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 influence of ionic strength or the concentration of K+ ([K+]) of the aqueous phase on the spontaneous transfer of cholesterol between negatively charged bilayer vesicles composed of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylserine (DPPS) (1:1, mole:mole) was studied using a pyrene-labelled cholesterol analogue, 1-pyrenemethyl-3 beta-hydroxy-22,23-bisnor-5-cholenate (PMC), as the probe. The decrease in PMC excimer fluorescence was best fitted to a bi-exponential function. Increasing [K+] from 0.1 M to 0.3 M had little effect on the shorter half-time (1.4 +/- 0.2 min) but increased the longer half-time from 16.3 +/- 1.9 min to 26.7 +/- 2.1 min. Fluorescence quenching and titration of an interface-located fluorophore, 1-anilinonaphthalene-8-sulfonic acid (ANS) revealed an increase in interfacial hydrophobicity upon increasing in ionic strength. The physical state of the acyl chains was not affected by ionic strength as indicated by a constant PMC excimer:monomer fluorescence intensity ratio. However, an increase in enthalpy change of the lipid phase transition from 15.7 kJ/mol ([K+] = 0.1 M) to 21.3 kJ/mol ([K+] = 0.3 M), together with a slight increase in the transition temperature, implies that interactions between adjacent molecules in the charged lipid bilayer vesicles became stronger at higher ionic strength. Our results suggest that the van der Waals attraction between PMC and phospholipid molecules could be affected by conformation changes in the charged head group region brought about by changes of ionic strength in the aqueous phase, with consequent effects on the desorption of cholesterol from the bilayer surface.  相似文献   

2.
We recently described site-specific pyrene labeling of RNA to monitor Mg(2+)-dependent equilibrium formation of tertiary structure. Here we extend these studies to follow the folding kinetics of the 160-nucleotide P4-P6 domain of the Tetrahymena group I intron RNA, using stopped-flow fluorescence with approximately 1 ms time resolution. Pyrene-labeled P4-P6 was prepared using a new phosphoramidite that allows high-yield automated synthesis of oligoribonucleotides with pyrene incorporated at a specific 2'-amino-2'-deoxyuridine residue. P4-P6 forms its higher-order tertiary structure rapidly, with k(obs) = 15-31 s(-1) (t(1/2) approximately 20-50 ms) at 35 degrees C and [Mg(2+)] approximately 10 mM in Tris-borate (TB) buffer. The folding rate increases strongly with temperature from 4 to 45 degrees C, demonstrating a large activation enthalpy DeltaH(double dagger) approximately 26 kcal/mol; the activation entropy DeltaS(double dagger) is large and positive. In low ionic strength 10 mM sodium cacodylate buffer at 35 degrees C, a slow (t(1/2) approximately 1 s) folding component is also observed. The folding kinetics are both ionic strength- and temperature-dependent; the slow phase vanishes upon increasing [Na(+)] in the cacodylate buffer, and the kinetics switch completely from fast at 30 degrees C to slow at 40 degrees C. Using synchrotron hydroxyl radical footprinting, we confirm that fluorescence monitors the same kinetic events as hydroxyl radical cleavage, and we show that the previously reported slow P4-P6 folding kinetics apply only to low ionic strength conditions. One model to explain the fast and slow folding kinetics postulates that some tertiary interactions are present even without Mg(2+) in the initial state. The fast kinetic phase reflects folding that is facilitated by these interactions, whereas the slow kinetics are observed when these interactions are disrupted at lower ionic strength and higher temperature.  相似文献   

3.
We have shown the dependence of the deamidation half-times of the peptides, GlyLeuGlnAlaGly and GlyArgGlnAlaGly upon pH, temperature, and ionic strength. Increase in temperature or ionic strength, variation of pH to pH′s higher or lower than pH 6, and the use of phosphate buffer rather than Tris buffer at high pH all decrease the half-time of dcamidation. Temperature increase of 20°C or pH change of 2 pH units decreases the half-time about fivefold, while increase of one ionic strength unit decreases the half-time about twofold. In pH 7.4, I = 0.2, 37.0°C phosphate buffer, the deamidation half-times are 663 ± 74 and 389 ± 56 days respectively for the two peptides, GlyLeuGlnAlaGly and GlyArgGlnAlaGly.These experiments should serve as a warning to peptide and protein experimenters that even the more stable glutaminyl residues are unstable with respect to deamidation in certain solvent conditions. These experiments also provide, along with previously reported experiments on asparaginyl peptides (7), some quantitative data to help with the extrapolation of in vitro deamidation experiments to in vivo deamidation conditions.  相似文献   

