首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 78 毫秒
1.
The Norway spruce genome provides key insights into the evolution of plant genomes, leading to testable new hypotheses about conifer, gymnosperm, and vascular plant evolution.In the past year a burst of plant genome sequences have been published, providing enhanced phylogenetic coverage of green plants (Figure (Figure1)1) and inclusion of new agricultural, ecological, and evolutionary models. Collectively, these sequences are revealing some extraordinary structural and evolutionary attributes in plant genomes. Perhaps most surprising is the exceptionally high frequency of whole-genome duplication (WGD): nearly every genome that has been analyzed has borne the signature of one or more WGDs, with particularly notable events having occurred in the common ancestors of seed plants, of angiosperms, and of core eudicots (the latter ''WGD'' represents two WGDs in close succession) [1,2]. Given this tendency for plant genomes to duplicate and then return to an essentially diploid genetic system (an example is the cotton genomes, which have accumulated the effects of perhaps 15 WGDs [3]), the conservation of genomes in terms of gene number, chromosomal organization, and gene content is astonishing. From the publication of the first plant genome, Arabidopsis thaliana [4], the number of inferred genes has been between 25,000 and 30,000, with many gene families shared across all land plants, although the number of members and patterns of expansion and contraction vary. Furthermore, conserved synteny has been detected across the genomes of diverse angiosperms, despite WGDs, diploidization, and millions of years of evolution.Open in a separate windowFigure 1Simplified phylogeny of land plants, showing major clades and their component lineages. Asterisks indicate species (or lineage) for which whole-genome sequence (or sequences) is (are) available. Increases and decreases in genome size are shown by arrows.Despite the proliferation of genome sequences available for angiosperms, genome-level data for both ferns (and their relatives, collectively termed monilophytes; Figure Figure1)1) and gymnosperms have been conspicuously lacking - until recently, with the publication of the genome sequence of the gymnosperm Norway spruce (Picea abies) [5]. The large genome sizes for both monilophytes and gymnosperms have discouraged attempts at genome sequencing and assembly, whereas the smaller genome size of angiosperms has resulted in more genome sequences being available (Table (Table1)1) [6]. Because of this limited phylogenetic sample, our understanding of the timing and phylogenetic positions of WGDs, the core number of plant genes, possible conserved syntenic regions, and patterns of expansion and contraction of gene families across both tracheophytes (vascular plants) and across all land plants is imperfect. This sampling problem is particularly acute in analyses of the genes and genomes of seed plants; many hundreds of genes are present in angiosperms that are not present in mosses or lycophytes, but whether these genes arose in the common ancestor of seed plants or of angiosperms cannot be determined without a gymnosperm genome sequence. The Norway spruce genome therefore offers tremendous power, not only for understanding the structure and evolution of conifer genomes, but also as a reference for interpreting gene and genome evolution in angiosperms.

Table 1

Genome sizes in land plants
LineageRange (1C; pg)Mean
Gymnosperms
  Conifers
    Pinaceae9.5-36.023.7
    Cupressaceae8.3-32.112.8
    Sciadopitys 20.8n/a
  Gnetales
    Ephedraceae8.9-15.78.9
    Gnetaceae2.3-4.02.3
    Cycadaceae12.6-14.813.4
    Ginkgo biloba11.75n/a
Monilophytes
    Ophioglossaceae10.2-65.631.05
    Equisetaceae12.9-30422.0
    Psilotum72.7n/a
  Leptosporangiate ferns
    Polypodiaceae7.5-19.77.5
    Aspleniaceae4.1-9.16.2
    Athyriaceae6.3-9.37.6
    Dryopteridaceae6.8-23.611.7
  Water ferns
    Azolla0.77n/a
Angiosperms
    Oryza sativa 0.50n/a
    Amborella trichopoda0.89n/a
    Arabidopsis thaliana0.16n/a
    Zea mays2.73n/a
Open in a separate windown/a, not applicable. Data based on [6].  相似文献   

2.
Peptidoglycan from Deinococcus radiodurans was analyzed by high-performance liquid chromatography and mass spectrometry. The monomeric subunit was: N-acetylglucosamine–N-acetylmuramic acid–l-Ala–d-Glu-(γ)–l-Orn-[(δ)Gly-Gly]–d-Ala–d-Ala. Cross-linkage was mediated by (Gly)2 bridges, and glycan strands were terminated in (1→6)anhydro-muramic acid residues. Structural relations with the phylogenetically close Thermus thermophilus are discussed.The gram-positive bacterium Deinococcus radiodurans is remarkable because of its extreme resistance to ionizing radiation (14). Phylogenetically the closest relatives of Deinococcus are the extreme thermophiles of the genus Thermus (4, 11). In 16S rRNA phylogenetic trees, the genera Thermus and Deinococcus group together as one of the older branches in bacterial evolution (11). Both microorganisms have complex cell envelopes with outer membranes, S-layers, and ornithine-Gly-containing mureins (7, 12, 19, 20, 22, 23). However, Deinococcus and Thermus differ in their response to the Gram reaction, having positive and negative reactions, respectively (4, 14). The murein structure for Thermus thermophilus HB8 has been recently elucidated (19). Here we report the murein structure of Deinococcus radiodurans with similar detail.D. radiodurans Sark (23) was used in the present study. Cultures were grown in Luria-Bertani medium (13) at 30°C with aeration. Murein was purified and subjected to amino acid and high-performance liquid chromatography (HPLC) analyses as previously described (6, 9, 10, 19). For further analysis muropeptides were purified, lyophilized, and desalted as reported elsewhere (6, 19). Purified muropeptides were subjected to plasma desorption linear time-of-flight mass spectrometry (PDMS) as described previously (1, 5, 16, 19). Positive and negative ion mass spectra were obtained on a short linear 252californium time-of-flight instrument (BioIon AB, Uppsala, Sweden). The acceleration voltage was between 17 and 19 kV, and spectra were accumulated for 1 to 10 million fission events. Calibration of the mass spectra was done in the positive ion mode with H+ and Na+ ions and in the negative ion mode with H and CN ions. Calculated m/z values are based on average masses.Amino acid analysis of muramidase (Cellosyl; Hoechst, Frankfurt am Main, Germany)-digested sacculi (50 μg) revealed Glu, Orn, Ala, and Gly as the only amino acids in the muramidase-solubilized material. Less than 3% of the total Orn remained in the muramidase-insoluble fraction, indicating an essentially complete solubilization of murein.Muramidase-digested murein samples (200 μg) were analyzed by HPLC as described in reference 19. The muropeptide pattern (Fig. (Fig.1)1) was relatively simple, with five dominating components (DR5 and DR10 to DR13 [Fig. 1]). The muropeptides resolved by HPLC were collected, desalted, and subjected to PDMS. The results are presented in Table Table11 compared with the m/z values calculated for best-matching muropeptides made up of N-acetylglucosamine (GlucNAc), N-acetylmuramic acid (MurNAc), and the amino acids detected in the murein. The more likely structures are shown in Fig. Fig.1.1. According to the m/z values, muropeptides DR1 to DR7 and DR9 were monomers; DR8, DR10, and DR11 were dimers; and DR12 and DR13 were trimers. The best-fitting structures for DR3 to DR8, DR11, and DR13 coincided with muropeptides previously characterized in T. thermophilus HB8 (19) and had identical retention times in comparative HPLC runs. The minor muropeptide DR7 (Fig. (Fig.1)1) was the only one detected with a d-Ala–d-Ala dipeptide and most likely represents the basic monomeric subunit. The composition of the major cross-linked species DR11 and DR13 confirmed that cross-linking is mediated by (Gly)2 bridges, as proposed previously (20). Open in a separate windowFIG. 1HPLC muropeptide elution patterns of murein purified from D. radiodurans. Muramidase-digested murein samples were subjected to HPLC analysis, and the A204 of the eluate was recorded. The most likely structures for each muroeptide as deduced by PDMS are shown. The position of residues in brackets is the most likely one as deduced from the structures of other muropeptides but could not be formally demonstrated. R = GlucNac–MurNac–l-Ala–d-Glu-(γ)→.

TABLE 1

Calculated and measured m/z values for the molecular ions of the major muropeptides from D. radiodurans
MuropeptideaIonm/z
ΔmbError (%)cMuropeptide composition
Muropeptide abundance (mol%)
CalculatedMeasuredNAGdNAMeGluOrnAlaGly
DR1[M+H]+699.69700.10.410.0611101012.0
DR2[M+H]+927.94928.30.360.041111125.7
DR3[M+Na]+1,006.971,007.50.530.051111133.0
DR4[M+Na]+963.95964.60.650.071111212.5
DR5[M+H]+999.02999.80.780.0811112227.7
[M−H]997.00997.30.300.03
DR6[M+Na]+1,078.51,078.80.750.071111232.4
DR7[M+H]+1,070.091,071.00.900.081111322.2
DR8[M+Na]+1,520.531,521.61.080.071122442.2
DR9[M+Na]+701.64702.10.460.0311f10105.0
DR10[M+H]+1,907.941,907.80.140.0122223410.1
[M−H]1,905.921,906.60.680.04
DR11[M+H]+1,979.011,979.10.090.0122224419.1
[M−H]1,977.001,977.30.300.02
DR12[M+H]+2,887.932,886.5−1.43−0.053333564.4
[M−H]2,885.912,885.8−0.11−0.01
DR13[M+H]+2,959.002,957.8−1.20−0.043333663.6
[M−H]2,956.992,955.9−1.09−0.04
Open in a separate windowaDR5 and DR10 to DR13 were analyzed in both the positive and negative ion modes. Muropeptides DR1 to DR4 and DR6 to DR9 were analyzed in the positive mode only due to the small amounts of sample available. bMass difference between measured and calculated quasimolecular ion values. c[(Measured mass−calculated mass)/calculated mass] × 100. dN-Acetylglucosamine. eN-Acetylmuramitol. f(1→6)Anhydro-N-acetylmuramic acid. Structural assignments of muropeptides DR1, DR2, DR8 to DR10, and DR12 deserve special comments. The low m/z value measured for DR1 (700.1) fitted very well with the value calculated for GlucNAc–MurNAc–l-Ala–d-Glu (699.69). Even smaller was the mass deduced for DR9 from the m/z value of the molecular ion of the sodium adduct (702.1) (Fig. (Fig.2).2). The mass difference between DR1 and DR9 (19.9 mass units) was very close indeed to the calculated difference between N-acetylmuramitol and the (1→6)anhydro form of MurNAc (20.04 mass units). Therefore, DR9 was identified as GlucNAc–(1→6)anhydro-MurNAc–l-Ala–d-Glu (Fig. (Fig.1).1). Muropeptides with (1→6)anhydro muramic acid have been identified in mureins from diverse origins (10, 15, 17, 19), indicating that it might be a common feature among peptidoglycan-containing microorganisms. Open in a separate windowFIG. 2Positive-ion linear PDMS of muropeptide DR9. Muropeptide DR9 was purified, desalted by HPLC, and subjected to PDMS to determine the molecular mass. The masses for the dominant molecular ions are indicated.The measured m/z value for the [M+Na]+ ion of DR8 was 1,521.6, very close to the mass calculated for a cross-linked dimer without one disaccharide moiety (1,520.53) (Fig. (Fig.1;1; Table Table1).1). Such muropeptides, also identified in T. thermophilus HB8 and other bacteria (18, 19), are most likely generated by the enzymatic clevage of MurNAc–l-Ala amide bonds in murein by an N-acetylmuramyl–l-alanine amidase (21). In particular, DR8 could derive from DR11. The difference between measured m/z values for DR8 and DR11 was 478.7, which fits with the mass contribution of a disaccharide moiety (480.5) within the mass accuracy of the instrument.The m/z values for muropeptides DR2, DR10, and DR12 supported the argument for structures in which the two d-Ala residues from the d-Ala–d-Ala C-terminal dipeptide were lost, leaving Orn as the C-terminal amino acid.The position of one Gly residue in muropeptides DR2, DR8, and DR10 to DR13 could not be formally demonstrated. One of the Gly residues could be at either the N- or the C-terminal positions. However, the N-terminal position seems more likely. The structure of the basic muropeptide (DR7), with a (Gly)2 acylating the δ-NH2 group of Orn, suggests that major muropeptides should present a (Gly)2 dipeptide. The scarcity of DR3 and DR6, which unambiguously have Gly as the C-terminal amino acid (Fig. (Fig.1),1), supports our assumption.Molar proportions for each muropeptide were calculated as proposed by Glauner et al. (10) and are shown in Table Table1.1. For calculations the structures of DR10 to DR13 were assumed to be those shown in Fig. Fig.1.1. The degree of cross-linkage calculated was 47.2%. Trimeric muropeptides were rather abundant (8 mol%) and made a substantial contribution to total cross-linkage. However, higher-order oligomers were not detected, in contrast with other gram-positive bacteria, such as Staphylococcus aureus, which is rich in such oligomers (8). The proportion of muropeptides with (1→6)anhydro-muramic acid (5 mol%) corresponded to a mean glycan strand length of 20 disaccharide units, which is in the range of values published for other bacteria (10, 17).The results of our study indicate that mureins from D. radiodurans and T. thermophilus HB8 (19) are certainly related in their basic structures but have distinct muropeptide compositions. In accordance with the phylogenetic proximity of Thermus and Deinococcus (11), both mureins are built up from the same basic monomeric subunit (DR7 in Fig. Fig.1),1), are cross-linked by (Gly)2 bridges, and have (1→6)anhydro-muramic acid at the termini of glycan strands. Most interestingly, Deinococcus and Thermus are the only microorganisms identified at present with the murein chemotype A3β as defined by Schleifer and Kandler (20). Nevertheless, the differences in muropeptide composition were substantial. Murein from D. radiodurans was poor in d-Ala–d-Ala- and d-Ala–Gly-terminated muropeptides (2.2 and 2.4 mol%, respectively) but abundant in Orn-terminated muropeptides (23.8 mol%) and in muropeptides with a peptide chain reduced to the dipeptide l-Ala–d-Glu (18 mol%). In contrast, neither Orn- nor Glu-terminated muropeptides have been detected in T. thermophilus HB8 murein, which is highly enriched in muropeptides with d-Ala–d-Ala and d-Ala–Gly (19). Furthermore, no traces of phenyl acetate-containing muropeptides, a landmark for T. thermophilus HB8 murein (19), were found in D. radiodurans. Cross-linkage was definitely higher in D. radiodurans than in T. thermophilus HB8 (47.4 and 27%, respectively), largely due to the higher proportion of trimers in the former.The similarity in murein basic structure suggests that the difference between D. radiodurans and T. thermophilus HB8 with respect to the Gram reaction may simply be a consequence of the difference in the thickness of cell walls (2, 3, 23). Interestingly, D. radiodurans murein turned out to be relatively simple for a gram-positive organism, possibly reflecting the primitive nature of this genus as deduced from phylogenetic trees (11). Our results illustrate the phylogenetic proximity between Deinococcus and Thermus at the cell wall level but also point out the structural divergences originated by the evolutionary history of each genus.  相似文献   

