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 187 毫秒
1.
2.
Canada geese (Branta canadensis) are prevalent in North America and may contribute to fecal pollution of water systems where they congregate. This work provides two novel real-time PCR assays (CGOF1-Bac and CGOF2-Bac) allowing for the specific and sensitive detection of Bacteroides 16S rRNA gene markers present within Canada goose feces.The Canada goose (Branta canadensis) is a prevalent waterfowl species in North America. The population density of Canada geese has doubled during the past 15 years, and the population was estimated to be close to 3 million in 2007 (4). Canada geese often congregate within urban settings, likely due to available water sources, predator-free grasslands, and readily available food supplied by humans (6). They are suspected to contribute to pollution of aquatic environments due to the large amounts of fecal matter that can be transported into the water. This can create a public health threat if the fecal droppings contain pathogenic microorganisms (6, 7, 9, 10, 12, 13, 19). Therefore, tracking transient fecal pollution of water due to fecal inputs from waterfowl, such as Canada geese, is of importance for protecting public health.PCR detection of host-specific 16S rRNA gene sequences from Bacteroidales of fecal origin has been described as a promising microbial source-tracking (MST) approach due to its rapidity and high specificity (2, 3). Recently, Lu et al. (15) characterized the fecal microbial community from Canada geese by constructing a 16S rRNA gene sequence database using primers designed to amplify all bacterial 16S rRNA gene sequences. The authors reported that the majority of the 16S rRNA gene sequences obtained were related to Clostridia or Bacilli and to a lesser degree Bacteroidetes, which represent possible targets for host-specific source-tracking assays.The main objective of this study was to identify novel Bacteroidales 16S rRNA gene sequences that are specific to Canada goose feces and design primers and TaqMan fluorescent probes for sensitive and specific quantification of Canada goose fecal contamination in water sources.Primers 32F and 708R from Bernhard and Field (2) were used to construct a Bacteroidales-specific 16S rRNA gene clone library from Canada goose fecal samples (n = 15) collected from grass lawns surrounding Wascana Lake (Regina, SK, Canada) in May 2009 (for a detailed protocol, see File S1 in the supplemental material). Two hundred eighty-eight clones were randomly selected and subjected to DNA sequencing (at the Plant Biotechnology Institute DNA Technologies Unit, Saskatoon, SK, Canada). Representative sequences of each operational taxonomic unit (OTU) were recovered using an approach similar to that described by Mieszkin et al. (16). Sequences that were less than 93% similar to 16S rRNA gene sequences from nontarget host species in GenBank were used in multiple alignments to identify regions of DNA sequence that were putatively goose specific. Subsequently, two TaqMan fluorescent probe sets (targeting markers designated CGOF1-Bac and CGOF2-Bac) were designed using the RealTimeDesign software provided by Biosearch Technologies (http://www.biosearchtech.com/). The newly designed primer and probe set for the CGOF1-Bac assay included CG1F (5′-GTAGGCCGTGTTTTAAGTCAGC-3′) and CG1R (5′-AGTTCCGCCTGCCTTGTCTA-3′) and a TaqMan probe (5′-6-carboxyfluorescein [FAM]-CCGTGCCGTTATACTGAGACACTTGAG-Black Hole Quencher 1 [BHQ-1]-3′), and the CGOF2-Bac assay had primers CG2F (5′-ACTCAGGGATAGCCTTTCGA-3′) and CG2R (5′-ACCGATGAATCTTTCTTTGTCTCC-3′) and a TaqMan probe (5′-FAM-AATACCTGATGCCTTTGTTTCCCTGCA-BHQ-1-3′). Oligonucleotide specificities for the Canada goose-associated Bacteroides 16S rRNA primers were verified through in silico analysis using BLASTN (1) and the probe match program of the Ribosomal Database Project (release 10) (5). Host specificity was further confirmed using DNA extracts from 6 raw human sewage samples from various geographical locations in Saskatchewan and 386 fecal samples originating from 17 different animal species in Saskatchewan, including samples from Canada geese (n = 101) (Table (Table1).1). An existing nested PCR assay for detecting Canada goose feces (15) (targeting genetic marker CG-Prev f5) (see Table S1 in the supplemental material) was also tested for specificity using the individual fecal and raw sewage samples (Table (Table1).1). All fecal DNA extracts were obtained from 0.25 g of fecal material by using the PowerSoil DNA extraction kit (Mo Bio Inc., Carlsbad, CA) (File S1 in the supplemental material provides details on the sample collection).

TABLE 1.

Specificities of the CGOF1-Bac, CGOF2-Bac, and CG-Prev f5 PCR assays for different species present in Saskatchewan, Canada
Host group or sample typeNo. of samplesNo. positive for Bacteroidales marker:
CGOF1-BacCGOF2-BacCG-Prev f5All-Bac
Individual human feces2500125
Raw human sewage60006
Cows4100041
Pigs4800148
Chickens3400834
Geese10158515995a
Gulls1600614
Pigeons2510222
Ducks1000010
Swans10001
Moose1000010
Deer
    White tailed1000010
    Mule1000010
    Fallow1000010
Caribou1000010
Bison1000010
Goats1000010
Horses1500015
Total392595177381
Open in a separate windowaThe 6 goose samples that tested negative for the All-Bac marker also tested negative for the three goose markers.The majority of the Canada goose feces analyzed in this study (94%; 95 of 101) carried the Bacteroidales order-specific genetic marker designated All-Bac, with a relatively high median concentration of 8.2 log10 copies g1 wet feces (Table (Table11 and Fig. Fig.1).1). The high prevalence and abundance of Bacteroidales in Canada goose feces suggested that detecting members of this order could be useful in identifying fecal contamination associated with Canada goose populations.Open in a separate windowFIG. 1.Concentrations of the Bacteroidales (All-Bac, CGOF1-Bac, and CGOF2-Bac) genetic markers in feces from various individual Canada geese.The composition of the Bacteroidales community in Canada goose feces (n = 15) was found to be relatively diverse since 52 OTUs (with a cutoff of 98% similarity) were identified among 211 nonchimeric 16S rRNA gene sequences. Phylogenetic analysis of the 52 OTUs (labeled CGOF1 to CGOF52) revealed that 43 (representing 84% of the 16S rRNA gene sequences) were Bacteroides like and that 9 (representing 16% of the 16S rRNA gene sequences) were likely to be members of the Prevotella-specific cluster (see Fig. S2 in the supplemental material). Similarly, Jeter et al. (11) reported that 75.7% of the Bacteroidales 16S rRNA clone library sequences generated from goose fecal samples were Bacteroides like. The majority of the Bacteroides- and Prevotella-like OTUs were dispersed among a wide range of previously characterized sequences from various hosts and did not occur in distinct clusters suitable for the design of Canada goose-associated real-time quantitative PCR (qPCR) assays (see Fig. S2 in the supplemental material). However, two single Bacteroides-like OTU sequences (CGOF1 and CGOF2) contained putative goose-specific DNA regions that were identified by in silico analysis (using BLASTN, the probe match program of the Ribosomal Database Project, and multiple alignment). The primers and probe for the CGOF1-Bac and CGOF2-Bac assays were designed with no mismatches to the clones CGOF1 and CGOF2, respectively.The CGOF2-Bac assay demonstrated no cross-amplification with fecal DNA from other host groups, while cross-amplification for the CGOF1-Bac assay was limited to one pigeon fecal sample (1 of 25, i.e., 4% of the samples) (Table (Table1).1). Since the abundance in the pigeon sample was low (3.3 log10 marker copies g1 feces) and detection occurred late in the qPCR (with a threshold cycle [CT] value of 37.1), it is unlikely that this false amplification would negatively impact the use of the assay as a tool for detection of Canada goose-specific fecal pollution in environmental samples. In comparison, the nested PCR CG-Prev f5 assay described by Lu and colleagues (15) demonstrated non-host-specific DNA amplification with fecal DNA samples from several animals, including samples from humans, pigeons, gulls, and agriculturally relevant pigs and chickens (Table (Table11).Both CGOF1-Bac and CGOF2-Bac assays showed limits of quantification (less than 10 copies of target DNA per reaction) similar to those of other host-specific Bacteroidales real-time qPCR assays (14, 16, 18). The sensitivities of the CGOF1-Bac and CGOF2-Bac assays were 57% (with 58 of 101 samples testing positive) and 50% (with 51 of 101 samples testing positive) for Canada goose feces, respectively (Table (Table1).1). A similar sensitivity of 58% (with 59 of 101 samples testing positive) was obtained using the CG-Prev f5 PCR assay. The combined use of the three assays increased the detection level to 72% (73 of 101) (Fig. (Fig.2).2). Importantly, all markers were detected within groups of Canada goose feces collected each month from May to September, indicating relative temporal stability of the markers. The CG-Prev f5 PCR assay is an end point assay, and therefore the abundance of the gene marker in Canada goose fecal samples could not be determined. However, development of the CGOF1-Bac and CGOF2-Bac qPCR approach allowed for the quantification of the host-specific CGOF1-Bac and CGOF2-Bac markers. In the feces of some individual Canada geese, the concentrations of CGOF1-Bac and CGOF2-Bac were high, reaching levels up to 8.8 and 7.9 log10 copies g1, respectively (Fig. (Fig.11).Open in a separate windowFIG. 2.Venn diagram for Canada goose fecal samples testing positive with the CGOF1-Bac, CGOF2-Bac, and/or CG-Prev f5 PCR assay. The number outside the circles indicates the number of Canada goose fecal samples for which none of the markers were detected.The potential of the Canada goose-specific Bacteroides qPCR assays to detect Canada goose fecal pollution in an environmental context was tested using water samples collected weekly during September to November 2009 from 8 shoreline sampling sites at Wascana Lake (see File S1 and Fig. S1 in the supplemental material). Wascana Lake is an urban lake, located in the center of Regina, that is routinely frequented by Canada geese. In brief, a single water sample of approximately 1 liter was taken from the surface water at each sampling site. Each water sample was analyzed for Escherichia coli enumeration using the Colilert-18/Quanti-Tray detection system (IDEXX Laboratories, Westbrook, ME) (8) and subjected to DNA extraction (with a PowerSoil DNA extraction kit [Mo Bio Inc., Carlsbad, CA]) for the detection of Bacteroidales 16S rRNA genetic markers using the Bacteroidales order-specific (All-Bac) qPCR assay (14), the two Canada goose-specific (CGOF1-Bac and CGOF2-Bac) qPCR assays developed in this study, and the human-specific (BacH) qPCR assay (17). All real-time and conventional PCR procedures as well as subsequent data analysis are described in the supplemental material and methods. The E. coli and All-Bac quantification data demonstrated that Wascana Lake was regularly subjected to some form of fecal pollution (Table (Table2).2). The All-Bac genetic marker was consistently detected in high concentrations (6 to 7 log10 copies 100 ml1) in all the water samples, while E. coli concentrations fluctuated according to the sampling dates and sites, ranging from 0 to a most probable number (MPN) of more than 2,000 100 ml1. High concentrations of E. coli were consistently observed when near-shore water experienced strong wave action under windy conditions or when dense communities of birds were present at a given site and time point.

TABLE 2.

