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 562 毫秒
1.
Nitric oxide desorption and reaction kinetics are compared on the (111), (110),and (100) planes of platinum using temperature programmed desorption mass spectrometry. NO exhibits large crystallographic anisotropies with the (100) plane having stronger bonding and much higher decomposition activity than the (110) or (111) planes. The desorption activation energies for the major tightly bound states are 36, 33.5, and 25 kcal mole?1 on the (100), (110), and (111) planes respectively. Pre-exponential factors for these states on the (110) and (111) planes are 1 × 1016±0.5s?1. The major tightly bound state on the (100) plane dissociates to yield 50% N2 and O2, but all other states all planes desorb without significant decomposition. The fraction decomposed is less than 2% on the Pt(111) surface.  相似文献   

2.
The growth modes and interaction of vapor-deposited Cu on a clean Pt(111) surface have been monitored by Auger electron spectroscopy (AES), low energy electron diffraction (LEED), and work function measurements. The LEED data indicate that below 475 K Cu grows in p(1 × 1) islands in the first monolayer with the interatomic Cu spacing the same as the Pt(111) substrate. The second monolayer of Cu grows in epitaxial, rotationally commensurate Cu(111) planes with the CuCu distance the same as bulk Cu. For substrate temperatures below ~ 475 K, the variation of work function and “cross-over beam voltage” with Cu coverage show characteristic features at one monolayer that are quite useful for calibration of θCu. Above 525 K, Cu-Pt alloy formation was observed in AES and LEED data. Thermal desorption spectroscopy of H2 and CO has demonstrated that simple site blocking of the Pt(111) surface by vapor-deposited Cu occurs linearly with chemisorption being essentially eliminated at θCu = 1.0–1.15. Conclusions drawn from this work correlate very favorably with the well-known effects of under potentially deposited copper on the electrochemistry of the H22H+ couple at platinum electrodes.  相似文献   

3.
The thermal and electro impact behaviour of NO adsorbed on Pt(111) and Pt(110) have been studied by LEED, Auger spectroscopy, and thermal desorption. NO was found to adsorb non-dissociatively and with very similar low coverage adsorption enthalpies on the two surfaces at 300 K. In both cases, heating the adlayer resulted in partial dissociation and led to the appearance of N2 and O2 in the desorption spectra. The (111) surface was found to be significantly more active in inducing the thermal dissociation of NO, and on this surface the molecule was also rapidly desorbed and dissociated under electron impact. Cross sections for these processes were obtained, together with the desorption cross section for atomically bound N formed by dissociation of adsorbed NO. Electron impact effects were found to be much less important on the (110) surface. The results are considered in relation to those already obtained by Ertl et al. for NO adsorption on Ni(111) and Pd(111), and in particular, the unusual desorption kinetics of N2 production are considered explicitly. Where appropriate, comparisons are made with the behaviour of CO on Pt(111) and Pt(110), and the adsorption kinetics of NO on the (110) surface have been examined.  相似文献   

4.
Mössbauer studies of spin reorientation transitions in the high magnetostrictive materials RFe2(R = rareearth) were performed. The transitions studied were [100] → [110] (Ho0.9Tb0.1Fe2 at 240 K), [100] → [111] (Dy0.7Tb0.3Fe2 at 220 K), [110] → [100] (Ho0.5Er0.5Fe2 at 135 K and Ho0.3Tm0.7Fe2 at 110 K), [111] → [100] (Dy0.2Er0.8Fe2 at 140 K) and [110] → [111] (Ho0.7Tb0.3Fe2 at 120 K). The first four transitions were found to be of second order, continuous reorientation of magnetization. In Dy0.2Er0.8Fe2 and Ho0.7Tb0.3Fe2 a sharp discontinuous first order transition is observed. In all systems the Mössbauer recoil free fraction, its anisotropy, the isomer shift and hyperfine interaction parameters change sharply through the spin reorientation transition. All phenomena observed can be understood in terms of changes in static magnetostrictive distortions combined with critical phonon softening at the spin reorientation phase transitions. The system Ho1-xTbxFe2 is unique in behaviour as Ho0.9Tb0.1Fe2 shows the highest magnetostrictive effects, whereas Ho0.7 Tb0.3Fe2 shows almost none.  相似文献   

