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.
Extracts of Pseudomonas B4 grown with l-β-lysine (3,6-diaminohexanoate) as the main energy source are shown to contain a 3-keto-6-acetamidohexanoate cleavage enzyme that converts 3-keto-6-acetamidohexanoate and acetyl · CoA reversibly to 4-acetamidobutyryl · CoA and acetoacetate. The enzyme catalyzes the third step in β-lysine degradation. In unfractionated extracts cleavage enzyme activity is generally assayed spectrophotometrically by coupling the forward reaction with excess 4-acetamidobutyryl · CoA thiolesterase, derived from the same organism, and measuring the rate of CoASH formation by reaction with 5,5-dithiobis(2-nitrobenzoic acid). Enzyme freed of thiolesterase is conveniently assayed by using 4-acetamidobutyryl · CoA and acetoacetate as substrates and measuring acetyl · CoA formation by means of citrate synthase reaction in the presence of 5,5-dithiobis(2-nitrobenzoic acid). The cleavage enzyme has been purified 38-fold to a specific activity of 237 mU/mg. The stoichiometry, equilibrium constant, molecular weight, and various kinetic properties of the enzymatic reaction have been determined. The substrate specificity of the Pseudomonas enzyme differs markedly from that of the analogous 3-keto-5-aminohexanoate cleavage enzyme of Clostridium subterminale strain SB4 and is broader. In the forward reaction 3-ketohexanoate can replace 3-keto-6-acetamidohexanoate, and propionyl · CoA can replace acetyl · CoA as a substrate. In the backward reaction, 4-acetamidobutyryl · CoA can be replaced by any of several CoA thiolesters including the butyryl, valeryl, 4-propionamidobutyryl, 3-acetamidopropionyl, and β-alanyl derivatives, and acetoacetate can be replaced by 2-methylacetoacetate. The products of these reactions have been characterized. Unlike the cleavage enzyme of Clostridium subterminale strain SB4, the Pseudomonas enzyme is not stimulated by Co2+ or Mn2+ and is not inhibited by EDTA, 5,5-dithiobis(2-nitrobenzoic acid), or p-chloromercuribenzoate. Tracer experiments indicate that carbon atoms 1 and 2 of acetoacetate are derived from carbon atoms 1 and 2 of 3-keto-6-acetamidohexanoate, and carbon atoms 3 and 4 of acetoacetate are derived from the acetyl group of acetyl · CoA. The cleavage enzyme is not formed in detectable amounts when Pseudomonas B4 is grown in a peptone-yeast extract medium.  相似文献   

2.
《Plant science》2001,160(2):229-236
The yellow coloration of snapdragon (Antirrhinum majus) flowers is mainly provided by the 6-glucosides of aureusidin and bracteatin. However, the biochemical mechanism of aurone biosynthesis is not well understood. In this study, we have identified aurone-biosynthesizing activity in the extracts of yellow snapdragon flowers. Incubation of 2′,4′,6′,4-tetrahydroxychalcone (THC) with an enzyme preparation in the presence of H2O2 caused the enzymatic formation of a single product, aureusidin, without the formation of a previously proposed 2-(α-hydroxybenzyl)coumaranone intermediate. The formation of aureusidin from THC was specifically observed with yellow flowers as well as aurone-accumulating flowers of other colors. The pH optimum for the enzymatic formation of aureusidin was around 5.4. Stoichiometric studies showed that one mole of aureusidin formation was accompanied by the consumption of one mole of oxygen with no detectable consumption of H2O2, which may work as an enzyme activator. The oxidative formation of aureusidin from THC could be explained in terms of the action of a single enzyme, an internal monooxygenase catalyzing the 3-hydroxylation and oxidative cyclization of THC. Incubation of 2′,4′,6′,3,4-pentahydroxychalcone (PHC) with an enzyme yielded both aureusidin and bracteatin at an approximate molar ratio of 6:1. In this case, H2O2 was not required for enzyme activity but rather inhibited the reaction. The 4′-glucosides of THC and PHC could also act as substrates for the formation of the 6-glucosides of aurones. These results suggest that aureusidin can be produced from either THC or PHC, whereas bracteatin is not produced through the 5′-hydroxylation of aureusidin but arise solely from PHC.  相似文献   

