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 109 毫秒
1.
The elimination of halide ion from either 5-bromo- or 5-iodo-5,6-dihydrouracil to yield uracil is a slow reaction which, in the case of 5-iodo-5,6-dihydrouracil, is 400 times slower than the enzymatic release of 125I? from 5-[125I]iodouracil. The elimination of HBr from 5-bromo-5,6-dihydrouracil is subject to general base catalysis by tris(hydroxymethyl)aminomethane (k2Tris base = 11 × 10?4M?1 min?1, 37°C, ionic strength 1.0 M). At pH values near and above physiological, both the bromo- and iododihydropyrimidines are subject to hydrolysis of the dihydropyrimidine ring, a reaction which parallels halide elimination to yield uracil. The resulting 2-halo-3-ureidopropionate then cyclizes via intramolecular attack of the ureido oxygen atom to yield halide ion and 2-amino-2-oxazoline-5-carboxylic acid as final products. In dilute hydroxide ion, the kinetics of 5-bromo-5,6-dihydrouracil hydrolysis (25°C, ionic strength 1.0 M) show a change in rate-determining step as a function of increasing hydroxide ion concentration, a result which, as in the case of 5,6-dihydrouracil, can be explained in terms of the formation of a tetrahedral addition intermediate. The data are discussed relative to enzymatically catalyzed halopyrimidine dehalogenation.  相似文献   

2.
S L Johnson  P T Tuazon 《Biochemistry》1977,16(6):1175-1183
The rate of the primary acid modification reaction of 1,4-dihydronicotinamide adenine dinucleotide (NADH) and 1,4-dihydro-3-acetylpyridine adenine dinucleotide (APADH) and their analogues has been studied over a wide pH range (pH 1-7) with a variety of general acid catalysts. The rate depends on [H+] at moderate pH and becomes independent of [H+] at low pH. This behavior is attributed to substrate protonation at the carbonyl group (pK of NADH = 0.6). The reaction is general acid catalyzed; large solvent deuterium isotope effects are observed for the general acid and lyonium ion terms. Most buffers cause a linear rate increase with increasing buffer concentration, but certain buffers cause a hyperbolic rate increase. The nonlinear buffer effects are due to complexation of the buffer with the substrate, rather than to a change in rate-limiting step. The rate-limiting step is a proton transfer from the general acid species to the C5 position of the substrate. Anomerization is not a necessary first step in the case of the primary acid modification reaction of beta-NADH, in which beta to alpha anomerization takes place.  相似文献   

3.
Incubations of allene oxide synthases of flax or maize with the E,E-isomers of the 13- and 9-hydroperoxides of linoleic acid (E,E-13- and E,E-9-HPOD, respectively) at pH 7.5 afforded substantial yields of trans-disubstituted cyclopentenones. Under the conditions used, (Z,E)-HPODs were converted mainly into -ketols and afforded only trace amount of cyclopentenones. These findings indicated that changing the double bond geometry from Z to E dramatically increased the rate of formation of the pericyclic pentadienyl cation intermediate necessary for electrocyclization of 18:2-allene oxides and thus the yield of cyclopentenones. The well-known cyclization of the homoallylic allene oxide (12,13-EOT) derived from -linolenic acid 13-hydroperoxide (E,Z-13-HPOT) into cis-12-oxo-10,15-phytodienoic acid was suppressed at pH below neutral and was not observable at pH 4.5. In contrast, cyclization of the allene oxide ((9E)-12,13-EOD) derived from (E,E)-13-HPOD was slightly favoured at low pH. The finding that the cyclizations of 12,13-EOT and (9E)-12,13-EOD were differently affected by changes in pH suggested that the mechanisms of cyclization of these allene oxides are distinct.  相似文献   

4.
S-Carboxymethylcysteine, formed by the reaction of iodoacetic acid with cysteine, was found to undergo intramolecular cyclization to yield 3-oxo-(2H,3H,5H,6H-1,4-thiazine)-5-carboxylic acid. The cyclization was studied under various conditions and the product was isolated and characterized. S-Carboxyethylcysteine, formed by the reaction of 3-bromopropionic acid with cysteine, did not undergo the cyclization reaction. The use of 3-bromopropionic acid was examined as an alternative to iodoacetic acid for the protection and determination of protein thiol groups.  相似文献   

