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 93 毫秒
1.
2.
The presence of Vibrio parahaemolyticus in 123 oyster samples collected from an estuary on the southern coast of Sao Paulo state, Brazil, was investigated. Of the 123 samples, 99.2% were positive with densities ranging from <3 to 105 most probable number (MPN)/g. Densities correlated significantly with water temperature (r = 0.48; P < 0.001) but not with salinity (r = −0.09; P = 0.34). The effect of harvest site on counts was not significant (P > 0.05). These data provide information for the assessment of exposure of V. parahaemolyticus in oysters at harvest.Infections caused by Vibrio parahaemolyticus have been reported in several countries (1, 3-5, 14, 15, 16, 17, 18, 19, 20, 22, 24, 26). Among other pathogenic features, V. parahaemolyticus strains produce a thermostable hemolysin, known as thermostable direct hemolysin (TDH), as well as TRH (a TDH-related hemolysin) (25, 29). However, not all strains are pathogenic, as less than 1% of food or environmental strains produce TDH or TRH (2, 7, 9, 10, 11).The most important vehicle for this microorganism is raw or partially cooked shellfish (8, 13, 25, 29). In this study, the densities of V. parahaemolyticus in oysters collected in six oyster bed sites in the estuary of Cananeia (25°S; 48°W) in the southern coastal area of Sao Paulo state, Brazil (Fig. (Fig.1)1) between May 2004 and June 2005 were determined using the most probable number (MPN) technique by the method of De Paola and Kaysner (12). Each sample consisted of 15 oysters, pooled in a plastic bag, and transported in a cold box to the laboratory located in the city of Sao Paulo, Brazil. The temperature during transportation did not exceed 13°C, and the travel time was around 5 h. In the laboratory, the oysters were kept under refrigeration (4 to 8°C) and analyzed within 24 h of collection. Oysters were cleaned and shucked by the method of Cook et al. (6). Identification of V. parahaemolyticus was based on traditional and API 20E strip biochemical tests (bioMérieux, France), using V. parahaemolyticus ATCC 17802 as the reference strain. The observed prevalence of this bacterium was high, as the microorganism was detected in 99.2% (122/123) of the samples and the densities varied between 0.78 and 5.04 log MPN/g.Open in a separate windowFIG. 1.Locations of oyster bed sites in the Cananeia estuary on the southern coast of Sao Paulo state, Brazil. (Courtesy of E. E. de Miranda and A. C. Coutinho [Embrapa Monitoramento por Satélite] [http://www.cdbrasil.cnpm.embrapa.br].)Strategies for the control of V. parahaemolyticus in oysters depend on understanding the seasonal and geographical distribution and the effects of environmental parameters on the growth of this pathogen. To verify the influence of salinity and temperature of seawater on the density of V. parahaemolyticus, samples of water (n = 123) were collected from the same depth of oyster beds using 250-ml plastic flasks. Salinity was determined using a salinometer (model RS10; Rosemount Analytical, Cedar Grove, NJ), and the temperature was determined at the time of collection using a digital thermometer (Hanna Instruments). The results are shown in Fig. Fig.22.Open in a separate windowFIG. 2.Total densities of Vibrio parahaemolyticus in oysters from the southern coast of Sao Paulo state, Brazil. Each bar or point represents the arithmetic mean of six sites, and each error bar represents the standard deviation.Total V. parahaemolyticus densities did not correlate significantly with water salinity, as determined by Pearson coefficient (r = −0.09; P = 0.34). However, the mean salinity varied significantly according to the sampling site and season (P < 0.05) (Table (Table1).1). The highest mean salinity (24.2 ppt) was detected at site 5 and was 1.4 times higher than at site 2 (17.3 ppt), the lowest mean salinity detected in this study.

TABLE 1.

Seasonal distribution of the total density of Vibrio parahaemolyticus in oysters, water temperature, and salinity in the southern coastal area of Sao Paulo state, Brazil
VariableSeasonNo. of samplesMeanaSDRange
Vibrio parahaemolyticusWinter352.44 A1.06<0.48-4.38
    density (log10Spring293.26 B1.171.04-5.04
    MPN/g)Summer243.47 B0.751.54-5.04
Fall353.48 B1.011.04-5.04
Temp (°C)Winter3520.1 A1.914.4-24.0
Spring2923.6 B1.820.0-26.0
Summer2426.7 C1.424.1-29.2
Fall3523.9 B2.220.6-28.3
Salinity (ppt)Winter3522.3 A4.512.2-29.8
Spring2920.2 AB4.411.2-29.4
Summer2418.2 B4.35.3-25.2
Fall3521.8 A3.98.7-28.2
Open in a separate windowaValues with different letters are significantly different (P < 0.05).The weak correlation between water salinity and V. parahaemolyticus densities in oysters suggests that salinity per se is a secondary factor for growth of this bacterium, as are turbidity and chlorophyll content in water (27, 30). These results agree with those obtained by Deepanjali et al. (9) and Martinez-Urtiga et al. (21), who did not find correlation between these two parameters. However, they are in contrast with the results reported by DePaola et al. (11), who observed correlation (P < 0.05) between salinity and total density of V. parahaemolyticus.The results of this study corroborate existing evidence (10, 11, 27, 30) indicating that the temperature of seawater has a significant correlation (r = 0.48; P < 0.001) on the densities of V. parahaemolyticus in oysters, but they are at odds with results reported by Deepanjali et al. (9), who observed no statistically significant correlation with seawater temperature. The temperature variations observed in the present study (15°C) were lower than those observed by DePaola et al. (22°C) (11) but higher than those reported by Deepanjali et al. (10°C) (9).The relationship between V. parahaemolyticus density and water temperature and salinity were analyzed by multiple linear regression. Results showed that salinity was not significant either for linear effects or for squared effects (P > 0.05). For temperature, while the parameter of linear effect was significant (P < 0.05), the squared effect was not (P > 0.05). Considering the goodness of fit of the model, the following linear regression described the density in oysters the best (Fig. (Fig.3):3): log10 MPN V. parahaemolyticus/g = −0.944 + (0.175 × temperature). The lack of model fitness test was not significant and was considered adequate to express the relationship between V. parahaemolyticus density and seawater temperature, in spite of the low R2 (0.23).Open in a separate windowFIG. 3.Goodness of fit regression model of V. parahaemolyticus density in oysters and water temperature.The effect of temperature was further summarized by rank correlation and the use of a smoothing technique (moving average) in which densities corresponding to temperatures within a range of 1°C were pooled to estimate an arithmetic mean of densities in successive intervals. The moving average was calculated using a length of three values. Although seawater temperature and V. parahaemolyticus densities were correlated in the present study, the mean densities reached a plateau at temperatures above 24°C and below 20°C (Fig. (Fig.4)4) where the density was not significantly influenced by temperature, consistent with observations reported also by DePaola et al. (11). Our findings could explain the lack of correlation among those parameters detected in tropical oysters by Deepanjali et al. (9) when the temperature varied from 25 to 35°C.Open in a separate windowFIG. 4.Relationship between the mean density of V. parahaemolyticus in oysters and seawater temperature in Cananeia estuary, Sao Paulo state, Brazil.The influence of the season of the year and site of collection on the mean densities was assessed by analysis of variance and Tukey''s test, when necessary. As shown in Table Table1,1, the V. parahaemolyticus densities were similar in the samples collected during spring, summer, and autumn but differed significantly (P < 0.05) in those collected during winter. Densities among samples collected during summer varied less compared to other seasons. Densities above 105 MPN/g were detected in six (4.9%) oyster samples (three samples during spring, two samples during summer, and one sample during autumn) collected when the temperature was higher than 24°C and the salinity was higher than 15 ppt. The effect of harvest site on densities was not significant (P > 0.05).Previous studies performed with oysters collected in the same region in Brazil have shown a low incidence of pathogenic V. parahaemolyticus (23, 28). Similar results were observed in the present study, as only one oyster sample (0.8%) and only one isolate among 2,243 isolates tested (0.044%) were Kanagawa and tdh positive. Besides the Kanagawa reaction (24), all strains have been tested for tlh, tdh, and trh genes using PCR (12). The pathogen-positive sample presented with a low density of V. parahaemolyticus (3 MPN/g) and was collected during winter, when the temperature was 21°C. Due to the low incidence of pathogenic strains in the samples, correlation between pathogenicity and water temperature or salinity could not be determined.This study indicates that the presence of V. parahaemolyticus in oysters cultivated in the southern coast of Sao Paulo state, Brazil, is high, but pathogenic strains are seldom detected. These results on the ecology and characteristics of V. parahaemolyticus are valuable for future risk assessments related to this pathogen in oysters at harvest.  相似文献   

