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 0 毫秒
1.
The carbon monoxide compounds of the fully reduced and mixed valence forms of cytochrome bo from Escherichia coli were laser photolysed under anaerobic conditions at room temperature. The carbon monoxide recombined with characteristic rate constants of 50 s-1 or 35 s-1 in the fully reduced and mixed valence forms, respectively. Rates of CO recombination with the fully reduced enzyme were examined in a variety of mutant forms of cytochrome bo, produced by site-directed mutagenesis. A method was developed to deconvolute cytochromes bo and bd, leading to some reassessment of histidine ligands to the metals. Significant changes in the rate constant of recombination of carbon monoxide occurred in many of these mutants and these results could be rationalised generally in terms of our current working model of the folding structure of subunit I. In the mixed valence form of the enzyme the transient photolysis spectra in the visible region are consistent with a rapid electron redistribution from the binuclear centre to the low-spin haem. This electron transfer is biphasic, with rate constants of around 10(5) and 8000 s-1. The process was also examined in the His-333-Leu mutant, in which a putative histidine ligand to CuB is replaced by leucine, and which results in the loss of the CuB. It appeared that rapid haem-haem electron transfer could still occur. The observation that CuB is apparently not required for rapid haem-haem electron transfer is consistent with the recently proposed model in which the two haems are positioned on opposite sides of transmembrane helix X in subunit I of the oxidase.  相似文献   

2.
The reaction of the quinol oxidase cytochrome bo3 from Escherichia coli with ubiquinol-2 (UQ2H2) was carried out using substoichiometric (0.5 equiv) amounts of substrate. Reactions were monitored through the use of freeze-quench EPR spectroscopy. Under 1 atm of argon, semiquinone was formed at the QB site of the enzyme with a formation rate constant of 140 s-1; the QB semiquinone EPR signal decayed with a rate constant of about 5 s-1. Heme b and CuB were reduced within the 10-ms dead time of the freeze-quench experiment and remained at a constant level of reduction over the 1-s time course of the experiment. Quantitation of the reduction levels of QB and heme b during this reaction yielded a reduction potential of 30-60 mV for heme b. Under a dioxygen atmosphere, the rates of semiquinone formation and its subsequent decay were not altered significantly. However, accurate quantitation of the EPR signals for heme b and heme o3 could not be made, due to interference from dioxygen. In the reaction between the QB-depleted enzyme and UQ2H2 under substoichiometric conditions, there was no observable change in the EPR spectra of the enzyme over the time course of the reaction, suggesting an electron transfer from heme b to the binuclear site in the absence of QB which occurs within the dead time of the freeze-quench apparatus. Analysis of the thermodynamics and kinetics of electron transfers in this enzyme suggests that a Q-cycle mechanism for proton translocation is more likely than a cytochrome c oxidase-type ion-pump mechanism.  相似文献   

3.
Photodissociation of fully reduced, carbonmonoxy cytochrome bo3 causes ultrafast transfer of carbon monoxide (C triple bond O) from heme iron to CuB in the binuclear site. At low temperatures, the C triple bond O remains bound to CuB for extended times. Here, we show that the binding of C triple bond O to CuB perturbs the IR stretch of an un-ionized carboxylic acid residue, which is identified as Glu286 by mutation to Asp or to Cys. Before photodissociation, the carbonyl (C=O)-stretching frequency of this carboxylic acid residue is 1726 cm-1 for Glu286 and 1759 cm-1 for Glu286Asp. These frequencies are definitive evidence for un-ionized R-COOH and suggest that the carboxylic acids are hydrogen-bonded, though more extensively in Glu286. In Glu286Cys, this IR feature is lost altogether. We ascribe the frequency shifts in the C=O IR absorptions to the effects of binding photodissociated C triple bond O to CuB, which are relay ed to the 286 locus. Conversely, the 2065 cm-1 C triple bond O stretch of CuB-CO is markedly affected by both mutations. These effects are ascribed to changes in the Lewis acidity of CuB, or to displacement of a CuB histidine ligand by C triple bond O. C triple bond O binding to CuB also induces a downshift of an IR band which can be attributed to an aromatic C-H stretch, possibly of histidine imidazole, at about 3140 cm-1. The results suggest an easily polarizable, through-bond connectivity between one of the histidine CuB ligands and the carboxylic group of Glu286. A chain of bound water molecules may provide such a connection, which is of interest in the context of the proton pump mechanism of the heme-copper oxidases.  相似文献   