5.
The biological function of the post-translationally methylated amino acid gamma-N-methylasparagine (gamma-NMA) in proteins is unknown. We are examining the premise that amide methylation protects against deamidation. The free amino acids Asn, gamma-NMA, Gln, and delta-N-methylglutamine (delta-NMG) were incubated at elevated temperature and a variety of pH conditions to assay for deamidation. Gln disappears 12- to 14-fold more rapidly than delta-NMG, and Asn hydrolyzes to Asp and NH3 as expected. However, the gamma-NMA deamidation rate is severely overestimated by simply measuring the disappearance of starting material because gamma-NMA undergoes a cyclization reaction in preference to deamidation. At pH 1 the predominant gamma-NMA reaction is formation of stable 3-amino-N-methylsuccinimide (NMS) and this occurs greater than 10-fold faster than Asn deamidation. At pH 4.0, 7.4, and 9.0 NMS is readily formed but it is unstable and partitions between the parent compound, gamma-NMA, and a second species, alpha-N-methylasparagine. At pH 7.4 and 9.0 gamma-NMA disappears 4-fold slower than Asn but the methyl amide hydrolysis rate is diminished by as much as 13-fold. The Asn incubations over the pH range 1-9 yield scant evidence of a succinimide intermediate. It is concluded that the amide methylation provides a unique reaction pathway and stabilization for the N-methylsuccinimide species. Amino acid analysis by o-phthalaldehyde postcolumn reaction fails to detect isoasparagine, alpha-N-methylasparagine, and NMS. Amino acid analysis by precolumn derivatization with phenyl isothiocyanate destroys NMS and therefore cannot quantitate this compound. The ninhydrin postcolumn derivatization method is able to detect and quantitate all of these amino acid species.  相似文献   

6.
E W Hafner  D Wellner 《Biochemistry》1979,18(3):411-417
The reactivity of the imino acids formed in the D- or L-amino acid oxidase reaction was studied. It was found that: (1) When imino acids reacted with the alpha-amino group of glycine or other amino acids, transimination yielded derivatives less stable to hydrolysis than the parent imino acids. In contrast, when imino acids reacted with the epsilon-amino group of lysine or other primary amines, transimination yielded derivatives more stable to hydrolysis than the parent imino acids. (2) Imino acids react rapidly with hydrazine and semicarbazide, forming stable hydrazones and semicarbazones. At pH 7.7, the rate of reaction of the imino acid analogue of leucine with semicarbazide was 10(4) times greater than that of the corresponding keto acid. The reaction of imino acids with these reagents is rapid enough to permit one to follow spectrophotometrically the amino acid oxidase reaction. Imino acids also reacted with cyanide to yield stable adducts. (3) The rate of hydrolysis of the imino acid analogue of leucine was independent of pH above pH 8.5. At lower pH values, the rate of hydrolysis increased with decreasing pH. At 25 degrees C and in the absence of added amino compounds, this imino acid had a half-life of 22 s at pH 8.5. Its half-life was 9.9 s at pH 7.9.  相似文献   

7.
Tentoxin is a naturally occurring phytotoxic cyclic tetrapeptide excreted by fungi of the Alternaria alternata family. The four total syntheses of tentoxin published to date give poor total yields, mainly owing to two difficulties, the introduction of the dehydro amino acid and more especially the cyclization step. Here we describe a method that stereospecifically introduces Z-dehydrophenylalanine (deltaZPhe) by a modified Erlenmeyer aldolization reaction. The linear tetrapeptide, Boc-R1Ala-Leu-R2deltaZPhe-G1y-OMe (R1, R2: CH3, 14CH3), the precursor of tentoxin, was obtained in a 72% yield from Boc-Leu-Gly-OH. This linear tetrapeptide, labelled with carbon-14, was used for a comparative study of four cyclization reagents DPPA, DCC-PfpOH, HBTU and HATU. This last was the most effective and gave tentoxin in a 81% cyclization yield. The activated ester formed with this reagent displayed an enhanced capacity for cyclization, permitting cyclization in concentrated medium (10 mM). This new synthetic route gave tentoxin in a 60% yield from Boc-Leu-Gly-OH and offers a means of achieving the synthesis of hitherto elusive analogues.  相似文献   

