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.
Isopenicillin N synthase (IPNS) catalyses the synthesis of isopenicillin N (IPN), the biosynthetic precursor to penicillin and cephalosporin antibiotics. IPNS is a non‐heme iron(II) oxidase that mediates the oxidative cyclisation of the tripeptide δ‐L ‐α‐aminoadipoyl‐L ‐cysteinyl‐D ‐valine (ACV) to IPN with a concomitant reduction of molecular oxygen to water. Solution‐phase incubation experiments have shown that, although IPNS can turn over analogues with a diverse range of hydrocarbon side chains in the third (valinyl) position of its substrate, the enzyme is much less tolerant of polar residues in this position. Thus, although IPNS converts δ‐L ‐α‐aminoadipoyl‐L ‐cysteinyl‐D ‐isoleucine (ACI) and AC‐D ‐allo‐isoleucine (ACaI) to penam products, the isosteric sulfur‐containing peptides AC‐D ‐thiaisoleucine (ACtI) and AC‐D ‐thia‐allo‐isoleucine (ACtaI) are not turned over. To determine why these peptides are not substrates, we crystallized ACtaI with IPNS. We report the synthesis of ACtaI and the crystal structure of the IPNS:FeII:ACtaI complex to 1.79 Å resolution. This structure reveals direct ligation of the thioether side chain to iron: the sulfide sulfur sits 2.66 Å from the metal, squarely in the oxygen binding site. This result articulates a structural basis for the failure of IPNS to turn over these substrates.  相似文献   

2.
Isopenicillin N synthase (IPNS) is a nonheme iron(II)‐dependent oxidase that catalyses the central step in penicillin biosynthesis, conversion of the tripeptide δ‐L ‐α‐aminoadipoyl‐L ‐cysteinyl‐D ‐valine (ACV) to isopenicillin N (IPN). This report describes mechanistic studies using the analogue δ‐(L ‐α‐aminoadipoyl)‐(3S‐methyl)‐L ‐cysteine D ‐α‐hydroxyisovaleryl ester (ASmCOV), designed to intercept the catalytic cycle at an early stage. ASmCOV incorporates two modifications from the natural substrate: the second and third residues are joined by an ester, so this analogue lacks the key amide of ACV and cannot form a β‐lactam; and the cysteinyl residue is substituted at its β‐carbon, bearing a (3S)‐methyl group. It was anticipated that this methyl group will impinge directly on the site in which the co‐substrate dioxygen binds. The novel depsipeptide ASmCOV was prepared in 13 steps and crystallised with IPNS anaerobically. The 1.65 Å structure of the IPNS–FeII–ASmCOV complex reveals that the additional β‐methyl group is not oriented directly into the oxygen binding site, but does increase steric demand in the active site and increases disorder in the position of the isovaleryl side chain. Crystals of IPNS–FeII–ASmCOV were incubated with high‐pressure oxygen gas, driving substrate turnover to a single product, an ene‐thiol/C‐hydroxylated depsipeptide. A mechanism is proposed for the reaction of ASmCOV with IPNS, linking this result to previous crystallographic studies with related depsipeptides and solution‐phase experiments with cysteine‐methylated tripeptides. This result demonstrates that a (3S)‐methyl group at the substrate cysteinyl β‐carbon is not in itself a block to IPNS activity as previously proposed, and sheds further light on the steric complexities of IPNS catalysis.  相似文献   

3.
A process to obtain L ‐valine has been developed using fluidized and packed bed reactors with L ‐aminoacylase (from hog kidney) immobilized by covalent binding. L ‐Valine production using the immobilized derivative of L ‐aminoacylase in fluidized and packed bed reactors was studied at three different substrate concentrations and two different flow rates. Higher productions were obtained in the packed bed reactor in all cases. The different solubilities of L ‐valine and acetyl‐D ‐valine in ethanol were used to purify L ‐amino acid from the reactor effluents. The amount of added ethanol did not influence the separation yields, although the purity of L ‐valine was strongly affected by this parameter. The last step involved was racemization of the unhydrolyzed acetyl‐D ‐valine, which was then used as substrate in a new reaction cycle. © 1999 Society of Chemical Industry  相似文献   

