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 46 毫秒
1.
Clostridium difficile, a major cause of antibiotic-associated diarrhea, produces highly resistant spores that contaminate hospital environments and facilitate efficient disease transmission. We purified C. difficile spores using a novel method and show that they exhibit significant resistance to harsh physical or chemical treatments and are also highly infectious, with <7 environmental spores per cm2 reproducibly establishing a persistent infection in exposed mice. Mass spectrometric analysis identified ∼336 spore-associated polypeptides, with a significant proportion linked to translation, sporulation/germination, and protein stabilization/degradation. In addition, proteins from several distinct metabolic pathways associated with energy production were identified. Comparison of the C. difficile spore proteome to those of other clostridial species defined 88 proteins as the clostridial spore “core” and 29 proteins as C. difficile spore specific, including proteins that could contribute to spore-host interactions. Thus, our results provide the first molecular definition of C. difficile spores, opening up new opportunities for the development of diagnostic and therapeutic approaches.Clostridium difficile is a gram-positive, spore-forming, anaerobic bacterium that can asymptomatically colonize the intestinal tracts of humans and other mammals (3, 30, 39). Antibiotic treatment can result in C. difficile overgrowth and can lead to clinical disease, ranging from diarrhea to life-threatening pseudomembranous colitis, particularly in immunocompromised hosts (2, 4, 7). In recent years, C. difficile has emerged as the major cause of nosocomial antibiotic-induced diarrhea, and it is frequently associated with outbreaks (21, 22). A contributing factor is that C. difficile can be highly infectious and difficult to contain, especially when susceptible patients are present in the same hospital setting (13).Person-to-person transmission of C. difficile is associated with the excretion of highly resistant spores in the feces of infected patients, creating an environmental reservoir that can confound many infection control measures (29, 44). Bacterial spores, which are metabolically dormant cells that are formed following asymmetric cell division, normally have thick concentric external layers, the spore coat and cortex, that protect the internal cytoplasm (15, 42). Upon germination, spores lose their protective external layers and resume vegetative growth (24, 27, 36). Bacillus spores and the spores of most Clostridium species germinate in response to amino acids, carbohydrates, or potassium ions (24, 36). In contrast, C. difficile spores show an increased level of germination in response to cholate derivatives found in bile (40, 41). Thus, spores are well adapted for survival and dispersal under a wide range of environmental conditions but will germinate in the presence of specific molecular signals (24, 36).While the spores of a number of Bacillus species, such as Bacillus subtilis and Bacillus anthracis, and those of other Clostridium species, such as Clostridium perfringens (15, 20), have been well characterized, research on C. difficile spores has been relatively limited. A greater understanding of C. difficile spore biology could be exploited to rationalize disinfection regimes, molecular diagnostics, and the development of targeted treatments such as vaccines. Here we describe a novel method to isolate highly purified C. difficile spores that maintain their resistance and infectious characteristics, thus providing a unique opportunity to study C. difficile spores in the absence of vegetative cells. A thorough proteomic and genomic analysis of the spore provides novel insight into the unique composition and predictive biological properties of C. difficile spores that should underpin future research into this high-profile but poorly understood pathogen.  相似文献   

2.
To cause disease, Clostridium difficile spores must germinate in the host gastrointestinal tract. Germination is initiated upon exposure to glycine and certain bile acids, e.g., taurocholate. Chenodeoxycholate, another bile acid, inhibits taurocholate-mediated germination. By applying Michaelis-Menten kinetic analysis to C. difficile spore germination, we found that chenodeoxycholate is a competitive inhibitor of taurocholate-mediated germination and appears to interact with the spores with greater apparent affinity than does taurocholate. We also report that several analogs of chenodeoxycholate are even more effective inhibitors. Some of these compounds resist 7α-dehydroxylation by Clostridium scindens, a core member of the normal human colonic microbiota, suggesting that they are more stable than chenodeoxycholate in the colonic environment.Clostridium difficile is a Gram-positive, spore-forming, anaerobic bacterium that is pathogenic for both humans and animals (33, 44). Infections caused by C. difficile range from mild diarrhea to more life-threatening conditions, such as pseudomembranous colitis (33). In the classic case, prior antibiotic treatment that disrupts the normally protective colonic flora makes patients susceptible to C. difficile infection (CDI) (35, 53). Other antibiotics, such as vancomycin and metronidazole, are the most commonly used treatments for CDI (54). However, because these antibiotics also disrupt the colonic flora, 10 to 40% of patients whose symptoms have been ameliorated suffer from relapsing CDI (15, 24). The annual treatment-associated cost for CDI in the United States is estimated to be between $750 million and $3.2 billion (8, 9, 16, 31). Moreover, the number of fatal cases of CDI has been increasing rapidly (14, 39). Thus, there is an urgent need to find alternative therapies for CDI.C. difficile infection likely is initiated by infection with the spore form of C. difficile (12). C. difficile elicits disease through the actions of two secreted toxins, TcdA and TcdB (48). TcdB was recently shown to be critical for pathogenesis in an animal model of disease (18). Since the toxins are produced by vegetative cells, not by spores (17), germination and outgrowth are prerequisites for pathogenesis.Spore germination is triggered by the interaction of small molecules, called germinants, with receptors within the spore inner membrane. These germinants vary by bacterial species and can include ions, amino acids, sugars, nucleotides, surfactants, or combinations thereof (43). The recognition of germinants triggers irreversible germination, leading to Ca2+-dipicolinic acid release, the uptake of water, the degradation of the cortex, and, eventually, the outgrowth of the vegetative bacterium (43). The germination receptors that C. difficile uses to sense the environment have not been identified. Based on homology searches, C. difficile germination receptors must be very different from known germination receptors (42), but they appear to be proteinaceous (13).Taurocholate, a primary bile acid, has been used for approximately 30 years by researchers and clinical microbiologists to increase colony formation by C. difficile spores from patient and environmental samples (3, 49, 51, 52). This suggested that C. difficile spores interact with bile acids along the gastrointestinal (GI) tract and that spores use a host-derived signal to initiate germination.The liver synthesizes the two major primary bile acids, cholate and chenodeoxycholate (40). These compounds are modified by conjugation with either taurine (to give taurocholate or taurochenodeoxycholate) or glycine (producing glycocholate or glycochenodeoxycholate). Upon secretion into the digestive tract, bile aids in the absorption of fat and cholesterol; much of the secreted bile is actively absorbed and recycled back to the liver for reutilization (40). Though efficient, enterohepatic recirculation is not complete; bile enters the cecum of the large intestine at a concentration of approximately 2 mM (30).In the cecum, bile is modified by the normal, benign colonic flora. First, bile salt hydrolases found on the surfaces of many bacterial species remove the conjugated amino acid, producing the deconjugated primary bile acids cholate and chenodeoxycholate (40). These deconjugated primary bile acids are further metabolized by only a few species of intestinal bacteria, including Clostridium scindens. C. scindens actively transports unconjugated primary bile acids into the cytoplasm, where they are 7α-dehydroxylated, converting cholate to deoxycholate and chenodeoxycholate to lithocholate (21, 40). The disruption of the colonic flora by antibiotic treatment abolishes 7α-dehydroxylation activity (41).Building upon the work on Wilson and others (51, 52), we demonstrated that taurocholate and glycine, acting together, trigger the loss of the birefringence of C. difficile spores (45). All cholate derivatives (taurocholate, glycocholate, cholate, and deoxycholate) stimulate the germination of C. difficile spores (45). Recently it was shown that taurocholate binding is prerequisite to glycine binding (37). At physiologically relevant concentrations, chenodeoxycholate inhibits taurocholate-mediated germination (46) and also inhibits C. difficile vegetative growth, as does deoxycholate (45). In fact, C. difficile spores use the relative concentrations of the various bile acids as cues for germination within the host (10).Since chenodeoxycholate is absorbed by the colonic epithelium and metabolized to lithocholate by the colonic flora (25, 40), the use of chenodeoxycholate as a therapy against C. difficile disease might be hindered by its absorption and conversion to lithocholate.Here, we further characterize the interaction of C. difficile spores with various bile acids and demonstrate that chenodeoxycholate is a competitive inhibitor of taurocholate-mediated germination. Further, we identify chemical analogs of chenodeoxycholate that are more potent inhibitors of germination and that resist 7α-dehydroxylation by the colonic flora, potentially increasing their stability and effectiveness as inhibitors of C. difficile spore germination in the colonic environment.  相似文献   

3.
Superdormant spores of Bacillus cereus and Bacillus subtilis germinated just as well as dormant spores with pressures of 150 or 500 MPa and with or without heat activation. Superdormant B. subtilis spores also germinated as well as dormant spores with peptidoglycan fragments or bryostatin, a Ser/Thr protein kinase activator.Spores of Bacillus species are formed in sporulation, a process that is generally triggered by starvation for one or more nutrients (13, 19). These spores are metabolically dormant and extremely resistant to a large variety of environmental stresses, including heat, radiation, and toxic chemicals, and as a consequence of these properties, these spores can remain viable in their dormant state for many years (13, 18, 19). However, spores are constantly sensing their environment, and if nutrients return, the spores can rapidly return to growth through the process of spore germination (17). Spore germination is generally triggered by specific nutrients that bind to nutrient germinant receptors, with this binding alone somehow triggering germination. However, spore germination can also be triggered by many non-nutrient agents, including cationic surfactants such as dodecylamine, a 1:1 complex of Ca2+ with pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA], a major spore small molecule), very high pressures, specific peptidoglycan fragments, and bryostatin, an activator of Ser/Thr protein kinases (17, 19, 20). For nutrient germinants in particular, spore germination is also potentiated by a prior sublethal heat treatment termed heat activation (17).While normally the great majority of spores in populations germinate relatively rapidly in response to nutrient germinants, a small percentage of spores germinate extremely slowly. These spores that are refractory to nutrient germination have been termed superdormant spores and are a major concern for the food industry (8). Recently superdormant spores of three Bacillus species have been isolated by repeated germination of spore populations with specific nutrient germinants and isolation of remaining dormant spores (5, 6). These superdormant spores germinate extremely poorly with the nutrient germinants used in superdormant spore isolation, as well as with other nutrient germinants. All of the specific defects leading to spore superdormancy are not known, although an increased level of receptors for specific nutrient germinants decreases levels of superdormant spores obtained with the nutrients that are ligands for these receptors (5). Superdormant spores also have significantly higher temperature optima for heat activation of nutrient germination than the spore population as a whole (7).In contrast to the poor germination of superdormant spores with nutrient germinants, superdormant spores germinate normally with dodecylamine and Ca-DPA (5, 6). This is consistent with possible roles of nutrient germinant receptor levels and/or heat activation temperature optima in affecting spore superdormancy, since neither dodecylamine nor Ca-DPA triggers Bacillus spore germination through nutrient germinant receptors, and germination with these agents is also not stimulated by heat activation (11, 15, 17). However, the effects of high pressures, peptidoglycan fragments, and bryostatin, all of which almost certainly trigger spore germination by mechanisms at least somewhat different than triggering of germination by nutrients, dodecylamine, and Ca-DPA (2, 3, 11, 15, 20, 22, 23), have not been tested for their effects on superdormant spores. Consequently, we have compared the germination of dormant and superdormant spores of two Bacillus species by high-pressures, peptidoglycan fragments, and bryostatin.The spores used in this work were from Bacillus subtilis PS533 (16), a derivative of strain 168 that also carries plasmid pUB110, providing resistance to kanamycin (10 μg/ml), and Bacillus cereus T (originally obtained from H. O. Halvorson). Spores of these strains were prepared and purified as described previously (6, 10, 12). Superdormant spores of B. subtilis were prepared by germination following heat activation at 75°C for 30 min by two germination treatments at 37°C with 10 mM l-valine for 2 h, followed by isolation of remaining dormant spores, all as described previously (5, 10, 12). These superdormant spores germinated extremely poorly with 10 mM valine at 37°C, giving ≤10% germination in 2 h at 37°C, while the initial spore population exhibited >95% germination under the same conditions (data not shown). Superdormant B. cereus spores were isolated similarly, although heat activation was at 65°C for 30 min and the germinant was 5 mM inosine as described previously (6). These superdormant B. cereus spores exhibited <5% germination with inosine in 2 h at 37°C compared to the >95% germination of the initial dormant spores under the same conditions (data not shown).  相似文献   

