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 453 毫秒
1.
Cluster abundance of Li n + (n≤19), Na n + (n≤25), Si n z+ (n≤8 forz=1, 3≤n≤7 forz=2), Ge n z+ (n≤11 forz=1, 3≤n≤9 forz=2,n=4 forz=3), Sn n z+ (n≤7 forz=1, 3≤n≤9 forz=2,n=4 forz=3) and Pb n z+ (n≤6 forz=1, 5≤n≤7 forz=2) ejected from a liquid metal ion source has been investigated by mass spectrometry. The abundance spectra of alkali metal clusters showed distinct maxima and steps atn=3, 7, 9, 13 and 19 for Li, and atn=3, 5, 11, 13 and 19 for Na. Mass spectra of Si, Ge and Sn clusters were very similar each other, showing intensity drops aftern=4 and 6 (and alson=10 for Ge) for singly charged clusters. The magic numbers observed are discussed in terms of stability of charged clusters.  相似文献   

2.
Experimental study of the kinetics of mechanosynthesis of TlCl nanoparticles in the reaction 2NaCl + Tl2SO4 + zNa2SO4 = = (z + 1)Na2SO4 + 2TlCl with z = z*1 = z* = 11.25 and comparison of kinetic parameters for this reaction with those determined theoretically for the model reaction KBr + TlCl + zKCl = (z + 1)KCl + TlBr with z = z*1 = 13.5 made it possible to estimate the mass transfer (mixing of particles) coefficient in a mechanochemical reactor by the mobile milling tools.  相似文献   

3.
Seeded beam expansions of high-temperature sodium vapour in helium, neon, argon, krypton and nitrogen carrier gases are examined. Photoionization coupled with mass spectroscopic detection is used to determine beam constituents. Contrary to non-seeded supersonic expansions, in which the abundance of sodium clusters observed decreases exponentially with increasing cluster size, large abundances of higher clusters (Nax, x ?65) are observed. Maxima centered at m/z = 161, m/z = 437 and m/z = 874 occur under certain conditions.  相似文献   

4.
Positive and negative ion electrospray mass spectra obtained from 50 mM phosphoric acid solutions presented a large number of phosphoric acid clusters: [(H3PO4)n?+?zH] z+ or [(H3PO4)n – zH] z– , with n up to 200 and z up to 4 for positively charged clusters, and n up to 270 and z up to 7 for negatively charged cluster ions. Ion mobility experiments allowed very explicit separation of the different charge states. Because of the increased pressures involved in ion mobility experiments, dissociation to smaller clusters was observed both in the trap and transfer areas. Voltages along the ion path could be optimized so as to minimize this effect, which can be directly associated with the cleavage of hydrogen bonds. Having excluded the ion mobility times that resulted from dissociated ions, each cluster ion appeared at a single drift time. These drift times showed a linear progression with the number of phosphoric atoms for cluster ions of the same charge state. Cross section calculations were carried out with MOBCAL on DFT optimized geometries with different hydrogen locations and with three types of atomic charges. DFT geometry optimizations yielded roughly spherical structures. Our results for nitrogen gas interaction cross sections showed that values were dependent on the atomic charges definition used in the MOBCAL calculation. This pinpointed the necessity to define a clear theoretical framework before any comparative interpretations can be attempted with uncharacterized compounds.
Figure
?  相似文献   

5.
Donnan dialysis with an ion exchange membrane was investigated for ions of different valence. The effective diffusion coefficients (De) of various kinds of ions in the membrane were obtained by fitting of the equation derived from the Nernst–Planck equation to three or more sets of experimental data for Donnan dialysis. It became apparent that the value of De/Ds of monovalent ions (e.g., K+ or Na+ ions) at zA=1 and zB=2 (feed ions are monovalent ones and driving ions are bivalent ones) remained constant at ca. 1/210 and that of bivalent ions (e.g., Ca2+, Cu2+, or Mg2+ ions) remained constant at ca. 1/526 where Ds denotes the diffusion coefficient of ions at infinite dilution in water calculated from the Nernst–Einstein equation, and zA and zB represent the valences of the feed and driving ions, respectively. De/Ds of monovalent ions (e.g., H+, K+, or Na+ ions) at zA=2 and zB=1 (feed ions are bivalent ones and driving ions are monovalent ones) was constant at ca. 1/23.3 and that of bivalent ions remained constant at ca. 1/58.4. It was proved that De/D using De at zA=1 and zB=2 was constant at 1/3.0 and that at zA=2 and zB=1 remained constant at 3.0 where D represents the diffusion coefficient of ions in the membrane at zA=zB (the valences of both feed and driving ions are equal). Therefore, it was found that a large flux of ions could be obtained using the monovalent driving ions in Donnan dialysis. On the other hand, the small flux can be obtained using bi- or higher-valent driving ions.  相似文献   

