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 560 毫秒
1.
Ye Y  Liu M  Kao JL  Marshall GR 《Biopolymers》2008,89(1):72-85
Phosphinic compounds have potential as amide-bond mimetics in the development of novel peptidomimetics, enzyme inhibitors, and metal-binding ligands. Novel pseudo-oligopeptides with two phosphinic acid groups embedded in the peptide backbone serving as amide-bond surrogates, Psi[P(O,OH)--CH(2)], were targeted. A series of linear and cyclic pseudo-oligopeptides with two phosphinic acid groups arrayed at different positions in the peptide sequence were designed, including Ac--Phe--{(R,S)--AlaPsi[P(O,OH)--CH(2)]Gly}(2)--NH(2) (P2), Ac--NH--(R,S)--AlaPsi[P(O,OH)--CH(2)]Gly--Phe--(R,S)--AlaPsi[P(O,OH)--CH(2)]Gly--NH(2) (P3), Ac--NH--(R,S)--AlaPsi[P(O,OH)--CH(2)]Gly--Phe--Phe--(R,S) --AlaPsi[P(O,OH)--CH(2)]Gly--NH(2) (P4), cyclo{NH--(R,S)--AlaPsi[P(O,OH)--CH(2)]Gly--Phe}(2) (P5), and cyclo[NH--(R,S)--AlaPsi[P(O,OH)--CH(2)]Gly--Phe--Phe](2) (P6). They were synthesized via conventional Fmoc chemistry on solid support utilizing Fmoc-protected phosphinic acid-containing pseudo-dipeptide fragment, i.e. Fmoc--(R,S)--AlaPsi[P(O,OCH(3))--CH(2)]Gly--OH. The pseudo-peptides containing two phosphinic acid groups exhibited the highest binding affinity and selectivity for Fe(III) among the 10-metal ions screened by ESI-MS analysis--Cu(II), Zn(II), Co(II), Ni(II), Mn(II), Fe(II), Fe(III), Al(III), Ga(III), and Gd(III). P4 and P6 with 11-atom linkages between the two phosphinic acids preferred intramolecular metal binding to form 1:1 ligand/metal complexes. As revealed by competition experiments, P4 showed the highest relative binding affinity among the six compounds tested. Noteworthy, P4 also showed higher relative binding affinity than similar dihydroxamate-containing pseudo-peptides reported previously. The novel structural prototype and facile synthesis along with selective and potent Fe(III) binding strongly suggest that pseudo-peptides containing the two or more phosphinic groups as amide-bond surrogates deserve further exploration in medicinal chemistry.  相似文献   

2.
A number of earlier unknown phosphonate analogues of aspartyl adenylate with anhydride oxygen substituted by --CH2--, and the carbonyl group substituted by --CH(OH)- or --CH(NH2)-groups were synthesized. These compounds were used to study the reaction mechanism of asparagine synthetases from white lupine and E. coli. The aspartyl adenylate analogues proved to be powerful competitive inhibitors (Ki = 10(-7) M) of the bacterial enzyme. In the case of white lupine enzyme catalyzing the aspartate-independent ATP--[32P]PPi exchange, the above compounds displayed a non-competitive type of inhibition with respect to aspartate and ATP, Ki = 10(-4) M. It is likely that for the latter enzyme the first intermediate is different from an aspartyl adenylate derivative.  相似文献   

3.
A series of carboxy-alkylamidated and N-acetylated amino acids and peptides were synthesized and examined for their ability to inhibit human leukocyte elastase. The Boc-amino acid alkylamides were found to be potent specific and competitive inhibitors of this enzyme. They were found not to or only poorly inhibit several other serine proteinases such as bovine trypsin, alpha-chymotrypsin, porcine pancreatic elastase and human leukocyte cathepsin G at concentrations less than 10(-4) M. Specificity and maximum inhibition of human leukocyte elastase were achieved when the N-terminus of the amino acid was protected by a t-butyloxy-carbonyl (Boc) group, the oligopeptide fragment consisted of valine residues and when the alkyl chain was between 10 and 12 carbon atoms in length and attached to the C-terminus of the peptide fragment. Highest inhibition was obtained with the compound Boc-[Val]3-NH[CH2]11--CH3 (Ki = 0.21 microM). These specific inhibitors were also found to be non-toxic after oral administration to mice and rats (LD50 greater than 3.0 g/kg body weight).  相似文献   

