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 31 毫秒
1.
The complex formation of S- and R-ofloxacin with the self-complementary oligonucleotides, namely d[ATAGCGCTAT](2), d[GCGATATCGC](2) and d[ATAICICTAT](2), were investigated by the molecular dynamics (MD) simulation. Four starting positions, including two intercalation positions with different insertion directions and two minor groove binding positions, were considered. The total energy of both S- and R-ofloxacin-d[ATAGCGCTAT](2) complex, in which ofloxacin binds in the minor groove of the oligonucleotide, were lower than any intercalation binding mode. For both enantiomers, formation of the complex with GC oligonucleotide is more favorable than AT and IC oligonucleotides. When S- and R-ofloxacin are compared, the S-enantiomer exhibits more favorable total energy and torsion angles in the complex formation. This result is in agreement with the experimental observation [Hwangbo et al., Eur J Pharm Sci 18, 197 (2003)]. In the complex, both enantiomers form two hydrogen bonds: one between the carbonyl group of ofloxacin and the amine group of G16 and the other between the fluorine group and the G6 amine for S-ofloxacin. However, only one hydrogen bond is formed between endocyclic hydrogen atom at the C2 position of adenine and inosine base and carbonyl group of ofloxacin, which may be the reason for the GC preferentiality of ofloxacin.  相似文献   

2.
A sensitive, simple and accurate method for determination of enantiomers of ofloxacin in microsomal incubates was developed by chiral ligand-exchange RP-HPLC with fluorescence detection to examine stereoselective metabolism of ofloxacin in the glucuronidation process. The C18 stationary phase was used as analytical column. The solution of chiral mobile phase additive was made up of 6 mM l-phenylalamine mixed with 3 mM CuSO4 in water. Mobile phase consisted of the solution of chiral mobile phase additive–methanol (86:14).The fluorescence detector was operated at λex 330 nm and λem 505 nm. The flow-rate of mobile phase was set at 1.0 ml/min. The achiral ODS column offers good separation of the two enantiomers in less than 25 min. The recovery of the assay was 97.9±6.1% (n=10) for S-ofloxacin and 99.6±6.0% (n=10) for R-ofloxacin. The method provides a high sensitivity and good precision (RSD<10%). The LOD was 0.6 μM for both enantiomers and the LOQ was 5.70±0.45 μM (n=8) for S-ofloxacin and 5.66±0.47 μM (n=8) for R-ofloxacin. The standard curves showed excellent linearity over the concentration range 5.5–2078 μM for S-(−)-ofloxacin and R-(+)-ofloxacin. The enantioselective method developed has been applied to determine the stereoselectivity of glucuronidation metabolism of ofloxacin optical isomers in rat liver microsomes.  相似文献   

3.
A capillary electrophoresis method for the simultaneous separation and enantioseparation of the antibacterial drug ofloxacin and its metabolites desmethyl ofloxacin and ofloxacin N-oxide in human urine has been developed and validated. Enantioseparation was achieved by adding sulfobutyl β-cyclodextrin to the running buffer. The detection of the analytes was performed by laser-induced fluorescence (LIF) detection using a HeCd-laser with an excitation wavelength of 325 nm. In comparison with conventional UV detection, LIF detection provides higher sensitivity and selectivity. The separation can be performed after direct injection of urine into the capillary without any sample preparation, because no matrix compounds interfere with the assay. Additionally, the high sensitivity of this method allows the quantification of the very low concentrations of enantiomers of both metabolites. The limit of quantification was 250 ng/ml for ofloxacin enantiomers and 100 ng/ml for each metabolites’ enantiomers. This method was applied to the analysis of human urine samples collected from a volunteer after oral administration of 200 mg of (±)-ofloxacin to elucidate stereoselective differences in the formation and excretion of the metabolites. It could be demonstrated that the renal excretion of the S-configured metabolites, especially S-desmethyl ofloxacin, within the first 20 h after dosage, is significantly lower than that of the R-enantiomers.  相似文献   

