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.
New chiral porphyrin imine was synthesized from (S)‐3‐benzyl‐2‐methyl‐4‐phenylbutanal according to dipyrromethane method using trifluoroacetic acid, BF3 etherate, and p‐chloranil. Manganese complex of this chiral porphyrin imine ligand was used as catalyst in the asymmetric epoxidation of styrene derivatives possessing different substituents. Styrene derivatives possessing electron withdrawing groups gave the corresponding chiral epoxides in high yield up to 98% and ee up to 99%. The mechanism for the catalytic asymmetric epoxidation was also discussed based on transfer of oxygen.  相似文献   

2.
(NIn)‐Formyl protective group of tryptophan has been introduced as a base/nucleophile‐labile protective group. It has long been known that a free ‐amino group of the peptide can serve as a nucleophile: an irreversible formyl NIn → NH2 transfer is consistently observed when deformylation is performed last on an otherwise deprotected peptide that possesses free ‐amino group. Obviously, this particular side reaction should be expected any time free amino group is exposed to Trp(For), but, at the best of our knowledge, has never been reported in the course of Boc‐SPPS. In the present communication, we describe a set of appropriately designed model experiments that permitted to detect the title side reaction both in solution and in solid‐phase reactions. We observed intermolecular formyl group transfer with a model compound, Trp(For)‐NH2. Importantly, we also observed this migration on solid support with the rate roughly estimated to be up to 1% of residues per minute. We also observed that the formyl‐group transfer reaction occurred in a sequence‐dependent manner and was suppressed to a non‐detectable level using ‘in situ neutralization’ technique. Because this side reaction is sequence dependent, there might be situations when the rate of the formation of Nα‐formyl termination by‐products is significant. In other cases, the Nα‐For truncated by‐products would not contaminate the final peptide significantly but still could be a source of microheterogeneity. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
Flexible perovskite solar cells (f‐PSCs) have attracted great attention due to their promising commercial prospects. However, the performance of f‐PSCs is generally worse than that of their rigid counterparts. Herein, it is found that the unsatisfactory performance of planar heterojunction (PHJ) f‐PSCs can be attributed to the undesirable morphology of electron transport layer (ETL), which results from the rough surface of the flexible substrate. Precise control over the thickness and morphology of ETL tin dioxide (SnO2) not only reduces the reflectance of the indium tin oxide (ITO) on polyethylene 2,6‐naphthalate (PEN) substrate and enhances photon collection, but also decreases the trap‐state densities of perovskite films and the charge transfer resistance, leading to a great enhancement of device performance. Consequently, the f‐PSCs, with a structure of PEN/ITO/SnO2/perovskite/Spiro‐OMeTAD/Ag, exhibit a power conversion efficiency (PCE) up to 19.51% and a steady output of 19.01%. Furthermore, the f‐PSCs show a robust bending resistance and maintain about 95% of initial PCE after 6000 bending cycles at a bending radius of 8 mm, and they present an outstanding long‐term stability and retain about 90% of the initial performance after >1000 h storage in air (10% relative humidity) without encapsulation.  相似文献   

4.
Two narrow bandgap non‐fullerene acceptors (NBG‐NFAs), namely, COTIC‐4F and SiOTIC‐4F, are designed and synthesized for the fabrication of efficient near‐infrared organic solar cells (OSCs). The chemical structures of the NBG‐NFAs contain a D′‐D‐D′ electron‐rich internal core based on a cyclopentadithiophene (or dithienosilole) (D) and alkoxythienyl (D′) core, end‐capped with the highly electron‐deficient unit 2‐(5,6‐difluoro‐3‐oxo‐2,3‐dihydro‐1H‐inden‐1‐ylidene)malononitrile (A), ultimately providing a A‐D′‐D‐D′‐A molecular configuration that enhances the intramolecular charge transfer characteristics of the excited states. One can thereby reduce the optical bandgap (Egopt) to as low as ≈1.10 eV, one of the smallest values for NFAs reported to date. In bulk‐heterojunction (BHJ) OSCs, NBG‐NFA blends with the polymer donor PTB7‐Th yield power conversion efficiencies (PCE) of up to 9.0%, which is particularly high when compared against a range of NBG BHJ blends. Most significantly, it is found that, despite the small energy loss (Egopt ? eVOC) of 0.52 eV, the PTB7‐Th/NBG‐NFA bulk heterojunction blends can yield short‐circuit current densities of up to 22.8 mA cm?2, suggesting that the design and application of NBG‐NFA materials have substantial potential to further improve the PCE of OSCs.  相似文献   

