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.
Glycoside hydrolase of Cel48F from Clostridium cellulolyticum is an important processive cellulose, which can hydrolyze cellulose into cellobiose. Molecular dynamics simulations were used to investigate the hydrolysis mechanism of cellulose. The two conformations of the Cel48F‐cellotetrose complex in which the cellotetroses are bound at different sites (known as the sliding conformation and the hydrolyzing conformation) were simulated. By comparing these two conformations, a water‐control mechanism is proposed, in which the hydrolysis proceeds by providing a water molecule for every other glucosidic linkage. The roles of certain key residues are determined: Glu55 and Asp230 are the most probable candidates for acid and base, respectively, in the mechanism of inverting anomeric carbon. Met414 and Trp417 constitute the water‐control system. Glu44 might keep the substrate at a certain location within the active site or help the substrate chain to move from the sliding conformation to the hydrolyzing conformation. The other hydrophobic residues around the substrate can decrease the sliding energy barrier or provide a hydrophobic environment to resist entry of the surrounding water molecules into the active site, except for those coming from a specific water channel. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

2.
Mulakala C  Reilly PJ 《Proteins》2005,61(3):590-596
AutoDock is a small-molecule docking program that uses an energy function to score docked ligands. Here AutoDock's grid-based method for energy evaluation was exploited to evaluate the force exerted by Fusarium oxysporum Cel7B on the atoms of docked cellooligosaccharides and a thiooligosaccharide substrate analog. Coupled with the interaction energies evaluated for each docked ligand, these forces give insight into the dynamics of the ligand in the active site, and help to elucidate the relative importance of specific enzyme-substrate interactions in stabilizing the substrate transition-state conformation. The processive force on the docked substrate in the F. oxysporum Cel7B active site is less than half of that on the docked substrate in the Hypocrea jecorina Cel7A active site. Hydrogen bonding interactions of the enzyme with the C2 hydroxyl group of the glucosyl residue in subsite -2 and with the C3 hydroxyl group of the glucosyl residue in subsite +1 are the most significant in stabilizing the distorted14B transition-state conformation of the glucosyl residue in subsite -1. The force calculations also help to elucidate the mechanism that prevents the active site from fouling.  相似文献   

3.
Mutagenesis experiments suggest that Asp79 in cellulase Cel6A (E2) from Thermobifida fusca has a catalytic role, in spite of the fact that this residue is more than 13 A from the scissile bond in models of the enzyme-substrate complex built upon the crystal structure of the protein. This suggests that there is a substantial conformational shift in the protein upon substrate binding. Molecular mechanics simulations were used to investigate possible alternate conformations of the protein bound to a tetrasaccharide substrate, primarily involving shifts of the loop containing Asp79, and to model the role of water in the active site complex for both the native conformation and alternative low-energy conformations. Several alternative conformations of reasonable energy have been identified, including one in which the overall energy of the enzyme-substrate complex in solution is lower than that of the conformation in the crystal structure. This conformation was found to be stable in molecular dynamics simulations with a cellotetraose substrate and water. In simulations of the substrate complexed with the native protein conformation, the sugar ring in the -1 binding site was observed to make a spontaneous transition from the (4)C(1) conformation to a twist-boat conformer, consistent with generally accepted glycosidase mechanisms. Also, from these simulations Tyr73 and Arg78 were found to have important roles in the active site. Based on the results of these various MD simulations, a new catalytic mechanism is proposed. Using this mechanism, predictions about the effects of changes in Arg78 were made which were confirmed by site-directed mutagenesis.  相似文献   

