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 46 毫秒
1.
The crystal structure of the dimeric catalytic domain (residues 118-424) of human PheOH (hPheOH), cocrystallized with the oxidized form of the cofactor (7,8-dihydro-L-biopterin, BH(2)), has been determined at 2.0 A resolution. The pterin binds in the second coordination sphere of the catalytic iron (the C4a atom is 6.1 A away), and interacts through several hydrogen bonds to two water molecules coordinated to the iron, as well as to the main chain carbonyl oxygens of Ala322, Gly247, and Leu249 and the main chain amide of Leu249. Some important conformational changes are seen in the active site upon pterin binding. The loop between residues 245 and 250 moves in the direction of the iron, and thus allows for several important hydrogen bonds to the pterin ring to be formed. The pterin cofactor is in an ideal orientation for dioxygen to bind in a bridging position between the iron and the pterin. The pterin ring forms an aromatic pi-stacking interaction with Phe254, and Tyr325 contributes to the positioning of the pterin ring and its dihydroxypropyl side chain by hydrophobic interactions. Of particular interest in the hPheOH x BH(2) binary complex structure is the finding that Glu286 hydrogen bonds to one of the water molecules coordinated to the iron as well as to a water molecule which hydrogen bonds to N3 of the pterin ring. Site-specific mutations of Glu286 (E286A and E286Q), Phe254 (F254A and F254L), and Tyr325 (Y325F) have confirmed the important contribution of Glu286 and Phe254 to the normal positioning of the pterin cofactor and catalytic activity of hPheOH. Tyr325 also contributes to the correct positioning of the pterin, but has no direct function in the catalytic reaction, in agreement with the results obtained with rat TyrOH [Daubner, S. C., and Fitzpatrick, P. F. (1998) Biochemistry 37, 16440-16444]. Superposition of the binary hPheOH.BH(2) complex onto the crystal structure of the ligand-free rat PheOH (which contains the regulatory and catalytic domains) [Kobe, B., Jennings, I. G., House, C. M., Michell, B. J., Goodwill, K. E., Santarsiero, B. D., Stevens, R. C., Cotton, R. G. H., and Kemp, B. E. (1999) Nat. Struct. Biol. 6, 442-448] reveals that the C2'-hydroxyl group of BH(2) is sufficiently close to form hydrogen bonds to Ser23 in the regulatory domain. Similar interactions are seen with the hPheOH.adrenaline complex and Ser23. These interactions suggest a structural explanation for the specific regulatory properties of the dihydroxypropyl side chain of BH(4) (negative effector) in the full-length enzyme in terms of phosphorylation of Ser16 and activation by L-Phe.  相似文献   

2.
Phenylalanine hydroxylase (PAH) is a tetrahydrobiopterin-dependent enzyme that catalyzes the hydroxylation of L-phenylalanine (L-Phe) to L-tyrosine using dioxygen as an additional substrate. The requirement of PAH for a cofactor is absolute, but several cofactor analogs are able to substitute the natural cofactor in catalysis. However, it is only the natural cofactor 6R-tetrahydrobiopterin (6R-BH(4)) that induces a negative regulatory effect on the enzyme. In order to get further insights on the molecular basis for this specificity, we studied the structure of the cofactor-enzyme complex and the conformational changes induced by cofactor binding by molecular dynamics simulations. Simulations were carried out on the enzyme alone and complexed with 6R-BH(4) and with two cofactor analogs, 6S-BH(4) and 6-methyl-tetrahydropterin (6M-PH(4)). In the resting unbound enzyme Tyr377 in the catalytic domain is hydrogen bonded to both Ser23 and Glu21 of the autoregulatory N-terminal sequence. This hydrogen bonding network is disturbed by the binding of BH(4), which interacts with Ser23. By doing so, 6R-BH(4) facilitates an interaction between Glu21 and the active site iron, further pulling the N-terminal into the active site of PAH and blocking the L-Phe binding site. Thus, in the 6R-BH(4) complexed enzyme, the N-terminal functions as an intrinsic amino acid regulatory sequence (IARS). Neither 6M-PH(4) nor 6S-BH(4) can interact favorably with Ser23, and do not induce an inhibitory effect on PAH. These simulations thus explain the previous findings that the two hydroxyl groups in the side chain of the 6R epimer of BH(4) are essential for the inhibitory regulatory effect on PAH.  相似文献   

