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 281 毫秒
1.
The pH dependence of hydrogen exchange in proteins   总被引:3,自引:0,他引:3  
The static accessibility modified discrete charge model for electrostatic interactions in proteins is extended to the prediction of the pH dependence of hydrogen exchange reactions. The exchange rate profiles of buried amide protons are shown to follow the calculated pH dependence of the electrostatic component of protein stability. Rate profiles are calculated for individual buried amide protons in ribonuclease S and bovine pancreatic trypsin inhibitor. The electrostatic free energy of stabilization of the protein and the energy required to bring the catalytic ion to an exchange site are expressed as an apparent, pH-dependent contribution to the activation energy. Changes in the electrostatic stabilization of the proteins affect the calculated exchange rate for buried amide protons by more than 1000, while local field effects raise or lower the predicted exchange rates by less than 100. The pH dependence of exchangeable protons at the protein surface, such as the C-2 imidazole protons, is shown to follow the estimated energy required to introduce the catalytic ion at the exchange site. These calculations are discussed in terms of current models for proton exchange which incorporate the dynamic nature of the structure to explain exchange data from the interior of a protein.  相似文献   

2.
Possible effects of changes in net charge on protein hydrogen exchange rates were investigated by desalting hen egg-white lysozyme, which allowed its net charge to increase with decreasing pH in the acid region. Chloride ion-binding ratios, expressed as ratios of free to total Cl?, were measured with a chloride-specific electrode at pH 5 on a 2.4% solution of a five-time-desalted product. This ratio was used to show a 97% reduction of the 11% Cl? present in a commercial lysozyme preparation upon three passes of the enzyme through a column of ion-retardation resin. Net charges on the purified product were assigned from a combination of electrophoretic mobility and proton titration data gathered under minimal ionic strength conditions. The net charge on the desalted product increased by 1.64 units between pH 5.0 and 3.0. Hydrogendeuterium exchange studies on the purified lysozyme in D2O were obtained using the near-infrared region of a Cary 14R spectrophotometer. The rate-pD profile for k2, the rate constant for the intermediate class of exchanging hydrogens, showed a decrease in the apparent pD of minimum exchange rate of 0.3 units, when compared to that obtained earlier in 0.2 m added NaCl. However, the rate of exchange at pD minimum and the number of hydrogens in the class remained largely unaffected. A similar shift was observed for the rate-pD profile of the class 1 hydrogens. Thus, the effect of an increase in net positive charge is to shift the rate-pD profile to a lower pD. Moreover, the effect extended to the interior peptide hydrogens of this globular protein. Consequently, the exchange rates of all the observable hydrogens are altered by the net charge changes, and the effect appeared uniform. The shift can be accounted for quantitatively by applying electrostatic interaction terms to the acid and base catalytic constants characterizing the exchange process. The calculated electrostatic interaction factors in minimal salt and 0.2 m added NaCl were found to be 29 and 18% lower, respectively, than those obtained theoretically. Therefore, under conditions where changes in net charge may occur for a globular protein, the effect on hydrogen exchange rates can be estimated fairly well theoretically, especially at moderate ionic strengths.  相似文献   

3.
To investigate the determinants of protein hydrogen exchange (HX), HX rates of most of the backbone amide hydrogens of Staphylococcal nuclease were measured by NMR methods. A modified analysis was used to improve accuracy for the faster hydrogens. HX rates of both near surface and well buried hydrogens are spread over more than 7 orders of magnitude. These results were compared with previous hypotheses for HX rate determination. Contrary to a common assumption, proximity to the surface of the native protein does not usually produce fast exchange. The slow HX rates for unprotected surface hydrogens are not well explained by local electrostatic field. The ability of buried hydrogens to exchange is not explained by a solvent penetration mechanism. The exchange rates of structurally protected hydrogens are not well predicted by algorithms that depend only on local interactions or only on transient unfolding reactions. These observations identify some of the present difficulties of HX rate prediction and suggest the need for returning to a detailed hydrogen by hydrogen analysis to examine the bases of structure-rate relationships, as described in the companion paper (Skinner et al., Protein Sci 2012;21:996-1005).  相似文献   

