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

TABLE 1.

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

4.
5.
6.
The human immunodeficiency virus type 1 (HIV-1) variants that are transmitted to newly infected individuals are the primary targets of interventions, such as vaccines and microbicides, aimed at preventing new infections. Newly acquired subtype A, B, and C variants have been the focus of neutralization studies, although many of these viruses, particularly of subtypes A and B, represent viruses circulating more than a decade ago. In order to better represent the global diversity of transmitted HIV-1 variants, an additional 31 sexually transmitted Kenyan HIV-1 env genes, representing several recent infections with subtype A, as well as subtypes A/D, C, and D, were cloned, and their neutralization profiles were characterized. Most env variants were resistant to neutralization by the monoclonal antibodies (MAbs) b12, 4E10, 2F5, and 2G12, suggesting that targeting the epitopes of these MAbs may not be effective against variants that are spreading in areas of endemicity. However, significant cross-subtype neutralization by plasma was observed, indicating that there may be other epitopes, not yet defined by the limited available MAbs, which could be recognized more broadly.Most effective viral vaccines are thought to provide protection primarily by stimulating neutralizing antibodies (NAbs) to clear cell-free virus (25, 27). Because protection by NAbs requires recognition of common viral epitopes, the extreme genetic diversity of human immunodeficiency virus type 1 (HIV-1) presents a particular challenge to NAb-based vaccine approaches. Therefore, a critical starting point for studies of immune-mediated protection against HIV-1 is a collection of newly transmitted HIV-1 variants, particularly from areas of endemicity, such as sub-Saharan Africa, in order to determine whether vaccines are appropriately targeted to common epitopes from these relevant transmitted strains.During HIV-1 transmission, a bottleneck allows only one or a few variants to be transmitted to a newly infected individual (6, 9, 16, 29, 34, 37, 39), and the sensitivity of these early transmitted strains to antibody-mediated neutralization is therefore of particular interest. Newly transmitted HIV-1 variants have demonstrated significant heterogeneity in their neutralization phenotypes both within and between subtypes (2, 3, 6-8, 11, 13-15, 22, 30, 32, 36). Panels of sexually transmitted HIV-1 envelope variants (based on the envelope gene, env) have been characterized, including subtype B variants from North America, Trinidad, and Europe, subtype C variants from South Africa and Zambia, and subtype A variants from Kenya collected between 1994 and 1996 (2, 14, 15). Here, we characterize an additional 31 envelope variants from 14 subjects with sexually transmitted HIV-1 who were infected in Kenya, where subtypes A, C, and D circulate, between 1993 and 2005 (24, 31).The env genes were cloned from samples drawn 14 to 391 (median, 65) days postinfection from individuals enrolled in a prospective cohort of high-risk women in Mombasa, Kenya (19-21). Demographic characteristics of the subjects are summarized in Table Table1;1; the timing of first infection was determined by both HIV-1 serology and HIV RNA testing as described previously (12). All of the subjects were presumably infected by male-to-female transmission and displayed a range of plasma viral loads at the time of env gene cloning (Table (Table1).1). For most individuals, full-length env genes were cloned from uncultured peripheral blood mononuclear cell (PBMC) DNA, though for two individuals, clones were obtained from DNA following short-term coculture with donor PBMCs (Table (Table1).1). env genes were cloned by single-copy nested PCR with primers and PCR conditions as described previously (4, 17). We tested env genes for their ability to mediate infection by transfecting env plasmid DNA into 293T cells along with an env-deficient HIV-1 subtype A proviral plasmid, Q23Δenv, to make pseudoviral particles (17). More than 80 env clones were obtained from 16 subjects; less than one-half were functional on the basis of the infectivity of pseudoviral particles in a single-round infection of TZM-bl cells (AIDS Research and Reference Reagent Program, National Institutes of Health), as observed previously for env genes cloned from proviral sequences (17); a lower fraction of functional env genes have been reported from plasma (18). We focused on the proviral sequences here because they presumably best represent the sequence closest to that of the transmitted strains. The 31 functional env variants are described in Table Table11.

TABLE 1.

Demographic characteristics, diversities, gp120 variable-region lengths, numbers of PNGS, and accession numbers of cloned env variants
SubjectVirus subtypeSample date (mo/day/yr)dpiaPlasma VLbSourcecIndividual env clonePairwise difference (%)dVariable-loop length (aa)
No. of PNGS
GenBank accession no.
V1/V2V3V4V5gp120gp41gp41 ecto
QB726A04/16/967061,940ucPBMCQB726.70M.ENV.B30.16633536102244FJ866111
QB726.70M.ENV.C4633536102244FJ866112
QF495A05/16/0623217,050ucPBMCQF495.23M.ENV.A10.121073537113044FJ866113
QF495.23M.ENV.A31073537113044FJ866114
QF495.23M.ENV.B21133537113144FJ866115
QF495.23M.ENV.D11133537113144FJ866116
QG984A07/12/042130,300ucPBMCQG984.21M.ENV.A3NA693436112433FJ866117
QH209A10/13/051428,600ucPBMCQH209.14M.ENV.A2NA723529112444FJ866118
QH343A09/08/052140,750,000ucPBMCQH343.21M.ENV.A100.19773532152644FJ866119
QH343.21M.ENV.B5773532152644FJ866120
QH359A10/05/052132,120ucPBMCQH359.21M.ENV.C11.4843536102944FJ866121
QH359.21M.ENV.D1733535102644FJ866122
QH359.21M.ENV.E2723540132844FJ866123
QA790eA/D06/10/9620448,100ccPBMCQA790.204I.ENV.A40.36773533112544FJ866124
QA790.204I.ENV.C1773533112644FJ866125
QA790.204I.ENV.C8773533112444FJ866126
QA790.204I.ENV.E2773533112544FJ866127
QG393A2/D06/23/046017,360ucPBMCQG393.60M.ENV.A10.7603431102455FJ866128
QG393.60M.ENV.B7573431102455FJ866129
QG393.60M.ENV.B8573431102455FJ866130
QB099eC02/10/9539127,280ucPBMCQB099.391M.ENV.B10.43653529102544FJ866131
QB099.391M.ENV.C8653529102544FJ866132
QC406C07/08/9770692,320ucPBMCQC406.70M.ENV.F3NA643520112254FJ866133
QA013D10/11/95701,527,700ccPBMCQA013.70I.ENV.H10.16603429122544FJ866134
QA013.70I.ENV.M12603429122544FJ866135
QA465D08/19/935937,750ucPBMCQA465.59M.ENV.A10.24653530112844FJ866136
QA465.59M.ENV.D1653530112744FJ866137
QB857D10/16/9711014,640ucPBMCQB857.23I.ENV.B3NA683432112654FJ866138
QD435D04/06/9910017,470ucPBMCQD435.100M.ENV.A40.88693429122654FJ866139
QD435.100M.ENV.B5673429112454FJ866140
QD435.100M.ENV.E1693429122654FJ866141
Open in a separate windowadpi, days postinfection as defined by RNA testing (12).bVL, viral load on the sample date in which env genes were cloned.cucPBMC, uncultured PBMCs; ccPBMC, cocultured PBMCs.dAverage pairwise distance between the full-length env variants from a given subject. NA, not applicable because there was only one variant available from the subject.eenv variants from these two subjects were cloned from >6 months postinfection, as noted, and should not be considered true early env variants.The full-length, functional env genes were sequenced and aligned to generate a maximum likelihood phylogenetic tree with reference sequences from the Los Alamos National Laboratory HIV database, as described previously (26). Viral env clones from the same subject clustered together, and a wide spectrum of genetic diversity was observed overall (Fig. (Fig.1).1). Some women, such as subject QF495, were infected with a relatively homogeneous viral population, with average pairwise differences of only 0.12% between env variants (Table (Table11 and Fig. Fig.1).1). However, as observed previously in this cohort (16, 28, 29, 33-35), other individuals, such as subjects QH359 and QD435, were infected with more heterogeneous viral populations with average pairwise differences of 1.4% and 0.88% between variants, respectively (Table (Table11 and Fig. Fig.1).1). env genes from subtypes A (13 variants), C (3 variants), and D (8 variants), as well as A/D recombinants (4 variants) and A2/D recombinants (3 variants), were represented (Fig. (Fig.1).1). The viral subtypes were confirmed using the NCBI genotyping database (http://www.ncbi.nlm.nih.gov/).Open in a separate windowFIG. 1.Maximum likelihood phylogenetic tree of full-length sequences from early subtype A, C, D, and A/D recombinant env variants in Kenya. The 31 novel env clones from Kenyan early infections and reference sequences for subtypes A, B, C, D, and K from the Los Alamos HIV database (http://www.hiv.lanl.gov/content/index) are displayed. The phylogenetic tree was rooted with subtype K env sequences. Values at nodes indicate the percentage of bootstraps in which the cluster the right was found; only values of 70% or greater are shown.The deduced amino acid sequences revealed that all functional variants had an uninterrupted open reading frame in env except for variant QB099.391I.ENV.C8, which had a frameshift mutation within the cytoplasmic tail of gp41. There was significant heterogeneity in the length of the protein variable loops, particularly V1/V2, which ranged from 57 amino acids (aa) to 113 aa (Table (Table1).1). The V3, V4, and V5 loops also varied in length, though less dramatically (Table (Table1).1). Variants from the same subject were generally similar in their variable-loop lengths. Moderate variation was also observed in the number and position of potential N-linked glycosylation sites (PNGS) (Table (Table11).Previous analyses indicated that early subtype C env proteins had shorter variable loops than did early subtype B env proteins (13), suggesting that there are different env protein features between subtypes. Thus, to compare variable-loop lengths and the numbers of PNGS between subtypes using this expanded group of early env variants, we evaluated the 31 newly cloned variants plus an additional 15 subtype A variants (2), 19 subtype B variants (14), and 18 subtype C variants (15) from other early virus panels. In order to avoid bias, when more than one env variant was available from a subject, the average loop length or PNGS number for that subject''s env proteins was used. We did not observe significant differences in V1/V2 length, V5 length, or the numbers of PNGS between subtypes by the Kruskal-Wallis equality-of-populations rank test (Table (Table2)2) . However, there were significant differences between the V3 and V4 loop lengths of the subtypes after adjusting for multiple comparisons (Table (Table2).2). The differences in V3 length appeared to be a result of shorter V3 loops in subtype D env proteins than in early subtype B (P = 0.006) or C (P < 0.001) env proteins (Table (Table2).2). The differences in V4 length were caused by shorter V4 loops in subtype C env proteins in comparison to both subtype A and B env proteins (P < 0.001; Table Table22).

