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.
Syntheses and biodegradation of statistical copolymers of D ,L ‐lactide (D ,L ‐LA) with trimethylene carbonate (TMC), rac‐1‐methyltrimethylene carbonate (1‐MTMC) and 2,2‐dimethyltrimethylene carbonate (2,2‐DTMC) were investigated at various monomer ratios using SmMe(C5Me5)2THF as an initiator at 80 °C for 24 h in toluene. Biodegradations of poly(D ,L ‐LA‐co‐racemo‐1‐MTMC) (95/5) and poly(D ,L ‐LA‐co‐2,2‐DTMC) (98/2) with a compost at 60 °C proceed rapidly. Enzymatic degradations of these polymers were also performed using cholesterol esterase, lipoprotein lipase and proteinase K. Only poly(D ,L ‐LA‐co‐TMC) was biodegraded with cholesterol esterase, while poly(TMC), poly(1‐MTMC), poly(2,2‐DTMC) and poly(D ,L ‐LA) were barely degraded with these enzymes. Biodegradations of poly(D ,L ‐LA‐co‐TMC) (87/13) and poly(D ,L ‐LA‐co‐racemo‐1‐MTMC) (95/5) are rapid using proteinase K. Physical properties of these copolymers were also described. © 2003 Society of Chemical Industry  相似文献   

2.
Recently, we reported on a transaldolase B variant (TalB F178Y) that is able to use dihydroxyacetone (DHA) as donor in aldol reactions. In a second round of protein engineering, we aimed at improving the affinity of this variant towards nonphosphorylated acceptor aldehydes, that is, glyceraldehyde (GA). The anion binding site was identified in the X‐ray structure of TalB F178Y where a sulfate ion from the buffer was bound in the active site. Therefore, we performed site‐directed saturation mutagenesis at three residues forming the putative phosphate binding site, Arg181, Ser226 and Arg228. The focused libraries were screened for the formation of D ‐fructose from DHA and d,l ‐GA by using an adjusted colour assay. The best results with respect to the synthesis of D ‐fructose were achieved with the TalB F178Y/R181E variant, which exhibited an at least fivefold increase in affinity towards d,l ‐GA (KM=24 mM ). We demonstrated that this double mutant can use D ‐GA, glycolaldehyde and the L ‐isomer, L ‐GA, as acceptor substrates. This resulted in preparative synthesis of D ‐fructose, D ‐xylulose and L ‐sorbose when DHA was used as donor. Hence, we engineered a DHA‐dependent aldolase that can synthesise the formation of polyhydroxylated compounds from simple and cheap substrates at preparative scale.  相似文献   

3.
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   

4.
Directly starting from lactic acid (LA) and poly(ethylene glycol) (PEG), biodegradable material polylactic acid‐polyethylene glycol (PLEG) was synthesized via melt copolycondensation. The optimal synthetic conditions, including prepolymerization method, catalyst kinds and quantity, copolymerization temperature and time, LA stereochemical configuration, feed weight ratio mLA/mPEG and Mn of PEG, were all discussed in detail. When D ,L ‐LA and PEG (Mn = 1000 Da) prepolymerized together as feed weight ratio mD ,l‐LA/mPEG = 90/10, 15 h copolycondensation under 165°C and 70 Pa, and 0.5 wt % SnO as catalyst, gave D ,L ‐PLEG1000 with the highest [η] of 0.40 dL/g, and the corresponding MW was 41,700 Da. Using L ‐LA instead of D ,L ‐LA, 10 h polymerization under 165°C and 70 Pa, and 0.5 wt % SnO as catalyst, gave L ‐PLEG1000 with the highest [η] of 0.21 dL/g and MW of 15,600 Da. Serial D ,L ‐PLEG with different feed weight ratio and Mn of PEG were synthesized via the simple and practical direct melt copolycondensation, and characterized with FTIR, 1H NMR, GPC, DSC, XRD, and contact angle testing. D ,L ‐PELG not only had higher MW than PDLLA, PLLA and L ‐PELG, but also better hydrophilicity than PDLLA. The novel one‐step method could be an alternative route to the synthesis of hydrophilic drug delivery carrier PLEG instead of the traditional two‐step method using lactide as intermediate. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 577–587, 2006  相似文献   