4.
Subtilisin is an unusual example of a monomeric protein with a substantial kinetic barrier to folding and unfolding. Here we document for the first time the in vitro folding of the mature form of subtilisin. Subtilisin was modified by site-directed mutagenesis to be proteolytically inactive, allowing the impediments to folding to be systematically examined. First, the thermodynamics and kinetics of calcium binding to the high-affinity calcium A-site have been measured by microcalorimetry and fluorescence spectroscopy. Binding is an enthalpically driven process with an association constant (Ka) equal to 7 x 10(6) M-1. Furthermore, the kinetic barrier to calcium removal from the A-site (23 kcal/mol) is substantially larger than the standard free energy of binding (9.3 kcal/mol). The kinetics of calcium dissociation from subtilisin (e.g., in excess EDTA) are accordingly very slow (t1/2 = 1.3 h at 25 degrees C). Second, to measure the kinetics of subtilisin folding independent of calcium binding, the high-affinity calcium binding site was deleted from the protein. At low ionic strength (I = 0.01) refolding of this mutant requires several days. The folding rate is accelerated almost 100-fold by a 10-fold increase in ionic strength, indicating that part of the free energy of activation may be electrostatic. At relatively high ionic strength (I = 0.5) refolding of the mutant subtilisin is complete in less than 1 h at 25 degrees C. We suggest that part of the electrostatic contribution to the activation free energy for folding subtilisin is related to the highly charged region of the protein comprising the weak ion binding site (site B).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Analysis of the results of calorimetric study of reconstituted collagen (type I) fibrils, in particular, the half-width of the temperature transition, shows that the collagen packing density in the fibrils and the size of cooperative blocks therein depend on the assembly temperature and on the initial collagen concentration. The least dense fibrils are formed at subphysiological temperatures (25° or 30°C) and low concentration (0.3 mg/ml). The extent of ordering does not change upon doubling the concentration but increases upon quadrupling it. At physiological temperature (35°C) the fibrils are densely packed regardless of collagen concentration. The enthalpy of fibril assembly is minimal at 35°C, 1.2 mg/ml, and ionic strength of 0.17 M. The influence of temperature on particular steps of fibrillogenesis and the role of water in these processes are discussed.  相似文献   

6.
Mode of high temperature injury to wheat during grain development   总被引:5,自引:0,他引:5  
High temperature stress adversely affects wheat growth in many important production regions, but the mode of injury is unclear. Wheat ( Triticum aestivum L. cv. Newton) was grown under controlled conditions to determine the relative magnitude and sequences of responses of source and sink processes to high temperature stress during grain development. Regimes of 25°C day/15°C night, 30°C day/20°C night, and 35°C day/25°C night from 5 days after anthesis to maturity differentially affected source and sink processes. High temperatures accelerated the normal decline in viable leaf blade area and photosynthetic activities per unit leaf area. Electron transport, as measured by Hill reaction activity, declined earlier and faster than other photosynthetic processes at the optimum temperature of 25/15 °C and at elevated temperatures. Changes in RUBP carboxylase activities were similar in direction but smaller in magnitude than changes in photosynthesic rate. Increased protease activity during senscence was markedly accentuated by high temperature stress. Specific protease activity increased 4-fold at 25/15 °C and 28-fold at 35/25 °C from 0 to 21 days after initiation of temperature treatments. Grain-filling rate decreased from the lowest to the highest temperature, but the change was smaller than the decrease in grain-filling duration at the same temperatures. We concluded that a major effect of high temperature is acceleration of senescence, including cessation of vegetative and reproductive growth, deterioration of photosynthetic activities, and degradation of proteinaceous constituents.  相似文献   

7.
The kinetics of respiration-dependent proton efflux and membrane energization have been studied in intact cells of logarithmic-phase Escherichia coli. Parallel measurements of the rate and extent of proton efflux into the external medium (half-time, about 10 s; ratio of H(+) to O, about 0.5) and the oxidation of E. coli cytochrome b (half-time, 相似文献   

