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 46 毫秒
1.
The X-ray induced desorption of H+ ions from NH3 layers adsorbed at T = 90 K on Ni(110) has been compared to the corresponding total electron yield (TY) in the photon energy range 390 to 900 eV. The H+ yield exhibits a jump at the N K-edge and the Ni L-edge which inversely varies with the NH3 layer thickness. The H+ Ni L-edge jump is closely correlated to the TY jump. Both vanish for the saturated NH3 multilayer, indicating that the observed Ni L-edge jump in the H+ yield is exclusively due to X-ray induced electron stimulated desorption (XESD). At the N K-edge, the near edge absorption fine structure of the H+ yield and TY of the saturated NH3 multilayer are distinctly different. This is interpreted as the H + yield being the superposition of direct photon stimulated ion desorption (PSID) and XESD. Based on the observed variation of the H+ yield near edge fine structure with varying NH3 layer thickness, a deconvolution of the PSID and XESD contributions is used to derive the relative contribution of PSID versus XESD to be 40% versus 60%, respectively. The relevance of this result for future PSID-SEXAFS studies is discussed. For monolayer NH3 on Ni(110) the polarization dependence of the N K-edge fine structure in the N(KVV) Auger yield indicates that the symmetry axis of NH3, is perpendicular to the surface.  相似文献   

2.
In the present work the ASED-MO method is applied to study the adsorption of cyclopentadienyl anion on a Ni(1 1 1) surface. The adsorption with the centre of the aromatic ring placed above the hollow position has been identified to be energetically the most favourable. The aromatic ring remains almost flat, the H atoms are tilted 17° away from the metal surface. We modelled the metal surface by a two-dimensional slab of finite thickness, with an overlayer of c-C5H5, one c-C5H5 per nine surface Ni atoms. The c-C5H5 molecule is attached to the surface with its five C atoms bonding mainly with three Ni atoms. The NiNi bond in the underlying surface and the CC bonds of c-C5H5 are weakened upon adsorption. We found that the band of Ni 5dz2 orbitals plays an important role in the bonding between c-C5H5 and the surface, as do the Ni 6s and 6pz bands.  相似文献   

3.
《Surface science》1986,171(1):111-134
The mechanism of ethanol decomposition on the Ni(111) surface has been investigated between 155 and 500 K. The sequence of bond scission steps which occur as ethanol undergoes dissociative reactions on this surface has been deduced using deuterium and 13C isotopic labels. Bond activation occurs in the order (1) OH, (2) CH2 (methylene CH), (3) CC, (4) CH3 (methyl CH). The products observed are CH3CHO(g), CH4(g), CO(g), H2(g) and surface carbon, C(a). The latter species exhibits a carbidic AES lineshape in the temperature range 450 to 670 K, at which temperature it dissolves into the Ni bulk. Acetaldehyde, CH3CHO, and methane, CH4, desorb with the same threshold temperature (260–265 K), and the formation of both of these products is controlled by scission of the methylene CH bond (CH2 group). The CH3 group is cleaved from the intermediate surface CH3CHO species to form CH3(ads). H2 exhibits a broad, doublet desorption peak from 300 to 450 K. The carbonoxygen bond in ethanol remains intact and CO ultimately desorbs in a single desorption limited process (Tp = 430 K). A small fraction of CO(a) species undergo exchange with the carbidic surface carbon in a minor process observed above 440 K.  相似文献   

4.
Surface structures and electronic properties of hypophosphite, H2PO2, molecularly adsorbed on Ni(1 1 1) and Cu(1 1 1) surfaces are investigated in this work by density functional theory at B3LYP/6-31++g(d, p) level. We employ a four-metal-atom cluster as the simplified model for the surface and have fully optimized the geometry and orientation of H2PO2 on the metal cluster. Six stable orientations have been discovered on both Ni (1 1 1) and Cu (1 1 1) surfaces. The most stable orientation of H2PO2 was found to have its two oxygen atoms interact the surface with two PO bonds pointing downward. Results of the Mulliken population analysis showed that the back donation from 3d orbitals of the transition metal substrate to the unfilled 3d orbital of the phosphorus atom in H2PO2 and 4s orbital's acceptance of electron donation from one lone pair of the oxygen atom in H2PO2 play very important roles in the H2PO2 adsorption on the transition metals. The averaged electron configuration of Ni in Ni4 cluster is 4s0.634p0.023d9.35 and that of Cu in Cu4 cluster is 4s1.004p0.033d9.97. Because of this subtle difference of electron configuration, the adsorption energy is larger on the Ni surface than on the Cu surface. The amount of charge transfers due to above two donations is larger from H2PO2 to the Ni surface than to the Cu surface, leading to a more positively charged P atom in NinH2PO2 than in CunH2PO2. These results indicate that the phosphorus atom in NinH2PO2 complex is easier to be attacked by a nucleophile such as OH and subsequent oxidation of H2PO2 can take place more favorably on Ni substrate than on Cu substrate.  相似文献   

