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.
Li H  Robertson AD  Jensen JH 《Proteins》2004,55(3):689-704
A computational methodology for protein pK(a) predictions, based on ab initio quantum mechanical treatment of part of the protein and linear Poisson-Boltzmann equation treatment of the bulk solvent, is presented. The method is used to predict and interpret the pK(a) values of the five carboxyl residues (Asp7, Glu10, Glu19, Asp27, and Glu43) in the serine protease inhibitor turkey ovomucoid third domain. All the predicted pK(a) values are within 0.5 pH units of experiment, with a root-mean-square deviation of 0.31 pH units. We show that the decreased pK(a) values observed for some of the residues are primarily due to hydrogen bonds to the carboxyl oxygens. Hydrogen bonds involving amide protons are shown to be particularly important, and the effect of hydrogen bonding is shown to be nonadditive. Hydrophobic effects are also shown to be important in raising the pK(a). Interactions with charged residues are shown to have relatively little effect on the carboxyl pK(a) values in this protein, in general agreement with experiment.  相似文献   

2.
Forsyth WR  Robertson AD 《Biochemistry》2000,39(27):8067-8072
A number of carboxyl groups in turkey ovomucoid third domain (OMTKY3) have low pK(a) values. A previous study suggested that neighboring amino groups were primarily responsible for the low carboxyl pK(a) values. However, the expected elevation in pK(a) values for these amino groups was not observed. In the present study, site-directed mutagenesis is used to investigate the origins of perturbed carboxyl pK(a) values in OMTKY3. Electrostatic calculations suggest that Lys 34 has large effects, 0.4-0.6 unit, on Asp 7, Glu 10, and Glu 19 which are 5-11 A away from Lys 34. Two-dimensional (1)H NMR techniques were used to determine pK(a) values of the acidic residues in OMTKY3 mutants in which Lys 34 has been replaced with threonine and glutamine. Surprisingly, the pK(a) values in the mutants are very close to those of the wild-type protein. The insensitivity of the acidic residues to replacement of Lys 34 suggests that long-range electrostatic interactions play less of a role in perturbing carboxyl pK(a) values than originally thought. We hypothesize that hydrogen bonds play a key role in perturbing some of the carboxyl ionization equilibria in OMTKY3.  相似文献   

3.
The ionizable groups in proteins with the lowest pKs are the carboxyl groups of aspartic acid side-chains. One of the lowest, pK=0.6, is observed for Asp76 in ribonuclease T1. This low pK appeared to result from hydrogen bonds to a water molecule and to the side-chains of Asn9, Tyr11, and Thr91. The results here confirm this by showing that the pK of Asp76 increases to 1.7 in N9A, to 4.0 in Y11F, to 4.2 in T91V, to 4.4 in N9A+Y11F, to 4.9 in N9A+T91V, to 5.9 in Y11F+T91V, and to 6.4 in the triple mutant: N9A+Y11F+T91V. In ribonuclease Sa, the lowest pK=2.4 for Asp33. This pK increases to 3.9 in T56A, which removes the hydrogen bond to Asp33, and to 4.4 in T56V, which removes the hydrogen bond and replaces the -OH group with a -CH(3) group. It is clear that hydrogen bonds are able to markedly lower the pK values of carboxyl groups in proteins. These same hydrogen bonds make large contributions to the conformational stability of the proteins. At pH 7, the stability of D76A ribonuclease T1 is 3.8 kcal mol(-1) less than wild-type, and the stability of D33A ribonuclease Sa is 4.1 kcal mol(-1) less than wild-type. There is a good correlation between the changes in the pK values and the changes in stability. The results suggest that the pK values for these buried carboxyl groups would be greater than 8 in the absence of hydrogen bonds, and that the hydrogen bonds and other interactions of the carboxyl groups contribute over 8 kcal mol(-1) to the stability.  相似文献   

