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.
The new Lewis acid Al(OTeF5)3 and its acetonitrile adduct CH3CN→Al(OTeF5)3 were obtained by a simple one-step synthesis in batches of up to 15 g. Al(OTeF5)3 and the adduct were characterized by vibrational spectroscopy (IR, Raman) and quantum-chemical calculations. Furthermore, five different salts of the new weakly coordinating anion [Al(OTeF5)4] were prepared in a two-step procedure. [Ph4P][Al(OTeF5)4], Cs[Al(OTeF5)4], [Ph3C][Al(OTeF5)4], as well as the protonated benzene derivatives [C9H13][Al(OTeF5)4] and [C6H7][Al(OTeF5)4] were characterized by low-temperature single-crystal X-ray diffraction and NMR spectroscopy. Arenium salts have rarely been characterized in the solid state and were synthesized in this work in a simplified fashion.  相似文献   

2.
A theoretical gas-phase “ligand-free” or “electron pair affinity” (EPA) approach, based on CCSD(T)/(SDB-)cc-pVTZ//MP2/(SDB-)cc-pVTZ electronic structure calculations, is introduced as a possible means for determining Lewis acidity trends among planar EX30/+ (E = B, C, Al, Si; X = F, Cl, Br, I) species. In this treatment, the free electron pair is considered to be an extreme Lewis base. The calculated EPA values are compared with experimental Lewis acidities, previously calculated fluoride ion affinity (FIA) and hydride ion affinity (HA) trends, and are found to exhibit reasonable correlations in all cases. The bonding in the planar and trigonal pyramidal conformations of EX30/+ and of the trigonal pyramidal Lewis base EX32−/− anions are assessed by use of natural bond orbital (NBO) and natural resonance theory (NRT) analyses. The NBO charges of the CX3+ (X = Cl, Br, I, OTeF5) cations are shown to correlate with the cation-anion and cation-solvent contacts in the recently determined crystal structures of [CCl3][Sb(OTeF5)6], [CBr3][Sb(OTeF5)6]·SO2ClF, [CI3][Al(OC(CF3)3)4], and [C(OTeF5)3][Sb(OTeF5)6]·3SO2ClF and known fluoro-carbocation structures. Topological electron localization function (ELF) basin lobe isosurfaces and volumes are used to rationalize the Lewis acidity trends and bond ionicities of the EX30/+ species, and Lewis basicities of the EX32−/− species.  相似文献   

3.
The pentafluoroorthotellurate group (teflate, OTeF5) is able to form species, for which only the fluoride analogues are known. Despite nickel fluorides being widely investigated, nickel teflates have remained elusive for decades. By reaction of [NiCl4]2− and neat ClOTeF5, we have synthesized the homoleptic [Ni(OTeF5)4]2− anion, which presents a distorted tetrahedral structure, unlike the polymeric [NiF4]2−. This high-spin complex has allowed the study of the electronic properties of the teflate group, which can be classified as a weak/medium-field ligand, and therefore behaves as the fluoride analogue also in ligand-field terms. The teflate ligands in [NEt4]2[Ni(OTeF5)4] are easily substituted, as shown by the formation of [Ni(NCMe)6][OTeF5]2 by dissolving it in acetonitrile. Nevertheless, careful reactions with other conventional ligands have enabled the crystallization of nickel teflate complexes with different coordination geometries, i.e. [NEt4]2[trans-Ni(OEt2)2(OTeF5)4] or [NEt4][Ni(bpyMe2)(OTeF5)3].  相似文献   

4.
Preparation and Electrochemistry of [Nb(OTeF5)6]? and [Ta(OTeF5)6]? Complexes Nb(OTeF5)5 and Ta(OTeF5)5 react with Cs[OTeF5], [Et4N][OTeF5], and [(n-Bu)4N][OTeF5] to the corresponding Cs[M(OTeF5)6], [Et4N][M(OTeF5)6], and [(n-Bu)4N][M(OTeF5)6] complexes, (M = Nb, Ta). The electrochemical reduction of the niobium complex occurs in CH2Cl2 at ?0,69 V and in acetonitrile at ?0,60 V (vs. SCE). The tantalum complex is reduced in CH2Cl2 at ?1,52 V and in acetonitrile at ?1,42 V (vs. SCE).  相似文献   

