Next Article in Journal
Genetic Characterization of Staphylococcus aureus, Staphylococcus argenteus, and Coagulase-Negative Staphylococci Colonizing Oral Cavity and Hand of Healthy Adults in Northern Japan
Previous Article in Journal
Characterization of Vaginal Microbiota in Third Trimester Premature Rupture of Membranes Patients through 16S rDNA Sequencing
Previous Article in Special Issue
Genomic and Biological Profile of a Novel Bacteriophage, Vibrio phage Virtus, Which Improves Survival of Sparus aurata Larvae Challenged with Vibrio harveyi
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibrio Phage Artemius, a Novel Phage Infecting Vibrio alginolyticus

by
Stavros Droubogiannis
1,2,†,
Lydia Pavlidi
1,†,
Maria Ioanna Tsertou
1,
Constantina Kokkari
1,
Dimitrios Skliros
3,
Emmanouil Flemetakis
3 and
Pantelis Katharios
1,*
1
Hellenic Centre for Marine Research, Institute of Marine Biology, Biotechnology & Aquaculture, 71500 Heraklion, Greece
2
Department of Biology, School of Sciences and Engineering, University of Crete, 71500 Heraklion, Greece
3
Department of Biotechnology, Agricultural University of Athens, 11855 Athens, Greece
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pathogens 2022, 11(8), 848; https://doi.org/10.3390/pathogens11080848
Submission received: 21 June 2022 / Revised: 21 July 2022 / Accepted: 27 July 2022 / Published: 28 July 2022

Abstract

:
Vibrio alginolyticus is an important pathogen of marine animals and has been the target of phage therapy applications in marine aquaculture for many years. Here, we report the isolation and partial characterization of a novel species of the Siphoviridae family, the Vibrio phage Artemius. The novel phage was species-specific and could only infect strains of V. alginolyticus. It could efficiently reduce the growth of the host bacterium at various multiplicities of infection as assessed by an in vitro lysis assay. It had a genome length of 43,349 base pairs. The complete genome has double-stranded DNA with a G + C content of 43.61%. In total, 57 ORFs were identified, of which 19 were assigned a predicted function. A genomic analysis indicated that Vibrio phage Artemius is lytic and does not harbor genes encoding toxins and antibiotic resistance determinants.

1. Introduction

Vibrio alginolyticus is one of the most important pathogenic bacteria of marine organisms and has become a serious threat to the aquaculture industry [1]. It has been reported to be a causative agent of vibriosis in fish [2,3], crustaceans [4,5] and mollusks [6] all over the world. The pathogen is particularly relevant to the Chinese aquaculture industry where it has been associated with severe economic damage [7]. In European marine aquaculture, it is a common inhabitant of fish and bivalve hatcheries and it has been linked to larval mortality when poor water quality favors its growth, mostly in live feed departments [8]. The management of bacterial infections in aquaculture is mainly conducted with antibiotics, a practice that is currently under close scrutiny because of the risk of the development of antimicrobial resistance and its effect on the environment and consumers. In addition, there are cases where the use of antibiotics or disinfectants is not advisable or cannot be considered at all. An aquaculture hatchery, for example, is a delicate environment where bacterial communities are in a dynamic balance and play a crucial role in the survival, development and performance of the developing fish or bivalve larvae [9]. Therefore, the destruction of this microbial balance may lead to dysbiosis with detrimental consequences [10]. In this regard, phage therapy may offer a unique advantage in the modulation of the microbial population because phages are highly host-specific and can selectively target pathogenic bacteria without affecting the beneficial ones. Phage therapy has become very attractive in aquaculture; this is evident from the increasing research efforts observed in the past years [11,12,13].
Here, we describe the isolation and characterization of a novel phage infecting Vibrio alginolyticus isolated from the live feeds of a fish hatchery. This work is part of a broader effort of our research team to isolate and characterize lytic bacteriophages against pathogenic vibrios relevant to marine aquaculture [8,14,15,16] toward the creation of potent phage cocktails that can be used as an alternative treatment for vibriosis.

2. Materials and Methods

2.1. Bacterial Strains

The bacterial strain used as a host (HCMR-2 Art1) was isolated from the water of a live feed Artemia salina culture tank of the Institute of Marine Biology, Biotechnology and Aquaculture of the Hellenic Centre for Marine Research in Heraklion, Crete (IMBBC-HCMR). HCMR-2 Art1 was identified as Vibrio alginolyticus using molecular methods (PCR amplifying 16S rRNA, MreB, toxR) [17,18,19] and whole genome sequencing. In addition to the original host, several clinical and environmental strains of the same species as well as other species of the genus from HCMR’s collection were used in order to assess the host range of the phage (Table 1). All bacterial strains were maintained in microbeads (MicroBank) at −80 °C; when working, they were grown in a lysogeny broth (LB) (23.4 gL−1 NaCl, 24.7 gL−1 MgSO4 × 7H2O, 1.5 gL−1 KCl, 1.43 gL−1 CaCl2 × 2H2O, 1% tryptone and 0.5% yeast extract) at 25 °C. All strains had been previously identified to a species level using sequencing.

2.2. Phage Isolation

The phage was isolated from the water of a live feed Artemia salina culture tank of IMBBC-HCMR following a standard enrichment procedure [20]. Briefly, 1 L of the sample water was enriched with 100 mL of 10 × LB medium, inoculated with 10 mL of an overnight culture of the host bacterium (Vibrio alginolyticus strain HCMR-2 Art1) and incubated at 25 °C for 24 h. Following filtration through 0.22 µm filters, 10 µL aliquots were plated by a standard double-layer agar method and incubated overnight at 25 °C to detect and enumerate the plaque forming units (PFU). The clearest plaque formed was picked and further purified by re-plating five times to ensure clonal phage stocks. The purified phage was named Vibrio phage Artemius and propagated to reach a titer of 1010 PFU mL−1 and stored at 4 °C.

2.3. Phage Morphology

The phage morphology was observed by transmission electron microscopy (TEM). An aliquot of the phage suspension with a titer of ~1010 PFU mL−1 was negatively stained with 4% w/v uranyl acetate. Vibrio phage Artemius was observed using a JOEL transmission electron microscope operated at 60 kV at the Electron Microscopy Lab of the University of Crete.

2.4. Host Range

The host range of Vibrio phage Artemius was evaluated by a spot test using the bacterial stains shown in Table 1. A total of 1 mL of fresh culture of each bacterial strain (107 CFU mL−1) was mixed with 3 mL of warm top agar (culture medium with 0.6% agar) and poured onto bottom LB 1/2 agar plates. When the top agar hardened, spots of 10 µL phage were made and examined after 24 h incubation for lysis.