Levels of E. coli and incidences of the Canada goose-specific (CGOF1-Bac and CGOF2-Bac), human-specific (BacH), and generic (All-Bac) Bacteroidales 16S rRNA markers at the different Wascana Lake sites sampled weeklya
SiteE. coli
All-Bac
CGOF1-Bac
CGOF2-Bac
BacH
No. of positive water samples/total no. of samples analyzed (%)Min level-max level (MPN 100 ml−1)Mean level (MPN 100 ml−1)No. of positive water samples/total no. of samples analyzed (%)Min level-max level (log copies 100 ml−1)Mean level (log copies 100 ml−1)No. of positive water samples/total no. of samples analyzed (%)Min level-max level (log copies 100 ml−1)Mean level (log copies 100 ml−1)No. of positive water samples/total no. of samples analyzed (%)Min level-max level (log copies 100 ml−1)Mean level (log copies 100 ml−1)No. of positive water samples/total no. of samples analyzedMin level-max level (log copies 100 ml−1)Mean level (log copies 100 ml−1)
W18/8 (100)6-19671.18/8 (100)6.2-8.16.96/8 (75)0-4.72.44/8 (50)0-41.72/80-3.71.7
W29/10 (90)0-1,12019410/10 (100)5.8-6.86.49/10 (90)0-3.72.68/10 (80)0-3.32.20/1000
W310/10 (100)6-1,55053410/10 (100)6-7.8710/10 (100)2.9-4.83.810/10 (100)2-4.53.40/1000
W410/10 (100)16-1,73252910/10 (100)6.4-7.6710/10 (100)3.2-4.63.910/10 (100)2.8-4.33.40/1000
W510/10 (100)2-2,42068710/10 (100)5.5-6.96.37/10 (70)0-3.21.75/10 (50)0-3.11.20/1000
W610/10 (100)3-1,99038910/10 (100)5.5-76.39/10 (90)0-4.32.86/10 (60)0-5.121/100-3.41.3
W77/7 (100)5-2,4204457/7 (100)5.7-7.876/7 (86)0-3.82.65/7 (71)0-4.42.42/70-5.12.8
W810/10 (100)17-98016010/10 (100)6.3-8.67.18/10 (80)0-4.62.87/10 (70)0-4.42.30/1000
Open in a separate windowaMin, minimum; max, maximum.The frequent detection of the genetic markers CGOF1-Bac (in 65 of 75 water samples [87%]), CGOF2-Bac (in 55 of 75 samples [73%]), and CG-Prev f5 (in 60 of 75 samples [79%]) and the infrequent detection of the human-specific Bacteroidales 16S rRNA gene marker BacH (17) (in 5 of 75 water samples [7%[) confirmed that Canada geese significantly contributed to the fecal pollution in Wascana Lake during the sampling period. Highest mean concentrations of both CGOF1-Bac and CGOF2-Bac markers were obtained at the sampling sites W3 (3.8 and 3.9 log10 copies 100 ml1) and W4 (3.4 log10 copies 100 ml1 for both), which are heavily frequented by Canada geese (Table (Table2),2), further confirming their significant contribution to fecal pollution at these particular sites. It is worth noting that concentrations of the CGOF1-Bac and CGOF2-Bac markers in water samples displayed a significant positive relationship with each other (correlation coefficient = 0.87; P < 0.0001), supporting the accuracy of both assays for identifying Canada goose-associated fecal pollution in freshwater.In conclusion, the CGOF1-Bac and CGOF2-Bac qPCR assays developed in this study are efficient tools for estimating freshwater fecal inputs from Canada goose populations. Preliminary results obtained during the course of the present study also confirmed that Canada geese can serve as reservoirs of Salmonella and Campylobacter species (see Fig. S3 in the supplemental material). Therefore, future work will investigate the cooccurence of these enteric pathogens with the Canada goose fecal markers in the environment.  相似文献   

3.
4.
5.
Sulfonamide-resistant Escherichia coli and Salmonella isolates from pigs and chickens in Ontario and Québec were screened for sul1, sul2, and sul3 by PCR. Each sul gene was distributed differently across populations, with a significant difference between distribution in commensal E. coli and Salmonella isolates and sul3 restricted mainly to porcine E. coli isolates.Resistance to sulfonamides is frequent in bacteria from farm animals (7, 8, 9, 10) and is usually caused by the acquisition of the genes sul1, sul2, and sul3 (20, 22). The objectives of this study were (i) to assess the distribution of these genes in Escherichia coli and Salmonella enterica isolates in swine and chickens from two major provinces in Canada, (ii) to assess whether differences occur in the distribution of these genes among bacterial species found within two different animal host species, and (iii) to assess whether significant differences in the distribution of these genes are present between the commensal E. coli strains used as indicators for surveillance of antimicrobial resistance and the zoonotic Salmonella pathogens found in the same ecological niche. In contrast to previous studies, a multivariable logistic regression model was used to analyze the data, control for confounding factors, and assess the interaction effect between animal and bacterial species in terms of the probability of an isolate carrying a specific sul gene. The distribution of sulfonamide resistance genes among sulfonamide-resistant E. coli (393 isolates from chickens and 311 from swine) and Salmonella (13 isolates from chickens and 221 from swine) isolates was assessed. These isolates were collected by the Canadian Integrated Program for Antimicrobial Resistance Surveillance (CIPARS) between 2003 and 2005 from ceca of apparently healthy animals at abattoirs in Ontario (n = 435) and Québec (n = 503). The methods used by CIPARS are presented in detail elsewhere (8-10). The isolates were screened with a previously published multiplex PCR for sul1, sul2, and sul3 (16). The sul1 and sul2 genes were found in E. coli and Salmonella isolates from both animal species. The sul3 gene was detected in both E. coli and Salmonella isolates from swine but only in E. coli isolates from chickens (Table (Table1).1). Three percent of the isolates had no detectable sul gene, 12.5% possessed two genes, and two isolates carried three genes (Table (Table1).1). Similar (2, 3, 14, 19) or higher (4, 11, 17) values for multiple genes have been reported by others. The overall higher prevalence of sul1 in Salmonella isolates and of sul2 and sul3 in E. coli isolates was in agreement with the results of previous studies (2-4, 11, 12, 14, 21).

TABLE 1.

Frequency of the three resistance genes sul1, sul2, and sul3 in sulfonamide-resistant E. coli and Salmonella isolates from chickens and swine in Ontario and Québec between 2003 and 2005
Bacterial speciesSource (no. of isolates)No. (%) of isolates with indicated gene(s)
sul1sul2sul3sul1 + sul2sul1+ sul3sul2 + sul3sul1 + sul2 + sul3None
E. coliSwine (311)61 (19.6)66 (21.2)132 (42.4)11 (3.5)9 (2.9)11 (3.5)2 (0.6)19 (6.1)
Chickens (393)103 (26.2)211 (53.7)9 (2.3)64 (16.3)0 (0.0)0 (0.0)0 (0.0)6 (1.5)
Total (704)164 (23.3)277 (39.3)141 (20.0)75 (10.7)9 (1.3)11 (1.6)2 (0.3)25 (3.6)
SalmonellaSwine (221)173 (78.3)20 (9.0)5 (2.3)7 (3.2)0 (0.0)14 (6.3)0 (0.0)2 (0.9)
Chickens (13)7 (53.8)5 (38.5)0 (0.0)0 (0.0)0 (0.0)0 (0.0)0 (0.0)1 (7.7)
Total (234)180 (76.9)25 (10.7)5 (2.1)7 (3.0)0 (0.0)14 (6.0)0 (0.0)3 (1.3)
Open in a separate windowThree logistic regression models (Table (Table2)2) for associations between the presence of each sul gene and the bacterial species, animal species, province of origin of the animals, and year of isolation were built using Stata 9 (StataCorp, College Station, TX). Statistical interactions between bacterial and animal species were assessed in the sul1 and sul2 models but not in the sul3 model (the sample size was insufficient). Tests were two tailed, with significance at a P value of ≤0.05. A significant interaction between bacterial and animal species was observed for sul1 (Table (Table2).2). The presence of statistical interactions indicates that the effect of each variable depends on the state of the other variable. Thus, the effects of bacterial and animal species on the distribution of sul1 cannot be interpreted independently. For instance, sul1 was more frequent in Salmonella isolates from swine than from chickens, but the reverse was true for E. coli isolates (Table (Table3).3). Genomic islands containing sul1 are present in some Salmonella serovars, including S. enterica serovar Typhimurium and S. Derby (1, 6, 18), which were the most frequent in the swine samples of this study (Table (Table4).4). This contrasts with E. coli strains, in which sul1 is usually associated with transposons and large transferable plasmids (22). Thus, the presence of significant statistical interactions in the sul1 model could be an indicator of the differential importance of horizontal gene transfer in E. coli strains versus clonal expansion of specific Salmonella serovars in the animal species investigated. No significant interaction was detected for sul2 (P = 0.66) (Table (Table2).2). This could be the consequence of the relatively small number of sul2-positive Salmonella isolates and the resultant lack of statistical power; however, it is also possible that sul2 has a different epidemiology than sul1. The sul2 gene has not been shown to be located on genomic islands and is usually plasmid borne (22). It may therefore be transferred more frequently than sul1 between E. coli and Salmonella and between bacteria from swine and chickens, leading to the absence of a significant interaction in the sul2 model. Serotyping, molecular typing, and assessment of gene transferability would be required to test these hypotheses.

TABLE 2.

Multivariable models for the distribution of the sul1, sul2, and sul3 sulfonamide resistance genes from swine and chickens at slaughter in Ontario and Québec between 2003 and 2005
Explanatory variable (referent group) or interactionOdds ratio (95% confidence interval)b
sul1sul2sul3
Salmonella (E. coli)1.54 (0.50-4.74)0.33 (0.21-0.51)0.10 (0.06-0.16)
Swine (chickens)0.49 (0.36-0.68)c0.16 (0.12-0.23)c39.57 (20.21-77.48)c
Québec (Ontario)0.80 (0.60-1.07)2.20 (1.61-2.99)c0.83 (0.56-1.23)
2004 (2003)0.77 (0.56-1.06)0.99 (0.71-1.40)1.84 (1.18-2.86)
2005 (2003)0.68 (0.45-1.04)1.36 (0.88-2.09)1.66 (0.99-2.79)
2005 (2004)0.88 (0.58-1.36)1.36 (0.88-2.11)0.90 (0.53-1.53)
Bacterial species × animal speciesa12.84 (3.79-43.46)cNINI
Open in a separate windowaStatistical interaction between data for bacterial species and animal species. The terms for this interaction are described in Table Table33.bExample of odds ratio interpretation: the odds of a porcine isolate carrying sul3 were 39.57 times higher than for an isolate obtained from chickens. NI, not included in the model. The sul2 model was not included because the P value for this interaction was equal to 0.66, and the sul3 model was not included because it did not converge when including this interaction.cP < 0.001.

TABLE 3.

Interaction terms for the association between bacterial and animal species for the presence of sul1
Contrast variablesOdds ratioa95% Confidence intervalP value
Salmonella in pigs vs Salmonella in chickens6.301.95-20.390.002
E. coli in pigs vs E. coli in chickens0.490.36-0.67<0.001
Salmonella in chickens vs E. coli in chickens1.540.50-4.740.451
Salmonella in pigs vs E. coli in pigs19.7912.28-31.87<0.001
Open in a separate windowaExample of interpretation: the interaction effect suggests that sul1-mediated sulfonamide resistance is 19.79 times more likely to occur in Salmonella in pigs than in E. coli in pigs.

TABLE 4.

Salmonella serovars and frequency of resistance genes detected in each serovar
Salmonella serovarNo. of isolatesaNo. of isolates with indicated genea
sul1sul2sul3
Agona14/014/00/00/0
Berta3/00/03/00/0
Bovismorbificans1/00/00/00/0
Brandenburg3/01/01/00/0
Derby63/059/04/02/0
Give O:15+1/01/00/00/0
Heidelberg2/40/30/02/0
4,12:−:−3/03/00/00/0
4,12:i:−2/02/00/00/0
4,5,12:−:−1/01/00/01/0
Rough-O:fg:−1/01/00/00/0
Rough-O:l,v:enz151/01/01/00/0
Infantis3/01/02/00/0
Johannesburg1/01/00/00/0
London2/01/00/01/0
Manhattan2/00/02/00/0
Mbandaka6/06/01/00/0
Ohio O:14+1/00/01/00/0
Putten1/01/00/00/0
Schwarzengrund0/60/10/50/0
Typhimurium110/3101/312/013/0
Total221/13194/727/519/0
Open in a separate windowaThe first and second numbers in each column represent swine and chicken isolates, respectively; the total number of tabulated occurrences of sulfonamide genes is larger than the number of resistant isolates investigated because some isolates carried several genes simultaneously.A significant increase in the prevalence of sul3 was detected between 2003 and 2004 (P = 0.007) but not between 2004 and 2005 (P = 0.055) nor at any time for sul1 and sul2 (Table (Table2).2). The sul3 gene has emerged recently (5), and its prevalence was probably still increasing in 2003. The odds of finding sul3 in Salmonella isolates were 10 times lower than for E. coli isolates and 40 times higher in swine than in chickens (Table (Table2).2). Other studies have also found high frequencies of sul3 in porcine E. coli isolates (5, 12, 20) and much lower frequencies in other sources (4, 11, 12, 14). Although these isolates were not typed, previous results have shown that sul3 is present in both pathogenic and commensal porcine E. coli isolates (5) of at least 13 different serotypes (P. Boerlin and R. M. Travis, unpublished data). The sul3 gene has spread extensively across porcine E. coli populations in North America and Europe but remains uncommon in other major farm animal species and in Salmonella populations (Table (Table1)1) (2, 13). It may require more time to spread to other populations. There may be biological and ecological barriers slowing its spread to bacteria of other animal species or coselection factors that favor its presence in porcine E. coli populations.Using the example of sulfonamides as a model for the application of multivariable statistical approaches to the study of antimicrobial resistance epidemiology, this study indicates that the relative frequencies of genes encoding resistance to the same antimicrobial either present in bacterial populations for decades or recently emerged can vary significantly between animal hosts and phylogenetically related bacterial species sharing the same ecological niche. Differences in the distribution of resistance determinants may remain hidden when assessing resistance phenotypes. Similar antimicrobial susceptibility results do not necessarily imply similar resistance genes. These findings highlight the need to further explore the interactions between commensals and pathogens and the ways in which commensal bacteria are interpreted as indicators of antimicrobial resistance in pathogens.  相似文献   

