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.
Triosephosphate isomerase (TIM), glycerol 3-phosphate dehydrogenase, and orotidine 5'-monophosphate decarboxylase each use the binding energy from the interaction of phosphite dianion with a flexible phosphate gripper loop to activate a second, phosphodianion-truncated, substrate towards enzyme-catalyzed proton transfer, hydride transfer, and decarboxylation, respectively. Studies on TIM suggest that the most important general effect of loop closure over the substrate phosphodianion, and the associated conformational changes, is to extrude water from the enzyme active site. This should cause a decrease in the effective active-site dielectric constant, and an increase in transition state stabilization from enhanced electrostatic interactions with polar amino acid side chains. The most important specific effect of these conformational changes is to increase the basicity of the carboxylate side chain of the active site glutamate base by its placement in a 'hydrophobic cage'.  相似文献   

2.
Amyes TL  Richard JP 《Biochemistry》2007,46(19):5841-5854
More than 80% of the rate acceleration for enzymatic catalysis of the aldose-ketose isomerization of (R)-glyceraldehyde 3-phosphate (GAP) by triosephosphate isomerase (TIM) can be attributed to the phosphodianion group of GAP [Amyes, T. L., O'Donoghue, A. C., and Richard, J. P. (2001) J. Am. Chem. Soc. 123, 11325-11326]. We examine here the necessity of the covalent connection between the phosphodianion and triose sugar portions of the substrate by "carving up" GAP into the minimal neutral two-carbon sugar glycolaldehyde and phosphite dianion pieces. This "two-part substrate" preserves both the alpha-hydroxycarbonyl and oxydianion portions of GAP. TIM catalyzes proton transfer from glycolaldehyde in D2O, resulting in deuterium incorporation that can be monitored by 1H NMR spectroscopy, with kcat/Km = 0.26 M-1 s-1. Exogenous phosphite dianion results in a very large increase in the observed second-order rate constant (kcat/Km)obsd for turnover of glycolaldehyde, and the dependence of (kcat/Km)obsd on [HPO32-] exhibits saturation. The data give kcat/Km = 185 M-1 s-1 for turnover of glycolaldehyde by TIM that is saturated with phosphite dianion so that the separate binding of phosphite dianion to TIM results in a 700-fold acceleration of proton transfer from carbon. The binding of phosphite dianion to the free enzyme (Kd = 38 mM) is 700-fold weaker than its binding to the fleeting complex of TIM with the altered substrate in the transition state (Kd = 53 muM); the total intrinsic binding energy of phosphite dianion in the transition state is 5.8 kcal/mol. We propose a physical model for catalysis by TIM in which the intrinsic binding energy of the substrate phosphodianion group is utilized to drive closing of the "mobile loop" and a protein conformational change that leads to formation of an active site environment that is optimally organized for stabilization of the transition state for proton transfer from alpha-carbonyl carbon.  相似文献   

