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 93 毫秒
1.
Copper amine oxidases (CuAO), from Escherichia coli (ECAO) and pea seedling (PSAO) were reacted with an excess of the hydrazine derivative 2-hydrazinopyridine (2HP) to form an initial, strongly absorbing adduct, (adduct 1; λmax 420–430 nm) formed by the covalent binding of 2HP with the active site cofactor 2,4,5-trihydroxyphenylalanine quinone (TPQ). Thermal incubation of buffered solutions of adduct 1 (pH 5.65–10.7) or addition of KOH solution (giving a final pH of 13–15) led isosbestically to a dramatic λmax shift yielding adduct 2 (λmax 520–530 nm). For both ECAO and PSAO, an increase in pH resulted in increased formation of adduct 2 with concomitant loss of adduct 1. Maximum adduct 2 formation occurred at pH 9.84 in ECAO and at pH 10.7 in PSAO. Beyond these pH levels, adduct 2 formation occurred to a much lesser extent which was independent of pH, suggesting enzyme denaturation. It is proposed that the conversion of adduct 1 to adduct 2 occurs as a result of hydrazone to azo conversion mediated by loss of a single proton, possibly to the active site base. It is further postulated that adduct formation and subsequent deprotonation can be likened to the substrate and product Schiff base complexes in the reductive half cycle of copper/TPQ containing amine oxidases. As part of this study an extinction coefficient at 280 nm was determined for ECAO by gravimetric analysis. This yielded a value of 2.1×105 M−1 cm−1 giving rise to the need of a correction factor when estimating the protein concentration from an absorbance reading at 280 nm. Using the estimated molecular mass of 160 kDa for the homodimeric ECAO, a correction factor of 0.76 must be applied.  相似文献   

2.
α-Aminoisobutyric acid is actively transported into yeast cells by the general amino acid transport system. The system exhibits a Km for α-aminoisobutyric acid of 270 μM, a Vmax of 24 nmol/min per mg cells (dry weight), and a pH optimum of 4.1–4.3. α-Aminoisobutyric acid is also transported by a minor system(s) with a Vmax of 1.7 nmol/min per mg cells. Transport occurs against a concentration gradient with the concentration ratio reaching over 1000:1 (in/out). The α-aminoisobutyric acid is not significantly metabolized or incorporated into protein after an 18 h incubation. α-Aminoisobutyric acid inhibits cell growth when a poor nitrogen source such as proline is provided but not with good nitrogen sources such as NH4+. During nitrogen starvation α-aminoisobutric acid strongly inhibits the synthesis of the nitrogen catabolite repression sensitive enzyme, asparaginase II. Studies with a mutant yeast strain (GDH-CR) suggest that α-aminoisobutyric acid inhibition of asparaginase II synthesis occurs because α-aminoisobutyric acid is an effective inhibitor of protein synthesis in nitrogen starved cells.  相似文献   

3.
New trypsin inhibitors Z-Lys-COCHO and Z-Lys-H have been synthesised. Ki values for Z-Lys-COCHO, Z-Lys-COOH, Z-Lys-H and Z-Arg-COOH have been determined. The glyoxal group (–COCHO) of Z-Lys-COCHO increases binding ~300 fold compared to Z-Lys-H. The α-carboxylate of Z-Lys-COOH has no significant effect on inhibitor binding. Z-Arg-COOH is shown to bind ~2 times more tightly than Z-Lys-COOH. Both Z-Lys-13COCHO and Z-Lys-CO13CHO have been synthesized. Using Z-Lys-13COCHO we have observed a signal at 107.4 ppm by 13C NMR which is assigned to a terahedral adduct formed between the hydroxyl group of the catalytic serine residue and the 13C-enriched keto-carbon of the inhibitor glyoxal group. Z-Lys-CO13CHO has been used to show that in this tetrahedral adduct the glyoxal aldehyde carbon is not hydrated and has a chemical shift of 205.3 ppm. Hemiketal stabilization is similar for trypsin, chymotrypsin and subtilisin Carlsberg. For trypsin hemiketal formation is optimal at pH 7.2 but decreases at pHs 5.0 and 10.3. The effective molarity of the active site serine hydroxyl group of trypsin is shown to be 25300 M. At pH 10.3 the free glyoxal inhibitor rapidly (t1/2=0.15 h) forms a Schiff base while at pH 7 Schiff base formation is much slower (t1/2=23 h). Subsequently a free enol species is formed which breaks down to form an alcohol product. These reactions are prevented in the presence of trypsin and when the inhibitor is bound to trypsin it undergoes an internal Cannizzaro reaction via a C2 to C1 alkyl shift producing an α-hydroxycarboxylic acid.  相似文献   

