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.
2,6-DIPN is a precursor of 2,6-naphthalene dicarboxylic acid (2,6-NDA) which is an important monomer for making advanced polyester materials and liquid crystalline polymers. The more advanced method for preparing 2,6-NDA is the direct oxidation of 2,6-dialkylnaphthalene (2,6-DAN). Since isopropyl group is more hindered than methyl and ethyl group, a β-selective isopropylation could be expected in case of using shape selective zeolite catalysts instead of the conventional Friedel-Crafts alkylation. A particularly preferred catalyst is synthetic Mordenite having a specific Si/Al ratio, which results in a high selectivity of 2,6-DIPN. It is more beneficial to the separation and purification of 2,6-DIPN from its isomers. In this paper the study on isopropylation of naphthalene over various zeolites and different dealuminated H-mordenites (HM) is reported.  相似文献   

2.
Recent development of advanced polymer materials such as polyethylene naphthalate, polybutylene naphthalate and liquid crystalline polymers has created an increasing demand for 2,6-dialkylnaphthalene, which has spurred interest in shape-selective naphthalene alkylation. This work deals with mordenite-catalyzed shape-selective naphthalene isopropylation to produce 2,6-diisopropylnaphthalene (2,6-DIPN). Effects of dealumination of mordenite on the structural and acidic characteristics and on the shape selectivity and activity were examined by physicochemical analysis, TPD, solid-state 27Al and 29Si MAS NMR, XRD, as well as catalytic alkylation reactions. Dealumination removes octahedral Al species as well as tetrahedral Al species, decreases the unit cell dimensions and reduces the number of strong acid sites in mordenites. Proper dealumination can improve selectivity to 2,6-DIPN from 33 to 61 % and significantly increases 2,6/2,7 ratio. Improved selectivity to 2,6-DIPN upon proper dealumination was attributed to the decrease in mordenite acidity, reduction in unit cell dimension and removal of some strong acid sites. However, neither the change in selectivity nor that in activity is a simple function of dealumination degree or SiO2/Al2O3 ratio. Some minor difference in the apparent framework SiO2/Al2O3 ratio can result in a major difference in activity or selectivity. There exist optimum conditions of dealumination as well as optimum reaction conditions for achieving higher selectivity to 2,6-DIPN.  相似文献   

3.
煤焦油萘择形异丙基化产物组成的气相色谱/质谱分析   总被引:3,自引:0,他引:3  
 采用毛细管气相色谱 /质谱 (CGC/MS)联用技术 ,经计算机质谱库检索并结合人工谱图解析 ,研究了在不同沸石分子筛催化剂上煤焦油萘择形异丙基化产物的组成。研究结果表明 ,煤焦油萘异丙基化产物主要由一取代萘、二取代萘以及少量三取代萘组成 ;除 HZSM-5活性较低以外 ,其它催化剂均有较高的反应活性 ,其中 HY和 Hβ反应活性最高 ,萘转化率分别达 75 .1 5 %和 5 5 .96% ,但β位产物的选择性较差 ;ZM类催化剂既有一定的活性 ,又有较高的β选择性 ,是萘择形异丙基化反应理想的备选催化剂。  相似文献   

4.
Pulmonary toxicity of naphthalene (NAP), 2-methylnaphthalene (2-MN), 2-isopropylnaphthalene (2-IPN) and 2,6-diisopropylnaphthalene (2,6-DIPN) was studied in mice. Twenty four h after the intraperitoneal (i.p.) administration of NAP (200 mg/kg (1.6 mmol] or 2-MN (400 mg/kg (2.8 mmol], pulmonary damage was detected. Prior treatment with diethyl maleate resulted in enhancement of NAP and 2-MN-induced bronchiolar damage. In contrast to the effects of NAP and 2-MN, injections of 2-IPN (3000 mg (17.6 mmol)/kg) and 2,6-DIPN (3000 mg (14.2 mmol)/kg) did not cause detectable pulmonary damage. Injections of NAP and 2-MN caused considerable depletion of pulmonary reduced glutathione (GSH), while injections of 2-IPN and 2,6-DIPN caused only a slight depletion. There were general decreases in the binding of the compounds to lung slices with increasing number of carbons of the alkyl substituent. Pretreatment with a cytochrome P-450 inducer (beta-naphthoflavone) increased the binding of NAP, 2-MN, and 2-IPN to lung slices. Treatments with NAP, 2-MN, 2-IPN and 2,6-DIPN did not affect the lipid peroxidation or phospholipid contents in the lung. These results suggest that the difference in pulmonary toxicity among NAP, 2-MN, 2-IPN, and 2,6-DIPN may be dependent on the ability of these compounds to irreversibly bind to lung tissue.  相似文献   

