Hindawi Publishing Corporation The Scientific World Journal Volume 2013, Article ID 165706, 13 pages http://dx.doi.org/10.1155/2013/165706

Research Article Bacterial Communities in Polluted Seabed Sediments: A Molecular Biology Assay in Leghorn Harbor Carolina Chiellini,1 Renato Iannelli,2 Franco Verni,1 and Giulio Petroni1 1 2

Department of Biology, Unit of Protistology-Zoology, University of Pisa, via Luca Ghini 13, 56126 Pisa, Italy Department of Engineering for Energy, Systems, Territory and Constructions, University of Pisa, via Carlo Francesco Gabba 22, 56122 Pisa, Italy

Correspondence should be addressed to Giulio Petroni; [email protected] Received 26 July 2013; Accepted 22 August 2013 Academic Editors: Y. Nishita, S. A. Stephenson, and Y. Xi Copyright © 2013 Carolina Chiellini et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Seabed sediments of commercial ports are often characterized by high pollution levels. Differences in number and distribution of bacteria in such areas can be related to distribution of pollutants in the port and to sediment conditions. In this study, the bacterial communities of five sites from Leghorn Harbor seabed were characterized, and the main bacterial groups were identified. T-RFLP was used for all samples; two 16S rRNA libraries and in silico digestion of clones were used to identify fingerprint profiles. Library data, phylogenetic analysis, and T-RFLP coupled with in silico digestion of the obtained sequences evidenced the dominance of Proteobacteria and the high percentage of Bacteroidetes in all sites. The approach highlighted similar bacterial communities between samples coming from the five sites, suggesting a modest differentiation among bacterial communities of different harbor seabed sediments and hence the capacity of bacterial communities to adapt to different levels and types of pollution.

1. Introduction Sea stretches of commercial ports are often characterized by high levels of pollution in sediments, low oxygen concentrations in the water column, and low biodiversity of benthic communities [1] that may cause the decrease of fauna in their seabed [2]. Further reasons of impact on harbor sea stretches often come from high concentrations of organic matter due to eutrophication, time variability of sediment deposition, allochthonous inputs, and low hydrodynamism [3, 4]. These reasons demonstrate that the activities conducted in ports lead to critical conditions in the environments, and the organisms develop several ways to adapt to the alterations of water and sediments parameters [5]. A crucial characteristic of harbor sites is the removal of organic matter. Consumption capacity of organic matter is mainly due to biotic uptake and degradation performed by microorganisms. This process is generally insufficient to maintain the equilibrium of ecosystems; in most cases, the presence of detrital organic matter (which increases anoxic conditions), heavy metals, polychlorinated biphenyls (PCBs, [6]), and other toxic compounds

plays also a crucial role in negatively impacting the seabed biocenosis [7, 8]. Although the origin and concentration of pollutants in seabed harbor sediments can be different, many studies demonstrate that, in most cases, pollution is mainly due to the presence of hydrocarbons, heavy metals [9], and organic matter. Possible methods for hindering the problem of contaminated aquatic sediments have been widely discussed [10]. Polluted seabed sediments can be sanitised via in situ actions, which always require an accurate knowledge of the local biocenosis [11]. Moreover, it was demonstrated that bacterial communities have great potential to be used as sensitive indicators of contamination in aquatic sediments [12]. New and accurate methods to study the composition of microbial communities in harbor seabed sediments are necessary for these reasons. Differences in number and distribution of bacteria, in such a large area as a harbor, can be related to pollutant distribution and general sediment state. Today, Leghorn’s Harbor is one of the most important ports in the Mediterranean Sea, linked with more than 300 ports worldwide. It is a multipurpose port that can cater for all kinds of vessels and handle all kinds of goods,

2 as well as passenger traffic. It sports a marine area of 1.6 km2 and a useable land area of 2.6 km2 . The Harbor offers 11 km of quays with more than 90 berths, with up to 13 m draught. The total extension covers one km2 surface outdoor and 70000 m2 indoor areas. In the present research, the bacterial community composition of Leghorn Harbor seabed sediments was studied with a molecular fingerprint approach and with the construction of libraries of the 16S rRNA gene. The Terminal Restriction Fragment Length Polymorphism (T-RFLP) approach was used for all samples to identify the community structure. in silico digestion of clones retrieved from two clone libraries (one for a less polluted sample and one for a more polluted sample) was used for the identification of bacterial Operational Taxonomic Units (OTUs) in the T-RFLP fingerprint profiles. The aim of the study was to highlight how the different conditions of port seabed influence the composition of microbial communities and provide hints of dynamics for nutrient removal.

2. Materials and Methods 2.1. Description of the Study Area: Collection and Storage of Samples. Figure 1 shows Leghorn’s Harbor area. The five sampling sites are located with tagged marks and briefly described within the figure. In each site, three samples were collected for a total of fifteen samples. According to the site assessment performed by the port authority [13] and the chemical parameters reported in Table 1, site one, which is located in the passenger terminal, represents the less polluted spot, while site two, which is located inside the “Darsena Calafatari” ship repairing section and has never been dredged for 60 years, has the highest concentrations of heavy metals and hydrocarbons. In this work, the samples named 2.1 (site 2, sample 1) and 1.2 (site 1, sample 2) were used. They were collected on the November 3, 2009, by a scuba diver and immediately brought to the laboratory for the extraction of the total DNA. 2.2. DNA Extraction, Construction of 16S rRNA Gene Libraries, Sequencing, T-RFLP Analysis, and In Silico Digestion of Sequenced Clones. Total DNA extraction was performed on all of the 15 samples using a soil master DNA extraction kit (Epicentre Biotechnologies, WI, USA). The library was built with samples 1.2 and 2.1 using the procedure described by Amann et al. [14]: 16S rRNA genes were directly amplified from extracted DNA using universal bacterial primers, 8F (5󸀠 -AGA GTT TGA T(CT)(AC) TGG CTC AG-3󸀠 ) and reverse 1492R (5󸀠 -GG(AGCT)(AT)AC CTT GTT ACG ACT T-3󸀠 ) [15]; the amplification products were cloned in a plasmid vector (pCRs2.1-TOPOs, TOPO TA Cloning Kit, Invitrogen, UK) and inserted in chemically competent cells (OneShot TOP10, Invitrogen, UK). The inserted fragments from a representative number of clones were then amplified by control PCR with M13F and M13R universal primers. The inserted genes showing a size similar to the 16S rRNA gene size were directly sequenced with primers M13F and M13R by the Macrogen Inc. sequencing service (Republic of Korea).