5.
The generalized susceptibility, χ(q), in Pd and Pt for q along the [100], [110], [111], and [120] directions was determined from their APW and RAPW energy band structures, respectively, using the analytic tetrahedron linear energy scheme of Rath and Freeman. The band structures were previously found to yield Fermi surface radii, temperature dependencies of the static magnetic susceptibility, χ(T), resistivity, and a spin lattice relaxation, T1T, in very good agreement with experiment. In the χ(q) calculations, we used 2048 tetrahedra in 1/48th irreducible BZ and the energy eigenvalues for bands 4, 5, and 6 which cross the Fermi energy as fitted to a Fourier series representation. The intraband parts of χ(q) at q = 0 for both metals are found to agree with the density of states at the Fermi energy to without 0.5%. Our results show that the dominant contribution to χintra arises from the dominant band 5 whose “jungle-gym” FS has strong nesting features; the main peak for Pd occurs at the same q value (= 0.65π/a) for q along the [0q0], [q, q, 0], and [q, q, q] directions. The locus of this main peak is a square in the (0, 0, 1) plane. The maximum of χintra for q along the [110] and [111] directions are 23% and 13%, respectively, higher than the value of χ(q) at q = 0. For q along the [010] and [120] directions, the peak is, however, lower than the value of χintra at q = 0. Hence, while phonon anomalies are predicted for the [110] and [111] directions, no anomaly is predicted for either the [100] or [120] direction. The predicted q value for the [110] anomaly, q = 0.65π/a is close to the experimental value of ~0.7 π/a. Although there may be a hint of an anomaly at 0.56 [111] in the measurements, a more detailed investigation of this region is called for. For platinum, χintra for q along the [010], [110] and [111] directions has main peaks which occur at q = 0.68 π/a, 0.75 π/a, and 0.85 π/a, respectively. Here too, this main peak comes from the nesting of the jungle-gym Fermi surface which is not, however, as flat as that of palladium. Anomalies are predicted (although weaker in Pt than in Pd) along [110] and [111] but not along [100] and [120]. The [110] anomaly is close to the measured q value (~0.7–0.8 π/a). Also in agreement with experiment, we predict a weaker [110] anomaly for Pt than for Pd. In both Pd and Pt, weaker anomalies are predicted for the [111] direction than for the [110] direction.  相似文献   

6.
The chemisorption of NO on clean and Na-dosed Ag(110) has been studied by LEED, Auger spectroscopy, and thermal desorption. On the clean surface, non-dissociative adsorption into the α-state occurs at 300 K with an initial sticking probability of ~0.1, and the surface is saturated at a coverage of about 125. Desorption occurs without decomposition, and is characterised by an enthalpy of Ed ~104 kJ mol?1 — comparable with that for NO desorption from transition metals. Surface defects do not seem to play a significant role in the chemistry of NO on clean Ag, and the presence of surface Na inhibits the adsorption of αNO. However, in the presence of both surface and subsurface Na, both the strength and the extent of NO adsorption are markedly increased and a new state (β1NO) with Ed ~121 kJ mol?1 appears. Adsorption into this state occurs with destruction of the Ag(110)-(1 × 2)Na ordered phase. Desorption of β1NO occurs with significant decomposition, N2 and N2O are observed as geseous products, and the system's behaviour towards NO resembles that of a transition metal. Incorporation of subsurface oxygen in addition to subsurface Na increases the desorption enthalpy (β2NO), maximum coverage, and surface reactivity of NO still further: only about half the adsorbed layer desorbs without decomposition. The bonding of NO to Ag is discussed, and comparisons are made with the properties of α and βNO on Pt(110).  相似文献   

7.
Selected thermal desorption and valence band photoemission data on the chemisorption of CO on PtCu(111) surfaces are presented. The main objective is to make a comparison with CO chemisorption on an annealed (1 × 3) reconstructed Pt0.98Cu0.02(110) surface. The (111) alloy surfaces are unreconstructed (1 × 1) surfaces, with average near-surface Cu concentrations ranging from ? 7.5% to ? 20% as indicated by the Cu 920 eV Auger signal. It is observed that the effect of alloying Pt(111) with Cu is to progressively lower the desorption peak temperature and hence the free energy of CO desorption from Pt sites. A second observation is that the energy distribution of the Cu 3d-derived states is little affected by CO adsorption on Cu sites at 155 K. Both these results offer a contrast to the results for CO/Pt0.98Cu0.02(110) reported earlier.  相似文献   

