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.
Incorporation of monatomic 2p ligands into the core of iron–sulfur clusters has been researched since the discovery of interstitial carbide in the FeMo cofactor of Mo‐dependent nitrogenase, but has proven to be a synthetic challenge. Herein, two distinct synthetic pathways are rationalized to install nitride ligands into targeted positions of W‐Fe‐S clusters, generating unprecedented nitride‐ligated iron–sulfur clusters, namely [(Tp*)2W2Fe64‐N)2S6L4]2? (Tp*=tris(3,5‐dimethyl‐1‐pyrazolyl)hydroborate(1?), L=Cl? or Br?). 57Fe Mössbauer study discloses metal oxidation states of WIV2FeII4FeIII2 with localized electron distribution, which is analogous to the mid‐valent iron centres of FeMo cofactor at resting state. Good agreement of Mössbauer data with the empirical linear relationship for Fe–S clusters indicates similar ligand behaviour of nitride and sulfide in such clusters, providing useful reference for reduced nitrogen in a nitrogenase‐like environment.  相似文献   

2.
Three FeIII2LnIII2 tetranuclear heterometallic clusters, [H4LGd(H2O)Tp*Fe(CN)3]2·8H2O·2MeOH (1) and [H4LLn(MeOH)Tp*Fe(CN)3]2·6MeOH·2MeCN (Ln?=?Tb and Dy for 2 and 3, respectively, H6L = N,N′-(2,6-pyridine-dicarboxyl)-disalicylhydrazide, Tp* = hydridotris(3,5dimethylpyrazol-1-yl)-borate), were synthesized by use of the [(Tp*)Fe(CN)3]? unit as a metalloligand toward LnCl3 and H6L species. Structural analyses reveal that FeIII and LnIII ions in all complexes are connected to each other by one cyanide to form a heterobinuclear unit of [Ln(H4L)][(Tp*)Fe(CN)3], which is dimerized through Ln–N–C?=?O–Ln interaction. Magnetic susceptibility measurements show weak antiferromagnetic interactions between cyano-bridged FeIII and GdIII ions and amide-bridged GdIII ions are operative. Complex 1 displays the magnetocaloric effect with ?ΔSmmax = 12.70 J·kg?1·K?1 at 4.0 K for ΔH?=?7 T. No single-molecule magnetic properties are observed for 2 or 3 down to 1.8 K.  相似文献   

3.
The reaction of N2 with trinuclear niobium and tungsten sulfide clusters Nb3Sn and W3Sn (n=0–3) was systematically studied by density functional theory calculations with TPSS functional and Def2-TZVP basis sets. Dissociations of N−N bonds on these clusters are all thermodynamically allowed but with different reactivity in kinetics. The reactivity of Nb3Sn is generally higher than that of W3Sn. In the favorite reaction pathways, the adsorbed N2 changes the adsorption sites from one metal atom to the bridge site of two metal atoms, then on the hollow site of three metal atoms, and at that place, the N−N bond dissociates. As the number of ligand S atoms increases, the reactivity of Nb3Sn decreases because of the hindering effect of S atoms, while W3S and W3S2 have the highest reactivity among four W3Sn clusters. The Mayer bond order, bond length, vibrational frequency, and electronic charges of the adsorbed N2 are analyzed along the reaction pathways to show the activation process of the N−N bond in reactions. The charge transfer from the clusters to the N2 antibonding orbitals plays an essential role in N−N bond activation, which is more significant in Nb3Sn than in W3Sn, leading to the higher reactivity of Nb3Sn. The reaction mechanisms found in this work may provide important theoretical guidance for the further rational design of related catalytic systems for nitrogen reduction reactions (NRR).  相似文献   

4.
Reactions of (Et4N)[Tp*WS3] [Tp* is hydridotris(3,5‐dimethylpyrazol‐1‐yl)borate] with CuSCN in MeCN in the presence of melamine afforded the title neutral dimeric cluster [Cu4W2(C15H22BN6)2(NCS)2S6(C2H3N)2] or [Tp*W(μ2‐S)23‐S)Cu(μ2‐SCN)(CuMeCN)]2, which has two butterfly‐shaped [Tp*WS3Cu2] cores bridged across a centre of inversion by two (CuSCN) anions. The S atoms of the bridging thiocyanate ligands interact with the H atoms of the methyl groups of the Tp* units of a neighbouring dimer to form a C—H...S hydrogen‐bonded chain. The N atoms of the thiocyanate anions interact with the H atoms of the methyl groups of the Tp* units of neighbouring chains, affording a two‐dimensional hydrogen‐bonded network.  相似文献   

