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 15 毫秒
1.
Photo-induced complex formation of tris-2,2′-bipyridine iron(II) complex ([Fe(bpy)3]2+) from the mixture of FeCl3 and 2,2′-bipyridine was achieved in silica gel containing 150-300 μm silica particles, derived from a complex emulsion with HCl aqueous solution and tetraethyl orthosilicate (TEOS). More than 95% of Fe(III) and 2,2′-bipyridine were incorporated in silica particles. Yellow-red color change, due to [Fe(bpy)3]2+, was observed by irradiation with 365 nm UV beam at 0.3 mW cm−2 for 120 s. The complex formation accompanies simultaneous spin transition from the high-spin state of Fe(III) to the low-spin state of Fe(II).  相似文献   

2.
The longitudinal magnetoresistance of a two-dimensional superconductor, β’-Et2Me2P[Pd(dmit)2]2 (dmit=C3S 5 2− ) under hydrostatic pressure was measured at low temperatures with the field applied perpendicular to the conducting layers. At 0.58 GPa, the field-dependent isothermal interlayer resistanceR╧ (H) exhibited a peak below the superconducting transition temperature T C . This peak effect can be explained by a model of resistively-shunted Josephson-Junctions. The peak is strongly suppressed at a higher pressure, 0.71 GPa.  相似文献   

3.
The Fe(II) spin crossover complex [Fe{H2B(pz)2}2(bipy)] (pz = pyrazol‐1‐yl, bipy = 2,2′‐bipyridine) can be locked in a largely low‐spin‐state configuration over a temperature range that includes temperatures well above the thermal spin crossover temperature of 160 K. This locking of the spin state is achieved for nanometer thin films of this complex in two distinct ways: through substrate interactions with dielectric substrates such as SiO2 and Al2O3, or in powder samples by mixing with the strongly dipolar zwitterionic p ‐benzoquinonemonoimine C6H2(—? NH2)2(—? O)2. Remarkably, it is found in both cases that incident X‐ray fluences then restore the [Fe{H2B(pz)2}2(bipy)] moiety to an electronic state characteristic of the high spin state at temperatures of 200 K to above room temperature; that is, well above the spin crossover transition temperature for the pristine powder, and well above the temperatures characteristic of light‐ or X‐ray‐induced excited‐spin‐state trapping. Heating slightly above room temperature allows the initial locked state to be restored. These findings, supported by theory, show how the spin crossover transition can be manipulated reversibly around room temperature by appropriate design of the electrostatic and chemical environment.  相似文献   

4.
This paper presents the solid-state electrical conductance properties of a series of complex bimetallic salts of the form [M(N–N)3][Cu(MNT)2] (M=Fe(II), Co(II), Ni(II), or Cd(II); N–N=1,10-phenanthroline (phen) or ethylenediamine (en); MNT2−=maleonitriledithiolate) and bridged heterobimetallic complexes Ag2[Cu(MNT)2] and Hg[Cu(MNT)2] that have been prepared by treatment of complex salt Na2[Cu(MNT)2] (generated in situ) with one equivalent of cationic complexes [M(N–N)3]X2 or Hg(CH3COO)2 and two-equivalent of AgNO3 in aqueous–methanol mixture and characterised by relevant spectroscopies (IR, EPR, UV-Visible) as well as by powder XRD spectra. Solution conductivity measurements in 10−3 M DMSO solution revealed 1:1 electrolytic behaviour of the bimetallic salts. Diamagnetism together with powder EPR spectra for Ag2[Cu(MNT)2] and Hg[Cu(MNT)2] show strong antiferromagnetic interaction between two adjacent copper(II) centers at room temperature. Majority of the complexes exhibited compressed pellet σrt in 8.19×10−11 to 5.37×10−7 S cm−1 range and show semiconductivity over the 303–383 K temperature range. Conductivity of both coordination polymers are appreciably higher compared to the bimetallic salts. For the salts [Co(phen)3][Cu(MNT)2], [Ni(phen)3][Cu(MNT)2] and bridged complexes Ag2[Cu(MNT)2] and Hg[Cu(MNT)2] the conductivity remarkably increases, i.e., 102 to 103 order of magnitude at elevated temperature showing some sort of phase transformation producing S⋯S intermolecular contact.  相似文献   

