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.
K2MCl5(M = La—Dy) and Rb2MCl5 (M = La—Eu): Ino-chlorides with Seven—Coordinated Rare Earths K2PrCl5 crystallizes with a = 1263.1(8), b = 875.6(3), c = 797.3(4) pm, Z = 4, orthorhombic, Pnma, isotypic to e. g. Y2HfS5. Monocapped trigonal prisms (C.N. = 7) are connected to chains via common edges in [010] direction according to [PrCl3/1tCl4/2k]2? with d?(Pr? Cl) = 281 pm. The chlorides K2MCl5 (M = La—Dy) and Rb2MCl5 (M = La—Eu) are isotypic to K2PrCl5. Thermal expansion of Rb2PrCl5 in [010] direction is remarkably smaller than parallel to (010).  相似文献   

2.
The equilibrium vaporization of CuInS2(s) was studied by Knudsen cell mass spectrometric techniques in the temperature range 902—1110 K. Based on these and independent congruency studies, CuInS2(s) decomposes under steady-state conditions according to the reaction 2CuInS2(s) = Cu2S(s) + In2S(g) + S2(g). During the initial transient period of the vaporization of CuInS2, the ion intensities of In2S+ and S2+ change and assume steady-state values after some time. A second-law evaluation of intensity versus temperature data, obtained under steady-state conditions, yielded a value for the heat of the above decomposition reaction (2 moles of CuInS2) of δH 630.9 ± 26 kJ. With this quantity and auxiliary thermochemical data, the enthalpy of formation of CuInS2(s) was computed to be δH = ?221.7 ± 13 kJ/mol.  相似文献   

3.
Cyclometalated Pd(II) complexes generally show inferior luminescence properties compared with their Pt(II) analogues. The established approach employing tridentate cyclometalating ligands has allowed us to create a series of square planar Pd(II) complexes [Pd( )X] from their protoligands H (2-(6-phenylpyridin-2-yl)thiazoles and -benzothiazoles; coligands X=Cl, Br, I) with extensive variations at the Carene group (phenyl, naphthyl, fluorenyl), the central Npyridine (pyridine, 4-phenylpyridine, 3,5-di-tert-butyl-4-phenylpyridine), and the peripheral Nthiazole (thiazole, benzothiazole) to probe for structural factors that might enhance efficient luminescence. Long-wavelength bands at 400–500 nm were assigned to transitions into mixed ligand-centred/metal-to-ligand charge transfer (MLCT) states based on time-dependent (TD)DFT calculations. The MLCT contributions are rather low, in agreement with relatively long lifetimes and high photoluminescence quantum yields of up to 0.79 recorded in frozen glassy solvent matrices at 77 K along with emission bands showing pronounced vibrational progressions and peaking at about 520 nm. No photoluminescence was observed at 298 K in solution. Variation of the ligand allowed to shift the experimental absorption energies from about 2.4 to 2.7 eV, in good agreement with the electrochemical band gaps (2.58 to 2.81 eV). The theoretical absorption and emission spectra excellently reproduced the experimental trends.  相似文献   

4.
Formation of Gaseous POCl Mass-spectrometric and Matrix-IR Investigations The formation of gaseous POCl by reaction of POCl3 with silver at about 1100 K has been shown by mass spectrometric and IR matrix investigations. The heat of formation and entropie has been calculated at ΔH298298 (POCl) = ?250.7 kJ/Mol, S298 (POCl) = 274.5 J/K · Mol. An approximate force constant calculation is presented as the three possible vibrational frequencies (1257.7 cm?1, 489.4 cm?1, 308.0 cm?1) and some isotopic shifts (16O/18O; 35Cl/37Cl) are observed in an Ar-matrix.  相似文献   

5.
6.
The heterogeneous ozonolysis of sodium oleate aerosols in an aerosol flow tube is reported under different relative humidity (RH%) conditions. Submicron sodium oleate particles were exposed to a known ozone concentration and the consumption of sodium oleate was monitored by infrared spectroscopy. When the experimental results are treated as a surface‐mediated reaction (i.e., following a Langmuir–Hinshelwood type mechanism), the following parameters are obtained: at low RH%, = (3 ± 1) × 10?16 cm3 molecule?1 and = (0.046 ± 0.006) s?1; at high RH%, = (6 ± 2) × 10?16 cm3 molecule?1 and = (0.21 ± 0.05) s?1. From these pseudo–first‐order coefficients, the reactive uptake coefficients for dry and aqueous sodium oleate aerosols are calculated as (1.5 ? 0.5) × 10?7 and (1.7 ? 0.7) × 10?6, respectively. Hydrated oleate aerosols display both an increase in the ozone trapping ability and an increase in the effective rate reaction at the droplet surface compared to dry aerosol surfaces. These observations may provide an explanation for some of the variability observed between lab studies of dry ozonolysis and real‐world, atmospheric lifetimes of oleic acid–related species.  相似文献   