6.
A differential mobility analyzer (DMA) is used in atmospheric pressure N2 to select a narrow range of electrical mobilities from a complex mix of cluster ions of composition (CA)n(C+)z. The clusters are introduced into the N2 gas by electrospraying concentrated (~20 mM) acetonitrile solutions of ionic liquids (molten salts) of composition CA (C+ = cation, A? = anion). Mass analysis of these mobility-selected ions reveals the occurrence of individual neutral ion-pair evaporation events from the smallest singly charged clusters: (CA)nC+→(CA)n? 1C++CA. Although bulk ionic liquids are effectively involatile at room temperature, up to six sequential evaporation events are observed. Because this requires far more internal energy than available in the original clusters, substantial heating (~10 eV) must take place in the ion guides leading to the mass analyzer. The observed increase in IL evaporation rate with decreasing size is drastic, in qualitative agreement with the exponential vapor pressure dependence predicted by Kelvin’s formula. A single evaporation event is barely detectable at n = 13, while two or more are prominent for n ≤ 9. Magic number clusters (CA)4C+ with singularly low volatilities are found in three of the four ionic liquids studied. Like their recently reported liquid phase prenucleation cluster analogs, these magic number clusters could play a key role as gas-phase nucleation seeds. All the singularly involatile clusters seen are cations, which may help understand commonly observed sign effects in ion-induced nucleation. No other charge-sign asymmetry is seen on cluster evaporation.  相似文献   

7.
(Bi1.90Eu0.10)(V1−zMoz)O5.5 (z = 0, 0.05, 0.10, 0.15 and 0.20) thin films with c-axis oriented were prepared on Pt(111)/Ti/SiO2/Si substrates by using chemical solution deposition method. The effect of Mo6+ concentration on the structure, luminescence properties and dielectric properties of the thin films were characterized systematically. X-ray diffraction data indicates that the thin films with low Mo6+-doping content can remain Bi2VO5.5 structure. When the Mo6+-doping content z reaches to 0.15, the thin films are a mixture of diphase with the main phase Bi2VO5.5 and secondary phase Bi2MoO6. Under UV irradiation, all the thin films emit a bright red or orange emission which origin from Eu3+. With increasing Mo6+-doping content z, the relative intensity of the Red and Orange emissions show obviously change. The value of Red/Orange ratio first decrease, and it reached minimum when z is 0.15, then it recover to the initial value. The variation trend of the Red/Orange ratio reflects the change of the lattice symmetry. Dielectric constant of the thin films increased with the increasing of the Mo6+ concentration while dielectric loss decreased. The decrease of the quantities of oxygen vacancies and the generation of Bi2MoO6 phase are responsible for the improvement of electric properties. These results explain that Eu3+ion can be used as an effective luminescent probe in (Bi1.90Eu0.10)(V1−zMoz)O5.5 (z = 0, 0.05, 0.10, 0.15 and 0.20) thin films, and the electric properties of the thin films can be improved by Mo6+ doping.  相似文献   

8.
A method of calculation of the number of atoms in the structure of nanoforms with pentagonal symmetry (fullerenes, nanoparticles, clusters) depending on the arrangement of atoms on the symmetry elements of the I h group has been developed. The formulas for calculation of the number of particles in all possible shells, including multilayer ones, are reported. The numbers of atoms in the shells of pentagonal symmetry are determined by four structurally invariant numbers and the “quantum number” of the order n of the group. The classification of all possible atomic shells S ? + 60z (z = 0, 1, ...) is presented, and the constructions of the basic shells S ? (? = 12, 20, 30, 50, 60) are given. For each basic shell, the sum rule is met: the sum of the coordination numbers of the elements of subshells is equal to 60. In clusters with magic numbers, basic shells are periodically repeated. In addition to the known shells of nanostructures, the formulas of new structures that are expected to be stable—B20O30, B60O90 (B2O3), and B90O 130 10+ (borate)—are reported for the first time. The same is valid for similar compounds of Group III elements.  相似文献   