3.
4.
A total of 905 enterohemorrhagic Escherichia coli (EHEC) O157:H7 isolates that were recovered from experimentally infected cattle, in addition to the inoculated strain, were analyzed by pulsed-field gel electrophoresis (PFGE). Twelve PFGE profiles other than that of the inoculated strain were observed. We successfully identified five distinct chromosomal deletions that affected the PFGE profiles using whole-genome PCR scanning and DNA sequencing analysis. The changes in PFGE profiles of EHEC O157:H7 isolates after passage through the intestinal tract of cattle were partially generated by deletion of chromosomal regions.Enterohemorrhagic Escherichia coli (EHEC) O157:H7 causes hemorrhagic colitis and hemolytic-uremic syndrome in humans worldwide (18). Cattle are considered the primary reservoir for this pathogen and play a central role in transmission to humans (6). Healthy cattle transiently carry EHEC O157:H7 and shed the bacteria in their feces (5, 7). Human infections have been associated with the consumption of contaminated meat and milk, direct contact with cattle, and the consumption of vegetables, fruits, and water contaminated with cattle manure (6).Because of its high discriminatory power, pulsed-field gel electrophoresis (PFGE) has been widely employed as a molecular typing method in many epidemiological investigations to identify various outbreaks and routes of transmission of EHEC O157:H7 (1, 12, 15, 17). Simpson''s index of diversity (9) was reported to be >0.985 in previous studies (1, 15), supporting the identification of richness (the number of types among isolates) and evenness (the relative distribution of individual strains among the different types) of molecular typing using PFGE.Instability of the PFGE patterns of EHEC O157:H7 isolates has been reported. Changes in PFGE patterns were observed among strains after repeated subculturing and prolonged storage at room temperature (11). Loss of Shiga toxin genes and a large-scale inversion within the genome were identified as genetic events generating changes in PFGE patterns in vitro (10, 13). Shifts in the genotypes of EHEC O157:H7 clinical isolates from patients and cattle have been reported (3, 14). This phenomenon was also observed in EHEC O157:H7 experimental infections of cattle. Spontaneous curing of a 90-kb plasmid resulted in the loss of two restricted fragments from the PFGE profiles of EHEC O157:H7 isolates obtained from experimentally infected cattle (2). The purpose of the present study was to identify the genetic events affecting the PFGE patterns of EHEC O157:H7 after passage through the intestinal tract of cattle, especially for restriction fragments that are >90 kb long.Four 5-month-old Holstein steers were housed individually in climate-controlled biosafety level 2 containment barns in accordance with the guidelines for animal experimentation defined by the National Institute of Animal Health of Japan. The pens had individual floor drains and were cleaned twice daily with water and disinfectant. All animals were healthy and culture negative for EHEC O157:H7 strains, as determined by a previously described technique (2), prior to inoculation.EHEC O157:H7 strain Sakai-215 (12, 23), which was isolated from an outbreak in Sakai, Osaka Prefecture, in 1996 was used for inoculation. This strain harbors the genes encoding Stx1 and Stx2. A spontaneous resistant strain was selected with nalidixic acid in order to facilitate the recovery of this strain from fecal samples. All calves were inoculated using a stomach tube with an exponential-phase culture (109 CFU) of the nalidixic acid-resistant Sakai-215 strain. Fecal samples were collected from the four calves daily for 45 days. Fecal culturing was performed as described previously (2). Eight non-sorbitol-fermenting colonies were selected daily from each animal and identified as EHEC O157:H7 colonies by routine diagnostic methods (25).All animals were clinically normal throughout the experimental period. The EHEC O157:H7-inoculated calves (calves 1 to 4) were culture positive for the organism 24 h after inoculation. Intermittent fecal shedding by the calves was observed until 27, 32, 26, and 39 days postinoculation for calves 1, 2, 3, and 4, respectively (Fig. (Fig.1).1). The numbers of EHEC O157:H7 isolates recovered from calves 1, 2, 3, and 4 were 200, 224, 200, and 281, respectively.Open in a separate windowFIG. 1.Changes in PFGE profiles of EHEC O157:H7 isolates recovered from calves 1 (A), 2 (B), 3 (C), and 4 (D). The absence of a bar indicates that no EHEC O157:H7 was detected. The open horizontal bars under the vertical bars indicate that the eight isolates obtained on a day were obtained from the enrichment culture.A total of 905 recovered isolates in addition to the inoculated strain were used for PFGE analysis. Genomic DNA from each EHEC O157:H7 isolate was prepared using the method of Persing (Mayo Clinic, Rochester, MN) described by Rice et al. (20). Agarose-embedded chromosomal DNA was cleaved with XbaI by following the manufacturer''s instructions. PFGE was performed in a 0.85% megabase agarose gel, using a CHEF DR III apparatus (Bio-Rad Laboratories). The pulse time was increased from 12 to 35 s for 18 h. The PFGE profiles of all of the EHEC O157:H7 isolates recovered from the four calves were compared with that of the inoculated strain. The number of band differences was determined by enumerating the loss and addition of fragments (22).Two hundred eighty-nine isolates had PFGE profiles different from that of the inoculated strain, and 12 distinct PFGE profiles were identified for these isolates (Table (Table1).1). The fact that only one to three band differences were observed for the 12 profiles suggested that these isolates were closely related (22) and were variants of the inoculated strain. In addition, the pens had individual floor drains and were cleaned twice daily with water and disinfectant, which reduced the likelihood of introduction of novel EHEC O157:H7 strains. We designated the PFGE profiles A to L. PFGE profiles A, C, and H were obtained for all four calves and accounted for 30.4% of the 905 isolates recovered. Different PFGE profiles were obtained for all animals at least 2 days postinoculation (Fig. (Fig.1).1). All eight isolates from calf 2 collected on day 15 postinoculation and from calf 3 collected on days 22 and 23 postinoculation had PFGE profiles different from that of the original isolate (Fig. (Fig.1).1). The isolates that had the same PFGE profile as the inoculated strain were detected again later.

TABLE 1.

