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 15 毫秒
1.
The product distributions for the reactions of (R)-glyceraldehyde 3-phosphate (GAP) in D(2)O at pD 7.5-7.9 catalyzed by triosephosphate isomerase (TIM) from chicken and rabbit muscle were determined by (1)H NMR spectroscopy. Three products were observed from the reactions catalyzed by TIM: dihydroxyacetone phosphate (DHAP) from isomerization with intramolecular transfer of hydrogen (49% of the enzymatic products), [1(R)-(2)H]-DHAP from isomerization with incorporation of deuterium from D(2)O into C-1 of DHAP (31% of the enzymatic products), and [2(R)-(2)H]-GAP from incorporation of deuterium from D(2)O into C-2 of GAP (21% of the enzymatic products). The similar yields of [1(R)-(2)H]-DHAP and [2(R)-(2)H]-GAP from partitioning of the enzyme-bound enediol(ate) intermediate between hydron transfer to C-1 and C-2 is consistent with earlier results, which showed that there are similar barriers for conversion of this intermediate to the alpha-hydroxy ketone and aldehyde products (Knowles, J. R., and Albery, W. J. (1977) Acc. Chem. Res. 10, 105-111). However, the observation that the TIM-catalyzed isomerization of GAP in D(2)O proceeds with 49% intramolecular transfer of the (1)H label from substrate to product DHAP stands in sharp contrast with the 相似文献   

2.
A method is presented for measurement of triglyceride (TG) synthesis that can be applied to slow-turnover lipids. The glycerol moiety of TG is labeled from 2H2O, and mass isotopomer distribution analysis (MIDA) is applied. Mice and rats were given 4-8% 2H2O in drinking water; TG-glycerol was isolated from adipose and liver during < or =12-wk of 2H2O labeling. Mass isotopomer abundances in the glycerol moiety of TG were measured by GC-MS. The combinatorial pattern of isotopomers revealed the number of H atoms in glycerol incorporating label from 2H2O (n) to be 3.8-4.0 of a possible 5 for adipose tissue and 4.6-4.8 for liver TG. Hepatic TG-glycerol in fact reached 97% predicted maximal value of label incorporation (4.4-4.6 x body 2H2O enrichment), indicating near-complete replacement of the liver TG pool. Label incorporation into adipose tissue revealed turnover of mesenteric TG to be faster (k = 0.21 day-1) than other depots (k = 0.04-0.06 day-1) in mice. TG isolated from subcutaneous depots of growing adult rats plateaued at 85-90% of calculated maximal values at 12 wk (k = 0.05 day-1), excluding significant dilution by unlabeled alpha-glycerol phosphate. Turnover of plasma TG, modeled from 2H incorporation over 60 min, was 0.06 min-1 (half-life 11.5 min). In summary, use of 2H2O labeling with MIDA of TG-glycerol allows measurement of new alpha-glycerol phosphate-derived TG synthesis and turnover. The hypothesis that mesenteric TG is more lipolytically active than other depots, previously difficult to prove by isotope dilution techniques, was confirmed by this label incorporation approach.  相似文献   

3.
The reaction of H2O2 with 4-substituted aryl alkyl sulfides (4-XC6H4SR), catalysed by lignin peroxidase (LiP) from Phanerochaete chrysosporium, leads to the formation of sulfoxides, accompanied by diaryl disulfides. The yields of sulfoxide are greater than 95% when X = OMe, but decrease significantly as the electron donating power of the substituent decreases. No reaction is observed for X = CN. The bulkiness of the R group has very little influence on the efficiency of the reaction, except for R = tBu. The reaction exhibits enantioselectivity (up to 62% enantiomeric excess with X = Br, with preferential formation of the sulfoxide with S configuration). Enantioselectivity decreases with increasing electron density of the sulfide. Experiments in H218O show partial or no incorporation of the labelled oxygen into the sulfoxide, with the extent of incorporation decreasing as the ring substituents become more electron-withdrawing. On the basis of these results, it is suggested that LiP compound I (formed by reaction between the native enzyme and H2O2), reacts with the sulfide to form a sulfide radical cation and LiP compound II. The radical cation is then converted to sulfoxide either by reaction with the medium or by a reaction with compound II, the competition between these two pathways depending on the stability of the radical cation.  相似文献   