4.
Peter R. Rich  Derek S. Bendall 《BBA》1980,591(1):153-161
1. In fresh chloroplasts, three b-type cytochromes exist. These are b-559HP (λmax, 559 nm; Em at pH 7, +370 mV; pH-independent Em), b-559LP (λmax, 559 nm; Em at pH 7, +20 mV; pH-independent Em) and b-563 (λmax, 563 nm; Em at pH 7, ?110 mV; pH-independent Em). b-559HP may be converted to a lower potential form (λmax, 559 nm; Em at pH 7, +110 mV; pH-independent Em).2. In catalytically active b-f particle preparations, three cytochromes exist. These are cytochrome f (λmax, 554 nm; Em at pH 7, +375 mV, pK on oxidised cytochrome at pH 9), b-563 (λmax, 563 nm; Em at pH 7, ?90 mV, small pH-dependence of Em) and a b-559 species (λmax, 559 nm, Em at pH 7, +85 mV; pH-independent Em).3. A positive method of demonstration and estimation of b-559LP in fresh chloroplasts is described which involves the use of menadiol as a selective reductant of b-559LP.  相似文献   

5.
A simple and reliable continuous assay for measurement of α-mannosidase activity is described and demonstrated for analysis with two recombinant human enzymes using the new substrate resorufin α-d-mannopyranoside (Res-Man). The product of enzyme reaction, resorufin, exhibits fluorescence emission at 585 nm with excitation at 571 nm and has a pKa of 5.8, allowing continuous measurement of fluorescence turnover at or near physiological pH values for human lysosomal and Drosophila Golgi α-mannosidases. The assay performed using recombinant Drosophila Golgi α-mannosidase (dGMII) has been shown to give the kinetic parameters Km of 200 μM and Vmax of 11 nmol/min per nmol dGMII. Methods for performing the assay using several concentrations of the known α-mannosidase inhibitor swainsonine are also presented, demonstrating a potential for use of the assay as a simple method for high-throughput screening of inhibitors potentially useful in cancer treatment.  相似文献   

6.
An α,β-dicarbonyl reductase activity was purified from Saccharomyces cerevisiae and identified as the cytosolic enzyme d-Arabinose dehydrogenase (ARA1) by MALDI-TOF/TOF. Size exclusion chromatography analysis of recombinant Ara1p revealed that this protein formed a homodimer. Ara1p catalyzed the reduction of the reactive α,β-dicarbonyl compounds methylglyoxal, diacetyl, and pentanedione in a NADPH dependant manner. Ara1p had apparent Km values of ∼ 14 mM, 7 mM and 4 mM for methylglyoxal, diacetyl and pentanedione respectively, with corresponding turnover rates of 4.4, 6.9 and 5.9 s− 1 at pH 7.0. pH profiling showed that Ara1p had a pH optimum of 4.5 for the diacetyl reduction reaction. Ara1p also catalyzed the NADP+ dependant oxidation of acetoin; however this back reaction only occurred at alkaline pH values. That Ara1p was important for degradation of α,β-dicarbonyl substrates was further supported by the observation that ara1-Δ knockout yeast mutants exhibited a decreased growth rate phenotype in media containing diacetyl.  相似文献   