4.
Natural polynucleotides including Micrococcus lysodeikticus and calf thymus DNA exhibit enantioselective recognition to S‐ofloxacin regulated by Cu2+. This is the first report that ofloxacin and Cu2+ have cooperative effects on the local distortions of polynucleotides. At the [Cu2+]/[base] ratio of 0.1, S‐ofloxacin is more liable to induce the locally distorted structures of polynucleotides, of which the association constant of S‐ofloxacin toward DNA‐Cu(II) is three times higher than that of the R‐enantiomer. The apparent increase of adsorption capability and cooperativity, as well as the change of adsorption mechanism were detected in the adsorption of ofloxacin enantiomers on polynucleotides upon Cu(II)‐coordination. This study not only discloses the effect of the chiral drug on the structural transition of long double‐stranded DNA, but provides fundamental data to develop a novel enantioseparation method based on natural polynucleotides. Chirality 27:306‐313, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

5.
Two racemic tetrathiafulvalene‐[2.2]paracyclophane electron donors EDT‐TTF‐[2.2]paracyclophane 1 and (COOMe)2‐TTF‐[2.2]paracyclophane 2 have been synthesized via the phosphite mediated cross coupling strategy. Chiral HPLC allowed the optical resolution of the (RP) and (SP) enantiomers for both compounds. Solid‐state structures of (RP)‐ 1 and (rac)‐ 2 have been determined by single crystal X‐ray analysis. Intermolecular π‐π and S???S interactions are disclosed in the packing. Single crystal X‐ray analysis of (RP)‐ 1 combined with experimental and theoretical circular dichroism spectra allowed the assignment of the absolute configuration of the enantiomers of 1 and 2 .  相似文献   

6.
The purpose of this study was to investigate whether CYP2C19 activity can be estimated from plasma concentrations of lansoprazole enantiomers 4 h (C4h) after single administration by oral and enteral routes. Sixty‐nine subjects, 22 homozygous extensive metabolizers (homEMs), 32 heterozygous EMs (hetEMs), and 15 poor metabolizers (PMs), participated in the study. After a single oral or enteral dose of racemic lansoprazole (30 mg), plasma concentrations of lansoprazole enantiomers were measured 4 h postdose. The R/S ratio of lansoprazole at 4 h differed significantly among the three groups (P < 0.0001) regardless of the administration route. The R/S ratio of lansoprazole in CYP2C19 PMs ranged from 3.0 to 13.7, whereas in homEMs and hetEMs the ratio ranged from 8.6 to 90 and 2.1 to 122, respectively. The relationship between (S)‐lansoprazole concentration and R/S ratio of lansoprazole at C4h is given by the following formula: log10 [R/S ratio] = 2.2 – 0.64 × log10 [C4h of (S)‐lansoprazole] (r = 0.867, P < 0.0001). Thus, phenotyping CYP2C19 using the R/S enantiomer ratio of lansoprazole seems unlikely. However, to obtain a pharmacological effect similar to that in CYP2C19 PMs, we can presume that lansoprazole has a sufficient effect in the patient with an R/S enantiomer ratio at 4 h ≤ 13.70 and (S)‐lansoprazole concentration at 4 h ≥ 50 ng/ml. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

7.
G Trinquier  K X Chen  N Gresh 《Biopolymers》1988,27(9):1491-1517
Intercalated complexes of the antitumor antibiotic nogalamycin (NGM) with the double-stranded oligonucleotides d(GCGC)2, d(ATAT)2, and d(ACAC) · d(GTGT) are investigated with the theoretical method SIBFA. The amino sugar part of the drug locates preferentially in the minor groove. An intrinsic preference for the d(ATAT)2 sequence over the d(ACAC) · d(GTGT) and d(GCGC)2 sequences is obtained, corresponding to relative energies 0, 11, and 15 kcal/mole, respectively. A mixed sugar-puckering pattern is preferred in the d(ATAT)2 complex while a uniform sugar-puckering pattern is preferred for the other sequences. No direct specific interaction involves the N+—H part of protonated NGM. The location of the amino sugar as well as the sequence selectivity is due to the global electrostatic interaction of the dimethylammonium group with the given groove. The two hydroxyl groups of the amino sugar and the carbonyl of the carbomethoxy group encounter partners for hydrogen bonding at the intercalation site, but these interactions do not appear to govern the base sequence selectivity. The nogalose part is not found to be directly involved in the binding or in the selectivity. The conformations of isolated and intercalated NGM are discussed.  相似文献   