4.
B Gomes  G Fendrich  R H Abeles 《Biochemistry》1981,20(6):1481-1490
Glutaryl-CoA dehydrogenase, a flavoprotein, catalyzes the reaction -OOCCH3CH2--CH2COSR (FAD leads to FADH2) leads to CH3CH = CHCOSR + CO2 (SR = CoA or pantetheine). With the isolated enzyme, a dye serves as the final electron acceptor. The enzyme from Pseudomonas fluorescens (ATCC 11250) has been purified to homogeneity. It was established with appropriate isotopic substitutions that the proton which is added to the gamma position of the product, subsequent to decarboxylation, is not derived from the solvent but is derived from the alpha position of the substrate. Under conditions where no net conversion of substrate occurs, i.e., in the absence of electron acceptor, the enzyme catalyzes the exchange of the beta hydrogen of the substrate with solvent protons. Butyryl-CoA dehydrogenase (M. elsedenii), which catalyzes an analogous reaction, catalyzes the exchange of both the alpha and beta hydrogens with solvent protons in the absence of electron acceptor. Glutaryl-CoA dehydrogenase and butyryl-CoA dehydrogenase are irreversibly inactivated by the substrate analogues 3-butynoylpantetheine and 3-pentynoylpantetheine. These inactivators do not form an adduct with the flavin and probably react with a nucleophile at the active site. Upon inactivation, the spectrum of the enzyme-bound flavin is essentially unchanged, and the flavin can be reduced by Na2S2O4. We suggest that inactivation involves intermediate allene formation. We proposed that these results support an oxidation mechanism for glutaryl-CoA dehydrogenase and butyryl-CoA dehydrogenase which is initiated by proton abstraction. With glutaryl-CoA dehydrogenase, the base, which abstracts the substrate alpha proton, is shielded from the solvent and is then used to protonate the carbanion (CH2--CH--CHCOSCoA) formed after oxidation and decarboxylation.  相似文献   

5.
The X-ray structure analysis of three compounds of interest as enzyme substrates is reported. They are the hydrated forms of (I) DL-2-amino-4-arsonobutanoic acid [HO-AsO2--CH2-CH2-CH(NH3+)-CO2H], (II) DL-2-amino-4-phosphonobutanoic acid [HO-PO2--CH2-CH2-CH(NH3+)-CO2H] and the hydrated barium salt of (III) D-3-phosphoglycerate [HO-PO2--O-CH2-CH(OH)-CO2-]. The structures were fully refined to R factors of 0.033, 0.053 and 0.046. For the compounds (I) and (II) the charge distribution was directly determined by locating all H atoms. The co-ordination around As and P is approximately tetrahedral, with the valency angle between the two charged O atoms enlarged to 112 degrees in compound (I), 166 degrees in compound (II) and 122 degrees in compound (III). The As-X bond distances are increased relative to P-X to accommodate the increased atomic radius. The analysis establishes that the compounds are structural analogues. Tables of co-ordinates for H atoms, anisotropic thermal parameters, bond lengths and bond angles for the three compounds have been deposited as Supplementary Publication SUP 50122 (5 pages) with the British Library Lending Division, Boston Spa, Wetherby, West Yorkshire LS23 7BQ, U.K., from whom copies can be obtained directly [see Biochem J. (1983) 209, 5].  相似文献   

