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 125 毫秒
1.
2.
Bacteroidales species were detected in (tap) water samples from treatment plants with three different PCR assays. 16S rRNA gene sequence analysis indicated that the sequences had an environmental rather than fecal origin. We conclude that assays for Bacteroidales 16S rRNA genes are not specific enough to discern fecal contamination of drinking water in the Netherlands.Drinking water in many countries is routinely monitored for recent fecal contamination by testing for fecal indicator organisms Escherichia coli, thermotolerant coliforms, and/or intestinal enterococci to demonstrate microbial safety (13, 21, 42). Although these indicator organisms have been used for many decades, they have some limitations: the number of E. coli/coliform/enterococcus bacteria in feces is relatively low (18, 38), and they sometimes might be able to grow in the environment (10, 11, 14, 27). Consequently, scientists have been searching for alternative indicator organisms to determine fecal contamination of water. In 1967, bacteria belonging to the genus Bacteroides were suggested as alternative indicator organisms (26). Bacteroides spp. might have some advantages over the traditional indicator organisms. The numbers of Bacteroides spp. in the intestinal tract of humans and animals are 10 to 100 times higher than the numbers of E. coli or intestinal enterococci (1, 2, 12, 26). However, the use of Bacteroides spp. as indicator organisms was hampered by the complex cultivation conditions required (1, 2). The introduction of molecular methods made it possible to detect bacterial species that belong to the order Bacteroidales, an order that includes the genus Bacteroides, without cultivation. As a result, real-time PCR methods were developed for the quantitative detection of Bacteroidales in surface and recreation water and the potential of Bacteroidales species as an indication of fecal contamination of recreational waters was demonstrated (6, 12, 16, 19, 20, 29). Bacteroidales species might be useful indicator organisms for fecal contamination of drinking water as well. However, methods to detect fecal contamination in drinking water should be more sensitive, because people ingest more drinking water and the quality assessments and standards for fecal contamination are stricter than for bathing water. Studies exploring real-time PCR for the detection of Bacteroidales genes in drinking water have not been published to our knowledge. The objective of our study was, therefore, to determine if assays for the detection of Bacteroidales 16S rRNA genes can be used to detect fecal contamination in drinking water.Unchlorinated tap water samples were obtained in November 2007 and February 2010 from one or more locations in the distribution systems of nine different drinking water treatment plants (plants A to I; Table Table1)1) that produced unchlorinated drinking water from confined (plants B, C, E, F, and G) and unconfined (plants A, D, H, and I) groundwater. The treatment plants are located in the central part of the Netherlands within 90 km of each other. In addition, untreated groundwater from extraction wells and/or untreated raw groundwater (mixture of groundwater from different extraction wells) was sampled in March 2008 (Table (Table1).1). Water samples (100 ml) were filtered over a 25-mm polycarbonate filter (0.22-μm pore size, type GTTP; Millipore, Netherlands) and a DNA fragment was added as internal control to determine the recovery efficiency of DNA isolation and PCR analysis (2a, 40). DNA was isolated using a FastDNA spin kit for soil (Qbiogene, United States) according to the supplier''s protocol. Primer sets AllBac 296f and AllBac 412r, resulting in a PCR product of 108 bp, were used in combination with TaqMan probe AllBac375Bhqr to quantitatively determine the number of Bacteroidales 16S rRNA gene copies in the water samples using a real-time PCR instrument (20). The PCR cycle after which the fluorescence signal of the amplified DNA was detected (threshold cycle [CT]) was used to quantify the concentration of 16S rRNA gene copies. Quantification was based on comparison of the sample CT value with the CT values of a calibration curve graphed using known copy numbers of the Bacteroidales 16S rRNA gene, as previously described (12, 20). The correlation coefficient of the calibration curve was 0.99, and the efficiency of the PCR 95 to 105%. Finally, the Bacteroidales cell number was calculated by using the recovery rate of the internal standard and assuming five 16S rRNA gene copy numbers per cell (5). The detection limit of this gene assay was 50 Bacteroidales cells 100 ml−1 (corresponding to 10 16S rRNA gene copies per reaction mixture). Furthermore, the 16S rRNA genes that were obtained from several water samples from treatment plant C with the AllBac and TotBac (12) primer sets were sequenced, and the nearest relatives were obtained from the GenBank database using BLAST searches.

TABLE 1.

Numbers of Bacteroidales cells in extraction wells, raw groundwater, and unchlorinated tap water of nine different groundwater plants in the Netherlandsa
PlantSource of sampleNo. (100 ml−1) of Bacteroidales cells in:
200720082010
ATap water 1b5,948 ± 950
Tap water 22,682 ± 1,4591,254 ± 216
Tap water 34,362 ± 947439 ± 136
Raw water96 ± 15
BTap water 13,553 ± 9815,302 ± 2,952
Tap water 24,487 ± 3912,119 ± 1,367
Tap water 37,862 ± 4,5883,896 ± 3,003
Raw water3,209 ± 833
CTap water 1661 ± 75386 ± 199
Tap water 21,051 ± 626
Tap water 3831 ± 584
Tap water 41,254 ± 216
Extraction well 11,126 ± 262
Extraction well 22,666 ± 51
Extraction well 3<50
Raw water90 ± 44
DTap water1,103 ± 291,254 ± 216
Raw water48 ± 16
ETap water1,302 ± 2221,254 ± 216
Extraction well 1671 ± 97
FTap water1,317 ± 198
Raw water<50
GTap water 1675 ± 92439 ± 300
Tap water 2216 ± 65249 ± 98
Tap water 3154 ± 6322 ± 137
Raw water<50
HTap water7,073 ± 845
Raw water511 ± 254
ITap water1,577 ± 176
Raw water420 ± 66
Open in a separate windowaValues are the average results and standard deviations from replicate PCRs on the same water sample using the AllBac primer set (20). In November 2007, the distribution systems (tap water) of plants A, B, and G were sampled at three different locations, whereas for the other plants, one location in the distribution system was sampled. In March 2008, raw water of plants A to G was sampled, as well as one (plant E) or three (plant C) different extraction wells. Finally, in February 2010, the distribution systems of plants A, B, C, D, E, and G were sampled again.bMore than one tap water sample from a treatment plant means that samples were taken at different locations in the distribution system.The Bacteroidales 16S rRNA gene, quantified with the AllBac primer set, was detected in all tap water samples in November 2007 and February 2010. The number of cells varied between 154 and 7,862 Bacteroidales cells 100 ml−1, and the numbers in tap water of each plant were similar in 2007 and 2010 (Table (Table1).1). The Bacteroidales counts were high compared to the number of E. coli that are occasionally observed in fecally contaminated drinking water (17a) but low compared to numbers observed in surface water (4, 20, 22). Water from the extraction wells and raw water used for unchlorinated drinking water production were analyzed, and Bacteroidales species were detected in 10 out of 15 samples (Table (Table1).1). These results would imply that the extracted groundwater, raw water, and tap water were fecally contaminated. According to the Dutch drinking water decree (2b), both raw and tap water from the nine different treatment plants are regularly analyzed for fecal contamination by monitoring for E. coli, F-specific RNA phages, and somatic coliphages. For at least the last 10 years, these indicator organisms have not been detected in these waters.Additional qualitative PCR analyses using TotBac and BacUni primer sets (12, 19) targeting other parts of the Bacteroidales 16S rRNA gene were performed to confirm the presence of Bacteroidales species in the water samples of November 2007 and March 2008. Nine or 10 of the 11 samples that were positive with the AllBac primer set were also positive with the TotBac and BacUni primer sets (data not shown). The BacUni primer set has a higher detection limit (30 gene copies per PCR; 19), which could explain the difference from the results with the AllBac primer set. The TotBac primer set has the same detection limit as the AllBac primer set (12), but small differences in PCR efficiencies might have resulted in different results, since some water samples showed Bacteroidales 16S rRNA gene copy numbers around the detection limit (Table (Table1).1). Nevertheless, the additional PCR analyses demonstrated that the detection of Bacteroidales species in tap, raw, and extracted well water with the AllBac primer set was not an artifact. The primer sets used were developed in three different studies (12, 19, 20) but have been applied in a number of recent studies to detect fecal contamination of surface water (3, 4, 16, 22, 33, 34). The results from most of these studies showed that 16S rRNA genes of Bacteroidales were present in all surface water samples tested. Only Sinigalliano et al. (34) observed that 2 out of 4 water samples were negative with the TotBac primer set. However, the detection limit of the assay was not specified in that study.The nine different treatment plants tested in our study produce unchlorinated drinking water from groundwater, which is considered to be of high hygienic quality. In addition, the extraction wells are protected from fecal contamination by a protection zone where no activities related to human waste or animal manure are allowed. In the Netherlands, this protection zone is based on a 60-day residence time of the water. Previous studies have demonstrated that a residence time of 60 days is highly effective in removing fecal bacteria and viruses (30, 31, 39). Moreover, the Bacteroidales numbers in tap water in November 2007 were significantly higher than the numbers in raw groundwater in March 2008 (Mann-Whitney U test; P < 0.01). Because the recovery efficiency of the internal control was the same between raw water and tap water samples, this result demonstrates that Bacteroidales cell numbers increased during treatment and/or drinking water distribution. This result could suggest that the water was fecally contaminated during drinking water treatment and/or distribution. However, it is unlikely that the integrity of nine different treatment trains and/or supply systems was affected in the sampling period. The statutory monitoring did not show the presence of E. coli at these sites. Another hypothesis is that the increase of Bacteroidales cell numbers in tap water was caused by the growth of Bacteroidales species in (drinking) water systems. In summary, it is unexpected that the majority of the tap water, raw water, and extracted groundwater samples were fecally contaminated. These unexpected observations raise the question of whether the PCR methods detect only fecal Bacteroidales species and, thus, if the gene assays are suitable to discern fecal contamination in drinking water in the Netherlands.Sequence analyses of the Bacteroidales 16S rRNA genes were performed to determine the relatedness of sequences from the different sampling sites to sequences from the nearest relatives in the GenBank database. All sequences contained the primer regions, indicating that nonspecific amplification had not occurred in the PCRs. Because the PCR product from the AllBac primer set was small (108 bp), many 16S rRNA gene sequences (100 to 5,000) in the GenBank database were identical to the Bacteroidales 16S rRNA gene sequences obtained from groundwater and unchlorinated tap water samples from plant C. These identical 16S rRNA gene sequences were in general obtained from fecal sources, but some of them came from environmental rather than fecal sources (Table (Table2).2). The AllBac 16S rRNA gene sequences from tap water and groundwater had relative high similarities (96.3 to 100%) to sequences from bacterial species of the genera Bacteroides, Prevotella, and Tannerella (Table (Table2),2), which all belong to the order Bacteroidales.

TABLE 2.

Nearest relatives in GenBank to the Bacteroidales 16S rRNA gene sequences obtained from groundwater and unchlorinated tap water from plant C using different primer setsa
Primer set used, source of sample, and OTUsbGenBank sequence accession no.Source of sequence (GenBank sequence accession no.)SimilaritycNearest cultivated bacterium in GenBank (sequence accession no.)Similarity
AllBac
    Extraction well 1 (3/6)GQ169588Rhizosphere (EF605968)108/108Prevotella oralis (AY323522)105/108
    Extraction well 1 (3/6)GQ169589Water from watershed (DQ886209)108/108Tannerella forsythia(AB035460)107/108
    Extraction well 2 (1/6)GQ169590Phyllosphere Brazilian forest (DQ221468)108/108Tannerella forsythia(AB035460)106/108
    Extraction well 2 (5/6)GQ169591Bovine rumen (EU348207)108/108Tannerella forsythia(AB035460)106/108
    Extraction well 3 (1/6)GQ169592Phyllosphere Brazilian forest (DQ221468)108/108Prevotella oralis (AY323522)104/108
    Extraction well 3 (5/6)GQ169593Prevotella corporis (L16465)108/108Prevotella corporis (L16465)108/108
    Raw water (3/6)GQ169594Spitsbergen permafrost (EF034756)108/108Tannerella forsythia(AB035460)106/108
    Raw water (3/6)GQ169595Hindgut beetle larvae (FJ374179)108/108Tannerella forsythia(AB035460)107/108
    Tap water (6/6)GQ169596Prevotella timonensis (DQ518919)108/108Prevotella timonensis (DQ518919)108/108
    Prevotella buccalis (L16476)Prevotella buccalis (L16476)
    Prevotella ruminicola (AF218617)Prevotella ruminicola (AF218617)
    Bacteroides vulgatus (NC_009614)Bacteroides vulgatus (NC_009614)
TotBac
    Extraction well 1 (1/10)GQ169597Deep subsurface groundwater (AB237705)339/369Salinimicrobium terrae (EU135614)315/370
    Extraction well 1 (1/10)GQ169598Songhuajiang River sediment (DQ444125)363/377Paludibacter propionicigenes (AB078842)357/376
    Extraction well 1 (4/10)GQ169599Freshwater pond sediment (DQ676447)352/360Paludibacter propionicigenes (AB078842)313/372
    Extraction well 1 (4/10)GQ169600Pine River sediment (DQ833352)364/371Bacteroides oleiciplenus (AB490803)334/375
    Extraction well 2 (4/10)GQ169601Groundwater (AF273319)364/371Xanthobacillum maris (AB362815)338/375
    Extraction well 2 (6/10)GQ169602Human saliva (AB028385)381/382Prevotella intermedia (AY689226)380/382
    Extraction well 3 (1/10)GQ169603Pig manure (AY816766)354/377Bacteroides thetaiotaomicron (AE015928)311/380
    Extraction well 3 (3/10)GQ169604Pig manure (AY816867)371/376Butyricimonas virosa (AB443949)307/379
    Extraction well 3 (6/10)GQ169605Swedish lake (AY509350)343/362Parabacteroides distasonis (AB238927)320/374
    Raw water (10/10)GQ169606Prevotella timonensis (AF218617)378/379Prevotella timonensis (AF218617)378/379
    Tap water (1/10)GQ169607Deep subsurface groundwater (AB237705)338/369Salinimicrobium terrae (EU135614)312/370
    Tap water (2/10)GQ169608Yukon River, AK(FJ694652)367/372Psychroserpens burtonensis (U62913)312/375
    Tap water (7/10)GQ169609Deep subsurface groundwater (AB237705)341/369Salinimicrobium terrae (EU135614)315/370
Open in a separate windowaPrimer sets AllBac (20) and TotBac (12) were used in PCRs of samples, and GenBank was searched for relatives using BLAST.bOTUs are indicated by the values in parentheses (number of sequences belonging to the OTU/total number of sequences analyzed).cNumber of base pairs identical in both sequences/total number of base pairs in sequences.16S rRNA gene sequences obtained with the TotBac primer set were longer (∼370 bp) and did not show 100% similarity with the nearest relatives in the GenBank database (Table (Table2).2). Sequences from the GenBank database that showed the highest similarity (91.6% to 99.7%) with the 16S rRNA gene sequences from tap water and groundwater from plant C were in general isolated from environmental sources (Table (Table2).2). The 16S rRNA gene sequences from cultivated bacterial species that showed the highest similarity to the 16S rRNA gene sequences obtained in our study belonged to different genera (Table (Table2).2). Some of these genera (Salinimicrobium, Xanthobacillum, and Psychroserpens) did not belong to the order Bacteroidales. However, the 16S rRNA gene sequences from bacterial species of these genera showed low similarities with the sequences obtained in this study (83.2% to 90.1%) and six mismatches to the TotBac primers. Thus, it is unlikely that DNA from bacterial species belonging to Salinimicrobium, Xanthobacillum, and Psychroserpens was amplified in the gene assay. More importantly, the majority of the nearest environmental clone sequences retrieved from the GenBank database showed no or a single mismatch with the AllBac and TotBac primer and probe sequences. Thus, these primer sets are capable of amplifying 16S rRNA genes from bacteria that have been observed in ecosystems outside the intestinal tract of humans and animals.16S rRNA gene sequences related to Prevotella species were commonly observed in extracted groundwater, raw water, and tap water (Table (Table2).2). The isolation of Prevotella paludivivens from rice roots in a rice field soil (35) demonstrated the environmental nature of some Prevotella species. In addition, primer sequences developed for the detection of fecal Bacteroidales species (8, 12, 19, 20, 25, 29) showed no or a single mismatch with 16S rRNA gene sequences from P. paludivivens, Xylanibacterium oryzae, Paludibacter propionicigenes, Proteiniphilum acetatigenes, and Petrimonas sulfuriphila that are present in the GenBank database. These five Bacteroidales species have all been isolated from ecosystems other than the gastrointestinal tract. Consequently, primer sets for 16S rRNA genes of Bacteroidales species cannot always be used to discern fecal contamination in water.A number of 16S rRNA gene sequences observed in groundwater and tap water fell in the genus Bacteroides. The presence of Bacteroides 16S rRNA gene sequences in groundwater and tap water might also suggest that some Bacteroides species are capable of growth in the environment. However, until now, type strains of Bacteroides species growing outside the animal intestinal tract have not been published. Another possible explanation is that the observed 16S rRNA gene sequences originate from Bacteroides species that inhabit the anoxic intestinal tract of insects. Previous studies have shown that bacterial species belonging to the genus Bacteroides are common inhabitants of the hindguts of insects (15, 23, 24, 28, 32). Some of the 16S rRNA gene sequences obtained with the AllBac primer set in our study showed 100% similarity to 16S rRNA gene sequences from the hindgut of insects. Moreover, a number of 16S rRNA gene sequences isolated from the hindguts of insects (15, 23, 24, 32) showed no or a single mismatch with the TotBac and AllBac primer and probe sequences. In conclusion, these primer sets are capable of detecting Bacteroides species from the hindgut of insects as well. Water insects are normal inhabitants of groundwater and drinking water distribution systems (7, 41) and might be a source of Bacteroides species in water. Bacteroides species from insect feces do not indicate fecal pollution by warm-blooded animals, and insects do not normally shed human fecal pathogenic microorganisms. Bacteroides species from insect feces, therefore, can hamper Bacteroides gene assays developed for the detection of water fecally contaminated by warm-blooded animals. Additional cultivation techniques in combination with molecular tools are required to demonstrate the persistence or growth of Bacteroides bacteria in groundwater and drinking water or whether Bacteroides bacteria are present in water insects. However, these experiments were beyond the scope of our study.The three extraction wells of plant C are located close to each other and extract water from the same aquifer. Subsequently, extracted water from the three wells is mixed and enters the treatment plant as raw water. We hypothesize that if a fecal source in the vicinity of the extraction field of plant C contaminated the groundwater, water from the extraction wells and raw water should (partly) have the same Bacteroidales species. Although a relatively limited amount of clones was sequenced per sample (16), the diversity of Bacteroidales operational taxonomic units (OTU) within a sample was low (Table (Table2).2). In contrast, unique 16S rRNA gene sequences were observed between the different water types (e.g., extracted groundwater, raw water, and tap water) and sequence overlap between water types was low. These results demonstrate that the Bacteroidales 16S rRNA gene sequences at the sampling locations were not from the same fecal source and imply once again that Bacteroidales species were environmental rather than fecal.Finally, we hypothesized that if the Bacteroidales species observed in tap water were of nonfecal origin, human- and/or bovine-specific Bacteroidales strains should not be present in tap water. We tested for the presence of human- or bovine-specific Bacteroidales strains by using source-specific 16S rRNA gene assays (5) on tap water samples from February 2010. The results showed that human- and bovine-specific Bacteroidales 16S rRNA genes could not be detected in tap water, whereas a PCR product was always detected with the positive control. Again, these results indicate that the Bacteroidales species observed in tap water were of nonfecal origin.Overall, the results from our study indicate that gene assays for Bacteroidales detected environmental rather than fecal Bacteroidales species in groundwater and tap water from treatment plants in the Netherlands. First, Bacteroidales 16S rRNA gene sequences obtained from water samples taken at plant C showed (high) similarity to clone sequences that were isolated from environmental sources. The majority of these clone sequences and several Bacteroides clone sequences from the hindguts of insects showed no or a single mismatch with AllBac, TotBac, and BacUni primer and probe sequences. Second, the primer and probe sequences used for the gene assays have no or a single mismatch with 16S rRNA gene sequences of environmental Bacteroidales species P. paludivivens, X. oryzae, P. propionicigenes, P. acetatigenes, and/or P. sulfuriphila (9, 17, 35-37). Third, Bacteroidales 16S rRNA gene sequences from raw water and water from extraction wells were unique, and sequence overlap between water types was low. It is expected that in the case of fecal contamination of groundwater, different water types from the same groundwater area have similar Bacteroidales species. Fourth, the quantitative assays for Bacteroidales 16S rRNA genes commonly used to detect fecal contamination (3, 4, 12, 16, 19, 20, 22, 33, 34) detected Bacteroidales species in deep groundwater and tap water that have no history of fecal contamination. Fifth, Bacteroidales gene copy numbers were significantly higher in tap water than in raw groundwater, demonstrating an increase or growth of Bacteroidales species during the treatment and/or distribution of drinking water. Finally, human- and bovine-specific Bacteroidales strains were not detected in tap water. Consequently, (quantitative) assays for general Bacteroidales 16S rRNA genes are not suitable to discern fecal contamination in groundwater and unchlorinated drinking water in the Netherlands.Nucleotide sequence accession numbers.The 16S rRNA gene sequences obtained in this study were deposited in the GenBank database under accession numbers GQ169588 to GQ169609.  相似文献   