6.
7.
Mycobacterium avium subsp. paratuberculosis, the causative agent of Johne''s disease in cattle, was identified in settled-dust samples of Dutch commercial dairy farms, both in the dairy barn and in the young stock housing. Bioaerosols may play a role in within-farm M. avium subsp. paratuberculosis transmission.Paratuberculosis is an infectious enteric disease caused by Mycobacterium avium subsp. paratuberculosis leading to economic losses in dairy cattle globally (2, 10). The main transmission route is the fecal/oral route from infectious adult cattle to susceptible calves (12).Preventive calf management was a key point in model studies (7), but 20-year implementation did not lead to farm-level eradication, suggesting uncontrolled routes of transmission (1, 7).Environmental samples were used to classify commercial dairy herds (3, 9, 11), based on long-term survival of M. avium subsp. paratuberculosis in the environment (16). Recently, bioaerosols containing viable M. avium subsp. paratuberculosis were identified in an experimental setting with 100% M. avium subsp. paratuberculosis prevalence (6) and may thus be a mode of transmission. Dust containing M. avium subsp. paratuberculosis might be ingested or inhaled by calves (4). Experimental M. avium subsp. paratuberculosis challenge studies in sheep successfully used inhalation (8). These transmission routes could hamper current control programs. Our objective was to study whether M. avium subsp. paratuberculosis could be detected in bioaerosols on commercial Dutch dairy farms.Dairy herds in three Dutch veterinary practices were sampled in 2009. All farms participated in a Dutch M. avium subsp. paratuberculosis monitoring program in 2008, either the Dutch Paratuberculosis Program (PPN; n = 2) or the Dutch Bulk Milk Quality Assurance Program (BMQAP; n = 22) (15). Both PPN herds were certified M. avium subsp. paratuberculosis-free. Herds corresponding to the BMQAP had at least one positive animal identified by enzyme-linked immunosorbent assay (ELISA) (Pourquier ELISA; Institut Pourquier, France). Farms were grouped into three M. avium subsp. paratuberculosis test prevalence levels (control, zero positive animals; group A, one positive animal; group B, two or more positive animals; Table Table11).

TABLE 1.

Overview of the results of the questionnaire about relevant M. avium subsp. paratuberculosis management practicesa
ParameterValue for groupb
Control (n = 2)A (n = 8)B (n = 14)
Mean herd size (SD)69 (15)67 (19)102 (26)
Median no. of ELISA-positive cows (maximum)0 (0)1 (1)3 (10)
No. of farms with:
    Cow brush in barn2513
    Cow barn cleaned in summer with high-pressure cleaner064
    Dry cows in young stock housing033
    Young stock housed separately178
    Young stock housing empty in summer000
    Young stock housing cleaned with high-pressure cleaner061
Open in a separate windowaResults of the questionnaire about relevant M. avium subsp. paratuberculosis management practices in 24 Dutch farms enrolled in this study with 0 (control), 1 (group A), or ≥2 (group B) ELISA-positive animals.bn, number of farms.Farms were visited twice during the housing period. Sampling locations were above the animal level inside the barn. At the first visit (sampling 1 [S1]), settled dust was collected with wipes and a short management questionnaire was taken. At the same time, five to seven electrostatic dust collectors (EDC; Zeeman, Alphen a/d Rhijn, Netherlands) were installed and collected after 4 weeks (sampling 2 [S2]) (6). Settled-dust samples were processed according to a previously described method (6). Results are presented as proportions of positive locations. McNemar''s χ2 test was performed to investigate whether S1 differed from S2.No M. avium subsp. paratuberculosis was detected by real-time PCR in any of the settled-dust samples at control farms (Fig. (Fig.1).1). M. avium subsp. paratuberculosis DNA was detected in dust samples at S1 and S2 in more than 50% of the group A and B farms, with seven farms consistently positive. M. avium subsp. paratuberculosis DNA was detected in the young stock area in 3/6 (S1) and 2/6 (S2) farms of group B with single-barn housing. M. avium subsp. paratuberculosis DNA was also detected in settled-dust samples from separate young stock housings in three farms, of which two cohoused dry cows.Open in a separate windowFIG. 1.Proportions of farms with M. avium subsp. paratuberculosis DNA detected in settled-dust samples collected at samplings 1 and 2. Black bar, control (n = 2); checked bar, group A (n = 8); white bar, group B (n = 14).At control farms, no viable M. avium subsp. paratuberculosis was detected in any of the collected dust samples (Fig. (Fig.2).2). Viable M. avium subsp. paratuberculosis was detected in 6 B farms at S1. At S2, viable bacteria were present in 3 A farms and in the majority of B farms (Table (Table2).2). On five farms in group B, viable M. avium subsp. paratuberculosis was detected at both samplings.Open in a separate windowFIG. 2.Proportions of farms with viable M. avium subsp. paratuberculosis detected in settled-dust samples collected at samplings 1 and 2. Black bar, control (n = 2); checked bar, group A (n = 8); white bar, group B (n = 14).

TABLE 2.

Detection of M. avium subsp. paratuberculosis DNA or viable M. avium subsp. paratuberculosis in 5 to 7 settled-dust samples collected at sampling 1 or 2
No. of positive dust samplesNo. of farms with:
M. avium subsp. paratuberculosis DNA
Viable M. avium subsp. paratuberculosis
Control (n = 2)
Group A (n = 8)
Group B (n = 14)
Control (n = 2)
Group A (n = 8)
Group B (n = 14)
S1S2S1S2S1S2S1S2S1S2S1S2
0224345228586
13446124
243112
31111112
412
Open in a separate windowViable M. avium subsp. paratuberculosis was detected in the young stock housing in 4 and 3 farms of group B with single-barn housing at S1 and S2, respectively. No viable M. avium subsp. paratuberculosis was detected in separate young stock housings.To our knowledge, this study is the first to confirm the presence of M. avium subsp. paratuberculosis DNA as well as viable M. avium subsp. paratuberculosis in settled-dust samples of commercial dairy farms. M. avium subsp. paratuberculosis dispersion by bioaerosols under experimental conditions was already described (6). These findings support the concept of dust-based environmental dispersion of M. avium subsp. paratuberculosis within farms.The relatively small number of farms and the convenience sampling are limitations of this study that could have introduced bias. However, this study is a proof of principle that viable M. avium subsp. paratuberculosis can be detected in settled-dust samples on farms with a low M. avium subsp. paratuberculosis prevalence. The environmental method also seems specific for M. avium subsp. paratuberculosis, since no M. avium subsp. paratuberculosis could be detected in any samples of known M. avium subsp. paratuberculosis-free herds.Paratuberculosis control measures aim to prevent fecal-oral contact between infectious shedding adults and susceptible calves as the main transmission route of M. avium subsp. paratuberculosis. Several studies showed that “calf hygiene improvement” decreased prevalence but did not eliminate the disease (1, 7, 14), suggesting the existence of other transmission routes. In utero transmission, transmission via milk, and calf-to-calf transmission have been described previously (1, 12, 13). Additionally, infection via ingestion and/or inhalation of bioaerosols may be possible (4, 8).Twenty-three of 24 herds were housed in free stalls with one tie-stall herd. Most farmers (n = 15) separated young stock from adult cattle as standard procedure. However, six of these farmers cohoused dry cows in the young stock housing occasionally, indicating the difficulties of consequently implementing management advice. Three farmers did not raise young stock on their farms. In almost all barns, cow brushes were present, as they were recommended to enhance cow well-being in group housings (5), but at the same time they contribute to aerosolization of dust. Animal movement on slatted floors also contributes to dust formation, especially during the winter housing period.Most farmers from group A farms, compared to only a few from group B farms, intended to clean their barns yearly, but only 50% met this aim. Young stock housings were never totally empty, but high-pressure cleaning was occasionally performed at 6/8 farms of group A and at 1 of group B. The numbers of farms in this study precluded statistical testing, but the difference in cleaning attitude seemed remarkable.Comparison of the two methods of dust collection showed no statistical difference. No M. avium subsp. paratuberculosis, neither DNA nor viable M. avium subsp. paratuberculosis, could be detected on known negative farms, whereas on farms of groups A and B, M. avium subsp. paratuberculosis DNA was present in comparable numbers of locations. Viable M. avium subsp. paratuberculosis was present only in group B farms at S1 and in both group A and B farms at S2. It seems that M. avium subsp. paratuberculosis can survive in dust for some time. Besides having a possible role in M. avium subsp. paratuberculosis transmission, dust might also be a useful predictor of M. avium subsp. paratuberculosis presence or M. avium subsp. paratuberculosis introduction on dairy farms, even on farms with low M. avium subsp. paratuberculosis prevalence.In conclusion, this study showed that dust on farms with a low M. avium subsp. paratuberculosis seroprevalence contained viable M. avium subsp. paratuberculosis, which indicated a role in M. avium subsp. paratuberculosis transmission. Further research is needed to study if and how infection with M. avium subsp. paratuberculosis-contaminated dust is possible. Additionally, dust sampling may be an alternative tool to monitor M. avium subsp. paratuberculosis status in control programs.  相似文献   

8.
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   

9.
10.
11.
The effects of the challenge dose and major histocompatibility complex (MHC) class IB alleles were analyzed in 112 Mauritian cynomolgus monkeys vaccinated (n = 67) or not vaccinated (n = 45) with Tat and challenged with simian/human immunodeficiency virus (SHIV) 89.6Pcy243. In the controls, the challenge dose (10 to 20 50% monkey infectious doses [MID50]) or MHC did not affect susceptibility to infection, peak viral load, or acute CD4 T-cell loss, whereas in the chronic phase of infection, the H1 haplotype correlated with a high viral load (P = 0.0280) and CD4 loss (P = 0.0343). Vaccination reduced the rate of infection acquisition at 10 MID50 (P < 0.0001), and contained acute CD4 loss at 15 MID50 (P = 0.0099). Haplotypes H2 and H6 were correlated with increased susceptibility (P = 0.0199) and resistance (P = 0.0087) to infection, respectively. Vaccination also contained CD4 depletion (P = 0.0391) during chronic infection, independently of the challenge dose or haplotype.Advances in typing of the major histocompatibility complex (MHC) of Mauritian cynomolgus macaques (14, 20, 26) have provided the opportunity to address the influence of host factors on vaccine studies (13). Retrospective analysis of 22 macaques vaccinated with Tat or a Tat-expressing adenoviral vector revealed that monkeys with the H6 or H3 MHC class IB haplotype were overrepresented among aviremic or controller animals, whereas macaques with the H2 or H5 haplotype clustered in the noncontrollers (12). More recently, the H6 haplotype was reported to correlate with control of chronic infection with simian immunodeficiency virus (SIV) mac251, regardless of vaccination (18).Here, we performed a retrospective analysis of 112 Mauritian cynomolgus macaques, which included the 22 animals studied previously (12), to evaluate the impact of the challenge dose and class IB haplotype on the acquisition and severity of simian/human immunodeficiency virus (SHIV) 89.6Pcy243 infection in 45 control monkeys and 67 monkeys vaccinated with Tat from different protocols (Table (Table11).

TABLE 1.

