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 546 毫秒
1.
The Raman spectra for (SiF2)x and GeF2 are reported. In the spectrum of (SiF2)x a band has been observed at 411 cm−1 and is most probably associated with SiSi bonding in the compound; this is the first direct evidence for such a bond. The high degree of similarity between the spectra of (SiF2)x and (CF2)x suggests that (SiF2)x is a zig-zag helical chain polymer similar to (CF2)x.  相似文献   

2.
We have previously reported that reduction of C6F13I in (DMF, LiClO4), yields RFHgRF, produced through the reduction of RFHgI itself formed at polarized mercury. The use of LiCl in the place of LiClO4 affords a possibility to suppress the organometallic route and yields the carbanion C6F13, formed through a direct bielectronic reduction. This feature is exploited to perform electrocarboxylation and sulfoxidation. In this way, RFCO2H and RFSO2Cl are easily electrosynthesized in macroscale experiments. The anion effect remarked between Cl? and ClO4? is studied in the case of Br?, I?, BF4? with various cations.  相似文献   

3.
Generalizations are made on the effect of the nature of a precursor (the original salt of a metal) on the sorption activity of hydrogels of oxidehydroxides (OHes) towards oxalate ions using the example of an OH obtained by the alkaline hydrolysis of chloride, perchlorate, and sulfate of Fe(III); chloride, sulfate, and nitrate of Al; and nitrate of Zr(IV). It is established that the sorption of C2O 4 2? on the studied OHes is described by the Langmuir equation. We find that the sorption activity depends on the nature of the precursor: Al2(SO4)3 > Al(NO3)3 > AlCl3 > ZrO(NO3)2 > Fe2(SO4)3 > FeCl3 > Fe(ClO4)3.  相似文献   

4.
The effect of the acidity of the medium on the hydroxylation and nitration of alkanes (RH) in 90–98% H2SO4 at 25°C is described quantitatively by a model taking account of the thermodynamic activity of the RH, H2O, and HSO4- particles. It was concluded that in the transition states the reagents H3O2+HSO4- and NO2+ HSO4- are present as HO+ and NO2+ ions without the bases H–O and HSO4-, the alkanes are present without hydrophobic shells, and the initial reaction products are ROH2+ and RNO2H+.  相似文献   

5.
The first step in the reduction of the dinitrogen ligand in (Cp2TiR)2N2 (R  C6H5, m-, p-CH3C6H4, C6F5, CH2C6H5) by sodium napthalene (NaC10H8) involves the removal of one Cp group per titanium atom. The resulting diimide precursor reacts with a second mole of NaC10H8 with formation of a hydrazine precursor. This compound is thermally unstable and decomposes to an ammonia precursor. A minor part of the hydrazine precursor abstracts a proton from the solvent.  相似文献   

6.
A number of products formed in reactions of cobalt(II) salts with monoethanolamine (HEtm) in a neutral medium were synthesized and studied. X-Ray diffraction study showed that the nitrate and acetate form the dimers [Co(HEtm)3][Co(Etm)3](NO3)3 and [Co(HEtm)3][Co(Etm)3](CH3COO)3 · 8H2O, respectively. In chloride solutions, cobalt is partially oxidized to give the trinuclear complex [CoII{CoIII(Etm)3}2]Cl3 · H2Etm · 2H2O. The reaction of the chelate [Co(Etm)3] · 3H2O with nitric acid is accompanied by complete protonation of the coordinated aminoethanolate ions, and the reaction with formic acid involves complete replacement of the coordinated ligand by acid residue anions and water molecules to give the coordination polymer {Co2(μ-HCOO)4(H2O)4} n .  相似文献   

7.
A series of perfluoropolyether bis‐carboxylic esters was synthesized and their hydrolytic stability investigated. Their formula is ROOCCF2O(CF2CF2O)p(CF2O)qCF2COOR, where p/q = 1.07 and p + q = 2.94. The alkyl group, R, varied both in terms of steric hindrance and electron‐withdrawing ability. Kinetic and thermodynamic data were obtained under homogeneous conditions and compared to a fully hydrogenated ester having a closely related structure CH3(CH2)3OOCCH2O(CH2CH2O)nCH2COO(CH2)3CH3, where n? = 10.6. Neutral ester hydrolysis (NEH) conditions were selected with methyl ethyl ketone as a solvent and a 3–4:1 water/ester ratio. The course of the reaction was monitored by 19F NMR or 1H NMR (when R = CH3CH2? ). Results indicated that the hydrolysis of fluorinated esters, with alkyl aliphatic substituents, is governed by steric hindrance of the substituents. Two distinctive kinetic regimes were observed. The first one, at low conversion, was characterized by lower kinetic constants and related to true NEH conditions. The second regime appeared at higher conversion when acidic autocatalysis dictated the reaction behavior. This is the only observed mechanism when esters more sensitive to the hydrolysis are considered. In these cases, polar factors prevail over steric considerations. Finally, all fluorinated esters of the class (I) showed a much higher reactivity than the hydrogenated ester whose hydrolysis took place only in the presence of a strong acidic catalyst. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 4266–4280, 2002  相似文献   