3.
4.
We analyzed the temporal and spatial diversity of the microbiota in a low-usage and a high-usage hospital tap. We identified a tap-specific colonization pattern, with potential human pathogens being overrepresented in the low-usage tap. We propose that founder effects and local adaptation caused the tap-specific colonization patterns. Our conclusion is that tap-specific colonization represents a potential challenge for water safety.Humans are exposed to and consume large amounts of tap water in their everyday life, with the tap water microbiota representing a potent reservoir for pathogens (8). Despite the potential impact, our knowledge about the ecological diversification processes of the tap water microbiota is limited (4, 11).The aim of the present work was to determine the temporal and spatial distribution patterns of the planktonic tap water microbiota. We compared the summer and winter microbiota from two hospital taps supplied from the same water source. We analyzed 16S rRNA gene clone libraries by using a novel alignment-independent approach for operational taxonomic unit (OTU) designation (6), while established OTU diversity and richness estimators were used for the ecological interpretations.Tap water samples (1 liter) from a high-usage kitchen and a low-usage toilet cold-water tap in Akershus University Hospital, Lørenskog, Norway, were collected in January and July 2006. The total DNA was isolated and the 16S rRNA gene PCR amplified and sequenced. Based on the sequences, we estimated the species richness and diversity, we calculated the distances between the communities, and trees were constructed to reflect the relatedness of the microbiota in the samples analyzed. Details about these analytical approaches are given in the materials and methods section in the supplemental material.Our initial analysis of species composition was done using the RDPII hierarchical classifier. We found that the majority of pathogen-related bacteria in our data set belonged to the class Gammaproteobacteria. The genera encompassed Legionella, Pseudomonas, and Vibrio (Table (Table1).1). We found a significant overrepresentation of pathogen-related bacteria in the toilet tap (P = 0.04), while there were no significant differences between summer and winter samples. Legionella showed the highest relative abundance for the pathogen-related bacteria. With respect to the total diversity, we found that Proteobacteria dominated the tap water microbiota (representing 86% of the taxa) (see Table S1 in the supplemental material). There was, however, a large portion (56%) of the taxa that could not be assigned to the genus level using this classifier.

TABLE 1.

Cloned sequences related to human pathogensa
Sampling placeSampling timePathogenNCBI accession no.Identity (%)
ToiletSummerEscherichia coliEF41861499
ToiletSummerEscherichia sp.EF07430799
ToiletSummerLegionella sp.AY92415595
ToiletSummerLegionella sp.AY92415395
ToiletSummerLegionella sp.AY92415396
ToiletWinterLegionella sp.AY92406196
ToiletWinterLegionella sp.AY92415897
ToiletWinterLegionella sp.AY92415897
KitchenWinterLegionella sp.AY92399697
ToiletSummerPseudomonas fluorescensEF41307398
ToiletSummerPseudomonas fluorescensEF41307398
KitchenSummerPseudomonas fluorescensDQ20773199
ToiletWinterVibrio sp.DQ40838898
ToiletWinterVibrio sp.AB27476098
KitchenWinterVibrio sp.DQ40838898
KitchenWinterVibrio lentusAY29293699
KitchenWinterVibrio sp.AM18376597
ToiletWinterStenotrophomonas maltophiliaAY83773099
KitchenWinterStenotrophomonas maltophiliaDQ42487098
ToiletWinterStreptococcus suisAF28457898
ToiletWinterStreptococcus suisAF28457898
Open in a separate windowaThe relatedness between the cloned sequences and potential pathogens was determined by BLAST searches of the NCBI database, carried out using default settings.To obtain a better resolution of the uncharacterized microbiota, we analyzed the data using a clustering approach that is not dependent on a predefined bacterial group (see the materials and methods section in the supplemental material for details). These analyses showed that there were three relatively tightly clustered groups in our data set (Fig. (Fig.1A).1A). The largest group (n = 590) was only distantly related to characterized betaproteobacteria within the order Rhodocyclales. We also identified another large betaproteocaterial group (n = 320) related to Polynucleobacter. Finally, a tight group (n = 145) related to the alphaproteobacterium Sphingomonas was identified.Open in a separate windowFIG. 1.Tap water microbiota diversity, determined by use of a principal component analysis coordinate system. (A) Each bacterium is classified by coordinates, with the following color code: brown squares, kitchen summer; red diamonds, toilet summer; green triangles, kitchen winter; and green circles, toilet winter. (B and C) Each square represents a 1 × 1 (B) or 5 × 5 (C) OTU. PC1, first principal component; PC2, second principal component.The tap-specific distributions of the bacterial groups were investigated using density distribution analyses. A dominant population related to Polynucleobacter was identified for the toilet summer samples, while for the winter samples there was a dominance of the Rhodocyclales-related bacteria. The kitchen summer samples revealed a dominance of Sphingomonas. The corresponding winter samples did not reveal distinct high-density bacterial populations (see Table S2 in the supplemental material).Hierarchical clustering for the 1 × 1 OTU density distribution confirmed the relatively low overlap for the microbiota in the samples analyzed (Fig. (Fig.2).2). We found that the microbiota clustered according to tap and not season.Open in a separate windowFIG. 2.Hierarchical clustering for the density distribution of the tap water microbiota. The density of 1 × 1 OTUs was used as a pseudospecies for hierarchical clustering. The tree for the Cord distance matrix is presented, while the distances calculated using the three distance matrices Cord, Brad Curtis, and Sneath Sokal, respectively, are shown for each branch.We have described the species diversity and richness of the microbiota in Table S3 in the supplemental material. For the low taxonomic level, these analyses showed that the diversity and species richness were greater for the winter samples than for the summer samples. Comparing the two taps, the diversity and richness were greater in the kitchen tap than in the toilet tap. In particular, the winter sample from the kitchen showed great richness and diversity. The high taxonomic level, however, did not reveal the same clear differences as did the low level, and the distributions were more even. Rarefaction analyses for the low taxonomic level confirmed the richness and diversity estimates (see Fig. S1 in the supplemental material).Our final analyses sought to fit the species rank distributions to common rank abundance curves. Generally, the rank abundance curves were best fitted to log series or truncated log normal distributions (see Table S4 in the supplemental material). The log series distribution could be fit to all of the samples except the kitchen summer samples at the low taxonomic level, while the truncated log normal distribution could not be fit to the kitchen samples at the high taxonomic level. Interestingly, however, the kitchen winter sample was best fit to a geometric curve at both the high and the low taxonomic level.Diversifying, adaptive biofilm barriers have been documented for tap water bacteria (7), and it is known that planktonic bacteria can interact with biofilms in an adaptive manner (3). On the other hand, tap usage leads to water flowthrough and replacement of the global with the local water population by stochastic founder effects (1).Therefore, we propose that parts of the local diversity observed can be explained by local adaptation (10) and parts by founder effects (9).Most prokaryote diversity measures assume log normal or log series OTU dominance density distributions (5). The kitchen winter sample, however, showed deviations from these patterns by being correlated to geometric distributions (in addition to the log series and truncated log normal distributions for the high taxonomic level). This sample also showed a much greater species richness than the other samples. A possible explanation is that the species richness of the tap water microbiota can be linked to usage and that the kitchen tap is driven toward a founder microbiota by high usage.Since our work indicates an overrepresentation of Legionella in the low-usage tap, it would be of high interest to determine whether the processes for local Legionella colonization can be related to tap usage. Understanding the ecological forces affecting Legionella and other pathogens are of great importance for human health. At the Akerhus University Hospital, this was exemplified by a Pseudomonas aeruginosa outbreak in an intensive care unit, where the outbreak could be traced back to a single tap (2).  相似文献   

5.
Twelve cluster groups of Escherichia coli O26 isolates found in three cattle farms were monitored in space and time. Cluster analysis suggests that only some O26:H11 strains had the potential for long-term persistence in hosts and farms. As judged by their virulence markers, bovine enterohemorrhagic O26:H11 isolates may represent a considerable risk for human infection.Shiga toxin (Stx)-producing Escherichia coli (STEC) strains comprise a group of zoonotic enteric pathogens (42). In humans, infections with some STEC serotypes result in hemorrhagic or nonhemorrhagic diarrhea, which can be complicated by hemolytic-uremic syndrome (HUS) (49). These STEC strains are also designated “enterohemorrhagic E. coli” (EHEC). Consequently, EHEC strains represent a subgroup of STEC with a high pathogenic potential for humans. Strains of the E. coli serogroup O26 were originally classified as enteropathogenic E. coli due to their association with outbreaks of infantile diarrhea in the 1940s. In 1977, Konowalchuk et al. (37) recognized that these bacteria produced Stx, and 10 years later, the Stx-producing E. coli O26:H11/H− strains were classified as EHEC. EHEC O26 strains constitute the most common non-O157 EHEC group associated with diarrhea and HUS in Europe (12, 21, 23, 24, 26, 27, 55, 60). Reports on an association between EHEC O26 and HUS or diarrhea from North America including the United States (15, 30, 33), South America (51, 57), Australia (22), and Asia (31, 32) provide further evidence for the worldwide spread of these organisms. Studies in Germany and Austria (26, 27) on sporadic HUS cases between 1996 and 2003 found that EHEC O26 accounted for 14% of all EHEC strains and for ∼40% of non-O157 EHEC strains obtained from these patients. A proportion of 11% EHEC O26 strains was detected in a case-control study in Germany (59) between 2001 and 2003. In the age group <3 years, the number of EHEC O26 cases was nearly equal to that of EHEC O157 cases, although the incidence of EHEC O26-associated disease is probably underestimated because of diagnostic limitations in comparison to the diagnosis of O157:H7/H− (18, 34). Moreover, EHEC O26 has spread globally (35). Beutin (6) described EHEC O26:H11/H−, among O103:H2, O111:H, O145:H28/H−, and O157:H7/H−, as the well-known pathogenic “gang of five,” and Bettelheim (5) warned that we ignore the non-O157 STEC strains at our peril.EHEC O26 strains produce Stx1, Stx2, or both (15, 63). Moreover, these strains contain the intimin-encoding eae gene (11, 63), a characteristic feature of EHEC (44). In addition, EHEC strains possess other markers associated with virulence, such as a large plasmid that carries further potential virulence genes, e.g., genes coding for EHEC hemolysin (EHEC-hlyA), a catalase-peroxidase (katP), and an extracellular serine protease (espP) (17, 52). The efa1 (E. coli factor for adherence 1) gene was identified as an intestinal colonization factor in EHEC (43). EHEC O26 represents a highly dynamic group of organisms that rapidly generate new pathogenic clones (7, 8, 63).Ruminants, especially cattle, are considered the primary reservoir for human infections with EHEC. Therefore, the aim of this study was the molecular characterization of bovine E. coli field isolates of serogroup O26 using a panel of typical virulence markers. The epidemiological situation in the beef herds from which the isolates were obtained and the spatial and temporal behavior of the clonal distribution of E. coli serogroup O26 were analyzed during the observation period. The potential risk of the isolates inducing disease in humans was assessed.In our study, 56 bovine E. coli O26:H11 isolates and one bovine O26:H32 isolate were analyzed for EHEC virulence-associated factors. The isolates had been obtained from three different beef farms during a long-term study. They were detected in eight different cattle in farm A over a period of 15 months (detected on 10 sampling days), in 3 different animals in farm C over a period of 8 months (detected on 3 sampling days), and in one cow on one sampling day in farm D (Table (Table1)1) (28).

TABLE 1.

