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 985 毫秒
1.
The gas‐phase reaction of CH3+ with NF3 was investigated by ion trap mass spectrometry (ITMS). The observed products include NF2+ and CH2F+. Under the same experimental conditions, SiH3+ reacts with NF3 and forms up to six ionic products, namely (in order of decreasing efficiency) NF2+, SiH2F+, SiHF2+, SiF+, SiHF+, and NHF+. The GeH3+ cation is instead totally unreactive toward NF3. The different reactivity of XH3+ (X = C, Si, Ge) toward NF3 has been rationalized by ab initio calculations performed at the MP2 and coupled cluster level of theory. In the reaction of both CH3+ and SiH3+, the kinetically relevant intermediate is the fluorine‐coordinated isomer H3X‐F‐NF2+ (X = C, Si). This species forms from the exoergic attack of XH3+ to one of the F atoms of NF3 and undergoes dissociation and isomerization processes which eventually result in the experimentally observed products. The nitrogen‐coordinated isomers H3X‐NF3+ (X = C, Si) were located as minimum‐energy structures but do not play an active role in the reaction mechanism. The inertness of GeH3+ toward NF3 is also explained by the endoergic character of the dissociation processes involving the H3Ge‐F‐NF2+ isomer. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

2.
Toy and Stringham recently reported [1] the synthesis of N2F+5 (CF3)3CO-, a salt containing the novel pentafluorohydrazinium cation. This cation would be of significant academic and practical interest [2] since it would constitute the first known example of a substituted NF+4 cation, i.e. an NF+4 cation in which a fluorine ligand is replaced by an NF2 group. According to the authors of [1], N2F+5(CF3)3CO- was formed in a very unusual reaction involving the transfer of a fluorine cation from (CF3)3COF to N2F4 according to:
  相似文献   

3.
Fluorination of Cyanuric Chloride and Low-Temperature Crystal Structure of [(ClCN)3F]+[AsF6]? The low-temperature fluorination of cyanuric chloride, (ClCN)3, with F2/AsF5 in SO2F2 solution yielded the salt [(ClCN)3F]+ [AsF6]? ( 1 ) essentially in quantitative yield. Compound 1 was identified by a low-temperature single crystal X-ray structure determination: R 3 c, trigonal, a = b = 10.4246(23) Å, c = 15.1850(24) Å, V = 1429.1(4) Å 3, Z = 6, RF = 0.056, Rw = 0.076 (for significant reflections), RF = 0.088, Rw = 0.079 (for all reflections). Fluorination of neat (ClCN)3 with [NF4]+ [Sb2F11]? yielded NF3, CClF3, SbF3, N2 and traces of CF4. A qualitative scale for the oxidizing strength of the oxidative fluorinators NF4+ and (XCN)3F+ (X = H, F, Cl) has been computed ab initio.  相似文献   

4.
NF3 plasma etching is used for dry cleaning of reactors after plasma-enhanced chemical vapor deposition of hydrogenated amorphous silicon from SiH4. The NF3 plasma chemistry, in a closed isothermal plasma box with silicon coated walls, is analyzed by mass spectrometry of gases. Silicon is etched as SiF4 by F atoms produced in the NF3 dissociation into F+NF2, or 2F+NF. The NF radicals recombine as N2 +2F whereas the long-lived NF2 radicals do not react with Si, but recombine as N2F4 This is the main limitation (or fluorine conversion into SiF4. The pressure increase at the end point of etching is attributed to the sudden increase of F atom concentration in the gas phase and the consequent recombination q( F atoms as F2.  相似文献   

5.
It is shown that enthalpies of nitrogen fluorides and oxo fluorides deviate markedly from enthalpies of corresponding hydroxo compounds, whereas there is good agreement (circa 2%) between carbon mono, di and trifluoro compounds and their hydroxo analogues. The extent of deviation correlates with the extra number of lone pair repulsions of fluoro over oxo compounds. The enthalpy of NF4+(g) can be extrapolated from the other (N,F) and (N,O,F) compounds. The enthalpies of solid NF4+ salts are close to those of corresponding NO2+ salts and from this an exothermic heat of formation for NF4+F? is predicted.  相似文献   

