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 incorporation of the non‐natural amino acids (4R)‐ and (4S)‐fluoroproline (Flp) has been successfully used to improve protein stability, but little is known about their effect on protein folding kinetics. Here we analyzed the influence of (4R)‐ and (4S)‐Flp on the rate‐limiting trans‐to‐cis isomerization of the Ile75–Pro76 peptide bond in the folding of Escherichia coli thioredoxin (Trx). While (4R)‐Flp at position 76 had essentially no effect on the isomerization rate in the context of the intact tertiary structure, (4S)‐Flp accelerated the folding reaction ninefold. Similarly, tenfold faster trans‐to‐cis isomerization of Ile75–(4S)‐Flp76 relative to Ile75–Pro76 was observed in the unfolded state of Trx. Our results show that the replacement of cis prolines by non‐natural proline analogues can be used for modulating the folding rates of proteins with cis prolyl‐peptide bonds in the native state.  相似文献   

2.
The minor cerebrosides from a Far‐Eastern glass sponge Aulosaccus sp. were analyzed as constituents of some multi‐component RP‐HPLC fractions. The structures of eighteen new and one known cerebrosides were elucidated on the basis of NMR spectroscopy, mass spectrometry, optical rotation data and chemical transformations. These β‐D‐glucopyranosyl‐(1→1)‐ceramides contain sphingoid bases N‐acylated with straight‐chain (2R)‐2‐hydroxy fatty acids, namely, (2S,3S,4R,11Z)‐2‐aminoeicos‐11‐ene‐1,3,4‐triol, acylated with 15E‐22:1, 16Z‐21:1, 15Z‐21:1, 15Z‐20:1, 15E‐20:1, 19:0, 18:0 acids, (2S,3S,4R)‐2‐amino‐13‐methyltetradecane‐1,3,4‐triol—with 19Z‐26:1, 16Z‐23:1, 23:0, 22:0 acids, (2S,3S,4R)‐2‐amino‐14‐methylpentadecane‐1,3,4‐triol—with 16Z‐23:1, 16E‐23:1, 15Z‐22:1, 22:0 acids, (2S,3S,4R)‐2‐amino‐14‐methylhexadecane‐1,3,4‐triol, linked to 16Z‐23:1, 15Z‐22:1 acids, (2S,3S,4R)‐2‐amino‐9‐methylhexadecane‐1,3,4‐triol—to 16Z‐23:1 acid, and (2S,3S,4R)‐2‐aminohexadecane‐1,3,4‐triol, attached to 15Z‐22:1 acid. The 13‐methyl and 9‐methyl‐branched trihydroxy sphingoid base backbones (C15 and C17, respectively) have not been found previously in sphingolipids. The ceramide parts, containing other backbones, present new variants of N‐acylation of the marine sphingoid bases with the 2‐hydroxy fatty acids. The combination of the instrumental and chemical methods used in this study improved the efficiency of the structural analysis of such complex cerebroside mixtures that gave more detailed information on glycosphingolipid metabolism of the organism.  相似文献   

3.
Dermorphin analogues, containing a (S)‐ and (R)‐4‐amino‐1,2,4,5‐tetrahydro‐2‐benzazepin‐3‐one scaffold (Aba) and the α‐methylated analogues as conformationally constrained phenylalanines, were prepared. Asymmetric phase‐transfer catalysis was unable to provide the (S)‐α‐Me‐o‐cyanophenylalanine precursor for (S)‐α‐MeAba in acceptable enantiomeric purity. However, by using a Schöllkopf chiral auxiliary, this intermediate was obtained in 88 % ee. [(S)‐Aba 3‐Gly 4]dermorphin retained μ‐opioid affinity but displayed an increased δ‐affinity. The corresponding R epimer was considerably less potent. In contrast, the [(R)‐α‐MeAba 3‐Gly 4]dermorphin isomer was more potent than its S epimer. Tar‐MD simulations of both non‐methylated [Aba 3‐Gly 4]dermorphin analogues showed a degree of folding at the C‐terminal residues toward the N terminus of the peptide, without however, adopting a stabilized β‐turn conformation. The α‐methylated analogues, on the other hand, exhibited a type I/I′ β‐turn conformation over the α‐MeAba 3 and Gly 4 residues, which was stabilized by a hydrogen bond involving Tyr 5‐HN and D ‐Ala 2‐CO.  相似文献   

