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.
Flurbiprofen (F) is a nonsteroidal anti‐inflammatory drug (NSAID) used therapeutically as the racemate of (R)‐enantiomer and (S)‐enantiomer. The inversion of RF to SF and vice versa was investigated in C57Bl/6 and SJL mice and Dark Agouti and Lewis rats. The enzyme α‐methylacyl‐CoA racemase (AMACR) is involved in the chiral inversion pathway that converts members of the 2‐arylpropionic acid NSAIDs from the R‐enantiomer to the S‐enantiomer. We studied C57Bl/6 mice deficient in AMACR postulating that they should show reduced inversion of RF to SF. In line with the data of others in mice, (R)‐inversion to (S)‐inversion was relatively high in both the C57Bl/6 and SJL mice (fraction inverted, FI = 37.7% and 24.7%, respectively). In contrast, in AMACR deficient mice, there was no measurable peak for SF after administration of RF. The results in both rat strains (Dark Agouti and Lewis rats, FI = 1.4% and 4.1%, respectively) confirm the low chiral inversion of the enantiomers of flurbiprofen in the rat, as observed by other authors in the Sprague‐Dawley strain (<5%). From the present results, we conclude that for the study of flurbiprofen enantiomers, the rat is more suitable than the mouse as a model for the human in which (R)‐inversion to (S)‐inversion is negligible.  相似文献   

2.
The separation of rac‐o‐chloromandelic acid 1 with enantiopure aryloxypropylamine via diastereomeric salt formation was investigated. (R)‐o‐chloromandelic acid (R)‐ 1 , a key intermediate for the antithrombotic agent clopidogrel, was obtained in 65% yield and 98% ee by Dutch resolution of rac‐ 1 with (S)‐2‐hydroxyl‐3‐(p‐chlorophenoxy) propylamine (S)‐ 5 as resolving agent and (S)‐2‐hydroxyl‐3‐(o‐nitrophenoxy) propylamine (S)‐ 4 as nucleation inhibitor. Chirality 24:1013–1017, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
Stereoselective metabolism of propranolol side‐chain glucuronidation was studied for two recombinant human uridine diphosphate glucuronosyltransferases (UGTs), UGT1A9 and UGT2B7. The S‐ and R‐propranolol side‐chain glucuronides produced in the incubation mixtures were assayed simultaneously by RP‐HPLC with fluorescent detector. The excitation and emission wavelengths were set at 310 nm and 339 nm, respectively. UGT1A9 prefers catalyzing S‐enantiomer to R‐enantiomer and the intrinsic clearance (CLint) ratios of S‐enantiomer to R‐enantiomer are 3.8 times and 6.5times for racemic propranolol and individual enantiomers, respectively. UGT2B7, however, catalyzes slightly less S‐enantiomer than R‐enantiomer and the CLint ratio of S‐enantiomer to R‐enantiomer is 0.8 times. The high concentration of racemic propranolol (>0.57 mmol/l) and individual enantiomers (>0.69 mmol/l) exhibited substrate inhibition of glucuronidation for UGT2B7, but only the S‐enantiomer (>0.44 mmol/l) in racemic propranolol exhibited substrate inhibition for UGT1A9. The substrate inhibition constants (Ksi) were all similar (P > 0.05). Drug–drug interactions were also found between S‐ and R‐enantiomer glucuronidation metabolisms by UGT1A9 and UGT2B7. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

4.
A simple, sensitive, and robust normal‐phase isocratic HPLC‐UV method was developed and validated for the enantiomeric separation of rasagiline mesylate and its (S)‐enantiomer. The rasagiline and its (S)‐enantiomer were resolved on a Chiralcel‐OJ‐H (4‐methylbenzoate cellulose coated on silica) column using a mobile phase consisting of n‐hexane:isopropyl alcohol:ethanol:diethyl amine (96:2:2:0.01) at a flow rate of 1.0 ml/min. The column temperature was maintained at 27 °C and elution was monitored at 215 nm. The resolution (Rs) between the enantiomers was found to be more than 2.0. The limit of detection and the limit of quantification of the (S)‐enantiomer were found to be 0.35 and 1.05 µg/ml, respectively. The developed method was validated as per ICH guidelines with respect to linearity, limit of detection and quantification, accuracy, precision, and robustness—and satisfactory results were obtained. The sample solution and mobile phase were found to be stable up to 48 h. The method is useful for routine evaluation of the quality of rasagiline mesylate in bulk drug‐manufacturing units. Chirality 25:324–327, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