9.
The structure of wüstite Fe1?zO is studied by neutron diffraction on one polycrystalline sample under equilibrium conditions at high temperature. The mean isotrope temperature factor B is expressed as a sum of two parts, BTh and BSt, which vary linearly with a single parameter, respectively temperature and z. A classification is established for clusters (mn) settled from m vacancies in octahedral sites and n FeIII ions in interstitial sites. Sixteen values have been experimentally determined for the vacancies to interstitials ratio ? = (z + t)t = mn. A constant value of ?, which is lower than 3, is observed. This result characterizes the short-range order. It eliminates several possibilities of clusters like those obeying the relation ? = (1 + 3n)n. Other clusters, namely (166) or (4014), might agree. The (83) and (94) clusters obtained from (41) clusters joined by an edge would be the more likely. An analysis of diffuse scattering eliminates the hypothesis of large domains with inverse spinel structure. The structural differences between the three varieties W1, W2 and W3 would not be found in a structural change of clusters.  相似文献   

10.
TheJ-dependence of the isotope shift in the terma 11 F of 4f 7 5d 2 6s between six stable Gd isotopes was found to be represented by these parameter values (in MHz):z 5d (160?158)=19.4(1.8),z 5d (160?157)=37.2(1.0),z 5d (160?156)=42.4(1.7),z 5d (160?155)=49.1(2.0),z 5d (160?154)=59.0(2.0). The normalization with the corresponding changes in the mean-square nuclear charge-radiiδr 2〉 yields values which are almost constant, mean value:z 5d /δr 2〉=134(14) MHZ/fm2. This indicates a second order IS interaction of the magnetic and the field shift operator.  相似文献   

11.
The reactions of IMes‐HCl with bimetallic iron‐copper clusters were investigated and it showed that interconversions among bimetallic FexCuy(CO)zn clusters were induced by IMes‐HCl. The reaction priority of IMes‐HCl with fragments in FexCuy(CO)zn was also investigated and the reaction behaviors of zero valent Fe(CO)4 and divalent Fe(CO)4 with IMes‐HCl were discussed.  相似文献   

12.
The kinetics of the solid-state mechanochemical synthesis of the nanosized product (TlCl) in the reaction 2NaCl + Tl2SO4 + zNa2SO4 = (z + 1)Na2SO4 + 2TlCl was studied experimentlaly. The method used was based on the dilution of the initial mixture of powdered reagents (2NaCl + Tl2SO4) with another exchange reaction product (Na2SO4) at the optimum theoretically estimated z value, z = z* = 11.25. Several special features of the development of this reaction were established. The parameters of the kinetic curve obtained for the mechanochemical synthesis of the desired product were compared with those of the kinetic curve determined theoretically for the model reaction KBr + TlCl + zKCl = (z + 1)KCl + TlBr with z = z 1 * = 13.5. This allowed us to experimentally estimate the mass transfer coefficient in a mechanochemical reactor by mobile milling bodies. This estimate was obtained for the first time. The dynamics of changes in the size of desired product nanoparticles depending on the time of mechanochemical activation in an AGO-2 ball planetary mill was studied.  相似文献   

