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 343 毫秒
1.
Leys D  Basran J  Scrutton NS 《The EMBO journal》2003,22(16):4038-4048
Here we report crystal structures of dimethylglycine oxidase (DMGO) from the bacterium Arthrobacter globiformis, a bifunctional enzyme that catalyzes the oxidation of N,N-dimethyl glycine and the formation of 5,10-methylene tetrahydrofolate. The N-terminal region binds FAD covalently and oxidizes dimethylglycine to a labile iminium intermediate. The C-terminal region binds tetrahydrofolate, comprises three domains arranged in a ring-like structure and is related to the T-protein of the glycine cleavage system. The complex with folinic acid indicates that this enzyme selectively activates the N10 amino group for initial attack on the substrate. Dead-end reactions with oxidized folate are avoided by the strict stereochemical constraints imposed by the folate-binding funnel. The active sites in DMGO are approximately 40 A apart, connected by a large irregular internal cavity. The tetrahydrofolate-binding funnel serves as a transient entry-exit port, and access to the internal cavity is controlled kinetically by tetrahydrofolate binding. The internal cavity enables sequestration of the reactive iminium intermediate prior to reaction with tetrahydrofolate and avoids formation of toxic formaldehyde. This mode of channelling in DMGO is distinct from other channelling mechanisms.  相似文献   

2.
It was observed previously that hydroxyguanidine is formed in the reaction of canavanine(2-amino-4-guanidinooxybutanoate) with amino acid oxidases. The present work shows that hydroxyguanidine is formed by a nonenzymatic beta,gamma-elimination reaction following enzymatic oxidation at the alpha-C and that the abstraction of the beta-H is general-base catalyzed. The elimination reaction requires the presence in the alpha-position of an anion-stabilizing group--the protonated imino group (iminium ion group) or the carbonyl group. The iminium ion group is more activating than the carbonyl group. Elimination is further facilitated by protonation of the guanidinooxy group. The other product formed in the elimination reaction was identified as vinylglyoxylate (2-oxo-3-butenoate), a very highly electrophilic substance. The product resulting from hydrolysis following oxidation was identified as alpha-keto-gamma-guanidinooxybutyrate (ketocanavanine). The ratio of hydroxyguanidine to ketocanavanine depended upon the concentration and degree of basicity of the basic catalyst and on pH. In the presence of semicarbazide, the elimination reaction was prevented because the imino group in the semicarbazone derivative of ketocanavanine is not significantly protonated. Incubation of canavanine with 5'-deoxypyridoxal also yielded hydroxyguanidine. Since the elimination reactions take place under mild conditions, they may occur in vivo following oxidation at the alpha-C of L-canavanine (ingested or formed endogenously) or of other amino acids with a good leaving group in the gamma-position (e.g., S-adenosylmethionine, methionine sulfoximine, homocyst(e)ine, or cysteine-homocysteine mixed disulfide) by an L-amino acid oxidase, a transaminase, or a dehydrogenase. Therefore, vinylglyoxylate may be a normal metabolite in mammals which at elevated concentrations may contribute to the in vivo toxicity of canavanine and of some of the other above-mentioned amino acids.  相似文献   

3.
Various dihydroxyacetone-phosphate (DHAP) analogues bearing an aromatic ring or β-dicarbonyl structures were synthesized. Their capacity to form a stabilized iminium ion or conjugated enamine in the reaction catalyzed by rabbit muscle aldolase (EC 4.1.2.13) were investigated by enzymatic kinetics and UV difference spectroscopic techniques. Whereas the aromatic derivative led to competitive inhibition without detectable iminium ion formation, slow reversible inhibitions of aldolase by β-dicarbonyl compounds was shown to have taken place. Conjugated enamine formation at the active site of the enzyme was detected by their specific absorbances close to 317 nm.  相似文献   

