首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The degradation of agmatine to succinate by Klebsiella aerogenes occurs in five steps. The enzyme catalyzing the first step, agmatinase, is induced by agmatine. The enzymes catalyzing the second and third steps, putrescine aminotransferase and 4-aminobutyraldehyde dehydrogenase, are induced by putrescine and also by their product, 4-aminobutyrate. The enzymes catalyzing the fourth and fifth steps, 4-aminobutyrate aminotransferase and succinate semialdehyde dehydrogenase, are induced by 4-aminobutyrate. This compound also serves as gratuitous inducer of the catabolic acetylornithine aminotransferase. The formation of the enzymes responsible for agmatine degradation is regulated not only by induction, but also by catabolite repression and activation by glutamine synthetase.  相似文献   

2.
The effects of inhibitors of diamine oxidase (EC 1.4.3.6), monoamine oxidase (EC 1.4.3.4) and 4-aminobutyrate aminotransferase (EC 2.6.1.19) on the catabolism of putrescine in mice in vivo were studied. Diamine oxidase inhibitors and carboxymethoxylamine (amino-oxyacetate) markedly inhibit the metabolism of [(14)C]putrescine to (14)CO(2), but affect different enzymes. Aminoguanidine specifically inhibits the mitochondrial and non-mitochondrial diamine oxidases, whereas carboxymethoxylamine specifically inhibits 4-aminobutyrate transamination by the mitochondrial pathway. Hydrazine inhibits at both sites, and results in increased concentrations of 4-aminobutyrate in brain and liver. Pretreatment of mice with carboxymethoxylamine and [(14)C]putrescine leads to the urinary excretion of amino[(14)C]butyrate. Carboxymethoxylamine does not affect the non-mitochondrial pathway of putrescine catabolism, as the product of oxidative deamination of putrescine in the extramitochondrial compartment is not further oxidized but is excreted in the urine as derivatives of 4-aminobutyraldehyde. Another catabolic pathway of putrescine involves monoamine oxidase, and the monoamine oxidase inhibitor, pargyline, decreases the metabolism of [(14)C]putrescine to (14)CO(2)in vivo. Catabolism of putrescine to CO(2)in vivo occurs along different pathways, both of which have 4-aminobutyrate as a common intermediate, in contrast with the non-mitochondrial catabolism of putrescine, which terminates in the excretion of 4-aminobutyraldehyde derivatives. The significance of the different pathways is discussed.  相似文献   

3.
Maria Chrysina  Vasili Petrouleas 《BBA》2010,1797(4):487-493
The oxygen evolving complex of Photosystem II undergoes four light-induced oxidation transitions, S0-S1,…,S3-(S4)S0 during its catalytic cycle. The oxidizing equivalents are stored at a (Mn)4Ca cluster, the site of water oxidation. EPR spectroscopy has yielded valuable information on the S states. S2 shows a notable heterogeneity with two spectral forms; a g = 2 (S = 1/2) multiline, and a g = 4.1 (S = 5/2) signal. These oscillate in parallel during the period-four cycle. Cyanobacteria show only the multiline signal, but upon advancement to S3 they exhibit the same characteristic g = 10 (S = 3) absorption with plant preparations, implying that this latter signal results from the multiline configuration. The fate of the g = 4.1 conformation during advancement to S3 is accordingly unknown. We searched for light-induced transient changes in the EPR spectra at temperatures below and above the half-inhibition temperature for the S2 to S3 transition (ca 230 K). We observed that, above about 220 K the g = 4.1 signal converts to a multiline form prior to advancement to S3. We cannot exclude that the conversion results from visible-light excitation of the Mn cluster itself. The fact however, that the conversion coincides with the onset of the S2 to S3 transition, suggests that it is triggered by the charge-separation process, possibly the oxidation of tyr Z and the accompanying proton relocations. It therefore appears that a configuration of (Mn)4Ca with a low-spin ground state advances to S3.  相似文献   

