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.
Extensive study of the electronic structure of Fe‐NO complexes using a variety of spectroscopic methods was attempted to understand how iron controls the binding and release of nitric oxide. The comparable energy levels of NO π* orbitals and Fe 3d orbitals complicate the bonding interaction within Fe? NO complexes and puzzle the quantitative assignment of NO oxidation state. Enemark–Feltham notation, {Fe(NO)x}n, was devised to circumvent this puzzle. This 40‐year puzzle is revisited using valence‐to‐core X‐ray emission spectroscopy (V2C XES) in combination with computational study. DFT calculation establishes a linear relationship between ΔEσ2s*‐σ2p of NO and its oxidation state. V2C Fe XES study of Fe? NO complexes reveals the ΔEσ2s*‐σ2p of NO derived from NO σ2s*/σ2p→Fe1s transitions and determines NO oxidation state in Fe? NO complexes. Quantitative assignment of NO oxidation state will correlate the feasible redox process of nitric oxide and Fe‐nitrosylation biology.  相似文献   

2.
First single crystals of Na2[{(ON)Ce4}Cl9] were obtained during synthesis attempts for the cerium nitride chloride Ce2NCl3(= {N2Ce4}Cl6). With a molar ratio of 8:10:3 for Ce, CeCl3 and NaN3 along with an excess of the flux NaCl, the mixture obviously contained a small amount of CeOCl that led to the formation of quinary Na2[{(ON)Ce4}Cl9]. It crystallizes in the monoclinic space group P21/m(no. 11) with two formula units in the unit cell with dimensions of a = 813.21(6) pm, b = 1146.13(9) pm, c = 942.86(7) pm and β = 107.504(5) °. As the dominating structural feature {ZCe4}9.5+ tetrahedra are fused through trans‐oriented edges to generate chains (Z = 1/2 O + 1/2 N) just as in the structurally isotypic compounds A2[{Z2R4}X9] (A = Na, K; R = Pr, Nd, Gd; X = Cl, Br; Z = O, O/N).  相似文献   

3.
Structural Chemistry and Impedance Spectroscopy of Mg2+ stabilized Na+/Pr3+-β″-Aluminas The structure with the ionic distribution in the conduction planes of a Pr3+-exchanged Mg2+ stabilized Na+-β″-alumina crystal, composition determined by electron microprobe analysis to Na1.19Pr0.13Mg0.49Al10.49O17 (degree of exchange ξ = 25%, related to Na+ content), has been investigated by single crystal X-ray diffraction methods at room temperature (R3 m, Z = 3, a = 561.8(2) pm, c = 3 361.5(11) pm). Pr3+ is slightly shifted in the ab-plane from the centrosymmetric 9 d (mO) into a 18 h site. Na+ occupies 18 h as well as 6 c (BR) positions. Ionic conductivity data of Mg2+ stabilized Na+/Pr3+-β″-Al2O3 crystals with varying degree of exchange determined by impedance spectroscopy are given in an Arrhenius diagram. With growing Pr3+ content the conductivity σ decreases with increasing activation energy.  相似文献   

4.
X‐ray crystal structures are reported for Na6[RuO2{TeO4(OH)2}2]·16H2O and Na5[Ag{TeO4(OH)2}2]·16H2O which contain respectively RuVI and AgIII coordinated to chelating bidentate tellurate ([TeO4(OH)2]4−) groups. Na6[RuO2{TeO4(OH)2}2]·16H2O: Space group P1¯, Z = 2, lattice dimensions at 120 K; a = 6.9865(1), b = 8.7196(2), c = 11.7395(2)Å, α = 74.008(1), β = 79.954(1), γ = 88.514(1)°; R1 = 0.025. Na5[Ag{TeO4(OH)2}2]·16H2O: Space group P1¯, Z = 2, lattice dimensions at 120 K; a = 5.888(1), b = 8.932(1), c = 12.561(2)Å, α = 98.219(6), β = 97.964(9), γ = 93.238(14)°; R1 = 0.047.  相似文献   