Temporal distribution of PFGE profiles of EHEC O157:H7 isolates recovered from experimentally infected cattle
PFGE profileNo. of isolates recovered at different times postinoculation from:
Total no. of isolates (%)
Calf 1
Calf 2
Calf 3
Calf 4
1 to 10 days11 to 20 days21 to 27 days1 to 10 days11 to 20 days21 to 30 days31 to 32 days1 to 10 days11 to 20 days21 to 36 days1 to 10 days11 to 20 days21 to 30 days31 to 39 days
Ina63562562465066046559454746616 (68.1)
A1119101372121324410191612181 (20.0)
B11 (0.1)
C132471362776572 (8.0)
D11 (0.1)
E11 (0.1)
F123 (0.3)
G11 (0.1)
H1314133221122 (2.4)
I1113 (0.3)
J11 (0.1)
K112 (0.2)
L11 (0.1)
Total no. of variants (%)b17 (21.3)24 (30.0)15 (37.5)18 (22.5)18 (25.0)22 (30.6)2 (25.0)20 (25.0)34 (42.5)35 (87.5)21 (26.3)27 (37.5)17 (26.6)19 (29.2)289 (31.9)
Open in a separate windowaPFGE profile of inoculated strain Sakai-215.bTotal numbers of isolates having PFGE profiles A to L.Kudva et al. (16) demonstrated that the difference in PFGE profiles between EHEC O157:H7 strains was due to distinct insertions or deletions that contained XbaI sites rather than to single-nucleotide polymorphisms in the XbaI sites themselves. To identify the locations of insertions or deletions in the genome of the EHEC O157:H7 isolates recovered from experimentally infected cattle, whole-genome PCR scanning (WGP scanning) was performed as described previously (19). Briefly, 549 pairs of PCR primers were used to amplify 549 segments covering the whole chromosome of EHEC O157:H7 strain RIMD 0509952, with overlaps of a certain length at every segment end. The inoculated strain (strain Sakai-215) and the strain whose genome was sequenced (RIMD 0509952) were isolated from the same outbreak in Japan in 1996 (23) and had same PFGE profile after XbaI digestion. All primer sequences are available at http://genome.gen-info.osaka-u.ac.jp/bacteria/o157/pcrscan.html. PCR were performed using genomic DNA as the template and long accurate PCR (LA-PCR) kits. The cycling conditions for the LA-PCR included an initial incubation at 96°C for 100 s, followed by 30 cycles of 96°C for 20 s and 69°C for 10 min.Prior to the WGP scanning of the isolates, we scanned an approximately 1.2-Mb region covered by 116 segments (71/72 to 146/147) of the EHEC O157:H7 genome using 24 strains, including inoculated strain Sakai-215, 4 isolates with the same PFGE profile as the inoculated strain, 3 isolates with PFGE profile A, 3 isolates with PFGE profile C, 2 isolates with PFGE profile F, 2 isolates with PFGE profile H, 2 isolates with PFGE profile I, and one isolate each with PFGE profiles B, D, E, G, J, K, and L. The main purpose of this preliminary scanning was to determine the extent of variation in the data for isolates having the same PFGE profiles.As shown in Fig. Fig.2,2, we successfully amplified products that were the expected sizes for 103 of the 116 segments for the 24 strains tested. No amplification in a segment was observed for the 24 strains. Polymorphism (expected amplification was observed in some but not all strains) was observed in 12 segments. Eleven of the 12 polymorphic segments consisted of two different sequentially unamplified regions. An IS629 insertion was also observed in a polymorphic segment in one strain. In other words, variation in the data for isolates with the same PFGE profile was not observed except for the isolates having the same PFGE profile as the inoculated strain. Hence, we performed WGP scanning using one isolate with each of the selected PFGE profiles.Open in a separate windowFIG. 2.Summary of the results of PCR scanning analysis of part of the EHEC O157:H7 genome using 24 strains recovered from experimentally infected cattle. The line at the top indicates data for the inoculated strain. The positions of Sp5 and Sp6 are indicated above the data lines. Segments showing polymorphism (expected amplification was observed in some strains but not in all strains) are indicated below the data lines. In, inoculated strain.The results of WGP scanning of the seven isolates with different PFGE profiles in addition to inoculated strain Sakai-215 are summarized in Fig. Fig.3.3. We successfully amplified products of the expected sizes for 530 of 549 segments for the eight strains tested. No amplification was observed for any of the eight strains for three segments (133.2/133.3, 164.4/164.5, and 164.5/164.6). Polymorphism was observed in 16 segments. Fourteen of the 16 polymorphic segments were located in four different regions.Open in a separate windowFIG. 3.Summary of the results of WGP scanning analysis of the EHEC O157:H7 isolates recovered from experimentally infected cattle and the inoculated strain. The positions of Sp5 and Sp13 are indicated above the data lines. Segments showing polymorphism (expected amplification was observed in some strains but not in all strains) are indicated below the data lines. In, inoculated strain.The 110/110.1-to-110.5/111 region in PFGE profile I, the 122/122.1-to-122.4/123 region in PFGE profile K, and the 199/199.1-to-199.2/200 region in PFGE profiles B, C, and G corresponded to prophages Sp5, Sp6, and Sp13, respectively. The 283/284-to-285/286 region in PFGE profile E and the 448/448.1-to-448.1/448.2 region in PFGE profile B corresponded to nonprophage regions on the chromosome. The sizes of the deletion sites of nonprophage regions 283/284 to 285/286 and 448/448.1 to 448.1/448.2 were 17 kb and 9.5 kb, respectively. We synthesized new primer pairs upstream and downstream of these five regions and performed LA-PCR (data not shown). The results of the sequencing analysis of the products indicated that the three prophage genomes were cured at their integration sites (Fig. 4A to C). It is not clear from this study whether deletion of the three prophages represented phage excisions or simple deletions. We identified short direct CCGCCA and GC repeats at both ends of the 17-kb and 9.5-kb deletion sites, respectively, compared with the sequence data for the Sakai-215 strain, although the deleted regions included one side of the direct repeats (Fig. 5D and E).Open in a separate windowFIG. 4.Schematic diagrams showing the relationships between deletions of chromosomal regions and changes in the sizes of restricted fragments. (A) The 467-kb restricted fragment of PFGE profile I was generated by deletion of prophage Sp5 located in the 530-kb fragments of the inoculated strain. (B) The 759-kb restricted fragment of PFGE profile K was generated by deletion of Sp6 located in the adjacent 530-kb and 278-kb fragments of the inoculated strain. (C) The 291-kb restricted fragment of PFGE profiles C, G, and H was generated by deletion of prophage Sp13 located in the adjacent 255-kb and 55-kb fragments of the inoculated strain. (D) The 188-kb restricted fragment of PFGE profile E was generated by deletion of the 17-kb chromosomal region in the 205-kb fragment of the inoculated strain. (E) The 334-kb restricted fragment of PFGE profile B was generated by deletion of the 9.5-kb region located in the adjacent 343-kb and 6.2-kb fragments of the inoculated strain.Open in a separate windowFIG. 5.Comparison of the PFGE profiles of the EHEC O157:H7 isolates recovered from experimentally infected cattle and the inoculated strain. Lane M, λ ladder used as a size marker; lane 1, inoculated strain; lanes 2 to 13, isolates with PFGE profiles A to L, respectively.The deleted 17-kb region contains 16 open reading frames, including formate hydrogenase-related genes (4), mutS (21), and rpoS (8), suggesting that the strain with PFGE profile E is more susceptible to environmental stresses than the inoculated strain. In fact, the isolate with PFGE profile E was more susceptible to low-pH, high-temperature, and high-osmolarity conditions or to the presence of deoxycholate in vitro than the other isolates obtained in this study (data not shown). The fact that this isolate was obtained 4 days after inoculation from calf 1 and could not be detected after that time suggested that the isolate with PFGE profile E could not survive in the intestine of the calf due to the loss of genes related to stress resistance. The deleted 9.5-kb region contains nine open reading frames whose functions are unknown. The strain with this deletion was isolated 1 day after inoculation from calf 3 and could not be detected after that time.Sp5 is one of the prophages in EHEC O157:H7 RIMD 0509952 carrying the stx2 gene. Deletion of this prophage affected the PFGE profile of inoculated strain Sakai-215. The loss of a 530-kb fragment and the gain of a 467-kb fragment due to deletion of the 63-kb prophage Sp5 were identified in PFGE profile I (Fig. (Fig.4A4A and and55).Sp6 is one of the lambda-like phages and has a single XbaI site in its genome. The loss of 530-kb and 278-kb fragments and the gain of a 759-kb fragment due to deletion of this phage were identified in PFGE profile K (Fig. (Fig.4B4B and and55).Sp13 is one of the P2-like phages that have a single XbaI site in the genome. The loss of 255-kb and 55-kb fragments and the gain of a 291-kb fragment due to deletion of this prophage were identified in PFGE profiles C, G, and H (Fig. (Fig.4C4C and and5).5). The same changes in PFGE profile B were not observed, although we found a sequentially unamplified region in which Sp13 was located in the genomes of isolates with PFGE profile B (Fig. (Fig.3).3). We detected part of the Sp13 sequence by Southern blot analysis; however, this part of the sequence was not detected in isolates with PFGE profiles C, G, and H (data not shown). One possible explanation for this phenomenon is that deletion of part of the Sp13 sequence included deletion of primer annealing sites. However, the details of mutation in this region for the isolates with PFGE profile B are not clear.Deletion of the two nonprophage regions also affected the PFGE profiles. The loss of a 205-kb fragment and the gain of a 188-kb fragment due to deletion of a 17-kb region were identified in PFGE profile E (Fig. (Fig.4D4D and and5).5). The loss of a 343-kb fragment and the gain of a 334-kb fragment due to deletion of a 9.5-kb region were identified in PFGE profile B (Fig. (Fig.4E4E and and55).Two single unamplified segments were both observed in the strain with PFGE profile F (106.3/106.4 and 204.2/204.3). We could not amplify these regions using additional primer pairs (data not shown). Insertion of DNA or large-scale inversion might have occurred in these regions. The other unamplified segments all corresponded to deletion of chromosomal regions. Recombination successfully occurred and cured three prophages and two other chromosomal regions. These data suggest that the changes in PFGE profiles after passage through the intestinal tract of cattle are generated in part by deletion of chromosomal regions. Obviously, deletion of five chromosomal regions does not explain the other changes in the PFGE profiles, including profiles A, D, F, J, and L. The genetic events behind such changes are not clear.Prior to drawing a conclusion, we need to consider the use of nalidixic acid, a potent inducer of bacteriophage induction (24), for selection of the isolates. In addition, most of the EHEC O157:H7 isolates obtained on day 8 postinoculation and later were isolated from enrichment cultures (Fig. (Fig.1).1). The possibility that the culturing process itself affected the deletion events affecting the PFGE profiles cannot be ruled out. Taken together, the results suggest that deletions can cause a single strain to mutate into several variants while it is passing through the gastrointestinal tract of a host, provided that the culture technique used does not contribute to this process. Hence, this study may explain why EHEC O157:H7 isolates with various PFGE profiles can be isolated from a single animal. What causes the deletion mutations and why the PFGE profiles show such patterns after passage through cattle are subjects for future studies.  相似文献   

