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 46 毫秒
1.
The thermodynamics and kinetics of the surface hydrogenation of adsorbed atomic carbon to methane, following the reaction sequence C+4 H?CH+3 H?CH2+2 H?CH3+H?CH4, are studied on Fe(100) by means of density functional theory. An assessment is made on whether the adsorption energies and overall energy profile are affected when zero‐point energy (ZPE) corrections are included. The C, CH and CH2 species are most stable at the fourfold hollow site, while CH3 prefers the twofold bridge site. Atomic hydrogen is adsorbed at both the twofold bridge and fourfold hollow sites. Methane is physisorbed on the surface and shows neither orientation nor site preference. It is easily desorbed to the gas phase once formed. The incorporation of ZPE corrections has a very slight, if any, effect on the adsorption energies and does not alter the trends with regards to the most stable adsorption sites. The successive addition of hydrogen to atomic carbon is endothermic up to the addition of the third hydrogen atom resulting in the methyl species, but exothermic in the final hydrogenation step, which leads to methane. The overall methanation reaction is endothermic when starting from atomic carbon and hydrogen on the surface. Zero‐point energy corrections are rarely provided in the literature. Since they are derived from C? H bonds with characteristic vibrations on the order of 2500–3000 cm?1, the equivalent ZPE of 1/2 is on the order of 0.2–0.3 eV and its effect on adsorption energy can in principle be significant. Particularly in reactions between CHx and H, the ZPE correction is expected to be significant, as additional C? H bonds are formed. In this instance, the methanation reaction energy of +0.77 eV increased to +1.45 eV with the inclusion of ZPE corrections, that is, less favourable. Therefore, it is crucial to include ZPE corrections when reporting reactions involving hydrogen‐containing species.  相似文献   

2.
We previously reported the dinuclear material [FeII2(ddpp)2(NCS)4] ? 4 CH2Cl2 ( 1? 4 CH2Cl2; ddpp=2,5‐di(2′,2′′‐dipyridylamino)pyridine) and its partially desolvated analogue ( 1? CH2Cl2), which undergo two‐ and one‐step spin‐crossover (SCO) transitions, respectively. Here, we manipulate the type and degree of solvation in this system and find that either a one‐ or two‐step spin transition can be specifically targeted. The chloroform clathrate 1? 4 CHCl3 undergoes a relatively abrupt one‐step SCO, in which the two equivalent FeII sites within the dinuclear molecule crossover simultaneously. Partial desolvation of 1? 4 CHCl3 to form 1? 3 CHCl3 and 1? CHCl3 occurs through single‐crystal‐to‐single‐crystal processes (monoclinic C2/c to P21/n to P21/n) in which the two equivalent FeII sites become inequivalent sites within the dinuclear molecule of each phase. Both 1? 3 CHCl3 and 1? CHCl3 undergo one‐step spin transitions, with the former having a significantly higher SCO temperature than 1? 4 CHCl3 and the latter, and each has a broader SCO transition than 1? 4 CHCl3, attributable to the overlap of two SCO steps in each case. Further magnetic manipulation can be carried out on these materials through reversibly resolvating the partially desolvated material with chloroform to produce the original one‐step SCO, or with dichloromethane to produce a two‐step SCO reminiscent of that seen for 1? 4 CH2Cl2. Furthermore, we investigate the light‐induced excited spin state trapping (LIESST) effect on 1? 4 CH2Cl2 and 1? CH2Cl2 and observe partial LIESST activity for the former and no activity for the latter.  相似文献   

