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 109 毫秒
1.
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   

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

TABLE 1.

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

TABLE 2.

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

3.
4.
5.
6.
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   

7.
8.
9.
10.
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   

11.
12.
Endoglucanase C5614-1 comprises a catalytic module (CM) and an X module (XM). The XM showed no significant homology with known carbohydrate-binding modules (CBMs). Recombinant full-length endoglucanase could bind Avicel, whereas the CM could not. The XM could bind various polysaccharides. The results demonstrated that the XM was a new CBM.Most cellulases are modular proteins that comprise two or more discrete modules, such as catalytic modules (CMs) and carbohydrate-binding modules (CBMs), each of which can function independently (9). CBMs are classified into 59 families based on their amino acid similarity in the CAZY database (http://www.cazy.org/fam/acc_CBM.html). The main functions of CBMs are to recognize and bind polysaccharides and to increase the hydrolytic activities of the enzymes against insoluble and soluble substrates (3). Endoglucanase C5614-1 (GenBank accession no. ACA61140), which was identified from the metagenome of the contents of buffalo rumen (5), is a modular enzyme comprising an N-terminal signal peptide (amino acids [aa] 1 to 20), a CM belonging to the glycosyl hydrolase family 5 (aa 40 to 334), and a C-terminal X module (XM) of unknown function (aa 335 to 537) (Fig. (Fig.1A).1A). No linker region rich in Ser/Pro/Thr was found between the CM and the XM. In this study, we aimed to ascertain the function of the XM in the endoglucanase C5614-1.Open in a separate windowFIG. 1.Endoglucanase C5614-1 and its derivatives. (A) Modular organization of C5614-1 and its truncated derivatives. Abbreviations: SP, signal peptide; GHF5, GHF5 catalytic module; X, the X module (XM). (B) SDS-PAGE of purified recombinant proteins. Protein samples were analyzed on a 10% gel. Lane 1, protein molecular mass standard (molecular masses are shown on the left); lane 2, rC5614-1; lane 3, rGHF5; lane 4, rX.The XM showed no significant homology to known CBMs. The XM shared 25% to 33% identities and 41% to 45% similarities with about 200 amino acids at the C terminus of a xylanase (GenBank accession no. AAC36862) from the ruminal bacterium Prevotella ruminicola, uncultured ruminal microbial cellulases (ABX76045, ACA61132, ACA61135, ACA61137, and ABB46200), and an uncultured bacterial bifunctional mannanase-xyloglucanase (ADA62505). None of these homologous polypeptides was confirmed to show carbohydrate-binding activity. An alignment of XM with these homologous sequences using ClustalW (http://www.ebi.ac.uk/Tools/clustalw) is shown in Fig. Fig.2.2. Seven conserved aromatic amino acid residues were found in the respective sequences. Aromatic amino acid residues in CBMs play critical roles in recognizing and binding polysaccharides (4).Open in a separate windowFIG. 2.Multiple sequence alignment of the XM (203 aa) in endoglucanase C5614-1 (aa 335 to 537), with its homologous sequences. The identical and similar amino acid residues are indicated by asterisks and dots, respectively, below the alignment. The conserved aromatic amino acid residues are indicated by arrows. GenBank accession no. AAC36862, Prevotella ruminicola xylanase (aa 376 to 584); ABB46200, ruminal uncultured bacterium endoglycosidase precursor protein (aa 719 to 917); ABX76045, ruminal uncultured microorganism endo-1,4-beta-d-glucanase (aa 342 to 552); ADA62505, ruminal uncultured bacterium bifunctional mannanase-xyloglucanase (aa 721 to 919); ACA61132, ruminal uncultured microorganism cellulase C29-2 (aa 341 to 553); ACA61135, ruminal uncultured microorganism cellulase C35-2 (aa 337 to 552); ACA61137, ruminal uncultured microorganism cellulase C67-1 (aa 335 to 546). The amino acid numbers in each of the parentheses above define the range of the homologous region in each sequence.A PCR-based approach was used to produce constructs expressing C5614-1 derivatives (Fig. (Fig.1A).1A). Plasmid C5614 (5), carrying the endoglucanase gene C5614-1, was used as a PCR template. The primer pairs used to amplify the portions of C5614-1 encoding C5614-1 amino acids 20 to 537 (recombinant C5614-1 [rC5614-1], comprising the CM and the XM), 20 to 348 (rGHF5, containing the CM only), and 349 to 537 (rX, containing the XM only) are, respectively, as follows: C5614-1F (5′CAGCCATGGAGGCACAAGATTTTGAGACTGCTACCGAA-3′) and C5614-1R (5′-GACCTCGAGTTGTGCTATGTATTTTTTGCCGTTCTGG-3′), C5614-1F and C5614-1CM (5′-GGGCTCGAGGGTTAATGTCTCAGCCAGGTCAGGCTG-3′), and C5614-1X (5′-GGGCCATGGCCAAAGCCTATCATGGCAGCGCGTTC-3′) and C5614-1R. The underlined sequences in the primers are NcoI and XhoI restriction sites.The digested PCR products were ligated into the same digested expression vector, pET-30a(+) (Novagen, Madison, WI). The recombinant plasmids were transformed into Escherichia coli Rosetta(DE3)pLysS (Novagen), and the cloned fragments were expressed as proteins with 6×His tags at both the N and C termini. The recombinant C5614-1 derivatives were purified by affinity chromatography with Cobalt immobilized metal chromatography resin (Clontech, Palo Alto, CA), according to the user manual. Each of the purified proteins produced a single band on an SDS-PAGE gel (8), and their molecular sizes were in agreement with those of the deduced polypeptides (Fig. (Fig.1B1B).The hydrolytic activities of rC5614-1 and rGHF5 toward carboxymethyl cellulose (CMC) were determined essentially as described by Duan et al. (5). The pH profiles of the enzymatic reactions of rC5614-1 and rGHF5 were similar, and both showed maximum activities at pH 5.0 (data not shown). However, rC5614-1 and rGHF5 showed different temperature profiles (Fig. (Fig.3A).3A). The rGHF5 showed narrower pH stability and lower temperature stability profiles than those of rC5614-1 (Fig. 3B and C). These results indicated that the XM is required for the stability of rC5614-1.Open in a separate windowFIG. 3.Effects of pH and temperature on the activities and stability of rC5614-1 and rGHF5. (A) Influence of temperature on the activities of rC5614-1 and rGHF5. Cellulase activity was measured at pH 5.0 in citrate-phosphate buffer at the indicated temperatures for 10 min. Values are expressed as percentages of maximal activity at 50°C and 35°C for rC5614-1 and rGHF5, respectively. (B) Influence of pH on rC5614-1 and rGHF5 stability. Purified enzyme was first incubated in citrate-phosphate buffer (pH 2.5 to 7.0), 0.1 M Tris-HCl buffer (pH 7.5 to 8.5), or 0.1 M glycine-NaOH buffer (pH 9.0 to 12) at 4°C for 24 h, and activity was measured under optimal conditions for 10 min. Values are expressed as percentages of maximal activity when the sample was incubated under pH 5.5 and pH 6.0 for rC5614-1 and rGHF5, respectively. (C) Thermal stability of recombinant rC5614-1 and rGHF5. Purified enzymes were first incubated at the indicated temperatures for 1 h; activity was then measured under optimal conditions for 10 min. Values are expressed as percentages of untreated-sample activity.The hydrolytic activities of rC5614-1 toward particle substrates, including birch wood xylan, lichenan, and acid-swollen cellulose (ASC), were about three times greater than those of rGHF5; however, rC5614-1 and rGHF5 showed similar hydrolytic activities toward soluble substrates (Table (Table1).1). This result indicated that the presence of the XM enhanced the hydrolytic activity of the enzyme toward insoluble substrates but not toward soluble substrates. Similar phenomena were reported for Irpex lacteus exocellulase I (7), Clostridium thermocellum Xyn10C (1), and Clostridium stercorarium Xyn10B (2).

TABLE 1.