5.
The four stereoisomers of azetidine‐2,3‐dicaroxylic acid (L ‐trans‐ADC, L ‐cis‐ADC, D ‐trans‐ADC, and D ‐cis‐ADC) were synthesized in a stereocontrolled fashion following two distinct strategies: one providing the two cis‐ADC enantiomers and one giving access to the two trans‐ADC enantiomers. The four azetidinic amino acids were characterized in a radioligand binding assay ([3H]CGP39653) at native NMDA receptors: L ‐trans‐ADC showed the highest affinity (Ki=10 μM ) followed by the D ‐cis‐ADC stereoisomer (21 μM ). In contrast, the two analogues L ‐cis‐ADC and D ‐trans‐ADC were low‐affinity ligands (>100 and 90 μM , respectively). Electrophysiological characterization of the ADC compounds at the four NMDA receptor subtypes NR1/NR2A, NR1/NR2B, NR1/NR2C, and NR1/NR2D expressed in Xenopus oocytes showed that L ‐trans‐ADC displayed the highest agonist potency at NR1/NR2D (EC50=50 μM ), which was 9.4‐, 3.4‐, and 1.9‐fold higher than the respective potencies at NR1/NR2A–C. D ‐cis‐ADC was shown to be a partial agonist at NR1/NR2C and NR1/NR2D with medium‐range micromolar potencies (EC50=720 and 230 μM , respectively). A subsequent in silico ligand–protein docking study suggested an unusual binding mode for these amino acids in the agonist binding site.  相似文献   

6.
BACKGROUND: γ‐Aminobutyric acid with several well‐known physiological functions is biosynthesized via the irreversible α‐decarboxylation of L ‐glutamate catalysed by glutamate decarboxylase (GAD). Although Streptococcus salivarius ssp. thermophilus has been widely applied to the dairy, the characterization of its GAD has not been reported. In this paper, the purification and the characterization of S. salivarius ssp. thermophilus GAD were investigated. RESULTS: GAD was purified 22‐fold from crude protein extracts with a yield of 7.8% in five steps. The final preparation gave a single band on SDS‐PAGE. The molecular weight of GAD determined by SDS‐PAGE and gel filtration was 46.9 kDa and 103.6 kDa, respectively, indicating that the enzyme exists as a dimmer of homological subunits. The optimum temperature and pH of GAD was 55 °C and pH 4.0, respectively. The enzyme reacted only with L ‐glutamate among 19 α‐amino acids with apparent Km at 2.3 mmol L?1 and did not react with D ‐glutamic acid. Activity of the enzyme could significantly be activated by 5 mmol L?1 of BaCl2 and inhibited by FeSO4, ZnSO4, CuSO4, MnSO4, Na2SO4, AgNO3, CoCl2, LiCl and KCl, respectively. The N‐terminal amino acid sequence of GAD was NH2‐MNEKLFREI. CONCLUSION: Both the characterization and the deduced amino sequence (ABI31651) showed the purified enzyme was a novel GAD. Copyright © 2008 Society of Chemical Industry  相似文献   

