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 29 毫秒
1.
As recently demonstrated by our group (da-Silva, W. S., Gómez-Puyou, A., Gómez-Puyou, M. T., Moreno-Sanchez, R., De Felice, F. G., de Meis, L., Oliveira, M. F., and Galina, A. (2004) J. Biol. Chem. 279, 39846-39855) mitochondrial hexokinase activity (mt-HK) plays a preventive antioxidant role because of steady-state ADP re-cycling through the inner mitochondrial membrane in rat brain. In the present work we show that ADP re-cycling accomplished by the mitochondrial creatine kinase (mt-CK) regulates reactive oxygen species (ROS) generation, particularly in high glucose concentrations. Activation of mt-CK by creatine (Cr) and ATP or ADP, induced a state 3-like respiration in isolated brain mitochondria and prevention of H(2)O(2) production obeyed the steady-state kinetics of the enzyme to phosphorylate Cr. The extension of the preventive antioxidant role of mt-CK depended on the phosphocreatine (PCr)/Cr ratio. Rat liver mitochondria, which lack mt-CK activity, only reduced state 4-induced H(2)O(2) generation when 1 order of magnitude more exogenous CK activity was added to the medium. Simulation of hyperglycemic conditions, by the inclusion of glucose 6-phosphate in mitochondria performing 2-deoxyglucose phosphorylation via mt-HK, induced H(2)O(2) production in a Cr-sensitive manner. Simulation of hyperglycemia in embryonic rat brain cortical neurons increased both DeltaPsi(m) and ROS production and both parameters were decreased by the previous inclusion of Cr. Taken together, the results presented here indicate that mitochondrial kinase activity performed a key role as a preventive antioxidant against oxidative stress, reducing mitochondrial ROS generation through an ADP-recycling mechanism.  相似文献   

2.
We have studied the binding of 125I-labeled high density lipoproteins (HDL3) to liver plasma membranes, which are thought to contain specific HDL receptor sites, using anti-peptide antibodies directed against two sites in the carboxyl-terminal region of human apoA-I. Two distinct antibody populations raised to peptides corresponding to amino acid residues 205-220 and 230-243, respectively, recognized regions of apoA-I that are exposed in the lipid environment of HDL3. However, anti-AI[230-243] IgG, but not anti-AI[205-220] IgG, recognized HDL2, suggesting that residues 205-220 of apoA-I are expressed differently in the two HDL populations. In addition, anti-AI[230-243] IgG showed strong cross-reactivity toward apoA-II. Epitope mapping studies showed that anti-AI[230-243] binds to an epitope located in the carboxyl-terminus of apoA-II, demonstrating significant structural homology between the carboxyl-terminal of apoA-II, demonstrating significant structural homology between the carboxyl-terminal regions of apoA-I and A-II, two candidate proteins for mediating the specific cellular interaction of HDL3. Fab fragments from anti-AI[205-220] and anti-AI[230-243] inhibited the binding of 125I-HDL3 to liver plasma membranes by approximately 80% and 60%, respectively. These findings are in agreement with our recent work using isolated CNBr fragments of apoA-I (Morrison, J., Fidge, N. H., and Tozuka, M. (1991) J. Biol. Chem. 266, 18780-18785), which suggest that the carboxyl-terminal region of apoA-I contains a binding domain which mediates the specific interaction of HDL3 with liver plasma membranes, possibly through the involvement of specific HDL receptors.  相似文献   

