首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The kinetics of hydrogen-deuterium exhcange in the polypeptide chain elongation factor Tu (EF Tu) from Escherichia coli and that from Thermus thermophilus HB8 has been examined in aqueous solutions at various pH and temperatures by means of infrared absorption measurements. The free EF-Tu from E. Coli has a greater reaction rate at all pH values and at every temperature than that of the GTP-bound or GDP-bound EF-Tu. The free EF-Tu from T. thermophilus, on the other hand, has an alomst equal reaction rate to that of EF-Tu-GDP in the temperature range 38-55 degrees C. For the peptide NH groups belonging to a medium-labile kinetic class, a small but definite difference in the rate of exchange reaction was observed between EF-Tu-GDP and EF-Tu-GTP for both E. coli and T. thermophilus. For less labile peptide NH groups, on the other hand, the rate of the exchange reaction with EF-Tu-GDP from T. thermophilus is only slightly affected by the pH of the solution at 38 degrees C and 45 degrees C, while the rate constant(k) with E. coli EF-Tu-GDP is pH-dependent (log k oc pH). For T. thermophilus EF-Tu, heat stability measurements, kinetics of the rates of GDP and GTP dissociation, and circular dichroic measurements have also been made. The molecular basis for the thermostability of T. thermophilus EF-Tu is discussed.  相似文献   

2.
The kinetics of the hydrogen-deuterium exchange reaction in bovine α-lactalbumin have been followed, by infrared absorption measurement, in aqueous solutions at various pH values and at various temperatures. A thermal transition which takes place at about 60 °C has been examined by ultraviolet absorption measurement and circular dichroism measurement.Outlines of the exchange kinetics and the thermal transition are quite similar to those observed for hen egg-white lysozyme, the amino acid sequence of which is known to be very similar to that of α-lactalbumin. Between these two proteins, however, differences have been found in the following respects. (1) The number of slowly exchanging peptide hydrogen atoms (35 in α-lactalbumin compared with 44 in egg-white lysozyme). (2) Kinetic profile of the slow exchange reaction. (3) The midpoint of the thermal transition (54 °C in water and 58 °C in deuterium oxide for α-lactalbumin, compared with 76 °C in both water and deuterium oxide for egg-white lysozyme). (4) The enthalpy and entropy changes in the transition (72 kcal/mol and 220 e.u., respectively, for α-lactalbumin, compared with 127 kcal/mol and 364 e.u. for egg-white lysozyme). (5) The circular dichroic spectrum of the “unfolded” molecule. (6) The effective amount of the unfolded forms estimated from the kinetic measurement at temperatures slightly lower than the transition temperature. (7) The effect of pH on the exchange kinetics.These differences between the proteins are interpreted in terms of the molecular structures and their fluctuations.  相似文献   

3.
The kinetics of the hydrogen-deuterium exchange reaction in a trypsin inhibitor (Kunitz) from soya bean have been followed by infrared absorption measurements in aqueous solutions at various temperatures and pH values. It was found that, in every case, 49% of the total peptide hydrogen atoms exchange relatively slowly. This amount corresponds to 83 peptide groups per molecule, and this is considered to be equal to the number of peptide NH groups involved in hydrogen bonds with the carbonyls of other peptide groups in the protein molecule in its native form. Each rate constant (k) determined at pH 2.75 for this category of the NH groups is in good agreement with the value expected from an idea that the breaking of the peptide-peptide hydrogen bonds takes place very slowly, and that this is the rate-determining process in the hydrogen-deuterium exchange reaction. Thus, by ultraviolet absorption measurements at 297 nm, the equilibrium constant of the native and denatured forms has been determined in the temperature range from 42 to 53.5 °C, as well as the reaction rate of reaching equilibrium from an off-equilibrium state. From these data the rate constant (k1) of the denaturation reaction is determined, and the k1 value is found to be practically equal to the hydrogen exchange rate constant (k). The Arrhenius plot of this rate constant (k) gives a straight line in the 25 to 55 °C region, and this gives a value of 48.6 kcal/mol for the activation energy of the denaturation reaction. The rate of this reaction is found to be very low at 25 °C; its half-life is about eleven days. Infrared absorption spectra observed in the amide I region suggest that the very slow denaturation of this protein is accompanied by a conformation change from an α-helix to a β-form. The number of the peptide groups involved in this αβ change is estimated to be 9 ± 3.  相似文献   

