Skip to main content

A shift in the virulence potential of Corynebacterium pseudotuberculosis biovar ovis after passage in a murine host demonstrated through comparative proteomics

Abstract

Background

Corynebacterium pseudotuberculosis biovar ovis, a facultative intracellular pathogen, is the etiologic agent of caseous lymphadenitis in small ruminants. During the infection process, C. pseudotuberculosis changes its gene expression to resist different types of stresses and to evade the immune system of the host. However, factors contributing to the infectious process of this pathogen are still poorly documented. To better understand the C. pseudotuberculosis infection process and to identify potential factors which could be involved in its virulence, experimental infection was carried out in a murine model using the strain 1002_ovis and followed by a comparative proteomic analysis of the strain before and after passage.

Results

The experimental infection assays revealed that strain 1002_ovis exhibits low virulence potential. However, the strain recovered from the spleen of infected mice and used in a new infection challenge showed a dramatic change in its virulence potential. Label-free proteomic analysis of the culture supernatants of strain 1002_ovis before and after passage in mice revealed that 118 proteins were differentially expressed. The proteome exclusive to the recovered strain contained important virulence factors such as CP40 proteinase and phospholipase D exotoxin, the major virulence factor of C. pseudotuberculosis. Also, the proteome from recovered condition revealed different classes of proteins involved in detoxification processes, pathogenesis and export pathways, indicating the presence of distinct mechanisms that could contribute in the infectious process of this pathogen.

Conclusions

This study shows that C. pseudotuberculosis modifies its proteomic profile in the laboratory versus infection conditions and adapts to the host context during the infection process. The screening proteomic performed us enable identify known virulence factors, as well as potential proteins that could be related to virulence this pathogen. These results enhance our understanding of the factors that might influence in the virulence of C. pseudotuberculosis.

Background

Corynebacterium pseudotuberculosis biovar ovis is a Gram-positive facultative intracellular pathogen. It is the etiologic agent of Caseous Lymphadenitis (CLA) in small ruminants, a disease characterized by abscess formation in lymph nodes and internal organs [1]. Cases of human infection caused by C. pseudotuberculosis have been reported and are associated with occupational exposure [1]. CLA is globally distributed and causes significant economic losses in goats, and sheep herds [2]. The pathogenic process of C. pseudotuberculosis in the host comprises two phases: (i) initial colonization and replication in lymph nodes that drain the site of infection, which is associated with pyogranuloma formation, and (ii) a secondary cycle of replication and dissemination via the lymphatic or circulatory systems. This dissemination is promoted by the action of phospholipase D (PLD) exotoxin, the major virulence factor of C. pseudotuberculosis, which allows this pathogen to contaminate visceral organs and lymph nodes, where it ultimately induces lesion formation [3–5].

Exported proteins reportedly favor the infection process in pathogenic bacteria; this class of proteins is involved in adhesion and invasion of host cells, nutrient acquisition, toxicity, and in the evasion of the host immune system [6]. Different strategies like the transposon mutagenesis have been adopted to identify C. pseudotuberculosis biovar ovis exported proteins [7]. Additionally, comparative proteomics has been applied to characterize the extracellular proteome of C. pseudotuberculosis biovar ovis, as well as, the extracellular immunoproteome (strains C231_ovis and 1002_ovis) [8–11]. In these studies, some proteins of the strain 1002_ovis, suspected to be virulence factors, were not detected suggesting this strain presents a low virulence. The surface proteome of C. pseudotuberculosis biovar ovis was also characterized using bacterial strains isolated from the lymph nodes of naturally infected sheep. This proteomic analysis allowed the identification of proteins that could favor the survival of this pathogen during the chronic phase of CLA [12].

The experimental passage of bacterial pathogens through in vitro or in an in vivo model is a strategy that has been applied to evaluate the virulence potential of several pathogens. By generating a confrontation between the pathogen and the dynamic network of host factors, including the immune system components, it helps to identify bacterial factors involved in virulence [12–19]. In this study, the strain 1002_ovis was experimentally inoculated in mice [20, 21] to identify factors which could contribute to virulence in C. pseudotuberculosis biovar ovis. Comparative proteomics of the culture supernatant from this strain collected before and after the experimental passage in mice was carried out to identify factors that might contribute to virulence of 1002_ovis.

Methods

Bacterial strains and growth conditions

The C. pseudotuberculosis biovar ovis strain 1002 (1002_ovis) was isolated from a goat in Brazil; this strain was cultivated under standard conditions in brain–heart infusion broth (BHI-HiMedia Laboratories Pvt. Ltd., India) at 37 °C. When necessary, 1.5% of agar was added to the medium for a solid culture. For extracellular proteomic analyses, 1002_ovis was grown in a chemically defined medium (CDM) [(Na2HPO4_7H2O (12.93 g/L), KH2PO4 (2.55 g/L), NH4Cl (1 g/L), MgSO4_7H2O (0.20 g/L), CaCl2 (0.02 g/L) and 0.05% (v/v) Tween 80], 4% (v/v) MEM Vitamins Solution (Invitrogen, Gaithersburg, MD, USA), 1% (v/v) MEM Amino Acids Solution (Invitrogen), 1% (v/v) MEM Non-Essential Amino Acids Solution (Invitrogen), and 1.2% (w/v) glucose at 37 °C [22].

