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 140 毫秒
1.
Clostridium sporogenes PA 3679 is a nonpathogenic, nontoxic model organism for proteolytic Clostridium botulinum used in the validation of conventional thermal food processes due to its ability to produce highly heat-resistant endospores. Because of its public safety importance, the uncertain taxonomic classification and genetic diversity of PA 3679 are concerns. Therefore, isolates of C. sporogenes PA 3679 were obtained from various sources and characterized using pulsed-field gel electrophoresis (PFGE) and whole-genome sequencing. The phylogenetic relatedness and genetic variability were assessed based on 16S rRNA gene sequencing and whole-genome single nucleotide polymorphism (SNP) analysis. All C. sporogenes PA 3679 isolates were categorized into two clades (clade I containing ATCC 7955 NCA3679 isolates 1961-2, 1990, and 2007 and clade II containing PA 3679 isolates NFL, UW, FDA, and Campbell and ATCC 7955 NCA3679 isolate 1961-4). The 16S maximum likelihood (ML) tree clustered both clades within proteolytic C. botulinum strains, with clade I forming a distinct cluster with other C. sporogenes non-PA 3679 strains. SNP analysis revealed that clade I isolates were more similar to the genomic reference PA 3679 (NCTC8594) genome (GenBank accession number AGAH00000000.1) than clade II isolates were. The genomic reference C. sporogenes PA 3679 (NCTC8594) genome and clade I C. sporogenes isolates were genetically distinct from those obtained from other sources (University of Wisconsin, National Food Laboratory, U.S. Food and Drug Administration, and Campbell''s Soup Company). Thermal destruction studies revealed that clade I isolates were more sensitive to high temperature than clade II isolates were. Considering the widespread use of C. sporogenes PA 3679 and its genetic information in numerous studies, the accurate identification and genetic characterization of C. sporogenes PA 3679 are of critical importance.  相似文献   

2.
Clostridium botulinum is a dangerous pathogen that forms the highly potent botulinum toxin, which when ingested causes a deadly neuroparalytic disease. The closely related Clostridium sporogenes is occasionally pathogenic, frequently associated with food spoilage and regarded as the non-toxigenic equivalent of Group I C. botulinum. Both species form highly resistant spores that are ubiquitous in the environment and which, under favourable growth conditions germinate to produce vegetative cells. To improve the control of botulinum neurotoxin-forming clostridia, it is imperative to comprehend the mechanisms by which spores germinate. Germination is initiated following the recognition of small molecules (germinants) by a specific germinant receptor (GR) located in the spore inner membrane. The present study precisely defines clostridial GRs, germinants and co-germinants. Group I C. botulinum ATCC3502 contains two tricistronic and one pentacistronic GR operons, while C. sporogenes ATCC15579 has three tricistronic and one tetracistronic GR operons. Insertional knockout mutants, allied with characterisation of recombinant GRs shows for the first time that amino acid stimulated germination in C. botulinum requires two tri-cistronic encoded GRs which act in synergy and cannot function individually. Spore germination in C. sporogenes requires one tri-cistronic GR. Two other GRs form part of a complex involved in controlling the rate of amino-acid stimulated germination. The suitability of using C. sporogenes as a substitute for C. botulinum in germination studies and food challenge tests is discussed.  相似文献   

3.
《Anaerobe》2002,8(3):89-100
Germination is the first step in the development of dormant spores into exponentially dividing cells. Spore germination in proteolytic Clostridium botulinum type B and in C. sporogenes was triggered strongly byL -alanine/L -lactate/NaHCO3, less well by L-alanine/NaHCO3, but not by inosine or the AGFK mixture (asparagine, glucose, fructose and potassium ions). This suggests the presence of a functional equivalent of the L-alanine receptor encoded by the gerA operon in Bacillus subtilis, while germination genes that respond to inosine (gerI in Bacillus cereus) and AGFK mixture (gerB/gerK in B. subtilis) appear to be absent. Putative gerA operons of proteolytic C. botulinum type B and C. sporogenes have been identified using PCR with degenerate primers and show a similar genetic arrangement with the gerAA and gerAB genes, encoding the GerAA and GerAB proteins. In both species, a third partial ORF may encode a GerAC homologue. The GerAA and GerAB proteins have homology with GerAA and GerAB proteins from other Clostridium and Bacillus species with the highest homology between the C. botulinum and C. sporogenes proteins. The GerAA proteins of these species have six putative transmembrane α -helices similar to other L -alanine-dependent germination GerAA proteins suggesting a membrane location. Expression studies in C. botulinum showed that the gerAA and gerAB genes are co-transcribed during sporulation.  相似文献   

