首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The aim of this project was to establish the minimal inhibitory concentration (MIC) of lactic acid for growth of Clostridium tyrobutyricum. A pH-auxostat was used to maintain a constant pH and to allow continuous growth at the highest possible rates at fixed, but adjustable concentrations of lactate. By raising the concentration of lactic acid and keeping the pH constant, the growth rate was shown to decrease linearly with increasing lactic acid concentration. The p K a of lactic acid, measured in the actual growth medium at 37°C, was 3.40 (±0.03). Based on this value, the MICundiss values for each pH were estimated. The MIC of total lactic acid (MICtot) ranged from 150 mmol l−1 to 1510 mmol l−1 at pH 4.6–6.25, respectively. The corresponding MIC values of undissociated lactic acid (MICundiss) ranged from 8.9 to 2.1 mmol l−1 at the same pH values. These results emphasize the importance of a rapid pH decrease and an equally rapid initial lactic acid fermentation of the ensilage, in order to sufficiently suppress clostridial growth.  相似文献   

2.
3.
Abstract Both lactic and acetic acids cause mixed inhibition of acid production in mutans streptococci. This inhibition is partly irreversible due to cell death, an important factor when considering acidogenicity and aciduricity of these organisms, and their role in the caries process. Other monocarboxylic end-products may be present and may also be important inhibitors of acid production in dental plaque. This study considered the effects of varying concentrations of the end-product formic acid on acid production rates in Streptococcus mutans R9, measured using the pH-stat. Undissociated formic acid caused mixed inhibition with constants of K iu (uncompetitive) of 6.07 ± 1.27 mmol 1−1 and K ic (competitive) of 0.2 ± 10.11 mmol I −1. Inhibition was found to be fully reversible, with no loss of cell viability. It is concluded that at those concentrations found in vivo, formate is not a significant inhibitor of acid production by S. mutans in dental plaque at any time, and is not important in determining the acidogenicity or aciduricity of this organism.  相似文献   

4.
Pathogens found in the environment of abattoirs may become adapted to lactic acid used to decontaminate meat. Such organisms are more acid tolerant than non-adapted parents and can contaminate meat after lactic acid decontamination (LAD). The fate of acid-adapted Yersinia enterocolitica and Listeria monocytogenes, inoculated on skin surface of pork bellies 2 h after LAD, was examined during chilled storage. LAD included dipping in 1%, 2% or 5% lactic acid solutions at 55°C for 120 s. LAD brought about sharp reductions in meat surface pH, but these recovered with time after LAD at ≈1–1·5 pH units below that of water-treated controls. Growth permitting pH at 4·8–5·2 was reached after 1% LAD in less than 0·5 d (pH 4·8–5·0), 2% LAD within 1·5 d (pH 4·9–5·1) and after 5% LAD (pH 5·0–5·2) within 4 d. During the lag on 2% LAD meat Y. enterocolitica counts decreased by 0·9 log10 cfu per cm2 and on 5% LAD the reduction was more than 1·4 log10 cfu per cm2. The reductions in L. monocytogenes were about a third of those in Y. enterocolitica . On 1% LAD the counts of both pathogens did not decrease significantly. The generation times of Y. enterocolitica and L. monocytogenes on 2–5% LAD meats were by up to twofold longer than on water-treated controls and on 1% LAD-treated meat they were similar to those on water-treated controls. Low temperature and acid-adapted L. monocytogenes and Y. enterocolitica that contaminate skin surface after hot 2–5% LAD did not cause an increased health hazard, although the number of Gram-negative spoilage organisms were drastically reduced by hot 2–5% LAD and intrinsic (lactic acid content, pH) conditions were created that may benefit the survival and the growth of acid-adapted organisms.  相似文献   

5.
The efficacy of nisin to control the food-borne pathogen Listeria monocytogenes in ricotta-type cheeses over long storage (70 d) at 6–8°C was determined. Cheeses were prepared from unpasteurized milk by direct acidification with acetic acid (final pH 5·9) and/or calcium chloride addition during heat treatment. Nisin was added in the commercial form of Nisaplin® pre-production to the milk. Each batch of cheese was inoculated with 102–103 cfu g−1 of a five-strain cocktail of L. monocytogenes before storage. Shelf-life analysis demonstrated that incorporation of nisin at a level of 2·5 mg l−1 could effectively inhibit the growth of L. monocytogenes for a period of 8 weeks or more (dependent on cheese type). Cheese made without the addition of nisin contained unsafe levels of the organism within 1–2 weeks of incubation. Measurement of initial and residual nisin indicated a high level of retention over the 10-week incubation period at 6–8°C, with only 10–32% nisin loss.  相似文献   

