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 437 毫秒
1.
The role of the HO4? anion in atmospheric chemistry and biology is a matter of debate, because it can be formed from, or be in equilibrium with, key species such as O3 + HO? or HO2 + O2?. The determination of the stability of HO4? in water therefore has the greatest relevance for better understanding the mechanism associated with oxidative cascades in aqueous solution. However, experiments are difficult to perform because of the short‐lived character of this species, and in this work we have employed DFT, CCSD(T) complete basis set (CBS), MRCI/aug‐cc‐pVTZ, and combined quantum mechanics/molecular mechanics (QM/MM) calculations to investigate this topic. We show that the HO4? anion has a planar structure in the gas phase, with a very large HOO? OO bond length (1.823 Å). In contrast, HO4? adopts a nonplanar configuration in aqueous solution, with huge geometrical changes (up to 0.232 Å for the HOO? OO bond length) with a very small energy cost. The formation of the HO4? anion is predicted to be endergonic by 5.53±1.44 and 2.14±0.37 kcal mol?1 with respect to the O3 + HO? and HO2 + O2? channels, respectively. Moreover, the combination of theoretical calculations with experimental free energies of solvation has allowed us to obtain accurate free energies for the main reactions involved in the aqueous decomposition of ozone. Thus, the oxygen transfer reaction (O3 + OH? → HO2 + O2?) is endergonic by 3.39±1.80 kcal mol?1, the electron transfer process (O3 + O2? → O3? + O2) is exergonic by 31.53±1.05 kcal mol?1, supporting the chain‐carrier role of the superoxide ion, and the reaction O3 + HO2? → OH + O2? + O2 is exergonic by 12.78±1.15 kcal mol?1, which is consistent with the fact that the addition of small amounts of HO2? (through H2O2) accelerates ozone decomposition in water. The combination of our results with previously reported thermokinetic data provides some insights into the potentially important role of the HO4? anion as a key reaction intermediate.  相似文献   

2.
Ab initio molecular orbital calculations with split-valence plus polarization basis sets and incorporating valence-electron correlation have been performed to determine the equilibrium structure of ethyloxonium ([CH3CH2OH2]+) and examine its modes of unimolecular dissociation. An asymmetric structure (1) is predicted to be the most stable form of ethyloxonium, but a second conformational isomer of Cs symmetry lies only 1.4 kJ mol?1 higher in energy than 1. Four unimolecular decomposition pathways for 1 have been examined involving loss of H2, CH4, H2O or C2H4. The most stable fragmentation products, lying 65 kJ mol?1 above 1, are associated with the H2 elimination reaction. However, large barriers of 257 and 223 kJ mol?1 have to be surmounted for H2 and CH4 loss, respectively. On the other hand, elimination of either C2H4 or H2O from ethyloxonium can proceed without a barrier to the reverse associations and, with total endothermicities of 130 and 160 kJ mol?1, respectively, these reactions are expected to dominate at lower energies. A second important equilibrium structure on the surface is a hydrogen-bridged complex, lying 53 kJ mol?1 above 1. This complex is involved in the C2H4 elimination reaction, acts as an intermediate in the proton-transfer reaction connecting [C2H5]+ +H2O and C2H4 + [H3O]+ and plays an important role in the isotopic scrambling that has been observed experimentally in the elimination of either H2O or C2H4 from ethyloxonium. The proton affinity of ethanol was calculated as 799 kJ mol?1, in close agreement with the experimental value of 794 kJ mol?1.  相似文献   

3.
Ab initio molecular orbital calculations with moderately large polarization basis sets and including valence-electron correlation have been used to examine the structure and dissociation mechanisms of protonated methanol [CH3OH2]+. Stable isomers and transition structures have been characterized using gradient techniques. Protonated methanol is found to be the only stable isomer in the [CH5O]+ potential surface. There is no evidence for a tightly-bound complex, [HOCH2]+…?H2, analogous to the preferred structure [CH3]+…?H2 of [CH5]+. Protonated methanol is found to possess a pyramidal arrangement of bonds at the oxygen atom with a barrier to inversion of 8kJ mol?1. The lowest energy fragmentation pathways are dissociation into methyl cation and water (predicted to require 284 kJ mol?1 with zero reverse activation energy) and loss of molecular hydrogen (endothermic by 138 kJ mol?1 but with a reverse activation barrier of 149 kJ mol?1). The results offer a possible explanation as to why production of [CH2OH]+ from the reaction of methyl cation with water is not observed. Other dissociation processes examined include loss of a hydrogen atom to yield the methylenoxonium radical cation or methanol radical cation (requiring 441 and 490 kJ mol?1, respectively) and loss of a proton to yield neutral methanol (requiring 784 kJ mol?1).  相似文献   

