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 31 毫秒
1.
K S Schmitz  R Pecora 《Biopolymers》1975,14(3):521-542
The quasi-elastic light-scattering (homodyne) time-correlation functions of calf thymus and λDNA are shown to contain contributions from at least two relaxation processes. A method of asymptotic analysis is described and used to obtain an estimate of the longest relaxation time as well as the “average” relaxation time and the mean-squared dispersion in this average. Most theories of scattering from macromolecules in the limit of inifinite dilution predict that the longest relaxation time is due to translational self-diffusion. The data obtained, however, indicate that the longest time is not simply related to the translational self-diffusion coefficient of unaggregated macromolecules. It is also shown that the longest relaxation time of λDNA decreases in the later stages of the denaturation transition region. Some possible mechanisms for the origin of this long time are discussed, including a model of restricted motion of a molecule by its neighbors.  相似文献   

2.
Viscoelastic behavior of erythrocyte membrane.   总被引:1,自引:0,他引:1       下载免费PDF全文
A nonlinear viscoelastic relation is developed to describe the viscoelastic properties of erythrocyte membrane. This constitutive equation is used in the analysis of the time-dependent aspiration of an erythrocyte membrane into a micropipette. Equations governing this motion are reduced to a nonlinear integral equation of the Volterra type. A numerical procedure based on a finite difference scheme is used to solve the integral equation and to match the experimental data. The data, aspiration length vs. time, is used to determine the relaxation function at each time step. The inverse problem of obtaining the time dependence of the aspiration length from a given relaxation function is also solved. Analytical results obtained are applied to the experimental data of Chien et al. 1978. Biophys. J. 24:463-487. A relaxation function similar to that of a four-parameter solid with a shear-thinning viscous term is proposed.  相似文献   

3.
The early stage of stress relaxation, up to 10 s after strain application, in compact bone was investigated in order to find the limit of the applicability of the empirical equation (Sasaki et al., 1993. Journal of Biomechanics 26, 1369-1376), E(t) = E0{A1exp[-(t/tau1)beta]+A2exp(-t/tau)}, A1+A2 = 1, 0相似文献   

4.
The spin-lattice relaxation time (T1) of the water protons of intrapelvic tissues was determined in vivo in women using an NMR imager. T1's of squamous cell carcinoma of the cervix are significantly longer than those of the normal cervix. A similar trend occurs in carcinomas of the endometrium and ovary. The results suggest that direct determination of T1 in vivo may be useful in the diagnosis and management of gynecologic malignancies.  相似文献   

5.
Precise measurement of the mechanical properties of a cell provides useful information about its structural organization and physiological state. It is interesting to understand the effect of individual components on the mechanical properties of the entire cell. In this study, we investigate the influence of the cytoskeletal actin on the viscoelastic properties of a cell. Actin-specific agents, including latrunculin A and jasplakinolide, are used to alter the organization of the cytoskeletal actin. Brassica oleracea protoplasts are treated with the drugs and deformed under an external electric potential. The relaxation processes of single protoplasts after electrodeformation are measured. The data are analyzed by a model-independent spectrum recovery algorithm. Two distinct characteristic time constants are obtained from the relaxation spectra. Treatment with latrunculin A increases both of the relaxation time constants. The longest relaxation times for control, latrunculin A treated, and jasplakinolide treated cells are determined to be 0.28, 1.0, and 0.21 s, respectively.  相似文献   

6.
Fluorescence polarization was used to determine the rotational relaxation time of fluorescein conjugated concanavalin A bound to the surface membrane of normal and malignant cells. The cells used were normal lymphocytes and malignant lymphoma cells, as examples of cells that are in suspension in vivo, and normal and simian virus 40-transformed fibroblasts, as examples of cells that form a solid tissue. The relaxation time of F-concanavalin A2 in phosphate-buffered saline, pH 7.2, at 24 °C was 58 nseconds. Under the same conditions, the relaxation times of F-concanavalin A bound to cells were: 70 nseconds for normal lymphocytes, 160 nseconds for malignant lymphoma cells, 120 nseconds for normal fibroblasts and 73 nseconds for simian virus 40-transformed fibroblasts. When cells were treated with trypsin or glutaraldehyde before adding F-concanavalin A, trypsin treatment produced a decrease, whereas glutaraldehyde fixation produced an increase of these values. Inhibition of cap formation of concanavalin A binding sites on normal lymphocytes by treating the cells with sodium azide did not change the rotational relaxation time of concanavalin A bound to the cells. These results indicate that the carbohydrate-containing structures on the cells that bind concanavalin A are mobile. In cells that are in suspension in vivo, malignant transformation is associated with reduction in mobility of these sites. However, in cells that form a solid tissue, malignant transformation is associated with an increase in mobility of these sites. Determination of the rotational relaxation time of fluorescent probes bound to specific sites on cell membranes can thus be used to quantitate receptor mobility in relation to cell behaviour.  相似文献   

