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 62 毫秒
1.
The activities of pure and mixed cultures of Desulfovibrio vulgaris and Methanosarcina barkeri in the exponential growth phase were monitored by measuring changes in dissolved-gas concentration by membrane-inlet mass spectrometry. M. barkeri grown under H2-CO2 or methanol produced limited amounts of methane and practically no hydrogen from either substrate. The addition of CO resulted in a transient H2 production concomitant with CO consumption. Hydrogen was then taken up, and CH4 production increased. All these events were suppressed by KCN, which inhibited carbon monoxide dehydrogenase activity. Therefore, with both substrates, H2 appeared to be an intermediate in CO reduction to CH4. The cells grown on H2-CO2 consumed 4 mol of CO and produced 1 mol of CH4. Methanol-grown cells reduced CH3OH with H2 resulting from carbon monoxide dehydrogenase activity, and the ratio was then 1 mol of CH4 to 1 mol of CO. Only 12CH4 and no 13CH4 was obtained from 13CO, indicating that CO could not be the direct precursor of CH4. In mixed cultures of D. vulgaris and M. barkeri on lactate, an initial burst of H2 was observed, followed by a lower level of production, whereas methane synthesis was linear with time. Addition of CO to the mixed culture also resulted in transient extra H2 production but had no inhibitory effect upon CH4 formation, even when the sulfate reducer was D. vulgaris Hildenborough, whose periplasmic iron hydrogenase is very sensitive to CO. The hydrogen transfer is therefore probably mediated by a less CO-sensitive nickel-iron hydrogenase from either of both species.  相似文献   

2.
The activities of pure and mixed cultures of Desulfovibrio vulgaris and Methanosarcina barkeri in the exponential growth phase were monitored by measuring changes in dissolved-gas concentration by membrane-inlet mass spectrometry. M. barkeri grown under H2-CO2 or methanol produced limited amounts of methane and practically no hydrogen from either substrate. The addition of CO resulted in a transient H2 production concomitant with CO consumption. Hydrogen was then taken up, and CH4 production increased. All these events were suppressed by KCN, which inhibited carbon monoxide dehydrogenase activity. Therefore, with both substrates, H2 appeared to be an intermediate in CO reduction to CH4. The cells grown on H2-CO2 consumed 4 mol of CO and produced 1 mol of CH4. Methanol-grown cells reduced CH3OH with H2 resulting from carbon monoxide dehydrogenase activity, and the ratio was then 1 mol of CH4 to 1 mol of CO. Only 12CH4 and no 13CH4 was obtained from 13CO, indicating that CO could not be the direct precursor of CH4. In mixed cultures of D. vulgaris and M. barkeri on lactate, an initial burst of H2 was observed, followed by a lower level of production, whereas methane synthesis was linear with time. Addition of CO to the mixed culture also resulted in transient extra H2 production but had no inhibitory effect upon CH4 formation, even when the sulfate reducer was D. vulgaris Hildenborough, whose periplasmic iron hydrogenase is very sensitive to CO. The hydrogen transfer is therefore probably mediated by a less CO-sensitive nickel-iron hydrogenase from either of both species.  相似文献   

3.
Cell suspensions of methanogenic bacteria (Methanosarcina barkeri, Methanospirillum hungatei, Methano-brevibacter arboriphilus, and Methanobacterium thermoautotrophicum) were found to form CO from CO2 and H2 according to the reaction: CO2 + H2----CO + H2O; delta G0 = +20 kJ/mol. Up to 15,000 ppm CO in the gas phase were reached which is significantly higher than the equilibrium concentration calculated from delta G0 (95 ppm under the experimental conditions). This indicated that CO2 reduction with H2 to CO is energy-driven and indeed the cells only generated CO when forming CH4. The coupling of the two reactions was studied in more detail with acetate-grown cells of M. barkeri using methanogenic substrates. The effects of the protonophore tetrachlorosalicylanilide (TCS) and of the proton-translocating ATPase inhibitor N,N'-dicyclohexylcarbodiimide (cHxN)2C were determined. TCS completely inhibited CO formation from CO2 and H2 without affecting methanogenesis from CH3OH and H2. In the presence of the protonophore the proton motive force delta p and the intracellular ATP concentration were very low. (cHxN)2C, which partially inhibited methanogenesis from CH3OH and H2, had no effect on CO2 reduction to CO. In the presence of (cHxN)2C delta p was high and the intracellular ATP content was low. These findings suggest that the endergonic formation of CO from CO2 and H2 is coupled to the exergonic formation of CH4 from CH3OH and H2 via the proton motive force and not via ATP. CO formation was not stimulated by the addition of sodium ions.  相似文献   

