首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 156 毫秒
1.
骨架杂原子对Y型分子筛结构稳定性的影响   总被引:1,自引:0,他引:1       下载免费PDF全文
沈志虹  鞠雅娜  陈然  齐欣  李聃 《分子催化》2007,21(5):396-400
采用水热晶化法合成了含骨架杂原子的Y分子筛(BY、TiY、FeY),并对经过离子交换及超稳处理后得到的超稳型USY、USBY、USTiY和USFeY分子筛进行了水热处理和差示扫描量热测定,考察了杂原子B、Ti、Fe对Y分子筛结构稳定性的影响.结果表明;杂原子Ti的引人,明显提高了Y分子筛的热及水热稳定性;而杂原子B、Fe的引入则降低了Y分子筛的热及水热稳定性,特别是FeY分子筛的水热稳定性下降较多.  相似文献   

2.
FCC催化剂中REHY分子筛的结构与酸性   总被引:13,自引:0,他引:13  
 用XRD,N2吸附,NH3-TPD和Py-IR等手段对REHY分子筛进行了表征,并通过多晶XRD法测定了稀土离子在Y分子筛骨架外的分布. 结果表明,在含水条件下,定位于Y型分子筛β笼SⅠ′ 位的RE3+与骨架氧及定位于SⅡ′ 位的H2O配位,稳定了分子筛的骨架,减少了分子筛酸性中的最强酸部分;定位的RE3+通过极化定位的配位水,增加了分子筛的中强酸,减缓了RE3+取代H+所引起的酸量下降. 从结构测定可推测出,在FCC催化剂中Y型分子筛上RE3+的最佳量应为每个晶胞含11个稀土离子,并完全定位于β笼的SⅠ′ 位.  相似文献   

3.
采用液相离子交换法制备了不同稀土含量的Y型分子筛(HY、USY和NaY),研究了稀土铈(Ce)阳离子在Y型分子筛上吸附-脱附烃类分子(苯)过程中的作用机理与影响。通过X射线荧光光谱仪(XRF)、智能质量分析仪(IGA)、脱附指数的计算和巨正则蒙特卡罗模拟计算等多种表征计算方法,对引入稀土物种后,Y型分子筛对苯的饱和吸附量、吸附作用力、脱附热力学参数、苯在Y分子筛上的吸附势能分布及扩散行为等方面进行了研究。结果表明,Ce离子对苯在Y分子筛上脱附活化能的降低、吸附作用力的减弱以及吸附态由团聚态向分散态转变等方面具有显著影响,该作用构成了CeY分子筛催化剂在流化催化裂化(FCC)过程中能够优化轻质产品选择性的重要因素。  相似文献   

4.
采用X射线粉末衍射(XRD)、固体核磁共振(MAS NMR)、扫描电镜(SEM)等分析手段研究了金属离子(Y、Ga、Cr、Zn、Cu)对Y型分子筛水热稳定性的影响。结果表明,金属离子Y的引入能抑制Y型分子筛晶胞收缩,避免骨架铝的脱除,显著提高了分子筛的水热稳定性;金属离子Zn、Cu的引入在一定程度上能较好地稳定分子筛骨架结构,但是其耐高温水热性能较差,经过800℃水热处理后,分子筛的骨架结构基本倒塌;金属离子Ga、Cr不同于Y、Zn、Cu,不能很好地稳定分子筛骨架结构,其耐高温水热性能介于两者之间。  相似文献   

5.
多产低碳烯烃及柴油用分子筛的设计   总被引:1,自引:0,他引:1  
目前,催化裂化的原料大多是蜡油和减压渣油的混合油,或全部的常压渣油 .其相对分子量范围很宽,组成十分复杂 .为了分别对待大小不同的分子和不同碳氢比化合物的裂化性能,现代催化裂化催化剂应具有梯度孔结构与梯度酸中心的优化结构配置 .而目前催化裂化的主导分子筛 USY二次孔含量还不能满足原料油日趋变重的需求,并且滞留在分子筛孔 /笼内的非骨架铝限制了其选择性的进一步提高,因此有必要对其结构进一步修饰 .  本研究在 USY基础上合成新的催化裂化多产低碳烯烃和柴油用分子筛催化剂 .新型 USY基分子筛从晶体结构、孔容及孔…  相似文献   