4.
Alanine-dependent reactions of 5'-deoxypyridoxal in water   总被引:1,自引:0,他引:1  
The non-enzymatic reaction of 5′-deoxypyridoxal (DPL) with l-alanine in water at 25 °C was investigated. DPL reacts with alanine to form an imine, which then undergoes deprotonation at the α-amino carbon of alanine to form a resonance delocalized DPL-stabilized carbanion. At early reaction times the only detectable products are pyruvate and the dimeric species formed by addition of the α-pyridine stabilized carbanion to DPL. No Claisen-type products of addition of the α-amino carbanion to DPL, as was previously reported to form from the reaction between DPL and glycine [K. Toth, T.L. Amyes, J.P. Richard, J.P.G. Malthouse, M.E. Nı´Beilliu, J. Am. Chem. Soc. 126 (2004) 10538–10539], are observed. The electrophile reacts instead at the α-pyridyl carbon. This dimer is in chemical equilibrium with reactants. At longer reaction times about 50% of DPL is converted to 5′-deoxypyridoxamine, the thermodynamically favored product of formal transamination of DPL.  相似文献   

5.
Rate constants for the hydrolysis of L-proline methyl ester to form proline and methanol in D(2)O buffered at neutral pD and 25 degrees C and the deuterium enrichment of the proline product determined by electrospray ionization mass spectrometry are reported. The data give k(DO) = 5.3 +/- 0.5 M(-1) s(-1) as the second-order rate constant for carbon deprotonation of N-protonated proline methyl ester by deuterioxide ion in D(2)O at 25 degrees C and I = 1.0 (KCl). The data provide good estimates of carbon acidities of pK(a) = 21 for N-protonated proline methyl ester and pK(a) = 29 for proline zwitterion in water and of the second-order rate constant k(HO) = 4.5 x 10(-5) M(-1) s(-1) for carbon deprotonation of proline zwitterion by hydroxide ion at 25 degrees C. There is no detectable acceleration of the deprotonation of N-protonated proline methyl ester by the Br?nsted base 3-quinuclidinone in water, and it is not clear that such Br?nsted catalysis would make a significant contribution to the rate acceleration for deprotonation of bound proline at proline racemase. A comparison of the first-order rate constants k(HO)[HO(-)] = 4.5 x 10(-11) s(-1) for deprotonation of free proline zwitterion in water at pH 8 and k(cat) = 2600 s(-1) for deprotonation of proline bound to the active site of proline racemase at pH 8 shows that the enzymatic rate acceleration for proline racemase is ca. 10(13)-fold. This corresponds to a 19 kcal/mol stabilization of the transition state for deprotonation of the enzyme-bound carbon acid substrate by interaction with the protein catalyst. It is suggested that (1) much of the rate acceleration of the enzymatic over the nonenzymatic reaction in water may result from transfer of the substrate proline zwitterion from the polar solvent water to a nonpolar enzyme active site and (2) the use of thiol anions rather than oxygen anions as Br?nsted bases at this putative nonpolar enzyme active site may be favored, because of the smaller energetic price for desolvation of thiol anions than for desolvation of the more strongly solvated oxygen anions.  相似文献   

6.
The pH dependence of the maximum velocity of the reaction catalyzed by diphosphopyridine nucleotide (DPN) dependent isocitrate dehydrogenase indicates the requirement for the basic form of an ionizable group in the enzyme-substrate complex with a pK of 6.6. This pK is unaltered from 10 to 33 degrees C, suggesting the ionization of a carboxyl rather than an imidazolium ion. The enzyme is inactivated upon incubation with 1-cyclohexyl-3-(2-morpholinoethyl)carbodiimide in the presence of glycinamide or glycine ethyl ester. This inactivation is dependent on pH and the rate constant (k) increases as the pH is decreased in the range 7.3 to 6.25. A plot of 1/(H+) vs. 1/k suggests that the enzyme is inactivated as a result of the modification of a single ionizable group in this pH range. The coenzyme DPN and substrate alpha-ketoglutarate do not affect the rate of inactivation. In contrast, manganous ion (2 mM) and isocitrate (60 mM) produce a sevenfold decrease in the rate constant. The allosteric activator ADP (1 mM) does not itself influence the rate of inactivation; however, it reduces the concentration of Mn2+ (1 mM) and isocitrate (20 mM) required to produce the same decrease in the inactivation constant. These observations imply that the modification occurs at the substrate-binding site. Experiments employing [1-14C]glycine ethyl ester show a net incorporation of 2 mol of glycine ethyl ester per subunit (40 000), concomitant with the complete inactivation of the enzyme. The radioactive modified enzyme, after removal of excess reagent by dialysis, was exhaustively digested with proteolytic enzymes. High voltage electrophoretic analyses of the hydrolysate at pH 6.4 and 3.5 yield two major radioactive spots with approximately equal intensity, which correspond to gamma-glutamylglycine and beta-aspartylglycine, the ultimate products of reaction with glutamic and aspartic acids, respectively. Modification in the presence of manganous ion and isocitrate results in significant reduction in the incorporation of radioactivity into the two dipeptides. These results suggest that carbodiimide attacks one glutamyl and one aspartyl residue per subunit of the enzyme and that the integrity of these residues is crucial for the enzymatic activity.  相似文献   