4.
Ammonia oxidizers (family Nitrobacteraceae) and methanotrophs (family Methylococcaceae) oxidize CO and CH4 to CO2 and NH4+ to NO2-. However, the relative contributions of the two groups of organisms to the metabolism of CO, CH4, and NH4+ in various environments are not known. In the ammonia oxidizers, ammonia monooxygenase, the enzyme responsible for the conversion of NH4+ to NH2OH, also catalyzes the oxidation of CH4 to CH3OH. Ammonia monooxygenase also mediates the transformation of CH3OH to CO2 and cell carbon, but the pathway by which this is done is not known. At least one species of ammonia oxidizer, Nitrosococcus oceanus, exhibits a Km for CH4 oxidation similar to that of methanotrophs. However, the highest rate of CH4 oxidation recorded in an ammonia oxidizer is still five times lower than rates in methanotrophs, and ammonia oxidizers are apparently unable to grow on CH4. Methanotrophs oxidize NH4+ to NH2OH via methane monooxygenase and NH4+ to NH2OH via methane monooxygenase and NH2OH to NO2- via an NH2OH oxidase which may resemble the enzyme found in ammonia oxidizers. Maximum rates of NH4+ oxidation are considerably lower than in ammonia oxidizers, and the affinity for NH4+ is generally lower than in ammonia oxidizers. NH4+ does not apparently support growth in methanotrophs. Both ammonia monooxygenase and methane monooxygenase oxidize CO to CO2, but CO cannot support growth in either ammonia oxidizers or methanotrophs. These organisms have affinities for CO which are comparable to those for their growth substrates and often higher than those in carboxydobacteria. The methane monooxygenases of methanotrophs exist in two forms: a soluble form and a particulate form. The soluble form is well characterized and appears unrelated to the particulate. Ammonia monooxygenase and the particulate methane monooxygenase share a number of similarities. Both enzymes contain copper and are membrane bound. They oxidize a variety of inorganic and organic compounds, and their inhibitor profiles are similar. Inhibitors thought to be specific to ammonia oxidizers have been used in environmental studies of nitrification. However, almost all of the numerous compounds found to inhibit ammonia oxidizers also inhibit methanotrophs, and most of the inhibitors act upon the monooxygenases. Many probably exert their effect by chelating copper, which is essential to the proper functioning of some monooxygenases. The lack of inhibitors specific for one or the other of the two groups of bacteria hampers the determination of their relative roles in nature.  相似文献   

5.
Prebiotic synthesis in atmospheres containing CH4, CO,and CO2   总被引:2,自引:0,他引:2  
The prebiotic synthesis of organic compounds using a spark discharge on various simulated primitive earth atmospheres at 25 degrees C has been studied. Methane mixtures contained H2 + CH4 + H2O + N2 + NH3 with H2/CH4 molar ratios from 0 to 4 and pNH3 = 0.1 torr. A similar set of experiments without added NH3 was performed. The yields of amino acids (1.2 to 4.7% based on the carbon) are approximately independent of the H2/CH4 ratio and whether NH3 was present, and a wide variety of amino acids are obtained. Mixtures of H2 + CO + H2O + N2 and H2 + CO2 + H2O + N2, with and without added NH3, all gave about 2% yields of amino acids at H2/CO and H2/CO2 ratios of 2 to 4. For a H2/CO2 ratio of 0, the yield of amino acids is extremely low (10(-3)%). Glycine is almost the only amino acid produced from CO and CO2 model atmospheres. These results show that the maximum yield is about the same for the three carbon sources at high H2/carbon ratios, but that CH4 is superior at low H2/carbon ratios. In addition, CH4 gives a much greater variety of amino acids than either CO or CO2. If it is assumed that an abundance of amino acids more complex than glycine was required for the origin of life, then these results indicate the requirement for CH4 in the primitive atmosphere.  相似文献   