4.
H Fritzsche 《Biopolymers》1967,5(9):863-870
The base composition of DNA from microorganisms varies over a wide range of guanine + cytosine (G-C) content. We have examined the infrared spectra of a number of DNA samples isolated from different microorganisms and have found marked differences in their infrared spectra. These differences have been related to the G-C content of the investigated DNA samples. A comparison between the infrared spectra of nucleosides, synthetic polynucleotides, apurinic acid, and apyrimidinic acid has permitted a more extensive assignment of infrared bands to distinct base residues. The relative intensity of the absorption bands at 1485 and 1505 cm.?1 depends on the G-C content of the deuterated DNA sample. On the basis of these results a new method for the determination of the G-C content in DNA has been developed from the results. Its advantage is that the measurement is rapid and the DNA samples are not changed by the procedure.  相似文献   

5.
We have undertaken a new and more detailed Fourier-transform infrared (FTIR) spectroscopic study of alpha-lactalbumin (in D2O solution) aimed at correlating its secondary structures to observed Amide I' infrared bands. The spectra reported here were interpreted in light of the recently determined crystal structure of alpha-lactalbumin and by comparison with the spectra and structure of the homologous protein lysozyme. Of particular importance is the new evidence supporting the assignment of the band at 1639 cm-1 to 3(10)-helices. This assignment is in excellent agreement with one based on theoretical and experimental studies of 3(10)-helical polypeptides. The frequency observed for 3(10)-helices is distinctly different from that at which alpha-helices are typically found (viz., around 1655 cm-1). In the present study, two bands are clearly resolved in the latter region at 1651 and 1659 cm-1. Both are apparently associated with alpha-helices. These results suggest that for D2O solutions of globular proteins. FTIR spectroscopy can be a facile method for detecting the presence of these two different types of helical conformation and distinguishing between them. This provides a distinct advantage over ultraviolet circular dichroism spectroscopy (UV-CD). This work also provides a basis for future studies of alpha-lactalbumin which examine the effects of environment (e.g., pH, temperature) and ligands (e.g., Ca2+, Mn2+) on its conformation.  相似文献   

6.
1H--2H exchange kinetics of the peptide hydrogens in corticotropin have been examined in 2H2O and CF3C2H2O2H solutions by means of infrared absorption measurements. In aqueous solution, around pH 3, the experimental data suggest a partially ordered structure, since in the two corticotropins 1--24 and 1--32 about 6 slowly exchanging peptide protons are numbered. These might belong to the N-terminal part of the molecule. The C-terminal 25--32 octapeptide segment appears to be unordered and slightly destabilizes the overall hormone conformation. For corticotropin1--24 in CF3C2H2O2H, the qualitative interpretation of infrared spectra and the quantitative analysis of exchange data give evidence of a strong stabilization: a predominantly alpha-helical structure is induced by trifluoroethanol.  相似文献   