4.
The product distributions for the reactions of dihydroxyacetone phosphate (DHAP) in D(2)O at pD 7.9 catalyzed by triosephosphate isomerase (TIM) from chicken and rabbit muscle were determined by (1)H NMR spectroscopy using glyceraldehyde 3-phosphate dehydrogenase to trap the first-formed products of the thermodynamically unfavorable isomerization reaction, (R)-glyceraldehyde 3-phosphate (GAP) and [2(R)-(2)H]-GAP (d-GAP). Three products were observed from the reactions catalyzed by TIM: GAP from isomerization with intramolecular transfer of hydrogen (18% of the enzymatic products), d-GAP from isomerization with incorporation of deuterium from D(2)O into C-2 of GAP (43% of the enzymatic products), and [1(R)-(2)H]-DHAP (d-DHAP) from incorporation of deuterium from D(2)O into C-1 of DHAP (40% of the enzymatic products). The ratios of the yields of the deuterium-labeled products d-DHAP and d-GAP from partitioning of the intermediate of the TIM-catalyzed reactions of GAP and DHAP in D(2)O are 1.48 and 0.93, respectively. This provides evidence that the reaction of these two substrates does not proceed through a single, common, reaction intermediate but, rather, through distinct intermediates that differ in the bonding and arrangement of catalytic residues at the enediolate O-1 and O-2 oxyanions formed on deprotonation of GAP and DHAP, respectively.  相似文献   

5.
Reduction of NO and NO2-by whole cells of eight strains of denitrifying bacteria known to contain either heme cd1 or copper-containing nitrite reductases (NiRs) has been examined in the presence of H218O. All organisms containing heme cd1 NiRs exhibited relatively large extents of exchange between NO2- and H218O (39-100%), as monitored by the 18O content of product N2O. Organisms containing copper NiRs gave highly variable results, with Achromobacter cycloclastes and Pseudomonas aureofaciens exhibiting no 18O incorporation and Rhodopseudomonas sphaeroides and Alcaligenes entrophus exhibiting complete exchange between NO2- and H218O. Organisms containing heme cd1 NiRs exhibited significant but lower levels of exchange between NO and H218O than between NO2- and H218O, while organisms containing copper NiRs gave significantly higher amounts of 18O incorporation than observed for the heme cd1 organisms. These results demonstrate the existence of an NO-derived species capable of undergoing O-atom exchange with H218O during the reduction of NO. Trapping experiments with 15NO, 14N3-, and crude extracts of R. sphaeroides support the electrophilic nature of this intermediate and suggest its formulation as an enzyme nitrosyl, E-NO+, analogous to that observed during reduction of NO2-. The observation of lower levels of 18O incorporation with NO2- than with NO as substrate for A. cycloclastes and P. aureofaciens indicates that, for these organisms at least, a sequential pathway involving free NO as an intermediate is significantly less important than a direct pathway in which N2O is formed via reaction of two NO2- ions on a single enzyme.  相似文献   

6.
Deferriferrioxamine B (H3DFB) is a linear trihydroxamic acid siderophore with molecular formula NH2(CH2)5[N(OH)C(O)(CH2)2C(O)NH(CH2)5]2N(OH)C(O)CH3 that forms a kinetically and thermodynamically stable complex with iron(III), ferrioxamine B. Under the conditions of our study (pH = 4.30, 25 degrees C), ferrioxamine B, Fe(HDFB)+, is hexacoordinated and the terminal amine group is protonated. Addition of simple hydroxamic acids, R1C(O)N(OH)R2 (R1 = CH3, R2 = H; R1 = C6H5, R2 = H; R1 = R2 = CH3), to an aqueous solution of ferrioxamine B at pH = 4.30, 25.0 degrees C, I = 2.0, results in the formation of ternary complexes Fe(H2DFB)A+ and Fe(H3DFB)A2+, and tris complexes FeA3, where A- represents the bidendate hydroxamate anion R1C(O)N(O)R2-. The addition of a molar excess of ethylenediaminetetraacetic acid (EDTA) to an aqueous solution of ferrioxamine B at pH 4.30 results in a slow exchange of iron(III) to eventually completely form Fe(EDTA)- and H4DFB+. The addition of a hydroxamic acid, HA, catalyzes the rate of this iron exchange reaction: (formula; see text) A four parallel path mechanism is proposed for reaction (1) in which catalysis occurs via transient formation of the ternary and tris complexes Fe(H2DFB) A+, Fe(H3DFB)A2+, and FeA3. Rate and equilibrium constants for the various reaction paths to products were obtained and the influence of hydroxamic acid structure on catalytic efficiency is discussed. The importance of a low energy pathway for iron dissociation from a siderophore complex in influencing microbial iron bio-availability is discussed. The system represented by reaction (1) is proposed as a possible model for in vivo catalyzed release of iron from its siderophore complex at the cell wall or interior, where EDTA represents the intracellular storage depot or membrane-bound carrier and HA represents a low molecular weight hydroxamate-based metabolite capable of catalyzing interligand iron exchange.  相似文献   