8.
Adsorption of NO and O2 on Rh(111) has been studied by TPD and XPS. Both gases adsorb molecularly at 120 K. At low coverages (θNO < 0.3) NO dissociates completely upon heating to form N2 and O2 which have peak desorption temperatures at 710 and 1310 K., respectively. At higher NO coverages NO desorbs at 455 K and a new N2 state obeying first order kinetics appears at 470 K. At saturation, 55% of the adsorbed NO decomposes. Preadsorbed oxygen inhibits NO decomposition and produces new N2 and NO desorption states, both at 400 K. The saturation coverage of NO on Rh(111) is approximately 0.67 of the surface atom density. Oxygen on Rh(111) has two strongly bound states with peak temperatures of 840 and 1125 K with a saturation coverage ratio of 1:2. Desorption parameters for the 1125 peak vary strongly with coverage and, assuming second-order kinetics, yield an activation energy of 85 ± 5 kcalmol and a pre-exponential factor of 2.0 cm2 s?1 in the limit of zero coverage. A molecular state desorbing at 150 K and the 840 K state fill concurrently. The saturation coverage of atomic oxygen on Rh(111) is approximately 0.83 times the surface atom density. The behavior of NO on Rh and Pt low index planes is compared.  相似文献   

9.
K.E. Lu  R.R. Rye 《Surface science》1974,45(2):677-695
The adsorption and flash desorption of hydrogen and the equilibration of H2 and D2 has been studied on the (110), (211), (111) and (100) planes of platinum. Desorption from Pt (211), a stepped surface composed of (111) and (100) ledges, yields a desorption spectrum which apparently is a composite of desorption from the individual ledges. Pt (110) is quite similar to the tungsten structural analog, W (211), in that both yield two-peak desorption spectra, and on both planes adsorption kinetics are dramatically different for filling of the two states. On all four planes adsorption kinetics are apparently proportional to (1 ? θ)2, and estimates of the initial sticking probabilities show them to decrease in the order: (110) > (211) > (100) > (111). Equilibration activity follows approximately the same order [(110) > (211) > (111) > (100)] with a factor of ~ 5 difference between the most and least active planes; no extraordinary activity is observed for the stepped surface, Pt(211). Below ~ 570 K equilibration of H2 and D2 is activated by less than 2 kcal/mole with the magnitude dependent on the specific face, and above this temperature the reaction is nonactivated. The non-activated case apparently results from absorption followed by statistical mixing on the surface. Calculated rates for HD production per cm2 based on this model are in excellent agreement with the experimental values for Pt(110) and Pt(211), and in somewhat poorer agreement in the case of Pt (111) and Pt (100). This latter is probably due to the greater inaccuracy in the values of the sticking coefficients on these planes.  相似文献   

10.
The formation of a Cu monolayer on Pt(100), (110) and (111) was investigated by optical and electrochemical techniques. The adsorption isotherms as obtained by cyclic current-potential curves clearly show that the monolayer is deposited in various steps at underpotentials. Differential reflectance spectroscopy at normal incidence was used to detect structural changes of the adsorbate as the coverage increases. For Cu on Pt(110) a pronounced anisotropy in ΔRRwas observed as the electric field vector of the linearly polarized light was rotated. From these measurements it was deduced that Cu in the submonolayer range is deposited onto Pt(110) in rows along the [11&#x0304;0] direction of the substrate. No such anisotropy was found for Cu on Pt(100) and Pt(111) and for surface oxide formation on all three low index faces of Pt. The spectral dependence of the normalized reflectance change,ΔRR, for the Cu monolayer on Pt(hkl) is shown and discussed.  相似文献   