2.5. In Vitro Efficacy

The efficacy of Vibrio phage Artemius on the host strain was assessed by an in vitro lysis assay. Sterile 96-well plates were used and loaded with 180 μL of a freshly prepared culture of the host bacterium. The plate was placed in a TECAN microplate reader (Infinite PRO 200) equipped with a temperature control and incubated at 25 °C with orbital shaking. When the bacterial culture was at the exponential phase (~107 CFU mL−1), it was infected with 20 μL of the phage preparation at 4 different MOIs (0.1, 1, 10 and 100 in triplicate). Three wells were not infected and were used as the control. The growth curve of the cultures was monitored in real-time over a minimum of 18 h and OD600 measurements were recorded every 10 min.

2.6. Phage Stability at Different Temperatures and pH

The thermal stability of the novel phage was examined by exposing 500 μL aliquots of the phage (titer: 107 PFU mL−1) to different temperatures (4, 25, 35, 45, 65, 70 and 80 °C). The aliquots were incubated at each temperature for 1 h and then rested at room temperature (RT) for 10 min. Each aliquot was then serially diluted and spotted (10 μL/spot) on a host bacterial lawn (HCMR-2 Art1). After a 24 h incubation of the agar plates, the phage titer was determined for each temperature and 4 °C was used as the control.
The sensitivity of the phage in acidic and alkaline environments was determined by exposing the phage to different pH values according to Pan et al. [21]. A scale of pH ranging from 1 to 10 was made by adding NaOH or HCl to the LB broth agar. The phage was suspended at each different pH value with a final titer of 107 PFU mL−1 and stored at 4 °C. After 2 h, and then 10 min of rest at RT, the aliquots were serially diluted and spotted onto the host bacterial lawn. After a 24 h incubation of the agar plates, the phage titer was determined for each pH value and pH = 7 was used as the control. Both the thermal and pH stability was tested in triplicate.

2.7. Adsorption Time and One-Step Growth

For the adsorption time and one-step growth, we followed the protocol described in Misol et al. [16] according to Clokie et al. [22] with a few modifications in triplicate. For the adsorption time, a fresh host culture at the exponential phase (~107 CFU mL−1) was infected with the phage at a multiplicity of infection (MOI) of 0.01. The same amount of phage was added to a tube containing only LB and served as the control. Immediately after the infection, 500 μL of the infected culture was placed in tubes containing chloroform and stored in ice (4 °C). The same procedure was repeated for 20 min with a 2 min interval. When all time points were collected, the aliquots were centrifuged at 13,000 rpm for 3 min. They were then serially diluted and spotted onto the host bacterial lawn on LB ½ agar plates. The phage titer was determined after 24 h of incubation of the agar plates at 25 °C.
For the one-step growth, 1 mL of freshly grown culture at the exponential phase (~107 CFU mL−1) was centrifuged at 13,000 rpm for 3 min. The supernatant was discarded and the pellet was washed with saline. After two repetitions of this step, the culture was finally resuspended in LB, infected with the phage at a MOI of 0.01 and rested for 15 min at RT. Following centrifugation at 13,000 rpm for 3 min, the supernatant was discarded and the pellet was dissolved in 1 mL saline. The infected culture was then transferred to a new tube containing 25 mL LB and the aliquots were immediately removed and placed in empty Eppendorf vials. The aliquots were then centrifuged at 13,000 rpm for 3 min and the supernatant was transferred to a new 2 mL Eppendorf vial, serially diluted and spotted onto the host bacterial lawn on LB ½ agar plates. This procedure was repeated every 10 min for 120 min. The phage titer was determined after 24 h of incubation of the agar plates at 25 °C.

2.8. Genomic Analysis

The DNA extraction of Vibrio phage Artemius was performed using the phenol–chloroform method according to Higuera et al. [20]. The extracted DNA was visualized for quality via 1% agarose gel electrophoresis at 80 kV for 1 h with a 50 kbp ladder. Milli-Q® Reference Water (Merck KGaA, Darmstadt, Germany) was used as a negative control. At least 5 μg of high-purity bacterial DNA, the quality of which was tested using a BioAnalyzer (Bio-Rad, Chicago, CA, USA), was used to generate a paired-end 300 PE genomic library with an optimized size selection using magnetic bead purifications based on the standard Illumina protocol and by using a Nextera XT Library Construction Kit (Illumina, San Diego, CA, USA). To calculate the average size of the library, we used a Tapestastion 4200 system (Agilent, New York, CA, USA). The insert size was estimated as the average size of the library minus the Illumina adapter size and was found to be 834 bp. The sequencing was performed using an Illumina NovaSeq 6000 sequencing platform (Illumina, San Diego, CA, USA) according to the manufacturer’s protocol at Life Sequencing (Valencia, Spain), which allowed us to obtain at least 1 million paired reads sequences. Possibly contaminated, primer, N-terminus and 3′-, 5′-low quality reads were trimmed off (threshold: 0.05). The raw reads were quality inspected and were assembled in Geneious Prime using the Geneious assembler. RASTk and Glimmer were used as the gene predictors for the structural annotation through the PATRIC webserver [23] and Galaxy webserver [24]. The presence of a phage start codon (ATG/GTG or TTG) was manually validated and a Shine–Dalgarno feature was added to all features that had a detectable match. The functional annotation of the predicted proteins was conducted manually using the NCBI Basic Local Alignment Search Tool (BLAST) [25] adjusted at a non-redundant (nr) protein database as well as the Gene Ontology [26], InterPro [27] and TΜHMM 2.0 [28,29] databases. The NCBI Conserved Domain Database (NCBI CDD) [30] was used to detect the conserved regions within the predicted proteins. Genes associated with integration, virulence and antibiotic resistance in the phage genome were searched for using the INTEGRALL Database webserver [31] and Virulence Factor Database (VFDB) [32] as well as the VirulenceFinder and ResFinder webservers [33]. The phage genome was scanned for the presence of tRNAs using the ARAGORN tool [34] through the Galaxy server [24]. The phage lifestyle was also assessed using the Phage AI platform (PhageAI S.A. Artificial Intelligence & Bioinformatics for Phage Research, https://phage.ai/, accessed on 1 May 2022). The genome of Vibrio phage Artemius with annotated predicted ORFs was then visualized in a circular representation with Geneious software (Geneious v9.1, Biomatters, Auckland, New Zealand, http://www.geneious.com, accessed on 1 May 2022) and CGview.

2.9. Phylogeny

A ViPTree was used to build a viral proteomic tree by comparing the proteome of Vibrio phage Artemius with 4982 dsDNA phage proteomes [35]. The phylogenetic relationship between Artemius and other similar Vibrio phages was determined using Molecular Evolutionary Genetics Analysis (MEGA X) software (Table 2) [36]. A total of 14 large terminase subunits of the described Vibrio phages were downloaded from the NCBI VIRUS database and were aligned with the large terminase subunit of Vibrio phage Artemius using the MUSCLE algorithm [37]. The gaps in the amino acid sequence alignments were trimmed. A maximum likelihood phylogenetic tree was constructed using the TN93 model [38] with a bootstrap test = 1000. The Interactive Tree of Life webserver [39] was used to visualize the constructed tree.

