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.
The polymerization of tobacco mosaic virus protein has been found to proceed through metastable states under conditions where initially one of the two polymerization-linked protons is bound. These metastable polymers have been characterized and are found to be helical rods, which resemble the structure of equilibrium helical rods that form when both polymerization-linked protons are bound. At pH 6.5 and 20 °C the true equilibrium distribution of these helical rods has been shown to consist of sedimenting species that are much smaller, 24 to 34 S, than described previously, 100 to 200 S. The larger, non-equilibrium rods are produced by an overshoot in polymerization that results from the slow formation of 20 S nuclei followed by a very rapid elongation reaction. Generally, this sequence of rate processes is sensitive to the rate at which a reaction is initiated. In the present case it is the rate of heating or the rate of change of the pH that determines the reaction path and therefore the rate of attainment of equilibrium. In addition to the formation of metastable helical rods during polymerization overshoot, metastable 20 S aggregates can form when either equilibrium or non-equilibrium helical rods are depolymerized by cooling to 5 to 7 °C at pH 6.5. These 20 S aggregates are presumably two-turn disks or helices and can serve as nuclei for helical rod formation in subsequent polymerization reactions. Both helical rod and 20 S metastability are extremely sensitive to pH but, under carefully controlled conditions, the metastability is quite reproducible and reproducible nucleation-controlled polymerization kinetics can be observed even when polymerization-depolymerization cycling is carried out between branches of a hysteresis loop. Temperature- or pH-induced polymerization of tobacco mosaic virus protein can be made to proceed by the slow formation of 20 S, two-turn helix, nuclei followed by the rapid addition of one or more species comprising the 4 S protein. These results confirm a previously proposed kinetic mechanism for the non-equilibrium polymerization reaction (Scheele &; Schuster, 1974).  相似文献   

2.
The polymerization of alfalfa mosaic virus (AMV) protein in the presence of homologous nucleic acids and a number of other natural and synthetic nucleic acids was studied. The conditions for optimal assembly were found to be pH 6.0 and low ionic strength (I = 0.1 M) at room temperature, irrespective of the type of nucleic acid. The resulting nucleoprotein particles exhibited the same structural characteristics as the virus. This information emerged from optical diffraction and computer filtering of electron micrographs from the reconstituted particles. Irrespective of the type of nucleic acid present the polymerization of the protein resulting in a nucleoprotein particle is a cooperative process. Evidence for this was obtained by nitrocellulose filter binding assay, sodium dodecylsulphate/polyacrylamide gel electrophoresis, sedimentation velocity and electron microscopy of the reaction mixtures. The rates and efficiencies of reconstitution were of the same order of magnitude for a number of ribonucleic acids. Sedimentation data derived from AMV protein and AMV RNA mixtures suggested the existence of a specific nucleation product in the first stage of assembly. The results are discussed in terms of a tentative model of the assembly, in which at least two different steps (nucleation and elongation) can be distinguished, each characterized by an association constant.  相似文献   

3.
C J Westover  C L Stevens 《Biochemistry》1977,16(26):5819-5824
The lowest stages of polymerization of tobacco mosaic virus protein were studied by means of high-speed sedimentation equilibrium experiments. Several distinct modes of polymerization were found. At pH 7.1 the expected monomer-trimer-higher polymer equilibrium was observed--very little dimer was detected at this pH. At pH 7.5, however, a strong dimerization was observed--neither monomer nor trimer was detected at this pH. An octamer appeared to be the only species present other than the dimer. When 0.01 M beta-mercaptoethanol was added to the solvent pH 7.5, the dimer was dissociated, resulting in a monomer-trimer association. The dimerization may be the basis for the larger "doubled" polymers formed by the protein at alkaline pH, while the octamer may correspond to the 8S peak frequently observed in sedimentation velocity experiments at alkaline pH. On the other hand, the monomer-trimer-higher polymer equilibrium may correspond to the single helix formed by the protein at slightly acid pH and to the combination of 4S and 20S peaks seen in sedimentation velocity experiments at slightly acid pH.  相似文献   

