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 15 毫秒
1.
Oligomers of (R)‐3‐hydroxybutanoate (OHB) have previously been shown to transport cations through lipid bilayers. The ion‐transport activity has been attributed to the formation of hydrophobic aggregates or pores, which have been identified by fluorescence‐microscopy measurements of membrane‐incorporated fluorescence‐labeled OHBs. To obtain more information about these aggregates, we describe here the synthesis of the specifically F‐labeled HB oligomers II – IV for structural investigation by means of solid‐state 19F‐NMR spectroscopic techniques.  相似文献   

2.
Oligomers of 3‐hydroxyalkanoic acids that contain two, three, and six residues with and without O‐terminal (tBu)Ph2Si and C‐terminal PhCH2 protection have been synthesized in such a way that the side chains on the oligoester backbone were those of the proteinogenic amino acids Ala (Me), Val (CHMe2), and Leu (CH2CHMe2). The enantiomerically pure 3‐hydroxyalkanoates were obtained by Noyori hydrogenation of the corresponding 3‐oxo‐alkanoates with [Ru((R)‐binap)Cl2](binap=2,2′bis(diphenylphosphanyl)‐1,1′‐binaphthalene)/H2 (Scheme 1), and the coupling was achieved under the conditions (pyridine/(COCl)2, CH2Cl2, −78°) previously employed for the synthesis of various oligo(3‐hydroxybutanoic acids) (Schemes 2 and 3). The Cotton effects in the CD spectra of the new oligoesters provided no hints about chiral conformation (cf. a helix) in MeOH, MeCN, octan‐1‐ol, or CF3CH2OH solutions (Figs. 1 and 2). Detailed NMR investigations in CDCl3 solution (Figs. 36, and Tables 15) of the hexa(3‐hydroxyalkanoic acid) with the side chains of Val (HC), Ala (HB), Leu (HH), Val, Ala, Leu (from O‐ to C‐terminus; 3 ) gave, on the NMR time‐scale, no evidence for the presence of any significant amount of a 21‐ or a 31‐helical conformation, comparable to those identified in stretched fibers of poly[(R)‐3‐hydroxybutanoic acid], or in lamellar crystallites and in single crystals of linear and cyclic oligo[(R)‐3‐hydroxybutanoic acids], or in the corresponding β‐peptide(s) (the oligo(3‐aminoalkanoic acid) analogs; 1 – 3 ). Thus, the extremely high flexibility (averaged or ‘random‐coil' conformation) of the polyester chain (CO−O rotational barrier ca. 13 kcal/mol; no hydrogen bonding), as compared to polyamide chains (CO−NH barrier ca. 18 kcal/mol; hydrogen bonding) has been demonstrated once again. The possible importance of this structural flexibility, which goes along with amphiphilic properties, for the role of PHB in biology, in evolution, and in prebiotic chemistry is discussed. Structural similarities of natural potassium‐channeling proteins and complexes of oligo(3‐hydroxybutanoates) with Na+, K+, or Ba2+ are alluded to (Figs. 79).  相似文献   

3.
The benzyl esters 1 – 3 of oligo[(R)-3-hydroxybutanoic acids] (OHB) containing 2, 16, or 32 HB units were coupled at the hydroxy terminus with arginine (by esterification with carbodiimide), with glucose (by acetalization with glucosyl trichloroacetimidate), and with 7-(dimethylamino)coumarin-4-acetic acid and biotin (by amide formation through a glycine linker) to give, after deprotection(s), the corresponding `labelled' OHB acids 7 – 9 , 12 , 13 , 25 , 26 , 33 , and 34 (Schemes 1, 4, and 5). The respective novel 16- and 32mer derivatives exhibit distinct water solubility (Table) or may be detected (in minute amounts) by fluorescence spectroscopy, properties required for biochemical investigations.  相似文献   

4.
Morphological changes of solution‐grown poly[(R)‐3‐hydroxybutyrate] lamellar crystals during heating were directly investigated by atomic force microscopy. The thickening of lamellar crystals was further visualized by enzymatic degradation of less‐ordered crystal regions in thermally treated lamellar crystals. The morphological changes of lamellar crystals induced by thermal treatment are due to recrystallization.  相似文献   

