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 reaction of benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-isopropylidene-β-lactoside with 1,11-ditosyloxy-3,6,9-trioxaundecane gave benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-isopropylidene-3,2′-O--(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (2, 47%). Acid hydrolysis of 2 and condensation of the product with 1,14-ditosyloxy-3,6,9,12-tetra-oxatetradecane afforded benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-(3,6,9,12-tetraoxa-tetradecane-1,14-diyl)-3,2′-O-(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (29%). Similarly, the reaction of benzyl 2,6,2′,4′,6′-penta-O-benzyl-β-lactoside with Ts[OCH2CH2]4OTs gave benzyl 2,6,2′,4′,6′-penta-O-benzyl-3,3′-O-(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (78%). 1H-N.m.r. spectroscopy has been used to study the formation of host-guest complexes with some of these macrocyclic compounds and benzyl ammonium thiocyanate.  相似文献   

2.
Steroid 21-hydroxylase activity has been identified in many tissues, including liver. But it is possible that the enzyme found in the liver is different from adrenal 21-hydroxylase. In the adrenal cortex, steroid 21-hydroxylase activity is increased by corticotropin (ACTH); the effect of ACTH is mediated by cyclic AMP (cAMP), and presumably involves a cAMP-dependent protein kinase (PKA). It is not yet clear, however, how extra-adrenal steroid 21-hydroxylase activity is regulated. In the present study, we examined the effect of N6, 2′-O-dibutyryl adenosine 3′,5′-cyclic monophosphate (dbcAMP), forskolin, N-[2-(methylamino)ethyl]5-isoquinolinesulfonamide (H-8) and 12-O-tetradecanoylphorbol-13-acetate (TPA) on steroid 21-hydroxylase activity in primary cultures of rat hepatocytes to determine the nature of regulation of extra-adrenal steroid 21-hydroxylase activity. Steroid 21-hydroxylase activity in hepatocytes incubated with 10−11M dbcAMP for 24 h was 1.6 times higher than that in control hepatocytes untreated with dbcAMP. On the other hand, steroid 21-hydroxylase activity decreased by 20 and 50% when the cells were incubated with 10−5 and 10−3 M dbcAMP, respectively. The stimulatory effect of 10−11 M dbcAMP was not blocked by 10−5 M H-8 (PKA inhibitor), but the inhibitory effect of 10−5 or 10−3 M cAMP was. TPA did not alter the activity of steroid 21-hydroxylase. These findings indicate that the steroid 21-hydroxylase in rat liver is regulated by mechanisms different from those in the adrenal glands.  相似文献   

3.
Methods for determining the differential susceptibility of human organs to DNA damage have not yet been explored to any large extent due to technical constraints. The development of comprehensive analytical approaches by which to detect intertissue variations in DNA damage susceptibility may advance our understanding of the roles of DNA adducts in cancer etiology and as exposure biomarkers at least. A strategy designed for the detection and comparison of multiple DNA adducts from different tissue samples was applied to assess esophageal and peripherally- and centrally-located lung tissue DNA obtained from the same person. This adductome approach utilized LC/ESI-MS/MS analysis methods designed to detect the neutral loss of 2′-deoxyribose from positively ionized 2′-deoxynucleoside adducts transmitting the [M+H]+ > [M+H−116]+ transition over 374 transitions. In the final analyses, adductome maps were produced which facilitated the visualization of putative DNA adducts and their relative levels of occurrence and allowed for comprehensive comparisons between samples, including a calf thymus DNA negative control. The largest putative adducts were distributed similarly across the samples, however, differences in the relative amounts of putative adducts in lung and esophagus tissue were also revealed. The largest-occurring lung tissue DNA putative adducts were 90% similar (n = 50), while putative adducts in esophagus tissue DNA were shown to be 80 and 84% similar to central and peripheral lung tissue DNA respectively. Seven DNA adducts, N2-ethyl-2′-deoxyguanosine (N2-ethyl-dG), 1,N6-etheno-2′-deoxyadenosine (dA), -S- and -R-methyl-γ-hydroxy-1,N2-propano-2′-deoxyguanosine (1,N2-PdG1, 1,N2-PdG2), 3-(2′-deoxyribosyl)-5,6,7,8-tetrahydro-8-hydroxy-pyrimido[1,2-a]purine-(3H)-one (8-OH-PdG) and the two stereoisomers of 3-(2′-deoxyribosyl)-5,6,7,8-tetrahydro-6-hydroxypyrimido[1,2-a]purine-(3H)-one (6-OH-PdG) were unambiguously detected in all tissue DNA samples by comparison to authentic adduct standards and stable isotope dilution and their identities were matched to putative adducts detected in the adductome maps.  相似文献   