4.
Pathway and enzyme redundancy in putrescine catabolism in Escherichia coli   总被引:2,自引:0,他引:2  
Putrescine as the sole carbon source requires a novel catabolic pathway with glutamylated intermediates. Nitrogen limitation does not induce genes of this glutamylated putrescine (GP) pathway but instead induces genes for a putrescine catabolic pathway that starts with a transaminase-dependent deamination. We determined pathway utilization with putrescine as the sole nitrogen source by examining mutants with defects in both pathways. Blocks in both the GP and transaminase pathways were required to prevent growth with putrescine as the sole nitrogen source. Genetic and biochemical analyses showed redundant enzymes for γ-aminobutyraldehyde dehydrogenase (PatD/YdcW and PuuC), γ-aminobutyrate transaminase (GabT and PuuE), and succinic semialdehyde dehydrogenase (GabD and PuuC). PuuC is a nonspecific aldehyde dehydrogenase that oxidizes all the aldehydes in putrescine catabolism. A puuP mutant failed to use putrescine as the nitrogen source, which implies one major transporter for putrescine as the sole nitrogen source. Analysis of regulation of the GP pathway shows induction by putrescine and not by a product of putrescine catabolism and shows that putrescine accumulates in puuA, puuB, and puuC mutants but not in any other mutant. We conclude that two independent sets of enzymes can completely degrade putrescine to succinate and that their relative importance depends on the environment.  相似文献   

5.
Hua Mei  Qi Chen  Qi Sun 《Inorganica chimica acta》2010,363(10):2265-5141
The synthesis and structure of a new 1-D molybdenum-arsenic compound based on the bi-capped Keggin anion [MoVI6MoV6O36(AsO4)(MoVO)2] have been reported, and its catalytic property has been examined. The title compound was characterized by IR, TG and X-ray diffraction analysis. Single crystal X-ray diffraction shows that it crystallizes in cubic crystal system, space group Pn-3m with cell dimension: a = 11.749(2) Å, V = 1622.0(5) Å3, Z = 2. Its structure has a 1-D infinite chain, in which the bi-capped Keggin anions are connected by sharing one terminal oxygen atom from the caps. The compound shows a moderate styrene conversion (48%), the major product for the oxidation of styrene is benzaldehyde (85.2%).  相似文献   

6.
Trypanosoma cruzi dihydroorotate dehydrogenase (TcDHODH) catalyzes the oxidation of l-dihydroorotate to orotate with concomitant reduction of fumarate to succinate in the de novo pyrimidine biosynthetic pathway. Based on the important need to characterize catalytic mechanism of TcDHODH, we have tailored a protocol to measure TcDHODH kinetic parameters based on isothermal titration calorimetry. Enzymatic assays lead to Michaelis-Menten curves that enable the Michaelis constant (KM) and maximum velocity (Vmax) for both of the TcDHODH substrates: dihydroorotate (KM = 8.6 ± 2.6 μM and Vmax = 4.1 ± 0.7 μM s-1) and fumarate (KM = 120 ± 9 μM and Vmax = 6.71 ± 0.15 μM s-1). TcDHODH activity was investigated using dimethyl sulfoxide (10%, v/v) and Triton X-100 (0.5%, v/v), which seem to facilitate the substrate binding process with a small decrease in KM. Arrhenius plot analysis allowed the determination of thermodynamic parameters of activation for substrates and gave some insights into the enzyme mechanism. Activation entropy was the main contributor to the Gibbs free energy in the formation of the transition state. A factor that might contribute to the unfavorable entropy is the hindered access of substrates to the TcDHODH active site where a loop at its entrance regulates the open-close channel for substrate access.  相似文献   