3. Results

The phage morphology of the virions as observed with TEM classified the novel phage to the Siphoviridae family. The head was 48.7 ± 0.9 nm wide and the tail was 107.0 ± 2.9 nm long (average ± S.E, n = 15) (Figure 1).

3.1. In Vitro Efficacy

Vibrio phage Artemius was able to lyse the host bacterial population in vitro using approximate MOIs ranging between 0.1 and 100 (Figure 2), with the highest MOI yielding a more rapid decrease in the bacterial titer. The bacterial titer remained low (close to the detection limit) over a period of approximately 10 h.

3.2. Host Range

According to the host range assay, Vibrio phage Artemius was able to only lyse strains that belonged to V. alginolyticus and none of the other bacteria that were tested in the study (Table 3).

3.3. Thermal and pH Stability

We exposed Artemius to different temperatures ranging from 25 °C to 80 °C, with T = 4 °C serving as the control (Figure 3a), to investigate its thermal stability. We found that Artemius was stable up to 65 °C with no significant difference in the titer compared with the control. A reduction of the titer (one-way ANOVA, p < 0.05) was observed when the phage was exposed to 70 °C and 80 °C. The phage was not inactivated at high temperatures. We then exposed the phage to acidic and alkaline pH values ranging from pH = 1 to pH = 10 and compared the titer of the phage with the titer at pH = 7, which served as the control (Figure 3b). The phage was inactivated when exposed to pH = 1 and pH = 2 (one-way ANOVA, p < 0.001), but remained stable from pH = 3 to pH = 10 with no statistically significant difference compared with the control.

3.4. Adsorption Time and One-Step Growth

The adsorption time assay showed that 90% of the Vibrio phage Artemius virions required 10 min to bind to the bacterial host (Figure 4a). Artemius had a latent period of 20 min and a rise period between 20 and 60 min. The plateau phase, when no more phages were released from the host cell, was reached at 60 min. The burst size was calculated to be 779 PFU cell−1.

3.5. Genomic Analysis

The genome size of Vibrio phage Artemius was 43,349 bp with a GC content of 43.61%. The genome arrangement was dense, as suggested by the 1.31 genes per kbp. The Rapid Annotation using Subsystem Technology (RASTk) server and Glimmer.hmm 3.0 revealed that 57 ORFs were present in the genome. Each individual ORF was manually inspected in order to validate the presence of start codons (ATG/GTG or TTG). There was no presence of tRNA in the genome, as shown by ARAGORN. A total of 50 ORFs used a start codon of ATG, 6 ORFs used GTG and 1 used TTG. A search of the NCBI nr database showed that 45 ORFs (78.95%) had significant hits (expected value ≤ 10−3) with an average similarity of 52.67%. Overall, 19 ORFs (10.83%) were assigned a function based on the protein homology. No genes associated with integration, virulence or antibiotic resistance were detected. The genome of Vibrio phage Artemius was not modularly organized (Figure 5). However, several gene encoding proteins required for phage assembly (ORF 2, ORF 5, ORF 6, ORF 8, ORF 10, ORF 13, ORF 16, ORF 17 and ORF 22) were arranged in subclusters as well as a few genes encoding for DNA replication and nucleotide metabolism proteins (ORF 25, ORF 33, ORF 51 and ORF 54). The genes that were functionally annotated are shown in Table 4.
The proteins required for phage morphogenesis included major capsid protein (ORF 8), tail-length tape measure protein (ORF 17), tail tube protein (ORF 13), tail tubular protein (ORF 22) and neck protein (ORF 16). In addition, the small terminase subunit (ORF 2), large terminase subunit (ORF 5), stopper protein (ORF 10) and phage portal protein (ORF 6), which are involved in DNA packaging for tailed phages, were identified. Proteins for DNA replication, recombination and repair were also detected: these were DNA polymerase (ORF 23), HNH endonuclease (ORF 51), DNA polymerases, DNA helicase (ORF 33, DNA primase (ORF 25) and other regulatory elements (ORF 54). Finally, several transmembrane proteins were detected (ORF 21, 35, 37 and 39) (Figure 6).

3.6. Phylogenetic Analysis

A wide genome proteomic tree analysis validated that Vibrio phage Artemius belonged to the Siphoviridae taxonomic family (Figure 7). In addition, it was predicted to infect hosts from the Gammaproteobacteria class, which includes the Vibrionaceae family. A closer view of the tree focusing on the nearest relatives (Figure 8) showed that Vibrio phage Artemius clustered together with Shewanella phage 1/44 (Siphoviridae) and Escherichia phage PTXU04 (Podoviridae).
Phylogeny using the large terminase subunits of the Vibrio alginolyticus phages (Figure 9) showed that Vibrio phage Artemius shared a common ancestor with Vibrio phage phi-St2 and Vibrio phage phi-Grn1. However, the phages were rather distant, resulting in low bootstrap values (<75%); thus, the node was not well-supported. A large terminase subunit gene was used for the phylogeny because it is considered to be a signature, well-conserved gene among the phages and is a strong molecular motor associated with phage packaging [40].