5.
6.
7.
A fitness cost due to imbalanced replichores has been proposed to provoke chromosome rearrangements in Salmonella enterica serovars. To determine the impact of replichore imbalance on fitness, the relative fitness of isogenic Salmonella strains containing transposon-held duplications of various sizes and at various chromosomal locations was determined. Although duplication of certain genes influenced fitness, a replichore imbalance of up to 16° did not affect fitness.The bacterial chromosome is a dynamic molecule that can undergo various types of rearrangements, including inversions, translocations, and duplications. These rearrangements can alter gene order and change replichore length. Replichores are defined as the halves of the chromosome between the origin of replication and the terminus region in the vicinity of the dif site (4, 6, 10, 15). In most bacteria, both replichores are approximately the same length, with each comprising 180° around the circular chromosome. In this state, the replichores and DNA replication are balanced. Imbalance is introduced when one replichore becomes longer than the other. For example, if the replichores comprise 200° and 160° around the circular chromosome, the replichores are 20° imbalanced. Various studies over the years have investigated the effect of imbalanced DNA replication on fitness in Escherichia coli (8, 9, 16, 17). In a recent analysis, E. coli strains that contained asymmetrical interreplichore inversions were found to have growth defects when the imbalance was at least 50° (8). While these studies have demonstrated that imbalanced replichores can affect fitness, the approaches used to introduce the imbalance either do not occur or are extremely rare in nature.Duplications play important evolutionary roles because the increase in gene copy number can facilitate adaptation to certain growth conditions (20), as well as being a source for the evolution of new genes (3). Typically duplications occur at frequencies between 10−3 and 10−5 in unselected cultures (2) but are lost at rates of up to 1,000-fold more often if they do not provide a selective advantage (19). Duplications also result in replichore imbalance. However, the effect of duplications on fitness in relation to replichore balance has not been investigated.A major hurdle in studying the effects of duplications on fitness is that if the duplication is detrimental, haploid revertants will outgrow the parental strain containing the duplication. To circumvent this problem, we used a set of 11 isogenic Salmonella enterica strains with transposon-held duplications (5) (Fig. (Fig.11 and Table Table1).1). Culturing these strains in the presence of chloramphenicol selects for the duplication because if the duplication collapses, the transposon is lost and the cells become chloramphenicol sensitive. As the size of the duplications in these strains varies, so does the amount of introduced replichore imbalance, ranging from 5° to 23°. In addition, fitness effects due to the location of the duplication versus the size of the duplication were also discerned, as the collection includes duplications of similar sizes located in different regions of the chromosome.Open in a separate windowFIG. 1.Genetic map of S. enterica serovar Typhimurium LT2 showing genes used as endpoints in constructing transposon-held duplications of the regions between genes. Balanced replichores are indicated by the symmetry of the oriC-Ter axis. Duplications increase the length of one replichore relative to the other, imbalancing axis symmetry.

TABLE 1.

Properties of the S. enterica strains used in this study
StrainAliasGenotypeDuplication location (min)Duplication size (kbp)Replichore imbalance (°)Reference
MST1LT2Wild type1.911
MST3813SV4200Dup [trp248 MudP hisD9953]38-44332.111.55
MST3814SV3193Dup [hisH9962 MudP cysA1586]44-53399.713.75
MST3815SV4015Dup [cysA1586 MudP purG2149]53-56158.45.75
MST3816SV4193Dup [purG2149 MudP argA9001]56-64430.714.75
MST3817SV4194Dup [argA9000 MudP cysG1573]64-75484.616.35
MST3818SV1601Dup [cysG1573 MudP ilvA2642]75-85486.416.45
MST3819SV4195Dup [ilvA2648 MudP purA1881]85-95495.316.75
MST3820SV4142Dup [purA1881 MudP thr469]95-0248.78.85
MST3821SV1604Dup [thr469 MudP proA692]0-8367.112.65
MST3822SV1603Dup [proA692 MudQ purE2164]8-12230.28.15
MST3823SV1611Dup [purE2514 MudP purB1879]12-27723.223.35
MST1529TT11183srl-203::Tn10d(Cam)1
TYT4480rrfH::pCE36This work
MST5198rrfH::pCE36 srl-203::Tn10d(Cam)This work
Open in a separate windowThe relative fitness of these strains was investigated by measuring growth rates and by assaying growth in a mixed culture with an isogenic competitor strain. Single colony isolates streaked from frozen stocks were used to inoculate broth cultures. Growth was measured in E medium supplemented with 0.2% glucose (minimal medium), Luria-Bertani medium (LB), or LB supplemented with 1× E-salts and 0.2% glucose (LBEDO) (18). Competition assays were done in LB. Media were supplemented with chloramphenicol (20 mg/ml) to maintain the duplications, and the solid media used in competition experiments also contained 5-bromo-4-chloro-3-indolyl-β-d-galactopyranoside (X-Gal; 40 mg/ml) to differentiate between the duplication and competitor strains.In minimal medium, most of the strains had a generation time of approximately 40 min, except for MST3813, MST3823, and MST3819, which grew somewhat slower (Fig. (Fig.2).2). In LB, the average generation times of all of the duplication-bearing strains were longer than that of the wild type. When grown in LBEDO, all of the strains had generation times of 11 to 15 min. Observed differences in growth rate were not due to lower viability as determined by plate counts. There was no correlation between duplication size and generation time in any of the three media.Open in a separate windowFIG. 2.Generation times of strains as a function of duplication size. Strains were grown in minimal medium (♦), LB (▪), or LBEDO (▴) aerobically at 37°C. Error bars show standard deviations.Competition indices were determined from mixed cultures with the isogenic competitor strain MST5198 and either the wild-type strain or one of the duplication-bearing strains. The competitor strain contained an integrated plasmid encoding a promoterless lacZ gene (7) driven by the rrnH promoter and Tn10d(Cam) conferring chloramphenicol resistance integrated into the srl locus (Table (Table1).1). Strains were grown to saturation overnight, and 10−4 dilutions were used to inoculate mixed cultures. Samples taken at t = 0 (input) and subsequent time points (output) were diluted and spot plated in triplicate. The competition index (C.I.) was calculated as follows: C.I. = (no. output Lac CFU/no. output Lac+ CFU)/(no. input Lac CFU/no. input Lac+ CFU).Data are shown for 6 and 12 h, but readings were taken at time points of up to 1 week. There was no correlation between the duplication size and the C.I. (Fig. (Fig.3).3). Most of the duplication-bearing strains competed as well as or slightly better than the wild type against the competitor strain. Two exceptions were MST3819 and MST3823, which both competed significantly more poorly against the competitor strain than did the wild-type strain (Student''s t test P < 0.05). While these two strains contain the largest duplications, the size of the duplication in MST3819 was close to the size of the duplications in MST3817 and MST3818. Strains MST3817 and MST3818 were either similar or slightly better competitors than the wild type, indicating that the fitness defect in MST3819 is not due to replichore imbalance.Open in a separate windowFIG. 3.Competition indices of the wild-type and duplication-bearing strains after 6 (▪) and 12 h (░⃞) of growth in mixed cultures with MST5198 (n = 6). Strains are in order of increasing duplication size, with the locations of duplications indicated at the top. Indices statistically significantly different (P < 0.05) from that of the wild-type strain (MST1) are marked (*). Error bars show standard deviations.To confirm that the duplications in MST3819 and MST3823 were responsible for the observed growth defects, isolates of each strain with collapsed duplications were obtained by growing cultures of each strain without chloramphenicol and then screening for chloramphenicol-sensitive derivatives. The loss of the duplication restored wild-type growth to both strains. The frequencies of duplication collapse in cultures of MST3818 and MST3819 were also compared. Since these two strains have duplications of similar sizes, the collapse frequencies of the duplicated regions, and concomitant loss of chloramphenicol resistance, were expected to be similar. While the reversion frequency of MST3818 cells was 3.5 × 10−3 cells/generation, MST3819 cells reverted with a frequency of 2.8 × 10−2 cells/generation, resulting in 97 to 98% of MST3819 cells being chloramphenicol sensitive and haploid after reaching stationary phase.In conclusion, the results of this study show that a replichore imbalance of up to 16° introduced by transposon-held duplications in the chromosome does not have a measurable effect on the fitness of S. enterica. The duplicated regions in these strains are larger than the islands acquired via horizontal gene transfer, which have been purported to disrupt replichore balance sufficiently to promote chromosome rearrangements. A duplication that introduced a replichore imbalance of 23° did have a growth defect, but whether the defect is from replichore imbalance or the gene content of the duplication could not be distinguished. This study suggests that the disruption of replichore balance by acquisition of horizontally transferred genes is not the cause of chromosome rearrangements in host-specific S. enterica serovars as hypothesized (12-14).  相似文献   

