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 15 毫秒
1.
A well-characterized experimental system was designed to evaluate the effect of porous media on macroscopic transport coefficients which are used to characterize the migration of bacterial populations. Bacterial density profiles of Pseudomonas putida PRS2000 were determined in the presence and absence of a chemical attractant (3-chlorobenzoate) gradient within sand columns having a narrow distribution of particle diameters. These experimental profiles were compared with theoretical predictions to evaluate the macroscopic transport coefficients. The effective random motility coefficient, used to quantify migration due to a random process in a porous medium, decreased nearly 20-fold as grain size in the columns decreased from 800 to 80 (mu)m. The effective random motility coefficient (mu)(infeff) was related to the random motility coefficient (mu), measured in a bulk aqueous system, according to (mu)(infeff) = ((epsilon)/(tau))(mu) with porosity (epsilon) and tortuosity (tau). Over the times and distances examined in these experiments, bacterial density profiles were unaffected by the presence of an attractant gradient. Theoretical profiles with the aqueous phase value of the chemotactic sensitivity coefficient (used to quantify migration due to a directed process) were consistent with this result and suggested that any chemotactic effect on bacterial migration was below the detection limits of our assay.  相似文献   

2.
We investigate the possibility that atomic burials, as measured by their distances from the structural geometrical center, contain sufficient information to determine the tertiary structure of globular proteins. We report Monte Carlo simulated annealing results of all-atom hard-sphere models in continuous space for four small proteins: the all-beta WW-domain 1E0L, the alpha/beta protein-G 1IGD, the all-alpha engrailed homeo-domain 1ENH, and the alpha + beta engineered monomeric form of the Cro protein 1ORC. We used as energy function the sum over all atoms, labeled by i, of |R(i) - R(i) (*)|, where R(i) is the atomic distance from the center of coordinates, or central distance, and R(i) (*) is the "ideal" central distance obtained from the native structure. Hydrogen bonds were taken into consideration by the assignment of two ideal distances for backbone atoms forming hydrogen bonds in the native structure depending on the formation of a geometrically defined bond, independently of bond partner. Lowest energy final conformations turned out to be very similar to the native structure for the four proteins under investigation and a strong correlation was observed between energy and distance root mean square deviation (DRMS) from the native in the case of all-beta 1E0L and alpha/beta 1IGD. For all alpha 1ENH and alpha + beta 1ORC the overall correlation between energy and DRMS among final conformations was not as high because some trajectories resulted in high DRMS but low energy final conformations in which alpha-helices adopted a non-native mutual orientation. Comparison between central distances and actual accessible surface areas corroborated the implicit assumption of correlation between these two quantities. The Z-score obtained with this native-centric potential in the discrimination of native 1ORC from a set of random compact structures confirmed that it contains a much smaller amount of native information when compared to a traditional contact Go potential but indicated that simple sequence-dependent burial potentials still need some improvement in order to attain a similar discriminability. Taken together, our results suggest that central distances, in conjunction to physically motivated hydrogen bond constraints, contain sufficient information to determine the native conformation of these small proteins and that a solution to the folding problem for globular proteins could arise from sufficiently accurate burial predictions from sequence followed by minimization of a burial-dependent energy function.  相似文献   

3.
Energetics of charge-charge interactions in proteins   总被引:21,自引:0,他引:21  
M K Gilson  B H Honig 《Proteins》1988,3(1):32-52
Electrostatic interactions between pairs of atoms in proteins are calculated with a model based on the linearized Poisson-Boltzmann equation. The equation is solved accurately by a method that takes into account the detailed shape of the protein. This paper presents applications to several systems. Experimental data for the interaction of ionized residues with an active site histidine in subtilisin BPN' allow the model to be tested, using various assumptions for the electrical properties of the protein and solvent. The electrostatic stabilization of the active site thiolate of rhodanese is analyzed, with attention to the influence of alpha-helices. Finally, relationships between electrostatic potential and charge-charge distance are reported for large and small globular proteins. The above results are compared with those of simpler electrostatic models, including Coulomb's law with both a distance-dependent dielectric constant (epsilon = R) and a fixed dielectric constant (epsilon = 2), and Tanford-Kirkwood theory. The primary conclusions are as follows: 1) The Poisson-Boltzmann model agrees with the subtilisin data over a range of ionic strengths; 2) two alpha-helices generate a large potential in the active site of rhodanese; 3) epsilon = R overestimates weak electrostatic interactions but yields relatively good results for strong ones; 4) Tanford-Kirkwood theory is a useful approximation to detailed solutions of the linearized Poisson-Boltzmann equation in globular proteins; and 5) the modified Tanford-Kirkwood theory over-screens the measured electrostatic interactions in subtilisin.  相似文献   