7.
5-Aminolevulinate synthase catalyzes the pyridoxal 5'-phosphate-dependent condensation of glycine and succinyl-CoA to produce carbon dioxide, CoA, and 5-aminolevulinate, in a reaction cycle involving the mechanistically unusual successive cleavage of two amino acid substrate alpha-carbon bonds. Single and multiple turnover rapid scanning stopped-flow experiments have been conducted from pH 6.8-9.2 and 5-35 degrees C, and the results, interpreted within the framework of the recently solved crystal structures, allow refined characterization of the central kinetic and chemical steps of the reaction cycle. Quinonoid intermediate formation occurs with an apparent pK(a) of 7.7 +/- 0.1, which is assigned to His-207 acid-catalyzed decarboxylation of the alpha-amino-beta-ketoadipate intermediate to form an enol that is in rapid equilibrium with the 5-aminolevulinate-bound quinonoid species. Quinonoid intermediate decay occurs in two kinetic steps, the first of which is acid-catalyzed with a pK(a) of 8.1 +/- 0.1, and is assigned to protonation of the enol by Lys-313 to generate the product-bound external aldimine. The second step of quinonoid decay defines k(cat) and is relatively pH-independent and is assigned to opening of the active site loop to allow ALA dissociation. The data support important refinements to both the chemical and kinetic mechanisms and indicate that 5-aminolevulinate synthase operates under the stereoelectronic control predicted by Dunathan's hypothesis.  相似文献   

8.
Thiol oxidation by hypochlorous acid and chloramines is a favorable reaction and may be responsible for alterations in regulatory or signaling pathways in cells exposed to neutrophil oxidants. In order to establish the mechanism for such changes, it is necessary to appreciate whether these oxidants are selective for different thiols as compared with other scavengers. We have measured rate constants for reactions of amino acid chloramines with a range of thiols, methionine, and ascorbate, using a combination of stopped-flow and competitive kinetics. For HOCl, rate constants are too fast to measure directly by our system and values relative to reduced glutathione were determined by competition with methionine. For taurine chloramine, the rate constants for reaction with 5-thio-2-nitrobenzoic acid, GSH, methionine, and ascorbate at pH 7.4 were 970, 115, 39, and 13 M(-1) s(-1), respectively. Values for 10 thiols varied by a factor of 20 and showed an inverse relationship to the pK(a) of the thiol group. Rate constants for chloramines of glycine and N-alpha-acetyl-lysine also showed these relationships. Rates increased with decreasing pH, suggesting a mechanism involving acid catalysis. For hypochlorous acid, rates of reaction with 5-thio-2-nitrobenzoic acid, GSH, cysteine, and most of the other thiols were very similar. Relative reactivities varied by less than 5 and there was no dependence on thiol pK(a). Chloramines have the potential to be selective for different cellular thiols depending on their pK(a). For HOCl to be selective, other factors must be important, or its reactions could be secondary to chloramine formation.  相似文献   