3.
The products of the gas‐phase reactions of the OH radical with n‐butyl methyl ether and 2‐isopropoxyethanol in the presence of NO have been investigated at 298 ± 2 K and 740 Torr total pressure of air by gas chromatography and in situ atmospheric pressure ionization tandem mass spectrometry. The products observed from n‐butyl methyl ether were methyl formate, propanal, butanal, methyl butyrate, and CH3C(O)CH2CH2OCH3 and/or CH3CH2C(O)CH2OCH3, with molar formation yields of 0.51 ± 0.11, 0.43 ± 0.06, 0.045 ± 0.010, ∼0.016, and 0.19 ± 0.04, respectively. Additional products of molecular weight 118, 149 and 165 were observed by API‐MS/MS analyses, with those of molecular weight 149 and 165 being identified as organic nitrates. The products observed and quantified from 2‐isopropoxyethanol were isopropyl formate and 2‐hydroxyethyl acetate, with molar formation yields of 0.57 ± 0.05 and 0.44 ± 0.05, respectively. For both compounds, the majority of the reaction products and reaction pathways are accounted for, and detailed reaction mechanisms are presented. The results of this product study are combined with previous literature product data to investigate the tropospheric reactions of R1R2C(Ȯ)OR radicals formed from ethers and glycol ethers, leading to a revised estimation method for the calculation of reaction rates of alkoxy radicals. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 501–513, 1999  相似文献   

4.
The previously unresolved issue of polymerization of allyl monomers CH2?CHCH2X is overcome by a palladium‐catalyzed insertion polymerization of diallyl ether as a monomer. An enhanced 2,1‐insertion of diallyl ether as compared to mono‐allyl ether retards the formation of an unreactive five‐membered cyclic O‐chelate (after 1,2‐insertion) that otherwise hinders further polymerization, and also enhances incorporation in ethylene polymers (20.4 mol %). Cyclic ether repeat units are formed selectively (96 %–99 %) by an intramolecular insertion of the second allyl moiety of the monomer. These features even enable a homopolymerization to yield polymers (poly‐diallyl ether) with degrees of polymerization of DPn≈44.  相似文献   

5.
The formation of reverse‐vesicular structures of the polyoxometalate‐containing hybrid surfactants [nBu4N]3[MnMo6O18{(OCH2)3? CNHCO(CH2)n?2CH3}2] (Mn‐Anderson‐Cn, n=6, 16) in nonpolar medium was achieved by titrating toluene into Mn‐Anderson‐Cn/acetonitrile (MeCN) solution. Stepwise change of the solvent polarity induces self‐association of the hydrophilic Mn‐Anderson cluster on the hybrid amphiphiles. The reverse‐vesicle formation was characterized by laser light scattering and further confirmed by transmission electron microscopy techniques, and the vesicle sizes increase with increasing toluene contents. The assembly process was accelerated at an elevated temperature. The length of the alkyl tails on the hybrid surfactants has a minor effect on the vesicle sizes, because the strong attraction between the polyoxometalate clusters is more dominant in the reverse‐vesicle formation.  相似文献   

6.
The alkaline hydrolysis reaction rates of 1,n‐bis(4‐cyanopyridinium)alkane derivatives Cnbis(CP)2+ with n = 3, 6, and 8 were studied and compared to the reaction rate of the N‐methyl‐4‐cyanopyridinium (MCP+). C6bis(CP)2+ and C8bis(CP)2+ obeyed the first‐order kinetic law. However for C3bis(CP)2+ data fitted to a consecutive two‐step model reaction, the observed rate constants (kobs) of C8bis(CP)2+ and C6bis(CP)2+ are approximately 50% and 100%, respectively, higher than those for MCP+, an effect mainly assigned to the higher charge density of these two derivatives. For C3bis(CP)2+, the kobs of the second (slow) step is almost twofold the value observed for C6bis(CP)2+, whereas the first (fast) step is approximately six times higher. As for MCP, the hydrolysis of Cnbis(CP)2+ generates pyridone (Po) and carbamidopyridinium (A+) units. For C3bis(CP)2+, however at pHs above 11.5, one additional product is formed. From the existence of the new product and the kinetic evidence, a “sandwiched‐type” complex with the OH? inserted between the rings is proposed. This structural effect in the C3bis(CP)2+ due to the conformational effect justifies the (i) two kinetic steps, (ii) high rate constants, (iii) high Po/A+ ratios, (iv) observed temperature and salt effects, and (v) the formation of the new product.  相似文献   