Experimental infection of strain 1002_ovis in a murine model (in vivo assay)

The standardization of the parameters for infection was performed according to Moraes et al. [20] and Ribeiro et al. [21]. Female BALB/c mice between six and eight weeks old were used in all experiments. They were provided by the Animal Care Facility of the Biological Sciences Institute from the Federal University of Minas Gerais and were handled by the guidelines of the UFMG Ethics Committee on Animal Testing (Permit Number: CETEA 103/2011). For the bacterial passage assay using the murine model, two groups of three mice each was infected via intraperitoneal injection with 106 colony forming units (CFU) of strain 1002_ovis. Thirty-six hours after infection, all animals were sacrificed. Their spleens were aseptically removed to recover the bacterial strain, as described below: the spleen removed from each animal was then, individually macerated in sterile saline solution (0.9% NaCl2), seeded onto BHI agar plates and incubated for 48 h at 37 °C. Subsequently, one recovered bacterial colony was cultured in BHI broth. The recovered bacteria were then referred to as Recovered (Rc). For the bacterial virulence assay, we used the freshly recovered bacteria and bacteria that did not contact the murine host as a control, which is referred to as Control (Ct). Groups of five mice were infected with Rc and Ct, via intraperitoneal injection of a suspension containing 106 CFU or 105 CFU. The animals’ survival rates were calculated and represented in GraphPad Prism v.5.0 (GraphPad Software, San Diego, CA, USA) using the Kaplan-Meier survival function. The results of 1002_ovis CFU count in the organs were calculated using the two-way ANOVA test.

Preparation of proteins from culture filtrates for proteome analysis

For proteomic analysis, the Ct and Rc (three independently recovered colonies) that was obtained from infected mice spleens as described above were grown in CDM at OD600 = 0.8. The cultures were then centrifuged for 20 min at 2700 × g. The supernatants were then filtered using 0.22-μm filters, 30% (w/v) ammonium sulfate was added to the samples, and the pH of the mixtures was adjusted to 4.0. Next, 20 mL N-butanol was added to each sample. The samples were centrifuged for 10 min at 1350 xg and 4 °C. The interfacial precipitate was collected and resuspended in 1 mL of 20 mM Tris–HCl pH 7.2 [23]. Finally the concentration protein was determined by Bradford method [24].

2D-PAGE electrophoresis and Mass Spectrometry

The 2-DE procedure and in-gel protein digestion were performed as described previously [9, 10]. Approximately 300 μg of the protein extract from of each condition was dissolved in rehydration buffer (Urea 7 M, thiourea 2 M, CHAPS 2%, Tris–HCl 40 mM, bromophenol blue 0.002%, DTT 75 mM, IPG Buffer 1%). Samples were applied to 18 cm pH 3–10 N.L strips (GE Healthcare, Pittsburgh, USA). Isoelectric focusing (IEF) was performed using the apparatus IPGphor 2 (GE Healthcare) under the following voltages: 100 V 1 h, 500 V 2 h, 1000 V 2 h, 10,000 V 3 h, 10,000 V 6 h, 500 V 4 h. The IPG strips were placed on 12% acrylamide/bis acrylamide gels in an Ettan DaltSix II system (GE Healthcare). The gels were stained with Coomassie Blue G-250 staining solution, and 2-DE gels were scanned using an Image Scanner (GE Healthcare). The Image Master 2D Platinum 7 (GE Healthcare) software was used to analyze the generated images and all spots were matched and analyzed by gel-to-gel comparison. The quantification of the spots was calculated according percentage volume (% Vol) and spots with reproducible changes in abundance were considered to be differentially expressed. Protein spots were excised from the gels, and in-gel digestion was carried out using trypsin enzyme (Promega, Sequencing Grade Modified Trypsin, Madison, WI, USA). The peptides were then desalted and concentrated using ZIP TIP C18 tips (Eppendorf).

The samples were subsequently analyzed for MS and MS/MS modes, using an MALDI-TOF/TOF mass spectrometer Autoflex IIITM (Bruker Daltonics, Billerica USA). The equipment was controlled in a positive/reflector way using the Flex-ControlTM software (Brucker Daltonics). External calibration was performed using peptide standards samples (angiotensin II, angiotensin I, substance P, bombesin, ACTH clip 1–17, ACTH clip 18–39, somatostatin 28, bradykinin Fragment 1–7, Renin Substrate tetra decapeptide porcine) (Bruker Daltonics). The peptides were added to the alpha-cyano-4-hydroxycinnamic acid matrix, applied on an Anchor-ChipTM 600 plate (Brucker Daltonics) and analyzed by Autoflex III. The search parameters were as follows: enzyme; trypsin; fixed modification, carbamidomethylation (Cys); variable modifications, oxidation (Met); mass values, monoisotopic; maximum missed cleavages, 1; and peptide mass tolerance of 0.005% Da (50 ppm). The results obtained by MS/MS were used to identify proteins utilizing the MASCOT_ (http://www.matrixscience.com) program and compared with the genomic data of the Actinobacteria class deposited in the NCBI nr database.

2D nanoUPLC-HDMSE data acquisition and Data Processing

The protein extracts from three biological replicates of each condition were concentrated using spin columns with a 10 kDa threshold (Millipore, Billerica, MA, USA) to perform the label-free proteomic analysis. The protein was denatured (0.1% RapiGEST SF at 60 °C for 15 min) (Waters, Milford, CA, USA), reduced (10 mM DTT), alkylated (10 mM iodoacetamide) and enzymatically digested with trypsin (Promega). The digestion process was stopped by adding 10 μL of 5% TFA (Fluka, Buchs, Germany), and glycogen phosphorylase (Sigma, Aldrich, P00489) was added to the digested samples after digest at 20 fmol.uL−1 as an internal standard for normalization. Each replicate was injected using a two-dimensional reversed phase (2D RPxRP) nanoUPLC-MS (Nano Ultra Performance Liquid Chromatography Mass Spectrometry) approach with 171 multiplexed high definition mass spectrometry (HDMSE) label-free quantitation [25]. Qualitative and quantitative experiments were performed using both a 1 h reversed phase gradient from 7% to 40% (v/v) acetonitrile (0.1% v/v formic acid) at 500 nL.min−1 and a nanoACQUITY UPLC 2D RPxRP Technology system [26]. A nanoACQUITY 174 UPLC HSS T3 1.8 μm, 75 μm × 15 cm column (pH 3) was used with an RP XBridge BEH130 C18 5 μm 300 μm x 50 mm nanoflow column (pH 10). Typical on-column sample loads were 250 ng of the total protein digests for each of the 5 fractions (250 ng/fraction/load). All analyses were performed using nano electrospray ionization in the positive ion mode nanoESI (+) and a NanoLockSpray (Waters, Manchester, UK) ionization source. The mass spectrometer was calibrated using an MS/MS spectrum of [Glu1]-Fibrinopeptide B human (Glu-Fib) solution (100 fmol.uL-1) delivered through the NanoLockSpray source reference sprayer. Multiplexed data-independent (DIA) scanning with additional specificity and selectivity for non-linear ‘T-wave’ ion mobility (HDMSE) experiments were performed using a Synapt G2-S HDMS mass spectrometer (Waters, Manchester, UK).

