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.
Molecularly imprinted polymeric membranes were prepared from various oligopeptide tweezers by the adoption of N‐α‐tert‐butoxycarbonyl‐D ‐tryptophan (Boc‐D ‐Trp) or N‐α‐tert‐butoxycarbonyl‐L ‐tryptophan (Boc‐L ‐Trp) as a print molecule. The chiral recognition ability of the formed molecular recognition sites was dependent on the absolute configuration of the print molecule adopted in the membrane preparation (molecular imprinting) process, whereas the candidate oligopeptide tweezers consisted of the L ‐amino acid residues. In other words, the membranes imprinted by the D ‐isomer recognized the D ‐isomer in preference to the corresponding L ‐isomer, and vice versa. The affinity constant between the target molecule and the chiral recognition site from the oligopeptide tweezers was higher than that from a single‐strand oligopeptide derivative. Those membranes selectively transported the enantiomer, which was preferentially incorporated into the membrane by dialysis. The permselectivities for these membranes exceeded their adsorption selectivities. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 385–396, 2005  相似文献   

2.
《Electroanalysis》2004,16(11):955-960
Enantioselective resolution is realized by combining potentiometry with ligand exchange (CE) in a new method called chiral ligand exchange potentiometry (CLEP). A chiral selector, N‐carbobenzoxy‐L ‐aspartic acid (N‐CBZ‐L‐Asp), preferentially recognizes D ‐aspartic acid (D‐Asp) and undergoes ligand exchange with the enantiomeric labile coordination complexes of [Cu(II)(D‐Asp)2] or [Cu(II)(L‐Asp)2] to form a diastereoisomeric complex [(D‐Asp)Cu(II)(N‐CBZ‐L‐Asp)] (a) or [(L‐Asp)Cu(II)(N‐CBZ‐L‐Asp)] (b). Considerable stereoselectivity occurs in the formation of these diastereoisomeric complexes, and their net charges were ?2 (a) and 0 (b), respectively, resulting in different Nernst factor (electrode slope), thus enabling chiral D‐Asp to be distinguished by potentiometry without any pre‐ or postseparation processes.  相似文献   

3.
A new styrene derivative having D ‐mannaric moiety, Np‐vinylbenzyl‐D ‐mannaramic acid (VB‐D ‐ManaH, 8 ) was synthesized though the ring‐opening reaction of D ‐mannaro‐1,4:6,3‐dilactone (D ‐MDL) with p‐vinylbenzylamine. VB‐D ‐ManaH was copolymerized with acrylamide (AAm) to give novel polymers having D ‐mannaric moiety in the pendants, P(VB‐D ‐ManaH‐co‐AAm), 10 . The resulting glycomonomer and polymer ( 8 and 10 ) bearing D ‐mannaric pendants were found to inhibit the β‐glucuronidase activity, although the inhibition ability of the corresponding saccharodilactone (D ‐MDL) was known to be low. Additionally, the inhibition ability of P(VB‐D ‐ManaH‐co‐AAm), 10 , was almost the same as that of the glycopolymer having D ‐glucaric pendants, P(VB‐6‐D ‐GlcaH‐co‐AAm), 1 , which was one of the most effective inhibitors for β‐glucuronidase, reported in our previous work. Thus, 10 and 8 may be the first D ‐mannaric strong inhibitors to the β‐glucuronidase activity. The Lineweaver–Burk plot suggested that the inhibition mechanisms of 10 and 8 were more complicated than in the case of the competitive and uncompetitive inhibition of Np‐(vinylbenzyl)‐6‐D ‐glucaramic ( 11 ) and Np‐(vinylbenzyl)‐1‐D ‐glucaramic acids ( 12 ), respectively. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 2032–2042, 2009  相似文献   

4.
Two kinds of novel vinyl monomers having D ‐glucaric moieties leading to a new type of glycopolymeric inhibitors of β‐glucuronidase, Np‐vinylbenzyl‐6‐D ‐glucaramide (6 ) and potassium Np‐vinylbenzyl‐6‐D ‐glucaramid‐1‐ate (8 ), were synthesized by the reaction of D ‐glucaro‐6,3‐lactone (3 ) with p‐vinylbenzylamine (5 ) with no catalyst, and the subsequent treatment of the reaction mixture with acetic anhydride and potassium hydroxide aqueous solution, respectively. The radical copolymerization of 8 with acrylamide in various feed ratios at 60°C in 0.1 N potassium chloride aqueous solution gave water‐soluble copolymers (9 ) composed of a synthetic polymeric main chain and many pendant D ‐glucaric chains. The resulting glycopolymers (9 ) were found to inhibit the activity of β‐glucuronidase strongly through a model reaction with p‐nitrophenyl β‐D ‐glucuronide (10 ) in acetic buffer solution (pH 4.7). © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 303–312, 1999  相似文献   