Specific activities of rC5614-1 and rGHF5 toward various substrates
Test substrateSp act (U/mg protein)a
rC5614-1rGHF5
Barley glucan126.1 ± 4.5319.8 ± 19.9
Carboxymethyl cellulose72.5 ± 2.162.8 ± 4.4
Lichenan57.2 ± 3.120.6 ± 0.8
2-Hydroxyethyl cellulose24.6 ± 0.820.0 ± 0.1
Methyl cellulose8.8 ± 0.713.6 ± 0.03
Birch wood xylan6.7 ± 0.71.9 ± 0.16
Acid-swollen cellulose1.7 ± 0.050.5 ± 0.005
p-Nitrophenyl-d-cellobioside<0.005<0.002
Avicel (β-1,4-glucan)00
Laminarin00
p-Nitrophenyl-d-glucopyranoside00
Open in a separate windowaOne unit of enzyme activity was defined as the amount of enzyme releasing 1 μmol of glucose equivalent or p-nitrophenol per min from substrates.Binding of rC5614-1 and rGHF5 to Avicel was determined quantitatively, as described in the supplemental material. As shown in Table Table2,2, rC5614-1 bound to Avicel, and the binding capability of rC5614-1 increased with increasing Avicel concentration and with prolonged incubation time. Its binding capability was also influenced by the pH of the mixture solution. Approximately 90% of rC5614-1 bound to 4% cellulose in the mixture solution at pH 5.0 after incubation for 5 h, whereas the proportion of the enzyme bound to cellulose dropped slightly to 83% at pH 4.0 and dropped significantly to less than 65% at pHs ≥6. Bovine serum albumin (BSA) in the mixture solution only slightly affected the binding of rC5614-1 to Avicel, suggesting that its binding to Avicel was specific. However, 100% of rGHF5 remained in the supernatant of the binding mixture, showing that rGHF5 could not bind Avicel. These results demonstrated that the XM is absolutely required for the binding of rC5614-1 to Avicel, suggesting that the XM is a CBM.

TABLE 2.

Adsorption properties of rC5614-1 to Avicel
Additive(s) (pH)aBinding to Avicel (%) after incubation for:
1 hb5 h
1% Avicel25.0 ± 1.347.8 ± 1.8
2% Avicel38.1 ± 2.872.7 ± 1.2
4% Avicel55.3 ± 1.792.1 ± 1.2
8% Avicel65.1 ± 3.896.2 ± 1.2
4% Avicel + 0.08% BSAND79.8 ± 1.1
4% Avicel + 0.01% BSAND81.5 ± 1.9
4% Avicel (4.0)ND83.6 ± 0.4
4% Avicel (6.0)ND63.5 ± 3.2
4% Avicel (7.0)ND62.0 ± 3.5
Open in a separate windowaUnless otherwise stated, the mixture solution was at pH 5.0.bND, not determined.To further confirm the XM as a CBM, the binding of purified rX to insoluble polysaccharides was investigated by incubating the polypeptide with various polysaccharides and comparing the proteins in the supernatant fraction (unbound protein) and in the precipitated fraction (bound protein) by SDS-PAGE, as described in the supplemental material. rX could bind to the insoluble polysaccharides Avicel, ASC, chitin, lichenan, xylan from sugarcane bagasse, powder of sugarcane bagasse, and raw cassava starch. It bound slightly to agarose and Sephadex G-100 (Fig. (Fig.4,4, top). The affinities of rX for soluble polysaccharides were examined by native-affinity PAGE. The migration of rX was strongly retarded by inclusion of methylcellulose, 2-hydroxyethylcellulose, and barley glucan in gels, whereas it was only slightly affected by the presence of birch wood xylan and CMC (Fig. (Fig.4,4, bottom) and not affected by the inclusion of laminarin and soluble starch (data not shown). The affinities of the XM in C5614-1 to insoluble substrates were similar to those of CBM37 from Ruminococcus albus (10) and CBM54 from Clostridium thermocellum (6), which could also bind various insoluble substrates. However, the binding of CBM54 and CBM37 to soluble substrates was either negative or not tested. We propose that the X module in endoglucanase C5614-1 is a novel CBM.Open in a separate windowFIG. 4.Binding of rX to insoluble (top) and soluble (bottom) polysaccharides. In the experiment whose results are shown in the top panel, purified rX was incubated with insoluble polysaccharides, including Avicel (lanes A), ASC (lanes B), chitin (lanes C), lichenan (lanes D), agarose (lanes E), Sephadex G-100 (lanes F), xylan from sugarcane bagasse (lanes G), the powder of sugarcane bagasse (lanes H), and raw cassava starch (lanes I). CK is a control (the amount of protein used in the binding assay). After centrifugation, proteins in the precipitate (lanes 1) and the supernatant (lanes 2) were analyzed by SDS-PAGE. In the experiment whose results are shown in the bottom panel, purified rX and bovine serum albumin (BSA) were separated in nondenaturing polyacrylamide gels containing 0.1% (wt/vol) soluble polysaccharides, including methylcellulose (B), 2-hydroxyethylcellulose (C), barley glucan (D), birch wood xylan (E), and CMC (F). A gel without polysaccharide served as a reference (A). Lanes M contained BSA as a control.   相似文献   

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

TABLE 1.