8.
Animal-to-Animal Variation in Fecal Microbial Diversity among Beef Cattle   总被引:1,自引:0,他引:1  
The intestinal microbiota of beef cattle are important for animal health, food safety, and methane emissions. This full-length sequencing survey of 11,171 16S rRNA genes reveals animal-to-animal variation in communities that cannot be attributed to breed, gender, diet, age, or weather. Beef communities differ from those of dairy. Core bovine taxa are identified.The gastrointestinal tracts (GIT) of beef cattle are colonized by microorganisms that profoundly impact animal physiology, nutrition, health, and productivity (5). The GIT microbiota potentially impact food safety via pathogen shedding (13) by interacting with organisms such as Salmonella and competing for resources in the GIT. Cattle intestinal microbiota also play an important role in methane emissions, with U.S. beef cattle alone contributing an estimated 3.87 million metric tons of methane into the environment each year, both from rumen and large-intestine fermentations (7). Although the bovine fecal microbiota have been well characterized using culture-based methods, these techniques are necessarily limited to characterizing bacteria that can be grown in the laboratory. Culture-independent methods can reveal community members that are recalcitrant to culture. Only a handful of deep-sequencing studies have been done using culture-independent 16S rRNA-based methods (1, 11, 12, 14), all with dairy cattle, which have a fundamentally different diet and metabolism from beef cattle. Despite the potential contributions of the beef cattle GIT microbiota to animal health, food safety, and global warming, these communities remain poorly characterized. With the advent of pyrosequencing technology, researchers now have the tools to characterize these important communities. Pyrosequencing will allow rapid characterization of large-sample data sets (1). However, the taxonomic information generated by rapid sequencing is approximate by necessity (9), and full-length 16S-rRNA sequencing remains the “gold standard” method. Accordingly, we have characterized fecal bacteria from six feedlot cattle by full-length capillary sequence analysis of 11,171 16S rRNA gene clones (Fig. (Fig.11).Open in a separate windowFIG. 1.Bacterial diversity of six feedlot beef cattle. Gray bars represent the percentages of all 16S sequences that were assigned to each taxonomy. Colored dots represent the percentages of 16S sequences from each library that were assigned to each taxonomic group. Asterisks indicate unclassified members of the named taxon. Panel A shows the data for the first 99% of all the sequences. Panel B shows the data for the remaining 1% of sequences. Note differences in scales for panels A and B.Rectal grab fecal samples (n = 6) were collected according to institutional animal care guidelines. All animals were female cross-bred MARCIII beef heifers, 6 to 8 months of age, 214 to 241 kg, housed in the same feedlot pen for 2 months prior to fecal collection, and fed the same typical feedlot beef production growing rations consisting of 61.6% corn silage (41.3% dry matter), 15.2% alfalfa hay, 20.9% corn, and 2.3% liquid supplement.Total fecal DNA was isolated from homogenized samples using MoBio UltraClean fecal kit (Carlsbad, CA). PCR was performed using 27F and 1392R primers (11). Amplification consisted of 25 cycles, with an annealing temperature of 55°C. Amplicons from three reactions per sample were pooled (8), cloned using the Invitrogen TOPO TA cloning kit (Carlsbad, CA), and sequenced bidirectionally with M13 primers using an ABI 3700 sequencer (17). Low-quality and chimeric sequences (6) were excluded from further analysis. Distance matrices were compiled from ClustalW alignments (18) in PHYLIP (4). Pairwise estimates of shared richness were calculated using EstimateS, version 8.2 (R. K. Colwell; http://purl.oclc.org/estimates). DOTUR (16) was used to identify operational taxonomic units (OTUs) and to generate rarefaction curves (Fig. (Fig.2),2), richness and evenness estimates, and Shannon''s and Simpson''s diversity indices (Table (Table1).1). A 97% similarity cutoff and an 85% similarity cutoff for estimating OTUs were used to approximate species and class-level designations (15). Taxonomies were assigned to one member of each OTU using the RDP “classifier” tool (19), and the RDP taxonomic information was used for Fig. Fig.11 and and3.3. Common bovine taxa were identified based on inclusion in all three U.S. culture-independent studies (this study and references 1 and 11).Open in a separate windowFIG. 2.Rarefaction curves for six feedlot beef cattle. OTUs were assigned at the 85% DNA sequence similarity level. For comparison purposes, all six curves were truncated after 1,321 sequences.Open in a separate windowFIG. 3.Phylum-level distribution of bacterial sequences from six beef feedlot cattle. Asterisks indicate unclassified members of the named taxon.

TABLE 1.

Richness and diversity indices for 6 beef feedlot cattle
Library and animal (n)No. of OTUs observedSpecies richness (CI)a by:
Diversity (CI) by:
ChaoACEShannon''s indexSimpson''s index
97% DNA sequence similarity
    Animal 1 (2,485)198372 (294-515)329 (280-408)3.89 (3.83-3.95)0.0422
    Animal 2 (2,084)416600 (538-694)604 (552-675)5.40 (5.35-5.45)0.0066
    Animal 3 (1,710)6961,393 (1,224-1,615)1,418 (1,327-1,523)6.13 (6.08-6.18)0.0027
    Animal 4 (1,512)294526 (439-665)483 (425-566)4.71 (4.63-4.78)0.0237
    Animal 5 (2,059)314612 (495-805)488 (434-566)4.93 (4.88-4.99)0.0126
    Animal 6 (1,321)174320 (252-447)289 (244-361)4.18 (4.11-4.25)0.0286
85% DNA sequence similarity
    Animal 1 (2,485)4861 (51-99)62 (52-90)2.64 (2.59-2.68)0.1056
    Animal 2 (2,084)77107 (87-165)102 (87-139)3.38 (3.34-3.43)0.0505
    Animal 3 (1,710)130153 (139-186)151 (140-174)4.07 (4.02-4.12)0.0254
    Animal 4 (1,512)6675 (68-98)77 (70-96)2.71 (2.64-2.78)0.0931
    Animal 5 (2,059)6980 (72-109)84 (75-110)3.31 (3.26-3.36)0.0545
    Animal 6 (1,321)5465 (57-102)61 (56-76)2.90 (2.83-2.97)0.0939
Open in a separate windowaCI, confidence interval.The GIT community of beef feedlot cattle characterized in this study was found to share many taxa with the bovine GIT community described for dairy cattle (1, 11, 14), although the relative abundances of the major bacterial groups differed considerably. The fecal microbiota of beef cattle were dominated by members of the Firmicutes, with 62.8% of the OTUs belonging to this taxonomic group (Fig. (Fig.3).3). Bacteroidetes (29.5% of the OTUs) and Proteobacteria (4.4% of the OTUs) were also represented in feces (Fig. (Fig.3).3). A total of seven phyla were found in our six animals.Total estimated species richness values (Chao) for each of the six animals were 372, 600, 1,393, 526, 612, and 320 (Table (Table1).1). These cattle richness numbers are higher than those observed for three human subjects (164, 332, and 297) (2). The mean of Chao pairwise estimates of shared richness between any two of the six cattle fecal libraries was 230.Our findings, in addition to those from pyrosequencing studies (1), identify a core set of bovine GIT bacterial taxa, including the Bacteroidetes Prevotella and Bacteroides; the Firmicutes Faecalibacterium, Ruminococcus, Roseburia, and Clostridium; and the proteobacterium Succinovibrio (Fig. (Fig.1).1). These genera are consistently identified in bovine feces and likely compose part of the bovine resident microbiota. Although the potential exists for culture-independent methods to reveal minority microbial community members, 16S rRNA gene sequencing in dairy (1, 11) and beef cattle supports the list of core taxa identified using culture-based methods.Comparisons between our data set and recent studies done with dairy cattle (1, 11, 12) suggest that although beef and dairy cattle share many of the same major bacterial groups, the relative abundances of these groups in beef and dairy cattle may differ, and there may be differences between the two groups in the compositions of minority community members. The most common genus in beef cattle from our study was Prevotella, representing 24% of the total number of sequences evaluated. In comparison, Dowd et al. (1) found that Prevotella spp. represented only 5.5% of the total 16S genes sequenced from 20 dairy cattle, and Prevotella was not listed in the top 10 most frequently occurring OTUs in either of the studies from McGarvey et al. (11, 12). Likewise, Clostridium represented only 1.5% of the total beef sequences but 19% of the dairy pyrosequences (1). There were a number of bacterial sequences present in the beef cattle sequences but not reported in the dairy sequences, including Arthrobacter, Asteroleplasma, Bifidobacterium, Collinsella, Delftia, Eggerthella, Lactobacillus, Mitsuokella, Olsenella, and Propionibacterium (1, 11), although a number of these genera have been cultured from dairy animals in the past. It must be noted that all of these sequencing studies examined only a small number of animals, and each method has limitations which affect interpretation of the results. The full-length sequencing performed as part of this beef cattle study and two dairy studies (11, 12) relies on a PCR step which can potentially affect the relative numbers of each taxon observed due to PCR bias, while the pyrosequeincg method used in the 20-animal dairy study suffers from artifacts that potentially affect taxonomic assignment and richness estimates due to short read lengths and potential biases in evenness (how many of each group) due to primer and template mismatches (3). Nonetheless, these studies indicate that there may be fundamental differences between the gastrointestinal communities of beef and dairy cattle, they provide a comprehensive examination of the communities present in the specific animals tested, and they serve to provide important baseline information for further studies examining various factors which can impact cattle gastrointestinal communities.The taxonomic information generated by deep sequencing of beef cattle feces revealed considerable animal-to-animal variation in the operational taxonomic unit (OTU) composition of the individual libraries (Fig. (Fig.1).1). The OTU designation facilitates an analysis of the community data without forcing the assignment of sequences into an incomplete and imperfect bacterial taxonomic system. It relies on DNA sequence similarity to assign sequences to a particular OTU defined by the level of DNA sequence similarity. In total, 1,906 OTUs (97% OTU designation) were identified in the six libraries. Of these, only 24 OTUs (1.2%) (comprising 1,253 [11.2%] of sequences) were present in all six libraries, while 1,348 OTUs (69%) were found only in individual libraries. Of these, 1,064 OTUs (77%) were unique, represented by a solitary clone (range of 3% to 29% of the total clones from each individual animal). These data hint at considerable animal-to-animal variation in bacterial community structure at the species level that cannot be readily attributed to breed, gender, age, macroecologic factors such as weather conditions, or diet, given that the animals in this study were controlled for these variables, and support the conclusions of Manter et al. (10) that pooling samples can obscure rare phylotypes.Our results from beef cattle suggest that there may be differences in the bacterial community members present in the GIT of each individual animal that cannot be attributed to diet, breed, gender, age, or macroecologic factors such as weather and suggest the need for the high-resolution community sequencing of much larger numbers of animals before “core” minority community members can be identified. Considering the limited nature of the community surveys to date and all of the genetic, management, geographic, and temporal factors that can contribute to the composition of GIT microbiota, much work remains before we are able to understand and predict the community composition of any individual animal.  相似文献   

