首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Quartz crystal microbalance (QCM) biosensor was used for the chiral recognition of five pairs of enantiomers by using goat serum albumin (GSA) and rabbit serum albumin (RbSA) as chiral selectors. Serum albumin (SA) was immobilized on the QCM through the self‐assembled monolayer technique, and the surface concentration of GSA and RbSA were 8.8 × 10?12 mol cm?2 and 1.2 × 10?11 mol cm?2, respectively. The QCM biosensors showed excellent sensitivity and selectivity. Meanwhile, the chiral recognition of SA sensors was quite species dependent. There were differences between GSA and RbSA sensors in the ability and the preference of chiral recognition. To R,S‐1,2,3,4‐tetrahydro‐1‐naphthylamine (R,S‐1‐TNA), R,S‐1‐(4‐methoxyphenyl)ethylamine (R,S‐4‐MPEA), and R,S‐1‐(3‐methoxyphenyl)ethylamine (R,S‐3‐MPEA), the preference of the stereoselective SA‐drug binding of the two kinds of SA sensors were consistent. However, to R,S‐2‐octanol (R, S‐2‐OT) and R,S‐methyl lactate (R,S‐MEL), the two kinds of SA sensors had opposite chiral recognition preference. Moreover, the interactions of SA and the five pairs of enantiomers have been further investigated through ultraviolet (UV) and fluorescent (FL) spectra. The UV/FL results were in accordance with the consequence of QCM. Chirality 24:804–809, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

2.
Xiong Liu  Yu Ma  Longqi Xu  Qi Liu 《Chirality》2019,31(9):750-758
(S,S)‐DIOP, a common catalyst used in asymmetric reaction, was adopted as chiral extractant to separate 3‐chloro‐phenylglycine enantiomers in liquid‐liquid extraction. The factors affecting extraction efficiency were studied, including metal precursors, organic solvents, extraction temperature, chiral extractant concentration, and pH of aqueous phase. (S,S)‐DIOP‐Pd exhibited good ability to recognize 3‐chloro‐phenylglycine enantiomers, and the operational enantioselectivity (α) is 1.836. The highest performance factor (pf) was obtained under the condition of extraction temperature of 9.1°C, (S,S)‐DIOP‐Pd concentration of 1.7 mmol/L, and pH of aqueous phase of 7.0. In addition, the possible recognition mechanism of (S,S)‐DIOP‐Pd towards 3‐chloro‐phenylglycine enantiomers was discussed.  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.
The present study describes the development of two approaches for the determination of the enantiopurity of both enantiomers of indatraline. Initially, a method was developed using different chiral solvating agents (CSAs) for diastereomeric discrimination regarding signal separation in 1H nuclear magnetic resonance (NMR) spectroscopy, revealing MTPA as a promising choice for the differentiation of the indatraline enantiomers. This CSA was also tested for its ideal molar ratio, temperature, and solvent. Optimized conditions could be achieved that made determination of enantiopurity for (1R,3S)‐indatraline up to 98.9% enantiomeric excess (ee) possible. To quantify even higher enantiopurities, a high‐performance liquid chromatography (HPLC) method based on a modified β‐cyclodextrine phase was established. The influence of buffer type, concentration, pH value, percentage and kind of organic modifier, temperature, injection volume as well as sample solvent on chromatographic parameters was investigated. Afterwards, the reliability of the established HPLC method was demonstrated by validation according to the ICH guideline Q2(R1) regarding specificity, accuracy, precision, linearity, and quantitation limit. The developed method proved to be strictly linear within a concentration range of 1.25–1000 μM for the (1R,3S)‐enantiomer and 1.25‐750 μM for its mirror image that enables a reliable determination of enantiopurities up to 99.75% ee for the (1R,3S)‐enantiomer and up to 99.67% ee for the (1S,3R)‐enantiomer. Chirality 25:923–933, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

5.
Luminescent lanthanide (III) ions have been exploited for circularly polarized luminescence (CPL) for decades. However, very few of these studies have involved chiral samarium (III) complexes. Complexes are prepared by mixing axial chiral ligands (R/S))‐2,2’‐bis(diphenylphosphoryl)‐1,1′‐binaphthyl (BINAPO) with europium and samarium Tris (trifluoromethane sulfonate) (Eu (OTf)3 and Sm (OTf)3). Luminescence‐based titration shows that the complex formed is Ln((R/S)‐BINAPO)2(OTf)3, where Ln = Eu or Sm. The CPL spectra are reported for Eu((R/S)‐BINAPO)2(OTf)3 and Sm((R/S)‐BINAPO)2(OTf)3. The sign of the dissymmetry factors, gem, was dependent upon the chirality of the BINAPO ligand, and the magnitudes were relatively large. Of all of the complexes in this study, Sm((S)‐BINAPO)2(OTf)3 has the largest gem = 0.272, which is one of the largest recorded for a chiral Sm3+ complex. A theoretical three‐dimensional structural model of the complex that is consistent with the experimental observations is developed and refined. This report also shows that (R/S)‐BINAPO are the only reported ligands where gem (Sm3+) > gem (Eu3+).  相似文献   