4.

1. 1. Tightly bound ATP and ADP, found on the isolated mitochondrial ATPase, exchange only slowly at pH 8, but the exchange is increased as the pH is reduced. At pH 5.5, more than 60% of the bound nucleotide exchanges within 2.5 min.

2. 2. Preincubation of the isolated ATPase with ADP leads to about 50% inhibition of ATP hydrolysis when the enzyme is subsequently assayed in the absence of free ADP. This effect, which is reversed by preincubation with ATP, is absent on the membrane-bound ATPase. This inhibition seems to involve the replacement of tightly bound ATP by ADP.

3. 3. Using these two findings, the binding specificity of the tight nucleotide binding sites was determined. iso-Guanosine, 2′-deoxyadenosine and formycin nucleotides displaced ATP from the tight binding sites, while all other nucleotides tested did not. The specificities of the tight sites of the isolated and membrane-bound ATPase were similar, and higher than that of the hydrolytic site.

4. 4. The nucleotide specificities of ‘coupled processes’ nucleoside triphosphate-driven reversal of electron transfer, nucleoside triphosphate-32Pi exchange and phosphorylation were higher than that of the hydrolytic site of the ATPase and similar to that of the tight nucleotide binding sites.

5. 5. The different nucleotide specificities of uncoupled ATP hydrolysis and coupled processes can be explained even if both processes involve a single common site on the ATPase molecule. This model requires that energy can be ‘coupled’ only when it is released/utilised in the nucleotide binding steps of the mechanism.

6. 6. Adenosine β,γ-imidotriphosphate (AMP-PNP) is not a simple reversible inhibitor of the ATPase, since incubation requires preincubation and is not reversed when the compound is diluted out, or by addition of ATP. This compound inhibits the isolated and membrane-bound ATPase equally well. Its guanosine analogue does not act in this way.

7. 7. In submitochondrial particles, ADP inhibited uncoupled hydrolysis of ATP much more effectively than coupled hydrolysis, the latter being measured both directly (from ATP hydrolysis in the absence of uncoupler) or indirectly, by monitoring ATP-driven reduction of NAD+ by succinate.

8. 8. The effects of ADP and AMP-PNP were interpreted as providing evidence for two of the intermediates in the proposed scheme for coupled triphosphate hydrolysis.

Abbreviations: ε-ATP, N1,N6-ethenoadenosine triphosphate; 8-BrATP, 8-bromoadenosine triphosphate; AMP-PNP, adenosine β,γ-imidotriphosphate; GMP-PNP, guanosine β,γ-imidotriphosphate; N1,O-ATP, adenosine-N1-oxide triphosphate; rro-ATP 2,2′[1-(9-adenyl)-1′-(triphosphoryl-oxymethyl)-dihydroxydiethyl ether; and similarly for the respective diphosphates; NTP, NDP, nucleoside tri-, diphosphate; ANS, 1-anilino-8-naphthalene sulphonate; FCCP, carbonylcyanide p-trifluoromethoxyphenylhydrazone; HEPES, N-2-hydroxyethylpiperazine-N′-2-ethane sulphonic acid; MES, 2-(N-morpholino)-ethane sulphonic acid; TES, tris(hydroxymethyl)methylamino ethane sulphonic acid  相似文献   


