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.
2.
Over zeolite H‐ZSM‐5, the aromatics‐based hydrocarbon‐pool mechanism of methanol‐to‐olefins (MTO) reaction was studied by GC‐MS, solid‐state NMR spectroscopy, and theoretical calculations. Isotopic‐labeling experimental results demonstrated that polymethylbenzenes (MBs) are intimately correlated with the formation of olefin products in the initial stage. More importantly, three types of cyclopentenyl cations (1,3‐dimethylcyclopentenyl, 1,2,3‐trimethylcyclopentenyl, and 1,3,4‐trimethylcyclopentenyl cations) and a pentamethylbenzenium ion were for the first time identified by solid‐state NMR spectroscopy and DFT calculations under both co‐feeding ([13C6]benzene and methanol) conditions and typical MTO working (feeding [13C]methanol alone) conditions. The comparable reactivity of the MBs (from xylene to tetramethylbenzene) and the carbocations (trimethylcyclopentenyl and pentamethylbenzium ions) in the MTO reaction was revealed by 13C‐labeling experiments, evidencing that they work together through a paring mechanism to produce propene. The paring route in a full aromatics‐based catalytic cycle was also supported by theoretical DFT calculations.  相似文献   

3.
Hydrocarbon‐pool chemistry is important in methanol to olefins (MTO) conversion on acidic zeolite catalysts. The hydrocarbon‐pool (HP) species, such as methylbenzenes and cyclic carbocations, confined in zeolite channels during the reaction are essential in determining the reaction pathway. Herein, we experimentally demonstrate the formation of supramolecular reaction centers composed of organic hydrocarbon species and the inorganic zeolite framework in H‐ZSM‐5 zeolite by advanced 13C–27Al double‐resonance solid‐state NMR spectroscopy. Methylbenzenes and cyclic carbocations located near Brønsted acid/base sites form the supramolecular reaction centers in the zeolite channel. The internuclear spatial interaction/proximity between the 13C nuclei (associated with HP species) and the 27Al nuclei (associated with Brønsted acid/base sites) determines the reactivity of the HP species. The closer the HP species are to the zeolite framework Al, the higher their reactivity in the MTO reaction.  相似文献   

4.
《中国化学》2018,36(5):373-373
The cover picture shows the complexity of the reaction mechanism of zeolites catalyzed methanol‐to‐olefins (MTO) conversion. The MTO process plays a vital role in the production of light olefins from nonpetroleum resources. Despite of the successful industrialization of the MTO process in China, the detailed reaction mechanism is not yet well understood. The theoretical studies on the MTO hydrocarbon pool mechanism by the Group of Xie are summarized in the Chemistry Author Up Close by Xie et al. on page 381–386.

  相似文献   


5.
由于可以从非石油资源如煤、天然气、生物质等出发制备低碳烯烃,分子筛催化甲醇制烯烃(MTO)反应在学术界和工业界引起了广泛的研究兴趣. H-SAPO-34是目前表现优异性能的分子筛催化剂之一,其双烯(乙烯+丙烯)的选择性在80%以上,已经实现了工业化应用.为了提升MTO反应的选择性,以及调控乙烯丙烯的选择性之比,非常有必要从反应机理出发来优化设计新的催化剂.然而,由于MTO催化反应产物复杂多样,对MTO反应机理的认识还存在很大的争议.目前基本能够接受的是MTO催化反应沿着烃池机理进行.在此反应机理中,无机分子筛和有机烃池活性中心形成共催化剂,甲醇进攻有机活性中心生成烷基链,此烷基链断裂得到烯烃产物.目前提出的烃池活性中心主要包括多甲基苯和烯烃自身,它们分别沿着各自的循环反应网络(芳烃循环和烯烃循环)生成烯烃产物.有文献指出在H-ZSM-5分子筛中芳烃循环主要生成乙烯,而烯烃循环主要生成丙烯等产物.因此,系统研究分子筛结构对两条循环网络相对贡献程度的影响规律,从而阐述分子筛结构和MTO催化性能之间的关系具有重要的意义. H-SAPO-18是一类结构上与H-SAPO-34相类似的分子筛,其笼由八元环孔道互联.实验研究指出,其也具有优异的MTO催化性能.在本工作中,我们利用包含范德华相互作用校正的交换相关泛函(BEEF-vdW),系统研究了H-SAPO-18分子筛中的芳烃循环反应机理.所有计算用VASP程序包完成, H-SAPO-18用48T周期性结构模型表示.利用静态吸附和相互转化的自由能变化情况,我们首先确认了反应条件下H-SAPO-18中最稳定的多甲基苯的结构.计算结果指出,1,2,4,5-四甲基苯的吸附能最强,而六甲基苯是主要存在的多甲基苯组分.多甲基苯在分子筛孔道内的稳定性主要由两个相反的作用共同影响:范德华相互作用引起的吸引,以及分子筛孔道结构引起的排斥.在芳烃循环路线中,乙基侧链的增长是反应的关键基元步.吉布斯自由能分析指出芳烃循环路线中,在反应温度673 K下H-SAPO-18中的六甲基苯并不比五甲基苯,四甲基苯的活性高,这与H-SAPO-34分子筛中的结果相一致. H-SAPO-18中的四甲基苯、五甲基苯和六甲基苯的总吉布斯自由能垒分别是208,215,239 kJ/mol.六甲基苯循环路线所表现出的高反应能垒的一个原因,是由于分子筛几何限域效应引起的熵增加所致.通过与烯烃循环路线的动力学进行比较,本文芳烃循环路线动力学的工作可以为MTO催化反应机理的研究提供一些启示.  相似文献   

