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 21 毫秒
1.
The proton spin-lattice relaxation time, T1, is measured as a function of temperature in α -(COOH)2·-2H2O, K2HgCl4· H2O and LiCHO·H2O. The relaxation is caused by 180° flips of the water molecules about their 2-fold axes and good agreement is obtained between calculated and observed values of T1. Empiricly the flip rate follows a classical Arrhenius equation: P· exp (? ΔH(RT)). A literature survey of values of P and ΔH obtained from similar investigations on other hydrates is given. The survey shows that the preexponential factor, P, is a function of the activation enthalpy, ΔH. P increases from 1012 to 1017 Hz when ΔH changes from 2 to 17 kcalmole. Using a dynamical rate theory as formulated by Feit, we find the flip rate is given by: K2· √(ΔH)· exp (K1ΔH)· exp (?ΔH(RT>)). This expression can be fitted to the observed data using K1 = 0.69 molekcal and K2 = 2 × 1011 Hz · (kcalmole)?12. Thus both the frequency factor, K2√ (ΔH), and the entropic factor, exp (K1ΔH), have been obtained for flipping water molecules in hydrates. The values of K1 and K2 are shown to be physically reasonable.  相似文献   

2.
The X-ray structure (293 K) of UO2(H2PO4)2·3H2O has been refined (R = 0.062): Mr = 518g, space group: P21/c (Z = 4); a = 10.816(1) A?, b = 13.896(2) A?, c = 7.481(1) A?, β = 105.65(1)°, V = 1082.7(2) A?3; Dc = 3.17 Mg m?3. The structure consists of infinite chains along the (101) axis with U atoms bridged by two H2PO4 groups. The U atom is surrounded by a pentagonal bipyramid of oxygen atoms, one of them being an equatorial water molecule. The cohesion between the chains is ensured by hydrogen bonds involving the two last water molecules. An assignment of IR and Raman bands with isotopic substitution spectra is proposed. A phase transition at 128 K was made evident by DSC and spectroscopy. The room-temperature phase is characterized by a high disorder of the OH bond orientation while in the low-temperature phase H2O and POH species appear well oriented. The conductivity seems to occur by proton transfer and protonic-species rotation at the POH-water molecular interface between the chains. ac conductivity has been determined by means of the complex-impedance method (σRT ~ (3?12) × 10?5 Ω?1cm?1; E ~ 0.20 eV).  相似文献   

3.
Previous studies have shown that NiZrF6.6H2O can be described by an axial spin Hamiltonian with parameters g = 2.33 and Dk = ?3.14K, and exhibits ferromagnetic ordering at Tc = 164mK. Our room temperature X-ray measurements confirm that NiZrF6·.6H2O is isomorphous with NiSnCl6·.6H2O and give lattice parameters of a = 6.55 A? and α = 96°09′. Proton NMR measurements show that there exist important molecular motions at room temperature and that the space group remains R3 down to liquid helium temperatures. The positions of all twelve protons have been determined and are found to lie on a sphere of radius 2.9 Å centered on the nickel ion.  相似文献   

4.
The low temperature infrared absorption and room temperature laser Raman spectra of polycrystalline Ba(NO2)2 · H2O and its deuterated analogue are studied. The strongly bonded H-atom of the water molecule makes a highly bent H-bond while the weakly bonded H-atom exhibits a slightly bent (or possible bifurcated) bond consistent with recent structural data; the H-bond enthalpies are estimated to be 3.3 and 2.1 KCalmole respectively. The wagging, twisting and rocking modes of the water molecule have been assigned using several well known criteria. The force constants of these modes have also been computed. The librational splittings observed at low temperature are consistent with polarization data, and are being attributed to infrared active factor group components.  相似文献   

5.
This susceptibility of the quadratic, S = 12, XY model as determined by the high-temperature series expansion is presented and compared with experimental data on CoCl2.6H2O and CoBr2.6H2O. For the specific heat of CoCl2.6H2O a similar comparison is made. For kT/|J| > 1.5 a good agreement between theory and experiment is found, yielding J/k = ?2.05 ± 0.1 and ?2.45 ± 0.1 K for the intralayer exchange of the Cl and the Br salt, respectively. These values compare favourably with those available from other sources.  相似文献   

