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.
Goethite–titania (α-FeOOH–TiO2) composites were prepared by co-precipitation and mechanical milling. The structural, morphological and optical properties of as-synthesized composites were characterized by X-ray powder diffraction, scanning electron microscopy and UV-Vis diffuse reflectance spectroscopy, respectively. α-FeOOH–TiO2 composites and TiO2-P25, as reference, were evaluated as photocatalysts for the disinfection of Escherichia coli under UV or visible light in a stirred tank reactor. α-FeOOH–TiO2 exhibited better photocatalytic activity in the visible region than TiO2-P25. The mechanical activation increased the absorption in the visible range of TiO2-P25 and the photocatalytic activity of α-FeOOH–TiO2. In the experiments with UV light and α-FeOOH–TiO2, mechanically activated, a 5.4 log-reduction of bacteria was achieved after 240 min of treatment. Using visible light the α-FeOOH–TiO2 and the TiO2-P25 showed a 3.1 and a 0.7 log-reductions at 240 min, respectively. The disinfection mechanism was studied by ROS detection and scavenger experiments, demonstrating that the main ROS produced in the disinfection process were superoxide radical anion, singlet oxygen and hydroxyl radical.

A photocatalytic mechanism for FeOOH–TiO2 composite is proposed under UV-Vis light, the FeOOH–TiO2 composite showed higher photocatalytic activity than TiO2-P25.  相似文献   

2.
Coating-free TiO2@β-SiC photocatalytic composite foams gathered within a ready-to-use shell/core alveolar medium the photocatalytically active TiO2 phase and the β-SiC foam structure were prepared via a multi-step shape memory synthesis (SMS) replica method. They were fabricated following a sequential two-step carburization approach, in which an external TiC skin was synthesized at the surface of a β-SiC skeleton foam obtained from a pre-shaped polyurethane foam during a first carburization step. The adsorption behaviour of the shell/core TiO2@β-SiC composite foams towards the Diuron pollutant in water was tuned by submitting the carbide foams to a final calcination treatment within the 550–700 °C temperature range. The controlled calcination step allowed (i) the selective oxidation of the TiC shell into a TiO2 crystallite shell owing to the β-SiC resistance to oxidation and (ii) the amount of residual unreacted carbon in the foams to be tuned. The lower the calcination temperature, the more pronounced the adsorption profiles of the composites and the higher the Diuron amount removed by adsorption on the residual unreacted carbon. The ready-to-use TiO2@β-SiC composite foams were active in the degradation of the Diuron pesticide, without any further post-synthesis immobilization or synthesis process at the foam surface. They displayed good reusability with test cycles and benefitted from an enhanced stability in terms of the titania release to water.

Coating-free TiO2@β-SiC photocatalytic composite foams gathering within a ready-to-use shell/core alveolar medium the TiO2 photocatalyst and the β-SiC foam structure were prepared via a multi-step shape memory synthesis (SMS) replica method.  相似文献   

3.
One-pot green synthesis of propargylamines using ZnCl2 loaded TiO2 nanomaterial under solvent-free conditions has been effectively accomplished. The aromatic aldehydes, amines, and phenylacetylene were reacted at 100 °C in the presence of the resultant catalyst to form propargylamines. The nanocrystalline TiO2 was initially synthesized by a sol–gel method from titanium(iv) isopropoxide (TTIP) and further subjected to ZnCl2 loading by a wet impregnation method. X-ray diffraction (XRD) patterns revealed the formation of crystalline anatase phase TiO2. Field emission scanning electron microscopy (FESEM) showed the formation of agglomerated spheroid shaped particles having a size in the range of 25–45 nm. Transmission electron microscopy (TEM) validates cubical faceted and nanospheroid-like morphological features with clear faceted edges for the pure TiO2 sample. Surface loading of ZnCl2 on spheroid TiO2 nanoparticles is evident in the case of the ZnCl2 loaded TiO2 sample. X-ray photoelectron spectroscopy (XPS) confirmed the presence of Ti4+ and Zn2+ species in the ZnCl2 loaded TiO2 catalyst. Energy-dispersive X-ray (EDS) spectroscopy also confirmed the existence of Ti, O, Zn and Cl elements in the nanostructured catalyst. 15% ZnCl2 loaded TiO2 afforded the highest 97% yield for 3-(1-morpholino-3-phenylprop-2-ynyl)phenol, 2-(1-morpholino-3-phenylprop-2-ynyl)phenol and 4-(1,3-diphenylprop-2-ynyl)morpholine under solvent-free and aerobic conditions. The proposed nanostructure-based heterogeneous catalytic reaction protocol is sustainable, environment-friendly and offers economic viability in terms of recyclability of the catalyst.

A one-pot green synthesis of propargylamines using nanostructured ZnCl2–TiO2 under solvent-free conditions has been effectively accomplished. The proposed reaction protocol is sustainable, environment-friendly and offers economic viability.  相似文献   

4.
The first use of biomass-derived HMF in the one-pot Kabachnik–Fields reaction is reported here. A wide range of furan-based α-amino phosphonates were prepared in moderate to excellent yields under mild, effective and environmentally-benign conditions: iodine as a non-metal catalyst, biobased 2-MeTHF as the solvent and room or moderate temperature. The hydroxymethyl group of HMF persists in the Kabachnik–Fields products, widening the scope of further modification and derivatization compared to those arising from furfural. Issues involving the diastereoselectivity and double Kabachnik–Fields condensation were also faced.

A mild and efficient one-pot protocol for the synthesis of α-amino phosphonates directly from 5-HMF was described.