13.
A study on the interactions between CH3Hg+ and some S, N and O donor ligands (2-mercaptopropanoic acid (thiolactic acid (H2 TLA)), 3-mercaptopropanoic acid (H2 MPA), 2-mercaptosuccinic acid (thiomalic acid (H3 TMA)), d,l-penicillamine (H2 PSH), l-cysteine (H2 CYS), glutathione (H3 GSH), N,N′-bis(3-aminopropyl)-1-4-diaminobutane (spermine (SPER)), 1,2,3,4,5,6-benzenehexacarboxylic acid (mellitic acid (H6 MLT)) and ethylenediaminetetraacetic acid (H4 EDTA)) is reported. The speciation models in aqueous solution and the possible structures of the complexes formed are discussed on the basis of potentiometric, calorimetric, UV spectrophotometric and electrospray mass spectrometric results. For the CH3Hg+–S donor ligand systems, the formation of ML1–z and MLH2–z complex species is observed, together with a diprotonated MLH 2 3–z species for CYS 2?, PSH 2? and GSH 3? and the mixed hydrolytic one ML(OH)?z for TLA 2? and MPA 2?. The dependence of the stability on ionic strength and on temperature is also analysed. In the other CH3Hg+-L systems (L?=?MLT 6?, SPER and EDTA 4?), ML1–z , MLH2–z and MLH2 3–z complex species are formed, together with the MLH3 4–z species for SPER, the mixed hydrolytic ML(OH)–z one for SPER and EDTA, and the M2L2–z for EDTA only. On the basis of the speciation models proposed, the sequestering ability of the ligands towards methylmercury(II) cation is evaluated. All S donor ligands show a good sequestering power (at 10?11?mol?L?1 level, in the pH range 4 to 8) following the trend MPA 2??<?PSH 2??<?GSH 3??<?TLA 2??<?CYS 2??<?TMA 3?, while significantly lower is the sequestering ability of MLT, SPER and EDTA (at 10?3–10?5?mol?L?1 level, in the pH range 4 to 8).
Figure
Sum of fractions of CH3Hg+-L z species (L?=?S, O and N donor ligands vs. pL  相似文献   

14.
A sensitive method for the determination of free fatty acids using 1,2-benzo-carbazole-9-ethyl-p-toluenesulfonate (BCETS) as tagging reagent with fluorescence detection has been developed. BCETS could easily and quickly label fatty acids in the presence of the K2CO3 catalyst at 80 °C for 30 min in N,N-dimethylformamide solvent. In this study, fatty acids from the extracted Potentilla anserina L. plant sample were sensitively determined. The corresponding derivatives were separated on a reversed-phase Eclipse XDB-C8 column by LC in conjunction with gradient elution. The identification was carried out by post-column APCI-MS in positive-ion detection mode. BCETS-fatty acid derivatives gave an intense molecular ion peak at m/z [M+H]+, the collision-induced dissociation spectra of m/z [M+H]+ produced the specific fragment ions at m/z [M′+CH2CH2]+, m/z 216.6 and m/z [MH?H2O]+ (here, M′: corresponding molecular mass of the fatty acids). The fluorescence excitation and emission wavelengths of the derivatives were at λ ex 279 nm and λ em 380 nm, respectively. Linear correlation coefficients for all fatty acid derivatives are more than 0.9994. Detection limits, at a signal-to-noise ratio of 3:1, are 10.79–34.19 fmol for the labeled fatty acids.  相似文献   

15.
Measurements of the equilibrium oxygen content, electrical conductivity and thermopower in the perovskite-like solid solution La0.7Sr0.3Co1-zMnzO3−δ (z=0 and 0.25) as a function of the temperature and oxygen partial pressure are used to determine the temperature dependence of the conductivity and thermopower at different values of the oxygen deficiency. A model for a hopping conductor with screened charge disproportionation is applied for the data analysis in combination with trapping reactions of n- and p-type carriers on local oxygen vacancy clusters and manganese cations, respectively. Changes in the ratio of n-type to p-type mobility are due to variations in oxygen vacancy concentration and manganese content, while the energetic parameters governing charge disproportionation of the trivalent cobalt cations and formation of vacancy associates are shown to be essentially invariable. These calculated charge carrier site occupancies are used to model temperature variations of the electrical properties in La0.7Sr0.3Co1−zMnzO3−δ in favorable correspondence with experimental observations.  相似文献   

16.
The hyperfine structure of the53Cr resonance linesa 7 S 3 ?z 7 P 2,3,4 has been investigated by means of laser saturation spectroscopy. By comparison of the experimental signal curves with theoretically computed spectra the hitherto unknown sign of the magnetic hyperfine coupling constant in thea 7 S 3 ground state of53Cr could be determined unambigiously to be negative. Further the signs of the hfs coupling constants in thez 7 P states — so far only evaluated by theoretical reasoning — could be confirmed. Additionally the lifetimes of the statesz 7 P,z 5 P,f 7 D,z 5 F,e 7 D 5 andy 5 P 3 in the Cr I spectrum have been determined from the fluorescence decay after pulsed laser excitation.  相似文献   