5.
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   

6.
Planar Pd(LH)2 complexes (LH2 = H2N C S C S N H2, CH3HNCSCSNHCH3) form mixed polymeric complexes with Ni(II), Cu(II), Zn(II) and Cd(II) in alcalic media, where the planar Pd(LH)2 complexes act as tetradentates with N-coordination. The electronic spectra and thermal behaviour are discussed, a thorough investigation of the i.r. spectra is presented and special attention has been given to the H/D, CH3/CD3 and 58Ni/62Ni, 63Cu/65Cu and 64Zn/68Zn isotopic shifts.  相似文献   

7.
Photoelectron spectroscopic studies of the oxidation of Ni(111), Ni(100) and Ni(110) surfaces show that the oxidation process proceeds at 295 and 485 K in two distinct steps: a fast dissociative chemisorption of oxygen followed by oxide nucleation and lateral oxide growth to a limiting coverage of 3 NiO layers. The oxygen concentration in the 295 K saturated oxygen layer on Ni(111) was confirmed by 16O(d,p) 17O nuclear microanalysis. At 295 and 485 K the oxide growth rates are in the order Ni(110) > Ni(111) > Ni(100). At 77 K the oxygen uptake proceeds at the same rate on all three surfaces and shows a continually decreasing sticking coefficient to saturation at ~2.1 layers (based upon NiO). An O 1sb.e. = 529.7 eV is associated with NiO, and O ls b.e.'s of ~531.5 and 531.3 eV can be associated, respectively, with defect oxide (Ni2O3) or (in the presence of H2O) with an NiO(H) species. The binding energies (Ni 2p, O 1s) of this NiO(H) species are similar to those for Ni(OH)2. Defect oxides are produced by oxidation at 485 K, or by oxidation of damaged films (e.g. from Ar+ sputtering) and evaporated films. Wet oxidation (or exposure to air) of clean nickel surfaces and oxides, and exposure of thick oxide to hydrogen at high temperature results in an O 1s b.e. ~531.3 eV species. Nuclear microanalysis 2H(3He,p) 4He indicates the presence of protonated species in the latter samples. Oxidation at 77 K yields O 1s b.e.'s of 529.7 and ~531 eV; the nature of the high b.e. species is not known. Both clean and oxidised nickel surfaces show a low reactivity towards H2O; clean nickel surfaces are ~103 times less reactive to H2O than to oxygen.  相似文献   

8.
The interaction of H2 with clean, Ni and Nb doped Mg(0001) surface are investigated by first-principles calculations. Individual Ni and Nb atoms within the outermost surface can reduce the dissociation barrier of the hydrogen molecule. They, however, prefers to substitute for the Mg atoms within the second layer, leading to a weaker catalytic effect for the dissociation of H2, a bottleneck for the hydriding of MgH2. Interestingly, co-doping of Ni and Nb stabilizes Ni at the first layer, and results in a significant reduction of the dissociation barrier of H2 on the Mg surface, coupled with an increase of the diffusion barrier of H. Although codoped Ni and Nb shows no remarkable advantage over single Nb here, it implies that the catalytic effect could be optimized by co-doping of “modest” transition metals with balanced barriers for dissociation of H2 and diffusion of H on Mg surfaces.  相似文献   

9.
The chemisorption, condensation, desorption, and decomposition of methanol, both CH3OH and CH3OD, on a clean Ni(110) surface have been characterized using high resolution electron energy loss spectroscopy, temperature programmed reaction spectroscopy, and low energy electron diffraction. The vibrational spectrum of the saturated chemisorbed layer, 7.4 × 1014 molecules cm?2, is almost identical to the infrared spectrum of liquid or solid methanol. Condensation of multilayers of methanol is facile at 80 K. The only quasi-stable intermediate isolated during the decomposition is a methoxy species, CH3O, which decomposes thermally to CO and H. The evolution of both CO and H2 occurs in desorption limited processes.  相似文献   