5.
The quest for new materials is one of the main factors propelling recent advances in organic photovoltaics. Star‐shaped small molecules (SSMs) have been proven promising candidates as perspective donor material due to the increase in numbers of excitation pathways caused by the degeneracy of the lowest unoccupied molecular orbital (LUMO) level. In order to unravel the pathways of the initial photon‐to‐charge conversion, the photovoltaic blends based on three different SSMs with a generic structure of N(phenylene‐nthiophene‐dicyanovinyl‐alkyl)3 (n = 1–3), and [6,6]‐phenyl‐C71‐butyric acid methyl ester (PC71BM) acceptor are investigated by ultrafast photoinduced absorption spectroscopy assisted by density functional theory calculations. It is shown that both electron transfer from SSMs to PC71BM and hole transfer from PC71BM to SSMs are equally significant for generation of long‐lived charges. In contrast, intramolecular (intra‐SSM) charge separation results in geminate recombination and therefore constitutes a loss channel. Overall, up to 60% of long‐lived separated charges are generated at the optimal PC71BM concentrations. The obtained results suggest that further improvement of the SSM‐based solar cells is feasible via optimization of blend morphology and by suppressing the intra‐SSM recombination channel.  相似文献   

6.
7.
A dynamic model of leaf CO2 assimilation was developed as an extension of the canonical steady‐state model, by adding the effects of energy‐dependent non‐photochemical quenching (qE), chloroplast movement, photoinhibition, regulation of enzyme activity in the Calvin cycle, metabolite concentrations, and dynamic CO2 diffusion. The model was calibrated and tested successfully using published measurements of gas exchange and chlorophyll fluorescence on Arabidopsis thaliana ecotype Col‐0 and several photosynthetic mutants and transformants affecting the regulation of Rubisco activity (rca‐2 and rwt43), non‐photochemical quenching (npq4‐1 and npq1‐2), and sucrose synthesis (spsa1). The potential improvements on CO2 assimilation under fluctuating irradiance that can be achieved by removing the kinetic limitations on the regulation of enzyme activities, electron transport, and stomatal conductance were calculated in silico for different scenarios. The model predicted that the rates of activation of enzymes in the Calvin cycle and stomatal opening were the most limiting (up to 17% improvement) and that effects varied with the frequency of fluctuations. On the other hand, relaxation of qE and chloroplast movement had a strong effect on average low‐irradiance CO2 assimilation (up to 10% improvement). Strong synergies among processes were found, such that removing all kinetic limitations simultaneously resulted in improvements of up to 32%.  相似文献   

8.
Nine hundred and forty‐one samples were collected in Sa Keao, Thailand (n = 554) and Banteay Meanchey, Cambodia (n = 387) from July 2014 to January 2015. A total of 667 Escherichia coli isolates (381 isolates from Sa Keao and 286 isolates from Banteay Meanchey) were obtained and examined for antimicrobial susceptibility, class 1 integrons, ESBL genes and horizontal transfer of resistance determinants. Prevalence of E. coli in pig and broiler carcass samples from slaughterhouses and fresh markets was 36–85% in Sa Keao and 11–69% in Banteay Meanchey. The majority of these isolates were multidrug resistant (75.3%). Class 1 integrons were common in both Thai (47%) and Cambodian (62%) isolates, of which four resistance gene cassette arrays including aadA1, dfrA1‐aadA1, dfrA12‐aadA2 and aadA2‐linF were identified. Class 1 integrons in two broiler isolates from Sa Keao (dfrA12‐aadA2) and one broiler isolate from Banteay Meanchey (dfrA1‐aadA1) were horizontally transferable. Sixteen isolates were confirmed to be ESBL‐producing strains with ESBL gene blaCTX‐M‐15, broad spectrum β‐lactamase gene blaTEM‐1 and the AmpC gene blaCMY‐2 being detected. The blaTEM‐1 gene was most prevalent and located on a conjugative plasmid.  相似文献   

