Next Article in Journal
An Efficient Distributed Compressed Sensing Algorithm for Decentralized Sensor Network
Next Article in Special Issue
Supramolecular Recognition of Escherichia coli Bacteria by Fluorescent Oligo(Phenyleneethynylene)s with Mannopyranoside Termini Groups
Previous Article in Journal
Deep Count: Fruit Counting Based on Deep Simulated Learning
Previous Article in Special Issue
Strategy for Accurate Detection of Escherichia coli O157:H7 in Ground Pork Using a Lateral Flow Immunoassay
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization of N-Acyl Homoserine Lactones in Vibrio tasmaniensis LGP32 by a Biosensor-Based UHPLC-HRMS/MS Method

1
Sorbonne Universités, UPMC Univ Paris 6, CNRS, Laboratoire de Biodiversité et Biotechnologies Microbiennes (LBBM), Observatoire Océanologique, F-66650 Banyuls/Mer, France
2
Sorbonne Universités, UPMC Univ Paris 06, CNRS, Observatoire Océanologique de Banyuls (OOB), F-66650 Banyuls/Mer, France
*
Author to whom correspondence should be addressed.
Sensors 2017, 17(4), 906; https://doi.org/10.3390/s17040906
Submission received: 4 April 2017 / Revised: 14 April 2017 / Accepted: 17 April 2017 / Published: 20 April 2017
(This article belongs to the Special Issue Sensors for Toxic and Pathogen Detection)

Abstract

:
Since the discovery of quorum sensing (QS) in the 1970s, many studies have demonstrated that Vibrio species coordinate activities such as biofilm formation, virulence, pathogenesis, and bioluminescence, through a large group of molecules called N-acyl homoserine lactones (AHLs). However, despite the extensive knowledge on the involved molecules and the biological processes controlled by QS in a few selected Vibrio strains, less is known about the overall diversity of AHLs produced by a broader range of environmental strains. To investigate the prevalence of QS capability of Vibrio environmental strains we analyzed 87 Vibrio spp. strains from the Banyuls Bacterial Culture Collection (WDCM911) for their ability to produce AHLs. This screening was based on three biosensors, which cover a large spectrum of AHLs, and revealed that only 9% of the screened isolates produced AHLs in the defined experimental conditions. Among these AHL-producing strains, Vibrio tasmaniensis LGP32 is a well-known pathogen of bivalves. We further analyzed the diversity of AHLs produced by this strain using a sensitive bioguided UHPLC-HRMS/MS approach (Ultra-High-Performance Liquid Chromatography followed by High-Resolution tandem Mass Spectrometry) and we identified C10-HSL, OH-C12-HSL, oxo-C12-HSL and C14:1-HSL as QS molecules. This is the first report that documents the production of AHL by Vibrio tasmaniensis LGP32.

1. Introduction

Bacteria of the genus Vibrio are ubiquitous marine bacteria belonging to the Gammaproteobacteria class and this genus includes both non-pathogenic and pathogenic species. Among the 133 described species of Vibrio (www.bacterio.net), at least 12 (e.g., V. cholerae, V. vulnificus, V. parahaemolyticus) are well known to be highly pathogenic to humans and a large number of other strains are pathogens of a wide range of marine organisms [1,2,3]. These opportunistic bacteria are also able to colonize diverse substrates, to participate in biofilm production [4,5], and to associate with phytoplankton, zooplankton, marine vertebrates and invertebrates [6,7,8,9]. Quorum sensing (QS) has been shown to be involved in many of these processes and associations [10,11], whereby the population density regulates gene expression, through the emission of self-generated small signal molecules named autoinducers (AI), to obtain a concerted physiological response [12,13,14].
Type 1 autoinducers (AI-1) also called N-acyl homoserine lactones (AHLs) are signaling compounds used by many bacteria for communication among single or closely related species [15,16,17,18]. In Gram-negative bacteria, the AHLs are synthesized by one or more synthases whose LuxI and LuxM/AinS are the most widespread [19,20,21]. Since the first investigations on QS in the symbiotic and bioluminescent V. fischeri, other Vibrio species were shown to communicate using AHLs [22,23,24]. QS molecules and signaling pathways are well known for V. cholerae, V. fischeri, V. harveyi, V. vulnificus and V. anguillarum as are their roles in bioluminescence, pathogenicity and biofilm formation [25]. Nonetheless, the prevalence and diversity of AHL-based QS in other environmental strains of Vibrio remains poorly understood.
Direct in situ studies of AHLs production in the marine environment have been limited and hampered by the fact that AHLs are likely released locally in microenvironments of high cellular concentration and thus at low total concentrations in a water sample. Therefore, quantifying AHL production directly in the environment is extremely difficult [26]. To overcome this problem, the isolation and identification of environmental strains able to produce AHLs and the characterization of these compounds appear as the best alternative. For these purpose, several detection approaches have been previously developed and most of them are using bioreporters able to detect a wide range of AHLs [27,28,29]. Most of these bioreporters are genetically modified strains, where reporter genes such as violacein or phenazine production, β-galactosidase or green fluorescent protein (GFP), are under the control of AHL-based-quorum sensing inducing promoters [29,30]. The use of multiple reporter strains does not allow the characterization of individual AHL molecules but provides an activation pattern reflecting a specific AHLs production phenotype in the defined culture condition [22,31]. Thus, in the past 10 years, numerous studies characterized AHLs—in combination or not with reporters—using analytical chemistry tools. The characterization of known AHLs is most commonly achieved by thin-layer chromatography (TLC) or High Performance Liquid Chromatography (HPLC) and comparison to standards [31,32,33,34]. However, since the diversity of AHLs is far from completely described [35], the characterization of novel AHLs requires the combination of chromatography with structural analysis, like Gas Chromatography coupled with Mass Spectrometry (GC-MS; [36]), HPLC coupled with tandem Mass Spectrometry (HPLC-MS/MS) or Nuclear Magnetic Resonance (NMR) [37]. The recent development of more sensitive and accurate techniques such as Ultra-High Performance Liquid Chromatography (UHPLC) combined to High Resolution Mass Spectrometry (HRMS) have yielded a high diversity of novel AHL molecules [38,39].
Vibrio tasmaniensis strain LGP32, previously named Vibrio splendidus, has been widely used as a model organism for the study of host-pathogens relationship in bivalves [40,41,42]. This strain had been isolated from diseased oysters during mortality events in France and is a facultative intracellular pathogen that attach and invade oyster hemocytes [43]. The complete genome analysis of LGP32 has revealed that—like other pathogenic ones, such as V. harveyi—this strain, harbors QS systems based on three different autoinducers (AHL, CAI-1 and AI-2) [16]. Interestingly, De Decker et al. showed in 2013 a possible involvement of QS in the virulence mechanisms of LGP32 [44], but remarkably the AHL molecules associated with this strain have not yet been characterized.
The aim of this study was to first evaluate the AHL production among environmental Vibrio spp. strains isolated from a large diversity of marine environments. The result of this screening led us to focus on the chemical diversity of AHLs in Vibrio tasmaniensis LGP32 using an optimized protocol to potentially maximize the discovery of novel AHLs. This protocol includes the use of large extraction volumes (3 L) followed by a biosensor-based screening of HPLC fractions prior to structural characterization by UHPLC-HRMS/MS. In this study, we also assessed the limit of detection of various approaches and extended the range of tested AHLs compared to previous studies. Our methodological approach allowed, for the first time, the characterization of AHL compounds involved in the QS of the important bivalve pathogen Vibrio tasmaniensis LGP32.