4.
The ionization and [C4H7]+ appearance energies for a series of C4H7CI and C4H7Br isomers have been measured by dissociative photoionization mass spectrometry. Cationic heats of formation, based on the stationary electron convention, are derived. No threshold ion is observed with a heat of formation corresponding to the trans-1-methylallyl cation, although there is evidence for formation of the less stable cis isomer. A 298 K heat of formation of 871 kJ mol?1 is obtained for the cyclopropylcarbinyl cation, with the cyclobutyl cation having a higher value of 886 kJ mol?1. At the HF/6-31G** level, ab initio molecular orbital calculations show the 2-butenyl, isobutenyl and homoallyl cations to be stable forms of [C4H7]+, being less stable than the trans-1-methylallyl cation by 101 kJ mol?1, 159 kJ mol?1 and 164 kJ mol?1, respectively. However, threshold formation is not observed for any of these ions, the fragmentation of appropriate precursor molecules producing [C4H7]+ ions with lower energy structures.  相似文献   

5.
The Absolute rate constants for the gas-phase reactions of NO3 with HO2 and OH have been determined using the discharge flow laser magnetic resonance method (DF-LMR). Since OH was found to be produced in the reaction of HO2 with NO3, C2F3Cl was used to scavenge it. The overall rate constant, k1, for the reaction, HO2 + NO3 → products, was measured to be k1=(3.0 ± 0.7)×10?12 cm3 molecule?1 s?1 at (297 ± 2) K and P=(1.4 – 1.9) torr. This result is in reasonable agreement with the previous studies. Direct detection of HO2 and OH radicals and the use of three sources of NO3 enabled us to confirm the existence of the channel producing OH:HO2+NO3→OH+NO2+O2 (1a); the other possible channel is HO2+NO3→HNO3+O2 (1b). From our measurements and the computer simulations, the branching ratio, k1a/(k1a + k1b), was estimated to be (1.0). The rate coefficient for the reaction of OH with NO3 was determined to be (2.1 ± 1.0) × 10?11 cm3 molecule?1 s?1. © 1993 John Wiley & Sons, Inc.  相似文献   

6.
The kinetics and mechanism by which monochloramine is reduced by hydroxylamine in aqueous solution over the pH range of 5–8 are reported. The reaction proceeds via two different mechanisms depending upon whether the hydroxylamine is protonated or unprotonated. When the hydroxylamine is protonated, the reaction stoichiometry is 1:1. The reaction stoichiometry becomes 3:1 (hydroxylamine:monochloramine) when the hydroxylamine is unprotonated. The principle products under both conditions are Cl, NH+4, and N2O. The rate law is given by ?[d[NH2Cl]/dt] = k+[NH3OH+][NH2Cl] + k0[NH2OH][NH2Cl]. At an ionic strength of 1.2 M, at 25°C, and under pseudo‐first‐order conditions, k+= (1.03 ± 0.06) ×103 L · mol?1 · s?1 and k0=91 ± 15 L · mol?1 · s?1. Isotopic studies demonstrate that both nitrogen atoms in the N2O come from the NH2OH/NH3OH+. Activation parameters for the reaction determined at pH 5.1 and 8.0 at an ionic strength of 1.2 M were found to be ΔH? = 36 ± 3 kJ · mol–1 and Δ S? = ?66 ± 9 J · K?1 · mol?1, and Δ H? = 12 ± 2 kJ · mol?1 and Δ S? = ?168 ± 6 J · K?1 · mol?1, respectively, and confirm that the transition states are significantly different for the two reaction pathways. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 38: 124–135, 2006  相似文献   