5.
Salts of the weakly coordinating anions [Ga(OTeF5)4] as well as [Ga(Et)(OTeF5)3] and the neutral Ga2(Et)3(OTeF5)3 were synthesized and characterized by spectroscopic methods and single-crystal X-ray diffraction. Ga2(Et)3(OTeF5)3 was formed by treating GaEt3 with pentafluoroorthotelluric acid (HOTeF5) and reacted with PPh4Cl and CPh3Cl to [PPh4][Ga(Et)(OTeF5)3] and [CPh3][Ga(Et)(OTeF5)3]. In contrast, Ag[Ga(OTeF5)4] was prepared from AgOTeF5 and GaCl3 and was used as a versatile starting material for further reactions. Starting with Ag[Ga(OTeF5)4] the substrates [PPh4][Ga(OTeF5)4] and [CPh3][Ga(OTeF5)4] were formed from PPh4Cl and CPh3Cl.  相似文献   

6.
Xe(OTeF5)2 reacts with Sb(OTeF5)3 under the formation of [Xe2(OTeF5)3]+[Sb(OTeF5)6]-. From SO2ClF solution a yellow solvate [F5TeOXe]+·SO2ClF· [Sb(OTeF5)6]- is formed with the crystal data: a = 1028.1(1), b = 1040.9(1), c = 1780.2(3) pm, α = 98.07(1), β = 97.68(1), γ = 105.82(1)°, space group . The O-Xe···O fragment is essentially linear (176.1(2)°), and the two Xe-O distances are quite different 197.1(4) and 242.6(4) pm.  相似文献   

7.
Comparative Structural Studies on 4‐Dimethylaminopyridine‐Adducts Lewis acid‐base adducts of the type dmap—MMe3 (M = Al 1 , Ga 2 , In 3 , Tl 4 ) as well as dmap—AlCl3 ( 6 ) and dmap—Al(t‐Bu)3 ( 7 ) were synthesized by reaction of MR3 with 4‐dimethylamino‐pyridine (dmap) whereas dmap—AlH3 ( 5 ) was obtained from AlH3·Et2O. 1 — 7 were characterized by means of NMR (1H, 13C{1H}) and mass spectrometry and elemental analysis. In addition, their solid state structures were determined by single crystal X‐ray diffraction studies. A comparison of the structural parameters reveales the influence of both electronic (Lewis acidity of the group 13 atom) and steric interactions on the structure and stability of as prepared Lewis acid‐base adducts.  相似文献   

8.
We herein report detailed investigations into the interaction of Lewis acidic fluoroboranes, for example BF2Pf (Pf=perfluorophenyl) and BF2ArF (ArF=3,5‐bis(trifluoromethyl)phenyl), with Lewis basic platinum complexes such as [Pt(PEt3)3] and [Pt(PCy3)2] (Cy=cyclohexyl). Two presumed Lewis adducts could be identified in solution and corresponding secondary products of these Lewis adducts were characterized in the solid state. Furthermore, the concept of frustrated Lewis pairs (FLP) was applied to the activation of ethene in the system [Pt(BPf3)(CH2CH2)(dcpp)] (dcpp=1,3‐bis(dicyclohexylphosphino)propane; Pf=perfluorophenyl). Finally, DFT calculations were performed to determine the interaction between the platinum‐centered Lewis bases and the boron‐centered Lewis acids. Additionally, several possible mechanisms for the oxidative addition of the boranes BF3, BCl3, and BF2ArF to the model complex [Pt(PMe3)2] are presented.  相似文献   

9.
Yanan Zhu  Zhigang Yao  Fan Xu 《Tetrahedron》2018,74(31):4211-4219
Cationic lanthanide complexes [Ln(CH3CN)9]3+[(AlCl4)3]3–·CH3CN served as efficient catalysts for the tandem Friedel–Crafts alkylation/ketalization reaction of 2-hydroxychalcones with naphthols/substituted phenols to produce diaryl-fused 2,8-dioxabicyclo[3.3.1]nonanes in moderate to high yields. The high catalytic efficiency of the cationic lanthanide complex [Yb(CH3CN)9]3+[(AlCl4)3]3–·CH3CN compared with that of YbCl3 can be attributed to the increased Lewis acidity of the Yb species as a result of cation formation.  相似文献   

