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.
Dissociative and nondissociative electron attachment in the electron impact energy range 0–14 eV are reported for SOF2 SOF4, SO2F2, SF4, SO2, and SiF4 compounds which can be formed by electrical discharges in SF6. The electron energy dependences of the mass-identified negative ions were determined in a time-of-flight mass spectrometer. The ions studied include F and SOF 2 –* from SOF2; SOF 3 and F from SOF4; SO2F 2 –* , SO2F, F 2 , and F from SO2F2; SF 4 –* and F from SF4; O, SO, and S from SO2; and SiF 3 and F from SiF4. Thermochemical data have been determined from the threshold energies of some of the fragment negative ions. Lifetimes of the anions SOF 2 –* , SO2F 2 –* , and SF 4 –* are also reported.  相似文献   

2.
By-product formation in spark breakdown of SF6/O2 mixtures   总被引:2,自引:0,他引:2  
The yields of SOF4, SO2F2, SOF2, and SO2 have been measured as a function of O2 content in SF6/O2 mixtures, following spark discharges. All experiments were made at a spark energy of 8.7 J/spark, a total pressure of 133 kPa, and for O2 additions of 0, 1, 2, 5, 10, and 20% to SF6. Even for the case of no added O2, trace amounts of O2 and H2O result in the formation of the above by-products. However, addition of O2 significantly increases the yields of SOF4 and SO2F2, while SOF2 is only slightly affected. The net yields for SOF4 and SO2F2 formation range from 0.18×10–9 and 0.64×10–10 mol·J–1, respectively, at 1% O2 content to 10.45×10–9 and 7.15×10–10 mol·J–1, respectively, at 20% O2 content. The mechanism for SOF4 production appears to involve SF4, an important initial product of SF6, as a precursor. Comparison of the SOF4 and SO2F2 yield from spark discharges (arc and corona) shows that the yields from other discharges (arc and corona) shows that the yields can vary by at least three orders of magnitude, depending on the type of discharge and on other discharge parameters.  相似文献   

3.
Processes which occur in microwave discharges of dilute mixtures of SF6 and O2 in He have been examined using a flow reactor sampled by a mass spectrometer. Two classes of experiments were performed. In the first set of experiments, mixtures containing 6×1011 cm–3 SF6, 6×1016 cm–3 He, and O2 in the range (0–3.6)×1013 cm–3 were passed through a 20-W 2450-MHz microwave discharge. The gas mixtures arriving at a sample point downstream from the discharge were examined for SF6, SF4, SOF2, SOF4, SO2F2, SO2, F, and O. In the second class of experiments, rate coefficients were measured for the reactions of SF4 with O and O2 and for the reaction of SF with O. The rate coefficient for the reaction of SF with O was found to be (4.2±1.5)×10–11 cm–3 s–1. SF4 was found to react so slowly with both oxygen atoms and oxygen molecules that only upper limits could be placed on the rate coefficients for these reactions. These values were 2×10–14 cm3 s–1 and 5×10–15 cm3 s–1 for reactions with O and O2 respectively. The observed distribution of products from the discharged mixtures is discussed in terms of the measured rate coefficients.  相似文献   

4.
Reactions of both SF5 and SF2 with O(3 P) and molecular oxygen have been studied at 295 K in a gas flow reactor sampled by a mass spectrometer. For reactions with O(3 P), rate coefficients of (2.0±0.5)×10–11 cm3 s–1 and (10.8±2.0)×10–11 cm3 s–1 were obtained for SF5 and SF2 respectively. The rate coefficients for reactions with O2 are orders of magnitude lower, with an estimated upper limit of 5×10–16 cm3 s–1 for both SF5 and SF2. Reaction of SF2 with O(3 P) leads to the production of SOF which then reacts with O(3 P) with a rate coefficient of (7.9±2.0)×10–11 cm3 s–1. Both SO and SO2 are products in the reaction sequence initiated by reaction between SF2 and O(3 P). Although considerable uncertainty exists for the heat of formation of SOF, it appears that SO arises only from reaction between SOF and O atoms which is also the source of SO2. These results are discussed in terms of a reaction scheme proposed earlier to explain processes occurring during the plasma etching of Si in SF6/O2 plasmas. A comparison between the results obtained here and those reported earlier for reactions of both CF3 and CF2 with O and O2 shows that there is a marked similarity in the free radical chemistry which occurs in SF6/O2 and CF4/O2 plasmas.  相似文献   