5.
The reactivity of the (PPN)2[Fe8S6(NO)8] and (PPN)2[Fe6S6(NO)6] clusters is explored and new derivative clusters have been synthesized and structurally characterized. The unique (PPN)2Fe4S4(NO)6 “open-cubane” cluster with a chair like Fe4S4 core is obtained along with the mixed metal pentandite-like clusters (PPN)2[Mo2Fe6S6(NO)6(CO)6], (PPr3)2Cu2Fe6S6(NO)6, (PPr3)4Cu4Fe4S6(NO)4, (PPr3)2Ni2Fe6S6(NO)6, (PPr3)3Ni3Fe4S6(NO)4. The rich electrochemistry of the mixed metal clusters is presented as well.  相似文献   

6.
The “CPNR” ligand may be viewed as being isolobal with fulminate, CNO; however, attempts to prepare a complex of such a ligand resulted instead in a range of novel imino and aminophosphinocarbyne complexes. Sequential treatment of [Mo(≡CBr)(CO)2(Tp*)] (Tp*=hydrotris(dimethylpyrazolyl)borate) with nBuLi and ClP=NMes* (Mes*=C6H2tBu3-2,4,6) afforded mixtures of the complexes [Mo(≡CPnBuNHMes*)(CO)2(Tp*)] and traces of the bimetallic products [Mo22-C2P2O(NHMes)2}(CO)4(Tp*)2] and [Mo22-C2PNHMes)(CO)4(Tp*)2]. The reaction of [W(≡CBr)(CO)2(Tp*)] with nBuLi and ClP=NMes* afforded predominantly the mononuclear carbyne [W{≡CP(=NMes*)nBu2})(CO)2(Tp*)] and traces of the binuclear complex [W2(μ-C2PNHMes)(CO)4(Tp*)2] which is also obtained when tBuLi is used. Although not isolable, the intended complexes [M(≡CPNMes*)(CO)2(Tp*)] could be generated in situ and spectroscopically characterized via the reactions of the stannyl carbynes [M(≡CSnnBu3)(CO)2(Tp*)] and ClP=NMes*. The preceding observations are mechanistically interpreted with reference to a computational interrogation of the model complex [Mo(≡CP=NCH3)(CO)2(Tp*)], the LUMO of which has considerable phosphorus character.  相似文献   

7.
    
Four new chalcogenide molybdenum and tungsten cubane clusters (NH4)6[M4Q4(CN)12]·6H2O (M=Mo or W; Q=S or Se) were prepared by high-temperature reactions of the triangular M3O7Br4 complexes with KCN at 430 °C followed by crystallization from aqueous solutions of ammonium acetate. The molecular and crystal structures of (NH4)6[Mo4S4(CN)12]·6H2O, (NH4)6[W4S4(CN)12]·6H2O, and (NH4)6[W4Se4(CN)12]·6H2O were established by X-ray diffraction analysis. The mixed-valence cubane clusters are diamagnetic and isostructural and have the symmetryT d . The clusters were characterized by IR and electronic spectroscopy. The data of cyclic voltammetry demonstrated that the [M4Q4(CN)12] n clusters exist in three oxidation states from the most oxidized (n=6; 10 cluster electrons) to the most reduced electron-precise 12-electron species (n=8). Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 18–24, January, 2000.  相似文献   

8.
Four new chalcogenide molybdenum and tungsten cubane clusters (NH4)6[M4Q4(CN)12]·6H2O (M=Mo or W; Q=S or Se) were prepared by high-temperature reactions of the triangular M3O7Br4 complexes with KCN at 430 °C followed by crystallization from aqueous solutions of ammonium acetate. The molecular and crystal structures of (NH4)6[Mo4S4(CN)12]·6H2O, (NH4)6[W4S4(CN)12]·6H2O, and (NH4)6[W4Se4(CN)12]·6H2O were established by X-ray diffraction analysis. The mixed-valence cubane clusters are diamagnetic and isostructural and have the symmetryT d . The clusters were characterized by IR and electronic spectroscopy. The data of cyclic voltammetry demonstrated that the [M4Q4(CN)12] n clusters exist in three oxidation states from the most oxidized (n=6; 10 cluster electrons) to the most reduced electron-precise 12-electron species (n=8). Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 18–24, January, 2000.  相似文献   

