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 enantioselectivity for lipase-catalyzed esterifications of 2-(4-substituted phenoxy)propionic acids in organic solvents was found to be mainly controlled by both size (steric) and electronic effects of substituents: H, F, Cl, CF3 and CH3. For the similar substituents in size, CF3 and CH3, however, their electronic effects play an important role in controlling the enantioselectivity. A model for the enantiorecognition is proposed by the discussion based on the value of the Michaelis constant obtained for the enantiomers.  相似文献   

2.
Trypsin catalyzed hydrolysis of seven new chromogenic arginine substrates, N alpha-benzyloxycarbonyl-L-arginine-3-nitro-5X-anilide (X = H, CF3, SO2CH3, F, Cl, Br and I) were studied. These substrates are suitable for studying electronic effects on trypsin activity. The Km and kcat values for the hydrolysis of each substrate were determined and found to differ significantly for the various substrates. The Hammett plot of the catalytic rate constants gave a straight line with a negative rho value (-0.82) thus indicating that electron withdrawing substituents retard the trypsin catalyzed hydrolysis of the new anilide substrates.  相似文献   

3.
Chuang CY  Vassar VC  Ma Z  Geney R  Ojima I 《Chirality》2002,14(2-3):151-162
Regio- and enantioselectivity in the asymmetric aminohydroxylation (AA) reaction of O-substituted 4-hydroxy-2-butenoates as well as the mechanism of the reaction were studied. When the electronic properties of the phenyl group in a substrate were altered by using different substituents, two conflicting trends were observed: The O-benzoyl substrates showed greater regio- and enantioselectivity when an electron-donating substituent was attached at the C-4 position of the phenyl group, while the O-benzyl substrates exhibited better regio- and enantioselectivity with an electron-withdrawing substituent at the C-4 position of the phenyl moiety. Thus, these results have disclosed hitherto unknown remarkable electronic effects in the AA reaction. Detailed analysis of possible electronic interactions in the chiral catalyst-substrate complex has revealed the importance of dipolar aromatic-aromatic interactions between the aromatic substituent of the substrate and the nitrogen heteroaromatic moiety of the chiral ligand for effective regiocontrol as well as enantioface selection in the AA reaction. A plausible model of the key intermediate in the AA reaction of O-substituted 4-hydroxy-2-butenoates is proposed.  相似文献   

4.
Summary Characteristics of the enantioselective hydrolysis of the acetic ester of 4-hydroxy-3-methyl-2-(2-propynyl)-2-cyclopentenone (HMPC) by Arthrobacter lipase were investigated in a water/oil biphasic reaction mixture. Kinetic studies revealed that the strict enantioselectivity was entirely due to a difference in the catalytic constants for the enantiomeric substrates and that (S)-HMPC acetate acted as a competitive inhibitor. The comparison of enantioselectivity for the acetates of HMPC analogues indicated that hydrophobic substituents in the HMPC molecule were essential for the strict enantioselectivity.Biological preparation of an optically active alcohol. Part II  相似文献   

5.
Six new methylenephosphonate analogues of P1P4-bis-(5',5'-adenosyl) tetraphosphate, Ap4A, having P2-P3 carbon bridges CF2, CCl2 and CH2CH2 or P1-P2 and P3-P4 carbon bridges CF2, CCl2 and CH2CH2 in the tetraphosphate chain, were examined as substrates or inhibitors for two specific Ap4A-degrading enzymes: (asymmetrical) Ap4A hydrolase (EC 3.6.1.17) from yellow-lupin seeds and (symmetrical) Ap4A hydrolase (EC 3.6.1.41) from Escherichia coli. All analogues in which the central oxygen atom was replaced by a stable carbon bridge were hydrolysed by the asymmetrical hydrolase (CF2 greater than CCl2 greater than O greater than CHBr greater than CH2 greater than CH2CH2). As expected, these analogues were not hydrolysed by the symmetrical hydrolase, which was also unable to act on analogues having P1-P2 and P3-P4 carbon bridges.  相似文献   