5.
The plasma chemistry of SF6/O2 mixtures is particularly complicated because of the large number of possible reactions. Over a wide range of conditions, products including SF4, SOF4, SOF2, and SO2F2 can be formed but thre is considerable uncertainty about the major reactions which contribute to the formation of these species. In this work reactions of oxygen atoms with SOF2 and fluorine atoms with SOF2 and SO2 have been studied in order to determine the principal sources of SO2F2 in these plasmas. Reactions were studied at 295 K in a gas flow reactor sampled by a mass spectrometer. No reaction could be detected between oxygen atoms and SOF2, which for the conditions employed, means that the upper limit for the reaction rate coefficient is 1×10–14 cm3 sec–1. The reaction of fluorine atoms with SOF2 was studied with the helium bath gas number density ranging from 3.1×1016 to 2.0×1017 cm–3. Within this range the rate coefficient increased with increasing [He] from (4.1 to 10.8)×10–14 cm3 sec–1. SO2 was found to react with fluorine atoms with a rate coefficient which appeared to be independent of the helium bath gas number density over the range given above. The possibility that this reaction occurred entirely on the walls of the reactor is discussed.  相似文献   

6.
A fuel-lean, premixed, CH4O2 flame at atmospheric pressure was doped with 0.2 mol% of SF6 and 0.1 mol% of triflic anhydride (CF3SO2)2O. Primarily the anions, but also the cations, occurring in the flame reaction zone and in the burnt-gas region downstream were observed by sampling the flame through a nozzle into a mass spectrometer. The main objective was to ascertain the ability of these sulphur/fluorine (S/F) additives and their combustion products to scavenge free electrons by forming negative ions in the flame gas. With either additive present, both total anions and cations increased in the reaction zone by a factor of between three and four. In the burnt gas, the total cations were essentially unchanged but the total anions showed a three-fold increase. With SF6 additive only, major S/F cations were observed in the reaction zone (e.g.; SF+3, SF3O+, SF+3, etc.) although H3O+ with both additives was completely dominant downstream. Both additives produced major S/F anion signals in the flame reaction zone. Some of these were related to the individual additive structure (e.g. SF5 with SF6; CF3SO3 with (CF3SO2)2O). Others were essentially unrelated but were observed with both additives (e.g. FSO3, HSO4, etc.) and persisted throughout the burnt gas. Some important features of the ion chemistry are discussed.  相似文献   

7.
The cluster ions formed by the attachment of dimethylsulfoxide (DMSO) and methanol to the molecular negative ions of C7F14 and SF6 have been studied by a pulsed e-beam high pressure mass spectrometer (PHPMS) and by an atmospheric pressure ionization mass spectrometer (APIMS). The free energy change (ΔG°) for the clustering equilibria reaction, M + S MS, at 35 °C are found to be −7.7 and −7.s kcal/mol for S = DMSO and M = C7F14 and SF6, respectively, and −6.4 and −4.5 kcal/mol for S = methanol and M = C7F14 and SF6, respectively. While the cluster ions formed by DMSO are found to be stable against side reactions, those formed by methanol undergo decomposition processes in which the central core ion is fragmented. At 35 °C, the rate law for the decomposition of the SF6 (CH3OH)1 ion is second-order, involving the M (CH3OH)1 cluster ion and another methanol molecule. While the C7F14(CH3OH)1 ion also decomposes through this second-order process, a competing unimolecular mechanism is also operative at 35 °C. With increases in the PHPMS ion source temperature to 150 °C, the unimolecular decomposition process becomes progressively dominant for both of the M(CH3OH)1 cluster ions of C7F14 and SF6. Methanol cluster ions of the type MS2 are not observed under any of the conditions examined here. When methanol or water partial pressures of a few torr or higher are present in the buffer gas of the APIMS ion source, the decomposition reactions are very fast and only the fragment ions produced by these reactions are observed in the electron-capture (EC)-APIMS spectra of C7F14 and SF6. Also, in the methanol-containing APIMS ion source, the course of the SF6 decomposition reaction is altered so that fragment ions of the type F(S)n dominate the EC-APIMS spectrum of SF6 at all ion source temperatures. For C7F14, fragment ions of the type F(S)n become dominant at lower ion source temperatures. These previously unknown reactions are expected to be important in the analysis of perfluorinated compounds by mass spectrometric methods that utilize ionization by electron capture or negative chemical ionization. The nature of the fragment ions produced in these cluster-assisted reactions may also provide a new source of information concerning the structures of the molecular negative ions of SF6 and C7F14.  相似文献   