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

15.
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   

16.
DNA sequence analysis and genetic mapping of loci from mating-type-specific chromosomes of the smut fungus Microbotryum violaceum demonstrated that the nonrecombining mating-type-specific region in this species comprises ∼25% (∼1 Mb) of the chromosome length. Divergence between homologous mating-type-linked genes in this region varies between 0 and 8.6%, resembling the evolutionary strata of vertebrate and plant sex chromosomes.EVOLUTION of mating types or sex-determining systems often involves the suppression of recombination around the primary sex-determining or mating-type-determining locus. In animals and plants, it is often an entire or almost entire chromosome (Y or W in male or female heterogametic species, respectively) that ceases to recombine with its homologous (X or Z) chromosome (Charlesworth and Charlesworth 2000; Charlesworth 2008). Self-incompatibility loci in plants are also thought to be located in regions of suppressed recombination (Charlesworth et al. 2005; Kamau and Charlesworth 2005; Kamau et al. 2007; Li et al. 2007; Yang et al. 2007). Regardless of the phylogenetic position of a species, such nonrecombining regions are known to follow similar evolutionary trajectories. The nonrecombining region on the sex-specific chromosome expands in several steps, forming evolutionary strata—regions of different X/Y (or Z/W) divergence (Lahn and Page 1999; Handley et al. 2004; Sandstedt and Tucker 2004; Nicolas et al. 2005)—and genes in the nonrecombining regions gradually accumulate deleterious mutations that eventually render them dysfunctional (Charlesworth and Charlesworth 2005; Charlesworth 2008).Fungal mating-type systems are very diverse, with the number of mating types varying from two to several hundred (Casselton 2002). Like sex chromosomes in several animals and plants, suppressed recombination has evolved in regions near fungal mating-type loci, including in Ustilago hordei (Lee et al. 1999), Cryptococcus neoformans (Lengeler et al. 2002), and Neurospora tetrasperma (Menkis et al. 2008). These species have two mating types, but no morphologically distinct sexes. The mating-type locus (the region of suppressed recombination) of C. neoformans is small (∼100 kb) compared with known sex chromosomes and contains only ∼20 genes that, unlike many sex chromosomes (Y or W chromosomes), show no obvious signs of genetic degeneration (Lengeler et al. 2002; Fraser et al. 2004). Judging from the divergence between the homologous genes on the two mating-type-specific chromosomes, C. neoformans started to evolve sex chromosomes a long time ago because silent divergence between the two mating types in the most ancient region exceeds 100% (Fraser et al. 2004). Genes in the younger mating-type-specific region are much less diverged between the two sex chromosomes, suggesting that the evolution of the sex locus in C. neoformans might have proceeded through several steps. The nonrecombining region around the mating-type locus of N. tetrasperma is much larger than in C. neoformans (at least 6.6 Mb), and silent divergence between homologous genes on the mating-type-specific chromosomes ranges from zero to 9%, demonstrating that these mating-type-specific chromosomes evolved recently (Menkis et al. 2008).M. violaceum, which causes anther smut disease in Silene latifolia and other species in the family Caryophyllaceae, has two mating types, A1 and A2 (reviewed by Giraud et al. 2008), which are determined by the presence of mating-type-specific chromosomes (hereafter A1 and A2 chromosomes, or sex chromosomes) in the haploid stage of the life cycle (Hood 2002; Hood et al. 2004). The A1 and A2 chromosomes are distinguishable by size in pulsed-field electrophoresis, and it is possible to isolate individual chromosomes electrophoretically (Hood et al. 2004). Random fragments of A1 and A2 chromosomes have previously been isolated from mating-type-specific bands of pulsed-field separated chromosomes of M. violaceum (Hood et al. 2004). These fragments were assumed to be linked to mating type. The same method was used to isolate fragments of non-mating-type-specific chromosomes. On the basis of the analysis of their sequences, (Hood et al. 2004) proposed that mating-type-specific chromosomes in M. violaceum might be degenerate because they contained a lower proportion of protein-coding genes than other chromosomes. However, it was not determined whether the sequences isolated from the mating-type chromosomes originated from the mating-type-specific or from the recombining regions (Hood et al. 2004), and the relative sizes of these regions are not known for these M. violaceum chromosomes. We tested the mating-type specificity of 86 of these fragments and demonstrate that fewer than a quarter of these loci are located in the mating-type-specific region, suggesting that the nonrecombining region on the A1 and A2 chromosomes is quite small, while the rest of the chromosome probably recombines (like pseudoautosomal regions of sex chromosomes) and is therefore not expected to undergo genetic degeneration. Genetic mapping confirms the presence of two pseudoautosomal regions in the M. violaceum mating-type-specific chromosomes.As these chromosomes are mating type specific in the haploid stage of M. violaceum, mating-type-specific loci (or DNA fragments) can be identified by testing whether they are present exclusively in A1 or A2 haploid strains. We therefore prepared haploid A1 and A2 M. violaceum cultures from S. latifolia plants from two geographically remote locations (accessions Sl405 from Sweden and Sl127 from the French Pyrenees). Haploid sporidial cultures were isolated by a standard dilution method (Kaltz and Shykoff 1997; Oudemans and Alexander 1998). Mating types were determined by PCR amplification of each culture with primers designed for A1 and A2 pheromone receptor genes linked to A1 and A2 mating types (Yockteng et al. 2007). The primers were as follows: 5′-TGGCATCCCTCAATGTTTCC-3′ and 5′-CACCTTTTGATGAGAGGCCG-3′ for the A1 pheromone receptor (GenBank accession no. EF584742) and 5′-TGACGAGAGCATTCCTACCG-3′ and 5′-GAAGCGGAACTTGCCTTTCT-3′ for the A2 pheromone receptor (GenBank accession no. EF584741). Cultures with PCR product amplified only from an A1 or A2 pheromone receptor gene were selected for further use. The mating types of the cultures were verified by conjugating them in all combinations.The GenBank nucleotide database was searched using BLAST for sequences similar to those isolated by Hood et al. (2004). Sequences with similarity to transposable elements (TE) and other repeats were excluded. The resulting set of nonredundant sequences was used to design PCR primers for 98 fragments. Half of these were originally isolated from the A1 and half from the A2 chromosomes and are hereafter called A1-NNN or A2-NNN (where NNN is the locus number; supporting information, Table S1), which does not imply that these loci are A1 or A2 specific, but merely indicates that they were originally isolated from the A1 or A2 chromosomes. Amplification of these regions from new A1 and A2 M. violaceum cultures, independently isolated by ourselves, revealed that only 5 of the 49 loci isolated from the A1 chromosome are indeed A1 specific and only 6 of 49 isolated from the A2 chromosome are A2 specific. All other loci amplified from both A1 and A2 cultures. Figure 1 illustrates some of these results from the Swedish sample (Sl405).Open in a separate windowFigure 1.—Testing of mating-type specificity for loci isolated from A1 and A2 chromosomes. (a) PCR amplifications from haploid cultures from Sl405 using primers designed from six A1-originated loci. Loci in which a PCR product could be amplified only from A1 cultures (boxed) were classified as specific to mating type A1. (b) PCR tests of six A2-originated loci on the same set of haploids as in a. Loci in which a PCR product amplified only from A2 cultures (boxed) were classified as specific to mating type A2. Loci amplified from both A1 and A2 cultures are not mating type specific.The fragments that amplified from both A1 and A2 mating types may be in recombining regions, or they could be present in mating-type-specific regions on both A1 and A2 chromosomes. If they are in recombining regions, the A1- and A2-linked homologs should not be diverged from each other, but if they are in nonrecombining, mating-type-specific regions, the divergence of the A1- and A2-linked homologs should be roughly proportional to the time since recombination stopped in the region. We therefore sequenced and compared PCR fragments amplified from the two mating types of Sl405 or Sl127 cultures (GenBank accession nos. FI855822FI856001). Sequencing of PCR products showed that 12 (4 A1 and 8 A2) loci have more than one copy, and they were excluded from further analysis. Sequences of 61 loci were identical between the A1 and A2 strains, and four loci demonstrated low total divergence (0.24–0.61%) between the two mating types (otintseva and D. Filatov, unpublished results). Thus, these loci might be located in the recombining part of the mating-type-specific chromosomes. Ten of 75 loci that amplified in both mating types demonstrated multiple polymorphisms fixed between the mating types rather than between the locations. Given that the strains that we used in the analysis originated from two geographically distant locations, it is highly unlikely that multiple polymorphisms distinguishing the A1 and A2 sequences arose purely by chance; thus, these loci are probably located in the nonrecombining mating-type-specific region of the M. violaceum A1 and A2 chromosomes.

TABLE 1

Loci from mating-type-specific chromosomes of M. violaceum used for PCR analysis and genetic map construction
With nonzero A1/A2 divergenceb
LociMating type specific<1%>1%With zero A1/A2 divergencebTotal
A1a52 (1)3 (3)35 (3)45 (7)
A2a62 (0)7 (7)26 (3)41 (10)
Subtotal4 (1)10 (10)
Total1114 (11)61 (6)86 (17)
Open in a separate windowaA1, loci originated from the A1 sex chromosome; A2, loci originated from the A2 sex chromosome.bThe number of loci used for genetic map construction is in parentheses.To confirm the mating-type-specific or pseudoautosomal locations of the loci with and without A1/A2 divergence, we conducted genetic mapping in a family of 99 individuals, 50 of which were of mating type A1 and 49 of mating type A2. The family was generated by a cross between A1 and A2 M. violaceum strains from S. latifolia accessions Sl405 (Sweden) and Sl127 (France), respectively. The choice of strains from geographically distant locations was motivated by the hope of maximizing the number of DNA sequence differences between them that can be used as molecular genetic markers in segregation analysis. We inoculated S. latifolia seedlings with sporidial cultures of both mating types. For inoculation, petri dishes with 12-day-old seedlings of S. latifolia were flooded with 2.5 ml of inoculum suspension. Inoculum suspension consisted of equal volumes of the A1 and A2 sporidial cultures that were mixed and conjugated overnight at 14° under rotation (Biere and Honders 1996; Van Putten et al. 2003). Seedlings were potted 3 days after inoculation. Two months later, teliospores were collected from the flowers of the infected plant and grown in petri dishes on 3.6% potato dextrose agar medium. Haploid sporidia formed after meiosis were isolated and grown as separate cultures for DNA extraction. The mating types of single sporidia cultures were identified as described above. The loci analyzed in the segregation analysis were sequenced in the two parental haploid strains and in 99 (50 A1 and 49 A2) haploid strains that were generated in the cross. Single nucleotide differences between the parental strains were used as molecular genetic markers for segregation analysis in the progeny. The genetic map was constructed using MAPMAKER/EXP v3.0 (Lincoln et al. 1992) and MapDisto v1.7 (http://mapdisto.free.fr/).The resulting genetic map is shown in Figure 2. As expected, no recombination was observed between the 10 loci with diverged A1- and A2-linked copies. In addition, one marker with no A1/A2 divergence, A2-397, was also completely linked to the loci with significant A1/A2 divergence. This locus either may be very tightly linked to the nonrecombining mating-type-specific region or may have been added to that region more recently than the loci that had already accumulated some divergence between the alleles in the two mating types. The mating-type-specific pheromone receptor locus (Devier et al. 2009) and 11 mating-type-specific loci are also located in this nonrecombining region (Figure 2). Interestingly, the cluster of nonrecombining markers is flanked on both sides with markers that recombine in meiosis, demonstrating that there are pseudoautosomal regions on both ends of the mating-type-specific chromosomes.Open in a separate windowFigure 2.—Genetic map of the mating-type-determining chromosome in M. violaceum. Genetic distance (in centimorgans) and the relative positions of the markers are shown to the left and the right of the chromosome, respectively. The position of the nonrecombining region corresponds to the cluster of linked markers shown on the right of the figure. Total A1/A2 divergence is shown in parentheses. Eleven mating-type-specific markers (for which sequences are available from only one mating type), located in the nonrecombining mating-type-specific region, are not shown.Our results demonstrate that although the loci reported by Hood et al. (2004) were isolated from the A1 and A2 chromosomes, most of these loci are not located in the nonrecombining mating-type-specific regions. In fact, the nonrecombining region might be relatively small: of 86 tested fragments, only 21 appeared to be either mating type specific or linked to the mating-type locus. Assuming that these loci represent a random set of DNA fragments isolated from the A1 and A2 chromosomes, it is possible to estimate the size of the nonrecombining region using the binomial distribution: the nonrecombining region is expected to be 24.4% (95% CI: 16.7–33.6%) of the chromosome length. As the sizes of the A1 and A2 chromosomes are ∼3.4 and 4.2 Mb long (Hood 2002; Hood et al. 2004), the nonrecombining region might be ∼1 Mb long.Interestingly, total A1/A2 divergence for the 11 loci with A1- and A2-linked copies mapped to the nonrecombining region varied from 0% to 8.6% (Figure 2). In addition, 11 loci amplified from only one mating type. These genes could represent degenerated genes, some of which degenerated in A1 strains, and some in A2 strains. Alternatively, they might be highly diverged genes, such that the PCR primers amplify only one allele, and not the other. Variation in divergence may be the result of the stepwise cessation of recombination between the A1 and A2 chromosomes in M. violaceum, resembling the evolutionary strata reported for human, chicken, and white campion sex chromosomes (Lahn and Page 1999; Handley et al. 2004; Bergero et al. 2007). However, only the differences between the most and the least diverged loci are statistically significant (Devier et al. 2009), the M. violaceum mating-type region has at least three strata: one oldest stratum, including the pheromone receptor locus; a younger stratum with ∼5–9% A1/A2 divergence; and the youngest stratum with 1–4% divergence between the two mating types. There may also be an additional very recently evolved stratum containing the locus named A2-397, which is also present in all A1 strains tested, with no fixed differences between the A1 and A2 strains (
No. of sites analyzedWithin A1
Within A2
Fixed differences between A1 and A2A1/A2 divergence (%)
LociaSb totalSπ (%)cSπ (%)c
A1/A2 divergence <1%A1-23645630020.4410.44
A1-0456544000040.61
A2-568413220.4820.4800.24
A2-411480210.210010.31
A1/A2 divergence >1%A1-2176679000091.35
A1-12856990010.1881.49
A1-199618130010.16122.02
A2-4223449000092.62
A2-516470140000142.98
A2-404508200030.59173.64
A2-4355062220.3920.39183.95
A2-4734572310.2210.22214.81
A2-4573031710.3300165.54
A2-5755034750.9930.59398.55
Open in a separate windowaA1, loci originated from the A1 sex chromosome; A2, loci originated from the A2 sex chromosome.bS, number of polymorphic sites.cπ (%), average number of differences per 100 nucleotides.

