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 photochemical reaction of piperazine with C70 produces a mono‐adduct (N(CH2CH2)2NC70) in high yield (67 %) along with three bis‐adducts. These piperazine adducts can combine with various Lewis acids to form crystalline supramolecular aggregates suitable for X‐ray diffraction. The structure of the mono‐adduct was determined from examination of the adduct I2N(CH2CH2)2NI2C70 that was formed by reaction of N(CH2CH2)2NC70 with I2. Crystals of polymeric {Rh2(O2CCF3)4N(CH2CH2)2NC70}n?nC6H6 that formed from reaction of the mono‐adduct with Rh2(O2CCF3)4 contain a sinusoidal strand of alternating molecules of N(CH2CH2)2NC70 and Rh2(O2CCF3)4 connected through Rh?N bonds. Silver nitrate reacts with N(CH2CH2)2NC70 to form black crystals of {(Ag(NO3))4(N(CH2CH2)2NC70)4}n?7nCH2Cl2 that contain parallel, nearly linear chains of alternating (N(CH2CH2)2NC70 molecules and silver ions. Four of these {Ag(NO3)N(CH2CH2)2NC70}n chains adopt a structure that resembles a columnar micelle with the ionic silver nitrate portion in the center and the nearly non‐polar C70 cages encircling that core. Of the three bis‐adducts, one was definitively identified through crystallization in the presence of I2 as 12{N(CH2CH2)2N}2C70 with addends on opposite poles of the C70 cage and a structure with C2v symmetry. In 12{I2N(CH2CH2)2N}2C70, individual 12{I2N(CH2CH2)2N}2C70 units are further connected by secondary I2???N2 interactions to form chains that occur in layers within the crystal. Halogen bond formation between a Lewis base such as a tertiary amine and I2 is suggested as a method to produce ordered crystals with complex supramolecular structures from substances that are otherwise difficult to crystallize.  相似文献   

2.
Trimethylsilyldiethylamine Me3SiNEt2 and MoOCl4 (1:1) undergo a free radical redox reaction in CH2Cl2 or Et2O to form MoCl3O(HNEt2). Reduction occurs even in aprotic media like CCl4 and CS2 to give MoV complexes Mo2Cl6O2(N2Et4) and Mo2Cl6O2[(SCNEt2)2S2], respectively. A 2:1 reaction in nonionizing protic solvents undergoes redox cum cleavage to provide MoCl2O(NEt2) (HNEt2) but a reaction at reflux temperature in 1,2-dichloroethane leads to diethylammonium salt, [Et2NH2][MoCl4O(HNEt2)]. Higher molar reactions (3:1, 4:1) in CH2Cl2 or Et2O are associated with redox reaction as well as oxygen atom abstraction to form de-oxo MoIV complex MoCl3(NEt2)(HNEt2)2, whereas, a 3:1 reaction in CS2 forms Mo2Cl4O(S2CNEt2)4. Compounds have been characterized by elemental analyses, redox titration, magnetic moment, conductance, infrared, electronic absorption and 1H-NMR measurements.  相似文献   

3.
Complexes of formula [(H2N2O2)TiCl2] and [(H2N2O2)Ti(OiPr)2] (H2N2O2H2 = HOPh’CH2NH(CH2)2NHCH2Ph’OH, where Ph’ = 2,4-(CMe2Ph)C6H2) were synthesized by the reaction of the salan ligand precursor H2N2O2H2 with TiCl4 and Ti(OiPr)4, respectively, in high yields. The dichlorido complex [(H2N2O2)TiCl2] revealed to be an efficient catalyst for the reduction of benzaldehyde in toluene. Full conversion was observed after 24 h at 55 °C in THF. The same catalyst also converted phenylacetaldehyde and hydrocinnamaldehyde into the corresponding alkanes quantitatively.  相似文献   