5.
(S)-1,4-benzodioxan-2-carboxylic acid-1 is used as starting compound for the production of the more effective (S) enantiomer of the drug doxazosin mesylate. The catalytic ability of some commercial lipases for preparations of (S) enantiomer of 1 from (±) methyl 1,4-benzodioxin-2-carboxylate-2 is reported. Lipases from bacterial sources were more successful in resolving the ester than those from the yeast lipases. About 85% enantiomerically pure ester was achieved by lipase from alcaligenes sp.  相似文献   

6.
Chiral discrimination observed in high‐performance liquid chromatography (HPLC) with the novel chiral stationary phase (CSP‐18C6I) derived from (+)‐(R)‐18‐crown‐6 tetracarboxylic acid [(+)‐18C6H4] was investigated by X‐ray crystallographic analysis of the complex composed of the R‐enantiomer of 1‐(1‐naphthyl)ethylamine (1‐NEA) and (+)‐18C6H4. Mixtures of 1‐NEA (the R‐ or S‐enantiomer) and (+)‐18C6H4 were dissolved in methanol‐water (1:1) solution and allowed to stand for crystallization. The R‐enantiomer crystallized with (+)‐18C6H4 as a co‐crystal, although the S‐enantiomer did not. This result was in good agreement with the enantiomer elution order of 1‐NEA in CSP‐18C6I. The apparent binding constants (Ka) of the enantiomers to the (+)‐18C6H4 obtained from 1H‐NMR experiments also supported the above‐mentioned result. The X‐ray crystal structure of the 1:1 complex of the R‐enantiomer and (+)‐18C6H4 indicated the four sets of hydrogen bond association between the naphthylethylammonium cation and oxygen of polyether ring or carbonyl group of (+)‐18C6H4. Chirality 11:173–178, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

7.
《Chirality》2017,29(7):348-357
Imazethapyr (IM) is a chiral herbicide composed of an (−)‐R‐enantiomer and an (+)‐S‐enantiomer with differential herbicidal activity. In this study, the effects of microbial organisms, humidity, and temperature on the selective degradation of the (−)‐R‐ and (+)‐S‐enantiomers of IM were determined in silty loam (SL) and clay loam (CL) soil with different pH values. The (−)‐R‐enantiomer of IM was preferentially degraded in two soils under different microorganism, humidity, and temperature conditions. The average half‐lives of R‐IM ranged from 43 to 66.1 days and were significantly shorter (P <  0.05) than those of S‐IM, which ranged from 51.4 to 79.8 days. The enantiomer fraction (EF = (+)‐S‐enantiomer/((−)‐R‐enantiomer + (+)‐S‐enantiomer)) values were used to describe the enantioselectivity of degradation of IM were >0.5 (P <  0.05) in two unsterilized soils under different humidity and temperature conditions. The highest EF values were observed at unsterilized CL soil samples under 50% maximum water‐holding capacity (MWHC) and 25 °C environmental conditions. The EF values of the IM enantiomers were significantly higher (P <  0.05) in CL soils (higher pH = 5.81) and were 0.581 (unsterilized) and 0.575 (50% MWHC; 25 °C) compared with those recorded in SL soil (lower pH = 4.85). In addition, this study revealed that microbial organisms preferentially utilized the more herbicidal active IM enantiomer.  相似文献   

