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 250 毫秒
1.
The negative ion mass spectra of dicarboxylic acids show [M]?˙ and prominent [M – H]?ions. These ions can therefore be used to determine the molecular weight of dicarboxylic acids which do not give positive molecular ions. The [C2H3]? ion is a base peak in the spectra of maleic and fumaric acids. Isomeric phthalic acids are readily differentiated.  相似文献   

2.
The reaction paths of product formation in the partial oxidation of n-pentane on vanadium-phosphorus oxide (VPO) and VPO-Bi catalysts are considered. The condensed products of n-pentane oxidation were analyzed by chromatography-mass spectrometry, and the presence of C4 rather than C5 unsaturated hydrocarbons was detected. It was found that the concentration of phthalic anhydride in the products increased upon the addition of C4 olefins and butadiene to the n-pentane-air reaction mixture. With the use of a system with two in-series reactors, it was found that the addition of butadiene to a flow of n-butane oxidation products (maleic anhydride, CO, and CO2) resulted in the formation of phthalic anhydride. The oxidation of 1-butanol was studied, and butene and butadiene were found to be the primary products of reaction; at a higher temperature, maleic anhydride and then phthalic anhydride were formed. The experimental results supported the reaction scheme according to which the activation of n-pentane occurred with the elimination of a methyl group and the formation of C4 unsaturated hydrocarbons. The oxidation of these latter led to the formation of maleic anhydride. The Diels-Alder reaction between maleic anhydride and C4 unsaturated hydrocarbons is the main path of phthalic anhydride formation.  相似文献   

3.
Benzene and maleic anhydride react over solid acids, viz. CrO3/Magnesol and SO4 2-/ZrO2 catalysts to form phthalic anhydride and olefins, which in turn produce phthalate esters as end products. Based on the product distribution, a reaction pathway is proposed.  相似文献   

4.
Ortho Phthalamic acids under electron impact show a retrosynthetic reaction leading to both phthalic anhydride and amine complementary ions, the corresponding neutrals of which are the usual synthetic precursors of the original compounds. The single case of a primary amine derivative is examined, which shows the formation of [M? H2O] ions having the structure of the related N-substituted imide, by a process which parallels a well known thermal reaction. It also gives the species [C8H6NO2]+ (of the same nominal mass as phthalic anhydride), the structure of which is still under study. Ionic structures are supported by collision induced mass analyzed ion kinetic energy spectra.  相似文献   

5.
The analytical potential of negative ion chemical ionization (NICI) mass spectrometry utilizing dibromodifluoro-methane (CF2Br2) and iodomethane (CH3I)/methane (CH4) as reagent gases is examined. The NICI mass spectrum of CF2Br2 contains Br?, [HBr2]? and [CF2Br3]? anions. Weak acids (i.e. those acids with approximately ΔH°(acid) values between 1674 and 1464 kJ mol?1) react with Br? to produce minor yields of the hydrogen?bonded bromide attachment [MH + Br]? anion or are unreactive. Strong acids (i.e. those acids with approximately ΔH°(acid) > 1464 kJ mol?1) produce primarily [MH + Br]? anions with a minor yield of proton transfer [M ? H]? anion. The NICI spectrum of CH3I/CH4 is dominated by I?. Weak acids react with I? to yield minor amounts of [MH + 1]? or are unreactive. Strong acids produce only [MH + l]? anions. From a consideration of the gas-phase basicity of the halide anion and the binding energy of the hydrogen-bonded halide attachment adduct, thermochemical data are used as a potential guide to rationalize or predict the ions observed in NICI mass spectra.  相似文献   

6.
We report the first positive chemical ionization (PCI) fragmentation mechanisms of phthalates using triple‐quadrupole mass spectrometry and ab initio computational studies using density functional theories (DFT). Methane PCI spectra showed abundant [M + H]+, together with [M + C2H5]+ and [M + C3H5]+. Fragmentation of [M + H]+, [M + C2H5]+ and [M + C3H5]+ involved characteristic ions at m/z 149, 177 and 189, assigned as protonated phthalic anhydride and an adduct of phthalic anhydride with C2H5+ and C3H5+, respectively. Fragmentation of these ions provided more structural information from the PCI spectra. A multi‐pathway fragmentation was proposed for these ions leading to the protonated phthalic anhydride. DFT methods were used to calculate relative free energies and to determine structures of intermediate ions for these pathways. The first step of the fragmentation of [M + C2H5]+ and [M + C3H5]+ is the elimination of [R? H] from an ester group. The second ester group undergoes either a McLafferty rearrangement route or a neutral loss elimination of ROH. DFT calculations (B3LYP, B3PW91 and BPW91) using 6‐311G(d,p) basis sets showed that McLafferty rearrangement of dibutyl, di(‐n‐octyl) and di(2‐ethyl‐n‐hexyl) phthalates is an energetically more favorable pathway than loss of an alcohol moiety. Prominent ions in these pathways were confirmed with deuterium labeled phthalates. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

