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.
Zhao  Feng-Jiao  Jin  Yun  Liu  Zhongchuan  Guo  Chao  Li  Tong-Biao  Li  Zi-Yi  Wang  Ganggang  Wu  Zhong-Liu 《Applied microbiology and biotechnology》2017,101(23):8395-8404

ChKRED20 is an efficient and robust anti-Prelog ketoreductase that can catalyze the reduction of ketones to chiral alcohols as pharmaceutical intermediates with great industrial potential. To overcome its limitation on the bioreduction of ortho-substituted acetophenone derivatives, the X-ray crystal structure of the apo-enzyme of ChKRED20 was determined at a resolution of 1.85 Å and applied to the molecular modeling and reshaping of the catalytic cavity via three rounds of iterative saturation mutagenesis together with alanine scanning and recombination. The mutant Mut3B was achieved with expanded catalytic scope that covered all the nine substrates tested as compared with two substrates for the wild type. It exhibited 13–20-fold elevated k cat/K m values relative to the wild type or to the first gain-of-activity mutant, while retaining excellent stereoselectivity toward seven of the substrates (98–> 99% ee). Another mutant 29G10 displayed complementary selectivity for eight of the ortho-substituted acetophenone derivatives, with six of them delivering excellent stereoselectivity (90–99% ee). Its k cat/K m value toward 1-(2-fluorophenyl)ethanone was 5.6-fold of the wild type. The application of Mut3B in elevated substrate concentrations of 50–100 g/l was demonstrated in 50-ml reactions, achieving 75–> 99% conversion and > 99% ee.

  相似文献   

2.
Niu  Chengtuo  Zhu  Linjiang  Xu  Xin  Li  Qi 《Applied microbiology and biotechnology》2017,101(3):1085-1097

Higher thermostability is required for 1,3-1,4-β-glucanase to maintain high activity under harsh conditions in the brewing and animal feed industries. In this study, a comprehensive and comparative analysis of thermostability in bacterial β-glucanases was conducted through a method named spatial compartmentalization of mutational hotspots (SCMH), which combined alignment of homologous protein sequences, spatial compartmentalization, and molecular dynamic (MD) simulation. The overall/local flexibility of six homologous β-glucanases was calculated by MD simulation and linearly fitted with enzyme optimal enzymatic temperatures. The calcium region was predicted to be the crucial region for thermostability of bacterial 1,3-1,4-β-glucanases, and optimization of four residue sites in this region by iterative saturation mutagenesis greatly increased the thermostability of a mesophilic β-glucanase (BglT) from Bacillus terquilensis. The E46P/S43E/H205P/S40E mutant showed a 20 °C increase in optimal enzymatic temperature and a 13.8 °C rise in protein melting temperature (T m) compared to wild-type BglT. Its half-life values at 60 and 70 °C were 3.86-fold and 7.13-fold higher than those of wild-type BglT. The specific activity of E46P/S43E/H205P/S40E mutant was increased by 64.4 %, while its stability under acidic environment was improved. The rational design strategy used in this study might be applied to improve the thermostability of other industrial enzymes.

