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 46 毫秒
1.
A method of calculation of average heat capacities of phase transformation products of complex oxides is suggested. The method takes into account the physical state of products and the increase in the heat capacities of products due to the change of entropy at a phase transformation. Average heat capacities of products formed in a congruous melting of compounds (YCuO2 and Y4Ba3O9), in an incongruous melting of compounds (Y2Cu2O5, BaCuO2, BaCu2O2, Y2BaCuO5, YBa2Cu3O7, YBa2Cu3O6) and in a decomposition in a crystalline state of compounds (Y2BaO4, Y2Ba2O5, Y2Ba4O7, Ba2CuO3, Ba3Cu5O8, YBa2Cu3.5O7.5, YBa2Cu4O8, YBa2Cu5O9) was estimated by using three methods.  相似文献   

2.
The interaction between a long chain alkane, tetradecane (abbreviated H14), molecule and a semi-fluorinated alkane, 1-perfluorododecyl-hexadecane F(CF2)12(CH2)16H (abbreviated F12H16), molecule at the air/ H14 solution interface was studied by measuring the surface tension of the H14 solutions of F12H16 as a function of temperature and bulk concentration under atmospheric pressure. Pure liquid H14 freezes without forming a condensed film at its surface. Nevertheless, a very small amount of F12H16 initiates the surface freezing of H14. In contrast to the F12H16-hexadecane (abbreviated H16) system, the condensed monolayer of H14 has a finite solubility of F12H16 in the F12H16-H14 system. By further increasing the bulk concentration of F12H16, the F12 chains of the F12H16 molecules form the other closely packed condensed state. Hence, as in the case of the H16 system, the H14 system also exhibits a surface hetero-azeotrope behavior in the lower temperature region. Below the surface hetero-azeotropic point, the condensed H14 monolayer containing a small amount of F12H16 is completely replaced by the condensed monolayer of F12H16. At 2 °C, for example, a surface of H14 solution of F12H16 covered with a gaseous film of F12H16 is replaced by a condensed H14 monolayer containing an almost gaseous state of F12H16, and is then completely replaced by the condensed monolayer of F12H16 with increasing bulk concentration. Above the temperature of the triple point for the F12H16 monolayer, the F12H16-H14 system exhibits a gaseous, expanded, and condensed state.  相似文献   

3.
The complexes (η5-C5H5)Pd(η1-C5H5)PR3 which are prepared from [Cl(PR3)-Pd]2(μ-OCOCH3)2 and TlC5H5 are fluxional in solution. According to the 1H and 13C NMR spectra at various temperatures, two dynamic processes occur. The process with the higher activation energy is a π/σ (η51) exchange of the two different cyclopentadienyl ligands, whereas the second one with the lower activation energy presumably is a metallotropic rearrangement (1,2-shift). The coalescence temperature for the η51 exchange depends on the size of the phosphine. The X-ray structural analysis of (C5H5)2PdPPri3 proves that it exists as a “frozen” η5 + η1 structure in the crystal with the palladium approximately in a square-planar coordination. The η5-bonded cyclopentadienyl ring shows some unusual bonding patterns which are obviously electronic in nature. EHT-MO calculations for (η5-C5H5)PdCH3(PH3) indicate that in this model system alternating CC distances in the ring and a stronger bond of the metal to one of the five carbon atoms of the C5H5 ligand are to be expected. The calculations suggest that in similar complexes possessing a six-electron donor ligand like C5H5? and a metal fragment which is isolobal to PdCH3(PH3)+, analogous distortions should be observed. Some reactions of the compounds (η5-C5H5)Pd(η1-C5H5)PR3 are described.  相似文献   

4.
The interlanthanide compounds Er3SmS6, Er3SmSe6, and Er1.12Sm0.88Se3 have been synthesized from stoichiometric reactions of the elements in a KI salt flux at 1273, 1173, and 1123 K, respectively. Er3SmS6 and Er3SmSe6, which are isostructural and ordered, crystallize in space group P21/m in the ScEr3S6 structure type whereas Er1.12Sm0.88Se3, in which the Er and Sm atoms are disordered, crystallizes in space group Pnma in the U2S3 structure type. Er3SmS6 is a paramagnet with a μeff=11.25(1) μB/mol. From optical measurements a direct band gap of 2.0 eV for light perpendicular to the (100) crystal face of Er3SmSe6 is derived whereas for isostructural Er3SmS6 an optical transition at 2.2-2.4 eV and a broad absorption peak at lower energies are observed.  相似文献   