6.
DNA polymerase catalysis and fidelity studies typically compare incorporation of "right" versus "wrong" nucleotide bases where the leaving group is pyrophosphate. Here we use dGTP analogues replacing the beta,gamma-bridging O with CH2, CHF, CF2, or CCl2 to explore leaving-group effects on the nucleotidyl transfer mechanism and fidelity of DNA polymerase (pol) beta. T.G mismatches occur with fidelities similar to dGTP with the exception of the CH2 analogue, which is incorporated with 5-fold higher fidelity. All analogues are observed to bind opposite template C with Kds between 1 and 4 microM, and structural evidence suggests that the analogues bind in essentially the native conformation, making them suitable substrates for probing linear free energy relationships (LFERs) in transient-kinetics experiments. Importantly, Brnsted correlations of log(kpol) versus leaving-group pKa for both right and wrong base incorporation reveal similar sensitivities (betalg approximately -0.8) followed by departures from linearity, suggesting that a chemical step rather than enzyme conformational change is rate-limiting for either process. The location of the breaks relative to pKas of CF2, O, and the sterically bulky CCl2-bridging compounds suggests a modification-induced change in the mechanism by stabilization of leaving-group elimination. The results are addressed theoretically in terms of the energetics of successive primer 3'-O addition (bond forming) and pyrophosphate analogue elimination (bond breaking) reaction energy barriers.  相似文献   

7.
Abstract

A number of p-haloacetophenones were asymmetrically bioreduced to their corresponding (S)-alcohols by Mucor sp. CG10 with good conversion and excellent enantioselectivity. The results showed that the electronic effects of the halogen substituent (X-group) affected the conversion of the substrates and the enantioselectivity of the reaction. The trend observed was as the X-group at the para-position became more electron donating from F, to Cl, Br and I, the conversion of substrates decreased, while the enantioselectivity increased.  相似文献   

8.
The peroxyl radicals CF3O2., CCl3O2. and CBr3O2. were produced by radiolysis of aerated aqueous-alcohol solutions of CF3Br, CF3Cl, CCl4 or CBr4. Kinetic spectrophotometric pulse radiolysis experiments were carried out in the presence of various substrates: urate, ascorbate, xanthine, hydroquinone, p-methoxyphenol, phenol and chlorpromazine. Absolute rate constants for one-electron oxidation of these substrates by the alkylperoxyl radicals were found to vary from less than 10(5) to greater than 10(9) M-1 s-1, depending to some extent on the redox potential of the substrate. For all substrates the order of reactivity was CF3O2. greater than CBr3O2. greater than CCl3O2. . Because of its high reactivity, CF3O2., may have deleterious effects on biological systems. Its likely environmental precursor, CF3Br, which is used as a fire extinguisher and a refrigerant, was found to be reduced by a ferrous porphyrin model for cytochrome P-450 only very slowly and thus is not expected to have a major toxic effect if inhaled.  相似文献   

9.
10.
Catechol 2,3-dioxygenase [EC 1.13.11.2] from Pseudomonas putida mt-2 (Mpc) catalyzes the extradiol cleavage of catechol to produce 2-hydroxymuconate semialdehyde. The K(m) values for the catecholic substrate (K(mA)) and O(2) (K(mO2)), and catalytic constants (k(cat)) were kinetically determined for eight C3/C4-substituted catechols at 25 degrees C and pH 6.5 or 7.5. The first pK(a) values (pK(1)) were determined for eleven catechols (pK(1) = 7.26-9.47), correlated with Hammett substituent constants, and electron-withdrawing substituents significantly stabilized the monoanionic species of free catechols. Mpc preferred catechols with non-ionic substituents at the C3 or C4 position. 3-Phenylcatechol, a biphenyl, was cleaved, while 4-tert-butylcatechol was not. The logarithm of k(cat)/K(mA) (substrate specificity constant) exhibited a good linear correlation with pK(1), with the exception of those for 4-halocatechols. The logarithm of k(cat)/K(mO2) showed a good linear correlation with pK(1), with the exception of that of 3-phenylcatechol. These results demonstrate that catechol binding to the Mpc active site, the following O(2) binding, and the activation of the bound O(2) are all sensitive to electronic effects of the substituents. However, k(cat) did not correlate significantly with pK(1). The present study distinguishes clearly between the electronic and the steric effects of catecholic substrates in the reactivity of Mpc, and provides important insight into the mechanistic basis for a vast range of substrate specificities of extradiol dioxygenases.  相似文献   

