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 31 毫秒
1.
2.
The survival times of Caenorhabditis elegans worms infected with Legionella pneumophila from day 7.5 or later after hatching were shorter than those of uninfected worms. However, nematodes fed bifidobacteria prior to Legionella infection were resistant to Legionella. These nematodes may act as a unique alternative host for Legionella research.Legionella pneumophila, an environmental bacterium naturally found in fresh water, is the major causative agent of Legionnaires'' disease (7). Fresh water amoebas, a natural host of Legionella, have been used as an infection model to study invasion of Legionella into human macrophages and subsequent intracellular growth (15). However, analyses using these protozoa have inevitably concentrated on the intracellular lifestyle of L. pneumophila. The fate of Legionella organisms in nonmammalian metazoans had not been described (10) until a very recent report by Brassinga et al. (6).Numerous authors have reported Caenorhabditis elegans to be a suitable model to investigate virulence-associated factors of human pathogens (2, 8, 11, 14, 16, 20, 23, 24, 30, 31, 33). In the present study, we examined whether C. elegans can serve as an alternative host for L. pneumophila. Although the nematocidal activity of Legionella has been described recently, the nematodes in the previous study were infected with the pathogen on buffered charcoal yeast extract (BCYE) agar plates, which can support Legionella growth (6). In contrast, our experiments were independently performed on simple agar plates to exclude the possibility that the inoculated pathogen would have proliferated regardless of whether it had successfully infected the nematodes and derived nutrition from the hosts. Garsin et al. showed that nutrition available in agar plates does influence the virulence of pathogens on the medium (9). Furthermore, some pathogens produce toxic metabolites on nutrient medium in situ (3), and thus, we also avoided this possibility. Moreover, we focused on the effects of worm age, since Legionella is prone to infect elderly people.Age at infection is likely one of the most important determinants of disease morbidity and mortality (18). Since Legionella organisms are prone to infect elderly people opportunistically, infections in young and older nematodes were compared. Furthermore, survival curves were compared between worms fed Escherichia coli OP50 (OP), an international standard food for these organisms, and those fed bifidobacteria prior to infection with Legionella organisms, since lactic acid bacteria exert beneficial effects on human and animal health (21).  相似文献   

3.
Legionella pneumophila exhibits surface translocation when it is grown on a buffered charcoal yeast extract (BCYE) containing 0.5 to 1.0% agar. After 7 to 22 days of incubation, spreading legionellae appear in an amorphous, lobed pattern that is most manifest at 25 to 30°C. All nine L. pneumophila strains examined displayed the phenotype. Surface translocation was also exhibited by some, but not all, other Legionella species examined. L. pneumophila mutants that were lacking flagella and/or type IV pili behaved as the wild type did when plated on low-percentage agar, indicating that the surface translocation is not swarming or twitching motility. A translucent film was visible atop the BCYE agar, advancing ahead of the spreading legionellae. Based on its abilities to disperse water droplets and to promote the spreading of heterologous bacteria, the film appeared to manipulate surface tension and, as such, acted like a surfactant. Indeed, a sample obtained from the film rapidly dispersed when it was spotted onto a plastic surface. L. pneumophila type II secretion (Lsp) mutants, but not their complemented derivatives, were defective for both surface translocation and film production. In contrast, mutants defective for type IV secretion exhibited normal surface translocation. When lsp mutants were spotted onto film produced by the wild type, they were able to spread, suggesting that type II secretion promotes the elaboration of the Legionella surfactant. Together, these data indicate that L. pneumophila exhibits a form of surface translocation that is most akin to “sliding motility” and uniquely dependent upon type II secretion.The genus Legionella was established in 1977, following the isolation of Legionella pneumophila from patients with a form of pneumonia now known as Legionnaires'' disease (33). Today, L. pneumophila is recognized as a common cause of both community-acquired and nosocomial pneumonia (84). Legionellosis occurs sporadically and in large outbreaks, with the largest outbreak occurring as recently as 2003 and encompassing 800 suspected and 449 confirmed cases (43). L. pneumophila is especially pathogenic for the elderly and the immunocompromised, large and growing segments of the population (39, 84), and recent studies have been highlighting the growing significance of travel-associated Legionnaires'' disease (107). L. pneumophila is a gram-negative, gammaproteobacterium that is widespread in natural and manufactured water systems (22, 39, 93). Infection occurs after the inhalation of Legionella-contaminated water droplets originating from a wide variety of aerosol-generating devices (39). Alarmingly, outbreaks can occur following the airborne spread of L. pneumophila over distances of >10 km from cooling towers or scrubbers (86). Within its aquatic habitats, L. pneumophila survives over a wide temperature range and grows on surfaces, in biofilms, and as an intracellular parasite of protozoa (9, 39, 110). Within the mammalian lung, the organism has the ability to attach to and invade macrophages and epithelia (27, 106, 113). Among the processes that promote L. pneumophila growth in both the environment and the mammalian lung are Lsp type II protein secretion, Dot/Icm type IVB protein secretion, and Lvh type IVA protein secretion (5, 25, 31, 106). Other key surface features of L. pneumophila are polar flagella that promote swimming motility and type IV pili that help mediate adherence (53, 103, 113). In addition to exporting proteins onto its surface into the extracellular milieu, and/or into host cells, L. pneumophila also secretes a siderophore and pyomelanin pigment that help mediate iron assimilation (23). We now report that L. pneumophila has the ability to translocate or spread across an agar surface. This new form of Legionella “motility” did not require the action of flagella, pili, or type IV secretion but was associated with the export of a surfactantlike material and an intact type II secretion system.  相似文献   

4.
