Vous êtes sur la page 1sur 45

TABLE OF CONTENTS

TITLE..................................................................................................................... I PREFACE............................................................................................................. II JOURNAL I COMPARATIVE PROTEOMIC ANALYSIS OF EXTRACELLULAR PROTEINS OF EDWARDSIELLA TARDA ....................................................... 3 ABSTRACT ........................................................................................................ 3 DEFINING VIRULENCE .................................................................................. 4 CULTURE CONDITIONS ................................................................................. 4 1D PAGE PROFILE ANALYSIS ....................................................................... 5 2D PAGE PROFILE ANALYSIS ....................................................................... 6 MS ....................................................................................................................... 7 PROTEIN IDENTIFICATION AND EDMAN N-TERMINAL SEQUENCING. .................................................................................................. 8 GENE CLONING AND SEQUENCING. ........................................................ 12 PUTATIVE ROLES OF FLAGELLIN AND SSEB. ....................................... 13 CONCLUDING REMARKS. ........................................................................... 14 NUCLEOTIDE SEQUENCE ACCESSION NUMBERS. ............................... 15 ACKNOWLEDGMENTS ................................................................................. 15 REFERENCES ................................................................................................ 15 JOURNAL II OUTER MEMBRANE VESICLES AS A CANDIDATE VACCINE AGAINST EDWARDSIELLOSIS .................................................................... 19 ABSTRACT ...................................................................................................... 19 INTRODUCTION ............................................................................................. 20
1

RESULTS .......................................................................................................... 21 TEM Examination of OMVs ......................................................................... 21 Protein Profiling of OMVs ............................................................................ 22 Immune Responses In Olive Flounder........................................................... 26 Protective Effects of OMV Injection Into Olive Flounder ............................ 28 DISCUSSION ................................................................................................... 29 MATERIALS AND METHODS ...................................................................... 33 Bacterial Strain and Growth Conditions ........................................................ 33 OMV Preparation ........................................................................................... 33 Transmission Electron Microscopy (TEM) ................................................... 34 Preparation of Whole Cell Lysates, Periplasmic Proteins and Outer Membrane Proteins ........................................................................................ 35 Electrophoresis and In Gel Digestion ......................................................... 35 LC-ESI-MS/MS and Data Analysis ............................................................... 36 Fish and Immunization with OMVs or FKC ................................................. 37 Quantitative Real Time PCR ...................................................................... 38 Fish Vaccination and Challenge .................................................................... 39 REFERENCES ................................................................................................ 40

COMPARATIVE PROTEOMIC ANALYSIS OF EXTRACELLULAR PROTEINS OF EDWARDSIELLA TARDA Y. P. Tan1, Q. Lin1, X. H. Wang1, S. Joshi1, C. L. Hew1,2 and K. Y. Leung1,2,*

ABSTRACT A comparison of extracellular proteins of virulent and avirulent Edwardsiella tarda strains revealed several major, virulent-strain-specific proteins. Proteomic analysis identified two of the proteins in the virulent strain PPD130/91 as flagellin and SseB, which are virulence factors in bacterial pathogens. PCR amplification and DNA sequencing confirmed the presence of the genes that encode these proteins. Our results clearly demonstrated the potency of the proteomic approach in identifying virulence factors. Edwardsiella tarda is a gram-negative enteric bacterial pathogen of both animals (6) and humans (14). Fish mortality caused by E. tarda infection (28) led to severe losses in the aquaculture industry. The bacterium's virulence features include the ability to invade epithelial cells (16, 19), resistance to phagocytic killing (26), and production of enzymes such as hemolysins (5, 15) and chondroitinase (17). Functional genomic analysis is essential for a better understanding of the pathogenesis of E. tarda infections. The proteomic approach, which complements genomic research, has been employed to study the role of iron in regulating the pathophysiology of Mycobacterium tuberculosis (29) and the effect of environmental cues on gene expression within Salmonella pathogenicity island 2 (SPI2) in Salmonella enterica serovar Typhimurium (7). It was thus adopted here for the first time to compare and identify the secreted virulence factors of E. tarda.

DEFINING VIRULENCE Fourteen E. tarda strains were used for comparative proteomic analysis. The median 50% lethal doses (LD50s) of these strains were determined by using nave blue gourami, Trichogaster trichopterus (Pallas), as described previously (19). Of the 14 strains used, 6 were described as virulent (LD50 of <106.5) and 8 were described as avirulent (LD50 of >107.0) (Table 1).

CULTURE CONDITIONS Tryptic soy agar (TSA) medium (Difco) is used for growing E. tarda cultures in our laboratory. Brain heart infusion agar medium (Difco), which has also been used by others to culture E. tarda (27), was included for comparison. We first checked the extracellular protein (ECP) profiles of the two representative strains (virulent PPD130/91 and avirulent PPD125/87) cultured on TSA and brain heart infusion agar for 24 and 48 h at 25C. ECP was prepared as described by Leung and Stevenson (18), with slight modifications. The final filtered ECP obtained was desalted and concentrated with a Millipore Biomax-5K column, and the protein concentration was determined by the Bio-Rad protein assay. One4

dimensional (1D) polyacrylamide gel electrophoresis (PAGE) of the ECP was then performed according to standard procedures (22). Since the ECP production pattern was not affected by varying the culture media (data not shown), TSA was chosen for subsequent E. tarda cultures, and incubations of 24 instead of 48 h were used to avoid possible protein degradation due to a prolonged incubation period.

1D PAGE PROFILE ANALYSIS Once the culture conditions had been fixed, the ECP profiles of another five virulent and seven avirulent strains were surveyed (Fig. 1). Three virulent strains, PPD130/91, NUF251, and NE8003, exhibited very similar protein band patterns (Fig. 1A), with two unique major bands of approximately 55 and 21 kDa. The

Figure 1. Survey of the ECP profiles of virulent (A) and avirulent (B) E.

tarda strains (cultured for 24 h on TSA) with 1D PAGE (silver staining). The label at the top of each lane denotes the strain used in that lane.

remaining three virulent strains, on the other hand, showed band profiles that were rather unique to each strain but that were different from those of the first three strains. As for the avirulent strains (Fig. 1B), their protein profiles were very similar, exhibiting multiple bands but no major bands as in the case of the virulent strains. Virulent strains in general shared background band profiles similar to those of the avirulent strains except for the virulent strains' additional major bands, which may thus be virulent-strain specific.

2D PAGE PROFILE ANALYSIS To obtain better protein resolution, two-dimensional (2D) PAGE (Bio-Rad) was employed to further resolve the ECP profiles of the two representative E. tarda strains. We used the following conditions for isoelectric focusing of the protein samples: 100 V for 4 h, 500 V for 2 h, 10,000 V for 3 h, and 10,000 V for 9 h. Nine prominent extra spots (Fig. 2A, spots 1 and 2 and 4 to 10), distributed from the acidic (pH 3) to the neutral pH range, were seen in the ECP of the virulent strain but not in that of the avirulent strain (Fig. 2B). Silver-stained gels (4) showed similar background patterns in both of the representative strains (data not shown). E. tarda NUF251 and NE8003, which shared 1D-gel band profiles similar to those of PPD130/91, were also found to have 2D-gel patterns similar to those of PPD130/91 (data not shown). Protein spot 3 (boxed) appeared to be common to both representative strains. In a separate, concurrent experiment done in our laboratory, attenuated E. tarda PPD130/91 PhoA+ fusion mutants were found to have lost the extra major proteins in the ECP, indicative of these proteins' probable involvement in virulence (P. S. Srinivasa Rao and K. Y. Leung, unpublished data).

