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 46 毫秒
1.
A tetranuclear CeIV oxo cluster compound containing the Kläui tripodal ligand [Co(η5‐C5H5){P(O)(OEt)2}3]? (LOEt?) has been synthesized and its reactions with H2O2, CO2, NO, and Brønsted acids have been studied. The treatment of [Ce(LOEt)(NO3)3] with Et4NOH in acetonitrile afforded the tetranuclear CeIV oxo cluster [Ce4(LOEt)4O7H2] ( 1 ) containing an adamantane‐like {Ce42‐O)6} core with a μ4‐oxo ligand at the center. The reaction of 1 with H2O2 resulted in the formation of the peroxo cluster [Ce4(LOEt)44‐O)(μ2‐O2)42‐OH)2] ( 2 ). The treatment of 1 with CO2 and NO led to isolation of [Ce(LOEt)2(CO3)] and [Ce(LOEt)(NO3)3], respectively. The protonation of 1 with HCl, ROH (R=2,4,6‐trichlorophenyl), and Ph3SiOH yielded [Ce(LOEt)Cl3] ( 3 ), [Ce(LOEt)(OR)3] ( 4 ), and [Ce(LOEt)(OSiPh3)3] ( 5 ), respectively. The chloride ligands in 3 are labile and can be abstracted by silver(I) salts. The treatment of 3 with AgOTs (OTs?=tosylate) and Ag2O afforded [Ce(LOEt)(OTs)3] ( 6 ) and 1 , respectively. The electrochemistry of the Ce‐LOEt complexes has been studied by using cyclic voltammetry. The crystal structures of complexes 1 – 5 have been determined.  相似文献   

2.
Interaction of [LOEtZrF3] ( = [Co(η5-C5H5){P(O)(OEt)2}3]) (1) with 3 equivalents of bis(trimethylsilyl) sulfate afforded the ZrIV hydrogensulfato complex [(LOEt)2Zr2(SO4)2(HSO4)2] (2) that reacted with Et3N to give [Et3NH][LOEtZr(H2O)(SO4)2] (3). Treatment of complex 1 with 3 equivalents of trimethylsilyl acetate afforded [LOEtZr(OCOCH3)3] (4), whereas that with 1 and 2 equivalents of trimethylsilyl trimethylsiloxyacetate yielded [LOEtZrF(OCOCH2O)]2 (5) and [LOEtZr(OCOCH2OH)(OCOCH2O)]2 (6), respectively. The crystal structures of complexes 2 and 6 have been determined.  相似文献   

3.
Herein we present a systematic study of the structures and magnetic properties of six coordination compounds with mixed azide and zwitterionic carboxylate ligands, [M(N3)2(2‐mpc)] (2‐mpc=N‐methylpyridinium‐2‐carboxylate; M=Co for 1 and Mn for 2 ), [M(N3)2(4‐mpc)] (4‐mpc=N‐methylpyridinium‐4‐carboxylate; M=Co for 3 and Mn for 4 ), [Co3(N3)6(3‐mpc)2(CH3OH)2] ( 5 ), and [Mn3(N3)6(3‐mpc)2] ( 6 ; 3‐mpc=N‐methylpyridinium‐3‐carboxylate). Compounds 1 – 3 consist of one‐dimensional uniform chains with (μ‐EO‐N3)2(μ‐COO) triple bridges (EO=end‐on); 5 is also a chain compound but with alternating [(μ‐EO‐N3)2(μ‐COO)] triple and [(EO‐N3)2] double bridges; Compound 4 contains two‐dimensional layers with alternating [(μ‐EO‐N3)2(μ‐COO)] triple, [(μ‐EO‐N3)(μ‐COO)] double, and (EE‐N3) single bridges (EE=end‐to‐end); 6 is a layer compound in which chains similar to those in 5 are cross‐linked by a μ3‐1,1,3‐N3 azido group. Magnetically, the three CoII compounds ( 1 , 3 , and 5 ) all exhibit intrachain ferromagnetic interactions but show distinct bulk properties: 1 displays relaxation dynamics at very low temperature, 3 is an antiferromagnet with field‐induced metamagnetism due to weak antiferromagnetic interchain interactions, and 5 behaves as a noninnocent single‐chain magnet influenced by weak antiferromagnetic interchain interactions. The magnetic differences can be related to the interchain interactions through π–π stacking influenced by different substitution positions in the ligands and/or different magnitudes of intrachain coupling. All of the MnII compounds show overall intrachain/intralayer antiferromagnetic interactions. Compound 2 shows the usual one‐dimensional antiferromagnetism, whereas 4 and 6 exhibit different weak ferromagnetism due to spin canting below 13.8 and 4.6 K, respectively.  相似文献   

