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.
Enantiomeric and racemic oxazepam (OX), 3-O-methyloxazepam (MeOX), and 3-O-ethyloxazepam (EtOX) were used to study racemization, heteronucleophilic, and homonucleophilic substitution reactions in anhydrous acidic methanol and ethanol. Kinetics of racemization and nucleophilic substitution reactions in nondeuterated and deuterated solvents were determined by circular dichroism spectropolarimetry, chiral stationary phase high-performance liquid chromatography (HPLC), reversed-phase HPLC, and mass spectrometry. Several reactions occurred when (S)-OX, for example, was dissolved in acidic methanol: (1) (S)-OX itself underwent spontaneous racemization, (2) the 3-hydroxyl group of (S)-OX was stereoselectively substituted by the methoxy group of methanol to form MeOX enriched in (S)-MeOX, (3) the 3-methoxy group of (S)-MeOX was stereoselectively substituted by the methoxy group of methanol to form MeOX enriched in (S)-MeOX, and (4) the 3-methoxy group of (R)-MeOX was stereoselectively substituted by the methoxy group of methanol to form MeOX enriched in (R)-MeOX. Repetitive reactions 3 and 4 eventually resulted in a racemic MeOX. Similar reactions occurred for an enantiomeric OX in acidic ethanol. © 1996 Wiley-Liss, Inc.
  • 1 This article is a US Government work and, as such, is in the public domain in the United States of America.
  •   相似文献   

    2.
    Shen K. Yang 《Chirality》1995,7(5):365-375
    Oxazepam (OX), 3-O-methyloxazepam, 3-O-ethyloxazepam, temazepam (TMZ), 3-O-methyltemazepam, and 3-O-ethyltemazepam underwent acid-catalyzed nucleophilic substitution reaction (hydrolysis) in an acetonitrile–oxygen-18 water mixture to form either OX or TMZ in which the 3-hydroxyl group was either partially or fully labeled with an oxygen-18 atom. The dependence of the hydrolysis rates on solvent composition, temperature, ionic strength, and in deuterated solvent was studied by reversed-phase high-performance liquid chromatography (HPLC). The rates of racemization of enantiomeric compounds in acidic aqueous solutions were studied by both spectropolarimetry and chiral stationary phase HPLC. In acetonitrile: 2.5 M H2SO4 (4:1, v/v) at 50°C, enantiomers of OX and TMZ underwent racemization at rates ≥40-fold faster than the rates of hydrolysis. Enantiomeric 3-O-alkyl derivatives of OX and TMZ in acidic aqueous solutions did not themselves undergo racemization and it was their hydrolysis products (either OX or TMZ) that underwent racemization. © 1995 Wiley-Liss, Inc.  相似文献   

    3.
    Shen K. Yang 《Chirality》1994,6(3):175-184
    Enantiomers of 3-O-methyloxazepam (MeOX) and 3-O-ethyloxazepam (EtOX) were resolved by chiral stationary phase high-performance liquid chromatography (CSP-HPLC). Reaction kinetics and deuterium isotope effects of acid-catalyzed racemization of enantiomeric MeOX in ethanol and enantiomeric EtOX in methanol were studied by spectropolarimetry. The acid-catalyzed heteronucleophilic substitution reactions of racemic MeOX in ethanol and racemic EtOX in methanol were studied by reversed-phase HPLC. Thermodynamic parameters involved in the reactions were obtained by temperature-dependent reaction rates. The effects of solvent's dielectric constant on the heteronucleophilic substitution reactions were also determined. A nucleophilically solvated and transient C3 carbocation intermediate resulting from an N4-protonated enantiomer, derived from a 1,4-benzodiazcpine either in M (minus) or P (plus) conformation, is proposed to be an intermediate and responsible for the acid-catalyzed stereoselective nucleophilic substitution and the resulting racemization. © 1994 Wiley-Liss, Inc.  相似文献   