4.
The cytochrome bo3 ubiquinol oxidase contains at least one and possibly two binding sites for ubiquinol/ubiquinone. Previous studies used the photoreactive affinity label 3-[3H]azido-2-methyl-5-methoxy-6-geranyl-1,4-benzoquinone (azido-Q), a substrate analogue, to demonstrate that subunit II contributes to at least one of the quinol binding sites. In the current work, mass spectroscopy is used to identify a peptide within subunit II that is photolabeled by the azido-Q. Purified cytochrome bo3 was photolabeled as previously described using azido-Q that was not tritiated (i.e., not radiolabeled). Subunit II was then isolated from an SDS-PAGE gel and proteolyzed in situ with trypsin. The resulting peptides were eluted from the gel and then identified using matrix-assisted laser desorption ionization mass spectrometry. The resulting mass spectrum was compared to that obtained by analysis of subunit II that had not been exposed to the photolabel. Using the amino acid sequence, each peak in the mass spectrum of the unlabeled subunit II could be assigned to an expected trypsin fragment. Two additional peaks were observed in the mass spectrum of the photolabeled subunit with m/z 1931.9 and 2287.7. Subtraction of the mass of azido-Q from the peak at m/z 1931.9 results in a mass equivalent to that of a peptide consisting of amino acids 165-178. The assignment of the peak at m/z 2287.7 cannot be made unequivocally and may correspond either to the covalent attachment of azido-Q to peptide 254-270 or to a peptide resulting from incomplete proteolysis. The labeled peptide, 165-178, is within the water-soluble domain of subunit II, whose X-ray structure is known. This peptide is located near the site where CuA is located in the homologous cytochrome c oxidases and can be placed near the interface between subunits I and II.  相似文献   

5.
Flavocytochrome b2, which has been fully reduced using L-lactate, can be rapidly oxidized by 1 equiv using the laser-generated triplet state of 5-deazariboflavin. Parallel photoinduced oxidation occurs at the reduced heme and at the fully reduced FMN (FMNH2) prosthetic groups of different enzyme monomers, producing the anion semiquinone of FMN and a ferric heme. Following the initial oxidation reaction, rapid intramolecular reduction of the ferric heme occurs with concomitant oxidation of FMNH2, generating the neutral FMN semiquinone. The observed rate constant for this intramolecular electron transfer is 2200 s-1, which is 1 order of magnitude larger than the turnover number under these conditions. A slower reduction of the heme prosthetic group also occurs with an observed rate constant of approximately 10 s-1, perhaps due to intersubunit electron transfer from reduced FMN to heme. The rapid intramolecular electron transfer between the FMNH2 and ferric heme is eliminated upon addition of excess pyruvate (Ki = 3.8 mM). This latter result indicates that pyruvate inhibition of catalytic turnover apparently can occur at the FMNH2-->heme electron transfer step. These results markedly differ from those previously obtained (Walker, M. C., & Tollin, G. (1991) Biochemistry 30, 5546-5555) and confirmed here for electron transfer within the one-electron reduced enzyme and for the effect of pyruvate binding, suggesting that intramolecular communication between the heme and flavin prosthetic groups can be controlled by the redox state of the enzyme and by ligand binding to the active site.  相似文献   

6.
To characterize the structural features of the quinol oxidation site (the QL site) of the cytochrome bo complex, a heme-copper respiratory oxidase in Escherichia coli, we carried out structure-inhibitory potency analyses using 7 p-benzoquinones and 33 substituted phenols. Their effects on its ubiquinol-1 oxidase activity were compared with those on the cytochrome bd complex in E. coli and on cytochromes o and alpha 1 in Acetobacter aceti. They showed similar structural properties of the QL site, although cytochrome o was more sensitive to 4-cyanophenols, suggesting a specific interaction of the hydrogen bond-accepting cyano group with the binding pocket. Replacing one of the methyl groups of 2,6-dimethyl-p-benzoquinone, which is the most potent competitive inhibitor, with an ethyl group markedly decreased the inhibitory activity, indicating that the QL site specifically recognizes one C = O group with two methyl groups as the ortho-substituents. In substituted phenols, ortho-chlorine substituents were the most effective in recognition, and the electron-withdrawing ability of the para-substituent determined an inhibitory potency, probably by stabilizing an anionic form. Based on these observations, we postulate that the QL site of the cytochrome bo complex asymmetrically recognizes exogenous ligands and that this property accounts for the sequential electron transfer from ubiquinols to the low-spin heme.  相似文献   