7.
α-d-Galactosidases (α-d-galactoside galactohydrolase, EC 3.2.1.22) from normal coconut endosperm were isolated and partially purified by a combination of ammonium sulfate fractionation, SP-Sephadex C50–120 ion-exchange chromatography and Sephadex G-200 and G-100 gel filtration. Two molecular forms of the enzyme, designated as A and B, were eluted after SP-Sephadex C50–120 ion-exchange chromatography. α-d-Galactosidase A, which is the major isoenzyme, was partially purified 43-fold on Sephadex G-200 and has a MW of about 23 000 whereas α-d-galactosidase B was partially purified 23-fold on Sephadex G-100 and has a similar MW of about 26 600. Both isoenzymes exhibited optimum activity at pH 7.5. The apparent Km and Vmax of α-d-galactosidase A were obtained at 3.46 × 10?4M and 1.38 × 10?3 M p-nitrophenyl α-<d-galactoside, respectively. A distinct substrate inhibition was noted. The enzyme was inhibited strongly by d-galactose and to a lesser extent by myo-inositol, d-glucose-6-phosphate, l-arabinose, melibiose and iodoacetic acid. Similarly, makapuno α-d-galactosidase was localized in the 40–70 % (NH4)2SO4 cut but its optimum activity at pH 7.5 was considerably lower as compared to the normal. Its Km was obtained at 6.75 × 10?4 M p-nitrophenyl α-d-galactoside while the Vmax was noted at 5.28 × 10?3 M p-nitrophenyl α-d-galactoside. Based on the above kinetic data, the possible cause(s) of the deficiency of α-d-galactosidase activity in makapuno is discussed.  相似文献   

8.
Schizokinen, a citrate-containing dihydroxamate, is a siderophore produced by Bacillus megaterium and Anabaena sp. The involvement of the citrate α-hydroxycarboxylate moiety in iron chelation was investigated by comparing the iron binding behavior of schizokinen with that of acetylschizokinen, a derivative in which the citrate hydroxyl group was modified by acetylation. Ferric schizokinen was found to exhibit an absorption spectrum (λmax = 460 nm) characteristic of a dihydroxamate below pH 2.5, with an isosbestic shift to a citrate dihydroxamate spectrum (λmax = 395 nm) above pH 4. Ferric acetylschizokinen also had a dihydroxamate absorption spectrum (λmax = 465 nm) at low pH. However, its spectral shift (λmax = 420 nm) and intensity above pH 4 were more typical of a ferric trihydroxamate. The molecular weight and electrophoretic mobility of ferric acetylschizokinen are consistent with a dimeric Fe2 (acetylschizokinen)3 structure, whereas ferric schizokinen appears to exist as a monomeric 1:1 complex Despite the differences in molecular weight and α-hydroxycarboxylate coordination, both complexes are effective in promoting iron uptake in Anabaena.  相似文献   

9.
Acetoacetate (AA) and 2-methylacetoacetate (MAA) are accumulated in metabolic disorders such as diabetes and isoleucinemia. Here we examine the mechanism of AA and MAA aerobic oxidation initiated by myoglobin (Mb)/H2O2. We propose a chemiluminescent route involving a dioxetanone intermediate whose thermolysis yields triplet α-dicarbonyl species (methylglyoxal and diacetyl). The observed ultraweak chemiluminescence increased linearly on raising the concentration of either Mb (10-500 μM) or AA (10-100 mM). Oxygen uptake studies revealed that MAA is almost a 100-fold more reactive than AA. EPR spin-trapping studies with MNP/MAA revealed the intermediacy of an α-carbon-centered radical and acetyl radical. The latter radical, probably derived from triplet diacetyl, is totally suppressed by sorbate, a well-known quencher of triplet carbonyls. Furthermore, an EPR signal assignable to MNP-AA adduct was observed and confirmed by isotope effects. Oxygen consumption and α-dicarbonyl yield were shown to be dependent on AA or MAA concentrations (1-50 mM) and on H2O2 or tert-butOOH added to the Mb-containing reaction mixtures. That ferrylMb is involved in a peroxidase cycle acting on the substrates is suggested by the reaction pH profiles and immunospin-trapping experiments. The generation of radicals and triplet dicarbonyl products by Mb/H2O2/β-ketoacids may contribute to the adverse health effects of ketogenic unbalance.  相似文献   

