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.
Lys-258 of aspartate aminotransferase forms a Schiff base with pyridoxal phosphate and is responsible for catalysis of the 1,3-prototropic shift central to the transamination reaction sequence. Substitution of arginine for Lys-258 stabilizes the otherwise elusive quinonoid intermediate, as assessed by the long wavelength absorption bands observed in the reactions of this mutant with several amino acid substrates. The external aldimine intermediate is not detectable during reactions of this mutant with amino acids, although the inhibitor alpha-methylaspartate does slowly and stably form this species. These results suggest that external aldimine formation is one of the rate-determining steps of the reaction. The pyridoxamine-5'-phosphate-like enzyme form (330-nm absorption maximum) is unreactive toward keto acid substrates, and the coenzyme bound to this species is not dissociable from the protein.  相似文献   

2.
A Planas  J F Kirsch 《Biochemistry》1991,30(33):8268-8276
The active-site essential catalytic residue of aspartate aminotransferase, Lys 258, has been converted to Cys (K258C) by site-directed mutagenesis. This mutant retains less than 10(-6) of the wild-type activity with L-aspartate. The deleted general base was functionally replaced by selective (with respect to the other five cysteines in wild type) aminoethylation of the introduced Cys 258 with (2-bromoethyl)amine following reversible protection of the nontarget sulfhydryl groups at different stages of unfolding. The chemically elaborated mutant (K258C-EA) is 10(5) times more reactive than is K258C and has a kcat value of approximately 7% of that of wild type (WT). Km and KI values are similar to those for WT. The acidic pKa controlling V/KAsp is shifted from 7.3 (WT) to 6.0 (mutant). V/K values for amino acids are approximately 3% of those found for WT, whereas they are approximately 20% for keto acids. The value of DV increases from 1.6 for WT to 3.4 for the mutant, indicating that C alpha proton abstraction constitutes a more significant kinetic barrier for the latter enzyme. A smaller, but still significant, increase in D(V/KAsp) from 1.9 in WT to 3.0 in the mutant shows that the forward and reverse commitment factors are inverted by the mutation. The acidic limb of the V/KAsp versus pH profile, is lowered by 1.3 pH units, probably reflecting the similar difference in the basicity of the epsilon-NH2 group in gamma-thialysine versus that in lysine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
3alpha-hydroxysteroid dehydrogenase/carbonyl reductase from Comamonas testosteroni catalyzes the oxidation of androsterone with NAD(+) to form androstanedione and NADH with a concomitant releasing of protons to bulk solvent. To probe the proton transfer during the enzyme reaction, we used mutagenesis, chemical rescue, and kinetic isotope effects to investigate the release of protons. The kinetic isotope effects of (D)V and (D(2)O)V for wild-type enzyme are 1 and 2.1 at pL 10.4 (where L represents H, (2)H), respectively, and suggest a rate-limiting step in the intramolecular proton transfer. Substitution of alanine for Lys(159) changes the rate-limiting step to the hydride transfer, evidenced by an equal deuterium isotope effect of 1.8 on V(max) and V/K(androsterone) and no solvent kinetic isotope effect at saturating 3-(cyclohexylamino)propanesulfonic acid (CAPS). However, a value of 4.4 on V(max) is observed at 10 mm CAPS at pL 10.4, indicating a rate-limiting proton transfer. The rate of the proton transfer is blocked in the K159A and K159M mutants but can be rescued using exogenous proton acceptors, such as buffers, small primary amines, and azide. The Br?nsted relationship between the log(V/K(d)(-base)Et) of the external amine (corrected for molecular size effects) and pK(a) is linear for the K159A mutant-catalyzed reaction at pH 10.4 (beta = 0.85 +/- 0.09) at 5 mm CAPS. These results show that proton transfer to the external base with a late transition state occurred in a rate-limiting step. Furthermore, a proton inventory on V/Et is bowl-shaped for both the wild-type and K159A mutant enzymes and indicates a two-proton transfer in the transition state from Tyr(155) to Lys(159) via 2'-OH of ribose.  相似文献   

