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 477 毫秒
1.
Reduction of the bis-pyrazolyl pyridine complex [CrL]2 with 4 KC8, followed by addition of one azobenzene (overall mole ratio 1:4:1), PhNNPh, transfers reducing equivalents to three azobenzenes, to form [K3Cr(PhNNPh)3]. This has three κ2 PhNNPh2− ligands and K+ bound to nitrogen atoms of azobenzene. When the stoichiometry is modified to 1:4:3, the product is changed to [K2CrL(PhNNPh)2], which has C2 symmetry except for the intimate ion pairing of two K+ ions to reduced azobenzene nitrogen atoms, and to pyrazolate and phenyl rings. The origin of the observed delivery of reducing equivalents to several, not to a single N=N bond, is traced to the resistance of the one-electron-reduced substrate to receiving a second electron, and is thus a general phenomenon. [CrL]2 alone is shown to be a two-electron reductant towards benzo[c]cinnoline (BCC) resulting in a product of formula [Cr2L2(BCC)], in which the reducing equivalents originate purely from CrII. An analogous study of the reaction of [CrL]2 with azobenzene yields [Cr2L2(PhNNPh)(THF)], an adduct in which one THF has displaced one of four hydrazide nitrogen/Cr bonds. Together these illustrate different modes for the Cr2L2 unit to bind and reduce the N=N bond. Collectively, these results show that two divalent Cr, without added K0, have the ability to reduce the N=N bond. Further KC8 reduction of preformed Cr2L2(RNNR) inevitably gives products in which K+ stabilizes the charge in the increasingly electron-rich nitrogen atoms, in a phenomenon which mimics proton coupled electron transfer: K+ performs the role of H+. A least-squares fit of the two singly reduced DFT structures shows that the only major change is a re-orientation of one of the two phenyl rings in order to avoid repulsion with potassium but to still allow interaction of that phenyl π system with K+. This shows both the impact of K+, being modest to nitrogen/chromium interactions, but nevertheless accommodating some π donation of phenyl to potassium. Finally, delivering increasing equivalents of KC8 leads to complete cleavage of the N=N bond, and both N bind to three CrII. The varied impacts of the K+ electrophile on NN multiple bond reduction is discussed.  相似文献   

2.
The reaction of UO2(OAc)2 ⋅ 2H2O with the biologically inspired ligand 2-salicylidene glucosamine (H2 L1 ) results in the formation of the anionic trinuclear uranyl complex [(UO2)3(μ3-O)( L1 )3]2− ( 1 2−), which was isolated in good yield as its Cs-salt, [Cs]2 1 . Recrystallization of [Cs]2 1 in the presence of 18-crown-6 led to formation of a neutral ion pair of type [M(18-crown-6)]2 1 , which was also obtained for the alkali metal ions Rb+ and K+ (M=Cs, Rb, K). The related ligand, 2-(2-hydroxy-1-naphthylidene) glucosamine (H2 L2 ) in a similar procedure with Cs+ gave the corresponding complex [Cs(18-crown-6)]2[(UO2)3(μ3-O)( L2 )3 ([Cs(18-crown-6)]2 2 ). From X-ray investigations, the [(UO2)3O( Ln )3]2− anion (n=1, 2) in each complex is a discrete trinuclear uranyl species that coordinates to the alkali metal ion via three uranyl oxygen atoms. The coordination behavior of H2 L1 and H2 L2 towards UO22+ was investigated by NMR, UV/Vis spectroscopy and mass spectrometry, revealing the in situ formation of the 1 2− and 2 2−dianions in solution.  相似文献   

