首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 109 毫秒
1.
The potential spread of prion infectivity in secreta is a crucial concern for prion disease transmission. Here, serial protein misfolding cyclic amplification (sPMCA) allowed the detection of prions in milk from clinically affected animals as well as scrapie-exposed sheep at least 20 months before clinical onset of disease, irrespective of the immunohistochemical detection of protease-resistant PrPSc within lymphoreticular and central nervous system tissues. These data indicate the secretion of prions within milk during the early stages of disease progression and a role for milk in prion transmission. Furthermore, the application of sPMCA to milk samples offers a noninvasive methodology to detect scrapie during preclinical/subclinical disease.PrPSc, a disease-specific marker for prion diseases and the likely infectious agent, is widely distributed within the central nervous system (CNS) and lymphoreticular tissues (LRS) in ovine scrapie, human variant Creutzfeldt-Jakob disease (vCJD), and cervine chronic wasting disease (CWD) during both clinical and preclinical stages (4, 11, 25). Furthermore, while the LRS distribution of PrPSc is much more restricted in bovine spongiform encephalopathy (BSE), sheep experimentally infected with BSE display a PrPSc tissue distribution more akin to that of ovine scrapie (11).For rodent-adapted scrapie and cervine CWD, the disease agent is detected in excreta when animals are in the clinical stages of disease, a process likely to contribute to environmental reservoirs of infectivity and lateral disease transmission (5, 13, 21). Within an experimental rodent model, it has also been demonstrated that the shedding of PrPSc and concomitant infectivity in feces occurs during preclinical scrapie (21).Evidence now also demonstrates that milk provides a vehicle for the transmission for prion diseases. Scrapie-free lambs fed milk from clinical scrapie-affected ewes propagate PrPSc within their LRS (8). Additionally, a recent study using a transgenic mouse bioassay demonstrated the secretion of infectivity in milk from preclinical animals where scrapie infectivity was found in milk months before the onset of clinical signs in animals with an ARQ/VRQ PrP genotype (10). The presence of scrapie infectivity within milk was irrespective of mammary gland pathology or PrPSc accumulation, and these animals were estimated to have considerable accumulation of immunohistochemically (IHC) detectable PrPSc within the LRS at the time of sampling.Here, we applied serial protein misfolding cyclic amplification (sPMCA) to the in vitro detection of PrPSc within sheep milk (Fig. (Fig.1)1) (Table (Table11).Open in a separate windowFIG. 1.sPMCA analysis of ovine milk samples. Milk was clarified and seeded into brain homogenate from sheep unexposed to the scrapie agent. Samples underwent sPMCA, and products were digested with proteinase K before analysis of 10 μl of each sample. PrP was detected with monoclonal antibodies SHA31 and P4; molecular mass markers are indicated (kDa). Milk was sampled from animals not exposed to the scrapie agent (U), those displaying clinical signs of scrapie (C), or those exposed to a scrapie-positive farm environment but not displaying clinical disease (S). NS, non-seeded PMCA brain substrate subjected to identical sPMCA conditions at the same time as positive samples were analyzed. Samples from the four nonexposed animals were analyzed 18 to 20 times each by sPMCA. Samples from clinically affected or clinically normal scrapie-exposed animals were analyzed in triplicate. For this triplicate analysis of each sample, the sPMCA round at which samples became positive is indicated under the appropriate lane. n, negative at round 12. Each sample was PrPSc negative until the stated round and thereafter was positive.

TABLE 1.