9.
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   

10.
Alkoxide and carbonyl ligands complement each other because they both behave as “π buffers” to transition metals. Alkoxides, which are π donors, stabilize early transition metals in high oxidation states by donating electrons into vacant dπ orbitals, whereas carbonyls, which are π acceptors, stabilize later transition elements in their lower oxidation states by accepting electrons from filled dπ orbitals. Both ligands readily form bridges that span M? M bonds. In solution fluxional processes that involve bridge–terminal ligand exchange are common to both alkoxide and carbonyl ligands. The fragments [W(OR)3], [CpW(CO)2], [Co(CO)3], and CH are related by the isolobal analogy. Thus the compounds [(RO)3W ? W(OR)3], [Cp(CO)2W?W(CO)2Cp], hypothetical [(CO)3Co?Co(CO)3], and HC?CH are isolobal. Alkoxide and carbonyl cluster compounds often exhibit striking similarities with respect to substrate binding—e.g., [W33-CR)(OR′)9] versus [Co33-CR)(CO)9] and [W4(C)(NMe)(OiPr)12] versus [Fe4(C)(CO)13]—but differ with respect to M? M bonding. The carbonyl clusters use eg-type orbitals for M? M bonding whereas the alkoxide clusters employ t2g-type orbitals. Another point of difference involves electronic saturation. In general, each metal atom in a metal carbonyl cluster has an 18-electron count; thus, activation of the cluster often requires thermal or photochemical CO expulsion or M? M bond homolysis. Alkoxide clusters, on the other hand, behave as electronically unsaturated species because the π electrons are ligand-centered and the LUMO metal-centered. Also, access to the metal centers may be sterically controlled in metal alkoxide clusters by choice of alkoxide groups whereas ancillary ligands such as tertiary phosphanes or cyclopentadienes must be introduced if steric factors are to be modified in carbonyl clusters. A comparison of the reactivity of alkynes and ethylene with dinuclear alkoxide and carbonyl compounds is presented. For the carbonyl compounds CO ligand loss is a prerequisite for substrate uptake and subsequent activation. For [M2(OR)6] compounds (M = Mo and W) the nature of substrate uptake and activation is dependent upon the choice of M and R, leading to a more diverse chemistry.  相似文献   

11.
Nitrogenases catalyze the reduction of N2 to NH4+ at its cofactor site. Designated the M‐cluster, this [MoFe7S9C(R‐homocitrate)] cofactor is synthesized via the transformation of a [Fe4S4] cluster pair into an [Fe8S9C] precursor (designated the L‐cluster) prior to insertion of Mo and homocitrate. We report the characterization of an eight‐iron cofactor precursor (designated the L*‐cluster), which is proposed to have the composition [Fe8S8C] and lack the “9th sulfur” in the belt region of the L‐cluster. Our X‐ray absorption and electron spin echo envelope modulation (ESEEM) analyses strongly suggest that the L*‐cluster represents a structural homologue to the l ‐cluster except for the missing belt sulfur. The absence of a belt sulfur from the L*‐cluster may prove beneficial for labeling the catalytically important belt region, which could in turn facilitate investigations into the reaction mechanism of nitrogenases.  相似文献   

12.
Single‐electron oxidation of a diiron‐sulfur complex [Cp*Fe(μ‐bdt)FeCp*] ( 1 , Cp*=η5‐C5Me5; bdt=benzene‐1,2‐dithiolate) to [Cp*Fe(μ‐bdt)FeCp*]+ ( 2 ) has been experimentally conducted. The bdt ligand with redox‐active character has been computationally proposed to be a dianion (bdt2?) rather than previously proposed monoanion (bdt·?) radical in 1 though it has un‐equidistant aromatic C? C bond lengths. The ground state of 1 is predicted to be two low‐spin ferrous ions (SFe=0) and 2 has a medium‐spin ferric ion (SFe=1/2) and a low‐spin ferrous center (SFe=0), and the oxidation of 1 to 2 is calculated to be a single‐metal‐based process. Both complexes have no significant antiferromagnetic coupling character.  相似文献   