7.
The incorporation of alpha-aminoisobutyryl (Aib) residues into peptide sequences facilitates helical folding. Aib-containing sequences have been chosen for the design of rigid helical segments in a modular approach to the construction of a synthetic protein mimic. The helical conformation of the synthetic peptide Boc-Aib-(Val-Ala-Leu-Aib)3-OMe in crystals is established by X-ray diffraction. The 13-residue apolar peptide adopts a helical form in the crystal with seven alpha-type hydrogen bonds in the middle and 3(10)-type hydrogen bonds at either end. The helices stack in columns, zigzag rather than linear, by means of direct NH...OC head to tail hydrogen bonds. Leucyl side chains are extended on one side of the helix and valyl side chains on the other side. Water molecules form hydrogen bonds with several backbone carbonyl oxygens that also participate in alpha-helix hydrogen bonds. There is no apparent distortion of the helix caused by hydration. The space group is P2(1)2(1)2(1), with a = 9.964 (3) A, b = 20.117 (3) A, c = 39.311 (6) A, Z = 4, and dx = 1.127 g/cm3 for C64H106N13O16.1.33H2O. The final agreement factor R was 0.089 for 3667 data observed greater than 3 sigma(F) with a resolution of 0.9 A.  相似文献   

8.
The design, synthesis, characterization and self-assembling properties of a new class of amphiphilic peptides, constructed from a bifunctional polar core attached to totally hydrophobic arms, are presented. The first series of this class, represented by the general structure Py(Aibn)2 (Py=2,6-pyridine dicarbonyl unit; Aib=alpha, alpha'-dimethyl glycine; n=1-4), is prepared in a single step by the condensation of commercially available 2,6-pyridine dicarbonyl dichloride with the methyl ester of homo oligoAib peptide (Aibn-OMe) in the presence of triethyl amine. 1H NMR VT and ROESY studies indicated the presence of a common structural feature of 2-fold symmetry and an NH...N hydrogen bond for all the members. Whereas the Aib3 segment in Py(Aib3)2 showed only the onset of a 3(10)-helical structure, the presence of a well-formed 3(10)-helix in both Aib4 arms of Py(Aib4)2 was evident in the 1H NMR of the bispeptide. X-ray crystallographic studies have shown that in the solid state, whereas Py(Aib2)2 molecules organize into a sheet-like structure and Py(Aib3)2 molecules form a double-stranded string assembly, the tetra Aib bispeptide, Py(Aib4)2, is organized to form a tetrameric assembly which in turn extends into a continuous channel-like structure. The channel is totally hydrophobic in the interior and can selectively encapsulate lipophilic ester (CH3COOR, R=C2H5, C5H11) molecules, as shown by the crystal structures of the encapsulating channel. The crystal structure parameters are: 1b, Py(Aib2)2, C25H37N5O8, sp. gr. P2(1)2(1)2(1), a=9.170(1) A, b=16.215(2) A, c=20.091(3) A, R=4.80; 1c, Py(Aib3)2, C33H51N7O10H2O, sp. gr. P1, a=11.040(1) A, b=12.367(1) A, c=16.959(1) A, alpha =102.41 degrees, beta =97.29 degrees, gamma =110.83 degrees, R1=6.94; 1 da, Py(Aib4)2.et ac, C41H65N9O12.1.5H2O.C4H8O2, sp. gr. P1, a=16.064(4) A, b=16.156 A, c=21.655(5) A, alpha =90.14(1)degrees, beta=101.38(2) degrees, gamma=97.07(1)degrees, Z=4, R1=9.03; 1db, Py(Aib4)2.amylac, C41H65N9O12.H2O.C7H14O2, P2(1)/c, a=16.890(1) A, b=17.523(1)A, c=20.411(1) A, beta=98.18 degrees, Z=4, R=11.1 (with disorder).  相似文献   