3.
Bacteriophage-mediated horizontal gene transfer is one of the primary driving forces of bacterial evolution. The pac-type phages are generally thought to facilitate most of the phage-mediated gene transfer between closely related bacteria, including that of mobile genetic elements-encoded virulence genes. In this study, we report that staphylococcal cos-type phages transferred the Staphylococcus aureus pathogenicity island SaPIbov5 to non-aureus staphylococcal species and also to different genera. Our results describe the first intra- and intergeneric transfer of a pathogenicity island by a cos phage, and highlight a gene transfer mechanism that may have important implications for pathogen evolution.Classically, transducing phages use the pac site-headful system for DNA packaging. Packaging is initiated on concatemeric post-replicative DNA by terminase cleavage at the sequence-specific pac site, a genome slightly longer than unit length is packaged, and packaging is completed by non-sequence-specific cleavage (reviewed in Rao and Feiss, 2008). Generalized transduction results from the initiation of packaging at pac site homologs in host chromosomal or plasmid DNA, and typically represents ∼1% of the total number of phage particles. In the alternative cos site mechanism packaging is also initiated on concatemeric post-replicative DNA by terminase cleavage at a sequence-specific (cos) site. Here, however, packaging is completed by terminase cleavage at the next cos site, generating a precise monomer with the cohesive termini used for subsequent circularization (Rao and Feiss, 2008). Although cos site homologs may exist in host DNA, it is exceedingly rare that two such sites would be appropriately spaced. Consequently, cos phages, of which lambda is the prototype, do not engage in generalized transduction. For this reason, cos-site phages have been preferred for possible phage therapy, since they would not introduce adventitious host DNA into target organisms.The Staphylococcus aureus pathogenicity islands (SaPIs) are the best-characterized members of the phage-inducible chromosomal island family of mobile genetic elements (MGEs; Novick et al., 2010). SaPIs are ∼15 kb mobile elements that encode virulence factors and are parasitic on specific temperate (helper) phages. Helper phage proteins are required to lift their repression (Tormo-Más et al., 2010, 2013), thereby initiating their excision, circularization and replication. Phage-induced lysis releases vast numbers of infectious SaPI particles, resulting in high frequencies of transfer. Most SaPI helper phages identified to date are pac phages, and many well-studied SaPIs are packaged by the headful mechanism (Ruzin et al., 2001; Ubeda et al., 2007). Recently, we have reported that some SaPIs, of which the prototype is SaPIbov5 (Viana et al., 2010), carry phage cos sequences in their genomes, and can be efficiently packaged and transferred by cos phages to S. aureus strains at high frequencies (Quiles-Puchalt et al., 2014). Here we show that this transfer extends to non-aureus staphylococci and to Listeria monocytogenes.Since the pac phages transfer SaPIs to non-aureus staphylococci and to the Gram-positive pathogen Listeria monocytogenes (Maiques et al., 2007; Chen and Novick, 2009), we reasoned that cos phages might also be capable of intra- and intergeneric transfer. We tested this with SaPIbov5, into which we had previously inserted a tetracycline resistance (tetM) marker to enable selection, and with lysogens of two helper cos phages, φ12 and φSLT, carrying SaPIbov5 (strains JP11010 and JP11194, respectively; Supplementary Table 1). The prophages in these strains were induced with mitomycin C, and the resulting lysates were adjusted to 1 μg ml−1 DNase I and RNase A, filter sterilized (0.2 μm pore), and tested for SaPI transfer with tetracycline selection, as previously described (Ubeda et al., 2008). To test for trans-specific or trans-generic transduction, coagulase-negative staphylococci species and L. monocytogenes strains were used as recipients for SaPIbov5 transfer, respectively, as previously described (Maiques et al., 2007; Chen and Novick, 2009). As shown in Figure 1 and Supplementary Table 2). In contrast, deletion of the SaPIbov5 cos site (strains JP11229 and JP11230) did not affect SaPI replication (Supplementary Figure 1), but completely eliminated SaPIbov5 transfer (Supplementary Table 2). The TerS protein is essential for φ12 and SaPIbov5 DNA packaging, but not for phage-mediated lysis (Quiles-Puchalt et al., 2014). As expected, this mutation abolished SaPIbov5 transfer (Open in a separate windowFigure 1(a) Map of SaPIbov5. Arrows represent the localization and orientation of ORFs greater than 50 amino acids in length. Rectangles represent the position of the ori (in purple) or cos (in red) sites. Positions of different primers described in the text are shown. (b) Amplimers generated for detection of SaPIbov5 in the different recipient strains. Supplementary Table 2 lists the sequence of the different primers used. The element was detected in S. epidermidis JP829 (Se-1), S. epidermidis JP830 (Se-2), L. monocytogenes SK1351 (Lm-1), L. monocytogenes EGDe (Lm-2), S. xylosus C2a (Sx) and S. aureus JP4226 (Sa).

Table 1

Intra- and intergeneric SaPIbov5 transfera
Donor strain
  
PhageSaPIRecipient strainSaPI titreb
φ12SaPIbov5S. aureus JP42268.3 × 104
  S. epidermidis JP8292.4 × 104
  S. epidermidis JP8304.7 × 104
  L. monocytogenes SK13516.6 × 103
  L. monocytogenes EGDe2.1 × 104
  S. xylosus C2a7.1 × 104
    
φ12SaPIbov5 ΔcosS. aureus JP4226<10
  S. epidermidis JP829<10
  S. epidermidis JP830<10
  L. monocytogenes SK1351<10
  L. monocytogenes EGDe<10
  S. xylosus C2a<10
    
φ12 ΔterSSaPIbov5S. aureus JP4226<10
  S. epidermidis JP829<10
  S. epidermidis JP830<10
  L. monocytogenes SK1351<10
  L. monocytogenes EGDe<10
  S. xylosus C2a<10
    
φSLTSaPIbov5S. aureus JP42264.1 × 103
  S. epidermidis JP8291.1 × 103
  S. epidermidis JP8302.1 × 103
  L. monocytogenes SK13513.6 × 102
  L. monocytogenes EGDe3.1 × 103
  S. xylosus C2a4.0 × 103
    
φSLTSaPIbov5 ΔcosS. aureus JP4226<10
  S. epidermidis JP829<10
  S. epidermidis JP830<10
  L. monocytogenes SK1351<10
  L. monocytogenes EGDe<10
  S. xylosus C2a<10
Open in a separate windowAbbreviation: SAPI, Staphylococcus aureus pathogenicity island.aThe means of results from three independent experiments are shown. Variation was within ±5% in all cases.bNo. of transductants per ml induced culture.Because plaque formation is commonly used to determine phage host range, we next determined the ability of phages φ12 and φSLT to parasitize and form plaques on S. xylosus, S. epidermidis and L. monocytogenes strains. As shown in Supplementary Figure 2, phages φ12 and φSLT can parasitize and form plaques on their normal S. aureus hosts, but are completely unable to lyse the non-aureus strains. Therefore, as previously observed with pac phages (Chen and Novick, 2009), these results indicate that the overall host range of a cos phage may also be much wider if it includes infection without plaque formation.Previous studies have demonstrated pac phage-mediated transfer of MGEs between S. aureus and other bacterial species (Maiques et al., 2007; Chen and Novick, 2009; Uchiyama et al., 2014); however, no previous studies have described the natural intra- or intergeneric transfer of pathogenicity islands by cos phages. As bacterial pathogens become increasingly antibiotic resistant, lytic and poorly transducing phages, such as cos phages, have been proposed for phage therapy, on the grounds that they would not introduce adventitious host DNA into target organisms and that the phages are so restricted in host range that the resulting progeny are harmless and will not result in dysbiosis of human bacterial flora. Because plaque formation was once thought to determine the host range of a phage, the evolutionary impact of phages on bacterial strains they can transduce, but are unable to parasitize, has remained an unrecognized aspect of phage biology and pathogen evolution. Our results add to the recently recognized concept of ‘silent transfer'' of pathogenicity factors carried by MGEs (Maiques et al., 2007; Chen and Novick, 2009) by phages that cannot grow on the target organism. They extend this capability to cos phages, which have hitherto been unrecognized as mediators of natural genetic transfer.The potential for gene transfer of MGEs by this mechanism is limited by the ability of cos phages to adsorb and inject DNA into recipient strains, and also by the presence of suitable attachment sites in recipient genomes. However, since different bacterial genera express wall teichoic acid with similar structures, which can act as bacteriophage receptors governing the routes of horizontal gene transfer between major bacterial pathogens, horizontal gene transfer even across long phylogenetic distances is possible (Winstel et al., 2013). In addition, our previous results also demonstrated that the SaPI integrases have much lower sequence specificity than other typical integrases, and SaPIs readily integrate into alternative sites in the absence of the cognate attC site, such that any bacterium that can adsorb SaPI helper phage is a potential recipient (Chen and Novick, 2009). Thus, we anticipate that cos phages can have an important role in spreading MGEs carrying virulence and resistance genes. We also predict that cos sites will be found on many other MGEs, enabling cos phage-mediated transfer of any such element that can generate post-replicative concatemeric DNA.  相似文献   

4.
5.
6.
The transfer range of phage genes was investigated at the single-cell level by using an in situ DNA amplification technique. After absorption of phages, a phage T4 gene was maintained in the genomes of non-plaque-forming bacteria at frequencies of 10−2 gene copies per cell. The gene transfer decreased the mutation frequencies in nonhost recipients.Recently, whole-genome analyses have revealed that many bacterial genomes contain foreign genes, especially phage genes (9). The phage genes in bacterial genomes include genes for virulence or fitness factors such as extracellular toxins, superantigens, lipopolysaccharide-modifying enzymes, and proteins conferring serum resistance, etc. (1). These findings suggest that the horizontal transfer of phage genes has contributed significantly to the acquisition of new genetic traits and to the genetic diversity of bacteria (1, 9, 10). To truly appreciate the mechanisms behind phage-associated evolution, it is important to understand the frequency and range of transfer of phage genes.Most phage genomes consist of many genes derived from different origins (5, 8). Some genes are similar to those of other phages with phylogenetically different hosts or are found in the genomes of bacteria that are not the phage hosts. The mosaic nature of phage genomes has been known for some time, and a body of molecular genetic studies of phages to explain the mechanisms that drive this feature have been attempted previously (1, 5). More importantly, the horizontal transfer of phage genes has emerged as a major factor in the evolution of the phage genome. Since recombination between phage and phage/prophage can occur when these elements coexist in the same cell, coinfection with multiple phage species may result in the production of hybrid phage genomes (5). The pathways by which phages exchange genetic material vary dramatically in concert with host ranges. However, conventional plaque assays have shown that the host ranges of the phages studied are narrow. We hypothesized that phage genes can be transferred to more diverse species than previously thought.In order to accurately quantify DNA movement, gene targeting that does not require cultivation or gene expression is necessary (7). In situ DNA amplification methods allow the visualization of specific DNA sequences inside bacterial cells. In this study, we employed cycling primed in situ amplification-fluorescent in situ hybridization (CPRINS-FISH) to examine the possible range and frequency of the transfer of phage genes. CPRINS uses one primer and results in linear amplification of the target DNA inside cells, and multiply labeled fluorescent probe sets are applied for detection of the amplicons to improve the specificity and sensitivity of CPRINS (3). Previously, CPRINS-FISH did clarify the movement of DNA of a specific gene among Escherichia coli cells at the single-cell level (4).Enterobacterial phages P1 and T4 infect E. coli and have been well studied. P1 can exist as circular DNA within the bacterial cell as if it were a plasmid. Phage T4 is capable of undergoing only a lytic life cycle and not the lysogenic life cycle. Conventional methods using plaque assays have shown that the host of P1 and T4 is E. coli, but orthologous phage genes have been found in bacteria other than E. coli (6, 8). In the present study, strains of Enterobacteriaceae were allowed to grow on agar medium after the phage was adsorbed, and the maintenance of the transferred phage gene in the bacterial genomes was examined at the community level by quantitative real-time PCR and at the single-cell level by CPRINS-FISH.The following bacterial strains were used for maintenance experiments: Citrobacter freundii IFO 12681, Enterobacter aerogenes BM 2688, E. coli NBRC 12713, a Proteus mirabilis clinical isolate, Salmonella enterica serovar Enteritidis IID 640, and Yersinia enterocolitica IID 981. The bacterial strains were grown in Luria-Bertani (LB) medium (1% tryptone, 0.5% yeast extract, 0.5% NaCl; Nacalai Tesque Inc., Kyoto, Japan) at 37°C overnight.Stationary-phase cultures of 500 μl were incubated with 500 μl of SM buffer (50 mmol liter−1 Tris-HCl [pH 7.5], 100 mmol liter−1 NaCl, 8 mmol liter−1 MgSO4, 0.01% gelatin) containing the phage P1kc NBRC 20008 (2) or T4GT7 (11) at 37°C for 10 min at a multiplicity of infection of 1:1 (ratio of PFU of the phage to CFU of the recipient bacterium). The concentration of bacterial cells was adjusted to 109 cells ml−1. After 10 min of incubation, the diluted cell suspension (105 cells) was filtered through a polycarbonate filter (Advantec, Tokyo, Japan) with a pore size of 0.2 μm and a diameter of 25 mm. Cells trapped on the filter were cultured on LB agar medium at 37°C for 24 h. The filter was transferred into a microtube, and cells on the filter were suspended in 1 ml of sterile deionized water. The numbers of cells in the suspension and cells remaining on the filter were determined by using an epifluorescence microscope (see below) after staining of the samples with 1 μg ml−1 of 4′,6-diamidino-2-phenylindole (DAPI; Sigma-Aldrich Japan, Tokyo). The level of recovery of cells from the filter into sterile deionized water was about 99%. The cultured cells were subjected to real-time PCR and CPRINS-FISH.For real-time PCR, bacterial DNA was extracted using a QIAamp DNA isolation kit (Qiagen, Tokyo, Japan). The cell suspension was mixed with 10 mg ml−1 of lysozyme solution and incubated at 37°C for 1 h. DNA extraction was then performed according to the kit manufacturer''s instructions. Table Table11 lists the oligonucleotide primers for PCR and CPRINS and the polynucleotide probes used in the present study. Tail fiber genes from phages P1kc and T4GT7 were quantified by real-time PCR with a LightCycler system (Roche Diagnostics, Tokyo, Japan). LightCycler FastStart DNA master SYBR green I (Roche Diagnostics) was used with 5 mmol liter−1 Mg2+ and 0.5 μmol liter−1 (each) primers targeting the tail fiber genes of P1kc (P1-tail931f and P1-tail1148r) and T4GT7 (T4-tail2770f and T4-tail2983r). After a hot start for 10 min at 95°C, 40 cycles of PCR were run with denaturation at 94°C for 15 s, annealing at 60°C for 10 s, extension at 72°C for 10 s, and fluorescence detection at 83°C for 5 s. The known amounts of PCR products from the phage DNA (101 to 107 copies per reaction) were used for the standard curves to quantify the target DNA. To confirm the specificity of the reaction after real-time PCR, the PCR mixture was collected in a glass capillary and subjected to agarose gel electrophoresis in addition to a melting-curve analysis with the LightCycler system. The maintenance frequencies determined by real-time PCR were recorded as the copy number of the phage tail fiber gene per bacterial genome detected by staining with PicoGreen (Invitrogen, Tokyo, Japan) after cultivation of cells on LB agar medium for 24 h as described above. The frequencies were determined in triplicate for each sample. The increase in the phage gene copy number was determined by comparing the copy numbers in cells on the filter before and after cultivation. The phage gene copy number in cells on the filter was determined by the following formula: (total number of cells determined by DAPI staining) × (phage tail fiber gene copy number determined by real-time PCR)/(bacterial genome copy number determined by PicoGreen staining).