4.
Guinea pig liver transglutaminase (TGase) reacts with 0.1 mM N-Cbz-L-Glu(gamma-p-nitrophenyl ester)Gly (5, prepared herein, K(M) = 0.02 mM) to undergo rapid acylation that can be followed spectrophotometrically at 400 nm (pH 7.0, 25 degrees C). Deacylation of the transiently formed thiolester acyl enzyme intermediate via catalytic aminolysis was studied in the presence of six primary amines of widely varying basicity (pK(NH+) = 5.6-10.5). Steady-state kinetic studies were performed to measure k(cat) and K(M) values for each amine substrate. A Br?nsted plot constructed through the correlation of log(k(cat)/K(M)) and pK(NH+) for each amine substrate displays a linear free-energy relationship with a slope beta(nuc) = -0.37 +/- 0.08. The shallow negative slope is consistent with a general-base-catalyzed deacylation mechanism in which a proton is removed from the amine substrate during its rate-limiting nucleophilic attack on the thiolester carbonyl. Kinetic isotope effects were measured for four acceptor substrates (water, kie = 1.1 +/- 0.1; aminoacetonitrile, kie = 5.9 +/- 1.2; glycine methyl ester, kie = 3.4 +/- 0.7; N-Ac-L-lysine methyl ester, kie = 1.1 +/- 0.1) and are consistent with a proton in flight at the rate-limiting transition state. The active site general-base implicated by these kinetic results is believed to be His-334, of the highly conserved TGase Cys-His-Asp catalytic triad.  相似文献   

5.
A simple method for measuring the internal pH of chloroplasts during steady-state illuminations based on the stimulation of proton uptake by monofunctional amines was developed. Predictions of a mathematical derivation concerning the dependence of the stimulation on the amine concentration, the internal volume, the pK of the amine and the external pH have been verified experimentally. To circumvent uncoupling and swelling due to large internal accumulation of amines extrapolation of the stimulation to low amine concentrations was suggested and shown to lead to valid values. Alternatively, swelling could be largely reduced in a medium containing potassium aspartate and valinomycin.  相似文献   

6.
Saccharopine dehydrogenase [N6-(glutaryl-2)-L-lysine:NAD oxidoreductase (L-lysine forming)] catalyzes the final step in the alpha-aminoadipate pathway for lysine biosynthesis. It catalyzes the reversible pyridine nucleotide-dependent oxidative deamination of saccharopine to generate alpha-Kg and lysine using NAD+ as an oxidizing agent. The proton shuttle chemical mechanism is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. In the direction of lysine formation, once NAD+ and saccharopine bind, a group with a pKa of 6.2 accepts a proton from the secondary amine of saccharopine as it is oxidized. This protonated general base then does not participate in the reaction again until lysine is formed at the completion of the reaction. A general base with a pKa of 7.2 accepts a proton from H2O as it attacks the Schiff base carbon of saccharopine to form the carbinolamine intermediate. The same residue then serves as a general acid and donates a proton to the carbinolamine nitrogen to give the protonated carbinolamine. Collapse of the carbinolamine is then facilitated by the same group accepting a proton from the carbinolamine hydroxyl to generate alpha-Kg and lysine. The amine nitrogen is then protonated by the group that originally accepted a proton from the secondary amine of saccharopine, and products are released. In the reverse reaction direction, finite primary deuterium kinetic isotope effects were observed for all parameters with the exception of V2/K(NADH), consistent with a steady-state random mechanism and indicative of a contribution from hydride transfer to rate limitation. The pH dependence, as determined from the primary isotope effect on DV2 and D(V2/K(Lys)), suggests that a step other than hydride transfer becomes rate-limiting as the pH is increased. This step is likely protonation/deprotonation of the carbinolamine nitrogen formed as an intermediate in imine hydrolysis. The observed solvent isotope effect indicates that proton transfer also contributes to rate limitation. A concerted proton and hydride transfer is suggested by multiple substrate/solvent isotope effects, as well as a proton transfer in another step, likely hydrolysis of the carbinolamine. In agreement, dome-shaped proton inventories are observed for V2 and V2/K(Lys), suggesting that proton transfer exists in at least two sequential transition states.  相似文献   