Figure 2. ECP profiles of E. tarda PPD130/91 (virulent) (A) and

PPD125/87 (avirulent) (B) on a Coomassie blue-stained 2D gel with a broad range of pHs (3 to 10). Ten major protein spots (circled spots indicate proteins unique to the virulent strain; boxed spots [including spot 3A] indicate proteins common to both strains) were selected for MS analysis. The inset in panel A is an enlarged image of protein spots 4 to 10, and the number of each spot is as shown. These observations prompted us to further examine these proteins by mass spectrometry (MS). A common protein (spot 3) which appears to be up-regulated in the virulent strain compared to that in the avirulent strain (spot 3A) was also analyzed. MS Ten protein spots of interest were excised from the 2D gel and digested with trypsin according to the procedure described by Shevchenko and coworkers (24). Mass spectra of each spot were acquired with a matrix-assisted laser desorption ionization-time of flight (MALDI-TOF) mass spectrometer (Voyager-DE STR BioSpectrometry work station; Applied Biosystems) operating in the delayedextraction reflectron mode. In addition, nanoflow electrospray ionization (ESI) tandem MS was performed for the purified tryptic digests (with Millipore Zip-Tip

C18 pipette tips) with a quadrupole TOF mass spectrometer (Q-tof-2; Micromass), and partial amino acid sequences of the peptides were obtained. The mass spectra obtained by both MALDI-TOF MS and ESI tandem MS (data not shown) revealed four basic peak patterns of the 10 spots analyzed, which categorized the spots into four groups: spots 1 and 2 in group I; spot 3 in group II; spots 5, 7, and 9 in group III; and spots 4, 6, 8, and 10 in group IV. The presence of more than one member in three of the groups and the close proximity of the spots was suggestive of probable protein isoforms. Peptide mass fingerprints (PMF) of the tryptic peptides from MALDI-TOF MS data on the representative spots (spots 1, 3, 7, and 8), together with the isoelectric points and molecular weights, were used to search the National Center for Biotechnology Information (NCBI) protein database with the programs Profound peptide mapping (ProteoMetrics) at http://129.85.19.192/profound_bin/WebProFound.exe and MSFit at http://prospector.ucsf.edu/ucsfhtml4.0/msfit.htm. The ESI tandem MS data obtained were subjected to the NCBI database search by the Mascot search engine at http://www.matrixscience.com.

PROTEIN IDENTIFICATION SEQUENCING.

AND

EDMAN

N-TERMINAL

The database search with MALDI-TOF MS PMF data on the representative spots did not yield any positive protein identifications. When the ESI tandem MS data were subjected to the database search, only protein spot 3 was identified to be the flagellin protein. Results from the Mascot search of both the NCBI and bacterial databases showed that five peptides matched the Serratia marcescens 274 flagellin (accession no. P13713) (Fig. 3), with a significant total score of 160. Two of the peptides, DDAAGQAISNR and INSAKDDAAGQAISNR, appeared to be the same except for a missed cleavage in the fifth lysine residue (K) of the latter. For the peptide ISEQTDFNGVK, the third glutamic acid residue (E) was actually a glutamine (Q) residue according to the nucleotide sequence obtained. The residue alteration (from Q to E) in the tandem MS sequence is likely due to

the occurrence of deamidation. A similar flagellin protein (spot 3A) was also identified in the ECP of avirulent E. tarda PPD125/87 by comparison of the MALDI-TOF MS PMF and by the recognition of at least 13 common flagellin peaks (data not shown). Since only one protein was identified by MS, Edman N-terminal sequencing of

Figure 3. Alignment of the E. tarda PPD130/91 (ET) and S. marcescens 274

(SM) flagellin (FLG) amino acid sequences. Lightly shaded sequences show conserved regions, and residues in bold indicate the matched peptides obtained by the Mascot database search with the ESI tandem MS data for protein spot 3. The residue (Q) highlighted with darker shading is obtained from the translation of the nucleotide sequence. Deamidation has occurred, which changed Q to E, and thus the residue matches the E residue in the S. marcescens sequence. Dashes indicate gaps in the sequence and were created for alignment. Underlined S. marcescens sequences denote regions of the matched peptides that were used for the design of primers to find the corresponding gene in E. tarda, and the arrows indicate the primer direction (5 to 3).
9

the proteins, which were either blotted onto a polyvinylidene difluoride membrane (spot 1) or purified by high-performance liquid chromatography (spots 7 and 8), was performed with a Procise model 494 pulsed-liquid-phase protein sequencer (Applied Biosystems). Nineteen, 55, and 80 N-terminal amino acids were obtained for spots 1, 7, and 8, respectively. The N-terminal sequences and/or the amino acid sequences obtained by ESI tandem MS were edited according to the rules provided by Shevchenko and coworkers (23) and subjected to a database search on the EMBL server, http://dove.embl-heidelberg.de/Blast2/, with the new MS-BLAST program for MS. Only protein spot 8 was identified to be the SseB protein, a member of the secretion system effector proteins of Salmonella serovar Typhimurium. Positive identification was claimed based on the fact that the sum of the scores for the high-scoring pair of the first two hits (110) was higher than the required confirmatory threshold score (102) (Fig. 4) (23). Protein spots 1 and 7, which did not exhibit any sequence homology to known proteins in the database, may be novel virulence-associated factors that require further characterization.

10

Figure 4. Alignment of the amino acid sequences of the E. tarda (ET) and Salmonella serovar Typhimurium (ST) SseB proteins (SSEB) and of enteropathogenic E. coli (EP) EspA protein (ESPA). Conserved regions are shaded, and the matched E. tarda partial peptides obtained by the MS-BLAST database search are given in bold. Underlined bold sequences were obtained by using both N-terminal and Q-TOF manual sequencing, while the two other matched peptides (bold) were obtained by using Q-TOF manual sequencing alone. The high-scoring pair (HSP) score for each matched peptide is given in parentheses, and the superscript number indicates the ranking of the hit. Dashes denote gaps in the sequence and were created for alignment. A degenerate primer was designed based on the N-terminal sequence (underlined) to work together with the adaptor primer to amplify the sseB gene in E. tarda. The direction of the degenerate primer (5 to 3) is indicated by the arrow.

11

GENE CLONING AND SEQUENCING. In order to confirm the presence of sseB-like and flagellin genes in E. tarda PPD130/91, PCR amplification was carried out with the Advantage 2 polymerase mix (Clontech) and the PCR fragments obtained were cloned in the pGEM-T Easy vector system (Promega) and transformed into E. coli TOP10F. For the flagellin gene, a pair of primers (GENSET; Singapore Biotech) was first designed based on the nucleotide sequences of two of the matched peptides flanking the front (5-ACAGCCTGTCTCTGATGGCG-3) and back (5CTCATGTTGGACACTTCGG-3) portions of the flagellin homologue (Fig. 3) obtained from the Mascot search results. PCR was then performed using these two primers to amplify the relevant region in the E. tarda genomic DNA, with the following cycling conditions: 25 s at 94C, seven cycles of 15 s at 94C and 1 min at 72C, 32 cycles of 15 s at 94C and 1 min at 67C, and 4 min at 67C. For amplification of the sseB-like gene in E. tarda, a degenerate primer [5AA(C/T)AC(A/C/G/T)GA(C/T)TA(C/T)CA(C/T)GG(A/C/G/T)GG-3] and the adaptor primer (Clontech) were used for PCR assay of the EcoRV genomewalking library, with the same cycling conditions. The PvuII and StuI libraries (Clontech) were also used for further genome walking to obtain the complete sequence. DNA sequencing, sequence assembly, and analysis were described previously (26). Complete sequences of the flagellin (1,251-bp) and sseB-like (597-bp) genes, made up of 416 and 198 amino acids, respectively, were obtained. Alignment of the E. tarda flagellin homologue with the S. marcescens flagellin amino acid sequences (Fig. 3) showed that the N and C termini are well conserved, and the percentage of identity was 77.8%. Unlike flagellin, the SseB-like protein in E. tarda was found to have only 34% identity to the corresponding gene in Salmonella serovar Typhimurium (GenBank accession no. AAL20322). Compared to the EspA protein of enteropathogenic E. coli (accession no. AF022236), the percentage of identity was even lower (27%). The alignment pattern of these three proteins is shown in Fig. 4.

12

PUTATIVE ROLES OF FLAGELLIN AND SSEB. The flagellin protein identified in this study was homologous to that of S. marcescens strains 274 (9) and 8000 (1). The latter strain was earlier found to secrete a 37-kDa flagellin protein into the culture medium. E. tarda may likewise secrete a similar protein. Furthermore, Hirose and coworkers (13) recently identified a flagellin protein (Hag) from culture media while analyzing the extracellular proteins of Bacillus subtilis. Although flagellin has been implicated in bacterial pathogenesis with regard to adhesion (2), motility (8), and/or in-phase variation (25), its exact role in E. tarda has not been studied. The isolation of a similar flagellin in the avirulent PPD125/87 strain may indicate an indirect role in the pathogenesis of E. tarda infections. Interestingly, a motility-deficient mutant of E. tarda, PPD130/91, that also lacks catalase production was found to be attenuated (20). In Salmonella serovar Typhimurium, the SseB protein is a secretion system effector of the type III secretion system encoded by SPI2 (11, 12). A recent study showed that this protein, together with SseC and SseD, assembled into a translocon complex on the bacterial cell surface (21) to mediate other SPI2 effector protein translocations. Upon induction by acidic pH, SseB was rapidly secreted onto the bacterial cell surface (3). Due to its apparent surface localization, it appeared to be fairly prone to mechanical shearing, which allowed it to be isolated in the extracellular milieu (3, 21). In E. tarda, the secretion of this protein may not be pH dependent, as in the case of Salmonella serovar Typhimurium, since the cells were not subjected to pH changes prior to the isolation of ECP. However, subjection to mechanical forces such as washing and filtering in the ECP preparation procedure may have resulted in its release into the ECP supernatant if it is also surface localized as in Salmonella serovar Typhimurium.

