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 62 毫秒
1.
2.
3.
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   

4.
Melioidosis has been considered an emerging disease in Brazil since the first cases were reported to occur in the northeast region. This study investigated two municipalities in Ceará state where melioidosis cases have been confirmed to occur. Burkholderia pseudomallei was isolated in 26 (4.3%) of 600 samples in the dry and rainy seasons.Melioidosis is an endemic disease in Southeast Asia and northern Australia (2, 4) and also occurs sporadically in other parts of the world (3, 7). Human melioidosis was reported to occur in Brazil only in 2003, when a family outbreak afflicted four sisters in the rural part of the municipality of Tejuçuoca, Ceará state (14). After this episode, there was one reported case of melioidosis in 2004 in the rural area of Banabuiú, Ceará (14). And in 2005, a case of melioidosis associated with near drowning after a car accident was confirmed to occur in Aracoiaba, Ceará (11).The goal of this study was to investigate the Tejuçuoca and Banabuiú municipalities, where human cases of melioidosis have been confirmed to occur, and to gain a better understanding of the ecology of Burkholderia pseudomallei in this region.We chose as central points of the study the residences and surrounding areas of the melioidosis patients in the rural areas of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50′W) (Fig. (Fig.1).1). There are two well-defined seasons in each of these locations: one rainy (running from January to May) and one dry (from June to December). A total of 600 samples were collected at five sites in Tejuçuoca (T1, T2, T3, T4, and T5) and five in Banabuiú (B1, B2, B3, B4, and B5), distributed as follows (Fig. (Fig.2):2): backyards (B1 and T1), places shaded by trees (B2 and T2), water courses (B3 and T3), wet places (B4 and T4), and stock breeding areas (B5 and T5).Open in a separate windowFIG. 1.Municipalities of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50″W).Open in a separate windowFIG. 2.Soil sampling sites in Banabuiú and Tejuçuoca.Once a month for 12 months (a complete dry/rainy cycle), five samples were gathered at five different depths: at the surface and at 10, 20, 30 and 40 cm (Table (Table1).1). The samples were gathered according to the method used by Inglis et al. (9). Additionally, the sample processing and B. pseudomallei identification were carried out as previously reported (1, 8, 9).

TABLE 1.

Distribution of samples with isolates by site and soil depth
Sitesa and depth (cm)No. of B. pseudomallei isolates in samples from:
Banabuiú (n = 300)Tejuçuoca (n = 300)Total (n = 600)
B1/T13
    Surface2
    10
    201
    30
    40
B2/T21
    Surface1
    10
    20
    30
    40
B3/T315
    Surface2
    102
    204
    303
    404
B4/T45
    Surface
    101
    201
    3011
    401
B5/T52
    Surface
    10
    20
    302
    40
Total62026
Open in a separate windowaSites designated with B are in Banabuiú, and sites designated with T are in Tejuçuoca. See the text for details.The data on weather and soil composition were obtained from specialized government institutions, such as FUNCEME, IPECE, and EMBRAPA. The average annual temperature in both municipalities is between 26 and 28°C. In 2007, the annual rainfall in Tejuçuoca was 496.8 mm, and that in Banabuiú was 766.8 mm. There are a range of soil types in both Tejuçuoca and Banabuiú: noncalcic brown, sodic planossolic, red-yellow podzolic, and litholic. In Banabuiú, there are also alluvial and cambisol soils. The characteristic vegetation in both municipalities is caatinga (scrublands).There were isolates of B. pseudomallei in 26 (4.3%) of the 600 samples collected. The bacterium was isolated at a rate (3%) similar to that previously reported (9). The bacterium isolation occurred in both the dry (53.8%) and the rainy (46.2%) seasons. Tejuçuoca represented 76.9% (20/26) of the strains isolated. Four sites in Tejuçuoca (T1, T3, T4, and T5) and three in Banabuiú (B1, B2, and B4) presented isolates of the bacterium (Table (Table1).1). The isolation of the B. pseudomallei strains varied from the surface down to 40 cm. However, 17 of the 26 positive samples (65.3%) were found at depths between 20 and 40 cm (Table (Table1).1). Only two isolates were found at the surface during the dry season.A study in Vietnam (13) and one in Australia (9) reported the presence of B. pseudomallei near the houses of melioidosis patients. In our study, the same thing happened. Site T3 (15/26; 57.6%) was located 290 m from the patient''s house, as reported by the Rolim group (14).B. pseudomallei was isolated from a sheep paddock in Australia, where animals sought shelter below mango and fig trees (17). In our study, the bacterium was isolated at site T5, a goat corral alongside the house where the outbreak occurred in Tejuçuoca. Four sites in places shaded by trees yielded positive samples (30.7%) in both Tejuçuoca (palm trees) and Banabuiú (mango trees). Additionally, B. pseudomallei was isolated on three occasions from a cornfield (site 4B) located alongside the house of the melioidosis patient in Banabuiú.In the main areas of endemicity, the disease is more prevalent in the rainy season (4, 5, 16). The outbreak in Tejuçuoca was related to rainfall (14). Besides the association of cases of the disease with rainfall itself, the isolation of B. pseudomallei in soil and water was also demonstrated during the dry season (12, 15). An Australian study isolated strains from soil and water during the dry and rainy seasons (17). A Thai study also reported B. pseudomallei in the dry season (18). In our study, the isolation of B. pseudomallei took place either at the end of the wet season or in the dry months. Fourteen of the positive samples (53.8%) were collected during the dry season, albeit near a river or reservoir (sites T3 and B4).Physical, biological, and chemical soil features appear to influence the survival of B. pseudomallei (6, 10). In the present study, the soil was classified as litholic with sandy or clayey textures. It is susceptible to erosion, and when there is a lack of water, it is subject to salinization. During the dry season, the clay layer becomes dried, cracked, and very hard. During the rainy season, it becomes soggy and sticky. The isolation of B. pseudomallei in the dry season is possibly related to the capacity for adaptation of this soil, since the extreme conditions of lithosols do not prevent the bacterial growth and survival.It has been shown that B. pseudomallei is more often isolated at depths between 25 and 45 cm (17). In our study, 65.3% of the positive samples were taken at depths between 20 and 40 cm. Moreover, of these 17 samples, 10 (58.8%) were collected during the dry months. Also, unlike in other regions, two positive samples were taken from the surface in the period without rainfall.The rainfall in Tejuçuoca and Banabuiú is generally low, and temperatures do not vary significantly during the year. Therefore, the isolation of B. pseudomallei in these places occurs outside the rainfall, temperature, and moisture conditions observed in other regions of endemicity. Our data thus suggest that peculiar environmental features, such as soil composition, might favor the multiplication of B. pseudomallei in northeast Brazil.  相似文献   

5.

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 SpilloverEducationProductivitySavings
Microeconomic level    
Total fertility rateChild survivalEnrollment in schoolIncomeMoney
Change in fertilityChild wellbeing and behaviorYears of schoolingPurchasing powerAssets
Age at first birth/ teenage pregnanciesAnthropometryEarly drop outPerformance
Birth 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   

6.
Intra- and inter-generic transfer of pathogenicity island-encoded virulence genes by cos phages     
John Chen  Nuria Carpena  Nuria Quiles-Puchalt  Geeta Ram  Richard P Novick  José R Penadés 《The ISME journal》2015,9(5):1260-1263
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   

7.
Quantitative Fluorescence In Situ Hybridization of Microbial Communities in the Rumens of Cattle Fed Different Diets     
Yunhong Kong  Maolong He  Tim McAlister  Robert Seviour  Robert Forster 《Applied and environmental microbiology》2010,76(20):6933-6938
At present there is little quantitative information on the identity and composition of bacterial populations in the rumen microbial community. Quantitative fluorescence in situ hybridization using newly designed oligonucleotide probes was applied to identify the microbial populations in liquid and solid fractions of rumen digesta from cows fed barley silage or grass hay diets with or without flaxseed. Bacteroidetes, Firmicutes, and Proteobacteria were abundant in both fractions, constituting 31.8 to 87.3% of the total cell numbers. They belong mainly to the order Bacteroidales (0.1 to 19.2%), hybridizing with probe BAC1080; the families Lachnospiraceae (9.3 to 25.5%) and Ruminococcaceae (5.5 to 23.8%), hybridizing with LAC435 and RUM831, respectively; and the classes Deltaproteobacteria (5.8 to 28.3%) and Gammaproteobacteria (1.2 to 8.2%). All were more abundant in the rumen communities of cows fed diets containing silage (75.2 to 87.3%) than in those of cows fed diets containing hay (31.8 to 49.5%). The addition of flaxseed reduced their abundance in the rumens of cows fed silage-based diets (to 45.2 to 58.7%) but did not change markedly their abundance in the rumens of cows fed hay-based diets (31.8 to 49.5%). Fibrolytic species, including Fibrobacter succinogenes and Ruminococcus spp., and archaeal methanogens accounted for only a small proportion (0.4 to 2.1% and 0.2 to 0.6%, respectively) of total cell numbers. Depending on diet, between 37.0 and 91.6% of microbial cells specifically hybridized with the probes used in this study, allowing them to be identified in situ. The identities of other microbial populations (8.4 to 63.0%) remain unknown.The rumen is an anaerobic ecosystem used by herbivores to convert fibrous plant material into fermentation products that are in turn used as energy by the host. Fibrolytic degradation is accomplished by a complex microbial community which includes specialized fungi, protozoa, and bacteria (14). More than 200 bacterial species (5) have been isolated from rumen, and many of these have been phylogenetically and physiologically characterized. Several of these, including Fibrobacter succinogenes, Ruminococcus albus, and Ruminococcus flavefaciens, have the ability to hydrolyze cellulose in axenic culture (24). Despite the presence of these fibrolytic populations, a large portion of the fiber in low-quality forage diets passes through the rumen undigested. In the rumen, fibrolytic bacteria do not digest plant cell walls in isolation but rather interact with a consortium of bacteria (18). Although culture-dependent studies have improved our understanding of rumen microbiology, the importance of the isolates to the structure and function of the rumen microbial community, with the possible exception of the fibrolytic strains, is still unknown. Expanding our knowledge of the structure and function of the rumen microbial community may provide insights into approaches to improve the efficiency of fiber digestion and biofuel production (14).To provide a high-resolution view of the population structure of the rumen bacterial community, we used quantitative fluorescence in situ hybridization (qFISH) to investigate the composition and distribution of bacterial populations associated with the liquid and solid rumen contents from 12 ruminally cannulated Holstein dairy cows (3 cows were used for each diet) fed (for at least 21 days) grass hay or barley silage diets with or without flaxseed (Table (Table1).1). Six new 16S rRNA-targeted FISH probes (Table (Table2)2) for not only the fibrolytic groups but also other unclassified bacterial groups in the rumen were designed, using ARB software (17), against the rumen 16S rRNA gene sequences (data not shown) retrieved from the Ribosomal Database Project (RDP) database (6). The new probes target Bacteroidales-related clones (probe BAC1080) (phylum Bacteroidetes), Lachnospiraceae- and Ruminococcaceae-related clones (probes LAC435 and RUM831, respectively) (phylum Firmicutes), Butyrivibrio fibrisolvens-related clones (probe BFI826), and R. albus- and R. flavefaciens-related clones (probes RAL1436 and RFL155, respectively).

TABLE 1.

Composition of diets used in this study
IngredientDiet composition (% dry weight)
Hay-based dietHay and flaxseed dietSilage-based dietSilage and flaxseed diet
Alfalfa grass hay (chopped)47.547.500
Barley silage0047.547.5
Steamed rolled barley grain47.532.547.532.5
Ground flaxseeds015015
Other5555
Open in a separate window

TABLE 2.