3.
Malabanan MM  Go MK  Amyes TL  Richard JP 《Biochemistry》2011,50(25):5767-5779
Product yields for the reactions of (R)-glyceraldehyde 3-phosphate (GAP) in D2O at pD 7.9 catalyzed by wildtype triosephosphate isomerase from Trypanosoma brucei brucei (Tbb TIM) and a monomeric variant (monoTIM) of this wildtype enzyme were determined by (1)H NMR spectroscopy and were compared with the yields determined in earlier work for the reactions catalyzed by TIM from rabbit and chicken muscle [O'Donoghue, A. C., Amyes, T. L., and Richard, J. P. (2005), Biochemistry 44, 2610 - 2621]. Three products were observed from the reactions catalyzed by TIM: dihydroxyacetone phosphate (DHAP) from isomerization with intramolecular transfer of hydrogen, d-DHAP from isomerization with incorporation of deuterium from D2O into C-1 of DHAP, and d-GAP from incorporation of deuterium from D2O into C-2 of GAP. The yield of DHAP formed by intramolecular transfer of hydrogen decreases from 49% for the muscle enzymes to 40% for wildtype Tbb TIM to 34% for monoTIM. There is no significant difference in the ratio of the yields of d-DHAP and d-GAP for wildtype TIM from muscle sources and Trypanosoma brucei brucei, but partitioning of the enediolate intermediate of the monoTIM reaction to form d-DHAP is less favorable ((k(C1))(D)/(k(C2))(D) = 1.1) than for the wildtype enzyme ((k(C1))(D)/(k(C2))(D) = 1.7). Product yields for the wildtype Tbb TIM and monoTIM-catalyzed reactions of glycolaldehyde labeled with carbon-13 at the carbonyl carbon ([1-(13)C]-GA) at pD 7.0 in the presence of phosphite dianion and in its absence were determined by (1)H NMR spectroscopy [Go, M. K., Amyes, T. L., and Richard, J. P. (2009) Biochemistry 48, 5769-5778]. There is no detectable difference in the yields of the products of wildtype muscle and Tbb TIM-catalyzed reactions of [1-(13)C]-GA in D2O. The kinetic parameters for phosphite dianion activation of the reactions of [1-(13)C]-GA catalyzed by wildtype Tbb TIM are similar to those reported for the enzyme from rabbit muscle [Amyes, T. L. and Richard, J. P. (2007) Biochemistry 46, 5841-5854], but there is no detectable dianion activation of the reaction catalyzed by monoTIM. The engineered disruption of subunit contacts at monoTIM causes movement of the essential side chains of Lys-13 and His-95 away from the catalytic active positions. We suggest that this places an increased demand that the intrinsic binding energy of phosphite dianion be utilized to drive the change in the conformation of monoTIM back to the active structure for wildtype TIM.  相似文献   

4.
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   

5.
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   

6.
Mutants of orotidine 5'-monophosphate decarboxylase containing all possible single (Q215A, Y217F, and R235A), double, and triple substitutions of the side chains that interact with the phosphodianion group of the substrate orotidine 5'-monophosphate have been prepared. Essentially the entire effect of these mutations on the decarboxylation of the truncated neutral substrate 1-(β-d-erythrofuranosyl)orotic acid that lacks a phosphodianion group is expressed as a decrease in the third-order rate constant for activation by phosphite dianion. The results are consistent with a model in which phosphodianion binding interactions are utilized to stabilize a rare closed enzyme form that exhibits a high catalytic activity for decarboxylation.  相似文献   

7.
Kinetic analysis of decarboxylation catalyzed by S154A, Q215A, and S154A/Q215A mutant yeast orotidine 5'-monophosphate decarboxylases with orotidine 5'-monophosphate (OMP) and with a truncated nucleoside substrate (EO) activated by phosphite dianion shows (1) the side chain of Ser-154 stabilizes the transition state through interactions with the pyrimidine rings of OMP or EO, (2) the side chain of Gln-215 interacts with the phosphodianion group of OMP or with phosphite dianion, and (3) the interloop hydrogen bond between the side chains of Ser-154 and Gln-215 orients the amide side chain of Gln-215 to interact with the phosphodianion group of OMP or with phosphite dianion.  相似文献   

8.
The importance of the fully conserved active site proline, Pro168, for the reaction mechanism of triosephosphate isomerase (TIM) has been investigated by studying the enzymatic and crystallographic properties of the P168A variant of trypanosomal TIM. In TIM, Pro168 follows the key catalytic residue Glu167, situated at the beginning of the flexible active site loop (loop 6). Turnover numbers of the P168A variant for its substrates are reduced approximately 50-fold, whereas the Km values are approximately 2 times lower. The affinity of the P168A variant for the transition state analogue 2-phosphoglycolate (2PG) is reduced 5-fold. The crystal structures of unliganded and liganded (2PG) P168A show that the phosphate moiety of 2PG is bound similarly as in wild-type TIM, whereas the interactions of the carboxylic acid moiety with the side chain of the catalytic Glu167 differ. The unique properties of the proline side chain at position 168 are required to transmit ligand binding to the conformational change of Glu167: the side chain of Glu167 flips from the inactive swung-out to the active swung-in conformation on ligand binding in wild-type TIM, whereas in the mutant this conformational change does not occur. Further structural comparisons show that in the wild-type enzyme the concerted movement of loop 6 and loop 7 from unliganded-open to liganded-closed appears to be facilitated by the interactions of the phosphate moiety with loop 7. Apparently, the rotation of 90 degrees of the Gly211-Gly212 peptide plane of loop 7 plays a key role in this concerted movement.  相似文献   

