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.
Hydride abstraction from diarylamines with the trityl ion is explored in an attempt to generate a stable diarylnitrenium ion, Ar2N+. Sequential H-atom abstraction reactions ensue. The first H-atom abstraction leads to intensely colored aminium radical cations, Ar2NH.+, some of which are quite stable. However, most undergo a second H-atom abstraction leading to ammonium ions, Ar2NH2+. In the absence of a ready source of H-atoms, a unique self-abstraction reaction occurs when Ar=Me5C6, leading to a novel iminium radical cation, Ar=N.+Ar, which decays via a second self H-atom abstraction reaction to give a stable iminium ion, Ar=N+HAr. These products differ substantially from those derived via photochemically produced diarylnitrenium ions.  相似文献   

2.
The reaction between Ar2+ and C2H2 has been studied, at centre-of-mass collision energies ranging from 3 to 7 eV, using a position-sensitive coincidence technique to detect the monocation pairs, which are formed. Sixteen different reaction channels generating pairs of monocations have been observed, these channels arise from double-electron-transfer, single-electron-transfer and chemical reactions forming ArC+. Examination of the scattering diagrams and energetic information extracted from the coincidence data indicate that double-electron-transfer is a direct process, which does not involve a collision complex, and the derived energetics point towards a concerted, not stepwise, mechanism for the two-electron-transfer. As is commonly observed, single-electron-transfer from C2H2 to Ar2+ takes place via a direct mechanism, again not involving complexation. Most of the C2H2+ products that are formed in the single-electron-transfer reactions possess significant (12–15 eV) internal energy and fragment rapidly within the electric field of the partner Ar+ ion. The chemical reactions appear to proceed via a direct mechanism involving the initial formation of ArCH+, which subsequently fragments to form ArC+.  相似文献   

3.
An evolutionary algorithm was used to search for the low-energy structures of Li+Arn and Li+Krn (n = 1 − 14). Two functions were used to describe the interaction potential at the CCSD(T)/aug-cc-pVQZ level of theory: one is based on a sum of all pair potentials, whereas the other includes three-body interactions. In general, the global minimum structures are similar for both Li+Arn and Li+Krn. Modifications in the octahedral structure of the first solvation shell lead to a high-energy penalty. Conversely, the second solvation shell shows a panoply of minima with similar energies that are likely to be interconverted. Post-optimization at the MP2 level confirmed that, for n = 2 and 3, one has to include three-body terms in the potential to reproduce the low-energy structures. Additionally, MP2 calculations indicate that energy reorder of the global minimum structure observed for Li+Kr8 is related to the Kr3 Axilrod-Teller-Muto term included in the potential.  相似文献   

4.
Observations on metastable peaks resulting from the unimolecular decomposition of ion clusters show that intensity variations as a function of cluster size can reveal the presence of stable cluster configurations. This technique has been used to confirm that (H2O)21H+ and (D2O)21D+ are stable ion clusters, and the method also provides evidence to suggest that Ar19+ is a particularly stable species.  相似文献   

5.
Determination of copper using inductively coupled plasma mass spectrometry (ICP-MS) suffers from polyatomic overlays originating from Na+ and Mg2+ matrix elements due to the formation of 40Ar23Na+ and 40Ar25Mg+ in the mass-to-charge ratios of 63 and 65, respectively. The collision/reaction cell technology belongs to the most modern methods used to overcome polyatomic interferences. Gas-filled collision/reaction cell can cause an additional mass bias effect influencing analytical precision of the method. In this study, the additional mass bias effect of the hexapole collision/reaction cell ICP-MS was studied on an example of n(65Cu)/n(63Cu) isotope ratio. As a result, a method for suppressing polyatomic interference on the mass-to-charge ratio of 63 and 65 was introduced and additional mass bias of the collision/reaction cell was lowered to an acceptable level.  相似文献   

6.
Photoionisation experiments were performed with heterogeneous Ar-Xe-clusters produced by supersonic expansion of argon gas with small quantities of xenon added to it. A threshold-electron photoionisation (TEPICO) technique was used to obtain time of flight cluster mass spectra. These mass spectra show particularly strong intensities for Ar12Xe+ and Ar18Xe+ which are attributed to the extraordinary stabilities of these cluster ions. Maxima in the ionic size distribution around Ar7Xe+ are related to a particular abundance ofneutral Ar12Xe which is fragmented after ionization. These stabilities are explained in terms of geometries consisting of a central Xe atom or ion surrounded by shells of Ar atoms. Filled shells exhibit particular strong bonding because these exhibit the largest number of atom-atom bonds. This conclusion is supported by simple theoretical calculations. The ionization process is discussed in terms of two direct and one indirect ionization channels the latter one proceeding via an intermediate electronic excitation of the Ar component in the neutral cluster.  相似文献   

