首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
X. -C. Guo  R. J. Madix   《Surface science》2004,550(1-3):81-92
The adsorption of oxygen and carbon dioxide on cesium-reconstructed Ag(1 1 0) surface has been studied with scanning tunneling microscopy (STM) and temperature programmed desorption (TPD). At 0.1 ML Cs coverage the whole surface exhibits a mixture of (1 × 2) and (1 × 3) reconstructed structures, indicating that Cs atoms exert a cooperative effect on the surface structures. Real-time STM observation shows that silver atoms on the Cs-covered surface are highly mobile on the nanometer scale at 300 K. The Cs-reconstructed Ag(1 1 0) surface alters the structure formed by dissociative adsorption of oxygen from p(2 × 1) or c(6 × 2) to a p(3 × 5) structure which incorporates 1/3 ML Ag atoms, resulting in the formation of nanometer-sized (10–20 nm) islands. The Cs-induced reconstruction facilitates the adsorption of CO2, which does not adsorb on unreconstructed, clean Ag(1 1 0). CO2 adsorption leads to the formation of locally ordered (2 × 1) structures and linear (2 × 2) structures distributed inhomogeneously on the surface. Adsorbed CO2 desorbs from the Cs-covered surface without accompanied O2 desorption, ruling out carbonate as an intermediate. As a possible alternative, an oxalate-type surface complex [OOC–COO] is suggested, supported by the occurrence of extensive isotope exchange between oxygen atoms among CO2(a). Direct interaction between CO2 and Cs may become significant at higher Cs coverage (>0.3 ML).  相似文献   

2.
C. Su  D. Tang  D. Heskett   《Surface science》1994,310(1-3):45-51
Using the technique of angle-resolved inverse photoemission, we have measured the dispersion of an unoccupied Cu(110) surface state for the clean Cu(110) surface and for the (1 × 2) reconstructed Na/Cu(110) surface along the symmetry lines. The dispersion of the crystal-induced surface state of clean Cu(110) at 2.05 eV above the Fermi energy at the point of the SBZ is free-electron-like with an effective mass of (1.0 ± 0.2)me at the point, which is in good agreement with other experimental results as well as a theoretical calculation. This surface state shifts to 2.5 eV above the Fermi energy for the (1 × 2) phase of Na/Cu(110) with a coverage of 0.25 ML, and the dispersion along the direction is considerably reduced compared to the clean surface. On the other hand, the dispersion of this state for (1 × 2) Na/Cu(110) (0.25 ML) along the direction is close to that of clean Cu(110). We account for these results within a missing-row picture of the Na-induced reconstruction.  相似文献   

3.
The adsorption of water of Ni(110) has been studied by nuclear reaction analysis (NRA), thermal desorption spectroscopy (TDS), LEED and work function measurements (Δφ). The major findings of this study are: (1) the saturation coverage of the first chemisorbed layer of water is slightly less than 0.5 water molecules per surface Ni atom or 0.5 ML (1 ML = 1 MONOLAYER = 1.14 × 1015 molecules cm−2) and the layer exhibits a c(2 × 2) LEED pattern; (2) this water desorbs in three separate desorption states; (3) the slightly less strongly bound, second layer of water can be distinguished from subsequent “ice” layers by a discrete work function change. These results are discussed in terms of a recently published model of Benndorf and Madey [C. Benndorf and T.E. Madey, Surf. Sci. 194 (1988) 63].  相似文献   

4.
The reactions of H2S with predosed surface oxygen on Ni(110) surfaces were studied for a variety of coverage conditions. The primary reaction product is H2O, but the details of the water formation and desorption depends on the coverage of both O and H2S.

For high coverages of oxygen (p(2 × 1)−O; 0.5 ML), the reaction to form water is quantitative. The loss of oxygen from the surface (as measured by AES) is equal to the increase in sulfur coverage. XPS and HREELS measurements indicate the presence of chemisorbed H2O immediately following large exposures of H2S on the oxygen predosed surface at 110 K. Deuterium incorporation results suggest that the primary mechanism for these coverage conditions involves direct transfer of hydrogen from SH or H2S moieties to the oxygen.

