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 15 毫秒
1.
Six mutations determining resistance to amethopterin were examined for their effects on the active transport of the drug. In strains bearing each of the mutations and exhibiting resistance levels varying from 10- to 100-fold, transport at limiting concentrations of H(3)-amethopterin was reduced from 2.5 to 10 times the rate characteristic of the wild type. Kinetic analysis of transport showed an increase in the value for K(m) of the system in all of the mutants. Values for the wild-type system were 0.9 x 10(-6)m and for the mutants varied between 2.5 x 10(-6)m and 9.0 x 10(-6)m. Values for V(max) were approximately the same for each system. The mutant transport systems also exhibited a shift in pH optimum from near 6.0 (wild-type) to below 5.0. The results were interpreted as an alteration in the binding properties of the permease in the mutant strains.  相似文献   

2.
The properties of folinate and 5-methyltetrahydrofolate (5-CH(3)-H(4)PteGlu) transport mechanism of Pediococcus cerevisiae were studied. The uptake was dependent on temperature, pH (optimum for both compounds at pH 6.0), and glucose. Iodoacetate, potassium fluoride, and sodium azide inhibited the uptake. 5-CH(3)-H(4)-PteGlu was apparently not metabolized but folinate was metabolized. Metabolism of folinate was reduced by preincubation of cells with fluorodeoxyuridine. The transport system for folinate and 5-CH(3)-H(4)PteGlu were specific for the l-isomers. Pteroylglutamate, aminopterin, and amethopterin did not interfere with the uptake. Tetrahydrofolate competed with the uptake of folinate. The transport of folinate and 5-CH(3)-H(4)PteGlu at 37 C conformed to Michaelis-Menten kinetics; apparent K(m) for both compounds was 4.0 x 10(-7)m, and the V(max) for folinate was 1.0 x 10(-10) moles per min per mg (dry weight) and for 5-CH(3)-H(4)PteGlu it was 1.6 x 10(-10) moles per min per mg (dry weight). Both compounds accumulated in the intracellular pool at a concentration about 80- to 140-fold higher than that in the external medium. Folinate inhibited competitively the uptake of 5-CH(3)-H(4)PteGlu with a K(i) of 0.4 x 10(-7)m. Unlike 5-CH(3)-H(4)PteGlu, which accumulated only at 37 C, folinate was also taken up at 0 C by a glucose- and temperature-independent process, which was not affected by the metabolic inhibitors mentioned above. Since at 0 C the intracellular concentration of folinate was also considerably higher than the external, binding of the substrate to some cellular component is assumed. The finding of an efficient transport system for l-5-CH(3)-H(4)PteGlu is of special interest, since this compound has no growth-promoting activity for P. cerevisiae.  相似文献   

3.
High concentration of inositol 1,4,5-trisphosphate in sea urchin sperm   总被引:1,自引:0,他引:1  
We measured inositol 1,4,5-trisphosphate (InsP3) content of sea urchin gametes by using a specific protein binding assay, and found that a spermatozoon contains 4 x 10(-19) to 1 x 10(-18) moles of InsP3 before the acrosome reaction. Since the acrosome reaction has previously been shown to increase the InsP3 content of sperm severalfold, our measurement indicates that a spermatozoon contains at least 2 x 10(-18) moles of InsP3 at fertilization, corresponding to a concentration in the spermatozoon of about 1 mM. The threshold for activation of eggs by injection of InsP3 dissolved in a much larger volume of solution has been found to be about 3 x 10(-18) moles, corresponding to a concentration in the injectate of 1 microM. This suggests that sea urchin sperm may contain enough InsP3 to activate eggs. With an electroporation method, we also showed that sperm extract acts on eggs only from inside, consistent with a primary messenger role for InsP3.  相似文献   