5.
A number of N,N′-bis(4-substituted phenyl)-1,7-diaza-12-crown-4 and N,N′-bis(4-substituted phenyl)-1, 10-diaza-18-crown-6 (where the substituents are OCH3, CH3, H, Cl, respectively) have been prepared by cyclization reaction of a ditosylate with the appropriately substituted diol. These new macrocyclic ligands have been characterized by means of elemental analysis, IR, 1H NMR and MS spectra. The crystal structures of N,N′-bis(4-chlorophenyl)-1,10-diaza-18-crown-6 (21) and its complex with barium thiocyanate Ba(SCN)2 (22) have been determined by single crystal X-ray diffraction. The crystallographic data are as follows: 21: C24H32Cl2N2O4, orthorhombic, P212121, A=4.852(1), B=11.989(2), C=41.231(8) Å, V=2398.7(8) Å3, Z=4; 22: C26H32Cl2N4O4S2Ba, monoclinic, P21/c, A=8.801(2), B=11.653(9), C=15.756(6) Å, ß=105.96(3)°, V=1553.7(14) Å3, Z=2. In the complex, the Ba atom is eight-coordinate (O(1), O(2), O(1)′, O(2)′, N(1), N(1)′, N(21), N(21)′) to form a distorted D6h geometry with the Ba atom at the center of crystallographic symmetry.  相似文献   

6.
The carcinogenic and mutagenic N-nitroso compounds produce GC to AT and TA to GC transition mutations because they alkylate O6 of guanine and O4 of thymine. It has been generally assumed that these mutations occur because O6-alkylguanine forms a stable mispair with thymine and O4-alkylthymine forms a mispair with guanine. Recent studies have shown that this view is mistaken and that the alkylG·T and alkylT·G mispairs are not more stable than their alkylG·C or alkylT·A counterparts. Two possible explanations based on recent structural studies are put forward to account for the miscoding. The first possibility is that the DNA polymerase might mistake O6-alkylguanine for adenine, and O4-alkylthymine for cytosine, because of the physical similarity of these bases. O6-Methylguanine and adenine are similarly lipophilic and X-ray crystallography of the nucleosides has shown a close similarity in bond angles and lengths between O6-methylguanine and adenine, and between O4-methylthymine and cytosine. The second possible explanation is that the important factor in the miscoding is that the alkylG·T and alkylT·G mispairs retain the Watson-Crick alignment with N1 of the purine juxtaposed to N3 of the pyrimidine while the alkylG·C and alkylT·A pairs adopt a wobble conformation. 31P NMR of DNA duplexes show that the phosphodiester links both 3′ and 5′ to the C have to be distorted to accomodate the O6-ethylguanine:C pair, whereas there is less distortion of the phosphodiesters 3′ and 5′ to the T in an ethylG·T pair. Recent kinetic measurements show that the essential aspect of base selection in DNA synthesis is the ease of formation of the phosphodiester links on both the 3′ and 5′ side of the incoming base. The Watson-Crick alignment of the alkylG·T and alkylT·G mispairs may facilitate formation of these phosphodiester links, and this alignment rather than the strength of the base pairs and the extent of hydrogen bonding between them may be the crucial factor in the miscoding. If either hypothesis is correct it suggests that previously too much emphasis has been placed on the stability of the normal pairs in the replication of DNA.  相似文献   

7.

1. 1. Cyanide inhibits the catalytic activity of cytochrome aa3 in both polarographic and spectrophotometric assay systems with an apparent velocity constant of 4·103 M−1·s−1 and a Ki that varies from 0.1 to 1.0 μM at 22 °C, pH 7·3.