7.
Cytochrome bd oxidase is a bacterial terminal oxidase that contains three cofactors: a low-spin heme (b558), a high-spin heme (b595), and a chlorin d. The center of dioxygen reduction has been proposed to be a binuclear b595/d site, whereas b558 is mainly involved in transferring electrons from ubiquinol to the oxidase. Information on the nature of the axial ligands of the three heme centers has come from site-directed mutagenesis and spectroscopy, which have implicated a His/Met coordination for b558 (Spinner, F., Cheesman, M. R., Thomson, A. J., Kaysser, T., Gennis, R. B., Peng, Q., & Peterson, J. (1995) Biochem. J. 308, 641-644; Kaysser, T. M., Ghaim, J. B., Georgiou, C., & Gennis, R. B. (1995) Biochemistry 34, 13491-13501), but the ligands to b595 and d are not known with certainty. In this work, the three heme chromophores of the fully reduced cytochrome bd oxidase are studied individually by selective enhancement of their resonance Raman (rR) spectra at particular excitation wavelengths. The rR spectrum obtained with 413.1-nm excitation is dominated by the bands of the 5cHS b595(2+) cofactor. Excitation close to 560 nm yields a rR spectrum dominated by the 6cLS b558(2+) heme. Wavelengths between these values enhance contributions from both b595(2+) and b558(2+) chromophores. The rR bands of the ferrous chlorin become the major features with red laser excitation (595-650 nm). The rR data indicate that d2+ is a 5cHS system whose axial ligand is either a weakly coordinating protein donor or a water molecule. In the low-frequency region of the 441.6-nm spectrum, we assign a rR band at 225 cm-1 to the (b595)Fe(II)-N(His) stretching vibration, based on its 1.2-cm(-1) upshift in the 54Fe-labeled enzyme. This observation provides the first physical evidence that the proximal ligand of b595 is a histidine. Site-directed mutagenesis had suggested that His 19 is associated with either b595 or d (Fang, H., Lin, R. -J., & Gennis, R. B. (1989) J. Biol. Chem. 264, 8026-8032). On the basis of the present study, we propose that the proximal ligand of b595 is His 19. We have also studied the reaction of cyanide with the fully reduced cytochrome bd oxidase. In approximately 700-fold excess cyanide (approximately 35 mM), the 629-nm UV/vis band of d2+ is blue-shifted to 625 nm and diminished in intensity. However, the rR spectra at each of three different gamma(0) (413.1, 514.5, and 647.1 nm) are identical with or without cyanide, thus indicating that both b595 and d remain as 5cHS species in the presence of CN-. This observation leads to the proposal that a native ligand of ferrous chlorin d is replaced by CN- to form the 5cHS d2+ cyano adduct. These findings corroborate our companion study of the "as-isolated" enzyme in which we proposed a 5cHS d3+ cyano adduct (Sun, J., Osborne, J. P., Kahlow, M. A., Kaysser, T. M., Hill, J. J., Gennis, R. B., & Loehr, T. M. (1995) Biochemistry 34, 12144-12151). To further characterize the unusual and unexpected nature of these proposed high-spin cyanide adducts, we have obtained EPR spectral evidence that binding of cyanide to fully oxidized cytochrome bd oxidase perturbs a spin-state equilibrium in the chlorin d3+ to yield entirely the high-spin form of the cofactor.  相似文献   

