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.
The reaction of NCA's with some amino acids having a nucleophilic functional group on the side chain was studied in a heterogeneous reaction medium (acetonitrile-water). Glutamic acid and aspartic acid, having a free carboxyl group on the side chain, were successfully used to synthesize oligopeptides without interactions of the γ- and β-carboxyl group with NCA's. Two products were obtained by the reaction of NCA with L -lysine, which contains a free amino group on the side chain. ε-Protected lysine was used to prepare α-peptides as a nucleophile in the reaction. No racemization was observed in the synthesis of peptides by the NCA method in the heterogeneous solvent system. Oligopeptides with some polar side chains were synthesized by the NCA method.  相似文献   

2.
In the polymerizations of alanine, γ-ethyl glutamate, and leucine N-carboxyanhydrides (NCA's) initiated by tertiary amines and some secondary amines such as N-methyl-L -alanine dialkylamide, a stereoselectivity was observed: the polymerization rates of L - and D -NCA's were identical to each other and larger than that of DL -NCA. However, this selectivity was not observed in the polymerizations of valine and isoleucine NCA's initiated by N-methyl-L -alanine dialkylamide. The stereoselective polymerizations of valine and isoleucine NCA's were induced only with tetriary amines such as tri-n-butylamine. N-Methyl-L -alanine di-alkylamide has been shown to initiate the polymerization of usual α-amino acid NCA according to the activated-NCA mechanism, but it initiated the polymerizations of valine and isoleucine NCA's according to the primary amine-type mechanism. This is because in the latter NCA's the N–H group is masked by the adjacent Cβ-branched alkyl substituent against the approach of the secondary amine. Poly(DL -alanine)s produced in the stereoselective polymerization had higher viscosities and were more stereoblock-like than those produced without the stereoselectivity. These experimental results indicate that the stereoselective polymerization is possible only when the polymerization proceeds through the activated-NCA mechanism.  相似文献   

3.
K R Anumula  R P Schulz  N Back 《Peptides》1992,13(4):663-669
Highly fluorescent N-methylanthranilyl (Mantyl) peptide derivatives were prepared by a one-step reaction with N-methylisatoic anhydride (MIA) for quantitative detection in HPLC. Reactions were carried out in an organic medium of acetonitrile-triethylamine, in aqueous alkaline sodium carbonate and sodium phosphate buffers. 4-Dimethylaminopyridine (DMAP) catalyzed specific mantylation of -NH2 groups of peptides in the organic reaction medium. The DMAP had no effect in the aqueous buffered reaction systems. Proline amino-terminal peptides reacted equally well with MIA. Mantyl-bradykinin had excitation and fluorescence maxima at 350 nm and 426 nm in water and water/acetonitrile (ACN)/trifluoroacetic acid (TFA) solvent mixtures, respectively. Fluorescence intensity increased with an increase in ACN concentration and decreased with an increase in acid content. Mantyl kinins were completely resolved on a C18 reversed phase HPLC column using an ACN-0.1% TFA gradient and their behavior on the column was similar to having an extra amino acid. Di-Mantyl derivatives obtained with Lys-BK and Met-Lys-BK did not exhibit fluorescence appreciably higher than Mantyl-BK. Fluorescence detection of Mantyl kinins was about 50-100 times more sensitive (lower limits of 0.1 to 0.5 picomole) than UV detection of the phenylisothiocyanate-derivatized kinins under typical HPLC conditions.  相似文献   

4.
The polymerizations of D ,-L β-phenylalanine NCA, p–nitro-D ,L -β-phenylalanine NCA, and o,p-dinitro-D ,L -β-phenylalanine NCA were investigated, homopolymers and copolymers of N-vinyl-2-ethylimidazole or 2-Vinylpyridine being used as catalysts. When N-vinylpyrrolidone and N,N-diethylacrylamide, which are capable of forming hydrogen bonds with the NCA's, were used as comonomers with N-vinyl-2-ethylimidazole, the copolymer catalysts were found to bring about a faster polymerization than poly-N-vinyl-2-ethylimidazole. However, when styrene, which has no particular interaction with the NCA's, was used as a comonomer with 2-vinylpyridine, the copolymer catalyst was found to give a slower polymerization than poly-2-vinylpyridine. Electronic spectroscopy showed that the charge-transfer complex between copolymer catalysts and the NCA's plays an important role in the polymerization. The experimental results are discussed in terms of the effectiveness of the copolymer catalysts for forming hydrogen bonds or charge-transfer complexes with the NCA's.  相似文献   

