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.
Fuculose-1-phosphate aldolase (FucA) is a useful biocatalyst with potential applications in chiral synthesis. In this study, the overall kinetic mechanism of FucA from the archaeon Methanococcus jannaschii was studied. The K(m) values of dihydroxyacetone phosphate (DHAP) and dl-glyceraldehyde were 0.09 and 0.74 mM, respectively. Dead-end inhibition by trimethyl phosphonoacetate and dl-threose were competitive and uncompetitive with respect to DHAP and dl-glyceraldehyde. Inhibition patterns obtained using reaction products were noncompetitive vs. DHAP and competitive vs. dl-glyceraldehyde. The equilibrium constant was 8.309×10(-3) M as assessed by varying the [DHAP]/[product] ratio at a fixed dl-glyceraldehyde concentration and by measuring the change in DHAP concentration after equilibrium was reached. This constant is consistent with the K(eq) value obtained from (13)C NMR (15.625×10(-3) M). The resultant inhibition kinetics may suggest the insights of kinetic mechanism of the FucA catalyzed reaction.  相似文献   

2.
The structures of human glyoxalase I in complexes with S-(N-hydroxy-N-p-iodophenylcarbamoyl)glutathione (HIPC-GSH) and S-p-nitrobenzyloxycarbonylglutathione (NBC-GSH) have been determined at 2.0 and 1.72 A resolution, respectively. HIPC-GSH is a transition state analogue mimicking the enediolate intermediate that forms along the reaction pathway of glyoxalase I. In the structure, the hydroxycarbamoyl function is directly coordinated to the active site zinc ion. In contrast, the equivalent group in the NBC-GSH complex is approximately 6 A from the metal in a conformation that may resemble the product complex with S-D-lactoylglutathione. In this complex, two water molecules occupy the liganding positions at the zinc ion occupied by the hydroxycarbamoyl function in the enediolate analogue complex. Coordination of the transition state analogue to the metal enables a loop to close down over the active site, relative to its position in the product-like structure, allowing the glycine residue of the glutathione moiety to hydrogen bond with the protein. The structure of the complex with the enediolate analogue supports an "inner sphere mechanism" in which the GSH-methylglyoxal thiohemiacetal substrate is converted to product via a cis-enediolate intermediate. The zinc ion is envisioned to play an electrophilic role in catalysis by directly coordinating this intermediate. In addition, the carboxyl of Glu 172 is proposed to be displaced from the inner coordination sphere of the metal ion during substrate binding, thus allowing this group to facilitate proton transfer between the adjacent carbon atoms of the substrate. This proposal is supported by the observation that in the complex with the enediolate analogue the carboxyl group of Glu 172 is 3.3 A from the metal and is in an ideal position for reprotonation of the transition state intermediate. In contrast, Glu 172 is directly coordinated to the zinc ion in the complexes with S-benzylglutathione and with NBC-GSH.  相似文献   

3.
4.
The three-dimensional structure of the complex of ribulose-1,5-bisphosphate carboxylase from Rhodospirillum rubrum, CO2, Mg2+, and ribulose bisphosphate has been determined with x-ray crystallographic methods to 2.6-A resolution. Ribulose-1,5-bisphosphate binds across the active site with the two phosphate groups in the two phosphate binding sites of the beta/alpha barrel. The oxygen atoms of the carbamate and the side chain of Asp-193 provide the protein ligands to the bound Mg2+ ion. The C2 and the C3 or C4 oxygen atoms of the substrate are also within the first coordination sphere of the metal ion. At the present resolution of the electron density maps, two slightly different conformations of the substrate, with the C3 hydroxyl group "cis" or "trans" to the C2 oxygen, can be built into the observed electron density. The two different conformations suggest two different mechanisms of proton abstraction in the first step of catalysis, the enolization of the ribulose 1,5-bisphosphate. Two loop regions, which are disordered in the crystals of the nonactivated enzyme, could be built into their respective electron density. A comparison with the structure of the quaternary complex of the spinach enzyme shows that despite the different conformations of loop 6, the positions of the Mg2+ ion, and most atoms of the substrate are very similar when superimposed on each other. There are, however, some significant differences at the active site, especially in the metal coordination sphere.  相似文献   