5.
Experimental and theoretical B3PW91/6-31G* spectra of diisopropylnaphthalene (DIPN) were compared. For the 1,3- and 2,6-DIPN isomers, which were isolated as pure compounds, the theoretical IR spectra were scaled down and were shown to fit the experimental spectra very well. The same scaling factor was used for comparison theoretical and experimental spectra of isomers present in unresolved mixtures of isomers, i.e. 1,4-, 1,5-, 1,6-, 1,7-, and 2,7-DIPNs. For three isomers, 1,2-, 1,8-, and 2,3-DIPN, the experimental IR spectra, unknown so far, were predicted.  相似文献   

6.
2,6-Diisopropylnaphthalene (2,6-DIPN), as the precursor of important monomer 2,6-naphthalene dicarboxylic acid, could be produced by the shape-selective isopropylation of naphthalene with propene resulting in an isomeric mixture having different alkylation levels. Since the boiling points of DIPNs were very close and the differences of melting points in-between isomers were quite distinctive, the static melt crystallization was applied to separate and purify 2,6-DIPN from its isomers. 2,6-DIPN with purity ≥99% was produced through a process of three stages: crystallization→sweating→melting. The phase diagram of 2,6-DIPN-2,7-DIPN binary system was plotted to opti- mize the temperature control of crystallization. By repeated crystallization of melts with different concentration levels, the yield of pure 2,6-DIPN could be enhanced to 87%. No solvent was necessary. Keywords 2,6-diisopropylnaphthalene, static melt crystallization, crystallization rate, sweating rate, eutectic temperature  相似文献   

7.
Isopropylation of naphthalene over modified-mordenite zeolites   总被引:2,自引:0,他引:2  
Isopropylation of naphthalene with propylene was carried out over Ce-, Si-, La-, and Mg-modified mordenite zeolites. The reaction was performed in a 500 mL batch reactor at 300°C, 8 kg/cm2, and naphthalene to catalyst weight ratio of 10:1, with excess propylene. It was attempted to eliminate the external acid sites of zeolites by these modifications. The results indicate that the CVD silanation and magnesium or cerium modification with Mg/Al or Ce/Al atomic ratios less than 0.5 can improve the 2,6-DIPN selectivity of mordenites. The improvement of catalyst performance is ascribed to the selective deactivation of acid sites on the external surface of HM zeolites.  相似文献   

8.
2,6-Dimethylnaphthalene is an important dimethylnaphthalene isomer which can be used in the production of polyethylene naphthalate. The novelty of this study is to reveal Langmuir–Hinshelwood and Eley–Rideal reaction rate equations for the methylation of naphthalene over Fe/ZSM-5 zeolite catalysts besides the proposed reaction rate equation. To investigate the kinetics and mechanisms of naphthalene methylation, the methylation experiments were carried out in a gas–solid catalytic fixed-bed reactor in the presence of Fe/ZSM-5 zeolite catalysts at two different temperatures (450 and 500 °C) and five different weight hourly space velocities (0.5, 1.0, 1.5, 2.0, 2.5, 3.0 h?1). A naphthalene:methanol:1,2,4-trimethyl benzene mixture having a 1:3:10 molar ratio was used as a feed stream. The methylation products were identified by using GC–MS. For the methylation kinetics of naphthalene, the reaction rates depending on the naphthalene and methanol concentration were determined. Furthermore, the effects of temperature and weight hourly space velocity on the conversion of naphthalene, the selectivity of 2,6-dimethylnaphthalene, and the ratio of 2-methylnaphthalene/1-methylnaphthalene were determined. The results of this study demonstrate that the Langmuir–Hinshelwood reaction mechanism for naphthalene methylation is more compatible at 450 °C and the Eley–Rideal reaction mechanism at 500 °C. Moreover, in addition to 2,6-dimethylnaphthalene, other dimethylnaphthalene and tri-methylnaphthalene isomers were formed in the methylation of naphthalene. The conversion of naphthalene reached approximately 70%. Moreover, the highest selectivity of 2,6-dimethylnaphthalene was almost 40%. The ratios of 2-methylnaphthalene/1-methylnaphalene demonstrate that the methylation of naphthalene to 2-methynaphthalene is much higher than to 1-methynaphthalene.  相似文献   

