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 fundamental vibrations of 13 cyclic SnSe8-n (n = 7—2) molecules have been calculated using a modified Urey—Bradley force field with 9–14 independent force constants whose values have been adapted from those of Se8 and S8. Calculated wavenumbers are compared to those obtained by Raman spectroscopy for sulfur—selenium phases prepared by recrystallizing quenched molten mixtures of the elements as previously described. Agreement between the observed spectra and calculated wavenumbers is closest by assuming the existence of selenium—selenium bonds and the absence of isolated selenium atoms in SnSe8-n molecules. It is assumed that sulfur—selenium phases are mixed crystals with the following components in varying concentrations: 1,2-S6Se2, 1,2,3-S5Se3, 1,2,3,4-S4Se4, 1,2,5,6-S4Se4, 1,2,3-S3Se5 and 1,2-S2Se6. The presence of S8 and Se8 in some of the phases is indicated by the Raman spectra.  相似文献   

2.
A self-consistent-field-Xα-scattered-wave molecular orbital calculation was carried out on the [CpMoS(μ-S)]2(Cp = η5-C5H5) complex. The calculated results were used to rationalize the observed photochemical isomerization of the title complex to [CpMo(μ-S)][μ-S2]. It is proposed that a terminal sulfur (St) → Mo charge-transfer excitation is responsible for the isomerization, which is an intramolecular redox; i.e. Mo(V) is reduced to Mo(IV) and S2− is oxidized to S22− , a result consistent with the charge-transfer character of the excitation. Specifically, the transition responsible for the isomerization is proposed to be 16bu → 18ag (1Ag1Bu). The 18ag orbital is primarily Mo in character but it is also Mo---St π-antibonding; cleavage of the Mo---St π-bond facilitates the isomerization.  相似文献   

3.
The homogeneous width and frequency of S1 ← S0 0-0 transitions of free-base porphin in site B of n-decane are studied by photochemical hole-burning (T = 1.2–4.2 K). A localized phonon mode of 7 cm?1 is identified as a phonon sideband and holes burnt into it yield a lifetime of 115 ± 20 ps. The results are consistent with the exchange model for slow exchange.  相似文献   

4.
The rate of the thermal rearrangement of (S) 2 chloromethyl-1-ethylpyrrolidine [(S)-1a] to (R)-3-chloro-1-ethylpiperidine [(R) 2a] has been examined at three temperatures in benzene by PMR and polarimetry. The rearrangement was shown to be completely stereospecific and to obey a simple first order rate law. The calculated Ea ΔH3 and ΔS3 were 22 ± 2 kcalmole (25°), 21 ± 2.5 kcalmole (25°) and - 10 ± 2 e.u. (0°K) respectively. The effect of solvents having differing dielectric constants was also studied. A transition state 9'a and an ion pair intermediate 3a are suggested for the rearrangement. The stereochemical course of the reactions of (S)-1a, (R)-2a and (S)-2a with hydroxide and methoxide ions have been shown to be 100% stereospecific with an uncertainty of about 1%. The absolute configurations of all optically active reactants and products [(S)- and (R)-4a, (S)-4b (R)- and (S)-5a, (R)-5b, (S,S')-6a, (S,R')-7a and (R,R')-8a] were established by chemical correlations with known compounds or by ORD and chemical inference. The ring opening of both the primary and secondary aziridinium ion positions of 1-azonia-1-ethylbicyclo [3.1.0]hexane [(S)-3a] by nucleophiles proceeds entirely by SN2 processes. The conversion of (R)-1-ethyl-3-hydroxypiperidine [(R)-5a] to (S)-2a. HCl with thionyl chloride in chloroform proceeds by inversion with 4.8% racemization, whereas the thermal rearrangement of (S)-1a to (R)-2a occurs with complete retention of absolute configuration.  相似文献   

5.
Abstract

P4S10 suffers a fast oxido-reductive dissociation when melted, losing one of its four terminal S atoms to yield P4S9, which is characterized by nmr, irand Raman spectroscopy. This dissociated state is largely retained in the cooled melts, which appear to be made up of unaltered P4S10 crystals embedded in an amorphous matrix containing the dissociation products; such melts give rise easily to supersaturated solutions in CS2, from which crystals of pure P4S9—form I separate on standing.

The same holds true for commercial “P2S5,” for which a definite positive correlation was found between the P and P4S9 contents.