  相似文献   

3.
Light-induced reaction dynamics of isolated photosynthetic membranes obtained from wild-type (WT) and reaction center (RC)-subunit deletion strains SPUHK1 (an H-subunit deletion mutant) and SKΔLM (an (L+M) deletion mutant) of the purple non-sulphur bacterium Rhodospirillum rubrum have been investigated by femtosecond transient absorption spectroscopy. Upon excitation of the spirilloxanthin (Spx) S2 state at 546 nm, of the bacteriochlorophyll Soret band at 388 nm and probing spectral regions, which are characteristic for carotenoids, similar dynamics in the SPUHK1, SKΔLM and WT strains could be observed. The excitation of Spx S2 is followed by the simultaneous population of the lower singlet excited states S1 and S* which decay with lifetimes of 1.4 and 5 ps, respectively for the mutants, and 1.4 and 4 ps, respectively, for the wild-type. The excitation of the BChl Soret band is followed by relaxation into BChl lower excited states which compete with excitation energy transfer BChl-to-Spx. The deexcitation pathway BChl(Soret) → Spx(S2) → Spx(S1) occurs with the same transition rate for all investigated samples (WT, SPUHK1 and SKΔLM). The kinetic traces measured for the Spx S1 → SN transition display similar behaviour for all samples showing a positive signal which increases within the first 400 fs (i.e. the time needed for the excitation energy to reach the Spx S1 excited state) and decays with a lifetime of about 1.5 ps. This suggests that the Spx excited state dynamics in the investigated complexes do not differ significantly. Moreover, a longer excited state lifetime of BChl for SPUHK1 in comparison to WT was observed, consistent with a photochemical quenching channel present in the presence of RC. For long delay times, photobleaching of the RC special pair and an electrochromic blue shift of the monomeric BChl a can be observed only for the WT but not for the mutants. The close similarity of the excited state decay processes of all strains indicates that the pigment geometry of the LH1 complex in native membranes is unaffected by the presence of an RC and allows us to draw a model representation of the WT, SKΔLM and SPUHK1 PSU complexes.  相似文献   

4.

Esterases comprise a group of enzymes that catalyze the cleavage and synthesis of ester bonds. They are important in biotechnological applications owing to their enantioselectivity, regioselectivity, broad substrate specificity, and the fact that they do not require cofactors. In a previous study, we isolated the esterase Est25 from a metagenomic library. Est25 showed catalytic activity toward the (R,S)-ketoprofen ethyl ester but had low enantioselectivity toward the (S)-ketoprofen ethyl ester. Because (S)-ketoprofen has stronger anti-inflammatory effects and fewer side effects than (R)-ketoprofen, enantioselectivity of this esterase is important. In this study, we generated Est25 mutants with improved enantioselectivity toward the (S)-ketoprofen ethyl ester; improved enantioselectivity of mutants was established by analysis of their crystal structures. The enantioselectivity of mutants was influenced by substitution of Phe72 and Leu255. Substituting these residues changed the size of the binding pocket and the entrance hole that leads to the active site. The enantioselectivity of Est25 (E = 1.1 ± 0.0) was improved in the mutants F72G (E = 1.9 ± 0.2), L255W (E = 16.1 ± 1.1), and F72G/L255W (E = 60.1 ± 0.5). Finally, characterization of Est25 mutants was performed by determining the optimum reaction conditions, thermostability, effect of additives, and substrate specificity after substituting Phe72 and Leu255.

  相似文献   

5.
Methylaspartate ammonia lyase (MAL; EC 4.3.1.2) catalyzes the reversible addition of ammonia to mesaconate to give (2S,3S)-3-methylaspartate and (2S,3R)-3-methylaspartate as products. MAL is of considerable biocatalytic interest because of its potential use for the asymmetric synthesis of substituted aspartic acids, which are important building blocks for synthetic enzymes, peptides, chemicals, and pharmaceuticals. Here, we have cloned the gene encoding MAL from the thermophilic bacterium Carboxydothermus hydrogenoformans Z-2901. The enzyme (named Ch-MAL) was overproduced in Escherichia coli and purified to homogeneity by immobilized metal affinity chromatography. Ch-MAL is a dimer in solution, consisting of two identical subunits (∼49 kDa each), and requires Mg2+ and K+ ions for maximum activity. The optimum pH and temperature for the deamination of (2S,3S)-3-methylaspartic acid are 9.0 and 70°C (k cat = 78 s−1 and K m = 16 mM). Heat inactivation assays showed that Ch-MAL is stable at 50°C for >4 h, which is the highest thermal stability observed among known MALs. Ch-MAL accepts fumarate, mesaconate, ethylfumarate, and propylfumarate as substrates in the ammonia addition reaction. The enzyme also processes methylamine, ethylamine, hydrazine, hydroxylamine, and methoxylamine as nucleophiles that can replace ammonia in the addition to mesaconate, resulting in the corresponding N-substituted methylaspartic acids with excellent diastereomeric excess (>98% de). This newly identified thermostable MAL appears to be a potentially attractive biocatalyst for the stereoselective synthesis of aspartic acid derivatives on large (industrial) scale.  相似文献   

