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.
Self nitrogen, oxygen and air-broadened half-widths of the 115-GHz line of CO have been measured at various temperatures between 293 K (room temperature) and 220 K. The temperature dependence of the broadening parameter CCO-XW is described by a power law CCO-XW (T) = CCO-XW(293 K)(T/293)-n co-x. The values of CCOW (293 K) and nCO-X are presented for each broadening gas X, X - CO, N2 and O2. The usual relation CCO-airW (T) = 0.78CCO−N2W(T) + 0.21CCO−O2 W(T) is found to be valid in the temperature and pressure ranges of the present experiments.  相似文献   

2.
Infrared and Raman spectra (10–3100 cm?1) of the layered ZnPSs compound and of intercalates with [Co(η5 ? C5H5)2+] and [Cr(η6 ? C6H6)2+] cations in the polycrystalline state have been recorded 300–10 K temperature range. A complete assignment of the spectra is proposed in terms of PSs group motions, Zn2+ ion translations and [Co(η5 ? C5H5)2+] or [Cr(η6 ? C6H6)2+] internal vibrations. New low frequency for the [Co(η5-C5H5)2+] intercalate at low temperature are assigned to librational and torsional modes of C5H5 rings. Moreover, the preresonance Raman spectra of this intercalate show a selective enhancement for the metal-ligand vibrations when the charge transfer band of the cobalticenium is approached. One concludes that guest molecules are intercalated under their cationic form, are weakly interacting with the host lattice and seem to be dynamically disordered at room temperature.  相似文献   

3.
4.
A modified method is proposed for preparing fullerene compounds with alkali metals in a solution. The compounds synthesized have the general formula Me n C60(THF)x, where Me = Li or Na; n=1–4, 6, 8, or 12; and THF = tetrahydrofuran. The use of preliminarily synthesized additives MeC10H8 makes it possible to prepare fullerene compounds with an exact stoichiometric ratio between C 60 n? and Me +. The IR and EPR spectra of the compounds prepared are analyzed and compared with the spectra of their analogs available in the literature. The intramolecular modes T u (1)-T u (4) for the C 60 n? anion are assigned. The splitting of the T u (1) mode into a doublet at room temperature for Me n C60(THF)x (n=1, 2, 4) compounds indicates that the fullerene anion has a distorted structure. An increase in the intensity of the T u (2) mode, a noticeable shift of the T u (4) mode toward the long-wavelength range, and an anomalous increase in the intensity of the latter mode for the Li3C60(THF)x complex suggest that, in the fullerene anion, the coupling of vibrational modes occurs through the charge-phonon mechanism. The measured EPR spectra of lithium-and sodium-containing fullerene compounds are characteristic of C 60 ? anions. The g factors for these compounds are almost identical and do not depend on temperature. The g factor for the C 60 n? anion depends on the nature of the metal and differs from the g factor for the C 60 ? anion.  相似文献   

5.
A study of energy transfer from optically excited Sm3+ to Nd3+ in borate glass has been performed. Contrary to the observations made by Cabezas and DeShazer, the Sm3+ → Nd3+ energy transfer has been observed as non-radiative. Energy transfer probabilities (Pda) and transfer efficiencies (ηT) have been calculated from our measurements of donor fluorescence intensity and decay times. The mechanism governing the transfer is electrostatic dipole-dipole in nature, contrary to the conclusions made by Nakazawa and Shionoya. At low acceptor (Nd3+) concentrations a linear dependence of Pda on acceptor concentration (C) has been observed which suggests the migration of excitation energy among donors. At high acceptor (Nd3+) concentration a plot of Pda vs (Co + C)2, where Co is donor ion concentration, presents a linear dependence which is consistent with the Fong-Drestler theory of dipole-dipole energy transfer mechanism and interaction of one donor (Sm3+) with two acceptors (Nd3+).  相似文献   