2. Materials and Methods

2.1. Culture Collection and Strains Identification

The Banyuls Bacterial Culture Collection (BBCC) is referenced in the World Data Center for Microorganisms as WDCM911 and harbors more than 2000 bacterial strains isolated from different geographical sites which are mostly heterotrophic marine bacteria identified on the 16S rRNA gene sequence [45]. Genomic DNA of each of the strains was extracted with the Wizard Genomic DNA purification kit (Promega, Charbonnières-les-Bains, France) as previously described [46]. PCR was performed using the universal primers targeting bacteria, 27Fmod (AGRGTTTGATC-MTGGCTCAG) [47] and 1492Rmod (TACGGYTACCTTGTTAYGACTT) [48]. PCR products were purified using the Agencourt AMPureXp purification kit (Beckman Coulter, Villepinte, France) and sequenced as described previously with an AB3130xl genetic analyzer (Applied Biosystems, Courtaboeuf, France). All molecular biology instrumentation was available through the Bio2Mar platform at the Observatoire Oceanologique de Banyuls-sur-Mer. The 16S rRNA gene sequences were compared to sequences within the NCBI nt database using the Basic Local Alignment Search Tool—2 sequences [49,50]. For the QS screening we selected all strains in the BBCC Culture Collection with a similarity percent above 98% to known Vibrio species (i.e., Table 1). The List of the 87 tested strains with their origin, their identification by 16S rRNA gene sequence and their GenBank accession numbers can be found in Table S1 of the Supplementary Material.

2.2. Biosensor Assays

The detection of AHLs in culture supernatants and HPLC fractions (see details in Section 2.3) followed previously described protocols using the biosensors Pseudomonas putida (P. putida), Escherichia coli (E. coli) and Chromobacterium violaceum (C. violaceum). The biosensors Pseudomonas putida F117 (pRK-C12; Kmr; ppuI::npt) and Escherichia coli MT102 (pJBA132) were used for the detection of AHLs in liquid medium and Chromobacterium violaceum CV026 was used for the detection of AHLs in solid medium [51,52,53]. Briefly, 50 µL of culture supernatants, obtained by centrifugation of 2 mL of culture in Marine Broth (MB), grown overnight at 25 °C under shaking (100 RPM), and 20 µL of HPLC fractions at 10 mg·mL−1 in dimethyl sulfoxide (DMSO) diluted in Luria Bertani (LB) broth (1/4 v:v), were tested in triplicate. For the biosensors E. coli and P. putida, microplates were incubated respectively at 30 °C and 37 °C and OD was measured at 535 nm after 0, 5 and 24 h of incubation. OD620 was also measured to control for biosensor cell growth [30,53,54,55]. For the biosensor C. violaceum, culture plates were incubated at 30 °C and purple zones of violacein production were inspected after 24 h [51]. For all tests, negative controls consisted of biosensor cultures without supernatant or fractions, and sterile LB medium. Biosensor cultures with addition of commercial AHLs (C6-HSL and oxo-C10-HSL, Cayman Chemical, Ann Arbor, MI, USA) were used as positive controls. To determine the AHL concentration range of detection of each biosensor, a dilution range from 0.01 to 25,000 nM of 27 commercial AHL standards (Cayman Chemical, see Table 2) was tested following the protocols described above.

2.3. AHL Extraction and HPLC Fractionation of LGP32

LGP32 was cultured in 3 L of Marine Broth (Difco, Le pont de Claix, France) at 25 °C under shaking (100 RPM) for 24 h (representing a late exponential phase, pH 7.5). A liquid-liquid extraction of the culture was performed with ethyl acetate (1/3 v:v) in a separatory funnel. The organic phase was evaporated to dryness and the extract was re-suspended in 1 mL of HPLC grade DMSO. The extract was fractionated using a separative HPLC system with two Varian Prep Star pumps, a manual injector, a Dionex Ultimate 3000 RS variable wavelength detector and a Dionex Ultimate 3000 fraction collector (Thermo Scientific, Courtaboeuf, France). The column was a Phenomenex Luna C18 (21.2 mm × 250 mm), with 5 μm particle size, and the flow rate was set to 20 mL·min−1. The mobile phase consisted of HPLC grade H2O and CH3CN at different proportions starting at 70:30 for 3 min, followed by a 12 min linear gradient from 70:30 to 0:100, followed by 100% CH3CN for 10 min. 22 fractions were collected every minute between 3 and 25 min. The solvent was removed with a HT-4X system (Genevac, Biopharma Technologies France, Lyon, France), each fraction was dissolved in 100 μL DMSO and diluted at 1/4 with LB medium (v/v) to perform the biosensor tests. Positive fractions were further analyzed by UHPLC-HRMS/MS. pH was controlled at each step of our experimental process and maintained in between 6 and 7.

2.4. AHL Detection by UHPLC-HRMS/MS

Prior to injection, fractions were diluted at 1 mg·mL−1 and 5 µL were injected. UHPLC-MS analyses were performed with a Dionex Ultimate 3000 UHPLC-HESI HRMS Q-Exactive focus system (Thermo Scientific) controlled by the Xcalibur software. The column was a Hypersil GOLD C18 (2.1 mm × 150 mm) with 1.9 µm particle size (Thermo Scientific). The column oven was set to 50 °C. The flow rate was maintained at 0.8 mL·min−1. The mobile phase was composed of 0.1% formic acid in water (eluent A) and 0.1% formic acid in acetonitrile (B). A gradient profile was used, starting with 100% of A, and keeping this composition constant for 5 min. The proportion of B was linearly increased to 100% in 5 min, and was left at 100% for 5 min. Settings for the ion source were: 20 aux gas flow rate, 75 sheath gas flow rate, 4 μA spray current, 3 kV spray voltage, 350 °C capillary temperature, 450 °C heater temperature, and 40 S-lens RF and nitrogen was used as nebulizing gas by the ion trap source.
Firstly, MS and MS/MS profiles were recorded alternating between a full scan (scan range 130 to 900 m/z) and All Ion Fragmentation (AIF) mode [scan range 60 to 600 m/z, normalized-collision energy (NCE) 25] to determine molecular weights and identify chromatographic peaks generating fragment ions at m/z 102.0555. Mass resolution was set at 35,000, AGC target was 1 × 106 and 5 × 104 respectively for the full scan and AIF mode, and injection time was 40 ms. The study of mass spectra obtained for our standard molecules and the study of Patel et al. [35] reveal that fragment ions at m/z 102.055, 84.045, 74.06 and 56.05 are specific of the homoserine lactone (HSL) moiety and these signals were chosen as the specific ions indicating the presence of AHL-type compounds. In a second step, we used a SIM (resolution 35,000) and dd-ms2 (resolution 17,500) mode to confirm the AHLs identification. MS/MS scans were isolated using an isolation width of 3.0 Da, fragmentations were performed at 17,500 with a collision energy of 20 eV. The limit of detection (LOD) of our UHPLC-HRMS method was established based on the standard deviation of the response to the 27 AHLs at 50 nmol·L−1 and the slope of a 10-fold concentration range (R2 values > 0.99). Briefly, a calibration curve was constructed using a simple linear regression analysis from the injection of AHL standard solution mixtures at concentration ranging from 5 to 500 nmol·L−1. Each AHL was injected 10 times at 50 nmol·L−1 and LOD was expressed as 3.3 times the standard deviation (LOD = 3.3SD).

