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 250 毫秒
1.
Our goal was to gain a better understanding of the contribution of the burial of polar groups and their hydrogen bonds to the conformational stability of proteins. We measured the change in stability, Δ(ΔG), for a series of hydrogen bonding mutants in four proteins: villin headpiece subdomain (VHP) containing 36 residues, a surface protein from Borrelia burgdorferi (VlsE) containing 341 residues, and two proteins previously studied in our laboratory, ribonucleases Sa (RNase Sa) and T1 (RNase T1). Crystal structures were determined for three of the hydrogen bonding mutants of RNase Sa: S24A, Y51F, and T95A. The structures are very similar to wild type RNase Sa and the hydrogen bonding partners form intermolecular hydrogen bonds to water in all three mutants. We compare our results with previous studies of similar mutants in other proteins and reach the following conclusions. (1) Hydrogen bonds contribute favorably to protein stability. (2) The contribution of hydrogen bonds to protein stability is strongly context dependent. (3) Hydrogen bonds by side chains and peptide groups make similar contributions to protein stability. (4) Polar group burial can make a favorable contribution to protein stability even if the polar groups are not hydrogen bonded. (5) The contribution of hydrogen bonds to protein stability is similar for VHP, a small protein, and VlsE, a large protein.  相似文献   

2.
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   

3.
For 30 years, the prevailing view has been that the hydrophobic effect contributes considerably more than hydrogen bonding to the conformational stability of globular proteins. The results and reasoning presented here suggest that hydrogen bonding and the hydrophobic effect make comparable contributions to the conformational stability of ribonuclease T1 (RNase T1). When RNase T1 folds, 86 intramolecular hydrogen bonds with an average length of 2.95 A are formed. Twelve mutants of RNase T1 [Tyr----Phe (5), Ser----Ala (3), and Asn----Ala (4)] have been prepared that remove 17 of the hydrogen bonds with an average length of 2.93 A. On the basis of urea and thermal unfolding studies of these mutants, the average decrease in conformational stability due to hydrogen bonding is 1.3 kcal/mol per hydrogen bond. This estimate is in good agreement with results from several related systems. Thus, we estimate that hydrogen bonding contributes about 110 kcal/mol to the conformational stability of RNase T1 and that this is comparable to the contribution of the hydrophobic effect. Accepting the idea that intramolecular hydrogen bonds contribute 1.3 +/- 0.6 kcal/mol to the stability of systems in an aqueous environment makes it easier to understand the stability of the "molten globule" states of proteins, and the alpha-helical conformations of small peptides.  相似文献   

4.
We previously suggested that proteins gain more stability from the burial and hydrogen bonding of polar groups than from the burial of nonpolar groups (Pace, C. N. (2001) Biochemistry 40, 310-313). To study this further, we prepared eight Thr-to-Val mutants of RNase Sa, four in which the Thr side chain is hydrogen-bonded and four in which it is not. We measured the stability of these mutants by analyzing their thermal denaturation curves. The four hydrogen-bonded Thr side chains contribute 1.3 +/- 0.9 kcal/mol to the stability; those that are not still contribute 0.4 +/- 0.9 kcal/mol to the stability. For 40 Thr-to-Val mutants of 11 proteins, the average decrease in stability is 1.0 +/- 1.0 kcal/mol when the Thr side chain is hydrogen-bonded and 0.0 +/- 0.5 kcal/mol when it is not. This is clear evidence that hydrogen bonds contribute favorably to protein stability. In addition, we prepared four Val-to-Thr mutants of RNase Sa, measured their stability, and determined their crystal structures. In all cases, the mutants are less stable than the wild-type protein, with the decreases in stability ranging from 0.5 to 4.4 kcal/mol. For 41 Val-to-Thr mutants of 11 proteins, the average decrease in stability is 1.8 +/- 1.3 kcal/mol and is unfavorable for 40 of 41 mutants. This shows that placing an [bond]OH group at a site designed for a [bond]CH3 group is very unfavorable. So, [bond]OH groups can contribute favorably to protein stability, even if they are not hydrogen-bonded, if the site was selected for an [bond]OH group, but they will make an unfavorable contribution to stability, even if they are hydrogen-bonded, when they are placed at a site selected for a [bond]CH3 group. The contribution that polar groups make to protein stability depends strongly on their environment.  相似文献   

