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 15 毫秒
1.
Sac7d and Sso7d are homologous, hyperthermophile proteins with a high density of charged surface residues and potential ion pairs. To determine the relative importance of specific amino acid side-chains in defining the stability and function of these Archaeal chromatin proteins, pK(a) values were measured for the acidic residues in both proteins using (13)C NMR chemical shifts. The stability of Sso7d enabled titrations to pH 1 under low-salt conditions. Two aspartate residues in Sso7d (D16 and D35) and a single glutamate residue (G54) showed significantly perturbed pK(a) values in low salt, indicating that the observed pH-dependence of stability was primarily due to these three residues. The pH-dependence of backbone amide NMR resonances demonstrated that perturbation of all three pK(a) values was primarily the result of side-chain to backbone amide hydrogen bonds. Few of the significantly perturbed acidic pK(a) values in Sac7d and Sso7d could be attributed to primarily ion pair or electrostatic interactions. A smaller perturbation of E48 (E47 in Sac7d) was ascribed to an ion pair interaction that may be important in defining the DNA binding surface. The small number (three) of significantly altered pK(a) values was in good agreement with a linkage analysis of the temperature, pH, and salt-dependence of folding. The linkage of the ionization of two or more side-chains to protein folding led to apparent cooperativity in the pH-dependence of folding, although each group titrated independently with a Hill coefficient near unity. These results demonstrate that the acid pH-dependence of protein stability in these hyperthermophile proteins is due to independent titration of acidic residues with pK(a) values perturbed primarily by hydrogen bonding of the side-chain to the backbone. This work demonstrates the need for caution in using structural data alone to argue the importance of ion pairs in stabilizing hyperthermophile proteins.  相似文献   

2.
3.
Experimentally determined mean pK(a) values of carboxyl residues located at the N-termini of alpha-helices are lower than their overall mean values. Here, we perform three types of analyses to account for this phenomenon. We estimate the magnitude of the helix macrodipole to determine its potential role in lowering carboxyl pK(a) values at the N-termini. No correlation between the magnitude of the macrodipole and the pK(a) values is observed. Using the pK(a) program propKa we compare the molecular surroundings of 18 N-termini carboxyl residues versus 233 protein carboxyl groups from a previously studied database. Although pK(a) lowering interactions at the N-termini are similar in nature to those encountered in other protein regions, pK(a) lowering backbone and side-chain hydrogen bonds appear in greater number at the N-termini. For both Asp and Glu, there are about 0.5 more hydrogen bonds per residue at the N-termini than in other protein regions, which can be used to explain their lower than average pK(a) values. Using a QM-based pK(a) prediction model, we investigate the chemical environment of the two lowest Asp and the two lowest Glu pK(a) values at the N-termini so as to quantify the effect of various pK(a) determinants. We show that local interactions suffice to account for the acidity of carboxyl residues at the N-termini. The effect of the helix dipole on carboxyl pK(a) values, if any, is marginal. Backbone amide hydrogen bonds constitute the single biggest contributor to the lowest carboxyl pK(a) values at the N-termini. Their estimated pK(a) lowering effects range from about 1.0 to 1.9 pK(a) units.  相似文献   

4.
Previous studies of ubiquitin disclosed numerous charge-charge interactions on the protein's surface. To investigate how neighboring residues influence the strength of these interactions, double-mutant cycles are combined with pK(a) determinations by 2D NMR. More specifically, the environment around the Asp21-Lys29 ion pair has been altered through mutations at position 25, which is an asparagine in mammalian ubiquitin and a positively-charged residue in many other ubiquitin-like proteins. The pK(a) value of Asp21 decreases by 0.4 to 0.7 pH unit when Asn25 is substituted with a positively charged residue, suggesting a new and favorable ion pair interaction between positions 21 and 25. However, analysis of double mutants reveals that the favorable interaction between Asp21 and Lys29 is weakened when position 25 is a positively charged residue. Interestingly, while the pK(a) value of His25 in the N25H variant agrees with model compound values, additional mutants reveal that this agreement is fortuitous, resulting from a balance of favorable and unfavorable interactions; similar results were observed previously for Glu34 in ubiquitin and His8 in staphylococcal nuclease. Ionizable groups may thus have pK(a) values similar to model compound values and yet still be involved in significant interactions with other protein groups. One surprising result of introducing positively charged residues at position 25 is a new interaction between Lys29 and Glu18, an interaction not present in wild-type ubiquitin. This unanticipated result illustrates a key advantage of using NMR to determine pK(a) values for many residues simultaneously in the variant proteins. Overall, the strength of an interaction between two residues at the surface of ubiquitin is sensitive to the identity of neighboring residues. The results also demonstrate that relatively conservative and common point mutations such as substitutions of polar with charged residues and vice versa can have effects on interactions beyond the site of mutation per se.  相似文献   

