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.
Deuterium isotope effects [D(V/K)] and stereoselectivity of ethanol oxidation in cytochrome P-450 containing systems and in the xanthine-xanthine oxidase system were compared with those of yeast alcohol dehydrogenase. The isotope effects were determined by using both a noncompetitive method, including incubation of unlabeled or [1,1-2H2]ethanol at various concentrations, and a competitive method, where 1:1 mixtures of [1-13C]- and [2H6]ethanol or [2,2,2-2H3]- and [1,1-2H2]ethanol were incubated and the acetaldehyde formed was analyzed by gas chromatography/mass spectrometry. The D(V/K) isotope effects of the cytochrome P-450 dependent ethanol oxidation were about 4 with liver microsomes from imidazole-, phenobarbital- or acetone-treated rabbits or with microsomes from acetone- or ethanol-treated rats. Similar isotope effects were reached with reconstituted membranes containing the rabbit ethanol-inducible cytochrome P-450 (LMeb), whereas control rat microsomes and membranes containing rabbit phenobarbital-inducible P-450 LM2 oxidized the alcohol with D(V/K) of about 2.8 and 1.8, respectively. Addition of FeIIIEDTA either to microsomes from phenobarbital-treated rabbits or to membranes containing P-450 LMeb significantly lowered the isotope effect, which approached that of the xanthine-xanthine oxidase system (1.4), whereas desferrioxamine had no significant effect. Incubations of all cytochrome P-450 containing systems or the xanthine-xanthine oxidase systems with (1R)- and (1S)-[1-2H]ethanol, revealed, taking the isotope effects into account, that 44-66% of the ethanol oxidized had lost the 1-pro-R hydrogen.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The reactions of RO(2)* radicals with Fe(H(2)O)(6)(2+) were studied, R[double bond]H; CH(3); CH(2)COOH; CH(2)CN; CH(2)C(CH(3))(2)OH; CH(2)OH; CHCl(2)/CCl(3). All these processes involve the following reactions: Fe(H(2)O)(6)(2+)+RO(2)*<==>(H(2)O)(5)Fe(III)[bond]OOR(2+) K(1) approximately 250 M(-1); (H(2)O)(5)Fe(III)[bond]OOR(2+)+H(3)O(+)/H(2)O-->Fe(H(2)O)(6)(3+)+ROOH+H(2)O/OH(-); (H(2)O)(5)Fe(III)[bond]OOR(2+)+2Fe(H(2)O)(6)(2+)-->3Fe(H(2)O)(6)(3+)+ROH; 2 RO(2)*-->Products; RO(2)*+(H(2)O)(5)Fe(III)[bond]OOR(2+)-->Fe(H(2)O)(6)(2+)+products. The values of k(1) and k(3) [reaction is clearly not an elementary reaction] approach the ligand exchange rate of Fe(H(2)O)(6)(2+), i.e. these reactions follow an inner sphere mechanism and the rate determining step is the ligand exchange step. The rate of reaction is several orders of magnitude faster than that of the Fenton reaction. Surprisingly enough the K(1) values are nearly independent of the redox potential of the radical and are considerably higher than calculated from the relevant redox potentials. These results indicate that the ROO(-) ligands considerably stabilise the Fe(III) complex, this stabilisation is smaller for radicals with electron withdrawing groups which raise the redox potential of the radical but decrease the basicity of the ROO(-) ligands, two effects which seem to nearly cancel each other. Finally, the results clearly indicate that reaction (5) is relatively fast and affects the nature of the final products. The contribution of these reactions to oxidation processes involving 'Fenton-like' processes is discussed.  相似文献   