10.
Recent reports of extremely low retinoid stores in fish living in contaminated river systems prompted an initial investigation of the mechanisms of hepatic storage and mobilization in brook trout. Enzyme characterization in microsomes revealed a lecithin:retinol acyltransferase activity (LRAT) optimum in the alkaline range (pH 9.0; Vmax=0.6 nmol per mg prot. h−1; Km=10.2 μM) which is not known to occur in mammals, in addition to a secondary optimum at pH 6.5 typical of mammals. Acyl CoA:retinol acyltransferase (ARAT) kinetic parameters were quite different to those of mammals. The substrate affinity of trout ARAT (Km=1.6 μM) was approximately 22-fold greater than that of the rat while maximal velocity (Vmax=0.2 nmol per mg prot. h−1) was 18-fold less. Retinyl ester hydrolase activity (REH) was optimal under acid conditions (pH 4.2; Vmax=6.6 nmol per mg prot. h−1; Km=0.6 mM), was inhibited by a bile salt analogue and was greater in males than females. This REH was tentatively categorized as a bile salt-independent, acid retinyl ester hydrolase (BSI-AREH). REH was inhibited in a dose-dependent manner following in vivo exposure to a representative environmental contaminant the coplanar polychlorinated biphenyl (PCB), 3,3′,4,4′-tetrachlorobiphenyl (TCBP). Inhibition may be an indirect effect because enzyme activity was not affected by in vitro exposure of control microsomes. REH inhibition in the brook trout may affect the uptake of retinyl esters (REs) from chylomicron remnants as well as the mobilization of stored REs.  相似文献   

11.
The iron-containing violet acid phosphatases from beef spleen and pig allantoic fluid have been purified to homogeneity. Molecular weight determinations by zonal gel filtration, SDS-gel electrophoresis, and ultracentrifugation support values close to 40,000 for both enzymes, necessitating reappraisal of literature values. Similarly, the equivalent weight for iron is close to 20,000 for both enzymes, indicating the presence of two iron atoms per molecule of enzyme. The enzymes also have very similar ultraviolet and visible spectra, with λmax values close to 550 nm, and ?550 values(in terms of iron) of 2.04 × 103 and 2.00 × 103 for the beef spleen and pig allantoic fluid enzymes respectively.  相似文献   

12.
The potato tuber lipoxygenase preparations convert α-linolenic acid not only to 9(S)-HPOTE, but also to some more polar metabolites. Two of these polar products, I and II, with ultraviolet absorbance maxima at 267 nm were purified by HPLC. It was found that metabolites I and II have, respectively, one and two hydroperoxy groups. Products of NaBH4 reduction of both I and II were identified by their chemical ionization and electron impact mass spectra and by 1H-NMR spectra as 9,16-dihydroxy-10(E), 12(Z), 14(E)-octadecatrienoic acid. The obtained results suggest that compound II is 9,16-dihydroperoxy-10(E), 12(Z), 14(E)-octadecatrienoic acid and product I is a mixture of two positional isomers, 9-hydroxy-16-hydroperoxy-10(E),12(Z),14(E)-octadecatrienoic and 9-hydroperoxy-16-hydroxy-10(E),12(Z), 14(E)-octadecatrienoic acids. Lipoxygenase converts efficiently [14C]9-HOTE into product I. Also, both metabolites I and II are the products of double dioxygenation. The second oxygenation at C-16 position as well as the first one at C-9 is controlled by lipoxygenase.  相似文献   

13.
This work demonstrated a 10-day batch culture experiment to test the physiology and toxicity of harmful dinoflagellate Karenia mikimotoi in response to ocean acidification (OA) under two different phosphate concentrations. Cells were previously acclimated in OA (pH = 7.8 and CO2 = 1100 μatm) condition for about three months before testing the responses of K. mikimotoi cells to a two-factorial combinations experimentation. This work measured the variation in physiological parameters (growth, rETR) and toxicity (hemolytic activity and its toxicity to zebrafish embryos) in four treatments, representing two factorial combinations of CO2 (450 and 1100 μatm) and phosphate concentration (37.75 and 4.67 umol l−1). Results: OA stimulated the faster growth, and the highest rETRmax in high phosphate (HP) treatment, low phosphate (LP) and a combination of high CO2 and low phosphate (HC*LP) inhibited the growth and Ek in comparison to low CO2*high phosphate (LCHP) treatment. The embryotoxicity of K. mikimotoi cells enhanced in all high CO2 (HC) conditions irrespective of phosphate concentration, but the EC50 of hemolytic activity increased in all high CO2 (HC) and low phosphate (LP) treatments in comparison of LCHP. Ocean acidification (high CO2 and lower pH) was probably the main factor that affected the rETRmax, hemolytic activity and embryotoxicity, but low phosphate was the main factor that affected the growth, α, and Ek. There were significant interactive effects of OA and low phosphate (LP) on growth, rETRmax, and hemolytic activity, but there were no significant effects on α, Ek, and embryotoxicity. If these results are extrapolated to the aquatic environment, it can be hypothesized that the K. mikimotoi cells were impacted significantly by future changing ocean (e.g., ocean acidification and nutrient stoichiometry).  相似文献   