Examples of redox dissociation of other P/S compounds are quoted for the sake of comparison.  相似文献   

6.
The determination of the enthalpy ΔH and entropy ΔS of the isomerization bicyclopropy l(trans)? bicyclopropyl(gauche) in the liquid phase by the IR intensity method is described. It is assumed that the ratio of the integral absorption coefficients of the two reference bands at 1351 cm?1 (gauche) and 1291 cm?1 (trans), which both belong to the same type of vibration, is temperature independent. The two values ΔH = ?160 ± 40 cal Mol?1 and ΔS = ?0.4 ± 0.5 cal (Mol · Grad)?1 respectively. ΔSU = ?1.8 ± 0.5 cal (Mol · Grad)?1 thus determined agree well with the corresponding results obtained from NMR and electron diffraction measurements. However, from the pair of reference bands at 695 cm?1 (gauche) and 1291 cm?1 (trans), which do not belong to the same type of vibration, strongly differing values for ΔH and ΔSresult under the same assumption as above, which apparently is not applicable in this case.It is shown through these data that the “Fateley-Test” does not provide a suitable tool to decide whether the absorption coefficients of the reference bands are temperature independent or not. The reason for this insignificance is the relatively poor accuracy and reproducibility of measured IR band intensities obtainable up to now.The relative density of bicyclopropyl between ?60°C and + 50°C was determined.  相似文献   

7.
The temperature dependence of the fluorescence and fluorescence excitation spectra of all-trans diphenyl hexathene (DPH) and octatetraene (DPO) in six solvents confirms the S1(1Ag*) and S2(1Bu*) state assignment, and determines their energy difference ΔE. The S1 fluorescence rate parameter kF depends on ΔE, the solvent refractive index n, the S2 (n = 1) fluorescence rate parameter kF20 (2.23 × 108 s?1 for DPH, 2.33 × 108 s?1 for DPO), and the S2-S1 coupling matrix element V (745 cm?1 for DPH, 500 cm?1 for DPO). The S1 fluorescence is induced by 1Bu*-1Ag* potential interaction (PI), via a bu vibrational mode (≈ 900 cm?1), and not by vibronic coupling. The main S1 radiationless transition, rate parameter kR, is thermally-activated internal rotation through an angle θ about the central ethylenic bond(s). The PI distorts the S1 (θ) potential surface and thus influences kR.  相似文献   

8.
The molecular structure and conformation of 2,3-dichloro-1-propene have been determined by gas-phase electron diffraction at nozzle temperatures of 24, 90 and 273°C. The molecules exist as a mixture of two conformers with the chlorine atoms anti (torsion angle ∠φ = 0°) or gauche (∠φ = 109°) to each other and with the anti form the more stable. The composition (mole fraction) of the vapor with uncertainties estimated at 2σ was found to be 0.55 (0.08), 0.49 (0.08) and 0.41 (0.10) at 24, 90 and 273°, respectively. These values correspond to an energy difference with estimated standard deviation ΔE° = E°g-E°a = 0.7 ± 0.3 kcal mol?1 and an entropy difference ΔS° = S°g-S°a = 0.6 ± 0.9 cal mol?1 K?1. Some of the diffraction results, together with spectroscopic observations, permit the evaluation of an approximate torsional potential function of the form 2V = V1 (1 - cos φ) + V2 (1 - cos 2φ) + V3 (1 - cos 3φ); the results are V1 = 4.4 ± 0.5, V2 = ?2.9 ± 0.5 and V3 = 4.8 ± 0.2, all in kcal mol?1. The results at 24°C for the distance (ra) and angle (∠α) parameters, with estimated uncertainties of 2σ, are: r(Csp2-H) = 1.098(0.020)Å, r(Csp3-H) = 1.103(0.020)Å, r(CC) = 1.334(0.009)Å, r(C-C) = 1.504(0.013)Å, r(Csp2-Cl) = 1.752(0.021)Å, r(Csp3-Cl) = 1.776(0.020)Å, ∠C-CC = 127.6(1.1)°, ∠Csp3-Csp2-Cl = 110.2(1.0), ∠Csp2-Csp3-Cl = 113.1(1.2)°, ∠H-Csp3-H = 109.5° (assumed), ∠CC-H = 120.0° (assumed) and ∠φ = 108.9(3.4)°.  相似文献   

9.
The unsymmetrical mono-tertiary stibines dimethyl(α-picolyl)stibine (picstib), dimethyl(8-quinolyl)stibine (quinstib), and (R;S)-methylphenyl(8-quinolyl)stibine (R;S-quinstib) have been synthesised and the square-planar complexes [MX2(picstib)], [MX2(quinstib)] (where M = Pd or Pt and X = Cl, Br, I or SCN) and [MCl2(R;S-quinstib)] (where M = Pd or Pt) isolated. The thiocyanato derivatives display linkage isomerism. The octahedral complexes [M(CO)4-(picstib)] and [M(CO)4(quinstib)] have also been prepared from the metal hexacarbonyls and the appropriate ligands by UV irradiation in tetrahydrofuran.  相似文献   