8.
The mechanism of the L-threo-3-methylaspartate ammonia-lyase (EC 4.3.1.2) reaction has been probed using deuterium and solvent isotope effects with three different substrates, (2S,3S)-3-methylaspartic acid, (2S)-aspartic acid and (2S,3R)-3-methylaspartic acid. Each substrate appears to form a covalent adduct with the enzyme through the amination of a dehydroalanine (DehydAla-173) residue. The true substrates are N-protonated and at low pH, the alkylammonium groups are deprotonated internally in a closed solvent-excluded pocket after K+ ion, an essential cofactor, has become bound to the enzyme. At high pH, the amino groups of the substrates are able to react with the dehydroalanine residue prior to K+ ion binding. This property of the system gives rise to complex kinetics at pH 9.0 or greater and causes the formation of dead-end complexes which lack Mg2+ ion, another essential cofactor. The enzyme-substrate adduct is subsequently deaminated in two elimination processes. Hydrazines act as alternative substrates in the reverse reaction direction in the presence of fumaric acid derivatives, but cause irreversible inhibition in their absence. Borohydride and cyanide are not inhibitors. N-Ethylmaleimide also irreversibly inactivates the enzyme and labels residue Cys-361. The inactivation process is enhanced in the presence of cofactor Mg2+ ions and Cys-361 appears to serve as a base for the removal of the C-3 proton from the natural substrate, (2S,3S)-3-methylaspartic acid. The dehydroalanine residue appears to be protected in the resting form of the enzyme by generation of an internal thioether cross-link. The binding of the substrate and K+ ion appear to cause a conformational change which requires hydroxide ion. This is linked to reversal of the thioether protection step and generation of the base for substrate deprotonation at C-3. The deamination reaction displays high reverse reaction commitments and independent evidence from primary deuterium isotope effect data indicates that a thiolate acts as the base for deprotonation at C-3.  相似文献   

9.
A J Zaug  J R Kent  T R Cech 《Biochemistry》1985,24(22):6211-6218
During self-splicing of the Tetrahymena rRNA precursor, the intervening sequence (IVS) is excised as a unique linear molecule and subsequently cyclized. Cyclization involves formation of a phosphodiester bond between the 3' end and nucleotide 16 of the linear RNA, with release of an oligonucleotide containing the first 15 nucleotides. We find that the rate of cyclization is independent of pH in the range 4.7-9.0. A minor site of cyclization at nucleotide 20 is characterized. Cyclization to this site becomes more prominent at higher pHs, although under all conditions examined it is minor compared to cyclization at nucleotide 16. The circular IVS RNAs are unstable, undergoing hydrolysis at the phosphodiester bond that was formed during cyclization. We find that the rate of site-specific hydrolysis is first order with respect to hydroxide ion concentration, with a rate constant 10(3)-10(4)-fold greater than that of hydrolysis of strained cyclic phosphate esters. On the basis of these results, we propose that circular IVS RNA hydrolysis involves direct attack of OH- on the phosphate at the ligation junction, that particular phosphate being made particularly reactive by the folding of the RNA molecule. Cyclization, on the other hand, appears to occur by direct attack of the 3'-terminal hydroxyl group of the linear IVS RNA without prior deprotonation.  相似文献   