6.
Sperm structure and motility of the freshwater teleost Cottus gobio   总被引:1,自引:0,他引:1  
When motility of spermatozoa of Cottos gobio was initiated with distilled water, the motility rate decreased to 0% within 1 min, and significant signs of osmotic alterations were observed at the end of the motility period. By contrast, in 50 mmol 1−1 NaCl solution, the motility rate persisted for 120–140 min. In both distilled water and in 50 mmol 1−1 NaCl solution, the main swimming type of spermatozoa was linear motion during the whole motility period. The initial swimming velocity (50.0 ± 2.1 μm s−1) measured 10 s after motility initiation was similar in both distilled water and in 50 mmol 1−1 NaCl solution. In distilled water, the velocity decreased to <20 μm s−1 (locally motile) during the first minute of the motility phase. In 50 mmol 1−1 NaCl solutions, it remained at a constant level during the first 60 min of the motility period, but then started to decrease to <20 μm s−1 after 120 min. When 5 mmol 1−1 potassium cyanide, antimycin or atractyloside was added to the 50 mmol 1−1 NaCl solution, the motility period was reduced to ≤2min. Ten millimoles per litre 2-deoxy-D-glucose, malonate or a mixture of 5 mmol 1−1 atractyloside and 5 mmol 1−1 carnithine did not effect the duration of the motility period. This indicates that sperm energy metabolism depends mainly on respiration rate and fatty acid metabolism.  相似文献   

7.
ADSORPTION OF FULVIC ACID ON ALGAL SURFACES AND ITS EFFECT ON CARBON UPTAKE   总被引:1,自引:0,他引:1  
Adsorption of Suwannee River fulvic acid (SRFA) to algal surfaces of three green algae was studied at environmentally relevant pH values (4 –7) and SRFA concentrations (5–100 mg·L 1). The influence of adsorbed SRFA on carbon uptake of Scenedesmus subspicatus Chodat was also examined. Although no adsorption was observed at neutral pH values (pH 6 and 7), at pH 4 up to 31 mg SRFA·m 2 and at pH 5 up to 4 mg SRFA·m 2 was adsorbed to the algal surfaces. Electrophoretic mobility measurements of S. subspicatus demonstrated an increase in the negative surface charge of the alga in the presence of SRFA at pH 4. The adsorbed SRFA also influenced 14C uptake in S. subspicatus; in this case, enhanced carbon uptake could be related to the amount of adsorbed SRFA. The binding of humic substances by algal surfaces was interpreted as the result of hydrogen bonding and hydrophobic interactions.  相似文献   

8.
It has been shown that Listeria monocytogenes produces acetoin from glucose under aerobic conditions. A defined medium with glucose as the sole carbon source was used in an aerobic shake flask culture to reliably produce acetoin. Acetoin, the reactive compound in the Voges–Proskauer test, was assayable in the medium and was used to quantify the metabolic response when inhibitors were added to the medium. Inhibitors such as lactic, acetic, propionic and benzoic acids were used to demonstrate the utility of acetoin production as an indicator of metabolic disruption. With increasing levels of inhibitor, the metabolic and growth responses were measured by acetoin production and optical density change, respectively. Both measurements decreased in a similar manner with increasing inhibitor concentrations. The data also showed the apparent mode of action of the inhibitors. A bacteriostatic effect was observed for the protonated organic acids, acetic (4 mmol l−1) and propionic (4 mmol l−1), whereas protonated lactic (4 mmol l−1) and benzoic (0·16 mmol l−1) acids gave an irreversible (apparent bacteriocidal) effect. Lactic, acetic, and propionic acids showed stimulation of metabolic activity at low concentrations, but benzoic did not. Acetoin production is a novel method for quantifying and assessing the mode of action of inhibitors against L. monocytogenes . This system can be used to screen inhibitors for applications in food safety.  相似文献   