10.
The S1← So electronic transitions of toluene involving also some internal rotational levels were observed for the first time in the multiphoton ionization spectrum in a supersonic jet. A large population in several low-lying internal rotational levels and a strong coupling between electronic motion and the internal rotation are suggested.  相似文献   

11.
The ultraviolet and vacuum-ultraviolet photolyses of GeH4, GeH3Cl, GeH2Cl2, GeHCl3, GeCl4, GeH3Br, and GeH2Br2 in argon and carbon monoxide matrices have been performed between 4° and 24° K. The results of these experiments support the successful isolation and characterization of a variety of germanium free-radical species produced as a result of primary and secondary photolytic processes. In order to control, characterize and understand the various elementary processes (photochemical and fragment diffusion) occurring in the matrix, both in situ photolysis as well as simultaneous deposition and photolysis were performed as a function of photolyzing radiation, temperature (of the cold window), and concentration (M/ R = matrix/reactive material).The identification and structure of the photochemically produced free radicals were obtained from the vibrational (infrared) spectrum before and after photolysis. In some cases, complete vibrational assignments were possible. Force field (normal coordinate) calculations were also performed in order to corroborate these assignments, within the limitation of their application to matrix isolated spectra.Finally, the structural and bonding properties of these free radicals, as reflected by their vibrational frequencies, are compared with similar stable and unstable Group IVA species in an effort to understand the major effect(s) controlling their geometry. The geometries obtained on the basis of the experimental observations are compared with those predicted by Walsh's semiempirical MO treatment and Self-Consistent Extended Huckel calculations. In addition, a simple thermodynamic argument in conjunction with elementary quantum mechanics is used as a tool for predicting the structure of some of these simple free-radical species.  相似文献   

12.
Chloroacetyl chloride is studied by gas-phase electron diffraction at nozzle-tip tempera- tures of 18, 110 and 215°C. The molecules exist as a mixture of anti and gauche confor- mers with the anti form the more stable. The composition (mole fraction) of the vapor with uncertainties estimated at 2σ is found to be 0.770 (0.070), 0.673 (0.086) and 0.572 (0.086) at 18, 110 and 215°C, respectively. These values correspond to an energy difference with estimated standard deviation ΔEo = Eog -Eoa = 1.3 ± 0.4 kcal mol?1 and an entropy difference ΔSo = Sog -Soa = 0.7 ± 1.1 cal mol?1 K?1. Certain of the diffraction results permit the evaluation of an approximate torsional potential function of the form 2V = V1(1 - cos φ) + V2(1 - cos 2φ) + V3(1 - cos 3φ); the results are V1 = 1.19 ± 0.33, V2 = 0.56 ± 0.20 and V3 = 0.94 ± 0.12, all in kcal mol?1. The results for the distance (ra), angle (∠α) and r.m.s. amplitude parameters obtained at the three temperatures are entirely consistent. At 18°C the more important parameters are, with estimated uncertainties of 2σ, r(C-H) = 1.062(0.030) Å, r(CO) = 1.182(0.004) Å, r(C-C) = 1.521(0.009) Å. r(CO-Cl) = 1.772(0.016) Å, r(CH2-Cl) = 1.782(0.018) Å, ∠C-C-0 = 126.9(0.9)°, ∠CH2-CO-C1 = 110.0(0.7)°,∠CO-CH2-C1 = 112.9(1–7)°, ∠H-C-H = 109.5° (assumed), ∠φ (gauche torsion angle relative to 0° for the anti form) = 116.4(7.7)°, δ (r.m.s. amplitude of torsional vibration in the anti conformer) == 17.5(4.2)°.  相似文献   

13.
The kinetics of the mercuration of 2-methylazobenzene in methanol were studied. The thermodynamic data found were ?Eact = 22.7 kcal mol?1, ?H1 = 22.0 kcal mol?1, and ?S1 = ?12.3 eu. In comparison with a value of ?S1 ? ?20 eu for the mercuration of benzene, this lowered entropy is taken as evidence for complex formation between mercuric acetate and 2-methylazobenzene before and during the rate determining step of electrophilic substitution.  相似文献   

14.
The title compound, Cu(S2CNEt2)2, behaves at low temperatures (1–20 K) as a normal spin-1/2 molecule, with 〈g〉 =2.06 and the Curie-Weiss θ = +0.25 K. This result contradicts an earlier investigation that led to the suggestion that the crystallographically-occurring dimers are coupled ferromagnetically.  相似文献   