9.
The thymidine mimics isocarbostyril nucleosides and difluorophenyl nucleosides were tested as deoxynucleoside kinase substrates using recombinant human cytosolic thymidine kinase (TK1) and deoxycytidine kinase (dCK), and mitochondrial thymidine kinase (TK2) and deoxyguanosine kinase (dGK). The isocarbostyril nucleoside compound 1‐(2‐deoxy‐β‐D‐ribofuranosyl)‐isocarbostyril (EN1) was a poor substrate with all the enzymes. The phosphorylation rates of EN1 with TK1 and TK2 were < 1% relative to Thd, where as the phosphorylation rates for EN1 were 1.4% and 1.1% with dCK and dGK relative to dCyd and dGuo, respectively. The analogue 1‐(2‐deoxy‐β‐D‐ribofuranosyl)‐7‐iodoisocarbostyril (EN2) showed poor relative‐phosphorylation efficiencies (k cat /K m ) with both TK1 and dGK, but not with TK2. The k cat /K m value for EN2 with TK2 was 12.6% relative to that for Thd. Of the difluorophenyl nucleosides, 5‐(1′‐(2′‐deoxy‐β‐D‐ribofuranosyl))‐2,4‐difluorotoluene (JW1) and 1‐(1′‐(2′‐deoxy‐β‐D‐ribofuranosyl))‐2,4‐difluoro‐5‐iodobenzene (JW2) were substrates for TK1 with phosphorylation efficiencies of about 5% relative to that for Thd. Both analogues were considerably more efficient substrates for TK2, with k cat /K m values of 45% relative to that for Thd. 2,5‐Difluoro‐4‐[1‐(2‐deoxy‐β‐L‐ribofuranosyl)]‐aniline (JW5), a L‐nucleoside mimic, was phosphorylated up to 15% as efficiently as deoxycytidine by dCK. These data provide a possible explanation for the previously reported lack of cytotoxicity of the isocarbostyril‐ and difluorophenyl nucleosides, but potential mitochondrial effects of EN2, JW1 and JW2 should be further investigated.  相似文献   

10.
It was recently demonstrated that a bioelectrochemical system (BES) with a redox mediator allowed Pseudomonas putida to perform anoxic metabolism, converting sugar to sugar acids with high yield. However, the low productivity currently limits the application of this technology. To improve productivity, the strain was optimized through improved expression of glucose dehydrogenase (GCD) and gluconate dehydrogenase (GAD). In addition, quantitative real‐time RT‐PCR analysis revealed the intrinsic self‐regulation of GCD and GAD. Utilizing this self‐regulation system, the single overexpression strain (GCD) gave an outstanding performance in the electron transfer rate and 2‐ketogluconic acid (2KGA) productivity. The peak anodic current density, specific glucose uptake rate and 2KGA producing rate were 0.12 mA/cm2, 0.27 ± 0.02 mmol/gCDW/hr and 0.25 ± 0.02 mmol/gCDW/hr, which were 327%, 477%, and 644% of the values of wild‐type P. putida KT2440, respectively. This work demonstrates that expression of periplasmic dehydrogenases involved in electron transfer can significantly improve productivity in the BES.  相似文献   