4.
The bridging fluoroolefin ligands in the complexes [Ir2(CH3)(CO)2(μ‐olefin)(dppm)2][OTf] (olefin=tetrafluoroethylene, 1,1‐difluoroethylene; dppm=μ‐Ph2PCH2PPh2; OTf?=CF3SO3?) are susceptible to facile fluoride ion abstraction. Both fluoroolefin complexes react with trimethylsilyltriflate (Me3SiOTf) to give the corresponding fluorovinyl products by abstraction of a single fluoride ion. Although the trifluorovinyl ligand is bound to one metal, the monofluorovinyl group is bridging, bound to one metal through carbon and to the other metal through a dative bond from fluorine. Addition of two equivalents of Me3SiOTf to the tetrafluoroethylene‐bridged species gives the difluorovinylidene‐bridged product [Ir2(CH3)(OTf)(CO)2(μ‐OTf)(μ‐C?CF2)(dppm)2][OTf]. The 1,1‐difluoroethylene species is exceedingly reactive, reacting with water to give 2‐fluoropropene and [Ir2(CO)2(μ‐OH)(dppm)2][OTf] and with carbon monoxide to give [Ir2(CO)3(μ‐κ12‐C?CCH3)(dppm)2][OTf] together with two equivalents of HF. The trifluorovinyl product [Ir21‐C2F3)(OTf)(CO)2(μ‐H)(μ‐CH2)(dppm)2][OTf], obtained through single C? F bond activation of the tetrafluoroethylene‐bridged complex, reacts with H2 to form trifluoroethylene, allowing the facile replacement of one fluorine in C2F4 with hydrogen.  相似文献   

5.
Treatment of [LOEtTi(OTf)3] (, OTf = triflate) with S-binapO2 (binap = 2,2′-bis(diphenylphosphinoyl)-1,1′-binaphthyl) afforded the terminal hydroxo complex [LOEtTi(S-binapO2)(OH)][OTf]2 (1). Treatment of [LOEtTi(OTf)3] with K(tpip) (tpip = [N(Ph2PO)2]) afforded [LOEtTi(tpip)(OTf)][OTf] (2) that reacted with CsOH to give [LOEtTi(tpip)(OH)][OTf] (3). The structures of 1 and 2 have been determined.  相似文献   

6.
Four copper(II) complexes and one copper(I) complex with pyridine-containing pyridylalkylamide ligands N-(pyridin-2-ylmethyl)pyrazine-2-carboxamide (HLpz) and N-(2-(pyridin-2-yl)ethyl)pyrazine-2-carboxamide (HLpz?) were synthesized and characterized. The X-ray crystal structures of [Cu2(Lpz)2(4,4?-bipy)(OTf)2] (1, OTf?=?trifluoromethanesulfonate, 4,4?-bipy?=?4,4?-bipyridine) and [Cu(Lpz)(py)2]OTf·H2O (2, py?=?pyridine) revealed binuclear and mononuclear molecular species, respectively, while [Cu(Lpz)(μ2-1,1-N3)]n (3), [Cu(Lpz?)(μ2-1,3-N3)]n (4), and [Cu(HLpz)Cl]n (5) are coordination polymer 1-D chains in the solid state.  相似文献   