4.
A new method based on protein fragmentation and directly coupled microbore high-performance liquid chromatography-fast atom bombardment mass spectrometry (HPLC-FABMS) is described for determining the rates at which peptide amide hydrogens in proteins undergo isotopic exchange. Horse heart cytochrome c was incubated in D2O as a function of time and temperature to effect isotopic exchange, transferred into slow exchange conditions (pH 2-3, 0 degrees C), and fragmented with pepsin. The number of peptide amide deuterons present in the proteolytic peptides was deduced from their molecular weights, which were determined following analysis of the digest by HPLC-FABMS. The present results demonstrate that the exchange rates of amide hydrogens in cytochrome c range from very rapid (k > 140 h-1) to very slow (k < 0.002 h-1). The deuterium content of specific segments of the protein was determined as a function of incubation temperature and used to indicate participation of these segments in conformational changes associated with heating of cytochrome c. For the present HPLC-FABMS system, approximately 5 nmol of protein were used for each determination. Results of this investigation indicate that the combination of protein fragmentation and HPLC-FABMS is relatively free of constraints associated with other analytical methods used for this purpose and may be a general method for determining hydrogen exchange rates in specific segments of proteins.  相似文献   

5.
Amide exchange rates were measured for Pyrococcus furiosus (Pf) rubredoxin substituted with either Zn(II), Ga(III), or Ge(IV). Base-catalyzed exchange rate constants increase up to 3000-fold per unit charge for the highly protected amides surrounding the active site metal, yielding apparent residue-specific conformational energy decreases of more than 8 kcal/mol in a comparison of the Zn(II)- and Ge(IV)-substituted proteins. However, the exchange kinetics for many of the other amides of the protein are insensitive to these metal substitutions. These differential rates are inversely correlated with the distance between the amide nitrogen and the metal in the X-ray structure, out to a distance of at least 12 A, consistent with an electrostatic potential-dependent shifting of the amide nitrogen pK. This strongly correlated distance dependence is consistent with a nativelike structure for the exchange-competent conformations. The electric field potential within the interior of the rubredoxin structure gives rise to a change of as much as a million-fold in the rate for the exchange-competent state of the individual amide hydrogens. Nevertheless, the strength of these electrostatic interactions in Pf rubredoxin appears to be comparable to those previously reported within other proteins. As a result, contrary to the conventional analysis of hydrogen exchange data, for exchange processes that occur via nonglobal transitions, the residual conformational structure will often modulate the observed rates. Although this necessarily complicates the estimation of the conformational equilibria of these exchange-competent states, this dependence on residual structure can provide insight into the conformation of these transient states.  相似文献   

6.
Mass spectrometry is used to probe the kinetics of hydrogen–deuterium exchange in lysozyme in pH 5, 6 and 7.4. An analysis based on a Verhulst growth model is proposed and effectively applied to the kinetics of the hydrogen exchange. The data are described by a power-like function which is based on a time-dependence of the exchange rate. Experimental data ranging over many time scales is considered and accurate fits of a power-like function are obtained. Results of fittings show correlation between faster hydrogen–deuterium exchange and increase of pH. Furthermore a model is presented that discriminates between easily exchangeable hydrogens (located in close proximity to the protein surface) and those protected from the exchange (located in the protein interior). A possible interpretation of the model and its biological significance are discussed.  相似文献   

