraction of these compounds has been identified and studied. In the present study, bacteria associated with the marine sponge T. swinhoei obtained from southern Taiwan were systematically studied and screened for production of antibiotics. Of a total of 700 cultured isolates, including 606 bacteria and 94 actinomycetes, 105 strains (15%) showed antibiotic activity towards at least one indicator microbe used in primary screening. A significantly higher percentage of antibiotic-producing isolates was detected in the actinomycetal isolates (42.55%) than in the bacterial isolates (10.73%). The fraction (12%) of antimicrobial-producing bacteria in our results was consistent with several previous studies; for example, Hentschel et al. (2001) studied the antimicrobial activity of the sponge Aplysina, and found that 11.35% of the bacterial isolates possessed antimicrobial activity; and Zhang et al. (2009) showed that 18% of the bacterial isolates obtained from four sponge species were antibiotic producers. In general, the percentage of microbial isolates with antimicrobial activity in sponges varies between 10 and 50% (Flemer et al. 2012; Hentschel et al. 2001; Kennedy et al. 2009; Santos et al. 2010). It should be noted that the percentage can be higher if the number of test strains is increased and more diverse conditions are used for the assay (Kennedy et al. 2009).

Phylogenetic analysis (Fig. 3a, b) showed that Streptomyces and Pseudoalteromonas were the two major genera that contained 78.43% of the antibiotic producers in this study. This result was consistent with the study by Kennedy et al. (2009), which showed that most of the isolates from Demospongiae sponge Haliclona simulans that exhibited antimicrobial activity belonged to the genera Streptomyces (36.5%) and Pseudoalteromonas (23.1%). In this study, most of the Pseudoalteromonas isolates obtained from T. swinhoei inhibited at least one Gram-positive and one Gram-negative indicator bacteria, whereas only three of the nine tested Pseudoalteromonas isolates showed antifungal activities against C. albicans. Pseudoalteromonas contains several species that are known to produce biologically active compounds with antimicrobial and antifouling effects (Wilson et al. 2010). They are often found in symbiosis with marine eukaryotic hosts, such as sponges, coral, and tunicates (Hentschel et al. 2001). In previous culture-based studies, Hentschel et al. (2001) and Kennedy et al. (2009) isolated several Pseudoalteromonas species from Irish and Mediterranean sponges, respectively, and found that they exhibited no antimicrobial activity against either E. coli or S. aureus, and only antimicrobial activity against a Gram-negative marine bacterium, respectively, which was significantly less activity than that demonstrated by our results. These results might indicate that different isolates from different sponges might adapt to dissimilar host environments and produce varying antibiotics, suggesting the need to isolate bacteria from different marine sponges.

Overall, the bacterial assemblage of sponge T. swinhoei was composed mainly of Acidobacteria, Chloroflexi, Planctomycetes, Proteobacteria, Actinobacteria, Cyanobacteria, Firmicutes, Bacteroidetes, Verrucomicrobia, and Deferribacteres, as well as candidate phylum TM6 and TM7 (Fig. 1). This high level of microbial richness at the phylum level was consistent with several previous culture-independent surveys of microbial communities of Demospongiae sponges, which are known to host a diverse range of bacteria. The bacterial communities associated with T. swinhoei were recently described for samples obtained from Palau, Israel, and Japan (Hentschel et al. 2002). Comparison of the bacterial flora of these samples with those identified in this study at the phylum level showed that some of the major phyla, such as Acidobacteria, Chloroflexi, Proteobacteria, Actinobacteria, and Cyanobacteria, were maintained, while Planctomycetes and Firmicutes were not, suggesting that the sponge host might have the ability to establish new bacterial flora to adapt to different environments.

OTU1, the most abundant 16S rRNA OTU of the sponge T. swinhoei, which represented almost half (approximately 47%) of the total sequence (Table 2), showed the best hit with a sequence (FJ543132, 98% similarity) obtained from sponge Acanthostrongylophora sp., belonging to the phylum Acidobacteria and class Holophagae. In addition, 11 of the 20 most abundant OTUs also belonged to Acidobacteria, representing 71.54% of the total sequence associated with T. swinhoei. This group of bacteria was also the most abundant bacterial phylum in Palauan and Israeli T. swinhoei (Hentschel et al. 2002), both representing about 40% of the total sequence; this was very different from the 71.54% observed in the present study, suggesting that geographic location or environmental factors might play roles in sponge-associated microbial communities, rather than species-specific factors. O’Connor-Sanchez et al. (2014) reported that a single Acidobacteria OTU dominated the bacterial community of Demospongiae sponge Hyrtios sp. collected in the Mexican Caribbean, with a relative abundance of 80%, which was consistent with our results.