6.
Wu  Zhiyun  Deng  Wenfeng  Tong  Yapei  Liao  Qian  Xin  Dongmin  Yu  Huashun  Feng  Juan  Tang  Lixia 《Applied microbiology and biotechnology》2017,101(8):3201-3211

As a crucial factor for biocatalysts, protein thermostability often arises from a combination of factors that are often difficult to rationalize. In this work, the thermostable nature of halohydrin dehalogenase from Agrobacterium radiobacter AD1 (HheC) was systematically explored using a combinatorial directed evolution approach. For this, a mutagenesis library of HheC mutants was first constructed using error-prone PCR with low mutagenesis frequency. After screening approximately 2000 colonies, six mutants with eight mutation sites were obtained. Those mutation sites were subsequently combined by adopting several rounds of iterative saturation mutagenesis (ISM) approach. After four rounds of saturation mutagenesis, one best mutant ISM-4 with a 3400-fold improvement in half-life (t 1/2) inactivation at 65 °C, 18 °C increase in apparent T m value, and 20 °C increase in optimum temperature was obtained, compared to wild-type HheC. To the best of our knowledge, the mutant represents the most thermostable HheC variant reported up to now. Moreover, the mutant was as active as wild-type enzyme for the substrate 1,3-dichloro-2-propanol, and they remained most enantioselectivity of wild-type enzyme in the kinetic resolution of rac-2-chloro-1-phenolethanol, exhibiting a great potential for industrial applications. Our structural investigation highlights that surface loop regions are hot spots for modulating the thermostability of HheC, the residues located at these regions contribute to the thermostability of HheC in a cooperative way, and protein rigidity and oligomeric interface connections contribute to the thermostability of HheC. All of these essential factors could be used for further design of an even more thermostable HheC, which, in turn, could greatly facilitate the application of the enzyme as a biocatalyst.

  相似文献   

7.
Direct transesterification of (R,S)-1-chloro-3-(3,4-difluorophenoxy)-2-propanol (rac-CDPP) (a key intermediate in the synthesis of the chiral drug (S)-lubeluzole) with vinyl butyrate by lipases from Pseudomonas aeruginosa (P. aeruginosa) MTCC 5113 was performed in hexane with ionic liquids (ILs) 1-butyl-3-methyl imidazolium hexafluorophosphate [BMIm][PF6] and 1-butyl-3-methyl imidazolium tetrafluoroborate [BMIm][BF4] as co-solvents. The maximum conversion (>49%) and enantiomeric excess (ee > 99.9%) was achieved in 6 h of incubation at 30 °C with [BMIm][PF6] as co-solvent in a two-phase system. The enzyme was able to perform with the same specificity even at 60 °C in the presence of ILs. It was possible to use lipases repeatedly for more than 10 times while still maintaining absolute enantioselectivity and reactivity. Stability studies on lipases from P. aeruginosa in ILs revealed the fact that the enzyme constancy and the reactivity in catalyzing transesterification of rac-CDPP into (S)-1-chloro-3-(3,4-difluorophenoxy)-2-butanoate was of the order of [BMIm][PF6] > [BMIm][BF4] in two-phase system.  相似文献   

8.
9.

