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 62 毫秒
1.
2.
A 30-probe assay was developed for simultaneous classification of Listeria monocytogenes isolates by lineage (I to IV), major serogroup (4b, 1/2b, 1/2a, and 1/2c), and epidemic clone (EC) type (ECI, ECIa, ECII, and ECIII). The assay was designed to facilitate rapid strain characterization and the integration of subtype data into risk-based inspection programs.Listeria monocytogenes is a facultative intracellular pathogen that can cause serious invasive illness (listeriosis) in humans and other animals. L. monocytogenes is responsible for over 25% of food-borne-disease-related deaths attributable to known pathogens and is a leading cause of food recalls due to microbial adulteration (12, 21). However, not all L. monocytogenes subtypes contribute equally to human illness, and substantial differences in the ecologies and virulence attributes of different L. monocytogenes subtypes have been identified (9, 13, 14, 23, 24, 33, 35, 36). Among the four major evolutionary lineages of L. monocytogenes, only lineages I and II are commonly isolated from contaminated food and human listeriosis patients (19, 27, 29, 33). Lineage I strains are overrepresented among human listeriosis isolates, particularly those associated with epidemic outbreaks, whereas lineage II strains are overrepresented in foods and the environment (13, 14, 24). Lineage III strains account for approximately 1% of human listeriosis cases but are common among animal listeriosis isolates and appear to be a host-adapted group that is poorly adapted to food-processing environments (6, 34-36). The ecological and virulence attributes of lineage IV are poorly understood, as this lineage is rare and was only recently described based on a small number of strains (19, 26, 29, 33).L. monocytogenes is differentiated into 13 serotypes; however, four major serogroups (4b, 1/2b, 1/2a, and 1/2c) from within lineages I and II account for more than 98% of human and food isolates (16, 31). Serogroups refer to evolutionary complexes typified by a predominant serotype but which include very rare serotypes that represent minor evolutionary variants (7, 9, 33). Phylogenetic analyses have indicated that rare serotypes may have evolved recently, or even multiple times, from one of the major serotypes (9), and numerous molecular methods fail to discriminate minor serotypes as independent groups (1, 4, 7, 9, 18, 22, 33, 38, 39). Serotyping is one of the most common methods for L. monocytogenes subtyping, and serogroup classifications are a useful component of strain characterization because ecotype divisions appear largely congruent with serogroup distinctions (16, 34). Serogroup 4b strains are of particular public health concern because contamination with these strains appears to increase the probability that a ready-to-eat (RTE) food will be implicated in listeriosis (16, 28). Serogroup 4b strains account for approximately 40% of sporadic listeriosis and also are responsible for the majority of listeriosis outbreaks despite being relatively rare contaminants of food products (9, 13, 17, 30, 34). In addition, serogroup 4b strains are associated with more severe clinical presentations and higher mortality rates than other serogroups (11, 16, 20, 31, 34). Serogroups 1/2a and 1/2b are overrepresented among food isolates but also contribute significantly to human listeriosis, whereas serogroup 1/2c rarely causes human illness and may pose a lower risk of listeriosis for humans (16). Serogroup-specific differences in association with human listeriosis are consistent with the prevalence of virulence-attenuating mutations in inlA within these serogroups (32, 34); however, a number of additional factors likely contribute to these differences.Four previously described epidemic clones (ECs; ECI, ECIa, ECII, and ECIII) of L. monocytogenes have been implicated in numerous listeriosis outbreaks and have contributed significantly to sporadic illness (15, 34). ECI, ECIa, and ECII are distinct groups within serogroup 4b that were each responsible for repeated outbreaks of listeriosis in the United States and Europe. ECIII is a lineage II clone of serotype 1/2a that persisted in the same processing facility for more than a decade prior to causing a multistate outbreak linked to contaminated turkey (15, 25). While there has been speculation that epidemic clones possess unique adaptations that explain their frequent involvement in listeriosis outbreaks (9, 34, 37), it is not clear that epidemic clones are more virulent than other strains with the same serotype. However, contamination of RTE food with EC strains would be cause for increased concern due to the previous involvement of these clones in major outbreaks of listeriosis (16).As a result of the L. monocytogenes subtype-specific differences in ecology, virulence, and association with human illness, molecular subtyping technologies have the potential to inform assessments of relative risk and to improve risk-based inspection programs. The objective of the present study was to develop a single assay for rapid and accurate classification of L. monocytogenes isolates by lineage, major serogroup, and epidemic clone in order to facilitate strain characterization and the integration of subtype data into inspection programs that are based on assessment of relative risk.A database of more than 5.3 Mb of comparative DNA sequences from 238 L. monocytogenes isolates (9, 33-35) was scanned for single nucleotide polymorphisms that could be used to differentiate lineages, major serogroups, and epidemic clones via a targeted multilocus genotyping (TMLGT) approach. The acronym TMLGT is used to distinguish this approach from previously published multilocus genotyping (MLGT) assays that were lineage specific and designed for haplotype discrimination (9, 33). To provide for simultaneous interrogation of the selected polymorphisms via TMLGT, six genomic regions (Table (Table1)1) were coamplified in a multiplex PCR. While the previous MLGT assays were based on three lineage-specific multiplexes and required prior identification of lineage identity, TMLGT was designed to target variation across all of the lineages simultaneously and is based on a unique set of amplicons. PCR was performed in 50-μl volumes with 1× High Fidelity PCR buffer (Invitrogen Life Technologies), 2 mM MgSO4, 100 μM deoxynucleoside triphosphate (dNTP), 300 nM primer, 1.5 U Platinum Taq high-fidelity DNA polymerase (Invitrogen Life Technologies), and 100 ng of genomic DNA. PCR consisted of an initial denaturation of 90 s at 96°C, followed by 40 cycles of 30 s at 94°C, 30 s at 50°C, and 90 s at 68°C. Amplification products were purified using Montage PCR cleanup filter plates (Millipore) and served as a template for allele-specific primer extension (ASPE) reactions utilizing subtype-specific probes.

TABLE 1.

