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 437 毫秒
1.
The adsorption of chlorine on the Ag(111) surface has been studied using LEED, Auger and temperature programmed desorption. Chlorine adsorbs dissociately with an initial sticking probability of ~ 0.4, and a precursor state is implicated in the chemisorption process. The chlorine appears to form a close-packed monolayer with the same packing density as in AgCl(111), and is epitaxially related to the substrate mesh. Chlorine continues to adsorb above a monolayer in coverage, though the sticking probability drops precipitately, being ~ 0.01 after the adsorption of 5 monolayers at 300 K. There is little increase in the chlorine Auger signal above one monolayer coverage at 300 K, but when adsorption is carried out at 240 K the chlorine signal is more than doubled. This is interpreted as being due to the formation of a layer structure of alternate Cl and Ag layers at the lower temperature, while adsorption at 300 K results in dissolution of subsurface Cl into the bulk of the crystal. Upon heating, the low temperature layer structure is destroyed, the chlorine signal diminishes to a limiting value at 450 K equivalent to the value for one adsorbed monolayer — apparently due to the dissolution of the near surface Cl layers into the bulk. However, the chlorine re-emerges at the surface at ~ 600 K, probably due to an exothermic heat of solution of Cl in the silver lattice. Desorption from the multilayers peaks at 670 K and both AgCl and Ag are desorbed coincidently with kinetics identical to those for the sublimation of bulk AgCl (ΔH = 235 kJ mol?1, ΔS = 90 JK?1 mol?1). After the multilayers have desorbed, the final Cl layer desorbs in a higher temperature peak ( ~ 760 K) as AgCl (no silver desorption) which shows complex desorption kinetics indicative of the strong influence of a precursor state in the desorption process.  相似文献   

2.
Reflection-absorption infrared spectroscopy has been combined with thermal desorption and surface coverage measurements to study nitrogen adsorption on a {111}-oriented platinum ribbon under ultrahigh vacuum conditions. Desorption spectra show a single peak (at 180 K) after adsorption at 120 K, giving a coverage-independent activation energy for desorption'of ~40 kJmol?1. The initial sticking probability at this temperature is 0.15, and the maximum uptake was ~1.1 × 1014 molecule cm?2. The adsorbed nitrogen was readily displaced by CO, h2 and O2. An infrared absorption band was observed with a peak located at 2238 ± 1 cm?1, and a halfwidth of 9 cm?1, with a molecular intensity comparable to that reported for CO on Pt{111}. The results are compared with data for chemisorption on other group VIII metals. An earlier assignment of infrared active nitrogen to B5 sites on these metals is brought into question by the present results.  相似文献   

3.
The adsorption of NO and its reaction with H2 over polycrystalline Pd were investigated using flash desorption technique and ultraviolet photoelectron spectroscopy under 10?5 Pa pressure range of reactants and surface temperatures between 300 and 900 K. NO was adsorbed dissociatively onto the Pd surface above 500 K, and the heat of dissociative adsorption was ca. 126 kJ/mol. Atomic nitrogen was observed to accumulate on the Pd surface during the NO-H2 reaction, whose desorption rate exhibited second order kinetics and is expressed as follows: Vd = 10?9.8 ± 0.3exp(?67(kJ/mol)/RT) (cm2/atom·s). Hydrogenation of the adsorbed nitrogen proceeded rapidly at 485 K. It was confirmed from these results that formation of N2 and NH3 in the NO-H2 reaction proceeds through this atomically adsorbed nitrogen. Pd-N bond energy and enthalpies of some intermediate states of the NO-H2 reaction were estimated.  相似文献   