4.
The inactivation of spores of four low-acid food spoilage organisms by high pressure thermal (HPT) and thermal-only processing was compared on the basis of equivalent thermal lethality calculated at a reference temperature of 121.1°C (F(z)(121.1)(°)(C, 0.1 MPa or 600 MPa)) and characterized as synergistic, not different or protective. In addition, the relative resistances of spores of the different spoilage microorganisms to HPT processing were compared. Processing was performed and inactivation was compared in both laboratory and pilot scale systems and in model (diluted) and actual food products. Where statistical comparisons could be made, at least 4 times and up to around 190 times more inactivation (log(10) reduction/minute at F(T)(z)(121.1)(°)(C)) of spores of Bacillus amyloliquefaciens, Bacillus sporothermodurans, and Geobacillus stearothermophilus was achieved using HPT, indicating a strong synergistic effect of high pressure and heat. Bacillus coagulans spores were also synergistically inactivated in diluted and undiluted Bolognese sauce but were protected by pressure against thermal inactivation in undiluted cream sauce. Irrespective of the response characterization, B. coagulans and B. sporothermodurans were identified as the most HPT-resistant isolates in the pilot scale and laboratory scale studies, respectively, and G. stearothermophilus as the least in both studies and all products. This is the first study to comprehensively quantitatively characterize the responses of a range of spores of spoilage microorganisms as synergistic (or otherwise) using an integrated thermal-lethality approach (F(T)(z)). The use of the F(T)(z) approach is ultimately important for the translation of commercial minimum microbiologically safe and stable thermal processes to HPT processes.  相似文献   

5.
Uracil was used by growing cultures of Clostridium sporogenes, and by proteolytic strains of C. botulinum types A and B. Uracil was not used by C. bifermentans; C. botulinum, type B (non-proteolytic); C. botulinum, type F (non-proteolytic); C. botulinum, type E; C. butyricum; C. cochlearium; C. difficile; C. histolyticum; C. oedematiens, type A; C. paraputrificum; C. scatologenes; C. septicum; C. sordellii; C. sticklandii; C. tertium; C. tetani; C. tetanomorphum; C. welchii, types A, B, C, E and 4 untyped strains. The growth of C. sporogenes was not increased by uracil; it was reduced to dihydrouracil. Experiments with washed cells of C. sporogenes showed that the uracil-reducing system was inducible. Washed cell suspensions incubated under hydrogen with uracil, thymine, iso-barbituric acid, 5-amino uracil and cytosine consumed 1 mole H2/mole pyrimidine. The reduction product of cytosine was dihydrouracil indicating that it was deaminated before reduction. The reduction products of the remaining pyrimidines were the corresponding dihydro derivatives. Extracts of C. sporogenes reduced uracil in the presence of NADPH2 but not NADH2.  相似文献   

6.
Production of Hydrocinnamic Acid by Clostridia   总被引:7,自引:2,他引:5       下载免费PDF全文
Hydrocinnamic acid was found in acid extracts of spent growth medium from cultures of Clostridium sporogenes. The acid was identified by mass spectrometry and its identity was confirmed by gas chromatography. The acid was produced in relatively large amounts (2 to 3 μmoles/ml of medium) by C. sporogenes, toxigenic types A, B, D, and F of C. botulinum, and some strains of C. bifermentans. Other strains of C. bifermentans and strains of C. sordellii and C. caproicum produced only small amounts (0.1 to 0.4 μmoles/ml) of the acid. The acid was not detected in spent medium from toxigenic types C and E of C. botulinum or from 25 other strains representing eight Clostridium species. Resting cell suspensions exposed to l-phenylalanine produced hydrocinnamic and cinnamic acid; the latter compound probably functions as an intermediate in the metabolism of l-phenylalanine.  相似文献   