9.
The 13C chemical shifts of several 85% 13C-enriched amino acids and small peptides were studied as a function of pH. The results show that the chemical shifts of carbon atoms of ionizable groups vary significantly within the zone of their pK. Generally with the pH GOING FROM 7 to 1 all the deltaC are shifted more or less upfield with the exception of the carbonyl group carbon of the second last residue which is shifted slightly downfield. This suggests the formation of an hydrogen bond at acid pH involving in a seven-membered ring the C=O in question and the COOH terminal. The percentage of cis and trans conformers of glycyl-L-proline and glycyl-L-prolylglycine were studied as a function of pH. The trans form is always preponderant whatever the pH. The accessibility of the carbonyl group to protonation of the proline residue strongly influences the cis-trans equilibrium. Thus, with the pH varying from 7 to 1, the trans isomer changes from 61 to 85% for glycyl-L-proline and only from 77 to 80% for glycyl-L-prolylglycine. The proton NMR studies underline the important differences existing between the two molecular forms of glycyl-L-proline. The cis conformation is characterized with regard to the trans form by the non-equivalence of the alpha-protons of the glycine residue, by a lower pK(1) and by a larger deltadeltaHalpha of the proline residue as a function of pH. These results could suggest an end-to-end interaction in the cis form of the glycyl-L-proline molecule. The 13C-13C coupling constants were also studied as a function of pH. The results show that J(Co-Calpha) of a C-terminal residue, varying from 5 to 6 Hz and reflecting thhe pK of the carboxylate group, is a linear function of delta(Co) and delta(Calpha) as in the case of the amino acids. The total variation of the electron density of those two carbons in an amino acid is approximately 40% weaker than in a C-terminal residue. The charge distribution along the Calpha-C(o) bond, however, is practically the same in both cases. Finally the ratios of the conversion rate constants of the two isomers cis-trans of glycyl-proline were calculated at different pH values; the relations between the isomer percentages and delta(Co), delta(Calpha) on the one hand and the J(Co-Calpha) on the other were established.  相似文献   

10.
The prephenate dehydrogenase activity of the bifunctional enzyme chorismate mutase-prephenate dehydrogenase from Escherichia coli catalyzes the oxidative decarboxylation of both prephenate and deoxoprephenate, which lacks the keto group in the side chain (V 78% and V/K 18% those of prephenate). Hydride transfer is to the B side of NAD, and the acetylpyridine and pyridinecarboxaldehyde analogues of NAD have V/K values 40 and 9% and V values 107 and 13% those of NAD. Since the 13C isotope effect on the decarboxylation is 1.0103 with deuterated and 1.0033 with unlabeled deoxoprephenate (the deuterium isotope effect on V/K is 2.34), the mechanism is concerted, and if CO2 has no reverse commitment, the intrinsic 13C and deuterium isotope effects are 1.0155 (corresponding to a very early transition state for C-C bond cleavage) and 7.3, and the forward commitment is 3.7. With deoxodihydroprephenate (lacking one double bond in the ring), oxidation occurs without decarboxylation, and one enantiomer has a V/K value 23-fold higher than the other (deuterium isotope effects are 3.6 and 4.1 for fast and slow isomers; V for the fast isomer is 5% and V/K 0.7% those of prephenate). The fully saturated analogue of deoxoprephenate is a very slow substrate (V 0.07% and V/K approximately 10(-5%) those of prephenate). pH profiles show a group with pK = 8.3 that must be protonated for substrate binding and a catalytic group with pK = 6.5 that is a cationic acid (likely histidine). This group facilitates hydride transfer by beginning to accept the proton from the 4-hydroxyl group of prephenate prior to the beginning of C-C cleavage (or fully accepting it in the oxidation of the analogues with only one double bond or none in the ring). In contrast with the enzymatic reaction, the acid-catalyzed decarboxylation of prephenate and deoxoprephenate (t1/2 of 3.7 min at low pH) is a stepwise reaction with a carbonium ion intermediate, since 18O is incorporated into substrate and its epi isomer during reaction in H218O. pH profiles show that the hydroxyl group must be protonated and the carboxyl (pK approximately 4.2) ionized for carbonium ion formation. The carbonium ion formed from prephenate decarboxylates 1.75 times faster than it reacts with water (giving 1.8 times as much prephenate as epi isomer). The observed 13C isotope effect of 1.0082 thus corresponds to an intrinsic isotope effect of 1.023, indicating an early transition state for the decarboxylation step. epi-Prephenate is at least 20 times more stable to acid than prephenate because it exists largely as an internal hemiketal.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
The reaction of beta-N-methylaminoalanine (BMAA) with L-amino acid oxidase (L-AAO) in the presence of catalase yields ammonia and beta-N-methylaminopyruvate, which was trapped as its 2,4-dinitrophenylhydrazone, as products. Incubation of BMAA with L-AAO in the presence of semicarbazide led to the formation of a semicarbazone, indicating intermediate iminium ion formation; when potassium cyanide (5 mM) was added, semicarbazone formation was blocked. The formation of beta-N-methylaminopyruvate was decreased by omission of catalase and was reduced in the presence of hydrogen peroxide (100 mM). These results indicate that BMAA is converted by L-AAO to the corresponding alpha-imino acid, which undergoes hydrolysis to beta-N-methylaminopyruvate. The alpha-keto acid is readily oxidized to N-methylglycine by hydrogen peroxide.  相似文献   