7.
The sequential segregation of Sn and Sb to the surface of a Cu(111) single crystal was measured in the temperature range 400–1100 K by Auger electron spectroscopy. It was found that Sn with the higher diffusion coefficient first segregates to the surface and then is replaced by the slower‐segregating Sb. The results were fitted by a ternary segregation model yielding segregation energies (ΔGSn = 76.3 kJ mol?1, ΔGSb = 95.9 kJ mol?1), interaction parameters (ΩSnCu = 3.8 kJ mol?1, ΩSbCu = 16.2 kJ mol?1, ΩSnSb = ?5.3 kJ mol?1) and diffusion coefficients (D0(Sn) = 1.8 × 10?5 m2 s?1, ESn = 173 kJ mol?1, D0(Sb) = 6.0 × 10?5 m2 s?1, ESb = 205 kJ mol?1) for both species. The validity of the interaction coefficients and segregation energies was verified using the Guttman equations for equilibrium segregation in ternary systems. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

8.
The oxidative addition of BF3 to a platinum(0) bis(phosphine) complex [Pt(PMe3)2] ( 1 ) was investigated by density functional calculations. Both the cis and trans pathways for the oxidative addition of BF3 to 1 are endergonic (ΔG°=26.8 and 35.7 kcal mol?1, respectively) and require large Gibbs activation energies (ΔG°=56.3 and 38.9 kcal mol?1, respectively). A second borane plays crucial roles in accelerating the activation; the trans oxidative addition of BF3 to 1 in the presence of a second BF3 molecule occurs with ΔG° and ΔG° values of 10.1 and ?4.7 kcal mol?1, respectively. ΔG° becomes very small and ΔG° becomes negative. A charge transfer (CT), F→BF3, occurs from the dissociating fluoride to the second non‐coordinated BF3. This CT interaction stabilizes both the transition state and the product. The B?F σ‐bond cleavage of BF2ArF (ArF=3,5‐bis(trifluoromethyl)phenyl) and the B?Cl σ‐bond cleavage of BCl3 by 1 are accelerated by the participation of the second borane. The calculations predict that trans oxidative addition of SiF4 to 1 easily occurs in the presence of a second SiF4 molecule via the formation of a hypervalent Si species.  相似文献   

9.
10.
The kinetics of decomposition of [Alg · Mn VIO42?] intermediate complex have been investigated spectrophotometrically at a constant ionic strength of 0.5 mol dm?3. The decomposition reaction was found to be first-order in the intermediate concentration. The results showed that the rate of reaction was base-catalyzed. The kinetic parameters have been evaluated and found to be ΔS? = ?103.88±6.18 J mol?1 K?1, ΔH? = 51.61 ± 1.02 kJ mol?1, and ΔG? = 82.57 ± 2.86 kJ mol?1, respectively. A reaction mechanism consistent with the results is discussed. © 1993 John Wiley & Sons, Inc.  相似文献   

11.
The surface segregation of In and S from a dilute Cu(In,S) ternary alloy were measured using Auger electron spectroscopy coupled with a linear programmed heater. The alloy was linearly heated and cooled at constant rates. Segregation data of a linear heat run showed surface segregation of In that reached a maximum surface coverage of 25% followed by S, which reached a coverage of 30%. It was found that after In had reached a maximum surface coverage, it started to desegregate as soon as the S enriched the surface until In was completely replaced by S. The segregation parameters, namely, the pre‐exponential factor (D0), activation energy (Q), segregation energy (ΔG?) and interaction energy (Ω) were extracted from the measured segregation data for both In and S segregation in Cu by simulating the measured segregation data with a theoretical segregation model (modified Darken model). The segregation parameters obtained for In segregation in Cu are D0 = 1.8 ± 0.5 × 10?5 m2 s?1, Q = 184.3 ± 1.0 kJ.mol?1, ΔG? = ?61.4 ± 1.4 kJ.mol‐1, ΩCu?In = 3.0 ± 0.4 kJ.mol?1; for S segregation in Cu the parameters are D0 = 8.9 ± 0.5 × 10?3 m2 s?1, Q = 212.8 ± 3.0 kJ.mol?1, ΔG? = ?120.0 ± 3.5 kJ.mol?1, ΩCu?S = 23.0 ± 2.0 kJ mol?1 and the In and S interaction parameter is ΩIn?S = ?4.0 ± 0.5 kJ.mol?1. The initial parameters used for the Darken calculations were extracted from fits performed with the Fick's and Guttmann model. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