9.
S Ghisla  C Thorpe  V Massey 《Biochemistry》1984,23(14):3154-3161
Butyryl-CoA dehydrogenase from Megasphera elsdenii catalyzes the exchange of the alpha- and beta-hydrogens of substrate with solvent [Gomes, B., Fendrich, G., & Abeles, R. H. (1981) Biochemistry 20, 1481-1490]. The stoichiometry of this exchange was determined by using 3H2O label as 1.94 +/- 0.1 per substrate molecule. The rate of 3H label incorporation into substrate under anaerobic conditions is monophasic, indicating that both the alpha- and beta-hydrogens exchange at the same rate. The exchange in 2H2O leads to incorporation of one 2H each into the alpha- and the beta-positions of butyryl-CoA, as determined by companion 1H NMR experiments and confirmed by mass spectroscopic analysis. In contrast, with general acyl-CoA dehydrogenase from pig kidney, only exchange of the alpha-hydrogen was found. The beta-hydrogen is the one that is transferred (reversibly) to the flavin 5-position during substrate dehydrogenation. This was demonstrated by reacting 5-3H- and 5-2H-reduced 5-deaza-FAD-general acyl-CoA dehydrogenase with crotonyl-CoA. Only one face of the reduced flavin analogue is capable of transferring hydrogen to substrate. The rate of this reaction is 11.1 s-1 for 5-deaza-FAD-enzyme and 2.2 s-1 for [5-2H]deaza-FAD-enzyme, yielding an isotope effect of 5. These values compare with a rate of 2.6 s-1 for the reaction of native reduced enzyme with crotonyl-CoA. The two reduced enzymes (normal vs. 5-deaza-FAD-enzyme) thus react at similar rates, indicating a similar mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The vitamin K dependent carboxylase of liver microsomes is involved in the posttranslational modification of certain serine protease zymogens which are critical components of the blood clotting cascade. During coupled carboxylation/oxygenation this carboxylase converts glutamate residues, dihydrovitamin K, CO2, and O2 to a gamma-carboxyglutamyl (Gla) residue, vitamin K (2R,3S)-epoxide, and H2O with a stoichiometry of 1:1 for all substrates and products. In this paper we investigate the role of molecular oxygen in the reaction by following the course of the oxygen atoms using 18O2. Two different mass spectroscopic techniques, electron ionization positive ion mass spectrometry and supercritical fluid chromatography-negative ion chemical ionization mass spectrometry, were used to quantitate the amount of 18O incorporation into the various oxygens of the vitamin K epoxide product. We found that 0.95 mol atoms of oxygen were incorporated into the epoxide oxygen, 0.05 mol atoms of oxygen were incorporated into the quinone oxygen of vitamin K epoxide, and the remaining ca. 1.0 mol atoms of oxygen were incorporated into H2O. No incorporation of oxygen into vitamin K epoxide from 50% H2(18)O was observed. Thus, the carboxylase operates as a dioxygenase 5% of the time during carboxylation/oxygenation. The relevance of these findings with respect to the nonenzymic "basicity enhancement" model proposed by Ham and Dowd [(1990) J. Am. Chem. Soc. 112, 1660-1661] is discussed.  相似文献   

11.
Optimizations of cell productivity and utility of gaseous substrates were investigated for Alcaligenes hydrogenophilus. Theoretical analyses for the cell productivity and substrate utility were discussed in the light of experimental data. It was proved from the theoretical analyses that the maximum cell productivity was at the following conditions: D = 0.21 h(-1), R(1) ( = H(2) partial pressure/O(2) partial pressure) = 2.25 and R(2) ( =CO(2) partial pressure/O(2) partial pressure) = 0.09, i.e., a molal ratio, H(2):O(2):CO(2) = 67.4:29.9:2,7. The Condition which gave the maximum cell productivity also gave high utilities of gaseous substrates. The growth yield for oxygen increased with increasing R(1) in chemostat culture, and the growth yield for hydrogen slightly increased with increasing R(1). Theoretical cell productivity and substrate and substrate utility were in good agreement with experimental values, taking into account the change of growth yield for oxygen.  相似文献   

