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.
DesII is a radical S‐adenosylmethionine (SAM) enzyme that catalyzes the C4‐deamination of TDP‐4‐amino‐4,6‐dideoxyglucose through a C3 radical intermediate. However, if the C4 amino group is replaced with a hydroxy group (to give TDP‐quinovose), the hydroxy group at C3 is oxidized to a ketone with no C4‐dehydration. It is hypothesized that hyperconjugation between the C4 C? N/O bond and the partially filled p orbital at C3 of the radical intermediate modulates the degree to which elimination competes with dehydrogenation. To investigate this hypothesis, the reaction of DesII with the C4‐epimer of TDP‐quinovose (TDP‐fucose) was examined. The reaction primarily results in the formation of TDP‐6‐deoxygulose and likely regeneration of TDP‐fucose. The remainder of the substrate radical partitions roughly equally between C3‐dehydrogenation and C4‐dehydration. Thus, changing the stereochemistry at C4 permits a more balanced competition between elimination and dehydrogenation.  相似文献   

2.
3.
The synthesis of linear and branched oligothiophenes of well‐defined structures is performed with regioselective deprotonation of 3‐substituted thiophenes and nickel‐catalyzed cross‐coupling of the thus formed metalated species with a bromothiophene. The reaction of 3‐hexylthiophene with EtMgCl and 2,2,6,6‐tetramethylpiperidine (TMP‐H, 10 mol %) induces the metalation selectively at the 5‐position by use of the catalytically generated hindered magnesium amide (TMPMgCl) and the subsequent reaction of a 2‐halo‐3‐hexylthiophene (bromide or chloride) in the presence of a nickel catalyst affords a head‐to‐tail (HT)‐type dimer. By repeating the same sequence, the linear oligothiophene up to a 4‐mer is synthesized in good yield. The reaction of 3‐hexylthiophene with 2,3‐dibromothiophene also takes place to afford a branched oligothiophene 3‐mer in quantitative yield. The obtained 3‐mer is also metalated at the sterically less‐hindered position in a regioselective manner furnishing a 7‐mer in >99 % yield after a further coupling reaction with 2,3‐dibromothiophene. These dendrimers react with several multifunctionalized organic electrophiles, leading to a variety of branched oligothiophenes.  相似文献   

4.
The highly selective tuning of the isomerization from 1‐arylalka‐1,2‐dien‐1‐yllithium to 1‐arylalka‐1,2‐dien‐3‐yllithium has been realized in the deprotonation of 1‐arylalk‐1‐yne (conditions A and B) and carbolithiation of 1‐arylbut‐3‐en‐1‐yne with alkyllithium (conditions C and D). Subsequent transmetallation and Pd‐catalyzed Negishi coupling reactions afforded 1,1‐diaryl or 1,3‐diaryl allenes with high selectivity. Deuterium‐labeling cross experiments indicated that an intermolecular lithiation process occurred in both 1,3‐lithium shift conditions (conditions B and D). 1‐Arylalka‐1,2‐diene was confirmed experimentally to be the intermediate. A computational study at the B3LYP level for the isomerization indicated that the acidity of H at the 3‐position is higher than that of the H at the 1‐position of 1‐phenyl‐1,2‐butadiene. Under conditions B, iPr2NH acts as a proton carrier to finish the 1,3‐lithium shift. The overall activation barrier for the rate‐determining step in the solvated models is ≈21.0 kcal mol?1, indicating that the isomerization is reasonable at room temperature. For the isomerization under conditions D, DFT calculations indicated that the addition of TMEDA (tetramethylethylenediamine) and HMPA (hexamethylphosphoramide) changes the global minimum of the system; among the possible mechanisms (P1–P5) considered, the mechanism catalyzed by dilithiated species (P5) is the most probable one. The overall activation barriers for isomerization in THF and TMEDA solvated models are 22.6 and 19.7 kcal mol?1, respectively, proving that the isomerization may proceed at RT in THF or at ?78 °C with TMEDA, due to the fact that the solvation of the additives may increase the concentration of 1‐phenyl‐1,2‐butadienyllithium monomer by a deaggregation effect.  相似文献   

