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 375 毫秒
1.
The application of hypergolic ionic liquids as propellant fuels is a newly emerging area in the fields of chemistry and propulsion science. Herein, a new class of [imidazolyl?amine?BH2]+‐cation‐based ionic liquids, which included fuel‐rich anions, such as dicyanamide (N(CN)2?) and cyanoborohydride (BH3CN?) anions, were synthesized and characterized. As expected, all of the ionic liquids exhibited spontaneous combustion upon contact with the oxidizer 100 % HNO3. The densities of these ionic liquids varied from 0.99–1.12 g cm?3, and the heats of formation, predicted based on Gaussian 09 calculations, were between ?707.7 and 241.8 kJ mol?1. Among them, the salt of compound 5 , that is, (1‐allyl‐1H‐imidazole‐3‐yl)?(trimethylamine)?dihydroboronium dicyanamide, exhibited the lowest viscosity (168 MPa s), good thermal properties (TgTd>130 °C), and the shortest ignition‐delay time (18 ms) with 100 % HNO3. These ionic fuels, as “green” replacements for toxic hydrazine‐derivatives, may have potential applications as bipropellant formulations.  相似文献   

2.
Spectroscopic Properties of HCl Adducts of the Di(phthalocyaninato(2–))lanthanide Acids Thin films of bis(triphenylphosphine)iminiumdi(phthalocyaninato(2–))lanthanidates(III), (PNP)[Ln(Pc2)2] (Ln = La…(? Ce, Pm)…Lu) react with hydrogen chloride yielding the green acid adduct [HLn(Pc2?)2] · xHCl. The typical π–π* transitions of the Pc2? ligand are observed in the UV-VIS spectra (B: ~ 14500 cm?1; Q: ~ 30300 cm?1); these are broadened and shifted to lower energy with respect to those of the precursor. A N? H stretching vibration at ~ 3170 cm?1 as well as a H? N? C deformation vibration at ~ 1200 cm?1 in the MIR spectra are diagnostic for these HCl adducts.  相似文献   

3.
Reactions of bis(phosphinimino)amines LH and L′H with Me2S ? BH2Cl afforded chloroborane complexes LBHCl ( 1 ) and L′BHCl ( 2 ), and the reaction of L′H with BH3 ? Me2S gave a dihydridoborane complex L′BH2 ( 3 ) (LH=[{(2,4,6‐Me3C6H2N)P(Ph2)}2N]H and L′H=[{(2,6‐iPr2C6H3N)P(Ph2)}2N]H). Furthermore, abstraction of a hydride ion from L′BH2 ( 3 ) and LBH2 ( 4 ) mediated by Lewis acid B(C6F5)3 or the weakly coordinating ion pair [Ph3C][B(C6F5)4] smoothly yielded a series of borenium hydride cations: [L′BH]+[HB(C6F5)3]? ( 5 ), [L′BH]+[B(C6F5)4]? ( 6 ), [LBH]+[HB(C6F5)3]? ( 7 ), and [LBH]+[B(C6F5)4]? ( 8 ). Synthesis of a chloroborenium species [LBCl]+[BCl4]? ( 9 ) without involvement of a weakly coordinating anion was also demonstrated from a reaction of LBH2 ( 4 ) with three equivalents of BCl3. It is clear from this study that the sterically bulky strong donor bis(phosphinimino)amide ligand plays a crucial role in facilitating the synthesis and stabilization of these three‐coordinated cationic species of boron. Therefore, the present synthetic approach is not dependent on the requirement of weakly coordinating anions; even simple BCl4? can act as a counteranion with borenium cations. The high Lewis acidity of the boron atom in complex 8 enables the formation of an adduct with 4‐dimethylaminopyridine (DMAP), [LBH ? (DMAP)]+[B(C6F5)4]? ( 10 ). The solid‐state structures of complexes 1 , 5 , and 9 were investigated by means of single‐crystal X‐ray structural analysis.  相似文献   

