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.
Weakly polar–polar isosteric pairs of 12-vertex p-carborane [closo-1,12-C2B10H12] (1[12]) and monocarbaborate [closo-1-CB11H12]? (2[12]) nematic liquid crystals, in which the difference in the calculated molecular dipole moment is 11.3 D, were synthesised, and the effect of the dipole moment on nematic phase stability was investigated. The trend observed for the 12-vertex series ([12]) was identical to that of the previously investigated 10-vertex series ([10]) containing [closo-1,10-C2B8H10] (1[10]) and [closo-1-CB9H10]? (2[10]): the uniform increase in the molecular dipole moment in the pairs of mesogens does not correspond to a uniform change in the clearing temperature, TNI. This demonstrates the role of a remote substituent in modulating the intermolecular dipole–dipole interactions. The magnitude of such interactions was calculated (using density functional theory methods) for a pair of polar (2[12]d2[12]d) and an analogous pair of weakly polar (1[12]d1[12]d) molecules. All results for the 12-vertex series ([12]) were analysed relative to the 10-vertex analogues ([10]).  相似文献   

2.
Abstract

The reactions of cyclotriphosphazene (1) with 2-(2-hydroxyethylamino)-ethanol (2) were investigated. 2-(2-hydroxyethylamino)-ethanol (2) is a tri-functional reagent consisting of both aliphatic hydroxyl and the secondary amino groups and its nucleophilic substitution reactions with cylotriphosphazene can lead to different product types; open chain, spiro, ansa, bridged and their mixtures. The reactions with one, two and three equimolar ratios of 2-(2-hydroxyethylamino)-ethanol, in the presence of NaH at 0–10?°C and at room temperature gave the following cyclotriphosphazene derivatives: one mono-spiro, N3P3Cl4[O–(CH2)2–NH–(CH2)2–O] (3, 1:1, r.t.); its isomer mono-ansa (5, 1:1, r.t.); one dispiro, N3P3Cl2[O–(CH2)2–NH–(CH2)2–O]2 (4, 1:1, r.t.); its isomer spiro-ansa (6, 1:2, r.t.); and one single-bridged compound with spiro substituted units, N3P3Cl3[O–(CH2)2–NH–(CH2)2–O]3N3P3Cl3 (7, 1:3, at 0–10?°C); as well as single-, N3P3Cl5[O–(CH2)2–NH–(CH2)2–O]N3P3Cl5 (8, 1:2, r.t.), double-, N3P3Cl4[O–(CH2)2–NH–(CH2)2–O]2N3P3Cl4 (9, 1:2, r.t.), and tri-bridged, N3P3Cl3[O–(CH2)2–NH–(CH2)2–O]3N3P3Cl3 (10, 1:3, at 0–10?°C) derivatives. Triple-bridged derivative is the major product in this system. The structures of the novel-derived compounds were characterized by TLC-MS, FT-IR, elemental analysis, 1H, and 31P NMR spectral.  相似文献   

3.
《Polyhedron》1999,18(26):3533-3544
[Tris(1,3-dithiole-2-thione-4,5-dithiolato)stannate]2−, [Q]2[Sn(C3S5)3], [tris(1,3-dithiole-2-one-4,5-dithiolato)stannate]2−, [Q]2[Sn(C3S4O)3], and [tris(1,3-dithiole-2-thione-4,5-diselenolato)stannate]2− [Q]2[Sn(C3S3Se2)3], complexes, have been synthesised and characterised. Crystal structure determinations of [Q]2[Sn(C3S5)3] (Q=1,4-dimethylpyridinium, monoclinic and orthorhombic forms; NMe4, NEt4, and PPh4) and [NEt4]2[Sn(C3S4O)3] revealed variations in the overall dianion structures. The geometry about tin in each case is essentially octahedral with the chelate bite angles in the range 80.7(5)–87.45(4)°: the range of Sn–S distances is 2.5207(18)–2.571(17) Å. A statistical analysis, carried out on the crystal structure data for the six complexes, indicated that the most critical factors in controlling the overall shape of the dianion were the distances of the Sn atom from the dithiolate ligand planes [Sn–OOP]. Interanionic S⋯S interactions, within the sum of the van der Waals’ radii for two S atoms, are affected by the size of the cation, Q; the secondary connectivity is 3-dimensional for the smallest cations, Q=1,4-dimethylpyridinium and NMe4, in chains for the somewhat larger cation, NEt4 and is absent for the still larger, PPh4 cation.  相似文献   

