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.
《中国化学》2017,35(10):1529-1539
A series of mesoporous Nb and Nb‐W oxides were employed as highly active solid acid catalysts for the conversion of glucose to 5‐hydroxymethylfurfural (HMF ). The results of solid state 31P MAS NMR spectroscopy with adsorbed trimethylphosphine as probe molecule show that the addition of W in niobium oxide increases the number of Brønsted acid sites and decreases the number of Lewis acid sites. The catalytic performance for Nb‐W oxides varied with the ratio of Brønsted to Lewis acid sites and high glucose conversion was observed over Nb5W5 and Nb7W3 oxides with high ratios of Brønsted to Lewis acid sites. All Nb‐W oxides show a relatively high selectivity of HMF , whereas no HMF forms over sulfuric acid due to its pure Brønsted acidity. The results indicate fast isomerization of glucose to fructose over Lewis acid sites followed by dehydration of fructose to HMF over Brønsted acid sites. Moreover, comparing to the reaction occurred in aqueous media, the 2‐butanol/H2O system enhances the HMF selectivity and stabilizes the activity of the catalysts which gives the highest HMF selectivity of 52% over Nb7W3 oxide. The 2‐butanol/H2O catalytic system can also be employed in conversion of sucrose, achieving HMF selectivity of 46% over Nb5W5 oxide.  相似文献   

2.
Operando X‐ray absorption experiments and density functional theory (DFT) calculations are reported that elucidate the role of copper redox chemistry in the selective catalytic reduction (SCR) of NO over Cu‐exchanged SSZ‐13. Catalysts prepared to contain only isolated, exchanged CuII ions evidence both CuII and CuI ions under standard SCR conditions at 473 K. Reactant cutoff experiments show that NO and NH3 together are necessary for CuII reduction to CuI. DFT calculations show that NO‐assisted NH3 dissociation is both energetically favorable and accounts for the observed CuII reduction. The calculations predict in situ generation of Brønsted sites proximal to CuI upon reduction, which we quantify in separate titration experiments. Both NO and O2 are necessary for oxidation of CuI to CuII, which DFT suggests to occur by a NO2 intermediate. Reaction of Cu‐bound NO2 with proximal NH4+ completes the catalytic cycle. N2 is produced in both reduction and oxidation half‐cycles.  相似文献   

3.
Supercritical drying of alginate gels is an efficient way to prepare aerogels with high surface area (>300 m2 · g−1). FTIR spectroscopy allows to monitor the adsorption of NH3 from the gas phase onto the acid sites of the alginate. Free carboxylic groups are effective Brønsted sites, whereas the divalent cations used in the ionotropic gelation present the properties of Lewis sites. The ratio between Brønsted and Lewis sites provides infomation on the role of pH in alginate gelation and suggests that non-buffered gelation by transition-metal cations is a mixed ionotropic-acid process.  相似文献   

4.
Bifunctional catalysts bearing two catalytic sites, Lewis acidic organometallic titanocene and Brønsted acidic COOH, have been assembled in situ from Cp2TiCl2 with carboxylic acid ligands, showing high catalytic activity over an intramolecular Mannich reaction towards synthesis of 2‐aryl‐2,3‐dihydroquinolin‐4(1H)‐ones. The determination of the bifunctional catalyst Cp2Ti(C8H4NO6)2 was elucidated by single X‐ray HR‐MS and investigation of catalytic behavior. In particular, masking the Brønsted acidic COOH catalytic site with dormant COOMe lowered the reaction yield greatly, indicating that two catalytic sites work together to maintain high catalytic efficiency.  相似文献   