6.
The key step in the conversion of methane to polyolefins is the catalytic conversion of methanol to light olefins. The most recent formulations of a reaction mechanism for this process are based on the idea of a complex hydrocarbon‐pool network, in which certain organic species in the zeolite pores are methylated and from which light olefins are eliminated. Two major mechanisms have been proposed to date—the paring mechanism and the side‐chain mechanism—recently joined by a third, the alkene mechanism. Recently we succeeded in simulating a full catalytic cycle for the first of these in ZSM‐5, with inclusion of the zeolite framework and contents. In this paper, we will investigate crucial reaction steps of the second proposal (the side‐chain route) using both small and large zeolite cluster models of ZSM‐5. The deprotonation step, which forms an exocyclic double bond, depends crucially on the number and positioning of the other methyl groups but also on steric effects that are typical for the zeolite lattice. Because of steric considerations, we find exocyclic bond formation in the ortho position to the geminal methyl group to be more favourable than exocyclic bond formation in the para position. The side‐chain growth proceeds relatively easily but the major bottleneck is identified as subsequent de‐alkylation to produce ethene. These results suggest that the current formulation of the side‐chain route in ZSM‐5 may actually be a deactivating route to coke precursors rather than an active ethene‐producing hydrocarbon‐pool route. Other routes may be operating in alternative zeotype materials like the silico‐aluminophosphate SAPO‐34.  相似文献   

7.
The formation of hydrocarbon pool (HCP) species during methanol‐to‐olefin (MTO) and ethanol‐to‐olefin (ETO) processes have been studied on individual micron‐sized SAPO‐34 crystals with a combination of in situ UV/Vis, confocal fluorescence, and synchrotron‐based IR microspectroscopic techniques. With in situ UV/Vis microspectroscopy, the intensity changes of the λ=400 nm absorption band, ascribed to polyalkylated benzene (PAB) carbocations, have been monitored and fitted with a first‐order kinetics at low reaction temperatures. The calculated activation energy (Ea) for MTO, approximately 98 kJ mol?1, shows a strong correlation with the theoretical values for the methylation of aromatics. This provides evidence that methylation reactions are the rate‐determining steps for the formation of PAB. In contrast for ETO, the Ea value is approximately 60 kJ mol?1, which is comparable to the Ea values for the condensation of light olefins into aromatics. Confocal fluorescence microscopy demonstrates that during MTO the formation of the initial HCP species are concentrated in the outer rim of the SAPO‐34 crystal when the reaction temperature is at 600 K or lower, whereas larger HCP species are gradually formed inwards the crystal at higher temperatures. In the case of ETO, the observed egg‐white distribution of HCP at 509 K suggests that the ETO process is kinetically controlled, whereas the square‐shaped HCP distribution at 650 K is indicative of a diffusion‐controlled process. Finally, synchrotron‐based IR microspectroscopy revealed a higher degree of alkylation for aromatics for MTO as compared to ETO, whereas high reaction temperatures favor dealkylation processes for both the MTO and ETO processes.  相似文献   

8.
The methanol to olefins conversion over zeolite catalysts is a commercialized process to produce light olefins like ethene and propene but its mechanism is not well understood. We herein investigated the formation of ethene in the methanol to olefins reaction over the H‐ZSM‐5 zeolite. Three types of ethylcyclopentenyl carbocations, that is, the 1‐methyl‐3‐ethylcyclopentenyl, the 1,4‐dimethyl‐3‐ethylcyclopentenyl, and the 1,5‐dimethyl‐3‐ethylcyclopentenyl cation were unambiguously identified under working conditions by both solid‐state and liquid‐state NMR spectroscopy as well as GC‐MS analysis. These carbocations were found to be well correlated to ethene and lower methylbenzenes (xylene and trimethylbenzene). An aromatics‐based paring route provides rationale for the transformation of lower methylbenzenes to ethene through ethylcyclopentenyl cations as the key hydrocarbon‐pool intermediates.  相似文献   