15.
Among the many sources of error which may affect the results for the enthalpy ΔH and entropy ΔS of isomerization determined by the IR-intensity method, the influence of the temperature dependence of the absorptivities of the reference bands, αg(T) and αt(T), is studied theoretically. Based on the general case I, the following three special cases II–IV are examined, in which increasingly stronger constraints are imposed on the compound-specific functions αg(T) and αT(T): II. Constant ratio αg(T)/αT(T), III. αg(T) and αT(T) both temperature independent and different, IV. αg(T) and αT(T) both temperature independent and equal. For each case the formulae for the evaluation of ΔH and ΔS as well as the kind and the number of the quantities to be measured are given. Moreover, in the cases I and II new methods for the analysis of the measured data are suggested. In each case the plots of the normalized band intensities Ag and At and the temperature T versus each other in different two-dimensional diagrams have been examined to find out whether one of the simpler special cases is applicable for the respective pair of reference bands. It is shown that these plots are not a suitable tool because of the relatively poor accuracy and reproducibility of measured IR-band intensities obtainable up to now.  相似文献   

16.
At a temperature above 1173 K the spinel-type structure of CdHo2S4 becomes unstable and gradually transforms into a rocksalt-type structure. During transition both Cd2+ and Ho3+ cations shift. Electron diffraction patterns obtained during transition show that the Cd, which shifted from a tetrahedral to an octahedral site, distorts the sulfur lattice. This distortion becomes visible as a modulation of the sulfur matrix. The modulation originally occurs in many crystallographic directions, however, upon annealing the 〈111〉 direction becomes preferred. Small domains of the high-temperature structure of CdHo2S4 appear to occur also in samples that were annealed well below 1173 K for a long period.  相似文献   

17.
Molecular structures and energies have been calculated in the MNDO approximation, for P4S3 and its molecular ion P4S3+, and for the mass spectral fragment pairs: (P3S3+ + P), (P3S2+ + PS), (P3S+ + PS2), (P2S3+ + P2), (P2S2+ + P2S), (P2S+ + P2S2), (P2S2), PS3+ + P3), (PS2+ + P3S), (PS+ + P3S2), and (PS+ + P2S + PS). Three distinct energy minima were found for each of P2S2+ and P2S2, and two minima for each of P2S+, P2S, PS3+, PS3+, PS2+, PS2, P3+ and P3. The fragments arising from P4 and P4+ were also investigated. The structures are discussed in terms of the Jahn—Teller effect, whose predictions are fulfilled without exception.  相似文献   

18.
Arguments based on bond-stretching frequencies, nuclear quadrupole coupling constants and bond lengths indicate that bonds having the same strength factor can have different electronic structures. It is then argued that a more useful probe for obtaining an insight into bond character is provided by infrared band intensities which can be analyzed to yield values for the constants occurring in bond moment functions. A new bond moment theory, in which the concept of “rehybridization moments” is introduced, is described. Using this theory, self-consistent values are obtained for the bond moment constants governing the band intensities for the molecules HCN, C6H6 and C2H4.  相似文献   

19.
The molecular structure of allyl silane has been studied by gas-phase electron diffraction. The experimental radial distribution curve has only four prominent peaks, resulting in serious resolution problems in the structure determination. A single conformer whose dimensions resemble those of related molecules fits the diffraction data. The torsion angle φsiccc is102 ± 1°, measured from the conformation having Si-C and CC eclipsed.  相似文献   

20.
The chemistry and thermodynamics of vaporization of CdGa2S4(s), CdGa8S13(s), and Ga2S3(s) were studied by computer-automated, simultaneous Knudsen-effusion and torsion-effusion, vapor pressure measurements in the temperature range 967–1280 K. The vaporization was incongruent with loss of Cd(g) + 1/2 S2(g) and production of CdGa8S13(s), a previously unknown compound, in equilibrium with CdGa2S4(s), until the solid became CdGa8S13 only. Then, incongruent vaporization continued with production of Ga2S3(s) until the solid was Ga2S3 only. The latter vaporized congruently. The ΔH°(298 K) of combination of one mole of CdS(s) with one mole of Ga2S3(s) to give CdGa2S4(s) was ?22.6 ± 0.9 kJ mole?1. The 2H2(298 K) of combination of one mole of CdS(s) with four moles of Ga2S3(s) to give CdGa8S13(s) was ?25.5 ± 1.1 kJ mole?1. The 2H2(298K) of CdGa8S13(s) with respect to disproportionation into CdGa2S4(s) and 3 Ga2S3(s) was ?2.8 ± 0.6 kJ mole?1. CdGa8S13(s) was not observed at room temperature. The 2H2(298 K) of vaporization of the residual Ga2S3(s) was 663.4 ± 0.8 kJ mole?1, which compared well with a value of 661.4 ± 0.3 kJ mole?1 already available from the literature. Implications of small variations in stoichiometry of compounds in this study were observed and are discussed.  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