7.
Thirty-one soil samples were examined for the presence of organisms capable of inhibiting growth and toxin production of strains of Clostridium botulinum type A. Such organisms were found in eight samples of soil. Inhibiting strains of C. perfringens were found in five samples, of C. sporogenes in three and of Bacillus cereus in three. Three of the C. perfringens strains produced an inhibitor effective on all 11 strains of C. botulinum type A against which they were tested, seven of eight proteolytic type B strains, one nonproteolytic type B strain, five of nine type E strains and all seven type F strains, whether proteolytic or nonproteolytic. They did not inhibit any of 26 type C strains, 6 type D strains, 4 type E strains, or 24 C. sporogenes strains. In mixed culture, an inhibitor strain of C. perfringens repressed growth and toxin production by a C. botulinum type A strain even though it was outnumbered by the latter about 40 times. It also repressed growth and toxin production of C. botulinum in mixed culture of soils in which this latter organism naturally occurred when cooked meat medium but not when trypticase medium was used.  相似文献   

8.

Background

Proteolytic Clostridium botulinum is the causative agent of botulism, a severe neuroparalytic illness. Given the severity of botulism, surprisingly little is known of the population structure, biology, phylogeny or evolution of C. botulinum. The recent determination of the genome sequence of C. botulinum has allowed comparative genomic indexing using a DNA microarray.

Results

Whole genome microarray analysis revealed that 63% of the coding sequences (CDSs) present in reference strain ATCC 3502 were common to all 61 widely-representative strains of proteolytic C. botulinum and the closely related C. sporogenes tested. This indicates a relatively stable genome. There was, however, evidence for recombination and genetic exchange, in particular within the neurotoxin gene and cluster (including transfer of neurotoxin genes to C. sporogenes), and the flagellar glycosylation island (FGI). These two loci appear to have evolved independently from each other, and from the remainder of the genetic complement. A number of strains were atypical; for example, while 10 out of 14 strains that formed type A1 toxin gave almost identical profiles in whole genome, neurotoxin cluster and FGI analyses, the other four strains showed divergent properties. Furthermore, a new neurotoxin sub-type (A5) has been discovered in strains from heroin-associated wound botulism cases. For the first time, differences in glycosylation profiles of the flagella could be linked to differences in the gene content of the FGI.

Conclusion

Proteolytic C. botulinum has a stable genome backbone containing specific regions of genetic heterogeneity. These include the neurotoxin gene cluster and the FGI, each having evolved independently of each other and the remainder of the genetic complement. Analysis of these genetic components provides a high degree of discrimination of strains of proteolytic C. botulinum, and is suitable for clinical and forensic investigations of botulism outbreaks.  相似文献   

9.
Botulism is diagnosed by detecting botulinum neurotoxin and Clostridium botulinum cells in the patient and in suspected food samples. In this study, a multiplex PCR assay for the detection of Clostridium botulinum types A, B, E, and F in food and fecal material was developed. The method employs four new primer pairs with equal melting temperatures, each being specific to botulinum neurotoxin gene type A, B, E, or F, and enables a simultaneous detection of the four serotypes. A total of 43 C. botulinum strains and 18 strains of other bacterial species were tested. DNA amplification fragments of 782 bp for C. botulinum type A alone, 205 bp for type B alone, 389 bp for type E alone, and 543 bp for type F alone were obtained. Other bacterial species, including C. sporogenes and the nontoxigenic nonproteolytic C. botulinum-like organisms, did not yield a PCR product. Sensitivity of the PCR for types A, E, and F was 102 cells and for type B was 10 cells per reaction mixture. With a two-step enrichment, the detection limit in food and fecal samples varied from 10−2 spore/g for types A, B, and F to 10−1 spore/g of sample material for type E. Of 72 natural food samples investigated, two were shown to contain C. botulinum type A, two contained type B, and one contained type E. The assay is sensitive and specific and provides a marked improvement in the PCR diagnostics of C. botulinum.  相似文献   