8.
By minimizing the global variance in the 1-reduced local-energy matrix E1( X 1; X 1′), subject to the normalization of the 1-reduced density matrix ρ1( X 1; X 1′), one derives an integral matrix equation for E1( X 1; X 1′) as a functional of ρ1( X 1; X 1′) at the location ( X 1; X 1′) of an arbitrary member of an N (≥ 2)-particle system. The implications for the possible local improvement in the accuracy of approximate wave functions through the imposition of global constraints are briefly discussed.  相似文献   

9.
The MINDO/3 technique gives geometries for (CH4)+, (CCl4)+ and the intermediate ions (CHnCl4 ? n)+ (n = 1, 2, 3) which have symmetries in precise accord with the predictions of the Jahn—Teller effect. The ground state of (CH4)+ has D2d symmetry, with a C3v structure ca. 45.6 kJ mol?1 higher. (CCl4)+ has a C2v ground state, with a D2d structure ca. 144 kJ mol?1 higher: no bound state of C3v symmetry could be found. (CH3Cl)+ and (CHCl3)+ both have Cs symmetry, and (CH2Cl2)+ has C2v symmetry. The analogous fluoro ions are discussed briefly.  相似文献   

10.
The first three reactions of the Calcote mechanism for soot formation, that is, C3H 3 + +C2H2→C5H 5 + , C5H 5 + →C5H 3 + H2, and C5H 3 + +C2H2→C7H 5 + , have been studied based on chemi-ions withdrawn directly from a premixed methane-oxygen flame by supersonic molecular beam sampling. The first reaction is reversible and involves the formation of a specific encounter complex sensitive to pressure and ion kinetic energy. The second reaction appears to require large amounts of internal energy in the C5H 5 + ion to proceed. The third reaction is reversible; however, in contrast to the initiating reaction, the C5H 3 + ion formed from the [C7H 5 + ]* complex exhibits a much lower reactivity. The conclusions are based on ion-molecule reactions as well as collision activation mass spectrometry of isolated chemi-ions. In addition, the product distributions as functions of pressure and ion kinetic energy were studied.  相似文献   

11.
The optimal mixing coefficient C of the exchange energy Ex and the electron-electron interaction part of the exchange-correlation energy W1xc in the formula for the total exchange-correlation energy Exc was expressed through the ratio of the kinetic Tc and potential Wc contributions to the correlation energy Ec. This expression can be derived from a Heavyside step function model of the dependence of Wλxc on the coupling parameter of the electron interaction λ. Values of Tc and Wc obtained from ab initio wave functions were used to estimate C for a number of atoms and molecules. A strong dependence of Tc, Wc, and C on the bond distance was demonstrated for the case of the H2 molecule. Tc and C approach zero in the bond-dissociation limit; so for an electron-pair bond, the admixing of exact exchange to obtain an accurate Exc is strongly dependent on the bond length and has to disappear for weak interaction/large bond distances. The potential of the exchange-correlation hole constructed for H2 from an ab initio second-order density matrix was compared with its generalized gradient approximation (GGA). © 1996 John Wiley & Sons, Inc.  相似文献   

12.
Density functional theory (PBE with a modified Dirac-Coulomb-Breit Hamiltonian) is used to simulate the adsorption of hydrocarbons (C2H2, C2H4, C2H6) on the surface of a sorbent containing Ag0, Agδ+, and AgO sites. The dynamics of change in the structural characteristics of Ag n (n ≤ 10) is analyzed and the adsorption of oxygen on Ag8 and Ag10 is studied to select the adsorption site model. Studying the interaction of hydrocarbons with Ag8, Ag10, Ag 10 + , Ag10O, and Ag10O2 clusters reveals that the presence of oxygen leads to an increase in the activation of unsaturated hydrocarbons, and the adsorption energy of C2H2 increases tenfold. It is found that the role of adsorbed oxygen is not only to form adsorption sites of hydrocarbons (Agδ+) but also to bind C2H2 and C2H4 directly to the sorbent’s surface.  相似文献   