4.
A method that allows the quantitative determination of reaction volumes from sedimentation velocity experiments in an analytical ultracentrifuge is presented. Combined with a second method for detecting pressure-induced depolymerization, general characteristics of polymer distributions may be probed. We show that it is possible to determine if a sample is in an equilibrium or metastable state of subunit association. Our approach to probe macromolecular aggregation systems by small pressure perturbations is not restricted to the use of centrifuges. This method has been applied to characterize certain aspects of the polymerization of tobacco mosaic virus coat protein (TMVP). There are at least two helical polymer conformations in RNA-free coat protein rods. The smaller, helix I, polymers are limited to sizes below about 70 subunits (four to five helical turns) and undergo some kind of cooperative conformational change before further subunits may be added indefinitely. In contrast to helix I, the larger helix II polymers occur as broader and skewed size distributions. Under moderately strong polymerization conditions, the equilibrium state can contain both types of helical rods. The reaction volume for the addition of trimers is -220 ml/mol for both types of helical polymers. These results are compared with the results of previous thermodynamic analyses of TMVP polymerization.  相似文献   

5.
A novel method has been developed for the detection and study of tRNA-like moieties in viral RNAs. Tobacco mosaic virus RNA is an acceptable substrate for crude Escherichia coli ribothymidine-forming tRNA methyltransferase. Under optimum reaction conditions at least 85% of the methylation product is ribothymidine (rT). The reaction is essentially quantitative, 1 mol of rT being formed per mol of tobacco mosaic virus RNA. The optimum reaction conditions include the presence of 6.6 micrometers S-adenosyl-L-[Me-3H]methionine, 25 micrometers spermine, 25 mM ammonium acetate, and 50 mM HEPES, pH 8.0. Sequence analysis of (Me-3H)-labeled tobacco mosaic virus RNA shows that all of the methylation occurs at a single site and strongly suggests that this site is the 32nd residue from the 3'-end of tobacco mosaic virus RNA. This site closely resembles the normal position of rT in transfer RNA.  相似文献   

6.
The association behaviour of the coat protein of alfalfa mosaic virus strain VRU was studied by sedimentation analysis and electron microscopy. The results of this study were compared with the data obtained from similar studies with the coat protein of strain 425 (Driedonks et al., 1977). In the depolymerized state VRU protein is likely a dinier of the 24,050 molecular weight polypeptide chain. The main association product is a tubular structure with a diameter of about 180 Å. The optimum conditions for the reaction were polyphosphate-containing buffer at pH 6·5. Optical diffraction analysis of negatively stained specimens revealed a helical arrangement of the protein subunits in these assemblies. The same type of reaction product was found when the association reaction was carried out in the presence of polynucleotides. The length of the VRU particles is abnormally long compared to other alfalfa mosaic virus strains. This phenomenon can be ascribed to the tendency of the protein to polymerize into tubular rather than spherical particles.  相似文献   

7.
Hydrogen ion titration curves of tobacco mosaic virus protein have been measured in various conditions of protein concentration, temperature, ionic strength, and rate of pH change. The polymers present at each stage are deduced from turbidity and sedimentation data, plus published information. A simple semi-quantitative analysis of the curves is given, and the pK values of the two abnormal carboxylates in single helix are estimated as 6.4 and about 7.0. Disks, and some faster-forming unknown polymers in the same size range, have been abnormal carboxylate with pK 6.9. These results are most easily interpreted in terms of electrostatic interactions between carboxylates, probably at the axial ends of the protein subunits.  相似文献   

8.
Holmes ribgrass virus (HRV), because of serological results, is regarded as a distantly related strain of tobacco mosaic virus (TMV). HRV protein differs substantially in amino acid sequence from TMV protein, especially in that it contains one histidine residue and three methionine residues, compared to none of either for TMV protein. Ultracentrifugation and hydrogen ion titration data on HRV protein, similar to those obtained previously for the early stage polymerization of TMV and E66 proteins, demonstrated some similarities and more distinct differences from those of the other two proteins. The major similarities are that the early polymerization of HRV protein is entropy driven and the first major polymerized product is a 20 S component, presumably a double disk or two-turn helix, as in the case of the other proteins. The major differences are that the unpolymerized HRV protein sediments at 3 S rather than at the 4 S for the others; it is presumably a dimer of the polypeptide chain. The enthalpy of polymerization per mole of A protein, delta H*, is 18,400 cal for HRV protein, compared to about 30,000 for TMV protein. One mol of H+ ion/mol HRV A protein, compared to 1.5 for TMV and E66 proteins, is bound during polymerization to the 20 S state. Contrasted with the other proteins, very little if any electrical work contribution was detected for the HRV protein. A major difference was found in hydrogen ion titration. Unpolymerized HRV protein binds hydrogen ions significantly in the unpolymerized A protein state, unlike the A proteins from the other two viruses.  相似文献   

