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.
Asp residues are significantly under represented in beta-sheet regions of proteins, especially in the middle of beta-strands, as found by a number of studies using statistical, modeling, or experimental methods. To further understand the reasons for this under representation of Asp, we prepared and analyzed mutants of a beta-domain. Two Gln residues of the immunoglobulin light-chain variable domain (V(L)) of protein Len were replaced with Asp, and then the effects of these changes on protein stability and protein structure were studied. The replacement of Q38D, located at the end of a beta-strand, and that of Q89D, located in the middle of a beta-strand, reduced the stability of the parent immunoglobulin V(L) domain by 2.0 kcal/mol and 5.3 kcal/mol, respectively. Because the Q89D mutant of the wild-type V(L)-Len domain was too unstable to be expressed as a soluble protein, we prepared the Q89D mutant in a triple mutant background, V(L)-Len M4L/Y27dD/T94H, which was 4.2 kcal/mol more stable than the wild-type V(L)-Len domain. The structures of mutants V(L)-Len Q38D and V(L)-Len Q89D/M4L/Y27dD/T94H were determined by X-ray diffraction at 1.6 A resolution. We found no major perturbances in the structures of these Q-->D mutant proteins relative to structures of the parent proteins. The observed stability changes have to be accounted for by cumulative effects of the following several factors: (1) by changes in main-chain dihedral angles and in side-chain rotomers, (2) by close contacts between some atoms, and, most significantly, (3) by the unfavorable electrostatic interactions between the Asp side chain and the carbonyls of the main chain. We show that the Asn side chain, which is of similar size but neutral, is less destabilizing. The detrimental effect of Asp within a beta-sheet of an immunoglobulin-type domain can have very serious consequences. A somatic mutation of a beta-strand residue to Asp could prevent the expression of the domain both in vitro and in vivo, or it could contribute to the pathogenic potential of the protein in vivo.  相似文献   

2.
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   

3.
Redzic JS  Bowler BE 《Biochemistry》2005,44(8):2900-2908
Cooperativity mediated through hydrogen bond networks in yeast iso-1-cytochrome c was studied using a thermodynamic triple mutant cycle. Three known stabilizing mutations, Asn 26 to His, Asn 52 to Ile, and Tyr 67 to Phe, were used to construct the triple mutant cycle. The side chain of His 26, a wild-type residue, forms two hydrogen bonds that bridge two substructures of the wild-type protein, and Tyr 67 and Asn 52 are part of an extensive buried hydrogen bond network. The stabilities of all variants in the triple mutant cycle were determined by guanidine hydrochloride denaturation methods and used to determine the pairwise, Delta(2)G(int), and triple interaction energies. His 26 and Ile 52 interact cooperatively (Delta(2)G(int) is 1-2 kcal/mol), whereas the two other pairs of mutations interact anticooperatively (Delta(2)G(int) is -0.5 to -1.5 kcal/mol). Previously reported structural data for iso-1-cytochrome c variants containing these mutations show that changes in the strength of the His 26 to Glu 44 hydrogen bond, apparently caused by changes in main chain dynamics, provide a mechanism for the long distance (His 26 to Phe 67 and His 26 to Ile 52) propagation of pairwise interaction energies. Opposing changes in the strength of the His 26 to Glu 44 hydrogen bond caused by the N52I and Y67F mutations generate a negative triple interaction energy (-0.9 +/-0.7 kcal/mol) that combined with cancellation of cooperative and anticooperative pairwise interactions produce apparent additivity for the stabilizing effects of the single mutations in the triple mutant variant.  相似文献   

4.
The goal of this article is to summarize what has been learned about the major forces stabilizing proteins since the late 1980s when site-directed mutagenesis became possible. The following conclusions are derived from experimental studies of hydrophobic and hydrogen bonding variants. (1) Based on studies of 138 hydrophobic interaction variants in 11 proteins, burying a –CH2− group on folding contributes 1.1 ± 0.5 kcal/mol to protein stability. (2) The burial of non-polar side chains contributes to protein stability in two ways: first, a term that depends on the removal of the side chains from water and, more importantly, the enhanced London dispersion forces that result from the tight packing in the protein interior. (3) Based on studies of 151 hydrogen bonding variants in 15 proteins, forming a hydrogen bond on folding contributes 1.1 ± 0.8 kcal/mol to protein stability. (4) The contribution of hydrogen bonds to protein stability is strongly context dependent. (5) Hydrogen bonds by side chains and peptide groups make similar contributions to protein stability. (6) Polar group burial can make a favorable contribution to protein stability even if the polar group is not hydrogen bonded. (7) Hydrophobic interactions and hydrogen bonds both make large contributions to protein stability.  相似文献   

