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.
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   

2.
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   

3.
Cj0859c variants fspA1 and fspA2 from 669 human, poultry, and bovine Campylobacter jejuni strains were associated with certain hosts and multilocus sequence typing (MLST) types. Among the human and poultry strains, fspA1 was significantly (P < 0.001) more common than fspA2. FspA2 amino acid sequences were the most diverse and were often truncated.Campylobacter jejuni is the leading cause of bacterial gastroenteritis worldwide and responsible for more than 90% of Campylobacter infections (7). Case-control studies have identified consumption or handling of raw and undercooked poultry meat, drinking unpasteurized milk, and swimming in natural water sources as risk factors for acquiring domestic campylobacteriosis in Finland (7, 9). Multilocus sequence typing (MLST) has been employed to study the molecular epidemiology of Campylobacter (4) and can contribute to virulotyping when combined with known virulence factors (5). FspA proteins are small, acidic, flagellum-secreted nonflagellar proteins of C. jejuni that are encoded by Cj0859c, which is expressed by a σ28 promoter (8). Both FspA1 and FspA2 were shown to be immunogenic in mice and protected against disease after challenge with a homologous strain (1). However, FspA1 also protected against illness after challenge with a heterologous strain, whereas FspA2 failed to do the same at a significant level. Neither FspA1 nor FspA2 protected against colonization (1). On the other hand, FspA2 has been shown to induce apoptosis in INT407 cells, a feature not exhibited by FspA1 (8). Therefore, our aim was to study the distributions of fspA1 and fspA2 among MLST types of Finnish human, chicken, and bovine strains.In total, 367 human isolates, 183 chicken isolates, and 119 bovine isolates (n = 669) were included in the analyses (3). PCR primers for Cj0859c were used as described previously (8). Primer pgo6.13 (5′-TTGTTGCAGTTCCAGCATCGGT-3′) was designed to sequence fspA1. Fisher''s exact test or a chi-square test was used to assess the associations between sequence types (STs) and Cj0859c. The SignalP 3.0 server was used for prediction of signal peptides (2).The fspA1 and fspA2 variants were found in 62.6% and 37.4% of the strains, respectively. In 0.3% of the strains, neither isoform was found. Among the human and chicken strains, fspA1 was significantly more common, whereas fspA2 was significantly more frequent among the bovine isolates (Table (Table1).1). Among the MLST clonal complexes (CCs), fspA1 was associated with the ST-22, ST-45, ST-283, and ST-677 CCs and fspA2 was associated with the ST-21, ST-52, ST-61, ST-206, ST-692, and ST-1332 CCs and ST-58, ST-475, and ST-4001. Although strong CC associations of fspA1 and fspA2 were found, the ST-48 complex showed a heterogeneous distribution of fspA1 and fspA2. Most isolates carried fspA2, and ST-475 was associated with fspA2. On the contrary, ST-48 commonly carried fspA1 (Table (Table1).1). In our previous studies, ST-48 was found in human isolates only (6), while ST-475 was found in both human and bovine isolates (3, 6). The strict host associations and striking difference between fspA variants in human ST-48 isolates and human/bovine ST-475 isolates suggest that fspA could be important in host adaptation.

TABLE 1.