4.
Synthesis, Crystal Structure, Vibrational Spectra, and Normal Coordinate Analysis of (Ph4P)2[OsN(N3)5] and 15N NMR Chemical Shifts of Nitridoosmates(VI, VIII) The treatment of (Ph4P)[OsNCl4] with NaN3 yields (Ph4P)2[OsN(N3)5], which crystal structure has been determined by single crystal X‐ray diffraction analysis (monoclinic, space group P 21/a, a = 20.484(6), b = 11.168(1), c = 20.666(4) Å, β = 97.35(3)°, Z = 4). The IR and Raman vibrations were assigned by a normal coordinate analysis based on the molecular parameters of the X‐ray determination. The valence force constants are fd(Os≡N) = 8.52, fd(Os–Nα) = 1.99, fd(Nα–Nβ) = 12.42, fd(Nβ–Nγ) = 12.73 and for the azido ligand in trans‐position to the nitrido group fd(Os–Nα · ) = 1.84, fd(Nα · –Nβ · ) = 11.91, fd(Nβ · –Nγ · ) = 12.18 mdyn/Å. The 15N NMR spectra of various nitridoosmates reveal the chemical shifts δ(15N) for K[OsO315N] = 387.6, K2[Os15NCl5] = 446.7, (Ph4P)[Os15NCl4] = 352.9, [(n‐C6H13)4N]2[Os15N(N3)5] = 307.3 and for [(n‐Pr)4N]2[Os15N(15NCO)5] = 483,7 (Os≡N), –417,7 (OsNCOeq) und –392,8 ppm (OsNCOax).  相似文献   

5.
Abstract : γ-Butyrolactone and γ-butyrolactam were reacted in the superacidic systems XF/MF5 (X=H, D; M=As, Sb). Salts of the monoprotonated species of γ-butyrolactone were obtained in terms of [(CH2)3OCOH]+[AsF6], [(CH2)3OCOH]+[SbF6] and [(CH2)3OCOD]+[AsF6] and the analogous lactam salts in terms of [(CH2)3NHCOH]+[AsF6], [(CH2)3NHCOH]+[SbF6] and [(CH2)3NDCOD]+[AsF6]. The salts were characterized by low temperature Raman and infrared spectroscopy and for both protonated hexafluoridoarsenates, [(CH2)3OCOH]+[AsF6] and [(CH2)3NHCOH]+[AsF6], single-crystal X-ray structure analyses were conducted. In addition to the experimental results, quantum chemical calculations were performed on the B3LYP/aug-cc-pVTZ level of theory. As in both crystal structures C⋅⋅⋅F contacts were observed, the nature of these contacts is discussed with Mapped Electrostatic Potential as a rate of strength.  相似文献   

