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 46 毫秒
1.
The presence of low concentrations of methanol or isopropyl alcohol (2-5%, v/v) in the assay medium stabilizes the latency of dynein 1 from sea urchin sperm flagella, with about a 50% decrease in ATPase level compared to that in the absence of solvent. Somewhat higher concentrations (10-20%, v/v) of these solvents in the assay give a 5-10-fold activation of ATPase activity. Dioxane, formamide, and dimethylformamide, on the other hand, always activate the ATPase activity, with a 5-10-fold increase observed at about 15% (v/v). The activation of latent ATPase activity by solvents is reversible for short exposures, especially in the presence of ATP and at low temperature, but the activation becomes irreversible upon more prolonged exposure. The rate constant for irreversible activation by 16% methanol at 21 degrees C is 0.08 min-1, compared to rates of 0.44 and 0.02 min-1 for activation by 0.05% Triton X-100 at 21 and 0 degree C, respectively. The slowness of this reversible activation induced by methanol and by Triton X-100 suggests that it is the result of large-scale conformational changes in the structure of the dynein. However, the activation by methanol occurs without the dissociation of the alpha and beta subunits of dynein that is observed with Triton X-100. The presence of 1 mM MgATP, or of 100 microM MgATP and 10 microM vanadate substantially protects latent dynein from activation by 0.05% Triton X-100.  相似文献   

2.
《Bioresource technology》2000,71(2):167-172
The effect of organic solvents on the chlorination activity of chloroperoxidase (CPO) was identified for use in biotransformations with CPO. CPO was found to chlorinate monochlorodimedon (MCD) in the presence of organic solvents with log P values less than 0. The relative rates of chlorination with chloride ion in the presence of H2O2, buffer and 2.5–20% of either dimethyl sulfoxide, N,N-dimethyl formamide, methanol or acetonitrile, were in the range of 10–58% of that in buffer (pH 2.8) at the same reactant concentrations. The presence of such organic solvents was found to alter CPO catalysis by altering the protein conformation and the local environment at the active site. CPO did not display chlorination activity in the presence of organic solvents which had log P values greater than 0.  相似文献   

3.
The effect of methanol on the folding of staphylococcal nuclease has been investigated. Equilibrium thermal unfolding transitions were monitored by fluorescence emission. The transition was very sensitive to the presence of methanol (at pH 7.0), the Tm decreased from above 50 degrees C for aqueous solution to below 0 degree C for 70% methanol. The transitions were fully reversible and conformed to two-state behavior. A linear relationship was observed between the hydrophobicity of the solvent and both the Tm and the change in delta G for unfolding. The effect of pH on the transition in 50% methanol at 0 degree C was essentially the same as for aqueous solution, with a cooperative transition in the vicinity of apparent pH (pH*) 4. The unfolding transition was determined as a function of guanidine thiocyanate in aqueous and 50% methanol solvents. The midpoints of the transitions were 0.30 and 0.20 M, respectively, at 2.1 degrees C. The kinetics of folding at 0 degree C were compared in aqueous, 50% methanol and 0.30 M guanidine thiocyanate solvents, by monitoring changes in the tryptophan fluorescence intensity. Triphasic kinetics for refolding in both aqueous and 50% methanol solutions were observed in stopped-flow experiments. In both solvent systems the slowest phase is ascribed to proline isomerization. The kinetics of refolding were monitored at subzero temperatures in 50% methanol at pH* 7.0 in manual mixing experiments. Biphasic kinetics were observed at temperatures between 0 and -35 degrees C. A third, faster phase, was inferred from the missing amplitude. The energies of activation were 20.0 and 17.2 kcal mol-1, respectively, for the two slower phases. At -33.8 degrees C, the observed pseudo first-order rate constants were 1.2 x 10(-3) and 2.1 x 10(-5) s-1. At temperatures above -35 degrees C, the sum of the observed amplitudes was essentially constant at 70-75% of the expected total amplitude. At lower temperatures the amplitude of the refolding reaction decreased, and the native state was not formed (unless the temperature was increased), due to the formation of a trapped intermediate state. This intermediate has circular dichroism and fluorescence properties consistent with a compact state with some molten globule characteristics.  相似文献   