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

12.
Florfenicol resistance was analyzed in 230 enteric pig isolates collected between 1998 and 2006. PCR, plasmid profiling, Southern blot hybridization, and a mixed-broth conjugation assay suggested the intra- and interspecies plasmid-mediated transfer of florfenicol resistance among the isolates that exhibited MICs for florfenicol between 4 to 128 mg/liter.Florfenicol, a fluorinated chloramphenicol derivative, is a broad-spectrum antimicrobial agent active against a wide range of both gram-positive and gram-negative bacteria. In South Korea, it was initially approved for the treatment of bovine and porcine respiratory disease in 1999 (10, 16). Recent reports have shown increasing use of florfenicol in the treatment of target respiratory pathogens as well as in Escherichia coli infection in bovine, porcine, and poultry production (4, 6, 11, 18). This may have lead to the emergence and spread of florfenicol resistance in a wide range of gram-negative and gram-positive bacteria (5, 8, 15, 17, 18). Florfenicol resistance in enteric microbes, such as E. coli, Klebsiella pneumoniae, and Salmonella spp., is of special concern because they share a common gut environment and are liable to transferring the resistance genes through mobile genetic structures and plasmids bearing antibiotic-resistant determinants (1, 2, 17). Florfenicol resistance is mediated by the floR gene in gram-negative bacteria, and chromosomal location of this gene, especially in the pentadrug-resistant gene cluster of Salmonella enterica serovar Typhimurium phage type (PT) DT104, has attracted wide interest in light of its contribution to multidrug resistance and the development of DT104 detection methods using PCR (1, 7, 9).Though recent reports have shown increased consumption of florfenicol in farms, no reports are available on the prevalence of florfenicol resistance among microbes of enteric origin from South Korean farms (10, 12, 19). Past work on florfenicol resistance has reported a wide range of MICs for florfenicol for E. coli animal isolates, with and without the floR gene (12, 14, 17, 18). The floR gene, however, has been identified in most E. coli isolates with MICs for florfenicol of >8 mg/liter (6, 15, 18). However, some strains with MICs between 8 mg/liter and 16 mg/liter have been reported to have other mechanisms of reduced susceptibility to florfenicol (15, 17). Currently, CLSI breakpoints are approved to indicate florfenicol resistance only for bovine and porcine respiratory disease pathogens (Pasteurella multocida, Mannheimia haemolytica, and Histophilus somni); no approved CLSI breakpoint is currently available for enteric bacteria except for S. enterica serovar Cholerasuis.In light of this, our study focused on an analysis of phenotypic and genotypic florfenicol resistance in enterobacteria isolated from the samples of clinically sick animals (pigs) between the years 1998 and 2006. We further tested whether or not this resistance was plasmid mediated. Conjugation experiments were performed to evaluate the ease with which resistance-associated plasmids would move across species or genera of enterobacteria.E. coli (n = 121), S. enterica serovar Typhimurium (n = 71), S. enterica serovar Enteritidis (n = 12), and K. pneumoniae (n = 26) strains were obtained from the feces, intestines, lungs, and lymph nodes of pigs with mixed clinical signs of digestive and respiratory disorders after necropsy. Identification of the suspected colonies, isolated from selective media incubated overnight with necropsied samples, was made by Vitek (Vitek system; bioMérieux, Marcy l''Etoile, France). All Salmonella spp. were serotyped by slide agglutination and tube agglutination with Salmonella O and H group antisera, respectively (Difco Co., Franklin Lakes, NJ), at the National Veterinary Research and Quarantine Service (NVRQS; Anyang, South Korea). Phage typing was performed for the isolates showing ACSSuTF resistance at NVRQS in accordance with the guidelines provided by the Public Health Laboratory Service (PHLS), London, United Kingdom.MICs of antimicrobial agents were determined by the microbroth dilution method, using microtiter plates that contained florfenicol concentrations of 0.5 to 256 mg/liter in serial twofold dilutions (16). Evaluation of the MIC was performed according to the recommendations of the CLSI (13). The MIC was considered to correspond to the first dilution at which no growth was detectable. E. coli ATCC 25922, Pseudomonas aeruginosa ATCC 27853, and Salmonella serovar Typhimurium DT104 ATCC 2501 were used for quality control.PCR was performed using genomic DNA (Wizard genomic DNA purification kit; Promega, Madison, WI) of all the isolates with MICs for florfenicol of ≥4 mg/liter (see Table Table2)2) as a template and a set of floR-specific oligonucleotide primers, Flo-F (5′-CTGATCGCTCCTTTCGACAT-3′) and Flo-R (5′-CCGTGGCGTAACAAATCAC-3′) (GenBank accession no. DQ647028.2). Amplified PCR product with the expected size of 1,083 bp from one of the isolates was cloned in the pQE-UA 30 vector system (Qiagen) using the manufacturer''s protocol. The identity of floR gene was confirmed by sequencing. For all other PCR-positive isolates, identity of the floR gene was confirmed by the size and Southern blot hybridization result of the PCR products with the biotin-labeled floR probe as described below.

TABLE 2.

Susceptibilities of antimicrobial agents for the donor and the transconjugant strains
StrainMIC (mg/liter)a
KANTESuMSTRFCCRIF
E. coli 03/16>256>256>256128162
Transconjugant12812812812816>256
E. coli 04/18>256>256>256>256324
Transconjugant>256>256>2566416>256
Salmonella serovar Enteritidis 03/22128128>256>25632<0.5
Transconjugant12864>256>25616128
K. pneumoniae 04/22>25664>256>25632<0.5
Transconjugant12832128>25616128
Open in a separate windowaKAN, kanamycin; TE, tetracycline; SuM, sulfamethoxazole; STR, streptomycin; FCC, florfenicol; RIF, rifampin.Mixed-broth culture mating was performed to observe the transferability of the florfenicol resistance gene, as described by Kang et al. (7). All isolates with the floR gene (see Table Table2)2) were included as putative donors in a conjugation experiment. E. coli RG488 Rifr, kindly provided by Je Chul Lee (Kyungpook National University), was used as the recipient to detect the transfer of resistance gene. The transconjugants were selected on MacConkey agar supplemented with florfenicol (2 mg/liter) and rifampin (rifampicin) (100 mg/liter). Transfer frequency was calculated as the number of transconjugants per recipient.Transfers of resistance to florfenicol, tetracycline, streptomycin, kanamycin, and sulfamethoxazole in transconjugants were confirmed by MIC, following the procedures described above. Presence of the florfenicol resistance gene, floR, in the transconjugants was confirmed by PCR and DNA hybridizations.Single colonies of the transconjugants (E. coli RG488 Rifr) of E. coli 03/16, E. coli 04/18, Salmonella serovar Enteritidis 03/22, and K. pneumoniae 04/22 and single colonies of isolates that failed to transfer the florfenicol resistance (E. coli 04/1, Salmonella serovar Typhimurium 04/13, and Salmonella serovar Enteritidis 04/8) in broth conjugation assay were picked from the MacConkey agar plates with and without antibiotics. Plasmid DNA was extracted using the midi extraction kit (Qiagen, Valencia, CA) following the manufacturer''s protocol and was digested by the EcoRI restriction enzyme. The digested plasmids and PCR-amplified products for all the floR-positive strains (see Table Table2)2) were electrophoresed through agarose gels and transferred to a positively charged nylon membrane (GE Healthcare, Little Chalfont, England). Hybridization experiment was performed by using a psoralen-biotin (BrightStar psoralen-biotin nonisotopic labeling kit; Ambion Inc, Austin, TX)-labeled PCR product of the floR gene. Detection was performed using a BrightStar BioDetect nonisotopic detection kit (Ambion Inc., Austin, TX) following the manufacturer''s protocol.Reports on the uses of antimicrobial agents have shown that the use of florfenicol gradually increased from 387 kg/year in 2001 to 17,159 kg in 2005 (10). More than half of this increase was used in the pig industry alone, followed by poultry and bovine farms (10). This figure could increase for subsequent years because the in vitro antimicrobial activity of florfenicol against respiratory pathogens is very effective, and no report of resistance in target pathogens is available from South Korea (16). In this study, 7.43% (9/121) of E. coli, 8.45% (6/71) of Salmonella serovar Typhimurium, 16.6% (2/12) of Salmonella serovar Enteritidis, and 7.69% (2/26) of K. pneumoniae strains exhibited MICs for florfenicol that ranged from 4 to 128 mg/liter. The floR gene was amplified by 14 out of 19 isolates that exhibited MIC for florfenicol of ≥4 mg/liter (Fig. (Fig.1).1). Two E. coli (MIC, 4 and 8 mg/liter) and three Salmonella serovar Typhimurium (MIC, 16 mg/liter each) isolates (Table (Table1)1) did not amplify the floR gene. The reduced susceptibility of these isolates to florfenicol might be due to the involvement of other mechanisms or resistance genes (15, 17).Open in a separate windowFIG. 1.PCR amplification of the floR gene (upper panel). Lane M, 100-bp DNA marker (iNtRON Biotechnology, South Korea); lanes 1 to 4, E. coli (03/16; MIC, 16 mg/liter), E. coli (04/18; MIC, 32 mg/liter), Salmonella serovar Enteritidis (03/22; MIC, 32 mg/liter), and K. pneumoniae (04/22; MIC, 32 mg/liter); lane N, negative control; lanes 5 to 10, E. coli (04/1; MIC, 16 mg/liter), Salmonella serovar Typhimurium (04/13; MIC, 16 mg/liter), Salmonella serovar Enteritidis (04/8; MIC, 32 mg/liter), E. coli (03/17; MIC, 16 mg/liter), E. coli (04/19; MIC, 32 mg/liter), and K. pneumoniae (04/21; MIC, 64 mg/liter). Identity confirmation of floR gene in these isolates was performed based on the size and Southern blot hybridization of PCR products (lower panel).

TABLE 1.

Susceptibilities, amplification, and hybridization patterns of floR in E. coli, Salmonella serovar Typhimurium, Salmonella serovar Enteritidis, and K. pneumoniae isolates
StrainscYr/no.FFC MIC (mg/liter)PCR resultaTFFCC/RIF MIC (TCs) (mg/liter)Southern blot hybridizationb
E. coli04/116+
04/332+
05/932+
05/134
03/148
03/1616+1.3 × 10−516/>256+
03/1716+
04/1832+1.4 × 10−516/>256+
04/1932+
ST (PT193)04/1316
04/2616
04/1516
ST (U302)05/24128+
ST (PT120)05/21128+
ST (DT104)03/4132+
SE04/832+
03/2232+1.3 × 10−416/128+
KP04/2164+
04/2232+1.2 × 10−416/128+
E. coli (RG488 Rifr)<0.5
Open in a separate windowaPCR amplification of floR gene, using whole-cell genomic DNA.bHybridization signal was in the plasmid DNA of the transconjugants (E. coli RG4888 Rif) of E. coli 03/16 (MIC of 16 mg/liter), E. coli 04/18 (MIC of 16 mg/liter), Salmonella serovar Enteritidis 03/22 (MIC of 16 mg/liter), and K. pneumoniae 04/22 (MIC of 16 mg/liter).cST, Salmonella serovar Typhimurium; SE, Salmonella serovar Enteritidis; KP, Klebsiella pneumoniae; FCC, florfenicol; RIF, rifampin; TF, transfer frequency; TCs, transconjugant.Past work on florfenicol resistance in E. coli animal isolates has reported various ranges of MICs. North American and European E. coli strains carrying the floR gene from animal origin have been reported to have MICs for florfenicol of 16 to ≥256 mg/liter and of ≥128 mg/liter, respectively (15). However, MIC values for florfenicol of isolates from South Korea (this study) and E. coli isolates from China were substantially less than those described earlier (6, 11, 15). This indicates that the MIC for florfenicol could vary geographically, so a comparative study of the florfenicol resistance gene from different geographical regions might help elucidate its mechanism.Conjugation assay showed the transfer of reduced florfenicol susceptibility by 4 out of 14 strains. These transconjugants that harbored the plasmid of 23 kb were also resistant to tetracycline, streptomycin, kanamycin, and sulfamethoxazole. Results of the MIC determination of these antibiotics for donor strains (E. coli 03/16, E. coli 04/18, Salmonella serovar Enteritidis 03/22, and K. pneumoniae 04/22) and their transconjugants are listed in Table Table2.2. Out of those 14 strains, 10 isolates amplified the floR gene in PCR and failed to transfer floR gene in conjugation experiments. From those 10 isolates, plasmid extraction was carried out for three randomly selected E. coli (04/1; MIC, 16 mg/liter), Salmonella serovar Typhimurium (04/13; MIC, 16 mg/liter), and Salmonella serovar Enteritidis (04/8; MIC, 32 mg/liter) strains. None of these three strains showed floR-specific probe hybridization in the plasmid profile and Southern blot (Fig. (Fig.2,2, lanes 1 to 3) indicating the chromosomal location of the floR gene in these isolates. The floR-specific probe hybridization of plasmid DNA of the transconjugants (E. coli RG488 Rifr) of E. coli 03/16, E. coli 04/18, Salmonella serovar Enteritidis 03/22, and K. pneumoniae 04/22 indicated the presence of both single (Fig. (Fig.2,2, left panel, lanes 4, 6, and 7) and multiple copies (Fig. (Fig.2,2, left panel, lane 5) of this gene. Likewise, the hybridization signal in the different digested fragments of plasmid DNA indicated the different orientations of the floR gene in the conjugative plasmid of these strains. The floR gene has also been extensively described in Salmonella serovar Typhimurium epidemic strain DT104. However, only one such PT was identified among the Salmonella serovar Typhimurium (MIC, 32 mg/liter) isolates in our study. Two other floR-positive Salmonella serovar Typhimurium isolates were identified as PT 302 and 120. Three Salmonella serovar Typhimurium isolates, identified as PT 193 (MIC, 16 mg/liter), did not amplify the floR gene (Table (Table1).1). None of these Salmonella serovar Typhimurium PTs transferred florfenicol resistance in the broth conjugation experiments.Open in a separate windowFIG. 2.Southern blot hybridization (left) and EcoRI restriction (right) profiles of plasmids extracted from E. coli, Salmonella serovar Typhimurium, Salmonella serovar Enteritidis, and K. pneumoniae isolates with different MICs. Lanes M, 1-kb DNA marker; lanes 1 to 3, E. coli (04/1; MIC, 16 mg/liter), Salmonella serovar Typhimurium (04/13; MIC, 16 mg/liter), and Salmonella serovar Enteritidis (04/8; MIC, 32 mg/liter) that showed no transfer of florfenicol resistance in broth conjugation experiments; lanes 4 to 7, transconjugants (E. coli RG4888 Rif) of E. coli (03/16; MIC, 16 mg/liter), E. coli (04/18; MIC, 16 mg/liter), Salmonella serovar Enteritidis (03/22; MIC, 16 mg/liter), and K. pneumoniae (04/22; MIC, 16 mg/liter). The hybridization signal in the different digested fragments of plasmid DNA indicated the different orientations of the floR gene in the conjugative plasmid of these strains. Likewise, the hybridization signal in the plasmid DNA of the transconjugant (E. coli RG4888 Rif) of E. coli (04/18; MIC, 16 mg/liter) (lane 4) indicated the multiple copies of this gene.Our findings are in agreement with those of previous reports on florfenicol resistance in E. coli and support the clinical relevance of an MIC breakpoint of 32 mg/liter in E. coli and other enteric bacteria as proposed by Singer et al. (17). Likewise, plasmid profiling, Southern blotting, phage typing, and conjugation experiment results indicated that emergence and dissemination of floR genes among the Enterobacteriaceae are not due to the prevalence of DT104 or other closely related PTs (4). Indeed, it could be inferred that floR-bearing promiscuous plasmids in these groups of enteric microbes are in circulation due to the local selection pressure imposed by the use of antimicrobial agents on farms (7, 10, 12).  相似文献   