Typing of E. coli O26 isolates
Sampling day, source, and isolateSerotypeVirulence profile by:
fliC PCR-RFLPstx1 genestx2 geneStx1 (toxin)Stx2 (toxin)Subtype(s)
efa1 genebEHEC-hlyA genekatP geneespP genePlasmid size(s) in kbCluster
stx1/stx2eaetirespAespB
Day 15
    Animal 6 (farm A)
        WH-01/06/002-1O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/06/002-2O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/06/002-3O26:H11H11++stx1ββββ+/++++110, 127
    Animal 8 (farm A)
        WH-01/08/002-2O26:H11H11++stx1ββββ+/++++110, 127
    Animal 26 (farm A)
        WH-01/26/001-2O26:H11H11++stx1ββββ+/++++130, 127
        WH-01/26/001-5O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/26/001-6O26:H11H11++stx1ββββ+/++++110, 127
        WH-01/26/001-7O26:H11H11++stx1ββββ+/−+++110, 127
Day 29
    Animal 2 (farm A)
        WH-01/02/003-1O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-2O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-5O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-6O26:H11H11++stx1ββββ+/+++110, 126
        WH-01/02/003-7O26:H11H11++stx1ββββ+/++++110, 126
        WH-01/02/003-8O26:H11H11++stx1ββββ−/++++110, 126
        WH-01/02/003-9O26:H11H11++stx1ββββ+/++++1106
        WH-01/02/003-10O26:H11H11++stx1ββββ+/++++1106
    Animal 26 (farm A)
        WH-01/26/002-2O26:H11H11++stx1ββββ+/++++130, 125
        WH-01/26/002-5O26:H11H11++stx1ββββ+/++++130, 125
        WH-01/26/002-8O26:H11H11++stx1ββββ+/++++130, 125
        WH-01/26/002-9O26:H11H11++stx1ββββ+/++110, 125
        WH-01/26/002-10O26:H11H11++stx1ββββ+/++++130, 125
Day 64
    Animal 20 (farm A)
        WH-01/20/005-3O26:H11H11++stx1ββββ+/+130, 2.52
Day 78
    Animal 29 (farm A)
        WH-01/29/002-1O26:H11H11++stx1ββββ+/−+130, 12, 2.54
        WH-01/29/002-2O26:H11H11++stx1ββββ+/++++130, 12, 2.54
        WH-01/29/002-3O26:H11H11++stx1ββββ+/++++130, 12, 2.54
        WH-01/29/002-4O26:H11H11++stx1ββββ+/++++130, 12, 2.54
        WH-01/29/002-5O26:H11H11++stx1ββββ+/++130, 12, 2.54
Day 106
    Animal 27 (farm A)
        WH-01/27/005-2O26:H11H11++stx1ββββ+/−+++145, 110, 123
        WH-01/27/005-5O26:H11H11++stx1ββββ+/++++130, 12, 2.55
        WH-01/27/005-6O26:H11H11++stx1ββββ+/+130, 12, 2.55
Day 113
    Animal 7 (farm C)
        WH-04/07/001-2O26:H11H11++++stx1/stx2ββββ+/+++55, 35, 2.511
        WH-04/07/001-4O26:H11H11++++stx1/stx2ββββ+/++++5512
        WH-04/07/001-6O26:H11H11++++stx1/stx2ββββ+/++++5512
Day 170
    Animal 22 (farm C)
        WH-04/22/001-1O26:H11H11++stx1ββββ+/++++110, 12, 6.312
        WH-04/22/001-4O26:H11H11++stx1ββββ+/++++110, 12, 6.312
        WH-04/22/001-5O26:H11H11++stx1ββββ+/++++110, 12, 6.312
Day 176
    Animal 14 (farm D)
        WH-03/14/004-8O26:H11H11++stx1ββββ+/+++11010
Day 218
    Animal 27 (farm A)
        WH-01/27/009-1O26:H11H11++++stx1/stx2ββββ+/++++110, 129
        WH-01/27/009-2O26:H11H11++++stx1/stx2ββββ+/++++110, 129
        WH-01/27/009-3O26:H11H11++++stx1/stx2ββββ+/++++110, 128
        WH-01/27/009-8O26:H11H11++++stx1/stx2ββββ+/++110, 128
        WH-01/27/009-9O26:H11H11++++stx1/stx2ββββ+/++++110, 129
Day 309
    Animal 29 (farm A)
        WH-01/29/010-1O26:H11H11++stx1ββββ+/++++110, 35, 124
        WH-01/29/010-2O26:H11H11++stx1ββββ+/++130, 55, 358
        WH-01/29/010-3O26:H11H11++stx1ββββ+/++++130, 35, 128
Day 365
    Animal 8 (farm C)
        WH-04/08/008-6O26:H11H11++stx1ββββ+/++++110, 5512
Day 379
    Animal 9 (farm A)
        WH-01/09/016-2O26:H32H32++stx1/stx2−/−145, 130, 1.81
    Animal 27 (farm A)
        WH-01/27/014-3O26:H11H11++stx1ββββ+/++++110, 129
        WH-01/27/014-4O26:H11H11++stx1ββββ+/++++110, 129
        WH-01/27/014-5O26:H11H11++stx1ββββ+/++++110, 128
Day 407
    Animal 29 (farm A)
        WH-01/29/013-4O26:H11H11++stx1ββββ+/++++110, 12, 2.58
        WH-01/29/013-7O26:H11H11++stx1ββββ+/++++110, 12, 2.58
Day 478
    Animal 27 (farm A)
        WH-01/27/017-1O26:H11H11++++stx1/stx2ββββ+/++++110, 128
        WH-01/27/017-5O26:H11H11++++stx1/stx2ββββ+/++++110, 128
        WH-01/27/017-6O26:H11H11++++stx1/stx2ββββ+/++++1108
        WH-01/27/017-7O26:H11H11++++stx1/stx2ββββ+/++++1108
        WH-01/27/017-10O26:H11H11+++stx1ββββ+/++++130, 12, 2.58
Open in a separate windowastx1/stx2, gene stx1 or stx2.befa1 was detected by two hybridizations (with lifA1-lifA2 and lifA3-lifA4 probes). +/+, complete gene; +/− or −/+, incomplete gene; −/−, efa1 negative.The serotyping of the O26 isolates was confirmed by the results of the fliC PCR-restriction fragment length polymorphism (RFLP) analysis performed according to Fields et al. (25), with slight modifications described by Zhang et al. (62). All O26:H11 isolates showed the H11 pattern described by Zhang et al. (62). In contrast, the O26:H32 isolate demonstrated a different fliC RFLP pattern that was identical to the H32 pattern described by the same authors. It has been demonstrated that EHEC O26:H11 strains belong to at least four different sequence types (STs) in the common clone complex 29 (39). In the multilocus sequence typing analysis for E. coli (61), the tested five EHEC O26:H11 isolates (WH-01/02/003-1, WH-01/20/005-3, WH-01/27/009-9, WH-03/14/004-8, and WH-04/22/001-1) of different farms and clusters were characterized as two sequence types (ST 21 and ST 396). The isolates from farms A and C belong to ST 21, the most frequent ST of EHEC O26:H11 isolates found in humans and animals (39), but the single isolate from farm D was characterized as ST 396.Typing and subtyping of genes (stx1 and/or stx2, eae, tir, espA, espB, EHEC-hlyA, katP, and espP) associated with EHEC were performed with LightCycler fluorescence PCR (48) and different block-cycler PCRs. To identify the subtypes of the stx2 genes and of the locus of enterocyte effacement-encoding genes eae, tir, espA, and espB, the PCR products were digested by different restriction endonucleases (19, 26, 46). The complete pattern of virulence markers was detected in most bovine isolates examined in our study. An stx1 gene was present in all O26 isolates. In addition, an stx2 gene was found in nine O26:H11 isolates in farm A and in three isolates of the same type in farm C, as well as in the O26:H32 isolate. Both Stx1 and Stx2 were closely related to families of Stx1 and Stx2 variants or alleles. EHEC isolates with stx2 genes are significantly more often associated with HUS and other severe disease manifestations than isolates with an stx1 gene, which are more frequently associated with uncomplicated diarrhea and healthy individuals (13). In contrast to STEC strains harboring stx2 gene variants, however, STEC strains of the stx2 genotype were statistically significantly associated with HUS (26). The stx2 genotype was found in all O26 isolates with an stx2 gene, while the GK3/GK4 amplification products after digestion with HaeIII and FokI restriction enzymes showed the typical pattern for this genotype described by Friedrich et al. (26). The nucleotide sequences of the A and B subunits of the stx2 gene of the selected bovine O26:H11 isolate WH-01/27/017-1 (GenBank accession no. EU700491) were identical to the stx2 genes of different sorbitol-fermenting EHEC O157:H− strains associated with human HUS cases and other EHEC infections in Germany (10) and 99.3% identical in their DNA sequences to the stx2 gene of the EHEC type strain EDL933, a typical O157:H7 isolate from an HUS patient. A characteristic stx1 genotype was present in all O26 isolates. The nucleotide sequences of the A and B subunits of the stx1 gene of the tested bovine O26:H11 isolate WH-01/27/017-1 (GenBank accession no. EU700490) were nearly identical to those of the stx1 genes of the EHEC O26:H11 reference type strains H19 and DEC10B, which had been associated with human disease outbreaks in Canada and Australia. Nucleotide exchanges typical for stx1c and stx1d subtypes as described by Kuczius et al. (38) were not found. All bovine O26:H11 strains produced an Stx1 with high cytotoxicity for Vero cells tested by Stx enzyme-linked immunosorbent assay and Vero cell neutralization assay (53). The Stx2 cytotoxicity for Vero cells was also very high in the O26:H11 isolates.Not only factors influencing the basic and inducible Stx production are important in STEC pathogenesis. It has been suggested that the eae and EHEC-hlyA genes are likely contributors to STEC pathogenicity (2, 3, 13, 50). Ritchie et al. (50) found both genes in all analyzed HUS-associated STEC isolates. In all O26:H11 isolates we obtained, stx genes were present in combination with eae genes. Only the O26:H32 isolate lacked an eae gene. To date, 10 distinct variants of eae have been described (1, 19, 36, 45, 47). Some serotypes were closely associated with a particular intimin variant: the O157 serogroup was linked to γ-eae, the O26 serogroup to β-eae, and the O103 serogroup to ɛ-eae (4, 19, 20, 58). Our study confirms these associations. All bovine O26:H11 isolates were also typed as members of the β-eae subgroup. A translocated intimin receptor gene (tir gene) and the type III secreted proteins encoded by the espA and espB genes were found in all 56 O26:H11 isolates but not in the O26:H32 isolate. These other tested locus of enterocyte effacement-associated genes belonged to the β-subgroups. These results are in accord with the results of China et al. (19), who detected the pathotypes β-eae, β-tir, β-espA, and β-espB in all investigated human O26 strains. Like the eae gene, the EHEC-hlyA gene was found in association with severe clinical disease in humans (52). Aldick et al. (2) showed that EHEC hemolysin is toxic (cytolytic) to human microvascular endothelial cells and may thus contribute to the pathogenesis of HUS. In our study, the EHEC-hlyA gene was detected in 50 of the 56 bovine E. coli O26:H11 isolates which harbored virulence-associated plasmids of different sizes (Table (Table1).1). The presence of virulence-associated plasmids corresponded to the occurrence of additional virulence markers such as the espP and katP genes (17). The katP gene and the espP gene were detected in 49 and 50 of the 56 O26:H11 isolates, respectively. The espP gene was missing in six of the seven bovine O26:H11 isolates in which the katP genes were also absent. Both genes were not found in the O26:H32 isolate (Table (Table1).1). Although we found large plasmids of the same size in O26:H11 isolates, they lacked one or more of the plasmid-associated virulence factors (Table (Table1).1). Two DNA probes were used to detect the efa1 genes by colony hybridization. (DNA probes were labeled with digoxigenin [DIG] with lifA1-lifA2 and lifA3-lifA4 primers [14] using the PCR DIG probe synthesis kit [Roche Diagnostics, Mannheim, Germany]; DIG Easy Hyb solution [Roche] was used for prehybridization and hybridization.) Positive results with both DNA probes were obtained for 52 of 56 E. coli O26:H11 isolates. A positive signal was only found in three isolates with the lifA1-lifA2 DNA probe and in one isolate with the lifA3-lifA4 probe. An efa1 gene was not detected in the O26:H32 isolate (Table (Table11).We also analyzed the spatial and temporal behavior of the O26:H11/H32 isolates in the beef herds by cluster analysis (conducted in PAUP* for Windows version 4.0, 2008 [http://paup.csit.fsu.edu/about.html]). This was performed with distance matrices using the neighbor-joining algorithm, an agglomerative cluster method which generates a phylogenetic tree. The distance matrices were calculated by pairwise comparisons of the fragmentation patterns produced by genomic typing through pulsed-field gel electrophoresis analysis with four restriction endonucleases (XbaI, NotI, BlnI, and SpeI) and the presence or absence of potential virulence markers (Fig. (Fig.11 and Table Table1).1). To this end, the total character difference was used, which counts the pairwise differences between two given patterns. During a monitoring program of 3 years in four cattle farms (29), different O26:H11 cluster groups and one O26:H32 isolate were detected in three different farms. The genetic distance of the O26:H32 isolate was very high relative to the O26:H11 isolates. Therefore, the O26:H32 isolate was outgrouped. The O26:H11 isolates of each farm represented independent cluster groups. The single isolate from farm D fitted better to the isolates from farm C than to those from farm A. This finding is in accord with the geographical distance between the farms. The fact that the farms were located in neighboring villages may suggest that direct or indirect connections between the farms were possible (e.g., by person contacts or animal trade). However, the isolates from farm C and farm D belonged to different sequence types (ST 21 and ST 396), which may argue against a direct connection. Interestingly, O26:H11 isolates with and without stx2 genes were detected in the same clusters. This phenomenon was observed in both farm A and farm C. In farm A, the isolates with additional stx2 genes were found in animal 27 and were grouped in clusters 8 and 9 (day 218). An stx2 gene was repeatedly found (four isolates) in the same animal (animal 27). The isolates grouped in cluster 8 on a later day of sampling (day 478). All other O26:H11 isolates grouped in the same clusters and obtained from the same animals (27 and 29) on different sampling days lacked an stx2 gene. Also, the isolates obtained from animal 27 on previous sampling days, which grouped in clusters 3 and 5, exhibited no stx2 genes. In farm C, the three isolates with additional stx2 genes obtained from animal 7 grouped in clusters 11 and 12. An stx2 gene was absent from all other O26:H11 isolates grouped in the same cluster 12 on later sampling days, and no other isolates of cluster 11 were found later on. However, we detected members of many clusters over relatively long periods (clusters 5, 8, and 9 in farm A and cluster 12 in farm C), but members of other clusters were only found on single occasions. This patchy temporal pattern is apparently not a unique property of O26:H11, as we found similar results for cluster groups of other EHEC serotypes of bovine origin (28). The isolates grouped in the dominant cluster 8 were found on 5 of 9 sampling days over a period of 10 months. In contrast, we found the members of clusters 4, 5, 9, and 12 only on two nonconsecutive sampling days. The period during which isolates of these groups were not detected was particularly long for cluster 4 (231 days). We also observed the coexistence of different clusters over long periods in the same farm and in the same cattle (clusters 8 and 9), while one of the clusters dominated. Transmission of clusters between cattle was also observed. These results suggest that some of the EHEC O26:H11 strains had the potential for a longer persistence in the host population, while others had not. The reasons for this difference are not yet clear. Perhaps the incomplete efa1 gene found in isolates of clusters which were only detected once might explain why some strains disappeared rapidly. Efa1 has been discussed as a potential E. coli colonization factor for the bovine intestine used by non-O157 STEC, including O26 (54, 56). The O165:H25 cluster detected during a longer period in farm B may have disappeared after it had lost its efa1 gene (28). The precise biological activity of Efa1 in EHEC O26 is not yet known, but it has been demonstrated that the molecule is a non-Stx virulence determinant which can increase the virulence of EHEC O26 in humans (8).Open in a separate windowFIG. 1.Neighbor-joining tree of bovine E. coli O26:H11/H32 strains based on the restriction pattern obtained after digestion with XbaI, NotI, BlnI, and SpeI.We distinguished 12 different clusters, but complete genetic identity was only found in two isolates. The variations in the O26:H11 clusters may be due to increasing competition between the bacterial populations of the various subtypes in the bovine intestine or to potential interactions between EHEC O26:H11 and the host.The ephemeral occurrence of additional stx2 genes in different clusters and farms may be the result of recombination events due to horizontal gene transfer (16). The loss of stx genes may occur rapidly in the course of an infection, but the reincorporation by induction of an stx-carrying bacteriophage into the O26:H11 strains is possible at any time (9, 40). Nevertheless, an additional stx2 gene may increase the dangerousness of the respective EHEC O26:H11 strains. While all patients involved in an outbreak caused by an EHEC O26:H11 strain harboring the gene encoding Stx2 developed HUS (41), the persons affected by another outbreak caused by an EHEC O26:H11 strain that produced exclusively Stx1 had only uncomplicated diarrhea (60).In conclusion, our results showed that bovine O26:H11 isolates can carry virulence factors of EHEC that are strongly associated with EHEC-related disease in humans, particularly with severe clinical manifestations such as hemorrhagic colitis and HUS. Therefore, strains of bovine origin may represent a considerable risk for human infection. Moreover, some clusters of EHEC O26:H11 persisted in cattle and farms over longer periods, which may increase the risk of transmission to other animals and humans even further.  相似文献   

6.
7.
A real-time quantitative PCR-based detection assay targeting the dnaJ gene (encoding heat shock protein 40) of the coral pathogen Vibrio coralliilyticus was developed. The assay is sensitive, detecting as little as 1 CFU per ml in seawater and 104 CFU per cm2 of coral tissue. Moreover, inhibition by DNA and cells derived from bacteria other than V. coralliilyticus was minimal. This assay represents a novel approach to coral disease diagnosis that will advance the field of coral disease research.Vibrio coralliilyticus has recently emerged as a coral pathogen of concern on reefs throughout the Indo-Pacific. It was first implicated as the etiological agent responsible for bleaching and tissue lysis of the coral Pocillopora damicornis on Zanzibar reefs (2). More recently, V. coralliilyticus has been identified as the causative agent of white syndrome (WS) outbreaks on several Pacific reefs (14). WS is a collective term describing coral diseases characterized by a spreading band of tissue loss exposing white skeleton on Indo-Pacific scleractinian corals (16). V. coralliilyticus is an emerging model pathogen for understanding the mechanisms linking bacterial infection and coral disease (13) and therefore provides an ideal model for the development of diagnostic assays to detect coral disease. Current coral disease diagnostic methods, which are based primarily upon field-based observations of macroscopic disease signs, often detect disease only at the latest stages of infection, when control measures are least effective. The development of diagnostic tools targeting pathogens underlying coral disease pathologies may provide early indications of infection, aid the identification of disease vectors and reservoirs, and assist managers in developing strategies to prevent the spread of coral disease outbreaks. In this paper, we describe the development and validation of a TaqMan-based real-time quantitative PCR (qPCR) assay that targets a segment of the V. coralliilyticus heat shock protein 40-encoding gene (dnaJ).Nucleotide sequences of the dnaJ gene were retrieved from relevant Vibrio species, including V. coralliilyticus (LMG 20984), using the National Center for Biotechnology Information''s (NCBI) Entrez Nucleotide Database search tool (http://www.ncbi.nlm.nih.gov/). Gene sequences of strains not available in public databases (V. coralliilyticus strains LMG 21348, LMG 21349, LMG 21350, LMG 10953, LMG 20538, LMG 23696, LMG 23691, LMG 23693, LMG 23692, and LMG 23694) were obtained through extraction of total DNA using a Promega Wizard Prep DNA Purification Kit (Promega, Sydney, Australia), PCR amplification, and sequencing using primers and thermal cycling parameters described by Nhung et al. (8). A 128-bp region (nucleotides 363 to 490) containing high concentrations of single nucleotide polymorphisms (SNPs), which were conserved within V. coralliilyticus strains but differed from non-V. coralliilyticus strains, was identified, and oligonucleotide primers Vc_dnaJ_F1 (5′-CGG TTC GYG GTG TTT CAA AA-3′) and Vc_dnaJ_R1 (5′-AAC CTG ACC ATG ACC GTG ACA-3′) and a TaqMan probe, Vc_dnaJ_TMP (5′-6-FAM-CAG TGG CGC GAA G-MGBNFQ-3′; 6-FAM is 6-carboxyfluorescein and MGBNFQ is molecular groove binding nonfluorescent quencher), were designed to target this region. The qPCR assay was optimized and validated using DNA extracted from V. coralliilyticus isolates, nontarget Vibrio species, and other bacterial species grown in marine broth (MB) (Table (Table1),1), under the following optimal conditions: 1× TaqMan buffer A, 0.5 U of AmpliTaq Gold DNA polymerase, 200 μM deoxynucleotide triphosphates (with 400 μM dUTP replacing deoxythymidine triphosphate), 0.2 U of AmpErase uracil N-glycosylase (UNG), 3 mM MgCl2, 0.6 μM each primer, 0.2 μM fluorophore-labeled TaqMan, 1 μl of template, and sterile MilliQ water for a total reaction volume to 20 μl. All assays were conducted on a RotoGene 300 (Corbett Research, Sydney, Australia) real-time analyzer with the following cycling parameters: 50°C for 120 s (UNG activation) and 95°C for 10 min (AmpliTaq Gold DNA polymerase activation), followed by 40 cycles of 95°C for 15 s (denaturation) and 60°C for 60 s (annealing/extension). During the annealing/extension phase of each thermal cycle, fluorescence was measured in the FAM channel (470-nm excitation and 510-nm detection).

TABLE 1.