5.
X-ray Photoelectron Spectroscopy and Raman spectroscopy are combined to evaluate the stoichiometry and charge state of building units in conductive films containing TTF, ET or TMTSF donors and [Ni(dmit)2] or [M(dcbdt)2] acceptors.  相似文献   

6.
The kinetics of the reaction of Np(V) with Fe(II) in dilute perchloric and nitric acid solutions containing H2C2O4 was studied by spectrophotometry. In the range pH 1–2, the reaction rate is described by the equation d[Np(V)]/dt = k[Np(V)][Fe(II)][H2C2O4]2[H+]−1.6, k = 182 mol−1.4 l1.4 s−1. The activation energy in the range 25–45°C is 26 kJ mol−1. The reaction mechanism involves formation of Fe(II) and Np(V) oxalate complexes, followed by their reaction with the participation of the H+ ion.  相似文献   

7.
The kinetics of reduction of Pu(IV) and Np(VI) with butanal oxime in undiluted TBP containing HNO3 was studied spectrophotometrically. In the range [HNO3] = 0.08-0.75 M the rate of Pu(IV) reduction is described by the equation -d[Pu(IV)]/dt = k[Pu(IV)]2[C3H7CHNOH]/{[Pu(III)][HNO3]2} with the rate constant k = 0.068±0.017 mol l-1 min-1 at 20°C. The kinetic equation of the reduction of Np(VI) to Np(V) in the range [HNO3] = 0.01-0.27 M is -d[Np(VI)]/dt = k[Np(VI)][C3H7CHNOH][H2O]2/[HNO3]0.5, where k = 0.058±0.007 l2.5 mol-2.5 min-1 at 25°C, and the activation energy is 79±9 kJ mol-1.  相似文献   

8.
The realization of spin‐crossover (SCO)‐based applications requires study of the spin‐state switching characteristics of SCO complex molecules within nanostructured environments, especially on surfaces. Except for a very few cases, the SCO of a surface‐bound thin molecular film is either quenched or heavily altered due to: (i) molecule–surface interactions and (ii) differing intermolecular interactions in films relative to the bulk. By fabricating SCO complexes on a weakly interacting surface, the interfacial quenching problem is tackled. However, engineering intermolecular interactions in thin SCO active films is rather difficult. Here, a molecular self‐assembly strategy is proposed to fabricate thin spin‐switchable surface‐bound films with programmable intermolecular interactions. Molecular engineering of the parent complex system [Fe(H2B(pz)2)2(bpy)] (pz = pyrazole, bpy = 2,2′‐bipyridine) with a dodecyl (C12) alkyl chain yields a classical amphiphile‐like functional and vacuum‐sublimable charge‐neutral FeII complex, [Fe(H2B(pz)2)2(C12‐bpy)] (C12‐bpy = dodecyl[2,2′‐bipyridine]‐5‐carboxylate). Both the bulk powder and 10 nm thin films sublimed onto either quartz glass or SiOx surfaces of the complex show comparable spin‐state switching characteristics mediated by similar lamellar bilayer like self‐assembly/molecular interactions. This unprecedented observation augurs well for the development of SCO‐based applications, especially in molecular spintronics.  相似文献   

9.
Thin films of the molecular spin‐crossover complex [Fe(HB(1,2,4‐triazol‐1‐yl)3)2] undergo spin transition above room temperature, which can be exploited in sensors, actuators, and information processing devices. Variable temperature viscoelastic mapping of the films by atomic force microscopy reveals a pronounced decrease of the elastic modulus when going from the low spin (5.2 ± 0.4 GPa) to the high spin (3.6 ± 0.2 GPa) state, which is also accompanied by increasing energy dissipation. This technique allows imaging, with high spatial resolution, of the formation of high spin puddles around film defects, which is ascribed to local strain relaxation. On the other hand, no clustering process due to cooperative phenomena was observed. This experimental approach sets the stage for the investigation of spin transition at the nanoscale, including phase nucleation and evolution as well as local strain effects.  相似文献   