3.
For the purpose of developing highly sensitive and convenient determination of plasmalogens, the high-performance liquid chromatography (HPLC) method using radioactive iodine ((125)I) was investigated. Radioactive triiodide (1-) ion ((125)I(3)(-)), which is an actual iodine form capable of reacting with vinyl ether bond ([bond]CH(2)[bond]O[bond]CH[double bond]CH[bond]) of plasmalogens, could be safely and efficiently produced by oxidizing a commercial radioactive sodium iodine (Na(125)I) with hydrogen peroxide (H(2)O(2)) under acid condition (pH 5.5-6.0), which is called iodine-125 reagent. I(3)(-) specifically reacted with plasmalogens at the molar ratio of 1:1 in methanol, and 1 or 2 mol of plasmalogens was involved in the binding with iodine per iodine atom, resulting in the formation of stable iodine-binding phospholipids. The HPLC system with Diol column and acetonitrile/water as a mobile phase was available for separating iodine-binding phospholipids from nonbinding free iodine and for separately eluting iodine-binding phospholipids derived from choline and ethanolamine plasmalogens. Using iodine-125 reagent (1.85 MBq/ml), plasmalogens were detectable at high sensitivity of 10,000-15,000 cpm/nmol, which is more than 1000-fold higher sensitivity than the classical determination with nonradioactive iodine. Plasmalogen concentrations in human plasma were measured with the HPLC system and determined as, on average, 129.1+/-31.3 microM (n=8) in a 1.2 content ratio of choline to ethanolamine plasmalogens, a concentration that nearly agrees with the value reported previously.  相似文献   

4.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

5.
Elimination of [2H]ethanol in vivo as studied by gas chromatography/mass spectrometry occurred at about half the rate in deer mice reported to lack alcohol dehydrogenase (ADH-) compared with ADH+ deer mice and exhibited kinetic isotope effects on Vmax and Km (D(V/K] of 2.2 +/- 0.1 and 3.2 +/- 0.8 in the two strains, respectively. To an equal extent in both strains, ethanol elimination was accompanied by an ethanol-acetaldehyde exchange with an intermolecular transfer of hydrogen atoms, indicating the occurrence of dehydrogenase activity. This exchange was also observed in perfused deer mouse livers. Based on calculations it was estimated that at least 50% of ethanol elimination in ADH- deer mice was caused by the action of dehydrogenase systems. NADPH-supported cytochrome P-450-dependent ethanol oxidation in liver microsomes from ADH+ and ADH- deer mice was not stereoselective and occurred with a D(V/K) of 3.6. The D(V/K) value of catalase-dependent oxidation was 1.8, whereas a kinetic isotope effect of cytosolic ADH in the ADH+ strain was 3.2. Mitochondria from both ADH+ and ADH- deer mice catalyzed NAD+-dependent ethanol oxidation and NADH-dependent acetaldehyde reduction. The kinetic isotope effects of NAD+-dependent ethanol oxidation in the mitochondrial fraction from ADH+ and ADH- deer mice were 2.0 +/- 0.1 and 2.3 +/- 0.3, respectively. The results indicate only a minor contribution by cytochrome P-450 to ethanol elimination, whereas the isotope effects are consistent with ethanol oxidation by the catalase-H2O2 system in ADH- deer mice in addition to the dehydrogenase systems.  相似文献   