3.
Phenylalanine hydroxylase (PAH) is a pterin-dependent non-heme metalloenzyme that catalyzes the oxidation of phenylalanine to tyrosine, which is the rate-limiting step in the catabolism of Phe. Chromobacterium violaceum phenylalanine hydroxylase (cPAH) has been prepared and its steady-state mechanism has been investigated. The enzyme requires iron for maximal activity. Initial rate measurements, done in the presence of the 6,7-dimethyl-5,6,7,8-tetrahydropterin (DMPH(4)) cofactor, yielded an average apparent k(cat) of 36+/-1 s(-1). The apparent K(M) values measured for the substrates DMPH(4), L-Phe, and O(2) are 44+/-7, 59+/-10, and 76+/-7 microM, respectively. Steady-state kinetic analyses using double-reciprocal plots revealed line patterns consistent with a sequential ter-bi mechanism in which L-Phe is the middle substrate in the order of binding. The occurrence of a line intersection on the double-reciprocal plot abscissa when either pterin or O(2) is saturated suggests that, prior to O(2) binding, DMPH(4) and L-Phe are in associative pre-equilibrium with cPAH. Together with an inhibition study using the oxidized cofactor, 7,8-dimethyl-6,7-dihydropterin, it is conclusive that the mechanism is fully ordered, with DMPH(4) binding the active site first, L-Phe second, and O(2) last. This represents the first conclusive steady-state mechanism for a PAH enzyme.  相似文献   

4.
Phenylalanine hydroxylase (PAH) is a tetrahydrobiopterin and non-heme iron-dependent enzyme that hydroxylates L-Phe to l-Tyr using molecular oxygen as additional substrate. A dysfunction of this enzyme leads to phenylketonuria (PKU). The conformation and distances to the catalytic iron of both L-Phe and the cofactor analogue L-erythro-7,8-dihydrobiopterin (BH2) simultaneously bound to recombinant human PAH have been estimated by (1)H NMR. The resulting bound conformers of both ligands have been fitted into the crystal structure of the catalytic domain by molecular docking. In the docked structure L-Phe binds to the enzyme through interactions with Arg270, Ser349 and Trp326. The mode of coordination of Glu330 to the iron moiety seems to determine the amino acid substrate specificity in PAH and in the homologous enzyme tyrosine hydroxylase. The pterin ring of BH2 pi-stacks with Phe254, and the N3 and the amine group at C2 hydrogen bond with the carboxylic group of Glu286. The ring also establishes specific contacts with His264 and Leu249. The distance between the O4 atom of BH2 and the iron (2.6(+/-0.3) A) is compatible with coordination, a finding that is important for the understanding of the mechanism of the enzyme. The hydroxyl groups in the side-chain at C6 hydrogen bond with the carbonyl group of Ala322 and the hydroxyl group of Ser251, an interaction that seems to have implications for the regulation of the enzyme by substrate and cofactor. Some frequent mutations causing PKU are located at residues involved in substrate and cofactor binding. The sites for hydroxylation, C4 in L-Phe and C4a in the pterin are located at a distance of 4.2 and 4.3 A from the iron moiety, respectively, and at 6.3 A from each other. These distances are adequate for the intercalation of iron-coordinated molecular oxygen, in agreement with a mechanistic role of the iron moiety both in the binding and activation of dioxygen and in the hydroxylation reaction.  相似文献   

