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 normal mode theory is developed for the Brownian dynamics of weakly bending rods with preset hydrodynamic interactions. The rod is replaced by a chain of contiguous spheres whose radius is chosen to yield the appropriate uniform translational and rotational diffusion coefficients. Despite the inclusion of preset hydrodynamic interactions in the dynamical operator, its normal modes are not coupled by the potential energy, so their amplitudes remain pairwise “orthogonal” under equilibrium averaging. The uniform translational and rotational diffusion coefficients obtained from Langevin theory are shown to be identical to those obtained from the Kirkwood algorithm, despite their rather different appearance. An expression is given for the mean squared angular displacement 〈Δxm(t)2〉 of the mth bond vector around the instantaneous x axis (perpendicular to the end-to-end vector z). Necessary algorithms are presented for the numerical evaluation of all quantities. The normal mode theory is compared with Brownian dynamics simulations for the same model by examining 3〈Δxm(t)2〉 for the central bond vector of rods comprising 10 and 30 subunits with various persistence lengths. The normal mode theory works very well for all times for L/P ? 0.6, where P = κ/kBT is the persistence length and κ is the bending rigidity. With increasing flexibility, the domain of validity of the normal mode theory is restricted to shorter times, where violations of the weak bending approximation are less severe. However, increasing the length of the rod from 10 to 30 subunits yields improved agreement with the simulations for the same and even longer times. This latter effect is tentatively attributed to the greater fluctuating tension in the longer chains, which acts to retard the rotational relaxation in the simulations, but is not taken into account in the present normal mode theory.  相似文献   

2.
An Exact theoretical expression for the apparent diffusion coefficient Dapp(K) of a thin rigid rod with arbitrary anisotropy of its translational diffusion diffusion coefficient is derived from the first cumulant of its dynamic structure factor. Dapp(K) is predicted to reach a limiting plateau value at extermely large values of KL, where K is the scattering vector and L the rod length. Howerver, that limiting plateau value is approached only very slowly along a quasi-plateau with a very gradual slope. Dynamic light-scattering studies have been performed on tobacco mosaic virus from K2 = (0.4–20) × 1010 cm?2 using 632-8-nm laser radiation. The present data yield D0 = (4.19 ± 0.10) × 10?8 cm2/s (corrected to 20,w conditions) and, with literature data to establish L = 2980 Å and the rotational diffusion coefficient DR = 318s?1, yield also Δ ≡ D ? D = (1.79 ± 0.38) × 10?8 cm2/s. The experimental data closely follow the curve of Dapp(K) vs K2 calcuated for these parameters. The present value of D0 substantially exceeds all previous dynamic light-scattering values, but is in good aggreement with previous sedimentation data, which were confirmed for the presemt sample. The anisotropy ratio Δ/D0 = 0.43 ± 0.09 is in accord with theoretical predictions based on the modified Kirkwood algorithm, despite the fact the D0 lies significantly below its corresponding theoretical value. The present data largely predlude the possibility that both D0 and Δ/D0 could simultaneously match their theoretical predictions. We present a detailed comparison of the experimental data with the calculations of Tirado and Garcia de la Torre based on the modified Kirkwood algorithm and with the Broersma formulas.  相似文献   

3.
Cyclic (1 --> 4)-alpha-D-glucan chains with or without excluded volume have been collected from a huge number (about 10(7)) of linear amylosic chains generated by the Monte Carlo method with a conformational energy map for maltose, and their mean-square radii of gyration and translational diffusion coefficients D (based on the Kirkwood formula) have been computed as functions of x (the number of glucose residues in a range from 7 to 300) and the excluded-volume strength represented by the effective hard-core radius. Both /x and D in the unperturbed state weakly oscillate for x < 30 and the helical nature of amylose appears more pronouncedly in cyclic chains than in linear chains. As x increases, these properties approach the values expected for Gaussian rings. Though excluded-volume effects on them are always larger in cycloamylose than in the corresponding linear amylose, the ratios of and the hydrodynamic radius of the former to the respective properties of the latter in good solvents can be slightly lower than or comparable to the (asymptotic) Gaussian-chain values when x is not sufficiently large. An interpolation expression is constructed for the relation between the gyration-radius expansion factors for linear and cyclic chains from the present Monte Carlo data and the early proposed asymptotic relation with the aid of the first-order perturbation theories.  相似文献   

