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 78 毫秒
1.
V N Ivanov 《Mikrobiologiia》1986,55(5):768-774
The balance of energy-rich compounds (ERC) was drawn up for the growth of Thiobacillus ferrooxidans in a medium with ferrous ions as an energy source. The balance items and the phosphorylating efficiency of oxidation (P/2e-) were calculated basing on the experimental yield values using the ERC balance equation. At a specific growth rate of 0.1 h-1, 55% of ferrous ions are used for the synthesis of cell biomass, 7.5% for maintainance, 4% of the ions are oxidized to reduce NAD+, and 34% are used to produce ERC necessary for the reduction. Here, 24% of ERC are used for the synthesis of monomers from CO2, 42% for the production of NADH, 24% for the biomass synthesis from monomers, and 10% for maintaining cell activity. The P/2e- for the oxidation of ferrous ions is 0.19 mole of ERC per 2e-. This is possible only when the [Fe3+]/[Fe2+] ratio in the cell periplasm is 1 X 10(3)-1 X 10(4).  相似文献   

2.
Studies on the relationship between cell synthesis and energy utilization in Hydrogenomonas eutropha have shown that the amount of oxidative energy required for synthetic reactions depends on the conditions of growth. The energy of hydrogen oxidation was most efficiently used when growth conditions were optimal (continuous culture, cells in exponential growth phase) and when the rate of growth was limited by H(2) or O(2) supply. Under these conditions, 2 to 2.5 atoms of oxygen were consumed by the oxyhydrogen reaction for the concomitant conversion of 1 mole of CO(2) to cell matter. This conversion efficiency, expressed as the O/C energyyield value, was observed with continuous cultures. A less efficient conversion was found with batch cultures. With limiting concentrations of CO(2) the rate of hydrogen oxidation was relatively high, and the O/C value was dependent on the growth rate. With nonlimiting concentrations of CO(2), the rate of hydrogen oxidation was strictly proportional to the rate of CO(2) fixation, and the O/C value was independent of growth rate. This proportionality between the rate of H(2) oxidation and the rate of CO(2) fixation suggested that energy supply regulates the (maximum) rate of growth. From the energy-yield measurements, we concluded that the oxidation of 1 mole of H(2) yields the equivalent of 2 moles of adenosine triphosphate for H. eutropha, and that at least 5 moles of this high-energy phosphate is required for the conversion of 1 mole of CO(2) into cellular constituents.  相似文献   

3.
Anaerobic glucose oxidation was coupled to xylose reduction in a nonfermentative Escherichia coli strain expressing NADPH-dependent xylose reductase. Xylitol production serves as the primary means of NAD(P)(+) regeneration, as glucose is converted primarily to acetate and CO(2). The membrane-bound transhydrogenase PntAB is required to achieve the maximum theoretical yield of four moles of xylitol per mole of glucose consumed.  相似文献   

4.
The anaerobic acetogenic bacterium Acetobacterium woodii can conserve energy by oxidation of various substrates coupled to either carbonate or caffeate respiration. We used a cell suspension system to study the regulation and kinetics of induction of caffeate respiration. After addition of caffeate to suspensions of fructose-grown cells, there was a lag phase of about 90 min before caffeate reduction commenced. However, in the presence of tetracycline caffeate was not reduced, indicating that de novo protein synthesis is required for the ability to respire caffeate. Induction also took place in the presence of CO(2), and once a culture was induced, caffeate and CO(2) were used simultaneously as electron acceptors. Induction of caffeate reduction was also observed with H(2) plus CO(2) as the substrate, but the lag phase was much longer. Again, caffeate and CO(2) were used simultaneously as electron acceptors. In contrast, during oxidation of methyl groups derived from methanol or betaine, acetogenesis was the preferred energy-conserving pathway, and caffeate reduction started only after acetogenesis was completed. The differential flow of reductants was also observed with suspensions of resting cells in which caffeate reduction was induced prior to harvest of the cells. These cell suspensions utilized caffeate and CO(2) simultaneously with fructose or hydrogen as electron donors, but CO(2) was preferred over caffeate during methyl group oxidation. Caffeate-induced resting cells could reduce caffeate and also p-coumarate or ferulate with hydrogen as the electron donor. p-Coumarate or ferulate also served as an inducer for caffeate reduction. Interestingly, caffeate-induced cells reduced ferulate in the absence of an external reductant, indicating that caffeate also induces the enzymes required for oxidation of the methyl group of ferulate.  相似文献   