5.
Potassium‐ion batteries are promising for low‐cost and large‐scale energy storage applications, but the major obstacle to their application is the lack of safe and effective electrolytes. A phosphate‐based fire retardant such as triethyl phosphate is now shown to work as a single solvent with potassium bis(fluorosulfonyl)imide at 0.9 m , in contrast to previous Li and Na systems where phosphates cannot work at low concentrations. This electrolyte is optimized at 2 m , where it exhibits the advantages of low cost, low viscosity, and high conductivity, as well as the formation of a uniform and robust salt‐derived solid‐electrolyte interphase layer, leading to non‐dendritic K‐metal plating/stripping with Coulombic efficiency of 99.6 % and a highly reversible graphite anode.  相似文献   

6.
The first preparation of acridin‐9(10H)‐ones carrying a tertiary thiocarbamoyl group at C(10), i.e., N,N‐dialkyl‐9‐oxoacridine‐10(9H)‐carbothioamides 9 , is described. The method is based on the reaction of (2‐halophenyl)(2‐isothiocyanatophenyl)methanones 7 , prepared from (2‐aminophenyl)(2‐halophenyl)methanones 5 by a convenient three‐step sequence, with secondary amines in DMF at room temperature to generate the corresponding thiourea derivatives 8 in situ, which are treated with NaH at 100–120° to provide the desired products in one‐pot reactions in generally good yields.  相似文献   

7.
The order in molecular monolayers is a crucial aspect for their technological application. However, the preparation of defined monolayers by spin‐coating is a challenge, since the involved processes are far from thermodynamic equilibrium. In the work reported herein, the dynamic formation of dioctyl‐benzothienobenzothiophene monolayers is explored as a function of temperature by using X‐ray scattering techniques and atomic force microscopy. Starting with a disordered monolayer after the spin‐coating process, post‐deposition self‐reassembly at room temperature transforms the initially amorphous layer into a well‐ordered bilayer structure with a molecular herringbone packing, whereas at elevated temperature the formation of crystalline islands occurs. At the temperature of the liquid‐crystalline crystal–smectic transition, rewetting of the surface follows resulting in a complete homogeneous monolayer. By subsequent controlled cooling to room temperature, cooling‐rate‐dependent kinetics is observed; at rapid cooling, a stable monolayer is preserved at room temperature, whereas slow cooling causes bilayer structures. Increasing the understanding and control of monolayer formation is of high relevance for achieving ordered functional monolayers with defined two‐dimensional packing, for future applications in the field of organic electronics.  相似文献   

8.
Radical ring‐opening polymerization of cyclic ketene acetals is a means to achieve novel types of aliphatic polyesters. 2‐methylene‐1,3‐dioxe‐5‐pene is a seven‐membered cyclic ketene acetal containing an unsaturation in the 5‐position in the ring structure. The double bond functionality enables further reactions subsequent to polymerization. The monomer 2‐methylene‐1,3‐dioxe‐5‐pene was synthesized and polymerized in bulk by free radical polymerization at different temperatures, to determine the structure of the products and propose a reaction mechanism. The reaction mechanism is dependent on the reaction temperature. At higher temperatures, ring‐opening takes place to a great extent followed by a new cyclization process to form the stable five‐membered cyclic ester 3‐vinyl‐1,4‐butyrolactone as the main reaction product. Thereby, propagation is suppressed and only small amounts of other oligomeric products are formed. At lower temperatures, the cyclic ester formation is reduced and oligomeric products containing both ring‐opened and ring‐retained repeating units are produced at higher yield. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 4587–4601, 2009  相似文献   