Recently, the production of chemicals from renewable biomass has attracted growing interests due to the dwindling reserves of fossil resources and the increasing awareness of environmental concerns.1 5-Hydroxymethylfurfural (HMF), a promising primary biomass-derived platform chemical readily obtained from acid-catalyzed dehydration of six-carbon carbohydrates, displays a strong potential in organic synthesis.2 Besides the well-developed conversions of HMF towards monomers and biofuels via oxidation or reduction reactions,3 some remarkable strategies converting HMF to high value-added fine chemicals have been disclosed.4 Nevertheless, the specific reactivity and reduced stability of HMF, in comparison with the pentose-derived furfural homolog, have limited its use in synthetic applications.5 Furthermore, its commercial availability, though not anymore a barrier nowadays, has limited the number of investigations in the past. In this regard, developing efficient and economic routes to existing or novel fine chemicals from HMF, with its different reactivity compared to simpler aldehydes, is still a challenge.Multi-component reactions (MCRs) are extremely convenient and efficient strategies to prepare highly functionalized compounds from simple starting materials by one-pot procedures. Since they have many advantages, such as high atom economy, high convergence, time and energy saving, MCRs have gained much attention in modern synthetic organic chemistry.6 Nevertheless, to our knowledge, the direct utilization of HMF in MCR processes has been only rarely explored.α-Amino phosphonates, due to the structural analogy to natural α-amino acids and their significant biological activities, such as antitumor, antitubercular, cytotoxic activities, and so on,7 have been the subject of considerable interest in the past decades both in synthetic organic and medicinal chemistry.8 Among several methods for preparing α-amino phosphonates, the Kabachnik–Fields reaction, a one-pot condensation of an aldehyde, an amine and a dialkyl phosphite is the most effective and convenient strategy.9 A large number of conditions have been reported for the acid-catalyzed (Lewis/Brønsted)10 and catalyst-free11 Kabachnik–Fields reaction, affording the α-amino phosphonates from various aldehydes. However, none of studies have included HMF as a substrate in their scope although the products from HMF can offer more possibilities for further functionalization, thanks to its CH2OH appendage. The sole synthesis of α-amino phosphonates from HMF was reported by Cottier and Skowroński in a two-step reaction strategy, consisting in pre-formation of the imine which upon isolation reacted with dialkyl phosphites as nucleophilic species at high temperature using trifluoroacetic acid as catalyst.12 Thus, a milder and more direct procedure for the synthesis of α-amino phosphonates from HMF is still to be developed.As part of our on-going interest on the application of HMF towards fine chemicals and on green and sustainable chemistry,13 we explored the possibility to synthesize furan-based α-amino phosphonates via the one-pot Kabachnik–Fields condensation, directly from HMF.For this study, we selected molecular iodine as a mild and effective Lewis acid catalyst, as often used in multicomponent synthesis because of its operational simplicity, low cost and toxicity and likely to be compatible with HMF sensitivity to acidic conditions.14 Wu and co-workers have confirmed its efficiency in Kabachnik–Fields reactions of simple aldehydes such as benzaldehyde and furfural.15 The primary set of experimental conditions has been fixed as 5 mol% iodine in ethanol [0.5 M] with equimolar stoichiometric ratio of all partners (HMF, aniline, diethyl phosphite). The corresponding Kabachnik–Fields product 4a was obtained in 71% isolated yield after 24 h, together with around 11% of unreacted HMF and 6% of the intermediate imine (
EntryCat. loadingSolvent [0.5 M]Temp.Ratio 1a/2a/3aTimeIsolated yield
15 mol%EtOH25 °C1 : 1 : 124 h71%
25 mol%MeCN25 °C1 : 1 : 124 h60%
35 mol%DCM25 °C1 : 1 : 124 h31%
45 mol%THF25 °C1 : 1 : 124 h84%
55 mol%2-MeTHF25 °C1 : 1 : 124 h74%
65 mol%THF25 °C1 : 1 : 1.524 h90%
7 5 mol% 2-MeTHF 25 °C 1 : 1 : 1.5 8 h 91%
82.5 mol%2-MeTHF25 °C1 : 1 : 1.58 h77%
91 mol%2-MeTHF25 °C1 : 1 : 1.58 h61%
102-MeTHF25 °C1 : 1 : 1.58 h54% (80%)b
115 mol%2-MeTHF50 °C1 : 1 : 1.54 h83%
125 mol%2-MeTHF78 °C1 : 1 : 13 h71%
Open in a separate windowaThe reaction was carried out in a sealed tube with HMF, aniline, diethyl phosphite, solvent and iodine, stirred at corresponding temperature for indicated time.b24 h.Based on this preliminary result, the reaction conditions were optimized, first by studying the influence of the solvent. THF was found to provide a better yield than EtOH, MeCN and DCM, affording product 4a in 84% yield (16 we decided to continue the investigation with 2-MeTHF as the solvent.Decreasing the catalyst loading to 2.5 mol% and 1 mol% led to slightly slower reactions (77% and 61% respectively) (Scheme 1 is depicted the scope of amines used in the reaction.Open in a separate windowScheme 1The Kabachnik–Fields reaction of HMF and different amines.a,b aThe reaction was carried out with HMF (1 mmol), amine (1 mmol), diethyl phosphite (1.5 mmol) with I2 (5 mol%) in 2-MeTHF (2 mL), stirred at 25 °C for indicated time. bIsolated yield. cAt 50 °C.Whatever the electron-donating or electron-withdrawing nature of the para substituent (methoxy-, chloro-, bromo-, iodo- and nitro-) on the aniline, the corresponding α-amino phosphonates 4b–4f were obtained in good to excellent yields (71–90%). An exception was observed for p-iodo-aniline requiring a 50 °C temperature for producing 4e in 77% yield. The same tendency was noticed for meta-substituted anilines, presumed to display low electronic influence on the reactivity (yields of 93% for 4g and 87% for 4h), and in a more unexpected way for 2-chloroaniline (82% for 4i). These results revealed that the substituted group on phenyl ring of aniline has globally a low impact on the reaction. Compared to anilines, aliphatic amines were found consistently as less reactive. In the case of aliphatic amines, elevated temperature (50 °C) was required to promote the reaction. Benzylamine and furfurylamine provided the corresponding α-amino phosphonates 4j and 4k in moderate yields, respectively 71% and 70%. Similar results were obtained for n-butylamine, cyclohexylamine and allylamine (4l–4n). Non-protected tryptamine afforded compound 4o in 57% yield. The product possibly arising from the reaction of the pyrrolic amine of tryptamine was not observed. tert-Butyl glycinate also worked under the conditions but gave a poor yield of 4p (31%). N-Methyl aniline, as an example of secondary amine, was also less reactive than aniline, giving 4q in 58% yield. When the chiral amine (R)-α-methylbenzylamine was used, the mixture of products was obtained 4r in 72% yield, from which the two isomers could not be separated entirely by column chromatography. A moderate diastereoselectivity was observed, with a 3.3 : 1 ratio of two diastereoisomers observed on the base of 31P NMR spectra. Similarly, (S)-α-methylbenzylamine gave the products 4s as a 3.5 : 1 mixture of two diastereoisomers in 70% yield.The nature of the dialkyl phosphite was also examined but to a minor extent due to the low diversity of commercially available phosphite reagents (Scheme 2). Dimethyl-, diisopropyl- and dibenzyl-phosphites afforded the corresponding products (4t–4v) in a range of yields of 86–89%. Surprisingly, phosphite with two strongly electron-withdrawing CF3– groups could also be used affording 4w in a modest 54% yield under the optimized conditions.Open in a separate windowScheme 2The Kabachnik–Fields reaction of HMF and commercially available phosphites.In order to expand the application of HMF, a pre-prepared 5,5′-[oxybis(methylene)]bis-2-furfural via self-etherification of HMF was subjected to the optimized conditions, resulting into the expected double Kabachnik–Fields product 4x in 86% yield. Alternatively, using p-phenylenediamine instead of aniline led to the other type of double Kabachnik–Fields product 4y. The results above undoubtedly indicate the possible application of the strategy towards highly functional polymers via Kabachnik–Fields polycondensation of 5,5′-[oxybis(methylene)]bis-2-furfural and suitable diamines (Scheme 3).17Open in a separate windowScheme 3The double Kabachnik–Fields reaction.A couple of derivatizations on hydroxyl group of the Kabachnik–Fields product were investigated using 4t as model substrate (Scheme 4). The aldehyde 4aa could be prepared in 87% yield by oxidation of 4t using Dess–Martin periodinane (DMP). The hydroxyl group of 4t could be also converted into an azido group after treatment with diphenylphosphoryl azide in the presence of DBU in 42% yield (compound 4ab). The acrylate 4ac was also easily obtained in a good yield. Diversification and optimization of these reactions are now in progress in the lab for further illustrating the usefulness of the hydroxymethyl appendage and providing a library of new α-amino phosphonates.Open in a separate windowScheme 4The derivatizations on hydroxyl group.Usually in a multicomponent reaction, the mechanism is not distinct because the reaction may undergo different pathways depending on which reactants react at first step. In order to gain insight into the mechanism of the Kabachnik–Fields reaction in our case, a series of control stepwise experiments were carried out (Scheme 5 and for more details see ESI). Mixing HMF (1 mmol) and aniline (1 mmol) in 2-MeTHF yielded the imine rapidly with and without iodine, with around 90% conversion observed in the crude NMR after 40 min in both cases (Exp. A and B). Subsequent addition of diethyl phosphite (1.5 mmol) and I2 (5 mol%) to the solution of the in situ formed imine (HMF, aniline, 1 h) afforded cleanly 4a after 8 h as seen by NMR (Exp. G). On the other hand, no reaction occurred when HMF (1 mmol) and diethyl phosphite (1.5 mmol) were mixed, either in the presence or absence of iodine (Exp. C and D). The above results indicate that the reaction likely undergoes the imine pathway, followed by nucleophilic attack by the phosphite to afford the α-amino phosphonate.8e,18 It is also known that iodine, as a Lewis acid, can activate imines in nucleophilic addition reactions.8e,19 This imine pathway was corroborated by the observation of the imine in the crude NMR of the three-component reaction mixture in the absence of iodine (Exp. E). In the presence of iodine, the proton of CH Created by potrace 1.16, written by Peter Selinger 2001-2019 N is shifted from 8.17 ppm to 8.42 ppm which made it difficult to identify, but the imine component was clearly detected in the crude reaction mixture by MS (imine plus H+: 202.0), thus also supporting the imine pathway (Exp. F).Open in a separate windowScheme 5Control experiments.  相似文献   

