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 375 毫秒
1.
Lipases immobilized on polypropylene powders have been used as the biocatalyst in the enantioselective hydrolysis of (S)-naproxen from racemic naproxen thioesters in isooctane, in which trioctylamine was added to perform in situ racemization of the remaining (R)-thioester substrate. A detailed study of the kinetics for hydrolysis and racemization indicates that increasing the trioctylamine concentration can activate and stabilize the lipase as well as enhance the racemization and non-stereoselective hydrolysis of the thioester. Effects of the aqueous pH value and trioctylamine concentration on (S)-naproxen dissociation and partitioning in the aqueous phase as well as the transportation in a hollow fiber membrane were further investigated. Good agreements between the experimental data and theoretical results were obtained when the dynamic kinetic resolution process was integrated with a hollow fiber membrane to reactively extract the desired (S)-naproxen out of the reaction medium.  相似文献   

2.
A lipase-catalyzed enantioselective hydrolysis process under conditions of continuous in situ racemization of substrate with trioctylamine as the catalyst was developed for the production of (S)-suprofen from (R,S)-suprofen 2,2,2-trifluoroethyl thioester in isooctane. A detailed investigation of trioctylamine concentration on the enzyme activation and stability as well as the kinetic behaviors of the thioester in racemization and enzymatic reaction was conducted, in which good agreement between the experimental data and theoretical results was observed. A complete conversion of the racemate for the desired (S)-suprofen in 95% ee(P) was obtained. Moreover, the recovery of the acid product by extraction and reuse of the organic solution were reported.  相似文献   

3.
An efficient method of peptide thioester synthesis is described. The reaction is based on an N‐4,5‐dimethoxy‐2‐mercaptobenzyl (Dmmb) auxiliary‐assisted NS acyl shift reaction after assembling a peptide chain by Fmoc‐solid phase peptide synthesis. The Dmmb‐assisted NS acyl shift reaction proceeded efficiently under mildly acidic conditions, and the peptide thioester was obtained by treating the resulting S‐peptide with sodium 2‐mercaptoethanesulfonate. No detectable epimerization of the amino acid residue adjacent to the thioester moiety in the case of Leu was found. The reactions were also amenable to the on‐resin preparation of peptide thioesters. The utility was demonstrated by the synthesis of a 41‐mer peptide thioester, a phosphorylated peptide thioester and a 33‐mer peptide thioester containing a trimethylated lysine residue. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Methyl (R)‐N‐(2,6‐dimethylphenyl)alaninate ((R)‐DMPM) is a key chiral intermediate for the production of (R)‐metalaxyl, which is one of the best‐selling fungicides. A new strain, Pseudochrobactrum asaccharolyticum WZZ003, was identified as a biocatalyst for the enantioselective hydrolysis of (R,S)‐DMPM. The key parameters including pH, temperature, rotation speed and substrate concentrations were optimized in the enantioselective hydrolysis of (R,S)‐DMPM. After the 48 h hydrolysis of 256 mM (R,S)‐DMPM under the optimized reaction conditions, the enantiomeric excess of product (e.e.p) was up to 99% and the conversion was nearly 50%. Subsequently, the unhydrolyzed (S)‐DMPM was converted to (R,S)‐DMPM through the n‐butanal‐catalyzed racemization. Furthermore, stereoselective hydrolysis of (R,S)‐DMPM catalyzed by whole cells of P. asaccharolyticum WZZ003 was scaled up to kilogram‐scale, offering (R)‐MAP‐acid with 98.6% e.e.p and 48.0% yield. Moreover, (R)‐metalaxyl was prepared at kilogram scale after subsequent esterification and coupling reactions. Therefore, a practical production process of (R)‐DMPM and (R)‐metalaxyl with the prospect of industrialization was developed in this study. © 2018 American Institute of Chemical Engineers Biotechnol. Prog., 34:921–928, 2018  相似文献   

5.
A Candida rugosa lipase immobilized on polypropylene powder was employed as the biocatalyst for the enantioselective hydrolysis of (R, S)-suprofen 2,2,2-trifluorothioester in cyclohexane, in which trioctylamine was added as the catalyst to perform in situ racemization of the remaining (R)-thioester. A hollow-fiber membrane was also integrated with the dynamic kinetic resolution process in order to continuously extract the desired (S)-suprofen into an aqueous solution containing NaOH. A kinetic model for the whole process (operating in batch and feed-batch modes) was developed, in which enzymatic hydrolysis and deactivation, lipase activation, racemization and non-enantioselective hydrolysis of the substrate by trioctylamine, and reactive extraction of (R)- and (S)-suprofen into the aqueous phase in the membrane were considered. Theoretical predictions from the model for the time-course variations of substrate and product concentrations in each phase were compared with experimental data.  相似文献   