7.
The miscibility, crystallization behavior, and component interactions of two binary blends, poly(L ‐lactide) (L ‐PLA)/poly(vinylpyrrolidone) (PVP) and poly(D ,L ‐lactide) (DL ‐PLA)/PVP, were studied with differential scanning calorimetry and Fourier transform infrared (FTIR) spectroscopy. The composition‐dependent changes of the glass‐transition temperature (Tg) and degree of crystallinity (Xc) of the L ‐PLA phase indicated that L ‐PLA and PVP were immiscible over the composition range investigated. However, the sharp decrease of Xc with increasing PVP content in the second heating run demonstrated that the cold crystallization process of L ‐PLA was remarkably restricted by PVP. In DL ‐PLA/PVP blends, the existence of two series of isolated Tg's indicated that DL ‐PLA and PVP were phase‐separated, but evidence showed that there was some degree of interaction at the interface of the two phase, especially for the blends with low DL ‐PLA contents. FTIR measurements showed that there was no appreciable change in the spectra of L ‐PLA/PVP with respect to the coaddition of each component spectrum, implying the immiscibility of the two polymers. In contrast to L ‐PLA, the intermolecular interaction between DL ‐PLA and PVP was detected by FTIR; this was evidenced by the observation of a high‐frequency shift of the C?O stretching vibration band of PVP with increasing DL ‐PLA content, which suggested some degree of miscibility. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 88: 973–979, 2003  相似文献   

8.
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   

9.
Levoglucosan kinase (LGK) catalyzes the simultaneous hydrolysis and phosphorylation of levoglucosan (1,6‐anhydro‐β‐d ‐glucopyranose) in the presence of Mg2+–ATP. For the Lipomyces starkeyi LGK, we show here with real‐time in situ NMR spectroscopy at 10 °C and pH 7.0 that the enzymatic reaction proceeds with inversion of anomeric stereochemistry, resulting in the formation of α‐d ‐glucose‐6‐phosphate in a manner reminiscent of an inverting β‐glycoside hydrolase. Kinetic characterization revealed the Mg2+ concentration for optimum activity (20–50 mm ), the apparent binding of levoglucosan (Km=180 mm ) and ATP (Km=1.0 mm ), as well as the inhibition by ADP (Ki=0.45 mm ) and d ‐glucose‐6‐phosphate (IC50=56 mm ). The enzyme was highly specific for levoglucosan and exhibited weak ATPase activity in the absence of substrate. The equilibrium conversion of levoglucosan and ATP lay far on the product side, and no enzymatic back reaction from d ‐glucose‐6‐phosphate and ADP was observed under a broad range of conditions. 6‐Phospho‐α‐d ‐glucopyranosyl fluoride and 6‐phospho‐1,5‐anhydro‐2‐deoxy‐d ‐arabino‐hex‐1‐enitol (6‐phospho‐d ‐glucal) were synthesized as probes for the enzymatic mechanism but proved inactive with the enzyme in the presence of ADP. The pyranose ring flip 4C11C4 required for 1,6‐anhydro‐product synthesis from d ‐glucose‐6‐phosphate probably presents a major thermodynamic restriction to the back reaction of the enzyme.  相似文献   

10.
Polysulfones bearing a derivative of alanyl residue employed as chiral selectors were prepared by polymer modification. The specific rotation ([α]D) of the polysulfone with a derivative of D ‐alanyl residue (PSf‐Ac‐D ‐Ala) was determined to be 2.87 deg · cm2 · g?1 (c = 1.00 g · dL?1 in DMF) and that with L ‐alanyl residue (PSf‐Ac‐L ‐Ala) to be ‐2.36 deg · cm2 · g?1 (c = 1.00 g · dL?1 in DMF). The membrane from PSf‐Ac‐D ‐Ala preferentially adsorbed the D ‐isomer of Glu from racemic mixture of Glu and vice versa. Chiral separation ability was studied by applying a potential difference as a driving force for membrane transport. The permselectivity of PSf‐Ac‐D ‐Ala toward D ‐Glu (αD/L) was determined to be 1.40, and that of PSf‐Ac‐L ‐Ala toward the L ‐isomer (αL/D) to be 1.48 at 18.0 V, reflecting their adsorption selectivity.

