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 62 毫秒
1.
Mass-selected antimony cluster ions Sb n + (n = 3-12) and bismuth cluster ions Bi {ntn} + (n = 3-8) are allowed to collide with the surface of highly oriented pyrolytic graphite at energies up to 350 eV. The resulting fragment ions are analysed in a time-of-flight mass spectrometer. Two main fragmentation channels can be identified. At low impact energies both Sb n + and Bi n + cluster ions lose neutral tetramer and dimer units upon collision. Above about 150 eV impact energy Sb 3 + becomes the predominant fragment ion of all investigated antimony clusters. The enhanced stability of these fragment clusters can be explained in the framework of the polyhedral skeletal electron pair theory. In contrast, Bi n + cluster scattering leads to the formation of Bi 3 + , Bi 2 + and Bi+ with nearly equal abundances, if the collision energy exceeds 75 eV. The integral scattering yield is substantially higher in this case as compared to Sb n + clusters.  相似文献   

2.
Reaction dynamics of sodium cluster ions, Na n + (n = 2–9), in collision with molecular oxygen, O2 was investigated by measuring the absolute dissociation cross sections and the branching fractions by using a tandem mass spectrometer equipped with several octapole ion guides. The mass spectrum of the product ions show that the dominant reaction channels are production of oxide ions, NakOi (i =1, 2), and intact ions, Na p + (p < n). With increase in the collision energy, the cross section for the production of the oxide ions decreased, while that for the production of the intact ions increased. The collision-energy dependences of the cross section for the oxide formation reveals that electron harpooning from the molecule to Na n + preludes the oxideion formation. On the other hand, the collision-energy dependences of the cross sections for the intact ion formation is explained by a hard-sphere-collision model similar to the collisional dissociation of Na n + by rare-gas impact.  相似文献   

3.
Fission of doubly charged silver clusters is investigated by the method of shell corrections. The following fission events are considered: Ag 22 2+ → Ag n + + Ag 22 ?n + , (n=11, 10, 9, 8); Ag 21 2+ → Ag n + + Ag 21 ?n + , (n=10, 9, 8, 7); Ag 18 2+ → Ag n + + Ag 18 ?n + , (n=9, 8, 7, 6). It is found that the shell correction energy is comparable to or larger than the deformation energy of the liquid drop. Threshold energies for the fission events are calculated and compared with the experimental abundance spectra obtained by Katakuse et al. (1990). Correspondence between the calculated threshold energies with the shell corrections and the experimental abundance is very good, showing products from lower threshold fission channels yield more abundance. The threshold energies without the shell corrections are almost constant irrespective of the fission channels and cannot explain the experimental abundance. Abundance of some products are too small to be accounted for only by the threshold energies. The low abundance of those products may be explained by the presence of competing fission channels that have similar minimal energy paths. It is found in fission of Ag 18 2+ that the shell correction overwhelms the Coulomb energy and the fission channel to Ag8 + Ag 10 2+ is preferred over the fission channel to Ag 8 + + Ag 10 + .  相似文献   

4.
The results of a detailed study of the photodissociation of carbon cluster ions, C 3 + to C 20 + , are presented and discussed. The experiments were performed using internally cold cluster ions derived from pulsed laser evaporation of a graphite target rod in a helium buffer gas followed by supersonic expansion. The mass selected clusters were photodissociated using 248 nm and 351 nm light from an excimer laser. Photofragment branching ratios, photodissociation cross sections and data on the laser fluence dependence of photodissociation are reported. For almost all initial clusters, C n + , the dominant photodissociation pathway was observed to be loss of a C3 unit to give a C n?3 + ion. This observation is interpreted as indicating that dissociation occurs by a statistical unimolecular process rather than by direct photodissociation. The photodissociation was found to be linear with laser fluence forn>5 with 248 nm and 351 nm light; quadratic forn=5 for 248 nm and 351 nm; and linear forn=4 at 248 nm. Dissociation energies for the carbon cluster ions implied by these results are discussed. The photodissociation cross sections were found to change dramatically with cluster size and with the wavelength of the photodissociating light.  相似文献   