5.
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   

6.
Anabaena apoflavodoxin contains a single histidine residue (H34) that interacts with two aromatic residues (F7 and Y47). The histidine and phenylalanine rings are almost coplanar and they can establish a cation-pi interaction when the histidine is protonated. The histidine and tyrosine side-chains are engaged in a hydrogen bond, which is their only contact. We analyse the energetics of these interactions using p Ka-shift analysis, double-mutant cycle analysis at two pH values, and X-ray crystallography. The H/F interaction is very weak when the histidine is neutral, but it is strengthened by 0.5 kcal mol-1on histidine protonation. Supporting this fact, the histidine p Kain a F7L mutant is 0.4 pH units lower than in wild-type. The strength of the H/Y hydrogen bond is 0.7 kcal mol-1when the histidine is charged, and it becomes stronger (1.3 kcal mol-1) when the histidine is neutral. This is consistent with our observation that the (H34)Nepsilon2-OH(Y47) distance is slightly shorter in the apoflavodoxin structure at pH 9.0 than in the previously reported structure at pH 6.0. It is also consistent with a histidine p Kavalue 0.6 pH units higher in a Y47F mutant than in the wild-type protein. We suggest that the higher stability of the neutral hydrogen bond could be due to a higher desolvation penalty of the charged hydrogen bond that would offset its more favourable enthalpy of formation. The relationship between hydrogen bond strength and the contribution of hydrogen bonds to protein stability is discussed.  相似文献   

7.
Ribonuclease Sa (RNase Sa) contains no tryptophan (Trp) residues. We have added single Trp residues to RNase Sa at sites where Trp is found in four other microbial ribonucleases, yielding the following variants of RNase Sa: Y52W, Y55W, T76W, and Y81W. We have determined crystal structures of T76W and Y81W at 1.1 and 1.0 A resolution, respectively. We have studied the fluorescence properties and stabilities of the four variants and compared them to wild-type RNase Sa and the other ribonucleases on which they were based. Our results should help others in selecting sites for adding Trp residues to proteins. The most interesting findings are: 1), Y52W is 2.9 kcal/mol less stable than RNase Sa and the fluorescence intensity emission maximum is blue-shifted to 309 nm. Only a Trp in azurin is blue-shifted to a greater extent (308 nm). This blue shift is considerably greater than observed for Trp71 in barnase, the Trp on which Y52W is based. 2), Y55W is 2.1 kcal/mol less stable than RNase Sa and the tryptophan fluorescence is almost completely quenched. In contrast, Trp59 in RNase T1, on which Y55W is based, has a 10-fold greater fluorescence emission intensity. 3), T76W is 0.7 kcal/mol more stable than RNase Sa, indicating that the Trp side chain has more favorable interactions with the protein than the threonine side chain. The fluorescence properties of folded Y76W are similar to those of the unfolded protein, showing that the tryptophan side chain in the folded protein is largely exposed to solvent. This is confirmed by the crystal structure of the T76W which shows that the side chain of the Trp is only approximately 7% buried. 4), Y81W is 0.4 kcal/mol less stable than RNase Sa. Based on the crystal structure of Y81W, the side chain of the Trp is 87% buried. Although all of the Trp side chains in the variants contribute to the unusual positive circular dichroism band observed near 235 nm for RNase Sa, the contribution is greatest for Y81W.  相似文献   