7.
The reaction of the trans‐hyponitrito complex [Ru2(CO)4(μ‐η2‐ONNO)(μ‐H)(μ‐PtBu2)(μ‐dppen)] ( 1 , dppen = Ph2PC(=CH2)PPh2) with tetrafluorido boric acid afforded the new complex salt [Ru2(CO)4(μ‐η2‐ONNOH)(μ‐H)(μ‐PtBu2)(μ‐dppen)]BF4 ( 2 ) containing the monoprotonate hyponitrous acid as the ligand in the cationic complex. Complex 1 showed a nucleophilic reactivity towards the trimethyloxonium cation resulting in the monoester derivative of the hyponitrous acid [Ru2(CO)4(μ‐η2‐ONNOMe)(μ‐H)(μ‐PtBu2)(μ‐dppen)]BF4 ( 3 ). During heating of compound 2 in ethanol under reflux for a short time nitrous oxide was liberated affording unexpectedly a new tridentate 2, 2‐bis(diphenylphosphanyl)ethanolato ligand formed by an intramolecular attack of an intermediate hydroxido ligand towards the unsaturated carbon carbon double bond in the bridging dppen ligand. Thus the complex salt [Ru2(CO)4{μ‐η3‐OCH2CH(PPh2)2}(μ‐H)(μ‐PtBu2)]BF4 ( 4 ) was formed in good yields. The new compounds 2 , 3 , and 4 were characterized by spectroscopic means as well as their molecular structures were determined in the crystal.  相似文献   

8.
The three zinc sulfate complexes presented herein display three completely different coordination modes, viz tri­aqua(1,10‐phenanthroline‐N,N′)(sulfato‐O)­zinc(II) hydrate, [Zn(SO4)(C12H8N2)(H2O)3]·H2O (octahedral, monomeric), bis(μ‐sulfato‐O:O′)­bis[(2,9‐di­methyl‐1,10‐phenanthroline‐N,N′)­zinc(II)], [Zn2(SO4)2(C14H12N2)2] (tetrahedral, dimeric), and catena‐poly­[[di­aqua(2,2′‐bipyridyl‐N,N′)­zinc(II)]‐μ‐(sul­fato‐O:O′)], [Zn(SO4)(C10H8N2)(H2O)2]n (octahedral, polymeric, twofold crystallographic symmetry). In the first, the sulfate is monodentate, while in the other two it acts as a bidentate bridge between two different Zn centers. There is a variety of sulfate S—O bond lengths, depending on the different coordination conditions and hydrogen‐bonding interactions.  相似文献   

9.
Bimetallic macrocyclic complexes have attracted the attention of chemists and various organic ligands have been used as molecular building blocks, but supramolecular complexes based on semi‐rigid organic ligands containing 1,2,4‐triazole have remained rare until recently. It is easier to obtain novel topologies by making use of asymmetric semi‐rigid ligands in the self‐assembly process than by making use of rigid ligands. A new semi‐rigid ligand, 3‐[(pyridin‐4‐ylmethyl)sulfanyl]‐5‐(quinolin‐2‐yl)‐4H‐1,2,4‐triazol‐4‐amine (L), has been synthesized and used to generate two novel bimetallic macrocycle complexes, namely bis{μ‐3‐[(pyridin‐4‐ylmethyl)sulfanyl]‐5‐(quinolin‐2‐yl)‐4H‐1,2,4‐triazol‐4‐amine}bis[(methanol‐κO)(nitrato‐κ2O,O′)nickel(II)] dinitrate, [Ni2(NO3)2(C17H14N6S)2(CH3OH)2](NO3)2, (I), and bis{μ‐3‐[(pyridin‐4‐ylmethyl)sulfanyl]‐5‐(quinolin‐2‐yl)‐4H‐1,2,4‐triazol‐4‐amine}bis[(methanol‐κO)(nitrato‐κ2O,O′)zinc(II)] dinitrate, [Zn2(NO3)2(C17H14N6S)2(CH3OH)2](NO3)2, (II), by solution reactions with the inorganic salts M(NO3)2 (M = Ni and Zn, respectively) in mixed solvents. In (I), two NiII cations with the same coordination environment are linked by L ligands through Ni—N bonds to form a bimetallic ring. Compound (I) is extended into a two‐dimensional network in the crystallographic ac plane via N—H…O, O—H…N and O—H…O hydrogen bonds, and neighbouring two‐dimensional planes are parallel and form a three‐dimensional structure via π–π stacking. Compound (II) contains two bimetallic rings with the same coordination environment of the ZnII cations. The ZnII cations are bridged by L ligands through Zn—N bonds to form the bimetallic rings. One type of bimetallic ring constructs a one‐dimensional nanotube via O—H…O and N—H…O hydrogen bonds along the crystallographic a direction, and the other constructs zero‐dimensional molecular cages via O—H…O and N—H…O hydrogen bonds. They are interlinked into a two‐dimensional network in the ac plane through extensive N—H…O hydrogen bonds, and a three‐dimensional supramolecular architecture is formed via π–π interactions between the centroids of the benzene rings of the quinoline ring systems.  相似文献   