8.
The absolute yields of gaseous oxyfluorides SOF2, SO2F2, and SOF4 from negative, point-plane corona discharges in pressurized gas mixtures of SF6 with O2 and H2O enriched with18O2 and H2 18O have been measured using a gas chromatograph-mass spectrometer. The predominant SF6 oxidation mechanisms have been revealed from a determination of the relative18O and16O isotope content of the observed oxyfluoride by-product. The results are consistent with previously proposed production mechanisms and indicate that SOF2 and SO2F2 derive oxygen predominantly from H2O and O2, respectively, in slow, gas-phase reactions involving SF4, SF3, and SF2 that occur outside of the discharge region. The species SOF4 derives oxygen from both H2O and O2 through fast reactions in the active discharge region involving free radicals or ions such as OH and O, with SF5 and SF4.  相似文献   

9.
Characteristics of the -induced chain reaction between sulfur dioxide and molecular oxygen in perchloric and sulfuric acid media in the presence of Ce(III) ions have been studied. The concentration effects of dissolved oxygen (0.2·10–3–9.4·10–3 mol/dm3, sulfur dioxide (0.3·10–1–2.0·10–1 mol/dm3 and Ce(III) (0.2·10–3–4.8·10–3 mol/dm3) and dose rate (0.26·1019–1.0·1019 eV/dm3·s) on the radiation — chemical yield of oxygen consumption G(–O2) and accumulation of sulfate G(HSO 4 ), have been investigated. The reaction proceeds with G(–O2) reaching 102–103 molecule/100eV in a catalytic regime. The reaction rate in perchloric acid medium is 3–4 times lower than that in the sulfuric acid medium and depends on the SO2, O2 and Ce(III) concentrations, the reaction order varying from 1.0 to 0 and/or in the reverse direction. The mechanism of the process involves chain propagation with 3 stages and 3 intermediates: SO3H, HSO5 and Ce(IV). The catalytic effect is caused by the interaction of HSO4 with Ce(IV) ions followed by their reduction when interacting with SO2, yielding SO3H radicals. Chain termination may be due to one or two of the three intermediates or due to all three particles, the kinetics depending on this. Kinetic equations describing the experimental data have been obtained.  相似文献   

10.
The structure of nearly saturated or supersaturated aqueous solutions of NaCI [6.18 mol (kg H2O)–1], KCI [4.56 mol (kg H2O)–1], KF [16.15 mol (kg H2O)–1] and CsF [31.96 mol (kg H2O)–1] has been investigated by means of solution X-ray diffraction at 25°C. In the NaCI and KCI solutions about 30% and 60%, respectively, of the ions form ion pairs and the Na+–Cl and K+–Cl distances have been determined to be 282 and 315 pm, respectively. The average hydration numbers of Na+ and Cl ions are 4.6 and 5.3, respectively, in the NaCI solution and those of K+ and Cl ions in the KCI solution are both 5.8. In the KF solution, clusters containing some cations and anions, besides 1:1 (K+–F) ion pairs, are formed. The K+–F interatomic distance has been determined to be 269 pm, and nonbonding K+...K+ and F...F distances in the clusters are 388 and 432 pm, respectively, and the average coordination numbers n KF , n KK and n FF have been estimated to be 2.3, 1.9, and 1.6, respectively. In the highly supersaturated CsF solution an appreciable amount of clusters containing several caesium and fluoride ions are formed. The Cs+–F distance in the cluster has been determined to be 312 pm, while the nonbonding Cs+...Cs+ and F...F distances are estimated to be 442 and 548 pm, respectively, the distances being about and times the Cs+–F distance, respectively. The coordination numbers n CsF , n CsCs , and n FF in the first coordination sphere of each ion are 3.3, 2.3 and 5.3, respectively, and the result shows the formation of clusters of higher order than 1:1 and 2:2 ion pairs. These ion pairs and clusters may be regarded as embryos for the formation of nuclei of crystals and the results obtained in the present diffraction study support observations for the nucleation of the alkali halide crystals studied by molecular dynamics simulations previously examined.  相似文献   

