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 422 毫秒
1.
N‐Acetyl‐D ‐neuraminic acid (Neu5Ac) was efficiently synthesized from lactate and a mixture of N‐acetyl‐D ‐glucosamine (GlcNAc) and N‐acetyl‐D ‐mannosamine (ManNAc) by whole cells. The biotransformation utilized Escherichia coli cells (Neu5Ac aldolase), Pseudomonas stutzeri cells (lactate oxidase components), GlcNAc/ManNAc and lactate. By this process, 18.32±0.56 g/liter Neu5Ac were obtained from 65.61±2.70 g/liter lactate as an initial substrate input. Neu5Ac (98.4±0.4 % purity, 80.87±0.79 % recovery yield) was purified by anionic exchange chromatography. Our results demonstrate that the reported Neu5Ac biosynthetic process can compare favorably with natural product extraction or chemical synthesis processes.  相似文献   

2.
The ordered honeycomb structures of poly(L ‐lactic‐co‐glycolic acid) and poly(D ,L ‐lactic‐co‐glycolic acid) fabricated in a humid atmosphere were reported in this paper. It was found that surfactants were important in the formation of honeycomb films of hydrophobic polymer. The affecting factors, such as the environment temperature, the atmosphere humidity, and the concentration of the polymer solution of the honeycomb porous structure, were also tested. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 100: 1013–1018, 2006  相似文献   

3.
Using racemic tert‐leucine amide as sole nitrogen source in minimal medium, 162 strains were isolated by enrichment techniques and shown to contain amidase activity. Among these isolates three D ‐amidase producers were found and identified as Variovorax paradoxus (two strains) and Klebsiella spec. The D ‐amidase from Variovorax paradoxus was purified to homogeneity by three chromatographic steps. With dl ‐Tle‐amide as substrate Michaelis Menten kinetics were observed with a KM of 0.74 mM, a KI of 640 mM and a Vmax of 1.4 U/mg. The amidase has a broad pH‐optimum between 7 and 9.5 and a temperature optimum at 47–49 °C. The amidase hydrolyzed amino acid amides as well as carboxamides and 2‐hydroxy acid amides. The stereoselectivity of the reaction was variable, however. Hydrolyzing dl ‐Tle‐amide the enantiomeric ratio E was >200 resulting in D ‐Tle with an ee of >99% and up to 47% conversion. Similar results were obtained with dl ‐Leu‐amide and dl ‐Val‐amide while dl ‐Phe‐amide was hydrolyzed with an enantiomeric ratio E of only 5.  相似文献   

4.
BACKGROUND: In syntheses of biodegradable and bioresorbable polymers, efficient metal‐free catalysts are very desirable as the resulting products may be more biocompatible. As an attempt to find new metal‐free catalysts, 1,3‐dialkylimidazolium salts, the most commonly used ionic liquids or organic melt salts, were used as single‐component catalysts in the melt polycondensation of L ‐lactic acid for the first time. The resulting poly(L ‐lactic acid) (PLLA) was characterized using gel permeation chromatography, 13C NMR, DSC and polarimetry. RESULTS: It has been found that less bulky substituents on the imidazolium ring are conducive to catalytic activity. PLLA with molar mass of about 20 000 g mol?1 was synthesized at high yield (over 70%) in the presence of various 1,3‐dialkylimidazolium salts. The product exhibits satisfactory color (white to slightly yellow), optical purity (89–95%) and crystallinity (40–55%). A possible catalytic mechanism is proposed. CONCLUSION: As compared with the well‐known binary catalyst system SnCl2 · 2H2O/toluene sulfonic acid, the catalysts used in this investigation are better in terms of increasing PLLA yield and preventing discoloration and comparable in terms of racemization. Copyright © 2008 Society of Chemical Industry  相似文献   

5.
The effects caused by alkaline treatment on the susceptibility of waste cardboard to enzymatic hydrolysis have been studied. Optimised conditions leading to extensive saccharification of both cellulose (870 g kg?1 conversion) and hemicelluloses (845 g kg?1 conversion) were identified. Samples treated under selected operational conditions were employed for producing D ‐lactic acid by simultaneous saccharification and fermentation (SSF) in media containing cellulases, β‐glucosidase and Lactobacillus coryniformis ssp torquens cells. SSF fed‐batch experiments led to D ‐lactic acid concentrations up to 23.4 g dm?3 at a product yield of 514 g lactic acid kg?1 of potential glucose and a volumetric productivity of 0.48 g dm?3 h?1. Copyright © 2004 Society of Chemical Industry  相似文献   