A second mechanism involving reaction of surface hydroxyl groups with surface hydrogen was also identified. This mechanism is particularly important for high coverages of oxygen (0.5 ML) and low coverages of H2S (0.15 ML), where water desorption was observed at 235 K, but was not observed spectroscopically at 110 K. The sequential addition of two surface hydrogen atoms to surface oxygen is not an important mechanism in this system.

These reactions were modeled using a bond-order conservation method, and the model successfully reproduced the important mechanistic conclusions.  相似文献   


5.
6.
I. D. Cocks  Q. Guo  E. M. Williams   《Surface science》1997,390(1-3):119-125
The TiO2(110)-(1 × 1) surface and its reconstruction as a (1 × 2) form have been studied with low energy electron diffraction (LEED), electron stimulated desorption ion angular distribution (ESDIAD) and scanning tunnelling microscopy (STM). Oxygen ion desorption occurs within a lobe perpendicular to the (1 × 1) surface, changing to two off-normal lobes for the (1 × 2) reconstruction. This transformation in the ESDIAD pattern is consistent with the added Ti2O3 row model of the (1 × 2) reconstruction proposed by Onishi and Iwasawa. STM studies of the stoichiometric and electron irradiated surfaces reinforce the association of the O+ ESD contribution with majority sites at the surface. Adsorption of acetic acid on the (1 × 1) surface produces a (2 × 1) overlayed and induces a reconstruction of the underlying substrate. ESDIAD reveals H+ ions emitted off-normally from dissociatively adsorbed acetate, and along the surface normal from surface hydroxyls. Adsorption of acetic acid on the (1 × 2) surface does not modify the LEED pattern, but ESDIAD reveals H+ desorption with a weaker off-normal contribution consistent with the Ti2O3 model of the reconstruction.  相似文献   

7.
B.J. Mrstik 《Surface science》1983,124(1):253-266
The initial stages of growth of epitaxial Ge overlayers on the GaAs(100) surface have been studied by LEED and AES on overlayers from 0.1 monolayers (ML) to 10 ML in thickness. It is found that a coverage of about 0.2 ML converts the initial clean surface reconstruction into a single domain (1 × 2) reconstruction with a surface atomic geometry very similar to that of clean Ge. Further growth does not significantly change the arrangement of atoms at the surface. Growth from 1 to 4 ML proceeds by a double layer growth mechanism which maintains the single (1 × 2) domain. Auger measurements indicate that the growing surface has a 12 ML As enrichment, and that the interface is not abrupt, but has a mixed GeGa or GeAs transition layer.  相似文献   

8.
The adsorption of oxygen on the Rh(110) surface has been studied by a variety of techniques. Low-energy electron diffraction shows the following patterns: (2 × 1)p2mg at 1 ML coverage and temperatures between 125 and 300 K; (2 × 2)p2mg at 0.5 ML coverage after heating to above 470 K; c(2 × 8) and complex streaked c(2 × 2n) patterns at coverages above 0.5 after heating to 470 K. These results are in partial agreement with previous work. Models for the first two structures are suggested. In the (2 × 2) structure, the oxygen is found to be much less reactive with CO at room temperature than in the (2 × 1) structure, suggesting that it is subsurface. A metastable (1 × 2) structure was produced from the (2 × 2) by reduction of the oxygen by CO at 450 K, and is interpreted as a surface reconstruction.  相似文献   

9.
PbI2 has been adsorbed on the clean InSb(001)-(4 × 1) reconstructed surface and on the InSb(001)-(1 × 3)-Pb lead covered reconstructed surface. On the clean surface epitaxial growth occurred with the unit mesh of the layered PbI2 aligning exactly with both the substrate [110] and [1 0] directions. On desorption a reaction occurred between the last layer of PbI2, and the substrate, forming a series of structures which finished with a well-formed (1 × 3)-Pb structure in which the surface is depleted/enriched in In/Sb compared to the clean (4 × 1). The Pb in this structure is thought to partially replace surface In. Epitaxial adsorption also occurred on the (1 × 3)-Pb surface generating a single, well-formed structure with the hexagonal net of the PbI2 aligned with just the [1 0] substrate direction. The structures and reactions are discussed and a row matching model is proposed to explain the single epitaxial orientation of PbI2 on the (1 × 3)-Pb surface.  相似文献   