3.
《Analytical biochemistry》1987,165(2):341-348
A flow microcalorimetric method was developed for the analysis of enzymatic activities in crude tissue homogenates. It can be applied whenever a heat exchange is involved in an enzymatic reaction. The consequent sensitivity obviously depends on the enthalpy variation observed. Dihydrofolate reductase was chosen as an example; this enzyme is the molecular target of methotrexate, a widely used anticancer agent. This calorimetric method, whose sensitivity limit is 1.48 × 10−4 units of dihydrofolate reductase per milliliter of reactant medium, allows enzyme activity measurements in tissues with low dihydrofolate reductase levels. A few examples of measurements in animal tissues are given. These measurements are of some interest; indeed, increased activity and increased levels of this enzyme are two of the mechanisms which may explain resistance to methotrexate.  相似文献   

4.
Dihydrofolate reductase from chicken liver has a single sulfhydryl group which reacts stoichiometrically and specifically with a wide variety of organic mercury compounds to yield an enzyme derivative which exhibits up to 10-fold the activity of the unmodified form when measured at pH 6.5, the optimum for the modified enzyme. The sulfhydryl group is apparently not at the active site since a 25-fold excess of either major cosubstrate, dihydrofolate or TPNH, affects neither the rate nor extent of the modification reaction. The reaction is essentially instantaneous and yields an enzyme with altered kinetic properties for all the substrate pairs examined (TPNH/dihydrofolate, TPNH/ folate, and DPNH/dihydrofolate) when tested near their pH optima. V values increased 3- to 10-fold when TPNH was cofactor; Km values increased 10- to 15-fold for the TPNH/dihydrofolate pair. The mercurial-activated enzyme, unlike the native form, exhibits a markedly increased sensitivity to heat, proteolysis, and the ionic environment, losing approximately 50% of its activity under conditions where there is no loss of activity in the native form. However, substrates can afford protection, the order of effectiveness being identical with the relative affinities of the substrates for the native enzyme (Subramanian, S., and Kaufman, B. T. (1978) Proc. Nat. Acad. Sci. USA75, 3201). Thus, dihydrofolate, with the largest binding constant is the most efficient, protecting completely against trypsin digestion when present at a 1:1 ratio with enzyme. Heating the mercury enzyme in the absence of substrates gives rise to a stable but altered conformation characterized by a time course which shows marked hysteresis. The striking similarity of the properties of the mercurial-activated dihydrofolate reductase to the reductase activated by 4 m urea, a reagent known to affect the tertiary structure of proteins, suggests that covalent binding of organic mercurials to the sulfhydryl group results in a similar conformational change characterized by a marked facilitation of the dihydrofolate reductase reaction.  相似文献   