5.
The structure of a class II fructose-1,6-bisphosphate aldolase in complex with the substrate analogue and inhibitor phosphoglycolohydroxamate (PGH) has been determined using X-ray diffraction terms to a resolution of 2.0 A (1 A=0.1 nm). The crystals are trigonal, space group P3121 with a=b=78.24 A, c=289.69 A. The asymmetric unit is a homodimer of (alpha/beta)8 barrels and the model has refined to give R-work 19.2 %, R-free (based on 5 % of the data) 23.0 %. PGH resembles the ene-diolate transition state of the physiological substrate dihydroxyacetone phosphate. It is well ordered and bound in a deep polar cavity at the C-terminal end of the (alpha/beta)8 barrel, where it chelates the catalytic zinc ion using hydroxyl and enolate oxygen atoms. Trigonal bipyramidal coordination of the zinc ion is completed by three histidine residues. The complex network of hydrogen bonds at the catalytic centre is required to organise the position of key functional groups and metal ion ligands. A well-defined monovalent cation-binding site is observed following significant re-organisation of loop structures. This assists the formation of a phosphate-binding site on one side of the barrel that tethers PGH in the catalytic site. The positions of functional groups of substrate and putative interactions with key amino acid residues are identified. Knowledge of the complex structure complements the results of spectroscopic and site-directed mutagenesis studies, and contributes to our understanding of the mechanism and substrate specificity of this family of enzymes. A reaction mechanism distinct from that proposed for other class II aldolases is discussed. The results suggest that the class II aldolases should be sub-divided into two groups on the basis of both distinct folds and mechanism.  相似文献   

6.
Yeast Cox4 is a zinc binding subunit of cytochrome c oxidase. Cox4 is the only cofactor-containing subunit that is not directly part of the catalytic core of the enzyme located in the mitochondrial inner membrane. The Zn(II) site is shown to be distinct from the bovine ortholog, as it results from the x-ray structure of the entire cytochrome c oxidase in having a single histidyl residue and three conserved cysteines residues in the coordination sphere. Substitutions at the Cys ligand positions result in non-functional Cox4 proteins that fail to lead to cytochrome oxidase assembly. Limited function exists in His-119 mutants when overexpressed. Zn(II) binding in Cox4 is, therefore, important for the stability of the complex. The solution structure of yeast Cox4 elucidated by multidimensional NMR reveals a C-terminal globular domain consisting of two beta sheets analogous to the bovine ortholog except the loop containing the coordinating His in the yeast protein and the fourth Cys in the bovine protein are in different positions in the two structures. The conformation of this loop is dictated by the different sequence position of the fourth coordinating zinc ligand. The Zn(II) ion is buried within the domain, consistent with its role in structural stability. Potential functions of this matrix-facing subunit are discussed.  相似文献   

7.
Vogel A  Schilling O  Meyer-Klaucke W 《Biochemistry》2004,43(32):10379-10386
Zinc phosphodiesterase (ZiPD) is a member of the metallo-beta-lactamase family with a binuclear zinc binding site. As an experimental attempt to identify the metal ligands of Escherichia coli ZiPD and to investigate their function in catalysis, we mutationally exchanged candidate metal coordinating residues and performed kinetic and X-ray absorption spectroscopic analysis of the mutant proteins. All mutants (H66E, H69A, H141A, D212A, D212C, H231A, H248A, and H270A) show significantly lower catalytic rates toward the substrate bis(p-nitrophenyl)phosphate. Substrate binding, represented by the kinetic value K', remains unchanged for six mutants, whereas it is increased 3-4-fold for H231A and H270A. Accordingly, these two residues are supposed to be involved in substrate binding, whereas the others are more important for catalytic turnover and thus are assumed to be involved in zinc ligation. Structural insight into the metal binding of D212 was gained by zinc K-edge extended X-ray absorption fine structure (EXAFS). The sulfur coordination number of the cysteine mutant was found to be 1, demonstrating binding to both zinc metals in a bridging mode. Taken together with two residues from a strictly conserved sequence region within the metallo-beta-lactamase family, the metal site of ZiPD is proposed with H64, H66, and H141 coordinating ZnA, D68, H69, and H248 coordinating ZnB, and D212 bridging both metals. Surprisingly, the same coordination sphere is found in glyoxalase II. This is further substantiated by comparable EXAFS spectra of both native enzymes. This is the first example of the same metal site in two members of the metallo-beta-lactamase domain proteins catalyzing different reactions. The kinetic analysis of mutants provides unexpected insights into the reaction mechanism of ZiPD.  相似文献   