4.
Clostridium sordellii is a spore-forming, obligately anaerobic, Gram-positive bacterium that can cause toxic shock syndrome after gynecological procedures. Although the incidence of C. sordellii infection is low, it is fatal in most cases. Since spore germination is believed to be the first step in the establishment of Bacilli and Clostridia infections, we analyzed the requirements for C. sordellii spore germination in vitro. Our data showed that C. sordellii spores require three structurally different amino acids and bicarbonate for maximum germination. Unlike the case for Bacilli species, d-alanine had no effect on C. sordellii spore germination. C. sordellii spores germinated only in a narrow pH range between 5.7 and 6.5. In contrast, C. sordellii spore germination was significantly less sensitive to temperature changes than that of the Bacilli. The analysis of the kinetics of C. sordellii spore germination showed strong allosteric behavior in the binding of l-phenylalanine and l-alanine but not in that of bicarbonate or l-arginine. By comparing germinant apparent binding affinities to their known in vivo concentrations, we postulated a mechanism for differential C. sordellii spore activation in the female reproductive tract.Clostridium sordellii is an anaerobic, Gram-positive, spore-forming bacterium that is commonly found in soil and in the intestines of animals (4). Many C. sordellii strains are nonpathogenic; however, virulent strains cause lethal infections in several animal species, such as hemorrhagic enteritis in foals, sheep, and cattle (5, 10, 16, 28), omphalitis in foals (43), and wound infection in humans (4, 35).C. sordellii also can cause life-threatening necrotizing infections after gynecological procedures (4). In addition, fatal cases of C. sordellii endometritis following medical abortion with a mifepristone-misoprostol combination have been reported recently (13, 19, 56). The increased use of mifepristone-misoprostol for medical abortion may result in larger numbers of C. sordellii infections (38, 40).Although C. sordellii rarely has been identified in the genital tract, a correlation between gynecological procedures and C. sordellii-mediated toxic shock syndrome is apparent (19). Pregnancy, childbirth, or abortion may predispose some women to acquire C. sordellii in the vaginal tract (19). Under these conditions, C. sordellii infections result in an almost 100% mortality rate.Since there is no national system for tracking and reporting complications associated with gynecological procedures, the identification of the true rates of reproductive tract infections in women is not readily available (8). Therefore, the number of known C. sordellii-associated infections, although low, may be underreported (19, 29). Furthermore, unsafe abortion practices in developing countries cause large mortality rates due to complicating infections (24, 34). In many cases, however, the causative agent of the abortion-associated sepsis have not been characterized (24). Thus, the worldwide morbidity and mortality associated with C. sordellii infections is not currently known.C. sordellii produces several virulence factors. The two major toxins are the lethal toxin (TcsL) and the hemorrhagic toxin (37, 46). The lethal toxin produced by C. sordellii is causally involved in enteritis of domestic animals and in systemic toxicity following infections of humans (46). Furthermore, TcsL is associated with rapid mortality in C. sordellii endometritis rodent models (26). Interestingly, TcsL cytopathic effects are increased at low pH, a characteristic found in the vaginal tract (48). The hemorrhagic toxin is not well characterized, but it has been reported to cause dermal and intestinal necrosis in guinea pigs (6, 52).C. sordellii, like other Bacilli and Clostridia species, has the ability to form metabolically dormant spores that are extremely resistant to environmental stresses, such as heat, radiation, and toxic chemicals (42, 55). Upon encountering a suitable environment, spores germinate into vegetative cells, the form that is responsible for toxin production and disease onset (39, 54).In most cases, the germination process initially is triggered by the detection of low-molecular-weight germinants by a sensitive biosensor (39, 54). This sensor consists of a proteinaceous germination (Ger) receptor encoded, in general, by a tricistronic operon. Spore germination requirements have been studied most extensively for Bacilli and can be initiated by a variety of factors, including amino acids, sugars, and nucleosides (20, 30).Spore germination in the Clostridia generally requires combinations of multiple germinants. The germination of spores of proteolytic Clostridium botulinum types A and B was triggered by a defined three-component mixture comprised of l-alanine (or l-cysteine), l-lactate (or sodium thioglycolate), and sodium bicarbonate (3). In contrast, the optimum germination of spores of nonproteolytic C. botulinum types B, E, and F required binary combinations of l-alanine-l-lactate, l-cysteine-l-lactate, and l-serine-l-lactate (45).Clostridium difficile is a human pathogen that can cause fulminant colitis (11). Interestingly, C. difficile does not encode any known Ger receptors (53). However, it is likely that germination receptors exist, because C. difficile spores must germinate in order to complete their life cycle. While C. difficile germination receptors remain elusive, the spores of C. difficile germinate in rich medium supplemented with bile salts (62). More recently, taurocholate (a bile salt) and glycine (an amino acid) were shown to act as cogerminants for C. difficile spore germination (57, 61).Clostridium bifermentans is a close relative of C. sordellii (14). The minimum requirement for C. bifermentans spore germination was the presence of l-alanine, l-phenylalanine, and l-lactate (59). In addition, an unknown factor present in yeast extract was suggested to enhance germination (59). However, the Ger receptors involved in C. bifermentans spore germination are not known.Even though many Bacilli and Clostridia species use similar metabolites as germinants, the mechanisms of germinant recognition remain to be elucidated. Unfortunately, the multimeric interactions of Ger receptor complexes and the hydrophobic nature of the Ger receptor subunits have hindered our understanding of the mechanism of germinant recognition.To understand the molecular determinants of germinant recognition, we recently applied kinetic methods to study bacterial spore germination (1, 2, 18). Spore germination can be analyzed quantitatively by fitting optical density (OD) decreases to the Michaelis-Menten equation (2). The kinetic parameters obtained allow the determination of the apparent binding affinity (Km) of spores for the different cogerminants and the maximum rate of spore germination (Vmax). In these instances, Km refers to the concentration of substrate required to reach half of the maximal germination rate. These parameters can, in turn, be used to determine the mechanism of germination and potential interactions between germination receptors. Furthermore, by comparing apparent Km values to germinant concentrations in vivo, models for spore-germinant complex distribution can be proposed, and rate-limiting steps for the germination process can be derived. Thus, kinetic analysis can yield information on spore activation even if the identities of the germination receptors are not known.Using this procedure, we were able to determine the mechanism for Bacillus anthracis germination with inosine and l-alanine. In turn, this information was used to design nucleoside analogs that inhibit B. anthracis spore germination in vitro and protect macrophages from anthrax cytotoxicity (2).Since C. sordellii germination receptors have not been identified, we used chemical probes and kinetic methods to investigate the conditions necessary for spore germination. We found that C. sordellii spores germinate better at slightly acidic pH. Furthermore, germination rates varied slightly from 25 to 40°C. We also found that C. sordellii spores have an absolute requirement for a small amino acid, a basic amino acid, an aromatic amino acid, and bicarbonate (NaHCO3) for efficient germination. Kinetic analysis showed allosteric interaction for the putative l-phenylalanine and l-alanine germination receptors. In contrast, l-arginine or bicarbonate recognition followed typical Michaelis-Menten kinetics. The implication of germinant recognition and host environment is discussed.  相似文献   