7.
Centrosymmetric linear [Ar-H-Ar]+ and asymmetric linear [Ar---Ar-H]+ are two stable configurations of [Ar2H]+. Based on the global potential energy surface of [Ar2H]+ provided by our group recently, we calculated the vibrational spectra of [Ar---Ar-H]+ with total angular momentum J = 0 by time-dependent quantum mechanical method, and the influence of quantum tunneling effect on vibrational spectra was found. With the help of the observation on the eigenstate functions and the modified potential energy surface, assignments were made to the spectra. The strong coupling between the excited bending mode of [Ar-H-Ar]+ and the vibrational states of [Ar---Ar-H]+ was discussed.  相似文献   

8.
Mass spectra of Nb+ and Rh+ complexes with argon ligands exhibit `magic' peaks Nb+Ar4 and Rh+Ar6, similar to observations for V+Ar4 and Co+Ar6, indicating coordinative saturation. A consistent explanation is obtained by assuming that the rare gas ligands seek out electron density minima in the valence shell of the ion, which permit a closer approach to the metal core and a stronger charge-induced dipole bond. Ab initio density functional calculations, which predict stable square planar complexes for the d4 ions and octahedral for the d8 species, support this interpretation and show that rare gas complexes of d4 metal ions fit perfectly well into the coordination chemical framework based on the Jahn–Teller effect.  相似文献   

9.
A DIM model using ab initio input for the diatomic interactions predicts a collinear bound Ar3 + molecule (in agreement with ab initio calculations) and stable clusters Arn + consisting of an Ar3 + ion embedded in (n?3) neutral atoms. These results support existing theories that dynamical size selection may be more relevant in interpreting experimental results than the relative stabilities of clusters in their minimum energy configurations.  相似文献   

10.
The structure, binding energy, and vibrational frequencies have been determined for ScCO2+. The inserted OSc+CO structure in the 1A′ state is the most stable isomer and lies 43.2 kcal/mol below the ground-state Sc++ CO2 asymptote. The linear η1-O Sc+(SINGLE BOND)OCO 3Δ state is bound by a charge-quadrupole interaction and has a binding energy of 13.9 kcal/mol. © 1997 John Wiley & Sons, Inc. Int J Quant Chem 63: 523–528, 1997  相似文献   

11.
The formation of Ar 2 + ions has been investigated by means of the threshold photoelectron photoion coincidence (TPEPICO) technique. Two pathways for the formation of Ar 2 + ions are important. One is a direct path via excitation of Rydberg states of Ar2 with consecutive autoionization. The other path is dissociative ionization of larger argon clusters, in this case argon trimers. These two pathways lead to Ar 2 + ions with different internal energy. The pathways are easily distinguished in the TPEPICO-TOF spectra by the kinetic energy released (KER) in the dissociative ionization. The KER for the reaction Ar 3 + → Ar 2 + + Ar was measured as a function of the photon energy and compared to the KER expected from statistical theory. The agreement is satisfying and confirms that Ar 3 + ions do indeed dissociate at the thermochemical threshold. At higher photon energy the excited2Π(3/2)g state of Ar 3 + is also detected from a second component in the KER. By applying a kinetic energy discrimination it is possible to measure cluster ion spectra in the presence of larger clusters but essentially without interference from the latter.  相似文献   

12.
Picosecond multiphoton ionization of (NO)mArn clusters produced in a supersonic expansion of NO/Ar gas mixtures has been studied using time-of-flight mass spectrometry. Two-photon ionization with 266 nm photons show that dilute gas mixtures (1% NO/Ar) yield clusters limited to m≤7, but with as many as 37 argon atoms. Magic numbers are observed for NO+Ar12, NO+Ar18, (NO) 2 + Ar17, NO+Ar22, and (NO) 2 + Ar21 and are understood in terms of solvation of the NO+ and (NO) 2 + by argon in icosahedral arrangements. Four-photon ionization with 532 nm light produces dissociation of the clusters to yield only NO+Arn with n up to 54. This distribution exhibits an additional magic number at n=54, consistent with the completion of a second solvation sphere about the NO+. The known wavelength dependence for photodissociation of (NO) 2 + and (NO) 3 + and comparison of MPI spectra obtained with 266, 355, and 532 nm light indicate that the dissociation is occurring in the cluster ions.  相似文献   