4.
Saturation mutagenesis at sites lining the binding pockets of enzymes constitutes a viable protein engineering technique for enhancing or inverting stereoselectivity. Statistical analysis shows that oversampling in the screening step (the bottleneck) increases astronomically as the number of residues in the randomization site increases, which is the reason why reduced amino acid alphabets have been employed, in addition to splitting large sites into smaller ones. Limonene epoxide hydrolase (LEH) has previously served as the experimental platform in these methodological efforts, enabling comparisons between single‐code saturation mutagenesis (SCSM) and triple‐code saturation mutagenesis (TCSM); these employ either only one or three amino acids, respectively, as building blocks. In this study the comparative platform is extended by exploring the efficacy of double‐code saturation mutagenesis (DCSM), in which the reduced amino acid alphabet consists of two members, chosen according to the principles of rational design on the basis of structural information. The hydrolytic desymmetrization of cyclohexene oxide is used as the model reaction, with formation of either (R,R)‐ or (S,S)‐cyclohexane‐1,2‐diol. DCSM proves to be clearly superior to the likewise tested SCSM, affording both R,R‐ and S,S‐selective mutants. These variants are also good catalysts in reactions of further substrates. Docking computations reveal the basis of enantioselectivity.  相似文献   

5.
The enantiomers of the potent σ1 ligand fluspidine ( 1 ) were prepared by using chiral preparative HPLC. Synthesis of racemic tosylate 2 and subsequent separation of enantiomers yielded (R)‐ 2 and (S)‐ 2 in excellent enantiomeric purities. The fluspidine enantiomers (R)‐ 1 and (S)‐ 1 were synthesized from (R)‐ 2 and (S)‐ 2 by nucleophilic substitution with tetra‐n‐butylammonium fluoride, affording (R)‐ 1 with 99.6 % ee and (S)‐ 1 with 96.4 % ee. Tosylates (R)‐ 2 and (S)‐ 2 can also serve as precursors for the radiosynthesis of enantiomerically pure radiotracers [18F](R)‐ 1 and [18F](S)‐ 1 . The absolute configuration of the pure enantiomers was elucidated by comparison of their CD spectra with a calculated CD spectrum of a simplified model compound. In receptor binding studies, both enantiomers displayed very high σ1 receptor affinity and selectivity against the σ2 receptor. (R)‐Fluspidine ((R)‐ 1 ) is the eutomer, with a Ki value of 0.57 nM and a eudysmic ratio of 4. Incubation of (R)‐ 1 and (S)‐ 1 with rat liver microsomes led to the identification of seven and eight metabolites, respectively. Although the S‐configured enantiomer formed additional metabolite (S)‐ 1‐3 , it is metabolically more stable than (R)‐ 1 .  相似文献   

