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.
The pyrolysis kinetics of ethyl 3-hydroxy-3-methylbutanoate have been examined over the temperature range of 286–330°C and pressure range of 29–108 Torr. In a seasoned vessel and in the presence of the free radical inhibitor cyclohexene or toluene the reaction is homogeneous, unimolecular and obeys a first-order rate law. The elimination products are mainly acetone and ethyl acetate, and very small amounts of ethyl 3-butenoate, acetic acid, ethylene and H2O. The rate coefficient is expressed by the following equation: log k1(s–1)=(12.39±0.46)–(174.5±5.2) kJ mol–1 (2.303RT)–1. The mechanism appears to proceed via a six-membered cyclic transition state, where polarization of the (CH3)C(OH)+...-CH2COOCH2CH3 bond is rate determining.  相似文献   

2.
Reaction rate coefficients have been measured at 295 K for both CF3 and CF2 with atomic and molecular fluorine. The reaction between CF3 and F was studied over a gas number density range of (2.4–23)×1016 cm–3 with helium as the bath gas. The measured rate coefficient increased from (1.1–1.7)×10–11 cm3 s–1 as the gas number density increased over this range. In contrast to this relatively small change in rate coefficient with gas number density, the rate coefficient for CF2+F increased from (0.4–2.3)×10–12 cm3 s–1 as the helium gas number density increased from (3.4–28.4)×1016 cm–3. Even for the highest bath gas number density employed, the rate coefficient was still more than an order of magnitude lower than earlier measurements of this coefficient performed at comparable gas number densities.Both these association reactions are examined from the standpoint of the Gorin model for association of radicals and use is made of unimolecular dissociation theory to examine the expected dependence on gas number density. The calculations reveal that CF3+F can be explained satisfactorily in these terms but CF2+F is not well described by the simple Gorin model for association.CF3 was found to react with molecular fluorine with a rate coefficient of (7±2)×10–14 cm3 s–1 whereas only an upper limit of 2×10–15 cm3 s–1 could be placed on the rate coefficient for the reaction between CF2 and F2. The values obtained for this set of reactions mean that the reaction between CF3 and F will play an important role in plasmas containing CF4. The high rate coefficient will mean that, under certain conditions, this particular reaction will control the amount of CF4 consumed. On the other hand, the much lower rate coefficient for reactions between CF2 and F means that CF2 will attain much higher concentrations than CF3 in plasmas where these combination reactions are dominant.  相似文献   

3.
The present study shows the electrochemistry, electrogenerated chemiluminescence and solid-state electroluminescence of [Ru(bpy)2(4,4-(CH3(CH2)12COO)bpy)](ClO4)2 as thin film deposited onto ITO. Cyclic voltammetry in aqueous solution of the Ru-LC film shows two reversible waves with peak potentials at about 1.27 V and –0.97 V. The apparent diffusion coefficient of the slowest species was found to be 4–6×10–10 cm2/s, with a concentration of active sites of about 0.4–0.5 M.Relatively strong ECL was observed during both oxidation and reduction, in both potential scans and potential step experiments. The reduced species appears to be much less stable than the oxidized one, as already known for other tris(bipyridine) Ru(II) derivatives in aqueous media.Two-electrode solid-state devices prepared by Ga-In printing on the tops of Ru-LC films are similar to the Ru(bpy)3 ones. The currents that flow through these devices are however about three orders of magnitude smaller than those for Ru(bpy)3 devices as a result of their low electron and hole mobilities.Dedicated to Zbigniew Galus on the occasion of his 70th birthday.  相似文献   

4.
Poly[(R)-(–)-3-(l-pyrrolyl)propyl-N-(3,5-dinitrobenzoyl)-α-phenylglycinate] films were deposited on ITO electrodes using potentiodynamic and galvanostatic methods. Polymerization occurred as a charge dependent process at 1.0 V vs. Ag/Ag+(CH3CN) and was not affected by the presence of nitro groups in the monomer. The surface morphology of the film and its electrochemical properties were studied as a function of deposition charge (Qdep) and deposition method. Film thickness increased in a quasi-linear manner with respect to Qdep within the range 40–80 mC cm2. The galvanostatic method provided easier control of Qdep compared with potentiodynamic deposition, and produced a more adherent film with homogeneous grain geometry. Cyclic voltammetry revealed a well defined redox couple at the anodic region, attributable to polymer p-doping, and a poorly defined redox pair at the cathodic region, attributable to the reduction of the nitro group.  相似文献   

