Therefore, it remains unclear whether treatment of MO-DCs with GA

Therefore, it remains unclear whether treatment of MO-DCs with GA at that high dose abolished stimulation-dependent upregulation of surface markers, or only partially inhibited upregulation, as was observed for most molecules in our work for a ten-fold lower dose of GA applied. In agreement with impaired upregulation of the cytoskeletal protein Fscn1, required for dendrite formation [22] and migration [41], MO-DCs cotreated with GA in the course of stimulation were characterized by a lower migratory activity than the corresponding control group. This functional defect may reflect in part impaired actin polymerization,

shown to require HSP90 activity [42]. MO-DCs treated with GA during stimulation, in accordance with reduced upregulation of DC activation

markers and proinflammatory cytokines, exhibited lower allo CD4+ T cell activation capacity Mitomycin C cost as compared with stimulated control MO-DCs. Consequently, the corresponding DC/T cell cocultures contained lower levels of the Th1/Th2 effector cytokines [43] IFN-γ, and IL-5. In general, stimulation of MO-DCs results in the activation of a number of signaling pathways, and a number of key regulators have been reported to constitute client proteins of HSP90. In this regard, STAT1 has been identified as a genuine HSP90 MG-132 cell line target [44]. Here we show that GA-treated HEK293T cells displayed impaired STAT1/2 activity under basal conditions, and impaired Nintedanib (BIBF 1120) upregulation in response to stimulation. In stimulated DCs, STAT1 has been demonstrated to mediate increased expression of activation markers like CD40 [45], and its inhibition may contribute to impaired DC maturation. Moreover, MAPK members JNK [46], and p38 [47] have been shown to positively regulate DC activation, and both kinases interact with HSP90 (JNK [48], p38 [49]). Both MAPK are known to activate PKC, which in turn mediates phosphorylation-dependent activation

of TFs of the AP-1 family that are important i.e. for expression of MMP-9 in stimulated DCs as a prerequisite for emigration from the periphery [50]. In line with the relevance of HSP90-mediated protein maturation of either MAPK, we observed impaired upregulation of AP-1 activity in HEK293T cells cotreated with GA and the maturation cocktail. Besides, stimulation-dependent MAPK activation is known increase of NF-κB activity [13], based on transient degradation of the endogenous inhibitor IκB-α [34], and in case of APCs also on elevated expression and activity of the NF-κB family member RelB [51]. In case of DCs, RelB is essential for stimulation-dependent increases of activation marker expression and consequently the T cell stimulatory capacity [33]. Therefore, our finding of GA-dependently impaired RelB expression in stimulated Mo-DCs may explain in part the detrimental effects of this agent on the phenotype and function of stimulated Mo-DCs.

By HPTLC immunostaining or RIA, mAb MEST-3 showed reactivity with

By HPTLC immunostaining or RIA, mAb MEST-3 showed reactivity with GIPCs isolated from mycelium forms of P. brasiliensis and hyphae of A. fumigatus and A. nidulans (Figure 1A-C), but it is noteworthy that no fluorescence was observed

with mycelium forms of P. brasiliensis and hyphae of A. fumigatus and A. nidulans (not shown). As expected, by immunostaining and RIA (Figures 1A-C), no reactivity of MEST-3 was observed with mycelium forms of S. schenckii and H. capsulatum. Negative controls using an irrelevant mAb showed no fluorescence (not shown). Figure 3 Indirect immunofluorescence. Indirect immunofluorescence of yeast forms of P. brasiliensis (Pb), H capsulatum (Hc) and S. schenckii (Ss), with mAb MEST-3. A- fluorescence. B- phase contrast. Effect of monoclonal antibodies on fungal growth By counting the total number of colony forming units (CFUs), the effect of mAbs MEST-1, -2 and -3 at different GSK-3 assay concentrations on fungal growth was analyzed. Under