Using the 16S rDNA-based molecular method, bacteria in the phylum Acidobacteria have been found to be highly diverse, and are consistently detected in different habitats, including soil, sediment, freshwater, marine, and wastewater environments (Barns et al. 2007; Ward et al. 2009), which suggests that they play an important ecological role and have varied functions. They are particularly abundant in soil and sediment habitats, comprising 10 to 50% of the total bacterial abundance (Barns et al. 2007). Previous studies of microbes associated with marine invertebrates have also revealed that Acidobacteria can inhibit several sponge (Hardoim and Costa 2014; O’Connor-Sanchez et al. 2014; Schmitt et al. 2012; Simister et al. 2013) and coral species (Closek et al. 2014; Lin et al. 2016; Ng et al. 2015). Currently, our knowledge of the ecological functions served by this group of bacteria is still scarce (O’Connor-Sanchez et al. 2014). The high abundance of Acidobacteria in the present study might be representative of adaption to the environmental conditions at Nanwan Bay in southern Taiwan. In addition, the Acidobacteria symbiont might produce secondary metabolites to inhibit the growth of other microbes (O’Connor-Sanchez et al. 2014). However, we did not isolate any Acidobacteria with antimicrobial activity from T. swinhoei, probably because Acidobacteria are very difficult to cultivate and slow-growing, suggesting that different culturing strategies may needed to culture Acidobacteria.

We observed that Planctomycetaceae, a member of the Planctomycetes family, was the second most abundant family identified in T. swinhoei, with a relative read abundance of 10.67% (in identified families) (Supplementary File 1). Planctomycetes have been found in a variety of marine, freshwater, and terrestrial environments (Pimentel-Elardo et al. 2003), and have also been found to be associated with marine eukaryotes (Kellogg et al. 2016; Mohamed et al. 2010). Recently, several members of this family were found to mediate anaerobic ammonia oxidation (anammox), which is an important biological nitrogen removal process that has been postulated to generate 50% of the nitrogen in the atmosphere (Jetten 2008). Although the diversity of Planctomycetaceae (including 178 OTUs or five genera) was high in T. swinhoei, only 13 OTUs (22 reads), which were all identified as OM190 (Silva taxonomy), were found to be distantly related to anammox bacteria (Ye et al. 2016). The OM190 lineage has been found to be associated with kelp Laminaria hyperborea (Bengtsson and Ovreas 2010) and marine sponge Mycale laxissima (Mohamed et al. 2010). The presence of anammox OTUs might be important for the nitrogen cycle in T. swinhoei, and clearly warrants further study.

Phylum Chloroflexi (green non-sulfur bacteria) was the second most abundant phylum (19.6% of total reads) in this study. A total 142 Chloroflexi OTUs, which accounts for 17.36% of all OTUs, was identified, suggesting the diversity of Chloroflexi community in T. swinhoei was also high. This group of bacteria was also found to be a dominant bacterial phylum in several marine sponges, such as Rhopaloeides odorabile (39–41% of total reads) (Simister et al. 2012) and Ecionemia alata (19–28%) (Cardenas et al. 2014). Recently, Schmitt et al. (2011) compared the Chloroflexi community structure between high microbial abundance (HMA) and low microbial abundance (LMA) sponges and found that Chloroflexi bacteria are more diverse and abundant in HMA than in LMA sponges, which is consistent with our results for T. swinhoei (also HMA). At present, our knowledge regarding Chloroflexi in sponges is still largely based on the application of cultivation-independent methods (Schmitt et al. 2011). Further clarifying the role of sponge-associated Chloroflexi and determining its ecological significance remains a big challenge for future research.