9.
The product distributions for the reactions of (R)-glyceraldehyde 3-phosphate (GAP) in D(2)O at pD 7.5-7.9 catalyzed by triosephosphate isomerase (TIM) from chicken and rabbit muscle were determined by (1)H NMR spectroscopy. Three products were observed from the reactions catalyzed by TIM: dihydroxyacetone phosphate (DHAP) from isomerization with intramolecular transfer of hydrogen (49% of the enzymatic products), [1(R)-(2)H]-DHAP from isomerization with incorporation of deuterium from D(2)O into C-1 of DHAP (31% of the enzymatic products), and [2(R)-(2)H]-GAP from incorporation of deuterium from D(2)O into C-2 of GAP (21% of the enzymatic products). The similar yields of [1(R)-(2)H]-DHAP and [2(R)-(2)H]-GAP from partitioning of the enzyme-bound enediol(ate) intermediate between hydron transfer to C-1 and C-2 is consistent with earlier results, which showed that there are similar barriers for conversion of this intermediate to the alpha-hydroxy ketone and aldehyde products (Knowles, J. R., and Albery, W. J. (1977) Acc. Chem. Res. 10, 105-111). However, the observation that the TIM-catalyzed isomerization of GAP in D(2)O proceeds with 49% intramolecular transfer of the (1)H label from substrate to product DHAP stands in sharp contrast with the 相似文献   

10.
The replacement of Arg-54 by Ala in the active site of Escherichia coli aspartate transcarbamoylase causes a 17,000-fold loss of activity but does not significantly influence the binding of substrates or substrate analogs (Stebbins, J.W., Xu, W., & Kantrowitz, E.R., 1989, Biochemistry 28, 2592-2600). In the X-ray structure of the wild-type enzyme, Arg-54 interacts with both the anhydride oxygen and a phosphate oxygen of carbamoyl phosphate (CP) (Gouaux, J.E. & Lipscomb, W.N., 1988, Proc. Natl. Acad. Sci. USA 85, 4205-4208). The Arg-54-->Ala enzyme was crystallized in the presence of the transition state analog N-phosphonacetyl-L-aspartate (PALA), data were collected to a resolution limit of 2.8 A, and the structure was solved by molecular replacement. The analysis of the refined structure (R factor = 0.18) indicates that the substitution did not cause any significant alterations to the active site, except that the side chain of the arginine was replaced by two water molecules. 31P-NMR studies indicate that the binding of CP to the wild-type catalytic subunit produces an upfield chemical shift that cannot reflect a significant change in the ionization state of the CP but rather indicates that there are perturbations in the electronic environment around the phosphate moiety when CP binds to the enzyme. The pH dependence of this upfield shift for bound CP indicates that the catalytic subunit undergoes a conformational change with a pKa approximately 7.7 upon CP binding. Furthermore, the linewidth of the 31P signal of CP bound to the Arg-54-->Ala enzyme is significantly narrower than that of CP bound to the wild-type catalytic subunit at any pH, although the change in chemical shift for the CP bound to the mutant enzyme is unaltered. 31P-NMR studies of PALA complexed to the wild-type catalytic subunit indicate that the phosphonate group of the bound PALA exists as the dianion at pH 7.0 and 8.8, whereas in the Arg-54-->Ala catalytic subunit the phosphonate group of the bound PALA exists as the monoanion at pH 7.0 and 8.8. Thus, the side chain of Arg-54 is essential for the proper ionization of the phosphonate group of PALA and by analogy the phosphate group in the transition state. These data support the previously proposed proton transfer mechanism, in which a fully ionized phosphate group in the transition state accepts a proton during catalysis.  相似文献   

