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.
Sharrow SD  Novotny MV  Stone MJ 《Biochemistry》2003,42(20):6302-6309
The mouse pheromone 2-sec-butyl-4,5-dihydrothiazole (SBT) binds to an occluded, nonpolar cavity in the mouse major urinary protein-I (MUP-I). The thermodynamics of this interaction have been characterized using isothermal titration calorimetry (ITC). MUP-I-SBT binding is accompanied by a large favorable enthalpy change (DeltaH = -11.2 kcal/mol at 25 degrees C), an unfavorable entropy change (-TDeltaS = 2.8 kcal/mol at 25 degrees C), and a negative heat capacity change [DeltaC(p)() = -165 cal/(mol K)]. Thermodynamic analysis of binding between MUP-I and several 2-alkyl-4,5-dihydrothiazole ligands indicated that the alkyl chain contributes more favorably to the enthalpy and less favorably to the entropy of binding than would be expected on the basis of the hydrophobic desolvation of short-chain alcohols. However, solvent transfer experiments indicated that desolvation of SBT is accompanied by a net unfavorable change in enthalpy (DeltaH = +1.0 kcal/mol) and favorable change in entropy (-TDeltaS = -1.8 kcal/mol). These results are discussed in terms of the possible physical origins of the binding thermodynamics, including (1) hydrophobic desolvation of both the protein and the ligand, (2) formation of a buried water-mediated hydrogen bond network between the protein and ligand, (3) formation of strong van der Waals interactions, and (4) changes in the structure, dynamics, and/or hydration of the protein upon binding.  相似文献   

2.
3.
Chen M  Jiang M  Sun Y  Guo ZF  Guo Z 《Biochemistry》2011,50(26):5893-5904
1,4-Dihydroxy-2-naphthoyl-coenzyme A (DHNA-CoA) synthase, or MenB, catalyzes an intramolecular Claisen condensation involving two oxyanion intermediates in the biosynthetic pathway of menaquinone, an essential respiration electron transporter in many microorganisms. Here we report the finding that the DHNA-CoA product and its analogues bind and inhibit the synthase from Escherichia coli with significant ultraviolet--visible spectral changes, which are similar to the changes induced by deprotonation of the free inhibitors in a basic solution. Dissection of the structure--affinity relationships of the inhibitors identifies the hydroxyl groups at positions 1 (C1-OH) and 4 (C4-OH) of DHNA-CoA or their equivalents as the dominant and minor sites, respectively, for the enzyme--ligand interaction that polarizes or deprotonates the bound ligands to cause the observed spectral changes. In the meantime, spectroscopic studies with active site mutants indicate that C4-OH of the enzyme-bound DHNA-CoA interacts with conserved polar residues Arg-91, Tyr-97, and Tyr-258 likely through a hydrogen bonding network that also includes Ser-161. In addition, site-directed mutation of the conserved Asp-163 to alanine causes a complete loss of the ligand binding ability of the protein, suggesting that the Asp-163 side chain is most likely hydrogen-bonded to C1-OH of DHNA-CoA to provide the dominant polarizing effect. Moreover, this mutation also completely eliminates the enzyme activity, strongly supporting the possibility that the Asp-163 side chain provides a strong stabilizing hydrogen bond to the tetrahedral oxyanion, which takes a position similar to that of C1-OH of the enzyme-bound DHNA-CoA and is the second high-energy intermediate in the intracellular Claisen condensation reaction. Interestingly, both Arg-91 and Tyr-97 are located in a disordered loop forming part of the active site of all available DHNA-CoA synthase structures. Their involvement in the interaction with the small molecule ligands suggests that the disordered loop is folded in interaction with the substrates or reaction intermediates, supporting an induced-fit catalytic mechanism for the enzyme.  相似文献   