5.
Bimetallic three‐dimensional amorphous mesoporous materials, Al‐Zr‐TUD‐1 materials, were synthesised by using a surfactant‐free, one‐pot procedure employing triethanolamine (TEA) as a complexing reagent. The amount of aluminium and zirconium was varied in order to study the effect of these metals on the Brønsted and Lewis acidity, as well as on the resulting catalytic activity of the material. The materials were characterised by various techniques, including elemental analysis, X‐ray diffraction, high‐resolution TEM, N2 physisorption, temperature‐programmed desorption (TPD) of NH3, and 27Al MAS NMR, XPS and FT‐IR spectroscopy using pyridine and CO as probe molecules. Al‐Zr‐TUD‐1 materials are mesoporous with surface areas ranging from 700–900 m2 g?1, an average pore size of around 4 nm and a pore volume of around 0.70 cm3 g?1. The synthesised Al‐Zr‐TUD‐1 materials were tested as catalyst materials in the Lewis acid catalysed Meerwein–Ponndorf–Verley reduction of 4‐tert‐butylcyclohexanone, the intermolecular Prins synthesis of nopol and in the intramolecular Prins cyclisation of citronellal. Although Al‐Zr‐TUD‐1 catalysts possess a lower amount of acid sites than their monometallic counterparts, according to TPD of NH3, these materials outperformed those of the monometallic Al‐TUD‐1 as well as Zr‐TUD‐1 in the Prins cyclisation of citronellal. This proves the existence of synergistic properties of Al‐Zr‐TUD‐1. Due to the intramolecular nature of the Prins cyclisation of citronellal, the hydrophilic surface of the catalyst as well as the presence of both Brønsted and Lewis acid sites synergy could be obtained with bimetallic Al‐Zr‐TUD‐1. Besides spectroscopic investigation of the active sites of the catalyst material a thorough testing of the catalyst in different types of reactions is crucial in identifying its specific active sites.  相似文献   

6.
The reaction of γ‐alumina with tetraethylorthosilicate (TEOS) vapor at low temperatures selectively yields monomeric SiOx species on the alumina surface. These isolated (‐AlO)3Si(OH) sites are characterized by PXRD, XPS, DRIFTS of adsorbed NH3, CO, and pyridine, and 29Si and 27Al DNP‐enhanced solid‐state NMR spectroscopy. The formation of isolated sites suggests that TEOS reacts preferentially at strong Lewis acid sites on the γ‐Al2O3 surface, functionalizing the surface with “mild” Brønsted acid sites. For liquid‐phase catalytic cyclohexanol dehydration, these SiOx sites exhibit up to 3.5‐fold higher specific activity than the parent alumina with identical selectivity.  相似文献   

7.
The CeO2, CeO2‐ZrO2 and CeO2‐WO3 catalysts were prepared by hydrothermal method and used to the selective catalytic reduction of NOx by NH3. The addition of ZrO2 or WO3 into CeO2 was favorable for pore structure, and then improved the number of active sites. Besides, the introduction of ZrO2 into CeO2 could improve the Lewis acid sites while WO3 could contribute to the generation of Brønsted and Lewis acid sites, which could improve the catalytic performance and N2 selectivity. The CeO2‐WO3 catalyst exhibited optimal catalytic activity with above 90% NOx conversion performance at 220–425 °C and approximately 100% N2 selectivity at 150–425 °C.  相似文献   