Legionella pneumophila proliferates in aquatic habitats within free-living protozoa, 17 species of which have been identified as hosts by using in vitro experiments. The present study aimed at identifying protozoan hosts for L. pneumophila by using a biofilm batch test (BBT). Samples (600 ml) collected from 21 engineered freshwater systems, with added polyethylene cylinders to promote biofilm formation, were inoculated with L. pneumophila and subsequently incubated at 37°C for 20 days. Growth of L. pneumophila was observed in 16 of 18 water types when the host protozoan Hartmannella vermiformis was added. Twelve of the tested water types supported growth of L. pneumophila or indigenous Legionella anisa without added H. vermiformis. In 12 of 19 BBT flasks H. vermiformis was indicated as a host, based on the ratio between maximum concentrations of L. pneumophila and H. vermiformis, determined with quantitative PCR (Q-PCR), and the composition of clone libraries of partial 18S rRNA gene fragments. Analyses of 609 eukaryotic clones from the BBTs revealed that 68 operational taxonomic units (OTUs) showed the highest similarity to free-living protozoa. Forty percent of the sequences clustering with protozoa showed ≥99.5% similarity to H. vermiformis. None of the other protozoa serving as hosts in in vitro studies were detected in the BBTs. In several tests with growth of L. pneumophila, the protozoa Diphylleia rotans, Echinamoeba thermarum, and Neoparamoeba sp. were identified as candidate hosts. In vitro studies are needed to confirm their role as hosts for L. pneumophila. Unidentified protozoa were implicated as hosts for uncultured Legionella spp. grown in BBT flasks at 15°C.Legionella pneumophila, the causative agent of Legionnaires'' disease, is a common inhabitant of natural freshwater environments and human-made water systems, including cooling towers, whirlpools, air-conditioning systems, and installations for warm tap water (14). In the aquatic environment L. pneumophila proliferates within certain free-living protozoa, which serve as its hosts (15, 30, 59). Environmental factors favoring the growth and survival of L. pneumophila in freshwater systems include a water temperature between 20°C and 45°C (41, 60) and the presence of biofilms and sediments on which the protozoan hosts can graze (30, 41, 56).Rowbotham (44) was the first to report the growth of L. pneumophila within free-living amoebae, which belonged to the genera Acanthamoeba and Naegleria. In vitro studies with cocultures have revealed that 14 species of amoebae, viz., Acanthamoeba spp. (1, 35, 44, 53), Balamuthia mandrillaris (47), Echinamoeba exundans (15), Hartmannella spp. (43), Naegleria spp. (38, 44, 53), and Vahlkampfia jugosa (43); the slime mold Dictyostelium discoideum (20, 48); and two species of the ciliate genus Tetrahymena (15, 26) can serve as hosts for L. pneumophila. Recently, it has been reported that L. pneumophila can also replicate within the intestinal tract of the microbiovorous nematode Caenorhabditis elegans (3).A number of the free-living protozoa mentioned above and others, e.g., Vannella spp. and Saccamoeba spp., have been observed in aquatic environments from which L. pneumophila was cultivated or in which it was detected with PCR (4, 42, 51, 52). However, it remains unknown which of these protozoa actually serve as hosts for L. pneumophila in the aquatic environment, including human-made water systems. Moreover, it cannot be excluded that free-living protozoa other than those tested in vitro can serve as hosts for L. pneumophila as well. Information is also lacking about protozoan hosts for Legionella anisa (13, 49), which is frequently present in water installations in temperate regions (11, 62). Furthermore, it is unknown which free-living protozoa serve as hosts for uncultured Legionella bacteria that can grow at temperatures of about 15°C (61; B. A. Wullings, G. Bakker, and D. van der Kooij, submitted for publication).L. pneumophila can proliferate in samples of surface water, effluent of wastewater treatment plants, potable water, and water from cooling towers incubated at 25°C, 35°C, or 37°C (28, 45, 56). Consequently, incubation of freshwater samples can be used to amplify protozoan hosts for L. pneumophila and other Legionella spp. In this study, different human-made water types were investigated using a biofilm batch test (BBT) system to (i) amplify and subsequently identify predominating, known, and yet-undescribed hosts for L. pneumophila and (ii) identify potential protozoan hosts for Legionella bacteria that can grow at 15°C.  相似文献   

5.
Legionella pneumophila is a ubiquitous inhabitant of environmental water reservoirs. The bacteria infect a wide variety of protozoa and, after accidental inhalation, human alveolar macrophages, which can lead to severe pneumonia. The capability to thrive in phagocytic hosts is dependent on the Dot/Icm type IV secretion system (T4SS), which translocates multiple effector proteins into the host cell. In this study, we determined the draft genome sequence of L. pneumophila strain 130b (Wadsworth). We found that the 130b genome encodes a unique set of T4SSs, namely, the Dot/Icm T4SS, a Trb-1-like T4SS, and two Lvh T4SS gene clusters. Sequence analysis substantiated that a core set of 107 Dot/Icm T4SS effectors was conserved among the sequenced L. pneumophila strains Philadelphia-1, Lens, Paris, Corby, Alcoy, and 130b. We also identified new effector candidates and validated the translocation of 10 novel Dot/Icm T4SS effectors that are not present in L. pneumophila strain Philadelphia-1. We examined the prevalence of the new effector genes among 87 environmental and clinical L. pneumophila isolates. Five of the new effectors were identified in 34 to 62% of the isolates, while less than 15% of the strains tested positive for the other five genes. Collectively, our data show that the core set of conserved Dot/Icm T4SS effector proteins is supplemented by a variable repertoire of accessory effectors that may partly account for differences in the virulences and prevalences of particular L. pneumophila strains.Many bacterial pathogens use specialized protein secretion systems to deliver into host cells virulence effector proteins that interfere with the antimicrobial responses of the host and facilitate the survival of the pathogen (5, 10, 22, 76). The components of these secretion systems are highly conserved. Comparative bioinformatic analysis of pathogen genomes revealed an ever-increasing number of proteins that are likely to be translocated virulence effectors. Only a few effectors have been characterized, and their biochemical functions are unknown, yet the identification of translocated effector proteins and their mechanism of action is fundamental to understanding the pathogenesis of many bacterial infections.Legionella pneumophila is the etiological agent of Legionnaires’ disease, which is an acute form of pneumonia (34, 66). L. pneumophila serogroup 1 accounts for more than 90% of all cases worldwide. Although L. pneumophila is an environmental organism, its ability to survive and replicate in amoebae, such as Acanthamoeba castellanii, has equipped the organism with the capacity to replicate in human cells (45, 58, 68, 80). Following the inhalation of aerosols containing L. pneumophila into the human lung, the bacteria promote their uptake by alveolar macrophages and epithelial cells (21, 44, 71), where they replicate within an intracellular vacuole that avoids fusion with the endocytic pathway (46, 47). L. pneumophila evades endosome fusion by establishing a replicative vacuole that shares many characteristics with the endoplasmic reticulum (ER) (48, 53, 89). The formation of the unique Legionella-containing vacuole (LCV) requires the Dot (defective in organelle trafficking)/Icm (intracellular multiplication) type IV secretion system (T4SS) (85, 91).Type IV secretion systems are versatile multiprotein complexes that can transport DNA and proteins to recipient bacteria or host cells (19, 36). Based on structural and organizational similarity, three main T4SS classes have been distinguished: T4SSA, T4SSB, and genomic island-associated T4SS (GI-T4SS) (3, 51). The genetic organization and components of T4SSA have high similarity to the classical VirB4/VirD4 transfer DNA (T-DNA) transfer system of Agrobacterium tumefaciens (3). In the sequenced L. pneumophila strains, three distinct T4SSAs with different prevalences among strains have been described: Lvh, Trb-1, and Trb-2 (37, 83, 86). The Lvh (Legionella vir homologues) T4SSA is not required for intracellular bacterial replication in macrophages and amoebae but seems to contribute to infection at lower temperatures and inclusion in Acanthamoeba castellanii cysts (6, 78, 86).The Dot/Icm T4SSB secretes and translocates multiple bacterial effector proteins into the vacuolar membrane and cytosol of the host cell (31, 70). The functions of the great majority of these proteins are unknown. Several effectors have similarity to eukaryotic proteins or carry eukaryotic motifs (7, 16, 25). They are predicted to allow L. pneumophila to manipulate host cell processes by functional mimicry (31, 70). Many of the effectors have paralogues or belong to related protein families that are likely to have overlapping functions.Comparative analysis of the recent L. pneumophila genome sequences has revealed their diversity and plasticity (16, 18, 88). This plasticity enables the bacterium to acquire new genetic factors, including new effector proteins that enhance bacterial replication and survival in eukaryotic cells. This has resulted in a diverse species in which 7 to 11% of the genes in each L. pneumophila isolate are strain specific (38). Some of the diversity occurs among genes encoding Dot/Icm effectors, including those within the same family. For example some ankyrin repeat and F-box effector genes are highly conserved, while others vary considerably between L. pneumophila isolates (16, 41, 62, 73, 75). Even though it is not experimentally proven, the subsequent selection of Dot/Icm effectors among different L. pneumophila isolates might reflect their usefulness in host-pathogen interactions, whereby different effector repertoires are maintained during adaptation to different environmental niches or hosts. This may then translate into differences in virulence during opportunistic infection.In this study, we sequenced the genome of L. pneumophila serogroup 1 strain 130b (ATCC BAA-74, also known as Wadsworth or AA100) (29, 30) and analyzed the sequence for T4SSs and novel Dot/Icm effectors. This analysis established that the strain encodes a unique combination of T4SSs and a set of Dot/Icm effectors that had not been described previously but that are present in a range of clinical and environmental L. pneumophila isolates. The new effectors represent the latest members of an ever-growing list of T4SS substrates and presumably reflect the great capacity of L. pneumophila for adaptation to a variety of hosts.  相似文献   

6.