4.
Sedimentation coefficient of polyoma virus DNA   总被引:5,自引:0,他引:5  
H B Gray 《Biopolymers》1967,5(10):1009-1019
The sedimentation coefficient of the twisted circular form of polyoma virus DNA is calculated from the Kirkwood sedimentation–diffusion equation, the structure being assumed to be a rigid double superhelix. Agreement with the experimental sedimentation coefficient can be obtained, with the use of an experimental value for the number of superhelical turns, when the pitch of the superhelix is intermediate between its minimal and maximal possible values. Another model, which has been proposed for polyoma DNA at low ionic strengths, may be visualized as a superhelical structure wound about a torus. Calculations of sedimentation coefficients for this model agree qualitatively with experimental data at ionic strengths Below 10?2M.  相似文献   

5.
Optical activity of aromatic chromophores. I. O, m, and p-tyrosine   总被引:1,自引:0,他引:1  
T M Hooker  J A Schellman 《Biopolymers》1970,9(11):1319-1348
The optical activity and the conformational energy of the amino acids tyrosine, o-tyrosine, and m-tyrosine have been calculated as a function of molecular conformation. A new graphic technique, rotatory strength–conformational energy maps, was developed for the presentation of these calculations. Experimental circular dichroism spectra were determined by utilization of a new experimental technique that involves complete computer control of a Cary spectropolarimeter. This permitted repetitive scanning with signal averaging over extended time periods and resulted in a greatly enhanced signal to noise ratio. On the basis of these calculations, it, was concluded that the Kirkwood “coupled oscillator” mechanism is capable of explaining the observed optical activity of the aromatic chromophores of the molecules that were investigated.  相似文献   

6.
The translational friction coefficients, rotational friction coefficient, and intrinsic viscosity of rigid regular structures composed of up to eight identical spherical subunits have been accurately calculated. The aim of this calculation is to interpret the hydrodynamic properties of oligomeric subunit proteins. To avoid the well-known failure of the theory in the evaluation of rotational coefficients and intrinsic viscosities, each subunit is hydrodynamically modeled as a polyhedral array of smaller spheres. The analysis of several alternatives suggests that a cubic array is the best choice. The reliability of this strategy is checked by comparison of the calculated values for all the transport properties of a sphere and the translational friction coefficients of a dimer with their exact values. Finally, the hydrodynamic properties of a number of subunit structures with varying number of subunits and different geometries are tabulated.  相似文献   

7.
Abstract

The molecular simulation technique of stochastic dynamics (SD) is tested by application to the immunosuppressive drug cyclosporin A (CPA). Two stochastic dynamics simulations are performed, one (SDCCl4 ) with atomic friction coefficients proportional to the viscosity of the nonpolar solvent CCl4, and one (SDH2O) with atomic friction coefficients corresponding to an aqueous solution. The atomic friction coefficients are also taken proportional to an approximate expression for the atomic accessible surface area. The properties of both stochastic dynamics simulations are compared to those of two full molecular dynamics (MD) simulations of cyclosporin A, one in a box with 591 CCl4 molecules, and one in a box with 632 H2O molecules.

The properties of cyclosporin A as found in the molecular dynamics simulation in CCl4 are well reproduced by the SDCCl4 simulation. This indicates that the neglect of a mean force reresenting the average solvent effects on the solute is justified in the case of nonpolar solvents. For polar solvents, like water, this mean force may not be neglected. The SDH2O simulation of cyclosporin A clearly fails to reproduce the amount of hydrogen bonding found in the molecular dynamics stimulation of cyclosporin A in water.