5.
The crystal structures of dimagnesium disodium decavanadate icosahydrate, Mg2Na2V10O28·20H2O, (I), and trimagnesium decavanadate octacosahydrate, Mg3V10O28·28H2O, (II), have been determined by single‐crystal X‐ray diffraction. They crystallize with monoclinic (C2/c) and triclinic () symmetry, respectively. All the Mg2+ cations in (I) and (II) are octahedrally coordinated by six water mol­ecules. The Na+ cations in (I) are coordinated by three water mol­ecules and three O atoms of the decavanadate anions, and link the latter into a three‐dimensional network. The decavanadate anions in (II) are not linked to one another.  相似文献   

6.
This article investigated the low-energy structures of Al6Na mC (m = 2, 4, 6, 8) clusters and their electronic structures by using genetic algorithm combined with density functional theory and configuration interaction methods. The computations show that the C atoms prefer sitting at the center, whereas the Na atoms tend to locate at the outside of the clusters. The valence molecular orbitals (MOs) agree well with the prediction of the jellium model. The stronger attraction of the central carbon to the valence electrons will depress the potential energies locally, which makes the 2S level go obviously lower and the 2P and 1D orbitals form a sub-band. The 26 valence electrons in Al6Na4C form closed 1S21P62S21D102P6 shells and correspond to a new magic structure. The MOs and electron localization function show that the sodium cores are exposed at the outside of the valence electrons and form naked cations. The contraction of the valence electrons because of the carbon doping enhances the charges on the Al6C moieties, and the Na+ cores on the peripheries are ionically bonded to the Zintl anions (Al6C)q−. The Al6Na4C has a tetrahedral structure with symmetry Td, and it may be used as building blocks to synthesize Zintl solid.  相似文献   

7.
Polyoxopalladates (POPs) are a class of self-assembling palladium-oxide clusters that span a variety of sizes, shapes and compositions. The largest of this family, {Pd84}Ac, is constructed from 14 building units of {Pd6} and lined on the inner and outer torus by 28 acetate ligands. Due to its high water solubility, large hydrophobic cavity and distinct 1H NMR fingerprint {Pd84}Ac is an ideal molecule for exploring supramolecular behaviour with small organic molecules in aqueous media. Molecular visualisation studies highlighted potential binding sites between {Pd84}Ac and these species. Nuclear Magnetic Resonance (NMR) techniques, including 1H NMR, 1H Diffusion Ordered Spectroscopy (DOSY) and Nuclear Overhauser Spectroscopy (NOESY), were employed to study the supramolecular chemistry of this system. Here, we provide conclusive evidence that {Pd84}Ac forms a 1 : 7 host-guest complex with benzyl viologen (BV2+) in aqueous solution.  相似文献   

8.
As opposed to the reversible redox reaction ({Fe(NO)2}10 reduced‐form DNIC [(NO)2Fe(S(CH2)3S)]2? ( 1 )?{Fe(NO)2}9 oxidized‐form [(NO)2Fe(S(CH2)3S)]?), the chemical oxidation of the {Fe(NO)2}10 DNIC [(NO)2Fe(S(CH2)2S)]2? ( 2 ) generates the dinuclear {Fe(NO)2}9–{Fe(NO)2}9 complex [(NO)2Fe(μ‐SC2H4S)2Fe(NO)2]2? ( 3 ) bridged by two terminal [SC2H4S]2? ligands. On the basis of the Fe K‐edge pre‐edge energy and S K‐edge XAS, the oxidation of complex 1 yielding [(NO)2Fe(S(CH2)3S)]? is predominantly a metal‐based oxidation. The smaller S1‐Fe1‐S2 bond angle of 94.1(1)° observed in complex 1 (S1‐Fe1‐S2 88.6(1)° in complex 2 ), compared to the bigger bond angle of 100.9(1)° in the {Fe(NO)2}9 DNIC [(NO)2Fe(S(CH2)3S)]?, may be ascribed to the electron‐rich {Fe(NO)2}10 DNIC preferring a restricted bite angle to alleviate the electronic donation of the chelating thiolate to the electron‐rich {Fe(NO)2}10 core. The extended transition state and natural orbitals for chemical valence (ETS‐NOCV) analysis on the edt‐/pdt‐chelated {Fe(NO)2}9 and {Fe(NO)2}10 DNICs demonstrates how two key bonding interactions, that is, a Fe?S covalent σ bond and thiolate to the Fe d charge donation, between the chelating thiolate ligand and the {Fe(NO)2}9/10 core could be modulated by the backbone lengths of the chelating thiolate ligands to tune the electrochemical redox potential (E1/2=?1.64 V for complex 1 and E1/2=?1.33 V for complex 2 ) and to dictate structural rearrangement/chemical transformations (S‐Fe‐S bite angle and monomeric vs. dimeric DNICs).  相似文献   