11.
Previous crystallographic studies on glycogen phosphorylase have described the different conformational states of the protein (T and R) that represent the allosteric transition and have shown how the properties of the 5'-phosphate group of the cofactor pyridoxal phosphate are influenced by these conformational states. The present work reports a study on glycogen phosphorylase b (GPb) complexed with a modified cofactor, pyridoxal 5'-diphosphate (PLPP), in place of the natural cofactor. Solution studies (Withers, S.G., Madsen, N.B., & Sykes, B.D., 1982, Biochemistry 21, 6716-6722) have shown that PLPP promotes R-state properties of the enzyme indicating that the cofactor can influence the conformational state of the protein. GPb complexed with pyridoxal 5'-diphosphate (PLPP) has been crystallized in the presence of IMP and ammonium sulfate in the monoclinic R-state crystal form and the structure refined from X-ray data to 2.8 A resolution to a crystallographic R value of 0.21. The global tertiary and quaternary structure in the vicinity of the Ser 14 and the IMP sites are nearly identical to those observed for the R-state GPb-AMP complex. At the catalytic site the second phosphate of PLPP is accommodated with essentially no change in structure from the R-state structure and is involved in interactions with the side chains of two lysine residues (Lys 568 and Lys 574) and the main chain nitrogen of Arg 569. Superposition of the T-state structure shows that were the PLPP to be incorporated into the T-state structure there would be a close contact with the 280s loop (residues 282-285) that would encourage the T to R allosteric transition. The second phosphate of the PLPP occupies a site that is distinct from other dianionic binding sites that have been observed for glucose-1-phosphate and sulfate (in the R state) and for heptulose-2-phosphate (in the T state). The results indicate mobility in the dianion recognition site, and the precise position is dependent on other linkages to the dianion. In the modified cofactor the second phosphate site is constrained by the covalent link to the first phosphate of PLPP. The observed position in the crystal suggests that it is too far from the substrate site to represent a site for catalysis.  相似文献   

12.
NMR titration curves are reported for the 4 histidine residues of ribonuclease A in sodium acetate and for ribonuclease S in sodium acetate, phosphate, and sulfate solutions. Evidence is presented that the imidazole side chain of histidine residue 48 undergoes a conformational change, probably also involving the carboxyl side chain of aspartic acid residue 14. This group is considered to be responsible for the low pH inflection with pKa 4.2 present in the NMR titration curve of the C-2 proton resonance of histidine 48. The NMR titration curves of the active site histidine residues 12 and 119 also exhibit inflections at low pH values, although there is no carboxyl group within 9 A of the imidazole side chain of histidine residue 12 in the structure of ribonuclease S determined by x-ray crystallography (Wyckoff, H. W., Tsernoglou, D., Hanson, A. W. Knox, J. R., Lee, B., and Richards, F. M. (1970) J. Biol. Chem. 245, 305-328). Curve fitting was carried out on 11 sets of NMR titration data using a model in which the 3 histidine residues 12, 119, and 48 are assumed to be affected by a common carboxyl group. The results obtained indicate that such a model with fewer parameters gives as good a representation of the data as the model in which each histidine residue is assumed to interact separately with a different carboxyl group. Therefore, it is concluded that the ionization of aspartic acid residue 14 is indirectly experienced by the active site histidine residues through the conformational change at histidine 48. A model assuming mutual interaction of the active site histidine residues does not account for the low pH inflections in these curves.  相似文献   

13.
Triose phosphate isomerase is a dimeric enzyme of molecular mass 56 000 which catalyses the interconversion of dihydroxyacetone phosphate (DHAP) and D-glyceraldehyde-3-phosphate. The crystal structure of the enzyme from chicken muscle has been determined at a resolution of 2.5 A, and an independent determination of the structure of the yeast enzyme has just been completed at 3 A resolution. The conformation of the polypeptide chain is essentially identical in the two structures, and consists of an inner cylinder of eight strands of parallel beta-pleated sheet, with mostly helical segments connecting each strand. The active site is a pocket containing glutamic acid 165, which is believed to act as a base in the reaction. Crystallographic studies of the binding of DHAP to both the chicken and the yeast enzymes reveal a common mode of binding and suggest a mechanisms for catalysis involving polarization of the substrate carbonyl group.  相似文献   