5.
The interactions of the sulfonium ions (CH3)3S+, (CH3)2S+CH2CO2 , and (CH3)2S+-CH2CH2CO2 with up to four water molecules have been studied by ab initio molecular orbital methods. Complexes of (CH3)3S+ with one to three water molecules involve strong electrostatic sulfur-oxygen interactions; in contrast, the sulfide (CH3)2S interacts with water molecules via weak S-H hydrogen bonds, suggesting that methyl-group transfer from (CH3)3S+ in aqueous solution involves a significant alteration of the hydration pattern around the sulfur atom. Two conformers of (CH3)2S+CH2CO2 were found that display sulfur-oxygen distances which are approximately 0.3 å less than the sum of the sulfur and oxygen van der Waals radii, indicating a strong intramolecular electrostatic interaction. For the complexes (CH3)2S+CH2CO2 ·nH2O(n =1–4), water interacts primarily with the carboxylate group via hydrogen bonds, rather than electrostatically with the sulfur atom, although in complexes with the three- and four-water complexes, the proximity of the positively charged sulfur atom to the carboxylate group significantly alters the hydration pattern compared to that in the corresponding of complexes CH3SCH2CO2 · Thus, methyl transfer from (CH3)2S+CH2CO2 to an acceptor in aqueous solution also involves substantial changes in the hydration pattern around the carboxylate group.  相似文献   

6.
The purpose of this report is to quantitatively find the cause for the elongation of the R-C bond in R-COO (R = H, CH3 and C2H5) and the shortening of the C-O bond in CH3-O upon deprotonation in the gas phase. These elongations and shortenings result from the contributions of R---CO2 and H---CH2=O as resonance structures to the systems. Because these structures must make only a small contribution in the crystal, the R-C bond lengths of R-COO (R= H and CH3) in the crystal structure are shorter than those in the gas phase.  相似文献   

7.
Unlike micelles of straight hydrocarbon chain-surfactants, isoprenoid surfactants, CH3 [CH(CH3)CH2CH2CH2]3 CH(CH3)CH2–R (R=CH2N+ (CH3)3 Br, CH2OPO3H Na+, CH2OSO 3 Na+, CO 2 Na+), gave large globular and cellular assemblies in water which could be observed directly by transmission electron microscopy; critical micelle concentration of 0.31.4×10–3 M at 20°C, aggregation number of 215×104, and diameter of 200–2000 Å. A basic structure of the assemblies was a thin layer with a thickness (about 30 Å) which was close to the molecular length of the surfactants. The assemblies were decomposed during gel column chromatography; viz., they were not as stable as the liposomes of lecithins. The morphology was discussed in conjunction with a steric effect of the isoprenoid chain.  相似文献   

8.
The electrochemical behaviour of some Ni γ-monosubstituted dibenzotetraaza [14] annulene complexes has been investigated. The oxidation in CH2Cl2 of the complex containing a 4-carboxybenzyl group leads to the corresponding γ-γ dimer whose electrochemical properties have been studied. The electrode surface can be coated by thin films of this dimer using CH3CN instead of CH2Cl2; however, the resulting modified electrode is poorly stable. The oxidation of the complex containing 1-(4-carboxybenzyl)pyrrole as γ substituent involves γ-γ dimer formation before the formation of a regular polypyrrole film. The film displays reversible electrochemical reduction of the metal centre (Ni(II)/Ni(I)) and two successive oxidations of the macrocycle (Mc/Mc•+ and Mc•+/Mc2+). The complex containing a bromo(4-carboxybenzyl) group offers an unusual feature in that polymeric films can be obtained following an original procedure based on alternated dimerizations. This is a consequence of the formation of two different dimers obtained by anodic and cathodic processes.  相似文献   

9.
Patterning of TiO2 thin films was successfully obtained on different self-assembled monolayers (SAMs) in aqueous solution by micro-contact printing (μCP) method. The substrates were immersed in an aqueous solution containing titanium sulfate (Ti(SO4)2) and hydrogen peroxide for deposition at 80 °C. The growth rates on various surfaces were as follows: sulfonic (–SO3H) > amino (–NH2) > methyl (–CH3) > hydroxyl (–OH). According to the XPS results, SAMs with the terminal groups of –SO3H and –NH2 were favorable for the deposition. The TiO2 film deposited on the SAM with the terminal group of –CH3 could be easily peeled off. Clearly, TiO2 patterns were obtained on the prepatterned surfaces of –SO3H/–CH3 and –NH2/–CH3 SAMs. The deposition mechanism might be relevant to electrostatic interaction, Stern layer, lone pair electrons and Van der Vaals forces. The TiO2 film was anatase after annealing at 500 °C and comprised particles with an average diameter of ca. 10 nm.  相似文献   