4.
Z Zheng  R W Ashton  F Ni  H A Scheraga 《Biochemistry》1992,31(18):4426-4431
Fibrinogen Lille, a congenital dysfibrinogenemia, has been reported to arise from a mutation from Asp to Asn at position 7 of the A alpha chain of human fibrinogen, thereby reducing the thrombin-catalyzed rate of hydrolysis of the Arg(16)-Gly(17) peptide bond of this chain. Synthetic peptides of relevant portions of the wild-type and mutant A alpha chains were prepared, and the thrombin-catalyzed rates of hydrolysis of their Arg(16)-Gly(17) peptide bonds were determined. In addition, transferred NOE measurements were made to deduce their conformations, when complexed to bovine thrombin. The kinetics data showed little difference in the hydrolysis rates between the wild-type and mutant peptides, and the NMR data indicate no difference in the bound conformation of these two peptides. Therefore, electrostatic (or salt-bridge) interactions between Asp(7) and thrombin do not influence the bound conformations of these peptides. Asp(7) may interact with a remote residue of fibrinogen, not present in these synthetic peptides, or there may be additional mutations beyond A alpha (1-20) which have not been detected in fibrinogen Lille. Alternatively, when thrombin binds to fibrinogen at its secondary binding site, its primary (active) site may display different reactivities toward wild-type fibrinogen and fibrinogen Lille.  相似文献   

5.
Fungal cellobiohydrolases act at liquid-solid interfaces. They have the ability to hydrolyze cellulose chains of a crystalline substrate because of their two-domain structure, i.e. cellulose-binding domain and catalytic domain, and unique active site architecture. However, the details of the action of the two domains on crystalline cellulose are still unclear. Here, we present real time observations of Trichoderma reesei (Tr) cellobiohydrolase I (Cel7A) molecules sliding on crystalline cellulose, obtained with a high speed atomic force microscope. The average velocity of the sliding movement on crystalline cellulose was 3.5 nm/s, and interestingly, the catalytic domain without the cellulose-binding domain moved with a velocity similar to that of the intact TrCel7A enzyme. However, no sliding of a catalytically inactive enzyme (mutant E212Q) or a variant lacking tryptophan at the entrance of the active site tunnel (mutant W40A) could be detected. This indicates that, besides the hydrolysis of glycosidic bonds, the loading of a cellulose chain into the active site tunnel is also essential for the enzyme movement.  相似文献   

6.
One prominent feature of Trichoderma reesei (Tr) endoglucanases catalyzed cellulose hydrolysis is that the reaction slows down quickly after it starts (within minutes). But the mechanism of the slowdown is not well understood. A structural model of Tr- Cel7B catalytic domain bound to cellulose was built computationally and the potentially important binding residues were identified and tested experimentally. The 13 tested mutants show different binding properties in the adsorption to phosphoric acid swollen cellulose and filter paper. Though the partitioning parameter to filter paper is about 10 times smaller than that to phosphoric acid swollen cellulose, a positive correlation is shown for two substrates. The kinetic studies show that the reactions slow down quickly for both substrates. This slowdown is not correlated to the binding constant but anticorrelated to the enzyme initial activity. The amount of reducing sugars released after 24 h by Cel7B in phosphoric acid swollen cellulose, Avicel and filter paper cellulose hydrolysis is correlated with the enzyme activity against a soluble substrate p-nitrophenyl lactoside. Six of the 13 tested mutants, including N47A, N52D, S99A, N323D, S324A, and S346A, yield ∼15–35% more reducing sugars than the wild type (WT) Cel7B in phosphoric acid swollen cellulose and filter paper hydrolysis. This study reveals that the slowdown of the reaction is not due to the binding of the enzyme to cellulose. The activity of Tr- Cel7B against the insoluble substrate cellulose is determined by the enzyme’s capability in hydrolyzing the soluble substrate.  相似文献   

7.
Tobi D  Elber R  Thirumalai D 《Biopolymers》2003,68(3):359-369
The conformational equilibrium of a blocked valine peptide in water and aqueous urea solution is studied using molecular dynamics simulations. Pair correlation functions indicate enhanced concentration of urea near the peptide. Stronger hydrogen bonding of urea-peptide compared to water-peptide is observed with preference for helical conformation. The potential of mean force, computed using umbrella sampling, shows only small differences between urea and water solvation that are difficult to quantify. The changes in solvent structure around the peptide are explained by favorable electrostatic interactions (hydrogen bonds) of urea with the peptide backbone. There is no evidence for significant changes in hydrophobic interactions in the two conformations of the peptide in urea solution. Our simulations suggest that urea denatures proteins by preferentially forming hydrogen bonds to the peptide backbone, reducing the barrier for exposing protein residues to the solvent, and reaching the unfolded state.  相似文献   