11.
Aimed at achieving ideal morphology, illuminating morphology–performance relationship, and further improving the power conversion efficiency (PCE) of ternary polymer solar cells (TSCs), a ternary system is designed based on PTB7‐Th:PffBT4T‐2OD:PC71BM in this work. The PffBT4T‐2OD owns large absorption cross section, proper energy levels, and good crystallinity, which enhances exciton generation, charge dissociation and transport and suppresses charge recombination, thus remarkably increasing the short‐circuit current density (J sc) and fill factor (FF). Finally, a notable PCE of 10.72% is obtained for the TSCs with 15% weight ratio of PffBT4T‐2OD. As for the working mechanism, it confirmed the energy transfer from PffBT4T‐2OD to PTB7‐Th, which contributes to the improved exciton generation. And morphology characterization indicates that the devices with 15% PffBT4T‐2OD possess both appropriate domain size (25 nm) and enhanced domain purity. Under this condition, it affords numerous D/A interface for exciton dissociation and good bicontinuous nanostructure for charge transport simultaneously. As a result, the device with 15% PffBT4T‐2OD exhibits improved exciton generation, enhanced charge dissociation possibility, elevated hole mobility and inhibited charge recombination, leading to elevated J sc (19.02 mA cm?2) and FF (72.62%) simultaneously. This work indicates that morphology optimization as well as energy transfer plays a significant role in improving TSC performance.  相似文献   

12.
Photoelectrodes without a p–n junction are often limited in efficiency by charge recombination at semiconductor surfaces and slow charge transfer to electrocatalysts. This study reports that tin oxide (SnOx) layers applied to n‐Si wafers after forming a thin chemically oxidized SiOx layer can passivate the Si surface while producing ≈620 mV photovoltage under 100 mW cm?2 of simulated sunlight. The SnOx layer makes ohmic contacts to Ni, Ir, or Pt films that act as precatalysts for the oxygen‐evolution reaction (OER) in 1.0 m KOH(aq) or 1.0 m H2SO4(aq). Ideal regenerative solar‐to‐O2(g) efficiencies of 4.1% and 3.7%, respectively, are obtained in 1.0 m KOH(aq) with Ni or in 1.0 m H2SO4(aq) with Pt/IrOx layers as OER catalysts. Stable photocurrents for >100 h are obtained for electrodes with patterned catalyst layers in both 1.0 m KOH(aq) and 1.0 m H2SO4(aq).  相似文献   

13.
The biodegradation of the sulfonated azo dyes, Acid Orange 7 (AO7) and Acid Red 88 (AR88), by a bacterial consortium isolated from water and soil samples obtained from sites receiving discharges from textile industries, was evaluated. For a better removal of azo dyes and their biodegradation byproducts, an aerobically operated two‐stage rectangular packed‐bed biofilm reactor (2S‐RPBR) was constructed. Because the consortium's metabolic activity is affected by oxygen, the effect of the interstitial air flow rate QGI on 2S‐RPBR's zonal values of the oxygen mass transfer coefficient kLa was estimated. In the operational conditions probed in the bioreactor, the kLa values varied from 3 to 60 h?1, which roughly correspond to volumetric oxygen transfer rates, dcL/dt, ranging from 20 to 375 mg O2 L?1h?1. Complete biodegradation of azo dyes was attained at loading rates BV,AZ up to 40 mg L?1d?1. At higher BV,AZ values (80 mg L?1 d?1), dye decolorization and biodegradation of the intermediaries 4‐amino‐naphthalenesulphonic acid (4‐ANS) and 1‐amino‐2‐naphthol (1‐A2N) was almost complete. However, a diminution in COD and TOC removal efficiencies was observed in correspondence to the 4‐aminobenzenesulfonic acid (4‐ABS) accumulation in the bioreactor. Although the oxygen transport rate improved the azo dye mineralization, the results suggest that the removal efficiency of azo dyes was affected by biofilm detachment at relatively high QGI and BV,AZ values. After 225 days of continuous operation of the 2S‐RFBR, eight bacterial strains were isolated from the biofilm attached to the porous support. The identified genera were: Arthrobacter, Variovorax, Agrococcus, Sphingomonas, Sphingopyxis, Methylobacterium, Mesorhizobium, and Microbacterium.  相似文献   