Timeline of exposure of animals to a scrapie-positive farm environment, sample collection, and scrapie status
Animala (PrP genotypeb)Age at exposurecDays postexposure at lactationDays postlactation to clinical scrapiedClinical statusePrPSc detection at postmortemfPrPSc detection in milk (positive tests/total tests)g
1349/08 (VRQ/VRQ)Not exposedNAh,iNANegativeNegative0/20
K489 (VRQ/VRQ)Not exposedNAiNANegativeNA (still alive)0/18
0618/06 (VRQ/VRQ)Not exposedNAiNANegativeNegative0/20
1348/08 (VRQ/VRQ)Not exposedNAiNANegativeNegative0/20
0695/07 (VRQ/VRQ)Birth666-6800PositivePositive3/3
0334/07 (VRQ/VRQ)Birth661-6660PositivePositive3/3
0335/07 (VRQ/VRQ)Birth666-6740PositivePositive3/3
0350/07 (VRQ/VRQ)Birth663-6760PositivePositive3/3
0333/07 (VRQ/VRQ)Birth667-6750PositivePositive3/3
0142/07 (VRQ/VRQ)Birth665-6730PositivePositive3/3
0326/07 (VRQ/VRQ)Birth670-6760PositivePositive2/3
0199/07 (VRQ/VRQ)Birth6640PositivePositive2/3
0692/07 (ARQ/VRQ)∼480 days1,003>450NegativePositive2/3
0480/07 (ARQ/VRQ)∼480 days1,003>355NegativePositive3/3
0349/07 (ARQ/VRQ)∼480 days1,003>348NegativePositive3/3
0822/07 (ARQ/VRQ)Birth760>564NegativeNegative2/3
2295 (AHQ/VRQ)∼120 days1,376>621NegativeNegative3/3
3148 (ARR/VRQ)Birth1,288>621NegativeNA (still alive)2/3
1514 (ARR/VRQ)Unknown597>621NegativeNA (still alive)2/3
1518 (ARR/VRQ)Unknown597>621NegativeNA (still alive)1/3
1244 (ARR/VRQ)Birth1,130>621NegativeNA (still alive)3/3
Open in a separate windowaAll animal procedures were performed under Home Office (United Kingdom) and local ethical review committee approval and compliance with the Animal (Scientific Procedures) Act of 1986.bAmino acid residues at positions 136, 154, and 171 of the PRNP gene.cIntroduction into a scrapie-affected flock.dDays postlactation to postmortem or as of 12 December 2008 for animals that were still alive at the time this paper was written.eClinical disease usually included head tremors and pruritus with associated wool loss and nervousness. The indicated clinical status was applicable throughout lactation to either postmortem or as of 12 December 2008 for animals that were still alive at the time this paper was written.fPrPSc was analyzed by IHC, Western blot analysis, or enzyme-linked immunosorbent assay. All animals with a positive result contained PrPSc within both brain and lymphatic tissues.gsPMCA was used for PrPSc detection and the results are tallied within this column. Replica analysis of a single milk sample from each animal was carried out. For animals 0695/07, 0334/07, 0335/07, 0350/07, 0333/07, 0142/07, and 0326/07, multiple milk samples were collected during the lactation period indicated and multiple samples from an individual animal were pooled before analysis.hNA, not applicable.iNonexposed animals were 750 to 1,110 days old at lactation and where applicable were 1,200 to 1,650 days old at postmortem.PMCA was first described by Saborio and colleagues (20) and allows the amplification of minute quantities of PrPSc (18). In rodent scrapie models, this methodology has detected PrPSc in both blood (2, 18) and brain (22) material in the clinical and preclinical stages of disease as well as in urine excreted during clinical disease (14). This technique has recently been applied to the high-level amplification of PrPSc from natural hosts of prion diseases, including vCJD (7), CWD (9), and scrapie (23). Fresh ovine milk was obtained from individual sheep at least 7 days postpartum. Milk was collected from individual animals into sterile containers and stored on ice for shipping. Within 48 h of collection, milk samples were stored at −80°C. Colostrum was not analyzed. After thawing milk samples, samples from the same individual animal were pooled and EDTA, Nonidet P-40, and sodium deoxycholate were added to final concentrations of 50 mM, 0.5% (vol/vol), and 0.5% wt/vol, respectively. Samples (1 ml) were centrifuged for 10 min at 16,000 × g. After cooling on ice for 5 min, clarified milk supernatant was withdrawn from under the solidified fat layer.sPMCA was carried out as described by Thorne and Terry (23), who demonstrated that samples from a range of animals containing at least one VRQ PrP allele could be amplified by this technique. Clarified milk supernatant was diluted 1 in 10 into PMCA brain homogenate substrate (10% [wt/vol] brain homogenate from a VRQ/VRQ PrP genotype animal within 150 mM NaCl, 4 mM EDTA, pH 8.0, 1.0% [wt/vol] Triton X-100, and miniprotease inhibitor; Roche) to a final volume of 100 μl. Samples contained within sealed 0.2-ml PCR tubes were placed in a rack within an ultrasonicating water bath (model 3000; Misonix) that held the bottom of the tubes 0.4 cm above the sonicator horn. Water was added to the water bath up to the rack surface, immersing the sonicator horn. The water bath was held at 37°C, and sonications were performed for 40 s at 200 W, equivalent to 80% of the maximum power output of the machine. Sonications were repeated once every 30 min for 24 h (one PMCA round), after which the amplified samples were diluted 1 in 3 with PMCA substrate in a final volume of 100 μl and the sample was subjected to further rounds of PMCA. Twelve PMCA rounds were performed for each sample, a total of 576 sonications over 12 days. PMCA samples were digested with 50 μg/ml proteinase K for 1 h at 37°C before analysis of 10 μl of each sample by Western blotting using 12% (wt/vol) acrylamide NuPAGE precast Bis-Tris gels (as described in reference 15). All clinical scrapie-affected animals or those exposed to the scrapie agent were challenged by introduction into the Ripley flock (Veterinary Laboratories Agency, United Kingdom), where natural scrapie is endemic with a high incidence since 1996. Ryder and coworkers (17) reported that all animals with PrP genotypes VRQ/VRQ and ARQ/VRQ that were born into this flock developed scrapie, with incubation periods of 21 to 28 months and 28 to 39 months, respectively. When ARQ/VRQ animals were introduced into the flock at 6 to 26 months of age, 77% of the animals had subclinical scrapie 24 to 30 months later, as detected by IHC analysis of the LRS. Here, PrPSc was detected in the milk from clinically affected animals at a rate of 92% (24 analyses; triplicate analyses of samples from 8 animals) and from scrapie-exposed, clinically normal sheep at a rate of 78% (27 analyses; triplicate analyses of samples from 9 animals) (Fig. (Fig.1)1) (Table (Table1).1). All scrapie-exposed sheep, both clinically affected and clinically normal, tested positive for PrPSc in at least one sPMCA reaction. PrPSc was amplified from the milk of sheep with VRQ/VRQ, ARR/VRQ, ARQ/VRQ, and AHQ/VRQ PrP genotypes (Table (Table1).1). It required at least four to eight rounds of sPMCA to produce detectable PrPSc within a milk sample from each of the scrapie-exposed sheep (Fig. (Fig.1).1). Replica analysis of a pooled milk sample from each individual sheep occasionally demonstrated high variability in the round that samples became positive for PrPSc (Fig. (Fig.1);1); this result may indicate the presence of very low levels of PrPSc (19) and/or heterogeneity within milk samples. Analyses of ovine milk from a New Zealand-derived scrapie-free flock kept under strict biosecurity conditions (ADAS, United Kingdom) did not amplify PrPSc within 12 rounds of sPMCA (78 analyses; up to 20 replica analyses of samples from 4 animals). For each of the sPMCA analyses, both positive and negative samples were analyzed concurrently within the same run on the same sonicator. These data demonstrate that PrPSc amplified from samples from scrapie-exposed animals is not due to spontaneous PrPSc formation or cross-contamination between samples within the sPMCA procedure. It is of note that prions were shed within milk from clinically normal, scrapie-exposed animals with multiple PrP genotypes. The ARQ/VRQ genotype is indicative of a high level of disease penetrance and widespread preclinical PrPSc accumulation within the LRS system, whereas AHQ/VRQ and ARR/VRQ genotype animals typically have much lower disease penetrance (24) and LRS involvement (11). This indicates the secretion of prions within milk regardless of high-level PrPSc accumulation within the LRS and also the very likely detection of subclinical as well as preclinical disease in some of these animals.No clinical scrapie-affected animals displayed clinical mastitis, and PrPSc was not detected within mammary gland tissue from five sheep with clinical scrapie (Fig. (Fig.22 and data not shown). This is in agreement with the study by Lacroux et al. (10), indicating that while the accumulation of PrPSc within mammary gland tissue can occur, it is not a prerequisite for its deposition within milk. Here, postmortem detection of PrPSc was carried out by routine diagnosis using IHC and Western blot analysis of the obex. Exceptions were animals 1349/08 and 1348/08, where obex tissue was analyzed by Bio-Rad TeSeE enzyme-linked immunosorbent assay (Table (Table1).1). Postmortem IHC examination of palatine tonsil, ileal Peyer''s patches, medial retropharyngeal lymph node, and mesenteric lymph node tissue was also carried out. Scrapie-exposed animals were shown to secrete PrPSc within their milk irrespective of whether they could be confirmed as scrapie positive by postmortem immunoassay detection of PrPSc within the CNS and LRS (Table (Table1).1). This discrepancy in PrPSc detection may well reflect the greater sensitivity of sPMCA compared to immunoassay detection of PrPSc; these results also indicate that PrPSc is secreted during the early stages of disease progression. Scrapie-exposed animals had PrPSc detected within their milk at least 20 months prior to possible clinical onset of disease, and this was not apparently influenced by the PrP genotype.Open in a separate windowFIG. 2.Detection of protease-resistant PrPSc within CNS and mammary gland tissues of animals displaying clinical scrapie. Tissues were prepared as 10% or 40% (wt/vol) homogenates for spinal cord and mammary gland tissue, respectively, as described previously elsewhere (15). Native or proteinase K-digested homogenates (25 μg/ml; 1 h at 37°C) were analyzed as indicated. Protease-resistant PrPSc was readily detectable within spinal cord tissue (SC; lanes 1 to 2) but was not detectable within mammary gland samples (MG; lanes 3 to 6). Either 0.33 mg (0350/07) or 0.165 mg (0326/07 and 0344/07) of spinal cord tissue and 1.32 mg (lanes 3 and 4) and 6 mg (lane 5) of mammary gland tissue was analyzed per lane. Protease-resistant PrPSc was still undetectable from 20 mg of mammary gland tissue following precipitation with sodium phosphotungstic acid (25) prior to analysis (lane 6). PrPSc within scrapie-positive brain tissue (63 μg) was readily detected by this method after spiking into 20 mg of mammary gland homogenate (B, lane 7). Full-length and fragmented protease-sensitive PrPC was readily detected within mammary gland tissue (lane 3). Animal numbers are indicated. PrP was detected with monoclonal antibody SHA31; molecular mass markers are indicated (kDa).These data clearly demonstrate that the secretion of PrPSc within milk occurs in natural scrapie. There are several routes through which the prion protein could be secreted into milk. Evidence suggests that within ovine mammary gland tissue, PrPC is actively produced within epithelial cells, and its secretion is most likely by exocytosis and the apocrine secretion of fat globules (3). It is unknown whether PrPSc is produced within epithelial cells and secreted into milk through similar mechanisms. Alternative mechanisms are through vesicular transcytosis or paracellular transport of PrPSc from the blood. It is established that prion-infected animals harbor infectivity and PrPSc within the blood during preclinical disease (6, 18) and that blood components are secreted within milk, including cell types known to colocalize with PrPSc within ovine mammary glands (12).Results indicate the potential transmission of scrapie in the milk of infected sheep for a prolonged period prior to clinical onset. As well as ewe-to-lamb disease transmission, this process is also likely to contribute to lateral transmission, as lambs fed milk from clinically infected ewes were the source for the transmission of scrapie between lambs within the first few months after birth (8). It is unknown whether other prion diseases result in the secretion of prions within milk. CWD, vCJD, and experimental ovine BSE share similarities with scrapie in the tissue distribution of infectivity, and it seems plausible that an analogous secretion mechanism may occur. Given the extended preclinical stages and the purported importance of subclinical states for these diseases (16), such an outcome would have significant implications for the transmission of prion diseases from apparently healthy animals and humans.With regard to ovine milk and milk products, scrapie is not transmissible to humans, and to date there is no evidence of the natural occurrence of ovine BSE. As such, the reported findings do not indicate the likely introduction of zoonotic prions from sheep into the human food chain. Nevertheless, the presented data do indicate caution in the risk assessment associated with such foods. Also, it is unknown if analogous shedding of prions into milk occurs with bovine BSE; evidence from previous epidemiological and bioassay studies would suggest that such a scenario seems unlikely to cause clinical disease (1, 26). However, the present report demonstrates that prions are secreted within the milk of sheep with PrP genotypes not typically associated with LRS accumulation of PrPSc and that prions were secreted from animals devoid of IHC-detectable PrPSc within their LRS. Such PrPSc tissue distribution is similar to bovine BSE, and given the importance of bovine milk in the human diet, the potential presence of low levels of prions within bovine milk warrants further investigation.Finally, analyzing milk samples by sPMCA offers a methodology with a clear potential for the identification of clinically sick animals and those with preclinical/subclinical scrapie. Such a noninvasive live-animal assay has the potential to contribute to the epidemiological study, management, and control of prion diseases within farmed animals.  相似文献   

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

TABLE 1.

