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 31 毫秒
1.
The apparent binding energy for the interaction of the 3-hydroxyl group of chloramphenicol (CM) with the proposed general base (His-195) in chloramphenicol acetyltransferase (CAT) was determined by comparison of the dissociation constants of CM and 3-deoxyCM with CAT. The delta Gapp for this hydrogen bond to the N-3 of the imidazole ring is 1.5 kcal mol-1. Extending the use of modified ligands, in an approach which is complementary to that of site-directed mutagenesis, the binding affinity of each of a family of 3-halo-3-deoxychloramphenicol derivatives was observed to increase in the series F less than Cl less than Br less than I and is dominated by hydrophobic considerations. There is a linear free energy relationship between the dissociation constants for binding to CAT and an empirical hydrophobicity scale derived from reverse-phase HPLC retention times. The existence of such a relationship allows a true estimate of the total energetic contribution of interactions between the 3-hydroxyl group of CM and its contacts at the active site of CAT to be made on the basis of a regression analysis. The calculated value of delta Gbind (2.7 kcal mol-1) must include not only the hydrogen bond but also some favorable van der Waals interactions. The results demonstrate some of the advantages of an analysis of the energetics of ligand binding using modified ligands, in an approach that is formally analogous with and complementary to the use of site-directed mutations.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
J K Hwang  A Warshel 《Biochemistry》1987,26(10):2669-2673
The catalytic free energy and binding free energies of the native and the Asn-155----Thr, Asn-155----Leu, and Asn-155----Ala mutants of subtilisin are calculated by the empirical valence bond method and a free energy perturbation method. Two simple procedures are used; one "mutates" the substrate, and the other "mutates" the enzyme. The calculated changes in free energies (delta delta G not equal to cat and delta delta Gbind) between the mutant and native enzymes are within 1 kcal/mol of the corresponding observed values. This indicates that we are approaching a quantitative structure-function correlation. The calculated changes in catalytic free energies are almost entirely due to the electrostatic interaction between the enzyme-water system and the charges of the reacting system. This supports the idea that the electrostatic free energy associated with the changes of charges of the reacting system is the key factor in enzyme catalysis.  相似文献   

3.
The catalytic motif (YSASK) at the active site of 11beta-hydroxysteroid dehydrogenase type 1 (11beta-HSD1) is conserved across different species. The crystal structures of the human, guinea pig and mouse enzymes have been resolved to help identify the non-conserved residues at the active site. A tyrosine residue (Y177) upstream of the catalytic motif in human 11beta-HSD1 represents the largest difference at the active sites between the human and the rodent enzyme where the corresponding residue is glutamine. Although Y177 was postulated as a potential hydrogen bond donor in substrate binding in crystal structure-based modeling, no experimental evidence is available to support this notion. Here, we report that Y177 is not a hydrogen bond donor in substrate binding because removal of the hydroxyl group from its side chain by mutagenesis (Y177F) did not significantly change the Km value for cortisone. However, removal of the hydrophobic side chain by changing tyrosine to alanine (Y177A) or substitution with a hydrophilic side chain by changing tyrosine to glutamine (Y177Q) increased Km values for cortisone. These data suggest that Y177 is involved in substrate binding through its hydrophobic side chain but not by hydrogen bonding. In addition, the three mutations had little effect on the binding of the rodent substrate 11-dehydrocorticosterone, suggesting that Y177 does not confer substrate specificity. However, the same mutations reduced the affinity of the licorice derived 11beta-HSD1 inhibitor glycyrrhetinic acid by about 6- to 10-fold. Interestingly, the affinity of carbenoxolone, the hemisuccinate ester of glycyrrhetinic acid with a similar potency against the wildtype enzyme, was not drastically affected by the same mutations at Y177. These data suggest that Y177 has a unique role in inhibitor binding. Molecular modeling with glycyrrhetinic acid led to findings consistent with the experimental data and provided potential interaction mechanisms. Our data suggest that Y177 plays an important role in both substrate and inhibitor binding but it is unlikely a hydrogen bond donor for the substrate.  相似文献   