Enantiomerically pure 3,3,3-trifluoro-2-hydroxy-2-methylpropionic acids are important chiral building blocks for a series of pharmaceuticals. Here, a bacteria strain with 3,3,3-trifluoro-2-hydroxy-2-methylpropanamide-degrading ability was screened and identified as Burkholderia phytofirmans ZJB-15079, from which a novel amidase (Bp-Ami) was cloned and demonstrated to be capable of kinetic resolution of rac-3,3,3-trifluoro-2-hydroxy-2-methylpropanamide to optically pure (R)-3,3,3-trifluoro-2-hydroxy-2-methylpropionic acid. Phylogenetic analysis revealed that Bp-Ami was closely located to the acetamidase/formamidase (FmdA_AmdA) family, and it shared high homology with acetamidases. Bp-Ami was found to be the first cobalt-dependent FmdA_AmdA family amidase. The enzyme activity was significantly increased by 37.7-fold in the presence of 1 mM Co2+, with a specific activity of 753.5 U/mg, K m value of 24.73 mM, and k cat /K m value of 22.47 mM−1 s−1. As an enzyme from mesophile, Bp-Ami exhibited extreme thermostability with a half-life of 47.93 h at 80 °C, which was even superior to other reported amidases from thermophiles. The whole cell catalysis of 200 g/L 3,3,3-trifluoro-2-hydroxy-2-methylpropanamide by Escherichia coli harboring Bp-Ami (5 g/L) resulted in 44 % yield and an enantiomeric excess (ee p) of 95 % within 10 min (E = 86). The high substrate tolerance, high specific activity, and extreme thermostability demonstrated the great potential of Bp-Ami for efficient biocatalytic synthesis of (R)-3,3,3-trifluoro-2-hydroxy-2-methylpropionic acid.

  相似文献   

10.
A novel microbial method of synthesizing (S)-1-phenyl-1,3-propanediol [(S)-PPD] was developed in this study. Our laboratory stock cultures were screened for microorganisms that stereospecifically produced (S)-PPD from 3-hydroxy-1-phenylpropane-1-one (HPPO) using an intact cell system. Of the 828 strains examined (321 bacteria, 233 yeasts and 274 molds), certain strains of Williopsis saturnus var. mrakii and Cryptococcus albidus were found to produce (S)-PPD with over 99% enantiomeric excess (e.e.). Screening identified W. saturnus var. mrakii AJ-5620 as the most productive strain, and this strain was used for further experiments. The (S)-PPD-producing reaction using intact W. saturnus var. mrakii AJ-5620 cells was carried out by successive feeding of HPPO. A total (S)-PPD yield of 9.9 g/l was produced in 20 h. The molar yield was 81% and the optical purity of the (S)-PPD produced was over 99% e.e.  相似文献   

11.

Objectives

To discover novel ketoreductases (KRED) from soil metagenome preparation of chiral alcohols.

Results

Three putative KRED were cloned, heterologously expressed in Eschericha coli and characterized based on the sequence analysis of soil metagenome. All the three enzymes (KRED424, KRED432, and KRED433) had maximum activity at 55 °C and pH 7. KRED424 had a broader substrate spectrum compared with the other two. Three prochiral carbonyl compounds were used to evaluate the abilities of enantioselective reductions of the KRED. For N-Boc-3-pyrrolidone, all enzymes produced an (S)-type alcohol in enantiomeric excess (>99 % ee). For ethyl 2-oxo-4-phenylbutyrate, KRED424 showed a higher conversion (91.5 %) and enantioselectivity (S-type, >99 % ee) than KRED432 and KRED433. For ethyl 4-chloroacetoacetate (COBE), both of KRED424 and KRED433 completely converted 20 mM substrate and KRED433 could obtain an (R)-alcohol with 94 % ee.

Conclusions

The three ketoreductases have potential in the preparation of pharmaceuticals and fine chemicals.
  相似文献   