A total of 32 strains of Legionella pneumophila were used to optimize pulsed-field gel electrophoresis (PFGE) for subtyping of L. pneumophila. Twenty-six isolates of L. pneumophila with various origins and 11 isolates from five different water systems were used as the panels. For optimization of electrophoretic parameters (EPs) of SfiI PFGE, 26 isolates were analyzed with SfiI digestion, using four EPs yielding the same D value. The EP of a switch time of 5 to 50 s for 21 h had the smallest similarity coefficients and was declared the optimal EP for SfiI PFGE of L. pneumophila. By software analysis and pilot study, AscI was chosen as another PFGE enzyme. AscI PFGE could cluster the isolates from each water system into the same or very similar patterns and had a high degree of typing concordance with other molecular methods. In evaluating the discriminatory power of AscI with the panel of 26 isolates, AscI PFGE gave one single pattern and a D value of 100%. AscI PFGE had a high discriminatory power and a high degree of consistency with epidemiological data and other molecular typing methods for L. pneumophila subtyping, and hence, AscI could be used as a restriction enzyme in PFGE subtyping of L. pneumophila.Legionella pneumophila is an environmental organism that can cause disease in humans and is increasingly recognized as an important pathogen causing nosocomial pneumonia. Potable water systems (14, 26), spa water (28), and cooling towers (7, 13) are among the sources implicated in outbreaks of Legionnaires’ disease. Transmission of bacteria from the environment to humans occurs via inhalation or aspiration of Legionella-containing aerosols (3, 5). Strain differentiation is necessary for the identification of sources of contamination and determination of routes of transmission; this could in turn enable us to more accurately detect outbreaks and limit the spread of L. pneumophila infections. A variety of subtyping techniques have been used to identify and characterize L. pneumophila strains, including monoclonal antibody (MAb) analysis (16, 19), ribotyping (4), amplified fragment length polymorphism (AFLP) analysis (9, 22), PCR-based methods (15, 24), sequence-based typing (SBT) (9, 16), and pulsed-field gel electrophoresis (PFGE) (1, 6).Preliminary reports demonstrated that PFGE is a highly discriminative epidemiological marker for subtyping of L. pneumophila (6, 11, 23, 25), and a number of L. pneumophila PFGE protocols have been described in the literature (1, 2, 4, 14); however, most laboratories that use PFGE to subtype L. pneumophila cannot compare their results because the protocols differ from each other in critical parameters, such as the restriction enzymes and electrophoresis conditions used to generate the DNA fingerprints. To enhance our ability to monitor this pathogen, there is an urgent need for a standardized L. pneumophila PFGE protocol which can readily be implemented in different laboratories for information interpretation.An optimal PFGE protocol produces a suitable number of restriction fragments and gives distinct patterns by agarose gel electrophoresis, with these determined by the restriction enzymes and the electrophoretic parameters (EPs) used. SfiI is the most frequently used enzyme in conventional PFGE protocols for L. pneumophila, and there are several different EPs for SfiI digestion used by investigators for characterization and epidemiological studies. For a certain restriction enzyme, selection of the EP with the smallest similarity coefficients will increase the discriminatory power of PFGE. As the first phase of this study, we compared the similarity coefficients obtained for four EPs with SfiI digestion and determined the one with the maximal discriminatory power.There were some problems found in practical applications of epidemiological investigation of L. pneumophila by PFGE with single SfiI digestion, such as having epidemiologically unrelated strains exhibit the same patterns (30) and the appearance of “ghost” or “phantom” bands. Combination use of two enzymes would give a higher discriminatory power and more accurate results (10, 29). Thus, as the second phase of this study, we selected another suitable enzyme and compared it with SfiI to evaluate the possibility of its use in characterization and epidemiological studies of L. pneumophila.  相似文献   

7.
8.