8.
Bovine pancreatic deoxyribonuclease I (DNase I) is an endonuclease which cleaves double-stranded DNA. Cocrystal structures of DNase I with oligonucleotides have revealed interactions between the side chains of several amino acids (N74, R111, N170, S206, T207, and Y211) and the DNA phosphates. The effects these interactions have on enzyme catalysis and DNA hydrolysis selectivity have been investigated by site-directed mutagenesis. Mutations to R111, N170, T207, and Y211 severely compromised activity toward both DNA and a small chromophoric substrate. A hydrogen bond between R111 (which interacts with the phosphate immediately 5' to the cutting site) and the essential amino acid H134 is probably required to maintain this histidine in the correct orientation for efficient hydrolysis. Both T207 and Y211 bind to the phosphate immediately 3' to the cleavage site. Additionally, T207 is involved in binding an essential, structural, calcium ion, and Y211 is the nearest neighbor to D212, a critical catalytic residue. N170 interacts with the scissile phosphate and appears to play a direct role in the catalytic mechanism. The mutation N74D, which interacts with a phosphate twice removed from the scissile group, strongly reduced DNA hydrolysis. However, a comparison of DNase I variants from several species suggests that certain amino acids, which allow interaction with phosphates (positively charged or hydrogen bonding), are tolerated. S206, which binds to a DNA phosphate two positions away from the cleavage site, appears to play a relatively unimportant role. None of the enzyme variants, including a triple mutation in which N74, R111, and Y211 were altered, affected DNA hydrolysis selectivity. This suggests that phosphate binding residues play no role in the selection of DNA substrates.  相似文献   

9.
Despite much work, many key aspects of the mechanism of the dihydrofolate reductase (DHFR) catalyzed reduction of dihydrofolate remain unresolved. In bacterial forms of DHFR both substrate and water access to the active site are controlled by the conformation of the mobile M20 loop. In vertebrate DHFRs only one conformation of the residues corresponding to the M20 loop has been observed. Access to the active site was proposed to be controlled by residue 31. MD simulations of chicken DHFR complexed with substrates and cofactor revealed a closing of the side chain of Tyr 31 over the active site on binding of dihydrofolate. This conformational change was dependent on the presence of glutamate on the para-aminobenzoylamide moiety of dihydrofolate. In its absence, the conformation remained open. Although water could enter the active site and hydrogen bond to N5 of dihydrofolate, indicating the feasibility of water as the proton donor, this was not controlled by the conformation of Tyr 31. The water accessibility of the active site was low for both conformations of Tyr 31. However, when hydride was transferred from NADPH to C6 of dihydrofolate before protonation, the average time during which water was found in hydrogen bonding distance to N5 of dihydrofolate in the active site increased almost fivefold. These results indicated that water can serve as the Broensted acid for the protonation of N5 of dihydrofolate during the DHFR catalyzed reduction.  相似文献   

10.
Hydrogen bonding interactions between transmembrane helices stabilize the visual pigment rhodopsin in an inactive conformation in the dark. The crystal structure of rhodopsin has previously revealed that Glu122 and Trp126 on transmembrane helix H3 form a complex hydrogen bonding network with Tyr206 and His211 on H5, while the indole nitrogen of Trp265 on H6 forms a water-mediated hydrogen bond with Asn302 on H7. Here, we use solid-state magic angle spinning NMR spectroscopy to probe the changes in hydrogen bonding upon rhodopsin activation. The NMR chemical shifts of 15N-labeled tryptophan are consistent with the indole nitrogens of Trp126 and Trp265 becoming more weakly hydrogen bonded between rhodopsin and metarhodopsin II. The NMR chemical shifts of 15N-labeled histidine show that His211 is neutral; the unprotonated imidazole nitrogen is not coordinated to zinc in rhodopsin and becomes more strongly hydrogen bonded in metarhodopsin II. Moreover, measurements of rhodopsin containing 13C-labeled histidine show that a strong hydrogen bond between the side-chain of Glu122 and the backbone carbonyl of His211 is disrupted in metarhodopsin II. The implications of these observations for the activation mechanism of rhodopsin are discussed.  相似文献   