8.
Proudfoot EM  Mackay JP  Karuso P 《Biochemistry》2001,40(15):4867-4878
The molecular recognition of oligonucleotides by chiral ruthenium complexes has been probed by NMR spectroscopy using the template Delta-cis-alpha- and Delta-cis-beta-[Ru(RR-picchxnMe(2)) (bidentate)](2+), where the bidentate ligand is one of phen (1,10-phenanthroline), dpq (dipyrido[3,2-f:2',3'-h]quinoxaline), or phi (9,10-phenanthrenequinone diimine) and picchxnMe(2)() is N,N'-dimethyl-N,N'-di(2-picolyl)-1,2-diaminocyclohexane. By varying only the bidentate ligand in a series of complexes, it was shown that the bidentate alone can alter binding modes. DNA binding studies of the Delta-cis-alpha-[Ru(RR-picchxnMe(2))(phen)](2+) complex indicate fast exchange kinetics on the chemical shift time scale and a "partial intercalation" mode of binding. This complex binds to [d(CGCGATCGCG)](2) and [d(ATATCGATAT)](2) at AT, TA, and GA sites from the minor groove, as well as to the ends of the oligonucleotide at low temperature. Studies of the Delta-cis-beta-[Ru(RR-picchxnMe(2))(phen)](2+) complex with [d(CGCGATCGCG)](2) showed that the complex binds only weakly to the ends of the oligonucleotide. The interaction of Delta-cis-alpha-[Ru(RR-picchxnMe(2))(dpq)](2+) with [d(CGCGATCGCG)](2) showed intermediate exchange kinetics and evidence of minor groove intercalation at the GA base step. In contrast to the phen and dpq complexes, Delta-cis-alpha- and Delta-cis-beta-[Ru(RR-picchxnMe(2))(phi)](2+) showed evidence of major groove binding independent of the metal ion configuration. DNA stabilization induced by complex binding to [d(CGCGATCGCG)](2) (measured as DeltaT(m)) increases in the order phen < dpq and DNA affinity in the order phen < dpq < phi. The groove binding preferences exhibited by the different bidentate ligands is explained with the aid of molecular modeling experiments.  相似文献   

9.
Herein a molecular mechanic study of the interaction of a lethal chemical warfare agent, O-ethyl S-[2-(diisopropylamino)ethyl]methylphosphonothioate (also called VX), with Torpedo californica acetylcholinesterase (TcAChE) is discussed. This compound inhibits the enzyme by phosphonylating the active site serine. The chirality of the phosphorus atom induces an enantiomeric inhibitory effect resulting in an enhanced anticholinesterasic activity of the SP isomer (VXS) versus its RP counterpart (VXR). As formation of the enzyme-inhibitor Michaelis complex is known to be a crucial step in the inhibitory pathway, this complex was addressed by stochastic boundary molecular dynamics and quantum mechanical calculations. For this purpose two models of interaction were analyzed: in the first, the leaving group of VX was oriented toward the anionic subsite of TcAChE, in a similar way as it has been suggested for the natural substrate acetylcholine; in the second, it was oriented toward the gorge entrance, placing the active site serine in a suitable position for a backside attack on the phosphorus atom. This last model was consistent with experimental data related to the high inhibitory effect of this compound and the difference in activity observed for the two enantiomers. Proteins 28:543–555, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

10.
Racemic sulfonylated 2,5-diaryltetrahydrofuran [L-668,750, (±)-trans-2-[3-methoxy-5-(2-hydroxy)ethylsulfonyl-4-n-propoxy]-phenyl-5-(3,4,5-trimethoxyphenyl)-tetrahydrofuran, I] is a potent, specific and orally active platelet-activating factor (PAF) receptor antagonist. Its (—)-(2S,5S) enantiomer [L-680,573, (S)-I] exhibited higher PAF antagonistic potency than the (+)-(2R,5R) enantiomer [L-680,574, (R)-I] in vitro and in animal models. For assay of drug concentrations in plasma of rats dosed intravenously or orally with tritium-labeled I, we have developed a high-performance liquid chromatographic (HPLC) method which directly resolved the two enantiomers. The column contained α1-acid glycoprotein as the chiral stationary phase and was eluted with phosphate buffer, methanol and ethanol at neutral pH. The concentration of each enantiomer in the plasma was then determined by reverse isotope dilution assay. Results showed that the plasma clearance rate of the more potent (S)-I enantiomer was more than ten-fold faster than that of the (R)-I enantiomer; the enantioselective clearance resulted in nearly ten-fold higher concentrations of the latter in plasma at all time points regardless of the dosing route. This paper describes the HPLC chiral resolution method and its application in plasma analysis.  相似文献   