6.
Capillary zone electrophoresis (CZE) was used for determination of dissociation constant of phosphinate group in phosphinic pseudopeptides, i.e. peptides where one peptide bond is substituted by phosphinic acid moiety -PO2--CH2-. The dissociation constants were determined for a set of newly synthesized pseudopeptides derived from a structure N-Ac-Val-Ala(psi)(PO2--CH2)Leu-His-NH2 by nonlinear regression of experimentally measured pH dependence of their effective electrophoretic mobilities. CZE experiments were carried out in Tris-phosphate background electrolytes in the pH range 1.4-3.2. The pseudopeptides were synthesized as a mixture of four diastereomers, the separation of which was achieved in most cases. Moreover, differences of the effective mobilities of the pseudopeptide diastereomers enabled simultaneous determination of the dissociation constant of their phosphinate group without necessity of previous isolation of individual isomers.  相似文献   

7.
Cyclic analogues of substance P of the formula cyclo-[Glu-Phe-Phe-Gly-Leu-Met-NH(CH2)nNH-], where n = 3-10, 12, and open-chain analogues (XVIIIa, b) H-Glu.(NHR)-Phe-Phe-Gly-Leu-Met-NHR, where R = -CH3, -CH2CH2CH3, were synthesized. By NMR spectroscopy it was found that cyclo-compounds with n = 3-8 have regularly arranged structures, stabilized by intramolecular hydrogen bonds. Substances of this type showed less than or equal to 0.1% of the substance P activity on the guinea pig ileum, but some of them antagonize the natural peptide (for compound with n = 5 IC50 = 3.2.10(-6) M). The open-chain compounds proved to have rather high myotropic activity, viz., 22% (R = -CH3) and 8% (R = -CH2CH2CH3) of the substance P activity.  相似文献   

8.
The organometallic precursor (NEt(4))(2)[ReBr(3)(CO)(3)] was reacted with bidendate dithioethers (L) of the general formula H(3)C-S-CH(2)CH(2)-S-R (R = -CH(2)CH(2)COOH, CH(2)-C&tbd1;CH) and R'-S-CH(2)CH(2)-S-R' (R' = CH(3)CH(2)-, CH(3)CH(2)-OH, and CH(2)COOH) in methanol to form stable rhenium(I) tricarbonyl complexes of the general composition [ReBr(CO)(3)L]. Under these conditions, the functional groups do not participate in the coordination. As a prototypic representative of this type of Re compounds, the propargylic group bearing complex [ReBr(CO(3))(H(3)C-S-CH(2)CH(2)-S-CH(2)C&tbd1;CH)] Re2 was studied by X-ray diffraction analysis. Its molecular structure exhibits a slightly distorted octahedron with facial coordination of the carbonyl ligands. The potentially tetradentate ligand HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH was reacted with the trinitrato precursor [Re(NO(3))(3)(CO)(3)](2-) to yield a cationic complex [Re(CO)(3)(HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH)]NO(3) Re8 which shows the coordination of one hydroxy group. Re8 has been characterized by correct elemental analysis, infrared spectroscopy, capillary electrophoresis, and X-ray diffraction analysis. Ligand exchange reaction of the carboxylic group bearing ligands H(3)C-S-CH(2)CH(2)-S-CH(2)CH(2)-COOH and HOOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH with (NEt(4))(2)[ReBr(3)(CO)(3)] in water and with equimolar amounts of NaOH led to complexes in which the bromide is replaced by the carboxylic group. The X-ray structure analysis of the complex [Re(CO)(3)(OOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH)] Re6 shows the second carboxylic group noncoordinated offering an ideal site for functionalization or coupling a biomolecule. The no-carrier-added preparation of the analogous (99m)Tc(I) carbonyl thioether complexes could be performed using the precursor fac-[(99m)Tc(H(2)O)(3)(CO)(3)](+), with yields up to 90%. The behavior of the chlorine containing (99m)Tc complex [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 in aqueous solution at physiological pH value was investigated. In saline, the chromatographically separated compound was stable for at least 120 min. However, in chloride-free aqueous solution, a water-coordinated cationic species Tc1a of the proposed composition [(99m)Tc(H(2)O)(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))](+) occurred. The cationic charge of the conversion product was confirmed by capillary electrophoresis. By the introduction of a carboxylic group into the thioether ligand as a third donor group, the conversion could be suppressed and thus the neutrality of the complex preserved. Biodistribution studies in the rat demonstrated for the neutral complexes [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 and [(99m)TcCl(CO)(3)(CH(2)-S-CH(2)CH(2)-S-CH(2)-C&tbd1;CH)] Tc2 a significant initial brain uptake (1.03 +/- 0.25% and 0.78 +/- 0.08% ID/organ at 5 min. p.i.). Challenge experiments with glutathione clearly indicated that no transchelation reaction occurs in vivo.  相似文献   