9.
After a prolonged effort over many years, the route for the formation of a direct carbon?carbon (C?C) bond during the methanol‐to‐hydrocarbon (MTH) process has very recently been unveiled. However, the relevance of the “direct mechanism”‐derived molecules (that is, methyl acetate) during MTH, and subsequent transformation routes to the conventional hydrocarbon pool (HCP) species, are yet to be established. This important piece of the MTH chemistry puzzle is not only essential from a fundamental perspective, but is also important to maximize catalytic performance. The MTH process was probed over a commercially relevant H‐SAPO‐34 catalyst, using a combination of advanced solid‐state NMR spectroscopy and operando UV/Vis diffuse reflectance spectroscopy coupled to an on‐line mass spectrometer. Spectroscopic evidence is provided for the formation of (olefinic and aromatic) HCP species, which are indeed derived exclusively from the direct C?C bond‐containing acetyl group of methyl acetate. New mechanistic insights have been obtained from the MTH process, including the identification of hydrocarbon‐based co‐catalytic organic reaction centers.  相似文献   

10.
The catalytic activity of large zeolite H‐ZSM‐5 crystals in methanol (MTO) and ethanol‐to‐olefins (ETO) conversions was investigated and, using operando UV/Vis measurements, the catalytic activity and deactivation was correlated with the formation of coke. These findings were related to in situ single crystal UV/Vis and confocal fluorescence micro‐spectroscopy, allowing the observation of the spatiotemporal formation of intermediates and coke species during the MTO and ETO conversions. It was observed that rapid deactivation at elevated temperatures was due to the fast formation of aromatics at the periphery of the H‐ZSM‐5 crystals, which are transformed into more poly‐aromatic coke species at the external surface, preventing the diffusion of reactants and products into and out of the H‐ZSM‐5 crystal. Furthermore, we were able to correlate the operando UV/Vis spectroscopy results observed during catalytic testing with the single crystal in situ results.  相似文献   

11.
王传明  王仰东  谢在库 《催化学报》2018,39(7):1272-1279
低碳烯烃(乙烯、丙烯等)是重要的基本有机原料, 一般通过蒸汽裂解或催化裂解生成得到.基于中国的资源结构特点, 发展非石油资源路线合成低碳烯烃具有重要的战略意义. 其中从煤、天然气等资源出发, 通过甲醇合成低碳烯烃就提供了这样一条可替代的路线. 因此分子筛催化甲醇制烯烃(MTO)反应在过去几十年获得了广泛的关注和研究. 为了获得高的产物选择性, 一般要求MTO分子筛催化材料具有较小的孔道结构以及合适的笼结构, H-SAPO-34和H-SAPO-18分子筛就具有这样的空间结构特点. 但是MTO催化反应产物分布多样复杂, 因此需要深入认识MTO催化反应机理, 从而优化设计分子筛结构和反应条件.目前已经形成的共识认为, MTO催化反应沿着烃池反应机理进行, 但是烃池活性中心的结构还存在很多争议. 我们曾系统研究了H-SAPO-18分子筛中多甲基苯的分布, 以及催化MTO反应的芳烃循环路线, 指出多甲基苯路线的总吉布斯自由能垒高于200 kJ/mol (673 K). 本文以四甲基乙烯(TME)作为代表性的烯烃烃池活性中心, 系统研究了H-SAPO-18分子筛催化MTO反应的烯烃循环路线. TME循环路线的总吉布斯自由能垒不大于150 kJ/mol, 远小于芳烃循环的总能垒. 因此, 烯烃本身有很大可能是H-SAPO-18催化MTO反应的烃池活性中心. 我们也指出了芳烃循环和烯烃循环路线的相似性, 这包括基元反应的相似性和中间体结构的相似性. 或者可以说, 芳烃循环和烯烃循环路线机理上没有区别, 关键是为了得到具有烷基(侧)链的裂解前驱体, 最后通过裂解生成低碳烯烃. 在烯烃循环路线中, 产物选择性与裂解前驱体(高碳烯烃、碳正离子等)的分布以及裂解动力学有关. 计算发现生成乙烯和丙烯的裂解基元反应能垒与裂解前驱体的碳数之间存在线性关系. 本文进一步强调了分子筛催化MTO反应中烯烃活性中心的重要性, 并且清楚指出了烯烃循环和芳烃循环的机理相似性.  相似文献   