4.
L ‐α‐Glycerylphosphorylcholine (L ‐α‐GPC) was successfully prepared from phosphatidylcholine (PC) of food‐grade soy lecithin powder using a novel enzymatic reaction in an aqueous medium. 94.5% yield of L ‐α‐GPC was obtained under the optimal conditions of 55°C, 6.67 mg/mL substrate, 2 mM CaCl2, and 33.4 U/mL phospholipase A1 (Lecitase Ultra). L ‐α‐GPC at 98% purity, 73.4% (wt%) recovery, and specific rotation ( ) of ?2.5° was achieved by silica gel column chromatography. Owing to its excellent catalytic efficiency, low cost, and ready availability, phospholipase A1 (Lecitase Ultra) provides a very satisfactory option for converting PC to L ‐α‐GPC. Practical applications: L ‐α‐Glycerylphosphorylcholine (L ‐α‐GPC) has been studied recently for its potential use as a supplement that may support neurological functions, but it is only found in trace amounts in nature. The present results indicate that Lecitase Ultra can be used for producing L ‐α‐GPC from aqueous PC and suggest encouraging prospects for practical or industrial applications utilizing its notable catalytic performance, economy, and convenience.  相似文献   

5.
L ‐α‐Aminoadipic acid reductases catalyze the ATP‐ and NADPH‐dependent reduction of L ‐α‐aminoadipic acid to the corresponding 6‐semialdehyde during fungal L ‐lysine biosynthesis. These reductases resemble peptide synthetases with regard to their multidomain composition but feature a unique domain of elusive function—now referred to as an adenylation activating (ADA) domain—that extends the reductase N‐terminally. Truncated enzymes based on NPS3, the L ‐α‐aminoadipic acid reductase of the basidiomycete Ceriporiopsis subvermispora, lacking the ADA domain either partially or entirely were tested for activity in vitro, together with an ADA‐adenylation didomain and the ADA domainless adenylation domain. We provide evidence that the ADA domain is required for substrate adenylation: that is, the initial step of the catalytic turnover. Our biochemical data are supported by in silico modeling that identified the ADA domain as a partial peptide synthetase condensation domain.  相似文献   

6.
Biodegradable polymers/oligomers were successfully synthesized through a ring‐opening polymerization of ε‐caprolactone and L ,L ‐lactide, initiated by L ‐arginine and L ‐citrulline. The α‐amino acid initiators are natural, operationally simple, inexpensive, environmentally friendly and safe for human health. The polymerizations were performed with no solvents and without introducing any metal impurities. The chemical structures of the polymers obtained were elucidated using 1H NMR, 13C NMR and Fourier transform infrared spectroscopies. In addition, incorporation of α‐amino acid molecules into the polymer chain was confirmed using matrix‐assisted laser desorption ionization time‐of‐flight mass spectrometry. Due to the significant biological activity of L ‐arginine and L ‐citrulline, these α‐amino acid initiators may open a new route for the synthesis of functional polymers especially for pharmaceutical applications. Copyright © 2011 Society of Chemical Industry  相似文献   

7.
Based on the combined use of dimethylformamide (DMF) modulation and neighboring group participation, three iterative one‐pot α‐glycosylation methods, i.e., one‐pot (α,α)‐, one‐pot (β,α)‐, and one‐pot (α,β)‐glycosylations, were developed. These methods are applicable to a range of thioglycosyl donors, confer stereocontrol in α‐/β‐glycosidic bond formation, and thus provide for rapid access to oligosaccharides with various permutations of anomeric configurations. The utility of these one‐pot glycosylation methods is demonstrated in the synthesis of eight non‐natural and natural oligosaccharide targets, including the core 1 serine conjugate, core 8 serine conjugate, the D ‐Gal‐α(1→3)‐D ‐Glc‐α(1→3)‐L ‐Rha trisaccharide unit of the cell wall component in Streptococcus pneumoniae, and the D ‐Glc‐α(1→2)‐D ‐Glc‐α(1→3)‐D ‐Glc trisaccharide terminus of the N‐linked glycan precursor. Confirmation of the anomeric configurations of these oligosaccharides is evidenced by 1H, 13C, 13C‐non‐proton decoupling, and heteronuclear correlation 2D NMR experiments. Global deprotection of selected oligosaccharide targets is illustrated.  相似文献   