13

Salmonella serovar Typhimurium SPI2 is important for systemic infection, intracellular survival, and replication (10). A mutation in sseB led to the attenuation of serovar Typhimurium cells and failure of the cells to accumulate within macrophages (12). Since E. tarda is biochemically similar to Salmonella (14) and it is known to survive and replicate within macrophages (26), it may possess an SPI2-like pathogenicity island involved in virulence, as indicated by the identification of the SseB-like protein. The isolation of this protein in the ECP may also be suggestive of a translocon role similar to that of SseB in serovar Typhimurium. Gene knockout experiments are essential for establishing the protein's exact in vivo function. Southern blot analysis performed to survey the distribution of the sseB-like gene in virulent and avirulent E. tarda strains showed that it is present in all six virulent strains and only one avirulent strain (ET82015) used in this study (data not shown). This survey thus suggests the probable involvement of this gene in E. tarda virulence.

CONCLUDING REMARKS. Comparison of ECP profiles from representative virulent and avirulent E. tarda strains led to the identification of two potential extracellular virulence-associated proteins out of the four categories of proteins isolated. The precise involvement of these two proteins will have to be established by gene knockout experiments in later studies. Despite the lack of genome information, the work done here has clearly demonstrated the effectiveness of the comparative proteomic approach in the identification of important virulence factors. The understanding of the pathogenesis of E. tarda infections will be enhanced, and the knowledge gained in the present study will facilitate the identification of targets for the development of therapies against infections caused by this bacterium.

14

NUCLEOTIDE SEQUENCE ACCESSION NUMBERS. GenBank accession numbers for the sequences of the flagellin and sseB-like genes are AF487406 and AF498017, respectively.

ACKNOWLEDGMENTS We are grateful to the National University of Singapore for providing the research grant for this work. We thank John Grizzle from Auburn University, Auburn, Ala.; H. Wakabayashi from the University of Tokyo, Tokyo, Japan; and T. T. Ngiam and H. Loh from the Agri-food and Veterinary Authority (AVA) of Singapore for providing us with E. tarda strains from the United States, Japan, and Singapore, respectively.

REFERENCES 1. Akatsuka, H., E. Kawai, K. Omori, and T. Shibatani. 1995. Divergence of a flagellin protein in Serratia marcescens. Gene 163:157-158. 2. Attridge, S. R., and D. Rowley. 1983. The role of flagellum in the adherence of Vibrio cholerae. J. Infect. Dis. 147: 864-872. 3. Beuzon, C. R., G. Banks, J. Deiwick, M. Hensel, and D. W. Holden. 1999. pH-dependent secretion of SseB, a product of the SPI-2 type III secretion system of Salmonella typhimurium. Mol. Microbiol. 33:806-816. 4. Blum, H., H. Beier, and H. J. Gross. 1987. Improved silver staining of plant proteins, RNA and DNA in polyacrylamide gels. Electrophoresis 8:93-99.

15

5. Chen, J. D., and S. L. Huang. 1996. Hemolysin from Edwardsiella tarda strain ET16 isolated from eel Anguilla japonica identified as a holeforming toxin. Fish. Sci. 62:538-542. 6. Cook, R. A., and J. P. Tappe. 1985. Chronic enteritis associated with Edwardsiella tarda infection in Rockhopper penguins. J. Am. Vet. Med. Assoc. 187:1219-1220. 7. Deiwick, J., and M. Hensel. 1999. Regulation of virulence genes by environmental signals in Salmonella typhimurium. Electrophoresis 20:813-817. 8. Drake, D., and T. C. Montie. 1988. Flagella, motility, and invasive virulence of Pseudomonas aeruginosa. J. Gen. Microbiol. 134:43-52. 9. Harshey, R. M., G. Estepa, and H. Yanagi. 1989. Cloning and nucleotide sequence of a flagellin-coding gene (hag) from Serratia marcescens 274. Gene 79:1-8. 10. Hensel, M. 2000. Salmonella pathogenicity island 2. Mol. Microbiol. 36:1015-1023. 11. Hensel, M., J. E. Shea, A. J. Bumler, C. Gleeson, F. Blattner, and D. W. Holden. 1997. Analysis of the boundaries of Salmonella pathogenicity island 2 and the corresponding chromosomal region of Escherichia coli K-12. J. Bacteriol. 179:1105-1111. 12. Hensel, M., J. E. Shea, S. R. Waterman, R. Mundy, T. Nikolaus, G. Banks, A. Vazquez-Torres, C. Gleeson, F. C. Fang, and D. W. Holden. 1998. Genes encoding putative effector proteins of the type III secretion system of Salmonella pathogenicity island 2 are required for bacterial virulence and proliferation in macrophages. Mol. Microbiol. 30:163-174. 13. Hirose, I., K. Sano, I. Shioda, M. Kumano, K. Nakamura, and K. Yamane. 2000. Proteome analysis of Bacillus subtilis extracellular proteins: a two-dimensional protein electrophoretic study. Microbiology 146:65-75. 14. Janda, J. M., and S. L. Abbott. 1993. Infections associated with the genus Edwardsiella: the role of Edwardsiella tarda in human disease. Clin. Infect. Dis. 17: 742-748.

16

15. Janda, J. M., and S. L. Abbott. 1993. Expression of an iron-regulated hemolysin from Edwardsiella tarda. FEMS Microbiol. Lett. 111:275-280. 16. Janda, J. M., S. L. Abbott, and L. S. Oshiro. 1991. Penetration and replication of Edwardsiella spp. in HEp-2 cells. Infect. Immun. 59:154161. 17. Janda, J. M., S. L. Abbott, S. Kroske-Bystrom, W. K. W. Cheung, C. Powers, R. P. Kokka, and K. Tamura. 1991. Pathogenic properties of Edwardsiella species. J. Clin. Microbiol. 29:1997-2001. 18. Leung, K. Y., and R. M. W. Stevenson. 1988. Characteristics and distribution of extracellular proteases from Aeromonas hydrophila. J. Gen. Microbiol. 134:151-160. 19. Ling, S. H. M., X. H. Wang, L. Xie, T. M. Lim, and K. Y. Leung. 2000. Use of green fluorescent protein (GFP) to track the invasion pathways of Edwardsiella tarda in in vivo and in vitro fish models. Microbiology 146:7-19. 20. Mathew, J. A., Y. P. Tan, P. S. Srinivasa Rao, T. M. Lim, and K. Y. Leung. 2001. Edwardsiella tarda mutants defective in siderophore production, motility, serum resistance and catalase activity. Microbiology 147:449-457. 21. Nikolaus, T., J. Deiwick, C. Rappl, J. A. Freeman, W. Schrder, S. I. Miller, and M. Hensel. 2001. SseBCD proteins are secreted by the type III secretion system of Salmonella pathogenicity island 2 and function as a translocon. J. Bacteriol. 183:6036-6045. 22. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, N.Y. 23. Shevchenko, A., S. Sunyaev, A. Loboda, A. Shevchenko, P. Bork, W. Ens, and K. G. Standing. 2001. Charting the proteomes of organisms with unsequenced genomes by MALDI-quadrupole time-of-flight mass

spectrometry and BLAST homology searching. Anal. Chem. 73:19171926.