10.
The kinetics of reduction of AFB(1) to AFB(2) and AFG(1) to AFG(2) by lactic acid has been investigated in dilute aqueous acidic solutions (pH 3.35-4.50) as a function of the concentrations of lactic acid, AFB(1), AFG(1) and hydrogen ion at 37 degrees C. The rate of the reaction was found to be first order with respect to the concentrations of lactic acid and aflatoxins and independent on hydrogen ion concentration. The experimental results are interpreted in terms of mechanisms involving an initial formation of transient oxonium intermediate, which tends to polarize the olefinic (C=C) carbon, which in turn causes the hydride abstraction from alpha-carbon atom of lactic acid in rate determining step. The proposed mechanisms involve an overall transfer of two protons and two electrons from lactic acid to AFB(1) and AFG(1) to give the corresponding reduced less toxic products AFB(2) and AFG(2) and the oxidised product pyruvic acid.  相似文献   

11.
The physical and chemical reaction pathways of the metabolite model compound benzo(a)pyrene-7,8-diol-9,10-epoxide (BPDE) in aqueous (double-stranded) DNA solutions was investigated as a function of temperature (0-30 degrees C), pH (7.0-9.5), sodium chloride concentration (0-1.5M) and DNA concentration in order to clarify the relationships between the multiple reaction mechanisms of this diol epoxide in the presence of nucleic acids. The reaction pathways are (1) noncovalent intercalative complex formation with DNA, characterized by the equilibrium constant K, and Xb the fraction of molecules physically bound; (2) accelerated hydrolysis of BPDE bound to DNA; (3) covalent binding to DNA; and (4) hydrolysis of free BPDE(kh). The DNA-induced hydrolysis of BPDE to tetraols and the covalent binding to DNA are parallel pseudo-first-order reactions. Following the rapid (millisecond time scale) noncovalent complex formation between BPDE and DNA, a much slower (approximately minutes) H+-dependent (either specific or general acid catalysis) formation of a DNA-bound triol carbonium ion (rate constant k3) occurs. At pH 7.0 the activation energy of k3 is 8.7 +/- 0.9 kcal/mol, which is lower than the activation energy of hydrolysis of free BPDE in buffer solution (14.2 +/- 0.7 kcal/mol), and which thus partially accounts for the acceleration of hydrolysis of BPDE upon complexation with DNA. The formation of the triol carbonium ion is followed by a rapid reaction with either water to form tetraols (rate constant kT), or covalent binding to DNA (kc). The fraction of BPDE molecules which undergo covalent binding is fcov approximately equal to kc/(kc + kT) = 0.10 and is independent of the overall BPDE reaction rate constant k = kh(1 - Xb) + k3Xb if Xb----1.0, or is independent of Xb as long as k3Xb much greater than kh(1 - Xb). Thus, at Xb = 0.9, fcov is independent of pH (7.0-9.5) even though k exhibits a 70-fold variation in this pH range and k----kh above pH 9 (k3 = kh). Similarly, fcov is independent of temperature (0-30 degrees C), while k varies by a factor of approx. 3. In the range of 0-1.5 M NaCl, fcov decreases from 0.10 to 0.04. These variations are attributed to a combination of salt-induced variations in the factors k3, Xb and the ratio kc/kT.  相似文献   

12.
The decomposition of 2-chloroethylphosphonic acid in aqueous solution has been studied at pH values from 6 to 9 and at temperatures in the 30 to 55 C range. The rate of decomposition is estimated from the rate of formation of ethylene. The rate is proportional to the concentration of the phosphonate dianion and is independent of the hydroxyl ion concentration. The rate constant at 40 C is 1.9 × 10−4 sec−1 and the activation energy is 29.8 kcal mol−1. The rate of reaction is not affected significantly by the presence of potassium iodide or urea (substances which increase the rate of leaf abscission in trees sprayed by 2-chloroethylphosphonic acid). The rate decreases slightly in the presence of low concentrations of magnesium and calcium ions.  相似文献   