Species, strain, and threshold cycle for all bacterial strains testeda
SpeciesStrainbOriginHost organismCT ± SEMcdnaJ gene sequence accession no.Reference
Vibrio coralliilyticusLMG 23696Nelly Bay, Magnetic Island, AustraliaMontipora aequituberculata12.43 ± 0.20HM21557014
LMG 23691Majuro Atoll, Republic of Marshall IslandsAcropora cytherea14.07 ± 1.33HM21557114
LMG 23693Nikko Bay, PalauPachyseris speciosa10.83 ± 2.76HM21557214
LMG 23692Nikko Bay, PalauPachyseris speciosa9.40 ± 0.36HM21557314
LMG 23694Nikko Bay, PalauPachyseris speciosad12.54 ± 0.24HM21557414
LMG 20984TIndian Ocean, Zanzibar, TanzaniaPocillopora damicornis12.80 ± 0.71HM2155752
LMG 21348Red Sea, Eilat, IsraelPocillopora damicornis13.81 ± 0.49HM2155763
LMG 21349Red Sea, Eilat,Pocillopora damicornis12.98 ± 0.94HM2155773
LMG 21350Red Sea, Eilat,Pocillopora damicornis11.49 ± 0.19HM2155783
LMG 10953Kent, United KingdomCrassostrea gigas (oyster) larvae10.53 ± 0.40HM2155793
LMG 20538Atlantic Ocean, Florianópolis, BrazilNodipecten nodosus (bivalve) larvae12.13 ± 0.50HM2155803
C1Caribbean Sea, La Parguera, Puerto RicoPseudopterogorgia americana14.53 ± 0.28HM21556815
C2Caribbean Sea, La Parguera, Puerto RicoPseudopterogorgia americanaNAHM21556915
Vibrio alginolyticusATCC 1774933.74 ± 0.33
Vibio brasiliensisDSM 1718437.84†
Vibrio calviensisDSM 1434727.06 ± 0.52
Vibrio campbelliiATCC 25920T39.10†
Enterovibrio campbelliiLMG 2136337.33 ± 2.41
Alliivibrio fischeriDSM 50731.36 ± 1.42
Vibrio fortisDSM 19133NA
Vibrio furnissiiDSM 19622NA
Vibrio harveyiDSM 19623NA
Vibrio natriegensATCC 1404828.56 ± 0.60
Vibrio neptuniusLMG 20536NA
Vibrio ordaliiATCC 3350925.56 ± 0.41
Vibrio parahaemolyticusATCC 17802NA
Vibrio proteolyticusATCC 1533830.00 ± 0.89††
Vibrio rotiferianusLMG 21460NA
Vibrio splendidusATCC 3312532.31 ± 0.82
Vibrio tubiashiiATCC 19109NA
Vibrio xuiiLMG 21346NA
Escherichia coliATCC 25922NA
Psychrobacter sp.AIMS 1618NA
Shewanella sp.AIMS C04125.34 ± 0.45
Open in a separate windowaOrigin, host organism, and dnaJ gene sequence accession numbers are shown for V. coralliilyticus strains.bStrain designations beginning with LMG were derived from the Belgian Coordinated Collections of Microorganisms, ATCC strains are from the American Type Culture Collection, DSM strains are from the Deutsche Sammlung von Mikroorganismen und Zellkulturen GmbH culture collection, AIMS strains are from the Australian Institute of Marine Science culture collection, and C1 and C2 were provided by Pamela Morris.c†, amplification in one of three reactions; ††, amplification in two of three reactions; NA, no amplification.dIsolated from seawater above coral.The qPCR assay specifically detected 12 out of 13 isolated V. coralliilyticus strains tested in this study (Table (Table1).1). The exception was one Caribbean strain (C2), which failed to give specific amplification despite repeated attempts. Positive detection of the target gene segment was determined by the increase in fluorescent signal beyond the fluorescence threshold value (normalized fluorescence, 0.010) at a specific cycle, referred to as the threshold cycle (CT). Specific detection was further confirmed by gel electrophoresis, which revealed a PCR product of the correct theoretical size (128 bp) (data not shown), and DNA sequencing, which confirmed the target amplified product to be a segment of the dnaJ gene. No amplification with the assay was detected for 13 other closely related Vibrio strains, including the closely related Vibrio neptunius and two non-Vibrio species (Table (Table1).1). A total of five other Vibrio strains and one non-Vibrio strain (Shewanella sp.) exhibited CT values less than the cutoff of 32 cycles. However, CT values for these strains (mean ± standard error of the mean [SEM], 27.96 ± 2.40) were all much higher than those for V. coralliilyticus strains (12.30 ± 1.52), and no amplicons were evident in post-qPCR gel electrophoresis (data not shown).The detection limit for purified V. coralliilyticus genomic DNA was 0.1 pg of DNA, determined by performing 10-fold serial dilutions (100 ng to 0.01 pg per reaction), followed by qPCR amplification. Similarly, qPCR assays of serial dilutions of V. coralliilyticus (LMG 23696) cells cultured overnight in MB (108 CFU ml−1 to extinction) were able to detect as few as 104 CFU (Fig. (Fig.1).1). Standard curves revealed a strong linear negative correlation between CT values and both DNA and cell concentrations of V. coralliilyticus over several orders of magnitude, with r2 values of 0.998 and 0.953 for DNA and cells, respectively (Fig. (Fig.11).Open in a separate windowFIG. 1.Standard curves delineating threshold (CT) values of fluorescence for indicators of pathogen presence: (A) concentration of V. coralliilyticus DNA and (B) number of V. coralliilyticus cells in pure culture. Error bars indicate standard error of the mean for three replicate qPCRs.Little interference of the qPCR assay was observed when purified V. coralliilyticus (LMG 23696) DNA (10 ng) was combined with 10-fold serial dilutions (0.01 to 100 ng per reaction) of non-V. coralliilyticus DNA (i.e., Vibrio campbellii [ATCC 25920T]). Over the entire range of nontarget DNA concentrations tested, the resulting CT values (mean ± SEM, 17.76 ± 0.53) were not significantly different from those of a control treatment containing 10 ng of V. coralliilyticus DNA and no nonspecific DNA (16.75 ± 0.18; analysis of variance [ANOVA], P = 0.51) (Table (Table2).2). Detection of V. coralliilyticus (LMG 23696) bacterial cells (104, 105, 106, 107, or 108 CFU per ml) in a background of non-V. coralliilyticus cells (i.e., V. campbellii [ATCC 25920T] at 0, 10, 104, or 107 CFU per ml) showed little reduction in assay sensitivity (see Fig. S1 in the supplemental material). For example, when V. coralliilyticus was seeded at 107 cells with similarly high concentrations of nontarget cells, little inhibition of the assay was observed.

TABLE 2.

Effect of nontarget bacterial DNA on the detection of 10 ng of purified V. coralliilyticus DNA
Amt of nontarget DNA (ng)CT (mean ± SEM)
10016.97 ± 0.33
1016.9 ± 0.08
116.74 ± 0.10
0.117 ± 0.09
0.0116.37 ± 0.43
0a16.75 ± 0.18
NTCb35.04 ± 0.02
Open in a separate windowaV. coralliilyticus (LMG 23696) DNA (10 ng) free of nontarget DNA and cells served as positive controls.bA qPCR mixture containing no bacterial DNA served as a no-template, or negative, control (NTC).The assay''s detection limit in seawater was tested by inoculating 10-fold serial dilutions of V. coralliilyticus (LMG 23696) cultures (grown overnight in MB medium, pelleted at 14,000 rpm for 10 min, and washed twice with sterile phosphate-buffered saline [PBS]) into 1 liter of seawater (equivalent final concentrations were 106 to 1 CFU ml−1). The entire volume of V. coralliilyticus-seeded seawater was filtered through a Sterivex-GP filter (Millipore), and DNA was extracted using the method described by Schauer et al. (11). The lowest detection limit for V. coralliilyticus cells seeded into seawater was 1 CFU ml−1 (Fig. (Fig.2),2), with no detection in a 1-liter volume of an unseeded seawater negative control. Standard curves revealed a strong correlation between CT values and the concentrations of V. coralliilyticus bacteria seeded into the seawater over several orders of magnitude (r2 of 0.968) (Fig. (Fig.22).Open in a separate windowFIG. 2.Standard curves showing CT values of the fluorescent signal versus the number of V. coralliilyticus cells per ml seawater (▿), and cells per cm2 of M. aequituberculata tissue, with (○) or without (·) enrichment. Each dot represents an independent experiment. Error bars indicate standard error of the mean for three replicate qPCR runs.The detection limit in seeded coral tissue homogenate was determined by seeding 10-fold dilutions (1010 to 103 CFU ml−1) of pelleted, PBS-washed and resuspended (in 10 ml of sterile PBS) V. coralliilyticus cells onto healthy fragments (∼10 cm2) of the coral Montipora aequituberculata collected from Nelly Bay (Magnetic Island, Australia). Corals were collected in March 2009 and maintained in holding tanks supplied with flowthrough ambient seawater. Resuspended cells were inoculated onto M. aequituberculata fragments, each contained in an individual 3.8-liter plastic bag, allowed to sit at room temperature for 30 min, and then air brushed with compressed air until only white skeleton remained. One-milliliter aliquots of the resulting slurry (PBS, bacteria, and coral tissue) was vortexed for 10 min at 14,000 rpm, and DNA was extracted using a PowerPlant DNA Isolation Kit (Mo Bio, Carlsbad, CA). The lowest detection limits for V. coralliilyticus cells seeded onto coral fragments was 104 CFU per cm2 of coral tissue (Fig. (Fig.2).2). Again, standard curves revealed a strong correlation between CT values and the concentrations of seeded bacteria over several orders of magnitude (r2 of 0.981) (Fig. (Fig.2).2). When a 1-ml aliquot of the slurry was also inoculated into 25 ml of MB and enriched for 6 h at 28°C (with shaking at 170 rpm), the detection limit increased by 1 order of magnitude, to 103 CFU of V. coralliilyticus per cm2 of coral tissue (Fig. (Fig.2).2). The slope of the standard curve reveals some inhibition, particularly at the highest V. coralliilyticus concentrations, which could result from lower replication rates in the cultures with the highest bacterial densities (i.e., 109 CFU). However, since this effect is most pronounced only at the highest bacterial concentrations, the detection limit is still valid. In all trials, unseeded coral fragments and enrichment cultures derived from uninoculated coral fragments served as negative controls.The current study describes the first assay developed to detect and quantify a coral pathogen using a real-time quantitative PCR (qPCR) approach. While previous studies have utilized antibodies or fluorescent in situ hybridization (FISH) to detect coral pathogens (1, 6), the combination of high sensitivity and specificity, low contamination risk, and ease and speed of performance (5) make qPCR technology an ideal choice for rapid pathogen detection in complex hosts, such as corals. The assay developed is highly sensitive for V. coralliilyticus, detecting as few as 1 CFU ml−1 of seawater and 104 CFU cm−2 of coral tissue (103 CFU cm−2 of coral tissue with a 6-h enrichment). These detection limits are likely to be within biologically relevant pathogen concentrations. For example, antibodies for specific detection of the coral bleaching pathogen Vibrio shiloi showed that bacterial densities reached 8.4 × 108 cells cm−3 1 month prior to maximum visual bleaching signs on the coral Oculina patagonica (6). Each seeded seawater and coral (enriched and nonenriched) dilution assay was performed in triplicate. The linearity of the resulting standard curves indicates consistent extraction efficiencies over V. coralliilyticus concentrations spanning 6 orders of magnitude (Fig. (Fig.2)2) and provides strong support for the robustness of the assay. In addition, the presence of competing, non-V. coralliilyticus bacterial cells and DNA had a minimal impact on the detection of V. coralliilyticus. This is an important consideration for accurate detection within the complex coral holobiont, where the target organism is present within a matrix of other microbial and host cells.V. coralliilyticus, like V. shiloi (10), is becoming a model pathogen for the study of coral disease. Recent research efforts have characterized the organism''s genome (W. R. Johnson et al., submitted for publication), proteome (N. E. Kimes et al., submitted for publication), resistome (15), and metabolome (4) and enhanced our understanding of the genetic (7, 9) and physiological (7, 13) basis of its virulence. Before effective management response plans can be formulated, however, continuing research on the genetic and cellular aspects of V. coralliilyticus must be complemented with knowledge of the epidemiology of this pathogen, including information on its distribution, incidence of infection, and rates of transmission throughout populations. The V. coralliilyticus-specific qPCR assay developed in this study will provide important insights into the dynamics of pathogen invasion and spread within populations (6) while also aiding in the identification of disease vectors and reservoirs (12). These capabilities will play an important role in advancing the field of coral disease research and effective management of coral reefs worldwide.   相似文献   

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

TABLE 1.

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

9.
10.
GTP cyclohydrolase I (GCYH-I) is an essential Zn2+-dependent enzyme that catalyzes the first step of the de novo folate biosynthetic pathway in bacteria and plants, the 7-deazapurine biosynthetic pathway in Bacteria and Archaea, and the biopterin pathway in mammals. We recently reported the discovery of a new prokaryotic-specific GCYH-I (GCYH-IB) that displays no sequence identity to the canonical enzyme and is present in ∼25% of bacteria, the majority of which lack the canonical GCYH-I (renamed GCYH-IA). Genomic and genetic analyses indicate that in those organisms possessing both enzymes, e.g., Bacillus subtilis, GCYH-IA and -IB are functionally redundant, but differentially expressed. Whereas GCYH-IA is constitutively expressed, GCYH-IB is expressed only under Zn2+-limiting conditions. These observations are consistent with the hypothesis that GCYH-IB functions to allow folate biosynthesis during Zn2+ starvation. Here, we present biochemical and structural data showing that bacterial GCYH-IB, like GCYH-IA, belongs to the tunneling-fold (T-fold) superfamily. However, the GCYH-IA and -IB enzymes exhibit significant differences in global structure and active-site architecture. While GCYH-IA is a unimodular, homodecameric, Zn2+-dependent enzyme, GCYH-IB is a bimodular, homotetrameric enzyme activated by a variety of divalent cations. The structure of GCYH-IB and the broad metal dependence exhibited by this enzyme further underscore the mechanistic plasticity that is emerging for the T-fold superfamily. Notably, while humans possess the canonical GCYH-IA enzyme, many clinically important human pathogens possess only the GCYH-IB enzyme, suggesting that this enzyme is a potential new molecular target for antibacterial development.The Zn2+-dependent enzyme GTP cyclohydrolase I (GCYH-I; EC 3.5.4.16) is the first enzyme of the de novo tetrahydrofolate (THF) biosynthesis pathway (Fig. (Fig.1)1) (38). THF is an essential cofactor in one-carbon transfer reactions in the synthesis of purines, thymidylate, pantothenate, glycine, serine, and methionine in all kingdoms of life (38), and formylmethionyl-tRNA in bacteria (7). Recently, it has also been shown that GCYH-I is required for the biosynthesis of the 7-deazaguanosine-modified tRNA nucleosides queuosine and archaeosine produced in Bacteria and Archaea (44), respectively, as well as the 7-deazaadenosine metabolites produced in some Streptomyces species (33). GCYH-I is encoded in Escherichia coli by the folE gene (28) and catalyzes the conversion of GTP to 7,8-dihydroneopterin triphosphate (55), a complex reaction that begins with hydrolytic opening of the purine ring at C-8 of GTP to generate an N-formyl intermediate, followed by deformylation and subsequent rearrangement and cyclization of the ribosyl moiety to generate the pterin ring in THF (Fig. (Fig.1).1). Notably, the enzyme is dependent on an essential active-site Zn2+ that serves to activate a water molecule for nucleophilic attack at C-8 in the first step of the reaction (2).Open in a separate windowFIG. 1.Reaction catalyzed by GCYH-I, and metabolic fate of 7,8-dihydroneopterin triphosphate.A homologous GCYH-I is found in mammals and other higher eukaryotes, where it catalyzes the first step of the biopterin (BH4) pathway (Fig. (Fig.1),1), an essential cofactor in the biosynthesis of tyrosine and neurotransmitters, such as serotonin and l-3,4-dihydroxyphenylalanine (3, 52). Recently, a distinct class of GCYH-I enzymes, GCYH-IB (encoded by the folE2 gene), was discovered in microbes (26% of sequenced Bacteria and most Archaea) (12), including several clinically important human pathogens, e.g., Neisseria and Staphylococcus species. Notably, GCYH-IB is absent in eukaryotes.The distribution of folE (gene product renamed GCYH-IA) and folE2 (GCYH-IB) in bacteria is diverse (12). The majority of organisms possess either a folE (65%; e.g., Escherichia coli) or a folE2 (14%; e.g., Neisseria gonorrhoeae) gene. A significant number (12%; e.g., B. subtilis) possess both genes (a subset of 50 bacterial species is shown in Table Table1),1), and 9% lack both genes, although members of the latter group are mainly intracellular or symbiotic bacteria that rely on external sources of folate. The majority of Archaea possess only a folE2 gene, and the encoded GCYH-IB appears to be necessary only for the biosynthesis of the modified tRNA nucleoside archaeosine (44) except in the few halophilic Archaea that are known to synthesize folates, such as Haloferax volcanii, where GCYH-IB is involved in both archaeosine and folate formation (13, 44).

TABLE 1.