4.
Stabilization of an oxyanion transition state is important to catalysis of peptide bond hydrolysis in all proteases. For subtilisin BPN', a bacterial serine protease, structural data suggest that two hydrogen bonds stabilize the tetrahedral-like oxyanion intermediate: one from the main chain NH of Ser221 and another from the side chain NH2 of Asn155. Molecular dynamic studies (Rao, S., N., Singh, U., C. Bush, P. A., and Kollman, P. A. (1987) Nature 328, 551-554) have indicated the gamma-hydroxyl of Thr220 may be a third hydrogen bond donor even though it is 4A away in the static x-ray structure. We have probed the role of Thr220 by replacing it with serine, cysteine, valine, or alanine by site-directed mutagenesis. These substitutions were intended to alter the size and hydrogen bonding ability of residue 220. Removal of the gamma-hydroxyl group reduced the transition state stabilization energy (delta delta GT) by 1.8-2.1 kcal/mol depending upon the substitution. By comparison, removal of the gamma-methyl group in the Thr220 to serine mutation only decreased delta GT by 0.5 kcal/mol. The gamma-hydroxyl of Thr220 is most important for catalysis, not substrate binding, because virtually all of the effects were on kcat and not KM. The role of the Thr220 hydroxyl is functionally independent from the amide NH2 of Asn155 because the free energy effects of double alanine mutants at these two positions are additive. These data indicate that a distal hydrogen bond donor, namely the hydroxyl of Thr220, plays a functionally important role in stabilizing the oxyanion transition state in subtilisin which is independent of Asn155.  相似文献   

5.
We present a calculation of the relative changes in binding free energy between the complex of ribonuclease T1 (RNase Tr) with its inhibitor 2'-guanosine monophosphate (2'GMP) and that of RNase T1-2'-adenosine monophosphate (2'AMP) by means of a thermodynamic perturbation method implemented with molecular dynamics. Using the available crystal structure of the RNase T1-2'GMP complex, the structure of the RNase T1-2'AMP complex was obtained as a final structure of the perturbation calculation. The calculated difference in the free energy of binding (delta delta Gbind) was 2.76 kcal/mol. This compares well with the experimental value of 3.07 kcal/mol. The encouraging agreement in delta delta Gbind suggests that the interactions of inhibitors with the enzyme are reasonably represented. Energy component analyses of the two complexes reveal that the active site of RNase T1 electrostatically stabilizes the binding of 2'GMP more than that of 2'AMP by 44 kcal/mol, while the van der Waals' interactions are similar in the two complexes. The analyses suggest that the mutation from Glu46 to Gln may lead to a preference of RNase T1 for adenine in contrast to the guanine preference of the wild-type enzyme. Although the molecular dynamics equilibration moves the atoms of the RNase T1-2'GMP system about 0.9 A from their X-ray positions and the mutation of the G to A in the active site increases the deviation from the X-ray structure, the mutation of the A back to G reduces the deviation. This and the agreement found for delta delta Gbind suggest that the molecular dynamics/free energy perturbation method will be useful for both energetic and structural analysis of protein-ligand interactions.  相似文献   