4.
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   

5.
To explore electrostatic interactions in ubiquitin, pK(a) values have been determined by NMR for all 12 carboxyl groups in wild-type ubiquitin and in variants where single lysines have been replaced by neutral residues. Aspartate pK(a) values in ubiquitin range from 3.1 to 3.8 and are generally less than model compound values. Most aspartate pK(a) values are within 0.2 pH unit of those predicted with a simple Tanford-Kirkwood model. Glutamate pK(a) values range from 3.8 to 4.5, close to model compound values and differing by 0.1-0.8 pH unit from calculated values. To determine the role of positive charges in modulating carboxyl pK(a) values, we mutated lysines at positions 11, 29, and 33 to glutamine and threonine. NMR studies with these six single-site mutants reveal significant interactions of Lys 11 and Lys 29 with Glu 34 and Asp 21, respectively: pK(a) values for Glu 34 and Asp 21 increase by approximately 0.5-0.8 pH unit, similar to predicted values, when the lysines are replaced by neutral residues. In contrast, the predicted interaction between Lys 33 and Glu 34 is not observed experimentally. In some instances, substitution of lysine by glutamine and threonine did not lead to the same changes in carboxyl pK(a) values. These may reflect new short-range interactions between the mutated residues and the carboxyl groups. Carboxyl pK(a) shifts > 0.5 pH unit result from mutations at groups that are <5 A from the carboxyl group. No interactions are observed at >10 A.  相似文献   

6.
Koide A  Jordan MR  Horner SR  Batori V  Koide S 《Biochemistry》2001,40(34):10326-10333
It is generally considered that electrostatic interactions on the protein surface, such as ion pairs, contribute little to protein stability, although they may play important roles in conformational specificity. We found that the tenth fibronectin type III domain of human fibronectin (FNfn10) is more stable at acidic pH than neutral pH, with an apparent midpoint of transition near pH 4. Determination of pK(a)'s for all the side chain carboxyl groups of Asp and Glu residues revealed that Asp 23 and Glu 9 have an upshifted pK(a). These residues and Asp 7 form a negatively charged patch on the surface of FNfn10, with Asp 7 centrally located between Asp 23 and Glu 9, suggesting repulsive electrostatic interactions among these residues at neutral pH. Mutant proteins, D7N and D7K, in which Asp 7 was replaced with Asn and Lys, respectively, exhibited a modest but significant increase in stability at neutral pH, compared to the wild type, and they no longer showed pH dependence of stability. The pK(a)'s of Asp 23 and Glu 9 in these mutant proteins shifted closer to their respective unperturbed values, indicating that the unfavorable electrostatic interactions have been reduced in the mutant proteins. Interestingly, the wild-type and mutant proteins were all stabilized to a similar degree by the addition of 1 M sodium chloride at both neutral and acidic pH, suggesting that the repulsive interactions between the carboxyl groups cannot be effectively shielded by 1 M sodium chloride. These results indicate that repulsive interactions between like charges on the protein surface can destabilize a protein, and protein stability can be significantly improved by relieving these interactions.  相似文献   

7.
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   

8.
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   

9.
Lysozymes from family 22 of glycoside hydrolases are usually part of the defense system against bacteria. However in ruminant artiodactyls and saprophagous insects, lysozymes are involved in the digestion of bacteria. Here, we report the first crystallographic structure of a digestive lysozyme in its native and complexed forms, the structure of lysozyme 1 from Musca domestica larvae midgut (MdL1). Structural and biochemical data presented for MdL1 are analyzed in light of digestive lysozymes' traits. The structural core is similar, but a careful analysis of a structural alignment generated with other lysozymes c reveals that significant differences occur in coil regions. The loop from MdL1 defined by residues 98-100 has one deletion previous to residue Gln100, which leads to a less exposed conformation and might justify the resistance to proteolysis observed for MdL1. In addition, Gln100 is directly involved in a few hydrogen bonds to the ligand in a yet unobserved substrate binding mode. The pK(a)s of the MdL1 catalytic residues (Glu32 and Asp50) are lower (6.40 and 3.09, respectively) than those from Gallus gallus egg lysozyme (GgL, hen egg white lysozyme-HEWL) (6.61 and 3.85, respectively). A unique feature of MdL1 is a hydrogen bond between Thr107 Ogamma and Glu32 carboxylate group, which combined with the presence of Ser106 contributes to decrease the pK(a) of Glu32. Furthermore, in MdL1 the presence of Asn46 preventing the occurrence of an electrostatic repulsion with Asp50 and the increment in the solvent exposition of Asp50 due to Pro42 insertion contribute to reduce the pK(a) of Asp50. These structural elements affecting the pK(a)s of the catalytic residues should contribute to the acidic pH optimum presented by MdL1.  相似文献   