5.
Synthesis of core–shell N-TiO2@CuOx with enhanced visible light photocatalytic performance     
Shu Wang  Rufei Huo  Rui Zhang  Yuchuan Zheng  Changjiang Li  Le Pan 《RSC advances》2018,8(44):24866
In this paper, a core–shell N-TiO2@CuOx nanomaterial with increased visible light photocatalytic activity was successfully synthesized using a simple method. By synthesizing ammonium titanyl oxalate as a precursor, N-doped TiO2 can be prepared, then the core–shell structure of N-TiO2@CuOx with a catalyst loading of Cu on its surface was prepared using a precipitation method. It was characterized in detail using XRD, TEM, BET, XPS and H2-TPR, while its photocatalytic activity was evaluated using the probe reaction of the degradation of methyl orange. We found that the core–shell N-TiO2@CuOx nanomaterial can lessen the TiO2 energy band-gap width due to the N-doping, as well as remarkably improving the photo-degradation activity due to a certain loading of Cu on the surfaces of N-TiO2 supports. Therefore, a preparation method for a novel N, Cu co-doped TiO2 photocatalyst with a core–shell structure and efficient photocatalytic performance has been provided.

In this paper, a core–shell N-TiO2@CuOx nanomaterial with increased visible light photocatalytic activity was successfully synthesized using a simple method.  相似文献   

6.
Cationic palladium(ii)-catalyzed synthesis of substituted pyridines from α,β-unsaturated oxime ethers     
Takahiro Yamada  Yoshimitsu Hashimoto  Kosaku Tanaka  III  Nobuyoshi Morita  Osamu Tamura 《RSC advances》2022,12(33):21548
An efficient method for the synthesis of multi-substituted pyridines from β-aryl-substituted α,β-unsaturated oxime ethers and alkenes via Pd-catalyzed C–H activation has been developed. The method, using Pd(OAc)2 and a sterically hindered pyridine ligand, provides access to various multi-substituted pyridines with complete regioselectivity. Mechanistic studies suggest that the pyridine products are formed by Pd-catalyzed electrophilic C–H alkenylation of α,β-unsaturated oxime followed by aza-6π-electrocyclization. The utility of this method is showcased by the synthesis of 4-aryl-substituted pyridine derivatives, which are difficult to synthesize efficiently using previously reported Rh-catalyzed strategies with alkenes.

An efficient method for the synthesis of multi-substituted pyridines from α,β-unsaturated oxime ethers via cationic Pd(ii)-catalyzed C–H activation has been developed.  相似文献   

7.
An efficient access to β-ketosulfones via β-sulfonylvinylamines: metal–organic framework catalysis for the direct C–S coupling of sodium sulfinates with oxime acetates     
Tuong A. To  Chau B. Tran  Ngoc T. H. Nguyen  Hai H. T. Nguyen  Anh T. Nguyen  Anh N. Q. Phan  Nam T. S. Phan 《RSC advances》2018,8(31):17477
A copper-based framework Cu2(OBA)2(BPY) was synthesized and used as a recyclable heterogeneous catalyst for the synthesis of β-sulfonylvinylamines from sodium sulfinates and oxime acetates via direct C–S coupling reaction. The transformation was remarkably affected by the solvent, and chlorobenzene emerged as the best option. This Cu-MOF displayed higher activity than numerous conventional homogeneous and MOF-based catalysts. The catalyst was reutilized many times in the synthesis of β-sulfonylvinylamines without considerably deteriorating in catalytic efficiency. These β-sulfonylvinylamines were readily converted to the corresponding β-ketosulfones via a hydrolysis step with aqueous HCl solution. To the best of our knowledge, this direct C–S coupling reaction to achieve β-sulfonylvinylamines was not previously conducted with a heterogeneous catalyst.

Cu2(OBA)2(BPY) was used as catalyst for the synthesis of β-sulfonylvinylamines from sodium sulfinates and oxime acetates. These β-sulfonylvinylamines were readily converted to corresponding β-ketosulfones via a hydrolysis step.  相似文献   

8.
Enhanced charge separation in TiO2/nanocarbon hybrid photocatalysts through coupling with short carbon nanotubes     
Ahmed Al Mayyahi  Brian M. Everhart  Tej B. Shrestha  Tyson C. Back  Placidus B. Amama 《RSC advances》2021,11(19):11702
The interfacial contact between TiO2 and graphitic carbon in a hybrid composite plays a critical role in electron transfer behavior, and in turn, its photocatalytic efficiency. Herein, we report a new approach for improving the interfacial contact and delaying charge carrier recombination in the hybrid by wrapping short single-wall carbon nanotubes (SWCNTs) on TiO2 particles (100 nm) via a hydration-condensation technique. Short SWCNTs with an average length of 125 ± 90 nm were obtained from an ultrasonication-assisted cutting process of pristine SWCNTs (1–3 μm in length). In comparison to conventional TiO2–SWCNT composites synthesized from long SWCNTs (1.2 ± 0.7 μm), TiO2 wrapped with short SWCNTs showed longer lifetimes of photogenerated electrons and holes, as well as a superior photocatalytic activity in the gas-phase degradation of acetaldehyde. In addition, upon comparison with a TiO2–nanographene “quasi-core–shell” structure, TiO2-short SWCNT structures offer better electron-capturing efficiency and slightly higher photocatalytic performance, revealing the impact of the dimensions of graphitic structures on the interfacial transfer of electrons and light penetration to TiO2. The engineering of the TiO2–SWCNT structure is expected to benefit photocatalytic degradation of other volatile organic compounds, and provide alternative pathways to further improve the efficiency of other carbon-based photocatalysts.