7.
CF3 radicals were generated by the photolysis of perfluoroacetic anhydride. In the presence of pentafluorobenzene, the CF3 radicals react according to the following mechanism: It was found that the addition reaction (3) becomes reversible above ca. 453 K. The addition rate parameters have been revised and they satisfactorily agree with those reported previously. At temperatures higher than 593 K, only true H-abstraction occurs. The rate constant kH for reaction (5) is given by: where θ = 2.303 RT kJmol?1 and kc is the rate constant for combination of CF3 radicals. The reactions of CF3 with benzene and pentafluorobenzene are compared.  相似文献   

8.
Low-energy reactive collisions between the negative molecular ion of a tetrachlorodibenzo-p-dioxin (TCDD) and oxygen inside the collision cell of a triple-stage quadrupole mass spectrometer produce a substitution ion [M ? Cl + O]?, a phenoxide ion [C6H4-nO2Cln], [M ? HCl], and Cl? by which 1,2,3,4-, 1,2,3,6/1,2,3,7- and 2,3,7,8-TCDD isomers can be distinguished either directly or on the basis of intensity ratios. The collision conditions have an important effect on the relative abundances. Energy- and pressure-resolved curves show that the ions formed by a collisionally activated reaction (CAR) process, i.e. [M ? Cl + O]? and [C6H4-n,O2Cln], are favoured by a high pressure of oxygen (3-6 mTorr) (1 Torr = 133.3 Pa) and a low collision energy (0.1-7 eV), whereas the ions formed by a collisionally activated dissociation (CAD) process, i.e. [M ? HCl] and Cl?, are favoured by high pressure and high energy. By choosing a relatively low collision energy (5 eV) and high pressure (4 mTorr), the CAR and CAD ions can be clearly detected.  相似文献   

9.
The novel tetrameric gadolinium(III) compound [Gd4(OH)4(CF3COO)8(H2O)4] · 2.5 H2O was synthesized and structurally characterized by X‐ray crystallography. The Gd3+ ions are bridged by hydroxide ions and carboxylate groups to tetramers with Gd3+‐Gd3+ distances between 384.2(2) and 388.1(2) pm. The compound crystallizes in the monoclinic space group C2/c (Z = 4). The magnetic behaviour of [Gd4(OH)4(CF3COO)8(H2O)4] · 2.5 H2O was investigated in the temperature range of 2 to 300 K. The magnetic data of this compound indicate antiferromagnetic interactions (Jex = ?0.0197 cm?1).  相似文献   

10.
The ions [CF3CO2]+ and [CH3CO2]+ give peaks of small abundance in conventional positive ion spectra. These ions can be produced by collision-induced charge stripping of the corresponding stable negative ions. Six and ten fragment ions respectively are observed in the spectra of [CF3CO2]+ and [CH3CO2]+.  相似文献   

11.
On the Lewis Acidity of Fluorinated Sulfonium Ions NMR investigations show, that sulfonium salts [(CF3)nSF3?n]+ AsF6? ( 1–3 , n = 0–2) add CH3CN under formation of ψ-pentacoordinated sulfuranonium ions [(CF3)nSF3?n · NCCH3]+ ( 1a – 3a ,) with the donor in an axial position. In solution NSF3 ( 4 ,) forms similar salts [(CF3)nSF3?n · NSF3]+ AsF6? ( 1b-3b ,) with weaker donor-acceptor interactions. With NSF2NMe2 ( 5 ,) the step of the primary addition products is passed very quickly, by fluoride-migration from 1 , and 2 , persulfuranonium ions [(CH3)2NSF3NSF2]+ ( 6 ,) and [(CH3)2NSF3NSFCF3]+ ( 7 ,), respectively, are formed, while from 3 , only decomposition products (Me2NSF2+, CF3SSCF3, CF4) were obtained.  相似文献   