8.
Boswell S  Mathew J  Beach M  Osuna R  Colón W 《Biochemistry》2004,43(10):2964-2977
The diverse roles of tyrosine residues in proteins may be attributed to their dual hydrophobic and polar nature, which can result in hydrophobic and ring stacking interactions, as well as hydrogen bonding. The small homodimeric DNA binding protein, factor for inversion stimulation (FIS), contains four tyrosine residues located at positions 38, 51, 69, and 95, each involved in specific intra- or intermolecular interactions. To investigate their contributions to the stability, flexibility, and spectroscopic properties of FIS, each one was independently mutated to phenylalanine. Equilibrium denaturation experiments show that Tyr95 and Tyr51 stabilize FIS by about 2 and 1 kcal/mol, respectively, as a result of their involvement in a hydrogen bond-salt bridge network. In contrast, Tyr38 destabilizes FIS by about 1 kcal/mol due to the placement of a hydroxyl group in a hydrophobic environment. The stability of FIS was not altered when the solvent-exposed Tyr69 was mutated. Limited proteolysis with trypsin and V8 proteases was used to monitor the flexibility of the C-terminus (residues 71-98) and the dimer core (residues 26-70), respectively. The results for Y95F and Y51F FIS revealed a different proteolytic susceptibility of the dimer core compared to the C-terminus, suggesting an increased flexibility of the latter. DNA binding affinity of the various FIS mutants was only modestly affected and correlated inversely with the C-terminal flexibility probed by trypsin proteolysis. Deconvolution of the fluorescence contribution of each mutant revealed that it varies in intensity and direction for each tyrosine in WT FIS, highlighting the role of specific interactions and the local environment in determining the fluorescence of tyrosine residues. The significant changes in stability, flexibility, and signals observed for the Y51F and Y95F mutations are attributed to their coupled participation in the hydrogen bond-salt bridge network. These results highlight the importance of tyrosine hydrogen-bonding and packing interactions for the stability of FIS and demonstrate the varying roles that tyrosine residues can play on the structural and spectroscopic properties of even small proteins.  相似文献   

9.
Increasing the conformational stability of proteins is an important goal for both basic research and industrial applications. In vitro selection has been used successfully to increase protein stability, but more often site‐directed mutagenesis is used to optimize the various forces that contribute to protein stability. In previous studies, we showed that improving electrostatic interactions on the protein surface and improving the β‐turn sequences were good general strategies for increasing protein stability, and used them to increase the stability of RNase Sa. By incorporating seven of these mutations in RNase Sa, we increased the stability by 5.3 kcal/mol. Adding one more mutation, D79F, gave a total increase in stability of 7.7 kcal/mol, and a melting temperature 28°C higher than the wild‐type enzyme. Surprisingly, the D79F mutation lowers the change in heat capacity for folding, ΔCp, by 0.6 kcal/mol/K. This suggests that this mutation stabilizes structure in the denatured state ensemble. We made other mutants that give some insight into the structure present in the denatured state. Finally, the thermodynamics of folding of these stabilized variants of RNase Sa are compared with those observed for proteins from thermophiles.  相似文献   

10.
It is difficult to increase protein stability by adding hydrogen bonds or burying nonpolar surface. The results described here show that reversing the charge on a side chain on the surface of a protein is a useful way of increasing stability. Ribonuclease T1 is an acidic protein with a pI approximately 3.5 and a net charge of approximately -6 at pH 7. The side chain of Asp49 is hyperexposed, not hydrogen bonded, and 8 A from the nearest charged group. The stability of Asp49Ala is 0.5 kcal/mol greater than wild-type at pH 7 and 0.4 kcal/mol less at pH 2.5. The stability of Asp49His is 1.1 kcal/mol greater than wild-type at pH 6, where the histidine 49 side chain (pKa = 7.2) is positively charged. Similar results were obtained with ribonuclease Sa where Asp25Lys is 0.9 kcal/mol and Glu74Lys is 1.1 kcal/mol more stable than the wild-type enzyme. These results suggest that protein stability can be increased by improving the coulombic interactions among charged groups on the protein surface. In addition, the stability of RNase T1 decreases as more hydrophobic aromatic residues are substituted for Ala49, indicating a reverse hydrophobic effect.  相似文献   