2. 2. When cyanide is added to the ascorbate-cytochrome c-cytochromeaa3−O2 system a biphasic reduction of cytochrome c occurs corresponding to an initial Ki of 0.8 μM and a final Ki of about 0.1 μM for the cytochrome aa3−cyanide reaction.

3. 3. The inhibited species (a2+a33+HCN) is formed when a2+a33+ reacts with HCN, when a2+a32+HCN reacts with oxygen, or when a3+a33+HCN (cyano-cytochrome aa3) is reduced. Cyanide dissociates from a2+a33+HCN at a rate of 2·10−3 s−1 at 22 °C, pH 7.3.

4. 4. The results are interpreted in terms of a scheme in which one mole of cyanide binds more tightly and more rapidly to a2+a33+ than to a3+a33+.

Abbreviations: TMPD, N,N,N′,N′-tetramethyl-p-phenylenediamine  相似文献   


8.
9.
The lithiation of indole, using a slight excess of n-butyl lithium in THF, followed by methylation and reaction with [Cr(CO)6] in refluxing dibutyl ether, resulted in the formation of [Cr(η6-N-methylindole)(CO)3] (1a) and [Cr(η6-N-methyl-2-methylindole)(CO)3] (1b). In contrast, lithiation of quinoline in THF, silylation and the subsequent reaction with [Cr(CO)6] under similar reaction conditions, afforded [Cr(η6-N-trimethylsilyl-2-butyl-1,2-dihydroquinoline)(CO)3] (2) and [Cr(η6-{2-butyl-1,2,3,4-tetrahydroquinoline})(CO)3] (3). The formation of [Cr(η6-2,2′-bis{N-methylindolyl})(CO)3] (4) implied lithiation at the 2-position of 1a. However, metallation at the 7-position was also indicated during the same reaction. In the presence of [Mn(CO)5Br], product 4 and the transmetallation product [Cr(η6-{7-(N-methylindolyl)Mn(CO)5})(CO)3] (5) were isolated. Reaction with titanocene dichloride gave [Cr(η6-{2-(N-methylindolyl)TiCp2Cl})(CO)3] (6), which slowly converted into [TiCp2{Cr(η6-2-(N-methylindolyl)(CO)3}2] (7).  相似文献   

10.
Various sulfidic anions and the oxidizing cations [Ru(NH3)6]3+ and N,N′-dimethyl-4,4′-bipyridinium2+ (paraquat2+) form ion pairs in aqueous solutions which display outer-sphere charge-transfer (CT) absorptions. The CT energies are used to establish a series of sulfidic anions with increasing CT donor strength: SCN2O3 2−4 3−3S3−2 −2S2 −4 2−.  相似文献   

11.
Total syntheses of (±)-ovalicin, its C4(S*)-isomer 44, and C5-side chain intermediate 46 were accomplished via an intramolecular Heck reaction of (Z)-3-(tert-butyldimethylsilyloxy)-1-iodo-1,6-heptadiene and a catalytic amount of palladium acetate. Subsequent epoxidation, dihydroxylation, methylation, and oxidation led to (3S*,5R*,6R*)-5-methoxy-6-(tert-butyldimethylsilyloxy)-1-oxaspiro[2.5]octan-4-one (2), a reported intermediate. The addition of a side chain with cis-1-lithio-1,5-dimethyl-1,4-hexadiene (27) followed by oxidation afforded (±)-ovalicin. The functional group manipulation afforded a number of regio- and stereoisomers, which allow the synthesis of analogs for bioevaluation. The structure of 44 was firmly established via a single-crystal X-ray analysis. The stereochemistry at C4 generated from the addition reactions of alkenyllithium with ketones 2, 40, and 45 is dictated by C6-alkoxy functionality. Anti-trypanosomal activities of various ovalicin analogs and synthetic intermediates were evaluated, and C5-side chain analog, 46, shows the strongest activity. Compound 44 shows antiproliferative effect against HL-60 tumor cells in vitro. Compounds 46 and a precursor, (3S*,4R*,5R*,6R*)-5-methoxy-4-[(E)-(1′,5′-dimethylhexa-1′,4′-dienyl)]-6-(tert-butyldimethylsilyloxy)-1-oxaspiro[2.5]octan-4-ol (28), may be explored for the development of anti-parasitic drugs.  相似文献   