The eight most dominant Chloroflexi OTUs identified were OTU3, OTU5, OTU6, OTU4, OTU9, OTU7, OTU21, and OTU17, comprising 12.07% of total reads (Table 2). Although two Chloroflexi classes, Caldilineae and Anaerolineae, were identified in this study, all the eight Chloroflexi OTUs belong to genius Caldilinea (class Caldilineae). Caldilineae is a non-photosynthetic and ubiquitous class with 16S rRNA gene sequences reported in soils (Breuker et al. 2011), marine sediments (Schmitt et al. 2011), activated sludge (Kindaichi et al. 2012), and marine animals, such as sponge (Schmitt et al. 2011). BLAST results showed these eight OTUs were all closely related to bacterial sequences in marine sponges. Interesting, similar to Planctomycetes OTUs, members of this class were also found to exist in anammox reactor (Kindaichi et al. 2012), suggested that anammox reaction could be an important pathway for ammonium removal in host. To the author’s knowledge, there has been no published study of antibiotics producing Chloroflexi. However, several PKS genes were identified in a Chloroflexi genome (Kiss et al. 2011), suggesting the possibility of using Chloroflexi as antibiotics producer.

In the present study, eight isolates of Pseudoalteromonas and one isolate of Streptomycetes were found to be shared between the cultured isolates and 454 sequences. This discrepancy between the major groups of culture-based isolates and deep pyrosequencing sequences was in agreement with the results of several previous studies (Kennedy et al. 2009; Montalvo et al. 2014). The inconsistency is possibly due to the percentage of bacteria isolated was a very small part of the bacterial community of T. swinhoei, and therefore cannot detect in pyrosequencing sequences (Kennedy et al. 2009). Our study highlights that the culture-based method can address some of the limitations of NGS sequencing, and is worthwhile not only in order to isolate bacteria for biotechnology usage, but also to increase the detection of bacterial community diversity in sponges.

Eight isolates (Table 4) that exhibited strong or wide-spectrum antibiotic activities and cytotoxic activity against cancer cells provided further evidence that culturable sponge microflora is an important source of biologically active compounds. For isolate GIC10-1, the 16S rRNA gene BLAST results showed a 100% identity with Streptomycetes sp. OPMA00072 (Genbank accession no. AB896819). The crude extract of this isolate possessed antimicrobial activity against two indicator microbes and cytotoxic activity against three cancer cells, which suggested that it is a good candidate for further natural product isolation and characterization. In fact, a new compound, bafilomycin M, was recently isolated from the culture broth of GIC10-1 (Chen et al. 2016). Thus, our study demonstrated that microbes associated with the marine sponge T. swinhoei can be a source of natural antimicrobial agents, and these antibiotic-producing bacteria may play an ecologically important role in the relationship between microbial symbiont and sponge host.

References

  • Abdelmohsen UR, Bayer K, Hentschel U (2014) Diversity, abundance and natural products of marine sponge-associated actinomycetes. Nat Prod Rep 31:381–399

    Article  CAS  PubMed  Google Scholar 

  • Barns SM, Cain EC, Sommerville L, Kuske CR (2007) Acidobacteria phylum sequences in uranium-contaminated subsurface sediments greatly expand the known diversity within the phylum. Appl Environ Microbiol 73:3113–3116

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Bayer K, Schmitt S, Hentschel U (2008) Physiology, phylogeny and in situ evidence for bacterial and archaeal nitrifiers in the marine sponge Aplysina aerophoba. Environ Microbiol 10:2942–2955

    Article  CAS  PubMed  Google Scholar 

  • Bengtsson MM, Ovreas L (2010) Planctomycetes dominate biofilms on surfaces of the kelp Laminaria hyperborea. BMC Microbiol 10:261

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Bewley CA, Holland ND, Faulkner DJ (1996) Two classes of metabolites from Theonella swinhoei are localized in distinct populations of bacterial symbionts. Experientia 52:716–722

    Article  CAS  PubMed  Google Scholar 

  • Bredholdt H, Galatenko OA, Engelhardt K, Fjaervik E, Terekhova LP, Zotchev SB (2007) Rare actinomycete bacteria from the shallow water sediments of the Trondheim fjord, Norway: isolation, diversity and biological activity. Environ Microbiol 9:2756–2764