7.
T Endo  T Ueda  H Yamada  T Imoto 《Biochemistry》1987,26(7):1838-1845
Nuclear magnetic resonance analyses have been made of the individual hydrogen-deuterium exchange rates of tryptophan indole N-1 hydrogens in native lysozyme and its chemically modified derivatives including lysozyme with an ester cross-linkage between Glu-35 and Trp-108, lysozyme with an internal amide cross-linking between the epsilon-amino group of Lys-13 and the alpha-carboxyl group of Leu-129, and lysozyme with the beta-aspartyl sequence at Asp-101. The pH dependence curves of the exchange rates for Trp-63 and Trp-108 are different from those expected for tryptophan. The pH dependence curve for Trp-108 exchange exhibits the effects from molecular aggregation at pH above 5 and from a transition between the two conformational fluctuations at around pH 4. The exchange rates for tryptophan residues in native lysozyme and modified derivatives are not correlated with the thermodynamic or kinetic parameters in protein denaturation, suggesting that the fluctuations responsible for the exchange are not global ones. The exchange rates for tryptophan residues remote from the modification site are perturbed. Such tryptophan residues are found to be involved in a small but distinct conformational change due to the modification. Therefore, the perturbations of the N-1 hydrogen exchange rates are related to the minor change in local conformation or in conformational strain induced by the chemical modification.  相似文献   

8.
The acid-catalyzed hydrogen exchange rate constants kH, and the base-catalyzed rate constants kOH, have been determined (in the preceding paper) for the 25 most rapidly exchanging NH groups of bovine pancreatic trypsin inhibitor. Most of these NH groups are at the protein-solvent interface. The correlation of kH, but not kOH, with the static accessibility and hydrogen bonding of the peptide carbonyl O atom indicates that the mechanism of acid catalysis in proteins involves O-protonation. Agreement between the ionic strength dependence observed for kH and kOH and the ionic strength dependence calculated for an O-protonation mechanism supports this conclusion. N-protonation for acid catalysis, as well as N-deprotonation for base catalysis, have traditionally been assumed in the mechanism of the chemical step in peptide amide proton exchange. A preference for the alternative O-protonation mechanism has far-reaching implications in the interpretation of protein hydrogen exchange kinetics. With an O-protonation mechanism, acid-catalyzed rates of surface NH groups are primarily a function of the average solvent accessibility of the carbonyl O atoms in the dynamic solution structure, while base-catalyzed rates of surface NH groups measure solvent accessibility of the peptide N. The relative dynamic accessibilities of peptide O atoms, as measured by relative values of kH (corrected for electrostatic effects), correlate with O static accessibilities in the crystal structure. A lower correlation of static accessibility of N atoms with kOH is observed for surface NH groups in peptide groups in which the carbonyl O is not hydrogen bonded. For some surface NH groups, the observed pH of minimum rate, pHmin, deviates widely from the pHmin of model compounds. This is explained as the combined result of electrostatic effects and of the differences in accessibility of the carbonyl O and N atoms that result in a change in the relative values of kH and kOH as compared to those of model peptides. A mechanism whereby exchange of interior sites is catalyzed by interactions of catalysis ions with protein surface atoms via charge transfer is suggested.  相似文献   

9.
The hydrogen exchange behavior of rhodopsin was re-examined by studies of the protein in the disc membrane and after solubilization in octyl glucoside. The methods used measure either the peptide hydrogens alone (hydrogen-deuterium exchange by infrared spectroscopy) or all slowly exchanging hydrogens (hydrogen-tritium exchange by hel filtration). Under mild exchange conditions, disc membranes and solubilized lipid-free proteins show very similar exchange behavior, indicating the absence of slowly exchanging lipid protons. At high temperature, exchange of an additional large group of very slow peptide NH can be detected. The total number of slow hydrogens significantly exceeds the amide content, and apparently includes slowly exchanging protons from perhaps 40% of the protein's non-amide side chains. This is thought to require the involvement of many polar side chains in internal H-bonding. The exchange rates of the non-amide side chains sites have not been determined. However, to the extent that these contribute to the fast time region of the measured kinetic H-exchange curve, previously identified with exposed, non-H-bonded peptides, the estimate of freely exposed rhodopsin peptides must be reduced. The fraction of free peptides could range from a remarkably high value of 70% down to about 45%.  相似文献   