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

3.
4.
We have previously shown that Fhit tumor suppressor protein interacts with Hsp60 chaperone machinery and ferredoxin reductase (Fdxr) protein. Fhit-effector interactions are associated with a Fhit-dependent increase in Fdxr stability, followed by generation of reactive oxygen species and apoptosis induction under conditions of oxidative stress. To define Fhit structural features that affect interactions, downstream signaling, and biological outcomes, we used cancer cells expressing Fhit mutants with amino acid substitutions that alter enzymatic activity, enzyme substrate binding, or phosphorylation at tyrosine 114. Gastric cancer cell clones stably expressing mutants that do not bind substrate or cannot be phosphorylated showed decreased binding to Hsp60 and Fdxr and reduced mitochondrial localization. Expression of Fhit or mutants that bind interactor proteins results in oxidative damage and accumulation of cells in G2/M or sub-G1 fractions after peroxide treatment; noninteracting mutants are defective in these biological effects. Gastric cancer clones expressing noncomplexing Fhit mutants show reduction of Fhit tumor suppressor activity, confirming that substrate binding, interaction with heat shock proteins, mitochondrial localization, and interaction with Fdxr are important for Fhit tumor suppressor function.Fhit protein is a powerful tumor suppressor that is frequently lost or reduced in cancer cells because of rearrangement of the exquisitely DNA damage-sensitive fragile FHIT gene. Restoration of Fhit expression suppresses tumorigenicity of cancer cells of various types, and the ability to induce apoptosis in cancer cells in vitro is reduced by specific Fhit mutations (1, 2).Through studies of signal pathways affected by Fhit expression, by searches for Fhit protein effectors, and by in vitro analyses of Fhit activity, we and others have defined Fhit enzymatic activity in vitro (3), apoptotic activity in cells and tumors (46), and most recently identification of a Fhit protein complex that affects Fhit stability, mitochondrial localization, and interaction with ferredoxin reductase (Fdxr)5 (7). The complex includes Hsp60 and Hsp10 that mediate Fhit stability and may affect import into mitochondria, where Fhit interacts with Fdxr, which is responsible for transferring electrons from NADPH to cytochrome P450 via ferredoxin. Virally mediated Fhit restoration in Fhit-deficient cancer cells increases production of intracellular reactive oxygen species (ROS), followed by increased apoptosis of cancer cells under oxidative stress conditions; conversely, Fhit-negative cells escape apoptosis, likely carrying oxidative DNA damage that contributes to accumulation of mutations.The Fhit protein sequence, showing high homology to the histidine triad (HIT) family of proteins, suggested that the protein product would hydrolyze diadenosine tetraphosphate or diadenosine triphosphate (Ap3A) (8), and in vitro studies showed that Ap3A was cleaved into ADP and AMP by Fhit. The catalytic histidine triad within Fhit was essential for catalytic activity (3), and a Fhit mutant that substituted Asn for His at the central histidine (H96N mutant) was catalytically inactive, although it bound substrate well (3). Early tumor suppression studies showed that cancer cells stably transfected with wild type (WT) or H96N mutant Fhit were suppressed for tumor growth in nude mice. This suggested the hypothesis that the Fhit-substrate complex sends the tumor suppression signal (9, 10). To test this hypothesis, a series of FHIT alleles was designed to reduce substrate-binding and/or hydrolytic rates and was characterized by quantitative cell-death assays on cancer cells virally infected with each allele. The allele series covered defects as great as 100,000-fold in kcat and increases as large as 30-fold in Km. Mutants with 2–7-fold increases in Km had significantly reduced apoptotic indices and the mutant with a 30-fold increase in Km retained little apoptotic function. Thus, the proapoptotic function of Fhit, which is likely associated with tumor suppressor function, is limited by substrate binding and is unrelated to substrate hydrolysis (11).Fhit, a homodimeric protein of 147 amino acids, is a target of tyrosine phosphorylation by the Src family protein kinases, which can phosphorylate Tyr-114 of Fhit in vitro and in vivo (12). After co-expression of Fhit with the Elk tyrosine kinase in Escherichia coli to generate phosphorylated forms of Fhit, unphosphorylated, mono-, and diphosphorylated Fhit were purified, and enzyme kinetics studies showed that monophosphorylated Fhit exhibited monophasic kinetics with Km and kcat values ∼2- and ∼7-fold lower, respectively, than for unphosphorylated Fhit. Diphosphorylated Fhit exhibited biphasic kinetics; one site had Km and kcat values ∼2- and ∼140-fold lower, respectively, than for unphosphorylated Fhit; the second site had a Km ∼60-fold higher and a kcat ∼6-fold lower than for unphosphorylated Fhit (13). Thus, it was possible that the alterations in Km and kcat values for phosphorylated forms of Fhit might favor formation and lifetime of the Fhit-Ap3A complex and enhance tumor suppressor activity (see
Fhit forms
Kinetic parameters
% Sub-G1
Direct binding
Subcellular location
Co-IP in vivo
8-OHdG
Apoptosis
Tumor suppressor
Km (mm)kcat (s–1)A549MKN74Hsp60FdxrHsp60Fdxr
Fhit WT 1.6 +/– 0.19 2.7 +/– 0.95 43 24 Yes Yes Cyt & mito Yes Yes Yes Yes Yes
Catalyt mutants
   H96D Up 2-fold Down >2 × 104 29 NT NT NT Cyt & mito Yes Yes NT Yes NT
   H96N Up 2-fold Down >5 × 105 31 14.4 NT NT Cyt & mito Yes Yes Yes Yes Yes
Loop mutants
   Y114A Up 23-fold Down 2-fold 3.7 NT NT NT Cyt +/– +/– +/– No No
   Y114D NT NT 2.9 6 NT NT Cyt +/– +/– No –/+
   Y114E NT NT NT NT NT NT Cyt & mito –/+ –/+ No NT
   Y114F Up 5-fold Up 1.1-fold 11.5 3 NT NT Cyt & mito –/+ –/+ No No
   Y114W Up 5-fold Up 1.4-fold NT NT NT NT Cyt & mito –/+ NT NT
   del113–117 Up 10-fold Down 38-fold 5 NT NT NT NT NT NT No NT
Other mutants
   L25W Up 7-fold Down 4-fold 15 NT NT NT Cyt NT –/+
   I10W,L25W Up 32-fold Down 6-fold 11 NT NT NT NT NT NT NT NT NT
   F5W Up 3.3 fold NT NT 5 NT NT NT NT NT +/– No NT
Purified pFhit
   pFhit Down 0.4-fold Down 7-fold NA NA –/+ Yes NA NA NA NA NA NA
   ppFhit Down 0.4-fold Down > 100-fold NA NA –/+ Yes NA NA NA NA NA NA
Up 60-fold Down 6-fold
Open in a separate windowTo explore the in vivo importance of the Tyr-114 phosphorylation site and define Fhit-mediated signaling events, Semba et al. (14) compared the differential biological effects of Ad-FHIT WT and Ad-FHIT Tyr-114 mutant expression in human lung cancer cells. Caspase-dependent apoptosis was effectively induced only by WT Fhit protein. However, the biological significance of phosphorylation at Tyr-114 has been difficult to study because the endogenous phosphorylated forms have very short half-lives; activation of epidermal growth facto receptor family members induces Fhit phosphorylation by Src and proteasome degradation of phosphorylated Fhit (15).Although there are possible connections among the various pathways known to be altered in Fhit-deficient cells, apoptosis, DNA damage-response checkpoint activation, ROS production, and related biological effects of Fhit loss or overexpression, details of the pathway(s) leading from Fhit overexpression to cell death and tumor suppression have not been delineated. Now that a Fhit signaling complex has been identified, we set out to examine which structural features of Fhit protein might participate in individual steps of the pathway leading from Fhit overexpression through complex formation, subcellular localization, interaction with mitochondrial Fdxr, DNA damage induction, cell cycle changes, apoptosis, and ultimately tumor suppression. The underlying hypotheses were as follows: substrate-binding mutants would behave similarly to WT; nonsubstrate-binding mutants would be defective in some step of the pathway, perhaps complexing with heat shock proteins or Fdxr or perhaps induction of DNA damage; and Tyr-114 mutants, which also affect formation or stability of the enzyme-substrate complex, would also be defective in executing some step of the Fhit overexpression pathway to cell death. One goal was to identify specific mutants that exhibited deficiency in specific steps of the pathway, so that such mutants could be used to dissect each step in more detail. Using in vitro Fhit and Fhit-effector protein interactions, we aimed to determine the following: 1) which proteins of the complex interact directly with Fhit, and 2) the biological role of these interactions in vivo. Using cancer cells expressing exogenous WT and mutant Fhit proteins, we were able to examine the structural features of Fhit that affect the direct interaction with its effectors, participate in ROS production, and are necessary for tumor suppression activity.  相似文献   

5.
MitoMiner, an Integrated Database for the Storage and Analysis of Mitochondrial Proteomics Data     
Anthony C. Smith  Alan J. Robinson 《Molecular & cellular proteomics : MCP》2009,8(6):1324-1337
  相似文献   

6.
Plasmid pAMS1-Encoded,Bacteriocin-Related “Siblicide” in Enterococcus faecalis     
Christine M. Sedgley  Don B. Clewell  Susan E. Flannagan 《Journal of bacteriology》2009,191(9):3183-3188
  相似文献   

7.
Impact of Viral Dose and Major Histocompatibility Complex Class IB Haplotype on Viral Outcome in Mauritian Cynomolgus Monkeys Vaccinated with Tat upon Challenge with Simian/Human Immunodeficiency Virus SHIV89.6P     
Aurelio Cafaro  Stefania Bellino  Fausto Titti  Maria Teresa Maggiorella  Leonardo Sernicola  Roger W. Wiseman  David Venzon  Julie A. Karl  David O'Connor  Paolo Monini  Marjorie Robert-Guroff  Barbara Ensoli 《Journal of virology》2010,84(17):8953-8958
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   

8.
Nooks and Crannies in Type VI Secretion Regulation     
Christophe S. Bernard  Yannick R. Brunet  Erwan Gueguen  Eric Cascales 《Journal of bacteriology》2010,192(15):3850-3860
  相似文献   

9.
Impact of remuneration and organizational factors on completing preventive manoeuvres in primary care practices     
Dahrouge S  Hogg WE  Russell G  Tuna M  Geneau R  Muldoon LK  Kristjansson E  Fletcher J 《CMAJ》2012,184(2):E135-E143

Background:

Several jurisdictions attempting to reform primary care have focused on changes in physician remuneration. The goals of this study were to compare the delivery of preventive services by practices in four primary care funding models and to identify organizational factors associated with superior preventive care.

Methods:

In a cross-sectional study, we included 137 primary care practices in the province of Ontario (35 fee-for-service practices, 35 with salaried physicians [community health centres], 35 practices in the new capitation model [family health networks] and 32 practices in the established capitation model [health services organizations]). We surveyed 288 family physicians. We reviewed 4108 randomly selected patient charts and assigned prevention scores based on the proportion of eligible preventive manoeuvres delivered for each patient.

Results:

A total of 3284 patients were eligible for at least one of six preventive manoeuvres. After adjusting for patient profile and contextual factors, we found that, compared with prevention scores in practices in the new capitation model, scores were significantly lower in fee-for-service practices (β estimate for effect on prevention score = −6.3, 95% confidence interval [CI] −11.9 to −0.6) and practices in the established capitation model (β = −9.1, 95% CI −14.9 to −3.3) but not for those with salaried remuneration (β = −0.8, 95% CI −6.5 to 4.8). After accounting for physician characteristics and organizational structure, the type of funding model was no longer a statistically significant factor. Compared with reference practices, those with at least one female family physician (β = 8.0, 95% CI 4.2 to 11.8), a panel size of fewer than 1600 patients per full-time equivalent family physician (β = 6.8, 95% CI 3.1 to 10.6) and an electronic reminder system (β = 4.6, 95% CI 0.4 to 8.7) had superior prevention scores. The effect of these three factors was largely but not always consistent across the funding models; it was largely consistent across the preventive manoeuvres.