6.
Rates of exchange catalysed by alcohol dehydrogenase were determined in vivo in order to find rate-limiting steps in ethanol metabolism. Mixtures of [1,1-2H2]- and [2,2,2-2H3]ethanol were injected in rats with bile fistulas. The concentrations in bile of ethanols having different numbers of 2H atoms were determined by g.l.c.-m.s. after the addition of [2H6]ethanol as internal standard and formation of the 3,5-dinitrobenzoates. Extensive formation of [2H4]ethanol indicated that acetaldehyde formed from [2,2,2-2H3]ethanol was reduced to ethanol and that NADH used in this reduction was partly derived from oxidation of [1,1-2H2]ethanol. The rate of acetaldehyde reduction, the degree of labelling of bound NADH and the isotope effect on ethanol oxidation were calculated by fitting models to the found concentrations of ethanols labelled with 1-42H atoms. Control experiments with only [2,2,2-2H3]ethanol showed that there was no loss of the C-2 hydrogens by exchange. The isotope effect on ethanol oxidation appeared to be about 3. Experiments with (1S)-[1-2H]- and [2,2,2-2H3]ethanol indicated that the isotope effect on acetaldehyde oxidation was much smaller. The results indicated that both the rate of reduction of acetaldehyde and the rate of association of NADH with alcohol dehydrogenase were nearly as high as or higher than the net ethanol oxidation. Thus, the rate of ethanol oxidation in vivo is determined by the rates of acetaldehyde oxidation, the rate of dissociation of NADH from alcohol dehydrogenase, and by the rate of reoxidation of cytosolic NADH. In cyanamide-treated rats, the elimination of ethanol was slow but the rates in the oxidoreduction were high, indicating more complete rate-limitation by the oxidation of acetaldehyde.  相似文献   

7.
Kumar NN  Swamy KC 《Chirality》2008,20(6):781-789
Diastereoselective synthesis and characterization of chiral unsymmetrical tris-spirocyclic cyclotriphosphazenes based on chiral 1,1'-bi-2-naphthol (BINOL) are reported. Specifically, the chiral compounds (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)Cl(2) [(-)-4] and (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](OCH(2)CH(2)NMe)(2) [(-)-5] are prepared by starting with the chiral mono-spiro compound (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)]Cl(4) [(-)-3]. Synthesis of four other chiral spirocyclics, N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](OCH(2)CH(2) NMe)(O-2,2'C(6)H(4)-C(6)H(4)O)[(-)-6 and (+)-6], N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](NMe(2))(4) [(-)-7], N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)(NMeCH(2)CH(2)OH)(2) [(-)-8 and (+)-8], and N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)[NHCH(2)CH(2)CH(2)Si(OEt)(3)](2) (9) is also reported herein. Compounds 4-6 are obtained in the solid state diastereoselectively and their X-ray structures have been determined and discussed. The diastereoselectivity is also shown by structural characterization of two distinct isomers in the case of 6 [(-)-6 and (+)-6, respectively] by starting with precursor of 3 having (R) or (S)-BINOL residue. The (1)H NMR spectra of 7 and 8 exhibit doublets with virtual coupling for the methyl protons, consistent with the chiral nature of the binaphthoxy residue. The potential of 9, which hydrolyzes readily in CDCl(3) solution, as a useful precursor for chiral polymer applications is highlighted.  相似文献   

8.
The oxidation of 1,1,2,2-tetrachloroethane to dichloroacetic acid was investigated with rat liver microsomes and purified cytochrome P-450. Deuterium substitution had no effect on Km values, but both the inter- and intramolecular isotope effects (kH/kD) on Vmax were in the range 5.7-6.1. The equivalence of the inter- and intramolecular values indicates that 6.0 may be a good estimate of the intrinsic isotope effect. The intermolecular kH/kD value for the conversion of 1,1,2,2-trichloroethane and its 1-2H analog to chloroacetic acid was 5.5. These data, and the finding that 1 atom of 18O was incorporated into the product when TCEA was oxidized in an 18O2 atmosphere, support an oxidative dechlorination mechanism that involves hydrogen atom abstraction by the P-450 intermediate oxo complex.  相似文献   