The interfacial contact between TiO2 and graphitic carbon in a hybrid composite plays a critical role in electron transfer behavior, and in turn, its photocatalytic efficiency.  相似文献   

9.
A novel superparamagnetic powerful guanidine-functionalized γ-Fe2O3 based sulfonic acid recyclable and efficient heterogeneous catalyst for microwave-assisted rapid synthesis of quinazolin-4(3H)-one derivatives in Green media     
Fateme Haji Norouzi  Naser Foroughifar  Alireza Khajeh-Amiri  Hoda Pasdar 《RSC advances》2021,11(48):29948
The novel organic–inorganic nanohybrid superparamagnetic (γ-Fe2O3@CPTMS–guanidine@SO3H) nanocatalyst modified with sulfonic acid represents an efficient and green catalyst for the one-pot synthesis of quinazolin-4(3H)-one derivatives via three-component condensation reaction between anthranilic acid, acetic anhydride and different amines under microwave irradiation and solvent-free conditions (4a–q). XRD, FT-IR, FE-SEM, TGA, VSM and EDX were used to characterize this new magnetic organocatalyst. Outstanding performance, short response time (15–30 min), simple operation, easy work-up procedure, and avoidance of toxic catalysts can be regarded as its significant advantages. Moreover, it can be easily separated from the reaction solution through magnetic decantation using an external magnet, and recycled at least six times without notable reduction in its activity.

A novel organic–inorganic nanohybrid superparamagnetic nanocatalyst (γ-Fe2O3@CPTMS–guanidine@SO3H) represents an efficient and green catalyst for the one-pot synthesis of quinazolin-4(3H)-one derivatives via a three-component condensation reaction.  相似文献   

10.
Microwave synthesised Pd–TiO2 for photocatalytic ammonia production     
Jake M. Walls  Jagdeep S. Sagu  K. G. Upul Wijayantha 《RSC advances》2019,9(11):6387
Palladium doped anatase TiO2 nanoparticles were synthesised by a rapid (3 min) one-pot microwave synthesis technique at low temperature and pressure. After being fully characterised by SEM, XRD, Raman, XPS and EDX, photocatalytic nitrate reduction and ammonia production were studied over various dopant levels between 0–3.97 wt% Pd and compared to similar previous literature. Improved yields of ammonia were observed with most dopant levels when compared to non-doped microwave synthesised TiO2 with 2.65 wt% found to be the optimum dopant level producing 21.2 μmol NH3. Electrochemical impedance spectroscopy of TiO2 and Pd–TiO2 photoelectrodes revealed improvements in charge transfer characteristics at high Pd dopant levels.

A rapid 3 minute one-pot microwave synthesis of Pd–TiO2 showing improved activity for photocatalytic nitrate reduction into ammonia.  相似文献   

11.
Enhanced photocatalytic water splitting of a SILAR deposited α-Fe2O3 film on TiO2 nanoparticles     
Zahra-Sadat Pourbakhsh  Kyana Mohammadi  Ahmad Moshaii  Maryam Azimzadehirani  Amir Hosseinmardi 《RSC advances》2019,9(55):31860
We have investigated the effect of deposition of a α-Fe2O3 thin layer on a substrate of TiO2 nanoparticles for photoelectrochemical (PEC) water splitting. The TiO2 layer was coated on an FTO substrate using the paste of TiO2 nanoparticles. The α-Fe2O3 layer was deposited on the TiO2 thin film, using the method of Successive Ionic Layer Adsorption and Reaction (SILAR) with different cycles. Various characterizations including XRD, EDX and FE-SEM confirm the formation of α-Fe2O3 and TiO2 nanoparticles on the electrode. The UV-visible absorption spectrum confirms a remarkable enhancement of the absorption of the α-Fe2O3/TiO2/FTO composite relative to the bare TiO2/FTO. In addition, the photocurrents of the composite samples are remarkably higher than the bare TiO2/FTO. This is mainly due to the low band gap of α-Fe2O3, which extends the absorption spectrum of the α-Fe2O3/TiO2 composite toward the visible region. In addition, the impedance spectroscopy analysis shows that the recombination rate of the charge carriers in the α-Fe2O3/TiO2 is lower than that for the bare TiO2. The best PEC performance of the α-Fe2O3/TiO2 sample was achieved by the sample of 70 cycles of α-Fe2O3 deposition with about 7.5 times higher photocurrent relative to the bare TiO2.

Optimization of photoelectrochemical water splitting by a composite of SILAR-deposited α-Fe2O3 thin film on a substrate of TiO2 nanoparticles.  相似文献   

12.
Copper-catalyzed synthesis of α-ketoamides using water and dioxygen as the oxygen source     
Yuanyuan Xiao  Zijuan Yi  Xianyong Yu  Fang Xiao 《RSC advances》2020,10(49):29114
The reaction employing H2O and O2 as the co-oxygen source in the catalytic synthesis of α-ketoamides is described. This copper-catalyzed reaction is carried out in a tandem manner constituted by the hydroamination of alkyne, hydration of vinyl–Cu complex and subsequent oxidation. Isotope labeling and radical capture experiments reveal that the oxygen atom of α-ketone at α-ketoamides derives from O2 and the oxygen atom of amide group originates from H2O.

The reaction employing H2O and O2 as the co-oxygen source in the catalytic synthesis of α-ketoamides is described. This Cu-catalyzed reaction is carried out in a tandem manner constituted by hydroamination of alkyne, hydration of vinyl–Cu complex and subsequent oxidation.  相似文献   

13.
Visible light-induced oxidative α-hydroxylation of β-dicarbonyl compounds catalyzed by ethylenediamine–copper(ii)     
Yujie He  Hao Yin  Yifeng Wang  Mingming Chu  Yiming Li 《RSC advances》2023,13(12):7843
We have developed an efficient oxidative α-hydroxylation of β-keto esters with firstly using the structurally simple ethylenediamine–copper(ii) as a catalyst for β-keto esters activation and using visible light as the driving force for generating more active singlet oxygen (1O2) from triplet state oxygen (3O2) in the air, providing a series of α-hydroxy β-keto esters in excellent yields (up to 99%) under extremely low photosensitizer loading (0.01 mol%) and catalyst loading (1 mol%) within a short time. Moreover, the gram-scale synthesis showed the practical utility of this protocol.

An efficient visible light-induced oxidative α-hydroxylation of β-dicarbonyl compounds has been developed using structurally simple ethylenediamine–copper(ii) as a catalyst.  相似文献   

14.
Fullerene–porphyrin hybrid nanoparticles that generate activated oxygen by photoirradiation     
Kouta Sugikawa  Kosuke Masuda  Kentaro Kozawa  Riku Kawasaki  Atsushi Ikeda 《RSC advances》2021,11(3):1564
The preparation of water-dispersible hybrid nanoparticles comprising fullerene and porphyrin from cyclodextrin complexes is described. In the presence of polyethylene glycol, C60 fullerene and porphyrin were expelled from the cyclodextrin cavity to form fullerene–porphyrin hybrid nanoparticles in water. The fullerene–porphyrin hybrid nanoparticles exhibit improved singlet oxygen generation ability under photoirradiation compared with that of C60 nanoparticles.

Hybrid nanoparticles comprising fullerene and porphyrin are formed via guest exchange reaction of cyclodextrin complexes. The hybrid nanoparticles exhibit singlet oxygen generation ability under photoirradiation.