5.
The simultaneous nutrient germination of hundreds of individual wild-type spores of three Bacillus species and a number of Bacillus subtilis strains has been measured by two new methods, and rates of release of the great majority of the large pool of dipicolinic acid (DPA) from individual spores of B. subtilis strains has been measured by Raman spectroscopy with laser tweezers. The results from these analyses and published data have allowed a number of significant conclusions about the germination of spores of Bacillus species as follows. (i) The time needed for release of the great majority of a Bacillus spore''s DPA once rapid DPA release had begun (ΔTrelease) during nutrient germination was independent of the concentration of nutrient germinant used, the level of the germinant receptors (GRs) that recognize nutrient germinants used and heat activation prior to germination. Values for ΔTrelease were generally 0.5 to 3 min at 25 to 37°C for individual wild-type spores. (ii) Despite the conclusion above, germination of individual spores in populations was very heterogeneous, with some spores in wild-type populations completing germination ≥15-fold slower than others. (iii) The major factor in the heterogeneity in germination of individual spores in populations was the highly variable lag time, Tlag, between mixing spores with nutrient germinants and the beginning of ΔTrelease. (iv) A number of factors decrease spores'' Tlag values including heat activation, increased levels of GRs/spore, and higher levels of nutrient germinants. These latter factors appear to affect the level of activated GRs/spore during nutrient germination. (v) The conclusions above lead to the simple prediction that a major factor causing heterogeneity in Bacillus spore germination is the number of functional GRs in individual spores, a number that presumably varies significantly between spores in populations.Spores of various Bacillus species are metabolically dormant and can survive for years in this state (30). However, spores constantly sense their environment, and if appropriate small molecules termed germinants are present, spores can rapidly return to life in the process of germination followed by outgrowth (25, 29, 30). The germinants that most likely trigger spore germination in the environment are low-molecular-weight nutrient molecules, the identities of which are strain and species specific, including amino acids, sugars, and purine nucleosides. Metabolism of these nutrient germinants is not needed for the triggering of spore germination. Rather, these germinants are recognized by germinant receptors (GRs) located in the spore''s inner membrane that recognize their cognate germinants in a stereospecific manner (17, 24, 25, 29). Spores have a number of such GRs, with three functional GRs in Bacillus subtilis spores and even more in Bacillus anthracis, Bacillus cereus, and Bacillus megaterium spores (6, 29, 30). Binding of nutrient germinants to some single GRs is sufficient to trigger spore germination, for example the triggering of B. subtilis spore germination by binding of l-alanine or l-valine to the GerA GR. However, many GRs cooperate such that binding of germinants by ≥2 different GRs is needed to trigger germination (2, 29): for example, the triggering of B. subtilis spore germination by the binding of components of a mixture of l-asparagine, d-glucose, d-fructose, and K+ ions (AGFK) to the GerB and GerK GRs. The binding of nutrient germinants to GRs triggers subsequent events in germination, although how this is accomplished is not known.The first readily measured biochemical event after addition of nutrient germinants to Bacillus spores is the rapid release of the spore''s large depot (∼10% of spore dry weight) of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) plus its chelated divalent cations, predominantly Ca2+ (Ca-DPA), from the spore core (25, 29). Ca-DPA release then results in the activation of two redundant cortex-lytic enzymes (CLEs), CwlJ and SleB, which hydrolyze the spore''s peptidoglycan cortex layer (16, 22, 27, 29). CwlJ is activated by Ca-DPA as it is released from the spore while SleB is activated only after most DPA is released (17, 20, 22, 26, 27). Cortex hydrolysis ultimately allows the spore core to expand and take up more water, raising the core water content from the 35 to 45% of wet weight in the dormant spore to the 80% of wet weight characteristic of growing cells. Full hydration of the spore core then allows enzyme action, metabolism, and macromolecular synthesis to resume in the now fully germinated spore.Germination of spores in populations is very heterogeneous, with some spores germinating rapidly and some extremely slowly (4, 5, 9, 11, 13-15, 19, 26, 31, 32). Where it has been studied, the reason for this heterogeneity has been suggested to be due to a variable lag period (Tlag) between the time of mixing spores with a germinant and the time at which rapid DPA release begins, since once rapid DPA release begins, the time required for release of almost all DPA as well as for subsequent cortex hydrolysis is generally rather short compared to Tlag values in individual spores (5, 11, 13-15, 19, 26, 31, 32). The times required for DPA release and cortex hydrolysis are also similar in wild-type spores with both very short and long Tlag values (5, 15, 19, 27). The reasons for the variability in Tlag times between individual spores in populations are not known, although there are reports that both activation of spores for germination by a sublethal heat treatment (heat activation) as well as increasing concentrations of nutrient germinants can shorten Tlag values (12, 14, 15, 18, 32). However, there has been no detailed study of the causes of the variability in Tlag values between very large numbers of individual spores in populations.In order to study the heterogeneity in spore germination thoroughly, methods are needed to follow the germination of hundreds of individual spores over several hours. Initial studies of the germination of individual spores examined a single spore in a phase-contrast microscope and followed the germination of this spore by changes in the core''s refractive index due to DPA release and core swelling (14, 15, 32, 34). However, this method is labor-intensive for gathering data with hundreds of individual spores. More recently, confocal microscopy and then surface adsorption and optical tweezers have been used to capture single spores, and germination events have been followed by methods such as Raman spectroscopy to directly measure DPA release, as well as phase-contrast microscopy and elastic light scattering (3, 5, 9, 10, 19, 26). While the latter recent advances have allowed accumulation of much information about germination, collection of this type of data for large numbers of individual spores is still labor-intensive, although use of dual optical traps (35) and perhaps multiple traps in the future may alleviate this problem. However, phase-contrast microscopy plus appropriate computer software has recently allowed the monitoring of many hundreds of individual spores for several hours, with automated assessment of various changes in the cells during the period of observation (19). In the present work, we have used both phase-contrast and differential interference contrast (DIC) microscopy to monitor the germination of many hundreds of individual spores of three Bacillus species adhered on either an agarose pad or a glass coverslip for 1 to 2 h. This work, as well as examination of times needed for release of most DPA once rapid DPA release has begun during germination of individual spores under a variety of conditions, has allowed detailed examination of the effects of heat activation, nutrient germinant concentration, GR numbers per spore, and individual CLEs on spore germination heterogeneity and on values of Tlag for individual spores.  相似文献   

6.
Spores of Bacillus subtilis contain a number of small, acid-soluble spore proteins (SASP) which comprise up to 20% of total spore core protein. The multiple α/β-type SASP have been shown to confer resistance to UV radiation, heat, peroxides, and other sporicidal treatments. In this study, SASP-defective mutants of B. subtilis and spores deficient in dacB, a mutation leading to an increased core water content, were used to study the relative contributions of SASP and increased core water content to spore resistance to germicidal 254-nm and simulated environmental UV exposure (280 to 400 nm, 290 to 400 nm, and 320 to 400 nm). Spores of strains carrying mutations in sspA, sspB, and both sspA and sspB (lacking the major SASP-α and/or SASP-β) were significantly more sensitive to 254-nm and all polychromatic UV exposures, whereas the UV resistance of spores of the sspE strain (lacking SASP-γ) was essentially identical to that of the wild type. Spores of the dacB-defective strain were as resistant to 254-nm UV-C radiation as wild-type spores. However, spores of the dacB strain were significantly more sensitive than wild-type spores to environmental UV treatments of >280 nm. Air-dried spores of the dacB mutant strain had a significantly higher water content than air-dried wild-type spores. Our results indicate that α/β-type SASP and decreased spore core water content play an essential role in spore resistance to environmentally relevant UV wavelengths whereas SASP-γ does not.Spores of Bacillus spp. are highly resistant to inactivation by different physical stresses, such as toxic chemicals and biocidal agents, desiccation, pressure and temperature extremes, and high fluences of UV or ionizing radiation (reviewed in references 33, 34, and 48). Under stressful environmental conditions, cells of Bacillus spp. produce endospores that can stay dormant for extended periods. The reason for the high resistance of bacterial spores to environmental extremes lies in the structure of the spore. Spores possess thick layers of highly cross-linked coat proteins, a modified peptidoglycan spore cortex, a low core water content, and abundant intracellular constituents, such as the calcium chelate of dipicolinic acid and α/β-type small, acid-soluble spore proteins (α/β-type SASP), the last two of which protect spore DNA (6, 42, 46, 48, 52). DNA damage accumulated during spore dormancy is also efficiently repaired during spore germination (33, 47, 48). UV-induced DNA photoproducts are repaired by spore photoproduct lyase and nucleotide excision repair, DNA double-strand breaks (DSB) by nonhomologous end joining, and oxidative stress-induced apurinic/apyrimidinic (AP) sites by AP endonucleases and base excision repair (15, 26-29, 34, 43, 53, 57).Monochromatic 254-nm UV radiation has been used as an efficient and cost-effective means of disinfecting surfaces, building air, and drinking water supplies (31). Commonly used test organisms for inactivation studies are bacterial spores, usually spores of Bacillus subtilis, due to their high degree of resistance to various sporicidal treatments, reproducible inactivation response, and safety (1, 8, 19, 31, 48). Depending on the Bacillus species analyzed, spores are 10 to 50 times more resistant than growing cells to 254-nm UV radiation. In addition, most of the laboratory studies of spore inactivation and radiation biology have been performed using monochromatic 254-nm UV radiation (33, 34). Although 254-nm UV-C radiation is a convenient germicidal treatment and relevant to disinfection procedures, results obtained by using 254-nm UV-C are not truly representative of results obtained using UV wavelengths that endospores encounter in their natural environments (34, 42, 50, 51, 59). However, sunlight reaching the Earth''s surface is not monochromatic 254-nm radiation but a mixture of UV, visible, and infrared radiation, with the UV portion spanning approximately 290 to 400 nm (33, 34, 36). Thus, our knowledge of spore UV resistance has been constructed largely using a wavelength of UV radiation not normally reaching the Earth''s surface, even though ample evidence exists that both DNA photochemistry and microbial responses to UV are strongly wavelength dependent (2, 30, 33, 36).Of recent interest in our laboratories has been the exploration of factors that confer on B. subtilis spores resistance to environmentally relevant extreme conditions, particularly solar UV radiation and extreme desiccation (23, 28, 30, 34 36, 48, 52). It has been reported that α/β-type SASP but not SASP-γ play a major role in spore resistance to 254-nm UV-C radiation (20, 21) and to wet heat, dry heat, and oxidizing agents (48). In contrast, increased spore water content was reported to affect B. subtilis spore resistance to moist heat and hydrogen peroxide but not to 254-nm UV-C (12, 40, 48). However, the possible roles of SASP-α, -β, and -γ and core water content in spore resistance to environmentally relevant solar UV wavelengths have not been explored. Therefore, in this study, we have used B. subtilis strains carrying mutations in the sspA, sspB, sspE, sspA and sspB, or dacB gene to investigate the contributions of SASP and increased core water content to the resistance of B. subtilis spores to 254-nm UV-C and environmentally relevant polychromatic UV radiation encountered on Earth''s surface.  相似文献   

7.
Clostridium difficile is a major cause of antibiotic-associated diarrheal disease in many parts of the world. In recent years, distinct genetic variants of C. difficile that cause severe disease and persist within health care settings have emerged. Highly resistant and infectious C. difficile spores are proposed to be the main vectors of environmental persistence and host transmission, so methods to accurately monitor spores and their inactivation are urgently needed. Here we describe simple quantitative methods, based on purified C. difficile spores and a murine transmission model, for evaluating health care disinfection regimens. We demonstrate that disinfectants that contain strong oxidizing active ingredients, such as hydrogen peroxide, are very effective in inactivating pure spores and blocking spore-mediated transmission. Complete inactivation of 106 pure C. difficile spores on indicator strips, a six-log reduction, and a standard measure of stringent disinfection regimens require at least 5 min of exposure to hydrogen peroxide vapor (HPV; 400 ppm). In contrast, a 1-min treatment with HPV was required to disinfect an environment that was heavily contaminated with C. difficile spores (17 to 29 spores/cm2) and block host transmission. Thus, pure C. difficile spores facilitate practical methods for evaluating the efficacy of C. difficile spore disinfection regimens and bringing scientific acumen to C. difficile infection control.Clostridium difficile is a Gram-positive, spore-forming, anaerobic bacterium that is a major cause of health care-acquired infections and antibiotic-associated diarrhea (2). In recent years, several genetic variants of C. difficile have emerged as important health care pathogens (6). Perhaps most notable is the “hypervirulent” variant, commonly referred to as PCR ribotype 027/restriction endonuclease analysis (REA) group BI, that produces elevated levels of toxins TcdA and TcdB (17, 19). Other virulent ribotypes that display extensive heterogeneity among their toxin protein sequences (26) and gene activities (8) have emerged. Using whole-genome sequencing, we demonstrated that there are broad genetic differences between the entire genomes of several common variants, including ribotype/REA group variants 012/R, 017/CF, and 027/BI used in this study (12, 27, 31). In contrast, phylogeographic analysis of 027/BI isolates from Europe and the United States demonstrates that this clade is extremely clonal and implies recent transcontinental spread of hypervirulent C. difficile (12).C. difficile is distinct from many other health care pathogens because it produces highly infectious spores that are shed into the environment (25, 28). C. difficile spores can resist disinfection regimens that normally inactivate other health care pathogens, such as methicillin-resistant Staphylococcus aureus and vancomycin-resistant enterococci, therefore challenging current infection control measures (2). A multifaceted approach is normally used to control C. difficile in health care facilities (32). Interventions include antimicrobial stewardship, increased clinical awareness, patient isolation (11), and enhanced environmental disinfection regimens based on hydrogen peroxide (H2O2) vapor (HPV) (4). While attempts to break the spore-mediated infection cycle and interrupt these efficient routes of transmission are important for infection control measures, there is little quantitative evidence indicating which interventions are most effective (7). Here we describe the exploitation of pure C. difficile spores (16) and a murine transmission model (15) in simple, practical methods to quantitatively monitor the impact of health care disinfection regimens on C. difficile viability. These methods can be used to optimize disinfection regimens targeted at C. difficile.  相似文献   