A total of 560 Legionella species were isolated from environmental water sources from public facilities from June to September 2008 throughout South Korea. The distribution of Legionella isolates was investigated according to geographical region, facility type, and sample type. The genetic diversity of 104 isolates of Legionella pneumophila serogroup 1 (sg 1) was analyzed by sequence-based typing (SBT). L. pneumophila was distributed broadly throughout Korea, accounting for 85.0% of the isolates, and L. pneumophila sg 1 predominated in all of the public facilities except for the springs. Legionella anisa and Legionella bozemanii predominated among non-L. pneumophila species (48.1% and 21.0%, respectively). The second most dominant strain differed depending on the facility type: L. anisa was the second most dominant strain in the buildings (10.8%), L. pneumophila sg 5 in public baths (21.6%), L. pneumophila sg 6 in factories (12.0%), and L. pneumophila sg 7 in hospitals (13.1%). In the SBT analysis, 104 L. pneumophila sg 1 isolates were differentiated into 26 sequence types (STs) and categorized into 3 clonal groups (CGs) and 10 singleton STs via the eBURST V3 program. ST1, a potential founder of major CG1, was commonly distributed (48.1%). The dominant ST in hot water was ST-K1 (7, 12, 17, 3, 35, 11, 11), which was designated in this study (36.1%). The second most dominant strain differed depending on the type of facility from which the samples were obtained. The unique allelic profile of ST-K1, obtained from hot water, was not found in the European Working Group for Legionella Infections (EWGLI) SBT database.Legionella species, ubiquitous Gram-negative bacteria, are found in a variety of artificial water systems, natural freshwaters, and soils. Currently, the Legionella genus includes 52 species and more than 70 different serogroups, and more than 20 species have been proven to be causative agents of Legionnaires'' disease (LD). The species Legionella pneumophila accounts for approximately 90% of confirmed cases of legionellosis, and L. pneumophila serogroup 1 (sg 1) has been recognized as the most important agent in this regard, as that specific strain was initially implicated as the pathogen causative of LD in 1977 (15; http://www.bacterio.cict.fr/l/legionellaceae.html). The other non-L. pneumophila sg 1 strains, sg 2 to 15, accounted for 7.4% of cases, and Legionella longbeachae (3.9%) and Legionella bozemanii (2.4%) have also been associated with the pathogen of LD. In particular, L. longbeachae has been recognized as accounting for 30.4% of community-acquired Legionella isolates in Australia and New Zealand (53).The most common transmission mechanism of legionellosis is the inhalation of aerosols from the water systems of artificial facilities, including large buildings, hotels, hospitals, public baths, spas, or decorative fountains contaminated by Legionella species (1). Therefore, hot water and water from cooling towers have been perceived as sources of infection in cases of community-acquired, nosocomially acquired, or travel-associated LD (15, 26, 31, 37, 38, 39, 41, 43). Thus, it is important from a public health perspective to continually survey environmental water systems for the presence of Legionella species (2, 34, 35). In particular, hot-water systems used as public baths, such as springs, spas, or tubs, have become a popular means of recreation in a lot of countries, including South Korea. The contamination of hot-water systems has gradually become recognized as an important risk factor all over the world (4, 12, 18, 23, 42, 50), as sources of legionellosis have been detected increasingly since 1982 (52) and many cases of nosocomially acquired (32, 51) and community-acquired (6, 7, 48) LD have been detected in Legionella-contaminated hot-water systems or hot springs.In South Korea, several cases of nosocomial infection and community-acquired pneumonia have occasionally been reported (9, 45) since the first recognized outbreak in South Korea in 1984, which was associated with Legionella gormanii (27). Since 2006, the Korean National Infectious Disease Surveillance (NIDS) program (http://dis.cdc.go.kr/) has reported an average of 20 cases of LD per year (29). In South Korea, surveys of Legionella acquired from environmental water in public facilities such as hot springs and public baths has been gradually enhanced since 2007. An annual training program for the detection of Legionella species from environmental water systems and clinical specimens is currently conducted for the personnel of 16 Provincial Institute of Health and Environment locations (PIHEs) throughout South Korea. Recently, the rate of detection of environmental Legionella bacteria has been gradually increasing (8.1% in 2006, 9.4% in 2007, and 10.3% in 2008).The principal objectives of this study were to assess the current distribution of Legionella species from environmental water sources from public facilities such as buildings, hotels, public baths, springs, hospitals, or factories throughout South Korea. Additionally, the molecular typing of L. pneumophila sg 1 isolates was conducted using sequence-based typing (SBT) to assess the genetic diversity among the isolates.  相似文献   

9.
Legionella longbeachae causes most cases of legionellosis in Australia and may be underreported worldwide due to the lack of L. longbeachae-specific diagnostic tests. L. longbeachae displays distinctive differences in intracellular trafficking, caspase 1 activation, and infection in mouse models compared to Legionella pneumophila, yet these two species have indistinguishable clinical presentations in humans. Unlike other legionellae, which inhabit freshwater systems, L. longbeachae is found predominantly in moist soil. In this study, we sequenced and annotated the genome of an L. longbeachae clinical isolate from Oregon, isolate D-4968, and compared it to the previously published genomes of L. pneumophila. The results revealed that the D-4968 genome is larger than the L. pneumophila genome and has a gene order that is different from that of the L. pneumophila genome. Genes encoding structural components of type II, type IV Lvh, and type IV Icm/Dot secretion systems are conserved. In contrast, only 42/140 homologs of genes encoding L. pneumophila Icm/Dot substrates have been found in the D-4968 genome. L. longbeachae encodes numerous proteins with eukaryotic motifs and eukaryote-like proteins unique to this species, including 16 ankyrin repeat-containing proteins and a novel U-box protein. We predict that these proteins are secreted by the L. longbeachae Icm/Dot secretion system. In contrast to the L. pneumophila genome, the L. longbeachae D-4968 genome does not contain flagellar biosynthesis genes, yet it contains a chemotaxis operon. The lack of a flagellum explains the failure of L. longbeachae to activate caspase 1 and trigger pyroptosis in murine macrophages. These unique features of L. longbeachae may reflect adaptation of this species to life in soil.Isolation of Legionella longbeachae was first reported in 1981 after isolation from patients with pneumonia in the United States (11, 59). Although L. longbeachae is not a common respiratory pathogen in either North America or Europe, where Legionella pneumophila infections are predominant, it accounts for more than 50% of legionellosis cases in Australia and is also prevalent in New Zealand and Thailand (10, 12, 60, 66, 68, 77, 93, 94). Legionnaires'' disease induced by L. longbeachae infection is clinically indistinguishable from the disease caused by L. pneumophila (65). However, L. longbeachae infections have been associated with gardening and the use of potting soil, whereas the disease caused by other species is linked to freshwater sources (4, 65). L. longbeachae can survive for up to 9 months in moist potting soil at room temperature, in contrast to other Legionella species, which inhabit natural and manmade freshwater systems worldwide (34, 83, 84).In addition to the differences in habitat, L. longbeachae differs from L. pneumophila in its virulence in murine models of infection. L. longbeachae replicates in the lungs of A/J, C57BL/6, and BALB/c mice (6), whereas most inbred mice, including C57BL/6 and BALB strains, are resistant to L. pneumophila (61). These differences in murine host susceptibility are likely due to different abilities to activate caspase 1-mediated pyroptosis in macrophages. While L. pneumophila rapidly triggers pyroptosis in C57BL/6 mouse macrophages, L. longbeachae does not do this (6).Intracellular trafficking of L. longbeachae in mammalian macrophages also follows a route distinct from that of L. pneumophila. After phagocytosis, the L. pneumophila-containing vacuole (LCV) excludes early and late endosomal markers, such as early endosomal antigen 1 (EEA1), Rab5, LAMP-1, LAMP-2, and the mannose 6-phosphate receptor (M6PR) (5, 89). In L. pneumophila the Dot/Icm type IV secretion system is required for prevention of phagosome-lysosome fusion and for intracellular replication (47). Conversely, the L. longbeachae-containing vacuole acquires the early endosomal marker EEA1 and the late endosomal markers LAMP-2 and M6PR (5). It has been suggested that L. longbeachae intracellular trafficking resembles that of the facultative intracellular pathogen Brucella abortus, since a Brucella-containing vacuole also acquires early and late endosomal markers soon after infection (5). Despite the difference in intracellular trafficking between L. longbeachae and L. pneumophila, L. longbeachae rescues Dot/Icm-deficient L. pneumophila when these two organisms coinhabit LCV (5).Results of the studies cited above indicate that L. longbeachae differs from other legionellae in terms of habitat, host specificity, and intracellular trafficking. In this paper, we describe an analysis of the sequenced and annotated genome of L. longbeachae clinical isolate D-4968 compared with published genomes of L. pneumophila strains Corby, Lens, Paris, and Philadelphia-1 (16, 17, 38). Specifically, we compared genes involved in gene regulation, protein secretion systems, and motility in order to identify genes responsible for making L. longbeachae unique among the legionellae.  相似文献   

10.