9.
2,6-Dimethylnaphthalene (2,6-DMN) is one of the crucial intermediates for the synthesis of polybutylenenaphthalate and polyethylene naphthalate (PEN). The complex synthesis procedure and the high cost of 2,6-DMN production significantly reduce the commercialisation of PEN even though PEN demonstrates superior properties compared with polyethylene terephthalate. 2,6-DMN can be produced by methylation of 2-methylnaphthalene (2-MN) and/or naphthalene, disproportionation of 2-MN, and/or isomerisation of dimethylnaphthalenes (DMNs). In this study, synthesis of 2,6-triad DMN isomers consisting of 2,6-DMN, 1,6-DMN, and 1,5-DMN have been investigated with the disproportionation of 2-MN over unmodified and Zr-modified mesoporous MCM-41 zeolite catalysts. In contrast to other DMN isomers, both 1,5-DMN and 1,6-DMN can be effectively isomerised to be profitable 2,6-DMN. The disproportionation of 2-MN experiments were carried out in a catalytic fixed-bed reactor in the presence of 1 g of catalyst at a temperature range of 350–500 °C and weight hourly space velocity between 1 to 3 h?1. The results demonstrated that mesoporous MCM-41 zeolite catalyst has a selective pore shape for 2,6-triad DMN isomers, which may allow a decrease in the production cost of 2,6-DMN. Additionally, 2,6-DMN was successfully synthesised by the disproportionation of 2-MN over MCM-41 zeolite catalyst. Furthermore, both the conversion of 2-MN and the selectivity of 2,6-DMN were considerably enhanced by the Zr impregnation on MCM-41.  相似文献   

10.
《Fluid Phase Equilibria》1999,166(1):101-110
An equation proposed by Darken, including the thermodynamic factor and tracer diffusion coefficients of solvent and solute, was adopted to correlate the diffusion coefficients for naphthalene and dimethylnaphthalene (DMN) isomers in supercritical carbon dioxide and the correlated results were compared with the experimental data. IML equation of state with mixing rules and combining rules containing two adjustable interaction parameters were used for calculation of the thermodynamic factor. By using the interaction parameters adjusted to the solubility data, the concentration dependence of diffusion coefficients and their anomaly near the critical point of carbon dioxide can be quantitatively represented. In order to improve the reliability of experimental results, some re-measured diffusion coefficient data for naphthalene, 2,6- and 2,7-DMN at 308.2 K, and new data for naphthalene at 318.2 K and for 2,3-DMN at 308.2 K are presented.  相似文献   

11.
《Liquid crystals》1999,26(12):1771-1776
The synthesis and phase behaviour of a series of broadened oligoaryls, in which replacement of one or two of the phenyl groups of bi- and ter-phenyl mesogens has been made by 1,4-substituted naphthalene, is reported. The novel materials, which are not mesogenic, can be compared with the liquid crystalline 2,6-substituted isomers.  相似文献   