the conditions described in Methods, it was determined for P. brasiliensis, H. capsulatum and S. schenckii, Selleck ABT199 a total of 57 ± 4, 41 ± 3 and 79 ± 4 CFUs, respectively. As shown in Figure 4A, mAbs MEST-1 and -3 were effective in inhibiting P. brasiliensis and H. capsulatum CFUs in a dose-dependent manner. mAb MEST-1 was able to inhibit P. brasiliensis and H. capsulatum CFU by about 38% and 45%, respectively, while MEST-3 inhibited P. brasiliensis, H. capsulatum and S. schenckii CFUs by about 30%, 55% and 65%, respectively (*p < 0.05). Conversely, as expected, MEST-1 was not able to inhibit S. schenckii CFU, since this fungus does not present glycolipids containing terminal residues of β-D-galactofuranose [22, 23]. It should

be noted that MEST-2 did not present significant CFU inhibitory activity in none of the three fungi used in this study. Confirming these results, P. brasiliensis, H. capsulatum and S. schenckii were grown in media containing mAbs for 48 h, after that, MTT 3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide was added to measure the growth rate. As observed in Figure 4B, MEST-1 and -3 inhibited significantly the growth of P. brasiliensis and H. capsulatum, whereas for S. schenckii, Oxymatrine only MEST-3 was able to inhibit fungal growth. Figure 4 Effect of monoclonal antibodies on fungal growth. Panel A, Yeast forms of P. brasiliensis, H. capsulatum and S. schenckii were incubated for 24 h with mAbs, or a control IgG or left alone, at 37°C. Yeasts were transferred to a petri dish containing PGY or BHI-agar medium, and incubated for 2 days at 37°C. Colony forming units (CFUs) were counted, and expressed as percentage of those incubated with an irrelevant mAb, considered as 100% of CFU. Panel B, MTT assay of fungi after incubation with mAbs MEST-1, -2, and -3. Yeast forms of P. brasiliensis, H. capsulatum and S. schenckii were incubated with mAbs, a control IgG or left alone.

(YP_004116848) 59 tet(A) 41265-42464 Tetracycline efflux protein

(YP_004116848) 59 tet(A) 41265-42464 Tetracycline efflux protein pQKp331H (ABS19074) 100 tetR 42592-43233 Repressor protein for Tet(A) pQKp331H 100 pN3_052 43438-43941 Unknown No good match   pN3_053 44147-44563 Unknown pLVPK (NP_943518) 59 tnp orfA 44921-45265 IS911 transposase, truncated Shigella flexneri 2a str. 2457 T (NP_835957) 80 pN3_055 45468-46295 Putative bacitracin resistance protein

Acinetobacter sp. DR1 (YP_003733303) 62 pN3_056 46450-47589 Putative amino acid racemase Pectobacterium carotovorum PC1 (YP_003017826)

Epigenetics inhibitor 73 pN3_057 47686-48597 Putative LysR-type regulator Shewanella halifaxensis HAW-EB4 (YP_001674862) 56 pN3_058 48594-49526 Putative amino acid dehydrogenase/cyclodeaminase Pectobacterium carotovorum subsp. brasiliensis PBR1692 (ZP_03825565) 72 pN3_059 50018-50623 Putative sodium:dicarboxylate symporter Burkholderia dolosa AUO158 (ZP_04944635) 56 tnpA 50681-51385 IS26 transposase pKOX105 100 hsdM 51636-53192 Type I restriction enzyme KU-57788 purchase EcoprrI M protein Escherichia coli B185 (ZP_06660389) 90 pN3_062 53656-54165 Unknown pKOX105 90 1 Where more than one protein shares the exact

same identity with pN3 an example is given The effect of the genetic composition of the plasmid on its fitness impact The fitness impacts of the related plasmids RP1 and pUB307 and R46 and N3 on E. coli 345-2RifC were compared. pUB307 is a derivative of RP1 which has lost the Tn1 transposon. The fitness impact of the Tn1 transposon itself has been demonstrated to be variable depending on the insertion site, with some insertion sites conferring a fitness benefit [24]. Here, pUB307 had a small fitness cost of 1.9 ± 0.8% per generation, significantly buy C59 lower than that of RP1 of -3.3 ± 0.9% per generation (students t-test p = 0.041). In animals, carriage of neither RPI nor pUB307 influenced the ability of E. coli 345-2RifC to colonize the pig gut compared to the plasmid-free 345-2RifC (ANOVA F value = 0.77, p = 0.471). R46 was previously determined to confer a fitness cost of – 3.3 ± 1.7% per generation [24] in the laboratory, whilst no significant fitness cost in pigs was detected.