5.
A microchip electrophoresis (MCE) method with LIF detection was presented for quantification of D ‐aspartic acid (D ‐Asp) and D ‐glutamate (D ‐Glu) in biological samples. D ‐Asp and D ‐Glu were determined after precolumn derivatization with FITC. The chiral separation was performed on a glass/PDMS hybrid microfluidic chip using γ‐CD as chiral selector in the running buffer. High sensitive detection was obtained by the LIF detection. The LODs (S/N = 3) for D ‐Asp and D ‐Glu were 6.0×10–8 and 4.0×10–8 M, respectively. Using this method, the levels of D ‐Asp and D ‐Glu in rat brain and human cerebrospinal fluid (CSF) were determined.  相似文献   

6.
A novel, simple and sensitive method for the enantioseparation and determination of DL ‐tetrahydropalmatine (DL ‐THP) was developed using ACE in combination with partial filling technique and field‐amplified sample injection. A chiral selector, i.e. BSA, was used for the enantioseparation of DL ‐THP in ACE. Effects of BSA concentration, pH and separation voltage on the effectiveness of the enantiomer separation were evaluated. In an optimal condition, D ‐ and L ‐THP were completely enantio‐separated in less than 9 min by partially filling an electrophoretic capillary with 50 μmol/L BSA (50 mbar, 100 s) and carrying out an electrophoresis with 20 mmol/L phosphate buffer (pH 7.4) at 15 kV. The sensitivity was further improved by making use of field‐amplified sample injection to lower the LOD (defined as S/N=3) down to 6 ng/mL. Real samples were also tested and promising results for the determination of DL ‐THP enantiomers were obtained.  相似文献   

7.
d ‐Aspartate (d ‐Asp) and N‐methyl‐d ‐aspartate (NMDA) occur in the neuroendocrine systems of vertebrates and invertebrates, where they play a role in hormone release and synthesis, neurotransmission, and memory and learning. N‐methyl‐d ‐glutamate (NMDG) has also been detected in marine bivalves. Several methods have been used to detect these amino acids, but they require pretreatment of tissue samples with o‐phthaldialdehyde (OPA) to remove primary amino acids that interfere with the detection of NMDA and NMDG. We report here a one‐step derivatization procedure with the chiral reagent N‐α‐(5‐fluoro‐2,4‐dinitrophenyl)‐(d or l )‐valine amide, FDNP‐Val‐NH2, a close analog of Marfey's reagent but with better resolution and higher molar absorptivity. The diastereomers formed were separated by HPLC on an ODS‐Hypersil column eluted with TFA/water–TFA/MeCN. UV absorption at 340 nm permitted detection levels as low as 5–10 pmol. d ‐Asp, NMDA and NMDG peaks were not obscured by other primary or secondary amino acids; hence pretreatment of tissues with OPA was not required. This method is highly reliable and fast (less than 40 min HPLC run). Using this method, we detected d ‐Asp, NMDA and NMDG in several biological tissues (octopus brain, optical lobe and bucchal mass; foot and mantle of the mollusk Scapharca broughtonii), confirming the results of other researchers. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Metal–organic frameworks are promising porous materials. Chiral metal–organic frameworks have attracted considerable attention in controlling enantioselectivity. In this study, a homochiral metal–organic framework [Co2(D‐cam)2(TMDPy)] (D‐cam = d ‐camphorates, TMDPy = 4,4′‐trimethylenedipyridine) with a non‐interpenetrating primitive cubic net has been used as a chiral stationary phase in high‐performance liquid chromatography. It has allowed the successful separation of six positional isomers and six chiral compounds. The good selectivity and baseline separation, or at least 60% valley separation, confirmed its excellent molecular recognition characteristics. The relative standard deviations for the retention time of run‐to‐run and column‐to‐column were less than 1.8 and 3.1%, respectively. These results demonstrate that [Co2(D‐cam)2(TMDPy)] may represent a promising chiral stationary phase for use in high‐performance liquid chromatography.  相似文献   