8.
Bacillus cereus spores are assembled with a series of concentric layers that protect them from a wide range of environmental stresses. The outermost layer, or exosporium, is a bag-like structure that interacts with the environment and is composed of more than 20 proteins and glycoproteins. Here, we identified a new spore protein, ExsM, from a β-mercaptoethanol extract of B. cereus ATCC 4342 spores. Subcellular localization of an ExsM-green fluorescent protein (GFP) protein revealed a dynamic pattern of fluorescence that follows the site of formation of the exosporium around the forespore. Under scanning electron microscopy, exsM null mutant spores were smaller and rounder than wild-type spores, which had an extended exosporium (spore length for the wt, 2.40 ± 0.56 μm, versus that for the exsM mutant, 1.66 ± 0.38 μm [P < 0.001]). Thin-section electron microscopy revealed that exsM mutant spores were encased by a double-layer exosporium, both layers of which were composed of a basal layer and a hair-like nap. Mutant exsM spores were more resistant to lysozyme treatment and germinated with higher efficiency than wild-type spores, and they had a delay in outgrowth. Insertional mutagenesis of exsM in Bacillus anthracis ΔSterne resulted in a partial second exosporium and in smaller spores. In all, these findings suggest that ExsM plays a critical role in the formation of the exosporium.Bacillus cereus and Bacillus anthracis are closely related members of the Bacillus cereus group (47). Although B. cereus is mainly an apathogenic organism, certain isolates can cause two different types of food poisoning, emetic syndrome and diarrheal disease (18). The emetic syndrome is caused by ingestion of cereulide, a heat-resistant toxin produced by vegetative cells contaminating the food (30), while the diarrheal disease occurs when spores germinate in the intestinal tract. Spores are also the infective agent in anthrax, a disease caused by B. anthracis (64).B. cereus and B. anthracis differentiate into spores when faced with nutrient deprivation. The spore is a dormant cell type that can remain viable for decades until favorable conditions induce germination and the resumption of vegetative growth. The remarkable resistance properties of the spore result from its unique architecture, consisting of a series of concentric protective layers (51). The spore core contains the genetic material and is surrounded by the cortex, a thick layer of modified peptidoglycan that promotes a highly dehydrated state. Encasing the core and the cortex, the coat is a multilayer protein shell that provides mechanical and chemical resistance. In addition, both the cortex and coat contribute to spore germination (17). Separated from the coat by an interspace, the exosporium encloses the rest of the spore, and it is composed of an inner basal layer and an outer hair-like nap (25).Being the most external layer of the spore, the exosporium interacts directly with the environment and as such provides a semipermeable barrier that may exclude large molecules, like antibodies and hydrolytic enzymes (3, 23, 24, 54). However, the exosporium does not appear to contribute to the typical resistance properties of the spore (6, 35, 60). Also, the exosporium is not necessary in anthrax pathogenesis when tested under laboratory conditions (7, 27, 59), although it is able to down-modulate the innate immune response to spores and mediate adhesion to host tissues (4, 8, 43, 44). The exosporium may also help the spore avoid premature germination in unsustainable environments, since it contains two enzymes, alanine racemase (Alr) and inosine hydrolase (Iunh), that can inactivate low quantities of the germinants l-alanine and inosine, respectively (6, 48, 55, 61). However, regulation of germination by the exosporium is poorly understood. Mutation of exosporial proteins has resulted in only negligible and inconsistent germination phenotypes (2, 5, 27, 28, 52, 54).The exosporium is composed of at least 20 proteins and glycoproteins in tight or loose association (48, 53, 57, 61, 65). These proteins are synthesized in the mother cell and always start self-assembly at the forespore pole near the middle of the mother cell, concurrently with the cortex and coat formation (42). Exosporium assembly is discontinuous and starts with a synthesis of a substructure known as the cap, which likely contains only a subset of the proteins present in the exosporium (55). After cap formation, construction of the rest of the exosporium requires the expression of ExsY (6). BclA is the main component of the hair-like nap on the external side of the exosporium, and it is linked to the basal layer through interaction with ExsFA/BxpB (54, 58). In addition, CotE participates in the correct attachment of the exosporium to the spore (27).Despite these findings, exosporium assembly continues to be a poorly understood process, and many questions remain regarding its composition and the regulation of its synthesis. In this study, we characterized a new spore protein, ExsM, which plays a key role in assembly of the exosporium. In B. cereus, inactivation of exsM resulted in spores with an unusual double-layer exosporium, and a similar phenotype was also observed in B. anthracis exsM null mutant spores. Finally, double-layer exosporium spores allowed us to study the role of the exosporium in germination and outgrowth.  相似文献   

9.
Spores of Bacillus species are said to be committed when they continue through nutrient germination even when germinants are removed or their binding to spores'' nutrient germinant receptors (GRs) is both reversed and inhibited. Measurement of commitment and the subsequent release of dipicolinic acid (DPA) during nutrient germination of spores of Bacillus cereus and Bacillus subtilis showed that heat activation, increased nutrient germinant concentrations, and higher average levels of GRs/spore significantly decreased the times needed for commitment, as well as lag times between commitment and DPA release. These lag times were also decreased dramatically by the action of one of the spores'' two redundant cortex lytic enzymes (CLEs), CwlJ, but not by the other CLE, SleB, and CwlJ action did not affect the timing of commitment. The timing of commitment and the lag time between commitment and DPA release were also dependent on the specific GR activated to cause spore germination. For spore populations, the lag times between commitment and DPA release were increased significantly in spores that germinated late compared to those that germinated early, and individual spores that germinated late may have had lower appropriate GR levels/spore than spores that germinated early. These findings together provide new insight into the commitment step in spore germination and suggest several factors that may contribute to the large heterogeneity among the timings of various events in the germination of individual spores in spore populations.Spores of Bacillus species can remain dormant for long times and are extremely resistant to a variety of environmental stresses (26). However, under appropriate conditions, normally upon the binding of specific nutrients to spores'' nutrient germinant receptors (GRs), spores can come back to active growth through a process called germination followed by outgrowth (19, 20, 25, 26). Germination of Bacillus subtilis spores can be triggered by l-alanine or l-valine or a combination of l-asparagine, d-glucose, d-fructose, and K+ (AGFK). These nutrient germinants trigger germination by binding to and interacting with GRs that have been localized to the spore''s inner membrane (12, 20). l-Alanine and l-valine bind to the GerA GR, while the AGFK mixture triggers germination by interacting with both the GerB and GerK GRs (25). Normally, l-asparagine alone does not trigger B. subtilis spore germination. However, a mutant form of the GerB GR, termed GerB*, displays altered germinant specificity such that l-asparagine alone will trigger the germination of gerB* mutant spores (1, 18).A number of events occur in a defined sequence during spore germination. Initially, exposure of spores to nutrient germinants causes a reaction that commits spores to germinate, even if the germinant is removed or displaced from its cognate GR (7, 10, 21, 27, 28). This commitment step is followed by release of monovalent cations, as well as the spore core''s large pool of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) along with divalent cations, predominantly Ca2+, that are chelated with DPA (Ca-DPA). In Bacillus spores, the release of Ca-DPA triggers the hydrolysis of spores'' peptidoglycan cortex by either of two cortex lytic enzymes (CLEs), CwlJ and SleB (11, 16, 23). CwlJ is activated during germination by Ca-DPA as it is being released from individual spores, while SleB activation requires that most Ca-DPA be released (14, 16, 17). Cortex hydrolysis, in turn, allows the spore core to expand and fully hydrate, which leads to activation of enzymes and initiation of metabolism in the spore core (21, 25).As noted above, commitment is the first event that can be assessed during spore germination, although the precise mechanism of commitment is not known. Since much has been learned about proteins important in spore germination in the many years since commitment was last studied (25, 26), it seemed worth reexamining commitment, with the goal of determining those factors that influence this step in the germination process. Knowledge of factors important in determining kinetics of commitment could then lead to an understanding of what is involved in this reaction.Kinetic analysis of spore germination, as well as commitment, has mostly been based on the decrease in optical density at 600 nm (OD600) of spore suspensions, which monitors a combination of events that occur well after commitment, including DPA release, cortex hydrolysis, and core swelling (25-27). In the current work, we have used a germination assay that measures DPA release, an early event in spore germination, and have automated this assay to allow routine measurement of commitment, as well as DPA release from large numbers of spore samples simultaneously. This assay has allowed comparison of the kinetics of DPA release and commitment during germination and study of the effects of heat activation, germinant concentration, GR levels, and CLEs on commitment.  相似文献   

10.
Understanding the molecular basis of Clostridium difficile infection is a prerequisite to the development of effective countermeasures. Although there are methods for constructing gene-specific mutants of C. difficile, currently there is no effective method for generating libraries of random mutants. In this study, we developed a novel mariner-based transposon system for in vivo random mutagenesis of C. difficile R20291, the BI/NAP1/027 epidemic strain at the center of the C. difficile outbreaks in Stoke Mandeville, United Kingdom, in 2003 to 2004 and 2004 to 2005. Transposition occurred at a frequency of 4.5 (±0.4) × 10−4 per cell to give stable insertions at random genomic loci, which were defined only by the nucleotide sequence TA. Furthermore, mutants with just a single transposon insertion were generated in an overwhelming majority (98.3% in this study). Phenotypic screening of a C. difficile R20291 random mutant library yielded a sporulation/germination-defective clone with an insertion in the germination-specific protease gene cspBA and an auxotroph with an insertion in the pyrimidine biosynthesis gene pyrB. These results validate our mariner-based transposon system for use in forward genetic studies of C. difficile.Clostridium difficile infection is widely recognized as the leading cause of health care-associated diarrhea in North America and Europe. Infection usually follows antibiotic treatment, which disrupts the native gastrointestinal microflora and thus allows C. difficile to proliferate. The emergence of so-called “epidemic” or “hypervirulent” strains of C. difficile over the last 5 to 10 years has compounded an already serious problem. Classed as BI/NAP1/027, these epidemic strains are believed to cause a more severe disease and lead to increased mortality and relapse rates (11, 20, 24).Understanding the genetic and molecular basis of C. difficile infection will be a crucial step in the development of effective countermeasures. Methods for directed gene inactivation in C. difficile have recently been described (7, 21). This has opened the way for reverse genetic studies, in which the exact role of a specific gene, hypothesized to be important in a given phenotype, can be elucidated experimentally. By way of contrast, forward genetic studies aim to identify the genetic basis of a particular phenotype without making any assumptions about the genes involved. In forward genetic studies, transposons are often used to generate libraries of random insertion mutants. Libraries are then screened to identify mutants that are defective in a particular phenotype. Identification of the gene or genes which have been inactivated by transposon insertion then implicates them as having a role in that particular phenotype. Recently, just such an approach was used to identify a novel toxin-regulatory locus in Clostridium perfringens (29). This study elegantly demonstrated the power of forward genetic studies in bacterial pathogens.A number of transposon mutagenesis systems have been described for Gram-positive bacteria (2, 3, 15, 16, 29, 32). Two different systems have recently been developed for use in C. perfringens (15, 29). Both are in vitro mutagenesis systems which rely on being able to transform the recipient organism. As such, they are not suitable for use in C. difficile because in the laboratory at present, recombinant DNA can be transferred into C. difficile only via conjugation. The conjugative transposons Tn916 and Tn5397 have been studied in C. difficile, but both have been found either to have a strong target site preference or to yield multiple insertions in individual clones (9, 30). Therefore, neither is well suited to generating libraries of random C. difficile mutants.We reasoned that a mariner-based transposon mutagenesis system would be an effective tool for generating libraries of random C. difficile mutants. The mariner-transposable element Himar1 has been shown to insert randomly into the genomes of many bacterial species (3, 6, 16, 17, 32). The cognate Himar1 transposase is the only factor required for transposition, which occurs via a cut-and-paste mechanism (13, 14). The transposon itself is defined by inverted terminal repeats (ITRs) at either end and inserts into a TA target site. This is highly appropriate for an organism with a low-GC content such as C. difficile. In this study, we have developed a novel mariner-based transposon system for in vivo random mutagenesis of C. difficile. Moreover, we have demonstrated the system in C. difficile R20291, the BI/NAP1/027 epidemic strain at the center of the C. difficile outbreaks in Stoke Mandeville, United Kingdom, in 2003 to 2004 and 2004 to 2005. This new genetic tool opens the way for forward genetic studies of C. difficile.  相似文献   