7.
Endohedral metalloborofullerenes (EMBFs) are novel boron analogues of the famous endohedral metallofullerenes (EMFs). Many EMBFs have been proposed by theoretical calculations thus far. However, in sharp contrast to EMFs, which trap most of the lanthanides with f electrons inside the cages, the corresponding lanthanide‐based EMBFs have never been reported. In this work, the encapsulation of Eu and Gd in the B38 and B40 fullerenes was studied by means of density functional theory calculations. Our results revealed that Gd@B38(9A), Eu@B40(8B2), and Gd@B40(7A″) all favor the endohedral configuration, and the electronic structures can be described as Gd3+@ , Eu2+@ , and Gd3+@ with jailed f electron spins. The large binding energies and sizable HOMO–LUMO gaps suggest that they may be achieved experimentally. They feature σ and π double aromaticity, and their excellent stabilities were confirmed by the Born–Oppenheimer molecular dynamics simulations. Finally, the infrared and UV/vis spectra were simulated to assist experimental characterization.  相似文献   

8.
A joint theoretical and experimental study on 32 endohedral silafullerane derivatives [X@Si20Y20] (X=F-I; Y=F-I, H, Me, Et) and -[Cl@Si20H12Y8] (Y=F-I) is presented. First, we evaluated the structure-determining template effect of Cl in a systematic series of concave silapolyquinane model systems. Second, we investigated the X→Si20 interaction energy ( ) as a function of X and Y and found the largest values for electron-withdrawing exohedral substituents Y. Given that X ions can be considered as Lewis bases and empty Si20Y20 clusters as Lewis acids, we classify our inseparable host–guest complexes [X@Si20Y20] as “confined Lewis pairs”. Third, 35Cl NMR spectroscopy proved to be highly diagnostic for an experimental assessment of the Cl→Si20 interaction as the paramagnetic shielding and, in turn, (35Cl) of the endohedral Cl ion correlate inversely with . Finally, we disclose the synthesis of [PPN][Cl@Si20Y20] (Y=Me, Et, Br) and provide a thorough characterization of these new silafulleranes.  相似文献   

9.
Complex [PtMe2(PMe2Ar )] ( 1 ), which contains a tethered terphenyl phosphine (Ar =2,6‐(2,6‐i Pr2C6H3)2C6H3), reacts with [H(Et2O)2]BArF (BArF=B[3,5‐(CF3)2C6H3]4) to give the solvent (S) complex [PtMe(S)(PMe2Ar )]+ ( 2⋅S ). Although the solvent molecule is easily displaced by a Lewis base (e.g., CO or C2H4) to afford the corresponding adducts, treatment of 2⋅S with C2H2 yielded instead the allyl complex [Pt(η3‐C3H5)(PMe2Ar )]+ ( 6 ) via the alkyne intermediate [PtMe(η2‐C2H2)(PMe2Ar )]+ ( 5 ). Deuteration experiments with C2D2, and kinetic and theoretical investigations demonstrated that the conversion of 5 into 6 involves a PtII‐promoted HC≡CH to :C=CH2 tautomerization in preference over acetylene migratory insertion into the Pt−Me bond.  相似文献   

10.
The controlled folding of a single polymer chain is for the first time realized by metal‐ complexation. α,ω‐Bromine functional linear polymers are prepared via activators regenerated by electron transfer (ARGET) ATRP (,SEC = 5900 g mol−1, Đ = 1.07 and 12 000 g mol−1, Đ = 1.06) and the end groups of the polymers are subsequently converted to azide functionalities. A copper‐catalyzed azide–alkyne cycloaddition (CuAAC) reaction is carried out in the presence of a novel triphenylphosphine ligand and the polymers to afford homotelechelic bis‐triphenylphosphine polymeric‐macroligands (MLs) (,SEC = 6600 g mol−1, Đ = 1.07, and 12 800 g mol−1, Đ = 1.06). Single‐chain metal complexes (SCMCs) are formed in the presence of Pd(II) ions in highly diluted solution at ambient temperature. The results derived via 1H and 31P{1H} NMR experiments, SEC, and DLS unambiguously evidence the efficient formation of SCMCs via metal ligand complexation.