Distribution and candidate Zur-dependent regulation of alternative GCYH-I genes in bacteriaa
OrganismcPresence of:
folEfolE2
Enterobacteria
    Escherichia coli+
    Salmonella typhimurium+
    Yersinia pestis+
    Klebsiella pneumoniaeb++a
    Serratia marcescens++a
    Erwinia carotovora+
    Photorhabdus luminescens+
    Proteus mirabilis+
Gammaproteobacteria
    Vibrio cholerae+
    Acinetobacter sp. strain ADP1++a
    Pseudomonas aeruginosa++a
    Pseudomonas entomophila L48++a
    Pseudomonas fluorescens Pf-5++a
    Pseudomonas syringae++a
    Pseudomonas putida++a
    Hahella chejuensis KCTC 2396++a
    Chromohalobacter salexigens DSM 3043++a
    Methylococcus capsulatus++a
    Xanthomonas axonopodis++a
    Xanthomonas campestris++a
    Xylella fastidiosa++a
    Idiomarina loihiensis+
    Colwellia psychrerythraea++
    Pseudoalteromonas atlantica T6c++a
    Pseudoalteromonas haloplanktis TAC125++
    Alteromonas macleodi+
    Nitrosococcus oceani++
    Legionella pneumophila+
    Francisella tularensis+
Betaproteobacteria
    Chromobacterium violaceum+
    Neisseria gonorrhoeae+
    Burkholderia cepacia R18194++
    Burkholderia cenocepacia AU 1054++
    Burkholderia xenovorans+
    Burkholderia mallei+
    Bordetella pertussis+
    Ralstonia eutropha JMP134+
    Ralstonia metallidurans++
    Ralstonia solanacearum+
    Methylobacillus flagellatus+
    Nitrosomonas europaea+
    Azoarcus sp.++
Bacilli/Clostridia
    Bacillus subtilisd++
    Bacillus licheniformis++
    Bacillus cereus+
    Bacillus halodurans++
    Bacillus clausii+
    Geobacillus kaustophilus+
    Oceanobacillus iheyensis+
    Staphylococcus aureus+
Open in a separate windowaGenes that are preceded by candidate Zur binding sites.bZur-regulated cluster is on the virulence plasmid pLVPK.cExamples of organisms with no folE genes are in boldface type.dZn-dependent regulation of B. subtilis folE2 by Zur was experimentally verified (17).Expression of the Bacillus subtilis folE2 gene, yciA, is controlled by the Zn2+-dependent Zur repressor and is upregulated under Zn2+-limiting conditions (17). This led us to propose that the GCYH-IB family utilizes a metal other than Zn2+ to allow growth in Zn2+-limiting environments, a hypothesis strengthened by the observation that an archaeal ortholog from Methanocaldococcus jannaschii has recently been shown to be Fe2+ dependent (22). To test this hypothesis, we investigated the physiological role of GCYH-IB in B. subtilis, an organism that contains both isozymes, as well as the metal dependence of B. subtilis GCYH-IB in vitro. To gain a structural understanding of the metal dependence of GCYH-IB, we determined high-resolution crystal structures of Zn2+- and Mn2+-bound forms of the N. gonorrhoeae ortholog. Notably, although the GCYH-IA and -IB enzymes belong to the tunneling-fold (T-fold) superfamily, there are significant differences in their global and active-site architecture. These studies shed light on the physiological significance of the alternative folate biosynthesis isozymes in bacteria exposed to various metal environments, and offer a structural understanding of the differential metal dependence of GCYH-IA and -IB.  相似文献   

11.
12.
Escherichia coli isolates (72 commensal and 10 O157:H7 isolates) were compared with regard to physiological and growth parameters related to their ability to survive and persist in the gastrointestinal tract and found to be similar. We propose that nonhuman hosts in E. coli O157:H7 strains function similarly to other E. coli strains in regard to attributes relevant to gastrointestinal colonization.Escherichia coli is well known for its ecological versatility (15). A life cycle which includes both gastrointestinal and environmental stages has been stressed by both Savageau (15) and Adamowicz et al. (1). The gastrointestinal stage would be subjected to acid and detergent stress. The environmental stage is implicit in E. coli having transport systems for fungal siderophores (4) as well as pyrroloquinoline quinone-dependent periplasmic glucose utilization (1) because their presence indicates evolution in a location containing fungal siderophores and pyrroloquinoline quinone (1).Since its recognition as a food-borne pathogen, there have been numerous outbreaks of food-borne infection due to E. coli O157:H7, in both ground beef and vegetable crops (6, 13). Cattle are widely considered to be the primary reservoir of E. coli O157:H7 (14), but E. coli O157:H7 does not appear to cause disease in cattle. To what extent is E. coli O157:H7 physiologically unique compared to the other naturally occurring E. coli strains? We feel that the uniqueness of E. coli O157:H7 should be evaluated against a backdrop of other wild-type E. coli strains, and in this regard, we chose the 72-strain ECOR reference collection originally described by Ochman and Selander (10). These strains were chosen from a collection of 2,600 E. coli isolates to provide diversity with regard to host species, geographical distribution, and electromorph profiles at 11 enzyme loci (10).In our study we compared the 72 strains of the ECOR collection against 10 strains of E. coli O157:H7 and six strains of E. coli which had been in laboratory use for many years (Table (Table1).1). The in vitro comparisons were made with regard to factors potentially relevant to the bacteria''s ability to colonize animal guts, i.e., acid tolerance, detergent tolerance, and the presence of the Entner-Doudoroff (ED) pathway (Table (Table2).2). Our longstanding interest in the ED pathway (11) derives in part from work by Paul Cohen''s group (16, 17) showing that the ED pathway is important for E. coli colonization of the mouse large intestine. Growth was assessed by replica plating 88 strains of E. coli under 40 conditions (Table (Table2).2). These included two LB controls (aerobic and anaerobic), 14 for detergent stress (sodium dodecyl sulfate [SDS], hexadecyltrimethylammonium bromide [CTAB], and benzalkonium chloride, both aerobic and anaerobic), 16 for acid stress (pH 6.5, 6.0, 5.0, 4.6, 4.3, 4.2, 4.1, and 4.0), four for the ability to grow in a defined minimal medium (M63 glucose salts with and without thiamine), and four for the presence or absence of a functional ED pathway (M63 with gluconate or glucuronate). All tests were done with duplicate plates in two or three separate trials. The data are available in Tables S1 to S14 in the supplemental material, and they are summarized in Table Table22.

TABLE 1.

E. coli strains used in this study
E. coli strain (n)Source
ECOR strains (72)Thomas Whittman
Laboratory adapted (6)
    K-12 DavisPaul Blum
    CG5C 4401Paul Blum
    K-12 StanfordPaul Blum
    W3110Paul Blum
    BTyler Kokjohn
    AB 1157Tyler Kokjohn
O157:H7 (10)
    FRIK 528Andrew Benson
    ATCC 43895Andrew Benson
    MC 1061Andrew Benson
    C536Tim Cebula
    C503Tim Cebula
    C535Tim Cebula
    ATCC 43889William Cray, Jr.
    ATCC 43890William Cray, Jr.
    ATCC 43888Willaim Cray, Jr.
    ATCC 43894William Cray, Jr.
Open in a separate window

TABLE 2.

Physiological comparison of 88 strains of Escherichia coli
Growth medium or conditionOxygencNo. of strains with type of growthb
ECOR strains (n = 72)
Laboratory strains (n = 6)
O157:H7 strains (n = 10)
GoodPoorNoneVariableGoodPoorNoneVariableGoodPoorNoneVariable
LB controlaBoth72000600010000
1% SDSAerobic6930060008002
5% SDSAerobic6840060008200
1% SDSAnaerobic53154023101702
5% SDSAnaerobic0684004200704
CTABd (all)Both00720006000100
0.05% BACAerobic31158202220091
0.2% BACAerobic01710105000100
0.05% BACAnaerobic2367001500091
0.2% BACAnaerobic00720006000100
pH 6.5Both72000600010000
pH 6Both72000600010000
pH 5Both7020060009001
pH 4.6Both70200600010000
pH 4.3Aerobic14015731203205
pH 4.3Anaerobic6930031201100
pH 4.1 or 4.2Aerobic00720NDgND
pH 4.0Both0072000600091
M63 with supplemente
    GlucoseAerobicf6912050109010
    GlucoseAnaerobicf7002050109010
    GluconateBoth6912050109010
    GlucuronateAerobic6822050109010
    GlucuronateAnaerobic6912050109010
Open in a separate windowaEight LB controls were run, two for each set of LB experiments: SDS, CTAB, benzalkonium chloride (BAC), and pH stress.bGrowth was measured as either +++, +, or 0 (good, poor, and none, respectively), with +++ being the growth achieved on the LB control plates. “Variable” means that two or three replicates did not agree. All experiments were done at 37°C.c“Anaerobic” refers to use of an Oxoid anaerobic chamber. Aerobic and anaerobic growth data are presented together when the results were identical and separately when the results were not the same or the anaerobic set had not been done. LB plates were measured after 1 (aerobic) or 2 (anaerobic) days, and the M63 plates were measured after 2 or 3 days.dCTAB used at 0.05, 0.2%, and 0.4%.eM63 defined medium (3) was supplemented with glucose, gluconate, or glucuronate, all at 0.2%.fIdentical results were obtained with and without 0.0001% thiamine.gND, not determined.  相似文献   

13.
14.
Amino acid modifications of the Thermobifida fusca Cel9A-68 catalytic domain or carbohydrate binding module 3c (CBM3c) were combined to create enzymes with changed amino acids in both domains. Bacterial crystalline cellulose (BC) and swollen cellulose (SWC) assays of the expressed and purified enzymes showed that three combinations resulted in 150% and 200% increased activity, respectively, and also increased synergistic activity with other cellulases. Several other combinations resulted in drastically lowered activity, giving insight into the need for a balance between the binding in the catalytic cleft on either side of the cleavage site, as well as coordination between binding affinity for the catalytic domain and CBM3c. The same combinations of amino acid variants in the whole enzyme, Cel9A-90, did not increase BC or SWC activity but did have higher filter paper (FP) activity at 12% digestion.Cellulases catalyze the breakdown of cellulose into simple sugars that can be fermented to ethanol. The large amount of natural cellulose available is an exciting potential source of fuels and chemicals. However, the detailed molecular mechanisms of crystalline cellulose degradation by glycoside hydrolases are still not well understood and their low efficiency is a major barrier to cellulosic ethanol production.Thermobifida fusca is a filamentous soil bacterium that grows at 50°C in defined medium and can utilize cellulose as its sole carbon source. It is a major degrader of plant cell walls in heated organic materials such as compost piles and rotting hay and produces a set of enzymes that includes six different cellulases, three xylanases, a xyloglucanase, and two CBM33 binding proteins (12). Among them are three endocellulases, Cel9B, Cel6A, and Cel5A (7, 8), two exocellulases, Cel48A and Cel6B (6, 19), and a processive endocellulase, Cel9A (5, 7).T. fusca Cel9A-90 (Uniprot P26221 and YP_290232) is a multidomain enzyme consisting of a family 9 catalytic domain (CD) rigidly attached by a short linker to a family 3c cellulose binding module (CBM3c), followed by a fibronectin III-like domain and a family 2 CBM (CBM2). Cel9A-68 consists of the family 9 CD and CBM3c. The crystal structure of this species (Fig. (Fig.1)1) was determined by X-ray crystallography at 1.9 Å resolution (Protein Data Bank [PDB] code 4tf4) (15). Previous work has shown that E424 is the catalytic acid and D58 is the catalytic base (11, 20). H125 and Y206 were shown to play an important role in activity by forming a hydrogen bonding network with D58, an important supporting residue, D55, and Glc(−1)O1. Several enzymes with amino acid changes in subsites Glc(−1) to Glc(−4) had less than 20% activity on bacterial cellulose (BC) and markedly reduced processivity. It was proposed that these modifications disturb the coordination between product release and the subsequent binding of a cellulose chain into subsites Glc(−1) to Glc(−4) (11). Another variant enzyme with a deletion of a group of amino acids forming a block at the end of the catalytic cleft, Cel9A-68 Δ(T245-L251)R252K (DEL), showed slightly improved filter paper (FP) activity and binding to BC (20).Open in a separate windowFIG. 1.Crystal structure of Cel9A-68 (PDB code 4tf4) showing the locations of the variant residues, catalytic acid E424, catalytic base D58, hydrogen bonding network residues D55, H125, and Y206, and six glucose residues, Glc(−4) to Glc(+2). Part of the linker is visible in dark blue.The CBM3c domain is critical for hydrolysis and processivity. Cel9A-51, an enzyme with the family 9 CD and the linker but without CBM3c, had low activity on carboxymethyl cellulose (CMC), BC, and swollen cellulose (SWC) and showed no processivity (4). The role of CBM3c was investigated by mutagenesis, and one modified enzyme, R557A/E559A, had impaired activity on all of these substrates but normal binding and processivity (11). Variants with changes at five other CBM3c residues were found to slightly lower the activity of the modified enzymes, while Cel9A-68 enzymes containing either F476A, D513A, or I514H were found to have slightly increased binding and processivity (11) (see Table Table1).1). In the present work, CBM3c has been investigated more extensively to identify residues involved in substrate binding and processivity, understand the role of CBM3c more clearly, and study the coordination between the CD and CBM3c. An additional goal was to combine amino acid variants showing increased crystalline cellulose activity to see if this further increased activity. Finally, we have investigated whether the changes that improved the activity of Cel9A-68 also enhanced the activity of intact Cel9A-90.

TABLE 1.

Activities of Cel9A-68 CBM3c variant enzymes and CD variant enzymes used to create the double variants
EnzymeActivity (% of wild type) on:
% Processivity% BC bindingReference
CMCSWCBCFPa
Wild type10010010010010015This work
R378K9891103931392011
DELb981011011289620
F476A97105791001452111
D513A1001151211071192011
I514H104911121041102311
Y520A1087833a79871411
R557A1039860a9390This work
E559A869030a7094This work
R557A+E559A907515a751061511
Q561A1035651a7874This work
R563A977052a931292011
Open in a separate windowaThe target percent digestion could not be reached; activity was calculated using 1.5 μM enzyme.bDEL refers to deletion of T245 to L251 and R252K.  相似文献   

15.
To determine whether and which spirochetes are cleared from Ixodes ricinus ticks during feeding on ruminants, ticks were removed from goats and cattle grazing on tick-infested pastures. Although about a quarter of ticks questing on the pasture were infected by spirochetes, no molted ticks that had previously engorged to repletion on ruminants harbored Lyme disease spirochetes. Borrelia miyamotoi spirochetes, however, appear not to be eliminated. Thus, the more subadult ticks are diverted from reservoir-competent hosts to zooprophylactic ruminants, the smaller the risk of infection by Lyme disease spirochetes is.Various vertebrates serve as reservoir hosts for the tick-borne agents of Lyme disease. A competent reservoir host acquires Lyme disease spirochetes when an infected tick feeds on it and maintains them to become and remain infectious for feeding ticks (10). It appears that each of the seven genospecies of Borrelia burgdorferi sensu lato prevalent in Central Europe is associated with particular reservoir hosts. Whereas rodents serve as a reservoir for B. afzelii and the recently differentiated but not yet validated “Borrelia bavariensis,” birds maintain B. garinii and B. valaisiana (3, 4). B. lusitaniae and B. spielmanii, on the other hand, seem to be limited to lizards and dormice, respectively (9, 12, 13). Ticks harboring rodent-associated spirochetes from their larval blood meal may lose the infectious burden when feeding as nymphs on a bird and vice versa (5). It appears that solely B. burgdorferi sensu stricto constitutes an intermediate position, since it may be perpetuated by birds and rodents (10, 11). As a generalist, B. burgdorferi sensu stricto appears to be less efficiently adapted to rodents than is the specialist B. afzelii. A host that is competent for one genospecies seems less competent or incompetent for another.The Central European vector tick, Ixodes ricinus, not only feeds on small animals. Wild ruminants, such as red, roe, and fallow deer, are frequently infested by all three stages of this tick (2, 6, 15). Interestingly, virtually no spirochetes were detected microscopically in ticks recovered from shot deer. On pastures, where domesticated ruminants graze at an extensive density, spirochetal infection in questing ticks is less prevalent than in nearby nonpastured sites (8). These ruminants appear to exert a zooprophylactic effect. Ruminants, although feeding numerous ticks, appear to be incompetent hosts for Lyme disease spirochetes. It is not known whether the incompetence of ruminants eliminates spirochetes in the feeding tick and whether it extends to each of the Lyme disease genospecies.To determine whether and which spirochetes are cleared from ticks feeding on ruminants, ticks were removed from goats and cattle grazing on tick-infested pastures and examined at various developmental stages for Lyme disease genospecies and B. miyamotoi. Infection rates in ruminant-derived ticks were compared to that in ticks questing on the pastures.The cattle study site was located southwest of the city of Flensburg, Germany, at the German-Danish border. The former training area of the German armed forces is used as low-intensity pasture, covering about 400 ha. Galloway cattle, in herds of mother cows, and Konik horses are allowed to graze year-round and are rotated on grazing patches. Most cattle which were examined for feeding ticks grazed in a 40-ha area which has been pastured since October 2004 and from which cattle and horses are excluded each year from April through June to permit rare plants to bloom and seed. The approximate grazing density of 0.25 livestock units (LU)/ha throughout the rest of the year fails to keep the vegetation short. The goat site was located about 50 km southeast of Stuttgart, Germany, near the village of Gruibingen in the Swabian highlands. Beech and juniper heath characterize the southern-facing mountain slopes, where goats were allowed to graze in a rotating regime during the vegetation period. The sites were in use as pastures for different lengths of time, with the oldest dating back to 2004.To obtain feeding ticks from cattle and goats, two approaches were used. For the yearly blood sampling in the spring, cattle were corralled into squeeze chutes. The head of each animal was examined for ticks. Feeding ticks were carefully removed with forceps, and replete ticks were gently rubbed off onto a sheet of fabric positioned under the cow''s head. From April through October 2006 and 2007 and in May of 2008, tame goats were examined individually for feeding ticks monthly and feeding ticks were removed with forceps. Ticks recovered from an individual animal were confined in screened vials and stored at 22°C to permit molting and/or until they were examined for spirochetes. Questing ticks were collected monthly from April through October 2008 in the cattle site and from April through October 2005 through 2007 at the goat site. They were collected by means of a flannel flag, identified to stage and species by microscopy, and preserved in 80% ethanol. To detect and identify the various spirochetes that may be present in questing or host-derived ticks, DNA from individual ticks was isolated, and a 600-nucleotide fragment of the gene carrying the 16S rRNA gene was amplified by nested PCR and sequenced as described previously (12). This method detects as little as a single spirochete even in the presence of tick and ruminant DNA. Each resulting sequence was compared with sequences of the same gene fragment representing various spirochetal genospecies. The following sequences were used for comparison: GenBank accession numbers X85196 and X85203 for B. burgdorferi sensu stricto, X85190, X85192, and X85194 for B. afzelii, X85193, X85199, and M64311 for B. garinii, X98228 and X98229 for B. lusitaniae, X98232 and X98233 for B. valaisiana, AY147008 for B. spielmanii, and AY253149 for B. miyamotoi. A complete match, permitting no more than two nucleotide changes, was required.Ticks removed while feeding on cattle or goats were examined for spirochetal DNA by nested PCR. Nineteen larvae were obtained during their feeding on goats, but none of the 17 engorged larvae and 2 resulting nymphs contained spirochetal DNA (Table (Table1).1). Of the 416 nymphal ticks that were obtained from 80 goats, only 9% developed to the adult stage, because most of the nymphs were only partially fed. None of the 37 resulting adults contained spirochetal DNA. However, three partially fed nymphal ticks were infected by Lyme disease spirochetes (0.8%), one each by B. afzelii, B. valaisiana, and B. lusitaniae. In three additional nymphs (0.8%), DNA of B. miyamotoi was detected. Of the 415 engorged nymphal ticks obtained from 42 cattle, as many as 319 (77%) molted to the adult stage, because mostly replete ticks had been collected from the cattle''s heads. None of the cattle-derived molted ticks harbored DNA of Lyme disease spirochetes. Four ticks, a nymph and three adults (one male and two females), contained DNA of B. miyamotoi. Of 306 partially engorged females removed from 68 goats, spirochetal DNA was detected in 9 females (2.9%); three harbored B. afzelii, four B. miyamotoi, and one each B. garinii and B. lusitaniae. In addition, 30 females which had fully engorged on cattle were tested for spirochetal DNA after egg-laying. None of these contained spirochetal DNA. Although DNA of Lyme disease spirochetes was detected in a rare partially fed tick, no molted tick that had previously engorged to repletion on a ruminant was infected by Lyme disease spirochetes. In contrast, B. miyamotoi appears to be present in ruminant-fed ticks regardless of their feeding state.

TABLE 1.