12.
This study investigated the effects of intensity and duration of exercise on lymphocyte proliferation as a measure of immunologic function in men of defined fitness. Three fitness groups--low [maximal O2 uptake (VO2max) = 44.9 +/- 1.5 ml O2.kg-1.min-1 and sedentary], moderate (VO2max = 55.2 +/- 1.6 ml O2.kg-1.min-1 and recreationally active), and high (VO2max = 63.3 +/- 1.8 ml O2.kg-1.min-1 and endurance trained)--and a mixed control group (VO2max = 52.4 +/- 2.3 ml O2.kg-1.min-1) participated in the study. Subjects completed four randomly ordered cycle ergometer rides: ride 1, 30 min at 65% VO2max; ride 2, 60 min at 30% VO2max; ride 3, 60 min at 75% VO2max; and ride 4, 120 min at 65% VO2max. Blood samples were obtained at various times before and after the exercise sessions. Lymphocyte responses to the T cell mitogen concanavalin A were determined at each sample time through the incorporation of radiolabeled thymidine [( 3H]TdR). Despite differences in resting levels of [3H]TdR uptake, a consistent depression in mitogenesis was present 2 h after an exercise bout in all fitness groups. The magnitude of the reduction in T cell mitogenesis was not affected by an increase in exercise duration. A trend toward greater reduction was present in the highly fit group when exercise intensity was increased. The reduction in lymphocyte proliferation to the concanavalin A mitogen after exercise was a short-term phenomenon with recovery to resting (preexercise) values 24 h after cessation of the work bout. These data suggest that single sessions of submaximal exercise transiently reduce lymphocyte function in men and that this effect occurs irrespective of subject fitness level.  相似文献   

13.
B A Amato  S J Benkovic 《Biochemistry》1975,14(22):4877-4887
Cyclic N,N'-diphenylethylenephosphorodiamidates, 1, have been found to yield up to 70% of the N-formyldiphenylethylenediamines when reacted in 1 M formate buffers (50 degrees, 25% v/v dioxane-H2O, mu = 1.0). The mechanism proposed for formylation on the basis of kinetic, product ratio, and 18O-labeling studies apparently involves cyclization of a formyl phosphoramidate-amine adduct derived from formate attack on 1H+, the process probably being catalyzed by intramolecular proton transfer between the phosphoryl moiety and aniline nitrogen. An oxygen is transferred from formate to inorganic phosphate in the net reaction. The initial site of formylation with dissymmetric 1 (p-CH3O, p-Cl) was shown by 18O incorporation studies to be the p-anisidine nitrogen, which also corresponds to the isomer of thermodynamic control. Attack by formate on 1H+, therefore, results in expulsion of the more basic nitrogen. Certain mechanistic aspects of the reaction might occur in the enzyme-catalyzed ATP-dependent formylation of tetrahydrofolic acid.  相似文献   

14.
A study is presented on the pH dependence of proton translocation in the oxidative and reductive phases of the catalytic cycle of purified cytochrome c oxidase (COX) from beef heart reconstituted in phospholipid vesicles (COV). Protons were shown to be released from COV both in the oxidative and reductive phases. In the oxidation by O2 of the fully reduced oxidase, the H+/COX ratio for proton release from COV (R --> O transition) decreased from approximately 2.4 at pH 6.5 to approximately 1.8 at pH 8.5. In the direct reduction of the fully oxidized enzyme (O --> R transition), the H+/COX ratio for proton release from COV increased from approximately 0.3 at pH 6.5 to approximately 1.6 at pH 8.5. Anaerobic oxidation by ferricyanide of the fully reduced oxidase, reconstituted in COV or in the soluble case, resulted in H+ release which exhibited, in both cases, an H+/COX ratio of 1.7-1.9 in the pH range 6.5-8.5. This H+ release associated with ferricyanide oxidation of the oxidase, in the absence of oxygen, originates evidently from deprotonation of acidic groups in the enzyme cooperatively linked to the redox state of the metal centers (redox Bohr protons). The additional H+ release (O2 versus ferricyanide oxidation) approaching 1 H+/COX at pH < or = 6.5 is associated with the reduction of O2 by the reduced metal centers. At pH > or = 8.5, this additional proton release takes place in the reductive phase of the catalytic cycle of the oxidase. The H+/COX ratio for proton release from COV in the overall catalytic cycle, oxidation by O2 of the fully reduced oxidase directly followed by re-reduction (R --> O --> R transition), exhibited a bell-shaped pH dependence approaching 4 at pH 7.2. A mechanism for the involvement in the proton pump of the oxidase of H+/e- cooperative coupling at the metal centers (redox Bohr effects) and protonmotive steps of reduction of O2 to H2O is presented.  相似文献   