The Scientific World Journal For T-RFLP analysis, the amplification of the 16S rRNA gene was performed with the same procedures of the library construction, and primer 8F was labeled with FAM fluorochrome (Applied Biosystem CA, USA). The template was digested with two different restriction endonucleases: BsuRI (GG∧ CC, 0.2 u/𝜇L, Fermentas, Canada) and RsaI (GT∧ AC, 0.2 u/𝜇L, Fermentas). Digested DNA was precipitated with cold ethanol 100% to eliminate salts at 4∘ C and 10,000 RCF. For each reaction, a mix was prepared with 1.2 𝜇L of loading buffer (GeneScan 600 LIZ, Applied Biosystem), a maximum of 5.5 𝜇L of sample (calculated after cold ethanol precipitation on the bases of its final concentration), and 13.3 𝜇L of deionized formamide (Applichem, Germany). Capillary electrophoresis was performed with Abi Prism 310 Genetic Analyzer (Applied Biosystem); T-RFLP profiles were analyzed using GeneScan analysis software (Applied Biosystem), and the data matrix was transformed for statistics as described in Iannelli et al. [13]. Nonmetric Multidimensional Scaling (NMDS) was performed on the whole T-RFLP dataset with the Bray-Curtis coefficient and the Shannon diversity index was also calculated. All statistical analyses were performed using PAST software v.2.15 (Paleontological Statistic, [16]). Further details on the applied techniques are reported in Chiellini et al. [17]. All sequences obtained from the construction of both libraries were digested in silico by searching the restriction site on Arb database, as recognized by restriction enzymes BsuRI and RsaI in order to know the size of the terminal restriction fragments representative of the OTUs. With the in silico digestion of retrieved clones, we could correlate the peaks on T-RFLP fingerprints of the 15 samples from Leghorn Harbor seabed (3 replicates in 5 different sites) to specific OTUs. The coverage percentages of the peaks corresponding to the recognized OTUs and corresponding bacterial phyla and classes were calculated for each electropherogram. 2.3. Detection of Chimeric Sequences and Phylogenetic Analysis. All the retrieved sequences were scanned with the Bellerophon server [18] to identify chimeras. We decided to analyze chimeric sequences by “cutting” them in proximity of the recombination site identified by Bellerophon software and considering the obtained fragments as independent sequences belonging to different organisms. Fragments were treated in the same way as complete sequences from other screened clones; they were included in 16S rRNA library analysis, but they were not included in the construction of phylogenetic trees, due to their short length. NCBI BLAST analysis [19] was used to determine a preliminary affiliation of clone sequences. After BLAST analysis, sequences were inserted in SILVA 104 database [20] and aligned using the appropriate tool from the ARB program package [21]. A deeper phylogenetic analysis was performed on the Proteobacteria phylum, because (i) it included the most represented 15 T-RFLP profiles in the two libraries, and (ii) it was one of the main groups previously retrieved in marine sediments (e.g., [37, 38]). Two phylogenetic reconstructions on different subclasses of Proteobacteria were performed independently. The analysis focuses in particular on two groups of sequences: the first one including bacteria belonging to the Deltaproteobacteria

The Scientific World Journal

3

Site 3

N

W

Site 4

E S

0

1

0.5 (km)

Site 5

Site 2

Site 1

Figure 1: Representation of the Leghorn’s Harbor area. The tagged marks represent the sampling sites in the seabed of the area, which are briefly described as follows. Site 1: ferry boat departure area (sandy loam; water depth 11 m). Site 2: seldom dredged shipbuilding area (loam; water depth 4 m; last dredging 60 years ago). Site 3: container terminal (silt loam; water depth 13 m; located by the open sea mouth of the Navicelli canal). Site 4: cargo ferry transit (loam; water depth 6 m). Site 5: chemical processing and oil refinery terminal (silty clay loam; water depth 9 m; located by the mouth of Ugione stream, which crosses the inland industrial area and collects some spare wastewater discharges).

Table 1: Chemical parameters of the five sampling sites [13]. Samples 1.1 1.2 1.3 2.1 2.2 2.3 3.1 3.2 3.3 4.1 4.2 4.3 5.1 5.2 5.3

Cd Ni Pb Cr Cu Zn TPH TOC TN TP NO3 − Cl− SO4 2− mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgdw mg/kgfw g/kgfw mg/kgfw 1.06 37.7 33.2 25.2 75.0 313 1433 6102 197 917 831 10 3587 0.63 25.2 271.0 34.7 73.5 353 1066 8938 225 1024 975 12 4211 0.71 39.6 104.0 15.8 268.0 407 1030 7720 304 1321 903 11 3899 1.16 41.0 249.0 48.8 80.5 308 1915 25460 2170 696 1161 17 3192 1.42 75.4 221.0 35.4 375.0 854 2031 33540 1580 759 1363 19 3747 1.31 94.9 171.0 27.7 485.0 884 1798 27500 1890 722 1262 18 3469 1.22 101.0 47.1 32.1 74.4 257 900 19664 1019 568 708 16 3417 1.44 98.4 36.5 35.6 71.5 336 766 25736 1325 651 832 19 4012 1.20 118.0 37.8 29.2 71.4 228 633 21200 1631 455 770 17 3714 1.09 63.1 58.4 16.9 123.0 568 1765 19572 730 616 1929 17 3848 1.20 67.0 54.3 18.7 91.3 277 1563 21100 878 504 2265 20 4517 1.11 67.4 90.0 25.0 86.4 247 1666 16628 602 570 2097 18 4183 1.36 100.0 62.0 34.6 119.0 500 1266 17148 1993 864 491 17 2807 1.59 103.0 65.6 39.4 127.0 476 1331 26652 1660 724 576 20 3295 1.48 135.0 67.7 42.6 135.0 368 1032 23900 1327 844 534 19 3051

TPH: total petroleum hydrocarbons; TOC: total organic carbon; TN: total N; TP: total P; dw: dry weight; fw: fresh sample weight. All parameters are means of three replicates.

4

The Scientific World Journal

subclass using Epsilonproteobacteria as an outgroup, and the second one focusing on the Gammaproteobacteria group with some sequences of Alphaproteobacteria used as an outgroup. Both trees were constructed using maximum likelihood algorithm, 100 bootstraps, and a filter specifically designed for the selection of sequences in each tree, considering only positions conserved in at least 10% of sequences. A distance matrix, constructed using the neighbor joining algorithm, was also calculated and examined for clone sequences of the two samples that clustered together with described and cultivated bacterial species, in order to understand which clades represented the species-level or the genus-level groups. In the construction and interpretation of similarity matrices, we used the 95% limit to represent the threshold for genus definition and the limit of 97% to represent the threshold for OTUs [22]. 2.4. Statistical Analysis of Clone Libraries. OTUs were identified in each library using MOTHUR software [23]. Richness and alpha diversity indices, including the Chao 1 estimator [24] and the Shannon index, were calculated for each library at different cutoff levels (0.01, 0.03, and 0.05). The Chao 1 estimator evaluates the richness of a total species as Chao 1 = Sobs +

𝑛12 , 2𝑛2

(1)

where Sobs is the number of observed species, 𝑛1 is the number of singletons (species captured once), and 𝑛2 is the number of doubletons (species captured twice). Shannon’s diversity index [25] was obtained as 𝐻 = −Σ𝑝𝑖 ln 𝑝𝑖 ,

(2)

where 𝑝𝑖 is the population of each species 𝑖, and the resulting product is summed up across species and multiplied by −1 [26]. The MOTHUR software was also applied to calculate the number of OTUs shared among the two different libraries at different cutoff levels. The LIBSHUFF software [27], based on the Jukes-Cantor pairwise distance matrix, was also applied to find out similarities in the two libraries, as described by Zhang et al. [28].