10.
The57 Fe Mőssbauer resonance has been applied to study polyacetylene films doped chemically and electrochemically to different doping levels. Both methods lead to the insertion of a high spin FeIII complex with Mőssbauer parameters characteristic of FeCl?4. Degradation of [CH(FeCl4)y]x in air leads to gradual transformation of FeIII into a high spin FeII.  相似文献   

11.
Abstract

Recent studies on the physical properties of the triangular system based on the Pd(dmit)2 salts (dmit=1,3-dithiole-2-thione-4,5-dithiolate) are reviewed. Quantum chemical architectures of the Pd(dmit)2 molecule and its dimer are introduced with emphasis on the strong dimerization of a two-level system, which provides unique physical properties of the salts. The magnetic properties are outlined in view of the magneto-structural correlation specific to the frustrated spin systems. Some newly discovered ground states and their origins are discussed, for which the valence bond formation plays a key role. Among them, the two-level structure is crucial for the novel charge-separated state found in two salts. The valence bond ordering, similar to the spin-Peierls transition, has been found in a two-dimensional frustrated spin system. The physical aspects and possible relation to the pressure-induced superconductivity are discussed.  相似文献   

12.
Oxidation of U(IV) with nitric acid in 30% solution of TBP in n-dodecane is catalyzed by Tc ions; the rate-determining steps are 3U(IV) + 2Tc(VII) → 3U(VI) + 2Tc(IV) and Tc(IV) + Tc(VII) → Tc(V) + Tc(VI). Oxidation of U(IV) is inhibited by the reaction product, HNO2, which partially binds Tc(IV) ions (TcO2+) in an inert complex. The overall rate equation of U(IV) oxidation is-d[U(IV)]/dt = k 1[U(IV)][Tc][HNO3]?3 ? k 4[U(IV)]2[HNO2]2[HNO3]?1, where k 1 = 4.8 ± 1.0 mol21?2 min?1 and k 4 = (2.4 ± 1.0) × 105 12 mol?2 min?1 at 25°C, [H2O] = 0.4 M ([Tc] is the total Tc concentration in the reaction mixture). Water and U(VI) have no effect on the reaction rate.  相似文献   

13.
Sorption of 85Sr, 137Cs, 22Na, and 152Eu on solid mixed potassium neodymium ferrocyanide KNd[Fe(CN)6]·4H2O from neutral, acidic, and alkaline media and also coprecipitation of these radionuclides with KNd[Fe(CN)6]·4H2O in its formation from a homogeneous solution were studied. It was found that 85Sr and 22Na do not noticeably coprecipitate with solid KNd[Fe(CN)6]·4H2O and are not sorbed by this substance. In aqueous medium, depending on the cesium concentration in solution, from 80 to 98% of 137Cs coprecipitates with solid KNd[Fe(CN)6]·4H2O. In this case, the distribution coefficient Kd depends on both the cesium concentration in solution and solution pH. Within 30 min of contact of the solid and liquid phases, the degree of recovery of 137Cs from aqueous solution with KNd[Fe(CN)6]·4H2O is approximately 95.0% of the initial amount. 152Eu coprecipitates with solid KNd[Fe(CN)6]·4H2O during its formation from a homogeneous solution to 98–99.9%. The degree of recovery of 152Eu from aqueous solution with KNd[Fe(CN)6]·4H2O precipitate within 60 min of contact of the solid and liquid phases is 70.3% of the initial amount.  相似文献   