11.
12.
Superdormant spores of Bacillus subtilis and Bacillus megaterium were isolated in 4 to 12% yields following germination with high nutrient levels that activated one or two germinant receptors. These superdormant spores did not germinate with the initial nutrients or those that stimulated other germinant receptors, and the superdormant spores'' defect was not genetic. The superdormant spores did, however, germinate with Ca2+-dipicolinic acid or dodecylamine. Although these superdormant spores did not germinate with high levels of nutrients that activated one or two nutrient germinant receptors, they germinated with nutrient mixtures that activated more receptors, and using high levels of nutrient mixtures activating more germinant receptors decreased superdormant spore yields. The use of moderate nutrient levels to isolate superdormant spores increased their yields; the resultant spores germinated poorly with the initial moderate nutrient concentrations, but they germinated well with high nutrient concentrations. These findings suggest that the levels of superdormant spores in populations depend on the germination conditions used, with fewer superdormant spores isolated when better germination conditions are used. These findings further suggest that superdormant spores require an increased signal for triggering spore germination compared to most spores in populations. One factor determining whether a spore is superdormant is its level of germinant receptors, since spore populations with higher levels of germinant receptors yielded lower levels of superdormant spores. A second important factor may be heat activation of spore populations, since yields of superdormant spores from non-heat-activated spore populations were higher than those from optimally activated spores.Spores of various Bacillus species are formed in sporulation and are metabolically dormant and very resistant to environmental stress factors (21, 37). While such spores can remain in this dormant, resistant state for long periods, they can return to life rapidly through the process of germination, during which the spore''s dormancy and extreme resistance are lost (36). Spore germination has long been of intrinsic interest, and continues to attract applied interest, because (i) spores of a number of Bacillus species are major agents of food spoilage and food-borne disease and (ii) spores of Bacillus anthracis are a major bioterrorism agent. Since spores are much easier to kill after they have germinated, it would be advantageous to trigger germination of spores in foods or the environment and then readily inactivate the much less resistant germinated spores. However, this simple strategy has been largely nullified because germination of spore populations is heterogeneous, with some spores, often called superdormant spores, germinating extremely slowly and potentially coming back to life long after treatments are applied to inactivate germinated spores (8, 9, 16). The concern over superdormant spores in populations also affects decisions such as how long individuals exposed to B. anthracis spores should continue to take antibiotics, since spores could remain dormant in an individual for long periods and then germinate and cause disease (3, 11).In many species, spore germination can be increased by a prior activation step, generally a sublethal heat treatment, although the changes taking place during heat activation are not known (16). Spore germination in Bacillus species is normally triggered by nutrients such as glucose, amino acids, or purine ribosides (27, 36). These agents bind to germinant receptors located in the spore''s inner membrane that are specific for particular nutrients. In Bacillus subtilis, the GerA receptor responds to l-alanine or l-valine, while the GerB and GerK receptors act cooperatively to respond to a mixture of l-asparagine (or l-alanine), d-glucose, d-fructose and K+ ions (AGFK [or Ala-GFK]) (1, 27, 36). There are even more functional germinant receptors in Bacillus megaterium spores, and these respond to d-glucose, l-proline, l-leucine, l-valine, or even salts, such as KBr (6). Glucose appears to trigger germination of B. megaterium spores through either of two germinant receptors, GerU or GerVB, while l-proline triggers germination through only the GerVB receptor, and KBr germination is greatly decreased by the loss of either GerU or GerVB (6). Nutrient binding to the germinant receptors triggers the release of small molecules from the spore core, most notably the huge depot (∼10% of spore dry weight) of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) present in spores predominantly as a 1:1 diluted chelate with Ca2+ (Ca-DPA) (35, 36). Ca-DPA release then triggers the activation of one of two redundant cortex lytic enzymes (CLEs) that degrade the spore''s peptidoglycan cortex, and cortex degradation completes spore germination and allows progression into outgrowth and then vegetative growth (27, 33, 36).Spore germination can also be triggered by nonnutrient agents, including Ca-DPA and cationic surfactants (27, 33, 36). With B. subtilis spores, Ca-DPA triggers germination by activating one particular CLE, termed CwlJ, and bypasses the spore''s germinant receptors. Germination by the cationic surfactant dodecylamine also bypasses the germinant receptors, and this agent appears to release small molecules including Ca-DPA from the spore core either by opening a normal channel in the spore''s inner membrane for Ca-DPA and other small molecules or by creating such a channel (31, 38, 39).Almost all work on the specifics of the germination of spores of Bacillus species has focused on the majority of spores in populations, and little detailed attention has been paid to that minority of spores that either fail to germinate or germinate extremely slowly. However, it is these latter spores that are most important in unraveling the cause of superdormancy and perhaps suggesting a means to germinate and thus easily inactivate such superdormant spores. Consequently, we have undertaken the task of isolating superdormant spores from spore populations, using buoyant density centrifugation to separate dormant spores from germinated spores. The properties of these purified superdormant spores were then studied, and this information has suggested some reason(s) for spore superdormancy.  相似文献   

13.
Purified superdormant spores of Bacillus cereus, B. megaterium, and B. subtilis isolated after optimal heat activation of dormant spores and subsequent germination with inosine, d-glucose, or l-valine, respectively, germinate very poorly with the original germinants used to remove dormant spores from spore populations, thus allowing isolation of the superdormant spores, and even with alternate germinants. However, these superdormant spores exhibited significant germination with the original or alternate germinants if the spores were heat activated at temperatures 8 to 15°C higher than the optimal temperatures for the original dormant spores, although the levels of superdormant spore germination were not as great as those of dormant spores. Use of mixtures of original and alternate germinants lowered the heat activation temperature optima for both dormant and superdormant spores. The superdormant spores had higher wet-heat resistance and lower core water content than the original dormant spore populations, and the environment of dipicolinic acid in the core of superdormant spores as determined by Raman spectroscopy of individual spores differed from that in dormant spores. These results provide new information about the germination, heat activation optima, and wet-heat resistance of superdormant spores and the heterogeneity in these properties between individual members of dormant spore populations.Spores of Bacillus species are formed in sporulation and are metabolically dormant and extremely resistant to a variety of stress factors (31, 32). While spores can remain dormant for long periods, if given the proper stimulus, they can rapidly “return to life” in the process of spore germination followed by outgrowth (30). Since spores are generally present in significant amounts on many foodstuffs and growing cells of a number of Bacillus species are significant agents of food spoilage and food-borne disease (32), there is continued applied interest in spore resistance and germination. While dormant spores can be killed by a treatment such as wet heat, this requires high temperatures that are costly and detrimental to food quality. Consequently, there has long been interest in triggering spore germination in foodstuffs, since germinated spores have lost the extreme resistance of dormant spores and are relatively easy to kill. However, this strategy has been difficult to apply because of the significant heterogeneity in germination rates between individual spores in populations. One reflection of this heterogeneity is the extremely variable lag times following addition of germinants but prior to initiation of germination events; while these lag times can vary from 10 to 30 min for most spores in populations, some spores have lag times of many hours or even many days (2, 12, 13, 15, 25). The spores that are extremely slow to germinate have been termed superdormant spores, and populations of superdormant spores have recently been isolated from three Bacillus species, and their germination properties characterized (9, 10). These superdormant spores germinate extremely poorly with the original germinants used to remove dormant spores from spore populations, thus allowing superdormant spore isolation, and also poorly with a number of other germinants, in particular, germinants that target nutrient germinant receptors different than those activated to isolate the superdormant spores. However, the superdormant spores germinate reasonably well with mixtures of nutrient germinants that target multiple germinant receptors. All reasons for spore superdormancy are not known, but one contributing factor is the number of nutrient germinant receptors in the spore''s inner membrane that trigger spore germination by binding to nutrient germinants (9). The levels of these receptors are most likely in the tens of molecules per spore (24), and thus stochastic variation in receptor numbers might result in some spores with such low receptor numbers that these spores germinate very poorly (23). Indeed, 20- to 200-fold elevated levels of at least one nutrient germinant receptor greatly decreases yields of superdormant spores of Bacillus subtilis (9).Spores of Bacillus species generally exhibit a requirement for an activation step in order to exhibit maximum germination (17). Usually this activation is a sublethal heat treatment that for a spore population exhibits an optimum of 60 to 100°C depending on the species. Spores are also extremely resistant to wet heat, generally requiring temperatures of 80 to 110°C to achieve rapid spore killing, with the major factor influencing the wet-heat resistance of spores of mesophilic strains being the spore core''s water content, which can be as low as 30% of wet weight as water in a fully hydrated spore (8, 19, 27, 28, 31). Invariably, increases in core water content are associated with a decrease in spore wet-heat resistance (8, 19, 22, 25). While spore populations most often exhibit log-linear kinetics of wet-heat killing, the observation of tailing in such killing curves at high levels of killing is not uncommon, suggesting there is significant heterogeneity in the wet-heat resistances of individual spores in populations (27, 28). While there has been no comparable work suggesting that there is also heterogeneity in the temperature optima for heat activation of individual spores in populations, this certainly seems possible and indeed was suggested as one cause of spore superdormancy, as yields of superdormant spores from spore populations that are not heat activated are much higher (9, 10). Consequently, the current work was initiated to test the hypothesis that superdormant spores require heat activation temperatures that are higher than those of the original dormant spores. Once this was found to be the case, the wet-heat resistance and core water content of the superdormant and original dormant spores were compared, and the environment of the spore core''s major small molecule, pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) was assessed by Raman spectroscopy of individual spores.  相似文献   