8.
Pyromellitic dianhydride (1,2,4,5‐benzenetetracarboxylic acid 1,2,4,5‐dianhydide) was reacted with L ‐valine in a mixture of acetic acid and pyridine (3:2) at room temperature, and then was refluxed at 90–100 °C, N,N′‐(pyromellitoyl)‐bis‐L ‐valine diacid was obtained in quantitative yield. The imide–acid was converted to N,N′‐(pyromellitoyl)‐bis‐L ‐valine diacid chloride by reaction with thionyl chloride. Rapid and highly efficient synthesis of a number of poly(amide–imide)s was achieved under microwave irradiation using a domestic microwave oven by polycondensation of N,N′‐(pyromellitoyl)‐bis‐L ‐valine diacid chloride with six different derivatives of 5,5‐disubstituted hydantoin compounds in the presence of a small amount of a polar organic medium that acts as a primary microwave absorber. A suitable organic medium was o‐cresol. The polycondensation proceeded rapidly, compared with conventional melt polycondensation and solution polycondensation and was almost completed within 8 min, giving a series of poly(amide–imide)s with inherent viscosities in the range 0.15–0.36 dl g?1. The resulting poly(amide–imide)s were obtained in high yield and are optically active and thermally stable. All of the above compounds were fully characterized by Fourier‐transform infrared (FT‐IR) spectroscopy, elemental analysis, inherent viscosity (ηinh) measurements, solubility testing and specific rotation measurements. The thermal properties of the poly(amide–imide)s were investigated by using thermogravimetric analysis. Copyright © 2004 Society of Chemical Industry  相似文献   

9.
β‐Lactam synthetase (β‐LS) is the paradigm of a growing class of enzymes that form the critical β‐lactam ring in the clavam and carbapenem antibiotics. β‐LS catalyzes a two‐stage reaction in which N2‐(2‐carboxyethyl)‐L ‐arginine is first adenylated, and then undergoes intramolecular ring closure. It was previously shown that the forward kinetic commitment to β‐lactam formation is high, and that the overall rate of reaction is partially limited to a protein conformational change rather than to the chemical step alone of closing the strained ring. β‐Lactam formation was evaluated on the basis of X‐ray crystal structures, site‐specific mutation, and kinetic and computational studies. The combined evidence clearly points to a reaction coordinate involving the formation of a tetrahedral transition state/intermediate stabilized by a conserved Lys. The combination of substrate preorganization, a well‐stabilized transition state and an excellent leaving group facilitates this acyl substitution to account for the strong forward commitment to catalysis and to lower the barrier of four‐membered ring formation to the magnitude of a protein conformational change.  相似文献   

10.
Copolymerization of mixtures of L ‐lactide and ε‐caprolactone has been initiated by diphenylzinc. The reaction conditions were investigated, to discover the effects on yield, molecular weight and microstructure of copolymers obtained. The temperature used varied between 50 and 120 °C, the molar ratio of monomer to initiator ranged between 90 and 1440 mol/mol, and the molar ratio of ε‐caprolactone to L ‐lactide employed was between 100/0 and 0/100 mol/mol. Copolymers were characterized by 1H‐NMR, 13C‐NMR, DSC and gel permeation chromatography. The results indicate that incorporation of L ‐lactide to the growing chain is preferred and ε‐caprolactone is copolymerized after most of the L ‐lactide has been depleted. The microstructure of obtained copolyesters was affected considerably by transesterification reactions. It was observed that increasing reaction temperature, reaction time and concentration initiator was advantageous to the transesterification. The crystallinity of copolyester obtained was determined by differential scanning calorimetry. The results are in good agreement with both molar composition and sequence distribution of copolyesters. Copyright © 2006 Society of Chemical Industry  相似文献   