8.
张贺  邹永刚  彭悦 《催化学报》2017,38(1):160-167
由发电厂等固定源和柴油机等移动源排放的一氧化氮(NO)造成的环境污染问题日益严重.随着严苛的排放法规出台,NO排放控制技术受到越来越多关注.NH3选择性催化还原(SCR)技术是目前去除NO应用最为广泛的方法之一.商业催化剂V2O5-WO3/TiO2在300–400℃温度窗口内显示出优越的NO去除效率,但仍存在一些问题,如钒氧化物的毒性以及在高温时形成N2O和SO3.因此,开发出低钒或无钒的新型催化剂是解决上述问题的关键.CeO2和含铈材料是重要的催化剂载体,具有良好的还原能力和氧存储功能,因而广泛应用于催化领域.CeO2添加到商用催化剂中不仅可以降低钒用量,而且可以提高催化剂抗碱金属中毒能力.CeO2-WO3催化剂在200℃以上时比商用催化剂具有更宽的温度窗口,并展现出较高的抗SO2和碱金属中毒能力.CeO2-ZrO2催化剂通过添加过渡金属元素可以提升其SCR活性,在较宽的温度窗口内具有较高的催化活性.废气中SO2可导致催化剂失活,在实际应用中催化剂硫中毒是较为常见的催化剂失效原因.通常情况下,锰基和铁基催化剂最容易硫中毒.然而CeO2催化剂在硫酸化处理后却展现出良好的SCR活性.催化剂硫酸化主要包括气相、液相和前驱体硫化三种方法.三种方法各有异同,但在催化剂表面形成的硫物种都是SO42–.硫酸化可以增强Ce基催化剂的SCR活性,但是对于硫化引起的催化剂表面酸性、氧化还原性以及NO吸附脱附性质的详细研究报道较少.本文通过液相法对CeO2-ZrO2(CeZr)催化剂进行了硫酸化.XRD结果表明,硫酸化并未对催化剂结晶结构产生影响.TPD和TPR结果表明,硫酸化后催化剂(S-CeZr)表面酸性增强,但抑制了其氧化性.通过原位红外光谱技术系统研究了催化剂在SCR反应过程中表面物种的变化,结果发现,CeZr和S-CeZr的催化机理相同,不同的SCR活性主要是由表面酸性和氧化性引起的.CeO2基催化剂在不同温度窗口遵循不同反应机理.CeZr催化剂具有较强的氧化还原性,使其对NO和NH3具有很强的氧化能力,所以其在低于200℃时具有较好的SCR活性.而S-CeZr催化剂具有更多的Br?nsted酸性位,导致NO不易吸附在催化剂表面,所以其在低温时SCR活性较差,但在高温时(>200℃)具有优良的SCR活性.通过SCR活性和反应机理研究,发现在高温时(>200℃),表面酸性尤其是强酸Br?nsted酸性位在SCR反应中起到决定性作用;而在低温时(<200℃),酸性位对NH3分子较强的键合作用导致NH3难以被氧化,所以较强的酸性位对SCR活性具有抑制作用,而氧化还原性在低温时对SCR反应起到主要作用.同时,在高温时,较高的氧化性可使NH3被O2直接氧化,导致N2选择性降低.  相似文献   

9.
Commercial bioethanol can be readily converted into ethylene by a dehydration process using solid acids, such as Brønsted acidic H‐ZSM‐5 zeolites, and thus, it is an ideal candidate to replace petroleum and coal for the sustainable production of ethylene. Now, strong Lewis acidic extra‐framework three‐coordinate Al3+ species were introduced into H‐ZSM‐5 zeolites to improve their catalytic activity. Remarkably, Al3+ species working with Brønsted acid sites can accelerate ethanol dehydration at a much lower reaction temperature and shorten the unsteady‐state period within 1–2 h, compared to >9 h for those without Al3+ species, which can significantly enhance the ethanol dehydration efficiency and reduce the cost. The reaction mechanism, studied by solid‐state NMR, shows that strong Lewis acidic EFAl‐Al3+ species can collaborate with Brønsted acid sites and promote ethanol dehydration either directly or indirectly via an aromatics‐based cycle to produce ethylene.  相似文献   