4.
The adsorption, desorption, surface structural chemistry, and electron impact properties of CO on Rh(110) have been studied by LEED, Auger spectroscopy, thermal desorption, and surface potential measurements. At 300 K, CO adsorbs into a single chemisorbed state whose desorption energy (Ed) is ~130kJmol-1. The initial sticking probability is unity, and at saturation coverage a (2 × 1)plgl ordered phase reaches its maximum degree of perfection, thus demonstrating that this CO structure is common to the (110) faces of all the cubic platinum group metals. The saturated adlayer corresponds to θ = 1 and shows a surface potential of Δ? = +0.97 V. Under electron impact, desorption and dissociation of CO occur with about equal probability, the relevant cross sections being ~10-22 m2 in each case. Slow thermal dissociation of CO occurs at high temperature and pressure, leaving a deposit of C and O atoms on the surface. The thermal, electron impact, and Δ? properties of Rh(110)CO resemble those of Ni(110)CO rather closely, and are very different from those of Pt(110)CO. Surface carbon is shown to inhibit CO chemisorption, whereas surface oxygen appears to lead to the formation of a new more tightly bound form of CO with a considerably enhanced desorption energy (Ed ~ 183 kJmol-1). Similar oxygen-induced high temperature CO states have been reported recently on Co(0001) and Ru(101̄1).  相似文献   

5.
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   

6.
Transitions were observed by heat capacity measurements at 74.6 K, 195.2 K, and 303 K. They are a soft mode transition (ΔHt = 30 J mol?1, ΔSt = 0.42 J K?1 mol?1), a first-order commensurate-incommensurate transition (ΔHt = 6.2 J mol?1ΔSt = 0.032 J K?1 mol?1), and a second-order incommensurate-normal transition, respectively.  相似文献   

7.
Molecular sulphur undergoes rapid dissociative chemisorption on Ag(111) with an essentially constant sticking probability of unity up to the completion of the first layer of S atoms. At this stage a (√39 R 16.1° × √39 R? 16.1°) structure is formed in which the S atom arrangement and spacing is similar to that in the (100) plane of γ-Ag2S (the high temperature form of silver sulphide). Further dosing with S2 leads to continued rapid uptake of sulphur and the appearance of a (√7 × √7) R 10.9° structure, the Auger, Δφ and thermal desorption data all indicate that fast formation of Ag2S now occurs. Very well-ordered growth of γ-Ag2S(111) is now observed, and low-temperature S2 desorption spectra appear which show that the activation energy for S2 desorption is ~175 kJ mol?1 ; this value is in excellent agreement with that observed for the enthalpy of decomposition of bulk Ag2S (2 Ag2S(s) → 4 Ag(s) + S2(g), ΔH = +179 kJmol?1). All the properties of the Ag(111)-S system imply that the material characterised by the √39 structure (i.e. the first adsorbed layer of S) is very different from bulk Ag2S. This is discussed and compared with the results of other studies on metal-sulphur systems.  相似文献   

8.
The heat capacity of the layer compounds tetrachlorobis (n-propylammonium) manganese II and tetrachlorobis (n-propylammonium) cadmium II, (CH3CH2CH2NH3)2MnCl4 and (CH3CH2CH2NH3)2CdCl4 respectively, has been measured over the temperature range 10 K ?T ? 300 K.Two known structural phase transitions were observed for the Mn compound in this temperature region: at T = 112.8 ± 0.1 K (ΔHt= 586 ± 2 J mol?1; ΔSt = 5.47 ± 0.02 J K?1mol?1) and at T =164.3 ± (ΔHt = 496 ± 7 J mol?1; ΔSt =3.29 ± 0.05 J K?1mol?1). The lower transition is known to be from a monoclinic structure to a tetragonal structure, while the upper is from the tetragonal phase to an orthorhombic one. From comparison with the results for the corresponding methyl Mn compound it is deduced that the lower transition primarily involves changes in H-bonding while the upper transition involves motion in the propyl chain.A new structural phase transition was observed in the Cd compound at T= 105.5 ± 0.1 K (ΔHt= 1472.3 ± 0.1 J mol?1; ΔSt = 13.956 ± 0.001 J K?1mol?1), in addition to two transitions that have been observed previously by other techniques. The higher of these transitions(T = 178.7 ± 0.3 K; ΔHt = 982 ± 4 J mol?1 ΔSt = 6.16 ± 0.02 J K? mol?1) is known to be between two orthorhombic structures, while the structural changes at the lower transition (T= 156.8 ± 0.2 K; ΔHt = 598 ± 5 J mol?1, ΔSt = 3.85 ± 0.03 J K?1 mol?1) and at the new transition are not known. It is proposed that these two transitions correspond respectively to the tetragonal to orthorhombic and monoclinic to tetragonal transitions in the propyl Mn compounds.In addition to the structural phase transitions (CH3CH2CH2NH3)2MnCl4 magnetically orders at t? 130 K. The magnetic contribution to the heat capacity is deduced from the heat capacity of the corresponding diamagnetic Cd compound and is of the form expected for a quasi 2-dimensional Heisenberg antiferromagnet.  相似文献   