8.
We have previously provided evidence for a dinuclear zinc site in rabbit skeletal muscle AMPD compatible with a (micro-aqua)(micro-carboxylato)dizinc(II) core with an average of two histidine residues at each metal site. XAS of the zinc binding site of the enzyme in the presence of PRN favors a model where PRN is added to the coordination sphere of one of the two zinc ions increasing its coordination number to five. The uncompetitive nature of the inhibition of AMPD by fluoride reveals that the anion probably displaces the nucleophile water molecule terminally coordinated to the catalytic Zn(1) ion at the enzyme C-terminus, following the binding of AMP at the Zn(2) ion located at N-terminus of the enzyme. Thus, the two Zn ions in the AMPD metallocenter operate together as a single catalytic unit, but have independent function, one of them (Zn(1)) acting to polarize the nucleophile water molecule, whilst the other (Zn(2)) acts transiently as a receptor for an activating substrate molecule. The addition of fluoride to AMPD also abolishes the cooperative behaviour induced in the enzyme by the inhibitory effect of ATP at acidic pH that probably resides in the competition with the substrate for an adenine nucleotide specific regulatory site located in the Zn(2) ion binding region and which is responsible for the positive homotropic cooperativity behaviour of AMPD.  相似文献   

9.
Native thermolysin binds a single catalytically essential zinc ion that is tetrahedrally coordinated by three protein ligands and a water molecule. During catalysis the zinc ligation is thought to change from fourfold to fivefold. Substitution of the active-site zinc with Cd2+, Mn2+, Fe2+, and Co2+ alters the catalytic activity (Holmquist B, Vallee BL, 1974, J Biol Chem 249:4601-4607). Excess zinc inhibits the enzyme. To investigate the structural basis of these changes in activity, we have determined the structures of a series of metal-substituted thermolysins at 1.7-1.9 A resolution. The structure of the Co(2+)-substituted enzyme is shown to be very similar to that of wild type except that two solvent molecules are liganded to the metal at positions that are thought to be occupied by the two oxygens of the hydrated scissile peptide in the transition state. Thus, the enhanced activity toward some substrates of the cobalt-relative to the zinc-substituted enzyme may be due to enhanced stabilization of the transition state. The ability of Zn2+ and Co2+ to accept tetrahedral coordination in the Michaelis complex, as well as fivefold coordination in the transition state, may also contribute to their effectiveness in catalysis. The Cd(2+)- and Mn(2+)-substituted thermolysins display conformational changes that disrupt the active site to varying degrees and could explain the associated reduction of activity. The conformational changes involve not only the essential catalytic residue, Glu 143, but also concerted side-chain rotations in the adjacent residues Met 120 and Leu 144. Some of these side-chain movements are similar to adjustments that have been observed previously in association with the "hinge-bending" motion that is presumed to occur during catalysis by the zinc endoproteases. In the presence of excess zinc, a second zinc ion is observed to bind at His 231 within 3.2 A of the zinc bound to native thermolysin, explaining the inhibitory effect.  相似文献   