11.
A novel enzymatic production system of optically pure β‐hydroxy α‐amino acids was developed. Two enzymes were used for the system: an N‐succinyl L ‐amino acid β‐hydroxylase (SadA) belonging to the iron(II)/α‐ketoglutarate‐dependent dioxygenase superfamily and an N‐succinyl L ‐amino acid desuccinylase (LasA). The genes encoding the two enzymes are part of a gene set responsible for the biosynthesis of peptidyl compounds found in the Burkholderia ambifaria AMMD genome. SadA stereoselectively hydroxylated several N‐succinyl aliphatic L ‐amino acids and produced N‐succinyl β‐hydroxy L ‐amino acids, such as N‐succinyl‐L ‐β‐hydroxyvaline, N‐succinyl‐L ‐threonine, (2S,3R)‐N‐succinyl‐L ‐β‐hydroxyisoleucine, and N‐succinyl‐L ‐threo‐β‐hydroxyleucine. LasA catalyzed the desuccinylation of various N‐succinyl‐L ‐amino acids. Surprisingly, LasA is the first amide bond‐forming enzyme belonging to the amidohydrolase superfamily, and has succinylation activity towards the amino group of L ‐leucine. By combining SadA and LasA in a preparative scale production using N‐succinyl‐L ‐leucine as substrate, 2.3 mmol of L ‐threo‐β‐hydroxyleucine were successfully produced with 93% conversion and over 99% of diastereomeric excess. Consequently, the new production system described in this study has advantages in optical purity and reaction efficiency for application in the mass production of several β‐hydroxy α‐amino acids.

  相似文献   


12.
Comparison of the production of L ‐phenylacetylcarbinol (L ‐PAC) from benzaldehyde by cells of Torulaspora delbrueckii immobilized in calcium alginate and barium alginate showed that cells immobilized in calcium alginate beads were usable for up to six cycles compared with nine cycles for barium alginate‐immobilized cells. Barium alginate‐immobilized cells yielded a total of 2.17 g of L ‐PAC in nine cycles whereas in the case of calcium alginate‐immobilized cells, a total of 1.17 g of L ‐PAC was formed when 400 mg benzaldehyde and 400 mm3 of acetaldehyde were used in each cycle as substrate and co‐substrate respectively. Copyright © 2003 Society of Chemical Industry  相似文献   

13.
We report the first heterologous production of a fungal rutinosidase (6‐O‐α‐L ‐rhamnopyranosyl‐β‐D ‐glucopyranosidase) in Pichia pastoris. The recombinant rutinosidase was purified from the culture medium to apparent homogeneity and biochemically characterized. The enzyme reacts with rutin and cleaves the glycosidic linkage between the disaccharide rutinose and the aglycone. Furthermore, it exhibits high transglycosylation activity, transferring rutinose from rutin as a glycosyl donor onto various alcohols and phenols. The utility of the recombinant rutinosidase was demonstrated by its use for the synthesis of a broad spectrum of rutinosides of primary (saturated and unsaturated), secondary, acyclic and phenolic alcohols as well as for the preparation of free rutinose. Moreover, the α‐L ‐rhamnosidase‐catalyzed synthesis of a chromogenic substrate for a rutinosidase assay – para‐nitrophenyl β‐rutinoside – is described.

  相似文献   


14.
With D,L ‐lactic acid and Nϵ‐carbobenzoyloxy‐L ‐lysine [Lys(Z)] as the starting monomer material and tin dichloride as the catalyst, the drug carrier material poly(lactic acid‐coNϵ‐carbobenzoyloxy‐L ‐lysine) was synthesized via direct melt polycondensation. The copolymer was systematically characterized with intrinsic viscosity testing, Fourier transform infrared spectroscopy, 1H‐NMR, gel permeation chromatography, differential scanning calorimetry, and X‐ray diffraction. The influences of different feed molar ratios were examined. With increasing molar feed content of Lys(Z), the intrinsic viscosity, weight‐average molecular weight, and polydispersity index (weight‐average molecular weight/number‐average molecular weight) gradually decreased. Because of the introduction of Lys(Z) with a big aromatic ring into the copolymer, the glass‐transition temperature gradually increased with increasing feed charge of Lys(Z), and all of the copolymers were amorphous. The copolymers, with weight‐average molecular weights from 10,500 to 6900 Da, were obtained and could reach the molecular weight level of poly(lactic acid) modified by Lys(Z) via the ring‐opening polymerization of the cyclic intermediates, such as lactide and morpholine‐2,5‐dione. However, a few terminal carboxyl groups might have been deprotected during the polymerization reaction under high temperatures. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