6.
We have previously shown that the β‐aminopeptidases BapA from Sphingosinicella xenopeptidilytica and DmpA from Ochrobactrum anthropi can catalyze reactions with non‐natural β3‐peptides and β3‐amino acid amides. Here we report that these exceptional enzymes are also able to utilize synthetic dipeptides with N‐terminal β2‐amino acid residues as substrates under aqueous conditions. The suitability of a β2‐peptide as a substrate for BapA or DmpA was strongly dependent on the size of the Cα substituent of the N‐terminal β2‐amino acid. BapA was shown to convert a diastereomeric mixture of the β2‐peptide H‐β2hPhe‐β2hAla‐OH, but did not act on diastereomerically pure β23‐dipeptides containing an N‐terminal β2‐homoalanine. In contrast, DmpA was only active with the latter dipeptides as substrates. BapA‐catalyzed transformation of the diastereomeric mixture of H‐β2hPhe‐β2hAla‐OH proceeded along two highly S‐enantioselective reaction routes, one leading to substrate hydrolysis and the other to the synthesis of coupling products. The synthetic route predominated even at neutral pH. A rise in pH of three log units shifted the synthesis‐to‐hydrolysis ratio (vS/vH) further towards peptide formation. Because the equilibrium of the reaction lies on the side of hydrolysis, prolonged incubation resulted in the cleavage of all peptides that carried an N‐terminal β‐amino acid of S configuration. After completion of the enzymatic reaction, only the S enantiomer of β2‐homophenylalanine was detected (ee>99 % for H‐(S)‐β2‐hPhe‐OH, E>500); this confirmed the high enantioselectivity of the reaction. Our findings suggest interesting new applications of the enzymes BapA and DmpA for the production of enantiopure β2‐amino acids and the enantioselective coupling of N‐terminal β2‐amino acids to peptides.  相似文献   

7.
The local hydrophobicity of an amino acid residue in a peptide sequence can be determined by measuring the hydrophobicity index (φ0) by reversed-phase (RP) HPLC. Herein, the impact on the local hydrophobicity of the replacement of an amide by a monofluoroalkene unit in short peptides is discussed. Monofluoroalkene-containing dipeptides and tripeptides were synthesized, as well as their natural parent compounds, and the hydrophobicity indexes of these short peptides and peptidomimetics were determined. Comparison between the natural parent peptides and their alkene-containing analogues was made, and the dependence of the peptidomimetic analogues’ behaviour on the pH and the solvent was studied. It was found that the presence of a monofluoroalkene unit enhanced a peptide's hydrophobicity.  相似文献   

8.
A novel fluorinated bis(ether amine) monomer, 2,6‐bis(4‐amino‐2‐trifluoromethylphenoxy) naphthalene, was prepared through the nucleophilic aromatic substitution reaction of 2‐chloro‐5‐nitrobenzotrifluoride and 2,6‐dihydroxynaphthalene in the presence of potassium carbonate, followed by catalytic reduction with hydrazine and Pd/C in ethanol. A series of novel trifluoromethylated polyimides were synthesized from the diamine with various commercially available aromatic tetracarboxylic dianhydrides using a two‐stage process with thermal imidization of poly(amic acid) films. Most of the resulting polyimides were highly soluble in a variety of organic solvents and could afford transparent and tough films via solution casting. These polyimides exhibited moderately high glass transition temperatures (Tgs) of 249–311 °C, high thermal stability and good mechanical properties. Low moisture (0.19–0.85 %), low dielectric constants (2.49–3.59 at 10 kHz), and low color intensity were also observed. For a comparative study, a series of analogous polyimides based on 2,6‐bis(4‐aminophenoxy)naphthalene were also prepared and characterized. Copyright © 2005 Society of Chemical Industry  相似文献   

9.
A method for the synthesis of two (23R)‐ and (23S)‐epimeric pairs of 23‐fluoro‐3α,7α,12α‐trihydroxy‐5β‐cholan‐24‐oic acid and 23‐fluoro‐3α,7α‐dihydroxy‐5β‐cholan‐24‐oic acid is described. The key intermediates, 23,24‐dinor‐22‐aldehyde peracetates were prepared from cholic and chenodeoxycholic acids via the 24‐nor‐22‐ene, 24‐nor‐22ξ,23‐epoxy, and 23,24‐dinor‐22‐aldehyde derivatives. The Horner–Wadsworth–Emmons reaction of the 23,24‐dinor‐22‐aldehydes using triethyl 2‐fluoro‐2‐phosphonoacetate in the presence of LiCl and 1,8‐diazabicyclo[5,4,0]undec‐7‐ene (DBU), and subsequent hydrogenation of the resulting 23ξ‐fluoro‐22‐ene ethyl esters, followed by hydrolysis, gave a mixture of the epimeric (23R)‐ and (23S)‐fluorinated bile acids which were resolved efficiently by preparative RP‐HPLC. The stereochemical configuration of the fluorine atom at C‐23 in the newly synthesized compounds was confirmed directly by the X‐ray crystallographic data. The 1H and 13C NMR spectral differences between the (23R)‐ and (23S)‐epimers were also discussed.  相似文献   