6.
T N Wells  A R Fersht 《Biochemistry》1986,25(8):1881-1886
The utilization of enzyme-substrate binding energy in catalysis has been investigated by experiments on mutant tyrosyl-tRNA synthetases that have been generated by site-directed mutagenesis. The mutants are poorer enzymes because they lack side chains that form hydrogen bonds with ATP and tyrosine during stages of the reaction. The hydrogen bonds are not directly involved in the chemical processes but are at some distance from the seat of reaction. The free energy profiles for the formation of enzyme-bound tyrosyl adenylate and the equilibria between the substrates and products were determined from a combination of pre-steady-state kinetics and equilibrium binding methods. By comparison of the profile of each mutant with wild-type enzyme, a picture is built up of how the course of reaction is affected by the influence of each side chain on the energies of the complexes of the enzyme with substrates, transition states, and intermediates (tyrosyl adenylate). As the activation reaction proceeds, the apparent binding energies of certain side chains with the tyrosine and nucleotide moieties increase, being weakest in the enzyme-substrate complex, stronger in the transition state, and strongest in the enzyme-intermediate complex. Most marked is the interaction of Cys-35 with the 3'-hydroxyl of the ribose. Removal of the side chain of Cys-35 leads to no change in the dissociation constant of ATP but causes a 10-fold lowering of the catalytic rate constant. It contributes no net apparent binding energy in the E X Tyr X ATP complex and stabilizes the transition state by 1.2 kcal/mol and the E X Tyr-AMP complex by 1.6 kcal/mol.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The imidazole of His-195 plays an essential role in the proposed general base mechanism of chloramphenicol acetyltransferase (CAT). The structure of the binary complex of CATIII and chloramphenicol suggests that two unusual interactions might determine the conformation of the side chain of His-195: (i) an intraresidue hydrogen bond between its main chain carbonyl and the protonated N delta 1 of the imidazole ring and (ii) face-to-face van der Waals contact between the His-195 imidazole group and the aromatic side chain of Tyr-25. Tyr-25 also makes a hydrogen bond, via its phenolic hydroxyl, to the carbonyl oxygen of the substrate chloramphenicol. Replacement of Tyr-25 of CATIII by phenylalanine results in a modest increase in the Km for chloramphenicol (from 11.6 to 14.6 microM) and a 2-fold fall in kcat (599 to 258 s-1), indicative of a free energy contribution to transition state binding of 0.6 kcal mol-1 for the hydrogen bond between Tyr-25 and chloramphenicol. In contrast, substitution of Tyr-25 by alanine yields an enzyme that is dramatically impaired in its ability to bind chloramphenicol (Km = 173 microM). As kcat for Ala-25 CAT is also reduced (130 s-1), the loss of the aryl group results in a 69-fold decrease in kcat/Km, corresponding to a free energy contribution to binding and catalysis of 2.5 kcal mol-1. In addition to the loss of the hydrogen bond between Tyr-25 and chloramphenicol, the loss of substrate affinity in Ala-25 CAT may be a direct consequence of reduced hydrophobicity of the chloramphenicol-binding site and/or the loss of critical constraints on the precise conformation of the catalytic imidazole. However, as with wild type CAT, inactivation of Ala-25 CAT by the affinity reagent 3-(bromoacetyl) chloramphenicol is accompanied by modification solely at N epsilon 2 of His-195. Hence, the results demonstrate that tautomeric stabilization of the imidazole ring persists in the absence of van der Waals interactions with the side chain of Tyr-25, probably as a consequence of hydrogen bonding between the protonated N delta 1 and the carbonyl oxygen of His-195.  相似文献   

8.
Aktas DF  Cook PF 《Biochemistry》2008,47(8):2539-2546
The mitochondrial NAD-malic enzyme catalyzes the oxidative decarboxylation of malate to pyruvate and CO2. The role of the dinucleotide substrate in oxidative decarboxylation is probed in this study using site-directed mutagenesis to change key residues that line the dinucleotide binding site. Mutant enzymes were characterized using initial rate kinetics, and isotope effects were used to obtain information on the contribution of these residues to binding energy and catalysis. Results obtained for the N479 mutant enzymes indicate that the hydrogen bond donated by N479 to the carboxamide side chain of the nicotinamide ring is important for proper orientation in the hydride transfer step. The stepwise oxidative decarboxylation mechanism observed for the wt enzyme changed to a concerted one, which is totally rate limiting, for the N479Q mutant enzyme. In this case, it is likely that the longer glutamine side chain causes reorientation of malate such that it binds in a conformation that is optimal for concerted oxidative decarboxylation. Converting N479 to the shorter serine side chain gives very similar values of KNAD, Kmalate, and isotope effects relative to wt, but V/Et is decreased 2 000-fold. Data suggest an increased freedom of rotation, resulting in nonproductively bound cofactor. Changes were also made to two residues, S433 and N434, which interact with the nicotinamide ribose of NAD. In addition, N434 donates a hydrogen bond to the beta-carboxylate of malate. The KNAD for the S433A mutant enzyme increased by 80-fold, indicating that this residue provides significant binding affinity for the dinucleotide. With N434A, the interaction of the residue with malate is lost, causing the malate to reorient itself, leading to a slower decarboxylation step. The longer glutamine and methionine side chains stick into the active site and cause a change in the position of malate and/or NAD resulting in more than a 104-fold decrease in V/Et for these mutant enzymes. Overall, data indicate that subtle changes in the orientation of the cofactor and substrate dramatically influence the reaction rate.  相似文献   