9.
The β‐H‐elimination in the (IPr)AuEt complex and its microscopic reverse, the insertion of ethene into (IPr)AuH, were investigated in a combined experimental and computational study. Our DFT‐D3 calculations predict free‐energy barriers of 49.7 and 36.4 kcal mol?1 for the elimination and insertion process, respectively, which permit an estimation of the rate constants for these reactions according to classical transition‐state theory. The elimination/insertion pathway is found to involve a high‐energy ethene hydride species and is not significantly affected by continuum solvent effects. The high barriers found in the theoretical study were then confirmed experimentally by measuring decomposition temperatures for several different (IPr)AuI‐alkyl complexes which, with a slow decomposition at 180 °C, are significantly higher than those of other transition‐metal alkyl complexes. In addition, at the same temperature, the decomposition of (IPr)AuPh and (IPr)AuMe, both of which cannot undergo β‐H‐elimination, indicates that the pathway for the observed decomposition at 180 °C is not a β‐H‐elimination. According to the calculations, the latter should not occur at temperatures below 200 °C. The microscopic reverse of the β‐H‐elimination, the insertion of ethene into the (IPr)AuH could neither be observed at pressures up to 8 bar at RT nor at 1 bar at 80 °C. The same is true for the strain‐activated norbornene.  相似文献   

10.
Strain developed in crystals in response to stimuli causes mechanical response. Methods to tune such mechanical response is important for practical applications. Crystals of a monomer having azide and alkyne units pre‐organized in a ready‐to‐react orientation, undergo thermal topochemical dimerization and show rate‐dependent mechanical response. When the reaction rate is fast, the crystals explode violently. When the reaction rate is slow, the crystals absorb water from the surroundings contemporaneously with the reaction to form the dimer‐hydrate in a single‐crystal‐to‐single‐crystal (SCSC) manner. Crystals of the dimer‐hydrate upon dehydration also undergo explosion. Thus, at slow reaction rate, the strain gets stored in crystals by hydration and the explosion can be harvested, at will, by dehydration. Use of this rate‐dependent explosion in the automatic activation of a remedial electrical circuit in case of a sudden rise in temperature has been demonstrated.  相似文献   

11.
We investigate the cause of amplification of light‐energy conversion when coupling a nc‐TiO2 film to a TiO2 inverse opal by comparing it to an inverse TiO2 glass (i‐TiO2‐g) fabricated with the exact monodisperse air–hole size as an inverse opal with a stop band at 600 nm (600‐i‐TiO2‐o). A significant twofold average gain in the photon‐to‐current conversion efficiency is measured to the red of the stop band at the 600‐i‐TiO2‐o/nc‐TiO2 bilayer under front‐wall and back‐wall illumination, greater than the gain within the stop band. A smaller amplification is measured under front‐wall illumination—and no gain is measured under back‐wall illumination—for i‐TiO2‐g/nc‐TiO2 at these energies. The photonic crystal therefore causes trapping of light through the bilayer, not only within the gap but also to the red, at frequencies within its dielectric band. This light‐trapping effect is found to be dependent on structural order, as a highly disordered inverse glass film with the same air–hole size and thickness does not yield the same gain. A drop in the transmission of light is measured within the same frequencies to the red of the stop band upon adding nc‐TiO2 to 600‐i‐TiO2‐o, consistent with light trapping in the bilayer.  相似文献   

12.
4‐Amino‐3,5‐dinitropyrazole (LLM‐116, 1 ) undergoes trimerization to 4‐diazo‐3,5‐bis(4‐amino‐3,5‐dinitropyrazol‐1‐yl) pyrazole (LLM‐226) upon heating. A mechanism is proposed and discussed. LLM‐226 is a new diazo‐based energetic material, thermally stable, and insensitive to impact, friction, and spark. The material may be prepared by heating 1 and 4‐diazo‐3,5‐dinitropyrazole ( 2 ) in a mixture of toluene and butyl acetate at 110°C or heating 1 alone in dichlorobenzene at 160°C. The characterization of LLM‐226 including X‐ray crystallographic analysis and small‐scale safety properties will be discussed.  相似文献   