Borrelia genospecies detected in I. ricinus ticks that had engorged as larvae, nymphs, or adults on goats or cattle
HostNo. of host animalsTicks
% of infected ticks harboring Borrelia sp.b
StageStateNo. examined% infected% infected by LDa spirocheteafzgarvallusmiy
Goats10LarvaEngorged170.00.0
2NymphMolted20.00.0
80NymphEngorged3791.60.816.70.016.716.750.0
22AdultMolted370.00.0
68AdultEngorged3062.91.633.311.10.011.144.4
Cattle36NymphEngorged961.00.00.00.00.00.0100
42AdultMolted3190.90.00.00.00.00.0100
10AdultEngorged300.00.0
Total144c1,1861.40.723.55.95.911.852.9
Open in a separate windowaLD, Lyme disease.bafz, B. afzelii; gar, B. garinii; val, B. valaisiana; lus, B. lusitaniae; miy, B. miyamotoi.cValue is not the sum of the above numbers because individual hosts were infested by various tick stages.The prevalence of spirochetal infection was determined in questing ticks collected on the pastures on which the cattle or the goats had roamed. A third of the nymphs and nearly a fifth of the adult ticks that quested on the cattle pasture in northern Germany contained spirochetal DNA (Table (Table2).2). The majority of these nymphs and half the infected adults were infected by B. afzelii. About a fifth of the nymphs and a quarter of the adult ticks questing on the goat pastures in southern Germany were infected by spirochetes. B. afzelii and B. lusitaniae infected most of these ticks. Thus, the cattle and goats in the study sites must have been exposed to numerous vector ticks infected by spirochetes.

TABLE 2.

Relative prevalences of Borrelia genospecies in questing nymphal and adult I. ricinus ticks sampled on goat or cattle pastures in Germany
Grazing animals on pastureTicks
% of infected ticks harboring Borrelia sp.a
StageNo. examined% infectedafzgarvalburlusspibismiy>1 genospecies
GoatsNymph55717.241.76.35.24.236.50.00.08.32.1
Adult51125.013.37.06.33.961.70.80.010.23.1
CattleNymph41332.490.30.70.70.00.00.04.56.02.2
Adult6717.950.016.78.30.00.00.016.78.30.0
Open in a separate windowaafz, B. afzelii; gar, B. garinii; val, B. valaisiana; bur, B. burgdorferi; lus, B. lusitaniae; spi, B. spielmanii; bis, B. bissettii-like (1, 7); miy, B. miyamotoi.Ticks infected by Lyme disease spirochetes appear to lose their infection when feeding on goats or cattle. If the blood meal on ruminants had no effect on the spirochetal burden, about 130 and 70 of the analyzed nymphs derived from cattle and goats, respectively, should have contained spirochetal DNA. The two infected cattle-derived ticks harbored solely spirochetes not related to those causing Lyme disease. Most of the ticks removed from goats were partially fed and appeared to be somewhat more likely to contain spirochetal DNA. Whether the detected DNA indicates viable spirochetes is not known. Either feeding on goats fails to eliminate spirochetes as effectively as does a blood meal on cattle or, more plausibly, engorgement to repletion is required for a complete elimination of DNA from Lyme disease spirochetes. If no spirochetal DNA is detected, the tick cannot contain viable spirochetes and thus is not infectious in its host-seeking stage.Wild and domestic ruminants appear to be reservoir incompetent for Lyme disease spirochetes. They do not constitute reservoirs for this pathogen, because no larval tick feeding on them acquires Lyme disease spirochetes. Of 176 engorged I. ricinus larvae or resulting nymphs that had been collected from roe, fallow, red deer, and wild sheep in a Central European site in an earlier study, spirochetes were detected by dark-field microscopy in only two ticks (6). Similarly, only 2 of nearly 200 Ixodes dammini nymphs resulting from larvae that had engorged on white-tailed deer in northeast America contained spirochetes detectable by direct immunofluorescence (15). No spirochetes were detected by phase-contrast microscopy in more than 200 Swedish nymphs that derived from roe-deer-fed larvae (2). Considering that B. miyamotoi morphologically resembles Lyme disease spirochetes, it is likely that all of the spirochetes detected microscopically in ruminant-derived ticks during these earlier studies were not related to B. burgdorferi sensu lato. Not only do larvae fail to acquire Lyme disease spirochetes from ruminants, but infected nymphs also appear to lose their spirochetal load when feeding on these animals, as the present study demonstrates. In ticks that had fully engorged on cattle, the only spirochetal DNA that was detected was that of B. miyamotoi. Also, the American strain of B. miyamotoi was discovered in larvae resulting from field-collected adult females that had routinely been fed on sheep (14). The previous observation that the prevalence of B. miyamotoi in a cattle pasture was not significantly reduced compared to that in the nonpastured site nearby further exemplifies the differential effect of ruminants on these two kinds of spirochetes (8). Whereas Lyme disease spirochetes are eliminated when their tick vector feeds on a ruminant, B. miyamotoi appears not to be affected by such a blood meal.Ruminants reduce the prevalence of infected ticks on a pasture. For the present study, sites were chosen that had only recently come into use as pastures and where cattle were excluded during the peak season of tick activity. The spirochetal prevalence was thus similar to that in the surrounding areas where no domestic ruminants roamed (data not shown) and permitted us to compare infection rates before and after the blood meal on ruminants. The effect of the grazing schedule and of the grazing duration that is required to result in reduced prevalence still needs to be determined. Domestic ruminants employed in landscape management appear to exert their zooprophylactic effect in multiple ways, by eliminating spirochetes from vector ticks feeding on them and by reducing the ecotonal vegetation, thereby limiting coverage and food sources of reservoir hosts while simultaneously rendering the microclimate less suitable for vector ticks. This study''s observations indicate that Lyme disease spirochetes are eliminated from the tick during its blood meal on a ruminant. The mechanism by which Lyme disease spirochetes are cleared during the tick''s blood meal is under investigation. Evidently, Lyme disease spirochetes are destroyed in a way that renders their DNA no longer detectable by means of nested PCR. A simulation model indicates that the availability of incompetent hosts for subadult tick stages would reduce the prevalence of infection (16). Therefore, the more subadult ticks are diverted from reservoir competent birds or mice to incompetent ruminants, the smaller the risk of infection with the agent of Lyme disease is.  相似文献   

16.
Vertebrate genomic assemblies were analyzed for endogenous sequences related to any known viruses with single-stranded DNA genomes. Numerous high-confidence examples related to the Circoviridae and two genera in the family Parvoviridae, the parvoviruses and dependoviruses, were found and were broadly distributed among 31 of the 49 vertebrate species tested. Our analyses indicate that the ages of both virus families may exceed 40 to 50 million years. Shared features of the replication strategies of these viruses may explain the high incidence of the integrations.It has long been appreciated that retroviruses can contribute significantly to the genetic makeup of host organisms. Genes related to certain other viruses with single-stranded RNA genomes, formerly considered to be most unlikely candidates for such contribution, have recently been detected throughout the vertebrate phylogenetic tree (1, 6, 13). Here, we report that viruses with single-stranded DNA (ssDNA) genomes have also contributed to the genetic makeup of many organisms, stretching back as far as the Paleocene period and possibly the late Cretaceous period of evolution.Determining the evolutionary ages of viruses can be problematic, as their mutation rates may be high and their replication may be rapid but also sporadic. To establish a lower age limit for currently circulating ssDNA viruses, we analyzed 49 published vertebrate genomic assemblies for the presence of sequences derived from the NCBI RefSeq database of 2,382 proteins from known viruses in this category, representing a total of 23 classified genera from 7 virus families. Our survey uncovered numerous high-confidence examples of endogenous sequences related to the Circoviridae and to two genera in the family Parvoviridae: the parvoviruses and dependoviruses (Fig. (Fig.11).Open in a separate windowFIG. 1.Phylogenetic tree of vertebrate organisms and history of ssDNA virus integrations. Times of integration of ancestral dependoviruses (yellow icosahedrons), parvoviruses (blue icosahedrons), and circoviruses (triangles) are approximate.The Dependovirus and Parvovirus genomes are typically 4 to 6 kb in length, include 2 major open reading frames (encoding replicase proteins [Rep and NS1, respectively] and capsid proteins [Cap and VP1, respectively]), and have characteristic hairpin structures at both ends (Fig. (Fig.2).2). For replication, these viruses depend on host enzymes that are recruited by the viral replicase proteins to the hairpin regions, where self-primed viral DNA synthesis is initiated (2). Circovirus genomes are typically ∼2-kb circles. DNA of the type species, porcine circovirus 1 (PCV-1), contains a stem-loop structure within the origin of replication (Fig. (Fig.2),2), and the largest open reading frame includes sequences that are homologous to the Parvovirus replicase open reading frame (9, 11). The circoviruses also depend on host enzymes for replication, and DNA synthesis is self-primed from a 3′-OH end formed by endonucleolytic cleavage of the stem-loop structure (4). The frequency of Dependovirus infection is estimated to be as high as 90% within an individual''s lifetime. None of the dependoviruses have been associated with human disease, but related viruses in the family Parvoviridae (e.g., erythrovirus B19 and possibly human bocavirus) are pathogenic for humans, and members of both the Parvoviridae and the Circoviridae can cause a variety of animal diseases (2, 4).Open in a separate windowFIG. 2.Schematics illustrating the structure and organization of Parvoviridae and Circoviridae genomes and origins of several of the longest-integrated ancestral viral sequences found in vertebrates. Integrations were aligned to the Dependovirus adeno-associated virus 2 (AAV2), the Parvovirus minute virus of mice (MVM), and the Circovirus porcine circovirus 1 (PCV-1). The inverted terminal repeat (ITR) sequences in the Dependovirus and Parvovirus genomes are depicted on an expanded scale. A linear representation of the circular genome of PCV-1 is shown with the 10-bp stem-loop structure on an expanded scale. Horizontal lines beneath the maps indicate the lengths of similar sequences that could be identified by BLAST. The numbers indicate the locations of amino acids in the viral proteins where the sequence similarities in the endogenous insertions start and end. The actual ancestral virus-derived integrated sequences may extend beyond the indicated regions.With some ancestral endogenous sequences that we identified, phylogenetic comparisons can be used to estimate age. For example, as a Dependovirus-like sequence is present at the same location in the genomes of mice and rats, the ancestral virus must have existed before their divergence, more than 20 million years ago. Some Circovirus- and Dependovirus-related integrations also predate the split between dog and panda, about 42 million years ago. However, in most other cases, we rely on an indirect method for estimating age (1). As genomic sequences evolve, they accumulate new stop codons and insertion/deletion-induced frameshifts. The rates of these events can be tied directly to the rates of neutral sequence drift and, therefore, the time of evolution. To apply this method, we first performed a BLAST search of vertebrate genomes for all known ssDNA virus proteins (BLAST options, -p tblastn -M BLOSUM62 -e 1e−4). Candidate sequences were then recorded, along with 5 kb of flanking regions, and then again aligned against the database of ssDNA viruses to find the most complete alignment (BLAST options, -t blastx -F F -w 15 -t 1500 -Z 150 -G 13 -E 1 -e 1e−2). Detected alignments were then compared with a neutral model of genome evolution, as described in the supplemental material, and the numbers of stop codons and frameshifts were converted into the expected genomic drift undergone by the sequences. The age of integration was then estimated from the known phylogeny of vertebrates (7, 10). Using these methods, we discovered that as many as 110 ssDNA virus-related sequences have been integrated into the 49 vertebrate genomes considered, during a time period ranging from the present to over 40 to 60 million years ago (Table (Table1;1; see also Tables S1 to S3 in the supplemental material).

TABLE 1.

Selected endogenous sequences in vertebrate genomes related to single-stranded DNA viruses
Virus group and vertebrate speciesInitial genomic search using TBLASTN
Best sequence homology identified using BLASTX
Predicted nucleotide drift (%)Integration labelAge (million yr) or timing of integration based on sequence aging
Chromosomal or scaffold locationProteinBLAST E value/% sequence identityMost similar virusaProteinCoordinatesNo. of stop codons/frameshifts
Circoviruses
    CatScaffold_62068Rep6E−05/37Canary circovirusRep4-2833/7 in 268 aab14.2fcECLG-182
Scaffold_24038Rep6E−06/51Columbid circovirusRep44-3174/5 in 231 aac15.2fcECLG-287
    DogChr5dRep7E−16/46Raven circovirusRep16-2636/5 in 250 aa17.6cfECLG-198
Chr22Rep1E−14/43Beak and feather disease virusRep7-2642/1 in 261 aac4.5cfECLG-254
    OpossumChr3Rep4E−46/44Finch circovirusRep2-2910/2 in 282 aa2.3mdECLG12
Cap6-360/0 in 30 aa
Dependoviruses
    DogChrXRep6E−05/55AAV5Rep239-4453/4 in 200 aa14.0cfEDLG-178
    DolphinGeneScaffold1475Rep8E−39/39Avian AAV DA1Rep79-4863/4 in 379 aac6.6ttEDLG-255
Cap4E−61/47Cap1-7384/7 in 678 aac
    ElephantScaffold_4Rep0/55AAV5Rep3-5890/0 in 579 aa0.0laEDLGRecent
    HyraxGeneScaffold5020Cap3E−34/53AAV3Cap485-7350/5 in 256 aa7.0pcEDLG-129
Scaffold_19252Rep9E−72/47Bovine AAVRep2-3488/4 in 348 aa14.3pcEDLG-260
    MegabatScaffold_5601Rep2E−13/31AAV2Rep315-4791/5 in 175 aa13.1pvEDLG-376
    MicrobatGeneScaffold2026Rep1E−117/50AAV2Rep1-6172/5 in 612 aa5.8mlEDLG-127
Cap9E−33/51Cap1-7312/9 in 509 aac
Scaffold_146492Cap6E−32/42AAV2Cap479-7320/3 in 252 aa4.2mlEDLG-219
    MouseChr1Rep2E−06/34AAV2Rep4-2063/5 in 191 aa17.1mmEDLG-139
Chr3Rep2E−24/31AAV5Rep71-47812/7 in 389 aa16.5mmEDLG-237
Cap2E−22/45Cap22-72412/10 in 649aac
Chr8Rep1E−08/46AAV2Rep314-4733/3 in 147 aa13.8mmEDLG-331
Cap1-1371/2 in 114 aa
    PandaScaffold2359Rep2E−06/37Bovine AAVRep238-4262/3 in 186 aa10.4amEDLG-159
    PikaScaffold_9941Rep4E−14/28AAV5Rep126-4152/2 in 282 aa5.4opEDLG14
    PlatypusChr2Rep9E−10/35Bovine AAVRep297-4374/3 in 138 aa17.1oaEDLG-179
Cap272-4191/2 in 150 aac
Contig12430Rep2E−09/47Bovine AAVRep353-4503/1 in 123 aa12.0oaEDLG-255
Cap2E−05/32Cap253-3672/1 in 116 aa
    RabbitChr10Rep3E−97/39AAV2Rep1-6193/9 in 613 aa9.3ocEDLG43
Cap5E−50/45Cap1-72310/9 in 675 aa
    RatChr13Rep2E−09/33AAV2Rep4-1752/4 in 177 aa13.3rnEDLG-128
Chr2Rep4E−18/40AAV5Rep1-46112/12 in 454 aa22.7rnEDLG-251
Chr19Rep2E−07/33AAV5Rep329-4642/4 in 136 aa16.1rnEDLG-335
Cap31-1332/1 in 93 aa
    TarsierScaffold_178326Rep4E−14/23AAV5Rep96-4652/3 in 356 aa5.3tsEDLG23
Parvoviruses
    Guinea pigScaffold_188Rep3E−24/46Porcine parvovirusRep313-5675/3 in 250 aa12.3cpEPLG-140
Cap1E−16/36Cap10-68911/12 in 672 aa
Scaffold_27Rep1E−50/39Canine parvovirusRep11-6401/4 in 616 aa5.3cpEPLG-217
Cap1E−38/39Porcine parvovirusCap3-7192/14 in 700 aa
    TenrecScaffold_260946Rep2E−20/38LuIII virusRep406-5984/4 in 190 aa19.0etEPLG-260
Cap11-63916/15 in 595 aa
    RatChr5Rep6E−10/56Canine parvovirusRep1-2820/0 in 312 aa0.6rnEPLGRecent
Cap0/62Cap637-6670/2 in 760 aa
Rep0/631-751
    OpossumChr3Rep2E−39/33LuIII virusRep7-57011/3 in 502 aa10.9mdEPLG-256
Cap7E−8/33Cap11-72914/7 in 704 aa
Chr6Rep6E−58/44Porcine parvovirusRep16-5633/7 in 534 aac4.6mdEPLG-324
Cap6E−60/38Cap10-7152/5 in 707 aac
    WallabyScaffold_108040Rep4E−74/62Canine parvovirusRep341-6450/0 in 287 aa1.3meEPLG-37
Cap8E−37/32Cap35-7380/4 in 687 aa
Scaffold_72496Rep2E−61/42Porcine parvovirusRep23-5674/3 in 531 aa5.7meEPLG-630
Cap2E−31/38Cap10-5326/4 in 514 aa
Scaffold_88340Rep7E−37/55Mouse parvovirus 1Rep344-5660/3 in 223 aa6.7meEPLG-1636
Cap7E−22/33Cap11-7136/9 in 700 aa
Open in a separate windowaSome ambiguity in choosing the most similar virus is possible. We generally used the alignment with the lowest E value in the BLAST results. However, one or two points in the exponent of an E value were sometimes sacrificed to achieve a longer sequence alignment.baa, amino acids.cThese sequences have long insertions compared to the present-day viruses. In all cases tested, these insertions originated from short interspersed elements (SINEs). These insertions were excluded from the counts of stop codons and frameshifts and the estimation of integration age.dChr, chromosome.It is important to recognize that there is an intrinsic limit on how far back in time we can reach to identify ancient endogenous viral sequences. First, the sequences must be identified with confidence by BLAST or similar programs. This requirement places a lower limit on sequence identity at about 20 to 30% of amino acids, or about 75% of nucleotides (nucleotides evolve nearly 2.5 times slower than the amino acid sequence they encode). Second, the related, present-day virus must have evolved at a rate that is not much higher than that of the endogenous sequences. The viruses for which ancestral endogenous sequences were identified in this study exhibit sequence drift similar to that associated with mammalian genomes. Setting this rate at 0.14% per million years of evolution (8), we arrive at 90 million years as the theoretical limit for the oldest sequences that can be identified using our methods. This limit drops to less than 35 million years for endogenous viral sequences in rodents and even lower for sequences related to viruses that evolve faster than mammalian genomes.The most widespread integrations found in our survey are derived from the dependoviruses. These include nearly complete genomes related to adeno-associated virus (AAV) in microbat, wallaby, dolphin, rabbit, mouse, and baboon (Fig. (Fig.2).2). We did not detect inverted terminal repeats in several integrations tested, even though repeats are common in the present-day dependoviruses. This result could be explained by sequence decay or the absence of such structures in the ancestral viruses. However, we do see sequences that resemble degraded hairpin structures to which Dependovirus Rep proteins bind, with an example from microbat integration mlEDLG-1 shown in Fig. Fig.3.3. The second most widespread endogenous sequences are related to the parvoviruses. They are found in 6 of 49 vertebrate species considered, with nearly complete genomes in rat, opossum, wallaby, and guinea pig (Fig. (Fig.22).Open in a separate windowFIG. 3.Hairpin structure of the inverted terminal repeat of adeno-associated virus 2 (left) and a candidate degraded hairpin structure located close to the 5′ end of the mlEDLG-1 integration in microbats (right). Structures and mountain plots were generated using default parameters of the RNAfold program (5), with nucleotide coloring representing base-pairing probabilities: blue is below average, green is average, and red is above average. Mountain plots represent hairpin structures based on minimum free energy (mfe) calculations and partition function (pf) calculations, as well as the centroid structure (5). Height is expressed in numbers of nucleotides; position represents nucleotide.The Dependovirus AAV2 has strong bias for integration into human chromosome 19 during infection, driven by a host sequence that is recognized by the viral Rep protein(s). Rep mediates the formation of a synapse between viral and cellular sequences, and the cellular sequences are nicked to serve as an origin of viral replication (14). The related integrations in mice and rats, located in the same chromosomal locations, might be explained by such a mechanism. However, the extent of endogenous sequence decay and the frequency of stop codons indicate that these integrations occurred some 30 to 35 million years ago, implying that they are derived from a single event in a rodent ancestor rather than two independent integration events at the same location. Similarly, integrations EDLG-1 in dog and panda lie in chromosomal regions that can be readily aligned (based on University of California—Santa Cruz [UCSC] genome assemblies) and show sequence decay consistent with the age of the common ancestor, about 42 million years. Endogenous sequences related to the family Parvoviridae can thus be traced to over 40 million years back in time, and viral proteins related to this family have remained over 40% conserved.Sequences related to circoviruses were detected in five vertebrate species (Table (Table11 and Table S1 in the supplemental material). At least one of these sequences, the endogenous sequence in opossum, likely represents a recent integration. Several integrations in dog, cat, and panda, on the other hand, appear to date from at least 42 million years ago, which is the last time when pandas and dogs shared a common ancestor. We see evidence for this age in data from sequence degradation (Table (Table1),1), phylogenetic analyses of endogenous Circovirus-like genomes (see Fig. S2 in the supplemental material), and genomic synteny where integration ECLG-3 is surrounded by genes MTA3 and ARID5A in both dog and panda and integration ECLG-2 lies 35 to 43 kb downstream of gene UPF3A. In fact, Circovirus integrations may even precede the split between dogs and cats, about 55 million years ago, although the preliminary assembly and short genomic contigs for cats make synteny analysis impossible.The most common Circovirus-related sequences detected in vertebrate genomes are derived from the rep gene. We speculate that, like those of the Parvoviridae, the ancestral Circoviridae sequences might have been copied using a primer sequence in the host DNA that resembled the viral origin and was therefore recognized by the virus Rep protein. Higher incidence of rep gene identifications may represent higher conservation of this gene with time, or alternatively, possession of these sequences may impart some selective advantage to the host species. The largest Circovirus-related integration detected, in the opossum, comprises a short fragment of what may have been the cap gene immediately adjacent to and in the opposite orientation from the rep gene. This organization is similar to that of the present day Circovirus genome in which these genes share a promoter in the hairpin regions but are translated in opposite directions (Fig. (Fig.22).In summary, our results indicate that sequences derived from ancestral members of the families Parvoviridae and Circoviridae were integrated into their host''s genomes over the past 50 million years of evolution. Features of their replication strategies suggest mechanisms by which such integrations may have occurred. It is possible that some of the endogenous viral sequences could offer a selective advantage to the virus or the host. We note that rep open reading frame-derived proteins from some members of these families kill tumor cells selectively (3, 12). The genomic “fossils” we have discovered provide a unique glimpse into virus evolution but can give us only a lower estimate of the actual ages of these families. However, numerous recent integrations suggest that their germ line transfer has been continuing into present times.   相似文献   