5.
Polymerizations of D ,L -β-phenylalanine, p-nitro-D ,L -β- phenylalanine, and o,p-dinitro-D ,L -β-phenylalanine NCA's were carried out with the use of α-picoline or poly-2-vinyl-pyridine as initiator. Polymerizations induced by the polymer catalyst were always faster than those with α-picoline in the same base concentrations. Furthermore, the polymer effect was more marked when the number of nitro groups in the NCA's increased. It was considered that the polymer catalyst interacts with the NCA's primarily by hydrogen bonding and increases the effective concentration of NCA along the chain. The increase of the NCA concentration in the vicinity of the polymer catalyst wits also achieved through charge-transfer complexes between nitrophenyl groups in the NCA's and pyridine groups in the polymer catalyst. As the polymer chain is flexible, a collision between an adsorbed NCA and a pyridine unit in the same polymer chain is favored, thus increasing the polymerization rate.  相似文献   

6.
Enzymes are industrially applied under increasingly diverse environmental conditions that are dictated by the efforts to optimize overall process efficiency. Engineering the operational stability of biocatalysts to enhance their half-lives under the desired process conditions is a widely applied strategy to reduce costs. Here, we present a simple method to enhance enzyme stability in the presence of monophasic aqueous/organic solvent mixtures based on the concept of strengthening the enzyme's surface hydrogen-bond network by exchanging surface-located amino acid residues for arginine. Suitable residues are identified from sequence comparisons with homologous enzymes from thermophilic organisms and combined using a shuffling approach to obtain an enzyme variant with increased stability in monophasic aqueous/organic solvent mixtures. With this approach, we increase the stability of the broad-spectrum amino acid racemase of Pseudomonas putida DSM 3263 eightfold in mixtures with 40% methanol and sixfold in mixtures with 30% acetonitrile.  相似文献   

7.
Esterification of N-acetyl phenylalanine with ethanol catalyzed by immobilized α-chymotrypsin (EC 3.4.21.1) was studied in various reaction media. The effect of reaction medium polarity on enzymatic activity as well as equilibrium yield was measured. The reaction rate increased with increasing amounts of added water, reaching an optimum corresponding to water saturation of the reaction medium. Further additions of water resulted in decreased activity. Bell-shaped activity profiles were obtained for all reaction media tested. Reaction media consisting of pure solvents and of mixtures of solvents were used. The enzymatic activity and the equilibrium yield increased with decreased polarity of the medium. Maximum activity was found in a reaction medium consisting of 80% diisopropyl ether and 20% heptane. The enzymatic activity obtained at optimal water additions in the different solvents and solvents mixtures could be correlated to the solubility of water and the log P of the medium. The equilibrium yield of the reaction was much more closely correlated to the solubility of water than the log P. Much lower enzymatic activity was obtained when solvent mixtures producing water-miscible media were created, than in mixtures producing water-immiscible media, such as mixtures of acetonitrile and diisopropyl ether. The equilibrium yield could be increased by decreasing the water content in the reaction medium, which reduced the water activity.  相似文献   

8.
Circular dichroism, optical rotatory dispersion, and viscosity of hyaluronic acid at various solvents compositions, concentrations, and pH values have been studied. The data show a large change in the molecular properties in organic/water solvents such as ethanol, p-dioxane, or acetonitrile/water at pH ? pKa. At this pH range of aqueous solution, hyaluronic acid shows a CD minimum near 210 nm whereas in the presence of organic solvent it exhibits a strong negative dichroism (below 200 nm) and a positive band near 226 nm. It undergoes a sharp, cooperative transition with respect to pH and solvent. The observed CD features are assigned to the π-π* and n-π* transitions of the amide and carboxyl chromophores. The ORD results show a gradual blue shift of trough at 220 nm with increasing magnitude of rotation when the organic solvents and hydrogen ion concentrations are increased. A one-term Drude's equation was used to analyze the ORD data, and the result show a variation of dispersion parameters with different solvents in accordance with the observed CD changes. The intrinsic viscosity of hyaluronic acid in mixed solvent at pH 2.6 is lower than that of aqueous solution. All the observed property changes of hyaluronic acid are reversed on addition of foramide in mixed solvents indicating that the hydrogen bonds are involved in this transition. The observed spectroscopic and hydrodynamic features are attributed to a conformational change of hyaluronic acid in a mixed solvent involving intramolecular hydrogen bonding between the acetamido and carboxyl groups. The possible conformational state of hyaluronic acid in solution under various conditions is discussed in terms of the reported helical structure of hyaluronic acid from x-ray diffraction studies.  相似文献   