13.
Atomically precise molecular metal-oxo clusters provide ideal models to understand metal oxide surfaces, self-assembly, and form-function relationships. Devising strategies for synthesis and isolation of these molecular forms remains a challenge. Here, the synthesis of four Ln-Fe oxo clusters that feature the ϵ-{Fe13} Keggin cluster in their core is reported. The {Fe13} metal-oxo cluster motif is the building block of two important iron oxyhydroxyide phases in nature and technology, ferrihydrite (as the δ-isomer) and magnetite (the ϵ-isomer). The reported ϵ-{Fe13} Keggin isomer as an isolated molecule provides the opportunity to study the formation of ferrihydrite and magnetite from this building unit. The four currently reported isostructural lanthanide-iron-oxo clusters are fully formulated [Y12Fe33(TEOA)12(Hyp)63-OH)204-O)28(H2O)12](ClO4)23 ⋅ 50 H2O ( 1 , Y12Fe33 ), [Gd12Fe33(TEOA)12(Hyp)63-OH)204-O)32(H2O)12](ClO4)15 ⋅ 50 H2O ( 2 , Gd12Fe33 ) and [Ln16Fe29(TEOA)12(Hyp)63-OH)244-O)28(H2O)16](ClO4)16(NO3)3n H2O (Ln=Y for 3 , Y16Fe29 , n=37 and Ln=Gd for 4 , Gd16Fe29 n=25; Hyp=trans-4-Hydroxyl-l -proline and TEOA=triethanolamine). The next metal layer surrounding the ϵ-{Fe13} core within these clusters exhibits a similar arrangement as the magnetite lattice, and Fe and Ln can occupy the same positions. This provides the opportunity to construct a family of compounds and optimize magnetic exchange in these molecules through composition tuning. Small-angle X-ray scattering (SAXS) and high-resolution electrospray ionization mass spectrometry (HRESI-MS) show that these clusters are stable upon dissolution in both water and organic solvents, as a first step to performing further chemistry towards building magnetic arrays or investigating ferrihydrite and magnetite assembly from pre-nucleation clusters.  相似文献   

14.
In this work, a new nanocatalyst, Fe2W18Fe4@NiO@CTS, was synthesized by the reaction of sandwich‐type polyoxometalate (Fe2W18Fe4), nickel oxide (NiO), and chitosan (CTS) via sol–gel method. The assembled nanocatalyst was systematically characterized by FT‐IR, UV–vis, XRD, SEM, and EDX analysis. The catalytic activity of Fe2W18Fe4@NiO@CTS was tested on oxidative desulfurization (ODS) of real gasoline and model fuels. The experimental results revealed that the levels of sulfur content and mercaptan compounds of gasoline were lowered with 97% efficiency. Also, the Fe2W18Fe4@NiO@CTS nanocatalyst demonstrated an outstanding catalytic performance for the oxidation of dibenzothiophene (DBT) in the model fuel. The major factors that influence the desulfurization efficiency and the kinetic study of the ODS reactions were fully detailed and discussed. The probable ODS pathway was proposed via the electrophilic mechanism on the basis of the electrophilic characteristic of the metal‐oxo‐peroxo intermediates. The prepared nanocatalyst could be reused for 5 successive runs without any appreciable loss in its catalytic activity. As a result, the current study suggested the potential application of the Fe2W18Fe4@NiO@CTS hybrid nanocatalyst as an ideal candidate for removal of sulfur compounds from fuel.  相似文献   

15.
The structures and energies of the electronic ground states of the FeS0/?, FeS20/?, Fe2S20/?, Fe3S40/?, and Fe4S40/? neutral and anionic clusters have been computed systematically with nine computational methods in combination with seven basis sets. The computed adiabatic electronic affinities (AEA) have been compared with available experimental data. Most reasonable agreements between theory and experiment have been found for both hybrid B3LYP and B3PW91 functionals in conjugation with 6‐311+G* and QZVP basis sets. Detailed comparisons between the available experimental and computed AEA data at the B3LYP/6‐311+G* level identified the electronic ground state of 5Δ for FeS, 4Δ for FeS?, 5B2 for FeS2, 6A1 for FeS2?, 1A1 for Fe2S2, 8A′ for Fe2S2?, 5A′′ for Fe3S4, 6A′′ for Fe3S4?, 1A1 for Fe4S4, and 1A2 for Fe4S4?. In addition, Fe2S2, Fe3S4, Fe3S4?, Fe4S4, and Fe4S4? are antiferromagnetic at the B3LYP/6‐311+G* level. The magnetic properties are discussed on the basis of natural bond orbital analysis.  相似文献   