10.
We present molecular dynamics simulations on the active site region of dimeric triose phosphate isomerase (TIM) using the co-ordinates of native chicken muscle TIM as a starting point and performing simulations with no substrate, with dihydroxyacetone phosphate (DHAP), the natural substrate, and with dihydroxyacetone sulfate (DHAS), a substrate analog. Whereas most of the protein moves less than 1 A during the simulation, some residues in the active site loop move more than 8 A during the 10.5 picoseconds of dynamics for each of the simulations. Most interestingly, the nature of the loop motion depends on the substrate, with the largest motion found in the presence of DHAP, and only in the presence of DHAP does the loop move to "close off" the active site pocket. The final structure found for the DHAP-chicken TIM complex is qualitatively similar to that described by Alber et al. for DHAP-yeast TIM. Simulations on the monomeric protein gives insight into why the molecule is active only as a dimer.  相似文献   

11.
Thermoanaerobacter brockii alcohol dehydrogenase (TbADH) catalyzes the reversible oxidation of secondary alcohols to the corresponding ketones using NADP(+) as the cofactor. The active site of the enzyme contains a zinc ion that is tetrahedrally coordinated by four protein residues. The enzymatic reaction leads to the formation of a ternary enzyme-cofactor-substrate complex; and catalytic hydride ion transfer is believed to take place directly between the substrate and cofactor at the ternary complex. Although crystallographic data of TbADH and other alcohol dehydrogenases as well as their complexes are available, their mode of action remains to be determined. It is firmly established that the zinc ion is essential for catalysis. However, there is no clear agreement about the coordination environment of the metal ion and the competent reaction intermediates during catalysis. We used a combination of X-ray absorption, circular dichroism (CD), and fluorescence spectroscopy, together with structural analysis and modeling studies, to investigate the ternary complexes of TbADH that are bound to a transition-state analogue inhibitor. Our structural and spectroscopic studies indicated that the coordination sphere of the catalytic zinc site in TbADH undergoes conformational changes when it binds the inhibitor and forms a pentacoordinated complex at the zinc ion. These studies provide the first active site structure of bacterial ADH bound to a substrate analogue. Here, we suggest the active site structure of the central intermediate complex and, more specifically, propose the substrate-binding site in TbADH.  相似文献   

12.
G H Reed  T S Leyh 《Biochemistry》1980,19(24):5472-5480
The complete coordination scheme for Mn(II) in transition-state-analogue complexes with creatine kinase has been determined by electron paramagnetic resonance (EPR) spectroscopy. Perturbations in the EPR spectra for Mn(II) due to superhyperfine coupling to 17O of selectively labeled ligands have been used to identify oxygen ligands in the first coordination sphere of the metal ion. The results show that in the complex of enzyme-MnADP-formate-creatine, Mn(II) is bound to oxygen ligands from both the alpha- and beta-phosphate groups of ADP, to an oxygen from the carboxylate group of formate, and to three water molecules. In the complex with thiocyanate replacing formate as the stabilizing anion, previous infrared experiments [Reed, G. H., Barlow, C. H., & Burns, R. A., Jr. (1978) J. Biol. Chem. 253, 4153-4158] indicated that the nitrogen from thiocyanate was bound to the Mn(II). The magnitudes of the 17O superphyperfine coupling constants from the O- ligands of the ADP phosphate groups and from the formate carboxylate are approximately equal and are larger than that for the water ligands. The symmetry of the zero-field-splitting tensor for Mn(II) indicates that the oxygens from the alpha- and beta-phosphate groups of ADP and the ligand donor atom from the anion occupy mutually cis positions in the octahedral coordination geometry. Water proton relaxation time measurements show that the three water molecules which are bound to Mn(II) are not in free exchange with the bulk solvent. Hence, an enclosed structure at the active site is indicated. The results suggest that for creatine kinase the activating metal ion is bound to all three phosphate groups in the transition state of the reaction.  相似文献   