11.
Xylanases (EC 3.2.1.8 endo-1,4-glycosyl hydrolase) catalyze the hydrolysis of xylan, an abundant hemicellulose of plant cell walls. Access to the catalytic site of GH11 xylanases is regulated by movement of a short β-hairpin, the so-called thumb region, which can adopt open or closed conformations. A crystallographic study has shown that the D11F/R122D mutant of the GH11 xylanase A from Bacillus subtilis (BsXA) displays a stable “open” conformation, and here we report a molecular dynamics simulation study comparing this mutant with the native enzyme over a range of temperatures. The mutant open conformation was stable at 300 and 328 K, however it showed a transition to the closed state at 338 K. Analysis of dihedral angles identified thumb region residues Y113 and T123 as key hinge points which determine the open-closed transition at 338 K. Although the D11F/R122D mutations result in a reduction in local inter-intramolecular hydrogen bonding, the global energies of the open and closed conformations in the native enzyme are equivalent, suggesting that the two conformations are equally accessible. These results indicate that the thumb region shows a broader degree of energetically permissible conformations which regulate the access to the active site region. The R122D mutation contributes to the stability of the open conformation, but is not essential for thumb dynamics, i.e., the wild type enzyme can also adapt to the open conformation.  相似文献   

12.
13.
Ca2+-transporting adenosine triphosphatase (ATPase) of sarcoplasmic reticulum couples ATP hydrolysis with ion transport. Phosphorylation of the cytosolic region of the calcium-bound conformation (E1) of the protein leads to drastic conformational rearrangements of the transmembrane helices and the release of Ca2+. The resulting calcium-free conformation (E2) is less stable than the E1 form. The changes in van der Waals interactions and interhelical hydrogen bonding in the E1 and E2 conformations were compared. Conformational changes in the transmembrane region concomitant with the release of Ca2+ mainly affect the number of interhelical hydrogen bonds, which is reduced to half of that in E1 form, whereas the number of interhelical atomic pairwise contacts reflecting van der Waals interactions experience little change. The interhelical hydrogen bonds in Ca2+-transporting ATPase can be divided into two groups according to their roles: those that play a structural stabilizing role and those that are important for the correct geometry of the Ca2+ binding site. Interhelical hydrogen bonds in the transmembrane regions play important roles for the stability and specificity of helix-helix interactions in proteins where change of conformation is required for transport of ions or small molecules.  相似文献   

14.
Thermoactinomyces vulgaris R-47 alpha-amylase 2 (TVAII) can efficiently hydrolyze both starch and cyclomaltooligosaccharides (cyclodextrins). The crystal structure of an inactive mutant TVAII in a complex with maltohexaose was determined at a resolution of 2.1A. TVAII adopts a dimeric structure to form two catalytic sites, where substrates are found to bind. At the catalytic site, there are many hydrogen bonds between the enzyme and substrate at the non-reducing end from the hydrolyzing site, but few hydrogen bonds at the reducing end, where two aromatic residues, Trp356 and Tyr45, make effective interactions with a substrate. Trp356 drastically changes its side-chain conformation to achieve a strong stacking interaction with the substrate, and Tyr45 from another molecule forms a water-mediated hydrogen bond with the substrate. Kinetic analysis of the wild-type and mutant enzymes in which Trp356 and/or Tyr45 were replaced with Ala suggested that Trp356 and Tyr45 are essential to the catalytic reaction of the enzyme, and that the formation of a dimeric structure is indispensable for TVAII to hydrolyze both starch and cyclodextrins.  相似文献   