6.
Coordination equilibrium constants (K NiS) of some donor solvent molecules to 1,4,7,10-tetramethyl-1,4,7,10-tetraazacyclododecanenickel(II) ([Ni(Me4[12]aneN4)]2+) were determined in nitrobenzene (a noncoordinating bulk solvent). The first (K NiS1) and second stepwise coordination equilibrium constants (K NiS2) for 1,4,7,10-tetraazacyclododecanenickel(II) ([Ni([12]aneN4)]2+), 1,4,8,11-tetraazac yclotetradecane- nickel(II) ([Ni([14] aneN4)]2+), 1,4,8,11-tetrathiacyclotetra-decanenickel(II) ([Ni([14]aneS4)]2+) were also reinvestigated. The K NiS values for [Ni(Me4[12]aneN4)]2+ were compared to those of [Ni([12]aneN4)]2+, (1R,4S, 8R,11S)-1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecanenickel(II) (R,S,R,S-[Ni(Me4[14]aneN4)]2+), R,R,S,S-[Ni(Me4[14]aneN4)]2+, [Ni([14]aneN4)]2+, and [Ni([14]aneS4)]2+. Coordination of pyridine (Py), N,N,N′,N′-tetramethylurea (TMU), and N,N-dimethylacetamide (DMA) to [Ni(Me4[12]aneN4)]2+ was observed, although these donor solvent molecules did not coordinate to R,S,R,S-[Ni(Me4[14]aneN4)]2+. The K NiS values for Py, TMU, and DMA are 7.9, 2.8, and 9.0 dm3⋅mol−1, respectively. Some hydrogen-bonding waters were coordinated to R,S,R,S-[Ni(Me4[14]aneN4)]2+, but such waters did not coordinate to [Ni(Me4[12] aneN4)]2+. Also, the K NiS2 values were larger than the corresponding K NiS1 values for [Ni([14]aneS4)]2+. Furthermore, the K NiS1 values for [Ni([12]aneN4)]2+ were the largest among these nickel(II) complex cations. The K NiS, K NiS1, and K NiS2 values are discussed in terms of properties of the donor solvents and steric strains of these nickel(II) complex cations.  相似文献   

7.
The kinetics of dimerization of arylmercurials XC6H4HgCl (X ? H, p-CH3, m-CH3, p-Cl, m-OCH3, m-CO2C2H5, and o-OCH3) in the presence of [ClRh(CO)2]2 was studied in hexamethylphosphoramide (HMPA). The experimental rate law obtained is ?d[ArHgCl]/dt=k[ArHgCl]2. The kinetic parameters of these reactions have been reported, and the variation, therein, has been explained on the basis of steric effect of substituents. The apparent activation energy E is linearly proportional to pre-exponential factor lnA. A most plausible mechanism has been proposed on the basis of experimental results. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
Chloro- and Polyselenoselenates(II): Synthesis, Structure, and Properties of [Ph3(C2H4OH)P]2[SeCl4] · MeCN, [Ph4P]2[Se2Cl6], and [Ph4P]2[Se(Se5)2] By symproportionation of elemental selenium and SeCl4 in polar protic solvents the novel chloroselenates(+II), [SeCl4]2? and [Se2Cl6]2?, could be stabilized; they were crystallized with voluminous organic cations. They were characterized from complete X-ray structure analysis. Yellow-orange [Ph3(C2H4OH)P]2[SeCl4] · MeCN (space group P1 , a = 10.535(4), b = 12.204(5), c = 16.845(6) Å, α = 77.09(3)°, β = 76.40(3)°, γ = 82.75(3)° at 140 K) contains in its crystal structure monomeric [SeCl4]2? anions with square-planar coordination of Se(+II). The mean Se? Cl bond length is 2.441 Å. In yellow [Ph4P]2[Se2Cl6] (space group P1 , a = 10.269(3), b = 10.836(4), c = 10.872(3) Å, α = 80.26(3)°, β = 79.84(2)°, γ = 72.21(3)° at 140 K) a dinuclear centrosymmetric [Se2Cl6]2? anion, also with square-planar coordinated Se(+II), is observed. The average terminal and bridging Se? Cl bond distances are 2.273 and 2.680 Å, respectively. From redox reactions of elemental Se with boranate/thiolate in ethanol/DMF the bis(pentaselenido)selenate(+II) anion [Se(Se5)2]2? was prepared as a novel type of a mixed-valent chalcogenide. In dark-red-brown [Ph4P]2[Se(Se5)2] (space group P21/n, a = 12.748(4), b = 14.659(5), c = 14.036(5) Å, β = 108.53(3)° at 140 K) centrosymmetric molecular [Se(Se5)2]2? anions with square-planar coordination of the central Se(+II) by two bidentate pentaselenide ligands is observed (mean Se? Se bond lengths: 2.658 Å at Se(+II), 2.322 Å in [Se5]2?). The resulting six-membered chelate rings with chair conformation are spirocyclically linked through the central Se(+II). The vibrational spectra of the new anions are reported.  相似文献   