17

24. Shevchenko, A., M. Wilm, O. Vorm, and M. Mann. 1996. Mass spectrometric sequencing of proteins from silver-stained polyacrylamide gels. Anal. Chem. 68:850-858. 25. Silverman, M. J., and M. Simon. 1980. Phase variation: genetic analysis of switching mutants. Cell 19:845-854. 26. Srinivasa Rao, P. S., T. M. Lim, and K. Y. Leung. 2001. Opsonized virulent Edwardsiella tarda strains are able to adhere to and survive and replicate within fish phagocytes but fail to stimulate reactive oxygen intermediates. Infect. Immun. 69:5689-5697. 27. Strauss, E. J., N. Ghori, and S. Falkow. 1997. An Edwardsiella tarda strain containing a mutation in a gene with homology to shlB and hpmB is defective for entry into epithelial cells in culture. Infect. Immun. 65:39243932. 28. Thune, R. L., L. A. Stanley, and R. K. Cooper. 1993. Pathogenesis of gram-negative bacterial infections in warm water fish. Annu. Rev. Fish Dis. 3:37-68. 29. Wong, D. K., B.-Y. Lee, M. A. Horwitz, and B. W. Gibson. 1999. Identification of Fur, aconitase, and other proteins expressed by Mycobacterium tuberculosis under conditions of low and high

concentrations of iron by combined two-dimensional gel electrophoresis and mass spectrometry. Infect. Immun. 67:327-336. 30. http://iai.asm.org/content/70/11/6475.full

18

PLoS One. 2011; 6(3): e17629. Published online 2011 March 8. PMCID: PMC3050902 OUTER MEMBRANE VESICLES AS A CANDIDATE VACCINE AGAINST EDWARDSIELLOSIS Seong Bin Park, Ho Bin Jang, Seong Won Nho, In Seok Cha, Jun-ichi Hikima, Maki Ohtani, Takashi Aoki,* and Tae Sung Jung* doi: 10.1371/journal.pone.0017629

ABSTRACT Infection with Edwardsiella tarda, a Gram-negative bacterium, causes high morbidity and mortality in both marine and freshwater fish. Outer membrane vesicles (OMVs) released from Gram-negative bacteria are known to play important roles in bacterial pathogenesis and host immune responses, but no such roles for E. tarda OMVs have yet been described. In the present study, we investigated the proteomic composition of OMVs and the immunostimulatory effect of OMVs in a natural host, as well as the efficacy of OMVs when used as a vaccine against E. tarda infection. A total of 74 proteins, from diverse subcellular fractions, were identified in OMVs. These included a variety of important virulence factors, such as hemolysin, OmpA, porin, GAPDH, EseB, EseC, EseD, EvpC, EvpP, lipoprotein, flagellin, and fimbrial protein. When OMVs were administrated to olive flounder, significant induction of mRNAs encoding IL 1, IL 6, TNF, and IFN was observed, compared with the levels seen in fish injected with formalin killed E. tarda. In a vaccine trial, olive flounder given OMVs were more effectively protected (p<0.0001) than were control fish. Investigation of OMVs may be useful not only for understanding the pathogenesis of E. tarda but also in development of an effective vaccine against edwardsiellosis.

19

INTRODUCTION Outer membrane vesicles (OMVs) are spherical blebs of average diameter 10300 nm that are naturally released from Gram negative bacteria into the environment [1]. Although the budding mechanisms are unclear, it has been shown that OMVs are continuously produced during growth of various Gram negative bacteria including Escherichia coli, Helicobacter pylori, Neisseria meningitidis, Pseudoaltermonas antarctica, Pseudomonas aeruginosa, Shigella flexneri, and Vibrio cholerae [2] [7]. Such vesicles are known to contain lipopolysaccharide (LPS), lipoproteins, outer membrane, periplasmic, and cytoplasmic proteins, DNA, and RNA [1], [8] [10], and have been suggested to be involved in exclusion of competing bacteria, conveyance of proteins or genetic material to other bacteria, and presentation of virulence factors to the host [1]. Edwardsiella tarda is the causative agent of edwardsiellosis in a variety of cultured freshwater and marine fish, including channel catfish Ictalurus punctatus, olive flounder Paralichthys olivaceus, Japanese eel Anguilla japonica, red sea bream Pagrus major, mullet Mugil cephalus, and turbot Scophthalmus maximus [11] [16]. Edwardsiellosis has been implicated in the mass mortality of olive flounder, which is the main mariculture species of South Korea. Typical clinical symptoms of E. tarda infection in olive flounder are exophthalmia, enlargement of the spleen, malodorous ascites, and rectal hernia [17]. A number of studies have shown that vaccination using outer membrane proteins results in development of protective effects against E. tarda infection [18] [21]. In addition, the outer membrane proteins of E. tarda include several important virulence factors that play key roles in pathogenicity [17]. Virulence factors of E. tarda that have been investigated include dermatotoxin, hemolysins, catalase, outer membrane proteins, EseDs, and glyceraldehyde 3 phosphate dehydrogenase (GAPDH) [19], [22] [24].

20

Previous proteomic studies indicated that both outer membrane proteins and LPS in OMVs might play roles as pathogen-associated molecular patterns (PAMPs) delivered to the host innate immune system, and could thus elicit immune responses [25], [26]. Because OMVs have antigenic properties, such vesicles have been investigated as useful candidate vaccines against Gram negative bacterial infections [27] [29]. For example, a Neisseria meningitidis serogroup B vaccine was successfully developed using derived OMVs; 55 million doses have been administered to date [30]. However, none of OMV protein composition, antigenicity, or vaccine efficacy has been studied in bacteria pathogenic for fish, especially the bacterium E. tarda. Thus, we investigated the possibility of using OMVs released from E. tarda as a vaccine against edwardsiellosis, based on both a proteomic study and cytokine induction assays.

RESULTS TEM Examination of OMVs Numerous ovoid to round shaped blebs were evident on the surface of ED45 cells when thin sections were stained to show OMVs (Figure 1 - A). The supernatant concentrate contained OMVs that were round in shape, ranged from 10 40 nm in diameter, and contained electron dense substances (Figure 1 - B). Because cell debris and pili were observed upon negative staining, OMVs suspended in PBS were purified by discontinuous sucrose gradient centrifugation prior to protein analysis and in vivo immunogenicity testing. After centrifugation, two clear white bands were evident in the centrifugation tube, and the materials therein were examined by TEM (Figure 1 - C).

21

Figure 5. Transmission electron micrographs (TEMs) of outer membrane vesicles (OMVs) released from Edwardsiella tarda ED45. (A) Thin-section TEM of OMVs released from an ED45 cell. Bar = 50 nm. (B) Negatively stained TEM of concentrated OMVs of ED45. Bar = 50 nm. (C) OMVs purified on a sucrose density gradient. The lower fraction is composed of OMVs. doi:10.1371/journal.pone.0017629.g001 The upper band appeared to contain cell debris or aggregates, whereas the lower band was mainly OMVs, at a density of 1.185 g/ml. All further experiments were performed using these purified OMVs.

Protein Profiling of OMVs Purified OMVs were loaded onto 1D SDS-PAGE gels and proteins were visualized, after electrophoresis, using a silver staining method, and compared with proteins in WCL, PP, and OMP fractions. OMV proteins of molecular size 68, 31.5, 30, 24, 19, 14.5, and 14 kDa (Figure 2) were similar to those in the OMP fraction, but the 17 -, 37 -, and 54 kDa bands of OMVs were absent from the OMP sample. To identify the protein components of OMVs, the proteins were separated by 12.5% (w/v) SDS PAGE and the gels were cut into 12 slices. Proteins in individual slices were analyzed by LC ESI MS/MS; acquired peptide mass data were analyzed using the MASCOT Daemon interface.

22

A total of 74 proteins were identified in the E. tarda database (Table 1), and each slice contained at least 5 of these proteins (data not shown). Proteins were categorized into 15 different orthologous groups using the COG approach (Figure 3), indicating that the identified proteins were involved in both cellular processing and signaling (COG groups M, O, and N; 33% of proteins); information storage and processing (COG groups L, K, and J; 25% of proteins), and metabolism (COG groups I, H, G, F, E, and C; 21% of proteins). In total, 5% of proteins fell into poorly characterized categories (COG groups S and R), whereas 16% could not be identified in any COG grouping. The subcellular locations of the identified proteins were predicted using the PSORTb algorithm; this exercise suggested that 37 proteins were localized in the cytoplasmic space, 6 in the inner membrane, 16 in the outer membrane, and 9 in the extracellular space. The locations of six proteins could not be identified.

Figure 6. SDS-PAGE protein profiles of WCL, PPs, OMPs, and OMVs. Arrowheads indicate OMV polypeptides with molecular weights the same as those of OMPs. M: protein marker lane; WCL: whole cell lysate, PP: periplasmic proteins; OMPs: outer membrane proteins; OMVs: outer membrane vesicles. doi:10.1371/journal.pone.0017629.g002

23

24

25

Figure 7. Functional classification of OMVs according to COG functional categories. The pie chart shows the numbers and percentages of identified proteins in each COG grouping. Individual protein assignment to COGs is shown in Table 1. doi:10.1371/journal.pone.0017629.g003

Immune Responses In Olive Flounder Based on the proteomic data, QRT PCR was conducted to measure mRNA expression levels of IL - 1, IL 6, TNF, and IFN in olive flounder injected with FKC or OMVs, to investigate whether OMVs could elicit host immune responses. After intraperitoneal injection of OMVs or FKC, kidneys of injected fish were sampled at 0, 3, 7, and 12 hpi; and 1 and 5 dpi (Figure 4). Fish injected with OMVs showed increased cytokine levels compared with those measured at 0 hours. IL - 1 and IL 6 expression levels were induced 320 and 515 fold at 3 hpi, and these levels were maintained to 5 dpi. TNF and IFN synthesis levels increased 4.8- and 7.6-fold at 3 hpi, compared with zero time measurements.