Following the identification of proteins, the quantitative data were packaged using dedicated algorithms [27] and searching against a database with default parameters to account for ions [28]. The databases used were reversed on-the-fly during the database queries and appended to the original database to assess the false positive rate during identification. For proper spectra processing and database searching conditions, the ProteinLynxGlobalServer v.2.5.2 (PLGS) with IdentityE and ExpressionE informatics v.2.5.2 (Waters, Manchester, UK) was used. UniProtKB (release 2013_01) with manually reviewed annotations was used, and the search conditions were based on taxonomy (Corynebacterium pseudotuberculosis). One missed cleavage by trypsin was allowed be up to 1 and various modifications as carbamidomethyl (C), Acetyl N terminal, phosphoryl (STY) and oxidation (M) were allowed [29]. The proteins collected were organized by the PLGS ExpressionE tool algorithm into a statistically significant list that corresponded to higher or lower regulation ratios between the different groups. For protein quantitation, we used the PLGS v2.5.2 software with the IdentifyE algorithm using the Hi3 methodology. The search threshold to accept each spectrum was the default value for a false discovery rate 4%. The quantitation values were averaged over all samples, and the standard deviations of p < 0.05, which were determined using the ExpressionE software, refer to the differences between biological replicates.

Bioinformatic analysis

The proteins identified in 1002_ovis under both conditions were analyzed using the following prediction tools: SecretomeP 2.0 server, to predict proteins exported from non-classical systems (positive prediction score greater than to 0.5) [30] and PIPs software, to predict proteins in the pathogenicity islands [31]. Gene ontology (GO) functional annotations were generated using the Blast2GO tool [32].

Results

The main objective of this study was to assay the virulence of 1002_ovis in a murine model after passage through mice. We thus carried out an in vivo survival assay using BALB/c mice infected with bacteria that did not contact with murine model (Ct) and bacteria recovered (Rc) from mice spleens. In this assay using an infection inoculum of 106 CFU, all the animals infected with Rc died within 48 h after infection (Fig. 1a). On the other hand, the control group, infected with Ct, survived the evaluation period (6 days). Similarly, in an assay with a lower infective dose (105 CFU), a 100% mortality was observed four weeks post infection with the recovered bacteria (Fig. 1b). Comparison of the Ct and Rc numbers isolated from the spleen within five days of infection (Fig. 1c) showed that the serial passage process affected the potential for spleen colonization during the infection. After four weeks of infection in the assay with 105 CFU, bacteria were isolated from the spleen, liver, left and right kidney, only in mice infected with Rc (Fig. 1d). Finally, regarding the clinical signs, in the assay using 105 CFU, caseous lesions were detected in different organs (liver, left kidney and right kidney) of all the animals infected only with Rc (data not shown). Altogether, these results showed that the serial passage process in a murine model increased the virulence potential of strain 1002_ovis. In addition, these results confirmed the low virulence of this strain, which was previously suggested based on the composition of its extracellular proteome [8–10].

Fig. 1
figure 1

Survival of Balb/C mice infected with strain 1002_ovis. a The survival rate was measured to determine the virulence profile of strain 1002_ovis control and recovered in mice infected with 106 CFU of bacteria Ct = control condition, Rc recovered condition. b Survival rates of mice infected with 105 CFU of Ct and Rc. c CFU in the spleen of BALB/c mice infected with control and recovered condition for the first five days of infection. d CFU in the different organs (spleen, left kidney, right kidney and liver) of BALB/c mice infected with control and recovered condition after four weeks of infection. The mortality rates were measured daily. Results represent three independent experiments. P values of <0.05 were considered to be statistically significant, and asterisks indicate statistically significant differences

After passage in BALB/c mice, a dramatic change in the virulence potential of strain 1002_ovis was observed. We thus hypothesized that this phenotypic change was visible at the proteome level since C. pseudotuberculosis virulence relies on the production of a proteinaceous virulence factor. Thus, considering the importance of extracellular proteins for bacterial virulence, the proteomic analysis was conducted on the extracellular proteomes of 1002_ovis recovered from infected mice spleens in comparison to the control condition, using two proteomics approaches: 2-DE and 2D nanoUPLC-HDMSE. The electrophoretic resolution of the extracellular protein extract of Ct and Rc condition allowed the visualization of spots distributed over pH 3–10 (Fig. 2). A total of 14 spots were found to be differentially expressed between Ct and Rc condition, these spots were excised out of the gel, and identified by MS/MS (Table 1). In the LC/MS analysis, we used the label-free quantitative proteomic to evaluate the relative difference between the proteome of Rc and Ct condition. In this analysis, only proteins which presented p < 0.05 and differential expression (log2 ratios) equal or greater than a factor of 1.2 were considered, as described previously [33]. We detected a total of 118 expressed differentially proteins, between Ct and Rc condition (Fig. 3) (Table 2 and Additional file 1). Also, 48 proteins were assigned only to Ct (Additional file 2) and 32 proteins were exclusive to Rc (Table 3) The information about sequence coverage and a number of identified peptides for each protein sequence identified, as well as the information about the native peptide are available at Additional file 3: Table S3.