4.
We considered an irreversible biochemical intrachain reaction of supercoiled DNA as a random event that occurs, with certain probability, at the instant of collision between two reactive groups bound to distant DNA sites. Using the Brownian dynamics technique, we modeled this process for a supercoiled DNA molecule of 2.5 kb length in dilute aqueous solution at an NaCl concentration of 0.1 M. We calculated the mean reaction time tau(Sigma) as a function of the intrinsic second-order rate constant k(I), the reaction radius R, and the contour separation S of the reactive groups. At the diffusion-controlled limit (k(I) --> infinity), the kinetics of reaction are determined by the mean time tau(F) of the first collision. The dependence of tau(F) on R is close to inversely proportional, implying that the main contribution to the productive collisions is made by bending of the superhelix axis. At sufficiently small k(I), the mean reaction time can be satisfactory approximated by tau(Sigma) = tau(F)(app) + 1/(k(I)c(L)), where c(L) is the local concentration of one reactive group around the other, and tau is an adjustable parameter, which we called the apparent time of the first collision. The value of tau depends on R very weakly and is approximately equal to the mean time of the first collision caused by mutual reptation of two DNA strands forming the superhelix. The quasi-one-dimensional reptation process provides the majority of productive collisions at small k(I) values.  相似文献   

5.
The bichromophoric pigment bilirubin acts as a molecular exciton in its UV-visible and circular dichroism (CD) spectroscopy. In both polar and nonpolar solvents, an optically active analog, (beta R,beta 'R)-dimethylmesobilirubin-XIII alpha (1), exhibits intense bisignate CD Cotton effects in the region of its long wavelength UV-vis absorption near 400 nm: Delta epsilon(434)(max) + 337, Delta epsilon(389)(max) - 186 (CHCl(3)), and Delta epsilon(431)(max) + 285, Delta epsilon(386)(max) - 177 (CH(3)OH). However, introduction of an amine into a CHCl(3) solution of 1 causes the Cotton effect signs to become inverted, e.g., after addition of NH(3), Delta epsilon(433)(max) - 345, Delta epsilon(389)(max) + 243, and after addition of ethylene diamine, Delta epsilon(435)(max) - 420, Delta epsilon(390)(max) + 299. The sign inversions imply inversion of molecular chirality of the bilirubin and the phenomenon appears to be general for amines, including alpha,omega-diamines. 1,8-Diaminooctane was found to be more effective than longer or shorter chain analogs in producing CD sign inversion.  相似文献   