9.
After incorporation of purified sheep cardiac Ca(2+)-release channels into planar phospholipid bilayers, we have investigated the blocking effects of a series of monovalent (CH3-(CH2)n-1-N+(CH3)3) and divalent ((CH3)3N(+)-(CH2)n-N+(CH3)3) trimethylammonium derivatives under voltage clamp conditions. All the compounds tested produce voltage-dependent block from the cytoplasmic face of the channel. With divalent (Qn) derivatives the effective valence of block decreases with increasing chain length, reaching a plateau with a chain length of n > or = 7. No decline in effective valence is observed with the monovalent (Un) derivatives. A plausible interpretation of this phenomena suggests that for the 90% of the voltage drop measured, the increase in length following the addition of a CH2 in the chain spans 12.7% of the electrical field. Extrapolating this distance to include the remaining 10% suggests that the applied holding potential falls over a total distance of 10.4 A. In addition, at high positive holding potentials there is evidence for permeation of the trimethylammonium ions and a valency specific relief of block.  相似文献   

10.
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   

11.
The peptide CO-NH function was replaced by a trans carbon-carbon double bond or by a CH2-CH2 isostere in enkephalin analogues of DADLE, DCDCE-NH2 or DPDPE. In DADLE the 2-3 and the 3-4 peptide bond was modified, whereas in the cyclic analogues the Gly3-Phe4 bond was replaced by the isosteres Gly psi (E,CH = CH)Phe [5-amino-2-(phenylmethyl)-3(E)-pentenoic acid] or Gly psi (CH2CH2)Phe [5-amino-2-(phenylmethyl)pentanoic acid]. In general, the modification results in a drop in potency which is the largest for the flexible CH2-CH2 replacement. The Gly3 psi (E,CH = CH)Phe4 DCDCE-NH2 analogue retains considerable potency. These results confirm the importance of the peptide function at the 2-3 and 3-4 position in enkephalin analogues for biological potency.  相似文献   

12.
Permeabilities of uncharged ammonia (NH3), methylamine (CH3NH2), and ethylamine (CH3CH2NH2) in the gram-negative phototrophic bacterium Rhodobacter sphaeroides were measured directly in cells grown heterotrophically under aerobic conditions. The permeability of NH3 was 2.55 +/- 0.73 microns s-1 (n = 20), but the permeabilities of CH3NH2 (MA) and CH3CH2NH2 (EA) were higher, PMA = 17.8 +/- 2.8 microns s-1 (n = 50), PEA = 24.7 +/- 3.9 microns s-1 (n = 44). The relative permeabilities of amines were also determined from their effect on the pH gradient across the cell membrane at alkaline external pH. In aerobically grown R. sphaeroides, both techniques indicated that the permeability of CH3CH2NH2 was about 30% greater than that of CH3NH2 but that the permeability of NH3 was only about 1/5 that of CH3NH2. The relative permeabilities of NH3 (A) and CH3NH2 were different in R. sphaeroides cells grown under three different physiological conditions: (a) cells grown aerobically with ammonium sulfate (PA/PMA about 0.20), (b) cells grown anaerobically with ammonium sulfate as their nitrogen source (PA/PMA about 0.29), and (c) diazotrophic cells (PA/PMA about 0.38). NH3 was also found to be only about 1/3 as permeable as CH3NH2 in the alkalophilic gram-positive bacterium Bacillus firmus. The findings that permeability properties of NH3 and CH3NH2 are very different in different bacteria and vary according to the conditions under which the organism is grown need to be taken into account in the interpretation of experiments where [14C]methylamine is used as an ammonia analog.  相似文献   