4.
The mass spectra of the following acetylenic derivatives of iron, ruthenium and osmium carbonyls are reported: the iron compounds Fe2(CO)6[C2(C6H5)s2]2, Fe2(CO)6[C2(CH3)2]2 and Fe2(CO)6[C2(C2H5)2]2, the ruthenium compounds Ru2(CO)6[C2(C6H5)2]2, and Ru2(CO)6[C2(CH3)2]2 and the osmium compounds Os2(CO)6[C2(C6H5)2]2, Os2(CO)6[C2HC6H5]2 and Os2(CO)6[C2(CH3)2]2. Iron compounds exhibit breakdown schemes where binuclear, mononuclear and hydrocarbon ions are present. On the other hand, ruthenium and osmium compounds fragment in a similar way and give rise to singly and doubly charged binuclear ions. Phenylic derivatives of ruthenium and osmium also give weak triply charged ions. The results are discussed in terms of relative strengths of the metal-metal and metal-carbon bonds.  相似文献   

5.
The kinetic regularities of the heat release during the thermal decomposition of liquid NH4N(NO2)2 at 102.4–138.9 °C were studied. Kinetic data for decomposition of different forms of dinitramide and the influence of water on the rate of decomposition of NH4N(NO2)2 show that the contributions of the decomposition of N(NO2)2 and HN(NO2)2 to the initial decomposition rate of the reaction at temperatures about 100 °C are approximately equal. The decomposition has an autocatalytic character. The analysis of the effect of additives of HNO3 solutions and the dependence of the autocatalytic reaction rate constant on the gas volume in the system shows that the self-acceleration is due to an increase in the acidity of the NH4N(NO2)2 melt owing to the accumulation of HNO3 and the corresponding increase in the contribution of the HN(NO2)2 decomposition to the overall rate. The self-acceleration ceases due to the accumulation of NO3 ions decreasing the equilibrium concentration of HN(NO2)2 in the melt. For Part 2, see Ref. 1. Translated fromIzvestiya Akademii Nauk, Seriya Khimicheskaya, No. 3, pp. 395–401 March 1998.  相似文献   

6.
The structure of precursors is used to control the formation of six possible structural isomers that contain four structural units of PbSe and four structural units of NbSe2: [(PbSe)1.14]4[NbSe2]4, [(PbSe)1.14]3[NbSe2]3[(PbSe)1.14]1[NbSe2]1, [(PbSe)1.14]3[NbSe2]2[(PbSe)1.14]1[NbSe2]2, [(PbSe)1.14]2[NbSe2]3[(PbSe)1.14]2[NbSe2]1, [(PbSe)1.14]2[NbSe2]2[(PbSe)1.14]1[NbSe2]1[(PbSe)1.14]1[NbSe2]1, [(PbSe)1.14]2[NbSe2]1[(PbSe)1.14]1[NbSe2]2[(PbSe)1.14]1[NbSe2]1. The electrical properties of these compounds vary with the nanoarchitecture. For each pair of constituents, over 20 000 new compounds, each with a specific nanoarchitecture, are possible with the number of structural units equal to 10 or less. This provides opportunities to systematically correlate structure with properties and hence optimize performance.  相似文献   

7.
Solid alkali metal carbonates are universal passivation layer components of intercalation battery materials and common side products in metal‐O2 batteries, and are believed to form and decompose reversibly in metal‐O2/CO2 cells. In these cathodes, Li2CO3 decomposes to CO2 when exposed to potentials above 3.8 V vs. Li/Li+. However, O2 evolution, as would be expected according to the decomposition reaction 2 Li2CO3→4 Li++4 e?+2 CO2+O2, is not detected. O atoms are thus unaccounted for, which was previously ascribed to unidentified parasitic reactions. Here, we show that highly reactive singlet oxygen (1O2) forms upon oxidizing Li2CO3 in an aprotic electrolyte and therefore does not evolve as O2. These results have substantial implications for the long‐term cyclability of batteries: they underpin the importance of avoiding 1O2 in metal‐O2 batteries, question the possibility of a reversible metal‐O2/CO2 battery based on a carbonate discharge product, and help explain the interfacial reactivity of transition‐metal cathodes with residual Li2CO3.  相似文献   