11.
The rate constants and product ion branching ratios were measured for the reactions of various small negative ions with O2(X 3Σg) and O2(a 1Δg) in a selected ion flow tube (SIFT). Only NH2 and CH3O were found to react with O2(X) and both reactions were slow. CH3O reacted by hydride transfer, both with and without electron detachment. NH2 formed both OH, as observed previously, and O2, the latter via endothermic charge transfer. A temperature study revealed a negative temperature dependence for the former channel and Arrhenius behavior for the endothermic channel, resulting in an overall rate constant with a minimum at 500 K. SF6, SF4, SO3 and CO3 were found to react with O2(a 1Δg) with rate constants less than 10−11 cm3 s−1. NH2 reacted rapidly with O2(a 1Δg) by charge transfer. The reactions of HO2 and SO2 proceeded moderately with competition between Penning detachment and charge transfer. SO2 produced a SO4 cluster product in 2% of reactions and HO2 produced O3 in 13% of the reactions. CH3O proceeded essentially at the collision rate by hydride transfer, again both with and without electron detachment. These results show that charge transfer to O2(a 1Δg) occurs readily if the there are no restrictions on the ion beyond the reaction thermodynamics. The SO2 and HO2 reactions with O2(a) are the only known reactions involving Penning detachment besides the reaction with O2 studied previously [R.S. Berry, Phys. Chem. Chem. Phys., 7 (2005) 289–290].  相似文献   

12.
We studied on the function of the metal in the sulfated zirconia(SO42–/ZrO2) catalyst for the isomerization reaction of light paraffins. The addition of Pt to the SO42–/ZrO2 carrier could keep the high catalytic activity. The improvement in this isomerization activity is because Pt promotes removal of the coke precursor deposited on the catalyst surface. Though this catalytic function was observed in other transition metals, such as Pd, Ru, Ni, Rh and W, Pt exhibited the highest effect among them. It was further found that the Pd/SO42–/ZrO2–Al2O3 catalyst possessed a catalytic function for desulfurization of sulfur-containing light naphtha in addition to the skeletal isomerization. The sulfur tolerance of catalyst depended on the method of adding Pd, and the catalyst prepared by impregnation of the SO42–/ZrO2–Al2O3 with an aqueous solution of Pd exhibited the highest sulfur tolerance.Further, we investigated the improvement in sulfur tolerance of the Pt/SO42–/ZrO2–Al2O3 catalyst by impregnation of Pd. The results of EPMA analysis indicated that this catalyst was a hybrid-type one (Pt/SO42–/ZrO2–Pd/Al2O3) in which Pt/SO42–/ZrO2 particles and Pd/Al2O3 particles adjoined closely. This hybrid catalyst possessed a very high sulfur tolerance to the raw light naphtha that was obtained from the atmospheric distillation apparatus, although this light naphtha contained much sulfur. We assume that such a high sulfur tolerance in the hybrid catalyst is brought about by the isomerization function of Pt/SO42–/ZrO2 particles and the hydrodesulfurization function of Pd/Al2O3 particles. Besides, since the hybrid catalyst also provides high catalytic activity in the isomerization of HDS light naphtha, we suggest that the Pd/Al2O3 particles supply atomic hydrogen to the Pt/SO42–/ZrO2 particles by homolytic dissociation of gaseous hydrogen and also enhance the sulfur tolerance of Pt/SO42–/ZrO2 particles. Finally, we also propose the most suitable location of Pd and Pt in the metal-supported SO42–/ZrO2–Al2O3 catalyst.  相似文献   

13.
The production ofSOF 4 initiated by the reaction of F atoms withSOF 2 has been studied in a gas-flow reactor at 295 K for helium bath gas number densities in the range (3.0–27.0)×1016 cm–3. The effect of O atoms on the formation ofSOF 4 has been analyzed in terms of the competing reactionsSOF 3+FSOF4 andSOF 3+OSO 2 F 2+F. This analysis leads to the conclusion that the rate coefficients for these two processes are equal within an uncertainty of about 50%. Furthermore, both experiment and calculations indicate that the rate coefficient for reactions between F atoms andSOF 3 is close to its high-pressure limit under the conditions employed. The experiments set a lower limit on this rate coefficient of 5×10–11 cm3 s–1, while calculations based on unimolecular rate theory suggest that it may be greater than 1×10–10 cm3 s–1. These results make it clear that the two reactions shown above cannot explain the relative abundances ofSOF 4 andSO 2 F 2 which are observed inSF 6/O 2 plasmas. This suggests thatSF 2 is a major precursor in the sequence of reactions following the dissociation ofSF 6.  相似文献   