5.
There is rising interest in non-enzymatic cholesterol oxidation because the resulting oxysterols have biological activity and can be used as non-invasive markers of oxidative stress in vivo. The preferential site of oxidation of cholesterol by highly reactive species is at C7 having a relatively weak carbon–hydrogen bond. Cholesterol autoxidation is known to proceed via two distinct pathways, a free radical pathway driven by a chain reaction mechanism (type I autoxidation) and a non-free radical pathway (type II autoxidation). Oxysterols arising from type II autoxidation of cholesterol have no enzymatic correlates, and singlet oxygen (1ΔgO2) and ozone (O3) are the non-radical molecules involved in the mechanism. Four primary derivatives are possible in the reaction of cholesterol with singlet oxygen via ene addition and the formation of 5α-, 5β-, 6α- and 6β-hydroxycholesterol preceded by their respective hydroperoxyde intermediates. The reaction of ozone with cholesterol is very fast and gives rise to a complex array of oxysterols. The site of the initial ozone reaction is at the Δ5,6 –double bond and yields 1,2,3-trioxolane, a compound that rapidly decomposes into a series of unstable intermediates and end products. The downstream product 3β-hydroxy-5-oxo-5,6-secocholestan-6-al (sec-A, also called 5,6-secosterol), resulting from cleavage of the B ring, and its aldolization product (sec-B) have been proposed as a specific marker of ozone-associated tissue damage and ozone production in vivo. The relevance of specific ozone-modified cholesterol products is, however, hampered by the fact sec-A and sec-B can also arise from singlet oxygen via Hock cleavage of 5α-hydroperoxycholesterol or via a dioxietane intermediate. Whatever the mechanism may be, sec-A and sec-B have no enzymatic route of production in vivo and are reportedly bioactive, rendering them attractive biomarkers to elucidate oxidative stress-associated pathophysiological pathways and to develop pharmacological agents.  相似文献   

6.
Malic enzyme (S)-malate: NADP+ oxidoreductase (oxaloacetate-decar☐ylating, EC 1.1.1.40) purified from the thermoacidophilic archaebacterium Sulfolobus solfataricus, strain MT-4, catalyzed the metal-dependent decar☐ylation of oxaloacetate at optimum pH 7.6 at a rate comparable to the decar☐ylation of l-malate. The oxaloacetate decar☐ylase activity was stimulated about 50% by NADP but only in the presence of MgCl2, and was strongly inhibited by l-malate and NADPH which abolished the NADP activation. In the presence of MnCl2 and in the absence of NADP, the Michaelis constant and Vm for oxaloacetate were 1.7 mM and 2.3 μmol·min−1·mg−1, respectively. When MgCl2 replaced MnCl2, the kinetic parameters for oxaloacetate remained substantially unvaried, whereas the Km and Vm values for l-malate have been found to vary depending on the metal ion. The enzyme carried out the reverse reaction (malate synthesis) at about 70% of the forward reaction, at pH 7.2 and in the presence of relatively high concentrations of bicarbonate and pyruvate. Sulfhydryl residues (three cysteine residues per subunit) have been shown to be essential for the enzymatic activity of the Sulfolobus solfataricus malic enzyme. 5,5′-Dithiobis(2-nitrobenzoic acid), p-hydroxymercuribenzoate and N-ethylmaleimide caused the inactivation of the oxidative decar☐ylase activity, but at different rates. The inactivation of the overall activity by p-hydroxymercuribenzoate was partially prevented by NADP singly or in combination with both l-malate and MnCl2, and strongly enhanced by the car☐ylic acid substrates; NADP + malate + MnCl2 afforded total protection. The inactivation of the oxaloacetate decar☐ylase activity by p-hydroxymercuribenzoate treatment was found to occur at a slower rate than that of the oxidative decar☐ylase activity.  相似文献   

7.
The catalytically active form (Ea) of pyruvate formate-lyase in Escherichia coli cells is generated from an inactive form of the enzyme (Ei) through a post-translational process that requires a distinct activating enzyme and is linked to the cleavage of adenosylmethionine to methionine and 5′-deoxyadenosine. Ei and the activating enzyme were purified to homogeneity and structurally characterized. Ei has an α2 oligomeric structure (2 × 85 kDa) and contains no cofactor. The amino acid composition has been determined. Out of a total of six cysteinyl residues per subunit, one shows an unusually fast reaction with iodoacetate (k2 = 7 (m? s?) at pH 6.8, 30 °C), which is accompanied by loss of the activatability of the enzyme. The 1500-fold purified activating enzyme is a monomeric protein of 30 kDa. It contains a covalently bound, as yet unidentified chromophoric factor which has an optical absorption peak at 388 nm. Further studies of the in situ state of pyruvate formate-lyase detected a reversible backconversion of the active form Ea into Ei when anaerobic cells become nutrient-depleted.  相似文献   