13.
The presence of Campylobacter spp. was investigated in 41 Antarctic fur seals (Arctocephalus gazella) and 9 Weddell seals (Leptonychotes weddellii) at Deception Island, Antarctica. Infections were encountered in six Antarctic fur seals. The isolates, the first reported from marine mammals in the Antarctic region, were identified as Campylobacter insulaenigrae and Campylobacter lari.The Antarctic and sub-Antarctic regions are often regarded as pristine landscapes, unaffected by human activity. A limited number of surveys have been carried out to investigate the possible occurrence of zoonotic enteropathogens and if certain bacteria could be used as tools for evaluating biological pollution in this area (4, 11). In the case of Campylobacter species, there have been only three reports in the literature, but in all of them Campylobacter was isolated from marine seabirds but not from marine mammals. Campylobacter jejuni was isolated in Antarctic and sub-Antarctic areas from Macaroni penguins (Eudyptes chrysolophus) (4), and Campylobacter lari was isolated from Brown skuas, South Polar skuas, and Adelie penguins (2, 11).Reports of Campylobacter species isolated from marine mammals are rare. Campylobacter insulaenigrae was isolated from three harbor seals (Phoca vitulina) and a harbor porpoise (Phocoena phocoena) in Scotland (7). The isolation of C. jejuni, C. lari, and an unknown Campylobacter species from juvenile northern elephant seals (Mirounga angustirostris) in California was also reported (22). Finally, 71 isolates of C. insulaenigrae and 1 isolate similar to but distinct from both Campylobacter upsaliensis and Campylobacter helveticus were isolated from northern elephant seals in California (23). In the South Georgia Archipelago, fecal swabs were taken from 206 Antarctic fur seal pups, but no isolates could be obtained (4). In this study, we successfully isolated C. lari from 7.3% of Antarctic fur seals (Arctocephalus gazella) sampled and C. insulaenigrae from a further 7.3%. On the other hand, Campylobacter was not detected in the nine Weddell seals (Leptonychotes weddellii) sampled. To our knowledge, this is the first report on the isolation of C. lari and C. insulaenigrae from marine mammals in the Antarctic region.Fieldwork was conducted at Deception Island (latitude of 62°58′S and longitude of 60°40′W), in the South Shetland Islands. During January to February 2007, Antarctic fur seals (Arctocephalus gazella) and Weddell seals (Leptonychotes weddellii) were captured and fecal samples were collected by insertion of sterile cotton wool swabs into the rectum of the marine mammals. A total of 41 Antarctic fur seals and 9 Weddell seals were sampled. The distribution by ages was of 7 adults (over 4 years of age with breeding activity), 19 subadults (2 to 4 years of age), and 15 juvenile Antarctic fur seals (less than 2 years of age), and 8 adult Weddell seals and 1 juvenile. All animals presented a good body condition and showed no symptoms at the time of sampling.Three swabs were taken from each animal and were placed in FBP medium (8) with 0.5% active charcoal (Sigma Ltd.), Amies transport medium with charcoal, and Cary Blair transport medium, respectively. All samples were kept at +4 to 8°C until culture in the lab. The number of days between sampling and cultivation varied from 96 to 124 days, with a median value of 105 days.Each swab was placed in 10 ml of Campylobacter enrichment broth (Lab M) with 5% laked horse blood and CAT supplement (cefoperazone [8 μg/ml], teicoplanin [4 μg/ml], and amphotericin B [10 μg/ml]) at 37°C. The broth was incubated at 37°C for 48 h and 5 days in 3.5-liter anaerobic containers using CampyGen sachets (Oxoid), before an aliquot of 100 μl was plated onto CAT agar and the plates were incubated at 37°C for 72 h in a microaerobic atmosphere. In addition, a 47-mm-diameter cellulose membrane with 0.60-μm pores was placed on the surface of an anaerobe agar base (Oxoid) with 5% laked horse blood. Eight to 10 drops of enrichment broth (200 μl) were placed onto the surface of the membrane. The membrane was left for 20 to 30 min on the agar surface at room temperature until all of the fluid had passed through (20). The plates were incubated as described above, but for 5 days to isolate the less common, slower growing species.Isolates were examined by dark-field microscopy to determine morphology and motility and tested to determine whether oxidase was produced. For each sample, five isolates from each of the solid media that had typical morphology and motility and for which the oxidase test was positive were frozen at −80°C in FBP medium (8) until they were tested by phenotypic and genotypic methods.Original Campylobacter identification was done by Gram staining, catalase activity, hippurate hydrolysis, ability to hydrolyze indoxyl acetate, urease activity, H2S production on triple-sugar iron slants, growth at 25°C and 42°C in a microaerophilc environment, growth at 37°C in an aerobic atmosphere, and agglutination with Microscreen latex (Microgen, Camberley, United Kingdom).No differences between the strains were observed in any of the phenotypic tests used. All isolates showed a Gram-negative, slender, curved, seagull wing-like morphology under light microscopy and positive reactions in the catalase test. They were negative for hippurate and indoxyl acetate hydrolysis and urease and did not show H2S production. In addition, they grew at 42°C but did not grow at 25°C or 37°C in an aerobic atmosphere. Finally, all of them were positive in the agglutination test.Because phenotypic results commonly lead to misidentification of Campylobacter species, it is recommended that a molecular method be included in the identification scheme for Campylobacter (5, 15). Identification of the isolates was performed using 16S rRNA gene PCR and sequence analysis (15, 21). Forward and reverse conserved 16S rRNA eubacterial primers 8F (5′-AGAGTTTGATCCTGGCTCAG-3′) and 1492R (5′-GGTTACCTTGTTACGACTT-3′) were used to amplify the 16S rRNA according to the protocol described by Jang et al. (9). Forward and reverse sequencing reactions were performed by the Laboratorio Central de Veterinaria''s DNA sequencing facility (LCV Algete, Madrid, Spain). Three strains were identified as C. lari and the other three as C. insulaenigrae based on both forward and reverse sequence analysis.Molecular characterization of strains was carried out using a combination of pulsed-field gel electrophoresis (PFGE) using KpnI enzyme and multilocus sequence typing (MLST). Preparation of intact Campylobacter DNA for PFGE was performed following the Pulsenet protocol (17, 24). PFGE for the restriction enzyme KpnI (Takara, Conda, Spain) was performed following the protocol described by Ribot et al. (17). DNA fragments were resolved on 0.9% Seakem Gold agarose gels (Iberlabo, Spain) with a Bio-Rad CHEF DRIII system (Bio-Rad, Spain) at 14°C and 6 V/cm. Electrophoresis was carried out for 22 h with pulse times ramping from 4 s to 20 s. The fingerprinting experiments were analyzed using the InfoQuest FP software (Bio-Rad, Spain), and the dendrogram was constructed using the unweighted-pair group method using average linkages (UPGMA).MLST of C. lari strains was performed as described by Miller et al. (13). In the case of C. insulaenigrae strains, MLST was performed following the protocol described by Stoddard et al. (23). All amplicons were sequenced by the Sequencing Service of the Instituto de Salud Carlos III (Madrid, Spain). Sequence data were collated, and alleles were assigned using the Campylobacter PubMLST database (http://pubmlst.org/campylobacter/). Novel alleles and sequence types were submitted for allele and sequence type (ST) designations when appropriate.Regarding the age distribution of animals, C. lari was isolated from 1 of 7 adult (14.3%), 1 of 19 subadult (5.3%), and 1 of 15 juvenile (6.6%) Antarctic fur seals. C. insulaenigrae was isolated from 1 of 7 adults (14.3%) and 2 of 19 of subadults (10.5%) but not from juvenile animals (Table (Table1).1). All strains were obtained from the swabs kept in FBP transport medium.

TABLE 1.

Source of Campylobacter isolates
AnimalAge category and sexDate (mo/day/yr) of:
Campylobacter sp. and isolate no.
SamplingCulture
L 06/56Adult male2/15/075/30/07C. insulaenigrae FR-07
L 06/78Subadult male2/15/075/30/07C. insulaenigrae FR-15
L 06/102Subadult male2/22/075/30/07C. lari FR-28
L 06/134Juvenile male2/21/075/30/07C. lari FR-36
L 06/146Subadult male2/22/075/30/07C. insulaenigrae FR-38
L 06/48Adult male2/22/075/30/07C. lari FR-48
Open in a separate windowCampylobacter is very sensitive to excessive amounts of oxygen and has little capacity to survive in the environment. It is therefore possible that the prevalence of Campylobacter species in Antarctic fur seals is greater than that obtained in our survey and that we have isolated more-resistant strains with a larger ability to survive a prolonged transport. Nevertheless, we think that the freezing medium described by Gorman and Adley (8) modified by the addition of 0.5% of activated charcoal is a very good transport medium since the bacteria remained viable for 3 months at refrigeration temperature, whereas they did not survive in the transport media routinely used for the preservation of fecal samples such as Amies and Cary Blair media.PFGE is a useful tool for conducting epidemiological studies of Campylobacter species. We used digestion with KpnI because it has been reported to have greater power of discrimination than digestion with SmaI (16). All isolates showed very different patterns (Fig. (Fig.1),1), indicating different sources of infection and circulation of different clones on Deception Island. These data were confirmed by the results of MLST, in which each strain belonged to a different ST, none of which had been previously reported. We submitted to the MLST database 12 new sequences of alleles tested for C. insulaenigrae and 10 new sequences of C. lari obtained (Table (Table22).Open in a separate windowFIG. 1.UPGMA dendrogram of PFGE profiles.

