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 0 毫秒
1.
Frederick KK  Palfey BA 《Biochemistry》2005,44(40):13304-13314
p-Hydroxybenzoate hydroxylase (PHBH) is an FAD-dependent monooxygenase that catalyzes the hydroxylation of p-hydroxybenzoate (pOHB) to 3,4-dihydroxybenzoate in an NADPH-dependent reaction. Two structural features are coupled to control the reactivity of PHBH with NADPH: a proton-transfer network that allows protons to be passed between the sequestered active site and solvent and a flavin that adopts two positions: "in", where the flavin is near pOHB, and "out", where the flavin is near NADPH. PHBH uses the proton-transfer network to test for the presence of a suitable aromatic substrate before allowing the flavin to adopt the NADPH-accessible conformation. In this work, kinetic analysis of the His72Asn mutant, with a disrupted proton-transfer network, showed that flavin movement could occur in the presence or absence of NADPH but that NADPH stimulated movement to the reactive conformation required for hydride transfer. Substrate and solvent isotope effects on the transient kinetics of reduction of the His72Asn mutant showed that proton transfer was linked to flavin movement and that the conformational change occurred in a step separate from that of hydride transfer. Proton transfers during the reductive half-reaction were observed directly in the wild-type enzyme by performing experiments in the presence of a fluorescent pH-indicator dye in unbuffered solutions. NADPH binding caused rapid proton release from the enzyme, followed by proton uptake after flavin reduction. Solvent and substrate kinetic isotope effects showed that proton-coupled flavin movement and reduction also occurred in different steps in wild-type PHBH. These results allow a detailed kinetic scheme to be proposed for the reductive half-reaction of the wild-type enzyme. Three kinetic models considered for substrate-induced isomerization are analyzed in the Appendix.  相似文献   

2.
The fly homeodomain (HD) protein EXTRADENTICLE (EXD) is dependent on a second HD protein, HOMOTHORAX (HTH), for nuclear localization. We show here that in insect cells the mammalian homolog of EXD, PBX1A, shows a similar dependence on the HTH homologs MEIS1, 2, and 3 and the MEIS-like protein PREP1. Paradoxically, removal of residues N-terminal to the PBX1A HD abolishes interactions with MEIS/PREP but allows nuclear accumulation of PBX1A. We use deletion mapping and fusion to green fluorescent protein to map two cooperative nuclear localization signals (NLSs) in the PBX HD. The results of DNA-binding assays and pull-down experiments are consistent with a model whereby the PBX N-terminus binds to the HD and masks the two NLSs. In support of the model, a mutation in the PBX HD that disrupts contact with the N-terminus leads to constitutive nuclear localization. The HD mutation also increases sensitivity to protease digestion, consistent with a change in conformation. We propose that MEIS family proteins induce a conformational change in PBX that unmasks the NLS, leading to nuclear localization and increased DNA-binding activity. Consistent with this, PBX1 is nuclear only where Meis1 is expressed in the mouse limb bud.  相似文献   

3.
p-Hydroxybenzoate hydroxylase is a flavoprotein monooxygenase that catalyzes a reaction in two parts: reduction of the enzyme cofactor FAD by NADPH in response to binding p-hydroxybenzoate to the enzyme and reaction of reduced FAD with oxygen to form a hydroperoxide, which then oxygenates p-hydroxybenzoate. Three different reactions, each with specific requirements, are achieved by moving the position of the isoalloxazine ring in the protein structure. In this paper, we examine the operation of protein conformational changes and the significance of charge-transfer absorption bands associated with the reduction of FAD by NADPH when the substrate analogue, 5-hydroxypicolinate, is bound to the enzyme. It was discovered that the enzyme with picolinate bound was reduced at a rate similar to that with p-hydroxybenzoate bound at high pH. However, there was a large effect of pH upon the rate of reduction in the presence of picolinate with a pK(a) of 7.4, identical to the pK(a) of picolinate bound to the enzyme. The intensity of charge-transfer bands observed between FAD and NADPH during the reduction process correlated with the rate of flavin reduction. We conclude that high rates of reduction of the enzyme require (a) the isoalloxazine of the flavin be held by the protein in a solvent-exposed position and (b) the movement of a loop of protein so that the pyridine ring of NADPH can move into position to form a complex with the isoalloxazine that is competent for hydride transfer and that is indicated by a strong charge-transfer interaction.  相似文献   