6.
Anaerobic metabolism of immediate methane precursors in Lake Mendota.   总被引:10,自引:10,他引:0       下载免费PDF全文
Lake Mendota sediments and the immediate overlying water column were studied to better understand the metabolism of the methanogenic precursors H2/CO2 and acetate in nature. The pool size of acetate (3.5 microns M) was very small, and the acetate turnover time (0.22h) was very rapid. The dissolved inorganic carbon pool was shown to be large (6.4 to 8.3 mM), and the turnover time was slow (111 H.). CO2 was shown to account for 41 +/- 5.5% of the methane produced in sediment. Acetate and H2/CO2 were simultaneously converted to CH4. The addition of H2 to sediments resulted in an increase specific activity of CH4 from H(14)CO3- and a decrease in specific activity of CH4 from [2-14C]acetate. Acetate addition resulted in a decrease in specific activity of CH4 from H(14)CO3-. The metabolism of H(14)CO3- or [2-14C]acetate to 14CH4 was not inhibited by addition of acetate or H2. After greater than 99% of added [2-14C]acetate had been turned over, 42% of the label was recovered as 14CH4 20% was recovered as 14CO2 and 38% was incorporated into sediment. Inhibitor studies of [2-14C]acetate metabolism in sediments demonstrated that CHCl3 completely inhibited CH4 formation, but not CO2 production. Air and nitrate addition inhibited CH4 formation and stimulated CO2 production, whereas fluoroacetate addition totally inhibited acetate metabolism. The oxidation of [2-14C]acetate to 14CO2 was shown to decrease with time when sediment was incubated before the addition of label, suggesting depletion of low levels of an endogenous sediment electron acceptor. Acetate metabolism varied seasonally and was related to the concentration of sulfate in the lake and interstitial water. Methanogenesis occurred in the sediment and in the water immediately overlying the sediment during period of lake stratification and several centimeters below the sediment-water interface during lake turnovers. These data indicate that methanogenesis in Lake Mendota sediments was limited by "immediate" methane precursor availability (i.e., acetate and H2), by competition for these substrates by nonmethanogens, and by seasonal variations which altered sediment and water chemistry.  相似文献   

7.
Anaerobic metabolism of immediate methane precursors in Lake Mendota.   总被引:8,自引:0,他引:8  
Lake Mendota sediments and the immediate overlying water column were studied to better understand the metabolism of the methanogenic precursors H2/CO2 and acetate in nature. The pool size of acetate (3.5 microns M) was very small, and the acetate turnover time (0.22h) was very rapid. The dissolved inorganic carbon pool was shown to be large (6.4 to 8.3 mM), and the turnover time was slow (111 H.). CO2 was shown to account for 41 +/- 5.5% of the methane produced in sediment. Acetate and H2/CO2 were simultaneously converted to CH4. The addition of H2 to sediments resulted in an increase specific activity of CH4 from H(14)CO3- and a decrease in specific activity of CH4 from [2-14C]acetate. Acetate addition resulted in a decrease in specific activity of CH4 from H(14)CO3-. The metabolism of H(14)CO3- or [2-14C]acetate to 14CH4 was not inhibited by addition of acetate or H2. After greater than 99% of added [2-14C]acetate had been turned over, 42% of the label was recovered as 14CH4 20% was recovered as 14CO2 and 38% was incorporated into sediment. Inhibitor studies of [2-14C]acetate metabolism in sediments demonstrated that CHCl3 completely inhibited CH4 formation, but not CO2 production. Air and nitrate addition inhibited CH4 formation and stimulated CO2 production, whereas fluoroacetate addition totally inhibited acetate metabolism. The oxidation of [2-14C]acetate to 14CO2 was shown to decrease with time when sediment was incubated before the addition of label, suggesting depletion of low levels of an endogenous sediment electron acceptor. Acetate metabolism varied seasonally and was related to the concentration of sulfate in the lake and interstitial water. Methanogenesis occurred in the sediment and in the water immediately overlying the sediment during period of lake stratification and several centimeters below the sediment-water interface during lake turnovers. These data indicate that methanogenesis in Lake Mendota sediments was limited by "immediate" methane precursor availability (i.e., acetate and H2), by competition for these substrates by nonmethanogens, and by seasonal variations which altered sediment and water chemistry.  相似文献   

8.
Yields based on carbon are usually reported in prebiotic experiments, while energy yields (moles cal-1) are more useful in estimating the yields of products that would have been obtained from the primitive atmosphere of the earth. Energy yields for the synthesis of HCN and H2CO from a spark discharge were determined for various mixtures of CH4, CO, CO2, H2, H2O, N2 and NH3. The maximum yields of HCN and H2CO from CH4, CO, and CO2 as carbon sources are about 4 X 10(-8) moles cal-1.  相似文献   