9.
M A Vanoni  R G Matthews 《Biochemistry》1984,23(22):5272-5279
Previous work from this laboratory has established that the NADPH-menadione oxidoreductase reaction catalyzed by methylenetetrahydrofolate reductase from pig liver proceeds by Ping Pong Bi Bi kinetics and that the reductive half-reaction is rate limiting in steady-state turnover. We have now shown that methylenetetrahydrofolate reductase stereo-specifically removes the pro-S hydrogen from the 4-position of NADPH. During the oxidation of [4(S)-3H]NADPH, we observed a kinetic isotope on V/KNADPH of 10.8 +/- 0.4. When comparing the rates of oxidation of [4(S)-2H]NADPH and [4(S)-1H]NADPH, we measure kinetic isotope effects on V of 4.78 +/- 0.15 and on V/KNADPH of 4.54 +/- 0.59. When oxidation of [4(R)-2H]NADPH and [4(R)-1H]NADPH is compared, the secondary kinetic isotope effect on V is 1.04 +/- 0.01. When the NADPH-menadione oxidoreductase reaction is catalyzed in tritiated water, no incorporation of solvent tritium into residual NADPH is observed. We conclude from these observations that the oxidation of NADPH is largely or entirely rate limiting in the reductive half-reaction and, hence, in NADPH-menadione oxidoreductase turnover at saturating menadione concentration. In the presence of saturating NADPH, the flavin reduction proceeds with a rate constant of 160 S-1, which is at least 29-fold slower than estimates of the lower limit for the diffusion-limited rate constant characterizing NADPH binding to the enzyme under physiological conditions. Albery & Knowles have defined criteria for perfection in enzyme catalysis [Albery, W. J., & Knowles, J.R. (1976) Biochemistry 15, 5631-5640].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
11.
In this study, we irradiated the antioxidant kaempferol in ethanol and methanol solutions with gamma rays at doses ranging from 0.2-20 kGy. NMR and ES-MS spectroscopy were used to identify radiolysis products. Two depsides, [2-[(4'-hydroxybenzoyl)oxy]-4,6-dihydroxyphenyl](oxo) methyl acetate and [2-[(4'-hydroxybenzoyl)oxy]-4,6-dihydroxyphenyl](oxo) ethyl acetate, were the major compounds of kaempferol degradation in methanol and in ethanol, respectively. Other products formed in low concentrations were identified as [4-hydroxyphenyl](oxo) methyl acetate, [4-hydroxyphenyl](oxo) ethyl acetate, and depside [2-[(4'-hydroxybenzoyl)oxy]-4,6-dihydroxyphenyl](oxo) acetic acid. The formation of the latter was observed in both solvents. We propose degradation mechanisms that suggest that (.)CH(2)OH and CH(3)(.)CHOH, produced by solvent radiolysis, react with the 3-OH kaempferol group because of its high H-donor capacity. pi-Electron delocalization in the flavonoxy formed after the first H-transfer leads to C-ring opening and consequently to the formation of depsides. G calculation of the degradation products and of (.)CH(2)OH and CH(3)(.)CHOH radicals confirmed the proposed mechanism of kaempferol radiolysis. The rate constants for the reaction between kaempferol and these free radicals were also calculated. Formation of depside has also been observed in many studies of the oxidation of flavonoids; those studying human metabolism have suggested similar redox transformation of flavonols. The antioxidant activities of radiolysis products were evaluated and compared to those of kaempferol.  相似文献   

12.
A bioorganometallic approach to malaria therapy led to the discovery of ferroquine (FQ, SSR97193). To assess the importance of the electronic properties of the ferrocenyl group, cyclopentadienyltricarbonylrhenium analogues related to FQ, were synthesized. The reaction of [N-(7-chloro-4-quinolinyl)-1,2-ethanodiamine] with the cyrhetrenylaldehyde complexes (η(5)-C(5)H(4)CHO)Re(CO)(3) and [η(5)-1,2-C(5)H(3)(CH(2)OH)(CHO)]Re(CO)(3) produces the corresponding imine derivatives [η(5)-1,2-C(5)H(3)(R)(CHN-CH(2)CH(2)NH-QN)]Re(CO)(3) R=H 3a; R=CH(2)OH 3b; QN=N-(7-Cl-4-quinolinyl). Reduction of 3a and 3b with sodium borohydride in methanol yields quantitatively the amine complexes [η(5)-1,2-C(5)H(3)(R)(CH(2)-NH-CH(2)CH(2)NH-QN)]Re(CO)(3) R=H 4a; R=CH(2)OH 4b. To establish the role of the cyrethrenyl moiety in the antimalarial activity of this series, purely organic parent compounds were also synthesized and tested. Evaluation of antimalarial activity measured in vitro against the CQ-resistant strains (W2) and the CQ-susceptible strain (3D7) of Plasmodium falciparum indicates that these cyrhetrene conjugates are less active compared to their ferrocene and organic analogues. These data suggest an original mode-of-action of FQ and ferrocenyl analogues in relationship with the redox pharmacophore.  相似文献   