TABLE 2.

Summary of variable-loop lengths and the numbers of PNGS in gp120 and gp41 within early HIV-1 env variantsa
ParameterMedian value (25th percentile, 75th percentile) for subtype:
Kruskal- Wallis P valuebWilcoxon rank sum P values for individual comparisonsc
A (n = 11)B (n = 19)C (n = 20)D (n = 4)A vs. BA vs. CA vs. DB vs. CB vs. DC vs. D
Length
    V1/V270.3 (62, 76)70 (66, 70)65 (62, 76)66.5 (62, 69)0.210.7300.2820.2150.0510.1130.846
    V335 (34, 35)35 (35, 35)35 (34, 35)34 (34, 35)0.0010.2400.0160.1070.1410.006<0.001
    V432 (30, 36)33 (31, 34)26.5 (22, 29)29.5 (29, 31)0.00010.880<0.0010.148<0.0010.0230.056
    V511 (11, 11)10 (9, 11)10 (9, 11)11.5 (11, 12)0.0300.0960.0150.1840.6770.0990.021
No. of PNGS in:
    gp12024 (23, 28)25 (24, 26)24 (23, 25)26 (26, 27)0.200.6800.6920.2650.1460.1860.042
    gp414 (4, 5)5 (4, 5)5 (4, 5)4.5 (4, 5)0.200.0300.1790.4700.4100.4080.799
    gp41ecto4 (4, 4)4 (4, 4)4 (4, 5)4 (4, 4)0.0440.1070.0250.5500.0880.5070.201
Open in a separate windowaVariable-loop lengths and the numbers of PNGS in gp120 and gp41 within early HIV-1 env variants from subtypes A, B, C, and D characterized here and previously (2, 14, 15). n, number of samples.bKruskal-Wallis equality-of-populations rank test (based on multiple comparisons; P values of <0.007 were considered significant; significant values are presented in boldface).cWilcoxon rank sum test (based on multiple comparisons; P values of <0.008 were considered significant; significant values are presented in boldface).We then assessed the neutralization sensitivity of the pseudoviruses to antibodies in plasma from HIV-1-infected individuals and to HIV-1-specific MAbs by using the TZM-bl neutralization assay as described previously (2, 23, 38). Median inhibitory concentrations (IC50s) were defined as the reciprocal dilution of plasma or concentration of MAb that resulted in 50% inhibition of infection (2, 38). The Kenya pool was derived by pooling plasma collected between 1998 and 2000 from 30 HIV-1-infected individuals in Mombasa, Kenya, and the other three pools were derived by pooling plasma collected between 1993 and 1997 from 10 individuals from Nairobi, Kenya, and with an infection with a known subtype (A, C, or D) of HIV-1 as described previously (2).The env variants demonstrated a range of neutralization sensitivities to plasma samples, from neutralization resistant (defined as <50% neutralization with a 1:50 dilution of plasma) to neutralization sensitive with an IC50 of 333 (Fig. (Fig.2).2). Some clones, such as QF495.23M.ENV.A1, were relatively sensitive to all the plasma pools, with IC50s from 100 to 333, whereas other clones, such as QH343.21M.ENV.A10, were relatively resistant to these plasma pools, with IC50s from <50 to 85 (Fig. (Fig.2).2). The plasma pools did differ in their neutralization potencies. The Kenya pool, with a median IC50 of <50 across all viruses tested, was significantly less likely to neutralize these transmitted variants than were the subtype A, C, and D plasma pools, which had median IC50s of 110, 105, and 123, respectively (P values of <0.0001, 0.0001, and 0.001, respectively, by paired t test on log-transformed IC50s). The basis for these differences in neutralizing activity is not clear, although the location, timing, and level of immunodeficiency at the time of sample collection could have contributed to the differences in NAb levels between the pools.Open in a separate windowFIG. 2.Neutralization sensitivity of early subtype A, C, D, and A/D recombinant env variants to plasma samples and MAbs in relation to the sequences of the MAb binding sites. The env used to generate the pseudovirus tested is shown at the left, and the plasma pool or MAb tested is indicated at the top. The IC50s of each plasma sample or MAb against each viral pseudotype is shown, with darker shading indicating more potent neutralization, as defined at the bottom of the figure. Gray boxes indicate that <50% neutralization was observed at the highest dilution of plasma or concentration of MAb tested. Each IC50 shown is an average of the results from two independent neutralization assays, using pseudovirus generated in independent transfection experiments. The median IC50s from the 31 variants are shown at the bottom. Neutralization of the pseudovirus derived from the subtype B variant SF162 is shown as a control, and neutralizations of murine leukemia virus (MLV) and simian immunodeficiency virus clone 8 (SIV) are shown as negative controls. In the panels on the right, the sequences for the MAbs 2G12, 2F5, and 4E10 are displayed. For 2G12, the amino acid numbers for the five PNGS that are important for 2G12 binding are shown for each virus tested. A plus sign indicates that the PNGS at that site in the envelope sequence was preserved, and a minus sign indicates that the PNGS was deleted. A shift in the PNGS position is indicated by the amino acid position to which the PNGS shifted. All sequences were numbered relative to the HXB2 sequence. The two rightmost panels show data for the canonical 2F5 and 4E10 epitopes, with a period indicating that the amino acid is preserved.The env variants were significantly more susceptible to their subtype-matched plasma pool, with a higher mean IC50 for subtype-matched plasma samples than for unmatched plasma samples (138 versus 108, P = 0.0081, paired t test). However, a significant amount of cross-subtype neutralization was observed, as every env variant that was susceptible to the subtype-matched plasma pool was also susceptible to at least one of the other plasma pools (Fig. (Fig.2).2). Thus, although potency was enhanced when the plasma antibodies were produced in response to infection with the same subtype of HIV-1, there were shared neutralization determinants between subtypes, as has been observed previously (reviewed in reference 3).To identify potential correlates of neutralization sensitivity to the antibodies within these plasma pools, we included these 31 env variants and an additional 15 subtype A env variants we previously characterized from the same cohort with the same plasma pools (2). We did not observe a change in neutralization sensitivity during the evolution of the HIV-1 epidemic in Kenya, as no correlation was observed between neutralization sensitivity and the calendar date from which the env variants were isolated. In addition, no correlation was observed between the neutralization sensitivity of a variant to the plasma pools and the duration of estimated infection within that individual. Finally, there was no significant correlation between the neutralization sensitivity and variable-loop length or the number of PNGS. Thus, although changes in the variable-loop length or number of PNGS may alter the exposure of epitopes within the HIV-1 env protein, these changes do not appear to be the primary determinant of neutralization sensitivity.Despite relatively universal sensitivity to at least one of the pooled plasma samples, these transmitted Kenyan env variants were generally resistant to the MAbs 2G12 (provided by Hermann Katinger, Polymun Scientific) and b12 (provided by Dennis Burton, The Scripps Research Institute), as well as 2F5 and 4E10 (obtained from the AIDS Research and Reference Reagent Program, National Institutes of Health) (Fig. (Fig.2),2), though these MAbs neutralized the subtype B env variant SF162, with IC50s similar to those reported previously (1). Subtype D strains were the most susceptible to MAbs, with 4/8 variants neutralized with <20 μg/ml of 2F5 and 2/8 neutralized with <20 μg/ml of the other MAbs. This could reflect the fact that subtype D variants are more closely related to subtype B strains (Fig. (Fig.1)1) (see reference 10), and these MAbs were all derived from subtype B-infected individuals.Among all 31 variants, 2F5 was the most broadly neutralizing, with 15/31 variants from 8/14 subjects neutralized with <20 μg/ml of this MAb. Some 2F5-resistant env variants, such as QH209.14M.ENV.A2 and QB857.110I.ENV.B3, had mutations in the canonical 2F5 binding epitopes, though other 2F5-resistant env variants such as QF495.23M.ENV.A3 and QA790.204I.ENV.A4 maintained the canonical 2F5 epitope. The results with the MAb 4E10 were similar; 4E10 neutralized only seven variants from 4 of the 14 subjects, and the presence of mutations in the 4E10 epitope, which were common, did not predict neutralization sensitivity (Fig. (Fig.2).2). For instance, the env variants QH343.21M.ENV.A10 and QH343.21M.ENV.B5 contained identical N671S and D674S mutations and QH343.21M.ENV.B5 was highly sensitive to 4E10, while QH343.21M.ENV.A10 was resistant (Fig. (Fig.2).2). Thus, for the 2F5 and 4E10 epitopes, the presumed epitopes appear to be shielded in a subset of these early non-subtype B env variants, as has been previously observed (Fig. (Fig.2)2) (1, 2, 5, 14).The MAb b12 neutralized only two variants from two subtype D-infected individuals, with no neutralization of the subtype A, C, and A/D recombinant pseudoviruses. Only four variants from two subjects were neutralized by 2G12 at <20 μg/ml, and these were the only variants that maintained all five of the PNGS within the 2G12 epitope (Fig. (Fig.2).2). Overall, the median IC50 of all the MAbs against these transmitted variants was >20 μg/ml. None of the variants was susceptible to all four MAbs (Fig. (Fig.2),2), unlike many of the early subtype B env variants characterized previously (14).In summary, these newly characterized HIV-1 env clones represent a range of neutralization sensitivities and can be used to supplement existing panels of transmitted variants, in particular, adding the first subtype D and A/D recombinant variants. Some differences between subtypes in env structure following transmission were noted, though these differences did not correlate with neutralization sensitivity. Although the significant levels of cross-subtype neutralization sensitivity observed with plasma samples indicate that some neutralization determinants were shared across subtypes, the epitopes for the MAbs b12, 2G12, 2F5, and 4E10 did not appear to be among the shared determinants. Thus, despite the fact that significant attention has focused on using vaccination to develop antibodies that resemble these MAbs in their specificity, such antibodies may not neutralize the transmitted strains that are causing most new infections worldwide. These data therefore stress the importance of evaluating transmitted variants in endemic areas when designing immunogens and evaluating vaccine and microbicide strategies.  相似文献   