Percent distributions of fspA1 and fspA2 variants among 669 human, poultry, and bovine Campylobacter jejuni strains and their associations with hosts, STs, and CCs
Host or ST complex/ST (no. of isolates)% of strains witha:
P valueb
fspA1fspA2
Host
    All (669)64.335.4
    Human (367)69.530.0<0.001
    Poultry (183)79.220.8<0.001
    Bovine (119)25.274.8<0.0001
ST complex and STs
    ST-21 complex (151)2.697.4<0.0001
        ST-50 (76)NF100<0.0001
        ST-53 (19)NF100<0.0001
        ST-451 (9)NF100<0.0001
        ST-883 (11)NF100<0.0001
    ST-22 complex (22)100NF<0.0001
        ST-22 (11)100NF<0.01
        ST-1947 (9)100NF0.03
    ST-45 complex (268)99.30.7<0.0001
        ST-11 (7)100NFNA
        ST-45 (173)99.40.6<0.0001
        ST-137 (22)95.54.50.001
        ST-230 (14)100NF<0.0001
    ST-48 complex (18)44.455.6NA
        ST-48 (7)100NFNA
        ST-475 (8)NF100<0.001
    ST-52 complex (5)NF100<0.01
        ST-52 (4)NF1000.02
    ST-61 complex (21)NF100<0.0001
        ST-61 (11)NF100<0.0001
        ST-618 (3)NF1000.04
    ST-206 complex (5)NF100<0.01
    ST-283 complex (24)100NF<0.0001
        ST-267 (23)100NF<0.0001
    ST-677 complex (59)100NF<0.0001
        ST-677 (48)100NF<0.0001
        ST-794 (11)100NF<0.001
    ST-692 complex (3)NF1000.04
    ST-1034 complex (5)NF80NA
        ST-4001 (3)NF1000.04
    ST-1287 complex/ST-945 (8)100NFNA
    ST-1332 complex/ST-1332 (4)NF1000.02
    Unassigned STs
        ST-58 (6)NF100<0.01
        ST-586 (6)100NFNA
Open in a separate windowaIn 0.3% of the strains, neither isoform was found. NF, not found.bNA, not associated.A total of 28 isolates (representing 6 CCs and 13 STs) were sequenced for fspA1 and compared to reference strains NCTC 11168 and 81-176. All isolates in the ST-22 CC showed the same one-nucleotide (nt) difference with both NCTC 11168 and 81-176 strains, resulting in a Thr→Ala substitution in the predicted protein sequence (represented by isolate FB7437, GenBank accession number HQ104931; Fig. Fig.1).1). Eight other isolates in different CCs showed a 2-nt difference (isolate 1970, GenBank accession number HQ104932; Fig. Fig.1)1) compared to strains NCTC 11168 and 81-176, although this did not result in amino acid substitutions. All 28 isolates were predicted to encode a full-length FspA1 protein.Open in a separate windowFIG. 1.Comparison of FspA1 and FspA2 isoforms. FspA1 is represented by 81-176, FB7437, and 1970. FspA2 is represented by C. jejuni strains 76763 to 1960 (GenBank accession numbers HQ104933 to HQ104946). Scale bar represents amino acid divergence.In total, 62 isolates (representing 7 CCs and 35 STs) were subjected to fspA2 sequence analysis. Although a 100% sequence similarity between different STs was found for isolates in the ST-21, ST-45, ST-48, ST-61, and ST-206 CCs, fspA2 was generally more heterogeneous than fspA1 and we found 13 predicted FspA2 amino acid sequence variants in total (Fig. (Fig.1).1). In several isolates with uncommon and often unassigned (UA) STs, the proteins were truncated (Fig. (Fig.1),1), with most mutations being ST specific. For example, all ST-58 isolates showed a 13-bp deletion (isolate 3074_2; Fig. Fig.1),1), resulting in a premature stop codon. Also, all ST-1332 CC isolates were predicted to have a premature stop codon by the addition of a nucleotide between nt 112 and nt 113 (isolate 1960; Fig. Fig.1),1), a feature shared with two isolates typed as ST-4002 (UA). A T68A substitution in ST-1960 (isolate T-73494) also resulted in a premature stop codon. Interestingly, ST-1959 and ST-4003 (represented by isolate 4129) both lacked one triplet (nt 235 to 237), resulting in a shorter FspA2 protein. SignalP analysis showed the probability of a signal peptide between nt 22 and 23 (ACA-AA [between the underlined nucleotides]). An A24C substitution in two other strains, represented by isolate 76580, of ST-693 and ST-993 could possibly result in a truncated FspA2 protein as well.In conclusion, our results showed that FspA1 and FspA2 showed host and MLST associations. The immunogenic FspA1 seems to be conserved among C. jejuni strains, in contrast to the heterogeneous apoptosis-inducing FspA2, of which many isoforms were truncated. FspA proteins could serve as virulence factors for C. jejuni, although their roles herein are not clear at this time.  相似文献   