4. Discussion

Vibrio alginolyticus has been used frequently as a target for novel phage isolation. Until today, more than 40 different phages have been isolated against V. alginolyticus, which is suggestive of the importance of this opportunistic pathogen for the aquaculture industry [41]. Several reviews have been published on the topic of phage therapy in aquaculture and the existing challenges [11,42], but also specifically on the interaction of phages with vibrios [13].
Vibrio phage Artemius is a novel phage, as indicated by the BLAST search in the NCBI nr database where the closest phage at the nucleotide level was Vibrio phage 2.044.O._10N.261.51.B8 (MG592661.1) with a 79.6% similarity over a 2% query cover. A phylogenetic analysis of Artemius verified the novelty of this phage. According to the phylogeny obtained with the terminase large subunit using phages infecting the same host (V. alginolyticus), Vibrio phage Artemius was closer to Vibrio phage phi-St2 and Vibrio phage phi-Grn1, which belong to the genus of Schizotequatrovirus [43]. According to the ViPTree analysis, which uses the whole proteome of all described phages, Vibrio phage Artemius was closer to siphovirus 1/44 infecting Shewanella sp. Interestingly, this phage was isolated from iced water in the Baltic Sea [44]. The diversity of the phages and the mosaicism of their genomic arrangement make a phylogenetic analysis using traditional phylogenetic trees difficult and occasionally vague, especially when phage species are under-represented [45,46].
Two of the most crucial characteristics of a phage to be considered for therapeutic purposes are the inability to integrate in the host genome and the lack of unwanted genes such as toxins and antibiotic resistance determinants that could be transferred to its host through lysogenic conversion or recombination [47]. A genomic analysis of Vibrio phage Artemius showed that it was a lytic phage because no integrase or temperateness-related genes were identified. This was further corroborated by all the bioinformatic tools used such as INTEGRALL and Phage AI, which also indicated a lytic life cycle. Furthermore, no toxins or AMR-related genes were found, indicating that Vibrio phage Artemius could be considered to be safe for use in phage therapy applications. Gene-encoding predicted structural proteins were identified in the genome of the novel phage, including capsid protein, tail-length tape measure protein, tail tube protein, tail tubular protein and neck protein. An interesting gene found in the Vibrio phage Artemius genome was ORF29 encoding a PD-(D/E) XS nuclease. A search in the Conserved Domain Database (CDD) of NCBI revealed that this particular protein belongs to the superfamily of Cas4 nucleases. It has been suggested that phage-derived Cas4 leads to the acquisition of more host-derived spacers in type II-C CRISPR-Cas systems, enhancing the survival of the host and ultimately the bacteriophage whilst in a carrier state [48]. If this is the case, a further investigation is required in order to better understand the role of this particular gene, which will improve our knowledge of the dynamics of phage/host interactions.
High temperatures usually have a devastating effect on phage integrity, causing tail aggregation, detachment of the phage head and denaturation of the nucleic acid [49]. Interestingly, Vibrio phage Artemius withstood temperatures as high as 80 °C. It has been suggested that heat resistance may be the result of mutations or strong protein interactions [50] and it is usually a unique characteristic of the phage particle [51]. Although it is not common, phages that are resistant to high temperatures have been reported before [52], many of which were members of the of Siphoviridae family [53].
The host range of Vibrio phage Artemius was limited to strains of Vibrio alginolyticus as it was not able to infect the other congeneric species; thus, it was a species-specific phage. Broad host-range phages are considered to be ideal for phage therapy, especially in aquaculture where the diversity of the pathogenic strains and species is wide. Phages with a broad host range have been reported against Vibrio alginolyticus [14,41]. In terms of efficacy, Vibrio phage Artemius could efficiently prevent the growth of its host even at low MOIs, suggesting that it could be practically used in aquaculture where considerable quantities are required to treat large volumes of water. This was further corroborated by the large burst size of the novel phage.
In conclusion, Vibrio phage Artemius is a potential new species of the Siphoviridae family that infects V. alginolyticus and, based on its biological and genomic characteristics, could be considered for efficient and safe phage therapy applications.

Author Contributions

Writing and formal analysis, S.D., L.P., M.I.T., C.K., D.S., E.F., P.K.; writing—original draft preparation, S.D. and L.P.; funding acquisition, E.F. and P.K.; writing—review and editing, S.D. and P.K.; All authors have read and agreed to the published version of the manuscript.