6.
In contact with lipid bilayers and Ca2+-ions, the intracellular protein human annexin V (wild-type), Mr = 35,800, forms two types of cation-selective channels for the transport of Ca2+-, K+-, Na+- and Mg2+-ions, depending on the protein concentration [AN]. Type (I) channel events are large and predominant at high values [AN] > or = K = 5 nM at 296 K. At 50 mM Ca2+, symmetrical on both membrane sides, AN added at the cis side, the conductance is gCa(I) = 22 +/- 2 pS and at symmetrical 0.1 M K+-conditions: gK(I) = 32 +/- 3 pS, associated with two mean open-times tau1(I) = 0.68 +/- 0.2 ms and tau2(I) = 31 +/- 2 ms. Monoclonal anti-AN antibodies added to the trans-side first increase the mean open-times and then abolish the channel activity, suggesting that type (I) channels refer to a membrane spanning protein complex, probably a trimer T, which at [AN] > K changes its membrane organization to a higher oligomer, probably to the side-by-side double-trimer T2. The smaller type (II) channel events are predominant at low [AN] < or = K and refer to the (electroporative) adsorption complex of the monomer. The conductances g(i)(II) for symmetrical concentrations depend non-linearly on the voltage Um = Uext + U(AN), where U(AN) = 0.02 +/- 0.002 V is the electrostatic contribution of the Ca2+-AN complex and Uext the externally applied voltage. There is only one mean open-time tau(o)(II) which is voltage-dependent according to a functional of b x Um2 where b = 113.9 +/- 15 V(-2), yielding an activation Gibbs free energy of Ga = RT x b x Um2. The conformational flicker probability f(i)(II) in g(i)(II) = g(i)0(II) x gamma(i) x f(i)(II) is non-linearly voltage-dependent according to a functional of a x Um2. The Nernst term gamma(i) refers to asymmetrical ion concentrations. From a = 50 V(-2), independent of the ion type, we obtain f(i)0(II) = 0.03 +/- 0.002 and the conductances for the fully open-channel states: gCa0(II) = 69 +/- 3 pS (0.05 M Ca2+) and gK0(II) = 131 +/- 5 pS (1.2 M K+). From the electroporation term a = pi[r(p)2]epsilon0(epsilon(w) - epsilon(m))/(2 kTd) we determine the mean pore radius of the complex in its fully open state as r(p)= 0.86 +/- 0.05 nm. The adsorbed annexin V (Ca2+) monomer appears to electrostatically facilitate the electric pore formation at the contact interface between the protein and the lipid phase. The complex rapidly flickers and thus limits the ion transport in a voltage-dependent manner.  相似文献   

7.
Strain echocardiography (SE) provides the rate and extent of myocardial segment shortening and lengthening. Thus we hypothesized that SE will noninvasively provide estimates of shortening velocity (SV) and length change (delta L). We compared SE-derived strain rate (SR) and strain (epsilon) to force/length transducer-derived SV and percent delta L in isolated muscle strips at multiple load levels and under varying conditions. Electrically stimulated contractions in left ventricular muscle strips (n = 20) were simultaneously recorded with a force/length transducer (to measure SV and percent delta L) and an ultrasound transducer (to measure SR and epsilon). Measurements were made at baseline, after inotropic stimulation, and during hypoxia at multiple load clamp levels (simulating multiple levels of afterload). Despite a difference in absolute numbers, there was a close correlation between SR and SV at baseline (R2 = 0.95), with dobutamine treatment (R2 = 0.99), and during hypoxia (R2 = 0.99). SR was load dependent at baseline (r = 0.98), with dobutamine treatment (r = 0.99), and during hypoxia (r = 0.92). Similarly, there was a close correlation between epsilon and delta L at baseline (R2 = 0.99), with dobutamine treatment (R2 = 0.96), and during hypoxia (R2 = 0.87). Percent epsilon was load dependent at baseline (r = 0.98), with dobutamine treatment (r = 0.98), and during hypoxia (r = 0.94). Bland-Altman analysis revealed a systematic overestimation of SV by SE-derived SR at baseline and with dobutamine treatment. There was no bias with SR measurements during hypoxia or with epsilon measurements. SE closely tracks standard physiological parameters of regional contractile function, such as SV and delta L, under conditions of varying afterload.  相似文献   

8.
The energy cost of kayaking per unit distance (C(k), kJ x m(-1)) was assessed in eight middle- to high-class athletes (three males and five females; 45-76 kg body mass; 1.50-1.88 m height; 15-32 years of age) at submaximal and maximal speeds. At submaximal speeds, C(k) was measured by dividing the steady-state oxygen consumption (VO(2), l x s(-1)) by the speed (v, m x s(-1)), assuming an energy equivalent of 20.9 kJ x l O(-1)(2). At maximal speeds, C(k) was calculated from the ratio of the total metabolic energy expenditure (E, kJ) to the distance (d, m). E was assumed to be the sum of three terms, as originally proposed by Wilkie (1980): E = AnS + alphaVO(2max) x t-alphaVO(2max) x tau(1-e(-t x tau(-1))), were alpha is the energy equivalent of O(2) (20.9 kJ x l O(2)(-1)), tau is the time constant with which VO(2max) is attained at the onset of exercise at the muscular level, AnS is the amount of energy derived from anaerobic energy utilization, t is the performance time, and VO(2max) is the net maximal VO(2). Individual VO(2max) was obtained from the VO(2) measured during the last minute of the 1000-m or 2000-m maximal run. The average metabolic power output (E, kW) amounted to 141% and 102% of the individual maximal aerobic power (VO(2max)) from the shortest (250 m) to the longest (2000 m) distance, respectively. The average (SD) power provided by oxidative processes increased with the distance covered [from 0.64 (0.14) kW at 250 m to 1.02 (0.31) kW at 2000 m], whereas that provided by anaerobic sources showed the opposite trend. The net C(k) was a continuous power function of the speed over the entire range of velocities from 2.88 to 4.45 m x s(-1): C(k) = 0.02 x v(2.26) (r = 0.937, n = 32).  相似文献   