5.
Previously we reported that Lys, Asp, and Glu residues at positions 66 and 92 in staphylococcal nuclease (SNase) titrate with pK(a) values shifted by up to 5 pK(a) units in the direction that promotes the neutral state. In contrast, the internal Lys-38 in SNase titrates with a normal pK(a). The crystal structure of the L38K variant shows that the side chain of Lys-38 is buried. The ionizable moiety is approximately 7 A from solvent and ion paired with Glu-122. This suggests that the pK(a) value of Lys-38 is normal because the energetic penalty for dehydration is offset by a favorable Coulomb interaction. However, the pK(a) of Lys-38 was also normal when Glu-122 was replaced with Gln or with Ala. Continuum electrostatics calculations were unable to reproduce the pK(a) of Lys-38 unless the protein was treated with an artificially high dielectric constant, consistent with structural reorganization being responsible for the normal pK(a) value of Lys-38. This reorganization must be local because circular dichroism and NMR spectroscopy indicate that the L38K protein is native-like under all conditions studied. In molecular dynamics simulations, the ion pair between Lys-38 and Glu-122 is unstable. The simulations show that a minor rearrangement of a loop is sufficient to allow penetration of water to the amino moiety of Lys-38. This illustrates both the important roles of local flexibility and water penetration as determinants of pK(a) values of ionizable groups buried near the protein-water interface, and the challenges faced by structure-based pK(a) calculations in reproducing these effects.  相似文献   

6.
To define adequately enzyme activation/inhibition mechanisms as a function of pH, it is necessary to characterize the effector-induced pK shifts on both the free enzyme and on the enzyme-substrate complex. On the basis of our recent three-protons model for sucrase [Vasseur, van Melle, Frangne & Alvarado (1988) Biochem. J. 251, 667-675], we show how the 'fundamental' pK values, deduced from the classical double-logarithmic transformations, are insufficient to generate the required information. This insufficiency derives from the fact that, for sucrase, the acid ionization constant, K1, is a molecular constant that involves complex, V-type plus K-type, activatory and inhibitory kinetic effects. As a consequence, substrate-induced pK shifts cannot be interpreted correctly only by using the fundamental pK approach, because an unequal number of key protons is involved, depending on whether the free enzyme or the enzyme-substrate complex is considered. We demonstrate how this problem can be solved by using the 'theoretical' pK values, derived from the reciprocals of the Michaelis pH functions, i.e. Cha's fractional concentration factors. The procedure we propose, which is general, has the advantage of yielding all the macroscopic pK values for any given model, as calculated from the microscopic pK values. Furthermore, it permits predicting pK shifts as a function of [S] and/or [A] (where S is the substrate and A is the allosteric modifier), an objective that cannot be attained by using the double-logarithmic plot approach. Finally, we describe the relation existing between the fundamental and the theoretical pK values.  相似文献   

7.
We developed a Rosetta-based Monte Carlo method to calculate the pK(a) values of protein residues that commonly exhibit variable protonation states (Asp, Glu, Lys, His, and Tyr). We tested the technique by calculating pK(a) values for 264 residues from 34 proteins. The standard Rosetta score function, which is independent of any environmental conditions, failed to capture pK(a) shifts. After incorporating a Coulomb electrostatic potential and optimizing the solvation reference energies for pK(a) calculations, we employed a method that allowed side-chain flexibility and achieved a root mean-square deviation (RMSD) of 0.83 from experimental values (0.68 after discounting 11 predictions with an error over 2 pH units). Additional degrees of side-chain conformational freedom for the proximal residues facilitated the capture of charge-charge interactions in a few cases, resulting in an overall RMSD of 0.85 pH units. The addition of backbone flexibility increased the overall RMSD to 0.93 pH units but improved relative pK(a) predictions for proximal catalytic residues. The method also captures large pK(a) shifts of lysine and some glutamate point mutations in staphylococcal nuclease. Thus, a simple and fast method based on the Rosetta score function and limited conformational sampling produces pK(a) values that will be useful when rapid estimation is essential, such as in docking, design, and folding.  相似文献   