Funding

MIS 5010932: OP Fisheries and Maritime 2014–2020 Innovative measures, Innovation 2019.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The genome sequence of phage Vibrio phage Artemius is available in GenBank under accession number ON366409 and under Biosample and Bioproject accession numbers SAMN29862054 and PRJNA860748 respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kong, W.; Huang, L.; Su, Y.; Qin, Y.; Ma, Y.; Xu, X.; Lin, M.; Zheng, J.; Yan, Q. Investigation of Possible Molecular Mechanisms Underlying the Regulation of Adhesion in Vibrio Alginolyticus with Comparative Transcriptome Analysis. Antonie Leeuwenhoek Int. J. Gen. Mol. Microbiol. 2015, 107, 1197–1206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Balebona, M.C.; Andreu, M.J.; Bordas, M.A.; Zorrilla, I.; Moriñigo, M.A.; Borrego, J.J.; Morinigo, M.A.; Borrego, J.J. Pathogenicity of Vibrio Alginolyticus for Cultured Gilt-Head Sea Bream (Sparus aurata L.). Appl. Environ. Microbiol. 1998, 64, 4269–4275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rigos, G.; Katharios, P. Pathological Obstacles of Newly-Introduced Fish Species in Mediterranean Mariculture: A Review. Rev. Fish Biol. Fish. 2010, 20, 47–70. [Google Scholar] [CrossRef]
  4. Liu, C.H.; Cheng, W.; Hsu, J.P.; Chen, J.C. Vibrio Alginolyticus Infection in the White Shrimp Litopenaeus Vannamei Confirmed by Polymerase Chain Reaction and 16S RDNA Sequencing. Dis. Aquat. Org. 2004, 61, 169–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Selvin, J.; Lipton, A.P. Vibrio Alginolyticus Associated with White Spot Disease of Penaeus Monodon. Dis. Aquat. Org. 2003, 57, 147–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Luna-González, A.; Maeda-Martínez, A.N.; Sainz, J.C.; Ascencio-Valle, F. Comparative Susceptibility of Veliger Larvae of Four Bivalve Mollusks to a Vibrio Alginolyticus Strain. Dis. Aquat. Org. 2002, 49, 221–226. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, B.; Zhai, S.; Zhang, F.; Wang, H.; Ren, L.; Li, Y.; Li, Q.; Liu, S. Genome-Wide Association Study toward Efficient Selection Breeding of Resistance to Vibrio Alginolyticus in Pacific Oyster, Crassostrea Gigas. Aquaculture 2022, 548, 737592. [Google Scholar] [CrossRef]
  8. Katharios, P.; Kalatzis, P.G.; Kokkari, C.; Sarropoulou, E.; Middelboe, M. Isolation and Characterization of a N4-like Lytic Bacteriophage Infecting Vibrio splendidus, a Pathogen of Fish and Bivalves. PLoS ONE 2017, 12, e0190083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Infante-Villamil, S.; Huerlimann, R.; Jerry, D.R. Microbiome Diversity and Dysbiosis in Aquaculture. Rev. Aquac. 2021, 13, 1077–1096. [Google Scholar] [CrossRef]
  10. Dittmann, K.K.; Rasmussen, B.B.; Melchiorsen, J.; Sonnenschein, E.C.; Gram, L.; Bentzon-Tilia, M. Changes in the Microbiome of Mariculture Feed Organisms after Treatment with a Potentially Probiotic Strain of Phaeobacter Inhibens. Appl. Environ. Microbiol. 2020, 86, e00499-20. [Google Scholar] [CrossRef] [PubMed]
  11. Culot, A.; Grosset, N.; Gautier, M. Overcoming the Challenges of Phage Therapy for Industrial Aquaculture: A Review. Aquaculture 2019, 513, 734423. [Google Scholar] [CrossRef] [Green Version]
  12. Kowalska, J.D.; Kazimierczak, J.; Sowińska, P.M.; Wójcik, E.A.; Siwicki, A.K.; Dastych, J. Growing Trend of Fighting Infections in Aquaculture Environment—Opportunities and Challenges of Phage Therapy. Antibiotics 2020, 9, 301. [Google Scholar] [CrossRef] [PubMed]
  13. Kalatzis, P.; Castillo, D.; Katharios, P.; Middelboe, M. Bacteriophage Interactions with Marine Pathogenic Vibrios: Implications for Phage Therapy. Antibiotics 2018, 7, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kalatzis, P.G.; Bastías, R.; Kokkari, C.; Katharios, P. Isolation and Characterization of Two Lytic Bacteriophages, φ St2 and φ Grn1; Phage Therapy Application for Biological Control of Vibrio Alginolyticus in Aquaculture Live Feeds. PLoS ONE 2016, 11, e0151101. [Google Scholar] [CrossRef] [Green Version]
  15. Kokkari, C.; Sarropoulou, E.; Bastias, R.; Mandalakis, M.; Katharios, P. Isolation and Characterization of a Novel Bacteriophage Infecting Vibrio Alginolyticus. Arch. Microbiol. 2018, 200, 707–718. [Google Scholar] [CrossRef] [PubMed]
  16. Misol, G.N.; Kokkari, C.; Katharios, P. Biological and Genomic Characterization of a Novel Jumbo Bacteriophage, Vb_VhaM_pir03 with Broad Host Lytic Activity against Vibrio Harveyi. Pathogens 2020, 9, 1051. [Google Scholar] [CrossRef] [PubMed]
  17. Weisburg, W.G.; Barns, S.M.; Pelletier, D.A.; Lane, D.J. 16S Ribosomal DNA Amplification for Phylogenetic Study. J. Bacteriol. 1991, 173, 697–703. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Conejero, M.J.U.; Hedreyda, C.T. Isolation of Partial ToxR Gene of Vibrio Harveyi and Design of ToxR-Targeted PCR Primers for Species Detection. J. Appl. Microbiol. 2003, 95, 602–611. [Google Scholar] [CrossRef]
  19. Mougin, J.; Roquigny, R.; Travers, M.A.; Grard, T.; Bonnin-Jusserand, M.; le Bris, C. Development of a MreB-Targeted Real-Time PCR Method for the Quantitative Detection of Vibrio Harveyi in Seawater and Biofilm from Aquaculture Systems. Aquaculture 2020, 525, 735337. [Google Scholar] [CrossRef]
  20. Higuera, G.; Bastías, R.; Tsertsvadze, G.; Romero, J.; Espejo, R.T. Recently Discovered Vibrio Anguillarum Phages Can Protect against Experimentally Induced Vibriosis in Atlantic Salmon, Salmo Salar. Aquaculture 2013, 392, 128–133. [Google Scholar] [CrossRef]
  21. Pan, L.; Li, D.; Sun, Z.; Lin, W.; Hong, B.; Qin, W.; Xu, L.; Liu, W.; Zhou, Q.; Wang, F.; et al. First Characterization of a Hafnia Phage Reveals Extraordinarily Large Burst Size and Unusual Plaque Polymorphism. Front. Microbiol. 2022, 12, 754331. [Google Scholar] [CrossRef]
  22. Clokie, M.R.J.; Kropinski, A.M. Bacteriophages; Humana: Totowa, NJ, USA, 2009; Volume 1, ISBN 9781627032384. [Google Scholar]
  23. Davis, J.J.; Wattam, A.R.; Aziz, R.K.; Brettin, T.; Butler, R.; Butler, R.M.; Chlenski, P.; Conrad, N.; Dickerman, A.; Dietrich, E.M.; et al. The PATRIC Bioinformatics Resource Center: Expanding Data and Analysis Capabilities. Nucleic Acids Res. 2020, 48, D606–D612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Ramsey, J.; Rasche, H.; Maughmer, C.; Criscione, A.; Mijalis, E.; Liu, M.; Hu, J.C.; Young, R.; Gill, J.J. Galaxy and Apollo as a Biologist-Friendly Interface for High-Quality Cooperative Phage Genome Annotation. PLoS Comput. Biol. 2020, 16, e1008214. [Google Scholar] [CrossRef] [PubMed]
  25. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic Local Alignment Search Tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  26. Harris, M.A.; Clark, J.; Ireland, A.; Lomax, J.; Ashburner, M.; Foulger, R.; Eilbeck, K.; Lewis, S.; Marshall, B.; Mungall, C.; et al. The Gene Oncology (GO) Database and Informatics Resource. Nucleic Acids Res. 2004, 32, 258–261. [Google Scholar] [CrossRef] [Green Version]
