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 31 毫秒
1.
The diuretic drug ethacrynic acid (EA), both an inhibitor and substrate of pi class glutathione S‐transferase (GST P1‐1), has been tested in clinical trials as an adjuvant in chemotherapy. We recently studied the role of the active site residue Tyr‐108 in binding EA to the enzyme and found that the analysis was complicated by covalent binding of this drug to the highly reactive Cys‐47. Previous attempts to eliminate this binding by chemical modification yielded ambiguous results and therefore we decided here to produce a double mutant C47S/Y108V by site directed mutagenesis and further expression in Escherichia coli and the interaction of EA and its GSH conjugate (EASG) examined by calorimetric studies and X‐ray diffraction. Surprisingly, in the absence of Cys‐47, Cys‐101 (located at the dimer interface) becomes a target for modification by EA, albeit at a lower conjugation rate than Cys‐47. The Cys‐47 → Ser mutation in the double mutant enzyme induces a positive cooperativity between the two subunits when ligands with affinity to G‐site bind to enzyme. However, this mutation does not seem to affect the thermodynamic properties of ligand binding to the electrophilic binding site (H‐site) and the thermal or chemical stability of this double mutant does not significantly affect the unfolding mechanism in either the absence or presence of ligand. Crystal structures of apo and an EASG complex are essentially identical with a few exceptions in the H‐site and in the water network at the dimer interface. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

2.
Cystathionine β‐lyase (CBL) catalyzes the hydrolysis of L ‐cystathionine (L ‐Cth) to produce L ‐homocysteine, pyruvate, and ammonia. A series of active‐site mutants of Escherichia coli CBL (eCBL) was constructed to investigate the roles of residues R58, R59, D116, W340, and R372 in catalysis and inhibition by aminoethoxyvinylglycine (AVG). The effects of these mutations on the kcat/K for the β‐elimination reaction range from a reduction of only 3‐fold for D116A and D116N to 6 orders of magnitude for the R372L and R372A mutants. The order of importance of these residues for the hydrolysis of L ‐Cth is: R372 >> R58 > W340 ≈ R59 > D116. Comparison of the kinetic parameters for L ‐Cth hydrolysis with those for inhibition of eCBL by AVG demonstrates that residue R58 tethers the distal carboxylate group of the substrate and confirms that residues W340 and R372 interact with the α‐carboxylate moiety. The increase in the pKa of the acidic limb and decrease in the pKa of the basic limb of the kcat/K versus pH profiles of the R58K and R58A mutants, respectively, support a role for this residue in modulating the pKa of an active‐site residue.  相似文献   

3.
Cystathionine γ‐synthase (CGS) catalyzes the condensation of O‐succinyl‐L ‐homoserine (L ‐OSHS) and L ‐cysteine (L ‐Cys), to produce L ‐cystathionine (L ‐Cth) and succinate, in the first step of the bacterial transsulfuration pathway. In the absence of L ‐Cys, the enzyme catalyzes the futile α,γ‐elimination of L ‐OSHS, yielding succinate, α‐ketobutyrate, and ammonia. A series of 16 site‐directed variants of Escherichia coli CGS (eCGS) was constructed to probe the roles of active‐site residues D45, Y46, R48, R49, Y101, R106, N227, E325, S326, and R361. The effects of these substitutions on the catalytic efficiency of the α,γ‐elimination reaction range from a reduction of only ~2‐fold for R49K and the E325A,Q variants to 310‐ and 760‐fold for R361K and R48K, respectively. A similar trend is observed for the kcat/K of the physiological, α,γ‐replacement reaction. The results of this study suggest that the arginine residues at positions 48, 106 and 361 of eCGS, conserved in bacterial CGS sequences, tether the distal and α‐carboxylate moieties, respectively, of the L ‐OSHS substrate. In contrast, with the exception of the 13‐fold increase observed for R106A, the K is not markedly affected by the site‐directed replacement of the residues investigated. The decrease in kcat observed for the S326A variant reflects the role of this residue in tethering the side chain of K198, the catalytic base. Although no structures exist of eCGS bound to active‐site ligands, the roles of individual residues is consistent with the structures inhibitor complexes of related enzymes. Substitution of D45, E325, or Y101 enables a minor transamination activity for the substrate L ‐Ala.  相似文献   