4.
Cucumber hypocotyls were extracted and the extract centrifuged at 100,000g to yield a supernatant or cytosol fraction. Binding of [(3)H]-gibberellin(4) (GA(4)) to soluble macromolecular components present in the cytosol was demonstrated at 0 C by Sephadex chromatography. Binding assays performed with cytosol that had been preheated or incubated with protease, DNase, RNase, or phospholipase A or C indicated that heat and protease treatments disrupted the binding, which suggests that binding occurred to a protein. Equilibrium dialysis of a protein-enriched fraction prepared by ammonium sulfate precipitation also indicated binding of [(3)H]GA(4) to macromolecular components. [(3)H]GA(4) binding was pH-sensitive, saturable, reversible, and significantly affected by biologically active gibberellins, but not by inactive gibberellins or other plant hormones such as indoleacetic acid, abscisic acid, or kinetin. Thin layer chromatography indicated that [(3)H]GA(4), and not a metabolite, was the species bound. A kinetic analysis indicated that specific binding of [(3)H]GA(4) was due to a single class of binding sites having an estimated K(d) of 10(-7) molar and a concentration of 0.8 x 10(-12) moles gram(-1) fresh weight or 0.4 x 10(-12) moles milligram(-1) soluble protein.  相似文献   

5.
The effects of DBcAMP in doses from 1.5 x 10(-8) to 1.5 x 10(-3) M on the compartmental apparent surface area (ASA) and (5(-3H)uridine radioactivity concentration (URC), (methyl-3H)thymidine labelling index per 1 hour ([Me-3H]Tdr LI/h) and per cent mitotic index (MI%) and colchicine metaphase index (CMI%) of young rat differentiated hepatocytes in primary tissue culture were investigated by morphometric and radioautographie methods. In these cells DBcAMP was found to elicit: (1) progressive increments in the ASA of nucleoli, karyoplasm and cytoplasm; (2) peak increases in nucleolar URC at 1.5 x 10(-8) and 10(-5) M, but a slight decrease at 1.5 x 10(-3) M; (3) singificant increments in karyoplasmic and total nuclear URC at all doses, except at 1.5 x 10(-6) and 10(-4) M, when such parameters remained at control levels; (4) steady and progressive increases in cytoplasmic and total cell URC values; (5) marked increments in (Me-3H)Tdr LI/h, MI% and CMI% up to the dose of 1.5 x 10(-4) M, but at 1.5 x 10(-3) M these parameters were found to be either much less enhanced or to approach closely to control values. cAMP in doses from 1.5 x 10(-8) to 10(-4) M also markedly incremented the in vitro hepatocyte CMI%, while having a lesser stimulatory effect at 1.5 x 10(-3)M. Finally of the various possible metabolites of DBcAMP administered at 1.5 x 10(-8) M to liver cultures, N6- and O2'-MBcAMP and, again, cAMP significantly increased the CMI%, of cultured hepatocytes, whereas 5'-AMP, adenosine and allantoin had no significant effect and Na-butyrate slightly decreased it. The present observations strengthen the hypothesis that cAMP and its butyrated derivatives, by possibly amplifying the template activity of the liver chromatin, accelerate the flow of differentiated primary young rat hepatocytes into the various stages of the mitotic cell cycle.  相似文献   

6.
The binding activity of [3H]dexamethasone to the specific receptor was studied in the cytoplasmic fraction of a established fibroblast line derived from rat carrageenin granuloma in culture condition. Specific receptor to dexamethasone was demonstrated. Scatchard analysis revealed a single class of binding sites with a dissociation constant for [3H]dexamethasone of 3.64 - 10(-8) M and a concentration of binding sites of 0.825 pmol per mg cytosol protein. The number of cytoplasmic binding sites per cell was calculated at 1.15 - 10(5). Total binding activity to [3H]dexamethasone of the cytoplasmic fraction was enhanced when the cells were cultured in a medium containing salicylic acid was at 37 degrees C. The maximum enhancement was seen at the concentration of 10(-3)M and in 3h treatment of salicylic acid. This enhancement by salicylic acid was lost when cycloheximide was added to the culture medium at the same time. If salicyclic acid was added to the cell free system, it showed no effect on the binding activity. The other non-steroidal anti-inflammatory drugs; phenylbutazone and indomethacin,also enhanced the total binding activity to [3H]dexamethasone of the cytoplasmic fraction at the concentration of 2 - 10(-5) M and 2 - 10(-7) M, respectively.  相似文献   