3. Results 3.1. Construction of 16S rRNA Gene Libraries and Detection of Chimeric Sequences. Two libraries were constructed on samples 1.2 and 2.1. A total of 194 clones were sequenced, 101 of which taken from sample 1.2 and 93 from sample 2.1. All obtained sequences were submitted online on the DDBJ/EMBL/GenBank databases. They are available under the accession numbers from HE803828 to HE804037. The percentages of detected chimeric sequences were 11.9% in sample 1.2 (12 sequences) and 4.3% in sample 2.1 (4 sequences). All chimeric sequences were composed of partial sequences coming from only two different organisms. Considering fragments constituting chimeras as independent

sequences belonging to different organisms, the library from sample 1.2 was composed of a total of 115 sequences and the library from sample 2.1 of 95 sequences. The affiliation of nucleotide sequences was determined by BLAST analysis. In both libraries, the highest percentage of screened clones emerged as belonging to the Proteobacteria phylum (74% to library 1.2 and 73% to library 2.1, Figure 2(a)). Within this group, bacteria belonging to Alpha-, Gamma-, Delta-, and Epsilonproteobacteria classes were present in all libraries, while bacteria belonging to the Betaproteobacteria class were detected only in library 2.1 with low percent values (1%). In particular, the Gammaproteobacteria subclass was the most represented one in both 1.2 and 2.1 libraries (35% and 38%, resp.), followed by the Deltaproteobacteria subclass (resp., 24% and 20%). The second most represented bacterial taxon detected in both libraries was the Bacteroidetes phylum, represented by 12% of the total sequences in library 1.2, and by 6% of the sequences in library 2.1. There are some differences in the distribution of bacterial phyla detected in the two libraries, especially for minor groups such as the Lentisphaerae, Nitrospirae, and Thermotogae phyla, that were present only in sample 1.2 with percent values lower than 1%, and the Chlorobi, Verrucomicrobia, Deferribacteres, and Gemmatimonadetes phyla, that were detected in sample 2.1 with fractions ranging from 1% to 2%. 3.2. T-RFLP Analysis. NMDS analysis on T-RFLP data matrix highlights some groups of samples that correspond to the five sampling sites (Figure 3). The grouping of the triplicates is mainly evident for sites 1 (samples 1.1, 1.2, and 1.3) and 4 (samples 4.1, 4.2, and 4.3). Triplicates of sites 2 and 3 clustered together in the central part of the plot, and the triplicates from site 5 (samples 5.1, 5.2, and 5.3) were all located in the second quadrant of the plot, isolated from all the other samples. The Shannon diversity index calculated from the T-RFLP data matrix ranged from 3.17 to 3.75. Figure 4 presents some of the results of the attribution of the peaks obtained by in silico clone digestion, in comparison with the peaks detected in the microbial communities of the five sampling sites. The coverage percentages of the bacterial groups in T-RFLP profiles are shown in Figure 2(b). Not all the bacterial OTUs detected with the rDNA clone library could be retrieved in the T-RFLP electropherograms of the 15 samples. All OTUs retrieved in the T-RFLP electropherograms through in silico digestion belong to subclasses of the Proteobacteria phylum, with the only exception of the Betaproteobacteria class, which has not been detected. Figure 2(b) highlights a dominance of Gammaproteobacteria in sites 1 and 2 and a dominance of Deltaproteobacteria organisms in sites 3, 4, and 5. The less represented group of Proteobacteria was the Epsilonproteobacteria in all BsuRI samples. When considering the RsaI restriction endonuclease, the Alphaproteobacteria subclass emerged as the less represented group in each case. 3.3. Phylogenetic Analysis. In order to assess the relative abundance of Proteobacteria related sequences, we focused on the phylogenetic analysis of this group of organisms. In particular, two phylogenetic trees were built: the first one comprised all the retrieved Delta- and Epsilonproteobacteria clone

The Scientific World Journal

5

1.2

2.1

1% 2% 1% 1% 1% 1%

35% 3% 3%

10% 5%

74%

12%

2% 1% 1% 1%

8% 38% 3%

1% 1%

24%

1% 7% 6%

73%

1% 6%

20%

1%

1%

Spirochaetes Bacteroidetes Chlorobi Verrucomicrobia Lentisphaerae Planctomycetes Deferribacteres Others Chloroflexi Nitrospirae

Spirochaetes Bacteroidetes Chlorobi Verrucomicrobia Lentisphaerae Planctomycetes Deferribacteres Others Chloroflexi Nitrospirae

Gemmatimonadetes Acidobacteria Cyanobacteria Thermotogae Gammaproteobacteria Betaproteobacteria Alphaproteobacteria Epsilonproteobacteria Deltaproteobacteria

Gemmatimonadetes Acidobacteria Cyanobacteria Thermotogae Gammaproteobacteria Betaproteobacteria Alphaproteobacteria Epsilonproteobacteria Deltaproteobacteria

(a)

BsuRI

30

25

25

20

20

15

15

10

10

5

5

0

1

2

3

4

0

5

Epsilon (%)

Gamma (%) Delta (%)

RsaI

30

1

2

3

Alpha (%) Delta (%)

4

5

Epsilon/Bacteroidetes

(b)

Figure 2: (a) Comparison among main bacteria phyla detected in the two clone libraries; the larger pink sectors represent Proteobacteria. (b) Coverage percentages of the main bacterial groups attributed through in silico digestion of the 15 T-RFLP electropherograms. Each of the five sites is represented by the means of three replicates; the black bars represent the standard deviations. 5.1

0.32 0.24 0.16 2.2 0.08 3.2

Coordinate 2

4.2

4.1

−0.32 4.3

−0.24

−0.16

−0.08 3.1 3.3 −0.08 2.3 2.1 −0.16 −0.24 Coordinate 1

5.2

0.08

5.3 0.16

0.24 1.2 1.3 1.1

Figure 3: NMDS plot obtained from T-RFLP data analysis.

6

The Scientific World Journal

3000

3500

4000

4500

5000

5500

6000

6500

7000 Sample 1.1

450 400 350 300 250 200 150 100 50 0 (a) 3000

3500

4000

4500

5000

5500

6000

6500

7000 Sample 2.3

700 600 500 400 300 200 100 0 (b) 3000 2500 2250 2000 1750 1500 1250 1000 750 500 250 0

3500

4000

4500

5000

5500

6000

6500

7000 Sample 3.1

(c) 3000

3500

4000

4500

5000

5500

6000

6500

450 400 350 300 250 200 150 100 50 0

7000 Sample 4.3

(d) 3000

3500

4000

4500

5000

5500

6000

6500

7000 Sample 5.3

1800 1600 1400 1200 1000 800 600 400 200 0 Enzyme BsuRI Enzyme RsaI (e)

Figure 4: T-RFLP peak attribution after in silico digestion.