8.
In the presence of CoX2(PPh3)2/3 PPh3 and zinc metal conjugated alkenes (CH2CHCOOR, CH2CHCN, CH2CHSO2Ph and CH2CHCONEt2) undergo reductive tail-to-tail dimerization to yield the corresponding saturated linear products. Under similar reaction conditions, vinylarenes (ArCHCH2) give stereoselective head-to-tail dimerization products, trans-1,3-diarylbut-1-ene, in good to excellent yields.  相似文献   

9.
《Polyhedron》1986,5(9):1467-1473
Direct- and alternating-current polarograms of aqueous SO2 · OH2 solutions show four reduction waves, more than previously reported. Waves I and II probably result from the electroreduction of SO2 · OH2 and HSO3, respectively; these two waves completely overlap at pH 1 but are partially resolved at higher pH values due to different pH dependence. Reduction of SO2 · OH2 involves two electrons and two H+ ions and the initial product is probably sulfoxylic acid, H2SO2. This product can disproportionate to S0 and SO2 · OH2 in very acidic media (pH ≤ 1) and, in the limit, double the reduction current of SO2 · OH2. Reduction of HSO3 appears to occur via two paths: one is a two-electron three-H+ ion path and the other is a one-electron one-H+ ion path. The former dominates at pH ≤ 3 and probably produces H2SO2; the latter dominates at pH > 4 and may produce SO2. H2SO2 in less acidic media can react with HSO3 to yield dithionite species (such as H2S2O4, HS2O4 and S2O2−4) and HSO2 and SO2 by dissociation of the dithionite species. Waves III and IV are believed to result from reduction of HSO2 and SO2, respectively, to H2SO2 species.  相似文献   

10.
The PH bond of dialkylphosphites (dimethylphosphite, 5,5-dimethyl-1,3-dioxa-2-phosphorinane and 4,4,5,5-tetramethyl-1,3-dioxa-2-phospholane) oxidatively adds to irClL2(L = PPh3, AsPh3) and IrCl(PMe2Ph)3 generated in situ to give six-coordinate hydrido(dialkylphosphonato)iridium(III) complexes, e.g. IrHClL2[{(MeO)2-PO}2H] and IrHCl(PMe2Ph)3[PO(OMe)2]. Addition of triphenylphosphine to a solution containing [IrCl(C8H14)2]2 and dimethylphosphite in a 1:2 mol ratio gives a five-coordinate hydrido (dimethylphosphonato)iridium(III) complex IrHCl(PPh3)2{PO(OMe)2}, from which six-coordinate pyridine and acetylacetonato complexes IrHCl(PPh3)2(C5H5N){PO(OMe)2} and IrH(PPh3)2(acac){PO(OMe)2} can be obtained. The ligand arrangements in the various complexes are inferred from IR, 1H and 31P NMR data.  相似文献   

11.
An attempt has been made to design double‐stranded ladder‐like coordination polymers (CPs) of hemidirected PbII. Four CPs, [Pb(μ‐bpe)(O2C‐C6H5)2] ? 2H2O ( 1 ), [Pb2(μ‐bpe)2(μ‐O2C‐C6H5)2(O2C‐C6H5)2] ( 2 ), [Pb2(μ‐bpe)2(μ‐O2C‐p‐Tol)2(O2C‐p‐Tol)2] ? 1.5 H2O ( 3 ) and [Pb2(μ‐bpe)2(μ‐O2C‐m‐Tol)2(O2C‐m‐Tol)2] ( 4 ) (bpe=1,2‐bis(4′‐pyridyl)ethylene), have been synthesised and investigated for their solid‐state photoreactivity. CPs 2 – 4 , having a parallel orientation of bpe molecules in their ladder structures and being bridged by carboxylates, were found to be photoreactive, whereas CP 1 is a linear one‐dimensional (1D) CP with guest water molecules aggregating to form a hydrogen‐bonded 1D structure. The linear strands of 1 were found to pair up upon eliminating lattice water molecules by heating, which led to the solid‐state structural transformation of photostable linear 1D CP 1 into photoreactive ladder CP 2 . In the construction of the double‐stranded ladder‐like structures, the parallel alignment of C?C bonds in 2 – 4 is dictated by the chelating and μ2‐η21 bridging modes of the benzoate and toluate ligands. The role of solvents in the formation of such double‐stranded ladder‐like structures has also been investigated. A single‐crystal‐to‐single‐crystal transformation occurred when 4 was irradiated under UV light to form [Pb2(rctt‐tpcb)(μ‐O2C‐m‐Tol)2(O2C‐m‐Tol)2] ( 5 ).  相似文献   