10.
New Azido Complexes of Tantalum(V). Synthesis and Molecular Structure of the Dinuclear Compounds [Cp*TaCl(N3)(μ‐N3)]2(μ‐O) and [Cp*Ta(N3)3(μ‐N3)]2 (Cp* = Pentamethylcyclopentadienyl) The reaction of Cp*TaCl4 ( 1 ) with an excess of trimethylsilyl azide (Me3Si–N3) leads to azide‐rich dinuclear complexes which contain both terminal and bridging azido ligands. The oxo complex [Cp*TaCl(N3)(μ‐N3)]2(μ‐O) ( 4 ) was formed in dichloromethane in the presence of traces of water, whereas [Cp*Ta(N3)3(μ‐N3)]2 ( 5 ) was obtained from boiling toluene after several days. According to the X‐ray structure determinations the Ta…Ta distance in 4 (314,5 pm) is considerably shorter than in 5 (382,2 pm).  相似文献   

11.
A chemically non‐innocent pyrrole‐based trianionic (ONO)3? pincer ligand within [(pyr‐ONO)TiCl(thf)2] ( 2 ) can access the dianionic [(3H‐pyr‐ONO)TiCl2(thf)] ( 1‐THF ) and monoanionic [(3H,4H‐pyr‐ONO)TiCl2(OEt2)][B{3,5‐(CF3)2C6H3}4] ( 3‐Et2O ) states through remote protonation of the pyrrole γ‐C π‐bonds. The homoleptic [(3H‐pyr‐ONO)2Zr] ( 4 ) was synthesized and characterized by X‐ray diffraction and NMR spectroscopy in solution. The protonation of 4 by [H(OEt2)2][B{C6H3(CF3)2}4] yields [(3H,4H‐pyr‐ONO)(3H‐pyr‐ONO)Zr][B{3,5‐(CF3)2C6H3}4] ( 5 ), thus demonstrating the storage of three protons.  相似文献   

12.
Reactions of Cp*NbCl4 and Cp*TaCl4 with Trimethylsilyl‐azide, Me3Si‐N3. Molecular Structures of the Bis(azido)‐Oxo‐Bridged Complexes [Cp*NbCl(N3)(μ‐N3)]2(μ‐O) and [Cp*TaCl2(μ‐N3)]2(μ‐O) (Cp* = Pentamethylcyclopentadienyl) The chloro ligands in Cp*TaCl4 (1c) can be stepwise substituted for azido ligands by reactions with trimethylsilyl azide, Me3Si‐N3 (A) , to generate the complete series of the bis(azido)‐bridged dimers [Cp*TaCl3‐n(N3)n(μ‐N3)]2 ( n = 0 (2c) , n = 1 (3c) , n = 2 (4c) and n = 3 (5c) ). If the solvent CH2Cl2 contains traces of water, an additional oxo bridge is incorporated to give [Cp*‐TaCl2(μ‐N3)]2(μ‐O) (6c) or [Cp*TaCl(N3)(μ‐N3)]2(μ‐O) (7c) , respectively. Both 6c and 7c are also formed in stoichiometric reactions from [Cp*TaCl2(μ‐OH)]2(μ‐O) (8c) and A . Analogous reactions of Cp*NbCl4 (1b) with A were used to prepare the azide‐rich dinuclear products [Cp*NbCl3‐n(N3)n(μ‐N3)]2 (n = 2 (4b) , and n = 3 (5b) ), and [Cp*NbCl(N3)(μ‐N3)]2(μ‐O) (7b) . The mononuclear complex Cp*Ta(N3)Me3 (10c) is obtained from Cp*Ta(Cl)Me3 and A . All azido complexes were characterised by their IR as well as their 1H and 13C NMR spectra; X‐ray crystal structure analyses are available for 6c and 7b .  相似文献   

