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.
Cell-free extracts of Salmonella typhimurium synthesize a mutagenic azide metabolite from sodium azide and O-acetylserine. S. typhimurium mutant DW379 (O-acetylserine sulfhydrylase-deficient) extracts were neither able to carry out this reaction nor produce the mutagenic azide metabolite in vivo. The in vitro reaction was inhibited by sulfide but not by l-cysteine. The catalytic activity responsible for the mutagenic metabolite synthesis was stable to brief heating up to 55°C and had a pH optimum between 7–7.4. These results suggest that the enzyme O-acetylserine sulfhydrylase catalyzes the reaction of azide with O-acetylserine to form a mutagenic azide metabolite.  相似文献   

2.
The ability of L-cysteine to inhibit azide-metabolite synthesis and mutagenecity is investigated in Salmonella typhimurium TA1530 and cys E6 strains. L-cysteine specifically inhibits the synthesis of the mutagenic azide metabolite as other compounds containing SH group did not affect the production of this metabolite. Azide mutagenicity is completely inhibited by L-cysteine at a concentration (5 μmoles/plate) where the metabolite mutagenicity was not affected. O-Acetyl-L-serine can reverse the L-cysteine mediated inhibition of the metabolite synthesis and thus mutagenicity in the same strains. These results suggest that O-acetyl-L-serine may be required to synthesize the azide metabolite or its precursor.  相似文献   

3.
A reproducible and sensitive method is presented for quantitating sodium azide (NaN3) that exploits the fact that NaN3 inhibits Escherichia coli RNA synthesis. A linear correlation is observed between incorporation of [3H]uridine into a trichloroacetic acid-precipitable form and NaN3 concentration over a 31- to 2000-μg range of azide. This technique was used to determine the azide content of a complex enzyme solution where established colorimetric azide determinations proved to be unworkable. This technique when properly controlled should be applicable to a variety of similar solutions.  相似文献   

4.
Nitric oxide (NO) is involved in versatile functions in plant growth and development as a signaling molecule. To date, plants have been reported to produce NO following exposure to nitrite (N O 2 ? ) the amino acid L-arginine, hydroxylamine, or polyamines. Here we demonstrate azide-dependent NO production in plants. The water fern Azolla pinnata emitted NO into air upon exposure to sodium azide (NaN3). The NO production was dependent on azide concentration and was strongly inhibited by potassium cyanide (KCN). Incubation of A. pinnata with the catalase inhibitor 3-aminotriazole (3-AT) abolished the azide-dependent NO production. Although nitrite-dependent NO production was inhibited by sodium azide, azide-dependent NO production was not affected by nitrite. These results indicate that A. pinnata enzymatically produces NO using azide as a substrate. We suggest that plants are also capable of producing NO from azide by the action of catalase as previously reported in animals.  相似文献   

5.
Exposure of rounded, glass-adherent hemocytes from a Schistosoma mansoni-susceptible (PR albino) and S. mansoni-refractory (10-R2) stock of snails, Biomphalaria glabrata, to fluoresceinlabeled concanavalin A induces a redistribution of surface membrane Con A receptors. Receptor redistribution (patching and capping) on hemocytes from both snail stocks can be characterized as (1) rapid, with maximum cap formation occurring within 15 min of lectin treatment at 22°C, (2) sodium azide sensitive, but only at relatively high inhibitor concentrations (100–200 mm?N3 for capping and 200 mm?N3 for patching inhibition), (3) pronase sensitive (partial), but trypsin resistant, and (4) generally unaffected by exposure of snails to S. mansoni miracidia 60 or 180 min prior to extraction of hemolymph (hemocyte) samples for Con A testing. Although differences in the time course of receptor redistribution are exhibited between PR albino and 10-R2 snail hemocytes, the results of experiments involving sodium azide, proteolytic enzymes, and schistosome exposure strongly suggest that Con A-binding determinants and their associated membrane components on rounded hemocytes are very similar in both susceptible and refractory Biomphalaria stocks. It is concluded that if schistosome recognition in refractory 10-R2 snails is mediated through specific hemocyte membrane components, those components associated with Con A reactivity probably are not directly involved in the recognition process.  相似文献   