9.
Na3[BN2] and Na2K[BN2] were obtained as white polycrystalline powders from the reaction of the respective binary mixtures NaNH2:NaBH4 and NaNH2:KBH4 in molar ratio 2:1 at 873 K and 683 K, respectively, in an argon stream. According to the results of thermal analysis measurements, both compounds are thermally stable only up to 954 K (Na3[BN2]) and 712 K (Na2K[BN2]), respectively, decomposing under evolution of alkali metal and nitrogen to yield hexagonal BN as final residue, which was identified from powder patterns. The crystal structure of Na3[BN2] {β‐Li3[BN2] type; P21/c (No. 14); Z = 4} was confirmed and the unit cell parameters redetermined: a = 5.724(1) Å, b = 7.944(1) Å, c = 7.893(1) Å, β = 111.31(1)°. According to X‐ray powder data, Na2K[BN2] crystallizes isotypic to Na2KCuO2 in the tetragonal space group I4/mmm (No. 139) with a = 4.2359(1) Å, c = 10.3014(2) Å and Z = 2. The crystal structure of Na2K[BN2] is composed of linear [N–B–N]3– anions centering elongated M14 rhombic dodecahedra, which are formed by 8 sodium and 6 potassium atoms. The [BN2]@Na8/4K6/6 polyhedra are stacked along [001] and condensed via common tetragonal faces to generate a space‐filling 3D arrangement. The B–N bond lengths for the strictly linear [N–B–N]3– units are 1.357(4) Å. Vibrational spectra of the title compounds were measured and analyzed based on D∞h symmetry of the relevant [N–B–N]3– groups taking into account the site symmetry effects for Na3[BN2]. Both the wavenumbers, as well as the calculated valence force constants f(B–N) = 7.29 N · cm–1 (Na3[BN2]) and 7.33 N · cm–1 (Na2K[BN2]), respectively, are in good agreement with those of the known alkali and alkaline earth dinitridoborates.  相似文献   

10.
To investigate the influence of P/As substitution on structures and electrical properties, e.g. the effect on material densities, two new solid P/As‐doped solutions, Na2CoP1.60As0.40O7 (disodium cobalt diphosphorus arsenic heptaoxide) and Na2CoP1.07As0.93O7 (disodium cobalt phosphorus arsenic heptaoxide), with melilite‐like structures have been synthesized by solid‐state reactions. Their unit‐cell parameters are in agreement with Vegard's law. The obtained structural models were investigated by the bond valence sum (BVS) and charge distribution (CHARDI) validation tools and, for the latter, the structures are described as being built on anion‐centred polyhedra. The frameworks can be described as layered and formed by {[Co(P,As)2O7]2−} slabs, with alkali cations sandwiched between the layers and with the interlayer spaces increased due to P/As substitution. The BVS model was extended to a preliminary simulation of the sodium conduction properties in the studied structural type and suggests that the most probable sodium conduction pathways are bidimensional, at the (002) planes.  相似文献   

11.
Highly conductive solid electrolytes are crucial to the development of efficient all‐solid‐state batteries. Meanwhile, the ion conductivities of lithium solid electrolytes match those of liquid electrolytes used in commercial Li+ ion batteries. However, concerns about the future availability and the price of lithium made Na+ ion conductors come into the spotlight in recent years. Here we present the superionic conductor Na11Sn2PS12, which possesses a room temperature Na+ conductivity close to 4 mS cm?1, thus the highest value known to date for sulfide‐based solids. Structure determination based on synchrotron X‐ray powder diffraction data proves the existence of Na+ vacancies. As confirmed by bond valence site energy calculations, the vacancies interconnect ion migration pathways in a 3D manner, hence enabling high Na+ conductivity. The results indicate that sodium electrolytes are about to equal the performance of their lithium counterparts.  相似文献   

