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.
2.
Cody V  Galitsky N  Rak D  Luft JR  Pangborn W  Queener SF 《Biochemistry》1999,38(14):4303-4312
Structural data from two independent crystal forms (P212121 and P21) of the folate (FA) binary complex and from the ternary complex with the oxidized coenzyme, NADP+, and recombinant Pneumocystis carinii dihydrofolate reductase (pcDHFR) refined to an average of 2.15 A resolution, show the first evidence of ligand-induced conformational changes in the structure of pcDHFR. These data are also compared with the crystal structure of the ternary complex of methotrexate (MTX) with NADPH and pcDHFR in the monoclinic lattice with data to 2.5 A resolution. Comparison of the data for the FA binary complex of pcDHFR with those for the ternary structures reveals significant differences, with a >7 A movement of the loop region near residue 23 that results in a new "flap-open" position for the binary complex, and a "closed" position in the ternary complexes, similar to that reported for Escherichia coli (ec) DHFR complexes. In the orthorhombic lattice for the binary FA pcDHFR complex, there is also an unwinding of a short helical region near residue 47 that places hydrophobic residues Phe-46 and Phe-49 toward the outer surface, a conformation that is stabilized by intermolecular packing contacts. The pyrophosphate moiety of NADP+ in the ternary folate pcDHFR complexes shows significant differences in conformation compared with that observed in the MTX-NADPH-pcDHFR ternary complex. Additionally, comparison of the conformations among these four pcDHFR structures reveals evidence for subdomain movement that correlates with cofactor binding states. The larger binding site access in the new "flap-open" loop 23 conformation of the binary FA complex is consistent with the rapid release of cofactor from the product complex during catalysis as well as the more rapid release of substrate product from the binary complex as a result of the weaker contacts of the closed loop 23 conformation, compared to ecDHFR.  相似文献   

3.
The complex formed by bovine trypsinogen and the pancreatic trypsin inhibitor crystallizes in large crystals isomorphous with trypsin-PTI2 complex crystals Rühlmann et al. 1973. X-ray diffraction data to 1.9 Å resolution were collected in the absence and presence of Ile-Val dipeptide. Both trypsinogen complex structures have been crystallographically refined, using the refined trypsin-PTI complex Huber et al. 1974a as a starting model. The final R values are 0.25 and 0.26, respectively. The mean main-chain atom deviations between the three complex structures are about 0.15 Å. In contrast, the mean deviation between the complexed and the free trypsinogen Fehlhammer et al. 1977 is 0.28 Å, reflecting the influence of crystal packing and complexation. The trypsinogen component adopts a trypsin-like conformation upon PTI binding: The Asp194 side-chain turns around and the activation domain becomes rigid, forming the specificity pocket and the Ile16 binding cleft. The specific interactions between PTI and trypsin are also observed in the trypsinogen complex. As in free trypsinogen, the N-terminus including residues Val10 to Gly18 is mobile and sticks out into solution. Apart from the different arrangement of the N-termini in the two complexes, the only significant, but minor structural difference is the enhanced thermal mobility of the autolysis loop in the trypsinogen complex. Upon binding of the Ile-Val dipeptide, the autolysis loop becomes fixed as in the trypsin complex. The Ile-Val position is identical in the ternary and the trypsin complex.  相似文献   

4.
The 3D structure of recombinant bacterial carboxypeptidase T (CPT) in complex with N-BOC-L-leucine was determined at 1.38 Å resolution. Crystals for the X-ray study were grown in microgravity using the counter-diffusion technique. N-BOC-L-leucine and SO 4 2? ion bound in the enzyme active site were localized in the electron density map. Location of the leucine side chain in CPT-N-BOC-L-leucine complex allowed identification of the S1 subsite of the enzyme, and its structure was determined. Superposition of the structures of CPT-N-BOC-L-leucine complex and complexes of pancreatic carboxypeptidases A and B with substrate and inhibitors was carried out, and similarity of the S1 sub-sites in these three carboxypeptidases was revealed. It was found that SO 4 2? ion occupies the same position in the S1’ subsite as the C-terminal carboxy group of the substrate.  相似文献   