TABLE 3

P-values for the 2 × 2 G-tests for significance of differences in A1/A2 divergence between the loci in the nonrecombining region
LaSbLocusA2-397A1-217A1-128A1-199A2-422A2-516A2-404A2-435A2-473A2-457
5190A2-397
6679A1-2170.006
5698A1-1280.0060.93
61812A1-1990.00070.410.48
3449A2-4220.00030.170.210.51
47014A2-5160.000030.060.0860.280.76
50817A2-40400.0250.0380.150.550.75
50618A2-43500.0150.0240.1040.450.620.86
45721A2-47300.0010.0030.01630.150.210.340.43
30316A2-45700.00090.00170.00970.090.130.2030.260.69
50339A2-5750000.000010.0020.0020.0030.0060.0550.199
Open in a separate windowP-values <0.05 are in boldface type.aL, the length of the region compared.bS, the number of nucleotide differences observed.As most of the loci isolated from the A1 and A2 chromosomes recombine in meiosis, they are not expected to degenerate. Thus, the observation of a higher proportion of TEs in these loci, compared to other chromosomes (Hood et al. 2004), is unlikely to reflect genetic degeneration attributable to a lack of recombination in these loci. A higher abundance of TEs in the sequences isolated from the A1 and A2 chromosomes, as reported by Hood et al. (2004), may simply reflect variation in the TE density across the genome. Thus, it remains to be seen whether M. violaceum mating-type-specific regions degenerate, similar to vertebrate Y (or W) chromosomes, or remain largely intact, as in C. neoformans (Lengeler et al. 2002). If the latter were the case, it may suggest that nonrecombining regions in fungi do not necessarily follow the same degenerative path as animal Y and W chromosomes. The analysis of sequences from the M. violaceum genome (and perhaps other fungal genomes) will hopefully provide the answer to this question.The lack of degeneration of mating-type-specific regions in C. neoformans may be due to the relatively small size of the nonrecombining regions. The 20 genes present in this region may not be sufficient for the operation of such detrimental population genetic processes as background selection or Muller''s ratchet because the speed of these processes depends critically on the number of active genes linked together (Charlesworth 2008). Larger mating-type-specific regions in M. violaceum might contain more genes; thus, more active genetic degeneration may be expected in this species. Indeed, many strains of M. violaceum show haplolethality linked to one of the mating types (Hood and Antonivics 2000; Thomas et al. 2003; Tellier et al. 2005), which may reflect the accumulation of deleterious mutations in the nonrecombining regions around the mating-type loci. Mating-type specificity of the markers that amplified in only A1 or A2 strains in this study may also reflect genetic degeneration.Another factor that may potentially prevent degeneration of genes linked to mating-type loci in fungi is the haploid expression of genes in these regions. In animals, many Y-linked genes have functional homologs on the X chromosome, and loss of the Y-linked gene may be compensated for by expression of the X-linked homologs. The haploid stage in an animal''s life cycle is very short, and very few genes are actively expressed in animal gametes (Schultz et al. 2003). In plants, on the other hand, a significant proportion of the genome is expressed in pollen (da Costa-Nunes and Grossniklaus 2003), and so the loss of Y-linked genes expressed in gametes may be more detrimental than in animals. Indeed, most genes isolated from the white campion X chromosome have intact Y-linked copies (Filatov 2005; Bergero et al. 2007), but due to the small number of genes available, it is still unclear whether genetic degeneration of Y-linked genes is indeed slower in this species (and in plants generally) compared to animal Y chromosomes. Haploid expression could be an even more powerful force in fungi and other organisms with haploid sexes, such as bryophytes, as most genes are expressed in the haploid stage. Further analysis of genetic degeneration in nonrecombining sex- or mating-type-specific regions in fungi and bryophytes will help to shed light on this question.  相似文献   