6.
CdTe nanocrystals (NCs) were fabricated through an organic synthesis. The growth and properties of CdTe NCs depended strongly on the preparation conditions. In a reaction system of octadecene and tetradecylphosphonic acid (TDPA), the growth was slow. CdTe NCs with cubic‐like morphology were created in trioctylamine (TOA) using octadecylphosphonic acid (ODPA)‐CdO or TDPA‐CdO as precursors. The TOA and ODPA system gives rise to NCs with high photoluminescence (PL) efficiencies (12%). A CdxZn1‐xS shell coating on the CdTe core, gave rise to tunable dark red PL (630–670 nm). The morphology and PL properties of the CdTe cores were drastically affected by shell coating and this determined the properties of CdTe/CdxZn1‐xS NCs. Small CdTe NCs were easily coated with CdxZn1‐xS shells. The resulting core/shell NCs revealed a spherical morphology. However, shell growth became slow when large CdTe cores were used. This is ascribed to the cores with a cubic‐like morphology. CdS interlayer plays an important role for the formation of the CdTe/CdxZn1‐xS NCs because the experimental result indicated it is difficult to coat CdTe NCs with a ZnS shell. The core/shell NCs benefited from a CdxZn1‐xS composite shell because CdTe/CdS NCs created via a similar procedure revealed a low PL efficiency. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

7.
Native chemical ligation is widely used for the convergent synthesis of proteins. The peptide thioesters required for this process can be challenging to produce, particularly when using Fmoc‐based solid‐phase peptide synthesis. We have previously reported a route to peptide thioesters, following Fmoc solid‐phase peptide synthesis, via an NS acyl shift that is initiated by the presence of a C‐terminal cysteine residue, under mildly acidic conditions. Under typical reaction conditions, we occasionally observed significant thioester hydrolysis as a consequence of long reaction times (~48 h) and sought to accelerate the reaction. Here, we present a faster route to peptide thioesters, by replacing the C‐terminal cysteine residue with selenocysteine and initiating thioester formation via an NSe acyl shift. This modification allows thioester formation to take place at lower temperatures and on shorter time scales. We also demonstrate how application of this strategy also accelerates peptide cyclization, when a linear precursor is furnished with an N‐terminal cysteine and C‐terminal selenocysteine. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Stereoselective arylpropionyl-CoA thioester formation in vitro   总被引:3,自引:0,他引:3  
The inversion from R- to S-enantiomer that occurs for some arylpropionic acids may have both toxicological and therapeutic implications. To characterize some properties of this inversion, arylpropionyl-CoA thioester formation was studied in rat tissue homogenates and subcellular fractions for the enantiomers of fenoprofen, ibuprofen, and flurbiprofen. Thioesters were formed from (R)-fenoprofen (64%) and (R)-ibuprofen (33%) but not from the corresponding S-enantiomers or the enantiomers of flurbiprofen. This correlates with the extensive inversion of fenoprofen and ibuprofen and lack of inversion of flurbiprofen in vivo. Subcellular fractions from rat liver showed thioester formation to occur in mitochondria and microsomes but not cytosol. Once formed, the thioesters were readily racemized by whole rat liver homogenate, mitochondria, and cytosol, but only partially inverted (S:R = 0.3) in microsomes. Thioester formation from fenoprofen and ibuprofen was studied in tissue homogenate obtained from liver, diaphragm, kidney, lung, skeletal muscle, smooth muscle, fat, caecum, and intestines. The liver was at least 50-fold more efficient than the other tissues studied and would be expected to be a major organ of enantiomeric inversion. Our data support the hypothesis that R- to S-enantiomeric inversion of arylpropionic acids proceeds via the stereoselective formation of CoA thioesters followed by enzymatic racemization and hydrolysis of the thioesters to regenerate free acid.  相似文献   