12.
Steric interaction of reagents with zeolites was studied in isopropylation, sec‐butylation, and tert‐butylation of naphthalene (NP) over several large‐pore zeolites to elucidate the mechanism of selective catalysis. Selectivities for dialkylnaphthalene (DAN) isomers were influenced by the type of zeolite and bulkiness of alkylating agent. Selective formation of β,β‐ and 2,6‐diisopropylnaphthalene (DIPN) occurred only over H‐mordenite (MOR) in the isopropylation of NP using propene; bulky transition states of α,α‐ and α,β‐DIPN are excluded because of steric restriction by the channels, resulting in selective formation of β,β‐ and 2,6‐DIPN. However, low selectivities for β,β‐ and 2,6‐DIPN were observed over the zeolites, SSZ‐24 (AFI), SSZ‐55 (ATS), and SSZ‐42 (IFR) with 12‐membered‐ring (12‐MR) pore entrances of one‐dimensional channels, CIT‐5 (CFI), UTD‐1 (DON), and SSZ‐53 (SFH) with 14‐membered‐ring (14‐MR) pore entrances of one‐dimensional channels, and Y‐zeolite (FAU), zeolite β (BEA), and CIT‐1 (CON) with 12‐MR pore entrances of three‐dimensional channels, because their channels are too large for the exclusion of bulky isomers. Catalysis over these zeolites occurs under kinetic and/or thermodynamic control, resulting in predominant formation of α,α‐ and α,β‐DIPN at lower temperatures and an increase of the stable isomer β,β‐DIPN at higher temperatures. The selectivities for β,β‐ and 2,6‐DAN were enhanced with the increase in bulkiness of alkylating agents: 1‐butene for sec‐butylation and 2‐methylpropene for tert‐butylation. In particular, β,β‐di‐tert‐butylnaphthalene (DTBN) was selectively formed in the tert‐butylation. The selectivities for β,β‐ and 2,6‐DAN were enhanced even in large channels: the transition states of the least bulky isomers only fit the channels because other bulky isomers are excluded by steric restriction of the channels. However, tert‐butylation over FAU, BEA, and CON had selectivities for 2,6‐DTBN of around 50–60%, although selectivities for β,β‐DTBN were almost 100% selectivity; these zeolites can hardly recognize the differences between 2,6‐ and 2,7‐DTBN. The results indicate that the fitting of the least bulky isomers to zeolite channels, leading to the exclusion of other bulky isomers, is essential for highly shape‐selective catalysis.  相似文献   

13.
New methods were proposed for synthesizing 1,4,5,8-tetrakis(dimethylamino)naphthalene with an overall yield of 4 to 12% to replace the known procedure ensuring an overall yield of 2%. Catalytic hydrogenation was shown to be inapplicable for preparation of polyaminonaphthalenes from nitro compounds having 3 or 4 nitro gruops in the -positions. Nucleophilic amination of 1,5-dinitronaphthalene in the system NH2OH/NaOH/MeOH yields 1-amino-4-nitronaphthalene. The nitration of 1,5-bis(p-tolylsulfonylamino)naphthalene leads to formation of 2,6-dinitro rather than 4,8-dinitro derivative, as it was believed formerly. This was confirmed by transformation of the latter into 1,2,5,6-tetrakis(dimethylamino)naphthalene. 3-Nitro, 2,6-dinitro, 2,6-diamino, and 2,4,6,8-tetranitro derivatives of 1,5-bis(dimethylamino)naphthalene, nitro and amino derivatives of 1,4,5-tris(dimethylamino)naphthalene, and 4,5-diamino-1,8-bis(methylamino)naphthalene were synthesized. By treatment with perchloric acid 1,4,5,8-tetrakis(dimethylamino)naphthalene was oxidized to 2,3-dihydroperimidinium salt.  相似文献   

14.
An automated extractor-concentrator was used to extract metabolites of naphthalene, 2,6-dimethylnaphthalene, and benzo[a]pyrene from serum, bile and liver homogenate of rainbow trout (Salmo gairdneri). The extracts were analyzed by reversed-phase high-performance liquid chromatography (HPLC) with fluorescence detection. Recoveries of naphthalene and 2,6-dimethylnaphthalene metabolites from all matrices were generally greater than 90%; however, the recoveries of benzo[a]pyrene metabolites from serum ranged from 37-99%. In addition, conjugated metabolites of polycyclic aromatic hydrocarbons (PAHs) were separated from non-conjugated metabolites and parent PAHs by using two diol columns with normal-phase HPLC. The extraction and separation techniques were also applied to isolate metabolites in samples from fish fed 2,6-dimethylnaphthalene.  相似文献   