8.
Cathechol oxidase extracted from tea leaves was purified over 200-fold, using isoelectric focusing. The purified catechol oxidase was free of peroxidase and flavanol gallate esterase activities. Further, this enzyme was shown to have optimum activity near pH 5·7 and a Km of 2·3 × 10?3 M (at 25°) for (?)-epigallocatechin gallate. The purified enzyme was found to be capable of epimerizing tea flavanols at their C-2 position whether oxidation of the flavanol occurs (aerobic conditions) or not (anaerobic conditions). When oxygen is present, gallic acid is formed as a result of oxidation of either (?)-epigallocatechin gallate or (?)-epicatechin gallate. Formation of gallic acid is a side reaction of the oxidation of the flavanol gallates and is named oxidative degallation; no esterase per se is involved in this reaction.  相似文献   

9.
The latex from Vasconcellea quercifolia (“oak leaved papaya”), a member of the Caricaceae family, contains at least seven cysteine endopeptidases with high proteolytic activity, which helps to protect these plants against injury. In this study, we isolated and characterized the most basic of these cysteine endopeptidases, named VQ-VII. This new purified enzyme was homogeneous by bidimensional electrophoresis and MALDI-TOF mass spectrometry, and exhibited a molecular mass of 23,984 Da and an isoelectric point >11. The enzymatic activity of VQ-VII was completely inhibited by E-64 and iodoacetic acid, confirming that it belongs to the catalytic group of cysteine endopeptidases. By investigating the cleavage of the oxidized insulin B-chain to establish the hydrolytic specificity of VQ-VII, we found 13 cleavage sites on the substrate, revealing that it is a broad-specificity peptidase. The pH profiles toward p-Glu-Phe-Leu-p-nitroanilide (PFLNA) and casein showed that the optimum pH is about 6.8 for both substrates, and that in casein, it is active over a wide pH range (activity higher than 80 % between pH 6 and 9.5). Kinetic enzymatic assays were performed with the thiol peptidase substrate PFLNA (K m = 0.454 ± 0.046 mM, k cat = 1.57 ± 0.07 s?1, k cat/K m = 3.46 × 103 ± 14 s?1 M?1). The N-terminal sequence (21 amino acids) of VQ-VII showed an identity >70 % with 11 plant cysteine peptidases and the presence of highly conserved residues and motifs shared with the “papain-like” family of peptidases. VQ-VII proved to be a new latex enzyme of broad specificity, which can degrade extensively proteins of different nature in a wide pH range.  相似文献   

10.
Dihydrofolate reductase (EC 1.5.1.3), purified to homogeneity from an amethopterin-resistant subline (R6) of cultured L1210 murine leukemia cells, has been used to study enzyme-substrate and enzyme-inhibitor complexes. NADPH, NADP+acid-modified NADPH (λmax at 265 nm, elevated absorbance at 290 nm), 2′-phosphoadenosine-5′-diphosphate ribose, dihydrofolate, and amethopterin formed binary complexes with the enzyme. Ternary complexes could be formed by admixing the enzyme with: (a) NADPH and amethopterin; (b) NADP+ and tetahydrofolate; and (c) acid-modified NADPH and dihydrofolate. All of these complexes migrated as stable well-defined bands on polyacrylamide gel electrophoresis at pH 8.3. The bands could be visualized by staining both for enzyme activity and for protein. These binary and ternary complexes were also stable to extensive dialysis. Spectra of the dialyzed enzyme complexes indicated that each ligand was present at an equimolar ratio with the enzyme.  相似文献   