9.
The reaction of [(η5‐L3)Ru(PPh3)2Cl], where; L3 = C9H7 ( 1 ), C5Me5 (Cp*) ( 2 ) with acetonitrile in the presence of [NH4][PF6] yielded cationic complexes [(η5‐L3)Ru(PPh3)2(CH3CN)][PF6]; L3= C9H7 ([3]PF6) and L3 = C5Me5 ([4]PF6), respectively. Complexes [3]PF6 and [4]PF6 reacts with some polypyridyl ligands viz, 2,3‐bis (α‐pyridyl) pyrazine (bpp), 2,3‐bis (α‐pyridyl) quinoxaline (bpq) yielding the complexes of the formulation [(η5‐L3)Ru(PPh3)(L2)]PF6 where; L3 = C9H7, L2 = bpp, ([5]PF6), L3 = C9H7, L2 = bpq, ([6]PF6); L3 = C5Me5, L2 = bpp, ([7]PF6) and bpq, ([8]PF6), respectively. However reaction of [(η5‐C9H7)Ru(PPh3)2(CH3CN)][PF6] ([3]PF6) with the sterically demanding polypyridyl ligands, viz. 2,4,6‐tris(2‐pyridyl)‐1,3,5‐triazine (tptz) or tetra‐2‐pyridyl‐1,4‐pyrazine (tppz) leads to the formation of unexpected complexes [Ru(PPh3)2(L2)(CH3CN)][PF6]2; L2 = tppz ([9](PF6)2), tptz ([11](PF6)2) and [Ru(PPh3)2(L2)Cl][PF6]; L2 = tppz ([10]PF6), tptz ([12]PF6). The complexes were isolated as their hexafluorophosphate salts. They have been characterized on the basis of micro analytical and spectroscopic data. The crystal structures of the representative complexes were established by X‐ray crystallography.  相似文献   

10.
Preparation of Tetramethylammonium Azidosulfite and Tetramethylammonium Cyanate Sulfur Dioxide‐Adduct, [(CH3)4N]+[SO2N3], [(CH3)4N]+[SO2OCN] and Crystal Structure of [(CH3)4N]+[SO2N3] Tetramethylammonium azide forms with sulfur dioxide an azidosulfite salt. It is characterized by NMR and vibrational spectroscopy and the crystal structure analysis. [(CH3)4N]+[SO2N3] crystallizes in the monoclinic space group P21/c with a = 551.3(1) pm, b = 1095.2(1) pm, c = 1465.0(1) pm, β = 100.63(1)°, and four formula units in the unit cell. The crystal structure possesses a strong S–N interaction between the N3– anions and the SO2 molecules. The S–N distance of 200.5(2) pm is longer than a covalent single S–N bond. The structure is compared with ab initio calculated data. Furthermore an adduct of tetrametylammonium cyanate and sulfur dioxide is reported. It is characterised by NMR and vibrational spectroscopy. The structure is calculated by ab initio methods.  相似文献   

11.
The visible absorption spectrum of the water soluble polynuclear metallamacrocyclic LaIII‐CuII complex La(H2O)3[15‐MCCu(II)Phalaha‐5](Cl)3 ( 1 ) based on α‐phenylalaninehydroxamic acid appears to be solvent‐ and ion‐sensitive. The copper(II) d–d transitions of the complex 1 dissolved in methanol, ethanol, water, dimethylformamide, dimethylsulfoxide, pyridine, and N‐methylpyrrolidone were studied. The chromophoric behavior of complex 1 was investigated in the presence of the Cl, Br, I, HSO4, CO32–, HCO3, H2PO4, CN, SCN, and N3 anions. A considerable change of the d–d transition of the central copper(II) atom was observed for the strongly coordinating cyanide and azide anions. In the presence of HSO4, the d–d intensity of copper(II) also decreased significantly. The molecular structure of La2(H2O)7[15‐MCCu(II)Phalaha‐5]2(SO4)4 ( 2 ), obtained as result of the substitution of the coordinated water molecules in 1 by the SO42– anions, was investigated by X‐ray crystallography.  相似文献   