The Scientific World Journal sequences and their closer relatives (Figure 5); the second one included all the retrieved Gamma- and Alphaproteobacteria clone sequences and their closer relatives (Figure 6). The trees were constructed using PHYML with 100 bootstraps [29] from the ARB package. The majority of sequences in the trees are derived from studies about marine sediment samples. Some of our clones are closely related to symbiotic bacteria found in marine metazoan bodies [30, 31]. These sequences belong to the Gammaproteobacteria subclass; clone 147b from library 2.1 shows a 99% similarity with the already mentioned Spongiispira norvegica (AM117931, [31]), while the OTU composed of clones 228, 143, 162, 61, and 160, all coming from library 1.2, shows a 98% similarity with Endozoicomonas elysicola (AB196667, [30]). All the clones related to the Deltaproteobacteria subclass cluster together with species having a metabolism involving the presence of sulphur (most sulphate reducing microorganisms). The majority of published sequences that are closely related to our clone sequences belong to uncultured bacteria, with the exception of some Deltaproteobacteria, which belong to Desulfobulbus, Desulfosarcina, and Desulfonema genera, all providing the sulfate reduction in the marine environment. The similarity matrix evidenced that within the Deltaproteobacteria subclass our clone 188b from library 2.1 showed a 95% similarity with the described Desulfobulbus japonicus species (AB110549, [32]), while the 52b clone from library 2.1 showed a 96% similarity with the described Desulfosarcina ovata (Y17286, [33]). The 124b and 130b sequences from library 2.1 and the sequence 184 from library 1.2 are all parts of the same OTU. They cluster together with the described Epsilonproteobacteria species Arcobacter nitrofigilis (L14627, [34]), symbiotic bacteria characterized by a nitrogen fixing metabolism. 3.4. Statistical Analysis of Clone Libraries. Table 2 shows the results concerning richness and alpha diversity indices in each of the two constructed libraries at three different cutoff levels. The Shannon diversity index shows similar values in both sampling sites, at all cutoff levels. The richness index (Chao1 estimator) emerged as being higher in sample 1.2 at 0.01 cutoff level and lower in sample 1.2 at 0.03 and 0.05 cutoff values. The number of detected OTUs emerged as being very similar in both sites. The OTUs at a 0.03 cutoff value were represented by the same number in both cases; library 2.1 at 0.01 and 0.05 cutoff values shows less detected OTUs than library 1.2 at the same cutoff levels (Table 2). Table 3 presents the MOTHUR analysis of OTUs that are shared between the two different libraries. At 0.01 cutoff (specie level), there are 7 OTUs shared between sites 1.2 and 2.1; at the 0.03 cutoff there are 12 shared OTUs; at 0.05 cutoff, the number of shared OTUs, which usually represents the genus level, rises up to 17. Figure 7 presents the results of the LIBSHUFF analysis. The two lines, representing the homologous and heterologous curves, are almost totally overlapping, thus indicating that the samples are similar to each other, as confirmed by the high 𝑃 value (𝑃 > 0.025 with Bonferroni correction, as described in [27]).

7 Table 2: Alpha diversity indices for the two samples at different cutoff values. Sample cutoff

Observed richness (OTUs)

Chao 1 estimator

Shannon index

1.2

0.01 0.03 0.05

82 75 71

667,2 258,3 193,8

4,36 4,24 4,17

2.1

0.01 0.03 0.05

80 75 68

620,2 306,1 239,1

4,33 4,24 4,07

Table 3: OTUs shared between site 1.2 and 2.1 calculated with MOTHUR software. Cutoff 0.01 0.03 0.05

Observed richness shared between 1.2 and 2.1 7 12 17

4. Discussion The molecular approach is commonly used in the recent literature for the study of bacterial communities in marine sediments [35–38] and to study the influence of bacterial communities in removing pollutants from marine sediments [39]. The two libraries highlighted a substantial similarity concerning the abundance and diversity of the sequences represented in the two different sites, even though the chemical composition was different [13]. The similarity between the two clone libraries is shown by the alpha diversity indices calculated with MOTHUR software (Table 2). The Shannon diversity index calculated for both libraries at different cutoff levels, on the base of the DNA sequences obtained after the library construction, ranged from 4.07 to 4.36. The same Shannon diversity index calculated on the base of T-RFLP profiles of the 15 samples ranged from 3.17 to 3.75. This difference can be explained because the diversity in a sample is underestimated when only T-RFLP profiles are used. This approach does not register rare OTUs. The Shannon diversity index used in this work should provide more reliable values, although an increase of the sample size of the sequenced clones of each library could probably provide even more accurate values. The construction of the two libraries and the screening of clones coupled with T-RFLP peak identification highlighted the dominance of Proteobacteria related microorganisms in all the five sites. There are no significant differences among the coverage percentages of different subclasses of the Proteobacteria phylum among the two libraries and the 5 sites; these data are in agreement with previously published studies, where PCR-based techniques on marine sediment samples revealed a dominance of bacteria related to Proteobacteria phylum [35, 37] or to Firmicutes, Delta-, and Gammaproteobacteria [40, 41]. In other papers, the analysis of marine sediment showed a dominance of Actinobacteria [38]. In this study, no significant differences in the coverage percentage of different subclasses of the Proteobacteria phylum were

8

The Scientific World Journal

100%

100%

0.10

Uncultured bacterium, FM179902, 97% Uncultured bacterium, FN396770, 100% Uncultured bacterium, AM040136 94% 164 Agriport 1b 100% 16 Agriport 1b Uncultured bacterium, FN421322 19b Agriport 2a 100% Uncultured bacterium, AB530217 84% Uncultured bacterium, EU617853 Uncultured bacterium, GU230394 94% 181b Agriport 2a 76% Uncultured bacterium, FN396678 100% Uncultured bacterium, AY907763 Uncultured bacterium, AJ620500 7 Agriport 1b 98% 99% 230 Agriport 1b Uncultured bacterium, DQ394998 80% Uncultured bacterium, GQ259292 194 Agriport 1b 190b Agriport 2a 100% 158 Agriport 1b Uncultured bacterium, EU592471 71% 52b Agriport 2a Uncultured bacterium, FM242403 100% Uncultured bacterium, DQ351771 Desulfosarcina ovata, Y17286 96% 184b Agriport 2a 100% Desulfonema magnum, U45989 159b Agriport 2a 99% Uncultured bacterium, AY799901 90% Uncultured bacterium, FN429813 61b Agriport 2a 100% Uncultured bacterium, FJ748772 99% Uncultured bacterium, AM746084 173b Agriport 2a 98% Uncultured bacterium, GQ246438 100% Uncultured bacterium, DQ395025 Uncultured bacterium, EU386051 83% 221 Agriport 1b 100% Uncultured bacterium, GU302477 100% Uncultured bacterium, Y771932 90% 179 Agriport 1b 97% Uncultured bacterium, DQ351762 100% Uncultured bacterium, AY771964 73% Uncultured Desulfocapsa sp., DQ831542 100% 188b Agriport 2a 100% Desulfobulbus sp., U85473 98% Desulfobulbus japonicus, AB110549 63 Agriport 1b Uncultured bacterium, DQ831547 71% Uncultured bacterium, DQ831555 Uncultured bacterium, GQ246300 Uncultured bacterium, GQ246285 100% Uncultured bacterium, DQ831548 99% 51 Agriport 1b Uncultured bacterium, FJ416070 100% Uncultured bacterium, AB530214 95% 59b Agriport 2a 191b Agriport 2a 100% 188 Agriport 1b 100% Uncultured bacterium, FJ202627 Uncultured bacterium, EF123510 Desulfovibrio sp. r02, AB470955 Uncultured bacterium, FJ748774 100% Uncultured bacterium, EU617844 141 Agriport 1b Uncultured bacterium, EF999371 100% 74 Agriport 1b Uncultured bacterium, FM165238 199 Agriport 1b 100% Uncultured bacterium, DQ833497 99% Uncultured bacterium, FJ516915 Uncultured bacterium, DQ395007 100% 31a Agriport 2a, Uncultured bacterium, EU385735 Uncultured bacterium, DQ267496 100% Uncultured bacterium, FJ517013 99% 163b Agriport 2a Uncultured bacterium, AM997597 6 Agriport 1b 79% Uncultured bacterium, GU291346 100% 192b Agriport 2a Uncultured bacterium, FJ545548 88% Pelobacter sp. A3b3, AJ271656 136a Agriport 2a Uncultured bacterium, GQ249592 95% 117b Agriport 2a 34 Agriport 1b 5 Agriport 1b 94% 91% Uncultured bacterium, EU700147 100% Uncultured bacterium, GU180167 146 Agriport 1b 100% Uncultured bacterium, GQ246309 90% Uncultured bacterium, EU652627 95% 79a Agriport 2a 100% Uncultured bacterium, EU592455 100% Uncultured bacterium, EU488292 178 Agriport 1b 100% 2 Agriport 1b Uncultured bacterium, FJ202384 Uncultured bacterium, EU617864 97% 124b Agriport 2a 100% 130b Agriport 2a 184 Agriport 1b 98% Uncultured Arcobacter sp., AB262371 100% Arcobacter nitrofigilis, L14627 Arcobacter butzleri, AY621116 12 Agriport 1b 100% Uncultured bacterium, FJ717179 97% 157 Agriport 1b 100% Sulfurimonas autotrophica, AB088431 Uncultured bacterium, AB193949 Uncultured Campylobacterales bacterium, DQ234141 38a Agriport 2a Uncultured bacterium, AY218555 78% 137 Agriport 1b 158b Agriport 2a Uncultured bacterium, FJ717140 183 Agriport 1b