6.
The mechanism of the gas‐phase reactions of SiHn+ (n = 1,2) with NF3 were investigated by ab initio calculations at the MP2 and CAS‐MCSCF level of theory. In the reaction of SiH+, the kinetically relevant intermediates are the two isomeric forms of fluorine‐coordinated intermediate HSi‐F‐NF2+. These species arise from the exoergic attack of SiH+ to one of the F atoms of NF3 and undergo two competitive processes, namely an isomerization and subsequent dissociation into SiF+ + HNF2, and a singlet‐triplet crossing so to form the spin‐forbidden products HSiF+ + NF2. The reaction of SiH2+ with NF3 involves instead the concomitant formation of the nitrogen‐coordinated complex H2Si‐NF3+ and of the fluorine‐coordinated complex H2Si‐F‐NF2+. The latter isomer directly dissociates into NF2+ + H2SiF, whereas the former species preferably undergoes the passage through a conical intersection point so to form a H2SiF‐NF2+ isomer, which eventually dissociates into H2SiF+ and NF2. © 2012 Wiley Periodicals, Inc.  相似文献   

7.
Vibrational spectra and structural features of AuF5 complexes with nitrogen fluorides (NF3, N2F4) and oxofluorides (FNO, NF3O) are investigated. Vibrational frequency assignment in the solid phase and in solution of anhydrous HF was made. Distinctive features of vibrational spectra of X+AuF6 ? (X = NF4 +, N2F3 +, NO+, NOF2 +) related to structural transformations of cations and hexafluoroaurate anion due to the influence of the crystal field and cation-anion interactions are discussed.  相似文献   

8.
The reproducible synthesis of the unusual ionic aluminum compound [Tl3F2Al(OR)3]+[Al(OR)4] ( 1 ) is reported. In the reaction of Li[Al(OR)4] [R = C(H)(CF3)2] with TlF the initially desired Tl[Al(OR)4] only formed with an exact 1:1 stoichiometry, while an excess of TlF led to [Tl3F2Al(OR)3]+[Al(OR)4] ( 1 ). Additionally the x‐ray single crystal structure of the byproduct [(R‐OH)TlAl(OR)3(μ‐F)]2 ( 2 ) was determined. Compounds 1 and 2 were characterized by X‐ray single crystal structure determinations and 1 also by NMR spectroscopy and an elemental analysis. In 1 the [Tl3F2Al(OR)3]+ cation forms a trigonal bipyramid with a pentacoordinate aluminum atom. Three Tl+ cations cover the [F2Al(OR)3]2— dianion core and the charge of the resulting [Tl3F2Al(OR)3]+ cation is compensated by a weakly coordinating [Al(OR)4] anion. Compound 2 contains a centrosymmetric [Al(OR)3(μ‐F)]22— dianion core with pentacoordinate aluminum atoms building a distorted edge sharing double trigonal bipyramid. The [Al(OR)3(μ‐F)]22— dianion coordinates two [Tl(R‐OH)]+ cations giving the non charged molecular [(R‐OH)TlAl(OR)3(μ‐F)]2 ( 2 ). Based on BP86/SVP (DFT‐) and lattice enthalpy calculations a pathway of the reaction was proposed to rationalize the formation of the [M3F2Al(OR)3]+ cation upon reaction of Li[Al(OR)4] with MF for M = Tl but not for M = Cs (cf. Cs+ and Tl+ have very similar ionic radii). Using a suitable BorñHaber cycle and in agreement with the experiment, the enthalpies of the reaction of 2 M[Al(OR)4] with 2 MF giving [M3F2Al(OR)3]+[Al(OR)4] and MOR were shown to be favorable for M = Tl by 127 kJ/mol but endothermic for the formation of the hypothetical [Cs3F2Al(OR)3]+[Al(OR)4] by 95 kJ/mol. It is suggested that in the reaction leading to 1 initially Tl[Al(OR)4] is formed, followed by an abstraction of TlOR and Al(OR)3. The latter very strong Lewis acid reacts subsequently with an excess of TlF yielding 1 .  相似文献   