12.
Geometry of the CO2–H2O complex and reaction barriers leading to the formation of H2CO3were studied at the RHF/6-311++G**, MP2/6-311++G**, B3LYP/AUG-cc-pVDZ, B3LYP/AUG-cc-pVTZ, MP2/AUG-cc-pVDZ and CCD/AUG-cc-pVDZ levels of theory. The rotational barrier of the CO2–H2O complex and the reaction barrier leading to the formation of H2CO3–H2O from CO2–(H2O)2 were studied using the first three of the above-mentioned methods. Microsolvation of CO2 in water clusters having upto eight water molecules was studied using the B3LYP/AUG-cc-pVDZ method. Various methods except MP2/AUG-cc-pVDZ predict the equilibrium structure of the CO2–H2O complex to be symmetric while the MP2/AUG-cc-pVDZ method predicts it to be unsymmetric. Formation of H2CO3 from CO2–H2O is strongly catalyzed by the presence of a second water molecule. Atomic orbitals are strongly rehybridized in going from the equilibrium structures of the CO2–H2O and CO2–(H2O)2 complexes to the transition states involved in the formation of H2CO3 and H2CO3–H2O, respectively, as shown by hybridization displacement charges.  相似文献   

13.
The solubilities of the systems CeO2-SeO2-H2O and Ce2O3-SeO2-H2O were studied at 100°C. The field of crystallization of Ce(SeO3)2 was established in the system CeO2-SeO2-H2O, and fields of crystallization of Ce2(SeO3)3 and Ce2(SeO3)3H2SeO3 were established in the system Ce2O3-SeO2-H2O. The compound obtained were identified by means of chemical, X-ray and derivatograph analysis. The mechanism of thermal dissociation of Ce(SeO3)2, Ce2(SeO3)3 and Ce2(SeO3)3·H2SeO3 was studied. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

14.
A new efficient synthesis of functionalized perfluoroalkyl fluorophosphates by oxidative addition of Me2NCH2F to the electron‐deficient phosphanes (C2F5)nPF3?n (n=0–3) is reported. The initially formed zwitterionic, hexacoordinated phosphates [(C2F5)nF5?nP(CH2NMe2?CH2NMe2)] are converted into the corresponding phosphonium salts [(Me3PCH2NMe2]+[(C2F5)nF5?nP(CH2NMe2)]? by treatment with PMe3. In addition [(C2F5)3F2P(CH2NMe2?CH2NMe2)] can undergo a 1,3‐methyl shift from the internal to the terminal nitrogen—a structural characterization was achieved from the CF3 analogue. Reaction of [(C2F5)3F2P(CH2NMe?CH2NMe3)] and PMe3 gave rise to the formation of the zwitterionic phosphonium phosphate [(C2F5)3F2P(CH2NMe?CH2PMe3)], which was fully characterized by X‐ray diffraction analysis. Moreover, an efficient one‐pot synthesis of Cs+[(C2F5)3F2P(CH2NMe2)]? was pursued. This salt turned out to be a useful nucleophile in several alkylation reactions.  相似文献   

