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 93 毫秒
1.
The activity and conformational change of human placental cystatin (HPC), a low molecular weight thiol proteinase inhibitor (12,500) has been investigated in presence of guanidine hydrochloride (GdnHCl) and urea. The denaturation of HPC was followed by activity measurements, fluorescence spectroscopy and Circular Dichroism (CD) studies. Increasing the denaturant concentration significantly enhanced the inactivation and unfolding of HPC. The enzyme was 50% inactivated at 1.5 M GdnHCl or 3 M urea. Up to 1.5 M GdnHCl concentration there was quenching of fluorescence intensity compared to native form however at 2 M concentration intensity increased and emission maxima had 5 nm red shift with complete unfolding in 4–6 M range. The mid point of transition was in the region of 1.5–2 M. In case of urea denaturation, the fluorescence intensity increased gradually with increase in the concentration of denaturant. The protein unfolded completely in 6–8 M concentration of urea with a mid-point of transition at 3 M. CD spectroscopy shows that the ellipticity of HPC has increased compared to that of native up to 1.5 M GdnHCl and then there is gradual decrease in ellipticity from 2 to 5 M concentration. At 6 M GdnHCl the protein had random coil conformation. For urea the ellipticity decreases with increase in concentration showing a sigmoidal shaped transition curve with little change up to 1 M urea. The protein greatly loses its structure at 6 M urea and at 8 M it is a random coil. The urea induced denaturation follows two-state rule in which Native→Denatured state transition occurs in a single step whereas in case of GdnHCl, intermediates or non-native states are observed at lower concentrations of denaturant. These intermediate states are possibly due to stabilizing properties of guanidine cation (Gdn+) at lower concentrations, whereas at higher concentrations it acts as a classical denaturant.  相似文献   

2.
Deuterated rhamnolipids were produced using strain AD7 of Pseudomonas aeruginosa, which was progressively adapted to increasing levels of deuterium in D2O and carbon substrates. Fourteen different deuterated rhamnolipid structures, including structural isomers, were produced which is similar to normal protonated structures. There were two main products monorhamnolipid Rha-C10-C10 and dirhamnolipid Rha2-C10-C10. The levels of deuteration varied from 16% with 25% D2O + h-glycerol to 90% with 100% D2O + d-glycerol. When d-tetradecane was used with H2O, virtually all the deuterium appeared in the lipid chains while using h-tetradecane + D2O led to the majority of deuterium in the sugars. The adaptation to growth in deuterium appeared to be metabolic since no genetic changes could be found in the key rhamnolipid biosynthetic genes, the rhamnosyl transferases RhlB and RhlC. Deuterated sophorolipids were similarly produced using Candida bombicola and Candida apicola although in this case, no adaptation process was necessary. Up to 40 different sophorolipids were produced by these yeasts. However, unlike the rhamnolipids, use of D2O did not lead to any deuteration of the lipid chains, but direct incorporation into the lipid was achieved using d-isostearic acid. The results from these experiments show the feasibility of producing deuterated bioactive compounds from microorganisms coupled with the possibility of manipulating the pattern of labelling through judicious use of different deuterated substrates.  相似文献   

3.
The kinetics of the coil-to-helix transition of (dG-dC)3 in M NaCl, 45 mM sodium cacodylate, pH 7, were measured in H2O, D2O, 10 mol % ethanol, 10 mol % urea, and 10 mol % glycerol. At 43°C in H2O the recombination rate is 1.3 ± 0.2 × 107 M?1 s?1; the dissociation rate is 68 ± 10 s?1. The destabilization of the helix in 10 mol % ethanol and 10 mol % urea relative to water is primarily due to a large increase in the helix-dissociation rate. In 10 mol % glycerol, the destabilization of the helix is due to a decrease in the recombination rate and an increase in the dissociation rate. Above 20°C, two exponential decays longer than 1 μs are observed after a temperature jump. The slower relaxation time is 4–10 times faster than the bimolecular component and is independent of oligomer concentration. We attribute this relaxation to a rapid equilibrium between two helical states. At low temperatures and oligomer concentrations of 1 mM or greater, the helices aggregate in 1M NaCl. Experimental data are presented under conditions where aggregation is unimportant and evidence is given that the ΔH-determined spectroscopically is unaffected by aggregation.  相似文献   