7.
Ti[O(CH2)4OCH?CH2]4, used for the ring‐opening polymerization (ROP) of ε‐caprolactone, was synthesized through the ester‐exchange reaction of titanium n‐propoxide and 1,4‐butanediol vinyl ether, and its chemical structure was confirmed by nuclear magnetic resonance (1H NMR) and thermogravimetric analysis (TGA). The mechanism and kinetics of Ti[O(CH2)4OCH?CH2]4‐initiated bulk polymerization of ε‐caprolactone were investigated. The results demonstrate that Ti[O (CH2)4OCH?CH2]4‐initiated polymerization of ε‐caprolactone proceeds through the coordination‐insertion mechanism, and all the four alkoxide arms in Ti[O (CH2)4OCH?CH2]4 share a similar activity in initiating ROP of ε‐caprolactone. The polymerization process can be well predicted by the obtained kinetic parameters, and the activation energy is 106 KJ/mol. Then, the rheological method was employed to investigate the feasibility of producing the crosslinked poly(ε‐caprolactone)‐poly (n‐butyl acrylate) network by using Ti[O(CH2)4OCH?CH2]4 as the ROP initiator. The tensile test demonstrates that the in situ generated crosslinked PCL‐PBA network in PMMA matrix provides the possibility of ameliorating the tensile properties of PMMA. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 7773–7784, 2008  相似文献   

8.
Heterobimetallic Complexes of Lithium, Aluminum, and Gold with the N ‐[2‐ N ′, N ′‐(dimethylaminoethyl)‐ N ‐methyl‐aminoethyl]‐ferrocenyl Ligand (η5‐C5H5)Fe{η5‐C5H3[CH(CH3)N(CH3)CH2CH2NMe2]‐2} N‐[2‐N′,N′‐(dimethylaminoethyl)‐N‐methyl‐aminoethyl]ferrocene FcN,NH ( 1 ) reacts with nBuLi under formation of the lithium organyl (FcN,N)Li ( 2 ). At reactions of 2 with AlBr3 and AuCl · PPh3 the heterobimetallic organo derivatives (FcN,N)AlBr2 ( 3 ), (FcN,N)Au · PPh3 ( 4 ) are formed. A detailed characterization of 2 – 4 was carried out by single crystal x‐ray analyses as well as by NMR and Mößbauer spectroscopy.  相似文献   

9.
The effects of phase modulation due to homonuclear proton–proton coupling constants in HSQMBC‐IPAP and HMBC‐IPAP experiments are experimentally evaluated. We show that accurate values of small proton–carbon coupling constants, nJCH, can be extracted even for phase‐distorted cross‐peaks obtained from a selHSQMBC experiment applied simultaneously on two mutually J‐coupled protons. On the other hand, an assessment of the reliability of nJCH measurement from distorted cross‐peaks obtained in broadband IPAP versions of equivalent HMBC and HSQMBC experiments is also presented. Finally, we show that HMBC‐COSY experiments could be an excellent complement to HMBC for the measurement of small nJCH values. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

10.
The novel coordination polymers [Cu(Hoxonic)(H2O)]n ( 1 ) and [Cu(Hoxonic)(bpy)0.5]n ? 1.5 n H2O ( 2?H2O ) (H3oxonic: 4,6‐dihydroxy‐1,3,5‐triazine‐2‐carboxylic acid; bpy: 4,4′‐bipyridine) have been isolated and structurally characterised by ab initio X‐ray powder diffraction. The dense phase 1 contains 1D zig‐zag chains in which Hoxonic dianions bridge square‐pyramidal copper(II) ions, apically coordinated by water molecules. On the contrary, 2?H2O , prepared by solution and solventless methods, is based on 2D layers of octahedral copper(II) ions bridged by Hoxonic ligands, further pillared by bpy spacers. The resulting pro‐porous 3D network possesses small hydrated cavities. The reactivity, thermal, magnetic and adsorptive properties of these materials have been investigated. Notably, the adsorption studies on 2 show that this material possesses unusual adsorption behaviour. Indeed, guest uptake is facilitated by increasing the thermal energy of both the guest and the framework. Thus, neither N2 at 77 K nor CO2 at 195 K are incorporated, and CH4 is only minimally adsorbed at 273 K and high pressures (0.5 mmol g?1 at 2500 kPa). By contrast, CO2 is readily incorporated at 273 K (up to 2.5 mmol g?1 at 2500 kPa). The selectivity of 2 towards CO2 over CH4 has been investigated by means of variable‐temperature zero coverage adsorption experiments and measurement of breakthrough curves of CO2/CH4 mixtures. The results show the highly selective incorporation of CO2 in 2 , which can be rationalised on the basis of the framework flexibility and polar nature of its voids.  相似文献   