4.
Zheng Y  Dong J  Palfey BA  Carey PR 《Biochemistry》1999,38(51):16727-16732
X-ray crystallographic studies of several complexes involving FAD bound to p-hydroxybenzoate hydroxylase (PHBH) have revealed that the isoalloxazine ring system of FAD is capable of adopting in two positions on the protein. In one, the "in" form, the ring is surrounded by protein groups and has little contact with solvent; in the second, "out" form, the ring is largely solvent exposed. Using Raman difference spectroscopy, it has been possible to obtain Raman spectra for the flavin ring in both conformational states for different complexes in solution. The spectra consist of a rich assortment of isoalloxazine ring modes whose normal mode origin can be assigned by using density functional theory and ab initio calculations. Further insight into the sensitivity of these modes to changes in environment is provided by the Raman spectra of lumiflavin in the solid state, in DMSO and in aqueous solution. For the protein complexes, the Raman difference spectra of flavin bound to wt PHBH and wt PHBH plus substrate, p-hydroxybenzoate, provided examples of the "in" conformation. These data are compared to those for flavin bound to wt PHBH plus 2,4-dihydroxybenzoate, where X-ray analysis show that the flavin is "out". There are several spectral regions where characteristic differences exist for flavin in the "in" or "out" conformation, these occur near 1700, 1500, 1410, 1350, 1235, and 1145 cm(-)(1). These spectral features can be used as empirical marker bands to determine the populations of "in" and "out" for any complex of PHBH and to monitor changes in those populations with perturbations to the system, e.g., by changing temperature or pH. Thus, it will now be possible to determine the conformational state of the flavin in PHBH for those complexes that have resisted X-ray crystallographic analysis. Raman difference data are also presented for the Tyr222Phe mutant. The Raman data show that the isoalloxazine ring is predominantly "out" for Tyr222Phe. However, in the presence of the substrate p-hydroxybenzoate there is clear evidence from the Raman marker bands that a mixed population of "in" and "out" exists with the majority being in the "out" state. This is consistent with the conclusions drawn from crystallographic studies on this complex (Gatti, D. L., Palfey, B. A., Lah, M. S., Entsch, B., Massey, V., Ballou, D. P., and Ludwig, M. L. (1994) Science, 266, 110-114).  相似文献   

5.
Hampel KJ  Burke JM 《Biochemistry》2001,40(12):3723-3729
The catalysis of site-specific RNA cleavage and ligation by the hairpin ribozyme requires the formation of a tertiary interaction between two independently folded internal loop domains, A and B. Within the B domain, a tertiary structure has been identified, known as the loop E motif, that has been observed in many naturally occurring RNAs. One characteristic of this motif is a partial cross-strand stack of a G residue on a U residue. In a few cases, including loop B of the hairpin ribozyme, this unusual arrangement gives rise to photoreactivity. In the hairpin, G21 and U42 can be UV cross-linked. Here we show that docking of the two domains correlates very strongly with a loss of UV reactivity of these bases. The rate of the loss of photoreactivity during folding is in close agreement with the kinetics of interdomain docking as determined by hydroxyl-radical footprinting and fluorescence resonance energy transfer (FRET). Fixing the structure of the complex in the cross-linked form results in an inability of the two domains to dock and catalyze the cleavage reaction, suggesting that the conformational change is essential for catalysis.  相似文献   

6.
The simulation of enzymatic reactions, using computer models, is becoming a powerful tool in the most fundamental challenge in biochemistry: to relate the catalytic activity of enzymes to their structure. In the present study, various computed parameters were correlated with the natural logarithm of experimental rate constants for the hydroxylation of various substrate derivatives catalysed by wild-type para-hydroxybenzoate hydroxylase (PHBH) as well as for the hydroxylation of the native substrate (p-hydroxybenzoate) by PHBH reconstituted with a series of 8-substituted flavins. The following relative parameters have been calculated and tested: (a) energy barriers from combined quantum mechanical/molecular mechanical (QM/MM) (AM1/CHARMM) reaction pathway calculations, (b) gas-phase reaction enthalpies (AM1) and (c) differences between the HOMO and LUMO energies of the isolated substrate and cofactor molecules (AM1 and B3LYP/6-31+G(d)). The gas-phase approaches yielded good correlations, as long as similarly charged species are involved. The QM/MM approach resulted in a good correlation, even including differently charged species. This indicates that the QM/MM model accounts quite well for the solvation effects of the active site surroundings, which vary for differently charged species. The correlations obtained demonstrate quantitative structure activity relationships for an enzyme-catalysed reaction including, for the first time, substitutions on both substrate and cofactor.  相似文献   