9.
The methanation activity of W(110) was measured over a range of reactant partial pressures and temperatures (PH2 = 1–1000 Torr, PCO = 0.1–10 Torr, T = 475–820 K). Plotting the results in an Arrhenius fashion yielded a lower apparent activation energy (Ea = 56 kJmol?1) than previously determined for Ni(100) (Ea = 103 kJmol?1) with an activity surpassing that of Ni at lower temperatures. The H2 pressure dependence of the methanation activity was found to be much stronger for W(110) than for Ni(100), the surface becoming increasingly inactive at the lowest H2 pressures investigated. Auger electron spectroscopy revealed the active catalytic surface to be carbidic in nature.  相似文献   

10.
The adsorption, desorption, and surface structural properties of Na and NO on Ag(111), together with their coadsorption and surface reactivity, have been studied by LEED, Auger spectroscopy, and thermal desorption. On the clean surface, non-dissociative adsorption of NO into the a-state occurs at 300 K with an initial sticking probability of ~0.1, saturation occurring at a coverage of ~120. Desorption occurs reversibly without decomposition and is characterised by a desorption energy of Ed ~ 103 kJ mol?1. In the coverage regime 0 < θNa < 1, sodium adsorbs in registry with the Ag surface mesh and the desorption spectra show a single peak corresponding to Ed ~ 228 kJ mol?1. For multilayer coverages (1 < θ Na < 5) a new low temperature peak appears in the desorption spectra with Ed ~ 187 kJ mol?1. This is identified with Na desorption from an essentially Na surface, and the desorption energy indicates that Na atoms beyond the first chemisorbed layer are significantly influenced by the presence of the Ag substrate. The LEED results show that Na multilayers grow as a (√7 × √7) R19.2° overlayer, and are interpreted in a way which is consistent with the above conclusion. Coadsorption of Na and NO leads to the appearance of a more strongly bound and reactive chemisorbed state of NO (β-NO) with Ed ~ 121 kJ mol?1. β-NO appears to undego surface dissociation to yield adsorbed O and N atoms whose subsequent reactions lead to the formation of N2, N2O, and O2 as gaseous products. The reactive behaviour of the system is complicated by the effects of Na and O diffusion into the bulk of the specimen, but certain invariant features permit us to postulate an overall reaction mechanism, and the results obtained here are compared with other relevant work.  相似文献   

11.
The interaction of oxygen with a Pt(110) crystal surface has been investigated by thermal desorption mass spectroscopy, LEED and AES. Adsorption at room temperature produces a β-state which desorbs at ~800 K. Complete isotopic mixing occurs in desorption from this state and it populates with a sticking probability which varies as (1 ? θ)2, both observations consistent with dissociative adsorption. The desorption is second order at low coverage but becomes first order at high coverage. The saturationcoverage is 3.5 × 1014 mol cm?2. The spectra have been computer analysed to determine the fraction desorbing by first (β1) and second (β2) order kinetics as a function of total fractional coverage θ using this fraction as the only adjustable parameter. The β1 desorption commences at θ ~ 0.25 and β1 and β2 contribute equally to the desorption at saturation. The kinetic parameters for β1 desorption were calculated from the variation of peak temperature with heating rate as ν1 = 1.7 × 109 s?1 and E1 = 32 kcal mole?1 whereas two different methods of analysis gave consistent parameters ν2 = 6.5 × 10?7 cm2 mol?1 s?1 and E2 = 29 and 30 kcal mole?1 for β2 desorption. The kinetics of desorptior are discussed in terms of the statistics for occupation of near neighbour sites. While many fea tures of the results are consistent with this picture, it is concluded that simple models considering either completely mobile or immobile adlayers with either strong or zero adatom repulsion are not completely satisfactory. The thermal desorption surface coverage has been correlated with the AES measurements and it has been possible to use the AES data for PtO as an internal standard for calibration of the AES oxygen coverage determination. At low temperature (170 K) oxygen populates an additional molecular α-state. Adsorption into the α- and β-states is competitive for the same sites and pre-saturation of the β-state at 300 K excludes the α-state. This, together with the AES observation that the adsorption is enhanced and faster at 450 than 325 K suggests a low activation energy for adsorption into the β-state.  相似文献   