26

Fish injected with FKC showed differences in IL - 1 and IL 6 expression levels (compared with controls) at early time points, but neither the TNF nor IFN synthesis level was distinct from that of 0-hour control values. When expression levels of IL - 1, IL 6, TNF, and IFN in fish injected with OMVs were compared with those in fish injected with FKC, the cytokine levels of OMV injected fish

Figure 8. Relative induction of IL-1b (A), IL-6 (B), TNFa (C), and IFNc (D) in olive flounder injected with OMVs or FKC, estimated using quantitative real-time PCR. OMV: group injected with outer membrane vesicles; FKC: group injected with formalin-killed ED45. *P,0.05, **P,0.01, and ***P,0.001. Bars indicate standard deviations. N= 4. doi:10.1371/journal.pone.0017629.g004

were significantly higher, at early time points, than in fish injected with FKC. The level of IL - 1 expression in fish injected with OMVs differed significantly from that in fish injected with FKC, at 3, 7, and 12 hpi, whereas IL 6 levels were significantly different at 3 and 7 hpi. TNF and IFN expression levels in fish injected with OMVs differed, at 3 hpi, from those in fish injected with FKC.

27

mRNA TLR22 (toll like receptor 22) and TLR2 expression levels, which are pattern recognition receptors (PRRs) in olive flounder, were measured to determine whether the innate immune response could be elicited by OMVs [31]. TLR22 levels in fish injected with OMVs were significantly higher than in fish injected with FKC. However, TLR2 levels did not differ (Figure 5).

Figure 9. Relative induction of TLR22 (A) and TLR2 (B) in olive flounder injected with OMVs or FKC, estimated using quantitative real-time PCR. OMV: group injected with outer membrane vesicles; FKC: group injected with formalin-killed ED45. **P,0.01. Bars indicate standard deviations. N= 4. doi:10.1371/journal.pone.0017629.g005

Protective Effects of OMV Injection Into Olive Flounder OMVs were administered to olive flounder as a candidate vaccine against E. tarda infection; control fish were injected with PBS (Figure 6). Fish injected with FKC were used as positive controls. After acclimatization of fish for 28 days, 1.1104 CFU/ml ED45 was used as an intraperitoneal challenge, and survival rates were recorded over 31 days. Control fish died from 6 dpi and all fish were dead at 11 dpi. OMV vaccinated fish showed some mortality at 5 dpi; however, 70% of fish survived until 31 dpi. FKC-injected fish had a survival rate of 65%. OMV

28

showed an RPS of 70%; this differed significantly from that of control (PBS) (p<0.0001).

Figure 10. Survival rates of olive flounder challenged with ED45 four weeks after immunization. Control: PBS-injected group; FKC: group injected with formalin killed ED45; OMV: group injected with outer membrane vesicles. doi:10.1371/journal.pone.0017629.g006

DISCUSSION Several Gram-negative bacteria produce OMVs [1]. In the present study, E. tarda was also shown to produce round OMVs, 10 40 nm in diameter, when grown in a liquid medium. Preparation of pure OMVs from bacterial supernatants is essential when studying the roles played by OMVs in both bacteria and their hosts [9], [10]. Recently, pure OMVs from bacterial supernatants have been obtained by filtration followed by density gradient ultracentrifugation [8], [25]. In the present study, homogenously sized E. tarda OMVs were separated from contaminants using such methods. Flagella, pili, and aggregated proteins were present in OMV samples before sucrose density gradient purification, but were removed during this step.

29

Two representative proteomic analysis methods have been suggested for the study of OMVs. Of these, the first is two-dimensional electrophoresis (2 DE) followed by matrix-associated laser desorption/ionization time-of-flight (MALDI-TOF) spectrometry, and the other is 1D SDS-PAGE followed by LC-ESI-MS/MS [10], [25], [26]. 2-DE is most commonly used to establish proteomic maps of bacteria, but limitations of the technique include poor separation of high molecular weight and hydrophobic proteins [32]. However, 1D SDS PAGE in combination with LC ESI MS/MS was effective in analysis of hydrophobic outer membrane proteins contained in OMVs [6]. In the present work, a total of 74 proteins were identified, 16 of which were characteristic in the outer membrane, as indicated by the PSORTb algorithm. However, a proteomic survey of the E. tarda outer membrane identified only 21 proteins by 2 DE analysis, and only 1 protein was confined to this membrane [33]. Therefore, 1D SDS PAGE coupled with LC ESI MS/MS is more sensitive when used to analyze proteins of outer membranes or OMVs. Some reports have suggested that OMV proteomes consist mainly of outer membrane and periplasmic proteins [34], but other proteomic studies found that OMVs contained proteins of various origin, including all of cytoplasmic, inner membrane, outer membrane, and periplasmic proteins [8], [25], [35]. By 1D SDSPAGE, a 17 kDa band evident in OMVs could not be observed in outer membrane or periplasmic protein fractions. The PSORTb algorithm also indicated that proteins of OMVs originated not only from the outer membrane but also from cytoplasmic and extracellular compartments. These findings, together with those of previous proteomic studies, indicate that the OMV proteome includes proteins varying in subcellular origin, and other components. LC ESI MS/MS analysis of OMVs showed that several proteins involved in pathogenesis, including hemolysin, outer membrane protein A (OmpA), GAPDH, EseB, EseC, EseD, EvpC, EvpP, lipoprotein, flagellin, and fimbrial protein, were present.

30

Hemolysin is known to be required for cellular invasion and cytotoxicity [22]; hemolysin-negative E. tarda mutants could not invade human epithelial cell lines [36]. An ompA-negative E. coli strain invaded brain microvascular endothelial cells only with difficulty [37], [38]. GAPDH in the outer membrane is highly antigenic, and is considered to be a strong vaccine candidate to counter not only Gram-negative but also Gram-positive bacterial infections [19], [39]. EseB, EseC, EseD, EvpC, and EvpP are essential for E. tarda pathogenesis; these proteins are located on the cell surface and are responsible for pore formation [21], [40], [41]. Thus, OMVs released from E. tarda contain numerous virulence factors that may be important both for bacterial survival and to enhance immunogenicity in the natural host. QRT PCR demonstrated that the genes encoding all of IL - 1, IL 6, TNF, and IFN were induced in fish injected with OMVs of E. tarda, compared with fish injected with FKC, especially at early post-injection time points. Similarly, mice injected with OMVs of Bordetella pertussis showed upregulation of mRNAs encoding IL 6 and TNF, compared with the levels seen in animals injected with formalin killed B. pertussis [27]. It is known that the primary host immune response is mediated by the proinflammatory cytokines IL - 1, IL 6, and TNF [42]. Expression of such cytokines following injection of E. tarda OMVs indicates that the OMVs may initiate a proinflammatory cytokine cascade, including both recruitment and activation of macrophages and stimulation of an adaptive immune response. IFN, a Th1 cytokine, is also upregulated in fish injected with OMVs, and is known to enhance cell mediated immunity and antigen presentation to macrophages [42]. Such observations may indicate that OMVs trigger an elevated host immune response, thus provoking host adaptive immunity, even though protein levels in OMVs are lower than those in FKC.

31

Innate immunity is an important first line of defense against invading pathogens, and recognition of PAMPs is achieved principally by PRRs [43]. In the present study, virulence factors, including lipoprotein, flagellin, and peptidoglycan, contained within OMVs of E. tarda, may have played roles as PAMPs, thus interacting with the innate immune system. In addition, OMVs are known to possess LPS and bacterial DNA that function as ligands for host PRRs [44], [45]. OMVs have been reported to be recognized by TLR2 and TLR9, both of which participate in IL 6 production via myeloid differentiation factor 88 (MyD88) dependent pathways [45]. In the present study, TLR2 was not induced after injection with OMVs (compared with what was seen in positive control fish), whereas, in contrast, the TLR22 level was significantly increased. Therefore, the PRRs of olive flounder also participate in recognition of OMVs but the teleost detection mechanism for OMVs may differ from that of mammals. The proteomic data, and the demonstrated immunostimulatory effects of OMVs from E. tarda, showed that OMVs exhibited high protective efficacy against E. tarda infection of olive flounder, with an RPS of 70% (p<0.0001). Other reports have described successful inhibition of E. tarda infection in laboratory experiments using various antigen preparations including formalin-killed and ghost cells, recombinant proteins, and outer membrane fractions [18][21]. However, these advances have not been translated into successful commercial vaccines. Although OMVs were no more effective as a vaccine than was FKC, vaccines of the latter type sometimes function poorly [46]. The OMV work of the present study offers a new insight into vaccine formulation. Thus, an acellular vaccine, or the use of acellular materials as a vaccine adjuvant, may be helpful in development of an effective vaccine protecting against edwardsiellosis in fish. In conclusion, OMVs na turally released from E. tarda contained various important virulence factors originating from diverse subcellular locations. Such vesicles were able to induce synthesis of several proinflammatory cytokines, and may stimulate the host innate immune system, thus serving as PAMPs.