9.
Recently, amino acid ionic liquids (AAILs) have attracted much research interest. In this paper, we present the first application of AAILs in chiral separation based on the chiral ligand exchange principle. By using 1‐alkyl‐3‐methylimidazolium L ‐proline (L ‐Pro) as a chiral ligand coordinated with copper(II), four pairs of underivatized amino acid enantiomers—dl ‐phenylalanine (dl ‐Phe), dl ‐histidine (dl ‐His), dl ‐tryptophane (dl ‐Trp), and dl ‐tyrosine (dl ‐Tyr)—were successfully separated in two major chiral separation techniques, HPLC and capillary electrophoresis (CE), with higher enantioselectivity than conventionally used amino acid ligands (resolution (Rs)=3.26–10.81 for HPLC; Rs=1.34–4.27 for CE). Interestingly, increasing the alkyl chain length of the AAIL cation remarkably enhanced the enantioselectivity. It was inferred that the alkylmethylimidazolium cations and L ‐Pro form ion pairs on the surface of the stationary phase or on the inner surface of the capillary. The ternary copper complexes with L ‐Pro are consequently attached to the support surface, thus inducing an ion‐exchange type of retention for the dl ‐enantiomers. Therefore, the AAIL cation plays an essential role in the separation. This work demonstrates that AAILs are good alternatives to conventional amino acid ligands for ligand‐exchange‐based chiral separation. It also reveals the tremendous application potential of this new type of task‐specific ILs.  相似文献   

10.
Novel polyamide with chiral environment was obtained from aromatic diamine, 4,4′‐diaminodiphenylmethane (DADPM), and N‐α‐protected L ‐glutamic acid, N‐α‐benzoyl‐L ‐glutamic acid (Benzoyl‐L ‐Glu‐OH). The optical rotation ([α]D ) of the polyamide was determined to be 3.6° (c = 1.00 g/dL in DMF), implying that the optically active polyamide was obtained. The present polyamide gave a durable self‐standing membrane. The membrane selectively incorporated the D ‐isomer of Ac‐Trp from racemic mixture of Ac‐Trp. The adsorption selectivity toward Ac‐D ‐Trp was determined to be 1.95. It showed chiral separation ability by adopting potential difference as a driving force for membrane transport. The permselectivity was dependent on the potential difference, and at the applied potential difference of 3.0 V, the membrane selectively transported Ac‐D ‐Trp and the permselectivity toward Ac‐D ‐Trp was determined to be 1.84, which was close to the adsorption selectivity of 1.95. Contrary to this, the membrane showed opposite permselectivity at the applied potential difference of 2.0 V and the permselectivity toward the L ‐isomer reached 2.48. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 2530–2538, 2009  相似文献   

11.
The melt ring‐opening/condensation reaction of trans‐4‐hydroxy‐N‐benzyloxycarbonyl‐L‐proline (N‐CBz‐Hpr) with cyclic carbonate [trimethylene carbonate (tri‐MC) or tetramethylene carbonate (tetra‐MC)] at a wide range of molar fractions in the feed produced new degradable poly(ester‐carbonate)s. The influence of reaction conditions such as polymerization time and temperature on the yield and inherent viscosity of the copolymers was investigated. The polymerizations were carried out in bulk at 140 °C with 1.5 wt % stannous octoate as a catalyst for 30 h. The poly(ester‐carbonate)s obtained were characterized by Fourier transform infrared spectroscopy, 1H NMR, differential scanning calorimetry, gel permeation chromatography, and Ubbelohde viscometry. The copolymers synthesized exhibited moderate molecular weights with rather narrow molecular weight distributions. The values of the glass‐transition temperature (Tg) of the copolymers depend on the molar fractions of cyclic carbonate. For the poly(N‐CBz‐Hpr‐co‐tri‐MC) system, with a decreased tri‐MC content from 93 to 16 mol %, the Tg increased from ?10 to 60 °C. Similarly, for the poly(N‐CBz‐Hpr‐co‐tetra‐MC) system, when the tetra‐MC content decreased from 80 to 8 mol %, the Tg increased from ?18 to 52 °C. The relationship between the poly(N‐CBz‐Hpr‐co‐tri‐MC) Tg and the compositions was in approximation with the Fox equation. In vitro degradation of these poly(N‐CBz‐Hpr‐co‐tri‐MC)s was evaluated from weight‐loss measurements. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 1435–1443, 2003  相似文献   

12.
Crosslinked chiral nanoparticles were successfully synthesized via reversible addition‐fragmentation chain transfer (RAFT) miniemulsion polymerization of 6‐Op‐vinylbenzyl‐1,2:3,4‐di‐O‐isopropylidene‐D ‐galactopyranose (VBPG) using linear poly(VBPG) as the macro‐RAFT agent. The polymerization of VBPG in the absence of crosslinker was first studied and the kinetic results showed that the molecular weights of the obtained poly(VBPG) increased linearly with the monomer conversion and was in good consistency with the corresponding theoretical ones while there remained a relative narrow polydispersity. The effect of the amount of crosslinker, divinylbenzene, on the nanoparticle size and chiral separation properties of the obtained nanoparticles were investigated in detail using four racemates ±‐3‐Amino‐1,2‐propanediol, D ,L ‐arabinose, D ,L ‐tartaric acid, and D ,L ‐mandelic acid. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 1324–1331, 2010  相似文献   