14.
The Bacillus anthracis spore is the causative agent of the disease anthrax. The outermost structure of the B. anthracis spore, the exosporium, is a shell composed of approximately 20 proteins. The function of the exosporium remains poorly understood and is an area of active investigation. In this study, we analyzed the previously identified but uncharacterized exosporium protein ExsK. We found that, in contrast to other exosporium proteins, ExsK is present in at least two distinct locations, i.e., the spore surface as well as a more interior location underneath the exosporium. In spores that lack the exosporium basal layer protein ExsFA/BxpB, ExsK fails to encircle the spore and instead is present at only one spore pole, indicating that ExsK assembly to the spore is partially dependent on ExsFA/BxpB. In spores lacking the exosporium surface protein BclA, ExsK fails to mature into high-molecular-mass species observed in wild-type spores. These data suggest that the assembly and maturation of ExsK within the exosporium are dependent on ExsFA/BxpB and BclA. We also found that ExsK is not required for virulence in murine and guinea pig models but that it does inhibit germination. Based on these data, we propose a revised model of exosporium maturation and assembly and suggest a novel role for the exosporium in germination.During starvation, bacteria of the genus Bacillus differentiate into dormant, highly robust cell types called spores, thereby preserving their genomes during stressful and nutrient-poor conditions (10). Spores can withstand extremely harsh environmental insults, including toxic chemicals, UV radiation, and heat (31). When conditions again become favorable for cell survival, spores can return to vegetative cell growth through a process called germination (17, 18, 31, 49). Spores are formed in an approximately 8-h process during which the developing spore first forms as a compartment (the forespore) contained within the surrounding cell (the mother cell) (34). Ultimately, the mother cell envelope lyses, releasing the mature spore into the environment.Spores from all Bacillus species have similar architectures. At the spore interior is the core, which houses the spore chromosome. Surrounding the core is an inner membrane encased in a specialized peptidoglycan called the cortex and finally a series of outer layers that vary significantly among species (10). In some species, including Bacillus subtilis, the outermost structure is a protective layer called the coat, which guards the spore against reactive small molecules, degradative enzymes, and predation by other microbes (11, 17, 20, 38). Spores of other species, including the pathogens Bacillus anthracis, Bacillus cereus, and Bacillus thuringiensis and the nonpathogenic bacteria Bacillus megaterium and Bacillus odysseyi, have an additional structure called the exosporium, which surrounds the coat (24, 32, 47). The exosporium is composed of two structural units: the basal layer, which is a shell of proteins forming a hexagonal array, and a nap of hairlike protrusions extending outward from the basal layer (2, 32). A major component of the nap (and of the spore surface) is the collagen-like protein BclA (40, 43). The proteins that comprise the outer structures (the coat and exosporium) are synthesized in the mother cell cytoplasm, from which location they assemble onto the spore surface to form their respective structures (11).The function of the exosporium is poorly understood. Previous studies have implicated its contribution to germination, resistance to host cells and other stresses, adhesion to inert surfaces, and interactions with epithelial cells and macrophages (1, 6, 7, 13, 33, 41, 48; G. Chen, A. Driks, K. Tawfiq, M. Mallozzi, and S. Patil, submitted for publication). In most cases, however, the roles of individual exosporium proteins in each of these functions remain unclear, in part because the location of each protein within the exosporium is largely unknown.Interestingly, it appears that the exosporium is not essential for virulence of B. anthracis in several animal models (5, 7, 12, 13). Nonetheless, it is possible that in natural infections the exosporium plays a significant role. Because it is involved in attachment, the exosporium is also likely to have a significant impact on the persistence of B. anthracis spores in the environment.To gain insight into the molecular basis of exosporium assembly and function, we studied a previously identified but otherwise uncharacterized exosporium protein, ExsK. Using immunofluorescence microscopy (IFM), we found that ExsK is asymmetrically distributed on the surfaces of mature spores and is also present beneath the exosporium. In the absence of ExsFA/BxpB, ExsK was restricted to one spore pole, suggesting that the encirclement of the spore by ExsK depends on ExsFA/BxpB. Western blot analysis indicated that in mature spores ExsK is present in high-molecular-mass complexes, the formation of which is BclA dependent. Although ExsK is not required for several spore resistance properties or virulence, we found that it is required for normal germination. Our results provide a deeper understanding of the composition, function, and assembly of the B. anthracis exosporium and show that proteins comprising outer-spore structures can have multiple locations.  相似文献   

15.
The bacterial spore cortex is critical for spore stability and dormancy and must be hydrolyzed by germination-specific lytic enzymes (GSLEs), which allows complete germination and vegetative cell outgrowth. We created in-frame deletions of three genes that encode GSLEs that have been shown to be active in Bacillus anthracis germination: sleB, cwlJ1, and cwlJ2. Phenotypic analysis of individual null mutations showed that the removal of any one of these genes was not sufficient to disrupt spore germination in nutrient-rich media. This finding indicates that these genes have partially redundant functions. Double and triple deletions of these genes resulted in more significant defects. Although a small subset of ΔsleB ΔcwlJ1 spores germinate with wild-type kinetics, for the overall population there is a 3-order-of-magnitude decrease in the colony-forming efficiency compared with wild-type spores. ΔsleB ΔcwlJ1 ΔcwlJ2 spores are unable to complete germination in nutrient-rich conditions in vitro. Both ΔsleB ΔcwlJ1 and ΔsleB ΔcwlJ1 ΔcwlJ2 spores are significantly attenuated, but are not completely devoid of virulence, in a mouse model of inhalation anthrax. Although unable to germinate in standard nutrient-rich media, spores lacking SleB, CwlJ1, and CwlJ2 are able to germinate in whole blood and serum in vitro, which may explain the persistent low levels of virulence observed in mouse infections. This work contributes to our understanding of GSLE activation and function during germination. This information may result in identification of useful therapeutic targets for the disease anthrax, as well as provide insights into ways to induce the breakdown of the protective cortex layer, facilitating easier decontamination of resistant spores.Bacillus anthracis, a gram-positive spore-forming bacterium, is the causative agent of anthrax. The dormant spore form is the infectious particle and produces three different forms of the disease depending on the route of entry into a suitable host (8). When spores enter through a skin lesion and when they are ingested, they cause cutaneous and gastrointestinal anthrax, respectively. Spores entering through the lungs cause the most severe form of the disease, inhalation anthrax, which is often fatal even with aggressive antibiotic therapy (1, 8, 34). Because true pneumonias are rarely seen in victims, it is believed that inhaled spores do not germinate in the lung but are phagocytosed by alveolar macrophages and germinate intracellularly en route to the mediastinal lymph nodes, which leads to dissemination, septicemia, toxemia, and often death (1, 34). It has been shown that the spores are able to germinate and the bacteria are able to multiply inside macrophages both in cell culture and in the lungs of challenged animals (7, 11, 28, 29).Independent of the route of infection, spore germination inside a susceptible host is essential for disease. The highly stable spore form of the bacterium can remain viable under harsh environmental conditions for many decades (32). However, a spore can form a rapidly dividing vegetative cell upon entry into a host and recognition of specific chemical signals, or germinants, through specialized germinant receptors (32). The spore cortex, a thick layer of modified peptidoglycan (PG), contributes much of the spore''s environmental resistance as it is necessary to maintain dehydration of the spore core (25). This protective barrier is broken down following the activation of germination-specific lytic enzymes (GSLEs), allowing full core rehydration and cell outgrowth (32). Experimentally, germination can also be triggered by nongerminant treatments, such as lysozyme treatment, high pressure, exogenous Ca2+-dipicolinic acid treatment, and treatment with cationic surfactants (32). Several of these treatments likely cause spore cortex hydrolysis, triggering spore germination. This indicates the importance of cortex degradation in the spore germination process.Bacterial cell wall PG consists of polysaccharide chains of repeating N-acetylglucosamine and N-acetylmuramic acid, joined by β(1,4) glycosidic bonds (25). This basic structure is modified in several ways in spore cortex PG. In one major modification, 50% of the muramic acid residues (alternating every other residue) are converted to muramic-δ-lactam residues (25). This modification is essential for the specificity of GSLEs for degrading the cortex and prevents degradation of the bacterial cell wall during cortex hydrolysis (21).Previous work on the role of GSLEs in Bacillus subtilis and, recently, in B. anthracis has shown that the enzymes SleB and CwlJ have partially redundant roles and are necessary together for full cortex hydrolysis and spore germination (6, 14). SleB is a lytic transglycosylase that, when activated by an unknown mechanism, hydrolyzes the bond between N-acetylmuramic acid and N-acetylglucosamine (5). In both B. subtilis and B. anthracis, the sleB gene is found in a bicistronic operon with ypeB. Although the function of YpeB is not known, deletion of ypeB prevents SleB activity in spore germination, and sleB and ypeB mutants have similar phenotypes (5). Expression of both gene products is necessary for the presence of SleB in the cortex and inner membrane of mature spores (2, 5).Although no specific enzymatic activity has been attributed to CwlJ, it is required for full germination and it shares a homologous catalytic domain with SleB (20). In B. subtilis and Bacillus cereus, cwlJ is found in an operon with gerQ. Similar to the finding that ypeB is necessary for a functional SleB protein, gerQ is required for CwlJ activity (26). The B. anthracis genome contains two homologs of cwlJ (designated cwlJ1 and cwlJ2 [14]), whereas a single copy is present in B. subtilis and B. cereus. As it is in the related species, cwlJ1 is found in an operon with gerQ, but cwlJ2 is in a different locus and is not in an operon with a gerQ homolog (14). It has been shown that CwlJ is localized to the spore coat and that it is necessary for spore germination with exogenous Ca2+-dipicolinic acid treatment (3, 24).GSLE activation represents a critical step in the complex process of germination. The relatively small number of genes involved and the apparent essential nature of their activity make them attractive targets for new therapeutics, as well as environmental decontamination compounds. The objective of this study was to test by using genetic analysis the role of the GSLE genes sleB, cwlJ1, and cwlJ2 in B. anthracis spore germination. Mutants lacking these three genes were tested to determine their effects on in vitro germination kinetics and colony-forming efficiency. Additionally, the virulence of these mutant strains was examined by comparing mutant and wild-type spores in an in vivo mouse model of inhalational anthrax.  相似文献   

16.
17.
Recent studies have identified Clostridium difficile in food animals and retail meat, and concern has been raised about the potential for food to act as a source of C. difficile infection in humans. Previous studies of retail meat have relied on enrichment culture alone, thereby preventing any assessment of the level of contamination in meat. This study evaluated the prevalence of C. difficile contamination of retail ground beef and ground pork in Canada. Ground beef and ground pork were purchased from retail outlets in four Canadian provinces. Quantitative and enrichment culture was performed. Clostridium difficile was isolated from 28/230 (12%) samples overall: 14/115 (12%) ground beef samples and 14/115 (12%) ground pork samples (P = 1.0). For ground beef, 10/14 samples (71%) were positive by enrichment culture only. Of the 4 ground beef samples that were positive by direct culture, 20 spores/g were present in 2 while 120 and 240 spores/g were present in 1 each. For ground pork, 10/14 (71%) samples were positive by enrichment culture only. Of the 4 ground pork samples that were positive by direct culture, 20 spores/g were present in 3 while 60 spores/g were present in 1. Ribotype 078 predominated, consistent with some previous studies of C. difficile in food animals. Ribotype 027/North American pulsotype 1 was also identified in both retail beef and pork. This study has identified relatively common contamination of retail ground beef and pork with C. difficile spores; however, the levels of contamination were very low.Clostridium difficile is an important cause of enteric disease in humans. It is the most commonly diagnosed cause of hospital- and antimicrobial agent-associated diarrhea in people, and recent evidence suggests that it may be emerging as an important community-associated pathogen (2, 5). In addition to humans, C. difficile can be found in the intestinal tracts of a variety of animal species, including food animals, such as cattle and pigs (7, 10, 13). Clostridium difficile has also been found in retail meat (11, 12, 17), and concerns about the role of food in the epidemiology of community-associated C. difficile infection (CA-CDI) have been expressed (5, 8, 15).Initial studies have reported isolation of C. difficile from 4.6 to 45% of retail meat samples (11, 12, 17). However, all studies have used broth enrichment protocols, which could detect very low spore numbers and provide no information about the number of organisms present in a sample. No studies have evaluated numbers of C. difficile spores in food. While the infectious dose is not known, an understanding of the level of contamination may be an important factor in determining the relevance of contamination of food. Additionally, the use of different methods between studies hampers comparison of results. Recently a study was performed to evaluate different methods for qualitative and quantitative detection of C. difficile (21). This study determined that the detection threshold of enrichment culture could be at least as low as 10 spores/g of meat. It also determined that quantitative culture can accurately determine the level of contamination in experimentally inoculated meat samples, albeit with a higher detection threshold. The objective of this study was to determine the prevalence of C. difficile contamination of retail ground beef and ground pork using both qualitative and quantitative methods.  相似文献   