10.
The two most buried carboxyl groups in ribonuclease Sa (RNase Sa) are Asp33 (99% buried; pK 2.4) and Asp79 (85% buried; pK 7.4). Above these pK values, the stability of the D33A variant is 6kcal/mol less than wild-type RNase Sa, and the stability of the D79A variant is 3.3kcal/mol greater than wild-type RNase Sa. The key structural difference between the carboxyl groups is that Asp33 forms three intramolecular hydrogen bonds, and Asp79 forms no intramolecular hydrogen bond. Here, we focus on Asp79 and describe studies of 11 Asp79 variants. Most of the variants were at least 2kcal/mol more stable than wild-type RNase Sa, and the most interesting was D79F. At pH 3, below the pK of Asp79, RNase Sa is 0.3kcal/mol more stable than the D79F variant. At pH 8.5, above the pK of Asp79, RNase Sa is 3.7kcal/mol less stable than the D79F variant. The unfavorable contribution of Asp79 to the stability appears to result from the Born self-energy of burying the charge and, more importantly, from unfavorable charge-charge interactions. To counteract the effect of the negative charge on Asp79, we prepared the Q94K variant and the crystal structure showed that the amino group of the Lys formed a hydrogen-bonded ion pair (distance, 2.71A; angle, 100 degrees ) with the carboxyl group of Asp79. The stability of the Q94K variant was about the same as the wild-type at pH 3, where Asp79 is uncharged, but 1kcal/mol greater than that of wild-type RNase Sa at pH 8.5, where Asp79 is charged. Differences in hydrophobicity, steric strain, Born self-energy, and electrostatic interactions all appear to contribute to the range of stabilities observed in the variants. When it is possible, replacing buried, non-hydrogen bonded, ionizable side-chains with non-polar side-chains is an excellent means of increasing protein stability.  相似文献   

11.
Low-energy conformations of the S-peptide fragment (20 amino acid residues long) of ribonuclease A were studied by Monte Carlo simulated annealing. The obtained lowest-energy structures have alpha-helices with different size and location, depending distinctively on the ionizing states of acidic amino acid residues. The simulation started from completely random initial conformation and was performed without any bias toward a particular structure. The most conspicuous alpha-helices arose from the simulation when both Glu 9 and Asp 14 were assumed to be electrically neutral, whereas the resulting conformations became much less helical when Asp 14 rather than Glu 9 was allowed to have a negative charge. Together with experimental evidence that the alpha-helix in the S-peptide is most stable at pH 3.8, we consider the helix formation need the carboxyl group of Asp 14 to be electrically neutral in this weakly acidic condition. In contrast, a negative charge at Asp 14 appears to function in support of a view that this residue is crucial to helix termination owing to its possibility to form a salt bridge with His 12. These results indicate that the conformation of the S-peptide depends considerably on the ionizing state of Asp 14.  相似文献   

12.
A Giletto  C N Pace 《Biochemistry》1999,38(40):13379-13384
The side-chain carboxyl of Asp 76 in ribonuclease T1 (RNase T1) is buried, charged, non-ion-paired, and forms three good intramolecular hydrogen bonds (2.63, 2.69, and 2.89 A) and a 2.66 A hydrogen bond to a buried, conserved water molecule. When Asp 76 was replaced by Asn, Ser, and Ala, the conformational stability of the protein decreased by 3.1, 3.2, and 3.7 kcal/mol, respectively. The stability was measured as a function of pH for wild-type RNase T1 and the D76N mutant and showed that the pH dependence below pH 3 was almost entirely due to Asp 76. The pK of Asp 76 is 0.5 in the native state and 3.7 in the denatured state. Thus, the hydrogen bonding of the carboxyl group of Asp 76 contributes more than half of the net stability of RNase T1 at pH 7. In addition, the charged carboxyl of Asp 76 stabilizes structure in the denatured states of RNase T1 that is not present in D76N, D76S, and D76A.  相似文献   