  相似文献   


11.
Poly‐L ‐lactide (PLLA) is being widely considered for repair of damaged tissues, for controlled antibiotic release, and also as scaffolds for cultured cells. PLLA was blended with the lactide monomer in its two enantiomeric forms: D ‐lactide (D ‐la) and L ‐lactide (L ‐la) and with the cyclic dimmer D ,L ‐la, in order to enhance its flexibility and thereby overcome its inherent problem of brittleness. In this work, the crystallization, phase structure, and tensile properties of PLLA and PLLA plasticized with 5, 10, 15, and 20 wt% of D ‐la, L ‐la, and D ,L ‐la are explored. The three plasticizers used were effective in lowering the glass transition temperature (Tg) and the melting temperature (Tm) of PLLA, around 20°C for a plasticizer content of 20 wt%. The tensile strength and modulus of the blends decreased following the increasing content of plasticizers from approximately 58 MPa to values below 20 MPa, and from 1667 to 200 MPa, respectively. Aging the blends at storage ambient temperature revealed that the enhanced flexibility as well as the morphological stability was lost over time due to the migration of the plasticizer to the surface, this being less marked in the case of D ‐la as a result of interactions between the polymer and its enantiomeric monomer of complementary configuration. POLYM. ENG. SCI., 53:2073–2080, 2013. © 2013 Society of Plastics Engineers  相似文献   

12.
Transketolase (TK) from S. cerevisiae was successfully immobilized on layered double hydroxides (LDH) using simple, affordable and efficient adsorption and coprecipitation based immobilization procedures. Optimization of the preparation was performed using zinc aluminium nitrate (Zn2Al‐NO3) and magnesium aluminium nitrate (Mg2Al‐NO3) LDH as immobilization supports, and the protein‐to‐LDH weight ratio (Q) was varied. The highest immobilization yields (98–99%) and highest relative specific activities (4.2–4.4 U⋅mg−1 for the immobilized enzyme compared to 4.5 U⋅mg−1 for the free enzyme) were both achieved when using a protein‐to‐LDH weight ratio (Q) of 0.38. Efficient lyophilization of the LDH‐TK bionanocomposites thus synthesized was proven to allow easy use and storage of the supported TK with no significant loss of activity over a three‐month period. The kinetic parameters of the LDH‐TK enzyme were comparable to those of the free TK. The LDH‐TK enzyme was finally tested for the synthesis of L ‐erythrulose starting from hydroxypyruvate lithium salt (Li‐HPA) and glycolaldehyde (GA) as substrates. L ‐erythrulose was characterized and obtained with an isolated yield of 56% similar to that obtained with free TK. The reusability of the LDH‐TK biohybrid material was then investigated, and we found no loss of enzymatic activity over six cycles.  相似文献   

13.
The mycobacterial cell wall is a complex architecture, which has, as its major structural component, a lipidated polysaccharide covalently bound to peptidoglycan. This structure, termed the mycolyl–arabinogalactan–peptidoglycan complex, possesses a core galactan moiety composed of approximately 30 galactofuranosyl (Galf) resides attached via alternating β‐(1→6) and β‐(1→5) linkages. Recent studies have shown that the entire galactan is synthesized by the action of only two bifunctional galactofuranosyltransferases, GlfT1 and GlfT2. We report here saturation‐transfer difference (STD) NMR spectroscopy studies with GlfT2 using two trisaccharide acceptor substrates, β‐D ‐Galf‐(1→6)‐β‐D ‐Galf‐(1→5)‐β‐D ‐Galf‐O(CH2)7CH3 ( 2 ) and β‐D ‐Galf‐(1→5)‐β‐D ‐Galf‐(1→6)‐β‐D ‐Galf‐O(CH2)7CH3 ( 3 ), as well as the donor substrate for the enzyme, UDP‐Galf. Competition STD‐NMR titration experiments and saturation transfer double difference (STDD) experiments with 2 and 3 were undertaken to explore the bifunctionality of this enzyme, in particular to answer whether one or two active sites are responsible for the formation of both β‐(1→5)‐ and β‐(1→6)‐Galf linkages. It was demonstrated that 2 and 3 bind competitively at the same site; this suggests that GlfT2 has one active site pocket capable of catalyzing both β‐(1→5) and β‐(1→6) galactofuranosyl transfer reactions. The addition of UDP‐Galf to GlfT2 in the presence of either 2 or 3 generated a tetrasaccharide product; this indicates that the enzyme was catalytically active under the conditions at which the STD‐NMR experiments were carried out.  相似文献   

