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 46 毫秒
1.
A dipeptide taste ligand L -aspartyl-D -2-aminobutyric acid-(S)-α-ethylbenzylamide was found to be about 2000 times more potent than sucrose. To investigate the molecular basis of its potent sweet taste, we carried out conformational analysis of this molecule and several related analogues by NMR spectroscopy, computer simulations and X-ray crystallographic studies. The results of the studies support our earlier model that an ‘L’-shape molecular array is essential for eliciting sweet taste. In addition, we have identified an aromatic group located between the stem and the base of the ‘L-shape’, which is responsible for enhancement of sweetness potency. In this study, we also assessed the optimal size of the essential hydrophobic group (X) and the effects of the chirality of the second residue toward taste. ©1997 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
A less frequently employed Escherichia coli strain W, yet possessing useful metabolic characteristics such as less acetic acid production and high L ‐valine tolerance, was metabolically engineered for the production of L ‐valine. The ilvA gene was deleted to make more pyruvate, a key precursor for L ‐valine, available for enhanced L ‐valine biosynthesis. The lacI gene was deleted to allow constitutive expression of genes under the tac or trc promoter. The ilvBNmut genes encoding feedback‐resistant acetohydroxy acid synthase (AHAS) I and the L ‐valine biosynthetic ilvCED genes encoding acetohydroxy acid isomeroreductase, dihydroxy acid dehydratase, and branched chain amino acid aminotransferase, respectively, were amplified by plasmid‐based overexpression. The global regulator Lrp and L ‐valine exporter YgaZH were also amplified by plasmid‐based overexpression. The engineered E. coli W (ΔlacI ΔilvA) strain overexpressing the ilvBNmut, ilvCED, ygaZH, and lrp genes was able to produce an impressively high concentration of 60.7 g/L L ‐valine by fed‐batch culture in 29.5 h, resulting in a high volumetric productivity of 2.06 g/L/h. The most notable finding is that there was no other byproduct produced during L ‐valine production. The results obtained in this study suggest that E. coli W can be a good alternative to Corynebacterium glutamicum and E. coli K‐12, which have so far been the most efficient L ‐valine producer. Furthermore, it is expected that various bioproducts including other amino acids might be more efficiently produced by this revisited platform strain of E. coli. Bioeng. 2011; 108:1140–1147. © 2010 Wiley Periodicals, Inc.  相似文献   

3.
A computer‐aided engineering approach recently enabled to deeply reshape the active site of N. polysaccharea amylosucrase for recognition of non‐natural acceptor substrates. Libraries of variants were constructed and screened on sucrose allowing the identification of 17 mutants able to synthesize molecules from sole sucrose, which are not synthesized by the parental wild‐type enzyme. Three of the isolated mutants as well as the new products synthesized were characterized in details. Mutants contain between 7 and 11 mutations in the active site and the new molecules were identified as being a sucrose derivative, named erlose (α‐d ‐glucopyranosyl‐(1→4)‐α‐d ‐glucopyranosyl‐(1→2)‐β‐d ‐Fructose), and a new malto‐oligosaccharide named panose (α‐d ‐glucopyranosyl‐(1→6)‐α‐d ‐glucopyranosyl‐(1→4)‐α‐d ‐Glucose). These product specificities were never reported for none of the amylosucrases characterized to date, nor their engineered variants. Optimization of the production of these trisaccharides of potential interest as sweeteners or prebiotic molecules was carried out. Molecular modelling studies were also performed to shed some light on the molecular factors involved in the novel product specificities of these amylosucrase variants.  相似文献   

4.
We have previously reported the development of a 100% genetically defined engineered Escherichia coli strain capable of producing L ‐valine from glucose with a high yield of 0.38 g L ‐valine per gram glucose (0.58 mol L ‐valine per mol glucose) by batch culture. Here we report a systems biological strategy of employing flux response analysis in bioprocess development using L ‐valine production by fed‐batch culture as an example. Through the systems‐level analysis, the source of ATP was found to be important for efficient L ‐valine production. There existed a trade‐off between L ‐valine production and biomass formation, which was optimized for the most efficient L ‐valine production. Furthermore, acetic acid feeding strategy was optimized based on flux response analysis. The final fed‐batch cultivation strategy allowed production of 32.3 g/L L ‐valine, the highest concentration reported for E. coli. This approach of employing systems‐level analysis of metabolic fluxes in developing fed‐batch cultivation strategy would also be applicable in developing strategies for the efficient production of other bioproducts. Biotechnol. Bioeng. 2011; 108:934–946. © 2010 Wiley Periodicals, Inc.  相似文献   