17.
The higher order fields present in the quadrupole ion trap may have beneficial effects such as increases in mass resolution in the mass-selective instability or resonance ejection modes of operation, but may also result in losses of ions due to nonlinear resonances. In this work, the reduction in ion intensities observed in the mass spectra of polyethylene glycol (PEG 1000) has been utilized to monitor the ion losses resulting from these higher order fields during the rf voltage scans in both the forward and reverse directions. Extensive ion losses were observed in reverse rf voltage scans at q z=0.64 (a z=0), which corresponds to octopole resonance at β z=1/2. The losses depended upon rf voltage scan rate and ion mass being greater for lower scan rates and lower masses. For ions of m/z 877, losses of up to 60% of the stored ions were observed at low scan rates (<1×104 Da/s), but were minimal at higher scan rates. Thus, it is possible to avoid such losses during reverse scans by scanning the region q z=0.64 at rates in excess of 4×104 Da/s. In forward rf voltage scans, ion storage was considerably more reliable, with significant losses observed only at very high scan rates near the region q z=0.78 (hexapole resonance at β z=2/3).  相似文献   

18.
Paramagnetic absorption of Mo5+ has been studied in a polycrystalline TiO2 rutile lattice, The g tensor (gx = 1.897, gy = 1.920, gz = 1.857) and the hyperfine tensor (Ax = 32.7, (Ay = 51.2, (Az = 58.5 (in 10?4 cm?1)) are in agreement with those expected for an nd1 ion in an interstitial position.  相似文献   

19.
A new vanado-molybdate LiMg3VMo2O12 has been synthesized, the crystal structure determined an ionic conductivity measured. The solid solution Li2−zMg2+zVzMo3−zO12 was investigated and the structures of the z=0.5 and 1.0 compositions were refined by Rietveld analysis of powder X-ray (XRD) and powder neutron diffraction (ND) data. The structures were refined in the orthorhombic space group Pnma with a∼5.10, b∼10.4 and c∼17.6 Å, and are isostructural with the previously reported double molybdates Li2M2(MoO4)3 (M=M2+, z=0). The structures comprise of two unique (Li/Mg)O6 octahedra, (Li/Mg)O6 trigonal prisms and two unique (Mo/V)O4 tetrahedra. A well-defined 1:3 ratio of Li+:Mg2+ is observed in octahedral chains for LiMg3VMo2O12. Li+ preferentially occupies trigonal prisms and Mg2+ favours octahedral sheets. Excess V5+ adjacent to the octahedral sheets may indicate short-range order. Ionic conductivity measured by impedance spectroscopy (IS) and differential scanning calorimetry (DSC) measurements show the presence of a phase transition, at 500-600 °C, depending on x. A decrease in activation energy for Li+ ion conductivity occurs at the phase transition and the high temperature structure is a good Li+ ion conductor, with σ=1×10−3-4×10−2 S cm−1 and Ea=0.6 to 0.8 eV.  相似文献   

20.
The zero-point average structures of acetyl chloride and acetyl bromide have been determined by the combined use of their moments of inertia and average distances, obtained by means of microwave spectroscopy and electron diffraction. The rz parameters determined are as follows: rz(CO) = 1.185 ± 0.003 Å, rz(C-Cl) = 1.796 ± 0.002 Å, rz(C-C) = 1.505 ± 0.003 Å, rz(C-H) = 1.092 ± 0.005 Å, φz(OCCl) = 121.2 ± 0.6°, φz(CCCl) = 111.6 ± 0.6°, φz(HCH) = 108.8 ± 0.8° and tilt(CH3) = 1.3 ± 1.0°, for chloride; rz(CO) = 1.181 ± 0.003 Å, rz(C-Br) = 1.974 ± 0.003 Å, rz(C-C) = 1.516 ± 0.003 Å, φz(OCBr) = 122.3 ± 1.5°, φz(CCBr) = 111.0 ± 1.5°, φz(HCH) = 109.9 ± 1.1°, tilt(CH3) = 1.9 ± 1.0°, for bromide. The barriers V3 to internal rotation have been revised to 1260 and 1256 cal mol−1 for the chloride and bromide, respectively.  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