5.
Phosphorylation of phenylalanine hydroxylase (PAH) at Ser16 by cAMP-dependent protein kinase increases the basal activity of the enzyme and its resistance to tryptic proteolysis. The modeled structures of the full-length phosphorylated and unphosphorylated enzyme were subjected to molecular dynamics simulations, and we analyzed the energy of charge-charge interactions for individual ionizable residues in the final structures. These calculations showed that the conformational changes induced by incorporation of phosphate were localized and limited mostly to the region around the phosphoserine (Arg13-Asp17) and a region around the active site in the catalytic domain that includes residues involved in the binding of the iron and the substrate L-Phe (Arg270 and His285). The absence of a generalized conformational change was confirmed by differential scanning calorimetry, thermal-dependent circular dichroism, fluorescence spectroscopy, and limited chymotryptic proteolysis of the phosphorylated and unphosphorylated PAH. Our results explain the effect of phosphorylation of PAH on both the resistance to proteolysis specifically by trypsin-like enzymes and on the increase in catalytic efficiency.  相似文献   

6.
Tryptophan hydroxylase (TPH) carries out the 5-hydroxylation of L-Trp, which is the rate-limiting step in the synthesis of serotonin. We have prepared and characterized a stable N-terminally truncated form of human TPH that includes the catalytic domain (Delta90TPH). We have also determined the conformation and distances to the catalytic non-heme iron of both L-Trp and the tetrahydrobiopterin cofactor analogue L-erythro-7,8-dihydrobiopterin (BH2) bound to Delta90TPH by using 1H NMR spectroscopy. The bound conformers of the substrate and the pterin were then docked into the modeled three-dimensional structure of TPH. The resulting ternary TPH-BH2-L-Trp structure is very similar to that previously determined by the same methods for the complex of phenylalanine hydroxylase (PAH) with BH2 and L-Phe [Teigen, K., et al. (1999) J. Mol. Biol. 294, 807-823]. In the model, L-Trp binds to the enzyme through interactions with Arg257, Ser336, His272, Phe318, and Phe313, and the ring of BH2 interacts mainly with Phe241 and Glu273. The distances between the hydroxylation sites at C5 in L-Trp and C4a in the pterin, i.e., 6.1 +/- 0.4 A, and from each of these sites to the iron, i.e., 4.1 +/- 0.3 and 4.4 +/- 0.3 A, respectively, are also in agreement with the formation of a transient iron-4a-peroxytetrahydropterin in the reaction, as proposed for the other hydroxylases. The different conformation of the dihydroxypropyl chain of BH2 in PAH and TPH seems to be related to the presence of nonconserved residues, i.e., Tyr235 and Pro238 in TPH, at the cofactor binding site. Moreover, Phe313, which seems to interact with the substrate through ring stacking, corresponds to a Trp residue in both tyrosine hydroxylase and PAH (Trp326) and appears to be an important residue for influencing the substrate specificity in this family of enzymes. We show that the W326F mutation in PAH increases the relative preference for L-Trp as the substrate, while the F313W mutation in TPH increases the preference for L-Phe, possibly by a conserved active site volume effect.  相似文献   