5.
The reactivity of bis(dimethylamido) complexes of phenyl- and hydridogallium with ammonia, dimethylamine and 1,1-dimethylhydrazine is described. Synthesis of the starting gallium hydride, [HGa(NMe2)2]2, was achieved in nearly quantitative yield from the reaction of HGaCl2(quinuclidine) with LiNMe2. In neat ammonia or methylamine at room temperature both dimethylamido ligands in [HGa(NMe2)2]2 were substituted by a single equivalent of NH3 or MeNH2 to produce amorphous (HGaNH)n or (HGaNMe)n, respectively. In contrast, the reaction of [PhGa(NMe2)2]2 with neat Me2NNH2, at room temperature consumed two equivalents of the substituted hydrazine to form [PhGa(NHNMe2)2]2 in a 73% yield. Single crystal X-ray crystallographic analyses of [HGa(NMe2)2]2 and [PhGa(NHNMe2)2]2 establish that in the solid state both compounds adopt a cyclic Ga-N-Ga-N structure with a crystallographic center of symmetry located at the center of the ring.  相似文献   

6.
This paper reports on a mass spectrometric study of the neutral and ionic species in a low-pressure rf discharge sustained in a C2H4-SiH4 mixture diluted in helium. It is shown that C2H4 is readily decomposed into C2H 2 * and C2H3. The formation of secondary products such as C4H2, C4H4, and C4H6 is observed and confirms the presence of C2H2 in the discharge. Methylsilane (CH3SiH3) and ethylsilane (C2H5SiH3) are also synthesized in this discharge. It is also observed that the major ions C2H 4 + , C3H 5 + , SiH 3 + , Si2H 4 + , SiCH 3 + , SiC2H 3 + , and SiC2H 7 + are not representative of the direct ionization of neutral species. Their formation is thus interpreted on the basis of ion-molecule reactions.  相似文献   

7.
Results are presented of studying electrochemical properties of perovskite-like solid solutions (La0.5 + x Sr0.5 ? x )1 ? y Mn0.5Ti0.5O3 ? δ (x = 0–0.25, y = 0–0.03) synthesized using the citrate technique and studied as oxide anodic materials for solid oxide fuel cells (SOFC). X-ray diffraction (XRD) analysis is used to establish that the materials are stable in a wide range of oxygen chemical potential, stable in the presence of 5 ppm H2S in the range of intermediate temperatures, and also chemically compatible with the solid electrolyte of La0.8Sr0.2Ga0.8Mg0.15Co0.05O3 ? δ (LSGMC). It is shown that transition to a reducing atmosphere results in a decrease in electron conductivity that produced a significant effect on the electrochemical activity of porous electrodes. Model cells of planar SOFC on a supporting solid-electrolyte membrane (LSGMC) with anodes based on (La0.6Sr0.4)0.97Mn0.5Ti0.5O3 ? δ and (La0.75Sr0.25)0.97Mn0.5Ti0.5O3 ? δ and a cathode of Sm0.5Sr0.5CoO3 ? δ are manufactured and tested using the voltammetry technique.  相似文献   

8.
The reaction of {(HNEt3)2[Ag10(tBuC6H4S)12]}n, Ag2O, Na2MoO4, and m‐methoxybenzoic acid (Hmbc) in CH3OH/CH2Cl2 led to yellow crystals of [Ag4S4 (MoO4)5@Ag66] (SD/Ag70b; SD=SunDi) only, while in the presence of DMF, additional dark‐red crystals of [Ag10@ (MoO4)7@Ag60] (SD/Ag70a) were obtained. SD/Ag70b consists of five MoO42? ions wrapped by a shell of 66 Ag atoms, while SD/Ag70a contains a rare Ag10 kernel consisting of five tetrahedra sharing faces and edges, surrounded by seven MoO42? ions enclosed in a shell of 60 Ag atoms. The formation of the Ag10 kernel originates from a reduction reaction during the self‐assembly process that involves DMF. This work provides the structural information of a unique Ag10 kernel (five fused Ag4 tetrahedra) and paves an avenue to trap elusive silver species with hierarchical multi‐shell silver nanocluster assemblies with the help of anion templates.  相似文献   