Primers used in multiplex amplification for the TMLGT assay
AmpliconPositionaGene(s)PrimerSequence (5′-3′)b
INLa455381-456505inlAinl2-a1GTCCTTGATAGTCTACTG
inl2-a2ACCAAATTAGTAATCTAGCAC
INLb457726-458752inlBinl-f1dGAATTRTTTAGYCAAGAATGT
inlb-rCTACCGGRACTTTATAGTAYG
LMO325116-326096lmo0298-lmo0300lmo-a1AAGGCTTACAAGATGGCT
lmo1a-1rAAATAATAYGTGATACCGAC
VGCa205366-206622plcA, hlyplca-fCTCATCGTATCRTGTGTACC
hly-rTCTGGAAGGTCKTGTAGGTTC
VGCb208447-209465mplra_mpl-fGTGGAYAGAACTCATAAAGG
ra_mpl-rACTCCCTCCTYGTGATASGCT
VGCc209728-211239actAvgc1a-2fTTCMATRCCAGCAGAACG
vgc1a-2rGCAGACCTAATAGCAATGTTG
Open in a separate windowaCorresponding nucleotide positions in the complete genome sequence of L. monocytogenes strain EGD-e (GenBank accession number NC_003210).bSee IUPAC codes for definition of degenerate bases.ASPE was performed in multiplex reactions including 30 probes, with each lineage (I to IV), major serogroup (4b, 1/2b, 1/2a, and 1/2c), and epidemic clone (ECI, ECIa, ECII, and ECIII) targeted by two different probes (Table (Table2).2). In addition, positive-control probes were included to confirm the presence of each amplicon in the multiplex PCR. As serogroups and epidemic clones are nested within a particular lineage, probes for these groups were designed to be specific within the appropriate lineage and values for these probes were evaluated only for isolates of the appropriate lineage. For example, serogroup 1/2a probes were evaluated only for isolates that were positive for lineage II probes. ASPE probes were designed with a unique 5′ sequence tag specific to individual sets of xMAP fluorescent polystyrene microspheres (Luminex Corporation) used to sort extension products. Extension and hybridization reactions were performed as described previously (9) except microspheres were twice pelleted by centrifugation (4 min at 2,250 × g) and resuspended in 75 μl 1× TM buffer prior to being pelleted and resuspended in 100 μl 1× TM buffer containing 2 μg/ml streptavidin-R-phycoerythrin (Invitrogen Life Technologies). Samples were incubated for 15 min at 37°C prior to detecting the microsphere complexes with a Luminex 100 flow cytometer (Luminex Corporation). The median fluorescence intensity (MFI) from biotinylated extension products attached to 100 microspheres was measured for each probe. The average MFI from three template-free control samples was also determined and subtracted from the raw MFI of each sample to account for background fluorescence. Probe performance was initially evaluated via the index of discrimination (ID) as described by Ducey et al. (9), and probes with ID values less than 2.0 were redesigned.

TABLE 2.

TMLGT probes and probe performance data
ProbebTarget (n)cProbe sequencedIDeSensitivity (%)Specificity (%)
VGCb-21Lineage I (506)AATCCTTTCTTTAATCTCAAATCAgcggaagcttgggaagcggtc7.3100100
VGCa-94Lineage ICTTTCTATCTTTCTACTCAATAATcaacccgatgttcttcctgtc51.7100100
VGCc-8Lineage II (340)AATCCTTTTACATTCATTACTTACattagctgattcgctttcct14.1100100
INLb-51Lineage IITCATTTCAATCAATCATCAACAATagcgccaataaagctggc21.9100100
VGCb-19Lineage III (50)TCAATCAATTACTTACTCAAATACccgctattaaaatgtactcca31.0100100
VGCb-29Lineage IIIAATCTTACTACAAATCCTTTCTTTggtataccgctattaaaatgt45.1100100
LMO-17Lineage IV (10)CTTTAATCCTTTATCACTTTATCAgaaccaaacaatgttattggt11.8100100
VGCa-27Lineage IVCTTTTCAAATCAATACTCAACTTTttaacgacggtaacgtgccac58.3100100
INLb-84Serogroup 4b (213)TCAACTAACTAATCATCTATCAATggtaaaaatatgcgaatattg9.7100100
INLb-85Serogroup 4bATACTACATCATAATCAAACATCActcgtgaacaagctttcc5.5100100
INLb-16Serogroup 1/2b (293)AATCAATCTTCATTCAAATCATCAggtaaaaatatgcgtatctta11.7100100
INLb-100Serogroup 1/2bCTATCTTTAAACTACAAATCTAACgtgaataagctatcggtctat13.0100100
LMO-42Serogroup 1/2a (268)CTATCTTCATATTTCACTATAAACtggcgttgctgrctaagtttg6.6100100
VGCb-40Serogroup 1/2aCTTTCTACATTATTCACAACATTAaatcaagcsgctcatatgaag10.410098.6
LMO-9Serogroup 1/2c (72)TAATCTTCTATATCAACATCTTACtttactggtgaaatggcg13.5100100
VGCb-5Serogroup 1/2cCAATTCAAATCACAATAATCAATCaagattacgaatcgcttccac20.898.6100
LMO-10ECI (111)ATCATACATACATACAAATCTACAatgattaaaagtcagggaaag19.0100100
LMO-28ECICTACAAACAAACAAACATTATCAAaatcgaggcttacgaacgt23.7100100
VGCc-80ECIa (44)CTAACTAACAATAATCTAACTAACactacaacgaaaacagcgc10.7100100
VGCa-35ECIaCAATTTCATCATTCATTCATTTCAgttacttttatgtcgagt9.2100100
LMO-12ECII (35)TACACTTTCTTTCTTTCTTTCTTTataccgattatttggacggtt3.8100100
LMO-30ECIITTACCTTTATACCTTTCTTTTTACgacttgtagcagttgatttcaa7.5100100
VGCc-45ECIII (10)TCATTTCACAATTCAATTACTCAActcttatttgcttttgttggtc21.110099.4
INLa-3ECIIITACACTTTATCAAATCTTACAATCgagcttaatgaaaatcagcta17.010099.4
INLa-1INLa controlCTTTAATCTCAATCAATACAAATCagaagtggaagctgggaaNAaNANA
INLb-13INLb controlCAATAAACTATACTTCTTCACTAAtgcacctaaacctccgacNANANA
LMO-88LMO controlTTACTTCACTTTCTATTTACAATCccgtttccttatgccacaNANANA
VGCa-23VGCa controlTTCAATCATTCAAATCTCAACTTTcaagycctaagacgccaatcgNANANA
VGCb-25VGCb controlCTTTTCAATTACTTCAAATCTTCAgcatgcgttagttcatgrccaNANANA
VGCc-82VGCc controlTACATACACTAATAACATACTCATgactgcatgctagaatctaagNANANA
Open in a separate windowaNA, not applicable for positive amplicon control probes.bLuminex microsphere sets (Luminex Corporation) used for hybridization reactions are indicated following the hyphen.cn, number of isolates representing the target subtype among the 906 tested isolates.dThe 5′ sequence tag portions of extension probes are capitalized. See IUPAC codes for definitions of degenerate bases.eID, index of discrimination.Validation of the TMLGT assay was performed using 906 L. monocytogenes isolates for which the lineage, major serogroup, and epidemic clone type had been determined independently (see Table S1 in the supplemental material). A subset of 92 isolates, including at least five isolates from each lineage, serogroup, and epidemic clone type, was used to evaluate the discriminatory power of subtype-specific probes and the repeatability of the assay (see Table S1). Two independent runs of the 30-probe TMLGT assay produced identical results for these 92 isolates. In addition, genotypes matched expectations for all isolate/probe combinations, and the fluorescence intensities for positive genotypes (those targeted by a particular probe) were 3.8 to 58.3 (mean, 18.5) times as high as background values for isolates with negative genotypes (those not targeted by a particular probe) (Table (Table2).2). The performances of individual probes also were assessed in terms of sensitivity and specificity, where sensitivity is defined as the percentage of positive samples that produced positive results and specificity indicates the percentage of negative samples that produce negative results (5). Based on results from all 906 isolates analyzed by TMLGT, probe sensitivity was at least 98.6% and 23 of the 24 subtype-specific probes exhibited 100% sensitivity (Table (Table2).2). The specificities for all probes were also greater than 98.6%, and 21 of the 24 subtype-specific probes exhibited 100% specificity (Table (Table22).All but three of the 906 isolates in the validation panel were fully and accurately typed relative to lineage, serogroup, and epidemic clone by using the TMLGT assay (typeability, 99.9%; accuracy of isolate assignment, 99.8%). One of the lineage II isolates, NRRL B-33880, could not be assigned to a serogroup based on the TMLGT results because this isolate was positive for one of the serogroup 1/2a probes (VGCb-40) and one of the serogroup 1/2c probes (LMO-9). This isolate was previously identified as a member of serogroup 1/2c based on mapping lineage-specific MLGT data onto a multilocus phylogeny (34) but produced a serogroup 1/2a-specific banding pattern (data not shown) with the multiplex PCR assay described by Doumith et al. (7). Similar strains, including the common laboratory strain EGD-e, were found to have genomes that are more similar to serogroup 1/2c strains than to strains from the 1/2a serogroup (8, 33) and likely represent intermediates in the evolution of the 1/2c clade from 1/2a ancestors. There is a poor correlation between genomic and antigenic variation for such isolates (34), consistent with the ambiguous results produced by application of the TMLGT assay to NRRL B-33880. The two other problematic isolates, NRRL B-33555 and NRRL B-33559, were accurately identified based on TMLGT data as lineage II isolates from the 1/2a serogroup. However, these two isolates were positive for both ECIII-specific probes in the TMLGT assay but have lineage-specific MLGT haplotypes (Lm2.46), indicating that they are representatives of a sister group closely related to ECIII (33).In 2005, the Food Safety and Inspection Service (FSIS) implemented an approach to inspection that includes consideration of relative risk in order to determine L. monocytogenes sampling frequency among establishments that produce certain RTE products. This approach incorporates information on production volume, outgrowth potential in the product, steps taken to prevent postlethality contamination, and FSIS sampling history. However, L. monocytogenes subtype-specific variation in ecology and virulence indicates that information on the lineage, major serogroup, and epidemic clone identities of isolates could be used to inform assessments of relative risk and to improve inspection programs that are based on consideration of risk. Several PCR-based methods have been described for differentiation of various combinations of these subgroups (1-3, 5, 7, 10, 35, 37); however, these approaches have focused on a single subgroup or a smaller set of subgroups than is differentiated by TMLGT analysis. Although we previously developed a set of three MLGT assays that can be used to differentiate all of the major serogroups and epidemic clones of L. monocytogenes (9, 33, 34), those assays did not include probes for lineage discrimination and require identification of the lineage prior to application of one of three unique sets of probes. In addition, the MLGT assays were designed to maximize strain discrimination, as opposed to subgroup identification, and require the use of at least twice as many probes as is needed for TMLGT analysis. MLGT data analysis is also more complicated than analysis of TMLGT data, and serogroup or epidemic clone type identification via MLGT requires phylogenetic analyses to place novel haplotypes within an established phylogenetic framework.In the present study, we developed the first assay for simultaneous discrimination of the four lineages, the four major serogroups, and the four previously described epidemic clones of L. monocytogenes. The assay includes multiple markers for each of these subtype probes as well as control probes to ensure that negative probe data were not the result of amplification failure, providing a high degree of internal validation required for use in inspection programs that consider risk in making sampling decisions. In addition, the utility of the assay has been validated with a large and diverse panel of 906 isolates, including 567 isolates from FSIS surveillance of RTE products and processing facilities (see Table S1 in the supplemental material). Data produced by the TMLGT assay are amenable to high-throughput analysis, and a simple spreadsheet utility has been developed to semiautomate subtype identifications and to alert investigators to potentially conflicting probe data (available upon request). In addition to having a potential application in inspection programs, the TMLGT assay provides a rapid and accurate means of characterizing L. monocytogenes isolates from different environments, which would facilitate pathogen tracking and improve understanding of L. monocytogenes ecology.   相似文献   