3.
Complexes of Chromium Containing 1,5,9-Triazacyclododecane: Synthesis, Magnetism, and Crystal Structure of Tri-μ-hydroxo-bis[(1,5,9-triazacyclododecane) chromium (III)] tribromide · Dihydrate; Kinetic and Mechanism of its Bridge-cleavage with Hydroxide The oxidative decarbonylation of LCr(CO)3 (L = 1,5,9-triazacyclododecane) with bromine yields green LCrBr3. Base hydrolysis affords red [LCr(μ-OH)3CrL]3+, whereas in the presence of acetate ions [L2Cr2(μ-OH)2(CH3CO2)]3+ is formed. [LCr(μ-OH)3CrL]Br3· 2 H2O crystallizes in the orthorhombic space group P212121 with 8 formula units per unit cell. Two CrIII centers are connected via three OH? bridges; the spins of d3-electronic configuration are coupled intramolecularly, antiferromagnetically (2J = ?96 cm?1). With excess OH? but not with protons the tri-μ-hydroxo species is cleaved to give [L2Cr2(OH)2(μ-OH)2]2+. The kinetics of this reaction have been measured using the stopped flow technique. The mechanism is discussed.  相似文献   

4.
Digallane [L1Ga−GaL1] ( 1 , L1=dpp-bian=1,2-[(2,6-iPr2C6H3)NC]2C12H6) reacts with RN=C=O (R=Ph or Tos) by [2+4] cycloaddition of the isocyanate C=N bonds across both of its C=C−N−Ga fragments to afford [L1(O=C−NR)Ga−Ga(RN−C=O)L1] (R=Ph, 3 ; R=Tos, 4 ). The reactions with both isocyanates result in new C−C and N−Ga single bonds. In the case of allyl isocyanate, the [2+4] cycloaddition across one C=C−N−Ga fragment of 1 is accompanied by insertion of a second allyl isocyanate molecule into the Ga−N bond of the same fragment to afford compound [L1Ga−Ga(AllN− C=O)2L1] ( 5 ) (All=allyl). In the presence of Na metal, the related digallane [L2Ga−GaL2] ( 2 ; L2=dpp-dad=[(2,6-iPr2C6H3)NC(CH3)]2) is converted into the gallium(I) carbene analogue [L2Ga:] ( 2 A ), which undergoes a variety of reactions with isocyanate substrates. These include the cycloaddition of ethyl isocyanate to 2 A affording [Na2(THF)5]{L2Ga[EtN−C(O)]2GaL2} ( 6 ), cleavage of the N=C bond with release of 1 equiv. of CO to give [Na(THF)2]2[L2Ga(p-MeC6H4)(N−C(O))2−N(p-MeC6H4)]2 ( 7 ), cleavage of the C=O bond to yield the di-O-bridged digallium compound [Na(THF)3]2[L2Ga-(μ-O)2-GaL2] ( 8 ), and generation of the further addition product [Na2(THF)5][L2Ga(CyNCO2)]2 ( 9 ). Complexes 3 – 9 have been characterized by NMR (1H, 13C), IR spectroscopy, elemental analysis, and X-ray diffraction analysis. Their electronic structures have been examined by DFT calculations.  相似文献   

5.
Reduction of the bis-(pyrazolyl)pyridine complex [LCr]2 with stoichiometric KC8 in THF produces a species that is reactive with CO2 to produce an aggregate composed of paramagnetic K2L2Cr2(CO3) linked by KCl into a product of formula [K2L2Cr2(CO3)]4⋅2KCl. X-ray diffraction reveals a pincer hydrocarbon exterior and an inorganic interior composed of K+, Cl and carbonate oxygens. Every Cr is five coordinate and square pyramidal, with the axial N donor weakly bonded to Cr due to the Jahn–Teller effect of a high spin d4 configuration. Reaction with 13CO2 confirms that carbonate here is derived from CO2, that oxide is derived from CO2, and that CO is indeed released, since it is not a competent ligand to CrII. Guiding principles for selectivity in CO2 reduction are deduced from the diverse successful molecular constructs to date.  相似文献   