6.
以NaY沸石为原料,用SiCl_4同晶取代法制备了不同硅铝比且具有高结晶度的八面沸石,进一步将样品交换成铵型。还制备了超稳Y样品及铝交换型八面沸石。研究结果表明,随骨架硅铝比提高,八面沸石的骨架稳定性、热稳定性及水热稳定性均得到很大改善。骨架外铝会使较低硅铝比样品的水热稳定性变差。钠型及铵型样品在不同硅铝比时,对稳定性有不同影响。在骨架硅铝比相近情况下,SiC1_4脱铝沸石(DNH_4Y)较水热法脱铝沸石(USY)有好得多的热稳定性,然而二者却具有相近的水热稳定性。  相似文献   

7.
采用X射线粉末衍射(XRD)、魔角旋转固体核磁共振(MAS NMR)、程序升温脱附(TPD)以及吡啶-傅里叶变换红外(Py-FTIR)光谱等方法研究了稀土离子对Y型分子筛结构稳定性和酸性的影响.结果表明:稀土离子的引入不仅显著增强了Y型分子筛的结构稳定性,而且调变了分子筛酸性,使分子筛中等强度酸中心数量增多,强酸中心数量减少.结合密度泛函理论(DFT)计算结果,从理论上阐述了这种机制:进入分子筛β笼l′位的稀土离子与骨架O原子相互作用,增加了Al-O作用力,提高了分子筛的结构稳定性,从而避免了骨架Al的脱除和非骨架Al物种的生成,进而导致稀土氧Y(REHY)分子筛的Br(o)nsted (B)酸强度较USY分子筛的弱;但与HY分子筛相比,稀土离子的引入使骨架O1负电荷减弱,Al-O1键长变短,O1-H作用力减弱,导致REHY分子筛的B酸强度较HY分子筛的强;在酸中心数量方面,REHY分子筛的B酸中心数量比USY分子筛的多,但比HY分子筛的少.  相似文献   

8.
 针对催化裂化 (FCC) 废催化剂细粉带来的环境污染和未开发利用的现状, 以此为原料合成了小晶粒 Y 型分子筛, 并采用 X 射线衍射、扫描电镜、29Si和27Al 固体核磁共振和激光粒度分析等技术对样品进行了表征. 结果表明, 以 FCC 废催化剂细粉为原料可以合成粒径为 200 nm 左右的 Y 型分子筛, 且制备过程中原料的活化方法和导向剂对所得样品结构影响很大. 对原料采用碱熔活化, 并在合成体系中加入导向剂, 所得 Y 型分子筛的相对结晶度提高, 比表面积增大, 粒度减小.  相似文献   

9.
分子筛作为一类重要的无机多孔晶体材料,由于其规整的孔道结构以及优异的酸性质等特点,在催化剂、吸附剂和离子交换床等许多领域有着重要而广泛的应用.而现代分子筛制备方法的发展主要得益于有机结构导向剂(OSDA)在分子筛合成中的广泛使用.但是,大部分OSDA都具有剧毒、价格昂贵、制备方法繁琐等缺点,因而限制了其大规模应用.高硅Y型分子筛的合成研究也面临同样的问题.Y型分子筛具有十二元环孔口和三维孔道结构,是目前催化裂化催化剂中的主要活性组分.目前,通过常规合成方法无法获得硅铝比大于6.0的Y型分子筛,无法满足石油化工对其酸性的要求.目前工业上主要通过后处理法得到高硅Y沸石,但该方法繁杂的后处理过程、不均匀的化学分布、大量损失的结晶度以及严重的环境污染促使人们开发直接合成高硅Y型分子筛的新方法以替代后处理过程.此外,使用OSDA一步法合成的高硅铝比Y型分子筛具有优异的热和水热稳定性.因此,使用OSDA一步直接合成高硅Y型分子筛在材料合成和催化领域一直备受关注.然而,目前尚未见关于绿色OSDA用于高硅Y型分子筛合成的报道.本研究首次将氢氧化胆碱或氯化胆碱作为一种新型、绿色、廉价的OSDA引入到高硅Y分子筛的合成凝胶体系,成功合成了高结晶度且硅铝比大于6.0的高硅Y型分子筛.实验详细考察了合成条件对硅铝比的影响,并采用XRD, XRF, NMR,TG以及N2物理吸附等表征手段研究了合成样品的物理化学性质.表征结果证明,胆碱阳离子作为一个稳定的OSDA存在于分子筛的孔结构中,并且取代了部分Na~+以平衡分子筛骨架的负电荷,因此胆碱的使用可使样品的硅铝比提高并具有更加优异的热稳定性和水热稳定性.实验确定了Na~+和OSDA~+在高硅Y分子筛合成中的竞争关系.大量的实验证据表明, Na~+进料比例对FAU骨架硅铝比有决定性的影响.首次提出采用氢氧根离子型OSDA是一种直接有效提高骨架硅铝比的方法.  相似文献   