6.
The oxidation of methanol was studied on a Ag(110) single-crystal by temperature programmed reaction spectroscopy. The Ag(110) surface was preoxidized with oxygen-18, and deuterated methanol, CH3OD, was used to distinguish the hydroxyl hydrogen from the methyl hydrogens. Very little methanol chemisorbed on the oxygen-free Ag(110) surface, and the ability of the silver surface to dissociatively chemisorb methanol was greatly enhanced by surface oxygen. CH3OD was selectively oxidized upon adsorption at 180 K to adsorbed CH3O and D218O, and at high coverages the D218O was displaced from the Ag(110) surface. The methoxide species was the most abundant surface intermediate and decomposed via reaction channels at 250, 300 and 340 K to H2CO and hydrogen. Adsorbed H2CO also reacted with adsorbed CH3O to form H2COOCH3which subsequently yielded HCOOCH3 and hydrogen. The first-order rate constant for the dehydrogenation of D2COOCH3 to DCOOCH3 and deuterium was found to be (2.4 ± 2.0) × 1011 exp(?14.0 ± 0.5 kcalmole · RT)sec?1. This reaction is analogous to alkoxide transfer from metal alkoxides to aldehydes in the liquid phase. Excess surface oxygen atoms on the silver substrate resulted in the further oxidation of adsorbed H2CO to carbon dioxide and water. The oxidation of methanol on Ag(110) is compared to the previous study on Cu(110).  相似文献   

7.
The decomposition of HCOOD was studied on Ni(100). Low temperature adsorption of HCOOD resulted in the desorption of D2O, CO2, CO, and H2. The D2O was evolved below room temperature. CO2 and H2 were evolved in coincident peaks at a temperature above that at which h2 desorbed following H2 adsorption and well above that for CO2 desorption from CO2 adsorption; CO desorbed primarily in a desorption limited step. The decomposition of formic acid on the clean surface was found to yield equal amounts of H2, CO, and CO2 within experimental error. The kinetics and mechanism of the decomposition of formic acid on Ni (110) and Ni(100) single crystal surfaces were compared. The reaction proceeded by the dehydration of formic acid to formic anhydride on both surfaces. The anhydride intermediate condensed into islands due to attractive dipole-dipole interactions. Within the islands the rate of the decomposition reaction to form CO2 was given by:
Rate = 6 × 1015 exp{?[25,500 + ω(ccsat)]/RT} × c
, where c is the local surface concentration, csat is the saturation coverage for the particular crystal plane, and ω is the interaction potential. The interaction potential was determined to be 2.7 kcal/mole on Ni(110) and 1.4 kcal/mole on Ni(100); the difference observed was due to structural differences of the surfaces relating to the alignment of the dipole moments within the islands. These attractive interactions resulted in an autocatalytic reaction on Ni(110), whereas the interaction was not strong enough on Ni(100) to sustain the autocatalytic behavior. Formic acid decomposition oxidized the Ni(100) surface resulting in the formation of a stable surface oxide. The buildup of the oxide resulted in a change in the selectivity reducing the amount of CO formed. This trend indicated that on the oxide surface the decomposition proceeded via a formate intermediate as on Ni(110) O.  相似文献   

8.
Weighted average cross sections for quenching of the K(42P)-doublet by N2, H2, O2 and H2O, measured in flames, show no significant temperature dependence in the range from 1500 to 2500K. Doublet mixing cross sections for K(42P32?42P12) transitions were measured at 1720K for N2, O2, H2O. The ratios of both mixing cross sections were measured independently and were found to agree with the detailed balance condition within 2 per cent. It is shown that an ionic intermediate-state model cannot explain the large magnitude of N2? mixing cross sections.  相似文献   

9.
Adsorption of CO on Ni(111) surfaces was studied by means of LEED, UPS and thermal desorption spectroscopy. On an initially clean surface adsorbed CO forms a √3 × √3R30° structure at θ = 0.33 whose unit cell is continuously compressed with increasing coverage leading to a c4 × 2-structure at θ = 0.5. Beyond this coverage a more weakly bound phase characterized by a √72 × √72R19° LEED pattern is formed which is interpreted with a hexagonal close-packed arrangement (θ = 0.57) where all CO molecules are either in “bridge” or in single-site positions with a mutual distance of 3.3 Å. If CO is adsorbed on a surface precovered by oxygen (exhibiting an O 2 × 2 structure) a partially disordered coadsorbate 2 × 2 structure with θo = θco = 0.25 is formed where the CO adsorption energy is lowered by about 4 kcal/mole due to repulsive interactions. In this case the photoemission spectrum exhibits not a simple superposition of the features arising from the single-component adsorbates (i.e. maxima at 5.5 eV below the Fermi level with Oad, and at 7.8 (5σ + 1π) and 10.6 eV (4σ) with COad, respectively), but the peak derived from the CO 4σ level is shifted by about 0.3 eV towards higher ionization energies.  相似文献   