  27. Mitchell, A.L.; Attwood, T.K.; Babbitt, P.C.; Blum, M.; Bork, P.; Bridge, A.; Brown, S.D.; Chang, H.Y.; El-Gebali, S.; Fraser, M.I.; et al. InterPro in 2019: Improving Coverage, Classification and Access to Protein Sequence Annotations. Nucleic Acids Res. 2019, 47, D351–D360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Krogh, A.; Larsson, B.; Von Heijne, G.; Sonnhammer, E.L.L. Predicting Transmembrane Protein Topology with a Hidden Markov Model: Application to Complete Genomes. J. Mol. Biol. 2001, 305, 567–580. [Google Scholar] [CrossRef] [Green Version]
  29. Sonnhammer, E.L.L.; Krogh, A. A Hidden Markov Model for Predicting Transmembrane Helices in Protein Sequence. Proc. Int. Conf. Intell. Syst. Mol. Biol. 1998, 6, 175–182. [Google Scholar]
  30. Lu, S.; Wang, J.; Chitsaz, F.; Derbyshire, M.K.; Geer, R.C.; Gonzales, N.R.; Gwadz, M.; Hurwitz, D.I.; Marchler, G.H.; Song, J.S.; et al. CDD/SPARCLE: The Conserved Domain Database in 2020. Nucleic Acids Res. 2020, 48, D265–D268. [Google Scholar] [CrossRef] [Green Version]
  31. Moura, A.; Soares, M.; Pereira, C.; Leitão, N.; Henriques, I.; Correia, A. INTEGRALL: A Database and Search Engine for Integrons, Integrases and Gene Cassettes. Bioinformatics 2009, 25, 1096–1098. [Google Scholar] [CrossRef] [Green Version]
  32. Chen, L.; Yang, J.; Yu, J.; Yao, Z.; Sun, L.; Shen, Y.; Jin, Q. VFDB: A Reference Database for Bacterial Virulence Factors. Nucleic Acids Res. 2005, 33, 325–328. [Google Scholar] [CrossRef] [Green Version]
  33. Kleinheinz, K.A.; Joensen, K.G.; Larsen, M.V. Applying the ResFinder and VirulenceFinder Web-Services for Easy Identification of Acquired Antibiotic Resistance and E. Coli Virulence Genes in Bacteriophage and Prophage Nucleotide Sequences. Bacteriophage 2014, 4, e27943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Laslett, D.; Canback, B. ARAGORN, a Program to Detect TRNA Genes and TmRNA Genes in Nucleotide Sequences. Nucleic Acids Res. 2004, 32, 11–16. [Google Scholar] [CrossRef]
  35. Nishimura, Y.; Yoshida, T.; Kuronishi, M.; Uehara, H.; Ogata, H.; Goto, S. ViPTree: The Viral Proteomic Tree Server. Bioinformatics 2017, 33, 2379–2380. [Google Scholar] [CrossRef] [PubMed]
  36. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef] [PubMed]
  37. Edgar, R.C. MUSCLE: A Multiple Sequence Alignment Method with Reduced Time and Space Complexity. BMC Bioinform. 2004, 5, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Jones, D.T.; Taylor, W.R.; Thornton, J.M. The Rapid Generation of Mutation Data Matrices. Bioinformatics 1992, 8, 275–282. [Google Scholar] [CrossRef] [PubMed]
  39. Letunic, I.; Bork, P. Interactive Tree of Life (ITOL) v5: An Online Tool for Phylogenetic Tree Display and Annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef] [PubMed]
  40. Mitchell, M.S.; Matsuzaki, S.; Imai, S.; Rao, V.B. Sequence Analysis of Bacteriophage T4 DNA Packaging/Terminase Genes 16 and 17 Reveals a Common ATPase Center in the Large Subunit of Viral Terminases. Nucleic Acids Res. 2002, 30, 4009–4021. [Google Scholar] [CrossRef] [Green Version]
  41. Li, F.; Xing, S.; Fu, K.; Zhao, S.; Liu, J.; Tong, Y.; Zhou, L. Genomic and Biological Characterization of the Vibrio Alginolyticus-Infecting “Podoviridae” Bacteriophage, VB_ValP_IME271. Virus Genes 2019, 55, 218–226. [Google Scholar] [CrossRef]
  42. Ramos-Vivas, J.; Superio, J.; Galindo-Villegas, J.; Acosta, F. Phage Therapy as a Focused Management Strategy in Aquaculture. Int. J. Mol. Sci. 2021, 22, 10436. [Google Scholar] [CrossRef] [PubMed]
  43. Skliros, D.; Kalatzis, P.G.; Katharios, P.; Flemetakis, E. Comparative Functional Genomic Analysis of Two Vibrio Phages Reveals Complex Metabolic Interactions with the Host Cell. Front. Microbiol. 2016, 7, 1807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Luhtanen, A.M.; Eronen-Rasimus, E.; Kaartokallio, H.; Rintala, J.M.; Autio, R.; Roine, E. Isolation and Characterization of Phage-Host Systems from the Baltic Sea Ice. Extremophiles 2014, 18, 121–130. [Google Scholar] [CrossRef] [PubMed]
  45. Rohwer, F.; Edwards, R. The Phage Proteomic Tree: A Genome-Based Taxonomy for Phage. J. Bacteriol. 2002, 184, 4529–4535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Dion, M.B.; Oechslin, F.; Moineau, S. Phage Diversity, Genomics and Phylogeny. Nat. Rev. Microbiol. 2020, 18, 125–138. [Google Scholar] [CrossRef] [PubMed]
  47. Hyman, P. Phages for Phage Therapy: Isolation, Characterization, and Host Range Breadth. Pharmaceuticals 2019, 12, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Hooton, S.P.T.; Connerton, I.F. Campylobacter Jejuni Acquire New Host-Derived CRISPR Spacers When in Association with Bacteriophages Harboring a CRISPR-like Cas4 Protein. Front. Microbiol. 2015, 6, 744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Lasobras, J.; Muniesa, M.; Frías, J.; Lucena, F.; Jofre, J. Relationship between the Morphology of Bacteriophages and Their Persistence in the Environment. Water Sci. Technol. 1997, 35, 129–132. [Google Scholar] [CrossRef]
  50. Kadowaki, K.-I.; Shibata, T.; Takeuchi, K.; Himeno, M.; Sakai, H.; Komano, T. Identification of a Temperature-Resistant Bacteriophage X174 Mutant. J. Gen. Virol. 1987, 68, 2443–2447. [Google Scholar] [CrossRef]
  51. Ahmadi, H.; Radford, D.; Kropinski, A.M.; Lim, L.-T.; Balamurugan, S. Thermal-Stability and Reconstitution Ability of Listeria Phages P100 and A511. Front. Microbiol. 2017, 8, 2375. [Google Scholar] [CrossRef] [PubMed]
  52. Jończyk, E.; Kłak, M.; Międzybrodzki, R.; Górski, A. The Influence of External Factors on Bacteriophages-Review. Folia Microbiol. 2011, 56, 191–200. [Google Scholar] [CrossRef] [Green Version]
  53. Yu, M.X.; Slater, M.R.; Ackermann, H.W. Isolation and Characterization of Thermus Bacteriophages. Arch. Virol. 2006, 151, 663–679. [Google Scholar] [CrossRef]
Figure 1. (A) Transmission electron micrograph of Vibrio phage Artemius. (B) Details of the baseplate structure at the distal end of the tail.
Figure 1. (A) Transmission electron micrograph of Vibrio phage Artemius. (B) Details of the baseplate structure at the distal end of the tail.
Pathogens 11 00848 g001
Figure 2. In vitro cell lysis of Vibrio phage Artemius vs. host strain HCMR-2 Art1. The conditions tested were a control (HCMR-2 Art1 strain without any phage addition) and 4 different MOIs used for infecting HCMR-2 Art1 (MOI = 0.1, MOI = 1, MOI = 10 and MOI = 100). The values are means ± standard deviations of the three replicates.