4.
5.
6.
7.
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   

8.
9.
10.
11.
12.
13.
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   

14.
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   

15.
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   

16.
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   

17.
18.
The human immunodeficiency virus type 1 (HIV-1) variants that are transmitted to newly infected individuals are the primary targets of interventions, such as vaccines and microbicides, aimed at preventing new infections. Newly acquired subtype A, B, and C variants have been the focus of neutralization studies, although many of these viruses, particularly of subtypes A and B, represent viruses circulating more than a decade ago. In order to better represent the global diversity of transmitted HIV-1 variants, an additional 31 sexually transmitted Kenyan HIV-1 env genes, representing several recent infections with subtype A, as well as subtypes A/D, C, and D, were cloned, and their neutralization profiles were characterized. Most env variants were resistant to neutralization by the monoclonal antibodies (MAbs) b12, 4E10, 2F5, and 2G12, suggesting that targeting the epitopes of these MAbs may not be effective against variants that are spreading in areas of endemicity. However, significant cross-subtype neutralization by plasma was observed, indicating that there may be other epitopes, not yet defined by the limited available MAbs, which could be recognized more broadly.Most effective viral vaccines are thought to provide protection primarily by stimulating neutralizing antibodies (NAbs) to clear cell-free virus (25, 27). Because protection by NAbs requires recognition of common viral epitopes, the extreme genetic diversity of human immunodeficiency virus type 1 (HIV-1) presents a particular challenge to NAb-based vaccine approaches. Therefore, a critical starting point for studies of immune-mediated protection against HIV-1 is a collection of newly transmitted HIV-1 variants, particularly from areas of endemicity, such as sub-Saharan Africa, in order to determine whether vaccines are appropriately targeted to common epitopes from these relevant transmitted strains.During HIV-1 transmission, a bottleneck allows only one or a few variants to be transmitted to a newly infected individual (6, 9, 16, 29, 34, 37, 39), and the sensitivity of these early transmitted strains to antibody-mediated neutralization is therefore of particular interest. Newly transmitted HIV-1 variants have demonstrated significant heterogeneity in their neutralization phenotypes both within and between subtypes (2, 3, 6-8, 11, 13-15, 22, 30, 32, 36). Panels of sexually transmitted HIV-1 envelope variants (based on the envelope gene, env) have been characterized, including subtype B variants from North America, Trinidad, and Europe, subtype C variants from South Africa and Zambia, and subtype A variants from Kenya collected between 1994 and 1996 (2, 14, 15). Here, we characterize an additional 31 envelope variants from 14 subjects with sexually transmitted HIV-1 who were infected in Kenya, where subtypes A, C, and D circulate, between 1993 and 2005 (24, 31).The env genes were cloned from samples drawn 14 to 391 (median, 65) days postinfection from individuals enrolled in a prospective cohort of high-risk women in Mombasa, Kenya (19-21). Demographic characteristics of the subjects are summarized in Table Table1;1; the timing of first infection was determined by both HIV-1 serology and HIV RNA testing as described previously (12). All of the subjects were presumably infected by male-to-female transmission and displayed a range of plasma viral loads at the time of env gene cloning (Table (Table1).1). For most individuals, full-length env genes were cloned from uncultured peripheral blood mononuclear cell (PBMC) DNA, though for two individuals, clones were obtained from DNA following short-term coculture with donor PBMCs (Table (Table1).1). env genes were cloned by single-copy nested PCR with primers and PCR conditions as described previously (4, 17). We tested env genes for their ability to mediate infection by transfecting env plasmid DNA into 293T cells along with an env-deficient HIV-1 subtype A proviral plasmid, Q23Δenv, to make pseudoviral particles (17). More than 80 env clones were obtained from 16 subjects; less than one-half were functional on the basis of the infectivity of pseudoviral particles in a single-round infection of TZM-bl cells (AIDS Research and Reference Reagent Program, National Institutes of Health), as observed previously for env genes cloned from proviral sequences (17); a lower fraction of functional env genes have been reported from plasma (18). We focused on the proviral sequences here because they presumably best represent the sequence closest to that of the transmitted strains. The 31 functional env variants are described in Table Table11.