9.
The evaluation of lateral diffusion coefficients of membrane components by the technique of fluorescence recovery after photobleaching (FRAP) is often complicated by uncertainties in the values of the intensities F(O), immediately after bleaching, and F(infinity), after full recovery. These uncertainties arise from instrumental settling time immediately after bleaching and from cell, tissue, microscope, or laser beam movements at the long times required to measure F(infinity). We have developed a method for precise analysis of FRAP data that minimizes these problems. The method is based on the observation that a plot of the reciprocal function R(tau) = F(infinity)/[F(infinity)-F(tau)] is linear over a large time range when (a) the laser beam has a Gaussian profile, (b) recovery involves a single diffusion coefficient, and (c) there is no membrane flow. Moreover, the ratio of intercept to slope of the linear plot is equal to tau 1/2, the time required for the bleached fluorescence to rise to 50% of the full recovery value, F(infinity). The lateral diffusion coefficient D is related to tau 1/2 by tau 1/2 = beta w2/4D where beta is a defined parameter and w is the effective radius of the focused laser beam. These results are shown to indicate that the recovery of fluorescence F(tau) can be represented over a large range of percent bleach, and recovery time tau by the relatively simple expression F(tau) = [ F(o) + F(infinity) (tau/tau 1/2)]/[1 + tau/tau 1/2)]. FRAP data can therefore be easily evaluated by a nonlinear regression analysis with this equation or by a linear fit to the reciprocal function R(tau). It is shown that any error in F(infinity) can be easily detected in a plot of R(tau) vs. tau which deviates significantly from a straight line when F(infinity) is in error by as little as 5%. A scheme for evaluating D by linear analysis is presented. It is also shown that the linear reciprocal plot provides a simple method for detecting flow or multiple diffusion coefficients and for establishing conditions (data precision, differences in multiple diffusion coefficients, magnitude of flow rate compared to lateral diffusion) under which flow or multiple diffusion coefficients can be detected. These aspects are discussed in some detail.  相似文献   

10.
The affinities of Mg(2+) for various substituted carbonyl ligands were determined at the DFT (B3LYP/6-31+G(d)) and semi-empirical (PM6) levels of theory. Two sets of carbonyl ligands were studied: monosubstituted [aldehydes R-CHO and RPh-CHO] and homodisubstituted [ketones R(2)C=O and (RPh)(2)C=O], where R = NH(2), OCH(3), OH, CH(3), H, F, Cl, Br, CN, or NO(2)). In the (RPh)(2)CO case, the R group was bonded to the para position of a phenyl ring. The enthalpies of interaction between the ligands and a pentaaquomagnesium(II) complex were calculated to determine the affinity of each ligand for the Mg(2+) cation and to correlate with geometrical and electronic parameters. These parameters exhibited the same trends for all of the ligands studied, showing that the affinity of Mg(2+) for electron-donating ligands is higher than its affinity for electron-withdrawing ligands. In the complexes, electron-donating groups increase both the electrostatic and the covalent components of the Mg-ligand interaction. This behavior correlates with the Mg-O(carbonyl) distance and the ligand electron-donor strength.  相似文献   

