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 46 毫秒
1.
The reactions of the cobalozime Co (DH)-2 and the related Co(DOH)DOpn0 are strongly catalyzed by cationic micelles of cetyltrimethylammonium bromide (CTABr). The rate enhancements for the cobaloxime reaction and for that of Co (DOH)DOpn0 (in parentheses) are: EtBr approximately 1 (8.5); n-C5H11Cl, 220 (200); ClCH2CO-2, 420 (173); ClCH2CH2CO-2, 373. Anionic micelles of sodium lauryl sulfate inhibit the reaction of Co(DOH)DOpn0 with ClCH2CO-2, but do not affect that with n-C5H11Cl. The reactions of Co(DOH)DOpn0, like those of the cobaloxime are SN2 displacements and in the absence of surfactant in MeOH n-PrBr is more reactive than iso-PrBr.  相似文献   

2.
Complexes [Au(2)(H(2)Gy3DH)(2)]Cl(2) (1), [Au(H(2)Gy3Me)]Cl(3) (2) and [Au(H(2)Gy3Et)]Cl(3) (3) were obtained with glyoxaldehyde bis(thiosemicarbazone) (H(2)Gy3DH) and its N(3)-methyl (H(2)Gy3Me) and N(3)-ethyl (H(2)Gy3Et) derivatives. The bis(thiosemicarbazones) and their gold(I) and gold(III) complexes exhibited anti-proliferative activity against HL-60, Jurkat (leukemia) and MCF-7 (breast cancer) cells at 10 μmol L(-1). Complex (2) was able to in vitro inhibit thioredoxin reductase (TrxR) activity, which suggests that inhibition of TrxR could be part of its mechanism of action.  相似文献   

3.
Glycosyl-1,4-benzodiazepin-2,5-diones were prepared by coupling polyhydroxylated groups at N-1 of the corresponding benzodiazepine. The groups include 1-deoxy-D,L-xylit-1-yl, 6-deoxy-D-glucopyranos-6-yl, and 6-deoxy-3-OR-D-glucopyranos-6-yl (R = n-CnH(2n +1); n = 8, 12, and 16). The structural variations of the sugar group allowed comparison of such amphiphilic data as water solubility (Sw), critical micelle concentration (CMC), and corresponding surface tension (gamma) values. At 25 degrees C, unsubstituted benzodiazepines have Sw values from 0.9 to 4.2 10(-3) mol L(-1), whereas xylit-1-yl and 6-deoxy-D-glucopyranos-6-yl derivatives are, respectively, 7.4-25 and 58-204 times more soluble. Also, compounds with R = n-C8H17 are more soluble than corresponding benzodiazepines (1.4-5.8 times) and give micelles with CMC from 2.7 to 5.6 10(-3) mol L(-1) and corresponding gamma from 29 to 37 mN m(-1). In contrast, compounds with R = n-C12H25 and n-C16H33 are not soluble enough to reach the critical micelle concentration.  相似文献   

4.
We report the synthesis of new complexes with the general formula (R2SnX2)y.H2BiIm, where y = 1 or 2; R = Me, Et, Bun; X = Cl or Br (for R = Et) and H2BiIm = 2,2'-Biimidazole. The complexes have been characterized by elemental analysis and M?ssbauer, infra-red and 1H n.m.r. spectroscopy and tested (like the ligand, Me2SnCl2 and Et2SnCl2) against P388D1 leukemic cells.  相似文献   