Figure 5: Maximum likelihood phylogenetic tree made with PHYML and 100 bootstrap pseudoreplicates. The tree represents the phylogenetic position of characterized Delta- and Epsilonproteobacteria clone sequences together with closely related sequences present in the database. The characterized sequences are in bold.

The Scientific World Journal

9 100%

Uncultured bacterium, EU617749 Uncultured bacterium, AB294972 Uncultured bacterium, FJ712598 206b Agriport 2.1 94b Agriport 2.1 Uncultured bacterium, DQ351764 80% 5 Agriport 1.2 100% 100% 227 Agriport 1.2 39b Agriport 2.1 99% 60b Agriport 2.1 100% Uncultured bacterium, FJ717240 100% 100b Agriport 2.1 Uncultured bacterium, FJ717244 100% 73 Agriport 1.2 89% 21 Agriport 1.2 Uncultured bacterium, AM882548 Uncultured bacterium, GU119320 Uncultured bacterium, DQ394975 100% 2a Agriport 2.1 Beggiatoa sp., AF110276 225 Agriport 1.2 76% 8 Agriport 1.2 75% Uncultured bacterium, AY225632 99% 82a Agriport 2.1 100% Uncultured bacterium, FJ497588 Uncultured bacterium, GU369896 219 Agriport 1.2 100% 56b Agriport 2.1 Uncultured bacterium, FJ717231 85% Uncultured bacterium, AF223299 97% Uncultured bacterium, AY193138 120b Agriport 2.1 99% Uncultured bacterium, FJ358865 91% Uncultured bacterium, DQ351795 173 Agriport 1.2 97% Uncultured bacterium, DQ395021 80b Agriport 2.1 95% 100% Uncultured bacterium, GU061255 100% 45a Agriport 2.1 100% 140 Agriport 1.2 Uncultured bacterium, AM889165 92% 22b Agriport 2.1 100% Uncultured bacterium, GQ346961 80% Uncultured bacterium, GQ337082 95% 13b Agriport 2.1 100% Uncultured bacterium, EU488031 Uncultured bacterium, GQ246299 100% Uncultured bacterium, FJ545438 70% 86b Agriport 2.1 Uncultured bacterium, GU230336 75% 48a Agriport 2.1 79% 62b Agriport 2.1 Uncultured Pseudomonas sp., DQ395002 100% Uncultured bacterium, GQ246314 147 Agriport 1.2 93% 94% Uncultured bacterium, GQ249568 73% Uncultured bacterium, GU230337 100% Uncultured bacterium, GQ246429 99% 223 Agriport 1.2 Uncultured bacterium, FJ628326 100% 193 Agriport 1.2 100% Uncultured bacterium, GU369889 Uncultured bacterium, AB294964 Thiohalomonas nitratireducens, DQ836238 70% Uncultured bacterium, EU652539 97% 72a Agriport 2.1 82% 139b Agriport 2.1 100% Uncultured bacterium, GU230360 84% 169 Agriport 1.2 Uncultured bacterium, EU491143 100% 99% Uncultured bacterium, EU491145 3a Agriport 2.1 Uncultured bacterium, FJ497645 98% 100% 77b Agriport 2.1 Uncultured bacterium, GU061260 71% 46a Agriport 2.1 82% Uncultured bacterium, AB294952 100% 70 Agriport 1.2 78% Uncultured bacterium, AM882516 98% Uncultured bacterium, AM882511 Uncultured bacterium, AM882517 90% 171 Agriport 1.2 90% Uncultured bacterium, FJ497360 Uncultured bacterium, GU108533 100% 147b Agriport 2.1 82% Uncultured bacterium, FM958463 Spongiispira norvegica, AM117931 8b Agriport 2.1 99% Uncultured bacterium, FJ626892 100% Uncultured bacterium, FJ626896 93% Thalassomonas actiniarum, AB369380 100% 59 Agriport 1.2 100% Glaciecola sp. HTCC2999, ABST01000050 100% Uncultured bacterium, FN435248 Glaciecola punicea, U85853 65b Agriport 2.1 100% 39a Agriport 2.1 Uncultured bacterium, EF379837 Amphritea japonica, AB330881 88% 75 Agriport 1.2 100% Uncultured bacterium, FJ744816 Uncultured bacterium, EU050812 228 Agriport 1.2 143 Agriport 1.2 100% 162 Agriport 1.2 61 Agriport 1.2 160 Agriport 1.2 100% Uncultured bacterium, GU118490 elysicola, AB196667 97% 93Endozoicomonas Agriport 1.2 100% Uncultured bacterium, FN435240 Uncultured bacterium, EU287400 85b Agriport 2.1 100% Uncultured bacterium, FM242245 Uncultured bacterium, BD411343 98% 76b Agriport 2.1 98% Uncultured bacterium, FM242466 97% Uncultured bacterium, EU799754 100% Uncultured bacterium, AM997787 Uncultured bacterium, GU230343 90% Uncultured bacterium, EU488035 41b Agriport 2.1 Uncultured bacterium, EU488024 211 Agriport 1.2 87% Uncultured bacterium, EU328058 Uncultured bacterium, GU118250 100% 213 Agriport 1.2 144 Agriport 1.2 86% Uncultured bacterium, FJ717220 Uncultured bacterium, GU118337 100% 15b Agriport 2.1 Uncultured bacterium, EU287169 153 Agriport 1.2 77% Uncultured bacterium, FM242243 100% Uncultured bacterium, DQ351782 155b Agriport 2.1 Uncultured bacterium, AB271120 76% Uncultured bacterium AB271123 74% 81% Uncultured bacterium, AB271124 100% 182b Agriport 2.1 100% Uncultured bacterium, FJ712548 220 Agriport 1.2 100% Uncultured bacterium, AB015250 Uncultured bacterium, FJ787300 76% 119b Agriport 2.1 Uncultured bacterium, FM242244 70% 52a Agriport 2.1 55a Agriport 2.1 94% Uncultured bacterium, AB530202 222 Agriport 1.2 Uncultured bacterium, EU707297 100% Uncultured bacterium, EU652547 80% 206 Agriport 1.2 100% 99 Agriport 1.2 Uncultured bacterium, FJ717249 99% Roseobacter sp., AY258102 88% 161b Agriport 2.1 100% 58b Agriport 2.1 Loktanella maricola, EF202613 Uncultured bacterium, EU050750 23b Agriport 2.1 100% 205 Agriport 1.2 170 Agriport 1.2 100% Uncultured bacterium, AY162058 Ochrobactrum anthropi, U70978 100% Uncultured bacterium, F62377A9 4 Agriport 1.2 79% 91 Agriport 1.2 84% Uncultured bacterium, FJ497610 89% Uncultured bacterium, AY588951 97% Uncultured bacterium, FJ497661 148 Agriport 1.2 20 Agriport 1.2 100% Uncultured bacterium, EU652596 Uncultured bacterium, GQ259326 88% Erythrobacter longus, L01786 100% Uncultured bacterium, FJ545435 103a Agriport 2.1 100%

