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 unique five‐membered aminocyclitol core of the antitumor antibiotic pactamycin originates from d ‐glucose, so unprecedented enzymatic modifications of the sugar intermediate are involved in the biosynthesis. However, the order of the modification reactions remains elusive. Herein, we examined the timing of introduction of an amino group into certain sugar‐derived intermediates by using recombinant enzymes that were encoded in the pactamycin biosynthesis gene cluster. We found that the NAD+‐dependent alcohol dehydrogenase PctP and pyridoxal 5′‐phosphate dependent aminotransferase PctC converted N‐acetyl‐d ‐glucosaminyl‐3‐aminoacetophonone into 3′‐amino‐3′‐deoxy‐N‐acetyl‐d ‐glucosaminyl‐3‐aminoacetophenone. Further, N‐acetyl‐d ‐glucosaminyl‐3‐aminophenyl‐β‐oxopropanoic acid ethyl ester was converted into the corresponding 3′‐amino derivative. However, PctP did not oxidize most of the tested d ‐glucose derivatives, including UDP‐GlcNAc. Thus, modification of the GlcNAc moiety in pactamycin biosynthesis appears to occur after the glycosylation of aniline derivatives.  相似文献   

2.
In Streptomyces albogriseolus the indolethiophen alkaloid thienodolin is derived from tryptophan. The first step in thienodolin biosynthesis is the regioselective chlorination of tryptophan in the 6‐position of the indole ring. The second step is catalyzed by the aminotransferase ThdN. ThdN shows sequence homology (up to 69 % similarity) with known pyridoxal 5′‐phosphate‐dependent aminotransferases of the aspartate aminotransferase family from Gram‐positive bacteria. thdN was heterologously expressed in Pseudomonas fluorescens, and the enzyme was purified by nickel‐affinity chromatography. ThdN is a homodimeric enzyme with a mass of 90 600 kDa and catalyzes the conversion of l ‐tryptophan and a number of chlorinated and brominated l ‐tryptophans. The lowest KM values were found for 6‐bromo‐ and 6‐chlorotryptophan (40 and 66 μm , respectively). For l ‐tryptophan it was 454 μm, which explains why thienodolin is the major product and dechlorothienodolin is only a minor component. The turnover number (kcat) for 7‐chlorotryptophan (128 min?1) was higher than that for the natural substrate 6‐chlorotryptophan (88 min?1).  相似文献   

3.
The present work assesses in vitro the role of human Stanniocalcin 1 (hSTC‐1) in glucose metabolism in white retroperitoneal adipose tissue (WRAT) from fed rat. In the fed state, hSTC1 increases the incorporation of 14C from glucose into lipids in the rat WRAT. The increase in lipogenesis capacity supports the hypothesis that the activity of the glycerol‐3‐phosphate‐generating pathway (glycolysis) from glucose is regulated by hSTC‐1. The effect of hSTC‐1 on de novo fatty acid synthesis and on glucose oxidation in WRAT is supported by an 85 % increase in 14CO2 production from 14C‐glucose. The incubation of WRAT in the presence of hSTC‐1 maintained the ADP/ATP ratio close to the control group. The presence of hSTC‐1 in the incubation medium did not inhibit the lipolytic effect of epinephrine. In conclusion, hSTC‐1 is one of the hormonal factors that control glucose metabolism in WRAT in the fed state.  相似文献   