7.
Chemical implantation of Group 4 cations [Ti(III), Ti(IV), Zr(IV), Hf(IV)] has been carried out under mild conditions by the reaction of polycyclopentadienyl- (MCpn; M = Ti, n = 3, 4; M = Zr, Hf, n = 4), mixed cyclopentadienyl/N,N-dialkylcarbamato (MLx(O2CNEt2)y; M = Ti, L = Cp, C5Me5 (Cp*), x = 2, y = 1; M = Hf, L = Cp, x = 1, y = 3), and N,N-dialkylcarbamato (M(O2CNR2)n, M = Ti, n = 3, R = iPr; M = Ti, Hf, n = 4, R = Et; M = Zr, n = 4, R = iPr) derivatives, with the silanol groups of amorphous silica. Cyclopentadiene/pentamethylcyclopentadiene and/or carbon dioxide and the secondary amine are released in the process. The amount of implanted cations depends on the metal and on the ligands, the pentamethylcyclopentadienyl complex being less reactive than the unsubstituted congener. The starting complexes and the final products have been characterized by EPR or by 13C CP-MAS NMR spectroscopy.  相似文献   

8.
Three strains of new strictly anaerobic, Grampositive, non-sporeforming bacteria were isolated from various anoxic sediment samples with putrescine as sole carbon and energy source. Optimal growth in carbonate-buffered defined medium occurred at 37°C at pH 7.2–7.6. The DNA base ratio of strain NorPut1 was 29.6±1 mol% guanine plus cytosine. In addition to a surface layer and the peptidoglycan layer, the cell wall contained a second innermost layer with a periodic arrangement of subunits. All strains fermented putrescine to acetate, butyrate, and molecular hydrogen; the latter originated from both oxidative putrescine deamination and 4-aminobutyraldehyde oxidation. In defined mixed cultures with methanogens or homoacetogenic bacteria, methane or additional acetate were formed due to interspecies hydrogen transfer. Also 4-aminobutyrate and 4-hydroxybutyrate were fermented to acetate and butyrate, but no hydrogen was released from these substrates. No sugars, organic acids, other primary amines or amino acids were used as substrates. Neither sulfate, thiosulfate, sulfur, nitrate nor fumarate was reduced. Most of the enzymes involved in putrescine degradation could be demonstrated in cell-free extracts. A pathway of putrescine fermentation via 4-aminobutyrate and crotonyl-CoA with subsequent dismutation to acetate and butyrate is suggested.  相似文献   

9.
10.
Copper(I) cyanide reacts with various liquid amines and sulfides (L) under solvent-less conditions to form (CuCN)Ln, n = 0.5, 0.57, 0.75, 0.8, 1, 1.25, 1.5, 2. New X-ray structures are reported for L = Py (Py = pyridine, n = 0.57), 2-MePy (n = 1), 3-EtPy (n = 1.5), 2-ClPy (n = 1), 3-ClPy (n = 2), 3-MeOPy (n = 2), 4-tBuPy (n = 1.5), piperidine (n = 1.25), N-methylmorpholine (n = 1), N,N-dimethylcyclohexylamine (n = 1), 1-methylimidazole (n = 3), Me2S (n = 1), and tetrahydrothiophene (n = 1). The amine structures (except for the monomeric 1-methylimidazole complex) reveal 1D CuCN chains decorated with 0-2 L per metal atom. Chain structures observed include zigzag, helical and figure-8 helical. The CuCN-sulfide structures show sulfur-bridging of CuCN chains. In some cases (CuCN)L?1.5 species are transformed to (CuCN)L under vacuum. Thermal analysis shows facile release of ligand, yielding CuCN. Most of the (CuCN)Ln products are photoluminescent, emitting in the visible region. In some cases, coordination of very similar amines results in remarkably different emission spectra.  相似文献   