3.
4.
5.
6.
7.
Predator-prey relationships among prokaryotes have received little attention but are likely to be important determinants of the composition, structure, and dynamics of microbial communities. Many species of the soil-dwelling myxobacteria are predators of other microbes, but their predation range is poorly characterized. To better understand the predatory capabilities of myxobacteria in nature, we analyzed the predation performance of numerous Myxococcus isolates across 12 diverse species of bacteria. All predator isolates could utilize most potential prey species to effectively fuel colony expansion, although one species hindered predator swarming relative to a control treatment with no growth substrate. Predator strains varied significantly in their relative performance across prey types, but most variation in predatory performance was determined by prey type, with Gram-negative prey species supporting more Myxococcus growth than Gram-positive species. There was evidence for specialized predator performance in some predator-prey combinations. Such specialization may reduce resource competition among sympatric strains in natural habitats. The broad prey range of the Myxococcus genus coupled with its ubiquity in the soil suggests that myxobacteria are likely to have very important ecological and evolutionary effects on many species of soil prokaryotes.Predation plays a major role in shaping both the ecology and evolution of biological communities. The population and evolutionary dynamics of predators and their prey are often tightly coupled and can greatly influence the dynamics of other organisms as well (1). Predation has been invoked as a major cause of diversity in ecosystems (11, 12). For example, predators may mediate coexistence between superior and inferior competitors (2, 13), and differential trajectories of predator-prey coevolution can lead to divergence between separate populations (70).Predation has been investigated extensively in higher organisms but relatively little among prokaryotes. Predation between prokaryotes is one of the most ancient forms of predation (27), and it has been proposed that this process may have been the origin of eukaryotic cells (16). Prokaryotes are key players in primary biomass production (44) and global nutrient cycling (22), and predation of some prokaryotes by others is likely to significantly affect these processes. Most studies of predatory prokaryotes have focused on Bdellovibrionaceae species (e.g., see references 51, 55, and 67). These small deltaproteobacteria prey on other Gram-negative cells, using flagella to swim rapidly until they collide with a prey cell. After collision, the predator cells then enter the periplasmic space of the prey cell, consume the host cell from within, elongate, and divide into new cells that are released upon host cell lysis (41). Although often described as predatory, the Bdellovibrionaceae may also be considered to be parasitic, as they typically depend (apart from host-independent strains that have been observed [60]) on the infection and death of their host for their reproduction (47).In this study, we examined predation among the myxobacteria, which are also deltaproteobacteria but constitute a monophyletic clade divergent from the Bdellovibrionaceae (17). Myxobacteria are found in most terrestrial soils and in many aquatic environments as well (17, 53, 74). Many myxobacteria, including the model species Myxococcus xanthus, exhibit several complex social traits, including fruiting body formation and spore formation (14, 18, 34, 62, 71), cooperative swarming with two motility systems (64, 87), and group (or “wolf pack”) predation on both bacteria and fungi (4, 5, 8, 9, 15, 50). Using representatives of the genus Myxococcus, we tested for both intra- and interspecific variation in myxobacterial predatory performance across a broad range of prey types. Moreover, we examined whether prey vary substantially in the degree to which they support predatory growth by the myxobacteria and whether patterns of variation in predator performance are constant or variable across prey environments. The latter outcome may reflect adaptive specialization and help to maintain diversity in natural populations (57, 59).Although closely related to the Bdellovibrionaceae (both are deltaproteobacteria), myxobacteria employ a highly divergent mode of predation. Myxobacteria use gliding motility (64) to search the soil matrix for prey and produce a wide range of antibiotics and lytic compounds that kill and decompose prey cells and break down complex polymers, thereby releasing substrates for growth (66). Myxobacterial predation is cooperative both in its “searching” component (6, 31, 82; for details on cooperative swarming, see reference 64) and in its “handling” component (10, 29, 31, 32), in which secreted enzymes turn prey cells into consumable growth substrates (56, 83). There is evidence that M. xanthus employs chemotaxis-like genes in its attack on prey cells (5) and that predation is stimulated by close contact with prey cells (48).Recent studies have revealed great genetic and phenotypic diversity within natural populations of M. xanthus, on both global (79) and local (down to centimeter) scales (78). Phenotypic diversity includes variation in social compatibility (24, 81), the density and nutrient thresholds triggering development (33, 38), developmental timing (38), motility rates and patterns (80), and secondary metabolite production (40). Although natural populations are spatially structured and both genetic diversity and population differentiation decrease with spatial scale (79), substantial genetic diversity is present even among centimeter-scale isolates (78). No study has yet systematically investigated quantitative natural variation in myxobacterial predation phenotypes across a large number of predator genotypes.Given the previous discovery of large variation in all examined phenotypes, even among genetically extremely similar strains, we anticipated extensive predatory variation as well. Using a phylogenetically broad range of prey, we compared and contrasted the predatory performance of 16 natural M. xanthus isolates, sampled from global to local scales, as well as the commonly studied laboratory reference strain DK1622 and representatives of three additional Myxococcus species: M. flavescens (86), M. macrosporus (42), and M. virescens (63) (Table (Table1).1). In particular, we measured myxobacterial swarm expansion rates on prey lawns spread on buffered agar (31, 50) and on control plates with no nutrients or with prehydrolyzed growth substrate.