9.
Conclusions Electron-impact methods have been used to study positive (NF+, NF2 +, N2F2 +) and negative (F, F2 , N2F, NF2 ) ion formation in the mass spectra of tetrafluorohydrazine and its decomposition products.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 6, pp. 1306–1311, June, 1978.The authors would like to thank V. Ya. Rosolovskii for a discussion of the results obtained in this work.  相似文献   

10.
Electronically excited NF in both the a1Δ and b 1Σ+ states hasbeen observed from the reaction of fluorine atoms with HN3. The results suggest that fluorine atoms first abstract the hydrogen atom from HN3, then react with the remaining N3 to form NF(a1Δ). NF*(b1Σ+) is produced by a subsequent energy pooling reaction between NF(a1Δ) and vibrationally excited HF. The rate of the F + N3 reaction is estimated to be ≈ 1012 and 3 mole?1 s?1.  相似文献   

11.
Thermally induced dehydrogenation of the H‐bridged cation L2B2H5+ (L=Lewis base) is proposed to be the key step in the intramolecular C? H borylation of tertiary amine boranes activated with catalytic amounts of strong “hydridophiles”. Loss of H2 from L2B2H5+ generates the highly reactive cation L2B2H3+, which in its sp2‐sp3 diborane(4) form then undergoes either an intramolecular C? H insertion with B? B bond cleavage, or captures BH3 to produce L2B3H6+. The effect of the counterion stability on the outcome of the reaction is illustrated by formation of LBH2C6F5 complexes through disproportionation of L2B2H5+ HB(C6F5)3?.  相似文献   

12.
The products of the reaction between the electrophilic alkenylxenonium cation [1-Xe+–C6F9] and the halide anions I?, Br?, Cl? and F? depend on the hardness of the halide anion. With the soft halides I? and Br? Xe(II) is formally displaced by halogen as well in basic MeCN as in superacidic (AHF1), whereas with hard fluoride and chloride no reaction takes place in AHF. In MeCN F? initiates the formation of alkenyl radicals, which abstract hydrogen from the solvent, whereas Cl? exhibits borderline character: RH and RCl formation. Possible reaction paths are discussed. The reactivity of the arylxenonium cation [C6F5Xe]+ in AHF toward halide ions is reported and the relative electrophilicity of the cations [C6F5Xe]+ and [1-Xe+–C6F9] is determined by the competitive reaction with Cl?. In addition the synthesis of cyclohexene 1-CF3–C6F9 from C6F5CF3 and XeF2 is performed and its electrophilicity is compared with that of the aromatic compound C6F5CF3.  相似文献   

13.
Salts that contain radical cations of benzidine (BZ), 3,3′,5,5′‐tetramethylbenzidine (TMB), 2,2′,6,6′‐tetraisopropylbenzidine (TPB), and 4,4′‐terphenyldiamine (DATP) have been isolated with weakly coordinating anions [Al(ORF)4]? (ORF=OC(CF3)3) or SbF6?. They were prepared by reaction of the respective silver(I) salts with stoichiometric amounts of benzidine or its alkyl‐substituted derivatives in CH2Cl2. The salts were characterized by UV absorption and EPR spectroscopy as well as by their single‐crystal X‐ray structures. Variable‐temperature UV/Vis absorption spectra of BZ . +[Al(ORF)4]? and TMB . +[Al(ORF)4]? in acetonitrile indicate an equilibrium between monomeric free radical cations and a radical‐cation dimer. In contrast, the absorption spectrum of TPB . +SbF6? in acetonitrile indicates that the oxidation of TPB only resulted in a monomeric radical cation. Single‐crystal X‐ray diffraction studies show that in the solid state BZ and its methylation derivative (TMB) form radical‐cation π dimers upon oxidation, whereas that modified with isopropyl groups (TPB) becomes a monomeric free radical cation. By increasing the chain length, π stacks of π dimers are obtained for the radical cation of DATP. The single‐crystal conductivity measurements show that monomerized or π‐dimerized radicals (BZ . +, TMB . +, and TPB . +) are nonconductive, whereas the π‐stacked radical (DATP . +) is conductive. A conduction mechanism between chains through π stacks is proposed.  相似文献   