3.
The backbone dynamics of uniformly 15N-labeled interleukin-1 beta are investigated by using two-dimensional inverse detected heteronuclear 15N-1H NMR spectroscopy. 15N T1, T2, and NOE data at a spectrometer frequency of 600 MHz are obtained for 90% of the backbone amide groups. The data provide evidence for motions on three time scales. All the residues exhibit very fast motions on a time scale of approximately less than 20-50 ps that can be characterized by a single-order parameter with an average value of 0.82 +/- 0.05. For a model comprising free diffusion within a cone, these residue-specific order parameters translate to an average cone semiangle of 20.7 +/- 3.3 degrees. Thirty-two residues also display motions on a time scale of 0.5-4 ns, slightly less than the overall rotational correlation time of the protein (8.3 ns). These additional motions must be invoked to account for the discrepancy between experiment and the simplest theoretical formulation in which the internal motions are described by only two parameters, a generalized order parameter and an effective correlation time [Lipari, G., & Szabo, A. (1982a) J. Am. Chem. Soc. 104, 4546-4559]. In particular, while the simple formulation can account for the 15N T1 and T2 data, it fails to account for the 15N-1H NOE data and yields calculated values for the NOEs that are either too small or negative, whereas the observed NOEs are positive. With the introduction of two internal motions that are faster than the rotational correlation time and differ in time scales by at least 1-2 orders of magnitude [Clore, G. M., Szabo, A., Bax, A., Kay, L. E., Driscoll, P. C., & Gronenborn, A. M. (1990) J. Am. Chem. Soc. 112, 4989-4991], all the relaxation data for these 32 residues can be fitted by two order parameters and an effective correlation time for the slower of the two internal motions. A simple model for these two motions is one in which the very fast motion involves axially symmetric diffusion within a cone, while the slower motion comprises jumps between two different orientations of the NH vector. For such a model the jump angle (excluding the C-terminal residue) ranges from 15 degrees to 69 degrees with a mean value of 28.6 +/- 14.0 degrees. Another 42 residues are characterized by some sort of motion on the 30-ns-10-ms time scale, which results in 15N line broadening due to chemical exchange between different conformational substates with distinct 15N chemical shifts.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Three unrestrained stochastic dynamics simulations have been carried out on the RNA hairpin GGAC[UUCG] GUCC, using the AMBER94 force field (Cornell et al., 1995. J. Am. Chem. Soc. 117:5179-5197) in MacroModel 5.5 (Mohamadi et al., 1990. J. Comp. Chem. 11:440-467) and either the GB/SA continuum solvation model (Still et al., 1990. J. Am. Chem. Soc. 112:6127-6129) or a linear distance-dependent dielectric (1/R) treatment. The linear distance-dependent treatment results in severe distortion of the nucleic acid structure, restriction of all hydroxyl dihedrals, and collapse of the counterion atmosphere over the course of a 5-ns simulation. An additional vacuum simulation without counterions shows somewhat improved behavior. In contrast, the two GB/SA simulations (1.149 and 3.060 ns in length) give average structures within 1.2 A of the initial NMR structure and in excellent agreement with results of an earlier explicit solvent simulation (Miller and Kollman, 1997. J. Mol. Biol. 270:436-450). In a 3-ns GB/SA simulation starting with the incorrect UUCG tetraloop structure (Cheong et al., 1990. Nature. 346:680-682), this loop conformation converts to the correct loop geometry (Allain and Varani, 1995. J. Mol. Biol. 250:333-353), suggesting enhanced sampling relative to the previous explicit solvent simulation. Thermodynamic effects of 2'-deoxyribose substitutions of loop nucleotides were experimentally determined and are found to correlate with the fraction of time the ribose 2'-OH is hydrogen bonded and the distribution of the hydroxyl dihedral is observed in the GB/SA simulations. The GB/SA simulations thus appear to faithfully represent structural features of the RNA without the computational expense of explicit solvent.  相似文献   