8.
The cytochrome bo complex is a terminal ubiquinol oxidase in the aerobic respiratory chain of Escherichia coli and is encoded by the cyoABCDE operon. Recently, we have demonstrated that heme O at the high-spin heme-binding site is essential for redox-coupled proton pumping by the oxidase and suggested that the cyoE gene encodes a novel enzyme for heme O biosynthesis, protoheme IX farnesyltransferase (heme O synthase) (Saiki, K., Mogi, T., and Anraku, Y. (1992) Biochem. Biophys. Res. Commun. 189, 1491-1497). This study was focused to define the catalytic domain(s) of the CyoE protein via a site-directed mutagenesis approach. We have individually substituted 40 amino acid residues including 22 invariant residues with alanines and found that 23 mutant oxidases were nonfunctional and exhibited a specific loss of the CO binding activity at the site of the high-spin heme. Characterizations of the purified D65A, Y120A, and W172A mutant oxidases, which represent the mutations of different topological domains, revealed that their defects are attributable to substitution of protoheme IX for heme O present in the high-spin heme-binding site. Based on the above observations, we suggest that the conserved amino acid residues present in the cytoplasmic loops II/III and IV/V are part of the catalytic center of heme O synthase.  相似文献   

9.
We report on a system to improve expression of mature c-type cytochromes in Escherichia coli. It is based on the use of plasmid pEC86 that expresses the E. coli cytochrome c maturation genes ccmABCDEFGH constitutively, whereby the production of both endogenous and foreign c-type cytochromes was increased substantially. The periplasmic soluble domains of the c-type cytochrome subunits FixO and FixP of the Bradyrhizobium japonicum cbb3 oxidase could be expressed in E. coli only when pEC86 was provided in a degP-deficient strain. This shows that a stimulation of heme attachment by the Ccm maturase system combined with the diminished proteolytic activity in the periplasm can increase c-type cytochrome yields.  相似文献   

10.
Krabbe disease or globoid cell leukodystrophy (GLD) is an autosomal recessive disorder resulting from the defective lysosomal hydrolysis of specific galactolipids found primarily in myelin. This leads to severe neurological symptoms including seizures, hypotonia, blindness, and death, usually before 2 years of age in human patients. In addition to human patients, several animals, including dog, mouse, and monkey, have the same disease caused by a deficiency of galactocerebrosidase (GALC) activity. In this article we describe studies in cairn and West Highland white terriers (WHWT) affected with GLD. Through a screening test based on the molecular defect found in these breeds, over 50 cairn terrier carriers have been identified and a colony of five carrier dogs has been established. Affected dogs from this colony plus an affected WHWT were available for study. An affected WHWT was evaluated by magnetic resonance imaging at 6 and 11 months of age and pronounced changes in the T-2 weighted fast spin-echo images were found. Biochemical and pathological evaluation of the same dog after euthanasia at 12 months of age showed a large accumulation of psychosine in the brain and white matter filled with globoid cells. Some comparisons were made to younger affected and carrier dogs. Studies have shown successful transduction of cultured skin fibroblasts from an affected dog and normal canine bone marrow using a retroviral vector containing the human GALC cDNA. Successful treatment of this canine model will lead to studies in some humans with GLD.  相似文献   

11.
The phantom scatter correction factor Sp of megavoltage photon beams can be accurately described using a three-Gaussian fit. The model leads to six parameters, with which Sp(r) is described as a smooth function of the field radius r for beam qualities in the range from 60Co up to 25 MV. The parameters allow Sp values to be calculated at intermediate beam energies and for any field shape. Calculated Sp(X, Y) values for rectangular fields (X, Y) can be subsequently used as reference values to compare with measured Sp(X, Y) values, for example when appraising a new beam.  相似文献   

12.
Determining the way in which deleterious mutations interact in their effects on fitness is crucial to numerous areas in population genetics and evolutionary biology. For example, if each additional mutation leads to a greater decrease in log fitness than the last (synergistic epistasis), then the evolution of sex and recombination may be favored to facilitate the elimination of deleterious mutations. However, there is a severe shortage of relevant data. Three relatively simple experimental methods to test for epistasis between deleterious mutations in haploid species have recently been proposed. These methods involve crossing individuals and examining the mean and/or skew in log fitness of the offspring and parents. The main aim of this paper is to formalize these methods, and determine the most effective way in which tests for epistasis could be carried out. We show that only one of these methods is likely to give useful results: crossing individuals that have very different numbers of deleterious mutations, and comparing the mean log fitness of the parents with that of their offspring. We also reconsider experimental data collected on Chlamydomonas moewussi using two of the three methods. Finally, we suggest how the test could be applied to diploid species.  相似文献   