Oligonucleotide probes and their target populations used in this study for FISH analyses
Probe nameaTarget rRNADesigned target(s)% FAbReference
EUB338 (00159)16SDomain Bacteria0-5016
EUB338II (00160)16SPhylum Planctomycetes0-5016
EUB338III (00161)16SPhylum Verrucomicrobia0-5016
NONEUB (00243)16SControl probe complementary to EUB3380-5016
ALF968 (00021)16SClass Alphaproteobacteria, phylum Proteobacteria2016
BET42a (00034)23SClass Betaproteobacteria, phylum Proteobacteria3516
GAM42a (00174)23SClass Gammaproteobacteria, phylum Proteobacteria3516
SRB385 (00300)16SClass Deltaproteobacteria, phylum Proteobacteria3516
SRB385Db (00301)16SClass Deltaproteobacteria, phylum Proteobacteria3516
HGC69a (00182)23SPhylum Actinobacteria2516
GNSB941 (00718)16SPhylum Chloroflexi3516
CFX1223 (00719)16SPhylum Chloroflexi3516
SPIRO1400 (01004)16SSubgroup of family Spirochaetaceae2016
TM7-905 (00600)16SCandidate phylum TM72016
LGC354A (00195)16SPhylum Firmicutes3516
LGC354B (00196)16SPhylum Firmicutes3516
LGC354C (00197)16SPhylum Firmicutes3516
RUM83116SRumen clones in family Ruminococcaceae, phylum Firmicutes35This study
RAL143616SRuminococcus albus-related clones, phylum Firmicutes20This study
RFL15516SRuminococcus flavefaciens-related clones, phylum Firmicutes45This study
LAC43516SClones in family Lachnospiraceae, phylum Firmicutes35This study
BFI82616SButyrivibrio fibrisolvens-related clones, phylum Firmicutes35This study
BAC108016SClones in order Bacteroidales, phylum Bacteroidetes20This study
Fibr225 (00005)16SFibrobacter succinogenes-related clones, phylum Fibrobacteres20c16
ARCH915 (00027)16SDomain Archaea2016
Open in a separate windowaThe numbers in parentheses after the probe names represent the probe accession numbers in probeBase (16).bFA, formamide concentration used in the FISH buffer.cThe optimum formamide concentration for the probe was determined in this study.The optimal formamide concentrations (OFC) of the new probes used in FISH were assessed in different ways. Probes RUM831 and BAC1080 were assessed by using pure cultures of Ruminococcus and Prevotella strains with zero and one mismatch (Fig. (Fig.1)1) to the probes. The OFC of probes LAC435 and BFI826 were assessed using Clone-FISH (21) with zero and one mismatch 16S rRNA clone (Fig. (Fig.1)1) by following the procedure described previously (9, 10). The highest formamide concentration (tested in 5% stepwise increases) at which a clear fluorescent signal was observed with the reference bacterium or competent cells with zero mismatches after FISH probing, but not with bacteria or competent cells with one mismatch, was selected. The OFC of probes FIB225 (designed by Stahl et al. [23]), RFL155, and RAL1436 were assessed using only pure cultures of F. succinogenes, R. flavefaciens, and R. albus, respectively, all having perfect matches to each probe (Fig. (Fig.1).1). The highest formamide concentration (tested in 5% stepwise increases) at which a clear fluorescent signal was observed with the reference bacterium after FISH probing was selected. These probes were employed with other available probes (Table (Table2)2) chosen from probeBase (16) based on the alignment and classification of the 16S rRNA gene sequences retrieved from rumen communities.Open in a separate windowFIG. 1.Alignments of the probe sequences and their target sites and sequences of corresponding sites in reference bacteria or clones. The probe names in parentheses after the abbreviated names are according to Oligonucleotide Probe Database nomenclature (2). Only the nucleotides that are different from target sequences are shown. E, empty space; R., Ruminococcus; P., Prevotella; F., Fibrobacter.The digest samples from the top, bottom, and middle of the rumen were collected through a cannula, thoroughly mixed, and fractioned as liquid fraction (LiqF) and solid fraction (SolF). On-site, about 100 ml was transferred to a heavy-wall 250-ml beaker and squeezed using a Bodum coffee maker plunger (Bodum Inc., Triengen, Switzerland). The extruded liquid samples (containing the planktonic cells) were fixed in ethanol and paraformaldehyde (PFA) for FISH probing (3). The remaining liquid was discarded, and the squeezed particulate samples (used to collect particulate-attached cells) were washed with 100 ml phosphate buffer (5.23 g/liter K2HPO4, 2.27 g/liter KH2PO4, 3.00 g/liter NaHCO3, and 20 ml/liter 2.5% cysteine HCl) by stirring gently with a spatula, followed by squeezing again and decanting. Washed particulate samples (5 g) were then fixed for FISH as described above.After fixation, the particulate samples plus the fixation solution were transferred into a stomacher bag and “stomached” (Stomacher 400 Circulator, Seaward England) at 230 rpm for 6 min. Treated samples were then transferred into a clean 250-ml beaker and squeezed again. Microscopic examination of the squeezed residues after DAPI (4′,6-diamidino-2-phenylindole) staining (100 μl [0.003 mg/ml] for 10 min) showed only a few bacterial cells attached on the plant fibers, indicating that most bacterial cells had been “stomached” into the liquid (data not shown). To recover cells, filtrates were centrifuged (5,000 × g), and the cell pellet was washed three times with phosphate buffer before being used for FISH probing. On the day of sampling, each cow was sampled twice, at 1100 h and 1600 h. The liquid FISH samples obtained from the 3 cows fed with the same diet (at two different sampling times) were mixed, as were the particulate FISH samples, and used in qFISH analysis. FISH was carried out according to Amann (3). FISH was carried out on glass coverslips (24 by 60 mm) coated with gelatin (9). DAPI staining of biomass samples was carried out after FISH probing. FISH and DAPI images were captured with a Zeiss epifluorescence microscope (Zeiss PM III) equipped with a Canon 5D Mark II camera. Raw images captured randomly were transferred into gray TIF images and sharpened in Adobe Photoshop CS3. Cells stained with DAPI and hybridized to the probes were enumerated using the function provided in ImageJ (1). The percent compositions of these probe-defined groups (against all DAPI-stained cells in the same microscopic field) in the different fractions of rumen contents from cows fed different diets are presented in Table Table33.

TABLE 3.

Distribution and composition of FISH probe-defined groups in rumen microbial communities in cows fed with different diets
Probe-defined microbial groupComposition (mean value [%] ± SD)a
Hay-based diet
Hay and flaxseed diet
Silage-based diet
Silage and flaxseed diet
LiqFSolFLiqFSolFLiqFSolFLiqFSolF
BAC10809.6 ± 1.330.1 ± 0.0219.2 ± 3.714.2 ± 0.7214.2 ± 3.1118.8 ± 3.8814.4 ± 2.8916.7 ± 4.33
ALF9680.2 ± 0.020.2 ± 0.020.2 ± 0.030.2 ± 0.040.7 ± 0.141.5 ± 0.410.1 ± 0.010.1 ± 0.01
BET42a000.6 ± 0.011.2 ± 0.270.1 ± 0.01<0.10.4 ± 0.060.2 ± 0.04
GAM42a3.2 ± 0.534.4 ± 0.574.2 ± 0.764.5 ± 0.672.0 ± 0.321.2 ± 0.238.2 ± 1.235.3 ± 0.95
SRBmix5.8 ± 0.8811.6 ± 2.439.0 ± 1.5210.1 ± 2.5628.3 ± 4.4323.3 ± 4.547.7 ± 0.7813.2 ± 2.22
CHLmix1.7 ± 0.2700.5 ± 0.010 ± 00.2 ± 0.020.4 ± 0.070.1 ± 0.010.1 ± 0.02
SPIRO14000.5 ± 0.091.9 ± 0.321.7 ± 0.332.0 ± 0.211.4 ± 0.311.9 ± 0.330.4 ± 0.030.4 ± 0.07
TM7-9050.6 ± 0.080.8 ± 0.070.5 ± 0.010.1 ± 0.031.5 ± 0.230.2 ± 0.020.6 ± 0.020.3 ± 0.08
HGC69a1.3 ± 0.282.1 ± 0.310.3 ± 0.060.3 ± 0.050.4 ± 0.030.1 ± 0.020.5 ± 0.090.2 ± 0.02
RUM8315.5 ± 0.135.7 ± 0.895.8 ± 0.738.9 ± 1.3218.0 ± 4.1323.8 ± 3.115.6 ± 1.147.4 ± 1.32
RAL14360.4 ± 0.060.3 ± 0.030.2 ± 0.060.2 ± 0.030.3 ± 0.050.6 ± 0.090.7 ± 0.130.6 ± 0.12
RFL1550.7 ± 0.110.2 ± 0.030.3 ± 0.070.7 ± 0.190.1 ± 0.010.8 ± 0.110.5 ± 0.061.2 ± 0.34
LAC43525.5 ± 3.9810.0 ± 1.519.6 ± 1.3111.7 ± 1.6712.6 ± 2.5620.2 ± 3.239.3 ± 1.5116.1 ± 3.31
BFI8260.3 ± 0.060.4 ± 0.050.4 ± 0.060.7 ± 0.120.5 ± 0.050.3 ± 0.082.4 ± 0.370.2 ± 0.02
Fibr225000.2 ± 0.040.1 ± 0.020.8 ± 0.140.7 ± 0.140.4 ± 0.110.1 ± 0.04
ARCH9150.3 ± 0.080.2 ± 0.070.6 ± 0.010.3 ± 0.070.6 ± 0.090.1 ± 0.020.4 ± 0.050.4 ± 0.06
Total hybridizedb54.13752.443.780.991.64860.7
Otherc45.96347.656.319.18.45239.3
Open in a separate windowaThe two numbers represent the mean value (%) and the standard deviation of individual probe-defined microbial groups in a specified rumen digest fraction, which were calculated based on 3 mean values, each consisting of 20 enumerations.bThe numbers represent the sum of percentages of all individual probe-defined microbial groups in a specified rumen digest fraction. The percentages obtained with FISH probes RAL1436, RFL155, and BFI826 were not counted in the sum because the bacterial cells hybridizing with the former two probes also hybridized with RUM831, and the bacterial cells hybridizing with the last probe also hybridized with probe LAC435.cThe numbers represent the percentages of microorganisms which were not identified by FISH in a specified rumen digest fraction.We provided quantitative data by using qFISH to show that Bacteroidetes, Firmicutes, and Proteobacteria were abundant in both the LiqF and the SolF, constituting 31.8 to 87.3% of the total cell numbers. These FISH data add weight to the view that Firmicutes and Bacteroidetes might be dominant in rumens, as suggested previously from their high ratios retrieved from 16S rRNA clone libraries (e.g., see references 12, 26, and 27). However, information emerging from 16S rRNA gene clone library data cannot be used to reach conclusions on the quantitative composition of the rumen bacterial community. Bacteria may have 1 to 14 copies of rRNA genes, and several biases are known to be associated with their PCR amplification (8).These 3 dominant bacterial groups have been identified at a high-resolution level. They belong mainly to the order Bacteroidales (0.1 to 19.2%), hybridizing with probe BAC1080 (Fig. (Fig.22 A); the families Lachnospiraceae (9.3 to 25.5%) and Ruminococcaceae (5.5 to 23.8%), hybridizing with LAC435 (Fig. (Fig.2E)2E) and RUM831 (Fig. (Fig.2D),2D), respectively; and the classes Deltaproteobacteria (5.8 to 28.3%) and Gammaproteobacteria (1.2 to 8.2%), hybridizing with SRBmix (equal moles of SRB385 and SRB385Db) (Fig. (Fig.2C)2C) and GAM42a (Fig. (Fig.2B),2B), respectively. All were more abundant in the microbial communities in the rumens of cows fed diets containing silage (75.2 to 87.3%) than in those in the rumens of cows fed diets containing hay (31.8 to 49.5%). These results show how diets containing different forages (hay or silage) may influence the distribution of the microbial populations, which is in line with data by Tajima et al. (25). We also found in this study that the addition of flaxseed (to inhibit methane emission) reduced their abundance in the rumens of cows fed silage-based diets (to 45.2 to 58.7%) but did not change markedly their abundance in the rumens of cows fed hay-based diets (31.8 to 49.5%), suggesting that adding flaxseed to these diets also affected rumen microbial community composition, although the extent of its influence reflected the forage used, being more profound with a silage-based diet than when hay was used.Open in a separate windowFIG. 2.Images of digest samples from the rumens of cows fed hay- or silage-based diets with and without flaxseed after color combination. Images from probes are labeled in red, and those from DAPI staining are in green. The yellow (combination of red and green), including those partly colored cells in panels A to F, hybridized with probes BAC1080, GAM42a, SRBmix, RUM831, LAC435, and ARCH915, respectively. A few cells (arrows) hybridizing with SRBmix (C) were not stained by DAPI. Bars, 10 μm.We also present evidence here to suggest that Proteobacteria are common members of the microbial community, with sulfur-reducing bacteria (SRB) belonging to Deltaproteobacteria in particular being readily detected (up to 28% of the total cells) in both the LiqF and the SolF of rumen contents from cows fed the four different diets examined here. SRB have seldom been retrieved in clone libraries obtained from rumen samples. Lin et al. (15) have estimated SRB abundance in the rumen using DNA hybridization and concluded that they were of minor importance (0.7 to 0.8% of the total rRNA). Our estimates are much higher than those for every diet regime examined, possibly reflecting the coverage of the probes used in the two different studies. The probe mixture SRBmix used here targets most members of the Deltaproteobacteria, while those of Lin et al. (15) covered mainly members of the Desulfobacteraceae, Desulfovibrionaceae, and Desulfobulbaceae. We also recognized that the probe mixture SRBmix perfectly matched with the 16S rRNA genes of some bacteria other than SRB in Deltaproteobacteria. The possibility of overestimation of SRB cannot be ruled out. Interestingly, our data suggest that Gammaproteobacteria were abundant in some of the rumen communities we examined by FISH, comprising 1.2 to 8.2% of total cells.The other unexpected finding was that the fibrolytic bacteria and archaeal methanogens accounted for only a minor fraction of the communities. Of the three characterized fibrolytic bacterial species, F. succinogenes was not detected in the rumen digesta from cattle fed the hay-based diet but was present in the remainder of the diets. In contrast, R. albus and R. flavefaciens were present in both the LiqF and the SolF of the rumen digesta from cows fed all four diets. Although the importance of these bacteria within the rumen microbial community cannot be denied, these three populations accounted for only 0.7 to 2.1% of the total microbial cells. This numerical range compares well with that determined previously for F. succinogenes (0.1 to 6.9% of total rRNA) (4, 23) and Ruminococcus spp. (1.5 to 2.9% of total rRNA) (11), considering that different animals and diets were used in those studies and that different specificities of the probes and different detection methods were used. However, this is much lower than the 9% (of total rRNA) detected by Michalet-Doreau et al. (19) in their work. The abundance of fibrolytic B. fibrisolvens-related species was also low, being present at <1% in all fractions, except in the LiqF in cows fed the mixture of silage and flaxseed, where they contributed 2.4% of total cells.Methanogens hybridized to ARCH915 (Fig. (Fig.2F)2F) were present (0.1 to 0.6%) in all rumen samples examined by FISH, which is close to or within the range (0.3 to 3.3%) estimated in other studies (15, 22). Interestingly, no marked difference in abundance of the methanogens could be seen between the samples from the rumens of cows fed diets with flaxseed and those from the rumens of cows fed diets without flaxseed, although it has been reported (7) that the addition of fatty acids could decrease methane production in the rumen. This may be due to the presence of methanogens with different activities in different rumen samples or the inability of probe ARCH915 to hybridize to all methanogens in the rumen samples examined here.Bacteria belonging to Chloroflexi, TM7, Spirochetes, and Actinobacteria hybridizing with CHLmix, TM7-905, SPRO1400, and HGC69a, respectively, accounted for only a minor fraction of the total cell numbers observed. In most cases, their abundances in each fraction did not change markedly with diet, always being present in small numbers (0 to 1%), suggesting that they have a minor role there. This conclusion, however, has to be confirmed since many (8.4 to 63.0%, depending on diet) of the bacteria could not be identified in the rumens of cows fed with all diets except the silage-based diet (Table (Table33).FISH with the probes designed in this study failed to identify all of the bacterial cells. This is because the probes do not target all rumen 16S rRNA gene sequences and/or the true extent of rumen biodiversity has not been revealed from cloning analyses. This indicates that our current understanding of the quantitative composition of the rumen microbial community is far from complete. Moreover, no physiological data were generated in this study to suggest what the role(s) of most of the dominant populations (except the SRB hybridized with probe SRBmix) identified by FISH might be, meaning that it is still not possible to link their abundance to their in situ function. Furthermore, each FISH-probed population probably includes bacteria with different phenotypes. Clearly, much needs to be done before the structure and function of the rumen microbial community are fully understood.FISH is a useful tool in the investigation of microbial composition in complex ecosystems (3). However, FISH probes targeting rumen bacterial populations are limited. By comparison with other culture-independent methods, e.g., quantitative PCR, FISH has several advantages (8). In particular, in combination with histochemical staining methods (20) and microautoradiography (MAR-FISH) (13), the in situ ecophysiology of a targeted population can be determined under specified electron acceptor conditions. These techniques may provide important clues as to the functional role of microbial populations within complex communities, like that of the rumen. The possession of the FISH probes described in this paper could allow such studies to be undertaken in herbivore rumens.  相似文献   