5.
Mechanism for nucleotide incorporation into steady-state microtubules   总被引:2,自引:0,他引:2  
M Caplow  B P Brylawski  R Reid 《Biochemistry》1984,23(26):6745-6752
We have extended our previous theoretical analysis of the kinetics for radioactive GTP incorporation into steady-state microtubules [Zeeberg, B., Reid, R., & Caplow, M. (1980) J. Biol. Chem. 255, 9891-9899] to include the effects of a kinetic barrier for equilibration of labeled GTP with the tubulin E site. This binding has been found to be relatively slow; the half-time for GTP dissociation is approximately 25 s (k = 0.028 s-1). The slow binding of radioactive GTP apparently accounts for the following observations: (a) more radioactive nucleotide is incorporated into steady-state microtubules in the first 20 s when tubulin-[3H]GTP is used in a pulse than when [3H]GTP is used; (b) when steady-state microtubules are pulsed for 20 s with tubulin-[3H]GTP and then chased with excess nonradioactive GTP, radioactive nucleotide incorporation is not stopped immediately. Quantitative analysis of these results indicates that our steady-state microtubules do not contain significant amounts (greater than 1%) of GDP or GTP which can exchange with added GTP. The principal route for labeled nucleotide incorporation appears to be from tubulin-[3H]GTP subunit uptake, by diffusional and treadmilling processes.  相似文献   

6.
Molecular dynamics simulations were performed on models of the dodecamer DNA double-stranded segment, [d(CGCGAATTCGCG)](2), in which each of the adenine residues, individually or jointly, was replaced by the water-mimicking analog 2'-deoxy-7-(hydroxy-methyl)-7-deazaadenosine (hm(7)c(7)dA) [Rockhill, J.K., Wilson,S.R. and Gumport,R.I. (1996) J. Am. Chem. Soc.,118, 10065-10068]. The simulations, when compared with those of the dodecamer itself, show that incorporation of the analog affects neither the overall DNA structure nor its hydrogen-bonding and stacking interactions when it replaces a single individual base. Furthermore, the water molecules near the bases in the singly-substituted oligonucleotides are similarly unaffected. Double substitutions lead to differences in all the aforementioned parameters with respect to the reference sequence. The results suggest that the analog provides a good mimic of specific 'ordered' water molecules observed in contact with DNA itself and at the interface between protein and DNA in specific complexes.  相似文献   

7.
Earlier work from this laboratory led to the proposal that the cytoplasmic pH of streptococcal cells is regulated solely by changes in the amount and activity of a proton-translocating ATPase, F1F0 complex [Kobayashi, H., Suzuki, T. & Unemoto, T. (1986) J. Biol. Chem. 261, 627-630]. We have now examined the proposal with the aid of computer simulation. We find that an increase in the amount of the H+-ATPase is necessary for pH regulation and is sufficient to maintain a constant steady-state cytoplasmic pH. An increase in H+-ATPase activity is insufficient by itself to maintain a constant cytoplasmic pH, but suppresses the initial fluctuation of the pH. When both variations were allowed, the simulated cytoplasmic pH remained constant despite large perturbations, suggesting that this regulatory system has ample capacity to compensate for pH changes. The present work shows that a computer simulation is a useful way to examine a model for biological regulatory system; application of the simulation to other regulatory systems is discussed.  相似文献   

8.
Baysal C  Meirovitch H 《Biopolymers》2000,53(5):423-433
Using a recently developed statistical mechanics methodology, the solution structures and populations of the cyclic pentapeptide cyclo(D-Pro(1)-Ala(2)-Ala(3)-Ala(4)-Ala(5)) in DMSO are obtained ab initio, i.e., without using experimental restraints. An important ingredient of this methodology is a novel optimization of implicit solvation parameters, which in our previous publication [Baysal, C.; Meirovitch, H. J Am Chem Soc 1998, 120, 800-812] has been applied to a cyclic hexapeptide in DMSO. The molecule has been described by the simplified energy function E(tot) = E(GRO) + summation operator(k) sigma(k)A(k), where E(GRO) is the GROMOS force-field energy, sigma(k) and A(k) are the atomic solvation parameter (ASP) and the solvent accessible surface area of atom k. This methodology, which relies on an extensive conformational search, Monte Carlo simulations, and free energy calculations, is applied here with E(tot) based on the ASPs derived in our previous work, and for comparison also with E(GRO) alone. For both models, entropy effects are found to be significant. For E(tot), the theoretical values of proton-proton distances and (3)J coupling constants agree very well with the NMR results [Mierke, D. F.; Kurz, M.; Kessler, H. J Am Chem Soc 1994, 116, 1042-1049], while the results for E(GRO) are significantly worse. This suggests that our ASPs might be transferrable to other cyclic peptides in DMSO as well, making our methodology a reliable tool for an ab initio structure prediction; obviously, if necessary, parts of this methodology can also be incorporated in a best-fit analysis where experimental restraints are used.  相似文献   