15.
The 1H NMR relaxation effects produced by paramagnetic Cr(III) complexes on nucleoside 5'-mono- and -triphosphates in D2O solution at pH' = 3 were measured. The paramagnetic probes were [Cr(III)(H2O)6]3+, [Cr(III)(H2O)3(HATP)], [Cr(III)(H2O)3(HCTP)] and [Cr(III)(H2O)3(UTP)-, while the matrix nucleotides (0.1 M) were H2AMP, HIMP-, and H2ATP2-. For the aromatic base protons, the ratios of the transverse to longitudinal paramagnetic relaxation rates (R2p/R1p) for the [Cr(III)(H2O)6]3+/H2ATP2-, [Cr(III)(H2O)3(HATP)]/H2ATP2-, [Cr(III)(H2O)3(HCTP)]/H2ATP2 and [Cr(III)(H2O)3(UTP)]-/H2ATP2 systems were below 2.33 so the dipolar term predominates. For a given nucleotide, R1p for the purine H(8) signal was larger than for the H(2) signal with the [Cr(III)(H2O)6]3+ probe, while R1p for the H(2) signal was larger with all the other Cr(III) probes. Molecular mechanics computations on the [Cr(III)(H2O)4(HPP)(alpha,beta)], [Cr(III)(NH3)4(HPP)(alpha,beta)], [Co(III)(NH3)3(H2PPP)(alpha,beta,gamma)] and [Co(III)(NH3)4(HPP)(alpha,beta)] complexes gave calculated energy-minimized geometries in good agreement with those reported in crystal structures. The molecular mechanics force constants found were then used to calculate the geometry of the inner sphere [Cr(III)(H2O)6]3+ and [Cr(III)(H2O)3(HATP)(alpha,beta,gamma)] complexes as well as the structures of the outer sphere [Cr(III)(H2O)6]3(+)-(H2AMP) and [Cr(III)(H2O)6]-(HIMP)- species. The gas-phase structure of the [Cr(III)(H2O)3(HATP)(alpha,beta,gamma)] complex shows the existence of a hydrogen bond interaction between a water ligand and the adenine N(7)(O...N = 2.82 A). The structure is also stabilized by intramolecular hydrogen bonds involving the -O(2')H group and the adenine N(3) (O...N = 2.80 A) as well as phosphate oxygen atoms and a water molecule (O...O = 2.47 A). The metal center has an almost regular octahedral coordination geometry. The structures of the two outer-sphere species reveal that the phosphate group interacts strongly with the hexa-aquochromium probe. In both complexes, the nucleotides have a similar "anti" conformation around the N(9)-C(1') glycosidic bond. However, a very important difference characterizes the two structures. For the (HIMP)- complex, strong hydrogen bond interactions exist between one and two water ligands and the inosine N(7) and O(6) atoms, respectively (O...O = 2.63 A; O...N = 2.72, 2.70 A). For the H2AMP complex, the [Cr(III)(H2O)6]3+ cation does not interact with N(7) since it is far from the purine system. Hydrogen bonds occur between water ligands and phosphate oxygens. The Cr-H(8) and Cr-H(2) distances revealed by the energy-minimized geometries for the two outer sphere species were used to calculate the R1p values for the H(8) and H(2) signals for comparison with the observed R1p values: 0.92(c), 1.04(ob) (H(8)) and 0.06(c), 0.35(ob) (H(2)) for H2AMP; and 3.76(c), 4.53(ob) (H(8)) and 0.16(c), 0.77(ob) s-1 (H(2)) for HIMP-.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
The reaction of copper(II) complexes supported by a series of beta-diketiminate ligands ((R1,R2)L, [(Dipp)N-C(R(2))-C(R(1))-C(R(2))-N(Dipp)](-), Dipp=2,6-diisopropylphenyl; see ) and H(2)O(2) has been examined spectroscopically at a low temperature. The beta-diketiminatocopper(II) complexes with R(2)=H (no substituent on the beta-carbon) provided a copper-oxygen intermediate that exhibited the same spectroscopic features as those of the bis(mu-oxo)dicopper(III) complex generated by the reaction of corresponding beta-diketiminatocopper(I) complex and O(2). On the other hand, the beta-diketiminatocopper(II) complexes with methyl substituent on the beta-carbon (R(2)=Me) did not produce such an intermediate in the same reaction. The beta-diketiminatocopper(II) complexes carrying an electron-withdrawing substituent on the alpha-carbon (R(1)=NO(2) or CN) but no beta-substituent (R(2)=H) exhibited a high catalytic activity in the oxygenation reaction of alkanes with H(2)O(2). Mechanism of the catalytic oxygenation reaction as well as the substituent effects of the ligands on the copper(II)-H(2)O(2) reactivity is discussed.  相似文献   