TABLE 1.

Demographic characteristics, diversities, gp120 variable-region lengths, numbers of PNGS, and accession numbers of cloned env variants
SubjectVirus subtypeSample date (mo/day/yr)dpiaPlasma VLbSourcecIndividual env clonePairwise difference (%)dVariable-loop length (aa)
No. of PNGS
GenBank accession no.
V1/V2V3V4V5gp120gp41gp41 ecto
QB726A04/16/967061,940ucPBMCQB726.70M.ENV.B30.16633536102244FJ866111
QB726.70M.ENV.C4633536102244FJ866112
QF495A05/16/0623217,050ucPBMCQF495.23M.ENV.A10.121073537113044FJ866113
QF495.23M.ENV.A31073537113044FJ866114
QF495.23M.ENV.B21133537113144FJ866115
QF495.23M.ENV.D11133537113144FJ866116
QG984A07/12/042130,300ucPBMCQG984.21M.ENV.A3NA693436112433FJ866117
QH209A10/13/051428,600ucPBMCQH209.14M.ENV.A2NA723529112444FJ866118
QH343A09/08/052140,750,000ucPBMCQH343.21M.ENV.A100.19773532152644FJ866119
QH343.21M.ENV.B5773532152644FJ866120
QH359A10/05/052132,120ucPBMCQH359.21M.ENV.C11.4843536102944FJ866121
QH359.21M.ENV.D1733535102644FJ866122
QH359.21M.ENV.E2723540132844FJ866123
QA790eA/D06/10/9620448,100ccPBMCQA790.204I.ENV.A40.36773533112544FJ866124
QA790.204I.ENV.C1773533112644FJ866125
QA790.204I.ENV.C8773533112444FJ866126
QA790.204I.ENV.E2773533112544FJ866127
QG393A2/D06/23/046017,360ucPBMCQG393.60M.ENV.A10.7603431102455FJ866128
QG393.60M.ENV.B7573431102455FJ866129
QG393.60M.ENV.B8573431102455FJ866130
QB099eC02/10/9539127,280ucPBMCQB099.391M.ENV.B10.43653529102544FJ866131
QB099.391M.ENV.C8653529102544FJ866132
QC406C07/08/9770692,320ucPBMCQC406.70M.ENV.F3NA643520112254FJ866133
QA013D10/11/95701,527,700ccPBMCQA013.70I.ENV.H10.16603429122544FJ866134
QA013.70I.ENV.M12603429122544FJ866135
QA465D08/19/935937,750ucPBMCQA465.59M.ENV.A10.24653530112844FJ866136
QA465.59M.ENV.D1653530112744FJ866137
QB857D10/16/9711014,640ucPBMCQB857.23I.ENV.B3NA683432112654FJ866138
QD435D04/06/9910017,470ucPBMCQD435.100M.ENV.A40.88693429122654FJ866139
QD435.100M.ENV.B5673429112454FJ866140
QD435.100M.ENV.E1693429122654FJ866141
Open in a separate windowadpi, days postinfection as defined by RNA testing (12).bVL, viral load on the sample date in which env genes were cloned.cucPBMC, uncultured PBMCs; ccPBMC, cocultured PBMCs.dAverage pairwise distance between the full-length env variants from a given subject. NA, not applicable because there was only one variant available from the subject.eenv variants from these two subjects were cloned from >6 months postinfection, as noted, and should not be considered true early env variants.The full-length, functional env genes were sequenced and aligned to generate a maximum likelihood phylogenetic tree with reference sequences from the Los Alamos National Laboratory HIV database, as described previously (26). Viral env clones from the same subject clustered together, and a wide spectrum of genetic diversity was observed overall (Fig. (Fig.1).1). Some women, such as subject QF495, were infected with a relatively homogeneous viral population, with average pairwise differences of only 0.12% between env variants (Table (Table11 and Fig. Fig.1).1). However, as observed previously in this cohort (16, 28, 29, 33-35), other individuals, such