11.
CO adsorption on Pt(111) and vicinal Pt(111) surfaces has been studied by means of work function variation and He scattering measurements. AES and LEED were used mainly for correlations with other work. Special attention has been paid to the low coverage regime (θco < 0.1) with emphasis on surface structural dependencies. The minimum of the work function versus CO exposure curve occurs at a coverage less than 11% on “kink-free” surfaces. This is much lower than the hitherto commonly accepted value of 33%, and does not relate to any observed LEED superstructure. The value of Δφmin depends strongly on the surface structure. For an “ideal” Pt(111) surface with a step density less than 10?3 at a temperature of 300 K, Δφmin = ?240 meV. The scattering cross section Σ of CO adsorbed on Pt(111) for 63 meV He is typically > 250 Å2, i.e. much larger than expected from the Van der Waals radii of He and CO. For two nominal Pt(111) surfaces with step densities of 10?2 and less than 10?3, respectively, the measured Σ values varied by a factor of three. This can be explained by preferential CO occupation of defect sites, which are already not “seen” by thermal helium. By comparing results on a stepped (997) and a kinked (12 11 9) Pt surface with similar defect densities, the kinks are proven to play a decisive role. They probably form saddles in the recently proposed activation barrier for migration between terrace and step sites.  相似文献   

12.
Electron spin resonance (ESR) experiments have been carried out at cryogenic temperatures (4.2 ? T ? 35 K) and room temperatures at 9.0 and 20.9 GHz on the Pb0 and Pb1 (commonly referred to as Pb) spin-active defects residing at the Si/SiO2 interface. The ESR lineshapes were shown to display gaussian characteristics with inhomogeneous line broadenings amounting to 0.7 ± 0.1 and 0.2 ± 0.1 mT at K band for Pb0, and Pb1 respectively, whereas the oscillator strength of both signals followed the paramagnetic law (~ T?1) down to 4.2 K within experimental error. In general the observed Pb spectrum appeared to have fewer peaks than in other observations, at most displaying two distinguishable lines. Mostly however, only one somewhat broad signal (of measured peak-to-peak linewidth >Bptp ? 1.5 ± 0.15 mT and g = 2.0058 ± 0.0002 for $?B 6 [001] was observed. By fully incorporating ΔBptp data for the first time in these observations and using computer simulations, it has been shown that the pervailing experimental spectrum always contains the signals from both kind of centers although mostly these are not separately discernable. Further, it emerged that the actual appearance of the experimental spectra is dirigated by the presence of a distribution of the SiIII unsaturated bond orientations around the ones normally prescribed by the Si crystallinity at the interface. It is found that for both centers this “angle” distribution predominantly occurs vertically with respect to the (001) interface plane. Ion implantation of 1014 As+ cm?2 at 60 or 80 keV into the oxide layer of the Si/SiO2 structure is shown to randomize the Pb dangling bond (DB) orientations (resulting in an isotropic g value behavior) but the effect of this is totally eliminated by subsequent annealing at 1000°C in N2 ambient. It is argued that ESR has become a very sensitive means to study the “purity” of interfacial DB positions with respect to the Si single-crystal prescribed positions and to enable the display of collective fingerprints of the interface defects.  相似文献   