8.
Intraspecies Signaling Involving the Diffusible Signal Factor BDSF (cis-2-Dodecenoic Acid) Influences Virulence in Burkholderia cenocepacia     
Robert P. Ryan  Yvonne McCarthy  Steven A. Watt  Karsten Niehaus  J. Maxwell Dow 《Journal of bacteriology》2009,191(15):5013-5019
  相似文献   

9.
RecA-Independent DNA Damage Induction of Mycobacterium tuberculosis ruvC Despite an Appropriately Located SOS Box     
Lisa F. Dawson  Joanna Dillury  Elaine O. Davis 《Journal of bacteriology》2010,192(2):599-603
  相似文献   

10.
RNA-Based Investigation of Ammonia-Oxidizing Archaea in Hot Springs of Yunnan Province,China     
Hongchen Jiang  Qiuyuan Huang  Hailiang Dong  Peng Wang  Fengping Wang  Wenjun Li  Chuanlun Zhang 《Applied and environmental microbiology》2010,76(13):4538-4541
  相似文献   

11.
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   

12.
Widespread Distribution of Cell Defense against d-Aminoacyl-tRNAs     
Sandra Wydau  Guillaume van der Rest  Caroline Aubard  Pierre Plateau    Sylvain Blanquet 《The Journal of biological chemistry》2009,284(21):14096-14104
Several l-aminoacyl-tRNA synthetases can transfer a d-amino acid onto their cognate tRNA(s). This harmful reaction is counteracted by the enzyme d-aminoacyl-tRNA deacylase. Two distinct deacylases were already identified in bacteria (DTD1) and in archaea (DTD2), respectively. Evidence was given that DTD1 homologs also exist in nearly all eukaryotes, whereas DTD2 homologs occur in plants. On the other hand, several bacteria, including most cyanobacteria, lack genes encoding a DTD1 homolog. Here we show that Synechocystis sp. PCC6803 produces a third type of deacylase (DTD3). Inactivation of the corresponding gene (dtd3) renders the growth of Synechocystis sp. hypersensitive to the presence of d-tyrosine. Based on the available genomes, DTD3-like proteins are predicted to occur in all cyanobacteria. Moreover, one or several dtd3-like genes can be recognized in all cellular types, arguing in favor of the nearubiquity of an enzymatic function involved in the defense of translational systems against invasion by d-amino acids.Although they are detected in various living organisms (reviewed in Ref. 1), d-amino acids are thought not to be incorporated into proteins, because of the stereospecificity of aminoacyl-tRNA synthetases and of the translational machinery, including EF-Tu and the ribosome (2). However, the discrimination between l- and d-amino acids by aminoacyl-tRNA synthetases is not equal to 100%. Significant d-aminoacylation of their cognate tRNAs by Escherichia coli tyrosyl-, tryptophanyl-, aspartyl-, lysyl-, and histidyl-tRNA synthetases has been characterized in vitro (39). Recently, using a bacterium, transfer of d-tyrosine onto tRNATyr was shown to occur in vivo (10).With such misacylation reactions, the resulting d-aminoacyl-tRNAs form a pool of metabolically inactive molecules, at best. At worst, d-aminoacylated tRNAs infiltrate the protein synthesis machinery. Although the latter harmful possibility has not yet been firmly established, several cells were shown to possess a d-tyrosyl-tRNA deacylase, or DTD, that should help them counteract the accumulation of d-aminoacyl-tRNAs. This enzyme shows a broad specificity, being able to remove various d-aminoacyl moieties from the 3′-end of a tRNA (46, 11). Such a function makes the deacylase a member of the family of enzymes capable of editing in trans mis-aminoacylated tRNAs. This family includes several homologs of aminoacyl-tRNA synthetase editing domains (12), as well as peptidyl-tRNA hydrolase (13, 14).Two distinct deacylases have already been discovered. The first one, called DTD1, is predicted to occur in most bacteria and eukaryotes (see d-amino acids, including d-tyrosine (6). In fact, in an E. coli Δdtd strain grown in the presence of 2.4 mm d-tyrosine, as much as 40% of the cellular tRNATyr pool becomes esterified with d-tyrosine (10).

TABLE 1

Distribution of DTD1 and DTD2 homologs in various phylogenetic groupsHomologs of DTD1 and DTD2 were searched for using a genomic Blast analysis against complete genomes in the NCBI Database (www.ncbi.nlm.nih.gov). Values in the table are number of species. For instance, E. coli is counted only once in γ-proteobacteria despite the fact that several E. coli strains have been sequenced.
DTD1DTD2DTD1 + DTD2None
Bacteria
    Acidobacteria 2 0 0 0
    Actinobacteria 27 0 0 8
    Aquificae 1 0 0 0
    Bacteroidetes/Chlorobi 12 0 0 5
    Chlamydiae 1 0 0 6
    Chloroflexi 4 0 0 0
    Cyanobacteria 5 0 0 16
    Deinococcus/Thermus 4 0 0 0
    Firmicutes
        Bacillales 19 0 0 0
        Clostridia 19 0 0 0
        Lactobacillales 23 0 0 0
        Mollicutes 0 0 0 15
    Fusobacteria/Planctomycetes 2 0 0 0
    Proteobacteria
        α 6 0 0 55
        β 24 0 0 11
        γ 80 0 0 8
        δ 15 0 0 0
        ε 1 0 0 12
    Spirochaetes 0 0 0 7
    Thermotogae 5 0 0 0
Archaea
    Crenarchaeota 0 13 0 0
    Euryarchaeota 1 26 0 2
    Nanoarchaeota 0 0 0 1
Eukaryota
    Dictyosteliida 1 0 0 0
    Fungi/Metazoa
        Fungi 13 0 0 1
        Metazoa 19 0 0 0
    Kinetoplastida 3 0 0 0
    Viridiplantae 4 4 4 0
Open in a separate windowHomologs of dtd/DTD1 are not found in the available archaeal genomes except that of Methanosphaera stadtmanae. A search for deacylase activity in Sulfolobus solfataricus and Pyrococcus abyssi led to the detection of another enzyme (DTD2), completely different from the DTD1 protein (15). Importing dtd2 into E. coli functionally compensates for dtd deprivation. As shown in 16).Several cells contain neither dtd nor dtd2 homologs (d-tyrosyl-tRNA deacylase (DTD3). This protein, encoded by dtd3, behaves as a metalloenzyme. Sensitivity of the growth of Synechocystis to external d-tyrosine is strongly exacerbated by the disruption of dtd3. Moreover, expression of the Synechocystis DTD3 in a Δdtd E. coli strain, from a plasmid, restores the resistance of the bacterium to d-tyrosine. Finally, using the available genomes, we examined the occurrence of DTD3 in the living world. The prevalence of DTD3-like proteins is surprisingly high. It suggests that the defense of protein synthesis against d-amino acids is universal.  相似文献   

13.
Identification of Enhancer Binding Proteins Important for Myxococcus xanthus Development     
Krista M. Giglio  Jessica Eisenstatt  Anthony G. Garza 《Journal of bacteriology》2010,192(1):360-364
  相似文献   