10.
Through elaboration of its botulinum toxins, Clostridium botulinum produces clinical syndromes of infant botulism, wound botulism, and other invasive infections. Using comparative genomic analysis, an orphan nine-gene cluster was identified in C. botulinum and the related foodborne pathogen Clostridium sporogenes that resembled the biosynthetic machinery for streptolysin S, a key virulence factor from group A Streptococcus responsible for its hallmark β-hemolytic phenotype. Genetic complementation, in vitro reconstitution, mass spectral analysis, and plasmid intergrational mutagenesis demonstrate that the streptolysin S-like gene cluster from Clostridium sp. is responsible for the biogenesis of a novel post-translationally modified hemolytic toxin, clostridiolysin S.  相似文献   

11.
Endospores of proteolytic type B Clostridium botulinum TMW 2.357 and Bacillus amyloliquefaciens TMW 2.479 are currently described as the most high-pressure-resistant bacterial spores relevant to food intoxication and spoilage in combined pressure-temperature applications. The effects of combined pressure (0.1 to 1,400 MPa) and temperature (70 to 120°C) treatments were determined for these spores. A process employing isothermal holding times was established to distinguish pressure from temperature effects. An increase in pressure (600 to 1,400 MPa) and an increase in temperature (90 to 110°C) accelerated the inactivation of C. botulinum spores. However, incubation at 100°C, 110°C, or 120°C with ambient pressure resulted in faster spore reduction than treatment with 600 or 800 MPa at the same temperature. This pressure-mediated spore protection was also observed at 120°C and 800, 1,000, or 1,200 MPa with the more heat-tolerant B. amyloliquefaciens TMW 2.479 spores. Inactivation curves for both strains showed a pronounced pressure-dependent tailing, which indicates that a small fraction of the spore populations survives conditions of up to 120°C and 1.4 GPa in isothermal treatments. Because of this tailing and the fact that pressure-temperature combinations stabilizing bacterial endospores vary from strain to strain, food safety must be ensured in case-by-case studies demonstrating inactivation or nongrowth of C. botulinum with realistic contamination rates in the respective pressurized food and equipment.  相似文献   

12.
Cryptosporidium parvum, which is resistant to chlorine concentrations typically used in water treatment, is recognized as a significant waterborne pathogen. Recent studies have demonstrated that chlorine dioxide is a more efficient disinfectant than free chlorine against Cryptosporidium oocysts. It is not known, however, if oocysts from different suppliers are equally sensitive to chlorine dioxide. This study used both a most-probable-number–cell culture infectivity assay and in vitro excystation to evaluate chlorine dioxide inactivation kinetics in laboratory water at pH 8 and 21°C. The two viability methods produced significantly different results (P < 0.05). Products of disinfectant concentration and contact time (Ct values) of 1,000 mg · min/liter were needed to inactivate approximately 0.5 log10 and 2.0 log10 units (99% inactivation) of C. parvum as measured by in vitro excystation and cell infectivity, respectively, suggesting that excystation is not an adequate viability assay. Purified oocysts originating from three different suppliers were evaluated and showed marked differences with respect to their resistance to inactivation when using chlorine dioxide. Ct values of 75, 550, and 1,000 mg · min/liter were required to achieve approximately 2.0 log10 units of inactivation with oocysts from different sources. Finally, the study compared the relationship between easily measured indicators, including Bacillus subtilis (aerobic) spores and Clostridium sporogenes (anaerobic) spores, and C. parvum oocysts. The bacterial spores were found to be more sensitive to chlorine dioxide than C. parvum oocysts and therefore could not be used as direct indicators of C. parvum inactivation for this disinfectant. In conclusion, it is suggested that future studies address issues such as oocyst purification protocols and the genetic diversity of C. parvum, since these factors might affect oocyst disinfection sensitivity.  相似文献   