10.
Nickel (Ni) and cobalt (Co) metal nanowires were fabricated by using an electrochemical deposition method based on an anodic alumina oxide (Al2O3) nanoporous template. The electrolyte consisted of NiSO4 · 6H2O and H3BO3 in distilled water for the fabrication of Ni nanowires, and of CoSO4 · 7H2O with H3BO3 in distilled water for the fabrication of the Co ones. From SEM and TEM images, the diameter and length of both the Ni and Co nanowires were measured to be ∼ 200 nm and 5–10 μm, respectively. We observed the oxidation layers in nanometer scale on the surface of the Ni and Co nanowires through HR–TEM images. The 3 MeV Cl2+ ions were irradiated onto the Ni and Co nanowires with a dose of 1 × 1015 ions/cm2. The surface morphologies of the pristine and the 3 MeV Cl2+ ion-irradiated Ni and Co nanowires were compared by means of SEM, AFM, and HR–TEM experiments. The atomic concentrations of the pristine and the 3 MeV Cl2+ ion-irradiated Ni and Co nanowires were investigated through XPS experiments. From the results of the HR–TEM and XPS experiments, we observed that the oxidation layers on the surface of the Ni and Co nanowires were reduced through 3 MeV Cl2+ ion irradiation.  相似文献   

11.
The interaction of C2H2 with Ni surfaces has been studied by the Hartree-Fock-Slater-LCAO method (with core pseudopotentials). Different adsorption sites (π, di-σ, μ2, μ3) at the Ni(111) surface have been modelled by clusters of 1 to 4 Ni atoms; the structure of C2H2 and the Ni-C distance have been varied (3 structures, 2 distances). The acetylene-metal bonding can be interpreted in terms of π to metal donation and, especially, metal to π1 back donation effects which considerably weaken the C-C bond. These effects become increasingly important when more metal atoms are directly involved in the adsorption bonding: π < di-σ < μ2 < μ3. The calculated shifts in the ionization energies are in fair agreement with the experimentally observed shifts (by UPS) for C2H2 adsorbed on Ni(111) (and other Ni surfaces); these shifts do not depend very sensitively on the bonding situation, however, so that we could not assign the structure of adsorbed C2H2 solely on this basis. From the comparison between the measured C-C stretch frequency (by ELS) and the calculated C-C overlap populations, using a relation calibrated on Ni-acetylene complexes, we find that μ3 bonding of C2H2 with a Ni-C distance of about 1.9 Å is most probable on the Ni(111) surface; the CCH angle is estimated to be somewhat smaller than 150°. We have suggested an explanation for the surface specific dissociation of C2H2: C2 fragments (C-H bond breaking) have been observed on stepped Ni surfaces (at low temperature), CH fragments (C-C bond breaking) have been found on ideal surfaces (at higher temperatures).  相似文献   

12.
《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   

13.
Yields of ion impact induced electrons from very pure Ni(110) and Ni(111) surfaces have been measured. In several tilt planes the angle of incidence of a 5 keV H+, H+2 or H+3 ion beam is varied from perpendicular to grazing incidence. Below = 75° the yield increases as sec but shows characteristic depressions when the beam is incident along crystallographically low indexed lattice directions. This is explained by kinetic electron emission with respect to the projectile transparency of the crystal lattice.  相似文献   

14.
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   

15.
We use a periodic density functional theory (DFT) code to study the adsorption of CH3 and H, as well as their co-adsorption on a Ni(111) surface with and without Ni ad-atom, at a surface coverage of 0.25 monolayer (ML). We systematically investigate the site preference for CH3 and H. Then we combine CH3 and H in many co-adsorbed configurations on both surfaces. Methyl and hydrogen adsorption on a flat Ni(111) surface favours the hollow site over the top site. The presence of a Ni ad-atom stabilizes the adsorption of CH3 better than a flat surface, while hydrogen is more stable on a flat Ni(111) surface. When H and CH3 are co-adsorbed at nearest Ni neighbours on the (111) surface, their interaction is always repulsive. However, the dissociative adsorption of CH4 is stabilised when the fragments are infinitely separated. For the co-adsorbed fragments CH3 and H, in the presence of an ad-atom, the repulsive interaction is lowered, so that the dissociative form of CH4 is locally stable.  相似文献   

16.
The mechanism of H2 dissociative adsorption on Mn-modified Ni(111) surface is investigated and explained using spin-polarized density functional theory (DFT). Potential energy surface (PES) is used to determine the efficient reaction pathway of H2 on the surface. The dissociative adsorption of H2 in the hollow sites with its center-of-mass (CM) positioned on top of Ni atom has low activation barrier. This is lower compared if its CM is on top of the Mn atom. The difference in the reactivity of H2 with Ni and Mn as the CM is corroborated by the positions of the bonding and antibonding orbitals of H2 as it approaches the surface which is verified from local density of states (LDOS). The greater density of states in the region around the Fermi level of the dzz, dxz, and dyz orbitals of the Ni atom explains the low activation barrier obtained for the dissociation of H2 on top of the Ni atom in the Mn-modified Ni(111) surface.  相似文献   