14.
Thermally induced dehydrogenation of the H‐bridged cation L2B2H5+ (L=Lewis base) is proposed to be the key step in the intramolecular C H borylation of tertiary amine boranes activated with catalytic amounts of strong “hydridophiles”. Loss of H2 from L2B2H5+ generates the highly reactive cation L2B2H3+, which in its sp2‐sp3 diborane(4) form then undergoes either an intramolecular C H insertion with B B bond cleavage, or captures BH3 to produce L2B3H6+. The effect of the counterion stability on the outcome of the reaction is illustrated by formation of LBH2C6F5 complexes through disproportionation of L2B2H5+ HB(C6F5)3.  相似文献   

15.
The valence threshold photoelectron spectrum of NF3 is reported for the first time in the literature, and threshold photoelectron–photoion coincidence (TPEPICO) spectroscopy has measured, state-selectively, the decay dynamics of the valence states of NF3+ in the range 13–23 eV. Vacuum–UV radiation from the Daresbury synchrotron source dispersed by a 1 m Seya-Namioka monochromator photoionises the parent molecules. Electrons and ions are detected by threshold electron analysis and time-of-flight mass spectrometry, respectively. TPEPICO spectra are recorded continuously as a function of photon energy, allowing coincidence ion yields of the fragment ions and the breakdown diagram to be obtained. A comparison of the integrated threshold photoelectron and the total ion signals as a function of energy suggests that, in the range 16–19 eV, autoionisation via Rydberg states of NF3 makes a significant contribution to the production of threshold electrons. The 50% crossover energy for production of NF2+ from NF3+ is determined to be 14.10±0.05 eV. The first onsets for NF2+ and NF+ production are 13.95±0.05 and 17.6±0.1 eV, respectively. The majority of the Franck–Condon region of the ground state of NF3+ is stable with respect to dissociation to NF2+, whereas the unresolved states and most of the state dissociate exclusively to NF2+. The and states dissociate to NF+. Translational kinetic energy releases have been measured in NF2+ and NF+ at the energies of the Franck–Condon maxima of the valence states of NF3+. The results are compared with models assuming statistical and impulsive dissociation. The Ã/ states of NF3+ dissociate directly from the excited-state potential energy surface to NF2+, whereas the higher-lying state probably dissociates off the ground-state surface following rapid internal conversion. It is not possible to correlate unambiguously the formation of NF+ with either F2 or 2F, although on energetic grounds the latter products are more likely. Assuming that the neutral products are 2F, no information is obtained whether the two N–F bonds break simultaneously or sequentially.  相似文献   

16.
Gas phase irradiation of N2F4 (NF2) in the presence of hexafluoroacetone imine (I), N-chlorohexafluoroacetone imine (II), or N-bromohexafluoroacetone imine (III) resulted in the formation of products that correspond to either perhalogenation of the unsaturation or conversion of the substrate to a saturated halocarbon. The mechanism suggested involves the formation of an imino radical that reacts with N2F4(NF2) to produce N,N-difluorohydrazone, (CF3)2CNNF2. A bimolecular homolytic displacement (SH2′) by Cl and F on the hydrazone forms an intermediate diazene which leads to the observed products. N-fluorohexafluoroacetone imine is inert to F atoms and NF2 under the reaction conditions.  相似文献   

17.
The X‐ray irradiation of binary mixtures of alkyl iodides R?I (R=CH3, C2H5, or i‐C3H7 radicals) and NF3 produces R?NF2 and R?F. Based on calculations performed at the CCSD(T), MRCI(SD+Q), G3B3, and G3 levels of theory, the former product arises from a bimolecular homolytic substitution reaction (SH2) by the alkyl radicals R, which attack the N atom of NF3. This mechanism is consistent with the suppression of R?NF2 by addition of O2 (an efficient alkyl radical scavenger) to the reaction mixture. The R?F product arises from the attack of R to the F atom of NF3, but additional contributing channels are conceivably involved. The F‐atom abstraction is, indeed, considerably more exothermic than the SH2 reaction, but the involved energy barriers are comparable, and the two processes are comparably fast.  相似文献   