6.
Reaction of [Pt(DMSO)2Cl2] or [Pd(MeCN)2Cl2] with the electron-rich LH=N,N’-bis(4-dimethylaminophenyl)ethanimidamide yielded mononuclear [PtL2] ( 1 ) but dinuclear [Pd2L4] ( 2 ), a paddle-wheel complex. The neutral compounds were characterized through experiments (crystal structures, electrochemistry, UV-vis-NIR spectroscopy, magnetic resonance) and TD-DFT calculations as metal(II) species with noninnocent ligands L. The reversibly accessible cations [PtL2]+ and [Pd2L4]+ were also studied, the latter as [Pd2L4][B{3,5-(CF3)2C6H3}4] single crystals. Experimental and computational investigations were directed at the elucidation of the electronic structures, establishing the correct oxidation states within the alternatives [PtII(L)2] or [Pt.(L )2], [PtII(L0.5−)2]+ or [PtIII(L)2]+, [(PdII)2(μ-L)4] or [(Pd1.5)2(μ-L0.75−)4], and [(Pd2.5)2(μ-L)4]+ or [(PdII)2(μ-L0.75−)4]+. In each case, the first alternative was shown to be most appropriate. Remarkable results include the preference of platinum for mononuclear planar [PtL2] with an N-Pt-N bite angle of 62.8(2)° in contrast to [Pd2L4], and the dimetal (Pd24+→Pd25+) instead of ligand (L→L ) oxidation of the dinuclear palladium compound.  相似文献   

7.
Treatment of CrCl3(THF)3 with KPzTp in THF affords of the compound K[Cr(PzTp)Cl3], and the K+ in this complex can be replaced by Et4N+ in CH2Cl2. Well-defined green crystals of [Et4N]r(PzTp)Cl3] (I) suitable for X-ray diffraction are obtained at −20°C. In the anion the metal center shows a distorted octahedral geometry with the tetra(pyrazolyl) borate bonded as three N-donor tripod ligands and three chloride atoms completing the coordination sphere.  相似文献   

8.
The ability of the tetraaza‐dithiophenolate ligand H2L2 (H2L2 = N,N′‐Bis‐[2‐thio‐3‐aminomethyl‐5‐tert‐butyl‐benzyl]propane‐1,3‐diamine) to form dinuclear chromium(III) complexes has been examined. Reaction of CrIICl2 with H2L2 in methanol in the presence of base followed by air‐oxidation afforded cis,cis‐[(L2)CrIII2(μ‐OH)(Cl)2]+ ( 1a ) and trans,trans‐[(L2)CrIII2(μ‐OH)(Cl)2]+ ( 1b ). Both compounds contain a confacial bioctahedral N2ClCrIII(μ‐SR)2(μ‐OH)CrIIIClN2 core. The isomers differ in the mutual orientation of the coligands and the conformation of the supporting ligand. In 1a both Cl? ligands are cis to the bridging OH function. In 1b they are in trans‐positions. Reaction of the hydroxo‐bridged complexes with HCl yielded the chloro‐bridged cations cis,cis‐[(L2)CrIII2(μ‐Cl)(Cl)2]+ ( 2a ) and trans,trans‐[(L2)CrIII2(μ‐Cl)(Cl)2]Cl ( 2b ), respectively. These bridge substitutions proceed with retention of the structures of the parent complexes 1a and 1b .  相似文献   

9.
The first example of a diphosphaborolediide, the benzo-fused [C6H4P2BPh]2− ( 12− ), is prepared from ortho-bis(phosphino)benzene (C6H4{PH2}) and dichlorophenylborane, via a sequential lithiation approach. The dilithio-salt can be obtained as an oligomeric THF solvate or discrete TMEDA adduct, both of which are fully characterized, including by X-ray diffraction. Alongside NICS calculations, data strongly suggest some aromaticity within 12− , which is further supported by preliminary coordination studies that demonstrate η5-coordination to a zerovalent molybdenum center, as observed crystallographically for the oligomeric [{Mo(CO)35- 1 )}{μ-η1-Mo(CO)3(TMEDA)}2] ⋅ [μ-Li(THF)][μ-Li(TMEDA)].  相似文献   