11.
Free-living protozoan communities in water supplies may include hosts for Legionella pneumophila and other undesired bacteria, as well as pathogens. This study aimed at identifying free-living protozoa in two unchlorinated groundwater supplies, using cultivation-independent molecular approaches. For this purpose, samples (<20°C) of treated water, distributed water, and distribution system biofilms were collected from supply A, with a low concentration of natural organic matter (NOM) (<0.5 ppm of C), and from supply B, with a high NOM concentration (7.9 ppm of C). Eukaryotic communities were studied using terminal restriction fragment length polymorphism and clone library analyses of partial 18S rRNA gene fragments and a Hartmannella vermiformis-specific quantitative PCR (qPCR). In both supplies, highly diverse eukaryotic communities were observed, including free-living protozoa, fungi, and metazoa. Sequences of protozoa clustered with Amoebozoa (10 operational taxonomic units [OTUs]), Cercozoa (39 OTUs), Choanozoa (26 OTUs), Ciliophora (29 OTUs), Euglenozoa (13 OTUs), Myzozoa (5 OTUs), and Stramenopiles (5 OTUs). A large variety of protozoa were present in both supplies, but the estimated values for protozoan richness did not differ significantly. H. vermiformis was observed in both supplies but was not a predominant protozoan. One OTU with the highest similarity to Acanthamoeba polyphaga, an opportunistic human pathogen and a host for undesired bacteria, was observed in supply A. The high level of NOM in supply B corresponded with an elevated level of active biomass and with elevated concentrations of H. vermiformis in distributed water. Hence, the application of qPCR may be promising in elucidating the relationship between drinking water quality and the presence of specific protozoa.Free-living protozoa are ubiquitous in natural freshwater environments (7, 38, 51, 71) but also proliferate in engineered water systems, including water treatment systems (3, 47, 70), distribution systems (6, 75), and tap water installations inside buildings (54, 69). Concentrations of protozoa, determined using cultivation methods and microscopy, range from <1 to 104 cells liter−1 in treated water (3, 47, 70, 75) and from <1 to 7 × 105 cells liter−1 in distribution systems (6, 61, 64, 75). Genera of free-living protozoa commonly observed in these systems and in tap water installations include Acanthamoeba, Echinamoeba, Hartmannella, Platyamoeba, Vahlkampfia, and Vannella (47, 58, 69, 70). In warm water systems, certain free-living protozoa, e.g., Acanthamoeba spp. (57), Balamuthia mandrillaris (62), Echinamoeba exandans (16), Hartmannella spp. (39, 56), Naegleria spp. (49, 57), Tetrahymena spp. (18, 33), and Vahlkampfia jugosa (56), serve as hosts for Legionella pneumophila, the etiologic agent of Legionnaires'' disease. High concentrations of L. pneumophila are generally associated with the proliferation of host protozoa in biofilms (38, 53). In addition, other amoeba-resistant, potentially pathogenic bacteria, e.g., Burkholderia spp. (28) and Mycobacterium spp. (37), have been observed in man-made aquatic environments (24). Free-living protozoa may enhance the multiplication of bacteria, serve as a transmission vector, or serve as a shelter against unfavorable environmental conditions, such as the presence of disinfectants. Furthermore, certain free-living protozoa are human pathogens, e.g., Naegleria fowleri (81), Balamuthia mandrillaris (77), and Acanthamoeba spp. (12) can cause encephalitis. Acanthamoeba spp. have also been associated with keratitis in persons wearing contact lenses (31).Free-living protozoa feed on bacteria, algae, fungi, other protozoa, and organic detritus in biofilms or in the planktonic phase, thereby affecting the structure of microbial communities. In turn, the community of free-living protozoa depends on the diversity and abundance of bacteria in the biofilm and in the planktonic phase (26, 50, 51, 55, 63, 65). Water quality is a critical factor for biofilm formation in distribution systems and tap water installations and therefore will affect the abundance and diversity of free-living protozoa in these systems (72, 78). However, information about the presence and identity of free-living protozoa in water supplies in relation to the quality of treated water is scarce, which may be attributed to the limitations of microscopic techniques and cultivation methods for detection and identification of these organisms, e.g., low detection limits and selectivity for specific groups (19).In this study, we applied a variety of cultivation-independent techniques, viz., quantitative PCR, terminal restriction fragment length polymorphism (T-RFLP) analysis, and cloning and sequencing of eukaryotic 18S rRNA gene fragments, for the detection and identification of free-living protozoa predominating in two unchlorinated groundwater supplies. The concentrations of dissolved natural organic matter (NOM) in treated water at the plant were <0.5 mg C liter−1 and 7.9 mg C liter−1, covering the entire range of NOM concentrations in drinking water in The Netherlands. The objectives of the study were (i) to elucidate the identities of and diversity in the free-living protozoa predominating in these two different water supplies and (ii) to trace the presence of host protozoa for L. pneumophila and pathogenic free-living protozoa. The study revealed that treated water and biofilms in the distribution systems of both water supplies contained a large variety of free-living protozoa, including protozoan hosts for Legionella bacteria.  相似文献   

12.