12.
After intracellular in vitro exposure to the mutagenic and carcinogenic N-nitroso compounds N-methyl-N-nitrosourea (MeNU) or N-ethyl-N-nitrosourea (EtNU), respectively, the average relative amounts of the premutational lesion O6-alkylguanine represent about 6% and 8% of all alkylation products formed in genomic DNA. At the level of individual DNA molecules gunine-O6 alkylation does nor occur at random; rather, the probability of a substitution reaction at the nucleophilic O6 atom is influenced by nucleotide sequence, DNA conformation, and chromatin structure. In the present study, 5 different double-stranded polydeoxynucleotides and 15 double-stranded oligodeoxynucleotides (24-mers) were reacted with MeNU or EtNU in vitro under standardized conditions. Using a competitive radioimmunoassay in conjunction with an anti-(O6-2′-deoxyguanosine) monoclonal antibody, the frequency of guanine-O6 alkylation was found to be strongly dependent on the nature of the nucleotides flanking guanine on the 5 and 3′ sides. Thus, a 5′ neighboring guanine, followed by 5 adenine and 5′ cytosine, provided an up to 10-fold more ‘permissive’ condition for O6-alkylation of the central guanine than a 5′ thymine (with a 5-methylcytocine in the 5′ position being only slightly less inhibitory). Thymine and cytosine were more ‘permissive’ when placed 3′ in comparison with their affects in the 5′ flanking position.  相似文献   

13.
14.
Panax ginseng root and cell cultures were shown to biotransform paeonol (1) into its 2-O-β-d-glucopyranoside (2). P. ginseng root cultures were also able to biotransform paeonol (1) into its 2-O-β-d-xylopyranoside (3), 2-O-β-d-glucopyranosyl(1 → 6)-β-d-glucopyranoside (4) and 2-O-β-d-xylopyranosyl(1 → 6)-β-d-glucopyranoside (5), and its demethylated derivate, 2′,4′-dihydroxyacetophenone (6). Compounds 3 and 4 are new glycosides. It is the first example that the administrated compound was converted into its xylopyranoside by plant biotransformation.  相似文献   

15.
16.
[RuII(Me2edda)(H2O)2] (1), Me2edda2− = N,N′-dimethylethylenediaminediacetate, exhibits a sterically-controlled molecular recognition in forming η2 and η4 olefin complexes. 1 exists with an N2O2 in-plane set of chelate donors and axial H2O ligands. The two CH3 functionalities of Me2edda2− are poised above and below the N2O2 plane of the glycinato rings. Studies herein of the 2,2′-bipyridine complex, [RuII(Me2edda)(bpy)], with bidentate bpy chelation as established via 1H NMR and electrochemical methods show 1 to be ligated in the S,S configuration with the glycinato rings in-plane as a cis-O form. 1 is sterically discriminating in forming η2 complexes with smaller olefins (ethylene, 2-propene, cis-2-butene, methyl vinyl ketone and 3-cyclohexene-1-methanol), but rejects larger decorated ring structures and branched olefins (1,2-dimethyluracil, cyclohexene-1-one 2-methyl-2-propene). η2 complexes of 1 have characteristic RuII/III DPP waves near 0.55 V which vary slightly with olefin structure. Potentially bidendate dienes (1,3-butadiene, 1,3-cyclohexadiene and 2,5-norbornadiene (nbd) form η4 complexes as shown by RuII/III waves between 0.94 and 1.30 V, indicate of a highly stabilized RuII center by π-backboning. An η2η4 ‘equilibrium’ with apparent K = 22 at 25 °C is observed for nbd coordinated to 1. (The η2 and η4 distribution may be a kinetic one and not a thermodynamic one). To allow formation of the cis η4 complexes, 1 must undergo a shift of one or both glycinato donors from the N2O2 plane into the axial site away from the dimethyl functionalities. η4 chelation by 1,3-butadiene has been confirmed by 1H NMR spectral assignments of two [RuII(Me2edda)] isomers, one in the axial rans-O glycinato configuration, e.g. 1,3-butadiene is bidentate in the original N2O2 plane and a second unsymmetrical glycinato arrangement with in-plane and axial glycinato as well as in-plane and axial η4-1,3-butadiene coordination. [RuII(hedta)(H2O)] (2), hedta3− = N-hydrpxyethylenediaminetriacetate, is less discriminating for olefin structures, forming η2 complexes with all eleven olefins and dienes mentioned for studies with 1. However, 2 does not undergo displacement of a carboxylate donor by the second olefin unit of a diene [RuII(hedta)(diene)] complexes possess a pendant non-coordinated olefin and on η2-bound olefin in the complex, indicated by a normal RuII(pac)(olefin)RuII/III wave near 0.55 V.  相似文献   