6.
One-pot reaction of cobalt(II) nitrate hexahydrate Co(NO3)2 · 6H2O with H2salpn (N,N′-bis(salicylidene)-1,3-diaminopropane) in presence of a large excess of sodium azide (NaN3) gives the new Co(III) compound {Na[CoIII(μ-salpn)(μ1,1-N3)2]}n (1), which was characterized by single crystal X-ray diffraction analysis. The crystal structure shows polymeric 1D complex generated by the hexadentate Schiff base salpn2− and two crystallographically different azide ligands. The two nitrogen atoms of the salpn ligand are bonded to the cobalt(III) ion while each phenoxo oxygen atom is bonded to the same Co(III) ion and to two equivalent sodium ions. Each azide ligand acts with the end-on bridging coordination mode between Co(III) and Na(I) ions. The Co(III) ion adopts a distorted octahedral geometry arising from two oxygen and two nitrogen atoms of the salpn ligand and from two nitrogen atoms of the two crystallographically different azide ligands in trans positions. Such [Co(salpn)(N3)2] entities are connected each other by sodium ions through four oxygen atoms of two equivalent Schiff base ligands and two nitrogen atom of the two different azide ligands to generate the 1D structure of 1.  相似文献   

7.
The solubilization and subsequent separation of the hepatic microsomal ethanol-oxidizing system from alcohol dehydrogenase and catalase activities by DEAE-cellulose column chromatography is described. Absence of alcohol dehydrogenase in the column eluates exhibiting microsomal ethanol-oxidizing system activity was demonstrated by the failure of NAD+ to promote ethanol oxidation at pH 9.6. Differentiation of the microsomal ethanol-oxidizing system from alcohol dehydrogenase was further shown by the apparent Km for ethanol (7.2 mm, insensitivity of the microsomal ethanol-oxidizing system to the alcohol dehydrogenase inhibitor pyrazole (0.1 mm) and by the failure of added alcohol dehydrogenase to increase the ethanol oxidation. Absence of catalatic activity in these fractions was demonstrated by spectrophotometric and polarographic assay. Differentiation of the microsomal ethanol-oxidizing system from the peroxidatic activity of catalase was shown by the apparent Km for oxygen (8.3 μm), insensitivity of the microsomal ethanol-oxidizing system to the catalase inhibitors azide and cyanide, and by the lack of a H2O2-generating system (glucose-glucose oxidase) to sustain ethanol oxidation in the eluates. The oxidation of ethanol to acetaldehyde by the alcohol dehydrogenase- and catalase-free fractions required NADPH and oxygen and was inhibited by CO. The column eluates showing microsomal ethanol-oxidizing system activity contained cytochrome P-450, NADPH-cytochrome c reductase, and phospholipids and also metabolized aminopyrine, benzphetamine, and aniline.  相似文献   

8.
A kinetic study of the oxidation of (hydroxyethyl)ferrocene (HEF) by [2-pyridylmethylbis(2-ethyl-thioethyl)ainine]copper(II) (Cu(pmas)2+) is reported, with the objective of documenting the influence of the two thioether sulfur ligands on the electron transfer rate. Both reactants exhibit a first-order dependence at pH 6, I = 0.1 M(NaNO3); k(25°C) = 1.3 × 104M−1sec−1, ΔH3 = 10.1 kcal/mole, ΔS3 = −6 eu. The apparent Cu(pmas)2+/+ self-exchange electron transfer rate constant calculated from this reaction on the basis of relative Marcus theory (4.7 × 101M−1 sec−1) agrees well with previous findings on ferrocytochrome c, Fe(CN)64−, and Ru(NH3)5py2+ oxidations. Spectrophotometric titrations of Cu(pmas)2+ and Cu(tmpa)2+ (tmpa = tris(2-pyridylmethyl)amine) with azide ion showed that both Cu(pmas)N3)+ (Kf1 = 3.1 × 103M−1) and Cu(pmas)(N3)2 (Kf2 = 3.5 × 101M−1) but Cu(tmpa)(N3)+ (Kf = 6.6 × 102M−1) are formed up to 0.15 M N3 (25°C, pH 6, I = 0.2 M), suggesting that a thioether sulfur atom is displaced in the uptake of a second N3 ion by Cu(pmas)(N3)+. The effect of thioether sulfur displacement by azide ion on the HEF-Cu(pmas)2+ reaction rate may be understood entirely through the tendency of N3 to shift the position of the redox equilibrium towards the reactant side, without invoking any special role for the sulfur ligand in promoting electron transfer reactivity.  相似文献   