11.
The enantioselective hydroformylation catalyzed by [(R,R)-Diop]Pt(SnCl3)Cl7 and the enantioselective hydroisopropoxycarbonylation catalyzed by [(R,R)-Diop]PdCl28 or by [(R,R)-Diop-dbp]PdCl29 of some para-substituted 2-phenylpropenes (para substituents = NO2, H, CH3O, Cl, CF3) was investigated in order to recognize possible electronic influences on the regioselectivity and on the enantioface selection which take place in such carbonylation reactions. The catalytic systems used gave no carbonylation products when the nitro compound was the substrate. 7 and 8 show similar regioselectivities, the less branched isomer being exclusively formed for all substrates except p-methoxy-2-phenylpropene which gave small amounts of the alternative regioisomer. The enantioselectivity depends on the σp effect of the substituent in both cases; the differences are, however, rather small and the trend is opposite in the two cases. The regioselectivity displayed by 9 is still in favour of the less branched isomer but it is high only in the case of p-trifluoromethyl-2-phenylpropene. Larger differences with respect to the other catalytic systems were also observed for enantioselectivity but the trend for both regioselectivity and enantioselectivity is not linear.  相似文献   

12.
Fluorotelomer alcohols (FTOHs; CF(3)(CF(2))(x)C(2)H(4)OH; where x=3, 5, 7, 9) are a novel class of polyfluorinated contaminants, recently detected in the North American atmosphere, that are possible precursors to the series of perfluoroalkyl carboxylates (PFCAs) in human blood. An in vivo rat study validated earlier independent work that poly- and per-fluoroalkyl carboxylates were metabolites of FTOHs, but our detection of several novel metabolites prompted us to examine their pathways in greater detail using isolated rat hepatocytes. Using 8:2 FTOH (i.e. where x=7) as a model compound, the metabolic products formed by isolated rat hepatocytes were identified, and three synthesized intermediates were incubated separately to elucidate the metabolic pathways. For 8:2 FTOH, a major fate was direct conjugation to form the O-glucuronide and O-sulfate. Using 2,4-dinitrophenylhydrazine (DNPH) trapping, the immediate oxidation product of 8:2 FTOH was identified as 8:2 fluorotelomer aldehyde (8:2 FTAL; CF(3)(CF(2))(7)CH(2)C(H)O). 8:2 FTAL was transient and eliminated HF non-enzymatically to yield 8:2 fluorotelomer alpha,beta-unsaturated aldehyde (8:2 FTUAL; CF(3)(CF(2))(6)CFCHC(H)O) which was also short-lived and reacted GSH and perhaps other endogenous nucleophiles. Four polyfluorinated acid intermediates were also detected, including 8:2 fluorotelomer carboxylate (8:2 FTCA; CF(3)(CF(2))(7)CH(2)C(O)O(-)), 8:2 fluorotelomer alpha,beta-unsaturated carboxylate (8:2 FTUCA; CF(3)(CF(2))(6)CFCHC(O)O(-)), tetrahydroperfluorodecanoate (CF(3)(CF(2))(6)(CH(2))(2)CO(2)(-)), and dihydroperfluorodecenoate (CF(3)(CF(2))(6)CHCHCO(2)(-)). The pathways leading to 8:2 FTCA and FTUCA involve oxidation of 8:2 FTAL, however, the pathways leading to the latter two polyfluorinated acids remain inconclusive. The fate of the unsaturated metabolites, 8:2 FTUAL and FTUCA, included conjugation with GSH and dehydrofluorination to yield alpha,beta-unsaturated GSH conjugates, and GS-8:2 FTUAL which was subsequently reduced to the corresponding alcohol. Perfluorooctanoate (PFOA) and minor amounts of perfluorononanoate (PFNA) were confirmed as metabolites of 8:2 FTOH, and the respective roles of beta- and alpha-oxidation mechanisms are discussed. The analogous acids, aldehydes, and conjugated metabolites of 4:2, 6:2, and 10:2 FTOH (i.e. where x=3, 5, and 9, respectively) were also detected, and metabolite profiles among FTOHs generally differed only in the length of their perfluoroalkyl chains. Preincubation with aminobenzotriazole, but not pyrazole, inhibited the formation of metabolites from all FTOHs, suggesting that their oxidation was catalyzed by P450, not alcohol dehydrogenase.  相似文献   