    4.
    An extensive study of the behavior of three chiral stationary phases (CSP) used in liquid chromatography (LC) is presented for the stereoselective determination of methadone. The following chromatographic columns were selected: a cellulose, Chiralcel OJ; a modified cyclodextrin, Cyclobond I 2000 RSP, and a protein, Chiral‐AGP. Retention factors, enantioselectivity, efficiency, and resolution were tested by modifying the composition of the mobile phase as well as the temperature. The mechanism for the chiral recognition of methadone on each support was discussed. Optimal chromatographic parameters were obtained for the three supports tested, and methadone enantiomers were separated in less than 20 minutes. The cellulose‐based column gave the best resolution, but this CSP was not adapted to clinical analyses of methadone. Under optimized conditions, the cyclodextrin‐ and protein‐based columns allowed an excellent separation of methadone enantiomers, but no interference with the primary metabolite was found only with Chiral‐AGP. Chirality 11:319–325, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

    5.
    In this study an enantioseparation method for rac‐bambuterol (5‐(2‐(tert‐butylamino)‐1‐hydroxyethyl)‐1,3‐phenylene bis(dimethylcarbamate)) via diastereoisomeric salt formation with o‐chloromandelic acid was developed. The enantiomeric excess (ee) values and chemical purities of the desired products were confirmed by high‐performance liquid chromatography (HPLC) using chiral stationary phase and reverse‐phase HPLC analyses, respectively. The ee values and the chemical purities both exceeded 99%. Animal experiments showed that (R)‐bambuterol was a potent inhibitor for histamine‐induced asthma reactions. (S)‐bambuterol was ineffective in relaxing the airways. Both enantiomers increased heart rates in beagles. Therefore, replacing rac‐bambuterol with (R)‐bambuterol could be beneficial for asthma patients. Chirality 28:306–312, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

    6.
    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   

    7.
    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   

    8.
    A modified macrocyclic glycopeptide‐based chiral stationary phase (CSP), prepared via Edman degradation of vancomycin, was evaluated as a chiral selector for the first time. Its applicability was compared with other macrocyclic glycopeptide‐based CSPs: TeicoShell and VancoShell. In addition, another modified macrocyclic glycopeptide‐based CSP, NicoShell, was further examined. Initial evaluation was focused on the complementary behavior with these glycopeptides. A screening procedure was used based on previous work for the enantiomeric separation of 50 chiral compounds including amino acids, pesticides, stimulants, and a variety of pharmaceuticals. Fast and efficient chiral separations resulted by using superficially porous (core‐shell) particle supports. Overall, the vancomycin Edman degradation product (EDP) resembled TeicoShell with high enantioselectivity for acidic compounds in the polar ionic mode. The simultaneous enantiomeric separation of 5 racemic profens using liquid chromatography‐mass spectrometry with EDP was performed in approximately 3 minutes. Other highlights include simultaneous liquid chromatography separations of rac‐amphetamine and rac‐methamphetamine with VancoShell, rac‐pseudoephedrine and rac‐ephedrine with NicoShell, and rac‐dichlorprop and rac‐haloxyfop with TeicoShell.  相似文献   

    9.
    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   

    10.
    Bonded polysaccharide‐derived chiral stationary phases were found to be useful for the preparation of the four stereoisomers of the cyclopropane analogue of phenylalanine (c3Phe) as well as for the direct determination of the enantiomeric purity of c3Phe derivatives by HPLC. Three chiral stationary phases, consisting of cellulose and amylose derivatives chemically bonded on allylsilica gel, were tested. The mixed 10‐undecenoate/3,5‐dimethylphenylcarbamate of cellulose, 10‐undecenoate/3,5‐dimethylphenylcarbamate of amylose and 10‐undecenoate/p‐methylbenzoate of cellulose were the starting polysaccharide derivatives for CSP‐1, CSP‐2, and CSP‐3, respectively. Using mixtures of n‐hexane/chloroform/2‐propanol as mobile phase on a semi‐preparative column (150 mm × 20 mm ID) containing CSP‐2, we separated about 1.7 g of racemic cis‐methyl 1‐tert‐butoxycarbonylamino‐2‐phenylcyclopropanecarboxylate (cis‐ 6 ) and 1.2 g of racemic trans‐methyl‐1‐tert‐butoxycarbonylamino‐2‐phenylcycloprop‐anecarboxylate (trans‐ 6 ) by successive injections. Chirality 11:583–590, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