17.
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   

18.
The effects of the challenge dose and major histocompatibility complex (MHC) class IB alleles were analyzed in 112 Mauritian cynomolgus monkeys vaccinated (n = 67) or not vaccinated (n = 45) with Tat and challenged with simian/human immunodeficiency virus (SHIV) 89.6Pcy243. In the controls, the challenge dose (10 to 20 50% monkey infectious doses [MID50]) or MHC did not affect susceptibility to infection, peak viral load, or acute CD4 T-cell loss, whereas in the chronic phase of infection, the H1 haplotype correlated with a high viral load (P = 0.0280) and CD4 loss (P = 0.0343). Vaccination reduced the rate of infection acquisition at 10 MID50 (P < 0.0001), and contained acute CD4 loss at 15 MID50 (P = 0.0099). Haplotypes H2 and H6 were correlated with increased susceptibility (P = 0.0199) and resistance (P = 0.0087) to infection, respectively. Vaccination also contained CD4 depletion (P = 0.0391) during chronic infection, independently of the challenge dose or haplotype.Advances in typing of the major histocompatibility complex (MHC) of Mauritian cynomolgus macaques (14, 20, 26) have provided the opportunity to address the influence of host factors on vaccine studies (13). Retrospective analysis of 22 macaques vaccinated with Tat or a Tat-expressing adenoviral vector revealed that monkeys with the H6 or H3 MHC class IB haplotype were overrepresented among aviremic or controller animals, whereas macaques with the H2 or H5 haplotype clustered in the noncontrollers (12). More recently, the H6 haplotype was reported to correlate with control of chronic infection with simian immunodeficiency virus (SIV) mac251, regardless of vaccination (18).Here, we performed a retrospective analysis of 112 Mauritian cynomolgus macaques, which included the 22 animals studied previously (12), to evaluate the impact of the challenge dose and class IB haplotype on the acquisition and severity of simian/human immunodeficiency virus (SHIV) 89.6Pcy243 infection in 45 control monkeys and 67 monkeys vaccinated with Tat from different protocols (Table (Table11).

TABLE 1.

Summary of treatment, challenge dose, and outcome of infection in cynomolgus monkeys
Protocol codeNo. of monkeysImmunogen (dose)aAdjuvantbSchedule of immunization (wk)RoutecChallenged (MID50)Virological outcomee
Reference(s) or source
ACV
ISS-ST6Tat (10)Alum or RIBI0, 2, 6, 12, 15, 21, 28, 32, 36s.c., i.m.104114, 17
ISS-ST1Tat (6)None0, 5, 12, 17, 22, 27, 32, 38, 42, 48i.d.101004, 17
ISS-PCV3pCV-tat (1 mg)Bupivacaine + methylparaben0, 2, 6, 11, 15, 21, 28, 32, 36i.m.103006
ISS-ID3Tat (6)none0, 4, 8, 12, 16, 20, 24, 28, 39, 43, 60i.d.10111B. Ensoli, unpublished data
ISS-TR6Tat (10)Alum-Iscom0, 2, 6, 11, 16, 21, 28, 32, 36s.c., i.d., i.m.10420Ensoli, unpublished
ISS-TGf3Tat (10)Alum0, 4, 12, 22s.c.1503Ensoli, unpublished
ISS-TG3Tatcys22 (10)Alum1503Ensoli, unpublished
ISS-TG4Tatcys22 (10) + Gag (60)Alum1504Ensoli, unpublished
ISS-TG4Tat (10) + Gag (60)Alum1504Ensoli, unpublished
ISS-MP3Tat (10)H1D-Alum0, 4, 12, 18, 21, 38s.c., i.m.15021Ensoli, unpublished
ISS-MP3Tat (10)Alums.c.15003Ensoli, unpublished
ISS-GS6Tat (10)H1D-Alum0, 4, 12, 18, 21, 36s.c., i.m.15132Ensoli, unpublished
NCI-Ad-tat/Tat7Ad-tat (5 × 108 PFU), Tat (10)Alum0, 12, 24, 36i.n., i.t., s.c.15232Ensoli, unpublished
NCI-Tat9Tat (6 and 10)Alum/Iscom0, 2, 6, 11, 15, 21, 28, 32, 36s.c., i.d., i.m.1524312
ISS-NPT3pCV-tat (1 mg)Bupivacaine + methylparaben-Iscom0, 2, 8, 13, 17, 22, 28, 46, 71i.m.20003Ensoli, unpublished
ISS-NPT3pCV-tatcys22 (1 mg)Bupivacaine + methylparaben-Iscom0, 2, 8, 13, 17, 22, 28, 46, 71i.m.20111
    Total vaccinated67191731
        Naive11NoneNoneNAgNA10 or 15137
        Control34None, Ad, or pCV-0Alum, RIBI, H1D, Iscom or bupivacaine + methylparaben-Iscoms.c., i.d., i.n., i.t., i.m.10, 15, or 2051316
    Total controls4561623
    Total112253354
Open in a separate windowaAll animals were inoculated with the indicated dose of Tat plasmid DNA (pCV-tat [8], adenovirus-tat [Ad-tat] [27]) or protein, Gag protein, or empty vectors (pCV-0, adenovirus [Ad]) by the indicated route. Doses are in micrograms unless indicated otherwise.bAlum, aluminum phosphate (4); RIBI oil-in-water emulsions containing squalene, bacterial monophosphoryl lipid A, and refined mycobacterial products (4); Iscom, immune-stimulating complex (4); H1D are biocompatible anionic polymeric microparticles used for vaccine delivery (10, 12, 25a).cs.c., subcutaneous; i.m., intramuscular; i.d., intradermal; i.n., intranasal; i.t., intratracheal.dAll animals were inoculated intravenously with the indicated dose of the same SHIV89.6.Pcy243 stock.eAccording to the virological outcome upon challenge, monkeys were grouped as aviremic (A), controllers (C), or viremic (V).fBecause of the short follow-up, controller status could not be determined and all infected monkeys of the ISS-TG protocol were therefore considered viremic.gNA, not applicable.  相似文献   

19.
Specific therapy is not available for hantavirus cardiopulmonary syndrome caused by Andes virus (ANDV). Peptides capable of blocking ANDV infection in vitro were identified using antibodies against ANDV surface glycoproteins Gn and Gc to competitively elute a cyclic nonapeptide-bearing phage display library from purified ANDV particles. Phage was examined for ANDV infection inhibition in vitro, and nonapeptides were synthesized based on the most-potent phage sequences. Three peptides showed levels of viral inhibition which were significantly increased by combination treatment with anti-Gn- and anti-Gc-targeting peptides. These peptides will be valuable tools for further development of both peptide and nonpeptide therapeutic agents.Andes virus (ANDV), an NIAID category A agent linked to hantavirus cardiopulmonary syndrome (HCPS), belongs to the family Bunyaviridae and the genus Hantavirus and is carried by Oligoryzomys longicaudatus rodents (11). HCPS is characterized by pulmonary edema caused by capillary leak, with death often resulting from cardiogenic shock (9, 16). ANDV HCPS has a case fatality rate approaching 40%, and ANDV is the only hantavirus demonstrated to be capable of direct person-to-person transmission (15, 21). There is currently no specific therapy available for treatment of ANDV infection and HCPS.Peptide ligands that target a specific protein surface can have broad applications as therapeutics by blocking specific protein-protein interactions, such as preventing viral engagement of host cell receptors and thus preventing infection. Phage display libraries provide a powerful and inexpensive tool to identify such peptides. Here, we used selection of a cyclic nonapeptide-bearing phage library to identify peptides capable of binding the transmembrane surface glycoproteins of ANDV, Gn and Gc, and blocking infection in vitro.To identify peptide sequences capable of recognizing ANDV, we panned a cysteine-constrained cyclic nonapeptide-bearing phage display library (New England Biolabs) against density gradient-purified, UV-treated ANDV strain CHI-7913 (a gift from Hector Galeno, Santiago, Chile) (17, 18). To increase the specificity of the peptides identified, we eluted phage by using monoclonal antibodies (Austral Biologicals) prepared against recombinant fragments of ANDV Gn (residues 1 to 353) or Gc (residues 182 to 491) glycoproteins (antibodies 6B9/F5 and 6C5/D12, respectively). Peptide sequences were determined for phage from iterative rounds of panning, and the ability of phage to inhibit ANDV infection of Vero E6 cells was determined by immunofluorescent assay (IFA) (7). Primary IFA detection antibodies were rabbit polyclonal anti-Sin Nombre hantavirus (SNV) nucleoprotein (N) antibodies which exhibit potent cross-reactivity against other hantavirus N antigens (3). ReoPro, a commercially available Fab fragment which partially blocks infection of hantaviruses in vitro by binding the entry receptor integrin β3 (5), was used as a positive control (80 μg/ml) along with the original antibody used for phage elution (5 μg/ml). As the maximum effectiveness of ReoPro in inhibiting hantavirus entry approaches 80%, we set this as a threshold for maximal expected efficacy for normalization. The most-potent phage identified by elution with the anti-Gn antibody 6B9/F5 bore the peptide CPSNVNNIC and inhibited hantavirus entry by greater than 60% (61%) (Table (Table1).1). From phage eluted with the anti-Gc antibody 6C5/D12, those bearing peptides CPMSQNPTC and CPKLHPGGC also inhibited entry by greater than 60% (66% and 72%, respectively).

TABLE 1.