Fig. 2
figure 2

Two-dimensional electrophoresis of the extracellular proteins 1002_ovis after following passage process: a Control condition. b Recovered condition. Red circle: spot proteins identified by MS/MS

Table 1 List of proteins identified in 1002_ovis control and recovered by 2D-PAGE-MS/MS
Fig. 3
figure 3

Volcano Plot show Log(2) Fold Change of the differentially expressed proteins detected by label-free proteomics between the recovered and control condition. Green: Up-regulated proteins; Grey: unchanged proteins; Red: Down-regulated proteins

Table 2 Proteins differentially produced among the recovered and control condition
Table 3 List of proteins identified in the exclusive proteome of recovered-condition

The proteins identified in both conditions were analyzed by SecretomeP [29] to assess whether these proteins could be exported by non-classical secretion systems. Among the expressed differentially proteins 31% (37 proteins) were predicted as secreted through non-classical secretion systems. In turn, when analyzed the exclusive proteome of each condition 19% (6 proteins) and 27% (13 proteins) were considered to be exported by non-classical secretion systems for recovered and control condition, respectively. The PIPS tool was used to evaluate whether the genes that encode the proteins which were differentially expressed and identified in the exclusive proteome of the Rc condition are included in predicted pathogenicity islands. According these analysis 16 proteins was encoded by genes located on a predicted pathogenicity island; these proteins are related to cellular metabolism, pathogenesis, transport pathway, stress response and unknown function (Additional file 4). To classify the proteins identified in functional groups, we used the Blast2Go tool [31]; according to this analysis, the proteins were grouped into 17 biological processes (Fig. 4). Among these proteins, we identified processes that are directly involved in bacterial virulence, such as protein transport, pathogenesis, cell adhesion and stress response (Table 2).

Fig. 4
figure 4

Biological processes differentially regulated in 1002_ovis after passage in mice. Analysis of the differentially expressed proteins grouped into biological processes for strain 1002_ovis after passage in mice

Important factors directly linked to C. pseudotuberculosis virulence, like the PLD phospholipase, as well as, the CP40 protease were detected only in the proteome of recovered 1002_ovis (Tables 1 and 3). Also, components of several secretion systems were also activated in the bacteria recovered. These include proteins related to hemin uptake, ATP-binding cassette (ABC) transporters and the Opp transporter, like OppA, OppC, and OppD. Proteins related to detoxification process were also specifically identified in the Rc supernatant: e.g. the glutaredoxin-like protein NrdH, which belongs to the NrdH-redoxins, a family of small protein disulfide oxidoreductases [34], mycothiol glutathione reductase present in Actinobacteria [35] and copper resistance protein CopC (Tables 2 and 3). In addition, we have identified 31 proteins in the recovered condition that also were detected in a strain of C. pseudotuberculosis isolated directly from ovine lymph nodes [12] (Tables 2 and 3). Proteins involved in the resistance to antimicrobial agents, such as penicillin-binding proteins, metallo-beta-lactamase, and penicillin-binding protein transpeptidase and proteases like Clp protease involved in the expression of cytotoxins in Staphylococcus aureus and Listeria monocytogenes [36, 37] were found induced in Rc supernatant.

Discussion

To investigate the protein factors that could influence the adaptive processes of C. pseudotuberculosis biovar ovis during the infection process, we combined a unique bacterial passage experiment in mice with proteomic analyses of 1002_ovis culture supernatants, collected before and after passage. In the first analysis, we observed that strain 1002_ovis (isolated from caprine) exhibited a low virulence potential, which is consistent with previous reports indicating the low virulence potential of this strain [38, 39]. Although a recent in silico analysis of the 1002_ovis genome predicted various genes involved in virulence [40], studies examining the exoproteome of this strain under laboratory growth conditions failed to detect many of these virulence proteins (e.g., PLD exotoxin or proteins involved in the pathway of cell invasion, detoxification) [8–10].

One explanation for this relies on the fact that after being first isolated, strains 1002_ovis have been maintained, in vitro, under laboratory conditions with extensive passages on the culture medium, which may alter the gene expression profile of the strain, especially for effectors related to bacterial virulence. This phenomenon has also been reported in other pathogens such as Mycobacterium bovis, Helicobacter pylori, S. aureus, and L. monocytogenes. In vitro passages of these bacteria on culture medium altered both bacterial physiology and virulence profile [41–44]. However, we showed that the bacterial passage process in a murine model changed the virulence potential of strain 1002_ovis. Previous reports on experimental serial passages showed that pathogens such as H. pylori, Escherichia coli, Xenorhabdus nematophila, Arcobacter butzleri, and Salmonella enterica also exhibited altered virulence profiles after in vivo passage in a host, which helped identifying factors that contribute to infectious process [14–19]. Thus, as observed in these pathogens, the recovered condition also showed increased capacity to persist into host, when compared with control condition. The altered physiology and virulence status observed in 1002_ovis is supported by our proteomic analyses, where several proteins involved in processes favoring infection and host adaptation were differentially expressed after passage in mice.

Although our study focused on the C. pseudotuberculosis extracellular proteins, cytoplasmic proteins were also detected in the proteomic analyses. The presence of cytoplasmic proteins in the extracellular fraction is reported in several other proteomic studies [8–10, 12, 45]. It may be partially due to cell lysis and thus, be considered artifacts. However, cytoplasmic proteins in the culture supernatant may act as moonlighting proteins and be exported via a non-classical secretion pathway [30, 46]. The moonlighting proteins are described both Gram-positive and Gram-negative bacteria, and can be detected in different subcellular locations (cytoplasm, membrane, cell surface, and extracellular environment) and exhibit distinct functional behavior depending on the host cell type [46, 47]. Interestingly, some proteins, such as Chromosome partitioning protein ParB, Phosphoenolpyruvate carboxykinase GTP, Methylmalonyl CoA carboxyltransferase 12S subunit, Acetate kinase, and Enolase, induced in the Rc supernatants were identified only in the membrane shaving of C. pseudotuberculosis harvested directly from ovine lymph nodes [12].