6.
A novel L ‐pantolactone hydrolase, Lph, from Agrobacterium tumefaciens Lu681 was characterized, which stereospecifically hydrolyses L ‐pantolactone to L ‐pantoic acid yielding D ‐pantolactone with > 95% enantiomeric excess. The enzyme was found to be a 30 kDa‐Zn2+‐hydrolase with a Km for L ‐pantolactone of 7 mM and a Vmax of 30 U/mg. The corresponding lph gene was identified as an 807 bp ORF and cloned into E. coli. It was overexpressed under control of Ptac and Prha yielding enzyme activities of up to 600 U/g dry weight. Resolution of d,l ‐pantolactone in repeated batches with isolated Lph and enzyme recovery by membrane filtration gave D ‐pantolactone with 50% yield and 90–95% ee over 6 days. Covalent immobilization to EupergitC led to a stable biocatalyst easy to handle in a repeated batch production of D ‐pantolactone. Further improvements in the activity of Lph were achieved by directed evolution of the enzyme. Activities of mutants F62S, K197D and F100L were increased 2.3, 1.7, and 1.5 fold, respectively.  相似文献   

7.
Here we report the best artificial metalloenzyme to date for the selective oxidation of aromatic alkenes; it was obtained by noncovalent insertion of MnIIImeso‐tetrakis(p‐carboxyphenyl)porphyrin [Mn(TpCPP), 1 ‐Mn] into a host protein, xylanase 10A from Streptomyces lividans (Xln10A). Two metallic complexes—N,N′‐ethylene bis(2‐hydroxybenzylimine)‐5,5′‐dicarboxylic acid MnIII [(Mn‐salen), 2 ‐Mn] and 1 ‐Mn—were associated with Xln10A, and the two hybrid biocatalysts were characterised by UV–visible spectroscopy, circular dichroism and molecular modelling. Only the artificial metalloenzyme based on 1 ‐Mn and Xln10A was studied for its catalytic properties in the oxidation of various substituted styrene derivatives by KHSO5: after optimisation, the 1 ‐Mn ‐ Xln10A artificial metalloenzyme was able to catalyse the oxidation of para‐methoxystyrene by KHSO5 with a 16 % yield and the best enantioselectivity (80 % in favour of the R isomer) ever reported for an artificial metalloenzyme.  相似文献   

8.
Lipase B from Candida antarctica (CAL‐B) catalyzes the slow, but highly enantioselective (E>200), ring‐opening alcoholysis of two bicyclic and two 4‐aryl‐substituted β‐lactams. Surprisingly, the rate of the reaction varies with the nature of the alcohols and was fastest with either enantiomer of 2‐octanol. A 0.5‐g scale reaction with 2‐octanol as the nucleophile in diisopropyl ether at 60 °C yielded the unreacted β‐lactam in 39–46% yield (maximum yield is 50%) with ≥96% ee. The product β‐amino acid esters reacted further by polymerization (not isolated or characterized) or by hydrolysis due to small amounts of water in the reaction mixture yielding β‐amino acids (7–11% yield, ≥96% ee). The favored enantiomer of all four β‐lactams had similar 3‐D orientation of substituents, as did most previously reported β‐lactams and β‐lactones in similar ring‐opening reactions. Computer modeling of the ring opening of 4‐phenylazetidin‐2‐one suggests that the reaction proceeds via an unusual substrate‐assisted transition state, where the substrate alcohol bridges between the catalytic histidine and the nitrogen of the β‐lactam. Computer modeling also suggested that the molecular basis for the high enantioselectivity is a severe steric clash between Ile189 in CAL‐B and the phenyl substituent on the slow‐reacting enantiomer of the β‐lactam.  相似文献   

9.
BACKGROUND: R‐mandelic acid is an important chiral pharmaceutical intermediate, which is commonly obtained by biotransformation. This work has focused on using novel chiral recognition technology, aqueous two‐phase extraction, for the chiral separation of mandelic acid. RESULTS: The copper (II) formed a 2:1 complex with β‐CD in an alkaline solution, which was isolated from solution by the addition of ethanol. The complex structure was characterized by IR and UV spectroscopy. The chiral recognition system was established by adding Cu2‐β‐CD into the triton‐114 aqueous two‐phase extraction system, which preferentially recognizes the (R)‐enantiomer rather than the (S)‐enantiomer. Factors affecting the extraction mechanism were analyzed, namely the concentration of Cu2‐β‐CD and tritonX‐114, the types of salts, pH, and temperature. It was found that the concentration of Cu2‐β‐CD and temperature were the most important influencing factors for chiral separation of mandedlic acid. The experimental results showed that the ee values increased with pH and concentration of trition‐114, and the maximum ee was 67.91%. The addition of inorganic salt had a strong influence on ee, which decreased when salt was added into the aqueous two‐phase extraction system. CONCLUSION: A novel chiral recognition technology ‐ aqueous two phase extraction is reported in this paper.The tritonX‐114 aqueous two phase system have a good recognition ability for mandelic acid. Copyright © 2012 Society of Chemical Industry  相似文献   