11.
A new type of planar chiral (Rp)‐ and (Sp)‐4,7,12,15‐tetrasubstituted [2.2]paracyclophanes was prepared from racemic 4,7,12,15‐tetrabromo[2.2]paracyclophane as the starting substrate. Regioselective lithiation and transformations afforded racemic bis‐(para)‐pseudo‐meta‐type [2.2]paracyclophane (4,15‐dibromo‐7,12‐dihydroxy[2.2]paracyclophane). Its optical resolution was performed by the diastereomer method using a chiral camphanoyl group as the chiral auxiliary. The diastereoisomers were readily isolated by simple silica gel column chromatography, and the successive hydrolysis afforded (Rp)‐ and (Sp)‐bis‐(para)‐pseudo‐meta‐type [2.2]paracyclophanes ((Rp)‐ and (Sp)‐4,15‐dibromo‐7,12‐dihydroxy[2.2]paracyclophanes). They can be used as pseudo‐meta‐substituted chiral building blocks.  相似文献   

12.
Fluoxetine is used clinically as a racemic mixture of (+)‐(S) and (–)‐(R) enantiomers for the treatment of depression. CYP2D6 catalyzes the metabolism of both fluoxetine enantiomers. We aimed to evaluate whether exposure to gasoline results in CYP2D inhibition. Male Wistar rats exposed to filtered air (n = 36; control group) or to 600 ppm of gasoline (n = 36) in a nose‐only inhalation exposure chamber for 6 weeks (6 h/day, 5 days/week) received a single oral 10‐mg/kg dose of racemic fluoxetine. Fluoxetine enantiomers in plasma samples were analyzed by a validated analytical method using LC‐MS/MS. The separation of fluoxetine enantiomers was performed in a Chirobiotic V column using as the mobile phase a mixture of ethanol:ammonium acetate 15 mM. Higher plasma concentrations of the (+)‐(S)‐fluoxetine enantiomer were found in the control group (enantiomeric ratio AUC(+)‐(S)/(–)‐(R) = 1.68). In animals exposed to gasoline, we observed an increase in AUC0‐∞ for both enantiomers, with a sharper increase seen for the (–)‐(R)‐fluoxetine enantiomer (enantiomeric ratio AUC(+)‐(S)/(–)‐(R) = 1.07), resulting in a loss of enantioselectivity. Exposure to gasoline was found to result in the loss of enantioselectivity of fluoxetine, with the predominant reduction occurring in the clearance of the (–)‐(R)‐fluoxetine enantiomer (55% vs. 30%). Chirality 25:206–210, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
So far, no detailed view has been expressed regarding the interactions between vancomycin and racemic compounds including mandelic acid. In the current study, a chiral stationary phase was prepared by using 3-aminopropyltriethoxysilane and succinic anhydride to graft carboxylated silica microspheres and subsequently by activating the carboxylic acid group for vancomycin immobilization. Characterization by elemental analysis, Fourier transform infrared spectroscopy, solid-state nuclear magnetic resonance, and thermogravimetric analysis demonstrated effective functionalization of the silica surface. R and S enantiomers of mandelic acid were separated by the synthetic vancomycin column. Finally, the interaction between vancomycin and R/S mandelic acid enantiomers was simulated by Auto-dock Vina. The binding energies of interactions between R and S enantiomers and vancomycin chiral stationary phase were different. In the most probable interaction, the difference in mandelic acid binding energy was approximately 0.2 kcal/mol. In addition, circular dichroism spectra of vancomycin interacting with R and S enantiomers showed different patterns. Therefore, R and S mandelic acid enantiomers may occupy various binding pockets and interact with different vancomycin functions. These observations emphasized the different retention of R and S mandelic acid enantiomers in vancomycin chiral column.  相似文献   

14.
The biosynthetic origins of the hydrogen, nitrogen and oxygen atoms in the pyrrolidone ring of violacein'were established by an anaylses of the 1H, 13C NMR and MS spectra of its isotope-enriched metabolites. Feeding experiments of [2-2H] and [3-2H2]tryptophans have revealed that the hydrogen in the pyrrolidone ring was derived from the methylene protons at the 3-position of tryptophan. The stereochemical fate of the prochiral hydrogens was determined to be in the retention of the pro-S hydrogen by these feeding experiments using [3R-2H] and [3S”-2H]tryptophans. The incorporation experiment of [α-15N]tryptophan demonstrated that the nitrogen atom in the ring originated from the α-amino group of tryptophan. The incorporation experiment of 18O2 gas verified that all the oxygen atoms of violacein were derived from the molecular oxygen.  相似文献   