0 or 4 1, in the presence or absence of BA precursors Transcript

0 or 4.1, in the presence or absence of BA precursors. Transcriptional levels were calculated relative to the mRNA levels GW-572016 chemical structure of an unstressed sample for each condition tested, using the expression

of the tuf gene as internal control (see Methods). A similar pattern of expression for both genes was observed in unstressed and stressed samples for all conditions tested (Figure 3). mRNAs corresponding to tyrDC (Figure 3A) or aguA1 (Figure 3B) were induced only if the bacterium had been challenged with tyrosine or agmatine. Under all conditions tested, higher levels of tyrDC and aguA1 transcripts were detected when both BAs precursors were present (approximately 9-fold increase in unstressed cultures

and 11-fold under gastric stress at pH 4.1). Furthermore, it should be noted that transcriptional levels of the two genes in the control cultures were not reduced Stem Cell Compound Library cell line under conditions of gastric stress. Figure 3 Relative expression of tdc (A) and aguA1 (B) genes. Total RNAs were extracted at mid-exponential phase prior treatment (untreated) and after saliva plus gastric stress at either pH 5.0 (G pH 5.0) or pH 4.1 (G pH 4.1), in presence of 4.38 mM agmatine, 10 mM tyrosine or both, or in their absence. mRNA levels were quantified as n-fold differences by comparing to RNA samples from their respective unstressed cultures (mRNA value=1). Relative levels of expression in absence of BA-precursors for untreated/G pH 5.0/G pH 4.1 were 1/0.7/0.4 in (A) and 1/0.6/0.3 in (B). Each experiment IKBKE was performed in triplicate. Vertical bars represent the standard deviation. Differences were assessed

by Anova test. Different superscript letters associated with values of either tyrDC or aguA1 mRNA levels indicate statistically significant differences (P < 0.05). These results show a transcriptional induction of tyrDC and aguA1 mediated by the respective BA-precursors under saliva and gastric stresses similar to that previously observed for IOEB 9809 under wine stress conditions [29]. The increased transcription of both genes in the presence of tyrosine plus agmatine strongly suggests a previously undetected synchronous regulation of both BA pathways, which deserves further investigation. Considering the overall results pertaining to BA production (Table 1), cell survival (Figure 1) and transcriptional analysis (Figure 3), it appears that induction of BA biosynthetic pathway at the transcriptional level by the presence of the BA precursor under mild gastric conditions results in increase of the bacterial survival. Behaviour of L. brevis IOEB 9809 in the presence of human Caco-2 intestinal epithelial cells Our results revealed that at pH 4.1 there is an approximately 35% survival of IOEB 9809 (in the presence of agmatine and tyramine) and an approximately 0.4% survival at pH 3.0 (Figure 1).

Hybridization of tiling arrays Fluorescently labeled cDNA was hyb

Hybridization of tiling arrays Fluorescently labeled cDNA was hybridized to CombiMatrix arrays as previously described[8]. In addition to the Cy5-labeled sample described above, a common Cy3-labeled sample was used as a counterpoint reference on each array. Images of the hybridized arrays were acquired with a GenePix 4000B scanner (Axon Instruments) buy PR-171 controlled by the GenePix 4.0 program (Molecular Devices). Each array was scanned three times using the following PMT settings for the 635 nm laser: 400, 450, 540. Images were gridded with GenePix 4.0 and the median foreground intensity for each feature was used as the input for subsequent analysis. Based

on the negative control probes, signal/noise was constant for the three scans, so all subsequent analysis was carried out using the lowest PMT scan. Probe detection on tiling arrays Background intensity see more was estimated based on the