13.
14.
Stem‐cell behavior is regulated by the material properties of the surrounding extracellular matrix, which has important implications for the design of tissue‐engineering scaffolds. However, our understanding of the material properties of stem‐cell scaffolds is limited to nanoscopic‐to‐macroscopic length scales. Herein, a solid‐state NMR approach is presented that provides atomic‐scale information on complex stem‐cell substrates at near physiological conditions and at natural isotope abundance. Using self‐assembled peptidic scaffolds designed for nervous‐tissue regeneration, we show at atomic scale how scaffold‐assembly degree, mechanics, and homogeneity correlate with favorable stem cell behavior. Integration of solid‐state NMR data with molecular dynamics simulations reveals a highly ordered fibrillar structure as the most favorable stem‐cell scaffold. This could improve the design of tissue‐engineering scaffolds and other self‐assembled biomaterials.  相似文献   

15.
Franka Kálmán 《Electrophoresis》2016,37(22):2913-2921
3‐(2‐furoyl)quinoline‐2‐carboxaldehyde (FQ) is a sensitive fluorogenic dye, used for derivatization of proteins for SDS‐CGE with LIF detection (SDS‐CGE‐LIF) at silver staining sensitivity (ng/mL). FQ labels proteins at primary amines, found at lysines and N‐termini, which vary in number and accessibility for different proteins. This work investigates the accuracy of estimation of protein concentration with SDS‐CGE‐LIF in real biological samples, where a different protein must be used as a standard. Sixteen purified proteins varying in molecular weight, structure, and sequence were labeled with FQ at constant mass concentration applying a commonly used procedure for SDS‐CGE‐LIF. The fluorescence of these proteins was measured using a spectrofluorometer and found to vary with a RSD of 36%. This compares favorably with other less sensitive methods for estimation of protein concentration such as SDS‐CGE‐UV and SDS‐PAGE‐Coomassie and is vastly superior to the equivalently sensitive silver stain. Investigation into the number of labels bound with UHPLC‐ESI‐QTOF‐MS revealed large variations in the labeling efficiency (percentage of labels to the number of labeling sites given by the sequence) for different proteins (from 3 to 30%). This explains the observation that fluorescence per mole of protein was not proportional to the number of lysines in the sequence.  相似文献   

16.
The low‐bandgap polymer poly{[4,4‐bis(2‐ethylhexyl)‐cyclopenta‐(2,1‐b;3,4‐b′)dithiophen]‐2,6‐diyl‐alt‐(2,1,3‐benzo‐thiadiazole)?4,7‐diyl} (PCPDTBT) is widely used for organic solar cell applications. Here, we present a comprehensive study of the optical properties as a function of temperature for PCPDTBT in solution and in thin films with two distinct morphologies. Using absorption and photoluminescence spectroscopy as well as Franck‐Condon analyses, we show that PCPDTBT in solution undergoes a phase transformation at 300 K from a disordered to a truly aggregated state on cooling. The saturation value of aggregates in solution is reached in PCPDTBT thin films at any temperature. In addition, we demonstrate that the photophysical properties of the aggregates in films are similar to those in solution and that a low percentage of thermally activated excimer states is present in the films at temperatures above 200 K. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2015 , 53, 1416–1430  相似文献   

17.
Drying is a critical step to prolong the storage time in natural medicine processing but it changes the chemical characteristics of the product. In this study, research was performed to characterize the metabolomic changes in toad venom induced by vacuum‐drying at 60°C and air‐drying at room temperature by ultra high performance liquid chromatography coupled with pattern recognition approaches. In total 52 metabolites, down‐regulated or up‐regulated, were identified as potential chemical markers. Compared with fresh toad venom, vacuum‐drying at 60°C succeeded in raising the conjugated‐type bufadienolide content significantly, while the content of free‐type bufadienolides were slightly reduced. On the other hand, toad venom air‐dried at room temperature presented a relatively low amount of bufadienolides compared with fresh venom. For example, the content of several known anti‐tumor components (gamabufotalin, bufotalin, cinobufagin, etc.) were significantly reduced. The 3‐(4,5‐dimethylthiazol‐2‐yl)‐2,5‐diphenyltetrazolium bromide bioassay further showed that venom air‐dried at room temperature had weaker anti‐tumor activity on human hepatocellular carcinoma SMMC‐7721 proliferation in vitro than samples vacuum‐dried at 60°C. These results showed that the great metabolomic changes of toad venom occurred during the drying process, suggesting that a proper drying procedure is important for sustaining the chemical quality of natural medicines.  相似文献   