7.
Histamine, the result of histidine decarboxylation, has been associated with allergic reactions due to the consumption of certain foods. Other biogenic amines, such as putrescine and cadaverine, have been related to quality deterioration in foods. A quantitative miniaturized method for the detection of biogenic amines produced by microorganisms in culture media, was designed. The reaction takes place in microplates containing microquantities of inoculated media and reagents. Amine production is determined spectrophotometrically by monitoring changes in the acid phase of the pH indicator at 405 nm. Using the following amino acids: histidine, phenylalanine, tyrosine, tryptophan, arginine, lysine and ornithine, 44 microorganisms were tested for amine production. Sensitivity of the method is 10 μM of amine.  相似文献   

8.
D L Smith  S C Almo  M D Toney  D Ringe 《Biochemistry》1989,28(20):8161-8167
The three-dimensional structure of a mutant of the aspartate aminotransferase from Escherichia coli, in which the active-site lysine has been substituted by alanine (K258A), has been determined at 2.8-A resolution by X-ray diffraction. The mutant enzyme contains pyridoxamine phosphate as cofactor. The structure is compared to that of the mitochondrial aspartate aminotransferase. The most striking differences, aside from the absence of the lysine side chain, occur in the positions of the pyridoxamine group and of tryptophan 140.  相似文献   

9.
Saccharopine dehydrogenase (SDH) catalyzes the final reaction in the α-aminoadipate pathway, the conversion of l-saccharopine to l-lysine (Lys) and α-ketoglutarate (α-kg) using NAD? as an oxidant. The enzyme utilizes a general acid-base mechanism to conduct its reaction with a base proposed to accept a proton from the secondary amine of saccharopine in the oxidation step and a group proposed to activate water to hydrolyze the resulting imine. Crystal structures of an open apo form and a closed form of the enzyme with saccharopine and NADH bound have been determined at 2.0 and 2.2 ? resolution, respectively. In the ternary complex, a significant movement of domain I relative to domain II that closes the active site cleft between the two domains and brings H96 and K77 into the proximity of the substrate binding site is observed. The hydride transfer distance is 3.6 ?, and the side chains of H96 and K77 are properly positioned to act as acid-base catalysts. Preparation of the K77M and H96Q single-mutant and K77M/H96Q double-mutant enzymes provides data consistent with their role as the general acid-base catalysts in the SDH reaction. The side chain of K77 initially accepts a proton from the ε-amine of the substrate Lys and eventually donates it to the imino nitrogen as it is reduced to a secondary amine in the hydride transfer step, and H96 protonates the carbonyl oxygen as the carbinolamine is formed. The K77M, H976Q, and K77M/H96Q mutant enzymes give 145-, 28-, and 700-fold decreases in V/E(t) and >103-fold increases in V?/K(Lys)E(t) and V?/K(α-kg)E(t) (the double mutation gives >10?-fold decreases in the second-order rate constants). In addition, the K77M mutant enzyme exhibits a primary deuterium kinetic isotope effect of 2.0 and an inverse solvent deuterium isotope effect of 0.77 on V?/K(Lys). A value of 2.0 was also observed for (D)(V?/K(Lys))(D?O) when the primary deuterium kinetic isotope effect was repeated in D?O, consistent with a rate-limiting hydride transfer step. A viscosity effect of 0.8 was observed on V?/K(Lys), indicating the solvent deuterium isotope effect resulted from stabilization of an enzyme form prior to hydride transfer. A small normal solvent isotope effect is observed on V, which decreases slightly when repeated with NADD, consistent with a contribution from product release to rate limitation. In addition, V?/K(Lys)E(t) is pH-independent, which is consistent with the loss of an acid-base catalyst and perturbation of the pK(a) of the second catalytic group to a higher pH, likely a result of a change in the overall charge of the active site. The primary deuterium kinetic isotope effect for H96Q, measured in H?O or D?O, is within error equal to 1. A solvent deuterium isotope effect of 2.4 is observed with NADH or NADD as the dinucleotide substrate. Data suggest rate-limiting imine formation, consistent with the proposed role of H96 in protonating the leaving hydroxyl as the imine is formed. The pH-rate profile for V?/K(Lys)E(t) exhibits the pK(a) for K77, perturbed to a value of ~9, which must be unprotonated to accept a proton from the ε-amine of the substrate Lys so that it can act as a nucleophile. Overall, data are consistent with a role for K77 acting as the base that accepts a proton from the ε-amine of the substrate lysine prior to nucleophilic attack on the α-oxo group of α-ketoglutarate, and finally donating a proton to the imine nitrogen as it is reduced to give saccharopine. In addition, data indicate a role for H96 acting as a general acid-base catalyst in the formation of the imine between the ε-amine of lysine and the α-oxo group of α-ketoglutarate.  相似文献   