10.
The solid-liquid phase diagrams of binary mixtures of tetrabutylammonium halogenated carboxylates with water were examined in order to confirm the formation of clathrate-like hydrates. It was found that, among thirteen carboxylates examined, four carboxylates having CH2FCOO, CHF2COO, CF3COO, and CH2ClCOO, formed a hydrate with hydration numbers around 30 and seven carboxylates having CHCl2COO, CCl3COO, CH2BrCOO, CHBr2COO, CBr3COO, CH3CHClCOO, and CH3CHBrCOO formed a hydrate with hydration numbers around 23. The latter hydrate has not been reported earlier. The melting points of these newly found hydrates were fairly high: they lie between 10 and 16°C. The effect of Cl and Br atoms attached to the carbon atom of the-position of a carboxylate anion both on the type of hydrate formed and on its stability was greatly different from that of a CH3 group attached to the same position of the carboxylate anion.Dedicated to Dr D. W. Davidson in honor of his great contributions to the sciences of inclusion phenomena.  相似文献   

11.
Raman OD stretching spectra of alcoholic LiX (X = Cl, Br, I, ClO4 , NO3 , and CH3COO) solutions (alcohols = methanol and ethanol) were measured in the liquid state at room temperature and in the glassy state at liquid nitrogen temperature. The effects of the anions on the Raman OD stretching spectra in these alcoholic solutions are investigated and the structural changes of the solutions are discussed. It is shown that the structure-breaking effects of anions on the intrinsic alcoholic structure increase in the order: Cl < Br < I < ClO4 . From spectral changes, it seems that CH3COOLi exerts little effect on the liquid structure of the alcohol.  相似文献   

12.
The concentration dependence of the205Tl chemical shifts of Tl+ and of (CH3)2Tl+ ions was determined in several solvents with NO 3 and ClO 4 counterions. In general, increased ion pairing caused a low-frequency shift of the205Tl resonance, with the exceptions of (CH3)2TlNO3 inn-butylamine and TlNO3 in N,N-dimethylformamide (DMF) and in hexamethylphosphorotriamide (HMPA). In HMPA,205Tl linewidths of both Tl+ and (CH3)2Tl+ increased dramatically with dilution below 0.1M. Analysis of the data allowed ion-pair formation constants and205Tl chemical shifts for the ion-paired cation and for the free (solvated) cation to be estimated for some of the solvents.  相似文献   

13.
Fourier transform infrared (FT-IR) and Raman spectroscopy were employed to study the hydrothermal stability and the influence of surface functional groups on the surface wettability of methyl-modified silica films. The surface free energy parameters of the silica films were determined using the Lifshitz-van der Waals/acid–base approach. The thermal decomposition mechanisms of the CH3 groups in the methyl-modified silica material are proposed. The results show that with the increase of methyltriethoxysilane (MTES)/tetraethylorthosilicate (TEOS) ratio, the surface free energy and surface wettability of the silica films decrease greatly. This is mainly because of the contribution of the acid–base term; the intensity of Si–CH3 groups increases at the expense of the intensity of O–H groups in the samples. The surfaces of the methyl-modified silica films exhibited predominantly monopolar electron-donicity. The contact angle on the silica film surface reaches its maximum value when calcination is performed at 350 °C. Thermogravimetric analysis implies that some low molecular weight species, such as H2, CH4, and C, are eliminated upon thermal decomposition of the –CH3 groups. The Si–CH3 and –CH3 vibrational bands diminish in intensity as the calcination temperature is increased, disappearing completely when the calcination temperature is increased to 600 °C. When the calcination temperature is increased to 750 °C, the free carbon and CSi4 species will be formed.  相似文献   