14.
Summary

Wheat bran contains several ester-linked dehydrodimers of ferulic acid, which were detected and quantified after sequential alkaline hydrolysis. The major dimers released were: trans-5-[(E)-2-carboxyvinyl]-2-(4-hydroxy-3-methoxy-phenyl)-7-methoxy-2,3-dihydrobenzofuran-3-carboxylic acid (5–8-BendiFA), (Z)-β-(4-[(E)-2-carboxyvinyl]-2-methoxy-phenoxy)-4-hydroxy-3-methoxycinnamic acid (8-O-4-diFA) and (E,E)-4,4′-dihydroxy-5,5′-dimethoxy-3,3′-bicinnamic acid (5–5-diFA). trans-7-hydroxy-1-(4-hydroxy-3methoxyphenyl)-6-methoxy-1,2-dihydro-naphthalene-2,3-dicarboxylic acid (8–8-diFA cyclic form) and 4,4′-dihydroxy-3,3′-dimethoxy-β,β'-bicinnamic acid (8–8-diFA non cyclic form) were not detected. One of the most abundant dimers, 8-O-4-diFA, was purified from de-starched wheat bran after alkaline hydrolysis and preparative HPLC. The resultant product was identical to the chemically synthesised 8-O-4-dimer by TLC and HPLC as confirmed by 1H-NMR and mass spectrometry. The absorption maxima and absorption coefficients for the synthetic compound in ethanol were: λmax: 323 nm, λmin: 258 nm, ελmax (M?1cm?1): 24800 ± 2100 and ε280 (M?1cm?1): 19700 ± 1100. The antioxidant properties of 8-O-4-diFA were assessed using: (a) inhibition of ascorbate/iron-induced peroxidation of phosphatidylcholine liposomes and; (b) scavenging of the radical cation of 2,2′-azinobis (3-ethyl-benzothiazoline-6-sulphonate) (ABTS) relative to the water-soluble vitamin E analogue, Trolox C. The 8-O-4-diFA was a better antioxidant than ferulic acid in both lipid and aqueous phases. This is the first report of the antioxidant activity of a natural diferulate obtained from a plant.  相似文献   

15.
《BBA》1987,893(2):275-288
The membrane-bound ATP synthase from chloroplasts can occur in different redox and activation states. In the absence of reductants the enzyme usually is oxidized and inactive, Eoxi. Illumination in the presence of dithiothreitol leads to an active, reduced enzyme, Ereda. If this form is stored in the dark in the presence of dithiothreitol an inactive, reduced enzyme Eredi is formed. The rates of ATP synthesis and ATP hydrolysis catalyzed by the different enzyme species are measured as a function of ΔpH (Δψ = 0 mV). The ΔpH was generated with an acid-base transition using a rapid-mixing quenched flow apparatus. The following results were obtained. (1) The oxidized ATP synthase catalyzes high rates of ATP synthesis, voxmax = 400 ATP per CF0F1 per s. The half-maximal rate is obtained at ΔpH = 3.4. (2) The active, reduced ATP synthase catalyzes high rates of ATP synthesis, vredmax = 400 ATP per CF0F1 per s. The half-maximal rate is obtained at ΔpH = 2.7. It catalyzes also high rates of ATP hydrolysis vredmax = −90 ATP per CF0F per s at ΔpH = 0. (3) The inactive species (both oxidized and reduced) catalyze neither ATP synthesis nor ATP hydrolysis. The activation/inactivation of the reduced enzyme is completely reversible. (4) The activation of the reduced, inactive enzyme is measured as a function of ΔpH by measuring the rate of ATP hydrolysis catalyzed by the active species. Half-maximal activation is observed at ΔpH = 2.2. (5) On the basis of these results a reaction scheme is proposed relating the redox reaction, the activation and the catalytic reaction of the chloroplast ATP synthase.  相似文献   