7.
We used FRET to examine the kinetics and thermodynamics of structural changes associated with ADP release in myosin V, which is thought to be a strain-sensitive step in many muscle and non-muscle myosins. We also explored essential dynamics using FIRST/FRODA starting with three different myosin V X-ray crystal structures to examine intrinsic flexibility and correlated motions. Our steady-state and time-resolved FRET analysis demonstrates a temperature-dependent reversible conformational change in the nucleotide-binding pocket (NBP). Our kinetic results demonstrate that the NBP goes from a closed to an open conformation prior to the release of ADP, while the actin-binding cleft remains closed. Interestingly, we find that the temperature dependence of the maximum actin-activated myosin V ATPase rate is similar to the pocket opening step, demonstrating that this is the rate-limiting structural transition in the ATPase cycle. Thermodynamic analysis demonstrates that the transition from the open to closed NBP conformation is unfavorable because of a decrease in entropy. The intrinsic flexibility analysis is consistent with conformational entropy playing a role in this transition as the MV.ADP structure is highly flexible compared to the MV.APO structure. Our experimental and modeling studies support the conclusion of a novel post-power-stroke actomyosin.ADP state in which the NBP and actin-binding cleft are closed. The novel state may be important for strain sensitivity as the transition from the closed to open NBP conformation may be altered by lever arm position.  相似文献   

8.
The oxygen transfer to p-hydroxybenzoate catalyzed by p-hydroxybenzoate hydroxylase (PHBH) has been shown to occur via a C4a-hydroperoxide of the flavin. Two factors are likely to be important in facilitating the transfer of oxygen from the C4a-hydroperoxide to the substrate. (a) The positive electrostatic potential of the active site partially stabilizes the negative charge centered on the oxygen of the flavin-C4a-alkoxide leaving group during the transition state [Ortiz-Maldonado, M., Ballou, D. P., and Massey, V. (1999) Biochemistry 38, 8124-8137]. (b) The hydrogen-bonding network ionizes the substrate to promote its nucleophilic attack on the electrophilic C4a-hydroperoxide intermediate [Entsch, B., Palfey, B. A., Ballou, D. P., and Massey, V. (1991) J. Biol. Chem. 266, 17341-17349]. This ionization is also aided by the positive electrostatic potential of the active site [Moran, G. R., Entsch, B., Palfey, B. A., and Ballou, D. P. (1997) Biochemistry 36, 7548-7556]. Substituents on the flavin can specifically affect the stability of the alkoxide leaving-group, whereas changes to specific enzyme residues can affect the charge in the active site and the hydrogen-bonding network. We have used wild-type (WT) PHBH and several mutant forms, all with normal FAD and with 8-Cl-FAD substituted for FAD, to assess the relative contributions of the two effects. Lys297Met and Asn300Asp have decreased positive charge in the active site, and these variants engender approximately 35-fold slower hydroxylation rates than the WT enzyme. Substitution of 8-Cl-FAD in these mutant forms gives approximately 1.8-fold increases in hydroxylation rates, compared with a > or =4.8-fold increase for WT with this flavin. The hydroxylation catalyzed by Tyr385Phe, a mutant enzyme form with a disrupted hydrogen-bonding network that compromises the ionization of the substrate without changing the positive charge of the active site, is stimulated 1.5-fold by substituting the enzyme with 8-Cl-FAD. The substrate, tetrafluoro-p-hydroxybenzoate, is fully ionized in WT PHBH, but this phenolate is a poor nucleophile because of the electron-withdrawing effects of the fluorine substituents. With tetrafluoro-p-hydroxybenzoate as the substrate, substitution of FAD with 8-Cl-FAD in the WT enzyme stabilizes the leaving alkoxide and leads to a 2.3-fold increase in the hydroxylation rate compared to that with FAD. Either the use of substrates that do not communicate with the proton network or the mutation of amino acid residues that perturb this interaction may prevent a necessary conformational change that allows proper orientation between reactants during the hydroxylation reaction or permits the essential protonation of the initially formed nascent flavin-C4a-peroxide anion. Thus, both activation of substrate by the proton network and stabilization of the leaving alkoxide appear to be important for oxygen transfer catalyzed by PHBH. The full effect of the substituents on the flavin (4.8-fold) can only be realized when the optimal transition state can be achieved, and this optimal state is not fully realized with the mutant forms.  相似文献   