Figure 2. In vitro cell lysis of Vibrio phage Artemius vs. host strain HCMR-2 Art1. The conditions tested were a control (HCMR-2 Art1 strain without any phage addition) and 4 different MOIs used for infecting HCMR-2 Art1 (MOI = 0.1, MOI = 1, MOI = 10 and MOI = 100). The values are means ± standard deviations of the three replicates.
Pathogens 11 00848 g002
Figure 3. Thermal and pH stability. Vibrio phage Artemius exposed to (a) different temperatures, with T = 4 °C used as the control, and (b) different pH values, with pH = 7 used as the control. Error bars represent the standard deviation of the mean (n = 3). Statistical significance is indicated by different letters above the bars; p < 0.05.
Figure 3. Thermal and pH stability. Vibrio phage Artemius exposed to (a) different temperatures, with T = 4 °C used as the control, and (b) different pH values, with pH = 7 used as the control. Error bars represent the standard deviation of the mean (n = 3). Statistical significance is indicated by different letters above the bars; p < 0.05.
Pathogens 11 00848 g003
Figure 4. (a) Adsorption time assay: at 10 min, 90% of phage Artemius was irreversibly adsorbed by the host bacteria. (b) One-step growth curve: latency period for Artemius was 20 min and the burst size was 779 PFU cell−1. Error bars represent the standard deviation of the mean (n = 3).
Figure 4. (a) Adsorption time assay: at 10 min, 90% of phage Artemius was irreversibly adsorbed by the host bacteria. (b) One-step growth curve: latency period for Artemius was 20 min and the burst size was 779 PFU cell−1. Error bars represent the standard deviation of the mean (n = 3).
Pathogens 11 00848 g004
Figure 5. Visual representation of the Vibrio phage Artemius genome in which the genome GC content is shown by the inner black line and the GC skew guanine over-representation by the inner purple/green line. The predicted ORFs are shown as arrows. The color of the ORFs refers to the annotated biochemical function: phage assembly proteins (brown); DNA replication, repair and recombination associated proteins (purple); auxiliary metabolic proteins (light green); and hypothetical (dark green).
Figure 5. Visual representation of the Vibrio phage Artemius genome in which the genome GC content is shown by the inner black line and the GC skew guanine over-representation by the inner purple/green line. The predicted ORFs are shown as arrows. The color of the ORFs refers to the annotated biochemical function: phage assembly proteins (brown); DNA replication, repair and recombination associated proteins (purple); auxiliary metabolic proteins (light green); and hypothetical (dark green).
Pathogens 11 00848 g005
Figure 6. (A) Number of ORFs in a forward and reverse direction. (B) Number of annotated and non-annotated ORFs. (C) Length of ORFs.
Figure 6. (A) Number of ORFs in a forward and reverse direction. (B) Number of annotated and non-annotated ORFs. (C) Length of ORFs.
Pathogens 11 00848 g006
Figure 7. Prediction of taxa and host groups for Vibrio phage Artemius according to the proteomic tree produced by ViPTree. The novel phage was determined to belong to the Siphoviridae family and infect the Gammaproteobacteria group (red star and line). The Vibrio phage Artemius (asterisk) proteome was compared with 4892 dsDNA phage proteomes. The tree was rooted using “midpoint rooting”. The branch length scale was calculated as log values. The inner and outer ring indicate the taxonomic virus family and host group, respectively.
Figure 7. Prediction of taxa and host groups for Vibrio phage Artemius according to the proteomic tree produced by ViPTree. The novel phage was determined to belong to the Siphoviridae family and infect the Gammaproteobacteria group (red star and line). The Vibrio phage Artemius (asterisk) proteome was compared with 4892 dsDNA phage proteomes. The tree was rooted using “midpoint rooting”. The branch length scale was calculated as log values. The inner and outer ring indicate the taxonomic virus family and host group, respectively.
Pathogens 11 00848 g007
Figure 8. A closer view of the proteomic tree analysis by ViPTree focusing on the closer-related phages of Vibrio phage Artemius.
Figure 8. A closer view of the proteomic tree analysis by ViPTree focusing on the closer-related phages of Vibrio phage Artemius.
Pathogens 11 00848 g008
Figure 9. Phylogenetic tree of Vibrio phage Artemius with other Vibrio alginolyticus phages. The nucleotide sequences of large terminase subunits of phages were downloaded from the NCBI database and aligned using MUSCLE and a maximum likelihood (bootstrap = 1000). A rooted phylogenetic tree was constructed using MEGA X. Gene sequences with less than 20% similarity with the obtained sequence were arbitrarily set as an outgroup. The bootstrap support value is denoted in each branch as a percentage.
Figure 9. Phylogenetic tree of Vibrio phage Artemius with other Vibrio alginolyticus phages. The nucleotide sequences of large terminase subunits of phages were downloaded from the NCBI database and aligned using MUSCLE and a maximum likelihood (bootstrap = 1000). A rooted phylogenetic tree was constructed using MEGA X. Gene sequences with less than 20% similarity with the obtained sequence were arbitrarily set as an outgroup. The bootstrap support value is denoted in each branch as a percentage.
Pathogens 11 00848 g009
Table 1. Bacterial strains used in the host range assay (T: strain type; h: host). “Environmental” refers to an environmental isolation source whereas “clinical” refers to bacteria that were isolated from disease cases in various fish.
Table 1. Bacterial strains used in the host range assay (T: strain type; h: host). “Environmental” refers to an environmental isolation source whereas “clinical” refers to bacteria that were isolated from disease cases in various fish.
StrainSpeciesLocalityIsolation Source
HCMR-2 Art1hV. alginolyticusGreeceEnvironmental
HCMR-1 Art1V. alginolyticusGreeceEnvironmental
HCMR-1 Art3V. alginolyticusGreeceEnvironmental
LAR41V. alginolyticusGreeceEnvironmental
LAR50V. alginolyticusGreeceEnvironmental
LAR73V. alginolyticusGreeceEnvironmental
LAR74V. alginolyticusGreeceEnvironmental
LAR76V. alginolyticusGreeceEnvironmental
Gal 019V. alginolyticusGreeceEnvironmental
LAR170V. alginolyticusGreeceClinical
Gal 074V. alginolyticusGreeceClinical
Gal 048V. alginolyticusGreeceClinical
DSMZ2171V. alginolyticusJapanCollection (T)
VhSerFrEV. harveyiGreeceClinical
DSM 19623V. harveyiMassachusettsClinical (T)
VH2V. harveyiGreeceClinical
VIB391V. campbelliiThailandClinical
LAR52V. tubiashiiGreeceEnvironmental
MAN113V. splendidusGreeceEnvironmental
LAR194V. mediterraneiGreeceEnvironmental
Table 2. Phages infecting Vibrio alginolyticus used for the phylogenetic analysis with a large terminase subunit.