9.
The kinetic behaviour of tyrosinase is very complex because the enzymatic oxidation of monophenol and o-diphenol to o-quinones occurs simultaneously with the coupled non-enzymatic reactions of the latter. Both reaction types are included in the kinetic mechanism proposed for tyrosinase (Mechanism I [J. Biol. Chem. 267 (1992) 3801-3810]). We previously confirmed the validity of the rate equations by the oxidation of numerous monophenols and o-diphenols catalysed by tyrosinase from different fruits and vegetables. Other authors have proposed a simplified reaction mechanism for tyrosinase (Mechanism II [Theor. Biol. 203 (2000) 1-12]), although without deducing the rate equations. In this paper, we report new experimental work that provides the lag period value, the steady-state rate, o-diphenol concentration released to the reaction medium. The contrast between these experimental data and the respective numerical simulations of both mechanisms demonstrates the feasibility of Mechanism I. The need for the steps omitted from Mechanism II to interpret the experimental data for tyrosinase, based on the rate equations previously deduced for Mechanism I is explained.  相似文献   

10.
For the analysis of coding mechanisms in the insect olfactory system, a fully connected network of synchronously updated McCulloch and Pitts neurons (MC-P type) was developed [Quenet, B., Horn, D., 2003. The dynamic neural filter: a binary model of spatio-temporal coding. Neural Comput. 15 (2), 309-329]. Considering the update time as an intrinsic clock, this "Dynamic Neural Filter" (DNF), which maps regions of input space into spatio-temporal sequences of neuronal activity, is able to produce exact binary codes extracted from the synchronized activities recorded at the level of projection neurons (PN) in the locust antennal lobe (AL) in response to different odors [Wehr, M., Laurent, G., 1996. Odor encoding by temporal sequences of firing in oscillating neural assemblies. Nature 384, 162-166]. Here, in a first step, we separate the populations of PN and local inhibitory neurons (LN) and use the DNF as a guide for simulations based on biological plausible neurons (Hodgkin-Huxley: H-H type). We show that a parsimonious network of 10 H-H neurons generates action potentials whose timing represents the required codes. In a second step, we construct a new type of DNF in order to study the population dynamics when different delays are taken into account. We find synaptic matrices which lead to both the emergence of robust oscillations and spatio-temporal patterns, using a formal criterion, based on a Normalized Euclidian Distance (NED), in order to measure the use of the temporal dimension as a coding dimension by the DNF. Similarly to biological PN, the activity of excitatory neurons in the model can be both phase-locked to different cycles of oscillations which remind local field potential (LFP), and nevertheless exhibit dynamic behavior complex enough to be the basis of spatio-temporal codes.  相似文献   

11.
Lateral diffusion is an essential process for the functioning of biological membranes. Solid-state nuclear magnetic resonance (NMR) is, a priori, a well-suited technique to study lateral diffusion within a heterogeneous environment such as the cell membrane. Moreover, restriction of lateral motions by lateral heterogeneities can be used as a means to characterize their geometry. The goal of this work is to understand the advantages and limitations of solid-state NMR exchange experiments in the study of obstructed lateral diffusion in model membranes. For this purpose, simulations of lateral diffusion on a sphere with varying numbers and sizes of immobile obstacles and different percolation properties were performed. From the results of these simulations, two-dimensional 31P NMR exchange maps and time-dependent autocorrelation functions were calculated. The results indicate that the technique is highly sensitive to percolation properties, total obstacle area, and, within certain limits, obstacle size. A practical example is shown, namely the study of the well-characterized DMPC-DSPC binary mixture. The comparison of experimental and simulated results yielded obstacle sizes in the range of hundreds of nanometers, therefore bridging the gap between previously published NMR and fluorescence recovery after photobleaching results. The method could also be applied to the study of membrane protein lateral diffusion in model membranes.  相似文献   