9.
10.
Reconstitution of tobacco mosaic virus from its constituents, the coat protein and RNA, was investigated by means of ultracentrifugation and circular dichroism measurement. Tobacco mosaic virus protein forms a 20S double-layer disc under conditions favorable for tobacco mosaic virus reconstitution. Dibromination of the tyrosine 139 residue of tobacco mosaic virus protein prevents formation of the 20S disc.Acidification of the tobacco mosaic virus protein solution causes 20S discs to polymerize into long helical rods. Changes in the CD spectra of tobacco mosaic virus protein in the near-ultraviolet region suggest that stacking of the aromatic sidechains of amino acid residues stabilizes the helical rod. The dibrominated tobacco mosaic virus protein also has the ability of rod elongation under acidic condition. CD studies reveal that assembly of tobacco mosaic virus particles from its constituents is stabilized by the stacking effect between the base residues of RNA and the aromatic residues of tobacco mosaic virus protein.Cucumber green mottle mosaic virus protein, which acts as a substituent for tobacco mosaic virus protein in tobacco mosaic virus reconstitution, was also investigated.  相似文献   

11.
Protein of the tobacco mosaic virus mutant E66 has lysine replacing asparagine of the type strain, vulgare, at position 140. Thus, E66 protein should have one more positive or one less net negative charge than vulgare at pH 6 to 7. To investigate the effect of charge, a comparative study of the polymerization of E66 and vulgare proteins at pH 6.0, 6.2, 6.4, 6.6, and 6.8 at ionic strengths 0.15, 0.10, and 0.05 was made by turbidimetry. Polymerization of E66 protein always proceeded at a lower temperature than vulgare. However, the extent of polymerization was much lower in E66, especially at the higher ionic strengths. Sedimentation velocity results paralleled those from turbidity measurements in that E66 protein polymerizes at lower temperatures than vulgare; the 20 S component is more abundant in E66 protein. Osmotic pressure measurements also show that E66 protein is more polymerized than vulgare, especially at lower pH values. Hydrogen ion titrations of E66 protein were carried out from pH 8 to 5 and back to pH 8 in 0.10 m KCl at three temperatures, 4, 10, and 15 °C. These titrations were reversible when carried out slowly. The isoionic point is near pH 5; thus the charge at pH 7.5 is ?3. The reversible titration results were correlated with the aggregates present at the various pH values and temperatures, determined from the areas under the schlieren peaks in sedimentation velocity experiments. It is found that hydrogen ion binding at the three pH values is correlated with the disappearance of the smallest aggregates and is independent of the type of higher polymer formed. To investigate the effect of ionic strength and pH on the characteristic temperature corresponding to an optical density increment of 0.01 by the method used previously for vulgare, two sets of turbidity measurements were carried out. In the first one the ionic strength was changed from 0.025 to 0.15 in increments of 0.025 at pH 6.0 and 6.4. In the other set, the ionic strength was kept constant at 0.10 and the pH changed from 5.9 to 6.7 in increments of 0.1 pH units. When the analysis of these data was carried out, ΔH1 = 30 kcal/mol was obtained. For the salting out constant a value of 1.7 was found, compared to 2.2 for vulgare, a result consistent with the fact that E66 should be less hydrophobic than vulgare. The electrical work term ΔWel also turns out to be about one-half that for vulgare, which is expected from the lower net negative charge on E66 protein.  相似文献   

12.
The rates of polymerization and depolymerization of sickle cell hemoglobin   总被引:4,自引:0,他引:4  
The polymerization and depolymerization of concentrated solutions of sickle cell deoxyhemoglobin were initiated by raising and lowering the temperature, and the time courses of the reactions monitored by the change in apparent turbidity. The polymerization reaction exhibits a marked lag phase followed by a rapid increase in turbidity, and is dependent on a very high power of the hemoglobin concentration, roughly the fifteenth. The depolymerization reaction exhibits no such lag, and is much less dependent on concentration. The implications of these results for polymerization models are discussed.  相似文献   

13.
Horseradish peroxidase catalyzed the H2O2-dependent oxidation and polymerization of acetaminophen. Six acetaminophen polymers were isolated from horseradish peroxidase reaction mixtures by semipreparative high pressure liquid chromatography. Chemical structures were determined by a combination of electron impact and chemical ionization mass spectrometry and 500-MHz proton magnetic resonance spectroscopy. Two dimers, three trimers, and one tetramer were identified. The polymers formed primarily through a covalent bond between carbons ortho to the hydroxyl group, and to a lesser extent, between the carbon ortho to the hydroxyl group and the amino group of another acetaminophen molecule. Greater than 99% of the polymerization reaction products were quenched by the addition of 2.0 mM ascorbate. High acetaminophen concentration favored dimer formation, whereas low acetaminophen concentration favored formation of trimers and tetramers. Since approximately 1 mol of H2O2 was consumed per mol of covalent ligand formed between acetaminophen molecules, these products probably result from free radical termination reactions.  相似文献   