14.
Mycorrhiza formation represents a significant carbon (C) acquisition alternative for orchid species, particularly those that remain achlorophyllous through all life stages. As it is known that orchid mycorrhizas facilitate nutrient transfer (most notably of C), it has not been resolved if C transfer occurs only after lysis of mycorrhizal structures (fungal pelotons) or also across the mycorrhizal interface of pre‐lysed pelotons. We used high‐resolution secondary ion mass spectrometry (nanoSIMS) and labelling with enriched 13CO2 to trace C transfers, at subcellular scale, across mycorrhizal interfaces formed by Rhizanthella gardneri, an achlorphyllous orchid. Carbon was successfully traced in to the fungal portion of orchid mycorrhizas. However, we did not detect C movement across intact mycorrhizal interfaces up to 216 h post 13CO2 labelling. Our findings provide support for the hypothesis that C transfer from the mycorrhizal fungus to orchid, at least for R. gardneri, likely occurs after lysis of the fungal peloton.  相似文献   

15.
The limits of maximizing the open‐circuit voltage Voc in solar cells based on poly[2,7‐(9,9‐didecylfluorene)‐alt‐5,5‐(4,7‐di‐2‐thienyl‐2,1,3‐benzothiadiazole)] (PF10TBT) as a donor using different fullerene derivatives as acceptor are investigated. Bulk heterojunction solar cells with PF10TBT and [6,6]‐phenyl‐C61‐butyric acid methyl ester (PCBM) give a Voc over 1 V and a power conversion efficiency of 4.2%. Devices in which PF10TBT is blended with fullerene bisadduct derivatives give an even higher Voc, but also a strong decrease in short circuit current (Jsc). The higher Voc is attributed to the higher LUMO of the acceptors in comparison to PCBM. By investigating the photophysics of PF10TBT:fullerene blends using near‐IR photo‐ and electroluminescence, time‐resolved photoluminescence, and photoinduced absorption we find that the charge transfer (CT) state is not formed efficiently when using fullerene bisadducts. Hence, engineering acceptor materials with a LUMO level that is as high as possible can increase Voc, but will only provide a higher power conversion efficiency, when the quantum efficiency for charge transfer is preserved. To quantify this, we determine the CT energy (ECT) and optical band gap (Eg), defined as the lowest first singlet state energy ES1 of either the donor or acceptor, for each of the blends and find a clear correlation between the free energy for photoinduced electron transfer and Jsc. We find that Eg ? qVoc > 0.6 eV is a simple, but general criterion for efficient charge generation in donor‐acceptor blends.  相似文献   