3. Results

3.1. Biosensors Strains & UHPLC-HRMS/MS AHL Detection Limits

To define the specificity and the sensitivity of the three biosensors strains, we evaluated their responses to 27 commercially available AHL molecules (Table 2). The C. violaceum biosensor (CV026) showed a highly specific response to short chain AHLs (<10 carbons in the acyl chain) with a high sensitivity for C6-HSL (2.5 nmol·L−1), C7-HSL (1 nmol·L−1) and OH-C10-HSL (2.5 nmol·L−1). The biosensors E. coli MT102 detected mostly short acyl side chain AHLs and P. putida F117 showed the lowest specificity, by detecting 23 of 27 AHLs, while exhibited highest sensitivity to 3-oxo-HSL (LOD < 0.001 nmol·L−1; Table 2). Overall, GFP-based biosensors, MT102 and F117, showed the highest sensitivity levels to the most tested molecules. Finally, C4-HSL was only detected by C. violaceum.
Considering the overlap between the biosensors CV026 and MT102 and the suitability of GFP-based biosensors for high-throughput analyses, our fractionation and structural determination by UHPLC-HRMS/MS was solely based on detection by E. coli MT102 and P. putida F117 as biosensors. We evaluated the LOD of our UHPLC/HRMS-MS devices for 27 commercially available AHLs also used as AHL standards in our study and these detection limits were between 2.39 nmol·L−1 for the 3-OH-C12 HSL and 36.58 nmol·L−1 for the 3-OH-C14 HSL. Overall, short chain AHLs (with the notable exception of C4-HSL) presented lower LODs compared to long acyl side chain AHLs. All MS/MS fragmentation spectra of the 27 AHL standards can be found in Figure S1 of the Supplementary Material.

3.2. Screening of AHL-Producing Vibrio Strains

A total of 87 Vibrio spp. strains were selected from the BBCC Culture Collection of c.a. 2000 strains. The strains were previously isolated from seawater, crater lake water, sea urchin, green alga, sponge, Rodophyta, macrophytes, Urochordata or jellyfish and from different geographical sites. These strains are closely related to 28 described species, as shown in Table 1. Isolates closely related to the Splendidus clade represented 40% of the strains, the highest number of isolates was related to V. gigantis (16 isolates) mostly isolated from benthic and pelagic macro-organisms followed by V. splendidus (8 isolates) which were rather isolated from seawater. The 26 remaining species varied between 1 and 7 isolates. A 16S rRNA phylogenetic analysis of the strains and the GenBank accession numbers are provided in the Supplementary Material Figure S2 and Table S1.
In order to detect a broad diversity of AHLs, the isolates were tested for their capability to produce AHLs using three biosensors, P. putida F117 (pKR-C12), E. coli MT102 (pJBA-132) and C. violaceum (CV026; Table 1). The vast majority of the analyzed isolates (91%) did not produced AHLs able to activate any of our three biosensors in the defined experimental conditions. Eight isolates (9%) closely related to the species V. mytili, V. metschnikovii, V. scophtalmi, V. tasmaniensis and V. ordalii activated at least one of the three bioreporter strains. Strain BBCC 1015 (V. ordalii), CIP 107715 (V. tasmaniensis, LGP32) and BBCC 2428 (V. mytili) activated the biosensor C. violaceum CV026. The strain BBCC 1015 (V. ordalii), BBCC 1026 and 1055 (V. metschnikovii), BBCC 1228, 1237, and 1238 (V. scophtalmi), CIP 107715 (V. tasmaniensis, LGP32) activated the biosensor E. coli MT102. All these strains, except for LGP32, also activated the biosensor P. putida F117. In addition to the fact that LGP32 is likely to produce a diverse panel of AHLs, this strain is also a well-known pathogen of marine invertebrates including some with commercial interest [1]. It therefore appeared important to describe the AHLs produced by this strain.

3.3. Vibrio Tasmaniensis LGP32: AHLs Characterization by UHPLC-HRMS/MS

The culture supernatant was extracted with ethyl acetate and fractionated into 22 fractions, identified as LGP32_a to LGP32_v. These fractions were then tested for AHL production using the biosensor strains E. coli MT102 and P. putida F117. A total of four fractions (LGP32_l, LGP32_m, LGP32_o and LGP32_p) were positive with at least one of the biosensors. UHPLC-HRMS and UHPLC-HRMS/MS analyses were then performed to identify AHLs. The presence of AHLs in the fractions were revealed by the detection in the fragmentation patterns of at least one of the following characteristic lactone ring fragments (m/z 102.055, 84.045, 74.061 and 56.050) [35]. In the condition tested, four different AHLs were detected including unsubstituted, oxo and hydroxy AHLs at the third carbon atom (Table 3). The AHLs detected for this strain were C10-HSL (N-decanoyl homoserine lactone), OH-C12-HSL (N-3-hydroxy-dodecanoyl homoserine lactone), oxo-C12-HSL (N-3-oxo-dodecanoyl homoserine lactone) and C14:1-HSL (N-tetradecenoyl homoserine lactone). These identifications were supported by the analysis of 27 standard AHLs (Table 4), the retention time and the exact mass of the [M + H]+ pseudo-molecular ion (precision 3 ppm). However, the exact double bond position on the acyl side chain of the C14:1-HSL has not been determined as it is not easily achievable by mass spectroscopy and such determination would have required derivatization methods or high-field NMR.

4. Discussion

4.1. Diversity of AHL Producing Strains

We investigated AHL production among 87 Vibrio spp. strains using three different bioreporter strains. Remarkably, while the production of AHL quorum-sensing signal molecules has been widely reported among Vibrio, only a small percentage of our Vibrio spp. strains (9%) were shown to produce AHLs.
This result was different from the observations made by Garcia-Aljaro [22] and Purohit et al., [56] who found that the majority (85%) of Vibrio spp. strains in their collection were AHL producers. Similarly, Yang et al. focused on 25 strains and found 23 positive for AHL production [31]. By contrast, and in line with our study, Rasmussen et al., found only 32 positive strains (10%) among the 301 tested in their culture collection [24]. Different non-exclusive hypotheses can be made to explain this low percentage of AHL producing Vibrio: (1) the growth conditions might not have been optimal for all Vibrio spp. strains to reach the necessary density to produce AHLs or the threshold is variable among strains [57]; (2) The concentrations of produced AHLs might be below the limit of detection of our biosensors; (3) The strains produce novel or undetected AHLs, that are not activating our biosensors and (4) The strains do not contain the machinery necessary to produce AHLs. Since we followed the culture growth and we did not observe significant differences in OD between positive and negative strains, hypothesis (1) is somewhat less supported by our results than the remainder explanations. The differences between the activation profiles among strains reflect a diversity of produced AHLs and that can be affected by the genetic diversity of AHL synthases but also the presence of one or more synthases in their genome [58], in agreement with previous observations showing high intraspecific genetic diversity in the genus Vibrio [59]. However, interactions and ecological processes that drive or are affected by this phenotypic variation are still poorly understood and warrant further studies [60,61].