13.
Stereochemical aspects of the formation of double bonds in abscisic acid   总被引:2,自引:1,他引:1  
The stereochemistry of the hydrogen elimination that occurs during the formation of the Delta(4)- and Delta(2)'-double bonds of abscisic acid has been determined from the (14)C/(3)H ratios in abscisic acid biosynthesized by avocado fruit from [2-(14)C,(2R)-2-(3)H(1)]-, [2-(14)C,(2S)-2-(3)H(1)]- and [2-(14)C,(5S)-5-(3)H(1)]-mevalonate. Setting the (14)C/(3)H ratio at 3:3 for [2-(14)C,(2R)-2-(3)H(1)]mevalonate, the corresponding ratio in derived methyl abscisate was 3:2.28; the analogous ratio for methyl abscisate from [2-(14)C,(2S)-2-(3)H(1)]mevalonate was 3:1.63. Removal of the 3'-hydrogen atom of abscisic acid by base-catalysed exchange altered the ratios to 3:1.55 and 3:1.44 respectively. It was concluded that this 3'-hydrogen atom is derived from the pro-2R-hydrogen atom of mevalonate. Removal of the 4-hydrogen atom from methyl abscisate by formation of a derivative, a lactone, lacking this hydrogen atom changed the ratio to 3:1.04 for material derived from [2-(14)C,(2R)-2-(3)H(1)]-mevalonate and to 3:1.05 for [2-(14)C,(2S)-2-(3)H(1)]mevalonate, showing that this hydrogen atom also is derived from the pro-2R-hydrogen atom of mevalonate. These ratios of the lactones are consistent with their retaining one (3)H atom at the 6'-methyl position of abscisic acid from the [(2R)-2-(3)H(1)]- and [(2S)-2-(3)H(1)]-mevalonate. The presence of some label at positions 3' and 4 when [(2S)-2-(3)H(1)]mevalonate was the precursor is attributed to the action of isopentenyl pyrophosphate isomerase. The hydrogen atom at C-5 of abscisic acid is derived from the pro-5S-hydrogen atom of mevalonate.  相似文献   

14.
When the stereospecifically deuterated dopamine enantiomers, (R)- and (S)-[alpha-2H1]dopamine, are incubated with amine oxidases, the deuterium atom may be either retained to form monodeuterated 3,4-dihydroxyphenylacetaldehyde, or eliminated to produce the nondeuterated or protio-aldehyde product. These two aldehydes can be separated from one another and identified by high-performance liquid chromatography with electrochemical detection. Three types of stereospecific abstraction of a hydrogen from the alpha-carbon of dopamine during deamination have been observed. In the first type, the pro-R hydrogen is removed from the alpha-carbon. Enzymes in this category are mitochondrial monoamine oxidases A and B, as isolated from different tissues and species. The second type of deamination involves the abstraction of pro-S hydrogen from the alpha-carbon of dopamine. Soluble enzymes, such as rat aorta benzylamine oxidase or diamine oxidase from hog kidney and pea seedling, have been found to belong to this group. Bovine plasma amine oxidase exhibits the third type of deamination where no absolute stereospecificity is required. This enzyme catalyzes the oxidation of either (S)- or (R)-[alpha-2H1]dopamine, preferably breaking the C-H bond rather than the C-2H bond in both cases. The kinetic deuterium isotope effect during the deamination of dopamine catalyzed by the different amine oxidases varies greatly; VH/VD ranges from 1.5 to 5.5. The high magnitude of the isotope effect suggests that hydrogen abstraction may be the rate-limiting step (i.e., in reactions catalyzed by benzylamine oxidase and monoamine oxidase). When the isotope effect is low (i.e., for diamine oxidases from hog kidney or pea seedling), it is uncertain if the breaking of the bond is rate limiting.  相似文献   