TABLE 1.

List of myxobacteria used, with geographical origin
Organism abbreviation used in textSpeciesStrainGeographic originReference(s)
A9Myxococcus xanthusA9Tübingen, Germany78
A23Myxococcus xanthusA23Tübingen, Germany78
A30Myxococcus xanthusA30Tübingen, Germany78
A41Myxococcus xanthusA41Tübingen, Germany78
A46Myxococcus xanthusA46Tübingen, Germany78
A47Myxococcus xanthusA47Tübingen, Germany78
A75Myxococcus xanthusA75Tübingen, Germany78
A85Myxococcus xanthusA85Tübingen, Germany78
TVMyxococcus xanthusTvärminneTvärminne, Finland79
PAKMyxococcus xanthusPaklenicaPaklenica, Croatia79
MADMyxococcus xanthusMadeira 1Madeira, Portugal79
WARMyxococcus xanthusWarwick 1Warwick, UK79
TORMyxococcus xanthusToronto 1Toronto, Ontario, Canada79
SUL2Myxococcus xanthusSulawesi 2Sulawesi, Indonesia79
KALMyxococcus xanthusKalalauKalalau, HI79
DAVMyxococcus xanthusDavis 1ADavis, CA79
GJV1Myxococcus xanthusGJV 1Unknown35, 72
MXFL1Myxococcus flavescensMx fl1Unknown65
MXV2Myxococcus virescensMx v2Unknown65
CCM8Myxococcus macrosporusCc m8Unknown65
Open in a separate window  相似文献   

8.
9.
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   

10.
11.
12.
A total of 210 Salmonella isolates, representing 64 different serovars, were isolated from imported seafood samples, and 55/210 isolates were found to be resistant to at least one antibiotic. Class 1 integrons from three multidrug-resistant Salmonella enterica strains (Salmonella enterica serovars Newport [strain 62], Typhimurium var. Copenhagen [strain 629], and Lansing [strain 803], originating from Hong Kong, the Philippines, and Taiwan, respectively) were characterized. Southern hybridization of plasmids isolated from these strains, using a class 1 integron probe, showed that trimethoprim-sulfamethoxazole and streptomycin resistance genes were located on a megaplasmid in strain 629. Our study indicates that imported seafood could be a reservoir for Salmonella isolates resistant to multiple antibiotics.Salmonella spp. are recognized as major food-borne pathogens of humans worldwide. In the United States, there are an estimated 800,000 to 4 million Salmonella infections annually, and approximately 500 of the cases are fatal (8, 26). A variety of foods have been implicated as vehicles transmitting salmonellosis to humans, including poultry, beef, pork, eggs, milk, cheese, fish, shellfish, fruits, juice, and vegetables (1, 4, 9, 12, 23). Previous studies by field laboratories of the U.S. Food and Drug Administration have shown the prevalences of Salmonella isolates in imported and domestic seafood as 7.2% and 1.3%, respectively (6, 11, 27).Mobile genetic elements, such as plasmids, transposons, and integrons, which disseminate antibiotic resistance genes by horizontal or vertical transfer, as part of either resistance plasmids or conjugative transposons, play an important role in the evolution and dissemination of multidrug resistance (2, 3, 10, 17). Salmonella genomic island 1 (SGI1), the first genomic island reported to contain an antibiotic resistance gene cluster, was identified in the multidrug-resistant Salmonella enterica serovar Typhimurium strain DT 104 (21).Most studies of the prevalence and characterization of antimicrobial resistance genes and integrons in Salmonella spp. have focused on strains from clinical and veterinary sources. However, little is known about the occurrence of SGI1 and its variants in Salmonella spp. isolated from seafood. We have screened a set of drug-resistant S. enterica strains from seafood belonging to 64 different serovars for SGI1 and class 1 integron conserved sequences (CS). We report the presence of a class I variant integron carrying the dfrXII and aadA2 genes on a megaplasmid in serovar Typhimurium var. Copenhagen and on the chromosome in Salmonella enterica serovar Lansing. We also found the variant class 1 integron carrying the dfrA1 and orfC genes in Salmonella enterica serovar Newport strains from seafood.A total of 210 Salmonella enterica strains isolated from seafood imported into the United States between 2000 and 2005 were identified and serotyped by the Pacific Regional Laboratory-Southwest of the FDA, Irvine, CA. The Salmonella strains represented 20 serogroups (Table (Table1)1) from various imported seafood items. The Salmonella strains were tested with 16 antibiotics (14) commonly used in either human or veterinary medicine on Mueller-Hinton agar (Difco Laboratories, Detroit, MI), using a disk diffusion method. The sensitivity and resistance were determined by the criteria of the Clinical and Laboratory Standards Institute (1999).

TABLE 1.

Salmonella serotypes isolated from imported foods
No. of strainsS. enterica serovar(s) or Salmonella group(s)
39Weltevreden
16Newport
13Saintpaul
10Senftenberg
8Lexington
7Virchow
6Enteritidis, Bareily
5Bovismorbificans, Brunei, Java, Hvittingfoss
4Paratyphi B var. Java, Thompson
3Aberdeen, Cubana, Stanley, Derby, Lansing
2Montevideo, Hadar, Agona, San Diego, Braenderup, Lanka, Salmonella enterica subsp. diarizonae, Oslo, Bareily variant, Salmonella monophasic group C2
1Ouakam, Cannstatt, Albany, Newport/Bardo, Adelaide, S. enterica subsp. diarizonae, Houten, Giza, Miami, Onderstepoof, Infantis, Salmonella monophasic group D1, Mbandaka, Salmonella monophasic group G2, Ohio, Rutgers, Salmonella monophasic group D2, Amsterdam, Salmonella enterica subsp. IV serotype 43:z4z23, Paratyphi B var. Java, Wentworth, Potsdam, Muenster var. 15+, 34+, Lexington var. 15+, Weltevreden var. 15+, S. enterica subsp. I, Madella, Alachua, London, Singapore, Uphill, Thielallee, Typhimurium var. Copenhagen
Open in a separate windowAll Salmonella strains that were resistant to three or four antibiotics and trimethoprim were screened by PCR for the presence of class 1 integrons, using the CSL1 and CSR1 primers (Table (Table2)2) (14). To confirm other antibiotic resistance genes, we used primers and PCR methods described previously (13, 14, 16). To identify SGI1 in multidrug-resistant strains, PCR was performed by using primers U7-L12/LJ-R1 and 104-RJ/104-D (Table (Table2),2), corresponding to the left and right junctions of SGI1 in the Salmonella chromosome, respectively (16). For a positive control, serovar Typhimurium DT104 strain DT7 (13) was used. As a negative control, Escherichia coli cells or DNA was used. A reagent blank included in each PCR contained distilled water instead of template DNA. For sequencing, the PCR-amplified integrons were purified and cloned into plasmid vector pCR2.1 (Invitrogen Corp., Carlsbad, CA). The clones were investigated for the presence of inserts by isolating the recombinant plasmid, which was confirmed by digestion with the restriction enzyme EcoRI. Sequencing of both strands was performed. DNA sequences were analyzed with Lasergene (DNASTAR, Inc., Madison, WI) software. Oligonucleotide primers and probes were purchased from MWG Biotech (High Point, NC).