8.
J Mo  M E Holtzer  A Holtzer 《Biopolymers》1991,31(12):1417-1427
Stopped flow CD (SFCD) kinetic studies of self-assembly of coiled coils of rabbit alpha alpha-tropomyosin and of nonpolymerizable alpha alpha-tropomyosin (NPTm) are reported. The protein was denatured in 6 M urea buffer, then renatured by 10-fold dilution into benign saline buffer. Folding was monitored by SFCD in the backbone region (222 nm). Protein chains are shown to be totally unfolded (and separated in the reduced species) in the initial denaturing medium and fully folded as two-chain coiled coils in the final benign medium. In all cases of folding in benign buffer of totally unfolded chains, two phases were found in the folding process: a fast phase (less than 0.04 s, the SFCD dead time), in which an intermediate state with about 70% of the equilibrium ellipticity forms; followed by a slower, observable phase that completes the folding. The slow phase is first order (k-1 = 1.6 s at 20 degrees C), signifying that chain association for reduced samples occurs in the fast phase. In contrast, folding in benign buffer from an initial state with 70% of the equilibrium ellipticity is all fast, suggesting that the folding intermediate is not an equilibrium species. Cross-linking at Cys-190 increases the helix content of the fast-formed intermediate state to about 85% of the equilibrium value, but leaves the rate constant of the slow phase unchanged. In NPTm, which does not form high aggregates at low ionic strength, the rate of the observable phase is almost independent of ionic strength in the range of approximately 0.15-0.6 M, but is reduced one to two orders of magnitude by further reduction to 0.026 M. In folding from totally unfolded chains, the rate is reduced less than one order of magnitude by changing the final state to about 50% folded. In contrast to folding, unfolding of alpha alpha-tropomyosin from the native state is all fast.  相似文献   

9.
The effects of low temperature storage on the physiology of cut rose flowers ( Rosa hybridaL. cv. Mercedes) were studied. Extension of cold storage or increase in temperature (from 3 to 8°C) was accompanied by shortening of vase life and advancement of petal senescence, as reflected in an advance in the timing of the rise in ethylene production and an increase in membrane permeability (ion leakage). Although storage at a relative humidity (RH) of 65% reduced petal water content by 20% in comparison with flowers stored at 95% RH, it did not shorten vase life. The progression of petal senescence was measured during storage at 3°C and during aging at 22°C. Both ethylene production rates and membrane microviscosity measured by fluorescence depolarization increased with time at 3°C and at 22°C, but more slowly at 3°C. At 3°C membrane permeability measured by ion leakage did not increase. Following cold storage the rate of ethylene production in the petals was increased by up to eight times the rate in unstored flowers. Silver thiosulphate extended the vase life of both stored and fresh flowers equally by 2 days, but did not increase the life of stored flowers to that of treated fresh flowers. It is concluded that the primary effect of cold storage on roses is to slow down senescence and that the continued slow senescence leads to shorter vase life. The possible occurrence of sequential processes during senescence and the effects of temperature on these processes is discussed.  相似文献   

10.
A direct measurement of the unzippering rate of a nucleic acid double helix   总被引:1,自引:0,他引:1  
The rate of double helix unzippering was determined directly by application of a fast temperature jump method to a nucleotide system of partly unzippered helices formed from oligoriboadenylates and oligoribouridylates of equal chain lengths (14 and 18 nucleotide residues). These helices showed a relaxation process in the time range of 0.1 to 0.3 μsec, that is assigned to the unzippering reaction. Measurements at 0.05 M and 0.1 M [Na+] demonstrated a rather small dependence upon the ionic strength. Increase of temperature increases the rate of unzippering. Simulation of the unzippering relaxation by a zipper model yielded a rate constant of base pair formation adjacent to a helix sequence of 8 × 106 sec?1 at 25°C associated with an activation enthalpy of 4 kcalmole. This elementary rate constant is higher than that obtained from a simulation of the overall recombination and dissociation rates of entire helices. The difference is attributed to reduced electrostatic and steric hindrance effects for base pair equilibration at helix ends.  相似文献   