7.
A new adenosine-selective membrane electrode using rabbit thymus tissue as catalyst is described. A typical response slope of 51.2 mV per concentration decade is observed over a linear range which extends from 3.16 x 10(-5) M to 5.62 x 10(-3) M. Detection limits of 2.99 x 10(-5) M have been established. Measured response times are 7 min. The coefficient of variation ranged from 1 to 5.62% (n = 7, m = 5). Fourteen compounds were specifically tested as possible interferents, but no significant response was observed. The standard recoveries of adenosine were from 95.3 to 104.0% (m = 5, n = 5), and the recoveries of adenosine in rabbit blood ranged from 94.0 to 108.4% (n = 3, m = 5) over the linear range. This tissue-based biosensor has excellent sensitivity and selectivity, and has additional advantages of simplicity and low cost. The biosensor can be used to measure directly the concentration of adenosine in body fluid samples without sample processing.  相似文献   

8.
Both ascorbic acid and copper were strong prooxidants in the oxidation of linoleate in a buffered (pH 7.0) aqueous dispersion at 37 degrees C. Minimum concentrations at which catalytic activity was detected were 1.3 x 10(-7) m for copper and 1.8 x 10(-6) m for ascorbic acid. For concentrations up to 10(-3) m, the increase in rate of oxidation with increase in concentration of catalyst was greater for ascorbic acid than for copper. Ascorbic acid had maximum catalytic activity at 2.0 x 10(-3) m, but was still prooxidant at the highest concentration tested (5.0 x 10(-2) m). Dehydroascorbic acid was a weaker prooxidant than ascorbic acid. Further degradation products of ascorbic acid were not prooxidant. In early stages of the oxidation autocatalytic behavior was observed with copper, but not with ascorbic acid. Ascorbic acid functioned as a true catalyst, i.e., it accelerated the reaction but it was not oxidized simultaneously with the linoleate. It is proposed that the dehydroascorbic acid radical initiates the linoleate oxidation reaction.  相似文献   

9.
Aleem, M. I. H. (Research Institute for Advanced Studies, Baltimore, Md.). Thiosulfate oxidation and electron transport in Thiobacillus novellus. J. Bacteriol. 90:95-101. 1965.-A cell-free soluble enzyme system capable of oxidizing thiosulfate was obtained from Thiobacillus novellus adapted to grow autotrophically. The enzyme systems of autotrophically grown cells brought about the transfer of electrons from thiosulfate to molecular oxygen via cytochromes of the c and a types; the reactions were catalyzed jointly by thiosulfate oxidase and thiosulfate cytochrome c reductase. The levels of both of these enzymes were markedly reduced in the heterotrophically grown organism. Cell-free extracts from the autotrophically grown T. novellus catalyzed formate oxidation and enzymatically reduced cytochrome c with formate. Both formate oxidation and cytochrome c reduction activities were abolished under heterotrophic conditions. The thiosulfate-activating enzyme S(2)O(3) (-2)-cytochrome c reductase, as well as thiosulfate oxidase, was localized chiefly in the soluble cell-free fractions, and the former enzyme was purified more than 200-fold by ammonium sulfate fractionation and calcium phosphate gel adsorption procedures. Optimal activity of the purified enzyme occurred at pH 8.0 in the presence of 1.67 x 10(-1)m S(2)O(3) (-2) and 2.5 x 10(-4)m cytochrome c. The thiosulfate oxidase operated optimally at pH 7.5 and thiosulfate concentrations of 1.33 x 10(-3) to 3.33 x 10(-2)m in the presence of added cytochrome c at a concentration of 5 x 10(-4)m. Both enzymes were markedly sensitive to cyanide and to a lesser extent to some metal-binding agents. Although a 10(-3)m concentration of p-hydroxymercuribenzoate had no effect on S(2)O(3) (-2)-cytochrome c reductase, it caused a 50% inhibition of S(2)O(3) (-2) oxidase, which was completely reversed in the presence of 10(-3)m reduced glutathione. Carbon monoxide also inhibited S(2)O(3) (-2) oxidase; the inhibition was completely reversed by light.  相似文献   