6.
The pressure-tuning Raman spectra of the two methylbenzoate complexes, (η6-C6H5CO2CH3)Cr(CO)2(CX) (X = O, S), have been examined up to ~35 kbar. Structural changes occurred for both complexes in the 10–15 kbar pressure range, most probably as the result of second-order phase transitions. From the observed pressure dependences, replacement of a CO group in the piano-stool (η6-C6H5CO2CH3)Cr(CO)3 molecule by a CS group has a marked influence on the Cr-arene ring vibrational modes, and the arene ring clearly plays a role in determining the nature of the Cr-CO and Cr-CS bonding interactions.  相似文献   

7.
The He(II) spectra of the unsubstituted metallocenes {M(η-C5H5)2}, M  V, Cr, Mn, Fe, Co, Ni and Ru, and of {Mn(η-C5H4Me)2} are reported; both He(I) and He(II) spectra of some decamethylmetallocenes {M(η-C5Me5)2}, where M  Mg, V, Cr, Mn, Fe, Co and Ni, are also given. Intensity changes between the He(I) and He(II) spectra are shown to provide a reliable guide to band assignment. A ligand field treatment of the decamethylmanganocene cation, including limited configuration interaction, gives values for Δ2, B and C; these values are also in good agreement with the photoelectron spectra of {M(η-C5Me5)2} where M  V, Cr and Fe. Overlap between the ligand and metal “d” band structures prevents complete assignment in the cases of Co and Ni.  相似文献   

8.
Low-temperature neutron-diffraction and magnetization measurements carried out on poly- and monocrystalline UIr are explained by a simple collinear ferromagnetic structure with magnetic moments of 0.6(3)μB/U atom oriented along [010] of the monoclinic cell described in space group P21. The bulk magnetic behavior of UIr (TC = 46 K) is not substantially altered by a partial substitution of Ir by Rh, Pt (increase of TC) or Os (decrease of TC) and it is virtually reproduced in the isoelectronic compound UOs0.5Pt0.5.  相似文献   

9.
The electron spin-lattice relaxation times (T 1) of a variety of semiquinone ions in hydrogen bonding solvents have been measured by the pulsed saturation recovery technique as a function of temperature (T) and viscosity (η) of the solvent. Also linewidths (ΔH) have been measured in suitable cases in such solvents at low radical concentrations (~10?4 M). It is observed that (i) the temperature and viscosity dependence ofT 1 can be fitted to an equation of the form 1/T 1=A(T/η)+Bexp(-ΔE/RT) whereA andB are constants and ΔE is an activation energy of the order of 1 kcal mole?1 for these systems; (ii)T 1 is essentially independent of the radical concentration within the range 10?3 to 5×10?2 M; (iii) the concentration independent part of the linewidth (ΔH) increases linearly with (η/T) at sufficiently low temperatures, and (iv) the (η/T) dependent part ofT 1 is sensitive to the size of the semiquinone as well as that of the solvent molecule, whereas the linewidth which is proportional to (η/T) at high viscosity, low temperature region is not sensitive to the size of the semiquinone and that of the solvent. Based on these observations, it is postulated that in hydrogen bonding solvents, three types of motion contribute significantly to electron spin relaxation:
  1. A restricted small step diffusional motion, not involving large changes in the orientation of the molecule, leading to the dominant viscosity dependent contributions toT 1 and ΔH, due to spin rotation interaction;
  2. a large amplitude reorientation of the semiquinone, coupled to translational diffusion, resulting in viscosity dependent contributions toT 1 and ΔH, throughg-modulation;
  3. a hindred rotation of the semiquinone within the solvent cage, contributing toT 1 due to spin rotation interaction.
The fact thatT 1 is not sensitive to the concentration of the radicals, is ascribed to the formation of the solvent cage that prevents the close approach of radicals, thereby rendering radical-radical interactions to be weak mechanisms for relaxation, even at relatively high radical concentrations.  相似文献   

10.
LF‐Muon Spin Relaxation data are reported for the organometallic compounds Pb(C6H5)4, (C6H6)Cr(CO)3 and (C5H5)2Ru. In each case the change in relaxation rate with temperature shows a peak analogous to the T_1 minimum in NMR. The activation parameters were calculated, and the mechanism of muon spin relaxation in the case of (C6H6)Cr(CO)3 is shown to be the reorientation motion of the benzene ring. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