11.
Polymerization of 2‐pentene with [ArN?C(An)C(An)·NAr)NiBr2 (Ar?2,6‐iPr2C6H3)] ( 1‐Ni) /M‐MAO catalyst was investigated. A reactivity between trans‐2‐pentene and cis‐2‐pentene on the polymerization was quite different, and trans‐2‐pentene polymerized with 1‐Ni /M‐MAO catalyst to give a high molecular weight polymer. On the other hand, the polymerization of cis‐2‐butene with 1‐Ni /M‐MAO catalyst did not give any polymeric products. In the polymerization of mixture of trans‐ and cis‐2‐pentene with 1‐Ni /M‐MAO catalyst, the Mn of the polymer increased with an increase of the polymer yields. However, the relationship between polymer yield and the Mn of the polymer did not give a strict straight line, and the Mw/Mn also increased with increasing polymer yield. This suggests that side reactions were induced during the polymerization. The structures of the polymer obtained from the polymerization of 2‐ pentene with 1‐Ni /M‐MAO catalyst consists of ? CH2? CH2? CH(CH2CH3)? , ? CH2? CH2? CH2? CH(CH3)? , ? CH2? CH(CH2CH2CH3)? , and methylene sequence ? (CH2)n? (n ≥ 5) units, which is related to the chain walking mechanism. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2858–2863, 2008  相似文献   

12.
Herein, we report the syntheses of silicon‐ and tin‐containing open‐chain and eight‐membered‐ring compounds Me2Si(CH2SnMe2X)2 ( 2 , X=Me; 3 , X=Cl; 4 , X=F), CH2(SnMe2CH2I)2 ( 7 ), CH2(SnMe2CH2Cl)2 ( 8 ), cyclo‐Me2Sn(CH2SnMe2CH2)2SiMe2 ( 6 ), cyclo‐(Me2SnCH2)4 ( 9 ), cyclo‐Me(2?n)XnSn(CH2SiMe2CH2)2SnXnMe(2?n) ( 5 , n=0; 10 , n = 1, X= Cl; 11 , n=1, X= F; 12 , n=2, X= Cl), and the chloride and fluoride complexes NEt4[cyclo‐ Me(Cl)Sn(CH2SiMe2CH2)2Sn(Cl)Me?F] ( 13 ), PPh4[cyclo‐Me(Cl)Sn(CH2SiMe2CH2)2Sn(Cl)Me?Cl] ( 14 ), NEt4[cyclo‐Me(F)Sn(CH2SiMe2CH2)2Sn(F)Me?F] ( 15 ), [NEt4]2[cyclo‐Cl2Sn(CH2SiMe2CH2)2SnCl2?2 Cl] ( 16 ), M[Me2Si(CH2Sn(Cl)Me2)2?Cl] ( 17 a , M=PPh4; 17 b , M=NEt4), NEt4[Me2Si(CH2Sn(Cl)Me2)2?F] ( 18 ), NEt4[Me2Si(CH2Sn(F)Me2)2?F] ( 19 ), and PPh4[Me2Si(CH2Sn(Cl)Me2)2?Br] ( 20 ). The compounds were characterised by electrospray mass‐spectrometric, IR and 1H, 13C, 19F, 29Si, and 119Sn NMR spectroscopic analysis, and, except for 15 and 18 , single‐crystal X‐ray diffraction studies.  相似文献   