10.
Dempsey, Walter B. (University of Florida, Gainesville). Synthesis of pyridoxine by a pyridoxal auxotroph of Escherichia coli. J. Bacteriol. 92:333-337. 1966.-A pyridoxal auxotroph of Escherichia coli B produced pyridoxol and pyridoxol 5'-phosphate during starvation for pyridoxal. The identification of these compounds was made both by bioassay and by ion-exchange chromatography. Pyridoxol 5'-phosphate oxidase activity was absent in extracts of the auxotroph. The rate of synthesis of total pyridoxine by a pyridoxal-starved culture of this auxotroph was 6.0 x 10(-6) moles per mg per hr. Cellular content of pyridoxine was constant at 4.0 x 10(-10) moles/mg.  相似文献   

11.
Bilirubin glucuronyltransferase. Specific assay and kinetic studies   总被引:5,自引:5,他引:0       下载免费PDF全文
1. Bilirubin glucuronide was synthesized in vitro in a system containing a rat liver microsomal fraction, UDP-glucuronic acid, Mg(2+) and bilirubin. The enzymic synthesis was accomplished without the addition of a bilirubin carrier. 2. Azobilirubin and azobilirubin glucuronide were separated by t.l.c. and paper chromatography and the measurement of the conjugate provided a specific assay for bilirubin UDP-glucuronyltransferase (EC 2.4.1.17). 3. This diazo compound was labelled when [U-(14)C]UDP-glucuronic acid was employed in the transglucuronidation reaction. 4. Identity of the glucuronide nature of the product was further confirmed by hydrolysis with beta-glucuronidase prepared from limpets and Helix pomatia. In each instance azobilirubin and glucuronic acid were liberated. 5. There was a close correlation between the bilirubin glucuronyl-transferase activity as measured by two procedures, colorimetric and radioisotopic. The specific activities so measured were 19nmol of bilirubin ;equivalents' conjugated/h per mg of protein and 16.9-18.4nmol of UDP-glucuronic acid incorporated/h per mg of protein, respectively. On this basis, it was concluded that the major product formed in vitro was bilirubin monoglucuronide; this represents about 77% of the total products formed. 6. The K(m) values for bilirubin and UDP-glucuronic acid at pH8.2 are 3.3x10(-4)m and 1.67x10(-3)m, respectively. 7. The addition of Mg(2+) at a final concentration of 5mm to the reaction mixture increased the rate of conjugation by 5.6-fold in the microsomal preparation that had been subjected to overnight dialysis against 10mm-EDTA (disodium salt). 8. Diethyl-nitrosamine at a final concentration of 1-20mm has no effect on the glucuronidation of bilirubin in vitro.  相似文献   

12.
7-Amino-4-methylcoumarin-3-acetic acid (AMCA) has been found to be a useful fluorophore for immunofluorescence. The present study describes a spectrophotometric method for determining the ratio of moles AMCA to moles protein (or the f/p ratio) in an AMCA-conjugated IgG. The concentration of a substance absorbing light can be determined spectrophotometrically using Beer's Law: Absorbance = Concentration x Extinction coefficient. From Beer's law, one can derive the following formula for determining the f/p ratio of AMCA-IgG conjugates: f/p = (epsilon 280IgG).A350 - (epsilon 350IgG).A280/(epsilon 350AMCA).A280 - (epsilon 280AMCA).A350 where A is the optical density of the conjugate at the given wavelength and epsilon is the extinction coefficient of a substance at the wavelength specified. Using conjugates of model proteins, it was found that the extinction coefficients of the AMCA moiety of AMCA-conjugated protein were 1.90 x 10(4) at 350 nm and 8.29 x 10(3) at 280 nm. Similarly, it was found that the extinction coefficients of swine IgG were 1.56 x 10(3) at 350 nm and 1.26 x 10(5) at 280 nm. Thus, for AMCA-conjugated swine IgG: f/p = (1.26 x 10(5)).A350 - (1.56 x 10(3)).A280/(1.47 x 10(4)).A280 - (6.42 x 10(3)).A350 [corrected]. Based on this formula, the f/p ratios of some AMCA-IgG conjugates useful for immunohistochemistry have been found to range between 6 and 24.  相似文献   