14.
Isolation of VanA-Type Vancomycin-Resistant Enterococcus Strains from Domestic Poultry Products with Enrichment by Incubation in Buffered Peptone Water at 42°C     
Tetsuya Harada  Masashi Kanki  Takao Kawai  Masumi Taguchi  Tsutomu Asao  Yuko Kumeda 《Applied and environmental microbiology》2010,76(15):5317-5320
Eight VanA-type enterococcal strains were isolated from 8 of 171 domestic poultry products by using enrichment by incubation in buffered peptone water at 35°C and 42°C. The pulsed-field gel electrophoresis patterns of all six VanA-type Enterococcus faecalis isolates were nearly indistinguishable, indicating the presence of a specific clone in Japan.The use of avoparcin as a growth promoter and prophylactic agent in feed has been associated with the occurrence of vancomycin-resistant enterococci (VRE) in farm animals in the European Union (EU) (2, 3, 4, 11). In Japan, avoparcin was also used as a food additive on animal farms for approximately 7 years until it was banned in March 1997 (10, 22). However, there has been no documentation in the literature on the isolation of VanA-type VRE from domestic poultry products in Japan (8, 10). Several traditional methods for detecting enterococci, including VRE, in poultry products include an enrichment step using buffered peptone water (BPW) (7, 12, 14, 21). Although BPW is typically incubated at 35 to 37°C, based on the reported optimal growth temperature for enterococci of 35°C (19, 20), the growth of VRE at this temperature can be inhibited by antagonistic activities of background microflora in samples contaminated with low levels of VRE, as 35 to 37°C is also the optimal growth temperature for most mesophilic microorganisms. Several investigators have demonstrated that incubation at an elevated temperature using selective enrichment medium was useful in the isolation of enterococcal strains from retail meat (9, 17, 18). To our knowledge, however, no study has evaluated cultivation temperature during the enrichment process in the isolation of VanA-type VRE from poultry products to examine the influence of background microorganisms on VanA-type VRE growth in BPW. In this study, we carried out isolation of VanA-type VRE from retail Japanese domestic chicken products by using two different enrichment temperatures (35°C and 42°C) and characterized VanA-type VRE isolates by antimicrobial susceptibility testing and pulsed-field gel electrophoresis (PFGE). We also assessed the effects of incubation temperature on the growth of VanA-type VRE under pure culture conditions, on the rates of recovery of inoculated VanA-type VRE strains from artificially contaminated poultry samples, and on the inhibition of suspected enterococcal strains by background microorganisms in poultry samples.A total of 171 retail domestic poultry products obtained in Osaka, Japan, between September 2008 and April 2009 were examined for the presence of VanA-type VRE. Twenty-five grams of each sample was removed aseptically and placed in a sterile plastic bag (30 by 19 cm), to which 225 ml of BPW (Eiken Chemical Co., Ltd., Tokyo, Japan) was added. The sample was then incubated for 24 h at either 35°C or 42°C. Ten microliters of the culture supernatants was spread plated on Enterococcosel agar (Becton, Dickinson and Company, Sparks, MD) plates containing 4 μg/ml vancomycin (Wako Pure Chemical Industries, Ltd., Hyogo, Japan) and incubated at 36°C for 48 h. Screening for the vanA gene was done using the colony-sweep PCR method with phosphate-buffered saline (PBS) washed-cell suspension, Z Taq polymerase (Takara Biochemicals, Shiga, Japan), and previously described primers (1, 16). PCR-positive sweeps were streaked for isolated VanA-type VRE colonies on an Enterococcosel agar plate containing 32 μg/ml vancomycin. The plates were incubated at 36°C for 48 h, and suspected VRE colonies were confirmed as enterococci by conventional biochemical methods (8). In addition, identification of Enterococcus faecalis and Enterococcus faecium harboring the vanA gene was performed by a multiple PCR assay previously described by Depardieu et al. (6). For the five VanA-positive poultry product samples obtained in October 2008 and April 2009, the number of VanA-type VRE was determined by plating 1-ml volumes of samples diluted 10-fold in sterile peptone saline onto Enterococcosel agar plates containing 32 μg/ml vancomycin. Colonies were counted after incubation at 36°C for 48 h. MICs of vanA-genotype VRE isolates to vancomycin and teicoplanin were determined using the Etest method (AB Biodisk, Solna, Sweden) in accordance with the criteria of the Clinical and Laboratory Standards Institute (5). Subtyping of VanA-type VRE isolates was performed by PFGE with SmaI-digested chromosomal DNA (8, 15). PFGE was performed on a 1.0% SeaKem Gold agarose gel (Cambrex Bio Science, Rockland, ME) and a CHEF-DRIII apparatus (Bio-Rad Laboratories, Hercules, CA). The PFGE standard strain Salmonella enterica serovar Braenderup H9812 was obtained from the PulseNet program, Enteric Diseases Laboratory Branch, Centers for Disease Control and Prevention, through the National Institute of Infectious Diseases (Japan).The two representative VanA-type VRE strains were precultured overnight in Trypticase soy broth (Becton, Dickinson and Company) at 36°C. Each culture medium was diluted in BPW to obtain 2 log CFU/ml, and 100 μl of the culture dilution was inoculated into 40 ml of brain heart infusion (BHI) broth (Becton, Dickinson and Company) in a 200-ml Erlenmeyer flask. The flask was incubated at 35°C or 42°C with constant shaking (160 rpm) for 24 h, and the optical density at 600 nm (OD600) of the cultures was monitored every 30 min using an OD-Monitor A & S (Taitec Co., Ltd., Saitama, Japan). Each strain was examined in triplicate. All VanA-type isolates in this study were inoculated onto the poultry samples to compare the effect of BPW incubation at 42°C on the isolation of VanA-type VRE with the effect of incubation at 35°C. The strains were precultured at 36°C for 24 h under static conditions and then serially diluted in BPW to obtain a cell concentration of approximately 3 log CFU/ml. One hundred microliters of each bacterial cell suspension was individually inoculated onto two 25-g portions of each poultry sample in sterile plastic bags. Each inoculated poultry sample was added to 225 ml of BPW and homogenized for 1 min. One sample was incubated at 35°C for 24 h, and the second was incubated at 42°C for 24 h. Colony-sweep PCR analyses with Enterococcosel agar containing 4 μg/ml vancomycin for the presence of the vanA gene were then performed as described above. The artificially contaminated poultry sample experiment was performed in triplicate. To evaluate the success of cultivation using the different enrichment temperatures, data regarding the identification of the vanA gene from the colony-sweep PCR were analyzed using Fisher''s exact test. To examine the effect of enrichment temperature on inhibition of suspected enterococcal strains by background microflora, growth kinetics were compared among these microorganisms in three different poultry samples. For each sample, two 25-g portions were placed into sterile plastic bags to which 225 ml of BPW was added. After homogenization, the samples were incubated at 35°C and 42°C. Each culture was sampled every 3 h for the initial 15 h and after incubation for 24 h. The number of background microorganisms was determined by plating 1-ml volumes of samples serially diluted in PBS onto standard agar (Eiken) plates. Colonies were counted after incubation at 36°C for 24 h. The growth of enterococcal strains was monitored by plating 100-μl volumes of appropriate dilutions onto EF agar plates (Nissui Pharmaceutical Co., Ltd., Tokyo, Japan), and characteristic enterococcal colonies were counted after incubation at 36°C for 48 h.Colony-sweep PCR analyses revealed that 8 of 171 domestic poultry samples were positive for the vanA gene. Isolation of VRE strains from the PCR-positive sweeps resulted in the identification of six vanA-genotype E. faecalis strains and two vanA-genotype E. faecium strains. Of the eight vanA-genotype VRE-positive samples, five were identified from only the 42°C BPW enrichment culture, while only a single isolate was uniquely identified from incubation at 35°C (Table (Table1).1). The number of VanA-type VRE was beneath the detection limit (i.e., less than 10 CFU/g) in each of the five VRE-positive poultry samples obtained in October 2008 and April 2009, suggesting that VanA-type VRE levels might be markedly low in domestic chickens in Japan. All vanA-genotype E. faecalis and E. faecium strains had vancomycin and teicoplanin MICs of >256 μg/ml and ≥32 μg/ml, respectively, confirming that these strains corresponded to the VanA phenotype (Table (Table1).1). Although the PFGE patterns of two VanA-type E. faecium strains were distinct (Fig. (Fig.1,1, lanes 7 and 8), the patterns of all E. faecalis isolates were almost indistinguishable (lanes 1 to 6). Given that the six VanA-type E. faecalis strains were isolated from distinct poultry samples that were obtained from 5 different retail shops, these results indicated that a specific E. faecalis clone harboring the vanA gene may be widely disseminated in Japan. The dominant species of VanA-type VRE isolated from poultry sources has been shown to differ depending on region. In New Zealand, greater than 80% of VRE isolates from broiler fecal samples were identified as vanA-genotype E. faecalis strains with very similar PFGE patterns, indicating that once established, a single, dominant clonal strain is capable of widespread contamination, and supporting the idea of clonal VRE dissemination rather than horizontal transfer of vancomycin resistance elements (13). Our findings also suggest that the epidemiological features of VanA-type VRE strains associated with poultry production environments in Japan may be similar to those observed in New Zealand. To our knowledge, this is the first report of the isolation of VanA-type VRE from retail domestic poultry products in Japan.Open in a separate windowFIG. 1.DNA fingerprinting of VanA-type VRE (E. faecalis and E. faecium strains) isolated from retail domestic poultry samples. Isolated genomic DNA of each isolate was digested with SmaI and subjected to PFGE. Lane 1, E. faecalis Ha8; lane 2, E. faecalis Ha16; lane 3, E. faecalis Ha36; lane 4, E. faecalis Ha55; lane 5, E. faecalis Ha58; lane 6, E. faecalis Ha61; lane 7, E. faecium Ha20; lane 8, E. faecium Ha26; lanes M, XbaI-digested PFGE patterns of DNA size standard Salmonella enterica serovar Braenderup H9812. Numbers to the left of the gel image indicate the molecular size in kilobase pairs.

TABLE 1.

Characteristics of vancomycin-resistant enterococcal strains isolated from domestic poultry samples in this study
Strain no.Enterococcus speciesVancomycin resistance geneMIC (μg/ml)
Enrichment tempa (°C)Isolation date
VancomycinTeicoplanin
Ha8E. faecalisvanA>256>25642October 2008
Ha16E. faecalisvanA>256>25642October 2008
Ha36E. faecalisvanA>256>25642March 2009
Ha55E. faecalisvanA>25612842April 2009
Ha58E. faecalisvanA>2563235April 2009
Ha61E. faecalisvanA>25625635/42April 2009
Ha20E. faeciumvanA>2563242November 2008
Ha26E. faeciumvanA>2563235/42December 2008
Open in a separate windowaAbbreviations: 35, a strain isolated only from BPW enrichment culture incubated at 35°C; 42, a strain isolated only from BPW enrichment culture incubated at 42°C; 35/42, a strain isolated from BPW enrichment culture incubated at 35°C and 42°C.The growth kinetics of E. faecalis Ha8 and E. faecium Ha20 cultured under pure culture conditions in BHI broth incubated at 35°C or 42°C are shown in Fig. Fig.2.2. For both isolates, initiation of the exponential growth phase was shortened by approximately 2 to 3 h at the 42°C incubation temperature. In poultry samples artificially contaminated with VanA-type VRE isolates, a significant difference in the detection of the vanA gene using colony-sweep PCR with BPW was observed between incubation at 35°C and incubation at 42°C, with detection rates of 11/24 and 19/24, respectively (P < 0.05) (Table (Table2),2), indicating that the use of the 42°C enrichment step increased the recovery rate of the VanA-type VRE strains, which was significantly higher than that at 35°C. Figure Figure33 illustrates the growth kinetics of background microflora and the suspected enterococcal strains isolated from three independent retail poultry products in BPW enrichment cultures incubated at 35°C and 42°C. The initial levels of background microflora and suspected enterococcal strains in samples 1, 2, and 3 were 4.3 and 1.0 log CFU/ml, 3.2 and 0.7 log CFU/ml, and 5.1 and 2.9 log CFU/ml, respectively. After the BPW enrichment step, the maximum densities of background microflora in cultures incubated at 35°C were similar to those at 42°C. In contrast, the maximum levels of suspected enterococcal strains in BPW enrichment cultures incubated at 35°C and 42°C in samples 1, 2, and 3 were 6.0 and 7.5 log CFU/ml, 5.6 and 7.9 log CFU/ml, and 7.3 and 8.0 log CFU/ml, respectively (Fig. (Fig.3).3). These results suggested that the impact of elevated incubation temperature on the enrichment of VRE may be greater in poultry with low levels of enterococcal contamination, such as the VanA-positive samples obtained in this study. We conclude that BPW incubation at 42°C is superior to incubation at 35°C for the enrichment and detection of VanA-type VRE in poultry samples.Open in a separate windowFIG. 2.Growth curves of E. faecalis Ha8 (A) and E. faecium Ha20 (B) cultured in BHI broth incubated at either 35°C (▪) or 42°C (○). Results obtained from three independent experiments are presented as means ± standard deviations.Open in a separate windowFIG. 3.Growth curves of background microflora and suspected enterococcal strains in three different poultry samples cultured in BPW incubated at either 35°C or 42°C. ▴, background microflora; □, suspected enterococcal strains. Solid lines show the data for incubation at 35°C, and dotted lines show those for incubation at 42°C.

TABLE 2.