14.
New kinds of organic derivatives of layer structured -zirconium phosphate Zr(HPO4)2·2H2O are prepared by the exchange of the interlayer phosphate groups with phosphoric ester groups having oxyethylene chains, CH3–(OCH2CH2)n–OPO3 2– (n=1–3). Half of the interlayer phosphate groups are exchanged topochemically, the oxyethylene chains being grafted onto the phosphate layers through the ester bonds in the resulting derivatives. The derivatives behave like crown ethers, and form complexes with alkali salts of soft base anions such as SCN and I. Alkali salts of hard base anions such as Br and NO3 do not form complexes with the derivatives. The alkali iodide complexes of the organic derivatives can be used for the halogen exchange reaction. n-Butyl bromide is converted into n-butyl iodide in the presence of the alkali iodide complexes. The reactivities for the halogen exchange reaction increase with the number of the oxyethylene units.  相似文献   

15.
29Si NMR peaks due to species with the double four-membered ring siloxane backbone composed of both Si(O)4/2 and CH3Si(O)3/2 units, (CH3) n Si8O 20 – n /(8 – n) – (n=1–3), formed by co-hydrolysis of tetraethoxysilane and methyltriethoxysilane in the presence of tetramethylammonium ions in methanol have been assigned. It has been found that 29Si NMR peaks due to Si(OSi)3(O) units shift to lower frequencies by replacement of the adjacent Si(O)4/2 units by CH3Si(O)3/2 units, in other words, with increasing m value in Si[OSi(O)3]3 – m [OSi(CH3) (O)2] m (O) (m=0–2). Peaks from CH3 Si(OSi)3 units in the species have also appeared as separated due to the kind of neighbor structural units. On the basis of the assignments, positions of CH3Si(O)3/2 units in the cubic octameric siloxane framework of (CH3) n Si8O 20 – n /(8 – n) – (n=2, 3), for both of which three isomers are present, have been estimated.  相似文献   

16.
Silica from leached chrysotile fibers (SILO) was silanized with trialkoxyaminosilanes to yield inorganic–organic hybrids designated SILx (x=1–3). The greatest amounts of the immobilized agents were quantified as 2.14, 1.90, and 2.18 mmol g−1 on SIL1, SIL2, and SIL3 for –(CH2)3NH2,–(CH2)3NH(CH2)2NH2, and –(CH2)3NH(CH2)2NH(CH2)2NH2 groups attached to the inorganic support. The infrared spectra for all modified silicas showed the absence of the Si–OH deformation mode, originally found at 950 cm−1, and the appearance of asymmetric and symmetric C–H stretching bands at 2950 and 2840 cm−1. Other important bands associated with the organic moieties were assigned to νas(NH) at 3478 and νsym(NH) at 3418 cm−1. The NMR spectrum of the solid precursor material suggested two different kinds of silicon atoms: silanol and siloxane groups, between −90 and 110 ppm; however, additional species of silicon that contain the organic moieties bonded to silicon at −58 and −66 ppm appeared after chemical modification. These modified silicas showed a high adsorption capacity for cobalt and copper cations in aqueous solution, in contrast to the original SILO matrix, confirming the unequivocal anchoring of silylating agents on the silica surface.  相似文献   

17.
Micro fabricated sensors based on nanomechanical motion with piezoresistive electrical readout have become a promising biochemical sensing tool. The conventional microcantilever materials are mostly silicon-based. The sensitivity of the sensor depends on Young's modulus of the structural material, thickness of the cantilever as well as on the gauge factor of the piezoresistor. UV patternable polymers such as SU-8 have a very low Young's modulus compared to the silicon-based materials. Polymer cantilevers with a piezoresistive material having a large gauge factor and a lower Young's modulus are therefore highly suited for sensing applications. In this work, a spin coatable and photopatternable mixture of carbon black (CB) and SU-8, with proper dispersion characteristics, has been demonstrated as a piezoresistive thin film for polymer microcantilevers. Results on percolation experiments of SU-8/CB composite and fabrication of piezoresistive SU-8 microcantilevers using this composite are presented. With our controlled dispersion experiments, we could get a uniform piezoresistive thin film of thickness less than 1.2 μm and resistivity of 2.7 Ω cm using 10 wt% of CB in SU-8. The overall thickness of the SU-8/composite/SU-8 is approximately 3 μm. We further present results on the electromechanical characterization and biofunctionalization of the cantilever structures for biochemical sensing applications. These cantilevers show a deflection sensitivity of 0.55 ppm/nm. Since the surface stress sensitivity is 4.1 × 10−3 [N/m]1, these cantilevers can well be used for detection of protein markers for pathological applications.  相似文献   