13.
Sustainable management of toilet waste must prevent disease transmission but allow reuse of plant nutrients. Inactivation of uterus-derived Ascaris suum eggs was studied in relation to ammonia in source-separated urine without additives and in human feces to which urea had been added, in order to evaluate ammonia-based sanitation for production of safe fertilizers from human excreta. Urine was used concentrated or diluted 1:1 and 1:3 with tap water at 4, 14, 24, and 34°C. Fecal material, with and without ash, was treated with 1% or 2% (wt/wt) urea at 24 and 34°C. At 34°C eggs were inactivated in less than 10 days in urine and in amended feces. At 24°C only feces with 2% (wt/wt) urea or 1% (wt/wt) urea at high pH (10) inactivated all eggs within 1 month, and no inactivation was observed after 75 days in urine diluted 1:3 (18 ± 11 mM NH3). At temperatures of ≥24°C, NH3 proved to be an efficient sanitizing agent in urine and feces at concentrations of ≥60 mM. Treating fecal material at 34°C can give a 6-log10 egg inactivation within 1 month, whereas at 24°C 6 months of treatment is necessary for the same level of egg inactivation. At temperatures of 14°C and below, inactivation rates were low, with viable eggs after 6 months even in concentrated urine.  相似文献   

14.
Some Observations on Bacterial Thermal Death Time Curves   总被引:2,自引:0,他引:2       下载免费PDF全文
Thermal death rate data were obtained for spores of Clostridium sporogenes P.A. 3679 and Bacillus subtilis var. niger, and for cells of Salmonella senftenberg 775W. The survival curves for P.A. 3679 were approximately linear, but for B. subtilis var. niger or S. senftenberg 775W they were sigmoidal. Decimal reduction times were derived from the regression slopes of the apparent linear portion of the survival curves, and from these a phantom thermal death time (TDT) curve was constructed. In general, the phantom TDT curves were linear for B. subtilis var. niger and S. senftenberg 775W and nonlinear for P.A. 3679.  相似文献   

15.
In order to survive a temperature downshift, bacteria have to sense the changing environment and adjust their metabolism and structure. Two-component signal transduction systems (TCSs) play a central role in sensing and responding to many different environmental stimuli. Although the nonproteolytic (group II) Clostridium botulinum represents a major hazard in chilled foods, the cold adaption mechanisms of group II C. botulinum organisms are not known. Here, we show that the CLO3403/CLO3404 TCS of C. botulinum E1 Beluga is involved in the cold shock response and growth at 12°C. Cold shock induced the expression of the genes encoding the histidine kinase (clo3403) and the response regulator (clo3404) by more than 100-fold after 5 h relative to their expression in a nonshocked culture at the corresponding time point. The involvement of CLO3403/CLO3404 in growth at low temperature was demonstrated by impaired growth of the insertional clo3403 and clo3404 knockout mutants at 12°C compared to the growth of the wild-type culture. Additionally, the inactivation of clo3403 had a negative effect on motility. The growth efficiency at 12°C of the TCS mutants and the motility of the kinase mutants were restored by introducing a plasmid harboring the operon of the CLO3403/CLO3404 TCS. The results suggest that the CLO3403/CLO3404 TCS is important for the cold tolerance of C. botulinum E1 Beluga.  相似文献   