5.
Conformation of coenzyme fragments when bound to lactate dehydrogenase   总被引:7,自引:0,他引:7  
The conformations of adenosine, 5′-AMP and 5′-ADP when bound to dogfish M4 lactate dehydrogenase at pH 7.8 or greater have been determined at 2.8 Å resolution to investigate the events on coenzyme binding. The coenzyme fragments AMP and ADP induce a conformational change in lactate dehydrogenase at pH values less than 6.0 in the same way as do NAD+, NADH or ADPR at any pH value. The structure of NAD+ when bound to lactate dehydrogenase had previously been determined at 5.0 Å resolution. The structures of the bound adenosine, AMP, ADP and NAD+ are compared with the preliminary structure of NAD in a 3.0 Å resolution map of the ternary complex LDH-NAD—pyruvate. Small but significant changes in the binding of the phosphates could be important in the folding of the protein loop over the substrate binding pocket.  相似文献   

6.
The crystal structure of ternary and binary substrate complexes of the catalytic subunit of cAMP-dependent protein kinase has been refined at 2.2 and 2.25 A resolution, respectively. The ternary complex contains ADP and a 20-residue substrate peptide, whereas the binary complex contains the phosphorylated substrate peptide. These 2 structures were refined to crystallographic R-factors of 17.5 and 18.1%, respectively. In the ternary complex, the hydroxyl oxygen OG of the serine at the P-site is 2.7 A from the OD1 atom of Asp 166. This is the first crystallographic evidence showing the direct interaction of this invariant carboxylate with a peptide substrate, and supports the predicted role of Asp 166 as a catalytic base and as an agent to position the serine -OH for nucleophilic attack. A comparison of the substrate and inhibitor ternary complexes places the hydroxyl oxygen of the serine 2.7 A from the gamma-phosphate of ATP and supports a direct in-line mechanism for phosphotransfer. In the binary complex, the phosphate on the Ser interacts directly with the epsilon N of Lys 168, another conserved residue. In the ternary complex containing ATP and the inhibitor peptide, Lys 168 interacts electrostatically with the gamma-phosphate of ATP (Zheng J, Knighton DR, Ten Eyck LF, Karlsson R, Xuong NH, Taylor SS, Sowadski JM, 1993, Biochemistry 32:2154-2161). Thus, Lys 168 remains closely associated with the phosphate in both complexes. A comparison of this binary complex structure with the recently solved structure of the ternary complex containing ATP and inhibitor peptide also reveals that the phosphate atom traverses a distance of about 1.5 A following nucleophilic attack by serine and transfer to the peptide. No major conformational changes of active site residues are seen when the substrate and product complexes are compared, although the binary complex with the phosphopeptide reveals localized changes in conformation in the region corresponding to the glycine-rich loop. The high B-factors for this loop support the conclusion that this structural motif is a highly mobile segment of the protein.  相似文献   

7.
Summary The three-dimensional structure of porcine pancreatic PLA2 (PLA2), present in a 40 kDa ternary complex with micelles and a competitive inhibitor, has been determined using multidimensional heteronuclear NMR spectroscopy. The structure of the protein (124 residues) is based on 1854 constraints, comprising 1792 distance and 62 torsion angle constraints. A total of 18 structures was calculated using a combined approach of distance geometry and restrained molecular dynamics. The atomic rms distribution about the mean coordinate positions for residues 1–62 and 72–124 is 0.75±0.09 Å for the backbone atoms and 1.14±0.10 Å for all atoms. The rms difference between the averaged minimized NMR structures of the free PLA2 and PLA2 in the ternary complex is 3.5 Å for the backbone atoms and 4.0 Å for all atoms. Large differences occur for the calcium-binding loop and the surface loop from residues 62 through 72. The most important difference is found for the first three residues of the N-terminal -helix. Whereas free in solution Ala1, Leu2 and Trp3 are disordered, with the -amino group of Ala1 pointing out into the solvent, in the ternary complex these residues have an -helical conformation with the -amino group buried inside the protein. As a consequence, the important conserved hydrogen bonding network which is also seen in the crystal structures is present only in the ternary complex, but not in free PLA2. Thus, the NMR structure of the N-terminal region (as well as the calcium-binding loop and the surface loop) of PLA2 in the ternary complex resembles that of the crystal structure. Comparison of the NMR structures of the free enzyme and the enzyme in the ternary complex indicates that conformational changes play a role in the interfacial activation of PLA2.  相似文献   