8.
The oxidation-reduction midpoint potentials, Em, of the FAD and active site disulfide couples of Escherichia coli thioredoxin reductase have been determined from pH 5.5 to 8.5. The FAD and disulfide couples have similar Em values and thus a linked equilibrium of four microscopic enzyme oxidation-reduction states exists. The binding of phenylmercuric acetate to one enzyme form could be monitored which allowed solving the four microscopic Em values. The Em values at pH 7.0 and 12 degrees C of the four couples of thioredoxin reductase are: (S)2-enzyme-FAD/FADH2 = -0.243 V, (SH)2-enzyme-FAD/FADH2 = -0.260 V, (FAD)-enzyme-(S)2/(SH)2 = -0.254 V, and (FADH2)-enzyme-(S)2/(SH)2 = -0.271 V. Thus, at pH 7.0, the FAD and disulfide moieties have a 0.017-V negative interaction and Em values which are different by 0.011 V. The delta Em/delta pH of the FAD couples E2m and E3m are about 0.060 V/pH throughout the pH range studied, showing an approximately 2-proton stoichiometry of reduction of the enzyme FAD. The delta Em/delta pH of the disulfide couples E1m and E4m are about 0.052 V/pH from pH 5.5 to 8.5, showing an apparently nonintegral proton stoichiometry of reduction of 1.8 in this pH range. This proton stoichiometry suggests the presence of a base with an ionization behavior that is linked to the oxidation-reduction state of the disulfide. A novel method is presented for determining the pK values on oxidized and reduced enzyme which agrees with the less accurate classical method. The proton stoichiometry results are consistent with the presence of a thiol-base ion pair in which the pK of the base is elevated from 7.6 in disulfide containing enzyme to greater than 8.5 upon forming an ion pair with a thiol anion of pK 7.0 generated upon reduction of the disulfide. The fluorescence of the FAD in thioredoxin reductase decreases as the pH is lowered with a pK of 7.0, direct evidence for a base near the FAD probably distinct from the base interacting with the dithiol.  相似文献   

9.
Using the basic ionization constants for a pair of isomers, m1ApA and Apm1A, and the measured values for the overlapping pK values of ApA, the microscopic ionization constants and intramolecular stacking quotients for the monoprotonated ApA were estimated. The results indicate that, in contrast to the case of GpG, ApA did not exhibit preferential protonation on either site of 3'- and 5'-linked nucleoside bases and neither enhanced nor diminished stacking was observed for ApA and ApA as compared to ApA.  相似文献   

10.
In order to design any new potential drug, it is crucial to know their corresponding pK(a) since their protonation state will be critical in the ligand-receptor interaction and it will play an essential role in their pharmacokinetic profile. Several authors have developed approaches for the computational determination of pK(a) which involve the use of a thermodynamic cycle relating pK(a) to the gas-phase proton basicity via the solvation energies of the products and the reactants. Such methods are very dependent on the solvation model used and the nature of the system. The theoretical pK(a) of a number of agonists and antagonists of the alpha1A-adrenoceptor has been computed and the performance of this approach has been tested through comparison with the available and/or measured experimental pK(a) values.  相似文献   

11.
If the titration pK values of a dibasic acid are called pK +/- logp, then its molecular pK values are pK +/- log(p+1/p). If the pH values at which the concentration of its monoprotonated form is half-maximal are called pK +/- logq, then its molecular pK values are pK +/- log(q-4+1/q).  相似文献   