5.
Oxygenation of carbon monoxide by bovine heart cytochrome c oxidase   总被引:1,自引:0,他引:1  
Cytochrome c oxidase (ferrocytochrome c:oxygen oxidoreductase, EC 1.9.3.1), as the terminal enzyme of the mammalian mitochondrial electron transport chain, has long been known to catalyze the reduction of dioxygen to water. We have found that when reductively activated in the presence of dioxygen, the enzyme will also catalyze the oxidation of carbon monoxide to its dioxide. Two moles of carbon dioxide is produced per mole of dioxygen, and similar rates of production are observed for 1- and 2-electron-reduced enzyme. If 13CO and O2 are used to initiate the reaction, then only 13CO2 is detected as a product. With 18O2 and 12CO, only unlabeled and singly labeled carbon dioxide are found. No direct evidence was obtained for a water-gas reaction (CO + H2O----CO2 + H2) of the oxidase with CO. The CO oxygenase activity is inhibited by cyanide, azide, and formate and is not due to the presence of bacteria. Studies with scavengers of partially reduced dioxygen show that catalase decreases the rate of CO oxygenation.  相似文献   

6.
Saturated and unsaturated fatty aldehydes were synthesized 99+% pure with yields of up to 80% by the reduction of 1-acylaziridines with lithium aluminium hydride, and in yields of up to 87% by oxidation of the corresponding alcohol with 1-chlorobenzotriazole. It was found for the reduction that optimum aldehyde yield was obtained with a mole ratio of reactants, consisting of acid chloride-ethylenimine-triethylamine-LiAlH(4), equal to 1:2:2:2. Optimum conditions for alcohol oxidation were found to be a mole ratio of oxidant to alcohol of 1:1.3 with refluxing for 45 min in methylene chloride containing 25% pyridine. Methods for the purification of the final product are also described. Purity criteria were thin-layer and gas-liquid chromatography and infrared and nuclear magnetic resonance spectroscopy.  相似文献   

7.
Pyridoxal phosphate (PLP) is an organic cofactor found in all transaminase enzymes. In this study PLP was used to replace the enzymatic deamination step in the Ehrlich pathway, for the oxidative conversion of amino acids into 2-keto acids. PLP functions in an enzymeindependent manner. It was further used in the synthesis of higher alcohols through a sequential enzymatic reduction in vitro and in vivo. PLP-dependent oxidation was investigated against five representative amino acids: valine, leucine, isoleucine, norvaline, and phenylalanine. In vitro amino acid oxidation resulted in approximately 45 ~ 75% [mole/mole] of each 2-keto acid conversion and in vitro ammonia formation was less than 2-keto acid formation, with 20% of conversion yields. Whole cell E. coli expressing reduction enzymes KivD/ADH with both single amino acid and amino acid mixture (4% yeast extract) gave the highest yield (30 ~ 55%) in the presence of the PLP-Cu complex and following enzymatic reactions.  相似文献   

8.
Pyruvate Metabolism in Sarcina maxima   总被引:6,自引:4,他引:2       下载免费PDF全文
The mechanisms of pyruvate cleavage and hydrogen production by Sarcina maxima were studied. It was found that a phosphoroclastic system for pyruvate oxidation, similar to that occurring in saccharolytic clostridia, is present in S. maxima. Cleavage of pyruvate by extracts of the latter organism resulted in the formation of acetyl phosphate, CO(2), and electrons which were transferred to ferredoxin. Formate was not an intermediate in this system. Pyruvate oxidation was coupled with ferredoxin-dependent nicotinamide adenine dinucleotide phosphate (NADP) reduction. A hydrogenase, active in particulate extracts of S. maxima, did not accept electrons from reduced ferredoxin. Formate was detected as a fermentation product when S. maxima was grown in media buffered with CaCO(3). Whole cells and extracts degraded formate to H(2) and CO(2). The evidence suggests that electrons generated by ferredoxin-linked pyruvate oxidation by S. maxima are not used for H(2) production, but that they serve for the reduction of NADP. Reduced NADP may be utilized by the organisms for synthesis of cell material. Production of H(2) by S. maxima may occur through a pyruvate clastic system similar to that present in coliform bacteria.  相似文献   