4.
Trimethylamine‐N‐oxide (TMAO) is a naturally occurring osmolyte that stabilizes proteins against denaturation. Although the impact of TMAO on the folding thermodynamics of many proteins has been well characterized, far fewer studies have investigated its effects on protein folding kinetics. In particular, no previous studies have used Φ‐value analysis to determine whether TMAO may alter the structure of the folding transition state. Here we have measured the effects on folding kinetics of 16 different amino acid substitutions distributed across the structure of the Fyn SH3 domain both in the presence and absence of TMAO. The folding and unfolding rates in TMAO, on average, improved to equivalent degrees, with a twofold increase in the protein folding rate accompanied by a twofold decrease in the unfolding rate. Importantly, TMAO caused little alteration to the Φ‐values of the mutants tested, implying that this compound minimally perturbs the folding transition state structure. Furthermore, the solvent accessibility of the transition state was not altered as reflected in an absence of a TMAO‐induced change in the denaturant β factors. Through TMAO‐induced folding studies, a β factor of 0.5 was calculated for this compound, suggesting that the protein backbone, which is the target of action of TMAO, is 50% exposed in the transition state as compared to the native state. This finding is consistent with the equivalent effects of TMAO on the folding and unfolding rates. Through thermodynamic analysis of mutants, we also discovered that the stabilizing effect of TMAO is lessened with increasing temperature.  相似文献   

5.
Equilibrium unfolding (folding) studies reveal that the autoregulatory RNA pseudoknots derived from the bacteriophage T2 and T4 gene 32 mRNAs exhibit significant stabilization by increasing concentrations of divalent metal ions in solution. In this report, the apparent affinities of exchange inert trivalent Co(NH3) have been determined, relative to divalent Mg2+, for the folded, partially folded (Kf), and fully unfolded (Ku) conformations of these molecules. A general nonspecific, delocalized ion binding model was developed and applied to the analysis of the metal ion concentration dependence of individual two‐state unfolding transitions. Trivalent Co(NH3) was found to associate with the fully folded and partially unfolded pseudoknotted forms of these RNAs with a Kf of 5–8 × 104 M−1 in a background of 0.10 M K+, or 3‐ to 5‐fold larger than the Kf obtained for two model RNA hairpins and hairpin unfolding intermediates, and ≈ 40–50‐fold larger than Kf for Mg2+. The magnitude of Kf was found to be strongly dependent on the monovalent salt concentration in a manner qualitatively consistent with polyelectrolyte theory, with Kf reaching 1.2 × 105 M−1 in 50 mM K+. Two RNA hairpins were found to have affinities for Co(NH3) and Ru(NH3) of 1–2 ×104 M−1, or ≈ 15‐fold larger than the Kf of ∼ 1000 M−1 observed for Mg2+. Additionally, the Ku of 4,800 M−1 for the trivalent ligands is ≈ 8‐fold larger than the Ku of 600 M−1 observed for Mg2+. These findings suggest that the T2 and T4 gene 32 mRNA pseudoknots possess a site(s) for Mg2+ and Co(NH3) binding of significantly higher affinity than a “duplexlike” delocalized ion binding site that is strongly linked to the thermodynamic stability of these molecules. Imino proton perturbation nmr spectroscopy suggests that this site(s) lies near the base of the pseudoknot stem S2, near a patch of high negative electrostatic potential associated with the region where the single loop L1 adenosine crosses the major groove of stem S2. © 1999 John Wiley & Sons, Inc. Biopoly 50: 443–458, 1999  相似文献   

6.
Excited‐state intramolecular proton transfer (ESIPT) and dual luminescence behaviour of 3‐hydroxyflavone (3‐HF) have been utilized to monitor its binding to liposomal membranes prepared from egg yolk phosphatydilcholine (EYPC). Additionally, absorption spectrophotometric assay has been performed to evaluate the antioxidant activity of 3‐HF against lipid peroxidation in this membrane system. When 3‐HF molecules are partitioned into EYPC liposomes, a weak long‐wavelength absorption band with λ ~410 nm appears in addition to the principal absorption at ~λ = 345 nm. Selective excitation of the 410 nm band produces the characteristic emission (λ~460 nm) of the ground‐state anionic species, whereas excitation at the higher energy absorption band leads to dual emission with predominatly ESIPT tautomer fluorescence (λ = 528 nm). Both ESIPT tautomer and the anionic species exhibit fairly high fluorescence anisotropy (r) values (r = 0.122 and 0.180, respectively). Biexponential fluorescence decay kinetics are observed for the ESIPT tautomer as well as the ground‐state anionic forms, indicating heterogeneity in the microenvironments of the corresponding emitting species. Furthermore, we demonstrate that lipid peroxidation of EYPC liposomes is significantly inhibited upon 3‐HF binding, suggesting that 3‐HF can be potentially useful as an inhibitor of peroxidative damage of cell membranes. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