TABLE 2.

Alelle numbers and sequence types of Campylobacter isolates
Species and isolate no.STAllele no.a
aspA or adkatpAglnAglyApgipgmtkt
C. insulaenigrae
    FR-7412 (aspA)16*12*215*15*11*
    FR-15424 (aspA)1011*12*14*15*12*
    FR-38437 (aspA)17*11*13*14*313*
C. lari
    FR-281752* (adk)57*250*56*51*31*
    FR-361652* (adk)57*2256*52*31*
    FR-481853* (adk)58*1257*52*32*
Open in a separate windowaAsterisks indicate new alleles.The introduction of C. lari in the Antarctic fur seal colonies may have occurred through seabirds. C. lari has been isolated from Adelie penguins (Pygoscelis adeliae), kelp gull (Larus dominicanus), Brown skuas (Stercorarius antarctica lonnbergi), and South Polar skuas (Stercorarius maccormicki) in Hope Bay (11) and in the Antarctic Peninsula (2). Gulls can travel between South America and Antarctica and are potential carriers of enteric pathogens (1). Thus, C. lari has been isolated from kelp gulls in southern Chile (6). Also, South Polar skuas have been reported in Greenland and the Aleutian Islands and Brown skuas move around the Antarctic coast. Therefore, it is possible that these birds acquire infectious organisms when they move to areas with high levels of human activity. These birds have been reported on Deception Island (10), and it is common to find skuas and giant petrels on beaches where Antarctic fur seal colonies rest. The carrier birds could eliminate Campylobacter and pollute these areas. Alternatively, these birds could be occasional prey for Antarctic fur seals.C. insulaenigrae is a new Campylobacter species whose host range might be restricted to marine mammals (23). It could be hypothesized that C. insulaenigrae evolved from C. lari based on the presence of both species in sea lions and their sharing other characteristics such as the absence of the citrate synthase gene (23). In addition, considering that C. insulaenigrae has not been isolated from seabirds or shellfish and the migration ranges of sea lions are generally not very large, Antarctic fur seals could have been initially infected with C. lari, and subsequently this species has evolved, adapting to mammals. Alternatively both species could share an ancestor and have adapted to different hosts.The Antarctic fur seals captured showed no weight loss, diarrhea, or other symptoms at the time of sampling. However, due to the nature of our study, it is not possible to know whether the animal had been ill before the time of collection and was subsequently a carrier. Taking into account previous reports (7, 23) and our results, pinnipeds could possibly act as reservoir of C. insulaenigrae.The presence of Campylobacter in Antarctic fur seals could also be important due to the zoonotic potential of both species (5, 12, 18, 19). Therefore, researchers should continue to exercise caution when working with these animals. In addition, C. lari has been involved in waterborne outbreaks (3) and some reports have identified this species as the most frequently isolated from surface water (25). Most of the Antarctic stations'' catchwater from lakes generated by meltwater and the water treatment cannot be accomplished by chemical products to prevent marine pollution. In general, water is not treated or is treated only by filtration and UV light. Antarctic fur seals can nevertheless pollute the water of these lakes and/or infect other species such as penguins and other birds, which in turn could also act as a source of infection for humans. Furthermore, Obiri-Danso et al. (14) have reported that C. lari survives for longer in surface waters than C. jejuni and Campylobacter coli, so it would have a greater chance of surviving the water treatment. Finally, in case of infection, the therapy may be complicated because in many of the stations there are only basic medical services.In summary, we describe here the first isolation and characterization of two species of Campylobacter, C. lari and C. insulaenigrae, from Antarctic fur seals. Further studies are needed to determine the prevalence of Campylobacter spp. in Antarctic pinnipeds, the possible sources of infection and if the presence of Campylobacter in marine mammals could be a risk for human illness or could be a result of microbial pollution associated with human activity.  相似文献   

14.
Twenty-one salts were tested for their effects on the growth of Pectobacterium carotovorum subsp. carotovorum and Pectobacterium atrosepticum. In liquid medium, 11 salts (0.2 M) exhibited strong inhibition of bacterial growth. The inhibitory action of salts relates to the water-ionizing capacity and the lipophilicity of their constituent ions.Different biochemical mechanisms have been put forth to explain the antimicrobial activity of organic and inorganic salts, including inhibition of several steps of the energy metabolism (benzoate, bicarbonate, propionate, sorbate, and sulfite salts) (2, 3, 11, 16, 17, 19, 25) and complexation to DNA and RNA (aluminum and sulfites) (12, 13, 15, 20, 27, 28). However, little is known about the physicochemical basis for the general antimicrobial action of salts. The objective of this work was to gain an understanding of the relationship between the inhibitory action of salts on bacterial growth and their physicochemical properties by using the bacteria Pectobacterium carotovorum subsp. carotovorum (formerly Erwinia carotovora subsp. carotovora) and Pectobacterium atrosepticum (formerly Erwinia carotovora subsp. atroseptica). These bacteria are responsible for soft rot, a disease of economic importance affecting numerous stored vegetable crops (14, 22).Pectobacterium carotovorum subsp. carotovorum (strain Ecc 1367) and P. atrosepticum (strain Eca 709), provided by the Laboratoire de Diagnostic en Phytoprotection (MAPAQ, Québec, Canada), were grown in 250-ml flasks containing 50 ml of 20% tryptic soy broth (Difco Laboratories, Becton Dickinson, Sparks, MD) amended with salts (200 mM) or unamended (control), by incubation at 24°C with agitation (150 rpm; Lab-Line Instruments Inc., Melrose Park, IL) for 24 h. The pHs of the media were not adjusted but varied with the type of salts, unless stated otherwise. Flasks were inoculated with 100 μl of each bacterial suspension (1 × 107 CFU/ml). Bacterial growth was determined by turbidimetry at 600 nm with a UV/visible spectrophotometer (Ultrospec 2000; Pharmacia Biotech Ltd, Cambridge, United Kingdom), using appropriate blanks. Results were expressed as the percentage of growth inhibition compared with the growth of the control. A completely randomized experimental design with three replicates was used, the experimental unit being a flask. Analysis of variance was carried out with the GLM (general linear model) procedure of SAS (SAS Institute, Cary, NC) software. When they were significant (P < 0.05), treatment means were compared using Fisher''s protected least-significant-difference test.Among the 21 salts tested, sodium carbonate, sodium metabisulfite, trisodium phosphate, aluminum lactate, aluminum chloride, sodium bicarbonate, sodium propionate, ammonium acetate, aluminum dihydroxy acetate, potassium sorbate, and sodium benzoate exhibited strong inhibition (≥97%) of the growth of both P. carotovorum subsp. carotovorum and P. atrosepticum (Table (Table1).1). Calcium chloride, sodium formate, sodium acetate, ammonium hydrogen phosphate, and sodium hydrogen phosphate exhibited a moderately inhibitory effect; sodium lactate and tartrate had no effect. On the other hand, ammonium chloride, potassium chloride, and sodium chloride stimulated the growth of P. atrosepticum.

TABLE 1.

Effect of salts on the growth of P. atrosepticum and P. carotovorum subsp. carotovorum
Salt (0.2 M)apHbOsmotic pressure (atm)cGrowth inhibition (%)d
P. atrosepticumP. carotovorum subsp. carotovorum
Aluminum dihydroxy acetate [Al(OH)2C2H3O2]4.99.79100 a100 a
Aluminum chloride (AlCl3·6H2O)2.519.57100 a100 a
Aluminum lactate [Al(C3H5O3)3]3.419.57100 a100 a
Ammonium acetate (NH4C2H3O2)7.29.79100 a100 a
Ammonium chloride (NH4Cl)7.09.79−18 dND
Ammonium hydrogen phosphate [(NH4)2HPO4]8.314.6843 b23 c
Calcium chloride (CaCl2·2H2O)5.814.6885 a70 b
Potassium chloride (KCl)7.39.79−27 dND
Potassium sorbate (KC6H7O2)7.79.79100 a97 a
Sodium acetate (NaC2H3O2·3H2O)7.49.7963 bND
Sodium benzoate (NaC7H5O2)7.49.79100 a100 a
Sodium bicarbonate (NaHCO3)8.19.79100 a100 a
Sodium carbonate (Na2CO3)10.614.68100 a100 a
Sodium chloride (NaCl)7.29.79−29 dND
Sodium formate (NaCHO2)7.39.7924 cND
Sodium lactate (C3H5O3Na)7.39.793 cND
Sodium metabisulfite (Na2S2O5)4.519.57100 a100 a
Sodium hydrogen phosphate (Na2HPO4)8.714.6869 b61 b
Sodium propionate (NaC3H5O2)7.49.79100 a99 a
Sodium tartrate (Na2C4H4O6·2H2O)7.314.682 cND
Trisodium phosphate (Na3PO4·12H2O)11.919.57100 a100 a
Open in a separate windowaSalts were purchased from Sigma Chemical Co. (St. Louis, MO), except for ammonium acetate (BDH Inc., Toronto, Canada), sodium chloride (BDH), sodium bicarbonate (BDH), and aluminum lactate (Aldrich Chemical, Milwaukee, WI).bpH of the medium amended with each salt.cOsmotic pressure of the salt solution was calculated using van’t Hoff''s equation, Π = iRTc, where R is the gas constant, T is the absolute temperature (K), c is the concentration of the salt (mol/liter), and i is the number of ions into which the salt dissociates in solution.dPercentage of growth inhibition compared to growth of the control. Each value represents the mean of three replicates. Values in the same column followed by the same letter are not significantly different according to Fisher''s protected least-significant-difference test (P > 0.05). ND, not determined. Negative values signify bacterial growth stimulation.Several factors in the salt solutions can contribute to bacterial growth inhibition. Elevated osmolarity due to salt addition may trigger the osmoregulatory process, causing an increased maintenance metabolism and leading to reduction in bacterial growth. Thus, we calculated the osmotic pressure (Π) of salt solutions using van''t Hoff''s equation (26). As shown in Table Table1,1, salts with comparable osmolarities displayed complete or no bacterial growth inhibition, indicating that osmotic stress or reduction in water activity alone may not have brought about the inhibition of the bacterial growth. Therefore, other factors may play a role.The acidity or alkalinity of the medium resulting from the addition of some of the salts can have profoundly adverse effects on bacterial growth. Extreme pH conditions can lead to denaturation of proteins like enzymes present on the cell surface, depolarization of transport for essential ions and nutrients, modification of cytoplasmic pH, and DNA damage (12, 18). Table Table11 shows that the addition of aluminum lactate, aluminum chloride, and sodium metabisulfite, whose ΔpHs (ΔpH = |7.5 [the optimal pH for growth] − the pH of the salt-amended medium|) are ≥3, strongly acidified the medium, whereas the addition of sodium carbonate and trisodium phosphate strongly increased the pH (ΔpH ≥ 3.1). Except for ammonium acetate, sodium acetate, sodium bicarbonate, and the preservative salts (potassium sorbate, sodium benzoate, and sodium propionate), whose ΔpHs are <1, all the other salts generally display inhibitory effects when ΔpH values are ≥1 (Fig. (Fig.1).1). Based on this result, the effect of the highly acidic or alkaline salts (which strongly affected the pH of the medium) on the growth of P. atrosepticum was evaluated at pH 7.5. Sodium carbonate and sodium metabisulfite completely inhibited bacterial growth at pH 7.5, as they did at pHs 10.6 and 4.5, respectively; trisodium phosphate (pH 11.9) exhibited a slightly lower inhibitory effect (growth inhibition of 83.2%) at pH 7.5. These observations suggest that growth inhibition by sodium carbonate, sodium metabisulfite, and trisodium phosphate cannot be attributed solely to extreme pH and passive proton transfer (extreme pH) across the bacterial membrane. Since aluminum salts precipitate at pH 7.5 (due to formation of hydrated aluminum hydroxide), it was not possible to test their inhibitory effect at pH 7.5.Open in a separate windowFIG. 1.Relationship between ΔpH (|7.5 [the optimal pH for growth] − the pH of the salt-amended medium|) and growth inhibition of Pectobacterium atrosepticum. 1, Sodium chloride; 2, potassium chloride; 3, ammonium chloride; 4, sodium tartrate; 5, sodium lactate; 6, sodium formate; 7, ammonium hydrogen phosphate; 8, sodium acetate; 9, sodium hydrogen phosphate; 10, calcium chloride; 11, ammonium acetate; 12, sodium benzoate; 13, sodium propionate; 14, potassium sorbate; 15, sodium bicarbonate; 16, aluminum dihydroxy acetate; 17, sodium metabisulfite; 18, sodium carbonate; 19, aluminum lactate; 20, aluminum chloride; 21, trisodium phosphate.The dissociation of salts in aqueous medium generates ionic species which can participate in proton exchange reactions with water molecules. The capacity of an ion to dissociate water is an intrinsic characteristic, determined by its pK value (pKa for acidic species or pKb for basic ones) (4, 21, 24). For an ionic strength of >0.1 M, pKa and pKb values of the ions are more accurate when they are defined as apparent constants (pK′a or pK′b) in terms of the activities of hydronium and hydroxyl ions, ionic species concentrations and activity coefficients (6). Thus, for the acidic ions, we have the equation ), and for the basic anions, pK′b = pKb + log(γHB/γB), where pK′a and pK′b are the apparent acidity constant and basicity constant, respectively; is the activity coefficient of the conjugate base (B); and γHB is that of the acidic (HB) species. The activity coefficient (γ) of the species i can be expressed as a function of ionic strength (μ), using the Güntelberg approximation of the Debye-Hückel equation (21), as follows: −log γi=[(0.51Zi2 μ1/2)/(1 + μ1/2)], where Zi is the charge on the species i, and μ is the ionic strength. Thus, log(/γHB) = [(0.51μ1/2)/(1 + μ1/2)] (), and log(γHB/) = −[(0.51μ1/2)/(1 + μ1/2)] ().Polytropic acid-potentiating ions (bicarbonate, carbonate, monohydrogen phosphate, phosphate, sulfite, and tartrate) in an aqueous solution can exist as (n + 1) possible species for which the parent acid is HnA. These species may coexist in equilibrium under certain pH conditions. For these ions, pK′a or pK′b were expressed as the means of the coexisting species at a specified pH. Calculated values for pK′a of acidic anions and cations and calculated values for pK′b of basic anions are presented in Table Table2.2. Figure Figure2A2A shows a sigmoidal relationship between the inhibitory effect of salts on bacterial growth and the pK′b value of the basic ions (with a common cation, sodium or potassium, in the salt) and the pK′a value of the acidic ions (with a common anion, chloride, in the salt). The plot exhibits a sharp linear relationship in the pK′ range of 8.0 to 12.0. Below the pK′ value of 8.0, inhibition is maximal, whereas above the pK′ value of 11.0, ions appear to stimulate growth (growth was maximal above the pK′ value of 12). This result demonstrates that the capacity of the constitutive ions of the salts to either donate or subtract protons to water molecules, either in the growth environment (as reflected in the modification of the medium pH) or in the developing cells, generally plays a role in their inhibitory action. The consequent transmembrane pH gradient generated leads to a passive H+ transport across the microbial membrane and to acidification (in the case of ions with low pK′a) or alkalinization (in the case of ions with low pK′b) of the cytoplasm, once the capacity for proton-coupled active transport is outstripped. In both cases, proton exchange with outer membrane proteins will destabilize these proteins, their interaction with membrane lipids, and ultimately, their function in solute transport, leading to growth inhibition. The modification of cytoplasmic pH can also alter nucleic acid structures and functions and contribute to growth inhibition (18).Open in a separate windowFIG. 2.(A) Relationship between the growth inhibition of Pectobacterium atrosepticum and the apparent basicity constant (pK′b,•) of basic anions with common Na+ (or K+) cations in the salt, the apparent acidity constant (pK′a,○) of acidic bisulfite anion (HSO3), and the cations with common Cl ions in the salt. (B) Relationship between the growth inhibition of Pectobacterium atrosepticum and the addition parameter (pK′ + pPo/w) combining the partition coefficient (Po/w) and pK′b (•) of basic anions (common cation, Na+ or K+, in the salt) or pK′a (○) of cations (common anion, Cl, in the salt) and the acidic bisulfite anion (HSO3).