10.
Catalysis by Escherichia coli and Porphyromonas gingivalis iron superoxide dismutase was activated by addition of primary amines, as measured by pulse radiolysis and stopped-flow spectrophotometry. This activation was saturable for most amines investigated, and a free energy plot of the apparent second-order rate constant of activation was linear as a function of the pK(a) of the amine, indicating activation by proton transfer. Amines provide an alternate rather than the only pathway for proton transfer, and catalysis was appreciable in the absence of amines. Solvent hydrogen isotope effects were near unity for amine activation, which is consistent with rate-contributing proton transfer if the pK(a) of the proton acceptor on the enzyme is not in the region of the pK(a) values of the amines studied, from 7.8 to 10.6. The activation of catalysis by these amines was uncompetitive with respect to superoxide, interpreted as proton transfer in a ternary complex of amine with the enzyme-bound peroxide dianion.  相似文献   

11.
Zheng R  Blanchard JS 《Biochemistry》2000,39(51):16244-16251
Ketopantoate reductase (EC 1.1.1.169) catalyzes the NADPH-dependent reduction of alpha-ketopantoate to D-(-)-pantoate in the biosynthesis of pantothenate. The pH dependence of V and V/K for the E. coli enzyme suggests the involvement of a general acid/base in the catalytic mechanism. To identify residues involved in catalysis and substrate binding, we mutated the following six strictly conserved residues to Ala: Lys72, Lys176, Glu210, Glu240, Asp248, and Glu256. Of these, the K176A and E256A mutant enzymes showed 233- and 42-fold decreases in V(max), and 336- and 63-fold increases in the K(m) value of ketopantoate, respectively, while the other mutants exhibited WT kinetic properties. The V(max) for the K176A and E256A mutant enzymes was markedly increased, up to 25% and 75% of the wild-type level, by exogenously added primary amines and formate, respectively. The rescue efficiencies for the K176A and E256A mutant enzymes were dependent on the molecular volume of rescue agents, as anticipated for a finite active site volume. The protonated form of the amine is responsible for recovery of activity, suggesting that Lys176 functions as a general acid in catalysis of ketopantoate reduction. The rescue efficiencies for the K176A mutant by primary amines were independent of the pK(a) value of the rescue agents (Bronsted coefficient, alpha = -0.004 +/-0.008). Insensitivity to acid strength suggests that the chemical reaction is not rate-limiting, consistent with (a) the catalytic efficiency of the wild-type enzyme (k(cat)/K(m) = 2x10(6) M(-1) s(-1) and (b) the small primary deuterium kinetic isotope effects, (D)V = 1.3 and (D)V/K = 1.5, observed for the wild-type enzyme. Larger primary deuterium isotope effects on V and V/K were observed for the K176A mutant ((D)V = 3.0, (D)V/K = 3.7) but decreased nearly to WT values as the concentration of ethylamine was increased. The nearly WT activity of the E256A mutant in the presence of formate argues for an important role for this residue in substrate binding. The double mutant (K176A/E256A) has no detectable ketopantoate reductase activity. These results indicate that Lys176 and Glu256 of the E. coli ketopantoate reductase are active site residues, and we propose specific roles for each in binding ketopantoate and catalysis.  相似文献   