    11.
    An on-column HPLC procedure using a chiral stationary phase (CSP) was developed for the determination of rate constants and free energy barriers of enantiomerization of (+/-)IDRA21. Subsequently, the HPLC method was applied for investigation of two structurally related chiral compounds. The individual enantiomers of the studied compounds were isolated in parallel by preparative HPLC and rate constants and free energy barriers of enantiomerization were determined in different solvents. The on-column enantiomerization data revealed that CSP induces different rate constants for the two enantiomers. The results generated off-line were used to determine the influence of solvents on the racemization of (+) and (-) IDRA21 and to gain further insight into the enantiomerization mechanism of chiral 3,4-dihydro-1,2,4-benzothiadiazine 1,1-dioxide type compounds.  相似文献   

    12.
    Cycloaminoalkylnaphthalene 3 shows interesting opioid‐like analgesic properties. It possesses two chiral centers and can exist as two racemic pairs and four diastereomers. Since the binding of opioids with receptors is stereoselective, it was important to have the two racemic pairs as well as the four diastereomers. In this paper the synthesis of the (2R,3S/2S,3R) racemate and the (2R,3S) and (2S,3R) enantiomers of the 1,2‐dimethyl‐3‐[2‐(6‐hydroxynaphthyl)]‐3‐hydroxypyrrolidine 3 is considered and the determination of absolute configuration is described. The (2R,3S/2S,3R)‐ 3 racemate and the (2R,3S)‐ 3 and (2S,3R)‐ 3 enantiomers were prepared by reaction of the racemic and optically active 1,2‐dimethyl‐3‐pyrrolidone 2, respectively, with the lithiation product obtained from 2‐bromo‐6‐tetrahydropyranyloxy‐naphthalene 1 and acidic hydrolysis. The above‐mentioned enantiomers of 3 were also obtained by optical resolution via fractional crystallization of the salts with d ‐ and l ‐tartaric acids. The configuration of the optically active compounds was determined by X‐ray analysis of a crystal of (−)‐(2S,3R)‐ 3 · HCl · H2O. The pharmacological test HPT showed that (−)‐(2S,3R)‐ 3 · HCl · H2O enantiomer is able to induce opioid‐like analgesia with a relative potency 1.5 times that of (2R,3S/2S,3R)‐ 3 and ∼1.5 times that of morphine. Chirality 11:21–28, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

    13.
    Chromatographic separation of the enantiomers of parent compounds dimethyl α‐hydroxyallyl phosphonate 1a and 1‐(dimethoxyphosphoryl) allyl methyl carbonate 1b was demonstrated by high‐performance liquid chromatography (HPLC) using Chiralpak AS‐H and ad ‐H chiral stationary phases (CSP), respectively, using a combination of UV, polarimetric, and refractive index detectors. A comparison was made of the separation efficiency and elution order of enantiomeric α‐hydroxyallyl phosphonates and their carbonate derivatives on commercially available polysaccharide AS, ad , OD, IC‐3, and Whelk‐O 1 CSPs. In general, the α‐hydroxyallyl phosphonates were resolved on the AS‐H CSP, whereas the carbonate derivatives 1b and 2b were preferentially resolved on the ad ‐H CSP. The impact of aryl substitution on the resolution of analytes 1d , 1e , 1f and 2 , 3 , 4 , 5 , 6 , 7 , 8 was evaluated. Thermodynamic parameters determined for enantioselective adsorption hydroxyphosphonates 1a and 4 on the AS‐H CSP and carbonate 1b on the ad ‐H CSP demonstrated enthalpic control for separation of the enantiomers. Chirality 28:656–662, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