10.
Mixed matrix membranes (MMMs) were made by incorporating vinyltrimethoxysilane (VTMS)‐modified Silicalite‐1 zeolite nanoparticles (V‐Silicalite‐1 NPs) into fluorinated polybenzoxazine (F‐PBZ) modified polydimethylsiloxane (PDMS) polymer through in situ polymerization method. The membrane morphology, surface wettability, and pervaporation performance were systematically investigated. The addition of F‐PBZ into PDMS membranes resulted in substantially improved flux and marginal increase of separation factor, which is the result of higher free volume and higher hydrophobicity caused by the addition of F‐PBZ. The modification of Silicalite‐1 NPs improved the interfacial contact between zeolite crystals and polymer phase. The incorporation of hydrophobic V‐Silicalite‐1 zeolite NPs into the PDMS membranes led to much higher separation factor but reduced flux, which is the result of increased hydrophobicity and reduced free volume. The three‐component MMMs with V‐Silicalite‐1 zeolite NPs in the F‐PBZ fluorinated PDMS exhibited separation factor of 28.7 and flux of 0.207 kg m?2 h?1 for 5 wt % ethanol aqueous solution at 50 °C, while the pure PDMS membranes only had separation factor of 4.8 and flux of 0.088 kg m?2 h?1. The substantial increase of both flux and separation factor were attributed to the higher hydrophobicity and free volume caused by the incorporation of both hydrophobic zeolite crystals and F‐PBZ polymer into the PDMS membranes. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134 , 44753.  相似文献   

11.
In order to develop an optimal subunit as a T‐recognition element in hairpin polyamides, 15 novel chirality‐modified polyamides containing (R)‐α,β‐diaminopropionic acid (Rβ), (S)‐α,β‐diaminopropionic acid (Sβ), (1R,3S)‐3‐aminocyclopentanecarboxylic acid (RSCp), (1S,3R)‐3‐amino‐cyclopentanecarboxylic acid (RSCp), (1R,3R)‐3‐aminocyclopentanecarboxylic acid (RRCp) and (1S,3S)‐3‐amino‐cyclopentanecarboxylic acid (SSCp) residues were synthesized. Their binding characteristics to DNA sequences 5′‐TGC N CAT‐3′/3′‐ACG N′ GTA‐5′ ( N?N′ =A ? T, T ? A, G ? C and C ? G) were systemically studied by surface plasmon resonance (SPR) and molecular simulation (MSim) techniques. SPR showed that polyamide 4 , AcIm‐Sβ‐ImPy‐γ‐ImPy‐β‐Py‐βDp (β/Sβ pair), bound to a DNA sequence containing a core binding site of 5′‐TGC A CAT‐3′ with a dissociation equilibrium constant (KD) of 4.5×10?8 m. This was a tenfold improvement in specificity over 5′‐TGCTCAT‐3′ (KD=4.5×10?7 M ). MSim studies supported the SPR results. More importantly, for the first time, we found that chiral 3‐aminocyclopentanecarboxylic acids in polyamides can be employed as base readers with only a small decrease in binding affinity to DNA. In particular, SPR showed that polyamide 9 (RRCp/β pair) had a 15‐fold binding preference for 5′‐TGCTCAT‐3′ over 5′‐TGCACAT‐3′. A large difference in standard free energy change for A ? T over T ? A was determined (ΔΔGo=5.9 kJ mol?1), as was a twofold decrease in interaction energy by MSim. Moreover, a 1:1 stoichiometry ( 9 to 5′‐TGC T CAT‐3′/3′‐ACG A GTA‐5′) was shown by MSim to be optimal for the chiral five‐membered cycle to fit the minor groove. Collectively, the study suggests that the (S)‐α‐amino‐β‐aminopropionic acid and (1R,3R)‐3‐aminocyclopentanecarboxylic acid can serve as a T‐recognition element, and the stereochemistry and the nature of these subunits significantly influence binding properties in these recognition events. Subunit (1R,3R)‐3‐aminocyclopentanecarboxylic acid broadens our scope to design novel polyamides.  相似文献   