9.
The nicotinic acetylcholine receptor and related Cys-loop receptors are ligand-gated ion channels that mediate fast synaptic transmission throughout the central and peripheral nervous system. A highly conserved aspartate residue (D89) that is near the agonist binding site but does not directly contact the ligand plays a critical part in receptor function. Here we probe the role of D89 using unnatural amino acid mutagenesis coupled with electrophysiology. Homology modeling implicates several hydrogen bonds involving D89. We find that no single hydrogen bond is essential to proper receptor function. Apparently, the side chain of D89 establishes a redundant network of hydrogen bonds; these bonds preorganize the agonist binding site by positioning a critical tryptophan residue that directly contacts the ligand. Earlier studies of the D89N mutant led to the proposal that a negative charge at this position is essential for receptor function. However, we find that receptors with neutral side chains at position 89 can function well, if the side chain is less perturbing than the amide of asparagine (nitro or keto groups allow function) or if a compensating backbone mutation is introduced to relieve unfavorable electrostatics.  相似文献   

10.
The major histocompatibility complex (MHC) class II binding requirements for solvent-exposed peptide residues were systematically studied using amino acid and peptoid substitutions. In a peptoid residue, the side chain is present on the backbone nitrogen atom as opposed to the alpha-carbon atom in an amino acid residue. To investigate the effect of this side chain shifting on MHC binding, three amino acids in the central part of the peptide sticking out of the binding groove were replaced by corresponding peptoid residues. Two peptoid-peptide hybrids showed large affinity decreases in the MHC-peptide binding assay. To investigate this affinity loss, the individual contributions to MHC binding affinity of the side chain (position), the putative hydrogen bond, and the flexibility were dissected. We conclude that the side chain position as well as the backbone nitrogen atom hydrogen bonding features of solvent-exposed residues in the peptide can be important for MHC binding affinity.  相似文献   

11.
This article presents a molecular dynamics (MD) study of the cdk2 enzyme and its two complexes with the inhibitors isopentenyladenine and roscovitine using the Cornell et al. force field from the AMBER software package. The results show that inserting an inhibitor into the enzyme active site does not considerably change enzyme structure but it seemingly changes the distribution of internal motions. The inhibitor causes differences in the domain motions in free cdk2 and in its complexes. It was found out that repulsion of roscovitine N9 substituent causes conformational change on Lys 33 side chain. Isopentenyladenine forms with Lys 33 side chain terminal amino group a hydrogen bond. It implies that the cavity, where N9 substituent of roscovitine is buried, can adopt larger substituent due to Lys 33 side chain flexibility. The composition of electrostatic and van der Waals interactions between the inhibitor and the enzyme were also calculated along both cdk2/inhibitor MD trajectories together with MM-PB/GBSA analysis. These results show that isopentenyladenine-like inhibitors could be more effective after modifications leading to an increase in their van der Waals contact with the enzyme. We suggest that a way leading to better inhibitors occupying isopentenyladenine binding mode could be: to keep N9 and N7 purine positions free, to keep 3,3-dimethylallylamino group at C6 position, and to add, e.g., benzylamino group at C2 position. The results support the idea that the isopentenyladenine binding mode can be used for cdk2 inhibitors design and that all possibilities to improve this binding mode were not uncovered yet.  相似文献   

12.
Arginine residues are commonly found in the active sites of enzymes catalyzing phosphoryl transfer reactions. Numerous site-directed mutagenesis experiments establish the importance of these residues for efficient catalysis, but their role in catalysis is not clear. To examine the role of arginine residues in the phosphoryl transfer reaction, we have measured the consequences of mutations to arginine 166 in Escherichia coli alkaline phosphatase on hydrolysis of ethyl phosphate, on individual reaction steps in the hydrolysis of the covalent enzyme-phosphoryl intermediate, and on thio substitution effects. The results show that the role of the arginine side chain extends beyond its positive charge, as the Arg166Lys mutant is as compromised in activity as Arg166Ser. Through measurement of individual reaction steps, we construct a free energy profile for the hydrolysis of the enzyme-phosphate intermediate. This analysis indicates that the arginine side chain strengthens binding by approximately 3 kcal/mol and provides an additional 1-2 kcal/mol stabilization of the chemical transition state. A 2.1 A X-ray diffraction structure of Arg166Ser AP is presented, which shows little difference in enzyme structure compared to the wild-type enzyme but shows a significant reorientation of the bound phosphate. Altogether, these results support a model in which the arginine contributes to catalysis through binding interactions and through additional transition state stabilization that may arise from complementarity of the guanidinum group to the geometry of the trigonal bipyramidal transition state.  相似文献   