6.
A simple, precise, accurate, robust chiral high‐performance liquid chromatographic (chiral HPLC) method was developed for estimation of (S)‐piperidin‐3‐amine (S‐isomer) in (R)‐piperidin‐3‐amine dihydrochloride (R‐AMP). As AMP is a high‐melting solid and nonchromophoric compound, development of a suitable chiral method is a challenging task. The proposed chiral HPLC‐UV method involves a precolumn derivatization technique with para toluene sulphonyl chloride (PTSC) in the presence of a base to introduce chromophore into analytes. It utilizes chiralpak AD‐H column with a simple mobile phase of 0.1% diethyl amine in ethanol with a 0.5 mL/min flow rate. Analytes were monitored by using a UV detector at 228 nm. The resolution between the two enantiomers was more than 4.0. The developed method was validated as per current International Conference on Harmonization (ICH) guidelines. Chirality 26:775–779, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
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   

8.
The drug chirality is attracting increasing attention because of different biological activities, metabolic pathways, and toxicities of chiral enantiomers. The chiral separation has been a great challenge. Optimized high‐performance liquid chromatography (HPLC) methods based on vancomycin chiral stationary phase (CSP) were developed for the enantioseparation of propranolol, atenolol, metoprolol, venlafaxine, fluoxetine, and amlodipine. The retention and enantioseparation properties of these analytes were investigated in the variety of mobile phase additives, flow rate, and column temperature. As a result, the optimal chromatographic condition was achieved using methanol as a main mobile phase with triethylamine (TEA) and glacial acetic acid (HOAc) added as modifiers in a volume ratio of 0.01% at a flow rate of 0.3 mL/minute and at a column temperature of 5°C. The thermodynamic parameters (eg, ΔH, ΔΔH, and ΔΔS) from linear van 't Hoff plots revealed that the retention of investigated pharmaceuticals on vancomycin CSP was an exothermic process. The nonlinear behavior of lnk′ against 1/T for propranolol, atenolol, and metoprolol suggested the presence of multiple binding mechanisms for these analytes on CSP with variation of temperature. The simulated interaction processes between vancomycin and pharmaceutical enantiomers using molecular docking technique and binding energy calculations indicated that the calculated magnitudes of steady combination energy (ΔG) coincided with experimental elution order for most of these enantiomers.  相似文献   

9.
Triadimenol is a widely used triazole fungicide and consists of four stereoisomers with 1R,2S, 1S,2R, 1R,2R, and 1S,2S configurations. The trans‐enantiomeric pair (1R,2S‐isomer and 1S,2R‐isomer) is also called triadimenol‐A and the cis‐enantiomeric pair (1R,2R‐isomer and 1S,2S‐isomer) triadimenol‐B. In this study, the stereoselective degradation and chiral stability of triadimenol in two soils were investigated in details. The dissipation of technical triadimenol, a 6:1 mixture of triadimenol‐A and triadimenol‐B, showed significant epimerization from triadimenol‐A to triadimenol‐B occurred along with the dissipation process. The degradation exhibited some stereoselectivity, resulting in a concentration order of 1S,2S > 1R,2R > 1R,2S > 1S,2R or 1S,2S > 1R,2R > 1S,2R > 1R,2S at the end of the 100 days incubation for Baoding soil or Wuhan soil, respectively. Further incubation of triadimenol‐B revealed no epimerization, i.e. triadimenol‐B was configurationally stable in soil, and 1R,2R‐triadimenol degraded slightly slower in the former part and slightly faster in the later part of the incubation than 1S,2S‐triadimenol. Moreover, by incubation of enantiopure 1S,2R‐triadimenol and 1R,2S‐triadimenol, the results documented the epimerization for each enantiomer occurred at both C‐1 and C‐2 positions. Finally, the present work also documented that the enantiomerization reaction for all the four stereoisomers was nearly negligible in the soils. Chirality 25:355‐360:, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