17.
The ejection of H2O, O2, H2 and H from water ice at 30–140 K, bombarded by 0.5–6 keV H+ and Ne+ was studied experimentally. Neon ions in this energy range deposit their energy in the ice by nuclear collisions, whereas with protons of 0.5 to 6 keV the energy deposition mechanism shifts gradually from predominantly nuclear collisions to predominantly electronic processes. The existing theory of nuclear sputtering predicts very well the yield of ejected water molecules and the experimental results in the region of electronic processes agree well with the experimental results of Lanzerotti, Brown and Johnson. However, the major mass loss from water by ion bombardment is via the ejection of O2, H2 and H atoms, which exceed the ejection of water molecules. O2 and H2 production is markedly enhanced at temperatures exceeding ~100 K, whereas H2O and H production are temperature independent, suggesting that O2 and H2 are produced in the bulk of the ice whereas H2O and H atoms are ejected from the surface or near surface layers.  相似文献   

18.
《Surface science》1993,294(3):L945-L951
This paper reports the results of a theoretical study of Na, H and C subsurface atomic species in nickel and demonstrates how these interstitial atoms influence the reactivity of the Ni(111) surface and the structure of carbon species adsorbed on the surface. The benzene molecule, C6H6, in planar and nonplanar geometries, is used to probe bonding at the surface. Adsorption energies are calculated by ab initio configuration interaction techniques modelling the surface as an embedded cluster. Adsorption energies of planar C6H6 at the most stable, three-fold, adsorption site are 18 kcal/mol for the Ni(111) surface, and 10, 19 and 44 kcal/mol in the presence of the Na, H and C interstitials, respectively. The energies required for the planar to puckered distortion are 99 kcal/mol on Ni(111), 69 kcal/mol with the Na interstitial, 83 kcal/mol with H, and 134 kcal/mol with C compared to 198 kcal/mol for distortion of C6H6 in the gas phase. The possible relevance of these results to the nucleation of diamond on nickel are discussed. The results indicate that subsurface Na stabilizes tetrahedrally bonded carbon subunits of the diamond structure while subsurface C may make it easier for the overlayer to revert to a planar graphite structure.  相似文献   

19.
《Solid State Ionics》2006,177(15-16):1355-1359
We explore the hydrogen anode reaction chemistry at the Ni–zirconia triple phase boundary in solid oxide fuel cells by using hybrid density functional quantum chemistry calculations and cluster models. The activation energy for H spillover is calculated to be the same order of magnitude as experimental estimates at the reversible potential. Proton transport on the oxide surface is shown to be activated by strongly held hydrogen-bonded water molecules: in the absence of H2O the activation energy is calculated to be 4.98 eV and the water molecule reduces the activation energy to 0.25 eV. Substitutional Y3+ (for Zr4+) is shown to slow proton diffusion when present in the zirconia surface.  相似文献   

20.
D. Pillay  M.D. Johannes 《Surface science》2008,602(16):2752-2757
Adsorption strengths of hydrogen and sulfur both individually and together as co-adsorbates were investigated on Pt(1 1 1), Ni(1 1 1) and Pt3Ni(1 1 1) surfaces using density functional theory in order to determine the effect of metal alloying on sulfur tolerance. The adsorption strengths of H and S singly follow the same trend: Ni(1 1 1) > Pt(1 1 1) > Pt3Ni(1 1 1), which correlates well with the respective d-band center positions of each surface. We find that the main effect of alloying is to distort both the sub-layer structure and the Pt overlayer resulting in a lowered d-band. For all three surfaces, the d-band shifts downward non-linearly as a function of S coverage. Nearly identical decreases in d-band position were calculated for each surface, leading to an expectation that subsequent adsorption of H would scale with surface type similarly to single species adsorption. In contradiction to this expectation, there was no clearly discernable difference between the energies of coadsorbed H on Pt(1 1 1) and Ni(1 1 1) and only a slightly lowered energy on Pt3Ni(1 1 1). This provides evidence that coadsorbed species in close proximity interact directly through itinerant mobile electrons and through electrostatic repulsion rather than solely through the electronic structure of the surface. The combination of the lowered d-band position (arising from distorted geometry) and direct co-adsorbate interactions on Pt3Ni(1 1 1) leads to a lower energy barrier for H2S formation on the surface compared to pure Pt(1 1 1). Thus, alloying Pt with Ni both decreases the likelihood of S adsorption and favors S removal through H2S formation.  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