12.
Methanol‐to‐olefin (MTO) catalysis is a very active field of research because there is a wide variety of sometimes conflicting mechanistic proposals. An example is the ongoing discussion on the initial C?C bond formation from methanol during the induction period of the MTO process. By employing a combination of solid‐state NMR spectroscopy with UV/Vis diffuse reflectance spectroscopy and mass spectrometry on an active H‐SAPO‐34 catalyst, we provide spectroscopic evidence for the formation of surface acetate and methyl acetate, as well as dimethoxymethane during the MTO process. As a consequence, new insights in the formation of the first C?C bond are provided, suggesting a direct mechanism may be operative, at least in the early stages of the MTO reaction.  相似文献   

13.
甲醇制烯烃(MTO)作为一条由煤、天然气及生物质等含碳资源制备重要化学品的非石油路线,近年来备受人们关注。分子筛作为MTO的催化剂,其催化性能和MTO反应行为与其骨架结构和酸性特征密切相关,而认识这些关系对研发新型高效MTO催化剂和改进反应工艺具有重要意义。为此,研究简述了近年来有关甲醇转化制烯烃过程中分子筛催化活性及反应机理的理论和实验研究进展。重点讨论了不同分子筛在MTO过程中烃池物种、反应路线以及催化动力学方面的差异,分析了分子筛催化剂的骨架结构及酸性对其MTO催化性能的影响。  相似文献   

14.
Although industrialized, the mechanism for catalytic upgrading of bioethanol over solid‐acid catalysts (that is, the ethanol‐to‐hydrocarbons (ETH) reaction) has not yet been fully resolved. Moreover, mechanistic understanding of the ETH reaction relies heavily on its well‐known “sister‐reaction” the methanol‐to‐hydrocarbons (MTH) process. However, the MTH process possesses a C1‐entity reactant and cannot, therefore, shed any light on the homologation reaction sequence. The reaction and deactivation mechanism of the zeolite H‐ZSM‐5‐catalyzed ETH process was elucidated using a combination of complementary solid‐state NMR and operando UV/Vis diffuse reflectance spectroscopy, coupled with on‐line mass spectrometry. This approach establishes the existence of a homologation reaction sequence through analysis of the pattern of the identified reactive and deactivated species. Furthermore, and in contrast to the MTH process, the deficiency of any olefinic‐hydrocarbon pool species (that is, the olefin cycle) during the ETH process is also noted.  相似文献   

15.
Formation of coke in large H‐ZSM‐5 and H‐SAPO‐34 crystals during the methanol‐to‐olefin (MTO) reaction has been studied in a space‐ and time‐resolved manner. This has been made possible by applying a high‐temperature in‐situ cell in combination with micro‐spectroscopic techniques. The buildup of optically active carbonaceous species allows detection with UV/Vis microscopy, while a confocal fluorescence microscope in an upright configuration visualises the formation of coke molecules and their precursors inside the catalyst grains. In H‐ZSM‐5, coke is initially formed at the triangular crystal edges, in which straight channel openings reach directly the external crystal surface. At reaction temperatures ranging from 530 to 745 K, two absorption bands at around 415 and 550 nm were detected due to coke or its precursors. Confocal fluorescence microscopy reveals fluorescent carbonaceous species that initially form in the near‐surface area and gradually diffuse inwards the crystal in which internal intergrowth boundaries hinder a facile penetration for the more bulky aromatic compounds. In the case of H‐SAPO‐34 crystals, an absorption band at around 400 nm arises during the reaction. This band grows in intensity with time and then decreases if the reaction is carried out between 530 and 575 K, whereas at higher temperatures its intensity remains steady with time on stream. Formation of the fluorescent species during the course of the reaction is limited to the near‐surface region of the H‐SAPO‐34 crystals, thereby creating diffusion limitations for the coke front moving towards the middle of the crystal during the MTO reaction. The two applied micro‐spectroscopic techniques introduced allow us to distinguish between graphite‐like coke deposited on the external crystal surface and aromatic species formed inside the zeolite channels. The use of the methods can be extended to a wide variety of catalytic reactions and materials in which carbonaceous deposits are formed.  相似文献   