15.
Abstract— The quenching of intramolecular exciplex fluorescence in 2,6-bis(methylamino)naphtha]ene, (II), 2,6-bis(/V-methyl, methylamino)naphthalene, (III) and 2,6-bis(N-dimethy], methylamino)naphthalene, (IV) by water was investigated. Exciplex fluorescence intensity decreased continuously with increase in water concentration; while monomer fluorescence intensity remained constant up to 25% of water, followed by a rapid increase in intensity.
Absorption spectra showed that specific interaction occurred between water and ground state amine molecules at high concentrations (25% by volume) of water. These observations were interpreted in terms of a static quenching involving the interaction of water with the ground-state amine and a dynamic quenching involving the interactions of water with the exciplex. Our observations explain the greater efficiency of exciplex-fluorescence quenching by protic polar solvents.  相似文献   

16.
Two types of naphthalene trimers linked by 1,3,4-oxadiazole spacers were synthesized and investigated for their physical and electronic properties. 2,6- and 2,7-isomers on central naphthalene moieties were obtained in the forms of pale yellow solids and colorless crystals, respectively. The melting point of the 2,6-isomer was higher than that of the 2,7-isomer. An X-ray crystallographic analysis revealed a π-stacked column with a short intermolecular distance in the crystals of the 2,6-isomer. The absorption maximum of the 2,6-isomer was red-shifted as compared to that of the 2,7-isomer, indicating a π-conjugation between di-2-naphthyloxadiazole moieties in the 2,6-isomer. The quantum yields of the 2,6- and 2,7-isomers were measured to be 0.97 and 0.74, relative to that of 2,5-di-2-naphthyl-1,3,4-oxadiazole (0.85). Molecular orbital (MO) calculations demonstrated that the 2,6-isomer had a higher electron affinity than the 2,7-isomer. Thus, the crosslinking of building blocks is important for the design of functional materials.  相似文献   

17.
The one-step synthesis of 2,6-naphthalenedicarboxylic acid from naphthalene with carbon tetrachloride, copper powder and aqueous alkali has been achieved under mild conditions by the use of β-cyclodextrin as catalyst, producing 2,6-naphthalenedicarboxylic acid in 65 mol-% yield with 79% selectivity.  相似文献   

18.
2,6-Dichloronaphthalene dianhydride has been synthesized by a modified procedure. The imidization of this dichlorinated anhydride with amines and subsequent stepwise nucleophilic exchange of the chlorine atoms by alkyl- or arylamines afforded a series of hitherto unknown monoamino- and diamino-substituted naphthalene diimides. An alternative route for the synthesis of diamino-substituted naphthalene diimides is also reported. Optical and electrochemical properties of the newly synthesized amino-functionalized naphthalene diimides were studied in detail. The absorption maxima (530-620 nm) of these dyes are appreciably bathochromically shifted compared to those of the corresponding core-unsubstituted compounds. At the naphthalene core alkylamino-substituted diimides exhibit fluorescence quantum yields up to 60%.  相似文献   

19.
采用化学液相沉积法,经异丁基三乙氧基硅烷修饰并用高温水蒸气处理得到了改性HY沸石.采用X射线衍射、低温N2吸附和脉冲式质量分析技术研究了改性样品骨架结构、比表面积、孔结构参数和吸附性质的变化,并考察了HY沸石及其改性后样品对萘与叔丁醇烷基化制备2,6-二叔丁基萘(2,6-DTBN)反应的催化性能.结果表明,改性后HY沸石的骨架结构基本不变,但比表面积增大,平均孔径缩小,孔口尺寸得到了一定调变.在改性后HY沸石催化剂上萘与叔丁醇烷基化反应活性下降,但催化剂择形性能明显提高,其2,6-DTBN/2,7-DTBN比可以达到6.62。  相似文献   

20.
The pure 24-membered cyclic dimer ( 2 ) of ethylene 2,6-naphthalate was isolated in 9% yield by reaction of bis(2′-hydroxyethyl) 2,6-naphthalate ( 5 ) with 2,6-naphthalenedicarbonyl chloride ( 4 ) in THF with pyridine as catalyst under high dilution conditions. The macrocycle was characterized by NMR, MS, elemental analysis, and X-ray crystallography. The crystal structure reveals a significant distortion of the naphthalene rings, forming an ellipsoidal cavity as a result of the ring strain, suggesting facile ring-opening polymerization.  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