5.
6.
7.
The synthesis and single crystal X‐ray structure determination are reported for the 2,2′ : 6′,2″‐terpyridine (= tpy) adduct of bismuth(III) nitrate. The hydroxide‐bridged dimer [(η2‐NO3)2(tpy)Bi(μ‐OH)2Bi(tpy)(η2‐NO3)2] with nine‐coordinate geometry about Bi was the only isolable product from all crystallization attempts in varying ratios of Bi(NO3) : terpy.; [(η2‐NO3)2(tpy)Bi(μ‐OH)2Bi(tpy) · (η2‐NO3)2] is triclinic, P 1, a = 7.941(8), b = 10.732(9), c = 11.235(9) Å; α = 63.05(1), β = 85.01(1), γ = 79.26(1)°, Z = 1, dimer, R = 0.058 for N0 = 2319.  相似文献   

8.
The utilization of poly[(R)‐3‐hydroxybutyric acid] (PHB) biopolymer for a device that uses charging process in friction to convert mechanical energy into electric power is reported. The triboelectric generator (TEG) is fabricated by stacking a drop cast PHB film between indium tin oxide coated poly(ethylene terephthalate) (PET) and PET sheet. The charge transfer takes place through an established general rule according to which the material with higher dielectric constant becomes positively charged. Furthermore, the utilization of such TEG as pressure sensor is illustrated. TEGs have the potential of harvesting energy from touch screen, mechanical vibration, and more, with great applications in self‐powered sensors for heat and environmental monitoring and even large‐scale applications. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2014 , 52, 859–863  相似文献   

9.
(E)‐4‐Hydroxy‐3‐methylbut‐2‐enyl diphosphate ( 1 ) is a key intermediate of the deoxyxylulose phosphate pathway of isoprenoid biosynthesis and a precursor of the plant hormone zeatin. The availability of this intermediate with various labeling patterns is pivotal for its use in biosynthetic studies. The number of positions, however, that can be easily labeled by chemical synthesis is limited, and the synthesis by means of recombinant enzymes is laborious and time consuming. We demonstrated that chromoplasts from Capsicum annuum, whose enzyme activity was impaired by freeze‐thawing, accumulate 1 . This observation built the basis for the development of a cell‐free system allowing the synthesis of this intermediate with labels in various positions. With 2C‐methyl‐D ‐erythritol 2,4‐cyclodiphosphate ( 5 ) as substrate, yields were in the range of 50%.  相似文献   

10.
The configuration of the chiral ring atoms of the title compound, C26H26N2O, obtained in an enantioselective synthesis, has been established relative to the known R configuration of the α‐methyl­benzyl moieties. The crystal packing involves a two‐dimensional network of C—H?π interactions between the aromatic rings.  相似文献   

11.
Two series of segmented poly(ester‐urethane)s were synthesized from bacterial poly[(R)‐3‐hydroxybutyrate]‐diol (PHB‐diol), as hard segments, and either poly(ε‐caprolactone)‐diol (PCL‐diol) or poly(butylene adipate)‐diol (PBA‐diol), as soft segments, using 1,6‐hexamethylene diisocyanate as a chain extender. The hard‐segment content varied from 0 to 50 wt.‐%. These materials were characterized using 1H NMR spectroscopy and GPC. The polymers obtained were investigated calorimetrically and dielectrically. DSC showed that the Tg of either the PCL or PBA soft segments are shifted to higher temperatures with increasing PHB hard‐segment content, revealing that either the PCL or PBA are mixed with small amounts of PHB in the amorphous domains. The results also showed that the crystallization of soft or hard segments was physically constrained by the microstructure of the other crystalline phase, which results in a decrease in the degree of crystallinity of either the soft or hard segments upon increase of the other component. The dielectric spectra of poly(ester‐urethane)s, based on PCL and PHB, showed two primary relaxation processes, designated as αS and αH, which correspond to glass–rubber transitions of PCL soft and PHB hard segments, respectively. Whereas in the case of other poly(ester‐urethane)s, derived from PBA and PHB, only one relaxation process was observed, which broadens and shifts to higher temperature with increasing PHB hard‐segment content. It was concluded from these results that our investigated materials exhibit micro‐phase separation of the hard and soft segments in the amorphous domains.  相似文献   

12.
The title compound, [Fe(C8H11ClO2)(CO)3], has been synthesized, isolated and characterized by single‐crystal X‐ray diffraction. The mol­ecule crystallizes in the orthorhombic space group P212121. The metal–ligand arrangement is typical of (1,3‐diene)­tri­carbonyl­iron complexes.  相似文献   