5.
Aims: This study focused on the cloning, expression and characterization of recombinant α‐l ‐arabinosidases from Bifidobacterium longum H‐1. Methods and Results: α‐l ‐Arabinofuranosidase (AfuB‐H1) and bifunctional α‐l ‐arabinopyranosidase/β‐d ‐galactosidase (Apy‐H1) from B. longum H‐1 were identified by Southern blotting, and their recombinant enzymes were overexpressed in Escherichia coli BL21 (DE3). Recombinant AfuB‐H1 (rAfuB‐H1) was purified by single‐step Ni2+‐affinity column chromatography, whereas recombinant Apy‐H1 (rApy‐H1) was purified by serial Q‐HP and Ni2+‐affinity column chromatography. Enzymatic properties and substrate specificities of the two enzymes were assessed, and their kinetic constants were calculated. According to the results, rAfuB‐H1 hydrolysed p‐nitrophenyl‐α‐l ‐arabinofuranoside (pNP‐αL‐Af) and ginsenoside Rc, but did not hydrolyse p‐nitrophenyl‐α‐l ‐arabinopyranoside (pNP‐αL‐Ap). On the other hand, rApy‐H1 hydrolysed pNP‐αL‐Ap, p‐nitrophenyl‐β‐d ‐galactopyranoside (pNP‐βD‐Ga) and ginsenoside Rb2. Conclusions: Ginsenoside‐metabolizing bifidobacterial rAfuB‐H1 and rApy‐H1 were successfully cloned, expressed, and characterized. rAfuB‐H1 specifically recognized the α‐l ‐arabinofuranoside, whereas rApy‐H1 had dual functions, that is, it could hydrolyse both β‐d ‐galactopyranoside and α‐l ‐arabinopyranoside. Significance and Impact of the Study: These findings suggest that the biochemical properties and substrate specificities of these recombinant enzymes differ from those of previously identified α‐l ‐arabinosidases from Bifidobacterium breve K‐110 and Clostridium cellulovorans.  相似文献   

6.
Four diastereomeric‐Leu‐Leu‐Aib‐Leu‐Leu‐Aib‐peptides, Boc‐D ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (1), Boc‐L ‐Leu‐D ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (2), Boc‐L ‐Leu‐L ‐Leu‐Aib‐D ‐Leu‐L ‐Leu‐Aib‐OMe (3), and Boc‐L ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐D ‐Leu‐Aib‐OMe (4), were synthesized. The crystals of the four hexapeptides were characterized by X‐ray crystallographic analysis. Two diastereomeric hexapeptides 1 and 2 having D ‐Leu(1) or D ‐Leu(2) were folded into right‐handed (P) 3 10 ‐helical structures, while peptide 3 having D ‐Leu(4) was folded into a turn structure nucleated by type III′ and I$' \bf{\beta}$ ‐turns, and peptide 4 having D ‐Leu(5) was folded into a left‐handed (M) 3 10 ‐helical structure. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
One chiral L ‐valine (L ‐Val) was inserted into the C‐terminal position of achiral peptide segments constructed from α‐aminoisobutyric acid (Aib) and α,β‐dehydrophenylalanine (ΔZPhe) residues. The IR, 1H NMR and CD spectra indicated that the dominant conformations of the pentapeptide Boc‐Aib‐ΔPhe‐(Aib)2‐L ‐Val‐NH‐Bn (3) and the hexapeptide Boc‐Aib‐ΔPhe‐(Aib)3‐L ‐Val‐NH‐Bn (4) in solution were both right‐handed (P) 310‐helical structures. X‐ray crystallographic analyses of 3 and 4 revealed that only a right‐handed (P) 310‐helical structure was present in their crystalline states. The conformation of 4 was also studied by molecular‐mechanics calculations. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Cofactor recycling is known to be crucial for amino acid synthesis. Hence, cofactor supply was now analyzed for L ‐valine to identify new targets for an improvement of production. The central carbon metabolism was analyzed by stoichiometric modeling to estimate the influence of cofactors and to quantify the theoretical yield of L ‐valine on glucose. Three different optimal routes for L ‐valine biosynthesis were identified by elementary mode (EM) analysis. The modes differed mainly in the manner of NADPH regeneration, substantiating that the cofactor supply may be crucial for efficient L ‐valine production. Although the isocitrate dehydrogenase as an NADPH source within the tricarboxylic acid cycle only enables an L ‐valine yield of YVal/Glc = 0.5 mol L ‐valine/mol glucose (mol Val/mol Glc), the pentose phosphate pathway seems to be the most promising NADPH source. Based on the theoretical calculation of EMs, the gene encoding phosphoglucoisomerase (PGI) was deleted to achieve this EM with a theoretical yield YVal/Glc = 0.86 mol Val/mol Glc during the production phase. The intracellular NADPH concentration was significantly increased in the PGI‐deficient mutant. L ‐Valine yield increased from 0.49 ± 0.13 to 0.67 ± 0.03 mol Val/mol Glc, and, concomitantly, the formation of by‐products such as pyruvate was reduced. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