11.
To interpret effects of urea and guanidinium (GuH(+)) salts on processes that involve large changes in protein water-accessible surface area (ASA), and to predict these effects from structural information, a thermodynamic characterization of the interactions of these solutes with different types of protein surface is required. In the present work we quantify the interactions of urea, GuHCl, GuHSCN, and, for comparison, KCl with native bovine serum albumin (BSA) surface, using vapor pressure osmometry (VPO) to obtain preferential interaction coefficients (Gamma(mu3)) as functions of nondenaturing concentrations of these solutes (0-1 molal). From analysis of Gamma(mu3) using the local-bulk domain model, we obtain concentration-independent partition coefficients K(nat)(P) that characterize the accumulation of these solutes near native protein (BSA) surface: K(nat)(P,urea)= 1.10 +/- 0.04, K(nat)(P,SCN(-)) = 2.4 +/- 0.2, K(nat)(P,GuH(+)) = 1.60 +/- 0.08, relative to K(nat)(P,K(+)) identical with 1 and K(nat)(P,Cl(-)) = 1.0 +/- 0.08. The relative magnitudes of K(nat)(P) are consistent with the relative effectiveness of these solutes as perturbants of protein processes. From a comparison of partition coefficients for these solutes and native surface (K(nat)(P)) with those determined by us previously for unfolded protein and alanine-based peptide surface K(unf)(P), we dissect K(P) into contributions from polar peptide backbone and other types of protein surface. For globular protein-urea interactions, we find K(nat)(P,urea) = K(unf)(P,urea). We propose that this equality arises because polar peptide backbone is the same fraction (0.13) of total ASA for both classes of surface. The analysis presented here quantifies and provides a physical basis for understanding Hofmeister effects of salt ions and the effects of uncharged solutes on protein processes in terms of K(P) and the change in protein ASA.  相似文献   

12.
Visualization of flowing neutrophils colliding with adherent 1-mum-diameter beads presenting P-selectin allowed the simultaneous measurement of collision efficiency (epsilon), membrane tethering fraction (f), membrane tether growth dynamics, and PSGL-1/P-selectin binding lifetime. For 1391 collisions analyzed over venous wall shear rates from 25 to 200 s(-1), epsilon decreased from 0.17 to 0.004, whereas f increased from 0.15 to 0.70, and the average projected membrane tether length, L(tether)(m), increased from 0.35 mum to approximately 2.0 mum over this shear range. At all shear rates tested, adhesive collisions lacking membrane tethers had average bond lifetimes less than those observed for collisions with tethers. For adhesive collisions that failed to form membrane tethers, the regressed Bell parameters (consistent with single bond Monte Carlo simulation) were zero-stress off-rate, k(off)(0) = 0.56 s(-1) and reactive compliance, r = 0.10 nm, similar to published atomic force microscopy (AFM) measurements. For all adhesion events (+/- tethers), the bond lifetime distributions were more similar to those obtained by rolling assay and best simulated by Monte Carlo with the above Bell parameters and an average of 1.48 bonds (n = 1 bond (67%), n = 2 (22%), and n = 3-5 (11%)). For collisions at 100 s(-1), pretreatment of neutrophils with actin depolymerizing agents, latrunculin or cytochalasin D, had no effect on epsilon, but increased L(tether)(m) by 1.74- or 2.65-fold and prolonged the average tether lifetime by 1.41- or 1.65-fold, respectively. Jasplakinolide, an actin polymerizing agent known to cause blebbing, yielded results similar to the depolymerizing agents. Conversely, cholesterol-depletion with methyl-beta-cyclodextrin or formaldehyde fixation had no effect on epsilon, but reduced L(tether)(m) by 66% or 97% and reduced the average tether lifetime by 30% or 42%, respectively. The neutrophil-bead collision assay combines advantages of atomic force microscopy (small contact zone), aggregometry (discrete interactions), micropipette manipulation (tether visualization), and rolling assays (physiologic flow loading). Membrane tether growth can be enhanced or reduced pharmacologically with consequent effects on PSGL-1/P-selectin lifetimes.  相似文献   