12.
The heat capacity of 1T-TaS2 has been measured over the temperature range including the successive phase transitions (140 K–370 K) by an adiabatic calorimeter. There are three transitions in the measured temperature range, two first-order transitions (at about 226 K (T1) and about 353.5 K (T3)) and one small anomaly at about 283 K (T2) with a broad peak. The transition enthalpies are as follows; ΔH1=52±5 cal·mol-1, ΔH2=7.5±2 cal· mol-1 and ΔH3=122±8cal·mol-1.  相似文献   

13.
At 300 K oxygen chemisorbs on Ag(331) with a low sticking probability, and the surface eventually facets to form a (110)?(2 × 1) O structure with ΔΦ = +0.7 eV. This facetting is completely reversible upon O2 desorption at ~570 K. The electron impact properties of the adlayer, together with the LEED and desorption data, suggest that the transition from the (110) facetted surface to the (331) surface occurs at an oxygen coverage of about two-thirds the saturation value. Chemisorbed oxygen reacts rapidly with gaseous CO at 300 K, the reaction probability per impinging CO molecule being ~0.1. At 300 K chlorine adsorbs via a mobile precursor state and with a sticking probability of unity. The surface saturates to form a (6 × 1) structure with ΔΦ = +1.6 eV. This is interpreted in terms of a buckled close-packed layer of Cl atoms whose interatomic spacing is similar to those for Cl overlayers on Ag(111) and Ag(100). Desorption occurs exclusively as Cl atoms with Ed ~ 213 kJ mol?1; a comparison of the Auger, ΔΦ, and desorption data suggests that the Cl adlayer undergoes significant depolarisation at high coverages. The interaction of chlorine with the oxygen predosed surface, and the converse oxygen-chlorine reaction are examined.  相似文献   

14.
《Surface science》1987,182(3):499-520
Photoelectron spectroscopy (UPS), thermal desorption spectroscopy (TDS), isotope exchange experiments, work function change (δφ) and LEED were used to study the adsorption and dissociation behavior of H2O on a clean and oxygen precovered stepped Ni(s)[12(111) × (111)] surface. On the clean Ni(111) terraces fractional monolayers of H2O are adsorbed weakly in a single adsorption state with a desorption peak temperature of 180 K, just above that of the ice multilayer desorption peak (Tm = 155 K). In the angular resolved UPS spectra three H2O induced emission maxima at 6.2, 8.5 and 12.3 eV below EF were found for θ ≈ 0.5. Angular and polarization dependent UPS measurements show that the C2v symmetry of the H2O gas-phase molecule is not conserved for H2O(ad) on Ni(s)(111). Although the Δφ suggest a bonding of H2O to Ni via the negative end of the H2O dipole, the O atom, no hints for a preferred orientation of the H2O molecular axes were found in the UPS, neither for the existence of water dimers nor for a long range ordered H2O bilayer. These results give evidence that the molecular H2O axis is more or less inclined with respect to the surface normal with an azimuthally random distribution. H2O adsorption at step sites of the Ni(s)(111) surface leads in TDS to a desorption maximum at Tm = 225 K; the binding energy of H2O to Ni is enhanced by about 30% compared to H2O adsorbed on the terraces. Oxygen precoverage causes a significant increase of the H2O desorption energy from the Ni(111) terraces by about 50%, suggesting a strong interaction between H2O and O(ad). Work function measurements for H2O+O demonstrate an increase of the effective H2O dipole moment which suggests a reorientation of the H2O dipole in the presence of O(ad), from inclined to a more perpendicular position. Although TDS and Δφ suggest a significant lateral interaction between H2O+O(ad), no changes in the molecular binding energies in UPS and no “isotope exchange” between 18O(ad) and H216O(ad) could be observed. Also, dissociation of H2O could neither be detected on the oxygen precovered Ni(s)(111) nor on the clean terraces.  相似文献   