13.
[NiFe] hydrogenases catalyze the reversible formation of H2. The [NiFe] heterobimetallic active site is rich in redox states. Here, we investigate the key catalytic state Ni?CC of Desulfovibrio vulgaris Miyazaki F hydrogenase using a cluster model that includes the truncated amino acids of the entire second coordination sphere of the enzyme. The optimized geometries, computed g?tensors, hyperfine coupling constants, and IR stretching frequencies all agree well with experimental values. For the hydride in the bridging position, only a single minimum on the potential energy surface is found, indicating that the hydride bridges and binds to both nickel and iron. The influence of the second coordination sphere on the electronic structure is investigated by comparing results from the large cluster models with truncated models. The largest interactions of the second coordination sphere with the active site concern the hydrogen bonds with the cyanide ligands, which modulate the bond between iron and these ligands. Secondly, the electronic structure of the active site is found to be sensitive to the protonation state of His88. This residue forms a hydrogen bond with the spin-carrying sulfur atom of Cys549, which in turn tunes the spin density at the nickel and coordinating sulfur atoms. In addition, the unequal distribution of spin density over the equatorial cysteine residues results from different orientations of the cysteine side chains, which are kept in their particular orientation by the secondary structure of the protein.  相似文献   

14.
Tagatose-1,6-bisphosphate aldolase from Streptococcus pyogenes is a class I aldolase that exhibits a remarkable lack of chiral discrimination with respect to the configuration of hydroxyl groups at both C3 and C4 positions. The enzyme catalyzes the reversible cleavage of four diastereoisomers (fructose 1,6-bisphosphate (FBP), psicose 1,6-bisphosphate, sorbose 1,6-bisphosphate, and tagatose 1,6-bisphosphate) to dihydroxyacetone phosphate (DHAP) and d-glyceraldehyde 3-phosphate with high catalytic efficiency. To investigate its enzymatic mechanism, high resolution crystal structures were determined of both native enzyme and native enzyme in complex with dihydroxyacetone-P. The electron density map revealed a (α/β)8 fold in each dimeric subunit. Flash-cooled crystals of native enzyme soaked with dihydroxyacetone phosphate trapped a covalent intermediate with carbanionic character at Lys205, different from the enamine mesomer bound in stereospecific class I FBP aldolase. Structural analysis indicates extensive active site conservation with respect to class I FBP aldolases, including conserved conformational responses to DHAP binding and conserved stereospecific proton transfer at the DHAP C3 carbon mediated by a proximal water molecule. Exchange reactions with tritiated water and tritium-labeled DHAP at C3 hydrogen were carried out in both solution and crystalline state to assess stereochemical control at C3. The kinetic studies show labeling at both pro-R and pro-S C3 positions of DHAP yet detritiation only at the C3 pro-S-labeled position. Detritiation of the C3 pro-R label was not detected and is consistent with preferential cis-trans isomerism about the C2–C3 bond in the carbanion as the mechanism responsible for C3 epimerization in tagatose-1,6-bisphosphate aldolase.  相似文献   

15.
The use of substrate analogues as inhibitors provides a way to understand and manipulate enzyme function. Here we report two 1 A resolution crystal structures of liver alcohol dehydrogenase in complex with NADH and two inhibitors: dimethyl sulfoxide and isobutyramide. Both structures present a dynamic state of inhibition. In the dimethyl sulfoxide complex structure, the inhibitor is caught in transition on its way to the active site using a flash-freezing protocol and a cadmium-substituted enzyme. One inhibitor molecule is partly located in the first and partly in the second coordination sphere of the active site metal. A hydroxide ion bound to the active site metal lies close to the pyridine ring of NADH, which is puckered in a twisted boat conformation. The cadmium ion is coordinated by both the hydroxide ion and the inhibitor molecule, providing structural evidence of a coordination switch at the active site metal ion. The structure of the isobutyramide complex reveals the partial formation of an adduct between the isobutyramide inhibitor and NADH. It provides evidence of the contribution of a shift from the keto to the enol tautomer during aldehyde reduction. The different positions of the inhibitors further refine the knowledge of the dynamics of the enzyme mechanism and explain how the crowded active site can facilitate the presence of a substrate and a metal-bound hydroxide ion.  相似文献   