Table 2. Phages infecting Vibrio alginolyticus used for the phylogenetic analysis with a large terminase subunit.
PhageFamilyAccession NumberIsolation CountryDate of
Isolation
Vb ValS X1 DemerecviridaeQKN88544.1China2020-06-15
Valm-yong1 MyoviridaeQGF21295.1China2020-03-09
Va2MyoviridaeQPB08501.1China2020-11-15
phi-Grn1MyoviridaeALP46971Greece2016-03-01
phi-St2MyoviridaeALP47351Greece2016-03-01
V039CMyoviridaeQIQ61300China2020-04-01
Vb ValS PJ32SiphoviridaeQNJ59211.1China2020-08-30
Vb ValP VA-RY-4SiphoviridaeUFD98283China2021-11-24
BUCT194SchitoviridaeUAW01163China2021-09-22
USC-1MyoviridaeQDH47503Australia2019-07-07
5 TSL-2019MyoviridaeQCW23225Australia2019-07-10
NF SiphoviridaeQGZ13228China2019-12-25
Vb ValP IME271PodoviridaeASR73874China2017-08-08
Vb ValP IME234 PodoviridaeQUE30183China2021-04-28
Table 3. Host range of the Artemius phage (h: host strain; +++: complete clearing; +: substantial turbidity throughout the cleared zone; −: no lysis).
Table 3. Host range of the Artemius phage (h: host strain; +++: complete clearing; +: substantial turbidity throughout the cleared zone; −: no lysis).
StrainSpeciesLysis
HCMR-2 Art1hV. alginolyticus+++
HCMR-1 Art1V. alginolyticus+++
HCMR-1 Art3V. alginolyticus+++
LAR41V. alginolyticus+
LAR50V. alginolyticus+
LAR73V. alginolyticus+
LAR74V. alginolyticus+
LAR76V. alginolyticus+
Gal019V. alginolyticus+
LAR170V. alginolyticus+
Gal074V. alginolyticus+
Gal 048V. alginolyticus+
DSMZ2171V. alginolyticus+
VhSerFreV. harveyi
DSM 19623V. harveyi
VH2V. harveyi
VIB391V. campbellii
LAR52V. tubiashii
MAN113V. splendidus
LAR194V. mediterranei
Table 4. Summary of Vibrio phage Artemius ORFs that were annotated with relevant information based on significant amino acid sequences and protein structural homologies (E-value ≤ 10−3).
Table 4. Summary of Vibrio phage Artemius ORFs that were annotated with relevant information based on significant amino acid sequences and protein structural homologies (E-value ≤ 10−3).
Predicted FunctionStartEndLengthDirectionNCBI BLAST Best HitE-Value
ORF1Hypothetical protein154876722ForwardAUR84443.13.00 × 10−99
ORF2Terminase small subunit8781279401ForwardAUR84444.12.00 × 10−47
ORF3Hypothetical protein14361786350Forwardno hit
ORF4Hypothetical protein17732015242Forwardno hit
ORF5Terminase large subunit200538131808ForwardAUS01868.10
ORF6Portal protein380650531247ForwardAUS01869.10
ORF7Hypothetical protein50315696665ForwardQZI87804.13.00 × 10−81
ORF8Major capsid protein578071351355ForwardPIY67122.16.00 × 10−155
ORF9Hypothetical protein72227743521ForwardPIY67121.11.00 × 10−17
ORF10Stopper protein77338077344ForwardUNY40138.11.00 × 10−11
ORF11Hypothetical protein80708882812ForwardQZI86159.15.00 × 10−80
ORF12Hypothetical protein88839248365ForwardPIY67118.15.00 × 10−10
ORF13Tail tube protein92549952698ForwardUNY40117.12.00 × 10−65
ORF14Hypothetical protein995410,355401Forwardno hit
ORF15Hypothetical protein10,46010,582122Forwardno hit
ORF16Neck protein10,58211,103521ForwardAUS01881.11.00 × 10−83
ORF17Tail-length tape measure protein11,10013,4782378ForwardAUR85334.13.00 × 10−100
ORF18Hypothetical protein13,47813,975497ForwardQGF20993.14.00 × 10−62
ORF19Hypothetical protein13,97214,487515ForwardUNY40123.14.00 × 10−40
ORF20Hypothetical protein14,48414,825341ForwardQGF20994.19.00 × 10-40
ORF21TMhelix-containing protein14,81317,5992786ForwardQGF21014.10.00 × 100
ORF22Tail tubular protein17,59620,1572561ForwardQGF21011.17.00 × 10−23
ORF23Putative DNA polymerase22,43820,1982240ReverseAUR81370.10.00 × 100
ORF24Hypothetical protein23,66322,4941169ReverseAUR81371.12.00 × 10−15
ORF25DNA primase26,25123,7592492ReverseAUR81372.10.00 × 100
ORF26Hypothetical protein26,50626,261245ReverseAUS01892.17.00 × 10−4
ORF27Hypothetical protein26,94026,506434ReverseAUR83587.12.00 × 10−72
ORF28Hypothetical protein28,09426,9461148ReverseAUR85121.15.00 × 10−114
ORF29PD-(D/E) XS nuclease29,23528,0991136ReverseAUS01895.18.00 × 10−133
ORF30Hypothetical protein30,17329,241932ReverseAUS01896.15.00 × 10−34
ORF31Hypothetical protein30,32930,604275Forwardno hit
ORF32Hypothetical protein30,60130,867266Forwardno hit
ORF33Helicase30,86432,7501886ForwardAUR81379.10.00 × 100
ORF34Hypothetical protein32,74733,091344ForwardUAW01179.13.00 × 10−16
ORF35N-acetylmuramoy-L-alanine amidase33,28133,742461ForwardAUS01901.15.50 × 10−76
ORF36TMhelix-containing protein33,74733,923176ForwardAUS01902.11.00 × 10−8
ORF37TMhelix-containing protein33,92834,230302ForwardAUS01903.16.00 × 10−3
ORF38Hypothetical protein34,25634,387131ForwardAUS01904.15.00 × 10−8
ORF39Hypothetical protein34,72334,409314ReverseAUR81384.11.00 × 10−14
ORF40Hypothetical protein35,29534,972323ReverseAUS01906.12.00 × 10−12
ORF41Hypothetical protein35,69235,276416Reverseno hit
ORF42Hypothetical protein36,46235,695767ReverseAUS01910.12.00 × 10−56
ORF43Hypothetical protein37,23136,452779Reverseno hit
ORF44Hypothetical protein37,50837,221287Reverseno hit
ORF45Hypothetical protein37,84037,544296Reverseno hit
ORF46Hypothetical protein38,38537,837548Reverseno hit
ORF47Hypothetical protein38,89638,372524ReverseAUR81476.16.00 × 10−13
ORF48Hypothetical protein39,32238,915407ReverseQZI87821.18.00 × 10−35
ORF49Hypothetical protein40,24539,319926ReverseAUR81389.11.00 × 10−21
ORF50Hypothetical protein40,62740,884257Forwardno hit
ORF51HNH nuclease40,87741,305428ForwardAUR81392.16.00 × 10−48
ORF52Hypothetical protein41,76641,969203ForwardQDP60675.11.00 × 10−4
ORF53Hypothetical protein41,97742,300323ForwardQZI91906.16.00 × 10−25
ORF54DNA-binding domain protein42,29742,458161ForwardAUR81395.16.00 × 10−14
ORF55Hypothetical protein42,46142,838377ForwardYP_009103702.16.00 × 10−27
ORF56Hypothetical protein42,83542,954119ForwardYP_007675934.18.00 × 10−6
ORF57Hypothetical protein42,96743,173206ForwardWP_081230226.17.00 × 10−26
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Droubogiannis, S.; Pavlidi, L.; Tsertou, M.I.; Kokkari, C.; Skliros, D.; Flemetakis, E.; Katharios, P. Vibrio Phage Artemius, a Novel Phage Infecting Vibrio alginolyticus. Pathogens 2022, 11, 848. https://doi.org/10.3390/pathogens11080848

AMA Style

Droubogiannis S, Pavlidi L, Tsertou MI, Kokkari C, Skliros D, Flemetakis E, Katharios P. Vibrio Phage Artemius, a Novel Phage Infecting Vibrio alginolyticus. Pathogens. 2022; 11(8):848. https://doi.org/10.3390/pathogens11080848

Chicago/Turabian Style

Droubogiannis, Stavros, Lydia Pavlidi, Maria Ioanna Tsertou, Constantina Kokkari, Dimitrios Skliros, Emmanouil Flemetakis, and Pantelis Katharios. 2022. "Vibrio Phage Artemius, a Novel Phage Infecting Vibrio alginolyticus" Pathogens 11, no. 8: 848. https://doi.org/10.3390/pathogens11080848

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop