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.
ENDOR measurements at 25 K have been used to determine the hyperfine coupling tensors for all ten protons in the VO(H2O)5 2+ ion in single crystals of Mg(NH4)2(SO4)26H2O. The traceless components of all the tensors are close to axial and their use in a point dipole treatment enables a very plausible geometrical model of the complex ion to be constructed. Six of the protons in the equatorial water molecules have substantial positive isotropic couplings and it is suggested that these reflect the direct admixture of hydrogen 1s components into the singly occupied orbital.  相似文献   

2.
Comparative investigations of secondary ion emission, electron induced ion emission and flash filament signals from polycrystalline vanadium surfaces exposed to well-defined O2, H2, H2O and (O2 + H2) doses (<500 L) have been carried out. The vanadium target could be heated and bombarded by either electrons (300 eV) or ions (3 keV) under ultra high vacuum conditions (<10?10 Torr). The investigations were carried out with a computer controlled ultra high vacuum mass spectrometer. The experimental results establish exact reproducible spectra of well defined surface layers. They give detailed insight into the reactions between H2, O2 H2O and vanadium, and some interactions between these species. They further indicate the importance of bulk and surface diffusion as well as the influence of the probing ion and electron bombardment. A clear distinction between bulk oxygen, surface oxides, and adsorbed oxygen for the vanadium-oxygen interaction at room temperature could be established. For the interaction of hydrogen with clean and oxygen covered vanadium surfaces the formation of adsorbed hydrogen, bulk solution of hydrogen, and the formation of OH groups and H2O could be demonstrated. A detection limit below 10?5 of one single monolayer for metal bonded hydrogen could be established.  相似文献   

3.
《Surface science》1992,262(3):L139-L143
Selective reaction of copper dipivaloylmethanate (Cu(DPM)2, β-diketonate complex) with H2O adsorbed on the SrTiO3(100) surface is investigated. Adsorption of H2O on SrTiO3(100) occurs on Ti3+ sites produced by prior Ar+ ion bombardment. It is found that the β-diketonate complex, Cu(DPM)2, reacted selectively with the adsorbed H2O at room temperature. Further reaction of the adsorbed Cu(DPM)2 with H2O vapor results in the removal of the ligand, DPM, leaving elemental copper on the surface.  相似文献   

4.
Vibrational wavenumbers (IR and Raman) have been calculated for the diperiodate ion H4I2O102− on the basis of a DFT method and assigned to experimental wavenumbers obtained from CuH4I2O10·6H2O. To obtain vibrational wavenumbers in the range comparable to the experiment it was necessary to use the complex [(H4I2O10)7(Cu(H2O)6)6]2−. Smaller complexes lead to much too high wavenumbers for the O‐H stretching vibrations and to too small wavenumbers in the range of the internal vibrations of the anion. On the basis of the results of the calculations an assignment of the Raman lines observed for CuH4I2O10· 6H2O to the vibrational modes is given. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

5.
The electron paramagnetic resonance spectra of Cu2+ in [Cd(sac)2(H2O4]·2H2O and [Cd(sac)2(HydEt-en)2] (HydEt-en=N-(2-hydroxyethyl)-ethylenediamine) single crystals and powder were examined at room temperature. A detailed study of the spectra of the compounds indicates the replacement of Cd2+ in the host compounds with Cu2=. [Cd(sac)2(H2O)4]·2H2Oshows the presence of two sites for Cu2+ and [Cd(sac)2(HydEt-en)2] has a single site. The principal values for theg-tensor and the hyperfine tensor for Cu2+ in the two compounds were obtained. The Cu2+ ion was found to be mostly in the 3dx 2y 2 orbital and the ground-state wavefunction of [Cd(sac)2(HydEten]2] was constructed.  相似文献   

6.
The adsorption of H2O on Al(111) has been studied by ESDIAD (electron stimulated desorption ion angular distributions), LEED (low energy electron diffraction), AES (Auger electron spectroscopy) and thermal desorption in the temperature range 80–700 K. At 80 K, H2O is adsorbed predominantly in molecular form, and the ESDIAD patterns indicate that bonding occurs through the O atom, with the molecular axis tilted away from the surface normal. Some of the H2O adsorbed at 80 K on clean Al(111) can be desorbed in molecular form, but a considerable fraction dissociates upon heating into OHads and hydrogen, which leaves the surface as H2. Following adsorption of H2O onto oxygen-precovered Al(111), additional OHads is formed upon heating (perhaps via a hydrogen abstraction reaction), and H2 desorbs at temperatures considerably higher than that seen for H2O on clean Al(111). The general behavior of H2O adsorption on clean and oxygen-precovered Al(111) (θO ? monolayer) is rather similar at low temperature, but much higher reactivity for dissociative adsorption of H2O to form OH adsis noted on the oxygen-dosed surface around room temperature.  相似文献   