12.
The effect of surfactants on pyrene degradation in Pseudomonas fluorescens 29L was investigated. This strain produced 30.1 μM of rhamnolipid equivalents (RE) of biosurfactants on 50 mg of pyrene per liter of medium. The production of biosurfactants was significantly correlated with the water solubility (S w) of the substrate and the growth rate on it. When chrysene, with a S w of 2.8 × 10−3 mg per liter of water, was the carbon source, 13.1 μM of RE of biosurfactants were produced compared to 10.3 μM of RE of biosurfactants on acenaphthene with a S w of 1.9 mg per liter of water. No biosurfactants were produced on salicylic acid, catechol, and citrate. All of the strain 29L mutants which grew on pyrene produced biosurfactants while among the mutants which grew on naphthalene, only 88.4% produced biosurfactants. The rhamnolipid mixture, JBR425, inhibited the growth of Strain 29L wild type (WT) and all of its mutants on pyrene. However, these mutants were able to grow in the presence of pyrene when the growth medium was supplemented with 10−6 mg of emulsan per milliliter of medium. This study implies biosurfactants are produced by Strain 29L as a physiological response to the hydrophobicity of pyrene. The combined use of indigenous biosurfactants and the added biosurfactant, emulsan, is a biotechnology to enhance pyrene degradation by Pseudomonas fluorescens 29L.  相似文献   

13.
Burhenne N  Tischner R 《Planta》2000,211(3):440-445
 A method is presented to isolate mutants of Chlorella sorokiniana with defects in NO3 metabolism. Three nitrite-reductase (NIR; E.C.1.7.7.1)-deficient mutants were obtained from 500 pinpoint-colony-forming clones. The final screening was performed using NO3 , NO2 or NH+ 4 as N-source. The mutants isolated absorb NO3 with rates close to those measured for the wild type and they excrete NO2 into the medium. The ratio between NO3 uptake and NO2 excretion was 1:1. The sensitivity of NO3 uptake to NH+ 4 was reduced in the mutant strains as it was in the N-starved wild type of Chlorella. Nitrate reductase (NR; EC 1.6.6.1) expression and NR activity were slightly reduced compared to the wild type due to feedback regulation in the mutant strains. No NIR protein was found in the three mutants. However, NIR activity was obtained (50% of the wild-type) for one mutant strain. The NIR-deficient mutants and the already available NR-deficient mutants will be promising tools for investigations of the nitrate assimilation pathway on the molecular level and for studies searching for signaling of C and N metabolism by inorganic N-compounds. Received: 8 October 1999 / Accepted: 25 January 2000  相似文献   

14.
Ribose-5-phosphate isomerase from Clostridium thermocellum converted d-psicose to d-allose, which may be useful as a pharmaceutical compound, with no by-product. The 12 active-site residues, which were obtained by molecular modeling on the basis of the solved three-dimensional structure of the enzyme, were substituted individually with Ala. Among the 12 Ala-substituted mutants, only the R132A mutant exhibited an increase in d-psicose isomerization activity. The R132E mutant showed the highest activity when the residue at position 132 was substituted with Ala, Gln, Ile, Lys, Glu, or Asp. The maximal activity of the wild-type and R132E mutant enzymes for d-psicose was observed at pH 7.5 and 80°C. The half-lives of the wild-type enzyme at 60°C, 65°C, 70°C, 75°C, and 80°C were 11, 7.0, 4.2, 1.5, and 0.6 h, respectively, whereas those of the R132E mutant enzymes were 13, 8.2, 5.1, 3.1, and 0.9 h, respectively. The specific activity and catalytic efficiency (k cat/K m) of the R132E mutant for d-psicose were 1.4- and 1.5-fold higher than those of the wild-type enzyme, respectively. When the same amount of enzyme was used, the conversion yield of d-psicose to d-allose was 32% for the R132E mutant enzyme and 25% for the wild-type enzyme after 80 min.  相似文献   