8.
l-3-Hydroxyacyl-CoA dehydrogenase reversibly catalyzes the conversion of l-3-hydroxyacyl-CoA to 3-ketoacyl-CoA concomitant with the reduction of NAD(+) to NADH as part of the beta-oxidation spiral. In this report, crystal structures have been solved for the apoenzyme, binary complexes of the enzyme with reduced cofactor or 3-hydroxybutyryl-CoA substrate, and an abortive ternary complex of the enzyme with NAD(+) and acetoacetyl-CoA. The models illustrate positioning of cofactor and substrate within the active site of the enzyme. Comparison of these structures with the previous model of the enzyme-NAD(+) complex reveals that although significant shifting of the NAD(+)-binding domain relative to the C-terminal domain occurs in the ternary and substrate-bound complexes, there are few differences between the apoenzyme and cofactor-bound complexes. Analysis of these models clarifies the role of key amino acids implicated in catalysis and highlights additional critical residues. Furthermore, a novel charge transfer complex has been identified in the course of abortive ternary complex formation, and its characterization provides additional insight into aspects of the catalytic mechanism of l-3-hydroxyacyl-CoA dehydrogenase.  相似文献   

9.
Three crystal structures have been determined of active site specific substituted Cd(II) horse liver alcohol dehydrogenase and its complexes. Intensities were collected for the free, orthorhombic enzyme to 2.4-A resolution and for a triclinic binary complex with NADH to 2.7-A resolution. A ternary complex was crystallized from an equilibrium mixture of NAD+ and p-bromobenzyl alcohol. The microspectrophotometric analysis of these single crystals showed the protein-bound coenzyme to be largely NADH, which proves the complex to consist of CdII-LADH, NADH, and p-bromobenzyl alcohol. Intensity data for this abortive ternary complex were collected to 2.9-A resolution. The coordination geometry in the free Cd(II)-substituted enzyme is highly similar to that of the native enzyme. Cd(II) is bound to Cys-46, Cys-174, His-67, and a water molecule in a distorted tetrahedral geometry. Binding of coenzymes induces a conformational change similar to that in the native enzyme. The interactions between the coenzyme and the protein in the binary and ternary complexes are highly similar to those in the native ternary complexes. The substrate binds directly to the cadmium ion in a distorted tetrahedral geometry. No large, significant structural changes compared to the native ternary complex with coenzyme and p-bromobenzyl alcohol were found. The implications of these results for the use of active site specific Cd(II)-substituted horse liver alcohol dehydrogenase as a model system for the native enzyme are discussed.  相似文献   

10.
Cysteine protease SpeB is secreted from Streptococcus pyogenes and has been studied as a potential virulence factor since its identification almost 70 years ago. Here, we report the crystal structures of apo mature SpeB to 1.06 Å resolution as well as complexes with the general cysteine protease inhibitor trans-epoxysuccinyl-l-leucylamido(4-guanidino)butane and a novel substrate mimetic peptide inhibitor. These structures uncover conformational changes associated with maturation of SpeB from the inactive zymogen to its active form and identify the residues required for substrate binding. With the use of a newly developed fluorogenic tripeptide substrate to measure SpeB activity, we determined IC50 values for trans-epoxysuccinyl-l-leucylamido(4-guanidino)butane and our new peptide inhibitor and the effects of mutations within the C-terminal active site loop. The structures and mutational analysis suggest that the conformational movements of the glycine-rich C-terminal loop are important for the recognition and recruitment of biological substrates and release of hydrolyzed products.  相似文献   

11.
The catalytic (C) subunit of cAMP-dependent protein kinase [protein kinase A (PKA)] is a major target of cAMP signaling, and its regulation is of fundamental importance to biological processes. One mode of regulation is N-myristylation, which has eluded structural and functional characterization so far because most crystal structures are of the non-myristylated enzyme, are phosphorylated on Ser10, and generally lack electron density for the first 13 residues. We crystallized myristylated wild-type (WT) PKA and a K7C mutant as binary (bound to a substrate peptide) and ternary [bound to a substrate peptide and adenosine-5′-(β,γ-imido)triphosphate] complexes. There was clear electron density for the entire N-terminus in the binary complexes, both refined to 2.0 Å, and K7C ternary complex, refined to 1.35 Å. The N-termini in these three structures display a novel conformation with a previously unseen helix from residues 1 to 7. The K7C mutant appears to have a more stable N-terminus, and this correlated with a significant decrease in the B-factors for the N-terminus in the myr-K7C complexes compared to the WT binary complex. The N-terminus of the myristylated WT ternary complex, refined to 2.0 Å, was disordered as in previous structures. In addition to a more ordered N-terminus, the myristylated K7C mutant exhibited a 53% increase in kcat. The effect of nucleotide binding on the structure of the N-terminus in the WT protein and the kinetic changes in the K7C protein suggest that myristylation or occupancy of the myristyl binding pocket may serve as a site for allosteric regulation in the C-subunit.  相似文献   