13.
The binding of polyamines and of ethidium bromide to tRNA.   总被引:1,自引:0,他引:1  
The binding of spermidine and ethidium bromide to mixed tRNA and phenylalanine tRNA has been studied under equilibrium conditions. The numbers and classes of binding sites obtained have been compared to those found in complexes isolated by gel filtration a low ionic strength. The latter complexes contain 10-11 moles of either spermidine or ethidium per mole of tRNA; either cation is completely displaceable by the other. In ethidium complexes, the first 2-3 moles are bound in fluorescent binding sites; the remaining 7-8 molecules bind in non-fluorescent form. At least one of the binding sites for spermidine appears similar to a binding site for fluorescent ethidium. Similar results are found with E. coli formylmethionine tRNA. Spermine, in excess of 18-20 moles per mole tRNA, causes precipitation of the complex. Putrescine does not form isolable complexes with yeast tRNA and displaces ethidium less readily from preformed ethidium-tRNA complexes. Under equilibrium conditions, in the absence of Mg++, there are 16-17 moles of spermidine bound per mole of tRNA as determined by equilibrium dialysis. Of these, 2-3 bind with a Ksence of 9 mM Mg++, the total number of binding sites is decreased slightly and there appears to be only one class of sites with a Ka = 600 M(-1). Quantitatively similar results are obtained for the binding of spermidine to yeast phenylalanine tRNA. When the interaction between ethidium bromide and mixed tRNA is studied by equilibrium dialysis or spectrophotometric titration, two classes of binding sites are obtained: 2-3 molecules bind with an average Ka = 6.6 x 10(5) M(-1) and 14-15 molecules bind with an average Ka = 4.1 x 10(4) M(-1). Spermidine, spermine, and Mg++ compete effectively for both classes of ethidium sites and have the effect of reducing the apparent binding constants for ethidium. When the binding of ethidium is studied by fluorometry, there are 3-4 highly fluorescent sites per tRNA. These sites are also affected by spermidine, spermine and Mg++. Putrescine has little effect on any of the classes of binding sites. These data are consistent with those found under non-equilibrium conditions. They suggest that polyamines bind to fairly specific regions of tRNA and may be involved in the maintenance of certain structural features of tRNA.  相似文献   

14.
Effect of Gossypol on Some Oxidative Respiratory Enzymes   总被引:3,自引:2,他引:1  
Gossypol was examined in relation to its effect on certain enzymes and enzyme complexes associated with the tricarboxylic acid cycle and the electron transport system. Succinic dehydrogenase and cytochrome oxidase activity from sweet potato was completely inhibited by gossypol at 7.5 x 10(-3)m and 2.0 x 10(-3)m, respectively. Succinoxidase activity of the same preparations was fully inhibited at a lower concentration, 2.5 x 10(-4)m. This concentration did not affect either succinic dehydrogenase or cytochrome oxidase, the primary and terminal enzymes of the succinoxidase complex. The nature of the intermediate step or steps inhibited at this concentration is not yet known. Gossypol was further shown to inhibit phosphorylation at concentrations having no appreciable effect on oxidation. Inhibition in general was not reduced by increased substrate concentrations in the enzyme systems examined, with the exception of cytochrome c for cytochrome oxidase. Bovine serum albumin was partially effective in reducing gossypol inhibition, provided that it was present before enzyme exposure to gossypol.  相似文献   

15.
The cardiac effects of PAF and its antagonist BN 52021 have been investigated on the isolated perfused guinea-pig heart maintained at a constant hydrostatic perfusion pressure of 80 cm water. In this model, PAF (1 x 10(-11) to 1 x 10(-7) moles) induced a dose-dependent coronary vasoconstriction, a decrease in heart rate and a fall in contractile force. BN 52021 (1 x 10(-6) to 2 x 10(-4) M) dose-dependently inhibited the vasospasm induced by PAF (1 x 10(-10) moles). BN 52021 also antagonized the decrease in coronary flow and heart rate, but not that of contractile force induced by a high dose of PAF (1 x 10(-7) moles). This dose of PAF also significantly (p less than 0.001) provoked a marked release of TxB2 but did not alter the generation of 6 Keto PGF1 alpha, PGE2 or LTC4. The PAF-induced increase in TxB2 release was completely abolished by BN 52021.  相似文献   