13.
Hydration water is vital for various macromolecular biological activities, such as specific ligand recognition, enzyme activity, response to receptor binding, and energy transduction. Without hydration water, proteins would not fold correctly and would lack the conformational flexibility that animates their three-dimensional structures. Motions in globular, soluble proteins are thought to be governed to a certain extent by hydration-water dynamics, yet it is not known whether this relationship holds true for other protein classes in general and whether, in turn, the structural nature of a protein also influences water motions. Here, we provide insight into the coupling between hydration-water dynamics and atomic motions in intrinsically disordered proteins (IDP), a largely unexplored class of proteins that, in contrast to folded proteins, lack a well-defined three-dimensional structure. We investigated the human IDP tau, which is involved in the pathogenic processes accompanying Alzheimer disease. Combining neutron scattering and protein perdeuteration, we found similar atomic mean-square displacements over a large temperature range for the tau protein and its hydration water, indicating intimate coupling between them. This is in contrast to the behavior of folded proteins of similar molecular weight, such as the globular, soluble maltose-binding protein and the membrane protein bacteriorhodopsin, which display moderate to weak coupling, respectively. The extracted mean square displacements also reveal a greater motional flexibility of IDP compared with globular, folded proteins and more restricted water motions on the IDP surface. The results provide evidence that protein and hydration-water motions mutually affect and shape each other, and that there is a gradient of coupling across different protein classes that may play a functional role in macromolecular activity in a cellular context.  相似文献   

14.
A mathematical problem relating to membrane cylinders is stated and solved; its implications are illustrated and discussed. The problem concerns the volume distribution, in cylindrical coordinates, of the electric potential inside and outside a membrane cylinder of finite length (with sealed ends), during passive decay of an initially nonuniform membrane potential. The time constants for equalization with respect to the angle, theta, are shown to be typically about ten thousand times smaller than the time constant, tau(m) = R(m)C(m), for uniform passive membrane potential decay. The time constants for equalization with respect to length are shown to agree with those from one-dimensional cable theory; typically, they are smaller than tau(m) by a factor between 2 and 10. The relation of the membrane current density, I(m)(theta, x, t), to the values (at the outer membrane surface) of the extracellular potential phi(e)(r, theta, x, t) and of partial differential(2)phi(e)/ partial differentialx(2), is examined and it is shown that these quantities are not proportional to each other, in general; however, under certain specified conditions, all three of these quantities are proportional with each other and with phi(i)(r, theta, x, t) and partial differential(2)phi(i)/ partial differentialx(2) (at the inner membrane surface). The relation of these results to those of one-dimensional cable theory is discussed.  相似文献   

15.
Das B  Meirovitch H 《Proteins》2001,43(3):303-314
A novel procedure for optimizing the atomic solvation parameters (ASPs) sigma(i) developed recently for cyclic peptides is extended to surface loops in proteins. The loop is free to move, whereas the protein template is held fixed in its X-ray structure. The energy is E(tot) = E(FF)(epsilon = nr) + summation operator sigma(i)A(i), where E(FF)(epsilon = nr) is the force-field energy of the loop-loop and loop-template interactions, epsilon = nr is a distance-dependent dielectric constant, and n is an additional parameter to be optimized. A(i) is the solvent-accessible surface area of atom i. The optimal sigma(i) and n are those for which the loop structure with the global minimum of E(tot)(n, sigma(i)) becomes the experimental X-ray structure. Thus, the ASPs depend on the force field and are optimized in the protein environment, unlike commonly used ASPs such as those of Wesson and Eisenberg (Protein Sci 1992;1:227-235). The latter are based on the free energy of transfer of small molecules from the gas phase to water and have been traditionally combined with various force fields without further calibration. We found that for loops the all-atom AMBER force field performed better than OPLS and CHARMM22. Two sets of ASPs [based on AMBER (n = 2)], optimized independently for loops 64-71 and 89-97 of ribonuclease A, were similar and thus enabled the definition of a best-fit set. All these ASPs were negative (hydrophilic), including those for carbon. Very good (i.e., small) root-mean-square-deviation values from the X-ray loop structure were obtained with the three sets of ASPs, suggesting that the best-fit set would be transferable to loops in other proteins as well. The structure of loop 13-24 is relatively stretched and was insensitive to the effect of the ASPs.  相似文献   