9.
Cooperative binding of ATP and RNA to DEAD-box helicases induces the closed conformation of their helicase core, with extensive interactions across the domain interface. The bound RNA is bent, and its distortion may constitute the first step towards RNA unwinding. To dissect the role of the conformational change in the helicase core for RNA unwinding, we characterized the RNA-stimulated ATPase activity, RNA unwinding and the propensity to form the closed conformer for mutants of the DEAD box helicase YxiN. The ATPase-deficient K52Q mutant forms a closed conformer upon binding of ATP and RNA, but is deficient in RNA unwinding. A mutation in motif III slows down the catalytic cycle, but neither affects the propensity for the closed conformer nor its global conformation. Hence, the closure of the cleft in the helicase core is necessary but not sufficient for RNA unwinding. In contrast, the G303A mutation in motif V prevents a complete closure of the inter-domain cleft, affecting ATP binding and hydrolysis and is detrimental to unwinding. Possibly, the K52Q and motif III mutants still introduce a kink into the backbone of bound RNA, whereas G303A fails to kink the RNA substrate.  相似文献   

10.
We report Hammett correlations, using 8-substituted flavins, to clarify the mechanism of hydroxylation by p-hydroxybenzoate hydroxylase (PHBH). The 8-position of the FAD isoalloxazine ring was chosen for modifications, because in PHBH it has minimal interactions with the protein, and it is accessible to solvent and away from the site of hydroxylation. Although two intermediates, a flavin-C4a-hydroperoxide and a flavin-C4a-hydroxide, are known to participate in hydroxylation, the mechanism of oxygen transfer remains controversial. Mechanisms as diverse as electrophilic aromatic substitution, diradical formation, and isoalloxazine ring opening have been proposed. In the studies reported here, it was possible to monitor spectrally each of the individual steps involved in hydroxylation, because the FAD cofactor acts as a reporter group. Thus, with PHBH, substituted separately with nine derivatives of FAD altered in the 8-position, quantitative structure-reactivity relationships (QSAR) have been applied to probe the mechanisms of formation of the flavin-C4a-hydroperoxide, the conversion to the flavin-C4a-hydroxide with concomitant oxygen transfer to the substrate, and the dehydration of the flavin-C4a-hydroxide to form oxidized FAD. The individual chemical steps in the mechanism of PHBH were not altered when using any of the modified flavins, and normal products were obtained; however, the rates of individual steps were affected, and depended on the electronic properties of the 8-substituent. Increased hydroxylation rates were observed when a more electrophilic flavin-C4a-hydroperoxide (i.e., with an electron-withdrawing substituent at the 8-position) is bound to PHBH. On the basis of QSAR analysis, we conclude that the mechanism of the hydroxylation step is best described by electrophilic aromatic substitution.  相似文献   