12.
The primary goal of this study was to gain a better understanding of the effect of environment and ionic strength on the pK values of histidine residues in proteins. The salt-dependence of pK values for two histidine residues in ribonuclease Sa (RNase Sa) (pI=3.5) and a variant in which five acidic amino acids have been changed to lysine (5K) (pI=10.2) was measured and compared to pK values of model histidine-containing peptides. The pK of His53 is elevated by two pH units (pK=8.61) in RNase Sa and by nearly one pH unit (pK=7.39) in 5K at low salt relative to the pK of histidine in the model peptides (pK=6.6). The pK for His53 remains elevated in 1.5M NaCl (pK=7.89). The elevated pK for His53 is a result of screenable electrostatic interactions, particularly with Glu74, and a non-screenable hydrogen bond interaction with water. The pK of His85 in RNase Sa and 5K is slightly below the model pK at low salt and merges with this value at 1.5M NaCl. The pK of His85 reflects mainly effects of long-range Coulombic interactions that are screenable by salt. The tautomeric states of the neutral histidine residues are changed by charge reversal. The histidine pK values in RNase Sa are always higher than the pK values in the 5K variant. These results emphasize that the net charge of the protein influences the pK values of the histidine residues. Structure-based pK calculations capture the salt-dependence relatively well but are unable to predict absolute histidine pK values.  相似文献   

13.
Hydrogen bonded histidine-aspartate (His-Asp) pairs are critical constituents in several key enzymatic reactions. To date, the role that these pairs play in catalysis is best understood in serine and trypsin-like proteases, where structural and biochemical NMR studies have revealed important pK(a) values and hydrogen bonding patterns within the catalytic pocket. However, the role of the His-Asp pair in metal-assisted catalysis is less clear. Here, we apply liquid-state NMR to investigate the role of a critical histidine residue of apurinic endonuclease 1 (Ape1), a human DNA repair enzyme that cleaves adjacent to abasic sites in DNA using one or more divalent cations and an active-site His-Asp pair. The results of these studies suggest that the Ape1 His-Asp pair does not function as either a general base catalyst or a metal ligand. Rather, the pair likely stabilizes the pentavalent transition state necessary for phospho-transfer.  相似文献   

14.
Structural transitions of porin, a transmembrane protein   总被引:6,自引:0,他引:6  
Conformational transitions of porin were monitored using 3 independent criteria: (i) oligomeric state as observed by SDS-polyacrylamide gel electrophoresis; (ii) spectroscopic titrations (ultraviolet and circular dichroism) and (iii) chemical modifications. Four pH-dependent transitions were observed with half-maximal changes occurring at pH values of 1.6, 3.5, 11.2 and 12.4. Two of these pH values differ significantly from intrinsic pK values of the constituent amino acids of this membrane protein. Since porin is very polar despite its location predominantly within the outer membranes, this may be due to ion pair formation in the hydrophobic environment of the membrane.  相似文献   

15.
The pK values of the titratable groups in ribonuclease Sa (RNase Sa) (pI=3.5), and a charge-reversed variant with five carboxyl to lysine substitutions, 5K RNase Sa (pI=10.2), have been determined by NMR at 20 degrees C in 0.1M NaCl. In RNase Sa, 18 pK values and in 5K, 11 pK values were measured. The carboxyl group of Asp33, which is buried and forms three intramolecular hydrogen bonds in RNase Sa, has the lowest pK (2.4), whereas Asp79, which is also buried but does not form hydrogen bonds, has the most elevated pK (7.4). These results highlight the importance of desolvation and charge-dipole interactions in perturbing pK values of buried groups. Alkaline titration revealed that the terminal amine of RNase Sa and all eight tyrosine residues have significantly increased pK values relative to model compounds.A primary objective in this study was to investigate the influence of charge-charge interactions on the pK values by comparing results from RNase Sa with those from the 5K variant. The solution structures of the two proteins are very similar as revealed by NMR and other spectroscopic data, with only small changes at the N terminus and in the alpha-helix. Consequently, the ionizable groups will have similar environments in the two variants and desolvation and charge-dipole interactions will have comparable effects on the pK values of both. Their pK differences, therefore, are expected to be chiefly due to the different charge-charge interactions. As anticipated from its higher net charge, all measured pK values in 5K RNase are lowered relative to wild-type RNase Sa, with the largest decrease being 2.2 pH units for Glu14. The pK differences (pK(Sa)-pK(5K)) calculated using a simple model based on Coulomb's Law and a dielectric constant of 45 agree well with the experimental values. This demonstrates that the pK differences between wild-type and 5K RNase Sa are mainly due to changes in the electrostatic interactions between the ionizable groups. pK values calculated using Coulomb's Law also showed a good correlation (R=0.83) with experimental values. The more complex model based on a finite-difference solution to the Poisson-Boltzmann equation, which considers desolvation and charge-dipole interactions in addition to charge-charge interactions, was also used to calculate pK values. Surprisingly, these values are more poorly correlated (R=0.65) with the values from experiment. Taken together, the results are evidence that charge-charge interactions are the chief perturbant of the pK values of ionizable groups on the protein surface, which is where the majority of the ionizable groups are positioned in proteins.  相似文献   