16.
Atomically precise Cu‐rich bimetallic superatom clusters have been synthesized by adopting a galvanic exchange strategy. [Cu@Cu12(S2CNnBu2)6(C≡CPh)4][CuCl2] ( 1 ) was used as a template to generate compositionally uniform clusters [M@Cu12(S2CNnBu2)6 (C≡CPh)4][CuCl2], where M=Ag ( 2 ), Au ( 3 ). Structures of 1 , 2 and 3 were determined by single crystal X‐ray diffraction and the results were supported by ESI‐MS. The anatomies of clusters 1 – 3 are very similar, with a centred cuboctahedral cationic core that is surrounded by six di‐butyldithiocarbamate (dtc) and four phenylacetylide ligands. The doped Ag and Au atoms were found to preferentially occupy the centre of the 13‐atom cuboctahedral core. Experimental and theoretical analyses of the synthesized clusters revealed that both Ag and Au doping result in significant changes in cluster stability, optical characteristics and enhancement in luminescence properties.  相似文献   

17.
By the additional reaction of binuclear compounds (Me4N)2M2S4 (TDT)2 (M = Mo, W; H2TDT= H2CH3C6H3S2) with mononuclear cuprous complex, two new M-Cu-S clusters Mo2Cu2S4(TDT)2-(PPh3 )2·CH3CH2OH (1) and W2Cu2S4(TDT)2(PPh3)2·0.5CH3CH2OH-0.5H2O (2) have been prepared and characterized by IR, UV-Vis, 31P NMR spectroscopy, cyclic voltammetry and single crystal X-ray structure analysis . Both compounds crystallize in space group P 21 / n with lattice parameters a = 1.0956(3), b = 2.2072(3), c = 2.4340(3) nm, β= 100.36(2)°, V= 5.790(3) nm3 and Z = 4 for 1 and a = 1.0965(9),b= 2.2135(3), c = 2.4317(4) nm, β = 99.63(8)°, V= 5.819(8) nm3 and Z = 4 for 2. Both molecular structures contain a cubane-like cluster core [M2Cu2S4] (M = Mo, W) and their skeletons are almost the same except for M atoms. The syntheses, structures and spectrum characterizations of these two clusters are discussed. The third-order nonlinear optical (NLO) property of the two clusters was studied by the technique of forward degenerate four  相似文献   

18.
Reactions of a hexadentate ligand N,N,N’,N’-tetrakis(2-hydroxyethyl)ethylenediamine (H4edte) with different iron(III) salts in different solvents yielded three new twisted-saddle Fe12 clusters with adamantane-like [Fe4O6] inner core. Preliminary magnetic studies show that strong intracluster anti-ferromagnetic interaction exists in both 1 and 3, generating the S T = 0 spin ground state.  相似文献   

19.
Multimetallic clusters have long been investigated as molecular surrogates for reactive sites on metal surfaces. In the case of the μ4‐nitrido cluster [Fe44‐N)(CO)12]?, this analogy is limited owing to the electron‐withdrawing effect of carbonyl ligands on the iron nitride core. Described here is the synthesis and reactivity of [Fe44‐N)(CO)8(CNArMes2)4]?, an electron‐rich analogue of [Fe44‐N)(CO)12]?, where the interstitial nitride displays significant nucleophilicity. This characteristic enables rational expansion with main‐group and transition‐metal centers to yield unsaturated sites. The resulting clusters display surface‐like reactivity through coordination‐sphere‐dependent atom rearrangement and metal–metal cooperativity.  相似文献   

20.
We have designed new trithiols Temp(SH)3 and Tefp(SH)3 that can be synthesized conveniently in short steps and are useful for preparation of crystalline [3:1] site-differentiated [4Fe-4S] clusters suitable for X-ray structural analysis. The ethanethiolate clusters (PPh4)2[Fe4S4(SEt)(TempS3)] ( 4a ) and (PPh4)2[Fe4S4(SEt)(TefpS3)] ( 4b ) were prepared as precursors, and the unique iron sites were then selectively substituted. Upon reaction with H2S, (PPh4)2[Fe4S4(SH)(TempS3)] ( 6a ) and (PPh4)2[Fe4S4(SH)(TefpS3)] ( 6b ), which model the [4Fe-4S] cluster in the β subunit of (R)-2-hydroxyisocaproyl-CoA dehydratase, were synthesized. Clusters 6a and 6b were further converted to the sulfido-bridged double cubanes (PPh4)4[{Fe4S4(TempS3)}22-S)] ( 6b ) and (PPh4)4[{Fe4S4(TefpS3)}22-S)] ( 7b ), respectively, via intermolecular condensation with the release of H2S. Conversely, addition of H2S to 7a , 7b afforded the hydrosulfide clusters 6a , 6b . The molecular structures of the clusters reported herein were elucidated by X-ray crystallographic analysis. Their redox properties were investigated by cyclic voltammetry.  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