9.
In order to synthesize interferon‐γ glycoform having an oligosaccharide at the 97 position by a semi‐synthetic method, interferon‐γ‐polypeptide‐(1–94)‐α‐hydrazide was prepared by the specific Cys‐cyanylation of polypeptide‐(1–94)‐Cys‐His6 expressed from E. coli and subsequent hydrazinolysis in 22% yield (two steps). This polypeptide‐α‐hydrazide was then converted into corresponding polypeptide‐α‐thioester under NaNO2/acid conditions followed by thiolysis in 83% yield. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
The pyrolytic behaviour of (?)‐(S)‐nicotine in methanol was investigated using on‐line pyrolysis GC/MS to establish whether racemization to the R(+) antipode occurs and to identify other products of pyrolysis. The conditions used included pyrolysing the sample for 15 seconds in an atmosphere of 9% oxygen in nitrogen (275ml/min total flow) across the temperature range of 200°C–1000°C. A chiral Cyclodex‐B analytical column (30m × 0.25mm i.d. × 0.25 μm film thickness) was used to separate the enantiomers of nicotine, although the two enantiomer peaks were not baseline resolved. The results of the experiment shows that there is no increase in (+)‐(R)‐nicotine levels across a wide temperature range. This suggests that the elevated levels of (+)‐R‐nicotine observed in tobacco smoke (compared to tobacco leaf material) are not due to the pyrolytic auto‐racemization of (?)‐(S)‐nicotine but are a result of more complex interactions between (?)‐(S)‐nicotine and other smoke components. The pyrolysis of isotopically labelled nicotine established that nicotine undergoes thermal decomposition to β‐nicotyrine which in turn may decompose to other products. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

11.
Alloy CdTe1‐xSex quantum dots (QDs) have been fabricated by an organic route using Cd, Te and Se precursors in a mixture of trioctylamine and octadecylphosphonic acid at 280 °C. The variation of photoluminescence (PL) peak wavelength of the CdTe1‐xSex QDs compared with CdTe QDs confirmed the formation of an alloy structure. The Se component drastically affected the stability of CdTe1‐xSex QDs. A Cd0.5Zn0.5S shell coating on CdTe1‐xSex cores was carried out using oleic acid as a capping agent. CdTe1‐xSex/Cd0.5Zn0.5S core/shell QDs revealed dark red PL while a yellow PL peak was observed for the CdTe1‐xSex cores. The PL efficiency of the core/shell QDs was drastically increased (less than 1% for the cores and up to 65% for the core/shell QDs). The stability of QDs in various buffer solutions was investigated. Core/shell QDs can be used for biological applications because of their high stability, tunable PL and high PL efficiency. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

12.
β‐cyclodextrin (CD) and its derivatives HP‐β‐CD, DM‐β‐CD, and TM‐β‐CD have been employed as chiral selectors for the separation of three nonsteroidal antiinflammatory drugs (NSAIDs) and anticoagulant at relatively low concentration (8–15 mM) by capillary zone electrophoresis (CZE). In this study, baseline separation was achieved for ibuprofen, ketoprofen, naproxen, and warfarin. It was found that the addition of 0.1% hydroxypropyl methyl cellulose (HPMC) was effective for separation. Under these conditions, the S‐(+) enantiomer eluted before R‐(−) in terms of ibuprofen; the calculated energy values obtained from the molecular modeling correlated well with the elution order. An equation for calculating the pKa values by capillary electrophoresis was introduced, and the pKa values of the four chiral drugs at 25°C were obtained based on the equation. The value pKa + 0.5 is proposed to be the suitable pH of the background electrolyte for the separation of chiral compounds containing a carboxylic group. Chirality 11:56–62, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

13.
Most amino acids contain chiral centres and exist as both D‐enantiomer and L‐enantiomer. The optically pure enantiomer is often more valuable than the racemate. Enzymatic resolution provides an effective strategy to obtain optically pure amino acids but often results in large amounts of unwanted isomer. In this study, optically pure L‐glufosinate (L‐PPT) was obtained by coupling amidase‐mediated hydrolysis of N‐phenylacetyl‐D,L‐glufosinate with racemization of N‐phenylacetyl‐D‐glufosinate (NPDG), which exclusively exhibits effective herbicidal properties compared with its D‐enantiomer. To improve the yield of L‐PPT, the racemization reaction conditions were optimized, and through single‐factor experiments, the optimal reaction temperature, reaction time, and mole ratio of phenylacetic acid to NPDG were determined to be 150°C, 30 minutes, and 1.5, respectively. The response surface methodology was applied to further optimize the racemization conditions, and the final yield of L‐PPT reached 96.13% with optimum reaction temperature of 154°C, reaction time of 23 minutes, and phenylacetic acid/NPDG mole ratio of 1.7, respectively. Moreover, adding a small amount of acetic anhydride further raised the yield of L‐PPT to 97.02%.  相似文献   