9.
CO(2) changes the biochemistry of peroxynitrite basically in two ways: (i) nitrating species is the CO(3)(-) / ()NO(2) radical pair, and (ii) peroxynitrite diffusion distance is significantly reduced. For peroxynitrite generated extracellularly this last effect is particularly dramatic at low cell density because CO(3)(-) and ()NO(2) are short-lived and decay mostly in the extracellular space or at the cell surface/membrane. This study was aimed to distinguish between peroxynitrite-induced extra- and intracellular modifications of red blood cells (RBC). Our results show that at low cell density and in the presence of CO(2) peroxynitrite induced the oxidation of surface thiols, the formation of 3-nitrotyrosine and DMPO-RBC adducts, and the down-regulation of glycophorins A and C (biomarkers of senescence). Reactivation of glycolysis reversed only the oxidation of surface thiols. Without CO(2) peroxynitrite also induced the oxidation of hemoglobin and glutathione, the accumulation of lactate, a decrease in ATP, the clustering of band 3, the externalization of phosphatidylserine, and the activation of caspases 8 and 3 (biomarkers of apoptosis). The latter biomarkers were all reversed by reactivation of glycolysis. We hypothesize that cell senescence could (generally) be derived by irreversible radical-mediated oxidation of membrane targets, while the appearance of apoptotic biomarkers could be bolstered by oxidation of intracellular targets. These results suggest that, depending on extracellular homolysis or diffusion to the intracellular space, peroxynitrite prompts RBCs toward either senescence or apoptosis through different oxidation mechanisms.  相似文献   

10.
Seravalli J  Ragsdale SW 《Biochemistry》2008,47(26):6770-6781
Carbon monoxide dehydrogenase (CODH) catalyzes the reversible oxidation of CO to CO2 at a nickel-iron-sulfur cluster (the C-cluster). CO oxidation follows a ping-pong mechanism involving two-electron reduction of the C-cluster followed by electron transfer through an internal electron transfer chain to external electron acceptors. We describe 13C NMR studies demonstrating a CODH-catalyzed steady-state exchange reaction between CO and CO2 in the absence of external electron acceptors. This reaction is characterized by a CODH-dependent broadening of the 13CO NMR resonance; however, the chemical shift of the 13CO resonance is unchanged, indicating that the broadening is in the slow exchange limit of the NMR experiment. The 13CO line broadening occurs with a rate constant (1080 s-1 at 20 degrees C) that is approximately equal to that of CO oxidation. It is concluded that the observed exchange reaction is between 13CO and CODH-bound 13CO2 because 13CO line broadening is pH-independent (unlike steady-state CO oxidation), because it requires a functional C-cluster (but not a functional B-cluster) and because the 13CO2 line width does not broaden. Furthermore, a steady-state isotopic exchange reaction between 12CO and 13CO2 in solution was shown to occur at the same rate as that of CO2 reduction, which is approximately 750-fold slower than the rate of 13CO exchange broadening. The interaction between CODH and the inhibitor cyanide (CN-) was also probed by 13C NMR. A functional C-cluster is not required for 13CN- broadening (unlike for 13CO), and its exchange rate constant is 30-fold faster than that for 13CO. The combined results indicate that the 13CO exchange includes migration of CO to the C-cluster, and CO oxidation to CO2, but not release of CO2 or protons into the solvent. They also provide strong evidence of a CO2 binding site and of an internal proton transfer network in CODH. 13CN- exchange appears to monitor only movement of CN- between solution and its binding to and release from CODH.  相似文献   

11.
Abstract The formation of H2 by chemolithoautrophically growing Oligotropha carboxidovorans has been identified as the result of the oxidation of CO mediated by the cytoplasmic species of the molybdenum-containing CO dehydrogenase multienzyme complex as follows: CO + H 2 O → CO 2+ H 2. Purified CO dehydrogenase was shown to carry hydrogen uptake and formation activities in addition to its catabolic function which is the oxidation of CO. Among the electron donors supporting H2 formation were CO, NADH, reduced flavins and reduced viologen dyes. The reduction of protons to H2 by cytoplasmic CO dehydrogenase is interpreted as a detoxification reaction for electrons to prevent cell damage in O. carboxidovorans .  相似文献   