9.
A bioorganometallic approach to malaria therapy led to the discovery of ferroquine (FQ, SSR97193). To assess the importance of the electronic properties of the ferrocenyl group, cyclopentadienyltricarbonylrhenium analogues related to FQ, were synthesized. The reaction of [N-(7-chloro-4-quinolinyl)-1,2-ethanodiamine] with the cyrhetrenylaldehyde complexes (η(5)-C(5)H(4)CHO)Re(CO)(3) and [η(5)-1,2-C(5)H(3)(CH(2)OH)(CHO)]Re(CO)(3) produces the corresponding imine derivatives [η(5)-1,2-C(5)H(3)(R)(CHN-CH(2)CH(2)NH-QN)]Re(CO)(3) R=H 3a; R=CH(2)OH 3b; QN=N-(7-Cl-4-quinolinyl). Reduction of 3a and 3b with sodium borohydride in methanol yields quantitatively the amine complexes [η(5)-1,2-C(5)H(3)(R)(CH(2)-NH-CH(2)CH(2)NH-QN)]Re(CO)(3) R=H 4a; R=CH(2)OH 4b. To establish the role of the cyrethrenyl moiety in the antimalarial activity of this series, purely organic parent compounds were also synthesized and tested. Evaluation of antimalarial activity measured in vitro against the CQ-resistant strains (W2) and the CQ-susceptible strain (3D7) of Plasmodium falciparum indicates that these cyrhetrene conjugates are less active compared to their ferrocene and organic analogues. These data suggest an original mode-of-action of FQ and ferrocenyl analogues in relationship with the redox pharmacophore.  相似文献   

10.
11.
The stimulation of carbon dioxide reduction to methane by addition of 2-(methylthio)ethanesulfonate (CH3-S-CoM) to cell extracts of Methanobacterium thermoautotrophicum was investigated. Similar stimulation of CO2 reduction by CH3-S-CoM was found for cell extracts of Methanobacterium bryantii and Methanospirillum hungatei. The CH3-S-CoM requirement could be met by the methanogenic precursors formaldehyde, serine, or pyruvate, or by 2-(ethylthio)ethanesulfonate (CH3CH2-S-CoM), but not by other coenzyme M derivatives. Efficient reduction of CO2 to CH4 was favored by low concentrations of CH3-S-CoM and high concentrations of CO2. Sulfhydryl compounds were identified as effective inhibitors of CO2 reduction. Both an allosteric model and a free-radical model for the mechanism of CO2 activation and reduction are discussed.  相似文献   

12.
Veratric acids 14C-labelled in carboxyl group, 3-OCH3, 4-OCH3, or aromatic ring together with unlabelled veratric acid were supplemented in the cultures of the white-rot fungus Phlebia radiata. The effect of various carbon sources on the release of 14CO2 was studied. Veratric acid was readily decarboxylated, maximally already on day 1 from the addition of [14COOH]-veratric acid. High amounts (4%) of glucose slightly repressed the decarboxylation. In medium supplemented with cellulose the methoxyl group in position 4 was much more readily mineralized to CO2 than the group in position 3. The maximum evolution was achieved on day 5, two days from the addition. Cellulose did not repress methanol oxidation but repression of methanol oxidation by glucose was detected in media supplemented with [O14CH3]-veratric acids and 14CH3OH. However, glucose did not repress oxidation of H14CHO. The apparent uptake of 14C by fungal mycelium, especially from methoxyl groups, but also from the aromatic ring, may partially be due to the strong slime formation observed in cellobiose medium. Also in cellobiose medium apparent uptake of 14C from 14C-labelled methoxyl groups was observed.  相似文献   