10.
Detailed studies of the structures formed by the electrodeposition of atomic layers of Te on Au(1 1 1) surfaces from aqueous solutions were performed using in situ scanning tunneling microscopy (STM), as well as by UHV-EC techniques such as low energy electron diffraction and Auger electron spectroscopy. There are two features in the voltammetry that may be considered underpotential deposition (UPD). However, from the voltammetry, it is clear that the deposition process is kinetically slow, and from this study it appears that several atomic layer structures are actually formed at overpotentials. Prior to deposition, a surface excess of a tellurium oxide species coats the surface. This layer is then converted to a Au(1 1 1)(√3×√3)R30°–Te structure with an array of domain walls, at 1/3 ML. The initial structure appears to have a symmetric array of walls, resulting in a (13×13) periodicity, which then converts to a less symmetric structure where the domain walls form rhombi, with a larger periodicity. During the second UPD feature, the coverage increases, forming a (√7×√13) unit cell at 0.36 ML and then a (3×3) at 0.44 ML. Commensurate with the formation of these higher coverage structures, a roughening transition takes place, where the surface becomes pitted, resulting in about 40% of the surface being covered with single atom deep pits. This process appears to be related to the pits formed in the surfaces of self-assembled monolayers (SAM) of thiols on Au surfaces, and layers of Se and S on Au surfaces. Several theories have been suggested to account for these pits. The model that appears to best explain the pits is based on shrinking of the size of the underlying Au atoms, reconstructing the underlying Au. There also appears to be a high coverage structure, near 0.9 ML, that forms at potentials near where the (3×3) forms, but only by holding the potential for an extended period of time. Subsequent dissolution of this high coverage structure produces domains of disordered Te atoms, which gradually decrease in coverage until the (3×3) is again formed at 0.44 ML.  相似文献   

11.
K. W. Self  C. Yan  W. H. Weinberg   《Surface science》1997,380(2-3):408-416
Scanning tunneling microscopy and temperature-programmed desorption have been used to investigate the chemistry of water on Si(111)-(7 × 7) substrates which were misoriented 2° toward the [ 10] direction. Upon room temperature exposure to water, the adatoms of the (7 × 7) unit cell are still evident even after high exposures, implying that major modifications of the substrate do not occur. At high coverages, the distribution of reacted adatoms shifts from one controlled by dissociative adsorption across the adatom-rest atom pair to a statistical distribution based on the availability of dangling bonds. Desorption of the oxide layer which remains after water adsorption and the desorption of hydrogen have also been characterized. The oxide desorption occurs along well-defined wavefronts which originate at step edges and advance in directions consistent with the underlying substrate symmetry, primarily the [ 2] direction (i.e. the wave vector points in the [ 2] direction). In regions of the surface where the oxide has desorbed, the (7 × 7) unit cell can be seen clearly. Vacancies resulting from the loss of surface silicon atoms (via the etching) coalesce into islands in the clean regions of the terraces, but unlike desorption of oxide layers from Si(100), the desorption does not occur from the boundaries of these vacancy islands.  相似文献   

12.
C.D. MacPherson  D.Q. Hu  M. Doan  K.T. Leung   《Surface science》1994,310(1-3):231-242
Recently, we reported a thermal desorption study on the evolution of an intense mass 78 profile for the room-temperature exposure of cyclohexene to Si(111)7 × 7 surface, which was believed to give rise to the formation of benzene by a surface dehydrogenation reaction. Because mass 78 was also found to be the base ion in the gas-phase cracking patterns of both 1,3- and 1,4-cyclohexadiene, the dehydrogenation of cyclohexene on clean, sputtered and oxidized Si(111)7 × 7 surfaces has been re-examined in order to determine the origin of the intense mass 78 desorption profile; i.e. whether it was in fact due to the evolution of benzene or cyclohexadiene, or both. Moreover, a similar dehydrogenation reaction giving rise to toluene desorption between 350 and 600 K has been observed for the room-temperature exposure of 1-methyl-1,4-cyclohexadiene to clean and sputtered Si(111)7 × 7 surfaces. The effects of methyl substitution on the reactivity of these cyclic olefins towards Si(111)7 × 7 can be inferred from these studies. Furthermore, the catalytic activity of Si(111)7 × 7 was found to be enhanced significantly by extending the thermal desorption cycles to a higher temperature of 925 K. The dehydrogenation of these olefins on Si(111)7 × 7 also gave rise to a unique 7 × 1 low energy electron diffraction pattern. Possible factors that may play a role in any proposed model for the dehydrogenation reaction are discussed. Finally, evidence of other surface reactions including cyclohexene hydrogénation to cyclohexane will also be presented.  相似文献   