12.
Reduction of neutral metal clusters (Co4(CO)12, Ru3(CO)12, Fe3(CO)12, Ir4(CO)12, Rh6(CO)16, {CpMo(CO)3}2, {Mn(CO)5}2) by decamethylchromocene (Cp*2Cr) or sodium fluorenone ketyl in the presence of cryptand[2.2.2] and DB‐18‐crown‐6 was studied. Nine new salts with paramagnetic Cp*2Cr+, cryptand[2.2.2](Na+), and DB‐18‐crown‐6(Na+) cations and [Co6(CO)15]2– ( 1 , 2 ), [Ru6(CO)18]2– ( 3 – 4 ) dianions, [Rh11(CO)23]3– ( 6 ) trianions, and new [Ir8(CO)18]2– ( 5 ) dianions were obtained and structurally characterized. The increase of nuclearity of clusters under reduction was shown. Fe3(CO)12 preserves the Fe3 core under reduction forming the [Fe3(CO)11]2– dianions in 7 . The [CpMo(CO)3]2 and [Mn(CO)5]2 dimers dissociate under reduction forming mononuclear [CpMo(CO)3] ( 8 ) and [Mn(CO)5] ( 9 ) anions. In all anions the increase of negative charge on metal atoms shifts the bands attributed to carbonyl C–O stretching vibrations to smaller wavenumbers in agreement with the elongation of the C–O bonds in 1 – 9 . In contrast, the M–C(CO) bonds are noticeably shortened at the reduction. Magnetic susceptibility of the salts with Cp*2Cr+ is defined by high spin Cp*2Cr+ (S = 3/2) species, whereas all obtained anionic metal clusters and mononuclear anions are diamagnetic. Rather weak magnetic coupling between S = 3/2 spins is observed with Weiss temperature from –1 to –11 K. That is explained by rather long distances between Cp*2Cr+ and the absence of effective π–π interaction between them except compound 7 showing the largest Weiss temperature of –11 K. The {DB‐18‐crown‐6(Na+)}2[Co6(CO)15]2– units in 2 are organized in infinite 1D chains through the coordination of carbonyl groups of the Co6 clusters to the Na+ ions and π–π stacking between benzo groups of the DB‐18‐crown‐6(Na+) cations.  相似文献   

13.
An anionic hexanuclear NiII metallamacrocycle with endo and exo linking sites has been employed as a building block to generate a series of capsules and bowls of nanometric size. The supramolecular arrangement of the {Ni6} rings was tailored by the size of the alkali cations, showing the transition from {Ni6-M2-Ni6} capsules (M=LiI and NaI) to {Ni6-M} bowls (M=KI and CsI). The alkyl co-cations are determinant to stabilize the assemblies by means of CH⋅⋅⋅π interactions on the exo side of the metallamacrocycles. The effect on the topology of the supramolecular assemblies of the cation size, cation charge, Et3NH+ or Me4N+ counter cations has been analyzed. Magnetic measurements reveal the presence of ferromagnetic and antiferromagnetic interactions inside the rings that allow a S=0 ground state.  相似文献   

14.
Investigations of the Synthesis of [CpxSb{M(CO)5}2] (Cpx = Cp, Cp*; M = Cr, W) The reaction of CpSbCl2 with [Na2{Cr2(CO)10}] leads to the chlorostibinidene complex [ClSb{Cr(CO)5}2(thf)] ( 1 ), whereas the reaction of CpSbCl2 with [Na2{W2(CO)10}] results in the formation of the complexes [ClSb{W(CO)5}3] ( 2 ), [Na(thf)][Cl2Sb{W(CO)5}2] ( 3 ), [ClSb{W(CO)5}2(thf)] ( 4 ) and [Sb2{W(CO)5}3] ( 5 ). The stibinidene complex [CpSb{Cr(CO)5}2] ( 6 ) is obtained by the reaction of [ClSb{Cr(CO)5}2] with NaCp, while its Cp* analogue [Cp*Sb{Cr(CO)5}2] ( 7 ) is formed via the metathesis of Cp*SbCl2 with [Na2{Cr2(CO)10}]. The products 2 , 3 , 4 and 7 are additionally characterised by X‐ray structure analyses.  相似文献   