9.
Veillonellae cultures were grown on agar media supplemented with tartrate and examined for inhibitory effects on the growth of Listeria monocytogenes . Veillonellae cultures grown on media supplemented with 0 or 50 mmol l−1 of tartrate did not inhibit the growth of L. monocytogenes ; however, veillonellae grown on media supplemented with 100, 150 or 200 mmol l−1 of tartrate did inhibit the growth of L. monocytogenes . The inhibition of the growth of L. monocytogenes by the veillonellae was correlated with the increased production of acetate and propionate from tartrate by veillonellae and with the reduction of the pH of the media by L. monocytogenes .  相似文献   

10.
In this Study the effects of both pH and organic acids on Helicobacter pylori NCTC 11637 were tested. Lactobacillus acidophilus, Lact. casei, Lact. bulgaricus, Pediococcus pentosaceus and Bifidobacterium bifidus were assayed for their lactic acid production, pH and inhibition of H. pylori growth. A standard antimicrobial plate well diffusion assay was employed to examine inhibitory effects. Lactic, acetic and hydrochloric acids demonstrated inhibition of H. pylori growth in a concentration-dependent manner with the lactic acid demonstrating the greatest inhibition. This inhibition was due both to the pH of the solution and its concentration. Six strains of Lact. acidophilus and one strain of Lact. casei subsp. rhamnosus inhibited H. pylori growth where as Bifidobacterium bifidus, Ped. pentosaceus and Lact. bulgaricus did not. Concentrations of lactic acid produced by these strains ranged from 50 to 156 mmol 1−1 and correlated with H. pylori inhibition. The role of probiotic organisms and their metabolic by-products in the eradication of H. pylori in vivo remains to be determined.  相似文献   

11.
Yearling brown trout, Salmo trutta , were exposed to low mineral content water (nominal concentrations of 20μmol 1−1 magnesium, 7.7 μmol 1−1 potassium, 44 μmol 1−1 sodium) over a pH range of 4.0–5.2 with ambient calcium concentrations of 2.5–60 μmol 1−1. All fish died at pH 4.0 and 4.2 irrespective of ambient calcium concentration and also at pH 4.4 with only 2–3 μmol 1 −1 calcium (that is calcium-free water except for that leached from the diet or excreted by the fish). Good growth rates were obtained over the remaining treatments which extended down to pH 4.4 with as little as 7 μmol 1−1 calcium. When starved, weight loss was inversely correlated with pH. Effects on plasma chloride, percentage dry weight and calcium, potassium sodium, and phosphorus contents of skin, muscle and bone tissue were also investigated. These demonstrated pH effects on mineral metabolism in starved fish, but no effects were detected in fed fish.  相似文献   

12.
F. RUÍZ-TERÁN AND J.D. OWENS. 1996. The effect of pH on the heat resistance of Bacillus stearothermophilus spores at 100°C in the presence of 0.11 mol 1-1 lactic acid and 0.2 mol 1-1 sodium phosphate buffer was examined. At pH values of 7.0 and 6.0 spores survived 60 min exposure unharmed but at pH 4.3 and 3.0 they died with decimal reduction times (DRTs) of 27 min and 2.8 min, respectively. Death rates were similar in the presence or absence of hydrated soybean cotyledons. In the presence of phosphate buffer and cotyledons at mean pH 3.6 the DRT was 118 min but in the presence, in addition, of lactic acid it was 11 min. It is suggested that the enhanced death rate was due to toxic effects of undissociated lactic acid. Rhizopus oligosporus NRRL 2710 grew well on cotyledons, having pH values from 7.0 to 3.7, prepared by boiling for 60 min in the presence of 0.11 mol 1-1 lactic acid and 0.2 mol 1-1 phosphate buffer.  相似文献   

13.
Exposure of brown trout, Salmo trutta , to zinc under continuous flow conditions over 96 h showed that both water hardness and pH exert major influences on the toxicity of the metal. 96-h LC50 values for total zinc ranged from <0.14mg 1−1 in alkaline soft water (pH 8; lOmg 1−1 as CaCO3) to 3.20 mg 1−1 in acidic hard water (pH 5; 204 mg 1−1 as CaCO3). A variable reduction in zinc toxicity in hard water compared with soft water over the pH range 4–9 was attributed to high external calcium. Zinc toxicity was positively correlated with decreasing acidity over the pH range 5–7, the metal being most toxic at pH 8–9 where metal complexes predominate. Below pH 5 metal toxicity also increased, irrespective of hardness. Water hardness and pH interacted with zinc toxicity in a complex manner, apparently dependent on physical and chemical transformations of the metal, and as changes in uptake. detoxification and excretion by the fish.  相似文献   