11.
RNase T is a classical member of the DEDDh family of exonucleases with a unique sequence preference in that its 3′‐to‐5′ exonuclease activity is blocked by a 3′‐terminal dinucleotide CC in digesting both single‐stranded RNA and DNA. Our previous crystal structure analysis of RNase T‐DNA complexes show that four phenylalanine residues, F29, F77, F124, and F146, stack with the two 3′‐terminal nucleobases. To elucidate if the π–π stacking interactions between aromatic residues and nucleobases play a critical role in sequence‐specific protein–nucleic acid recognition, here we mutated two to four of the phenylalanine residues in RNase T to tryptophan (W mutants) and tyrosine (Y mutants). The Escherichia coli strains expressing either the W mutants or the Y mutants had slow growth phenotypes, suggesting that all of these mutants could not fully substitute the function of the wild‐type RNase T in vivo. DNA digestion assays revealed W mutants shared similar sequence specificity with wild‐type RNase T. However, the Y mutants exhibited altered sequence‐dependent activity, digesting ssDNA with both 3′‐end CC and GG sequences. Moreover, the W and Y mutants had reduced DNA‐binding activity and lower thermal stability as compared to wild‐type RNase T. Taken together, our results suggest that the four phenylalanine residues in RNase T not only play critical roles in sequence‐specific recognition, but also in overall protein stability. Our results provide the first evidence showing that the π−π stacking interactions between nucleobases and protein aromatic residues may guide the sequence‐specific activity for DNA and RNA enzymes.  相似文献   

12.
An intricate architecture of covalent bonds and noncovalent interactions appear to position the side chain of Lys 41 properly within the active site of bovine pancreatic ribonuclease A (RNase A). One of these interactions arises from Tyr 97, which is conserved in all 41 RNase A homologues of known sequence. Tyr 97 has a solvent-inaccessible side chain that donates a hydrogen bond to the main-chain oxygen of Lys 41. Here, the role of Tyr 97 was examined by replacing Tyr 97 with a phenylalanine, alanine, or glycine residue. All three mutant proteins have diminished catalytic activity, with the value of Kcat being perturbed more significantly than that of Km. The free energies with which Y97F, Y97A, and Y97G RNase A bind to the rate-limiting transition state during the cleavage of poly(cytidylic acid) are diminished by 0.74, 3.3, and 3.8 kcal/mol, respectively. These results show that even though Tyr 97 is remote from the active site, its side chain contributes to catalysis. The role of Tyr 97 in the thermal stability of RNase A is large. The conformational free energies of native Y97F, Y97A, and Y97G RNase A are decreased by 3.54, 12.0, and 11.7 kcal/mol, respectively. The unusually large decrease in stability caused by the Tyr-->Phe mutation could result from a decrease in the barrier to isomerization of the Lys 41-Pro 42 peptide bond.  相似文献   

13.
Laurents DV  Scholtz JM  Rico M  Pace CN  Bruix M 《Biochemistry》2005,44(21):7644-7655
The conformational stability of ribonuclease Sa (RNase Sa) has been measured at the per-residue level by NMR-monitored hydrogen exchange at pH* 5.5 and 30 degrees C. In these conditions, the exchange mechanism was found to be EXII. The conformational stability calculated from the slowest exchanging amide groups was found to be 8.8 kcal/mol, in close agreement with values determined by spectroscopic methods. RNase Sa is curiously rich in acidic residues (pI = 3.5) with most basic residues being concentrated in the active-site cleft. The effects of dissolved salts on the stability of RNase Sa was studied by thermal denaturation experiments in NaCl and GdmCl and by comparing hydrogen exchange rates in 0.25 M NaCl to water. The protein was found to be stabilized by salt, with the magnitude of the stabilization being influenced by the solvent exposure and local charge environment at individual amide groups. Amide hydrogen exchange was also measured in 0.25, 0.50, 0.75, and 1.00 M GdmCl to characterize the unfolding events that permit exchange. In contrast to other microbial ribonucleases studied to date, the most protected, globally exchanging amides in RNase Sa lie not chiefly in the central beta strands but in the 3/10 helix and an exterior beta strand. These structural elements are near the Cys7-Cys96 disulfide bond.  相似文献   