15.
The crystal structures of Na2Mg3(OH)2(SO4)3 · 4H2O and K2Mg3(OH)2(SO4)3 · 2H2O, were determined from conventional laboratory X‐ray powder diffraction data. Synthesis and crystal growth were made by mixing alkali metal sulfate, magnesium sulfate hydrate, and magnesium oxide with small amounts of water followed by heating at 150 °C. The compounds crystallize in space group Cmc21 (No. 36) with lattice parameters of a = 19.7351(3), b = 7.2228(2), c = 10.0285(2) Å for the sodium and a = 17.9427(2), b = 7.5184(1), c = 9.7945(1) Å for the potassium sample. The crystal structure consists of a linked MgO6–SO4 layered network, where the space between the layers is filled with either potassium (K+) or Na+‐2H2O units. The potassium‐bearing structure is isostructural to K2Co3(OH)2(SO4)3 · 2(H2O). The sodium compound has a similar crystal structure, where the bigger potassium ion is replaced by sodium ions and twice as many water molecules. Geometry optimization of the hydrogen positions were made with an empirical energy code.  相似文献   

16.
A new phosphate, sodium calcium magnesium tetrakis(phosphate), Na8Ca1.5Mg12.5(PO4)12, has been synthesized by a flux method. Its novel structure consists of MgOx (x = 5 and 6) polyhedra and MO7 (M = Mg or Na) octahedra linked directly through common corners or edges to form a rigid three‐dimensional skeleton, reinforced by corner‐sharing between identical Mg12MO48 units. The connection of these units by the PO4 tetrahedra induces cavities and crossing tunnels where the Na+ and Ca2+ cations are located. This structural model was supported by a 31P NMR spectroscopy study which confirmed the existence of 12 crystallographically independent sites for the P atoms.  相似文献   

17.
The intrinsic kinetics, unaffected by diffusional and mass transfer effects, of the air oxidation of Na2TiF6 and Na3TiF6 were determined by using a nonisothermal technique. The oxidation of these sodium fluorotitanates proceeds through two-step reactions involving the formation of oxyfluorotitanate, i.e. Na3TiOF5, as the intermediate. The oxidation rate shows a first-order dependence on the amount of the unreacted solids for each of the two-step reactions for both fluorotitanates. The activation energy for the further oxidation of Na3TiOF5 to a mixture of NaF + TiO2 was determined to be 52.4 kJ/mol and 55.3 kJ/raol for Na2TiF6 and Na3TiF6 as reactants, respectively.  相似文献   

18.
The single crystal of sodium manganese arsenate (1.72/3.28/12), Na1.72Mn3.28(AsO4)3, used for analysis was prepared by solid‐state reaction at 1073 K. The compound crystallizes in the monoclinic system in space group C2/c. The structure consists of a complex network of edge‐sharing MnO6 octahedral chains, linked together by AsO4 tetrahedra, forming two distinct channels, one containing Na+ cations and the other occupied statistically by Mn+ and Na+ cations.  相似文献   

19.
Syntheses and Crystal Structures of Ternary Carbides Na2PdC2 and Na2PtC2 Na2PdC2 and Na2PtC2 were synthesized by the reaction of sodium carbide with palladium and platinum respectively. The crystal structures could be solved from X-ray powder diffraction data (space group: P3 m1, Z = 1). Both compounds crystallize in a new structure type with [M(C2)2/22?] chains (M?Pd, Pt) as the characteristic structural unit. The existence of a C? C triple bond was confirmed by Raman spectroscopy.  相似文献   

20.
The structure of synthetic disodium magnesium disulfate decahydrate at 180 K consists of alternating layers of water‐coordinated [Mg(H2O)6]2+ octahedra and [Na2(SO4)2(H2O)4]2− sheets, parallel to [100]. The [Mg(H2O)6]2+ octahedra are joined to one another by a single hydrogen bond, the other hydrogen bonds being involved in inter‐layer linkage. The Mg2+ cation occupies a crystallographic inversion centre. The sodium–sulfate sheets consist of chains of water‐sharing [Na(H2O)6]+ octahedra along b, which are then connected by sulfate tetrahedra through corner‐sharing. The associated hydrogen bonds are the result of water–sulfate interactions within the sheets themselves. This is believed to be the first structure of a mixed monovalent/divalent cation sulfate decahydrate salt.  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