TABLE 1.

Probes and primers designed in this study
NameTargetTypeNucleotide sequence (5′-3′)
P1-tail931fTail fiber gene of phage P1PrimerAACGACCCGAATTACAGCAC
P1-tail1148rTail fiber gene of phage P1PrimerAGTGCTGCTGCAAGCTCATA
T4-tail2770fTail fiber gene of phage T4PrimerAGCACAAATGGTGAGCACAG
T4-tail2983rTail fiber gene of phage T4PrimerTTGCTACCGTGTGGGTATGA
T4-tail2664Tail fiber gene of phage T4ProbeGGCTTCAAGTACTGACTTAGGTACTAAAACCACATCAAGCTTTGACTATGGTACG
T4-tail2720Tail fiber gene of phage T4ProbeAAGGGAACTAACAGTACGGGTGGACACACTCACTCTGGTAGTGGTTCTA
T4-tail2769Tail fiber gene of phage T4ProbeTAGCACAAATGGTGAGCACAGCCACTACATCGAGGCATGGAATGG
T4-tail2818Tail fiber gene of phage T4ProbeGGTGTAGGTGGTAATAAGATGTCATCATATGCCATATCATACAGGGCGGG
T4-tail2869Tail fiber gene of phage T4ProbeGGGAGTAACACTAATGCAGCAGGGAACCACAGTCACACTTTCTCTTTTGGG
T4-tail2922Tail fiber gene of phage T4ProbeTAGCAGTGCTGGCGACCATTCCCACTCTGTAGGTATTGGTGCTCATA
Open in a separate windowCPRINS-FISH targeting the tail fiber gene of phage T4GT7 was performed as described by Kenzaka et al. (3, 4), except for the probe/primer sequences and thermal conditions. After cell wall permeabilization by lysozyme treatment (3), the CPRINS reaction was performed under the following conditions: a hot start at 95°C for 9 min, denaturation at 94°C for 1 min, annealing at 60°C for 30 s, and extension at 72°C for 1.5 min for primer T4-tail2983r. Amplification was repeated for 30 cycles by using a thermal cycler (PTC-200; Bio-Rad Laboratories, Inc.). After amplification, filters were rinsed with 0.1% Nonidet P-40 and sterile deionized water, dehydrated in 99% ethanol, and vacuum dried. Hybridization with Alexa Fluor 546-labeled polynucleotide probes (T4-tail2664, T4-tail2720, T4-tail2769, T4-tail2818, T4-tail2869, and T4-tail2922), washing, and DAPI staining were performed as described in a previous study (4). In order to exclude the possibility of nonspecific probe binding to cell structures other than the target DNA in the target cells, FISH using laboratory strains without amplification of target DNA and CPRINS-FISH targeting the tail fiber genes in E. coli strains that did not carry the genes were performed.In order to examine the infection ranges of phages, plaque assays and direct counting of phages were performed. Plaque assays were performed with LB soft agar (0.8% agar) as described by Kenzaka et al. (4). For the direct counting, phages were stained with 5× SYBR gold (Invitrogen, Tokyo, Japan) and trapped onto an Anodisc filter (Whatman Japan, Tokyo) with a pore size of 0.02 μm and a diameter of 25 mm.The cells or phage particles on the filters were observed under an epifluorescence microscope (E-400; Nikon, Tokyo, Japan) with the Nikon filter sets UV-2A (EX300-350, DM400, and BA420) for DAPI, B-2A (EX450-490, DM505, and BA520) for SYBR gold, and HQ-CY3 (G535/50, FT565, and BP610/75) for Alexa Fluor 546. Images were acquired using a Retiga 2000R cooled charge-coupled device camera (QImaging, Surrey, BC, Canada), and at least 2,000 DAPI- or SYBR gold-stained objects per sample were counted. The maintenance frequencies determined by CPRINS-FISH were recorded as the number of CPRINS-FISH-positive cells divided by the total direct count of recipient cells after cultivation as described above. The frequencies were determined in triplicate for each sample.After cultivation on LB agar medium for 24 h, the total number of cells on the filter as determined by DAPI staining increased by 8.7 × 102- to 1.1 × 104-fold (Table (Table2).2). Real-time PCR showed that the phage P1kc gene copy number increased only in plaque-forming strains (E. coli and E. aerogenes) and not in non-plaque-forming strains (Table (Table2).2). In contrast, the phage T4GT7 gene copy number increased in both plaque-forming and non-plaque-forming strains by 7.6 × 101- to 7.0 × 104-fold. The maintenance frequencies were more than 10−2 gene copies per bacterial genome (Table (Table2).2). Direct observation via epifluorescence microscopy showed that progeny phages were not produced in the non-plaque-forming strains (Table (Table2),2), and thus, fragments of phage genes were thought to integrate into the genomes of non-plaque-forming strains and replicate with the bacterial genomes.

TABLE 2.

Frequencies of maintenance of phage P1kc and T4GT7 genes in Enterobacteriaceae strains
PhageRecipientResult for infection range indicator:
Increase in total no. of cellscIncrease in phage gene copy no. (SD)dMaintenance frequency (SD) as determined bye:
Plaque formationaProduction of progenybReal-time PCRCPRINS-FISH
P1kcC. freundii7.0 × 103None<1.5 × 10−3ND
E. aerogenes++1.7 × 1037.7 × 103 (6.5 × 103)5.0 × 100 (4.2 × 100)ND
E. coli++7.2 × 1035.5 × 103 (2.7 × 103)9.1 × 10−1 (0.5 × 10−1)ND
P. mirabilis7.4 × 103None<1.5 × 10−3ND
S. Enteritidis8.4 × 103None<1.7 × 10−4ND
Y. enterocolitica4.6 × 103None<1.8 × 10−4ND
T4GT7C. freundii1.5 × 1037.5 × 103 (4.0 × 103)8.3 × 10−1 (4.4 × 10−1)8.6 × 10−2 (3.4 × 10−2)
E. aerogenes++8.7 × 1021.2 × 103 (0.8 × 103)8.0 × 10−1 (5.0 × 10−1)4.0 × 10−1 (0.7 × 10−1)
E. coli++1.1 × 1047.0 × 104 (2.7 × 104)8.0 × 101 (3.0 × 10)2.1 × 10−1 (0.4 × 10−1)
P. mirabilis4.0 × 1035.8 × 103 (4.2 × 103)3.3 × 10−1 (2.4 × 10−1)3.4 × 10−2 (2.2 × 10−2)
S. Enteritidis1.0 × 1047.6 × 101 (5.0 × 101)1.0 × 10−2 (0.7 × 10−2)8.8 × 10−2 (2.0 × 10−2)
Y. enterocolitica3.6 × 1031.6 × 104 (0.4 × 104)6.1 × 10−1 (1.6 × 10−1)2.2 × 10−2 (2.9 × 10−2)
Open in a separate windowaPlaque formation on soft agar was tested.bThe production of progeny phage particles was observed via epifluorescence microscopy.cThe increase (n-fold) in the total number of cells during bacterial growth for 24 h was determined via epifluorescence microscopy.dThe increase (n-fold) in the copy number of the phage tail fiber gene during bacterial growth for 24 h was determined by real-time PCR. Values in parentheses indicate standard deviations of results for triplicate samples.eMaintenance frequencies were determined by real-time PCR and CPRINS-FISH analyses targeting the phage tail fiber gene and are shown as the phage tail fiber gene copy numbers per bacterial genome and the numbers of gene-positive cells divided by the total numbers of cells, respectively. Values in parentheses indicate standard deviations of results for triplicate samples. ND, not determined.Real-time PCR provided a copy number for the target phage gene in the whole population, but the location of the target phage gene and the frequency of cells carrying the target gene were unclear. In addition, bacterial genomic DNA, which was measured using PicoGreen, included phage DNA, and thus the frequencies measured by dividing by the amount of bacterial genomic DNA were probably less accurate than those measured as described below. In order to confirm that the phage gene was located inside bacterial cells and determine a more accurate maintenance frequency for total cells, CPRINS-FISH targeting the tail fiber gene of phage T4GT7 was performed. CPRINS-FISH visualized the target phage gene in individual cells under an epifluorescence microscope (Fig. (Fig.1).1). It showed that the frequencies of maintenance of the tail fiber gene, expressed as the number of gene-positive cells divided by the total number of cells, were 2.1 × 10−1 to 4.0 × 10−1 for plaque-forming strains after growth on LB medium for 24 h (Table (Table2).2). Since phage T4GT7 is capable of undergoing only a lytic life cycle, CPRINS-FISH would detect cells in which the phage gene was replicating. For non-plaque-forming strains, the maintenance frequencies were 2.2 × 10−2 to 8.8 × 10−2 (Table (Table2).2). If the gene was amplified by the CPRINS reaction outside bacterial cells, the amplicon would not accumulate inside bacterial cells and they would not exhibit bright fluorescence. Therefore, CPRINS-FISH proved that a part of the phage T4GT7 gene was located inside cells of non-plaque-forming strains. The tail fiber gene is responsible for the phage tail structure. The DNA sequences of the phage genes responsible for phage morphology have been found in many bacterial genomes (1, 5).Open in a separate windowFIG. 1.Visualization of E. coli cells carrying the tail fiber gene transferred by phage T4GT7. (A) After being mixed with phages for 10 min, E. coli NBRC 12713 cells were cultured for 24 h and subjected to CPRINS-FISH targeting the phage gene. Only cells having amplified tail fiber gene products emitted the fluorescence of the Alexa Fluor 546-labeled probe under green excitation (exposure, 0.5 s). (B) All DAPI-stained bacterial cells were visualized under UV excitation (exposure time, 0.1 s).In order to explore the effect of integration of the phage gene into the bacterial genome on bacterial heredity, we determined the mutation frequency for a C. freundii strain that acquired the phage T4GT7 gene. Two colonies which acquired the phage T4GT7 gene were screened by colony PCR with T4-tail2770f and T4-tail2983r primers and designated Cik8-1 and Cik8-4. Mutation frequencies were determined with LB medium containing 150 μg ml−1of rifampin (rifampicin) or 10 μg ml−1of nalidixic acid. The mutation frequencies associated with nalidixic acid resistance decreased by 12- to 240-fold and the frequencies associated with rifampin resistance decreased by 40- to 83-fold compared to those for the parent strains (Fig. (Fig.2).2). Mutation increases genetic variation. The decreased mutation frequency would contribute to the genetic stability of the genome in individual cells but not to genetic variation in the population. Our results show that phage T4GT7 was capable of affecting the genomic properties of C. freundii, which was thought previously not to be the host, although the mechanism by which mutation frequencies decreased remains unknown. Further experiments are required to clarify the molecular mechanism by which mutation frequencies altered after gene transfer.Open in a separate windowFIG. 2.Mutation frequencies for T4GT7-infected C. freundii strains. Mutation frequencies were determined with LB agar medium containing nalidixic acid or rifampin. Cik8-1 and Cik8-4 were strains which acquired a phage gene transferred from phage T4GT7. Cik1 and Cik2 were the parent strains.In summary, during growth on agar medium after the phage was allowed to be adsorbed by strains of Enterobacteriaceae, the phage P1kc gene was not maintained in non-plaque-forming strains but the phage T4GT7 gene was maintained in more diverse species than previously expected. The transfer of foreign DNA molecules (DNA entry) into a bacterium is an important first step in genetic diversification through horizontal gene transfer. A previous study reported that phage P1kc is capable of injecting DNA into non-plaque-forming E. coli cells (4), but the phage P1kc gene was not maintained during bacterial growth in the present study. The results showing the difference in maintenance between phage P1kc and T4GT7 genes suggest that the maintenance of transferred phage genes depends on phage gene sequences or other phage factors. When maintained, the phage gene could alter the mutation frequency for bacteria that acquired the gene, affecting the genomic variability at the population level. Conventionally, phage-bacterium interaction has been studied with certain models consisting of a phage and a bacterium in which the phage can multiply (12, 13). Our results indicate the importance of the dynamic of phage genes among diverse bacteria that were previously thought not to be hosts and the hereditary impact of phage gene transfer on such bacteria.  相似文献   