15.
Detailed studies were carried out on the binding of the enantiomers of [PtCl2(mepyrr)] (mepyrr = N-methyl-2-aminomethylpyrrolidine) to dG, d(GpG) and a 52-mer oligonucleotide. The pyrrolidine ligand structure was found to be neither sufficiently rigid nor bulky to enforce a single chirality at the exocyclic amine site in this complex, resulting in the presence of diastereomers that complicated the binding studies. Reaction of the (GpG) dinucleotide with R- and S-[PtCl2(mepyrr)] resulted in formation of four [Pt{d(GpG)}(mepyrr)] isomers for each enantiomer as a consequence of the existence of two orientational isomers and two diastereomers. These isomers formed in different amounts most likely as a consequence of the unequal formation of the diastereomers together with stereoselectivity induced by interactions between the dinucleotide and the mepyrr ligand. The [PtCl2(mepyrr)] complexes displayed stereoselectivity and enantioselectivity in their reactions with a 52-mer duplex designed to allow formation of only GpG intrastrand adducts. All four bifunctional adducts formed for each enantiomer, providing further evidence of the lack of directing ability of the ligand in formation of the 1,2-intrastrand adduct. Significant amounts of monofunctional species remained in these assays suggesting that the introduction of the methyl substituent to the exocyclic amine inhibited ring-closure to the bifunctional adduct. This was not sufficient to achieve enantiospecificity, but in the case of the R-enantiomer, one of the bifunctional adducts formed in only small amounts.  相似文献   

16.
Twenty four 12-mer DNA duplexes, each containing a chiral phosphorothioate group successively replacing one of the internucleotide phosphate groups either in the EcoRII recognition site (5CCA/TGG) or near to it, were obtained for studying the interaction of the restriction endonuclease EcoRII with internucleotide DNA phosphates. Twelve of the 12-mer oligonucleotides were synthesized as R p and S p diastereomeric mixtures. Six of them were separated by reversed-phase HPLC using various buffers. Homogeneous diastereomers of the other oligonucleotides were obtained by enzymatic ligation of the R p and S p diastereomers of 5–7-mer oligonucleotides preliminarily separated by HPLC with the corresponding short oligonucleotides on a complementary DNA template.  相似文献   

17.
The enantioselective hydrolysis of (R,S)-3-acetoxymethyl-7,8-difluoro-2,3-dihydro-4H-[1,4]benzoxazine (I) with enzymes was investigated. Optically active I and its hydrolyzate, 7,8-difluoro-2,3-dihydro-3-hydroxymethyl-4H-[1,4]benzoxazine (II), are the intermediates for preparing optically active ofloxacins, whose racemate is known to be an excellent antibacterial agent. Lipoprotein lipase from Pseudomonas fluorescens (LPL Amano 3) was found to predominantly hydrolyze (S)-I, giving (R)-I in 54% e.e. and (R)-II in 44% e.e. On the other hand, lipase from Candida cylindracea was found to predominantly hydrolyze (R)-I, giving (S)-I in 24% e.e. and (S)-II in 20% e.e. Since, the optical purities of I and II thus obtained were not particularly high, these optically active I and II were converted into 3-acetoxymethyl-7,8-difluoro-2,3-dihydro-4-(3,5-dinitrobenzoyl)-4H-[1,4]benzoxazine (IV). After recrystallizing IV from ethyl acetate-hexane, (S)- and (R)-II were obtained with high enantiomeric excess by removing the crystallized racemic IV and subsequently hydrolyzing the resulting optically active IV with alkali. The reduction of II afforded 7,8-difluoro-2,3-dihydro-3-methyl-4H-[1,4]benzoxazine (III), for which the optical purity was estimated to be >96%e.e. by HPLC analysis. (R)- and (S)-ofloxacin were prepared from (R)- and (S)-III with retention of their configuration.  相似文献   