100%

76%

100%

100%

0.10

Figure 6: Maximum likelihood phylogenetic tree made with PHYML and 100 bootstrap pseudoreplicates. The tree represents the phylogenetic position of characterized Alpha- and Gammaproteobacteria clone sequences together with closely related sequences present in the database. The characterized sequences are in bold.

10

The Scientific World Journal 1

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2

P = 0.1341

0.1 0

0

0.1

0.2

0.3

0.4

1b 2a

Figure 7: Results of LIBSHUFF comparison between the 1.2 and 2.1 libraries. The solid line indicates the homologous coverage curve 𝐶𝑋 (𝐷), and the broken line indicates the heterologous coverage curve 𝐶𝑋𝑌 (𝐷). The 𝑃 value is 𝑃 > 0.025, thus indicating that the two libraries are similar to each other in OTU composition. The 𝑦-axis represents the coverage (𝐶) while the 𝑥-axis represents the evolutionary distance (𝐷).

observed among the five sites; this agrees with other studies in which different libraries of marine sediments collected in different sites showed similar bacterial compositions [35]. The Bacteroidetes phylum is the only further phylum with an appreciable presence in all sites, and the percentages of abundance are comparable to those found by other authors [35, 38]. This result can mean that bacteria belonging to these groups play the main role in nutrient recycling in the harbor seabed ecosystem. Considering the digestion profiles obtained with the BsuRI restriction endonuclease, Gammaproteobacteria emerged as being dominant in sites 1 and 2, while the sediments were dominated by Deltaproteobacteria organisms in sites 3, 4, and 5. Considering the RsaI restriction endonuclease, sites 2, 4, and 5 emerged as being dominated by Deltaproteobacteria microorganisms, while sites 1 and 3 were dominated by Epsilonproteobacteria and Bacteroidetes, which are groups of bacteria sharing the same peak interval. Some groups of bacteria could be detected only with one of the two adopted restriction enzymes. For instance, Alphaproteobacteria were not detected with BsuRI and Gammaproteobacteria were not detected with RsaI. This circumstance can be explained by the fact that the first restriction site in the gene sequences of those bacteria is located outside the interval 50–500 bp, which is the interval covered by our T-RFLP analysis. In fact, the use of more than one restriction enzyme is a recommended practice in this technique, in order to facilitate the resolution of bacterial populations [42, 43]. The fact that other OTUs identified through clone library construction were not identified in T-RFLP profiles can be in part explained by their poor representation in the sediment samples, causing the failure of their T-RFLP quantification. Overall, our results are in agreement with results of other authors that examined samples of port sediments.

In the study performed by Wu et al. [44], three different samples of marine sediments were investigated in order to characterize their bacterial communities. In all samples, the dominant phylum was Proteobacteria with a presence of 82%, 42%, and 42% in the three samples. Zhang et al. [28] analyzed four sites characterized by different pollution levels in Victoria Harbor (Hong Kong). In all sites, they found the dominance of bacteria belonging to the Proteobacteria phylum and, especially, to the same subclasses detected in our work. The dominance of Proteobacteria was also found in many other studies (e.g., [37, 45, 46]). The diversity index calculated by Zhang et al. [28] did not highlight significant differences among the four analyzed sites; this result is similar to the finding of this work, where the two analyzed libraries highlighted similar Shannon diversity values. Bacteria belonging to the Alphaproteobacteria subclass are important for hydrocarbon degradation [47]. The presence of these bacterial taxa in all five sites confirms a contamination of hydrocarbons in Leghorn seabed sediments [13]. A large portion of the sequences detected in our clone libraries belong to the Deltaproteobacteria subclass. With a deep phylogenetic analysis, we discovered that all these clones are closely related to bacteria characterized by a metabolism involved in the removal of sulphur. This observation is in agreement with other two works performed on marine port sediments [28, 48]. The percentage of Deltaproteobacteria detected in the two different samples is not significantly different (24% library 1.2 and 20% library 2.1), and TRFLP results suggest that their abundance should be roughly comparable among all the five sites. This result does not agree with the chemical results published in Iannelli et al. [13], which evidenced that the concentration of SO4 2− in site 2.1 (4399 mg/(kg dw)) was double that in site 1.2 (2321 mg/(kg dw)).

The Scientific World Journal An interesting finding concerns the Gammaproteobacteria subclass, in particular the sequences 228, 143, 162, 61, and 160, from library 1.2, and the sequence 147b from library 2.1. The first group of sequences, all representing the same OTU, shows a 98% similarity with Endozoicomonas elysicola, which is a typical symbiont of the Elysia ornata slug. Conversely, the sequence 147b is 99% similar to the sequence of Spongiispira norvegica, a typical symbiont of sponges. This observation possibly suggests that site 1.2 is richer in marine metazoans characterized by a symbiotic association with bacteria than site 2.1. This difference is probably due to the chemical characteristics of the sediments in Leghorn seabed area. Site 1.2 in fact, emerged as being less polluted than site 2.1, and probably more suitable for being colonized by marine metazoans harboring bacterial symbionts.

5. Conclusions In the present study, bacterial communities from five sites of Leghorn Harbor seabed were analyzed and identified through T-RFLP analysis, 16S rRNA library construction, and in silico digestion of retrieved clones. Some considerations emerged about organisms involved in the recycling of organic matter. Both the diversity indices and the phylogenetic affiliation of bacterial sequences obtained from the molecular screening highlighted a substantial similarity in the composition of the bacterial communities. This finding contrasts with the chemical characterization of an earlier study, where the same sites evidenced significant differences in the presence of nutrients and pollutants, and the T-RFLP analysis evidenced a heterogeneity among the different sites of the harbor. Retrieved sequences from the two libraries were more than sufficient to provide generic information on metabolism present in all of the 15 seabed sediment samples and to compare them with previous similar studies. Only a significantly higher sequencing coverage would have probably allowed to distinguish between the samples that looked rather similar in the presented analysis. The T-RFLP approach proved to be more efficient in highlighting the bacterial community structure in the whole harbor area, whereas the present work helped in clarifying the role of specific bacteria present in the studied samples and their related metabolisms.