    14.
    A stereoselective high‐performance liquid chromatographic (HPLC) method was developed and validated to determine S‐(?)‐ and R‐(+)‐propranolol in rat serum. Enantiomeric resolution was achieved on cellulose tris(3,5‐dimethylphenylcarbamate) immobilized onto spherical porous silica chiral stationary phase (CSP) known as Chiralpak IB. A simple analytical method was validated using a mobile phase consisted of n‐hexane‐ethanol‐triethylamine (95:5:0.4%, v/v/v) at a flow rate of 0.6 mL min‐1 and fluorescence detection set at excitation/emission wavelengths 290/375 nm. The calibration curves were linear over the range of 10–400 ng mL‐1 (R = 0.999) for each enantiomer with a detection limit of 3 ng mL‐1. The proposed method was validated in compliance with ICH guidelines in terms of linearity, accuracy, precision, limits of detection and quantitation, and other aspects of analytical validation. Actual quantification could be made for propranolol isomers in serum obtained from rats that had been intraperitoneally (i.p.) administered a single dose of the drug. The proposed method established in this study is simple and sensitive enough to be adopted in the fields of clinical and forensic toxicology. Molecular modeling studies including energy minimization and docking studies were first performed to illustrate the mechanism by which the active enantiomer binds to the β‐adrenergic receptor and second to find a suitable interpretation of how both enantiomers are interacting with cellulose tris(3,5‐dimethylphenylcarbamate) CSP during the process of resolution. The latter interaction was demonstrated by calculating the binding affinities and interaction distances between propranolol enantiomers and chiral selector. Chirality 26:194–199, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

    15.
    Chiral natural flavor compounds exhibit characteristic enantiomeric excesses due to stereoselective, enzymatically catalyzed reactions during biogenesis. Although the enzymatic formation of the strawberry key flavor compound 4-hydroxy-2,5-dimethyl-3(2H)-furanone (HDMF; Furaneol(R)) is anticipated, the naturally occurring compound is racemic. As racemization due to keto-enol-tautomerism of HDMF could account for this observation, HDMF was investigated by (1)H-NMR spectroscopy tracing the exchange of the proton bound to the furanone-ring at C2 with deuteron from the medium (D(2)O). In addition, the racemization rate of HDMF was directly determined by cyclodextrin-modified capillary electrophoresis of enantiomerically enriched HDMF stored at different pH values. Tautomerism and the racemization rate of HDMF was lowest at pH values between 4 and 5. However, tautomerism and thus racemization was catalyzed under stronger acidic conditions (pH 2) and especially at pH values greater than 7, the value published for plant cell cytosol. Approximately 50% of the protons at C2 were exchanged with deuteron within 1 h at pH 7.2. Therefore, in order to demonstrate the enzymatic formation of HDMF, incubation experiments with Zygosaccharomyces rouxii as well as strawberry protein extract were carried out under slightly acidic conditions (pH 5), the most suitable pH value for studies on the enantiomeric ratio of HDMF. In both experiments the formation of enantiomerically enriched HDMF could be demonstrated for the first time, whereas incubation experiments under neutral conditions resulted in the detection of racemic HDMF.  相似文献   

    16.
    S K Yang  M Mushtaq  P P Fu 《Chirality》1990,2(1):58-64
    1,12-Dimethylbenz[a]anthracene (1,12-DMBA) cis-5,6-dihydrodiol was synthesized by oxidation of 1,12-DMBA with osmium tetroxide in pyridine in low yield (less than or equal to 3%) and was purified by sequential use of reversed-phase and normal-phase HPLC. Two pairs of 1,12-DMBA cis-5,6-dihydrodiol enantiomers, derived from P (right-handed helix) and M (left-handed helix) conformers, were eluted as a single chromatographic peak on both reversed-phase and normal-phase HPLC. However, these four enantiomers were resolved by sequential use of two chiral stationary phase (CSP) HPLC columns. CSP (Pirkle type I) columns were packed with either (R)-N-(3,5-dinitrobenzoyl)phenylglycine or (S)-N-(3,5-dinitrobenzoyl)leucine, which is ionically bonded to gamma-aminopropylsilanized silica. Absolute configurations of enantiomers were determined by comparing their circular dichroism spectra with those of conformationally similar cis-5,6-dihydrodiol enantiomers of 4-methylbenz[a]anthracene and 7,12-dimethylbenz[a]anthracene with known absolute stereochemistry.  相似文献   