16.
In promoting oxidation of 0.02 m potassium linoleate in a buffered (pH 7.0) aqueous dispersion at 37 degrees C, ascorbic acid at low concentrations (1.8 x 10(-6) and 1.8 x 10(-5) m) in combination with copper (1.3 x 10(-7) to 1.3 x 10(-3) m) had greater catalytic activity than the additive activity of the two catalysts individually. Possible explanations for the enhanced catalysis include reduction of copper by ascorbic acid to the cuprous form, increased concentration of semidehydroascorbic acid radical, and formation of a metal-ascorbic acid-oxygen complex. Some combinations of ascorbic acid (1.8 x 10(-4) and 1.8 x 10(-3) m) and copper (1.3 x 10(-6) and 1.3 x 10(-3) m) inhibited the formation of conjugated dienes but not the oxidation of ascorbic acid, and caused rapid loss of part of the conjugated dienes that were already present. It is suggested that free-radical inhibitors formed by the combination of catalysts inhibit initiation of lipid oxidation but not copper-catalyzed oxidation of ascorbic acid. Effects of the inhibitory combinations on changes in UV absorption by conjugated dienes, and absorbance in the TBA test, indicate the presence of at least two conjugated dienes that differ in stability.  相似文献   

17.
Anaerobic filter matings of Butyrivibrio fibrisolvens H17c, CF3, D1, or GS113, representing different DNA relatedness groups, were done with Enterococcus faecalis CG110, which contains chromosomally inserted Tn916. Tetracycline-resistant transconjugants were obtained with each mating pair at average frequencies of 4.4 x 10(-6) (per recipient) and 5.2 x 10(-6) (per donor). The transfer frequencies of Tn916 into B. fibrisolvens varied 5- to 10-fold with mating time, strain, and growth stage. By using Southern hybridization with pAM120 as the probe, Tn916 was shown to insert at one or more separate chromosomal sites for each strain of B. fibrisolvens. Retransfer of Tn916 from B. fibrisolvens H17c or CF3 to E. faecalis OG1-X or JH 2-2 or to B. fibrisolvens D1 or GS113 could not be shown. Matings of E. faecalis RH110, which contains chromosomally inserted Tn916 delta E, with B. fibrisolvens 49, H17c, D1, CF3, GS113, or VV-1 resulted in erythromycin-resistant transconjugants at average frequencies of 5.3 x 10(-7) (per recipient) and 2.5 x 10(-7) (per donor). Tn916 delta E was shown by Southern hybridization with pAM120 to insert at one or more sites in the chromosome of each strain. B. fibrisolvens H17c was anaerobically filter mated with E. faecalis JH 2-SS, which contains pAM beta 1. Erythromycin-resistant transconjugants were obtained at frequencies of 2 x 10(-5) (per recipient) and 6 x 10(-5) (per donor). The presence of pAM beta 1 in these transconjugants could not be shown by agarose gel electrophoresis of plasmid minilysates but could be shown by Southern hybridization analysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Anaerobic filter matings of Butyrivibrio fibrisolvens H17c, CF3, D1, or GS113, representing different DNA relatedness groups, were done with Enterococcus faecalis CG110, which contains chromosomally inserted Tn916. Tetracycline-resistant transconjugants were obtained with each mating pair at average frequencies of 4.4 x 10(-6) (per recipient) and 5.2 x 10(-6) (per donor). The transfer frequencies of Tn916 into B. fibrisolvens varied 5- to 10-fold with mating time, strain, and growth stage. By using Southern hybridization with pAM120 as the probe, Tn916 was shown to insert at one or more separate chromosomal sites for each strain of B. fibrisolvens. Retransfer of Tn916 from B. fibrisolvens H17c or CF3 to E. faecalis OG1-X or JH 2-2 or to B. fibrisolvens D1 or GS113 could not be shown. Matings of E. faecalis RH110, which contains chromosomally inserted Tn916 delta E, with B. fibrisolvens 49, H17c, D1, CF3, GS113, or VV-1 resulted in erythromycin-resistant transconjugants at average frequencies of 5.3 x 10(-7) (per recipient) and 2.5 x 10(-7) (per donor). Tn916 delta E was shown by Southern hybridization with pAM120 to insert at one or more sites in the chromosome of each strain. B. fibrisolvens H17c was anaerobically filter mated with E. faecalis JH 2-SS, which contains pAM beta 1. Erythromycin-resistant transconjugants were obtained at frequencies of 2 x 10(-5) (per recipient) and 6 x 10(-5) (per donor). The presence of pAM beta 1 in these transconjugants could not be shown by agarose gel electrophoresis of plasmid minilysates but could be shown by Southern hybridization analysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
Using molecular techniques and microsensors for H(2)S and CH(4), we studied the population structure of and the activity distribution in anaerobic aggregates. The aggregates originated from three different types of reactors: a methanogenic reactor, a methanogenic-sulfidogenic reactor, and a sulfidogenic reactor. Microsensor measurements in methanogenic-sulfidogenic aggregates revealed that the activity of sulfate-reducing bacteria (2 to 3 mmol of S(2-) m(-3) s(-1) or 2 x 10(-9) mmol s(-1) per aggregate) was located in a surface layer of 50 to 100 microm thick. The sulfidogenic aggregates contained a wider sulfate-reducing zone (the first 200 to 300 microm from the aggregate surface) with a higher activity (1 to 6 mmol of S(2-) m(-3) s(-1) or 7 x 10(-9) mol s(-1) per aggregate). The methanogenic aggregates did not show significant sulfate-reducing activity. Methanogenic activity in the methanogenic-sulfidogenic aggregates (1 to 2 mmol of CH(4) m(-3) s(-1) or 10(-9) mmol s(-1) per aggregate) and the methanogenic aggregates (2 to 4 mmol of CH(4) m(-3) s(-1) or 5 x 10(-9) mmol s(-1) per aggregate) was located more inward, starting at ca. 100 microm from the aggregate surface. The methanogenic activity was not affected by 10 mM sulfate during a 1-day incubation. The sulfidogenic and methanogenic activities were independent of the type of electron donor (acetate, propionate, ethanol, or H(2)), but the substrates were metabolized in different zones. The localization of the populations corresponded to the microsensor data. A distinct layered structure was found in the methanogenic-sulfidogenic aggregates, with sulfate-reducing bacteria in the outer 50 to 100 microm, methanogens in the inner part, and Eubacteria spp. (partly syntrophic bacteria) filling the gap between sulfate-reducing and methanogenic bacteria. In methanogenic aggregates, few sulfate-reducing bacteria were detected, while methanogens were found in the core. In the sulfidogenic aggregates, sulfate-reducing bacteria were present in the outer 300 microm, and methanogens were distributed over the inner part in clusters with syntrophic bacteria.  相似文献   