4.
The role of protein residues in activating the substrate in the reaction catalyzed by the flavoprotein p-hydroxybenzoate hydroxylase was studied. X-ray crystallography (Schreuder, H. A., Prick, P.A.J., Wieringa, R.K., Vriend, G., Wilson, K.S., Hol, W.G. J., and Drenth, J. (1989) J. Mol. Biol. 208, 679-696) indicates that Tyr-201 and Tyr-385 form a hydrogen bond network with the 4-OH of p-hydroxybenzoate. Therefore, site directed mutants were constructed, converting each of these tyrosines into phenylalanines. Spectral (visible and fluorescence) properties, reduction potentials, and binding constants are very similar to those of wild type, indicating that there are no major structural changes in the mutants. In the absence of substrate, the mutants and wild type exhibit similar pH-dependent changes in the FAD spectrum. However, the enzyme-substrate complex of Tyr-201----Phe lacks an ionization observed in both wild type and Tyr-385----Phe, which preferentially bind the phenolate form of substrates. Tyr-201----Phe shows no preference, indicating that Tyr-201 is required to ionize the substrate. The mutants have less than 6% the activity of the wild type enzyme. The effects on catalysis were studied by stopped flow techniques. Reduction of FAD by NADPH is slower by 10-fold in Tyr-201----Phe and 100-fold in Tyr-385----Phe. When the reduced Tyr-201----Phe-p-hydroxybenzoate complex reacts with oxygen, a long-lived flavin-C(4a)-hydroperoxide is observed, which slowly eliminates H2O2 with very little hydroxylation. Thus, the role of Tyr-201 is to activate the substrate by stabilizing the phenolate. Tyr-385----Phe reacts with oxygen to form 25% oxidized enzyme, and 75% flavin hydroperoxide, which successfully hydroxylates the substrate. This mutant also hydroxylates the product (3, 4-dihydroxybenzoate) to form gallic acid.  相似文献   

5.
Barrick D 《Proteins》2000,39(4):291-308
The trans-substituted histidine to glycine mutant of sperm whale myoglobin (H93G Mb) is used to study energetics of proximal hydrogen bonding, proximal ligand-heme interactions, and coupling to distal ligand binding. Comparison of mono- and dimethylimidazole structural isomers shows that the hydrogen bond between the proximal ligand and the neighboring Ser92 hydroxyl (position F7) is stabilizing. The range of hydrogen bond stabilities measured here for different distal ligand complexes ranges from -0.7 kcal/mol (monomethylimidazole isomers to MbCO) to -4.1 kcal/mol (dimethylimidazole isomers to MbCN). This range of hydrogen bond stabilities, which is similar to that seen in protein mutagenesis unfolding studies, demonstrates the high sensitivity of the hydrogen bond to modest structural perturbations. The degree to which the 2-methyl group destabilizes proximal ligand binding is found to depend inversely on the total electronic spin. For monomethylimidazole proximal ligands, distal ligand binding weakens the proximal hydrogen bond compared to deoxyMb. Surprisingly, this trend is largely reversed for the dimethylimidazole proximal ligands. These results demonstrate strong coupling between the proximal protein matrix and distal ligand binding. These results provide an explanation for the strong avoidance of hydrogen bonding residues at position F7 in hemoglobin sequences.  相似文献   