10.
We report herein the synthesis and full characterization of the donor‐free Lewis superacids Al(ORF)3 with ORF=OC(CF3)3 ( 1 ) and OC(C5F10)C6F5 ( 2 ), the stabilization of 1 as adducts with the very weak Lewis bases PhF, 1,2‐F2C6H4, and SO2, as well as the internal C? F activation pathway of 1 leading to Al2(F)(ORF)5 ( 4 ) and trimeric [FAl(ORF)2]3 ( 5 , ORF=OC(CF3)3). Insights have been gained from NMR studies, single‐crystal structure determinations, and DFT calculations. The usefulness of these Lewis acids for halide abstractions has been demonstrated by reactions with trityl chloride (NMR; crystal structures). The trityl salts allow the introduction of new, heteroleptic weakly coordinating [Cl‐Al(ORF)3]? anions, for example, by hydride or alkyl abstraction reactions.  相似文献   

11.
The dimethylchloronium salt [Me2Cl][Al(OTeF5)4] is used to methylate electron-deficient aromatic systems in Friedel–Crafts type reactions as shown by the synthesis of N-methylated cations, such as [MeNC5F5]+, [MeNC5F4I]+, and [MeN3C3F3]+. To gain a better understanding of such fundamental Friedel–Crafts reactions, the role of the dimethylchloronium cation has been evaluated by quantum-chemical calculations.  相似文献   

12.
Treating 1,3-dichloro-2,4-bis[tris(trimethylsilyl)silyl]-cyclo-diphosphadiazane, [HypNPCl]2 ((Me3Si)3Si = Hyp), or N-(2,4,6-tri-tert-butylphenyl)imino(chloro)phosphane, Mes-NP-Cl (Mes = 2,4,6-tri-tert-butylphenyl), with Ag[Al(OCH(CF3)2)4] leads to the abstraction of [OCH(CF3)2] from the counter ion [Al(OCH(CF3)2)4] in a formal Lewis acid/Lewis base reaction. The final products Hyp2N2P2(Cl)(OCH(CF3)2), Mes-NP-OCH(CF3)2 and the dimeric Lewis acid [Al(OCH(CF3)2)3]2 have been characterized by means of X-ray analysis.  相似文献   

13.
The unexpected but facile preparation of the silver salt of the least coordinating [(RO)3Al‐F‐Al(OR)3]? anion (R=C(CF3)3) by reaction of Ag[Al(OR)4] with one equivalent of PCl3 is described. The mechanism of the formation of Ag[(RO)3Al‐F‐Al(OR)3] is explained based on the available experimental data as well as on quantum chemical calculations with the inclusion of entropy and COSMO solvation enthalpies. The crystal structures of (RO)3Al←OC4H8, Cs+[(RO)2(Me)Al‐F‐Al(Me)(OR)2]?, Ag(CH2Cl2)3+[(RO)3Al‐F‐Al(OR)3]? and Ag(η2‐P4)2+[(RO)3Al‐F‐Al(OR)3]? are described. From the collected data it will be shown that the [(RO)3Al‐F‐Al(OR)3]? anion is the least coordinating anion currently known. With respect to the fluoride ion affinity of two parent Lewis acids Al(OR)3 of 685 kJ mol?1, the ligand affinity (441 kJ mol?1), the proton and copper decomposition reactions (?983 and ?297 kJ mol?1) as well as HOMO level and HOMO–LUMO gap and in comparison with [Sb4F21]?, [Sb(OTeF5)6]?, [Al(OR)4]? as well as [B(RF)4]? (RF=CF3 or C6F5) the [(RO)3Al‐F‐Al(OR)3]? anion is among the best weakly coordinating anions (WCAs) according to each value. In contrast to most of the other cited anions, the [(RO)3Al‐F‐Al(OR)3] anion is available by a simple preparation in conventional inorganic laboratories. The least coordinating character of this anion was employed to clarify the question of the ground state geometry of the Ag(η2‐P4)2+ cation (D2h, D2 or D2d?). In agreement with computational data and NMR spectra it could be shown that the rotation along the Ag‐(P‐P‐centroid) vector has no barrier and that the structure adopted in the solid state depends on packing effects which lead to an almost D2h symmetric Ag(η2‐P4)2+ cation (0 to 10.6° torsion) for the more symmetrical [Al(OR)4]? anion, but to a D2 symmetric Ag(η2‐P4)2+ cation with a 44° twist angle of the two AgP2 planes for the less symmetrical [(RO)3Al‐F‐Al(OR)3]? anion. This implies that silver back bonding, suggested by quantum chemical population analyses to be of importance, is only weak.  相似文献   