12.
Lillehoj EB  Smith FG 《Plant physiology》1966,41(10):1553-1560
Ascorbic acid oxidase activity in Myrothecium verrucaria extracts resulted in O(2) uptake exceeding 0.5 mole per mole of ascorbic acid and in CO(2) evolution. Measurement of oxidized ascorbic acid at completion of the reaction demonstrated that an average of 10% of the oxidized product disappeared. A comparison of the gas exchange data with the amount of ascorbic acid not accounted for indicated that the reaction could not be explained by independent oxidase and oxygenase systems. Chromatographic examination of the reaction mixtures identified l-threonic acid. Experiments with ascorbic acid-1-(14)C showed that C-1 was partially decarboxylated during the oxidation. Test of the fungal extracts for enzymes that might explain the deviation from expected stoichiometry showed that phenolase, glutathione reductase, cytochrome oxidase, peroxidase and oxalic decarboxylase were not involved. Addition of azide in concentrations sufficient to block catalase increased excess O(2) consumption about 65%. No enzymes were found that could directly attack oxidized ascorbic acid. H(2)O(2) accumulated during oxidation in azide-blocked systems.The O(2) excess could be explained by assuming the enzyme had peroxidative capacity on a reductant other than ascorbic acid. An intermediate of ascorbic acid oxidation appeared to function as the substrate yielding CO(2) and l-threonic acid on degradation. The increase in excess O(2) utilized in azide-blocked systems and the H(2)O(2) accumulation also were explained by the proposed scheme.Another interpretation would involve production of free radicals during ascorbic acid oxidation. Evidence for this was the ability of extracts to oxidize DPNH in the presence of ascorbic acid. Oxygen radicals formed in such reactions were considered possible agents of degradation of ascorbic acid.  相似文献   

13.
Methanosarcina barkeri has recently been shown to produce a multisubunit membrane-bound [NiFe] hydrogenase designated Ech (Escherichia coli hydrogenase 3) hydrogenase. In the present study Ech hydrogenase was purified to apparent homogeneity in a high yield. The enzyme preparation obtained only contained the six polypeptides which had previously been shown to be encoded by the ech operon. The purified enzyme was found to contain 0.9 mol of Ni, 11.3 mol of nonheme-iron and 10.8 mol of acid-labile sulfur per mol of enzyme. Using the purified enzyme the kinetic parameters were determined. The enzyme catalyzed the H2 dependent reduction of a M. barkeri 2[4Fe-4S] ferredoxin with a specific activity of 50 U x mg protein-1 at pH 7.0 and exhibited an apparent Km for the ferredoxin of 1 microM. The enzyme also catalyzed hydrogen formation with the reduced ferredoxin as electron donor at a rate of 90 U x mg protein-1 at pH 7.0. The apparent Km for the reduced ferredoxin was 7.5 microM. Reduction or oxidation of the ferredoxin proceeded at similar rates as the reduction or oxidation of oxidized or reduced methylviologen, respectively. The apparent Km for H2 was 5 microM. The kinetic data strongly indicate that the ferredoxin is the physiological electron donor or acceptor of Ech hydrogenase. Ech hydrogenase amounts to about 3% of the total cell protein in acetate-grown, methanol-grown or H2/CO2-grown cells of M. barkeri, as calculated from quantitative Western blot experiments. The function of Ech hydrogenase is ascribed to ferredoxin-linked H2 production coupled to the oxidation of the carbonyl-group of acetyl-CoA to CO2 during growth on acetate, and to ferredoxin-linked H2 uptake coupled to the reduction of CO2 to the redox state of CO during growth on H2/CO2 or methanol.  相似文献   

14.
The rate of appearance of (14)CO(2) in expired air after the injection of a single dose of NaH(14)CO(3) has been determined in normal individuals both in the fasted and fed states. These data were combined with previously obtained results on the rate of disappearance of injected palmitate-(14)C from the bloodstream, to give a multicompartmental analysis of free fatty acid oxidation and esterification. The results confirm that glucose feeding promptly inhibits the rate of free fatty acid oxidation to CO(2). The "irreversible disposal rate," or irreversible flux of free fatty acids from the plasma, was also consistently reduced by glucose feeding. The diminution in irreversible disposal, not accounted for entirely by reduction of direct oxidation, must indicate suppression of other disposal mechanisms, including net esterification of free fatty acids. An average drop of 49% in "net esterification" when glucose was given may be compared with the 65% inhibition of rapid free fatty acid oxidation.  相似文献   