9.
Abstract

The effects of the choice of lipase, reaction medium, immobilization, presence of additives and temperature on conversion and stereoselectivity during a lipase catalysed decarboxylative aldol reaction were examined. It was shown that Candida antarctica lipase B (CALB) catalysed a decarboxylative aldol reaction between 4-nitrobenzaldehyde and ethyl acetoacetate in a 60% acetonitrile–40% aqueous buffer co-solvent mixture. Interestingly, free and immobilized forms of CALB showed opposite enantioselectivity in this media. The addition of 30 mol% imidazole increased the reaction rate from 8.5 to 55.7 μM min??1 mg??1. A 98% conversion could be achieved in 14 h (instead of 168 h) by adding imidazole. Other lipases also catalysed this reaction in different reaction media to a varying extent. With Mucor javanicus lipase in 30% DMSO, 20% enantiomeric excess (ee) of the (R)-product was observed. CALB also catalysed this reaction in nearly anhydrous acetonitrile. In the presence of cross-linked protein coated microcrystals of CALB, 90% conversion was obtained in this media in 24 h. A commercially available protease, alcalase, was also found to catalyse this reaction. While low water media gave poor conversion, the reaction in aqueous–60% acetonitrile co-solvent mixture gave 99% conversion in 72 h, provided imidazole was used as an additive.  相似文献   

10.
Microbial oxidations of n-tetradecane, tetradecanols and tetradecanoic acid were investigated by using intact cells of Corynebacterium equi, a hydrocarbon-assimilating bacterium, in an aqueous phase and organic solvents. The bacterial cells were hydrophobic and could be well dispersed in all organic solvents employed to give homogeneous reaction mixtures, and among them, isooctane was found to be the best for the reaction. n-Tetradecane and tetradecanoic acid were completely oxidized in the aqueous phase, but not in isooctane. In contrast, 1-tetradecanol was oxidized much more readily in isooctane than in the aqueous phase, and an oxidation product identified as myristyl myristate was accumulated in isooctane at the conversion rate of 80%. 2-Tetradecanol was also readily oxidized in isooctane, and 2-tetradecanone was obtained at the conversion rate of nearly 100%. Similar results were obtained when toluene and n-hexane were used as the solvent in place of isooctane, while no reaction was observed when chloroform was employed.  相似文献   

11.
The influence on lipase activity in water of a pretreatment on Candida rugosa lipase using water miscible and immiscible solvents was studied. The lipase activity in the hydrolysis of esteric substrates in aqueous media increases when the lipase was previously treated with various nearly anhydrous organic media. This activation, which was irreversible, was higher for longer pretreatment times. It was dependent on the pretreatment medium (water activity and solvent used). A relation between variations in the emission intensity and the activities of treated and untreated lipases was found. Activating pretreatment did not shift the peak of fluorescence emission but gave rise to variations in the secondary protein structure by increasing the helical nature. A similar increment in the hydrolysis rate in water can be obtained with the addition of an appropriate amount of solvent (acetonitrile or n-heptane) to the aqueous reaction medium.  相似文献   

12.
Two statistical methods were used for medium optimization for a hydrophilic solvent‐stable protease production by Serratia sp. SYBC H with duckweed as the nitrogen source. Orthogonal design was applied to find the significant variables, then response surface methodology (RSM), including Box–Behnken central composite experiments, was used to determine the optimal concentrations and interaction of the significant variables. Results demonstrated that duckweed powder, wheat flour, Tween 80, sodium chloride had significant effects on the solvent‐stable protease production. The interaction between duckweed and wheat flour was significant. The optimal level of the variables for the maximum protease production was duckweed 43.9 g/L, wheat flour 20 g/L, sodium chloride 0.08 M, Tween 80 1% v/v, initial pH 11.0, and inoculum size 7% v/v. The maximum protease activity reached 1922.8 U/mL in the optimized medium, with about 18.3‐fold higher than that in the unoptimized medium. Most importantly, the protease from Serratia sp. SYBC H has successfully catalyzed the specific acylation of sucrose in a two‐solvent medium consisting of pyridine and n‐hexane (1:1, v/v), and non‐specific acylation of sucrose in anhydrous DMSO. These results demonstrated that the protease from Serratia sp. SYBC H is a solvent‐stable protease and it could be an ideal biocatalyst for sugar esters syntheses in non‐aqueous media.  相似文献   