7.
The hydrogen–deuterium exchange reaction for the tryptophan residues in lysozyme have been followed in 4.5M LiBr at pH 7.2 in the temperature range of the unfolding transition by measuring the transmittance change at 293 nm. The exchange reaction proceeded in three phases at low temperature for native protein. The first and the second phases were ascribed to the H-D exchange reactions of three relatively exposed tryptophan residues on the molecular surface. The third phase corresponded to the H-D exchange reaction of the three tryptophan residues buried in the interior of the molecule. The H-D exchange reaction proceeded in two phases near the melting temperature and in a single phase at high temperature, where almost all molecules are unfolded. The H-D exchange of three tryptophan residues buried in folded molecules was caused by fluctuation between the folded and unfolded structure of the protein molecule. The rates of such a fluctuation were determined from the rates of the exchange reaction at various temperatures. These rates agreed very well with those determined from the temperature-jump method. This means that a protein molecule in solution fluctuates between the N- and D-states at every temperature within the transition region, where the N-form is the tightly folded native structure and the D-form the randomly coiled chain. From measurements of thermal unfolding of ester-108-lysozyme and the binding constant of (NAG)3 to ester-108-lysozyme, it was found that almost all cross-linked molecules are in the folded state near 50°C and pH 7.2 in 4.5M LiBr, where intact molecules are unfolded. We also studied the H-D exchange reaction of ester-108-lysozyme. In the temperature region of 43–50°C, about 70% of the exchangeable tryptophan residues of ester-108-lysozyme were exchanged within 1 s immediately after the mixing of D2O, in spite of the fact that almost all molecules are in the folded state. This was considered the premelting of the surface of a corss-linked molecule.  相似文献   

8.
The kinetics of hydrogen-deuterium exchange in hen egg-white lysozyme (muramidase) has been followed in aqueous solutions of various pH values and in solutions with various concentrations of lithium chloride, by an infrared absorption measurement. It was found that, in every case, 34% of the total peptide hydrogen atoms exchange relatively slowly with a rate of a first-order reaction. This amount corresponds to 44 peptide groups per molecule, and this is equal to the number of peptide NH-groups which are found to be involved in hydrogen bonds with the carbonyls of other peptide groups in the lysozyme molecule in the crystalline state. Each rate constant determined is in good agreement with the value expected from two simple assumptions. (1) The scheme of the isotope exchange reaction is N ? D → D (? N), where N is the native form of the molecule, D a denatured (unfolded) form, and ∗ indicates the deuterated products. (2) The N ? D fluctuation rate is much higher than the rate of the isotope exchange reaction D → D. It has been shown that the N ? D transition postulated here is the same as that which can be followed by circular dichroism measurement and by some other physical measurements. The effect of lithium chloride on the exchange reaction rate is solely attributable to the change in the N ? D equilibrium caused by the salt, whereas the effect of pH (in the 5 to 8 range) is wholly ascribed to the catalytic action of the OH anion on the D → D reaction rate. From the deuterium exchange rate observed, an effective value of the mole fraction of the D form is estimated to be 3 × 10−6in the solution with no lithium chloride at 20 °C and of pH = 5 to 8.  相似文献   

9.
The pH dependence of amide proton exchange rates have been measured for trp-repressor. One class of protons exchanges too fast to be measured in these experiments. Among the protons that have measurable hydrogen-deuterium exchange rates, two additional classes may be distinguished. The second class of protons are in elements of secondary structure that are mostly on the surface of the protein, and exchange linearly with increasing base concentration (log kex versus pH). The third class of amide protons is characterized by much higher protection against exchange at higher pH. These protons are located in the core of the protein, in helices B and C. The exchange rate in the core region does not increase linearly with pH, but rather goes through a minimum around pH 6. The mechanism of exchange for the slowly exchanging core protons is interpreted in terms of the two-process model of Hilton and Woodward (1979, Biochemistry 18:5834-5841), i.e., exchange through both a local mechanism that does not require unfolding of the protein, and a mechanism involving global unfolding of the protein. The increase in exchange rates at low pH is attributed to a partial unfolding of the repressor. It is concluded that the formation of secondary structure alone is insufficient to account for the high protection factors seen in the core of native proteins at higher pH, and that tertiary interactions are essential to stabilize the structure.  相似文献   

10.
Using the technique of frustrated multiple internal reflection infrared spectroscopy on germanium plates the hydrogen-deuterium (1H2H) exchange of protein films was studied. Provided that the films were kept under an atmosphere of at least 90% relative humidity the exchange characteristics were the same as in solutions. A quantitative comparison of the exchange pattern with protein structure data obtained from X-ray spectroscopy and tritium isotope studies proved the reliability of the film technique. As in solution the number of unexchanged hydrogens decreased with decreasing proton concentrations. From this and results on the ionization of outer protein surface carboxylate groups it is concluded that proteins at low pH exist in a compact structure, which is gradually opened with increasing pH by the ionization and subsequent hydration of the carboxyl groups.  相似文献   