13.
J.-W. He  P.R. Norton   《Surface science》1990,230(1-3):150-158
The co-adsorption of oxygen and deuterium at 100 K on a Pd(110) surface has been studied by measurements of the change in work function (Δφ) and by thermal desorption spectroscopy (TDS). When the surface with co-adsorbed species is heated, the adsorbates O and D react to form D2O which desorbs from the surface at T > 200 K. The D2O desorption peaks shift continuously to lower temperatures as the surface D coverage (θD) increases. The maximum production of D2O is estimated to be 0.26 ML (1 ML = 9.5 × 1014 atoms cm−2), resulting from reaction in a layer containing 0.65 ML D and 0.3 ML O. The maximum work function increase caused by adsorption of D to saturation onto oxygen precovered Pd(110) decreases almost linearly with ΔφO of the oxygen precovered surface. On a surface with pre-adsorbed D however, the maximum Δφ increase contributed by oxygen adsorption decreases abruptly at ΔφD > 200 mV. This sharp change occurs at θD > 1 ML and is believed to be associated with the development of the reconstructed (1 × 2) phase of D/Pd(110).  相似文献   

14.
The Sb adsorption process on the Si(1 1 1)–In(4×1) surface phase was studied in the temperature range 200–400 °C. The formation of a Si(1 1 1)–InSb (2×2) structure was observed between 0.5 and 0.7 ML of Sb. This reconstruction decomposes when the Sb coverage approaches 1 ML and Sb atoms rearrange to and (2×1) reconstructions; released In atoms agglomerate into islands of irregular shapes. During the phase transition process from InSb(2×2) to Sb (θSb>0.7 ML), we observed the formation of a metastable (4×2) structure. Possible atomic arrangements of the InSb(2×2) and metastable (4×2) phases were discussed.  相似文献   

15.
Periodic, self-consistent, density functional theory (GGA-PW91) calculations are performed for both surface and subsurface atomic hydrogen on and in Ni(1 1 1). At a low coverage (θ=0.25 ML), the binding energies (BEs) of a hydrogen atom in surface fcc, subsurface octahedral (first layer), and subsurface octahedral (second layer) sites are −2.89, −2.18, and −2.11 eV, respectively. The activation energy barriers for hydrogen diffusion from the surface to the first subsurface layer and from the first to the second subsurface layer are estimated to be 0.88 and 0.52 eV, respectively. In the entire coverage range studied, hydrogen occupies surface fcc and hcp sites and subsurface octahedral sites. In addition, the magnitude of the BE per hydrogen atom and the magnetization of the nickel slabs both decrease as hydrogen coverage increases. Vibrational frequencies of hydrogen at various surface and subsurface sites are calculated and are in reasonable agreement with experimental data. A phase stability calculation with a 2 × 2 surface unit cell shows that a p(2 × 2)-2H overlayer structure (θ=0.5 ML) and a p(1 × 1)-1H structure (θ=1.0 ML) are stable at low hydrogen pressures, in agreement with numerous experimental results. A very large increase in pressure is required to populate subsurface sites. After such an increase occurs, the first subsurface layer is filled completely.  相似文献   

16.
The effects of adsorbed H on the Mo1−xRex(110), x=0, 0.05, 0.15, and 0.25, surfaces have been investigated using low-energy electron diffraction (LEED) and high-resolution electron energy loss spectroscopy (HREELS). For the x=0.15 alloy only, a c(2×2) LEED pattern is observed at a coverage Θ0.25 ML. A (2×2) pattern is observed for H coverages around Θ0.5 ML from surfaces with x=0, 0.05, and 0.15. Both c(2×2) and (2×2) patterns are attributed to reconstruction of the substrate. At higher coverages, a (1×1) pattern is observed. For the alloy surface with x=0.25, only a (1×1) pattern is obtained for all H coverages. Two H vibrations are observed in HREELS spectra for all Re concentrations, which shift to higher energies at intermediate coverages. Both peaks exhibit an isotopic shift, confirming their assignment to hydrogen. For Re concentrations of x=0.15 and higher, a third HREELS peak appears at 50 meV as H (D) coverage approaches saturation. This peak does not shift in energy with isotopic substitution, yet cannot be explained by contamination. The intrinsic width of the loss peaks depends on the Re concentration in the surface region and becomes broader with increasing x. This broadening can be attributed to surface inhomogeneity, but may also reflect increased delocalization of the adsorbed hydrogen atom.  相似文献   