11.
Preparation of (B1−xCx)(Sr1−yBay)2Ca2Cu3Oz has been studied using high-pressure synthesis technique to improve sample quality. Samples were prepared from various starting compositions of 0.2≤x≤0.8 and 0≤y≤1. Nearly single-phased samples were obtained for (B0.6C0.4)(Sr1−yBay)2Ca2Cu3Oz (0≤y≤0.75) and (B1−xCx)-(Sr0.25Ba0.75)2Ca2Cu3Oz (0.3≤x≤0.6). We have found that the partial substitution of Ba is effective to improve the sample quality as well as to enhance the Tc. The C substitution was also demonstrated to affect the sample preparation and the physical properties. Based on the substitution study, a maximum Tc of 120 K was observed for the sample with a starting composition of (B0.65C0.35)(Sr0.3Ba0.7)2Ca2Cu3O9+δ. Critical current density (Jc) and irreversibility field (Hirr) were estimated from magnetization measurements. The Jc at 77 K in a field of 1 T was about 1.1×104 A/cm3 and the Hirr at 77 K was about 2.5 T. The Hirr was well-described by Hirr=a(1−T/Tc)n with a=39.1 and n=2.38.  相似文献   

12.
The thermal chemistry of allyl alcohol (CH2CHCH2OH) on a Ni(100) single-crystal surface was studied by the temperature programmed desorption (TPD) and the X-ray photoelectron spectroscopy (XPS). The allyl alcohol adsorbs molecularly on the metal surface at 100 K. Intact molecular desorption from the surface occurs at temperatures around 180 K, but some molecules exhibit chemical reactivity on the surface: activation of the OH, CC, and CO bonds produces η1(O)-allyloxy CH2CHCH2O(a), η2(C, C) allyl alcohol (C(a)H2C(a)HCH2OH), and η3(C, C, O)-alkoxide (C(a)H2C(a)CH2 O(a)) intermediates. Further thermal activation of allyl alcohol on the surface yields propylene (CH2CHCH3), 1-propanol (CH3CH2CH2OH), propanal (CH3CH2CHO), and combustion and dehydrogenation products (H2O, H2, and CO). Propylene desorbs from the surface at temperatures of around 270 K. Hydrogenation to the η3(C, C, O)-alkoxide intermediate leads to the production of propanal which desorbs from the surface around 320 K, while hydrogenation of the η2(C, C) allyl alcohol intermediate produces 1-propanol, which desorbs at around 310 K. The co-adsorption of hydrogen atoms on the surface enhances the formation of the saturated alcohol, while co-adsorption of oxygen enhances the formation of both the saturated alcohol and the saturated aldehydes.  相似文献   

13.
Shock-tube and flow-reactor experiments were used to study the thermal decomposition of diethyl carbonate (C2H5OC(O)OC2H5; DEC). The formation of CO2, C2H4, and C2H5OH was measured with gas chromatography/mass spectrometry (GC/MS) and high-repetition-rate time-of-flight mass spectrometry (HRR-TOF-MS) behind reflected shock waves. The same products were also detected by GC/MS in flow reactor experiments. All experiments combined span a temperature range of 663–1203 K at pressures between 1.0 and 2.0 bar. Time-resolved species concentration profiles from HRR-TOF-MS and product compositions from GC/MS measurements were simulated applying a detailed reaction mechanism for DEC combustion. A master-equation analysis was conducted based on computed energies from G4 calculations. Quantum chemical calculations confirm that DEC primarily decomposes by six-center elimination, C2H5OC(O)OC2H5 → C2H4 + C2H5OC(O)OH (1a), followed by rapid decomposition of the alkoxy acid, C2H5OC(O)OH → C2H5OH + CO2 (1b). Measured DEC decomposition rate constants k(T) at p ≈ 1.5 bar can be represented by the Arrhenius equation k(T) = 1013.64±0.12 exp(?204.24±1.95 kJ/mol/RT) s ? 1. Theoretical predictions for k1a were in good agreement with experimentally derived values. The theoretical analysis also included dipropyl carbonate (C3H7OC(O)OC3H7; DPC) decomposition and the reactivities of DEC and DPC are compared and discussed in the context of reactivity of dialkyl carbonates under pyrolytic conditions.  相似文献   