16.
Ten variable-number tandem-repeat (VNTR) regions identified within the complete genomic sequence of Clostridium botulinum strain ATCC 3502 were used to characterize 59 C. botulinum strains of the botulism neurotoxin A1 (BoNT/A1) to BoNT/A4 (BoNT/A1-A4) subtypes to determine their ability to discriminate among the serotype A strains. Two strains representing each of the C. botulinum serotypes B to G, including five bivalent strains, and two strains of the closely related species Clostridium sporogenes were also tested. Amplified fragment length polymorphism analyses revealed the genetic diversity among the serotypes and the high degree of similarity among many of the BoNT/A1 strains. The 10 VNTR markers amplified fragments within all of the serotype A strains but were less successful with strains of other serotypes. The composite multiple-locus VNTR analysis of the 59 BoNT/A1-A4 strains and 3 bivalent B strains identified 38 different genotypes. Thirty genotypes were identified among the 53 BoNT/A1 and BoNT/A1(B) strains, demonstrating discrimination below the subtype level. Contaminating DNA within crude toxin preparations of three BoNT/A subtypes (BoNT/A1 to BoNT/A3) also supported amplification of all of the VNTR regions. These markers provide clinical and forensics laboratories with a rapid, highly discriminatory tool to distinguish among C. botulinum BoNT/A1 strains for investigations of botulism outbreaks.  相似文献   

17.
Chapman et al. (B. Chapman, N. Jensen, T Ross, and M. B. Cole, Appl. Environ. Microbiol. 72:5165-5172, 2006) demonstrated that an increased NaCl concentration prolongs survival of Escherichia coli O157 SERL 2 in a broth model simulating the aqueous phase of a food dressing or sauce containing acetic acid. We examined the responses of five other E. coli strains and four Salmonella enterica strains to increasing concentrations of NaCl under conditions of lethal acidity and observed that the average “lag” time prior to inactivation decreases in the presence of hydrochloric acid but not in the presence of acetic acid. For E. coli in the presence of acetic acid, the lag time increased with increasing NaCl concentrations up to 2 to 4% at pH 4.0, up to 4 to 6% at pH 3.8, and up to 4 to 7% (wt/wt of water) NaCl at pH 3.6. Salmonella was inactivated more rapidly by combined acetic acid and NaCl stresses than E. coli, but increasing NaCl concentrations still decreased the lag time prior to inactivation in the presence of acetic acid; at pH 4.0 up to 1 to 4% NaCl was protective, and at pH 3.8 up to 1 to 2% NaCl delayed the onset of inactivation. Sublethal injury kinetics suggest that this complex response is a balance between the lethal effects of acetic acid, against which NaCl is apparently protective, and the lethal effects of the NaCl itself. Compared against 3% NaCl, 10% (wt/wt of water) sucrose with 0.5% NaCl (which has similar osmotic potential) was found to be equally protective against adverse acetic acid conditions. We propose that hypertonicity may directly affect the rate of diffusion of acetic acid into cells and hence cell survival.We previously observed that inactivation of Escherichia coli O157 SERL 2 by acetic acid at adverse pH in a broth model simulating the aqueous phase of acidic sauces and dressings was reduced by the presence of NaCl (4). Specifically, the time to a 3-log10-unit reduction (t3D) of E. coli SERL 2 as function of NaCl concentration was significantly nonmonotonic; that is, the t3D initially increased when NaCl was increased (from 1 to 3% [wt/wt] of solution), but the t3D decreased upon a further increase in NaCl concentration (to 8% [wt/wt] of solution) (4). The statistical significance of this “nonmonotonic” response increased with increasing exposure time from 24 to 72 h (at 23°C), primarily due to a proportionally greater increase in inactivation at 1% (wt/wt) NaCl with increasing treatment time than that which was observed at higher NaCl concentrations (4).The combination of acid and NaCl is a common example of the food industry''s “hurdle” approach, which is used to preserve a large and diverse range of foods, including acidic dressings and sauces, fermented meats, cheeses, and preserved vegetables. Given the widespread use of this hurdle combination in food manufacturing, the first aim of this study was to determine whether the observed protection of E. coli SERL 2 from acid inactivation by NaCl is common among E. coli and Salmonella enterica and at what NaCl concentration maximum protection is achieved. A second aim was to determine whether NaCl protection is specific against acid pH in general or against acetic acid in particular. Third, possible protection against acid inactivation by another osmolyte, sucrose, was assessed to resolve whether the effect is solute specific.When cells are placed in hypertonic environments, plasmolysis occurs as the cytoplasmic volume reduces due to water loss by osmosis. The thin peptidoglycan layer of gram-negative microorganisms is anchored to the cytoplasmic membrane and can be distended by plasmolysis or even ruptured when plasmolysis is more extreme. Decad and Nikaido (5) observed that the cytoplasmic volume in gram-negative microorganisms was reduced to ∼50% at ∼0.3 M NaCl but that the plasmolysis-induced cell wall damage was minimal. At 0.5 M (2.9%, wt/wt) NaCl, however, they observed cell wall damage in a large fraction of cells. Thus, in the experiments described here we explored the mechanism of the protective effect of NaCl, and specifically cell wall damage in E. coli populations simultaneously exposed to NaCl and either acetic or hydrochloric acid (HCl), by enumeration of both injured and noninjured survivors by culture on media with and without bile salts.  相似文献   