13.
The contribution of His64 to the function and stability of tissue plasminogen activator (t-PA) kringle-2 domain (His244 in t-PA numbering) has been studied by using microcalorimetric methods to compare the ligand binding and thermal denaturation behavior of wild-type kringle-2 and mutants having His64 replaced with Tyr or Phe. This site was examined because modeling studies suggested that the His64 side chain could play an important role in ligand binding by forming an ion-pair with the carboxylate of the ligand, L-lysine. Kringle-2 domains were expressed by secretion of the 174-263 portion of t-PA in E. coli and purified as previously described for the wild-type domain. Both mutant proteins retain affinity for L-lysine, although reduced three- to four-fold relative to wild-type, demonstrating that His64 does not interact with the ligand carboxylate through an ion-pair interaction or by hydrogen bonding. The H64Y substitution does result in an altered specificity of the lysine binding site with the mutant domain having greatest affinity for a ligand of 6.8 A chain length, whereas the wild-type domain prefers an 8.8 A long ligand. For both wild-type and mutant, the binding of the optimal chain length ligand is dominated by enthalpic effects (delta H = -6,000 to -7,000 cal/mol) and T delta S accounts for less than 15% of delta G. In addition, the H64Y mutant differs from wild-type in the effect of ligand alpha-amino group modification on binding affinity. Based on examination of the x-ray structure recently determined for wild-type kringle-2, the specificity changes accompanying the H64Y substitution probably result from changes in side chain interactions in the lysine binding site. Thermal denaturation experiments show that the H64Y mutant is also more stable than the wild-type protein with the difference in stabilization free energy (delta delta G) equal to 2.7 kcal/mol at 25 degrees C and pH 3. The increased stability of the mutant appears to be related to the difference in hydrophobicity between His and Tyr.  相似文献   

14.
Canyuk B  Focia PJ  Eakin AE 《Biochemistry》2001,40(9):2754-2765
The role of an invariant aspartic acid (Asp137) in hypoxanthine phosphoribosyltransferases (HPRTs) was examined by site-directed and saturation mutagenesis, functional analysis, and X-ray crystallography using the HPRT from Trypanosoma cruzi. Alanine substitution (D137A) resulted in a 30-fold decrease of k(cat), suggesting that Asp137 participates in catalysis. Saturation mutagenesis was used to generate a library of mutant HPRTs with random substitutions at position 137, and active enzymes were identified by complementation of a bacterial purine auxotroph. Functional analyses of the mutants, including determination of steady-state kinetic parameters and pH-rate dependence, indicate that glutamic acid or glutamine can replace the wild-type aspartate. However, the catalytic efficiency and pH-rate profile for the structural isosteric mutant, D137N, were similar to the D137A mutant. Crystal structures of four of the mutant enzymes were determined in ternary complex with substrate ligands. Structures of the D137E and D137Q mutants reveal potential hydrogen bonds, utilizing several bound water molecules in addition to protein atoms, that position these side chains within hydrogen bond distance of the bound purine analogue, similar in position to the aspartate in the wild-type structure. The crystal structure of the D137N mutant demonstrates that the Asn137 side chain does not form interactions with the purine substrate but instead forms novel interactions that cause the side chain to adopt a nonfunctional rotamer. The results from these structural and functional analyses demonstrate that HPRTs do not require a general base at position 137 for catalysis. Instead, hydrogen bonding sufficiently stabilizes the developing partial positive charge at the N7-atom of the purine substrate in the transition-state to promote catalysis.  相似文献   