5.
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   

6.
The 4,6-O-(1-methoxycarbonylethylidene), -(hydroxyisopropylidene), and -(methoxyisopropylidene) acetals of methyl 2,3-di-O-methyl-alpha-D-glucopyranoside were subjected to reductive cleavage in the presence of triethylsilane and trimethylsilyl methanesulfonate-boron trifluoride etherate (Me3SiOMs-BF3.Et2O), BF3.Et2O, or trimethylsilyl trifluoromethanesulfonate (Me3SiOSO2CF3) and the mole fractions of products were determined as a function of reaction time. The 4,6-(1-methoxycarbonylethylidene) acetal was quite stable to reductive-cleavage conditions but isomerization of the initial R,S mixture of diastereomers to the more-stable S diastereoisomer was noted. In addition, a slow, regiospecific, reductive ring-opening of the acetal was observed to give 6-O-[1-(methoxycarbonyl)ethyl] derivatives. The 4,6-(hydroxyisopropylidene) acetal was very unstable under reductive-cleavage conditions. Both Me3SiOMs-BF3.Et2O and Me3SiOSO2CF3 catalyzed complete removal of the group, via the intermediate 6-[1-(hydroxymethyl)ethyl] ether, but BF3.Et2O gave a mixture of products. The 4,6-(methoxyisopropylidene) acetal was also very labile under reductive-cleavage conditions; Me3SiOMs-BF3.Et2O catalyzed complete removal of the acetal, via the intermediate 6-[1-(methoxymethyl)ethyl]ether, but the intermediate ether was quite stable in the presence of either BF3.Et2O or Me3SiOSO2CF3. It is concluded from these studies that polysaccharides bearing 4,6-O-(1-carboxyethylidene) substituents can be analyzed directly by sequential permethylation and reductive cleavage. It is proposed that the identity of the substituted monomer and the positions of substitution of the acetal can be determined by sequential permethylation, ester reduction, and reductive cleavage.  相似文献   

7.
A series of pentanuclear gold(I)-copper(I) and -silver(I) mixed-metal alkynyl complexes, [(n)Bu(4)N][Au(3)M(2)(C triple bond CC(6)H(4)R-p)(6)] [M = Cu, R = OMe, O(n)Bu, O(n)Hex, Me, Et; M = Ag, R = Et, O(n)Hex] have been synthesized. The complexes were found to be emissive both in the solid state and in fluid solutions. DFT calculations at the B3LYP level of theory were performed on [Au(3)M(2)(C triple bond CC(6)H(4)Me-p)(6)](-) (M = Cu, Ag) to provide an understanding on the electronic structure of the complexes.  相似文献   

8.
The effect of phospholipid liposomes and surfactant micelles on the rate of nitric oxide release from zwitterionic diazeniumdiolates, R1R2N[N(O)NO]-, with significant hydrophobic structure, has been explored. The acid-catalyzed dissociation of NO has been examined in phosphate-buffered solutions of sodium dodecylsulfate (SDS) micelles and 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and 1,2-dipalmitoyl-sn-glycero-3-[phospho-(1-glycerol)] sodium salt (DPPG) phospholipid liposomes. The reaction behavior of dibenzylamine-, monobenzylamine-, and dibutylamine-derived substrates [1]: R1 = C6H5CH2, R2 = C6H5CH2 NH2+(CH2)2, 2: R1 = C6H5CH2, R2 = NH3+(CH2)2, and 3: R1 = n-butyl, R2 = n-butyl-NH2+(CH2)6] has been compared with that of SPER/NO, 4: R1 = H2N(CH2)3, R2 = H2N(CH2) 3NH2+(CH2)4]. Catalysis of NO release is observed in both micellar and liposome media. Hydrophobic interactions contribute to micellar binding for 1-3 and appear to be the main factor facilitating catalysis by charge neutral DPPC liposomes. Binding constants for the association of 1 and 3 with SDS micelles were 3-fold larger than those previously obtained with comparable zwitterionic substrates lacking their hydrophobic structure. Anionic DPPG liposomes were much more effective in catalyzing NO release than either DPPC liposomes or SDS micelles. DPPG liposomes (at 10 mM total lipid) induced a 30-fold increase in the NO dissociation rate of SPER/NO compared to 12- and 14-fold increases in that of 1 and 3.  相似文献   

9.
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   

10.
A series of new platinum(II) amine complexes containing 1,1- and 1,2-cyclobutanedicarboxylate ligands, cis-[PtA2(1,1-CBDCA)] (A = RNH2, where R = C2H5, n-C3H7, n-C4H9, n-C5H11, n-C6H13, c-C3H5, c-C5H9, c-C6H11; A2 = ethylenediamine, 1,3-diaminopropane), cis-[PtA2(1,2-CBDCA)] (A = NH3, RNH2 where R = CH3, C2H5, n-C3H7, n-C4H9, c-C3H5) and trans-[Pt(NH3)2(1,1-CBDCAH)2] (CBDCA, CBDCAH = dianion and monoanion of the dicarboxylic acid, respectively) have been synthesized by an improved route. These complexes are stable in aqueous solution and show good aqueous solubility. The [Pt(c-C3H5NH2)2(1,1-CBDCA)] can be isolated in white, grey and blue forms. The grey and blue forms exhibit ESR signals analogous to the so-called platinum blues. The existence of the blue form in aqueous solution is time and temperature dependent. Several of the complexes have been tested against leukaemia L1210 in male BDF mice and activity appears to decrease with the increase in length of the aliphatic chain (or increase in size of the alicyclic ring) of the primary amine. The Yoshida lymphoscarcoma screen, usually insensitive to platinum drugs, was found to respond well to [Pt(n-C4H9NH2)2(1,1-CBDCA)] in 5-day subcutaneously implanted tumours in female Wistar rats.  相似文献   