10.
A study of the adsorption/desorption behavior of CO, H2O, CO2 and H2 on Ni(110)(4 × 5)-C and Ni(110)-graphite was made in order to assess the importance of desorption as a rate-limiting step for the decomposition of formic acid and to identify available reaction channels for the decomposition. The carbide surface adsorbed CO and H2O in amounts comparable to the clean surface, whereas this surface, unlike clean Ni(110), did not appreciably adsorb H2. The binding energy of CO on the carbide was coverage sensitive, decreasing from 21 to 12 kcalmol as the CO coverage approached 1.1 × 1015 molecules cm?2 at 200K. The initial sticking probability and maximum coverage of CO on the carbide surface were close to that observed for clean Ni(110). The amount of H2, CO, CO2 and H2O adsorbed on the graphitized surface was insignificant relative to the clean surface. The kinetics of adsorption/desorption of the states observed are discussed.  相似文献   

11.
The Curie-Weiss constant θ of the quasi b.c.c., S = 12, Heisenberg ferromagnets, (NH4)2 Cu Br4 · 2H2 O and Rb2 Cu Br4 · 2H2 O, have been derived from measurements of the susceptibility in the paramagnetic region (4.2 < T < 100 K) and of the high-field magnetization curves at T = 4.2 K (TTc ? 2.3). The Tcθ values obtained point to the presence of further neighbour interactions considerably stronger than previously assumed. The effective number of equivalently interacting magnetic neighbours is estimated to be at least seventeen.  相似文献   

12.
The Tian Calvet microcalorimetric method has been improved in order to determine ΔHH(D), the partial molar enthalpy of mixing of hydrogen (deuterium) in the Ti-H2(D2) solid systems for compositions 0 < (H/Ti) < 1.85, at 713 K and in the α TiOy + H2 solid solutions (y = (O/Ti)) at 745 K. The combined calorimetric and equilibrium method allows a precise evaluation of the partial molar entropies. The results of this study differ substantially from earlier published data.  相似文献   

13.
Interference between one- and two-photon processes for e+e? annihilation into hadrons in a two-jet parton model leads to a charge asymmetry of detected final state hadrons along the directions of the incident leptons. The asymmetry near the lepton axis grows as 2ln212θ - 4ln12θ ln s/ΔE2, so that in spite of the O(α) suppression relative to the Born cross section, the asymmetry can amount to a 2% effect at θ = 2° for π± or K± inclusive measurements in a typical experiment. The precise size of the asymmetry depends on the parton charges. Our results are only meant to apply in the region ω ? 0.5,where the parton model appears to be relevant at present experimental energies.  相似文献   

14.
We have measured the isothermal magnetization M of antiferromagnetic K2[FeCl5(H2O)] as a function of applied magnetic field H, for fields up to 80 kOe and for several hydrostatic pressures up to P ~ 10 kbar. We obtained the P =1 atm exchange parameters of this material for an orthorhombic model. The spin-flop boundary of K2[FeCl5(H2O)] was studied for several pressures. For the spin-flop fields HSF, a relation of the form HSF(T) = HSF(0) (1 + BT32 + CT2 + DT 52) adequately describes the experimental points up to relatively high temperatures, for all pressures. The isotherms HSF vs. P show an abrupt change of slope at P ? 1 kbar, possibly due to a pressure-induced exchange of the intermediate and hard magnetic axes.  相似文献   

15.
Experimental data are presented for the angular dependence of the relative flux, the mean energy and the speed ratio of deuterium molecules desorbing from a Ni(111) crystal surface at a surface temperature of Ts = 1143 K and at sulphur coverages ranging between 30% and less than 2% of a monolayer.The angular flux distribution is sharply peaked in the forward direction (cosdθwith 3 ? d ? 5) and the mean energy 〈E〉 of the desorbate depends strongly on the desorption angle θ. For normal desorption (θ = 0°) 〈E〉2k is about 700 K higher than Ts and for glancing angles (θ = 80°) it decreases to about 400 K below Ts The results obtained on sulphur free and sulphur covered Ni(111) surfaces are compared with our former data on polycrystalline nickel. The main differences in the kinetic features can be ascribed to the surface roughness. Accordingly, the angular distributions of flux, mean energy, and speed ratio, which deviate strongly from the Knudson and Maxwellian law, do not seem to depend considerably on sulphur coverage and surface structure. A qualitative explanation for these deviations is presented using the principle of detailed balancing.  相似文献   