4.
Despite the unsurpassed selectivity that enzymes usually offer, biocatalytic transformations repeatedly fall short of the robustness and process efficiency demanded for production‐scale chemical synthesis. Nucleotide sugar‐dependent “Leloir” glycosyltransferases (GTs) are fine catalysts of glycosylation but there is concern as to whether reactions from this enzyme class are fit for industrial process development. We demonstrate in this study of sucrose synthase (SuSy; EC 2.4.1.13) that, in order to unlock the synthetic potential of the GT reaction, it was vital to combine a focused, kinetic characteristics‐based enzyme selection with a reaction design properly aligned to thermodynamic constraints. The equilibrium constant (Keq) for the conversion of sucrose and uridine 5′‐diphosphate (UDP) into the target product UDP‐α‐d ‐glucose and d ‐fructose decreased with increasing pH due to deprotonation of the β‐phosphate group of UDP above the pKa of ∼6.0. Proton uptake coupled to the glucosyl transfer made it essential that the pH was carefully controlled throughout the reaction. Comparing two SuSys from Acidithiobacillus caldus and Glycine max (soybean), substrate inhibition by UDP superseded catalytic efficiency as the prior selection criterion, demanding choice of the bacterial GT for use at high UDP concentrations. Reaction at the operational pH optimum, determined as 5.0, gave 255 mM (144 g L−1) of UDP‐glucose in 85% yield from UDP. Using an enzyme concentration of only 0.1 g L−1, a space‐time yield of 25 g L−1 h−1 was obtained. The mass‐based turnover number (g product formed per g enzyme added) reached a value of 1440 from a single batch conversion. Therefore, these parameters of the UDP‐glucose synthesis show that the reaction of a GT can be pushed to a process efficiency typically required for implementation in fine chemicals production.

  相似文献   


5.
Glucose oxidase (EC 1.1.3.4) was encapsulated in liposomes (prepared from phosphatidyl choline and cholesterol) by the dehydration–rehydration method. The enzymatic activities of native and liposomal glucose oxidase were followed by the amount of H2O2 obtained in the enzymatic β‐D ‐glucose oxidation. Some characteristics of the liposomal and free glucose oxidase were compared. The enzyme encapsulated in liposomes showed an apparent inhibition by glucose at concentrations higher than 0.28 mol dm?3 with a substrate inhibition constant of 0.95 ± 0.12 mol dm?3. The enzyme entrapped showed an apparent Km value higher than that of the free enzyme. The apparent Vmax of liposomal enzyme decreased by a factor of 0.35 with respect of that of the native enzyme. The optimum temperature of the free and entrapped enzymes remained similar but the liposomal enzyme showed maximal activity at a more acid pH (5.2). The thermal and proteolytic stabilities were enhanced by encapsulation in liposomes. The stabilization factors (relationship between half‐lives of entrapped form and free enzyme) at 45, 50 and 55 °C for liposomal glucose oxidase were 2.6, 1.6 and 1.6, respectively. Copyright © 2003 Society of Chemical Industry  相似文献   

6.
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   

7.
Nathan L. Ta  Thomas N. Seyfried 《Lipids》2015,50(12):1167-1184
Glucose and glutamine are essential energy metabolites for brain tumor growth and survival under both normoxic and hypoxic conditions. Both metabolites can contribute their carbons to lipid biosynthesis. We used uniformly labeled [14C]‐U‐d ‐glucose and [14C]‐U‐l ‐glutamine to examine the profile of de novo lipid biosynthesis in the VM‐M3 murine glioblastoma cells. The major lipids synthesized included phosphatidylcholine (PtdCho), phosphatidylethanolamine (EtnGpl), phosphatidylinositol (PtdIns), phosphatidylserine (PtdSer), sphingomyelin (CerP Cho), bis(monoacylglycero)phosphate (BMP)/phosphatidic acid (PtdOH), cholesterol (C), cardiolipin (Ptd2Gro), and gangliosides. Endogenous lipid synthesis, using either glucose or glutamine, was greater in media without fetal bovine serum (FBS) than in media containing 10 % FBS under normoxia. De novo lipid synthesis was greater using glucose carbons than glutamine carbons under normoxia. The reverse was observed for most lipids under hypoxia suggesting an attenuation of glucose entering the TCA cycle. Lactate was produced largely from glucose carbons with minimal lactate derived from glutamine under either normoxia or hypoxia. Accumulation of triacylglycerols (TAG), containing mostly saturated and mono‐unsaturated fatty acids, was observed under hypoxia using carbons from either glucose or glutamine. The data show that the incorporation of labeled glucose and glutamine into most synthesized lipids was dependent on the type of growth environment, and that the VM‐M3 glioblastoma cells could acquire lipids, especially cholesterol, from the external environment for growth and proliferation.  相似文献   