13.
Manikandan K  Ramakumar S 《Proteins》2004,56(4):768-781
A comprehensive database analysis of C--H...O hydrogen bonds in 3124 alpha-helices and their corresponding helix termini has been carried out from a nonredundant data set of high-resolution globular protein structures resolved at better than 2.0 A in order to investigate their role in the helix, the important protein secondary structural element. The possible occurrence of 5 --> 1 C--H...O hydrogen bond between the ith residue CH group and (i - 4)th residue C==O with C...O < or = 3.8 A is studied, considering as potential donors the main-chain Calpha and the side-chain carbon atoms Cbeta, Cgamma, Cdelta and Cepsilon. Similar analysis has been carried out for 4 --> 1 C--H...O hydrogen bonds, since the C--H...O hydrogen bonds found in helices are predominantly of type 5 --> 1 or 4 --> 1. A total of 17,367 (9310 of type 5 --> 1 and 8057 of type 4 --> 1) C--H...O hydrogen bonds are found to satisfy the selected criteria. The average stereochemical parameters for the data set suggest that the observed C--H...O hydrogen bonds are attractive interactions. Our analysis reveals that the Cgamma and Cbeta hydrogen atom(s) are frequently involved in such hydrogen bonds. A marked preference is noticed for aliphatic beta-branched residue Ile to participate in 5 --> 1 C--H...O hydrogen bonds involving methylene Cgamma 1 atom as donor in alpha-helices. This may be an enthalpic compensation for the greater loss of side-chain conformational entropy for beta-branched amino acids due to the constraint on side-chain torsion angle, namely, chi1, when they occur in helices. The preference of amino acids for 4 --> 1 C--H...O hydrogen bonds is found to be more for Asp, Cys, and for aromatic residues Trp, Phe, and His. Interestingly, overall propensity for C--H...O hydrogen bonds shows that a majority of the helix favoring residues such as Met, Glu, Arg, Lys, Leu, and Gln, which also have large side-chains, prefer to be involved in such types of weak attractive interactions in helices. The amino acid side-chains that participate in C--H...O interactions are found to shield the acceptor carbonyl oxygen atom from the solvent. In addition, C--H...O hydrogen bonds are present along with helix stabilizing salt bridges. A novel helix terminating interaction motif, X-Gly with Gly at C(cap) position having 5 --> 1 Calpha--H...O, and a chain reversal structural motif having 1 --> 5 Calpha-H...O have been identified and discussed. Our analysis highlights that a multitude of local C--H...O hydrogen bonds formed by a variety of amino acid side-chains and Calpha hydrogen atoms occur in helices and more so at the helix termini. It may be surmised that the main-chain Calpha and the side-chain CH that participate in C--H...O hydrogen bonds collectively augment the cohesive energy and thereby contribute together with the classical N--H...O hydrogen bonds and other interactions to the overall stability of helix and therefore of proteins.  相似文献   

14.
The first three residues at the N terminus of the alpha-helix are called N1, N2 and N3. We surveyed 2102 alpha-helix N termini in 298 high-resolution, non-homologous protein crystal structures for N1, N2 and N3 amino acid and side-chain rotamer propensities and hydrogen-bonding patterns. We find strong structural preferences that are unique to these sites. The rotamer distributions as a function of amino acid identity and position in the helix are often explained in terms of hydrogen-bonding interactions to the free N1, N2 and N3 backbone NH groups. Notably, the "good N2" amino acid residues Gln, Glu, Asp, Asn, Ser, Thr and His preferentially form i, i or i,i+1 hydrogen bonds to the backbone, though this is hindered by good N-caps (Asp, Asn, Ser, Thr and Cys) that compete for these hydrogen bond donors. We find a number of specific side-chain to side-chain interactions between N1 and N2 or between the N-cap and N2 or N3, such as Arg(N-cap) to Asp(N2). The strong energetic and structural preferences found for N1, N2 and N3, which differ greatly from positions within helix interiors, suggest that these sites should be treated explicitly in any consideration of helical structure in peptides or proteins.  相似文献   