5.
Cornish PV  Giedroc DP 《Biochemistry》2006,45(37):11162-11171
A 28-nucleotide mRNA pseudoknot that overlaps the P1 and P2 genes of sugarcane yellow leaf virus (ScYLV) stimulates -1 ribosomal frameshifting. The in vitro frameshifting efficiency is decreased >or=8-fold upon substitution of the 3'-most loop 2 nucleotide (C27) with adenosine, which accepts a hydrogen bond from the 2'-OH group of C14 in stem S1. The solution structures of the wild-type (WT) and C27A ScYLV RNA pseudoknots show that while the RNAs adopt virtually identical overall structures, there are significant structural differences at the helical junctions of the two RNAs. Specifically, C8(+) in loop L1 in the C8(+).(G12.C28) L1-S2 major groove base triple is displaced by approximately 2.3 A relative to the accepting stem 2 base pair (G12.C28) in the C27A RNA. Here, we use a double mutant cycle approach to analyze the pairwise coupling of the C8(+).(G12.C28)...C27.(C14-G7) and ...A27.(C14-G7) hydrogen bonds in the WT and C27A ScYLV RNAs, respectively, and compare these findings with previous results from the beet western yellows virus (BWYV) RNA. We find that the pairwise coupling free energy (delta(AB)(i)) is favorable for the WT RNA (-0.7 +/- 0.1 kcal/mol), thus revealing that formation of these two hydrogen bonds is positively cooperative. In contrast, delta(AB)(i) is 0.9 +/- 0.4 kcal/mol for the poorly functional C27A ScYLV RNA, indicative of nonadditive hydrogen bond formation. These results reveal that cooperative hydrogen bond formation across the helical stem junction in H-type pseudoknots correlates with enhanced frameshift stimulation by luteoviral mRNA pseudoknots.  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.
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   

8.
The aim of this study was to gain a better understanding of the contribution of hydrogen bonds by tyrosine -OH groups to protein stability. The amino acid sequences of RNases Sa and Sa3 are 69 % identical and each contains eight Tyr residues with seven at equivalent structural positions. We have measured the stability of the 16 tyrosine to phenylalanine mutants. For two equivalent mutants, the stability increases by 0.3 kcal/mol (RNase Sa Y30F) and 0.5 kcal/mol (RNase Sa3 Y33F) (1 kcal=4.184 kJ). For all of the other mutants, the stability decreases with the greatest decrease being 3.6 kcal/mol for RNase Sa Y52F. Seven of the 16 tyrosine residues form intramolecular hydrogen bonds and the average decrease in stability for these is 2.0(+/-1.0) kcal/mol. For the nine tyrosine residues that do not form intramolecular hydrogen bonds, the average decrease in stability is 0.4(+/-0.6) kcal/mol. Thus, most tyrosine -OH groups contribute favorably to protein stability even if they do not form intramolecular hydrogen bonds. Generally, the stability changes for equivalent positions in the two proteins are remarkably similar. Crystal structures were determined for two of the tyrosine to phenylalanine mutants of RNase Sa: Y80F (1.2 A), and Y86F (1.7 A). The structures are very similar to that of wild-type RNase Sa, and the hydrogen bonding partners of the tyrosine residues always form intermolecular hydrogen bonds to water in the mutants. These results provide further evidence that the hydrogen bonding and van der Waals interactions of polar groups in the tightly packed interior of folded proteins are more favorable than similar interactions with water in the unfolded protein, and that polar group burial makes a substantial contribution to protein stability.  相似文献   