16.
A microwave assisted facile synthesis of a fluorescent 6-O-naphthylacetyl agarose (NA-agarose) employing carbodiimide chemistry (dicyclohexylcarbodiimide/4-dimethylaminopyridine) has been described. NA-agarose was characterized by TGA, GPC, UV spectrophotometry, fluorescence spectroscopy, FT-IR, 1H and 13C NMR spectra, exhibiting that in NA-agarose the naphthylacetyl group was attached to the backbone of the agarose polymer. The hydrolysis of NA-agarose in heterogeneous aqueous phase showed that the 1-naphthyl acetic acid (NAA), a plant growth regulator, got released in a controlled manner, the release rate being dependent on the hydrophilicity of the polymer adduct as well as on pH and temperature. The fluorescence emission (λmax 332 nm) of NA-agarose (1 × 10−3 M) in ethylene glycol was significantly higher (ca. 82%) than that of the molar equivalent of NAA content in the product i.e. 0.08 mg in 1 × 10−3 M solution. The resulting polymer would be of potential utility as a sustained release plant growth regulator and sensory applications.  相似文献   

17.
Membranes were prepared from fresh, washed human erythrocytes by hemolysis and washing with 5 mm sodium phosphate buffer (pH 7.4). The mean residue ellipticity, [θ], of erythrocyte membrane circular dichroism was altered by prostaglandin E1 or prostaglandin F at 37 °C when observed from 250 nm to 190 nm. The decrease in negativity of [θ] with 10?6m prostaglandin E1 was 12.7% at 222 nm and 17.7% at 208 nm, and with 10?6m prostaglandin F 22.5% and 34.2%, respectively (P < 0.01). Similar changes in [θ] were observed at lower concentrations of prostaglandins. No strict relationship between amount of change of [θ] and prostaglandin concentrations of 3 × 10?5m to 3 × 10?12m was evident. A persistent alteration of [θ] with prostaglandin was observed at 37 °C. Transient change of [θ] occurred at 25 °C with prostaglandin. No change of [θ] was observed at 15 or 20 °C. Buffer or palmitic acid were without effect on membrane [θ]. Phosphatidyl inositol or methyl arachidonate caused an increase in negativity of membrane spectra. The observed alterations of membrane [θ] did not arise from changes in light scattering as the OD700–OD200 of membranes was not changed by prostaglandin. Effects of prostaglandin were not dependent on light path length. The prostaglandin E1 antagonist, 7-oxa-13-prostynoic acid, at 10?7m produced no change of [θ] of membrane spectra and prevented the otherwise demonstrable effects of 10?10m prostaglandin E1 on [θ]. The decrease in negativity of [θ] at 222 nm is indicative of a decrease in ellipticity of membrane protein. These studies suggest that prostaglandins may act by inducing a conformational change in membrane protein.  相似文献   

18.
The amino acid l-lysine is synthesized in Saccharomyces cerevisiae via the α-aminoadipate pathway. An as yet unidentified PLP-containing aminotransferase is thought to catalyze the formation of α-aminoadipate from α-ketoadipate in the l-lysine biosynthetic pathway that could be the yeast Aro8 gene product. A screen of several different amino acids and keto-acids showed that the enzyme uses l-tyrosine, l-phenylalanine, α-ketoadipate, and l-α-aminoadipate as substrates. The UV–visible spectrum of the aminotransferase exhibits maxima at 280 and 343 nm at pH 7.5. As the pH is decreased the peak at 343 nm (the unprotonated internal aldimine) disappears and two new peaks at 328 and 400 nm are observed representing the enolimine and ketoenamine tautomers of the protonated aldimine, respectively. Addition, at pH 7.1, of α-ketoadipate to free enzyme leads to disappearance of the absorbance at 343 nm and appearance of peaks at 328 and 424 nm. The V/Et and V/Kα-ketoadipateEt pH profiles are pH independent from pH 6.5 to 9.6, while the V/Kl-tyrosine pH-rate profile decreases below a single pKa of 7.0 ± 0.1. Data suggest the active enzyme form is with the internal aldimine unprotonated. We conclude the enzyme should be categorized as a α-aminoadipate aminotransferase.  相似文献   