20.
Two essentially isogenic strains of Escherichia coli K-12 were compared: D31 had chromosomally and D1-R1 episomally mediated resistance to ampicillin. The two strains had the same ability to form colonies on ampicillin plates, but in other tests they were quite different. In serial dilution tests as well as in exponentially growing cultures, D1-R1 was far more resistant to ampicillin than was D31. The inoculum effect with D1-R1 was large and with D31 was rather small. On plates, D31 was more resistant to penicillin G than was D1-R1. The penicillinase activity of buffer suspended cells against dl-ampicillin was 15 times higher for D1-R1 than for D31, but the two strains showed about the same rate of hydrolysis of penicillin G. With dl-ampicillin as substrate, for D1-R1 the apparent K(m) was 1.7 x 10(-4)m, whereas D31 gave a slightly sigmoid curve with a half-saturation concentration of about 5 x 10(-3)m. No induction of penicillinase activity was found. When the growth rate was varied by a factor of four, the amount of penicillinase per cell mass was constant in both D1-R1 and D31, whereas in two wild-type strains the amounts of penicillinase increased with increasing growth rates. With exponentially growing D1-R1, ampicillin disappearance started within 3 min, but at low ampicillin concentrations the rate was less than 10% of the rate of hydrolysis by buffer-suspended cells. Before D31 started hydrolysis, there was a lag period that lasted at least one generation and depended on the concentration of ampicillin. After this lag period, the rate of hydrolysis was 10 times higher than that observed with buffer-suspended cells. These differences between growing and nongrowing cells indicate that both the chromosomally and the episomally mediated penicillinases are controlled by some products present in growing cells.  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