12.
Around half of the neurons of a human brain are granule cells (approximately 10(11)granule neurons) [Kandel, E.R., Schwartz, J.H., Jessell, T.M., 2000. Principles of Neural Science. McGraw-Hill Professional Publishing, New York]. In order to study in detail the functional role of the intrinsic features of this cell we have developed a pre-compiled behavioural model based on the simplified granule-cell model of Bezzi et al. [Bezzi, M., Nieus, T., Arleo, A., D'Angelo, E., Coenen, O.J.-M.D., 2004. Information transfer at the mossy fiber-granule cell synapse of the cerebellum. 34th Annual Meeting. Society for Neuroscience, San Diego, CA, USA]. We can use an efficient event-driven simulation scheme based on lookup tables (EDLUT) [Ros, E., Carrillo, R.R., Ortigosa, E.M., Barbour, B., Ags, R., 2006. Event-driven simulation scheme for spiking neural networks using lookup tables to characterize neuronal dynamics. Neural Computation 18 (12), 2959-2993]. For this purpose it is necessary to compile into tables the data obtained through a massive numerical calculation of the simplified cell model. This allows network simulations requiring minimal numerical calculation. There are three major features that are considered functionally relevant in the simplified granule cell model: bursting, subthreshold oscillations and resonance. In this work we describe how the cell model is compiled into tables keeping these key properties of the neuron model.  相似文献   

13.
The expression of certain COOH-terminal truncation mutants of the epidermal growth factor receptor (EGFR) can lead to cell transformation, and with ligand stimulation, a broader spectrum of phosphorylated proteins appears compared with EGF-treated cells expressing wild-type EGFR. Accordingly, it has been proposed that elements within the COOH terminus may determine substrate specificity of the EGFR tyrosine kinase (Decker, S. J., Alexander, C., and Habib, T. (1992) J. Biol. Chem. 267, 1104-1108; Walton, G. M., Chen, W. S., Rosenfeld, M. G., and Gill, G. N. (1990) J. Biol. Chem. 265, 1750-1754). To address this hypothesis, we analyzed in vitro the steady-state kinetic parameters for phosphorylation of several substrates by both wild-type EGFR and an oncogenic EGFR mutant (the ct1022 mutant) truncated at residue 1022. The substrates included: (i) a phospholipase C-gamma fragment (residues 530-850); (ii) the 46-kDa isoform of the Shc adapter protein; (iii) a 13-residue peptide mimic for the region around the major autophosphorylation tyrosine and the Shc binding site (the Y1173 peptide); (iv) a poly(Glu,Tyr) 4:1 copolymer; and (v) the 8-residue peptide, angiotensin II. Our data demonstrate that the steady-state kinetic parameters for the ct1022 mutant differ from those of the wild-type enzyme, and the differences are substrate-dependent. These results support the concept that this oncogenic truncation/mutation alters EGFR substrate specificity, rather than causing a general alteration of activity. We performed the experiments using a non-radioactive fluorescence polarization assay that quantifies the degree of phosphorylation of peptide as well as natural substrates. The results are consistent with those from the traditional [gamma-32P]ATP/filtration assay.  相似文献   

14.
The dynamics of threonine side chains of the Tenebrio molitor antifreeze protein (TmAFP) were investigated using natural abundance (13)C NMR. In TmAFP, the array of threonine residues on one face of the protein is responsible for conferring its ability to bind crystalline ice and inhibit its growth. Heteronuclear longitudinal and transverse relaxation rates and the [(1)H]-(13)C NOE were determined in this study. The C alpha H relaxation measurements were compared to the previously measured (15)N backbone parameters and these are found to be in agreement. For the analysis of the threonine side chain motions, the model of restricted rotational diffusion about the chi(1) dihedral angle was employed [London and Avitabile (1978) J. Am. Chem. Soc., 100, 7159-7165]. We demonstrate that the motion experienced by the ice binding threonine side chains is highly restricted, with an approximate upper limit of less than +/-25 degrees.  相似文献   