10.
Vanadium‐containing saponite samples were synthesized in a one‐pot synthetic procedure with the aim of preparing samples for potential application as fillers for polymeric composites. These vanadium‐modified materials were prepared from an acid support by adopting a synthetic strategy that allowed us to introduce isolated structural V species (H/V‐SAP). The physicochemical properties of these materials were investigated by XRD analysis and by DR‐UV/Vis and FTIR spectroscopy of CO that was adsorbed at 100 K; these data were compared to those of a V‐modified saponite material that did not contain any Brønsted acid sites (Na/V‐SAP). The surface‐acid properties of both samples (together with the fully acidic H‐SAP material and the Na‐SAP solid) were studied in the catalytic isomerization of α‐pinene oxide. The V‐containing solids were tested in the oxidative dehydrogenation reaction of propene to evaluate their potential use as flame‐retardant fillers for polymer composites. The effect of tuning the presence of Lewis/Brønsted acid sites was carefully studied. The V‐containing saponite sample that contained a marked presence of Brønsted acid sites showed the most interesting performance in the oxidative dehydrogenation (ODH) reactions because they produced coke, even at 773 K. The catalytic data presented herein indicate that the H/V‐SAP material is potentially active as a flame‐retardant filler.  相似文献   

11.
Thermally stable Brønsted acid sites were generated on alumina‐supported niobium oxide (Nb2O5/Al2O3) by calcination at high temperatures, such as 1123 K. The results of structural characterization by using Fourier‐transform infrared (FTIR) spectroscopy, TEM, scanning transmission electron microscopy (STEM), and energy‐dispersive X‐ray (EDX) analysis indicated that the Nb2O5 monolayer domains were highly dispersed over alumina at low Nb2O5 loadings, such as 5 wt %, and no Brønsted acid sites were presents. The coverage of Nb2O5 monolayer domains over Al2O3 increased with increasing Nb2O5 loading and almost‐full coverage was obtained at a loading of 16 wt %. A sharp increase in the number of hydroxy groups, which acted as Brønsted acid sites, was observed at this loading level. The relationship between the acidic properties and the structure of the material suggested that the bridging hydroxy groups (Nb? O(H)? Nb), which were formed at the boundaries between the domains of the Nb2O5 monolayer, acted as thermally stable Brønsted acid sites.  相似文献   

12.
The triflimide iron(III) salt [Fe(NTf2)3] promotes the direct hydration of terminal and internal alkynes with very good Markovnikov regioselectivities and high yields. The enhanced carbophilic Lewis acidity of the FeIII cation mediated by the weakly‐coordinating triflimide anion is crucial for the catalytic activity. The iron(III) metal salt can be recycled in the form of the OPPh3/[Fe(NTf2)3] system with similar activity and selectivity. However, spectroscopic and kinetic studies show that [Fe(NTf2)3] hydrolyzes under the reaction conditions and that catalytically less active Brønsted species are formed, which points to a Lewis/Brønsted co‐catalysis. This triflimide‐based catalytic system is regioselective for the hydration of internal aryl‐alkynes and opens the door to a new synthetic route to alkyl ketophenones. As a proof of concept, the synthesis of two antipsychotics Haloperidol and Melperone, with general butyrophenone‐like structure, is shown.  相似文献   

13.
To gain insight into the underlying mechanisms of catalyst durability for the selective catalytic reduction (SCR) of NOx with an ammonia reductant, we employed scanning transmission X‐ray microscopy (STXM) to study Cu‐exchanged zeolites with the CHA and MFI framework structures before and after simulated 135 000‐mile aging. X‐ray absorption near‐edge structure (XANES) measurements were performed at the Al K‐ and Cu L‐edges. The local environment of framework Al, the oxidation state of Cu, and geometric changes were analyzed, showing a multi‐factor‐induced catalytic deactivation. In Cu‐exchanged MFI, a transformation of CuII to CuI and CuxOy was observed. We also found a spatial correlation between extra‐framework Al and deactivated Cu species near the surface of the zeolite as well as a weak positive correlation between the amount of CuI and tri‐coordinated Al. By inspecting both Al and Cu in fresh and aged Cu‐exchanged zeolites, we conclude that the importance of the preservation of isolated CuII sites trumps that of Brønsted acid sites for NH3‐SCR activity.  相似文献   