32

To demonstrate vaccine efficacy, we injected olive flounder with OMVs, and found that such fish were protected to a significantly higher extent than were control fish. The present study on OMVs may lead to development of more effective vaccines, and may enhance our understanding of the roles of OMVs in hosts.

MATERIALS AND METHODS Bacterial Strain and Growth Conditions ED45, a virulent E. tarda strain isolated from the spleen of infected olive flounder, was used in the present study; the LD50 value was 1.2103 CFU/ml (data not shown). ED45 was cultured on Tryptone Soya Agar (TSA; Oxoid, Hampshire, England) or in Tryptone Soya Broth (TSB; Oxoid), supplemented with 2% (w/v) NaCl (termed the TSA 2 and TSB 2 media, respectively). ED45 was grown on TSA 2 at 25C for 36 h; a single colony was inoculated into 25 ml TSB 2, and the culture grown with shaking to a cell density of 1109 CFU/ml. Next, 20 ml of this culture was added to 4 l TSB 2 and growth proceeded in a shaking incubator at 25C for 15 h. After centrifugation at 5,000g for 20 min, the supernatant was used for isolation of OMVs.

OMV Preparation OMVs were collected and purified from the supernatant as described previously, with several modifications [25], [47]. The supernatant containing OMVs were filtered through a 0.45 m pore sized hollow fiber cartridge and concentrated using a hollow cartridge 100 kDa in size cutoff, employing a Quixstand benchtop system (GE Healthcare, Uppsala, Sweden).

33

The concentrated supernatant was filtered through a 0.2 m pore sized vacuum filter to eliminate all remaining ED45 cells. Subsequently, OMVs were pelleted by ultracentrifugation at 150,000g for 3 h at 4C in an XL 90 ultracentrifuge running a 70Ti rotor (Beckman Coulter, Palo Alto, CA). Pelleted OMVs were resuspended in phosphate buffered saline (PBS; pH 7.2), layered onto a 30 60% (w/v) discontinuous sucrose gradient, and centrifuged at 200,000g for 20 h at 4C in a XL-90 ultracentrifuge running an SW41 Ti rotor, to obtain purified OMVs. Two clear bands were visualized, and the sucrose density in the vicinity of each band was measured by refractometry (Figure 1). After washing material from both bands (collected separately) with a 20 fold dilution of PBS, the materials were centrifuged at 150,000g for 3 h at 4C in a XL 90 ultracentrifuge. The final pellets were resuspended in PBS and protein concentrations were determined using a non-interfering assay (G Biosciences, St. Louis, MO). The final preparations were stored at 80C prior to use.

Transmission Electron Microscopy (TEM) ED45 cells were grown in TSB-2 at 25C until an OD600 value of 1.0 was attained, centrifuged at 5,000g for 30 min, and washed three times with PBS prior to preparation of ultrathin sections. After centrifugation at 5,000g for 20 min, pellets were pre-fixed in 2% (v/v) paraformaldehyde for 3 h, washed, and fixed in 4% (w/v) osmium tetroxide for 2 h. After washing with PBS, the fixed pellets were dehydrated in a series of 70 100% (v/v) ethanol baths and embedded in epoxy resin. Sections were cut using a diamond blade (Diatome, Biel, Switzerland) fitted to an ULTRACUT UCT (Leica, Vienna, Austria), mounted on carbon-coated copper grids, and stained with 3% (w/v) uranyl acetate and lead citrate. To visualize OMVs after negative staining, samples were placed on 400- mesh carbon coated grids for 2 min, washed with deionized sterile water (dd H20), and negatively stained with 3% (w/v) uranyl acetate for 30 sec. All TEM images were acquired using a Technai 12 (FEI, Hillsboro, OR) operating at an acceleration voltage of 120 kV.
34

Preparation of Whole Cell Lysates, Periplasmic Proteins and Outer Membrane Proteins Periplasmic proteins (PPs) and outer membrane proteins (OMPs) were purified as described previously [48]. Whole cell lysates (WCLs) of ED45 were obtained from cells grown in TSB 2 at 25C, pelleted at 5,000g for 30 min, and washed with PBS. Pelleted cell density was adjusted to 4 g/ml in 20% (w/v) sucrose dissolved in 20 mM Tris HCl (pH 8.0), also with 0.1 M EDTA and lysozyme (600 g/g cells). After incubation on ice for 40 min, 0.5 M MgCl 2 (0.16 ml/g cells) was added and spheroplasts were removed by centrifugation (9,500g for 20 min). The supernatant containing PPs was stored at 80C until use. Spheroplasts were resuspended in ice cold 10 mM Tris-HCl (pH 8.0) and sonicated for purification of OMPs. Following centrifugation at 8,000g for 5 min to remove cells, the supernatant was centrifuged at 40,000g for 1 h to pellet cell membrane material. Pellets were washed in 10 mM Tris HCl (pH 8.0), resuspended in dd H20, and freeze thawed. The membranes were incubated in 0.5% (w/v) Sarkosyl (sodium N-lauroylsarcosinate; Sigma, St. Louis, MO) at 25C for 20 min and OMP pellets were purified by centrifugation at 40,000g for 1 h. Purified OMPs were resuspended in 10 mM Tris-HCl (pH 8.0) and stored at 80C until use. Protein concentration was determined using a non-interfering protein assay (G-Biosciences, Hercules, CA).

Electrophoresis and In Gel Digestion SDS-PAGE was performed using a 15% (w/v) acrylamide separating gel, according to the method of Laemmli [49]. Briefly, each resuspended WCL, PP, OMP, and OMV protein sample was mixed with 5sample buffer (5 1 v/w ratio of buffer to sample). The buffer contained 60 mM Tris-HCl, 25% (v/v) glycerol, 2% (w/v) SDS, 14.4 mM mercaptoethanol, and 0.1% (w/v) bromophenol blue. Samples were boiled for 10 min, cooled on ice, and centrifuged at 16,000g for 20 min. Supernatants were subjected to SDS PAGE analysis.

35

Five microgram amounts of WCL, PP, OMP, and OMV proteins were loaded and proteins were detected by silver staining [50]. To achieve in gel digestion, 20 g amounts of OMVs were electrophoresed on a 12.5% (w/v) separating gel and stained with Bio safe Coomassie G 250 (BioRad, Hercules, CA). After cutting the gel into 12 slices, each slice was destained with a 500 l amount of 40% (v/v) ethanol in 75 mM ammonium bicarbonate (ABC) and treated with DTT (0.0039 g dithiothreitol in 5 ml 25 mM ABC) and IAA (0.0509 g iodoacetamide in 5 ml 25mM ABC) solutions. Each gel slice was dehydrated using 300 l acetonitrile (ACN) for 30 min at 37C, dried, and used for in-gel digestion employing 20 ng/ml of sequencing grade modified trypsin (Promega, Madison, WI) at 37C overnight. Tryptic peptides were extracted into 30 l 0.1% (v/v) formic acid and the solutions were sonicated for 10 min.