9.
Dithiolylium Chlorooxomolybdates(V): Synthesis and Crystal Structure of (C3Cl3S2)[MoOCl4] and (C3Cl3S2)[Mo2O2Cl7] The reaction of 3, 4, 5‐Trichlor‐1, 2‐dithiolylium chloride with MoOCl3 in dichlormethane under solvothermal conditions at 65 °C simultaneously yields the green tetrachlorooxomolybdate(V) (C3Cl3S2)[MoOCl4] and the yellow‐brown heptachlorodioxodimolybdate(V) (C3Cl3S2)[Mo2O2Cl7]. The crystal structures of both compounds contain nearly planar (C3Cl3S2)+ ions with a S—S bond length of 203 pm. The discrete [MoOCl4] ion in the structure of (C3Cl3S2)[MoOCl4] has the shape of a square pyramid with the oxygen atom at the apex. The molybdenum atom is displaced by 58 pm from the basal plane towards the oxygen atom. The [Mo2O2Cl7] ion in the structure of (C3Cl3S2)[Mo2O2Cl7] has the form of a face‐sharing double octahedron. It is formally composed of a [MoOCl4] ion and a MoOCl3 molecule connected by one symmetrical and two unsymmetrical chloro bridges. The molybdenum atoms placed in the centers of such connected octahedra are 357 pm apart, indicating no Mo—Mo bond.  相似文献   

10.
The accelerated formation of 2,3-diphenylquinoxalines in microdroplets generated in a nebulizer has been investigated by competition experiments in which equimolar quantities of 1,2-phenylenediamine, C6H4(NH2)2, and a 4-substituted homologue, XC6H3(NH2)2 [X = F, Cl, Br, CH3, CH3O, CO2CH3, CF3, CN or NO2], or a 4,5-disubstituted homologue, X2C6H2(NH2)2 [X = F, Cl, Br, or CH3], compete to condense with benzil, (C6H5CO)2. Electron-donating substituents (X = CH3 and CH3O) accelerate the reaction; in contrast, electron-attracting substituents (X = F, Cl, Br and particularly CO2CH3, CN, CF3 and NO2) retard it. A structure–reactivity relationship in the form of a Hammett correlation has been found by analyzing the ratio of 2,3-diphenylquinoxaline and the corresponding substituted-2,3-diphenylquinoxaline, giving a ρ value of −0.96, thus confirming that the electron density in the aromatic ring of the phenylenediamine component is reduced in the rate-limiting step in this accelerated condensation. This correlation shows that the phenylenediamine acts as a nucleophile in the reaction.  相似文献   

11.
Reactions of cuprous chloride with the phosphorus–nitrogen cage ligand 2,4,6,8,9,10-hexamethyl-2,4,6,8,9,10-hexaaza-1,3,5,7-tetraphosphaadamantane, P4(NCH3)6, in acetonitrile form distinct solids depending on the ligand-to-metal ratio. Three structurally characterized compounds include: a solvated 3-D network of formula [P4(NCH3)6]2(CuCl)3(CH3CN)2; a “ladder-type” polymer {[μ2-P4(NCH3)6]2(CuCl)2}; and a monomeric complex [P4(NCH3)6]2CuCl. Thermal decomposition of the solvated network results in formation of two more materials [P4(NCH3)6]2(CuCl)3, and [P4(NCH3)6](CuCl)2 that are not isolated from solution reactions. The variety of products isolated based solely on ligand-to-metal ratio suggest that this system participates in solution equilibria common to many phosphorus(III) ligands and multiple solubility equilibria.  相似文献   

12.
Reaction of (CH3NPF3)2 with equimolar amounts of N-methylhexamethyldisilazane yields a reaction product, which can be separated in a polymer and a crystalline fraction. High-vacuum sublimation of the crystalline part yields the already known compound (CH3N)4P3F7, the new spiro-isomer F3P(CH3N)2PF(CH3N)2PF3 and the spiro-compound F3P(CH3N)2PF(CH3N)2PF(CH3N)2PF3, an isomer of the known compound (CH3N)6P4F8.  相似文献   