9.
Trehalose (α‐D ‐glucopyranosyl‐(1,1)‐α‐D ‐glucopyranoside) is widely used in the food industry, thanks to its protective effect against freezing and dehydration. Analogs of trehalose have the additional benefit that they are not digested and thus do not contribute to our caloric intake. Such trehalose analogs can be produced with the enzyme trehalose phosphorylase, when it is applied in the reverse, synthetic mode. Despite the enzyme's broad acceptor specificity, its catalytic efficiency for alternative monosaccharides is much lower than for glucose. For galactose, this difference is shown here to be caused by a lower Km whereas the kcat for both substrates is equal. Consequently, increasing the affinity was attempted by enzyme engineering of the trehalose phosphorylase from Thermoanaerobacter brockii, using both semirational and random mutagenesis. While a semirational approach proved unsuccessful, high‐throughput screening of an error‐prone PCR library resulted in the discovery of three beneficial mutations that lowered Km two‐ to three‐fold. In addition, it was found that mutation of these positions also leads to an improved catalytic efficiency for mannose and fructose, suggesting their involvement in acceptor promiscuity. Combining the beneficial mutations did not further improve the affinity, and even resulted in a decreased catalytic activity and thermostability. Therefore, enzyme variant R448S is proposed as new biocatalyst for the industrial production of lactotrehalose (α‐D ‐glucopyranosyl‐(1,1)‐α‐D ‐galactopyranoside). © 2012 American Institute of Chemical Engineers Biotechnol. Prog., 2012  相似文献   

10.
New analogues of deltorphin I (DT I, Tyr‐d ‐Ala‐Phe‐Asp‐Val‐Val‐Gly‐NH2), with the d ‐Ala residue in position 2 replaced by α‐methyl‐β‐azido(amino, 1‐pyrrolidinyl, 1‐piperidinyl or 4‐morpholinyl)alanine, were synthesized by a combination of solid‐phase and solution methods. All ten new analogues were tested for receptor affinity and selectivity to μ‐ and δ‐opioid receptors. The affinity of analogues containing (R) or (S)‐α‐methyl‐β‐azidoalanine in position 2 to δ‐receptors strongly depended on the chirality of the α,α‐disubstituted residue. Peptide II , containing (S)‐α‐methyl‐β‐azidoalanine in position 2, displayed excellent δ‐receptor selectivity with its δ‐receptor affinity being only three times lower than that of DT I. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
The use of columns packed with sub‐2 μm particles in liquid chromatography with very high pressure conditions (known as UHPLC) was investigated for the fast enantioseparation of drugs. Two different procedures were evaluated and compared using amphetamine derivatives and β‐blockers as model compounds. In one case, cyclodextrins (CD) were directly added to the mobile phase and chiral separations were carried out in less than 5 min. However, this strategy suffered from several drawbacks linked to column lifetime and low chromatographic efficiencies. In the other case, the analysis of enantiomers was carried out after a derivatization procedure using two different reagents, 2,3,4‐tri‐O‐acetyl‐α‐D ‐arabinopyranosyl isothiocyanate (AITC) and N‐α‐(2,4‐dinitro‐5‐fluorophenyl)‐L ‐alaninamide (Marfey's reagent). Separation of several amphetamine derivatives contained within the same sample was achieved in 2–5 min with high efficiency and selectivity. The proposed approach was also successfully applied to the enantiomeric purity determination of (+)‐(S)‐amphetamine and (+)‐(S)‐methamphetamine. Similar results were obtained with β‐blockers, and the separation of 10 enantiomers was carried out in less than 3 min, whereas the individual separation of several β‐blocker enantiomers was performed in 1 min or less. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