8.
Biomimetic synthesis routes towards the important natural d ‐mannosyl donor guanosine 5′‐diphospho‐d ‐mannose (GDP‐Man) rely on kinase‐catalyzed nucleotide triphosphate (NTP)‐dependent phosphorylations of d ‐mannose (Man), to give d ‐mannose 6‐phosphate or α‐d ‐mannose 1‐phosphate (αMan 1‐P) as an intermediate product. A GDP‐Man synthesis not requiring the kinase/NTP system would be practical and cost‐effective. Here, we have developed a multienzyme cascade towards GDP‐Man, characterized in that αMan 1‐P was obtained by a diastereoselective phosphatase‐catalyzed phosphorylation of Man. α‐d ‐Glucose 1‐phosphate (αGlc 1‐P), prepared in situ through phosphorylase‐catalyzed conversion of sucrose in the presence of inorganic phosphate, was used as an expedient phosphoryl donor. The incipient αMan 1‐P and guanosine triphosphate (GTP) were converted into GDP‐Man by a highly manno compared to gluco selective nucleotidyltransferase. Pyrophosphatase was additionally required to hydrolyze the pyrophosphate released from the GTP, thus driving the reaction towards GDP‐Man. The enzymatic cascade was operated with the αMan 1‐P and the GDP‐Man formation decoupled from one another (sequential mode) or having all steps run concurrently (simultaneous mode). Detailed time course analysis revealed that kinetic pull due to the constant removal of the intermediate αMan 1‐P in simultaneous‐mode reactions was important to promote phosphorylation of Man from αGlc 1‐P in high efficiency, avoiding loss of sugar 1‐phosphates by hydrolysis. Under optimized conditions for the one‐pot transformation involving four enzymes, 100 mM (67 g L−1) GDP‐Man was prepared from 140 mM sucrose and phosphate, using 400 mM Man as the phosphoryl acceptor. The product was recovered by anion‐exchange and size‐exclusion chromatography in ≥95% purity in about 50% yield (100 mg). These results demonstrate for the first time the practical use of a phosphorylase‐phosphatase combi‐catalyst as an alternative to the canonical kinase for the anomeric phosphorylation of the sugar substrate in nucleoside diphospho‐sugar synthesis. Phosphorylation from inorganic phosphate via the intermediate αGlc 1‐P rather than from NTP, particularly GTP, appears advantageous specifically in cases where the sugar acceptor is a bulk commodity that can be applied in suitable excess to the phosphatase reaction.

  相似文献   


9.
The enzymatic epimerization of uridine 5′‐diphospho‐α‐D ‐glucose (UDP‐Glc, 1 ) and uridine 5′‐diphospho‐N‐acetyl‐α‐D ‐glucosamine (UDP‐GlcNAc, 2 ) and the subsequent oxidation of uridine 5′‐diphospho‐α‐D ‐galactose (UDP‐Gal, 3 ) and uridine 5′‐diphospho‐N‐acetyl‐α‐D ‐galactosamine (UDP‐GalNAc, 4 ) were combined with chemical biotinylation with biotin‐ε‐amidocaproylhydrazide in a one‐pot synthesis. Analysis by CE and NMR revealed a mixture (1.0:1.4) of the biotinylated nucleotide sugars uridine 5′‐diphospho‐6‐biotin‐ε‐amidocaproylhydrazino‐α‐D ‐galactose (UDP‐6‐biotinyl‐Gal, 7) and uridine 5′‐diphospho‐6‐biotin‐ε‐amidocaproylhydrazino‐α‐D ‐glucose (UDP‐6‐biotinyl‐Glc, 9 ), respectively, in a reaction started with 1 . One product, uridine 5′‐diphospho‐6‐biotin‐ε‐amidocaproylhydrazino‐N‐acetyl‐α‐D ‐galactosamine (UDP‐6‐biotinyl‐GalNAc, 8) was formed when the reaction was initiated with 2 . It could be demonstrated for the first time that a UDP‐Glc(NAc) 4′‐epimerase (Gne from Campylobacter jejuni) and galactose oxidase from Dactylium dendroides can be used simultaneously in enzymatic catalysis. This is of particular interest since the coaction of an enzyme demanding reductive conditions and an oxygen‐dependent oxidase is unexpected.  相似文献   