11.
An extremely slow pH dependent conformational equilibrium between a valence-delocalized and a valence-trapped species of the dinuclear CuA domain of cytochrome c oxidase from Thermus thermophilus has been identified and characterized using UV-visible absorption, circular dichroism, time-resolved fluorescence and electron paramagnetic resonance spectroscopy as well as by stopped-flow kinetic techniques. The results indicated that the nature of this pH dependent conformation change in the CuA domain in the Thermus protein was distinctly different from that observed in the mesophilic analogue from Paracoccus denitrificans and in the engineered CuA domain in azurin. pH jump kinetic studies suggested existence of a fast deprotonation equilibrium followed by slow conformational change in the protein, which is contrary to that observed in the case of the analogous protein from P. denitrificans. Continuous-flow electrospray mass spectral studies on H/D exchange in the TtCuA showed that approximately 75% of the protons are exchanged within the dead-time of the experiment supporting fast proton transfer kinetics in the protein. Analysis of temperature dependence of the kinetics of the conformational transition showed that the rigidity of the protein structure decreases with increase in temperature. The results indicated that though the rate of proton transfer at individual sites in the protein could be very fast, the conformational change that requires simultaneous breaking of several interactions in a segment of the structure might be slow in the thermostable protein.  相似文献   

12.
Water is a highly polar molecule that is capable of making four H-bonding linkages. Stability and specificity of folding of water-soluble protein macromolecules are determined by the interplay between water and functional groups of the protein. Yet, under some conditions, water can be replaced with sugar or other polar protic molecules with retention of protein structure. Infrared (IR) spectroscopy allows one to probe groups on the protein that interact with solvent, whether the solvent is water, sugar or glycerol. The basis of the measurement is that IR spectral lines of functional groups involved in H-bonding show characteristic spectral shifts with temperature excursion, reflecting the dipolar nature of the group and its ability to H-bond. For groups involved in H-bonding to water, the stretching mode absorption bands shift to lower frequency, whereas bending mode absorption bands shift to higher frequency as temperature decreases. The results indicate increasing H-bonding and decreasing entropy occurring as a function of temperature, even at cryogenic temperatures. The frequencies of the amide group modes are temperature dependent, showing that as temperature decreases, the amide group H-bonds to water strengthen. These results are relevant to protein stability as a function of temperature. The influence of solvent relaxation is demonstrated for tryptophan fluorescence over the same temperature range where the solvent was examined by infrared spectroscopy.  相似文献   

13.
The secondary structure of the purified 70-kDa protein Na+/Ca2+ exchanger, functionally reconstituted into asolectin lipid vesicles, was examined by Fourier transform infrared attenuated total reflection spectroscopy. Fourier transform infrared attenuated total reflection spectroscopy provided evidence that the protein is composed of 44% alpha-helices, 25% beta-sheets, 16% beta-turns, and 15% random structures, notably the proportion of alpha-helices is greater than that corresponding to the transmembrane domains predicted by exchanger hydropathy profile. Polarized infrared spectroscopy showed that the orientation of helices is almost perpendicular to the membrane. Tertiary structure modifications, induced by addition of Ca2+, were evaluated by deuterium/hydrogen exchange kinetic measurements for the reconstituted exchanger. This approach was previously proven as a useful tool for detection of tertiary structure modifications induced by an interaction between a protein and its specific ligand. Deuterium/hydrogen exchange kinetic measurements indicated that, in the absence of Ca2+, a large fraction of the protein (40%) is inaccessible to solvent. Addition of Ca2+ increased to 55% the inaccessibility to solvent, representing a major conformational change characterized by the shielding of at least 93 amino acids.  相似文献   