13.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

14.
A novel hydrogenase has recently been found in methanogenic archaea. It catalyzes the reversible dehydrogenation of methylenetetrahydromethanopterin (CH2 = H4MPT) to methenyltetrahydromethanopterin (CH identical to H4MPT+) and H2 and was therefore named H2-forming methylenetetrahydromethanopterin dehydrogenase. The hydrogenase, which is composed of only one polypeptide with an apparent molecular mass of 43 kDa, does not mediate the reduction of viologen dyes with either H2 or CH2 = H4MPT. We report here that the purified enzyme from Methanobacterium thermoautotrophicum exhibits the following other unique properties: (a) the colorless protein with a specific activity of 2000 U/mg (Vmax) did not contain iron-sulfur clusters, nickel, or flavins; (b) the activity was not inhibited by carbon monoxide, acetylene, nitrite, cyanide, or azide; (c) the enzyme did not catalyze an isotopic exchange between 3H2 and 1H+; (d) the enzyme catalyzed the reduction of CH identical to H4MPT+ with 3H2 generating [methylene-3H]CH2 = H4MPT; and (e) the primary structure contained at most four conserved cysteines as revealed by a comparison of the DNA-deduced amino acid sequence of the proteins from M. thermoautotrophicum and Methanopyrus kandleri. None of the four cysteines were closely spaced as would be indicative for a (NiFe) hydrogenase or a ferredoxin-type iron-sulfur protein. Properties of the H2-forming methylenetetrahydromethanopterin dehydrogenase from Methanobacterium wolfei are also described indicating that the enzyme from this methanogenic archaeon is very similar to the enzyme from M. thermoautotrophicum with respect both to molecular and catalytic properties.  相似文献   

15.
Mock WL  Cheng H 《Biochemistry》2000,39(45):13945-13952
Hydroxamic acids of structure RCON(OH)CH(2)CH(CH(2)C(6)H(5))CO(2)H induce micromolar competitive inhibition of catalysis for the enzyme carboxypeptidase A. Enzyme affinity depends on the nature of the acyl group, for RCO equaling HCO, CH(3)CO, FCH(2)CO, F(2)CHCO, F(3)CCO, CH(3)OCH(2)CO, or CH(3)OCO. In acid dissociation these residues yield hydroxamic acid pK(a) values that vary from 7.6 to 10.3. Profiles of inhibitory pK(i) plotted versus pH indicate characteristically a maximum effectiveness near neutrality. Weaker binding to enzyme is generally displayed in either acidic or alkaline solution, with the position of the alkaline limb of the profiles depending on the pK(a) of the inhibitor. A reverse-protonation pattern of association with the enzyme is indicated, in which the hydroxamate anion of the inhibitor displaces a relatively acidic H(2)O ligand (pK(a) of 6) from the active-site zinc ion of carboxypeptidase A. The metal-coordinating, N-substituted hydroxamic acid functional groups exist in solution as a mixture of syn and anti rotamers, with relative abundances that depend on their pK(a). A pyrrolidinone analogue having a conformationally syn-fixed cyclohydroxamic acid was not an especially potent inhibitor. Structure-activity relationships suggest design criteria for hydroxamic acid inhibitors in order to provide most effective binding with metalloenzymes.  相似文献   

16.
The O-methyl substituents of aromatic compounds constitute a C(1) growth substrate for a number of taxonomically diverse anaerobic acetogens. In this study, strain TH-001, an O-demethylating obligate anaerobe, was chosen to represent this physiological group, and the carbon flow when cells were grown on O-methyl substituents as a C(1) substrate was determined by C radiotracer techniques. O-[methyl-C]vanillate (4-hydroxy-3-methoxy-benzoate) was used as the labeled C(1) substrate. The data showed that for every O-methyl carbon converted to [C]acetate, two were oxidized to CO(2). Quantitation of the carbon recovered in the two products, acetate and CO(2), indicated that acetate was formed in part by the fixation of unlabeled CO(2). The specific activity of C in acetate was 70% of that in the O-methyl substrate, suggesting that only one carbon of acetate was derived from the O-methyl group. Thus, it is postulated that the carboxyl carbon of the product acetate is derived from CO(2) and the methyl carbon is derived from the O-methyl substituent of vanillate. The metabolism of O-[methyl-C]vanillate by strain TH-001 can be described as follows: 3CH(3)OC(7)H(5)O(3) + CO(2) + 4H(2)O --> CH(3)COOH + 2CO(2) + 10H + 10e + 3HOC(7)H(5)O(3).  相似文献   

17.
19F NMR and x-ray absorption experiments have been performed with both the isolated FeMo cofactor and the MoFe protein of nitrogenase in search of direct evidence for substrate or inhibitor binding. Using 19F NMR as a probe and p-CF3C6H4S- as the receptor ligand, the data show that the nitrogenase inhibitors CN- and CH3NC bind to the isolated FeMo cofactor-RFS- complex in N-methylformamide with a finite formation constant. Their binding increases the electronic relaxation time of the complex and increases the life-time of the FeMo cofactor-p-CF3C6H4S- bond, Parallel molybdenum K edge and extended x-ray absorption fine structure experiments show that CH3NC does not bind to molybdenum. Although CO and N3- both relieve CN- and CH3NC inhibition of electron flow through nitrogenase, unlike the latter, they do not appear to bind to isolated FeMo cofactor. In experiments with the dithionite-reduced MoFe protein, we did not detect any changes in the molybdenum K edge or extended x-ray absorption fine structure spectra upon addition of CO, N2, C2H2, NaCN, CH3NC, or azide demonstrating that either these substrates and inhibitors do not bind to molybdenum or that the FeMo cofactor site of nitrogenase is inaccessible to substrate binding except under turnover conditions.  相似文献   