16.
Tsang WY  Amyes TL  Richard JP 《Biochemistry》2008,47(16):4575-4582
The ratio of the second-order rate constants for reduction of dihydroxyacetone phosphate (DHAP) and of the neutral truncated substrate glycolaldehyde (GLY) by glycerol 3-phosphate dehydrogenase (NAD (+), GPDH) saturated with NADH is (1.0 x 10 (6) M (-1) s (-1))/(8.7 x 10 (-3) M (-1) s (-1)) = 1.1 x 10 (8), which was used to calculate an intrinsic phosphate binding energy of at least 11.0 kcal/mol. Phosphite dianion binds very weakly to GPDH ( K d > 0.1 M), but the bound dianion strongly activates GLY toward enzyme-catalyzed reduction by NADH. Thus, the large intrinsic phosphite binding energy is expressed only at the transition state for the GPDH-catalyzed reaction. The ratio of rate constants for the phosphite-activated and the unactivated GPDH-catalyzed reduction of GLY by NADH is (4300 M (-2) s (-1))/(8.7 x 10 (-3) M (-1) s (-1)) = 5 x 10 (5) M (-1), which was used to calculate an intrinsic phosphite binding energy of -7.7 kcal/mol for the association of phosphite dianion with the transition state complex for the GPDH-catalyzed reduction of GLY. Phosphite dianion has now been shown to activate bound substrates for enzyme-catalyzed proton transfer, decarboxylation, hydride transfer, and phosphoryl transfer reactions. Structural data provide strong evidence that enzymic activation by the binding of phosphite dianion occurs at a modular active site featuring (1) a binding pocket complementary to the reactive substrate fragment which contains all the active site residues needed to catalyze the reaction of the substrate piece or of the whole substrate and (2) a phosphate/phosphite dianion binding pocket that is completed by the movement of flexible protein loop(s) to surround the nonreacting oxydianion. We propose that loop motion and associated protein conformational changes that accompany the binding of phosphite dianion and/or phosphodianion substrates lead to encapsulation of the substrate and/or its pieces in the protein interior, and to placement of the active site residues in positions where they provide optimal stabilization of the transition state for the catalyzed reaction.  相似文献   

17.
The singnificance of the zinc hydroxide–Thr-199–Glu-106 hydrogen-bond network in the active site of human carbonic anhydrase II has been examined by X-ray crystallographic analyses of site-specific mutants. Mutants with Ala-199 and Ala-106 or Gln-106 have low catalytic activities, while a mutant with Asp-106 has almost full CO2 hydration activity. The structures of these four mutants, as well as that of the bicarbonate complex of the mutant with Ala-199, have been determined at 1.7 to 2.2 Å resolution. Removal of the γ atoms of residue 199 leads to distorted tetrahedral geometry at the zine ion, and a catalytically important zinc-bound water molecule has moved towards Glu-106. In the bicarbonate complex of the mutant with Ala-199 one oxygen atom from bicarbonate binds to zinc without displacing this water molecule. Tetrahedral coordination geometries are retained in the mutants at position 106. The mutants with Ala-106 and Gln-106 have a zinc-bound sulfate ion, whereas this sulfate site is only partially occupied in the mutant with Asp-106. The hydrogen-bond network seems to be “reversed” in the mutants with Ala-106 and Gln-106. The network is preserved as in native enzyme in the mutant with Asp-106 but the side chain of Asp-106 is more extended than that of Glu-106 in the native enzyme. These results illustrate the importance of Glu-106 and Thr-199 for controlling the precise coordination geometry of the zinc ion and its ligand preferences with results in an optimal orientation of a zine-bound hydroxide ion for an attack on the CO2 substrate. © 1993 Wiley-Liss, Inc.  相似文献   