14.
Here we have characterized the first transketolase (TK) from a thermophilic microorganism, Geobacillus stearothermophilus, which was expressed from a synthetic gene in Escherichia coli. The G. stearothermophilus TK (mTKgst) retained 100% activity for one week at 50 °C and for 3 days at 65 °C, and has an optimum temperature range around 60–70 °C, which will be useful for preparative applications and for future biocatalyst development. The thermostability of the mTKgst allowed us to carry out an easy, one‐step purification by heat shock treatment of crude cell extracts at 65 °C for 45 min, directly yielding 132 mg of pure mTKgst from 1 L of culture. The reaction rate of mTKgst with glycolaldehyde was 14 times higher at 70 °C than at 20 °C, and 4 times higher at 50 °C when compared to E. coli TK under identical conditions. When tested at 50 °C with other aldehydes as acceptors, mTKgst activity was approximately 3 times higher than those obtained at 20 °C. Applications of this new TK in biocatalysis were performed with hydroxypyruvate as donor and three different aldehydes as acceptors – glycolaldehyde, D ‐glyceraldehyde and butyraldehyde – from which the corresponding products L ‐erythrulose 1 , D ‐xylulose 2 and 1,3‐dihydroxyhexan‐2‐one 3 were obtained, respectively. The optical rotations for products 1 and 2 indicate that the stereospecificity of mTKgst is identical to that of other TK sources, leading to a (3S) configuration. With the non‐hydroxylated substrate, butanal, the ee value was 85% (3S), showing higher enantioselectivity than the E. coli TK (75% ee, 3S). Processes at elevated temperatures could offer opportunities to extend the applications of TK biocatalysis, by favoring hydrophobic aldehyde acceptor substrate solubility and tolerance towards non‐conventional media.  相似文献   

15.
Ring‐opening polymerization of L ‐ or D ‐lactide was realized at 140 °C for a period of 7 days in the presence of dihydroxyl poly(ethylene glycol) (PEG), with M?n = 4000 g mol?1, using zinc lactate as initiator. The resulting poly(L ‐lactide)–PEG–poly(L ‐lactide) and poly(D ‐lactide)–PEG–poly(D ‐lactide) triblock copolymers are water soluble with polylactide (PLA) block length ranging from 11 to 17 units. Both the tube inverting method and rheological measurements were used to evaluate the gelation properties of aqueous solutions containing single copolymers or L /D copolymer pairs. Stereocomplexation between poly(L ‐lactide) and poly(D ‐lactide) blocks is observed for mixed solutions. Hydrogel formation is detected in the case of relatively long PLA blocks (DP PLA = 17), but not for copolymers with shorter PLA blocks (DP PLA = 11–13) due to partial racemization of L ‐lactyl units. Racemization is largely reduced when the reaction time is shortened to 1 day. Under these conditions, DP PLA of 8 is sufficient for the stereocomplexation of PLA–PEG block copolymers, and DP PLA above 10 leads to the formation of hydrogels of PLA–PEG block copolymers. On the other hand, racemization appears as a general phenomenon in the (co)polymerization of L ‐lactide with Zn(Lac)2 as initiator, although it is negligible or undetectable in the case of high molar mass polymers. Therefore, racemization is the limiting factor for the stereocomplexation‐induced gelation of water‐soluble PLA–PEG block copolymers where the PLA block length generally ranges from 10 to 30. Reaction conditions including initiator, time and temperature should be strictly controlled to minimize racemization. Copyright © 2010 Society of Chemical Industry  相似文献   