15.
The active sites of caspases are composed of four mobile loops. A loop (L2) from one half of the dimer interacts with a loop (L2′) from the other half of the dimer to bind substrate. In an inactive form, the two L2′ loops form a cross‐dimer hydrogen‐bond network over the dimer interface. Although the L2′ loop has been implicated as playing a central role in the formation of the active‐site loop bundle, its precise role in catalysis has not been shown. A detailed understanding of the active and inactive conformations is essential to control the caspase function. We have interrogated the contributions of the residues in the L2′ loop to catalytic function and enzyme stability. In wild‐type and all mutants, active‐site binding results in substantial stabilization of the complex. One mutation, P214A, is significantly destabilized in the ligand‐free conformation, but is as stable as wild type when bound to substrate, indicating that caspase‐7 rests in different conformations in the absence and presence of substrate. Residues K212 and I213 in the L2′ loop are shown to be essential for substrate‐binding and thus proper catalytic function of the caspase. In the crystal structure of I213A, the void created by side‐chain deletion is compensated for by rearrangement of tyrosine 211 to fill the void, suggesting that the requirements of substrate‐binding are sufficiently strong to induce the active conformation. Thus, although the L2′ loop makes no direct contacts with substrate, it is essential for buttressing the substrate‐binding groove and is central to native catalytic efficiency.  相似文献   

16.
Eukaryotic low-molecular-weight protein tyrosine phosphatases (LMW PTPs) contain a conserved serine, a histidine with an elevated pKa, and an active site asparagine that together form a highly conserved hydrogen bonding network. This network stabilizes the active site phosphate binding loop for optimal substrate binding and catalysis. In the phosphatase from the bovine parasite Tritrichomonas foetus (TPTP), both the conserved serine (S37) and asparagine (N14) are present, but the conserved histidine has been replaced by a glutamine residue (Q67). Site-directed mutagenesis, kinetic, and spectroscopic experiments suggest that Q67 is located near the active site and is important for optimal catalytic activity. Kinetic experiments also suggest that S37 participates in the active site/hydrogen bonding network. Nuclear magnetic resonance spectroscopy was used to determine the three-dimensional structure of the TPTP enzyme and to further examine the roles of S37 and Q67. The backbone conformation of the TPTP phosphate binding loop is nearly superimposable with that of other tyrosine phosphatases, with N14 existing in a strained, left-handed conformation that is a hallmark of the active site hydrogen bonding network in the LMW PTPs. As expected, both S37 and Q67 are located at the active site, but in the consensus structure they are not within hydrogen bonding distance of N14. The hydrogen bond interactions that are observed in X-ray structures of LMW PTPs may in fact be transient in solution. Protein dynamics within the active site hydrogen bonding network appear to be affected by the presence of substrate or bound inhibitors such as inorganic phosphate.  相似文献   

17.
Backbone mimicry by the formation of closed-loop C7, C10 and C13 (mimics of gamma-, beta- and alpha-turns) conformations through side chain-main chain hydrogen bonds by polar groups is a frequent observation in protein structures. A data set of 250 non-homologous and high-resolution protein crystal structures was used to analyze these conformations for their characteristic features. Seven out of the nine polar residues (Ser, Thr, Asn, Asp, Gln, Glu and His) have hydrogen bonding groups in their side chains which can participate in such mimicry and as many as 15% of all these polar residues engage in such conformations. The distributions of dihedral angles of these mimics indicate that only certain combinations of the dihedral angles involved aid the formation of these mimics. The observed examples were categorized into various classes based on these combinations, resulting in well defined motifs. Asn and Asp residues show a very high capability to perform such backbone secondary structural mimicry. The most highly mimicked backbone structure is of the C10 conformation by the Asx residues. The mimics formed by His, Ser, Thr and Glx residues are also discussed. The role of such conformations in initiating the formation of regular secondary structures during the course of protein folding seems significant.  相似文献   