11.
The fate of terminal (nonreducing) alpha-D-glucopyranosyluronic groups under reductive cleavage conditions was investigated by using the Klebsiella K2 (strain NCTC-418) capsular polysaccharide. Treatment of the fully methylated polysaccharide (1) with triethylsilane and a mixture of trimethylsilyl methanesulfonate (Me3SiOSO2CH3) and boron trifluoride etherate (BF3.Et2O) as the catalyst, resulted in complete cleavage of all glycosidic linkages to yield the expected products, namely 3-O-acetyl-1,5-anhydro-2,4,6-tri-O-methyl-D-glucitol (2), 3,4-di-O-acetyl-1,5-anhydro-2,6-di-O-methyl-D-mannitol (3), 4-O-acetyl-1,5-anhydro-2,3,6-tri-O-methyl-D-glucitol (4), and methyl 2,6-anhydro-3,4,5-tri-O-methyl-L-gulonate. Treatment of 1 with trimethylsilyl trifluoromethanesulfonate (Me3SiOSO2CF3) as the catalyst resulted in incomplete cleavage of the glycosidic linkage of the methylated D-glucopyranosyluronic group, to yield 4-O-acetyl-1,5-anhydro-2,6-di-O-methyl- 3-O-(methyl2,3,4-tri-O-methyl-alpha-D-glucopyranosyluronate )-D-mannitol (9). Reductive cleavage of 1 in the presence of BF3.Et2O resulted in incomplete cleavage of all glycosidic linkages and gave rise to all four dimers (including 9) that could be formed from a tetrasaccharide repeating unit. The proposed structures of these dimers are based upon their composition, as established by chemical ionization mass spectrometry and by the reported structure of the polysaccharide. A small proportion of 1,5-anhydro-2,4,6-tri-O-methyl-3-O-(methyl 2,3,4-tri-O-methyl-alpha-D-glucopyranosyluronate)-D-mannitol (12) was also detected in the products of the BF3.Et2O-catalyzed reductive cleavage. The presence of 12 is chemical evidence for the phase of the tetrasaccharide repeating unit in the polysaccharide. The reductive cleavage of 1 was also accomplished after reduction of its ester groups with lithium aluminum hydride. Complete cleavage of all glycosidic linkages was observed when either Me3SiOSO2CF3 or Me3SiOSO2CH3-BF3.Et2O was used to catalyze reductive cleavage, and anhydroalditols 2, 3, 4, and 6-O-acetyl-1,5-anhydro-2,3,4-tri-O-methyl-D-glucitol were produced, as expected.  相似文献   

12.
Reaction of VOCl2 with 2-pyridineformamide thiosemicarbazone (H2Am4DH) and its N(4)-methyl (H2Am4Me), N(4)-ethyl (H2Am4Et) and N(4)-phenyl (H2Am4Ph) derivatives in ethanol gave as products [VO(H2Am4DH)Cl2] (1), [VO(H2Am4Me)Cl2] · 1/2HCl (2), [VO(H2Am4Et)Cl2] · HCl (3) and [VO(2Am4Ph)Cl] (4). Upon the dissolution of 1-4 in water, oxidation immediately occurs with the formation of [VO2(2Am4DH)] (5), [VO2(2Am4Me)] (6), [VO2(2Am4Et)] (7) and [VO2(2Am4Ph)] (8). The crystal and molecular structures of 5 and 6 were determined. Complexes 5-8 inhibited glycerol release in a similar way to that observed with insulin but showed a low enhancing effect on glucose uptake by rat adipocytes.  相似文献   