15.
Extracts of aerobically, CO-autotrophically grown cells of Pseudomonas carboxydovorans were shown to catalyze the oxidation of CO to CO(2) in the presence of methylene blue, pyocyanine, thionine, phenazine methosulfate, or toluylene blue under strictly anaerobic conditions. Viologen dyes and NAD(P)(+) were ineffective as electron acceptors. The same extracts catalyzed the oxidation of formate and of hydrogen gas; the spectrum of electron acceptors was identical for the three substrates, CO, formate, and H(2). The CO- and the formate-oxidizing activities were found to be soluble enzymes, whereas hydrogenase was membrane bound exclusively. The rates of oxidation of CO, formate, and H(2) were measured spectrophotometrically following the reduction of methylene blue. The rate of carbon monoxide oxidation followed simple Michaelis-Menten kinetics; the apparent K(m) for CO was 45 muM. The reaction rate was maximal at pH 7.0, and the temperature dependence followed the Arrhenius equation with an activation energy (DeltaH(0)) of 35.9 kJ/mol (8.6 kcal/mol). Neither free formate nor hydrogen gas is an intermediate of the CO oxidation reaction. This conclusion is based on the differential sensitivity of the activities of formate dehydrogenase, hydrogenase, and CO dehydrogenase to heat, hypophosphite, chlorate, cyanide, azide, and fluoride as well as on the failure to trap free formate or hydrogen gas in coupled optical assays. These results support the following equation for CO oxidation in P. carboxydovorans: CO + H(2)O --> CO(2) + 2 H(+) + 2e(-) The CO-oxidizing activity of P. carboxydovorans differed from that of Clostridium pasteurianum by not reducing viologen dyes and by a pH optimum curve that did not show an inflection point.  相似文献   

16.
J Heo  C R Staples  P W Ludden 《Biochemistry》2001,40(25):7604-7611
Carbon monoxide dehydrogenase (CODH) from Rhodospirillum rubrum catalyzes both the oxidation of CO and the reduction of CO(2). Studies of the redox dependence of CO(2) reduction by R. rubrum CODH show that (1) CODH is unable to catalyze CO(2) reduction at potentials greater than -300 mV; (2) the maximum activity is observed at potentials less than -480 mV; and (3) the midpoint potential (E(m)) of the transition from minimum to maximum CO(2) reduction activity occurs at approximately -339 mV. These results indicate that the C(red1) state of R. rubrum CODH (E(m) = -110 mV; g(zyx)() = 2.03, 1.88, 1.71) is not competent to reduce CO(2). Nernst analyses suggest that the reduction of CODH from the C(red1) state to the CO(2)-reducing form (C(unc), g(zyx)() = 2.04, 1.93, 1.89; E < approximately -300 mV) of the enzyme is a one-electron process. For the entire redox range, viologens stimulate CO(2) reduction by CODH more than 50-fold, and it is proposed that viologens accelerate the redox equilibration of redox buffers and [Fe(4)S(4)](B) during catalysis.  相似文献   

17.
Understanding how bioleaching systems respond to the availability of CO(2) is essential to developing operating conditions that select for optimum microbial performance. Therefore, the effect of inlet gas and associated dissolved CO(2) concentration on the growth, iron oxidation and CO(2) -fixation rates of pure cultures of Acidithiobacillus ferrooxidans and Leptospirillum ferriphilum was investigated in a batch stirred tank system. The minimum inlet CO(2) concentrations required to promote the growth of At. ferrooxidans and L. ferriphilum were 25 and 70 ppm, respectively, and corresponded to dissolved CO(2) concentrations of 0.71 and 1.57 μM (at 30°C and 37°C, respectively). An actively growing culture of L. ferriphilum was able to maintain growth at inlet CO(2) concentrations less than 30 ppm (0.31-0.45 μM in solution). The highest total new cell production and maximum specific growth rates from the stationary phase inocula were observed with CO(2) inlet concentrations less than that of air. In contrast, the amount of CO(2) fixed per new cell produced increased with increasing inlet CO(2) concentrations above 100 ppm. Where inlet gas CO(2) concentrations were increased above that of air the additional CO(2) was consumed by the organisms but did not lead to increased cell production or significantly increase performance in terms of iron oxidation. It is proposed that At. ferrooxidans has two CO(2) uptake mechanisms, a high affinity system operating at low available CO(2) concentrations, which is subject to substrate inhibition and a low affinity system operating at higher available CO(2) concentrations. L. ferriphilum has a single uptake system characterised by a moderate CO(2) affinity. At. ferrooxidans performed better than L. ferriphilum at lower CO(2) availabilities, and was less affected by CO(2) starvation. Finally, the results demonstrate the limitations of using CO(2) uptake or ferrous iron oxidation data as indirect measures of cell growth and performance across varying physiological conditions.  相似文献   