14.
Methods of 19F NMR and impedance spectroscopy are used to investigate the internal mobility of fluoride (ammonium) ions and electrophysical characteristics of complex trivalent antimony fluorides MSb4F13, MSb3F10, MSb2F7, M2Sb3F11, M3Sb4F15, and MSbF4 (M is an alkali cation, ammonium, thallium). The ion motion types in the cationic and anionic sublattices of the fluorides are determined at 150–500 K. The polymorphous transformations in the fluorides are usually phase transitions to a superionic state and their high ionic (superionic) conductivity (σ ≥ 10−4 to 10−2 S cm−1 at 400 K) is due to the diffusion motion of ions of fluoride, ammonium, and possibly sodium, potassium, and thallium. The high polarizability of thallium ions favors the development of high mobility of fluoride ions in the fluorides.__________Translated from Elektrokhimiya, Vol. 41, No. 5, 2005, pp. 560–572.Original Russian Text Copyright © 2005 by Kavun, Uvarov, Slobodyuk, Brovkina, Zemnukhova, Sergienko.  相似文献   

15.
Effects of sulphate (SO42–), thiosulphate (S2O32–) and hydrogensulphate (HSO4) ion additives on the pitting corrosion of pure aluminium in 1 M NaCl solution have been investigated at various solution temperatures Ts, 40–70 °C using potentiodynamic polarisation experiment, double current step experiment and scanning electron microscopy (SEM). From the analysis of the galvanostatic potential transients obtained from the double current step experiment, it was suggested that both SO42– and S2O32– ions retard the initiation of the etch pits, and that they also inhibit the passivation of the etch pits during the current interruption to promote the subsequent re-activation of the etch pits over the whole range of Ts. In contrast, it was found that HSO4 ions promote the initiation of the etch pits and at the same time, they enhance the passivation of the etch pits during the current interruption with rising Ts. From the SEM micrographs, it was revealed that as Ts increased the pit morphology changed from circular shape to irregular shape with rough surface in the presence of SO42– or S2O32– ions, but it changed to strip-like shape in the presence of HSO4 ions. The beneficial effects of anion additives on the increase in surface area were discussed on the basis of the morphological change of the etch pits in the presence of anion additives.  相似文献   

16.
Discrete electron-molecule processes relevant to SF6 etching plasmas are examined. Absolute, total scattering cross sections for 0.2–12-eV electrons on SF6, SO2, SOF2, SO2F2, SOF4, and SF4, as well as cross sections for negative-ion formation by attachment of electrons, have been measured. These are used to calculate dissociative-attachment rate coefficients as a function ofE/N for SF6 by-products in SF6.  相似文献   

17.
Summary A rapid method has been developed for the determination of phosphate by means of filter paper impregnated with lead iodide. A sample is added to the impregnated filter paper by means of a capillary, and after irrigation to cause migration of the ions a white spot is obtained as the lead iodide is converted into the phosphate. The weight of the spot is dependent on the pH and the quantity of phosphate present.The determination is possible in the presence of SCN, Cl, Br, NO3 , CO3 , I, IO3 , CH3COO, B4O7 2–, F, Sb2O7 4–, K+, Na+, NH4 +, OH, H+, succinic, citric and tartaric acids. The determination is impossible in the presence of C2O4 2–, SO4 2–, MoO4 2–, NO2 , SO3 2–, S2–, CrO4 2–, or CO3 2–.The method permits the determination of 7–100g of phosphate with an accuracy of 2%.
Zusammenfassung Ein schnelles Verfahren zur Phosphatbestimmung wird besehrieben, bei dem man sich eines mit Bleijodid imprägnierten Filterpapiers bedient. Die Probe wird mit einer Kapillare auf das Papier aufgebracht. Man erleichtert die Ionenbewegung durch geeignete Befeuchtung und erhält einen weißen Fleck infolge Umsetzung des Bleijodids in -phosphat. Das Gewicht des Fleckens hängt vom pH und von der Phosphatmenge ab.Die Bestimmung ist möglich in Gegenwart von SCN, Cl, Br, NO3 , CO3 , J, JO3 , CH3COO, B4O7 2–, F, Sb2O7 4–, K+, Na+, NH4 +, OH, H+, Bernsteinsäure, Zitronensäure und Weinsäure; sie ist nicht möglich bei Gegenwart von C2O4 2–, SO4 2–, MoO4 2–, NO2 , SO3 2–, S2–, CrO4 2– oder CO3 2–. 7 bis 100g Phosphat können mit einer Genauigkeit von 2% bestimmt werden.