9.
Halogenated organic compounds serve as terminal electron acceptors for anaerobic respiration in a diverse range of microorganisms. Here, we report on the widespread distribution and diversity of reductive dehalogenase homologous (rdhA) genes in marine subsurface sediments. A total of 32 putative rdhA phylotypes were detected in sediments from the southeast Pacific off Peru, the eastern equatorial Pacific, the Juan de Fuca Ridge flank off Oregon, and the northwest Pacific off Japan, collected at a maximum depth of 358 m below the seafloor. In addition, significant dehalogenation activity involving 2,4,6-tribromophenol and trichloroethene was observed in sediment slurry from the Nankai Trough Forearc Basin. These results suggest that dehalorespiration is an important energy-yielding pathway in the subseafloor microbial ecosystem.Scientific ocean drilling explorations have revealed that marine subsurface sediments harbor remarkable numbers of microbial cells that account for approximately 1/10 to 1/3 of all living biota on Earth (20, 25, 33). Thermodynamic calculations of pore-water chemistry suggest that subseafloor microbial activities are generally supported by nutrient and energy supplies from the seawater and/or underlying basaltic aquifers (6, 7). Although sulfate, nitrate, Fe(III), Mn(IV), and bicarbonate are known to be potential electron acceptors for anaerobic microbial respiration in marine subsurface sediments (5), the incidence of both the dissimilatory dehalorespiration pathway and microbial activity in halogenated organic substrates remains largely unknown.Previous molecular ecological studies using 16S rRNA gene sequences demonstrated that Chloroflexi is one of the most frequently detected phyla in subseafloor sediments of the Pacific Ocean margins (12-14). Some of the sequences within the Chloroflexi are closely related to sequences in the genus Dehalococcoides, which contains obligatory dehalorespiring bacteria that employ halogenated organic compounds as terminal electron acceptors (21, 29). The frequent detection of Dehalococcoides-related 16S rRNA genes from these environments implies the occurrence of dissimilatory dehalorespiration in marine subsurface sediments.In this study, we detected and phylogenetically analyzed the reductive dehalogenase homologous (rdhA) genes, key functional genes for dehalorespiration pathways, from frozen sediment core samples obtained by Ocean Drilling Program (ODP) Leg 201 (Peru margin and eastern equatorial Pacific) (7, 14); Integrated Ocean Drilling Program (IODP) Expedition 301 (Juan de Fuca Ridge flank) (8, 24); Chikyu Shakedown Expedition CK06-06 (Northwest Pacific off Japan) (20, 23); and IODP Expedition 315 (Nankai Trough Forearc Basin off Japan) (Table (Table1).1). DNA was extracted using an ISOIL bead-beating kit (Nippon Gene, Japan) and purified using a MagExtractor DNA fragment purification kit (Toyobo, Japan) according to the manufacturer''s instructions. To increase concentration, DNA was amplified by multiple displacement amplification using the phi29 polymerase supplied with a GenomiPhi kit (GE Healthcare, United Kingdom) (20). Putative rdhA genes were amplified by PCR using Ex Taq polymerase (TaKaRa, Japan) with degenerate primers RRF2 and B1R (17), dehaloF3, dehaloF4, dehaloF5, dehaloR2, dehaloR3, and dehaloR4 (32), and ceRD2S, ceRD2L, and RD7 (26) and the PCR conditions described in those studies. Amplicons of the approximate target size were gel purified and cloned into the pCR2.1 vector (Invitrogen, Japan). Sequence similarity was analyzed using FastGroupII web-based software (34), and sequences with a 95% identity were tentatively assigned to the same phylotype. Amino acid sequences were aligned by ClustalW (31), including known and putative reductive dehalogenase sequences in the genome of Dehalococcoides ethenogenes strain 195 (28), as well as several functionally characterized reductive dehalogenases from other species.

TABLE 1.