4.
The stability of diester-diterpenoid alkaloids (DDA) from plants of the genus Aconitum L. has been studied in different solvents and pH buffers. The HPLC/ESIMS method for analysing the concentration of DDA was established and DDA's decomposition products were elucidated by HPLC/ESI-MS/MS(n). In different solvents, e.g. dichloromethane, ether, methanol and distilled water, the decomposition pathways of DDA are quite different and their difference in stabilities depends on the difference of their structures, in which substituents at the N atom and substituents at C-3 are different. The pyrolytic products of DDA, such as deacetoxy aconitine-type alkaloids, have been observed in the above solvents, whereas 8-methoxy-14-benzoyl aconitine-type alkaloids have been obtained only in methanol. Furthermore, the experimental results demonstrate that the stability of DDA depends on pH values of the buffer. Aconine as hydrolysate has been only found in pH 10.0 buffer, and the other hydrolysates and the pyrolyzates of DDA, such as benzoylaconine and deacetoxy aconitine, have been observed in all pH aqueous solutions. The decomposition pathways of DDA in buffers are related to the substituent on the C-3 position. The decomposition pathway of aconitine is similar to that of mesaconitine, but different from that of hypaconitine.  相似文献   

5.
The stability of a 15-residue peptide has been investigated using CD spectroscopy and molecular simulation techniques. The sequence of the peptide was designed to include key features that are known to stabilize alpha-helices, including ion pairs, helix dipole capping, peptide bond capping, and aromatic interactions. The degree of helicity has been determined experimentally by CD in three solvents (aqueous buffer, methanol, and trifluoroethanol) and at two temperatures. Simulations of the peptide in the aqueous system have been performed over 500 ps at the same two temperatures using a fully explicit solvent model. Consistent with the CD data, the degree of helicity is decreased at the higher temperature. Our analysis of the simulation results has focused on competition between different side-chain/side-chain and side-chain/main-chain interactions, which can, in principle, stabilize the helix. The unfolding in aqueous solution occurs at the amino terminus because the side-chain interactions are insufficient to stabilize both the helix dipole and the peptide hydrogen bonds. Loss of capping of the peptide backbone leads to water insertion within the first peptide hydrogen bond and hence unfolding. In contrast, the carboxy terminus of the alpha-helix is stable in both simulations because the C-terminal lysine residue stabilizes the helix dipole, but at the expense of an ion pair.  相似文献   

6.
A study was made of the effects of common protein denaturants and water-miscible organic solvents on both the stability and activity of the malic enzyme [(S)-malate:NADP+ oxidoreductase (oxaloacetate-decarboxylating); EC 1.1.1.40] from the extreme thermoacidophilic archaebacterium Sulfolobus solfataricus. At 25 degrees C, the enzyme was not inactivated in 4 M urea or 0.05% SDS over 24 h, while the half-life was 30 min in 6 M guanidine hydrochloride and 5 h in 0.075% SDS. The enzyme stability in water-miscible organic solvents at 25 degrees C is somewhat surprising: after a 24-h incubation, the enzyme was completely active in 50% dimethylformamide; it lost 15% of its initial activity in 50% methanol or 15% ethanol. However, the resistance to organic solvents was greatly reduced at higher temperatures. The enzyme was able to catalyze the malate conversion even in the presence of 1.5% Triton X-100 or sodium deoxycholate. A number of solvents were found to stimulate the malic activity independent of time. Studies with 50% methanol revealed that the activation was reversible and inversely related to the temperature; moreover, the solvent was demonstrated to exclusively affect the maximal velocity of catalysis, the Km values for both substrates being unchanged. Investigation was made to find out whether there was a correlation between enzyme stability, as well as activation, and hydrophobicity of the organic medium. The residual malic activity after incubation in the water/organic medium correlated inversely with the logarithm of the partition coefficient in octanol/H2O of the mixture used as a hydrophobicity index. On the other hand, the extent of activation depended directly on the logarithm of the molar concentration of the organic solvent required for maximal enzymatic activation. Because of its remarkable resistance to organic solvents required for maximal enzymatic activation. Because of its remarkable resistance to organic solvents and protein denaturants in general, the malic enzyme from Sulfolobus solfataricus can be considered suitable for biotechnological applications.  相似文献   