7.
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   

8.
9.
10.
11.
The magnitude and character of adenovirus serotype 5 (Ad5)-specific T cells were determined in volunteers with and without preexisting neutralizing antibodies (NAs) to Ad5 who received replication-defective Ad5 (rAd5)-based human immunodeficiency virus vaccines. There was no correlation between T-cell responses and NAs to Ad5. There was no increase in magnitude or activation state of Ad5-specific CD4+ T cells at time points where antibodies to Ad5 and T-cell responses to the recombinant gene products could be measured. These data indicate that rAd5-based vaccines containing deletions in the E1, E3, and E4 regions do not induce appreciable expansion of vector-specific CD4+ T cells.Replication-defective adenoviruses (rAd) have been engineered to provide high levels of expression of foreign inserts with minimum expression of adenovirus proteins, making them excellent candidates for vaccine and gene therapy applications (3, 16). Despite promising immunogenicity, a prophylactic vaccine trial of a serotype 5 rAd (rAd5) vector expressing human immunodeficiency virus (HIV) Gag, Pol, and Nef genes (Step trial) was recently halted due to an increase in HIV infections among volunteers who had preexisting neutralizing antibodies (NAs) to Ad5 (7). This finding raises the possibility that the presence of Ad5-specific T-cell responses (specifically CD4+ T-cell responses) in subjects with preexisting Ad5 NAs could be boosted by rAd5 vaccines, thereby providing an expanded susceptible target cell population that could be more easily infected by HIV. If this mechanism were operative, it would have broad implications for the future use of rAd viruses, and indeed other virus vectors, as vaccines or therapeutic agents within HIV-susceptible populations (2, 12, 15). We therefore measured the frequency, magnitude, and activation status of rAd5-specific T cells in HIV-uninfected volunteers who had received rAd5-based HIV vaccines in the presence or absence of preexisting NAs to Ad5.We studied 31 volunteers enrolled in two NIAID Institutional Review Board-approved phase I clinical trials of rAd5-based HIV vaccines. VRC 006 was a dose escalation study evaluating a single inoculation of a rAd5 mixture expressing EnvA, EnvB, EnvC, and fusion protein Gag/PolB at 109, 1010, and 1011 total particle units (10). VRC 008 evaluated DNA priming by needle and syringe or Biojector, followed by rAd5 boosting. Both studies enrolled healthy, HIV-uninfected adults; used the same rAd5 products; and evaluated immunogenicity on the day of and 4 weeks after rAd5 immunization. Both of these trials involved rAd5 products that contained deletions in the E1, E3, and E4 regions (8, 10).NAs to Ad5 were determined for all volunteers as previously described (19). A 90% NA titer of 12 or more was considered positive and taken as evidence of preexisting humoral immunity to Ad5. Volunteers were chosen for assessment of Ad5-specific T-cell responses based upon the availability of peripheral blood mononuclear cell samples at key time points and the presence or absence of preexisting NAs to Ad5. Only volunteers who received the vaccine (not the placebo) were included. Table Table11 lists the volunteers who were tested for Ad5-specific T-cell responses and their NA titers to Ad5 before and after rAd5 vaccination. All volunteers, except for one (volunteer 12) who had a less-than-maximum NA titer to Ad5 before vaccination, had an increase in titer by 4 weeks after vaccination, indicating the successful “take” of the rAd5-based vaccine. There was no correlation between rAd5 dose and increase in Ad5 NA titer.

TABLE 1.