10.
Optically active poly(L ‐phenyllactic acid) (Ph‐PLLA), poly(L ‐lactic acid) (PLLA), and poly(L ‐phenyllactic acid‐co‐L ‐lactic acid) with weight‐average molecular weight exceeding 6 × 103 g mol?1 were successfully synthesized by acid catalyzed direct polycondensation of L ‐phenyllactic acid and/or L ‐lactic acid in the presence of 2.5–10 wt % of p‐toluenesulfonic acid. Their physical properties and crystallization behavior were investigated by differential scanning calorimetry, thermogravimetry, and polarimetry. The absolute value of specific optical rotation ([α]) for Ph‐PLLA (?38 deg dm?1 g?1 cm3) was much lower than that of [α] for PLLA (?150 deg dm?1 g?1 cm3), suggesting that the helical nature was reduced by incorporation of bulky phenyl group. PLLA was crystallizable during solvent evaporation, heating from room temperature, and cooling from the melt. Incorporation of a very low content of bulky phenyllactyl units even at 4 mol % suppressed the crystallization of L ‐lactyl unit sequences during heating and cooling, though the copolymers were crystallizable for L ‐phenylactyl units up to 6 mol % during solvent evaporation. The activation energy of thermal degradation (ΔEtd) for Ph‐PLLA (200 kJ mol?1) was higher than that for PLLA (158 kJ mol?1). The ΔEtd for the copolymers increased with an increase in L ‐phenyllactyl unit content. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

11.
A biocatalytic redox‐neutral reaction cascade was designed for the deracemisation of racemic mandelic acid to yield optically pure L ‐phenylglycine employing three enzymes. The cascade consisted of three steps: a racemisation, an enantioselective oxidation and a stereoselective reductive amination. The enantioselective oxidation of D ‐mandelic acid to the corresponding oxo acid was coupled with the stereoselective reductive amination of the latter; thus the oxidation as well as the reduction reactions were performed simultaneously. The formal hydrogen abstracted in the first step – the oxidation – was consumed in the reductive amination allowing a redox‐neutral cascade due to a cascade‐internal cofactor recycling. The enantiomers of the starting material were interconverted by a racemase (mandelate racemase) ensuring that in theory 100% of the starting material can be transformed. Using this set‐up racemic mandelic acid was transformed to optically pure L ‐phenylglycine (ee >97%) at 94% conversion without the requirement of any additional redox reagents in stoichiometric amounts.  相似文献   

12.
Physically crosslinked hydrogels are formed upon mixing equal amounts of aqueous solutions of dextran grafted with L ‐lactic acid oligomers and dextran grafted with D ‐lactic acid oligomers. In this paper, the dex‐lactate hydrogels are analyzed by X‐ray diffraction to provide evidence that hydrogel formation occurred through stereocomplex formation. First, the stereocomplex formation of monodisperse lactic acid oligomers was studied and comparison with published results of poly(lactic acid) was made. It was clearly demonstrated that stereocomplexes were formed between oligomers of opposite chirality. The data for these monodisperse lactic acid oligomers were used to interpret the results obtained with the dex‐lactate hydrogels. X‐ray diffraction measurements demonstrated that in the hydrogels stereocomplexes were formed between the lactic acid oligomers of opposite chirality. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 86: 289–293, 2002  相似文献   

13.
BACKGROUND: Biocatalysts have gained increasing attention because of their inherent advantages over chemical catalysts. However, the poor operational stability has always prevented their broad application. In this study, (R)‐mandelic acid was chosen as a model compound of alpha‐hydroxy acids. The objective was to obtain a new biocatalyst with desired operational stability for the preparation of (R)‐mandelic acid as well as other optically pure alpha‐hydroxy acids of pharmaceutical importance. RESULTS: Using a two‐step screening strategy, Saccharomyces ellipsoideus GIM2.105 was selected as an effective biocatalyst with high enantioselectivity and remarkable operational stability. After 20 cycles of reuse, whole cells of S. ellipsoideus GIM2.105 maintained its activity, and no obvious decrease in conversion or enantiomeric excess (ee) was observed. Furthermore, effects of various reaction parameters, including pH, temperature, co‐substrate (type, concentration), substrate concentration and reaction time, on the bioreduction were studied. Under optimal conditions, (R)‐mandelic acid and four substituted aromatic (R)‐alpha‐hydroxy acids were prepared in high ee (95–>99%) and good conversion (>90%). CONCLUSION: The high enantioselectivity, remarkable operational stability and mild reaction conditions showed S. ellipsoideus GIM2.105 to be an economical biocatalyst with great industrial application potential for the production of optically active alpha‐hydroxy acids. Copyright © 2009 Society of Chemical Industry  相似文献   