13.
Synthesis and deprotonation reactions of half‐sandwich iridium complexes bearing a vicinal dioxime ligand were studied. Treatment of [{Cp*IrCl(μ‐Cl)}2] (Cp*=η5‐C5Me5) with dimethylglyoxime (LH2) at an Ir:LH2 ratio of 1:1 afforded the cationic dioxime iridium complex [Cp*IrCl(LH2)]Cl ( 1 ). The chlorido complex 1 undergoes stepwise and reversible deprotonation with potassium carbonate to give the oxime–oximato complex [Cp*IrCl(LH)] ( 2 ) and the anionic dioximato(2?) complex K[Cp*IrCl(L)] ( 3 ) sequentially. Meanwhile, twofold deprotonation of the sulfato complex [Cp*Ir(SO4)(LH2)] ( 4 ) resulted in the formation of the oximato‐bridged dinuclear complex [{Cp*Ir(μ‐L)}2] ( 5 ). X‐ray analyses disclosed their supramolecular structures with one‐dimensional infinite chain ( 1 and 2 ), hexagonal open channels ( 3 ), and a tetrameric rhomboid ( 4 ) featuring multiple intermolecular hydrogen bonds and electrostatic interactions.  相似文献   

14.
Seven new mixed oxochalcogenate compounds in the systems MII/XVI/TeIV/O/(H), (MII = Ca, Cd, Sr; XVI = S, Se) were obtained under hydrothermal conditions (210 °C, one week). Crystal structure determinations based on single‐crystal X‐ray diffraction data revealed the compositions Ca3(SeO4)(TeO3)2, Ca3(SeO4)(Te3O8), Cd3(SeO4)(Te3O8), Cd3(H2O)(SO4)(Te3O8), Cd4(SO4)(TeO3)3, Cd5(SO4)2(TeO3)2(OH)2, and Sr3(H2O)2(SeO4)(TeO3)2 for these phases. Peculiar features of the crystal structures of Ca3(SeO4)(TeO3)2, Ca3(SeO4)(Te3O8), Cd3(SeO4)(Te3O8), Cd3(H2O)(SO4)(Te3O8), and Sr3(H2O)2(SeO4)(TeO3)2 are metal‐oxotellurate(IV) layers connected by bridging XO4 tetrahedra and/or by hydrogen‐bonding interactions involving hydroxyl or water groups, whereas Cd4(SO4)(TeO3)3 and Cd5(SO4)2(TeO3)2(OH)2 crystallize as framework structures. Common to all crystal structures is the stereoactivity of the TeIV electron lone pair for each oxotellurate(IV) unit, pointing either into the inter‐layer space, or into channels and cavities in the crystal structures.  相似文献   

15.
Reactions of cerium(III) nitrate, Ce(NO3)3?6 H2O, with different carboxylic acids, such as pivalic acid, benzoic acid, and 4‐methoxybenzoic acid, in the presence of a tridentate N,N,N‐donor ligand, diethylenetriamine (L1), under aerobic conditions yielded the corresponding cerium hexamers Ce6O8(O2CtBu)8(L1)4 ( 1 ), Ce6O8(O2CC6H5)8(L1)4 ( 2 ), and Ce6O8(O2CC6H4‐4‐OCH3)8(L1)4 ( 3 ). Hexamers 1 , 2 , and 3 contain the same octahedral CeIV6O8 core, in which all interstitial oxygen atoms are connected by μ3‐oxo bridging ligands. In contrast, treatment of the CeIV precursor (NH4)2Ce(NO3)6 (CAN) with pivalic acid and the ligand L1 under the same conditions afforded Ce6O4(OH)4(O2CtBu)12(L1)2 ( 4 ), exhibiting a deformed octahedral CeIV6O4(OH)4 core containing μ3‐oxo and μ3‐hydroxo moieties in defined positions. In contrast to the formation of 1 – 3 , the use of N‐methyldiethanolamine (L) in the reaction with Ce(NO3)3?6 H2O and pivalic acid afforded a previously reported CeIII dinuclear cluster, Ce2(O2CtBu)6L2, even in the presence of dioxygen. ESI‐MS analysis of the reaction mixture clearly indicated the importance of the ligand L1 in promoting oxidation of the CeIII aggregates, [Cen(O2CtBu)3n(L1)2], which is necessary for the formation of CeIV hexamers.  相似文献   