7.
A L Fink  D Kar  R Kotin 《Biochemistry》1987,26(26):8571-8579
In order to determine the necessary conditions to stabilize intermediates in ribonuclease A catalysis at subzero temperatures for structural studies, we have examined the suitability of alcohol-based cryosolvents. On the basis of thermal denaturation transition curves, the enzyme is in the native conformation in high concentrations of ethanol and methanol, provided the temperature is suitably low. The effects of methanol on the catalytic properties for the hydrolysis for mono- and dinucleotide substrates also are consistent with the absence of adverse effects of the cosolvent. Significant methanolysis occurs in the presence of methanol as cosolvent. The kinetics of 2',3'-CMP hydrolysis are complicated by severe competitive product inhibition, both in aqueous and in methanolic solvents, accounting for the previously observed effect of substrate concentration on the observed Km. Computer-aided analysis allowed the determination of the inhibition constant as a function of experimental parameters. The reaction of ribonuclease A with 2',3'-CMP was investigated at subzero temperatures. The turnover reaction could be made negligible at temperatures below -60 degrees C at pH 3-6 in 70% methanol and below -35 degrees C at pH 2.1. The rate of the catalytic reaction with crystalline enzyme was compared to that of enzyme in solution for both 2',3'-CMP and the dinucleotide CpC. The rates were 50- and 200-fold slower, respectively, in the crystal. These investigations allowed calculation of the necessary conditions for NMR and X-ray diffraction experiments on the trapped enzyme--substrate intermediate.  相似文献   

8.
R J Coll  A L Fink 《Cryobiology》1987,24(4):332-344
A comparative study of nucleophilic (methanol), aprotic (dimethyl sulfoxide), and protic but non nucleophilic (ethylene glycol, ethylene glycol/dimethylformamide) solvents on the catalytic and structural properties of human plasmin has been made. All four solvent systems are potentially suitable as cryosolvents for plasmin catalysis at subzero temperatures although the solubility of plasmin is limited in the methanol and dimethyl sulfoxide systems. Each cryosolvent system caused minor effects on the catalytic properties of the enzyme, which could be rationalized in terms of the known physical properties of the cosolvent. Solvent systems containing ethylene glycol induce a minor conformational change which increases the catalytic efficiency of plasmin. The cosolvent effects on Km and Ki indicate that electrostatic interactions dominate the binding of both substrates and inhibitors such as benzamidine. A change in slope of the Arrhenius plots for catalysis, reflecting a temperature-induced isomerization, is observed around 0 degree C; the energies of activation being 13 +/- 2 kcal mol-1 at higher temperatures and 19 +/- 2 kcal mol-1 at subzero temperatures, and essentially independent of solvent. Deacylation was shown to be the rate-limiting step in the hydrolysis of specific p-nitrophenyl ester substrates. Previous stopped-flow studies at room temperature provided observations suggesting that a tetrahedral intermediate could be detected in the plasmin-catalyzed hydrolysis of p-nitroanilide substrates. Experiments at subzero temperatures with such substrates failed to reveal any buildup of a tetrahedral intermediate under the experimental conditions.  相似文献   

9.
The rates of hydrolysis of acetyl phosphate in the presence of 0.1 M NaOH and of ATP in the presence of either 1 M HCl or 1 M NaOH were measured at different temperatures and in the presence of different concentrations of the organic solvents dimethyl sulfoxide or ethylene glycol. Under all conditions tested, there was a progressive increase in the rate constant of hydrolysis of both phosphate compounds as the water activity of the medium was decreased by the addition of organic solvents. At 25 degrees C, substitution of 70% of the water of the medium by dimethyl sulfoxide promoted an increase of two orders of magnitude in the rate constant of acetyl phosphate hydrolysis. In the presence of 80% and 90% dimethyl sulfoxide the rate of acetyl phosphate hydrolysis increased by more than two orders of magnitude and was so fast that it could not be measured with the method used. The effect of organic solvents on the rate of ATP hydrolysis was less pronounced than that observed for acetyl phosphate hydrolysis. At 30 degrees C, substitution of 90% of water by an organic solvent promoted a 4-6-fold increase of the rate of ATP hydrolysis. Acceleration of either acetyl phosphate or ATP hydrolysis rates was promoted by a decrease in both activation energies (Ea) and in entropies of activation delta S. The data obtained are discussed with reference to the mechanism of catalysis of enzymes involved in energy transduction such as the Ca2+-ATPase of sarcoplasmic reticulum and the F1-ATPase of mitochondria.  相似文献   