12.
In the presence of Na2CO3 (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐3‐(1,3‐dioxo‐butyl)oxymethyl‐1,2,3,4‐tetrahydrocarboline ( 1 ) were transformed into (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐3‐hydroxymethyl‐1,2,3,4‐tetrahydrocarboline ( 2 ), which were cyclized to (6S)‐3‐acetyl‐6‐hydroxymethyl‐4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizine ( 4 ), via(6S,12bS)‐ and (6S,12bR)‐3‐acetyl‐2‐hydroxyl‐6‐hydroxymethyl‐1,2,3,4,6,7,12,12b‐octahydro‐4‐oxoindolo[2,3‐a]quinoline ( 3 ). (6S)‐ 4 was coupled with Boc‐Gly, Boc‐L‐Asp(β‐benzyl ester), or Boc‐L‐Gln to give 6‐amino acid substituted (6S)‐3‐acetyl‐4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizines 5a , 5b , or 5c , respectively. After the removal of Boc from (6S)‐ 5a (6S)‐3‐acetyl‐6‐glycyl‐4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizine ( 6 ) was obtained. The anticancer activities of (6S)‐ 5 and (6S)‐ 6 in vitro were tested.  相似文献   

13.
Photopolymerization processes are often used in industrial applications because of their solvent‐free formulations and various advantages over conventional thermal processes. Fluorinated monomers and oligomers yield coatings of great interest because of the peculiar characteristics of fluorine atoms: these coatings show hydrophobicity, chemical stability, weathering resistance, etc. Novel UV‐curable fluorinated epoxy acrylate oligomers were synthesized from 1H,1H‐perfluorohexan‐1‐ol, 1,6‐hexamethylene diisocyanate (HDI) and epoxy acrylate (EA). The HDI plays the role of a spacer group in the side chain between the EA backbone chain and the fluorinated segment. This new spacer containing a urethane moiety with long alkyl groups can exhibit a self‐organization effect through the formation of strong hydrogen bonding. This resulted in a stiffening of the whole HDI urethane–perfluoalkyl chain to form nanostructure surface segregation. The designed fluorinated EA with fluoroalkyl (C5F11) units in the side chain exhibited a contact angle of about 151°, which is in the superhydrophobic range. Copyright © 2010 Society of Chemical Industry  相似文献   

14.
The asymmetric 1,4‐addition of phenylboronic acid to cyclohexenone were performed by using a low amount of rhodium/(R)‐(6,6′‐dimethoxybiphenyl‐2,2′‐diyl)bis[bis(3,4,5‐trifluorophenyl)phosphine] (MeO‐F12‐BIPHEP) catalyst. Because the catalyst shows thermal resistance at 100 °C, up to 0.00025 mol% Rh catalyst showed good catalytic activity. The highest turnover frequency (TOF) and turnover number (TON) observed were 53,000 h−1 and 320,000, respectively. The enantioselectivities of the products were maintained at a high level of 98% ee in these reactions. The Eyring plots gave the following kinetic parameters (ΔΔH=−4.0±0.1 kcal mol−1 and ΔΔS=−1.3±0.3 cal mol−1 K−1), indicating that the entropy contribution is relatively small. Both the result and consideration of the transition state in the insertion step at the B3LYP/6‐31G(d) [LANL2DZ for rhodium] levels indicated that the less σ‐donating electron‐poor (R)‐MeO‐F12‐BIPHEP could be creating a rigid chiral environment around the rhodium catalyst even at high temperature.  相似文献   