18.
Dual-trap laser tweezers Raman spectroscopy (LTRS) and elastic light scattering (ELS) were used to investigate dynamic processes during high-temperature treatment of individual spores of Bacillus cereus, Bacillus megaterium, and Bacillus subtilis in water. Major conclusions from these studies included the following. (i) After spores of all three species were added to water at 80 to 90°C, the level of the 1:1 complex of Ca2+ and dipicolinic acid (CaDPA; ∼25% of the dry weight of the spore core) in individual spores remained relatively constant during a highly variable lag time (Tlag), and then CaDPA was released within 1 to 2 min. (ii) The Tlag values prior to rapid CaDPA release and thus the times for wet-heat killing of individual spores of all three species were very heterogeneous. (iii) The heterogeneity in kinetics of wet-heat killing of individual spores was not due to differences in the microscopic physical environments during heat treatment. (iv) During the wet-heat treatment of spores of all three species, spore protein denaturation largely but not completely accompanied rapid CaDPA release, as some changes in protein structure preceded rapid CaDPA release. (v) Changes in the ELS from individual spores of all three species were strongly correlated with the release of CaDPA. The ELS intensities of B. cereus and B. megaterium spores decreased gradually and reached minima at T1 when ∼80% of spore CaDPA was released, then increased rapidly until T2 when full CaDPA release was complete, and then remained nearly constant. The ELS intensity of B. subtilis spores showed similar features, although the intensity changed minimally, if at all, prior to T1. (vi) Carotenoids in B. megaterium spores'' inner membranes exhibited two changes during heat treatment. First, the carotenoid''s two Raman bands at 1,155 and 1,516 cm−1 decreased rapidly to a low value and to zero, respectively, well before Tlag, and then the residual 1,155-cm−1 band disappeared, in parallel with the rapid CaDPA release beginning at Tlag.Bacterial spores of Bacillus species are formed in sporulation and are metabolically dormant and extremely resistant to a variety of harsh conditions, including heat, radiation, and many toxic chemicals (37). Since spores of these species are generally present in foodstuffs and cause food spoilage and food-borne disease (37, 38), there has long been interest in the mechanisms of both spore resistance and spore killing, especially for wet heat, the agent most commonly used to kill spores. The killing of dormant spores by wet heat generally requires temperatures about 40°C higher than those for the killing of growing cells of the same strain (37, 43). A number of factors influence spore wet-heat resistance, with a major factor being the spore core''s water content, as spores with higher core water content are less wet-heat resistant than are spores with lower core water (15, 25). The high level of pyridine-2,6-dicarboxylic acid (dipicolinic acid [DPA]) and the types of its associated divalent cations, predominantly Ca2+, that comprise ∼25% of the dry weight of the core also contribute to spore wet-heat resistance, although how low core water content and CaDPA protect spores against wet heat is not known. The protection of spore DNA against depurination by its saturation with a group of α/β-type small, acid-soluble spore proteins also contributes to spore wet-heat resistance (14, 23, 33, 37).Despite knowledge of a number of factors important in spore wet-heat resistance, the mechanism for wet-heat killing of spores is not known. Wet heat does not kill spores by DNA damage or oxidative damage (35, 37). Instead, spore killing by this agent is associated with protein denaturation and enzyme inactivation (2, 7, 44), although specific proteins for which damage causes spore death have not been identified. Wet-heat treatment also often results in the release of the spore core''s large depot of CaDPA. The mechanism for this CaDPA release is not known but is presumably associated with the rupture of the spore''s inner membrane (7). In addition, the relationship between protein denaturation and CaDPA release is not clear, although recent work suggests that significant protein denaturation can occur prior to CaDPA release (7). Almost all information on spore killing by moist heat has been obtained with spore populations, and essentially nothing is known about the behavior of individual spores exposed to potentially lethal temperatures in water. Given the likely heterogeneity of spores in populations, in particular in their wet-heat resistances (16, 18, 39, 40), it could be most informative to analyze the behavior of individual spores exposed to high temperatures in water.Raman spectroscopy is widely used in biochemical studies, as this technique has high sensitivity and responds rapidly to subtle changes in molecule structure (1, 22, 31). In addition, when Raman spectroscopy is combined with confocal microscopy and optical tweezers, the resultant laser tweezers Raman spectroscopy (LTRS) allows the nondestructive, noninvasive detection of biochemical processes at the single-cell level (9, 10, 19, 46). Indeed, LTRS has been used to analyze the DPA level and the germination of individual Bacillus spores (5, 19, 30). In order to obtain information more rapidly, dual- and multitrap laser tweezers have been developed to allow multiple individual cells or particles to be analyzed simultaneously (11, 13, 24, 27), and the dual trap has been used to measure the hydrodynamic cross-correlations of two particles (24). In addition to Raman scattering, the elastic light scattering (ELS) from trapped individual cells also provides valuable information on cell shape, orientation, refractive index, and morphology (12, 45) and has been used to monitor spore germination dynamics as well (30).In this work, we report studies of wet-heat treatment of individual spores of three different Bacillus species by dual-trap LTRS and ELS. A number of important processes related to wet-heat inactivation of spores, including CaDPA release and protein denaturation, and the correlation between these processes were investigated by monitoring changes in Raman scattering at CaDPA-, protein structure-, and phenylalanine-specific bands and changes in ELS intensity.  相似文献   

19.
Some cholate derivatives that are normal components of bile can act with glycine to induce the germination of Clostridium difficile spores, but at least one bile component, chenodeoxycholate, does not induce germination. Here we show that chenodeoxycholate inhibits the germination of C. difficile spores in response to cholate and taurocholate.The anaerobic human pathogen Clostridium difficile must be in the spore form to survive for extended periods of time outside the colonic environment (6). Spores are also the form of the organism most likely to be ingested by a host. To cause disease, however, C. difficile spores must germinate in the gastrointestinal tract and reach the anaerobic environment of the colon, where they can grow out as vegetative bacteria (2). The vegetative form produces two toxins that damage the colonic epithelium and lead to C. difficile-associated diseases, such as diarrhea, pseudomembranous colitis, and toxic megacolon (4, 15). Extending the work of Wilson and colleagues (17, 18), we have shown that certain bile salts and glycine act as cogerminants for C. difficile spores (13). Primary bile salts produced by the liver are composed mainly of cholate (CA) and chenodeoxycholate (CDCA) derivatives conjugated with either taurine or glycine (11). Since CA derivatives are found in the relatively aerobic proximal ileum (9), we reasoned that C. difficile might benefit if its germination were inhibited until the spores reached the anaerobic environment of the large intestine.Inhibitors of germination are typically structurally similar to the germinant whose activities they inhibit. For example, l-alanine-mediated germination of Bacillus subtilis spores is inhibited by d-alanine (16) and 6-thioguanosine inhibits inosine-mediated germination in Bacillus anthracis (1, 16). Since CA and CDCA are structurally similar but CA induces the germination of C. difficile spores (13) and CDCA does not, we tested whether CDCA could act as an inhibitor of germination. C. difficile strain CD196 (10) spores were produced and their concentration determined as described previously (13). After the vegetative bacteria were killed by incubation at 60°C for 20 min, spores were incubated in water containing various concentrations and combinations of bile salts for 10 min. Here we took advantage of the finding by Wilson et al. that C. difficile spores germinate very inefficiently on rich medium plates lacking bile salts (18) unless they are preincubated with bile salts (13, 17). After incubation, spores were serially diluted and plated on brain heart infusion agar supplemented with 5 g yeast extract per liter-0.1% l-cysteine (BHIS) (Difco) in the absence of any bile salt (BHIS contains enough glycine to act as a cogerminant). After overnight growth at 37°C, colonies were enumerated. As a positive or negative control, spores were plated on BHIS containing 0.1% taurocholate (TA) [BHIS(TA)] or on BHIS agar alone, respectively. Preincubation of spores with 0.1% TA in water resulted in the recovery of approximately 0.5% of the total number of spores as colonies compared to results for spores plated directly on BHIS(TA). These results are similar to our previous findings that spores germinate and grow out as colonies more efficiently on agar medium containing TA (13). As reported previously, 0.1% CDCA poorly stimulated colony formation by C. difficile spores (13), yielding only 0.006% spore recovery (Fig. (Fig.1A).1A). When TA and CDCA were combined, both at 0.1%, colony formation by C. difficile spores was reduced 21-fold to 0.024% compared to the effect of TA alone. This result indicates that CDCA blocks TA-stimulated colony formation and suggests that CDCA may be an inhibitor of C. difficile spore germination. Increasing the ratio of TA to CDCA suppressed the inhibitory effect of CDCA, increasing colony formation by spores (Fig. (Fig.1A).1A). Thus, CDCA seems to block colony formation by competing with TA.Open in a separate windowFIG. 1.CDCA inhibits colony formation by C. difficile spores in response to TA and CA. (A) Spores were prepared and preincubated with TA or CDCA or both in water for 10 min before serial dilution and plating on BHIS agar in the absence of TA. Spores plated on BHIS(TA) served as a positive control for 100% colony formation (CFU). Based on comparisons of total spore counts obtained by microscopy and by colony formation on BHIS(TA) plates, the efficiency of colony formation on BHIS(TA) was estimated at 83%. (B) Spores were prepared as described for panel A and exposed to CA or CDCA or both. Values shown are the averages for three independent experiments, and error bars represent one standard deviation from the mean.CA and other cholate derivatives (e.g., TA, glycocholate, and deoxycholate [DCA]) are also germinants for C. difficile spores (13, 17). To test if CDCA prevents colony formation induced by CA, spores were preexposed to 0.1% CA with and without CDCA. Exposure to CA alone resulted in approximately 1% spore recovery, whereas exposure to 0.1% CA and 0.1% CDCA together led to a decrease in colony formation to 0.075% (Fig. (Fig.1B).1B). The effect of CDCA on CA-mediated colony formation was relieved by increasing the concentration of CA to 1.0%, raising colony formation to 2.6% (Fig. (Fig.1B).1B). These results indicate that CDCA blocks colony formation induced by CA, as well as that induced by TA, and may be an inhibitor of germination by C. difficile spores that acts competitively in both cases.Spore germination per se is classically measured as a decrease in the optical density of a spore suspension occurring concomitantly with a release of Ca2+-dipicolinate from the spore core, rehydration of the core, and degradation of the cortex (8, 12). As determined by this assay, TA is the most effective bile salt for inducing rapid germination (13). To test if CDCA is an inhibitor of germination as opposed to an inhibitor of some other step between germination and colony formation, spores were purified as described previously (13). Spores did not germinate in BHIS medium alone or when this medium was supplemented with 0.1% CDCA (Fig. (Fig.2).2). When C. difficile spores were suspended in BHIS containing 0.1% TA, the optical density of the suspension rapidly decreased, indicating that the spores were germinating. However, the optical density of the spores suspended in BHIS with 0.1% TA plus 0.1% CDCA did not decrease over time, indicating that CDCA inhibited TA-mediated germination (Fig. (Fig.2).2). When the concentration of TA was increased from 0.1% to 1.0% in the presence of 0.1% CDCA, spores were able to germinate (Fig. (Fig.2).2). After overnight incubation in BHIS with 0.1% TA plus 0.1% CDCA, 84% of the spores remained phase bright, while only 11% of spores remained phase bright in BHIS with 1.0% TA plus 0.1% CDCA, indicating that CDCA blocks germination at a very early step. Thus, CDCA is an inhibitor of germination by C. difficile spores that functions by competing with TA and possibly with CA.Open in a separate windowFIG. 2.CDCA inhibits germination of Clostridium difficile spores. Spores were prepared as described previously (13). C. difficile spores were suspended in BHIS alone (•), BHIS plus 0.1% CDCA (▾), BHIS plus 0.1% TA (⧫), BHIS plus 0.1% TA-0.1% CDCA (▪), or BHIS plus 1.0% TA-0.1% CDCA (▴). The ratio of the OD600 at the various time points to the OD600 at time zero is plotted versus time. Data points are the averages of three independent experiments, and error bars represent one standard deviation from the mean.We previously suggested a role for bile salts in determining the ability of C. difficile to colonize and cause disease (13). In this model, germination of C. difficile spores depends on interaction with glycine and certain bile salts. We show here that the primary bile salt CDCA inhibits germination of C. difficile spores. As mentioned above, germination inhibitors are commonly structurally related to the germinant they inhibit. The structures of CA derivatives and CDCA derivatives are very similar; they differ only insofar as CDCA lacks the 12α hydroxyl group (11).CDCA and CA derivatives are present in approximately equal concentrations in the cecum (5). Under such conditions, CDCA would compete with CA derivatives for binding to putative germinant receptors on C. difficile spores. Mekhjian and colleagues measured the colonic absorption rates of CDCA, CA, and DCA that were introduced into the cecum and collected at the distal colon (7). They found that CDCA was absorbed by the colon at 10 times the rate for CA (7). Thus, when spores reach the distal large intestine, they encounter a decreased ratio of CDCA to CA. Such a change in ratio might allow CA derivatives to act as effective germinants. Thus, C. difficile spores would not be expected to germinate until they reach the colon, which also provides the anaerobic environment required for C. difficile growth.The colonic microflora, which is known to protect the host against C. difficile infection, plays a significant role in the metabolism of bile salts (3, 11). Many different species express on their cell surfaces bile salt hydrolases that serve to remove the conjugated tauryl or glycyl groups from primary bile salts (11). After deconjugation, CA and CDCA are further metabolized by a small percentage of the bacterial species in the cecum to the secondary bile salts deoxycholate and lithocholate, respectively (11, 14). Deoxycholate is an inhibitor of C. difficile growth (13, 17). CDCA inhibits both germination and growth (13). The use of CDCA either as prophylaxis or as a therapy for C. difficile-associated disease might be helpful for patients who are undergoing antibiotic regimens or who are colonized by this bacterium. For example, when an antibiotic that is known to be associated with an increased risk of inciting C. difficile-associated disease is administered, the coadministration of CDCA might protect that individual from colonization by C. difficile through inhibiting spore germination. Alternatively, administering CDCA to individuals who are already being given vancomycin or metronidazole for C. difficile-associated disease may have the benefit of preventing spore germination and further vegetative growth (13) after antibiotic therapy is stopped. This strategy may reduce the already significant risk of a relapse.  相似文献   