Water-dispersible colloidal fullerene assemblies, referred to as fullerene nanoparticles (NPs), have recently received increasing attention.1–3 Fullerene NPs are negatively charged and can be dispersed in water in the absence of any solubilizer. Fullerene NPs demonstrate promise within biological and medical applications, as radical scavengers and photosensitizers for photodynamic therapy. To further extend the applications of fullerene NPs, additional hybridization with desired functional molecules is required. Porphyrin and associated derivatives are highly promising candidates for hybridization with fullerenes to increase photoactivity.4 Numerous studies on the complexation of fullerenes with porphyrin molecules using synthetic organic chemistry5–7 or supramolecular chemistry8–10 have been reported. Although fullerene NPs have been intensively studied over the last decade, no reliable method to achieve the hybridization of porphyrins with fullerene NPs has been proposed.Poly(ethylene glycol) monomethyl ether (PEG) was recently observed to accelerate the decomposition of fullerene C60–γ-CD complexes in water, which leads to the rapid aggregation of C60 to form water-dispersible C60 NPs.11 In this method, C60–γ-CD complexes can exist as stable isolated molecules in water, enabling the precise size control and step-wise growth of C60 NPs.12,13 Herein, the preparation of hybrid NPs comprising C60 and hydrophobic porphyrin molecules are reported. C60–γ-CD and porphyrin-trimethyl-β-cyclodextrin (por–TMe-β-CD) complexes are mixed in water in the presence of PEG. Both complexes decompose through the interaction of PEG with the CDs, leading to the formation of C60–porphyrin hybrid NPs (denoted as C60–por NPs). The C60–por NPs are negatively charged and easily disperse in water. Additionally, the ability of C60–por NPs to generate activated oxygen is also evaluated.The C60–γ-CD complex14,15 and 1–TMe-β-CD complex (Fig. 1)16–18 were prepared according to a previously described procedure (see the ESI for details). The 1H NMR spectrum of the mixed solution comprising the C60–γ-CD complex and PEG after 1 h incubation at 80 °C shows that the peaks attributed to the C60–γ-CD complexes completely disappeared (Fig. S1). Hence, the 1H NMR data confirm the decomposition of the C60–γ-CD complexes and the formation of water-dispersible C60 NPs, as previously reported.11–13 The effect of PEG on 1–TMe-β-CD complexes was also investigated by 1H-NMR as shown in Fig. S2. After incubating the mixed solution of 1–TMe-β-CD complex and PEG ([1] = 0.1 mM, [PEG] = 5.0 g L−1) for 1 h at room temperature, peaks attributed to the 1–TMe-β-CD complex were still evident at 4.97 ppm and above 7.7 ppm (Fig. S2(i)). Hence, PEG has no influence upon the 1–TMe-β-CD structure at room temperature. Conversely, after incubating for 1 h at 80 °C, a dark purple precipitate formed and the aforementioned 1H NMR peaks completely disappeared (Fig. S2(ii)). In the absence of PEG, the 1–TMe-β-CD complex was stable in water both at room temperature (Fig. S2(iii)) and 80 °C (Fig. S2(iv)). These results suggest a decomposition route of 1–TMe-β-CD by interaction with PEG at 80 °C, with a concomitant formation of non-dispersible large aggregates.Open in a separate windowFig. 1Chemical structures of porphyrin derivatives used in this study.To obtain hybrid NPs comprising C60 and 1 (C60–1 NPs), PEG (Mw = 2000) was added to an aqueous solution containing C60–γ-CD and 1–TMe-β-CD complexes ([C60] = [1] = 0.1 mM, [PEG] = 5.0 g L−1), which were thereafter incubated at room temperature or 80 °C. The 1H NMR spectrum of the mixed solution at room temperature shows peaks attributed to γ-CD in the C60–γ-CD complex at 5.03 ppm, TMe-β-CD in the 1–TMe-β-CD complex at 4.97 ppm, and 1 in the 1–TMe-β-CD complex in the region of 7.6–8.5 ppm (Fig. 2a(i)). The data indicate that PEG fails to induce the decomposition of the C60–γ-CD and 1–TMe-β-CD complexes at room temperature. Conversely, after the mixture was heated at 80 °C for 1 h, the peaks attributed to the C60–γ-CD and 1–TMe-β-CD complexes completely disappeared (Fig. 2a(ii)). Hence, C60–γ-CD and 1–TMe-β-CD were decomposed at 80 °C, in the presence of PEG. The solution after being subjected to heat treatment at 80 °C for 1 h, was dark purple in the absence of any precipitate. The hydrodynamic diameter and ζ-potential of the reacted solution were 125 nm (polydispersity index = 0.21) and −20.2 mV, respectively. Water dispersible fullerene NPs typically exhibit negative ζ-potentials, the origin of which still requires elucidating.19,20 Hence, the formation of water-dispersible nano-composites, C60–1 NPs, is suggested.Open in a separate windowFig. 2(a) 1H NMR spectra of mixed solutions comprising fullerene C60–γ-cyclodextrin (CD) and 1–trimethyl (TMe)-β-CD complexes ([C60] = [1] = 0.1 mM) (i) before and (ii) after heating at 80 °C for 1 h, in the presence of polyethylene glycol (PEG) (5 g L−1). Open circles: free γ-CD, filled circles: C60–γ-CD, open diamonds: free TMe-β-CD, and filled diamonds: porphyrin–TMe-β-CD (por–TMe-β-CD) complex. The spectra at 7.6–8.5 ppm, are amplified five-fold. (b) Ultraviolet-visible (UV/Vis) absorption spectra of the mixed solution comprising C60–γ-CD and 1–TMe-β-CD complexes ([C60] = [1] = 0.1 mM) with PEG (5 g L−1), before (dashed line) and after (solid line) heating at 80 °C for 1 h. (c) UV/Vis absorption spectra of the mixed solution comprising the C60–γ-CD complex as a function of the 1–TMe-β-CD complex concentration ([C60] = 0.1 mM, [1] = 0–0.2 mM) with PEG (5 g L−1) after heating at 80 °C for 1 h.The por–TMe-β-CD complexes using 2–6 (Fig. 1), were also prepared adopting the same procedure as that for the 1–TMe-β-CD complex. Each por–TMe-β-CD complex solution was mixed with C60–γ-CD and PEG ([C60] = [por] = 0.1 mM, [PEG] = 5.0 g L−1). The 1H NMR spectrum of each individual mixed solution after being incubated for 1 h at 80 °C, is shown in Fig. S3. In the 1H NMR spectrum of the mixture comprising C60–γ-CD and 2–TMe-β-CD, the peaks attributed to these complexes at 5.03, 4.98, and 7.60–8.50 ppm, completely disappeared after being subjected to incubation for 1 h at 80 °C, without precipitation (Fig. S3a). Similar changes in the 1H NMR spectrum of the mixture comprising C60–γ-CD and 3–TMe-β-CD complexes are observed, as shown in Fig. S3b. The data suggest that the 2–TMe-β-CD and 3–TMe-β-CD complexes can be decomposed in a similar manner as the C60–γ-CD complexes, and imply the formation of C60–2 and C60–3 NPs.The 1H NMR spectra of the mixed solutions comprising C60–γ-CD and 4, 5, or 6–TMe-β-CD complexes failed to show peaks attributed to the C60–γ-CD complex, and peaks associated with the respective por–TMe-β-CD complexes were observed after incubation for 1 h at 80 °C (Fig. S3c–e, respectively). Hence, the data suggest that the C60–γ-CD complex decomposed in the presence of PEG, and the 4, 5, and 6–TMe-β-CD complexes were observed to be stable without decomposition at 80 °C. There have been reports suggesting the strong interaction of water-soluble tetraphenyl porphyrins with TMe-β-CDs.21,22 Polar substituents prompt the penetration of the polarized porphyrin rims into the TMe-β-CD cavity. Porphyrins 4–6 possess polar substituents, which are suggested to enable the formation of stable TMe-β-CD complexes. Furthermore, the size of the β-CD cavity is sufficiently narrow to prevent any strong interaction with PEG.23 Thus, PEG-induced decomposition of the 4, 5, and 6–TMe-β-CD complexes is not possible.The absorption behavior of C60–1 NPs was investigated using ultraviolet-visible (UV/Vis) spectroscopy. In the UV/Vis spectra, the characteristic peak of solvated C60–γ-CD, at 333 nm shifted to 344 nm after heating at 80 °C for 1 h (Fig. 2b). An additional broad absorption at 400–550 nm is also apparent, which is characteristic of solid-state crystalline C60 and arises from the electronic interactions between adjacent C60 molecules.24,25 The characteristic peak of the solvated 1–TMe-β-CD complex at 415 nm, shifted to 432 nm, with induced broadening after being subjected to heat treatment at 80 °C for 1 h (Fig. 2b). In the absence of C60–γ-CD complexes, the characteristic absorption peak attributed to the solvated 1–TMe-β-CD complex completely disappeared after heating for 1 h at 80 °C, in the presence of PEG (Fig. S4). The data show that 1, when expelled from the TMe-β-CD cavities, forms non-dispersible precipitates in the absence of C60. For 1 dispelled from the TMe-β-CD cavities to be stably dispersed in water, formation of co-aggregates with C60 may be a prerequisite. C60–2 and C60–3 NPs also show similar UV-Vis absorption spectra after being subjected to heating at 80 °C for 1 h, as shown in Fig. S5a and b, respectively.To further elucidate the composite formation of C60 and 1, the influence of 1 concentration on C60–1 NP formation was investigated by UV/Vis spectroscopy (Fig. 2c). The intensity of the absorption peak at 432 nm increased as a function of 1 concentration from 0.05 to 0.1 mM. Conversely, the absorption peak at 345 nm, derived from the formation of fullerene NPs, shifted to 338 nm, with increasing concentration of 1. This absorption peak derives from the fullerene nanoparticle size, and as the size decreased (i.e., the NPs became smaller), the peak blue-shifted.11 The results suggest that in the presence of 1, the fullerene interaction might be disturbed, or smaller C60 NPs might form. For C60–1 NPs formulated with 0.2 mM of 1 ([C60] = 0.1 mM, [1] = 0.2 mM), the absorption peak derived from the Soret band of 1 split into two peaks (Fig. 2c). The absorption peak at the shorter wavelength of 415 nm is consistent with that of the 1–TMe-β-CD complex. The absorption peak at the longer wavelength of 431 nm is almost consistent with the absorption peaks in the UV/Vis spectra of C60–1 NPs fabricated with 0.05 and 0.1 mM of 1. The findings indicate that in the sample comprising 0.2 mM of 1, a portion of the 1–TMe-β-CD complexes remained in the complex state after heating for 1 h at 80 °C, in the presence of PEG. The absorption peak at 338 nm, which reflects the state of fullerene NPs, is similar to that of C60–1 NPs fabricated with 0.1 mM of 1. When C60 and 1 form co-aggregated NPs, the ratio of 1 to C60 is thought to be limited to ∼1 : 1.Morphological observations of the hybrid NPs were also undertaken. In the absence of the por–TMe-β-CD complex, C60 NPs possessing fairly monodisperse size distributions were observed (Fig. S6a). The average diameter of the individual NPs, determined from the transmission electron microscopy (TEM) images, is 82 nm. C60 NPs have been previously reported to exhibit lattice fringes and diffraction patterns, which suggests that C60 NPs maintain the face-centered cubic (fcc) crystalline structure.11 C60–1 NPs prepared with 0.05 mM C60 and 0.1 mM 1, possessed irregular shapes (Fig. 3a and b, respectively). The average diameter of C60–1 NPs, determined by TEM, is 119 nm (Fig. 3a and S6b), demonstrating the larger C60–1 NP size than that of the C60 NPs (82 nm). Increasing the concentration of 1 to 0.1 mM results in the average diameter of C60–1 NPs to increase to 131 nm (Fig. 3b and S6c). Similar morphology is observed from the TEM micrographs of C60–2 and C60–3 NPs having average diameters of 109 nm and 144 nm, respectively (Fig. 3c and d, respectively). A lower PEG molecular weight or a lower reaction temperature during C60 NP formation via C60–γ-CD complexes have been reported to induce an increase in the diameter of the C60 NPs.11,12 Thus, the findings suggest that slower reaction conditions result in less nucleation and a larger NP formation. Porphyrin 3 possesses a methoxy substituent at the para position of the phenyl group and is more polar than 1 or 2. Previous reports have demonstrated that the higher the polarity of the phenyl group, the more stable the complex with TMe-β-CD,21,22 which indicates that 3–TMe-β-CD is more stable than 1– or 2–TMe-β-CD in water. Thus, the aforementioned decomposition, which results from the interaction with PEG, is slower in 3 with a concomitant increase in the NP size.Open in a separate windowFig. 3Transmission electron microscopy (TEM) images of C60–1 nanoparticles (NPs) prepared with (a) 0.05 and (b) 0.1 mM of the 1–TMe-β-CD complex. TEM images of (c) C60–2 and (d) C60–3 NPs. Scale bars in images (a–d) are 100 nm. (e) High resolution TEM micrograph and selected-area electron diffraction pattern of C60–1 NP. Scale bar is 10 nm. (f) 13C NMR spectra of (i) C60 NPs and (ii) C60–1 NPs. (g) Illustration of a C60–por NP. A portion of C60 form crystalline structures, while a portion of the porphyrin molecules interact with C60 at the molecular level.In the high-resolution TEM micrograph (Fig. 3e), the C60–1 NPs only exhibited partial lattice fringes, and hence did not show clear diffraction patterns compared with the C60 NPs (inset in Fig. 3e).11 The findings demonstrate the highly amorphous nature of the C60–1 NPs, and that the C60 crystal structure was only retained in part. 13C NMR spectra also provide important information about the structure of the C60–1 NPs. A characteristic C60 cluster signal at 142.4 ppm was detected in both C60 NPs and C60–1 NPs (Fig. 3f).26 The C60–1 NP dispersions also exhibited several small new signals at 141.5, 143.3, and 143.7 ppm, as shown in Fig. 3f(ii). When a fullerene and a porphyrin molecule form a stable complex in solution, the C60 signal shifts depending on the interaction type between the fullerene and the porphyrin molecule.27 Thus, the porphyrin molecule interacted with the aggregate of C60 in C60–1, as illustrated in Fig. 3g.Some porphyrin molecules can form a co-crystal with fullerene C60.28 To form a crystal structure, not only the interaction between molecules but also the relationship with the solvent, such as gradually changing the polarity of the solvent or removing the solvent, are important. In our system, porphyrin molecules that are pseudo-dissolved by TMe-β-CDs are added to water, which is a poor solvent for porphyrins, through the interaction of PEG with TMe-β-CDs. The porphyrin molecule should immediately aggregate and have difficulty forming a crystal structure. Furthermore, because water is also a poor solvent for fullerene C60, C60 also immediately aggregates in water. Thus, it should be extremely difficult for C60 and porphyrin molecules to regularly associate to form a co-crystal structure.The concentration of singlet oxygen molecules (1O2, Type-II energy transfer pathway) generated by photoirradiation was measured according to a chemical method using 9,10-anthracenediyl-bis(methylene) dimalonic acid (ABDA)15,29 as a marker to determine the biological activities of C60 NPs, C60–1 NPs, C60–2 NPs, and C60–3 NPs. The absorption of ABDA at the absorption maximum (380 nm) was monitored as a function of irradiation time ([C60] = 0.1 mM, [por] = 0 or 0.1 mM). Under visible-light irradiation at wavelengths > 620 nm, C60–1 NPs, C60–2 NPs, and C60–3 NPs generated higher levels of 1O2 than C60 (Fig. 4a). These results show that the 1O2 photoproduction abilities of the C60–por NPs were higher than that of the C60 NPs. There are no significant differences in the 1O2 photoproduction abilities of C60–1 NPs, C60–2 NPs, and C60–3 NPs, which suggests that the structure of the porphyrin has an insignificant influence on the ability of the hybrid NPs. The generation of formazan, via the reduction of nitroblue tetrazolium (NBT) by oxygen radicals (O2˙), is observed as an increase of absorption intensity at 560 nm.30 The reduction of NBT by O2˙ was scarcely detected in solutions containing C60–1 NPs, C60–2 NPs, and C60–3 NPs under photoirradiation, even though formazan was readily detected in the positive control sample in the presence of reduced nicotinamide adenine dinucleotide (NADH) (Fig. 4b). The results suggest that the reactive oxygen species produced by C60–1 NPs, C60–2 NPs, and C60–3 NPs are predominantly 1O2 generated by a Type-II reaction.18Open in a separate windowFig. 4(a) 1O2 generation by NPs. Bleaching of 9,10-anthracenediyl-bis(methylene) dimalonic acid (ABDA) was monitored as a function of the decrease in the absorbance at 380 nm, for C60 NPs (black circles and solid line), C60–1 NPs (red circles and solid line), C60–2 NPs (blue circles and solid line), and C60–3 NPs (green circles and solid line) ([C60] = 15 μM, [por] = 0 or 15 μM, [ABDA] = 25 μM). (b) O2˙ generation by NPs. The amount of formazan generated by the reduction of nitroblue tetrazolium (NBT) in the presence of O2˙ was analyzed by the absorbance at 560 nm, of C60–1 NPs (red circles) and C60–2 NPs (blue circles) in the absence (solid lines) and presence (dashed lines) of NADH ([C60] = 15 μM, [por] = 15 μM, [NBT] = 200 μM, [NADH] = 0 or 625 μM). All samples were photoirradiated at >620 nm, in O2-saturated aqueous solutions.In summary, the preparation of hybrid C60–porphyrin NPs was achieved via a guest exchange reaction comprising porphyrin CD complexes and C60. Seven C60–por NP derivatives with various moieties were prepared. CD porphyrin complexes possessing phenyl and methoxyphenyl moieties were decomposed in the presence of PEG at the same time as C60–γ-CD complexes and formed NPs with C60. Porphyrins containing a hydrophilic moiety form stable complexes with TMe-β-CD and fail to co-aggregate with C60. The C60–por NPs are negatively charged and are easily dispersed and stable in water. The 1O2 generation ability of C60–por NPs under photoirradiation (>620 nm) is greater than that of C60 NPs. The findings herein demonstrate a new method to fabricate fullerene–porphyrin composite materials, which provides a route to highly functional fullerene-based materials.  相似文献   