Interpretation:

No funding model was clearly associated with superior preventive care. Factors related to physician characteristics and practice structure were stronger predictors of performance. Practices with one or more female physicians, a smaller patient load and an electronic reminder system had superior prevention scores. Our findings raise questions about reform initiatives aimed at increasing patient numbers, but they support the adoption of information technology.Primary care providers are increasingly interested in ensuring that preventive health care be part of their work routines.1 This reorientation fits with the evidence that recommendations from family practitioners increase substantially the likelihood of patients undergoing preventive manoeuvres,2 whereas the lack of such recommendations has been linked with patient noncompliance.3,4Studies evaluating adherence to recommended preventive care suggest that the most pervasive barriers rest with the organization of the health care system and the practice itself, such as the absence of external financial incentives for the work done and the lack of a reminder system in the office.3,59Countries attempting to reform their delivery of primary care and improve the delivery of preventive services have often directed their efforts in finding alternatives to the traditional fee-for-service model, in which providers receive payment for each service provided. There are two predominant alternative funding models: capitation (providers receive a fixed lump-sum payment per patient per period, independent of the number of services performed) and salaried remuneration. Some health care systems blend components of fee for service with either of these models or offer additional incentives for reaching defined quality-of-care targets. Despite considerable rhetoric, there is little evidence to point to the remuneration models associated with superior delivery of primary care services.10 The complexity of health care systems makes the evaluation of models through international comparisons difficult.In Canada, the province of Ontario has four primary care funding models (11

Table 1:

Characteristics of the four primary care models in the province of Ontario in 2005/06
Fee for serviceCapitation


CharacteristicSalaried (community health centres)*Traditional*ReformedNew (family health networks)Established (health services organizations)
Year introduced1970s200420011970s

Group size, no. of physicians> 1 (no specific size requirement)1≥ 3≥ 3≥ 3

Physician remunerationSalaryFee for serviceFee for service and incentivesCapitation with 10% fee- for-service component, and incentivesCapitation and incentives

Patient enrolmentRequired; no limit on size of rosterNot requiredRequired; no limit on size of rosterRequired; disincentive to enrol > 2400Required; disincentive to enrol > 2400

Incentive for enhanced preventive care

 Influenza immunization (age ≥ 65 yr)NoneNoneNoneApril 2002July 2003

 Colorectal cancer screening (age 50–74 yr)NoneNoneApril 2006April 2006April 2006

 Breast cancer screening (age 50–70 yr)NoneNoneNoneApril 2002April 2003

 Cervical cancer screening (age 35–70 yr)NoneNoneNoneApril 2002April 2003
Open in a separate window*Community health centres and fee-for-service practices did not receive productivity or quality incentives. No model offered incentives for screening of visual or auditory impairment.Late in 2004, the Ontario Ministry of Health and Long-term Care created a reformed fee-for-service model — the family health group — to which fee-for-service practices could transition. We combined these two fee-for-service models for our analyses.Incentives for service enhancement of preventive manoeuvres, available through the Ministry of Health and Long-Term Care for the study period. Dates when the incentive bonuses came into effect are indicated in the cells. Incentives cover care delivered during the 30 months before the date the incentives became effective.Source: Adapted from the Ontario Medical Association document comparing models (www.oma.org/Member/Resources/Documents/2008PCRComparisonChart.pdf), and supplemented with other information found on the Ontario Medical Association website.We conducted this study to compare the delivery of preventive services by practices in the four funding models and to identify organizational factors associated with superior preventive care. This study is part of a larger evaluation of primary care models in Ontario funded by the Ontario Ministry of Health and Long-Term Care through its Primary Health Care Transition Fund.  相似文献   

10.
Widespread Distribution of Cell Defense against d-Aminoacyl-tRNAs     
Sandra Wydau  Guillaume van der Rest  Caroline Aubard  Pierre Plateau    Sylvain Blanquet 《The Journal of biological chemistry》2009,284(21):14096-14104
Several l-aminoacyl-tRNA synthetases can transfer a d-amino acid onto their cognate tRNA(s). This harmful reaction is counteracted by the enzyme d-aminoacyl-tRNA deacylase. Two distinct deacylases were already identified in bacteria (DTD1) and in archaea (DTD2), respectively. Evidence was given that DTD1 homologs also exist in nearly all eukaryotes, whereas DTD2 homologs occur in plants. On the other hand, several bacteria, including most cyanobacteria, lack genes encoding a DTD1 homolog. Here we show that Synechocystis sp. PCC6803 produces a third type of deacylase (DTD3). Inactivation of the corresponding gene (dtd3) renders the growth of Synechocystis sp. hypersensitive to the presence of d-tyrosine. Based on the available genomes, DTD3-like proteins are predicted to occur in all cyanobacteria. Moreover, one or several dtd3-like genes can be recognized in all cellular types, arguing in favor of the nearubiquity of an enzymatic function involved in the defense of translational systems against invasion by d-amino acids.Although they are detected in various living organisms (reviewed in Ref. 1), d-amino acids are thought not to be incorporated into proteins, because of the stereospecificity of aminoacyl-tRNA synthetases and of the translational machinery, including EF-Tu and the ribosome (2). However, the discrimination between l- and d-amino acids by aminoacyl-tRNA synthetases is not equal to 100%. Significant d-aminoacylation of their cognate tRNAs by Escherichia coli tyrosyl-, tryptophanyl-, aspartyl-, lysyl-, and histidyl-tRNA synthetases has been characterized in vitro (39). Recently, using a bacterium, transfer of d-tyrosine onto tRNATyr was shown to occur in vivo (10).With such misacylation reactions, the resulting d-aminoacyl-tRNAs form a pool of metabolically inactive molecules, at best. At worst, d-aminoacylated tRNAs infiltrate the protein synthesis machinery. Although the latter harmful possibility has not yet been firmly established, several cells were shown to possess a d-tyrosyl-tRNA deacylase, or DTD, that should help them counteract the accumulation of d-aminoacyl-tRNAs. This enzyme shows a broad specificity, being able to remove various d-aminoacyl moieties from the 3′-end of a tRNA (46, 11). Such a function makes the deacylase a member of the family of enzymes capable of editing in trans mis-aminoacylated tRNAs. This family includes several homologs of aminoacyl-tRNA synthetase editing domains (12), as well as peptidyl-tRNA hydrolase (13, 14).Two distinct deacylases have already been discovered. The first one, called DTD1, is predicted to occur in most bacteria and eukaryotes (see d-amino acids, including d-tyrosine (6). In fact, in an E. coli Δdtd strain grown in the presence of 2.4 mm d-tyrosine, as much as 40% of the cellular tRNATyr pool becomes esterified with d-tyrosine (10).

TABLE 1

Distribution of DTD1 and DTD2 homologs in various phylogenetic groupsHomologs of DTD1 and DTD2 were searched for using a genomic Blast analysis against complete genomes in the NCBI Database (www.ncbi.nlm.nih.gov). Values in the table are number of species. For instance, E. coli is counted only once in γ-proteobacteria despite the fact that several E. coli strains have been sequenced.
DTD1DTD2DTD1 + DTD2None
Bacteria
    Acidobacteria 2 0 0 0
    Actinobacteria 27 0 0 8
    Aquificae 1 0 0 0
    Bacteroidetes/Chlorobi 12 0 0 5
    Chlamydiae 1 0 0 6
    Chloroflexi 4 0 0 0
    Cyanobacteria 5 0 0 16
    Deinococcus/Thermus 4 0 0 0
    Firmicutes
        Bacillales 19 0 0 0
        Clostridia 19 0 0 0
        Lactobacillales 23 0 0 0
        Mollicutes 0 0 0 15
    Fusobacteria/Planctomycetes 2 0 0 0
    Proteobacteria
        α 6 0 0 55
        β 24 0 0 11
        γ 80 0 0 8
        δ 15 0 0 0
        ε 1 0 0 12
    Spirochaetes 0 0 0 7
    Thermotogae 5 0 0 0
Archaea
    Crenarchaeota 0 13 0 0
    Euryarchaeota 1 26 0 2
    Nanoarchaeota 0 0 0 1
Eukaryota
    Dictyosteliida 1 0 0 0
    Fungi/Metazoa
        Fungi 13 0 0 1
        Metazoa 19 0 0 0
    Kinetoplastida 3 0 0 0
    Viridiplantae 4 4 4 0
Open in a separate windowHomologs of dtd/DTD1 are not found in the available archaeal genomes except that of Methanosphaera stadtmanae. A search for deacylase activity in Sulfolobus solfataricus and Pyrococcus abyssi led to the detection of another enzyme (DTD2), completely different from the DTD1 protein (15). Importing dtd2 into E. coli functionally compensates for dtd deprivation. As shown in 16).Several cells contain neither dtd nor dtd2 homologs (d-tyrosyl-tRNA deacylase (DTD3). This protein, encoded by dtd3, behaves as a metalloenzyme. Sensitivity of the growth of Synechocystis to external d-tyrosine is strongly exacerbated by the disruption of dtd3. Moreover, expression of the Synechocystis DTD3 in a Δdtd E. coli strain, from a plasmid, restores the resistance of the bacterium to d-tyrosine. Finally, using the available genomes, we examined the occurrence of DTD3 in the living world. The prevalence of DTD3-like proteins is surprisingly high. It suggests that the defense of protein synthesis against d-amino acids is universal.  相似文献   

