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.
Azide mutagenicity involves the requisite formation of the putative novel aminoacid metabolite, beta-azidoalanine. The role of this metabolite, however, is unclear. In order to confirm the identity of this metabolite and provide additional information on possible stereochemical requirements for mutagenicity, authentic racemic and L-azidoalanine were synthesized by an unambiguous route and tested for mutagenicity in Salmonella typhimurium TA100, TA1535, hisG46 and Escherichia coli WP2-. A marked antipodal potency ratio was observed in strains TA100 and TA1535 when racemic and L-azidoalanine were compared. The mutagenic activity resided primarily in the L-isomer. The molar potency of L-azidoalanine in TA100 and TA1535 was nearly identical to that of azide. The lack of mutagenic response for racemic or L-azidoalanine in hisG46 and E. coli WP2- was like that reported for azide and is consistent with similar modes of action for these agents.  相似文献   

2.
Azide is metabolized to the proximate mutagen, L-azidoalanine in bacterial systems. While this novel mutagenic metabolite plays a key role in azide mutagenesis, the biochemistry of this role is unknown. The chemical synthesis of authentic racemic azidoalanine and several derivatives thereof allowed the exploration of structure-activity relationships with this unique mutagen. We found that whereas azide, azidoalanine and azidoalanine tert.-butyl ester were of comparable mutagenic potency, derivatives which lack the free amino group, such as azidopropionic acid and amino-blocked azidoalanine, were orders of magnitude less active. These findings demonstrate that the free amino group is essential for significant activity, while the carboxyl group may be less important. This conclusion together with the finding that DL-azidoalanine is a less potent mutagen than azide itself, suggests that the metabolite, while necessary for azide mutagenicity, may not be the ultimate mutagenic species. Instead, the data are consistent with the hypothesis that azidoalanine requires further bioactivation.  相似文献   

3.
Sadiq MF  Owais WM 《Mutation research》2000,469(2):253-257
The mutagenic and toxic activities of sodium azide (NaN(3) ) and its organic metabolite L-azidoalanine [N(3)-CH(2)-CH(NH)(2)-COOH] were examined in the different stages of spermatogenesis in Drosophila melanogaster. Both azide and azidoalanine were toxic to the injected males, but azidoalanine was significantly less toxic than sodium azide. Following the injection with 0.2 microl of these compounds in the hemocoel of young adult wild-type males, the minimum concentrations of these compounds with complete toxic effects (zero survival) were 40 mM sodium azide and 160 mM azidoalanine. Sex-linked recessive lethals were scored by the Muller-5 method in three successive broods, representing sperms (brood A), spermatids (brood B), and a compiled group of meiotic and premeiotic germ cell stages (brood C). The results provide strong experimental evidence that azidoalanine is significantly (p<0.01) mutagenic to all stages of spermatogenesis in Drosophila melanogaster. Sodium azide, however, was not significantly (p>0.05) mutagenic and did not increase the rate of sex-linked recessive lethals over those produced by the control group injected with 0.45% NaCl. These results indicate the requirement of metabolic activation of azide in Drosophila as a prerequisite for its mutagenic effects.  相似文献   

4.
The L- and D-isomers of azidoalanine (azide metabolite) have been chemically synthesized with 60% yield using corresponding N-(tert-butoxycarbonyl)-serine as starting materials. The mutagenic properties of synthesized L-azidoalanine are very similar to those of azide and in vivo synthesized azidoalanine. Synthetic D-azidoalanine shows very low mutagenic activity on Salmonella typhimurium TA1530 strain compared to that of the L-isomer. Thus a stereoselective process is involved in azidoalanine mutagenicity. The data presented in this study suggest that further biochemical activation is required for L-azidoalanine to produce its mutagenic activity.  相似文献   