13.
The effects of pH, solvent isotope, and primary isotope replacement on substrate dehydrogenation by Rhodotorula gracilis d-amino acid oxidase were investigated. The rate constant for enzyme-FAD reduction by d-alanine increases approximately fourfold with pH, reflecting apparent pKa values of approximately 6 and approximately 8, and reaches plateaus at high and low pH. Such profiles are observed in all presteady-state and steady-state kinetic experiments, using both d-alanine and d-asparagine as substrates, and are inconsistent with the operation of a base essential to catalysis. A solvent deuterium isotope effect of 3.1 +/- 1.1 is observed on the reaction with d-alanine at pH 6; it decreases to 1.2 +/- 0.2 at pH 10. The primary substrate isotope effect on the reduction rate with [2-D]d-alanine is 9.1 +/- 1.5 at low and 2.3 +/- 0.3 at high pH. At pH 6.0, the solvent isotope effect is 2.9 +/- 0.8 with [2-D]d-alanine, and the primary isotope effect is 8.4 +/- 2.4 in D2O. Thus, primary and solvent kinetic isotope effects (KIEs) are independent of the presence of the other isotope, i.e. the 'double' kinetic isotope effect is the product of the individual KIEs, consistent with a transition state in which rupture of the two bonds of the substrate to hydrogen is concerted. These results support a hydride transfer mechanism for the dehydrogenation reaction in d-amino acid oxidase and argue against the occurrence of any intermediates in the process. A pKa,app of approximately 8 is interpreted to arise from the microscopic ionization of the substrate amino acid alpha-amino group, but also includes contributions from kinetic parameters.  相似文献   

14.
Previously proposed mechanisms for the production of L-citrulline from L-arginine by macrophage nitric oxide (NO.) synthase involve either hydrolysis of arginine or hydration of an intermediate and thus predict incorporation of water oxygen into L-citrulline. Macrophage NO. synthase was incubated with L-arginine, NADPH, tetrahydrobiopterin, FAD, and dithiothreitol in H2(18)/16O2. L-Citrulline produced in this reaction was analyzed with gas chromatography/mass spectrometry. Its mass spectrum matched that of L-citrulline generated in H2(16)O/16O2. The base fragment ion of m/z 99 was shown to contain the ureido carbonyl group by using L-[guanidino-13C]arginine as substrate. When the enzyme reaction was performed in H2(16)O/18O2, the base fragment ion shifted to m/z 101 with L-[guanidino-12C]arginine as the substrate and to m/z 102 with L-[guanidino-13C]arginine. These results indicate that the ureido oxygen of the L-citrulline product of macrophage NO.synthase derives from dioxygen and not from water.  相似文献   