4.
The possibility of electron binding to five molecules (i.e., F3N → BH3, H2FN → BH3, HF2N → BH3, H3N → BH2F, H3N → BHF2) was studied at the coupled cluster level of theory with single, double, and noniterative triple excitations and compared to earlier results for H3N → BH3 and H3N → BF3. All these neutral complexes involve dative bonds that are responsible for significant polarization of these species that generates large dipole moments. As a consequence, all of the neutral systems studied, except F3N → BH3, support electronically stable dipole‐bound anionic states whose calculated vertical electron detachment energies are 648 cm?1 ([H2FN → BH3]?), 234 cm?1 ([HF2N → BH3]?), 1207 cm?1 ([H3N → BH2F]?), and 1484 cm?1 ([H3N → BHF2]?). In addition, we present numerical results for a model designed to mimic charge–transfer (CT) and show that the electron binding energy correlates with the magnitude of the charge flow in the CT complex. © 2003 Wiley Periodicals, Inc. Int J Quantum Chem, 2003  相似文献   

5.
Crystals of the novel title arsenic(III)–phthalocyanine complex, [As(C32H16N8)]2[As2I8] or [AsPc]2[As2I8], where Pc is the phthalocyaninate(2−) macrocycle, have been obtained from the reaction of pure powdered arsenic with phthalonitrile under oxidizing conditions (iodine vapour) at 463 K. The crystals are formed by separate but inter­acting [AsPc]+ cations and centrosymmetric [As2I8]2− anions. The As atom of the [AsPc]+ ion is bonded to the four isoindole N atoms of the Pc macrocycle and lies 0.762 (1) Å out of their plane. The anionic part of the complex consists of two [AsI4] units joined together into a centrosymmetric [As2I8]2− counter‐ion. The arrangement of oppositely charged moieties, viz. [AsPc]+ and [As2I8]2−, in the crystal structure is determined mainly by their ionic attractions and by π–π inter­actions between the aromatic phthalocyaninate(2−) macrocycles.  相似文献   

6.
Following the development in the synthesis of subvalent cluster compounds, we report on the use of three different classes of room-temperature ionic liquids for the synthesis of the pentabismuth-tris(tetragallate) salt, Bi5[GaCl4]3, characterized by X-ray diffraction. The Bi5[GaCl4]3 salt was prepared by reduction of BiCl3 using gallium metal in ionic liquid reaction media containing a strong Lewis acid, GaCl3. The ionic liquids; trihexyltetradecyl phosphonium chloride [Th-Td-P+]Cl?, 1-dodecyl-3-methylimidazolium chloride [Dod-Me-Im+]Cl? and N-butyl-N-methylpyrrolidinium chloride [Bu-Me-Pyrr+]Cl? from three of the main classes of ionic liquids were used in synthesis. Reactions using ionic liquids composed of the trihexyltetradecyl phosphonium cation [Th-Td-P+] and the anions; tetrafluoroborate [BF4 ?], bis(trifluoro-methyl sulfonyl) imide [(Tf)2N?] and hexafluorophosphate [PF6 ?] were also investigated.  相似文献   

7.
Alkylidynephosphanes and -arsanes. I [P ≡ C? S]?[Li(dme)3]+ – Synthesis and Structure O,O′-Diethyl thiocarbonate and bis(tetrahydrofuran)-lithium bis(trimethylsilyl)phosphanide dissolved in 1,2-dimethoxyethane, react below 0°C to give ethoxy trimethylsilane and tris(1,2-dimethoxyethane-O,O′)lithium 2λ3-phosphaethynylsulfanide – [P≡C? S]? [Li(dme)3]+ – ( 1a ). Apart from bis(trimethylsilyl)sulfane or carbon oxide sulfide, dark red concentrated solutions of λ3-phosphaalkyne 1 are also obtained from reactions of carbon disulfide with bis(tetrahydrofuran)-lithium bis(trimethylsilyl)phosphanide or with the homologous lithoxy-methylidynephosphane ( 2 ) [1]. The ir spectrum shows two absorptions at 1762 and 747 cm?1 characteristic for the P≡C and C? S stretching vibrations. The nmr parameters {δ(31P) ? 121.3; δ(13C) 190.8 ppm; 1JCP 18.2 Hz} resemble much more values of diorganylamino-2λ3-phosphaalkynes than those of bis(1,2-dimethoxyethane-O,O′)lithoxy-methylidyne-phosphane ( 2a ). As found by an X-ray structure analysis (P21/c; a = 1192.6(16); b = 1239.1(19); c = 1414.8(26) pm; β = 105.91(13)° at ?100 ± 3°C; Z = 4 formula units; wR = 0.064) of pale yellow crystals (mp. + 16°C) isolated from the reaction with O,O′-diethyl thiocarbonate, the solid is built up of separate [P≡C? S]? and [Li(dme)3]+ ions. Typical bond lengths and angles are: P≡C 155.5(11); C? S 162.0(11); Li? O 206.4(17) to 220.3(20) pm; P≡C? S 178.9(7)°.  相似文献   