13.
We show that water‐soluble monosodic salts of F‐alkyl phosphates CnF2n+1(CH2)2OP(O)(OH)2, with n=8 and 10 (F8H2Phos and F10H2Phos) form Gibbs films with exceptionally high dilational viscoelastic modules E that reach ~900 mN m?1 in the condensed phases. These E values are up to one order of magnitude larger than those recorded for phospholipid, protein and polymer films commonly considered as highly viscoelastic. F8H2Phos.1Na undergoes a transition between a liquid‐expanded and a liquid‐condensed phase. In the case of F10H2Phos.1Na, a transition occurs between a gas phase of surface domains, in which the molecules are densely packed, and a liquid‐condensed phase.  相似文献   

14.
A new modification of the Friedel–Crafts type intramolecular cyclization involving O‐protected ortho‐acetal diarylmethanols as a new type of reactant, was carried out for the first time in a medium containing a large amount of water at room temperature and enabled synthesis of a series of electron‐rich, hexahydroxylated 10‐O‐R‐substituted anthracenes, where R is an alkyl (Me, nBu, n‐C16H33) or arylalkyl group (CH2Ph, CH2‐2‐Napht, CH2C6H4CH2OAr) and also evaluation of their electronic and optoelectronic properties in solution, crystal, and solid thin film. In this transformation, a central 10‐O‐R‐substituted benzene ring was formed, fused to rings originating from two independent aromatic aldehydes. The reaction proceeded via two identified mechanisms involving acetal and/or free aldehyde groups. The acid sensitive acetal and dibenzyl alkoxy functions have never been used together in the intramolecular Friedel–Crafts type cyclization. The new compounds revealed deep blue fluorescence and quantum yields in solution around 0.3. The electrical properties investigated for thin films obtained by vacuum deposition on glass were 10‐O‐R‐substituent dependent and showed much faster transient current decay in the case of the 10‐O‐CH2Ph derivative than for the material with a 10‐O‐Me substituent (the lifetime of charge carriers was 25 times shorter in this case). The AFM images of thin films, Stokes shifts, and X‐ray analysis of π‐stacking interactions in crystals of the new materials have been also obtained.  相似文献   

15.
Purification of rare earth elements is challenging due to their chemical similarities. All of the deployed separation methods rely on thermodynamic properties, such as distribution equilibria in solvent extraction. Rare‐earth‐metal separations based on kinetic differences have not been examined. Herein, we demonstrate a new approach for rare‐earth‐element separations by exploiting differences in the oxidation rates within a series of rare earth compounds containing the redox‐active ligand [{2‐(t BuN(O))C6H4CH2}3N]3−. Using this method, a single‐step separation factor up to 261 was obtained for the separation of a 50:50 yttrium–lutetium mixture.  相似文献   

16.
Hydration of acetylene under steady‐state conditions around 450 K proceeds on Cd‐clinoptilolite without catalyst deactivation and formation of by‐products. Reaction rates were determined under steady‐state conditions at different partial pressures of acetylene, water, and acetaldehyde. In relation with the results, rate equations for different kinetic models were evaluated. Langmuir‐Hinshelwood kinetics was established. According to this model, acetylene and water must adsorb on similar sites, and the surface reaction between the adsorbed reactants is the rate‐determining step, which is followed by equilibrated desorption of the produced acetaldehyde.  相似文献   