Methanogens are of great importance in carbon cycling and alternative energy production, but quantitation with culture-based methods is time-consuming and biased against methanogen groups that are difficult to cultivate in a laboratory. For these reasons, methanogens are typically studied through culture-independent molecular techniques. We developed a SYBR green I quantitative PCR (qPCR) assay to quantify total numbers of methyl coenzyme M reductase α-subunit (mcrA) genes. TaqMan probes were also designed to target nine different phylogenetic groups of methanogens in qPCR assays. Total mcrA and mcrA levels of different methanogen phylogenetic groups were determined from six samples: four samples from anaerobic digesters used to treat either primarily cow or pig manure and two aliquots from an acidic peat sample stored at 4°C or 20°C. Only members of the Methanosaetaceae, Methanosarcina, Methanobacteriaceae, and Methanocorpusculaceae and Fen cluster were detected in the environmental samples. The three samples obtained from cow manure digesters were dominated by members of the genus Methanosarcina, whereas the sample from the pig manure digester contained detectable levels of only members of the Methanobacteriaceae. The acidic peat samples were dominated by both Methanosarcina spp. and members of the Fen cluster. In two of the manure digester samples only one methanogen group was detected, but in both of the acidic peat samples and two of the manure digester samples, multiple methanogen groups were detected. The TaqMan qPCR assays were successfully able to determine the environmental abundance of different phylogenetic groups of methanogens, including several groups with few or no cultivated members.Methanogens are integral to carbon cycling, catalyzing the production of methane and carbon dioxide, both potent greenhouse gases, during organic matter degradation in anaerobic soils and sediment (8). Methanogens are widespread in anaerobic environments, including tundra (36), freshwater lake and wetland sediments (9, 12), estuarine and marine sediments (2), acidic peatlands (4, 14), rice field soil (10, 16), animal guts (41), landfills (30), and anaerobic digesters treating animal manure (1), food processing wastewater (27), and municipal wastewater and solid waste (37, 57). Methane produced in anaerobic digesters may be captured and used for energy production, thus offsetting some or all of the cost of operation and reducing the global warming potential of methane release to the atmosphere.Methanogens are difficult to study through culture-based methods, and therefore many researchers have instead used culture-independent techniques to study methanogen populations. The 16S rRNA gene is the most widely used target for gene surveys, and a number of primers and probes have been developed to target methanogen groups (9, 11, 31, 36, 38, 40, 46, 48, 57). To eliminate potential problems with nonspecific amplification, some researchers have developed primers for the gene sequence of the α-subunit of the methyl coenzyme M reductase (mcrA) (17, 30, 49). The Mcr is exclusive to the methanogens with the exception of the methane-oxidizing Archaea (18) and shows mostly congruent phylogeny to the 16S rRNA gene, allowing mcrA analysis to be used in conjunction with, or independently of, that of the 16S rRNA gene (3, 30, 49). A number of researchers have examined methanogen communities with mcrA and have found uncultured clades quite different in sequence from cultured methanogen representatives (9, 10, 12, 14, 17, 22, 28, 47).Previous studies described methanogen communities by quantitation of different clades through the use of rRNA-targeted or rRNA gene-targeted probes with techniques such as dot blot hybridization (1, 27, 37, 38, 48) and fluorescent in situ hybridization (11, 40, 44, 57). Real-time quantitative PCR (qPCR) is an alternate technique capable of determining the copy number of a particular gene present in the DNA extracted from an environmental sample. Only a few studies have used qPCR to quantitatively examine different clades within methanogen communities, and most of these studies have exclusively targeted the 16S rRNA gene (19, 41, 42, 54-56). Far fewer researchers have used qPCR to quantify methanogen clades by targeting the mcrA (21, 34, 45), and these studies were limited to only a few phylogenetic groups.In this paper we present a methodology for determining methanogen gene copy numbers through the use of qPCR targeting the mcrA. Methanogens were quantified in total using methanogen-specific primers in SYBR green assays and also as members of nine different phylogenetic groups using TaqMan probes targeting specific subsets of methanogens.  相似文献   

13.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

14.
An intracellular multiplication F (IcmF) family protein is a conserved component of a newly identified type VI secretion system (T6SS) encoded in many animal and plant-associated Proteobacteria. We have previously identified ImpLM, an IcmF family protein that is required for the secretion of the T6SS substrate hemolysin-coregulated protein (Hcp) from the plant-pathogenic bacterium Agrobacterium tumefaciens. In this study, we characterized the topology of ImpLM and the importance of its nucleotide-binding Walker A motif involved in Hcp secretion from A. tumefaciens. A combination of β-lactamase-green fluorescent protein fusion and biochemical fractionation analyses revealed that ImpLM is an integral polytopic inner membrane protein comprising three transmembrane domains bordered by an N-terminal domain facing the cytoplasm and a C-terminal domain exposed to the periplasm. impLM mutants with substitutions or deletions in the Walker A motif failed to complement the impLM deletion mutant for Hcp secretion, which provided evidence that ImpLM may bind and/or hydrolyze nucleoside triphosphates to mediate T6SS machine assembly and/or substrate secretion. Protein-protein interaction and protein stability analyses indicated that there is a physical interaction between ImpLM and another essential T6SS component, ImpKL. Topology and biochemical fractionation analyses suggested that ImpKL is an integral bitopic inner membrane protein with an N-terminal domain facing the cytoplasm and a C-terminal OmpA-like domain exposed to the periplasm. Further comprehensive yeast two-hybrid assays dissecting ImpLM-ImpKL interaction domains suggested that ImpLM interacts with ImpKL via the N-terminal cytoplasmic domains of the proteins. In conclusion, ImpLM interacts with ImpKL, and its Walker A motif is required for its function in mediation of Hcp secretion from A. tumefaciens.Many pathogenic gram-negative bacteria employ protein secretion systems formed by macromolecular complexes to deliver proteins or protein-DNA complexes across the bacterial membrane. In addition to the general secretory (Sec) pathway (18, 52) and twin-arginine translocation (Tat) pathway (7, 34), which transport proteins across the inner membrane into the periplasm, at least six distinct protein secretion systems occur in gram-negative bacteria (28, 46, 66). These systems are able to secrete proteins from the cytoplasm or periplasm to the external environment or the host cell and include the well-documented type I to type V secretion systems (T1SS to T5SS) (10, 15, 23, 26, 30) and a recently discovered type VI secretion system (T6SS) (4, 8, 22, 41, 48, 49). These systems use ATPase or a proton motive force to energize assembly of the protein secretion machinery and/or substrate translocation (2, 6, 41, 44, 60).Agrobacterium tumefaciens is a soilborne pathogenic gram-negative bacterium that causes crown gall disease in a wide range of plants. Using an archetypal T4SS (9), A. tumefaciens translocates oncogenic transferred DNA and effector proteins to the host and ultimately integrates transferred DNA into the host genome. Because of its unique interkingdom DNA transfer, this bacterium has been extensively studied and used to transform foreign DNA into plants and fungi (11, 24, 40, 67). In addition to the T4SS, A. tumefaciens encodes several other secretion systems, including the Sec pathway, the Tat pathway, T1SS, T5SS, and the recently identified T6SS (72). T6SS is highly conserved and widely distributed in animal- and plant-associated Proteobacteria and plays an important role in the virulence of several human and animal pathogens (14, 19, 41, 48, 56, 63, 74). However, T6SS seems to play only a minor role or even a negative role in infection or virulence of the plant-associated pathogens or symbionts studied to date (5, 37-39, 72).T6SS was initially designated IAHP (IcmF-associated homologous protein) clusters (13). Before T6SS was documented by Pukatzki et al. in Vibrio cholerae (48), mutations in this gene cluster in the plant symbiont Rhizobium leguminosarum (5) and the fish pathogen Edwardsiella tarda (51) caused defects in protein secretion. In V. cholerae, T6SS was responsible for the loss of cytotoxicity for amoebae and for secretion of two proteins lacking a signal peptide, hemolysin-coregulated protein (Hcp) and valine-glycine repeat protein (VgrG). Secretion of Hcp is the hallmark of