13.
Although integral to remote marine atmospheric sulfur chemistry, the reaction between methylsulfinyl radical (CH3SO) and ozone poses challenges to theoretical treatments. The lone theoretical study on this reaction reported an unphysically large barrier of 66 kcal mol−1 for abstraction of an oxygen atom from O3 by CH3SO. Herein, we demonstrate that this result stems from improper use of MP2 with a single-reference, unrestricted Hartree-Fock (UHF) wavefunction. We characterized the potential energy surface using density functional theory (DFT), as well as multireference methodologies employing a complete active-space self-consistent field (CASSCF) reference. Our DFT PES shows, in contrast to previous work, that the reaction proceeds by forming an addition adduct [CH3S(O3)O] in a deep potential well of 37 kcal mol−1. An O−O bond of this adduct dissociates via a flat, low barrier of 1 kcal mol−1 to give CH3SO2+O2. The multireference computations show that the initial addition of CH3SO+O3 is barrierless. These results provide a more physically intuitive and accurate picture of this reaction than the previous theoretical study. In addition, our results imply that the CH3SO2 formed in this reaction can readily decompose to give SO2 as a major product, in alignment with the literature on CH3SO reactions.  相似文献   

14.
Phase relations up to the solidus line in part of the Sb-Zn-O system have been investigated over the entire concentration range of the α-Sb2O4-ZnO system in air (= 0.21 atm) using XRD and DTA/TG. The components of this system in air form ZnSb2O6 and Zn7Sb2O12. The results allow division of the system into three subsystems, i.e. α-Sb2O4-ZnSb2O6; ZnSb2O6Zn7Sb2O12 and Zn7Sb2O12-ZnO. The temperature ranges over which the ZnSb2O6 and Zn7Sb2O12 remain at equilibrium with other solid compounds depend on the gaseous atmosphere.   相似文献   

15.
The two compounds, Me4Si2 (C2H3)2 and Me8Si2 (C2H3)2 have been studied by the anti-symmetrized free electron molecular orbital method. Electron delocalisation over the whole chain via the silicon atoms occurs and a satisfactory account of the electronic spectra may be obtained.
Zusammenfassung Me4Si2 (C2H3)2 und Me8Si4 (C2H3)2 wurden nach der MO-Methode des freien Elektronengases behandelt. Delokalisierung der Elektronen über die Si-Atome der Kette wird festgestellt. Die berechneten Spektren sind zufriedenstellend.

Résumé On a étudié par la méthode des orbitales moléculaires d'électrons libres avec antisymétrisation les composés Me4Si2(C2H3)2 et Me8Si4(C2H3)2. Il se produit une délocalisation électronique le long de toute la chaÎne par l'intermédiaire des atomes de silicium, et l'on peut rendre compte d'une manière satisfaisante des spectres électroniques.


We wish to thank the S.R.C. for a maintenance grant to one of us (D.R.A.).  相似文献   

16.
The decomposition of thin surface oxide films on polycrystalline palladium Pd(poly) at 500–1300 K was investigated by mathematical modeling. This process was analyzed in terms of a model including O2 desorption from the chemisorbed oxygen layer (Oads) and the passage of oxygen inserted under the surface layer of the metal (Oabs) and oxygen dissolved in metal subsurface layers (Odis) to the surface. O2 desorption was modeled on a surface with a square lattice of adsorption sites, with account taken of the energy of the lateral repulsive interactions between adjacent Oads atoms (εaa). At εaa = 10 kJ/mol and when the activation energy of O2 desorption for a chemisorbed-oxygen surface coverage of θ ≈ 0 is Edes0 = 230 kJ/mol, the calculated spectra are in agreement with the oxygen temperature-programmed desorption (TPD) spectra obtained for Pd(poly) at θ ≤ 0.5. The passage of Oabs and Odis atoms to the surface was calculated using a first-order equation, with account taken of the activation energy for these atoms coming out to the surface (E2 and E3, respectively). As the oxide film is heated, O2 desorption is accompanied by the passage of Oabs and then Odis to the surface, which leads to an increase in the Oads surface coverage and, accordingly, to a buildup of lateral surroundings in the adsorbed layer. Owing to this fact and to the repulsive interactions between Oads atoms, the bonds between Oads and the surface weaken and Edes decreases. As a consequence, the O2 desorption rate increases and a low-temperature peak with Tmax ≈ 710 K, which is due to the passage of Oabs atoms to the surface, and then a high-temperature peak with Tmax ≈ 770 K, which is due to the passage of Odis atoms to the surface, appear in the TPD spectrum. At εaa = 10 kJ/mol, Edes0 = 230 kJ/mol, E2 = 145 kJ/mol, and E3 = 160 kJ/mol and when the number of inserted oxygen monolayers is θabs ≤ 0.3 and the number of oxygen monolayers dissolved in subsurface layers is θdis ≤ 10, the TPD spectra calculated for the given model are in agreement with the O2 TPD spectra that are observed for Pd(poly) and are due to the decomposition of surface oxide films.  相似文献   