7.
Abstract

Phenylalanine hydroxylase (PAH) is a tetrahydrobiopterin-dependent enzyme that catalyzes the hydroxylation of L-phenylalanine (L-Phe) to L-tyrosine using dioxygen as an additional substrate. The requirement of PAH for a cofactor is absolute, but several cofactor analogs are able to substitute the natural cofactor in catalysis. However, it is only the natural cofactor 6R-tetrahydrobiopterin (6R-BH4) that induces a negative regulatory effect on the enzyme. In order to get further insights on the molecular basis for this specificity, we studied the structure of the cofactor-enzyme complex and the conformational changes induced by cofactor binding by molecular dynamics simulations. Simulations were carried out on the enzyme alone and complexed with 6R-BH4 and with two cofactor analogs, 6S-BH4 and 6-methyl-tetrahydropterin (6M-PH4). In the resting unbound enzyme Tyr377 in the catalytic domain is hydrogen bonded to both Ser23 and Glu21 of the autoregulatory N-terminal sequence. This hydrogen bonding network is disturbed by the binding of BH4, which interacts with Ser23. By doing so, 6R-BH4 facilitates an interaction between Glu21 and the active site iron, further pulling the N-terminal into the active site of PAH and blocking the L-Phe binding site. Thus, in the 6R-BH4 complexed enzyme, the N-terminal functions as an intrinsic amino acid regulatory sequence (IARS). Neither 6M-PH4 nor 6S-BH4 can interact favorably with Ser23, and do not induce an inhibitory effect on PAH. These simulations thus explain the previous findings that the two hydroxyl groups in the side chain of the 6R epimer of BH4 are essential for the inhibitory regulatory effect on PAH.  相似文献   

8.
The crystal structures of the catalytic domain (DeltaN1-102/DeltaC428-452) of human phenylalanine hydroxylase (hPheOH) in its catalytically competent Fe(II) form and binary complex with the reduced pterin cofactor 6(R)-L-erythro-5,6,7,8-tetrahydrobiopterin (BH4) have been determined to 1.7 and 1.5 A, respectively. When compared with the structures reported for various catalytically inactive Fe(III) forms, several important differences have been observed, notably at the active site. Thus, the non-liganded hPheOH-Fe(II) structure revealed well defined electron density for only one of the three water molecules reported to be coordinated to the iron in the high-spin Fe(III) form, as well as poor electron density for parts of the coordinating side-chain of Glu330. The reduced cofactor (BH4), which adopts the expected half-semi chair conformation, is bound in the second coordination sphere of the catalytic iron with a C4a-iron distance of 5.9 A. BH4 binds at the same site as L-erythro-7,8-dihydrobiopterin (BH2) in the binary hPheOH-Fe(III)-BH2 complex forming an aromatic pi-stacking interaction with Phe254 and a network of hydrogen bonds. However, compared to that structure the pterin ring is displaced about 0.5 A and rotated about 10 degrees, and the torsion angle between the hydroxyl groups of the cofactor in the dihydroxypropyl side-chain has changed by approximately 120 degrees enabling O2' to make a strong hydrogen bond (2.4 A) with the side-chain oxygen of Ser251. Carbon atoms in the dihydroxypropyl side-chain make several hydrophobic contacts with the protein. The iron is six-coordinated in the binary complex, but the overall coordination geometry is slightly different from that of the Fe(III) form. Most important was the finding that the binding of BH4 causes the Glu330 ligand to change its coordination to the iron when comparing with non-liganded hPheOH-Fe(III) and the binary hPheOH-Fe(III)-BH2 complex.  相似文献   

9.
The crystal structures of the catalytic domain of human phenylalanine hydroxylase (hPheOH) in complex with the physiological cofactor 6(R)-L-erythro-5,6,7,8-tetrahydrobiopterin (BH(4)) and the substrate analogues 3-(2-thienyl)-L-alanine (THA) or L-norleucine (NLE) have been determined at 2.0A resolution. The ternary THA complex confirms a previous 2.5A structure, and the ternary NLE complex shows that similar large conformational changes occur on binding of NLE as those observed for THA. Both structures demonstrate that substrate binding triggers structural changes throughout the entire protomer, including the displacement of Tyr138 from a surface position to a buried position at the active site, with a maximum displacement of 20.7A for its hydroxyl group. Two hinge-bending regions, centred at Leu197 and Asn223, act in consort upon substrate binding to create further large structural changes for parts of the C terminus. Thus, THA/L-Phe binding to the active site is likely to represent the epicentre of the global conformational changes observed in the full-length tetrameric enzyme. The carboxyl and amino groups of THA and NLE are positioned identically in the two structures, supporting the conclusion that these groups are of key importance in substrate binding, thus explaining the broad non-physiological substrate specificity observed for artificially activated forms of the enzyme. However, the specific activity with NLE as the substrate was only about 5% of that with THA, which is explained by the different affinities of binding and different catalytic turnover.  相似文献   