T6SS. Interestingly, mutation of hcp blocks the secretion of VgrG proteins (VgrG-1, VgrG-2, and VgrG-3), and, conversely, vgrG-1 and vgrG-2 are both required for secretion of the Hcp and VgrG proteins from V. cholerae (47, 48). Similarly, a requirement of Hcp for VgrG secretion and a requirement of VgrG for Hcp secretion have also been shown for E. tarda (74). Because Hcp forms a hexameric ring (41) stacked in a tube-like structure in vitro (3, 35) and VgrG has a predicted trimeric phage tail spike-like structure similar to that of the T4 phage gp5-gp27 complex (47), Hcp and VgrG have been postulated to form an extracellular translocon. This model is further supported by two recent crystallography studies showing that Hcp, VgrG, and a T4 phage gp25-like protein resembled membrane penetration tails of bacteriophages (35, 45).Little is known about the topology and structure of T6SS machinery subunits and the distinction between genes encoding machinery subunits and genes encoding regulatory proteins. Posttranslational regulation via the phosphorylation of Fha1 by a serine-threonine kinase (PpkA) is required for Hcp secretion from Pseudomonas aeruginosa (42). Genetic evidence for P. aeruginosa suggested that the T6SS may utilize a ClpV-like AAA+ ATPase to provide the energy for machinery assembly or substrate translocation (41). A recent study of V. cholerae suggested that ClpV ATPase activity is responsible for remodeling the VipA/VipB tubules which are crucial for type VI substrate secretion (6). An outer membrane lipoprotein, SciN, is an essential T6SS component for mediating Hcp secretion from enteroaggregative Escherichia coli (1). A systematic study of the T6SS machinery in E. tarda revealed that 13 of 16 genes in the evp gene cluster are essential for secretion of T6S substrates (74), which suggests the core components of the T6SS. Interestingly, most of the core components conserved in T6SS are predicted soluble proteins without recognizable signal peptide and transmembrane (TM) domains.The intracellular multiplication F (IcmF) and H (IcmH) proteins are among the few core components with obvious TM domains (8). In Legionella pneumophila Dot/Icm T4SSb, IcmF and IcmH are both membrane localized and partially required for L. pneumophila replication in macrophages (58, 70, 75). IcmF and IcmH are thought to interact with each other in stabilizing the T4SS complex in L. pneumophila (58). In T6SS, IcmF is one of the essential components required for secretion of Hcp from several animal pathogens, including V. cholerae (48), Aeromonas hydrophila (63), E. tarda (74), and P. aeruginosa (41), as well as the plant pathogens A. tumefaciens (72) and Pectobacterium atrosepticum (39). In E. tarda, IcmF (EvpO) interacted with IcmH (EvpN), EvpL, and EvpA in a yeast two-hybrid assay, and its putative nucleotide-binding site (Walker A motif) was not essential for secretion of T6SS substrates (74).In this study, we characterized the topology and interactions of the IcmF and IcmH family proteins ImpLM and ImpKL, which are two essential components of the T6SS of A. tumefaciens. We adapted the nomenclature proposed by Cascales (8), using the annotated gene designation followed by the letter indicated by Shalom et al. (59). Our data indicate that ImpLM and ImpKL are both integral inner membrane proteins and interact with each other via their N-terminal domains residing in the cytoplasm. We also provide genetic evidence showing that ImpLM may function as a nucleoside triphosphate (NTP)-binding protein or nucleoside triphosphatase to mediate T6S machinery assembly and/or substrate secretion.  相似文献   

15.
16.
Phenoxyalkanoic acid (PAA) herbicides are widely used in agriculture. Biotic degradation of such herbicides occurs in soils and is initiated by α-ketoglutarate- and Fe2+-dependent dioxygenases encoded by tfdA-like genes (i.e., tfdA and tfdAα). Novel primers and quantitative kinetic PCR (qPCR) assays were developed to analyze the diversity and abundance of tfdA-like genes in soil. Five primer sets targeting tfdA-like genes were designed and evaluated. Primer sets 3 to 5 specifically amplified tfdA-like genes from soil, and a total of 437 sequences were retrieved. Coverages of gene libraries were 62 to 100%, up to 122 genotypes were detected, and up to 389 genotypes were predicted to occur in the gene libraries as indicated by the richness estimator Chao1. Phylogenetic analysis of in silico-translated tfdA-like genes indicated that soil tfdA-like genes were related to those of group 2 and 3 Bradyrhizobium spp., Sphingomonas spp., and uncultured soil bacteria. Soil-derived tfdA-like genes were assigned to 11 clusters, 4 of which were composed of novel sequences from this study, indicating that soil harbors novel and diverse tfdA-like genes. Correlation analysis of 16S rRNA and tfdA-like gene similarity indicated that any two bacteria with D > 20% of group 2 tfdA-like gene-derived protein sequences belong to different species. Thus, data indicate that the soil analyzed harbors at least 48 novel bacterial species containing group 2 tfdA-like genes. Novel qPCR assays were established to quantify such new tfdA-like genes. Copy numbers of tfdA-like genes were 1.0 × 106 to 65 × 106 per gram (dry weight) soil in four different soils, indicating that hitherto-unknown, diverse tfdA-like genes are abundant in soils.Phenoxyalkanoic acid (PAA) herbicides such as MCPA (4-chloro-2-methyl-phenoxyacetic acid) and 2,4-D (2,4-dichlorophenoxyacetic acid) are widely used to control broad-leaf weeds in agricultural as well as nonagricultural areas (19, 77). Degradation occurs primarily under oxic conditions in soil, and microorganisms play a key role in the degradation of such herbicides in soil (62, 64). Although relatively rapidly degraded in soil (32, 45), both MCPA and 2,4-D are potential groundwater contaminants (10, 56, 70), accentuating the importance of bacterial PAA herbicide-degrading bacteria in soils (e.g., references 3, 5, 6, 20, 41, 59, and 78).Degradation can occur cometabolically or be associated with energy conservation (15, 54). The first step in the degradation of 2,4-D and MCPA is initiated by the product of cadAB or tfdA-like genes (29, 30, 35, 67), which constitutes an α-ketoglutarate (α-KG)- and Fe2+-dependent dioxygenase. TfdA removes the acetate side chain of 2,4-D and MCPA to produce 2,4-dichlorophenol and 4-chloro-2-methylphenol, respectively, and glyoxylate while oxidizing α-ketoglutarate to CO2 and succinate (16, 17).Organisms capable of PAA herbicide degradation are phylogenetically diverse and belong to the Alpha-, Beta-, and Gammproteobacteria and the Bacteroidetes/Chlorobi group (e.g., references 2, 14, 29-34, 39, 60, 68, and 71). These bacteria harbor tfdA-like genes (i.e., tfdA or tfdAα) and are categorized into three groups on an evolutionary and physiological basis (34). The first group consists of beta- and gammaproteobacteria and can be further divided into three distinct classes based on their tfdA genes (30, 46). Class I tfdA genes are closely related to those of Cupriavidus necator JMP134 (formerly Ralstonia eutropha). Class II tfdA genes consist of those of Burkholderia sp. strain RASC and a few strains that are 76% identical to class I tfdA genes. Class III tfdA genes are 77% identical to class I and 80% identical to class II tfdA genes and linked to MCPA degradation in soil (3). The second group consists of alphaproteobacteria, which are closely related to Bradyrhizobium spp. with tfdAα genes having 60% identity to tfdA of group 1 (18, 29, 34). The third group also harbors the tfdAα genes and consists of Sphingomonas spp. within the alphaproteobacteria (30).Diverse PAA herbicide degraders of all three groups were identified in soil by cultivation-dependent studies (32, 34, 41, 78). Besides CadAB, TfdA and certain TfdAα proteins catalyze the conversion of PAA herbicides (29, 30, 35). All groups of tfdA-like genes are potentially linked to the degradation of PAA herbicides, although alternative primary functions of group 2 and 3 TfdAs have been proposed (30, 35). However, recent cultivation-independent studies focused on 16S rRNA genes or solely on group 1 tfdA sequences in soil (e.g., references 3-5, 13, and 41). Whether group 2 and 3 tfdA-like genes are also quantitatively linked to the degradation of PAA herbicides in soils is unknown. Thus, tools to target a broad range of tfdA-like genes are needed to resolve such an issue. Primers used to assess the diversity of tfdA-like sequences used in previous studies were based on the alignment of approximately 50% or less of available sequences to date (3, 20, 29, 32, 39, 47, 58, 73). Primers specifically targeting all major groups of tfdA-like genes to assess and quantify a broad diversity of potential PAA degraders in soil are unavailable. Thus, the objectives of this study were (i) to develop primers specific for all three groups of tfdA-like genes, (ii) to establish quantitative kinetic PCR (qPCR) assays based on such primers for different soil samples, and (iii) to assess the diversity and abundance of tfdA-like genes in soil.  相似文献   

17.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

18.