13.
Anaerobic metabolism of immediate methane precursors in Lake Mendota.   总被引:10,自引:10,他引:0       下载免费PDF全文
Lake Mendota sediments and the immediate overlying water column were studied to better understand the metabolism of the methanogenic precursors H2/CO2 and acetate in nature. The pool size of acetate (3.5 microns M) was very small, and the acetate turnover time (0.22h) was very rapid. The dissolved inorganic carbon pool was shown to be large (6.4 to 8.3 mM), and the turnover time was slow (111 H.). CO2 was shown to account for 41 +/- 5.5% of the methane produced in sediment. Acetate and H2/CO2 were simultaneously converted to CH4. The addition of H2 to sediments resulted in an increase specific activity of CH4 from H(14)CO3- and a decrease in specific activity of CH4 from [2-14C]acetate. Acetate addition resulted in a decrease in specific activity of CH4 from H(14)CO3-. The metabolism of H(14)CO3- or [2-14C]acetate to 14CH4 was not inhibited by addition of acetate or H2. After greater than 99% of added [2-14C]acetate had been turned over, 42% of the label was recovered as 14CH4 20% was recovered as 14CO2 and 38% was incorporated into sediment. Inhibitor studies of [2-14C]acetate metabolism in sediments demonstrated that CHCl3 completely inhibited CH4 formation, but not CO2 production. Air and nitrate addition inhibited CH4 formation and stimulated CO2 production, whereas fluoroacetate addition totally inhibited acetate metabolism. The oxidation of [2-14C]acetate to 14CO2 was shown to decrease with time when sediment was incubated before the addition of label, suggesting depletion of low levels of an endogenous sediment electron acceptor. Acetate metabolism varied seasonally and was related to the concentration of sulfate in the lake and interstitial water. Methanogenesis occurred in the sediment and in the water immediately overlying the sediment during period of lake stratification and several centimeters below the sediment-water interface during lake turnovers. These data indicate that methanogenesis in Lake Mendota sediments was limited by "immediate" methane precursor availability (i.e., acetate and H2), by competition for these substrates by nonmethanogens, and by seasonal variations which altered sediment and water chemistry.  相似文献   

14.
Anaerobic metabolism of immediate methane precursors in Lake Mendota.   总被引:8,自引:0,他引:8  
Lake Mendota sediments and the immediate overlying water column were studied to better understand the metabolism of the methanogenic precursors H2/CO2 and acetate in nature. The pool size of acetate (3.5 microns M) was very small, and the acetate turnover time (0.22h) was very rapid. The dissolved inorganic carbon pool was shown to be large (6.4 to 8.3 mM), and the turnover time was slow (111 H.). CO2 was shown to account for 41 +/- 5.5% of the methane produced in sediment. Acetate and H2/CO2 were simultaneously converted to CH4. The addition of H2 to sediments resulted in an increase specific activity of CH4 from H(14)CO3- and a decrease in specific activity of CH4 from [2-14C]acetate. Acetate addition resulted in a decrease in specific activity of CH4 from H(14)CO3-. The metabolism of H(14)CO3- or [2-14C]acetate to 14CH4 was not inhibited by addition of acetate or H2. After greater than 99% of added [2-14C]acetate had been turned over, 42% of the label was recovered as 14CH4 20% was recovered as 14CO2 and 38% was incorporated into sediment. Inhibitor studies of [2-14C]acetate metabolism in sediments demonstrated that CHCl3 completely inhibited CH4 formation, but not CO2 production. Air and nitrate addition inhibited CH4 formation and stimulated CO2 production, whereas fluoroacetate addition totally inhibited acetate metabolism. The oxidation of [2-14C]acetate to 14CO2 was shown to decrease with time when sediment was incubated before the addition of label, suggesting depletion of low levels of an endogenous sediment electron acceptor. Acetate metabolism varied seasonally and was related to the concentration of sulfate in the lake and interstitial water. Methanogenesis occurred in the sediment and in the water immediately overlying the sediment during period of lake stratification and several centimeters below the sediment-water interface during lake turnovers. These data indicate that methanogenesis in Lake Mendota sediments was limited by "immediate" methane precursor availability (i.e., acetate and H2), by competition for these substrates by nonmethanogens, and by seasonal variations which altered sediment and water chemistry.  相似文献   