15.
The density functional theory (DFT) calculations are carried out to study the mechanism details and the ensemble effect of methanol dehydrogenation over Pt(3) and PtAu(2) clusters, which present the smallest models of pure Pt clusters and bimetallic PtAu clusters. The energy diagrams are drawn out along both the initial O-H and C-H bond scission pathways via the four sequential dehydrogenation processes, respectively, i.e., CH(3)OH → CH(2)OH → CH(2)O → CHO → CO and CH(3)OH → CH(3)O → CH(2)O → CHO → CO, respectively. It is revealed that the reaction kinetics over PtAu(2) is significantly different from that over Pt(3). For the Pt(3)-mediated reaction, the C-H bond scission pathway, where an ensemble composed of two Pt atoms is required to complete methanol dehydrogenation, is energetically more favorable than the O-H bond scission pathway, and the maximum barrier along this pathway is calculated to be 12.99 kcal mol(-1). In contrast, PtAu(2) cluster facilitates the reaction starting from the O-H bond scission, where the Pt atom acts as the active center throughout each elementary step of methanol dehydrogenation, and the initial O-H bond scission with a barrier of 21.42 kcal mol(-1) is the bottom-neck step of methanol decomposition. Importantly, it is shown that the complete dehydrogenation product of methanol, CO, can more easily dissociate from PtAu(2) cluster than from Pt(3) cluster. The calculated results over the model clusters provide assistance to some extent for understanding the improved catalytic activity of bimetal PtAu catalysts toward methanol oxidation in comparison with pure Pt catalysts.  相似文献   

16.
Bioethanol is one of the world’s most extensively produced biofuels. However, it is difficult to purify due to the formation of the ethanol–water azeotrope. Knowledge of the azeotrope structure at the molecular level can help to improve existing purification methods. In order to achieve a better understanding of this azeotrope structure, the characterization of (ethanol)5–water heterohexamers was carried out by analyzing the results of electronic structure calculations performed at the B3LYP/6-31+G(d) level. Hexamerization energies were found to range between ?36.8 and ?25.8 kcal/mol. Topological analysis of the electron density confirmed the existence of primary (OH…O) hydrogen bonds (HBs), secondary (CH…O) HBs, and H…H interactions in these clusters. Comparison with three different solvated alcohol systems featuring the same types of atom–atom interactions permitted the following order of stability to be determined: (methanol)5–water > (methanol)6 > (ethanol)5–water > (ethanol)6. These findings, together with accompanying geometric and spectroscopic analyses, show that similar cooperative effects exist among the primary HBs for structures with the same arrangement of primary HBs, regardless of the nature of the molecules involved. This result provides an indication that the molecular ratio can be considered to determine the unusual behavior of the ethanol–water system. The investigation also highlights the presence of several types of weak interaction in addition to primary HBs.
Graphical Abstract Water-ethanol clusters exhibit a variety of interaction types between their atoms, such as primary OH...O (blue), secondary CH...O (green) and H...H (yellow) interactions as revealed by Quantum Chemical Topology
  相似文献   