TABLE 2.

Primer pairs for integron PCR and sequencing
PrimerSequence (5′-3′)LocationPCR product size (bp)
CSL1GGC ATC CAA GCA GCA AGC5′ CS
CSR1AAG CAG ACT TGA CCT GAT3′ CS
U7-L12ACA CCT TGA GCA GGG CAA AGthdF500
LJ-R1AGT TCT AAA GGT TCG TAG TCG
104-RJTGA CGA GCT GAA GCG AAT TGS044
104DACC AGG GCA AAA CTA CAC AGyidY
aadA2FTGT TGG TTA CTG TGG CCG TAaadA2380
aadA2RGCT GCG AGT TCC ATA GCT TC
Open in a separate windowPlasmid DNA was isolated using an alkaline lysis method with modifications described previously (19). Plasmids were separated by electrophoresis in 1× Tris-acetate-EDTA buffer at 64 V for 2 h on 1.0% agarose gels, stained with 40 μl of ethidium bromide (0.625 mg/ml) for visualization, and then transferred and cross-linked to positively charged nylon membranes (Roche, Indianapolis, IN). The resulting blots were hybridized at 65°C for 18 h with digoxigenin-labeled DNA probes (1.2-kb and 1.9-kb PCR-amplified products), using CSL1 and CSR1 primers specific for class 1 integrons (22).  相似文献   

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

TABLE 1.

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

TABLE 2.

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

TABLE 3.

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

TABLE 4.

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

14.
Recent studies indicate that sexual transmission of human immunodeficiency virus type 1 (HIV-1) generally results from productive infection by only one virus, a finding attributable to the mucosal barrier. Surprisingly, a recent study of injection drug users (IDUs) from St. Petersburg, Russia, also found most subjects to be acutely infected by a single virus. Here, we show by single-genome amplification and sequencing in a different IDU cohort that 60% of IDU subjects were infected by more than one virus, including one subject who was acutely infected by at least 16 viruses. Multivariant transmission was more common in IDUs than in heterosexuals (60% versus 19%; odds ratio, 6.14; 95% confidence interval [CI], 1.37 to 31.27; P = 0.008). These findings highlight the diversity in HIV-1 infection risks among different IDU cohorts and the challenges faced by vaccines in protecting against this mode of infection.Elucidation of virus-host interactions during and immediately following the transmission event is one of the great challenges and opportunities in human immunodeficiency virus (HIV)/AIDS prevention research (14-16, 31, 34, 45). Recent innovations involving single-genome amplification (SGA), direct amplicon sequencing, and phylogenetic inference based on a model of random virus evolution (18-20, 43) have allowed for the identification of transmitted/founder viruses that actually cross from donor to recipient, leading to productive HIV type 1 (HIV-1) infection. Our laboratory and others have made the surprising finding that HIV-1 transmission results from productive infection by a single transmitted/founder virus (or virally infected cell) in ∼80% of HIV-infected heterosexuals and in ∼60% of HIV-infected men who have sex with men (MSM) (1, 13, 18, 24). These studies thus provided a precise quantitative estimate for the long-recognized genetic bottleneck in HIV-1 transmission (6, 11-13, 17, 25, 28, 30, 35, 38, 42, 47-49) and a plausible explanation for the low acquisition rate per coital act and for graded infection risks associated with different exposure routes and behaviors (15, 36).In contrast to sexual transmission of HIV-1, virus transmission resulting from injection drug use has received relatively little attention (2, 3, 29, 42) despite the fact that injection drug use-associated transmission accounts for as many as 10% of new infections globally (26, 46). We hypothesized that SGA strategies developed for identifying transmitted/founder viruses following mucosal acquisition are applicable to deciphering transmission events following intravenous inoculation and that, due to the absence of a mucosal barrier, injection drug users (IDUs) exhibit a higher frequency of multiple-variant transmission and a wider range in numbers of transmitted viruses than do acutely infected heterosexual subjects. We obtained evidence in support of these hypotheses from the simian immunodeficiency virus (SIV)-Indian rhesus macaque infection model, where we showed that discrete low-diversity viral lineages emanating from single or multiple transmitted/founder viruses could be identified following intravenous inoculation and that the rectal mucosal barrier to infection was 2,000- to 20,000-fold greater than with intravenous inoculation (19). However, we also recognized potentially important differences between virus transmission in Indian rhesus macaques and virus transmission in humans that could complicate an IDU acquisition study. For example, in the SIV macaque model, the virus inocula can be well characterized genetically and the route and timing of virus exposure in relation to plasma sampling precisely defined, whereas in IDUs, the virus inoculum is generally undefined and the timing of virus infection only approximated based on clinical history and seroconversion testing (8). In addition, IDUs may have additional routes of potential virus acquisition due to concomitant sexual activity. Finally, there is a paucity of IDU cohorts for whom incident infection is monitored sufficiently frequently and clinical samples are collected often enough to allow for the identification and enumeration of transmitted/founder viruses. To address these special challenges, we proposed a pilot study of 10 IDU subjects designed to determine with 95% confidence if the proportion of multivariant transmissions in IDUs was more than 2-fold greater than the 20% frequency established for heterosexual transmission (1, 13, 18, 24). A secondary objective of the study was to determine whether the range in numbers of transmitted/founder viruses in IDUs exceeded the 1-to-6 range observed in heterosexuals (1, 13, 18, 24). To ensure comparability among the studies, we employed SGA-direct amplicon sequencing approaches, statistical methods, and power calculations identical to those that we had used previously to enumerate transmitted/founder viruses in heterosexual and MSM cohorts (1, 13, 18, 20, 24).We first surveyed investigators representing acute-infection cohorts in the United States, Canada, Russia, and China; only one cohort—the Montreal Primary HIV Infection Cohort (41)—had IDU clinical samples and clinical data available for study. The Montreal cohort of subjects with acute and early-stage HIV-1 infection was established in 1996 and recruits subjects from both academic and private medical centers throughout the city. Injection drug use is an important contributing factor to Montreal''s HIV burden, with IDUs comprising approximately 20% of the city''s AIDS cases and 35% of the cohort (21, 40, 41). A large proportion of Montreal''s IDUs use injection cocaine, with 50 to 69% of subjects reporting cocaine as their injection drug of choice (4, 5, 9, 22, 23).Subjects with documented serological evidence of recent HIV-1 infection and a concurrent history of injection drug use were selected for study. These individuals had few or no reported risk factors for sexual HIV-1 acquisition. Clinical history and laboratory tests of HIV-1 viremia and antibody seroconversion were used to determine the Fiebig clinical stage (8) and to estimate the date of infection (Table (Table1).1). One subject was determined to be in Fiebig stage III, one subject was in Fiebig stage IV, five subjects were in Fiebig stage V, and three subjects were in Fiebig stage VI. We performed SGA-direct amplicon sequencing on stored plasma samples and obtained a total of 391 3′ half-genomes (median, 25 per subject; range, 19 to 167). Nine of these sequences contained large deletions or were G-to-A hypermutated and were excluded from subsequent analysis. Sequences were aligned, visually inspected using the Highlighter tool (www.hiv.lanl.gov/content/sequence/HIGHLIGHT/highlighter.html), and analyzed by neighbor-joining (NJ) phylogenetic-tree construction. A composite NJ tree of full-length gp160 env sequences from all 10 subjects (Fig. (Fig.1A)1A) revealed distinct patient-specific monophyletic lineages, each with high bootstrap support and separated from the others by a mean genetic distance of 10.79% (median, 11.29%; range, 3.00 to 13.42%). Maximum within-patient env gene diversity ranged from 0.23% to 3.34% (Table (Table1).1). Four subjects displayed distinctly lower within-patient maximum env diversities (0.23 to 0.49%) than the other six subjects (1.48% to 3.34%). The lower maximum env diversities in the former group are consistent with infection either by a single virus or by multiple closely related viruses, while the higher diversities can be explained only by transmission of more than one virus based on empirical observations (1, 13, 18, 24) and mathematical modeling (18, 20).Open in a separate windowFIG. 1.NJ trees and Highlighter plots of HIV-1 gp160 env sequences. (A) Composite tree of 382 gp160 env sequences from all study subjects. The numerals at the nodes indicate bootstrap values for which statistical support exceeded 70%. (B) Subject ACT54869022 sequences suggest productive infection by a single virus (V1). (C) Subject HDNDRPI032 sequences suggest productive infection by as many as three viruses. (D) Subject HDNDRPI001 sequences suggest productive infection by at least five viruses with extensive interlineage recombination. Sequences are color coded to indicate viral progeny from distinct transmitted/founder viruses. Recombinant virus sequences are depicted in black. Methods for SGA, sequencing, model analysis, Highlighter plotting, and identification of transmitted/founder virus lineages are described elsewhere (18, 20, 24, 44). The horizontal scale bars represent genetic distance. nt, nucleotide.