Comparison of vanA gene detection by colony-sweep PCR after enrichment incubation with BPW at 35°C and 42°C
Inoculum strainInoculum dose (CFU/25 g)No. of positive samples/no. of samples tested with incubation temp:
35°C42°C
E. faecalis Ha8300-3102/33/3
E. faecalis Ha16130-6002/33/3
E. faecalis Ha36100-2901/32/3
E. faecalis Ha55630-1,4002/33/3
E. faecalis Ha58260-6602/33/3
E. faecalis Ha61780-1,7002/33/3
E. faecium Ha20260-4500/32/3
E. faecium Ha26350-5100/30/3
Total11/2419/24
Open in a separate window  相似文献   

15.
Detection and Identification of tdh- and trh-Positive Vibrio parahaemolyticus Strains from Four Species of Cultured Bivalve Molluscs on the Spanish Mediterranean Coast     
Ana Roque  Carmen Lopez-Joven  Beatriz Lacuesta  Laurence Elandaloussi  Sariqa Wagley  M. Dolores Furones  Imanol Ruiz-Zarzuela  Ignacio de Blas  Rachel Rangdale  Bruno Gomez-Gil 《Applied and environmental microbiology》2009,75(23):7574-7577
Presented here is the first report describing the detection of potentially diarrheal Vibrio parahaemolyticus strains isolated from cultured bivalves on the Mediterranean coast, providing data on the presence of both tdh- and trh-positive isolates. Potentially diarrheal V. parahaemolyticus strains were isolated from four species of bivalves collected from both bays of the Ebro delta, Spain.Gastroenteritis caused by Vibrio parahaemolyticus has been reported worldwide, though only sporadic cases have been reported in Europe (7, 14). The bacterium can be naturally present in seafood, but pathogenic isolates capable of inducing gastroenteritis in humans are rare in environmental samples (2 to 3%) (15) and are often not detected (10, 19, 20).The virulence of V. parahaemolyticus is based on the presence of a thermostable direct hemolysin (tdh) and/or the thermostable direct hemolysin-related gene (trh) (1, 5). Both are associated with gastrointestinal illnesses (2, 9).Spain is not only the second-largest producer in the world of live bivalve molluscs but also one of the largest consumers of bivalve molluscs, and Catalonia is the second-most important bivalve producer of the Spanish Autonomous Regions. Currently, the cultivation of bivalves in this area is concentrated in the delta region of the Ebro River. The risk of potentially pathogenic Vibrio spp. in products placed on the market is not assessed by existing legislative indices of food safety in the European Union, which emphasizes the need for a better knowledge of the prevalence of diarrheal vibrios in seafood products. The aim of this study was to investigate the distribution and pathogenic potential of V. parahaemolyticus in bivalve species exploited in the bays of the Ebro delta.Thirty animals of each species of Mytilus galloprovincialis, Crassostrea gigas, Ruditapes decussatus, and Ruditapes philippinarum were collected. They were sampled from six sites of the culture area, three in each bay of the Ebro River delta, at the beginning (40°37′112"N, 0°37′092"E [Alfacs]; 40°46′723"N, 0°43′943"E [Fangar]), middle (40°37′125"N, 0°38′570"E [Alfacs]; 40°46′666"N, 0°45′855"E [Fangar]), and end (40°37′309"N, 0°39′934"E [Alfacs]; 40°46′338"N, 0°44′941"E [Fangar]) of the culture polygon. Clams were sampled from only one site per bay as follows: in the Alfacs Bay from a natural bed of R. decussatus (40°37′44"N, 0°38′0"E) and in the Fangar Bay from an aquaculture bed of R. philippinarum (40°47′3"N, 0°43′8"E). In total, 367 samples were analyzed in 2006 (180 oysters, 127 mussels, 30 carpet shell clams, and 30 Manila clams) and 417 samples were analyzed in 2008 (178 oysters, 179 mussels, 30 carpet shell clams, and 30 Manila clams).All animals were individually processed and homogenized, and 1 ml of the homogenate was inoculated into 9 ml of alkaline peptone water (Scharlau, Spain). Following a 6-h incubation at 37°C, one loopful of the contents of each tube of alkaline peptone water was streaked onto CHROMagar vibrio plates (CHROMagar, France) and incubated for 18 h at 37°C. Mauve-purple colonies were purified, and each purified isolate was cryopreserved at −80°C (135 isolates in 2006 and 96 in 2008). From the initial homogenate portion, 100 μl was inoculated onto marine agar (Scharlau, Spain) and onto thiosulfate citrate-bile salts-sucrose agar (Scharlau, Spain) for total heterotrophic marine bacteria counts and total vibrio counts, respectively (Table (Table11).

TABLE 1.

Vibrio parahaemolyticus isolates, serotypes, and origins and total number of vibrios/heterotrophic bacteria contained in the bivalvea
IsolateDate of collectionOrganism and site of originTemp (°C)Salinity (‰)Gene(s)SerotypeBacterial count using indicated medium (CFU ml−1)
TCBS agarMarine agar
I7458 August 2006Mg-F24.537tdhND1.5 × 1041.2 × 104
I79314 August 2006Cg-A2535tdhND9.2 × 1028.5 × 103
I80514 August 2006Cg-A2535tdhO2:KUT7.2 × 1029 × 103
I80614 August 2006Cg-A2535tdh and trhO3:K331.9 × 1034.6 × 103
I80914 August 2006Cg-A2535tdhO2:K288 × 1047.3 × 102
I6784 July 2006Rd-A28.636tdhO2:K283.1 × 1052.5 × 105
I6284 July 2006Rd-A28.636tdhO4:KUT2.9 × 1048.4 × 104
I7758 August 2006Cg-A24.537tdhND4.21 × 1031.1 × 104
I6914 July 2006Rd-A28.636trhO1:K322.2 × 1052.6 × 105
I71227 July 2006Mg-A29.435.5trhO1:KUT8.6 × 1038.4 × 103
I7658 August 2006Cg-F24.537trhO4:K341 × 104Uncountable
I98022 July 2008Cg-A26.733.5tdhO1:K322.7 × 1041.3 × 104
I98122 July 2008Cg-A26.733.5trhO1:KUT1 × 1042.2 × 104
I99322 July 2008Cg-A26.733.5tdhO5:K173 × 1031.1 × 104
I99429 July 2008Mg-A27.737trhO3:KUT3.4 × 1037 × 103
I10315 August 2008Cg-F27.737tdhO5:KUT5.5 × 1043.3 × 104
I10345 August 2008Cg-F27.737tdhO3:KUT8.7 × 1044 × 104
I10405 August 2008Cg-F27.737tdhO3:KUT1.6 × 1043.2 × 104
I10425 August 2008Cg-F27.737tdh and trhND2.8 ×1043 × 104
I10505 August 2008Cg-F27.737tdhO1:KUT4.7 × 1047.3 × 104
I106320 August 2008Mg-F25.936tdhO3:KUT7.9 ×1041.4 × 104
I106520 August 2008Mg-F25.936tdhO2:KUT2.2 × 1031.2 × 104
I106820 August 2008Mg-F25.936tdhO5:KUT2.6 × 1045.2 × 104
I106920 August 2008Mg-F25.936tdhO3:KUT2.4 × 1035.3 × 104
I107320 August 2008Mg-F25.936tdhO5:KUT2.3 × 1037.5 × 103
I107420 August 2008Mg-F25.936tdhO3:KUT7.6 × 1046.9 × 104
I107720 August 2008Mg-F25.936tdhO4:KUT1.7 × 1031.6 × 103
I107920 August 2008Mg-F25.936trhO3:KUT2.5 × 1031.1 × 104
I109220 August 2008Mg-F25.936tdhND1.7 × 1031.6 × 103
I113025 August 2008Rd-A26.435tdhND1.7 × 1043.8 × 104
I114325 August 2008Rd-A26.435tdhND1.1 × 1041.9 × 104
I116525 August 2008Rd-A26.435trhO2:KUT4.4 × 1046.8 × 104
I113325 August 2008Rp-F25.536.5tdhND3.4 × 1044 × 104
I113425 August 2008Rp-F25.536.5tdhND3.9 × 1045.8 × 104
I115825 August 2008Rp-F25.536.5trhO4:KUT6.6 × 1044.7 × 104
I116125 August 2008Rp-F25.536.5trhO3:KUT2.2 × 1046.6 × 104
Open in a separate windowaMg, Mytilus galloprovincialis; Cg, Crassostrea gigas; Rd, Ruditapes decussatus; Rp, R. phillipinarum; A, Alfacs; F, Fangar; ND, not determined; TCBS, thiosulfate citrate-bile salts-sucrose.Total DNA was extracted from each purified isolate using the Wizard genomic DNA purification kit (Promega), following the instructions of the manufacturer. A one-step PCR analysis was performed to identify/confirm which isolates were tl positive (species marker for V. parahaemolyticus). Further detection of the tdh or trh gene was carried out on all positive tl strains. All PCR analyses were carried out using the primers described by Bej et al. (2) with the following amplification conditions on the thermocycler (Eppendorf Mastercycler Personal): an initial denaturation at 95°C for 8 min, followed by 40 cycles of a 1-min denaturation at 94°C, annealing at 55°C for 1 min, elongation at 72° for 1 min, and a final extension of 10 min at 72°C. Positive and negative controls were included in all reaction mixtures: two positive controls, tl and tdh CAIM 1400 and trh CAIM 1772 (Collection of Aquatic Important Microorganisms [http://www.ciad.mx/caim/CAIM.html]), and negative control DNA-free molecular grade water (Sigma-Aldrich, Spain). Expected amplicons were visualized in 2% agarose gels stained with ethidium bromide.Fifty-eight isolates contained the gene tl in 2006 and 96 in 2008, which confirmed their identity as V. parahaemolyticus. In 2006, the distribution of the 58 isolates was as follows: 7 from 127 mussels, 34 from 180 oysters, and 17 from 30 R. decussatus clams. No tl-positive isolates were found in R. philippinarum. PCR analysis of the tl-positive isolates for the presence of the tdh or trh gene indicated that eight isolates contained the tdh gene and four contained the trh gene. In 2008, the source of the confirmed V. parahaemolyticus isolates was as follows: 31 from 88 oysters, 44 from 89 mussels, 9 from 30 R. decussatus clams, and 12 from 30 R. philippinarum clams. Of these, 17 were found to contain the tdh gene and 7 contained the trh gene. Two isolates (I806 and I1042) contained both toxigenic genes, tdh and trh.Putative tdh- and trh-positive PCR products were purified using the QIAquick PCR purification kit (Qiagen) following the manufacturer''s instructions and were sequenced bidirectionally by Macrogen Inc. Sequences were aligned using BioEdit (8) and analyzed using BLAST (National Center for Biotechnology Information). None of the toxigenic isolates was found positive by PCR analysis for the presence of open reading frame 8 of the phage 237 (16), a marker for the pandemic strain O3:K6.The isolates were fingerprinted by repetitive extragenic palindromic PCR (rep-PCR) as described previously (3), and the resulting electrophoretic band patterns were analyzed with the GelCompar II software (v4.5; Applied Maths). The similarity matrix was calculated with the Jaccard coefficient with a band position tolerance of 0.8%, and the dendrogram was constructed with the Ward algorithm. A high level of genomic diversity was found among the 32 toxigenic isolates characterized by rep-PCR. Three clonal groups were identified (those having identical rep-PCR band patterns) (Fig. 1a to c).Open in a separate windowFIG. 1.rep-PCR dendrogram of toxigenic isolates of V. parahaemolyticus isolated in the Ebro delta. Letters denote clonal groups of isolates.In vitro antibiotic susceptibility tests were performed using the diffusion disc test following a previously described protocol (18). The antibiotics used were gentamicin (10 μg), oxolinic acid (10 μg), amoxicillin (25 μg), polymyxin B (300 UI), vancomycin (30 μg), trimethoprim sulfamethoxazole (1.25/23.75 μg), nitrofurantoin (300 μg), doxycyclin (30 μg), ceftazidime (30 μg), streptomycin (10 μg), neomycin (30 UI), penicillin (6 μg), flumequine (30 μg), tetracycline (30 μg), ampicillin (10 μg), kanamycin (30 μg), ciprofloxacin (5 μg), and sulfonamide (300 μg). All tests were performed in duplicate. A Student t test for two samples with unequal variance was performed to compare the sensitivity of all 2006 isolates against the sensitivity of 2008 isolates for each antibiotic (Microsoft Office Excel 97-2003). Antibiogram results revealed a lower susceptibility in 2008 than in 2006, indicating a possible shift in overall susceptibility. Results from the t test indicated that significantly lower susceptibility in 2008 was detected (P ≤ 0.05; n = 36) for the following antibiotics: vancomycin, polymyxin B, ampicillin, amoxicillin, gentamicin, neomycin, trimethoprim sulfamethoxazole, nitrofurantoin, doxycyclin, ceftazidime, tetracycline, flumequine, and ciprofloxacin.The serological types for 27 strains were determined by the agglutination method using commercially available V. parahaemolyticus antisera (Denka Seiken Ltd.; Cosmos Biomedical Ltd, United Kingdom) following the manufacturer''s instructions. Potentially toxigenic V. parahaemolyticus isolates collected in 2006 were serologically heterogeneous (8 out of the 11 isolates) (Table (Table1).1). In isolates collected in 2008, results were more homogenous, with seven serotypes found among 19 isolates analyzed. The O3:K6 serotype was not detected in any of the strains analyzed, in agreement with the open reading frame 8 PCR results.The present study is the first to report the detection of potentially diarrheal V. parahaemolyticus strains isolated from cultured bivalves on Spanish Mediterranean coasts, providing data on the presence of both tdh- and trh-positive isolates. V. parahaemolyticus has previously been detected in several European countries (4, 13, 21, 22). A recent study carried out in Spain detected tdh-positive V. parahaemolyticus strains from patients who had consumed fresh oysters in a market in Galicia on the Atlantic coast of Spain (12) and potentially pathogenic V. parahaemolyticus strains have also been reported in France (17). These studies indicate that the risk of infections caused by V. parahaemolyticus in Europe is low compared to that in America or Asia (15). However, this risk could have been underestimated, since V. parahaemolyticus is not included in the current European surveillance programs, such as the European Network for Epidemiological Surveillance and Control of Communicable Diseases.Toxigenic V. parahaemolyticus strains detected in this study were genomically and serologically heterogeneous. The pandemic serotype O3:K6 was not detected, and although attempts to isolate O3:K6 from the environment and from seafood have not always been successful in previous studies reviewed by Nair and coauthors (15), this finding seems to be in agreement with the fact that no outbreak of diarrhea was observed in the area. Interestingly, isolates I806 and I1042 have been found positive for both tdh and trh in PCR tests. The coexistence of tdh and trh genes has already been reported in isolates from Japan, the United States, and Mexico (3, 6, 11, 19, 23). To our knowledge, no occurrence of an environmental isolate positive for both tdh and trh had previously been reported in Europe. All isolates tested were slightly different in their antibiotic resistance profiles. Typically, a high level of resistance could be determined. The detection of tdh- and/or trh-positive V. parahaemolyticus strains for the first time on the Mediterranean coast emphasizes the need to monitor for the presence of potentially diarrheal vibrios and bacterial gastroenteritis, and these data should be taken into consideration to revise the European legislation on the requirements for shellfish harvested for consumption in order to include the surveillance of these pathogens in Europe.  相似文献   