13.
Ohashi A 《Chirality》2002,14(7):573-577
Using a series of the rhodium complexes with (1S,2S)-1-(R(1))methylphosphino-2-(R(2))(R(3))phosphinoethane (R(1), R(2) and R(3) = 1-adamantyl, t-butyl, cyclohexyl, cyclopentyl, methyl; abbreviated as unsymmetrical BisP*), very high enantioselectivities were observed when the di- or tri- substituted and tetra-substituted dehydro-alpha-amino acid derivatives were used as the substrates. The main factor to give high enantioselectivity is the repulsive interaction between the functional groups of the substrate and the bulky substituents of the unsymmetrical BisP*. Since the unsymmetrical BisP* has two independent chiral phosphorous atoms in the vicinity of the active site, the higher enantioselectivity than those by the C2 symmetric BisP* complexes can be obtained. Moreover, the fine-tuning to obtain extremely high enantioselectivity may be possible by changing the combination of the substituents on the two phosphorous atoms of the unsymmetrical BisP*.  相似文献   

14.
Bacterial O6-alkylguanine-DNA alkyltransferase (AGT) removes alkyl group from O6-alkylguanine and O4-alkylthymine residues in DNA, both of which are considered to be DNA damages most related to the induction of cancer and/or mutation. The repair process involves alkyl-transfer of an O-alkyl group to the active site of the enzyme, where an SH-group of cysteine residue plays the role of alkyl acceptor. In order to elucidate the chemical characteristics of substrates for this enzyme, dealkylation rates of O6-alkyldeoxyguanosine, O4-alkylthymidine and related compounds were measured using an alkyl-transfer system. Thiophenol-triethylamine system was employed as an alkyl acceptor and twenty-one O-alkyl compounds were tested. Dealkylation proceeded with pseudo first order kinetics. The half-life of O6-methyldeoxyguanosine (MedG) was 122 h and no remarkable dependence on N-9 substituents (H, CH3 and deoxyribose) was observed. A compound lacking 2-NH2 group underwent demethylation about three times faster than O6-methylguanines did, while, a compound lacking imidazole moiety underwent demethylation about 2.5 times more slowly. The half-life of O4-methylthymidine (MedT) was 38 h and no remarkable dependence on N-1 (H, CH3 and deoxyribose) and C-5 (H and CH3) substituents was observed. Deethylation proceeded much more slowly than demethylation. Substitution of selenophenol for thiophenol resulted in a 4.5 times faster MedG demethylation rate. Demethylation rates were moderately correlated with values for NMR chemical shift of CH3 group, an indicator of electron density, although the correlation curves of a series of MedG and MedT derivatives were quite different. This result suggests that some different rate-determining factors other than electron density are playing a role. These findings may be of help in resolving the details of the mechanisms of enzymic repair by bacterial and mammalian AGT.  相似文献   

15.
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   

16.
Reaction of elemental copper and zinc powder mixtures with glycine (NH2.CH2COOH; HA) or aspartic acid (NH2CHCOOHCH2COOH; H2B) (in 1:1:2 ratio, respectively) in the presence of excess hydrogen peroxide (H2O2) at 50 degrees C, results in the formation of a new mixed metal peroxy carbonate compound corresponding to formula [Cu(Zn)2(O2(2-) (CO3)2(H2O)4], while the same reaction with elemental copper powder alone yields merely peroxy amino acid compounds having the formula [Cu(O2(2-)) (HA)2(H2O)] and [Cu(O2(2-)) (H2B) (H2O)2] for glycine and aspartic acid, respectively. These compounds have been characterized by elemental analysis, ESR, and electronic and IR spectra. It is interesting to note that both amino acids are converted to carbonate in the presence of zinc alone. A method analogous to that described above, for the reaction of elemental copper, zinc powder mixtures with succinic acid [(CH2COOH)2] or acetic acid (CH3COOH) in excess H2O2, on the other hand, gave a product essentially comprising copper succinate or acetate, respectively. These observations suggest an interesting and perhaps important phenomenon by which only the simple amino acids such as glycine and aspartic acid are converted to carbonates while their corresponding carboxylic acids form only their respective salts.  相似文献   