median intensities of a control set of known antisense and intergenic regions, a method similar to the use of median intensities of known introns in the analysis of rice tiling data[6]. Specifically, the background intensity was estimated as the median intensity of the positive control probes corresponding to the intergenic (untranscribed) regions flanking CBP1 and TYR1 and the antisense (untranscribed) probes for CBP1, TYR1, and TEF1. A tiling probe was considered detected if it had intensity greater than the background intensity estimated for the corresponding array. 58% of the tiling probes were considered detected by this method. Transcript detection on tiling arrays In H. capsulatum, introns are small enough to make detection of

complete transcripts feasible (in contrast to, e.g., Homo sapiens) but are large and irregular enough to make such detection non-trivial (in contrast to, e.g., Escherichia coli or Saccharomyces cerevisiae). For this study, we traded resolution for improved signal to noise and defined transcripts as genomic loci ≥ 200 bp for which the normalized density of detected probes was ADP ribosylation factor greater than 65% of the normalized density of all probes. Smoothed densities were calculated with the density function in R[25] using a bandwidth of 500 bp, and transcripts were truncated such that transcript ends coincided with detected tiles. In order to avoid regions of the tiling path that were rendered sparse due to repeat masking, transcript detection was restricted to regions spanning at least 10 kb of genome sequence with a minimum tiling density of 1 probe per 250 bp (1/5 th of the target tiling density). 6,172 transcripts were detected. The length distribution (in terms of genomic locus) for detected and predicted transcripts is shown in Figure 4. Known transcripts showed a mild 3′ bias, meaning that signal intensity was enriched at the 3′ end of the gene, as expected given the method of sample preparation.

However, the force increase is not significant when the speed cha

However, the force increase is not significant when the speed changes from 1 to 10 m/s. Second, within the range of the indenter travel distance of 10 Å, the three curves under dry or wet indentation overlap each other and the indentation force almost linearly

increases with the travel distance. As the indenter tip further advances, the three curves start to deviate from each other. Figure 12 Effect of indentation speed on indentation force evolution. (a) Dry condition for cases 6, 2, and 4. (b) Wet condition for cases 5, 1, and 3. Moreover, Ku-0059436 solubility dmso we also analyze how the indentation speed affects friction behaviors along the indenter/work interface. Figure 13 shows the normal and friction force distributions under dry condition for cases 6, 2, and 4. It can be seen that under dry indentation, the normal force of case 4 (100 m/s speed) is significantly higher than those of cases 6 and 2 (1 and 10 m/s, respectively) at surface locations close to the indenter tip. The difference diminishes at the position about 2.5 nm to the indenter tip, in which all three indentation speeds have approximately the same normal force. When the surface position to the indenter tip further increases, the normal force at 100 m/s becomes smaller than those at 1 and 10 m/s, and the 1 m/s curve is overall

slightly lower than the 10 m/s curve in terms of normal force. The trend in normal force is consistent with that observed in indentation force comparison, as shown in Figure 12a. In terms of friction force distributions, the three curves have a similar shape, and the Torin 1 cost peak friction force is located around 3.4 to 4.4 nm to the indenter tip depending on the indentation speed. Also, the overall (total) friction force decreases with the increase of indentation speed. Figure 13 Indentation speed effect on (a) normal and (b) friction force distributions under dry indentation. In the mean time, Figure 14 compares the normal and friction

distributions under 6-phosphogluconolactonase wet indentation at the indentation speeds of 1 m/s (case 5), 10 m/s (case 1), and 100 m/s (case 3). Compared with Figure 13a, similar observations can be made among the three normal force curves under wet indentation. Also, the friction force curves in Figure 14b have fairly consistent shapes, and the peak friction force is always located at around 4.4 nm to the indenter tip. Figure 14 Indentation speed effect on (a) normal and (b) friction force distributions under wet indentation. Conclusions This research investigates nano-indentation processes with the existence of water molecules by using the numerical approach of MD simulation. The potential tribological benefits of water or other liquids, as well as the influence on material property measurements, are intriguing to nano-indentation. This also applies to other tool-based precision manufacturing processes. By configuring 3D indentation of single-crystal copper with a diamond indenter, six simulation cases are developed.