14.
The optimal synthetic conditions of poly(lactic acid‐co‐glycolic acid) (PLGA) via melt copolycondensation directly from L ‐lactic acid (L ‐LA) and glycolic acid (GA) with a feed molar ratio of 50/50 are discussed; the important drug‐delivery carrier PLGA50/50 is used as a special example. With reaction conditions of 165°C and 70 Pa and with 0.5 wt % SnCl2 as the catalyst, 10 h of polymerization gave the L ‐PLGA50/50 with the biggest intrinsic viscosity ([η]), 0.1993 dL/g. The optimal synthetic conditions were verified by the synthesis of D,L ‐PLGA50/50 with D,L ‐lactic acid (D,L ‐LA) instead of L ‐LA, but the biggest [η] was 0.2382 dL/g. Under the same synthetic conditions with L ‐LA and D,L ‐LA as starting materials, serial PLGA with different molar feed ratios, including 100/0, 90/10, 70/30, 50/50, 30/70, 10/90, and 0/100, were synthesized via simple and practical direct melt copolycondensation, and their solubilities were investigated. When the glycolic acid feed molar percentage was equal to or more than 70%, solubilities in tetrahydrofuran and CHCl3 became worse, and some samples were even wholly insoluble. These biodegradable polymers were also systematically characterized with gel permeation chromatography, Fourier transform infrared spectroscopy, 1H‐NMR spectroscopy, differential scanning calorimetry, and X‐ray diffraction. PLGA synthesized from L ‐LA and D,L ‐LA had many differences in weight‐average molecular weight (Mw), glass‐transition temperature, crystallinity, and composition. When the molar feed ratio of LA to GA was 50/50, both the [η] and Mw values of D,L ‐PLGA were higher than those of L ‐PLGA. With D,L ‐LA as the starting material, the structure of the PLGA copolymer was relatively simple, and its properties were apt to be controlled by its GA chain segment. When the feed molar percentage of the monomer (LA or GA) was more than or equal to 90%, the copolymer was apt to be crystalline, and the aptness was more obvious for the L ‐LA monomer. The composition percentage of GA in PLGA was not only higher than the feed molar percentage of GA, but also, the GA percentage in D,L ‐PLGA was higher than in L ‐PLGA. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 99: 244–252, 2006  相似文献   

15.
Starting from D,L ‐acid and SnCl2 as catalyst, poly(D,L ‐lactic acid) (PDLLA) was directly synthesized by melt polycondensation. Under the appropriate conditions such as 0.5 wt % SnCl2, 170–180°C, 70 Pa, and 10 h, the viscosity‐average molecular weight (Mη) of PDLLA was 4100 Da. PDLLA produced by the most practical method was used as the drug‐delivery material for erythromycin and ciprofloxacin. The optimal conditions for the preparation of erythromycin–poly(D,L ‐lactic acid)–microsphere (ERY–PDLLA–MS) for lung targeting was investigated, and further confirmed by good reappearance tests. DSC and SEM demonstrated that ERY–PDLLA–MS had good spherical shape. The release in vitro of ERY–PDLLA–MS was effective and the half‐time (T1/2) was 51.0 h. After 175 h, the accumulated release percentage was 80.0%. The test in vivo showed that ERY–PDLLA–MS was more easily distributed in rabbit lung tissue. When PDLLA was applied in an antibacterial ciprofloxacin drug‐delivery microsphere (CIP–PDLLA–MS), CIP–PDLLA–MS was also characterized with DSC and SEM, and the release T1/2 in vitro was 24.9 h. After 53.2 h, the accumulated release percentage reached 84.0%, which indicated that CIP–PDLLA–MS was advantageous to long‐term release. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 91: 2143–2150, 2004  相似文献   