16.
SiO films obtained by sputtering in an ArO mixture with an oxygen partial pressure less than 3% are similar to a-Si films: the resistivity is proportional to exp (T0/T)14 and T0 increases with oxygen content and decreases with increasing Au concentration (? 3.7 at.%). On the other hand, above an oxygen partial pressure of 5% one obtains insulating amorphous SiO2 films. Conductivity appears in such films for Au? 13 at.% (? the percolation threshold) and then the resistivity is proportional to exp (T0/T)12. The same behavior is observed when oxygen is replaced by hydrogen.  相似文献   

17.
The diffusion of hydrogen in palladium (HPd = 0.73) has been investigated from 170 to 300K by measurements of the proton spin-lattice relaxation time in the rotating frame, T1?. In contrast to previous T1 measurements, a single activation energy of 0.225 eV is obtained, in agreement with the high-temperature T1 data and with internal friction experiments at about 120K.  相似文献   

18.
In search for structural contributions to the low temperature anomaly we report high resolution resistance and magnetoresistance measurements (0.02 K ? T ? 20 K) of amorphous splats of Gd67Co33 and Pd80Si20. For both alloys, the resistivity ?(H = 0, T) has a minimum at T ~ 10 K and increases with decreasing T. The ferromagnetic Gd67Co33 shows a strong negative field dependence of Δ??(0), saturating at H ~ 2T for T = 4.2 K but no measurable change in ???T below 10 K is observed.The diamagnetic Pd80Si20 exhibits a positive field dependent magnetoresistance [Δ??(0)](H) at low temperatures. Additionally, a field dependent part in ???T is found which is probably due to paramagnetic impurities (~ 1 ppm Fe). However, there is also a field independent contribution in the amorphous state of Pd80Si20, which vanishes after crystallization. We attribute this to non-magnetic scattering induced by the disordered structure.  相似文献   

19.
A new modification of molecular beam relaxation spectrometry (MBRS) is described: the temperature jump method for studying catalytic surface processes on metal foils. The temperature of the catalyst foil is maintained by direct ohmic heating; a constant particle beam is directed towards the catalyst surface. A jump of the surface temperature caused by a high current pulse generates a response of the fluxes of desorption. The decay of the desorption intensity after the temperature jump contains the relaxation times of the elementary steps involved. The mathematical treatments of unimolecular and bimolecular surface reactions, of sequences of two and three unimolecular steps and of a sequential reaction accompanied by the redesorption of the reactant are given. The application of the new method is shown by a study of the catalytic decomposition of CH3)OH on polycrystalline Ni: CO and H2 are the sole reaction products. The limit of the catalytic activity — apart from the low sticking probability of the reactant — must be seen in the abstraction of the first methyl hydrogen from the transient methoxy species. In the temperature range between 320 and 550 K the reaction mechanism can be described as follows:
Rate constants in dependence from surface temperature T are: k1 = 4.2 × 104 exp(?22.4RTkJmol) s?1; k3 = 2.4 × 109 exp(?75RTkJmol) s?1; k4 = 1.2 × 1013 exp(?104RTkJmol) s?1; η = 0.2. Typical surface residence times of the intermediates are: 110 ? τ1 ? 15 ms at 320 ? T ? 450 K; 210 ? τ3 ? 6 ms at 450 ? T ? 550 K; 98 ? τ4 ? 6 ms at 450 ? T ? 500 K.  相似文献   

20.
Single crystal neutron diffraction measurements of K2Pt(CN)4Br0.3 · 3H2O (KCP) above room temperature, to the point of irreversible crystal breakdown (318–323°K), and at liquid nitrogen temperature (77°K), give no indication of a crystallographic phase change. Full three-dimensional data collected at 77°K (a = 9.848(5) A?, c = 5.714(3) A?, and space group P4mm) indicate that the structure is essentially unchanged from that at room temperature except for increased hydrogen bonding association between H2O(2) and Br(1). The possible relationship between the hydrogen bonding changes and Br(1) site ordering is 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