14.
In this work, sulfated zirconia (SZ) was prepared by a solvent-free method, ammonium sulfate ((NH4)2SO4) and sulfamic acid (NH2SO3H) were as sulfur sources, respectively. The resulting catalysts were characterized by powder X-ray diffraction (XRD), thermogravimetric analysis and differential scanning calorimetry (TGA-DSC), N2 adsorption–desorption isotherms, inductively coupled plasma (ICP) and pyridine-FTIR. The characterization of the catalysts revealed that SZ prepared with (NH4)2SO4 as sulfur source contained more Lewis acid sites and less Brønsted acid sites. The effects of sulfur sources on the catalytic activity of catalysts were studied. Catalytic testing showed that SZ prepared with (NH4)2SO4 as sulfur source had a better catalytic performance for removing trace olefins from aromatics.  相似文献   

15.
The catalytic properties of ZnAlVO mixed oxides derived from decavanadate-exchanged ZnAl–layered double hydroxide (LDH) precursors prepared by a sol–gel method (ZnAlVO–LDHx,y) were investigated in the oxidative dehydrogenation of propane and compared with those of supported catalysts obtained by conventional impregnation of NH4VO3 on ZnO (ZnVO-I,y) and ZnAlO mixed oxide (ZnAlVO-I,y) supports. The effects of composition and calcination time on the catalytic behavior were particularly examined. Higher propane conversions were achieved at higher vanadium content and calcination time of the precursors. The LDH-derived catalysts were the most active ones in all the temperature range studied (300–425 °C). The order of activity for propane conversion for the different catalyst families varies as ZnAlVO–LDHx,y > ZnAlVO-I > ZnVO-I and follows the strength of the Lewis and Brønsted acid sites determined by monitoring of pyridine adsorption by Fourier transform infrared spectroscopy, whereas the propene selectivities are close together in agreement with the similar densities of basic sites determined by CO2–temperature-programmed desorption measurements. It was indeed established that the acidity, rather than the nature of the crystalline phases, the reducibility, or the specific surface area of the samples, governs the catalytic activity.  相似文献   

16.
Diarylvinylidenecyclopropanes undergo a novel rearrangement in the presence of the Brønsted acid Tf2NH (Tf: trifluoromethanesulfonyl) to give the corresponding naphthalene derivatives in good to high yields upon heating, whereas in the presence of the Brønsted acid toluene‐4‐sulfonic acid (p‐TSA), the corresponding triene derivatives are afforded in moderate to good yields under mild conditions. Corresponding mechanistic studies on the basis of density functional theory (DFT) with the Gaussian03 program by using the B3LYP method have revealed that the pKa value of the Brønsted acid, as well as the entropy and solvent effects, plays a significant role in this reaction; these factors can discriminate the differences in the reactivity and regioselectivity among the Brønsted acids used in this reaction. In the presence of Lewis acid Sn(OTf)2, a butatrienecyclopane can produce the corresponding ring‐opened products in moderate yields.  相似文献   

17.
CF3-Containing α-arylpropargyl-α-amino carboxylates and α-arylpropargyl-α-amino phosphonates on treatment with the Brønsted or Lewis acids undergo cyclization involving the free NH2 group and triple bond to yield the corresponding dehydroprolines and their phosphorus analogs.  相似文献   

18.
Without using any acid or base catalyst, complexing agent or zirconium alkoxides, ZrO2–SiO2 mixed oxide with the ZrO2 content of 50 mol% was prepared by combination of sol–gel and alcohol-aqueous heating method using zirconyl nitrate and tetraethoxysilane as starting materials. The structural and surface acidic properties were characterized by FT-IR, XRD, NH3-TPD and pyridine adsorption FTIR. Compared with another mixed oxide with the same ZrO2 content prepared by mechanical grinding, the obtained ZrO2–SiO2 mixed oxide was homogeneously mixed in molecular level. The existed Zr–O–Si hetero-linkages strongly retarded the ZrO2 particle growth. The obtained mixed oxide maintained amorphous phase until it was calcined at 1,173 K for 3 h when crystallization of tetragonal zirconia took place. NH3-TPD and pyridine adsorption FTIR showed that both Brønsted and Lewis acidity were largely developed in the mixed oxide and most of the acidic sites belonged to the medium acidity. Because of the existence of abundant medium acidity, the mixed oxide showed catalytic activity for tetrahydrofuran polymerization. Furthermore, the produced poly tetramethylene ether glycol had moderate average molecular weight around 2,000. Neither the pure oxides nor the mixed oxide prepared by the mechanical grinding presented catalytic activity for this reaction.  相似文献   