A comparison with a molecular dynamics simulation of cyclosporin A in vacuo shows that both the SDCCl4 and the SDH2O simulation come closer to the properties of the molecular dynamics simulations in CCl4 and in H2O than a molecular dynamics simulation in vacuo.  相似文献   

8.
A. Patkowski  S. Jen  B. Chu 《Biopolymers》1978,17(11):2643-2662
We have measured the translational (DT) and rotational (DR) diffusion coefficients of bulk tRNA from baker's yeast during the thermal unfolding process by means of photon-correlation spectroscopy. It should be noted that our estimate of the rotational diffusion coefficient represented, for the first time, measurements on a small macromolecule in solution by the photoelectron time-of-arrival technique with a delay-time resolution of 1 nsec. The melting curves expressed in terms of δDT vs temperature were consistent with the literature data in revealing the melting steps and their dependence on NaCl concentration. Additionally, it was possible to prove the existence of an intermediate, more compact structure during the initial steps of the thermal unfolding process. We found that the temperature ranges over which this intermediate structure appears depend strongly on salt concentration. By utilizing both translational and rotational diffusion coefficients and Perrin's equations for ellipsoids of revolution, we have computed the values of the equivalent length and width of tRNA molecules in solution at four different temperatures for NaCl concentrations of 0.2, 0.5, and 1M. The approximate model of ellipsoids of revolution also permits us to obtain an estimate of the radius of gyration, which is in very good agreement with literature data measured by means of small-angle x-ray scattering. Furthermore, we have measured the shape and size changes of tRNA with varying NaCl concentrations at room temperatures (25°C). The molecule becomes smaller and more spherical when NaCl concentration increases. As a result of partial melting at 70°C, the macromolecule is surprisingly elongated with an approximate axial ratio of 8:1 and has dimensions of about 180/22Å. Such information on conformational changes by a simultaneous determination of rotational and translational diffusion coefficients illustrates the potential of this approach, not available by other methods.  相似文献   

9.
Abstract

Ab initio quantum-chemical calculations with inclusion of electron correlation made since 1994 (such reliable calculations were not feasible before) significantly modified our view on interactions of nucleic acid bases. These calculations allowed to perform the first reliable comparison of the strength of stacked and hydrogen bonded pairs of nucleic acid bases, and to characterize the nature of the base-base interactions. Although hydrogen-bonded complexes of nucleobases are primarily stabilized by the electrostatic interaction, the dispersion attraction is also important. The stacked pairs are stabilized by dispersion attraction, however, the mutual orientation of stacked bases is determined rather by the electrostatic energy. Some popular theories of stacking were ruled out: The theory based on attractive interactions of polar exocyclic groups of bases with delocalized electrons of the aromatic rings (Bugg et al., Biopolymers 10, 175 (1971).), and the II-II interactions model (C.A. Hunter, J. Mol. Biol. 230, 1025 (1993)). The calculations demonstrated that amino groups of nucleobases are very flexible and intrinsically nonplanar, allowing hydrogen-bond-like interactions which are oriented out of the plane of the nucleobase. Many H-bonded DNA base pairs are intrinsically nonplanar. Higher-level ab initio calculations provide a unique set of reliable and consistent data for parametrization and verification of empirical potentials. In this article, we present a short survey of the recent calculations, and discuss their significance and limitations. This summary is written for readers which are not experts in computational quantum chemistry.  相似文献   

10.
11.
M T Record 《Biopolymers》1967,5(10):975-992
An approximate analytical expression for the electrostatic free energy of a polynucleo-tide in any of its possible ordered or random conformations is derived by integration of the screened-Coulomb potential energy function over all charge pairs in the structure. The electrostatic free energy of any form is found to be a linear function of the logarithm of the monovalent counterion concentration, in the range of low salt concentrations. Hence the electrostatic free energy difference between ordered and disordered forms in a polynucleotide structural transition is a linear function of the logarithm of the monovalent counterion concentration. A free energy balance applied to a two-state model for the transition then yields a linear dependence of the transition temperature Tm upon the logarithm of the counterion concentration. Calculation of the quantity dTm/d log M, where M is the monovalent counterion concentration, shows it to be a characteristic constant for a given transition, with a magnitude and sign proportional to the charge density difference between the ordered and disordered forms. Use of any one of several alternate, simple assumptions yields predicted dTm/d log M values in good agreement with experimental data for various polynucleotide transitions.  相似文献   