17.
Novel GH10 Xylanase,with a Fibronectin Type 3 Domain,from Cellulosimicrobium sp. Strain HY-13, a Bacterium in the Gut of Eisenia fetida     
Do Young Kim  Mi Kyoung Han  Doo-Sang Park  Jong Suk Lee  Hyun-Woo Oh  Dong-Ha Shin  Tae-Sook Jeong  Sung Uk Kim  Kyung Sook Bae  Kwang-Hee Son  Ho-Yong Park 《Applied and environmental microbiology》2009,75(22):7275-7279
The gene encoding a novel modular xylanase from Cellulosimicrobium sp. strain HY-13 was identified and expressed in Escherichia coli, and its truncated gene product was characterized. The enzyme consisted of three distinct functional domains, an N-terminal catalytic GH10 domain, a fibronectin type 3 domain, and C-terminal carbohydrate-binding module 2.Most known microbial xylanases, which decompose primarily β-1,4-xylosic polysaccharides in an endo fashion, are currently affiliated with the two glycoside hydrolase (GH) families 10 and 11. Compared to GH11 xylanases, GH10 xylanases generally have a molecular mass of >30 kDa and an acidic pI. In addition, GH10 xylanases are frequently found in nature as modular enzymes that consist of a catalytic GH10 domain with one or more substrate-binding domains, such as a cellulose-binding domain, carbohydrate-binding module (CBM), or xylan-binding domain (1, 5, 12). However, no modular xylanase with a fibronectin type 3 (Fn3) domain has been characterized to date, even though Fn3 modules are often found in bacterial carbohydrolases such as cellulases, amylases, pullulanases, polygalacturonidases, and chitinases. It is now believed that the Fn3 domains in bacterial carbohydrolases participate in promotion of the hydrolysis of carbohydrate substrates by modifying their surfaces (10, 17).Gut microorganisms from invertebrates have recently attracted a great deal of attention as sources of novel fibrolytic enzymes with unique molecular structures and distinct substrate specificities (2-4, 7, 15). However, no study has been conducted to evaluate xylanolytic enzymes from the gut microorganisms of earthworms that may participate in the digestion of cellulosic or hemicellulosic foods taken up by the hosts. Here, we report a novel GH10 xylanase (XylK1) with an Fn3 domain from Cellulosimicrobium sp. strain HY-13 KCTC 11302BP (11), which was isolated from the digestive tract of the earthworm Eisenia fetida.Amplification of a partial sequence of the Cellulosimicrobium sp. strain HY-13 xylanase gene from the genomic DNA was conducted using the degenerate primers designed on the basis of conserved regions (WDVVNE and ITELDI) in the GH10 xylanases. The upstream primer (KF) was 5′-TGGGACGTCSTCAACGAG-3′, and the downstream primer (KR) was 5′-GATGTCGAGCTCSGTGAT-3′, which produced a 342-bp DNA fragment. Cloning of the full xylK1 gene was performed by repeated genomic walking and nested-PCR methods using a DNA Walking SpeedUp premix kit (Seegene). To overproduce mature XylK1, its encoding gene was cloned into the NdeI/HindIII sites of a pET-28a(+) vector (Novagen). Likewise, a partial sequence containing the GH10 domain (Ala34 to Leu345) of XylK1 was amplified using the primers tKF (5′-CATATGGCCACCGAGCCGCTCG-3′) and tKR (5′-AAGCTTTCAGGACCTCGGCGATCGC-3′) and subsequently also cloned into the same expression vector. When overexpressed in recombinant Escherichia coli BL21 cells harboring pET-28a(+)/xylK1, most recombinant proteins (rXylK1) were produced as inactive inclusion bodies. Therefore, after solubilization of the isolated inclusion bodies, on-column refolding and purification of rXylK1 was conducted using a HisTrap HP (GE Healthcare, Sweden) (5-ml) column attached to a fast-performance liquid chromatography (LC) system (Amersham Pharmacia Biotech, Sweden) according to the manufacturer''s instructions. The active rXylK1 proteins were then purified to electrophoretic homogeneity by gel permeation chromatography using a HiLoad 26/60 Superdex 200 prep-grade (Amersham Biosciences, Sweden) column, as previously described (11). Like rXylK1, the recombinant proteins (rXylK1ΔFn3) without both an Fn3 domain and CBM 2 were also purified by the method described above because they were produced as insoluble inclusion bodies. The relative molecular mass of the denatured rXylK1 was evaluated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis on a 12% gel, and the protein concentrations were assayed by using the Bradford reagent (Bio-Rad). Matrix-assisted laser desorption ionization-time-of-flight mass spectrometry (MALDI-TOF MS) analysis was conducted using an Ultraflex III MALDI-TOF mass spectrometer (Bruker Daltonics, Germany) at the Korea Basic Science Institute (Daejeon, South Korea). The binding capacity of recombinant enzymes with/without the Fn3 domain to carbohydrate polymers was determined as described elsewhere (4). Xylanase activity was routinely assayed by measuring the amount of reducing sugars released from birch wood xylan by using the 3,5-dinitrosalicylic acid reagent. The standard assay mixture (0.5 ml) consisted of birch wood xylan (1.0%) or p-nitrophenyl (PNP)-sugar derivatives (5 mM) with suitably diluted enzyme solution (0.05 ml) in 50 mM sodium phosphate buffer (pH 6.0), and the catalytic reaction was performed at 55°C for 10 min. One international unit of xylanase activity for xylans or PNP-sugar derivatives was defined as the amount of enzyme required to produce 1 μmol of reducing sugar or PNP, respectively, per min under standard assay conditions. Enzymatic hydrolysis of birch wood xylan (10 mg) (Sigma Co.), xylooligosaccharides (1 mg each) (Megazyme International Ireland, Ireland), and cellooligosaccharides (1 mg each) (Seikagaku Biobusiness Co., Japan) was conducted using purified rXylK1 (2 μg) in 0.1 ml of 50 mM sodium phosphate buffer (pH 6.0) for 3 or 6 h at 37°C, during which time the enzyme remained fairly stable. The reaction mixture was then heated to 100°C for 5 min to stop the enzyme reaction. The hydrolysis products were identified by LC-MS, as previously described (11).The isolated XylK1 gene (GenBank accession no. FJ859907) contained a 1,671-bp open reading frame that encodes a protein of 556 amino acids with a deduced molecular mass of 58,296 Da and a calculated pI of 4.59. It was predicted that the signal sequence cleavage site of premature XylK1 was between Ala33 and Ala34, which may generate a mature XylK1 of 523 amino acids with a deduced molecular mass of 54,843 Da and a calculated pI of 4.49 (Fig. (Fig.1).1). The results of a protein BLAST survey revealed that XylK1 was a unique modular xylanase composed of an N-terminal catalytic GH10 (Leu38-to-Asp330) domain, an Fn3 (Pro359-to-Gly430) domain, and a C-terminal CBM 2 (Cys454 to Cys553), which was very comparable to the domain architectures of other related GH10 enzymes (Fig. (Fig.2).2). To the best of our knowledge, no xylanase with domain architecture identical to that of XylK1 with an Fn3 domain has been reported to date, although an uncharacterized modular xylanase (GenBank accession no. ABQ06877) consisting of an N-terminal Fn3 domain and a C-terminal GH10 domain from Flavobacterium johnsoniae UW101 was previously identified through a genome survey. As shown in Fig. Fig.1,1, the catalytic GH10 domain of XylK1 showed the highest sequence identity (67%) with that of the Cellulomonas fimi xylanase (AAA56792) among other GH10 enzymes available in the NCBI database. However, its CBM 2 was 64% identical to that of Cellulomonas fimi GH6 cellulase (AAC36898). The highest sequence identity (70%) of the Fn3 domain in XylK1 was obtained when it was compared to that of the Acidothermus cellulolyticus 11B GH48 enzyme (ABK52390), which degrades cellulose. The two conserved residues of Glu161 (acid/base catalyst) and Glu266 (catalytic nucleophile) were predicted in the active site of premature XylK1.Open in a separate windowFIG. 1.Alignment of the deduced amino acid sequence of GH10 xylanase from Cellulosimicrobium sp. strain HY-13 with those of other GH10 xylanases. Shown are sequences (GenBank accession numbers) of Cellulosimicrobium sp. strain HY-13 (Csp) xylanase (FJ859907), Cellulomonas fimi (Cfi) xylanase (AAA56792), Streptomyces coelicolor (Sco) A3 xylanase (CAB61191), Streptomyces ambofaciens (Sam) xylanase (CAJ88420), Acidothermus cellulolyticus 11B (Ace) xylanase (ABK51955), and Thermobifida alba (Tal) xylanase (CAB02654). The identical and similar amino acids are shown by black and gray boxes, respectively. The predicted signal peptide is indicated by a black bar. The internal peptide sequences used in the design of degenerate oligonucleotides for PCR are marked by arrows. Highly conserved amino acid residues that play an essential role in the catalytic reaction are indicated by asterisks. GH10, Fn3, and CBM 2 domains are outlined by solid, dashed, and dotted lines, respectively.Open in a separate windowFIG. 2.Domain architectures of Cellulosimicrobium sp. strain HY-13 xylanase and the following related bacterial GH10 xylanases (GenBank accession numbers): Cellulosimicrobium sp. strain HY-13 xylanase (FJ859907) (A), Streptomyces thermocarboxydus HY-15 xylanase (EU880430) (B), Cellulomonas fimi xylanase (AAZ76373) (C), Streptomyces thermoviolaceus xylanase (BAD02382) (D), Pseudomonas fluorescens xylanase (P23030) (E), and Cellvibrio japonicus xylanase (YP 001982932) (F).The molecular mass of the purified enzyme was estimated to be approximately 42.0 kDa, which was smaller than the deduced molecular mass (57,138 Da) of intact His-tagged rXylK1, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (data not shown). In addition, MALDI-TOF MS analysis revealed that the purified His-tagged rXylK1 with a calculated molecular mass of 45,169 Da was a smaller protein than the intact rXylK1. These results indicate that rXylK1 was formed by proteolytic cleavage at the C terminus region because the enzyme was able to tightly bind to a His tag column. Based on the calculated molecular mass (45,169 Da) of the truncated rXylK1, it is assumed that the intact rXylK1 was processed at the Val439-Thr440 site in a hinge region between the Fn3 domain and the C-terminal CBM 2 of the premature XylK1. The deduced molecular mass (45,179 Da) of rXylK1 with the Val439 residue at the C-terminal extremity was very close to the molecular mass (45,169 Da) of the enzyme calculated by MALDI-TOF MS analysis. A similar C-terminal processing of some modular xylanases with a cellulose-binding domain by proteases has also been observed when they are expressed in E. coli (8, 14). It is likely that the C-terminal truncation does not induce a significant alteration of the binding affinity of rXylK1 with an Fn3 domain to carbohydrate polymers since the truncated enzyme could still bind to both Avicel and insoluble oat spelt xylan (Table (Table1).1). In this case, only weak catalytic activity of the rXylK1 (<3% of its original activity [0.5 IU]) was recovered from an enzyme solution after binding. It was of great interest that no rXylK1ΔFn3 was bound to Avicel, although the enzyme could still not only bind to insoluble oat spelt xylan but also catalyze the hydrolysis of xylosic polymers. The specific activity (27 IU/mg) of rXylK1ΔFn3 for birch wood xylan was evaluated to be approximately 19% of that (143 IU/mg) of rXylK1 with the Fn3 domain for the same substrate. Taken together, the binding ability of the C-terminal CBM 2-lacking rXylK1 to Avicel and insoluble oat spelt xylan clearly suggests that the Fn3 domain plays an important role in enzyme-substrate binding because rXylK1ΔFn3 was not bound to Avicel. A significant decrease in the catalytic activity of rXylK1ΔFn3 induced by deletion of the Fn3 domain also suggests that the Fn3 domain may take part in the promotion of the catalytic hydrolysis of xylosic substrates by modifying their surfaces, as shown in other GHs (10, 17). The maximum catalytic activity of rXylK1 toward birch wood xylan was observed at pH 6.0 and 55°C, and it maintained over 80% of its highest activity at a relatively broad pH range of 5.0 to 9.0 during the reaction period of 15 min. These high activities of rXylK1 in alkaline pH ranges suggest that it is a peculiar enzyme that can be distinguished from xylanases of other invertebrate-symbiotic microorganisms, which showed weak hydrolytic activities toward xylan at the same alkaline pHs (4, 11, 15). At 55°C, the half-life of rXylK1 was approximately 10 min, which indicates that it is a typical mesophilic enzyme. Compared to many other xylanases that were completely inhibited by Hg2+ (11, 13), rXylK1 was partially inhibited (by 40% relative to its original activity) by 1 mM Hg2+. In addition, the enzyme was relatively suppressed by <25%, relative to its original activity, in the presence of some divalent cations at a concentration of 1 mM, in the order of Ca2+ > Cu2+ > Ba2+. No significant alterations of rXylK1 activity by Mn2+ and Co2+ were interesting to note because the xylanases from Streptomyces sp. strain S9 [12] and Aeromonas caviae ME-1 [14] have been negatively affected by the compound. In this study, the catalytic activity of rXylK1 increased by approximately 1.4-fold when the reaction was conducted in the presence of 1 mM Fe2+. The promotion of rXylK1 activity by Fe2+ was comparable to previous observations of xylanase inhibition by the metal ion (6, 15). The rXylK1 was relatively unaffected by sulfhydryl reagents (5 mM) such as sodium azide, iodoacetamide, and N-ethylmaleimide, while the enzyme lost 68% of its original activity when preincubated with 5 mM EDTA for 10 min. The complete inhibition of rXylK1 by 5 mM N-bromosuccinimide was in good agreement with the fact that three Trp residues in the highly conserved region of the GH10 enzymes are critically involved in enzyme-substrate interaction, as shown for Streptomyces lividans (16) and Geobacillus stearothermophilus T-6 (18) GH10 xylanases. It was predicted that the three residues Trp118, Trp306, and Trp314 in premature XylK1 might be responsible for catalysis and substrate binding of the enzyme. It is also noteworthy that the catalytic activity of His-tagged rXylK1 increased significantly, by approximately 1.8-fold, when the reaction was conducted in the presence of Tween 80 or Triton X-100 at a concentration of 0.5%. It is assumed that the nonionic-detergent-induced activation of His-tagged rXylK1 might be due to the direct interaction of the recombinant enzyme with the Tween 80 or Triton X-100 molecule, which may lead to an alteration of the enzyme-substrate interaction. Indeed, the stimulation of His-tagged rXylK1 activity was insignificant when the enzyme reaction was conducted in the presence of the detergents without preincubation with the same compounds for 10 min. As with rXylK1, it has been reported that the catalytic activity of a His-tagged esterase from Bacillus megaterium 20-1 expressed in E. coli is greatly stimulated by various nonionic detergents (9).