Ad5 serostatus before and after vaccination
VolunteerPrior DNA immunizationrAd5 dose (PUa)Ad5 NA titer
PrevaccinePostvaccine
1No1011<12739
2No1011<12834
3No1011<124,787
4No1011<12806
5No1011<121,033
6No1010<12130
7No1010<121,354
8Yes1010<121,387
9Yes1010<12575
10Yes1010<12170
11Yes1010<12>8,748
12Yes1010<12<12
13No101130>8,748
14No10946>8,748
15No10970328
16No1010176>8,748
17No10104786,198
18No1092,472>8,748
19No1093,502>8,748
20No10104,820>8,748
21No1095,078>8,748
22No10116,162>8,748
23No109>8,748>8,748
24No1011>8,748>8,748
25Yes1010643>8,748
26Yes1010942>8,748
27Yes10101,510>8,748
28Yes10101,611>8,748
29Yes10102,934>8,748
30Yes1010>8,748>8,748
31Yes1010>8,748>8,748
Open in a separate windowaPU, particle units.HIV-specific T-cell responses were measured by multiparameter flow cytometry after 6 h of stimulation with peptides (15-mers overlapping by 11) corresponding to the HIV EnvA protein (one of the vaccine inserts expressed in the Ad5 vectors), as previously described (13). Overlapping peptides corresponding to the major Ad5 surface protein (hexon), the Ad5 early regulatory protein (E2A), and Ad5 ORF1, -2, and -3 proteins were used to assess Ad5-specific T-cell responses, and additional markers of cell viability (ViViD), T-cell memory (CD45RO and CD27), and activation/division (CCR5, CD38, HLA-DR, and Ki67) were added to the panel for these assessments. Antibodies and fluorochromes used in this panel were CCR5-Cy7-phycoerythrin (PE), CD38-allophycocyanin, Ki67-fluorescein isothiocyanate, and CD3-Cy7-allophycocyanin, all from BD PharMingen; CD8-Cy55-PE from BD Biosciences; CD27-Cy5-PE and CD45RO-Texas Red-PE, both from Beckman Coulter; CD4-Cy5.5-PE from Caltag; CD14- and CD19-PacificBlue, CD57-QDot545, and HLA-DR-Alexa680, conjugated according to standard protocols [http://drmr.com/abcon/index.html]); gamma interferon-PE and interleukin-2-PE from BD Biosciences; and a violet amine dye from Invitrogen. Cells were analyzed on an LSRII instrument (Becton Dickinson), and data analysis was performed using FlowJo, version 8.1.1 (TreeStar). The gating strategy is shown in Fig. Fig.1A1A.Open in a separate windowFIG. 1.CD4+ and CD8+ T-cell responses to Ad5. (A) Gating tree used to determine antigen-specific T-cell frequencies. Single CD3+ ViViD CD14 CD19 cells were gated on CD4 or CD8 cells. Naïve CD27+ CD45RO cells were gated out, and the frequency of cells expressing gamma interferon (IFNg) and/or interleukin-2 (IL2) was determined. FSC-A, forward scatter area; FSC-H, forward scatter height; SSC-A, side scatter area. (B) Frequencies of CD4+ and CD8+ T-cell responses after stimulation with Ad5 hexon or E2A peptides were plotted against the prevaccination Ad5 NA titer. The prevaccine T-cell response was used. (C) Frequencies of CD4+ and CD8+ T-cell responses to Ad5 hexon, E2A, and HIV EnvA before and 4 weeks after rAd5 vaccination are shown for subjects with (Ad5 NA titer of >12) and without (Ad5 NA titer of <12) preexisting NAs to Ad5. Boxed areas represent interquartile ranges, and horizontal lines represent medians.Previously, we had found no T-cell responses to Ad5 ORF1, -2, or -3, so data from these antigen stimulations are not shown. As shown in Fig. Fig.1B,1B, T-cell responses to Ad5 hexon and E2A were detected, but there was no association between the NA response to Ad5 and the T-cell responses to these Ad5 proteins. Volunteers with an absence of NAs to Ad5 often had very strong CD4+ and CD8+ T-cell responses to Ad5 proteins. This probably reflects the degree of protein sequence homology between different adenovirus serotypes (11) and suggests that T-cell responses to adenoviruses may be significantly cross-reactive, while NAs are serotype specific. It also indicates that the NA response to Ad5 cannot be used as a surrogate for either a CD4+ or a CD8+ T-cell response to that adenovirus serotype.We next asked whether Ad5-specific T-cell responses were boosted by a single rAd5 vaccination in subjects with or without preexisting NAs to Ad5. At the time point 4 weeks after vaccination, there was clear evidence of boosting of the insert-specific (EnvA) CD4+ and CD8+ T-cell responses in volunteers with and without preexisting NAs to Ad5 (Fig. (Fig.1C).1C). The results of the Ad5-specific responses were consistent across volunteers who had received prior DNA immunization (VRC 008) and those who had not (VRC 006), so the results are combined in Fig. Fig.1C1C and show no increase in Ad5 hexon- or E2A-specific CD4+ T-cell responses after rAd5 immunization irrespective of Ad5 NA status. There was evidence of an increase in the CD8+ T-cell response to Ad5 hexon (P = 0.004 by paired t test), but not that to E2A, after rAd5 vaccination. These results, while showing evidence of adenovirus-specific CD8+ T-cell boosting by rAd5 vaccination, do not indicate an expansion of Ad5-specific CD4+ T cells that could serve as a substrate for HIV infection in subjects with or without NAs to Ad5.Having failed to demonstrate an expansion of Ad5-specific CD4+ T cells after vaccination, we assessed whether the activation profile of the unexpanded Ad5-specific CD4+ T cells was changed by vaccination. The gating tree is shown in Fig. Fig.2A.2A. Ad5 hexon- and E2A-specific CD4+ T cells expressed activation markers CCR5, CD38, and HLA-DR and a marker of recent cell division, Ki67, more frequently than did total memory CD4+ T cells (Fig. (Fig.2B).2B). However, none of these markers were significantly increased on total or Ad5-specific CD4+ T cells after vaccination in volunteers with or without preexisting NAs to Ad5.Open in a separate windowFIG. 2.Vaccine-induced activation of Ad5-specific CD4+ T cells. (A) Total CD4+ memory cells or Ad5-specific CD4+ memory cells (as gated in Fig. Fig.1A)1A) were further defined by expression of Ki67, CD38, CCR5, and HLA-DR. (B) Percentages of Ad5 hexon-specific cells, E2A-specific cells, or total memory CD4+ T cells that express CCR5, CD38, HLA-DR, or Ki67 before and 4 weeks after rAd5 vaccination are shown for subjects with (Ad5 NA titer of >12) (left) and without (Ad5 NA titer of >12) (right) preexisting NAs to Ad5. The phenotype was assessed only for those responders for whom at least 10 cytokine-positive events were counted. None of the comparisons of pre- and postvaccination marker expression were significant at a P value of 0.02 by paired t test. Boxed areas represent interquartile ranges, and horizontal lines represent medians.Expansion of Ad5-specific T cells after rAd5-based vaccination or gene therapy has been reported by others (14, 20, 21). Those studies evaluated Ad5-specific responses to rAd5 vectors with only the adenovirus E1 gene deleted (as used in the Step trial vaccines). The vectors used here contained deletions of the adenovirus E1, E3, and E4 genes (8, 10). While adenovirus gene deletions can render the vectors replication defective (6, 9), they do not necessarily completely shut off all adenovirus protein expression (20, 21). To demonstrate the importance of E4 deletions in limiting expression of adenovirus gene products, we measured the level of adenovirus protein synthesis in infected A549 cells as previously described (1, 4, 5). Cells were infected with adenovirus vectors with E1 and E3 deletions or with E1, E3, and E4 deletions at the same multiplicity of infection (10 focus-forming units per cell). At 24 h postinfection, [35S]methionine was added for 1 h. Levels of total and adenovirus protein synthesis in the infected and mock-infected cells were compared (Fig. (Fig.3).3). Adenovirus early protein single-stranded DNA binding protein, as well as late gene products hexon, penton, and fiber, was immunoprecipitated, fractionated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and resolved by autoradiography. The results show that the amount of newly synthesized adenovirus proteins in cells infected with adenovirus with E1, E3, and E4 deletions is significantly lower than that for an adenovirus vector with E1 and E3 deletions. Therefore, our inability to detect a vaccine-induced increase in the frequency and character of the Ad5-specific T-cell response could relate to the very low levels of adenovirus proteins that were probably expressed in vivo by the rAd5 vectors with multiple deletions.Open in a separate windowFIG. 3.Ad5 protein expression in vitro after infection with different Ad5 vectors. A549 cells were infected with adenovirus vectors with E1 and E3 deletions or with E1, E3, and E4 deletions and [35S]methionine labeled, and levels of total and adenovirus protein synthesis in the infected and mock-infected cells were compared after sodium dodecyl sulfate-polyacrylamide gel electrophoresis and autoradiography. Markers for the adenovirus early protein single-stranded DNA binding protein (DBP) and capsid proteins hexon, penton base, and fiber are shown.We were therefore unable to demonstrate (i) that Ad5-specific CD4+ T cells were restricted to subjects with preexisting Ad5 NAs, (ii) that rAd5 vaccination expanded or increased the activation of Ad5-specific CD4+ T cells, or (iii) that there was a substantial effect on the magnitude or character of the Ad5-specific CD4+ T-cell response to vaccination based upon preexisting NAs to Ad5. While the kinetics of Ad5-specific T-cell responses after rAd5-based vaccination are not known, it is clear that insert-specific responses are increased at 4 weeks after vaccination and subsequently contract (10). It is therefore reasonable to assume that if Ad5-specific responses were similarly affected, they would be detected at the 4-week-postvaccination time point.It is possible that rAd5 vaccines expand a preexisting mucosal T-cell response to Ad5 that is not reflected within the blood. While we do not have mucosal samples from our vaccine volunteers to directly address this possibility, it is likely that expansion of a mucosal response would be reflected to some degree within the blood.The mechanism underlying the increase in HIV infections in vaccinees with NAs to Ad5 in the Step trial is yet to be determined (2, 7, 12, 15, 17). Confounding factors and alternative hypotheses have recently been proposed to account for the increased acquisition (7, 12, 15, 18). Until there is a better understanding of the processes involved, future studies of rAd5-based products should proceed with appropriate safety considerations and monitoring of adenovirus-specific responses. In addition, the use of vaccine regimens involving single injections of vectors with multiple deletions may help mitigate risk.  相似文献   