10.
针对催化裂化(FCC)废催化剂的回收利用问题,提出了一种废催化剂再利用的方法,即以FCC废催化剂为铝源,合成时只补充部分硅源,采用自制的高效NaY沸石导向剂,水热合成NaY分子筛。同时,以普通的化工原料合成了对比试样Y型分子筛。讨论了不同的FCC废催化剂预处理方式对合成产物性能的影响,发现以经过碱熔活化处理的废催化剂为原料合成的Y分子筛拥有更高的结晶度和纯度。采用X射线衍射、热分析、程序升温脱附法(NH3-TPD)和N2静态容量吸附法对结晶产物和对比样品的晶体结构,热稳定性、酸性质、比表面积以及孔分布进行了表征。结果显示,以FCC废催化剂为原料完全可以合成出与普通原料性能接近的NaY分子筛。其BET比表面积可以达到615 m2·g-1,孔体积可达0.38 cm3·g-1,孔径集中在0.51 nm左右。  相似文献   

11.
Pd was loaded on ultra stable Y (USY) zeolites prepared by steaming NH(4)-Y zeolite under different conditions. Heck reactions were carried out over the prepared Pd/USY. We found that H? bubbling was effective in improving not only the catalytic activity of Pd/USY, but also that of other supported Pd catalysts and Pd(OAc)?. Moreover, the catalytic activity of Pd/USY could be optimized by choosing appropriate steaming conditions for the preparation of the USY zeolites; Pd loaded on USY prepared at 873 K with 100% H?O gave the highest activity (TOF = 61,000 h?1), which was higher than that of Pd loaded on other kinds of supports. The prepared Pd/USY catalysts were applicable to the Heck reactions using various kinds of substrates including bromo- and chloro-substituted aromatic and heteroaromatic compounds. Characterization of the acid properties of the USY zeolites revealed that the strong acid site (OH(strong)) generated as a result of steaming had a profound effect on the catalytic activity of Pd.  相似文献   

12.
用XPS考察了USY和FSY超稳Y型沸石及其起始原料NH_4Y沸石的表面组成,并结合Ar~+刻蚀技术与其体相组成进行了比较。结果表明,NH_4Y沸石的铝分布基本均匀一致,经超稳化处理后,沸石表面组成发生了很大的变化。用(NH_4)_2SiF_6液相脱铝补硅法制备的FSY表面呈缺铝特征,其表面铝原子浓度比体相低8~10%;而用传统水热法制备的USY具有典型的富铝表面,其表面铝原子浓度比体相高30%左右。USY表面富铝主要是由于USY“次表面”(Subsurface)附近的非骨架铝类迁移到表面所致。内层的非骨架铝类在常规水热超稳处理条件下一般则很难迁移,仍然留有在USY晶体内部。  相似文献   

13.
The preparation of different samples of vanadia supported on ultrastable zeolite (VO(x)/USY) is discussed. The samples were prepared in order to obtain highly dispersed V-species, avoiding the formation of crystalline vanadia and the destruction of the zeolite framework. Two methods were employed for preparing VO(x)/USY samples: an organic route using V(AcAc)3 and an inorganic route using NH4VO3. The characterization of the samples was performed with XRD, TPR, NH3-TPD, and N2 isotherms. From these results it is concluded that when VO(x) is supported on the surface of USY from acidic aqueous solution of ammonium metavanadate, the destruction of the zeolite framework is accomplished. For higher pH values in the impregnating solution, undesired V2O5 is formed on the USY surface. On the other hand, VO(x)/USY prepared from the organic precursor shows no destruction of the USY structure. In addition, highly dispersed VO(x) are formed, though for relatively high V loadings (6%) an obstruction of the zeolite windows takes place. The samples are tested as catalysts for gas phase dehydrogenation of n-butane to olefins. The catalysts prepared from NH4VO3 are almost inactive for the reaction. On the other hand, both samples prepared from V(AcAc)3 present initial conversion levels in the 8-12% range. However, the selectivity depends on the V loading, the catalysts with 6% loading being the most selective (75%). The catalytic patterns of the samples (activity and selectivity) are in agreement with the physicochemical features of the VO(x)/USY surface.  相似文献   