15.
Charged polypeptides containing sulfonate groups were prepared by transesterification of poly(γ‐methyl L ‐glutamate) with isethionic acid. The coil–helix transition of the sulfonated polypeptides was investigated in aqueous alcohols. Marked counter‐ion specificity was observed for helix formation: Li+ < Na+ < Cs+ ≦ Rb+ ≦ K+; this was different to that for poly(L ‐glutamate) (PLG): Cs+ ? K+ < Li+ < Na+. Specific helix stabilization by counter‐ion mixing, which has been found for the PLG system, was not observed for the sulfonated polypeptides. The counter‐ion‐ and solvent‐specific helix formation is discussed and compared with that in PLG. © 2001 Society of Chemical Industry  相似文献   

16.
Four poly(γ‐benzyl‐L ‐glutamate) (PBLG) derivatives bearing at one end specific groups were synthesized by ring‐opening polymerization of the corresponding γ‐benzyl‐L ‐glutamate N‐carboxyanhydride using different amine‐terminated initiators. These moieties were chosen to introduce, on demand, specific functionalities in nanoparticles of pharmaceutical interest. The PBLG and PBLG derivatives were characterized by 1H NMR, viscosimetry, Fourier transform infrared spectroscopy and differential scanning calorimetry. Nanoparticles smaller than 100 nm in diameter could be easily prepared from these PBLG derivatives by slight modification of a known nanoprecipitation technique. Copyright © 2006 Society of Chemical Industry  相似文献   

17.
Phase behavior of octahydro‐1,3,5,7‐tetranitro‐1,3,5,7‐tetrazocine (HMX) is investigated by X‐ray powder diffraction (XRD). The XRD patterns at elevated temperature show that there is a co‐existing temperature range of β‐ and δ‐phase during the phase transition process. Additionally, mechanical forces can catalyze the conversion from δ‐ back to β‐phase. Based on the diffraction patterns of β‐ and δ‐phase at different temperatures, we calculate the coefficients of thermal expansion by Rietveld refinement. For β‐HMX, the linear coefficients of thermal expansion of a‐axis and b‐axis are about 1.37×10−5 and 1.25×10−4 °C−1. A slight decrease in c‐axis with temperature is also observed, and the value is about −0.63×10−5 °C−1. The volume coefficient of thermal expansion is about 1.60×10−4 °C−1, with a 2.2% change from 30 to 170 °C. For δ‐HMX, the linear coefficients of thermal expansion of a‐axis and c‐axis are found to be 5.39×10−5 and 2.38×10−5 °C−1, respectively. The volume coefficient of thermal expansion is about 1.33×10−4 °C−1, with a 2.6% change from 30 to 230 °C. The results indicate that β‐HMX has a similar volume coefficient of thermal expansion compared with δ‐HMX, and there is about 10.5% expansion from β‐HMX at 30 °C to δ‐HMX at 230 °C, of which about 7% may be attributed to the reconstructive transition.  相似文献   