8.
Ruthenium(II)-Phthalocyaninates(1–): Synthesis and Properties of (Halo)(carbonyl)phthalocyaninato(1–)ruthenium(II) Brown-violet (halo)(carbonyl)phthalocyaninato(1–)ruthenium(II), [Ru(X)(CO)Pc?] (X = Cl, Br) is prepared by oxidation of [Ru(X)(CO)Pc2?]? with the corresponding halogen or dibenzoylperoxide. The eff. magnetic moment μeff = 1.74 (X = Cl), 1.68 μB (Br) confirms the presence of a low-spin RuII complex of the Pc? radical. Accordingly, only the first ring oxidation at ~0.64 V and the first ring reduction at ~ ?1.19 V is observed in the cyclovoltammogram of [Ru(X)(CO)Pc2?]?. The UV-VIS-NIR spectra characterizing a monomeric Pc? radical with intense π-π* transitions at 14500, 19800, 25100 and 33900 cm?1 are compared with those of [Ru(Cl)2Pc?] and of monomeric as well as dimeric [Zn(Cl)Pc?]. The IR and resonance Raman(RR) spectra are characteristic for a Pc? radical, too. Diagnostic in-plane vibrations of the Pc? ligand are in the IR spectrum at 1071, 1359, 1445 cm?1 and in the RR spectrum (λ0 = 488.0 nm) at 567, 1597 cm?1. v(C? O) at 1950 cm?1 and v(Ru? X) at 260 (X = Cl) resp. 184 cm?1 (X = Br) are observed only in the IR spectrum.  相似文献   

9.
Ten [C8C1Im]+ (1‐methyl‐3‐octylimidazolium)‐based ionic liquids with anions Cl?, Br?, I?, [NO3]?, [BF4]?, [TfO]?, [PF6]?, [Tf2N]?, [Pf2N]?, and [FAP]? (TfO=trifluoromethylsulfonate, Tf2N=bis(trifluoromethylsulfonyl)imide, Pf2N=bis(pentafluoroethylsulfonyl)imide, FAP=tris(pentafluoroethyl)trifluorophosphate) and two [C8C1C1Im]+ (1,2‐dimethyl‐3‐octylimidazolium)‐based ionic liquids with anions Br? and [Tf2N]? were investigated by using X‐ray photoelectron spectroscopy (XPS), NMR spectroscopy and theoretical calculations. While 1H NMR spectroscopy is found to probe very specifically the strongest hydrogen‐bond interaction between the hydrogen attached to the C2 position and the anion, a comparative XPS study provides first direct experimental evidence for cation–anion charge‐transfer phenomena in ionic liquids as a function of the ionic liquid’s anion. These charge‐transfer effects are found to be surprisingly similar for [C8C1Im]+ and [C8C1C1Im]+ salts of the same anion, which in combination with theoretical calculations leads to the conclusion that hydrogen bonding and charge transfer occur independently from each other, but are both more pronounced for small and more strongly coordinating anions, and are greatly reduced in the case of large and weakly coordinating anions.  相似文献   