7.
Metal–ligand bond enthalpy data can afford invaluable insights into important reaction patterns in organometallic chemistry and catalysis. In this paper, the Fe–O and Fe–S homolytic bond dissociation energies [ΔHhomo(Fe–O)'s and ΔHhomo(Fe–S)'s] of two series of para‐substituted phenoxydicarbonyl(η5‐cyclopentadienyl) iron [p‐G‐C6H4OFp ( 1 )] and (para‐substituted benzenethiolato)dicarbonyl(η5‐cyclopentadienyl) iron [p‐G‐C6H4SFp ( 2 )] were studied using Hartree–Fock and density functional theory (DFT) methods with large basis sets. In this study, Fp is (η5‐C5H5)Fe(CO)2, and G are NO2, CN, COMe, CO2Me, CF3, Br, Cl, F, H, Me, MeO, and NMe2. The results show that DFT methods can provide the best price/performance ratio and accurate predictions of ΔHhomo(Fe–O)'s and ΔHhomo(Fe–S)'s. The remote substituent effects on ΔHhomo(Fe–O)'s and ΔHhomo(Fe–S)'s [ΔΔHhomo(Fe–O)'s and ΔΔHhomo(Fe–S)'s] can also be satisfactorily predicted. The good correlations [r = 0.98 (g, 1), 0.98 (g, 2)] of ΔΔHhomo(Fe–O)'s and ΔΔHhomo(Fe–S)'s in series 1 and 2 with the substituent σp+ constants imply that the para‐substituent effects on ΔHhomo(Fe–O)'s and ΔHhomo(Fe–S)'s originate mainly from polar effects, but those on radical stability originate from both spin delocalization and polar effects. ΔΔHhomo(Fe–O)'s ( 1 ) and ΔΔHhomo(Fe–S)'s ( 2 ) conform to the captodative principle. Insight from this work may help the design of more effective catalytic processes. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

8.
The proton conductivity in H3OUO2AsO4·3H2O (HUAs) has been measured below the transition temperature at 299 K. A consistent picture of the elementary process taking place is developed from separate electrochemical, spectroscopic and calorimetric measurements. The conductivity is proposed to occur by the “vehicle mechanism” of proton transport, i.e. the cooperative motion of H3O+ and H2O. This mechanism is compared with the Grotthuss and the simple ion hopping mechanism.  相似文献   

9.
The IR spectra and polarization Raman spectra of Kal(SO4)2·12(H2O) and Kal(SO4)2·12[H2O)0.3(D2O)0.7] crystals at 93 K and room temperature have been obtained experimentally. The vibrational spectra of structural elements of potassium alum — the complexes [Al(H2O)6 3+ and [Al(D2O)6]3+ — have been calculated. The vibrational spectra have been interpreted based on the calculation and factor-group analysis data. The spectral data obtained point to the fact that, in the crystals considered, the sulfate ions are partially disordered and there exist two crystallographically different types of water molecules.  相似文献   

10.
EPR investigations using Cu2+ ion as a probe have been performed on supersaturated sucrose solution with percent concentration c = 66 as a function of temperature T, and at room temperature as a function of c. The motionally averaged spectrum of [Cu(H2O)6]2+ was used to monitor the changes in intermolecular interactions that occur as a function of [c, T]. A drastic increase in the line width, symptomatic of increase in the rotational correlation time of [Cu(H2O)6]2+, is observed between 293 and 288 K. The motionally averaged spectrum disappears below 281 K. The motionally averaged spectrum is also absent in the room temperature spectra of the solution with c= 85. Even in the [c, T] range where [Cu(H2O)6]2 is found to be nearly static, these molecules appear to have an orientational fluctuation manifesting in the m 1 dependence of the line width of the parallel component.  相似文献   

11.
Raman spectra of vajdakite, [(Mo6+O2)2(H2O)2As O5]·H2O, were studied and interpreted in terms of the structure of the mineral. The Raman spectra were compared with the published infrared spectrum of vajdakite. The presence of dimolybdenyl and diarsenite units and of hydrogen bonded water molecules was inferred from the Raman spectra which supported the known and published crystal structure of vajdakite. Mo O and O H···O bond lengths were calculated from the Raman spectra. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