10.
Novel indolocarbazole derivative 12‐(α‐L ‐arabinopyranosyl)indolo[2,3‐α]pyrrolo[3,4‐c]carbazole‐5,7‐dione (AIC) demonstrated high potency (at submicromolar concentrations) against the NCI panel of human tumor cell lines and transplanted tumors in vivo. In search of tentative targets for AIC, we found that the drug formed high affinity intercalative complexes with d(AT)20, d(GC)20 and calf thymus DNA (binding constants (1.6×106) M ?1Ka≤(3.3×106) M ?1). The drug intercalated preferentially into GC pairs of the duplex. Importantly, the concentrations at which AIC formed the intercalative complexes with DNA (C≤1 μM ) were identical to the concentrations that triggered p53‐dependent gene reporter transactivation, the replication block, the inhibition of topoisomerase I‐mediated DNA relaxation and death of HCT116 human colon carcinoma cells. We conclude that the formation of high affinity intercalative complexes with DNA is an important factor for anticancer efficacy of AIC.  相似文献   

11.
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   

12.
Environmental and enzymatic degradations were investigated on a series of copolycarbonates consisting of equimolar amounts of 1,4 : 3,6‐dianhydrohexitols (1,4 : 3,6‐dianhydro‐D ‐glucitol (1a) and 1,4 : 3,6‐dianhydro‐D ‐mannitol (1b)) and alkylene diols (1,4‐butanediol, 1,6‐hexanediol, 1,8‐octanediol, and 1,10‐decanediol) or oligo(ethylene glycol)s (di‐, tri‐, and tetraethylene glycols). Fourteen different copolycarbonates with number average molecular weights in the range of 1.1–4.2 × 104 were prepared by solution polycondensation as described in our previous article. Biodegradability of the copolycarbonates was assessed by soil burial degradation tests in composted soil at 27 °C and by enzymatic degradation tests in a phosphate buffer solution at 37 °C. In general, biodegradability of the copolycarbonates increased with increasing chain lengths of the methylene groups of alkylene diols or of the oxyethylene groups of the oligo(ethylene glycol)s. SEM observations of the film surfaces of polymers recovered from soil burial indicated that the copolycarbonates were degraded by microorganisms in soil. In enzymatic degradation, the copolycarbonates containing alkylene diol components showed high degradability with Pseudomonas sp. lipase, whereas the copolycarbonates containing oligo(ethylene glycol) components were not degraded at all. The enzymatic degradability of the copolycarbonates is discussed with reference to the geometrical structure around the carbonate linkages and the microstructure and hydrophobicity of the polymer chains. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 98: 1679–1687, 2005  相似文献   