18.
Formation of free radicals in golden hamster embryo (GHE) cells in the frozen living state by gamma irradiation has been studied by electron spin resonance spectroscopy at 4.2 and 77 K. The relative yields of H atoms, OH radicals, and organic radicals trapped in the irradiated GHE cells are 12, 72, and 16%, respectively, of total radical yields. When dimethylsulfoxide (DMSO) is added to GHE cells at 77 K, a large quantity of CH2SOCH3 radicals (DMSO radicals) are formed after gamma irradiation. The yields of OH radicals are not affected by the addition of DMSO. When the GHE cell-DMSO mixtures are irradiated with gamma rays at 77 K and then warmed to 111 K, the OH radicals decay, whereas the DMSO radicals do not increase complementarily. Moreover, the decay rates of the OH radicals at 111 K do not depend upon the concentration of DMSO. Thus OH radicals do not react with DMSO during warming of the irradiated sample. When H atoms are produced by gamma irradiation of acid ice at 60 K, the decay rates of the H atoms at 77 K increase with increasing DMSO concentration, indicating that DMSO reacts with H atoms (CH3SOCH3 + H----.CH2SOCH3 + H2) at 77 K by quantum-mechanical tunneling. When the GHE cell-DMSO mixture is irradiated with gamma rays at 77 or 4.2 K in the dark, DMSO ions are produced in addition to DMSO radicals. Therefore it is concluded that DMSO does not scavenge OH radicals, but does capture H atoms, holes and/or electrons in the gamma-irradiated cells, resulting in the remarkable formation of DMSO radicals. This scavenger effect of DMSO may be related to the radioprotection of DMSO against cell killing described in the companion paper (Watanabe et al., Radiat. Res., this issue).  相似文献   

19.
R Ahmad  Z Wu  D A Armstrong 《Biochemistry》1983,22(8):1806-1810
The kinetics of reaction of oxidized lumiflavin (F0) with the radicals .CO2(-), CH3CHOH, and (CH3)2COH have been investigated at pH 7 and 24 +/- 1 degree C by the pulse radiolysis technique. The radicals have been shown to react with lumiflavin with second-order rate constants of 36 +/- 4, 26 +/- 3, and 20 +/- 3 in units of 10(8) M-1 s-1, respectively. These rate constants are close to the diffusion limit. The main product in each case was the lumiflavin semiquinone radical FH.. By utilization of long pulses (approximately 100 mus), it was shown that the reaction FH. + .AH(alpha) leads to FH- + A(alpha) + H+ [.AH(alpha) = .CO2(-), CH3CHOH, or (CH3)2COH] proceeded for all three types of .AH(alpha) radical with second-order rate constants of 17 (+4,-3), 9 (+5,-3), and 9 (+4,-3), respectively, in the above units. The beta-carbon radical .CH2CH(OH)CH3 added to .FH, forming an alkylated flavin, while the .CH2CH2OH radical appeared to be capable of addition or hydrogen atom donation to .FH.  相似文献   

20.
A series of derivatives of the 2-deoxy analogue of beta-KDO (2,6-anhydro-3-deoxy-D-glycero-D-talo-octonic acid; ammonium salt, 2) has been synthesised as potential inhibitors of CMP-KDO synthetase, starting from methyl 2,6-anhydro-3-deoxy-4,5:7,8-di-O-isopropylidene-D-glycero-D-talo- octonate and replacing the CO2Me group attached to C-2 variously by CONH2, CONHOH, CH2OH, CH2PO(OH)(O-NH4+), COCH2PO(OH)(O-H3N+pheny), CH2CO2-NH4+, CON-HCH2CO2-NH4+, CONHBn, CONHHexyl, CO2Bn, and CO2Hexyl. Of these derivatives, the hydroxamic acid (CONHOH) was the best inhibitor of CMP-KDO synthetase, but was less potent than 2.  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