17.
A new approach to ligand design for the sequestration of metal-oxo cations has been called stereognostic coordination chemistry, in that the ligand incorporates a traditional Lewis base coordination to the metal center and a hydrogen bond donor to interact with the oxo group. This paper reports the synthesis of ligands that are more rigid and sterically predisposed to bind the targeted UO22+ cation. These are the tripod ligands tris-N,N′,N′′-[2-(2-carboxy-phenoxy)ethyl]-1,4,7-triazacyclononane bis-hydrochloride (ETAC · 2HCl) and tris-N,N′,N′′-[2-(2-carboxy-4-decyl-phenoxy)ethyl]-1,4,7-triazacyclononane tris-hydrochloride (DETAC · 3HCl), which chelate uranyl with a tris-carboxylate coordination sphere and provide a hydrogen bond donor through a protonated amine on the triazacyclononane macrocycle to interact with one uranyl oxo atom. Structural models predict that upon uranyl binding the hydrogen bond donor must point directly towards the oxo atom, enforcing a stereognostic interaction. Both ETAC and DETAC chelate the uranyl ion; DETAC is a powerful extractant and will quantitatively extract uranyl into an organic phase at pH 1.9 and above. The extraction coefficient is estimated to be 1014 in neutral aqueous conditions. Vibrational spectra of 18O labeled UO22+ have been used to probe the stereognostic coordination to uranyl utilizing hydrogen bonding.  相似文献   

18.
1′-O-Mesyl-6,6′-di-O-tritylsucrose and the corresponding 1′-O-tosyl derivative were prepared from 6,6′-di-O-tritylsucrose by selective sulphonylation. Both sulphonates underwent intramolecular cyclisation reactions, to give 2,1′-anhydrosucrose in high yields rather than the isomeric 1′,4′-anhydride. Sequential benzoylation, detritylation, and mesylation of the 2,1′-anhydride afforded 2,1′-anhydro-6,6′-di-O-mesylsucrose tetrabenzoate which, in the presence of base, gave 2,1′:3,6:3′,6′-trianhydrosucrose that was not identical with the product previously claimed to have this structure. Several derivatives of 2,1′-anhydrosucrose were prepared possessing different functional groups at either the 6,6′- or 4,6′-positions. Dimolar mesitylene-sulphonylation of 3,3′,4′6′-tetra-O-acetylsucrose gave the 6,1′-disulphonate, which, in the presence of alkali, gave 2,1′:3,6-dianhydrosucrose, which was transformed into the 2,1′:3,6:3′,6′-trianhydride by sequential bromination at C-6′ (carbon tetrabromide-triphenylphosphine) and base-catalysed cyclisation. Treatment of 3,3′,4′,6′-tetra-O-benzoylsucrose with sulphuryl chloride furnished the 4,6,1′-trichloro derivative, which, on alkaline hydrolysis, was converted into 2,1′:3,6-dianhydro-4-chloro-4-deoxy-galacto-sucrose.  相似文献   