5.
The Bacillus subtilis ATCC 6633 strain synthesizes a mutagenic metabolite from sodium azide and O-acetylserine. Mutagenicity of azide was decreased in growth media containing 10(-4) M glutathione, L-cysteine or L-djenkolic acid whereas dithiothritol (DTT) added at the same concentration did not reduce the mutagenicity of azide. Likewise, glutathione, L-cysteine, L-djenkolic acid, and DTT were found to have no effect in reducing the mutagenicity of the in vitro produced metabolite using bacterial cell-free extract. These results suggest that O-acetyl-serine sulfhydrylase catalyzes the reaction of azide and O-acetylserine to form a mutagenic metabolite, which is ninhydrin positive and migrates in TLC to an Rf value similar to that of azidoalanine in both acidic and basic solvent systems.  相似文献   

6.
A mutagenic azide metabolite was purified from the medium in which Salmonella typhimurium cells were grown in the presence of azide. This metabolite was identified to be azidoalanine based on infrared and mass spectroscopy and elemental analysis. This compound appeared to be identical to the mutagenic compound synthesized in vitro from azide and O-acetylserine by partially purified O-acetylserine sulfhydrylase. The metabolite (azidoalanine) mutagenic efficiency and spectrum in S. typhimurium was similar to that of inorganic azide. The compounds 2-azidoethylamine, 2-bromoethylamine, 3-bromopropionic acid and N-(azidomethyl) phthalimide were also mutagenic with a similar spectrum to azide and azidoalanine, but with lower efficiency. The compounds 3-azidopropylamine, 4-azidobutylamine, 3-chloroalanine and ethylamine were only weakly or nonmutagenic. Numerous other chloro, bromo and azido phthalimide derivatives tested were nonmutagenic. It is suggested that the lack of azide mutagenicity (and perhaps carcinogenicity) in mammalian cells may be due to their inability to convert azide to azidoalanine.  相似文献   

7.
Sodium azide is a potent mutagen of maize (Zea mays L.) kernels that may have potential as a point mutagen for inducing biochemical mutations in maize tissue cultures. Azide mutagenicity was evaluated in friable, embryogenic maize callus and a nonregenerable maize suspension culture by determining the number of resistant variant cell lines able to grow on media containing inhibitory concentrations of lysine plus threonine (LT). The number of LT-resistant variants selected from either culture type did not increase in response to azide treatment. In addition, there was no increase in somatic mutations in more than 100 plants regenerated from azide treated LT-resistant lines. The levels of mutagenic metabolite of azide (presumably azidoalanine), were determined by bioassay in the two azide-treated maize callus types and compared to levels of mutagenic metabolite in embryos isolated from azide-treated kernels. The two types of maize tissue cultures and isolated embryos contained similar levels of mutagenic metabolite 4 h after azide treatment indicating similar uptake and conversion of azide to mutagenic metabolite in the three tissues. Mutagenic metabolite in azide-treated embryos did not significantly decrease after 40 h. However, mutagenic metabolite levels in both azide-treated tissue cultures decreased to near background levels within 20 h providing evidence for rapid metabolism of the azide mutagenic metabolite. The lack of evidence for azide mutagenicity in maize callus and its known potent mutagenicity in kernels appears to be associated with specific differences in azide metabolism between callus tissues and kernel embryos.  相似文献   

8.
Sodium azide is unique among mutagens. It is highly mutagenic in many plant and bacterial species but marginally mutagenic in mammalian cells. A possible explanation for this difference in mutagenic efficiency may lie in the inability of mammalian cells to convert azide to the putative ultimate mutagen. Normal human fibroblasts and Chinese hamster cells or cell-free extracts from these cell lines were treated with azide and the sonicates tested for mutagenicity in Salmonella strain TA1530. The data suggest that neither cell line was capable of converting azide to a mutagenic intermediate. In addition, both cell lines expressed the enzyme O-acetylserine(thio)-lyase which is responsible for the conversion of azide to azidoalanine, the putative mutagenic intermediate. Although mammalian cells possess the enzyme responsible for the conversion of azide to azidoalanine, they appear incapable of converting azide into a mutagenic intermediate in appreciable quantities. Further, the data support the conclusion that azide may be further modified in mammalian cells to an intermediate that is not genotoxic.  相似文献   