15.
Rates of NO uptake by five denitrifying bacteria were estimated by NO-electrode and gas chromatography methods under conditions of rather low cell densities and [NOaq]. The rates so measured, VmaxNO, represent lower limits for the true value of that parameter, but nevertheless exceed Vmax for nitrite uptake, VmaxNi, by a factor of two typically. Previous estimates under suboptimal conditions had placed VmaxNO at 0.3-0.5 of VmaxNi (St. John, R. T., and Hollocher, T. C. (1977) J. Biol. Chem. 252, 212-218; Garber, E. A. E., and Hollocher, T.C. (1981) J. Biol. Chem. 256, 5459-5465). The steady-state [NOaq] during denitrification of nitrite by nitrate-grown cells was less than or equal to 1 microM. The above observations, taken with a recent direct estimate for the KmNO for NO uptake of 0.4 microM (Zafiriou, O. C., Hanley, Q. S., and Snyder, G. (1989) J. Biol. Chem. 264, 5694-5699), would allow NO to be a free intermediate between nitrite and N2O with steady-state concentrations of less than or equal to 0.4 microM. As the result of special conditions during cell growth or differential inhibition by azide, it was possible to establish systems that accumulated NO during denitrification of nitrite. In all such cases, VmaxNO less than VmaxNi, and the time required to reach the maximum [NOaq] corresponded closely to the time needed to exhaust the nitrite. A semiquantitative isotope experiment with Paracoccus denitrificans demonstrated the trapping of 15NO from 15NO2- in a pool of NOaq. A quantitative isotope method using low densities of the same bacterium showed that 15N from 15NO2- and 14N from NOg combine randomly to form N2O during the simultaneous denitrification of 15NO2- and NO. The result requires that the pathways from nitrite and NO share a common mononitrogen intermediate. Results to the contrary obtained at high cell densities (first two references cited above) are now believed to have been due to technical artifacts. The present results are consistent with the view that NO is under kinetic control as a free intermediate in denitrification and serve to remove previously imagined constraints on this view.  相似文献   

16.
An extension of the anisotropic united atoms intermolecular potential model is proposed for nitriles. The electrostatic part of the intermolecular potential is calculated using atomic charges obtained by a simple Mulliken population analysis. The repulsion-dispersion interaction parameters for methyl and methylene groups are taken from transferable AUA4 literature parameters [Ungerer et al., J. Chem. Phys., 2000, 112, 5499]. Non-bonding Lennard-Jones intermolecular potential parameters are regressed for the carbon and nitrogen atoms of the nitrile group (–C≡N) from experimental vapor-liquid equilibrium data of acetonitrile. Gibbs Ensemble Monte Carlo simulations and experimental data agreement is very good for acetonitrile, and better than previous molecular potential proposed by Hloucha et al. [J. Chem. Phys., 2000, 113, 5401]. The transferability of the resulting potential is then successfully tested, without any further readjustment, to predict vapor-liquid phase equilibrium of propionitrile and n-butyronitrile. Figure Saturated vapour pressure of nitriles calculated in this work by molecular simulation compared to experimental data: a) for acetonitrile and b) for both propionitrile and butyronitrile  相似文献   

17.
An automatic method for deriving steady-state rate equations.   总被引:6,自引:6,他引:0       下载免费PDF全文
A method is described for systematically deriving steady-state rate equations. It is based on the schematic method of King & Altman [J. Phys. Chem. (1956) 60, 1375-1378], but is expressed in purely algebraic terms. It is suitable for implementation as a computer program, and a program has been written in FORTRAN IV and deposited as Supplementary Publication SUP 50078 (12 pages) at the British Library (Lending Division), Boston Spa, Wetherby, West Yorkshire LS23 7BQ, U.K., from whom copies can be obtained on the terms indicated in Biochem. J. (1977) 161, 1-2.  相似文献   