13.
Lipase‐catalyzed esterification and properties of synthesized carbohydrate esters were investigated. Methyl α‐d ‐glucopyranoside was the acyl group acceptor and different carbon atom chain lengths of aliphatic carboxylic acids (C12, C14 and C16) as the acyl group donors were applied in the esterification. Physico‐chemical studies on the synthesized carbohydrate esters were carried out. It was found that melting point for the methyl 6‐O‐hexadecanoyl‐α‐d ‐glucopyranoside was the highest consecutively followed by methyl 6‐O‐tetradecanoyl‐α‐d ‐glucopyranoside and methyl 6‐O‐dodecanoyl‐α‐d ‐glucopyranoside. Liquid crystal properties of the synthesized carbohydrate ester synthesized were evaluated via optical polarized microscopy. It was found that the liquid crystal textures for mono‐substituted carbohydrate esters were of the smectic phase. In a quaternary system (carbohydrate ester/n‐butanol/n‐hexadecane/water), a maximum 34 % of water (by mass) was contained in the monophasic region of methyl 6‐O‐tetradecanoyl‐α‐d ‐glucopyranoside and a maximum of 52 % water (by mass) was contained in a monophasic methyl 6‐O‐dodecanoyl‐α‐d ‐glucopyranoside. For methyl‐6‐O‐dodecanoyl‐α‐d ‐glucopyranoside, its concentration at aggregation was 5.2 × 10?4 mM, with minimum air/water surface tension of 26 mN m?1. The Gibbs energy of micellization was calculated at ?50 kJ mol?1. The maximum adsorption density of methyl 6‐O‐dodecanoyl‐α‐d ‐glucopyranoside was determined at 4 × 10?6 mol m?2 while its minimum area per surfactant molecule at the air/water surface was 47 Å2.  相似文献   

14.
Stereo diblock polylactides (SDB‐PLAs) composed of relatively short poly(d ‐lactide) (PDLA) segments and relatively long poly(l ‐lactide) (PLLA) segments were synthesized to have a wide number‐average molecular weight (Mn) range of 2.5 × 104–2.0 × 105 g mol?1 and d ‐lactyl unit content of 0.9–38.6%. The effects of incorporated short PDLA segments (Mn = 2.0 × 103–7.7 × 103 g mol?1) on crystallization behavior of the SDB‐PLAs were first investigated during heating after complete melting and quenching or during slow cooling after complete melting. Stereocomplex (SC) crystallites can be formed at d ‐lactyl unit content as low as 4.3 and 5.8% for heating and slow cooling, respectively, and for Mn of PDLA segments as low as 2.0 × 103 and 3.5 × 103 g mol?1, respectively. With decreasing Mn and increasing d ‐lactyl unit content, the cold crystallization temperature during heating decreased and the crystallization temperature during slow cooling increased. With increasing d ‐lactyl unit content, the melting enthalpy (ΔHm) of SC crystallites during heating and the crystallinity (Xc) of SC crystallites after slow cooling increased, whereas ΔHm of PLLA homo‐crystallites during heating and Xc of PLLA homo‐crystallites after slow cooling decreased. The total ΔHm of SC crystallites and PLLA homo‐crystallites during heating and the total Xc after slow cooling became a minimum at d ‐lactyl unit content of 10–15% and gave a maximum at d ‐lactyl unit content of 0%. Despite the accelerated crystallization of some of SDB‐PLAs, the low values of total ΔHm and Xc at d ‐lactyl unit content of 10–15% are attributable to the formation of two crystalline species of SC crystallites and PLLA homo‐crystallites.  相似文献   