13.
This paper extends the research of the utilization of borate coordination complexes in chiral separation by counter‐current chromatography (CCC). Racemic propafenone was successfully enantioseparated by CCC with di‐n‐butyl l ‐tartrate combined with boric acid as the chiral selector. The two‐phase solvent system was composed of chloroform/ 0.05 mol/L acetate buffer pH 3.4 containing 0.10 mol/L boric acid (1:1, v/v), in which 0.10 mol/L di‐n‐butyl l ‐tartrate was added in the organic phase. The influence of factors in the enantioseparation of propafenone were investigated and optimized. A total of 92 mg of racemic propafenone was completely enantioseparated using high‐speed CCC in a single run, yielding 40–42 mg of (R)‐ and (S)‐propafenone enantiomers with an HPLC purity over 90–95%. The recovery for propafenone enantiomers from fractions of CCC was in the range of 85–90%.  相似文献   

14.
Mixed mode (MM) separation using a combination of MEKC and polyelectrolyte multilayer (PEM) coatings is herein reported for the separation of achiral and chiral analytes. Many analytes are difficult to separate by MEKC and PEM coatings alone. Therefore, the implementation of a MM separation provides several advantages for overcoming the limitations of these well‐established methods. In this study, it was observed that achiral separations using MEKC and PEM coatings individually resulted in partial resolution of eight very similar aryl ketones when the molecular micelle (sodium poly(N‐undecanoyl‐L ‐glycinate)) concentration was varied from 0.25 to 1.00% w/v and the bilayer number varied from 2 to 4. However, when MM separation was introduced, baseline resolution was achieved for all eight analytes. In the case of chiral separations, temazepam, aminoglutethimide, benzoin, benzoin methyl ether, and coumachlor were separated using the three separation techniques. For chiral separations, the chiral molecular micelle, sodium poly(N‐undecanoyl‐L ‐leucylvalinate), was employed at concentrations of 0.25–1.50% w/v for both MEKC and PEM coatings. Overall, the results revealed partial separation with MEKC and PEM coatings individually. However, MM separation enabled baseline separation of each chiral mixture. The separation of achiral and chiral compounds from different compound classes demonstrates the versatility of this MM approach.  相似文献   

15.
Copolymerization of ornithine‐ and lysine‐derived N‐propargylamides, N‐α‐tert‐butoxycarbonyl‐N‐δ‐fluorenylmethoxycarbonyl‐L ‐ornithine N′‐propargylamide ( 1 ), N‐α‐tert‐butoxycarbonyl‐N‐ε‐fluorenylmethoxycarbonyl‐L ‐lysine N′‐propargylamide ( 2 ), N‐α‐fluorenylmethoxycarbonyl‐N‐δ‐tert‐butoxycarbonyl‐L ‐ornithine N′‐propargylamide ( 3 ), and N‐α‐fluorenylmethoxycarbonyl‐N‐ε‐tert‐butoxycarbonyl‐L ‐lysine N′‐propargylamide (4) with dipropargyl adipate was carried out using (nbd)Rh+6‐C6H5B?(C6H5)3] as a catalyst in THF to obtain polymer gels in 80–93% yields. The gels adsorbed N‐benzyloxycarbonyl L ‐alanine, N‐benzyloxycarbonyl L ‐alanine methyl ester, and (S)‐(+)‐1‐phenyl‐1,2‐ethanediol preferably than the corresponding optical isomers. The order of chiral discrimination was poly( 1 ) > poly( 4 ) > poly( 2 ), poly( 3 ) gels. The fluorenylmethoxycarbonyl groups of the gels could be partly removed by piperidine treatment, leading to increase of adsorptivity but decrease of chiral recognition ability. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4175–4182, 2008  相似文献   