17.
Deconvolutions of measured absorption line profiles in the 1n0 (n = 0 to 5) and the 320 bands of the Ã2A2X?2B1 electronic transition of ClO2 reveal subnanosecond lifetimes for all rotational levels of the 2A2 state. Observed ratios of radiationless rates from spin-doublet components identify direct spin-orbit coupling of the 2A2 state with 2A1 and/or 2B1 vibronic states as a predominant predissociation mechanism. Variations of rates with ν′1 locate an intersection of a second potential surface with that of the 2A2 state.  相似文献   

18.
Density functional theory methods (B3LYP and BP86) indicate that the preferred structures for such early transition metal derivatives are (η8-C8H8)M(η4-C8H8) (M = Ti, V, Cr) with one octahapto η8-C8H8 ring and one tetrahapto η4-C8H8 ring. In such structures only 12 of the 16 carbon atoms of the two C8H8 rings are bonded to the metal, leading to 16-, 17-, and 18-electron complexes, respectively, in accord with the experimentally known structures for the Ti and V derivatives. The preferred structures for the Mn and Fe derivatives are (η6-C8H8)M(η4-C8H8) (M = Mn, Fe) with one hexahapto and one tetrahapto C8H8 ring and thus having 17- and 18-electron configurations, respectively, in accord with experimental data on the iron complex. The lowest energy structure for the cobalt complex is (η4-C8H8)Co(η2,2-C8H8) with two different types of tetrahapto C8H8 rings and a 17-electron metal configuration. The nickel complex (C8H8)2Ni appears to prefer a structure with a 16-electron configuration and two trihapto C8H8 rings, similar to the known (η3-C3H5)2Ni rather than a bis(tetrahapto) structure with the favored 18-electron configuration. These theoretical studies indicate that in (C8H8)2M derivatives of the first row transition metals, the number of carbon atoms in the pair of C8H8 rings involved in the bonding to the central metal atom gives the metal atoms 16-, 17-, or 18-electron configurations.  相似文献   

19.
Summary Characteristic directions of the mass-spectrometric fragmentation of teucrins H are the elimination of -vinylfuran and a carboxy group, and also the formation of furan-containing ions with the compositions C11H12O3 (m/e 192), C10H10O3 (m/e 178), C11H11O2 (m/e 175), C11H10O2; (m/e 174), C10H9O2 (m/e 161), C10H8O2 (m/e 160), C9H10O2 (m/e 150), C10H11O (m/e 147), C9H9O (m/e 133), C6H7O (m/e 95), C6H6O (m/e 94) CsHzO (m/e 81).Translated from Khimiya Prirodnykh Soedinenii, No. 6, pp. 727–736, November-December, 1978. Original article submitted May 3, 1978.  相似文献   

20.
The theoretical knowledge about the zinc-zinc bond has been recently expanded after the proposal of a zinc-zinc double bond in several [Zn2(L)4] compounds (Angew. Chem. Int. Ed. 2017 , 56, 10151-10155). Prompted by these results, we have selected the [Zn2(CO)4] species, isolobally related to ethylene, and theoretically investigated the possible η2-Zn2-coordination to several first-row transition metal fragments. The [Zn2(CO)4] coordination to the metal fragment produces an elongation of the dizinc bond and a concomitant pyramidalization of the [Zn(CO)2] unit. These structural parameters are indicative of π-backdonation from the metal to the coordinated dizinc moiety, as occurred with ethylene ligand. A quantum theory of atoms in molecules study of the Zn Zn bond shows a decrease of ρBCP, ∇2ρBCPZn∩Znρ and delocalization indexes δ(Zn,Zn), relative to corresponding values in the parent [Zn2(CO)4] molecule. The Zn Zn and M Zn bonds in these [(η2-Zn2(CO)4)M(L)n] complexes can be described as shared interactions with an important covalent component where the Zn Zn bond is preserved, albeit weakened, upon coordination.  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