20.
Molecular-genetic and muropeptide analysis techniques have been applied to examine the function in vivo of the Bacillus megaterium QM B1551 SleB and SleL proteins. In common with Bacillus subtilis and Bacillus anthracis, the presence of anhydromuropeptides in B. megaterium germination exudates, which is indicative of lytic transglycosylase activity, is associated with an intact sleB structural gene. B. megaterium sleB cwlJ double mutant strains complemented with engineered SleB variants in which the predicted N- or C-terminal domain has been deleted (SleB-ΔN or SleB-ΔC) efficiently initiate and hydrolyze the cortex, generating anhydromuropeptides in the process. Additionally, sleB cwlJ strains complemented with SleB-ΔN or SleB-ΔC, in which glutamate and aspartate residues have individually been changed to alanine, all retain the ability to hydrolyze the cortex to various degrees during germination, with concomitant release of anhydromuropeptides to the surrounding medium. These data indicate that while the presence of either the N- or C-terminal domain of B. megaterium SleB is sufficient for initiation of cortex hydrolysis and the generation of anhydromuropeptides, the perceived lytic transglycosylase activity may be derived from an enzyme(s), perhaps exclusively or in addition to SleB, which has yet to be identified. B. megaterium SleL appears to be associated with the epimerase-type activity observed previously in B. subtilis, differing from the glucosaminidase function that is apparent in B. cereus/B. anthracis.Spores of the genera Bacillus and Clostridium emerge from dormancy via the process of germination. The germination process comprises a series of sequential biophysical and biochemical reactions that result irreversibly in the spore losing its properties of metabolic dormancy and extreme resistance to various chemical and physical treatments (24, 34). Germination is initiated by the presumed binding of small molecular germinants, commonly amino acids or sugars, to cognate receptors located within the spore inner membrane (25, 28). In a process that is poorly understood at the molecular level, this interaction leads to a change in the permeability of the inner membrane, resulting in the release of various solutes from the spore core, including metal ions, calcium dipicolinate (Ca-DPA), and some amino acids (32, 33, 35). A degree of rehydration of the core is evident at or around the same time, although this is insufficient to permit a significant degree of vegetative metabolism (9, 31). These events, which appear common to all Bacillus species where examined, comprise stage I of germination (31, 32, 34).The major event in stage II of the germination process from a biochemical perspective involves depolymerization of the spore cortex. The spore cortex is a thick layer of peptidoglycan, characterized by the spore-specific muramic acid lactam (MAL) moiety (37, 38), which, together with the thin inner layer of germ cell wall peptidoglycan (36), forms contiguous layers that entirely envelope the spore protoplast. While the germ cell wall forms the initial cell wall during vegetative outgrowth, the spore cortex serves primarily to maintain the relatively dehydrated status of the spore protoplast during dormancy (13). Dissolution of the cortex permits complete hydration of the spore core and resumption of vegetative metabolism, leading ultimately to shedding of the spore coat and the emergence of a new vegetative cell (34).A number of studies have indicated that spores of various Bacillus species employ two cortex-lytic enzymes (CLEs), SleB and CwlJ, to initiate hydrolysis of the cortex during stage II of the germination process (16, 19, 32). These enzymes are semiredundant; hence, strains bearing null mutations in either structural gene can still degrade the cortex sufficiently to complete germination, whereas double mutant strains do not appear capable of degrading the cortex at all, resulting typically in a decrease of several orders of magnitude in colony-forming ability (15, 19, 32). Other enzymes, including Bacillus cereus/Bacillus anthracis SleL, are also involved in stage II of germination, apparently hydrolyzing peptidoglycan products of SleB and/or CwlJ to smaller peptidoglycan fragments that can more easily permeate through the spore coats to the surrounding germination medium (21).Studies with SleB and SleL purified from dormant and germinating spores indicate that whereas the latter enzyme degrades only cortical fragments of peptidoglycan (7), SleB has a requirement for intact peptidoglycan that has adopted the precise architecture present within the spore (12, 22). These substrate requirements appear to be important in maintenance of the respective autolysins, which are present in the spore in a mature form, in an inactive state during dormancy. Additionally, whereas the molecular mechanism of activation of SleB remains unclear—a change in cortical stress/architecture induced by stage I events has been hypothesized (12)—the efflux of Ca-DPA from the spore core to the cortex/coat boundary where CwlJ is localized (5) appears to be the mechanism by which this CLE is activated. CwlJ can also be activated by high concentrations of exogenous Ca-DPA, presenting an alternative germination pathway that bypasses the germinant receptors (27).The hydrolytic bond specificity of various CLEs has been examined by both direct and indirect biochemical means. Direct assays are typically conducted by incubation of purified or recombinant enzymes with peptidoglycan fragments or suspensions of spores in which the cortex is rendered accessible by first chemically compromising the permeability of the spore coats (7, 12, 22). Subsequent assays for the generation of reducing groups and/or free amino groups can yield information on the probable hydrolytic bond specificity of the respective enzyme(s) being assayed.More recently, the high-performance liquid chromatography/mass spectrometry (HPLC/MS)-based muropeptide analysis technique has been applied to characterize CLE activity during germination of various spore-forming species (2, 4, 10). This methodology has the resolution to reveal fine structural changes that occur to the peptidoglycan in vivo during germination, and when used in combination with CLE null mutant strains, it can be used to indirectly correlate the generation of certain classes of muropeptides, and therefore the hydrolytic bond specificity, with defined CLEs. Muropeptide analysis has revealed, for example, that an intact copy of the sleB gene in B. subtilis and B. anthracis is required for the presence of anhydromuropeptides in the germination exudates of these respective species, indicating that SleB is a lytic transglycosylase or generates substrate for subsequent lytic transglycosylase activity (6, 16). Conversely, B. cereus SleB was characterized as a probable amidase after enzyme purified from germinating spores was found to liberate a large amount of free amino groups when incubated with coat-stripped spores as a substrate (22). The hydrolytic bond specificity of SleB therefore remains ambiguous and perhaps varies between different species.Contrary to these observations, the overall structural architecture of SleB appears to be well conserved between different Bacillus species. Alignment of the primary amino acid sequence from different species indicates that the mature protein comprises an N-terminal domain that is connected to the C-terminal domain by a linker region that is variable in length and amino acid composition (Fig. (Fig.1).1). The N-terminal domain is thought to comprise the peptidoglycan binding domain by virtue of two direct sequence repeats that are reminiscent of cell wall-binding motifs observed in other proteins (26). The C-terminal domain shows homology with that of the other major Bacillus CLE, CwlJ, which lacks a corresponding peptidoglycan binding domain and is therefore thought to comprise the catalytic domain (19), although there is as yet no experimental evidence to substantiate this idea.Open in a separate windowFIG. 1.ClustalW alignment of SleB from various Bacillus species. Residues predicted to comprise putative structural domains are denoted. Stars indicate charged residues that were subjected to amino acid substitution in this work. BM, B. megaterium QM B1551; BC, B. cereus W; BCl, B. clausii KSM-K16; BS, B. subtilis 168.In the current study, we have investigated the molecular function of SleB during germination of Bacillus megaterium QM B1551 spores, employing engineered SleB N- and C-terminal deletion strains, site-directed mutagenesis (SDM), and muropeptide analyses. In addition to revealing several cortex-modifying activities during germination of this species, the presented data indicate that while the presence of either the N- or C-terminal domain of SleB is sufficient for the generation of anhydromuropeptides during germination, this may be an indirect effect, and at least a degree of lytic transglycosylase activity may result from the activity of another as yet unidentified enzyme.  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