12.
The kinetics of base hydrolysis of (αβ S)-(o -methoxy benzoato) (tetraethylenepentamine)cobalt(III) obeyed the rate law: kobs = kOH[OH?], in the range 0.05 ? [OH?]T, mol dm?3 ? 1.0, I = 1.0 mol dm?3, and 20.0–40.0°C. At 25°C, kOH = 13.4 ± 0.4 dm3 mol?1 s?1, ΔH = 93 ± 2 kJ mol?1 and ΔS = 90 ± 5 JK?1 mol?1. Several anions of varying charge and basicity, CH3CO2?, SO32?, SO42?, CO32?, C2O42?, CH2(CO2)22?, PO43?, and citrate3? had no effect on the rate while phthalate2?, NTA3?, EDTA4?, and DTPA5? accelerated the process via formation of the reactive ion pairs. The anionic (SDS), cationic (CTAB), and neutral (Triton X-100) micelles, however, retarded the reaction, the effect being in the order SDS> CTAB > Triton X-100. The importance of electrostatic and hydrophobic effects of the micelles on the selective partitioning of the reactants between the micellar and bulk aqueous pseudo-phases which control the rate are discussed. © 1994 John Wiley & Sons, Inc.  相似文献   

13.
We report that 2,6‐lutidine?trichloroborane (Lut?BCl3) reacts with H2 in toluene, bromobenzene, dichloromethane, and Lut solvents producing the neutral hydride, Lut?BHCl2. The mechanism was modeled with density functional theory, and energies of stationary states were calculated at the G3(MP2)B3 level of theory. Lut?BCl3 was calculated to react with H2 and form the ion pair, [LutH+][HBCl3?], with a barrier of ΔH=24.7 kcal mol?1G=29.8 kcal mol?1). Metathesis with a second molecule of Lut?BCl3 produced Lut?BHCl2 and [LutH+][BCl4?]. The overall reaction is exothermic by 6.0 kcal mol?1rG°=?1.1). Alternate pathways were explored involving the borenium cation (LutBCl2+) and the four‐membered boracycle [(CH2{NC5H3Me})BCl2]. Barriers for addition of H2 across the Lut/LutBCl2+ pair and the boracycle B?C bond are substantially higher (ΔG=42.1 and 49.4 kcal mol?1, respectively), such that these pathways are excluded. The barrier for addition of H2 to the boracycle B?N bond is comparable (ΔH=28.5 and ΔG=32 kcal mol?1). Conversion of the intermediate 2‐(BHCl2CH2)‐6‐Me(C5H3NH) to Lut?BHCl2 may occur by intermolecular steps involving proton/hydride transfers to Lut/BCl3. Intramolecular protodeboronation, which could form Lut?BHCl2 directly, is prohibited by a high barrier (ΔH=52, ΔG=51 kcal mol?1).  相似文献   

14.
The formation of 1 : 2 titanium(III) complex with chromotropic acid (4, 5-dihydroxy-2, 7-naphthalene-disulfonic acid) was observed by spectrophotometric measurements at various ionic strengths. An expression, [Ti(III)]/D=1/Δ? + αH2+/KΔ?[H2R2?]2, was derived for the determination of the formation constant, K=7.2×102 liter2 mol?2 for the Ti(III).(HR)2 ion in the pH range of 1.3–1.8 at constant ionic strength, I=0.2 M, at 25°C. The thermodynamic data for the reaction, Ti(III)+2H3R2?=Ti(III) (HR)2+2H+, were calculated to be ΔG° = ?16 kJ mol?1 ΔH° = 18 kJ mol?1, ΔS° = 110 JK?1 mol?1, at 25°C.  相似文献   