4.2. AHL Diversity of Vibrio Tasmaniensis LGP32

The observation of AHL production in Vibrio tasmaniensis LGP32 is notable, as this strain is a well-known pathogen and producer of outer membrane vesicles (OMVs), OmpU, porins and metalloproteases [41,62,63], all involved in the virulence in oyster larvae [44]. We detected four produced AHLs: C10-HSL, 3-OH-C12-HSL, 3-oxo-C12-HSL and C14:1-HSL and this is the first report of C10-HSL, 3-oxo-C12-HSL and C14:1-HSL in Vibrio strains belonging to the Splendidus clade. On the other hand 3-OH-C12-HSL have already been reported [24,54]. Among these four AHLs, three have already been identified in the putative pathogens V. campbellii, V. furnissii, V. fluvialis and V. anguillarum [32,64,65,66]. More broadly among Gram-negative Proteobacteria, 3-oxo-C12-HSL controls biofilm production in Pseudomonas aeruginosa and has a determining immunomodulatory activity of the human host [67], 3-OH-C12-HSL is involved in virulence factor production by Acinetobacter baumannii [68], and finally C14:1-HSL participate in the establishment of a necrosis phenotype in Agrobacterium vitis [69]. Considering the AHLs regulation of virulence mechanisms in other Vibrio species [11,70,71] and other Proteobacteria, our results might add to the understanding of the role of AHLs in the physiology and the pathogenicity of these microorganisms.

4.3. Method Performance

In addition to the results described above, our work also provides data on AHL detection limits by AHL bioreporter strains and by UHPLC-HRMS/MS. Such data, and more especially their comparison, is crucial for future studies of AHL production by Proteobacteria. Previous reports have already characterized the limits of detection for these biosensors [36,51,53]. However, in this study, we extended our work to a larger panel of AHLs including long acyl side chain compounds (>14 carbons in the acyl chain), that have not been determined before. While numerous studies have demonstrated that the ability to detect OH-HSL is unique to Agrobacterium tumefaciens NTL4 (pZLR4) [22,72], we showed that Pseudomonas putida F117 is also able to detect that type of AHLs with similar detection limits (between 13.43 and 0.125 nmol·L−1).
Our UHPLC-HRMS protocol yielded a good separation and well defined peaks for 26 AHL standards, with a mass accuracy for all standards below 3 ppm and a median LOD of 10.58 nmol·L−1. Finally, the combination of two GFP-based biosensors, E. coli MT102 and P. putida F117, was responsive to over 90% of the AHL standards and exhibited a lower limit of detection than analytical UHPLC-HRMS methods [24,35]. Our protocol included larger culture volumes compared to those used in similar studies (75 µL to 50 mL [24,35,56]), an activity-based screening of HPLC fractions, and UHPLC-HRMS/MS structural determination that can in theory increase the detection of rarer and/or novel AHL structures when compared with these previously published approaches. The simple fact that HPLC fractions are less complex than supernatants or raw extracts might significantly increase the discovery of novel QS-receptor agonists in the future. In the current study, we did not quantify the AHLs production, but this method is fully compatible with quantification.
The non-targeted UHPLC/HRMS/MS method reported here allows, prior to NMR structure determination, the identification of novel AHLs without any corresponding standards. The AHL identification is based on the study of MS/MS fragmentation patterns and the search of characteristic fragment ions corresponding to the lactone ring fragmentation [35]. In theory, this method could be more suitable for comparative studies of AHL production. Unfortunately, we did not uncover novel AHLs in LGP32 to demonstrate this potential but several ongoing studies with other gram-negative bacterial strains in our group have already yielded novel AHL structures using the same approach.

5. Conclusions

In this work, we confirmed that AHL are produced by different species of Vibrio and that this production varies among different strains of the same species, pointing to the need of further studies to understand the biological origin as well as ecological significance of this intraspecific variation. To our knowledge, this is the first study that demonstrated AHLs production of Vibrio tasmaniensis LGP32 a pathogenic bacterium involved in oyster mortality. Four AHLs, namely C10-HSL, OH-C12-HSL, oxo-C12-HSL and C14:1-HSL were detected and identified by our novel approach combining a HPLC fractionation followed by an activity-based AHL identification by UHPLC-HRMS/MS. This study should be useful for the understanding of the virulence and physiology mechanisms of LGP32. In addition, the limit of detection of a large panel of AHL standards was established and a broader response range was highlighted for the biosensors Escherichia coli MT102 and Pseudomonas putida F117 compared to previous studies.

Supplementary Materials

The following are available online at https://www.mdpi.com/1424-8220/17/4/906/s1, Figure S1: Fragmentation MS/MS spectra of AHL standards, Figure S2: Maximum likelihood tree of 16S rDNA gene sequence (559 bp) of the 87 isolates and 35 type strains of Vibrio using the Kimura 2 parameter (K2+G+I, Mega), Table S1: List of the 87 tested strains with their origin, their identification by 16S rRNA gene sequence and their GenBank accession numbers.

Acknowledgments

We thank the French Ministry of Education, for a PhD fellowship to Lea Girard through the Doctoral School 227 of the University Pierre et Marie Curie and the National Museum of Natural History, Paris. We thank the CNRS (grant EC2CO-ROSEOCOM) and the Sorbonne Universités (grant ENDOQUO) for funding this research. We are grateful to the BIO2MAR platform (http://bio2mar.obs-banyuls.fr) for access to instrumentation and particularly Karine Escoubeyrou for providing support technical aid. We thank Alice Rodrigues for her support in analytical chemistry and Nicole Batailler for her technical assistance in some steps of the experimental work.