12.
The aim of this work was to prepare L ‐ and D ‐(adamant‐1‐yl)‐Gly‐L ‐Ala‐D ‐isoGln peptides in order to study their adjuvant (immunostimulating) activities. Adjuvant activity of adamant‐1‐yl tripeptides was tested in the mouse model using ovalbumin as an antigen and in comparison to the peptidoglycan monomer (PGM; β‐D ‐GlcNAc‐(1→4)‐D ‐MurNAc‐L ‐Ala‐D ‐isoGln‐mesoDAP(εNH2)‐D ‐Ala‐D ‐Ala) and structurally related adamant‐2‐yl tripeptides.  相似文献   

13.
Three new steroid saponins (3β,25R)‐spirost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[β‐D ‐glucopyranosyl‐(1→4)‐6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 1 ), (3β,22R,25R)‐26‐(β‐D ‐glucopyranosyloxy)‐22‐hydroxyfurost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 3 ), and (3β,22R,25R)‐26‐(β‐D ‐glucopyranosyloxy)‐22‐hydroxyfurost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[β‐D ‐glucopyranosyl‐(1→4)‐6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 5 ), as well as the new pregnane glycoside (3β,16β)‐3‐{[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranosyl]oxy}‐20‐oxopregn‐5‐en‐16‐yl (4R)‐5‐(β‐D ‐glucopyranosyloxy)‐4‐methylpentanoate ( 6 ), were isolated from the rhizomes of Tacca integrifolia together with two known (25R) configurated steroid saponins (3β,25R)‐spirost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 2 ) and (3β,22R,25R)‐26‐(β‐D ‐glucopyranosyloxy)‐22‐methoxyfurost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 4 ). The cytotoxic activity of the isolated compounds was evaluated in HeLa cells and showed the highest cytotoxicity value for compound 2 with an IC50 of 1.2±0.4 μM . Intriguingly, while compounds 1 – 5 exhibited similar cytotoxic properties between 1.2±0.4 ( 2 ) and 4.0±0.6 μM ( 5 ), only compound 2 showed a significant microtubule‐stabilizing activity in vitro.  相似文献   

14.
The half reactions of ω‐aminotransferase (ω‐AT) from Vibrio fluvialis JS17 (ω‐ATVf) were carried out using purified pyridoxal 5′‐phosphate‐enzyme (PLP‐Enz) and pyridoxamine 5′‐phosphate‐enzyme (PMP‐Enz) complexes to investigate the relative activities of substrates. In the reaction generating PMP‐Enz from PLP‐Enz using L ‐alanine as an amine donor, L ‐alanine showed about 70% of the initial reaction rate of (S)‐α‐methylbenzylamine ((S)‐α‐MBA). However, in the subsequent half reaction recycling PLP‐Enz from PMP‐Enz using acetophenone as an amine acceptor, acetophenone showed nearly negligible reactivity compared to pyruvate. These results indicate that the main bottleneck in the asymmetric synthesis of (S)‐α‐MBA lies not in the amination of PLP by alanine, but in the amination of acetophenone by PMP‐Enz, where conformational restraints of the enzyme structure is likely to be the main reason for limiting the amine group transfer from PMP‐Enz to acetophenone. Based upon those half reaction experiments using the two amino acceptors of different activity, it appears that the relative activities of the two amine donors and the two acceptors involved in the ω‐AT reactions can roughly determine the asymmetric synthesis yield of the target chiral amine compound. Predicted conversion yields of several target chiral amines were calculated and compared with the experimental conversion yields. Approximately, a positive linear correlation (Pearson's correlation coefficient = 0.92) was observed between the calculated values and the experimental conversion yields. To overcome the low (S)‐α‐MBA productivity of ω‐ATVf caused by the possible disadvantageous structural constraints for acetophenone, new ω‐ATs showing higher affinity to benzene ring of acetophenone than ω‐ATVf were computationally screened using comparative modeling and protein‐ligand docking. ω‐ATs from Streptomyces avermitilis MA‐4680 (SAV2612) and Agrobacterium tumefaciens str. C58 (Atu4761) were selected, and the two screened ω‐ATs showed higher asymmetric synthesis reaction rate of (S)‐α‐MBA and lower (S)‐α‐MBA degradation reaction rate than ω‐ATVf. To verify the higher conversion yield of the variants of ω‐ATs, the reaction with 50 mM acetophenone and 50 mM alanine was performed with coupling of lactate dehydrogenase and two‐phase reaction system. SAV2612 and Atu4761 showed 70% and 59% enhanced yield in the synthesis of (S)‐α‐MBA compared to that of ω‐ATVf, respectively. Biotechnol. Bioeng. 2011;108: 253–263. © 2010 Wiley Periodicals, Inc.  相似文献   