13.
The reaction of a dichloromethane solution of a mixture of cis,trans-[PtCl2(SMe2)2] with a tetrahydrofuran solution of SnBr2 resulted in oxidation of platinum(II) with halogen exchange producing cis,trans-[PtBr4(SMe2)2]. Reaction of a mixture of cis,trans-[PtCl2(SEt2)2], potassium tetrachloroplatinate(II) or potassium hexachloroplatinate(IV) with SnBr2 in hydrochloric acid solution resulted in formation of predominantly anionic five-coordinate trichlorostannyl platinum(II) complexes. Reaction of potassium tetrabromoplatinate(II) with SnCl2 in hydrobromic acid in the presence of tetraphenylphosphonium bromide affords cis-[PPh4]2[PtBr2(SnBr3)2]. The insertion of SnCl2 into Pt–Cl bond of platinum(II) complexes cis-[PtCl2(L2)] {L2 = (PPh3)2; (PMe3)2; {P(OMe)3}2; dppm (bis(diphenylphosphino)methane); dppa (bis(diphenylphosphino)amine); and dppe (1,2-bis(diphenylphosphino)ethane)} is described.  相似文献   

14.
Reduction of [(C5H5)CoI2]2 by sodium amalgam in toluene in the presence of 1,3-butadiene, 1,3-cyclohexadiene or 1,5-cyclooctadiene affords the corresponding cyclopentadienylcobalt(I) diolefin complexes in high yields. Reduction of [(C5H5)CoI2]2 in the presence of 2-butyne yields the binuclear metallocyclic compound (C5H5)2Co2(C4(CH3)4), previously characterized as a structurally fluxional catalyst for alkyne cyclotrimerisation, as the major product; a trinuclear dicarbyne compound, (C5H5)3Co3(C-CH3)2, is obtained as a minor product. With diphenylacetylene, the analogous phenylcarbyne derivative (C5H5)3Co3(C-C6H5)2, previously obtained from thermal reaction with (C5H5)Co(CO)2, is obtained along with the major product, the tetraphenylcyclobutadiene complex (C5H5)Co(C4(C6H5)4). Pathways and intermediates for these reactions are discussed.  相似文献   

15.
The betain like carbodiphosphorane CO2 adduct O2CC(PPh3)2 ( 1a ) can serve as a ligand versus hard Lewis acids from main group compounds. Thus, reaction of 1a with InCl3, InI3 and SnCl2 in polar solvents leads to the addition compounds [Cl3In{O2CC(PPh3)2}] ( 2 ), [Cl2SnO2CC(PPh3)2}] ( 3 ) and the salt like compound [I2In{O2CC(PPh3)2}2]I ( 4 ) in good yields. Whereas in the indium compounds 1a acts as a chelating ligand, in the tin compound the molecule coordinates with one oxygen atom only as a monodentate ligand. 4 has a pyramidal structure with a stereochemical active pair of electrons. All compounds could be characterized by X‐ray analyses and the usual spectroscopic methods.  相似文献   

16.
The reactions between either BiBr2Ph and Na/K[Mo(CO)3(η? C5H5)], [BiPh{Mo(CO)3(η? C5H5)}2] and BiBr2Ph, or BiBr3 and BiPh3 and [Bi{Mo(CO)3(η? C5H5)}3] afford the complex [BiBrPh{Mo(CO)3(η? C5H5)}] 1 which has been characterised by X-ray crystallography. Complex 1 comprises a bismuth centre bonded to a bromine atom, a phenyl group and a Mo(CO)3(η? C5H5) fragment together with a longer secondary intermolecular interaction between a bromine from an adjacent molecule which results in a one-dimensional polymeric structure. Addition of a source of bromide anion to 1 affords the anionic complex [BiBr2Ph{Mo(CO)3(η? C5H5)}]? 3 ? although prolonged reaction results in the complex [BiBr2{Mo(CO)3(η? C5H5)}2]? 5 ? which was characterised by X-ray crystallography as its [Ph4P]+ salt. Complex 5 ? comprises a mononuclear bismuth centre bonded to two bromine atoms and two Mo(CO)3(η? C5H5) fragments in a geometry which lies between equatorially vacant trigonal bipyramidal and tetrahedral. The complex [PPN]2[Bi2Cl6{Mo(CO)3(η? C5H5)}2] 8 has also been synthesised and characterised by X-ray crystallography. A dimeric dianion is observed which can be viewed as two edge-shared square-based pyramids with chlorine atoms in the basal planes and Mo(CO)3(η? C5H5) fragments in the apical positions on opposite sides of the Bi2Cl6 plane.  相似文献   