15.
The complete sequence for the constant and 3' untranslated regions of a mouse gamma 2a immunoglobulin heavy chain mRNA is reported. The sequence is 1093 nucleotides long coding for the CH1 (amino-acids 118-214), the Hinge (215-230), the CH2 (231-340) and the CH3 (341-447). The 3' untranslated region is 103 nucleotides long preceding the poly(A). The nucleotide sequence predicts as in the case for gamma 1 and gamma 2b heavy chains an additional lysine residue before the termination codon. This sequence has been compared to the corresponding sequences of gamma 1 and gamma 2b heavy chain mRNAs. These sequences are respectively 75% and 84% homologous. The CH2 domains of gamma 2a and gamma 2b are 95% homologous at the nucleotide level. The cross-over point of a gamma 2a - gamma 2b heavy chain variant is located in a segment of 73 perfectly matching nucleotides. The 3' non coding regions of gamma 2a and gamma 2b are 89% homologous.  相似文献   

16.
The glutathione (GSH)-dependent dichloromethane dehalogenase from Methylophilus sp. strain DM11 catalyzes the dechlorination of CH(2)Cl(2) to formaldehyde via a highly reactive, genotoxic intermediate, S-(chloromethyl)glutathione (GS-CH(2)Cl). The catalytic mechanism of the enzyme toward a series of dihalomethane and monohaloethane substrates suggests that the initial addition of GSH to the alkylhalides is fast and that the rate-limiting step in turnover is the release of either the peptide product or formaldehyde. With the exception of CH(2)ClF, which forms a relatively stable GS-CH(2)F intermediate, the turnover numbers for a series of dihalomethanes fall in a very narrow range (1-3 s(-1)). The pre-steady-state kinetics of the DM11-catalyzed addition of GSH to CH(3)CH(2)Br exhibits a burst of S-(ethyl)-glutathione (k(b) = 96 +/- 56 s(-1)) followed by a steady state with k(cat) = 0.13 +/- 0.01 s(-1). The turnover numbers for CH(3)CH(2)Cl, CH(3)CH(2)Br, and CH(3)CH(2)I are identical, indicating a common rate-limiting step. The turnover numbers of the enzyme with CH(3)CH(2)Br and CH(3)CH(2)I are dependent on viscosity and are very close to the measured off-rate of GSEt. The turnover number with CH(2)I(2) is also dependent on viscosity, suggesting that a diffusive step is rate-limiting with dihaloalkanes as well. The rate constants for solvolysis of CH(3)SCH(2)Cl, a model for GS-CH(2)Cl, range between 1 s(-1) (1:1 dioxane/water) and 64 s(-1) (1:10 dioxane/water). Solvolysis of the S-(halomethyl)glutathione intermediates may also occur in the active site of the enzyme preventing the release of the genotoxic species. Together, the results indicate that dissociation of the GS-CH(2)X or GS-CH(2)OH intermediates from the enzyme may be a relatively rare event.  相似文献   