LC-ESI-MS/MS and Data Analysis LC-ESI-MS/MS (liquid chromatography electrospray ionization tandem mass spectrometry) analysis was carried out using a Thermo Finnigan Proteome X workstation with an LTQ linear ion trap and a MS apparatus equipped with NSI sources (Thermo Electron, San Jose, CA). Twelve microliter amounts of peptide mixtures were injected and loaded onto peptide trap cartridges (Agilent, Palo Alto, CA). Trapped peptides were eluted onto a 10 cm long reverse-phase PicoFrit column packed in house with 5 m C18 resin of pore size 300 , and the peptides were next separated on an RP column using gradient elution. The mobile phase solutions were H2O and ACN, both containing 0.1% (v/v) formic acid, and delivered at a constant flow rate of 0.2 l/min. The gradient commenced with 2.0% (v/v) ACN, rising linearly to 60% (v/v) ACN over 50 min, next increasing to 80% (v/v) ACN over the next 5 min, with a change to 100% H2O for the final 15 min. A data-dependent acquisition mode (m/z 300 1,800) was enabled, and each survey MS scan was followed by five MS/MS scans with the 30 sec dynamic exclusion option set. The spray voltage was 1.9 kV and the ion transfer tube temperature 195C. The normalized collision energy was set to 35%.
36

The mzData file of tandem mass spectra was used to search the E. tarda database of NCBI downloaded on 9 July 2010, using the Mascot Deamon interface (Version 2.2.2, Matrix Science Inc., London, UK) supported by the Korea Basic Science Institute (KBSI; Yusung gu, Daejeon, Korea). The peptide tolerance of the parent ion was adjusted to be 1 Da and the MS/MS tolerance was 0.8 Da. During analysis, carbamidomethyl (C) modification was fixed whereas oxidation modification (M) was not. One missed cleavage was permitted, and peptide charges were set at 2+ and 3+. Individual ion scores of more than 15 (p<0.05) were considered reliable and were subsequently used. Functional annotations of OMV proteins were performed using the clusters of orthologous groups (COGs) functional classification (http://www.ncbi.nih.gov/COG) [51]. The subcellular localization of identified proteins was predicted using PSORTb version 3.0 (http://www.psort.org) [52].

Fish and Immunization with OMVs or FKC Nave olive flounders (with an approximate weight of 55 g) were purchased from a commercial fish farm in Korea and acclimatized for 2 weeks at 21C in aerated seawater. Formalin killed E. tarda cells (FKC) were prepared by addition of 1% (v/v) formalin to an ED45 culture, followed by adjustment to an OD600 value of 1.0 using PBS. OMVs were prepared as described above and used after confirmation that no colonies formed when OMVs were plated onto TSA 2. Aliquots (0.1 ml) of FKC (1.2108 CFU/ml, 344 g), or 10 g amounts of OMVs, were intraperitoneally injected into olive flounder. The administred dosage was determined based on previous studies on immunization with OMVs originating from several gram negative bacteria [25], [27], [53], [54]. To evaluate expression of immune response-related molecules in fish, the kidneys of immunized olive flounder were dissected 0, 3, 7, and 12 hours post-infection (hpi), and 1.5 days post-infection (dpi), and stored in RNA-later solution (Ambion, Austin, TX) until all samples were available (N = 4).

37

All experimental protocols were approved by the Institutional Animal Care and Use Committee (IACUC) at Gyeongsang National University, Jinju, Republic of Korea (Approval Number: GNU LA - 17).

Quantitative Real Time PCR Total RNA was isolated from kidney samples using the TRIzol reagent (Invitrogen, Scotland, UK) according to the manufacturer's protocol. After purification of 1 g aliquots of mRNA using deoxyribonuclease I (Invitrogen), cDNAs were synthesized with the assistance of a high capacity cDNA reverse transcriptase kit (Applied Biosystems, Foster city, CA) according to the manufacturer's protocol. cDNA samples were diluted 1 4 (v/v) into nuclease free water prior to amplification using quantitative real-time PCR (QRT - PCR). Primers were designed employing the Primer Express software of Real Time PCR version 3.0 (Applied Biosystems). Primer names, sequences, and GenBank accession numbers are listed in Table 2. QRT PCR was achieved using StepOne Plus Real Time PCR (Applied Biosystems) with a Fast-Start Universal SYBR Green Master Mix (Roche, Indianapolis, IN), according to the manufacturers' instructions. The amplification steps consisted of initiation at 95C for 10 min, 40 cycles of denaturation at 95C for 15 sec each followed by annealing at 60C for 1 min, with next a melting curve analysis step featuring 1 cycle at 95C for 15 sec, a hold at 60C for 1 min, and a further hold at 95C for 15 sec, to confirm that only a single amplicon was present. All reactions were performed in duplicate using olive flounder -actin mRNA as an internal control. The relative standard curve method was applied to calculate standard errors of the mean (SEM). In statistical analysis, Student's t test was used to determine differences between tests using FKC and OMV; p<0.05 was viewed as significant.

38

Table 2 Oligonucleotide Sequences Used In The Present Study.

Fish Vaccination and Challenge Before vaccination, 10 nave olive flounders in each group were intraperitoneally injected with ED45 (1.2104109 CFU/ml) and mortality was recorded over the next 28 days. A group of fish injected with 1.2104 CFU/ml of ED45 showed 70% mortality, and this dose was thus chosen as the challenge (data not shown). Twenty fish were vaccinated with 0.1 ml amounts of FKC (1.2108 CFU/ml; 344 g) or 10 g amounts of OMV, whereas fish injected with PBS served as controls. After acclimatization for 28 days, fish were challenged with 100 l amounts of ED45 (1.1104 CFU/ml) by intraperitoneal injection, and mortality was recorded over the following 31 days. Vaccine efficacy was calculated as relative percentage survival (RPS = [1 minus vaccine group mortality/control group mortality]100) [55]. In statistical analysis, Fisher's exact test was applied to compare survival differences between vaccinated and control groups at the p<0.0001 level.

39

REFERENCES 1. Kuehn MJ, Kesty NC. Bacterial outer membrane vesicles and the hostpathogen interaction. Genes Dev. 2005;19:26452655. [PubMed] 2. Beveridge TJ. Structures of gram-negative cell walls and their derived membrane vesicles. J Bacteriol. 1999;181:47254733. [PMC free article] [PubMed] 3. Devoe IW, Gilchrist JE. Release of endotoxin in the form of cell wall blebs during in vitro growth of Neisseria meningitidis. J Exp Med. 1973;138:11561167. [PMC free article] [PubMed] 4. Hoekstra D, van der Laan JW, de Leij L, Witholt B. Release of outer membrane fragments from normally growing Escherichia coli. Biochim Biophys Acta. 1976;455:889899. [PubMed] 5. Fiocca R, Necchi V, Sommi P, Ricci V, Telford J, et al. Release of Helicobacter pylori vacuolating cytotoxin by both a specific secretion pathway and budding of outer membrane vesicles. uptake of released toxin and vesicles by gastric epithelium. J Pathol. 1999;188:220226. [PubMed] 6. Kadurugamuwa JL, Beveridge TJ. Membrane vesicles derived from Pseudomonas aeruginosa and Shigella flexneri can be integrated into the surfaces of other gram-negative bacteria. Microbiology. 1999;145:2051 2060. [PubMed] 7. Kondo K, Takade A, Amako K. Release of the outer membrane vesicles from Vibrio cholerae and Vibrio parahaemolyticus. Microbiol Immunol. 1993;37:149152. [PubMed] 8. Nevot M, Deroncel V, Messner P, Guinea J, Mercad E. Characterization of outer membrane vesicles released by the psychrotolerant bacterium Pseudoalteromonas antarctica NF3. Environ Microbiol. 2006;8:15231533. [PubMed] 9. Wai SN, Lindmark B, Soderblom T, Takade A, Westermark M, et al. Vesicle-mediated export and assembly of pore-forming oligomers of the enterobacterial ClyA cytotoxin. Cell. 2003;115:2535. [PubMed]

40

10. Bauman SJ, Kuehn MJ. Purification of outer membrane vesicles from Pseudomonas aeruginosa and their activation of an IL-8 response. Microb Infect. 2006;8:24002408. 11. Meyer FP, Bullock GL. Edwardsiella tarda, a new pathogen of channel catfish (Ictalurus punctatus). Appl Microbiol. 1973;25:155156. [PMC free article] [PubMed] 12. Egusa S. Some bacterial diseases of freshwater fishes in Japan. Fish Pathol. 1976;10:103114. 13. Nakatsugawa T. Edwardsiella tarda isolated from cultured young flounder. Fish Pathol. 1983;18:99101. 14. Kusuda R, Toyoshima T, Iwamura Y, Sako H. Edwardsiella tarda from an epizootic of mullets (Mugil cephalus) in okitsu bay. Bull Jap Soc Sci Fish. 1976;42:271275. 15. Nougayrede PH, Vuillaume A, Vigneulle M, Faivre B, Luengo S, et al. First isolation of Edwardsiella tarda from diseased turbot (Scophthalmus maximus) reared in a sea farm in the bay of biscay. EAFP Bulletin. 1994;14:128129. 16. Herman RL, Bullock GL. Pathology caused by the bacterium Edwardsiella tarda in striped bass. Trans Am Fish Soc1. 1986;15:232235. 17. Mohanty BR, Sahoo PK. Edwardsiellosis in fish: A brief review. J Biosci. 2007;32:13311344. [PubMed] 18. Castro N, Toranzo AE, Nez S, Magarios B. Development of an effective Edwardsiella tarda vaccine for cultured turbot (Scophthalmus maximus). Fish Shellfish Immunol. 2008;25:208212. [PubMed] 19. Kawai K, Liu Y, Ohnishi K, Oshima S. A conserved 37 kDa outer membrane protein of Edwardsiella tarda is an effective vaccine candidate. Vaccine. 2004;22:34113418. [PubMed] 20. Kwon SR, Nam YK, Kim SK, Kim KH. Protection of tilapia (Oreochromis mosambicus) from edwardsiellosis by vaccination with Edwardsiella tarda ghosts. Fish Shellfish Immunol. 2006;20:621626. [PubMed]