19.
N-Methyl-N′-nitro-N-nitrosoguanidine (MNNG) reacts with 12 nucleophilic sites in DNA to induce a variety of lesions, but O6-methylguanine (O6-MeG) and O4-methylthymine are the most effective premutagenic lesions produced, mispairing with thymine and guanine, respectively. O6-MeG is repaired by O6-alkylguanine-DNA alkyltransferase (AGT), which removes the methyl group from the O6 position and transfers it to itself, rendering the transferase inactive. When diploid human fibroblasts were exposed to 25 μM, O6-benzylguanine (O6-BzG) in the medium for 3 h, their level of AGT activity was dramatically reduced, to a level of at most 1.6% of the control. Populations of cells pretreated with this level of O6-BzG for 2 h or not pretreated, were exposed to MNNG at a concentration of 2, 4 or 6 μM in the presence or absence of O6-BzG and assayed for survival of colony-forming ability and the frequency of 6-thioguanine-resistant cells (mutations induced in the HPRT gene). O6-BzG (25 μM) was also present in the appropriate half of the cells during the 24 h immediately follwing exposure to MNNG. This 27-h exposure to O6-BzG alone had no cytotoxic or mutagenic effect on the cells but significantly increased the cytotoxicity and mutagenecity of MNNG, increasing the mutant frequency to that found previously in human cells constitutively devoid of AGT activity. At doses of 2 μM and 4 μM MNNG, the mutant frequency observed with the AGT-depleted cells was 120 × 10−6 and 240 × 10−6, respectively; in the cells with abundant AGT activity, these values were 10 × 10−6 and 20 × 10−6, respectively. DNA-sequence analysis of the coding region of the HPRT gene in 36 independent mutants obtained from MNNG-treated AGT-depleted populations and 36 from the control populations showed that even though AGT repair lowered the frequency of mutants by more than 90%, it did not affect the kinds of mutations induced by MNNG nor the strand distribution of the premutagenic guanine lesions. In mutants from the AGT-depleted cells, there were 26 base substitutions and 13 putative splice site mutations; in the control, there were 25 base substitutions and 11 splice site mutations. All but two substitutions involved G · C with 92% being G · C → A · T. In both sets, of the premutagenic lesions were located in the nontranscribed strand. Many ‘hot spots’ were seen, and there was evidence that AGT repaired more lesions from the 5′ half of the gene than from the 3′ half.  相似文献   

20.
Per-N-formylation of aminoglycoside (aminocyclitol) antibiotics followed by mild hydrolysis with aqueous ammonia gave mono-N-deformylated derivatives. Each positional isomer of the mono-N-deformylated derivatives thus obtained was separated by column chromatography on Amberlite CG-50 (NH4+ ). Acylation of mono-N-deformylated derivatives gave the corresponding mono-N-acylated derivatives. The N-formyl groups of the mono-N-acylates were removed by the treatment with dilute aqueous hydrazine acetate, whereas the newly introduced N-acyl group was stable under these conditions. The 1-N-formyl group of the deoxystreptamine moiety of per-N-formylated aminoglycoside antibiotics containing neamine (or 3′-deoxyneamine) is more readily deformylated than the 3-N-formyl group. In this report, isolation and structural-elucidation studies, including 13C-n.m.r. spectral assignments, of positional isomers of tri-N-formyl derivatives of xylostasin (1), 3′-deoxyxylostasin (2), kanamycin A (3), and neamine (4) are described. This selective N-acylation provides a useful method for the preparation of 1-N-modified derivatives, and the synthesis of 3′-deoxybutirosin A (2f) from 2 is described in detail as an example.  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