12.
13.
Simian immunodeficiency virus (SIV) infection of natural-host species, such as sooty mangabeys (SMs), is characterized by a high level of viral replication and a low level of generalized immune activation, despite evidence of an adaptive immune response. Here the ability of SIV-infected SMs to mount neutralizing antibodies (Nab) against autologous virus was compared to that of human immunodeficiency virus type 1 (HIV-1) subtype C-infected subjects. While high levels of Nab were observed in HIV-1 infection, samples obtained at comparable time points from SM exhibited relatively low titers of autologous Nab. Nevertheless, SM plasma with higher Nab titers also contained elevated peripheral CD4+ T-cell levels, suggesting a potential immunologic benefit for SMs. These data indicate that AIDS resistance in these primates is not due to high Nab titers and raise the possibility that low levels of Nab might be an inherent feature of natural-host SIV infections.More than 40 species of African nonhuman primates (NHPs) naturally harbor CD4+-tropic lentiviruses that are collectively known as simian immunodeficiency viruses (SIVs) and represent the ancestors of the human pathogens human immunodeficiency virus type 1 (HIV-1) and HIV-2. Interestingly, African NHPs infected with their cognate SIV generally do not progress to AIDS, despite high levels of sustained virus replication, with the only known exception being chimpanzee SIV (SIVcpz)-infected chimpanzees (16). Among the natural hosts for SIV infection, the sooty mangabey ([SM] Cercocebus atys) is of particular interest, because cross-species transmission of SM SIV (SIVsm) from this natural host into humans initiated the HIV-2 epidemic in West Africa (17). In addition, SIVsm (herein referred to as SIV) is the ancestor of the rhesus macaque SIV (SIVmac) viruses that are used in disease pathogenesis and vaccination studies in the rhesus macaque model (17). Both naturally infected and experimentally inoculated SMs remain healthy, maintain CD4+ T cells, and do not progress to AIDS-like disease, despite sustained high levels of virus replication (31).Nonpathogenic infection of SMs is characterized by low levels of immune activation during the chronic phase of infection, which are reached after a transient immune activation that occurs during primary infection (reviewed in reference 31). These findings have led to the hypothesis that the absence of generalized immune activation in SIV-infected SMs during the chronic phase of infection is an important feature that favors the preservation of CD4+ T-cell homeostasis, thereby avoiding disease progression (31). However, most of these earlier studies focused on T cells and innate immune cells, with a significant gap existing in our understanding of whether humoral immunity might also differ between pathogenic and nonpathogenic infections. In HIV-1-infected patients, B cells produce neutralizing antibodies against the infecting (autologous) virus, which drives viral escape, continuous de novo antibody production (26-28, 32), and B-cell dysfunction (24). The striking differences in both the clinical outcomes of infection and the levels of immune activation between SIV-infected SMs and HIV-1-infected humans prompted us to compare the neutralizing antibody (Nab) response against the autologous virus in these two populations. To this end, we utilized a pseudovirus assay that has been used extensively by our group and others to evaluate Nab against HIV-1 and SIV envelope (Env) glycoproteins (15, 19, 22, 26, 28, 32, 33; also unpublished data). All SMs were housed at the Yerkes National Primate Research Center (Atlanta, GA) and maintained in accordance with National Institutes of Health guidelines. The Emory University Animal Care and Use Committee approved these studies. Details of the Zambia Emory HIV Research Project (ZEHRP) have been described elsewhere (2, 10, 21). The Emory University Institutional Review Board and the University of Zambia School of Medicine Research Ethics Committee approved informed-consent and human subject protocols. None of the subjects received antiretroviral therapy during the evaluation period.In HIV-1 infection, autologous Nabs develop to relatively high titers against the newly transmitted virus within the first few months (15, 19, 26-28, 32). Here we sought to test whether a similar increase in Nab titer occurs during nonpathogenic SIV infection of SMs. Samples were obtained from five animals that were inoculated intravenously with plasma from a naturally infected SM as part of a previous study (30). Multiple, biologically functional Envs were cloned from plasma collected at day 14 postinoculation (Table (Table1),1), and Nab activity was evaluated in plasma collected at 6 months postinoculation. To facilitate comparison with early HIV-1 infection, Nab activity in plasma was also evaluated between 2 and 9 months against Envs that were cloned between 31 and 88 estimated days after infection from four subtype C HIV-1-infected seroconverters in Zambia (Table (Table1).1). Figure Figure1A1A demonstrates that Nab activity in plasma diluted 1:100 was readily detectable in all HIV-1-infected subjects at levels approaching 100% neutralization. However, Nab activity in the SM plasma was significantly lower than in the human subjects (median, 10% versus 93%, respectively; P = 0.02). Binding antibody was detected in all five SMs at titers greater than 1:51,200 by enzyme-linked immunosorbent assay (ELISA), demonstrating that all monkeys had seroconverted by 6 months and maintained high titers of binding antibody throughout the evaluation period (Fig. (Fig.1B).1B). Thus, the low level of Nab was not due to a diminished humoral immune response.Open in a separate windowFIG. 1.Autologous Nab activity and B-cell proliferation during experimental infection of SMs. (A) Neutralization activity levels in plasma from five SMs (filled black circles), which were experimentally inoculated with plasma from a naturally SIV-infected SM, and four HIV-1-infected Zambian subjects (half-filled squares), who were recently infected through heterosexual contact, are shown. The horizontal bars represent the median for each group. To assess neutralizing activity, pseudoviruses were created by expressing each cloned Env with an HIV-1 env-deficient backbone (ΔSG3). JC53-BL (Tzm-bl) cells were infected with each pseudovirus in the presence or absence of serially diluted autologous plasma. Each point represents the average level of neutralization at a 1:100 dilution of plasma for at least two Env clones (see Table Table11 for number of Envs tested). Each neutralization assay was performed twice independently, using duplicate wells. Statistical significance between the groups was determined by a Mann-Whitney test, using GraphPad Prism 5. Longitudinal measurements of endpoint antibody ELISA titers in plasma (filled green circles) (23) (B), autologous neutralization activity in plasma (filled blue diamonds) (C), percentages of Ki-67+ CD20+ cells in blood (filled black triangles) (D), and percentages of CD20+ cells in blood (filled red squares) (E) are shown for the five experimentally inoculated SMs combined. In panel C, each point represents average neutralization at a 1:100 dilution of plasma over time for at least two day 14 Env clones from each SM. For panels D and E, PBMCs were gated by forward and side scatter, and the CD3 CD20+ population was assessed for Ki-67 staining (D) by flow cytometry. SP34-2 was used to stain CD3, L27 was used for CD20, and B56 was used for Ki-67 (all from BD Biosciences). Error bars represent the standard errors of the means (SEMs). Plasma viral load peaked at day 14 (data not shown). Filled symbols in panels A through E indicate data generated from experimentally infected SMs.

TABLE 1.

Autologous Nab activity in experimentally SIV-infected SM and acutely HIV-1-infected humans
Subject IDaVirusNo. of mo postinfection Nab activity was evaluatedNo. of days postinfection Envs were cloned from plasmaNo. of Envs tested% neutralization at a 1:100 dilution of plasma
FuvSIVsm-Fuo614416.3
FSsSIVsm-Fuo614310.6
FWvSIVsm-Fuo614510.5
FFsSIVsm-Fuo614210.3
FRsSIVsm-Fuo61439.3
185FHIV-1533494.6
153MHIV-1988594.3
221MHIV-1631691.5
205FHIV-1248587.1
Open in a separate windowaID, identification.The low level of Nab activity observed in the five experimentally inoculated SMs persisted for 16 months and did not exceed 50% at a 1:100 dilution of plasma at any time point tested (Fig. (Fig.1C).1C). In contrast, the high levels of Nab activity in the HIV-1-infected subjects persisted for over 2 years, often exceeding 50% inhibitory titers of 1:3,000 against the early virus, as is characteristic of early subtype C HIV-1 infection (15, 19, 26, 28). Figure Figure1D1D demonstrates that a transient increase in proliferating B cells, as measured by positive Ki-67 staining (12), occurred in the SMs and peaked around day 30 postinfection and then declined to a level just above baseline by day 60. Analysis using a Wilcoxon signed-rank test for paired samples showed that the percentages of Ki-67-positive (Ki-67+) B cells were higher at days 21 and 30 than at day −5, reaching borderline significance at both time points (P = 0.06). In contrast, the percentages of Ki-67+ B cells on days 60 and 475 were not significantly different from that on day −5 (P = 0.8 and 0.3, respectively). An early but transient decrease in the percentage of circulating CD20+ B cells was also observed during the initial 20 days of infection (Fig. (Fig.1E).1E). Thus, the B-cell compartment within the SM underwent changes consistent with immune activation followed by resolution. Based on these results, it does not appear that a global defect in the B-cell response in the SM can account for the low-level Nab response elicited.To investigate Nab responses during established infection, we extended this analysis to a panel of 11 naturally SIV-infected SMs in the Yerkes colony and 5 chronically HIV-1-infected subjects in Zambia. Envs were cloned from these monkeys and human subjects using peripheral blood mononuclear cell (PBMC) DNA or plasma samples, and sensitivity to Nab was evaluated. Because Nab activity against contemporaneous Env is often low or undetectable in HIV-1 infection (1, 5, 14, 25, 27, 28, 32), we evaluated plasma collected between 6 and 55 months after the Envs were cloned from each individual. Table Table22 shows that the SM Envs reflected the four SIV subtypes that circulate in the Yerkes colony (3). Figure Figure2A2A demonstrates that Nab activity in the chronically HIV-1-infected subjects was high (median, 91%), whereas in the naturally SIV-infected SMs it was again significantly lower (median, 14%; P = 0.003). Nevertheless, Nab activity in the naturally infected SMs exhibited a considerable range, from undetectable to 84% neutralization (Fig. (Fig.2A).2A). This observation prompted us to investigate whether parameters associated with disease progression in HIV-1 infection were correlated with the level of Nab activity. Figure Figure2B2B demonstrates that the number of CD4+ T cells was positively correlated with the potency of neutralization (r = 0.69; P = 0.02), while the plasma viral load showed a trend toward an inverse correlation with neutralization (Fig. (Fig.2C)2C) (r = −0.54; P = 0.08). A correlation between plasma viral load and autologous Nab titer in established HIV-1 infection has not been observed (9).Open in a separate windowFIG. 2.Autologous Nab activity and its correlation with CD4+ count and plasma viral load during established natural infection of SMs. (A) Neutralization activity levels in plasma from 11 naturally SIV-infected SMs in the Yerkes colony (open circles) and 5 chronically HIV-1-infected human subjects from Zambia (half-filled squares) are shown. Statistical significance between groups was determined by a Mann-Whitney test using GraphPad Prism 5. Correlation between Nab activity and CD3+ CD4+ T cell counts or plasma viral load in naturally infected SMs (open circles) is shown in panels (B) and (C), respectively. The percent neutralization at a 1:100 dilution of plasma (shown in panel A) is plotted along the x axis. Each CD4+ T cell count and viral load value represents the average of three measurements from samples collected from the 11 SMs approximately 1 year apart. The significance of each correlation was determined using a nonparametric Spearman test. Open circles indicate data from naturally infected SMs.