11.
In order to investigate the effects of temperature and ionic strength on the N-B-transition and the alkaline denaturation of the human serum albumin, the pH-dependences of fluorescence position and relative yield of Trp-24 and of protein bound dye ANS were measured. The measurements were carried out at temperatures from 10 to 45 degrees C and ionic strengths (NaCl) from 0.001 to 0.2. The pH-induced structural transitions have different realization in environments of tryptophanyl and tightly bound ANS. The alkaline denaturation does not change the Trp-214 fluorescence. The N-B-transition gives rise to the slight polarity and/or mobility lowering in the Trp-214 environment (the shorter-wave-length spectral shift). Increase in the temperature and ionic strength induces the shift of the transition midpoint from ca. 8 to 8.7 and reduces the spectral shift amplitude. At low ionic strengths, the new structural transition in the Trp-214 environment is observed at pH change from 6.7 to 5.7. This transition is not observable using ANS fluorescence. The N-B-transition is accompanied by an enhancement and longer-wavelength shift of the ANS fluorescence spectra. The transition midpoint is independent of temperature, but is shifted to lower pH values at a decrease of ionic strength value. At ionic strengths less than or equal to 0.01 the shorter-wavelength spectral shift is seen at pH from 7.5 to 9, which seems to reflect the disulfide B-A-isomerisation. The alkaline denaturation gives rise to the sharp quenching of ANS fluorescence, probably due to the ANS binding site decomposition.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The temperature dependence of the pressure-induced equilibrium unfolding of staphylococcal nuclease (Snase) was determined by fluorescence of the single tryptophan residue, FTIR absorption for the amide I' and tyrosine O-H bands, and small-angle X-ray scattering (SAXS). The results from these three techniques were similar, although the stability as measured by fluorescence was slightly lower than that measured by FTIR and SAXS. The resulting phase diagram exhibits the well-known curvature for heat and cold denaturation of proteins, due to the large decrease in heat capacity upon folding. The volume change for unfolding became less negative with increasing temperatures, consistent with a larger thermal expansivity for the unfolded state than for the folded state. Fluorescence-detected pressure-jump kinetics measurements revealed that the curvature in the phase diagram is due primarily to the rate constant for folding, indicating a loss in heat capacity for the transition state relative to the unfolded state. The similar temperature dependence of the equilibrium and activation volume changes for folding indicates that the thermal expansivities of the folded and transition states are similar. This, along with the fact that the activation volume for folding is positive over the temperature range examined, the nonlinear dependence of the folding rate constant upon temperature implicates significant dehydration in the rate-limiting step for folding of Snase.  相似文献   

13.
Abstract: Secretion of both epinephrine and norepinephrine by cultured chromaffin cells was studied at temperatures ranging from 0°C to 37°C. The percentage of epinephrine secreted was always lower than that of norepinephrine when the cells were stimulated with either acetylcholine or high K+ at any temperature. When the cells were stimulated with acetylcholine or carbachol the percentage of catecholamine secreted at 10 min increased with temperature from 4°C to 24°C and then decreased from 24°C to 37°C. Potassium-stimulated cells secreted increasing amounts of catecholamine as the temperature was increased to 37°C. We found, however, that the initial rates of secretion increased continuously as temperature increased throughout the range for both carbachol-and K+-stimulated cells. The temperature maximum of acetylcholine-stimulated secretion is caused by a faster shut-off of secretion at higher temperature. The Arrhenius plots of initial rates show an inflection point at approximately 17°C for carbachol-stimulated cells. The plot for K+-stimulated cells is a straight line over the entire temperature range. The transition could be caused by a conformational change in the cholinergic receptor/ion channel molecule.  相似文献   

14.
Unconjugated bilirubin (bilirubin-IX alpha), the hydrophobic end product of heme degradation, is esterified in the hepatocyte endoplasmic reticulum to water-soluble conjugates prior to excretion in bile. To characterize the process of intracellular bilirubin transport, the kinetic and thermodynamic activation parameters for the spontaneous transfer of bilirubin between small unilamellar egg lecithin vesicles were determined. Bilirubin-IX alpha was added to donor vesicles labeled with the fluorescent phospholipid probe, (5-(dimethylamino)naphthalene-1-sulfonyl) dipalmitoyl-L-alpha-phosphatidylethanolamine (dansyl-PE). When bound to the donor vesicles, bilirubin quenches the dansyl probe fluorescence through resonance energy transfer. The movement of bilirubin from dansyl-labeled donor vesicles to unlabeled acceptor vesicles was monitored directly by the reemergence of dansyl fluorescence over time. Vesicle fusion and intervesicle transfer of the dansyl-PE probe were excluded by quasielastic light scattering and fluorescence resonance energy transfer studies. Stopped-flow analysis demonstrated that the transfer of bilirubin was described by a single-exponential function with a mean half-time of 2.0 +/- 0.1 ms (+/- SD) at 37 degrees C. The rate of bilirubin transfer was independent of acceptor vesicle concentration and decreased with increasing buffer ionic strength, indicating that intermembrane transfer occurred via aqueous diffusion, rather than vesicle collisions. The free energy of activation (delta G++) for the dissociation of bilirubin from donor vesicles was 14.2 kcal.mol-1. These studies suggest that bilirubin is associated with phospholipid bilayers at the membrane-water interface. We postulate that the movement of unconjugated bilirubin between intracellular membranes occurs via spontaneous transfer through the aqueous phase.  相似文献   