11.
Genome-Wide Characterization of the SloR Metalloregulome in Streptococcus mutans     
Kevin P. O'Rourke  Jeremy D. Shaw  Mitchell W. Pesesky  Brian T. Cook  Susanne M. Roberts  Jeffrey P. Bond  Grace A. Spatafora 《Journal of bacteriology》2010,192(5):1433-1443
  相似文献   

12.
Sequences from Ancestral Single-Stranded DNA Viruses in Vertebrate Genomes: the Parvoviridae and Circoviridae Are More than 40 to 50 Million Years Old     
Vladimir A. Belyi  Arnold J. Levine  Anna Marie Skalka 《Journal of virology》2010,84(23):12458-12462
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   

13.
Hepatitis C Virus (HCV) Sequence Variation Induces an HCV-Specific T-Cell Phenotype Analogous to Spontaneous Resolution     
Victoria Kasprowicz  Yu-Hoi Kang  Michaela Lucas  Julian Schulze zur Wiesch  Thomas Kuntzen  Vicki Fleming  Brian E. Nolan  Steven Longworth  Andrew Berical  Bertram Bengsch  Robert Thimme  Lia Lewis-Ximenez  Todd M. Allen  Arthur Y. Kim  Paul Klenerman  Georg M. Lauer 《Journal of virology》2010,84(3):1656-1663
Hepatitis C virus (HCV)-specific CD8+ T cells in persistent HCV infection are low in frequency and paradoxically show a phenotype associated with controlled infections, expressing the memory marker CD127. We addressed to what extent this phenotype is dependent on the presence of cognate antigen. We analyzed virus-specific responses in acute and chronic HCV infections and sequenced autologous virus. We show that CD127 expression is associated with decreased antigenic stimulation after either viral clearance or viral variation. Our data indicate that most CD8 T-cell responses in chronic HCV infection do not target the circulating virus and that the appearance of HCV-specific CD127+ T cells is driven by viral variation.Hepatitis C virus (HCV) persists in the majority of acutely infected individuals, potentially leading to chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma. The cellular immune response has been shown to play a significant role in viral control and protection from liver disease. Phenotypic and functional studies of virus-specific T cells have attempted to define the determinants of a successful versus an unsuccessful T-cell response in viral infections (10). So far these studies have failed to identify consistent distinguishing features between a T-cell response that results in self-limiting versus chronic HCV infection; similarly, the impact of viral persistence on HCV-specific memory T-cell formation is poorly understood.Interleukin-7 (IL-7) receptor alpha chain (CD127) is a key molecule associated with the maintenance of memory T-cell populations. Expression of CD127 on CD8 T cells is typically only observed when the respective antigen is controlled and in the presence of significant CD4+ T-cell help (9). Accordingly, cells specific for persistent viruses (e.g., HIV, cytomegalovirus [CMV], and Epstein-Barr virus [EBV]) have been shown to express low levels of CD127 (6, 12, 14) and to be dependent on antigen restimulation for their maintenance. In contrast, T cells specific for acute resolving virus infections, such as influenza virus, respiratory syncytial virus (RSV), hepatitis B virus (HBV), and vaccinia virus typically acquire expression of CD127 rapidly with the control of viremia (5, 12, 14). Results for HCV have been inconclusive. The expected increase in CD127 levels in acute resolving but not acute persisting infection has been found, while a substantial proportion of cells with high CD127 expression have been observed in long-established chronic infection (2). We tried to reconcile these observations by studying both subjects with acute and chronic HCV infection and identified the presence of antigen as the determinant of CD127 expression.Using HLA-peptide multimers we analyzed CD8+ HCV-specific T-cell responses and CD127 expression levels in acute and chronic HCV infection. We assessed a cohort of 18 chronically infected subjects as well as 9 individuals with previously resolved infection. In addition, we longitudinally studied 9 acutely infected subjects (5 individuals who resolved infection spontaneously and 4 individuals who remain chronically infected) (Tables (Tables11 and and2).2). Informed consent in writing was obtained from each patient, and the study protocol conformed to the ethical guidelines of the 1975 Declaration of Helsinki, as reflected in a priori approval from the local institutional review boards. HLA-multimeric complexes were obtained commercially from Proimmune (Oxford, United Kingdom) and Beckman Coulter (CA). The staining and analysis procedure was as described previously (10). Peripheral blood mononuclear cells (PBMCs) were stained with the following antibodies: CD3 from Caltag; CD8, CD27, CCR7, CD127, and CD38 from BD Pharmingen; and PD-1 (kindly provided by Gordon Freeman). Primer sets were designed for different genotypes based on alignments of all available sequences from the public HCV database (http://hcvpub.ibcp.fr). Sequence analysis was performed as previously described (8).

TABLE 1.

Patient information and autologous sequence analysis for patients with chronic and resolved HCV infection
CodeGenotypeStatusEpitope(s) targetedSequencea
02-031bChronicA1 NS3 1436-1444P: ATDALMTGY
A: no sequence
00-261bChronicA1 NS3 1436-1444P: ATDALMTGY
A: no sequence
99-242aChronicA2 NS3 1073-1083P: CINGVCWTV
No recognitionA: S-S--L---
A2 NS3 1406-1415P: KLVALGINAV
No recognitionA: A-RGM-L---
A2 NS5B 2594-2602P: ALYDVVTKL
A: no sequence
1111aChronicA2 NS3 1073-1083P: CINGVCWTV
A: ---------
A2 NS5 2594-2602P: ALYDVVTKL
A: ---------
00X3aChronicA2 NS5 2594-2602P: ALYDVVTKL
No recognitionA: -----IQ--
O3Qb1aChronicA1 NS3 1436-1444P: ATDALMTGY
DiminishedA: --------F
03Sb1aChronicA1 NS3 1436-1444P: ATDALMTGY
DiminishedA: --------F
02A1aChronicA1 NS3 1436-1444P: ATDALMTGY
A: no sequence
01N1aChronicA1 NS3 1436-1444P: ATDALMTGY
DiminishedA: --------F
03H1aChronicA2 NS3 1073-1083P: CINGVCWTV
Full recognitionA: ----A----
01-391aChronicA1 NS3 1436-1444P: ATDALMTGY
DiminishedA: --------F
03-45b1aChronicA1 NS3 1436-1444P: ATDALMTGY
DiminishedA: --------F
06P3aChronicA1 NS3 1436-1444P: ATDALMTGY
DiminishedA: --------F
GS127-11aChronicA2 NS3 1073-1083P: CINGVCWTV
A: ---------
GS127-61aChronicA2 NS3 1073-1083P: CINGVCWTV
A: ---------
GS127-81bChronicA2 NS3 1073-1083P: CINGVCWTV
A: ---------
GS127-161aChronicA2 NS3 1073-1083P: CINGVCWTV
A: ---------
GS127-201aChronicA2 NS3 1073-1083P: CINGVCWTV
A: ---------
04D4ResolvedA2 NS5 1987-1996P: VLSDFKTWKL
01-49b1ResolvedA2 NS5 1987-1996P: VLSDFKTWKL
A2 NS3 1406-1415P: KLVALGINAV
01-311ResolvedA1 NS3 1436-1444P: ATDALMTGY
B57 NS5 2629-2637P: KSKKTPMGF
04N1ResolvedA1 NS3 1436-1444P: ATDALMTGY
01E4ResolvedA2 NS5 1987-1996P: VLSDFKTWKL
98A1ResolvedA2 NS3 1073-1083P: CINGVCWTV
00-10c1ResolvedA24 NS4 1745-1754P: VIAPAVQTNW
O2Z1ResolvedA1 NS3 1436-1444P: ATDALMTGY
99-211ResolvedB7 CORE 41-49P: GPRLGVRAT
OOR1ResolvedB35 NS3 1359-1367P: HPNIEEVAL
Open in a separate windowaP, prototype; A, autologous. Identical residues are shown by dashes.bHIV coinfection.cHBV coinfection.

TABLE 2.