13.
[(Cp4i Rh)2(μ‐Cl)3] [Rh(CO)2Cl2] (Cp4i = tetraisopropyl‐cyclopenta‐dienyl) has been prepared and its crystal is in the space group of Pbar with a= 0.9417 (8), b = 1.4806 (3), c = 1.5062 (2) nm, a = 92.980(10), β = 97.42(3), γ = 93.98 (3)°, V = 2.0735(18) nm3 and Z = 2. The crystal structure consists of a cation of [(η5‐Cp4i) Rh (III)(μ‐Cl)3 Rh (III) (η5‐Cp4i)]+ and an anion of [Rh (I) (CO)2 Cl2]. The two bulky tetraisopropylcyclopentadienyl ligands are in the ecliptic conformation with angle of 10.19° between two cyclopentadienyl ring planes.  相似文献   

14.
《先进技术聚合物》2018,29(1):30-40
Our daily life needs depend on plastics, as they are cheap and durable, so they become the most commonly used synthetic chemical products. But from an environmentalist's point of view, a major concern related to these plastics is their non‐biodegradable nature. Driven by growing demand to search for sustainable solutions to dispose off generating huge volume of synthetic plastic wastes, shifted the mind of researcher towards the use of biodegradable plastics which can be completely disposed‐off by microbial enzymatic degradation. These biodegradable plastics or “bioplastics” are also synthesized by microbes under certain stressed environmental conditions out of which poly(R‐3‐hydroxybutyrate) (PHB) is the most ubiquitous and best known representatives of polyhydroxyalkanoate family. The PHB is most intensively used for the innovative biomedical applications owing to suitable combination of biocompatibility, transport characteristics, and mechanical properties. These challenging aspects of PHB can be used for designing of novel medical devices, in tissue engineering, and for systematic sustained drug delivery. Lots of research reports on PHB degrading enzymes and their producing microorganisms including biochemical aspects are available but in scattered form. So this review highlighted all the relevant information of PHB and PHB‐degrading enzymes starting with basic classification, synthesis, mechanism, and applications that are environment friendly and are of public interest.  相似文献   

15.
Poly(4‐vinylpyridine) was determined to possess conductivity in the experiment. In order to understand properties of the polymer, a series of 4‐vinylpyridine oligomers were designed. The structures of these oligomers were optimized using density function theory (DFT) at B3LYP/6‐31G(d) level. The energy gaps and thermal stabilities of the oligomers were decreased when the chain lengths were increased. These properties were also decreased owing to the protonation of the pyridine ring. The holes were easily injected into the oligomers in the presence of hydrochloride. The electrons were conducted in the side chain composed of the pyridine rings rather than the main chain owing to the saturation of the main chain. The 13C nuclear magnetic resonance (NMR) spectra and nucleus independent chemical shifts (NICS) of these compounds were calculated at B3LYP/6‐31G(d) level. The chemical shifts of the carbon atoms connected with the nitrogen atoms in the protonated pyridines were moved upfield in comparison with those of the pyridines. The addition of hydrochloride on the pyridine ring in the oligomers led to the increase of the aromaticities, namely the aromaticities of the oligomers were obviously improved when the pyridine rings were protonated.  相似文献   

16.
17.
A novel 3D polymeric heteropolynuclear sodium(I) lead(II) complex containing different ligands, [NaPb(ClO4)(en)(NO2)2] was synthesized and characterized by elemental analysis and IR, and 1H‐, 13C‐, and 207Pb‐NMR spectroscopy. The single‐crystal X‐ray data of [NaPb(ClO4)(en)(NO2)2]n established that the complex is a three‐dimensional polymer, [(en)Pb(μ3‐ONO)2Na(μ3‐ONO)2Na(μ‐O2ClO2)Na]n. The Pb and Na atoms have four‐ and eight‐coordinate geometry, respectively. The lone pair of electrons at the PbII atom is ‘stereochemically active’.  相似文献   