15.
Global minimum beryllium hydride sheet with novel negative Poisson's ratio: first-principles calculations     
Feng Li  Urs Aeberhard  Hong Wu  Man Qiao  Yafei Li 《RSC advances》2018,8(35):19432
As one of the most prominent metal-hydrides, beryllium hydride has received much attention over the past several decades, since 1978, and is considered as an important hydrogen storage material. By reducing the dimensionality from 3 to 2, the beryllium hydride monolayer is isoelectronic with graphene; thus the existence of its two-dimensional (2D) form is theoretically feasible and experimentally expected. However, little is known about its 2D form. In this work, by a global minimum search with the particle swarm optimization method via density functional theory computations, we predicted two new stable structures for the beryllium hydride sheets, named α–BeH2 and β–BeH2 monolayers. Both structures have more favorable thermodynamic stability than the recently reported planar square form (Nanoscale, 2017, 9, 8740), due to the forming of multicenter delocalized Be–H bonds. Utilizing the recently developed SSAdNDP method, we revealed that three-center-two-electron (3c–2e) delocalized Be–H bonds are formed in the α–BeH2 monolayer, while for the β–BeH2 monolayer, novel four-center-two-electron (4c–2e) delocalized bonds are observed in the 2D system for the first time. These unique multicenter chemical bonds endow both α– and β–BeH2 with high structural stabilities, which are further confirmed by the absence of imaginary modes in their phonon spectra, the favorable formation energies comparable to bulk and cluster beryllium hydride, and the high mechanical strength. These results indicate the potential for experimental synthesis. Furthermore, both α– and β–BeH2 are wide-bandgap semiconductors, in which the α–BeH2 has unusual mechanical properties with a negative Poisson''s ratio of −0.19. If synthesized, it would attract interest both in experiment and theory, and be a new member of the 2D family isoelectronic with graphene.