14.
The adsorption of CO on Rh(111) has been studied by thermal desorption mass spectrometry and low-energy electron diffraction (LEED). At temperatures below 180 K, CO adsorbs via a mobile precursor mechanism with sticking coefficient near unity. The activation energy for first-order CO desorption is 31.6 kcal/mole (νd = 1013.6s?1) in the limit of zero coverage.As CO coverage increases, a (√3 ×√3)R30u overlayer is produced and then destroyed with subsequent formation of an overlayer yielding a (2 × 2) LEED pattern in the full coverage limit. These LEED observations allow the absolute assignment of the full CO coverage as 0.75 CO molecules per surface Rh atom. The limiting LEED behavior suggests that at full CO coverage two CO binding states are present together.  相似文献   

15.
The unit cell parameters a and c of nonirradiated [N(C2H5)4]2ZnBr4 crystals in the temperature region 90–300 K and of samples irradiated with γ rays to doses of 106 and 5 × 106 R in the 270-to 300-K interval were measured using x-ray diffraction. The data obtained were used to derive the thermal expansion coefficients αa and αc. It is shown that the parameter a increases and the parameter c decreases with increasing temperature. In the vicinity of the phase transition (PT) at T = 285 K, the temperature dependences of a(T) and c(T) reveal anomalies in the form of jumps and the αa(T) and αc(T) curves have a maximum and a minimum, respectively. The heat capacity of nonirradiated and irradiated [N(C2H5)4]2ZnBr4 samples was measured by adiabatic calorimetry. A maximum was found in the C p(T) curve at T = 285 K. Both x-ray diffraction and heat capacity measurements showed that the PT temperature decreased after γ irradiation.  相似文献   

16.
The temperature dependence of the heat capacity at a constant pressure C p 0 = f(T) for the dimerized phase of the C60 fullerene in the temperature range 300–575 K and the thermodynamic characteristics for depolymerization of this phase under normal pressure are investigated using precision differential scanning calorimetry. It is established that thermal depolymerization is a kinetically hindered process. The final products of thermal depolymerization are identified as a partially crystalline monomer face-centered cubic phase of C60 with a degree of crystallinity α = 67 mol %. The results obtained in this study and our previous experimental data on the low-temperature heat capacity are used in the calculations of standard thermodynamic functions for the (C60)2 crystalline dimer, namely, the heat capacity C p 0 (T), the enthalpy H 0(T) ? H 0(0), the entropy S 0(T), and the Gibbs function G 0(T) ? H 0(0) in the temperature range from T → 0 to 394 K.  相似文献   

17.
There are two inequivalent spin 1/2 local baryon field operators that can be constructed from 3 quarks. A priori theJ P =1/2+ baryons can couple to any linear combination of these operators. We show however that the coupling of the 1/2+ baryons to these operators is determined by the value of theSU(3) ratio ofF toD type pseudoscalar-baryon couplings. The experimental value of this ratio implies, for example, that the proton couples strongly to (u T C γ d)u and weakly to (u T C d)γ s u. This is of interest in QCD sum rule applications.  相似文献   

18.
The closeness of low-lying T1u and T1g levels of C 60 ? could enable their mixing under an odd parity vibration of (T1 u + T1 g ? (hg + τ1 u)type. In addition, the two levels are susceptible to Jahn-Teller interaction due to five-fold degenerate hg vibrations. This complex problem of (T1u+T1g)?(hg1u) vibronic interaction is transformed to a form similar to T2g ? (εg + τ2g) vibronic problem of octahedral symmetry. The problem is analysed in an infinite coupling model and compared with the experimental spectroscopic results for the C 60 ? radical. The resulting parameters are used to calculate the pair-binding energy and superconducting transition temperature in C 60 n? fullerides. Vibronic mixing with the T1g level is found to be responsible for maximising the pair-binding energy at the doping level n=3. It is also found to be an important source of Tc enhancement.  相似文献   