19.
20.
Adhesive pili on the surface of the serotype M1 Streptococcus pyogenes strain SF370 are composed of a major backbone subunit (Spy0128) and two minor subunits (Spy0125 and Spy0130), joined covalently by a pilin polymerase (Spy0129). Previous studies using recombinant proteins showed that both minor subunits bind to human pharyngeal (Detroit) cells (A. G. Manetti et al., Mol. Microbiol. 64:968-983, 2007), suggesting both may act as pilus-presented adhesins. While confirming these binding properties, studies described here indicate that Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role as a wall linker. Pili were localized predominantly to cell wall fractions of the wild-type S. pyogenes parent strain and a spy0125 deletion mutant. In contrast, they were found almost exclusively in culture supernatants in both spy0130 and srtA deletion mutants, indicating that the housekeeping sortase (SrtA) attaches pili to the cell wall by using Spy0130 as a linker protein. Adhesion assays with antisera specific for individual subunits showed that only anti-rSpy0125 serum inhibited adhesion of wild-type S. pyogenes to human keratinocytes and tonsil epithelium to a significant extent. Spy0125 was localized to the tip of pili, based on a combination of mutant analysis and liquid chromatography-tandem mass spectrometry analysis of purified pili. Assays comparing parent and mutant strains confirmed its role as the adhesin. Unexpectedly, apparent spontaneous cleavage of a labile, proline-rich (8 of 14 residues) sequence separating the N-terminal ∼1/3 and C-terminal ∼2/3 of Spy0125 leads to loss of the N-terminal region, but analysis of internal spy0125 deletion mutants confirmed that this has no significant effect on adhesion.The group A Streptococcus (S. pyogenes) is an exclusively human pathogen that commonly colonizes either the pharynx or skin, where local spread can give rise to various inflammatory conditions such as pharyngitis, tonsillitis, sinusitis, or erysipelas. Although often mild and self-limiting, GAS infections are occasionally very severe and sometimes lead to life-threatening diseases, such as necrotizing fasciitis or streptococcal toxic shock syndrome. A wide variety of cell surface components and extracellular products have been shown or suggested to play important roles in S. pyogenes virulence, including cell surface pili (1, 6, 32). Pili expressed by the serotype M1 S. pyogenes strain SF370 mediate specific adhesion to intact human tonsil epithelia and to primary human keratinocytes, as well as cultured keratinocyte-derived HaCaT cells, but not to Hep-2 or A549 cells (1). They also contribute to adhesion to a human pharyngeal cell line (Detroit cells) and to biofilm formation (29).Over the past 5 years, pili have been discovered on an increasing number of important Gram-positive bacterial pathogens, including Bacillus cereus (4), Bacillus anthracis (4, 5), Corynebacterium diphtheriae (13, 14, 19, 26, 27, 44, 46, 47), Streptococcus agalactiae (7, 23, 38), and Streptococcus pneumoniae (2, 3, 24, 25, 34), as well as S. pyogenes (1, 29, 32). All these species produce pili that are composed of a single major subunit plus either one or two minor subunits. During assembly, the individual subunits are covalently linked to each other via intermolecular isopeptide bonds, catalyzed by specialized membrane-associated transpeptidases that may be described as pilin polymerases (4, 7, 25, 41, 44, 46). These are related to the classical housekeeping sortase (usually, but not always, designated SrtA) that is responsible for anchoring many proteins to Gram-positive bacterial cell walls (30, 31, 33). The C-terminal ends of sortase target proteins include a cell wall sorting (CWS) motif consisting, in most cases, of Leu-Pro-X-Thr-Gly (LPXTG, where X can be any amino acid) (11, 40). Sortases cleave this substrate between the Thr and Gly residues and produce an intermolecular isopeptide bond linking the Thr to a free amino group provided by a specific target. In attaching proteins to the cell wall, the target amino group is provided by the lipid II peptidoglycan precursor (30, 36, 40). In joining pilus subunits, the target is the ɛ-amino group in the side chain of a specific Lys residue in the second subunit (14, 18, 19). Current models of pilus biogenesis envisage repeated transpeptidation reactions adding additional subunits to the base of the growing pilus, until the terminal subunit is eventually linked covalently via an intermolecular isopeptide bond to the cell wall (28, 41, 45).The major subunit (sometimes called the backbone or shaft subunit) extends along the length of the pilus and appears to play a structural role, while minor subunits have been detected either at the tip, the base, and/or at occasional intervals along the shaft, depending on the species (4, 23, 24, 32, 47). In S. pneumoniae and S. agalactiae one of the minor subunits acts as an adhesin, while the second appears to act as a linker between the base of the assembled pilus and the cell wall (7, 15, 22, 34, 35). It was originally suggested that both minor subunits of C. diphtheriae pili could act as adhesins (27). However, recent data showed one of these has a wall linker role (26, 44) and may therefore not function as an adhesin.S. pyogenes strain SF370 pili are composed of a major (backbone) subunit, termed Spy0128, plus two minor subunits, called Spy0125 and Spy0130 (1, 32). All three are required for efficient adhesion to target cells (1). Studies employing purified recombinant proteins have shown that both of the minor subunits, but not the major subunit, bind to Detroit cells (29), suggesting both might act as pilus-presented adhesins. Here we report studies employing a combination of recombinant proteins, specific antisera, and allelic replacement mutants which show that only Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role in linking pili to the cell wall.  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