TABLE 1.

Binding of rXylK1 and rXylK1ΔFn3 to hydrophobic polysaccharides
SubstrateResidual xylanase activity after binding (total IU)a
rXylK1rXylK1ΔFn3
Control0.500.50
Avicel≤0.010.49 ± 0.02
Insoluble oat spelt xylan0.05 ± 0.01≤0.01
Open in a separate windowaResidual xylanase activity was assayed using birch wood xylan.Of the evaluated xylosic materials, oat spelt xylan was most efficiently hydrolyzed by rXylK1; however, the enzyme was not capable of degrading glucose-based polysaccharides, which is indicative of a lack of other GH activities (Table (Table2).2). It should also be noted that the catalytic activity of rXylK1 toward PNP-cellobioside was approximately 1.7-fold higher than that (193 IU/mg) of the enzyme toward oat spelt xylan. In this study, the cleavage activity of rXylK1 for PNP-cellobioside was approximately 48 IU/mg, which is much higher than the activity (<10 IU/mg) of other known xylanases for the same substrate (6, 11). However, LC-MS analysis revealed that cellooligosaccharides of cellobiose to cellotetraose were not susceptible to rXylK1 (data not shown). Taken together, these results indicate that rXylK1 is a true endo-β-1,4-xylanase that lacks cellulase activity. Interestingly, rXylK1 was found to have transxylosylation activity (approximately 7.5% of its maximum hydrolytic activity for oat spelt xylan) that enabled the cleavage of PNP-xylopyranoside (Table (Table2).2). A similar transxylosylation reaction by rXylK1 was also observed when xylotriose (X3) and xylotetraose (X4) were subjected to hydrolytic reaction by the enzyme (Table (Table3).3). Specifically, a series of xylooligosaccharides (X4 to X7) were produced after the enzymatic hydrolysis of X3 for 3 h at 37°C, although X2 and X3 were identified as the major products. Similarly, the hydrolysis of X4 by rXylK1 resulted in the production of a mixture that contained longer xylooligosaccharides (42.3%) of X5 to X8, which suggests that these xylooligomers were produced by an rXylK1-catalyzed transxylosylation reaction. However, no X1 was detected as the hydrolysis product of X2, X3, or X4. The ability of rXylK1 to catalyze the synthesis of longer xylooligosaccharides from X3 or X4 was of interest because microbial xylanases generally produced shorter xylooligosaccharides, such as X2 and/or X3, from the same substrates (2, 15). Additionally, rXylK1 primarily degraded birch wood xylan to X2 (65.1%) and X3 (29.5%) together with small amounts of X4 (5.4%) when the enzyme reaction was conducted for 6 h at 37°C.

TABLE 2.

Hydrolysis activity of rXylK1 for different substrates
SubstrateRelative activity (%)a
Birch wood xylan74.1 ± 2.8
Beech wood xylan85.8 ± 3.5
Oat spelt xylan100.0
Soluble starchND
AvicelND
Carboxy methylcelluloseND
PNP-cellobioside171.7 ± 4.9
PNP-glucopyranoside<0.5
PNP-xylopyranoside7.5 ± 0.6
Open in a separate windowaRelative activity was obtained from three repeated experiments. ND, not detected.

TABLE 3.