18.
Poly(L ‐lactide‐co‐ε‐caprolactone)‐b‐poly(L ‐lactide) [P(LL‐co‐CL)‐b‐PLL] diblock copolyesters were synthesized in a two‐step process with 1‐dodecanol (DDC) and stannous octoate as the initiating system. In the first‐step reaction, a 50:50 mol % amorphous poly(L ‐lactide‐co‐ε‐caprolactone) [P(LL‐co‐CL)] copolyester was synthesized via the bulk copolymerization of L ‐lactide and ε‐caprolactone, which was followed by the polymerization of the PLL crystalline block at the end chain in the second‐step reaction. The yielded copolyesters were characterized with dilute‐solution viscometry, gel permeation chromatography, 1H‐ and 13C‐NMR, and differential scanning calorimetry methods. The molecular weights of the P(LL‐co‐CL) copolyesters from the first‐step reaction were controlled by the DDC concentrations, whereas in the second‐step reaction, the molecular weights of the P(LL‐co‐CL)‐b‐PLL diblock copolyesters depended on the starting P(LL‐co‐CL) copolyester molecular weights and L ‐lactide/prepolymer molar ratios. The starting P(LL‐co‐CL) copolyester molecular weights and PLL block lengths seemed to be the main factors affecting specific thermal properties, including the melting temperature (Tm), heat of melting (ΔHm), crystallizing temperature (Tc), and heat of crystallizing (ΔHc), of the final P(LL‐co‐CL)‐b‐PLL diblock copolyester products. Tm, ΔHm, Tc, and ΔHc increased when the PLL block lengths increased. However, these thermal properties of the diblock copolyesters also decreased when the P(LL‐co‐CL) block lengths increased. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2007  相似文献   

19.
The DNA adsorption properties of poly‐L ‐lysine‐immobilized poly(2‐hydroxyethyl methacrylate) (pHEMA) membrane were investigated. The pHEMA membrane was prepared by UV‐initiated photopolymerization and activated with epichlorohydrin. Poly‐L ‐lysine was then immobilized on the activated pHEMA membrane by covalent bonding, via a direct chemical reaction between the amino group of poly‐L ‐lysine and the epoxy group of pHEMA. The poly‐L ‐lysine content of the membrane was determined as 1537 mg m?2. The poly‐L ‐lysine‐immobilized membrane was utilized as an adsorbent in DNA adsorption experiments. The maximum adsorption of DNA on the poly‐L ‐lysine‐immobilized pHEMA membrane was observed at 4 °C from phosphate‐buffered salt solution (pH 7.4, 0.1 M; NaCl 0.5 M) containing different amounts of DNA. The non‐specific adsorption of DNA on the plain pHEMA membrane was low (about 263 mg m?2). Higher DNA adsorption values (up to 5849 mg m?2) were obtained in which the poly‐L ‐lysine‐immobilized pHEMA membrane was used. Copyright © 2003 Society of Chemical Industry  相似文献   

20.
BACKGROUND: Biodegradable block copolymers have attracted particular attention in both fundamental and applied research because of their unique chain architecture, biodegradability and biocompatibility. Hence, biodegradable poly[((R )‐3 ‐hydroxybutyrate)‐block‐(D ,L ‐lactide)‐block‐(ε‐caprolactone)] (PHB‐PLA‐PCL) triblock copolymers were synthesized, characterized and evaluated for their biocompatibility. RESULTS: The results from nuclear magnetic resonance spectroscopy, gel permeation chromatography and thermogravimetric analysis showed that the novel triblock copolymers were successfully synthesized. Differential scanning calorimetry and wide‐angle X‐ray diffraction showed that the crystallinity of PHB in the copolymers decreased compared with methyl‐PHB (LMPHB) oligomer precursor. Blood compatibility experiments showed that the blood coagulation time became longer accompanied by a reduced number of platelets adhering to films of the copolymers with decreasing PHB content in the triblocks. Murine osteoblast MC3T3‐E1 cells cultured on the triblock copolymer films spread and proliferated significantly better compared with their growth on homopolymers of PHB, PLA and PCL, respectively. CONCLUSION: For the first time, PHB‐PLA‐PCL triblock copolymers were synthesized using low molecular weight LMPHB oligomer as the macroinitiator through ring‐opening polymerization with D ,L ‐lactide and ε‐caprolactone. The triblock copolymers exhibited flexible properties with good biocompatibility; they could be developed into promising biomedical materials for in vivo applications. Copyright © 2008 Society of Chemical Industry  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