7.
Fluorescence anisotropy decay measurements were performed on melittin in water and in membranes of dimyristoylphosphatidylcholine. The fluorescence of the single tryptophan residue of melittin and of a pyrene label attached to melittin was detected. In water, the slowest relaxation process in the anisotropy decay occurs with a relaxation time of 1.5 or 5.5 ns in the case of low or high ionic strength and corresponds to rotational diffusion of monomeric or tetrameric melittin. Superimposed on this slow process are fast processes in the subnanosecond range reflecting fluctuations of the fluorophores relative to the polypeptide backbone. In membranes, the fast relaxation processes are not much altered. A slow process with a relaxation time of 35 ns is observed and assigned to orientational fluctuations of the melittin helices in membranes.  相似文献   

8.
Generally accepted manometric criteria for the diagnosis of achalasia are absent peristalsis and incomplete lower esophageal sphincter (LES) relaxation. However, in some patients with otherwise typical features of achalasia, esophageal manometry shows complete LES relaxation during swallowing. To establish whether such apparently complete LES relaxations are functionally adequate, we quantified changes in resistance to flow at the esophagogastric junction (EGJ) during wet swallowing. We studied seven achalasia patients with manometrically complete (>80%) LES relaxation, eight achalasia patients with incomplete (<40%) LES relaxation, and eight healthy volunteers. Complete LES relaxation on standard manometry (open-tip catheters) was confirmed in five of the seven achalasia patients by a Dentsleeve. Changes in EGJ resistance to flow were quantified using a pneumatic resistometer. Manometrically, the relaxation time span was significantly longer in patients with complete LES relaxation than in those with incomplete relaxation (7. 3 +/- 0.5 vs. 4.4 +/- 0.7 s; P < 0.05). The fall in EGJ resistance from basal values during swallowing was markedly reduced in both achalasia groups (21 +/- 8% in those with manometrically complete relaxation and 4 +/- 2% in those with incomplete relaxation) by comparison with healthy individuals, in whom resistance fell by 90 +/- 3% (P < 0.05 vs. both achalasia groups). The duration of EGJ resistance drop was also much shorter in achalasia with (0.7 +/- 0.2 s) and without (0.2 +/- 0.1 s) complete LES relaxation compared with healthy control values (6.6 +/- 1.2 s). Our results reveal that the apparently complete LES relaxation observed manometrically in some patients with achalasia is functionally inadequate since it is not associated with the normal profound fall in EGJ resistance to flow.  相似文献   

9.
The spin-lattice relaxation time, T1, and the dipolar energy relaxation time, TD, were measured as a function of temperature. The materials studied were samples of anhydrous L-dipalmitoyl lecithin, DL-dipalmitoyl lecithin, L-dimyristoyl lecithin, DL-dimyristoyl lecithin and their monohydrates, and of anhydrous egg yolk lecithin. It is shown that TD is a much more sensitive parameter than T1 for the determination of the Chapman phase transition. Comparison between T1 and TD provides information about new types of slow molecular motions below and above the phase transition temperature. It is suggested that the relaxation mechanisms for T1 and TD in the gel phase are governed by segmental motion in the phospholipid molecule. A new metastable phase was detected in dimyristoyl lecithin monohydrates. This phase could only be detected from the dipolar energy relaxation times.  相似文献   

10.
A proton nuclear magnetic resonance study of the four histidine residues of thermally unfolded ribonuclease A has provided evidence that two of the residues are in regions of residual structure, whereas the other two are freely exposed to solvent. Histidine-48 and, tentatively, histidine-105 occupy an environment at 69 degrees characterized by residual structure and display a pK value of 5.75 and a spin-lattice relaxation time of about 0.8 sec at pH 5.5. Histidine-12 and, tentatively, histidine-119 are in an environment at 69 degrees which is freely accessible to solvent and show a pK value of 5.96 and a spin-lattice relaxation time of about 1.1 sec at pH 5.5.  相似文献   