8.
《Chirality》2017,29(10):623-633
3‐Ethyl‐3‐phenylpyrrolidin‐2‐one ( EPP) is an experimental anticonvulsant based on the newly proposed α‐substituted amide group pharmacophore. These compounds show robust activity in animal models of drug‐resistant epilepsy and are thus promising for clinical development. In order to understand pharmaceutically relevant properties of such compounds, we are conducting an extensive investigation of their structures in the solid state. In this article, we report chiral high‐performance liquid chromatography (HPLC) separation, determination of absolute configuration of enantiomers, and crystal structures of EPP. Preparative resolution of EPP enantiomers by chiral HPLC was accomplished on the Chiralcel OJ stationary phase in the polar‐organic mode. Using a combination of electronic CD spectroscopy and anomalous dispersion of X‐rays we established that the first‐eluted enantiomer corresponds to (+)‐(R )‐EPP, while the second‐eluted enantiomer corresponds to (−)‐(S )‐EPP. We also demonstrated that, in the crystalline state, enantiopure and racemic forms of this anticonvulsant have considerable differences in their supramolecular organization and patterns of hydrogen bonding. These stereospecific structural differences can be related to the differences in melting points and, correspondingly, solubility and bioavailability.  相似文献   

9.
In order to define an enantioselective nuclear magnetic resonance (NMR) method for the antiasthmatic drug montelukast, a series of nine easily available products were evaluated as NMR chiral solvating agents (CSAs): D‐dibenzoyltartaric acid, D‐ditoluoyltartaric acid, (+)‐camphorsulfonic acid, (S)‐BINOL, (S)‐3,3’‐diphenyl‐2,2’‐binaphthyl‐1,1’‐diol, (R)‐3,3'′‐di‐9‐anthracenyl‐1,1'′‐bi‐2‐naphthol, (R)‐3,3'′‐di‐9‐phenanthrenyl‐1,1'′‐bi‐2‐naphthol, Pirkle's alcohol, and (?)‐cinchonidine. It was proved that most of the studied agents constitute diastereomeric complexes with both drug enantiomers in CD2Cl2 or CDCl3 solutions, thus permitting the direct 1H NMR detection of the unwanted S‐enantiomer, even at levels of 0.75%. (?)‐Cinchonidine was found to be the more convenient CSA in terms of NMR enantiodiscrimination power and ease of experimental requirements. The final method was validated and applied to the fast monitoring of the optical purity of montelukast “in‐process” samples, circumventing the need for tedious and slower analytical procedures like enantioselective chromatography or capillary electrophoresis. In addition, a method for the enantiopurity control of the commercial drug (montelukast sodium salt) was also established using (S)‐BINOL as NMR CSA. Chirality 25: 780–786, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

10.
In this study, the stereoselective pharmacokinetics of doxazosin enantiomers and their pharmacokinetic interaction were studied in rats. Enantiomer concentrations in plasma were measured using chiral high‐pressure liquid chromatography (HPLC) with fluorescence detection after oral or intravenous administration of (–)‐(R)‐doxazosin 3.0 mg/kg, (+)‐(S)‐doxazosin 3.0 mg/kg, and rac‐doxazosin 6.0 mg/kg. AUC values of (+)‐(S)‐doxazosin were always larger than those of (–)‐(R)‐doxazosin, regardless of oral or intravenous administration. The maximum plasma concentration (Cmax) value of (–)‐(R)‐doxazosin after oral administration was significantly higher when given alone (110.5 ± 46.4 ng/mL) versus in racemate (53.2 ± 19.7 ng/mL), whereas the Cmax value of (+)‐(S)‐doxazosin did not change significantly. The area under the curve (AUC) and Cmax values for (+)‐(S)‐doxazosin after intravenous administration were significantly lower, and its Cl value significantly higher, when given alone versus in racemate. We speculate that (–)‐(R)‐doxazosin increases (+)‐(S)‐doxazosin exposure probably by inhibiting the elimination of (+)‐(S)‐doxazosin, and the enantiomers may be competitively absorbed from the gastrointestinal tract. In conclusion, doxazosin pharmacokinetics are substantially stereospecific and enantiomer–enantiomer interaction occurs after rac‐administration. Chirality 27:738–744, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