as subjects QH359 and QD435, were infected with more heterogeneous viral populations with average pairwise differences of 1.4% and 0.88% between variants, respectively (Table (Table11 and Fig. Fig.1).1). env genes from subtypes A (13 variants), C (3 variants), and D (8 variants), as well as A/D recombinants (4 variants) and A2/D recombinants (3 variants), were represented (Fig. (Fig.1).1). The viral subtypes were confirmed using the NCBI genotyping database (http://www.ncbi.nlm.nih.gov/).Open in a separate windowFIG. 1.Maximum likelihood phylogenetic tree of full-length sequences from early subtype A, C, D, and A/D recombinant env variants in Kenya. The 31 novel env clones from Kenyan early infections and reference sequences for subtypes A, B, C, D, and K from the Los Alamos HIV database (http://www.hiv.lanl.gov/content/index) are displayed. The phylogenetic tree was rooted with subtype K env sequences. Values at nodes indicate the percentage of bootstraps in which the cluster the right was found; only values of 70% or greater are shown.The deduced amino acid sequences revealed that all functional variants had an uninterrupted open reading frame in env except for variant QB099.391I.ENV.C8, which had a frameshift mutation within the cytoplasmic tail of gp41. There was significant heterogeneity in the length of the protein variable loops, particularly V1/V2, which ranged from 57 amino acids (aa) to 113 aa (Table (Table1).1). The V3, V4, and V5 loops also varied in length, though less dramatically (Table (Table1).1). Variants from the same subject were generally similar in their variable-loop lengths. Moderate variation was also observed in the number and position of potential N-linked glycosylation sites (PNGS) (Table (Table11).Previous analyses indicated that early subtype C env proteins had shorter variable loops than did early subtype B env proteins (13), suggesting that there are different env protein features between subtypes. Thus, to compare variable-loop lengths and the numbers of PNGS between subtypes using this expanded group of early env variants, we evaluated the 31 newly cloned variants plus an additional 15 subtype A variants (2), 19 subtype B variants (14), and 18 subtype C variants (15) from other early virus panels. In order to avoid bias, when more than one env variant was available from a subject, the average loop length or PNGS number for that subject''s env proteins was used. We did not observe significant differences in V1/V2 length, V5 length, or the numbers of PNGS between subtypes by the Kruskal-Wallis equality-of-populations rank test (Table (Table2)2) . However, there were significant differences between the V3 and V4 loop lengths of the subtypes after adjusting for multiple comparisons (Table (Table2).2). The differences in V3 length appeared to be a result of shorter V3 loops in subtype D env proteins than in early subtype B (P = 0.006) or C (P < 0.001) env proteins (Table (Table2).2). The differences in V4 length were caused by shorter V4 loops in subtype C env proteins in comparison to both subtype A and B env proteins (P < 0.001; Table Table22).