The passage process in mice was also able to induce other proteins identified in Rc supernatants, and which contribute to the adhesion process. Proteins with an LPTXG domain, which characterizes the cell-wall anchored proteins, were identified and included monomers of membrane pilus. This latter class of proteins is described in pathogenic Corynebacterium species and may contribute especially in the process of cellular adhesion [48]. In Campylobacter jejuni, serial passages in mice induce the expression of invasiveness and increase the capacity of cell invasion [13]. Components of the Opp system were induced by the passage process, too. The Opp system facilitates the uptake of extracellular peptides, which are further used as carbon and nitrogen sources for bacterial nutrition [49]. Proteins that comprise the Opp system also were induced in a field isolated of C. pseudotuberculosis biovar ovis, when compared with the strain C231_ovis a laboratory reference strain [12, 50]. In the pathogen Mycobacterium avium the OppA gene was highly expressed during the infection in a mouse model [51]. We have identified known secreted virulence factors as CP40 serine protease, which previously shown to be necessary for C. pseudotuberculosis virulence potential and to induce an immune response [52, 53].

An important factor that precedes the chronic stage of infection by C. pseudotuberculosis is the capacity of this pathogen to disseminate within the host, which consequently favors the establishment of the disease [3]. In C. pseudotuberculosis, this process is mediated by the action of PLD exotoxin, a major virulence factor of this pathogen [54, 55] that catalyzes the dissociation of sphingomyelin and increases vascular permeability, which contributes to the dissemination process of C. pseudotuberculosis in the host. Here, PLD was only detected in the proteome of the Rc condition. This result is noteworthy because, a previous proteomic study performed by our research group, PLD was not identified in the extracellular proteome of 1002_ovis [8–10]. McKean et al. [5] showed that pld expression is expressed by different environmental factors, thus during the infection and recuperation process 1002_ovis was exposed to different environmental and stimulus, which may have affected the pld expression. A study showed that a pld mutant strain is indeed unable to disseminate and yields reduced virulence [55]. Here, we observed the presence of caseous lesions in different organs only at the end of experimental infection, only in the group of mice infected with the Rc condition. Altogether, the observations suggest that the expression of PLD can be modified by the passage in the host and can thus change the virulence potential of 1002_ovis.

Another attribute of PLD is its capacity to alter the viability of macrophage cells during the infection [5]. However, before promoting macrophages lysis, C. pseudotuberculosis has to be able to resist the hostile environment inside macrophages mainly against reactive oxygen species (ROS) and reactive nitrogen species (RNS). Thus, the induction of proteins involved in detoxification processes in Rc could be contributed for its resistance against ROS and RNS. The inductions of proteins related to oxidative stress also were observed in Shigella flexneri, after recuperation process in an in vivo infection model. We detected the mycothione glutathione reductase, a component of the mycothiol system, which is present in Mycobacterium and Rhodococcus genera. This system is used as an alternative mechanism of disulphide reduction and contributes to the cytosolic redox homeostasis and the resistance to ROS [35]. Glutaredoxin-like protein, NrdH, which plays an important role in the resistance to ROS, and is present in C. glutamicum [34] and M. tuberculosis [56] was also detected.

On the other hand, some proteins like dihydroxybiphenyl dioxygenase, Metallo beta lactamase superfamily protein, Formamidopyrimidine DNA glycosylase, MerR family transcriptional regulator, which were induced by 1002_ovis during the exposition to nitric oxide [57, 58] were also found induced in this study in the recovered condition. These proteins are related to different processes of resistance to nitrosative stress, DNA repair, antibiotic resistance, and transcription, these results show a set of proteins involved in the adaptation process of 1002_ovis to nitric oxide, which could contribute to the pathogenic process of this pathogen. Another type of defense of the host immune system against bacterial infection is the utilization of copper [59]. Here, CopC, a protein related to copper resistance, was detected in recovered 1002_ovis. In M. tuberculosis, proteins involved in copper resistance are essential to virulence [60, 61]. Thus, the association of this factor related to an antioxidant system with PLD could promote an effective pathway of defense against the action of the innate immune system and consequently contributes to virulence process of C. pseudotuberculosis.

Conclusion

In conclusion, the virulence potential and proteomic profiles of strain 1002_ovis undergo dramatic changes after recovery from experimentally infected mice. The proteomic screening outlined, after the serial passage in murine model showed a set of proteins that were induced in the recovered condition. Into this group were detected known secreted virulence factors, as well as some proteins which could contribute in its virulence. Therefore, more study is necessary to show the true role of these proteins in the virulence of C. pseudotuberculosis. Altogether, our results demonstrate that in vitro passages alter the expression of C. pseudotuberculosis exoproteome leading to a reduced virulence and that a single passage in vivo, in a murine model, can induce significant changes in the C. pseudotuberculosis extracellular proteome, contributing to the increase in virulence of this pathogen.

References

  1. Dorella FA, Pacheco LG, Oliveira SC, Miyoshi A, Azevedo V. Corynebacterium pseudotuberculosis: microbiology, biochemical properties, pathogenesis and molecular studies of virulence. Vet Res. 2006;37:201–18.

    Article  CAS  PubMed  Google Scholar 

  2. Paton MW, Walker SB, Rose IR, Watt GF. Prevalence of caseous lymphadenitis and usage of caseous lymphadenitis vaccines in sheep flocks. Aust Vet J. 2003;81:91–5.

    Article  CAS  PubMed  Google Scholar 

  3. Batey RG. Pathogenesis of caseous lymphadenitis in sheep and goats. Aust Vet J. 1986;63:269–72.

    Article  CAS  PubMed  Google Scholar 

  4. Pépin M, Pittet JC, Olivier M, Gohin I. Cellular composition of Corynebacterium pseudotuberculosis pyogranulomas in sheep. J Leukoc Biol. 1994;56:666–70.

    PubMed  Google Scholar 

  5. McKean SC, Davies JK, Moore RJ. Expression of phospholipase D, the major virulence factor of Corynebacterium pseudotuberculosis, is regulated by multiple environmental factors and plays a role in macrophage death. Microbiology. 2007;153:2203–11.