12.
We have extended our previous theories of the translational and rotational frictional properties of multisubunit complexes to calculate the intrinsic viscosity of such structures. Our theory is similar to those recently construced by McCammon and Deutch, and by Nakajima and Wada, in that it uses a modified hydrodynamic interaction tensor and solves the system of simultaneous interaction equations by digital computation rather than by successive approximations. However, there are some differences in the formulation and averaging of these equations. Extensive numerical comparison is made between this theory and others that are available—associated with the names of Hearst and Tagami, Abdel-Khalik and Bird, and Tsuda—using as a basis exact results for prolate ellipsoids of revolution. For large axial ratios, only our theory asymptotically approaches the correct limit; but for small axial ratios, only the Tsuda “shell-model” theory is adequate, because the other theories neglect the preponderant influence of the sphere located at the center of rotation. Intrinsic viscosities, translational frictional coefficients, and Scheraga-Mandelkern β parameters, are tabulated for a large number of polygonal and polyhedral subunit structures, with up to eight elements, using both our theory and Tsuda's. Particular application is made to hemerythrin and aspartate transcarbamylase. Finally, the viscosities and friction coefficients o once-broken rods are calculated and compared with an approximate theory by Wilenski.  相似文献   

13.
H X Zhou 《Biophysical journal》1995,69(6):2286-2297
A general method for calculating translational friction and intrinsic viscosity is developed through exploiting relations between hydrodynamics and electrostatics. An approximate relation xi = 6 pi eta 0C between the translational friction coefficient xi of a particle (eta 0: solvent viscosity) and its capacitance C was derived previously. This involved orientationally preaveraging the Oseen tensor, but the result was found to be very accurate. Based on preaveraging, we find that the intrinsic viscosity [eta] of a particle can be estimated from its polarizability alpha through [eta] = 3/4 alpha + 1/4 Vp, where Vp is the volume of the particle. Both the capacitance and the polarizability can be obtained in a single calculation using the boundary-element technique. An efficient approach is thus found for estimating [eta], a quantity that is very useful in practice because of its sensitivity to particle shape but is notoriously difficult to calculate. Illustrative calculations on ellipsoids, cylinders, and dumbbells demonstrate both the accuracy of the approximate relations and the efficiency of the present method.  相似文献   

14.
15.
The translational and rotational diffusion coefficients have been calculated for a simple, segmentally flexible model: the hinged dumbbell (HD). In the HD, two spherical subunits are attached to an universal joint by means of frictionless connectors. In addition to the case in which hydrodynamic interactions are neglected (NI), we have also considered two more cases, including hydrodynamic interaction by means of the Kirkwood-Riseman approximate treatment (KR) and using accurate procedure based in the series expansions for the two-sphere diffusion tensor (SE). Expressions for the friction coefficients of the HD are given for the three cases, and the diffusion coefficients are evaluted inverting the 9 × 9 resistance matrix, for two HDs with different dimensions. The KR treatment, which includes a contribution from the finite volume of the subunits, is shown to be an excellent approximation to the more rigorous procedure. In the NI case for rotation, the various coefficients present different deviations with respect to the SE results. A rough estimate of the errors of the NI relaxation times indicates that they may be smaller than 15% for a HD with identical beads. However, the influence of hydrodynamic interaction should be more important for the rotational diffusivity of a small sphere attached to a larger one. The error of the NI result for the translational diffusion coefficient is of about 25% for the two HDs.  相似文献   

16.
A drop-interaction model devised by the authors, which successfully predicts experimentally observed drop-size distributions in stirred batch vessels, has been used to assess the importance of droplet coalescence and redispersion in hydrocarbon fermentation yields. Calculations, which have been made with both n-alkane and gas–oil fermentations, have been compared with previous calculations based on simpler and, hence, less realistic droplet interaction mechanisms. The comparison shows that the differences in yield are not spectacularly different.  相似文献   