17.
In an M-T-O model system (M is a polyvalent metal; T = Ge or Si), we consider initial stages of formation of cyclic MT clusters and the mechanism of their modification by T tetrahedra. The polyhedron ratio T/M in clusters increases progressively during modeling from one in M2T2 to two (M2T2 + 2T = M2T4), three (M2T2 + 2T2 = M2T6), and four (M2T2 + 2T + 2T2 = M2T8). These types of clusters were used to find precursor clusters for T-condensed structures of Na2Pr6Ge8O26, Na4Sc2Ge4O13, and Na5ScGe4O12. The TOPOS program package was used to carry out the complete 3D reconstruction of the self-assembly of Na,TR germanates: precursor cluster → primary chain → microlayer → microframework (supraprecursor) → ... framework. In all structures, as previously in six orthotetrahedral Na,TR germanate structures, the basic invariant type of four-polyhedral cyclic precursor cluster M2T2 was identified; this cluster is built of TR polyhedra, with CN = 6 or 7, linked via orthotetrahedra. The features of the generation of a Ge radical were considered in the form of a Ge2O7 chain and a Ge4O12 ring in various layers of the Na2Pr6Ge8O26 composite structure, a Ge4O13 chain in Na4Sc2Ge4O13, and a Ge12O36 ring in the Na5ScGe4O12 superionic conductor. Original Russian Text ? G.D. Ilyushin, L.N. Dem’yanets, 2009, published in Zhurnal Neorganicheskoi Khimii, 2009, Vol. 54, No. 3, pp. 484–496.  相似文献   

18.
C2‐C70(CF3)8 was found to be a very promising substrate in the Bingel and the Bingel–Hirsch reactions combining perfect regioselectivity with much higher reactivity compared to its analogs. The reactions with diethyl malonate yield a single isomer of the monoadduct C70(CF3)8[C(CO2Et)2] and a single C2‐symmetrical bisadduct C70(CF3)8[C(CO2Et)2]2. The Bingel–Hirsch variation is particularly interesting in that it additionally affords, in a similar regioselective manner, the unexpected alkylated derivatives C70(CF3)8[CH(CO2Et)2]H and C70(CF3)8[C(CO2Et)2][CH(CO2Et)2]H. The novel compounds have been isolated and structurally characterized by means of 1H and 19F NMR spectroscopy as well as single‐crystal X‐ray diffraction. The mechanistic and regiochemical aspects of the reaction are explained with the aid of DFT calculations.  相似文献   

19.
Reaction of 2,2-Dimethylpropylidynephosphine with Molybdenum Pentachloride; Crystal Structure of [Mo2Cl6(α,α′-dipyridyl)3] 2,2-Dimethylpropylidynephosphine and molybdenum pentachloride dissolved in POCl3 react with oxydation of the phosphorus and reduction of the molybdenum atom to give the alkyne complex [Mo2Cl4(μ-Cl)2(μ-H9C4? C?C? C4H9)(OPCl3)2]. Addition of α,α′-dipyridyl or of methyltriphenylphosphonium chloride in dichloromethane results in a displacement of the ligands POCl3 and H9C4? C?C? C4H9 from this complex and in the formation of [Mo2Cl6(dipy)3] or [(H5C6? )3P? CH3]3[Mo2Cl9]. Besides the latter compound small amounts of [(H5C6? )3P? CH3]2[MoCl6] can be isolated from the reaction mixture. [Mo2Cl6(dipy)3] which has already been prepared by other methods crystallizes in the monoclinic space group P21/c with {a = 1612; b = 148; c = 1296 pm; γ 109.3°; Z = 4} at 20°C. As shown by a crystal structure determination the complex is built up from [MoCl2(dipy)2]+ cations and [MoCl4(dipy)]? anions. The molybdenum atoms are both octahedrally surrounded. With average values of 238 and 243 pm the Mo? Cl bond distances in the cation, where a cis-arrangement of the chlorine atoms is observed, and in the anion differ significantly from each other. [Mo2Cl6(dipy)3] which has already been prepared by other methods crystallizes in the monoclinic space group P21/c with {a = 1612; b = 148; c = 1296 pm; γ = 109.3°; Z = 4} at 20°C. As shown by a crystal structure determination the complex is built up from [MoCl2(dipy)2]+ cations and [MoCl4(dipy)]? anions. The molybdenum atoms are both octahedrally surrounded. With average values of 238 and 243 pm the Mo? Cl bond distances in the cation, where a cis-arrangement of the chlorine atoms is observed, and in the anion differ significantly from each other.  相似文献   

20.
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   

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