16.
Previous studies of the low molecular mass family 11 xylanase from Bacillus circulans show that the ionization state of the nucleophile (Glu78, pK(a) 4.6) and the acid/base catalyst (Glu172, pK(a) 6.7) gives rise to its pH-dependent activity profile. Inspection of the crystal structure of BCX reveals that Glu78 and Glu172 are in very similar environments and are surrounded by several chemically equivalent and highly conserved active site residues. Hence, there are no obvious reasons why their apparent pK(a) values are different. To address this question, a mutagenic approach was implemented to determine what features establish the pK(a) values (measured directly by (13)C NMR and indirectly by pH-dependent activity profiles) of these two catalytic carboxylic acids. Analysis of several BCX variants indicates that the ionized form of Glu78 is preferentially stabilized over that of Glu172 in part by stronger hydrogen bonds contributed by two well-ordered residues, namely, Tyr69 and Gln127. In addition, theoretical pK(a) calculations show that Glu78 has a lower pK(a) value than Glu172 due to a smaller desolvation energy and more favorable background interactions with permanent partial charges and ionizable groups within the protein. The pK(a) value of Glu172 is in turn elevated due to electrostatic repulsion from the negatively charged glutamate at position 78. The results also indicate that all of the conserved active site residues act concertedly in establishing the pK(a) values of Glu78 and Glu172, with no particular residue being singly more important than any of the others. In general, residues that contribute positive charges and hydrogen bonds serve to lower the pK(a) values of Glu78 and Glu172. The degree to which a hydrogen bond lowers a pK(a) value is largely dependent on the length of the hydrogen bond (shorter bonds lower pK(a) values more) and the chemical nature of the donor (COOH > OH > CONH(2)). In contrast, neighboring carboxyl groups can either lower or raise the pK(a) values of the catalytic glutamic acids depending upon the electrostatic linkage of the ionization constants of the residues involved in the interaction. While the pH optimum of BCX can be shifted from -1.1 to +0.6 pH units by mutating neighboring residues within the active site, activity is usually compromised due to the loss of important ground and/or transition state interactions. These results suggest that the pH optima of an enzyme might be best engineered by making strategic amino acid substitutions, at positions outside of the "core" active site, that electrostatically influence catalytic residues without perturbing their immediate structural environment.  相似文献   

17.
The 1H-NMR titration curves of chemical shifts versus pH were observed for imidazole, N1-methylimidazole, L-histidine, N1-methyl-L-histidine, N3-methyl-L-histidine, and other related compounds. With these results, the macroscopic pK values of these compounds were determined by a computer curve-fitting for a simple dissociation sequence. From the pK values of imidazole and N1-methylimidazole, the perturbation for the pK of the imidazole ring due to the substitution of a proton with a methyl group was estimated as -0.21 pH unit. The microscopic pK values of the individual tautomers of the imidazole ring were estimated with the pK values of N1-methyl-L-histidine, N3-methyl-L-histidine, and perturbation due to methyl substitution. The estimated pK values were 6.73 for the N1-H tautomer and 6.12 for the N3-H tautomer. These values were in good agreement with those obtained using carboxymethyl derivatives instead of methyl derivatives. Furthermore, the macroscopic pK value (6.02) calculated using the estimated microscopic pK values agreed with that (6.03) observed for the imidazole ring of L-histidine. Thus the method in this work was indicated to be self-consistent. The microscopic pK values of tautomers were also obtained for N alpha-acetyl-L-histidine and N alpha-acetyl-L-histidine methylamide. The molar ratios of tautomers were calculated on the basis of the microscopic pK values of tautomers. The intrinsic (or unperturbed) pK value of imidazole ring and perturbations due to the CO2- and NH3+ were obtained for each of the N1-H and N3-H tautomers.  相似文献   