9.
Azide mutagenicity in susceptible non-mammalian systems involves the requisite formation of L-azidoalanine, a novel mutagenic amino acid. The biochemical mechanism(s) of azidoalanine-induced mutagenesis, however, is not known. Previous studies of the structural requirements for azidoalanine mutagenicity suggested the importance of free L-amino acid character, and that bioactivation of azidoalanine to the ultimate mutagenic species is required. To gain more insight into possible enzymatic processing, the alpha-methyl analogue, alpha-methyl-azidoalanine, and the homologue, 2-amino-4-azidobutanoic acid, were synthesized and tested for mutagenic potency in Salmonella typhimurium strain TA1530. In addition, azidoacetic acid, a possible azidoalanine metabolite, was prepared and tested. The results show that alpha-methyl substitution effectively blocks the mutagenic effects of azidoalanine with alpha-methyl-azidoalanine being nearly devoid of mutagenic activity. In contrast, homologation of azidoalanine to yield 2-amino-4-azidobutanoic acid produces a marked increase in molar mutagenic potency. As with azidoalanine, the mutagenic activity of this homologue is associated with the L-isomer. Azidoacetic acid, however, was only very weakly mutagenic when tested as either the free acid or ethyl ester. This low mutagenic potency may indicate that bioactivation does not involve the entry of azide-containing azidoalanine catabolite into the Kreb's cycle. The high potency of 2-amino-4-azidobutanoic acid may be indicative of more efficient bioactivation and/or greater intrinsic activity. Importantly, the latter finding clearly shows that potent azido-amino acid mutagenicity is not limited to azidoalanine alone.  相似文献   

10.
Klebsiella pneumoniae nitrogenase reduced azide, at 30 degrees C and pH 6.8-8.2, to yield ammonia (NH3), dinitrogen (N2) and hydrazine (N2H4). Reduction of (15N = 14N = 14N)-followed by mass-spectrometric analysis showed that no new nitrogen-nitrogen bonds were formed. During azide reduction, added 15N2H4 did not contribute 15N to NH3, indicating lack of equilibration between enzyme-bound intermediates giving rise to N2H4 and N2H4 in solution. When azide reduction to N2H4 was partially inhibited by 15N2, label appeared in NH3 but not in N2H4. Product balances combined with the labelling data indicate that azide is reduced according to the following equations: (formula: see text); N2 was a competitive inhibitor and CO a non-competitive inhibitor of azide reduction to N2H4. The percentage of total electron flux used for H2 evolution concomitant with azide reduction fell from 26% at pH 6.8 to 0% at pH 8.2. Pre-steady-state kinetic data suggest that N2H4 is formed by the cleavage of the alpha-beta nitrogen-nitrogen bond to bound azide to leave a nitride (= N) intermediate that subsequently yields NH3.  相似文献   

11.
The role of copper in pig kidney diamine oxidase has been probed by examining the effects of potential Cu(II) ligands on the spectroscopic and catalytic properties of the enzyme. In the presence of azide and thiocyanate, new absorption bands are evident at 410 nm (epsilon = 6300 M-1 cm-1) and 365 nm (epsilon = 3000 M-1 cm-1), respectively. These bands are assigned as ligand-to-metal charge-transfer transitions, N3-/SCN- leads to Cu(II). One anion/Cu(II) is coordinated in an equitorial position. Anion binding can be completely reversed by dialysis. The equilibrium constants for diamine oxidase-anion complex formation are 134 M-1 (N3-) and 55 M-1 (SCN-). Azide and thiocyanate are linear uncompetitive inhibitors with respect to the amine substrate when O2 is present at saturating concentrations. Taken together, the data are consistent with a functional role for Cu(II) in diamine oxidase catalysis.  相似文献   