19.
The incubation of [5,6-3H]prostaglandin E1 ([3H]PGE1) with guinea pig kidney cortex microsomes in the presence of NADPH in an atmosphere of air, resulted in chromatographically polar metabolites. The incubation products were treated with base which converted PGE1 derivatives into PGB1 derivatives, with a λmax = 278 nm and the products were analyzed by TLC and high pressure-liquid chromatography (HPLC). Based on UV absorption, mobility on TLC and retention time in HPLC, as compared with authentic compounds, it was concluded that the two polar UV-absorbing peaks in HPLC represented 19-hydroxy-PGB1 (19-OH-PGB1) and 20-hydroxy-PGB1 (20-OH-PGB1). Further identification of the metabolites was obtained by derivatizing the incubation products as methyl esters and t-butyldimethylsilyl ethers, followed by co-injection with similarly derivatized authentic compounds in HPLC and gas chromatography. Finally, the derivatized metabolites were identified by comparing their mass fragmentation with that of similarly derivatized authentic compounds. There was an absolute requirement for NADPH, and NADH did not significantly support the hydroxylation of PGE1. Inhibitors of microsomal monooxygenase (SKF 525A, metyrapone, and cytochrome c) inhibited the hydroxylation of PGE1 by kidney cortex microsomes. By contrast, carbon monoxide at a CO:O2 ratio of 5:1 did not inhibit the hydroxylation of PGE1, pointing to a low or lack of CO sensitivity of the hydroxylation of PGE1. The addition of PGE1 or laurate to guinea pig kidney cortex microsomes elicited Type I spectral changes. The spectral dissociation constant (Ks) for PGE1 was 2.4 × 10?4m. The kinetic constants for 19- and 20-hydroxylations of PGE1 were determined. The KM values for the 19- and 20-hydroxylation pathways were found to be identical, being 3.3 × 10?4m, suggesting that the same enzyme is involved in both hydroxylations; however, the Vmax values for 19-hydroxylation and 20-hydroxylation of PGE1 were 50 nmol/hr and 20.8 nmol/hr respectively. These results demonstrate that PGE1 is a substrate for the kidney cortex microsomal monooxygenase. The similarities and differences of the kidney monooxygenase in the guinea pig with that in the rat are discussed.  相似文献   

20.
Clostridium acetobutylicum P262 cells that were growing on lactate and acetate had an NAD-independent lactate dehydrogenase (iLDH) activity of 200 nmol mg protein−1 min−1. Ammonium sulfate precipitation and DEAE cellulose caused a 35-fold purification. Gel filtration indicated that the iLDH had a molecular weight of approximately 55 kDa, but two bands were always observed. Phenyl sepharose could not separate the two proteins, and hydroxyapatite caused a complete loss of activity. The semi-purified iLDH had a Vmax of 13,000 nmol mg protein−1 min−1 and a K m value of 3.5 mM for D-lactate. The Vmax and K m values for L-lactate were 300 nmol mg protein−1 min−1 and 0.7 mM. The iLDH had a pH optimum of 7.5, was not activated by fructose-1,6-bisphosphate (FDP), and could be coupled to either 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) or dichlorophenol-indophenol (DCPIP), but not methyl viologen (MV) or benzyl viologen (BV). The iLDH did not have strong absorbance between 500 and 300 nm, and trichloroacetic acid or acid ammonium sulfate extracts had virtually no fluorescence at 450 nm. The crude extracts also had MTT-linked butyryl-CoA dehydrogenase activity (60 nmol mg protein−1 min−1). The NAD-independent butyryl-CoA dehydrogenase eluted from DEAE-cellulose as two fractions. The yellow fraction was extremely unstable, but the green fraction could be stored for short periods of time at 5°C. The green-colored butyryl-CoA dehydrogenase had strong absorption at 450 nm, and gel filtration indicated that it had a molecular weight of 90 kDa. The NAD-independent butyryl-CoA dehydrogenase could be coupled to MTT, DCPIP, or MV, but not BV. Because the NAD-independent lactate and butyryl-CoA dehydrogenase could both be linked to low potential carriers, these two enzymes may function as oxidation-reduction system in vivo. Received: 24 July 1996 / Accepted: 10 September 1996  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