13.
M F Parisi  R H Abeles 《Biochemistry》1992,31(39):9429-9435
A series of fluorinated alpha-keto acid derivatives [PhCHFCOCO2R,PhCH2CHFCOCO2R,PhCF2-COCO2R, and PhCH2CF2COCO2R (R = H, Me, and Et)] was synthesized. They were inhibitors of chymotrypsin, with Ki values ranging from 4700 to 15 microM. Benzylpyruvic derivatives were generally more potent than the corresponding phenylpyruvic analogs. Esters of the first series were also more potent than their corresponding acids, and potency increased with the number of fluorine atoms. By replacing the ethoxy group of PhCH2CF2COCO2Et (15b) with an amino acid chain (i.e., alanyl-leucyl-arginine methyl ester hydrochloride and alanyl-leucyl-valine ethyl ester), the resultant peptides PhCH2CF2COCO-Ala-Leu-Arg-OMe.HCl.H2O (20) and PhCH2CF2COCO-Ala-Leu-Val-OEt.H2O (23) were found to be slow-binding inhibitors of chymotrypsin with considerably lower Ki values (0.19 and 3.6 microM, respectively). 19F NMR studies indicate, in the case of 20, the presence of an enzyme-inhibitor complex with a hemiketal structure similar to those observed between trifluoromethyl ketones and chymotrypsin. The results illustrate that effective protease inhibitors can be designed by enhancing the electrophilic character of the reactive carbonyl group (with an electron-withdrawing group placed on each side of the carbonyl group). Their potency and/or selectivity can also be improved by taking advantage of binding interactions at S' subsites of the protease.  相似文献   

14.
The parallel lability trend for the anation reactions of RRh(DH)2H2O (RCH3, CH3CH2, CF3CH2, ClCH2) and RCo(DH)2H2O complexes suggests a dissociative activation process for the reactions of the organorhodoximes. The high lability of these complexes, arising from the stabilization of the transition state, is not entirely due to the trans-effect of the R group, but, at least partially, to the labilizing effect of the equatorial macrocycle.  相似文献   

15.
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   

16.
The non-covalent interaction of the complexes [Pt(bpy)(R,R'NCSNR',R')(2)]Cl(2) (bpy=2,2'-bipyridine; R=R'=R'=R'=H; R=Me, R'=R'=R'=H; R=n-Bu, R'=R'=R'=H; R=p-tolyl, R'=R'=R'=H; R=Et, R'=H, R'=Et, R'=H) with calf thymus DNA has been studied at pH 7 and 25 degrees C. The processes give rise to: (i) reversible bathochromic shifts and strong hypochromicity of the absorption bands of the complexes, (ii) induced circular dichroism and (iii) an increase both in the melting temperature and viscosity of the DNA comparable to that observed for other well known metallointercalators. The binding constants, K(B), have been determined spectrophotometrically using the McGhee von Hippel equation. Plot of logK(B) vs -log[Na(+)] for the complex with unsubstituted thiourea gives a straight line with a slope value close to that expected for a dicationic intercalator. The binding affinity of the various complexes for DNA is independent of the thiourea nature; this suggests that the intercalation occurs through stacking of the bpy moiety while the ancillary ligands lie outside the nucleobases far away from the sugar phosphate backbone. The data show also that the electronic effects of the ligand substituents are not transmitted to the intercalating unit.  相似文献   

17.
Treatment of MHCl(CO)(PPh3)3 (M=Ru, Os) with (CH2=CH)SnR3 is a good general route to the coordinatively unsaturated osmium and ruthenium stannyl complexes M(SnR3)Cl(CO)(PPh3)2 (1: M=Ru, R=Me; 2: M=Ru, R = n-butyl; 3: M=Ru, R = p-tolyl; 4: M=Os, R=Me). These coordinatively unsaturated complexes readily add CO and CN-p-tolyl to form the coordinatively saturated compounds M(SnR3)Cl(CO)L(PPh3)2 (5: M=Ru, R=Me, L=CO; 6: M=;Ru, R = n-butyl, L=CO; 7: M=Ru, R = p-tolyl, L=CO; 8: M=Os, R=Me, L=CO; 9: M=Ru, R=Me, L=CN-p-tolyl; 10: M=Ru, R = n-butyl, L=CN-p-tolyl; 11: M=Os, R=Me, L=CN-p-tolyl). In addition, the chloride ligand in Ru(SnR3)Cl(CO)(PPh3)2 proves to be labile and treatment with the potentially bidentate anionic ligands, dimethyldithiocarbamate or diethyldithiocarbamate, affords the coordinatively saturated compounds Ru(SnR3)(η2-S2CNR′2)(CO)(PPh3)2 (12: R=Me, R′ = Me; 13: R=Me, R′ = Et; 14: R = n-butyl, R′ = Me; 15: R = p-tolyl, R′ = Me; 16: R = p-tolyl, R′ = Et). Chloride is also displaced by carboxylates forming the six-coordinate compounds Ru(SnR3)(η2-O2CR′)(CO)(PPh3)2 (17: R=Me, R′ = H; 18: R=Me, R′ = Me; 19: R=Me, R′ = Ph; 20: R = n-butyl, R′ = Me; 21: R = p-tolyl, R′ = Me). IR and 1H NMR spectral data for all the new compounds and 31P and 119Sn NMR spectral data for selected compounds are reported.  相似文献   