10.
Src-homology-2 domains are small, 100 amino acid protein modules that are present in a number of signal transduction proteins. Previous NMR studies of SH2 domain dynamics indicate that peptide binding decreases protein motions in the pico- to nanosecond, and perhaps slower, time range. We suggest that amide hydrogen exchange and mass spectrometry may be useful for detecting changes in protein dynamics because hydrogen exchange rates are relatively insensitive to the time domains of the dynamics. In the present study, hydrogen exchange and mass spectrometry were used to probe hematopoietic cell kinase SH2 that was either free or bound to a 12-residue high-affinity peptide. Hydrogen exchange rates were determined by exposing free and bound SH2 to D(2)O, fragmenting the SH2 with pepsin, and determining the deuterium level in the peptic fragments. Binding generally decreased hydrogen exchange along much of the SH2 backbone, indicating a widespread reduction in dynamics. Alterations in the exchange of the most rapidly exchanging amide hydrogens, which was detected following acid quench and analysis by mass spectrometry, were used to locate differences in low-amplitude motion when SH2 was bound to the peptide. In addition, the results indicate that hydrogen exchange from the folded form of SH2 is an important process along the entire SH2 backbone.  相似文献   

11.
A medium resolution hydrogen exchange method (Rosa & Richards, 1979) has been used to measure the average rates of amide hydrogen exchange for known segments of the S-protein portion of ribonuclease-S. The analytical procedure permitted exchange rates to be monitored for seven S-protein fragments distributed throughout the structure, including regions of α-helix and β-sheet. Kinetics were measured as a function of pH, temperature and S-peptide binding.The pH dependence of exchange from isolated S-protein between pH 2·8 and pH 7·0 was found to deviate significantly from a first-order dependence on hydroxide ion concentration. The protection against exchange with increasing pH appeared to be closely related to the electrostatic stabilization of S-protein. It is suggested that such favorable electrostatic interactions result in increased energy barriers to the conformational fluctuations that provide solvent access to the time-average crystallographic structure. This explanation of the observed correlation between stability and exchange kinetics is also consistent with the calculated apparent activation energies for exchange from S-protein between 5·5 and 20 °C.S-peptide binding dramatically slows exchange from many S-protein sites, even those distant from the area of S-peptide contact. Interestingly, the effects of complex formation are not evenly propagated throughout S-protein. The most significantly perturbed sites (≥103-fold reduction in exchange rate constants) lie within fragments derived from regions of secondary structure. Exchange from several other fragments is not significantly affected. The S-peptide—S-protein dissociation constant at neutral pH is so small that the measured exchange must have occurred from the complex and not from the dissociated parts.  相似文献   

12.
Amide hydrogen/deuterium exchange rates were measured as a function of pH and urea for 37 slowly exchanging amides in the beta-trefoil protein hisactophilin. The rank order of exchange rates is generally maintained under different solution conditions, and trends in the pH and urea dependence of exchange rates are correlated with the rank order of exchange rates. The observed trends are consistent with the expected behavior for exchange of different amides via global and/or local unfolding. Analysis of the pH dependence of exchange in terms of rate constants for structural opening and closing reveals a wide range of rates in different parts of the hisactophilin structure. The slowest exchanging amides have the slowest opening and closing rates. Many of the slowest exchanging amides are located in trefoil 2, but there are also some slow exchanging amides in trefoils 1 and 3. Slow exchangers tend to be near the interface between the beta-barrel and the beta-hairpin triplet portions of this single-domain structure. The pattern of exchange behaviour in hisactophilin is similar to that observed previously in interleukin-1 beta, indicating that exchange properties may be conserved among beta-trefoil proteins. Comparisons of opening and closing rates in hisactophilin with rates obtained for other proteins reveal clear trends for opening rates; however, trends in closing rates are less apparent, perhaps due to inaccuracies in the values used for intrinsic exchange rates in the data fitting. On the basis of the pH and urea dependence of exchange rates and optical measurements of stability and folding, EX2 is the main exchange mechanism in hisactophilin, but there is also evidence for varying levels of EX1 exchange at low and high pH and high urea concentrations. Equilibrium intermediates in which subglobal portions of structure are cooperatively disrupted are not apparent from analysis of the urea dependence of exchange rates. There is, however, a strong correlation between the Gibbs free energy of opening and the denaturant dependence of opening for all amides, which suggests exchange from a continuum of states with different levels of structure. Intermediates are not very prominent either in equilibrium exchange experiments or in quenched-flow kinetic studies; hence, hisactophilin may not form partially folded states as readily as IL-1 beta and other beta-trefoil proteins.  相似文献   