11.
Conformational change in rat liver phenylalanine hydroxylase associated with activation by phenylalanine or N-(1-anilinonaphth-4-yl)maleimide was investigated by measuring fluorescence spectra and fluorescence lifetimes of tryptophanyl residues as well as the probe fluorophore conjugated with SH groups of the hydroxylase. The fluorescence spectrum of tryptophan exhibited its maximum at 342 nm. It shifted by 8 nm toward longer wavelength accompanied by an increase in its intensity, by preincubation with 1 mM phenylalanine. The fluorescence intensity of tryptophan increased by 36% upon the activation. On the other hand, the binding of (6R)-L-erythro-tetrahydrobiopterin, a natural cofactor of the enzyme, induced a decrease in the fluorescence intensity by 79% without a shift of the maximum wavelength. The fluorescence lifetime of tryptophan of phenylalanine hydroxylase exhibited two components with lifetimes of 1.7 and 4.1 ns. The values of the lifetimes changed to 1.4 and 5.6 ns, respectively, upon the activation. It is considered that the change in the longer lifetime is correlated with the shift of the emission peak upon the activation. The values of both the lifetimes decreased to 0.64 and 3.6 ns upon the binding of (6R)-L-erythro-tetrahydrobiopterin, which is coincident with the decrease in the fluorescence intensity. Conjugation of N-(1-anilinonaphth-4-yl)maleimide with SH of phenylalanine hydroxylase brought about a decrease in both the fluorescence intensity and the value of the shorter lifetime of the tryptophanyl residues, while the longer lifetime remained unchanged.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Cell-free translation of beta-tubulin mRNA generates full length beta-tubulin polypeptides distributed in three molecular forms: a high molecular weight lysate-associated form, the free beta-tubulin subunit, and the alpha beta-heterodimer (Yaffe, M.B., Farr, G. W., and Sternlicht, H. (1988) J. Biol. Chem. 263, 16023-16031). A quantitative assay system for these three forms was developed and used to measure the rates of incorporation/exchange of the newly synthesized free beta-subunit and the high molecular weight form into tubulin heterodimers following incubation of the 35S-translation products with unlabeled bovine tubulin dimer. This exchange process was found to be slow and strongly temperature-dependent. The half-lives for exchange ranged from 12.5 min at 37 degrees C to 17.5 h at 0 degree C with a measured activation energy of 22.5 kcal/mol. Microtubule-associated proteins appeared to play no role in the exchange process, since identical exchange rates were observed regardless of whether microtubule protein or phosphocellulose-purified tubulin was used as the source of tubulin dimer. Surprisingly, the exchange rates were found to be independent of dimer concentration. We interpret these results as evidence for a rate-limiting, slow conformational change that occurs within the newly synthesized beta-subunits prior to their association with alpha-tubulin to generate the alpha beta-hetero-dimer.  相似文献   

13.
para-Hydroxybenzoate hydroxylase is a flavoprotein monooxygenase that catalyzes a reaction in two parts: reduction of the enzyme cofactor, FAD, by NADPH in response to binding p-hydroxybenzoate to the enzyme, then oxidation of reduced FAD by oxygen to form a hydroperoxide, which oxygenates p-hydroxybenzoate to form 3,4-dihydroxybenzoate. These diverse reactions all occur within a single polypeptide and are achieved through conformational rearrangements of the isoalloxazine ring and protein residues within the protein structure. In this review, we examine the complex dynamic behavior of the protein that enables regulated fast and specific catalysis to occur. Original research papers (principally from the past 15 years) provide the information that is used to develop a comprehensive overview of the catalytic process. Much of this information has come from detailed analysis of many specific mutants of the enzyme using rapid reaction technology, biophysical measurements, and high-resolution structures obtained by X-ray crystallography. We describe how three conformations of the enzyme provide a foundation for the catalytic cycle. One conformation has a closed active site for the conduct of the oxygen reactions, which must occur in the absence of solvent. The second conformation has a partly open active site for exchange of substrate and product, and the third conformation has a closed protein structure with the isoalloxazine ring rotated out to the surface for reaction with NADPH, which binds in a surface cleft. A fundamental feature of the enzyme is a H-bond network that connects the phenolic group of the substrate in the buried active site to the surface of the protein. This network serves to protonate and deprotonate the substrate and product in the active site to promote catalysis and regulate the coordination of conformational states for efficient catalysis.  相似文献   

14.
15.
The kinetics of the phospholipase A2-catalyzed hydrolysis of bilayer vesicles and mixed micelles of several oxyglycero and thioglycero analogues of phospholipids have been studied. The results with vesicles show that, depending on the source of the enzyme, the rates of hydrolysis of the oxy-containing long-chain phosphatidylmethanols are 2.5- to 28-fold higher compared to the rates of hydrolysis of the analogous thio substrates. The oxygen to sulfur substitution does not significantly alter the affinities of the enzymes for the reaction products or calcium. Since it is unlikely that sulfur substitution changes the rate constants for the formation and dissociation of the enzyme-product complex by the same factor, the element effects seen in the rates of hydrolysis of the oxy- and thioester phospholipids in vesicles are primarily due to a change in the rate constant for the chemical step of the catalytic turnover cycle. For bovine pancreatic phospholipase A2, various mutants with lower catalytic activity were used to show that the value of the element effect does not increase in the mutants. These results establish that, for the pancreatic phospholipase A2, the element effect is fully expressed, and the chemical step is fully rate-limiting for both oxyglycero and thioglycero phospholipids in vesicles. It was found that the element effect decreases from 7 to 1 when long-chain phosphatidylmethanols are present in micelles of a neutral diluent. This result suggests that the chemical step is not rate-limiting during the hydrolysis of these mixed micelle substrates.  相似文献   