16.
An asymmetric tail‐to‐tail cross‐hydroalkenylation of vinylarenes with terminal olefins was achieved by catalysis with NiH complexes bearing chiral N‐heterocyclic carbenes (NHCs). The reaction provides branched gem‐disubstituted olefins with high enantioselectivity (up to 94 % ee) and chemoselectivity (cross/homo product ratio: up to 99:1). Electronic effects of the substituents on the vinylarenes and on the N‐aryl groups of the NHC ligands, but not a π,π‐stacking mechanism, assist the steric effect and influence the outcome of the cross‐hydroalkenylation.  相似文献   

17.
The catalytic, deactivation, and regeneration characteristics of large coffin‐shaped H‐ZSM‐5 crystals were investigated during the methanol‐to‐hydrocarbons (MTH) reaction at 350 and 500 °C. Online gas‐phase effluent analysis and examination of retained material thereof were used to explore the bulk properties of large coffin‐shaped zeolite H‐ZSM‐5 crystals in a fixed‐bed reactor to introduce them as model catalysts for the MTH reaction. These findings were related to observations made at the individual particle level by using polarization‐dependent UV‐visible microspectroscopy and mass spectrometric techniques after reaction in an in situ microspectroscopy reaction cell. Excellent agreement between the spectroscopic measurements and the analysis of hydrocarbon deposits by means of retained hydrocarbon analysis and time‐of‐flight secondary‐ion mass spectrometry of spent catalyst materials was observed. The obtained data reveal a shift towards more condensed coke deposits on the outer zeolite surface at higher reaction temperatures. Zeolites in the fixed‐bed reactor setup underwent more coke deposition than those reacted in the in situ microspectroscopy reaction cell. Regeneration studies of the large zeolite crystals were performed by oxidation in O2/inert gas mixtures at 550 °C. UV‐visible microspectroscopic measurements using the oligomerization of styrene derivatives as probe reaction indicated that the fraction of strong acid sites decreased during regeneration. This change was accompanied by a slight decrease in the initial conversion obtained after regeneration. H‐ZSM‐5 deactivated more rapidly at higher reaction temperature.  相似文献   

18.
Non‐oxidative dehydroaromatization of methane (MDA) is a promising catalytic process for direct valorization of natural gas to liquid hydrocarbons. The application of this reaction in practical technology is hindered by a lack of understanding about the mechanism and nature of the active sites in benchmark zeolite‐based Mo/ZSM‐5 catalysts, which precludes the solution of problems such as rapid catalyst deactivation. By applying spectroscopy and microscopy, it is shown that the active centers in Mo/ZSM‐5 are partially reduced single‐atom Mo sites stabilized by the zeolite framework. By combining a pulse reaction technique with isotope labeling of methane, MDA is shown to be governed by a hydrocarbon pool mechanism in which benzene is derived from secondary reactions of confined polyaromatic carbon species with the initial products of methane activation.  相似文献   

19.
Addition of 10 mol‐% of diphenyl diselenide to hydrostannylation reactions involving electron‐rich olefins results in a dramatic improvement in yield. For example, reaction of α‐{[(tert‐butyl)dimethylsilyl]oxy}styrene ( 1 ) with triphenylstannane ( 2a ; 1.1 equiv.) in the presence of PhSeSePh and 2,2′‐azobis[2‐methylpropanenitrile] (AIBN) affords {2‐{[(tert‐butyl)dimethylsilyl]oxy}‐2‐phenylethyl}triphenylstannane ( 3a ) in 95% yield after 2 h. This reaction presumably benefits, by the increased rate of H‐atom transfer, from the in situ generated polarity‐reversal catalyst, benzeneselenol.  相似文献   

20.
The reactions of fluorobenzene, 3-fluorotoluene, and three isomers of difluorotoluene, chlorobenzene, and bromobenzene with excesses of methanol were investigated on the large-pore catalysts HBeta (*BEA) and HSAPO-5 (AFI), and on the medium-pore HZSM-5 (MFI). Flow reactor studies in pulse mode with GC-MS detection revealed that the fluorobenzene derivatives were readily methylated at, for example, 375 degrees C, but not even pentamethylfluorobenzene was obviously active as a reaction center for methanol-to-olefin (MTO) catalysis. Carbon-labeling studies revealed that small amounts of methylbenzenes were formed by defluorination, and these aromatic hydrocarbons seemed to account for the small yields of olefins (and their secondary reaction products) observed. Loss of one fluorine was also evident in the products for one of the difluorotoluene isomers. On HSAPO-5 the activity order for ring-methylation of halobenzenes was F > Cl > Br. On HZSM-5, chlorobenzene and especially bromobenzene lost halogen through a route forming halomethane. These largely negative results will nevertheless be useful in testing theoretical models of the detailed reaction steps in the hydrocarbon pool mechanism for MTO catalysis.  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