13.
Uchida K  Markley JL  Kainosho M 《Biochemistry》2005,44(35):11811-11820
A novel method for monitoring proton-deuteron (H/D) exchange at backbone amides is based on the observation of H/D isotope effects on the (13)C NMR signals from peptide carbonyls. The line shape of the carbonyl (13)C(i) signal is influenced by differential H/D occupancy at the two adjacent amides: the H(N)(i)(+1) (beta site) and the H(N)(i) (gamma site). At a carbon frequency of 75.4 MHz, the H --> D isotope shifts on the (13)C signal are about 5-7 Hz for exchange at the beta site and 2 Hz or less for exchange at the gamma site. Because the effects at the two sites are additive, the time dependence of the line shape of a particular carbonyl resonance can report not only the exchange rates at the individual sites but also the level of dual exchange. Therefore, the data can be analyzed to determine the rate (k(c)) and degree of correlated exchange (X(betagamma)) at the two sites. We have applied this approach to the investigation of the pH dependence of hydrogen exchange at several adjacent residues in Streptomyces subtilisin inhibitor (SSI). Two selectively labeled SSI proteins were produced: one with selective (13)C' labeling at all valyl residues and one with selective (13)C' labeling at all leucyl residues. This permitted the direct observation by one-dimensional (13)C NMR of selected carbonyl signals from residues with slowly exchanging amides at the i and i + 1 positions. The residues investigated were located in an alpha helix and in a five-stranded antiparallel beta sheet. Samples of the two labeled proteins were prepared at various pH values, and (13)C NMR spectra were collected at 50 degrees C prior to and at various times after transferring the sample from H(2)O to (2)H(2)O. Most of the slowly exchanging amides studied were intramolecular hydrogen-bond donors. In agreement with prior studies, the results indicated that the exchange rates of the amide hydrogens in proteins are governed not only by hydrogen bonding but also by other factors. For example, the amide hydrogen of Thr34 exchanges rapidly even though it is an intramolecular hydrogen-bond donor. Over nearly the whole pH range studied, the apparent rates of uncorrelated exchange (k(beta) and k(gamma)) were proportional to [OH(-)] and the apparent rates of correlated exchange at two adjacent sites (k(c)) were roughly proportional to [OH(-)](2). This enabled us to extract the pH-independent exchange rates (k(beta) degrees , k(gamma) degrees , and k(c) degrees ). In all cases in which correlated exchange could be measured, the observed sigmoidal pH dependence of X(betagamma) could be replicated roughly from the derived pH-independent rates.  相似文献   

14.
The rate of exchange of the labile hydrogens of lysozyme was measured by out-exchange of tritium from the protein in solution and from powder samples of varied hydration level, for pH 2, 3, 5, 7, and 10 at 25 degrees C. The dependence of exchange of powder samples on the level of hydration was the same for all pHs. Exchange increased strongly with increased hydration until reaching a rate of exchange that is constant above 0.15 g of H2O/g of protein (120 mol of H2O/mol of protein). This hydration level corresponds to coverage of less than half the protein surface with a monolayer of water. No additional hydrogen exchange was observed for protein powders with higher water content. Considered in conjunction with other lysozyme hydration data [Rupley, J. A., Gratton, E., & Careri, G. (1983) Trends Biochem. Sci. (Pers. Ed.) 8, 18-22], this observation indicates that internal protein dynamics are not strongly coupled to surface properties. The use of powder samples offers control of water activity through regulation of water vapor pressure. The dependence of the exchange rate on water activity was about fourth order. The order was pH independent and was constant from 114 to 8 mol of hydrogen remaining unexchanged/mol of lysozyme. These results indicate that the rate-determining step for protein hydrogen exchange is similar for all backbone amides and involves few water molecules. Powder samples were hydrated either by isopiestic equilibration, with a half-time for hydration of about 1 h, or by addition of solvent to rapidly reach final hydration. Samples hydrated slowly by isopiestic equilibration exhibited more exchange than was observed for samples of the same water content that had been hydrated rapidly by solvent addition. This difference can be explained by salt and pH effects on the nearly dry protein. Such effects would be expected to contribute more strongly during the isopiestic equilibration process. Solution hydrogen exchange measurements made for comparison with the powder measurements are in good agreement with published data. Rank order was proven the same for all pHs by solution pH jump experiments. The effect of ionic strength on hydrogen exchange was examined at pH 2 and pH 5 for protein solutions containing up to 1.0 M added salt. The influence of ionic strength was similar for both pHs and was complex in that the rate increased, but not monotonically, with increased ionic strength.  相似文献   