16.
Lactose transport in membrane vesicles containing lactose permease with a single Cys residue in place of Val 315 is inactivated by N-ethylmaleimide in a manner that is stimulated by substrate or by a H+ electrochemical gradient (delta microH+; Sahin-Tóth M, Kaback HR, 1993, Protein Sci 2:1024-1033). The findings are confirmed and extended in this communication. Purified, reconstituted Val 315-->Cys permease reacts with N-ethylmaleimide or hydrophobic fluorescent maleimides but not with a membrane impermeant thiol reagent, and beta-galactosides specifically stimulate the rate of labeling. Furthermore, the reactivity of purified Val 315-->Cys permease is enhanced by imposition of a membrane potential (delta psi, interior negative). The results indicate that either ligand binding or delta psi induces a conformational change in the permease that brings the N-terminus of helix X into an environment that is more accessible from the lipid phase.  相似文献   

17.
The energies of binding of seven ligands by p-hydroxybenzoate hydroxylase (PHBH) were calculated theoretically. Direct enzyme–ligand interaction energies were calculated using the ab initio quantum mechanical model assembly of the active site at the 3-21G level. Solvation energies of the ligands needed in the evaluation of the binding energies were calculated with the semiempirical AM1–SM2 method and the long-range electrostatic interaction energies between the ligands and the protein matrix classically using the static charge distributions of the ligands and the protein. Energies for proton-transfer between the ligands OH or SH substituent at position 4 and the active-site tyrosine within the ab initio model assemblies were calculated and compared to the corresponding pKas in aqueous solution. Excluding 3,4-dihydroxybenzoate, the natural product of PHBH, a linear relationship between the calculated binding energies and the experimental binding free energies was found with a correlation coefficient of 0.90. Contributions of the direct enzyme–ligand interaction energies, solvation energies and the long-range electrostatic interaction energies to the calculated binding energies were analyzed. The proton-transfer energies of the ligands with substituents ortho to the ionized OH were found to be perturbed less in the model calculations than the energies of their meta isomers as deduced from the corresponding pKas. © 1995 Wiley-Liss, Inc.  相似文献   

18.
J R Petithory  W P Jencks 《Biochemistry》1986,25(16):4493-4497
The calcium adenosinetriphosphatase of sarcoplasmic reticulum, preincubated with Ca2+ on the vesicle exterior (cE X Ca2), reacts with 0.3-0.5 mM Mg X ATP to form covalent phosphoenzyme (E approximately P X Ca2) with an observed rate constant of 220 s-1 (pH 7.0, 25 degrees C, 100 mM KCl, 5 mM MgSO4, 23 microM free external Ca2+, intact SR vesicles passively loaded with 20 mM Ca2+). If the phosphoryl-transfer step were rate-limiting, with kf = 220 s-1, the approach to equilibrium in the presence of ADP, to give 50% EP and kf = kr, would follow kobsd = kf + kr = 440 s-1. The reaction of cE X Ca2 with 0.8-1.2 mM ATP plus 0.25 mM ADP proceeds to 50% completion with kobsd = 270 s-1. This result shows that phosphoryl transfer from bound ATP to the enzyme is not the rate-limiting step for phosphoenzyme formation from cE X Ca2. The result is consistent with a rate-limiting conformational change of the cE X Ca2 X ATP intermediate followed by rapid (greater than or equal to 1000 s-1) phosphoryl transfer. Calcium dissociates from cE X Ca2 X ATP with kobsd = 80 s-1 and ATP dissociates with kobsd = 120 s-1 when cE X Ca2 X ATP is formed by the addition of ATP to cE X Ca2. However, when E X Ca2 X ATP is formed in the reverse direction, from the reaction of E approximately P X Ca2 and ADP, Ca2+ dissociates with kobsd = 45 s-1 and ATP dissociates with kobsd = 35 s-1. This shows that different E X Ca2 X ATP intermediates are generated in the forward and reverse directions, which are interconverted by a conformational change.  相似文献   