10.
The kinetic properties of cruciform extrusion in supercoiled DNA molecules fall into two main classes. C-type cruciforms extrude in the absence of added salt, at relatively low temperatures, with large activation energies, while S-type cruciforms exhibit no extrusion in the absence of salt, and maximal rates at 50 mM NaCl, with activation energies about one quarter those of the C-type. These diverse properties are believed to reflect two distinct pathways for the extrusion process, and are determined by the nature of the sequences which form the context of the inverted repeat. C-type kinetics are conferred by A + T rich sequences, implying a role of helix stability in the selection. In this study we have shown that: 1. Helix-destabilising solvents (dimethyl formamide and formamide) facilitate extrusion by normally S-type molecules at low temperatures in the absence of salt. 2. C-type extrusion is strongly suppressed by low concentrations (2-4 microM) distamycin, at which concentrations S-type extrusion is enhanced. 3. Some extrusion occurs in a C-type construct in the presence of 50 mM NaCl. This is increased by addition of 3 microM distamycin, under which conditions extrusion becomes effectively S-type. Thus S-type constructs can behave in a quasi-C-type manner in the presence of helix-destabilising solvents, and C-type extrusion is suppressed by binding a compound which stabilises A + T rich regions of DNA. Helix destabilisation leads to C-type behaviour, while helix stabilisation results in S-type properties. These studies demonstrate the influence of contextual helix stability on the selection of kinetic mechanism of cruciform extrusion.  相似文献   

11.
The stability of tenuazonic acid solution at different temperatures and storage times was studied using methanol, methanol-water (8:2 v/v), benzene and benzene-acetonitrile (98:2 v/v) as solvents. Solutions were analysed by a spectrometric method TeA U.V.-spectrum was recorded. Results indicated that the optimum temperature for long-time storage period of tenuazonic acid solution in any solvent assayed is -20°C. Benzene and benzene-acetonitrile (98:2 v/v) could be advised to make tenuazonic acid solution which will be stored less than 2 months at 4°C. Methanol and methanolwater (8:2 v/v) are not recommended because a low stability of TeA solution in this solvents.  相似文献   

12.
The effect of extraction solvent and time on the measured indole-3-acetic acid (IAA) level was investigated in plant materials having different contents of lAA-conjugates, Tissues from pine ( Pinus sylvestris L.). tobacco ( Nicotiana tabacum L.), and maize ( Zea mays L.) were extracted for 1–9 h with Na-phosphate buffer (pH 7.5). 80% methanol and 70% acetone. IAA was measured by combined gas chromatography-selected ion minitoring-mass spectromctry (GC-SIM-MS) with [13C6]-IAA as an internal standard.
Extraction of maize seedlings with buffer gave a higher estimate of free IAA than did extraction with methanol or acetone, which produced similar values. The increase in free IAA after buffer extraction was paralleled by a stoichiometric decrease in lAA-ester conjugates, indicating that free IAA was formed during buffer extraction by hydrolysis of these conjugates, which are abundant in maize seedlings. The amount of hydrolysis during a 1-h extraction period was estimated to be ca 3% of the total lAA-ester pool. However, in the pine extraxylary tissues and tobacco in-ternodes which lack a significant lAA-ester pool, buffer extraction resulted in the same IAA estimate as extraction with the organic solvents, but produced a cleaner extract. For all the plant materials investigated, a 1-h extraction period was sufficient for equilibrating the internal standard with the endogenous IAA pool.  相似文献   

13.
Standardized specimens with reproducible staining properties were fabricated from extracts of biological objects (bovine liver, nucleoprotamine and defatted muscle). The standard specimens were stained with two formulations of the Romanowsky-Giemsa stain (RG), using the same azure B and eosin Y. One formulation used methanol and Sorensen's buffer and the other DMSO and Hepes buffer as solvents. The standard specimens were stained either in the composite stain or in the individual dyes dissolved in the same solvents and at the same concentration as the composite stain. Solution spectroscopy demonstrated different spectra for the two formulations with some wavelength regions varying by more than an order of magnitude. The RG spectra were also very different from those of the individual dyes dissolved at the RG concentration in the respective solvents. The stained standard specimens were analyzed by microspectrophotometry and were found to have spectra similar to those of cell smears. Furthermore, the standard specimens were shown to be a repeatable substrate for stain uptake. The transmitted light intensity from random fields of the same standardized specimen varied +/- 5%. When specimens were stained at the same time, the specimen-to-specimen variation depended on preparation conditions and the measurement wavelength, but was as good as +/- 5% for some conditions. The quantitative stain performance of both formulations was studied and compared. The standardized specimens provide a tool for the quantitative study of staining processes and specimen preparation procedures and for stain calibration.  相似文献   