14.
Y沸石脱铝有多种方法。由于在脱铝过程中发生水解、聚合、重排等反应,分子筛的孔结构将发生很大的变化,如形成空穴,甚至可导致骨架结构单元的崩塌,形成二次孔。由于不同的脱铝方法的脱铝机理不同,即使它们的脱铝程度一致,产生的孔结构也将不同,这必将导致它们的酸性质和催化性能的差异。目前关于Y沸石孔结构的研究报导还很少,尤其对SSY,US-SSY沸石孔结构的报导更为鲜见。本文用低温N_2吸附-脱附等温线和层厚t等方法,对四种不同方法制备的脱铝Y沸石(USY,DAY,SSY和US-SSY)的微孔、二次孔及其孔径分布(2.0~60.0nm)进行了初步研究,得到了一些有用的信息,为进一步研究其酸性质和催化性能打下了基础。  相似文献   

15.
探索了非缓冲条件下,用酒石酸对超稳Y型分子筛进一步脱铝改性的可能性。系统地考察了反应时间、反应温度、酒石酸溶液加入量、溶剂加入量、滴加速度及溶液pH值对处理过程的影响,并对改性后的分子筛样品采用X射线衍射测试、差热分析等手段进行了结构与性能的表征。结果表明,所用的分子筛脱铝方法在合适的处理条件下能制备出硅铝比达12左右的超稳Y型分子筛。改性后的超稳Y型分子筛的晶体结构热稳定性随着骨架硅铝比的提高而增加。  相似文献   

16.
Fluorescence methodologies have been utilized to examine micropolarity, intramolecular motion, and singlet quenching in the intraparticle void volume of zeolites X, Y, and ultrastable Y (USY) interfaced with bathing polar solvents. Micropolarity was assessed from the 3-to-1 band ratio (III/I) of the fluorescence spectrum of pyrene (PY) and from lambda(max) of the fluorescence spectrum of 1-pyrenecarboxaldehyde (1-PCA). In zeolites bathed in anhydrous solvents, both PY and 1-PCA reported increased micropolarity according to the trend USY < bulk solvent < NaX approximately NaY. For example, in NaY (USY), III/I ranged from 0.44 (0.98) in acetonitrile to 0.52 (1.34) in n-hexanol, compared to 0.60, 1.06, and 1.62 in bulk acetonitrile (ACN), n-hexanol, and n-hexane, respectively. The polarity studies reveal that the ionic nature of NaX and NaY and the hydrophobic nature of USY strongly influence the microenvironment of the arene despite the presence of desorbing polar solvents. Constraints on intramolecular motion were examined in polar-solvated NaX through measurements of the fluorescence lifetime of trans-stilbene. Lifetimes ranged from 113 ps in NaX-ACN to 671 ps in NaX-tert-butyl alcohol. The latter value is close to that observed in bulk glycerol. Diffusion-controlled quenching of PY fluorescence by O2 and a series of nitrocompounds dissolved in solvents bathing the zeolite was examined by a time-resolved approach. For all of the quenchers and solvents studied, quenching was more efficient in USY compared to NaX and NaY. Interestingly, the rate of O2 quenching in USY-MeOH was only 12 times lower than that in bulk MeOH. In contrast, in NaY-MeOH and NaX-MeOH the rate of O2 quenching was too low to be measured. The rate constants in these systems were therefore taken as the rate constant for diffusion-controlled quenching of trapped electrons measured previously. These values were 600 times and 10(5) times lower than the rate of fluorescence quenching in USY-MeOH, respectively. The O2 quenching studies show that dispersive interactions of polar solvents with the cavity walls dominate in USY because of the hydrophobic nature of the USY surface. In NaX and NaY, stronger ion-dipole and hydrogen bonding interactions dominate and lead to more restricted access and lowered quenching efficiency. Perrin (or static) quenching of pyrene fluorescence was also examined to infer the concentration of nitromethane (NM) in the void volume of NaX and NaY bathed in MeOH, ACN, or H2O. The results indicate that access of NM to the interior of NaY is more inhibited in ACN compared to MeOH, presumably because of the higher dipole moment of ACN and its resulting stronger association with the zeolite surface. At similar levels of static quenching equated to a similar NM concentration in the zeolite, dynamic quenching by NM varied by no more than a factor of 2 in all systems compared. This implies that the rate of NM diffusion in solvated zeolite interiors is similar regardless of zeolite or solvent properties. In contrast to O2 diffusion in zeolites, NM exhibits a high dipole moment and can therefore migrate through polar-solvated zeolite apertures by adsorbing to the zeolite. Overall, the results of this study show a close relationship between the behavior of probes and quenchers in the confines of polar-solvated zeolite interiors and the chemical properties of the zeolite. Differences between weakly and strongly interacting surfaces are revealed clearly in the results.  相似文献   