18.
The reactivity of ZnII dialkyl species ZnMe2 with a cyclic(alkyl)(amino)carbene, 1-[2,6-bis(1-methylethyl)phenyl]-3,3,5,5-tetramethyl-2-pyrrolidinylidene (CAAC, 1 ), was studied and extended to the preparation of robust CAAC-supported ZnII Lewis acidic organocations. CAAC adduct of ZnMe2 ( 2 ), formed from a 1:1 mixture of 1 and ZnMe2, is unstable at room temperature and readily undergoes a CAAC carbene insertion into the Zn−Me bond to produce the ZnX2-type species (CAAC-Me)ZnMe ( 3 ), a reactivity further supported by DFT calculations. Despite its limited stability, adduct 2 was cleanly ionized to robust two-coordinate (CAAC)ZnMe+ cation ( 5+ ) and derived into (CAAC)ZnC6F5+ ( 7+ ), both isolated as B(C6F5)4 salts, showing the ability of CAAC for the stabilization of reactive [ZnMe]+ and [ZnC6F5]+ moieties. Due to the lability of the CAAC−ZnMe2 bond, the formation of bis(CAAC) adduct (CAAC)2ZnMe+ cation ( 6+ ) was also observed and the corresponding salt [ 6 ][B(C6F5)4] was structurally characterized. As estimated from experimental and calculations data, cations 5+ and 7+ are highly Lewis acidic species and the stronger Lewis acid 7+ effectively mediates alkene, alkyne and CO2 hydrosilylation catalysis. All supporting data hints at Lewis acid type activation–functionalization processes. Despite a lower energy LUMO in 5+ and 7+ , their observed reactivity is comparable to those of N-heterocyclic carbene (NHC) analogues, in line with charge-controlled reactions for carbene-stabilized ZnII organocations.  相似文献   

19.
Attempts to prepare previously unknown simple and very Lewis acidic [RZn]+[Al(ORF)4]? salts from ZnR2, AlR3, and HO?RF delivered the ion‐like RZn(Al(ORF)4) (R=Me, Et; RF=C(CF3)3) with a coordinated counterion, but never the ionic compound. Increasing the steric bulk in RZn+ to R=CH2CMe3, CH2SiMe3, or Cp*, thus attempting to induce ionization, failed and led only to reaction mixtures including anion decomposition. However, ionization of the ion‐like EtZn(Al(ORF)4) compound with arenes yielded the [EtZn(arene)2]+[Al(ORF)4]? salts with arene=toluene, mesitylene, or o‐difluorobenzene (o‐DFB)/toluene. In contrast to the ion‐like EtZn(η3‐C6H6)(CHB11Cl11), which co‐crystallizes with one benzene molecule, the less coordinating nature of the [Al(ORF)4]? anion allowed the ionization and preparation of the purely organometallic [EtZn(arene)2]+ cation. These stable materials have further applications as, for example, initiators of isobutene polymerization. DFT calculations to compare the Lewis acidities of the zinc cations to those of a large number of organometallic cations were performed on the basis of fluoride ion affinity. The complexation energetics of EtZn+ with arenes and THF was assessed and related to the experiments.  相似文献   

20.
Upon reaction of gaseous Me3SiF with the in situ prepared Lewis acid Al(ORF)3, the stable ion‐like silylium compound Me3Si‐F‐Al(ORF)3 1 forms. The Janus‐headed 1 is a readily available smart Lewis acid that differentiates between hard and soft nucleophiles, but also polymerizes isobutene effectively. Thus, in reactions of 1 with soft nucleophiles (Nu), such as phosphanes, the silylium side interacts in an orbital‐controlled manner, with formation of [Me3Si?Nu]+ and the weakly coordinating [F?Al(ORF)3] or [(FRO)3Al‐F‐Al(ORF)3] anions. If exchanged for hard nucleophiles, such as primary alcohols, the aluminum side reacts in a charge‐controlled manner, with release of FSiMe3 gas and formation of the adduct R(H)O?Al(ORF)3. Compound 1 very effectively initiates polymerization of 8 to 21 mL of liquid C4H8 in 50 mL of CH2Cl2 already at temperatures between ?57 and ?30 °C with initiator loads as low as 10 mg in a few seconds with 100 % yield but broad polydispersities.  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