17.
Seven cyclic (1 --> 4)-alpha-D-glucan (cycloamylose) samples ranging in weight-average molecular weight from 5 x 10(3) to 1.8 x 10(4) and gamma-cyclodextrin have been studied by sedimentation equilibrium in dimethylsulfoxide (at 25 degrees C) and by dynamic light scattering in 0.5 N aqueous sodium hydroxide (at 25 degrees C), a good solvent for linear amylose. The measured translational diffusion coefficients D in the aqueous NaOH agree fairly closely with previous Monte Carlo results for cyclic (1 --> 4)-alpha-D-glucan chains with excluded volume, when correction is made for the effects of bead diameter and fluctuating hydrodynamic interaction (HI) on the Kirkwood theory on which the computation of D was based. These D data are also explained almost quantitatively by Yamakawa and Fujii's expression for the associated KP ring (based on the Kratky-Porod wormlike chain) with the molecular parameters for linear amylose if the fluctuating HI and excluded-volume effects are taken into account. It is concluded that the translational diffusion behavior of cycloamylose in the aqueous NaOH is consistent with the conformational characteristics derived from the conformational energy of maltose and dilute-solution data for linear amylose.  相似文献   

18.
Ice friction during speed skating.   总被引:2,自引:0,他引:2  
During speed skating, the external power output delivered by the athlete is predominantly used to overcome the air and ice frictional forces. Special skates were developed and used to measure the ice frictional forces during actual speed skating. The mean coefficients of friction for the straights and curves were, respectively, 0.0046 and 0.0059. The minimum value of the coefficient of ice friction was measured at an ice surface temperature of about -7 degrees C. It was found that the coefficient of friction increases with increasing speed. In the literature, it is suggested that the relatively low friction in skating results from a thin film of liquid water on the ice surface. Theories about the presence of water between the rubbing surfaces are focused on the formation of water by pressure-melting, melting due to frictional heating and on the 'liquid-like' properties of the ice surface. From our measurements and calculations, it is concluded that the liquid-like surface properties of ice seem to be a reasonable explanation for the low friction during speed skating.  相似文献   

19.
The method of Ermak and McCammon [Ermak, D. L. & McCammon, J. A. (1978) J. Chem. Phys. 69 , 1352–1360] is used to simulate the Brownian dynamics of a system of identical, interacting beads. In the present study, we use the method to obtain transport coefficients for a variety of rigid and flexible structures modeled as arrays of spherical subunits. Constraints are enforced using the SHAKE algorithm or a modification, SHAKE-HI, that is described for the first time. In SHAKE-HI, hydrodynamic interactions between subunits are taken into account when the constraints are enforced. Use of SHAKE-HI yields transport coefficients that are in perfect agreement with those obtained by other methods. The primary advantage of the present method is its generality. We also propose that multistep Brownian dynamics may be important in simulating actual experiments (such as fluorescence depolarization) on well-defined model systems that possess an arbitrary degree of internal flexibility.  相似文献   

20.
The modelling of thermodynamic properties of liquids from local density fluctuations is relevant to many chemical and biological processes. The Kirkwood–Buff (KB) theory connects the microscopic structure of isotropic liquids with macroscopic properties such as partial derivatives of activity coefficients, partial molar volumes and compressibilities. Originally, KB integrals were formulated for open and infinite systems which are difficult to access with standard Molecular Dynamics (MD) simulations. Recently, KB integrals for finite and open systems were formulated (J Phys Chem Lett. 2013;4:235). From the scaling of KB integrals for finite subvolumes, embedded in larger reservoirs, with the inverse of the size of these subvolumes, estimates for KB integrals in the thermodynamic limit are obtained. Two system size effects are observed in MD simulations: (1) effects due to the size of the simulation box and the size of the finite subvolume embedded in the simulation box, and (2) effects due to computing radial distribution functions (RDF) from a closed and finite system. In this study, we investigate the two effects in detail by computing KB integrals using the following methods: (1) Monte Carlo simulations of finite subvolumes of a liquid with an analytic RDF and (2) MD simulations of a WCA mixture for various simulation box sizes, but at the same thermodynamic state. We investigate the effect of the size of the simulation box and quantify the differences compared to KB integrals computed in the thermodynamic limit. We demonstrate that calculations of KB integrals should not be extended beyond half the size of the simulation box. For finite-size effects related to the RDF, we find that the Van der Vegt correction (J Chem Theory Comput. 2013;9:1347) yields the most accurate results.  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