15.
N2 is the second position in the alpha-helix. All 20 amino acids were placed in the N2 position of a synthetic helical peptide (CH(3)CO-[AXAAAAKAAAAKAAGY]-NH(2)) and the helix content was measured by circular dichroism spectroscopy at 273K. The dependence of peptide helicity on N2 residue identity has been used to determine a free-energy scale by analysis with a modified Lifson-Roig helix-coil theory that includes a parameter for the N2 energy (n2). The rank order of DeltaDeltaG((relative to Ala)) is Glu(-), Asp(-) > Ala > Glu(0), Leu, Val, Gln, Thr, Ile, Ser, Met, Asp(0), His(0), Arg, Cys, Lys, Phe > Asn, > Gly, His(+), Pro, Tyr. The results correlate very well with N2 propensities in proteins, moderately well with N1 and helix interior preferences, and not at all with N-cap preferences. The strongest energetic effects result from interactions with the helix dipole, which favors negative charges at the helix N terminus. Hydrogen bonds to side chains at N2, such as Gln, Ser, and Thr, are weak, despite occurring frequently in protein crystal structures, in contrast to the N-cap position. This is because N-cap hydrogen bonds are close to linear, whereas N2 hydrogen bonds have poor geometry. These results can be used to modify protein stability rationally, help design helices, and improve prediction of helix location and stability.  相似文献   

16.
EmrE is an Escherichia coli multidrug transporter that confers resistance to a variety of toxins by removing them in exchange for hydrogen ions. The detergent-solubilized protein binds tetraphenylphosphonium (TPP(+)) with a K(D) of 10 nM. One mole of ligand is bound per approximately 3 mol of EmrE, suggesting that there is one binding site per trimer. The steep pH dependence of binding suggests that one or more residues, with an apparent pK of approximately 7.5, release protons prior to ligand binding. A conservative Asp replacement (E14D) at position 14 of the only membrane-embedded charged residue shows little transport activity, but binds TPP(+) at levels similar to those of the wild-type protein. The apparent pK of the Asp shifts to <5.0. The data are consistent with a mechanism requiring Glu14 for both substrate and proton recognition. We propose a model in which two of the three Glu14s in the postulated trimeric EmrE homooligomer deprotonate upon ligand binding. The ligand is released on the other face of the membrane after binding of protons to Glu14.  相似文献   

17.
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   

18.
Lactose/H(+) symport by lactose permease of Escherichia coli involves interactions between four irreplaceable charged residues in transmembrane helices that play essential roles in H(+) translocation and coupling [Glu269 (helix VIII) with His322 (helix X) and Arg302 (helix IX) with Glu325 (helix X)], as well as Glu126 (helix IV) and Arg144 (helix V) which are obligatory for substrate binding. The conservative mutation Glu325-->Asp causes a 10-fold reduction in the V(max) for active lactose transport and markedly decreased lactose-induced H(+) influx with no effect on exchange or counterflow, neither of which involves H(+) symport. Thus, shortening the side chain may weaken the interaction of the carboxyl group at position 325 with the guanidino group of Arg302. Therefore, Gly-scanning mutagenesis of helices IX and X and the intervening loop was employed systematically with mutant Glu325-->Asp in an effort to rescue function by introducing conformational flexibility between the two helices. Five Gly replacement mutants in the Glu325-->Asp background are identified that exhibit significantly higher transport activity. Furthermore, mutant Val316-->Gly/Glu325-->Asp catalyzes active transport, efflux, and lactose-induced H(+) influx with kinetic properties approaching those of wild-type permease. It is proposed that introduction of conformational flexibility at the interface between helices IX and X improves juxtapositioning between Arg302 and Asp325 during turnover, thereby allowing more effective deprotonation of the permease on the inner surface of the membrane [Sahin-Tóth, M., Karlin, A., and Kaback, H. R. (2000) Proc. Natl. Acad. Sci. U.S.A. 97, 10729-10732.  相似文献   