12.
Dihydropyrimidine amidohydrolase (EC 3.5.2.2) catalyzes the reversible hydrolysis of 5,6-dihydropyrimidines to the corresponding beta-ureido acids. Previous work has shown that incubation of this Zn2+ metalloenzyme with 2,6-dipicolinic acid, 8-hydroxyquinoline-5-sulfonic acid, or o-phenanthroline results in inactivation by Zn2+ removal by a reaction pathway involving formation of a ternary enzyme-Zn2+-chelator complex which subsequently dissociates to yield apoenzyme and the Zn2+-chelate (K. P. Brooks, E. A. Jones, B. D. Kim, and E. G. Sander, (1983) Arch. Biochem. Biophys. 226, 469-483). In the present work, the pH dependence of chelator inactivation is studied. The equilibrium constant for formation of the ternary complex is strongly pH dependent and increases with decreasing pH for all three chelators. There is a positive correlation between the value of the equilibrium constant observed for each chelator and the value of its stability constant for formation of Zn2+-chelate. The affinity of the chelators for the enzyme increases in the order 8-hydroxyquinoline-5-sulfonic acid greater than o-phenanthroline greater than 2,6-dipicolinic acid. The first-order rate constant for breakdown of the ternary complex to yield apoenzyme and Zn2+-chelate is invariant with pH for a given chelator but is different for each chelator, increasing in the reverse order. The pH dependence of the inactivation shows that two ionizable groups on the enzyme are involved in the inactivation. On the other hand, the steady-state kinetic behavior of the enzyme is well-described by ionization of a single group with a pK of 6.0 in the free enzyme. The basic form of the group is required for catalysis; protonation of the group decreases both Vmax and the apparent affinity for substrate. Conversely, binding of substrate decreases the pK of this group to about 5. L-Dihydroorotic acid is shown to be a competitive inhibitor of dihydropyrimidine amidohydrolase. Binding of L-dihydroorotic acid increases the pK of the ionizable group to 6.5. The agreement between the pK in the enzyme-L-dihydroorotic acid complex and the higher pK observed in the pH dependence of inactivation by chelators suggests that the same group is involved in the binding of acid, and chelators. The different effects of substrate and L-dihydroorotic acid on the pK suggest that the binding modes of these two ligands may be different and suggest a structural basis for the mutally exclusive substrate specificities of dihydropyrimidine amidohydrolase and dihydroorotase.  相似文献   