Acknowledgments The authors acknowledge the Livorno Port Authority for their support in supplying information, historical data, the database of the ongoing environmental site assessment, and technical assistance for samplings. S. Gabrielli is gratefully acknowledged for his help with photographic artwork. Dr. Maria Nella Dell’Orbo is gratefully acknowledged for her help with the English revision. This study was performed within the CIP Eco-Innovation Project ECO/08/239065AGRIPORT, “Agricultural Reuse of Polluted Dredged Sediments.”

11

References [1] F. J. Estacio, E. M. Garc´ıa-Adiego, D. A. Fa et al., “Ecological analysis in a polluted area of Algeciras Bay (southern Spain): external “versus” internal outfalls and environmental implications,” Marine Pollution Bulletin, vol. 34, no. 10, pp. 780–793, 1997. [2] N. Dhainaut-Courtois, C. Pruvot, A. Empis, and K. Baudet, “Les peuplements macrozoobenthiques, indicateurs des qualit´es physicochimiques et chimiques des s´ediments portuaires: example du port de Boulogne-Sur-Mer (Manche),” Journ´ees Annuelles de la Soci´et´e Zoologique de France, vol. 125, no. 1, pp. 49–62, 2000. [3] E. Danulat, P. Muniz, J. Garc´ıa-Alonso, and B. Yannicelli, “First assessment of the highly contaminated harbour of Montevideo, Uruguay,” Marine Pollution Bulletin, vol. 44, no. 6, pp. 554–565, 2002. [4] J. M. Guerra-Garc´ıa and J. C. Garc´ıa-G´omez, “Oxygen levels versus chemical pollutants: do they have similar influence on macrofaunal assemblages? A case study in a harbour with two opposing entrances,” Environmental Pollution, vol. 135, no. 2, pp. 281–291, 2005. [5] G. Cognetti and F. Maltagliati, “Port activities and international environmental control in the Mediterranean,” Marine Pollution Bulletin, vol. 50, no. 2, pp. 119–120, 2005. [6] M. Azad, P. Peterma, K. Echols et al., “Polychlorinated biphenyls (PCBs) and organochlorine pesticides (OCPs) in harbor sediments from Sea Lots, Port-of-Spain, Trinidad and Tobago,” Marine Pollution Bulletin, vol. 62, no. 6, pp. 1324–1332, 2011. [7] G. Cornelissen, A. Pettersen, E. Nesse, E. Eek, A. Helland, and G. D. Breedveld, “The contribution of urban runoff to organic contaminant levels in harbour sediments near two Norwegian cities,” Marine Pollution Bulletin, vol. 56, no. 3, pp. 565–573, 2008. [8] K. Kaufmann, M. Christophersen, A. Buttler, H. Harms, and P. H¨ohener, “Microbial community response to petroleum hydrocarbon contamination in the unsaturated zone at the experimental field site Værløse, Denmark,” FEMS Microbiology Ecology, vol. 48, no. 3, pp. 387–399, 2004. [9] L. M. Buruaem, M. A. Hortellani, J. E. Sarkis, L. V. Costa-Lotufo, and D. M. S. Abessa, “Contamination of port zone sediments by metals from Large Marine Ecosystems of Brazil,” Marine Pollution Bulletin, vol. 64, no. 3, pp. 479–488, 2012. [10] P. M. Chapman and M. Smith, “Assessing, managing and monitoring contaminated aquatic sediments,” Marine Pollution Bulletin, vol. 64, no. 10, pp. 2000–2004, 2012. [11] S. Doni, C. Macci, E. Peruzzi, R. Iannelli, B. Ceccanti, and G. Masciandaro, “Decontamination and functional reclamation of dredged brackish sediments,” Biodegradation, vol. 24, no. 4, pp. 499–512, 2013. [12] M. Y. Sun, K. A. Dafforn, M. V. Brown, and E. L. Johnston, “Bacterial communities are sensitive indicators of contaminant stress,” Marine Pollution Bulletin, vol. 64, no. 5, pp. 1029–1038, 2012. [13] R. Iannelli, V. Bianchi, C. Macci et al., “Assessment of pollution impact on biological activity and structure of seabed bacterial communities in the Port of Livorno (Italy),” Science of the Total Environment, vol. 426, pp. 56–64, 2012. [14] R. I. Amann, W. Ludwig, and K.-H. Schleifer, “Phylogenetic identification and in situ detection of individual microbial cells without cultivation,” Microbiological Reviews, vol. 59, no. 1, pp. 143–169, 1995.

12 [15] D. J. Lane, “16S/23S rRNA sequencing,” in Nucleic Acid Techniques in Bacterial Systematics, E. Stackebrandt and M. Goodfellow, Eds., pp. 115–147, John Wiley & Sons, New York, NY, USA, 1991. [16] Ø. Hammer, D. A. T. Harper, and P. D. Ryan, “Past: paleontological statistics software package for education and data analysis,” Palaeontologia Electronica, vol. 4, no. 1, 2001. [17] C. Chiellini, R. Iannelli, L. Modeo, V. Bianchi, and G. Petroni, “Biofouling of reverse osmosis membranes used in river water purification for drinking purposes: analysis of microbial populations,” Biofouling, vol. 28, no. 9, pp. 969–984, 2012. [18] P. Hugenholtz and T. Huber, “Chimeric 16S rDNA sequences of diverse origin are accumulating in the public databases,” International Journal of Systematic and Evolutionary Microbiology, vol. 53, no. 1, pp. 289–293, 2003. [19] S. F. Altschul, T. L. Madden, A. A. Sch¨affer et al., “Gapped BLAST and PSI-BLAST: a new generation of protein database search programs,” Nucleic Acids Research, vol. 25, no. 17, pp. 3389–3402, 1997. [20] E. Pruesse, C. Quast, K. Knittel et al., “SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB,” Nucleic Acids Research, vol. 35, no. 21, pp. 7188–7196, 2007. [21] W. Ludwig, O. Strunk, R. Westram et al., “ARB: a software environment for sequence data,” Nucleic Acids Research, vol. 32, no. 4, pp. 1363–1371, 2004. [22] E. Stackebrandt, W. Frederiksen, G. M. Garrity et al., “Report of the ad hoc committee for the re-evaluation of the species definition in bacteriology,” International Journal of Systematic and Evolutionary Microbiology, vol. 52, no. 3, pp. 1043–1047, 2002. [23] P. D. Schloss, S. L. Westcott, T. Ryabin et al., “Introducing mothur: open-source, platform-independent, communitysupported software for describing and comparing microbial communities,” Applied and Environmental Microbiology, vol. 75, no. 23, pp. 7537–7541, 2009. [24] A. Chao, “Non-parametric estimation of the number of classes in a population,” Scandinavian Journal of Statistics, vol. 11, pp. 265–270, 1984. [25] C. E. Shannon, “The mathematical theory of communication,” in The Mathematical Theory of Communication, C. E. Shannon and W. Weaver, Eds., pp. 29–125, University of Illinois Press, Urbana, Ill, USA, 1949. [26] A. E. Magurran, Ecological Diversity and Its Measurement, Chapman and Hall, London, UK, 1998. [27] D. R. Singleton, M. A. Furlong, S. L. Rathbun, and W. B. Whitman, “Quantitative comparisons of 16S rRNA gene sequence libraries from environmental samples,” Applied and Environmental Microbiology, vol. 67, no. 9, pp. 4374–4376, 2001. [28] W. Zhang, J.-S. Ki, and P.-Y. Qian, “Microbial diversity in polluted harbor sediments I: bacterial community assessment based on four clone libraries of 16S rDNA,” Estuarine, Coastal and Shelf Science, vol. 76, no. 3, pp. 668–681, 2008. [29] S. Guindon and O. Gascuel, “A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood,” Systematic Biology, vol. 52, no. 5, pp. 696–704, 2003. [30] M. Kurahashi and A. Yokota, “Endozoicomonas elysicola gen. nov., sp. nov., a 𝛾-proteobacterium isolated from the sea slug Elysia ornata,” Systematic and Applied Microbiology, vol. 30, no. 3, pp. 202–206, 2007.