11.
Chiral distinction in the proton pump inhibitor drugs omeprazole and in its chiral‐switch esomeprazole magnesium was studied employing the Density Functional Theory (DFT) method. At B3LYP/6‐311G(d,p), the 6‐methoxy???6‐methoxy and 5‐methoxy???5‐methoxy homochiral and heterochiral dimers were calculated. The chiral distinction free energies (ΔΔG298,(RS‐SS)) between the cyclic C2‐(S,S)‐ and Ci‐(R,S)‐dimers with two intermolecular hydrogen bonds are 3.8, 1.9 (with BSSE counterpoise correction), and –6.9 (with D3 dispersion and BSSE counterpoise corrections) kJ/mol. Adding water as an implicit solvent (polarized continuum model [PCM] model) resulted in a chiral distinction energy of –3.3 kJ/mol, indicating a reversal of the order of the relative stabilities of C2‐(S,S)‐ and Ci‐(R,S)‐dimers. The chiral distinction free energies between the corresponding (less stable) C1‐dimers with one intermolecular hydrogen bond are –9.3, –5.8 (with BSSE CC), 17.6 (D3 + BSSE CC), and –3.2 (H2O) kJ/mol. The results highlight the contention that omeprazole is not just a superposition of its enantiomer constituents. They are consistent with the pharmacological evidence of enantiomer–enantiomer interactions in omeprazole versus esomeprazole and the differences between the drugs omeprazole and esomeprazole magnesium and support the lodged application for regulatory supplementary protection certificate (SPC) exclusivity for the esomeprazole‐related combination drug Vimovo. Chirality 26:214–227, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

12.
It was shown that racemic (±)‐ 2 [1′‐benzyl‐3‐(3‐fluoropropyl)‐3H‐spiro[[2]benzofuran‐1,4′‐piperidine], WMS‐1813 ] represents a promising positron emission tomography (PET) tracer for the investigation of centrally located σ1 receptors. To study the pharmacological activity of the enantiomers of 2 , a preparative HPLC separation of (R)‐2 and (S)‐2 was performed. The absolute configuration of the enantiomers was determined by CD‐spectroscopy together with theoretical calculations of the CD‐spectrum of a model compound. In receptor binding studies with the radioligand [3H]‐(+)‐pentazocine, (S)‐2 was thrice more potent than its (R)‐configured enantiomer (R)‐2 . The metabolic degradation of the more potent (S)‐enantiomer was considerably slower than the metabolism of (R)‐2 . The structures of the main metabolites of both enantiomers were elucidated by determination of the exact mass using an Orbitrap‐LC‐MS system. These experiments showed a stereoselective biotransformation of the enantiomers of 2 . Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
The acetylcholinesterase inhibition by enantiomers of exo‐ and endo‐2‐norbornyl‐Nn‐butylcarbamates shows high stereoselelectivity. For the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐exo‐2‐norbornyl‐Nn‐butylcarbamates, the R‐enantiomer is more potent than the S‐enantiomer. But, for the acetylcholinesterase inhibitions by (R)‐(+)‐ and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates, the S‐enantiomer is more potent than the R‐enantiomer. Optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norbornyl‐Nn‐butylcarbamates are synthesized from condensations of optically pure (R)‐(+)‐exo‐, (S)‐(?)‐exo‐, (R)‐(+)‐endo‐, and (S)‐(?)‐endo‐2‐norborneols with n‐butyl isocyanate, respectively. Optically pure norborneols are obtained from kinetic resolutions of their racemic esters by lipase catalysis in organic solvent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