15.
The interaction of the ethidium cation with calf thymus DNA is investigated in solutions of different ionic strength and temperature by observation of the enhancement of fluorescence of ethidium upon intercalation in the duplex structure. The quantum yield of the fluorescence of the intercalated dye is found to increase either upon lowering the Na+ concentration or upon increasing the temperature. The existence of a correlation between the geometry of the intercalation complex and the features of the secondary structure of DNA is suggested. Binding isotherms under corresponding environmental conditions are also quantitated by fluorescence enhancement and interpreted in terms of the neighbor exclusion model. Large contributions from change in hydration to the thermodynamics of binding are demonstrated by the temperature dependences of the equilibrium constants. The neighbor exclusion range is found to be practically independent of the salt concentration but its value increases from an average of 2.4 around room temperature to 4-5 at 80 degrees C, as inferred from the binding curves in 0.15 and 0.5 M [Na+] or from the DNA hypochromism vs temperature profiles of complexes at 10(-3) M [Na+]. All the data point to a possible sequence-conformation specificity in the intercalation of ethidium which in heterogeneous DNA is mediated by environmental changes.  相似文献   

16.
Changes in the temperature dependence of the photosynthetic rate depending on growth temperature were investigated for a temperate evergreen tree, Quercus myrsinaefolia . Plants were grown at 250 μ mol quanta m–2 s–1 under two temperature conditions, 15 and 30 °C. The optimal temperature that maximizes the light-saturated rate of photosynthesis at 350 μ L L–1 CO2 was found to be 20–25 and 30–35 °C for leaves grown at 15 and 30 °C, respectively. We focused on two processes, carboxylation and regeneration of ribulose-1,5-bisphosphate (RuBP), which potentially limit photosynthetic rates. Because the former process is known to limit photosynthesis at lower CO2 concentrations while the latter limits it at higher CO2 concentrations, we determined the temperature dependence of the photosynthetic rate at 200 and 1000 μ L L–1 CO2 under saturated light. It was revealed that the temperature dependence of both processes varied depending on the growth temperature. Using a biochemical model, we estimated the capacity of the two processes at various temperatures under ambient CO2 concentration. It was suggested that, in leaves grown at low temperature (15 °C), the photosynthetic rate was limited solely by RuBP carboxylation under any temperature. On the other hand, it was suggested that, in leaves grown at high temperature (30 °C), the photosynthetic rate was limited by RuBP regeneration below 22 °C, but limited by RuBP carboxylation above 22 °C. We concluded that: (1) the changes in the temperature dependence of carboxylation and regeneration of RuBP and (2) the changes in the balance of these two processes altered the temperature dependence of the photosynthetic rate.  相似文献   

17.
G R Parr  G G Hammes 《Biochemistry》1976,15(4):857-862
The kinetics of dissociation and reassembly of rabbit skeletal muscle phosphofructokinase has been studied using fluorescence, stopped-flow fluorescence and enzyme activity measurements. The dissociation of the fully active tetramer in 0.8 M guanidine hydrochloride (0.1 M potassium phosphate, pH 8.0) occurs in three kinetic phases as measured by changes in the protein fluorescence emission intensity: dissociation of tetramer to dimer with a relaxation time of a few milliseconds; dissociation of dimer to monomer with a relaxation time of a few seconds; and a conformational change of the monomer with a relaxation time of a few minutes. All three phases exhibit first-order kinetics; ATP (0.05 mM) retards the second step but does not influence the rate of the other two processes. The rate of the second process increases with decreasing temperature; this may be due to the involvement of hydrophobic interactions in the stabilization of the dimeric enzyme. A further unfolding of the monomer polypeptide chain occurs at higher guanidine concentrations, and the relaxation time associated with this process was found to be 83 ms in 2.5 M guanidine, 0.1 M potassium phosphate (pH 8.0) at 23 degrees C. The phosphofructokinase monomers were reassembled from 0.8 M guanidine chloride by 1:10 dilution of the guanidine hydrochloride concentration and yielded a protein with 70-94% of the original activity, depending on the protein concentration. The reactivation process follows second-order kinetics; ATP (5 mM) increases the rate of reactivation without altering the reaction order, while fructose 6-phosphate does not influence the rate of reaction. The rate-determining step is probably the association of monomers to form the dimer.  相似文献   