TABLE 2.

Summary of variable-loop lengths and the numbers of PNGS in gp120 and gp41 within early HIV-1 env variantsa
ParameterMedian value (25th percentile, 75th percentile) for subtype:
Kruskal- Wallis P valuebWilcoxon rank sum P values for individual comparisonsc
A (n = 11)B (n = 19)C (n = 20)D (n = 4)A vs. BA vs. CA vs. DB vs. CB vs. DC vs. D
Length
    V1/V270.3 (62, 76)70 (66, 70)65 (62, 76)66.5 (62, 69)0.210.7300.2820.2150.0510.1130.846
    V335 (34, 35)35 (35, 35)35 (34, 35)34 (34, 35)0.0010.2400.0160.1070.1410.006<0.001
    V432 (30, 36)33 (31, 34)26.5 (22, 29)29.5 (29, 31)0.00010.880<0.0010.148<0.0010.0230.056
    V511 (11, 11)10 (9, 11)10 (9, 11)11.5 (11, 12)0.0300.0960.0150.1840.6770.0990.021
No. of PNGS in:
    gp12024 (23, 28)25 (24, 26)24 (23, 25)26 (26, 27)0.200.6800.6920.2650.1460.1860.042
    gp414 (4, 5)5 (4, 5)5 (4, 5)4.5 (4, 5)0.200.0300.1790.4700.4100.4080.799
    gp41ecto4 (4, 4)4 (4, 4)4 (4, 5)4 (4, 4)0.0440.1070.0250.5500.0880.5070.201
Open in a separate windowaVariable-loop lengths and the numbers of PNGS in gp120 and gp41 within early HIV-1 env variants from subtypes A, B, C, and D characterized here and previously (2, 14, 15). n, number of samples.bKruskal-Wallis equality-of-populations rank test (based on multiple comparisons; P values of <0.007 were considered significant; significant values are presented in boldface).cWilcoxon rank sum test (based on multiple comparisons; P values of <0.008 were considered significant; significant values are presented in boldface).We then assessed the neutralization sensitivity of the pseudoviruses to antibodies in plasma from HIV-1-infected individuals and to HIV-1-specific MAbs by using the TZM-bl neutralization assay as described previously (2, 23, 38). Median inhibitory concentrations (IC50s) were defined as the reciprocal dilution of plasma or concentration of MAb that resulted in 50% inhibition of infection (2, 38). The Kenya pool was derived by pooling plasma collected between 1998 and 2000 from 30 HIV-1-infected individuals in Mombasa, Kenya, and the other three pools were derived by pooling plasma collected between 1993 and 1997 from 10 individuals from Nairobi, Kenya, and with an infection with a known subtype (A, C, or D) of HIV-1 as described previously (2).The env variants demonstrated a range of neutralization sensitivities to plasma samples, from neutralization resistant (defined as <50% neutralization with a 1:50 dilution of plasma) to neutralization sensitive with an IC50 of 333 (Fig. (Fig.2).2). Some clones, such as QF495.23M.ENV.A1, were relatively sensitive to all the plasma pools, with IC50s from 100 to 333, whereas other clones, such as QH343.21M.ENV.A10, were relatively resistant to these plasma pools, with IC50s from <50 to 85 (Fig. (Fig.2).2). The plasma pools did differ in their neutralization potencies. The Kenya pool, with a median IC50 of <50 across all viruses tested, was significantly less likely to neutralize these transmitted variants than were the subtype A, C, and D plasma pools, which had median IC50s of 110, 105, and 123, respectively (P values of <0.0001, 0.0001, and 0.001, respectively, by paired t test on log-transformed IC50s). The basis for these differences in neutralizing activity is not clear, although the location, timing, and level of immunodeficiency at the time of sample collection could have contributed to the differences in NAb levels between the pools.Open in a separate windowFIG. 2.Neutralization sensitivity of early subtype A, C, D, and A/D recombinant env variants to plasma samples and MAbs in relation to the sequences of the MAb binding sites. The env used to generate the pseudovirus tested is shown at the left, and the plasma pool or MAb tested is indicated at the top. The IC50s of each plasma sample or MAb against each viral pseudotype is shown, with darker shading indicating more potent neutralization, as defined at the bottom of the figure. Gray boxes indicate that <50% neutralization was observed at the highest dilution of plasma or concentration of MAb tested. Each IC50 shown is an average of the results from two independent neutralization assays, using pseudovirus generated in independent transfection experiments. The median IC50s from the 31 variants are shown at the bottom. Neutralization of the pseudovirus derived from the subtype B variant SF162 is shown as a control, and neutralizations of murine leukemia virus (MLV) and simian immunodeficiency virus clone 8 (SIV) are shown as negative controls. In the panels on the right, the sequences for the MAbs 2G12, 2F5, and 4E10 are displayed. For 2G12, the amino acid numbers for the five PNGS that are important for 2G12 binding are shown for each virus tested. A plus sign indicates that the PNGS at that site in the envelope sequence was preserved, and a minus sign indicates that the PNGS was deleted. A shift in the PNGS position is indicated by the amino acid position to which the PNGS shifted. All sequences were numbered relative to the HXB2 sequence. The two rightmost panels show data for the canonical 2F5 and 4E10 epitopes, with a period indicating that the amino acid is preserved.The env variants were significantly more susceptible to their subtype-matched plasma pool, with a higher mean IC50 for subtype-matched plasma samples