9.
A class of temperature-sensitive (ts) mutants of T4 lysozyme with reduced activity at 30 degrees C and no activity at 43 degrees C has been selected. These mutants, designated "tight" ts mutants, differ from most other T4 lysozyme mutants that are active at 43 degrees C, but only manifest their ts lesion by a reduced halo size around phage plaques after exposure of the growth plates to chloroform vapors. For example, in the series of T4 lysozyme mutants at position 157, the original randomly selected mutant, T1571, is the least stable of the series, yet, apart from the halo assay and subsequent in vitro protein stability measurements, this mutant is indistinguishable from wild type (WT) even at 43 degrees C. Two mutants were identified: L91P and L66P. Both insert proline residues into alpha-helical regions of the WT protein structure. The stabilities (delta delta G) as determined by urea denaturation are 8.2 kcal/mol for L91P and 7.1 kcal/mol for L66P. CD spectra indicate that no major conformational changes have occurred in the mutant structures. The structures of the mutants were modeled with a 40-ps molecular dynamics simulation using explicit solvent. For L91P, the reduction of stability appears to be due to an unsatisfied hydrogen bond in the alpha-helix and to a new buried cavity. For L66P, the reduction of stability appears to be due to a disruption of the interdomain alpha-helix, at least two unsatisfied hydrogen bonds, and a newly formed solvent-filled pocket that protrudes into the hydrophobic core, possibly reducing the stabilizing contribution of a partially buried intrachain salt bridge.  相似文献   

10.
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   

11.
Our goal was to gain a better understanding of the contribution of hydrophobic interactions to protein stability. We measured the change in conformational stability, Δ(ΔG), for hydrophobic mutants of four proteins: villin headpiece subdomain (VHP) with 36 residues, a surface protein from Borrelia burgdorferi (VlsE) with 341 residues, and two proteins previously studied in our laboratory, ribonucleases Sa and T1. We compared our results with those of previous studies and reached the following conclusions: (1) Hydrophobic interactions contribute less to the stability of a small protein, VHP (0.6 ± 0.3 kcal/mol per -CH2- group), than to the stability of a large protein, VlsE (1.6 ± 0.3 kcal/mol per -CH2- group). (2) Hydrophobic interactions make the major contribution to the stability of VHP (40 kcal/mol) and the major contributors are (in kilocalories per mole) Phe18 (3.9), Met13 (3.1), Phe7 (2.9), Phe11 (2.7), and Leu21 (2.7). (3) Based on the Δ(ΔG) values for 148 hydrophobic mutants in 13 proteins, burying a -CH2- group on folding contributes, on average, 1.1 ± 0.5 kcal/mol to protein stability. (4) The experimental Δ(ΔG) values for aliphatic side chains (Ala, Val, Ile, and Leu) are in good agreement with their ΔGtr values from water to cyclohexane. (5) For 22 proteins with 36 to 534 residues, hydrophobic interactions contribute 60 ± 4% and hydrogen bonds contribute 40 ± 4% to protein stability. (6) Conformational entropy contributes about 2.4 kcal/mol per residue to protein instability. The globular conformation of proteins is stabilized predominantly by hydrophobic interactions.  相似文献   

12.
Mottamal M  Lazaridis T 《Biochemistry》2005,44(5):1607-1613
Structural analyses of membrane proteins reveal a large number of C(alpha)-H...O contacts between transmembrane helices, presumed to be hydrogen bonds. Recent experiments produced conflicting results for the contribution of such hydrogen bonds to membrane protein stability. An FTIR study estimated an energy of -0.88 kcal/mol for the G79-C(alpha)-H...I76-O hydrogen bond in glycophorin A, whereas a mutagenesis study showed that the A51-C(alpha)-H...T24-O(gamma) hydrogen bond does not stabilize bacteriorhodopsin. Here, we reconcile these results using molecular mechanics calculations and an implicit membrane model (IMM1). With explicit hydrogen atoms, the potential energy of the G79-C(alpha)-H...I76-O interaction in GpA ranges from -0.54 to -0.9 kcal/mol and its contribution to stability (effective energy) from -0.49 to -0.83 kcal/mol, depending on the structural model used. The average values of these quantities in GpA-like motifs are similar. In bR, the corresponding numbers for the A51-C(alpha)-H...T24-O(gamma) interaction are +0.15 and +0.32 kcal/mol. The difference results from the different arrangement of the interacting groups and specifically the position of the acceptor with respect to the C(alpha) and N atoms. This conclusion likely applies to soluble proteins as well.  相似文献   