14.
Anaerobic l-lactate degradation by Lactobacillus plantarum   总被引:5,自引:0,他引:5  
Abstract Lactobacillus plantarum strains used as silage inoculants were investigated for their ability to metabolize lactic acid anaerobically after prolonged incubation (7–30 days) when glucose was absent from the medium. When citrate was present in the medium together with glucose during the initial fermentation, the lactic acid produced was degraded. Citrate was concomitantly degraded, resulting in accumulation of formic, acetic and succinic acids along with CO2. The anaerobic degradation was confirmed by the use of l 14C(U) labelled lactate. The existence of pyruvate formate lyase in L. plantarum was indicated by using 14C-labelled pyruvate and HPLC identification of end-products. The 1-14C-carboxylic acid group of pyruvate was converted to formic acid, and the 3-14C was found in acetic acid. The key enzyme(s) in this metabolic pathway appears to require anaerobic conditions and induction by citrate.  相似文献   

15.
In strictly anaerobic conditions in a culture medium adjusted to pH 5·2 with HCl and incubated at 30°C, inocula containing < 10 vegetative bacteria of Clostridium botulinum ZK3 (type A) multiplied to give > 108 bacteria per ml in 3 d. Growth from an inoculum of between 10 and 100 spores occurred after a delay of 10–20 weeks. Citric acid concentrations of 10–50 mmol/l at pH 5·2 inhibited growth from both vegetative bacteria and spore inocula, a concentration of 50 mmol/l increasing the number of vegetative bacteria or of spores required to produce growth by a factor of approximately 106. The citric acid also reduced the concentration of free Ca2+ in the medium. The inhibitory effect of citric acid on vegetative bacteria at pH 5·2 could be prevented by the addition of Ca2+ or Mg2+ and greatly reduced by Fe2+ and Mn2+. The addition of Ca2+, but not of the remaining divalent metal ions, restored the concentration of free Ca2+ in the medium to that in the citrate-free medium. The inhibitory effect of citric acid on growth from a spore inoculum was only partially prevented by Ca2+. Citric acid (50 mmol/l) did not inhibit growth of strain ZK3 at pH 6 despite the greater chelating activity of citrate at pH 6 than at pH 5·2. The effect of citric acid and Ca2+ at pH 5·2 on vegetative bacteria of strains VL1 (type A) and 2346 and B6 (proteolytic type B) was similar to that on strain ZK3.  相似文献   

16.
Yearling brown trout, Salmo trutta L., were exposed to various concentrations of inorganic aluminium (0–3.7 μM1−1) over a pH range of 4.3–6.5 in a flow-through bioassay apparatus using synthetic test media. Low pH, in the absence of aluminium, produced little effect on growth or survival except at the lowest pH tested (4.3). At pH less than 5.5, concentrations of total aluminium in excess of 1 μM 1−1 (27μg 1−1) were found to retard growth. The effects of a given aluminium concentration were markedly reduced at pH above 5.5.
The change in aluminium toxicity with pH must be related to changes in aluminium chemistry. When growth rates are correlated with the different aluminium species, calculated using thermodynamic equilibrium constants given in the literature, it appears that the Al(OH)2 + species is the most toxic, with a small contribution also coming from polymeric complexes.  相似文献   

17.
The survival of Cryptosporidium parvum during ensilage of perennial ryegrass was examined in laboratory silos with herbage prepared in one of three different ways; either untreated, inoculated with a strain of Lactobacillus plantarum or by direct acidification with formic acid. The pH values of all silages initially fell below 4.5, but only formic acid-treated silage remained stable at less than pH 4 after 106 d, with the pH of the untreated and inoculant-treated silages rising to above 6. The formic acid-treated silage had a high lactic acid concentration (109 g kg-1 dry matter (DM)) and low concentrations of propionic and butyric acids after 106 d. However, the untreated and inoculant-treated silages showed an inverse relationship, with low lactic acid concentrations and high concentrations of acetic, propionic and butyric acids. These silages also contained ammonia-N concentrations in excess of 9 g kg-1 DM. In terms of the viability of Cryptosporidium parvum oocysts very few differences were seen after 14 d of ensilage with ca 50% remaining viable, irrespective of treatment and total numbers had declined from the initial level of 5.9 × 104 to 1 x 104 g-1 fresh matter. Total oocyst numbers remained approximately the same until the end of the ensiling period, with the percentage of viable oocysts declining to 46, 41 and 32% respectively for formic acid, inoculant and untreated silages. The results are discussed in terms of changes occurring during the silage fermentation, in particular the products which may influence the survival of Cryptosporidium and implications for agricultural practice and the health of silage fed livestock.  相似文献   