18.
The anaerobic film pouch technique was used to quantitate and isolate clostridial spores in 2,358 samples of raw meat (1,078 of chicken, 624 of beef, 656 of pork). Of 19,727 putrefactive anaerobic (PA) sporeformers isolated, 1 was confirmed by mouse protection testing to be Clostridium botulinum type C. This isolate was obtained from a Western Canada chicken sample which contained 5.33 clostridia per gram. These data indicate a very low incidence of botulinal contamination in raw meats at the packing-plant level (0.042% of 2,358 samples) and an almost 20,000:1 ratio of nonbotulinal PA sporeformers to mesophilic C. botulinum spores. The mean level of PA contamination was 2.8 PA sporeformers per gram of meat; 77% of the samples contained three or less PA sporeformers per gram. Small but statistically significant differences in the incidence of clostridial spores were noted for season, geographical region, and type of meat.  相似文献   

19.
The sporicidal properties of hydrogen peroxide were evaluated at concentrations of 10 to 41% and at temperatures of 24 to 76 C. The organisms tested and their relative resistance at 24 C to 25.8% H2O2 were: Bacillus subtilis SA 22 > B. subtilis var. globigii > B. coagulans > B. stearothermophilus > Clostridium sp. putrefactive anaerobe 3679 > S. aureus, with „D” values of 7.3, 2, 1.8, 1.5, 0.8., and 0.2 min, respectively. Heat shocking spores prior to hydrogen peroxide treatment decreased their resistance. Wet spores were more resistant than dry spores when good mixing was achieved during hydrogen peroxide treatment. Inactivation curves followed first-order kinetics except for a lag period where the inactivation rate was very slow. Increasing the H2O2 concentration and the temperature reduced the lag period.  相似文献   

20.
The effects of pulsed electric field (PEF) treatment and processing factors on the inactivation kinetics of Listeria innocua NCTC 11289 were investigated by using a pilot plant PEF unit with a flow rate of 200 liters/h. The electric field strength, pulse length, number of pulses, and inlet temperature were the most significant process factors influencing the inactivation kinetics. Product factors (pH and conductivity) also influenced the inactivation kinetics. In phosphate buffer at pH 4.0 and 0.5 S/m at 40°C, a 3.0-V/μm PEF treatment at an inlet temperature of 40°C resulted in ≥6.3 log inactivation of strain NCTC 11289 at 49.5°C. A synergistic effect between temperature and PEF inactivation was also observed. The inactivation obtained with PEF was compared to the inactivation obtained with heat. We found that heat inactivation was less effective than PEF inactivation under similar time and temperature conditions. L. innocua cells which were incubated for a prolonged time in the stationary phase were more resistant to the PEF treatment, indicating that the physiological state of the microorganism plays a role in inactivation by PEF. Sublethal injury of cells was observed after PEF treatment, and the injury was more severe when the level of treatment was increased. Overall, our results indicate that it may be possible to use PEF in future applications in order to produce safe products.  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