11.
Cystathionine γ-lyase (CGL) catalyzes the hydrolysis of l-cystathionine (l-Cth), producing l-cysteine (l-Cys), α-ketobutyrate and ammonia, in the second step of the reverse transsulfuration pathway, which converts l-homocysteine (l-Hcys) to l-Cys. Site-directed variants substituting residues E48 and E333 with alanine, aspartate and glutamine were characterized to probe the roles of these acidic residues, conserved in fungal and mammalian CGL sequences, in the active-site of CGL from Saccharomyces cerevisiae (yCGL). The pH optimum of variants containing the alanine or glutamine substitutions of E333 is increased by 0.4–1.2 pH units, likely due to repositioning of the cofactor and modification of the pKa of the pyridinium nitrogen. The pH profile of yCGL-E48A/E333A resembles that of Escherichia coli cystathionine β-lyase. The effect of substituting E48, E333 or both residues is the 1.3–3, 26–58 and 124–568-fold reduction, respectively, of the catalytic efficiency of l-Cth hydrolysis. The Kml-Cth of E333 substitution variants is increased ~ 17-fold, while Kml-OAS is within 2.5-fold of the wild-type enzyme, indicating that residue E333 interacts with the distal amine moiety of l-Cth, which is not present in the alternative substrate O-acetyl-l-serine. The catalytic efficiency of yCGL for α,γ-elimination of O-succinyl-l-homoserine (kcat/Kml-OSHS = 7 ± 2), which possesses a distal carboxylate, but lacks an amino group, is 300-fold lower than that of the physiological l-Cth substrate (kcat/Kml-Cth = 2100 ± 100) and 260-fold higher than that of l-Hcys (kcat/Kml-Hcys = 0.027 ± 0.005), which lacks both distal polar moieties. The results of this study suggest that the glutamate residue at position 333 is a determinant of specificity.  相似文献   

12.
It is postulated that elevated tissue concentrations of cortisol may be associated with the development of metabolic syndrome, obesity, and type 2 diabetes. The 11β-hydroxysteroid dehydrogenase type 1 (11β-HSD1) enzyme regenerates cortisol from inactive cortisone in tissues such as liver and adipose. To better understand the pivotal role of 11β-HSD1 in disease development, an in vivo microdialysis assay coupled with liquid chromatography/tandem mass spectrometry (LC/MS/MS) analysis using stable isotope-labeled (SIL) cortisone as a substrate was developed. This assay overcomes the limitations of existing methodologies that suffer from radioactivity exposure and analytical assay sensitivity and specificity concerns. Analyte extraction efficiencies (Ed) were evaluated by retrodialysis. The conversion of SIL-cortisone to SIL-cortisol in rhesus monkey adipose tissue was studied. Solutions containing 100, 500, and 1000 ng/mL SIL-cortisone were locally delivered through an implanted 30-mm microdialysis probe in adipose tissue. At the delivery rate of 1.0 and 0.5 μL/min, Ed values for SIL-cortisone were between 58.7 ± 5.6% (n = 4) and 72.7 ± 1.3% (n = 4), whereas at 0.3 μL/min Ed reached nearly 100%. The presence of 11β-HSD1 activities in adipose tissue was demonstrated by production of SIL-cortisol during SIL-cortisone infusion. This methodology could be applied to cortisol metabolism studies in tissues of other mammalian species.  相似文献   

13.
Designed retroaldolases have utilized a nucleophilic lysine to promote carbon–carbon bond cleavage of β-hydroxy-ketones via a covalent Schiff base intermediate. Previous computational designs have incorporated a water molecule to facilitate formation and breakdown of the carbinolamine intermediate to give the Schiff base and to function as a general acid/base. Here we investigate an alternative active-site design in which the catalytic water molecule was replaced by the side chain of a glutamic acid. Five out of seven designs expressed solubly and exhibited catalytic efficiencies similar to previously designed retroaldolases for the conversion of 4-hydroxy-4-(6-methoxy-2-naphthyl)-2-butanone to 6-methoxy-2-naphthaldehyde and acetone. After one round of site-directed saturation mutagenesis, improved variants of the two best designs, RA114 and RA117, exhibited among the highest kcat (> 10− 3 s− 1) and kcat/KM (11–25 M− 1 s− 1) values observed for retroaldolase designs prior to comprehensive directed evolution. In both cases, the > 105-fold rate accelerations that were achieved are within 1–3 orders of magnitude of the rate enhancements reported for the best catalysts for related reactions, including catalytic antibodies (kcat/kuncat = 106 to 108) and an extensively evolved computational design (kcat/kuncat > 107). The catalytic sites, revealed by X-ray structures of optimized versions of the two active designs, are in close agreement with the design models except for the catalytic lysine in RA114. We further improved the variants by computational remodeling of the loops and yeast display selection for reactivity of the catalytic lysine with a diketone probe, obtaining an additional order of magnitude enhancement in activity with both approaches.  相似文献   