TABLE 2.

Autologous Nab activity in naturally SIV-infected SMs and HIV-1-infected humans with established infections
Subject IDaVirusEnv subtypeNo. of mo between plasma collection and Env cloningNo. of Envs tested% neutralization at 1:100 dilution of plasma
FWkSIVsm228584.4
FNnSIVsm131463.5
FFvSIVsm150459.7
FFmSIVsm130442.0
FNgSIVsm548428.0
FBnSIVsm349213.9
FDoSIVsm36512.0
FZoSIVsm12838.8
FOhSIVsm1624.7
FPnSIVsm13250.6
FFjSIVsm15420.0
109MHIV-1C6591.4
55MHIV-1C15891.3
135FHIV-1C16497.4
106MHIV-1C17579.4
153FHIV-1C55599.0
Open in a separate windowaID, identification.This study is the first to directly compare the Nab response against the autologous virus in nonpathogenic SIV versus HIV-1 infection, including evaluation of both the early, developing Nab response in acute infection and the mature response in chronic infection. A significant difference in the magnitude of Nab activity was apparent during both early and later time points, with relatively strong but ultimately ineffective neutralization activity developing and persisting into chronic infection in humans but not in SMs. Although the SIV and HIV-1 samples were obtained during similar stages of infection, the disparity in the magnitude of autologous Nab activity during early infection could in part reflect differences such as the route of infection (intravenous versus mucosal) or the complexity of the founder virus (a single variant in HIV-1 versus multiple variants in SIV). In addition, the production of SIV Env pseudoviruses in human 293T cells could have altered the glycosylation pattern or the proteins that are embedded within the virion, decreasing the neutralization susceptibility of the SIV Env pseudoviruses. However, production of a subset of these pseudoviruses in an African green monkey-derived cell line (COS-1) did not alter their Nab sensitivity (data not shown).Despite the lack of potent autologous Nab, both naturally and experimentally SIV-infected SMs produce antibodies that bind Env in ELISAs or Western blotting (4, 6, 13, 18, 23). It is possible that the SIV Env glycoproteins elicit a different profile of Nab than does HIV-1 Env. The potential for structural and biological differences between SIV and HIV-1 Envs has not been thoroughly investigated, although they would not be unexpected due to the low level of amino acid sequence conservation between them. SIVsm/HIV-2 lineage-derived Envs (i.e., the SIVmac series) show a “wide evolutionary distance” and lack of cross-reactivity with SIVcpz/HIV-1-derived Envs, with an overall sequence identity in gp120 of ∼25% across HIV-1, HIV-2, and SIVsm (7, 8). Clear biological differences in immunogenicity have been described for HIV-1 group M subtypes, which all derive from a common SIV ancestor (reviewed in reference 20). Furthermore, SM IgG antibody molecules have less flexibility in the hinge region than human IgG, which could lead to a failure of the SM antibodies to recognize recessed neutralization targets such as the receptor binding domains (29). Thus, HIV-1 Env could elicit neutralizing antibodies that are qualitatively different from those induced by SIV Env.Early resolution of immune activation could be a key feature that distinguishes nonpathogenic from pathogenic infection (12, 31). The data presented here are consistent with that hypothesis, in that signs of early B-cell proliferation were present in the experimentally infected SMs but were resolved and did not result in potent neutralizing activity. However, later in infection, the naturally infected SMs did develop low-to-moderate levels of Nab activity, and these levels were positively correlated with the number of peripheral CD4+ T cells. This finding suggests that synergy between CD4+ T cells and B cells is maintained in this nonpathogenic setting. Other biologic factors could contribute to this correlation; however, differences in age and viral subtype in this cohort of SMs could not explain this finding (data not shown).Taken together, these results indicate that a low level of autologous Nab activity is a novel and previously unappreciated feature of nonpathogenic SIV infection of SMs. The fact that high-titer Nabs are not necessary to avoid disease progression during SIV infection of SMs is consistent with the notion that the apathogenicity of natural SIV infections is not the result of particularly effective adaptive immune responses against the virus (11). It is possible that this low level of autologous Nab activity in SMs stems in part from antibody recognition of targets that are poorly exposed on the native SIV Env glycoproteins. A low level of neutralizing activity in SM may therefore have a protective effect because it does not drive viral escape or induce chronic immune activation in the B-cell compartment. Moreover, a low level of immune activation in B cells and/or preservation of CD4+ T cells could enhance the quality of the neutralizing antibody response. It will be important, in future work, to assess how this low level of autologous Nab activity in SIV-infected SMs meshes with the lower levels of immune activation and dysregulation observed in these animals. Understanding the qualitative and quantitative differences in the Nab response during pathogenic versus nonpathogenic infection could provide critical information regarding protection from AIDS.  相似文献   