18.
The preparation and characterisation of the trichlorotris(alkylnicotinate)chromium(III) complexes of general formula CrCl3(py·3COOR)3·nH2O, where R = Me, Et, Pr and Bu are reported, n being 3.5, 1.0, 0 and 0 respectively. It is concluded that the ligation of the three chloride ions and that of the three nitrogen atoms is consistent with a C2u arrangement in each case.  相似文献   

19.
Complexes [Bi(2Fo4Ph)Cl(2) ] (1), [Bi(2Ac4Ph)Cl(2) ] (2), [Bi(2Bz4Ph)Cl(2) ] (3), [Bi(H(2) Gy3DH)Cl(3) ] (4), [Bi(H(2) Gy4Et)(OH)(2) Cl] (5), and [Bi(H(2) Gy4Ph)Cl(3) ] (6) were prepared with pyridine-2-carbaldehyde 4-phenylthiosemicarbazone (H2Fo4Ph), 1-(pyridin-2-yl)ethanone 4-phenylthiosemicarbazone (H2Ac4Ph), phenyl(pyridin-2-yl)methanone 4-phenylthiosemicarbazone (H2Bz4Ph), as well as with glyoxaldehyde bis(thiosemicarbazone) (H(2) Gy4DH) and its 4-Et (H(2) Gy4Et) and 4-Ph (H(2) Gy4Ph) derivatives. The complexes exhibited antibacterial activities against Staphylococcus aureus, Staphylococcus epidermidis, Enterococcus faecalis, and Pseudomonas aeruginosa. Coordination to Bi(III) proved to be an effective strategy to increase the antibacterial activity of the thiosemicarbazones and bis(thiosemicarbazones).  相似文献   

20.
A new catalyst for reductive cleavage of methylated glycans   总被引:1,自引:0,他引:1  
Several per-O-methylated D-glucans and D-fructans were used as models in an attempt to identify new catalysts for carrying out reductive cleavage. Included in these model studies were several D-glucans that contained 4-linked D-glucopyranosyl residues as well as one having a 4-linked D-glucitol residue, as both types of residue had previously been found to give rise to substantial proportions of artifactual products. These studies led to the development of a new catalyst for carrying out reductive cleavage, namely, a mixture of 5 equivalents of trimethylsilyl methanesulfonate (Me3SiOSO2Me) and 1 equivalent of boron trifluoride etherate (BF3 . Et2O) per equivalent of acetal. This new catalyst was found to accomplish the reductive cleavage of per-O-methylated, 4-linked D-glucopyranosyl residues and 4-linked D-glucitol residues, to give only the expected derivatives of 1,5-anhydro-D-glucitol and D-glucitol, respectively. The mixture of Me3SiOSO2Me and BF3 . Et2O also catalyzed reductive cleavage of the D-fructofuranosyl residues of per-O-methylated sucrose and inulin, to give only the expected derivatives of 2,5-anhydro-D-mannitol and 2,5-anhydro-D-glucitol. Indeed, when used alone, Me3SiOSO2Me also rapidly catalyzed the reductive cleavage of D-fructofuranosyl residues, but, under the same conditions, D-glucopyranosyl residues were unaffected. The results of these and other model studies demonstrated that catalysis of reductive cleavage by the mixture of Me3SiOSO2Me and BF3 . Et2O occurs in a synergistic manner. Examination of the mixture of Me3SiOSO2Me and BF3 . Et2O by 1H-n.m.r. spectroscopy demonstrated that a reaction occurs to generate trimethylsily fluoride and species of the type F2BOSO2Me, FB(OSO2Me)2, or B(OSO2Me)3 via ligand exchange.  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