11.
The effect of sequence heterogeneity on DNA melting kinetics   总被引:1,自引:0,他引:1  
We consider kinetics of the cooperative melting of DNA sections situated at the edge of the helix. Accurate calculations based on the real sequences of such sections demonstrate that their internal heterogeneity has a drastic effect on the melting kinetics. Allowance for the internal heterogeneity increases the relaxation time by several orders of magnitude as compared with a model based on the assumption of equal base-pair stability within a section. The relaxation times obtained are in good agreement with the experimental data of Suyama and Wada (A. Suyama and A. Wada, Biopolymers, 23, 409 (1984)). An analysis of the melting process revealed some simple sequence characteristics that determine its rate. An examination of the temperature dependence of the relaxation time led to a distinct interpretation of the apparent activation energies of the denaturation and renaturation. The relaxation time proved to reach its maximum near the equilibrium melting point of the section examined.  相似文献   

12.
13.
The temperature and cell volume dependence of the NMR water proton line-width, spin-lattice, and spin-spin relaxation times have been studied for normal and sickle erythrocytes as well as hemoglobin A and hemoglobin S solutions. Upon deoxygenation, the spin-spin relaxation time (T2) decreases by a factor of 2 for sickle cells and hemoglobin S solutions but remains relatively constant for normal cells and hemoglobin A solutions. The spin-lattice relaxation time (T1) shows no significant change upon deoxygenation for normal or sickle packed red cells. Studies of the change in the NMR linewidth, T1 and T2 as the cell hydration is changed indicate that these parameters are affected only slightly by a 10-20% cell dehydration. This result suggests that the reported 10% cell dehydration observed with sickling is not important in the altered NMR properties. Low temperature studies of the linewidth and T1 for oxy and deoxy hemoglobin A and hemoglobin S solutions suggest that the "bound" water possesses similar properties for all four species. The low temperature linewidth ranges from about 250 Hz at -15 degrees C to 500 Hz at -36 degrees C and analysis of the NMR curves yield hydration values near 0.4 g water/g hemoglobin for all four species. The low temperature T1 data go through a minimum at -35 degrees C for measurements at 44.4 MHz and -50 degrees C for measurements at 17.1 MHz and are similar for oxy and deoxy hemoglobin A and hemoglobin S. These similarities in the low temperature NMR data for oxy and deoxy hemoglobin A and hemoglobin S suggest a hydrophobically driven sickling mechanism. The room temperature and low temperature relaxation time data for normal and sickle cells are interpreted in terms of a three-state model for intracellular water. In the context of this model the relaxation time data imply that type III, or irrotationally bound water, is altered during the sickling process.  相似文献   

14.
The complex permittivities of L-alpha-lysolecithin in the absence and presence of the gramicidin A ion channel were measured over the temperature range 0-60 degrees C and over the frequency range 1-1000 MHz. One dielectric relaxation/loss has been observed. It is located at 103.3 MHz (1.54 ns) for a micellar 0.4 M L-alpha-lysolecithin solution at 20 degrees C, whereas it is shifted to 71.7 MHz (2.22 ns) for a lamellar L-alpha-lysolecithin-gramicidin A aqueous solution (0.4 M L-alpha-lysolecithin, 0.0308 M gramicidin A) at 20 degrees C. The dielectric relaxation decreases and the relaxation time increases when gramicidin A is incorporated into L-alpha-lysolecithin. These dielectric changes are related, in part, to the micellar-to-lamellar lipid phase transition induced by the incorporation of gramicidin A into lysolecithin. We suggest that the diffuse rotational motion of the polar head group of L-alpha-lysolecithin contributes to the dielectric relaxation/loss at around 100 MHz.  相似文献   

15.
C H Luan  R D Harris  D W Urry 《Biopolymers》1988,27(11):1787-1793
Dielectric relaxation studies of bovine ligamentum nuchae are reported over the frequency range of 1 MHz to 1 GHz and over the temperature range of 23–48°C. A temperature-dependent relaxation process was observed at low megahertz-frequency with the correlation time of around 40 ns. The result is quite similar to that of a synthetic polypentapeptide (VPGVG) and of α-elastin. The relaxation is proposed to arise in part from the peptide libration within the polypentapeptide of bovine ligamentum nuchae.  相似文献   

16.
Abstract

The thermodynamic integration (TI) method for calculating free energy differences has three inherent problems: statistics, numerical integration, and relaxation. In this paper the latter is analyzed for the nonequilibrium TI method introduced by Postma which combines molecular dynamics simulation and TI in a very effective way. A nontrivial extrapolation technique is presented to remove the relaxation error and to calculate the underlying relaxation time. It is shown that the optimal choice of grid points, derived in a previous paper, for minimizing statistical errors not only removes integration errors, but also minimizes relaxation errors. The methods are applied in a calculation of the free energy of cavity formation in water.  相似文献   