13.
It has been suggested that two groups of Escherichia coli genes, the ccm genes located in the 47-min region and the nrfEFG genes in the 92-min region of the chromosome, are involved in cytochrome c biosynthesis during anaerobic growth. The involvement of the products of these genes in cytochrome c synthesis, assembly and secretion has now been investigated. Despite their similarity to other bacterial cytochrome c assembly proteins, NrfE, F and G were found not to be required for the biosynthesis of any of the c-type cytochromes in E. coli. Furthermore, these proteins were not required for the secretion of the periplasmic cytochromes, cytochrome C550 and cytochrome C552, or for the correct targeting of the NapC and NrfB cytochromes to the cytoplasmic membrane. NrfE and NrfG are required for formate-dependent nitrite reduction (the Nrf pathway), which involves at least two c-type cytochromes, cytochrome C552 and NrfB, but NrfF is not essential for this pathway. Genes similar to nrfE, nrfF and nrfG are present in the E. coli nap-ccm locus at minute 47. CcmF is similar to NrfE, the N-terminal region of CcmH is similar to NrfF and the C-terminal portion of CcmH is similar to NrfG. In contrast to NrfF, the N-terminal, NrfF-like portion of CcmH is essential for the synthesis of all c-type cytochromes. Conversely, the NrfG-like C-terminal region of CcmH is not essential for cytochrome c biosynthesis. The data are consistent with proposals from this and other laboratories that CcmF and CcmH form part of a haem lyase complex required to attach haem c to C-X-X-C-H haem-binding domains. In contrast, NrfE and NrfG are proposed to fulfill a more specialised role in the assembly of the formate-dependent nitrite reductase.  相似文献   

14.
Enrichment of the growth medium with iron partially relieves the phenotypic deficits imposed on Escherichia coli by lack of both manganese and iron superoxide dismutases. Thus iron supplementation increased the aerobic growth rate, decreased the leakage of sulfite, and diminished sensitivity toward paraquat. Iron supplementation increased the activities of several [4Fe-4S]-containing dehydratases, and this was seen even in the presence of 50 microg/ml of rifampicin, an amount which completely inhibited growth. Assessing the O-2 scavenging activity by means of lucigenin luminescence indicated that the iron-enriched sodAsodB cells had gained some means of eliminating O-2, which was not detectable as superoxide dismutase activity in cell extracts. It is noteworthy that iron-enriched cells were not more sensitive toward the lethality of H2O2 despite having the usual amount of catalase activity. This indicates that iron taken into the cells from the medium is not available for Fenton chemistry, but is available for reconstitution of iron-sulfur clusters. We suppose that oxidation of the [4Fe-4S] clusters of dehydratases by O-2 and their subsequent reductive reconstitution provides a mechanism for scavenging O-2 and that speeding this reductive reconstitution by iron enrichment both spared other targets from O-2 attack and maintained adequate levels of these enzymes to meet the metabolic needs of the cells.  相似文献   

15.
A hypothetical model for electron transfer complex between cytochrome c3 and the flavodoxin from the sulfate-reducing bacteria Desulfovibrio vulgaris has been proposed, based on electrostatic potential field calculations and NMR data [Stewart, D. E., LeGall, J. , Moura, I., Moura, J. J. G., Peck, H. D., Jr., Xavier, A. V., Weiner, P. K., & Wampler, J. E. (1988) Biochemistry 27, 2444-2450]. This modeled complex relies primarily on the formation of five ion pairs between lysine residues of the cytochrome and acidic residues surrounding the flavin mononucleotide cofactor of the flavodoxin. In this study, the role of several acidic residues of the flavodoxin in the formation of this complex and in electron transfer between these two proteins was evaluated. A total of 17 flavodoxin mutants were studied in which 10 acidic amino acids--Asp62, Asp63, Glu66, Asp69, Asp70, Asp95, Glu99, Asp106, Asp127, and Asp129--had been permanently neutralized either individually or in various combinations by substitution with their amide amino acid equivalent (i.e., asparate to asparagine, glutamate to glutamine) through site-directed mutagenesis. The kinetic data for the transfer of electrons from reduced cytochrome c3 to the various flavodoxin mutants do not conform well to a simple bimolecular mechanism involving the formation of an intermediate electron transfer complex. Instead, a minimal electron transfer mechanism is proposed in which an initial complex is formed that is stabilized by intermolecular electrostatic interactions but is relatively inefficient in terms of electron transfer. This step is followed by a rate-limiting reorganization of that complex leading to efficient electron transfer. The apparent rate of this reorganization step was enhanced by the disruption of the initial electrostatic interactions through the neutralization of certain acidic amino acid residues leading to faster overall observed electron transfer rates at low ionic strengths. Of the five acidic residues involved in ion pairing in the modeled complex proposed by Stewart et al. (1988), the kinetic data strongly implicate Asp62, Glu66, and Asp95 in the formation of the electrostatic interactions that control electron transfer. Less certainty is provided by this study for the involvement of Asp69 and Asp129, although the data do not exclude their participation. It was not possible to determine whether the modeled complex represents the optimal configuration for electron transfer obtained after the reorganization step or actually represents the initial complex. The data do provide evidence for the importance of electrostatic interactions in electron transfer between these two proteins and for the existence of alternative binding modes involving acidic residues on the surface of the flavodoxin other than those proposed in that model.  相似文献   