Sample locations and results of PCR amplification of rdhA
Sampling site (expedition name)LocationWater depth (m)Core sectionSediment depth (mbsf)rdh amplification resulta
1226 (ODP Leg 201)Eastern equatorial Pacific3,2971-33.2++
6-346.7++
1227 (ODP Leg 201)Southeast Pacific off Peru4271-10.3+
3-216.6+
5D-542.0
9-375.1+
1230 (ODP Leg 201)Southeast Pacific off Peru5,0861-10.3++
10-373.8
27-3209.3
1301 (IODP Expedition 301)Northeast Pacific Juan de Fuca Ridge flank off Oregon2,6561-22.5+
6-651.2
11-190.8
1D-2132.5
C9001 (JAMSTEC Chikyu Shakedown Expedition CK06-06)Northwest Pacific off Japan1,1801-11.0++
2-513.5++
9-478.5+
21-4191.5+
24-4216.8++
25-6228.9
38-7346.3
40-3358.6+
C0002 (IODP Expedition 315)Nankai Trough Forearc Basin off Japan1,9371-31.9+
1-64.7
2-49.2+
2-813.4
3-520.2+
4-530.0
8-366.6+
16-4155.4
Open in a separate windowa−, PCR product of expected size not amplified; +, PCR product of expected size weakly amplified; ++, PCR product of expected size amplified and confirmed by sequencing analysis.Putative rdhA genes were successfully detected by primer set RRF2-B1R in samples from the eastern equatorial Pacific (ODP site 1226, 3.2 and 46.7 m below the seafloor [mbsf]), the Peru margin (ODP site 1227, 0.3, 16.6, and 75.1 mbsf, and ODP site 1230, 0.3 mbsf), the Juan de Fuca Ridge flank (IODP site 1301, 2.5 mbsf), offshore from the Shimokita Peninsula of Japan (CK06-06 site C9001, 1.0, 13.5, 78.5, 191.5, 216.8, and 358.6 mbsf), and the Nankai Trough Forearc Basin off the Kii Peninsula of Japan (IODP site C0002, 1.9, 9.2, 20.2, and 66.6 mbsf) (Table (Table1).1). No amplification was observed in samples from several deep horizons at sites 1227, 1230, 1301, C9001, and C0002 (Table (Table1).1). A total of 92 clones of subseafloor putative rdhA genes were sequenced and classified into 32 phylotypes (Fig. (Fig.1).1). Phylogenetic analysis revealed that all of the detected putative rdhA sequences were related to those of Dehalococcoides.Open in a separate windowFIG. 1.Phylogenetic tree based on the deduced amino acid sequences of rdhA genes, including sequences from marine subsurface sediments. Putative rdhA sequences from marine subsurface sediments (rdhA clones 1 to 32) are marked in red, while those of the Dehalococcoides genome are marked in blue. Clonal frequencies and sequence accession numbers are indicated in parentheses. Bootstrap values from 50% to 84% and 85% to 100% are indicated by open and solid circles at the branches, respectively. Asterisks indicate the following functionally characterized rdhA genes: pceA and prdA, tetrachloroethene reductive dehalogenase; tceA, trichloroethene reductive dehalogenase; vcrA and bvcA, vinyl chloride reductive dehalogenase; dcaA, 1,2-dichloroethane reductive dehalogenase; cprA, chlorophenol reductive dehalogenase; and cbrA, chlorobenzene reductive dehalogenase. The tree was constructed by a neighbor-joining (NJ) method based on an alignment of almost-complete rdhA amino acid sequences with pairwise gap deletion on MEGA version 4.0 software (30). The resulting tree was displayed using Interactive Tree Of Life (19). The scale bar represents 0.1 substitutions per amino acid position.In the alignment of the subseafloor rdhA sequences, we observed two Fe-S cluster-binding motifs as a conserved structure of previously reported reductive dehalogenases (29). The sequences were amplified with primer RRF2 containing the N-terminal twin arginine translocation (Tat) signal sequence and primer B1R containing the rdhB genes encoding a putative dehalogenase membrane anchor protein (17). Thus, the dehalogenases of subseafloor bacteria have a structural framework similar to that of known dehalogenases from terrestrial Dehalococcoides species. However, BLASTP analysis showed that similarities among subseafloor rdhA sequences and previously reported dehalogenase sequences were generally low, ranging from 33.06% to 64.27%. Some sequences were affiliated, with relatively high bootstrap values, with subseafloor rdhA clusters I and II, which are clearly distinct from the rdhA sequences of Dehalococcoides and other known species (Fig. (Fig.1).1). In addition, we were unable to detect subseafloor rdhA genes using other primer sets targeting cprA- and pceA-like genes (26, 32). These results indicate that most subseafloor rdhA genes are distinct from those reported from terrestrial environments, a trend that corroborates the results of a metagenomic survey of subseafloor microbial communities at the Peruvian site (3). However, it is worth noting that the RRF2 and B1R primers used in this study are based on the rdhA sequences present in Dehalococcoides (17) and that sequence retrieval is probably biased by primer mismatch. It is thus likely that there are still unexplored functional genes related to the dehalorespiration pathways in marine subsurface sediments.An interesting finding of the functional gene survey is that the subseafloor rdhA homologues are preferentially detected in shallow sediments. At site C9001 off Japan, the sedimentation ratio is considerably higher than at other sites (54 to 95 cm per 1,000 years) (unpublished data), and rdhA genes were successfully detected in horizons as deep as 358 mbsf (Table (Table1).1). The rdhA genes were also detected in sediments from the open ocean at site 1226, which contained very low concentrations (<0.2%) of organic matter (7). This may be because halogenated compounds are derived not only from terrestrial environments but also from the seawater overlying the sediments. In addition, a diverse range of marine organisms, such as phytoplankton, mollusks, algae, polychaetes, jellyfish, and sponges, are known to produce halogenated organic compounds (11). For example, the amount of brominated organic compounds in the ocean has been estimated at 1 to 2 million tons per year (10). Since these halogenated compounds are generally recalcitrant or not metabolizable by aerobic microorganisms in the seawater column (15), they are effectively buried in marine subsurface sediments. In fact, debromination of brominated phenols in marine, estuarine, or intertidal strait sediments has been reported (4, 9, 16, 22), and a brominated phenol-dehalogenating microbial community has been observed in the marine sponge Aplysina aerophoba, which produces bromophenolic metabolites (1).We also observed reductive dehalogenation activity in subseafloor sediment slurry from site C0002 in the Nankai Trough (Fig. (Fig.2;2; also see the supplemental material). The slurry sample was prepared by mixing sediment samples from 1.9, 4.7, 9.2, 13.4, 20.2, 30.0, 66.6, and 155.4 mbsf. During the initial incubation with 2,4,6-tribromophenol (2,4,6-TBP) for 179 days, 2,4,6-TBP was completely converted to phenol. We then supplemented the same incubation slurry with 2,4,6-TBP and once again observed dehalogenation activity (Fig. (Fig.2A).2A). During the incubation, 2,4-dibromophenol and 4-bromophenol were produced as intermediates (Fig. (Fig.2C),2C), suggesting that ortho debromination occurred in preference to para debromination, as observed previously in marine sponge habitats (1). The maximum phenol production rate during the second incubation was calculated to be 0.094 μM per 1 cm3 of sediment per day (Fig. (Fig.2A2A).Open in a separate windowFIG. 2.Dehalogenation activities of subseafloor microbes. (A) Debromination of 2,4,6-TBP in a subseafloor sediment slurry from site C0002 in the Nankai Trough Forearc Basin. Arrow indicates the timing of 2,4,6-TBP supplementation. (B) Dechlorination of TCE in the same slurry sample. Sterilized control sediment slurries did not exhibit phenol and/or cis-DCE production (data not shown). (C) Potential debromination pathway of 2,4,6-TBP (solid arrows) and (D) potential dechlorination pathway of TCE (solid arrows) observed. The pathways indicated by dashed arrows were not observed in this experiment.Using the same sediment slurry sample, we also observed dehalogenation activity of trichloroethene (TCE), a substantial pollutant in the natural environment. During an incubation lasting more than 200 days, TCE was almost entirely converted to cis-dichloroethene (cis-DCE) (Fig. (Fig.2B).2B). The subsequent dechlorination step of cis-DCE, which is presumably from cis-DCE to monochloroethene, was not observed during the incubation. The rate of cis-DCE production was calculated as 0.045 μM per 1 cm3 of sediment per day.In conclusion, the observed molecular and activity data suggest that metabolically active dehalorespiring microbes are well represented in marine subsurface sediments and that these microbes may be widely distributed in Pacific Ocean margin sediments. Given the relatively high in vitro activity rates, we expect that subseafloor dehalorespiring microbes play important ecological roles in the biogeochemical cycles of chlorine, iodine, and bromine, as well as in halogenated carbon substrates. The distribution of in situ activity rates, chemical and geophysical constraints, metabolic characteristics of the individual dehalorespiring phylotypes, and genetic and enzymatic mechanisms of the microbes remain to be clarified. Nevertheless, the findings of this study provide new evidence of microbial functioning in the subseafloor ecosystem.  相似文献   