TABLE 1.

Subject demographics and HIV-1 envelope analysis results
Subject identifierAge (yr)SexaFiebig stageEstimated no. of days postinfectionbCD4 countPlasma viral load (log)No. of SGA ampliconsDiversity of env genes (%)c
No. of transmitted/ founder viruses
MeanInterquartile rangeMaximumdModel predictionePhylogenetic estimatef
HDNDRPI03447MIII292407.881631.070.553.34>116
HDNDRPI02918FIV484404.34290.160.150.4911
HTM38524MV624065.37220.120.080.2711
CQLDR0342MV66NDg5.01210.080.080.2311
HDNDRPI00136MV286905.94250.900.631.91>15
HTM31939MV685204.43250.770.461.54>13
HDNDRPI03237MV731,0403.53191.482.993.34>13
ACTDM58020839MVI933874.53301.170.972.64>13
ACT5486902228MVI687233.43270.070.040.2411
PSL02446MVI823404.46210.820.631.57>13
Open in a separate windowaM, male; F, female.bNumbers of days postinfection were estimated on the basis of serological markers, clinical symptoms, or a history of a high-risk behavior leading to virus exposure.cDiversity measurements determined by PAUP* analysis.dThe model prediction of the maximum achievable env diversity 100 days after transmission is 0.60% (95% CI, 0.54 to 0.68%). Diversity values exceeding this range imply transmission and productive infection by more than one virus. Diversity values less than 0.54% can be explained by transmission of one virus or of multiple closely related viruses (18).eModel described in Keele et al. (18).fMinimum estimate of transmitted/founder viruses.gND, not determined.An example of productive clinical infection by a single virus is shown in phylogenetic tree and Highlighter plots from subject ACT54869022 (Fig. (Fig.1B).1B). A similar phylogenetic pattern of single-variant transmission was found in 4 of 10 IDU subjects (Table (Table1).1). Examples of multivariant transmission are shown for subject HDNDRPI032, for whom there was evidence of infection by 3 transmitted/founder viruses (Fig. (Fig.1C)1C) and for subject HDNDRPI001, for whom there was evidence of infection by at least 5 transmitted/founder viruses (Fig. (Fig.1D).1D). One IDU subject, HDNDRPI034, had evidence of multivariant transmission to an extent not previously seen in any of 225 subjects who acquired their infection by mucosal routes (1, 13, 18, 24) or in any of 13 IDUs, as recently reported by Masharsky and colleagues (29). We greatly extended the depth of our analysis in this subject to include 163 3′ half-genome sequences in order to increase the sensitivity of detection of low-frequency viral variants. Power calculations indicated that a sample size of 163 sequences gave us a >95% probability of sampling minor variants comprising as little as 2% of the virus population. By this approach, we found evidence of productive infection by at least 16 genetically distinct viruses (Fig. (Fig.2).2). Fourteen of these could be identified unambiguously based on the presence of discrete low-diversity viral lineages, each consisting of between 2 and 48 sequences. Two additional unique viral sequences with long branch lengths (3F8 and G10) exhibited diversity that was sufficiently great to indicate a distinct transmission event as opposed to divergence from other transmitted/founder lineages (see the legend to Fig. Fig.2).2). It is possible that still other unique sequences from this subject also represented transmitted/founder viruses, but we could not demonstrate this formally. We also could not determine if all 16 (or more) transmission events resulted from a single intravenous inoculation or from a series of inoculations separated by hours or days; however, it is likely that all transmitted viruses in this subject resulted from exposure to plasma from a single infected individual, since the maximum env diversity was only 3.34% (Fig. (Fig.1A).1A). It is also likely that transmission occurred within a brief window of time, since the period from transmission to the end of Fiebig stage III is typically only about 25 days (95% CI, 22 to 37 days) (18, 20) and the diversity observed in all transmitted/founder viral lineages in subject HDNDRPI034 was exceedingly low, consistent with model predictions for subjects with very recent infections (18, 20).Open in a separate windowFIG. 2.NJ tree and Highlighter plot of HIV-1 3′ half-genome sequences from subject HDNDRPI034. Sequences emanating from 16 transmitted/founder viruses are color coded. Fourteen transmitted/founder viral lineages comprised of 2 or more identical or nearly identical sequences could be readily distinguished from recombinant sequences (depicted in black), which invariably appeared as unique sequences containing interspersed segments shared with other transmitted/founder virus lineages. The two sequences with the longest branch lengths (3F8 and G10) were interpreted to represent rare progeny of discrete transmitted/founder viruses because their unique polymorphisms far exceeded the maximum diversity estimated to occur in the first 30 days of infection (0.22%; CI, 0.15 to 0.31%) (18) and far exceeded the diversity observed within the other transmitted/founder virus lineages. The horizontal scale bar represents genetic distance.Lastly, we compared the multiplicity of HIV-1 transmission in the Montreal IDU subjects with that of non-IDU subjects for whom identical SGA methods had been employed. In this combined-cohort analysis, we found the frequency of multiple-variant transmission in heterosexuals to be 19% (34 of 175) and in MSM 38% (19 of 50) (Table (Table2)2) (24). The current study was powered to detect a >2-fold difference in multivariant transmission between IDUs and heterosexual subjects; in fact, we observed a 3-fold-higher frequency of multiple-variant transmission in Montreal IDUs (6 of 10 subjects [60%]) than in heterosexuals (odds ratio, 6.14; 95% CI, 1.37 to 31.27; Fisher exact test, P = 0.008) and a 1.5-fold-higher frequency in Montreal IDUs than in MSM (odds ratio, 2.41; 95% CI, 0.50 to 13.20; P = 0.294, not significant). In addition, we found that the range of numbers of transmitted/founder viruses was greater in IDUs (range, 1 to 16 viruses; median, 3) than in either heterosexuals (range, 1 to 6 viruses; median, 1) or MSM (range, 1 to 10 viruses; median, 1). The finding of larger numbers of transmitted/founder viruses in IDUs was not simply the result of more intensive sampling, since the numbers of sequences analyzed in all studies were comparable. Moreover, it is notable that in studies reported elsewhere, we sampled as many as 239 sequences by SGA or as many as 500,000 sequences by 454 pyrosequencing from four acutely infected MSM subjects and in each case found evidence of productive clinical infection by only a single virus (24; W. Fischer, B. Keele, G. Shaw, and B. Korber, unpublished). These results thus suggest that IDUs may be infected by more viruses and by a greater range of viruses than is the case following mucosal transmission. On this count, our findings differ from those reported by Masharsky and coworkers for an IDU cohort from St. Petersburg, Russia (29). Their study found a low frequency of multiple virus transmissions (31%), not significantly different from that of acutely infected heterosexuals, and a low number of transmitted/founder viruses (range, 1 to 3 viruses; median, 1). Because the SGA methods employed in both studies were identical, the numbers of sequences analyzed per subject were comparable (median of 25 sequences in Montreal versus 33 in St. Petersburg), and because the discriminating power of the SGA-direct sequencing method was sufficient to distinguish transmitted/founder viruses differing by as few as 3 nucleotides, or <0.1% of nucleotides (Fig. (Fig.2,2, compare lineages V4 and V5), it is unlikely that differences in the genetic diversity of HIV-1 in the two IDU populations explain the differences in findings between the two studies. Instead, we suspect that the explanation lies in the small cohort sizes (10 versus 13 subjects) and the particular risk behaviors of the IDUs in each cohort. The Russian cohort is heavily weighted toward heroine use, whereas the Montreal cohort is weighted toward injection cocaine use, the latter being associated with more frequent drug administration and the attendant infection risks of needle sharing (4).