    Article  CAS  PubMed  Google Scholar 

  • Breuker A, Koweker G, Blazejak A, Schippers A (2011) The deep biosphere in terrestrial sediments in the Chesapeake bay area, Virginia, USA. Front Microbiol 2:156

    Article  PubMed  PubMed Central  Google Scholar 

  • Burgess JG, Jordan EM, Bregu M, Mearns-Spragg A, Boyd KG (1999) Microbial antagonism: a neglected avenue of natural products research. J Biotechnol 70:27–32

    Article  CAS  PubMed  Google Scholar 

  • Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, Madden TL (2009) BLAST+: architecture and applications. BMC Bioinformatics 10:421

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Cardenas CA, Bell JJ, Davy SK, Hoggard M, Taylor MW (2014) Influence of environmental variation on symbiotic bacterial communities of two temperate sponges. FEMS Microbiol Ecol 88:516–527

    Article  CAS  PubMed  Google Scholar 

  • Chen Y-H, Kuo J, Sung P-J et al (2012) Isolation of marine bacteria with antimicrobial activities from cultured and field-collected soft corals. World J Microbiol Biotechnol 28:3269–3279

    Article  CAS  PubMed  Google Scholar 

  • Chen Y-H, Lu M-C, Chung H-M et al (2016) Bafilomycin M, a new cytotoxic bafilomycin produced by a Streptomyces sp. isolated from a marine sponge Theonella sp. Tetrahedron Lett 57:4863–4865

    Article  CAS  Google Scholar 

  • Closek CJ, Sunagawa S, DeSalvo MK et al (2014) Coral transcriptome and bacterial community profiles reveal distinct Yellow Band Disease states in Orbicella faveolata. ISME J 8:2411–2422

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Cole JR, Chai B, Farris RJ et al (2005) The Ribosomal Database Project (RDP-II): sequences and tools for high-throughput rRNA analysis. Nucleic Acids Res 33:D294–D296

    Article  CAS  PubMed  Google Scholar 

  • Flatt PM, Gautschi JT, Thacker RW, Musafija-Girt M, Crews P, Gerwick WH (2005) Identification of the cellular site of polychlorinated peptide biosynthesis in the marine sponge Dysidea (Lamellodysidea) herbacea and symbiotic cyanobacterium Oscillatoria spongeliae by CARD-FISH analysis. Mar Biol 147:761–774

    Article  CAS  Google Scholar 

  • Flemer B, Kennedy J, Margassery LM, Morrissey JP, O’Gara F, Dobson AD (2012) Diversity and antimicrobial activities of microbes from two Irish marine sponges, Suberites carnosus and Leucosolenia sp. J Appl Microbiol 112:289–301

    Article  CAS  PubMed  Google Scholar 

  • Gerwick WH, Moore BS (2012) Lessons from the past and charting the future of marine natural products drug discovery and chemical biology. Biol Chem 19:85–98

    Article  CAS  Google Scholar 

  • Graca AP, Bondoso J, Gaspar H et al (2013) Antimicrobial activity of heterotrophic bacterial communities from the marine sponge Erylus discophorus (Astrophorida, Geodiidae). PLoS One 8:e78992

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Hardoim CC, Costa R (2014) Temporal dynamics of prokaryotic communities in the marine sponge Sarcotragus spinosulus. Mol Ecol 23:3097–3112

    Article  CAS  PubMed  Google Scholar 

  • Hentschel U, Schmid M, Wagner M, Fieseler L, Gernert C, Hacker J (2001) Isolation and phylogenetic analysis of bacteria with antimicrobial activities from the Mediterranean sponges Aplysina aerophoba and Aplysina cavernicola. FEMS Microbiol Ecol 35:305–312

    Article  CAS  PubMed  Google Scholar 

  • Hentschel U, Hopke J, Horn M, Friedrich AB, Wagner M, Hacker J, Moore BS (2002) Molecular evidence for a uniform microbial community in sponges from different oceans. Appl Environ Microbiol 68:4431–4440

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Hill M, Hill A, Lopez N, Harriott O (2006) Sponge-specific bacterial symbionts in the Caribbean sponge, Chondrilla nucula (Demospongiae, Chondrosida). Mar Biol 148:1221–1230

    Article  Google Scholar 

  • Hoffmann F, Radax R, Woebken D et al (2009) Complex nitrogen cycling in the sponge Geodia barretti. Environ Microbiol 11:2228–2243