12.
Monodisperse polymeric nanospheres, which consist of polystyrene cores and poly(ethylene glycol) (PEG) branches on their surfaces, were prepared by the dispersion copolymerization of styrene (St) with PEG macromonomers that had a methacryloyl (MMA-PEG) or p-vinylbenzyl (St-PEG) end group in various organic solvent/water media. Electron spectroscopy for chemical analysis (ESCA) of the nanosphere surfaces indicated that PEG macromonomer chains were favorably located on their surfaces. The morphologies of the nanospheres were observed via a scanning electron micrograph (SEM), and particle sizes were estimated by a submicron particle analyzer. When both the concentration of macromonomers and molecular weight were higher, small nanospheres in diameter were obtained. Larger nanospheres in diameter were obtained using macromonomers with low molecular weight at lower concentration. The functions that correlate the diameter (Dn) on different concentration units were Dn = K[St]0.64[MMA-PEG]−0.53±0.01[I]−0.49 and Dn = K[St]0.80[St-PEG]−0.69±0.01[I]−0.22, where [I], [St], [MMA-PEG], and [St-PEG] are initiator, styrene, MMA-PEG, and St-PEG macromonomer concentration in feed, respectively. When the reaction parameters such as the molecular weight of the macromonomers were properly chosen, the particle size could be controlled in a range from ca. 80 to 3100 nm. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2155–2166, 1999  相似文献   

13.
Early transition metal complexes employing a diamido N-heterocyclic carbene (NHC) ligand set (denoted [NCN]) render the centrally disposed NHC moiety stable to dissociation. Aminolysis reactions with the mesityl-substituted ligand precursor (Mes[NCN]H2) and M(NMe2)4 (M = Zr, Hf) provide bis(amido)-NHC-metal complexes that can be further converted to chloro and alkyl derivatives. Activation of Mes[NCN]M(CH3)2 with [Ph3C][B(C6F5)4] yields {Mes[NCN]MCH3}{B(C6F5)4}, which is surprisingly inactive for the polymerization of 1-hexene. The zirconium cation did, however, show moderate ability to catalytically polymerize ethylene. The hafnium dialkyls are thermally stable with the exception of the diethyl complex, Mes[NCN]Hf(CH2CH3)2, which undergoes β-hydrogen transfer and subsequent C–H bond activation with an ortho-methyl substituent on the mesityl group. The hafnium dialkyl complexes also insert carbon monoxide and substituted isocyanides to yield η2-acyls and η2-iminoacyls, respectively. In some circumstances, further C–C bond coupling occurs to yield enediolates and eneamidolate metallocycles. The molecular structures of Mes[NCN]Hf(CH2CHMe2)2, Mes[NCN]Hf(η2-(2,6-Me2C6H3NCCH3)(CH3), Mes[NCN]Hf(η2-(2,6-Me2C6H3NCCH3)2, Mes[NCN]Hf(OC(CH3)C(CH3)NXy), and [Mes[NCN]Hf(OC(iBu)C(iBu)O)]2 are included.  相似文献   