Literature data shows that although SecA is essential for bacteri

Literature data shows that although SecA is essential for bacteria, its SecB-binding domain is dispensable for protein secretion and cell viability [43, 44]. Thus, we consider that the secA mutants that were picked up in our suppressor screen are impaired only in SecB-dependent protein secretion and in respect of the cell lysis phenotype they resemble secB-knockouts. Finally, unique insertions of transposon into PP1585 and PP4236, coding for putative antidote protein of a toxin-antitoxin system and a thiol:disulfide interchange protein, respectively,

also resulted in white non-lysing colonies of the colR mutant. In conclusion, inactivation of different MLN0128 datasheet genes prevented lysis of the colR mutant and most of these genes encode either membrane proteins or are implicated in regulating membrane proteins. Analysis of the outer membrane composition of the non-lysing transposon derivatives

of the colR mutant The results of the suppressor analysis predict that the colR mutant cannot maintain membrane protein homeostasis. This is supported by two phenomena. First, the reduction of protein secretion by the inactivation of the SecB-dependent protein export suppresses cell lysis. Second, the disruption of genes for the outer membrane porins, OprB1 and OprF, also eliminated the lysis indicating that the outer membrane (OM) composition may be unbalanced in the colR-deficient P. putida. In order to address this issue we compared the pattern of OM selleck proteins of the wild-type

and the colR mutant as well as the suppression mutants of the colR strain. Data in Figure 3 demonstrate that the overall OM protein pattern of the wild-type and the colR strains is similar. The PP1585, PP4236, secA and secB derivatives Ribose-5-phosphate isomerase of the colR mutant also have OM protein profiles that are quite similar to the wild-type. However, as expected, OM protein preparations of the colRoprB1 and colRoprF mutants respectively lacked OprB1 and OprF channel proteins. Note that OprF is represented by several differently migrating forms. This is consistent with previous data on several OM porins, including OprF of P. aeruginosa, showing that these proteins are prone to modification by heat and β-mercaptoethanol treatment that is carried out for the solubilization of proteins before applying to the gel [45]. Given that sigX and oprF genes comprise one operon and that OprF is positively regulated by SigX in P. aeruginosa and P. fluorescens [41], it was expected that all four different colRsigX knockout strains have significantly lowered OprF amount in their OM (Figure 3, only two colRsigX derivatives are presented). However, while three sigX derivatives of the colR mutant (minitransposon insertions after nucleotides 251, 304 and 336 of the sigX gene) revealed only modestly reduced expression of OprF (Figure 3, only colRsigX 336 is presented), the colRsigX strain with most distal transposon insertion in sigX, displayed drastically decreased OprF level (Figure 3, see colRsigX 480).

Figure 4 Phylogeny of RNA phages The phylogenetic analysis was b

Figure 4 Phylogeny of RNA phages. The phylogenetic analysis was based on the complete genomic RNA sequences (left) and amino acid sequences of the replicase (right) which is the most conserved of all ssRNA phage proteins. Trees were constructed by unweighted pair group method with arithmetic mean (UPGMA) and tested using the bootstrap method with 500 replicates. The bootstrap values are expressed as percentages next to the nodes. RNA and protein sequences were aligned using MUSCLE [49] and BTK inhibitor the phylogenetic trees were constructed in program MEGA5 [50]. Although all Leviviridae phages use pili for attachment, there is a marked difference between the types

of pili they utilize. The type IV pili used by phages AP205, ϕCb5 and PP7 are produced via a genome-encoded type II secretion pathway [51], whereas the plasmid-borne conjugative pili that the other phages utilize belong to a type IV secretion system [52]. Both systems share some functional similarities, like a retractable pilus and a membrane pore, but are thought to have evolved independently [53]. Therefore a jump from one to the other type of pili had to occur at some point in the Leviviridae history. Our phylogenetic analysis suggests that the ancestral phage infected cells via type IV pili, like phages AP205, ϕCb5 and PP7 are doing today and a PP7-like virus then might have evolved the ability to bind to some kind of conjugative