Peptide-bearing phage eluted from ANDV
Phage% Inhibition (SD)aP valueb
Phage bearing the following peptides eluted with anti-Gn antibody 6B9/F5
    Group 1 (<30% inhibition)
        CDQRTTRLC8.45 (15.34)0.0002
        CPHDPNHPC9.94 (7.72)0.333
        CQSQTRNHC11.76 (13.25)0.0001
        CLQDMRQFC13.26 (9.92)0.0014
        CLPTDPIQC15.70 (14.05)0.0005
        CPDHPFLRC16.65 (15.22)0.8523
        CSTRAENQC17.56 (16.50)0.0004
        CPSHLDAFC18.98 (20.06)0.0017
        CKTGHMRIC20.84 (7.47)0.0563
        CVRTPTHHC20.89 (27.07)0.1483
        CSGVINTTC21.57 (19.61)0.0643
        CPLASTRTC21.65 (5.98)0.004
        CSQFPPRLC22.19 (8.26)0.0004
        CLLNKQNAC22.34 (7.78)0.001
        CKFPLNAAC22.89 (6.15)0.0001
        CSLTPHRSC23.63 (16.74)0.0563
        CKPWPMYSC23.71 (6.68)0.0643
        CLQHDALNC24.01 (7.60)1
        CNANKPKMC24.67 (11.67)0.0004
        CPKHVLKVC25.30 (28.36)0.0003
        CTPDKKSFC26.91 (11.15)0.399
        CHGKAALAC27.22 (32.53)0.005
        CNLMGNPHC28.08 (21.35)0.0011
        CLKNWFQPC28.64 (18.49)0.0016
        CKEYGRQMC28.76 (29.33)0.0362
        CQPSDPHLC29.44 (31.22)0.0183
        CSHLPPNRC29.70 (17.37)0.0061
    Group 2 (30-59% inhibition)
        CSPLLRTVC33.05 (20.26)0.0023
        CHKGHTWNC34.17 (12.50)0.0795
        CINASHAHC35.62 (13.03)0.3193
        CWPPSSRTC36.75 (26.95)0.0006
        CPSSPFNHC37.78 (7.11)0.0001
        CEHLSHAAC38.47 (7.60)0.0115
        CQDRKTSQC38.74 (9.12)0.1802
        CTDVYRPTC38.90 (25.03)0.006
        CGEKSAQLC39.11 (27.52)0.0013
        CSAAERLNC40.13 (6.33)0.0033
        CFRTLEHLC42.07 (5.01)0.0608
        CEKLHTASC43.60 (27.92)0.1684
        CSLHSHKGC45.11 (49.81)0.0864
        CNSHSPVHC45.40 (28.80)0.0115
        CMQSAAAHC48.88 (44.40)0.5794
        CPAASHPRC51.84 (17.09)0.1935
        CKSLGSSQC53.90 (13.34)0.0145
    Group 3 (60-79% inhibition)
        CPSNVNNIC61.11 (25.41)0.1245
Negative control0 (6.15)
6B9/F5 (5 μg/ml)26.77 (5.33)
ReoPro (80 μg/ml)79.86 (4.88)
Phage bearing the following peptides eluted with anti-Gc antibody 6C5/D12
    Group 1 (<30% inhibition)
        CHPGSSSRC1.01 (7.03)0.0557
        CSLSPLGRC10.56 (13.62)0.7895
        CTARYTQHC12.86 (3.83)0.3193
        CHGVYALHC12.91 (7.32)0.0003
        CLQHNEREC16.79 (13.72)0.0958
        CHPSTHRYC17.23 (14.53)0.0011
        CPGNWWSTC19.34(9.91)0.1483
        CGMLNWNRC19.48 (19.42)0.0777
        CPHTQFWQC20.44 (13.65)0.0008
        CTPTMHNHC20.92 (11.68)0.0001
        CDQVAGYSC21.79 (23.60)0.0063
        CIPMMTEFC24.33 (9.28)0.2999
        CERPYSRLC24.38 (9.09)0.0041
        CPSLHTREC25.06 (22.78)0.1202
        CSPLQIPYC26.30 (34.29)0.4673
        CTTMTRMTC (×2)29.27 (8.65)0.0001
    Group 2 (30-59% inhibition)
        CNKPFSLPC30.09 (5.59)0.4384
        CHNLESGTC31.63 (26.67)0.751
        CNSVPPYQC31.96 (6.51)0.0903
        CSDSWLPRC32.95 (28.54)0.259
        CSAPFTKSC33.40 (10.64)0.0052
        CEGLPNIDC35.63 (19.90)0.0853
        CTSTHTKTC36.28 (13.42)0.132
        CLSIHSSVC36.40 (16.44)0.8981
        CPWSTQYAC36.81 (32.81)0.5725
        CTGSNLPIC36.83 (31.64)0.0307
        CSLAPANTC39.73 (4.03)0.1664
        CGLKTNPAC39.75 (16.98)0.2084
        CRDTTPWWC40.08 (18.52)0.0004
        CHTNASPHC40.26 (4.77)0.5904
        CTSMAYHHC41.89 (8.61)0.259
        CSLSSPRIC42.13 (29.75)0.2463
        CVSLEHQNC45.54 (6.55)0.5065
        CRVTQTHTC46.55 (8.45)0.3676
        CPTTKSNVC49.28 (14.00)0.3898
        CSPGPHRVC49.50 (42.60)0.0115
        CKSTSNVYC51.20 (4.60)0.0611
        CTVGPTRSC57.30 (11.31)0.0176
    Group 3 (60-79% inhibition)
        CPMSQNPTC65.60 (13.49)0.014
        CPKLHPGGC71.88 (27.11)0.0059
Negative control0.26 (4.53)
6C5/D12 (5 μg/ml)22.62 (8.40)
ReoPro (80 μg/ml)80.02 (76.64)
Open in a separate windowaStandard deviations of four experiments are shown in parentheses. Peptide-bearing phage were added at 109 phage/μl.bP values for the pairwise amino acid alignment score of each peptide versus that of integrin β3 were determined using an unpaired Student''s t test. P values considered statistically significant are shown in bold.To determine whether the peptide sequences of any of the identified inhibitory phage showed homology to integrin β3, a known entry receptor for pathogenic hantaviruses (6, 7), we used the Gap program to perform a pairwise amino acid alignment of each peptide versus the extracellular portion of integrin β3 and determined P values for the alignments. Of 45 phage eluted with the anti-Gn antibody, 6B9/F5, 27 of the peptide sequences showed homology to integrin β3 (P < 0.05), and 9 were highly significant (P ≤ 0.0005) (Fig. (Fig.1A).1A). Of the latter, CKFPLNAAC and CSQFPPRLC map to the hybrid domain (Fig. (Fig.1B),1B), which is proximal to the plexin-semaphorin-integrin domain (PSI) containing residue D39, shown to be critical for viral entry in vitro (19). Five sequences (CPSSPFNH, CPKHVLKVC, CNANKPKMC, CQSQTRNHC, and CDQRTTRLC) map to the I-like (or βA) domain near the binding site of ReoPro (2). Finally, CLPTDPIQC maps to the epidermal growth factor 4 (EGF-4) domain, and CSTRAENQC aligns to a portion of β3 untraceable in the crystal structure, specifically the linker region between the hybrid domain and EGF-1. Although this represents a disordered portion of the protein (22), the location of this loop proximal to the PSI domain is worth noting, due to the role of the PSI domain in facilitating viral entry (19). Therefore, 60% of phage eluted with the anti-Gn antibody showed some homology to integrin β3, and those with highly significant P values predominantly mapped to or proximal to regions of known interest in viral entry.Open in a separate windowFIG. 1.Inhibitory peptides identified through phage panning against ANDV show homology to integrin β3. (A) Alignment of phage peptide sequences with P values for integrin β3 pairwise alignment of less than 0.05. Residues comprising the signal peptide, transmembrane, and cytoplasmic domains, which were not included during pairwise alignment, are underlined. Residues 461 to 548, which are missing in the crystal structure, are italicized. Residues involved in the ReoPro binding site are highlighted in green (2). Residue D39 of the PSI domain is highlighted in yellow (19). Peptides are shown above the sequence of integrin β3, with antibody 6C5/D12-eluted sequences shown in blue text and sequences eluted with antibody 6B9/F5 shown in red. Peptide sequences with alignment P values of ≤0.0005 are highlighted in yellow. Percent inhibition of the peptide-bearing phage is shown in parentheses. (B) View of integrin αvβ3 (PDB ID 1U8C [23]). αv is shown in blue ribbon diagram, and β3 is shown in salmon-colored surface representation, with specific domains circled. Residues corresponding to the ReoPro binding site are shown in green, as in panel A, and D39 is shown in yellow. Regions corresponding to 6C5/D12-eluted peptides with P values of ≤0.0005 for alignment with integrin β3 (highlighted in panel A) are shown in blue, and those corresponding to 6B9/F5-eluted peptides with P values of ≤0.0005 for alignment with integrin β3 are shown in red. Alignment of peptide PLASTRT (P value of 0.0040) adjacent to D39 of the PSI domain is shown in magenta. Graphics were prepared using Pymol (DeLano Scientific LLC, San Carlos, CA).Of the 41 peptide-bearing phage eluted with the anti-Gc antibody 6C5/D12, 14 showed sequence homology to integrin β3 (P < 0.05), 4 of which had P values of ≤0.0005 (Fig. (Fig.1A).1A). Of the latter, sequence CTTMTRMTC mapped to the base of the I-like domain (Fig. (Fig.1B),1B), while CHGVYALHC and CRDTTPWWC mapped to the EGF-3 domain. Finally, sequence CTPTMHNHC mapped to the linker region untraceable in the crystal structure. Therefore, in contrast to peptide sequences identified by competition with the anti-Gn antibody, sequences identified by competition with the anti-Gc antibody 6C5/D12 appear to be mostly unrelated to integrin β3.As a low level of pathogenic hantavirus infection can be seen in cells lacking integrin β3, such as CHO cells (19), we asked if any of the identified peptide sequences could represent a previously unidentified receptor. We used the Basic Local Alignment Search Tool to search a current database of human protein sequences for potential alternate receptors represented by these peptides. However, none of the alignments identified proteins that are expressed at the cell surface, eliminating them as potential candidates for alternate viral entry receptors. This suggests that the majority of the peptides identified here likely represent novel sequences for binding ANDV surface glycoproteins.To determine whether synthetic peptides would also block infection, we synthesized cyclic peptides based on the 10 most-potent peptide-bearing phage. These peptides, in the context of phage presentation, showed levels of inhibition ranging from 44 to 72% (Table (Table2).2). When tested by IFA at 1 mM, four of the synthetic peptides showed inhibition levels significantly lower than those of the same peptide presented in the context of phage. This is not surprising, as steric factors due to the size of the phage and the multivalent presentation of peptide in the context of phage may both contribute to infection inhibition (8). However, there was no significant difference in inhibition by synthetic peptide versus peptide-bearing phage for six of the sequences, implying that inhibition in the context of phage was due solely to the nature of the peptide itself and not to steric factors or valency considerations contributed by the phage, which contrasts with our previous results, determined by using phage directed against αvβ3 integrin (10).

TABLE 2.

Synthetic cyclic peptides inhibit ANDV infection
TargetSample% Inhibition bya:
Peptide-bearing phageSynthetic peptide
GnCMQSAAAHC48.88 (44.40)59.66 (11.17)
GcCTVGPTRSC57.30 (11.31)46.47 (7.61)
GnCPSNVNNIC61.11 (25.41)44.14 (10.74)
GnCEKLHTASC43.60 (27.92)34.87 (9.26)
GcCPKLHPGGC71.88 (27.11)30.95 (7.73)b
GnCSLHSHKGC45.11 (49.81)29.79 (9.34)
GcCPMSQNPTC65.60 (13.49)18.19 (8.55)b
GnCKSLGSSQC53.90 (13.34)18.10 (7.55)b
GnCNSHSPVHC45.40 (28.80)15.52 (10.48)
GnCPAASHPRC51.84 (17.09)0 (10.72)b
Integrin β3ReoPro80.10 (7.72)
Gn6B9/F5 antibody42.72 (6.75)
Gc6C5/D12 antibody31.04 (7.81)
Open in a separate windowaStandard deviations of the results of at least four experiments are shown in parentheses.bMean percent inhibition between phage and synthetic peptide differs significantly (P < 0.05).The three most-potent synthetic peptides were examined for their ability to inhibit ANDV entry in a dose-dependent manner. The concentration of each peptide that produces 50% of its maximum potential inhibitory effect was determined. As shown in Fig. Fig.2A,2A, the 50% inhibitory concentration for each of the peptides was in the range of 10 μM, which from our experience is a reasonable potency for a lead compound to take forward for optimization.Open in a separate windowFIG. 2.Activities of synthetic peptides in inhibition of ANDV infection in vitro. (A) Peptides were examined for their ability to block ANDV infection of Vero E6 cells in a dose-dependent manner by IFA. (B) Peptides were tested in parallel for the ability to block infection of Vero E6 cells by ANDV, SNV, HTNV, and PHV. (C) Peptides were tested, singly or in combination, for the ability to block ANDV infection of Vero E6 cells. For all experiments, controls included media, ReoPro at 80 μg/ml, and monoclonal antibodies 6C5/D12 and 6B9/F5 at 5 μg/ml. All peptides were used at 1 mM. Data points represent n = 2 to 6, with error bars showing the standard errors of the means. Statistical analyses were performed on replicate samples using an unpaired Student''s t test.In order to determine the specificity of the three most-potent synthetic cyclic peptides in blocking ANDV, we examined them for inhibition of ANDV infection versus two other pathogenic hantaviruses, SNV and Hantaan virus (HTNV), or the nonpathogenic hantavirus Prospect Hill virus (PHV). As shown in Fig. Fig.2B,2B, ReoPro, which binds integrin β3, showed inhibition of infection by each of the pathogenic hantavirus strains, known to enter cells via β3, but not the nonpathogenic PHV, which enters via integrin β1 (6, 7). In contrast, peptides selected for the ability to bind ANDV were highly specific inhibitors of ANDV versus SNV, HTNV, or PHV. The specificities of peptides eluted by the anti-Gn monoclonal antibody are not surprising, as they are likely due to global differences in the Gn amino acid sequence. Specifically, sequence homologies between ANDV and SNV, HTNV, and PHV are 61%, 36%, and 51%, respectively, for the region corresponding to the immunogen for antibody 6B9/F5. Although homology between the immunogen for antibody 6C5/D12 and the corresponding Gc region of these viruses is somewhat higher (82% with SNV, 63% with HTNV, and 71% with PHV), the possibility that the monoclonal antibody used here recognizes a three-dimensional epitope lends itself to the high specificity of the peptides.The current model for cellular infection by hantaviruses (14) is as follows. Viral binding of the host cell surface target integrin is followed by receptor-mediated endocytosis and endosome acidification. Lowered pH induces conformational changes in Gn and/or Gc, which facilitate membrane fusion and viral release into the cytosol. As there is currently little information available about whether one glycoprotein is dominant in mediating infection, and as neutralizing epitopes have been found on both Gn and Gc glycoproteins (1, 4, 12, 13, 20), we examined whether combining anti-Gn- and anti-Gc-targeted synthetic peptides would lead to an increased infection blockade compared to those for single treatments. As shown in Fig. Fig.2C,2C, the combination of anti-Gn and anti-Gc peptides CMQSAAAHC and CTVGPTRSC resulted in a significant increase in infection inhibition (P = 0.0207 for CMQSAAAHC, and P = 0.0308 for CTVGPTRSC) compared to that resulting from single treatments. Although the high specificity of the peptides for ANDV makes it unlikely that this combination treatment will lead to more cross-reactivity with other pathogenic hantaviruses, this can be determined only by additional testing. Regardless, these data suggest a unique role for each of these viral proteins in the infection process as well as the benefits of targeting multiple viral epitopes for preventing infection.To our knowledge, the peptides reported here are the first identified that directly target ANDV, and this work further illustrates the power of coupling phage display and selective elution techniques in the identification of novel peptide sequences capable of specific protein-protein interactions from a large, random pool of peptide sequences. These novel peptide inhibitors (R. S. Larson, P. R. Hall, H. Njus, and B. Hjelle, U.S. patent application 61/205,211) provide leads for the development of more-potent peptide or nonpeptide organics for therapeutic use against HCPS.  相似文献   

20.
Basal cell carcinoma (BCC) is a very common malignant skin tumor that rarely metastatizes, but is often locally aggressive. Several factors, like large size (more than 3 cm), exposure to ultraviolet rays, histological variants, level of infiltration and perineural or perivascular invasion, are associated with a more aggressive clinical course. These morphological features seem to be more determinant in mideface localized BCC, which frequently show a significantly higher recurrence rate. An immunohistochemical profile, characterized by reactivity of tumor cells for p53, Ki67 and alpha-SMA has been associated with a more aggressive behaviour in large BCCs. The aim of this study was to verify if also little (<3 cm) basal cell carcinomas can express immunohistochemical markers typical for an aggressive behaviour.Basal cell carcinoma (BCC) is a very common malignant skin tumor that rarely metastatizes, even If Is often locally aggressive. Several factors, like large size (more than 3 cm), face localization, exposure to ultraviolet rays, histological variants, infiltration level and perineural or perivascular invasion, are associated with a more aggressive clinical course. In particular, the incidence of metastasis and/or death correlates with tumors greater than 3 cm in diameter in which setting patients are said to have 1–2 % risk of metastases that increases to 20–25% in lesions greater than 5 cm and to 50% in lesions greater than 10 cm in diameter (Snow et al., 1994). Histologically morpheiform, keratotic types and infiltrative growth of BCC are also considered features of the most aggressive course (Crowson, 2006). This can be explained by the fact that both the superficial and nodular variants of BCC are surrounded by a continuous basement membrane zone comprising collagens type IV and V admixed with laminin, while the aggressive growth variants (i.e. morpheiform, metatypical, and infiltrative growth subtypes) manifest the absence of basement membrane (Barsky et al., 1987).The molecular markers which characterize aggressive BCC include: increased expression of stromolysin (MMP-3) and collagenase-1 (MMP-1) (Cribier et al., 2001), decreased expression of syndecan-1 proteoglycan (Bayer-Garner et al., 2000) and of anti-apoptotic protein bcl-2 (Ramdial et al., 2000; Staibano et al., 2001).C-ras , c-fos (Urabe et al., 1994; Van der Schroeff et al., 1990) and p53 tumor supressor gene mutations (Auepemikiate et al., 2002) are indicative of an aggressive course.Focusing upon bcl-2 and p53 expression in BCC, there have been numerous studies documenting the utility of bcl-2 as a marker of favourable clinical behaviour while p53 expression may be a feature of a more aggressive outcome (Ramdial et al., 2000; Staibano et al., 2001; Bozdogan et al., 2002).An increased expression of cytoskeletal microfilaments like α–smooth muscle actin, frequently found in invasive BCC subtypes (Jones JCR et al., 1989), may explain an enhanced tumor mobility and deep tissue invasion through the stroma. (Cristian et al., 2001; Law et al., 2003). The aim of this preliminary study was to verify if also little (<3 cm) basal cell carcinomas may express aggressive immunohistochemical markers like p53, Ki67 and alpha-SMA. We used 31 excisional BCCs with tumor size less than 2 cm (ranging from 2 up to 20 mm) and with different skin localization (19 in the face, 6 in the trunk and 6 in the body extremities). All cases were immunostained for p53, BCL2, Ki67 and alpha-smooth muscle actin (α-SMA) (
AgeSexLocationHystotypeMax.DimDepthUlcEssInfp53Bcl-2Ki67AML
161MExtrKeratotic10×81No+++URD+++++-
261MFaceAdenoid10×94No+URD+++---
364MExtrSup mult11×130.8No+DRD+---
473MFaceNodular10×82Yes+DRD+++++++++
584MFaceNodular9×122Yes+DRD----
684MFaceAdenoid50.8No+URD+++---
784MExtrNodular13×103No+DRD+++++-
852FFaceNodular40.8No+URD+++-
976FFaceAdenoid10×44No+DRD+++-++-
1077FFaceMorph8×61Yes+++DRD+++---
1186MFaceMorph81Yes+DRD+++-++
1263FFaceAdenoid41No+URD+++++
1376FFaceNodular71.5No+DRD++++++-
1484MFaceNodular114Yes+++DRD+--+
1563FFaceKeratotic10×61.8No++DRD-+++-
1668FTrunkSup mult10×60.7No++URD++--
1767MFaceSup mult12×60.4No+URD+-+-
1867MExtrSup mult4×30.3No+URD+++++-
1932FExtrSup mult1×30.4No+URD+++-
2045MTrunkNodular7×52Yes+++URD+++-
2162MTrunkSup mult11×70.9No++URD-++-++
2265MTrunkAdenoid7×61.5No+URD+++++-
2372MTrunkNodular12×61No+URD+++-++
2486FFaceKeratotic20×113.1No++DRD+++-
2585MFaceNodular0.51.3No++DRD++++-
2674FExtrNodular4×40.9No+URD--+-
2771MFaceNodular6×121.7No+DRD--+-
2864FTrunkSup mult1.3×1.50.4No++URD+++---
2978FFaceNodular4×31.5No++DRD+++-+++
3080MFaceKeratotic4×41.6Yes+DRD--++++
Open in a separate window Our data show that p53 (75%), Bcl2 (50%) and Ki67 (63%) positivity was generally diffuse in the majority of cases. On the contrary, cytoplasmatic α-SMA expression was present only in 8 out of 31 cases (25,8%). All these 8 α-SMA positive BCCs, prevalently found in the mideface (6 out of 8), were characterized by an initial invasion beyond the dermis. Among these 6 face-localized α-SMA positive BCCs, 1 showed a sclerosing aggressive histotype, 1 a keratotic type and 4 a nodular histotype.These 8 little α-SMA-positive BCCs, compared to the others 23 α-SMA negative samples, all showed a major aggressiveness features: facial location, ulceration, morpheiform histotype and deeper infiltration into the dermis (Location
Histotype
Local aggressiveness
Immunohistochemistry
FaceKeratoticMorpheiformDepht of invasion Mean value(mm)UlcerationInfiltration of the dermisP53Bcl-2Ki678 α-SMA Positive cases75%12%12%1.650%63%75%50%63%23 α-SMA Negative cases56%13%4%1.413%48%78%43%65%
Open in a separate windowGiven the absence of a specific difference between α-SMA positive cases and α-SMA negative cases in the expression of aggressive immunohistochemical markers, except for a light reduction of bcl-2 in the α-SMA positive group (and2).2). By the analysis of the data, we selected the combination that could better define an aggressive behaviour even for little BCC: α-SMA, p53, Ki67 positivity and bcl-2 negativity. We considered p53 and ki67 markers of proliferation and cell-cycle alteration, combined with a loss of apoptotic activity expressed by Bcl-2 negativity, quite characteristic of aggressiveness; moreover α-SMA positivity probably reflects invasive potential and acquired mobility by neoplastic cells.This immunohistochemical profile (α-SMA, p53, Ki67 positivity and bcl-2 negativity) in our cases of BCC is present in two of them; one is a morpheiform BCC, that is an aggressive variant, while the other one is a nodular subtype (less aggressive).Therefore, our preliminary data suggest that only α-SMA positivity should be considered as an early diagnostic marker of potential aggressiveness in little BCC: all α-SMA positive little BCC in fact showed clinical and histological features of aggressiveness. Invasive potential is probably acquired by some BCCs not only when they reach large size, but it is probably present also when they have still little size, and can be revealed by α-SMA positivity in the neoplastic cells. Open in a separate windowFigure 1BCC, nodular type, HE, 10×. Open in a separate windowFigure 2BCC, nodular type, α-SMA positivity, 10×.  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