7.
The antioxidant behavior of a series of new synthesized substituted thiazolyl‐thiazolidine‐2,4‐dione compounds (TZDs) was examined using chemiluminescence and electron paramagnetic resonance spin trapping techniques. 5,5‐Dimethyl‐1‐pyrroline‐N‐oxide (DMPO) was used as the spin trap. The reactivity of TZDs with superoxide anion radical (O) and hydroxyl radical (HO?) was evaluated using potassium superoxide/18‐crown‐6 ether dissolved in dimethylsulfoxide, and the Fenton‐like reaction (Fe2+ + H2O2), respectively. The results showed that TZDs efficiently inhibited light emission from the O generating system at a concentration of 0.05–1 mmol L?1 (5–94% reductions were found at 1 mmol L?1 concentration). The TZD compounds showed inhibition of HO?‐dependent DMPO–OH spin adduct formation from DMPO (the amplitude decrease ranged from 8 to 82% at 1 mmol L?1 concentration). The findings showed that examined TZDs had effective activities as radical scavengers. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
The scavenging effects of eighteen thiazolyl thiazolidine‐2,4‐dione compounds (TTCs) on superoxide radical , hydroxyl radical HO?, and 1,1‐diphenyl‐2‐picrylhydrazyl (DPPH?) radical were evaluated by the chemiluminescence technique, electron spin resonance spectrometry (ESR) and visible spectrophotometry, respectively. The examined compounds were shown to have 27–59% scavenging ability, 19–69% HO? scavenging activity and 2–32% DPPH? scavenging ability. This property of the tested compound seems to be important in the prevention of various diseases of free radicals etiology. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

9.
Effects of hydrophobic and electrostatic interactions on the self‐assembling process of the ionic‐complementary peptide EMK16‐II are investigated by atomic force microscopy imaging, circular dichroism spectra, light scattering, and chromatography. It is found that the hydrophobicity of the peptide promotes the aggregation in pure water even at a very low concentration, resulting in a much lower critical aggregation concentration than that of another peptide, EAK16‐II. The effect of anions in solution with different valences on electrostatic interactions is also important. Monovalent anions (Cl? and Ac?) with a proper concentration can facilitate the formation of peptide fibrils, with Cl? of smaller size being more effective than Ac? of larger size. However, only small amounts of fibrils, but plenty of large amorphous aggregates, are found when the peptide solution is incubated with multivalent anions, such as SO, C6H5O, and HPO. More importantly, by gel filtration chromatography, the citrate anion, which induces a similar effect on the self‐assembling process of EMK16‐II as that of SO and HPO, can interact with two or more positively charged residues of the peptide and reside in the amorphous aggregates. This implies a “salt bridge” effect of multivalent anions on the peptide self‐assembling process, which can interpret a previous puzzle why divalent cations inhibit the formation of ordered nanofibrils of the ionic‐complementary peptides. Thus, our results clarify the important effects of hydrophobic and electrostatic interactions on the self‐assembling process of the ionic‐complementary peptides. These are greatly helpful for us to understand the mechanism of peptides' self‐assembling process and protein folding and aggregation. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 318–329, 2010. This article was originally published online as an acceptedpreprint. The “Published Online” date corresponds to the preprintversion. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

10.
M2 transmembrane domain channel (M2‐TMD) permeation properties are studied using molecular dynamics simulations of M2‐TMD (1NYJ) embedded in a lipid bilayer (DMPC) with 1 mol/kg NaCl or KCl saline solution. This study allows examination of spontaneous cation and anion entry into the selectivity filter. Three titration states of the M2‐TMD tetramer are modeled for which the four His37 residues, forming the selectivity filter, are net uncharged, +2 charged, or +3 charged. M2‐TMD structural properties from our simulations are compared with the properties of other models extracted from NMR and X‐ray studies. During 10 ns simulations, chloride ions occasionally occupy the positively‐charged selectivity filter region, and from umbrella sampling simulations, Cl? has a lower free‐energy barrier in the selectivity‐filter region than either Na+ or NH, and NH has a lower free‐energy barrier than Na+. For Na+ and Cl?, the free‐energy barriers are less than 5 kcal/mol, suggesting that the 1NYJ conformation would probably not be exquisitely proton selective. We also point out a rotameric configuration of Trp41 that could fully occlude the channel. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