7.
As translational clinical researchers familiar with the risk-benefit of hematopoietic stem cell transplantation in autoimmune diseases, we are intrigued by the recent report of umbilical cord mesenchymal stem cell (UC-MSC) transplantation in treatment-refractory systemic lupus erythematosus nephritis by Wang and colleagues. They report the results of an open-label single-arm multicenter phase I/II study. This stimulated us to examine whether collective data from this group provide sufficient evidence for the feasibility, safety, dose rationale, and potential efficacy of UC-MSCs to conduct a randomized controlled trial in such patients. Results, though confounded by variable baseline prednisone and immuno-suppressive treatment, appear to indicate near-term response rates of approximately 50%, which are comparable to those seen with hematopoietic stem cell transplantation but with less morbidity and mortality.Wang and colleagues have been in the forefront of evaluating the potential for mesenchymal stem cells (MSCs) to treat systemic lupus erythematosus (SLE), based on studies in murine autoimmune disease models, demonstrating immunomodulatory properties of MSCs [1]. We have reviewed the additional reports published in peer-reviewed journals from this group [25], together with two protocols available on ClinicalTrials.gov (NCT00698191 and NCT01741857) (Table 
Authors (date)ClinicalTrials.gov protocol numberStudy design/duration of follow-upNumber of patientsMSC type/regimenConditioningSafety: deaths/serious infectionPD marker a Efficacy
Sun et al. [2]NRSingle-arm/ median of 8.25 months (range of 3 to 28 months)16 (15 SLEN)UC, single infusionCYC 0.8 to 1.8 mg/kg intravenously, 2 to 4 days0/0Percentage of Treg cells increased at 3 months (P = 0.03)‘Decreasing SLEDAI and proteinuriab in all patients’
Liang et al. [3]NCT 00698191Single-arm/17.2 ± 9.5 months15 SLENBM, single infusionIncluded in protocol, but NR0/0Percentage of Treg cells increased at 1 week and 3 and 6 months (P <0.05)‘Decreasing SLEDAI and proteinuriab in all patients’
Wang et al. [4]NCT 00698191Unblinded-randomized, 2-arm/12 months58 (~88% SLEN)BM, UC, single versus 2× (7 days apart)CYC 10 mg/kg per day, day 4, 3, and 21/NRNDCR single: 16/30 (53%); double: 8/27 (29%)
Wang et al. [5]NRSingle-arm/mean of 27 months87 (84% SLEN)BM, UC, single infusion, 18 patients retreated at relapseCYC 10 mg/ kg/day, day 4, 3, and 25/NRNDCR in 23/83, relapse 10/83
Wang et al. [1] cNCT 01741857Single-arm40 (38 SLEN)UC, 2× infusion, 7 days apart)No3/4NDMCR 13/PCR 11, 7 relapse
Open in a separate window aPharmacodynamic (PD) markers = increased peripheral blood regulatory T (Treg) cells, balanced T helper 1 (Th1)/Th2 cytokines; b [2] baseline (BL) proteinuria 3.1 (±1.2) g/day versus 3 months, 1.3 (±0.9) g/day (P <0.001, n = 15); [3] BL proteinuria 2.7 (±1.2) g/day versus 6 months, 0.9 (±0.8) g/day (P <0.01, n = 12); cprotocol described at ClinicalTrials.gov consistent with this study, but not explicitly noted in report. BM, bone marrow; CR, complete remission; CYC, cyclophosphamide; MCR, major clinical response; MSC, mesenchymal stem cell; ND, not done; NR, not reported; PCR, partial clinical response; SLEDAI, Systemic Lupus Erythematosus Disease Activity Index; SLEN, systemic lupus erythematosus nephritis; UC, umbilical cord.Protocol NCT01741857, first posted on 26 November 2012 and updated 1 November 2013, appears to be the protocol for the study recently published in Arthritis Research & Therapy [1]. The report largely reflects the protocol, although dose is not described and patient entry criteria required a dose of prednisone of more than 20 mg/day. In the report, only 10 of 40 patients were receiving prednisone of more than 20 mg/day, and one dose - 1 × 106 cells per kg, infused twice 7 days apart - was evaluated. The umbilical cord mesenchymal stem cells (UC-MSCs) for infusion were centrally prepared by using a well-standardized, quality-controlled method, and infusions were well tolerated. One-year mortality (two patients with uncontrolled SLE) and morbidity (five serious infections) compare favorably to those seen with hematopoietic stem cell transplantation in patients with SLE [6].Previously reported studies by this group [25] evaluated stem cells derived from either bone marrow of healthy relatives or umbilical cords donated by healthy consenting mothers in patients with treatment-refractory SLE. Most patients had active SLE nephritis despite receiving prednisone of more than 20 mg and immuno-suppression with cyclophosphamide, mycophenylate mofetil, or leflunomide or a combination of these. In addition, they usually received ‘conditioning’ by cyclophosphamide 0.8 to 1.8 mg/kg per day for 3 days prior to transplant. In contrast, in the recently reported study [1], only UC-MSCs were transplanted, without ‘conditioning’, in patients with active treatment-refractory SLE nephritis receiving a more variable background of prednisone, often less than 20 mg/day, and immunosuppression. Although efficacy is difficult to evaluate because major clinical response (MCR) required that prednisone be tapered to less than 10 mg/day, while maintaining improvements in disease activity measured by British Isles Lupus Assessment Group and Systemic Lupus Erythematosus Disease Activity Index, conditioning is apparently not needed. However, because 18 of 40 patients were receiving prednisone of less than 20 mg/day at the start of the study, the efficacy criterion of ‘tapered’ prednisone is difficult to assess. Although 24 responses - 13 MCR and 11 partial clinical response (CR) - are reported, 6 MCR and 6 partial CR cannot be verified in the reported data, because prednisone tapering appears not to meet criteria for response. In addition, relapse occurred within a year in 6 patients, most receiving at least 20 mg prednisone at baseline. Nevertheless, there may be some evidence for potential efficacy or perhaps corticosteroid sparing, but dose regimen results are not sufficiently clear, or defined by relevant pharmacodynamic activity, to identify a specific UC-MSC dose to use for a randomized controlled trial (RCT).Review of previous reports may provide some additional insight (Table 4]. In the same study, patients were randomly assigned in an unblinded manner to receive a single infusion versus two infusions of MSC, 7 days apart; again, no significant difference in CR rate was observed. Of interest are two studies that reported (Table 2, 3]. No correlations with improvement in clinical status were reported, however.Overall, Wang and colleagues have demonstrated the feasibility of a multicenter trial, as they apparently recruited a sufficient number of patients in a reasonable period of time, and safety was acceptable. Apparently, conditioning pre-MSC dosing is not required, although this aspect of the treatment has not been studied in a controlled manner. It should be pointed out that the collective published results may reflect the fact that some of the same patients were reported in more than one study. For example, two publications [1, 4] refer to the same NCT00698191 protocol; thus, recruitment feasibility may be optimistic. The dose of stem cells, with biologic activity shown in two studies, may have been defined, although it is not absolutely clear (see above caveats). There are also some indications of an appropriate population, but again there seems to be some lack of clarity based on the recently published study [1] where the endpoints were not actually met.The lack of a well-rationalized dosing regimen, together with a lack of results from an appropriately designed, well-controlled study, makes it extremely difficult to develop a treatment approach with UC-MSCs. Nevertheless, we feel that these results should not be discarded without a proper RCT. Although there may be some challenges in designing a well-controlled, double-blind, placebo-controlled RCT in patients with prednisone-dependent active SLE nephritis, this should be attempted [7].  相似文献   

8.
Antimicrobial Activity of Simulated Solar Disinfection against Bacterial,Fungal, and Protozoan Pathogens and Its Enhancement by Riboflavin     
Wayne Heaselgrave  Simon Kilvington 《Applied and environmental microbiology》2010,76(17):6010-6012
Riboflavin significantly enhanced the efficacy of simulated solar disinfection (SODIS) at 150 watts per square meter (W m−2) against a variety of microorganisms, including Escherichia coli, Fusarium solani, Candida albicans, and Acanthamoeba polyphaga trophozoites (>3 to 4 log10 after 2 to 6 h; P < 0.001). With A. polyphaga cysts, the kill (3.5 log10 after 6 h) was obtained only in the presence of riboflavin and 250 W m−2 irradiance.Solar disinfection (SODIS) is an established and proven technique for the generation of safer drinking water (11). Water is collected into transparent plastic polyethylene terephthalate (PET) bottles and placed in direct sunlight for 6 to 8 h prior to consumption (14). The application of SODIS has been shown to be a simple and cost-effective method for reducing the incidence of gastrointestinal infection in communities where potable water is not available (2-4). Under laboratory conditions using simulated sunlight, SODIS has been shown to inactivate pathogenic bacteria, fungi, viruses, and protozoa (6, 12, 15). Although SODIS is not fully understood, it is believed to achieve microbial killing through a combination of DNA-damaging effects of ultraviolet (UV) radiation and thermal inactivation from solar heating (21).The combination of UVA radiation and riboflavin (vitamin B2) has recently been reported to have therapeutic application in the treatment of bacterial and fungal ocular pathogens (13, 17) and has also been proposed as a method for decontaminating donor blood products prior to transfusion (1). In the present study, we report that the addition of riboflavin significantly enhances the disinfectant efficacy of simulated SODIS against bacterial, fungal, and protozoan pathogens.Chemicals and media were obtained from Sigma (Dorset, United Kingdom), Oxoid (Basingstoke, United Kingdom), and BD (Oxford, United Kingdom). Pseudomonas aeruginosa (ATCC 9027), Staphylococcus aureus (ATCC 6538), Bacillus subtilis (ATCC 6633), Candida albicans (ATCC 10231), and Fusarium solani (ATCC 36031) were obtained from ATCC (through LGC Standards, United Kingdom). Escherichia coli (JM101) was obtained in house, and the Legionella pneumophila strain used was a recent environmental isolate.B. subtilis spores were produced from culture on a previously published defined sporulation medium (19). L. pneumophila was grown on buffered charcoal-yeast extract agar (5). All other bacteria were cultured on tryptone soy agar, and C. albicans was cultured on Sabouraud dextrose agar as described previously (9). Fusarium solani was cultured on potato dextrose agar, and conidia were prepared as reported previously (7). Acanthamoeba polyphaga (Ros) was isolated from an unpublished keratitis case at Moorfields Eye Hospital, London, United Kingdom, in 1991. Trophozoites were maintained and cysts prepared as described previously (8, 18).Assays were conducted in transparent 12-well tissue culture microtiter plates with UV-transparent lids (Helena Biosciences, United Kingdom). Test organisms (1 × 106/ml) were suspended in 3 ml of one-quarter-strength Ringer''s solution or natural freshwater (as pretreated water from a reservoir in United Kingdom) with or without riboflavin (250 μM). The plates were exposed to simulated sunlight at an optical output irradiance of 150 watts per square meter (W m−2) delivered from an HPR125 W quartz mercury arc lamp (Philips, Guildford, United Kingdom). Optical irradiances were measured using a calibrated broadband optical power meter (Melles Griot, Netherlands). Test plates were maintained at 30°C by partial submersion in a water bath.At timed intervals for bacteria and fungi, the aliquots were plated out by using a WASP spiral plater and colonies subsequently counted by using a ProtoCOL automated colony counter (Don Whitley, West Yorkshire, United Kingdom). Acanthamoeba trophozoite and cyst viabilities were determined as described previously (6). Statistical analysis was performed using a one-way analysis of variance (ANOVA) of data from triplicate experiments via the InStat statistical software package (GraphPad, La Jolla, CA).The efficacies of simulated sunlight at an optical output irradiance of 150 W m−2 alone (SODIS) and in the presence of 250 μM riboflavin (SODIS-R) against the test organisms are shown in Table Table1.1. With the exception of B. subtilis spores and A. polyphaga cysts, SODIS-R resulted in a significant increase in microbial killing compared to SODIS alone (P < 0.001). In most instances, SODIS-R achieved total inactivation by 2 h, compared to 6 h for SODIS alone (Table (Table1).1). For F. solani, C. albicans, ands A. polyphaga trophozoites, only SODIS-R achieved a complete organism kill after 4 to 6 h (P < 0.001). All control experiments in which the experiments were protected from the light source showed no reduction in organism viability over the time course (results not shown).

TABLE 1.

Efficacies of simulated SODIS for 6 h alone and with 250 μM riboflavin (SODIS-R)
OrganismConditionaLog10 reduction in viability at indicated h of exposureb
1246
E. coliSODIS0.0 ± 0.00.2 ± 0.15.7 ± 0.05.7 ± 0.0
SODIS-R1.1 ± 0.05.7 ± 0.05.7 ± 0.05.7 ± 0.0
L. pneumophilaSODIS0.7 ± 0.21.3 ± 0.34.8 ± 0.24.8 ± 0.2
SODIS-R4.4 ± 0.04.4 ± 0.04.4 ± 0.04.4 ± 0.0
P. aeruginosaSODIS0.7 ± 0.01.8 ± 0.04.9 ± 0.04.9 ± 0.0
SODIS-R5.0 ± 0.05.0 ± 0.05.0 ± 0.05.0 ± 0.0
S. aureusSODIS0.0 ± 0.00.0 ± 0.06.2 ± 0.06.2 ± 0.0
SODIS-R0.2 ± 0.16.3 ± 0.06.3 ± 0.06.3 ± 0.0
C. albicansSODIS0.2 ± 0.00.4 ± 0.10.5 ± 0.11.0 ± 0.1
SODIS-R0.1 ± 0.00.7 ± 0.15.3 ± 0.05.3 ± 0.0
F. solani conidiaSODIS0.2 ± 0.10.3 ± 0.00.2 ± 0.00.7 ± 0.1
SODIS-R0.3 ± 0.10.8 ± 0.11.3 ± 0.14.4 ± 0.0
B. subtilis sporesSODIS0.3 ± 0.00.2 ± 0.00.0 ± 0.00.1 ± 0.0
SODIS-R0.1 ± 0.10.2 ± 0.10.3 ± 0.30.1 ± 0.0
SODIS (250 W m−2)0.1 ± 0.00.1 ± 0.10.1 ± 0.10.0 ± 0.0
SODIS-R (250 W m−2)0.0 ± 0.00.0 ± 0.00.2 ± 0.00.4 ± 0.0
SODIS (320 W m−2)0.1 ± 0.10.1 ± 0.00.0 ± 0.14.3 ± 0.0
SODIS-R (320 W m−2)0.1 ± 0.00.1 ± 0.10.9 ± 0.04.3 ± 0.0
A. polyphaga trophozoitesSODIS0.4 ± 0.20.6 ± 0.10.6 ± 0.20.4 ± 0.1
SODIS-R0.3 ± 0.11.3 ± 0.12.3 ± 0.43.1 ± 0.2
SODIS, naturalc0.3 ± 0.10.4 ± 0.10.5 ± 0.20.3 ± 0.2
SODIS-R, naturalc0.2 ± 0.11.0 ± 0.22.2 ± 0.32.9 ± 0.3
A. polyphaga cystsSODIS0.4 ± 0.10.1 ± 0.30.3 ± 0.10.4 ± 0.2
SODIS-R0.4 ± 0.20.3 ± 0.20.5 ± 0.10.8 ± 0.3
SODIS (250 W m−2)0.0 ± 0.10.2 ± 0.30.2 ± 0.10.1 ± 0.2
SODIS-R (250 W m−2)0.4 ± 0.20.3 ± 0.20.8 ± 0.13.5 ± 0.3
SODIS (250 W m−2), naturalc0.0 ± 0.30.2 ± 0.10.1 ± 0.10.2 ± 0.1
SODIS-R (250 W m−2), naturalc0.1 ± 0.10.2 ± 0.20.6 ± 0.13.4 ± 0.2
Open in a separate windowaConditions are at an intensity of 150 W m−2 unless otherwise indicated.bThe values reported are means ± standard errors of the means from triplicate experiments.cAdditional experiments for this condition were performed using natural freshwater.The highly resistant A. polyphaga cysts and B. subtilis spores were unaffected by SODIS or SODIS-R at an optical irradiance of 150 W m−2. However, a significant reduction in cyst viability was observed at 6 h when the optical irradiance was increased to 250 W m−2 for SODIS-R only (P < 0.001; Table Table1).1). For spores, a kill was obtained only at 320 W m−2 after 6-h exposure, and no difference between SODIS and SODIS-R was observed (Table (Table1).1). Previously, we reported a >2-log kill at 6 h for Acanthamoeba cysts by using SODIS at the higher optical irradiance of 850 W m−2, compared to the 0.1-log10 kill observed here using the lower intensity of 250 W m−2 or the 3.5-log10 kill with SODIS-R.Inactivation experiments performed with Acanthamoeba cysts and trophozoites suspended in natural freshwater gave results comparable to those obtained with Ringer''s solution (P > 0.05; Table Table1).1). However, it is acknowledged that the findings of this study are based on laboratory-grade water and freshwater and that differences in water quality through changes in turbidity, pH, and mineral composition may significantly affect the performance of SODIS (20). Accordingly, further studies are indicated to evaluate the enhanced efficacy of SODIS-R by using natural waters of varying composition in the areas where SODIS is to be employed.Previous studies with SODIS under laboratory conditions have employed lamps delivering an optical irradiance of 850 W m−2 to reflect typical natural sunlight conditions (6, 11, 12, 15, 16). Here, we used an optical irradiance of 150 to 320 W m−2 to obtain slower organism inactivation and, hence, determine the potential enhancing effect of riboflavin on SODIS.In conclusion, this study has shown that the addition of riboflavin significantly enhances the efficacy of simulated SODIS against a range of microorganisms. The precise mechanism by which photoactivated riboflavin enhances antimicrobial activity is unknown, but studies have indicated that the process may be due, in part, to the generation of singlet oxygen, H2O2, superoxide, and hydroxyl free radicals (10). Further studies are warranted to assess the potential benefits from riboflavin-enhanced SODIS in reducing the incidence of gastrointestinal infection in communities where potable water is not available.  相似文献   