9.
The products obtained from the reaction of Pt(IV)Cl4(LL) compounds (LL denotes the chelating ligands ethylenediamine (en) and 2,2-dimethyl-1,3-diaminopropane (dmdap), or two cis- or trans-coordinated ammines) with 9-methylhypoxanthine (mHyp) at high temperature (80°C) have been characterized by proton NMR spectroscopy. It appeared that both platinum(II) and platinum(IV) adducts were present in the reaction mixtures. After cation-exchange chromatography, the Pt(II) compound could be characterized as Pt(II)(LL)(mHyp)2, whereas the Pt(TV) fractions appeared to contain mainly one or two adducts for the chelating diamine compound but more adducts for the ammine compounds. A 3J(195Pt-1H) coupling was observed for the Pt(IV), but not for the Pt(II) compounds at the used spectrometer frequency. This supplies a useful tool to discriminate between these two types of platinum adducts.  相似文献   

10.
11.
Tang Jun  Wu Shupin  Bai Juan  Sun Daye 《Planta》1996,198(4):510-516
A 21-kDa calmodulin (CaM)-binding protein and a 19-kDa calmodulin-binding protein were detected in 0.1 M CaCl2 extracts of Angelica dahurica L. suspension-cultured cells and carrot (Daucus carota L.) suspension-cultured cells, respectively, using a biotinylated cauliflower CaM gel-overlay technique in the presence of 1 mM Ca2+. No bands, or very weak bands, were shown on sodium dodecyl sulfate-polyacrylamide gel electrophoresis gels overlayed with biotinylated cauliflower CaM when 1 mM Ca2+ was replaced by 5 mM EGTA, indicating that the binding of these two CaM-binding proteins to CaM was dependent on Ca2+. Less 21-kDa CaM-binding protein was found in culture medium of Angelica dahurica suspension cells; however, a 21-kDa protein was abundant in the cell wall. We believe that the 21-kDa CaM-binding protein is mainly in the cell wall of Angelica dahurica. Based on its reaction with periodic acid-Schiff (PAS) reagent, this 21-kDa protein would appear to be a glycoprotein. The 21-kDa CaM-binding protein was purified by a procedure including Sephadex G-100 gel filtration and CM-Sepharose cation-exchange column chromatography. The purity reached 91% according to gel scanning. The purified 21-kDa CaM-binding protein inhibited the activity of CaM-dependent NAD kinase and the degree of inhibition increased with augmentation of the 21-kDa protein, which appeared to be the typical characteristic of CaM-binding protein.  相似文献   

12.
《Mutation Research Letters》1991,262(3):203-207
In order to elucidate the mechanisms of mutagenic activation of nitrobiphenyls by mammalian activation systems, 2,4,2′,4′-tetranitrobiphenyl was incubated with S9 and its mutagenic metabolites were separated by SiO2 and Al2O3 column chromatography. The most mutagenic diamino-dinitrobiphenyl was isolated from the reaction mixture of 2,4,2′ ,4′-tetranitrobiphenyl with S9 mix at 37°C for 48 h, and its mutability was 4646 revertants/50 ng in Salmonella typhimurium TA98 without S9 mix. The deamination product of this most mutagenic metabolite was identical to 2,4′-dinitrobiphenyl by gas chromatography-mass spectrometry. Therefore, the structure of the metabolite was determined as 2,4′-diamino-2′,4-dinitrobiphenyl by its chemical and physico-chemical properties.  相似文献   

13.
Two new polynuclear complexes of Cu(II), [(μ-1,1,3-N3)2{Cu2(me2tn)2(N3)2}]n (1) (me2tn=2,2-dimethylpropane-1,3-diamine) and [Cu2(μ-C2O4)(μ-N3)(ipr2en)2]n(ClO4)n (2) (ipr2en=N,N-di-isopropylethane-1,2-diamine) have been synthesized and structurally characterized by X-ray crystallography. The crystal structure of 1 displays a 2D network in which distorted octahedral copper(II) centers, chelated by a me2tn ligand and bound to a terminal azide, are connected through μ-1,1,3 bridging azide anions. The structure of 2 shows 1D chains comprising alternating [(ipr2en)Cu-Ox-Cu(ipr2en)] units and end-to-end azide ligand. The chains on mutual H-bonding interaction through ClO4, give rise to a 2D supramolecular architecture. The magnetic data of complexes were recorded in the temperature range, 300-2 K. In case of complex 1, the magnetic data are consistent with a ferromagnetic interaction through the end-on azide bridge (JFM=10 cm−1) and a weak antiferromagnetic interaction (zj=−0.8 cm−1) between the ferromagnetically coupled dimers and an average g-value of 2.05. The susceptibility data of 2 were fitted using an alternating AF-AF chain spin 1/2 law which leads to the following parameters Joxalate=−180 cm−1, Jazide=−43 cm−1 and g=2.25 cm−1.  相似文献   