12.
The products of the reaction between the electrophilic alkenylxenonium cation [1-Xe+–C6F9] and the halide anions I?, Br?, Cl? and F? depend on the hardness of the halide anion. With the soft halides I? and Br? Xe(II) is formally displaced by halogen as well in basic MeCN as in superacidic (AHF1), whereas with hard fluoride and chloride no reaction takes place in AHF. In MeCN F? initiates the formation of alkenyl radicals, which abstract hydrogen from the solvent, whereas Cl? exhibits borderline character: RH and RCl formation. Possible reaction paths are discussed. The reactivity of the arylxenonium cation [C6F5Xe]+ in AHF toward halide ions is reported and the relative electrophilicity of the cations [C6F5Xe]+ and [1-Xe+–C6F9] is determined by the competitive reaction with Cl?. In addition the synthesis of cyclohexene 1-CF3–C6F9 from C6F5CF3 and XeF2 is performed and its electrophilicity is compared with that of the aromatic compound C6F5CF3.  相似文献   

13.
Reactions of CF3Br with H atoms and OH radicals have been studied at room temperature at 1–2 torr pressures in a discharge flow reactor coupled to an EPR spectrometer. The rate constant of the reaction H + CF3Br → CF3 + HBr (1) was found to be k1 = (3.27 ± 0.34) × 10?14 cm3/molec·sec. For the reaction of OH with CF3Br (8) an upper limit of 1 × 10?15 cm3/molec·sec was determined for k8. When H atoms were in excess compared to NO2, used to produce OH radicals, a noticeable reactivity of OH was observed as a result of the reaction OH + HBr → H2O + Br, HBr being produced from reaction (1).  相似文献   

14.
Under the influence of azobisisobutyronitrile, benzoselenophene and maleic anhydride slowly polymerize to form an alternating copolymer. The two comonomers form a charge-transfer complex (KADc at 30° in chloroform: 0·22 l mol?1). These properties of benzoselenophene are similar to those of benzothiophene and indole but are different from those of benzofuran. It is possible that the charge-transfer complex plays a role in the copolymerization.  相似文献   

15.
Reactions in the gas phase of the 13- and 15-electron radical anions [Cr(CO)3]? ˙ and [Cr(CO)4]? ˙ with a series of 27 aldehydes, ketones, esters and ethers have been examined. Sequential alkane eliminations and metal-bonded CO ligand displacements were the principal reactions identified for the RCHO/[Cr(CO)3]? ˙ systems with the latter reaction also common to the RCHO/[Cr(CO)4]? ˙ systems. While [Cr(CO)4]? ˙ was generally unreactive towards ketones R · R'CO, the principal products identified for [Cr(CO)3]? ˙/ketone reactions were the metal-decarbonylated species, respectively [R · R'CO · Cr(CO)x]? ˙ with x = 0–3, and [R · (R' - H2)CO · Cr(CO)2]? ˙. The reaction of [Cr(CO)3]? ˙ with esters RCOOR' proceeds via metal insertion into the alkoxy C? O bond to give end products of the type [R'O · Cr · R(CO)2]? and [R'O? Cr(CO)3]? while the sole ionic products of dialkyl ether/[Cr(CO)3]? ˙ reactions were identified as the alkoxytricarbonylchromium species [RO · Cr(CO)3]?.  相似文献   

16.
The oxidation of o‐xylene to phthalic anhydride over Co‐Mn/H3PW12O40@TiO2 was investigated. The experimental results demonstrated that the prepared catalyst effectively catalyzed the oxidation of o‐xylene to phthalic anhydride. Also, the synergistic effect between three metals plays vital roles in this reaction. From a green chemistry point of view, this method is environmentally friendly due to carrying out the oxidation in a fixed‐bed reactor under solvent‐free condition and using molecular oxygen as a green and cheap oxidizing agent. The resulting solid catalysts were characterized by FT‐IR, XRD, XPS, ICP‐OES, FESEM, TEM, EDX, DR‐UV spectroscopy, BET and thermogravimetric analysis. The oxidation of o‐xylene yields four products: o‐tolualdehyde, phthaldialdehyde, phthalide and finally phthalic anhydride as the main product. The reaction conditions for oxidation of o‐xylene were optimized by varying the temperature, weight hourly space velocity and oxygen flow rate (contact time). The optimum weight percentage of phosphotungstic acid (HPW) and Co/Mn for phthalic anhydride production were 15 wt % and 2 wt%, respectively. The best Co/Mn ratio was found to be 10/1. Oxygen flow rate was very important on the phthalic anhydride formation. The optimum conditions for oxidation of o‐xylene were T = 370 °C, WHSV = 0.5 h?1 and oxygen flow rate = 10 mL min?1. Under optimized conditions, a maximum of 88.2% conversion and 75.5% selectivity to phthalic anhydride was achieved with the fresh catalyst. Moreover, reusability of the catalyst was studied and catalytic activity remained unchanged after at least five cycles.  相似文献   