pili and still sustain infectivity. Consequently, all of the specialized Branched chain aminotransferase plasmid-dependent RNA phages we know today would be descendants of this ancestral virus. Conclusions We have determined and characterized the genome sequence Autophagy Compound Library of IncM plasmid-dependent phage M and shown that it resembles the plasmid-specific leviviruses in many ways but has an atypical location of the lysis gene. It is a valuable addition to

the growing number of sequenced Leviviridae genomes and provides a better view on the diversity and evolution within this phage family. Methods Phage propagation and purification Bacteriophage M and its host E.coli J53(RIP69) were obtained from Félix d’Hérelle Reference Center for bacterial viruses, Laval University, Quebec, Canada (catalog numbers HER218 and HER1218, respectively). J53(RIP69) cells were grown in LB medium containing 6 μg/ml tetracycline overnight at 37 °C without agitation. To propagate the phage, 0.5 ml of the host cell suspension and 10 μl of phage lysate (approximately 1010 pfu/ml) were spotted on 1.5% LB agar plates, overlaid with 15-20 ml of molten 0.7% LB agar cooled to 45 °C, mixed by swirling and incubated overnight at 30 °C. The next morning, top agar layers from several plates were scraped off, transferred to centrifuge tubes and centrifuged for 20 minutes at 18500 g. Supernatant was collected and phage particles were precipitated by addition of sodium chloride and PEG 6000 to concentrations of 0.

For example, synthetic AI-2 directly stimulates Escherichia coli

For example, synthetic AI-2 directly stimulates Escherichia coli biofilm formation and controls biofilm architecture by stimulating bacterial motility [31]. Subsequently, several studies also indicated that AI-2 indeed controls biofilm formation [32–34]. In contrast,

some researchers reported that addition of AI-2 failed to restore biofilm phenotype of the parental strain [35–40], owing to the central metabolic effect of LuxS or difficulty in complementation of AI-2 GSK126 datasheet [41]. There exists a conserved luxS gene in S. aureus, and it has been proved to be functional for generating AI-2 [42]. Previous work indicated that AI-2-mediated QS modulated capsular polysaccharide synthesis and virulence in S. aureus[43], deletion of the luxS gene led to increased biofilm formation in Staphylococcus epidermis[20], and biofilm enhancement due to luxS repression was manifested by an increase in PIA [44]. In this study, we provide evidence that S. aureus ΔluxS strain formed stronger biofilms than the WT strain RN6390B, and that the luxS mutation was complemented by adding chemically synthesized DPD, the exogenous precursor of AI-2. AI-2 activated the transcription of icaR, and subsequently www.selleckchem.com/products/apo866-fk866.html led to decreased icaA transcription,

as determined by real-time RT-PCR analysis. Furthermore, the differences in biofilm-forming ability of S. aureus RN6911, ΔluxS strain, and the ΔagrΔluxS strain were also investigated. Our data suggest that Nintedanib chemical structure AI-2 could inhibit biofilm formation in S. aureus RN6390B through the IcaR-dependent regulation of the ica operon. Methods Bacterial strains, plasmids and DNA manipulations The bacterial strains and plasmids used in this study are described in Table 1. E. coli cells were grown in Luria-Bertani (LB) medium (Oxoid) with appropriate antibiotics for cloning selection. S. aureus strain RN4220, a cloning intermediate, was used for propagation of plasmids prior to transformation into other S. aureus strains.

S. aureus cells were grown at 37°C in tryptic soy broth containing 0.25% dextrose (TSBg) (Difco No. 211825). In the flow cell assay, biofilm bacteria were grown in tryptic soy broth without dextrose (TSB) (Difco No. 286220). Medium was supplemented when appropriate with ampicillin (150 μg/ml), kanamycin (50 μg/ml), erythromycin (2.5 μg/ml) and chloramphenicol (15 μg/ml). Table 1 Strains and plasmids used in this study Strain or plasmid Description Reference or source RN6390B Standard laboratory strain NARSAa RN4220 8325-4 r- NARSA ΔluxS RN6390B luxS::ermB This study RN6911 RN6390B derivative; agr locus replaced with tetM cassette NARSA ΔagrΔluxS RN6911 luxS::ermB, agr/luxS double mutant This study ΔluxSpluxS Complemented strain of ΔluxS; Apr Cmr This study RN6390BG RN6390B/pgfp This study ΔluxSG ΔluxS/pgfp This study RN6911G RN6911/pgfp This study ΔagrΔluxSG ΔagrΔluxS/pgfp This study NCTC8325 Standard Laboratory strain NARSA NCTC8325ΔluxS NCTC8325 luxS::ermB 60 E.