14.
1H NMR spectroscopy was used to follow the cleavage of sucrose by invertase. The parameters of the enzyme's kinetics, Km and Vmax, were directly determined from progress curves at only one concentration of the substrate. For comparison with the classical Michaelis-Menten analysis, the reaction progress was also monitored at various initial concentrations of 3.5 to 41.8 mM. Using the Lambert W function the parameters Km and Vmax were fitted to obtain the experimental progress curve and resulted in Km = 28 mM and Vmax = 13 μM/s. The result is almost identical to an initial rate analysis that, however, costs much more time and experimental effort. The effect of product inhibition was also investigated. Furthermore, we analyzed a much more complex reaction, the conversion of farnesyl diphosphate into (+)-germacrene D by the enzyme germacrene D synthase, yielding Km = 379 μM and kcat = 0.04 s− 1. The reaction involves an amphiphilic substrate forming micelles and a water insoluble product; using proper controls, the conversion can well be analyzed by the progress curve approach using the Lambert W function.  相似文献   

15.
The cholesterol metabolism pathway in Mycobacterium tuberculosis (M. tb) is a potential source of energy as well as secondary metabolite production that is important for survival of M. tb in the host macrophage. Oxidation and isomerization of 3β-hydroxysterols to 4-en-3-ones is requisite for sterol metabolism and the reaction is catalyzed by 3β-hydroxysteroid dehydrogenase (Rv1106c). Three series of 6-azasteroids and 4-azasteroids were employed to define the substrate preferences of M. tb 3β-hydroxysteroid dehydrogenase. 6-Azasteroids with large, hydrophobic side chains at the C17 position are the most effective inhibitors. Substitutions at C1, C2, C4 and N6 were poorly tolerated. Our structure-activity studies indicate that the 6-aza version of cholesterol is the best and tightest binding competitive inhibitor (Ki = 100 nM) of the steroid substrate and are consistent with cholesterol being the preferred substrate of M. tb 3β-hydroxysteroid dehydrogenase.  相似文献   

16.
《Inorganica chimica acta》2005,358(4):1141-1150
The synthesis of new oxaaza macrocyclic ligands (2-4) derived from O1,O7-bis(2-formylphenyl)-1,4,7-trioxaheptane and functionalized tris(2-aminoethyl)amine are described. Mononuclear copper(II) complexes were isolated in the reaction of the corresponding macrocyclic ligand and copper(II) perchlorate. The structure of the [Cu(2)](ClO4)2 complex was determined by X-ray diffraction analysis. The copper(II) ion is five-coordinated by all N5 donor atoms, efficiently encapsulated by the amine terminal pendant-arm, with a trigonal-bipyramidal geometry. The complexes are further characterized by UV-Vis, IR and EPR studies. The electronic reflectance spectra evidence that the coordination geometry for the Cu(II) complexes is trigonal-bipyramidal with the ligands 1 and 2 or distorted square-pyramidal with the ligands 3 and 4. The electronic spectra in MeCN solutions are different from those in the solid state, which suggest that some structural modification may occur in solution. The EPR spectrum of powder samples of the copper complex with 2 presents axial symmetry with hyperfine split at g// with the copper nuclei (I = 3/2), which is characteristic of weakly exchange coupled extended systems. The EPR parameters (g// = 2.230, A// = 156 × 10−4 cm−1 and g = 2.085) indicate a dx2-y2 ground state. The EPR spectra of the complexes with ligands 3 and 4 show EPR spectra with a poorly resolved hyperfine structure at g//. In contrast, the complex with ligand 2 shows no hyperfine split and a line shape which was simulated assuming rhombic g-tensor (g1 = 2.030, g2 = 2.115 and g3 = 2.190).  相似文献   