18.
Biochemistry of methanogenesis.   总被引:9,自引:0,他引:9  
Methane is a product of the energy-yielding pathways of the largest and most phylogenetically diverse group in the Archaea. These organisms have evolved three pathways that entail a novel and remarkable biochemistry. All of the pathways have in common a reduction of the methyl group of methyl-coenzyme M (CH3-S-CoM) to CH4. Seminal studies on the CO2-reduction pathway have revealed new cofactors and enzymes that catalyze the reduction of CO2 to the methyl level (CH3-S-CoM) with electrons from H2 or formate. Most of the methane produced in nature originates from the methyl group of acetate. CO dehydrogenase is a key enzyme catalyzing the decarbonylation of acetyl-CoA; the resulting methyl group is transferred to CH3-S-CoM, followed by reduction to methane using electrons derived from oxidation of the carbonyl group to CO2 by the CO dehydrogenase. Some organisms transfer the methyl group of methanol and methylamines to CH3-S-CoM; electrons for reduction of CH3-S-CoM to CH4 are provided by the oxidation of methyl groups to CO2.  相似文献   

19.
W Shin  P A Lindahl 《Biochemistry》1992,31(51):12870-12875
Adding 1,10-phenanthroline to carbon monoxide dehydrogenase from Clostridium thermoaceticum results in the complete loss of the NiFeC EPR signal and the CO/acetyl-CoA exchange activity. Other EPR signals characteristic of the enzyme (the gav = 1.94 and gav = 1.86 signals) and the CO oxidation activity are completely unaffected by the 1,10-phenanthroline treatment. This indicates that there are two catalytic sites on the enzyme; the NiFe complex is required for catalyzing the exchange and acetyl-CoA synthase reactions, while some other site is responsible for CO oxidation. The strength of CO binding to the NiFe complex was examined by titrating dithionite-reduced enzyme with CO. During the titration, the NiFeC EPR signal developed to a final spin intensity of 0.23 spin/alpha beta. The resulting CO titration curve (NiFeC spins/alpha beta vs CO pha beta) was fitted using two reactions: binding of CO to the oxidized NiFe complex, and reduction of the CO-bound species to a form that exhibits the NiFeC signal. Best fits yielded apparent binding constants between 6000 and 14,000 M-1 (Kd = 70-165 microM). This sizable range is due to uncertainty whether CO binds to all or only a small fraction (approximately 23%) of the NiFe complexes. Reduction of the CO-bound NiFe complex is apparently required to activate it for catalysis. The electron used for this reduction originates from the CO oxidation site, suggesting that delivery of a low-potential electron to the CO-bound NiFe complex is the physiological function of the CO oxidation reaction catalyzed by this enzyme.  相似文献   

20.
Peroxynitrite, the reactive species formed in vivo by the reaction of nitric oxide with superoxide anion, is capable of diffusing across erythrocyte membranes via anion channels and passive diffusion (A. Denicola, J. M. Souza, and R. Radi, Proc. Natl. Acad. Sci. USA 95, 3566-3571, 1998). However, peroxynitrite diffusion could be limited by extracellular targets, with the reaction with CO(2) (k(2) = 4.6 x 10(4) at 37 degrees C and pH 7.4) the most relevant. Herein, we studied the influence of physiological concentrations of CO(2) on peroxynitrite diffusion across intact red blood cells. The presence of CO(2) inhibited the oxidation of intracellular oxyhemoglobin by externally added peroxynitrite. However, the inhibition by CO(2) decreased at increasing red blood cell densities. At 45% hematocrit, 1.3 mM CO(2) (in equilibrium with 24 mM bicarbonate, at pH 7.4 and 25 degrees C) only inhibited 30% of intracellular oxyhemoglobin oxidation. This partial inhibition was also observed in red blood cells pretreated with the anion exchanger inhibitor 4,4'-diisothiocyanatostilbene-2,2'-disulfonic acid, ruling out a competition between peroxynitrite and bicarbonate for the transport through the anion channel. A theoretical model was developed to estimate the diffusion distance and half-life of extracellular peroxynitrite before reacting with intracellular oxyhemoglobin, at different red blood cell densities, and in the presence or absence of CO(2). The theoretical model correlated well with the experimental data. Our results indicate that, even in the presence of CO(2), peroxynitrite is able to diffuse and reach the inside of the erythrocyte.  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