15.
M D Fothergill  A R Fersht 《Biochemistry》1991,30(21):5157-5164
The crystal structures of two mutant tyrosyl-tRNA synthetases (TyrTS) are reported to test predictions from kinetic data about structural perturbations and also to aid in the interpretation of apparent strengths of hydrogen bonds measured by protein engineering. The enzyme-tyrosine and enzyme-tyrosyl adenylate complexes of the mutant, TyrTS(Cys----Gly-35), have been determined at 2.5- and 2.7-A resolution, respectively. Residue Cys-35 is in the ribose binding site. Small rearrangements in structure are seen in the enzyme-tyrosine complex that are localized around the cavity created by the mutation. The side chain of Thr-51 moves to occupy the cavity, and Ile-52 adopts two significantly populated conformations, one as in the native enzyme and a second unique to the mutant. On binding tyrosyl adenylate, Ile-52 in the mutant crystal structure preferentially occupies the conformation observed in the native structure. The side chain at Thr-51 becomes disordered. The double-mutant test, which was designed to detect interactions between residues, had previously shown a discrepancy of some 0.4 kcal/mol on mutating Cys-35 and Thr-51 separately and together. A crystal structure of a second mutant, delta TyrTS(Tyr----Phe-34), complexed with tyrosine has been determined at 2.7-A resolution. Tyr-34 in wild-type enzyme makes a hydrogen bond with the phenolic oxygen of the bound tyrosine substrate. The mutant crystal structure was solved to discover whether or not a water molecule binds to the substrate instead of the hydroxyl of Tyr-34 as the interpretation of apparent binding energies from site-directed mutagenesis experiments hinges crucially on whether there is access of water to the mutated region.  相似文献   

16.
In order to investigate the function of Asp-327, a bidentate ligand of one of the zinc atoms in Escherichia coli alkaline phosphatase, and the importance of this zinc atom in catalysis, site-specific mutagenesis was used to convert Asp-327 to either asparagine or alanine. The 10(7)-fold decrease in the kcat/Km ratio observed for the Asp-327----Ala enzyme compared to the wild-type enzyme indicates that the side chain of Asp-327 is important for zinc binding at the M1 site. However, only one of the two carboxyl oxygens of Asp-327 is essential for zinc binding, since the Asp-327----Asn enzyme shows approximately the same hydrolysis activity as the wild-type enzyme. The fact that the enzymatic activity of this mutant enzyme shows a dependence on zinc concentration suggests that the other carboxyl oxygen or the negative charge on the side chain of Asp-327 is important in binding of the zinc at the M1 site. However, the zinc hydroxyl must still be appropriately positioned to attack the phosphoserine in the Asp-327----Asn enzyme; therefore, the negative charge and at least one carboxyl oxygen of the side chain are not directly involved in positioning or deprotonating the zinc hydroxyl. 31P NMR studies indicate that the Asp-327----Asn enzyme exhibits transphosphorylation activity at both pH 8.0 and pH 10.0, but at a reduced level compared to the wild-type enzyme. The biphasic production of 2,4-dinitrophenylate in the pre-steady-state kinetics of the mutant enzymes at pH 5.5 suggests that the breaking of the phosphoenzyme covalent complex is rate-limiting for both mutant enzymes. These results suggest that the main function of the zinc atom at the M1 site in catalysis involves decomposition of the phosphoenzyme covalent complex and that it may be important in helping to stabilize the alcohol leaving group.  相似文献   

17.
We examined functional contributions of interdomain contacts within the nicotinic receptor ligand binding site using single channel kinetic analyses, site-directed mutagenesis, and a homology model of the major extracellular region. At the principal face of the binding site, the invariant alphaD89 forms a highly conserved interdomain contact near alphaT148, alphaW149, and alphaT150. Patch-clamp recordings show that the mutation alphaD89N markedly slows acetylcholine (ACh) binding to receptors in the resting closed state, but does not affect rates of channel opening and closing. Neither alphaT148L, alphaT150A, nor mutations at both positions substantially affects the kinetics of receptor activation, showing that hydroxyl side chains at these positions are not hydrogen bond donors for the strong acceptor alphaD89. However substituting a negative charge at alphaT148, but not at alphaT150, counteracts the effect of alphaD89N, demonstrating that a negative charge in the region of interdomain contact confers rapid association of ACh. Interpreted within the structural framework of ACh binding protein and a homology model of the receptor ligand binding site, these results implicate main chain amide groups in the domain harboring alphaW149 as principal hydrogen bond donors for alphaD89. The specific effect of alphaD89N on ACh association suggests that interdomain hydrogen bonding positions alphaW149 for optimal interaction with ACh.  相似文献   