14.
HLA-B*57-mediated selection pressure leads to a typical escape pathway in human immunodeficiency virus type 1 (HIV-1) CD8 epitopes such as TW10. Whether this T242N pathway is shared by all clades remains unknown. We therefore assessed the nature of HLA-B*57 selection in a large, observational Kenyan cohort where clades A1 and D predominate. While T242N was ubiquitous in clade D HLA-B*57+ subjects, this mutation was rare (15%) in clade A1. Instead, P243T and I247L were selected by clade A1-infected HLA-B*57 subjects but not by HLA-B*5801+ subjects. Our data suggest that clade A1 consensus proline at Gag residue 243 might represent an inherent block to T242N escape in clade A1. We confirmed immunologically that P243T and I247L likely represent escape mutations. HLA-B*57 evolution also differed between clades in the KF11 and IW9 epitopes. A better understanding of clade-specific evolution is important for the development of HIV vaccines in regions with multiple clades.Human immunodeficiency virus type 1 (HIV-1) displays extreme genetic diversity, with nine clades (subtypes) described in group M, and frequent genomic recombination among and within the clades (7, 44). HIV is also capable of rapid evolution, which can lead to mutational escape from immune control (43). Escape from CD8+ T-cell responses occurs frequently in HIV-1 infection through mutations that affect epitope processing, HLA class I binding, and/or T-cell receptor recognition (23). In early HIV-1 infection, the majority of amino acid substitutions are associated HLA class I alleles (1). The timing and consequences of mutational escape from CD8+ T-cell responses vary considerably (8, 22).HLA-B*57, and to a lesser extent HLA-B*5801, has been associated with slower progression to AIDS in several studies (18, 27, 39), and HLA-B*5701 was associated with a lower viremia set point in a genome-wide association study (16). Several attributes of HLA-B*57-restricted CD8+ T-cell responses may contribute to their protectiveness, including dominant responses in acute infection (2), recognition of protective epitopes in HIV-1 p24 (33), better recognition of epitope variation (45), and retention of proliferative capability in chronic infection (24).HLA-B*57/5801 also exert powerful selection pressure on HIV to avoid CD8+ T-cell recognition. This was first demonstrated in the HLA-B*57-restricted TW10 epitope (TSTLQEQIGW [Gag240-249]), which accounts for >30% of overall HIV-specific CD8+ T-cell responses in acutely infected HLA-B*57+ subjects (3). Escape in this epitope usually occurs early in infection, which coincidently is when HLA-B*57 is most protective (18). In clade B and C infections, >75 to 100% of HLA-B*57/5801+ subjects develop the T242N escape mutation, while HLA-B*57/5701-negative subjects rarely display polymorphism at this residue (5, 9, 10, 15, 32, 35, 38, 41). When T242N is transmitted to HLA-B*57/5801-negative subjects, it rapidly reverts to the consensus, suggesting that T242N is associated with a fitness defect (32, 35).While CD8+ T-cell cross-clade recognition has been tested extensively (6, 11, 19, 36, 48), few studies have addressed the possibility of clade-specific escape from CD8+ T-cell responses. This may be especially relevant where clade consensus sequences differ in immunologically relevant epitopes. Here we demonstrate in a large Kenyan cohort substantial differences in HLA-B*57/B*5801-mediated selection among HIV clades.Participants were enrolled from a Nairobi, Kenya-based cohort, and the relevant ethical review boards approved the study. HLA typing was performed as described previously (34). CD4 counts were measured longitudinally at biannual visits. Multiple and other clade infections were excluded. The HIV-1 p24 gene was amplified from proviral HIV DNA or RNA using a nested PCR approach and sequenced, and viral subtyping was carried out as described previously (42). Previously described HLA-B*57 epitopes IW9 (ISPRTLNAW), KF11 (KAFSPEVIPMF), and TW10 (TSTLQEQIGW) and selected variants were tested in immunological assays and described where relevant. Gamma interferon enzyme-linked immunospot (ELISPOT) assays were performed as described previously (37) using blood samples from HLA-B*57+ and -B*5801+ subjects. All peptides were tested at concentrations of 10 μg, 1 μg, 0.1 μg, and 0.01 μg/ml. Responses were considered positive if they were more than two times higher than that of the negative control and were measured at ≥100 spot-forming units ml−1. Fisher''s exact test and chi-square analyses were used to determine differences among groups in categorical analyses. Mann-Whitney U tests were used to compare response magnitudes and disease progression between groups.We confirmed the protective effects of HLA-B*57 in clade A1 infection (mean of 9.9 years versus 7.8 years until CD4 counts were <200, P = 0.041) (Fig. (Fig.1).1). Slow progressors were overrepresented in HLA-B*57+ clade A1+ subjects (52.2%) compared to both HLA-B*5801+ clade A1+ (13.3%, P = 0.02) and HLA-B*57/5801-negative clade A1+ (27.8%, P = 0.028) subjects (Fig. (Fig.1b).1b). In contrast to what has been shown for other clades (2, 27), protection was not observed for clade A1-infected HLA-B*5801+ subjects (mean of 6.5 years versus 7.8 years until CD4 counts were <200, P > 0.3) (Fig. (Fig.11).Open in a separate windowFIG. 1.HLA-B*57, but not HLA-B*5801, is associated with a lower rate of disease progression in clade A1-infected subjects than that of the overall cohort. (a) Number of years from cohort entry until sequential CD4 counts fell below 200/μl. (b) Slow progressors (>10 years with CD4 counts of >200) were also more common in HLA-B*57+ clade A1-infected subjects than in those expressing HLA-B*5801 or neither.Stratification of TW10 (Gag240-249) proviral sequences on the basis of HLA allele and clade revealed several differences in selection between clades A1 and D (Fig. (Fig.2a).2a). We observed the expected T242N substitution in 100% of HLA-B*57+ clade D-infected subjects (7/7), compared to only 14.7% variability at Gag residue 242 in HLA-B*57/5801-negative subjects (13/88, P = 3.26 × 10−9) (Fig. (Fig.2b).2b). In contrast, T242N was found infrequently in clade A1-infected HLA-B*57+ subjects (15%, 5/33, P = 0.0004). Instead, variants containing the mutations P243T and I247L were more frequently observed (both observed in 11/33 subjects). Overall, variation at residues 243 and 247 was more common in HLA-B*57+ subjects (51% and 15%, respectively; P = 2.92 × 10−6) than in HLA-B*57/5801-negative clade A1+ subjects (33% and 9%, respectively; P = 0.0008). Selection at both residues 243 and 247 was observed only in 2/33 HLA-B*57+ subjects, suggesting that these mutations are independent. Selection at residue 248, observed in clade B infection (32), was not evident in either clade A1 or D. While I247X selection has been described in other clades at low frequencies and in elite controllers (21, 40), HLA-B*57-mediated selection at Gag residue 243 has not yet been described. In summary, the T242N mutation, which is typical of other clades, does not appear to be the primary escape mutant in clade A1.Open in a separate windowFIG. 2.HLA-B*57-mediated selection in TW10 differs between clade A1 and clade D. (a) TW10 sequences were stratified by HLA-B*57, HLA-B*5801, or other alleles (HLA-B*57/5801) and compared between clades A1 and D, based on the clade B consensus TW10 sequence. Each subject is represented by one sequence, and the numbers of subjects with a given sequence are shown in parentheses. A summary of variation from the TW10 consensus at Gag residues 242, 243, 247, and 248 is shown for HLA-B*57+ (b) and -B*5801+ (c) subjects. (b and c) Clade D is shown at the top, and clade A1 is shown at the bottom. Variation is shown in dark gray, and consensus is shown in light gray. (d) The proportions of clade A1-infected subjects with selection at Gag residue 242 only, and those with selection at residues 242 and 243 in combination, are shown. *, P < 0.05; **, P < 0.005; ***, P < 0.0005.Previous studies have suggested that HLA-B*5801 places selection pressure on TW10, similar to that of HLA-B*57 (35). Similar to clades B and C, selection of T242N was evident in HLA-B*5801+ clade D-infected subjects (TW10 variation in 8/11 HLA-B*5801+ subjects versus 13/88 HLA-B*57/5801-negative subjects; P = 0.0069) (Fig. (Fig.2c).2c). Limited T242N selection was observed in clade A1-infected HLA-B*5801+ subjects, and in contrast to HLA-B*57, there were no HLA-B*5801-associated substitutions at residues 243 and 247 in clade A1 (P values of 0.75 and 0.29, respectively) (Fig. (Fig.2c).2c). In summary, these data suggest that in addition to HLA-B*5801 not being associated with protection in clade A1 (Fig. (Fig.1),1), HLA-B*5801 does not select the HLA-B*57-associated clade A1 TW10 escape mutations.Inclusion of all clade A1 sequences with T242X substitutions (regardless of the HLA allele) reveals that in every case (10/10), there is an accompanying residue 243 mutation. Polymorphisms at these sites correlate very strongly (P = 3.71 × 10−8) (Fig. (Fig.2d).2d). Together, these data suggest that residue 242 polymorphism in clade A1 is incompatible with proline at residue 243, which is the clade A1 consensus.We next assessed the immunological implications of novel clade A1 variants in HLA-B*57+ (n = 12) and -B*5801+ (n = 6) subjects infected primarily by clade A1. Clade A1-infected subjects commonly made anamnestic, low-avidity responses to TW10. The majority of HLA-B*57+ subjects who recognized clade A1 TW10 did not respond to P243T or I247L in ELISPOT assays (Fig. (Fig.3),3), supporting the hypothesis that these represent escape mutations. Those who did recognize P243T and I247L had lower magnitude responses than those who recognized clade A1 TW10 at the 10-μg/ml peptide concentration (P of 0.0005 for both) (Fig. (Fig.3a).3a). Similarly, these variants were not well recognized by CD8+ T cells from HLA-B*5801+ subjects (Fig. (Fig.3b).3b). For two HLA-B*57+ subjects, P243T and I247L responses had lower avidity than clade A1 TW10 responses (Fig. (Fig.3c).3c). These data support the hypothesis that P243T and I247L likely represent escape mutations.Open in a separate windowFIG. 3.Peptides with novel TW10 clade A1-selected mutations are poorly recognized in ex vivo gamma interferon ELISPOT avidity assays, suggestive of escape mutations. ELISPOT responses to TW10 and variants at 10 μg/ml peptide by HLA-B*57+ (a) and HLA-B*5801+ (b) subjects are shown. (c) The functional avidity of TW10 and variants for two HLA-B*57+ subjects is shown, suggesting that P243T and I247L are less recognized than TW10, particularly at lower peptide concentrations. Sequence names are described in the text. (d) Elispot responses to A1 TW10 and T242N correlated at 10 μg/ml. SFU/m, spot-forming units/million PBMCs.Recognition of the clade B/D consensus (TSTLQEQIGW) was diminished compared to that of clade A1 TW10. However, despite the presumed absence of this variant in these subjects'' autologous sequences, the clade B/D escape variant (TSNLQEQIGW [T242N]) was recognized at a magnitude similar to that of the consensus clade A1 TW10 (r = 0.71, P = 0.0099) (Fig. (Fig.3d).3d). No responses to T242N/G248A were observed (not shown), as described previously (32). These data suggest that clade A1 and B TW10, and their escape variants, are immunologically distinct from one another.We next assessed whether clade-specific selection was evident in other immunodominant HLA-B*57 p24 epitopes that are commonly targeted in chronic clade B infection (2). In clade D IW9 (ISPRTLNAW [Gag147-155]), variants containing the escape variant I147L (14) were more common in HLA-B*57+ subjects than in HLA-B*57/5801-negative subjects (86% and 30%, respectively; P = 0.0055) (Table (Table1).1). However, this variant was not selected in clade A1 (variation in 30% versus 21% subjects; P value was not significant), where leucine is the consensus. Interestingly, ELISPOT data indicated substantial cross-reactivity between 147L and 147I in clade A1-infected subjects (10 μg/ml, r = 0.987, P < 0.0001) (data not shown), suggesting that infection with an escape variant from one clade (clade D) does not necessarily preclude recognition of this epitope in another one (clade A1). Other amino acids (F, M, and P) were common in HLA-B*57+ subjects at residue 147 (>30% versus 3% in HLA-B*57/5801-negative subjects, P = 3.85 × 10−6). Although the immunological consequences are unknown, these HLA-associated substitutions could represent novel escape variants.