10.
Paclobutrazol, with two stereogenic centers, but gives only (2R, 3R) and (2S, 3S)‐enantiomers because of steric‐hindrance effects, is an important plant growth regulator in agriculture and horticulture. Enantioselective degradation of paclobutrazol was investigated in rat liver microsomes in vitro. The degradation kinetics and the enantiomer fraction were determined using a Lux Cellulose‐1 chiral column on a reverse‐phase liquid chromatography–tandem mass spectrometry system. The t1/2 of (2R, 3R)‐paclobutrazol is 18.60 min, while the t1/2 of (2S, 3S)‐paclobutrazol is 10.93 min. Such consequences clearly indicated that the degradation of paclobutrazol in rat liver microsomes was stereoselective and the degradation rate of (2S, 3S)‐paclobutrazol was much faster than (2R, 3R)‐paclobutrazol. In addition, significant differences between the two enantiomers were also observed in enzyme kinetic parameters. The Vmax of (2S, 3S)‐paclobutrazol was more than 2‐fold of (2R, 3R)‐paclobutrazol and the Clint of (2S, 3S)‐paclobutrazol was higher than that of (2R, 3R)‐paclobutrazol after incubation in rat liver microsomes. These results may have potential implications for better environmental and ecological risk assessment for paclobutrazol. Chirality 27:344–348, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

11.
The rotational barriers ΔG (T) of the four atropisomeric polychlorinated biphenyls (PCBs) 2,2′,3,5′,6-pentachlorobiphenyl (PCB 95), 2,2′3,3′,4,6′-hexachlorobiphenyl (PCB 132), 2,2′,3,3′,6,6′-hexachlorobiphenyl (PCB 136), and 2,2′,3,4′,5′,6-hexachlorobiphenyl (PCB 149) were determined via on-line enantiomerization kinetics by a new stopped-flow multidimensional gas chromatographic technique (stopped-flow MDGC) employing Chirasil-Dex as chiral stationary phase for enantiomer separation. The calculated rotational barriers ΔG (T) of the trichloro-ortho-substituted atropisomers are 184 ± 2 kJ/mol for PCB 95, 189 ± 4 kJ/mol for PCB 132, and 184 ± 1 kJ/mol for PCB 149 at 300°C. The rotational barrier ΔG (T) of tetrachloro-ortho-substituted PCB 136 is at least (or higher than) 210 kJ/mol at 320°C. Chirality 10:316–320, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

12.
《Chirality》2017,29(9):550-557
Resolution of chiral compounds has played an important role in the pharmaceutical field, involving detailed studies of pharmacokinetics, physiological, toxicological, and metabolic activities of enantiomers. Herein, a reliable method by high‐performance liquid chromatography (HPLC) coupled with an optical rotation detector was developed to separate isoborneol enantiomers. A cellulose tris(3, 5‐dimethylphenylcarbamate)‐coated chiral stationary phase showed the best separation performance for isoborneol enantiomers in the normal phase among four polysaccharide chiral packings. The effects of alcoholic modifiers and column temperature were studied in detail. Resolution of the isoborneol racemate displayed a downward trend along with an increase in the content of ethanol and column temperature, indicating that less ethanol in the mobile phase and lower temperature were favorable to this process. Moreover, two isoborneol enantiomers were obtained via a semipreparative chiral HPLC technique under optimum conditions, and further characterized by analytical HPLC, and experimental and calculated vibrational circular dichroism (VCD) spectroscopy, respectively. The solution VCD spectrum of the first‐eluted component was consistent with the Density Functional Theory (DFT) calculated pattern based on the SSS configuration, indicating that this enantiomer should be (1S , 2S , 4S )‐(+)‐isoborneol. Briefly, these results have provided reliable information to establish a method for analysis, preparative separation, and absolute configuration of chiral compounds without typical chromophoric groups.  相似文献   

13.
The increasing demand for biocatalysts in synthesizing enantiomerically pure chiral alcohols results from the outstanding characteristics of biocatalysts in reaction, economic, and ecological issues. Herein, fifteen yeast strains belonging to three food originated yeast species Candida zeylanoides, Pichia fermentans, and Saccharomyces uvarum were tested for their capability for asymmetric reduction of acetophenone to 1‐phenylethanol as biocatalysts. Of these strains, C. zeylanoides P1 showed an effective asymmetric reduction ability. Under optimized conditions, substituted acetophenones were converted to corresponding optically active secondary alcohols in up to 99% enantiomeric excess and at high yields. The preparative scale asymmetric bioreduction of 4‐nitroacetophenone ( 1m ) by C. zeylanoides P1 gave (S)‐1‐(4‐nitrophenyl)ethanol ( 2m ) with 89% yield and > 99% enantiomeric excess. Compound 2m has been obtained in an enantiomerically pure and inexpensive form. Additionally, these results indicate that C. zeylanoides P1 is a promising biocatalyst for the synthesis of chiral alcohols in industry.  相似文献   