10.
Previous studies of ferrous wild-type phenylalanine hydroxylase, [Fe(2+)]PAH(T)[], have shown the active site to be a six-coordinate distorted octahedral site. After the substrate and cofactor bind to the enzyme ([Fe(2+)]PAH(R)[L-Phe,5-deaza-6-MPH(4)]), the active site converts to a five-coordinate square pyramidal structure in which the identity of the missing ligand had not been previously determined. X-ray absorption spectroscopy (XAS) at the Fe K-edge further supports this coordination number change with the binding of both cosubstrates to the enzyme, and determines this to be due to the loss of a water ligand.  相似文献   

11.
The paramagnetic iron at the active site of highly purified, catalytically active phenylalanine hydroxylase was studied by EPR at 3.6 K and one-dimensional 1H-NMR spectroscopy at 293 K. The EPR-detectable iron of the bovine enzyme was found to be present as a high-spin form (S = 5/2) in different ligand field symmetries depending on medium conditions (buffer ions) and the presence of ligands known to bind at the active site. At 3.6 K and in phosphate buffer, the paramagnetic iron is coordinated in an environment of rhombic symmetry (g = 4.3), whereas Tris buffer favours an environment of axial ligand field symmetry (g = 6.7, 5.3 and 2.0). The latter axial type of signals resembles those observed at g = 7.0, 5.2 and 1.9 for the enzyme in phosphate buffer when L-noradrenaline is added as an active-site ligand (inhibitor). The same proportion of iron that coordinates to L-noradrenaline seems to be reduced by the pterin cofactor and participate in catalysis. Experimental evidence is presented that Tris inhibits the enzyme by interacting with the enzyme-bound ferric iron and decreases its rate of reduction by the tetrahydropterin cofactor. Preincubation with dithiothreitol also inhibits the enzyme activity and prevents the reduction of its catalytically active ferric iron by pterin cofactors as well as binding of catecholamines to the enzyme. 1H-NMR spectroscopy revealed that the substrate (L-phenylalanine) and L-noradrenaline bind close to the paramagnetic iron, and that the catecholamine displaces the substrate from its binding at the active site. The results support our recently proposed model for the cooperative binding of inhibitor and substrate at the active site [Martínez, A. et al. (1990) Eur. J. Biochem. 193, 211-219].  相似文献   

12.
Escherichia coli dihydrofolate reductase (DHFR) has several flexible loops surrounding the active site that play a functional role in substrate and cofactor binding and in catalysis. We have used heteronuclear NMR methods to probe the loop conformations in solution in complexes of DHFR formed during the catalytic cycle. To facilitate the NMR analysis, the enzyme was labeled selectively with [(15)N]alanine. The 13 alanine resonances provide a fingerprint of the protein structure and report on the active site loop conformations and binding of substrate, product, and cofactor. Spectra were recorded for binary and ternary complexes of wild-type DHFR bound to the substrate dihydrofolate (DHF), the product tetrahydrofolate (THF), the pseudosubstrate folate, reduced and oxidized NADPH cofactor, and the inactive cofactor analogue 5,6-dihydroNADPH. The data show that DHFR exists in solution in two dominant conformational states, with the active site loops adopting conformations that closely approximate the occluded or closed conformations identified in earlier X-ray crystallographic analyses. A minor population of a third conformer of unknown structure was observed for the apoenzyme and for the disordered binary complex with 5,6-dihydroNADPH. The reactive Michaelis complex, with both DHF and NADPH bound to the enzyme, could not be studied directly but was modeled by the ternary folate:NADP(+) and dihydrofolate:NADP(+) complexes. From the NMR data, we are able to characterize the active site loop conformation and the occupancy of the substrate and cofactor binding sites in all intermediates formed in the extended catalytic cycle. In the dominant kinetic pathway under steady-state conditions, only the holoenzyme (the binary NADPH complex) and the Michaelis complex adopt the closed loop conformation, and all product complexes are occluded. The catalytic cycle thus involves obligatory conformational transitions between the closed and occluded states. Parallel studies on the catalytically impaired G121V mutant DHFR show that formation of the closed state, in which the nicotinamide ring of the cofactor is inserted into the active site, is energetically disfavored. The G121V mutation, at a position distant from the active site, interferes with coupled loop movements and appears to impair catalysis by destabilizing the closed Michaelis complex and introducing an extra step into the kinetic pathway.  相似文献   