    Article  CAS  PubMed  Google Scholar 

  • Ivanitskaia LP, Singal EM, Bibikova MV, Vostrov SN (1978) Directed isolation of Micromonospora generic cultures on a selective medium with gentamycin. Antibiotiki 23:690–692

    CAS  PubMed  Google Scholar 

  • Jensen PR, Gontang E, Mafnas C, Mincer TJ, Fenical W (2005) Culturable marine actinomycete diversity from tropical Pacific Ocean sediments. Environ Microbiol 7:1039–1048

    Article  PubMed  Google Scholar 

  • Jetten MS (2008) The microbial nitrogen cycle. Environ Microbiol 10:2903–2909

    Article  CAS  PubMed  Google Scholar 

  • Jin L, Liu F, Sun W, Zhang F, Karuppiah V, Li Z (2014) Pezizomycotina dominates the fungal communities of South China Sea sponges Theonella swinhoei and Xestospongia testudinaria. FEMS Microbiol Ecol 90:935–945

    Article  CAS  PubMed  Google Scholar 

  • Kellogg CA, Ross SW, Brooke SD (2016) Bacterial community diversity of the deep-sea octocoral Paramuricea placomus. PeerJ 4:e2529

    Article  PubMed  PubMed Central  Google Scholar 

  • Kennedy J, Baker P, Piper C et al (2009) Isolation and analysis of bacteria with antimicrobial activities from the marine sponge Haliclona simulans collected from Irish waters. Mar Biotechnol 11:384–396

    Article  CAS  Google Scholar 

  • Keren R, Lavy A, Mayzel B, Ilan M (2015) Culturable associated-bacteria of the sponge Theonella swinhoei show tolerance to high arsenic concentrations. Front Microbiol 6:154

    Article  PubMed  PubMed Central  Google Scholar 

  • Kindaichi T, Yuri S, Ozaki N, Ohashi A (2012) Ecophysiological role and function of uncultured Chloroflexi in an anammox reactor. Water Sci Technol 66:2556–2561

    Article  CAS  PubMed  Google Scholar 

  • Kiss H, Nett M, Domin N et al (2011) Complete genome sequence of the filamentous gliding predatory bacterium Herpetosiphon aurantiacus type strain (114-95(T)). Stand Genomic Sci 5:356–370

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Lavy A, Keren R, Haber M, Schwartz I, Ilan M (2014) Implementing sponge physiological and genomic information to enhance the diversity of its culturable associated bacteria. FEMS Microbiol Ecol 87:486–502

    Article  CAS  PubMed  Google Scholar 

  • Leal MC, Puga J, Serodio J, Gomes NC, Calado R (2012) Trends in the discovery of new marine natural products from invertebrates over the last two decades—where and what are we bioprospecting? PLoS One 7:e30580

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Lin CH, Chuang CH, Twan WH et al (2016) Seasonal changes in bacterial communities associated with healthy and diseased Porites coral in southern Taiwan. Can J Microbiol 62:1021–1033

    Article  CAS  PubMed  Google Scholar 

  • Lu M-C, Du Y-C, Chuu J-J et al (2009) Active extracts of wild fruiting bodies of Antrodia camphorata (EEAC) induce leukemia HL 60 cells apoptosis partially through histone hypoacetylation and synergistically promote anticancer effect of trichostatin A. Arch Toxicol 83:121–129

    Article  CAS  PubMed  Google Scholar 

  • Mehbub MF, Lei J, Franco C, Zhang W (2014) Marine sponge derived natural products between 2001 and 2010: trends and opportunities for discovery of bioactives. Mar Drugs 12:4539–4577

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Mohamed NM, Saito K, Tal Y, Hill RT (2010) Diversity of aerobic and anaerobic ammonia-oxidizing bacteria in marine sponges. ISME J 4:38–48

  • Montalvo NF, Davis J, Vicente J, Pittiglio R, Ravel J, Hill RT (2014) Integration of culture-based and molecular analysis of a complex sponge-associated bacterial community. PLoS One 9:e90517

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Ng JC, Chan Y, Tun HM, Leung FC, Shin PK, Chiu JM (2015) Pyrosequencing of the bacteria associated with Platygyra carnosus corals with skeletal growth anomalies reveals differences in bacterial community composition in apparently healthy and diseased tissues. Front Microbiol 6:1142