    Article  CAS  PubMed  Google Scholar 

  6. Green ER, Mecsas J. Bacterial Secretion Systems – An overview. Microbiol Spectr. 2016;4:1. Hilbi H, Haas A. Secretive bacterial pathogens and the secretory pathway. Traffic. 2012; 13:1187–1197.

    CAS  Google Scholar 

  7. Dorella FA, Estevam EM, Pacheco LG, Guimarães CT, Lana UG, Gomes EA, et al. In vivo insertional mutagenesis in Corynebacterium pseudotuberculosis: an efficient means to identify DNA sequences encoding exported proteins. Appl Environ Microbiol. 2006;72:7368–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Pacheco LG, Slade SE, Seyffert N, Santos AR, Castro TL, Silva WM, et al. A combined approach for comparative exoproteome analysis of Corynebacterium pseudotuberculosis. BMC Microbiol. 2011;17:12.

  9. Silva WM, Seyffert N, Santos AV, Castro TL, Pacheco LG, Santos AR, et al. Identification of 11 new exoproteins in Corynebacterium pseudotuberculosis by comparative analysis of the exoproteome. Microb Pathog. 2013;16:37–42.

    Article  Google Scholar 

  10. Silva WM, Seyffert N, Ciprandi A, Santos AV, Castro TL, Pacheco LG, et al. Differential Exoproteome analysis of two Corynebacterium pseudotuberculosis biovar ovis strains isolated from goat (1002) and sheep. Curr Microbiol. 2013;67:460–5.

    Article  CAS  PubMed  Google Scholar 

  11. Seyffert N, Silva RF, Jardin J, Silva WM, Castro TL, Tartaglia NR, et al. Serological proteome analysis of Corynebacterium pseudotuberculosis isolated from different hosts reveals novel candidates for prophylactics to control caseous lymphadenitis. Vet Microbiol. 2014;174:255–60.

    Article  CAS  PubMed  Google Scholar 

  12. Rees MA, Kleifeld O, Crellin PK, Ho B, Stinear TP, Smith AI, Coppel RL. Proteomic Characterization of a Natural Host-Pathogen Interaction: Repertoire of in vivo Expressed Bacterial and Host Surface-Associated Proteins. J Proteome Res. 2015;2:120–32.

    Article  Google Scholar 

  13. Fernández H, Vivanco T, Eller G. Expression of invasiveness of Campylobacter jejuni ssp. jejuni after serial intraperitoneal passages in mice. J Vet Med B Infect Dis Vet Public Health. 2000;47:635–9.

    Article  PubMed  Google Scholar 

  14. Bleich A, Kohn I, Glage S, Beil W, Wagner S, Mahler M. Multiple in vivo passages enhance the ability of clinical Helicobacter pylori isolate to colonize the stomach of Mongolian gerbils and to induce gastritis. Lab Anim. 2005;39:221–9.

    Article  CAS  PubMed  Google Scholar 

  15. Chapuis É, Pagès S, Emelianoff V, Givauda A, Ferdy JB. Virulence and pathogen multiplication: a serial passage experiment in the hypervirulent bacterial insect-pathogen Xenorhabdus nematophila. PLoS One. 2011;31:e15872.

    Article  Google Scholar 

  16. Fernandez-Brando RJ, Miliwebsky E, Mejías MP, Baschkier A, Panek CA, Abrey-Recalde MJ, et al. Shiga toxin-producing Escherichia coli O157: H7 shows an increased pathogenicity in mice after the passage through the gastrointestinal tract of the same host. J Med Microbiol. 2012;61:852–9.

    Article  CAS  PubMed  Google Scholar 

  17. Fernández H, Flores SP, Villanueva M, Medina G, Carrizo M. Enhancing adherence of Arcobacter butzleri after serial intraperitoneal passages in mice. Rev Argent Microbiol. 2013;45:75–9.

    PubMed  Google Scholar 

  18. Koskiniemi S, Gibbons HS, Sandegren L, Anwar N, Ouellette G, Broomall S, et al. Pathoadaptive mutations in Salmonella enterica isolated after serial passage in mice. PLoS One. 2013;25:e70147.

    Article  Google Scholar 

  19. Liu X, Lu L, Liu X, Pan C, Feng E, Wang D, Zhu L, Wang H. Comparative proteomics of Shigella flexneri 2a strain using a rabbit ileal loop model reveals key proteins for bacterial adaptation in host niches. Int J Infect Dis. 2015;40:28–33.

    Article  CAS  PubMed  Google Scholar 

  20. Moraes PM, Seyffert N, Silva WM, Castro TL, Silva RF, Lima DD, et al. Characterization of the Opp peptide transporter of Corynebacterium pseudotuberculosis and its role in virulence and pathogenicity. Biomed Res Int. 2014;2014:489782.

  21. Ribeiro D, Rocha FS, Leite KM, Soares SC, Silva A, Portela RW, et al. An iron acquisition-deficient mutant of Corynebacterium pseudotuberculosis efficiently protects mice against challenge. Vet Res. 2014;45:28.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Moura-Costa LF, Paule BJA, Freire SM, Nascimento I, Schaer R, Regis LF, et al. Chemically defined synthetic medium for Corynebacterium pseudotuberculosis culture. Rev Bras Saúde Prod An. 2002;3:1–9.

    Google Scholar 

  23. Paule BJ, Meyer R, Moura-Costa LF, Bahia RC, Carminati R, Regis LF, et al. Three-phase partitioning as an efficient method for extraction/concentration of immunoreactive excreted-secreted proteins of Corynebacterium pseudotuberculosis. Protein Expr Purif. 2004;34:311–166.

    Article  CAS  PubMed  Google Scholar 

  24. Bradford MM. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem. 1976;72:248–54.

    Article  CAS  PubMed  Google Scholar 

  25. Silva JC, Gorenstein MV, Li GZ, Vissers JP, Geromanos SJ. Absolute quantification of proteins by LCMSE: a virtue of parallel MS acquisition. Mol Cell Proteomics. 2006;5:144–56.