12.
By means of Fourier transform microwave spectroscopy of a supersonic molecular beam, we have detected the more abundant rare isotopic species of monobridged Si(H)SiH, and determined of its geometrical structure by isotopic substitution. Unexpected abundance anomalies of the singly deuterated isotopic species with respect to normal and Si(D)SiD were observed with equal mixtures of SiH4 and SiD4. This nonstatistical distribution is consistent with Si(H)SiH formation which is dominated by the reaction of a bare Si atom or ion with either SiH2 or SiH4. Dedicated laboratory searches were also undertaken for the remaining singlet isomer, H2SiSi, so far without success. From the absence of lines at the expected frequencies we conclude that H2SiSi is at least 50 times less abundant than Si(H)SiH. Because it is only slightly less stable than Si(H)SiH and because the barrier to isomerization is calculated to be quite low, H2SiSi may rearrange on the timescale of the supersonic expansion to Si(H)SiH.  相似文献   

13.
The Raman scattering cross section (RSCS) is an important parameter in the applications of Raman spectroscopy to make quantitative analysis. To date, the dependence of the RSCS on concentration has remained unclear. Nitrate aerosols can easily achieve a supersaturated state, which provides a way to obtain the RSCS especially under this state. In this study, Raman spectra of NaNO3 and Mg(NO3)2 solutions are obtained with molar water‐to‐solute ratios (WSRs) ranging from 84.2 to 2.30 and 93.8 to 7.32, respectively. With decreasing WSR, a shift to higher wavenumbers of the symmetric stretching band of nitrate ion, i.e. ν1(NO3), is observed, indicating the formation of various ion pairs. Meanwhile, the area ratio between the strongly and weakly hydrogen‐bonded components of water O H stretching envelope, i.e. ν(H2O), reduces as the WSR decreases, implying the transformation of water molecules from strong hydrogen‐bonding structures to the weak ones. However, a good linear relationship is revealed between the integrated intensity ratio of the ν(H2O) band to ν1(NO3) band and WSR. The results suggest that the RSCSs of NO3 and H2O are insensitive to the structures of both ion pairs and hydrogen‐bonding structures. This observation points to the possibility of conducting quantitative analysis through the area ratio of the ν(H2O) band to the ν1(NO3) band with Raman spectra without considering the formation of ion pairs and the variation of the hydrogen‐bonding structure. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

14.
State-resolved measurements are presented for vibrational excitation of H2, N2, O2 and CO2 by H? impact in the collision energy rangeE cm=20–180 eV and for scattering in the forward direction (0±0.5°). The data obtained from the measurements are the relative intensities and differential cross sections for vibrational excitation up toν′=4, the transition probabilitiesP 0→ν′ and the vibrational energy transferΔE vib. For the systems H?-H2, N2, O2 the vibrational inelasticity increases in the order H2-N2-O2. The mechanism for vibrational excitation in these systems is due to transient charge transfer from the H? ion into antibonding orbitals of the target molecule which provides a bond stretching force during the collision. For H2 and N2, the results are compared with corresponding measurements for H+ scattering where the interaction mechanism is quite different. In the case of CO2, vibrational excitation in forward scattering is caused primarily by the long-range dipole interaction. The spectra are very similar for H?-CO2 and H+-CO2. Finer details are attributed to the influence of transient charge transfer and valence interactions.  相似文献   

15.

The crystal structure of di-(L-serine) phosphate monohydrate [C3O3NH7]2H3PO4H2O is determined by single-crystal x-ray diffraction. The intensities of x-ray reflections are measured at temperatures of 295 and 203 K. The crystal structure is refined using two sets of intensities. It is established that, in the structure, symmetrically nonequivalent molecules of L-serine occur in two forms, namely, the monoprotonated positively charged molecule CH2(OH)CH(NH3)+COOH and the zwitterion CH2(OH)CH(NH3)+COO?, which are linked with each other and with the H2PO ?4 ion through a hydrogen-bond system involving water molecules.

  相似文献   

16.
The paper deals with a study of the proton nuclear magnetic resonance (NMR) of crystallization water in isomorphous monohydrates MgSO4. 1 H2O and FeSO4. 1 H2O in the temperature range 123–313 K. The NMR second moment for diamagnetic MgSO4. 1 H2O shows only a weak dependence on temperature but the one for paramagnetic FeSO4. 1 H2O is rather strong. Results obtained for FeSO4. 1 H2O are in a good agreement with the Kroon's theory of NMR in paramagnetics. The Curie-Weiss constant and the effective magnetic moment of Fe2+ ions in FeSO4. 1 H2O are derived from the temperature dependence of NMR second moment. The motion of molecules of crystallization water in these hydrates is discussed on the basis of temperature dependences of the width and second moment of NMR spectra.  相似文献   