14.
Sorption and coprecipitation of U(VI) from aqueous solution containing various complexing anions (CO325-, SO42−, H2EDTA2−) with the Ni(OH)2 solid phase at 25°C was studied. Uranium(VI) is not noticeably sorbed on the Ni(OH)2 solid phase from aqueous solutions containing CO32− and SO42−. The distribution coefficients K d are less than 1.0 ml g−1 throughout the examined range of [U(VI)]: [L] ratios (L = CO32−, SO42−) at V/m ≥ 100 ml g−1 and contact time of the solid and liquid phases of 60 min. In the presence and in the absence of H2EDTA2−, the degree of the U(VI) sorption is essentially the same (K d ∼90–140 ml g−1 at V/m ≥ 100 ml g−1). Uranium(VI) does not coprecipitate with Ni(OH)2 from aqueous solutions containing SO42− and H2EDTA2−. The distribution coefficients K d are less than 0.001 ml g−1 at V/m ≥ 200 ml g−1 and contact time of the solid and liquid phases of 60 min. In solutions containing CO32−, the U(VI) capture by the Ni(OH)2 precipitate depends on the [CO32−]: [U(VI)] ratio. The higher the [CO32−]: [U(VI)] ratio, the more strongly U(VI) coprecipitates with Ni(OH)2.  相似文献   

15.
Coprecipitation of 60Co from aqueous solutions with solid phases of various coordination compounds was studied. Microamounts of 60Co do not noticeably coprecipitate with solid phases of [M(CE)]BPh4 (M = Na+, Cs+; CE = 12-crown-4, 15-crown-5, 18-crown-6) and of CsBPh4. 60Co can be efficiently removed from aqueous solutions containing 0.1 and 4.0 M NaNO3, 0.1 M EDTA, or 1.0 M H2C2O4 by coprecipitation with the KFe[Fe(CN)6] solid phase formed by successive addition of K4[Fe(CN)6] and Fe(NO3)3 into these solutions. The degree of 60Co removal varies from ~80 to ~99.9% depending on the experimental conditions. This procedure allows simultaneous removal of 137Cs from the same solution with no less than 97% efficiency.  相似文献   

16.
《晶体工程》2001,4(2-3):141-157
Crystal structures, absolute configurations, and crystalline packing features of nine complexes, namely, cis-α-Λ-(RR)(δλδ)-[Co(trien)(D-histidinato)](ClO4)2·2H2O 1, cis-β1-Δ-(RR)(λδδ)- [Co(trien)(L-asparaginato)](ClO4)2 2, cis-β2-Λ(SS)(λλδ)-[Co(trien)(L-valinato)](ClO4)2·H2O 3, cis-β2-Δ-(RR)(δδλ)[Co(trien)(D-methioninato)](ClO4)2 and its enantiomer 4, trans(N,t-N)-[Co(tren)(DL-leucinato)]X2, (X=ClO4 5, BF4 6, PF6 7), and trans(N,t-N)-[Co(tren)(DL- methioninato)]X2 (X=Br 8, Cl(BF4) 9) (trien=triethylenetetramine, tren=tris(2-aminoethyl)amine), have been determined. All compounds were prepared from racemic DL-amino acid. 13 crystallize as conglomerates. 57 are isomorphous and crystallize as the so-called conglomeratic solids. While 4, 8 and 9 undergo racemic crystallization. In 17, the carboxylic oxygen of the amino acid forms double hydrogen bonds with the amino hydrogen atoms of N4 and/or amino acidato of an adjacent cation. By these hydrogen bonds, cations having the same chirality are linked together into helical string. In the isomorphous 8 and 9, the cation containing L-methionine interacts with a cation containing D-methioninine, through hydrogen bonds, to form a racemic pair, and no spiral string arrangements are observed.  相似文献   

17.
Fe(III) ion can strongly inhibit the sulphidation amine flotation of smithsonite. However, its modification mechanism on smithsonite surface is still obscure. In this work, a systematic study of the modification of Fe(III) ion on smithsonite (1 0 1) surface was performed using DFT calculation. The optimal number of H2O ligands for Fe(III) ion hydrates in aqueous conditions was probed, and [Fe(OH)2(H2O)4]+ and [Fe(OH)4]? were identified as the major modification species, then their adsorption and bonding mechanisms were further revealed by analyzing the frontier orbitals, density of state, Mulliken population, and electron density. The calculated adsorption structures were consistent with the former experiment, and we found the O site that bonded to the C atom on smithsonite surface was the most favorable position for [Fe(OH)2(H2O)4]+ and [Fe(OH)4]? adsorptions. Besides, their adsorption mechanisms on smithsonite surface were principally due to the combined effect of FeO bond and hydrogen bonding. Simultaneously, hydrogen bonding greatly enhanced the stability of the adsorption structures. Moreover, the dominant orbital contribution for the bonding of FeO was primarily due to the orbital hybridization between Fe 3d and O 2p orbitals. This work can help in deeper understanding of the depression of Fe(III) ion on the sulphidation amine flotation of smithsonite.  相似文献   