  相似文献   


11.
Dimethylamine (DMA) ignition delay times and OH time histories during the oxidation process were investigated behind reflected shock waves. The ignition delay time measurements cover the temperature range of 1181–1498 K, with pressures near 0.9, 1.5, and 2.8 atm, and equivalence ratios of 0.5, 1, and 2, in 4% oxygen/argon. The ignition delay time data feature low scatter and can be correlated to a single expression with 2 ~ 0.99: τign = 7.30 × 10?4 ?0.68 Φ0.45 exp(18,265/), where τign is in μs, in atm, and in K. OH time histories were measured using laser absorption of the R1(5) line of the A‐X(0,0) transition near 306.7 nm, in stoichiometric mixtures of 500 ppm DMA/O2/argon. The mechanism developed by Li et al. was used initially to simulate the measured DMA ignition delay times and the OH time histories. The Li et al. mechanism was then updated by adding the DMA unimolecular decomposition channel: DMA = CH3NH + CH3, with the reaction rate constant estimated by analogy to dimethyl ether decomposition, previously investigated by Cook et al. The reactions of DMA + OH were also updated based on recent work in our laboratory. The simulation results using the modified Li et al. mechanism are in good agreement with both the ignition delay times and OH time‐history data.  相似文献   

12.
Hyperbranched polymers formed through step polymerization of AB2‐type monomer with equal reactivity for both B groups in a continuous flow stirred‐tank reactor (CSTR) are investigated theoretically. The weight fraction distribution at high molecular weight tail follows a power law, W (P ) ∝ P −1/ξ for ξ ≤ 0.5 with , where is the mean residence time. The degree of branching (DB) at the large degree of polymerization (P ) limit is DB P →∞ = 0.6 irrespective of the ξ‐value, which is larger than the case for the corresponding batch polymerization that gives DB P →∞ = 0.5. The relationship between the radius of gyration 〈s 20 and P shows that the hyperbranched polymers formed in a CSTR are very compact, and the 〈s 20‐values for large polymers are even smaller than the smallest possible case for a batch reactor with DB P →∞ = 1. For large polymers, the power law 〈s 20P 1/3 holds, which is 〈s 20P 1/2 for batch polymerization.

  相似文献   


13.
The theoretical data for the half-lantern complexes [{Pt( )(μ- )}2] [ 1 – 3 ; is cyclometalated 2-Ph-benzothiazole; is 2-SH-pyridine ( 1 ), 2-SH-benzoxazole ( 2 ), 2-SH-tetrafluorobenzothiazole ( 3 )] indicate that the Pt ⋅⋅⋅ Pt orbital interaction increases the nucleophilicity of the outer d orbitals to provide assembly with electrophilic species. Complexes 1 – 3 were co-crystallized with bifunctional halogen bonding (XB) donors to give adducts ( 1 – 3 )2 ⋅ (1,4-diiodotetrafluorobenzene) and infinite polymeric [ 1⋅ 1,1′-diiodoperfluorodiphenyl]n. X-ray crystallography revealed that the supramolecular assembly is achieved through (Aryl)I ⋅⋅⋅ d [PtII] XBs between iodine σ-holes and lone pairs of the positively charged (PtII)2 centers acting as nucleophilic sites. The polymer includes a curved linear chain ⋅⋅⋅ Pt2 ⋅⋅⋅ I(areneF)I ⋅⋅⋅ Pt2 ⋅⋅⋅ involving XB between iodine atoms of the perfluoroarene linkers and (PtII)2 moieties. The 195Pt NMR, UV/Vis, and CV studies indicate that XB is preserved in CH(D)2Cl2 solutions.  相似文献   

14.
Designing and characterizing the compounds with exotic structures and bonding that seemingly contrast the traditional chemical rules are a never‐ending goal. Although the silicon chemistry is dominated by the tetrahedral picture, many examples with the planar tetracoordinate‐Si skeletons have been discovered, among which simple species usually contain the 17/18 valence electrons. In this work, we report hitherto the most extensive structural search for the pentaatomic ptSi with 14 valence electrons, that is, (n + m = 4; q = 0, ±1, ?2; X, Y = main group elements from H to Br). For 129 studied systems, 50 systems have the ptSi structure as the local minimum. Promisingly, nine systems, that is, , HSiY3 (Y = Al/Ga), Ca3SiAl?, Mg4Si2?, C2LiSi, Si3Y2 (Y = Li/Na/K), each have the global minimum ptSi. The former six systems represent the first prediction. Interestingly, in HSiY3 (Y = Al/Ga), the H‐atom is only bonded to the ptSi‐center via a localized 2c–2e σ bond. This sharply contradicts the known pentaatomic planar‐centered systems, in which the ligands are actively involved in the ligand–ligand bonding besides being bonded to the planar center. Therefore, we proposed here that to generalize the 14e‐ptSi, two strategies can be applied as (1) introducing the alkaline/alkaline‐earth elements and (2) breaking the peripheral bonding. In light of the very limited global ptSi examples, the presently designed six systems with 14e are expected to enrich the exotic ptSi chemistry and welcome future laboratory confirmation. © 2014 Wiley Periodicals, Inc.  相似文献   