than for unmatched plasma samples (138 versus 108, P = 0.0081, paired t test). However, a significant amount of cross-subtype neutralization was observed, as every env variant that was susceptible to the subtype-matched plasma pool was also susceptible to at least one of the other plasma pools (Fig. (Fig.2).2). Thus, although potency was enhanced when the plasma antibodies were produced in response to infection with the same subtype of HIV-1, there were shared neutralization determinants between subtypes, as has been observed previously (reviewed in reference 3).To identify potential correlates of neutralization sensitivity to the antibodies within these plasma pools, we included these 31 env variants and an additional 15 subtype A env variants we previously characterized from the same cohort with the same plasma pools (2). We did not observe a change in neutralization sensitivity during the evolution of the HIV-1 epidemic in Kenya, as no correlation was observed between neutralization sensitivity and the calendar date from which the env variants were isolated. In addition, no correlation was observed between the neutralization sensitivity of a variant to the plasma pools and the duration of estimated infection within that individual. Finally, there was no significant correlation between the neutralization sensitivity and variable-loop length or the number of PNGS. Thus, although changes in the variable-loop length or number of PNGS may alter the exposure of epitopes within the HIV-1 env protein, these changes do not appear to be the primary determinant of neutralization sensitivity.Despite relatively universal sensitivity to at least one of the pooled plasma samples, these transmitted Kenyan env variants were generally resistant to the MAbs 2G12 (provided by Hermann Katinger, Polymun Scientific) and b12 (provided by Dennis Burton, The Scripps Research Institute), as well as 2F5 and 4E10 (obtained from the AIDS Research and Reference Reagent Program, National Institutes of Health) (Fig. (Fig.2),2), though these MAbs neutralized the subtype B env variant SF162, with IC50s similar to those reported previously (1). Subtype D strains were the most susceptible to MAbs, with 4/8 variants neutralized with <20 μg/ml of 2F5 and 2/8 neutralized with <20 μg/ml of the other MAbs. This could reflect the fact that subtype D variants are more closely related to subtype B strains (Fig. (Fig.1)1) (see reference 10), and these MAbs were all derived from subtype B-infected individuals.Among all 31 variants, 2F5 was the most broadly neutralizing, with 15/31 variants from 8/14 subjects neutralized with <20 μg/ml of this MAb. Some 2F5-resistant env variants, such as QH209.14M.ENV.A2 and QB857.110I.ENV.B3, had mutations in the canonical 2F5 binding epitopes, though other 2F5-resistant env variants such as QF495.23M.ENV.A3 and QA790.204I.ENV.A4 maintained the canonical 2F5 epitope. The results with the MAb 4E10 were similar; 4E10 neutralized only seven variants from 4 of the 14 subjects, and the presence of mutations in the 4E10 epitope, which were common, did not predict neutralization sensitivity (Fig. (Fig.2).2). For instance, the env variants QH343.21M.ENV.A10 and QH343.21M.ENV.B5 contained identical N671S and D674S mutations and QH343.21M.ENV.B5 was highly sensitive to 4E10, while QH343.21M.ENV.A10 was resistant (Fig. (Fig.2).2). Thus, for the 2F5 and 4E10 epitopes, the presumed epitopes appear to be shielded in a subset of these early non-subtype B env variants, as has been previously observed (Fig. (Fig.2)2) (1, 2, 5, 14).The MAb b12 neutralized only two variants from two subtype D-infected individuals, with no neutralization of the subtype A, C, and A/D recombinant pseudoviruses. Only four variants from two subjects were neutralized by 2G12 at <20 μg/ml, and these were the only variants that maintained all five of the PNGS within the 2G12 epitope (Fig. (Fig.2).2). Overall, the median IC50 of all the MAbs against these transmitted variants was >20 μg/ml. None of the variants was susceptible to all four MAbs (Fig. (Fig.2),2), unlike many of the early subtype B env variants characterized previously (14).In summary, these newly characterized HIV-1 env clones represent a range of neutralization sensitivities and can be used to supplement existing panels of transmitted variants, in particular, adding the first subtype D and A/D recombinant variants. Some differences between subtypes in env structure following transmission were noted, though these differences did not correlate with neutralization sensitivity. Although the significant levels of cross-subtype neutralization sensitivity observed with plasma samples indicate that some neutralization determinants were shared across subtypes, the epitopes for the MAbs b12, 2G12, 2F5, and 4E10 did not appear to be among the shared determinants. Thus, despite the fact that significant attention has focused on using vaccination to develop antibodies that resemble these MAbs in their specificity, such antibodies may not neutralize the transmitted strains that are causing most new infections worldwide. These data therefore stress the importance of evaluating transmitted variants in endemic areas when designing immunogens and evaluating vaccine and microbicide strategies.  相似文献   