14.
A specially designed model reactor based on a 42‐L laboratory fermentor was equipped with six stirrers (Rushton turbines) and five cylindrical disks. In this model reactor, the mixing time, Θ90, turned out to be 13 times longer compared with the 42‐L standard laboratory fermentor fitted with two Rushton turbines and four wall‐fixed longitudinal baffles. To prove the suitability of the model reactor for scaledown studies of mixing‐time‐dependent processes, parallel exponential fed‐batch cultivations were carried out with the leucine‐auxotrophic strain, Corynebacterium glutamicum DSM 5715, serving as a microbial test system. L‐ Leucine, the process‐limiting substrate, was fed onto the liquid surface of both reactors. Cultivations were conducted using the same inoculum material and equal oxygen supply. The model reactor showed reduced sugar consumption (−14%), reduced ammonium consumption (−19%), and reduced biomass formation (−7%), which resulted in a decrease in L ‐lysine formation (−12%). These findings were reflected in less specific enzyme activity, which was determined for citrate synthase (CS), phosphoenolpyruvate carboxylase (PEP‐C), and aspartate kinase (AK). The reduced specific activity of CS correlated with lower CO2 evolution (−36%) during cultivation. The model reactor represents a valuable tool to simulate the conditions of poor mixing and inhomogeneous substrate distribution in bioreactors of industrial scale. © 1999 John Wiley & Sons, Inc. Biotechnol Bioeng 64: 599–606, 1999.  相似文献   

15.
Peptide segment couplings are now widely utilized in protein chemical synthesis. One of the key structures for the strategy is the peptide thioester. Peptide thioester condensation, in which a C‐terminal peptide thioester is selectively activated by silver ions then condensed with an amino component, is a powerful tool. But the amino acid adjacent to the thioester is at risk of epimerization. During the preparation of peptide thioesters by the Boc solid‐phase method, no substantial epimerization of the C‐terminal amino acid was detected. Epimerization was, however, observed during a thioester–thiol exchange reaction and segment condensation in DMSO in the presence of a base. In contrast, thioester–thiol exchange reactions in aqueous solutions gave no epimerization. The epimerization during segment condensation was significantly suppressed with a less polar solvent that is applicable to segments in thioester peptide condensation. These results were applied to a longer peptide thioester condensation. The epimer content of the coupling product of 89 residues was reduced from 27% to 6% in a condensation between segments of 45 and 44 residues for the thioester and the amino component, respectively. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Analogues of coenzyme A (CoA) and of CoA thioesters have been prepared in which the amide bond nearest the thiol group has been modified. An analogue of acetyl-CoA in which this amide bond is replaced with an ester linkage was a good substrate for the enzymes carnitine acetyltransferase, chloramphenicol acetyltransferase, and citrate synthase, with K(m) values 2- to 8-fold higher than those of acetyl-CoA and V(max) values from 14 to >80% those of the natural substrate. An analogue in which an extra methylene group was inserted between the amide bond and the thiol group showed less than 4-fold diminished binding to the three enzymes but exhibited less than 1% activity relative to acetyl-CoA with carnitine acetyltransferase and no measurable activity with the other two enzymes. Analogues of several CoA thioesters in which the amide bond was replaced with a hemithioacetal linkage exhibited no measurable activity with the appropriate enzymes. The results indicate that some aspects of the amide bond and proper distance between this amide and the thiol/thioester moiety are critical for activity of CoA ester-utilizing enzymes.  相似文献   

17.
Phosphofructokinase‐1 from Saccharomyces cerevisiae is composed of two types of subunits, α and β. Subunit‐specific monoclonal antibodies were raised to elucidate structural and functional properties of both subunits. One monoclonal antibody, α‐F3, binds to an epitope either at the C‐terminal or at the N‐terminal part of the α‐polypeptide chain. By screening a heptapeptide library with this monoclonal antibody, a set of heptapeptides was selected, which contained the consensus sequences D–A–F and D–S–F. Two heptapeptides with these motifs were synthesized in order assess their capacity to inhibit the binding of antibody α‐F3 to native phosphofructokinase‐1. The peptide G–I–K–D–A–F–L inhibited the binding more strongly (IC50 = 1.5 µM) than the peptide A–P–W–H–D–S–F (IC50 = 33.3 µM). Sequence matching revealed the presence of the D–A–F motif in the polypeptide chain of phosphofructokinase‐1 at amino acid position 172–174. As a control, the nonapeptide A–P–T–S–K–D–A–F–L which corresponds to the sequence of the putative epitope was tested in the inhibition assay. In view of the high inhibitory capacity (IC50 = 0.3 µM) it was concluded that this nonapeptide represents the continuous epitope of phosphofructokinase‐1 that is recognized by antibody α‐F3. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