15.
Methyl (1S,3S and 1R,3S)‐1‐(2, 2‐dimethoxyethyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxylate ( 3 ) was hydrolyzed in the presence of sodium hydroxide to give (1S,3S and 1R,3S)‐1‐(2,2‐dimethoxyethyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxylic acid ( 4 ), which was reduced with LiAlH4 to provide (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐3‐hydroxymethyl‐1,2,3,4‐tetrahydrocarbolines ( 10 ), and then amidated in ammonia containing methanol to obtain (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxamide ( 14 ). Acylation of (1S,3S and 1R,3S)‐ 3 , (1S,3S and 1R,3S)‐ 4 , (1S,3S)‐ 10 , (1R, 3S)‐ 10 , (1S, 3S)‐ 14 and (1R,3S)‐ 14 afforded the corresponding methyl (1S,3S and 1R,3S)‐1‐(2,2‐dimethoxyethyl)‐ 2‐(1,3‐dioxobutyl)‐1,2,3,4‐tetrahydrocarbolines‐3‐carboxylate ( 6 ), (1S,3S and 1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxylic acid ( 5 ), (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐3‐(1,3‐dioxobutyl)oxymethyl‐1,2,3,4‐tetrahydrocarboline ( 11 ), (1S,3S)‐ and (1R,3S)‐1‐(2,2‐dimethoxyethyl)‐2‐(1,3‐dioxobutyl)‐1,2,3,4‐tetrahydrocarboline‐3‐carboxamide ( 15 ), respectively. After Aldol reaction, dehydration and dehydrogenation the desired (6S)‐6‐substituted 4,6,7,12‐tetrahydro‐4‐oxoindolo[2,3‐a]quinolizines 8 , 9 , 12 , 13 , and 16 were obtained. Their anticancer activities in vitro were investigated.  相似文献   

16.
The new complexes RuHCl(PPh2CH2CHRNH2)2 and RuHCl(PPh2CH2CHRNH2)(R‐ binap), R=H (Pgly), R=Me [(R)‐Pala] were prepared by the substitution of the PPh3 ligands in RuHCl(PPh3)3 or RuHCl(PPh3)[(R)‐binap] with beta‐aminophosphines derived from amino acids. The complex trans‐RuHCl(Pgly)[(R)‐binap] has been characterized by X‐ray crystallography. The complex trans‐RuHCl[(S)‐Ppro]2 where (S)‐Ppro is derived from proline was also prepared and characterized by X‐ray crystallography. These were used as catalyst precursors in the presence of a base (KOPr‐i or KOBu‐t) for the hydrogenation of various ketones and imines to the respective alcohols and amines with H2 gas (1–11 atm) at room temperature. Acetophenone was hydrogenated to (S)‐1‐phenylethanol in low ee (up to 40%) when catalyzed by the enantiomerically pure complexes. These complexes are especially active in the hydrogenation of sterically congested and electronically deactivated ketones and imines and are selective for the hydrogenation of CO bonds over CC bonds.  相似文献   

17.
Sum frequency generation (SFG) vibrational spectroscopy has been used to characterize the interfacial structure of a series of model peptides at the hydrophobic polystyrene-buffer interface. The peptides contain two types of amino acids, one hydrophobic (X) and one hydrophilic (Y). Their sequences are Ac-XYYXXYXXYYXXYX-NH2 (XY14) and Ac-XYXYXYX-NH2 (XY7), respectively, where the X and Y combinations are: leucine (L) and lysine (K); alanine (A) and lysine (K); alanine (A) and arginine (R); and phenylalanine (F) and arginine (R), respectively. One additional peptide was synthesized and characterized, Ac-LKKLLKL-NH2, referred to as LK7 α. The XY14 peptides showed SFG spectra that were characteristic of the hydrophobic (X) amino acid of the peptide. Comparison with the 7-amino acid peptides shows that the molecular orientation of alanine is more sensitive to changes in sequence and chain length than leucine or phenylalanine. The hydrophilic amino acids are not observed in the SFG spectra of these peptides at the hydrophobic polystyrene interface (with the possible exception of the AR7 peptide), suggesting the hydrophilic amino acids studied here have a random orientation at this interface. The results of these studies are put into the context of recent SFG studies of proteins adsorbed onto hydrophobic surfaces. Furthermore, our approach to theoretical understanding of interfacial peptide structure is outlined. The results of a molecular dynamics simulation of the LK14 peptide on a hydrophobic interface are presented and discussed.  相似文献   