14.
Analkali tolerant α-l-rhamnosidase has been purified to homogeneity from the culture filtrate of a new fungal strain, Fusarium moniliforme MTCC-2088, using concentration by ultrafiltration and cation exchange chromatography on CM cellulose column. The molecular mass of the purified enzyme has been found to be 36.0 kDa using SDS-PAGE analysis. The Km value using p-nitrophenyl-α-l-rhamnopyranoside as the variable substrate in 0.2 M sodium phosphate buffer pH10.5 at50 °C was 0.50 mM. The catalytic rate constant was15.6 s−1giving the values of kcat/Km is 3.12 × 104M−1 s−1. The pH and temperature optima of the enzyme were 10.5 and 50 °C, respectively. The purified enzyme had better stability at 10 °C in basic pH medium. The enzyme derhamnosylated natural glycosides like naringin to prunin, rutin to isoquercitrin and hesperidin to hesperetin glucoside. The purified α-l-rhamnosidase has potential for enhancement of wine aroma.  相似文献   

15.
Peter Nicholls 《BBA》1975,396(1):24-35
1. Sulphide, like cyanide, is a slow-binding inhibitor of cytochromeaa3 with a high affinity (Kd < 0.1 μM).2. Unlike cyanide binding, the binding of sulphide is apparently independent of the redox state of components of the oxidase other than cytochromea3and shows no anomalous kinetics during complex formation.3. Sulphide binding to cytochrome a33+ is accompanied by a blue-shift in the α-peak of the reduced enzyme (a2+ a33+H2S), similar to but smaller than that induced by azide.4. The reduced sulphide-inhibited system shows a much higher Soret peak at 445 nm than the corresponding cyanide and azide complexes, suggesting that partial electron transfer from sulphide to haem may occur in the complex. No evidence was obtained for the formation of any sulfhaem derivatives of cytochromea3.5. The influence of energization on the spectrum of mitochondrial cytochrome oxidase, and the effects of calcium on the α-peak of isolated cytochromeaa3 (Wikstro¨m, M. K. F. (1974) Ann. N. Y. Acad. Sci. 227, 146–158) are distinct from the action of the cytochromea3ligands.6. A classification of peak shifts in the α-region in terms of isosteric and allosteric ligands is proposed.  相似文献   

16.
A new phlorizin derivative (2′-O-(β-d-glucopyranosyl)-4-azidophloretin, 4-azidophlorizin) has been synthesized and its affinity for the d-glucose, Na+ co-transport system in brush border vesicles from intestinal and renal membranes has been compared with that of phlorizin. The extent of the reversible interaction of the ligand with the transporter in dim light has been evaluated from three separate measurements: (1) Ki, the constant for fully-competitive inhibition of (Na+, Δψ)-dependent d-glucose uptake, (2) Kd, the dissociation constant of 4-azido[3H]phlorizin binding in the presence of an NaSCN inward gradient, and (3) Ki, the constant for fully-competitive inhibition of the specific ((Na+, Δψ)-dependent, d-glucose protectable) high-affinity [3H]phlorizin binding. In experiments with vesicles derived from rat kidney, all three constants (Ki, Kd and Ki) were essentially equal and ranged between 3.2 and 5.2 μM, that is, the azide derivative has almost the same affinity for this transporter as phlorizin itself. On the other hand, compared to phlorizin, the 4-azidophlorizin has a lower affinity for the transporter in vesicles prepared from rabbit; its Ki values are some 15–20-times larger than those determined with rat membranes. However, the affinity of the azide for the sugar transporter in membranes from either the intestine or kidney of the same animal species (rabbit or rat) was essentially the same. In spite of the lower affinity for the transporter in either membrane system from the rabbit, results described elsewhere (Hosang, M., Gibbs, E.M., Diedrich, D.F. and Semenza, G. (1981) FEBS Lett., 130, 244–248) indicate that 4-azidophlorizin is an effective photoaffinity label in this species also. Photolysis of the azide yields a reactive intermediate which reacts with a 72 kDa protein in rabbit intestine brush borders. Covalent labeling of this protein occurred under conditions which suggests that it is (a component of) the glucose transporter.  相似文献   