18.
Relationships between protein structure and ionization of carboxyl groups were investigated in 24 proteins of known structure and for which 115 aspartate and 97 glutamate pK(a) values are known. Mean pK(a) values for aspartates and glutamates are < or = 3.4 (+/-1.0) and 4.1 (+/-0.8), respectively. For aspartates, mean pK(a) values are 3.9 (+/-1.0) and 3.1 (+/-0.9) in acidic (pI < 5) and basic (pI > 8) proteins, respectively, while mean pK(a) values for glutamates are approximately 4.2 for acidic and basic proteins. Burial of carboxyl groups leads to dispersion in pK(a) values: pK(a) values for solvent-exposed groups show narrow distributions while values for buried groups range from < 2 to 6.7. Calculated electrostatic potentials at the carboxyl groups show modest correlations with experimental pK(a) values and these correlations are not improved by including simple surface-area-based terms to account for the effects of desolvation. Mean aspartate pK(a) values decrease with increasing numbers of hydrogen bonds but this is not observed at glutamates. Only 10 pK(a) values are > 5.5 and most are found in active sites or ligand-binding sites. These carboxyl groups are buried and usually accept no more than one hydrogen bond. Aspartates and glutamates at the N-termini of helices have mean pK(a) values of 2.8 (+/-0.5) and 3.4 (+/-0.6), respectively, about 0.6 units less than the overall mean values.  相似文献   

19.
Polyketide synthases (PKS) assemble structurally diverse natural products using a common mechanistic strategy that relies on a cysteine residue to anchor the polyketide during a series of decarboxylative condensation reactions that build the final reaction product. Crystallographic and functional studies of chalcone synthase (CHS), a plant-specific PKS, indicate that a cysteine-histidine pair (Cys(164)-His(303)) forms part of the catalytic machinery. Thiol-specific inactivation and the pH dependence of the malonyl-CoA decarboxylation reaction were used to evaluate the potential interaction between these two residues. Inactivation of CHS by iodoacetamide and iodoacetic acid targets Cys(164) in a pH-dependent manner (pK(a) = 5.50). The acidic pK(a) of Cys(164) suggests that an ionic interaction with His(303) stabilizes the thiolate anion. Consistent with this assertion, substitution of a glutamine for His(303) maintains catalytic activity but shifts the pK(a) of the thiol to 6.61. Although the H303A mutant was catalytically inactive, the pH-dependent incorporation of [(14)C]iodoacetamide into this mutant exhibits a pK(a) = 7.62. Subsequent analysis of the pH dependence of the malonyl-CoA decarboxylation reaction catalyzed by wild-type CHS and the H303Q and C164A mutants also supports the presence of an ion pair at the CHS active site. Structural and sequence conservation of a cysteine-histidine pair in the active sites of other PKS implies that a thiolate-imidazolium ion pair plays a central role in polyketide biosynthesis.  相似文献   

20.
The pK(A) values of (4RS)-[4-carboxy-5,8,11-tris(carboxymethyl)-1-phenyl-2-oxa-5,8,11-triazatridecan-13-oic acid] (BOPTA), a polyprotic molecule whose gadolinium complex is an important magnetic resonance imaging contrast agent for clinical use, have been determined in water, in physiologic solution (PS), in serum (S), and in cerebrospinal fluid (CSF), by means of 13C nuclear magnetic resonance spectroscopy data processed by a dedicated software package called DISCO. The aim of this study was to supply the BOPTA pK(A) values in media very similar to the in vivo environment and, consequently, to get a picture of the in vivo behavior of its Gd complex, whose thermodynamic stability is directly linked to the pK(A) values. The pK(A) values appeared to be almost equal both in D(2)O and in PS, while pK(1) and pK(5) values in CSF differ a little. In S, only pK(2) and pK(3) were calculated due to the narrow pH range used for data collection. However, these pK(A) values were found equal to those in the other media. These results represent the first direct spectroscopic evidence of a substantial invariability of BOPTA behavior in different media and they justify the extrapolation to biological fluids of the data obtained in water. The values also confirmed the high-quality performance of DISCO in calculating pK(A) values of polyprotic molecules in complex media.  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