14.
Cyanoborane adducts of the Lewis acids B(CN)3, BF(CN)2, and BH(CN)2 with pyridine and 4-cyanopyridine have been obtained in high yields. The syntheses were accomplished by oxidation of the readily available potassium salts of the cyano(hydrido)borate anions [BH(CN)3] ( MHB ), [BFH(CN)2] ( FHB ), and [BH2(CN)2] ( DHB ) with bromine in the presence of the respective pyridine derivative C5H5N or 4-CN-C5H4N as starting material. All six cyanoborane adducts have been characterized by NMR and vibrational spectroscopy, elemental analysis, and single-crystal X-ray diffraction. The reduction of the cyanoborane adducts has been investigated by cyclic voltammetry and the Lewis acidity of the different cyanoboranes has been assessed using the Gutmann-Beckett method. Selected experimental data and trends are compared to theoretical ones, for example fluoride ion affinities (FIAs).  相似文献   

15.
Monocationic bis‐allyl complexes [Ln(η3‐C3H5)2(thf)3]+[B(C6X5)4]? (Ln=Y, La, Nd; X=H, F) and dicationic mono‐allyl complexes of yttrium and the early lanthanides [Ln(η3‐C3H5)(thf)6]2+[BPh4]2? (Ln=La, Nd) were prepared by protonolysis of the tris‐allyl complexes [Ln(η3‐C3H5)3(diox)] (Ln=Y, La, Ce, Pr, Nd, Sm; diox=1,4‐dioxane) isolated as a 1,4‐dioxane‐bridged dimer (Ln=Ce) or THF adducts [Ln(η3‐C3H5)3(thf)2] (Ln=Ce, Pr). Allyl abstraction from the neutral tris‐allyl complex by a Lewis acid, ER3 (Al(CH2SiMe3)3, BPh3) gave the ion pair [Ln(η3‐C3H5)2(thf)3]+[ER31‐CH2CH?CH2)]? (Ln=Y, La; ER3=Al(CH2SiMe3)3, BPh3). Benzophenone inserts into the La? Callyl bond of [La(η3‐C3H5)2(thf)3]+[BPh4]? to form the alkoxy complex [La{OCPh2(CH2CH?CH2)}2(thf)3]+[BPh4]?. The monocationic half‐sandwich complexes [Ln(η5‐C5Me4SiMe3)(η3‐C3H5)(thf)2]+[B(C6X5)4]? (Ln=Y, La; X=H, F) were synthesized from the neutral precursors [Ln(η5‐C5Me4SiMe3)(η3‐C3H5)2(thf)] by protonolysis. For 1,3‐butadiene polymerization catalysis, the yttrium‐based systems were more active than the corresponding lanthanum or neodymium homologues, giving polybutadiene with approximately 90 % 1,4‐cis stereoselectivity.  相似文献   

16.
The methylamino diazonium cations [CH3N(H)N2]+ and [CF3N(H)N2]+ were prepared as their low‐temperature stable [AsF6]? salts by protonation of azidomethane and azidotrifluoromethane in superacidic systems. They were characterized by NMR and Raman spectroscopy. Unequivocal proof of the protonation site was obtained by the crystal structures of both salts, confirming the formation of alkylamino diazonium ions. The Lewis adducts CH3N3?AsF5 and CF3N3?AsF5 were also prepared and characterized by low‐temperature NMR and Raman spectroscopy, and also by X‐ray structure determination for CH3N3?AsF5. Electronic structure calculations were performed to provide additional insights. Attempted electrophilic amination of aromatics such as benzene and toluene with methyl‐ and trifluoromethylamino diazonium ions were unsuccessful.  相似文献   