6.
Here, we examine the thermodynamic penalty arising from burial of a polar group in a hydrophobic pocket that forms part of the binding-site of the major urinary protein (MUP-I). X-ray crystal structures of the complexes of octanol, nonanol and 1,8 octan-diol indicate that these ligands bind with similar orientations in the binding pocket. Each complex is characterised by a bridging water molecule between the hydroxyl group of Tyr120 and the hydroxyl group of each ligand. The additional hydroxyl group of 1,8 octan-diol is thereby forced to reside in a hydrophobic pocket, and isothermal titration calorimetry experiments indicate that this is accompanied by a standard free energy penalty of +21 kJ/mol with respect to octanol and +18 kJ/mol with respect to nonanol. Consideration of the solvation thermodynamics of each ligand enables the "intrinsic" (solute-solute) interaction energy to be determined, which indicates a favourable enthalpic component and an entropic component that is small or zero. These data indicate that the thermodynamic penalty to binding derived from the unfavourable desolvation of 1,8 octan-diol is partially offset by a favourable intrinsic contribution. Quantum chemical calculations suggest that this latter contribution derives from favourable solute-solute dispersion interactions.  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.
Phosphonoacetaldehyde hydrolase (phosphonatase) catalyzes the hydrolytic P-C bond cleavage of phosphonoacetaldehyde (Pald) to form orthophosphate and acetaldehyde. The reaction proceeds via a Schiff-base intermediate formed between Lys-53 and the Pald carbonyl. The x-ray crystal structures of the wild-type phosphonatase complexed with Mg(II) alone or with Mg(II) plus vinylsulfonate (a phosphonoethylenamine analog) were determined to 2.8 and 2.4 A, respectively. These structures were used to determine the identity and positions of active site residues surrounding the Lys-53 ammonium group and the Pald carbonyl. These include Cys-22, His-56, Tyr-128, and Met-49. Site-directed mutagenesis was then employed to determine whether or not these groups participate in catalysis. Based on rate contributions, Tyr-128 and Cys-22 were eliminated as potential catalytic groups. The Lys-53 epsilon-amino group, positioned for reaction with the Pald carbonyl, forms a hydrogen bond with water 120. Water 120 is also within hydrogen bond distance of an imidazole nitrogen of His-56 and the sulfur atom of Met-49. Kinetic constants for mutants indicated that His-56 (1000-fold reduction in k(cat)/K(m) upon Ala substitution) and Met-49 (17,000-fold reduction in k(cat)/K(m) upon Leu substitution) function in catalysis of Schiff-base formation. Based on these results, it is proposed that a network of hydrogen bonds among Lys-53, water 120, His-56, and Met-49 facilitate proton transfer from Lys-53 to the carbinolamine intermediate. Comparison of the vinylsulfonate complex versus unliganded structures indicated that association of the cap and core domains is essential for the positioning of the Lys-53 for attack at the Pald carbonyl and that substrate binding at the core domain stabilizes cap domain binding.  相似文献   

9.
Values of K, delta G(o), delta H(o), delta S(o) and delta C(po) for the binding reaction of small organic ligands forming 1:1 complexes with either alpha- or beta-cyclodextrin were obtained by titration calorimetry from 15 degrees C to 45 degrees C. A hydrogen bond or hydrophobic interaction was introduced by adding a single functional group to the ligand. The thermodynamics of binding with and without the added group are compared to estimate the contribution of the hydrogen bond or hydrophobic interaction. A change in the environment of a functional group is required to influence the binding thermodynamics, but molecular size-dependent solute-solvent interactions have no effect. For phenolic O-H-O hydrogen bond formation, delta H(o) varies from -2 to -1.4 kcal mol(-1) from 15 degrees C to 45 degrees C, and delta C(p) is increased by 18 cal K(-1) mol(-1). The hydrophobic interaction has an opposite effect: in alpha-cyclodextrin, delta C(po) = -13.3 cal K(-1) mol(-1) per ligand -CH(2)-, identical to values found for the transfer of a -CH(2)-group from water to a nonpolar environment. At room temperature, the hydrogen bond and the -CH(2)-interaction each contribute about -600 cal mol(-1) to the stability (delta G(o)) of the complex. With increased temperature, the hydrogen bond stability decreases (i.e., hydrogen bonds "melt"), but the stability of the hydrophobic interaction remains essentially constant.  相似文献   

10.
Periplasmic binding proteins of a new family particularly well represented in Bordetella pertussis have been called Bug receptors. One B.pertussis Bug protein is part of a tripartite tricarboxylate transporter while the functions of the other 77 are unknown. We report the first structure of a Bug receptor, BugD. It adopts the characteristic Venus flytrap motif observed in other periplasmic binding proteins, with two globular domains bisected by a deep cleft. BugD displays a closed conformation resulting from the fortuitous capture of a ligand, identified from the electron density as an aspartate. The structure reveals a distinctive alpha carboxylate-binding motif, involving two water molecules that bridge the carboxylate oxygen atoms to the protein. Both water molecules are hydrogen bonded to a common carbonyl group from Ala14, and each forms a hydrogen bond with one carboxylate oxygen atom of the ligand. Additional hydrogen bonds are found between the ligand alpha carboxylate oxygen atoms and protein backbone amide groups and with a threonine hydroxyl group. This specific ligand-binding motif is highly conserved in Bug proteins, indicating that they may all be receptors of amino acids or other carboxylated solutes, with a similar binding mode. The present structure thus unveils the bases of ligand binding in this large family of periplasmic binding proteins, several hundred members of which have been identified in various bacterial species.  相似文献   