18.
The (R)- and (S)-enantiomers of the chiral herbicide napropamide (NAP) show different biological activities and ecotoxicities. These two enantiomers behave differently in the environment due to enantioselective catabolism by microorganisms. However, the molecular mechanisms underlying this enantioselective catabolism remain largely unknown. In this study, the genes (snaH and snpd) involved in the catabolism of NAP were cloned from Sphingobium sp. B2, which was capable of catabolizing both NAP enantiomers. Compared with (R)-NAP, (S)-NAP was much more rapidly transformed by the amidase SnaH, which initially cleaved the amide bonds of (S)/(R)-NAP to form (S)/(R)-2-(1-naphthalenyloxy)-propanoic acid [(S)/(R)-NP] and diethylamine. The α-ketoglutarate-dependent dioxygenase Snpd, showing strict stereoselectivity for (S)-NP, further transformed (S)-NP to 1-naphthol and pyruvate. Molecular docking and site-directed mutagenesis analyses revealed that when the (S)-enantiomers of NAP and NP occupied the active sites, the distance between the ligand molecule and the coordination atom was shorter than that when the (R)-enantiomers occupied the active sites, which facilitated formation of the transition state complex. This study enhances our understanding of the preferential catabolism of the (S)-enantiomer of NAP on the molecular level.  相似文献   

19.
The aim of the present study was to elucidate the differences in the plasma concentration of two enantiomers of donepezil in Chinese patients with Alzheimer's disease (AD) and investigate in vitro stereoselective metabolism and transport. Donepezil enantiomers were separated and determined by LC‐MS/MS using D5‐donepezil as an internal standard on a Sepax Chiralomix SB‐5 column. In vitro stereoselective metabolism and transport of donepezil were investigated in human liver microsomes and MDCKII‐MDR1 cell monolayer. Pre‐dose (Css‐min) plasma concentrations were determined in 52 patients. The mean plasma level of (R)‐donepezil was 14.94 ng/ml and that of (S)‐donepezil was 23.37 ng/ml. One patient's plasma concentration of (R)‐donepezil was higher than (S)‐donepezil and the ratio is 1.51. The mean plasma levels of (S)‐donepezil were found to be higher than those of (R)‐donepezil in 51 patients and the ratio of plasma (R)‐ to (S)‐donepezil varies from 0.34 to 0.85. In the in vitro microsomal system, (R)‐donepezil degraded faster than (S)‐donepezil. Vmax of (R)‐donepezil was significantly higher than (S)‐donepezil. The P‐gp inhibition experiment shown that the Papp of the two enantiomers was higher than 200 and the efflux ratios were 1.11 and 0.99. The results of the P‐gp inhibition identification experiment showed IC50 values of 35.5 and 20.4 μM, respectively, for the two enantiomers. The results indicate that donepezil exhibits stereoselective hepatic metabolism that may explain the differences in the steady‐state plasma concentrations observed. Neither (R)‐ nor (S)‐donepezil was a P‐gp substance and the two enantiomers are highly permeable through the blood–brain barrier. Chirality 25:498–505, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
Robert H. White 《Chirality》1996,8(4):332-340
The configuration at the C-9 of methanopterin (MPT) has been determined by comparing the circular dichroism (CD) spectra of MPT and its hydrolytic fragment, 1-[4-[[1-(2-amino-7-methyl-4-hydroxy-6-pteridinyl)-ethyl]amino]phenyl]-1-deoxy-D -ribitol (HP-1), with the CD spectra of a series of model compounds of known stereochemistry. These compounds included (S)-6-[1-(4-carboxymethylanilino)ethyl]pterin, (S-6(1-hydroxyethyl)-7-methylpterin, (S-6-(1-hydroxyethyl)pterin, (R)-6-(1-phenoxyethyl)pterin, D (+)-neopterin, and L -biopterin. From this comparison it was concluded that MPT has the R configuration at C-9 and is thus configurationally related to D (+)-neopterin, which has the S configuration at C-1. From previous work establishing the relative stereochemistry at C-6, C-7, and C-9 of N5-N10-methenyl-5,6,7,8-tetrahydromethanopterin (N5-N10-methenyl-H4MPT) as R, S, and R, respectively, it is clear that the remaining asymmetric carbons at C-6 and C-7 of H4MPT have the S and S configuration, respectively. Comparison of these latter two positions to the equivalent carbons in 5,6,7,8-tetrahydrofolate (H4folate) show that the steps involved in the biological reduction of MPT to H4MPT occur with the same stereochemical outcome as those involved in the biological reduction of folate to H4folate. © 1996 Wiley-Liss, Inc.  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