17.
The cellular slime mold, Dictyostelium discoideum, contains at least two classes of phosphodiesterase activity. One class of enzymes hydrolyses cyclic AMP (cAMP) and cyclic GMP (cGMP) with approximately equal rates. Another enzyme, which is less than 5% of the total activity, specifically hydrolyses cGMP. The cGMP-specific enzyme does not bind to a Con A-Sepharose column, while all the cAMP-hydrolyzing activities are retarded by this column. The cGMP-specific enzyme is activated by low cGMP concentrations (10?8-10?6 M); the enzyme has normal Michaelis-Menten kinetics at high substrate concentrations with a Km of about 3–6 μM. The cGMP-binding sites for activation and for catalysis show different cyclic nucleotide specificity, but they are probably located on one protein with a molecular weight of about 70 000. The enzyme is stable only under specific conditions, and the activation property of the enzyme is lost relatively easy. Irreversible modifications occur at temperatures below 0° and above 30°C, and at pH below 6.0. Several other conditions such as high ion concentrations, temperatures just above 0°C and pH above 8.0 lead to reversibel modifications of enzyme activity.  相似文献   

18.
DNA (calf thymus) was reacted to completion with varying amounts of (195mpt)-cis-diamminedichloroplatinum(II) and then hydrolyzed in formic acid at 110°C for 15 min. The hydrolysate was then applied to an Aminex A6 cation-exchange column and eluted with potassium carbonate (0.01 M, pH 11). For a molar ratio of bound Pt per nucleotide (r) of 0.06 or less, most of the radioactivity eluted in the form of two products identified as (1) a bifunctional homoadduct formed between Pt and two guanine residues, and (2) a bifunctional heteroadduct formed between Pt and a residue each of adenine and guanine. The amount of heteroadduct was about 20% of that of the homoadduct. When r was greater than r = 0.06, several additional peaks were observed, one of which was tentatively identified as a monofunctional adduct of guanine and Pt. For all r values, a portion of the Pt, amounting to $?20%, eluted in the void volume and may reflect a partial breakdown of the adducts during hydrolysis.  相似文献   

19.
Cytochrome c oxidase is a member of the heme-copper family of oxygen reductases in which electron transfer is linked to the pumping of protons across the membrane. Neither the redox center(s) associated with proton pumping nor the pumping mechanism presumably common to all heme-copper oxidases has been established. A possible conformational coupling between the catalytic center (Fea33+–CuB2+) and a protein site has been identified earlier from ligand binding studies, whereas a structural change initiated by azide binding to the protein has been proposed to facilitate the access of cyanide to the catalytic center of the oxidized bovine enzyme. Here we show that cytochrome oxidase pretreated with a low concentration of azide exhibits a significant increase in the apparent rate of cyanide binding relative to that of free enzyme. However, this increase in rate does not reflect a conformational change enhancing the rapid formation of a Fea33+–CN–CuB2+ complex. Instead the cyanide-induced transition of a preformed Fea33+–N3–CuB2+ to the ternary complex of Fea33+–N3 CuB2+–CN is the most likely reason for the observed acceleration. Significantly, the slow rate of azide release from the ternary complex indicates that cyanide ligated to CuB blocks a channel between the catalytic site and the solvent. The results suggest that there is a pathway that originates at CuB and that, during catalysis, ligands present at this copper center control access to the iron of heme a3 from the bulk medium.  相似文献   

20.
To date, the development of photoaffinity ligands targeting the human serotonin transporter (hSERT), a key protein involved in disease states such as depression and anxiety, have been radioisotope-based (i.e., 3H or 125I). This letter instead highlights three derivatives of the selective serotonin reuptake inhibitor (SSRI) (S)-citalopram that were rationally designed and synthesized to contain a photoreactive benzophenone or an aryl azide for protein target capture via photoaffinity labeling and a terminal alkyne or an aliphatic azide for click chemistry-based proteomics. Specifically, clickable benzophenone-based (S)-citalopram photoprobe 6 (hSERT Ki?=?0.16?nM) displayed 11-fold higher binding affinity at hSERT when compared to (S)-citalopram (hSERT Ki?=?1.77?nM), and was subsequently shown to successfully undergo tandem photoaffinity labeling-biorthogonal conjugation using purified hSERT. Given clickable photoprobes can be used for various applications depending on which reporter is attached by click chemistry subsequent to photoaffinity labeling, photoprobe 6 is expected to find value in structure-function studies and other research applications involving hSERT (e.g., imaging).  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