Author Contributions

L.G., J.B., M.S. and R.L. conceived and designed the experiments; L.G., É.B. and L.I. performed the experiments; L.G., É.B. and D.S. analyzed the data; P.L. and L.I. contributed by giving access to Banyuls Bacterial Culture Collection (WDCM911); M.S., R.L. and J.B. gave the financial support and L.G., M.S., R.L. and J.B. wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Austin, B. Vibrios as causal agents of zoonoses. Vet. Microbiol. 2010, 140, 310–317. [Google Scholar] [CrossRef] [PubMed]
  2. Kwok, A.Y.; Wilson, J.T.; Coulthart, M.; Ng, L.-K.; Mutharia, L.; Chow, A.W. Phylogenetic study and identification of human pathogenic Vibrio species based on partial hsp 60 gene sequences. Can. J. Microbiol. 2002, 48, 903–910. [Google Scholar] [CrossRef] [PubMed]
  3. Sawabe, T.; Ogura, Y.; Matsumura, Y.; Feng, G.; Amin, A.R.; Mino, S.; Nakagawa, S.; Sawabe, T.; Kumar, R.; Fukui, Y.; et al. Updating the Vibrio clades defined by multilocus sequence phylogeny: Proposal of eight new clades, and the description of Vibrio tritonius sp. nov. Front. Microbiol. 2013, 4, 414. [Google Scholar] [CrossRef] [PubMed]
  4. Chimetto Tonon, L.A.; Silva, B.S.; Moreira, A.P.B.; Valle, C.; Alves, N.; Cavalcanti, G.; Garcia, G.; Lopes, R.M.; Francini-Filho, R.B.; de Moura, R.L.; et al. Diversity and ecological structure of vibrios in benthic and pelagic habitats along a latitudinal gradient in the Southwest Atlantic Ocean. PeerJ 2015, 3, e741. [Google Scholar] [CrossRef] [PubMed]
  5. Vezzulli, L.; Pezzati, E.; Stauder, M.; Stagnaro, L.; Venier, P.; Pruzzo, C. Aquatic ecology of the oyster pathogens Vibrio splendidus and Vibrio aestuarianus. Environ. Microbiol. 2015, 17, 1065–1080. [Google Scholar] [CrossRef] [PubMed]
  6. Ben-Haim, Y. Vibrio coralliilyticus sp. nov., a temperature-dependent pathogen of the coral Pocillopora damicornis. Int. J. Syst. Evol. Microbiol. 2003, 53, 309–315. [Google Scholar] [PubMed]
  7. Heidelberg, J.F.; Heidelberg, K.B.; Colwell, R.R. Bacteria of the -subclass Proteobacteria associated with zooplankton in Chesapeake bay. Appl. Environ. Microbiol. 2002, 68, 5498–5507. [Google Scholar] [CrossRef] [PubMed]
  8. Soto, W.; Gutierrez, J.; Remmenga, M.D.; Nishiguchi, M.K. Salinity and temperature effects on physiological responses of Vibrio fischeri from diverse ccological niches. Microb. Ecol. 2009, 57, 140–150. [Google Scholar] [CrossRef] [PubMed]
  9. Wendling, C.C.; Wegner, K.M. Adaptation to enemy shifts: Rapid resistance evolution to local Vibrio spp. in invasive Pacific oysters. Proc. R. Soc. B Biol. Sci. 2015, 282, 20142244. [Google Scholar] [CrossRef] [PubMed]
  10. Hammer, B.K.; Bassler, B.L. Quorum sensing controls biofilm formation in Vibrio cholerae. Mol. Microbiol. 2003, 50, 101–104. [Google Scholar] [CrossRef] [PubMed]
  11. Zhu, J.; Miller, M.B.; Vance, R.E.; Dziejman, M.; Bassler, B.L.; Mekalanos, J.J. Quorum-sensing regulators control virulence gene expression in Vibrio cholerae. Proc. Natl. Acad. Sci. USA 2002, 99, 3129–3134. [Google Scholar] [CrossRef] [PubMed]
  12. Eberhard, A.; Burlingame, A.L.; Eberhard, C.; Kenyon, G.L.; Nealson, K.H.; Oppenheimer, N.J. Structural identification of autoinducer of Photobacterium fischeri luciferase. Biochemistry 1981, 20, 2444–2449. [Google Scholar] [CrossRef] [PubMed]
  13. Fuqua, W.C.; Winans, S.C.; Greenberg, E.P. Quorum sensing in bacteria: The LuxR-LuxI family of cell density-responsive transcriptional regulators. J. Bacteriol. 1994, 176, 269–275. [Google Scholar] [CrossRef] [PubMed]
  14. Nealson, K.H.; Platt, T.; Hastings, J.W. Cellular control of the synthesis and activity of the bacterial luminescent system. J. Bacteriol. 1970, 104, 313–322. [Google Scholar] [PubMed]
  15. Bassler, B.L. How bacteria talk to each other: Regulation of gene expression by quorum sensing. Curr. Opin. Microbiol. 1999, 2, 582–587. [Google Scholar] [CrossRef]
  16. Henke, J.M.; Bassler, B.L. Bacterial social engagements. Trends Cell Biol. 2004, 14, 648–656. [Google Scholar] [CrossRef] [PubMed]
  17. Miller, M.B.; Bassler, B.L. Quorum sensing in bacteria. Annu. Rev. Microbiol. 2001, 55, 165–199. [Google Scholar] [CrossRef] [PubMed]
  18. Persat, A.; Nadell, C.D.; Kim, M.K.; Ingremeau, F.; Siryaporn, A.; Drescher, K.; Wingreen, N.S.; Bassler, B.L.; Gitai, Z.; Stone, H.A. The mechanical world of bacteria. Cell 2015, 161, 988–997. [Google Scholar] [CrossRef] [PubMed]
  19. Engebrecht, J.; Silverman, M. Identification of genes and gene products necessary for bacterial bioluminescence. Proc. Natl. Acad. Sci. USA 1984, 81, 4154–4158. [Google Scholar] [CrossRef] [PubMed]
  20. Fuqua, C.; Greenberg, E.P. Self perception in bacteria: Quorum sensing with acylated homoserine lactones. Curr. Opin. Microbiol. 1998, 1, 183–189. [Google Scholar] [CrossRef]
  21. Gilson, L.; Kuo, A.; Dunlap, P.V. AinS and a new family of autoinducer synthesis proteins. J. Bacteriol. 1995, 177, 6946–6951. [Google Scholar] [CrossRef] [PubMed]
  22. García-Aljaro, C.; Vargas-Cespedes, G.J.; Blanch, A.R. Detection of acylated homoserine lactones produced by Vibrio spp. and related species isolated from water and aquatic organisms. J. Appl. Microbiol. 2012, 112, 383–389. [Google Scholar] [CrossRef] [PubMed]
  23. Meighen, E.A. Bacterial bioluminescence: Organization, regulation, and application of the lux genes. FASEB J. 1993, 7, 1016–1022. [Google Scholar] [PubMed]
  24. Rasmussen, B.; Nielsen, K.; Machado, H.; Melchiorsen, J.; Gram, L.; Sonnenschein, E. Global and phylogenetic distribution of quorum sensing signals, acyl homoserine lactones, in the family of Vibrionaceae. Mar. Drugs 2014, 12, 5527–5546. [Google Scholar] [CrossRef] [PubMed]
  25. Milton, D.L. Quorum sensing in vibrios: Complexity for diversification. Int. J. Med. Microbiol. 2006, 296, 61–71. [Google Scholar] [CrossRef] [PubMed]
  26. Hmelo, L.; van Mooy, B. Kinetic constraints on acylated homoserine lactone-based quorum sensing in marine environments. Aquat. Microb. Ecol. 2009, 54, 127–133. [Google Scholar] [CrossRef]
  27. Burton, E.O.; Read, H.W.; Pellitteri, M.C.; Hickey, W.J. Identification of acyl-homoserine lactone signal molecules produced by Nitrosomonas europaea strain Schmidt. Appl. Environ. Microbiol. 2005, 71, 4906–4909. [Google Scholar] [CrossRef] [PubMed]
  28. Shaw, P.D.; Ping, G.; Daly, S.L.; Cha, C.; Cronan, J.E.; Rinehart, K.L.; Farrand, S.K. Detecting and characterizing N-acyl-homoserine lactone signal molecules by thin-layer chromatography. Proc. Natl. Acad. Sci. USA 1997, 94, 6036–6041. [Google Scholar] [CrossRef] [PubMed]
  29. Yong, Y.-C.; Zhong, J.-J. A genetically engineered whole-cell pigment-based bacterial biosensing system for quantification of N-butyryl homoserine lactone quorum sensing signal. Biosens. Bioelectron. 2009, 25, 41–47. [Google Scholar] [CrossRef] [PubMed]
  30. Steindler, L.; Venturi, V. Detection of quorum-sensing N-acyl-homoserine lactone signal molecules by bacterial biosensors. FEMS Microbiol. Lett. 2007, 266, 1–9. [Google Scholar] [CrossRef] [PubMed]
  31. Yang, Q.; Han, Y.; Zhang, X.-H. Detection of quorum sensing signal molecules in the family Vibrionaceae. J. Appl. Microbiol. 2011, 110, 1438–1448. [Google Scholar] [CrossRef] [PubMed]
  32. Buchholtz, C.; Nielsen, K.F.; Milton, D.L.; Larsen, J.L.; Gram, L. Profiling of acylated homoserine lactones of Vibrio anguillarum in vitro and in vivo: Influence of growth conditions and serotype. Syst. Appl. Microbiol. 2006, 29, 433–445. [Google Scholar] [CrossRef] [PubMed]
  33. Kalia, V.C. (Ed.) Quorum Sensing vs. Quorum Quenching: A Battle with No End in Sight; Springer: New Delhi, India, 2015. [Google Scholar]