16.
The effect of the protein matrix on the standard potential of a buried redox center has been investigated by using a selection of mutants and chemical derivatives in Saccharomyces cerevisiae cytochrome c isoform 1. Assuming only local structural perturbation and no alteration of the iron-ligation chemistry, Delta E(m)(0)' can be regarded as a measure of the difference in polypeptide solvation of the heme charge, which reflects the dielectric properties of the protein. The evaluation of an apparent dielectric constant (U(exp)/U(theo)) yields variable, and sometimes even negative, values if U(exp) = Delta G(0)redox. However, some consistent result are observed if U(exp) = Delta H(0)redox, with a measured epsilon(Delta Delta)(H)(redox) = 19 +/- 6. The variability is thus attributed to an entropic factor (epsilon(Delta Delta)(S)(redox)) that is investigated using a series of substitutions of Asn(52) and/or Tyr(67). In double mutants Y67F/N52I Y67F/N52V, where most of the hydrogen bond network in the heme crevice is eliminated, Delta S(redox) compares to the wild type. This indicates that a fully consistent hydrogen bond network has a similar polarizability as an apolar matrix. We therefore argue that the variability in net dielectric susceptibility arises from conformational polarizability, a factor that is not a function of atomic properties and coordinates and is therefore hard to predict using conventional physical relationships.  相似文献   

17.
A series of covalently bound peptide-protoporphyrin-peptide compounds, also carrying naphthalene (N) to allow a photophysical investigation, were synthesized. Their general formula is P(nN)(2), where P refers to protoporphyrin IX, and n to the number of amino acids in the sequence Boc-Leu-Leu-Lys-(Ala)(x) -Leu-Leu-Lys-OtBu of each backbone chain (x = 0-3; n = x + 6). Their structural features in methanol solution were investigated by ir and CD spectra, and by steady-state and time resolved fluorescence experiments as well. The ir spectra indicate that intramolecularly H-bonded conformations form, and CD data in both methanol and water-methanol mixture suggest the presence of alpha-helix structure. Quenching of excited naphthalene takes place by electronic energy transfer from singlet N* to P ground state. Fluorescence decays coupled with molecular mechanics calculations indicate that two conformers for each dimeric peptide are the major contributors to the observed phenomena. These conformers are characterized by a globular, protein-like structure, where the protoporphyrin resides in a central pocket, while the two N groups are externally situated. Of the four N linkages in the two conformers, three of them attain a very similar steric arrangement around the central P molecule, in terms of both center-to-center distance and mutual orientation, while the fourth experiences a different steric disposition as compared to the others. Experimental photophysical parameters satisfactorily compare with those obtained by theoretical calculations, within the F?rster mechanism for long-range energy transfer, only when the mutual orientation of the chromophores was also taken into account. This implies that interconversion among conformational substates of probes linkages is slow on the time scale of the energy transfer process.  相似文献   

18.
We propose that variations in fat and carbohydrate (CHO) oxidation by working muscle alter O(2) uptake (VO(2)) kinetics. This hypothesis provides two predictions: 1) the kinetics should comprise two exponential components, one fast and the other slow, and 2) their contribution should change with variations in fat and CHO oxidation, as predicted by steady-state respiratory exchange ratio (RER). The purpose of this study was to test these predictions by evaluating the VO(2) kinetic model: VO(2)(t) = alpha(R) + alpha(F)(1 - exp[(t - TD)/-tau(F)]) + alpha(C)(1 - exp[(t - TD)/-tau(C)]) for short-term, mild leg cycling in 38 women and 44 men, where VO(2)(t) describes the time course, alpha(R) is resting VO(2), t is time after onset of exercise, TD is time delay, alpha(F) and tau(F) are asymptote and time constant, respectively, for the fast (fat) oxidative term, and alpha(C) and tau(C) are the corresponding parameters for the slow (CHO) oxidative term. We found that 1) this biexponential model accurately described the VO(2) kinetics over a wide range of RERs, 2) the contribution of the fast (alpha(F), fat) component was inversely related to RER, whereas the slow (alpha(C), CHO) component was positively related to RER, and 3) this assignment of the fast and slow terms accurately predicted steady-state respiratory quotient and CO(2) output. Therefore, the kinetic model can quantify the dynamics of fat and CHO oxidation over the first 5-10 min of mild exercise in young adult men and women.  相似文献   