15.
N P Botting  D Gani 《Biochemistry》1992,31(5):1509-1520
The enzyme 3-methylaspartate ammonia-lyase (EC 4.3.1.2) catalyzes the exchange of the C-3 hydrogen of the substrate, (2S,3S)-3-methylaspartic acid, with solvent hydrogen. The mechanism of the exchange reaction was probed using (2S,3S)-3-methylaspartic acid and its C-3-deuteriated isotopomer. Incubations conducted in tritiated water allowed the rate of protium or deuterium wash-out from the substrates to be measured as tritium wash-in. The primary deuterium isotope effects for the exchange under essentially Vmax conditions ( [S] much greater than Km) were 1.6, 1.5, and 1.5 at pH 9.0, 7.6, and 6.5. The deamination reaction, measured spectrophotometrically on the same incubations, showed isotope effects of 1.7, 1.6, and 1.4 at pH 9.0, 7.6, and 6.5, in agreement with the values of DV and D(V/K) reported previously [Botting, N.P., Akhtar, M., Cohen, M.A., & Gani, D. (1988) Biochemistry 27, 2956-2959]. The ratio of the rate of exchange to the rate of deamination, however, varied widely with pH. Together with the identical values of the primary isotope effects for the two reactions, this result indicates that the partition between reaction pathways occurs after the slowest steps in the common part of the reaction coordinate pathway, almost certainly after the cleavage of the C-N bond at the level of the enzyme-ammonia-mesaconic acid complex, and not at the putative carbanion level as was previously suggested. The enzyme requires both K+ and Mg2+ ions for activity, although ammonium ion is also able to bind in the K+ site and act as an activator. Variation of the metal ion concentration alters the magnitude of the primary deuterium isotope effects. The variation of potassium ion concentration causes the most marked changes: at 1.6 mM K+, DV and D(V/K) are 1.7, whereas at 50 mM K+, DV and D(V/K) are reduced to 1.0. The isotope effects are also reduced at low K+ concentration due to the emergence of a slow-acting high K+ affinity monopotassium form of the enzyme. The binding order and role of the metal ion cofactors and their influence in determining the formal mechanism of the reaction is discussed, and the failure of previous workers to observe primary deuterium isotope effects for the deamination process is explained. The product desorption order was tested by product inhibition, alternative product inhibition, and isotope exchange experiments. Ammonia and mesaconic acid debind in a random fashion.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
The rate constants associated with the series of successive transient absorptions initiated by one-electron oxidation of 5,6-dihydroxyindole (DHI), 5,6-dihydroxyindole-2-carboxylic acid (DHICA), precursors of melanin, and N-methyl-5,6-dihydroxyindole (NMDHI), a model compound, have been studied by pulse radiolysis. The initial transient species resulting from N3. oxidation reaction at pH 7.3-7.4 are assigned as the corresponding semiquinones. In each case, these radicals decayed, probably by disproportionation, into products most readily monitored in the 400-430 nm region. For DHI, the decay in this region could be fitted by two parent concentration independent first-order processes. These may correspond to transformations between 5,6-indolequinone, and its quinone-imine and quinone-methide tautomers. With NMDHI, on the other hand, a single longer-lived product with a peak around 430 nm predominated after decay of the corresponding radical, due almost certainly to N-methyl-5,6-indolequinone. The data appear to exclude significant melanin polymerisation by condensation of semiquinones, reaction of semiquinones with dihydroxyindoles, self-addition of indolequinones or tautomers, or reaction of indolequinones or tautomers with the parent dihydroxyindoles. It is suggested that polymerisation of melanin may rather occur by stepwise addition of indolequinone methide/imine to reduced oligomeric species.  相似文献   

17.
S L Ausenhus  M H O'Leary 《Biochemistry》1992,31(28):6427-6431
In addition to the normal carboxylation reaction, phosphoenolpyruvate carboxylase from Zea mays catalyzes a HCO3(-)-dependent hydrolysis of phosphoenolpyruvate to pyruvate and Pi. Two independent methods were used to establish this reaction. First, the formation of pyruvate was coupled to lactate dehydrogenase in assay solutions containing high concentrations of L-glutamate and aspartate aminotransferase. Under these conditions, oxalacetic acid produced in the carboxylation reaction was efficiently transaminated, and decarboxylation to form spurious pyruvate was negligible. Second, sequential reduction of oxalacetate and pyruvate was achieved by initially running the reaction in the presence of malate dehydrogenase with NADH in excess over phosphoenolpyruvate. After the reaction was complete, lactate dehydrogenase was added, thus giving a measure of pyruvate concentration. At pH 8.0 in the presence of Mg2+, the rate of phosphoenolpyruvate hydrolysis was 3-7% of the total reaction rate. The hydrolysis reaction catalyzed by phosphoenolpyruvate carboxylase was strongly metal dependent, with rates decreasing in the order Ni2+ greater than Co2+ greater than Mn2+ greater than Mg2+ greater than Ca2+. These results suggest that the active site metal ion binds to the enolate oxygen, thus stabilizing the proposed enolate intermediate. The more stable the enolate, the less reactive it is toward carboxylation and the greater the opportunity for hydrolysis.  相似文献   