18.
The crystal structure of the phospholipase A2 (PLA2) heterodimer from Naja naja sagittifera reveals the presence of a new PLA2-like protein with eight disulphide bridges. The heterodimer is formed between a commonly observed group I PLA2 having seven characteristic disulfide bonds and a novel PLA2-like protein (Cys-PLA2) containing two extra cysteines at two highly conserved sites (positions 32 and 49) of structural and functional importance. The crystals of the heterodimer belong to tetragonal space group P41212 with cell dimensions, a = b = 77.7 A and c = 68.4 A corresponding to a solvent content of 33%, which is one of the lowest values observed so far in the PLA2 crystals. The structure has been solved with molecular replacement method and refined to a final R value of 21.6% [Rfree = 25.6%]. The electron density revealed the presence of cysteines 32 and 49 that are covalently linked to give rise to an eighth disulphide bridge in the PLA2-like monomer. A non-protein high-quality electron density was also observed at the substrate-binding site in the PLA2-like protein that has been interpreted as N-acetylglucosamine. The overall tertiary folds of the two monomers are similar having all features of PLA2-type folding. A zinc ion is detected at the interface of the heterodimer with fivefold coordination while another zinc ion was found on the surface of Cys-PLA2 with sixfold coordination. The conformations of the calcium-binding loops of both monomers are significantly different from each other as well as from those in other group I PLA2s. The N-acetylglucosamine molecule is favorably placed in the substrate-binding site of Cys-PLA2 and forms five hydrogen bonds and several van der Waals interactions with protein atoms, thus indicating a strong affinity. It also provides clue of the possible mechanism of sugar recognition by PLA2 and PLA2-like proteins. The formation of heterodimer seems to have been induced by zinc ion.  相似文献   

19.
A model calculation is carried out to study the potential energy profile of a sodium ion with several water molecules inside a simplified model of the gramicidin ion channel. The sodium ion is treated as a Lennard-Jones sphere with a point charge at its center. The Barnes polarizable water model is used to mimic the water molecules. A polarizable and deformable gramicidinlike channel is constructed based on the model obtained by Koeppe and Kimura. Potential minima and saddle points are located and the static energy barriers are computed. The potential minima at the two mouths of the channel exhibit an aqueous solvation structure very different from that at any of the interior minima. These sites are approximately 23.6 and 24.4 A apart for binding of a sodium ion and a cesium ion, respectively. Ionic motion from these exterior sites to the first interior minimum requires substantial rearrangement of the waters of solvation; this rearrangement may be the hydration/dehydration step in ionic permeation through the channel. Based on these results, a mechanism by which the sodium ion moves from the exterior binding site to the interior of the channel is proposed. Our model channel accommodates about eight water molecules and the transport of the ion and water within the channel is found to be single file. Results of less extensive calculations for Cs+ and Li+ ions in a channel with or without water are also reported.  相似文献   

20.
Despite the enormous interest that has been devoted to the study of farnesyltransferase, many questions concerning its catalytic mechanism remain unanswered. In particular, several doubts exist on the structure of the active-site zinc coordination sphere, more precisely on the nature of the fourth ligand, which is displaced during the catalytic reaction by a peptide thiolate. From available crystallographic structures, and mainly from x-ray absorption fine structure data, two possible alternatives emerge: a tightly zinc-bound water molecule or an almost symmetrical bidentate aspartate residue (Asp-297beta). In this study, high-level theoretical calculations, with different-sized active site models, were used to elucidate this aspect. Our results demonstrate that both coordination alternatives lie in a notably close energetic proximity, even though the bidentate hypothesis has a somewhat lower energy. The Gibbs reaction and activation energies for the mono-bidentate conversion, as well as the structure for the corresponding transition state, were also determined. Globally, these results indicate that at room temperature the mono-bidentate conversion is reversible and very fast, and that probably both states exist in equilibrium, which suggests that a carboxylate-shift mechanism may have a key role in the farnesylation process by assisting the coordination/displacement of ligands to the zinc ion, thereby controlling the enzyme activity. Based on this equilibrium hypothesis, an explanation for the existing contradictions between the crystallographic and x-ray absorption fine structure results is proposed.  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