11.
Cavity complementation has been observed in many proteins, where an appropriate small molecule binds to a cavity-forming mutant. Here, the binding of compounds to the W191G cavity mutant of cytochrome c peroxidase is characterized by X-ray crystallography and binding thermodynamics. Unlike cavities created by removal of hydrophobic side-chains, the W191G cavity does not bind neutral or hydrophobic compounds, but displays a strong specificity for heterocyclic cations, consistent with the role of the protein to stabilize a tryptophan radical at this site. Ligand dissociation constants for the protonated cationic state ranged from 6 microM for 2-amino-5-methylthiazole to 1 mM for neutral ligands, and binding was associated with a large enthalpy-entropy compensation. X-ray structures show that each of 18 compounds with binding behavior bind specifically within the artificial cavity and not elsewhere in the protein. The compounds make multiple hydrogen bonds to the cavity walls using a subset of the interactions seen between the protein and solvent in the absence of ligand. For all ligands, every atom that is capable of making a hydrogen bond does so with either protein or solvent. The most often seen interaction is to Asp235, and most compounds bind with a specific orientation that is defined by their ability to interact with this residue. Four of the ligands do not have conventional hydrogen bonding atoms, but were nevertheless observed to orient their most polar CH bond towards Asp235. Two of the larger ligands induce disorder in a surface loop between Pro190 and Asn195 that has been identified as a mobile gate to cavity access. Despite the predominance of hydrogen bonding and electrostatic interactions, the small variation in observed binding free energies were not correlated readily with the strength, type or number of hydrogen bonds or with calculated electrostatic energies alone. Thus, as with naturally occurring binding sites, affinities to W191G are likely to be due to a subtle balance of polar, non-polar, and solvation terms. These studies demonstrate how cavity complementation and judicious choice of site can be used to produce a protein template with an unusual ligand-binding specificity.  相似文献   

12.
Thermodynamic analysis is an effective tool in screening of lead-compounds for development of potential drug candidates. In most cases, a ligand achieve high affinity and specificity to a target protein by means of both favorable enthalpy and entropy terms, which can be reflected in binding profiles of Isothermal Titration Calorimetry (ITC). A favorable enthalpy change suggests the contribution of noncovalent contacts such as hydrogen bonding and van der Waals interaction between a ligand and its target protein. In general, optimization of binding enthalpy is more difficult than that of entropies in ligand-design; therefore, it is desirable to choose firstly a lead-compound based on its binding enthalpic gain. In this paper, we demonstrate the utility of thermodynamic approach to ligand screening using anti-ciguatoxin antibody 10C9 as a model of a target protein which possesses a large hydrophobic pocket. As a result of this screening, we have identified three compounds that could bind to the antigen-binding pocket of 10C9 with a few kcal/mol of favorable binding enthalpy. Comparison of their structure with the proper antigen ciguatoxin CTX3C revealed that 10C9 rigorously identifies their cyclic structure and a characteristic hydroxyl group. ITC measurement might be useful and powerful for a rational ligand screening and the optimization of the ligand; the enthalpic gain is an effective index for ligand-design studies.  相似文献   