    Article  CAS  PubMed  Google Scholar 

  26. Gilar M, Olivova P, Daly AE, Gebler JC. Two-dimensional separation of peptides using RP-RP-HPLC system with different pH in first and second separation dimensions. J Sep Sci. 2005;8:1694–703.

    Article  Google Scholar 

  27. Geromanos SJ, Vissers JP, Silva JC, Dorschel CA, Li GZ, Gorenstein MV, et al. The detection, correlation, and comparison of peptide precursor and product ions from data independent LC-MS with data dependant LC-MS/MS. Proteomics. 2009;9:1683–95.

    Article  CAS  PubMed  Google Scholar 

  28. Li GZ, Vissers JP, Silva JC, Golick D, Gorenstein MV, Geromanos SJ. Database searching and accounting of multiplexed precursor and product ion spectra from the data independent analysis of simple and complex peptide mixtures. Proteomics. 2009;9:1696–719.

    Article  CAS  PubMed  Google Scholar 

  29. Curty N, Kubitschek-Barreira PH, Neves GW, Gomes D, Pizzatti L, Abdelhay E. Discovering the infectome of human endothelial cells challenged with Aspergillus fumigatus applying a mass spectrometry label-free approach. J Proteomics. 2014;31:126–40.

    Article  Google Scholar 

  30. Bendtsen JD, Kiemer L, Fausboll A, Brunak S. Non-classical protein secretion in bacteria. BMC Microbiol. 2005;5:58.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Soares SC, Abreu VA, Ramos RT, Cerdeira L, Silva A, Baumbach J. PIPS: pathogenicity island prediction software. PLoS One. 2012;7:e30848.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Conesa A, Gotz S, García-Gómez JM, Terol J, Talón M, Robles M. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics. 2005;15:3674–6.

    Article  Google Scholar 

  33. Levin Y, Hradetzky E, Bahn S. Quantification of proteins using data-independent analysis (MSE) in simple and complex samples: a systematic evaluation. Proteomics. 2011;11:3273–87.

    Article  CAS  PubMed  Google Scholar 

  34. Si MR, Zhang L, Yang ZF, Xu YX, Liu YB, Jiang CY, et al. NrdH Redoxin enhances resistance to multiple oxidative stresses by acting as a peroxidase cofactor in Corynebacterium glutamicum. Appl Environ Microbiol. 2014;80:1750–62.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Newton GL, Buchmeier N, Fahey RC. Biosynthesis and functions of mycothiol, the unique protective thiol of Actinobacteria. Microbiol Mol Biol Rev. 2008;72:471–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Frees D, Qazi SN, Hill PJ, Ingmer H. Alternative roles of ClpX and ClpP in Staphylococcus aureus stress tolerance and virulence. Mol Microbiol. 2013;48:1565–78.

    Article  Google Scholar 

  37. Gaillot O, Pellegrini E, Bregenholt S, Nair S, Berche P. The ClpP serine protease is essential for the intracellular parasitism and virulence of Listeria monocytogenes. Mol Microbiol. 2000;35:1286–94.

    Article  CAS  PubMed  Google Scholar 

  38. Ribeiro OC, Silva JAH, Oliveira SC, Meyer R, Fernandes GB. Preliminary results on a living vaccine against caseous lymphadenitis. Pesq Agrop Brasileira. 1991;26:461–5.

    Google Scholar 

  39. Meyer R, Carminati R, Cerqueira RB, Vale V, Viegas S, Martinez T. Evaluation of the goats humoral immune response induced by the Corynebacterium pseudotuberculosis lyophilized live vaccine. Rev Cienc Méd Biol. 2002;1:42–8.

    Google Scholar 

  40. Ruiz JC, D’Afonseca V, Silva A, Ali A, Pinto AC, Santos AR. Evidence for reductive genome evolution and lateral acquisition of virulence functions in two Corynebacterium pseudotuberculosis strains. PLoS One. 2011;18:e18551.

    Article  Google Scholar 

  41. Nascimento IP, Leite LC. The effect of passaging in liquid media and storage on Mycobacterium bovis--BCG growth capacity and infectivity. FEMS Microbiol Lett. 2005;1:81–6.

    Article  Google Scholar 

  42. Hopkins RJ, Morris Jr JG, Papadimitriou JC, Drachenberg C, Smoot DT, James SP, Panigrahi P. Loss of Helicobacter pylori hemagglutination with serial laboratory passage and correlation of hemagglutination with gastric epithelial cell adherence. Pathobiology. 1996;64:247–54.

    Article  CAS  PubMed  Google Scholar 

  43. Somerville GA, Beres SB, Fitzgerald JR, DeLeo FR, Cole RL, Hoff JS, Musser JM. In vitro Serial Passage of Staphylococcus aureus: Changes in Physiology, Virulence Factor Production, and agr Nucleotide Sequence. J Bacteriol. 2002;184:1430–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Asakura H, Kawamoto K, Okada Y, Kasuga F, Makino S, Yamamoto S, Igimi S. Intra host passage alters SigB-dependent acid resistance and host cell-associated kinetics of Listeria monocytogenes. Infect Genet Evol. 2012;12:94–101.

    Article  CAS  PubMed  Google Scholar 

  45. Muthukrishnan G, Quinn GA, Lamers RP, Diaz C, Cole AL, Chen S, Cole AM. Exoproteome of Staphylococcus aureus reveals putative determinants of nasal carriage. J Proteome Res. 2011;1:2064–78.

    Article  Google Scholar 

  46. Henderson B, Martin A. Bacterial virulence in the moonlight: multitasking bacterial moonlighting proteins are virulence determinants in infectious disease. Infect Immun. 2011;79:3476–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Peng Z, Krey V, Wei H, Tan Q, Vogelmann R, Ehrmann MA, Vogel RF. Impact of actin on adhesion and translocation of Enterococcus faecalis. Arch Microbiol. 2014;196:109–17.