14.
Data are presented to show that two correlations of viscosity–concentration data are useful representations for data over wide ranges of molecular weight and up to at least moderately high concentrations for both good and fair solvents. Low molecular weight polymer solutions (below the critical entanglement molecular weight Mc) generally have higher viscosities than predicted by the correlations. One correlation is ηsp/c[η] versus k′[η], where ηsp is specific viscosity, c is polymer concentration, [η] is intrinsic viscosity, and k′ is the Huggins constant. A standard curve for good solvent systems has been defined up to k′[η]c ≈? 3. It can also be used for fair solvents up to k′[η]c ≈? 1.25· low estimates are obtained at higher values. A simpler and more useful correlation is ηR versus c[η], where ηR is relative viscosity. Fair solvent viscosities can be predicted from the good solvent curve up to c[η] ≈? 3, above which estimates are low. Poor solvent data can also be correlated as ηR versus c[η] for molecular weights below 1 to 2 × 105.  相似文献   

15.
The first solvent‐free cationic complexes of the divalent rare‐earth metals, [{RO}REII]+[A]? (REII=YbII, 1 ; EuII, 2 ) and [{LO}REII]+[A]? ([A]?=[H2N{B(C6F5)3}2]?; REII=YbII, 3 ; EuII, 4 ), have been prepared by using highly chelating monoanionic aminoether‐fluoroalkoxide ({RO}?) and aminoether‐phenolate ({LO}?) ligands. Complexes 1 and 2 are structurally related to their alkaline‐earth analogues [{RO}AE]+[A]? (AE=Ca, 5 ; Sr, 6 ). Yet, the two families behave very differently during catalysis of the ring‐opening polymerization (ROP) of L ‐lactide (L ‐LA) and trimethylene carbonate (TMC) performed under immortal conditions with excess BnOH as an exogenous chain‐transfer agent. The ligand was found to strongly influence the behavior of the REII complexes during ROP catalysis. The fluoroalkoxide REII catalysts 1 and 2 are not oxidized under ROP conditions, and compare very favorably with their Ca and Sr congeners 5 and 6 in terms of activity (turnover frequency (TOF) in the range 200–400 molL‐LA (molEu h?1)) and control over the parameters during the immortal ROP of L ‐LA (Mn,theorMn,SEC, Mw/Mn<1.05). The EuII‐phenolate 4 provided one of the most effective ROP cationic systems known to date for L ‐LA polymerization, exhibiting high activity (TOF up to 1 880 molL‐LA?(molEu h)?1) and good control (Mw/Mn=1.05). By contrast, upon addition of L ‐LA the YbII‐phenolate 3 immediately oxidizes to inactive REIII species. Yet, the cyclic carbonate TMC was rapidly polymerized by combinations of 3 (or even 1 ) and BnOH, revealing excellent activities (TOF=5000–7000 molTMC?(molEu h)?1) and unusually high control (Mn,theorMn,SEC, Mw/Mn<1.09); under identical conditions, the calcium derivative 5 was entirely inert toward TMC. Based on experimental and kinetic data, a new ligand‐assisted activated monomer ROP mechanism is suggested, in which the so‐called ancillary ligand plays a crucial role in the catalytic cycle. A coherent reaction pathway computed by DFT, compatible with the experimental data, supports the proposed scenario.  相似文献   

16.
Potassium hexafluoridotechnetate(IV), K2[TcF6], slowly reacts in aqueous solution with acetohydroxamic acid with formation of the ammine nitrosyltechnetium(I) complex [Tc(NO)(NH3)4F]+. The product crystallizes as mixed TcF62–/HF2 salt of the composition [Tc(NO)(NH3)4F]4[TcF6][HF2]2. [Tc(NO)(NH3)4F]+ represents the first nitrosyltechnetium complex with a fluorido ligand in its coordination sphere. The Tc–F bonds in two crystallographically independent species are 1.987(2) and 2.034(2) Å, respectively. This is slightly longer than in the [TcF6]2– counterion.  相似文献   