15.
Using harmonic and anharmonic DFT calculations, we have established a general correlation between B–H stretching frequencies and B–H bond lengths valid for the closoboranes (= 6 – 12), substituted closoboranes B12H12 – (with X = F, Cl, Br and = 1 – 3 and 9 – 12) and the carboranes and , suggesting that this correlation is also applicable to other similar species. It appears that the average B–H stretching frequency observed around 2500 cm−1 shift by about −100 cm−1 if the average B–H bond length increases by 1 pm. In contrast to , the B–H bond in closoboranes is practically covalent and the correlation evidenced between its stretching frequency and its length proves to be similar to the one observed for the C–H bond.  相似文献   

16.
The synthesis of the first 4d transition metal oxide–hydride, LaSr3NiRuO4H4, is prepared via topochemical anion exchange. Neutron diffraction data show that the hydride ions occupy the equatorial anion sites in the host lattice and as a result the Ru and Ni cations are located in a plane containing only hydride ligands, a unique structural feature with obvious parallels to the CuO2 sheets present in the superconducting cuprates. DFT calculations confirm the presence of S= Ni+ and S=0, Ru2+ centers, but neutron diffraction and μSR data show no evidence for long‐range magnetic order between the Ni centers down to 1.8 K. The observed weak inter‐cation magnetic coupling can be attributed to poor overlap between Ni 3d and H 1s in the super‐exchange pathways.  相似文献   

17.
Hyperbranched polymer formation during step polymerization of AB2 type monomer with equal reactivity of two B's is investigated theoretically, focusing the attention to the degree of branching (DB) and the mean square radius of gyration for the unperturbed chains, . It is found that the DB‐value at large degree of polymerization (P) limit, = 0.5 is unchanged during the whole course of polymerization. The average value of having the same P is invariant throughout the polymerization. The universal curve between and P agrees perfectly with that for the self‐condensing vinyl polymerization (SCVP), another method to synthesize hyperbranched polymers, when the reactivity ratio for SCVP, rSCVP, is 2.589 that gives = 0.5. The power law, is found for large values of P.

  相似文献   


18.
At ultrahigh pressure (>110 GPa), H2S is converted into a metallic phase that becomes superconducting with a record Tc of approximately 200 K. It has been proposed that the superconducting phase is body‐centered cubic H3S (Im m, a=3.089 Å) resulting from the decomposition reaction 3 H2S→2 H3S+S. The analogy between H2S and H2O led us to a very different conclusion. The well‐known dissociation of water into H3O+ and OH? increases by orders of magnitude under pressure. H2S is anticipated to behave similarly under pressure, with the dissociation process 2 H2S→H3S++SH? leading to the perovskite structure (SH?)(H3S+). This phase consists of corner‐sharing SH6 octahedra with SH? ions at each A site (the centers of the S8 cubes). DFT calculations show that the perovskite (SH?)(H3S+) is thermodynamically more stable than the Im m structure of H3S, and suggest that the A site hydrogen atoms are most likely fluxional even at Tc .  相似文献   

19.
The bonding problem in borazine (B3N3H6), boroxine (B3O3H3), and carborazine (B2N2C2H6) is successfully addressed through the consideration of the excited states of the constituent fragments, namely BH( ), NH( ), and CH( ). We propose the participation of resonant structures for all three species that help to explain the experimental findings. A discussion on the chemical pattern of the parental molecule benzene (C6H6) helps to make coherent the whole bonding analysis on the titled species.  相似文献   

20.
Ketal‐substituted bridged azobenzenes have been synthesized; these display a symmetrical boat conformation with the ketal in pseudo‐equatorial positions. These bridged Z‐azobenzenes (Z1) readily photoisomerize to the E‐isomer as well as another Z‐conformer (Z2) with ketal function on the pseudo‐axial position upon irradiation at 406 nm. The two diastereomeric conformers display distinct physicochemical characteristics. Spectroscopic and NMR investigations supported that interconversion of two conformers occurs via the E‐isomer, with good photochemical quantum yield (Φ =0.45±0.03, Φ =0.33±0.05, Φ =0.37±0.06 and Φ =0.36±0.04). The system shows high photostability and no thermal equilibrium between the two stable Z1 and Z2 conformers.  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