Antimicrob Agents Chemother 2010,54(11):4794–4798 PubMedCentralPu

Antimicrob Agents Chemother 2010,54(11):4794–4798.PubMedCentralPubMedCrossRef 7. Lari N, Rindi L, Bonanni D, Rastogi N, Sola C, Tortoli E, Garzelli C: Three-year longitudinal study of genotypes of Mycobacterium tuberculosis Vincristine concentration isolates in Tuscany, Italy. J Clin Microbiol 2007,45(6):1851–1857.PubMedCentralPubMedCrossRef 8. Gibson AL, Huard RC, Gey van Pittius NC, Lazzarini LC, Driscoll J,

Kurepina N, Zozio T, Sola C, Spindola SM, Kritski AL, et al.: Application of sensitive and specific molecular methods to uncover global dissemination of the major RDRio Sublineage of the Latin American-Mediterranean Mycobacterium tuberculosis spoligotype family. J Clin Microbiol 2008,46(4):1259–1267.PubMedCentralPubMedCrossRef 9. Lazzarini LC, Huard RC, Boechat NL, Gomes HM, Oelemann MC, Kurepina N, Shashkina E, Mello FC, Gibson AL, Virginio MJ, et al.: Discovery of a novel Mycobacterium tuberculosis lineage that is a major cause of tuberculosis

in Rio de Janeiro, Brazil. J Clin Microbiol 2007,45(12):3891–3902.PubMedCentralPubMedCrossRef 10. Cubillos-Ruiz A, Sandoval A, Ritacco V, Lopez B, Robledo J, Correa N, Hernandez-Neuta I, Zambrano MM, Del Portillo P: Genomic signatures of the haarlem lineage of Mycobacterium tuberculosis: implications of strain genetic variation in drug and vaccine development. J Clin Microbiol 2010,48(10):3614–3623.PubMedCentralPubMedCrossRef Angiogenesis inhibitor 11. Devaux I, Kremer K, Heersma H, Van Soolingen D: Clusters of multidrug-resistant Mycobacterium tuberculosis cases, Europe. Emerg Infect Dis 2009,15(7):1052–1060.PubMedCrossRef 12. Filliol I, Sola C, Rastogi N: Detection

of a previously unamplified spacer within the DR locus of Mycobacterium tuberculosis: epidemiological implications. J Clin Microbiol 2000,38(3):1231–1234.PubMedCentralPubMed Chloroambucil 13. Gutacker MM, Mathema B, Soini H, Shashkina E, Kreiswirth BN, Graviss EA, Musser JM: Single-nucleotide polymorphism-based population genetic analysis of Mycobacterium tuberculosis strains from 4 geographic sites. J Infect Dis 2006,193(1):121–128.PubMedCrossRef 14. Alland D, Lacher DW, Hazbon MH, Motiwala AS, Qi W, Fleischmann RD, Whittam TS: Role of large sequence polymorphisms (LSPs) in generating genomic diversity among clinical isolates of Mycobacterium tuberculosis and the utility of LSPs in phylogenetic analysis. J Clin Microbiol 2007,45(1):39–46.PubMedCentralPubMedCrossRef 15. Bouakaze C, Keyser C, de Martino SJ, Sougakoff W, Veziris N, Dabernat H, Ludes B: Identification and genotyping of Mycobacterium tuberculosis complex species by use of a SNaPshot Minisequencing-based assay. J Clin Microbiol 2010,48(5):1758–1766.PubMedCentralPubMedCrossRef 16. Filliol I, Motiwala AS, Cavatore M, Qi W, Hazbon MH, Bobadilla del Valle M, Fyfe J, Garcia-Garcia L, Rastogi N, Sola C, et al.