Patient information and autologous sequence analysis for patients with acute HCV infection
CodeGenotypeOutcomeEpitope targeted and time analyzedSequencea
5541aPersistingA2 NS3 1073-1083P: CINGVCWTV
wk 8A: ---------
wk 30A: ---------
03-321aPersistingB35 NS3 1359-1367P: HPNIEEVAL
wk 8A: ---------
No recognition (wk 36)A: S--------
04-111a (1st)Persisting (1st) Resolving (2nd)A2 NS5 2594-2602P: ALYDVVTKL
1b (2nd)A: no sequence
00231bPersistingA1 NS3 1436-1444P: ATDALMTGY
Diminished (wk 7)A: --------F
Diminished (wk 38)A: --------F
A2 NS3 1073-1083P: CINGVCWTV
wk 7A: ---------
wk 38A: ---------
A2 NS3 1406-1415P: KLVALGINAV
Full recognition (wk 7)A: --S-------
Full recognition (wk 38)A: --S-------
3201ResolvingA2 NS3 1273-1282P: GIDPNIRTGV
5991ResolvingA2 NS3 1073-1083P: CINGVCWTV
11441ResolvingA2 NS3 1073-1083P: CINGVCWTV
B35 NS3 1359-1367P: HPNIEEVAL
06L3aResolvingB7 CORE 41-49P: GPRLGVRAT
05Y1ResolvingA2 NS3 1073-1083P: CINGVCWTV
Open in a separate windowaP, prototype; A, autologous. Identical residues are shown by dashes.In established persistent infection, CD8+ T-cell responses against HCV are infrequently detected in blood using major histocompatibility complex (MHC) class I tetramers and are only observed in a small fraction of those sampled (10). We were able to examine the expression of CD127 on antigen-specific T cells in such a group of 18 individuals. We observed mostly high levels of CD127 expression (median, 66%) on these populations (Fig. (Fig.1a),1a), although expression was higher on HCV-specific T-cell populations from individuals with resolved infection (median, 97%; P = 0.0003) (Fig. 1a and c). Importantly, chronically infected individuals displayed CD127 expression levels over a much broader range than resolved individuals (9.5% to 100% versus 92 to 100%) (Fig. (Fig.1a1a).Open in a separate windowFIG. 1.Chronically infected individuals express a range of CD127 levels on HCV-specific T cells. (a) CD127 expression levels on HCV-specific T-cell populations in individuals with established chronic or resolved infection. While individuals with resolved infection (11 tetramer stains in 9 subjects) uniformly express high levels of CD127, chronically infected individuals (21 tetramer stains in 18 subjects) express a wide range of CD127 expression levels. (b) CD127 expression levels are seen to be highly dependent on sequence match with the autologous virus, based on analysis of 9 responses with diminished recognition of the autologous virus and 8 responses with intact epitopes. (c) CD127 expression levels on HCV-specific T-cell B7 CORE 41-49-specific T cells from individual 01-49 with resolved HCV infection (left-hand panel). Lower CD127 expression levels are observed on an EBV-specific T-cell population from the same individual (right-hand panel). APC-A, allophycocyanin-conjugated antibody. (d) Low CD127 levels are observed on A2 NS3 1073-1083 HCV-specific T cells from individual 111 with chronic HCV infection in whom sequencing revealed an intact autologous sequence.Given the relationship between CD127 expression and antigenic stimulation as well as the potential of HCV to escape the CD8 T-cell response through viral mutation, we sequenced the autologous circulating virus in subjects with chronic infection (Table (Table1).1). A perfect match between the optimal epitope sequence and the autologous virus was found for only 8 responses. These were the only T-cell populations with lower levels of CD127 expression (Fig. (Fig.1a,1a, b, and d). In contrast, HCV T-cell responses with CD127 expression levels comparable to those observed in resolved infection (>85%) were typically mismatched with the viral sequence, with some variants compatible with viral escape and others suggesting infection with a non-genotype 1 strain (10) (Fig. (Fig.1).1). Enzyme-linked immunospot (ELISPOT) assays using T-cell lines confirmed the complete abrogation of T-cell recognition and thus antigenic stimulation in cases of cross-genotype mismatch (10). Responses targeting the epitope A1-143D expressed somewhat lower levels of CD127 (between 70% and 85%). Viral escape (Y to F at position 9) in this epitope has been shown to be associated with significantly diminished but not fully abolished recognition (11a), and was found in all chronically infected subjects whose T cells targeted this epitope. Thus, expression of CD127 in the presence of viremia is closely associated with the capacity of the T cell to recognize the circulating virus.That a decrease in antigenic stimulation is indeed associated with the emergence of CD127-expressing CD8 T cells is further demonstrated in subject 111. This subject with chronic infection targeted fully conserved epitopes with T cells with low CD127 expression; with clearance of viremia under antiviral therapy, CD127-negative HCV-specific CD8 T cells were no longer detectable and were replaced by populations expressing CD127 (data not shown). Overall these data support the notion that CD127 expression on HCV-specific CD8+ T-cell populations is dependent on an absence of ongoing antigenic stimulation.To further evaluate the dynamic relationship between antigenic stimulation and CD127 expression, we also analyzed HCV-specific T-cell responses longitudinally during acute HCV infection (Fig. (Fig.2a).2a). CD127 expression was generally low or absent during the earliest time points. After resolution of infection, we see a contraction of the HCV-specific T-cell response together with a continuous increase in CD127 expression, until virtually all tetramer-positive cells express CD127 approximately 6 months after the onset of disease (Fig. (Fig.2a).2a). A similar increase in CD127 expression was not seen in one subject (no. 554) with untreated persisting infection that maintained a significant tetramer-positive T-cell population for an extended period of time (Fig. (Fig.2a).2a). Importantly, sequence analysis of the autologous virus demonstrated the conservation of this epitope throughout persistent infection (8). In contrast, subject 03-32 (with untreated persisting infection) developed a CD8 T-cell response targeting a B35-restricted epitope in NS3 from which the virus escaped (8). The T cells specific for this epitope acquired CD127 expression in a comparable manner to those controlling infection (Fig. (Fig.2a).2a). In other subjects with persisting infection, HCV-specific T-cells usually disappeared from blood before the time frame in which CD127 upregulation was observed in the other subjects.Open in a separate windowFIG. 2.CD127 expression levels during acute HCV infection. (a) CD127 expression levels on HCV-specific T cells during the acute phase of HCV infection (data shown for 5 individuals who resolve and two individuals who remain chronically infected). (b) HCV RNA viral load and CD127 expression levels on HCV-specific T cells (A2 NS3 1073-1083 and A1 NS3 1436-1444) for chronically infected individual 00-23. PEG-IFN-α, pegylated alpha interferon. (c) Fluorescence-activated cell sorter (FACS) plots showing longitudinal CD127 expression levels on HCV-specific T cells (A2 NS3 1073-1083 and A1 NS3 1436-1444) from individual 00-23.We also characterized the levels of CD127 expression on HCV-specific CD4+ T-cell populations with similar results: low levels were observed during the acute phase of infection and increased levels in individuals after infection was cleared (data not shown). CD127 expression on CD4 T cells could not be assessed in viral persistence since we failed to detect significant numbers of HCV-specific CD4+ T cells, in agreement with other reports.In our cohort of subjects with acute HCV infection, we had the opportunity to study the effect of reencounter with antigen on T cells with high CD127 expression in 3 subjects in whom HCV viremia returned after a period of viral control. Subject 00-23 experienced viral relapse after interferon treatment (11), while subjects 05-13 and 04-11 were reinfected with distinct viral isolates. In all subjects, reappearance of HCV antigen that corresponded to the HCV-specific T-cell population was associated with massive expansion of HCV-specific T-cell populations and a decrease in CD127 expression on these T cells (Fig. (Fig.22 and and3)3) (data not shown). In contrast, T-cell responses that did not recognize the current viral isolate did not respond with an expansion of the population or the downregulation of CD127. This was observed in 00-23, where the sequence of the A1-restricted epitope 143D was identical to the frequent escape mutation described above in chronically infected subjects associated with diminished T-cell recognition (Fig. (Fig.2b2b and and3a).3a). In 05-13, the viral isolate during the second episode of viremia contained a variant in one of the anchor residues of the epitope A2-61 (Fig. (Fig.2d).2d). These results show that CD127 expression on HCV-specific T cells follows the established principles observed in other viral infections.Open in a separate windowFIG. 3.Longitudinal phenotypic changes on HCV-specific T cells. (a) HCV RNA viral load and CD127 expression (%) levels on A2 NS5B 2594-2602 HCV-specific T cells for individual 04-11. This individual was administered antiviral therapy, which resulted in a sustained virological response. Following reinfection, the individual spontaneously cleared the virus. (b) Longitudinal frequency of A2 NS5B 2594-2602 HCV-specific T cells and PD-1 expression levels (mean fluorescent intensity [MFI]) for individual 04-11. (c) Longitudinal analysis of 04-11 reveals the progressive differentiation of HCV-specific A2 259F CD8+ T cells following repetitive antigenic stimulation. FACS plots show longitudinal CD127, CD27, CD57, and CCR7 expression levels on A2 NS5B 2594-2602 tetramer-positive cells from individual 04-11. PE-A, phycoerthrin-conjugated antibody.In addition to the changes in CD127 expression for T cells during reencounter with antigen, we detected comparable changes in other phenotypic markers shortly after exposure to viremia. First, we detected an increase in PD-1 and CD38 expression—both associated with recent T-cell activation. Additionally, we observed a loss of CD27 expression, a feature of repetitive antigenic stimulation (Fig. (Fig.3).3). The correlation of CD127 and CD27 expression further supports the notion that CD127 downregulation is a marker of continuous antigenic stimulation (1, 7).In conclusion we confirm that high CD127 expression levels are common for detectable HCV-specific CD8+ T-cell populations in chronic infection and find that this phenotype is based on the existence of viral sequence variants rather than on unique properties of HCV-specific T cells. This is further demonstrated by our data from acute HCV infection showing that viral escape as well as viral resolution is driving the upregulation of CD127. We also show that some, but not all, markers typically used to phenotypically describe virus-specific T cells show a similar dependence on cognate HCV antigen. Our data further highlight that sequencing of autologous virus is vital when interpreting data obtained in chronic HCV infection and raise the possibility that previous studies, focused on individuals with established chronic infection, may have been confounded by antigenic variation within epitopes or superinfection with different non-cross-reactive genotypes. Interestingly, it should be pointed out that this finding is supported by previous data from both the chimpanzee model of HCV and from human HBV infection (3, 13).Overall our data clearly demonstrate that the phenotype of HCV-specific CD8+ T cells is determined by the level of antigen-specific stimulation. The high number of CD127 positive virus-specific CD8+ T cells that is associated with the presence of viral escape mutations is a hallmark of chronic HCV infection that clearly separates HCV from other chronic viral infections (4, 14).  相似文献   