9.
TATA Binding Protein Discriminates between Different Lesions on DNA,Resulting in a Transcription Decrease     
Frédéric Coin  Philippe Frit  Benoit Viollet  Bernard Salles  Jean-Marc Egly 《Molecular and cellular biology》1998,18(7):3907-3914
  相似文献   

10.
Characterization of a Thermostable Short-Chain Alcohol Dehydrogenase from the Hyperthermophilic Archaeon Thermococcus sibiricus     
Tatiana N. Stekhanova  Andrey V. Mardanov  Ekaterina Y. Bezsudnova  Vadim M. Gumerov  Nikolai V. Ravin  Konstantin G. Skryabin  Vladimir O. Popov 《Applied and environmental microbiology》2010,76(12):4096-4098
Short-chain alcohol dehydrogenase, encoded by the gene Tsib_0319 from the hyperthermophilic archaeon Thermococcus sibiricus, was expressed in Escherichia coli, purified and characterized as an NADPH-dependent enantioselective oxidoreductase with broad substrate specificity. The enzyme exhibits extremely high thermophilicity, thermostability, and tolerance to organic solvents and salts.Alcohol dehydrogenases (ADHs; EC 1.1.1.1.) catalyze the interconversion of alcohols to their corresponding aldehydes or ketones by using different redox-mediating cofactors. NAD(P)-dependent ADHs, due to their broad substrate specificity and enantioselectivity, have attracted particular attention as catalysts in industrial processes (5). However, mesophilic ADHs are unstable at high temperatures, sensitive to organic solvents, and often lose activity during immobilization. In this relation, there is a considerable interest in ADHs from extremophilic microorganisms; among them, Archaea are of great interest. The representatives of all groups of NAD(P)-dependent ADHs have been detected in genomes of Archaea (11, 12); however, only a few enzymes have been characterized, and the great majority of them belong to medium-chain (3, 4, 14, 16, 19) or long-chain iron-activated ADHs (1, 8, 9). Up to now, a single short-chain archaeal ADH from Pyrococcus furiosus (10, 18) and only one archaeal aldo-keto reductase also from P. furiosus (11) have been characterized.Thermococcus sibiricus is a hyperthermophilic anaerobic archaeon isolated from a high-temperature oil reservoir capable of growth on complex organic substrates (15). The complete genome sequence of T. sibiricus has been recently determined and annotated (13). Several ADHs are encoded by the T. sibiricus genome, including three short-chain ADHs (Tsib_0319, Tsib_0703, and Tsib_1998) (13). In this report, we describe the cloning and expression of the Tsib_0319 gene from T. sibiricus and the purification and the biochemical characterization of its product, the thermostable short-chain ADH (TsAdh319).The Tsib_0319 gene encodes a protein with a size of 234 amino acids and the calculated molecular mass of 26.2 kDa. TsAdh319 has an 85% degree of sequence identity with short-chain ADH from P. furiosus (AdhA; PF_0074) (18). Besides AdhA, close homologs of TsAdh319 were found among different bacterial ADHs, but not archaeal ADHs. The gene flanked by the XhoI and BamHI sites was PCR amplified using two primers (sense primer, 5′-GTTCTCGAGATGAAGGTTGCTGTGATAACAGGG-3′, and antisense primer, 5′-GCTGGATCCTCAGTATTCTGGTCTCTGGTAGACGG-3′) and cloned into the pET-15b vector. TsAdh319 was overexpressed, with an N-terminal His6 tag in Escherichia coli Rosetta-gami (DE3) and purified to homogeneity by metallochelating chromatography (Hi-Trap chelating HP column; GE Healthcare) followed by gel filtration on Superdex 200 10/300 GL column (GE Healthcare) equilibrated in 50 mM Tris-HCl (pH 7.5) with 200 mM NaCl. The homogeneity and the correspondence to the calculated molecular mass of 28.7 kDa were verified by SDS-PAGE (7). The molecular mass of native TsAdh319 was 56 to 60 kDa, which confirmed the dimeric structure in solution.The standard ADH activity measurement was made spectrophotometrically at the optimal pH by following either the reduction of NADP (in 50 mM Gly-NaOH buffer; pH 10.5) or the oxidation of NADPH (in 0.1 M sodium phosphate buffer; pH 7.5) at 340 nm at 60°C. The enzyme exhibited a strong preference for NADP(H) and broad substrate specificity (Table (Table1).1). The highest oxidation rates were found with pentoses d-arabinose (2.0 U mg−1) and d-xylose (2.46 U mg−1), and the highest reduction rates were found with dimethylglyoxal (5.9 U mg−1) and pyruvaldehyde (2.2 U mg−1). The enzyme did not reduce sugars which were good substrates for the oxidation reaction. The kinetic parameters of TsAdh319 determined for the preferred substrates are shown in Table Table2.2. The enantioselectivity of the enzyme was estimated by measuring the conversion rates of 2-butanol enantiomers. TsAdh319 showed an evident preference, >2-fold, for (S)-2-butanol over (RS)-2-butanol. The enzyme stereoselectivity is confirmed by the preferred oxidation of d-arabinose over l-arabinose (Table (Table1).1). The fact that TsAdh319 is metal independent was supported by the absence of a significant effect of TsAdh319 preincubation with 10 mM Me2+ for 30 min before measuring the activity in the presence of 1 mM Me2+ or EDTA (Table (Table3).3). TsAdh319 also exhibited a halophilic property, so the enzyme activity increased in the presence of NaCl and KCl and the activation was maintained even at concentration of 4 M and 3 M, respectively (Table (Table33).

TABLE 1.

Substrate specificity of TsAdh319
SubstrateaRelative activity (%)
Oxidation reactionb
    Methanol0
    2-Methoxyethanol0
    Ethanol36
    1-Butanol80
    2-Propanol100
    (RS)-(±)-2-Butanol86
    (S)-(+)-2-Butanol196
    2-Pentanol67
    1-Phenylmethanol180
    1.3-Butanediol91
    Ethyleneglycol0
    Glycerol16
    d-Arabinose*200
    l-Arabinose*17
    d-Xylose*246
    d-Ribose*35
    d-Glucose*146
    d-Mannose*48
    d-Galactose*0
    Cellobiose*71
Reduction reactionc
    Pyruvaldehyde100
    Dimethylglyoxal270
    Glyoxylic acid36
    Acetone0
    Cyclopentanone0
    Cyclohexanone4
    3-Methyl-2-pentanone*13
    d-Arabinose*0
    d-Xylose*0
    d-Glucose*0
    Cellobiose*0
Open in a separate windowaSubstrates were present in 250 mM or 50 mM (*) concentrations.bRelative rates, measured under standard conditions, were calculated by defining the activity for 2-propanol as 100%, which corresponds to 1.0 U mg−1. Data are averages from triplicate experiments.cRelative rates, measured under standard conditions, were calculated by defining the activity for pyruvaldehyde as 100%, which corresponds to 2.2 U mg−1. Data are averages from triplicate experiments.

TABLE 2.

Apparent Km and Vmax values for TsAdh319
Coenzyme or substrateApparent Km (mM)Vmax (U mg−1)kcat (s−1)
NADPa0.022 ± 0.0020.94 ± 0.020.45 ± 0.01
NADPHb0.020 ± 0.0033.16 ± 0.111.51 ± 0.05
2-Propanol168 ± 291.10 ± 0.090.53 ± 0.04
d-Xylose54.4 ± 7.41.47 ± 0.090.70 ± 0.04
Pyruvaldehyde17.75 ± 3.384.26 ± 0.402.04 ± 0.19
Open in a separate windowaActivity was measured under standard conditions with 2-propanol. Data are averages from triplicate experiments.bActivity was measured under standard conditions with pyruvaldehyde. Data are averages from triplicate experiments.

TABLE 3.

Effect of various ions and EDTA on TsAdh319a
CompoundConcn (mM)Relative activity (%)
None0100
NaCl400206
600227
4,000230
KCl600147
2,000200
3,000194
MgCl21078
CoCl210105
NiSO410100
ZnSO41079
FeSO41074
EDTA1100
580
Open in a separate windowaThe activity was measured under standard conditions with 2-propanol; relative rates were calculated by defining the activity without salts as 100%, which corresponds to 0.9 U mg−1. Data are averages from duplicate experiments.The most essential distinctions of TsAdh319 are the thermophilicity and high thermostability of the enzyme. The optimum temperature for the 2-propanol oxidation catalyzed by TsAdh319 was not achieved. The initial reaction rate of oxidation increased up to 100°C (Fig. (Fig.1).1). The Arrhenius plot is a straight line, typical of a single rate-limited thermally activated process, but there is no obvious transition point due to the temperature-dependent conformational changes of the protein molecule. The activation energy for the oxidation of 2-propanol was estimated at 84.0 ± 5.8 kJ·mol−1. The thermostability of TsAdh319 was calculated from residual TsAdh319 activity after preincubation of 0.4 mg/ml enzyme solution in 50 mM Tris-HCl buffer (pH 7.5) containing 200 mM NaCl at 70, 80, 90, or 100°C. The preincubation at 70°C or 80°C for 1.5 h did not cause a decrease in the TsAdh319 activity, but provoked slight activation. The residual TsAdh319 activities began to decrease after 2 h of preincubation at 70°C or 80°C and were 10% and 15% down from the control, respectively. The determined half-life values of TsAdh319 were 2 h at 90°C and 1 h at 100°C.Open in a separate windowFIG. 1.Temperature dependence of the initial rate of the 2-propanol reduction by TsAdh319. The reaction was initiated by enzyme addition to a prewarmed 2-propanol-NADP mixture. The inset shows the Arrhenius plot of the same data.Protein thermostability often correlates with such important biotechnological properties as increased solvent tolerance (2). We tested the influence of organic solvents at a high concentration (50% [vol/vol]) on TsAdh319 by using either preincubation of the enzyme at a concentration of 0.2 mg/ml with solvents for 4 h at 55°C or solvent addition into the reaction mixture to distinguish the effect of solvent on the protein stability and on the enzyme activity. TsAdh319 showed significant solvent tolerance in both cases (Table (Table4),4), and the effects of solvents could be modulated by salts, acting apparently as molecular lyoprotectants (17). Furthermore, TsAdh319 maintained 57% of its activity in 25% (vol/vol) 2-propanol, which could be used as the cosubstrate in cofactor regeneration (6).

TABLE 4.

Influence of various solvents on TsAdh319 activitya
SolventRelative activity (%)bRelative activity (%)c
Buffer without NaClBuffer with 600 mM NaCl
None100100100
DMSOd98040
DMFAe1011341
Methanol98259
Acetonitrile9500
Ethyl acetate470*33*
Chloroform10579*81*
n-Hexane10560*118*
n-Decane3691*107*
Open in a separate windowaThe activity measured at the standard condition with 2-propanol as a substrate. Data are averages from triplicate experiments.bPreincubation for 4 h at 55°C in the presence of 50% (vol/vol) of solvent prior the activity assay.cWithout preincubation, solvent addition to the reaction mixture up to 50% (vol/vol) or using the buffer saturated by a solvent (*).dDMSO, dimethyl sulfoxide.eDMFA, dimethylformamide.From all the aforesaid we may suppose TsAdh319 or its improved variant to be interesting both for the investigation of structural features of protein tolerance and for biotechnological applications.  相似文献   

11.
DNA microarray analysis of rheumatoid arthritis susceptibility genes identified by genome-wide association studies     
Hidehiko Sugino  Hooi-Ming Lee  Norihiro Nishimoto 《Arthritis research & therapy》2010,12(2):401
In a recent interesting review, Alex Clarke and Timothy Vyse described the genetics of rheumatic disease [1]. In the past several years, genome-wide association studies (GWAS) have led to the identification of six high-risk rheumatoid arthritis (RA) susceptibility genes - namely, CD244, PADI4, SLC22A2, PTPN22, CTLA4, and STAT4 (summarized in [2]). In vitro studies using mutant alleles and cultured cells have revealed the individual upregulation of CD244, PADI4, SLC22A2, and PTPN22 [2-6]; however, studies on the expression of RA susceptibility genes in RA patients are rare. We therefore investigated the expression of the above-mentioned six RA susceptibility genes in 112 RA patients using DNA microarray analysis. This study aims to clarify whether DNA microarray analysis and GWAS produce comparable results with respect to RA susceptibility genes.Total RNA extracted from total peripheral blood cells obtained from 112 RA patients and 45 healthy individuals was used to prepare aminoallyl RNA. As a reference, mixed RNA from 45 healthy individuals was used. The aminoallyl RNA of each individual and the reference was subjected to Cy3 and Cy5 labeling, respectively, and was hybridized with an oligonucleotide-based DNA microarray. The data obtained were analyzed by nonparametric statistical group comparison. The intensities of the noprobe spots were used as the background. The median and standard deviation of the background intensity were calculated. The genes with an intensity value that was less than the median plus 2 standard deviation of the background intensity were identified as null. The Cy3/Cy5 ratios of all spots on the DNA microarray were normalized using the global ratio median. Only gene expression data that were collected from at least 80% of samples from each group were selected for further analysis. The unpaired Mann-Whitney test was used to determine statistically significant differences in the mRNA expression levels between the RA and healthy groups. Statistical significance was set at P < 0.05.The results of our DNA microarray analysis showed that the expressions of four out of the six RA susceptibility genes were significantly higher in RA patients than in healthy individuals (1.0 × 10-16 to 2.32 × 10-5) (Table (Table1).1). As described above, the upregulation of these four genes (CD244, PADI4, SLC22A2, and PTPN22) has been previously confirmed in in vitro studies. We found, however, that CTLA4 expression levels were similar between the RA and control groups, whereas STAT4 expression was significantly downregulated in the RA group (1.38 × 10-8). We investigated the expression of other RA susceptibility genes - namely, TRF1/C5 [7], CD40 [8], and CCL21 [8] - and found that their expressions were similar in both groups. The genetic risk factors for RA were recently reported to differ between Caucasian and Asian (Korean) populations [9]. The samples used in our microarray analysis were derived from the same Asian (Japanese) cohort. The expression profiles for these three genes may therefore not be consistent with the profiles determined by GWAS.