12.
Orotate phosphoribosyltransferase (OPRTase, EC 2.4.2.10) catalyzes the Mg2+-dependent condensation of orotic acid (OA) with PRPP (5-alpha-d-phosphorylribose 1-diphosphate) to yield diphosphate (PPi) and the nucleotide OMP (orotidine 5'-monophosphate). We have determined the structures of three forms of Saccharomyces cerevisiae OPRTase representing different structural and enzymatic intermediates. The structures include the apoenzyme (2.35 A resolution); a ternary complex of enzyme, Mg2+-PRPP, and OA (1.74 A resolution); and the binary product complex of enzyme with OMP (1.89 A resolution). While the overall structure of the S. cerevisiae OPRTase is similar to that of the Salmonella typhimurium enzyme, as judged by comparison of the two apoenzymes, large conformational transitions occur proceeding from the apoenzyme structure to those of the substrate and product complexes. Comparison of these structures reveals a rotation of the upper hood domain onto the bound ligands by an average of 19.5 degrees in the OMP structure and an average of 24.6 degrees in the OA/Mg2+-PRPP ternary complex. As expected, the conserved loop, composed of residues 104-116, moves extensively and adopts a single stable conformation during the catalytic cycle in order to sequester the substrates from bulk solvent in the ternary complex. The OA and Mg2+-PRPP molecules bound in the ternary complex are oriented for proper attack of the N1 atom of OA onto the C1 atom of the ribose ring. This orientation of substrates, combined with the positioning of the flexible loop, provides a clear picture of a catalytically poised reaction complex for type I phosphoribosyltransferases. The structural asymmetry present in these structures, as well as that found in a recent structure of the S. typhimurium enzyme, combined with the closure of the flexible loop from one subunit into the active site of the opposing subunit in the ternary complex is consistent with the kinetic data [McClard, R. W., et al. (2006) Biochemistry 45, 5330-5342] that demonstrate induced nonequivalence and cooperativity of OPRTase.  相似文献   

13.
The crystal structure of the complex formed by bovine trypsin and bovine pancreatic trypsin inhibitor has been refined with data to 1.9 Å resolution, using a procedure described by Deisenhofer &; Steigemann (1974) in their refinement of the crystal structure of the free inhibitor. This procedure involves cycles consisting of phase calculation using the current atomic model, Fourier synthesis using these phases and the observed structure factor amplitudes and Diamond's real-space refinement (Diamond, 1971,1974). At various stages, difference Fourier syntheses are calculated to detect and correct gross errors in the model and to localize solvent molecules.The refinement progressed smoothly, starting with the model obtained from the isomorphous Fourier map at 2.6 Å resolution. The R-factor is 0.23 for 20,500 significantly measured reflections to 1.9 Å resolution, using an over-all temperature factor of 20 Å2. The estimated standard deviation of atomic positions is 0.09 Å.An objective assessment of the upper limit of the error in the atomic coordinates of the final model is possible by comparing the inhibitor component in the model of the complex with the refined structure of the free inhibitor (Deisenhofer &; Steigemann, 1974). The mean deviation of main-chain atoms of the two molecular models in internal segments is 0.25 Å, of main-chain dihedral angles 5.1 ° and side-chain dihedral angles 6.5 °.A comparison of the trypsin component with α-chymotrypsin (Birktoft &; Blow, 1972) showed a mean deviation of main-chain atoms of 0.75 Å. The structures are closely similar and the various deletions and insertions cause local structural differences only.  相似文献   

14.
We report crystal structures of ternary complexes of maltodextrin phosphorylase with natural oligosaccharide and phosphate mimicking anions: nitrate, sulphate and vanadate. Electron density maps obtained from crystals grown in presence of Al(NO3)3 show a nitrate ion instead of the expected in the catalytic site. The trigonal is coplanar with the Arg569 guanidinium group and mimics three of the four oxygen atoms of phosphate. The ternary complex with sulphate shows a partial occupancy of the anionic site. The low affinity of the sulphate ion, observed when the α-glucosyl substrate is present in the catalytic channel, is ascribed to restricted space for the anion. Even lower occupancy is observed for the larger vanadate anion. The Malp/G5/ structure shows the partial occupancy of the oligosaccharide and the dislocation of the 380’s loop. This has been attributed to the formation of oligosaccharide vanadate derivatives (confirmed by capillary electrophoresis) that reduces their effective concentration. The difficulty to trap a ternary complex mimicking the ground state has been correlated to the apparent lower affinity that natural substrates show regarding the intermediates of the enzymatic reaction.  相似文献   