12.
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   

13.
Azide (10?3 M, solution buffered at pH 3) is more effective in inducing mutations in embryonic shoots of seeds germinated between 8 and 16 h than in non-germinated seeds and in seeds germinated between 0 and 8 h and 16 to 28 h. This peak of chlorophyll-deficient seedling mutation frequency coincides with maximum frequencies of seedling lethals and DNA replication in the cells of the embryonic shoot. The mutation data suggest azide may only act on replicating DNA.Azide induced no chromosome-aberration frequencies significantly above controls in (1) embryonic shoots of barley seeds germinated for 8–12 h, (2) microsores of barley and (3) human leukocytes. It appears to be a point-mutation mutagen.  相似文献   

14.
Azide binding to the blue copper oxidases laccase and ascorbate oxidase (AO) was investigated by electron paramagnetic resonance (EPR) and pulsed electron-nuclear double resonance (ENDOR) spectroscopies. As the laccase : azide molar ratio decreases from 1:1 to 1:7, the intensity of the type 2 (T2) Cu(II) EPR signal decreases and a signal at g approximately 1.9 appears. Temperature and microwave power dependent EPR measurements showed that this signal has a relatively short relaxation time and is therefore observed only below 40 K. A g approximately 1.97 signal, with similar saturation characteristics was found in the AO : azide (1:7) sample. The g < 2 signals in both proteins are assigned to an S = 1 dipolar coupled Cu(II) pair whereby the azide binding disrupts the anti-ferromagnetic coupling of the type 3 (T3) Cu(II) pair. Analysis of the position of the g < 2 signals suggests that the distance between the dipolar coupled Cu(II) pair is shorter in laccase than in AO. The proximity of T2 Cu(II) to the S = 1 Cu(II) pair enhances its relaxation rate, reducing its signal intensity relative to that of native protein. The disruption of the T3 anti-ferromagnetic coupling occurs only in part of the protein molecules, and in the remaining part a different azide binding mode is observed. The 130 K EPR spectra of AO and laccase with azide (1:7) exhibit, in addition to an unperturbed T2 Cu(II) signal, new features in the g parallel region that are attributed to a perturbed T2 in protein molecules where the anti-ferromagnetic coupling of T3 has not been disrupted. While these features are also apparent in the AO : azide sample at 10 K, they are absent in the EPR spectra of the laccase : azide sample measured in the range of 6-90 K. Moreover, pulsed ENDOR measurements carried out at 4.2 K on the latter exhibited only a reduction in the intensity of the 20 MHz peak of the 14N histidine coordinated to the T2 Cu(II) but did not resolve any significant changes that could indicate azide binding to this ion. The lack of T2 Cu(II) signal perturbation below 90 K in laccase may be due to temperature dependence of the coupling within the trinuclear : azide complex.  相似文献   

15.
1. The initial rapid phase of ATP hydrolysis by bovine heart submitochondrial particles or by soluble F1-ATPase is insensitive to anion activation (sulphite) or inhibition (azide). 2. The second slow phase of ATP hydrolysis is hyperbolically inhibited by azide (Ki approximately 10(-5) M); the inosine triphosphatase activity of submitochondrial particles or F1-ATPase is insensitive to azide or sulphite. 3. The rate of interconversion between rapid azide-insensitive and slow azide-sensitive phases of ATP hydrolysis does not depend on azide concentration, but strongly depends on ATP concentration. 4. Sulphite prevents the interconversion of the rapid initial phase of the reaction into the slower second phase, and also prevents and slowly reverses the inhibition by azide. 5. The presence of sulphite in the mixture when ADP reacts with ATPase of submitochondrial particles changes the pattern of the following activation process. 6. Azide blocks the activation of ATP-inhibited ATPase of submitochondrial particles by phosphoenolpyruvate and pyruvate kinase. 7. The results obtained suggest that the inhibiting effect of azide on mitochondrial ATPase is due to stabilization of inactive E*.ADP complex formed during ATP hydrolysis; the activation of ATPase by sulphite is also realized through the equilibrium between intermediate active E.ADP complex and inactive E*.ADP complex.  相似文献   