13.
The nature of the complexes and equilibria shown by solutions of protohaemin in dimethyl sulphoxide/water mixtures and in the presence of acid and base were studied by u.v.-visible spectrophotometry. In neutral solutions containing from 40 to 100% dimethyl sulphoxide, haemin is present as a monomeric complex in which the Cl-ion is not coordinated. Only a single pH-dependent equilibrium pK12 is observed over the range 40-80% dimethylsulphoxide, corresponding to formation of the mu-oxo dimer. As the dimethyl sulphoxide content is lowered below 35%, so the single equilibrium (pK12) is replaced by two equilibria (pK1 and pK2); with solutions of 5 microM-haemin, pK1 decreases (from pK12 7.55 in 65% dimethyl sulphoxide to pK1 approx. 1.5 in 0.01% dimethyl sulphoxide), whereas pK2 hardly changes (from pK12 7.55 in 65% to pK2 approx. 7.5 in 0.01%).  相似文献   

14.
Ouyang A  Skibo EB 《Biochemistry》2000,39(19):5817-5830
Described herein is a study of the reductive alkylation chemistry of mitosene antitumor agents. We employed a 13C-enriched electrophilic center to probe the fate of the iminium ion resulting from reductive activation. The 13C-labeled center permitted the identification of complex products resulting from alkylation reactions. In the case of DNA reductive alkylation, the type and number of alkylation sites were readily assessed by 13C NMR. Although there has been much excellent work done in the area of mitosene chemistry and biochemistry, the present study provides a number of new findings: (1) The major fate of the iminium ion is head-to-tail polymerization, even in dilute solutions. (2) Dithionite reductive activation results in the formation of mitosene sulfite esters as well as the previously observed sulfonate adducts. (3) The mitosene iminium ion alkylates the adenosine 6-amino group as well as the guanosine 2-amino group. The identification of the latter adduct was greatly facilitated by the 13C-label at the electrophilic center. (4) The mitosene iminium ion alkylates DNA at both nitrogen and oxygen centers without any apparent base selectivity. The complexity of mitosene reductive alkylation of DNA will require continued adduct isolation studies.  相似文献   

15.
A facile procedure for the small scale chemical synthesis of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPL) is reported. Under optimal conditions, 36% of the sn-glycero-3-phosphocholine was converted to DPL. This procedure is suitable for small scale preparation of DPL with very high specific radioactivity from labeled palmitic acid. Furthermore, the labeled DPL obtained will serve as precursor for the chemical or enzymatic synthesis of other labeled phospholipids.  相似文献   

16.
Hunter GA  Ferreira GC 《Biochemistry》1999,38(12):3711-3718
5-Aminolevulinate synthase catalyzes the condensation of glycine and succinyl-CoA to form CoA, carbon dioxide, and 5-aminolevulinate. This represents the first committed step of heme biosynthesis in animals and some bacteria. Lysine 313 (K313) of mature murine erythroid 5-aminolevulinate synthase forms a Schiff base linkage to the pyridoxal 5'-phosphate cofactor. In the presence of glycine and succinyl-CoA, a quinonoid intermediate absorption is transiently observed in the visible spectrum of purified murine erythroid ALAS. Mutant enzymes with K313 replaced by glycine, histidine, or arginine exhibit no spectral evidence of quinonoid intermediate formation in the presence of glycine and succinyl-CoA. The wild-type 5-aminolevulinate synthase additionally forms a stable quinonoid intermediate in the presence of the product, 5-aminolevulinate. Only conservative mutation of K313 to histidine or arginine produces a variant that forms a quinonoid intermediate with 5-aminolevulinate. The quinonoid intermediate absorption of these mutants is markedly less than that of the wild-type enzyme, however. Whereas the wild-type enzyme catalyzes loss of tritium from [2-3H2]-glycine, mutation of K313 to glycine results in loss of this activity. Titration of the quinonoid intermediate formed upon binding of 5-aminolevulinate to the wild-type enzyme indicated that the quinonoid intermediate forms by transfer of a single proton with a pK of 8.1 +/- 0.1. Conservative mutation of K313 to histidine raises this value to 8.6 +/- 0.1. We propose that K313 acts as a general base catalyst to effect quinonoid intermediate formation during the 5-aminolevulinate synthase catalytic cycle.  相似文献   