15.
A key step in plant photorespiration, the oxidation of glycolate to glyoxylate, is carried out by the peroxisomal flavoprotein glycolate oxidase (EC 1.1.3.15). The three-dimensional structure of this alpha/beta barrel protein has been refined to 2 A resolution (Lindqvist Y. 1989. J Mol Biol 209:151-166). FMN dependent glycolate oxidase is a member of the family of alpha-hydroxy acid oxidases. Here we describe the crystallization and structure determination of two inhibitor complexes of the enzyme, TKP (3-Decyl-2,5-dioxo-4-hydroxy-3-pyrroline) and TACA (4-Carboxy-5-(1-pentyl)hexylsulfanyl-1,2,3-triazole). The structure of the TACA complex has been refined to 2.6 A resolution and the TKP complex, solved with molecular replacement, to 2.2 A resolution. The Rfree for the TACA and TKP complexes are 24.2 and 25.1%, respectively. The overall structures are very similar to the unliganded holoenzyme, but a closer examination of the active site reveals differences in the positioning of the flavin isoalloxazine ring and a displaced flexible loop in the TKP complex. The two inhibitors differ in binding mode and hydrophobic interactions, and these differences are reflected by the very different Ki values for the inhibitors, 16 nM for TACA and 4.8 microM for TKP. Implications of the structures of these enzyme-inhibitor complexes for the model for substrate binding and catalysis proposed from the holo-enzyme structure are discussed.  相似文献   

16.
Kunitz-type trypsin inhibitors bind to the active pocket of trypsin causing its inhibition. Plant Kunitz-type inhibitors are thought to be important in defense, especially against insect pests. From sequence analysis of various Kunitz-type inhibitors from plants, we identified CaTI2 from chickpea as a unique variant lacking the functionally important arginine residue corresponding to the soybean trypsin inhibitor (STI) and having a distinct and unique inhibitory loop organization. To further explore the implications of these sequence variations, we obtained the crystal structure of recombinant CaTI2 at 2.8Å resolution. It is evident from the structure that the variations in the inhibitory loop facilitates non-substrate like binding of CaTI2 to trypsin, while the canonical inhibitor STI binds to trypsin in substrate like manner. Our results establish the unique mechanism of trypsin inhibition by CaTI2, which warrant further research into its substrate spectrum. Abbreviations BApNA Nα-Benzoyl-L-arginine 4-nitroanilide

BPT bovine pancreatic trypsin

CaTI2 Cicer arietinum L trypsin inhibitor 2

DrTI Delonix regia Trypsin inhibitor

EcTI Enterolobium contortisiliquum trypsin inhibitor

ETI Erythrina caffra trypsin inhibitor

KTI Kunitz type inhibitor

STI soybean trypsin inhibitor

TKI Tamarindus indica Kunitz inhibitor

Communicated By Ramaswamy H. Sarma  相似文献   


17.
Human α‐amino‐β‐carboxymuconate‐ε‐semialdehyde decarboxylase determines the fate of tryptophan metabolites in the kynurenine pathway by controlling the quinolinate levels for de novo nicotinamide adenine dinucleotide biosynthesis. The unstable nature of its substrate has made gaining insight into its reaction mechanism difficult. Our electron paramagnetic resonance (EPR) spectroscopic study on the Cu‐substituted human enzyme suggests that the native substrate does not directly ligate to the metal ion. Substrate binding did not result in a change of either the hyperfine structure or the super‐hyperfine structure of the EPR spectrum. We also determined the crystal structure of the human enzyme in its native catalytically active state (at 1.99 Å resolution), a substrate analogue‐bound form (2.50 Å resolution), and a selected active site mutant form with one of the putative substrate binding residues altered (2.32 Å resolution). These structures illustrate that each asymmetric unit contains three pairs of dimers. Consistent with the EPR findings, the ligand‐bound complex structure shows that the substrate analogue does not directly coordinate to the metal ion but is bound to the active site by two arginine residues through noncovalent interactions. Proteins 2015; 83:178–187. © 2014 Wiley Periodicals, Inc.  相似文献   

18.
19.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid–base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR–Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

20.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid-base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR-Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  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