16.
The gene coding for O-acetylserine sulfhydrylase (OASS) from E. coli K12 was cloned into the vector pBR322 plasmid and expressed in a cysk mutant strain of E. coli that is deficient in O-acetylserine sulfhydrylase (OASS-). The clone containing the OASS gene was selected by using tetracycline-ammonium bismuth citrate medium. Retransformation of the hybrid plasmid into competent cysk mutant cells resulted in the recovery of a clone containing normal levels of O-acetylserine sulfhydrylase. Negative selection of retransformed cysk cells on 1,2,4-triazole plates resulted in the complete inhibition of growth indicating the presence of a functional OASS gene. The ability of the new clone to convert azide to its mutagenic metabolite was tested. Cultures of the clone cells containing significant levels of OASS activity were able to produce a mutagenic product from azide and O-acetylserine as tested on Salmonella typhimurium TA1530. This cloning method could be applied also to clone the same gene from eukaryotic sources.  相似文献   

17.
Fisher K  Dilworth MJ  Newton WE 《Biochemistry》2006,45(13):4190-4198
The Mo-nitrogenase-catalyzed reduction of both cyanide and azide results in the production of excess NH3, which is an amount of NH3 over and above that expected to be formed from the well-recognized reactions. Several suggestions about the possible sources of excess NH3 have been made, but previous attempts to characterize these reactions have met with either limited (or no) success or controversy. Because V-nitrogenase has a propensity to release partially reduced intermediates, e.g., N2H4 during N2 reduction, it was selected to probe the reduction of cyanide and azide. Sensitive assay procedures were developed and employed to monitor the production of either HCHO or CH3OH (its further two-electron-reduced product) from HCN. Like Mo-nitrogenase, V-nitrogenase suffered electron-flux inhibition by CN- (but was much less sensitive than Mo-nitrogenase), but unlike the case for Mo-nitrogenase, MgATP hydrolysis was also inhibited by CN-. V-Nitrogenase also released more of the four-electron-reduced intermediate, CH3NH2, than did Mo-nitrogenase. At high NaCN concentrations, V-nitrogenase directed a significant percentage of electron flux into excess NH3, and under these conditions, substantial amounts of HCHO, but no CH3OH, were detected for the first time. With azide, in contrast to the case for Mo-nitrogenase, both total electron flux and MgATP hydrolysis with V-nitrogenase were inhibited. V-Nitrogenase, unlike Mo-nitrogenase, showed no preference between the two-electron reduction to N2-plus-NH3 and the six-electron reduction to N2H4-plus-NH3. V-Nitrogenase formed more excess NH3, but reduction of the N2 produced by the two-electron reduction of N3(-) was not its source. Rather, it was formed directly by the eight-electron reduction of N3(-). Unlike Mo-nitrogenase, CO could not completely eliminate either cyanide or azide reduction by V-nitrogenase. CO did, however, eliminate the inhibition of both electron flux and MgATP hydrolysis by CN-, but not that caused by azide. These different responses to CO suggest different sites or modes of interaction for these two substrates with V-nitrogenase.  相似文献   