13.
Phenylalanine hydroxylase converts phenylalanine to tyrosine, a rate-limiting step in phenylalanine catabolism and protein and neurotransmitter biosynthesis. It is tightly regulated by the substrates phenylalanine and tetrahydrobiopterin and by phosphorylation. We present the crystal structures of dephosphorylated and phosphorylated forms of a dimeric enzyme with catalytic and regulatory properties of the wild-type protein. The structures reveal a catalytic domain flexibly linked to a regulatory domain. The latter consists of an N-terminal autoregulatory sequence (containing Ser 16, which is the site of phosphorylation) that extends over the active site pocket, and an alpha-beta sandwich core that is, unexpectedly, structurally related to both pterin dehydratase and the regulatory domains of metabolic enzymes. Phosphorylation has no major structural effects in the absence of phenylalanine, suggesting that phenylalanine and phosphorylation act in concert to activate the enzyme through a combination of intrasteric and possibly allosteric mechanisms.  相似文献   

14.
Nitric oxide is a key signaling molecule in many biological processes, making regulation of nitric oxide levels highly desirable for human medicine and for advancing our understanding of basic physiology. Designing inhibitors to specifically target one of the three nitric oxide synthase (NOS) isozymes that form nitric oxide from the L-Arg substrate poses a significant challenge due to the overwhelmingly conserved active sites. We report here 10 new X-ray crystallographic structures of inducible and endothelial NOS oxygenase domains cocrystallized with chlorzoxazone and four nitroindazoles: 5-nitroindazole, 6-nitroindazole, 7-nitroindazole, and 3-bromo-7-nitroindazole. Each of these bicyclic aromatic inhibitors has only one hydrogen bond donor and therefore cannot form the bidentate hydrogen bonds that the L-Arg substrate makes with Glu371. Instead, all of these inhibitors induce a conformational change in Glu371, creating an active site with altered molecular recognition properties. The cost of this conformational change is approximately 1-2 kcal, based on our measured constants for inhibitor binding to the wild-type and E371A mutant proteins. These inhibitors derive affinity by pi-stacking above the heme and replacing both intramolecular (Glu371-Met368) and intermolecular (substrate-Trp366) hydrogen bonds to the beta-sheet architecture underlying the active site. When bound to NOS, high-affinity inhibitors in this class are planar, whereas weaker inhibitors are nonplanar. Isozyme differences were observed in the pterin cofactor site, the heme propionate, and inhibitor positions. Computational docking predictions match the crystallographic results, including the Glu371 conformational change and inhibitor-binding orientations, and support a combined crystallographic and computational approach to isozyme-specific NOS inhibitor analysis and design.  相似文献   