16.
Variation in Adult Plant Phenotypes and Partitioning among Seed and Stem-Borne Roots across Brachypodium distachyon Accessions to Exploit in Breeding Cereals for Well-Watered and Drought Environments     
Vincent Chochois  John P. Vogel  Gregory J. Rebetzke  Michelle Watt 《Plant physiology》2015,168(3):953-967
Seedling roots enable plant establishment. Their small phenotypes are measured routinely. Adult root systems are relevant to yield and efficiency, but phenotyping is challenging. Root length exceeds the volume of most pots. Field studies measure partial adult root systems through coring or use seedling roots as adult surrogates. Here, we phenotyped 79 diverse lines of the small grass model Brachypodium distachyon to adults in 50-cm-long tubes of soil with irrigation; a subset of 16 lines was droughted. Variation was large (total biomass, ×8; total root length [TRL], ×10; and root mass ratio, ×6), repeatable, and attributable to genetic factors (heritabilities ranged from approximately 50% for root growth to 82% for partitioning phenotypes). Lines were dissected into seed-borne tissues (stem and primary seminal axile roots) and stem-borne tissues (tillers and coleoptile and leaf node axile roots) plus branch roots. All lines developed one seminal root that varied, with branch roots, from 31% to 90% of TRL in the well-watered condition. With drought, 100% of TRL was seminal, regardless of line because nodal roots were almost always inhibited in drying topsoil. Irrigation stimulated nodal roots depending on genotype. Shoot size and tillers correlated positively with roots with irrigation, but partitioning depended on genotype and was plastic with drought. Adult root systems of B. distachyon have genetic variation to exploit to increase cereal yields through genes associated with partitioning among roots and their responsiveness to irrigation. Whole-plant phenotypes could enhance gain for droughted environments because root and shoot traits are coselected.Adult plant root systems are relevant to the size and efficiency of seed yield. They supply water and nutrients for the plant to acquire biomass, which is positively correlated to the harvest index (allocation to seed grain), and the stages of flowering and grain development. Modeling in wheat (Triticum aestivum) suggested that an extra 10 mm of water absorbed by such adult root systems during grain filling resulted in an increase of approximately 500 kg grain ha−1 (Manschadi et al., 2006). This was 25% above the average annual yield of wheat in rain-fed environments of Australia. This number was remarkably close to experimental data obtained in the field in Australia (Kirkegaard et al., 2007). Together, these modeling and field experiments have shown that adult root systems are critical for water absorption and grain yield in cereals, such as wheat, emphasizing the importance of characterizing adult root systems to identify phenotypes for productivity improvements.Most root phenotypes, however, have been described for seedling roots. Seedling roots are essential for plant establishment, and hence, the plant’s potential to set seed. For technical reasons, seedlings are more often screened than adult plants because of the ease of handling smaller plants and the high throughput. Seedling-stage phenotyping may also improve overall reproducibility of results because often, growth media are soil free. Seedling soil-free root phenotyping conditions are well suited to dissecting fine and sensitive mechanisms, such as lateral root initiation (Casimiro et al., 2003; Péret et al., 2009a, 2009b). A number of genes underlying root processes have been identified or characterized using seedlings, notably with the dicotyledonous models Arabidopsis (Arabidopsis thaliana; Mouchel et al., 2004; Fitz Gerald et al., 2006; Yokawa et al., 2013) and Medicago truncatula (Laffont et al., 2010) and the cereals maize (Zea mays; Hochholdinger et al., 2001) and rice (Oryza sativa; Inukai et al., 2005; Kitomi et al., 2008).Extrapolation from seedling to adult root systems presents major questions (Hochholdinger and Zimmermann, 2008; Chochois et al., 2012; Rich and Watt, 2013). Are phenotypes in seedling roots present in adult roots given developmental events associated with aging? Is expression of phenotypes correlated in seedling and adult roots if time compounds effects of growth rates and growth conditions on roots? Watt et al. (2013) showed in wheat seedlings that root traits in the laboratory and field correlated positively but that neither correlated with adult root traits in the field. Factors between seedling and adult roots seemed to be differences in developmental stage and the time that growing roots experience the environment.Seedling and adult root differences may be larger in grasses than dicotyledons. Grass root systems have two developmental components: seed-borne (seminal) roots, of which a number emerge at germination and continue to grow and branch throughout the plant life, and stem-borne (nodal or adventitious) roots, which emerge from around the three-leaf stage and continue to emerge, grow, and branch throughout the plant life. Phenotypes and traits of adult root systems of grasses, which include the major cereal crops wheat, rice, and maize, are difficult to predict in seedling screens and ideally identified from adult root systems first (Gamuyao et al., 2012).Phenotyping of adult roots is possible in the field using trenches (Maeght et al., 2013) or coring (Wasson et al., 2014). A portion of the root system is captured with these methods. Alternatively, entire adult root systems can be contained within pots dug into the ground before sowing. These need to be large; field wheat roots, for example, can reach depths greater than 1.5 m depending on genotype and environment. This method prevents root-root interactions that occur under normal field sowing of a plant canopy and is also a compromise.A solution to the problem of phenotyping adult cereal root systems is a model for monocotyledon grasses: Brachypodium distachyon. B. distachyon is a small-stature grass with a small genome that is fully sequenced (Vogel et al., 2010). It has molecular tools equivalent to those available in Arabidopsis (Draper et al., 2001; Brkljacic et al., 2011; Mur et al., 2011). The root system of B. distachyon reference line Bd21 is more similar to wheat than other model and crop grasses (Watt et al., 2009). It has a seed-borne primary seminal root (PSR) that emerges from the embryo at seed germination and multiple stem-borne coleoptile node axile roots (CNRs) and leaf node axile roots (LNRs), also known as crown roots or adventitious roots, that emerge at about three leaves through to grain development. Branch roots emerge from all root types. There are no known anatomical differences between root types of wheat and B. distachyon (Watt et al., 2009). In a recent study, we report postflowering root growth in B. distachyon line Bd21-3, showing that this model can be used to answer questions relevant to the adult root systems of grasses (Chochois et al., 2012).In this study, we used B. distachyon to identify adult plant phenotypes related to the partitioning among seed-borne and stem-borne shoots and roots for the genetic improvement of well-watered and droughted cereals (Fig. 1; Krassovsky, 1926; Navara et al., 1994), nitrogen, phosphorus (Tennant, 1976; Brady et al., 1995), oxygen (Wiengweera and Greenway, 2004), soil hardness (Acuna et al., 2007), and microorganisms (Sivasithamparam et al., 1978). Of note is the study by Krassovsky (1926), which was the first, to our knowledge, to show differences in function related to water. Krassovsky (1926) showed that seminal roots of wheat absorbed almost 2 times the water as nodal roots per unit dry weight but that nodal roots absorbed a more diluted nutrient solution than seminal roots. Krassovsky (1926) also showed by removing seminal or nodal roots as they emerged that “seminal roots serve the main stem, while nodal roots serve the tillers” (Krassovsky, 1926). Volkmar (1997) showed, more recently, in wheat that nodal and seminal roots may sense and respond to drought differently. In millet (Pennisetum glaucum) and sorghum (Sorghum bicolor), Rostamza et al. (2013) found that millet was able to grow nodal roots in a dryer soil than sorghum, possibly because of shoot and root vigor.Open in a separate windowFigure 1.B. distachyon plant scanned at the fourth leaf stage, with the root and shoot phenotypes studied indicated. Supplemental Table S1.
PhenotypeAbbreviationUnitRange of Variation
All Experiments (79 Lines and 582 Plants)Experiment 6 (36 Lines)
Whole plant
TDWTDWMilligrams88.6–773.8 (×8.7)285.6–438 (×1.5)
Shoot
SDWSDWMilligrams56.4–442.5 (×7.8)78.2–442.5 (×5.7)
 No. of tillersTillerNCount2.8–20.3 (×7.4)10–20.3 (×2)
Total root system
TRLTRLCentimeters1,050–10,770 (×10.3)2,090–5,140 (×2.5)
RDWRDWMilligrams28.9–312.17 (×10.8)62.2–179.1 (×2.9)
RootpcRootpcPercentage (of TDW)20.5–60.6 (×3)20.5–44.3 (×2.2)
R/SR/SUnitless ratio0.26–1.54 (×6)0.26–0.80 (×3.1)
PSRs
 Length (including branch roots)PSRLCentimeters549.1–4,024.6 (×7.3)716–2,984 (×4.2)
PSRpcPSRpcPercentage (of TRL)14.9–94.1 (×6.3)31.3–72.3 (×2.3)
 No. of axile rootsPSRcountCount11
 Length of axile rootPSRsumCentimeters17.45–52 (×3)17.45–30.3 (×1.7)
 Branch rootsPSRbranchCentimeters · (centimeters of axile root)−119.9–109.3 (×5.5)29.3–104.3 (×3.6)
CNRs
 Length (including branch roots)CNRLCentimeters0–3,856.70–2,266.5
CNRpcCNRpcPercentage (of TRL)0–57.10–49.8
 No. of axile rootsCNRcountCount0–20–2
 Cumulated length of axile rootsCNRsumCentimeters0–113.90–47.87
 Branch rootsCNRbranchCentimeters · (centimeters of axile root)−10–77.80–77.8
LNRs
 Length (including branch roots)LNRLCentimeters99.5–5,806.5 (×58.5)216.1–2,532.4 (×11.7)
LNRpcLNRpcPercentage (of TRL)4.2–72.7 (×17.5)6–64.8 (×10.9)
LNRcountLNRcountCount2–22.2 (×11.1)3.3–15.3 (×4.6)
LNRsumLNRsumCentimeters25.9–485.548–232 (×4.8)
 Branch rootsLNRbranchCentimeters · (centimeters of axile root)−12.1–25.4 (×12.1)3.2–15.9 (×5)
Open in a separate windowThe third reason for dissecting the different root types in this study was that they seem to have independent genetic regulation through major genes. Genes affecting specifically nodal root growth have been identified in maize (Hetz et al., 1996; Hochholdinger and Feix, 1998) and rice (Inukai et al., 2001, 2005; Liu et al., 2005, 2009; Zhao et al., 2009; Coudert et al., 2010; Gamuyao et al., 2012). Here, we also dissect branch (lateral) development on the seminal or nodal roots. Genes specific to branch roots have been identified in Arabidopsis (Casimiro et al., 2003; Péret et al., 2009a), rice (Hao and Ichii, 1999; Wang et al., 2006; Zheng et al., 2013), and maize (Hochholdinger and Feix, 1998; Hochholdinger et al., 2001; Woll et al., 2005).This study explored the hypothesis that adult root systems of B. distachyon contain genotypic variation that can be exploited through phenotyping and genotyping to increase cereal yields. A selection of 79 wild lines of B. distachyon from various parts of the Middle East (Fig. 2 shows the geographic origins of the lines) was phenotyped. They were selected for maximum genotypic diversity from 187 diploid lines analyzed with 43 simple sequence repeat markers (Vogel et al., 2009). We phenotyped shoots and mature root systems concurrently because B. distachyon is small enough to complete its life cycle in relatively small pots of soil with minimal influence of pot size compared with crops, such as wheat. We further phenotyped a subset of this population under irrigation (well watered) and drought to assess genotype response to water supply. By conducting whole-plant studies, we aimed to identify phenotypes that described partitioning among shoot and root components and within seed-borne and stem-borne roots. Phenotypes that have the potential to be beneficial to shoot and root components may speed up genetic gain in future.Open in a separate windowFigure 2.B. distachyon lines phenotyped in this study and their geographical origin. Capital letters in parentheses indicate the country of origin: Turkey (T), Spain (S), and Iraq (I; Vogel et al., 2009). a, Adi3, Adi7, Adi10, Adi12, Adi13, and Adi15; b, Bd21 and Bd21-3 are the reference lines of this study. Bd21 was the first sequenced line (Vogel et al., 2010) and root system (described in detail in Watt et al., 2009), and Bd21-3 is the most easily transformed line (Vogel and Hill, 2008) and parent of a T-DNA mutant population (Bragg et al., 2012); c, Gaz1, Gaz4, and Gaz7; d, Kah1, Kah2, and Kah3. e, Koz1, Koz3, and Koz5; f, Tek1 and Tek6; g, exact GPS coordinates are unknown for lines Men2 (S), Mur2 (S), Bd2.3 (I), Bd3-1 (I), and Abr1 (T).  相似文献   