4.
Differential scanning microcalorimetry was used to investigate the enthalpy (ΔHd) and the temperature (td) of thermal denaturation of normal (nondeuterated) (H-PC) and deuterated (D-PC) phycocyanins in D2O solvent. Values of td in D-PC are about 5–7°C lower than those in H-PC. The magnitudes of ΔHd in D-PC are only 21–32% of those in H-PC. During the protein unfolding, the heat-capacity changes (ΔCp) in D-PC are also lower than those in H-PC. CD was employed to evaluate the secondary structure and the urea denaturation of these proteins in D2O solvent. These proteins have about the same α-helix content. D-PC is less resistant to the denaturant urea than is H-PC. In general, the apparent free-energy change in the process of protein unfolding at zero denaturant concentration is higher in H-PC than in D-PC. Comparisons of the present results for D2O solvent with those previously reported for H2O reveal that solvent isotope effect essentially does not change the α-helix content in H-PC and D-PC. However, D-PC or H-PC has a higher random-coil content in its secondary structure in D2O than in H2O. Substitution of H2O with D2O as the solvent increases td in both D-PC and H-PC, lowers ΔHd in H-PC, and greatly lowers ΔHd in D-PC. The deuterium solvent isotope effect does not change ΔCp in H-PC but lowers ΔCp in D-PC. In the urea denaturation, the magnitudes of (Cu)1/2 in H-PC and D-PC are not affected by such a solvent effect, whereas those of ΔG are greatly increased. These results are correlated with the structure and stability of the proteins.  相似文献   

5.
M Boublik  N Brot  H Weissbach 《Biopolymers》1973,12(9):2083-2092
Ribosomal proteins L7 and L12 are the only acidic proteins found on the 50S ribosomal subunit of Escherichia coli. The effect of ionic strength, helix-promoting solvents and denaturating agents on the conformation of these proteins has been studied. It has been established that the helicity of L7 and L12 proteins (approx. 45–50% α helix) can be increased to 60–70% when they are exposed to helix-promoting solvents such as methanol or ethanol in the presence of 0.1M salt. High ionic strength by itself was without any effect on the conformation of the proteins. However, the solvent, 2,2,2-trifluoroethanol increased the content of α helices up to 80% even in the absence of salt. Denaturating agents like urea (6M) or guanidine HCl (6M), decreased the content of the ordered structure below 20%. All conformational changes induced by salt or solvents were completely reversible and characterized by a broad transition showing a low degree of cooperativity. This might indicate the presence of discrete segments with variations in amino acid sequences and ordered structures with different stabilities.  相似文献   

6.
The helix-coil transitions for poly(L -glutamic acid) (PGA) in 0.2M NaCl and in its mixture with dioxane were studied by the methods of spectropolarimetry, viscometry, and potentiometric titration at different temperatures from 8 to 50°C. The enthalpy and entropy differences between the helical and coillike states of uncharged PGA molecules were determined from the curves of potentiometric titration. The temperature dependence of the cooperativity parameter σ was determined by two methods: from the sharpness of transition and from the dependence of the intrinsic viscosity on the helical content in the transition region. In 0.2MNaCl, σ= (2.5 ± 0.5) × 10?3 and practically does not depend on temperature, i.e., the cooperativity of the helix-coil transition is connected mainly with the entropy decrease in initiating helical regions (ΔSi ≈ ?12 is mole of helical regions). On the contrary, initiation of a helical region in the water-organic solvent mixture is accompanied by a considerable enthalpy increase.  相似文献   

7.
Thiol proteinase inhibitors are crucial to proper functioning of all living tissues consequent to their cathepsin regulatory and myriad important biologic properties. Equilibrium denaturation of dimeric goat pancreas thiol proteinase inhibitor (PTPI), a cystatin superfamily variant has been studied by monitoring changes in the protein's spectroscopic and functional characteristics. Denaturation of PTPI in guanidine hydrochloride and urea resulted in altered intrinsic fluorescence emission spectrum, diminished negative circular dichroism, and loss of its papain inhibitory potential. Native like spectroscopic properties and inhibitory activity are only partially restored when denaturant is diluted from guanidine hydrochloride unfolded samples demonstrating that process is partially reversible. Coincidence of transition curves and dependence of transition midpoint (3.2M) on protein concentration in guanidine hydrochloride‐induced denaturation are consistent with a two‐state model involving a native like dimer and denatured monomer. On the contrary, urea‐induced unfolding of PTPI is a multiphasic process with indiscernible intermediates. The studies demonstrate that functional conformation and stability are governed by both ionic and hydrophobic interactions. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 708–717, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