15.
Attempts were carried out to enhance the surface hydrophilicity of poly(L ‐lactide), that is, poly(L ‐lactic acid) (PLLA) film, utilizing enzymatic, alkaline, and autocatalytic hydrolyses in a proteinase K/Tris–HCL buffered solution system (37°C), in a 0.01N NaOH solution (37°C), and in a phosphate‐buffered solution (100°C), respectively. Moreover, its chain‐scission mechanisms in these different media were studied. The advancing contact‐angle (θa) value of the amorphous‐made PLLA film decreased monotonically with the hydrolysis time from 100° to 75° and 80° without a significant molecular weight decrease, when enzymatic and alkaline hydrolyses were continued for 60 min and 8 h, respectively. In contrast, a negligible change in the θa value was observed for the PLLA films even after the autocatalytic hydrolysis was continured for 16 h, when their bulk Mn decreased from 1.2 × 105 to 2.2 × 104 g mol?1 or the number of hydrophilic terminal groups per unit weight increased from 1.7 × 10?5 to 9.1 × 10?5 mol g?1. These findings, together with the result of gravimetry, revealed that the enzymatic and alkaline hydrolyses are powerful enough to enhance the practical surface hydrophilicity of the PLLA films because of their surface‐erosion mechanisms and that its practical surface hydrophilicity is controllable by varying the hydrolysis time. Moreover, autocatalytic hydrolysis is inappropriate to enhance the surface hydrophilicity, because of its bulk‐erosion mechanism. Alkaline hydrolysis is the best to enhance the hydrophilicity of the PLLA films without hydrolysis of the film cores, while the enzymatic hydrolysis is appropriate and inappropriate to enhance the surface hydrophilicity of bulky and thin PLLA materials, respectively, because a significant weight loss occurs before saturation of θa value. The changes in the weight loss and θa values during hydrolysis showed that exo chain scission as well as endo chain scission occurs in the presence of proteinase K, while in the alkaline and phosphate‐buffered solutions, hydrolysis proceeds via endo chain scission. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 87: 1628–1633, 2003  相似文献   

16.
Chiral amino acids are important intermediates for the pharmaceutical industry. We have developed a novel one‐pot enzymatic method for D ‐amino acid synthesis by the dynamic kinetic resolution of N‐succinyl‐dl ‐amino acids using D ‐succinylase (DSA) and N‐succinylamino acid racemase (NSAR, EC 4.2.1.113). The DSA from Cupriavidus sp. P4‐10‐C, which hydrolyzes N‐succinyl‐D ‐amino acids enantioselectively to their corresponding D ‐amino acids, was identified for the first time by screening soil microorganisms. Subsequently, the DSA gene was cloned and overexpressed in Escherichia coli. DSA was shown to comprise two subunits with molecular masses of 26 kDa and 60 kDa. Additionally, the NSAR gene from Geobacillus stearothermphilus NCA1503, which racemizes N‐succinylamino acids, was also cloned and overexpressed in E. coli. The highly purified DSA and NSAR prepared from each recombinant E. coli were characterized and used for D ‐amino acid synthesis. A one‐pot enzymatic method converted 100 mM N‐succinyl‐dl ‐phenylalanine to D ‐phenylalanine in 91.1% conversion with 86.7% ee. This novel enzymatic method may be useful for the industrial production of many D ‐amino acids.

  相似文献   


17.
Production of L ‐methionine by immobilized pellets of Aspergillus oryzae in a packed bed reactor was investigated. Based on the determination of relative enzymatic activity in the immobilized pellets, the optimum pH and temperature for the resolution reaction were 8.0 and 60 °C, respectively. The effects of substrate concentration on the resolution reaction were also investigated and the kinetic constants (Km and Vm) of immobilized pellets were found to be 7.99 mmol dm?3 and 1.38 mmol dm?3 h?1, respectively. The maximum substrate concentration for the resolution reaction without inhibition was 0.2 mol dm?3. The L ‐methionine conversion rate reached 94% and 78% when substrate concentrations were 0.2 and 0.4 mol dm?3, respectively, at a flow rate of 7.5 cm3 h?1 using the small‐scale packed bed reactor developed. The half‐life of the L ‐aminoacylase in immobilized pellets was 70 days in continuous operation. All the results obtained in this paper exhibit a practical potential of using immobilized pellets of Aspergillus oryzae in the production of L ‐methionine. © 2002 Society of Chemical Industry  相似文献   