11.
The oxygen free radical scavenging activities of 15 chromonyl‐thiazolidine‐2,4‐dione compounds (CTDs) were examined in chemical systems producing superoxide anion radicals, O (potasium superoxide–18‐crown‐6 ether–DMSO), and hydroxyl radicals, HO? (a Fenton reaction: Fe(II)–H2O2–sodium trifluoroacetate, pH 6.15). Chemiluminescence and electron spin resonance (ESR) spectroscopy using 5,5‐dimethyl‐1‐pyrroline‐1‐oxide (DMPO) as spin trap were applied to evaluate antioxidant behaviour of CTDs towards the oxygen radicals. The results indicated that 11 of the 15 tested compounds showed a significant inhibitory effect on the chemiluminescence generated from the O‐generating system, ranging from 41 to 86%, and 13 CTDs quenched the ESR signal of the DMPO–OH spin adduct by 33–86%, at a concentration of 1 mmol L?1. Our findings demonstrate that CTDs could be good free radical scavengers. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

12.
13.
Direct enantiomer separation of hypericin, pseudohypericin, and protohypericin was accomplished by high‐performance liquid chromatography (HPLC) using immobilized polysaccharide‐type chiral stationary phases (CSPs). Enantioselectivities up to 1.30 were obtained in the polar‐organic elution mode whereby for hypericin and pseudohypericin Chiralpak IC [chiral selector being cellulose tris(3,5‐dichlorophenylcarbamate)] and for protohypericin Chiralpak IA (chiral selector being the 3,5‐dimethylphenylcarbamate of amylose) gave favorable results. Enantiomers were distinguished by on‐line electronic circular dichroism detection. Optimized enantioselective chromatographic conditions were the basis for determining stereodynamic parameters of the enantiomer interconversion process of hypericin and pseudohypericin. Rate constants delivered by computational simulation of dynamic HPLC elution profiles (stochastic model, consideration of peak tailing) were used to calculate averaged enantiomerization barriers (ΔG) of 97.6–99.6 kJ/mol for both compounds (investigated temperature range 25–45°C). Complementary variable temperature off‐column (i.e., in solution) racemization experiments delivered ΔG = 97.1–98.0 kJ/mol (27–45°C) for hypericin and ΔG = 98.9–101.4 kJ/mol (25–55°C) for pseudohypericin. An activation enthalpy of ΔH# = 86.0 kJ/mol and an activation entropy of ΔS# = ?37.7 J/(K mol) were calculated from hypericin racemization kinetics in solution, whereas for pseudohypericin these figures amounted to 74.1 kJ/mol and ?82.6 J/(K mol), respectively. Although the natural phenanthroperylene quinone pigments hypericin and pseudohypericin as well as their biological precursor protohypericin are chiral and can be separated by enantioselective HPLC low enantiomerization barriers seem to prevent the occurrence of an excess of one enantiomer under typical physiological conditions—at least as long as stereoselective intermolecular interactions with other chiral entities are absent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

14.
Living microbial cells are considered to be the catalyst of choice for selective terpene functionalization. However, such processes often suffer from side product formation and poor substrate mass transfer into cells. For the hydroxylation of (S)‐limonene to (S)‐perillyl alcohol by Pseudomonas putida KT2440 (pGEc47ΔB)(pCom8‐PFR1500), containing the cytochrome P450 monooxygenase CYP153A6, the side products perillyl aldehyde and perillic acid constituted up to 26% of the total amount of oxidized terpenes. In this study, it is shown that the reaction rate is substrate‐limited in the two‐liquid phase system used and that host intrinsic dehydrogenases and not CYP153A6 are responsible for the formation of the undesired side products. In contrast to P. putida KT2440, E. coli W3110 was found to catalyze perillyl aldehyde reduction to the alcohol and no oxidation to the acid. Furthermore, E. coli W3110 harboring CYP153A6 showed high limonene hydroxylation activities (7.1 U g). The outer membrane protein AlkL was found to enhance hydroxylation activities of E. coli twofold in aqueous single‐phase and fivefold in two‐liquid phase biotransformations. In the latter system, E. coli harboring CYP153A6 and AlkL produced up to 39.2 mmol (S)‐perillyl alcohol L within 26 h, whereas no perillic acid and minor amounts of perillyl aldehyde (8% of the total products) were formed. In conclusion, undesired perillyl alcohol oxidation was reduced by choosing E. coli's enzymatic background as a reaction environment and co‐expression of the alkL gene in E. coli represents a promising strategy to enhance terpene bioconversion rates. Biotechnol. Bioeng. 2013; 110: 1282–1292. © 2012 Wiley Periodicals, Inc.  相似文献   