As one of the most prominent metal-hydrides, beryllium hydride has received much attention over the past several decades, since 1978, and is considered as an important hydrogen storage material.  相似文献   

16.
TiO2–Au composite nanofibers for photocatalytic hydrogen evolution     
Xiaojiao Yang  Xuelian Wu  Jun Li  Ying Liu 《RSC advances》2019,9(50):29097
TiO2-based materials for photocatalytic hydrogen (H2) evolution have attracted much interest as a renewable approach for clean energy applications. TiO2–Au composite nanofibers (NFs) with an average fiber diameter of ∼160 nm have been fabricated by electrospinning combined with calcination treatment. In situ reduced gold nanoparticles (NPs) with uniform size (∼10 nm) are found to disperse homogenously in the TiO2 NF matrix. The TiO2–Au composite NFs catalyst can significantly enhance the photocatalytic H2 generation with an extremely high rate of 12 440 μmol g−1 h−1, corresponding to an adequate apparent quantum yield of 5.11% at 400 nm, which is 25 times and 10 times those of P25 (584 μmol g−1 h−1) and pure TiO2 NFs (1254 μmol g−1 h−1), respectively. Furthermore, detailed studies indicate that the H2 evolution efficiency of the TiO2–Au composite NF catalyst is highly dependent on the gold content. This work provides a strategy to develop highly efficient catalysts for H2 evolution.