14.
Myoglobin, a small globular heme protein that binds gaseous ligands such asO2, CO and NO reversibly at the heme iron, provides an excellent modelsystem for studying structural and dynamic aspects of protein reactions. Flashphotolysis experiments, performed over wide ranges in time and temperature, reveal a complex ligand binding reaction with multiple kinetic intermediates, resulting from protein relaxation and movements of the ligand within the protein. Our recent studies of carbonmonoxy-myoglobin (MbCO) mutant L29W, using time-resolved infrared spectroscopy in combination with x-ray crystallography, have correlated kinetic intermediates with photoproduct structures that are characterized by the CO residing in different internal protein cavities, so-called xenon holes. Here we have used Fourier transform infrared temperature derivative spectroscopy (FTIR-TDS) to further examine the role of internal cavities in the dynamics. Different cavities can be accessed by the CO ligands at different temperatures, and characteristic infrared absorption spectra have been obtained for the different locations of the CO ligand within the protein, enabling us to monitor ligand migration through the protein as well as conformational changes of the protein.  相似文献   

15.
Fourier-transform infrared (FTIR) spectroscopy has been used to study the thermally induced exchange characteristics of those backbone amide protons which persist H-D exchange at ambient conditions in ribonuclease A, in wild type ribonuclease T1 and some of its variants, and in the histone-like protein HBsu. The H-D exchange processes were induced by increasing the thermal energy of the protein solutions in two ways: (i) by linearly increasing the temperature, and (ii) by a temperature jump. To trace the H-D exchange in the proteins, various infrared absorption bands known to be sensitive to H-D exchange were used as specific monitors. Characteristic H-D exchange curves were obtained from which the endpoints (TH/D) of H-D exchange could be determined. The H-D exchange curves, the TH/D-values and the phase transition temperatures Tm were used to estimate the structural flexibility and stability of the given proteins. It is suggested that time-resolved FTIR spectroscopy can be used to determine global stability parameters of proteins.  相似文献   

16.
The accessible tyrosines of bovine insulin were studied by the photochemically induced dynamic nuclear polarization (photo-CIDNP) method. Tyrosine 1H nuclear polarization is observed in acidic, neutral, and basic solutions at all concentrations studied, in the absence of added salts as well as in the presence of 0.05-0.1 M chloride or phosphate. At pH 2.1 in the presence of chloride, at concentrations of 640 microM and above, most of the nuclear polarization at delta 6.82 originates from one group of tyrosines. On the basis of the crystallographic model, these are assumed to be the A14 tyrosines. We explored the possibility of a genuine concentration dependence of the photo-CIDNP intensity of insulin due to aggregation. In order to discern between such effects and trivial kinetic effects traceable to the optical irradiation method, the effects of concentration changes on polarization were examined in three apparently nonassociating trypsin inhibitor proteins. In insulin, the intensity of Tyr-A 14 polarization changes slowly at concentrations above 1 mM, suggesting that these residues are similarly accessible in all association states. At insulin concentrations below 320 microM, additional tyrosine emission signals were observed. These signals are probably due to B16 and B26 tyrosines of monomers. Polarization transfer effects from Tyr-A14 are evident in the tetramer and hexamer. Enhanced absorption effects in the two histidines (B5 and B10) of the insulin monomer were observed at pH 10 in the presence of 0.1 M phosphate.  相似文献   

17.
The kinetic of 1H leads to 3H exchange between water and C(8)H-groups of the guanylic residues in poly(G) . poly(C) and poly(dG) . poly(dC) was investigated within the temperature range from 30 to 90 degrees in 0.5 M NaCl (pH 7.2). It was shown that the exchange in freshly dissolved preparations at temperatures lower than 50 degrees proceeds faster than that in the case of GMP. According to the ylide mechanism of the exchange reaction the observed acceleration of the exchange is considered as a consequence of associates formation in poly(G) . poly(c) and poly(dG) . poly(dC) solutions at temperatures lower than 50 degrees. Associates are stabilized by intermolecular hydrogen bonds in which N(7) atoms of guanylic residues take part. The increase of the temperature is accompanied by gradual disappearance of the exchange acceleration. The retardation of exchange, which is characteristic of most non-associated double-stranded polynucleotides and nucleic acids is observed at the temperatures above 60 degrees. The retardation points to thermal destruction of the associates at temperatures higher than 50 degrees. The associates which are characterized by ordered structure including several "side by side" arranged double-stranded molecules were observed by electron microscopy. The addition of EDTA to solutions as well as the increase of temperature leads to destruction of the associates whereas the addition of Mg2+ makes the associates more stable.  相似文献   