16.
Electron transfer from the tetraheme cytochrome c to the special pair of bacteriochlorophylls (P) has been studied by flash absorption spectroscopy in reaction centers isolated from seven strains of the photosynthetic purple bacterium Rhodopseudomonas viridis, where the residue L162, located between the proximal heme c-559 and P, is Y (wild type), F, W, G, M, T, or L. Measurements were performed between 294 K and 8 K, under redox conditions in which the two high-potential hemes of the cytochrome were chemically reduced. At room temperature, the kinetics of P+ reduction include two phases in all of the strains: a dominant very fast phase (VF), and a minor fast phase (F). The VF phase has the following t(1/2): 90 ns (M), 130 ns (W), 135 ns (F), 189 ns (Y; wild type), 200 ns (G), 390 ns (L), and 430 ns (T). These data show that electron transfer is fast whatever the nature of the amino acid at position L162. The amplitudes of both phases decrease suddenly around 200 K in Y, F, and W. The effect of temperature on the extent of fast phases is different in mutants G, M, L, and T, in which electron transfer from c-559 to P+ takes place at cryogenic temperatures in a substantial fraction of the reaction centers (T, 48%; G, 38%; L, 23%, at 40 K; and M, 28%, at 60 K), producing a stable charge separated state. In these nonaromatic mutants the rate of VF electron transfer from cytochrome to P+ is nearly temperature-independent between 294 K and 8 K, remaining very fast at very low temperatures (123 ns at 60 K for M; 251 ns at 40 K for L; 190 ns at 8 K for G, and 458 ns at 8 K for T). In all cases, a decrease in amplitudes of the fast phases is paralleled by an increase in very slow reduction of P+, presumably by back-reaction with Q(A)-. The significance of these results is discussed in relation to electron transfer theories and to freezing at low temperatures of cytochrome structural reorganization.  相似文献   

17.
Electron-transfer reactions following the formation of P700(+)A1- have been studied in isolated Photosystem I complexes from Synechococcus elongatus between 300 and 5 K by flash absorption spectroscopy. (1) In the range from 300 to 200 K, A1- is reoxidized by electron transfer to the iron-sulfur cluster FX. The rate slows down with decreasing temperature, corresponding to an activation energy of 220 +/- 20 meV in this temperature range. Analyzing the temperature dependence of the rate in terms of nonadiabatic electron-transfer theory, one obtains a reorganization energy of about 1 eV and an edge-to-edge distance between A1 and FX of about 8 A assuming the same distance dependence of the electron-transfer rate as in purple bacterial reaction centers. (2) At temperatures below 150 K, different fractions of PS I complexes attributed to frozen conformational substates can be distinguished. A detailed analysis at 77 K gave the following results: (a) In about 45%, flash-induced electron transfer is limited to the formation and decay of the secondary pair P700(+)A1-. The charge recombination occurs with a t1/2 of about 170 micros. (b) In about 20%, the state P700(+)FX- is formed and recombines with complex kinetics (t1/2 = 5-100 ms). (c) In about 35%, irreversible formation of P700(+)FA- or P700(+)FB- is possible. (3) The transition from efficient forward electron transfer at higher temperatures to heterogeneous photochemistry at low temperatures has been investigated in different glass-forming solvents. The yield of forward electron transfer to the iron-sulfur clusters decreases in a narrow temperature interval. The temperature of the half-maximal effect varies between different solvents and appears to be correlated with their liquid to glass transition. It is proposed that reorganization processes in the surroundings of the reactants which are required for the stabilization of the charge-separated state become arrested near the glass transition. This freezing of protein motions and/or solvent reorganization may affect electron-transfer reactions through changes in the free-energy gap and the reorganization energy. (4) The rate of charge recombination between P700(+) and A1- increases slightly (about 1.5-fold) when the temperature is decreased from 300 to 5 K. This charge recombination characterized by a large driving force is much less influenced by the solvent properties than the forward electron-transfer steps from A1- to FX and FA/B.  相似文献   