10.
The reactivity of aryl monocarboxylic acids (benzoic, 1- or 2-naphtoic, 4’-methylbiphenyl-4-carboxylic, and anthracene-9-carboxylic acids) as complexing agents for the ethoxide niobium(V) (Nb(OEt)5 precursor has been investigated. A total of eight coordination complexes were isolated with distinct niobium(V) nuclearities as well as carboxylate complexation states. The use of benzoic acid gives a tetranuclear core Nb42-O)4(L)4(OEt)8] (L=benzoate ( 1 )) with four Nb−(μ2-O)−Nb linkages in a square plane configuration. A similar tetramer, 7 , was obtained with 2-naphtoic acid by using a 55 % humid atmosphere synthetic route. Two types of dinuclear brick were identified with one central Nb−(μ2-O)−Nb linkage; they differ in their complexation state, with one bridging carboxylate ([Nb22-O)(μ2-OEt)(L)(OEt)6], with L=1-naphtoate ( 3 ) or anthracene-9-carboxylate ( 5 )) or two bridging carboxylate groups ([Nb22-O)(L)2(OEt)6], with L=4’-methylbiphenyl-4-carboxylic ( 4 ) or anthracene-9-carboxylate ( 6 )). An octanuclear moiety [Nb82-O)12(L)81-L)4−x(OEt)4+x] (with L=2-naphtoate, x=0 or 2; 8 ) was obtained by using a solvothermal route in acetonitrile; it has a cubic configuration with niobium centers at each node, linked by 12 μ2-O groups. The formation of the niobium oxo clusters was characterized by infrared and liquid 1H NMR spectroscopy in order to analyze the esterification reaction, which induces the release of water molecules that further react through oxolation with niobium atoms, in different {Nb2O}, {Nb4O4} and {Nb8O12} nuclearities.  相似文献   

11.
The activation of CS2 is of interest in a broad range of fields and, more particularly, in the context of creating new C−C bonds. The reaction of the dinuclear ytterbium(II) complex [Yb2L4], 1 , [L=(OtBu)3SiO] with carbon disulfide led to the isolation of unprecedented reduction products. In particular, the crystallographic characterization of complex [Yb2L4(μ-C2S2)], 2 , provided the first example of an acetylenedithiolate ligand formed from metal reduction of CS2. Computational studies indicated that this unprecedented reactivity can be ascribed to the unusual binding mode of CS22− in the isolated “key intermediate” [Yb2L4(μ-CS2)], 3 , which results from the dinuclear nature of 1 .  相似文献   

12.
Trinuclear systems of formula [{Cr(LN3O2Ph)(CN)2}2M(H2LN3O2R)] (M=MnII and FeII, LN3O2R stands for pentadentate ligands) were prepared in order to assess the influence of the bending of the apical M−N≡C linkages on the magnetic anisotropy of the FeII derivatives and in turn on their Single-Molecule Magnet (SMM) behaviors. The cyanido-bridged [Cr2M] derivatives were obtained by assembling trans-dicyanido CrIII complex [Cr(LN3O2Ph)(CN)2] and divalent pentagonal bipyramid complexes [MII(H2LN3O2R)]2+ with various R substituents (R=NH2, cyclohexyl, S,S-mandelic) imparting different steric demand to the central moiety of the complexes. A comparative examination of the structural and magnetic properties showed an obvious effect of the deviation from straightness of the M−N≡C alignment on the slow relaxation of the magnetization exhibited by the [Cr2Fe] complexes. Theoretical calculations have highlighted important effects of the bending of the apical C−N−Fe linkages on both the magnetic anisotropy of the FeII center and the exchange interactions with the CrIII units.  相似文献   