15.
Aldo-keto reductases (AKRs) are important biocatalysts that can be used to synthesize chiral pharmaceutical alcohols. In this study, the catalytic activity and stereoselectivity of a NADPH-dependent AKR from Kluyveromyces dobzhanskii (KdAKR) toward t-butyl 6-chloro (5S)-hydroxy-3-oxohexanoate ((5S)-CHOH) were improved by mutating its residues in the loop regions around the substrate-binding pocket. And the thermostability of KdAKR was improved by a consensus sequence method targeted on the flexible regions. The best mutant M6 (Y28A/L58I/I63L/G223P/Y296W/W297H) exhibited a 67-fold higher catalytic efficiency compared to the wild-type (WT) KdAKR, and improved R-selectivity toward (5S)-CHOH (dep value from 47.6% to >99.5%). Moreover, M6 exhibited a 6.3-fold increase in half-life (t1/2) at 40°C compared to WT. Under the optimal conditions, M6 completely converted 200 g/L (5S)-CHOH to diastereomeric pure t-butyl 6-chloro-(3R, 5S)-dihydroxyhexanoate ((3R, 5S)-CDHH) within 8.0 h, with a space-time yield of 300.7 g/L/day. Our results deepen the understandings of the structure−function relationship of AKRs, providing a certain guidance for the modification of other AKRs.  相似文献   

16.
17.
Construction of 25-hydroxy-steroidal side chain substituted with iodine at C-22 was elaborated on a model PTAD-protected steroidal 5,7-diene and applied to a synthesis of (22R)- and (22S)-22-iodo-1α,25-dihydroxyvitamin D3. Configuration at C-22 in the iodinated vitamins, obtained by nucleophilic substitution of the corresponding 22S-tosylates with sodium iodide, was determined by comparison of their iodine-displacement processes and cyclizations leading to isomeric five-membered (22,25)-epoxy-1α-hydroxyvitamin D3 compounds. Also, 20(22)-dehydrosteroids have been obtained and their structures established by 1H NMR spectroscopy. When compared to the natural hormone, (E)-20(22)-dehydro-1α,25-dihydroxyvitamin D3 was found 4 times less potent in binding to the porcine intestinal vitamin D receptor (VDR) and 2 times less effective in differentiation of HL-60 cells. 22-Iodinated vitamin D analogues showed somewhat lower in vitro activity, whereas (22,25)-epoxy analogues were inactive. Interestingly, it was established that (22S)-22-iodo-1α,25-dihydroxyvitamin D3 was 3 times more potent than its (22R)-isomer in binding to VDR and four times more effective in HL-60 cell differentiation assay. The restricted mobility of the side chain of both 22-iodinated vitamin D compounds was analyzed by a systematic conformational search indicating different spatial regions occupied by their 25-oxygen atoms. Preliminary data on the in vivo calcemic activity of the synthesized vitamin D analogues indicate that (E)-20(22)-dehydro-1α,25-dihydroxyvitamin D3 and 22-iodo-1α,25-dihydroxyvitamin D3 isomers were ca. ten times less potent than the natural hormone 1α,25-(OH)2D3 both in intestinal calcium transport and bone calcium mobilization.  相似文献   

18.
Nicotinic acetylcholine receptors (nAChRs) play important roles in nervous system functions and are involved in a variety of diseases. We previously demonstrated that ginsenosides, the active ingredients of Panax ginseng, inhibit subsets of nAChR channel currents, but not α7, expressed in Xenopus laevis oocytes. Mutation of the highly conserved Leu247 to Thr247 in the transmembrane domain 2 (TM2) channel pore region of α7 nAChR induces alterations in channel gating properties and converts α7 nAChR antagonists into agonists. In the present study, we assessed how point mutations in the Leu247 residue leading to various amino acids affect 20(S)-ginsenoside Rg3 (Rg3) activity against the α7 nAChR. Mutation of L247 to L247A, L247D, L247E, L247I, L247S, and L247T, but not L247K, rendered mutant receptors sensitive to Rg3. We further characterized Rg3 regulation of L247T receptors. We found that Rg3 inhibition of mutant α7 nAChR channel currents was reversible and concentration-dependent. Rg3 inhibition was strongly voltage-dependent and noncompetitive manner. These results indicate that the interaction between Rg3 and mutant receptors might differ from its interaction with the wild-type receptor. To identify differences in Rg3 interactions between wild-type and L247T receptors, we utilized docked modeling. This modeling revealed that Rg3 forms hydrogen bonds with amino acids, such as Ser240 of subunit I and Thr244 of subunit II and V at the channel pore, whereas Rg3 localizes at the interface of the two wild-type receptor subunits. These results indicate that mutation of Leu247 to Thr247 induces conformational changes in the wild-type receptor and provides a binding pocket for Rg3 at the channel pore.  相似文献   