17.
Three crystalline polymorphs of the helical decapeptide, Boc-Aib-Ala-Leu-Ala-Leu-Aib-Leu-Ala-Leu-Aib-OMe, have been obtained. Antiparallel helix aggregation is observed in crystals grown from methanol (A), while completely parallel packing is observed in crystals from isopropanol (B) or an ethylene glycol-ethanol mixture (C). Crystals B and C are very similar in molecular conformation and packing. The packing motifs in crystals A and B consist of rows of parallel molecules, with an almost identical arrangement in both crystals. In crystal A, adjacent rows assemble with the helix axes pointed in opposite directions, whereas in crystal B all rows assemble with helix axes pointed in the same direction. Electrostatic interactions between helix dipoles do not appear to be a major determinant of packing modes. The structures also do not provide a ready rationalization of packing preferences in terms of side-chain interactions or solvation. The alpha-helix of the peptide in crystal A has seven 5----1 hydrogen bonds; the helix in crystal B is a mixed 3(10)/alpha-helix. The crystal parameters are as follows. Crystal A: C51H92N10O13.CH3OH, space group P2(1) with a = 10.498 (1) A, b = 18.189 (3) A, c = 16.475 (3) A, beta = 99.28 (1) degree, Z = 2, R = 9.6% for 1860 data. Crystal B: C51H92N10O13.C3H7OH, space group P2(1) with a = 10.534 (1) A, b = 28.571 (4) A, c = 11.055 (2) A, beta = 95.74 (1) degree, Z = 2, R = 6.5% for 3251 data. Crystal C: C51H92N10O13.C2H5OH, space group P2(1), with a = 10.450 (1) A, b = 28.442 (5) A, c = 11.020 (2) A, beta = 95.44(1) degree, Z = 2, R = 14.8% (isotropic) for 1948 data.  相似文献   

18.
Totah RA  Hanzlik RP 《Biochemistry》2004,43(24):7907-7914
Horseradish peroxidase (HRP) typically oxidizes aniline derivatives using hydrogen peroxide as the oxidant. The action of HRP on N-alkyl-N-phenylglycine derivatives 1b-1e (PhN(R)CH(2)COOH; R = Me, Et, n-Pr, i-Pr, respectively) is highly unusual if not unique. Under standard peroxidatic conditions (HRP/H(2)O(2)/air), the major product (ca. 70%) is the secondary aniline 2b-2e (PhNHR) resulting from the expected oxidative decarboxylation process, but a significant amount (ca. 30%) of the related tertiary aniline PhN(CH(3))R (3b-3e) arises from an unexpected nonoxidative decarboxylation process. Under anaerobic, peroxide-free conditions only the tertiary anilines 3b-3e are formed in a reaction that is extremely rapid compared to those in which H(2)O(2), molecular oxygen, or both are present. In D(2)O buffers, the product is exclusively the monodeutero tertiary aniline PhN(CH(2)D)R and the reaction is much slower (k(H(2)O)/k(D(2)O) = 5.7), suggesting that a proton transfer step is substantially rate-limiting in turnover. It is proposed that ferric HRP oxidizes 1 to a cation radical, which then decarboxylates to an alpha-amino radical having carbanion character on carbon; protonation of the latter, followed by electron capture from ferrous HRP, completes the cycle. Hydrogen peroxide and oxygen slow turnover by diverting ferric HRP toward the compound I/compound II forms or toward compound III, respectively. Finally, under peroxidatic conditions, 1a (R = cyclopropyl) inactivates HRP with concurrent formation of 2a but not N-phenylglycine, but under anaerobic, peroxide-free conditions, 1a inactivates HRP almost instantly with no detectable product formation.  相似文献   