Table 1

Candidate genes identified from rheumatoid arthritis genome-wide association studies
GeneGeneIDPMIDGene expression (up or down)Microarray P valuesa
CD24460555418794858Up1.0 × 10-16
PADI460534712833157Up2.32 × 10-5
SLC22A260260814608356Up1.94 × 10-6
PTPN2260071615208781Up9.66 × 10-8
CTLA412389016380915No change0.767
STAT460055817804842Up1.38 × 10-8
Open in a separate windowaP values determined by comparison between 112 rheumatoid arthritis patients and 45 healthy individuals.In this study, we revealed the correlation between five out of the six high-risk RA susceptibility genes using DNA microarray analysis. Prostate cancer susceptibility genes identified by GWAS were recently reported to be consistent with those identified by microarray analysis [10]. We therefore concluded that the combination of microarray analysis and GWAS would be a more effective approach for gene identification than the analysis of individual datasets. Moreover, the simultaneous use of both methods would allow for more accurate identification of RA candidate genes.  相似文献   

12.
Borna Disease Virus Nucleoprotein (p40) Is a Major Target for CD8+-T-Cell-Mediated Immune Response          下载免费PDF全文
Oliver Planz  Lothar Stitz 《Journal of virology》1999,73(2):1715-1718
  相似文献   

13.
Active Efflux of Organic Solvents by Pseudomonas putida S12 Is Induced by Solvents   总被引:2,自引:0,他引:2       下载免费PDF全文
Jasper Kieboom  Jonathan J. Dennis  Gerben J. Zylstra    Jan A. M. de Bont 《Journal of bacteriology》1998,180(24):6769-6772
  相似文献   

14.
Current Status of Hemostatic Agents and Sealants in Urologic Surgical Practice     
Sashi S Kommu  Robert McArthur  Amr M Emara  Utsav D Reddy  Christopher J Anderson  Neil J Barber  Raj A Persad  Christopher G Eden 《Reviews in urology》2015,17(3):150-159
  相似文献   

15.
Introducing Titratable Water to All-Atom Molecular Dynamics at Constant?pH     
Wei Chen  Jason?A. Wallace  Zhi Yue  Jana?K. Shen 《Biophysical journal》2013,105(4):L15-L17
Recent development of titratable coions has paved the way for realizing all-atom molecular dynamics at constant pH. To further improve physical realism, here we describe a technique in which proton titration of the solute is directly coupled to the interconversion between water and hydroxide or hydronium. We test the new method in replica-exchange continuous constant pH molecular dynamics simulations of three proteins, HP36, BBL, and HEWL. The calculated pKa values based on 10-ns sampling per replica have the average absolute and root-mean-square errors of 0.7 and 0.9 pH units, respectively. Introducing titratable water in molecular dynamics offers a means to model proton exchange between solute and solvent, thus opening a door to gaining new insights into the intricate details of biological phenomena involving proton translocation.Solution pH is an important factor in biology. Although neutral pH in extracellular medium accounts for balanced electrostatics and proper folding of protein structures, pH gradients across cell membranes induce large conformational changes that are necessary for biological functions, such as ATP synthesis and efflux of small molecules out of the cell. To gain detailed insights into pH-dependent conformational phenomena, several constant pH molecular dynamics (pHMD) methods, based on either discrete or continuous titration coordinates, have been developed in the last decade (1–4). In the continuous pHMD (CpHMD) framework (2,4), a set of titration coordinates {λi} are simultaneously propagated along with the conformational degrees of freedom. Although the original CpHMD method based on the generalized Born (GB) implicit-solvent models (2,4) offers quantitative prediction of pKa values and pH dependence of folding and conformational dynamics of proteins (5), its accuracy and applicability to highly charged systems and those with dominantly hydrophobic regions are limited due to the approximate nature of the underlying implicit-solvent models.Motivated by the above-mentioned need, three groups have made efforts to develop a CpHMD method using exclusively the explicit-solvent models (6–8). In our development, the titration of acidic and basic sites is coupled with that of coions to level the total charge of the system (8). To further improve physical realism, here we replace the coions by titratable water molecules, which not only absorb the excess charge but also enable direct modeling of solute-solvent proton exchange in classical molecular dynamics simulations.To illustrate the utility of the new methodology, we applied it to the titration simulations of three proteins that were previously used to benchmark the GB-based CpHMD. Although this work does not explore specific interactions between titratable waters and proteins, the methodology can be further tested or improved to provide a rigorous way for modeling proton transfer in molecular dynamics, which is a computationally efficient alternative to the empirical valence-bond theory-based methodologies (9,10).We define titration of water as:
  • 1.Loss of a proton to give a negatively charged hydroxide,
H2O ? OH? + H+, (1)or
  • 2.Gain of a proton to give a positively charged hydronium,
H2O + H+ ? H3O+.(2)We now couple the titration of hydroxide (Eq. 1) with that of an acidic site of the solute in the CpHMD simulation,HA+OHKaA+H2O.(3)The use of hydronium is avoided here to prevent a potential artifact due to prolonged attraction with A. Analogously, we couple the titration of hydronium (Eq. 2) with that of a basic site,BH++H2OKbH3O++B.(4)Thus, effectively, a proton is transferred between the solute and solvent. However, we should note that in CpHMD simulations, titratable protons are represented by covalently attached dummies (2,4). Through varying the atomic charges and van der Waals interactions, they are seen by other atoms in the protonated state but not in the unprotonated state (see Table S1 in the Supporting Material). Furthermore, the solution proton concentration is implicitly modeled through a free energy term (2,4).In CpHMD, the reference potential of mean force (PMF) for titration is that of the model compound (blocked single amino acid in water) along λ (2,4). In the presence of cotitrating water molecules, it is necessary to add the PMF for the conversion of water to hydroxide or hydronium. One-nanosecond NPT simulations at ambient pressure and temperature were performed to calculate the average force, 〈dU/d,θ〉 at given θ-values, which are related to λ by λ = sin2 θ (see Fig. S1 in the Supporting Material). Thermodynamic integration was then applied to calculate the PMF. We found that the average force can be accurately fit when assuming the PMF is quadratic in λ (Fig. 1). The same applies to the PMFs for titration of models Asp, Glu, and His. After testing on the titration of model compounds (see Table S2), we performed 10-ns all-atom CpHMD simulations with the pH replica-exchange protocol for three proteins: HP36, BBL and HEWL (see the Supporting Material for details). Most of the calculated pKa values were converged in 10 ns per replica (see Fig. S3). Results are summarized in Fig. S4. Based on the 10-ns data, the root-mean-square (RMS) and average absolute errors are 0.9 and 0.7 pH units, respectively, while the largest absolute error is 2.5 (Glu35 of HEWL). Linear regression of the calculation versus experiment gives R2 of 0.8 and slope of 1.2.Open in a separate windowFigure 1Average force and potential of mean force for converting a water molecule to hydroxide (A) and hydronium. (B) (Data points) Average forces. (Dashed curves) Best fits using a linear function, 2A(λB). (Solid curves) Corresponding potential of mean force.

Table 1

Calculated and experimental pKa values of three proteins
ResidueExperimenta
GBa
All-atom CpHMD
Time (ns)b0–10–55–100–10
HP36
 Asp443.10 (0.01)3.2 (0.1)2.03.02.6 (0.5)
 Glu453.95 (0.01)3.5 (0.1)4.34.54.4 (0.1)
 Asp463.45 (0.12)3.5 (0.1)2.43.73.1 (0.6)
 Glu724.37 (0.03)3.5 (0.1)4.44.44.4 (0.0)
BBL
 Asp1293.88 (0.02)3.2 (0.0)2.23.22.7 (0.5)
 Glu1414.46 (0.04)4.3 (0.0)4.04.44.2 (0.2)
 His1426.47 (0.04)7.1 (0.0)5.95.85.8 (0.0)
 Asp1453.65 (0.04)2.8 (0.2)3.03.13.1 (0.0)
 Glu1613.72 (0.05)3.6 (0.3)4.23.94.0 (0.2)
 Asp1623.18 (0.04)3.4 (0.3)2.93.53.2 (0.3)
 Glu1644.50 (0.03)4.5 (0.1)5.74.65.2 (0.6)
 His1665.39 (0.02)5.4 (0.1)4.44.44.4 (0.0)
HEWL
 Glu72.6 (0.2)2.6 (0.1)3.63.43.5 (0.1)
 His155.5 (0.2)5.3 (0.5)5.15.15.1 (0.0)
 Asp182.8 (0.3)2.9 (0.0)2.53.32.9 (0.4)
 Glu356.1 (0.4)4.4 (0.2)8.58.78.6 (0.1)
 Asp481.4 (0.2)2.8 (0.2)−0.11.10.6 (0.6)
 Asp523.6 (0.3)4.6 (0.0)5.45.65.5 (0.1)
 Asp661.2 (0.2)1.2 (0.4)−0.60.80.3 (0.7)
 Asp872.2 (0.1)2.0 (0.1)0.82.11.5 (0.7)
 Asp1014.5 (0.1)3.3 (0.3)6.15.75.9 (0.2)
 Asp1193.5 (0.3)2.5 (0.1)3.03.33.2 (0.1)
Maximum absolute deviation1.82.42.62.5
Average absolute deviation (RMS deviation)0.5 (0.7)1.0 (1.2)0.6 (0.9)0.7 (0.9)
Linear fit R2 (slope)0.7 (0.8)0.8 (1.4)0.7 (1.1)0.8 (1.2)
Open in a separate windowaTaken from Wallace and Shen (12). The pKa''s of BBL were recalculated.bSampling time per pH replica.Breaking the simulations in two halves, we noticed that the second 5-ns sampling gave better agreement with experiment. The RMS deviation is reduced from 1.2 to 0.9 pH units, while the average absolute deviation is reduced from 1.0 to 0.6 pH units. The linear regression against experimental data is also improved, with the slope decreasing from 1.4 to 1.1 although R2 remains the same. Comparing these second-half results with the GB-based simulations, we find that the RMS and average absolute deviations are about the same as the GB-CpHMD results; however, the all-atom simulations show a small systematic overestimation (regression slope >1), whereas GB simulations show a systematic underestimation (regression slope <1).The improvement in the second halves of the simulations are seen mainly for residues involved in attractive electrostatic interactions, including Asp44 and Asp46 of HP36, Asp129 of BBL, and Asp48, Asp66, and Asp87 of HEWL. These residues are initially locked in salt-bridges or hydrogen bonds. However, in the second 5 ns, the attractive interactions weakened, leading to a decrease in the calculated pKa shifts relative to the model values and better agreement with experiment. For instance, Asp44 was initially in a salt-bridge distance from Arg55. However, the salt-bridge positions were sampled less often in the second 5 ns (see Fig. S5), which explains the 1-unit reduction in the calculated pKa shift. Significant fluctuation in ion-pair interactions was also observed in the work by Alexov (11). The carboxyl oxygen of Asp46 was a hydrogen-bond acceptor with both the backbone amide and hydroxyl of Ser43. These hydrogen bonds were less frequently sampled in the second 5 ns (see Fig. S6), leading to a decrease of the pKa shift for Asp46 by 1.3 units. These results indicate that extensive conformational sampling is necessary to give an accurate estimate of the ratio between the charged and neutral populations.Limited conformational sampling is also a contributing factor to the overestimation of the pKa shifts for buried residues (Fig. S7 and Fig. S8). The increase in SASA is correlated with the more frequent sampling of the states with λ close to 1, i.e., the deprotonated form (see Fig. S9). However, because Glu35 was buried in the starting conformation and the transition between buried and exposed states is slow compared to the simulation length, the exposed state may not be sufficiently sampled, leading to overestimation of the pKa shift.In contrast to Glu35, the SASA of Asp52 in HEWL is almost identical for both protonation states. The lack of conformational fluctuation is due to the strong hydrogen bonding with the side-chain amino group of Asn46 and Asn59 (data not shown). Overestimation of the pKa shifts for buried residues can also be attributed to the limitation of the additive force field which underestimates dielectric response in protein environment (more discussion see Supporting Material) of the pKa shifts for buried residues.Finally, to ascertain if the presence of hydroxide/hydronium introduces artifacts, we studied the interaction between hydroxide/hydronium and the titratable sites/ions. Comparing the hydroxide/hydronium with respective chloride/sodium ions, we find that the spatial distributions are nearly identical (see plots of distance distributions and radial distribution functions in Figs. S10–S13). However, the relative occupancy of the hydroxide around the neutral Asp/Glu, positive histidine, or sodium ion is 2–3 times as that of a chloride. The water-bridged interaction between sodium and chloride ions becomes much weaker when chloride is replaced by hydroxide or sodium is replaced by hydronium. By contrast, the occupancy of the hydronium around the solute is similar to that of the sodium. Furthermore, similar pKa results for these proteins were obtained when coions were used instead of titratable waters (data not shown). Thus, we believe that potential artifacts related to the ionized forms of water are negligible. Work is underway to further understand the limitations of the methodology and to explore applications to protein dynamics coupled to proton transfer.In summary, we have developed and tested titratable water models for use in all-atom CpHMD simulations. Although the benchmark pKa calculations indicate a comparable accuracy as the GB-CpHMD method, the all-atom method offers physical rigor and most importantly, it is applicable to systems that cannot be studied with GB-based simulations such as lipids and nucleic acids. We anticipate that the accuracy of this methodology can be further improved by incorporating the new-generation force fields that account for polarization. The coupling between proton titration of water and solute offers a computationally efficient way to model proton transfer in molecular mechanics simulations.  相似文献   

16.
Does an Eye-Hand Coordination Test Have Added Value as Part of Talent Identification in Table Tennis? A Validity and Reproducibility Study     
Irene R. Faber  Frits G. J. Oosterveld  Maria W. G. Nijhuis-Van der Sanden 《PloS one》2014,9(1)
This study investigated the added value, i.e. discriminative and concurrent validity and reproducibility, of an eye-hand coordination test relevant to table tennis as part of talent identification. Forty-three table tennis players (7–12 years) from national (n = 13), regional (n = 11) and local training centres (n = 19) participated. During the eye-hand coordination test, children needed to throw a ball against a vertical positioned table tennis table with one hand and to catch the ball correctly with the other hand as frequently as possible in 30 seconds. Four different test versions were assessed varying the distance to the TotalNationalRegionalLocalTotal43131119Boys268810Girls17539Age (years)10.4±1.410.9±1.510.4±1.510.1±1.47 year olds1--18 year olds51139 year olds3-3-10 year olds1232711 year olds1151512 year olds11443Length (cm)149±11150±12150±12148±10Weight (kg )38±837±737±738±9Right-handed359917Left-handed8422Training (hours*week-1)6 (0–20)11 (7–20)7 (4–11)2 (0–3)Competition (points)173 (−52–430)297 (144–430)188 (72–317)36 (−52–130)Open in a separate windowData are frequencies, except for age, length and weight (years±SD), and training and competition (mean (range)).  相似文献   

17.
A Systematic Proteomic Analysis of Listeria monocytogenes House-keeping Protein Secretion Systems     
Sven Halbedel  Swantje Reiss  Birgit Hahn  Dirk Albrecht  Gopala Krishna Mannala  Trinad Chakraborty  Torsten Hain  Susanne Engelmann  Antje Flieger 《Molecular & cellular proteomics : MCP》2014,13(11):3063-3081
  相似文献   