18.
The influence of the solvent on the activation energy of the synchronous, concerted, narcissistic reaction: CH3F + F¦CH3F2&#x00A6;FCH3+F, has been investigated by means of CNDO/2 method. The reaction path has been fully followed in vacuo, and at a suitable point only, of reaction coordinate, in solution. The geometries of H2O, F, CH3F, and ¦CH3F2¦, in vacuo and in different hydrated cages, have been optimized.
Zusammenfassung Der Einfluß des Lösungsmittels auf die Aktivierungsenergie der synchronverlaufenden Reaktionen CH3F+F¦CH3F2¦FCH3+ F, wurde mit der CNDO/2 Methode geprüft. Der Reaktions-Ablauf wurde in Vakuum vollkommen und in der Lösung an einem geeigneten Punkt der Reaktionskoordinate durchgeführt. Die Geometrien von H2O, F, CH3F, und ¦CH3F2¦ im Vakuum und in verschiedenen hydrierten Käfigen wurden optimalisiert.

Résumé L'influence du solvent sur l'énergie d'activation de la réaction synchronique, concertée, narcissistique: CH3F+ F¦CH3F2¦FCH3+F a été examinée par la méthode CNDO/2. On a suivie l'entier chemin de réaction in vacuo, alors qu'on a observé un seul point convenable de la coordonée de réaction en solution. On a optimize les geometries de H2O, F, CH3F, et ¦CH3F2¦ in vacuo et en differentes cages hydratés.


Financial aid from Italian Centro Nazionale delle Ricerche is gratefully acknowledged.  相似文献   

19.
The reaction of dimeric rhodium precursor [Rh(CO)2Cl]2 with two molar equivalent of 1,1,1-tris(diphenylphosphinomethyl)ethane trichalcogenide ligands, [CH3C(CH2P(X)Ph2)3](L), where X = O(a), S(b) and Se(c) affords the complexes of the type [Rh(CO)2Cl(L)] (1a–1c). The complexes 1a–1c have been characterized by elemental analyses, mass spectrometry, IR and NMR (1H, 31P and 13C) spectroscopy and the ligands a–c are structurally determined by single crystal X-ray diffraction. 1a–1c undergo oxidative addition (OA) reactions with different electrophiles such as CH3I, C2H5I and C6H5CH2Cl to give Rh(III) complexes of the types [Rh(CO)(COR)ClXL] {R = –CH3 (2a–2c), –C2H5 (3a–3c); X = I and R = –CH2C6H5 (4a–4c); X = Cl}. Kinetic data for the reaction of a–c with CH3I indicate a first-order reaction. The catalytic activity of 1a–1c for the carbonylation of methanol to acetic acid and its ester is evaluated and a higher turn over number (TON = 1564–1723) is obtained compared to that of the well-known commercial species [Rh(CO)2I2] (TON = 1000) under the reaction conditions: temperature 130 ± 2 °C, pressure 30 ± 2 bar and time 1 h.  相似文献   

20.
The Raman (3500–40 cm–1) and infrared (3500–70 cm–1) spectra of gaseous and solid 2-methoxypropene, CH3O(CH3)C=CH2, and the isotopomers, CD3O(CH3)C=CH2 and CH3O(CD3)C=CD2 have been recorded. In addition, the Raman spectra of the liquids have been recorded with qualitative depolarization measurements. All of these data indicate that only one conformer is present in the fluid phases at ambient temperature and this form is the cis conformer, which remains in the solid. Assignments are provided for the fundamentals of all three isotopomers for the cis conformer with Cs symmetry. The far-infrared spectra of all three isotopic species have been recorded at a resolution of 0.1 cm–1 in the gas and 1.0 cm–1 in the solid. The parameters of the potential function governing the asymmetric torsion are determined to be V3 = 1485 ± 9 cm–1 and V6 = –55 ± 4 cm–1 for the d0 compound, where only two terms were determined, since a second conformer was not evident. The barrier to internal rotation for the methyl group attached to the oxygen atom is 1370 ± 8 cm–1 and the C—CH3 barrier is 772 ± 5 cm–1. Ab initio calculations with full electron correlation have been carried out by the perturbation method to second order to obtain the equilibrium structural parameters, harmonic force constants, fundamental frequencies, infrared intensities, Raman activities, depolarization values, and conformational stability. The predicted values have been compared to the experimental values where appropriate.  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