17.
The reaction mechanism for selective oxidation of propylamine on oxygen-covered gold has been studied by the density functional theory (DFT) and generalized gradient approximation (GGA) with slab model. Our calculation results indicated that the adsorption energy of propylamine decreases with the increasing oxygen coverage, that is -0.38, -0.20 and -0.10 eV on clean, 2/9 monolayer (ML) and 2/3 monolayer (ML) oxygen, respectively. The adsorption energies of the intermediates also have the trend of the gradual lower. The present work also indicated that the final product distribution depends on the oxygen coverage: propylamine undergoes N-H bond and C-H bond cleavage to produce propionitrile and water at low-oxygen-coverage (θ(o)?=?2/9 ML), and to yield propionitrile, propionaldehyde and water at high-oxygen-coverage (θ(o)?=?2/3 ML). The energy barrier of the first step of propyamine oxidation (CH(3)CH(2)CH(2)NH(2)?→?CH(3)CH(2)CH(2)NH) is 0.16 eV (θ(o)?=?2/9 ML) and 0.38 eV (θ(o)?=?2/3 ML). On the second step, the barrier energy is 0.16 (θ(o)?=?2/9 ML) and 0.25 (θ(o)?=?2/3 ML) eV of CH(3)CH(2)CH(2)NH?→?CH(3)CH(2)CH(2)N, next both C-H breakage and the barrier energy is 0.20 eV (CH(3)CH(2)CH(2)N?→?CH(3)CH(2)CHN) and 0.25 eV (CH(3)CH(2)CHN?→?CH(3)CH(2)CN) on low oxygen coverage, and 0.15 eV (CH(3)CH(2)CH(2)N?→?CH(3)CH(2)CHN) and 0.26 eV(CH(3)CH(2)CHN?→?CH(3)CH(2)CN) on the high oxygen coverage. The additional reaction step of CH(3)CH(2)CHN?→?CH(3)CH(2)CHO occurs on the high oxygen coverage, and the associated barrier is 0.41 eV. The calculation results show that the oxidation of propylamine can occur at room temperature due to the lower energy barrier. Furthermore, it was found that the energy barrier for the possible reaction steps at the low oxygen coverage is generally smaller than that on high oxygen coverage, which agrees with the experimental results.  相似文献   

18.
The synthesis of two new synthetic analogues of lecithin, two of phosphatidyl ethanolamine ("cephalin"), and one new phosphatidic acid analogue is described. They comprise one of each of the following types: the "isosteric" diether lecithin and cephalin analogues ROCH(2)CH(OR)- CH(2)CH(2)P(O) (O(-))OCH(2)CH(2)N(+)R'(3) (R = C(18)H(37); R' = H or CH(3)); and the "hydrocarbon" analogues of phosphatidic acid, lecithin, and cephalin, C(17)H(35)CH(2)CH(C(18)H(37))CH(2)P(O)(R) = (R'); [R = R' = OH; R = O(-), R' = OCH(2)CH(2)N(+)(CH(3))(3); and R = O(-), R' = OCH(2)CH(2)N(+)H(3)]. Infrared spectra and other properties of these compounds are described.  相似文献   

19.
Mercury was tested at the same concentration but under two different forms, organic CH3HgCl and inorganic HgCl2, in order to compare its relative inhibitory effect on in vitro microtubules polymerization. Induced by GTP and glycerol 8 M, tubulin polymerization was completely inhibited by HgCl2 10(-3) M while a 75.8% inhibition was measured for CH3HgCl2 10(-3) M.  相似文献   

20.
H-Pro-Pro-Asp-NH2 is a highly active and selective catalyst for asymmetric aldol reactions. Here, the versatility of H-Pro-Pro-Asp-NH2 has been further improved by immobilization on a solid support and functionalization with a short polyethylene glycol linker at the C-terminus. The development, synthesis, and the catalytic properties in aldol reactions of H-Pro-Pro-Asp-resin and H-Pro-Pro-Asp-Ahx-NH(CH2CH2O)3CH3 are described. For the solid-supported catalyst, TentaGel with a loading of 0.1-0.2 mmol g(-1) proved to be the optimal support. The solid-supported catalyst can be recycled at least three times without a significant drop in the catalytic activity or selectivity. Using the pegylated catalyst H-Pro-Pro-Asp-Ahx-NH(CH2CH2O)3CH3, only 0.5 mol % are necessary to obtain aldol products in up to 96% yield and 91% enantiomeric excess. In all cases, enantioselectivities are comparable to those obtained with the parent catalyst H-Pro-Pro-Asp-NH2. Thus, immobilization of H-Pro-Pro-Asp-NH2 on Tentagel as well as pegylation led to catalysts with selectivities comparable to the nonmodified catalyst, exhibiting additional distinct advantages such as facile reusability, ease of handling, higher solubility, and thereby greater versatility. handling, higher solubility, and thereby greater versatility.  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