13.
The self-assembly processes between binuclear [Zn2Ln]2+ complex cations and complex anions, [M(CN)2] [M(I) = Ag(I), Au(I)], generate new one-dimensional (1-D) coordination polymers: 1[{L1Zn23-OH)}2(H2O){μ-[Ag(CN)2]}](ClO4)3 THF 0.5MeOH 1, 1[{L1Zn23-OH)}2(H2O){μ-[Au(CN)2]}](ClO4)3 THF H2O 2, 1[{L2Zn2(μ-OH)}{μ-[Ag(CN)2]}][Ag(CN)2] H2O 3 (H2Ln are bicompartmental Schiff-base ligands resulting from condensation reactions between 2,6-diformyl-p-cresol with 2-aminomethyl-pyridine, and 2-aminoethyl-pyridine, respectively). The luminescence properties of the new heterometallic complexes have been investigated.  相似文献   

14.
The reaction of Na[CoIII(d -ebp)] (d -H4ebp = N,N′-ethylenebis[d -penicillamine]) with [(AuICl)2(dppe)] (dppe = 1,2-bis[diphenylphosphino]ethane) gave a cationic AuI4CoIII2 hexanuclear complex, [CoIII2(LAu4)]2+ ([ 1 ]2+), where [LAu4]4− is a cyclic tetragold(I) metalloligand with a 32-membered ring, [AuI4(dppe)2(d -ebp)2]4−. Complex [ 1 ]2+ crystallized with NO3 to produce a charge-separation (CS)-type ionic solid of [ 1 ](NO3)2. In [ 1 ](NO3)2, the complex cations are assembled to form cationic supramolecular hexamers of {[ 1 ]2+}6, which are closely packed in a face-centered cubic (fcc) lattice structure. The nitrate anions of [ 1 ](NO3)2 were accommodated in hydrophilic and hydrophobic tetrahedral interstices of the fcc structure to form tetrameric and hexameric nitrate clusters of {NO3}4 and {NO3}6, respectively. An analogous CS-type ionic solid formulated as [NiIICoIII(LAu4)](NO3) ([ 2 ](NO3)) was obtained when a 1:1 mixture of Na[CoIII(d -ebp)] and [NiII(d -H2ebp)] was reacted with [(AuICl)2(dppe)], accompanied by the conversion of the diamagnetic, square-planar [NiII(d -H2ebp)] to the paramagnetic, octahedral [NiII(d -ebp)]2−. While the overall fcc structure in [ 2 ](NO3) was similar to that of [ 1 ](NO3)2, none of the nitrate anions were accommodated in any hydrophobic tetrahedral interstice, reflecting the difference in the complex charges between [ 1 ]2+ and [ 2 ]+.  相似文献   

15.
Four calixarene-coordinated titanium-oxo clusters, namely, [Ti4(C6A)23-O)2(DMF)2] ( CIAC-258 ), [Ti8(H3C6A)4(C6H5PO3)82-O)4]4− ( CIAC-259 ), [Ti6(TC4A)32-O)3 (OiPr)6] ( CIAC-260 ) and [Ti6(H2TC4A)4(C6H5PO3)42-O)5(DMF)2]2− ( CIAC-261 ) were obtained, which feature sandwich-like, windmill-like, triangular, and tetrahedral structures, respectively. The polarity of the solvent determines the involvement of the auxiliary ligand phenylphosphonic acid in the formation of the products and their structures. Compounds CIAC-258 and CIAC-261 exhibit good catalytic performance in the selective oxidation of sulfides to sulfoxides with H2O2 as the oxidant, which can be attributed to the Ti active sites coordinated by exchangeable DMF molecules. Moreover, density functional theory (DFT) calculations revealed that the Ti-hydroperoxo species formed by the interaction of the Ti active site in CIAC-258 or CIAC-261 and H2O2 is the most likely catalytic active component during the catalytic sulfoxidation process.  相似文献   

16.
The reaction of 1,1,3,3‐tetraphenyl‐1,3‐disiloxandiol (LH2) with n‐butyllithium and CrCl2 results in a mononuclear chromium(II) complex ( 1 ) that further reacts with O2 at low temperatures to yield a mononuclear chromium(III) superoxide complex [L2CrO2(THF)][Li2(THF)3] ( 2 ). The crystal structure revealed that the chromium superoxido entity is stabilized by the coordination to an adjacent lithium cation. Complex 2 thus contains an unprecedented heterobimetallic [CrIII(μ‐O2)Li+] core; beyond this it is the first chromium superoxide for which a temperature‐dependent magnetic characterization could be achieved, and the first structurally characterized representative with chromium in an exclusive O‐donor environment.  相似文献   