15.
16.
Microfluorometric recordings showed that the inhibitory neurotransmitters γ‐aminobutyric acid (GABA) and glycine activated transient increases in the intracellular Cl concentration in neurons of the inferior colliculus (IC) from acutely isolated slices of the rat auditory midbrain. Current recordings in gramicidin‐perforated patch mode disclosed that GABA and glycine mainly evoked inward or biphasic currents. These currents were dependent on HCO and characterized by a continuous shift of their reversal potential (EGABA/gly) in the positive direction. In HCO‐buffered saline, GABA and glycine could also evoke an increase in the intracellular Ca2+ concentration. Ca2+ transients occurred only with large depolarizations and were blocked by Cd2+, suggesting an activation of voltage‐gated Ca2+ channels. However, in the absence of HCO, only a small rise, if any, in the intracellular Ca2+ concentration could be evoked by GABA or glycine. We suggest that the activation of GABAA or glycine receptors results in an acute accumulation of Cl that is enhanced by the depolarization owing to HCO efflux, thus shifting EGABA/gly to more positive values. A subsequent activation of these receptors would result in a strenghtened depolarization and an enlarged Ca2+ influx that might play a role in the stabilization of inhibitory synapses in the auditory pathway. © 1999 John Wiley & Sons, Inc. J Neurobiol 40: 386–396, 1999  相似文献   

17.
Purple non‐sulfur (PNS) bacteria produce hydrogen by photofermentation of organic acids in wastewater. However, NH in wastewater may inhibit hydrogen synthesis by repressing the expression and activity of nitrogenase, the enzyme catalyzing hydrogen production in PNS bacteria. In this study, the Rhodobacter sphaeroides 6016 glnA gene encoding glutamine synthetase (GS) was knocked out by homologous recombination, and the effects on hydrogen production and nitrogenase activity were examined. Using 3 mM glutamine as the nitrogen source, hydrogen production (1,245–1,588 mL hydrogen/L culture) and nitrogenase activity were detected in the mutant in the presence of relatively high NH concentrations (15–40 mM), whereas neither was detected in the wild‐type strain under the same conditions. Further analysis indicated that high NH concentrations greatly inhibited the expression of nifA and nitrogenase gene in the wild‐type strain but not in the glnA1? mutant. These observations suggest that GS is essential to NH repression of nitrogenase and that deletion of glnA1 results in the complete derepression of nitrogenase by preventing NH assimilation in vivo, thus relieving the inhibition of nifA and nitrogenase gene expression. Knocking out glnA1 therefore provides an efficient approach to removing the inhibitory effects of ammonium ions in R. sphaeroides and possibly in other hydrogen‐producing PNS bacteria. Biotechnol. Bioeng. 2010;106: 564–572. © 2010 Wiley Periodicals, Inc.  相似文献   

18.
The kinetics of ethidium's intercalative binding to DNA packaged in bacteriophage T7 and two T7 deletion mutants have been determined, using enhancement of fluorescence to quantitate binding. At a constant ethidium concentration, the results can be described as first-order binding with two different rate constants, k (= k1 + k?1) and k (= k2 + k?2). The larger rate constant (k) was at least four orders of magnitude smaller than the comparable first-order forward rate constant for binding to DNA released from its capsid. At 25°C values of k decreased as the amount of DNA packaged per internal volume increased. This latter observation indicates that the rate of ethidium's binding to packaged T7 DNA is limited by an event that occurs inside of the DNA-containing region of T7, not by the crossing of T7 capsid's outer shell. Arrhenius plots of kM are biphasic, indicating a transition for packaged DNA at a temperature of 20°C. The data indicate that k s are limited by either sieving of ethidium during its passage through the packaged DNA or subsequent hindered intercalation.  相似文献   