18.
BACKGROUND: In patients with reflux oesophagitis, endoscopic healing and symptom relief are considered important treatment goals in long-term care. AIM: To compare the effect of lansoprazole 15 and 30 mg daily on maintaining endoscopic healing and symptom relief in patients with moderate reflux oesophagitis. PATIENTS AND METHODS: In a single-centre, double-blind randomized clinical trial, 103 patients with grade 1 or 2 reflux oesophagitis who were endoscopically healed and asymptomatic after lansoprazole 30 mg daily for 12 weeks, were randomized to maintenance therapy with either lansoprazole 15 mg or 30 mg o.m. Endoscopy was repeated after 3, 6 and 12 months, and symptom relief assessed after 3, 6, 9 and 12 months. Relapse of oesophagitis or symptoms were considered end-points. RESULTS: After 12 months, 14/50 patients (28%) receiving lansoprazole 15 mg daily had suffered an endoscopic relapse compared to 8/53 patients (15%) treated with lansoprazole 30 mg daily. A life table analysis showed no statistically significant difference between the two groups (P = 0.086). Significantly more patients were kept in complete symptomatic remission in the 30 mg group (P < 0.01). In the 15 mg group, 23/50 (46%) had suffered either an endoscopic or symptomatic relapse on completion of the study, compared to 12/53 (23%) in the 30 mg group. A life table analysis showed this difference to be statistically significant (P = 0.010). Lansoprazole 15 and 30 mg daily were equally well tolerated. CONCLUSION: No statistically significant differences were found in endoscopic relapse rate or occurrence of adverse events, while lansoprazole 30 mg proved superior to 15 mg in maintaining patients in symptomatic relief and combined endoscopic and symptomatic remission.  相似文献   

19.
20.
Structural models of subunit epsilon of the ATP synthase from Escherichia coli have been determined recently by NMR [Wilkens et al. (1995) Nat. Struct. Biol. 2, 961-967] and by X-ray crystallography [Uhlin et al. (1997) Structure 5, 1219-1230], revealing a two-domain protein. In this study, six new epsilon mutants were constructed and analyzed: Y63A, D81A, T82A, and three truncated mutants, tr80(S), tr94(LAS), and tr117(AS). Seven mutants constructed previously were also analyzed: E31A, E59A, S65A, E70A, T77A, R58A, and D81A/R85A. Subunits were purified by isoelectric focusing from extracts of cells that overproduced these 13 mutants. F1 was prepared lacking subunit epsilon by immobilized-Ni affinity chromatography. Three mutants, E70A, S65A, and E31A, showed somewhat higher affinities and extents of inhibition than the wild type. Three mutants, T82A, R85A, and tr94(LAS), showed both lower affinities and extents of inhibition, over the concentration range tested. Two showed no inhibition, D81A and tr80(S). The others, T77A, Y63A, E59A, and tr117(AS), showed lower affinities than wild type, but the extents of inhibition were nearly normal. Results indicate that the C-terminal domain of subunit epsilon contributes to inhibition of ATP hydrolysis, but it is not necessary for ATP-driven proton translocation. Interactions with subunit gamma are likely to involve a surface containing residues S65, E70, T77, D81, and T82, while residues R85 and Y63 are likely to be important in the conformation of subunit epsilon.  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