  34. Li, X.; Fekete, A.; Englmann, M.; Götz, C.; Rothballer, M.; Frommberger, M.; Buddrus, K.; Fekete, J.; Cai, C.; Schröder, P.; et al. Development and application of a method for the analysis of N-acylhomoserine lactones by solid-phase extraction and ultra high pressure liquid chromatography. J. Chromatogr. A 2006, 1134, 186–193. [Google Scholar] [CrossRef] [PubMed]
  35. Patel, N.M.; Moore, J.D.; Blackwell, H.E.; Amador-Noguez, D. Identification of unanticipated and novel N-acyl L-homoserine lactones (AHLs) using a sensitive non-targeted LC-MS/MS method. PLoS ONE 2016, 11, e0163469. [Google Scholar] [CrossRef] [PubMed]
  36. Wagner-Döbler, I.; Thiel, V.; Eberl, L.; Allgaier, M.; Bodor, A.; Meyer, S.; Ebner, S.; Hennig, A.; Pukall, R.; Schulz, S. Discovery of complex mixtures of novel long-chain quorum sensing signals in free-living and host-associated marine Alphaproteobacteria. ChemBioChem 2005, 6, 2195–2206. [Google Scholar] [CrossRef] [PubMed]
  37. Cao, J.G.; Meighen, E.A. Purification and structural identification of an autoinducer for the luminescence system of Vibrio harveyi. J. Biol. Chem. 1989, 264, 21670–21676. [Google Scholar] [PubMed]
  38. Fekete, A.; Frommberger, M.; Rothballer, M.; Li, X.; Englmann, M.; Fekete, J.; Hartmann, A.; Eberl, L.; Schmitt-Kopplin, P. Identification of bacterial N-acylhomoserine lactones (AHLs) with a combination of ultra-performance liquid chromatography (UPLC), ultra-high-resolution mass spectrometry, and in-situ biosensors. Anal. Bioanal. Chem. 2007, 387, 455–467. [Google Scholar] [CrossRef] [PubMed]
  39. Wang, J.; Quan, C.; Wang, X.; Zhao, P.; Fan, S. Extraction, purification and identification of bacterial signal molecules based on N-acyl homoserine lactones: Extraction, purification and identification of HSLs. Microb. Biotechnol. 2011, 4, 479–490. [Google Scholar] [CrossRef] [PubMed]
  40. Balbi, T.; Fabbri, R.; Cortese, K.; Smerilli, A.; Ciacci, C.; Grande, C.; Vezzulli, L.; Pruzzo, C.; Canesi, L. Interactions between Mytilus galloprovincialis hemocytes and the bivalve pathogens Vibrio aestuarianus 01/032 and Vibrio splendidus LGP32. Fish Shellfish Immunol. 2013, 35, 1906–1915. [Google Scholar] [CrossRef] [PubMed]
  41. Duperthuy, M.; Schmitt, P.; Garzon, E.; Caro, A.; Rosa, R.D.; Le Roux, F.; Lautredou-Audouy, N.; Got, P.; Romestand, B.; de Lorgeril, J.; et al. Use of OmpU porins for attachment and invasion of Crassostrea gigas immune cells by the oyster pathogen Vibrio splendidus. Proc. Natl. Acad. Sci. USA 2011, 108, 2993–2998. [Google Scholar] [CrossRef] [PubMed]
  42. Mateo, D.; Spurmanis, A.; Siah, A.; Araya, M.; Kulka, M.; Berthe, F.; Johnson, G.; Greenwood, S. Changes induced by two strains of Vibrio splendidus in haemocyte subpopulations of Mya arenaria, detected by flow cytometry with LysoTracker. Dis. Aquat. Organ. 2009, 86, 253–262. [Google Scholar] [CrossRef] [PubMed]
  43. Le Roux, F.; Zouine, M.; Chakroun, N.; Binesse, J.; Saulnier, D.; Bouchier, C.; Zidane, N.; Ma, L.; Rusniok, C.; Lajus, A.; et al. Genome sequence of Vibrio splendidus: An abundant planctonic marine species with a large genotypic diversity. Environ. Microbiol. 2009, 11, 1959–1970. [Google Scholar] [CrossRef] [PubMed]
  44. De Decker, S.; Reynaud, Y.; Saulnier, D. First molecular evidence of cross-species induction of metalloprotease gene expression in Vibrio strains pathogenic for Pacific oyster Crassostrea gigas involving a quorum sensing system. Aquaculture 2013, 392–395, 1–7. [Google Scholar] [CrossRef]
  45. MOLA Collection. Available online: http://collection.obs-banyuls.fr/ (accessed on 20 April 2017).
  46. Fagervold, S.K.; Urios, L.; Intertaglia, L.; Batailler, N.; Lebaron, P.; Suzuki, M.T. Pleionea mediterranea gen. nov., sp. nov., a gammaproteobacterium isolated from coastal seawater. Int. J. Syst. Evol. Microbiol. 2013, 63, 2700–2705. [Google Scholar] [CrossRef] [PubMed]
  47. Eiler, A.; Bertilsson, S. Composition of freshwater bacterial communities associated with cyanobacterial blooms in four Swedish lakes. Environ. Microbiol. 2004, 6, 1228–1243. [Google Scholar] [CrossRef] [PubMed]
  48. Acinas, S.G.; Sarma-Rupavtarm, R.; Klepac-Ceraj, V.; Polz, M.F. PCR-induced sequence artifacts and bias: Insights from comparison of two 16S rRNA clone libraries constructed from the same sample. Appl. Environ. Microbiol. 2005, 71, 8966–8969. [Google Scholar] [CrossRef] [PubMed]
  49. Tatusova, T.A.; Madden, T.L. BLAST 2 Sequences, a new tool for comparing protein and nucleotide sequences. FEMS Microbiol. Lett. 1999, 174, 247–250. [Google Scholar] [CrossRef] [PubMed]
  50. Blast. Available online: https://blast.ncbi.nlm.nih.gov/Blast.cgi?PAGE_TYPE=BlastSearch (accessed on 20 April 2017).
  51. McClean, K.H.; Winson, M.K.; Fish, L.; Taylor, A.; Chhabra, S.R.; Camara, M.; Daykin, M.; Lamb, J.H.; Swift, S.; Bycroft, B.W.; et al. Quorum sensing and Chromobacterium violaceum: Exploitation of violacein production and inhibition for the detection of N-acylhomoserine lactones. Microbiology 1997, 143, 3703–3711. [Google Scholar] [CrossRef] [PubMed]
  52. Riedel, K.; Hentzer, M.; Geisenberger, O.; Huber, B.; Steidle, A.; Wu, H.; Høiby, N.; Givskov, M.; Molin, S.; Eberl, L. N-acylhomoserine-lactone-mediated communication between Pseudomonas aeruginosa and Burkholderia cepacia in mixed biofilms. Microbiology 2001, 147, 3249–3262. [Google Scholar] [CrossRef] [PubMed]
  53. Winson, M.K.; Swift, S.; Fish, L.; Throup, J.P.; Jørgensen, F.; Chhabra, S.R.; Bycroft, B.W.; Williams, P.; Stewart, G.S. Construction and analysis of luxCDABE-based plasmid sensors for investigating N-acyl homoserine lactone-mediated quorum sensing. FEMS Microbiol. Lett. 1998, 163, 185–192. [Google Scholar] [CrossRef] [PubMed]
  54. Andersen, J.B.; Heydorn, A.; Hentzer, M.; Eberl, L.; Geisenberger, O.; Christensen, B.B.; Molin, S.; Givskov, M. gfp-Based N-acyl homoserine-lactone sensor systems for detection of bacterial communication. Appl. Environ. Microbiol. 2001, 67, 575–585. [Google Scholar] [CrossRef] [PubMed]
  55. Ravn, L.; Christensen, A.B.; Molin, S.; Givskov, M.; Gram, L. Methods for detecting acylated homoserine lactones produced by Gram-negative bacteria and their application in studies of AHL-production kinetics. J. Microbiol. Methods 2001, 44, 239–251. [Google Scholar] [CrossRef]
  56. Purohit, A.A.; Johansen, J.A.; Hansen, H.; Leiros, H.-K.S.; Kashulin, A.; Karlsen, C.; Smalås, A.; Haugen, P.; Willassen, N.P. Presence of acyl-homoserine lactones in 57 members of the Vibrionaceae family. J. Appl. Microbiol. 2013, 115, 835–847. [Google Scholar] [CrossRef] [PubMed]
  57. Bhedi, C.D.; Prevatte, C.W.; Lookadoo, M.S.; Waikel, P.A.; Gillevet, P.M.; Sikaroodi, M.; Campagna, S.R.; Richardson, L.L. Elevated temperature enhances short to medium chain acyl homoserine lactone production by black band disease associated vibrios. FEMS Microbiol. Ecol. 2017. [Google Scholar] [CrossRef] [PubMed]
  58. Tait, K.; Hutchison, Z.; Thompson, F.L.; Munn, C.B. Quorum sensing signal production and inhibition by coral-associated vibrios: Quorum sensing and coral-associated vibrios. Environ. Microbiol. Rep. 2010, 2, 145–150. [Google Scholar] [CrossRef] [PubMed]
  59. Sawabe, T.; Koizumi, S.; Fukui, Y.; Nakagawa, S.; Ivanova, E.P.; Kita-Tsukamoto, K.; Kogure, K.; Thompson, F.L. Mutation is the main driving force in the diversification of the Vibrio splendidus clade. Microbes Environ. 2009, 24, 281–285. [Google Scholar] [CrossRef] [PubMed]
  60. Keller, L.; Surette, M.G. Communication in bacteria: An ecological and evolutionary perspective. Nat. Rev. Microbiol. 2006, 4, 249–258. [Google Scholar] [CrossRef] [PubMed]
  61. Platt, T.G.; Fuqua, C. What’s in a name? The semantics of quorum sensing. Trends Microbiol. 2010, 18, 383–387. [Google Scholar] [CrossRef] [PubMed]