14.
Plant material is a rich source of valuable compounds such as flavanones. Their different forms influence bioavailability and biological activity, causing problems with the selection of plant material for specific purposes. The purpose of this research was to determine selected flavanone (eriodictyol, naringenin, liquiritigenin, and hesperetin) enantiomer contents in free form and bonded to glycosides by an RP‐UHPLC‐ESI‐MS/MS method. Different parts (stems, leaves, and flowers) of goldenrod (Solidago virgaurea L.), lucerne (Medicago sativa L.), and phacelia (Phacelia tanacetifolia Benth.) were used. The highest content of eriodictyol was found in goldenrod flowers (13.1 μg/g), where it occurred mainly as the (S)‐enantiomer, and the greatest proportion of the total amount was bonded to glycosides. The richest source of naringenin was found to be lucerne leaves (4.7 μg/g), where it was mainly bonded to glycosides and with the (S)‐enantiomer as the dominant form. Liquiritigenin was determined only in lucerne, where the flowers contained the highest amount (1.2 μg/g), with the (R)‐enantiomer as dominant aglycone form and the (S)‐enantiomer as the dominant glycosylated form. The highest hesperetin content was determined in phacelia leaves (0.38 μg/g), where it was present in the form of a glycoside and only as the (S)‐enantiomer. A comparison of the different analyte forms occurring in different plant parts was performed for the first time.  相似文献   

15.
A new and accurate HPLC method using β‐cyclodextrin chemically bonded to spherical silica particles as chiral stationary phase (CSP) was developed and validated for determination of S‐clopidogrel and its impurities R‐enantiomer and S‐acid as a hydrolytic product. The effects of acetonitrile and methanol content in the mobile phase and temperature on the resolution and retention of enantiomers were investigated. A satisfactory resolution of S‐clopidogrel active form and its impurities was achieved on ChiraDex® column (5 μm, 4 × 250 mm) at a flow rate of 1.0 ml/min and 17°C using acetonitrile, methanol and 0.01 M potassium dihydrogen phosphate solution (15:5:80 v/v/v) as mobile phase. The detection wavelength was set at 220 nm. The method was validated in terms of accuracy, precision, linearity, and robustness. The limit of detection for R‐enantiomer and S‐acid were 0.75 and 0.09 μg/ml, respectively, injection volume being 20 μl. Finally, the molecular modeling of the inclusion complexes between the analytes and β‐cyclodextrin was performed to investigate the mechanism of the enantiorecognition and to study the quantitative structure–retention relationships. Chirality, 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

16.
The sex pheromone of the cloaked pug moth, Eupithecia abietaria Götze, an important cone‐feeding pest in spruce seed orchards in Europe, was investigated. Chemical and electrophysiological analyses of pheromone gland extracts of female moths and analogous analyses of synthetic hydrocarbons and epoxides of chain length C19 and C21 revealed (3Z,6Z,9Z)‐3,6,9‐nonadecatriene (3Z,6Z,9Z‐19:H) and 3Z,6Zcis‐9,10‐epoxynonadecadiene (3Z,6Zcis‐9,10‐epoxy‐19:H) as candidate pheromone components, which were found in a gland extract in a ratio of 95 : 5. In field trapping experiments, conspecific males were only attracted to a combination of 3Z,6Z,9Z‐19:H and the (9S,10R)‐enantiomer of 3Z,6Zcis‐9,10‐epoxy‐19:H. The (9R,10S)‐enantiomer was not attractive, which is in agreement with studies on other Eupithecia species, for which males have only been attracted by the (9S,10R)‐enantiomer of epoxides. Subsequent experiments showed that E. abietaria males were attracted to a wide range of ratios of the two active compounds and that trap catches increased with increasing dose of the binary blend. A two‐component bait containing 300 μg 3Z,6Z,9Z‐19:H and 33 μg of the (9S,10R)‐enantiomer of 3Z,6Zcis‐9,10‐epoxy‐19:H was efficient for monitoring E. abietaria in spruce seed orchards in southern Sweden, where this species has probably been overlooked as an important pest in the past. With sex pheromones recently identified for two other moths that are major pests on spruce cones, the spruce seed moth, Cydia strobilella L., and the spruce coneworm, Dioryctria abietella Denis & Schiffermüller, pheromone‐based monitoring can now be achieved for the whole guild of cone‐feeding moths in European spruce seed orchards.  相似文献   