19.
A new inhibitor, H-Ala-Ile-pyrrolidin-2-yl boronic acid, was developed as an inhibitor against prolyl tripeptidyl aminopeptidase with a Ki value of 88.1 nM. The structure of the prolyl tripeptidyl aminopeptidase complexed with the inhibitor (enzyme-inhibitor complex) was determined at 2.2 Å resolution. The inhibitor was bound to the active site through a covalent bond between Ser603 and the boron atom of the inhibitor. This structure should closely mimic the structure of the reaction intermediate between the enzyme and substrate. We previously proposed that two glutamate residues, Glu205 and Glu636, are involved in the recognition of substrates. In order to clarify the function of these glutamate residues in substrate recognition, three mutant enzymes, E205A, E205Q, and E636A were generated by site-directed mutagenesis. The E205A mutant was expressed as an inclusion body. The E205Q mutant was expressed in soluble form, but no activity was detected. Here, the structures of the E636A mutant and its complex with the inhibitor were determined. The inhibitor was located at almost the same position as in the wild-type enzyme-inhibitor complex. The amino group of the inhibitor interacted with Glu205 and the main-chain carbonyl group of Gln203. In addition, a water molecule in the place of Glu636 of the wild-type enzyme interacted with the amino group of the inhibitor. This water molecule was located near the position of Glu636 in the wild-type and formed a hydrogen bond with Gln203. The kcat/KM values of the E636A mutant toward the two substrates used were smaller than those of the wild-type by two orders of magnitude. The Ki value of our inhibitor for the E636A mutant was 48.8 μM, which was 554-fold higher than that against the wild-type enzyme. Consequently, it was concluded that Glu205 and Glu636 are significant residues for the N-terminal recognition of a substrate.  相似文献   

20.
Secoisolariciresinol diglucosides (SDGs) (S,S)-SDG-1 (major isomer in flaxseed) and (R,R)-SDG-2 (minor isomer in flaxseed) were synthesized from vanillin via secoisolariciresinol (6) and glucosyl donor 7 through a concise route that involved chromatographic separation of diastereomeric diglucoside derivatives (S,S)-8 and (R,R)-9. Synthetic (S,S)-SDG-1 and (R,R)-SDG-2 exhibited potent antioxidant properties (EC50 = 292.17 ± 27.71 μM and 331.94 ± 21.21 μM, respectively), which compared well with that of natural (S,S)-SDG-1 (EC50 = 275.24 ± 13.15 μM). These values are significantly lower than those of ascorbic acid (EC50 = 1129.32 ± 88.79 μM) and α-tocopherol (EC50 = 944.62 ± 148.00 μM). Compounds (S,S)-SDG-1 and (R,R)-SDG-2 also demonstrated powerful scavenging activities against hydroxyl [natural (S,S)-SDG-1: 3.68 ± 0.27; synthetic (S,S)-SDG-1: 2.09 ± 0.16; synthetic (R,R)-SDG-2: 1.96 ± 0.27], peroxyl [natural (S,S)-SDG-1: 2.55 ± 0.11; synthetic (S,S)-SDG-1: 2.20 ± 0.10; synthetic (R,R)-SDG-2: 3.03 ± 0.04] and DPPH [natural (S,S)-SDG-1: EC50 = 83.94 ± 2.80 μM; synthetic (S,S)-SDG-1: EC50 = 157.54 ± 21.30 μM; synthetic (R,R)-SDG-2: EC50 = 123.63 ± 8.67 μM] radicals. These results confirm previous studies with naturally occurring (S,S)-SDG-1 and establish both (S,S)-SDG-1 and (R,R)-SDG-2 as potent antioxidants and free radical scavengers for potential in vivo use.  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