14.
We have examined the influence of surface hydrogen bonds on the stability of proteins by studying the effects of mutations of human immunoglobulin light chain variable domain (VL). In addition to the variants Y27dD, N28F, and T94H of protein κIV Len that were previously described, we characterized mutants M4L, L27cN, L27cQ, and K39T, double mutant M4L/Y27dD, and triple mutant M4L/Y27dD/T94H. The triple mutant had an enhanced thermodynamic stability of 4.2 kcal/mol. We determined the structure of the triple mutant by x-ray diffraction and correlated the changes in stability due to the mutations with changes in the three-dimensional structure. Y27dD mutant had increased stability of Len by 2.7 kcal/mol, a large value for a single mutation. Asp27d present in CDR1 formed hydrogen bonds with the side-chain and main-chain atoms within the loop. In the case of the K39T mutant, which reduces stability by 2 kcal/mol, Lys39 in addition to forming a hydrogen bond with a carbonyl oxygen of a neighboring loop may also favorably influence the surface electrostatics of the molecule. We showed that hydrogen bonds between residues in surface loops can add to the overall stability of the VL domains. The contribution to stability is further increased if the surface residue makes more than one hydrogen bond or if it forms a hydrogen bond between neighboring turns or loops separated from each other in the amino acid sequence. Based on our experiments we suggest that stabilization of proteins might be systematically accomplished by introducing additional hydrogen bonds on the surface. These substitutions are more straightforward to predict than core-packing interactions and can be selected to avoid affecting the protein’s function.  相似文献   

15.
Wu XH  Chen RC  Gao Y  Wu YD 《Biochemistry》2010,49(47):10237-10245
We recently found that Asp-His-Ser/Thr-Trp hydrogen-bonded tetrads are widely and uniquely present in the WD40-repeat proteins. WDR5 protein is a seven WD40-repeat propeller with five such tetrads. To explore the effect of the tetrad on the structure and stability of WD40-repeat proteins, the wild-type WDR5 and its seven mutants involving the substitutions of tetrad residues have been isolated. The crystal structures of the wild-type WDR5 and its three WDR5 mutants have been determined by X-ray diffraction method. The mutations of the tetrad residues are found not to change the basic structural features. The denaturing profiles of the wild type and the seven mutants with the use of denaturant guanidine hydrochloride have been studied by circular dichroism spectroscopy to determine the folding free energies of these proteins. The folding free energies of the wild type and the S62A, S146A, S188A, D192E, W330F, W330Y, and D324E mutants are measured to be about -11.6, -2.7, -3.1, -2.9, -3.6, -7.1, -7.0, and -7.5 kcal/mol, respectively. These suggest that (1) the hydrogen bonds in these hydrogen bond networks are unusually strong; (2) each hydrogen-bonded tetrad provides over 12 kcal/mol stability to the protein; thus, the removal of any single tetrad would cause unfolding of the protein; (3) since there are five tetrads, the protein must be in a highly unstable state without the tetrads, which might be related to its biological functions.  相似文献   

16.
Molecular dynamics simulations were carried out to calculate free energy differences between the folded and unfolded states of wild type and mutant collagen model peptides. The calculated stability of the collagen models was compared with the severity of osteogenesis imperfecta. Free energy differences of Gly → Xaa (Xaa: Ser, Cys, Glu, and Asp) mutations between the wild type and the mutants at position 15 of the model peptide were 3.8, 4.2, 5.6, and 8.8 kcal/mol, respectively. The corresponding free energy differences of a second Gly mutation at the same position in different chains were, on average, 1.3, 1.5, 2.9, and 5.4 kcal/mol, respectively. Free energy simulations were also performed to estimate the relative stability between an oxidized form and a reduced form of the mutants containing two Cys residues, which indicated that the mutant of the collagen-like peptide containing an intramolecular disulfide bond was more stable than the mutant containing one Cys residue but less stable than the wild type. The calculated free energy differences between an oxidized and a reduced form of the mutants containing two Cys residues are 0.8 and 2.6 kcal/mol for the disulfide bonds between Chains A and B and between Chains A and C, respectively.  相似文献   

17.
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   

18.
The side-chains of phenylalanine and tyrosine residues in proteins are frequently found to be involved in pairwise interactions. These occur both within repeating elements of secondary structure and in tertiary and quaternary interactions. It has been suggested that they are important in protein folding and stability, and non-bonded potential energy calculations indicate that a typical aromatic-aromatic interaction has an energy of between -1 and -2 kcal/mol and contributes between -0.6 and -1.3 kcal/mol to protein stability. There is such an aromatic pair on the solvent-exposed face of the first alpha-helix of barnase, the small ribonuclease from Bacillus amyloliquefaciens. The edge of the aromatic ring of Tyr17 interacts with the face of that of Tyr13. The two residues have been mutated both singly and pairwise to alanine, and their free energies of unfolding determined by denaturation with urea. Application of the double-mutant cycle analysis gives an interaction energy of -1.3 kcal/mol for the aromatic pair in the folded protein relative to solvation by water in the unfolded protein. This value is similar to that calculated from the change in surface-accessible area between the rings on the formation of the pair. Analysis of a further double-mutant cycle in which the Tyr residues are mutated to Phe indicates that the aromatic-aromatic interactions of Tyr/Tyr and Phe/Phe make identical contributions to protein stability. However, Tyr is preferred to Phe by 0.3(+/- 0.04) kcal/mol at the solvent-exposed face of the alpha-helix.  相似文献   