13.
The contribution of the Ser45 hydrogen bond to biotin binding activation and equilibrium thermodynamics was investigated by biophysical and X-ray crystallographic studies. The S45A mutant exhibits a 1,700-fold greater dissociation rate and 907-fold lower equilibrium affinity for biotin relative to wild-type streptavidin at 37 degrees C, indicating a crucial role in binding energetics. The crystal structure of the biotin-bound mutant reveals only small changes from the wild-type bound structure, and the remaining hydrogen bonds to biotin retain approximately the same lengths. No additional water molecules are observed to replace the missing hydroxyl, in contrast to the previously studied D128A mutant. The equilibrium deltaG degrees, deltaH degrees, deltaS degrees, deltaC degrees(p), and activation deltaG++ of S45A at 37 degrees C are 13.7+/-0.1 kcal/mol, -21.1+/-0.5 kcal/mol, -23.7+/-1.8 cal/mol K, -223+/-12 cal/mol K, and 20.0+/-2.5 kcal/mol, respectively. Eyring analysis of the large temperature dependence of the S45A off-rate resolves the deltaH++ and deltaS++ of dissociation, 25.8+/-1.2 kcal/mol and 18.7+/-4.3 cal/mol K. The large increases of deltaH++ and deltaS++ in the mutant, relative to wild-type, indicate that Ser45 could form a hydrogen bond with biotin in the wild-type dissociation transition state, enthalpically stabilizing it, and constraining the transition state entropically. The postulated existence of a Ser45-mediated hydrogen bond in the wild-type streptavidin transition state is consistent with potential of mean force simulations of the dissociation pathway and with molecular dynamics simulations of biotin pullout, where Ser45 is seen to form a hydrogen bond with the ureido oxygen as biotin slips past this residue after breaking the native hydrogen bonds.  相似文献   

14.
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   

15.
Stabilization of collagen fibrils by hydroxyproline   总被引:1,自引:0,他引:1  
G Némethy  H A Scheraga 《Biochemistry》1986,25(11):3184-3188
The substitution of hydroxyproline for proline in position Y of the repeating Gly-X-Y tripeptide sequence of collagen-like poly(tripeptide)s (i.e., in the position in which Hyp occurs naturally) is predicted to enhance the stability of aggregates of triple helices, while the substitution of Hyp in position X (where no Hyp occurs naturally) is predicted to decrease the stability of aggregates. Earlier conformational energy computations have indicated that two triple helices composed of poly(Gly-Pro-Pro) polypeptide chains pack preferentially with a nearly parallel orientation of the helix axes [Nemethy, G., & Scheraga, H.A. (1984) Biopolymers 23, 2781-2799]. Conformational energy computations reported here indicate that the same packing arrangement is preferred for the packing of two poly(Gly-Pro-Hyp) triple helices. The OH groups of the Hyp residues can be accommodated in the space between the two packed triple helices without any steric hindrance. They actually contribute about 1.9 kcal/mol per Gly-Pro-Hyp tripeptide to the packing energy, as a result of the formation of weak hydrogen bonds and other favorable noncovalent interatomic interactions. On the other hand, the substitution of Hyp in position X weakens the packing by about 1.7 kcal/mol per Gly-Hyp-Pro tripeptide. Numerous published experimental studies have established that Hyp in position Y stabilizes an isolated triple helix relative to dissociated random coils, while Hyp in position X has the opposite effect. We propose that Hyp in position Y also enhances the stability of the assembly of collagen into microfibrils while, in position X, it decreases this stability.  相似文献   

16.
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   

17.
Hydrogen bonding stabilizes globular proteins.   总被引:8,自引:1,他引:7       下载免费PDF全文
It is clear that intramolecular hydrogen bonds are essential to the structure and stability of globular proteins. It is not clear, however, whether they make a net favorable contribution to this stability. Experimental and theoretical studies are at odds over this important question. Measurements of the change in conformational stability, delta (delta G), for the mutation of a hydrogen bonded residue to one incapable of hydrogen bonding suggest a stabilization of 1.0 kcal/mol per hydrogen bond. If the delta (delta G) values are corrected for differences in side-chain hydrophobicity and conformational entropy, then the estimated stabilization becomes 2.2 kcal/mol per hydrogen bond. These and other experimental studies discussed here are consistent and compelling: hydrogen bonding stabilizes globular proteins.  相似文献   