13.
A glutathione monomer has been docked into the active site cleft of T4 glutaredoxin (previously called T4 thioredoxin) using molecular graphics. The central part of the cleft is formed by the side chain of Tyr-16 on one side and the residues Thr-64, Met-65, and Pro-66 on the other. The entire glutathione molecule fits well into the cleft. A cis-peptide bond between the residues Met-65 and Pro-66 allows glutathione to bind in an anti-parallel fashion to residues 64-66. Hydrogen bonds can be formed between Met-65 and the glutathione cysteine. This binding positions the glutathione sulfur atom ideally for reaction with the glutaredoxin disulfide. In the model, glutathione can form a hydrogen bond to the hydroxyl group of Tyr-16. Charged interactions at opposite ends of the binding cleft are provided by His-12 and Asp-80. The negatively charged alpha-carboxyl group of glutathione may interact with a positive helix dipole of the protein. Fifteen mutant T4 glutaredoxins have been produced and assayed for glutathione binding by determining thioltransferase activity. Mutant proteins with substitutions in the sides of the cleft (Tyr-16, Pro-66) exhibited the most marked decreases in thioltransferase activity. Mutation of His-12 to a serine decreases the catalytic efficiency whereas substitution of Asp-80 by serine increases the catalytic efficiency. A double mutant, D80S;H12S, has much less affinity for glutathione than either single mutant. Substitution of Cys-14 produces an inactive protein, whereas C17S retains some thioltransferase activity.  相似文献   

14.
15.
Gibberellins (GAs) are phytohormones regulating various developmental processes in plants. In rice, the initial GA-signaling events involve the binding of a GA to the soluble GA receptor protein, GID1. Although X-ray structures for certain GID1/GA complexes have recently been determined, an examination of the complexes does not fully clarify how GID1s discriminate among different GAs. Herein, we present a study aimed at defining the types of forces important to binding via a combination of isothermal titration calorimetry (ITC) and computational docking studies that employed rice GID1 (OsGID1), OsGID1 mutants, which were designed to have a decreased possible number of hydrogen bonds with bound GA, and GA variants. We find that, in general, GA binding is enthalpically driven and that a hydrogen bond between the phenolic hydroxyl of OsGID1 Tyr134 and the C-3 hydroxyl of a GA is a defining structural element. A hydrogen-bond network that involves the C-6 carboxyl of a GA that directly hydrogen bonds the hydroxyl of Ser198 and indirectly, via a two-water-molecule network, the phenolic hydroxyl of Tyr329 and the NH of the amide side-chain of Asn255 is also important for GA binding. The binding of OsGID1 by GA(1) is the most enthalpically driven association found for the biologically active GAs evaluated in this study. This observation might be a consequence of a hydrogen bond formed between the hydroxyl at the C-13 position of GA(1) and the main chain carbonyl of OsGID1 Phe245. Our results demonstrate that by combining ITC experiments and computational methods much can be learned about the thermodynamics of ligand/protein binding.  相似文献   

16.
The crystal structure of a periplasmic l-aspartate/l-glutamate binding protein (DEBP) from Shigella flexneri complexed with an l-glutamate molecule has been determined and refined to an atomic resolution of 1.0 Å. There are two DEBP molecules in the asymmetric unit. The refined model contains 4462 non-hydrogen protein atoms, 730 water molecules, 2 bound glutamate molecules, and 2 Tris molecules from the buffer used in crystallization. The final Rcryst and Rfree factors are 13.61% and 16.89%, respectively. The structure has root-mean-square deviations of 0.016 Å from standard bond lengths and 2.35° from standard bond angles.The DEBP molecule is composed of two similarly folded domains separated by the ligand binding region. Both domains contain a central five-stranded β-sheet that is surrounded by several α-helices. The two domains are linked by two antiparallel β-strands. The overall shape of DEBP is that of an ellipsoid approximately 55 Å × 45 Å × 40 Å in size.The binding of ligand to DEBP is achieved mostly through hydrogen bonds between the glutamate and side-chain and main-chain groups of DEBP. Side chains of residues Arg24, Ser72, Arg75, Ser90, and His164 anchor the deprotonated γ-carboxylate group of the glutamate with six hydrogen bonds. Side chains of Arg75 and Arg90 form salt bridges with the deprotonated α-carboxylate group, while the main-chain amide groups of Thr92 and Thr140 form hydrogen bonds with the same group. The positively charged α-amino group of the l-glutamate forms salt bridge interaction with the side-chain carboxylate group of Asp182 and hydrogen bond interaction with main-chain carbonyl oxygen of Ser90. In addition to these hydrogen bond and electrostatic interactions, other interactions may also play important roles. For example, the two methylene groups from the glutamate form van der Waals interactions with hydrophobic side chains of DEBP.Comparisons with several other periplasmic amino acid binding proteins indicate that DEBP residues involved in the binding of α-amino and α-carboxylate groups of the ligand and the pattern of hydrogen bond formation between these groups are very well conserved, but the binding pocket around the ligand side chain is not, leading to the specificity of DEBP. We have identified structural features of DEBP that determine its ability of binding glutamate and aspartate, two molecules with different sizes, but discriminating against very similar glutamine and asparagine molecules.  相似文献   