13.
Titration of an aqueous solution of sodium poly-L-glutamate with a strong acid usually produces turbidity and precipitation before the equivalence point is reached. In 1M sodium p-toluenesulfonate and 1M sodium methylcyclohexanesulfonate aggregation was delayed for days to months. In very concentrated tetra-n-butylammonium chloride and bromide the polymeric acid dissolved and could be titrated with NaOH. However, potentiometric titration curves in these solutions did not yield information (by way of plots of apparent pK versus the degree of neutralization) about the helix-to-coil transition of the polymer. It was argued in addition that the apparent pK of a weak polyelectrolyte should not be calculated from titrations in concentrated salt solutions since it is a mixed or composite quantity. It contains not only the effect of the salt on the dissociation of the weak electrolyte but its effect on the activity coefficient of hydrogen ions as well. Circular dichroic spectra were therefore measured a t various degrees of neutralization of poly(L -glutamic acid) in a number of aqueous salt solutions and mixtures of organic solvents with water. It was found that the undissociated polymer in the concentrated methylcyclohexanesulfonate and quaternary ammonium halide solutions had the spectrum of a right-handed α helix. The n → π* band at 222 mμ was used as a measure of the fraction of polymer in the helical conformation. The value of ? Δε for the undissociated polyacid in these organic electrolyte solutions was 11.4. By means of this value and a number of assumptions, the fraction of helix (fh) as a function of α, the degree of polyacid neutralization, was calculated for the different solvent mixtures. An empirical equation was used to describe the variation of fh with α, fh = 1/(1 + e?a+bα), in which b represents the degree of cooperativeness of the transition, and a is a measure of the effect of the medium on the onset of the transition. The values of b did not differ very much from one another, suggesting that the cooperativeness of the transition was not sensitive to changes in the medium. On the other hand, the value of a (or its equivalent, the value of α at the halfway point in the transition) was more dependent on the solvent. Comparisons of these results with those of some other workers were made by means of the empirical equation.  相似文献   

14.
This paper describes a rapid method to identify the best solvent and carrier compound combinations with the highest extraction capability and the lowest microbial toxicity characteristics for product recovery from microbial fermentation. The extraction system has an aqueous phase, and an emulsion phase, which was a blend of sodium carbonate and organic phase [91% (v/v) organic solvent, 5% (v/v or wt/v) carrier compound, and 4% (v/v) surfactant Span 80]. Alamine 336, or tri-n-octylamine in n-heptane; Alamine 336, Alamine 304, or tributyl phosphate in hexane; and Alamine 304 or tributyl phosphate in iso-octane; Alamine 304 or Amberlite in xylene demonstrated high lactic acid extraction. For determination of bacterial toxicity of selected solvent and carrier compounds, Lactobacillus casei subsp. rhamnosus (ATCC 11443) was grown in LAF medium containing one of the selected organic solvent, carrier compound, and Span 80 in 250 ml flask at 37 °C and 125 rpm. Samples were collected regularly during 48 hour incubation, and measured for changes in cell density by absorbance at 620 nm, cell count using a fluorescent dye with flow cytometry, and lactic acid, and glucose concentrations by HPLC. Hexadecane:tributyl phosphate, n-dodecane:tri-n-octylamine, and kerosene:tri-n-octylphosphine oxide demonstrated the least microbial toxicity among the tested blends with excess solvent media. Whereas, hexanes:Alamine 304 and xylenes:Alamine 304 were nontoxic in solvent saturated media.This revised version was published online in October 2005 with corrections to the Cover Date.  相似文献   

15.
A high-performance liquid chromatographic (HPLC) method is described for the simultaneous determination of acetylsalicylic acid (ASA) and its main metabolite salicylic acid (SA) in human plasma. Acidified plasma is deproteinized with acetonitrile which is separated from the aqueous layer by adding sodium chloride. ASA and SA are extracted into the acetonitrile layer with high yield, and determined by reversed-phase HPLC (column: Novapak C18 4 μm silica,150×4mm I.D.; eluent: 740 ml water, 900 μl 85% orthophosphoric acid, 180 ml acetonitrile) and photometric detection (237 nm). 2-Methylbenzoic acid is used as internal standard. The method allows the determination of ASA and SA in human plasma as low as 100 ng/ml with good precision (better than 10%). The assay was used to determine the pharmacokinetic parameters of ASA and SA following oral administration of 100–500 mg ASA in healthy volunteers.  相似文献   