14.
The three tobacco (Nicotiana tabacum L.) S-adenosyl-L-methionine: o-diphenol-O-methyltransferases (OMTs; EC 2.1.1.6) were purified to homogeneity by affinity chromatography on adenosine-agarose. Amounts and catalytic actities of the enzymes were measured in tobacco leaves during the hypersensitive reaction to tobacco mosaic virus. The drastic increase in activity of each enzyme upon infection was shown to arise from the accumulation of enzymatic protein with constant specific enzymatic activity. Rates of OMT synthesis were determined from pulse-labeling experiments with L-[14C]leucine injected into the leaves. The specific radioactivities of the homogenous enzymes were compared in healthy and tobacco mosaic virus-infected tobacco. The results demonstrated that increase in OMT amounts is a consequence of de novo synthesis of the enzymes.Abbreviations DEAE diethylaminoethyl - OMT O-methyltransferase - SAM S-adenosyl-L-methionine - TMV tobacco mosaic virus  相似文献   

15.
The effect of the dipolar ions, glycine, glycylglycine, and glycylglycylglycine on the polymerization of tobacco mosaic virus (TMV) protein has been studied by the methods of light scattering and ultracentrifugation. All three dipolar ions promote polymerization. The major reaction in the early stage is transition from the 4 S to the 20 S state. As in the absence of dipolar ions, the polymerization is enhanced by an increase in temperature; it is endothermic and therefore entropy-driven. The effect of the dipolar ions can be understood in terms of their action as salting-out agents; they increase the activity coefficient of TMV A protein, the 4 S material, and thus shift the equilibrium toward the 20 S state. The salting-out constants, K, for the reaction in 0.10 ionic strength phosphate buffer at pH 6.7 was found by the light scattering method to be 1.6 for glycine, 2.5 for glycylglycine, and 2.5 for glycylglycylglycine. A value of 2.7 was obtained by the ultracentrifugation method for glycylglycine in phosphate buffer at 0.1 ionic strength and pH 6.8 at 10 degrees C. For both glycine and glycylglycine, K increases when the ionic strength of the phosphate buffer is decreased. This result suggests that electrolytes decrease the activity coefficient of the dipolar ions, a salting-in phenomenon. However, the salting-in constants evaluated from these results are substantially higher than those previously determined by solubility measurements. The effect of glycine and glycylglycine on polymerization was studied at pH values between 6.2 and 6.8. The effectiveness of both dipolar ions is approximately 50% greater at pH 6.8 than at pH 6.2. The variation of the extent of polymerization with pH in the presence of the dipolar ions is consistent with the interpretation that approximately one hydrogen ion is bound for half of the polypeptide units in the polymerized A protein.  相似文献   

16.
M F Carlier  D Pantaloni 《Biochemistry》1978,17(10):1908-1915
In vitro polymerization of pig brain tubulin, highly purified and deprived of microtubule-associated proteins, was followed by turbidimetry. Treatment of the data yielded the relation existing between the observed turbidity and the amount of polymer formed. This allowed a kinetic analysis, according to Oosawa's theories, of the polymerization process, which consisted of a slow spontaneous nucleation followed by the growth process. The apparent elongation rate constant was closely related to the nucleation process and exhibited a highly cooperative variation with tubulin concentration. The cooperativity was indicative of the size of the nucleus which appears to remain the same whether sheets or microtubules are formed. Magnesium ions appear to play a role in the polymorphism of tubulin polymers, the proportion of microtubules to sheets increasing with magnesium ion concentration. From kinetic experiments evidence was provided for GDP binding in competition with GTP, with a sixfold lower affinity. The tubulin-GDP complex could participate in microtubules elongation, but was not able to form nuclei. The critical concentration of tubulin in the presence of GDP was roughly twice as high as in the presence of GTP.  相似文献   