TABLE 2.

Multiplicity of HIV-1 infection in IDU, heterosexual, and MSM subjects
CohortReferenceVirus subtypeTotal no. of subjectsSingle-variant transmission
Multiple-variant transmission
P valueOdds ratio95% CIMedianRange
No. of subjects% of totalNo. of subjects% of total
HeterosexualsKeele et al. (18)B796582.301417.7011-4
Abrahams et al. (1)C695478.301521.7011-5
Haaland et al. (13)A or C272281.50518.5011-6
Total17514180.603419.400.008a6.141.37-31.2711-6
MSMKeele et al. (18)B221359.10940.9011-6
Li et al. (24)B281864.301035.7011-10
Total503162.001938.000.294b2.410.50-13.2011-10
IDUsBarB10440.00660.0031-16
Open in a separate windowaFisher''s exact test of multiple-variant transmission in heterosexuals versus in IDUs.bFisher''s exact test of multiple-variant transmission in MSM versus in IDUs.The results from the present study indicate that transmission of HIV-1 to IDUs can be associated with a high frequency of multiple-variant transmission and a broad range in the numbers of transmitted viruses. This wide variation in the multiplicity of HIV-1 infection in IDUs is likely due to the absence of a mucosal barrier to virus transmission (12, 19) and differences in the virus inocula (27, 29, 32, 39). The findings substantiate concerns raised in recent HIV-1 vaccine efficacy trials that different vaccine candidates may be more efficacious in preventing infection by some exposure routes than by others (7, 10, 33, 37). They further suggest that biological comparisons of molecularly cloned transmitted/founder viruses responsible for vaginal, rectal, penile, and intravenous infection could facilitate a mechanistic understanding of HIV-1 transmission and vaccine prevention (24, 44).  相似文献   

15.
Halogenated organic compounds serve as terminal electron acceptors for anaerobic respiration in a diverse range of microorganisms. Here, we report on the widespread distribution and diversity of reductive dehalogenase homologous (rdhA) genes in marine subsurface sediments. A total of 32 putative rdhA phylotypes were detected in sediments from the southeast Pacific off Peru, the eastern equatorial Pacific, the Juan de Fuca Ridge flank off Oregon, and the northwest Pacific off Japan, collected at a maximum depth of 358 m below the seafloor. In addition, significant dehalogenation activity involving 2,4,6-tribromophenol and trichloroethene was observed in sediment slurry from the Nankai Trough Forearc Basin. These results suggest that dehalorespiration is an important energy-yielding pathway in the subseafloor microbial ecosystem.Scientific ocean drilling explorations have revealed that marine subsurface sediments harbor remarkable numbers of microbial cells that account for approximately 1/10 to 1/3 of all living biota on Earth (20, 25, 33). Thermodynamic calculations of pore-water chemistry suggest that subseafloor microbial activities are generally supported by nutrient and energy supplies from the seawater and/or underlying basaltic aquifers (6, 7). Although sulfate, nitrate, Fe(III), Mn(IV), and bicarbonate are known to be potential electron acceptors for anaerobic microbial respiration in marine subsurface sediments (5), the incidence of both the dissimilatory dehalorespiration pathway and microbial activity in halogenated organic substrates remains largely unknown.Previous molecular ecological studies using 16S rRNA gene sequences demonstrated that Chloroflexi is one of the most frequently detected phyla in subseafloor sediments of the Pacific Ocean margins (12-14). Some of the sequences within the Chloroflexi are closely related to sequences in the genus Dehalococcoides, which contains obligatory dehalorespiring bacteria that employ halogenated organic compounds as terminal electron acceptors (21, 29). The frequent detection of Dehalococcoides-related 16S rRNA genes from these environments implies the occurrence of dissimilatory dehalorespiration in marine subsurface sediments.In this study, we detected and phylogenetically analyzed the reductive dehalogenase homologous (rdhA) genes, key functional genes for dehalorespiration pathways, from frozen sediment core samples obtained by Ocean Drilling Program (ODP) Leg 201 (Peru margin and eastern equatorial Pacific) (7, 14); Integrated Ocean Drilling Program (IODP) Expedition 301 (Juan de Fuca Ridge flank) (8, 24); Chikyu Shakedown Expedition CK06-06 (Northwest Pacific off Japan) (20, 23); and IODP Expedition 315 (Nankai Trough Forearc Basin off Japan) (Table (Table1).1). DNA was extracted using an ISOIL bead-beating kit (Nippon Gene, Japan) and purified using a MagExtractor DNA fragment purification kit (Toyobo, Japan) according to the manufacturer''s instructions. To increase concentration, DNA was amplified by multiple displacement amplification using the phi29 polymerase supplied with a GenomiPhi kit (GE Healthcare, United Kingdom) (20). Putative rdhA genes were amplified by PCR using Ex Taq polymerase (TaKaRa, Japan) with degenerate primers RRF2 and B1R (17), dehaloF3, dehaloF4, dehaloF5, dehaloR2, dehaloR3, and dehaloR4 (32), and ceRD2S, ceRD2L, and RD7 (26) and the PCR conditions described in those studies. Amplicons of the approximate target size were gel purified and cloned into the pCR2.1 vector (Invitrogen, Japan). Sequence similarity was analyzed using FastGroupII web-based software (34), and sequences with a 95% identity were tentatively assigned to the same phylotype. Amino acid sequences were aligned by ClustalW (31), including known and putative reductive dehalogenase sequences in the genome of Dehalococcoides ethenogenes strain 195 (28), as well as several functionally characterized reductive dehalogenases from other species.