15.
The mannosylated derivative of adamant‐1‐yl tripeptide (D ‐(Ad‐1‐yl)Gly‐L ‐Ala‐D ‐isoGln) was prepared to study the effects of mannosylation on adjuvant (immunostimulating) activity. Mannosylated adamant‐1‐yl tripeptide (Man‐OCH2CH(Me)CO‐D ‐(Ad‐1‐yl)Gly‐L ‐Ala‐D ‐isoGln) is a non‐pyrogenic, H2O‐soluble, and non‐toxic compound. Adjuvant activity of mannosylated adamantyl tripeptide was tested in the mouse model with ovalbumin as an antigen and in comparison to the parent tripeptide and peptidoglycan monomer (PGM, β‐D ‐GlcNAc‐(1→4)‐D ‐MurNAc‐L ‐Ala‐D ‐isoGln‐mesoDAP(εNH2)‐D ‐Ala‐D ‐Ala), a well‐known effective adjuvant. The mannosylation of adamantyl tripeptide caused the amplification of its immunostimulating activity in such a way that it was comparable to that of PGM.  相似文献   

16.
The 50‐residue snake venom protein L ‐omwaprin and its enantiomer D ‐omwaprin were prepared by total chemical synthesis. Radial diffusion assays were performed against Bacillus megaterium and Bacillus anthracis; both L ‐ and D ‐omwaprin showed antibacterial activity against B. megaterium. The native protein enantiomer, made of L ‐amino acids, failed to crystallize readily. However, when a racemic mixture containing equal amounts of L ‐ and D ‐omwaprin was used, diffraction quality crystals were obtained. The racemic protein sample crystallized in the centrosymmetric space group P21/c and its structure was determined at atomic resolution (1.33 Å) by a combination of Patterson and direct methods based on the strong scattering from the sulfur atoms in the eight cysteine residues per protein. Racemic crystallography once again proved to be a valuable method for obtaining crystals of recalcitrant proteins and for determining high‐resolution X‐ray structures by direct methods.  相似文献   

17.
Y.-W. Lu    J. Chen    H.-Y. Zheng    M. J. Adams    J.-P. Chen 《Journal of Phytopathology》2008,156(4):251-255
Coat protein genes of six members of the genus Allexivirus, Garlic viruses ‐A, ‐B, ‐C, ‐D, ‐E and ‐X, were over‐expressed in bacteria and used to prepare polyclonal antibodies. In western blot analysis, Garlic virus D (GarV‐D) and GarV‐E were closely related serologically and there were weaker relationships between some other pairs, particularly GarV‐E with GarV‐A, GarV‐X with GarV‐A, GarV‐X with GarV‐B and GarV‐X with GarV‐C. Sequence analysis suggested that there was no simple relationship between amino acid identity and serological properties.  相似文献   