11.
Induction of endonucleolytic DNA cleavage is an essential event that links the initiating stimuli to the final effects of cells. The cleavage efficiency and thus the final yield could be affected by many factors, including structures of DNA substrates, composite structures of enzymes–substrates or enzymes–nucleic analogs and so on. However, it is not clear whether a nucleotide derivative-substituted in DNA substrates can influence the efficiency of enzymatic cleavage. To investigate the effect of sugar pucker conformation on DNA–protein interactions, we used 2′-O-methyl modified nucleotides (OMeN) to modify DNA substrates of isocaudemers BamHI and BglII in this study, and used FRET assay as an efficient method for analysis of enzyme cleavage. Experimental results demonstrated that OMeN-substituted recognition sequences influenced the cleavage rates significantly in a position-dependent manner. OMeN substitutions can reduce the cleavage as expected. Surprisingly, OMeN substitutions can also enhance the cleavage rates. The kinetics parameters of V max and K m have been obtained by fitting the Michaelis-Menten kinetic equation. These 2'-OMe nucleotides could behave as a regulatory element to modulate the enzymatic activity in vitro, and this property could enrich our understanding about the endonuclease cleavage mechanism and enhance our ability to regulate the enzymatic cleavage efficiency for applications in synthetic biology.  相似文献   

12.
l-allo-Threonine aldolase (l-allo-threonine acetaldehyde-lyase), which exhibited specificity for l-allo-threonine but not for l-threonine, was purified from a cell-free extract of Aeromonas jandaei DK-39. The purified enzyme catalyzed the aldol cleavage reaction of l-allo-threonine (Km=1.45 mM, Vmax=45.2 μmol min−1 mg−1). The activity of the enzyme was inhibited by carbonyl reagents, which suggests that pyridoxal-5′-phosphate participates in the enzymatic reaction. The enzyme does not act on either l-serine or l-threonine, and thus it can be distinguished from serine hydroxy-methyltransferase (l-serine:tetrahydrofolate 5,10-hydroxy-methyltransferase, EC 2.1.2.1) or l-threonine aldolase (EC 4.1.2.5).  相似文献   

13.
Over 7000 microorganisms were screened to find an enzyme source for the hydrolysis of a C4 methyl ester blocking group on 7-aminodesacetoxycephalosporanic acid (7-ADCA). Only one culture, Streptomyces capillispira Mertz and Higgens nov. sp., produced an enzyme that catalysed the reaction. Enzyme synthesis in a defined mineral salts medium was repressed by NH3 and amino acids. Under optimum fermentation conditions, the maximum rate of substrate hydrolysis was 6 × 10?10 mol min?1 mg?1 cell. The enzyme was recovered from the mycelia and partially purified by gel filtration. Kinetic studies by pH-stat titration indicated that the pH optimum was 7.5–8.5, the temperature optimum was 25–30°C, and the substrate Km value was 2.3 mg ml?1. The reaction products, 7-ADCA and methanol, were weak competitive inhibitors of the enzyme with K1 values of 6.63 and 0.188 mg ml?1, respectively. The enzyme also hydrolysed cefaclor and cephalexin methyl esters but did not hydrolyse cephalosporin ethyl esters. With further improvements in enzyme yields and stability, enzymatic deblocking of cephalosporins could provide an alternative to chemical deblocking processes.  相似文献   

14.
The formal first step in in vitamin A metabolism is the conversion of its natural precursor β,β-carotene (C40) to retinaldehyde (C20). This reaction is catalyzed by the enzyme β,β-carotene-15,15′-monooxygenase (BCMO1). BCMO1 has been cloned from several vertebrate species, including humans. However, knowledge about this protein’s enzymatic and structural properties is scant. Here we expressed human BCMO1 in Spodoptera frugiperda 9 insect cells. Recombinant BCMO1 is a soluble protein that displayed Michaelis–Menten kinetics with a KM of 14 μM for β,β-carotene. Though addition of detergents failed to increase BCMO1 enzymatic activity, short chain aliphatic detergents such as C8E4 and C8E6 decreased enzymatic activity probably by interacting with the substrate binding site. Thus we purified BCMO1 in the absence of detergent. Purified BCMO1 was a monomeric enzymatically active soluble protein that did not require cofactors and displayed a turnover rate of about 8 molecules of β,β-carotene per second. The aqueous solubility of BCMO1 was confirmed in mouse liver and mammalian cells. Establishment of a protocol that yields highly active homogenous BCMO1 is an important step towards clarifying the lipophilic substrate interaction, reaction mechanism and structure of this vitamin A forming enzyme.  相似文献   