15.
The reaction of o-C6H4(AsMe2)2 with VCl4 in anhydrous CCl4 produces orange eight-coordinate [VCl4{o-C6H4(AsMe2)2}2], whilst in CH2Cl2 the product is the brown, six-coordinate [VCl4{o-C6H4(AsMe2)2}]. In dilute CH2Cl2 solution slow decomposition occurs to form the VIII complex [V2Cl6{o-C6H4(AsMe2)2}2]. Six-coordination is also found in [VCl4{MeC(CH2AsMe2)3}] and [VCl4{Et3As)2]. Hydrolysis of these complexes occurs readily to form vanadyl (VO2+) species, pure samples of which are obtained by reaction of [VOCl2(thf)2(H2O)] with the arsines to form green [VOCl2{o-C6H4(AsMe2)2}], [VOCl2{MeC(CH2AsMe2)3}(H2O)] and [VOCl2(Et3As)2]. Green [VOCl2(o-C6H4(PMe2)2}] is formed from [VOCl2(thf)2(H2O)] and the ligand. The [VOCl2{o-C6H4(PMe2)2}] decomposes in thf solution open to air to form the diphosphine dioxide complex [VO{o-C6H4(P(O)Me2)2}2(H2O)]Cl2, but in contrast, the products formed from similar treatment of [VCl4{o-C6H4(AsMe2)2}x] or [VOCl2{o-C6H4(AsMe2)2}] contain the novel arsenic(V) cation [o-C6H4(AsMe2Cl)(μ-O)(AsMe2)]+. X-ray crystal structures are reported for [V2Cl6{o-C6H4(AsMe2)2}2], [VO(H2O){o-C6H4(P(O)Me2)2}2]Cl2, [o-C6H4(AsMe2Cl)(μ-O)(AsMe2)]Cl·[VO(H2O)3Cl2] and powder neutron diffraction data for [VCl4{o-C6H4(AsMe2)2}2].  相似文献   

16.
Pure sym‐N2O4 isolated in solid Ne was obtained by passing cold neon gas over solid N2O4 at ?115 °C and quenching the resulting gaseous mixture at 6.3 K. Filtered UV irradiation (260–400 nm) converts sym‐N2O4 into trans‐ONONO2, a weakly interacting (NO2)2 radical pair, and traces of the cis‐N2O2?O2 complex. Besides the weakly bound ON?O2 complex, cis‐N2O2?O2 was also obtained by co‐deposition of NO and O2 in solid Ne at 6.3 K, and both complexes were characterised by their matrix IR spectra. Concomitantly formed cis‐N2O2 dissociated on exposure to filtered IR irradiation (400–8000 cm?1), and the cis‐N2O2?O2 complex rearranged to sym‐N2O4 and trans‐ONONO2. Experiments using 18O2 in place of 16O2 revealed a non‐concerted conversion of cis‐N2O2?O2 into these species, and gave access to four selectively di‐18O‐substituted trans‐ONONO2 isotopomers. No isotopic scrambling occurred. The IR spectra of sym‐N2O4 and of trans‐ONONO2 in solid Ne were recorded. IR fundamentals of trans‐ONONO2 were assigned based on experimental 16/18O isotopic shifts and guided by DFT calculations. Previously reported contradictory measurements on cis‐ and trans‐ONONO2 are discussed. Dinitroso peroxide, ONOONO, a proposed intermediate in the IR photoinduced rearrangement of cis‐N2O2?O2 to the various N2O4 species, was not detected. Its absence in the photolysis products indicates a low barrier (≤10 kJ mol?1) for its exothermic O? O bond homolysis into a (NO2)2 radical pair.  相似文献   