13.
Rearrangement of the molecular ions of tetracyclone, tetraphenylquinone, tetraphenylthiophene dioxide and pentaphenylcyclopentadienol prior to decomposition, rather than formation of a symmetrical [C4Ar4]+˙ ion, is considered as an alternative method for the production of [C2Ar2]+˙. The p-fluoro substituent is used to elicit information about the positional origins of the [C2Ar2]+˙ fragments. Operation of a double focusing mass spectrometer in the defocused mode reveals that the integrity of the molecular ion in different systems is affected to varying degrees, ranging from total lack of rearrangement to virtually complete scrambling.  相似文献   

14.
The translational mobility of polypropylenephosphate (PPP) molecules has been studied by the NMR pulsed field gradient technique. The translational diffusion of long chain polymer molecules of polymerization degree ¯m > 20 may be neglected, in NMR measurements, on account of their rigidity. The self-diffusion coefficient of short chain polymer molecules (¯m=8) in PPP gels (¯Ds=3 · 10–8 cm2/s) is about two orders of magnitude lower than that of water molecules in gels.Two hydrates, i. e. PPP(H+) · 3H2O and PPP(H+, Mg2+) · 5H2O represent the most stable structures, at temperature 293 K.  相似文献   

15.
In aqueous H2SO4, Ce(IV) ion oxidizes rapidly Arnold's base((p-Me2NC6H4)2CH2, Ar2CH2) to the protonated species of Michler's hydrol((p-Me2NC6H4)2CHOH, Ar2CHOH) and Michler's hydrol blue((p-Me2NC6H4)2CH+, Ar2CH+). With Ar2CH2 in excess, the rate law of the Ce(IV)-Ar2CH2 reaction in 0.100 M H2SO4 is expressed -d[Ce(IV)]/dt = kapp[Ar2CH2]0[Ce(IV)] with kapp = 199 ± 8M?1s?1 at25°C. When the consumption of Ce(IV) ion is nearly complete, the characteristic blue color of Ar2CH+ ion starts to appear; later it fades relatively slowly. The electron transfer of this reaction takes place on the nitrogen atom rather than on the methylene carbon atom. The dissociation of the binuclear complex [Ce(III)ArCHAr-Ce(III)] is responsible for the appearance of the Ar2CH+ dye whereas the protonation reaction causes the dye to fade. In highly acidic solution, the rate law of the protonation reaction of Michler's hydrol blue is -d[Ar2CH+]/dt = kobs[Ar2CH+] where Kobs = ((ac + 1)[H*] + bc[H+]2)/(a + b[H+]) (in HClO4) and kobs= ((ac + 1 + e[HSO4?])[H+] + bc[H+]2 + d[HSO4?] + q[HSO4?]2/[H+])/(a + b[H+] + f[HSO4?] + g[HSO4?]/[H+]) (in H2SO4), and at 25°C and μ = 0.1 M, a = 0.0870 M s, b = 0.655 s, c = 0.202 M?1s?1, d = 0.110, e = 0.0070 M?1, f = 0.156 s, g = 0.156 s, and q = 0.124. In highly basic solution, the rate law of the hydroxylation reaction of Michler's hydrol blue is -d[Ar2CH+]/dt = kOH[OH?]0[Ar2CH+] with kOH = 174 ± 1 M?1s?1 at 25°C and μ = 0.1 M. The protonation reaction of Michler's hydrol blue takes place predominantly via hydrolysis whereas its hydroxylation occurs predominantly via the path of direct OH attack.  相似文献   

16.
在B3LYP/6-311++G**水平上用极化连续介质模型(PCM)系统研究了金属离子(M+/2+=Na+,K+,Ca2+,Mg2+,Zn2+)和十三种鸟嘌呤异构体形成的配合物GnxM+/2+(n为鸟嘌呤异构体的编号,x表示M+/2+与鸟嘌呤异构体的结合位点)在气(g)液(a)两相中的稳定性顺序.着重探讨了液相中配合物的稳定性差异,并且从溶质-溶剂效应、结合能、形变能及异构体的相对能量等几个方面分析了造成稳定顺序发生变化的原因.报道了溶液中这五种金属离子与鸟嘌呤异构体结合形成的六种基态配合物:aG1N2,N3Na+,aG1N2,N3K+,aG1O6,N7Ca2+,aG1N2,N3Mg2+(aG1O6,N7Mg2+),aG2N3,N9Zn2+.可以看出,除了在Zn2+配合物中鸟嘌呤异构体为G2外,构成其余四种金属离子配合物的鸟嘌呤异构体都是G1,但结合位点不同.同时对气相中各类配合物稳定性也进行了系统的排序,并报道了几种较稳定的配合物,如:gG3N1,O6K+,gG5N1,O6K+,gG3N1,O6Ca2+/Mg2+,gG4O6,N7Ca2+/Mg2+.  相似文献   