TABLE 1.

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

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

TABLE 1.

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

17.
18.
The biological, serological, and genomic characterization of a paramyxovirus recently isolated from rockhopper penguins (Eudyptes chrysocome) suggested that this virus represented a new avian paramyxovirus (APMV) group, APMV10. This penguin virus resembled other APMVs by electron microscopy; however, its viral hemagglutination (HA) activity was not inhibited by antisera against any of the nine defined APMV serotypes. In addition, antiserum generated against this penguin virus did not inhibit the HA of representative viruses of the other APMV serotypes. Sequence data produced using random priming methods revealed a genomic structure typical of APMV. Phylogenetic evaluation of coding regions revealed that amino acid sequences of all six proteins were most closely related to APMV2 and APMV8. The calculation of evolutionary distances among proteins and distances at the nucleotide level confirmed that APMV2, APMV8, and the penguin virus all were sufficiently divergent from each other to be considered different serotypes. We propose that this isolate, named APMV10/penguin/Falkland Islands/324/2007, be the prototype virus for APMV10. Because of the known problems associated with serology, such as antiserum cross-reactivity and one-way immunogenicity, in addition to the reliance on the immune response to a single protein, the hemagglutinin-neuraminidase, as the sole base for viral classification, we suggest the need for new classification guidelines that incorporate genome sequence comparisons.Viruses from the Paramyxoviridae family have caused disease in humans and animals for centuries. Over the last 40 years, many paramyxoviruses isolated from animals and people have been newly described (16, 17, 22, 29, 31, 32, 36, 42, 44, 46, 49, 58, 59, 62-64). Viruses from this family are pleomorphic, enveloped, single-stranded, nonsegmented, negative-sense RNA viruses that demonstrate serological cross-reactivity with other paramyxoviruses related to them (30, 46). The subfamily Paramyxovirinae is divided into five genera: Respirovirus, Morbillivirus, Rubulavirus, Henipavirus, and Avulavirus (30). The Avulavirus genus contains nine distinct avian paramyxovirus (APMV) serotypes (Table (Table1),1), and information on the discovery of each has been reported elsewhere (4, 6, 7, 9, 12, 34, 41, 50, 51, 60, 68).

TABLE 1.

Characteristics of prototype viruses APMV1 to APMV9 and the penguin virus
StrainHostDiseaseDistributionFusion cleavagecGI accession no.
APMV1/Newcastle disease virus>250 speciesHigh mortalityWorldwideGRRQKRF45511218
InapparentWorldwideGGRQGRLa11545722
APMV2/Chicken/CA/Yucaipa/1956Turkey, chickens, psittacines, rails, passerinesDecrease in egg production and respiratory diseaseWorldwideDKPASRF169144527
APMV3/Turkey/WI/1968TurkeyMild respiratory disease and moderate egg decreaseWorldwidePRPSGRLa209484147
APMV3/Parakeet/Netherlands/449/1975Psittacines, passerines, flamingosNeurological, enteric, and respiratory diseaseWorldwideARPRGRLa171472314
APMV4/Duck/Hong Kong/D3/1975Duck, geese, chickensNone knownWorldwideVDIQPRF210076708
APMV5/Budgerigar/Japan/Kunitachi/1974Budgerigars, lorikeetsHigh mortality, enteric diseaseJapan, United Kingdom, AustraliaGKRKKRFa290563909
APMV6/Duck/Hong Kong/199/1977Ducks, geese, turkeysMild respiratory disease and increased mortality in turkeysWorldwidePAPEPRLb15081567
APMV7/Dove/TN/4/1975Pigeons, doves, turkeysMild respiratory disease in turkeysUnited States, England, JapanTLPSSRF224979458
APMV8/Goose/DE/1053/1976Ducks, geeseNone knownUnited States, JapanTYPQTRLa226343050
APMV9/Duck/NY/22/1978DucksNone knownWorldwideRIREGRIa217068693
APMV10/Penguin/Falkland Islands/324/2007Rockhopper penguinsNone KnownFalkland IslandsDKPSQRIa300432141
Open in a separate windowaRequires the addition of an exogenous protease.bProtease requirement depends on the isolate examined.cPutative.Six of these serotypes were classified in the latter half of the 1970s, when the most reliable assay available to classify paramyxoviruses was the hemagglutination inhibition (HI) assay (61). However, there are multiple problems associated with the use of serology, including the inability to classify some APMVs by comparing them to the sera of the nine defined APMVs alone (2, 8). In addition, one-way antigenicity and cross-reactivity between different serotypes have been documented for many years (4, 5, 14, 25, 29, 33, 34, 41, 51, 52, 60). The ability of APMVs, like other viruses, to show antigenic drift as it evolves over time (37, 43, 54) and the wide use and availability of precise molecular methods, such as PCR and genome sequencing, demonstrate the need for a more practical classification system.The genetic diversity of APMVs is still largely unexplored, as hundreds of avian species have never been surveyed for the presence of viruses that do not cause significant signs of disease or are not economically important. The emergence of H5N1 highly pathogenic avian influenza (HPAI) virus as the cause of the largest outbreak of a virulent virus in poultry in the past 100 years has spurred the development of surveillance programs to better understand the ecology of avian influenza (AI) viruses in aquatic birds around the globe, and in some instances it has provided opportunities for observing other viruses in wild bird populations (15, 53). In 2007, as part of a seabird health surveillance program in the Falkland Islands (Islas Malvinas), oral and cloacal swabs and serum were collected from rockhopper penguins (Eudyptes chrysocome) and environmental/fecal swab pools were collected from other seabirds.While AI virus has not yet been isolated from penguins in the sub-Antarctic and Antarctic areas, there have been two reports of serum antibodies positive to H7 and H10 from the Adélie species (11, 40). Rare isolations of APMV1, both virulent (45) and of low virulence (8), have been reported from Antarctic penguins. Sera positive for APMV1 and AMPV2 have also been reported (21, 24, 38, 40, 53). Since 1981, paramyxoviruses have been isolated from king penguins (Aptenodytes patagonicus), royal penguins (Eudyptes schlegeli), and Adélie penguins (Pygoscelis adeliae) from Antarctica and little blue penguins (Eudyptula minor) from Australia that cannot be identified as belonging to APMV1 to -9 and have not yet been classified (8, 11, 38-40). The morphology, biological and genomic characteristics, and antigenic relatedness of an APMV recently isolated from multiple penguin colonies on the Falkland Islands are reported here. Evidence that the virus belongs to a new serotype (APMV10) and a demonstration of the advantages of a whole genome system of analysis based on random sequencing followed by comparison of genetic distances are presented. Only after all APMVs are reported and classified will epidemiological information be known as to how the viruses are moving and spreading as the birds travel and interact with other avian species.  相似文献   

19.
20.
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   

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