TABLE 1.

p24 sequences in HLA-B*57+ subjects infected by clades A1 and D
CladeHLA-B alleleSubject no.No. of years infected prior to samplebPolymorphism at residue S146Epitope sequencea
IW9KF11TW10
LSPRTLNAWKAFSPEVIPMFTSTPQEQIGW
A157011665>130NF-----------------------------
570213921N----------------------NI------
59NDP----------------------NT---LA-
41>122T----------------------NV------
1419>41N------------------------------
616>102T---------------------------LA-
164718-P----------------------S---LQ-
613>33P-----------------------T------
718>71PM----------------------T---L--
1315>58P----------G------------TS---L--
57032125>0A------------N---------NL------
561>78P------------N-----------------
1926>39P------------c----------T------
1778>13-------------------------------
16090T---------------------------LQ-
525>135-----------G-n--------------L--
1368>57-M---------G-n----------T------
509>150TF---------G-N-----------------
260>163TF---------G-N--------------L--
111>133-----------G-N--------------LQ-
1111>34-----------G-N--------------LQ-
1638>27-----------G-N----------T------
2101>0P----------G-----------NT------
532>28-----------G-----------------A-
1669>23PP---------G-----------------A-
703>71TF---------G-------------------
1741>11TF---------G-------------------
1452>37P----------G-------------------
30>121P---------R------------------A-
1122>39P---------RG-Q-----------------
995>80PF--------RG-Q----------T------
1564>27P---------RG------------T------
5707330>164T----------G-N----------T------
D57011859P----------------------NL------
1852P----------------------NL---VA-
57031756P----------------------NL------
1423P-T--------------------NL------
1188PI---------G-N---------NL----A-
1894P----------N-----------NL----R-
199P-T--------S-----------NL---V--
Open in a separate windowaThe first row of epitope sequences shows the consensus sequences.bND, not done.In addition, a substitution at Gag residue 146 (A146P) represents an IW9 processing escape mutation in clades B and C (14), and this mutation was also selected by HLA-B*57 in both clades A1 and D (Table (Table1).1). In clade A1, substitutions at Gag residue 146 (primarily P and T) were more frequent in HLA-B*57+ subjects than in HLA-B*57/5801-negative subjects (13/33 and 10/221, respectively; P = 1.42 × 10−7) (Table (Table1).1). Therefore, although the consensus at residue 146 differs among clades, here escape at residue 146 occurs in HLA-B*57+ subjects infected by clades A1, B, C, and D.For KF11 (KAFSPEVIPMF [Gag162-172]), HLA-B*57-associated variation from the consensus was more common in clade A1 (67% versus 21%, P = 2.44 × 10−7) than in clade D (43% versus 17%, P = 0.012). Previously described A163G and A163G/S165N variants (13, 20) were most common in clade A1 (Table (Table1).1). In addition, the novel K162X substitution was present in clade A1. HLA-B*5703 and -B*5701 have previously been shown to display differences in KF11 selection (20, 47), and our data indicate that HLA-B*5702 also differs from HLA-B*5703 in terms of KF11 selection. While the KF11 consensus is present in the majority of HLA-B*5702+ subjects, it is rare in HLA-B*5703+ clade A1 infection (8/9 versus 2/22, P = 4.74 × 10−5).Mounting evidence suggests that HLA alleles are a major force in viral evolution (26). We show that in clade A1 p24, HLA-B*57 selection in three epitopes differs from earlier clade B and C data in several important aspects, while clade D selection resembles what has previously been shown. This included a low frequency of T242N in clade A1 TW10, with selection being more common at Gag residues 243 and 247, more extensive KF11 escape, and selection of different amino acids in IW9. Overall, selection was evident in the majority of HLA-B*57+ subjects (>90% of clade A1-infected subjects had selection in more than one epitope, and >75% of them had selection in more than two) (Table (Table1).1). Parallel escape in multiple epitopes demonstrates the need to avoid the pressure of CD8+ T-cell responses.One possible mechanism underlying the differences in TW10 selection is that TSNPQEQIGW (never observed) (Fig. (Fig.2d)2d) is not feasible virologically, such that T242N is possible only in conjunction with a preexisting residue 243 mutation (TSNXQEQIGW, observed in 15% of HLA-B*57+ subjects) (underlining shows mutation). One would expect to observe T242N at a higher frequency, given its dominance in HLA-B*57+ subjects infected by other clades. Therefore, while T242N has been implicated in HLA-B*57-mediated protection, this mutation is rare in clade A1. Because HLA-B*57 remains protective in clade A1, that protection may be mediated by novel mechanisms.Because TW10 is commonly recognized by 86% of clade A1 subjects, it is evident that clade A1 TW10 can bind HLA-B*57. We therefore hypothesize that TSTPQEQIGW may affect the interaction between epitope and cognate T-cell receptors, which in turn influences which escape mutations are optimal. This hypothesis is supported by our immunological data showing cross-reactivity between clade A1 TW10 and TSNLQEQIGW (underlining shows mutation), which imply that mutation at residue 242 may not lead to effective escape in clade A1 (Fig. (Fig.3d3d).In contrast to other clades (including clade D), HLA-B*5801 does not appear to place selection pressure on clade A1 TW10. A previous study in Rwanda similarly showed that in clade A1, HLA-B*5703 but not HLA-B*5801 was associated with lower HIV viral loads (30). Therefore, HLA-B*5801 was associated with neither protection nor selection in clade A1 TW10. Our data also show that HLA-B*5702- and -B*5703-mediated KF11 selection differs, despite these alleles differing at only one codon. Similar findings have been published for HLA-B7 supertype alleles (31). These data highlight the differences in immunological pressure within HLA supertype alleles, even though these alleles often present the same epitopes to the immune system.Previous reports have suggested that HIV evolution can differ among clades for a variety of reasons. HLA-B*1503 differed in its protectiveness in clade B- and clade C-infected cohorts, and the apparent mechanism is broader recognition of subdominant epitopes, which remain intact due to limited selection where HLA-B*1503 is less common (17). Similarly, Yu et al. showed that differences in KF11 evolution between clades B and C were largely the result of differences in immunological features of HLA-B*5701- and -B*5703-restricted responses, with the latter allele being more frequent in clade C-infected populations (47). The temporality of selection can also differ between clades; while TW10 and IW9 selection is similar between clade B and C, the order in which they are selected is opposite (12). Our data show that virological factors (i.e., sequence differences) can also lead to clade-specific escape. Other reports have found few differences in evolution among clades, including no differences in HLA-A2 Gag SL9 escape among clades A, B, and D (25), so the presence of clade-specific evolution will depend on the epitope and allele under study.Recent reports have suggested that Gag-specific CD8+ T-cell responses are protective in HIV infection (28), possibly because escape in Gag comes at a fitness cost. In support of this, infection by strains containing multiple Gag escape mutations was associated with lower set point viremia independent of HLA alleles in the recipients (21). One of the first demonstrations of Gag escape with fitness cost was T242N selection and reversion (32), and this substitution dominates in clade B- and clade C-infected HLA-B*57+ subjects in numerous cohorts (5, 9, 10, 15, 32, 35, 38, 41). Our data show that while clade D follows clades B and C, HLA-B*57-mediated evolution in clade A1 differs not only in TW10 but also in other p24 epitopes. Knowledge of clade-specific escape pathways will be important for vaccines that aim to cover multiple clades, particularly where clades differ in immunologically critical epitopes.  相似文献   

15.
16.
17.
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   

18.
19.
Melioidosis has been considered an emerging disease in Brazil since the first cases were reported to occur in the northeast region. This study investigated two municipalities in Ceará state where melioidosis cases have been confirmed to occur. Burkholderia pseudomallei was isolated in 26 (4.3%) of 600 samples in the dry and rainy seasons.Melioidosis is an endemic disease in Southeast Asia and northern Australia (2, 4) and also occurs sporadically in other parts of the world (3, 7). Human melioidosis was reported to occur in Brazil only in 2003, when a family outbreak afflicted four sisters in the rural part of the municipality of Tejuçuoca, Ceará state (14). After this episode, there was one reported case of melioidosis in 2004 in the rural area of Banabuiú, Ceará (14). And in 2005, a case of melioidosis associated with near drowning after a car accident was confirmed to occur in Aracoiaba, Ceará (11).The goal of this study was to investigate the Tejuçuoca and Banabuiú municipalities, where human cases of melioidosis have been confirmed to occur, and to gain a better understanding of the ecology of Burkholderia pseudomallei in this region.We chose as central points of the study the residences and surrounding areas of the melioidosis patients in the rural areas of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50′W) (Fig. (Fig.1).1). There are two well-defined seasons in each of these locations: one rainy (running from January to May) and one dry (from June to December). A total of 600 samples were collected at five sites in Tejuçuoca (T1, T2, T3, T4, and T5) and five in Banabuiú (B1, B2, B3, B4, and B5), distributed as follows (Fig. (Fig.2):2): backyards (B1 and T1), places shaded by trees (B2 and T2), water courses (B3 and T3), wet places (B4 and T4), and stock breeding areas (B5 and T5).Open in a separate windowFIG. 1.Municipalities of Banabuiú (5°18′35″S, 38°55′14″W) and Tejuçuoca (03°59′20″S, 39°34′50″W).Open in a separate windowFIG. 2.Soil sampling sites in Banabuiú and Tejuçuoca.Once a month for 12 months (a complete dry/rainy cycle), five samples were gathered at five different depths: at the surface and at 10, 20, 30 and 40 cm (Table (Table1).1). The samples were gathered according to the method used by Inglis et al. (9). Additionally, the sample processing and B. pseudomallei identification were carried out as previously reported (1, 8, 9).

TABLE 1.

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

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

TABLE 1.

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

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