14.
Enzymes achieve their catalytic proficiency by precisely positioning the substrate and catalytic residues with respect to each other. Atomic resolution crystallography is an excellent tool to study the important details of these geometric active‐site features. Here, we have investigated the reaction mechanism of triosephosphate isomerase (TIM) using atomic resolution crystallographic studies at 0.82‐Å resolution of leishmanial TIM complexed with the well‐studied reaction‐intermediate analog phosphoglycolohydroxamate (PGH). Remaining unresolved aspects of the reaction mechanism of TIM such as the protonation state of the first reaction intermediate and the properties of the hydrogen‐bonding interactions in the active site are being addressed. The hydroxamate moiety of PGH interacts via unusually short hydrogen bonds of its N1? O1 moiety with the carboxylate group of the catalytic glutamate (Glu167), for example, the distance of N1(PGH)‐OE2(Glu167) is 2.69 ± 0.01 Å and the distance of O1(PGH)‐OE1(Glu167) is 2.60 ± 0.01 Å. Structural comparisons show that the side chain of the catalytic base (Glu167) can move during the reaction cycle in a small cavity, located above the hydroxamate plane. The structure analysis suggests that the hydroxamate moiety of PGH is negatively charged. Therefore, the bound PGH mimics the negatively charged enediolate intermediate, which is formed immediately after the initial proton abstraction from DHAP by the catalytic glutamate. The new findings are discussed in the context of the current knowledge of the TIM reaction mechanism. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

15.
The catalytic mechanism of triosephosphate isomerase (TIM) was investigated with ab initio quantum mechanical calculations. Electrostatic interactions between the quantum mechanical active site and the protein and solvent environment were modeled using the finite difference Poission-Boltzman method. The complexes of TIM with the substrate dihydroxyacetone phosphate (DHAP), five possible intermediates and the product glyceraldehyde-3-phosphate (GAP) were optimized in the active-site model at the 3-21G(*) level and energy profile for the proton abstraction from DHAP by the active-site Glu167 was calculated at the MP2/3-21G(*)//3-21G(*) level. Calculated energetics of the enzyme reaction were found to be in reasonable agreement with the experimental findings. Calculations revealed that an enediol of the substrate is a probable intermediate in the enzyme reaction. It was suggested that the proton abstracted from the substrate by the active-site glutamate goes to the carbonyl oxygen of the substrate producing enediol intermediate either directly or after it is exchanged with solvent. © 1996 Wiley-Liss, Inc.  相似文献   

16.
The crystal structure of the binary complex of non-activated ribulose-1,5-bisphosphate carboxylase/oxygenase from Rhodospirillum rubrum and its product 3-phospho-D-glycerate has been determined to 2.9-A resolution. This structure determination confirms the proposed location of the active site (Schneider, G., Lindqvist, Y., Br?ndén, C.-I., and Lorimer, G. (1986) EMBO J. 5, 3409-3415) at the carboxyl end of the beta-strands of the alpha/beta-barrel in the carboxyl-terminal domain. One molecule of 3-phosphoglycerate is bound per active site. All oxygen atoms of 3-phosphoglycerate form hydrogen bonds to groups of the enzyme. The phosphate group interacts with the sidechains of residues Arg-288, His-321, and Ser-368, which are conserved between enzymes from different species as well as with the main chain nitrogens from residues Thr-322 and Gly-323. These amino acid residues constitute one of the two phosphate binding sites of the active site. The carboxyl group interacts with the side chains of His-287, Lys-191, and Asn-111. Implications of the activation process for the binding of 3-phosphoglycerate are discussed.  相似文献   

17.
The product distributions for the reactions of dihydroxyacetone phosphate (DHAP) in D(2)O at pD 7.9 catalyzed by triosephosphate isomerase (TIM) from chicken and rabbit muscle were determined by (1)H NMR spectroscopy using glyceraldehyde 3-phosphate dehydrogenase to trap the first-formed products of the thermodynamically unfavorable isomerization reaction, (R)-glyceraldehyde 3-phosphate (GAP) and [2(R)-(2)H]-GAP (d-GAP). Three products were observed from the reactions catalyzed by TIM: GAP from isomerization with intramolecular transfer of hydrogen (18% of the enzymatic products), d-GAP from isomerization with incorporation of deuterium from D(2)O into C-2 of GAP (43% of the enzymatic products), and [1(R)-(2)H]-DHAP (d-DHAP) from incorporation of deuterium from D(2)O into C-1 of DHAP (40% of the enzymatic products). The ratios of the yields of the deuterium-labeled products d-DHAP and d-GAP from partitioning of the intermediate of the TIM-catalyzed reactions of GAP and DHAP in D(2)O are 1.48 and 0.93, respectively. This provides evidence that the reaction of these two substrates does not proceed through a single, common, reaction intermediate but, rather, through distinct intermediates that differ in the bonding and arrangement of catalytic residues at the enediolate O-1 and O-2 oxyanions formed on deprotonation of GAP and DHAP, respectively.  相似文献   