18.
The luminescence and second order non linear optical (NLO) response of [Ir(ttpy)2(5-R-1,10-phen)][PF6] (ttpy = cyclometallated 3′-(2-pyridil)-2,2′:5′,2″-terthiophene, phen = phenanthroline; R = Me, NO2) and [Ir(pq)2(5-R-1,10-phen)][PF6] (pq = cyclometallated 2-phenylquinoline) have been investigated experimentally in CH2Cl2 solution and compared with that of [Ir(ppy)2(5-R-1,10-phen)][PF6] (ppy = cyclometallated 2-phenylpyridine), characterized by one of the highest second order NLO response ever reported for a metal complex. Substitution of ppy with the more π-delocalized pq does not affect significantly the luminescence and NLO properties. A slightly lower NLO response and a much poorer luminescence is observed for the related complexes with ttpy. In these complexes, DFT/TDDFT calculations show that the presence of ttpy induces a significant downshift of the HOMO energy, compared to ppy and pq. The NLO response is dominated by intense MLCT excited states, which are also assigned as originating the emission.  相似文献   

19.
Magnetic and electrical properties of a series of new inorganic complex based materials [n-Pr4N]2[Pb(ecda)2], [Et4N]4[Hg2(ecda)4] and M[M′(ecda)2] (M=Co(II), Ni(II), Cu(II), Cd(II) or Pb(II); M′=Hg(II) or Pb(II); ecda2−=1-ethoxycarbonyl-1-cyanoethylene-2,2-dithiolate) have been investigated using molecular spectroscopic (IR, Raman, EPR, electronic and mass) and conducting properties. Complete quenching of paramagnetism indicates strong Cu–Cu interaction in the layered polymeric array of Cu[Hg(ecda)2] and Cu[Pb(ecda)2] in the solid state. The 1H NMR spectrum of Ni[Pb(ecda)2] in DMSO-d6 gives narrow unshifted peaks due to diamagnetic, square-planar geometry around Ni(II) and shows absence of some octahedral units (μeff=1.49 BM) present in the solid state of this product. While the peak broadening indicates paramagnetic nature and absence of dominant Cu–Cu interaction in solution for Cu[Hg(ecda)2] and Cu[Pb(ecda)2]. The temperature dependent (305–373 K) compressed pellet conductivity together with activation energy (0.36–1.23 eV), show semiconducting behaviour of some of the bimetallic derivatives.  相似文献   

20.
Three new transition-metal dithiocarbamates involving ferrocene (Fc), namely [Co(FcCH2EtOHdtc)3] (Co), [M(FcCH2EtOHdtc)2] M = Ni (Ni), Cu (Cu) (EtOHdtc = N-ethanol dithiocarbamate), have been synthesized and characterized by microanalyses, FTIR, 1H and 13C NMR spectroscopies and single crystal X-ray diffraction technique. The peak broadening in the 1H spectrum of the copper complex indicates the paramagnetic behavior of this compound. The observed single quasi-reversible cyclic voltammograms for the complexes indicate the stabilization of a metal center (except copper) other than Fe in their characteristic oxidation state. These complexes have been used as photo-sensitizer in dye-sensitized solar cells which indicates that Co displays the best photosensitization property with an overall conversion efficiency of 3.25 ± 0.04%. The low cell efficiency of Ni and Cu complexes may be due to slow regeneration of the dye by iodine/iodide redox couple followed by charge injection into TiO2.  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