17.
In the system Ba/(Mg, Li)/Ge, two new Zintl phases with the composition Ba2Mg12Ge7.33 (P63/m, Z = 1, a = 1121.7(5) pm, c = 440.2(2) pm) and Ba6Mg17.4Li2.6Ge12O0.64 (P63/m, Z = 1, a = 1537.8(8) pm, c = 454.6(2) pm) are found and structurally characterized. Their structures are described with respect to the Zintl‐Klemm concept, structure directing rules, and chemical twinning. These new compounds contain as a specific structural feature cationic channels with partial anion occupation which allows to adjust the electron count. In Ba2Mg12Ge7.33, the channels are formed by Mg2+ cations and are partially filled with germanium dumb‐bells, while the channels in Ba6Mg17.4Li2.6Ge12O0.64 are formed by Li+ and Mg2+ cations and host O2— anions. The electronic structure of both compounds has been investigated using Extended‐Hückel calculations with special emphasis on the states of the cationic channels and their interstitial heteroatoms. The potentiality of using the electron localization function (ELF) to find missing atoms in structures has been tested and verified for both compounds.  相似文献   

18.
Tris(2,6-dimethoxyphenyl)amine has been synthesized and its molecular and crystal structure determined by X-ray diffraction. This structure completes the series of isosteric compounds Ar3Z, where Z=B, C., N, and Ar=2,6-dimethoxyphenyl. Structures for the tris(2-methoxy-6-methylphenyl) borane and tris(2,6-dimethoxyphenyl)methyl cation triiodide (Ar3C+I3 ) are also reported. The Ar3B and Ar3N structures are isomorphous. The triiodide and the earlier reported tetrafluoroborate salt (Ar3C+BF4 ) are also quite similar, as are the two boranes above and the known trimesitylborane, which all tend toward D3 symmetric conformations. In contrast, the radical Ar3C., intermediate between Ar3B and Ar3N, is markedly unsymmetrical. Taken together, these findings support an earlier conjecture that the solid-state conformation of Ar3C. does not represent a minimum energy structure for the free radical in solution. Crystal seeding by radical oxidation products is offered as an explanation for the radical's markedly unsymmetrical crystal geometry.  相似文献   

19.
One‐electron oxidation of the disilicon(0) compound Si2(Idipp)2 ( 1 , Idipp=1,3‐bis(2,6‐diisopropylphenyl)imidazolin‐2‐ylidene) with [Fe(C5Me5)2][B(ArF)4] (ArF=C6H3‐3,5‐(CF3)2) affords selectively the green radical salt [Si2(Idipp)2][B(ArF)4] ( 1 ‐[B(ArF)4). Oxidation of the centrosymmetric 1 occurs reversibly at a low redox potential (E1/2=?1.250 V vs. Fc+/Fc), and is accompanied by considerable structural changes as shown by single‐crystal X‐ray structural analysis of 1 ‐B(ArF)4. These include a shortening of the Si?Si bond, a widening of the Si‐Si‐CNHC angles, and a lowering of the symmetry, leading to a quite different conformation of the NHC substituents at the two inequivalent Si sites in 1+ . Comparative quantum chemical calculations of 1 and 1+ indicate that electron ejection occurs from the symmetric (n+) combination of the Si lone pairs (HOMO). EPR studies of 1 ‐B(ArF)4 in frozen solution verified the inequivalency of the two Si sites observed in the solid‐state, and point in agreement with the theoretical results to an almost equal distribution of the spin density over the two Si atoms, leading to quite similar 29Si hyperfine coupling tensors in 1+ . EPR studies of 1 ‐B(ArF)4 in liquid solution unraveled a topomerization with a low activation barrier that interconverts the two Si sites in 1+ .  相似文献   

20.
The title compounds 2 are prepared from the reaction of 1-(N, N-diaroyl)amino-5-bromomethyl-1,2,3-triazoles with aromatic amines. The fragmentation pattern upon electron impact at 70 eV of compounds 2 is studied. The molecular ion peak is present in all the spectra examined. Besides the [M-28]++, there is also a more abundant [M-29]+ peak, corresponding to a N2H loss of the molecular ion. The ion Ar2NH = CH2 is the base or the most prominent peak.  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