15.
The exchange of a large number of amide hydrogens in oxidized equine cytochrome c was measured by NMR and compared with structural parameters. Hydrogens known to exchange through local structural fluctuations and through larger unfolding reactions were separately considered. All hydrogens protected from exchange by factors greater than 10(3) are in defined H-bonds, and almost all H-bonded hydrogens including those at the protein surface were measured to exchange slowly. H-exchange rates do not correlate with H-bond strength (length) or crystallographic B factors. It appears that the transient structural fluctuation necessary to bring an exchangeable hydrogen into H-bonding contact with the H-exchange catalyst (OH(-)-ion) involves a fairly large separation of the H-bond donor and acceptor, several angstroms at least, and therefore depends on the relative resistance to distortion of immediately neighboring structure. Accordingly, H-exchange by way of local fluctuational pathways tends to be very slow for hydrogens that are neighbored by tightly anchored structure and for hydrogens that are well buried. The slowing of buried hydrogens may also reflect the need for additional motions that allow solvent access once the protecting H-bond is separated, although it is noteworthy that burial in a protein like cytochrome c does not exceed 4 angstroms. When local fluctuational pathways are very slow, exchange can become dominated by a different category of larger, cooperative, segmental unfolding reactions reaching up to global unfolding.  相似文献   

16.
A procedure to measure exchange rates of fast exchanging protein amide hydrogens by time-resolved NMR spectroscopy following in situ initiation of the reaction by diluting a native protein solution into an exchanging deuterated buffer is described. The method has been used to measure exchange rates of a small set of amide hydrogens of reduced cytochrome c, maintained in a strictly anaerobic atmosphere, in the presence of an otherwise inaccessible range of guanidinium deuterochloride concentrations. The results for the measured protons indicate that hydrogen exchange in the unfolding transition region of cytochrome c reach the EX2 limit, but emphasize the difficulty in interpretation of the exchange mechanism in protein hydrogen exchange studies. Comparison of free energies of structure opening for the measured hydrogens with the global unfolding free energy monitored by far-UV CD measurements has indicated the presence of at least one partially unfolded equilibrium species of reduced cytochrome c. The results provide the first report of measurement of free energy of opening of structure to exchange in the 0–2-kcal/mol range. Proteins 32:241–247, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

17.
D G Cross  A Brown  H F Fisher 《Biochemistry》1975,14(12):2745-2749
The pH dependence of the apparent first-order rate constant for the exchange of the exocyclic amino hydrogens of adenosine with deuterium from the solvent was measured by stopped-flow ultraviolet spectroscopy. This dependence shows acid catalysis, base catalysis, and spontaneous exchange at neutral pH values. A study of the effect of several buffers on the rates of exchange reveals both general acid and general base catalytic behavior for the exchange process. We propose a general mechanism for the exchange which requires N-1 protonated adenosine as an intermediate for the acid-catalyzed exchange and amidine anion for the base-catalyzed exchange. In both cases the rate-limiting step is the base-catalyzed abstraction of a proton from the exocyclic amino moiety. Evaluation of the rate constants predicts the equilibrium for the exocyclic amino/imino tautomers to be 6.3 times 10(3):1.  相似文献   