18.
Recent studies have suggested that the isomerization/racemization of aspartate residues in proteins increases in aged tissues. One such residue is Asp151 in lens‐specific αA‐crystallin. Although many isomerization/racemization sites have been reported in various proteins, the factors that lead to those modifications in proteins in vivo remain obscure. Therefore, an in vitro system is needed to assess the mechanisms of modifications of Asp under various conditions. Deamidation of Asn to Asp in proteins occurs more rapidly than isomerization/racemization of Asp, although the reaction passes through the same intermediate in both pathways. Here, therefore, we replaced Asp151 in human lens αA‐crystallin with Asn by using site‐directed mutagenesis. The recombinant protein was expressed in Escherichia coli and used to investigate the deamidation/isomerization/racemization of Asn151 after incubation at 50°C for various durations and under different pH. After incubation, the mutant αA‐crystallin was subjected to enzymatic digestion followed by liquid chromatography–MS/MS to evaluate the ratio of modifications in Asn151‐containing peptides. The Asp151Asn αA‐crystallin mutant showed rapid deamidation to Asp with the formation of specific Asp isomers. In particular, deamidation increased greatly under basic conditions. By contrast, subunit–subunit interactions between αA‐crystallin and αB‐crystallin had little effect on the modification of Asn151. Our findings suggest that the Asp151Asn αA‐crystallin mutant represents a good in vitro model protein to assess deamidation, isomerization, and the racemization intermediates. Furthermore, our in vitro results show a different trend from in vivo data, implying the presence of specific factors that induce racemization from L‐Asp to D‐Asp residues in vivo.  相似文献   

19.
The β‐adrenergic receptors mediate several physiological processes including heart rate (β1), bronchodilation (β2), and lipolysis (β3). Therefore, selectivity is important for a possible therapeutic agent acting via these receptors. Aryloxypropanolamines are β‐receptor agonists or antagonists, depending on the aryl group and its substituents. We therefore hypothesized that fluorine substitution on the aromatic ring in this class could lead to significant biological effects because of the unique chemical characteristics of fluorine. Because the target compound has a chiral center, we set out to synthesize the two enantiomers so that effects of stereochemistry on biological activity could be evaluated. Syntheses of the enantiomers were performed starting with commercially available fluoronaphthalene and subsequent use of the chiral synthon (2R)‐ or (2S)‐glycidyl 3‐nitrobenzenesulfonate, depending on the desired enantiomer. High‐pressure liquid chromatography (HPLC) methods were used to characterize %ee. Each enantiomer was synthesized. They exhibited nanomolar binding activities on β‐adrenergic receptors. The (S)‐enantiomer was found to be up to 310 times more potent than the (R). It was also found to be about five‐fold more selective for β2‐ than for β1‐receptors. The current report demonstrates the importance of stereochemistry for the fluoroaromatic β‐receptor ligands. Chirality 11:144–148, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

20.
Phosphonium and uronium salt‐based reagents enable efficient and effective coupling reactions and are indispensable in peptide chemistry, especially in machine‐assisted SPPS. However, after the activating and coupling steps with these reagents in the presence of tertiary amines, Fmoc derivatives of Cys are known to be considerably racemized during their incorporation. To avoid this side reaction, a coupling method mediated by phosphonium/uronium reagents with a weaker base, such as 2,4,6‐trimethylpyridine, than the ordinarily used DIEA or that by carbodiimide has been recommended. However, these methods are appreciably inferior to the standard protocol applied for SPPS, that is, a 1 min preactivation procedure of coupling with phosphonium or uronium reagents/DIEA in DMF, in terms of coupling efficiency, and also the former method cannot reduce racemization of Cys(Trt) to an acceptable level (<1.0%) even when the preactivation procedure is omitted. Here, the 4,4′‐dimethoxydiphenylmethyl and 4‐methoxybenzyloxymethyl groups were demonstrated to be acid‐labile S‐protecting groups that can suppress racemization of Cys to an acceptable level (<1.0%) when the respective Fmoc derivatives are incorporated via the standard SPPS protocol of phosphonium or uronium reagents with the aid of DIEA in DMF. Furthermore, these protecting groups significantly reduced the rate of racemization compared to the Trt group even in the case of microwave‐assisted SPPS performed at a high temperature. © 2013 The Authors. European Peptide Society published by John Wiley & Sons, Ltd.  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