19.
A series of Keggin‐type heteropolyacid‐based heterogeneous catalysts (Co‐/Fe‐/Cu‐POM‐octyl‐NH3‐SBA‐15) were synthesized via immobilized transition metal mono‐ substituted phosphotungstic acids (Co‐/Fe‐/Cu‐POM) on octyl‐amino‐co‐functionalized mesoporous silica SBA‐15 (octyl‐NH2‐SBA‐15). Characterization results indicated that Co‐/Fe‐/Cu‐POM units were highly dispersed in mesochannels of SBA‐15, and both types of Brønsted and Lewis acid sites existed in Co‐/Fe‐/Cu‐POM‐octyl‐NH3‐SBA‐15 catalysts. Co‐POM‐octyl‐NH3‐SBA‐15 catalyst showed excellent catalytic performance in H2O2‐mediated cyclohexene epoxidation with 83.8% of cyclohexene conversion, 92.8% of cyclohexene oxide selectivity, and 98/2 of epoxidation/allylic oxidation selectivity. The order of catalytic activity was Co‐POM‐octyl‐NH3‐SBA‐15 > Fe‐POM‐octyl‐NH3‐SBA‐15 > Cu‐POM‐octyl‐NH3‐SBA‐15. In order to obtain insights into the role of ‐octyl moieties during catalysis, an octyl‐free catalyst (Co‐POM‐NH3‐SBA‐15) was also synthesized. In comparison with Co‐POM‐NH3‐SBA‐15, Co‐POM‐octyl‐NH3‐SBA‐15 showed enhanced catalytic properties (viz. activity and selectivity) in cyclohexene epoxidation. Strong chemical bonding between ‐NH3+ anchored on the surface of SBA‐15 and heteropolyanions resulted in excellent stability of Co‐POM‐octyl‐NH3‐SBA‐15 catalyst, and it could be reused six times without considerable loss of activity.  相似文献   

20.
In this research, a novel organic–inorganic hybrid salt, namely, N1,N1,N2,N2‐tetramethyl‐N1,N2‐bis(sulfo)ethane‐1,2‐diaminium tetrachloroferrate ([TMBSED][FeCl4]2) was prepared and characterized by Fourier‐transform infrared spectroscopy (FT‐IR), energy‐dispersive X‐ray spectroscopy (EDX), elemental mapping, field emission scanning electron microscopy (FE‐SEM), X‐ray diffraction (XRD), thermal gravimetric (TG), differential thermal gravimetric (DTG), and vibrating‐sample magnetometry (VSM) analyses. Catalytic activity of the hybrid salt was tested for the synthesis of N,N′‐alkylidene bisamides through the reaction of benzamide (2 eq.) and aromatic aldehydes (1 eq.) under solvent‐free conditions in which the products were obtained in high yields and short reaction times. The catalyst was superior to many of the reported catalysts in terms of two or more of these factors: the reaction medium and temperature, yield, time, and turnover frequency (TOF). [TMBSED][FeCl4]2 is a Brønsted–Lewis acidic catalyst; there are two SO3H groups (as Brønsted acidic sites) and two tetrachloroferrate anions (as Lewis acidic sites) in its structure. Highly effectiveness of the catalyst for the synthesis of N,N′‐alkylidene bisamides can be attributed to synergy of the Brønsted and Lewis acids and also possessing two sites of each acid.  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