15.
A double affinity-labelling approach has been developed in order to convert an oligomeric enzyme with multiple active centres into a single-site enzyme.Tryptophanyl-transfer RNA synthetase (EC 6.1.1.2) from beef pancreas is a symmetric dimer, α2 An ATP analogue, γ-(p-azidoanilide)-ATP does not serve as a substrate for enzymatic aminoacylation of tRNATrp but acts as an effective competitive inhibitor in the absence of photochemical reaction, with K1 = 1 × 10?3m (Kmfor ATP = 2 × 10?4m). The covalent photoaddition of azido-ATP3 results in complete loss of enzymatic activity in both the ATP-[32P]pyrophosphate exchange reaction and tRNA aminoacylation. ATP completely protects the enzyme against inactivation. However, covalent binding of azido-ATP is also observed outside the active centres. The difference between covalent binding of the azido-ATP in the absence and presence of ATP corresponds to 2 moles of the ATP analogue per mole of the enzyme.Two binding sites for tRNATrp have been found from complex formation at pH 5.8 in the presence of Mg2+. The two tRNA molecules bind, with Kdis = 3.6 × 10?8m and Kdis = 0.9 × 10?6m, respectively, pointing to a strong negative co-operativity between the binding sites for tRNA.N-chlorambucilyl-tryptophanyl-tRNATrp and TRSase form a complex with Kdis = 5.5 × 10?8m at pH 5.8 in the presence of 10 mm-Mg2+. This value is similar to the value of Kdis for tryptophanyl-tRNA of 4.8 × 10?8m. Under the same conditions a 1:1 complex (in mol) is formed between the enzyme and Trp-tRNA or N-chlorambucilyl-Trp-tRNA. On incubation, a covalent bond is formed between N-chlorambucilyl-Trp-tRNA and TRSase; 1 mole of affinity reagent alkylates 1 mole of enzyme independently of the concentration of the modifier. The alkylation reaction is completely inhibited by the presence of tRNATrp whereas the tRNA devoid of tRNATrp does not affect the rate of alkylation. In the presence of either ATP or tryptophan, or a mixture of the two, the alkylation reaction is inhibited even though these ligands have no effect on the complex formation between TRSase and the tRNA analogue. Photoaddition of the azido-ATP completely prevents the reaction of the enzyme with the tRNA analogue, although the non-covalent complex formation is not affected.Exhaustive alkylation of TRSase partially inhibits the reaction of ATP [32P]pyrophosphate exchange and completely blocks the aminoacylation of tRNATrp. Cleavage of the tRNA which is covalently bound to TRSase restores both the ATP-[32P]pyrophosphate exchange and aminoacylation activity.The TRSase which is covalently-bound to R-Trp-tRNA is able to incorporate only one ATP molecule per dimeric enzyme into the active centre. This doubly modified enzyme is completely enzymatically inactive. Removal of the tRNA residue from the doubly modified enzyme results in the formation of the derivative with one blocked ATP site. Therefore, a “single-site” TRSase may be generated either by alkylation of the enzyme with Cl-R-Trp-tRNA or after the removal of covalently bound tRNA from the doubly labelled protein.Tryptophanyl-tRNA synthetase containing blocked ATP and/or tRNA binding site(s) seems to bo a useful tool for investigation of negative co-operativity and may help in the elucidation of the structure function relationships between the active centres.  相似文献   