8.
Interaction of l-lysine with Co(II) and Cu(II) ions has been studied using 1H- and 13C-NMR and solution absorption spectrometry. In l-lysine-Co2+ solution in D2O (100: 1 in concentration), coordination interaction of the α-amino and carboxyl groups with Co2+ occurs from the neutral to alkaline pD region, whereas no interaction of the ?-amino group was observed throughout the whole pD region. On the other hand, in l-lysine-Cu2+ solution, the ?-amino group also takes part in complexation in the higher pD region (pD≧10). Structural changes in complexation of l-lysine with the divalent cations along with pD variations in aqueous solution are discussed. Dissociation constants of the three functional groups were obtained by 1H-NMR chemical shifts; pKa1 = 2.2, pKa2 = 9.5 and pKa3 = 11.2.  相似文献   

9.
T V Barskaya  O B Ptitsyn 《Biopolymers》1971,10(11):2181-2197
The helix–coil transitions for poly-L -lysine (PL) were investigated by the methods of spectropolarimetry, viscometry and potentiometric titration in 0.2M NaCl at different temperatures as well as in 0.2MNaBr, 1MKCl, and in mixtures of 0.2MNaCl or NaBr with methanol at room temperature. The enthalpy and entropy differences between the helical and coillike states of uncharged PL molecules in 0.2.M NaCl were determined from the potentiometric titration curves. The cooperativity parameters σ for PL in different solvents were determined by two methods (from the sharpness of the transition and from the dependence of the intrinsic viscosity on the helical content in the transition region). In 0.2MNaCl σ has a value of (2.3 ± 0.5) × 10?4 and does not depend on temperature, i.e., the cooperativity of the helix-coil transition, as for PGA, is mainly of an entropy origin (the initiating of the helical region is accompanied by the entropy decrease ΔSi = ?12 eu/mole of helical regions). A comparison of the obtained results for PGA and PL with the molecular theories of the helix-coil transitions shows that the role of dipole-dipole interactions of nonneighboring peptide groups is greatly overestimated in these theories, leading to a considerable enthalpy contribution to the free energy of initiating helical regions which is not observed in the experiment.  相似文献   

10.
R McGuire  I Feldman 《Biopolymers》1975,14(5):1095-1102
The fluorescence parameters—lifetime, relative quantum yield, wavelength of maximum fluorescence intensity, half-width, and polarization—of 0.01% lysozyme were measured at 15°C in aqueous solution, in glycerol–water mixtures (0–90% v/v glycerol), in aqueous urea (0–8M) solutions, and in aqueous guanidine hydrochloride (0–6.4M) solutions. The changes in the static and dynamic quenching of lysozyme fluorescence, monitored by the quantum yield and lifetime measurements, were correlated with the other fluorescence parameters and compared with our earlier results with bovine serum albumin. The results were interpreted in terms of induced conformational changes. The various perturbants altered the fluorescence parameters of lysozyme and bovine serum albumin very differently. The differences were shown to be entirely consistent with our earlier conclusion that bovine serum albumin fluorophores are nonsurface residues and with the conclusion of others that lysozyme fluorophores are surface residues. Unlike their effects on bovine serum albumin, urea and guanidine hydrochloride affect lysozyme structure quite differently, both in nature and degree. We have suggested that the affect of urea on lysozyme fluorescence is an indirect result of reduction in the size of the cleft brought about by the structure-breaking action of urea on water in the cleft. 4M Urea is sufficient for this reaction. Large decreases in the polarization of the fluorescence of lysozyme in the 0.8–1.6M and 3.2–4.8M guanidine hydrochloride ranges demonstrated two guanidine hydrochloride-induced conformation changes. A red shift of the fluorescence maximum to 354 nm indicated that the second transition completely exposes all fluorescing tryptophan residues of lysozyme to mobile solvent water. However, even 6.4M guanidine hydrochloride did not completely unravel the lysozyme molecule at 15°C, as evidenced by its failure to cause any of the tyrosine residues to become fluorescent.  相似文献   