17.
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   

18.
《Chirality》2017,29(5):167-171
The racemic pterocarpanquinone LQB‐118 is active, in mice and hamsters, against tegumentary and visceral leishmaniasis. This compound also presents antiinflammatory and antineoplastic activity in mice. The low level of toxicity observed in these studies makes LQB‐118 a promising drug candidate. In order to conduct further biological testing to investigate enantioselectivity in the above‐mentioned activities, a multimilligram amount of each enantiomer of LQB‐118 was produced. Furthermore, vibrational circular dichroism (VCD) and Density Functional Theory (DFT) calculations were used to determine unambiguously their absolute configurations. The comparison of experimental and calculated VCD data led to the assignment of (−)‐LQB‐118 as 7aR ,12aR and, consequently, (+)‐LQB‐118 as 7aS 12aS .  相似文献   

19.
The pharmacokinetics of (?)‐N‐(trans‐4‐isopropylcyclohexanecarbonyl)‐D ‐phenylalanine (nateglinide) and its enantiomer (L‐enantiomer) was studied in Goto‐Kakizaki (GK) rats after intravenous administration of nateglinide or L‐enantiomer at a dose of 40 μmol/kg body weight. Nateglinide, its L‐enantiomer and their metabolites in serum, bile and urine were determined. The total clearance (CLtot) and the volume of distribution (Vd) was slightly higher for nateglinide than those for L‐enantiomer in control rats, although the differences were not statistically significant. The cumulative excretions of L‐M1 (major metabolite of L‐enantiomer) and L‐M2 (major metabolite of L‐enantiomer) into bile were almost the same as that of M1 (major metabolite of nateglinide)and M2 (major metabolite of nateglinide). In GK rats, CLtot and Vd were higher for nateglinide than those for L‐enantiomer. The cumulative excretion of L‐M1 and L‐M2 were not different from those of M1 and M2, respectively, into bile or urine. CLtot and Vd for nateglinide or L‐enantiomer in GK rats were not different from those in control rats. The total excretion of M1, M2, L‐M1, and L‐M2 into bile or urine in GK rats was not substantially different from that of control rats. These results suggest that the L‐enantiomer of nateglinide shows higher CLtot and Vd compared with nateglinide, especially in the diabetic state. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

20.
Two new chiral mononuclear Mn(III) complexes, [Mn L ( R )Cl (C2H5OH)]?C2H5OH ( 1 ) and [Mn L ( S ) (CH3OH)2]Cl?CH3OH ( 2 ), {H2 L = (R,R)‐or (S,S)‐N,N’‐bis‐(2‐hydroxy‐1‐naphthalidehydene)‐cyclohexanediamine} were synthesized and characterized by various physicochemical techniques. Bond valence sum (BVS) calculations and the Jahn‐Teller effect indicate that the Mn centers are in a +3 oxidation state. The statuses of the two complexes in the solution were confirmed as a pair of enantiomers by electrospray ionization, mass spectrometry (ESI‐MS) spectrum. The binding ability of the complexes with calf thymus CT‐DNA was investigated by spectroscopic and viscosity measurements. Both of the complexes could interact with CT‐DNA via an intercalative mode with the order of 1 ( R ‐enantiomer) > 2 ( S ‐enantiomer). Under the physiological conditions, the two compounds exhibit efficient DNA cleavage activities without any external agent, which also follows the order of R ‐enantiomer > S ‐enantiomer. Interestingly, the concentration‐dependent DNA cleavage experiments indicate an optimal concentration of 17.5 μM. In addition, the interaction of the compounds with bovine serum albumin (BSA) was also investigated, which indicated that the complexes could quench the intrinsic fluorescence of BSA by a static quenching mechanism. Chirality 27:142‐150, 2015. © 2014 Wiley Periodicals, 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