19.
The chemistry of active-site cysteine residues is central to the activity of thiol-disulfide oxidoreductases of the thioredoxin superfamily. In these reactions, a nucleophilic thiolate is required, but the associated pK(a) values differ vastly in the superfamily, from less than 4 in DsbA to greater than 7 in Trx. The factors that stabilize this thiolate are, however, not clearly established. The glutaredoxins (Grxs), which are members of this superfamily, contain a Cys-Pro-Tyr-Cys motif in their active site. In reduced Grxs, the pK(a) of the N-terminal active-site nucleophilic cysteine residue is lowered significantly, and the stabilization of the corresponding thiolate is expected to influence the redox potential of these enzymes. Here, we use a combination of long molecular dynamics (MD) simulations, pK(a) calculations, and experimental investigations to derive the structure and dynamics of the reduced active site from Escherichia coli Grx3, and investigate the factors that stabilize the thiolate. Several different MD simulations converged toward a consensus conformation for the active-site cysteine residues (Cys11 and Cys14), after a number of local conformational changes. Key features of the model were tested experimentally by measurement of NMR scalar coupling constants, and determination of pK(a) values of selected residues. The pK(a) values of the Grx3 active-site residues were calculated during the MD simulations, and support the underlying structural model. The structure of Grx3, in combination with the pK(a) calculations, indicate that the pK(a) of the N-terminal active-site cysteine residue in Grx3 is intermediate between that of its counterpart in DsbA and Trx. The pK(a) values in best agreement with experiment are obtained with a low (<4) protein dielectric constant. The calculated pK(a) values fluctuate significantly in response to protein dynamics, which underscores the importance of the details of the underlying structures when calculating pK(a) values. The thiolate of Cys11 is stabilized primarily by direct hydrogen bonding with the amide protons of Tyr13 and Cys14 and the thiol proton of Cys14, rather than by long-range interactions from charged groups or from a helix macrodipole. From the comparison of reduced Grx3 with other members of the thioredoxin superfamily, a unifying theme for the structural basis of thiol pK(a) differences in this superfamily begins to emerge.  相似文献   

20.
Photoactive yellow protein (PYP) is a blue light sensor present in the purple photosynthetic bacterium Ectothiorhodospira halophila, which undergoes a cyclic series of absorbance changes upon illumination at its lambda(max) of 446 nm. The anionic p-hydroxycinnamoyl chromophore of PYP is covalently bound as a thiol ester to Cys69, buried in a hydrophobic pocket, and hydrogen-bonded via its phenolate oxygen to Glu46 and Tyr42. The chromophore becomes protonated in the photobleached state (I(2)) after it undergoes trans-cis isomerization, which results in breaking of the H-bond between Glu46 and the chromophore and partial exposure of the phenolic ring to the solvent. In previous mutagenesis studies of a Glu46Gln mutant, we have shown that a key factor in controlling the color and photocycle kinetics of PYP is this H-bonding system. To further investigate this, we have now characterized Glu46Asp and Glu46Ala mutants. The ground-state absorption spectrum of the Glu46Asp mutant shows a pH-dependent equilibrium (pK = 8.6) between two species: a protonated (acidic) form (lambda(max) = 345 nm), and a slightly blue-shifted deprotonated (basic) form (lambda(max) = 444 nm). Both of these species are photoactive. A similar transition was also observed for the Glu46Ala mutant (pK = 7.9), resulting in two photoactive red-shifted forms: a basic species (lambda(max) = 465 nm) and a protonated species (lambda(max) = 365 nm). We attribute these spectral transitions to protonation/deprotonation of the phenolate oxygen of the chromophore. This is demonstrated by FT Raman spectra. Dark recovery kinetics (return to the unphotolyzed state) were found to vary appreciably between these various photoactive species. These spectral and kinetic properties indicate that the hydrogen bond between Glu46 and the chromophore hydroxyl group is a dominant factor in controlling the pK values of the chromophore and the glutamate carboxyl.  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