14.
To obtain milligram amounts of the enantiomers of benzoxazolinone derivatives to be tested for binding to adrenergic sites, analytical HPLC methods using derivatized amylose chiral stationary phases were developed for the direct enantioseparation of benzoxazolinone aminoalcohols and their aminoketone precursors, derivatives with one or two chirals centers. The separations were made using normal phase methodology with a mobile phase of n‐hexane‐alcohol (ethanol, 1‐propanol, or 2‐propanol) in various proportions, and silica‐based amylose (tris‐3, 5‐dimethylphenylcarbamate) Chiralpak AD and (tris‐(S)‐1‐phenylethylcarbamate) Chiralpak AS columns. The effects of concentration of various aliphatic alcohols in the mobile phase were studied. The best separation was achieved on Chiralpak AS, so preparative HPLC was set up with this chiral stationary phase using a mobile phase consisting of n‐hexane‐alcohol using isocratic conditions and multiple repetitive injections. Physicochemicals properties of enantiomers were reported The effect of structural features of the solutes on discrimination between the enantiomers was examined. Limit of detection (LD) and limit of quantification (LQ) were determined using both ultra‐violet (UV) and evaporative light‐scattering detection (ELSD). Chirality, 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

15.
Enantioseletive toxicities of chiral pesticides have become an environmental concern recently. In this study, we evaluated the enantiomeric separation of salithion on a suite of commercial chiral columns and assessed the toxicity of enantiomers toward butyrylcholinesterase and Daphnia magna. Satisfactory separations of salithion enantiomers could be achieved on all tested columns, that is, Chiralcel OD, Chiralcel OJ, and Chiralpak AD column. However, the Chiralpak AD column offered the best separation and was chosen to prepare micro‐scale of pure salithion enantiomers for subsequent bioassays. The first and second enantiomers eluted on the Chiralpak AD column were further confirmed to be (?)‐S‐salithion and (+)‐R‐salithion, respectively. The half inhibition concentrations to butyrylcholinesterase of racemate, (+)‐R‐salithion, and (?)‐S‐salithion were 33.09, 2.92, and 15.60 mg/l, respectively, showing (+)‐R‐enantiomer being about 5.0 times more potent than its (?)‐S‐form. However, the median lethal concentrations (96 h) of racemate, (+)‐R‐salithion, and (?)‐S‐salithion toward D. magna were 3.54, 1.10, and 0.36 μg/l, respectively, suggesting that (?)‐S‐salithion was about 3.0 times more toxic than (+)‐R‐form. Racemic salithion was less toxic than either of the enantiomers in both bioassays, suggesting that antagonistic interactions might occur between the enantiomers during the toxication action. This work reveals that the toxicity of salithion toward butyrylcholinesterase and D. magna is enantioselective, and this factor should be taken into consideration in the environmental risk assessment of salithion. Chirality 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

16.
Four new tirucallane triterpenoids, (21S,23R,24R)‐21,23‐epoxy‐21,24‐dihydroxy‐25‐methoxytirucall‐7‐en‐3‐one ( 2 ), (3S,21S,23R,24S)‐21,23‐epoxy‐21,25‐dimethoxytirucall‐7‐ene‐3,24‐diol ( 8 ), (21S,23R,24R)‐21,23‐epoxy‐24‐hydroxy‐21‐methoxytirucalla‐7,25‐dien‐3‐one ( 11 ), and (21S,23R,24R)‐21,23‐epoxy‐21,24‐dihydroxytirucalla‐7,25‐dien‐3‐one ( 12 ), along with 16 known analogues, 1 , 3  –  7 , 9  –  10 , and 13  –  20 , were isolated from the fruits of Melia azedarach. Their structures were elucidated by spectroscopic methods including 1D‐ and 2D‐NMR techniques and mass spectrometry. These compounds were evaluated for their cytotoxicities against HepG2 (liver), SGC7901 (stomach), K562 (leukemia), and HL60 (leukemia) cancer cell lines. Compound 20 exhibited potent cytotoxicity against HepG2 and SGC7901 cancer cells with the IC50 values of 6.9 and 6.9 μm , respectively.  相似文献   