18.
Aims:  Weak acids are widely used by the food industry to prevent spoilage and to inhibit the growth of pathogenic micro-organisms. In this study the inhibitory effects of three commonly used weak acids, acetic acid, benzoic acid and sorbic acid, on the growth of Listeria monocytogenes were investigated.
Methods and Results:  In a chemically defined medium at pH 6·4 benzoic acid had the greatest inhibitory effect (50% inhibition of growth at 4 mmol l−1), while acetate was the least inhibitory (50% inhibition of growth at 50 mmol l−1). Mutants lacking either sigmaB (Δ sigB ) or two of the glutamate decarboxylase systems (Δ gadAB ) were used to investigate the contribution these systems make to weak acid tolerance in L. monocytogenes .
Conclusions:  The stress-inducible sigma factor sigmaB (σB) was not required for protection against acetate and played only a minor role in tolerating benzoate and sorbate. The glutamate decarboxylase system, which plays an important role in tolerating inorganic acids, played no significant role in the ability of L. monocytogenes to tolerate these weak acids, and neither did the presence of glutamate in the growth medium.
Significance and Impact of the Study:  These results suggest that the effectiveness of weak acid preservatives in food will not be compromised by the presence of glutamate, at least under mildly acidic conditions.  相似文献   

19.
Lactobacillus amylovorus ATCC 33621 is an actively amylolytic bacterial strain which produces a cell-bound glucoamylase (EC 3.2.1.3). Conditions of growth and glucoamylase production were investigated using dextrose-free de Man-Rogosa-Sharpe (MRS) medium in a 1.5 I fermenter, with varying dextrin concentration (0.1–1.5% (w/v)), pH (4.5–6.5) and temperature (25–55°C). Cell extracts were prepared by subjecting cells to treatment with a French Pressure cell in order to release intracellular proteins. Glucoamylase activity was then assayed. The effects of pH (4.0–9.0), temperature (15–85°C) and substrate (dextrin and starch, 0–2% w/v) concentration on crude enzyme activity were investigated. Optimal growth was obtained in MRS medium containing 1% (w/v) dextrin, at pH 5.5 and 37°C. Glucoamylase production was maximal at the late logarithmic phase of growth, during 16–18 h. Crude enzyme had a pH optimum of 6.0 and temperature optimum of 60°C. With starch as the substrate, maximal activity was obtained at a concentration of 1.5% (w/v). The effects of ions and inhibitors on glucoamylase activity were also investigated. Enzyme activity was not significantly influenced by Ca2+ and EDTA at 1 mmol 1−1 concentration; however Pb2+ and Co2+ were found to inhibit the activity at concentrations of 1 mmol 1−1. The crude enzyme was found to be thermolabile when glucoamylase activity decreased after about 10 min exposure at 60°C. This property can be exploited in the brewing of low calorie beers where only mild pasteurization treatments are used to inactivate enzymes. The elimination of residual enzyme effect would prevent further maltodextrin degradation and sweetening during long-term storage, thus helping to stabilize the flavour of beer.  相似文献   

20.
Samples from a natural population of pike (Esox lucius L.) from the River Danube were used in a 12-month study to determine seasonal variations in biochemical parameters of pike blood sera, hepatosomatic index (HSI) and gonadosomatic index (GSI). The ranges of enzyme activities for sample means were: aspartate aminotransferase (ASAT) 252.0–583.8 U 1−1, alanine aminotransferase (ALAT) 4.9–11 -2 U1 and alkaline phosphatase (AP) 39.5–91.8 U1−1. The ranges of other parameters analysed in serum were total protein 27.7–40.1 g 1 1, urea 0.57–l.52 mmol 1 1 and creatinine 21.2–118.6 μmol 1 1. The range of sample means for HSI and GSI were 1.28–4.16 and 0.07–20.2 respectively. Temperature ranged from 4.5 to 23.5°C. The activity of serum AP was positively correlated to water temperature in males only, while urea and creatinine showed a positive correlation to water temperature in individuals of both sexes. GSI was correlated significantly with HSI in females. Total protein reached the lowest values during the spawning period, while creatinine levels depended on both the sex and season.  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