14.
The aggregation behavior of peptides Ac‐VQIVYK‐amide (AcPHF6) and Ac‐QIVYK‐amide (AcPHF5) from the amyloidogenic protein tau was examined by atomic force microscopy (AFM) and fluorescence microscopy. Although AcPHF5 did not show enhancement of thioflavin T (ThT) fluorescence in aqueous buffer, distinct aggregates were discernible when peptide was dissolved in organic solvents such as methanol (MeOH), trifluoroethanol (TFE), and hexafluoroisopropanol (HFIP) dried on mica and examined by AFM. Self‐association was evident even though the peptide did not have the propensity to form secondary structures in the organic solvents. In dried films, the peptide adopts predominantly β‐conformation which results in the formation of distinct aggregates. ThT fluorescence spectra and fluorescence images indicate the formation of fibrils when AcPHF6 solutions in organic solvents were diluted into buffer. AcPHF6 had the ability to organize into fibrillar structures when AFM samples were prepared from peptide dissolved in MeOH, TFE, HFIP, and also when diluted into buffer. AcPHF6 showed propensity for β‐structure in aqueous buffer. In MeOH and TFE, AcPHF6 showed helical and β‐structure. Morphology of the fibrils was dependent on peptide conformation in the organic solvents. The structures observed for AcPHF6 are formed rapidly and long incubation periods in the solvents are not necessary. The structures with varying morphologies observed for AcPHF5 and AcPHF6 appear to be mediated by surfaces such as mica and the organic solvents used for dissolution of the peptides. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
We studied the effect of organic solvents on the kinetics of porcine pancreatic lipase (pp) for the resolution of racemic glycidol through esterification with butyric acid. We quantified ppl hydration by measuring water sorption isotherms for the enzyme in the solvents/mixtures tested. The determination of initial rates as a function of enzyme hydration revealed that the enzyme exhibits maximum apparent activity in the solvents/mixtures at the same water content (9% to 11% w/w) within the associated experimental error. The maximum initial rates are different in all the media and correlate well with the logarithm of the molar solubility of water in the media, higher initial rates being observed in the solvents/mixtures with lower water solubilities. The data for the mixtures indicate that ppl apparent activity responds to bulk property of the solvent. Measurements of enzyme particle sizes in five of the solvents, as function of enzyme hydration, revealed that mean particle sizes increased with enzyme hydration in all the solvents, differences between solvents being more pronounced at enzyme hydration levels close to 10%. At this hydration level, solvents having a higher water content lead to lower reaction rates; these are the solvents where the mean enzyme particle sizes are greater. Calculation of the observable modulus indicates there are no internal diffusion limitations. The observed correlation between changes in initial rates and changes in external surface area of the enzyme particles suggests that interfacial activation of ppl is only effective at the external surface of the particles. Data obtained for the mixtures indicate that ppl enantioselectivity depends on specific solvent-enzyme interactions. We make reference to ppl hydration and activity in supercritical carbon dioxide. (c) 1994 John Wiley & Sons, Inc.  相似文献   

16.
T Hofmann  A L Fink  B M Dunn 《Biochemistry》1984,23(22):5247-5256
Intrinsic spectral and kinetic properties of penicillopepsin and its action on N-acetylalanylalanyllysyl-p-nitrophenylalanylalanylalanine amide have been investigated at subzero temperatures in aqueous methanol and dimethyl sulfoxide solutions in an attempt to find evidence for or against a covalent mechanism in the catalyzed hydrolysis of peptide bonds. The study of fluorescence and circular dichroism spectra as a function of solvent concentrations gave no evidence for any solvent-induced structural effects at temperatures below the thermal denaturation transition. The effect of temperature on the intrinsic fluorescence of penicillopepsin in either 60% (v/v) methanol or 50% (v/v) dimethyl sulfoxide did not indicate any temperature-induced structural changes. On the other hand, Arrhenius plots for the hydrolysis reaction over the range 0 to -50 degrees C showed downward curvature. A probable explanation for this phenomenon is that the reduction in flexibility of the enzyme due to thermal and viscosity factors leads to the stabilization of a nonproductive conformation. The pH optima of kcat/Km are shifted from 5.1 in aqueous solvents to 5.6 in 60% methanol and to 6.6 in 50% dimethyl sulfoxide. Aqueous methanol caused small decreases of Km and of Kcat; the decrease in the latter was greater than that brought about by the decrease in the water concentration. In aqueous dimethyl sulfoxide, there was no detectable change in kcat up to 15%, but Km increased by more than an order of magnitude. Above 15%, only kcat/Km could be measured. No evidence for the accumulation of either covalent amino or covalent acyl intermediates was obtained when penicillopepsin was incubated at -70 degrees C in 67% methanol with several substrates. Although negative, these experiments do not rule out conclusively the involvement of covalent intermediates in penicillopepsin-catalyzed reactions.  相似文献   