18.
Vacuolar membrane-derived vesicles isolated from Vigna radiata catalyze oxygen exchange between medium phosphate and water. On the basis of the inhibitor sensitivity and cation requirements of the exchange activity, it is almost exclusively attributable to the vacuolar H+-pyrophosphatase (V-PPase). The invariance of the partition coefficient and the results of kinetic modeling indicate that exchange proceeds via a single reaction pathway and results from the reversal of enzyme-bound pyrophosphate synthesis. Comparison of the exchange reactions catalyzed by V-PPase and soluble PPases suggests that the two classes of enzyme mediate Pi---HOH exchange by the same mechanism and that the intrinsic reversibility of the V-PPase is no greater than that of soluble PPases.  相似文献   

19.
Phosphate-phosphate exchange through the inorganic phosphate (Pi) carrier of rat liver mitochondria was investigated by a new rapid filtration technique, which does not require the use of transport inhibitors to stop the reaction and offers high time resolution (starting from 10 ms), thus allowing kinetic measurements on a fine time scale even at room temperature. At approximately 22 degrees C, isotopic equilibrium of [32P]Pi is achieved within 0.8-2.5 s--depending on the Pi concentration--and an initial linear phase, lasting for 400-500 ms, is observed. Complete inhibition of Pi exchange by an excess (33 nmol/mg) of mersalyl, a well-known organomercurial inhibitor, required 200 ms, pointing to the insufficiency of this reagent for effective inhibitor stop. On the other hand, investigation of the effect of mersalyl (allowed to react with mitochondria for at least 20 s) on the initial rate of Pi exchange supports earlier observations on the protective effect of this inhibitor; i.e., up to 3 nmol of mersalyl/mg of protein does not decrease the transport rate whereas these low concentrations protect approximately 50% of the transport capacity from irreversible inactivation by N-ethylmaleimide. In nonrespiring mitochondria, at pH 7.3, Pi exchange exhibited a Km of 1.6 mM and a Vmax of 3.0 mumol min-1 (mg of mitochondrial protein)-1. The increase of the membrane potential without any concomitant change of delta pH had no significant influence on the kinetic parameters. The maximal velocity of Pi transport is significantly higher than the maximal velocity of all the other components of oxidative phosphorylation at comparable temperatures. The possible physiological significance of this excess capacity is discussed.  相似文献   

20.
We have examined the hydrogen exchange properties of bovine insulin under solution conditions that cause it to aggregate and eventually form amyloid fibrils. The results have been obtained at the residue-specific level using peptic digestion and mass spectrometry. A total of 19 peptides were assigned to regions of the protein and their exchange properties monitored for a period of 24 hours. The results of the peptic digestion show that residues A13 to A21 and B11 to B30 are more susceptible to proteolysis than the N-terminal regions of the protein. A total of 15 slowly exchanging amides were observed for insulin under these solution conditions. Location of the protected amides was carried out using a peptic-digestion protocol at low pH. Chromatographic separation was not required. This enabled a direct comparison of the peptides within the same mass spectrum. From kinetic analysis of the rates slow exchange has been located to 4(+/-1) backbone amides in the A13-A19 helix and 6(+/-1) in the B chain helix. The remaining 5(+/-1) are assigned to helix A2-A8. Taken together the results from digestion and hydrogen exchange show that at low pH and relatively high concentrations the C termini of both chains are susceptible to proteolysis but that the solution structure contains the native state helices. More generally the results demonstrate that mass spectrometry can be applied to study site-specific hydrogen exchange properties of proteins even under conditions where they are known to be partially folded and aggregate extensively in solution.  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