17.
The adsorption of oxygen on the carbon-covered W(001) surface was studied by AES and low energy ion scattering. At low carbon and oxygen coverages, both species can be accommodated on the surface. At higher coverages, the oxygen displaces the carbon into the near-surface (selvedge) region. When oxygen is adsorbed on the W(001)-p(5 × 1)C surface (formed by exposure to more than 50 L of ethylene at 1500 K), carbon is displaced in a nearly one for one manner. Annealing the oxygen-covered p(5 × 1)C surface to 950 K removes up to 0.5 ML oxygen as CO. Interestingly, the surface carbon coverage is unchanged by CO desorption at 950 K and subsurface carbon is partially replenished by diffusion from the bulk. Oxygen adsorption in excess of 0.5 ML suppresses carbon segregation from the bulk at 950 K. The additional oxygen does not desorb until 1300 K. Surface carbon is restored by annealing to 1500 K.. The degree and rate of carbon segregation depend on the initial ethylene exposure even though the resulting W(001)-p(5 × 1)C surfaces are identical according to AES, LEED, and ion scattering.  相似文献   

18.
To test the model that was originally proposed for the Si(103)1 × 1-Al facets and was later on tested with STM to be correct for the Ge(103)1 × 1-In facets, in the present paper we have studied the Si(103)1 × 1-In surface by means of the QKLEED/CMTA technique. A unit cell of the model consists of an indium atom, which sits in an adatom position and forms three sp2-like bonds with bulk silicon atoms, and a surface silicon atom with a dangling bond. The model has passed the QKLEED/CMTA test and the best parameters of it have been obtained. It has been noticed in the experiment that the clean Si(103) surface has a surprisingly high thermal stability.  相似文献   

19.
The role of kinetics in the superstructure formation of the Sb/Si(0 0 1) system is studied using in situ surface sensitive techniques such as low energy electron diffraction, Auger electron spectroscopy and electron energy loss spectroscopy. Sb adsorbs epitaxially at room-temperature on a double-domain (DD) 2 × 1 reconstructed Si(0 0 1) surface at a flux rate of 0.06 ML/min. During desorption, multilayer Sb agglomerates on a stable Sb monolayer (ML) in a DD (2 × 1) phase before desorbing. The stable monolayer desorbs in the 600–850 °C temperature range, yielding DD (2 × 1), (8 × 4), c(4 × 4), DD (2 × 1) phases before retrieving the clean Si(0 0 1)-DD (2 × 1) surface. The stable 0.6-ML (8 × 4) phase here is a precursor phase to the recently reported 0.25-ML c(4 × 4) surface phase, and is reported for the first time.  相似文献   

20.
K. Kishi  A. Oka  N. Takagi  M. Nishijima  T. Aruga   《Surface science》2000,460(1-3):264-276
We have studied the growth mechanism of a Pd(100)-p(2×2)-p4g-Al surface alloy by scanning tunneling microscopy (STM). The surface alloy has a bilayer structure and is formed by annealing at 450–700 K (depending on the initial aluminum coverage) after the deposition of aluminum on Pd(100) at room temperature. The ratio of the surface-alloy coverage to the initial aluminum coverage is found to be constant (0.44) irrespective of the initial aluminum coverage from 0.5 monolayers (ML) up to 2 ML. The growth mechanism of the surface alloy is proposed on the basis of the STM measurements at various annealing temperatures. Upon annealing at 450 K, some of the surface aluminum atoms migrate into the bulk and, instead, palladium atoms come out to the surface. These palladium atoms react with aluminum atoms remaining on the surface to form a surface alloy. When the initial aluminum coverage is less than 1 ML, bilayer-high islands of the surface alloy with an average area of 100 nm2 are formed at 450–500 K, which diffuse on the terrace at 500–700 K and coalesce to form larger islands. A possible role of the percolation transition of aluminum islands in the formation of the surface alloy 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