13.
An analysis has been made of on- and off-specular electron energy loss spectra (EELS) from C2H4 and C2D4 adsorbed on a clean Ni(110) and also a carbided Ni(110) surface. The carbided surface was prepared by heating the clean Ni surface in ethylene to 573 K or above. EELS spectra were obtained using a Leybold-Heraeus spectrometer at a beam energy of 3.0 eV and with a resolution of ca. 6.5 meV (ca. 50 cm?1).The loss spectrum from ethylene at low temperatures (110 K) showed principal features at 3000 (w), 1468 (w), 1162 (s), 879 (w) and 403 cm?1 (s) (C2D4 adsorption) and 2186 (w), 1258 (ms), 944 (ms), 645 (w) and 400 cm?1 (s) (C2D4 adsorption). The overall pattern of wavenumbers and intensifies of the C2H4/C2D4 loss peaks is very similar in form (although systematically different in positions) to those previously observed on Ni(111) (ref.1) and Pt(111) (ref.2) surfaces at low temperatures. Like these earlier spectra,the EELS results for C2H4/C2D4 adsorbed on clean Ni(110) can be well interpreted in terms of a MCH2CH2M/MCD2CD2M species (M = metal) with the CC bond parallel to the surface.After adsorption on the carbided Ni(110) surfaces at 125 K,the main loss features occur at 3065 (m), 2992 (m), 1524 (ms), 1250 (s), 895 (s), and 314 cm?1 (vs) (C2H4 adsorption) and 2339 (m), 2242 (m), 1395 (s), 968 (s), 661 (m) and 314 cm?1 (vs). With the exceptions of reduced intensities of the bands at 895 cm?1 (C2H4) and 661 cm?1 (C2D4) this pattern of losses - particularly the 1550-1200 cm?1 features which can be assigned to coupled νCC and δCH2/δCD2 modes - is well related to similar results on Cu(100) (ref.3) and Pd(111) (ref.4) which have been interpreted convincingly in terms of the presence of π-bonded species, (C2H4)M or (C2D4)M on the surface. This structural assignment is supported by comparison with the vibrational spectra of Zeise's salt, K[PtCl3(C2H4)].H2O (refs.5&6).Spectral changes occur on warming C2H4 on the clean Ni(110) surface with a growth of a feature near 895 cm?1 at 200 K. At 300 K a rather poorly-defined spectrum occurs, which differs substantially from those found on (111) surfaces of Pt (ref.2), Rh (ref.7) or Pd (ref.8) at room temperature. These latter have been attributed to the ethylidyne, CH3.CM3, surface species (ref.9). For adsorption on Ni(110) there is clearly a mixture of species at room temperature.The analysis of the vibrational spectra of selected metal-cluster compounds of known structure with selected hydrocarbon ligands has helped substantially to assign the spectra of surface species in terms of bonding structures of the adsorbed species, as in the cases of the identification of (C2H4)M π-adsorbed (refs.5&6) and the ethylidyne CH3.CM3 species (ref.9). We have recently analysed the infrared and Raman spectra of the cluster compound (C2H2)Os3(CO)10 and its deuterium-containing analogue. The infrared frequency and intensity pattern for the A′ modes (CS symmetry) of the two isotopomers bears a remarkable resemblance to EELS spectra previously obtained at low temperature for C2H2/C2D2 adsorbed on Pt(111) (ref.2) and (after taking into account systematic frequency shifts) for Pd(111) (ref.4). There is good evidence for believing that the structure of the hydrocarbon ligand interacting with the osmium complex takes the form
where the arrow denotes a π-bond to the third metal atom. This strongly confirms the structure for the low-temperature acetylene species on Pt(111) as proposed by Ibach and Lehwald (ref.2).Finally the room-temperature spectra for ethylene adsorbed on finely-divided silica-supported Pt and Pd catalysts have previously been interpreted in terms of the presence of MCH2CH2M (ref.10) and π-bonded (C2H4)M species (ref.11). However comparisons with the more recent EELS spectra from ethylene on Pt(111) at room temperature (ref.2) now leads to a reassignment of the 2880 cm?1 band, on Pt, previously assigned to MCH2CH2M, together with a new, related,band at 1340 cm?1 (ref.12), to the ethylidyne species CH3CPt3 found on the single crystal surface.More detailed analyses of the spectra reported here will be published later. Acknowledgement is given to substantial assistance for this programme of research from the Science and Engineering Research Council.  相似文献   

14.
Nuclear microanalysis (NMA) has been used to determine the absolute coverages of oxygen and CO adsorbed on Pt(111). The saturation oxygen coverage at 300 K is 3.9 ± 0.4 × 1014 O atoms cm?2 (θ = 0.26 ± 0.03), confirming the assignment of the LEED pattern as p(2 × 2). The saturation CO coverage at 300 K is 7.4 ± 0.3 × 1014 CO cm?2 (θ = 0.49 ± 0.02). The low temperature saturation CO coverages on Pt(100), (110) and (111) surfaces are compared.  相似文献   

15.
The bonding and orientation of CH{in3}NCO on Pt{110} and Cu{110} are studied by HREELS, ARUPS, AES, and LEED. CH{in3}NCO is found to adsorb nondissociatively on both Cu{110} and Pt{110} at T = 160 K, bonding primarily through the 2π a′ orbital with the NCO group lying down on the surface and the methyl group largely unperturbed. We propose that the absence of a strong HREELS band at about 2290 cm?1, which is the liquid phase frequency for the NCO asymmetric stretching mode, combined with the presence of strong bands between 1000 and 1450 cm?1 provides a “finger-print” for NCO bonded to the surface in a lying-down configuration.  相似文献   