    17.
    pKa1 values of 3-methoxy-N-desmethyldiazepam in acetonitrile and methanol containing various acid concentrations were determined by spectrophotometry to be 3.5 and 1.3, respectively. Temperature-dependent racemization of enantiomeric 3-methoxy-N-desmethyldiazepam in methanol containing 0.5 M H2SO4 was studied by circular dichroism spectropolorimetry and the racemization reactions were found to follow apparent first-order kinetics. Thermodynamic parameters of the racemization reaction were found to be: Eact = 18.8 kcal/mol, and at 25°C: ΔH? = 18.3 kcal/mol, ΔS? = ?14.8 entropy unit, and ΔG? = 22.7 kcal/mol, respectively. The racemization had an isotope effect (kH/kD) of 1.6 at 42°C. Based on the results of this report and those of earlier reports by other investigators, a nucleophilically solvated C3 carbocation intermediate resulting from either a P (plus) or an M (minus) conformation is proposed to be an intermediate and responsible for the stereoselective nucleophilic substitution and the subsequent racemization of 3-methoxy-N-desmethyldiazepam enantiomers. © 1993 Wiley-Liss, Inc.  相似文献   

    18.
    Applications of the on‐line HPLC‐exciton CD analysis using (S)‐2‐tert‐butyl‐2‐methyl‐1,3‐benzodioxole‐4‐carboxylic acid [(S)‐TBMBC‐OH] that can simultaneously determine the enantiomeric compositions and the absolute configuration of cyclohexane‐1,2‐diols and diamines as well as acyclic vicinal diols and amino alcohols were studied. Di‐O‐ or di‐N,O‐(S)‐TBMBC derivatives of acyclic terminal vicinal diols, 2‐hydroxy‐1‐amines, and nonterminal vicinal diols gave symmetrical exciton CD spectra between enantiomers, indicating their absolute configurations. However, Di‐N,O‐(S)‐TBMBC derivatives of 2‐amino‐1‐ols did not always give symmetrical exciton CD spectra between enantiomers, but their 2‐phthalimido‐1‐O‐(S)‐TBMBC derivatives gave symmetrical exciton CD spectra, indicating their absolute configurations. All these (S)‐TBMBC derivatives were separated by normal‐phase HPLC and unequivocally determined by the on‐line HPLC‐exciton CD analysis without recourse to reference samples. Chirality 11:149–159, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

    19.
    Chiral high‐performance liquid chromatography (HPLC) separation of trans‐bis[2‐(2‐pyridyl)aminophenolato] dichlorocyclotriphosphazene 1 was achieved and the absolute configuration of (+)-1 was assigned to be S,S by single‐crystal X‐ray structural analysis. The optically pure 1,2‐diphenyl‐1,2‐ethanediolate derivatives (+)‐ 2a and (?)‐ 2b were synthesized by the reactions of (+)-1 and (-)-1 with (R,R)‐hydrobenzoin, respectively, in refluxing toluene in the presence of an excess amount of triethylamine and a catalytic amount of 4‐(dimethylamino)pyridine. The racemization of the enantiomers of 1 and the epimerization of diastereomers of 2 were not observed in refluxing toluene neither under acidic nor basic conditions. The stereochemistry of (+)-1 was confirmed by the crystal structure of (+)‐ 2a and bis[(4‐methyl‐2‐pyridyl)oxy]cyclotriphosphazene (+)-3 derived from (+)-1 . Chirality 28:556–561, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

    20.
    A simple and rapid chiral high‐performance liquid chromatography (HPLC) method was developed and validated for bioanalysis of clopidogrel enantiomers on rat dried blood spots (DBS). Clopidogrel enantiomers were extracted from DBS using ethanol: methanol (80:20, v/v) and separated on a Chiralcel OJ‐H column containing cellulose tris (4‐methly benzoate) as a polysaccharide stationary phase using n‐hexane–ethanol‐diethylamine (70:30, 0.1 v/v) as a mobile phase at a flow rate of 1.0 mL/min. The detection was carried out at 220 nm using a photodiode array (PDA) detector while the elution order of the enantiomers was determined by a polarimeter connected to PDA in series. The effect of hematocrit on extraction of clopidogrel enantiomers from DBS was evaluated and no interference from endogenous substances was noticed. The overall accuracy of (R) and (S) enantiomers of clopidogrel from DBS were 91.6 and 89.2%, respectively. The calibration curves were linear over the concentration range of 1–500 µg/mL for both enantiomers. The results show that the method is specific, precise, and reproducible (intra‐ and interday precision relative standard deviations (RSDs) <10.0%). The stability of racemic clopidogrel was performed under all storage conditions and the results were found to be well within the acceptance limits. Chirality 26:102–107, 2014.© 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