    PubMed  PubMed Central  Google Scholar 

  • Nithyanand P, Manju S, Karutha Pandian S (2011) Phylogenetic characterization of culturable actinomycetes associated with the mucus of the coral Acropora digitifera from Gulf of Mannar. FEMS Microbiol Lett 314:112–118

    Article  CAS  PubMed  Google Scholar 

  • O’Connor-Sanchez A, Rivera-Dominguez AJ, Santos-Briones Cde L, Lopez-Aguiar LK, Pena-Ramirez YJ, Prieto-Davo A (2014) Acidobacteria appear to dominate the microbiome of two sympatric Caribbean sponges and one Zoanthid. Biol Res 47:67

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Osinga R, Armstrong E, Grant Burgess J, Hoffmann F, Reitner J, Schumann-Kindel G (2001) Sponge–microbe associations and their importance for sponge bioprocess engineering. Hydrobiologia 461:55–62

    Article  Google Scholar 

  • Piel J, Hui D, Wen G, Butzke D, Platzer M, Fusetani N, Matsunaga S (2004) Antitumor polyketide biosynthesis by an uncultivated bacterial symbiont of the marine sponge Theonella swinhoei. Proc Natl Acad Sci U S A 101:16222–16227

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Pimentel-Elardo S, Wehrl M, Friedrich AB, Jensen PR, Hentschel U (2003) Isolation of planctomycetes from Aplysina sponges. Aquat Microb Ecol 33:239–245

    Article  Google Scholar 

  • Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, Glockner FO (2007) SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res 35:7188–7196

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Reveillaud J, Maignien L, Eren AM, Huber JA, Apprill A, Sogin ML, Vanreusel A (2014) Host-specificity among abundant and rare taxa in the sponge microbiome. ISME J 8:1198–1209

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Rodriguez-R LM, Konstantinidis KT (2014) Nonpareil: a redundancy-based approach to assess the level of coverage in metagenomic datasets. Bioinformatics 30:629–635

    Article  CAS  PubMed  Google Scholar 

  • Santos OC, Pontes PV, Santos JF, Muricy G, Giambiagi-deMarval M, Laport MS (2010) Isolation, characterization and phylogeny of sponge-associated bacteria with antimicrobial activities from Brazil. Res Microbiol 161:604–612

    Article  CAS  PubMed  Google Scholar 

  • Schloss PD, Westcott SL, Ryabin T et al (2009) Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl Environ Microbiol 75

  • Schloss PD, Gevers D, Westcott SL (2011) Reducing the effects of PCR amplification and sequencing artifacts on 16S rRNA-based studies. PLoS One 6:e27310

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Schmidt EW, Obraztsova AY, Davidson SK, Faulkner DJ, Haygood MG (2000) Identification of the antifungal peptide-containing symbiont of the marine sponge Theonella swinhoei as a novel δ-proteobacterium, “Candidatus Entotheonella palauensis”. Mar Biol 136:969–977

    Article  CAS  Google Scholar 

  • Schmitt S, Deines P, Behnam F, Wagner M, Taylor MW (2011) Chloroflexi bacteria are more diverse, abundant, and similar in high than in low microbial abundance sponges. FEMS Microbiol Ecol 78:497–510

    Article  CAS  PubMed  Google Scholar 

  • Schmitt S, Hentschel U, Taylor MW (2012) Deep sequencing reveals diversity and community structure of complex microbiota in five Mediterranean sponges. Hydrobiologia 687:341–351

    Article  CAS  Google Scholar 

  • Simister R, Taylor MW, Tsai P, Fan L, Bruxner TJ, Crowe ML, Webster N (2012) Thermal stress responses in the bacterial biosphere of the Great Barrier Reef sponge, Rhopaloeides odorabile. Environ Microbiol 14:3232–3246

    Article  CAS  PubMed  Google Scholar 

  • Simister R, Taylor MW, Rogers KM, Schupp PJ, Deines P (2013) Temporal molecular and isotopic analysis of active bacterial communities in two New Zealand sponges. FEMS Microbiol Ecol 85:195–205