19.
Structural and kinetic studies have revealed two flavin conformations in p-hydroxybenzoate hydroxylase (PHBH), the in-position and the out-position. Conversion between these two conformations is believed to be essential during catalysis. Although substrate hydroxylation occurs while the flavin in PHBH is in the in-conformation, the position of the flavin during reduction by NADPH is uncertain. To investigate the catalytic importance of the out-conformation of the flavin and to clarify the mechanism of flavin reduction in PHBH, we report quantitative structure-reactivity relationships (QSAR) using PHBH substituted separately with nine derivatives of FAD modified in the 8-position and four dihydronicotinamide analogues as reducing agents. The 8-position of the FAD isoalloxazine ring was chosen for modification because in PHBH it has minimal interactions with the protein and is accessible to solvent. The chemical sequence of events during catalysis by PHBH was not altered when using any of the modified flavins, and normal products were obtained. Although the rate of reduction of PHBH reconstituted with flavin derivatives is expected to be dependent on the redox potential of the flavin, no strict correlation was observed. Instead, the rate of reduction correlated with the kappa-substituent constant, which is based on size and hydrophobicity of the 8-substituent on the FAD. Substituents that sterically hinder attainment of the out-conformation decreased the rate of flavin reduction much more than expected on the basis of the redox potential of the flavin. The results of this QSAR analysis are consistent with the hypothesis that the flavin in PHBH must move to the out-conformation for proper formation of the charge-transfer complex between NADPH and FAD that is necessary for rapid flavin reduction.  相似文献   

20.
A M Hanel  W P Jencks 《Biochemistry》1990,29(21):5210-5220
The calcium-transport ATPase (CaATPase) of rabbit sarcoplasmic reticulum preincubated with 0.02 mM Ca2+ (cE.Ca2) is phosphorylated upon the addition of 0.25 mM LaCl3 and 0.3 mM [gamma-32P]ATP with an observed rate constant of 6.5 s-1 (40 mM MOPS, pH 7.0, 100 mM KCl, 25 degrees C). La.ATP binds to cE.Ca2 with a rate constant of 5 X 10(6) M-1 s-1, while ATP, Ca2+, and La3+ dissociate from cE.Ca2.La.ATP at less than or equal to 1 s-1. The reaction of ADP with phosphoenzyme (EP) formed from La.ATP is biphasic. An initial rapid loss of EP is followed by a slower first-order disappearance, which proceeds to an equilibrium mixture of EP.ADP and nonphosphorylated enzyme with bound ATP. The fraction of EP that reacts in the burst (alpha) and the first-order rate constant for the slow phase (kb) increase proportionally with increasing concentrations of ADP to give maximum values of 0.34 and 65 s-1, respectively, at saturating ADP (KADPS = 0.22 mM). The burst represents rapid phosphoryl transfer and demonstrates that ATP synthesis and hydrolysis on the enzyme are fast. The phosphorylation of cE.Ca2 by La.ATP at 6.5 s-1 and the kinetics for the reaction of EP with ADP are consistent with a rate-limiting conformational change in both directions. The conformational change converts cE.Ca2.La.ATP to the form of the enzyme that is activated for phosphoryl transfer, aE.Ca2.La.ATP, at 6.5 s-1; this is much slower than the analogous conformational change at 220 s-1 with Mg2+ as the catalytic ion [Petithory & Jencks (1986) Biochemistry 25, 4493]. The rate constant for the conversion of aE.Ca2.La.ATP to cE.Ca2.La.ATP is 170 s-1. ATP does not dissociate measurably from aE.Ca2.La.ATP. Labeled EP formed from cE.Ca2 and La.ATP with leaky vesicles undergoes hydrolysis at 0.06 s-1. It is concluded that the reaction mechanism of the CaATPase is remarkably similar with Mg.ATP and La.ATP; however, the strong binding of La.ATP slows both the conformational change that is rate limiting for EP formation and the dissociation of La.ATP. An interaction between La3+ at the catalytic site and the calcium transport sites decreases the rate of calcium dissociation by greater than 60-fold. When cE-Ca2 is mixed with 0.3 mM ATP and 1.0 mM Cacl2, the phosphoenzyme is formed with an observed rate constant of 3 s-1. The phosphoenzyme formed from Ca.ATP reacts with 2.0 mM ADP and labeled ATP with a rate constant of 30 s-1; there may be a small burst (alpha less than or equal to 0.05).  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