18.
Carbon donor hydrogen bonds are typically weak interactions that contribute less than 2 kcal/mol, and provide only modest stabilization in proteins. One exception is the class of hydrogen bonds donated by heterocyclic side chain carbons. Histidine is capable of particularly strong interactions through the Cε1 and Cδ2 carbons when the imidazole is protonated or bound to metal. Given the frequent occurrence of metal-bound histidines in metalloproteins, we characterized the energies of these interactions through DFT calculations on model compounds. Imidazole-water hydrogen bonding could vary from −11.0 to −17.0 kcal/mol, depending on the metal identity and oxidation state. A geometric search of metalloprotein structures in the PDB identified a number of candidate His C-H···O hydrogen bonds which may be important for folding or function. DFT calculations on model complexes of superoxide reductase show a carbon donor hydrogen bond positioning a water molecule above the active site.  相似文献   

19.
In globular proteins, there are intermolecular hydrogen bonds between protein and water molecules, and between water molecules, which are bound with the proteins, in addition to intramolecular hydrogen bonds. To estimate the contribution of these hydrogen bonds to the conformational stability of a protein, the thermodynamic parameters for denaturation and the crystal structures of five Thr to Val and five Thr to Ala mutant human lysozymes were determined. The denaturation Gibbs energy (DeltaG) of Thr to Val and Thr to Ala mutant proteins was changed from 4.0 to -5.6 kJ/mol and from 1.6 to -6.3 kJ/mol, respectively, compared with that of the wild-type protein. The contribution of hydrogen bonds to the stability (DeltaDeltaG(HB)) of the Thr and other mutant human lysozymes previously reported was extracted from the observed stability changes (DeltaDeltaG) with correction for changes in hydrophobicity and side chain conformational entropy between the wild-type and mutant structures. The estimation of the DeltaDeltaG(HB) values of all mutant proteins after removal of hydrogen bonds, including protein-water hydrogen bonds, indicates a favorable contribution of the intra- and intermolecular hydrogen bonds to the protein stability. The net contribution of an intramolecular hydrogen bond (DeltaG(HB[pp])), an intermolecular one between protein and ordered water molecules (DeltaG(HB[pw])), and an intermolecular one between ordered water molecules (DeltaG(HB[ww])) could be estimated to be 8. 5, 5.2, and 5.0 kJ/mol, respectively, for a 3 A long hydrogen bond. This result shows the different contributions to protein stability of intra- and intermolecular hydrogen bonds. The entropic cost due to the introduction of a water molecule (DeltaG(H)()2(O)) could be also estimated to be about 8 kJ/mol.  相似文献   

20.
Ribonuclease T1 has two disulfide bonds linking cysteine residues 2-10 and 6-103. We have prepared a derivative of ribonuclease T1 in which one disulfide bond is broken and the cysteine residues carboxymethylated, (2-10)-RCM-T1, and three derivatives in which both disulfides are broken and the cysteine residues reduced, R-T1, carboxamidomethylated, RCAM-T1, or carboxymethylated, RCM-T1. The RNA hydrolyzing activity of these proteins has been measured, and urea and thermal denaturation studies have been used to determine conformational stability. The activity, melting temperature, and conformational stability of the proteins are: ribonuclease T1 (100%, 59.3 degrees C, 10.2 kcal/mol), (2-10)-RCM-T1 (86%, 53.3 degrees C, 6.8 kcal/mol), R-T1 (53%, 27.2 degrees C, 3.0 kcal/mol), RCAM-T1 (43%, 21.2 degrees C, 1.5 kcal/mol), and RCM-T1 (35%, 16.6 degrees C, 0.9 kcal/mol). Thus, the conformational stability is decreased by 3.4 kcal/mol when one disulfide bond is broken and by 7.2-9.3 kcal/mol when both disulfide bonds are broken. It is quite remarkable that RNase T1 can fold and function with both disulfide bonds broken and the cysteine residues carboxymethylated. The large decrease in the stability is due mainly to an increase in the conformational entropy of the unfolded protein which results when the constraints of the disulfide bonds on the flexibility are removed. We propose a new equation for predicting the effect of a cross-link on the conformational entropy of a protein: delta Sconf = -2.1 - (3/2)R 1n n, where n is the number of residues between the side chains which are cross-linked. This equation gives much better agreement with experimental results than other forms of this equation which have been used previously.  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