11.
The denaturation behavior of bovine lens gamma-crystallin fractions II, III, and IV and their susceptibility to proteolysis in vitro was compared to determine whether differences in their stability could play a role in cataract formation. Tertiary and secondary structure changes induced by increasing concentrations of urea, guanidine hydrochloride, and sodium dodecyl sulfate and by increasingly alkaline pH were followed by near-UV and far-UV circular dichroism, Trp fluorescence emission, and exposure of sulfhydryl groups. Major differences were found in the denaturation and proteolysis behavior of the three gamma-crystallin fractions. In general, the unfolding of gamma-II and gamma-III crystallins is rather gradual, suggesting the presence of intermediate unfolding states; in contrast, the order-disorder transition of gamma-IV crystallin is abrupt. The gamma-IV crystallin fraction is the most stable in urea and guanidine hydrochloride, but is most susceptible to nonspecific proteolysis and alkaline pH denaturation. Differences in denaturation and proteolysis behavior are attributed to the inherent differences in the tertiary structures of these crystallins.  相似文献   

12.
13.
In recent years, many advances have been made in the understanding of functional and structural characteristics of protein evolution from denaturant-based studies that subject the protein to a change in the microenvironment. This paper reports the chemical denaturation of purified goat muscle cystatin (GMC) a thiol-proteinase inhibitor, using urea and guanidine hydrochloride (GdnHCl). The subtle conformational changes of GMC were monitored by intrinsic fluorescence, extrinsic fluorescence, and CD spectroscopic techniques. Further, the activity of GMC as a function of increasing concentration of denaturants was also studied. It was found that increasing the concentration of GdnHCl significantly enhances the inactivation and unfolding of the inhibitor (GMC). In urea-induced denaturation, the intrinsic and extrinsic fluorescence intensity reveals significant structural changes in the inhibitor. Further, it was found that at low concentrations of urea, up to 0.5–1.0 M, there was quenching of fluorescence intensity compared with the native form and a red shift of 5 nm was observed up to 5–8 M. The results presented in this paper suggest that GdnHCl-induced denaturation of GMC follows a simple two-state rule in which native → denatured state transition occurs in a single step. However denaturation with urea proceeds through an intermediate or non-native state.  相似文献   

14.
The thermal stability of an isometric plant virus, Turnip Yellow Mosaic Virus (TYMV), has been investigated at low and high hydrostatic pressure, using small angle neutron scattering. Contrast variation allowed us to separately observe the structural changes of the protein capsid and the RNA core. The experiments were performed in 0.05M Tris buffer at pD = 8.0 and in 0.05M bis-Tris buffer at pD = 6.0 containing different H2O/D2O mixtures (40% and 70% D2O). It was found that hydrostatic pressure enhances the stability of TYMV. The thermally induced uncoating of RNA as well as structural transitions of the protein capsid are shifted to higher temperature upon increasing the pressure from 5 × 106 Pa to 2 × 108 Pa.  相似文献   

15.
High-molecular-weight poly(0,0′-dicarbobenzoxy-L -β-3,4-dihydroxyphenyl-α-alanine) was prepared by the N-carboxyanhydride method. From the results obtained by a study of the optical rotation, nuclear magnetic resonance, and solution infrared absorption, the conformation of poly(0,0′-dicarbobenzoxy-L -β-3,4-dihydroxyphenyl-α-alanine) depended greatly on the solvent taking a right-handed helix with [θ]225 = ?13,600 ~ ?18,900 in alkyl halides, a left-handed helix with [θ]228 = 22,100 ~ 24,800 in cyclic ethers or trimethylphosphate, and a random coil structure in dichloroacetic acid, trifluoroacetic acid, or hexafluoroacetone sesquihydrate. The polypeptide underwent a right-handed helix-coil transition in chloroform/dichloroacetic acid (or trifluoroacetic acid) mixed solvents and a left-handed helix-coil transition in dioxane/dichloroacetic acid (or trifluoroacetic acid) mixed solvents. The results were compared with those of poly(0-carbobenzoxy-L -tyrosine).  相似文献   