  62. Binesse, J.; Delsert, C.; Saulnier, D.; Champomier-Verges, M.-C.; Zagorec, M.; Munier-Lehmann, H.; Mazel, D.; Le Roux, F. Metalloprotease vsm is the major determinant of toxicity for extracellular products of Vibrio splendidus. Appl. Environ. Microbiol. 2008, 74, 7108–7117. [Google Scholar] [CrossRef] [PubMed]
  63. Vanhove, A.S.; Duperthuy, M.; Charrière, G.M.; Le Roux, F.; Goudenège, D.; Gourbal, B.; Kieffer-Jaquinod, S.; Couté, Y.; Wai, S.N.; Destoumieux-Garzón, D. Outer membrane vesicles are vehicles for the delivery of Vibrio tasmaniensis virulence factors to oyster immune cells. Environ. Microbiol. 2015, 17, 1152–1165. [Google Scholar] [CrossRef] [PubMed]
  64. Derber, C.; Coudron, P.; Tarr, C.; Gladney, L.; Turnsek, M.; Shankaran, S.; Wong, E. Vibrio furnissii: An unusual cause of bacteremia and skin lesions after ingestion of seafood. J. Clin. Microbiol. 2011, 49, 2348–2349. [Google Scholar] [CrossRef] [PubMed]
  65. Haldar, S.; Chatterjee, S.; Sugimoto, N.; Das, S.; Chowdhury, N.; Hinenoya, A.; Asakura, M.; Yamasaki, S. Identification of Vibrio campbellii isolated from diseased farm-shrimps from south India and establishment of its pathogenic potential in an Artemia model. Microbiology 2011, 157, 179–188. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, Y.; Wang, H.; Liang, W.; Hay, A.J.; Zhong, Z.; Kan, B.; Zhu, J. Quorum sensing regulatory cascades control Vibrio fluvialis pathogenesis. J. Bacteriol. 2013, 195, 3583–3589. [Google Scholar] [CrossRef] [PubMed]
  67. Williams, P.; Camara, M.; Hardman, A.; Swift, S.; Milton, D.; Hope, V.J.; Winzer, K.; Middleton, B.; Pritchard, D.I.; Bycroft, B.W. Quorum sensing and the population-dependent control of virulence. Philos. Trans. R. Soc. B Biol. Sci. 2000, 355, 667–680. [Google Scholar] [CrossRef] [PubMed]
  68. Bhargava, N.; Singh, S.P.; Sharma, A.; Sharma, P.; Capalash, N. Attenuation of quorum sensing-mediated virulence of Acinetobacter baumannii by Glycyrrhiza glabra flavonoids. Future Microbiol. 2015, 10, 1953–1968. [Google Scholar] [CrossRef] [PubMed]
  69. Hao, G.; Burr, T.J. Regulation of long-chain N-acyl-homoserine lactones in Agrobacterium vitis. J. Bacteriol. 2006, 188, 2173–2183. [Google Scholar] [CrossRef] [PubMed]
  70. Frans, I.; Michiels, C.W.; Bossier, P.; Willems, K.A.; Lievens, B.; Rediers, H. Vibrio anguillarum as a fish pathogen: Virulence factors, diagnosis and prevention. J. Fish Dis. 2011, 34, 643–661. [Google Scholar] [CrossRef] [PubMed]
  71. Lilley, B.N.; Bassler, B.L. Regulation of quorum sensing in Vibrio harveyi by LuxO and Sigma-54. Mol. Microbiol. 2000, 36, 940–954. [Google Scholar] [CrossRef] [PubMed]
  72. Cha, C.; Gao, P.; Chen, Y.-C.; Shaw, P.D.; Farrand, S.K. Production of acyl-homoserine lactone quorum-sensing signals by Gram-negative plant-associated bacteria. Mol. Plant. Microbe Interact. 1998, 11, 1119–1129. [Google Scholar] [CrossRef] [PubMed]
Table 1. Activation patterns of Vibrio species when tested against three biosensor strains: F117, Pseudomonas putida (pKR-C12); MT102, Escherichia coli (pJBA-132) and CV026, Chromobacterium violaceum.
Table 1. Activation patterns of Vibrio species when tested against three biosensor strains: F117, Pseudomonas putida (pKR-C12); MT102, Escherichia coli (pJBA-132) and CV026, Chromobacterium violaceum.
Closest Relative SpeciesNumber of IsolatesBBCC CodeCV026MT102F117
Vibrio atlanticus12313---
Vibrio brasiliensis2493, 494---
Vibrio breoganii11958---
Vibrio campbellii362, 416, 2415---
Vibrio chagasii4583, 586, 640, 2353---
Vibrio cortegadensis2529, 1974---
Vibrio gallaecicus4528, 2315, 2319, 2327---
Vibrio gigantis16503, 530, 853, 1230, 1232, 1233, 1972, 1973, 1980, 1982, 1989, 2045, 2312, 2357, 2372, 2412---
Vibrio harveyi7558, 576, 579, 605, 615, 626, 2366---
Vibrio hemicentroti22269, 2311---
Vibrio ichthyoenteri2490, 491---
Vibrio lentus3495, 850, 851---
Vibrio maritimus12338---
Vibrio metschnikovii21026, 1055-++
Vibrio mytili12428+--
Vibrio natriegens1546---
Vibrio neptunius1496---
Vibrio ordalii11015+++
Vibrio owensii31143, 1169, 1955---
Vibrio pectenicida11971---
Vibrio pomeroyi2502, 1962---
Vibrio rumoiensis21210, 1211---
Vibrio scophtalmi31228, 1237, 1238-++
Vibrio scophtalmi42361, 2365, 2370, 2413---
Vibrio shilonii32339, 2351, 2363---
Vibrio sinaloensis12347---
Vibrio splendidus866, 67, 165, 239, 498, 500, 527, 852---
Vibrio tasmaniensis1526---
Vibrio tasmaniensis LGP3212197++-
Vibrio tubiashi4620, 1231, 2159, 2190---
Table 2. Limit of detection (nmol·L−1) of AHL standards using UHPLC-HRMS and three different biosensors: F117, Pseudomonas putida (pKR-C12); MT102, Escherichia coli (pJBA-132) and CV026, Chromobacterium violaceum. ND: not detected. MD: missing data.
Table 2. Limit of detection (nmol·L−1) of AHL standards using UHPLC-HRMS and three different biosensors: F117, Pseudomonas putida (pKR-C12); MT102, Escherichia coli (pJBA-132) and CV026, Chromobacterium violaceum. ND: not detected. MD: missing data.
Limit of Detection (nmol·L−1)
CV026MT102F117UHPLC-HRMS
C4-HSL250NDND>500
C6-HSL2.50.631312.383.64
OXO-C6-HSL10<0.001ND10.90
C7-HSL10.094ND5.33
C8-HSL51.125212.496.50
OXO-C8-HSL100.00240.766.15
OH-C8-HSL100ND7.077MD
C9-HSL51.932.897.37
C10-HSL10074.521.54.56
OXO-C10-HSL10000.07<0.0012.91
OH-C10-HSL2.5ND13.433.23
C11-HSLNDND<0.0019.11
C12-HSLNDND0.015.07
OXO-C12-HSLND0.702<0.00121.28
OH-C12-HSLNDND0.1252.39
C13-HSLNDND0.0047514.78
C14-HSLNDND0.60811.82
C14:1-HSLND0.3660.005359.41
OXO-C14-HSLND0.76<0.0016.79
OXO-C14:1-HSLNDND0.06064.71
OH-C14-HSLNDND0.49236.58
C15-HSLNDND0.09415.11
C16-HSLNDND12.0116.30
C16:1-HSLND0.10.02314.68
OXO-C16:1-HSLNDND6.286.75
C18-HSLNDNDND10.48
C18:1-HSLNDND7.828.56
Table 3. UHPLC-HRMS data and AHL identification in V. tasmaniensis LGP32. Rt: Retention time. Theoretical mass correspond to the pseudo-molecular ion [M + H]+.
Table 3. UHPLC-HRMS data and AHL identification in V. tasmaniensis LGP32. Rt: Retention time. Theoretical mass correspond to the pseudo-molecular ion [M + H]+.
FractionsRt (min)Observed MassMolecular FormulaDelta ppmIdentification
NameMolecular FormulaMolecular WeightTheoretical Mass
LGP32_l10.04298.2014C16H28NO4−0.554OXO-C12-HSLC16H27NO4297.1940298.2012
LGP32_m9.82300.2166C16H30NO42.182OH-C12-HSLC16H29NO4299.2096300.2169
LGP32_o9.90256.1914C14H26NO32.810C10-HSLC14H25NO3255.1834256.1907
LGP32_p10.50310.2383C18H32NO32.158C14:1-HSLC18H31NO3309.2303310.2376
Table 4. UHPLC-HRMS data of AHL Standards. Rt: Retention time. Theoretical mass correspond to the pseudo-molecular ion [M + H]+. MS/MS spectra for AHL standards can be found in Supplementary Information (Figure S1).
Table 4. UHPLC-HRMS data of AHL Standards. Rt: Retention time. Theoretical mass correspond to the pseudo-molecular ion [M + H]+. MS/MS spectra for AHL standards can be found in Supplementary Information (Figure S1).
AHL StandardMolecular FormulaTheorical MassObserved MassRt (min)
C4-HSLC8H13NO3172.0968172.09685.26
C6-HSLC10H17NO3200.1281200.12818.43
OXO-C6-HSLC10H15NO3214.1074214.10727.56
C7-HSLC11H19NO3214.1438214.14408.83
C8-HSLC12H21NO3228.1594228.15949.27
OXO-C8-HSLC12H19NO4242.1387242.13818.69
OH-C8-HSLC12H21NO4244.1543244.1548.55
C9-HSLC13H23NO3242.1751242.17489.57
C10-HSLC14H25NO3256.1907256.19079.90
OXO-C10-HSLC14H23NO4270.1700270.16999.43
OH-C10-HSLC14H25NO4272.1856272.18569.25
C11-HSLC15H27NO3270.2064270.206310.13
C12-HSLC16H29NO3284.2220284.222010.46
OXO-C12-HSLC16H27NO4298.2013298.201310.04
OH-C12-HSLC16H29NO4300.2169300.21699.87
C13-HSLC17H31NO3298.2377298.237710.63
C14-HSLC18H33NO3312.2533312.253310.93
C14:1-HSLC18H31NO3310.2377310.237010.51
OXO-C14:1-HSLC18H29NO4324.2169324.217010.23
OXO-C14-HSLC18H31NO4326.2326326.232210.56
OH-C14-HSLC18H33NO4328.2482328.248210.42
C15-HSLC19H35NO3326.2690326.268911.15
C16-HSLC20H37NO3340.2846340.284611.34
C16:1-HSLC20H35NO3338.2690338.270410.93
OXO-C16:1-HSLC20H33NO4352.2482352.249710.61
C18-HSLC22H42NO3368.3159368.315511.66
C18:1-HSLC22H39NO3366.3003366.300311.40