5.
The low-energy dissociation channels of mass selected silver cluster ions Ag n 2+ (n = 9–24) are determined by collision induced dissociation (CID) in a Penning trap. While all clusters of the size n ≥ 17 evaporate neutral monomers, most smaller clusters undergo asymmetric fission of the form Ag n 2+ Ag n?3 + + Ag {3} + . However, Ag 15 2+ and Ag 11 2+ emit monomers which indicates shell or odd-even effects. The observed fragmentation pathways are different from previous reports of measurements with sputtered Ag n 2+ .  相似文献   

6.
Both positive and negative phosphorus cluster ions were generated from the laser ablation of a red phosphorus sample. The mass distribution of phosphorus cluster ions was found to be very sensitive to the power density of the ablation laser. The P 7 + species exhibits the highest signal intensity in the recorded mass spectra of bare phosphorus cluster cations, as does P 5 - among the anions. Their special structural stability can be attributed to their planar configuration and their aromatic character. As the phosphorus cluster size increases, the odd/even alternation of the signal intensity becomes more pronounced. For the P n + species with n > 24, the relative abundance varies in the order of 8 and P n + with n = 8k + 1 (k = 3–11) are more intense than their neighbors. For comparison, some binary phosphide cluster ions, including CnP m - , SinP m - , BnP m + and AlnP m + , were produced as well. The mass distribution of binary phosphide cluster ions changes with different components. From analysis of the recorded mass spectra of the phosphide cluster ions, the larger clusters may be in a polyhedral configuration and tend to have all valence electrons paired.  相似文献   

7.
Quantum chemical ab initio calculations have been performed for the vertical excitation energies and oscillator strengths of all low-lying electronically excited states of small helium cluster ions, He n + ,n=2, ..., 7. The geometrical structures of the ions were fixed at the equilibrium geometries of the respective ground states, for He 4 + and He 5 + also one alternative structure was considered. The low-lying excited states can be classified into two categories: the electronic transition can occur either within the central He 2 + or He 3 + unit or from the peripheral weakly bound He atoms to this unit. The latter transitions are very weak (f≈0.001), closely spaced, with vertical excitation energies of about 5.7 eV. The He 2 + and He 3 + units have strong transitions at 9.93 and 5.55 eV, respectively; these transitions are only slightly blue-shifted if He 2 + or He 3 + are placed as “chromophores” into the centre of a larger He n + cluster. The large difference in the vertical excitation energy of the strong transition should enable an experimental decision of the question whether the cluster ions have He 2 + or He 3 + cores.  相似文献   

8.
Optical absorption spectra of cobalt cluster ions, Co n + , and vanadium cluster ions, V n + , were analyzed by a theoretical calculation based on the spin-polarized DV- method, and their electronic and geometric structures were obtained. Relative absorption cross section associated with each electronic transition was calculated; the calculation enables a qualitative comparison of calculated spectrum with a measured one not only in its transition energy but also in its intensity profile. This analysis shows that Co 4 + , Co 3 + , and V 4 + have, respectively, a tetrahedral structure with a bond distance of 2.00Å, an equilateral triangle with a bond distance of 2.30Å, and a distorted tetrahedral structure with five bonds having a distance of 2.34 Å and one of 2.89Å. The differences in the population between majority and minority spins (spin-difference) evaluated from the electronic structure thus obtained were 2.0, 1.7, and zero per atom in Co 3 + , Co 4 + , and V 4 + , respectively. These spin differences indicate a ferromagnetic and an antiferromagnetic spin-coupling in the cobalt and vanadium cluster ions, respectively.  相似文献   

9.
The production of H 3 + ions resulting from single collisions of mass-selected ionic hydrogen clusters, H n + (n=9, 25, 31), with helium at high velocity (1.55 times the Bohr velocity) has been studied. A strong double H 3 + ion production resulting from one incident cluster is observed. Moreover, evidence for a triple H 3 + fragment production is presented forn=25 and 31. Thus, in this energy range, the collision gives rise to multifragmentation processes. The formation of H 3 + ions takes place in the fragmentation of the multicharged cluster resulting from the collision.  相似文献   

10.
Reactivity of positively charged cobalt cluster ions (Co n + ,n=2?22), produce by laser vaporization, with various gas samples (CH4, N2, H2, C2H4, and C2H2) were systematically investigated by using a fast-flow reactor. The reactivity of Co n + with the various gas samples is qualitatively consistent with the adsorption rate of the gas to cobalt metal surfaces. Co n + highly reacts with C2H2 as characterized by the adsorption rate to metal surfaces, and it indicates no size dependence. In contrast, the reactions of Co n + with the other gas samples indicate a similar cluster size dependence; atn=4, 5, and 10?15, Co n + highly reacts. The difference can be explained by the amount of the activation energy for chemisorption reaction. Compared with neutral cobalt clusters, the size dependence is almost similar except for Co 4 + and Co 5 + . The reactivity enhancement of Co 4 + and Co 5 + indicates that the cobalt cluster ions are presumed to have an active site for chemisorption atn=4 and 5, induced by the influence of positive charge.  相似文献   