17.
The thermal denaturation of lysozyme dissolved in aqueous phosphate buffer (pH 5.1) and glycerol was studied by Fourier-transform infrared (FTIR) spectroscopy. In both solvents, a single temperature-induced conformational transition was observed but at the distinctly different temperatures of 73 °C in aqueous buffer and 94 ± 2 °C in glycerol. No changes in the secondary structure were observed in glycerol up to 90 °C. Thus, FTIR data were consistent with the formation of a highly ordered molten globule state at temperatures below 90 °C followed by lysozyme unfolding at higher temperatures in glycerol.  相似文献   

18.
Temperature dependence of the α-helix conformation of bee venom melittin in methanol-water mixed solvents has been examined by NMR, in order to elucidate conformation stability and a phase diagram. At high methanol concentration of 100 - ca. 80 wt.%, melittin forms a full α-helix conformation in the temperature range from 25 °C to 60 °C. At intermediate methanol concentration of ca. 80 - ca. 25 wt.%, it undergoes a thermal transformation from a full α-helix to a partial α-helix. In solutions of low methanol concentrations of ca. 25 - 0 wt.%, partial α-helix monomers and their self-aggregated conformers coexist at low temperatures, and the relative number of the monomers increases with increase in temperature. The monomers turn to a random coil state at high temperatures only below ca. 10 wt. % methanol concentrations. The thermal transitions are discussed from the viewpoint of stability of intra-molecular hydrogen bonds and inter-molecular hydrophobic interactions.  相似文献   

19.
A combinatorial experimental technique was used to identify salts and salt mixtures capable of activating penicillin amidase in organic solvents for the transesterification of phenoxyacetate methyl ester with 1-propanol. Penicillin amidase was lyophilized in the presence of various chloride and acetate salts within 96-deep-well plates and catalytic rates measured to determine lead candidates for highly salt-activated preparations. The kinetics of the most active formulations were then further evaluated. These studies revealed that a formulation consisting of 98% (w/w) of a 1:1 KAc:CsCl salt mixture, 1% (w/w) enzyme, and 1% (w/w) potassium phosphate buffer was approximately 35,000-fold more active than the salt-free formulation in hexane, as reflected in values of V(max)/K(m). This extraordinary activation could be extended to more polar solvents, including tert-amyl alcohol, and to formulations with lower total salt contents. A correlation was found between the kosmotropic/chaotropic behavior of the salts (as measured by the Jones-Dole B coefficients) and the observed activation. Strongly chaotropic cations combined with strongly kosmotropic anions yielded the greatest activation, and this is likely due to the influence of the ions on protein-water and protein-salt interactions.  相似文献   

20.
Y Pan  M S Briggs 《Biochemistry》1992,31(46):11405-11412
Ubiquitin adopts a non-native folded structure in 60% methanol solution at low pH. Two-dimensional nuclear magnetic resonance (2D NMR) was used to measure the hydrogen-exchange rates of backbone amide protons of ubiquitin in both native and methanol forms, and to characterize the structure of ubiquitin in the methanol state. Protection factors (the ratios of experimentally determined exchange rates to the rates calculated for an unfolded polypeptide) for protons in the native form of ubiquitin range from less than 10 to greater than 10(5). Most of the protons that are protected from exchange are located in regions of hydrogen-bonded secondary structure. The most strongly protected backbone amide protons are those of residues comprising the hydrophobic core. Hydrogen exchange from ubiquitin in methanol solution was too rapid to measure directly by 2D NMR, so a labeling scheme was employed, in which exchange with solvent occurred while the protein was in methanol solution. Exchange was quenched by dilution with aqueous buffer after the desired labeling time, and proton occupancies were measured by 1H NMR of the native form of the protein. Protection factors for protons in the methanol form of ubiquitin range from 2.6 to 42, with all protected protons located in hydrogen-bonded structure in the native form. Again, the most strongly protected protons are those of residues in the hydrophobic core. Comparison of the patterns of the hydrogen-exchange rates in the native and methanol forms indicates that almost all of the native secondary structure persists in the methanol form, but that it is almost uniformly destabilized by 4-6 kcal/mol.(ABSTRACT TRUNCATED AT 250 WORDS)  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