17.
Using the redox-active tetrathiafulvalene tetrabenzoate (TTFTB4−) as the linker, a series of stable and porous rare-earth metal–organic frameworks (RE-MOFs), [RE93-OH)133-O)(H2O)9(TTFTB)3] ( 1-RE , where RE=Y, Sm, Gd, Tb, Dy, Ho, and Er) were constructed. The RE93-OH)133-O) (H2O)9](CO2)12 clusters within 1-RE act as segregated single-molecule magnets (SMMs) displaying slow relaxation. Interestingly, upon oxidation by I2, the S=0 TTFTB4− linkers of 1-RE were converted into S= TTFTB.3− radical linkers which introduced exchange-coupling between SMMs and modulated the relaxation. Furthermore, the SMM property can be restored by reduction in N,N-dimethylformamide. These results highlight the advantage of MOFs in the construction of redox-switchable SMMs.  相似文献   

18.
Metal‐superoxo species are involved in a variety of enzymatic oxidation reactions, and multi‐electron oxidation of substrates is frequently observed in those enzymatic reactions. A CrIII‐superoxo complex, [CrIII(O2)(TMC)(Cl)]+ ( 1 ; TMC=1,4,8,11‐tetramethyl‐1,4,8,11‐tetraazacyclotetradecane), is described that acts as a novel three‐electron oxidant in the oxidation of dihydronicotinamide adenine dinucleotide (NADH) analogues. In the reactions of 1 with NADH analogues, a CrIV‐oxo complex, [CrIV(O)(TMC)(Cl)]+ ( 2 ), is formed by a heterolytic O−O bond cleavage of a putative CrII‐hydroperoxo complex, [CrII(OOH)(TMC)(Cl)], which is generated by hydride transfer from NADH analogues to 1 . The comparison of the reactivity of NADH analogues with 1 and p ‐chloranil (Cl4Q) indicates that oxidation of NADH analogues by 1 proceeds by proton‐coupled electron transfer with a very large tunneling effect (for example, with a kinetic isotope effect of 470 at 233 K), followed by rapid electron transfer.  相似文献   

19.
《Mendeleev Communications》2021,31(5):628-630
Solid phase thermolysis of pivalate complex [Fe3O(Piv)6(HPiv)3]Piv generates the [Fe3O(Piv)6]+ complex cation due to a deficiency of ligands in the coordination sphere of the metal ions. Crystallization of [Fe3O(Piv)6]+ from THF–EtOH leads to the heteroleptic complex [Fe3O(Piv)6(THF)(EtOH)(OH)] · 0.5 THF · 0.5 H2O in 69% yield, while the reaction of [Fe3O(Piv)6]+ with AgNO3 in toluene results in the complex [Fe4Ag4O2(Piv)12] · 2 PhMe with a rare combination of FeIII and AgI atoms. Crystal structures of the two new complexes have been established.  相似文献   

20.
Reaction of [UO2Cl2(THF)3] with 3 equivalents of LiC6Cl5 in Et2O resulted in the formation of first uranyl aryl complex [Li(Et2O)2(THF)][UO2(C6Cl5)3] ([Li][ 1 ]) in good yields. Subsequent dissolution of [Li][ 1 ] in THF resulted in conversion into [Li(THF)4][UO2(C6Cl5)3(THF)] ([Li][ 2 ]), also in good yields. DFT calculations reveal that the U−C bonds in [Li][ 1 ] and [Li][ 2 ] exhibit appreciable covalency. Additionally, the 13C NMR chemical shifts for their Cipso environments are strongly affected by spin-orbit coupling—a consequence of 5f orbital participation in the U−C bonds.  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