41

21. Wang B, Mo ZL, Xiao P, Li J, Zou YX, et al. EseD, a putative T3SS translocon component of Edwardsiella tarda, contributes to virulence in fish and is a candidate for vaccine development. Mar Biotechnol. 2010 doi: 10.1007/s10126-009-9255-5. 22. Janda JM, Abbott SL. Expression of an iron-regulated hemolysin by Edwardsiella tarda. FEMS Microbiol Lett. 1993;111:275280. [PubMed] 23. Srinivasa Rao PS, Yamada Y, Leung KY. A major catalase (KatB) that is required for resistance to H2O2 and phagocyte-mediated killing in Edwardsiella tarda. Microbiology. 2003;149:26352644. [PubMed] 24. Ullah MA, Arai T. Pathological activities of the naturally occurring strains of Edwardsiella tarda. Fish Pathol. 1983;18:6570. 25. Lee EY, Bang JY, Park GW, Choi DS, Kang JS, et al. Global proteomic profiling of native outer membrane vesicles derived from Escherichia coli. Proteomics. 2007;7:31433153. [PubMed] 26. Post DM, Zhang D, Eastvold JS, Teghanemt A, Gibson BW, et al. Biochemical and functional characterization of membrane blebs purified from Neisseria meningitidis serogroup B. J Biol Chem. 2005;280:38383 38394. [PubMed] 27. Roberts R, Moreno G, Bottero D, Gaillard ME, Fingermann M, et al. Outer membrane vesicles as acellular vaccine against pertussis. Vaccine. 2008;26:46394646. [PubMed] 28. Schild S, Nelson EJ, Bishop AL, Camilli A. Characterization of Vibrio cholerae outer membrane vesicles as a candidate vaccine for cholera. Infect Immun. 2009;77:472484. [PMC free article] [PubMed] 29. van den Dobbelsteen GPJM, van Dijken HH, Pillai S, van Alphen L. Immunogenicity of a combination vaccine containing pneumococcal conjugates and meningococcal PorA OMVs. Vaccine. 2007;25:2491 2496. [PubMed] 30. Holst J, Martin D, Arnold R, Huergo CC, Oster P, et al. Properties and clinical performance of vaccines containing outer membrane vesicles from Neisseria meningitidis. Vaccine. 2009;27:B3B12. [PubMed]

42

31. Hirono I, Takami M, Miyata M, Miyazaki T, Han HJ, et al. Characterization of gene structure and expression of two toll-like receptors from Japanese flounder, Paralichthys olivaceus. Immunogenetics. 2004;56:3846. [PubMed] 32. Wu CC, Yates JR. The application of mass spectrometry to membrane proteomics. Nat Biotechnol. 2003;21:262267. [PubMed] 33. Kumar G, Sharma P, Rathore G, Bisht D, Sengupta U. Proteomic analysis of outer membrane proteins of Edwardsiella tarda. J Appl Microbiol. 2010;108:22142221. [PubMed] 34. Horstman AL, Kuehn MJ. Enterotoxigenic Escherichia coli secretes active heat-labile enterotoxin via outer membrane vesicles. J Biol Chem. 2000;275:1248912496. [PubMed] 35. Galka F, Wai SN, Kusch H, Engelmann S, Hecker M, et al. Proteomic characterization of the whole secretome of Legionella pneumophila and functional analysis of outer membrane vesicles. Infect Immun. 2008;76:18251836. [PMC free article] [PubMed] 36. Strauss EJ, Ghori N, Falkow S. An Edwardsiella tarda strain containing a mutation in a gene with homology to shlB and hpmB is defective for entry into epithelial cells in culture. Infect Immun. 1997;65:39243932. [PMC free article] [PubMed] 37. Prasadarao NV, Wass CA, Kim KS. Endothelial cell GlcNAc beta 14GlcNAc epitopes for outer membrane protein A enhance traversal of Escherichia coli across the blood-brain barrier. Infect Immun. 1996;64:154160. [PMC free article] [PubMed] 38. Prasadarao NV, Wass CA, Weiser JN, Stins MF, Huang SH, et al. Outer membrane protein A of Escherichia coli contributes to invasion of brain microvascular endothelial cells. Infect Immun. 1996;64:146153. [PMC free article] [PubMed] 39. Bolton A, Song XM, Willson P, Fontaine MC, Potter AA, et al. Use of the surface proteins GapC and mig of Streptococcus dysgalactiae as potential protective antigens against bovine mastitis. Can J Microbiol. 2004;50:423 432. [PubMed]

43

40. Rao PSS, Yamada Y, Tan YP, Leung KY. Use of proteomics to identify novel virulence determinants that are required for Edwardsiella tarda pathogenesis. Mol Microbiol. 2004;53:573586. [PubMed] 41. Zheng J, Leung KY. Dissection of a type VI secretion system in Edwardsiella tarda. Mol Microbiol. 2007;66:11921206. [PubMed] 42. Whyte SK. The innate immune response of finfish-A review of current knowledge. Fish Shellfish Immunol. 2007;23:11271151. [PubMed] 43. Akira S, Uematsu S, Takeuchi O. Pathogen recognition and innate immunity. Cell. 2006;124:783801. [PubMed] 44. Alaniz RC, Deatherage BL, Lara JC, Cookson BT. Membrane vesicles are immunogenic facsimiles of Salmonella typhimurium that potently activate dendritic cells, prime B and T cell responses, and stimulate protective immunity in vivo. J Immunol. 2007;179:76927701. [PubMed] 45. Jendholm J, Mrgelin M, Mnsson A, Larsson C, Cardell LO, et al. B cell activation by outer membrane VesiclesA novel virulence mechanism. PLoS Pathog. 6:e1000724. [PMC free article] [PubMed] 46. Mekuchi T, Kiyokawa T, Honda K, Nakai T, Muroga K. Vaccination trials in the Japanese flounder against edwardsiellosis. Fish Pathol. 1995;30:251256. 47. Kadurugamuwa JL, Beveridge TJ. Virulence factors are released from pseudomonas aeruginosa in association with membrane vesicles during normal growth and exposure to gentamicin: A novel mechanism of enzyme secretion. J Bacteriol. 1995;177:39984008. [PMC free article] [PubMed] 48. Kesty NC, Kuehn MJ. Incorporation of heterologous outer membrane and periplasmic proteins into Escherichia coli outer membrane vesicles. J Biol Chem. 2004;279:20692076. [PubMed] 49. Laemmli UK. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature. 1970;227:680685. [PubMed] 50. Shin GW, Palaksha KJ, Yang HH, Shin YS, Kim YR, et al. Partial twodimensional gel electrophoresis (2-DE) maps of Streptococcus iniae

44

ATCC29178 and Lactococcus garvieae KG9408. Dis Aquat Organ. 2006;70:7179. [PubMed] 51. Tatusov RL, Natale DA, Garkavtsev IV, Tatusova TA, Shankavaram UT, et al. The COG database: New developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res. 2001;29:2228. [PMC free article] [PubMed] 52. Yu NY, Wagner JR, Laird MR, Melli G, Rey S, et al. PSORTb 3.0: Improved protein subcellular localization prediction with refined localization subcategories and predictive capabilities for all prokaryotes. Bioinformatics. 2010;26:16081615. [PMC free article] [PubMed] 53. Gonzlez S, Caballero E, Soria Y, Cobas K, Granadillo M, et al. Immunization with Neisseria meningitidis outer membrane vesicles prevents bacteremia in neonatal mice. Vaccine. 2006;24:16331643. [PubMed] 54. Aoki M, Kondo M, Nakatsuka Y, Kawai K, Oshima SI. Stationary phase culture supernatant containing membrane vesicles induced immunity to rainbow trout Oncorhynchus mykiss fry syndrome. Vaccine. 2007;25:561 569. [PubMed] 55. Amend DF. Potency testing of fish vaccines. Dev Biol Stand. 1981;49:447454. 56. http://www.ncbi.nlm.nih.gov/pmc/articles/PMC3050902/

45

Vous aimerez peut-être aussi