17.
In contrast to ruthenocene [Ru(η5‐C5H5)2] and dimethylruthenocene [Ru(η5‐C5H4Me)2] ( 7 ), chemical oxidation of highly strained, ring‐tilted [2]ruthenocenophane [Ru(η5‐C5H4)2(CH2)2] ( 5 ) and slightly strained [3]ruthenocenophane [Ru(η5‐C5H4)2(CH2)3] ( 6 ) with cationic oxidants containing the non‐coordinating [B(C6F5)4]? anion was found to afford stable and isolable metal?metal bonded dicationic dimer salts [Ru(η5‐C5H4)2(CH2)2]2[B(C6F5)4]2 ( 8 ) and [Ru(η5‐C5H4)2(CH2)3]2[B(C6F5)4]2 ( 17 ), respectively. Cyclic voltammetry and DFT studies indicated that the oxidation potential, propensity for dimerization, and strength of the resulting Ru?Ru bond is strongly dependent on the degree of tilt present in 5 and 6 and thereby degree of exposure of the Ru center. Cleavage of the Ru?Ru bond in 8 was achieved through reaction with the radical source [(CH3)2NC(S)S?SC(S)N(CH3)2] (thiram), affording unusual dimer [(CH3)2NCS2Ru(η5‐C5H4)(η3‐C5H4)C2H4]2[B(C6F5)4]2 ( 9 ) through a haptotropic η5–η3 ring‐slippage followed by an apparent [2+2] cyclodimerization of the cyclopentadienyl ligand. Analogs of possible intermediates in the reaction pathway [C6H5ERu(η5‐C5H4)2C2H4][B(C6F5)4] [E=S ( 15 ) or Se ( 16 )] were synthesized through reaction of 8 with C6H5E?EC6H5 (E=S or Se).  相似文献   

18.
The potential of a dicationic strontium ansa-arene complex for Lewis acid catalysis has been explored. The key to its synthesis was a simple salt metathesis from SrI2 and 2 Ag[Al(ORF)4], giving the base-free strontium-perfluoroalkoxyaluminate Sr[Al(ORF)4]2 (ORF=OC(CF3)3). Addition of an ansa-arene yielded the highly Lewis acidic, dicationic strontium ansa-arene complex. In preliminary experiments, the complex was successfully applied as a catalyst in CO2-reduction to CH4 and a surprisingly controlled isobutylene polymerization reaction.  相似文献   

19.
With P(CH3)3 as the probe molecule adsorbed on titanium silicalite (TS-1) zeolite, the special and important role of T12 site in MFI-type zeolite was clearly elucidated. There are altogether three active sites present in TS-1 zeolite with Ti at the T12 site. Owing to the preferential adsorption of probe molecules on the first Brönsted acidic site, the Ti12 center will probably fail to show Lewis acidity. The ionic [HP(CH3)3]+ species can be stabilized by the first or second Brönsted acidic site, with the former energetically favored. The latter was formed through the transfer of the ionic [HP(CH3)3]+ species from the first to the second Brönsted acidic site.  相似文献   

20.
High-yield syntheses up to molar scales for salts of [BH(CN)3] ( 2 ) and [BH2(CN)2] ( 3 ) starting from commercially available Na[BH4] (Na 5 ), Na[BH3(CN)] (Na 4 ), BCl3, (CH3)3SiCN, and KCN were developed. Direct conversion of Na 5 into K 2 was accomplished with (CH3)3SiCN and (CH3)3SiCl as a catalyst in an autoclave. Alternatively, Na 5 is converted into Na[BH{OC(O)R}3] (R=alkyl) that is more reactive towards (CH3)3SiCN and thus provides an easy access to salts of 2 . Some reaction intermediates were identified, for example, Na[BH(CN){OC(O)Et}2] (Na 7 b ) and Na[BH(CN)2{OC(O)Et}] (Na 8 b ). A third entry to 2 and 3 uses ether adducts of BHCl2 or BH2Cl such as the commercial 1,4-dioxane adducts that react with KCN and (CH3)3SiCN. Alkali metal salts of 2 and 3 are convenient starting materials for organic salts, especially for low viscosity ionic liquids (ILs). [EMIm] 3 has the lowest viscosity and highest conductivity with 10.2 mPa s and 32.6 mS cm−1 at 20 °C known for non-protic ILs. The ILs are thermally, chemically, and electrochemically robust. These properties are crucial for applications in electrochemical devices, for example, dye-sensitized solar cells (Grätzel cells).  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