10.
Due to its high hydrogen density (14.8 wt %) and low dehydrogenation peak temperature (130 °C), Zr(BH4)4 ? 8 NH3 is considered to be one of the most promising hydrogen‐storage materials. To further decrease its dehydrogenation temperature and suppress its ammonia release, a strategy of introducing LiBH4 and Mg(BH4)2 was applied to this system. Zr(BH4)4 ? 8 NH3–4 LiBH4 and Zr(BH4)4 ? 8 NH3–2 Mg(BH4)2 composites showed main dehydrogenation peaks centered at 81 and 106 °C as well as high hydrogen purities of 99.3 and 99.8 mol % H2, respectively. Isothermal measurements showed that 6.6 wt % (within 60 min) and 5.5 wt % (within 360 min) of hydrogen were released at 100 °C from Zr(BH4)4 ? 8 NH3–4 LiBH4 and Zr(BH4)4 ? 8 NH3–2 Mg(BH4)2, respectively. The lower dehydrogenation temperatures and improved hydrogen purities could be attributed to the formation of the diammoniate of diborane for Zr(BH4)4 ? 8 NH3–4 LiBH4, and the partial transfer of NH3 groups from Zr(BH4)4 ? 8 NH3 to Mg(BH4)2 for Zr(BH4)4 ? 8 NH3–2 Mg(BH4)2, which result in balanced numbers of BH4 and NH3 groups and a more active Hδ+ ??? ?δH interaction. These advanced dehydrogenation properties make these two composites promising candidates as hydrogen‐storage materials.  相似文献   

11.
The novel PtII–dibenzo‐18‐crown‐6 (DB18C6) title complex, μ‐[tetrakis­(thio­cyanato‐S)­platinum(II)]‐N:N′‐bis{[2,5,8,­15,18,21‐hexa­oxa­tri­cyclo­[20.4.0.19,14]­hexa­cosa‐1(22),9(14),10,12,23,25‐hexaene‐κ6O]­potassium(I)}, [K(C20H24O6)]2[Pt(SCN)4], has been isolated and characterized by X‐ray diffraction analysis. The structure analysis shows that the complex displays a quasi‐one‐dimensional infinite chain of two [K(DB18C6)]+ complex cations and a [Pt(SCN)4]2? anion, bridged by K+?π interactions between adjacent [K(DB18C6)]+ units.  相似文献   

12.
Phosphorane Iminato-Trichloroselenates(II): Syntheses and Crystal Structures of [SeCl(NPPh3)2]+SeCl3? and [Me3SiN(H)PMe3]2+[Se2Cl6]2? [SeCl(NPPh3)2]+SeCl3? has been synthesized by the reaction of Se2Cl2 with Me3SiNPPh3 in acetonitrile solution, forming orangered crystals, whereas red crystals of [Me3SiN(H)PMe3]2+[Se2Cl6]2? were obtained by the reaction of Me3SiNPMe3 with SeOCl2 in acetonitrile solution. Both complexes were characterized by X-ray structure determinations. [SeCl(NPPh3)2]+SeCl3?: Space group P21/n, Z = 4, structure solution with 7 489 observed unique reflections, R = 0.057. Lattice dimensions at ?60°C: a = 1 117.0; b = 2 241, c = 1 407.5 pm, β = 95.61°. In the cation [SeCl(NPPh3)2]+ the selenium atom is φ-tetrahedrally coordinated by the chlorine atom and by the nitrogen atoms of the phosphorane iminato ligands, whereas the anion SeCl3? has a T-shaped structure with φ-trigonal-bipyramidale surrounding of the selenium atom. [Me3SiN(H)PMe3]2+[Se2Cl6]2?: Space group P21/c, Z = 4, structure solution with 2 093 observed unique reflections, R = 0.080. Lattice dimensions at ?70°C: a = 956, b = 828, c = 1 973 pm, β = 93.80°. The structure consists of [Me3SiN(H)PMe3]+ ions and planar [Se2Cl6]2? anions, in which the selenium atoms are bridged nearly symmetrically by two chlorine atoms.  相似文献   