14.
Escherichia coli O157:H7 and Other E. coli Strains Share Physiological Properties Associated with Intestinal Colonization     
Lisa Jacobsen  Lisa Durso  Tyrell Conway  Kenneth W. Nickerson 《Applied and environmental microbiology》2009,75(13):4633-4635
Escherichia coli isolates (72 commensal and 10 O157:H7 isolates) were compared with regard to physiological and growth parameters related to their ability to survive and persist in the gastrointestinal tract and found to be similar. We propose that nonhuman hosts in E. coli O157:H7 strains function similarly to other E. coli strains in regard to attributes relevant to gastrointestinal colonization.Escherichia coli is well known for its ecological versatility (15). A life cycle which includes both gastrointestinal and environmental stages has been stressed by both Savageau (15) and Adamowicz et al. (1). The gastrointestinal stage would be subjected to acid and detergent stress. The environmental stage is implicit in E. coli having transport systems for fungal siderophores (4) as well as pyrroloquinoline quinone-dependent periplasmic glucose utilization (1) because their presence indicates evolution in a location containing fungal siderophores and pyrroloquinoline quinone (1).Since its recognition as a food-borne pathogen, there have been numerous outbreaks of food-borne infection due to E. coli O157:H7, in both ground beef and vegetable crops (6, 13). Cattle are widely considered to be the primary reservoir of E. coli O157:H7 (14), but E. coli O157:H7 does not appear to cause disease in cattle. To what extent is E. coli O157:H7 physiologically unique compared to the other naturally occurring E. coli strains? We feel that the uniqueness of E. coli O157:H7 should be evaluated against a backdrop of other wild-type E. coli strains, and in this regard, we chose the 72-strain ECOR reference collection originally described by Ochman and Selander (10). These strains were chosen from a collection of 2,600 E. coli isolates to provide diversity with regard to host species, geographical distribution, and electromorph profiles at 11 enzyme loci (10).In our study we compared the 72 strains of the ECOR collection against 10 strains of E. coli O157:H7 and six strains of E. coli which had been in laboratory use for many years (Table (Table1).1). The in vitro comparisons were made with regard to factors potentially relevant to the bacteria''s ability to colonize animal guts, i.e., acid tolerance, detergent tolerance, and the presence of the Entner-Doudoroff (ED) pathway (Table (Table2).2). Our longstanding interest in the ED pathway (11) derives in part from work by Paul Cohen''s group (16, 17) showing that the ED pathway is important for E. coli colonization of the mouse large intestine. Growth was assessed by replica plating 88 strains of E. coli under 40 conditions (Table (Table2).2). These included two LB controls (aerobic and anaerobic), 14 for detergent stress (sodium dodecyl sulfate [SDS], hexadecyltrimethylammonium bromide [CTAB], and benzalkonium chloride, both aerobic and anaerobic), 16 for acid stress (pH 6.5, 6.0, 5.0, 4.6, 4.3, 4.2, 4.1, and 4.0), four for the ability to grow in a defined minimal medium (M63 glucose salts with and without thiamine), and four for the presence or absence of a functional ED pathway (M63 with gluconate or glucuronate). All tests were done with duplicate plates in two or three separate trials. The data are available in Tables S1 to S14 in the supplemental material, and they are summarized in Table Table22.

TABLE 1.

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

TABLE 2.

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

15.
Detection and Identification of tdh- and trh-Positive Vibrio parahaemolyticus Strains from Four Species of Cultured Bivalve Molluscs on the Spanish Mediterranean Coast     
Ana Roque  Carmen Lopez-Joven  Beatriz Lacuesta  Laurence Elandaloussi  Sariqa Wagley  M. Dolores Furones  Imanol Ruiz-Zarzuela  Ignacio de Blas  Rachel Rangdale  Bruno Gomez-Gil 《Applied and environmental microbiology》2009,75(23):7574-7577
Presented here is the first report describing the detection of potentially diarrheal Vibrio parahaemolyticus strains isolated from cultured bivalves on the Mediterranean coast, providing data on the presence of both tdh- and trh-positive isolates. Potentially diarrheal V. parahaemolyticus strains were isolated from four species of bivalves collected from both bays of the Ebro delta, Spain.Gastroenteritis caused by Vibrio parahaemolyticus has been reported worldwide, though only sporadic cases have been reported in Europe (7, 14). The bacterium can be naturally present in seafood, but pathogenic isolates capable of inducing gastroenteritis in humans are rare in environmental samples (2 to 3%) (15) and are often not detected (10, 19, 20).The virulence of V. parahaemolyticus is based on the presence of a thermostable direct hemolysin (tdh) and/or the thermostable direct hemolysin-related gene (trh) (1, 5). Both are associated with gastrointestinal illnesses (2, 9).Spain is not only the second-largest producer in the world of live bivalve molluscs but also one of the largest consumers of bivalve molluscs, and Catalonia is the second-most important bivalve producer of the Spanish Autonomous Regions. Currently, the cultivation of bivalves in this area is concentrated in the delta region of the Ebro River. The risk of potentially pathogenic Vibrio spp. in products placed on the market is not assessed by existing legislative indices of food safety in the European Union, which emphasizes the need for a better knowledge of the prevalence of diarrheal vibrios in seafood products. The aim of this study was to investigate the distribution and pathogenic potential of V. parahaemolyticus in bivalve species exploited in the bays of the Ebro delta.Thirty animals of each species of Mytilus galloprovincialis, Crassostrea gigas, Ruditapes decussatus, and Ruditapes philippinarum were collected. They were sampled from six sites of the culture area, three in each bay of the Ebro River delta, at the beginning (40°37′112"N, 0°37′092"E [Alfacs]; 40°46′723"N, 0°43′943"E [Fangar]), middle (40°37′125"N, 0°38′570"E [Alfacs]; 40°46′666"N, 0°45′855"E [Fangar]), and end (40°37′309"N, 0°39′934"E [Alfacs]; 40°46′338"N, 0°44′941"E [Fangar]) of the culture polygon. Clams were sampled from only one site per bay as follows: in the Alfacs Bay from a natural bed of R. decussatus (40°37′44"N, 0°38′0"E) and in the Fangar Bay from an aquaculture bed of R. philippinarum (40°47′3"N, 0°43′8"E). In total, 367 samples were analyzed in 2006 (180 oysters, 127 mussels, 30 carpet shell clams, and 30 Manila clams) and 417 samples were analyzed in 2008 (178 oysters, 179 mussels, 30 carpet shell clams, and 30 Manila clams).All animals were individually processed and homogenized, and 1 ml of the homogenate was inoculated into 9 ml of alkaline peptone water (Scharlau, Spain). Following a 6-h incubation at 37°C, one loopful of the contents of each tube of alkaline peptone water was streaked onto CHROMagar vibrio plates (CHROMagar, France) and incubated for 18 h at 37°C. Mauve-purple colonies were purified, and each purified isolate was cryopreserved at −80°C (135 isolates in 2006 and 96 in 2008). From the initial homogenate portion, 100 μl was inoculated onto marine agar (Scharlau, Spain) and onto thiosulfate citrate-bile salts-sucrose agar (Scharlau, Spain) for total heterotrophic marine bacteria counts and total vibrio counts, respectively (Table (Table11).

TABLE 1.