18.
DdlN from Vancomycin-Producing Amycolatopsis orientalis C329.2 Is a VanA Homologue with d-Alanyl-d-Lactate Ligase Activity          下载免费PDF全文
C. Gary Marshall  Gerard D. Wright 《Journal of bacteriology》1998,180(21):5792-5795
Vancomycin-resistant enterococci acquire high-level resistance to glycopeptide antibiotics through the synthesis of peptidoglycan terminating in d-alanyl-d-lactate. A key enzyme in this process is a d-alanyl-d-alanine ligase homologue, VanA or VanB, which preferentially catalyzes the synthesis of the depsipeptide d-alanyl-d-lactate. We report the overexpression, purification, and enzymatic characterization of DdlN, a VanA and VanB homologue encoded by a gene of the vancomycin-producing organism Amycolatopsis orientalis C329.2. Evaluation of kinetic parameters for the synthesis of peptides and depsipeptides revealed a close relationship between VanA and DdlN in that depsipeptide formation was kinetically preferred at physiologic pH; however, the DdlN enzyme demonstrated a narrower substrate specificity and commensurately increased affinity for d-lactate in the C-terminal position over VanA. The results of these functional experiments also reinforce the results of previous studies that demonstrated that glycopeptide resistance enzymes from glycopeptide-producing bacteria are potential sources of resistance enzymes in clinically relevant bacteria.The origin of antibiotic resistance determinants is of significant interest for several reasons, including the prediction of the emergence and spread of resistance patterns, the design of new antimicrobial agents, and the identification of potential reservoirs for resistance elements. Antibiotic resistance can occur either through spontaneous mutation in the target or by the acquisition of external genetic elements such as plasmids or transposons which carry resistance genes (7). The origins of these acquired genes are varied, but it has long been recognized that potential reservoirs are antibiotic-producing organisms which naturally harbor antibiotic resistance genes to protect themselves from the actions of toxic compounds (6).High-level resistance to glycopeptide antibiotics such as vancomycin and teicoplanin in vancomycin-resistant enterococci (VRE) is conferred by the presence of three genes, vanH, vanA (or vanB), and vanX, which, along with auxiliary genes necessary for inducible gene expression, are found on transposons integrated into plasmids or the bacterial genome (1, 20). These three genes are essential to resistance and serve to change the C-terminal peptide portion of the peptidoglycan layer from d-alanyl-d-alanine (d-Ala-d-Ala) to d-alanyl-d-lactate (d-Ala-d-Lac). This change results in the loss of a critical hydrogen bond between vancomycin and the d-Ala-d-Ala terminus and in a 1,000-fold decrease in binding affinity between the antibiotic and the peptidoglycan layer, which is the basis for the bactericidal action of this class of compounds (5). The vanH gene encodes a d-lactate dehydrogenase which provides the requisite d-Lac (3, 5), while the vanX gene encodes a highly specific dd-peptidase which cleaves only d-Ala-d-Ala produced endogenously while leaving d-Ala-d-Lac intact (19, 21). The final gene, vanA or vanB, encodes an ATP-dependent d-Ala-d-Lac ligase (4, 8, 10). This enzyme has sequence homology with the chromosomal d-Ala-d-Ala ligases, which are essential for peptidoglycan synthesis but which generally lack the ability to synthesize d-Ala-d-Lac (9).We have recently cloned vanH, vanA, and vanX homologues from two glycopeptide antibiotic-synthesizing organisms: Amycolatopsis orientalis C329.2, which produces vancomycin, and Streptomyces toyocaensis NRRL 15009, which produces A47934 (14). In addition, the vanH-vanA-vanX gene cluster was identified in several other glycopeptide producers. We have also demonstrated that the VanA homologue from S. toyocaensis NRRL 15009 can synthesize d-Ala-d-Lac in vitro and in the glycopeptide-sensitive host Streptomyces lividans (15, 16). We now report the expression of the A. orientalis C329.2 VanA homologue DdlN in Escherichia coli, its purification, and its enzymatic characterization. These data reinforce the striking similarity between vancomycin resistance elements in VRE and glycopeptide-producing organisms and support the possibility of a common origin for these enzymes.

Expression, purification, and specificity of DdlN.

DdlN was overexpressed in E. coli under the control of the bacteriophage T7 promoter. The construct gave good yields of highly purified enzyme following a four-step purification procedure (Table (Table1;1; Fig. Fig.1).1). Like other dd-ligases, DdlN behaved like a dimer in solution (not shown).

TABLE 1

Purification of DdlN from E. coli BL21 (DE3)/pETDdlN
SampleProtein (mg)Activity (nmol/min)Sp act (nmol/ min/mg)Recovery (%)Purification (fold)
Lysate1248436.82100
Ammonium sulfate (20–50% saturation)67.678011.5921.7
Sephacryl S20011.682571.49811
Q Sepharose2.87422658839
Phenyl Superose0.429974835110
Open in a separate windowOpen in a separate windowFIG. 1Purification of DdlN from E. coli BL21 (DE3)/pETDdlN. Proteins were separated on an SDS–11% polyacrylamide gel and stained with Coomassie blue. Lane 1, molecular mass markers (masses are noted at the left in kilodaltons); lane 2, whole-cell lysate; lane 3, ammonium sulfate fraction (20 to 50% saturation); lane 4, Sephacryl S200; lane 5, Q Sepharose; lane 6, phenyl Superose.The amino acid substrate specificity of DdlN was assessed by incubation of 14C-d-Ala with all 20 common amino acids in the d configuration. Purified DdlN catalyzed the synthesis of d-Ala-d-Ala in addition to that of several other mixed dipeptides, including d-Ala-d-Met and d-Ala-d-Phe (Fig. (Fig.2).2). Thus, DdlN exhibits a substrate specificity which is similar to that of VanA (4), with the capacity to synthesize not only d-Ala-d-Ala but also mixed dipeptides with bulky side chains in the C-terminal position.Open in a separate windowFIG. 2Substrate specificity of DdlN. Autoradiogram from thin-layer chromatography analysis of DdlN substrate specificity. All reaction mixtures contained 2.5 mM d-Ala and 1 mM ATP, and the radiolabel was 14C-d-Ala, except where noted. Lane 1, d-Ala; lane 2, d-Lac with 14C-d-Lac label; lane 3, d,l-methionine; lane 4, dl-phenylalanine; lane 5, d-Hbut; lane 6, d-hydroxyvalerate. Letters indicate the following: A, d-Ala-d-Lac; B, d-Lac; C, d-Ala-d-Met; D, d-Ala-d-Phe; E, d-Ala-d-Hbut; F, d-Ala-d-hydroxyvalerate.Importantly, DdlN is a depsipeptide synthase with the ability to synthesize d-Ala-d-Lac, d-Ala-d-hydroxybutyrate (Hbut), and d-Ala-d-hydroxyvalerate (Fig. (Fig.2).2). However, unlike VanA (5), d-hydroxycaproate and d-phenyllactate are not substrates (not shown). Thus, DdlN is a broad-spectrum d-Ala-d-X ligase with depsipeptide synthase activity.

Characterization of d-Ala-d-X ligase activity.

Following the initial assessment of the specificity of the enzyme, several substrates were selected for quantitative analysis by evaluation of their steady-state kinetic parameters (Table (Table2).2). DdlN has two amino acid (or hydroxy acid) Km values. Steady-state kinetic plots indicated that, like other dd-ligases, the N-terminal Km (Km1) was significantly lower (higher specificity) than the C-terminal Km (Km2). Since the former value is expected to be independent of the C-terminal substrate, only Km2 values were determined and are reported here.

TABLE 2

Characterization of steady-state parameters of DdlN and VanA
LigaseSubstrateKm2 (mM)kcat (min−1)kcat/Km2 (M−1 s−1)
DdlNd-Ala21 ± 2229 ± 71.8 × 102
d-Lac0.4 ± 0.0555 ± 12.3 × 103
d-Hbut2.5 ± 0.332 ± 22.1 × 102
ATPa1.2 ± 0.271 ± 50.98 × 102
DdlMbd-Ala166 ± 27
d-Lac1.08 ± 0.10
VanAcd-Ala382951.3 × 102
d-Lac7.1942.2 × 102
d-Hbut0.601083.0 × 103
Open in a separate windowa Determined in the presence of 10 mM d-Lac. b Data from reference 16c Data from reference 5. DdlN showed good d-Ala-d-Ala ligase activity but with a very high and physiologically questionable Km2 (21 mM). On the other hand, d-Ala-d-Lac synthesis was excellent, with a 4-fold decrease in kcat, compared to d-Ala-d-Ala synthesis, which was offset by a 52-fold drop in Km that resulted in a >12-fold increase in specificity (kcat/Km2). d-Hbut was also a good substrate, with a kcat/Km2 comparable to that of d-Ala.Steady-state kinetic parameters for d-Ala-d-X formation showed trends similar to those found with both VanA and DdlN. For example, the kcat values between VanA and DdlN were virtually the same for most substrates. There were significant differences, however. For instance, while the Km2 values for d-Ala were very high for all three enzymes, DdlN does have greater affinity for d-Ala, with a 1.8- and 7.9-fold lower Km2 than those of VanA and DdlM, respectively. Additionally, the Km2 for d-Lac was 17.8- and 2.7-fold lower than those for VanA and DdlM. Thus, DdlN has a more restrictive specificity for the C-terminal residue than VanA, which is compensated for by a higher affinity for the critical substrate d-Lac.

pH dependence of peptide versus that of depsipeptide synthesis activity.

The partitioning of the syntheses of d-Ala-d-Ala and d-Ala-d-Hbut in VanA and other depsipeptide-competent dd-ligases has been shown to be pH dependent (17). Determination of the pH dependence of DdlN in synthesizing peptide versus depsipeptide (Fig. (Fig.3)3) directly paralleled the results obtained with VanA in similar experiments. At lower pHs (<7), d-Ala-d-Hbut synthesis predominates and is exclusive at a pH of <6 (Fig. (Fig.3).3). At pH 7.5, levels of synthesis of d-Ala-d-Hbut and d-Ala-d-Ala are relatively equal, while at a pH greater than 8, the capacity to synthesize peptide overtakes the capacity to synthesize depsipeptide, although the latter is never abolished. Open in a separate windowFIG. 3pH dependence of partitioning of the syntheses of peptide and depsipeptide by DdlN. (A) Autoradiogram of a thin-layer chromatography separation of the products of reaction mixtures containing 14C-D-Ala, unlabeled D-Ala, and d-Hbut. (B) Quantification of reaction products following phosphorimage analysis. Filled circles, D-Ala-d-Hbut; open circles, D-Ala-D-Ala.The partitioning of the formation of peptide versus depsipeptide as a function of pH by DdlM is comparable to that by VanA and depsipeptide-competent mutants of DdlB (17), which show essentially exclusively depsipeptide formation at lower pHs and increasing peptide formation as the pH increases. This implies a potential role for the protonated ammonium group of d-Ala2 in second-substrate recognition and suggests a mechanism for the discrimination between d-Ala and d-Lac at physiologic pH. The structural basis for this distinction remains obscure for DdlB and VanA or DdlN.

Concluding remarks.

Resistance to vancomycin and other glycopeptides is mediated through the synthesis of a peptidoglycan which does not terminate with the canonical d-Ala-d-Ala dipeptide. Thus, enterococci which exhibit the VanC phenotype, which consists of low-level, noninducible resistance to vancomycin only, have peptidoglycan terminating in d-Ala-d-Ser (19). On the other hand, bacteria which are constitutively resistant to high concentrations of glycopeptides, such as lactic acid bacteria and VRE exhibiting the VanA or VanB phenotype (high-level inducible resistance to vancomycin), incorporate the depsipeptide d-Ala-d-Lac into their cell walls (2, 12, 13). The enzymes responsible for the intracellular synthesis of d-Ala-d-Lac not surprisingly have significant amino acid sequence similarity with d-Ala-d-Ala ligases, which are responsible for d-Ala-d-Ala synthesis in all bacteria with a cell wall (9).The d-Ala-d-Lac synthases can be subdivided into two groups based on sequence homology: those found in the constitutively resistant lactic acid bacteria and those found in glycopeptide-producing organisms and VanA or VanB VRE (9, 14). The former have more similarity with exclusive d-Ala-d-Ala ligases. Indeed, single point mutations in d-Ala-d-Ala ligases which yield sequences more similar to those of lactic acid bacterium d-Ala-d-Lac ligases are sufficient to induce significant depsipeptide synthase activity in these enzymes (17). Similarly, mutational studies of the d-Ala-d-Lac ligase from Leuconostoc mesenteroides have demonstrated that the converse also holds (18). On the other hand, the molecular basis for depsipeptide synthesis by the VanA or VanB ligases is unknown, in large part due to the lack of protein structural information on which to base mutational studies, unlike the situation with d-Ala-d-Ala ligases, where the E. coli DdlB structure serves as a template for mechanistic research (11).Significantly, a major difference in the VanA or VanB ligases and other dd-ligases lies in the amino acid sequence of the ω-loop region, which closes off the active site of DdlB (11) and has been shown to contribute amino acid residues with the capacity to control the syntheses of d-Ala-d-Ala and d-Ala-d-Lac, notably, Tyr216 (17, 18). Until recently, the VanA and VanB ligases were exceptional in amino acid structure and had no known homologues. The sequencing of resistance genes from glycopeptide-producing bacteria has uncovered enzymes with >60% homology to VanA or VanB and which are virtually superimposable in the critical ω-loop region (14, 15). One of these, DdlM from S. toyocaensis NRRL 15009, has been shown to have d-Ala-d-Lac ligase ability (15, 16), although no rigorous analysis of this activity has been performed. The results presented here demonstrate that DdlN from the vancomycin producer A. orientalis C329.2 not only is a d-Ala-d-Lac ligase but also has significant functional homology with VanA. It is not known at present if, like S. toyocaensis NRRL 15009 (16), A. orientalis C329.2 also possess a d-Ala-d-Ala-exclusive ligase, though the presence of a vanX gene (14) suggests that it may.These studies demonstrate that DdlN cloned from a vancomycin-producing bacterium is a d-Ala-d-Lac ligase which has not only amino acid sequence homology with the dd-ligases from VRE but also functional homology. Thus, VanA, VanB, DdlN, and DdlM have likely evolved from similar origins. The fact that a vanH-vanA-vanX gene cluster can be found in other glycopeptide producers as well (14) suggests that the genes now found in VRE may have originated in glycopeptide-producing bacteria. Our finding that overexpressed, purified, DdlN shows many enzymatic characteristics similar (though not identical) to those of VanA suggests that the genes from glycopeptide-producing bacteria can be important in elucidating biochemical and protein structural aspects of the VRE proteins.  相似文献   

19.
Economic Benefits of Investing in Women’s Health: A Systematic Review     
Kristine Hus?y Onarheim  Johanne Helene Iversen  David E. Bloom 《PloS one》2016,11(3)

Background

Globally, the status of women’s health falls short of its potential. In addition to the deleterious ethical and human rights implications of this deficit, the negative economic impact may also be consequential, but these mechanisms are poorly understood. Building on the literature that highlights health as a driver of economic growth and poverty alleviation, we aim to systematically investigate the broader economic benefits of investing in women’s health.