13.
The object of this study was to measure the liquid–liquid equilibria (LLE) data of binary mixtures containing ionic liquids and citrus essential oil. We investigated linalool as the citrus essential oil, and 1-alkyl-3-methyl-imidazolium bis(trifluoromethanesulfonyl)imide ([C n MIM]+[TFSI]?) as the ionic liquid. Firstly, the experimental apparatus and procedure for the LLE measurement of mixtures containing ionic liquids were verified by measuring the LLE of the binary mixture 1-hexyl-3-methyl-imidazolium bis(trifluoromethanesulfonyl)imide ([C6MIM]+[TFSI]?) + 1-hexanol as a reference test system recommended by Marsh et al. (Pure Appl Chem 81:781–789, 2009). Next, the LLE data for IL + linalool were obtained, and the LLE data of two binary mixtures 1-butyl-3-methyl-imidazolium bis(trifluoromethanesulfonyl)imide ([C4MIM]+[TFSI]?) or [C6MIM]+[TFSI]? + linalool were determined. The experimental LLE data were satisfactorily represented by the non-random two-liquid model.  相似文献   

14.
Crystals of a new antimony(III) phthalocyanine complex with the formula [Sb(C32H16N8)]4(Sb4I16), or (SbPc)·[Sb4I16]4?, where Pc is phthalocyaninate, have been obtained by the reaction of pure powdered antimony with phthalo­nitrile under a stream of iodine vapour. The crystals are built up from separate but interacting (SbPc)+ cations and centrosymmetric [Sb4I16]4? anions. Each Sb atom of two independent (SbPc)+ units is bonded to the four iso­indole N atoms of the phthalocyaninate(2?) macrocycle and lies 1.0 Å out of the plane defined by four iso­indole N atoms. The anionic part of the complex consists of four SbI6 distorted octahedra joined together into a centrosymmetric [Sb4I16]4? anion. The arrangement of oppositely charged moieties in the crystal is mainly determined by ionic attraction and by a set of distinct donor–acceptor interactions between (SbPc)+ and [Sb4I16]4? ions.  相似文献   

15.
A Pd‐catalyzed Suzuki cross‐coupling of arylboronic acids with Yagupolskii–Umemoto reagents was explored. In contrary to trifluoromethylations, the Pd‐catalyzed reaction of R?B(OH)2 and [Ar2SCF3]+[OTf]? provided the arylation products (R?Ar) in good to high yields. The reaction confirms that the S?Ar bonds of [Ar2SCF3]+[OTf]? can be readily cleaved in the presence of Pd complexes. The relatively electron‐poor aryl groups of asymmetric [Ar1Ar2SCF3]+[OTf]? salts are more favorably transferred compared to the electron‐rich ones. This reaction represents the first report of utilization of [Ar2SCF3]+[OTf]? as arylation reagents in organic synthesis.  相似文献   

16.
[Na · Triglyme]2[S(BH3)4]: a Salt of the New Anion Tetrakis(borane)sulfate(2? ). Crystal Structure and Theoretical Investigation of the Structure Na[H3B-m?2-S(B2H5)] 1 is produced by the reaction between NaSH and THF · BH3, under dehydrogenation. 1 is also formed as the first 11B-NMR-spectroscopically detectable reaction product by the reaction between anhydrous Na2S and THF · BH3. Adducts of BH3 with the S2? ion are not detectable in THF. The anion [S(BH3)4]2? can however be obtained, by the addition of NaBH4 to 1 in diglyme or triglyme respectively: [Na — Triglyme]2[S(BH3)4] 2. 2 crystallizes in the monoclinic space group P21/n (Nr. 14). Structural data of 1 and 2 have been calculated by SCF methods. The anion of 2 may be viewed either as an adduct of B2H6 with S2?, or as a bridge substituted thia derivative of B2H7?; furthermore the anion of 2 is isoelectronic and isostructural with the SO ion.  相似文献   