10.
11.
12.
13.
14.
15.
16.
This study focuses on two representatives of experimentally uncharacterized haloalkane dehalogenases from the subfamily HLD-III. We report biochemical characterization of the expression products of haloalkane dehalogenase genes drbA from Rhodopirellula baltica SH1 and dmbC from Mycobacterium bovis 5033/66. The DrbA and DmbC enzymes show highly oligomeric structures and very low activities with typical substrates of haloalkane dehalogenases.Haloalkane dehalogenases (EC 3.8.1.5.) acting on halogenated aliphatic hydrocarbons catalyze carbon-halogen bond cleavage, leading to an alcohol, a halide ion, and a proton as the reaction products (7). Haloalkane dehalogenases originating from various bacterial strains have potential for application in bioremediation technologies (4, 6, 22), construction of biosensors (2), decontamination of warfare agents (17), and synthesis of optically pure compounds (19). Recent evolutionary study of haloalkane dehalogenase sequences revealed the existence of three subfamilies, denoted HLD-I, HLD-II, and HLD-III (3). In contrast to subfamilies HLD-I and HLD-II, the subfamily HLD-III is currently lacking experimentally characterized proteins. We have therefore focused on the isolation and study of two selected representatives of the HLD-III subfamily, DrbA and DmbC.The drbA gene was amplified by PCR using the cosmid pircos.a3g10 originating from marine bacterium Rhodopirellula baltica SH1, and the dmbC gene was amplified from DNA originating from obligatory pathogen Mycobacterium bovis 5033/66. Six-histidine tails were added to the C termini of DrbA and DmbC in a cloning step, enabling single-step purification using Ni-nitrilotriacetic acid resin. Haloalkane dehalogenase DrbA was expressed under the T7 promoter and purified, with a resulting yield of 0.1 mg of protein per gram of cell mass. Haloalkane dehalogenase DmbC was obtained by expression in Mycobacterium smegmatis, with a yield of 0.07 mg of purified protein per gram of cell mass.The correct folding and secondary structures of the newly prepared enzymes were verified by circular dichroism (CD) spectroscopy. Far-UV CD spectra were recorded for DrbA and DmbC enzymes and other, related haloalkane dehalogenases. All enzymes tested exhibited CD spectra with two negative features at 208 and 222 nm and one positive peak at 195 nm, which are characteristic of α-helical content (Fig. (Fig.1).1). This suggested that both new enzymes, DrbA and DmbC, were folded correctly. However, DmbC exhibited more intense negative maxima which differed from other haloalkane dehalogenases in the θ222208 ratio. This finding indicated a slight variation in the arrangement of secondary structure elements of the DmbC enzyme. Thermally induced denaturations of DrbA and DmbC were tested in parallel. Both enzymes showed changes in ellipticity during increasing temperature. The melting temperatures calculated from these curves were 45.8 ± 0.4°C for DmbC and 39.4 ± 0.1°C for DrbA. The thermostability results obtained for DrbA and DmbC were in good agreement with the range of melting temperatures determined for other, related haloalkane dehalogenases.Open in a separate windowFIG. 1.Far-UV CD spectra of DrbA, DmbC, and seven different biochemically characterized haloalkane dehalogenases. Protein concentration used for far-UV CD spectrum measurement was 0.2 mg/ml.The sizes of the purified proteins were estimated by electrophoresis under native conditions conducted using a 10% polyacrylamide gel (Fig. (Fig.2).2). More precise determination of the sizes of DrbA and DmbC was achieved by gel filtration chromatography performed on Sephacryl S-500 HR (GE Healthcare, Uppsala, Sweden), calibrated with blue dextran 2000, thyroglobulin (669 kDa), ferritin (440 kDa), catalase (240 kDa), conalbumin (75 kDa), and ovalbumin (43 kDa) (Fig. (Fig.3A).3A). Both DrbA and DmbC were eluted from the column in the fraction prior to blue dextran, indicating that both enzymes form oligomeric complexes of a size larger than 2,000 kDa (Fig. 3B and C). The haloalkane dehalogenases which have been biochemically characterized so far form monomers, except for DbjA isolated from Bradyrhizobium japonicum USDA110 (21), which shows monomeric, dimeric, and tetrameric forms according to the pH of the buffer (R. Chaloupkova, submitted for publication).Open in a separate windowFIG. 2.Native protein electrophoresis of DrbA and DmbC. Lane 1, carbonic anhydrase (29 kDa); lane 2, ovalbumin (43 kDa); lane 3, bovine albumin (67 kDa); lane 4, conalbumin (75 kDa); lane 5, catalase (240 kDa); lane 6, ferritin (440 kDa); lane 7, DrbA; lane 8, DmbC.Open in a separate windowFIG. 3.Gel filtration chromatogram of DrbA and DmbC. (A) The following calibration kit samples (0.5 ml of a concentration of 2 mg/ml protein loaded) were analyzed using 50 mM Tris-HCl with 150 mM NaCl, pH 7.5, as elution buffer: blue dextran (line 1, 9.6-ml fraction), thyroglobulin (line 2, 15.95-ml fraction), ferritin (line 3, 16.78-ml fraction), ovalbumin (line 4, 18.55-ml fraction), and RNase A (line 5, 20.08-ml fraction). (B and C) Haloalkane dehalogenase DrbA eluted in the 9.03-ml fraction (B), and haloalkane dehalogenase DmbC in the 9.31-ml fraction (C).The substrate specificities of DrbA and DmbC were investigated with a set of 30 selected chlorinated, brominated, and iodinated hydrocarbons. Standardized specific activities related to 1-chlorobutane (summarized in Table Table1)1) were compared with the activity profiles of other haloalkane dehalogenases (Fig. (Fig.4).4). DrbA and DmbC displayed similar activity patterns, with catalytic activities approximately two orders of magnitude lower than those of other known haloalkane dehalogenases (1, 5, 8-11, 13-16, 18, 20, 23). HLD-III subfamily enzymes showed a restricted specificity range and a preference for iodinated short-chain hydrocarbons. Both phenomena may be related to the composition of the catalytic pentad Asp-His-Asp+Asn-Trp, which is unique to the members of the HLD-III subfamily (3). The preference for substrates carrying an iodine substituent can be related to a pair of halide-binding residues and their spatial arrangement with the catalytic triad. These residues make up the catalytic pentad, playing a critical role in substrate binding, formation of the transition states, and the reaction intermediates of the dehalogenation reaction (12).Open in a separate windowFIG. 4.Substrate specificity profiles of DrbA, DmbC, and seven different biochemically characterized haloalkane dehalogenases. Activities were determined using a consistent set of 30 halogenated substrates (see Table Table1).1). Data were standardized by dividing each value by the sum of all activities determined for individual enzymes in order to mask the differences in absolute activities. Specific activities (in μmol·s−1·mg−1) with 1-chlorobutane are 0.0003 (DrbA), 0.0001 (DmbC), 0.0003 (DatA), 0.0133 (DbjA), 0.0010 (DbeA), 0.0128 (DhaA), 0.0231 (LinB), 0.0171 (DmbA), and 0.0117 (DhlA).

TABLE 1.

Specific activities of haloalkane dehalogenases DrbA and DmbC toward a set of 30 halogenated hydrocarbonsa
SubstrateDrbA
DmbC
Sp act (nmol product·s−1· mg−1 protein)Relative activity (%)Sp act (nmol product·s−1· mg−1 protein)Relative activity (%)
1-Chlorobutane0.2911000.122100
1-Chlorohexane0.129440.122100
1-Bromobutane0.081281.2211,000
1-Bromohexane0.181620.977800
1-Iodopropane0.143492.1981,800
1-Iodobutane0.5061742.5642,100
1-Iodohexane0.095330.244200
1,2-DichloroethaneNANANANA
1,3-DichloropropaneNANA0.01210
1,5-DichloropentaneNANA0.06150
1,2-Dibromoethane0.098340.855700
1,3-DibromopropaneNANA5.0074,100
1-Bromo-3-chloropropane0.00101.4651,200
1,3-Diiodopropane0.3581236.7165,500
2-Iodobutane0.0289NANA
1,2-DichloropropaneNANANANA
1,2-Dibromopropane0.148510.244200
2-Bromo-1-chloropropane0.091310.488400
1,2,3-TrichloropropaneNANANANA
Bis-(2-chloroethyl) etherNANANANA
ChlorocyclohexaneNANANANA
Bromocyclohexane0.0269NANA
(1-Bromomethyl)-cyclohexaneNANA0.08973
1-Bromo-2-chloroethane0.167570.11191
ChlorocyclopentaneNANANANA
4-Bromobutyronitrile0.200690.444364
1,2,3-TribromopropaneNANA0.222182
3-Chloro-2-methyl propeneNANANANA
2,3-Dichloropropene0.27695NANA
1,2-Dibromo-3-chloropropane0.01030.04436
Open in a separate windowaNA, no activity detected.Substrates 1-iodobutane and 1,3-diiodopropane, identified as the best substrates for haloalkane dehalogenases DrbA and DmbC, were used for measuring the dependency of enzyme activity on temperature and for determination of the pH optima. DrbA exhibited the highest activity with 1-iodobutane at 50°C, although above this temperature, the enzyme rapidly became inactivated. DmbC showed the highest activity toward 1,3-diiodopropane at 40°C, which is similar to the temperature determined with the haloalkane dehalogenases DmbA and DmbB (45°C), isolated from the same species (10). Irrespective of the reaction temperature, DrbA showed the maximum activity at pH 9.15. DrbA kept 80% of its activity throughout a relatively wide range of pH values (pH 7.00 and 9.91) compared to DmbC, which showed a sharp maximum at pH 8.30. The Michaelis-Menten kinetics of DrbA and DmbC determined by isothermal titration microcalorimetry were investigated with 1-iodobutane, which is an iodinated analogue of 1-chlorobutane routinely used for characterization of haloalkane dehalogenases. The low magnitudes of the Michaelis constants (Km = 0.063 ± 0.003 mM for DrbA and 0.018 ± 0.001 mM for DmbC) suggest a high affinity of both enzymes for 1-iodobutane. The catalytic constants determined with 1-iodobutane (kcat = 0.128 ± 0.002 s−1 for DrbA and 0.0715 ± 0.0004 s−1 for DmbC) suggest that the low specific activities observed during substrate screening are not due to poor affinity but are instead due to a low conversion rate.The biochemical characteristics of purified DrbA and DmbC suggest that these proteins represent novel enzymes differing from previously characterized haloalkane dehalogenases by (i) their unique ability to form oligomers and (ii) low levels of dehalogenating activity with typical substrates of haloalkane dehalogenases. This study further illustrates how genome sequencing projects and phylogenetic analyses contribute to the identification of novel enzymes. Characterization of DrbA and DmbC, belonging to the subfamily HLD-III, partially filled a gap in the knowledge of the haloalkane dehalogenase family and provided an additional insight into evolutionary relationships among its members.  相似文献   

17.
18.
19.
20.
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