16.
NO adsorbs on Pt(111) with a (temperature independent) initial sticking coefficient S0=0.88. The fraction of molecules not being chemisorbed is directly inelastically scattered back due to failure of translational energy accommodation. The nonlinear variation of s with coverage can well be described by a precursor-state model, the precursor state being formed by NO molecules translationally and rotationally accommodated in a physisorbed second layer. Dissociation is essentially restricted to defect sites and is negligible on perfect (111) planes. These defect sites (present in small concentration) are first populated and are also sampled by the modulated beam technique yielding an activation energy for desorption Ed = 33.1 kcal/mole and preexponential factor vd = 1015.5s?1. Isothermal desorption measurements yielded Ed and vd as a function of coverage: Ed rapidly drops from its initial value (at defect sites) to about 27 kcal/mole — which value is considered as representing the adsorption energy on a perfect (111) plane — and then decreases continuously due to effective repulsive interactions. Simultaneously vd is decreasing to about 1012 s?1 at θ = 0.25 which marks the equilibrium coverage to be reached at 300 K. If the surface is precovered with oxygen atoms the NO sticking coefficient is reduced to 0.6, and the desorption parameters are lowered to Ed = 17.1 kcal/mole and vd= 1012.6s?1 (at zero NO coverage).  相似文献   

17.
The adsorption and reaction of glycine on the surface of a rutile TiO2(011) single crystal has been studied by X-ray Photoelectron Spectroscopy (XPS) and Temperature Programmed Desorption (TPD) techniques. Special attention was given to the formation and stability of the zwitterion structure (+NH3–CH2–COO?) in comparison to that of the dissociated structure (NH2–CH2–COO?). Both species have been observed on the surface at 300 K. The zwitterion structure was found less stable than the dissociated structure. This is in line with other experimental results related to proline on rutile TiO2(110) single crystal [13, 14], glycine on rutile TiO2(110) single crystal [17, 24] and computational results related to glycine on rutile TiO2(110) single crystal [25]. By 500 K most of the zwitterion structure has been converted to the dissociated one. TPD results indicated that glycine reacts in a similar way to carboxylic acids on this surface with the main decomposition products being ketene (CH2=C=O). Other masses left unassigned for were also observed during TPD. The most intense being m/e 55 that might be due to =CH–C(O)N=or C(O)N=CH fragments.  相似文献   

18.
The kinetics of the reorientation of the molecular impurities16O2 ? and18O2 ? in the host lattices KCl, KBr and KJ has been investigated by means of ESR as a function of uniaxial stress in the temperature range between 1,3°K and 4,5°K. Below 4°K the reorientation rate is proportional to the temperature, indicating that one-phonon processes dominate at low temperatures. ForT>4°K the reorientation through excited librational levels becomes significant. The large isotope effect in the reorientation rate of16O2 ? and18O2 ? confirms that tunneling plays an important role. The analogy between the diffusion of polarons and the reorientation of molecules in solids is discussed in detail. The evaluation and interpretation of the experimental results in terms of the theory of Gosar and Pire yields values for the tunneling matrix elements, the potential barrier between two equilibrium orientations, the libration frequencies and an effective moment of inertia. The theory accounts well for the contribution of the one-phonon processes to the relaxation rate for stress along [001]. For stress along [111] the agreement is only qualitative.  相似文献   

19.
The Peierls-Nabarro barrier and stress of thea/2〈111〉 edge dislocation on {112} and {110} plane inα-Fe at O K is calculated within the Peierls-Nabarro model. The method proposed by Nabarro is used, however, the sine force law is replaced by more general force laws based on two central interionic potentials inα-Fe. The values of the Peierls-Nabarro stress corresponding to one of the chosen interionic potentials, 3·5×10?4 μ and 1×10?4 μ on {112} plane (in the twinning direction) and on {110} plane, respectively, seem to be good estimates of the stress necessary to move edge dislocations inα-Fe at O K.  相似文献   

20.
Magnetic measurements have been made along the [100], [110] and [111]-axes of PrCd single crystal in magnetic fields up to 70 kOe in the temperature range from 4.2 to 300 K. PrCd shows metamagnetism below Tt = 25 K in which two critical fields are observed in the magnetization processes along the [110] and [111]-axes and a critical field along the [100]-axis. The metamagnetic behaviour is interpreted by a noncollinear antiferromagnetic model with the Pr moments of ~ 2μB directed to the 〈111〉-axes.  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