18.
We have examined the folding and unfolding of the caspase recruitment domain of procaspase-1 (CP1-CARD), a member of the alpha-helical Greek key protein family. The equilibrium folding/unfolding of CP1-CARD is described by a two-state mechanism, and the results show CP1-CARD is marginally stable with a DeltaG(H2O) of 1.1 +/- 0.2 kcal/mole and an m-value of 0.65 +/- 0.06 kcal/mole/M (10 mM Tris-HCl at pH 8.0, 1 mM DTT, 25 degrees C). Consistent with the equilibrium folding data, CP1-CARD is a monomer in solution when examined by size exclusion chromatography. Single-mixing stopped-flow refolding and unfolding studies show that CP1-CARD folds and unfolds rapidly, with no detectable slow phases, and the reactions appear to reach equilibrium within 10 msec. However, double jump kinetic experiments demonstrate the presence of an unfolded-like intermediate during unfolding. The intermediate converts to the fully unfolded conformation with a half-time of 10 sec. Interrupted refolding studies demonstrate the presence of one or more nativelike intermediates during refolding, which convert to the native conformation with a half-time of about 60 sec. Overall, the data show that both unfolding and refolding processes are slow, and the pathways contain kinetically trapped species.  相似文献   

19.
Equatorial X-ray diffraction patterns from single skinned fibres from bony fish muscle (turbot) were obtained with the fibres at 6 degrees C bathed in relaxing solutions of 170 down to 26 mM ionic strength. Diffraction patterns from rigor fibres were also obtained as controls. Unlike fibres from rabbit muscle, which show very clear evidence of substantial crossbridge formation at low ionic strength in what is mechanically a rapid equilibrium ("weak-binding") state (Brenner et al., 1982), diffraction patterns from bony fish fibres showed only a small change in relative peak intensities at low ionic strength (26 mM) compared with normal (170 mM) ionic strength. However, there was a slight ordering of the filament lattice at low ionic strength. The specimen temperature used (about 6 degrees C) was not far from the normal physiological temperature of the fish. Likewise, only a small change was seen by Xu et al. (1987) in patterns from frog fibres at low ionic strength at 2 to 6 degrees C. (Rabbit fibres previously studied, where large changes were seen at temperatures of 5 to 20 degrees C, were about 17 to 32 degrees C below physiological.) The I11/I10 ratio for fish fibres at 26 mM ionic strength was actually lower than that for rabbit even at normal ionic strength. This may be associated with an intrinsic structural difference between these muscles or alternatively with the disordering of the crossbridge helix in rabbit muscle found at low temperature by Wray (1987), and could support the view that rabbit fibres at 5 degrees C and normal ionic strength may already have a significant population of weak-binding crossbridges.  相似文献   

20.
《Biophysical journal》2022,121(4):552-564
Our knowledge of the folding behavior of proteins from extremophiles is limited at this time. These proteins may more closely resemble the primordial proteins selected in early evolution under extreme conditions. The small archaeal modifier protein 1 (SAMP1) studied in this report is an 87-residue protein with a β-grasp fold found in the halophile Haloferax volcanii from the Dead Sea. To gain insight into the effects of salt on the stability and folding mechanism of SAMP1, we conducted equilibrium and kinetic folding experiments as a function of sodium chloride concentration. The results revealed that increasing ionic strength accelerates refolding and slows down unfolding of SAMP1, giving rise to a pronounced salt-induced stabilization. With increasing NaCl concentration, the rate of folding observed via a combination of continuous-flow (0.1–2 ms time range) and stopped-flow measurements (>2 ms) exhibited a >100-fold increase between 0.1 and 1.5 M NaCl and leveled off at higher concentrations. Using the Linderström-Lang smeared charge formalism to model electrostatic interactions in ground and transition states encountered during folding, we showed that the observed salt dependence is dominated by Debye-Hückel screening of electrostatic repulsion among numerous negatively charged residues. Comparisons are also drawn with three well-studied mesophilic members of the β-grasp superfamily: protein G, protein L, and ubiquitin. Interestingly, the folding rate of SAMP1 in 3 M sodium chloride is comparable to that of protein G, ubiquitin, and protein L at lower ionic strength. The results indicate the important role of electrostatic interactions in protein folding and imply that proteins have evolved to minimize unfavorable charge-charge interactions under their specific native conditions.  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