19.
Shokhen M  Khazanov N  Albeck A 《Proteins》2011,79(3):975-985
Various mechanisms for the reversible formation of a covalent tetrahedral complex (TC) between papain and peptidyl aldehyde inhibitors were simulated by DFT calculations, applying the quantum mechanical/self consistent reaction field (virtual solvent) [QM/SCRF(VS)] approach. Only one mechanism correlates with the experimental kinetic data. The His–Cys catalytic diad is in an N/SH protonation state in the noncovalent papain–aldehyde Michaelis complex. His159 functions as a general base catalyst, abstracting a proton from the Cys25, whereas the activated thiolate synchronously attacks the inhibitor's carbonyl group. The final product of papain inhibition is the protonated neutral form of the hemithioacetal TC(OH), in agreement with experimental data. The predicted activation barrier g = 5.2 kcal mol?1 is close to the experimental value of 6.9 kcal mol?1. An interpretation of the experimentally observed slow binding effect for peptidyl aldehyde inhibitors is presented. The calculated g is much lower than the rate determining activation barrier of hemithioacetal formation in water, g, in agreement with the concept that the preorganized electrostatic environment in the enzyme active site is the driving force of enzyme catalysis. We have rationalized the origin of the acidic and basic pKa's on the k2/KS versus pH bell‐shaped profile of papain inhibition by peptidyl aldehydes. Proteins 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

20.
EPR and water proton relaxation rate (1/T1) studies of partially (40%) and “fully” (90%) purified preparations of membrane-bound (Na++K+) activated ATPase from sheep kidney indicate one tight binding site for Mn2+ per enzyme dimer, with a dissociation constant (KD = 0.88 μM) in agreement with the kinetically determined activator constant, identifying this Mn2+-binding site as the active site of the ATPase. Competition studies indicate that Mg2+ binds at this site with a dissociation constant of 1 mM in agreement with its activator constant. Inorganic phosphate and methylphosphonate bind to the enzyme-Mn2+ complex with similar high affinities and decrease l/T1 of water protons due t o a decrease from four to three in the number of rapidly exchanging water protons in the coordination sphere of enzyme-bound Mn2+. The relative effectiveness of Na+ and K+ in facilitating ternary complex formation with HPO and CH3PO as a function of pH indicates that Na+ induces the phosphate monoanion t o interact with enzyme-bound Mn2+, while K+ causes the phosphate dianion to interact with the enzyme-bound Mn2+. Thus protonation of an enzyme-bound phosphoryl group would convert a K+-binding site to a Na+-binding site. Dissociation constants for K+ and Na+, estimated from NMR titrations, agreed with kinetically determined activator constants of these ions consistent with binding t o the active site. Parallel 32Pi-binding studies show negligible formation (< 7%) of a covalent E–P complex under these conditions, indicating that the NMR method has detected an additional noncovalent intermediate in ion transport. Ouabain, which increases the extent of phosphorylation of the enzyme to 24% at pH 7.5 and t o 106% at pH 6.1, produced further decreases in l/T 1 of water protons. Preliminary 31P-relaxation studies of CH3PO in the presence of ATPase and Mn2+ yield an Mn to P distance (6.9 ± 0.5 Å) suggesting a second sphere enzyme-Mn-ligand-CH3PO complex. Previous kinetic studies have shown that T1+ substitutes for K+ in the activation of the enzyme but competes with Na+ at higher levels. From the paramagnetic effect of Mn2+ at the active site on the enzyme on I/T1 of 205T1 bound at the Na+ site, a Mn2+ to T1+ distance of 4.0 ± 0.1 Å is calculated, suggesting the sharing of a common ligand atom by Mn2+ and T1+ on the ATPase. Addition of P. increases this distance to 5.4 Å consistent with the insertion of P between Mn2+ and T1+. These results are consistent with a mechanism for the \documentclass{article}\pagestyle{empty}\begin{document}$ (\mathop {\rm N}\limits^{\rm i} {\rm a}^{\rm + } {\rm + K}^ +) $\end{document}-ATPase and for ion transport in which the ionization state of Pi at a single enzyme active site controls the binding and transport of Na+ and K+, and indicate that the transport site for monovalent cations is very near the catalytic site of the ATTase. Our mechanism also accounts for the order of magnitude weaker binding of Na+ compared to K+.  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