18.
In most eubacteria, fungi, apicomplexa, plants and some metazoans, the active form of vitamin B6, PLP (pyridoxal 5-phosphate), is de novo synthesized from three substrates, R5P (ribose 5-phosphate), DHAP (dihydroxyacetone phosphate) and ammonia hydrolysed from glutamine by a complexed glutaminase. Of the three active sites of DXP (deoxyxylulose 5-phosphate)independent PLP synthase (Pdx1), the R5P isomerization site has been assigned, but the sites for DHAP isomerization and PLP formation remain unknown. In the present study, we present the crystal structures of yeast Pdx1/Snz1, in apo-, G3P (glyceraldehyde 3-phosphate)- and PLP-bound forms, at 2.3, 1.8 and 2.2 ? (1 ?=0.1 nm) respectively. Structural and biochemical analysis enabled us to assign the PLP-formation site, a G3P-binding site and a G3P-transfer site. We propose a putative catalytic mechanism for Pdx1/Snz1 in which R5P and DHAP are isomerized at two distinct sites and transferred along well-defined routes to a final destination for PLP synthesis.  相似文献   

19.
The key residue of the active site of triosephosphate isomerase (TIM) is the catalytic glutamate, which is proposed to be important (i) as a catalytic base, for initiating the reaction, as well as (ii) for the subsequent proton shuttling steps. The structural properties of this glutamate in the liganded complex have been investigated by studying the high resolution crystal structures of typanosomal TIM, complexed with three suicide inhibitors: (S)-glycidol phosphate ((S)-GOP, at 0.99 Å resolution), (R)-glycidol phosphate, ((R)-GOP, at 1.08 Å resolution), and bromohydroxyacetone phosphate (BHAP, at 1.97 Å resolution). The structures show that in the (S)-GOP active site this catalytic glutamate is in the well characterized, competent conformation. However, an unusual side chain conformation is observed in the (R)-GOP and BHAP complexes. In addition, Glu97, salt bridged to the catalytic lysine in the competent active site, adopts an unusual side chain conformation in these two latter complexes. The higher chemical reactivity of (S)-GOP compared with (R)-GOP, as known from solution studies, can be understood: the structures indicate that in the case of (S)-GOP, Glu167 can attack the terminal carbon of the epoxide in a stereoelectronically favored, nearly linear O–C–O arrangement, but this is not possible for the (R)-GOP isomer. These structures confirm the previously proposed conformational flexibility of the catalytic glutamate in its closed, liganded state. The importance of this conformational flexibility for the proton shuttling steps in the TIM catalytic cycle, which is apparently achieved by a sliding motion of the side chain carboxylate group above the enediolate plane, is also discussed.  相似文献   

20.
The crystal structures of l-fuculose-1-phosphate aldolase (FucA) with and without a ligated analogue of dihydroxyacetone phosphate (DHAP) and of a number of active center mutants have resulted in a model of the catalytic mechanism. This model has now been confirmed by structural analyses of further mutations at the zinc coordination sphere and at the phosphate site. In addition, these mutants have revealed new aspects of the catalysis: the hydroxyl group of Tyr113' (from a neighboring subunit), which sits just outside the zinc coordination sphere, steers DHAP towards a productive binding mode at the zinc ion; Glu73 contacts zinc in between the two ligand positions intended for the DHAP oxygen atoms and thus avoids blocking of these positions by a tetrahedrally coordinated hydroxy ion; the FucA polypeptide does not assume its minimum energy state but oscillates between two states of elevated energy as demonstrated by a mutant in a minimum energy state. The back and forth motion involves a mobile loop connecting the phosphate site with intersubunit motions and thus with the Brownian motion of the solvent. The phosphate group is bound strongly at a given distance to the zinc ion, which prevents the formation of too tight a DHAP:zinc complex. This observation explains our failure to find mutants that accept phosphate-free substitutes for DHAP. The FucA zinc coordination sphere is compared with that of carbonic anhydrase.  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