Résumé On a développé une méthode rapide pour le dosage des phosphates sur papier-filtre imprégné d'iodure de plomb. On dépose l'échantillon sur le papier-filtre imprégné, à l'aide d'un capillaire, et, après humidification pour provoquer la migration des ions, on obtient une tache blanche quand l'iodure de plomb est converti en phosphate. Le poids de la tache dépend du pH et de la quantité de phosphate présent.Le dosage est possible en présence de SCN, Cl, Br, NO3 , CO3 , I, IO3 , CH3COO, B4O7 2–, F, Sb2O7 4–, K+, Na+, NH4 +, OH, H+, et des acides succinique, citrique et tartrique. Il est impossible en présence de C2O4 2–, SO4 2–, MoO4 2–, NO2 , SO3 2–, S2–, CrO4 2– ou CO3 2–.La méthode permet le dosage de 7 à 100g de phosphate avec une précision de 2%.
  相似文献   

18.
Capillary zone electrophoresis (CZE) was used for the separation of the sulfur species SO3 2–, SO4 2–, S2O3 2– and S2O8 2–. Using an electrolyte system with 9.5 mmol L–1 potassium chromate as UV-absorbing probe and 1 mmol L–1 diethylenetriamine (DETA) as electroosmotic flow modifier, various possibilities for the stabilization of sulfite and electrophoretic separation of the sulfur anions were investigated. By adding 5% propanol as a stabilizer to both the working electrolyte and the sample solution, a good stabilization for sulfite and a separation of the sulfur anions in a short analysis time (4 min) was achieved. The advantages by using propanol instead of other stabilizers often used in analytical techniques are discussed. The electrophoretic separation of the sulfur anions was optimized with respect to the pH of the working electrolyte and concentration of the electroosmotic flow modifier (DETA). The detection limits achieved for SO3 2–, SO4 2–, S2O3 2– and S2O8 2– were 0.35, 0.25, 0.78 and 0.80 mg L–1, respectively. Received: 16 December 1999 / Revised: 7 March 2000 / Accepted: 14 March 2000  相似文献   

19.
The molecular structures, vibrational frequencies, and electron affinities of the SF5On/SF5On (n = 1–3) species have been examined with four hybrid density functional theory (DFT) methods. The basis set used in this work is of double-ζ plus polarization quality with additional diffuse s- and p-type functions, denoted DZP++. The geometries are fully optimized with each DFT method independently. The SF5On (n = 1–3) species should be potential greenhouse gases. The anion SF5O2 with Cs symmetry has a 3A″ electronic state, and the neutral SF5O3 with 2A″ electronic state has Cs symmetry. The anions SF5O2 and SF5O3 should be regarded as SF5·O2 and SF5O·O2 complexes, respectively. Three different types of the neutral–anion energy separation presented in this work are the adiabatic electron affinity (EAad), the vertical electron affinity (EAvert), and the vertical detachment energy (VDE). The EAad values predicted by the B3PW91 method are 5.22 (SF5O), 4.38 (SF5O2), and 3.61 eV (SF5O3). Compared with the experimental vibrational frequencies, the BHLYP method overestimates the frequencies, and the other three methods underestimate the frequencies. The bond dissociation energies De (SF5On → SF5Onm + Om) for the neutrals SF5On and De (SF5On → SF5Onm + Om and SF5On → SF5Onm + Om) for the anions SF5On are reported.  相似文献   

20.
The structure of the H3O+ ion embedded in a solid environment (H3O+X, X = Cl, NO 3 , ClO 4 ) is studied using a modified version of CNDO/2. In this calculation the effect of the first shells of nearest neighbours is taken into account and the effects of second nearest neighbours are introduced by a simulation procedure. Electronic effects are also included. The ion structure is more planar in nitrate than in perchlorate environment and the hydrogen bonds are slightly bent. Trends in structural parameters are compared with chemical properties of the hydrogen bonds and parallels the Hammett acidity scale HNO3 < HCl < HClO4.  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