17.
Nucleation of alpha 1-antichymotrypsin polymerization   总被引:3,自引:0,他引:3  
Alpha(1)-antichymotrypsin is an acute phase plasma protein and a member of the serpin superfamily. We show here that wildtype alpha(1)-antichymotrypsin forms polymers between the reactive center loop of one molecule and the beta-sheet A of a second at a rate that is dependent on protein concentration and the temperature of the reaction. The rate of polymerization was accelerated by seeding with polymers of alpha(1)-antichymotrypsin and a complex of alpha(1)-antichymotrypsin with an exogenous reactive loop peptide but not with reactive loop cleaved alpha(1)-antichymotrypsin or with polymers of other members of the serpin superfamily. Sonication of alpha(1)-antichymotrypsin polymers markedly increased the efficacy of seeding such that polymers were able to form under physiological conditions. Taken together, these data provide the first demonstration that serpin polymerization can result from seeding. This mechanism is analogous to the fibrillization of the Abeta(1-42) peptide and may be important in the deposition of alpha(1)-antichymotrypsin in the plaques of Alzheimer's disease.  相似文献   

18.
P E Prevelige  Jr  D Thomas    J King 《Biophysical journal》1993,64(3):824-835
The polymerization of protein subunits into precursor shells empty of DNA is a critical process in the assembly of double-stranded DNA viruses. For the well-characterized icosahedral procapsid of phage P22, coat and scaffolding protein subunits do not assemble separately but, upon mixing, copolymerize into double-shelled procapsids in vitro. The polymerization reaction displays the characteristics of a nucleation limited reaction: a paucity of intermediate assembly states, a critical concentration, and kinetics displaying a lag phase. Partially formed shell intermediates were directly visualized during the growth phase by electron microscopy of the reaction mixture. The morphology of these intermediates suggests that assembly is a highly directed process. The initial rate of this reaction depends on the fifth power of the coat subunit concentration and the second or third power of the scaffolding concentration, suggesting that pentamer of coat protein and dimers or trimers of scaffolding protein, respectively, participate in the rate-limiting step.  相似文献   

19.
Recombinant DNA derived tobacco mosaic virus (vulgare strain) coat protein (r-TMVP) was obtained by cloning and expression in Escherichia coli and was purified by column chromatography, self-assembly polymerization, and precipitation. SDS-PAGE, amino terminal sequencing, and immunoblotting with polyclonal antibodies raised against TMVP confirmed the identify and purity of the recombinant protein. Isoelectric focusing in 8 M urea and fast atom bombardment mass spectrometry demonstrated that the r-TMVP is not acetylated at the amino terminus, unlike the wild-type protein isolated from the tobacco plant derived virus. The characterization of r-TMVP with regard to its self-assembly properties revealed reversible endothermic polymerization as studied by analytical ultracentrifugation, circular dichroism, and electron microscopy. However, the details of the assembly process differed from those of the wild-type protein. At neutral pH, low ionic strength, and 20 degrees C, TMVP forms a 20S two-turn helical rod that acts as a nucleus for further assembly with RNA and additional TMVP to form TMV. Under more acidic conditions, this 20S structure also acts as a nucleus for protein self-assembly to form viruslike RNA-free rods. The r-TMVP that is not acetylated carries an extra positive charge at the amino terminus and does not appear to form the 20S nucleus. Instead, it forms a 28S four-layer structure, which resembles in size and structure the dimer of the bilayer disk formed by the wild-type protein at pH 8.0, high ionic strength, and 20 degrees C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Interactions of both purified tubulin and microtubule protein (tubulin plus associated proteins) with two commonly used sulfonate buffers were examined. 1,4-Piperazineethanesulfonate (Pipes) and 4-morpholineethanesulfonate (Mes) at high concentrations induce the polymerization of purified tubulin in reactions requiring only buffer, tubulin and GTP. While both reactions were temperature-dependent, cold-reversible and inhibited by GDP, colchicine or Ca2+, there were significant differences between them. Substantially lower tubulin and buffer concentrations were required for Pipes-induced polymerization; and turbidity was much more intense in the Pipes-induced than in the Mes-induced reaction at the same protein concentration. Electron microscopy demonstrated that for the most part typical smooth-walled microtubules were formed in Mes, while aberrant forms were the predominant structures formed in Pipes. When the polymerization of microtubule protein was examined as a function of buffer concentration, biphasic patterns were observed with both Pipes and Mes: polymerization occurred at both low and high, but not intermediate, buffer concentrations. The turbidity observed at high concentrations of Pipes greatly exceeded that at low concentrations. With Mes, equivalent turbidity developed at both high and low buffer concentrations. Although associated proteins copolymerized with tubulin at low buffer concentrations, they were excluded from the polymerized material at high buffer concentrations. Pipes and Mes were compared to sodium phosphate, Tris/HCl and imidazole/HCl buffers at 0.1 M in several polymerization systems using both purified tubulin and microtubule protein. The sulfonate buffers were invariably associated with more vigorous reactions than the other buffers.  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