17.
NMR water-proton spin-lattice relaxation times were studied as probes of water structure in human red blood cells and red blood cell suspensions. Normal saline had a relaxation time of about 3000 ms while packed red blood cells had a relaxation time of about 500 ms. The relaxation time of a red cell suspension at 50% hematocrit was about 750 ms showing that surface charges and polar groups of the red cell membrane effectively structure extracellular water. Incubation of red cells in hypotonic saline increases relaxation time whereas hypertonic saline decreases relaxation time. Relaxation times varied independently of mean corpuscular volume and mean corpuscular hemoglobin concentration in a sample population. Studies with lysates and resealed membrane ghosts show that hemoglobin is very effective in lowering water-proton relaxation time whereas resealed membrane ghosts in the absence of hemoglobin are less effective than intact red cells.  相似文献   

18.
An acid-destabilized form of apomyoglobin, the so-called E state, consists of a set of heterogeneous structures that are all characterized by a stable hydrophobic core composed of 30-40 residues at the intersection of the A, G, and H helices of the protein, with little other secondary structure and no other tertiary structure. Relaxation kinetics studies were carried out to characterize the dynamics of core melting and formation in this protein. The unfolding and/or refolding response is induced by a laser-induced temperature jump between the folded and unfolded forms of E, and structural changes are monitored using the infrared amide I' absorbance at 1648-1651 cm(-1) that reports on the formation of solvent-protected, native-like helix in the core and by fluorescence emission changes from apomyoglobin's Trp14, a measure of burial of the indole group of this residue. The fluorescence kinetics data are monoexponential with a relaxation time of 14 micros. However, infrared kinetics data are best fit to a biexponential function with relaxation times of 14 and 59 micros. These relaxation times are very fast, close to the limits placed on folding reactions by diffusion. The 14 micros relaxation time is weakly temperature dependent and thus represents a pathway that is energetically downhill. The appearance of this relaxation time in both the fluorescence and infrared measurements indicates that this folding event proceeds by a concomitant formation of compact secondary and tertiary structures. The 59 micros relaxation time is much more strongly temperature dependent and has no fluorescence counterpart, indicating an activated process with a large energy barrier wherein nonspecific hydrophobic interactions between helix A and the G and H helices cause some helix burial but Trp14 remains solvent exposed. These results are best fit by a multiple-pathway kinetic model when U collapses to form the various folded core structures of E. Thus, the results suggest very robust dynamics for core formation involving multiple folding pathways and provide significant insight into the primary processes of protein folding.  相似文献   

19.
The aim of this work is to examine the correlation between measured instability of model flavor compounds in glassy matrices with the calorimetric relaxation times of the matrices. Spray-dried carbohydrate matrices were chosen as the model compounds for this study. Enthalpy relaxation times were determined for spray-dried carbohydrate matrices using differential and isothermal calorimetric methods. The losses of the volatile methyl acetate, ethyl acetate and limonene, as well as formation of limonene oxidation products, were measured by gas chromatography. Storage conditions were 30 and 40 °C, with samples equilibrated with 11, 23, 33 and 43 % RH at each temperature. A comparison of the relaxation times for temperatures below Tg was made using Modulated DSC (MDSC) and a Thermal Activity Monitor (TAM). TAM yields significantly lower values for relaxation times implying that it is capturing some of the faster dynamics as well as dynamics that are activated near Tg. However, plots of relaxation times as determined by both techniques versus temperature appear to converge at Tg. An increase in the relative humidity results in moderately higher loss of volatiles (methyl acetate, ethyl acetate and limonene) and greater oxidation rates. In general, there is a good correlation between relaxation time and stability, with greater enthalpy relaxation time associated with better stability. Enthalpy relaxation time appears to be a useful predictor of stability for both loss of volatiles and oxidation of limonene.  相似文献   

20.
We studied the spin-echo signal of muscle water in a large time domain and found that the motion of the nuclear magnetic moment of tissue water cannot be characterized by a single spin-lattice relaxation time (T1). The relaxation time T1B, which is the T1 characterized by those protons with a slower relaxation rate, is influenced by the early post mortem changes in skeletal muscle. T1B increased with time after the tissue was taken from the animal and reached a maximum at 3 h. However, the weighted average of T1 of all water protons (T1A) did not change throughout the time course of the experiments.  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