17.
Preparation and Crystal Structure of (NH4)2[V(NH3)Cl5]. The Crystal Chemistry of the Compounds (NH4)2[V(NH3)Cl5], [Rh(NH3)5Cl]Cl2, and M2VXCl5 with M = K, NH4, Rb, Cs and X ? Cl, O (NH4)2[V(NH3)Cl5] crystallizes like [Rh(NH3)5Cl]Cl2 in the orthorhombic space group Pnma with Z = 4. The compounds are built up by isolated NH4+ or Cl? and complex MX5Y ions. The following distances have been observed: V? N: 213.8, V? Cl: 235.8–239.1, Rh? N: 207.1–208.5, Rh? Cl: 235.5 pm. Both structures differ from the K2PtCl6 type mainly in the ordering of the MX5Y polyhedra. The compounds M2VCl6 and M2VOCl5 with M = K, NH4, Rb, and Cs crystallize with exception of the orthorhombic K2VOCl5 in the K2PtCl6 type. The ordering of the MX5Y polyhedra in the compounds (NH4)2[V(NH3)Cl5], [Rh(NH3)5Cl]Cl2 and K2VOCl5 enables a closer packing.  相似文献   

18.
This research presents the highly regioselective syntheses of 1,2-dicarboxylated cyclopentadienide salts [Cat]2[C5H3(CO2)2H] by reaction of a variety of organic cation methylcarbonate salts [Cat]OCO2Me (Cat=NR4+, PR4+, Im+) with cyclopentadiene (CpH) or by simply reacting organic cation cyclopentadienides Cat[Cp] (Cat=NR4+, PR4+, Im+) with CO2. One characteristic feature of these dianionic ligands is the acidic proton delocalized in an intramolecular hydrogen bridge (IHB) between the two carboxyl groups, as studied by 1H NMR spectroscopy and XRD analyses. The reaction cannot be stopped after the first carboxylation. Therefore, we propose a Kolbe-Schmitt phenol-carboxylation related mechanism where the acidic proton of the monocarboxylic acid intermediate plays an ortho-directing and CO2 activating role for the second kinetically accelerated CO2 addition step exclusively in ortho position. The same and related thiocarboxylates [Cat]2[C5H3(COS)2H] are obtained by reaction of COS with Cat[Cp] (Cat=NR4+, PR4+, Im+). A preliminary study on [Cat]2[C5H3(CO2)2H] reveals, that its soft and hard coordination sites can selectively be addressed by soft Lewis acids (Mo0, Ru2+) and hard Lewis acids (Al3+, La3+).  相似文献   

19.
Tetraethylphosphonium azide, [P(C2H5)4 ]+[N3 ], was prepared from tetraethyl phosphonium bromide and silver azide. Single crystals of [P(C2H5)4 ]+[N3 ] were grown from dichloromethane/THF (10:1) solution. The structure was determined by single-crystal X-ray diffraction analysis. [P(C2H5)4 ]+[N3 ] crystallizes in the monoclinic space group C 2/c with Z = 4 and unit cell dimensions a = 12.961(6), b = 6.835(3), c = 12.378(6) Å, and β = 100.57(4)°. The attempted preparation of phosphonium azide [PH4]+[N3] from phosphonium iodide and silver azide lead instead to the formation of PH3 and HN3. The instability of [PH4]+[N3] with respect to PH3 and HN3 is in accord with thermodynamic considerations according to which the reaction PH3(g) and HN3(g) to yield [PH4 ]+[N3 ] is thermodynamically unfavorable. (Non SI units employed: kcal ≈ 4.184 J, Å = 10−10 m.) © 1998 John Wiley & Sons, Inc. Heteroatom Chem 9:129–132, 1998  相似文献   

20.
Various novel double metal cyanide (DMC) catalysts were successfully prepared by modifying the central metal (M) and one of cyanide ion (CN-) in Zna[M(CN)b]c complex. Such modifications have significant impact on the catalytic efficiency as well as the polymer selectivity for the reaction of PO/CO2. Zn–Ni(Ⅱ) DMC is a potential catalyst for alternating copolymerization of PO/CO2, and DMC catalysts based on Zn3[Co(CN)5X]2 (X = Br-and N3-) exhibit moderate efficiency for the production of polycarbonates. This research presents the preliminary exploration of novel DMC complex via chemical modification of its central metal and ligand.  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