19.
Ammonium D-glucarate, NH(4)(C(6)H(9)O(8)) [ammonium D-saccharate, NH(4)-SAC], has been synthesized, and its crystal structure solved by single-crystal X-ray diffraction methods. NH(4)-SAC crystallizes in the monoclinic space group P2(1) (#4) with cell parameters a = 4.8350(4) Angstroms, b = 11.0477(8) Angstroms, c = 16.7268(12) Angstroms, beta = 90.973(1) degrees, V = 894.34(12) Angstroms(3), Z = 3. The structure was refined by full-matrix least-squares on F(2) yielding final R-values (all data) R1 = 0.0353 and R(w)2 = 0.0870. The structure consists of alternating (NH(4))(+) and (C(6)H(11)O(6))(-) layers parallel to the bc plane. An extended network of N-H...O(SAC) and O(SAC)-H...O(SAC) hydrogen bonds provide the 3-D connectivity. The aqueous solubility (S(w)) has been shown to be pH independent at ambient conditions within the range 4.5 < pH < 10 with S(w) = 2.19 M/L, whose value is about a factor of two lower than that of the ammonium isosaccharate analogue.  相似文献   

20.
The oxidation of reduced cytochrome c oxidase by hydrogen peroxide was investigated with stopped-flow methods. It was reported by us previously (A.C.F. Gorren, H. Dekker and R. Wever (1986) Biochim. Biophys. Acta 852, 81-92) that at low H2O2 concentrations cytochrome a is oxidised simultaneously with cytochrome a3, but that at higher H2O2 concentrations the oxidation of cytochrome a is slower than that of cytochrome a3. We now report that for high peroxide concentrations (10-45 mM) the oxidation rate of cytochrome a increased linearly with the concentration of H2O2 (k = 700 M-1.S-1). Upon extrapolation to zero H2O2 concentration an intercept with a value of 16 s-1 (at 20 degrees C and pH 7.4) was found. A reaction sequence is described to explain these results; according to this model the rate constant (16 S-1) at zero H2O2 concentration represents the true value of the rate of electron transfer from cytochrome a to cytochrome a3 when the a3-CuB site is oxidised and unligated. However, when a complex of hydrogen peroxide with oxidised cytochrome a3 is formed, this rate is strongly enhanced. The slope (700 M-1.S-1) would then represent the rate of cytochrome a3(3+)-H2O2 complex formation. From experiments in which the pH was varied, we conclude that the reaction of H2O2 with cytochrome a3(2+) is independent of pH, whereas the electron-transfer rate from cytochrome a to cytochrome a3 gradually decreases with increasing pH. From the temperature dependence we could calculate values of 23 kJ.mol-1 and 45 kJ.mol-1 for the activation energies of the oxidations by H2O2 of cytochrome a3(2+) and cytochrome a2+, respectively. The similarity of the values that were obtained for cytochrome a oxidation both with H2O2 and with O2 as the electron acceptor suggests that the reactions share the same mechanism. In 2H2O the reactions studied decreased in rate. For the reaction of 2H2O2 with reduced cytochrome a3 in 2H2O, a small effect was found (15% decrease in rate constant). However, the internal electron-transfer rate from cytochrome a to cytochrome a3 decreased by 50%, Our results suggest that the internal electron transfer is associated with proton translocation.  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