TABLE 2.

Calculated apparent values for acidity, pK′a, and basicity, pK′ba
SaltBasic anion
Cation and acidic anion
pHIonic species or species in equilibriumpK′bpHIonic species or species in equilibriumpK′a
Sodium acetate7.4Acetate9.5
Sodium benzoate7.4Benzoate10.0
Sodium bicarbonate8.1H2CO3/HCO3b7.7
Sodium carbonate10.6HCO3/CO32−6.1
Sodium formate7.3Formate10.4
Sodium hydrogen phosphate8.7H2PO4/HPO42−9.8
Sodium lactate7.3Lactate11.1
Trisodium phosphate11.9HPO42−/PO43−5.3
Sodium propionate7.4Propionate9.3
Potassium sorbate7.7Sorbate9.4
Sodium tartrate7.3Tartrate2−10.6
Sodium chloride7.2Cl17.2
Sodium metabisulfite4.5SO2·H2O/HSO34.0
Aluminum chloride2.5Al3+6.2
Calcium chloride5.8Ca2+13.4
Potassium chloride7.3K+16.2
Sodium chloride7.2Na+15.0
Ammonium chloride7.0NH4+9.5
Open in a separate windowaCalculation of pK′ was performed according to Edsall and Wyman (6). pH values were measured at 0.2 M.bIncludes CO2·H2O and H2CO3.However, the water-ionizing capacity of the constituent ions of the salts and the consequent modification of the pH of the medium are not the sole factors accounting for growth inhibition, as suggested by the exceptional inhibitory actions of benzoate, propionate, and sorbate (Fig. (Fig.11 and and2A).2A). These ions provide a higher inhibition than is expected from their pK′ values (pK′b values of 10.0, 9.3, and 9.4, respectively), while the pH of their solution is optimal for bacterial growth (pHs of 7.4, 7.4, and 7.7, respectively). This suggests that they possess additional characteristics mediating their action, in addition to their water-ionization property. In fact, these preservative agents have been shown to be active either as undissociated acids (like other weak acids) or as anions (7, 8), due to their possibly hydrophobic nature which would allow them to interact with lipid constituents of the cell envelope of gram-negative bacteria such as Pectobacterium spp., and to modify their functionality (5), resulting in growth inhibition. They can also cross the cell envelope due to their lipophilicity, and their acidification inside the cell can cause additional adverse effects.Thus, we determined the octanol/water partition coefficient (Po/w), an indicator of the lipophilic character of a compound, for the effective salts with common sodium (or potassium) or chloride ions. The Po/w coefficients of the salts were determined in duplicate by using the general solvent-solvent separation procedure (9). Equal volumes (50 ml) of 1-octanol (Sigma Chemical Co., St. Louis, MO) and bidistilled water were poured into a separating flask and thoroughly shaken for 5 min. Four grams of each salt was then added, and the flask content was thoroughly mixed three times for 5 min each time, with a rest period of 5 min after each agitation. After complete separation (20 to 24 h at room temperature), the two phases were recovered separately in different flasks, and the concentration of the accompanying ion of the salt was measured in each phase by atomic absorption (model 3300 unit; Perkin-Elmer, Ueberlinger, Germany). The Po/w coefficient was calculated as the ratio of the concentration of ion in 1-octanol to the concentration of ion in the aqueous phase. Sodium benzoate was found to be the most lipophilic (Po/w = 1.41 × 10−2), followed by potassium sorbate (Po/w = 7.6 × 10−3) and sodium metabisulfite (Po/w = 2.0 × 10−4). Most other salts, sodium chloride (reference salt), sodium bicarbonate and carbonate, sodium propionate, sodium acetate, calcium chloride, and aluminum chloride mainly remained in the aqueous phase (Po/w = 2.0 × 10−5 to 5.0 × 10−5). This lipophilic characteristic of benzoate and sorbate ions would result from a reduced charge density in their molecules (due to the conjugated double bonds in their molecules). An addition parameter, pK′ + pPo/w, which combines the two properties of salts ions, i.e., the water-ionizing capacity (pK′) and the lipophilicity (pPo/w = −log Po/w), appears to provide a more general basis for the inhibitory effect of salts (Fig. (Fig.2B).2B). This suggests that while the dissociation constant of ions plays a major role in growth inhibition, as seen in Fig. Fig.2A,2A, the lipophilic character of the preservative-salt ions confers to them an added ability to penetrate the cell envelope and to inhibit bacterial growth (5, 10). The exclusion of ammonium (lower inhibition than expected from its pK′a value) and calcium (higher inhibition than expected from its pK′a value) ions from the sigmoidal pattern portrayed in Fig. Fig.2B2B would have resulted from their interactions with water and other molecules (NH4+) (1) or from cell membrane destabilization (Ca2+) (23).In conclusion, the study has shown that several salts (0.2 M concentration), including aluminum dihydroxy acetate, aluminum chloride, aluminum lactate, ammonium acetate, potassium sorbate, sodium benzoate, sodium metabisulfite, sodium bicarbonate, sodium carbonate, sodium propionate, and trisodium phosphate, strongly inhibited the growth of P. carotovorum subsp. carotovorum and P. atrosepticum. In addition, the study has established for the first time a basic sigmoidal relationship between the antimicrobial activity of the salts and the physicochemical characteristics of their constituent ions, namely their water-ionizing capacity and their lipophilicity. The constituent ions of the highly inhibiting salts generally displayed a high capacity to ionize water molecules (low pK′a or pK′b values) (Al3+, CO32−, PO43−, HCO3, and HSO3) or a high lipophilicity (benzoate and sorbate), and these two parameters in combination with known biochemical activities of salts ions would affect bacterial growth.  相似文献   

15.
Melioidosis has been considered an emerging disease in Brazil since the first cases were reported to occur in the northeast region. This study investigated two municipalities in Ceará state where melioidosis cases have been confirmed to occur. Burkholderia pseudomallei was isolated in 26 (4.3%) of 600 samples in the dry and rainy seasons.Melioidosis is an endemic disease in Southeast Asia and northern Australia (2, 4) and also occurs sporadically in other parts of the world (3, 7). Human melioidosis was reported to occur in Brazil only in 2003, when a family outbreak afflicted four sisters in the rural part of the municipality of Tejuçuoca, Ceará state (14). After this episode, there was one reported case of melioidosis in 2004 in the rural area of Banabuiú, Ceará (14). And in 2005, a case of melioidosis associated with near drowning after a car accident was confirmed to occur in Aracoiaba, Ceará (11).The goal of this study was to investigate the Tejuçuoca and Banabuiú municipalities, where human cases of melioidosis have been confirmed to occur, and to gain a better understanding of the ecology of Burkholderia pseudomallei in this region.We chose as central points of the study the residences and surrounding areas of the melioidosis patients in the rural areas of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50′W) (Fig. (Fig.1).1). There are two well-defined seasons in each of these locations: one rainy (running from January to May) and one dry (from June to December). A total of 600 samples were collected at five sites in Tejuçuoca (T1, T2, T3, T4, and T5) and five in Banabuiú (B1, B2, B3, B4, and B5), distributed as follows (Fig. (Fig.2):2): backyards (B1 and T1), places shaded by trees (B2 and T2), water courses (B3 and T3), wet places (B4 and T4), and stock breeding areas (B5 and T5).Open in a separate windowFIG. 1.Municipalities of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50″W).Open in a separate windowFIG. 2.Soil sampling sites in Banabuiú and Tejuçuoca.Once a month for 12 months (a complete dry/rainy cycle), five samples were gathered at five different depths: at the surface and at 10, 20, 30 and 40 cm (Table (Table1).1). The samples were gathered according to the method used by Inglis et al. (9). Additionally, the sample processing and B. pseudomallei identification were carried out as previously reported (1, 8, 9).

TABLE 1.