17.
DTA and TG studies in air were carried out for hydrothermally prepared rhombohedral double carbonates of dolomite type, CaMg(CO3)2, CaMn(CO3)2, CdMg(CO3)2, CdMn(CO3)2 and CdZn(CO3)2. The solid decomposition products in air have been compared to those obtained under hydrothermal conditions with CO2 pressure. The dolomite [CaMg(CO3)2] decomposes in two stages both in air as well as under high CO2 pressure. The other carbonates studied, follow a single stage decomposition in air and a two stage decomposition under hydrothermal condition. In air, the manganese containing carbonates CaMn(CO3)2 and CdMn(CO3)2, decompose to form mixed oxides of CaMnO3 and CdMnO3 respectively, while CdMg(CO3)2 and CdZn(CO3)2 decompose to their respective two mono oxides.
Zusammenfassung Mittels DTA und TG in Luft werden die hydrothermisch hergestellten rhomboedrischen Doppelkarbonate (Dolomittyp) CaMg(CO3)2, CaMn(CO3)2, CdMg(CO3)2, CdMn(CO3)2 und CdZn(CO3)2 untersucht. Die in Luft erhaltenen festen Zersetzungsprodukte wurden mit denen verglichen, die unter hydrothermischen Bedingungen mit CO2-Druck entstehen. Dolomit zersetzt sich sowohl in Luft als auch unter hohem CO2-Druck in zwei Schritten. Die übrigen untersuchten Karbonate zersetzen sich in Luft in einem, unter hydrothermischen Bedingungen in zwei Schritten. In Luft zersetzen sich die magnesiumhaltigen Karbonate CaMn(CO3)2 und CdMn(CO3)2 unter Bildung der Mischoxide CaMnO3 und CdMnO3, während aus CdMg(CO3)2 und CdZn(CO3)2 jeweils die entsprechenden beiden Monoxide entstehen.
  相似文献   

18.
In dissociation experiments of H2O2 under shock wave conditions, the spectra of H2O2 and HO2 have been observed in the UV at 2200 ≤ 2800 Å. By the use of these spectra the H2O2 decomposition in the presence of H2 and CO at 870 ≤ T ≤ 1000°K has been analyzed. It was found that in this temperature range, in contrast to low temperature behavior, reactions of H atoms with H2O2 and with HO2 are equally important. The rate of the reaction H + H2O2 ← HO2 + H2 was estimated in comparison with the rate of the reaction between H and HO2. Good agreement between calculated and measured concentration profiles of HO2 and H2O2 was obtained.  相似文献   

19.
The ligand, 1,2-bis(difluorophosphino)ethane, (PF2C2H4PF2), reacts with Ni(CO)4 in the gas phase and in solution to produce carbon monoxide and a polymer, [Ni(PF2C2H4PF2)2]x. PF2C2H4PF2 displaces norbornadiene from (C7H8)Mo(CO)4 to yield the relatively air-stable complex, Mo(CO)4-(PF2C2H4PF2). Analysis of the infrared spectrum of the monomeric complex indicates that the ligand exhibits π-acceptor strength equal to PF2C6H10PF2.  相似文献   

20.
The reaction of (p‐MeOC6H4)2TeO with two equivalents of HO3SCF3 and HO2PPh2 provided the tetraorganoditelluroxanes (F3CSO3)(p‐MeOC6H4)2TeOTe(p‐MeOC6H4)2(O3SCF3) ( 1 ) and (Ph2PO2)(p‐MeOC6H4)2TeOTe(p‐MeOC6H4)2(O2PPh2)·2 Ph2PO2H ( 2 ) in good yields. Compounds 1 and 2 were characterized by solution and solid‐state 31P and 125Te NMR spectroscopy, IR spectroscopy, electrospray mass spectrometry, conductivity measurements and single crystal X‐ray diffraction. In solution, compound 1 undergoes an electrolytic dissociation and reversibly reacts with traces of water to give the mononuclear cation [(p‐MeOC6H4)2TeOH]+ and triflate anions. Theoretical aspects of the protonation and hydration of model telluroxanes R2TeO (R = H, Me, Ph) were investigated by preliminary DFT calculations and compared to the corresponding selenoxanes R2SeO. The tellurium dihydroxides R2Te(OH)2 seem to be more stable than the hydrogen‐bonded complexes R2TeO·H2O.  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