19.
The nature and strength of metal–ligand bonds in organotransition‐metal complexes are crucial to the understanding of organometallic reactions and catalysis. Quantum chemical calculations at different levels of theory have been used to investigate heterolytic Fe–N bond energies of para‐substituted anilinyldicarbonyl(η5‐cyclopentadienyl)iron [p‐G‐C6H4NH(η5‐C5H5)Fe(CO)2, abbreviated as p‐G‐C6H4NHFp (1), where G = NO2, CN, COMe, CO2Me, CF3, Br, Cl, F, H, Me, MeO, and NMe2] and para‐substituted α‐acetylanilinyldicarbonyl(η5‐cyclopentadienyl)iron [p‐G‐C6H4N(COMe)(η5‐C5H5)Fe(CO)2, abbreviated as p‐G‐C6H4N(COMe)Fp (2)] complexes. The results show that BP86 and TPSSTPSS can provide the best price/performance ratio and more accurate predictions in the study of ΔHhet(Fe–N)'s. The linear correlations [r = 0.98 (g, 1a), 0.93 (g, 2b)] between the substituent effects of heterolytic Fe–N bond energies [ΔΔHhet(Fe–N)'s] of series 1 and 2 and the differences of acidic dissociation constants (ΔpKa) of N–H bonds of p‐G‐C6H4NH2 and p‐G‐C6H4NH(COMe) imply that the governing structural factors for these bond scissions are similar. And the linear correlations [r = ?0.99 (g, 1c), ?0.92 (g, 2d)] between ΔΔHhet(Fe–N)'s and the substituent σp? constants show that these correlations are in accordance with Hammett linear free energy relationships. The polar effects of these substituents and the basis set effects influence the accuracy of ΔHhet(Fe–N)'s. ΔΔHhet(Fe–N)'s(1, 2) follow the captodative principle. MEα‐COMe, para‐Gs include the influences of the whole molecules. The correlation of MEα‐COMe, para‐Gs with σp? is excellent. MEα‐COMe, para‐Gs rather than ΔΔHhet(Fe–N)'s in series 2 are more suitable indexes for the overall substituent effects on ΔHhet(Fe–N)'s(2). Insight from this work may help the design of more effective catalytic processes. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

20.
The C7H7 potential energy surface was studied from first principles to determine the benzyl radical decomposition mechanism. The investigated high temperature reaction pathway involves 15 accessible energy wells connected by 25 transition states. The analysis of the potential energy surface, performed determining kinetic constants of each elementary reaction using conventional transition state theory, evidenced that the reaction mechanism has as rate determining step the isomerization of the 1,3-cyclopentadiene, 5-vinyl radical to the 2-cyclopentene,5-ethenylidene radical and that the fastest reaction channel is dissociation to fulvenallene and hydrogen. This is in agreement with the literature evidences reporting that benzyl decomposes to hydrogen and a C7H6 species. The benzyl high-pressure decomposition rate constant estimated assuming equilibrium between the rate determining step transition state and benzyl is k1(T) = 1.44 × 1013T0.453exp(−38400/T) s−1, in good agreement with the literature data. As fulvenallene reactivity is mostly unknown, we investigated its reaction with hydrogen, which has been proposed in the literature as a possible decomposition route. The reaction proceeds fast both backward to form again benzyl and, if hydrogen adds to allene, forward toward the decomposition into the cyclopentadienyl radical and acetylene with high-pressure kinetic constants k2(T) = 8.82 × 108T1.20exp(1016/T) and k3(T) = 1.06 × 108T1.35exp(1716/T) cm3/mol/s, respectively. The computed rate constants were then inserted in a detailed kinetic mechanism and used to simulate shock tube literature experiments.  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