Share and Cite

MDPI and ACS Style

Girard, L.; Blanchet, É.; Intertaglia, L.; Baudart, J.; Stien, D.; Suzuki, M.; Lebaron, P.; Lami, R. Characterization of N-Acyl Homoserine Lactones in Vibrio tasmaniensis LGP32 by a Biosensor-Based UHPLC-HRMS/MS Method. Sensors 2017, 17, 906. https://doi.org/10.3390/s17040906

AMA Style

Girard L, Blanchet É, Intertaglia L, Baudart J, Stien D, Suzuki M, Lebaron P, Lami R. Characterization of N-Acyl Homoserine Lactones in Vibrio tasmaniensis LGP32 by a Biosensor-Based UHPLC-HRMS/MS Method. Sensors. 2017; 17(4):906. https://doi.org/10.3390/s17040906

Chicago/Turabian Style

Girard, Léa, Élodie Blanchet, Laurent Intertaglia, Julia Baudart, Didier Stien, Marcelino Suzuki, Philippe Lebaron, and Raphaël Lami. 2017. "Characterization of N-Acyl Homoserine Lactones in Vibrio tasmaniensis LGP32 by a Biosensor-Based UHPLC-HRMS/MS Method" Sensors 17, no. 4: 906. https://doi.org/10.3390/s17040906

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