18.
Fluorinated acrylate emulsions have been extensively applied for hydrophobic surface coatings. To obtain fluorinated emulsions of low surface energy, amphiphilic sulfonyl macro emulsifiers consisting of 2‐(perfluorohexyl) ethyl acrylate (PFHEA) and 2‐acrylanmido‐2‐methylpropanesulfonic acid (AMPS) were designed via radical polymerization and subsequently used for the emulsion copolymerization of PFHEA, methyl methacrylate (MMA), and methyl butyl acrylate (MBA). Under optimum synthesis conditions, the emulsifier displayed superior emulsification properties such as high monomer conversion, small particle size, and excellent stability compared with conventional emulsifiers. The corresponding emulsions with sulfonyl macro emulsifiers exhibited extreme low surface energy (9.8 mN/m) and outstanding hydrophobicity due to high contents of fluorinated chains, as well as thimbleful hydrophilic sulfonyl groups, which shows the great potential in water repellent modification application. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 44921.  相似文献   

19.
Expansion of the structural diversity of peptide antibiotics was performed through two different methods. Supplementation‐based incorporation (SPI) and stop‐codon suppression (SCS) approaches were used for co‐translational incorporation of isostructural and orthogonal noncanonical amino acids (ncAAs) into the lasso peptide capistruin. Two ncAAs were employed for the SPI method and five for the SCS method; each of them probing the incorporation of ncAAs in strategic positions of the molecule. Evaluation of the assembly by HR‐ESI‐MS proved more successful for the SCS method. Bio‐orthogonal chemistry was used for post‐biosynthetic modification of capistruin congener Cap_Alk10 containing the ncAA Alk (Nε‐Alloc‐L ‐lysine) instead of Ala. A second‐generation Hoveyda–Grubbs catalyst was used for an in vitro metathesis reaction with Cap_Alk10 and an allyl alcohol, which offers options for post‐biosynthetic modifications. The use of synthetic biology allows for the in vivo production of new peptide‐based antibiotics from an expanded amino acid repertoire.  相似文献   

20.
Nine new cerebrosides 1a–d , 2a , 2b , 3a–c were found in the extract of a Far‐Eastern glass sponge Aulosaccus sp. (class Hexactinellida). These β‐d ‐glucopyranosyl‐(1 → 1)‐ceramides contain sphingoid bases (2S,3S,4R,11Z)‐2‐aminoeicos‐11‐ene‐1,3,4‐triol (in 1a – d ), (2S,3S,4R,13Z)‐2‐aminoeicos‐13‐ene‐1,3,4‐triol (in 2a , b ) and (2S,3S,4R,13S*,14R*)‐2‐amino‐13,14‐methylene‐eicosane‐1,3,4‐triol (in 3a – c ), which are N‐acylated by (2R,15Z)‐2‐hydroxydocos‐15‐enoic (in 1a , 2a , 3a ), (2R,16Z)‐2‐hydroxytricos‐16‐enoic (in 1b , 2b , 3b ), (2R,17Z)‐2‐hydroxytetracos‐17‐enoic (in 1d ) and (2R)‐2‐hydroxydocosanoic (in 1c , 3c ) acids. The monoenoic and cyclopropane‐containing sphingoid bases of compounds 1a–d , 2a , 2b , 3a–c have not been found previously in any sphingolipids. The structures of the cerebrosides were elucidated on the basis of 1H‐, 13C‐NMR spectroscopy, mass spectrometry, optical rotation data and chemical transformations. A simplified method for the assignment of the absolute configuration of 2‐hydroxy fatty acids by GC analysis of their (2R)‐ and (2S)‐oct‐2‐yl esters was proposed.  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