16.
The natural silk sericin, recovered from Bombyx mori silk waste by degumming and degrading, is a water‐soluble peptide with different molecular masses, ranging from 20 to 60 kDa. It is composed of 15 sorts of amino acids, among which the polar amino acids with hydroxyl, carboxyl and amino groups such as aspartic acid, serine and lysine account for 72%. The covalent attachment of the silk sericin peptides to L ‐asparaginase (ASNase) produces silk sericin peptides–L ‐asparaginase (SS–ASNase) bioconjugates that are active, stable, have a lower immune response, and have extended half‐lives in vitro in human serum. The modified enzyme coupled with sericin protein retains 55.8% of the original activity of the native enzyme. The optimal pH of SS–ASNase derivatives shifts considerably, to 5.0 in comparison with pH 6.0–8.0 of the native form. The thermostability and resistance to trypsin digestion of the modified enzyme are greatly enhanced as compared with ASNase alone. The Michaelis constant (Km) of SS–ASNase is 65 times lower than that of the enzyme alone. This suggests that the affinity of the enzyme to its substrate L ‐asparagine greatly increases when bioconjugated with silk sericin. The in vivo experiments also show that the silk sericin peptides have no immunogenicity, and the antigenicity of the enzyme is obviously decreased when coupled covalently with the silk sericin peptides. Copyright © 2005 Society of Chemical Industry  相似文献   

17.
Amorphous‐made poly(L ‐lactide) [i.e., poly(L ‐lactic acid) (PLLA)], poly(L ‐lactide‐co‐D ‐lactide)[P(LLA‐DLA)](77/23), and P(LLA‐DLA)(50/50) films and PLLA films with different crystallinity (Xc) values were prepared, and the effects of molecular weight, D ‐lactide unit content (tacticity and optical purity), and crystallinity of poly(lactide) [i.e., poly(lactic acid) (PLA)] on the water vapor permeability was investigated. The changes in number‐average molecular weight (Mn) of PLLA films in the range of 9 × 104–5 × 105 g mol?1 and D ‐lactide unit content of PLA films in the range of 0–50% have insignificant effects on their water vapor transmission rate (WVTR). In contrast, the WVTR of PLLA films decreased monotonically with increasing Xc from 0 to 20%, while leveled off for Xc exceeding 30%. This is probably due to the higher resistance of “restricted” amorphous regions to water vapor permeation compared with that of the “free” amorphous regions. The free and restricted amorphous regions are major amorphous components of PLLA films for Xc ranges of 0–20% and exceeding 30%, respectively, resulting in the aforementioned dependence of WVTR on Xc. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci, 2006  相似文献   

18.
The chiral pyrrolidine‐substituted imidazolium cetyl‐PEG10‐sulfate (D ‐ProMe) derived from D ‐proline worked as an excellent activating agent of Burkholderia cepacia lipase; it is particularly interesting that the D ‐isomer of the imidazolium salt worked better than the L ‐isomer. This suggests that the imidazolium cation group directly interacts with the enzyme protein and causes preferable modification of the reactivity.  相似文献   

19.
The transketolase from Geobacillus stearothermophilus (TKGst) is a thermostable enzyme with notable high activity and stability at elevated temperatures, but it accepts non‐α‐hydroxylated aldehydes only with low efficiency. Here we report a protein engineering study of TKGst based on double‐site saturation mutagenesis either at Leu191 or at Phe435 in combination with Asp470; these are the residues responsible for substrate binding in the active site. Screening of the mutagenesis libraries resulted in several positive variants with activity towards propanal up to 7.4 times higher than that of the wild type. Variants F435L/D470E and L191V/D470I exhibited improved (73 % ee, 3S) and inverted (74 % ee, 3R) stereoselectivity, respectively, for propanal. L191V, L382F/E, F435L, and D470/D470I were concluded to be positive mutations at Leu191, Leu382, Phe435, and Asp470 both for activity and for stereoselectivity improvement. These results should benefit further engineering of TKGst for various applications in asymmetric carboligation.  相似文献   

20.
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   

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