19.
The F(1)F(o)-ATP synthases of alkaliphilic bacteria exhibit latent ATPase activity, and for the thermoalkaliphile Bacillus sp. strain TA2.A1, this activity is intrinsic to the F(1) moiety. To study the mechanism of ATPase inhibition, we developed a heterologous expression system in Escherichia coli to produce TA2F(1) complexes from this thermoalkaliphile. Like the native F(1)F(o)-ATP synthase, the recombinant TA2F(1) was blocked in ATP hydrolysis activity, and this activity was stimulated by the detergent lauryldimethylamine oxide. To determine if the C-terminal domain of the epsilon subunit acts as an inhibitor of ATPase activity and if an electrostatic interaction plays a role, a TA2F(1) mutant with either a truncated epsilon subunit [i.e., TA2F(1)(epsilon(DeltaC))] or substitution of basic residues in the second alpha-helix of epsilon with nonpolar alanines [i.e., TA2F(1)(epsilon(6A))] was constructed. Both mutants showed ATP hydrolysis activity at low and high concentrations of ATP. Treatment of the purified F(1)F(o)-ATP synthase and TA2F(1)(epsilon(WT)) complex with proteases revealed that the epsilon subunit was resistant to proteolytic digestion. In contrast, the epsilon subunit of TA2F(1)(epsilon(6A)) was completely degraded by trypsin, indicating that the C-terminal arm was in a conformation where it was no longer protected from proteolytic digestion. In addition, ATPase activity was not further activated by protease treatment when compared to the untreated control, supporting the observation that epsilon was responsible for inhibition of ATPase activity. To study the effect of the alanine substitutions in the epsilon subunit in the entire holoenzyme, we reconstituted recombinant TA2F(1) complexes with F(1)-stripped native membranes of strain TA2.A1. The reconstituted TA2F(o)F(1)(epsilon(WT)) was blocked in ATP hydrolysis and exhibited low levels of ATP-driven proton pumping consistent with the F(1)F(o)-ATP synthase in native membranes. Reconstituted TA2F(o)F(1)(epsilon(6A)) exhibited ATPase activity that correlated with increased ATP-driven proton pumping, confirming that the epsilon subunit also inhibits ATPase activity of TA2F(o)F(1).  相似文献   

20.
The ligand-binding dynamics and the reaction with O(2) of the fully (five-electron) reduced cytochrome caa(3) from the thermohalophilic bacterium Rhodothermus (R.) marinus were investigated. The enzyme is a proton pump which has all the residues of the proton-transfer pathways found in the mitochondrial-like enzymes conserved, except for one of the key elements of the D-pathway, the helix-VI glutamate [Glu(I-286), R. sphaeroides numbering]. In contrast to what has been suggested previously as general characteristics of thermophilic enzymes, during formation of the R. marinus caa(3)-CO complex, CO binds weakly to Cu(B), and is rapidly (k(Ba) = 450 s(-1)) trapped by irreversible (K(Ba) = 4.5 x 10(3)) binding to heme a(3). Upon reaction of the fully reduced enzyme with O(2), four kinetic phases were resolved during the first 10 ms after initiation of the reaction. On the basis of a comparison to reactions observed with the bovine enzyme, these phases were attributed to the following transitions between intermediates (pH 7.8, 1 mM O(2)): R --> A (tau congruent with 8 micros), A --> P(r) (tau congruent with 35 micros), P(r) --> F (tau congruent with 240 micros), F --> O (tau congruent with 2.5 ms), where the last two phases were associated with proton uptake from the bulk solution. Oxidation of heme c was observed only during the last two reaction steps. The slower transition times as compared to those observed with the bovine enzyme most likely reflect the replacement of Glu(I-286) of the helix-VI motif -XGHPEV- by a tyrosine in the R. marinus enzyme in the motif -YSHPXV-. The presence of an additional, fifth electron in the enzyme was reflected by two additional kinetic phases with time constants of approximately 20 and approximately 720 ms during which the fifth electron reequilibrated within the enzyme.  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