16.
Immobilized Cu(OAc)2‐bis(oxazolines) via hydrogen bonding by SBA‐15 was applied to asymmetric Henry reaction, and good enantioselectivities were obtained (up to 83% ee) between 2‐methoxybenzaldehyde and CH3NO2 in isopropyl alcohol (iPrOH). The catalyst could be reused seven times without any obvious loss in enantioselectivity. For the first time, this facile and clean immobilization method is applied to the use of bis(oxazolines) complexes. Chirality 24:1092‐1095, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
Acclimation of respiration to the light environments is important for a plant’s carbon balance. Respiratory rates of mature leaves of Alocasia odora, a typical shade‐tolerant species, were measured during the night for 14 d after reciprocal transfers between high‐ (330 µ mol m?2 s?1) and low‐light (20 µ mol m?2 s?1) environments. Following the transfer, both the rate of CO2 efflux and that of O2 uptake of A. odora leaves adjusted to the new light environments. The O2‐uptake rates changed more slowly than the CO2‐efflux rates under the new environments. Leaf mass per area also changed after the transfer. We analysed whether substrate availability or ATP‐consumption rates influence the respiratory acclimation. Since the addition of sucrose to leaf segments did not influence the O2‐uptake rates, the change of respiratory substrate availability was not responsible for the respiratory acclimation. The addition of an uncoupler induced increases in the O2‐uptake rates, and the degree of enhancement significantly decreased after the transfer from low to high irradiance. Thus, the change in ATP‐consumption rates was responsible for the changes in respiratory rates in the plants transferred from low to high light. Potential rates of O2 uptake, as measured in the presence of both the substrate and the uncoupler, changed after the transfer, and strongly correlated with the O2‐uptake rates, irrespective of the directions of transfer (r = 0·961). There was a strong correlation between maximal activities of NAD‐isocitrate dehydrogenase and the potential rates of O2 uptake (r = 0·933), but a weaker correlation between those of cytochrome c oxidase and the potential rates (r = 0·689). These data indicate that the changes of light environments altered the respiratory rates via the change of the respiratory ATP demand, and that the altered rates of respiration will induce the changes of the respiratory capacities.  相似文献   

18.
Triplet exciton formation in neat 7,7‐(4,4‐bis(2‐ethylhexyl)‐4H‐silolo[3,2‐b:4,5‐b′] dithiophene‐2,6‐diyl)bis(6‐fluoro‐4‐(5′‐hexyl‐[2,2′‐bithiophen]‐5‐yl)benzo[c][1,2,5]thiadiazole) (p‐DTS(FBTTh2)2) and blends with [6,6]‐Phenyl C70 butyric acid methyl ester (PC70BM), with and without the selective solvent additive 1,8‐diiodooctane, is investigated by means of spin sensitive photoluminescence measurements. For all three material systems, a significant amount of long living triplet excitons is detected, situated on the p‐DTS(FBTTh2)2 molecules. The characteristic zero‐field splitting parameters for this state are identified to be D = 42 mT (1177 MHz) and E = 5 mT (140 MHz). However, no triplet excitons located on PC70BM are detectable. Using electrically detected spin resonance, the presence of these triplet excitons is confirmed even at room temperature, highlighting that triplet excitons form during solar cell operation and influence the photocurrent and photovoltage. Surprisingly, the superior performing blend is found to have the largest triplet population. It is concluded, that the formation of triplet excitons from charge transfer states via electron back transfer has no crucial impact on device performance in p‐DTS(FBTTh2)2:PC70BM based solar cells.  相似文献   

19.
Hierarchical Cu2S microspheres wrapped by reduced graphene oxide (RGO) nanosheets are prepared via a one‐step solvothermal process. The amount of graphene oxide used in the synthesis process has a remarkable effect on the features of Cu2S microspheres. Compared to Pt and Cu2S electrodes, RGO‐Cu2S electrodes show better electrocatalytic activity, greater stability, lower charge‐transfer resistance, and higher exchange current density. As expected, RGO‐Cu2S electrodes exhibit superior performance when functioning as counter electrodes in CdS/CdSe quantum dot‐sensitized solar cells (QDSSCs) using a polysulfide electrolyte. A power conversion efficiency up to 3.85% is achieved for the QDSSC employing an optimized RGO‐Cu2S counter electrode, which is higher than those of the QDSSCs featuring Pt (2.14%) and Cu2S (3.39%) counter electrodes.  相似文献   

20.
New chiral amidophosphite ligand was synthesized and tested in the Rh‐catalyzed asymmetric hydrogenation of (Z)‐β‐(acylamino)acrylates in protic solvents and supercritical carbon dioxide (scCO2) The catalytic performance is affected greatly by the acidity of the solvents. Better enantioselectivity (up to 88% ee) was achieved in scCO2 containing 1,1,1,3,3,3‐hexafluoro‐2‐propanol, compared to neat protic solvents. Chirality, 2011. © 2011 Wiley‐Liss, Inc.  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