17.
Metal organic frameworks (MOFs) have been well-known and extensively researched due to the high storage /good selectivity for gas molecules. Herein, the structures and electron paramagnetic resonance (EPR) spectra for dicopper paddle wheel MOF compound (Cu22-O2CCH3)4 with various gas molecule are theoretically investigated by density functional theory (DFT) calculations. The adsorption energies and isotherms (including pure gas molecules and the mixed ones) are calculated for the gas molecules interacting with the unsaturated Cu22-O2CCH3)4. Both quantities exhibit the roughly consistent orders (e.g. H2S?>?NH3?>?CO2?>?CO?>?H2O?>?N2?>?NO?>?H2 for isotherms and H2S?>?NH3?>?N2?>?CO2?>?NO?>?H2O?>?H2?>?CO for adsorption energies), possibly suggesting that this material may act as a potential adsorbent of these gas molecules. The catalytic property of Cu22-O2CCH3)4 for oxidation of CO and NO into non-toxic molecules and splitting of H2O into H2 and O2 in the solvent condition are uniformly discussed. Simulation of Grand Canonical Monte Carlo (GCMC) in MS 8.0 and calculations in Langmuir model reveal that Cu22-O2CCH3)4 has good selectivity for CH4 in natural gas (CH4/CO2/N2) and SO2 in fog (SO2/NO/NO2/H2O/O2), which would exhibit potential environmentally friendly applications.  相似文献   

18.
R Ratheesh  G Suresh  V U Nayar 《Pramana》1995,44(5):461-470
The infrared and Raman spectra of NaNi2OH(H2O)(MoO4)2 and NaZn2OH (H2O)(MoO4)2 and their partially deuterated analogues are recorded and analysed on the basis of vibrations of MoO 4 2− tetrahedra and H2O molecules. The MoO 4 2− groups are found to be more distorted in NaNi2OH(H2O)(MoO4)2 than in the other compound. Bands indicating the presence of H3O+ ions are not observed in NaZn2OH(H2O)(MoO4)2 ruling out the possibility of the formulation of NaZn2OHO(MoO3OH)2. Hydrogen bonds of medium strength are present in both the compounds.  相似文献   

19.
The addition reaction of CH2OO?+?H2S → HSCH2OOH without and with catalyst X (X?=?H2O and (H2O)2) has been investigated by CCSD(T)-F12a/VTZ-F12//B3LYP/aug-cc-pVTZ method and canonical variational transition state theory with small curvature tunneling correction. When H2O was introduced in the CH2OO?+?H2S reaction, it not only acts as a catalyst for producing HSCH2OOH, but also plays as a reactant to forming HOCH2OOH. The formation channel of HSCH2OOH is more important than the formation channel of HOCH2OOH with its calculated rate constant larger by 11.0–43.2 times within the temperature 280–320?K. Then, (H2O)2 catalysed CH2OO?+?H2S → HSCH2OOH reaction has been taken into account with its rate lower 1.9–4.2 times than the reaction of CH2OO?+?H2S → HSCH2OOH with water. Also, CH2OO?+?H2S with H2O cannot compete with the CH2OO?+?H2S reaction without water. This is different from CH2OO?+?(H2O)2 reaction, which is about 4 orders of magnitude larger than the rate constant for CH2OO?+?H2O reaction. Such discrepancy is possible because C(CH2OO)···O(H2O) interaction has been enhanced more obviously by H2O as compared to that of C(CH2OO)···O(H2S) interaction.  相似文献   

20.
The single‐crystal Raman spectra of minerals brandholzite and bottinoite, formula M[Sb(OH)6]2•6H2O, where M is Mg+2 and Ni+2, respectively, and the non‐aligned Raman spectrum of mopungite, formula Na[Sb(OH)6], are presented for the first time. The mixed metal minerals comprise alternating layers of [Sb(OH)6]−1 octahedra and mixed [M(H2O)6]+2/[Sb(OH)6]−1 octahedra. Mopungite comprises hydrogen‐bonded layers of [Sb(OH)6]−1 octahedra linked within the layer by Na+ ions. The spectra of the three minerals were dominated by the Sb O symmetric stretch of the [Sb(OH)6]−1 octahedron, which occurs at approximately 620 cm−1. The Raman spectrum of mopungite showed many similarities to spectra of the di‐octahedral minerals, supporting the view that the Sb octahedra give rise to most of the Raman bands observed, particularly below 1200 cm−1. Assignments have been proposed on the basis of the spectral comparison between the minerals, prior literature and density functional theory (DFT) calculations of the vibrational spectra of the free [Sb(OH)6]−1 and [M(H2O)6]+2 octahedra by a model chemistry of B3LYP/6‐31G(d) and lanl2dz for the Sb atom. The single‐crystal spectra showed good mode separation, allowing most of the bands to be assigned to the symmetry species A or E. Copyright © 2010 John Wiley & Sons, Ltd.  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