15.
The catalytic activity of diamond‐supported gold nanoparticle (Au/D) samples prepared by the deposition/precipitation method have been correlated as a function of the pH and the reduction treatment. It was found that the most active material is the one prepared at pH 5 followed by subsequent thermal treatment at 300 °C under hydrogen. TEM images show that Au/D prepared under optimal conditions contain very small gold nanoparticles with sizes below 2 nm that are proposed to be responsible for the catalytic activity. Tests of productivity using large phenol (50 g L ?1) and H2O2 excesses (100 g L ?1) and reuse gives a minimum TON of 458,759 moles of phenol degraded per gold atom. Analysis of the organic compounds extracted from the deactivated solid catalyst indicates that the poisons are mostly hydroxylated dicarboxylic acids arising from the degradative oxidation of the phenyl ring. By determining the efficiency for phenol degradation and the amount of O2 evolved two different reactions of H2O2 decomposition (the Fenton reaction at acidic pH values and spurious O2 evolution at basic pH values) are proposed for Au/D catalysis. The activation energy of the two processes is very similar (ranging between 30 and 35 kJ mol?1). By using dimethylsulfoxide as a radical scavenger and Ntert‐butyl‐α‐phenylnitrone as a spin trap under aerated conditions, the EPR spectrum of the expected PBN? OCH3 adduct was detected, supporting the generation of HO., characteristic of Fenton chemistry in the process. Phenol degradation, on the other hand, exhibits the same activation energy as H2O2 decomposition at pH 4 (due to the barrierless attack of HO. to phenol), but increases the activation energy gradually up to about 90 kJ mol?1 at pH 7 and then undergoes a subsequent reduction as the pH increases reaching another minimum at pH 8.5 (49 kJ mol?1).  相似文献   

16.
Ethenol, 1-d-ethenol, O-d-ethenol and Z-2-d-ethenol were prepared by pyrolysis of corresponding 5-norbornenols at 800°C/2 × 10?6 Torr. The most important fragments in the electron impact mass spectrum of ethenol are [C2H3O]+ and CHO+ and CH3˙. The hydrogen atom eliminated from the molecular ion comes mainly from the hydroxyl group (68%) and to a lesser extent from C(1) (25%) and C(2) (7%). The loss of the hydroxyl hydrogen is preceded by rate-determining migration of the hydrogen atom from C(1) onto C(2) to yield CH3C?OH+˙ions that decompose to CH3CO+ and H˙. The loss of deuterium from O-d-ethenol shows a very small primary isotope effect (kH/kD=1.07), whereas a significant effect is observed for the loss of hydrogen from 1-d-ethenol (kH/kD=1.28). The appearance energy of [C2H2DO]+ from 1-d-ethenol, AE=11.32 eV, gives a critical energy for the hydrogen loss, E=203 kJ mol?1, which is 90 kJ mol?1 above the thermochemical threshold for CH3CO++H˙. The appearance energy of CDO+ from 1-d-ethenol was measured as 12.96±0.07 eV, which sets the barrier to isomerization to CH3CDO+˙ at 1121 kJ mol?1. The ionization energy of ethenol was found to be 9.22±0.03 eV.  相似文献   

17.
The water exchange reactions of [PuO2(OH2)5]2+ and [UO2(OH2)5]2+ were investigated with density functional theory (DFT) and wave function theory (WFT). Geometries and vibrational frequencies were calculated with DFT and CPCM hydration. The electronic energies were evaluated with general multiconfiguration quasi-degenerate second-order perturbation theory (GMC-QDPT2). Spin-orbit (SO) effects, computed with SO configuration interaction (SO–CI), are negligible. Both Actinyl(VI) ions react via an associative exchange mechanism, most likely Ia. The Gibbs activation energies (ΔG?) at 25 °C are 33–34 and 30–37 kJ mol?1 for [PuO2(OH2)5]2+ and [UO2(OH2)5]2+, respectively. ΔG? for dissociative mechanisms (D, Id) is higher by more than 15 kJ mol?1.  相似文献   