The Scientific World Journal [31] I. Kaesler, I. Graeber, M. S. Borchert et al., “Spongiispira norvegica gen. nov., sp. nov., a marine bacterium isolated from the boreal sponge Isops phlegraei,” International Journal of Systematic and Evolutionary Microbiology, vol. 58, no. 8, pp. 1815–1820, 2008. [32] D. Suzuki, A. Ueki, A. Amaishi, and K. Ueki, “Diversity of substrate utilization and growth characteristics of sulfatereducing bacteria isolated from estuarine sediment in Japan,” Journal of General and Applied Microbiology, vol. 53, no. 2, pp. 119–132, 2007. [33] G. Harms, K. Zengler, R. Rabus et al., “Anaerobic oxidation of o-xylene, m-xylene, and homologous alkylbenzenes by new types of sulfate-reducing bacteria,” Applied and Environmental Microbiology, vol. 65, no. 3, pp. 999–1004, 1999. [34] I. V. Wesley, L. Schroeder-Tucker, A. L. Baetz, F. E. Dewhirst, and B. J. Paster, “Arcobacter-specific and Arcobacter butzleri-specific 16S rRNA-based DNA probes,” Journal of Clinical Microbiology, vol. 33, no. 7, pp. 1691–1698, 1995. [35] Y. Zeng, Y. Zou, B. Chen et al., “Phylogenetic diversity of sediment bacteria in the northern Bering Sea,” Polar Biology, vol. 34, no. 6, pp. 907–919, 2011. [36] H. Wang, X. Jiang, Y. He, and H. Guan, “Spatial and seasonal variations in bacterial communities of the Yellow Sea by T-RFLP analysis,” Frontiers of Environmental Science and Engineering in China, vol. 3, no. 2, pp. 194–199, 2009. [37] M. J. Elsab´e, V. Br¨uchert, and B. M. Fuchs, “Vertical shifts in the microbial community structure of organic-rich Namibian shelf sediments,” African Journal of Microbiology Research, vol. 6, no. 17, pp. 3887–3897, 2012. [38] C. Nithya and S. K. Pandian, “Evaluation of bacterial diversity in palk bay sediments using terminal-restriction fragment length polymorphisms (T-RFLP),” Applied Biochemistry and Biotechnology, vol. 167, no. 6, pp. 1763–1777, 2012. [39] O. Pringault, H. Viret, and R. Duran, “Influence of microorganisms on the removal of nickel in tropical marine sediments (New Caledonia),” Marine Pollution Bulletin, vol. 61, no. 7-12, pp. 530–541, 2010. [40] T. K¨ochling, P. Lara-Mart´ın, E. Gonz´alez-Mazo, R. Amils, and J. L. Sanz, “Microbial community composition of anoxic marine sediments in the Bay of C´adiz (Spain),” International Microbiology, vol. 14, pp. 143–154, 2011. [41] D. Castle and D. L. Kirchman, “Composition of estuarine bacterial communities assessed by denaturing gradient gel electrophoresis and fluorescence in situ hybridization,” Limnology and Oceanography, vol. 2, pp. 303–314, 2004. [42] W.-T. Liu, T. L. Marsh, H. Cheng, and L. J. Forney, “Characterization of microbial diversity by determining terminal restriction fragment length polymorphisms of genes encoding 16S rRNA,” Applied and Environmental Microbiology, vol. 63, no. 11, pp. 4516–4522, 1997. [43] T. L. Marsh, “Terminal restriction fragment length polymorphism (T-RFLP): an emerging method for characterizing diversity among homologous populations of amplification products,” Current Opinion in Microbiology, vol. 2, no. 3, pp. 323–327, 1999. [44] H. Wu, Y. Guo, G. Wang, S. Dai, and X. Li, “Composition of bacterial communities in deep-sea sediments from the South China Sea, the Andaman sea and the Indian Ocean,” African Journal of Microbiology Research, vol. 5, no. 29, pp. 5273–5283, 2011. [45] M. V. Brown, G. K. Philip, J. A. Bunge et al., “Microbial community structure in the North Pacific ocean,” ISME Journal, vol. 3, no. 12, pp. 1374–1386, 2009.

The Scientific World Journal [46] R. Schauer, C. Bienhold, A. Ramette, and J. Harder, “Bacterial diversity and biogeography in deep-sea surface sediments of the South Atlantic Ocean,” ISME Journal, vol. 4, no. 2, pp. 159–170, 2010. [47] L. Rocchetti, F. Beolchini, K. B. Hallberg, D. B. Johnson, and A. Dell’ Anno, “Effects of prokaryotic diversity changes on hydrocarbon degradation rates and metal partitioning during bioremediation of contaminated anoxic marine sediments,” Marine Pollution Bulletin, vol. 64, no. 8, pp. 1688–1698, 2012. [48] W. Zhang, L.-S. Song, J.-S. Ki, C.-K. Lau, X.-D. Li, and P.Y. Qian, “Microbial diversity in polluted harbor sediments II: sulfate-reducing bacterial community assessment using terminal restriction fragment length polymorphism and clone library of dsrAB gene,” Estuarine, Coastal and Shelf Science, vol. 76, no. 3, pp. 682–691, 2008.

13

International Journal of

Peptides

BioMed Research International Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Advances in

Stem Cells International Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Virolog y Hindawi Publishing Corporation http://www.hindawi.com

International Journal of

Genomics

Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Nucleic Acids

Zoology

 International Journal of

Hindawi Publishing Corporation http://www.hindawi.com

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Volume 2014

Submit your manuscripts at http://www.hindawi.com The Scientific World Journal

Journal of

Signal Transduction Hindawi Publishing Corporation http://www.hindawi.com

Genetics Research International Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Anatomy Research International Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Enzyme Research

Archaea Hindawi Publishing Corporation http://www.hindawi.com

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Biochemistry Research International

International Journal of

Microbiology Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

International Journal of

Evolutionary Biology Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Molecular Biology International Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Advances in

Bioinformatics Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Journal of

Marine Biology Volume 2014

Hindawi Publishing Corporation http://www.hindawi.com

Volume 2014

Bacterial communities in polluted seabed sediments: a molecular biology assay in Leghorn Harbor.

Seabed sediments of commercial ports are often characterized by high pollution levels. Differences in number and distribution of bacteria in such area...
3MB Sizes 0 Downloads 0 Views