12.
Aromatic amine dehydrogenase uses a tryptophan tryptophylquinone (TTQ) cofactor to oxidatively deaminate primary aromatic amines. In the reductive half-reaction, a proton is transferred from the substrate C1 to betaAsp-128 O-2, in a reaction that proceeds by H-tunneling. Using solution studies, kinetic crystallography, and computational simulation we show that the mechanism of oxidation of aromatic carbinolamines is similar to amine oxidation, but that carbinolamine oxidation occurs at a substantially reduced rate. This has enabled us to determine for the first time the structure of the intermediate prior to the H-transfer/reduction step. The proton-betaAsp-128 O-2 distance is approximately 3.7A, in contrast to the distance of approximately 2.7A predicted for the intermediate formed with the corresponding primary amine substrate. This difference of approximately 1.0 A is due to an unexpected conformation of the substrate moiety, which is supported by molecular dynamic simulations and reflected in the approximately 10(7)-fold slower TTQ reduction rate with phenylaminoethanol compared with that with primary amines. A water molecule is observed near TTQ C-6 and is likely derived from the collapse of the preceding carbinolamine TTQ-adduct. We suggest this water molecule is involved in consecutive proton transfers following TTQ reduction, and is ultimately repositioned near the TTQ O-7 concomitant with protein rearrangement. For all carbinolamines tested, highly stable amide-TTQ adducts are formed following proton abstraction and TTQ reduction. Slow hydrolysis of the amide occurs after, rather than prior to, TTQ oxidation and leads ultimately to a carboxylic acid product.  相似文献   

13.
Chemical modification, mutagenesis, chemical rescue, and isotope effect studies are used to identify and probe the roles of several conserved amino acid groups in catalysis by human dihydroorotate dehydrogenase. Time- and pH-dependent inactivation of human dihydroorotate dehydrogenase by trinitrobenzenesulfonate implicates at least one critical lysyl residue in catalysis. Of four highly conserved lysines, only the cognate of Lys255 was previously suggested to have catalytic functionality. We now show that replacement of either Lys184 or Lys186 by mutagenesis does not impact, whereas substitution of Lys100 abolishes, enzymatic activity. However, activity is partially restored to K100C (or K100A) by inclusion of exogenous primary amines in reaction mixtures. This rescued activity saturates with respect to numerous amines and exhibits a steric discrimination reflected in K(d,(amine)) values. For all amines, rescued k(cat) values were only approximately 10% of wild type and independent of amine basicity. K(M) values for dihydroorotate and coenzyme Q(0) were similar to wild type. Thus, exogenous amines (as surrogates for Lys100) apparently complement a chemical, not binding, step(s) of catalysis, which does not entail proton transfer. In support of this postulate, solvent kinetic isotope effect analysis indicates that Lys100 stabilizes developing negative charge on the isoalloxazine ring of flavin mononucleotide during hydride transfer, as has been observed for a number of flavoprotein oxidoreductases. Ser215 of human dihydroarotate dehydrogenase (DHODase) was also studied because of its alignment with the putative active-site base Cys130 of Lactococcus lactisDHODase. Substantial retention of activity by S215C, yet complete loss of activity for S215A, is consistent with Ser215 serving as the active-site base in the human enzyme.  相似文献   

14.
Mammalian transglutaminase (TGase) catalyzes covalent cross-linking of peptide-bound lysine residues or incorporation of primary amines to limited glutamine residues in substrate proteins. Using an unbiased M13 phage display random peptide library, we developed a screening system to elucidate primary structures surrounding reactive glutamine residue(s) that are preferred by TGase. Screening was performed by selecting phage clones expressing peptides that incorporated biotin-labeled primary amine by the catalytic reactions of TGase 2 and activated Factor XIII (Factor XIIIa). We identified several amino acid sequences that were preferred as glutamine donor substrates, most of which have a marked tendency for individual TGases: TGase 2, QxPphiD(P), QxPphi, and QxxphiDP; Factor XIIIa, QxxphixWP (where x and phi represent a non-conserved and a hydrophobic amino acid, respectively). We further confirmed that the sequences were favored for transamidation using modified glutathione S-transferase (GST) for recombinant peptide-GST fusion proteins. Most of the fusion proteins exhibited a considerable increase in incorporation of primary amines over that of modified GST alone. Furthermore, we identified the amino acid sequences that demonstrated higher specificity and inhibitory activity in the cross-linking reactions by TGase 2 and Factor XIIIa.  相似文献   