17.
Bar-Coded Pyrosequencing of 16S rRNA Gene Amplicons Reveals Changes in Ileal Porcine Bacterial Communities Due to High Dietary Zinc Intake     
W. Vahjen  R. Pieper  J. Zentek 《Applied and environmental microbiology》2010,76(19):6689-6691
Feeding high levels of zinc oxide to piglets significantly increased the relative abundance of ileal Weissella spp., Leuconostoc spp., and Streptococcus spp., reduced the occurrence of Sarcina spp. and Neisseria spp., and led to numerical increases of all Gram-negative facultative anaerobic genera. High dietary zinc oxide intake has a major impact on the porcine ileal bacterial composition.Zinc oxide (ZnO) is used as a feed additive for diarrhea prophylaxis in piglets (23). However, the mode of action of ZnO is not fully understood. Besides its effects on the host (10, 30, 31), high dietary zinc levels may affect the diversity of intestinal microbial communities (2, 11, 20). The prevention of postweaning diarrhea in piglets due to high dietary ZnO intake may not be directly related to a reduction of pathogenic E. coli (8) but, rather, to the diversity of the coliform community (15). Studies on the impact of high ZnO levels on the porcine ileal bacterial community are scarce but nevertheless important, as bacterial diarrhea is initiated in the small intestine (9, 17). The small intestine is a very complex habitat with many different factors shaping the bacterial community. Studies on the ecophysiology (22) and maturation of the porcine ileal microbiota (13, 27) indicate a drastic impact directly after weaning and a gradual decline of modifications during the following 2 weeks. Thus, the time point for analysis chosen in this study (14 days postweaning) does reflect a more stable period of the ileal porcine microbiota. In this study, we used bar-coded pyrosequencing of 16S rRNA genes to gain further insight into the mode of action of pharmacological levels of ZnO in the gastrointestinal tract of young pigs.Total DNA was extracted from the ileal digesta of 40- to 42-day-old piglets using a commercial kit (Qiagen stool kit; Qiagen, Hilden, Germany) and PCR amplified with unique bar-coded primer sets targeting the V1-to-V3 and the V6-to-V8 hypervariable regions (see the supplemental material for detailed methods). The rationale behind this approach was derived from the fact that no single “universal” primer pair can completely cover a complex bacterial habitat (4, 24, 32, 33). Furthermore, these studies also show that in silico information on the coverage of selected primer sets diverges from empirical results, and hence, two hypervariable regions were chosen in this study to maximize the detection of phylogenetically diverse bacterial groups.Equimolar dilutions of all samples were combined into one master sample. Pyrosequencing was performed by Agowa (Berlin, Germany) on a Roche genome sequencer FLX system using a Titanium series PicoTiterPlate. The resulting data files were uploaded to the MG-RAST server (http://metagenomics.nmpdr.org/) (19) and processed with its SEED software tool using the RDP database (5) as the reference database. After automated sequence analysis, all sequences with less than five identical reads per sample were deleted in order to increase the confidence of sequence reads and reduce bias from possible sequencing errors (12, 16). Thus, 0.43% of all sequences were not considered (1,882 of 433,302 sequences). These sequences were assigned to a total of 238 genera, of which most only occurred in a few samples (see the supplemental material). Furthermore, all unclassified sequences were removed (8.7%; 41,467 of 474,769 sequences). Due to the use of the RDP reference database, the SEED software incorrectly assigned the majority of unclassified sequences as unclassified Deferribacterales (83%; 34,393 sequences), which were actually identified as 16S soybean or wheat chloroplasts by BLAST or as cyanobacterial chloroplasts by the RDP II seqmatch tool.The pyrosequencing results for the two primer combinations were merged by taking only sequences from the primer combination that yielded the higher number of reads for a specific sequence assignment in a sample. The remaining reads were used to calculate the relative contribution of assigned sequences to total sequence reads in a sample.The Firmicutes phylum dominated the small intestinal bacterial communities in both the control group and the group with high dietary ZnO intake, with 98.3% and 97.0% of total sequence reads, respectively. No significant influence of high dietary ZnO intake was found for the main phyla Proteobacteria (0.92% versus 1.84%), Actinobacteria (0.61% versus 0.75%), Bacteroidetes (0.15% versus 0.17%), and Fusobacteria (0.09% versus 0.12%).On the order level, a total of 20 bacterial orders were detected (data not shown). Lactobacillales dominated bacterial communities in the control and high-dietary-ZnO-intake groups, with 83.37% and 93.24% of total reads. Lactic acid bacteria are well known to dominate the bacterial community in the ileum of piglets (11, 22). No significant difference between the control group and the group with high dietary ZnO intake was observed on the order level, although high dietary ZnO intake led to a strong numerical decrease for Clostridiales (14.4 ± 24.0% [mean ± standard deviation] versus 2.8 ± 1.7%), as well as to numerical increases for Pseudomonadales (0.3 ± 0.3% versus 0.6 ± 0.6%) and Enterobacteriales (0.2 ± 0.2% versus 0.5 ± 0.6%).On the genus level, a total of 103 genera were detected. Table Table11 summarizes the main 31 genera which exceeded 0.05% of total reads (see the supplemental material for a complete list). Lactobacilli clearly dominated the bacterial communities in both trial groups, but they also were numerically lower due to high dietary ZnO intake.

TABLE 1.

Bacterial genera in the ileum of piglets fed diets supplemented with 200 or 3,000 ppm ZnO
GenusProportion (% ± SD) of ileal microbiota in groupa receiving:
200 ppm ZnO3,000 ppm ZnO
Lactobacillus59.3 ± 30.640.7 ± 19.1
Weissella11.6 ± 7.8 A24.1 ± 8.3 B
Sarcina11.4 ± 20.5 A0.84 ± 1.2 B
Leuconostoc4.7 ± 3.2 A9.4 ± 3.1 B
Streptococcus1.8 ± 1.6 A5.7 ± 5.1 B
Lactococcus1.6 ± 1.52.6 ± 3.1
Veillonella0.57 ± 0.630.34 ± 0.30
Gemella0.34 ± 0.67 A0.45 ± 0.25 B
Acinetobacter0.25 ± 0.210.44 ± 0.50
Clostridium0.25 ± 0.400.22 ± 0.21
Enterococcus0.19 ± 0.150.26 ± 0.24
Acidovorax0.14 ± 0.040.16 ± 0.19
Arcobacter0.14 ± 0.150.16 ± 0.17
Neisseria0.14b0.03 ± 0.01
Enterobacter0.13 ± 0.090.29 ± 0.34
Lachnospira0.12 ± 0.130.13 ± 0.03
Peptostreptococcus0.11 ± 0.100.07 ± 0.09
Chryseobacterium0.10 ± 0.070.15 ± 0.16
Actinomyces0.09 ± 0.040.15 ± 0.16
Anaerobacter0.07 ± 0.080.02 ± 0.01
Aerococcus0.07 ± 0.040.07 ± 0.04
Dorea0.07b0.05 ± 0.05
Fusobacterium0.06 ± 0.090.08 ± 0.11
Microbacterium0.06 ± 0.010.07 ± 0.04
Carnobacterium0.06 ± 0.020.08 ± 0.13
Granulicatella0.06 ± 0.020.09 ± 0.10
Staphylococcus0.06 ± 0.040.05 ± 0.02
Facklamia0.05 ± 0.060.03 ± 0.01
Comamonas0.05 ± 0.030.04 ± 0.02
Citrobacter0.05 ± 0.020.07 ± 0.08
Erysipelothrix0.05 ± 0.010.22 ± 0.40
Open in a separate windowan = 6 piglets per trial group. A,B, results are significantly different by Kruskal-Wallis test.bSingle sample.Significant changes due to high dietary ZnO intake were observed for other lactic acid bacteria, including Weissella spp., Leuconostoc spp., and Streptococcus spp. A significant and strong decrease was observed for Sarcina spp., which is a genus of acid-tolerant strictly anaerobic species found in the intestinal tract of piglets and other mammals (6, 28, 29). This genus thus appeared to be very sensitive to modifications induced by high dietary ZnO intake.An interesting result was observed for Gram-negative Proteobacteria, (i.e., enterobacteria and relatives). Although not statistically significant, virtually all detected proteobacteria increased numerically due to high dietary ZnO intake (Enterobacter spp., Microbacterium spp., Citrobacter spp., Neisseria spp., and Acinetobacter spp.). Apparently, enterobacteria gained colonization potential by high dietary ZnO intake. This is in good agreement with the results of studies by Hojberg et al. (11), Amezcua et al. (1), and Castillo et al. (3). Therefore, the frequently observed diarrhea-reducing effect of zinc oxide may not be directly related to a reduction of pathogenic E. coli strains. Considering a possible antagonistic activity of lactobacilli against enterobacteria (25), it can be speculated that a numerical decrease of dominant lactobacilli may lead to increased colonization with Gram-negative enterobacteria. On the other hand, specific plasmid-borne genes for resistance against heavy metals have been reported for both Gram-positive and Gram-negative bacteria present in the intestine (21, 26), and an increased resistance against Zn ions may exist for Gram-negative enterobacteria. Zinc oxide is an amphoteric molecule and shows a high solubility at acid pH. The low pH in the stomach of piglets (pH 3.5 to 4.5) transforms a considerable amount of insoluble ZnO into zinc ions (54 to 84% free Zn2+ at 150 ppm and 24 ppm ZnO, respectively) (7), and thus, high concentrations of toxic zinc ions exist in the stomach. The stomach of piglets harbors large numbers of lactic acid bacteria, especially lactobacilli. Zn ions may thus lead to a modification of the lactic acid bacterial community in the stomach, and the changes observed in the ileum could have been created in the stomach. A reduction of dominant lactobacilli may thus point to an increased adaptation potential of Gram-negative facultative anaerobes and a generally increased bacterial diversity.Additionally, the direct effects of dietary ZnO on intestinal tissues include altered expression of genes responsible for glutathione metabolism and apoptosis (30), enhanced gastric ghrelin secretion, which increases feed intake (31), and increased production of digestive enzymes (10). An analysis of the intestinal morphology was beyond the scope of this study, but although ZnO concentrations are markedly increased in intestinal tissue, the influence of ZnO on morphology is apparently not always observed (10, 14, 18). Consequently, any changes in epithelial cell turnover, feed intake, or digestive capacity may influence the composition of bacterial communities in the small intestine.In conclusion, this study has shown that high dietary zinc oxide has a major impact on ileal bacterial communities in piglets. Future studies on the impact of zinc oxide in pigs should include a detailed analysis of host responses in order to identify the cause for the observed modifications of intestinal bacterial communities.  相似文献   