    Article  PubMed  Google Scholar 

  • Stern NJ, Svetoch EA, Eruslanov BV et al (2006) Isolation of a Lactobacillus salivarius strain and purification of its bacteriocin, which is inhibitory to Campylobacter jejuni in the chicken gastrointestinal system. Antimicrob Agents Chemother 50:3111–3116

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S (2011) MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28:2731–2739

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Taylor MW, Radax R, Steger D, Wagner M (2007) Sponge-associated microorganisms: evolution, ecology, and biotechnological potential. Microbiol Mol Biol Rev 71:295–347

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Terekhova LP, Galatenko OA, Alferova IV, Preobragenskaya TP (1991) Comparative estimation of some bacterial growth inhibitors as selective agents for isolation of soil actinomycetes. Antibiot Chemother 36:5–8

    CAS  Google Scholar 

  • Thomas TRA, Kavlekar DP, LokaBharathi PA (2010) Marine drugs from sponge-microbe association—a review. Mar Drugs 8:1417–1468

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Thoms C, Horn M, Wagner M, Hentschel U, Proksch P (2003) Monitoring microbial diversity and natural product profiles of the sponge Aplysina cavernicola following transplantation. Mar Biol 142:685–692

    Article  CAS  Google Scholar 

  • Unson MD, Faulkner DJ (1993) Cyanobacterial symbiont biosynthesis of chlorinated metabolites from Dysidea herbacea (Porifera). Experientia 49:349–353

    Article  CAS  Google Scholar 

  • Unson MD, Holland ND, Faulkner DJ (1994) A brominated secondary metabolite synthesized by the cyanobacterial symbiont of a marine sponge and accumulation of the crystalline metabolite in the sponge tissue. Mar Biol 119:1–11

    Article  CAS  Google Scholar 

  • Ward NL, Challacombe JF, Janssen PH et al (2009) Three genomes from the phylum Acidobacteria provide insight into the lifestyles of these microorganisms in soils. Appl Environ Microbiol 75:2046–2056

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Webster NS, Hill RT (2001) The culturable microbial community of the Great Barrier Reef sponge Rhopaloeides odorabile is dominated by an alpha-proteobacterium. Mar Biol 138:843–851

  • Webster NS, Taylor MW (2012) Marine sponges and their microbial symbionts: love and other relationships. Environ Microbiol 14:335–346

    Article  CAS  PubMed  Google Scholar 

  • Wilkinson CR, Fay P (1979) Nitrogen fixation in coral reef sponges with symbiotic cyanobacteria. Nature 279:527–529

    Article  CAS  Google Scholar 

  • Wilkinson CR, Nowak M, Austin B, Colwell RR (1981) Specificity of bacterial symbionts in Mediterranean and Great Barrier Reef sponges. Microb Ecol 7:13–21

    Article  CAS  PubMed  Google Scholar 

  • Wilson GS, Raftos DA, Corrigan SL, Nair SV (2010) Diversity and antimicrobial activities of surface-attached marine bacteria from Sydney Harbour, Australia. Microbiol Res 165:300–311

    Article  CAS  PubMed  Google Scholar 

  • Ye Q, Wu Y, Zhu Z, Wang X, Li Z, Zhang J (2016) Bacterial diversity in the surface sediments of the hypoxic zone near the Changjiang Estuary and in the East China Sea. MicrobiologyOpen 5:323–339

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Zhang W, Li Z, Miao X, Zhang F (2009) The screening of antimicrobial bacteria with diverse novel nonribosomal peptide synthetase (NRPS) genes from South China Sea sponges. Mar Biotechnol 11:346–355

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by intramural funding from the National Museum of Marine Biology and Aquarium.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jimmy Kuo.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Ethical approval

All procedures performed in studies involving animals were in accordance with the ethical standards of the institution or practice at which the studies were conducted.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Supplementary File 1

Taxonomy assignments of the bacterial community associated with marine sponge Theonella swinhoei at different levels. (XLSX 53.9 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kuo, J., Yang, YT., Lu, MC. et al. Antimicrobial activity and diversity of bacteria associated with Taiwanese marine sponge Theonella swinhoei. Ann Microbiol 69, 253–265 (2019). https://doi.org/10.1007/s13213-018-1414-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13213-018-1414-3

Keywords