17.
In our effort towards measuring the parity violation energy difference between two enantiomers, a simple chiral oxorhenium complex 5 bearing enantiopure 2‐mercaptocyclohexan‐1‐ol has been prepared as a potential candidate species. Vibrational circular dichroism revealed a chiral environment surrounding the rhenium atom, even though the rhenium is not a stereogenic center itself, and enabled to assign the (1S,2S)‐(?) and (1R,2R)‐(+) absolute configuration for 5 . For both compound 5 and complex 4 , previously studied by us and bearing a propane‐2‐olato‐3‐thiolato ligand, relativistic calculations predict parity violating vibrational frequency differences of a few hundreds of millihertz, above the expected sensitivity attainable by a molecular beam Ramsey interferometer that we are constructing.  相似文献   

18.
The stratum corneum, the rate‐limiting barrier in transdermal drug delivery, is chiral in nature and enantiomers behave differently with respect to their transport across the skin, resulting in enantioselective permeation. The permeation characteristics of individual enantiomers of metoprolol free base (MB) were investigated using hairless mouse skin. The influence of chiral permeation enhancers, l‐menthol and (±)‐linalool, on the permeation of MB was also investigated. In the absence of enhancers, the permeation profiles of R‐ and S‐MB from donor solutions containing either RS‐MB or pure enantiomers are comparable (p < 0.05). In presence of enhancers, l‐menthol and (±)‐linalool, the flux values were increased 2.4‐ to 3.0‐fold, respectively, and the permeation profiles of R‐ and S‐MB from donor solutions containing RS‐MB are comparable (p < 0.05). However, when donor vehicle contains pure enantiomers, the permeation enhancing effect of l‐menthol on S‐MB was significantly higher (by 25%) than on R‐MB (p < 0.05). Further, in presence of l‐menthol, the flux of S‐MB from donor solution containing pure S‐MB was 35% higher than the flux of RS‐MB from racemate. No such effect was seen with (±)‐linalool. In all the investigations, no enantiomeric inversion was observed during the permeation process. The lag times were shorter in the case of l‐menthol compared with (±)‐linalool. Chirality 11:536–540, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

19.
As an important intermediate of prostaglandins and entecavir, optically pure Corey lactone diol (CLD) has great value in the pharmaceutical industry. In this work, the enantioseparation of (±)‐CLD was evaluated using high‐performance liquid (HPLC) and supercritical fluid chromatography (SFC). In HPLC, the separations of CLD enantiomers on polysaccharide‐based chiral stationary phases with both normal phase and polar organic phase were screened. And the conditions for the enantioseparation were optimized in HPLC and SFC, including the selection of mobile phase, temperature, back‐pressure, and other conditions. More important, it was found that the chiral resolutions were greatly enhanced by the increase of the coating amount of ADMPC (amylose tris‐(3,5‐dimethylphenylcarbamate)) under both HPLC and SFC conditions, which can lead to the increase of the productivity and the decrease of the solvent consumption. The preparations of optically pure CLD were evaluated on a semi‐preparative (2 × 25 cm) column packed with 30% ADMPC‐coated CSP under HPLC and SFC conditions. Preparative performances in terms of kkd are 1.536 kg racemate/kg CSP/day and 1.248 kg racemate/kg CSP/day in HPLC and SFC, respectively.  相似文献   

20.
Chiral considerations are found to be very much relevant in various aspects of forensic toxicology and pharmacology. In forensics, it has become increasingly important to identify the chirality of doping agents to avoid legal arguments and challenges to the analytical findings. The scope of this study was to develop an liquid chromatography–mass spectrometry (LCMS) method for the enantiomeric separation of typical illicit drugs such as ephedrines (ie, 1S,2R(+)‐ephedrine and 1R,2S(?)‐ephedrine) and pseudoephedrine (ie, R,R(?)‐pseudoephedrine and S,S(+)‐pseudoephedrine) by using normal phase chiral liquid chromatography–high‐resolution mass spectrometry technique. Results show that the Lux i‐amylose‐1 stationary phase has very broad and balancing‐enantio‐recognition properties towards ephedrine analogues, and this immobilized chiral stationary phase may offer a powerful tool for enantio‐separation of different types of pharmaceuticals in the normal phase mode. The type of mobile phase and organic modifier used appear to have dramatic influences on separation quality. Since the developed method was able to detect and separate the enantiomers at very low levels (in pico grams), this method opens easy access for the unambiguous identification of these illicit drugs and can be used for the routine screening of the biological samples in the antidoping laboratories.  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