The H2 production rate of TiO2–Au nanofibers is dramatically improved to 12 440 μmol g−1 h−1, 10 times that of pure TiO2.  相似文献   

17.
An expeditious synthesis of 6,7-dihydrodibenzo[b,j][4,7] phenanthroline derivatives as fluorescent materials     
Kevin George  Sathananthan Kannadasan 《RSC advances》2022,12(42):27246
A rapid and efficient method has been developed for the synthesis of 13,14-dimethyl-6,7-dihydrodibenzo[b,j][4,7]phenanthroline derivatives (3a–d) through the Friedländer condensation of 2-aminoarylketone with 1,4-cyclohexanedione under solvent-free conditions using p-toluenesulphonic acid. The synthetic utility of compounds 3a, 3b, and 3c was demonstrated by synthesizing compounds 6a–kvia Suzuki coupling, 8 by Buchwald–Hartwig amination, and 9a–bvia NBS bromination. Significantly, the emission band corresponding to the π–π* electronic transition of compounds 3a, 6a, 6d, 6f, and 8 showed a redshift with increasing polarity of the solvents. Molar extinction coefficient (ε), Stoke''s shift (Δ Created by potrace 1.16, written by Peter Selinger 2001-2019 ), and quantum yield (Φf) were calculated for all these compounds.

A rapid method has been developed for the synthesis of 13,14-dimethyl-6,7-dihydrodibenzo[b,j][4,7]phenanthroline derivatives (3a–d) through the Friedländer condensation of 2-aminoarylketone with 1,4-cyclohexanedione under neat conditions using p-TSA.  相似文献   

18.
Deposition of platinum on boron-doped TiO2/Ti nanotube arrays as an efficient and stable photocatalyst for hydrogen generation from water splitting     
Mengjia Sun  Yanli Jiang  Mei Tian  Huijun Yan  Ran Liu  Lijuan Yang 《RSC advances》2019,9(20):11443
An efficient photocatalyst of boron-doped titanium dioxide/titanium nanotube array-supported platinum particles (Pt–B/TiO2/Ti NTs) was fabricated for photocatalytic water splitting for hydrogen production through a two-step route. First, B/TiO2/Ti NTs were prepared by anodic oxidation using hydrofluoric acid as electrolyte and boric acid as a B source. Then, Pt particles were deposited on the surface of B/TiO2/Ti NTs by photo-assisted impregnation reduction. The structure and properties of Pt–B/TiO2/Ti NTs were characterized by various physical measurements which showed the successful fabrication of Pt–B/TiO2/Ti NTs. The Pt–B/TiO2/Ti NTs, with a B-doping content of 15 mmol L−1, showed the highest photocatalytic activity and exhibited a photocatalytic hydrogen-production rate of 384.9 μmol g−1 h−1, which was 9.2-fold higher than that of unmodified TiO2/Ti NTs (41.7 μmol g−1 h−1). This excellent photocatalytic performance was ascribed mainly to the synergistic effect of Pt and B, which could enhance the photocatalytic activity of TiO2/Ti NTs.

Pt–B/TiO2/Ti NTs, prepared by anodic oxidation and photo-deposition methods, showed excellent photocatalytic activity.  相似文献   

19.
Synthesis of hierarchical nanocrystalline β zeolite as efficient catalyst for alkylation of benzene with benzyl alcohol     
Pan Zhou  Meng-Nan Liu  Qun-Xing Luo  Jianbo Zhang  Huiyong Chen  Xiaoxun Ma  Qing-Qing Hao 《RSC advances》2022,12(8):4865
To develop an efficient solid acid catalysts for the Friedel–Crafts alkylation reaction, especially for involving bulky molecules, the direct synthesis of hierarchical nanocrystalline β zeolites were achieved by using amphiphilic organosilane ([(CH3O)3SiC3H6N(CH3)2C18H37]Cl, TPOAC) as collaborative structure-directing agent (SDA). The growth evolution of β crystals and the influence of TPOAC/SiO2 molar ratio on the mesoporous structure, crystal size, and acidic properties of β zeolites were investigated and discussed in detail. The characterization results reveal that intracrystalline mesopores and intercrystalline mesopores/macropores via the stacking of β nanocrystals were generated over the hierarchical β zeolites. Moreover, most of the strong acid sites were well remained compared with the conventional microporous β zeolite. Consequently, the hierarchical nanocrystalline β zeolite synthesized under the optimized synthesis conditions shows improved specific catalytic activity of acid sites (turnover number, TON) in alkylation of benzene with benzyl alcohol, which can be attributed to the integrated balance of considerable mesoporosity, accessibility of the acid sites, and well-remained strong acid sites in the hierarchical β zeolite.

Hierarchical β zeolite with enhanced transport and specific catalytic activity of acid sites in Friedel–Crafts alkylation was achieved by using amphiphilic organosilane surfactant as mesopores-directing agent and crystal growth inhibitor.  相似文献   

20.
Mn-based catalysts supported on γ-Al2O3, TiO2 and MCM-41: a comparison for low-temperature NO oxidation with low ratio of O3/NO     
Lijun Liu  Boxiong Shen  Meng Si  Peng Yuan  Fengju Lu  Hongpei Gao  Yan Yao  Cai Liang  Hongjie Xu 《RSC advances》2021,11(31):18945
Mn-Based catalysts supported on γ-Al2O3, TiO2 and MCM-41 synthesized by an impregnation method were compared to evaluate their NO catalytic oxidation performance with low ratio O3/NO at low temperature (80–200 °C). Activity tests showed that the participation of O3 remarkably promoted the NO oxidation. The catalytic oxidation performance of the three catalysts decreased in the following order: Mn/γ-Al2O3 > Mn/TiO2 > Mn/MCM-41, indicating that Mn/γ-Al2O3 exhibited the best catalytic activity. In addition, there was a clear synergistic effect between Mn/γ-Al2O3 and O3, followed by Mn/TiO2 and O3. The characterization results of XRD, EDS mapping, BET, H2-TPR, XPS and TG showed that Mn/γ-Al2O3 had good manganese dispersion, excellent redox properties, appropriate amounts of coexisting Mn3+ and Mn4+ and abundant chemically adsorbed oxygen, which ensured its good performance. In situ DRIFTS demonstrated the NO adsorption performance on the catalyst surface. As revealed by in situ DRIFTS experiments, the chemically adsorbed oxygen, mainly from the decomposition of O3, greatly promoted the NO adsorption and the formation of nitrates. The Mn-based catalysts showed stronger adsorption strength than the corresponding pure supports. Due to the abundant adsorption sites provided by pure γ-Al2O3, under the interaction of Mn and γ-Al2O3, the Mn/γ-Al2O3 catalyst exhibited the strongest NO adsorption performance among the three catalysts and produced lots of monodentate nitrates (–O–NO2) and bidentate nitrates (–O2NO), which were the vital intermediate species for NO2 formation. Moreover, the NO–TPD studies also demonstrated that Mn/γ-Al2O3 showed the best NO desorption performance among the three catalysts. The good NO adsorption and desorption characteristics of Mn/γ-Al2O3 improved its high catalytic activity. In addition, the activity test results also suggested that Mn/γ-Al2O3 exhibited good SO2 tolerance.

The Mn/γ-Al2O3 catalyst exhibited excellent performance for NO conversion in the presence of a low ratio of O3/NO, which was due to the coexistence of Mn3+ and Mn4+ and abundant chemically adsorbed oxygen.  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