17.
Two europium trifluoroacetate complexes, Eu(CF3COO)3·phen ( 1 ) and Eu(CF3COO)3·bpy ( 2 ) (where phen=1,10‐phenanthroline, bpy=2,2′‐bipyridine), were synthesized and characterized by elemental analysis, Fourier transform infrared spectroscopy (FT‐IR), photoluminescence (PL) spectroscopy and thermogravimetric analysis (TA). Single‐crystal X‐ray structure has been determined for the complex [Eu2(CF3COO)6·(phen)3·(H2O)2]·EtOH. The crystal structure of [Eu2(CF3COO)6·(phen)3·(H2O)2]·EtOH shows that two different coordination styles with europium ions coexist in the same crystal and have entirely different coordination geometries and numbers. This crystal can be considered as an 1:1 adduct of [Eu(CF3COO)3·(Phen)2·H2O]·EtOH (9‐coordination part) and Eu(CF3COO)3·phen·H2O (8‐coordination part). The excitation spectra of the two complexes demonstrate that the energy collected by "antenna ligands" is transferred to Eu3+ ions efficiently. The room‐temperature PL spectra of the complexes are composed of the typical Eu3+ ions red emission, due to transitions between 5D07FJ(J=0→4). The lifetimes of 5D0 of Eu3+ in the complexes were examined using time‐resolved spectroscopic analysis, and the lifetime values of Eu(CF3COO)3·phen and Eu(CF3COO)3·bpy were fitting with bi‐exponential (2987 and 353 µs) and monoexponential (3191 µs) curves, respectively. In order to elucidate the energy transfer process of the europium complexes, the energy levels of the relevant electronic states had been estimated. The thermal analyses indicate that they are all quite stable to heat.  相似文献   

18.
In this work, poly(glycidylmethacrylate‐divinylbenzene) microspheres were prepared and applied for the preparation of weak acid cation exchange stationary phases. Succinic anhydride, phthalic anhydride, and maleic anhydride were selected as carboxylation reagents to prepare three weak acid cation exchangers by direct chemical derivatization reaction without solvent or catalyst. The diameters and dispersity of the microspheres were characterized by scanning electron microscopy; the amount of accessible epoxy groups and mechanical stability were also measured. The weak acid cation exchangers were characterized by Fourier transform infrared spectroscopy; the content of carboxyl groups was measured by traditional acid base titration method. The chromatographic properties were characterized and compared by separating alkali, alkaline earth metal ions and ammonium and polar amines. The separation properties enhanced in the order of succinic anhydride, phthalic anhydride, and maleic anhydride modified poly(glycidylmethacrylate‐divinylbenzene) cation exchangers.  相似文献   

19.
The methylamino diazonium cations [CH3N(H)N2]+ and [CF3N(H)N2]+ were prepared as their low‐temperature stable [AsF6]? salts by protonation of azidomethane and azidotrifluoromethane in superacidic systems. They were characterized by NMR and Raman spectroscopy. Unequivocal proof of the protonation site was obtained by the crystal structures of both salts, confirming the formation of alkylamino diazonium ions. The Lewis adducts CH3N3?AsF5 and CF3N3?AsF5 were also prepared and characterized by low‐temperature NMR and Raman spectroscopy, and also by X‐ray structure determination for CH3N3?AsF5. Electronic structure calculations were performed to provide additional insights. Attempted electrophilic amination of aromatics such as benzene and toluene with methyl‐ and trifluoromethylamino diazonium ions were unsuccessful.  相似文献   

20.
The trans isomer of the organogold(III) difluoride complex [PPh4][(CF3)2AuF2] has been obtained in a stereoselective way and in excellent yield by reaction of [PPh4][CF3AuCF3] with XeF2 under mild conditions. The compound is both thermally stable and reactive. Thus, the fluoride ligands are stereospecifically replaced by any heavier halide or by cyanide, the cyanide affording [PPh4][trans‐(CF3)2Au(CN)2]. The organogold fluoride complexes [CF3AuFx]? (x=1, 2, 3) have been experimentally detected to arise upon collision‐induced dissociation of the [trans‐(CF3)2AuF2]? anion in the gas phase. Their structures have been calculated by DFT methods. In the isomeric forms identified for the open‐shell species [CF3AuF2]?, the spin density residing on the metal center is found to strongly depend on the precise stereochemistry. Based on crystallographic evidence, it is concluded that Auiii and Agiii have similar covalent radii, at least in their most common square‐planar geometry.  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