17.
Zheng H  Lipscomb JD 《Biochemistry》2006,45(6):1685-1692
The hydroxylase component (MMOH) of the soluble form of methane monooxygenase (sMMO) isolated from Methylosinus trichosporium OB3b catalyzes both the O2 activation and the CH4 oxidation reactions at the oxygen-bridged dinuclear iron cluster present in its buried active site. During the reaction cycle, the diiron cluster forms a bis-mu-oxo-(Fe(IV))2 intermediate termed compound Q (Q) that reacts directly with methane. Many adventitious substrates also react with Q, most at a relatively slow rate. We have proposed that Q reacts preferentially with CH4 because the sMMO regulatory component MMOB induces a size selective pore into the MMOH active site as the two components form a complex. Support for this proposal has come through the observation of a nonlinear Arrhenius plot for the CH4 oxidation, presumably due to a shift in rate-limiting step from substrate binding at low temperature to C-H bond cleavage at high temperature. Reactions of all substrates other than CH4 fail to exhibit a break in the Arrhenius plot because binding is always rate limiting in the temperature range explored. Here we show that it is possible to induce a break in the Arrhenius plot for the ethane reaction with Q by using an MMOB mutant termed DBL2 (S109A/T111A) in which residues at the MMOH-MMOB interface are reduced in size. We hypothesize that this increases the ethane binding rate and shifts the Arrhenius breakpoint into the observable temperature range. As a result of this shift, the kinetic and activation parameters of the C-H bond breaking reaction for both methane and ethane can be observed using the DBL2 mutant. A 2H-KIE is observed for both substrate oxidation reactions when using DBL2, whereas only CH4 oxidation exhibits an effect when using wild type MMOB, consistent with the C-H bond cleaving reaction becoming at least partially rate limiting for ethane. Analysis of the temperature dependence of the 2H-KIE for ethane and methane for reactions using both mutant and wild type forms of MMOB suggests that quantum tunneling plays a significant role in methane oxidation but not ethane oxidation.  相似文献   

18.
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   

19.
The lipase from Pseudomonas cepacia represents a widely applied catalyst for highly enantioselective resolution of chiral secondary alcohols. While its stereopreference is determined predominantly by the substrate structure, stereoselectivity depends on atomic details of interactions between substrate and lipase. Thirty secondary alcohols with published E values using P. cepacia lipase in hydrolysis or esterification reactions were selected, and models of their octanoic acid esters were docked to the open conformation of P. cepacia lipase. The two enantiomers of 27 substrates bound preferentially in either of two binding modes: the fast-reacting enantiomer in a productive mode and the slow-reacting enantiomer in a nonproductive mode. Nonproductive mode of fast-reacting enantiomers was prohibited by repulsive interactions. For the slow-reacting enantiomers in the productive binding mode, the substrate pushes the active site histidine away from its proper orientation, and the distance d(H(N epsilon) - O(alc)) between the histidine side chain and the alcohol oxygen increases, d(H(N epsilon) - O(alc)) was correlated to experimentally observed enantioselectivity: in substrates for which P. cepacia lipase has high enantioselectivity (E > 100), d(H(N epsilon) - O(alc)) is >2.2 A for slow-reacting enantiomers, thus preventing efficient catalysis of this enantiomer. In substrates of low enantioselectivity (E < 20), the distance d(H(N epsilon) - O(alc)) is less than 2.0 A, and slow- and fast-reacting enantiomers are catalyzed at similar rates. For substrates of medium enantioselectivity (20 < E < 100), d(H(N epsilon) - O(alc)) is around 2.1 A. This simple model can be applied to predict enantioselectivity of P. cepacia lipase toward a broad range of secondary alcohols.  相似文献   

20.
A group of unnatural 1-(2-deoxy-beta-D-ribofuranosyl)isocarbostyrils having a variety of C-7 substituents [H, 4,7-(NO2)2, I, CF3, CN, (E)-CH=CH-I, -C triple bond CH, -C triple bond C-I, -C triple bond C-Br, -C=C-Me], designed as nucleoside mimics, were synthesized for evaluation as anticancer and antiviral agents. This class of compounds exhibited weak cytotoxicity in a MTT assay (CC50 = 10(-3) to 10(-5) M range) with the 4,7-dinitro derivative being the most cytotoxic, relative to thymidine (CC50 = 10(-3) to 10(-5) M range), against a variety of cancer cell lines. The 4,7-dinitro, 7-I and 7-C triple bond CH compounds exhibited similar cytotoxicity against non-transfected (KBALB, 143B), and HSV-1 TK+ gene transfected (KBALB-STK, 143B-LTK) cancer cell lines possessing the herpes simplex virus type 1 (HSV-1) thymidine kinase gene (TK+). This observation indicates that these compounds are not substrates for HSV type-1 TK, and are therefore unlikely to be useful in gene therapy based on the HSV gene therapy paradigm.  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