16.
2-Keto-3-deoxygluconate-6P (KDPG) aldolase ofPseudomonas putida mediates the cleavage of, as well as the condensation of, pyruvate andd-glyceraldehyde-3P (GaP) yielding, 2-keto-3-deoxygalactonate-6P (KDPGal) as side reactions of normal catalysis. These are visualized at high levels of aldolase. KDPGal cleavage occurs with aV max that is 1/5000 that for KDPG cleavage. TheKm for KDPGal is 0.2 mM, with aK i of 0.85 mM. The E-KDPGal complex is reductively inactivated having aKd of 0.55 mM. TheV/K value for KDPG cleavage is 2.0×108 sec?1, while the value for KDPGal cleavage is 1220 sec?1. The difference in first-order rate constants of 164,000-fold argues that a step in the cleavage of KDPGal mediated by the enzyme is uncatalyzed. The enzyme is reductively inactivated by trapping the E-pyruvate, E-KDPG, or E-KDPGal complex. The enzyme can also be inactivated by reductive trapping of a catalytically nonproductive E-glyceraldehyde-3P complex. This latter occurs with aKd for GaP of 20 mM and a rate constant equivalent to a limiting half-time of 1110 sec at 1 mM cyanoborohydride. Reductive inactivation half-times in the presence of high GaP/KDPG ratios were the sum of both E-GaP and E-KDPG trapping by cyanoborohydride so that the inactivation rate due to KDPG could be determined. It was found at 1 mM cyanoborohydride that the limiting half-time for the E-KDPG complex was 2382 sec. The corresponding value for the E-KDPGal complex was 215 sec. Consequently, the E-KDPGal complex is 11 times more sensitive to reductive derivativation than is the E-KDPG complex. This is interpreted to show that the enzyme binds the KDPGal in a “normal” step forming a ketimine. However, turnover to the eneamine with resultant C-C bond cleavage is uncatalyzed. For the case of KDPGal synthesized by KDPG aldolase, it is argued that the pyruvate eneamine is bound to the active site, which can be attacked by GaP with its aldehyde carbon in the catalytically nonproductive conformation as a side reaction, presumably forming a tertiary complex. Spontaneous protonation of the resultant alcoholate anion would generate KDPGal. The data are interpreted to support an argument that catalytic proton turnover at the OH of C-4 of KDPG is required for normal catalysis, and that this step, which catalytically interconverts ketimine/eneamine, imposes steric constraints controlling the overall stereochemistry of the reaction.  相似文献   

17.
The uptake and transformation of phenylmercuric acetate (PMA), ethylmercury chloride (EMC), and methylmercury chloride (MMC) by rat liver, kidney, and brain slices were investigated. A slightly greater enzymatic activity for the cleavage of PMA and EMC was observed in the kidney than in the liver; brain tissue had the least activity.The average rates of C-Hg cleavage were 2.4 to 3.0 nmole g?1 h?1 for PMA and 0.9 to 1.0 nmole g?1 h?1 for EMC. No cleavage of C-Hg bond from MMC was observed by the kidney and liver.The cleavage of PMA was an enzymatic process and the enzyme was located in the nuclear fraction of both liver and kidney. The apparent Michaelis constant and the maximal velocity value for PMA conversion by rat liver slices were 625 μM and 19.2 nmole per g tissue per 2 h, respectively.It is postulated that the PMA molecule is first split at the C-Hg bond to yield benzene and mercuric ion.The cleavage of the EMC molecule appears to be more complex.  相似文献   

18.
The respiratory chain-linked external NADH dehydrogenase has been isolated from Candida utilis in highly purified form. The enzyme is soluble and has a molecular weight of approx. 1.5 · 106. The enzyme contains two moles of FMN per mole of enzyme and is composed of two large subunits of mol. wt. 270 000 and eight smaller subunits of mol. wt. 135 000. Iron and copper are present in the preparations, but appear to be contaminants. The enzyme catalyzes the oxidation of NADH and NADPH at nearly equal rates and reacts readily with 2,6-dichlorophenolindophenol, CoQ6 and CoQ1 derivatives as acceptors. Rotenone (10?5 M) and seconal (10?3 M) do not inhibit enzymatic activity.  相似文献   