LC analysis of the hydrolysis products of xylosic materials by rXylK1
SubstrateComposition (%)a of products formed by hydrolysis reaction
X2X3X4X5X6X7X8
X2100.0
X327.845.416.85.63.90.5
X412.826.318.618.114.58.41.3
Birch wood xylan65.129.55.4
Open in a separate windowaLC area percent.In conclusion, the novel gene (xylK1) encoding a modular GH10 xylanase that consists of three putative functional domains (an N-terminal GH10 domain, an Fn3 domain, and C-terminal CBM 2) was identified from an earthworm-symbiotic bacterium, Cellulosimicrobium sp. strain HY-13. The molecular architecture of XylK1 indicates that it is a unique GH10 enzyme with an Fn3 domain that has not previously been reported. In addition, the relatively high cleavage activity of rXylK1 toward PNP-cellobioside and its transxylosylation activity that enables it to produce longer xylooligosaccharides from X3 or X4 differentiate it from other known GH10 xylanases.  相似文献   

18.
Involvement of Two Transport Systems and a Specific Porin in the Uptake of Phthalate by Burkholderia spp.     
Hung-Kuang Chang  Jonathan J. Dennis  Gerben J. Zylstra 《Journal of bacteriology》2009,191(14):4671-4673
  相似文献   

19.
Spontaneous Quinolone Resistance in the Zoonotic Serovar of Vibrio vulnificus     
Francisco J. Roig  A. Llorens  B. Fouz  C. Amaro 《Applied and environmental microbiology》2009,75(8):2577-2580
This work demonstrates that Vibrio vulnificus biotype 2, serovar E, an eel pathogen able to infect humans, can become resistant to quinolone by specific mutations in gyrA (substitution of isoleucine for serine at position 83) and to some fluoroquinolones by additional mutations in parC (substitution of lysine for serine at position 85). Thus, to avoid the selection of resistant strains that are potentially pathogenic for humans, antibiotics other than quinolones must be used to treat vibriosis on farms.Vibrio vulnificus is an aquatic bacterium from warm and tropical ecosystems that causes vibriosis in humans and fish (http://www.cdc.gov/nczved/dfbmd/disease_listing/vibriov_gi.html) (33). The species is heterogeneous and has been subdivided into three biotypes and more than eight serovars (6, 15, 33; our unpublished results). While biotypes 1 and 3 are innocuous for fish, biotype 2 can infect nonimmune fish, mainly eels, by colonizing the gills, invading the bloodstream, and causing death by septicemia (23). The disease is rapidly transmitted through water and can result in significant economic losses to fish farmers. Surviving eels are immune to the disease and can act as carriers, transmitting vibriosis between farms. Interestingly, biotype 2 isolates belonging to serovar E have been isolated from human infections, suggesting that serovar E is zoonotic (2). This serovar is also the most virulent for fish and has been responsible for the closure of several farms due to massive losses of fish. A vaccine, named Vulnivaccine, has been developed from serovar E isolates and has been successfully tested in the field (14). Although the vaccine provides fish with long-term protection from vibriosis, at present its use is restricted to Spain. For this reason, in many fish farms around the world, vibriosis is treated with antibiotics, which are usually added to the food or water.Quinolones are considered the most effective antibiotics against human and fish vibriosis (19, 21, 31). These antibiotics can persist for a long time in the environment (20), which could favor the emergence of resistant strains under selective pressure. In fact, spontaneous resistances to quinolones by chromosomal mutations have been described for some gram-negative bacteria (10, 11, 17, 24, 25, 26). Therefore, improper antibiotic treatment of eel vibriosis or inadequate residue elimination at farms could favor the emergence of human-pathogenic serovar E strains resistant to quinolones by spontaneous mutations. Thus, the main objective of the present work was to find out if the zoonotic serovar of biotype 2 can become quinolone resistant under selective pressure and determine the molecular basis of this resistance.Very few reports on resistance to antibiotics in V. vulnificus have been published; most of them have been performed with biotype 1 isolates. For this reason, the first task of this study was to determine the antibiotic resistance patterns in a wide collection of V. vulnificus strains belonging to the three biotypes that had been isolated worldwide from different sources (see Table S1 in the supplemental material). Isolates were screened for antimicrobial susceptibility to the antibiotics listed in Table S1 in the supplemental material by the agar diffusion disk procedure of Bauer et al. (5), according to the standard guideline (9). The resistance pattern found for each isolate is shown in Table S1 in the supplemental material. Less than 14% of isolates were sensitive to all the antibiotics tested, and more than 65% were resistant to more than one antibiotic, irrespective of their biotypes or serovars. The most frequent resistances were to ampicillin-sulbactam (SAM; 65.6% of the strains) and nitrofurantoin (F; 60.8% of the strains), and the least frequent were to tetracycline (12%) and oxytetracycline (8%). In addition, 15% of the strains were resistant to nalidixic acid (NAL) and oxolinic acid (OA), and 75% of these strains came from fish farms (see Table S1 in the supplemental material). Thus, high percentages of strains of the three biotypes were shown to be resistant to one or more antibiotics, with percentages similar to those found in nonbiotyped environmental V. vulnificus isolates from Asia and North America (4, 27, 34). In those studies, resistance to antibiotics could not be related to human contamination. However, the percentage of quinolone-resistant strains found in our study is higher than that reported in other ones, probably due to the inclusion of fish farm isolates, where the majority of quinolone-resistant strains were concentrated. This fact suggests that quinolone resistance could be related to human contamination due to the improper use of these drugs in therapy against fish diseases, as has been previously suggested (18, 20). Although no specific resistance pattern was associated with particular biotypes or serovars, we found certain differences in resistance distribution, as shown in Table Table1.1. In this respect, biotype 3 displayed the narrowest spectrum of resistances and biotype 1 the widest. The latter biotype encompassed the highest number of strains with multiresistance (see Table S1 in the supplemental material). Within biotype 2, there were differences among serovars, with quinolone resistance being restricted to the zoonotic serovar (Table (Table11).

TABLE 1.

Percentage of resistant strains distributed by biotypes and serovars
V. vulnificusNo. of isolatesResistance distribution (%) for indicated antibiotica
SAMCTXENALFOTOASXT-TMPTE
Biotype 14975.524.514.330.683.78.230.628.68.2
Biotype 2 (whole)7258.313.912.54.247.29.74.24.213.9
Biotype 2
    Serovar E3630.312.139.127.315.29.1321.2
    Serovar A231009.118.2077.3009.14.6
    Nontypeable82914.325057.114.30014.3
    Serovar I5100202002020000
Biotype 3510002008000020
Open in a separate windowaCTX, cefotaxime; E, erythromycin; OT, oxytetracycline; SXT-TMP, sulfamethoxazole-trimethoprim; TE, tetracycline.The origin of resistance to quinolones in the zoonotic serovar was further investigated. To this end, spontaneous mutants of sensitive strains were selected from colonies growing within the inhibition halo around OA or NAL disks. Two strains (strain CG100 of biotype 1 and strain CECT 4604 of biotype 2, serovar E) developed isolated colonies within the inhibition zone. These colonies were purified, and maintenance of resistance was confirmed by serial incubations on medium without antibiotics. Using the disk diffusion method, CG100 was shown to be resistant to SAM and F and CECT 4604 to F (see Table S1 in the supplemental material). The MICs for OA, NAL, flumequine (UB), and ciprofloxacin (CIP) were determined by using the microplate assay according to the recommendations of the Clinical and Laboratory Standards Institute and the European Committee for Antimicrobial Susceptibility Testing of the European Society of Clinical Microbiology and Infectious Diseases (8, 12) and interpreted according to the European Committee for Antimicrobial Susceptibility Testing of the European Society of Clinical Microbiology and Infectious Diseases (13). The MICs for OA and NAL and for the fluoroquinolones UB and CIP exhibited by the mutants and their counterparts are shown in Table Table2.2. The inhibition zone diameters correlated well with MICs (data not shown). Mutants FR1, FR2, FR3, and FR4 were resistant to NAL and sensitive to the remaining quinolones, although they showed higher resistances than their parental strains (Table (Table2).2). Thus, these four mutants showed increases of 32- to 128-fold for NAL MICs, 4- to 8-fold for UB MICs, and 16-fold for CIP MICs (Table (Table2).2). The fifth mutant, FR5, was resistant to the two tested quinolones and to UB, a narrow-spectrum fluoroquinolone. This mutant, although sensitive to CIP, multiplied its MIC for this drug by 128 with respect to the parental strain (Table (Table22).

TABLE 2.