    Article  CAS  PubMed  Google Scholar 

  48. Rogers EA, Das A, Ton-That H. Adhesion by pathogenic corynebacteria. Adv Exp Med Biol. 2011;715:91–103.

    Article  CAS  PubMed  Google Scholar 

  49. Lazazzera BA, Solomon J, Grossman AD. An exported peptide functions intracellularly to contribute to cell density signaling in B. subtilis. Cell. 1997;13:917–25.

    Article  Google Scholar 

  50. Rees MA, Stinear TP, Goode RJ, Coppel RL, Smith AI, Kleifeld O. Changes in protein abundance are observed in bacterial isolates from a natural host. Front Cell Infect Microbiol. 2015;14(5):71.

    Google Scholar 

  51. Danelishvili L, Stang B, Bermudez LE. Identification of Mycobacterium avium genes expressed during in vivo infection and the role of the oligopeptide transporter OppA in virulence. Microb Pathog. 2014;76:67–76.

    Article  CAS  PubMed  Google Scholar 

  52. Wilson MJ, Brandon MR, Walker J. Molecular and biochemical characterization of a protective 40-kilodalton antigen from Corynebacterium pseudotuberculosis. Infect Immun. 1995;63:206–11.

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Silva JW, Droppa-Almeida D, Borsuk S, Azevedo V, Portela RW, Miyoshi A, et al. Corynebacterium pseudotuberculosis cp09 mutant and cp40 recombinant protein partially protect mice against caseous lymphadenitis. BMC Vet Res. 2014;20(10):965.

    Article  Google Scholar 

  54. Hodgson AL, Tachedjian M, Corner LA, Radford AJ. Protection of sheep against caseous lymphadenitis by use of a single oral dose of live recombinant Corynebacterium pseudotuberculosis. Infect Immun. 1994;62:5275–80.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. McNamara PJ, Bradley GA, Songer JG. Targeted mutagenesis of the phospholipase D gene results in decreased virulence of Corynebacterium pseudotuberculosis. Mol Microbiol. 1994;12:921–30.

    Article  CAS  PubMed  Google Scholar 

  56. Leiting WU, Jianping XI. Comparative genomics analysis of Mycobacterium NrdH redoxins. Microb Pathog. 2010;48:97–102.

    Article  CAS  PubMed  Google Scholar 

  57. Pacheco LG, Castro TL, Carvalho RD, Moraes PM, Dorella FA, Carvalho NB, et al. A Role for Sigma Factor σ(E) in Corynebacterium pseudotuberculosis Resistance to Nitric Oxide/Peroxide Stress. Front Microbiol. 2012;3:126.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Silva WM, Carvalho RD, Soares SC, Bastos IF, Folador EL, Souza GH, et al. Label-free proteomic analysis to confirm the predicted proteome of Corynebacterium pseudotuberculosis under nitrosative stress mediated by nitric oxide. BMC Genomics. 2014;15:1065.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Samanovic MI, Ding C, Thiele DJ, Darwin KH. Copper in microbial pathogenesis: meddling with the metal. Cell Host Microbe. 2012;16:106–15.

    Article  Google Scholar 

  60. Wolschendorf F, Ackart D, Shrestha TB, Hascall-Dove L, Nolan S, Lamichhane S, et al. Copper resistance is essential for virulence of Mycobacterium tuberculosis. Proc Natl Acad Sci U S A. 2011;25:1621–6.

    Article  Google Scholar 

  61. Rowland JL, Niederweis M. A multicopper oxidase is required for copper resistance in Mycobacterium tuberculosis. J Bacteriol. 2013;195:3724–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors would like to thank to Pará State Genomics and Proteomics Network and Waters Corporation, Brazil.

Funding

The work was supported by the Brazilian Federal Agency for the Support and Evaluation of Graduate Education (CAPES), Pará Research Foundation (FAPESPA), Minas Gerais Research Foundation (FAPEMIG) and the National Council for Scientific and Technological Development (CNPq). Yves Le Loir is the recipient of a PVE grant (71/2013) from Programa Ciências sem Fronteiras.

Availability of data and materials

The datasets supporting the results of this article were then concatenated into a *xlsx file at peptide and protein level to fulfill the requirements and is available at supplemental material including sequence coverage and a number of identified peptides for each protein sequence identified. It also includes the native peptide information. In addition other data are included within the article.

Authors’ contributions

VA, WMS, and FAD designed the experiments. WMS and FAD performed in vivo experiments. WMS, TLPC, and NS performed microbiological analyses and sample preparation for proteomic analysis. GHMFS and WMS conducted the proteomic analysis. WMS and SCS performed bioinformatics analysis of the data. YLL, AM, and HF contributed substantially to data interpretation and revisions. VA, AS, and YLL participated in all steps of the project as coordinators, and critically reviewed the manuscript. All authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Consent for publication

Not applicable.

Ethics approval

All animals used in this study were provided by the Animal Care Facility of the Biological Sciences Institute from the Federal University of Minas Gerais and were handled by the guidelines of the UFMG Ethics Committee on Animal Testing (Permit Number: CETEA 103/2011).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vasco Azevedo.

Additional files

Additional file 1: Table S1.

Complete list of proteins differentially produced between the recovered and control condition of strain 1002_ovis. (XLSX 44 kb)

Additional file 2: Table S2.

List of proteins identified in the exclusive proteome of control condition. (XLSX 12 kb)

Additional file 3: Table S3.

Total list of peptide and proteins identified by LC-MSE. (XLSX 3 mb)

Additional file 4: Table S4.

Proteins identified in the recovered condition detected in pathogenicity island. (XLSX 10 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Silva, W.M., Dorella, F.A., Soares, S.C. et al. A shift in the virulence potential of Corynebacterium pseudotuberculosis biovar ovis after passage in a murine host demonstrated through comparative proteomics. BMC Microbiol 17, 55 (2017). https://0-doi-org.brum.beds.ac.uk/10.1186/s12866-017-0925-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12866-017-0925-6

Keywords