17.
Aromatase (CYP19) is the target of several therapeutics used for breast cancer treatment and catalyzes the three-step conversion of androgens to estrogens, with an unusual C-C cleavage reaction in the third step. To better understand the CYP19 reaction, the oxy-ferrous complex of CYP19 with androstenedione substrate was cryotrapped, characterized by UV-vis spectroscopy, and cryoreduced to generate the next reaction cycle intermediate. EPR analysis revealed that the initial intermediate observed following cryoreduction is the unprotonated g1 = 2.254 peroxo-ferric intermediate, which is stable up to 180 K. Upon gradual cryoannealing, the low-spin (g1 = 2.39) product complex is formed, with no evidence for accumulation of the g1 = 2.30 hydroperoxo-ferric intermediate. The relative stabilization of the peroxo-ferric heme and the lack of observed hydroperoxo-ferric heme distinguish CYP19 from other P450s, suggesting that the proton delivery pathway is more hindered in CYP19 than in most other P450s.  相似文献   

18.
Two isomers of the N,O-coordinated acetylpyrrolyl complex [Ru(PPh3)2(CO)(NC4H3C(O)CH3)H] {cis-N,H (1) and trans-N,H (2)} have been prepared as models for catalytic intermediates in the Murai reaction. Complex 2 isomerises to 1 upon heating via a dissociative pathway (ΔH = 195 ± 41 kJ mol−1; ΔS = 232 ± 62 J mol−1 K−1); the mechanism of this process has been modeled using density functional calculations. Complex 2 displays moderate catalytic activity for the Murai coupling of 2′-methylacetophenone with trimethylvinylsilane, but 1 proved to be catalytically inactive under the same conditions.  相似文献   

19.
Ten novel adducts of the form AgClO4:PR3:L (1:1:1) (R = Ph, cy, o-tolyl; L = 2,2′-bipyridyl (‘bpy’), 2,2′-biquinoline (‘bq’), bis(2-pyridyl)amine (‘dpa’), bis(2-picolyl)amine (‘dpca’)) have been synthesized and characterized by analytical, spectroscopic (IR, far-IR, 1H and 31P NMR) and single crystal X-ray diffraction studies. The solid state molecular structures show that the complexes predominantly take the form [(R3P)AgL]+X, with a trigonal PAgN2 coordination environment, where the approach of the anion or the solvent may perturb the planarity of the silver environment. The ClO4 anion shows uni- or semi-bidentate coordination, except in the complexes AgClO4:PR3:dpca (1:1:1) (R = Ph and o-tolyl), where the anion remains uncoordinated and the dpca donor is a three-coordinate pincer-like ligand.  相似文献   

20.
Complexes of the general formula cis-[MX2(PTA)2] (M = Pd, Pt; X = Cl, Br, I; PTA = 1,3,5-triaza-7-phosphaadamantane) were used to study the catalytic intramolecular hydroamination/cyclization of 4-pentyn-1-amine into 2-methyl-pyrroline in water, methanol, and dimethyl sulfoxide (DMSO). Kinetic data were measured via 1H NMR under homogeneous conditions at 50 °C and showed the following trends in rate: (i) Fastest rates were observed in D2O. (ii) The Pd complexes of this study produced faster rates than the Pt complexes. (iii) The identity of the halide had no effect on the catalytic rate. Cyclization by the catalytic precursor cis-[PdCl2(PTA)2] (4) in D2O was zero-order in substrate and first-order in metal complex with ΔH = 20.0 ± 2.1 kcal/mol, ΔS = −7.4 ± 6.3 cal/mol K, and Ea = 20.6 ± 2.1 kcal/mol. The acetylide complex, trans-[Pt(CC(CH2)3NH2)2(PTA)2] (6) precipitated from a catalytic mixture involving cis-[PtBr2(PTA)2] (2). Spectroscopic and kinetic studies indicated that 6 and its cis analog, 7, were the predominant species in solution and that they were both active catalysts for the cyclization reaction. These data, in conjunction with the rate trends, indicated that the mechanism of the Pd(II) and Pt(II) catalyzed hydroamination of terminal alkynylamines in aqueous solution followed a unique mechanism with cyclization of an acetylenic-amine ligand being rate determining.  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