18.
We have performed all atom simulations of blocked peptides of the form (AAXAA)3, where X = Gln, Asn, Glu, Asp, Arg, and Lys with explicit water molecules to examine the interactions between side chains spaced i,i–5 in the sequence. Although side chains in this i,i–5 arrangement are commonly believed to be noninteracting, we have observed the formation of unusual i,i–5 main chain hydrogen bonding in such sequences with positively charged residues (Lys) as well as polar uncharged groups (Gln). Our results are consistent with the unusual percentage of hydrogen bonding curves produced by amide exchange measurements on the well-studied sequence acetyl-(AAQAA)3-amide in water (Shalongo, W., Dugad, L., Stellwagen, E. J. Am. Chem. Soc. 116:8288–8293, 1994). Analysis of our simulations indicated that the glutamine side chain showed the greatest propensity to support π helix formation and that the i,i–5 intramolecular hydrogen bonds were stabilized by water-bridging side chain interactions. This intermittent formation of the unusual π helix structure was observed for up to 23% of the total simulation time in some residues in (AAQAA)3. Control studies on peptides with glutamine side chains spaced i,i–3, i,i–4, and i,i–6 did not reveal similar unique structures, providing stronger evidence for the unique role side chain interactions with i,i–5 spacing. © 1997 Wiley-Liss Inc.  相似文献   

19.
Wilkinson AS  Ward S  Kania M  Page MG  Wharton CW 《Biochemistry》1999,38(13):3851-3856
Time-resolved infrared difference spectroscopy has been used to show that the carbonyl group of the acylenzyme reaction intermediate in the Citrobacter freundii beta-lactamase-catalyzed hydrolysis of methicillin can assume at least four conformations. A single-turnover experiment shows that all four conformations decline during deacylation with essentially the same rate constant. The conformers are thus in exchange on the reaction time scale, assuming that deacylation takes place only from the conformation which is most strongly hydrogen bonded or from a more minor species not visible in these experiments. All conformers have the same (10 cm-1) narrow bandwidth compared with a model ethyl ester in deuterium oxide (37 cm-1) which shows that all conformers are well ordered relative to free solution. The polarity of the carbonyl group environment in the conformers varies from 'ether-like' to strongly hydrogen bonding (20 kJ/mol), presumably in the oxyanion hole of the enzyme. From the absorption intensities, it is estimated that the conformers are populated approximately proportional to the hydrogen bonding strength at the carbonyl oxygen. A change in the difference spectrum at 1628 cm-1 consistent with a perturbation (relaxation) of protein beta-sheet occurs slightly faster than deacylation. Consideration of chemical model reactions strongly suggests that neither enamine nor imine formation in the acyl group is a plausible explanation of the change seen at 1628 cm-1. A turnover reaction supports the above conclusions and shows that the conformational relaxation occurs as the substrate is exhausted and the acylenzymes decline. The observation of multiple conformers is discussed in relation to the poor specificity of methicillin as a substrate of this beta-lactamase and in terms of X-ray crystallographic structures of acylenzymes where multiple forms are not apparently observed (or modeled). Infrared spectroscopy has shown itself to be a useful method for assessment of the uniqueness of enzyme-substrate interactions in physiological turnover conditions as well as for determination of ordering, hydrogen bonding, and protein perturbation.  相似文献   

20.
The active site loop of triosephosphate isomerase (TIM) exhibits a hinged-lid motion, alternating between the two well defined "open" and "closed" conformations. Until now the closed conformation had only been observed in protein complexes with substrate analogues. Here, we present the first rabbit muscle apo TIM structure, refined to 1.5A resolution, in which the active site loop is either in the open or in the closed conformation in different subunits of the enzyme. In the closed conformation described here, the lid loop residues participate in stabilizing hydrogen bonds characteristic of holo TIM structures, whereas chemical interactions observed in the open loop conformation are similar to those found in the apo structures of TIM. In the closed conformation, a number of water molecules are observed at the projected ligand atom positions that are hydrogen bonded to the active site residues. Additives used during crystallization (DMSO and Tris molecules and magnesium atoms) were modeled in the electron density maps. However, no specific binding of these molecules is observed at, or close to, the active site and the lid loop. To further investigate this unusual closed conformation of the apo enzyme, two more rabbit muscle TIM structures, one in the same and another in a different crystal form, were determined. These structures present the open lid conformation only, indicating that the closed conformation cannot be explained by crystal contact effects. To rationalize why the active site loop is closed in the absence of ligand in one of the subunits, extensive comparison with previously solved TIM structures was carried out, supported by the bulk of available experimental information about enzyme kinetics and reaction mechanism of TIM. The observation of both open and closed lid conformations in TIM crystals might be related to a persistent conformational heterogeneity of this protein in solution.  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