18.
A bioassay‐guided phytochemical analysis of the ethanolic extract of Grindelia argentina Deble & Oliveira ‐Deble (Asteraceae) allowed the isolation of a known flavone, hispidulin, and three new oleanane‐type saponins, 3‐Oβ‐D ‐xylopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐2β,3β,16α,23‐tetrahydroxyolean‐12‐en‐28‐oic acid 28‐Oβ‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐apiofuranosyl‐(1→3)‐β‐D ‐xylopyranosyl‐(1→3)‐α‐L ‐rhamnopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl ester ( 2 ), 3‐Oβ‐D ‐glucopyranosyl‐2β,3β,23‐trihydroxyolean‐12‐en‐28‐oic acid 28‐Oβ‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐apiofuranosyl‐(1→3)‐β‐D ‐xylopyranosyl‐(1→3)‐α‐L ‐rhamnopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl ester, ( 3 ) and 3‐Oβ‐D ‐xylopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐2β,3β,23‐trihydroxyolean‐12‐en‐28‐oic acid 28‐Oβ‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐apiofuranosyl‐(1→3)‐β‐D ‐xylopyranosyl‐(1→3)‐α‐L ‐rhamnopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl ester ( 4 ), named grindeliosides A–C, respectively. Their structures were determined by extensive 1D‐ and 2D‐NMR experiments along with mass spectrometry and chemical evidence. The isolated compounds were evaluated for their inhibitory activities against LPS/IFN‐γ‐induced NO production in RAW 264.7 macrophages and for their cytotoxic activities against the human leukemic cell line CCRF‐CEM and MRC‐5 lung fibroblasts. Hispidulin markedly reduced LPS/IFN‐γ‐induced NO production (IC50 51.4 μM ), while grindeliosides A–C were found to be cytotoxic, with grindelioside C being the most active against both CCRF‐CEM (IC50 4.2±0.1 μM ) and MRC‐5 (IC50 4.5±0.1 μM ) cell lines.  相似文献   

19.
Four cyclic octapeptides were designed from ascidiacyclamide [cyclo(–Ile–Oxz–D ‐Val– Thz–)2] (ASC, 1 ) to investigate the effects of oxazoline (Oxz) and thiazole (Thz) rings on the structures and cytotoxicities of the peptides. cyclo(–Ile–Thz–D ‐Val–Oxz–)2 ( 2 ) had the same number of Oxz and Thz rings as ASC, but the ring positions were switched. cyclo(–Ile–Oxz–D ‐Val–Thz–Ile–Thz–D ‐Val–Thz–) ( 3 ) and cyclo(–Ile–Thz–D ‐Val–Oxz–Ile–Thz–D ‐Val–Thz–) ( 4 ) contained one Oxz and three Thz rings within the molecule. All Oxz rings were substituted with Thz in cyclo(–Ile–Thz–D ‐Val–Thz–)2 ( 5 ). These analogues had new Oxz and Thz blocks forming the 24‐membered ring. Based on CD spectra and X‐ray diffraction analyses, the structures of all four analogues were classified as square ASC forms. But the structures of 2 and 5 differed from the original square form of 1 , and they showed no cytotoxicity. The structure of 3 was very similar to that of 1 , and 3 showed 10 times greater cytotoxicity than 1 . Although no definite structure of 4 was obtained, it showed three times greater cytotoxicity than 1 . It appears that the position and number of Oxz residues are essential determinants in the structure‐cytotoxicity relationship of ASC analogues.  相似文献   

20.
The interaction of trimannoside, α?benzyl 3, 6‐di‐O‐(α‐D ‐mannopyranosyl)‐α‐D ‐mannopyranoside, 1 with ASAI (Allium sativam agglutinin I, garlic lectin) was studied to reveal the conformational preferences of this ligand in bound‐state and detailed binding mode at atomic level. The binding phenomenon was then compared with another well‐known mannose‐binding lectin, ConA (Concanavalin A). Structural studies of the ligand in free state were done using NMR spectroscopy and Molecular Dynamics simulations. It is found that the substituted‐trimannoside can undergo conformational transitions in solution, with one major and one minor conformation per glycosidic linkage (α 1→3 and α 1→6). On the other hand in the bound‐state only one of the two major conformations was significantly populated. The role of phenyl ring in the binding process was explored. An extended binding site was observed for the trimannoside in ASAI utilizing the aromatic substituent, which is not seen in ConA. Binding data from difference absorption spectroscopy supported this fact that the binding of benzyl‐substituted ligand is tighter with ASAI than ConA. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 952–967, 2010.  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