18.
Equilibrium amide hydrogen exchange studies of barstar have been carried out at pH 6.7, 32° SDC using one- and two-dimensional nuclear magnetic resonance. An unusually large fraction of the backbone amide hydrogens of barstar exchange too fast to be measured, and the exchange rates of only fifteen slow-exchanging amide sites including indole amides of two tryptophans could be measured in the presence of 0 to 1.8 M guanidine hydrochloride (GdnHCl). Measurement of exchange occurring in tens of seconds in the unfolding transition region was possible by the use of a fast stopped-flow mixing method. The observed exchange rates have been simulated in the EX2 limit according to a two-process model that incorporates two exchange-competent states: a transiently unfolded state (U*) in which many amide hydrogens are completely accessible to solvent-exchange, and a near-native locally unfolded state (N*), in which only one or a few amide hydrogens are completely accessible to solvent-exchange. The two-process model appears to account for the observed exchange behavior over the entire range of GdnHCl concentrations studied. For several measurable slow-exchanging amide hydrogens, the free energies of production of exchange-competent states from the exchange-incompetent native state are significantly higher than the free-energy of production of the equilibrium unfolded state from the native state, when the latter is determined from circular dichroism- or fluorescence-monitored equilibrium unfolding curves. The result implies that U*, which forms transiently in the strongly native-like conditions used for the hydrogen exchange studies, is higher in energy than the equilibrium-unfolded state. The higher energy of this transiently unfolded exchange-competent state can be attributed to either proline isomerization or to the presence of residual structure. On the basis of the free energies of production of exchange-competent states, the measured amide sites of barstar appear to define two structural subdomains—a three-helix unit and a two-β-strand unit in the core of the protein. Proteins 30:295–308, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

19.
The exchange kinetics for the slowly exchanging amide hydrogens in three defensins, rabbit NP-2, rabbit NP-5, and human HNP-1, have been measured over a range of pH at 25°C using 1D and 2D NMR methods. These NHs have exchange rates 102 to 105 times slower than rates from unstructured model peptides. The observed distribution of exchange rates under these conditions can be rationalized by intramolecular hydrogen bonding of the individual NHs, solvent accessibility of the NHs, and local fluctuations in structure. The temperature dependencies of NH chemical shifts (NH temperature coefficients) were measured for the defensins and these values are consistent with the defensin structure. A comparison is made between NH exchange kinetics, NH solvent accessibility, and NH temperature coefficients of the defensins and other globular proteins. Titration of the histidine side chain in NP-2 was examined and the results are mapped to the three-dimensional structure. © 1994 Wiley-Liss, Inc.  相似文献   

20.
We have measured the tritium outexchange of subtilisin BPN'. A consistent and rather small group of hydrogens was isolated by their sensitivity to inhibitor binding. The viscosity dependence of exchange from these inhibitor protected hydrogens was then examined in 0.05 M MES buffer, pH 6.5 and 10 degrees C. The viscosity of the reaction medium was varied by added glycerol and ethylene glycol. The exchange rates were corrected to be compared at identical hydroxyl ion and water activity. The salient observation is the strikingly similar viscosity coupling behavior when compared to the deacylation step of ester hydrolysis catalyzed by the same enzyme (Ng and Rosenberg, Biophysical Chemistry, 39 (1991) 57). We have obtained a viscosity coupling constant of 0.68 -/+ 0.18 for hydrogen exchange in glycerol (cf. 0.65 -/+ 0.11 for deacylation in glycerol, sucrose, glucose and fructose); 1.67 -/+ 0.07 for outexchange (cf. 1.92 -/+ 0.09 for deacylation), in the presence of ethylene glycol. The two reactions are very chemically dissimilar, yet they show very similar viscosity coupling behavior. This together with the well established role of structural fluctuations in hydrogen exchange implies a similar role of structural fluctuations in the deacylation step of subtilisin BPN' catalyzed ester hydrolysis.  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