18.
Two putative prenyltransferase genes, SAML0654 and Strvi8510, were identified in Streptomyces ambofaciens and Streptomyces violaceusniger, respectively. Their deduced products share 63 % sequence identity. Biochemical investigations with recombinant proteins demonstrated that L ‐tryptophan and derivatives, including D ‐tryptophan, 4‐, 5‐, 6‐ and 7‐methyl‐dl ‐tryptophan, were well accepted by both enzymes in the presence of DMAPP. Structural elucidation of the isolated products revealed regiospecific prenylation at C‐6 of the indole ring and proved unequivocally the identification of two very similar 6‐dimethylallyltryptophan synthases (6‐DMATS). Detailed biochemical investigations with SAML0654 proved L ‐tryptophan to be the best substrate (Km 18 μm, turnover 0.3 s?1). Incubation with different prenyl donors showed that they also accepted GPP and catalyzed the same specific prenylation. Utilizing GPP as a prenyl donor has not been reported for tryptophan prenyltransferases previously. Both enzymes also catalyzed prenylation of some hydroxynaphthalenes; this has not previously been described for bacterial indole prenyltransferases. Interestingly, SAML0654 transferred prenyl moieties onto the unsubstituted ring of hydroxynaphthalenes.  相似文献   

19.
Novel carbohydrate‐based non‐ionic gemini surfactants consisting of two sugar head groups, two hydrophobic tails having chain lengths of C12, C14, and C16 and a flexible –(CH2)6– spacer were synthesized and investigated for their reverse micellar encapsulation properties. The head groups of the geminis comprise glucose entities (with reducing function blocked in a cyclic acetal group) connected through C‐6 to tertiary amines. These surfactants were explored for reverse micellar encapsulation of d ‐ and l ‐enantiomers of aromatic α‐amino acids viz. histidine (His), phenylalanine (Phe), tyrosine (Tyr) and tryptophan (Trp) in neat n‐hexane. Similar studies were carried out for encapsulation of nucleobases viz. adenine (Ade), guanine (Gua), thymine (Thy), cytosine (Cyt) and Uracil (Ura). Reverse micellar studies revealed that aromatic α‐amino acids were encapsulated in the sequence His>Tyr>Phe>Trp. In most cases, a difference in the degree of encapsulation of d ‐ and l ‐enantiomers of aromatic amino acids in reverse micellar phases of gemini amphiphiles in neat n‐hexane, was revealed. For Tyr, l ‐enantiomer was better encapsulated than its antipode, i.e., d ‐enantiomer but for Trp, d ‐enantiomer was better encapsulated then l ‐enantiomer. In the case of nucleobases, Ura was found selectively encapsulated by reverse micelles formed by these new amphiphiles.  相似文献   

20.
The substrate scope of the flavoprotein alditol oxidase (AldO) from Streptomyces coelicolor A3(2), recombinantly produced in Escherichia coli, was explored. While it has been established that AldO efficiently oxidizes alditols to D ‐aldoses, this study revealed that the enzyme is also active with a broad range of aliphatic and aromatic alcohols. Alcohols containing hydroxy groups at the C‐1 and C‐2 positions like 1,2,4‐butanetriol (Km=170 mM, kcat=4.4 s−1), 1,2‐pentanediol (Km=52 mM, kcat=0.85 s−1) and 1,2‐hexanediol (Km=97 mM, kcat=2.0 s−1) were readily accepted by AldO. Furthermore, the enzyme was highly enantioselective for the oxidation of 1,2‐diols [e.g., for 1‐phenyl‐1,2‐ethanediol the (R)‐enantiomer was preferred with an E‐value of 74]. For several diols the oxidation products were determined by GC‐MS and NMR. Interestingly, for all tested 1,2‐diols the products were found to be the α‐hydroxy acids instead of the expected α‐hydroxy aldehydes. Incubation of (R)‐1‐phenyl‐1,2‐ethanediol with 18O‐labelled water (H218O) revealed that a second enzymatic oxidation step occurs via the hydrate product intermediate. The relaxed substrate specificity, excellent enantioselectivity, and independence of coenzymes make AldO an attractive enzyme for the preparation of optically pure 1,2‐diols and α‐hydroxy acids.  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