19.
Proteins have evolved to compensate for detrimental mutations. However, compensatory mechanisms for protein defects are not well understood. Using ketosteroid isomerase (KSI), we investigated how second-site mutations could recover defective mutant function and stability. Previous results revealed that the Y30F mutation rescued the Y14F, Y55F and Y14F/Y55F mutants by increasing the catalytic activity by 23-, 3- and 1.3-fold, respectively, and the Y55F mutant by increasing the stability by 3.3 kcal/mol. To better understand these observations, we systematically investigated detailed structural and thermodynamic effects of the Y30F mutation on these mutants. Crystal structures of the Y14F/Y30F and Y14F/Y55F mutants were solved at 2.0 and 1.8 previoulsy solved structures of wild-type and other mutant KSIs. Structural analyses revealed that the Y30F mutation partially restored the active-site cleft of these mutant KSIs. The Y30F mutation also increased Y14F and Y14F/Y55F mutant stability by 3.2 and 4.3 kcal/mol, respectively, and the melting temperatures of the Y14F, Y55F and Y14F/Y55F mutants by 6.4°C, 5.1°C and 10.0°C, respectively. Compensatory effects of the Y30F mutation on stability might be due to improved hydrophobic interactions because removal of a hydroxyl group from Tyr30 induced local compaction by neighboring residue movement and enhanced interactions with surrounding hydrophobic residues in the active site. Taken together, our results suggest that perturbed active-site geometry recovery and favorable hydrophobic interactions mediate the role of Y30F as a secondsite suppressor.  相似文献   

20.
Kim DH  Jang DS  Nam GH  Choi G  Kim JS  Ha NC  Kim MS  Oh BH  Choi KY 《Biochemistry》2000,39(16):4581-4589
Delta(5)-3-Ketosteroid isomerase from Pseudomonas putida biotype B is one of the most proficient enzymes catalyzing an allylic isomerization reaction at rates comparable to the diffusion limit. The hydrogen-bond network (Asp99... Wat504...Tyr14...Tyr55...Tyr30) which links the two catalytic residues, Tyr14 and Asp99, to Tyr30, Tyr55, and a water molecule in the highly apolar active site has been characterized in an effort to identify its roles in function and stability. The DeltaG(U)(H2O) determined from equilibrium unfolding experiments reveals that the elimination of the hydroxyl group of Tyr14 or Tyr55 or the replacement of Asp99 with leucine results in a loss of conformational stability of 3.5-4.4 kcal/mol, suggesting that the hydrogen bonds of Tyr14, Tyr55, and Asp99 contribute significantly to stability. While decreasing the stability by about 6.5-7.9 kcal/mol, the Y55F/D99L or Y30F/D99L double mutation also reduced activity significantly, exhibiting a synergistic effect on k(cat) relative to the respective single mutations. These results indicate that the hydrogen-bond network is important for both stability and function. Additionally, they suggest that Tyr14 cannot function efficiently alone without additional support from the hydrogen bonds of Tyr55 and Asp99. The crystal structure of Y55F as determined at 1.9 A resolution shows that Tyr14 OH undergoes an alteration in orientation to form a new hydrogen bond with Tyr30. This observation supports the role of Tyr55 OH in positioning Tyr14 properly to optimize the hydrogen bond between Tyr14 and C3-O of the steroid substrate. No significant structural changes were observed in the crystal structures of Y30F and Y30F/Y55F, which allowed us to estimate approximately the interaction energies mediated by the hydrogen bonds Tyr30...Tyr55 and Tyr14...Tyr55. Taken together, our results demonstrate that the hydrogen-bond network provides the structural support that is needed for the enzyme to maintain the active-site geometry optimized for both function and stability.  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