18.
A novel triblock copolymer PS–PHB–PS based on the microbial polyester Poly[(R)‐3‐hydroxybutyrate)] (PHB) and poly(styrene) (PS) was prepared to be used as compatibilizer for the corresponding PHB/PS blends. It was prepared in a three‐step procedure consisting of (i) transesterification reaction between ethylene glycol and a high‐molecular‐weight PHB, (ii) synthesis of bromo‐terminated PHB macroinitiator, and (iii) atom transfer radical polymerization polymerization of styrene initiated by the PHB‐based macroinitiator. Fourier transform infrared, gel permeation chromatography, 1H‐, and 13C‐NMR spectroscopies were used to determine the molecular structure and/or end‐group functionalities at each step of the procedure. Although thermogravimetric analysis showed that the block copolymer underwent a stepwise thermal degradation and had better thermal stability than their respective homopolymers, differential scanning calorimetry displayed that the PHB block in the copolymer could not crystallize, and thus generating a total amorphous structure. Atomic force microscopy images indicated that the block copolymer was phase segregated in a well‐defined morphological structure with nanodomain size of ~40 nm. Contact angle measurements proved that the wettability properties of the block copolymer were in between those of the PHB and PS homopolymers. Blends analyzed for their morphology and thermal properties showed good miscibility and had well‐defined morphological features. Polymer blends exhibited lower crystallinity and decreased stiffness which was proportional to the amount of compatibilizer content in the blends. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

19.
Enzymatic degradation of poly[(R)‐3‐hydroxybutyrate] (P(3HB)) film by the poly(hydroxybutyrate) (PHB) depolymerase from Ralstonia picketti T1 was studied in 0.01 M phosphate buffer solution (pH 7.4) at 37 °C by using a quartz crystal microbalance (QCM) technique. Enzymatic degradation of P(3HB) film was quantitatively followed by QCM as a positive frequency shift. While, the amount of depolymerases adsorbed on the film could be evaluated as a negative frequency shift by using a mutant enzyme which had no hydrolytic activity in a catalytic site. The degradation rate increased with enzyme concentration to reach a maximum value at 1.0 μg · mL?1, and then the rate decreased at higher enzyme concentration. This enzyme concentration dependence could be quantitatively explained in terms of a change of coverage of the film surface by the adsorbed enzyme. When the wild‐type enzyme solution in a QCM cell was replaced with the mutant enzyme solution in the middle of the reaction, the degradation rate was reduced markedly, indicating that the wild‐type enzyme adsorbed on the P(3HB) surface is easily substituted by the mutant enzyme in the solution. On the other hand, replacement of the wild‐type enzyme solution with other proteins or buffer solutions did not affect the degradation rate at all, suggesting that the adsorbed enzyme was not desorbed from the film surface. Thus, the adsorbed PHB depolymerase is released from the P(3HB) surface only by interaction with the same depolymerase in solution.

Time courses of frequency changes (ΔF) or weight changes (Δw) observed during enzymatic degradation of P(3HB) film by PHB depolymerase from R. picketti T1 at 37 °C.  相似文献   


20.
The GROMOS96 molecular‐dynamics (MD) program and force field was used to calculate the conformations at 298 K in CHCl3 solution of two hexakis(3‐hydroxyalkanoic acids). One consists of (R)‐3‐hydroxybutanoate (HB) residues only: H−(OCH(Me)−CH2−CO)6−OH ( 1 ). The other one carries the side chains of valine, alanine, and leucine: H−(OCH(CHMe2)CH2−CO−O−CH(Me)−CH2−CO−O−CH(CH2 CHMe2)−CH2−CO)2−OH ( 2 ), with homochiral 3‐hydroxyalkanoate (HA) moieties. In both cases, the conformational equilibria were sampled 2500 times for 25 ns. Other than clusters of arrangements with inter‐residual hydrogen bonding (between the O‐ and C‐terminal OH and COOH groups, and with chain‐bound ester carbonyl O‐atoms; Fig. 6), there are no preferred backbone conformations in which the molecules 1 and 2 spend more than 5% of the time. Specifically, neither the 21‐ nor the 31‐helical conformation of the oligoester backbone (found in stretched fibers, in lamellar crystallites, and in single crystals of polymers PHB and of oligomers OHB) is sampled to any significant extent (Fig. 8 and 9), in spite of the high population, in both oligomers, of the (−)‐synclinal conformation around the C(2)−C(3) bond (angle ϕ2 in Fig. 2). In contrast to β‐oligopeptides, for which strongly preferred secondary structures are found after a few ns, and for which the number of conformations levels off with time, the number of conformational clusters of the corresponding oligoesters found by our force‐field MD calculations increases steadily over the observation time of 25 ns (Fig. 5). Thus, the conclusion from biological and physical‐chemical studies, according to which the PHB chain is extremely flexible, is confirmed by our computational investigation: in CHCl3 solution, the hexakis(3‐hydroxyalkanoate) chain samples its conformational space randomly!  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