Methods

Using the Preferred Reporting Items for Systematic Reviews and Meta-Analysis (PRISMA) guidelines, we systematically reviewed health, gender, and economic literature to identify studies that investigate the impact of women’s health on micro- and macroeconomic outcomes. We developed an extensive search algorithm and conducted searches using 10 unique databases spanning the timeframe 01/01/1970 to 01/04/2013. Articles were included if they reported on economic impacts stemming from changes in women’s health (table of outcome measures included in full review, Outcome measures   FertilityIntergenerational Health SpilloverEducationProductivitySavingsMicroeconomic level    Total fertility rateChild survivalEnrollment in schoolIncomeMoneyChange in fertilityChild wellbeing and behaviorYears of schoolingPurchasing powerAssetsAge at first birth/ teenage pregnanciesAnthropometryEarly drop outPerformanceBirth spacingImproved cognitive developmentPerformance in school Life expectancyHigher education  Adult health outcomesLiteracy  Nutrition   Intrauterine growth  Macroeconomic level    Open in a separate windowGross domestic product/gross national product, gross domestic product/gross national product growth, income per capita, labor force participation, per capita income.

Results

The existing literature indicates that healthier women and their children contribute to more productive and better-educated societies. This study documents an extensive literature confirming that women’s health is tied to long-term productivity: the development and economic performance of nations depends, in part, upon how each country protects and promotes the health of women. Providing opportunities for deliberate family planning; healthy mothers before, during, and after childbirth, and the health and productivity of subsequent generations can catalyze a cycle of positive societal development.

Conclusions

This review highlights the untapped potential of initiatives that aim to address women’s health. Societies that prioritize women’s health will likely have better population health overall, and will remain more productive for generations to come.  相似文献   

20.
Heterosexual Transmission of a Murine AIDS Virus     
Yoshiaki Okada  Eri Abe  Katsutoshi Komuro  Toshiaki Mizuochi 《Journal of virology》1998,72(3):2541-2543
Heterosexual transmission of a murine leukemia virus mixture named LP-BM5 MuLV, which is known as the murine AIDS virus, was investigated. Our results indicated that the heterosexual transmission of LP-BM5 MuLV occurs in both directions with high frequency and that the frequencies of virus transmission in the cervix and penis are higher than those in other genital organs. The results suggested that infection by LP-BM5 MuLV via heterosexual transmission may initially take place at particular retrovirus-sensitive sites (cells) in the genital organs.Human immunodeficiency virus (HIV) infection is now pandemic. In many countries, HIV has been spread mainly by heterosexual transmission (3, 5). For the prevention of HIV infection, as well as for the development of vaccines against HIV, it is of a great importance to understand the mechanisms of the heterosexual transmission of retroviruses. Since it is difficult to investigate the mechanisms of heterosexual transmission of HIV in humans experimentally, an animal model with a retrovirus which induces an acquired immunodeficiency syndrome like human AIDS would be useful. A murine leukemia virus mixture called LP-BM5 MuLV induces a severe acquired immunodeficiency syndrome termed murine AIDS (MAIDS) in susceptible strains of mice (10). The mixture includes a replication-competent ecotropic virus, mink cell focus-inducing virus, and a replication-defective virus which has been considered to be involved in the pathogenesis of MAIDS (4). With many similarities to human AIDS patients, mice infected with the LP-BM5 MuLV mixture develop splenomegaly, systemic lymphadenopathy, and severe immunodeficiency (4, 11). We previously reported that maternal transmission of LP-BM5 MuLV occurs via mother’s milk with high frequency (12). In the present study, we demonstrate that the heterosexual transmission of LP-BM5 MuLV also occurs with high frequency via genital organs.C57BL/10 (B10) mice were purchased from Japan SLC Inc., Shizuoka, Japan. All mice were specific-pathogen free and were housed in an air-conditioned room. They were given autoclaved water and sterilized pelleted feed. An SC-1 clone chronically infected with LP-BM5 MuLV, the G6 cell line, was kindly supplied by H. C. Morse III, National Institutes of Health, Bethesda, Md. Virus was prepared from the supernatant of G6 cells as previously described (12). The virus preparation was stored at −70°C until use. B10 mice were inoculated by the intraperitoneal route with 0.3 ml of the LP-BM5 MuLV preparation. To increase the frequency of sexual contacts and to avoid pregnancy in the female mice, all male mice were sterilized by vasectomy under anesthesia with pentobarbital (Nembutal). The vasectomized male mice were mated with female mice at least 4 weeks postoperation, since sperm are usually kept alive for 2 to 3 weeks in spermiducts. Excised genital organs were crushed with plastic sticks in 1 ml of lysis buffer containing 10 mM Tris-HCl (pH 8.0), 100 mM NaCl, 1 mM EDTA, 0.5% sodium dodecyl sulfate, and proteinase K (0.5 mg/ml). Spleen cells were lysed after hemolysis with 0.83% NH4Cl. Lysed samples were incubated at 50°C for 3 h. DNA was extracted three times with phenol-chloroform, precipitated with cold ethanol, treated with RNase and proteinase K, and dissolved in 0.1 ml of H2O. LP-BM5 MuLV defective virus genome was detected by Southern blot hybridization combined with PCR as described previously (12). In brief, template DNAs (1 μg per tube) were added to a cocktail adjusted to final concentrations of 10 mM Tris-HCl (pH 8.3), 50 mM KCl, 1.5 mM MgCl2, 0.01% gelatin, 200 μM deoxynucleoside triphosphate, 100 pmol of each primer (5′-CCTCTTCCTTTATCGACACT-3′ [sense] and 5′-ATTAGGGGGGGAATAGCTCG-3′ [antisense]), and 2 U of Taq DNA polymerase (Boehringer Mannheim) in a total volume of 100 μl and were subjected to 32 cycles of amplification. In each cycle of PCR, the mixture was denatured at 95°C for 1 min (5 min for the first cycle), annealed at 55°C for 3 min, and extended at 72°C for 1 min. The PCR-amplified products were subjected to gel electrophoresis (1.5% agarose) and transferred to a Hybond N+ membrane (Amersham) by the alkaline blotting method. Hybridization was achieved with a 5′ 32P-labeled probe (5′-TGTCAAAGGGACCAGTTAAG-3′) at 45°C overnight in 6× SSC (1× SSC is 0.15 M NaCl plus 0.015 M sodium citrate)–0.5% sodium dodecyl sulfate–100 μg of salmon sperm DNA per ml. Hybridized membranes were washed twice in 2× SSC at 37°C for 10 min and then in 0.5× SSC at 45°C for 30 min. DNA derived from uterine cervices of uninfected mice was used as a negative control. The limit of sensitivity was approximately 10 copies per tube, as assessed by Southern blot analysis with plasmid DNAs (1/10 of the PCR product).Concanavalin A (ConA) was obtained from Pharmacia Fine Chemicals, Uppsala, Sweden. Responder spleen cells (2 × 105) were cultured with ConA (5 μg/ml) in 96-well flat-bottomed microculture plates in 0.2 ml of culture medium at 37°C in 7.5% CO2. The culture medium consisted of RPMI 1640 supplemented with 10% fetal calf serum, penicillin (5,000 IU/100 ml), streptomycin (5,000 μg/100 ml), nonessential amino acids, sodium pyruvate (11.0 mg/100 ml), 2-mercaptoethanol (5 × 10−5 M), and l-glutamine (29.2 mg/100 ml). On day 2, cultures were pulsed with 1 μCi of [3H]thymidine and incubated for an additional 12 to 18 h. Incorporation of [3H]thymidine into responder spleen cells was quantitated by liquid scintillation counting. Determinations were performed in triplicate; standard errors of the means were generally <5% and therefore have not been indicated.As illustrated in Fig. Fig.1,1, in order to investigate the heterosexual transmission of LP-BM5 MuLV from male to female mice, normal male mice were inoculated with LP-BM5 MuLV and vasectomized 1 week later. At 5 weeks after virus inoculation, they were mated with uninfected female mice. After 8 weeks of breeding, female mice were sacrificed and their vaginae, cervices uteri, corpora uteri, inguinal lymph nodes, and spleens were removed and stored at −70°C until use. In the opposite direction, to investigate virus transmission from female to male, normal female mice were inoculated with LP-BM5 MuLV and then mated with uninfected, vasectomized male mice as described above. After 8 weeks of breeding, male mice were sacrificed and their penes, prepuces, inguinal lymph nodes and spleens were removed and stored at −70°C until use. Figure Figure22 shows the detection by PCR of the LP-BM5 defective virus genome in genital organs and spleens that were taken from mice mated with their virus-infected counterparts. It was demonstrated that although the defective virus genome was detected in both spleens and genital organs in some male mice (2 of 17 [see Table Table1]),1]), as shown in Fig. Fig.2,2, lanes 3 and 4, the defective virus genome was detected only in the genital organs, not the spleens (Fig. (Fig.2,2, lanes 5 and 6), from most of the male mice. In contrast, all of the female mice were positive for defective virus genome only in the genital organs (Fig. (Fig.2,2, lanes 1 and 2). None of the mice examined were positive for the virus genome only in the spleens (this issue is discussed below). It should be noted here that the efficacy of PCR amplification, which was measured by experiments using the mixture of genomic DNA and plasmid DNA containing the defective virus, did not differ among the genital organs and spleens. By using the above strategy, the heterosexual transmission of LP-BM5 MuLV was investigated according to the protocol shown in Fig. Fig.1.1. Open in a separate windowFIG. 1Experimental design for examination of heterosexual transmission of the MAIDS virus in B10 mice. i.p., intraperitoneal.Open in a separate windowFIG. 2Detection of the LP-BM5 MuLV defective virus genome by PCR in genital organs and spleens. The template DNAs (1 μg) derived from female or male mice which were bred with LP-BM5 MuLV-infected mice were amplified by PCR. Samples were prepared from either female (lanes 1 and 2) or male (lanes 3 to 6) mice. Lanes 1, 3, and 5, spleen; lane 2, uterine cervix; lanes 4 and 6, penis (from two representative male mice). The PCR products (5 μl) were applied to a 1.5% agarose gel and analyzed by Southern blotting with a probe for the defective virus (12).

TABLE 1

Heterosexual transmission of LP-BM5 MuLV
ExptClinical condition
Detection of LP-BM5 MuLV (no. positive/total [%])
MaleFemaleSpleenInguinal lymph nodeCervixCorpusVaginaPenisPrepuce
1MAIDSNormal0/25 (0)0/16 (0)9/25 (36)NDaND
2MAIDSNormal0/11 (0)ND3/11 (27)0/11 (0)1/11 (9)
3NormalMAIDS1/8 (12)3/8 (38)6/8 (75)0/8 (0)
4NormalMAIDS1/9 (11)ND5/9 (56)1/9 (11)
Open in a separate windowaND, not done. Twenty-five female mice that were mated with the virus-infected male mice were analyzed for the presence of LP-BM5 defective genome in their genital organs, lymph nodes, and spleens. As summarized in Table Table1,1, the defective virus genome was detected with high frequency in cervices (9 of 25). However, the defective virus genome was not detected in spleens at all (0 of 25). The female genital organs are divided into three parts, namely, the vagina, cervix of uterus, and corpus of uterus. As also shown in Table Table1,1, the cervix appears to be more sensitive to virus infection than the other organs. Since MAIDS virus was not detected in castrated female mice, which were kept with virus-infected male mice in the same cage, the virus infection occurred via heterosexual transmission rather than by nonheterosexual horizontal transmission (data not shown). In 17 male mice mated with the virus-infected female mice (Table (Table1),1), the defective virus genome was detected in penes with high frequency (11 of 17). The defective virus genome was detected in DNA prepared from spleens with much lower frequency (2 of 17). In male mice, the penis seems to be much more sensitive to virus infection than are the prepuce and spleen (Table (Table1).1). In experiments 1 and 3, we also examined the inguinal lymph nodes from 16 female mice and 8 male mice. The defective virus genome was detected in some of the male mice (3 of 8) but not at all in the female mice examined. These results suggest that the LP-BM5 MuLV mixture initially infects the cervix or penis and then spreads over the whole body, including the lymph nodes and spleen.To determine whether mice infected with LP-BM5 MuLV by heterosexual transmission in fact develop MAIDS, we examined both spleen weights and mitogen (ConA) responses of female mice at 10 months after mating. As shown in Table Table2,2, female mice which were infected with LP-BM5 MuLV by heterosexual transmission (i.e., the defective virus genome was detected in the cervix) developed MAIDS as assessed by splenomegaly and decreased mitogen response, although the symptoms were less severe than of mice directly infected with LP-BM5 MuLV via the intraperitoneal route. Therefore, the cells in the genital organs were not only infected by the MAIDS virus but also able to replicate and spread the virus.

TABLE 2

Development of MAIDS in heterosexually infected B10 mice
Clinical condition
Spleen wt (mg)Mitogen response (cpm)Detection of LP-BM5 MuLV
MaleFemaleSpleenCervix
NormalNormal10539,981
9219,317
MAIDSNormal13610,346++
1867,799++
2454,911++
Open in a separate windowThe main route of HIV infection is heterosexual transmission (3, 5). However, the mechanisms of heterosexual transmission of retroviruses have been ill defined. HIV infection has been thought to occur during sexual contacts through slight injuries in the genital organs and to subsequently spread over the whole body. Among the genital organs of females, the parts of direct contact with male genital organs and semen are the vagina and cervix of uterus. The vagina is covered by a thick stratified squamous epithelium, while the cervix is covered by a monolayer columnar epithelium in addition to a squamous epithelium (2, 7). Histological examination (13) showed the presence of HIV-infected cells in the cervices derived from HIV carrier females (those infected with HIV by drug injections rather than by heterosexual transmission). Furthermore, a previous study utilizing female chimpanzees demonstrated that transmission of HIV could occur by insertion of cotton containing HIV into the vagina (8). These results suggested the presence of retrovirus-sensitive cells in genital organs. In our study, the cervix and penis are shown to be sensitive sites for virus infection (Table (Table1).1). Our assumption that there might be retrovirus-sensitive cells in a particular genital organ is currently under investigation by using in situ hybridization and immunohistochemical analyses.The heterosexual LP-BM5 MuLV infection rate for females to males appeared to be higher than that for males to females (Table (Table1).1). The mating frequency of normal male mice with infected female mice is supposed to be higher than that of normal female mice with infected male mice, since normal female mice fall into false pregnancy after mating and therefore reject male mice for a few weeks. This difference may also be attributed to the longer retention of genital secretions containing LP-BM5 MuLV in the male genital organs because of their phimoses (9). In fact, the defective virus genome was detected in vaginal secretions (both in secreted fluid and cells) by PCR (data not shown). Alternatively, the penis might be a highly sensitive site for retrovirus infection. In this regard, it is interesting that the defective virus genome was detected with very low frequency (1 of 17 male mice) in the prepuce even though it is constantly in contact with the penis. It is worth mentioning that contamination by retroviruses in the seminal fluid may happen at the prostate, seminal vesicle, vas deferens, Cowper’s glands, or penile urethra, since the sterilized (vasectomized) mice were still capable of transmitting the viruses to female mice (1, 6).The animal model for heterosexual transmission of retroviruses presented here has practical advantages, including (i) the high frequency of virus transmission and (ii) the possibility of rapid and cost-effective screening for antiretroviral agents (drugs and vaccines, etc.). This model may provide valuable information relating to heterosexual transmission of retroviruses including HIV and may further contribute to the prevention of HIV infection and the development of a remedy for AIDS.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号