17.
A series of RuIV–alkylidenes based on unsymmetrical imidazolin‐2‐ylidenes, that is, [RuCl2{1‐(2,4,6‐trimethylphenyl)‐3‐R‐4,5‐dihydro‐(3H)‐imidazol‐1‐ylidene}(CHPh)(pyridin)] (R=CH2Ph ( 5 ), Ph ( 6 ), ethyl ( 7 ), methyl ( 8 )), have been synthesized. These and the parent initiators [RuCl2(PCy3){1‐(2,4,6‐trimethylphenyl)‐3‐R‐4,5‐dihydro‐(3H)‐imidazol‐1‐ylidene}(CHC6H5)] (R=CH2C6H5 ( 1 ), C6H5 ( 2 ), ethyl ( 3 )) were used for the alternating copolymerization of norborn‐2‐ene (NBE) with cis‐cyclooctene (COE) and cyclopentene (CPE), respectively. Alternating copolymers, that is, poly(NBE‐alt‐COE)n and poly(NBE‐alt‐CPE)n containing up to 97 and 91 % alternating diads, respectively, were obtained. The copolymerization parameters of the alternating copolymerization of NBE with CPE under the action of initiators 1 – 3 and 5 – 8 were determined by using both a zero‐ and first‐order Markov model. Finally, kinetic investigations using initiators 1 – 3 , 6 , and 7 were carried out. These revealed that in contrast to the 2nd‐generation Grubbs‐type initiators 1 – 3 the corresponding pyridine derivatives 6 and 7 represent fast and quantitative initiating systems. Hydrogenation of poly(NBE‐alt‐COE)n yielded a fully saturated, hydrocarbon‐based polymer. Its backbone can formally be derived by 1‐olefin polymerization of CPE (1,3‐insertion) followed by five ethylene units and thus serves as an excellent model compound for 1‐olefin polymerization‐derived copolymers.  相似文献   

18.
(Acetoxymethyl)silanes 2 , 7 a – c , and 10 a – c with at least one alkoxy group, of the general formula (AcOCH2)Si(OR)3?n(CH3)n (R: Me, Et, iPr; n=0, 1, 2), were synthesized from the corresponding (chloromethyl)silanes 1 , 6 a – c , and 9 a – c by treatment with potassium acetate under phase‐transfer‐catalysis conditions. These compounds were found to provide 2,2,5,5‐organo‐substituted 1,4‐dioxa‐2,5‐disilacyclohexanes 3 , 8 a – c , and 11 a – c if treated with organotin(IV) catalysts such as dioctyltin oxide. The reaction proceeds through transesterification of the acetoxy and alkoxy units followed by ring‐closure to form a dimeric six‐membered ring. The corresponding alkyl acetates are formed as the reaction by‐products. With these mild conditions, the method overcomes the drawbacks of previously reported synthetic routes to furnish 2,2,5,5‐tetramethyl‐1,4‐dioxa‐2,5‐disilacyclohexane ( 3 ) and even allows the synthesis of 1,4‐dioxa‐2,5‐disilacyclohexanes bearing hydrolytically labile alkoxy substituents at the silicon atom in good yields and high purity. These new materials were fully characterized by NMR spectroscopy, elemental analysis, mass spectrometry, and X‐ray analysis (trans‐ 8 a ).  相似文献   

19.
15N NMR data of a series of 3‐alkyl[aryl] substituted 5‐trichloromethyl‐1,2‐dimethyl‐1H‐pyrazolium chlorides (where the 3‐substituents are H, Me, Et, n‐Pr, n‐Bu, n‐Pe, n‐Hex, (CH2)5CO2Et, CH2Br, Ph and 4‐Br‐C6H4), are reported. The 15N substituent chemical shifts (SCS) parameters are determined and these data are compared with the 13C SCS values and data obtained by MO calculations. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

20.
Free‐standing macroporous air electrodes with enhanced interfacial contact, rapid mass transport, and tailored deposition space for large amounts of Li2O2 are essential for improving the rate performance of Li‐O2 batteries. An ordered mesoporous carbon membrane with continuous macroporous channels was prepared by inversely topological transformation from ZnO nanorod array. Utilized as a free‐standing air cathode for Li‐O2 battery, the hierarchically porous carbon membrane shows superior rate performance. However, the increased cross‐sectional area of the continuous macropores on the cathode surface leads to a kinetic overpotential with large voltage hysteresis and linear voltage variation against Butler–Volmer behavior. The kinetics were investigated based on the rate‐determining step of second electron transfer accompanied by migration of Li+ in solid or quasi‐solid intermediates. These discoveries shed light on the design of the air cathode for Li‐O2 batteries with high‐rate performance.  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