11.
The potential energy curves of the Na 19 + /Na quasimolecular (Jellium—Atom)+ system determined by Guissani and Sidis [3] are used to calculate the vibrational energy levels of a sodium atom stuck to a Na 19 + Jellium-like cluster and to determine excitation and charge transfer cross sections in Na 19 + + Na collisions in the 0.2 107 cm/s < v < 1.5 107 cm/s velocity range.  相似文献   

12.
Cluster ions are produced by ion bombardment of thick metal targets and mass selected in a Wien filter. The unimolecular decomposition of Al n + , Cu n + , Mo n + , W n + , and Pb n + is investigated under UHV conditions. The time evolution of the decay allows a glimpse into the cluster formation/fragmentation process. Highly excited metal cluster ions decompose mainly by evaporating single neutral atoms with rates reaching 100%. The collision induced fragmentation (CIF) of stable mass selected metal cluster ions in a low pressure Ar and O2 gas target will be compared to the unimolecular decay.  相似文献   

13.
In a previous work the equilibrium geometrical and electronic structures of Xe n + clusters had been established using a non-empirical model hamiltonian. The same model is used to determine the energetic barriers between the nearly degenerate isomers; the movement of the neutral atoms around the Xe 3 + or Xe 4 + ionized linear cores are quite easy (ΔE?0.9 kcal/mole), the changes from a Xe 3 + to a Xe 4 + core are more difficult (ΔE?2.0 kcal/mole). The energetically possible fissions from a vertical photoionization \(Xe_n \xrightarrow{{h v}}Xe_n^{v + } \to Xe_p^ + + Xe_{n - p} \) forn≦19,p=1–9 and 12–14 and mass exchanges Xe p + +Xe q →Xe p+m + +Xe q?m (m=1,2,3) from relaxed Xe p + clusters are given forp+m≦9 and 12–14 andq≦19. Surprisingly the reverse reactions are shown to occur for some values ofp andq. Numerous processes lead to Xe 13 + , which is especially stable.  相似文献   

14.
SCF and CEPA calculations are applied to study the structure of small He cluster ions, He n + ,n=2, 3, 4, 5 and some low-lying Rydberg states of He4. The effect of electron correlation upon the equilibrium structures and binding energies is discussed. He 3 + has a linear symmetric equilibrium geometry with a bond length of 2.35a 0 and a binding energyD e =0.165 eV with respect to He 2 + +He (experimentally:D 0=0.17 eV which corresponds toD e ≈0.20 eV). He 4 + is a very floppy molecular ion with several energetically very similar geometrical configurations. Our CEPA calculations yield a T-shaped form with a He 3 + centre (R e = 2.35a 0) and one inductively bound He atom (4.39a 0 from the central He atom of He 3 + ) as equilibrium structure. Its binding energy with respect to He 3 + +He is 0.031 eV. A linear symmetric configuration consisting of a He 2 + centre with a bond length of 2.10a 0 and two inductively bound He atoms (4.20a 0 from the centre of He 2 + ) is only 0.02–0.03 eV higher in energy. We expect that in larger He cluster ions structures with He 2 + and He 3 + centres andn?2 orn?3 inductively bound He atoms have nearly the same energies. In He4 a low-lying metastable Rydberg state (3 Π symmetry for linear He 4 * ,3 B 1 for the T-shaped form) exists which is slightly stronger bound with respect to He 3 * +He than the corresponding ion.  相似文献   