17.
18.
Lee FH  Haskell C  Charo IF  Boettiger D 《Biochemistry》2004,43(22):7179-7186
Receptor-ligand binding analyses have generally used soluble components to measure thermodynamic binding constants. In their biological context, adhesion receptors bind to an immobile ligand and the binding reaction is confined to the cell-substrate contact zone. We have developed a new procedure based on the spinning disk technology to measure the number of receptor-ligand bonds in the contact zone. Application of this methodology to the CX3CR1-fractalkine and the CXCR1-IL-8 receptor-ligand systems demonstrated that the level of binding to an immobilized ligand is reduced by several orders of magnitude in comparison to solution binding. A comparison of the solution binding and contact zone binding constants shows that the effect of ligand immobilization was similar for each system. In contrast, although the CXCR1-IL-8 bond had the higher affinity, the average bond strength was only 10% of that for the CX3CR1 bond. Because fractalkine can be expressed as a cell surface-bound protein, CX3CR1 has been proposed to function as an adhesion receptor. The higher bond strength suggests that the bond architecture has also evolved to serve an adhesion function.  相似文献   

19.
Structural and theoretical studies on the geometrical features of a hydrogen‐bond network occurring in the binding site of nicotinic acetylcholine receptors (nAChRs) and composed of interconnected WxPD (Trp‐x‐Pro‐Asp) and SWyz (Ser‐Trp‐yz) sequences from loops A and B, respectively, have been carried out. Multiple sequence alignments using as template the sequence of the apoform of Aplysia californica acetylcholine binding protein (Ac‐AChBP) show the strict conservation of serine and tryptophan residues of the loop B SWyz sequence. Considering a sample of 19 high resolution AChBP structures, the strong conformational preferences of the key tryptophan residue has been pointing out, whatever the form, free or bounded, of AChBP. The geometry of the motif hydrogen‐bond network has been characterized through the analyses of seven distances. The robustness of the various hydrogen‐bond interactions is pointed out, the one involving the aspartate carboxylate group and the serine residue being the shortest of the network. The role of a cooperative effect involving a NH(His145)…OH (Ser142) hydrogen bond is highlighted. Density functional theory calculations on several simplified models based on the motif hydrogen‐bond network allow probing the importance of the various hydrogen‐bond interactions. The removal of the Ser142 hydroxyl group induces strong structural rearrangements, in agreement with the structural observations. Molecular electrostatic potential calculations on model systems highlight the importance of a cooperative effect in the whole hydrogen‐bond network. More precisely, the key role of the Ser142 hydroxyl group, involved in several hydrogen bonds, is underlined. Proteins 2014; 82:2303–2317. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
In the sperm whale myoglobin mutant H93G, the proximal histidine is replaced by glycine, leaving a cavity in which exogenous imidazole can bind and ligate the heme iron (Barrick, D. (1994) Biochemistry 33, 6545-6554). Structural studies of this mutant suggest that serine 92 may play an important role in imidazole binding by serving as a hydrogen bond acceptor. Serine 92 is highly conserved in myoglobins, forming a well-characterized weak hydrogen bond with the proximal histidine in the native protein. We have probed the importance of this hydrogen bond through studies of the double mutants S92A/H93G and S92T/H93G incorporating exogenous imidazole and methylimidazoles. (1)H NMR spectra reveal that loss of the hydrogen bond in S92A/H93G does not affect the conformation of the bound imidazole. However, the binding constants for imidazoles to the ferrous nitrosyl complex of S92A/H93G are much weaker than in H93G. These results are discussed in terms of hydrogen bonding and steric packing within the proximal cavity. The results also highlight the importance of the trans diatomic ligand in altering the binding and sensitivity to perturbation of the ligand in the proximal cavity.  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