16.
The utility for carbon‐carbon bond formation of a multienzyme system composed of recombinant dihydroxyacetone kinase (DHAK) from Citrobacter freundii, the fructose bisphosphate aldolase from rabbit muscle (RAMA) and acetate kinase (AK) for adenosine triphosphate (ATP) regeneration has been studied. Several aldehydes with great structural diversity, including three α,β‐unsaturated aldehydes, have been analysed as acceptor substrates. It was found that α,β‐unsaturated aldehydes bearing an electron‐withdrawing group in the β position to the double bond with a trans configuration are good acceptors for RAMA in this multienzyme system. The aldol reaction proceeds with excellent D ‐threo enantioselectivity and the aldol adduct is obtained in good overall yield. The L ‐threo and D ‐erythro enantiomers are also accessible from rhamnulose 1‐phosphate aldolase (Rha‐1PA) and fuculose 1‐phosphate aldolase (Fuc‐1PA) catalysed reactions, respectively.  相似文献   

17.
The 4‐ulose and the 3‐ulose, both derived in two steps from the α‐methyl glycoside of N‐acetyl‐D ‐glucosamine (GlcNAc), act as organocatalysts in the asymmetric epoxidation of alkenes, with unprecedented complementary enantioselectivity. The best results are found with α,β‐unsaturated esters as substrates, with enantiomeric ratios up to 90:10 and 11:89, respectively.  相似文献   

18.
BACKGROUND: A variety of nitrogen sources were tested for lactic acid production. Corn steep liquor is a low‐cost by‐product that could replace some of the expensive nutrients of the general lactobacilli media. This work deals with the optimisation of the composition of a low‐cost medium for lactic acid production from apple pomace by Lactobacillus rhamnosus CECT‐288. RESULTS: Corn steep liquor (CSL) and yeast extract (YE) were evaluated as nutrient sources for lactic acid production from apple pomace. In comparison with media containing CSL, experiments with YE led to higher volumetric productivities but lower lactic acid concentrations and product yields. The presence of YE increased the production of acetic acid. In media containing 0.2 g CSL g?1 potential sugars, 29.5 g lactic acid L?1 was obtained after 24 h, at good yield (35.5 g per 100 g dry apple pomace) and productivity (1.23 g L?1 h?1), with a lactic acid/acetic acid mass ratio of 98 g g?1. CONCLUSION: The experimental results proved that CSL is a suitable nutrient source for lactic acid production from apple pomace, even though the volumetric productivity was lower than in experiments employing YE. Considered as a nutrient supplement, CSL presents advantages over YE not only in terms of cost but also in terms of product yield, final lactic acid concentration and lactic acid/acetic acid mass ratio. The utilisation of apple pomace as the raw material and CSL as the sole nutrient source (both cheap by‐products) enables the production of lactic acid by an economical, environmentally friendly process. Copyright © 2008 Society of Chemical Industry  相似文献   

19.
Multiblock copolymers consisting of poly(l ‐lactic acid) and poly(dimethyl siloxane) were prepared by the polycondensation of oligo(l ‐lactic acid) (OLLA) with dihydroxyl‐terminated oligo(dimethyl siloxane) and dicarboxyl‐terminated oligo(dimethyl siloxane). Copolymers with number‐average molecular weights of 18,000?33,000 Da and various content ratios of oligo(dimethyl siloxane) (ODMS) unit were obtained by changing the feed ratio of these oligomers. A film prepared from the copolymer with an ODMS content ratio of 0.37 exhibited two independent peaks at ?107°C and 37°C in the mechanical loss tangent for temperature dependence, suggesting the formation of microphase separation between the OLLA and ODMS segments. The film had a tensile strength of 3.2 MPa and a high elongation of 132%. The film also exhibited a high strain recovery even after repeated straining. The incorporation of dimethyl siloxane units as multiblock segments was confirmed to improve the flexibility of poly(l ‐lactic acid). © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 40211.  相似文献   

20.
The efficient synthesis of pure d ‐glycerate‐2‐phosphate is of great interest due to its importance as an enzyme substrate and metabolite. Therefore, we investigated a straightforward one‐step biocatalytic phosphorylation of glyceric acid. Glycerate‐2‐kinase from Thermotoga maritima was expressed in Escherichia coli, allowing easy purification. The selective glycerate‐2‐kinase‐catalyzed phosphorylation was followed by 31P NMR and showed excellent enantioselectivity towards phosphorylation of the d ‐enantiomer of glyceric acid. This straightforward phosphorylation reaction and subsequent product isolation enabled the preparation of enantiomerically pure d ‐glycerate 2‐phosphate. This phosphorylation reaction, using recombinant glycerate‐2‐kinase, yielded d ‐glycerate 2‐phosphate in fewer reaction steps and with higher purity than chemical routes.  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