15.
Cross sections for the production of O 2 ? in charge transfer collisions of fast molecular hydrogen ions (H 2 + , D 2 + , H 3 + , and D 3 + of 10 to 140 keV kinetic energy) with O2 molecules have been determined by means of a time-of-flight mass spectrometer analysing the slow negative product ions from the collisions. Within the measuring accuracy equivelocity H 2 + and D 2 + ions have the same cross sections for the generation of O 2 ? . The projectile velocity dependence curve of the cross section passes through a broad maximum with a peak value of about 6.5×10?18 cm2 around the Bohr velocity (25 keV/u) before showing an asymptotic decrease still within the limited energy range under investigation that is in inverse proportion to the square of velocity. Throughout the examined energy range H 3 + ions yield a cross section which is about 1.4 times larger than that of H 2 + ions of the same velocity. The fragment ion O? has been found to appear with cross sections between 10?19 and 10?18 cm2 upon collisional excitation in the energy range under investigation, with ever decreasing intensity when the energy of the positive hydrogen ions, the proton included, was increased.  相似文献   

16.
The hydrogen-bonded (N2H4) n clusters and the van der Waals (OCS) n clusters are size selected in a scattering experiment with a He beam up to the cluster sizen=6. By measuring the angular distributions of the scattered clusters the complete fragmentation pattern of electron impact ionization is obtained. For Hydrazine the two main fragment masses are the protonated species (N2H4) n?1H+ and with somewhat weaker intensities also the nominal ion mass (N2H4) n + . The largest intensity is observed for the monomer ion N2H 4 + to which clusters up ton=5 fragment. For carbonylsulfide, completely different results are obtained. Aside from the fragments of the OCS monomer and the van der Waals cluster fragments (OCS) 2 + and (OCS) 3 + signals at mass S 2 + , S 3 + and S2OCS+ are detected. This indicates a fast chemical reaction in the cluster according to: S + OCS → CO + S2 which occurs for clusters of sizen ≥ 2. Peaks at S 3 + and S2OCS+ are seen for the first time forn ≥ 5 according to a further reaction of S2 in the cluster.  相似文献   

17.
A minimum-basis diatomics-in-molecules (DIM) model previously developed for singly-ionized argon clusters is applied to neon clusters, Ne n + , forn=3, 4,...,22. A search for the global minimum energy of each cluster yields structures with the positive charge localised on a dimer-ion. This appears to be due largely to the linear unsymmetrical configuration which the model finds for Ne 3 + . For this reason, the structures of the clusters at their minimum energy are different from those for Ar n + computed with the same model. On the other hand, the behaviour of the charge distribution as a function of the geometrical configuration is similar to that for Ar n + , as are the overall shapes of the potential energy surfaces. The results are discussed in terms of the charge distributions and the ratios of equilibrium properties of the dimers and dimer-ions which constitute the input to the model.  相似文献   

18.
Photoelectrons from mass-identified jet-cooled tin and lead cluster anions (Sn n ? , Pb n ? ) are detached by ultraviolet laser light (=3.68 eV). The photoelectron energy spectra give the detachment energies of ground state cluster anions (electron affinities) as well as excitation energies of neutral clusters in the geometry of the anions. The energy spectra for Sn n ? are dominated by flat thresholds with ann-dependence similar to that of other group IV clusters. In contrast, for Pb n ? we find pronounced narrow lines close to threshold, generally followed by a 0.3–1.4 eV gap which indicates closed-shell behaviour of Pb n ? for nearly alln.  相似文献   

19.
The previously measured relative cross section function for electron impact ionization (EII) of neutral Ag2 has now been calibrated quantitatively by combining the electron impact ionization with in situ non resonant two photon ionization (NR2PI). By comparing the NR2PI saturation intensities measured for Ag 2 + and Ag+ with the corresponding EII intensities, the ratio between the electron impact ionization cross sections (EIICS) of neutral Ag2 and Ag was determined to be σAg2Ag=1.53 for an electron energy of 46 eV. This result agrees well with the geometricn 2/3-rule \((\sigma X_n \sim n^{2/3} )\) commonly proposed for the dependence of the EIICS of clustersX n on the cluster sizen.  相似文献   

20.
The formation of cluster ions when hydrogen molecular ions H 2 + and H 3 + are injected into a drift tube filled with helium gas at 4.4 K has been investigated. When H 2 + ions are injected, cluster ions HHe x + (x≦14) are produced. No production of H2He x + ions is observed. When H 3 + ions are injected, cluster ions HHe x + (x≦14) are produced as well as H3He x + (x≦13), and very small signals corresponding to H2He x + (3≦x≦10) are observed. Information on the stability of HHe x + and H3He x + is derived from the drift field dependence of the cluster size distributions. The cluster sizex=13 is found to be a magic number for HHe x + , and for H3He x + ,x=10 and 11.  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