15.
Heat capacity of α-NH4HgCl3 crystal has been measured with an adiabatic calorimeter from 11 to 300 K. A sharply peaked anomaly due to an order-disorder change of the ammonium ions was found at 54.97 ± 0.04 K. The entropy and enthalpy changes were estimated to be ΔS = 5.2 ± 1.0JK?1 mol?1 and ΔH = 342 ± 65 J mol?1. In accordance withthe structural two-dimensionality of α-NH4HgCl3 crystal, Onsager's solution of the two-dimensional Ising model was used in calculation of the transition temperature. On the assumption that the octopole-octopole interaction is responsible for the ordering of the ammonium ions in the present crystal and in ammonium chloride, the calculation gives 74.44 K for the transition temperature. Several possibilities were discussed for explaining the remaining discrepancy between the observed and calculated transition temperatures.  相似文献   

16.
At 300 K and in the coverage regime (0<θ<13) bromine chemisorbs rapidly on Pd(111); the sticking probability and dipole moment per adatom remain constant at 0.8 ± 0.2 and 1.2 D, respectively. This stage is marked by the appearance of a √3 structure: desorption occurs exclusively as atomic Br. At higher coverages, desorption of molecular Br2 begins (desorption energy ~130 kJ mol?1) as does the nucleation and growth of PdBr2 on the surface. This latter stage is signalled by the appearance of a √2 LEED pattern and the observation of PdBr2 as a desorption product (desorption energy ~37 kJ mol?1). Some PdBr2 is also lost by surface decomposition and subsequent evaporation of atomic Br. The data indicate that the transition state to Br adatom desorption is localised and that PdBr2(a) ? Br(a) interconversion occurs; these surface species do not appear to be in thermodynamic equilibrium during the desorption process.  相似文献   

17.
The formation of water by the reaction of preadsorbed oxygen with hydrogen on a Pt(111) surface has been characterized, using secondary ion mass spectroscopy, below the desorption temperature of H2O (180 K). The concentration of chemisorbed water was monitored during the reaction by following the SIMS H3O+ signal. Reaction profiles were measured over a temperature range of 120 to 153 K, and an H2 pressure range of 10-9 to 10-6 Torr. Under all conditions the reaction profiles were characterized by an induction time, a region of rapid reaction, and finally a steady decline in the rate. In the rapid region, an overall activation energy of 2.9 ± 0.3 kcalmol-1 and a half-order H2 pressure dependence were observed. At low initial oxygen concentrations the induction time increased and the maximum rate decreased. The reaction was slow in the absence of gas phase hydrogen, even when the surface coverage of hydrogen was relatively high. Water and hydrogen thermal desorption spectra, measured after stopping the reaction by removal of gas phase hydrogen, were complex functions of the H2 exposure, exhibiting several peaks between 170 and 400 K. However, after an exposure large enough to drive the reaction to completion, only one H2O peak at 173 K and one H2 peak at 350 K were observed. The results indicate that only a fraction of the total H(a) on the surface was readily available for reaction during H2 exposure at T ? 153 K. the remainder either recombined to form H2 or reacted with O(a) during the thermal desorption ramp. There is good evidence for a surface rearrangement during the induction period. A model is proposed which involves the formation of water clusters that accelerate the rate.  相似文献   