19.
An extracellular protein with strong absorption at 406 nm was purified from cell-free culture fluid of latex-grown Xanthomonas sp. strain 35Y. This protein was identical to the gene product of a recently characterized gene cloned from Xanthomonas sp., as revealed by determination of m/z values and sequencing of selected isolated peptides obtained after trypsin fingerprint analysis. The purified protein degraded both natural rubber latex and chemosynthetic poly(cis-1,4-isoprene) in vitro by oxidative cleavage of the double bonds of poly(cis-1,4-isoprene). 12-Oxo-4,8-dimethyltrideca-4,8-diene-1-al (m/z 236) was identified and unequivocally characterized as the major cleavage product, and there was a homologous series of minor metabolites that differed from the major degradation product only in the number of repetitive isoprene units between terminal functions, CHO-CH2— and —CH2-COCH3. An in vitro enzyme assay for oxidative rubber degradation was developed based on high-performance liquid chromatography analysis and spectroscopic detection of product carbonyl functions after derivatization with dinitrophenylhydrazone. Enzymatic cleavage of rubber by the purified protein was strictly dependent on the presence of oxygen; it did not require addition of any soluble cofactors or metal ions and was optimal around pH 7.0 at 40°C. Carbon monoxide and cyanide inhibited the reaction; addition of catalase had no effect, and peroxidase activity could not be detected. The purified protein was specific for natural rubber latex and chemosynthetic poly(cis-1,4-isoprene). Analysis of the amino acid sequence deduced from the cloned gene (roxA [rubber oxygenase]) revealed the presence of two heme-binding motifs (CXXCH) for covalent attachment of heme to the protein. Spectroscopic analysis confirmed the presence of heme, and approximately 2 mol of heme per mol of RoxA was found.  相似文献   

20.
Small-angle neutron scattering studies of Escherichia coli tyrosyl-tRNA synthetase indicate that in solution this enzyme is a dimer of Mr, 91 (±6) × 103 with a radius of gyration RG of 37.8 ± 1.1 Å.The increase in the scattering mass of the enzyme upon binding tRNATyr has been followed in 20 mm-imidazole · HCl (pH 7.6), 10 mm-MgCl2, 0.1 mm-EDTA, 10 mm-2-mercaptoethanol, 150 mm-KCl. A stoichiometry of one bound tRNA per dimeric enzyme molecule was found. The RG of the complex is equal to 41 ± 1 Å. Titration experiments in 74% 2H2O, close to the matching point of tRNA, show an RG of 38.5 ± 1 Å for the enzyme moiety in the complex. From these values, a minimum distance of 49 Å between the centre of mass of the bound tRNA and that of the enzyme was calculated.In low ionic strength conditions (20 mm-imidazole-HCl (pH 7.6), 10 mm-MgCl2, 0.1 mm-EDTA, 10 mm-2-mercaptoethanol) and at limiting tRNA concentrations with respect to the enzyme, titrations of the enzyme by tRNATyr are characterized by the appearance of aggregates, with a maximum scattered intensity at a stoichiometry of one tRNA per two enzyme molecules. At this point, the measured Mr and RG values are compatible with a compact 1:2, tRNA: enzyme complex. This complex forms with a remarkably high stability constant: (enzyme:tRNA:enzyme)/(enzyme:tRNA)(enzyme) of 0.1 to 0.3(× 106) m?1 (at 20 °C). Upon addition of more tRNA, the complex dissociates in favour of the 1:1, enzyme:tRNA complex, which has a higher stability constant (1 to 3 (× 106) m?1).  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