15.
The active site residue lysine 258 of chicken mitochondrial aspartate aminotransferase was replaced with a histidine residue by means of site-directed mutagenesis. The mutant protein was expressed in Escherichia coli and purified to homogeneity. Addition of 2-oxoglutarate to its pyridoxamine form changed the coenzyme absorption spectrum (lambda max = 330 nm) to that of the pyridoxal form (lambda max = 330/392 nm). The rate of this half-reaction of transamination (kcat = 4.0 x 10(-4)s-1) is five orders of magnitude slower than that of the wild-type enzyme. However, the reverse half-reaction, initiated by addition of aspartate or glutamate to the pyridoxal form of the mutant enzyme, is only three orders of magnitude slower than that of the wild-type enzyme, kmax of the observable rate-limiting elementary step, i.e. the conversion of the external aldimine to the pyridoxamine form, being 7.0 x 10(-2)s-1. Aspartate aminotransferase (Lys258----His) thus represents a pyridoxal-5'-phosphate-dependent enzyme with significant catalytic competence without an active site lysine residue. Apparently, covalent binding of the coenzyme, i.e. the internal aldimine linkage, is not essential for the enzymic transamination reaction, and a histidine residue can to some extent substitute for lysine 258 which is assumed to act as proton donor/acceptor in the aldimine-ketimine tautomerization.  相似文献   

16.
Plant monofunctional aspartate kinase is unique among all aspartate kinases, showing synergistic inhibition by lysine and S-adenosyl-l-methionine (SAM). The Arabidopsis genome contains three genes for monofunctional aspartate kinases. We show that aspartate kinase 2 and aspartate kinase 3 are inhibited only by lysine, and that aspartate kinase 1 is inhibited in a synergistic manner by lysine and SAM. In the absence of SAM, aspartate kinase 1 displayed low apparent affinity for lysine compared to aspartate kinase 2 and aspartate kinase 3. In the presence of SAM, the apparent affinity of aspartate kinase 1 for lysine increased considerably, with K(0.5) values for lysine inhibition similar to those of aspartate kinase 2 and aspartate kinase 3. For all three enzymes, the inhibition resulted from an increase in the apparent K(m) values for the substrates ATP and aspartate. The mechanism of aspartate kinase 1 synergistic inhibition was characterized. Inhibition by lysine alone was fast, whereas synergistic inhibition by lysine plus SAM was very slow. SAM by itself had no effect on the enzyme activity, in accordance with equilibrium binding analyses indicating that SAM binding to aspartate kinase 1 requires prior binding of lysine. The three-dimensional structure of the aspartate kinase 1-Lys-SAM complex has been solved [Mas-Droux C, Curien G, Robert-Genthon M, Laurencin M, Ferrer JL & Dumas R (2006) Plant Cell18, 1681-1692]. Taken together, the data suggest that, upon binding to the inactive aspartate kinase 1-Lys complex, SAM promotes a slow conformational transition leading to formation of a stable aspartate kinase 1-Lys-SAM complex. The increase in aspartate kinase 1 apparent affinity for lysine in the presence of SAM thus results from the displacement of the unfavorable equilibrium between aspartate kinase 1 and aspartate kinase 1-Lys towards the inactive form.  相似文献   

17.
He Z  Toney MD 《Biochemistry》2006,45(15):5019-5028
Chorismate-utilizing enzymes catalyze diverse reactions, providing critical physiological functions unique to plants, bacteria, fungi, and some parasites. Their absence in animals makes them excellent targets for antimicrobials and herbicides. 4-Amino-4-deoxychorismate synthase (ADCS) catalyzes the first step in folate biosynthsis and shares a common core mechanism with isochorismate synthase (IS) and anthranilate synthase (AS), in which nucleophile addition at C2 initiates these reactions. Evidence was presented previously [He, Z., Stigers Lavoie, K. D., Bartlett, P. A., and Toney, M. D. (2004) J. Am. Chem. Soc. 126, 2378-2385] that K274 is the nucleophile in ADCS, implying formation of a covalent intermediate. Herein, we report the direct detection of this covalent intermediate formed in ADCS-catalyzed reactions by ESI-MS. Difference spectra show the covalent intermediate has an absorption maximum at 310 nm. This was used to study the pre-steady-state kinetics of covalent intermediate formation under various conditions. Additionally, E258 in ADCS was shown to be critical to formation of the covalent intermediate by acting as a general acid catalyst for loss of the C4 hydroxyl group. The E258A/D mutants both exhibit very low activity. Acetate is a poor chemical rescue agent for E258D but an excellent one for E258A, with a 20000-fold and 3000-fold rate increase for Gln-dependent and NH(4)(+)-dependent activities, respectively. Lastly, A213 in IS (structurally homologous to K274 in ADCS) was changed to lysine in an attempt to convert IS to an ADCS-like enzyme. HPLC studies support the formation of a covalent intermediate with this mutant.  相似文献   