15.
The effect of spermine on tyrosine hydroxylase (TH) activity purified from bovine adrenal medulla was examined before and after phosphorylation by the catalytic subunit of cyclic AMP-dependent protein kinase (A-kinase). Before phosphorylation, spermine (less than 1 mM) inhibited the enzymatic activity, and negative cooperative effect of spermine on TH (Hill coefficient = 0.7) was observed from the kinetic analysis concerning 6-methyl-5,6,7,8-tetrahydropterin (6MPH4). Spermine interacted noncompetitively toward tyrosine and the Ki for spermine was calculated to be 68 microM. Phosphorylation abolished the ability of spermine to inhibit TH activity in a negative cooperative manner against the pterin cofactor, and also increased four-fold the Ki value against the substrate. These results suggest that spermine may inhibit TH activity by interacting with the pterin binding site of the enzyme molecule in a manner of negative cooperativity, and that this inhibition is reversed by the conformational change of regulatory domain of TH after phosphorylation by A-kinase.  相似文献   

16.
A Kamb  J S Finer-Moore  R M Stroud 《Biochemistry》1992,31(51):12876-12884
We have solved crystal structures of two complexes with Escherichia coli thymidylate synthase (TS) bound either to the cofactor analog N10-propargyl-5,8-dideazafolate (CB3717) or to a tighter binding polygutamyl derivative of CB3717. These structures suggest that cofactor binding alone is sufficient to induce the conformational change in TS; dUMP binding is not required. Because polyglutamyl folates are the primary cofactor form in vivo, and because they can bind more tightly than dUMP to TS, these structures may represent a key intermediate along the TS reaction pathway. These structures further suggest that the dUMP binding site is accessible in the TS-cofactor analog binary complexes. Conformational flexibility of the binary complex may permit dUMP to enter the active site of TS while the cofactor is bound. Alternatively, dUMP may enter the active site from the opposite side that the cofactor appears to enter; that is, through a portal flanked by arginines that also coordinate the phosphate group in the active site. Entry of dUMP through this portal may allow dUMP to bind to a TS-cofactor binary complex in which the complex has completed its conformational transition to the catalytically competent structure.  相似文献   

17.
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   

18.
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   

19.
A C Price  Y M Zhang  C O Rock  S W White 《Biochemistry》2001,40(43):12772-12781
The structure of beta-ketoacyl-[acyl carrier protein] reductase (FabG) from Escherichia coli was determined via the multiwavelength anomalous diffraction technique using a selenomethionine-labeled crystal containing 88 selenium sites in the asymmetric unit. The comparison of the E. coli FabG structure with the homologous Brassica napus FabG.NADP(+) binary complex reveals that cofactor binding causes a substantial conformational change in the protein. This conformational change puts all three active-site residues (Ser 138, Tyr 151, and Lys 155) into their active configurations and provides a structural mechanism for allosteric communication between the active sites in the homotetramer. FabG exhibits negative cooperative binding of NADPH, and this effect is enhanced by the presence of acyl carrier protein (ACP). NADPH binding also increases the affinity and decreases the maximum binding of ACP to FabG. Thus, unlike other members of the short-chain dehydrogenase/reductase superfamily, FabG undergoes a substantial conformational change upon cofactor binding that organizes the active-site triad and alters the affinity of the other substrate-binding sites in the tetrameric enzyme.  相似文献   

20.
Heteronuclear NMR methods have been used to probe the conformation of four complexes of Escherichia coli dihydrofolate reductase (DHFR) in solution. (1)H(N), (15)N, and (13)C(alpha) resonance assignments have been made for the ternary complex with folate and oxidized NADP(+) cofactor and the ternary complex with folate and a reduced cofactor analog, 5,6-dihydroNADPH. The backbone chemical shifts have been compared with those of the binary complex of DHFR with the substrate analog folate and the binary complex with NADPH (the holoenzyme). Analysis of (1)H(N) and (15)N chemical shifts has led to the identification of marker resonances that report on the active site conformation of the enzyme. Other backbone amide resonances report on the presence of ligands in the pterin binding pocket and in the adenosine and nicotinamide-ribose binding sites of the NADPH cofactor. The chemical shift data indicate that the enzyme populates two dominant structural states in solution, with the active site loops in either the closed or occluded conformations defined by X-ray crystallography; there is no evidence that the open conformation observed in some X-ray structures of E. coli DHFR are populated 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