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

16.
17.
18.
Vertebrate genomic assemblies were analyzed for endogenous sequences related to any known viruses with single-stranded DNA genomes. Numerous high-confidence examples related to the Circoviridae and two genera in the family Parvoviridae, the parvoviruses and dependoviruses, were found and were broadly distributed among 31 of the 49 vertebrate species tested. Our analyses indicate that the ages of both virus families may exceed 40 to 50 million years. Shared features of the replication strategies of these viruses may explain the high incidence of the integrations.It has long been appreciated that retroviruses can contribute significantly to the genetic makeup of host organisms. Genes related to certain other viruses with single-stranded RNA genomes, formerly considered to be most unlikely candidates for such contribution, have recently been detected throughout the vertebrate phylogenetic tree (1, 6, 13). Here, we report that viruses with single-stranded DNA (ssDNA) genomes have also contributed to the genetic makeup of many organisms, stretching back as far as the Paleocene period and possibly the late Cretaceous period of evolution.Determining the evolutionary ages of viruses can be problematic, as their mutation rates may be high and their replication may be rapid but also sporadic. To establish a lower age limit for currently circulating ssDNA viruses, we analyzed 49 published vertebrate genomic assemblies for the presence of sequences derived from the NCBI RefSeq database of 2,382 proteins from known viruses in this category, representing a total of 23 classified genera from 7 virus families. Our survey uncovered numerous high-confidence examples of endogenous sequences related to the Circoviridae and to two genera in the family Parvoviridae: the parvoviruses and dependoviruses (Fig. (Fig.11).Open in a separate windowFIG. 1.Phylogenetic tree of vertebrate organisms and history of ssDNA virus integrations. Times of integration of ancestral dependoviruses (yellow icosahedrons), parvoviruses (blue icosahedrons), and circoviruses (triangles) are approximate.The Dependovirus and Parvovirus genomes are typically 4 to 6 kb in length, include 2 major open reading frames (encoding replicase proteins [Rep and NS1, respectively] and capsid proteins [Cap and VP1, respectively]), and have characteristic hairpin structures at both ends (Fig. (Fig.2).2). For replication, these viruses depend on host enzymes that are recruited by the viral replicase proteins to the hairpin regions, where self-primed viral DNA synthesis is initiated (2). Circovirus genomes are typically ∼2-kb circles. DNA of the type species, porcine circovirus 1 (PCV-1), contains a stem-loop structure within the origin of replication (Fig. (Fig.2),2), and the largest open reading frame includes sequences that are homologous to the Parvovirus replicase open reading frame (9, 11). The circoviruses also depend on host enzymes for replication, and DNA synthesis is self-primed from a 3′-OH end formed by endonucleolytic cleavage of the stem-loop structure (4). The frequency of Dependovirus infection is estimated to be as high as 90% within an individual''s lifetime. None of the dependoviruses have been associated with human disease, but related viruses in the family Parvoviridae (e.g., erythrovirus B19 and possibly human bocavirus) are pathogenic for humans, and members of both the Parvoviridae and the Circoviridae can cause a variety of animal diseases (2, 4).Open in a separate windowFIG. 2.Schematics illustrating the structure and organization of Parvoviridae and Circoviridae genomes and origins of several of the longest-integrated ancestral viral sequences found in vertebrates. Integrations were aligned to the Dependovirus adeno-associated virus 2 (AAV2), the Parvovirus minute virus of mice (MVM), and the Circovirus porcine circovirus 1 (PCV-1). The inverted terminal repeat (ITR) sequences in the Dependovirus and Parvovirus genomes are depicted on an expanded scale. A linear representation of the circular genome of PCV-1 is shown with the 10-bp stem-loop structure on an expanded scale. Horizontal lines beneath the maps indicate the lengths of similar sequences that could be identified by BLAST. The numbers indicate the locations of amino acids in the viral proteins where the sequence similarities in the endogenous insertions start and end. The actual ancestral virus-derived integrated sequences may extend beyond the indicated regions.With some ancestral endogenous sequences that we identified, phylogenetic comparisons can be used to estimate age. For example, as a Dependovirus-like sequence is present at the same location in the genomes of mice and rats, the ancestral virus must have existed before their divergence, more than 20 million years ago. Some Circovirus- and Dependovirus-related integrations also predate the split between dog and panda, about 42 million years ago. However, in most other cases, we rely on an indirect method for estimating age (1). As genomic sequences evolve, they accumulate new stop codons and insertion/deletion-induced frameshifts. The rates of these events can be tied directly to the rates of neutral sequence drift and, therefore, the time of evolution. To apply this method, we first performed a BLAST search of vertebrate genomes for all known ssDNA virus proteins (BLAST options, -p tblastn -M BLOSUM62 -e 1e−4). Candidate sequences were then recorded, along with 5 kb of flanking regions, and then again aligned against the database of ssDNA viruses to find the most complete alignment (BLAST options, -t blastx -F F -w 15 -t 1500 -Z 150 -G 13 -E 1 -e 1e−2). Detected alignments were then compared with a neutral model of genome evolution, as described in the supplemental material, and the numbers of stop codons and frameshifts were converted into the expected genomic drift undergone by the sequences. The age of integration was then estimated from the known phylogeny of vertebrates (7, 10). Using these methods, we discovered that as many as 110 ssDNA virus-related sequences have been integrated into the 49 vertebrate genomes considered, during a time period ranging from the present to over 40 to 60 million years ago (Table (Table1;1; see also Tables S1 to S3 in the supplemental material).

TABLE 1.

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

19.
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   

20.
Alexey Yanchukov 《Genetics》2009,182(4):1117-1127
A model of genomic imprinting with complete inactivation of the imprinted allele is shown to be formally equivalent to the haploid model of parental selection. When single-locus dynamics are considered, an internal equilibrium is possible only if selection acts in the opposite directions in males and females. I study a two-locus version of the latter model, in which maternal and paternal effects are attributed to the single alleles at two different loci. A necessary condition for the allele frequency equilibria to remain on the linkage equilibrium surface is the multiplicative interaction between maternal and paternal fitness parameters. In this case the equilibrium dynamics are independent at both loci and results from the single-locus model apply. When fitness parameters are additive, analytic treatment was not possible but numerical simulations revealed that stable polymorphism characterized by association between loci is possible only in several special cases in which maternal and paternal fitness contributions are precisely balanced. As in the single-locus case, antagonistic selection in males and females is a necessary condition for the maintenance of polymorphism. I also show that the above two-locus results of the parental selection model are very sensitive to the inclusion of weak directional selection on the individual''s own genotypes.PARENTAL genetic effects refer to the influence of the mother''s and father''s genotypes on the phenotypes of their offspring, not attributable just to the transfer of genes. Examples have been documented across a wide range of areas of organism biology; see, for example, Wade (1998) and and22 in Rasanen and Kruuk (2007). Parental selection is a more formal concept used in theoretical modeling and concerns situations where the fitness of the offspring depends, besides other factors, on the genotypes of its parent(s) (generalizing from Kirkpatrick and Lande 1989).

TABLE 1

Frequencies of genotypes and fitness parameterizations in model 1
Gametes/haploidsFrequency before selectionFitness
ZygoteMaleFemale
(A)AApfpm1 − α1 − δ
(A)a1/2 A 1/2 apf(1 − pm)11
(a)A1/2 a 1/2 A(1 − pf)pm1 − α1 − δ
(a)aA(1 − pf)(1 − pm)11
Open in a separate windowParentheses in the first column indicate maternal genotype (parental selection model) or inactivation of the maternally derived allele (imprinting model). Whether selection occurs at the diploid (first column) or subsequent haploid (second column) stage does not change the resulting allele frequencies.

TABLE 2

Offspring genotypic proportions from different mating types, sorted among four phenotypic groups/combinations of maternal and paternal effects: model 2
Offspring genotypes/phenotypes
Parental genotypes
Paternal (φ = 1)
Joint (φ = 4)
MaleFemaleABAbaBAbABAbaBab
ABAB1
Ab
aB
ab(1−r)/2r/2r/2(1−r)/2
AbAB
Ab1
aBr/2(1−r)/2(1−r)/2r/2
ab
Offspring genotypes/phenotypes
Parental genotypes
Maternal (φ = 2)
None (φ = 3)
MaleFemaleABAbaBAbABAbaBab
aBAB
Abr/2(1 − r)/2(1 − r)/2r/2
aB1
ab
abAB(1 − r)/2(1 − r)/2
Ab
aB
ab1
Open in a separate windowAnother well-known parent-of-origin phenomenon is genomic imprinting. Here, the level of expression of one of the alleles depends on which parent it is inherited from. Often it is difficult to tell apart the phenotypic patterns due to parental effects and genomic imprinting, and thus a problem arises in the process of identifying the candidate genes for such effects (Hager et al. 2008). Analytic methods (Weinberg et al. 1998; Santure and Spencer 2006; Hager et al. 2008) have been developed to quantify subtle differences between the two. In this article, I point out that a simple mathematical model, first suggested for genomic imprinting at a diploid locus, can be interpreted, without any formal changes, to describe parental selection on haploids.While there has been much progress in understanding the evolution of genomic imprinting (Hunter 2007), including advances in modeling (Spencer 2000, 2008), the population genetics theory of parental effects received less attention. Existing major-locus effect models of parental selection are single-locus, two-allele, and mostly concern uniparental (maternal) selection (Wright 1969; Spencer 2003; Gavrilets and Rice 2006; Santure and Spencer 2006), with only one specific case where the fitness effects of both parents interact studied by Gavrilets and Rice (2006). No attempt to extend this theory into multilocus systems has yet been made. Considering a two-locus model with both parents playing a role in selection on the offspring is called for by the observation that many maternal and paternal effects aim at the different traits or different life stages of their progeny. Among birds, for example, body condition soon after hatching is largely determined by the mother, while paternally transmitted sexual display traits develop much later in life (Price 1998). Such effects are therefore unlikely to be regulated within a single locus. Sometimes the effects are on the same trait, but still attributed to different loci: expression of gene Avy that causes the “agouti” phenotype (yellow fur coat and obesity) in mice is enhanced by maternal epigenetic modification (Rakyan et al. 2003), while paternal mutations at the other locus, MommeD4, contribute to a reverse phenotypic pattern in the offspring (Rakyan et al. 2003). The epigenetic state of the murine AxinFu allele is both maternally and paternally inherited (Rakyan et al. 2003).Focusing selection on haploids reduces the number of genotypes that need to be taken into account, while preserving the main properties of the multilocus system. Genes with haploid expression and a potential of parental effects can be found in two major taxonomic kingdoms. A notable candidate is Spam1 in mice, which is expressed during spermogenesis and encodes a factor that enables sperm to penetrate the egg cumulus (Zheng et al. 2001). This gene remains a target for effectively haploid selection, because its product is not shared via cytoplasm bridges between developing spermatides. Mutations at Spam1 alter performance of the male gametes that carry it and might indirectly, perhaps by altering the timing of fertilization, affect the fitness of the zygote. The highest estimated number of mouse genes expressed in the male gametes is currently 2375 (Joseph and Kirkpatrick 2004), and one might expect some of them to have similar paternal effects. Plants go through a profound haploid stage in their life cycles, and genes involved at this stage have an inevitable effect on the fitness of the future generations. In angiosperms, seed development is known to be controlled by both maternal (Chaudhury and Berger 2001; Yadegari and Drews 2004) and paternal (Nowack et al. 2006) effect genes, expressed, respectively, in female and male gametophytes.Under haploid selection, there can be no overdominance, and thus polymorphism is much more difficult to maintain than in diploid selection models (summarized in Feldman 1971). Nevertheless, differential or antagonistic selection between sexes can lead to a new class of stable internal equilibria in the diploid systems (Owen 1953; Bodmer 1965; Mandel 1971; Kidwell et al. 1977; Reed 2007), and I make use of this property in the haploid models developed below. In the experiment by Chippindale and colleagues (Chippindale et al. 2001), ∼75% of the total fitness variation in the adult stage of Drosophila melanogaster was negatively correlated between males and females, which suggests that a substantial portion of the fruit fly expressed genome is under sexually antagonistic selection. I assume that the effect of either parent on the fitness of the individual depends on the sex of the latter, which in respect to modeling is equivalent to the assumption of differential viability between the sexes in the progeny of the same parent(s). Biological systems that satisfy the latter assumptions can be found among colonial green algae: many members of the order Volvocales are haploid except for the short zygotic stage, and during sexual reproduction, they are also dioecious and anisogametic. I return to this example in the discussion. The possibility that genes expressed in animal gametes may be under antagonistic selection between sexes has been discussed (Bernasconi et al. 2004). For example, a (hypothetical) mutation increasing the ATP production in mitochondria would be beneficial in sperm, because of the increased mobility of the latter, but neutral or detrimental in the egg, due to a higher level of oxidative damage to DNA (Zeh and Zeh 2007).My main purpose was to derive conditions for existence and stability of the internal equilibria of the model(s). I begin with a simple one-locus case, which can be analyzed explicitly, and show how these one-locus results can be extended to the case of two recombining loci with multiplicative fitness. Then, I assume an additive relation between the maternal and paternal effect parameters and study the special cases where parental effects are symmetric.  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