18.
《Surface science》1996,364(2):L580-L586
The adsorption and decomposition of formic acid on NiO(111)-p(2 × 2) films grown on Ni(111) single crystal surface were studied by temperature-programmed desorption (TPD) spectroscopy. Exposure of formic acid at 163 K resulted in both molecular adsorption and dissociation to formate. The adsorbed formate underwent further dissociation to H2, CO2 and CO. H2 and CO2 desorbed at the same temperatures of 340, 390 and 520 K, while CO desorbed at 415 and 520 K. The desorption features varied with the formic acid exposure. Two reaction channels were identified for the decomposition of formate under equilibrium with gas-phase formic acid with a pressure of 2.5 × 10−4Pa, one preferentially producing H2 and CO2 with an activation energy of 22 ± 2 kJ mol−1 and the other preferentially producing CO and H2O with an activation energy of 16 ± 2 kJ mol−1. The order of both reaction paths was 0.5 with respect to the pressure of formic acid.  相似文献   

19.
《Surface science》1990,236(3):L372-L376
A new low temperature displacement mechanism for CO on the Pt(111) surface has been observed in the presence of high pressures of hydrogen (0.001 to 0.1 Torr H2). Temperature-programmed fluorescence yield near-edge spectroscopy (TP FYNES) was used to continuously monitor the CO coverage as a function of temperature both with and without hydrogen. For hydrogen pressures above 0.01 Torr, removal of CO begins at 130 K (Ed = 10.6 kcal/mol) instead of near the desorption temperature of 400 K (Ed = 26 kcal/mol). The large decrease in CO desorption energy appears to be caused by substantial repulsive interactions in the compressed monolayer induced by coadsorbed hydrogen. The new low temperature CO desorption channel appears to be caused by displacement of the compressed CO adlayer by coadsorbed hydrogen. In addition, the desorption activation energy for the main desorption channel of CO near 400 K is lowered by ~ 1 kcal/mol for hydrogen pressures in the 0.001 to 0.1 Torr range. These new results clearly emphasize the importance of in-situ methods capable of performing kinetic experiments at high pressures on well characterized adsorbed monolayers on single crystal surfaces. High coverages of coadsorbed hydrogen resulting from substantial overpressures may substantially modify desorption activation energies and thus coverages and kinetic pathways available even for strongly chemisorbed species. These phenomena may play an important role in surface reactions which occur at high pressure.  相似文献   

20.
P. U. Singare 《Ionics》2016,22(8):1433-1443
The short-lived radiotracer isotopes were applied to study the kinetics and thermodynamic feasibility of iodide as well as bromide ion adsorption reactions using industrial-grade resin materials. Free energy of activation (ΔG ?) and energy of activation (E a) were calculated by using Arrhenius equation, enthalpy of activation (ΔH ?), and entropy of activation (ΔS ?) calculated by using the Eyring-Polanyi equation. These parameters were used to predict the thermodynamic feasibility of the two ion adsorption reactions performed by using Dowex SBR LC and Indion-810 resins. It was observed that during iodide ion adsorption reactions, the values of energy of activation (?18.79 kJ mol?1), enthalpy of activation (?21.37 kJ mol?1), free energy of activation (58.13 kJ mol?1), and entropy of activation (?0.26 kJ K?1 mol?1) calculated for Indion-810 resins were lower than the respective values of ?4.28 kJ mol?1, ?6.87 kJ mol?1, 64.97 kJ mol?1, and ?0.23 kJ K?1 mol?1 calculated for Dowex SBR LC under similar experimental conditions. Identical trends were observed for the two resins during bromide ion adsorption reactions. The low values of different thermodynamic parameters obtained for Indion-810 resins during both the ion adsorption reactions indicate that the reactions are thermodynamically more feasible using Indion-810 resins as compared to Dowex SBR LC resins. It is expected here that the present nondestructive technique can be extended further for different ions in the solution in order to predict the thermodynamic feasibility of different ion adsorption reactions for the range of resins which are widely used for treatment of industrial waste water effluent.  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