17.
The syntheses of lithium and alkaline earth metal complexes with the bis(borane‐diphenylphosphanyl)amido ligand ( 1 ‐ H ) of molecular formulas [{κ2‐N(PPh2(BH3))2}Li(THF)2] ( 2 ) and [{κ3‐N(PPh2(BH3))2}2M(THF)2] [(M = Ca ( 3 ), Sr ( 4 ), Ba ( 5 )] are reported. The lithium complex 2 was obtained by treatment of bis(borane‐diphenylphosphanyl)amine ( 1 ‐ H ) with lithium bis(trimethylsilyl)amide in a 1:1 molar ratio via the silylamine elimination method. The corresponding homoleptic alkaline earth metal complexes 3 – 5 were prepared by two synthetic routes – first, the treatment of metal bis(trimethylsilyl)amide and protio ligand 1 ‐ H via the elimination of silylamine, and second, through salt metathesis reaction involving respective metal diiodides and lithium salt 2 . The molecular structures of lithium complex 2 and barium complex 5 were established by single‐crystal X‐ray diffraction analysis. In the solid‐state structure of 2 , the lithium ion is ligated by amido nitrogen atoms and hydrogen atoms of the BH3 group in κ2‐coordination of the ligand 1 resulting in a distorted tetrahedral geometry around the lithium ion. However, in complex 5 , κ3‐coordination of the ligand 1 was observed, and the barium ion adopted a distorted octahedral arrangement. The metal complex 5 was tested as catalyst for the ring opening polymerization of ?‐caprolactone. High activity for the barium complex 5 towards ring opening polymerization (ROP) of ?‐caprolactone with a narrow polydispersity index was observed. Additionally, first‐principle calculations to investigate the structure and coordination properties of alkaline earth metal complexes 3 – 5 as a comparative study between the experimental and theoretical findings were described.  相似文献   

18.
Ionic liquids have become commonplace materials found in research laboratories the world over, and are increasingly utilised in studies featuring water as co‐solvent. It is reported herein that proton activities, aH+, originating from auto‐protolysis of H2O molecules, are significantly altered in mixtures with common ionic liquids comprised of Cl?, [HSO4]?, [CH3SO4]?, [CH3COO]?, [BF4]?, relative to pure water. paH+ values, recorded in partially aqueous media as ?log(aH+), are observed over a wide range (~0–13) as a result of hydrolysis (or acid dissociation) of liquid salt ions to their associated parent molecules (or conjugate bases). Brønsted–Lowry acid–base character of ionic liquid ions observed is rooted in equilibria known to govern the highly developed aqueous chemistry of classical organic and inorganic salts, as their well‐known aqueous pKs dictate. Classical salt behaviour observed for both protic and aprotic ions in the presence of water suggests appropriate attention need be given to relevant chemical systems in order to exploit, or avoid, the nature of the medium formed.  相似文献   

19.
The title compounds are salts of the general form (Q+)2[Zn(dmit)2]2?, where dmit corresponds to the ligand (C3S5)? present in both and Q+ to the counter‐cations (nBu4N)+ [or C16H36N+] and (Ph4As)+ [or C24H20As+], respectively. In the first case, Zn is in the 4e special positions of space group C2/c and hence the [Zn(dmit)2]2? dianion possesses twofold axial crystallographic symmetry. Including these, there are now 11 known examples of [Zn(dmit)2]2? or its analogues, with O replacing the exocyclic thione S, and [Zn(dmio)2]2? dianions in nine structures with various Q. Comparison of these reveals a remarkable variation in details of the conformation which the dianion may adopt in terms of Zn coordination, equivalence of the Zn—S bond lengths, displacement of Zn from the plane of the ligand and overall dianion shape.  相似文献   

20.
In this work, the geometrical and electronic properties of the mono cationic ionic liquid 1‐hexyl‐3‐methylimidazolium halides ([C6(mim)]+_X?, X=Cl, Br and I) and dicationic ionic liquid 1,3‐bis[3‐methylimidazolium‐1‐yl]hexane halides ([C6(mim)2X2], X=Cl, Br and I) were studied using the density functional theory (DFT). The most stable conformer of these two types ionic liquids (IL) are determined and compared with each other. Results show that in the most stable conformers, in both monocationic ILs and dicationic ILs, the Cl? and Br? anions prefer to locate almost in the plane of the imidazolium ring whereas the I? anion prefers nearly vertical location respect to the imidazolium ring plan. Comparison of hydrogen bonding and ionic interactions in these two types of ionic liquids reveals that these ionic liquids can be formed hydrogen bond by Cl? and Br? anion. The calculated thermodynamic functions show that the interaction of cation — anion pair in the dicationic ionic liquids are more than monocationic ionic liquids and these interactions decrease with increasing the halide anion atomic weight.  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