18.
Beta‐amyloid peptide (Aβ) is the major protein constituent found in senile plaques in Alzheimer's disease (AD). It is believed that Aβ plays a role in neurodegeneration associated with AD and that its toxicity is related to its structure or aggregation state. In this study, an approach based on chemical modification of primary amines and mass spectrometric (MS) detection was used to identify residues on Aβ peptide that were exposed or buried upon changes in peptide structure associated with aggregation. Results indicate that the N terminus was the most accessible primary amine in the fibril, followed by lysine 28, then lysine 16. A kinetic analysis of the data was then performed to quantify differences in accessibility between these modification sites. We estimated apparent equilibrium unfolding constants for each modified site of the peptide, and determined that the unfolding constant for the N terminus was approximately 100 times greater than that for K28, which was about six times greater than that for K16. Understanding Aβ peptide structure at the residue level is a first step in designing novel therapies for prevention of Aβ structural transitions and/or cell interactions associated with neurotoxicity in Alzheimer's disease. Biotechnol. Bioeng. 2009; 104: 181–192 © 2009 Wiley Periodicals, Inc.  相似文献   

19.
The ionization state of the phosphate group bound at the aspartate aminotransferase apoenzyme's active site has been investigated utilizing Fourier-transform infrared spectroscopy following the band corresponding to the symmetric stretching of the dianionic phosphate. Unlike free phosphate, when inorganic phosphate is bound at the enzyme's active site, the integrated intensity value of the dianionic band does not change with pH within the studied range, and this value is similar to that for free dianionic phosphate at pH 8.3. From these results, we propose a dianionic state for the phosphate ion bound to cytosolic aspartate aminotransferase throughout the pH range of 5.7-8.3. The presence of other anions such as acetate and chloride or the substrate aspartate and its analogues produces a pH-dependent phosphate removal from the active site which is favored at low pH values. Elimination of the charged primary amine at the active-site Lys-258, through formation of a Schiff base with pyridoxal or chemical modification by carbamylation, also produces a pH-independent phosphate release. These results are interpreted as Lys-258 together with the active-site alpha-helix and other residues may be involved in stabilizing phosphate as a dianion in the apoenzyme phosphate pocket which anchors the phosphate ester of pyridoxal phosphate in the holoenzyme. It is proposed that the dianionic phosphate contributes to the apoenzyme's thermal stability through formation of strong hydrogen bond and salt bridges with the amino acid residues forming the phosphate binding pocket with assistance of Lys-258, and other active-site cationic components.  相似文献   

20.
The chemical reaction of monomeric acrylamide with primary, secondary, and tertiary amines, used as buffer components in polyacrylamide gel electrophoresis systems, was investigated in the basic pH range. Adduct formation proceeded for several minutes up to weeks, depending on the reactivity of the amino groups. A pH shift in the reaction mixture due to an altered pK value of the reaction product was observed. However, a few primary amines (tris(hydroxymethyl)aminomethane, 2-amino-2-methyl-1,3-propanediol) and secondary amines 3-([2-hydroxy-1,1-bis(hydroxymethyl)ethyl]amino)-1-propanesulfonic acid, 3-(dimethyl(hydroxymethyl)methylamino)-2-hydroxypropanesulfonic acid) showed negligible shifts of pH. They are, therefore, useful as components in the polymerization mixture; whereas some tertiary amines showing complete pH stability as well (e.g., triethanolamine) are not suitable, as they acted as accelerators of gel polymerization. Acrylamide can also covalently bind to proteins by reacting with the epsilon-amino group of lysine residues, especially. Bovine serum albumin, having an acidic isoelectric point, and the basic protein cytochrome c were treated with different acrylamide concentrations at alkaline pH yielding modified protein molecules with altered electrophoretic mobilities in different polyacrylamide gel electrophoresis systems. This reaction gave rise to artifacts in alkaline polyacrylamide gels and isoelectric focusing systems when residual acrylamide monomers were still present in the gel matrix after the polymerization process ceased.  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