18.
The ferrocenyl‐nucleoside, 5‐ethynylferrocenyl‐2′‐deoxycytidine ( 1 ) has been prepared by Pd‐catalyzed cross‐coupling between ethynylferrocene and 5‐iodo‐2′‐deoxycytidine and incorporated into oligonucleotides by using automated solid‐phase synthesis at both silica supports (CPG) and modified single‐crystal silicon electrodes. Analysis of DNA oligonucleotides prepared and cleaved from conventional solid supports confirms that the ferrocenyl‐nucleoside remains intact during synthesis and deprotection and that the resulting strands may be oxidised and reduced in a chemically reversible manner. Melting curve data show that the ferrocenyl‐modified oligonucleotides form duplex structures with native complementary strands. The redox potential of fully solvated ferrocenyl 12‐mers, 350 mV versus SCE, was shifted by +40 mV to a more positive potential upon treatment with the complement contrary to the anticipated negative shift based on a simple electrostatic basis. Automated solid‐phase methods were also used to synthesise 12‐mer ferrocenyl‐containing oligonucleotides directly at chemically modified silicon <111> electrodes. Hybridisation to the surface‐bound ferrocenyl‐DNA caused a shift in the reduction potential of +34 mV to more positive values, indicating that, even when a ferrocenyl nucleoside is contained in a film, the increased density of anions from the phosphate backbone of the complement is still dominated by other factors, for example, the hydrophobic environment of the ferrocene moiety in the duplex or changes in the ferrocene–phosphate distances. The reduction potential is shifted >100 mV after hybridisation when the aqueous electrolyte is replaced by THF/LiClO4, a solvent of much lower dielectric constant; this is consistent with an explanation based on conformation‐induced changes in ferrocene–phosphate distances.  相似文献   

19.
A new matrix system for phosphorescent organic light‐emitting diodes (OLEDs) based on an electron transporting component attached to an inert polymer backbone, an electronically neutral co‐host, and a phosphorescent dye that serves as both emitter and hole conductor are presented. The inert co‐host is used either as small molecules or covalently connected to the same chain as the electron‐transporting host. The use of a small molecular inert co‐host in the active layer is shown to be highly advantageous in comparison to a purely polymeric matrix bearing the same functionalities. Analysis of the dye phosphorescence decay in pure polymer, small molecular co‐host film, and their blend lets to conclude that dye molecules distribute mostly in the small molecular co‐host phase, where the co‐host prevents agglomeration and self‐quenching of the phosphorescence as well as energy transfer to the electron transporting units. In addition, the co‐host accumulates at the anode interface where it acts as electron blocking layer and improves hole injection. This favorable phase separation between polymeric and small molecular components results in devices with efficiencies of about 47 cd/A at a luminance of 1000 cd/m2. Investigation of OLED degradation demonstrates the presence of two time regimes: one fast component that leads to a strong decrease at short times followed by a slower decrease at longer times. Unlike the long time degradation, the efficiency loss that occurs at short times is reversible and can be recovered by annealing of the device at 180 °C. We also show that the long‐time degradation must be related to a change of the optical and electrical bulk properties. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2013  相似文献   

20.
We report a lipid‐based strategy to visualize Golgi structure and dynamics at super‐resolution in live cells. The method is based on two novel reagents: a trans‐cyclooctene‐containing ceramide lipid (Cer‐TCO) and a highly reactive, tetrazine‐tagged near‐IR dye (SiR‐Tz). These reagents assemble via an extremely rapid “tetrazine‐click” reaction into Cer‐SiR, a highly photostable “vital dye” that enables prolonged live‐cell imaging of the Golgi apparatus by 3D confocal and STED microscopy. Cer‐SiR is nontoxic at concentrations as high as 2 μM and does not perturb the mobility of Golgi‐resident enzymes or the traffic of cargo from the endoplasmic reticulum through the Golgi and to the plasma membrane.  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