16.
A high-performance liquid chromatography (HPLC) method for determining sophoridine in rat plasma was developed for application in the pharmacokinetic studies. The plasma was deproteinized with acetonitrile that contained an internal standard (ephedrine) and was separated from the aqueous layer by adding sodium chloride and sodium carbonate. The HPLC assay was carried out using a YMC-ODS column. The mobile phase was methanol-ethanol-0.01 moll(-1) ammonium acetate buffer-triethylamine (10:0.5:89.5:0.03, v/v/v/v) (pH 6.80). The flow rate was 0.8 ml min(-1). The detection wavelength was set at 210 nm. The method was used to determine the concentration-time profiles of sophoridine in the plasma following oral administration or injection of sophoridine aqueous solution. The fractions of sophoridine reaching the systemic circulation were estimated for the first time by a deconvolution method.  相似文献   

17.
The effect of various organic solvents on the catalytic activity, stability and substrate specificity of alchohol dehydrogenase from Haloferax volcanii (HvADH2) was evaluated. The HvADH2 showed remarkable stability and catalysed the reaction in aqueous?Corganic medium containing dimethyl sulfoxide (DMSO) and methanol (MeOH). Tetrahydrofuran and acetonitrile were also investigated and adversely affected the stability of the enzyme. High concentration of salt, essential to maintain the enzymatic activity and structural integrity of the halophilic enzyme under standard conditions may be partially replaced by DMSO and MeOH. The presence of organic solvents did not induce gross changes in substrate specificity. DMSO offered a protective effect for the stability of the enzyme at nonoptimal pHs such as 6 and 10. Salt and solvent effects on the HvADH2 conformation and folding were examined through fluorescence spectroscopy. The fluorescence findings were consistent with the activity and stability results and corroborated the denaturing properties of some solvents. The intrinsic tolerance of this enzyme to organic solvent makes it highly attractive to industry.  相似文献   

18.
Biphasic aqueous‐organic systems are important reaction systems for catalytic processes. This is especially true for biocatalysis where the range of accessible products can be significantly extended. In such systems, the aqueous phase is the reactive phase in which the biocatalyst is dissolved and the organic phase is nonreactive and acts as substrate reservoir and as in situ product extraction solvent. Here, the choice of the nonreactive phase is highly important for the overall performance of the system. In this contribution, a systematic approach to solvent selection for biphasic aqueous‐organic systems is presented with respect to partition coefficients. The model reaction is the stereoselective carbon‐carbon coupling of two 3,5‐dimethoxy‐benzaldehyde molecules to (R)‐3,3',5,5'‐tetramethoxy‐benzoin catalyzed by benzaldehyde lyase (EC 4.1.2.38) from Pseudomonas fluorescens. A systematic approach to solvent selection consisting of two steps is proposed: Firstly, the conductor‐like screening model for real solvents (COSMO‐RS) is used to facilitate a fast solvent screening. Since this is an ab initio approach it allows a pre‐screening without laborious experimental input. The proposed ranking of solvents, based on the ratio of partition coefficients at infinite dilution, is a sound basis for the successive steps. Secondly, a dynamic model is fitted to experimental data in order to obtain detailed and reliable results for mass transfer and partition coefficients. Therefore, the method makes efficient use of the experimental data and substantiates quantitative results with guided experiments.  相似文献   

19.
A molybdenum trioxide thin film with smooth surface and uniform thickness was successfully achieved by an electrospray deposition method using an aqueous solution with a drastically low concentration of 0.05 wt%. Previous papers demonstrated that an additive solvent technique is useful for depositing the thin film by the electrospray deposition, and the high vapor pressure and a low surface tension of an additive solvent were found to be important factors. As a result, the smooth molybdenum trioxide thin film was obtained when the acetonitrile was used as the additive solvent. Furthermore, the vapor pressure of acetone is much higher than that of aqueous solution, and this indicates that the acetone is easily evaporated after spraying from the glass capillary. By optimizing a concentration of acetone in the molybdenum aqueous solution, a minimum root mean square roughness of the MoO3 thin film became 3.7 nm. In addition, an organic photovoltaic cell was also demonstrated using the molybdenum trioxide as a hole transport layer. Highest photoconversion efficiency was 1.72%, a value comparable to that using conventional thermal evaporation process even though the aqueous solution was used for the solution process. The photovonversion efficiency was not an optimized value, and the higher value can be achieved by optimizing the coating condition of the active layer.  相似文献   

20.
The method for the assay of β-galactosidase introduced by Lederberg (1950) was modified and used for studies of the enzyme in the soil.o-Nitrophenyl-β-d-galactoside served as substrate. Trismaleinate buffer was found to be more suitable than phosphate buffer in the assay. The enzyme reaction was stopped by adding sulphuric acid and the incubation mixture was alkalized with sodium carbonate so that the yellow colour ofo-nitrophenol could develop. The method is sensitive and specific and requires small quantities of soil and a short incubation time.  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