18.
T Becker  K Weber    N Johnsson 《The EMBO journal》1990,9(13):4207-4213
Annexin II (p36) interacts with its ligand p11 via the short stretch of 12 amino acids (Ac-S-T-V-H-E-I-L-C-K-L-S-L) situated at the N-terminus. We have now synthesized some 37 tetradecapeptides, which differ from the original p11 binding sequence (Ac1-14) by single amino acid substitutions. The relative affinity of each peptide for p11 was determined by fluorescence spectroscopy using a competitive binding assay. The binding behaviour of the different peptides confirms the model of an amphiphilic alpha-helix induced upon binding to p11. The apparent affinities delta delta Gbind of the mutant peptides revealed that the N-acetyl group of serine 1 and the hydrophobic side chains at positions 3, 6, 7 and 10 contribute most to the binding. The observed destabilization of the complex upon removal of signal methyl groups from the hydrophobic side of the helix is comparable with the destabilization of proteins in which methyl groups have been removed from the inner core. We conclude that upon binding to p11 the hydrophobic side of the amphiphatic alpha-helix becomes fully buried.  相似文献   

19.
The side chain of Gln143, a conserved residue in manganese superoxide dismutase (MnSOD), forms a hydrogen bond with the manganese-bound solvent and is critical in maintaining catalytic activity. The side chains of Tyr34 and Trp123 form hydrogen bonds with the carboxamide of Gln143. We have replaced Tyr34 and Trp123 with Phe in single and double mutants of human MnSOD and measured their catalytic activity by stopped-flow spectrophotometry and pulse radiolysis. The replacements of these side chains inhibited steps in the catalysis as much as 50-fold; in addition, they altered the gating between catalysis and formation of a peroxide complex to yield a more product-inhibited enzyme. The replacement of both Tyr34 and Trp123 in a double mutant showed that these two residues interact cooperatively in maintaining catalytic activity. The crystal structure of Y34F/W123F human MnSOD at 1.95 A resolution suggests that this effect is not related to a conformational change in the side chain of Gln143, which does not change orientation in Y34F/W123F, but rather to more subtle electronic effects due to the loss of hydrogen bonding to the carboxamide side chain of Gln143. Wild-type MnSOD containing Trp123 and Tyr34 has approximately the same thermal stability compared with mutants containing Phe at these positions, suggesting the hydrogen bonds formed by these residues have functional rather than structural roles.  相似文献   

20.
The mode of binding of the substrate analog 2'-deoxy-2'-fluoroguanylyl- (3',5')-cytidine (GfpC) to RNase T1 was determined by computer modelling studies. The results obtained are in good agreement with the observations of 1H-nmr studies. The modes of binding of the substrate analog GfpC and the substrate GpC to the enzyme RNase T1 have been compared. Though the guanine base favours to occupy the same site of the enzyme in both the complexes, significant differences are observed in the local environment around the 2'-substituent group of guanosine ribose moiety. In the RNase T1-GpC complex, the 2'-OH group is in close proximity to the side chain carboxylic acid of Glu58 which leads to the formation of a hydrogen bond. However, in the RNase T1-GfpC complex, 2'-fluorine is positioned away from Glu58 due to electrostatic repulsion and instead forms a hydrogen bond with His40 imidazolium group. The results obtained rule out the possibility of His40 serving as the base group in catalysis as suggested by 1H-nmr studies and further support the primary role assigned to Glu58 as the general base group by earlier computer modelling and the recent site directed mutagenesis studies. This study also implies that the 2'-deoxy-2'-fluoro substrate analog may not serve as a good model for determining the amino acid residue which serves as the general base group in ribonuclease catalysed reactions.  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