18.
Electropolymerized film of 3,3',3',3'-tetraaminophthalocyanatonickel(II) (p-Ni(II)TAPc) on glassy carbon (GC) electrode was used for the selective and stable determination of 3,4-dihydroxy-L-phenylalanine (L-dopa) in acetate buffer (pH 4.0) solution. Bare GC electrode fails to determine the concentration of L-dopa accurately in acetate buffer solution due to the cyclization reaction of dopaquinone to cyclodopa in solution. On the other hand, p-Ni(II)TAPc electrode successfully determines the concentration of L-dopa accurately because the cyclization reaction was prevented at this electrode. It was found that the electrochemical reaction of L-dopa at the modified electrode is faster than that at the bare GC electrode. This was confirmed from the higher heterogeneous electron transfer rate constant (k(0)) of L-dopa at p-Ni(II)TAPc electrode (3.35 x 10(-2) cms(-1)) when compared to that at the bare GC electrode (5.18 x 10(-3) cms(-1)). Further, it was found that p-Ni(II)TAPc electrode separates the signals of ascorbic acid (AA) and L-dopa in a mixture with a peak separation of 220 mV. Lowest detection limit of 100 nM was achieved at the modified electrode using amperometric method. Common physiological interferents like uric acid, glucose and urea does not show any interference within the potential window of L-dopa oxidation. The present electrode system was also successfully applied to estimate the concentration of L-dopa in the commercially available tablets.  相似文献   

19.
The tripeptide Bz-Arg-Gly-Asp(-OMe)-OH was synthesized by enzymatic method. Bz-Arg-Gly-OEt was synthesized by trypsin in ethanol containing 0.1 M Tris/HCl buffer (pH 8.0), and then H-Asp(-OMe)(2) was incorporated into the Bz-Arg-Gly-OEt using chymopapain in 0.25M CHES/NaOH buffer (pH = 9.0, EDTA 10 mM). The yield of Bz-Arg-Gly-OEt and Bz-Arg-Gly-Asp(-OMe)-OH were 80% and 70% using 1M Bz-Arg-OEt and 0.5M Bz-Arg-Gly-OEt, respectively. For Bz-Arg-Gly-OEt synthesis reaction at high concentrations of the substrates, the buffer content in ethanol was a key factor to determine the optimal reaction condition. In Bz-Arg-Gly-Asp(-OMe)-OH synthesis reaction, the yield was low in organic solvent due to various side products such as Bz-Arg-OH, Bz-Arg-Gly-OH, and Bz-Arg-Gly-Asp(-OMe)-Asp(-OMe)-OH, suggesting that chymopapain has a very broad substrate specificity of the S(1) site. The Bz-Arg-Gly-Asp(-OMe)-OH synthesis rate and its yield were dramatically elevated and the side reactions were reduced using only the CHES/NaOH buffer (pH = 9.0, EDTA 10 mM) as a reaction media. The final product Bz-Arg-Gly-Asp(-OMe)-OH was identified to be formed via C-terminal hydrolysis of Bz-Arg-Gly-Asp(-OMe)(2) after the nucleophile, H-Asp(-OMe)(2), was added.  相似文献   

20.
The kinetics and spectroscopic properties of the single polypeptide and proteolytically cleaved form of recombinant Fe(3+)Fe(2+) human purple acid phosphatase (recHPAP) exhibit significant differences, primarily due to a difference in pK(es,1) (the value of an acid dissociation constant of the ES complex). These differences are due to the presence or absence, respectively, of an interaction between an aspartate residue in an exposed loop of the protein and one or more active site residues. To further explore the origin of these differences, the ferrous ion of recHPAP has been replaced by zinc. Analysis of the reconstituted Fe(3+)Zn(2+)recHPAP reveals an unexpected catalytic activity versus pH profile, in that the optimal pH is 6.3, similar to that of the proteolytically cleaved form (6.5). Moreover, replacement of the ferrous ion by zinc increases the turnover number more than 10-fold; the pK(es) values are also shifted as expected for the change in the divalent metal ion. Although the EPR spectra of both single polypeptide and proteolytically cleaved Fe(3+)Zn(2+)-recHPAP are independent of pH over the range 4.5-6.2, the visible spectrum of Fe(3+)Zn(2+)-recHPAP is pH dependent. These results suggest that the properties and environment of the divalent metal are important in determining the catalytic properties of mammalian PAPs, and in particular that a solvent molecule coordinated to the divalent metal ion may play a critical role in the catalytic cycle of these enzymes.  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