18.
This report describes and compares different strategies to deactivate (endcap) epoxide groups and azide groups on bio-chromatographic support surfaces, before and after ligand attachment. Adsorbents possessing epoxide groups were deactivated using acidic hydrolysis or were endcapped with 2-mercaptoethanol or 2-ethanolamine. The influence of surface-bound 2-ethanolamine was demonstrated for the triazine-type affinity adsorbent B14-2LP-FractoAIMs-1, which was tested in combination with the weak anion exchange material 3-aminoquinuclidine-FractoAIMs-3 (AQ-FA3). Azide groups were modified with 2-propargylalcohol using Click-Chemistry. Besides the conventional one-pot Click reaction, an alternative approach was introduced. This optimized Click protocol was employed (i) for the preparation of the weak anion exchange material AdQ-triazole-Fractogel (AdQ-TRZ-FG) and (ii) for the endcapping of residual azide groups with 3-propargyl alcohol. Using the new Click reaction protocol the ligand immobilization rate was doubled from 250 to 500 μmol/g dry adsorbent. Furthermore, the modified support surface was proven to be inert towards the binding of immunoglobulin G (IgG) as well as feed impurities. A thorough evaluation of modified surfaces and adsorbents was performed with dynamic binding experiments using cell culture supernatant containing monoclonal human immunoglobulin G (h-IgG-1). Besides SDS-Page, a recently introduced Protein A-size exclusion HPLC method (PSEC-HPLC) was used to visualize the feed impurity composition and the IgG content of all collected sample fractions in simple PSEC-Plots. A surprising outcome of this study was the irreversible binding of IgG to azide modified surfaces. It was found that organic azide compounds, e.g. 1-azide-3-(2-propen-1-yloxy)-2-propanol (AGE-N3) promote antibody aggregation to a slightly higher extent than the inorganic sodium azide. The possibility that the Hofmeister Series of salt anions may be applicable to predict the properties of the corresponding organic compounds is discussed.  相似文献   

19.
It has been found that the N(P, R)-type intermediate of the photocycle is formed in the Asp-96-->Asn mutant at acidic pH. Azide, which strongly activates the M decay in this mutant, allows the N intermediate to be shown also at neutral pH. Under these conditions mutant N decays in a pH-independent fashion. In the presence of azide, the H+ uptake by Asp-96-->Asn mutant bacteriorhodopsin follows the M decay, whereas the N decay occurs at a much slower rate. Two electrogenic stages have been shown to be associated with the M--->bR step in the Asp-96--->Asn mutant photocycle. The faster and slower stages correlate with the M--->N and N--->bR transitions, respectively. In the Asp-96--->Asn mutant, high concentrations of azide are found to increase the M decay rate up to the values higher than those in the wild-type protein, both with or without azide. Such an effect is absent for the Asp-96-->Glu mutant. The activation energies for M--->N and N--->bR transitions in the wild-type protein are equal to 18 and 19 kcal x mole-1, respectively. In the Asp-96-->Asn mutant without azide, the activation energy of the M decay is only 5 kcal x mole-1, whereas in the presence of azide in this mutant the activation energies for M and N decays are 8 and 9 kcal x mole-1, respectively. A scheme of events accompanying the Schiff base reprotonation during the photocycle is discussed.  相似文献   

20.
Tolland JD  Thorneley RN 《Biochemistry》2005,44(27):9520-9527
Stopped-flow FTIR spectroscopy was used to monitor continuously the pre-steady- and steady-state phases of azide reduction by nitrogenase and the accompanying hydrolysis of ATP. This was characterized by a ca. 1.3 s lag phase that is explained by the number of Fe protein cycles required to effect the reductions of azide to N(2) + NH(3), N(2)H(4) + NH(3), or 3NH(3). Extrapolation of the steady-state time course for azide reduction to zero time showed that one azide binds within 200 ms to each FeMo cofactor. Inhibition of azide reduction by CO was established at times <400 ms, which was faster than the appearance of the first observable IR band assigned to CO (1904 cm(-)(1) detectable at ca. 1 s with maximum amplitude at ca. 7 s). IR bands associated with the rapidly formed (<400 ms) CO species that inhibits azide reduction were not observed over the range 1700-2100 cm(-)(1). This suggests either that the CO is initially bridging two or more Fe atoms or that a rapid reduction of CO to a formyl state occurs by insertion into a metal-hydride bond. The frequencies and time courses for the appearance and loss of the CO bands under hi- and lo-CO conditions were essentially unaffected by the presence of 20 mM azide, consistent with CO being a noncompetitive inhibitor of azide reduction and with azide and CO binding to different sites on the FeMo cofactor.  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