19.
20.
Melioidosis has been considered an emerging disease in Brazil since the first cases were reported to occur in the northeast region. This study investigated two municipalities in Ceará state where melioidosis cases have been confirmed to occur. Burkholderia pseudomallei was isolated in 26 (4.3%) of 600 samples in the dry and rainy seasons.Melioidosis is an endemic disease in Southeast Asia and northern Australia (2, 4) and also occurs sporadically in other parts of the world (3, 7). Human melioidosis was reported to occur in Brazil only in 2003, when a family outbreak afflicted four sisters in the rural part of the municipality of Tejuçuoca, Ceará state (14). After this episode, there was one reported case of melioidosis in 2004 in the rural area of Banabuiú, Ceará (14). And in 2005, a case of melioidosis associated with near drowning after a car accident was confirmed to occur in Aracoiaba, Ceará (11).The goal of this study was to investigate the Tejuçuoca and Banabuiú municipalities, where human cases of melioidosis have been confirmed to occur, and to gain a better understanding of the ecology of Burkholderia pseudomallei in this region.We chose as central points of the study the residences and surrounding areas of the melioidosis patients in the rural areas of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50′W) (Fig. (Fig.1).1). There are two well-defined seasons in each of these locations: one rainy (running from January to May) and one dry (from June to December). A total of 600 samples were collected at five sites in Tejuçuoca (T1, T2, T3, T4, and T5) and five in Banabuiú (B1, B2, B3, B4, and B5), distributed as follows (Fig. (Fig.2):2): backyards (B1 and T1), places shaded by trees (B2 and T2), water courses (B3 and T3), wet places (B4 and T4), and stock breeding areas (B5 and T5).Open in a separate windowFIG. 1.Municipalities of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50″W).Open in a separate windowFIG. 2.Soil sampling sites in Banabuiú and Tejuçuoca.Once a month for 12 months (a complete dry/rainy cycle), five samples were gathered at five different depths: at the surface and at 10, 20, 30 and 40 cm (Table (Table1).1). The samples were gathered according to the method used by Inglis et al. (9). Additionally, the sample processing and B. pseudomallei identification were carried out as previously reported (1, 8, 9).

TABLE 1.

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

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