17.
Three mutants of Coprinus cinereus peroxidase (CIP) were made to mimic the substrate entrance histidine 82-glutamic acid 146 pair of the substrate channel in lignin peroxidase (LIP). Compound I formation of LIP has a low pH optimum around pH 3, while optimal formation of CIP compound I is obtained at pH 6-11. The mutants were glycine 154-->glutamic acid (G154E), proline 90-->histidine (P90H) and the double mutant P90H-G154E. All three showed kinetics of compound I formation similar to that of wt CIP between pH 3 and 9. However, the stability of compound I was strongly affected by these mutations. In wt CIP compound I is stable for approximately 30 min, while compound I of the mutants were stable for 5 s or less. The P90H and P90H-G154E mutants showed pK(a) values for the alkaline transition at least one pH unit lower than for wt CIP and the G154E mutant. We suggest that the changed electrostatic field results in destabilisation of the oxidised heme in compound I and II and that the P90H residue increases the electrostatic potential in the distal cavity thereby decreasing the pK(a) for the alkaline transition.  相似文献   

18.
A facile procedure for the small scale chemical synthesis of l,2-dipalmitoyl-Sn -glycero-3-phosphocholine (DPL) is reported. Under optimal conditions, 36% of the sn-glycero-3-phosphocholine was converted to DPL. This procedure is suitable for small scale preparation of DPL with very high specific radioactivity from labeled palmitic acid. Furthermore, the labeled DPL obtained will serve as precursor for the chemical or enzymatic synthesis of other labeled phospholipids.  相似文献   

19.
The beta-barrel and beta-helix formation, as in porins and gramicidin, respectively, represent two distinct mechanisms for ion channel formation by beta-sheet proteins in membranes. The design of beta-barrel proteins is difficult due to incomplete understanding of the basic principles of folding. The design of gramicidin-like beta-helix relies on an alternating pattern of L- and D-amino acid sequences. Recently we noticed that a short beta-sheet peptide (xSxG)(6), can form porin-like channels via self-association in membranes. Here, we proposed that glycine to D-alanine substitutions of the N-formyl-(xSxG)(6) would transform the porin-like channel into a gramicidin-like beta(12)-helical channel. The requirement of an N-formyl group for channel activity, impermeability to cations with a diameter >4 A, high monovalent cation selectivity, and the absence of either voltage gating or subconductance states upon D-alanine substitution support the idea of a gramicidin-like channel. Moreover, the circular dichroism spectrum in membranes is different, indicating a change in regular beta-sheet backbone structure. The conversion of a complex porin-like channel into a gramicidin-like channel provides a link between two different mechanisms of beta-sheet channel formation in membranes and emphasizes the importance of glycine and D-amino acid residues in protein folding and function and in the engineering of ion channels.  相似文献   

20.
Delta(5)-3-Ketosteroid isomerase catalyzes cleavage and formation of a C-H bond at a diffusion-controlled limit. By determining the crystal structures of the enzyme in complex with each of three different inhibitors and by nuclear magnetic resonance (NMR) spectroscopic investigation, we evidenced the ionization of a hydroxyl group (pK(a) approximately 16.5) of an inhibitor, which forms a low barrier hydrogen bond (LBHB) with a catalytic residue Tyr(14) (pK(a) approximately 11.5), and the protonation of the catalytic residue Asp(38) with pK(a) of approximately 4.5 at pH 6.7 in the interaction with a carboxylate group of an inhibitor. The perturbation of the pK(a) values in both cases arises from the formation of favorable interactions between inhibitors and catalytic residues. The results indicate that the pK(a) difference between catalytic residue and substrate can be significantly reduced in the active site environment as a result of the formation of energetically favorable interactions during the course of enzyme reactions. The reduction in the pK(a) difference should facilitate the abstraction of a proton and thereby eliminate a large fraction of activation energy in general acid/base enzyme reactions. The pK(a) perturbation provides a mechanistic ground for the fast reactivity of many enzymes and for the understanding of how some enzymes are able to extract a proton from a C-H group with a pK(a) value as high as approximately 30.  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