18.
Construction and Characterization of Three Lactate Dehydrogenase-Negative Enterococcus faecalis V583 Mutants     
Maria J?nsson  Zhian Saleihan  Ingolf F. Nes  Helge Holo 《Applied and environmental microbiology》2009,75(14):4901-4903
  相似文献   

19.
Use of a Mixture of Surrogates for Infectious Bioagents in a Standard Approach to Assessing Disinfection of Environmental Surfaces     
Safaa Sabbah  Susan Springthorpe  Syed A. Sattar 《Applied and environmental microbiology》2010,76(17):6020-6022
We used a mixture of surrogates (Acinetobacter baumannii, Mycobacterium terrae, hepatitis A virus, and spores of Geobacillus stearothermophilus) for bioagents in a standardized approach to test environmental surface disinfectants. Each carrier containing 10 μl of mixture received 50 μl of a test chemical or saline at 22 ± 2°C. Disinfectant efficacy criteria were ≥6 log10 reduction for the bacteria and the spores and ≥3 log10 reduction for the virus. Peracetic acid (1,000 ppm) was effective in 5 min against the two bacteria and the spores but not against the virus. Chlorine dioxide (CD; 500 and 1,000 ppm) and domestic bleach (DB; 2,500, 3,500, and 5,000 ppm) were effective in 5 min, except for sporicidal activity, which needed 20 min of contact with either 1,000 ppm of CD or the two higher concentrations of DB.Disinfectant testing with a single type of organism does not represent field conditions, where bioagents or other pathogens may be mixed with other contaminants. Such an approach also cannot predict the true spectrum of microbicidal activity of a given chemical, while the identity of the target pathogen(s) is often unknown. We used a mixture of Acinetobacter baumannii, Mycobacterium terrae (15), hepatitis A virus (HAV) (4), and the spores of Geobacillus stearothermophilus as surrogates for infectious bioagents, with an added soil load on disks (1 cm in diameter; 0.75 mm thick) of brushed stainless steel (AISI no. 430; Muzeen & Blythe, Winnipeg, MB, Canada), to better simulate environmental surface disinfection (1, 11). Table Table11 gives details on the microbial strains, media used for their culture and recovery, and methods for preparing working stocks. The quantitative carrier test (QCT) method, ASTM standard E-2197 (1), was used to test the organisms singly and in a mixture. Each 200 μl of the inoculum contained 34 μl each of the four organisms, 40 μl of bovine mucin, 14 μl of yeast extract, and 10 μl of bovine serum albumin stocks.

TABLE 1.

Organisms in the mixture and their growth/recovery media and titers
Organism (ATCC no.)Growth/recovery medium or host cell lineProcedure for culture and prepn of stockViability titer in stock
Mycobacterium terrae pBEN genetically modified in-house (ATCC 15755)Middlebrook 7H11 agar, OADC,a and kanamycin (10 μg/ml); incubation 20 days at 36 ± 1°C7H9 broth with ADCb and glycerol; cells washed and resuspended in deionized water (8 ml) in a Bijoux bottle (Wheaton, Millville, NJ) with glass beads (Sigma-Aldrich; 3 mm in diam; catalog no. Z143928) and stored at 4°C3.7 × 109 CFU/ml
Geobacillus stearothermophilus (ATCC 12980)Trypticase soy agar plates incubated at 56°C for 48 hSpores heat shocked at 100°C for 45 min, washed in deionized H2O, and stored at 4°C1.5 × 108 CFU/ml
Acinetobacter baumannii (ATCC 19606)Trypticase soy agar plates incubated at 36 ± 1°C for 24 hInoculated into Trypticase soy broth and incubated for 24 h at 36 ± 1°C, broth centrifuged, and pellet resuspended in deionized H2O and stored at 4°C1.2 × 109 CFU/ml
Hepatitis A virus (ATCC VR-1402)FRhK-4 cells (CRL-1688) infected and incubated for 6 daysCells grown in MEMc with 7% (vol/vol) fetal bovine serum (Fisher; M33-500) and 1% nonessential amino acids (Gibco; 11140) at 36 ± 1°C, monolayers infected and incubated at 36 ± 1°C for 7 days in medium with no antibiotics, flasks frozen and thawed (thrice), cell lysate centrifuged, and supernatant aliquoted for storage at −80°C8 × 108 PFU/ml
Open in a separate windowaOADC, oleic acid-albumin-dextrose-catalase.bADC, albumin dextrose-catalase.cMEM, minimal essential medium.Disinfectants tested were peracetic acid (PAA; 500 and 1,000 ppm), chlorine dioxide (CD; 500 and 1,000 ppm), and domestic bleach (DB; 2,500, 3,300, and 5,000 ppm). Buffered saline (pH 7.2) was the control fluid, eluent, and diluent. Hard water (400 ppm CaCO3) was the diluent for disinfectants (1).Each disk received 10 μl of the inoculum, dried and covered with 50 μl of test substance, or saline at 22 ± 2°C. At the end of the contact time, each disk was eluted in a neutralizer and the eluates were assayed (1, 9, 11, 12). The neutralizer consisted of 1% dextrose (Difco), 0.7% lecithin (Alfa Aesar), 0.25% sodium bisulfite (J. T. Baker), 0.1% sodium thioglycolate (Sigma), 0.6% sodium thiosulfate (Analar), 0.2% l-cysteine (Sigma), 0.5% tryptone (Oxoid), and 0.1% Tween 80 (Bioshop) in buffered saline (pH 7.2). In each experiment, three control and three test carriers were used, and all experiments were repeated thrice. The performance criteria for the tested substances were ≥3.0 log10 reduction in PFU of the virus and ≥6.0 log10 reductions in the CFU for the other three organisms. When the mixture of test organisms was used, the components were separated by first passing the mixture through a membrane filter (0.22-μm pore diameter) to retain all the organisms except the virus. The filtrate was subjected to plaque assays for HAV in FRhK-4 cells. For the three bacteria, separate filters were placed on appropriate agar plates (Table (Table1)1) and incubated.The data for 5-min contact are given in Table Table2.2. All levels of the disinfectants tested met the criterion for M. terrae and A. baumannii when tested individually or in mixture. Only 1,000 ppm of PAA was effective against the spores. Both levels of PAA were ineffective against HAV, while the other disinfectants could reduce its titer between 3.5 and 4 log10. Only 1,000 ppm of PAA could consistently meet the criterion for sporicidal activity after 10 min (data not shown). Extending the contact time to 20 min allowed both levels of PAA and DB to meet the criterion for sporicidal activity, while 500 ppm of CD failed to do so; CD at 1,000 ppm barely met the criterion when tested alone against the spores but could not do so in the mixture (Fig. (Fig.11).Open in a separate windowFIG. 1.Reductions of G. stearothermophilus spores by the test formulations after 20 min of contact, individually and in a mixture at 22 ± 2°C.

TABLE 2.

Reductions by the test formulations in 5 min at 22 ± 2°C when tested against each organism individually and in a mixture
Disinfectant (concn [ppm])Mean log10 reduction ± SD of:
M. terrae
A. baumannii
G. stearothermophilus
Hepatitis A virus
IndividualMixtureIndividualMixtureIndividualMixtureIndividualMixture
Peracetic acid (500)8.18 ± 0.197.33 ± 0.167.19 ± 0.036.33 ± 0.034.03 ± 0.084.45 ± 0.98Not tested0.30 ± 0.01
Peracetic acid (1,000)8.18 ± 0.197.33 ± 0.167.19 ± 0.036.33 ± 0.038.03 ± 0.287.21 ± 0.590.58 ± 0.220.68 ± 0.09
Chlorine dioxide (500)8.18 ± 0.197.72 ± 0.217.22 ± 0.036.37 ± 0.131.47 ± 0.450.69 ± 0.054.30 ± 0.183.97 ± 0.19
Chlorine dioxide (1,000)8.18 ± 0.197.72 ± 0.217.22 ± 0.036.37 ± 0.133.07 ± 0.091.27 ± 0.054.30 ± 0.183.97 ± 0.19
Domestic bleach (2,500)8.18 ± 0.197.72 ± 0.217.22 ± 0.036.37 ± 0.130.27 ± 0.030.25 ± 0.024.41 ± 0.233.97 ± 0.29
Domestic bleach (3,500)8.18 ± 0.197.72 ± 0.217.22 ± 0.036.37 ± 0.130.27 ± 0.030.25 ± 0.024.41 ± 0.233.45 ± 0.09
Domestic bleach (5,000)8.18 ± 0.197.72 ± 0.217.22 ± 0.036.37 ± 0.130.28 ± 0.010.25 ± 0.024.41 ± 0.233.97 ± 0.29
Open in a separate windowThe study showed the feasibility of testing liquid chemicals against a mixture of suitable surrogates for infectious bioagents. This approach allowed standardized and simultaneous assessment of the spectrum of microbicidal activities of the test formulations under identical conditions that better simulate field conditions and that can be readily adapted to test foams and gaseous chemicals on other carrier materials. The surrogates selected covered the spectrum of microbicide resistances of all currently known classes of infectious bioagents.A. baumannii is among the more environmentally stable and microbicide-resistant vegetative bacteria known (7, 13). M. terrae represented pathogens with generally higher resistance to microbicides (3) and possibly drug-resistant Mycobacterium tuberculosis and category C agents (6). HAV, a small, nonenveloped virus known for its stability and microbicide resistance (9), represented select agents (CBW, biological weapons classification, 2001 [http://www.selectagents.gov/Select%20Agents%20and%20Toxins%20List.html]) and also food- and waterborne pathogens listed as biothreats (2, 10). The spores of G. stearothermophilus may be more resistant to oxidizing chemicals than the spores of Bacillus anthracis (8); their thermophilic nature made them safer to handle and easy to separate from the mixtures.The disinfectants were selected for their commercial availability and broad-spectrum and relatively rapid action (5, 14). The last criterion excluded all but oxidizers because other common active agents are limited as microbicides and/or require hours of contact for sporicidal action.For PAA tests, the recovery of infectious HAV in the absence of any viable spores is somewhat anomalous but not surprising. While we do not believe HAV to be more resistant than bacterial spores, the small size of the virus in the dried inocula likely afforded it significant protection. Compared to HAV, the mycobacterium proved more susceptible to all the disinfectants tested. This highlights a serious weakness in the traditional rankings of disinfectant susceptibility, where mycobacteria are often considered more resistant than nonenveloped viruses (5, 14).In the initial trials with the mixtures, the titer of A. baumannii dropped sharply; using virus pools without antibiotics resolved the issue. The ability of A. baumannii to grow on 7H11 agar and thus interfere with the recovery of M. terrae was addressed by replacing the standard strain of M. terrae with one containing a kanamycin resistance gene (15). Incorporation of enough kanamycin in 7H11 suppressed the growth of A. baumannii while allowing the mycobacterium to grow.Using a mixture of surrogates in QCT not only proved feasible but also highlighted the need to review certain long-held concepts about the relative sensitivities of classes of pathogens to disinfectants. The details reported should allow extension of the work to CL-3 and possibly CL-4 agents to confirm that the results obtained with the carefully chosen surrogates are indeed applicable to various classes of infectious bioagents.  相似文献   

20.
Identification of Pathogenic Vibrio Species by Multilocus PCR-Electrospray Ionization Mass Spectrometry and Its Application to Aquatic Environments of the Former Soviet Republic of Georgia     
Chris A. Whitehouse  Carson Baldwin  Rangarajan Sampath  Lawrence B. Blyn  Rachael Melton  Feng Li  Thomas A. Hall  Vanessa Harpin  Heather Matthews  Marina Tediashvili  Ekaterina Jaiani  Tamar Kokashvili  Nino Janelidze  Christopher Grim  Rita R. Colwell  Anwar Huq 《Applied and environmental microbiology》2010,76(6):1996-2001
  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