MICs for quinolones and fluoroquinolones and mutations in gyrA, gyrB, and parC detected in naturally and artificially induced resistant strains
Strain(s)MIC (μg ml−1) for indicated antibioticb
Gene mutationa
gyrA
gyrB
parC
Position
Codon changeaa changePosition
Codon changeaa changePosition
Codon changeaa change
NALOAUBCIPntaantaantaa
CG1000.5 (S)0.125 (S)0.0625 (S)0.0078 (S)
FR116 (R)1 (S)0.25 (S)0.125 (S)24883AGT→ATTS→INCNCNCNCNCNCNCNC
FR216 (R)1 (S)0.25 (S)0.125 (S)24883AGT→ATTS→INCNCNCNCNCNCNCNC
CECT 46040.25 (S)0.0625 (S)0.0625 (S)0.0078 (S)
FR332 (R)2 (S)0.5 (S)0.125 (S)24883AGT→ATTS→INCNCNCNCNCNCNCNC
FR432 (R)2 (S)0.5 (S)0.125 (S)24883AGT→ATTS→INCNCNCNCNCNCNCNC
FR5256 (R)16 (R)16 (R)1 (S)24883AGT→ATTS→I1156386GCA→ACAA→T25485TCA→TTAS→L
1236412CAG→CACQ→H
CECT 4602128 (R)8 (R)64 (R)1 (S)24883AGT→ATTS→INCNCNCNC25485TCA→TTAS→L
CECT 4603, CECT 4606, CECT 4608, PD-5, PD-12, JE32 (R)2 (S)<1 (S)<1 (S)24883AGT→ATTS→INCNCNCNCNCNCNCNC
CECT 486264 (R)2 (S)2 (S)<1 (S)24983AGT→AGAS→RNCNCNCNCNCNCNCNC
A2, A4, A5, A6, A7, PD-1, PD-364-128 (R)2 (S)4 (S)<1 (S)24883AGT→ATTS→INCNCNCNC338113GCA→GTAA→V
V1128 (R)4 (S)4 (S)<1 (S)24883AGT→ATTS→I1274425GAG→GGGE→GNCNCNCNC
1314438AAC→AAAN→K
Open in a separate windowaMutations in a nucleotide (nt) that gave rise to a codon change and to a change in amino acids (aa) are indicated. NC, no change detected.bThe resistance (R) or sensitivity (S) against the antibiotic determined according to the Clinical and Laboratory Standards Institute and the European Committee for Antimicrobial Susceptibility Testing of the European Society of Clinical Microbiology and Infectious Diseases (9, 13) is indicated in parentheses.For other gram-negative pathogens, quinolone resistance relies on spontaneous mutations in the gyrA, gyrB, parC, and parE genes that occur in a specific region of the protein known as the quinolone resistance-determining region (QRDR) (1, 11, 17, 24, 25, 26, 28). To test the hypothesis that mutations in these genes could also produce quinolone resistance in V. vulnificus, the QRDRs of these genes were sequenced in the naturally resistant strains and in the two sensitive strains that had developed resistances by selective pressure in vitro. The genomic DNA was extracted (3), and the QRDRs of gyrA, gyrB, parE, and parC were amplified using the primers shown in Table Table3,3, which were designed from the published genomes of biotype 1 strains YJ016 and CMCP6 (7, 22). PCR products of the predicted size were sequenced in an ABI 3730 sequencer (Applied Biosystems). Analysis of the QRDR sequences for gyrA, gyrB, parC, and parE of the mutants and the naturally resistant strains revealed that all naturally resistant strains, except one, shared a specific mutation at nucleotide position 248 with the laboratory-induced mutants (Table (Table2).2). This mutation gave rise to a change from serine to isoleucine at amino acid position 83. The exception was a mutation in the adjacent nucleotide that gave rise to a substitution of arginine for serine at the same amino acid position (Table (Table2).2). All the isolates that were resistant to the quinolone NAL had a unique mutation in the gyrA gene, irrespective of whether resistance was acquired naturally or in the laboratory (Table (Table2).2). This result strongly suggests that a point mutation in gyrA that gives rise to a change in nucleotide position 83 can confer resistance to NAL in V. vulnificus biotypes 1 and 2 and that this mutation could be produced by selective pressure under natural conditions. gyrA mutations consisting of a change from serine 83 to isoleucine have also been described in isolates of Aeromonas from water (17) and in diseased fish isolates of Vibrio anguillarum (26). Similarly, replacement of serine by arginine at amino acid position 83 in diseased fish isolates of Yersinia ruckeri (16) suggests that this mechanism of quinolone resistance is widespread among gram-negative pathogens. In all cases, these single mutations were also related to increased resistance to other quinolones (OA) and fluoroquinolones (UB and CIP) (Table (Table2),2), although the mutants remained sensitive according to the standards of the Clinical and Laboratory Standards Institute and the European Committee for Antimicrobial Susceptibility Testing of the European Society of Clinical Microbiology and Infectious Diseases (9, 13). A total of 50% of the naturally resistant strains, all of them of biotype 1, showed additional mutations that affected parC (a change in amino acid position 113) or gyrB (changes in amino acids at positions 425 and 438) (Table (Table2).2). These strains exhibited higher MICs for OA and fluoroquinolones (Table (Table2),2), although they were still sensitive to these drugs (9, 13). Finally, one isolate of biotype 2, serovar E, which was naturally resistant to quinolones and UB, showed a mutation in parC that gave rise to a substitution of leucine for serine at amino acid position 85 (Table (Table2).2). This mutation was shared only with the laboratory-induced mutant, also a biotype 2, serovar E mutant, which was resistant to the fluoroquinolone UB. The same mutation in parC had been previously described in diseased fish isolates of V. anguillarum that were highly resistant to quinolones (28), but this had not been related to fluoroquinolone resistance in Vibrio spp. nor in other gram-negative bacteria. These results strongly suggest that resistance to fluoroquinolones in V. vulnificus is related to specific mutations in gyrA and parC and that mutations in different positions for parC or in gyrB could contribute to increased resistance to quinolones and fluoroquinolones. Our results also agree with previous studies confirming that the acquisition of higher quinolone resistance is more probable when arising from a gyrA parC double mutation than from a gyrA gyrB double mutation (29).

TABLE 3.

Oligonucleotides used in this study
PrimerSequenceAnnealing temp (°C)Size (bp)
GyrAFGGCAACGACTGGAATAAACC55.8416
GyrARCAGCCATCAATCACTTCCGTC
ParCFCGCAAGTTCACCGAAGATGC56.6411
ParCRGGCATCCGCAACTTCACG
GyrBFCGACTTCTGGTGACGATGCG57.4642
GyrBRGACCGATACCACAACCTAGTG
ParEFGCCAGGTAAGTTGACCGATTG56.8512
ParERCACCCAGACCTTTGAATCGTTG
Open in a separate windowFinally, the evolutionary history for each protein was inferred from previously published DNA sequences of the whole genes from different Vibrio species after multiple sequence alignment with MEGA4 software (32) by applying the neighbor-joining method (30) with the Poisson correction (35). The distance tree for each whole protein showed a topology similar to the phylogenetic tree based on 16S rRNA analysis, with the two isolates of V. vulnificus forming a single group, closely related to Vibrio parahaemolyticus, Vibrio cholerae, V. anguillarum, and Vibrio harveyi (see Fig. S1A in the supplemental material). A second analysis was performed with the QRDR sequences of the different mutants and isolates of V. vulnificus (GenBank accession numbers FJ379836 to FJ379927) to infer the intraspecies relationships (see Fig. S1B in the supplemental material). This analysis showed that QRDRs of gyrA, gyrB, parC, and parE were highly homogeneous within V. vulnificus.In summary, the zoonotic serovar of V. vulnificus can mutate spontaneously to gain quinolone resistance, under selective pressure in vitro, due to specific mutations in gyrA that involve a substitution of isoleucine for serine at amino acid position 83. This mutation appears in biotype 2, serovar E diseased-fish isolates and biotype 1 strains, mostly recovered from fish farms. An additional mutation in parC, resulting in a substitution of lysine for serine at amino acid position 85, seems to endow partial fluoroquinolone resistance on biotype 2, serovar E strains. This kind of double mutation is present in diseased-fish isolates of the zoonotic serovar but not in resistant biotype 1 isolates, which show different mutations in gyrB or in parC that increase their resistance levels but do not make the strains resistant to fluoroquinolones. Thus, antibiotics other than quinolones should be used at fish farms to prevent the emergence and spread of quinolone resistances, especially to CIP, a drug widely recommended for human vibriosis treatment.  相似文献   

20.
New Design Strategy for Development of Specific Primer Sets for PCR-Based Detection of Chlorophyceae and Bacillariophyceae in Environmental Samples     
Claire Valiente Moro  Olivier Crouzet  Séréna Rasconi  Antoine Thouvenot  Gérard Coffe  Isabelle Batisson  Jacques Bohatier 《Applied and environmental microbiology》2009,75(17):5729-5733
  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