18.
The following peptides have been examined in this study: GLDFG(OH), caeridin 1.1 [GLLDGLLGLGGL(NH2)], 11 Ala citropin 1.1 [GLFDVIKKVAAVIGGL(NH2)], Crinia angiotensin [APGDRIYVHPF(OH)] and their isoAsp isomers. It is not possible to differentiate between Asp‐ and isoAsp‐containing peptides (used in this study) using negative ion electrospray mass spectrometry. This is because the isoAsp residue cleaves to give the same fragment anions as those formed by δ and γ backbone cleavage of Asp. The isoAsp fragmentations are as follows: RNHCH(CO2H)?CHCONHR′ → [RNH?(HO2CCH?CHCONHR′)] → RNH?+HO2CCH?CHCONHR′ and RNHCH(CO2H)?CHCONHR′ → [RNH?(HO2CCH?CHCONHR′] → ?O2CCH?CHCONHR′+RNH2. Calculations at the HF/6‐31+G(d)//AM1 level of theory indicate that the first of these isoAsp cleavage processes is endothermic (by +115 kJ mol?1), while the second is exothermic (?85 kJ mol?1). The barrier to the highest transition state is 42 kJ mol?1. No diagnostic cleavage cations were observed in the electrospray mass spectra of the MH+ ion of the Asp‐ and isoAsp‐containing peptides (used in this study) to allow differentiation between these two amino acid residues. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
Protonated silanoic acid, HSi(OH)2+, 1a +, is cleanly generated by the dissociative electron ionization of triethoxysilane, HSi(OC2H5)3, and tetraethoxysilane, Si(OC2H5)4. This follows from tandem mass spectrometric experiments and CBS-QB3 model chemistry calculations. The calculations predict that 1a +Hf(298 K) = 205 kJ mol−1) is separated by high barriers from its isomers HOSiOH2+, 1b + and HSi(O)OH2+, 1c +. Low-energy (metastable) ions 1a + dissociate by loss of H2O via the pathway 1a + → 1b + → SiOH+ + H2O. Analysis of the metastable peak for this process confirms that the isomerization step 1a + → 1b + is rate determining. The calculations further predict that the incipient ions 1b + communicate via a low barrier with the proton-bound dimer SiO···H···OH2+, 1d +. This dimer ion is much lower in energy than its counterpart OSi···H···OH2+, 1e +, which is calculated to be only marginally stable. A comparison of the potential energy diagram for the silicon-containing ions 1a +– 1e + with that of their carbon analogues reveals that the dissociation chemistries of HSi(OH)2+ and HC(OH)2+ are only superficially similar. Neutralization–reionization experiments confirm the theoretical prediction that the HSi(OH)2 radical (ΔHf(298 K) = −455 kJ mol−1) is a stable species in the rarefied gas phase. However, owing to a mismatch of Franck–Condon factors a large fraction of the neutralized ions dissociates by loss of H yielding Si(OH)2. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

20.
The use of nano‐carbon paste electrodes for the measurement of Gibbs energies of transfer between oil and aqueous phases is reported. In this method the oil of interest is used as the binder for the nano‐carbon paste electrodes and the molecule of interest is dissolved in the organic or aqueous phase. Voltammetry is performed over a period of time and used to monitor the transfer of the molecule between the two phases. The method is illustrated for the transfer of ferrocenemethanol between water and oil using the ferrocenemethanol / ferroceniummethanol (FcCH2OH/FcCH2OH+) redox couple. Three pairs of voltammetric peaks were observed in a 0.1 M KCl solution when the nano‐carbon paste electrode was modified by dissolution of FcCH2OH in the binder oil: P1 [E=0.23 V, 0.17 V vs. Ag/AgCl (1 M KCl)], P2 [E=0.36 V, 0.32 V vs. Ag/AgCl (1 M KCl)] and P3 [E=0.55 V, 0.46 V vs. Ag/AgCl (1 M KCl)]. These are assigned to the FcCH2OH species existing in the aqueous solution [FcCH2OH(aq)/FcCH2OH+(aq)], originating in the oil (o) [FcCH2OH(o)/FcCH2OH+(aq)] and to oxidation of adsorbed (ads) material on the nano‐carbon [FcCH2OH(ads)] respectively. When supporting electrolyte containing the anions Cl?, NO3? or SCN? was used, an expulsion of the oxidised ferrocene occurred and the difference in midpoint potentials (Emid) between the peaks P1 and P2 observed in these experiments allowed the calculation of the Gibbs energy (Δ) of transfer of ferrocenemethanol from water to oil. The average Δ value thus obtained was (?12.7±0.2) kJ mol?1. For more hydrophobic anions (X?=PF6?, AsF6?), the electron transfer is coupled to the transfer of the anion into the oil and the Δ for the transfer of the ion pair of FcCH2OH+ and X? ions from water to oil was found to be ?1.3 and ?3.9 kJ mol?1 for PF6? and AsF6? 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