18.
Modifications to the AMBER force field [W.D. Cornell, P. Cieplak, C.I. Bayly, I.R. Gould, K. Merz, D.M. Ferguson, D.C. Spellmeyer, T. Fox, J.W. Caldwell, P.A. Kollman, J. Am. Chem. Soc., 117 (1995) 5179-5197] have been made to improve our ability to reproduce observed molecular properties of alpha-linked carbohydrates when calculated using empirical potential-energy functions. Molecular structures and energies obtained using gradient-optimized density functional methods with ab initio basis sets (B3LYP/6-31G*) on ten minimum-energy conformations of maltose [F.A. Momany, J.L. Willett, J. Comp. Chem., submitted for publication] were used to refine the empirical potentials. Molecular dynamics simulations on beta-maltose (i.e., the beta anomer of maltose), cyclohexamylose (alpha-cyclodextrin), cycloheptamylose (beta-cyclodextrin) and larger cyclomaltooligosaccharide structures were carried out and compared with experimental structural studies to test the new potentials. Ring-puckering potential during dynamics as well as conformational transitions to 'flipped' structures were examined. Results of the tests described here suggest that the revised AMBER parameters (AMB99C) are very good for computational studies of alpha-(1-->4)-linked carbohydrates.  相似文献   

19.
Mobility in the mitochondrial electron transport chain   总被引:1,自引:0,他引:1  
The role of lateral diffusion in mitochondrial electron transport has been investigated by measuring the diffusion coefficients for lipid, cytochrome c, and cytochrome oxidase in membranes of giant mitoplasts from cuprizone-fed mice using the technique of fluorescence redistribution after photobleaching (FRAP). The diffusion coefficient of the phospholipid analogue N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine is dependent on the technique used to remove the outer mitochondrial membrane. A sonication technique yields mitoplasts with monophasic recovery of the lipid probe (D = 6 X 10(-9) cm2/s), while digitonin-treated mitochondria show biphasic recoveries (D1 = 5 X 10(-9) cm2/s; D2 = 1 X 10(-9) cm2/s). Digitonin appears to incorporate into mitoplasts, giving rise to decreased lipid mobility concomitant with increased rates of electron transfer from succinate to oxygen, in a manner reminiscent of the effects of cholesterol incorporation [Schneider, H., Lemasters, J. J., Hochli, M., & Hackenbrock, C. R. (1980) J. Biol. Chem. 255, 3748-3756]. FRAP measurements on tetramethylrhodamine cytochrome c modified at lysine-39 and on a mixture of active morpholinorhodamine derivatives of cytochrome c gave diffusion coefficients of (3.5-7) X 10(-10) cm2/s depending on the assay medium. With morpholinorhodamine-labeled antibodies purified on a cytochrome oxidase affinity column, the diffusion coefficient for cytochrome oxidase was determined to be 1.5 X 10(-10) cm2/s. The results are discussed in terms of a dynamic aggregate model in which an equilibrium exists between freely diffusing and associated electron-transfer components.  相似文献   

20.
A thermodynamic model is presented to describe the redox behaviour of the tetraheme cytochrome c3 from Desulfovibrio gigas. This molecule displays different intrinsic redox potentials for the four hemes and during the redox titration process, interactions among different hemes occur, thus altering the values of redox potentials according to which of the hemes are oxidized [Santos, H., Moura, J.J.G., Moura, I., LeGall, J. & Xavier, A.V. (1984) Eur. J. Biochem. 141, 283-296]. This complex cooperative behaviour [Xavier, A.V. (1986) J. Inorg. Biochem. 28, 239-243] has been analyzed here using an I2H4-interaction network [Cornish-Bowden, A. & Koshland, D.E. Jr (1970) J. Biol. Chem. 245, 6241-6250] coupled to a proton-linked equilibrium between two tertiary structures. Such a formalism, which requires a reduced number of parameters, is able to fully account quantitatively for the pH dependence of the NMR redox-titration curves. The 'redox-Bohr' effect is discussed in terms of the available structure and thermodynamic data and a functional mechanism is proposed.  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