16.
The conformational transitions of schizophyllan were studied in aqueous alkaline solutions by high-sensitivity differential scanning calorimetry (DSC) and optical rotation measurements. The temperature of half completion for reversible intramolecular conformational transition determined by DSC, centered at 7.4°C in water, increases to 37.2°C at 0.01M KOH with increasing alkaline concentration. The transition enthalpy per mole of the polysaccharide repeating unit is 2.62 ± 0.23 kJ mol−1 independent of the alkaline concentration. The cooperative unit size for the transition decreases with increasing alkaline concentration. Optical rotation was measured as a function of pH at 25 and 60°C. A sharp decrease in optical rotation was observed at pH = 13, which is ascribed to the triple helix-coil transition. From data obtained by DSC and optical rotation measurements, in combination with results reported previously, a phase diagram for the conformation of schizophyllan as a function of temperature and pH is proposed. The irreversibility of the triple helix to single coil transition, induced by strong alkali, was investigated as a function of polymer concentration by gel permeation chromatography and electron microscopy. The renatured samples at polymer concentrations < 1.0 mg/mL, which are prepared by dissolution in 0.25M KOH followed by neutralization with HCl, are observed as a mixture of globular, linear, and circular structures, and larger aggregates with less-defined morphology by electron microscopy. Higher concentrations lead to increased proportions of multichain clusters (aggregates). Subsequent annealing of the renatured samples at 115–120°C increases the proportion of circular species. The change in molecular weight distribution of samples that accompanies the renaturation and annealing mentioned above can be well interpreted in terms of the proportion of species having different morphology as observed by electron microscopy. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
H J Hinz  W Haar  T Ackermann 《Biopolymers》1970,9(8):923-936
The enthalpies of the helix-coil transitions of the ordered polynucleotide systems of poly(inosinic acid)–poly(cytidylic acid) [poly(I + C)], (helical duplex), and of poly (inosinic acid) [poly(I + I + I)], (proposed secondary structure: a triple-stranded helical complex), were determined by using an adiabatic twin-vessel differential calorimeter. Measuring the temperature course of the heat capacity of the aqueous polymer solutions, the enthalpy values for the dissociation of the helical duplex poly (I + C) and the three-stranded helical complex poly(I + 1 + 1), respectively, were obtained by evaluating the additional heat capacity involved in the conformational change of the polynucleotide system in the transition range. The ΔH values of the helix-coil transition of poly (I + C) resulting from the analysis of the calorimetric measurements vary between the limits 6.5 ± 0.4 kcal/mole (I + C) and 8.4 ± 0.4 kcal/mole (I + C). depending on the variation of the cation concentration ranging from 0.063 mole cations kg H2O to 1.003 mole cations/kg H2O. The calorimetric investigation of an aqueous poly I solution (cation concentration 1.0 mole/kg H2O) yielded the enthalpy value ΔH = 1.9 ± 0.4 kcal/mole (I), a result which has been interpreted qualitatively following current models of inter- and intramolecular forces of biologically significant macromolecules. Additional information on the transition behavior of poly(I+ C)Was obtained by ultraviolet and infrared absorption measurements.  相似文献   

18.
19.
Dielectric relaxation and viscosity measurements were performed on delipidated and relipidated samples of bovine serum albumin (BSA) at urea concentrations between O and 6M. By the combined interpretation of these two hydrodynamic methods the characterization of conformational changes of the molecule during urea denaturation is possible. The denaturation of delipidated BSA results from two mechanisms. The first one is a slow, time-dependent elongation of the molecule; the second one is a rapid swelling which becomes most pronounced at urea concentrations higher than 4M. For relipidated albumin, the slow elongation mechanism occurs but the presence of fatty acids protects the protein aganist molecular swelling. In both cases these conformational changes are accompanied by an increased disymmetry of charge repartition and a concomitant increase of the dipole moment. From these results it follows that lipidated albumin (as occurs under physiological conditions) is less sensitive to denaturation than delipidated albumin.  相似文献   

20.
The denaturation of proteins by guanidine hydrochloride was studied in the presence of different concentrations of stabilizing salts, namely potassium phosphate, ammonium sulfate, and potassium acetate. The denaturation transition was followed by observing changes in the peptide circular dichroism atpH 7.0 and 25°C. From these results the free energy of stabilization for the process native denatured was determined. It was found that the stabilizing power of the anions increased in the order acetate < sulfate < phosphate, in agreement with the anionic lyotropic series. Ribonuclease A, which is known to have a site that can bind either a phosphate or a sulfate ion, showed a larger stabilization by these anions than that for lysozyme, pepsinogen, and myoglobin.  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