17.
Structures of Pd/zeolites immersed in solvents were measured by in situ X‐ray absorption fine structure (XAFS). Systematic studies revealed that the selection of an appropriate support (USY‐zeolite), thermal treatment temperature of USY, solvent (o‐xylene), H2 partial pressure (6 %), and the use of a Pd amine complex affect the structure of Pd. As a result, we found that monomeric Pd can be obtained in the USY support with H2 bubbling in o‐xylene. The structural properties of Pd correlate well with its catalytic performance in the Suzuki–Miyaura coupling reactions; a very high TON of up to 11 000 000 was obtained over the monomeric Pd.  相似文献   

18.
The structures of alkali‐exchanged faujasite (X–FAU, X = Li+ or Na+ ion) and ZSM‐5 (Li–ZSM‐5) zeolites and their interactions with ethylene have been investigated by means of quantum cluster and embedded cluster approaches at the B3LYP/6‐31G(d, p) level of theory. Inclusion of the Madelung potential from the zeolite framework has a significant effect on the structure and interaction energies of the adsorption complexes and leads to differentiation of different types of zeolites (ZSM‐5 and FAU) that cannot be drawn from a typical quantum cluster model, H3SiO(X)Al(OH)2OSiH3. The Li–ZSM‐5 zeolite is predicted to have a higher Lewis acidity and thus higher ethylene adsorption energy than the Li–FAU zeolites (16.4 vs. 14.4 kcal/mol), in good agreement with the known acidity trend of these two zeolites. On the other hand, the cluster models give virtually the same adsorption energies for both zeolite complexes (8.9 vs. 9.1 kcal/mol). For the larger cation‐exchanged Na–FAU complex, the adsorption energy (11.6 kcal/mol) is predicted to be lower than that of Li–FAU zeolites, which compares well with the experimental estimate of about 9.6 kcal/mol for ethylene adsorption on a less acidic Na–X zeolite. © 2003 Wiley Periodicals, Inc. Int J Quantum Chem 94: 333–340, 2003  相似文献   

19.
We outline two methodologies to selectively characterize the Brønsted acidity of the external surface of FAU-type zeolites by IR and NMR spectroscopy of adsorbed basic probe molecules. The challenge and goal are to develop reliable and quantitative IR and NMR methodologies to investigate the accessibility of acidic sites in the large pore FAU-type zeolite Y and its mesoporous derivatives often referred to as ultra-stable Y (USY). The accessibility of their Brønsted acid sites to probe molecules (n-alkylamines, n-alkylpyridines, n-alkylphosphine- and phenylphosphine-oxides) of different molecular sizes is quantitatively monitored either by IR or 31P NMR spectroscopy. It is now possible, for the first time to quantitatively discriminate between the Brønsted acidity located in the microporosity and on the external surface of large pore zeolites. For instance, the number of external acid sites on a Y (LZY-64) zeolite represents 2 % of its total acid sites while that of a USY (CBV760) represents 4 % while the latter has a much lower framework Si/Al ratio.  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