16.
The enantiomeric separation of d ,l ‐tryptophan (Trp) and d ,l ‐kynurenine (KYN) was investigated by high‐performance liquid chromatography using pre‐column fluorescence derivatization with a chiral fluorescent labeling reagent, R(−)‐4‐(3‐isothiocyanatopyrrolidin‐1‐yl)‐7‐ (N,N‐dimethylaminosulfonyl)‐2,1,3‐benzoxadiazole [R(−)‐DBD‐PyNCS]. Using an octadecylsilica column, namely, an Inertsil ODS‐3 column (250 × 2.0 mm; i.d., 3 µm), four fluorescence peaks of D‐ and l ‐Trp as well as d ‐ and l ‐KYN derivatized with R(−)‐DBD‐PyNCS were clearly observed, and their chemical structures were confirmed by HPLC–time‐of‐flight–mass spectrometry. Simultaneous separation was achieved under the mobile phase condition of 1.5% acetic acid in H2O–CH3CN (60:40), and the separation factors of d ,l ‐Trp and d ,l ‐KYN derivatized with R(−)‐DBD‐PyNCS were 1.22 and 1.19, respectively. Fluorescence detection was carried out by setting the emission wavelength at 565 nm, and the excitation wavelength at 440 nm, and the detection limits were approximately 0.3–0.5 pmol (signal‐to‐noise ratio of 3). Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

17.
Sarcosine N‐carboxyanhydride, D,L ‐alanine N‐carboxyanhydride, D,L ‐phenylalanine N‐carboxyanhydride, and D,L ‐leucine N‐carboxyanhydride were polymerized with pyridine or N‐ethyldiisopropylamine as the catalyst. With pyridine, cyclic oligo‐ and polypeptides were obtained in addition to water‐initiated or water‐terminated chains. The cyclopeptides were the main products in the case of sarcosine N‐carboxyanhydride and D,L ‐phenylalanine N‐carboxyanhydride. The fraction of cycles was particularly high when N‐methylpyrrolidone was used as the reaction medium. These results suggested the existence of a pyridine‐catalyzed zwitterionic mechanism. However, cyclopeptides were also obtained with N‐ethyldiisopropylamine as the catalyst. In this case, N‐deprotonation of N‐carboxyanhydrides, followed by the formation of N‐acyl N‐carboxyanhydride chain ends, was the most likely initiation mechanism. Various chain‐growth mechanisms were examined. In the case of γ‐benzyl ester‐L ‐glutamate N‐carboxyanhydride, side reactions such as the formation of pyroglutamoyl end groups were detected. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 4680–4695, 2006  相似文献   

18.
A stereoregular poly(phenylacetylene) bearing the bulky aza‐18‐crown‐6 ether as the pendant (poly‐ 1 ) formed a predominantly one‐handed helical conformation upon complexation with various chiral compounds, such as amino acids, peptides, aminosugars, amines, and amino alcohols in water. The complexes exhibited an induced circular dichroism (ICD) in the UV–visible region of the polymer main chain. Therefore, poly‐ 1 can be used as a novel probe for determining the chirality of chiral compounds in water. The assay of 19 common free L ‐amino acids gave the same ICD sign at 0 °C except for L ‐phenylalanine. The effects of pH, temperature, guest concentration, and organic solvent content on the ICD during the complexation of poly‐ 1 with chiral compounds were also investigated. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 1004–1013, 2003  相似文献   

19.
A series of chiral polymers based on poly(N‐acryl) amino acids was synthesized using a convergent synthetic approach. These chiral polymers have been used as chiral additives to induce enantioselective crystallization of racemic or conglomerate amino acids in solutions. These polymeric additives showed strong capabilities to enhance highly enantioselective resolution during the crystallization of amino acids. In addition, these polymers caused unusual modifications of amino acid crystal morphologies. Furthermore, spherical microparticles of those same chiral polymers were also shown active in similar chiral discriminations during amino acid crystallizations occurring on microparticle surfaces. Our study demonstrates the high potential of chiral polymers and microparticles to resolve amino acids throughout crystallization processes. High enantiomeric excesses in one targeted enantiomer of amino acids can also be maximized via time‐dependent kinetic control of crystallizations. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 3009–3017, 2006  相似文献   

20.
New styryl‐type water‐insoluble and methacryloyl‐type water‐soluble monomers, N‐(p‐vinylbenzyl)‐1,2‐O‐isopropylidene‐6‐D ‐glucofuranuronamide and N‐(2‐methacryloylamino)ethyl‐1,2‐O‐isopropylidene‐6‐D ‐glucofuranuronamide, were synthesized from the most common acidic saccharide, D ‐glucuronic acid. Their radical homopolymerizations and copolymerizations with styrene and acrylamide were tried under various conditions. The isopropylidene groups in the resulting polymers were removed in a mixture of trifluoroacetic acid and water (2/1 v/v) to give the corresponding polymers with many pendant D ‐glucopyranuronyl groups with reactive reducing groups. The pendant reducing ends may be useful as potential binding sites under a hydrophilic atmosphere. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 39: 3893–3901, 2001  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