17.
M Farnum  M Palcic  J P Klinman 《Biochemistry》1986,25(8):1898-1904
The pH dependence of steady-state parameters for [1,1-1H2]- and [1,1-2H2]benzylamine oxidation and of tritium exchange from [2-3H]dopamine has been measured in the bovine plasma amine oxidase reaction. Deuterium isotope effects on kcat/Km for benzylamine are observed to be constant, near the intrinsic value of 13.5, over the experimental pH range, indicating that C-H bond cleavage is fully rate limiting for this parameter. As a consequence, pKa values derived from kcat/Km profiles, 8.0 +/- 0.1 (pK1) and 9.0 +/- 0.16 (pKs), can be ascribed to microscopic pKa values for the ionization of an essential active site residue (EB1) and substrate, respectively. Profiles for kcat and Dkcat show that EB1 undergoes a perturbation from 8.0 to 5.6 +/- 0.3 (pK1') in the presence of substrate; additionally, a second ionization, pK2 = 7.25 +/- 0.25, is observed to mediate but not be essential for enzyme reoxidation. The pH dependence of the ratio of tritium exchange to product formation for dopamine also indicates base catalysis with a pKexch = 5.5 +/- 0.01, which is within experimental error of pK1'. We conclude that the data presented herein support a single residue catalyzing both substrate oxidation and exchange, consistent with recent stereochemical results that implicate a syn relationship between these processes [Farnum, M., & Klinman, J.P. (1985) Fed. Proc., Fed. Am. Soc. Exp. Biol. 44, 1055]. This conclusion contrasts with earlier kinetic data in support of a large rate differential for the exchange of hydrogen from C-1 vs. C-2 of phenethylamine derivatives [Palcic, M.M., & Klinman, J.P. (1983) Biochemistry 22, 5957-5966].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The anaerobic fermentation of glucose by Leuconostoc mesenteroides via the reductive pentose phosphate pathway leads to the accumulation of lactic acid and ethanol. The isotope redistribution coefficients (a(ij)) that characterize the specific derivation of each hydrogen atom in ethanol in relation to the non-exchangeable hydrogen atoms in glucose and the medium water have been determined using quantitative (2)H NMR. First, it is confirmed that the hydrogens of the methylene group are related only to the 1 and 3 positions of glucose via the NAD(P)H pool and not to the 4 position, in contrast to ethanol produced by Saccharomyces cerevisiae. Second, it is found that the conversion factors (C(f)) for the transfer of hydrogen to the pro-S and pro-R positions of the methylene group are not equivalent: the C(f)-1-R:C(f)-1-S ratio is 2.1, whereas the C(f)-3-R:C(f)-3-S ratio is 0.8. It is shown that this non-equivalence is not determined by the stereochemistry of the terminal NADH- and NADPH-dependent alcohol dehydrogenases, but is dependent on the cofactor selectivities of the reductive and oxidative steps of the reduced nucleotide cycle.  相似文献   

19.
20.
The Cu(II) and Zn(II) complexes of phenoxyl radical species [M(II)(L1*)(NO3)]+ (M=Cu or Zn, L1H: 2-methylthio-4-tert-butyl-6-[[bis[2-(2-pyridyl)ethyl]amino]methyl]phenol ) and [M(II)(L2*)(NO3)]+ (M=Cu or Zn, L2H: 2,4-di-tert-butyl-6-[[bis[2-(2-pyridyl)ethyl]amino]methyl]phenol) are prepared as model complexes of the active form of galactose oxidase (GAO). Hydrogen atom abstraction of 1,4-cyclohexadiene and tert-butyl substituted phenols by the GAO model complexes proceeds very efficiently to give benzene and the corresponding phenoxyl radical or its C-C coupling dimer as the oxidation products, respectively. Kinetic analyses on the oxidation reactions have shown that the hydrogen atom abstraction of the phenol substrates is significantly enhanced by the coordinative interaction of the OH group to the metal ion center of the complex, providing valuable insight into the enzymatic mechanism of the alcohol oxidation. Details of the substrate-activation process have been discussed based on the activation parameters (deltaH* and deltaS*) of the reactions.  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