Vibrio parahaemolyticus isolates, serotypes, and origins and total number of vibrios/heterotrophic bacteria contained in the bivalvea
IsolateDate of collectionOrganism and site of originTemp (°C)Salinity (‰)Gene(s)SerotypeBacterial count using indicated medium (CFU ml−1)
TCBS agarMarine agar
I7458 August 2006Mg-F24.537tdhND1.5 × 1041.2 × 104
I79314 August 2006Cg-A2535tdhND9.2 × 1028.5 × 103
I80514 August 2006Cg-A2535tdhO2:KUT7.2 × 1029 × 103
I80614 August 2006Cg-A2535tdh and trhO3:K331.9 × 1034.6 × 103
I80914 August 2006Cg-A2535tdhO2:K288 × 1047.3 × 102
I6784 July 2006Rd-A28.636tdhO2:K283.1 × 1052.5 × 105
I6284 July 2006Rd-A28.636tdhO4:KUT2.9 × 1048.4 × 104
I7758 August 2006Cg-A24.537tdhND4.21 × 1031.1 × 104
I6914 July 2006Rd-A28.636trhO1:K322.2 × 1052.6 × 105
I71227 July 2006Mg-A29.435.5trhO1:KUT8.6 × 1038.4 × 103
I7658 August 2006Cg-F24.537trhO4:K341 × 104Uncountable
I98022 July 2008Cg-A26.733.5tdhO1:K322.7 × 1041.3 × 104
I98122 July 2008Cg-A26.733.5trhO1:KUT1 × 1042.2 × 104
I99322 July 2008Cg-A26.733.5tdhO5:K173 × 1031.1 × 104
I99429 July 2008Mg-A27.737trhO3:KUT3.4 × 1037 × 103
I10315 August 2008Cg-F27.737tdhO5:KUT5.5 × 1043.3 × 104
I10345 August 2008Cg-F27.737tdhO3:KUT8.7 × 1044 × 104
I10405 August 2008Cg-F27.737tdhO3:KUT1.6 × 1043.2 × 104
I10425 August 2008Cg-F27.737tdh and trhND2.8 ×1043 × 104
I10505 August 2008Cg-F27.737tdhO1:KUT4.7 × 1047.3 × 104
I106320 August 2008Mg-F25.936tdhO3:KUT7.9 ×1041.4 × 104
I106520 August 2008Mg-F25.936tdhO2:KUT2.2 × 1031.2 × 104
I106820 August 2008Mg-F25.936tdhO5:KUT2.6 × 1045.2 × 104
I106920 August 2008Mg-F25.936tdhO3:KUT2.4 × 1035.3 × 104
I107320 August 2008Mg-F25.936tdhO5:KUT2.3 × 1037.5 × 103
I107420 August 2008Mg-F25.936tdhO3:KUT7.6 × 1046.9 × 104
I107720 August 2008Mg-F25.936tdhO4:KUT1.7 × 1031.6 × 103
I107920 August 2008Mg-F25.936trhO3:KUT2.5 × 1031.1 × 104
I109220 August 2008Mg-F25.936tdhND1.7 × 1031.6 × 103
I113025 August 2008Rd-A26.435tdhND1.7 × 1043.8 × 104
I114325 August 2008Rd-A26.435tdhND1.1 × 1041.9 × 104
I116525 August 2008Rd-A26.435trhO2:KUT4.4 × 1046.8 × 104
I113325 August 2008Rp-F25.536.5tdhND3.4 × 1044 × 104
I113425 August 2008Rp-F25.536.5tdhND3.9 × 1045.8 × 104
I115825 August 2008Rp-F25.536.5trhO4:KUT6.6 × 1044.7 × 104
I116125 August 2008Rp-F25.536.5trhO3:KUT2.2 × 1046.6 × 104
Open in a separate windowaMg, Mytilus galloprovincialis; Cg, Crassostrea gigas; Rd, Ruditapes decussatus; Rp, R. phillipinarum; A, Alfacs; F, Fangar; ND, not determined; TCBS, thiosulfate citrate-bile salts-sucrose.Total DNA was extracted from each purified isolate using the Wizard genomic DNA purification kit (Promega), following the instructions of the manufacturer. A one-step PCR analysis was performed to identify/confirm which isolates were tl positive (species marker for V. parahaemolyticus). Further detection of the tdh or trh gene was carried out on all positive tl strains. All PCR analyses were carried out using the primers described by Bej et al. (2) with the following amplification conditions on the thermocycler (Eppendorf Mastercycler Personal): an initial denaturation at 95°C for 8 min, followed by 40 cycles of a 1-min denaturation at 94°C, annealing at 55°C for 1 min, elongation at 72° for 1 min, and a final extension of 10 min at 72°C. Positive and negative controls were included in all reaction mixtures: two positive controls, tl and tdh CAIM 1400 and trh CAIM 1772 (Collection of Aquatic Important Microorganisms [http://www.ciad.mx/caim/CAIM.html]), and negative control DNA-free molecular grade water (Sigma-Aldrich, Spain). Expected amplicons were visualized in 2% agarose gels stained with ethidium bromide.Fifty-eight isolates contained the gene tl in 2006 and 96 in 2008, which confirmed their identity as V. parahaemolyticus. In 2006, the distribution of the 58 isolates was as follows: 7 from 127 mussels, 34 from 180 oysters, and 17 from 30 R. decussatus clams. No tl-positive isolates were found in R. philippinarum. PCR analysis of the tl-positive isolates for the presence of the tdh or trh gene indicated that eight isolates contained the tdh gene and four contained the trh gene. In 2008, the source of the confirmed V. parahaemolyticus isolates was as follows: 31 from 88 oysters, 44 from 89 mussels, 9 from 30 R. decussatus clams, and 12 from 30 R. philippinarum clams. Of these, 17 were found to contain the tdh gene and 7 contained the trh gene. Two isolates (I806 and I1042) contained both toxigenic genes, tdh and trh.Putative tdh- and trh-positive PCR products were purified using the QIAquick PCR purification kit (Qiagen) following the manufacturer''s instructions and were sequenced bidirectionally by Macrogen Inc. Sequences were aligned using BioEdit (8) and analyzed using BLAST (National Center for Biotechnology Information). None of the toxigenic isolates was found positive by PCR analysis for the presence of open reading frame 8 of the phage 237 (16), a marker for the pandemic strain O3:K6.The isolates were fingerprinted by repetitive extragenic palindromic PCR (rep-PCR) as described previously (3), and the resulting electrophoretic band patterns were analyzed with the GelCompar II software (v4.5; Applied Maths). The similarity matrix was calculated with the Jaccard coefficient with a band position tolerance of 0.8%, and the dendrogram was constructed with the Ward algorithm. A high level of genomic diversity was found among the 32 toxigenic isolates characterized by rep-PCR. Three clonal groups were identified (those having identical rep-PCR band patterns) (Fig. 1a to c).Open in a separate windowFIG. 1.rep-PCR dendrogram of toxigenic isolates of V. parahaemolyticus isolated in the Ebro delta. Letters denote clonal groups of isolates.In vitro antibiotic susceptibility tests were performed using the diffusion disc test following a previously described protocol (18). The antibiotics used were gentamicin (10 μg), oxolinic acid (10 μg), amoxicillin (25 μg), polymyxin B (300 UI), vancomycin (30 μg), trimethoprim sulfamethoxazole (1.25/23.75 μg), nitrofurantoin (300 μg), doxycyclin (30 μg), ceftazidime (30 μg), streptomycin (10 μg), neomycin (30 UI), penicillin (6 μg), flumequine (30 μg), tetracycline (30 μg), ampicillin (10 μg), kanamycin (30 μg), ciprofloxacin (5 μg), and sulfonamide (300 μg). All tests were performed in duplicate. A Student t test for two samples with unequal variance was performed to compare the sensitivity of all 2006 isolates against the sensitivity of 2008 isolates for each antibiotic (Microsoft Office Excel 97-2003). Antibiogram results revealed a lower susceptibility in 2008 than in 2006, indicating a possible shift in overall susceptibility. Results from the t test indicated that significantly lower susceptibility in 2008 was detected (P ≤ 0.05; n = 36) for the following antibiotics: vancomycin, polymyxin B, ampicillin, amoxicillin, gentamicin, neomycin, trimethoprim sulfamethoxazole, nitrofurantoin, doxycyclin, ceftazidime, tetracycline, flumequine, and ciprofloxacin.The serological types for 27 strains were determined by the agglutination method using commercially available V. parahaemolyticus antisera (Denka Seiken Ltd.; Cosmos Biomedical Ltd, United Kingdom) following the manufacturer''s instructions. Potentially toxigenic V. parahaemolyticus isolates collected in 2006 were serologically heterogeneous (8 out of the 11 isolates) (Table (Table1).1). In isolates collected in 2008, results were more homogenous, with seven serotypes found among 19 isolates analyzed. The O3:K6 serotype was not detected in any of the strains analyzed, in agreement with the open reading frame 8 PCR results.The present study is the first to report the detection of potentially diarrheal V. parahaemolyticus strains isolated from cultured bivalves on Spanish Mediterranean coasts, providing data on the presence of both tdh- and trh-positive isolates. V. parahaemolyticus has previously been detected in several European countries (4, 13, 21, 22). A recent study carried out in Spain detected tdh-positive V. parahaemolyticus strains from patients who had consumed fresh oysters in a market in Galicia on the Atlantic coast of Spain (12) and potentially pathogenic V. parahaemolyticus strains have also been reported in France (17). These studies indicate that the risk of infections caused by V. parahaemolyticus in Europe is low compared to that in America or Asia (15). However, this risk could have been underestimated, since V. parahaemolyticus is not included in the current European surveillance programs, such as the European Network for Epidemiological Surveillance and Control of Communicable Diseases.Toxigenic V. parahaemolyticus strains detected in this study were genomically and serologically heterogeneous. The pandemic serotype O3:K6 was not detected, and although attempts to isolate O3:K6 from the environment and from seafood have not always been successful in previous studies reviewed by Nair and coauthors (15), this finding seems to be in agreement with the fact that no outbreak of diarrhea was observed in the area. Interestingly, isolates I806 and I1042 have been found positive for both tdh and trh in PCR tests. The coexistence of tdh and trh genes has already been reported in isolates from Japan, the United States, and Mexico (3, 6, 11, 19, 23). To our knowledge, no occurrence of an environmental isolate positive for both tdh and trh had previously been reported in Europe. All isolates tested were slightly different in their antibiotic resistance profiles. Typically, a high level of resistance could be determined. The detection of tdh- and/or trh-positive V. parahaemolyticus strains for the first time on the Mediterranean coast emphasizes the need to monitor for the presence of potentially diarrheal vibrios and bacterial gastroenteritis, and these data should be taken into consideration to revise the European legislation on the requirements for shellfish harvested for consumption in order to include the surveillance of these pathogens in Europe.  相似文献   

16.
Allelic Variation in Cell Wall Candidate Genes Affecting Solid Wood Properties in Natural Populations and Land Races of Pinus radiata     
S. K. Dillon  M. Nolan  W. Li  C. Bell  H. X. Wu  S. G. Southerton 《Genetics》2010,185(4):1477-1487
  相似文献   

17.
Dominant Bacteria and Biomass in the Kuytun 51 Glacier     
Shu-Rong Xiang  Tian-Cui Shang  Yong Chen  Ze-Fan Jing  Tandong Yao 《Applied and environmental microbiology》2009,75(22):7287-7290
  相似文献   

18.
RNA-Based Investigation of Ammonia-Oxidizing Archaea in Hot Springs of Yunnan Province,China     
Hongchen Jiang  Qiuyuan Huang  Hailiang Dong  Peng Wang  Fengping Wang  Wenjun Li  Chuanlun Zhang 《Applied and environmental microbiology》2010,76(13):4538-4541
  相似文献   

19.
Stability of Murine Cytomegalovirus Genome after In Vitro and In Vivo Passage     
Tammy P. Cheng  Mark C. Valentine  Jian Gao  Jeanette T. Pingel  Wayne M. Yokoyama 《Journal of virology》2010,84(5):2623-2628
  相似文献   

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

TABLE 1.

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

TABLE 2.

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

TABLE 3.

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

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