16.
The reaction of [CpRuCl(PPh3)2] (Cp=cyclopentadienyl) and [CpRuCl(dppe)] (dppe=Ph2PCH2CH2PPh2) with bis‐ and tris‐phosphine ligands 1,4‐(Ph2PC≡C)2C6H4 ( 1 ) and 1,3,5‐(Ph2PC≡C)3C6H3 ( 2 ), prepared by Ni‐catalysed cross‐coupling reactions between terminal alkynes and diphenylchlorophosphine, has been investigated. Using metal‐directed self‐assembly methodologies, two linear bimetallic complexes, [{CpRuCl(PPh3)}2(μ‐dppab)] ( 3 ) and [{CpRu(dppe)}2(μ‐dppab)](PF6)2 ( 4 ), and the mononuclear complex [CpRuCl(PPh3)(η1‐dppab)] ( 6 ), which contains a “dangling arm” ligand, were prepared (dppab=1,4‐bis[(diphenylphosphino)ethynyl]benzene). Moreover, by using the triphosphine 1,3,5‐tris[(diphenylphosphino)ethynyl]benzene (tppab), the trimetallic [{CpRuCl(PPh3)}33‐tppab)] ( 5 ) species was synthesised, which is the first example of a chiral‐at‐ruthenium complex containing three different stereogenic centres. Besides these open‐chain complexes, the neutral cyclic species [{CpRuCl(μ‐dppab)}2] ( 7 ) was also obtained under different experimental conditions. The coordination chemistry of such systems towards supramolecular assemblies was tested by reaction of the bimetallic precursor 3 with additional equivalents of ligand 2 . Two rigid macrocycles based on cis coordination of dppab to [CpRu(PPh3)] were obtained, that is, the dinuclear complex [{CpRu(PPh3)(μ‐dppab)}2](PF6)2 ( 8 ) and the tetranuclear square [{CpRu(PPh3)(μ‐dppab)}4](PF6)4 ( 9 ). The solid‐state structures of 7 and 8 have been determined by X‐ray diffraction analysis and show a different arrangement of the two parallel dppab ligands. All compounds were characterised by various methods including ESIMS, electrochemistry and by X‐band ESR spectroscopy in the case of the electrogenerated paramagnetic species.  相似文献   

17.
《中国化学会会志》2017,64(1):61-72
The stable tribridged dicopper(I) carboxylate complexes [Cu2(μ‐dppm)2(μ‐O2CR)]BF4 (RCO2 = formate (OFc), m1 ; acetate (OAc), m2 ; benzoate (OBAc), m3 ; o‐toluate (O2TAc), m4 ; p‐toluate (O4TAc), m5 ; 4‐phenylbutyrate (O4PBAc), m6 ; 2‐nitrobenzoate (O2NBAc), m7 ), abbreviated as MM, and neutral dipyridyl compounds (NN; NN = 4,4′‐bipyridine (bpy), 1,2‐bis(4‐pyridyl)ethane (bpa), trans ‐1,2‐bis(4‐pyridyl)ethylene (bpe), 4,4′‐trimethylenedipyridine (tmp)) can form dynamic equilibria in CH2Cl2. From the equilibrium mixtures containing MM and NN with MM/NN = 1:1, nine 2:1 oligomers ([( m1 )2(μ‐bpy)](BF4)2 ( o1a (BF4)2), [( m3 )2(μ‐bpe)](BF4)2 ( o3c (BF4)2), [( m3 )2(μ‐tmp)](BF4)2 ( o3d (BF4)2), [( m4 )2(μ‐bpe)](BF4)2 ( o4c (BF4)2), [( m5 )2(μ‐bpy)](BF4)2 ( o5a (BF4)2), [( m5 )2(μ‐tmp)](BF4)2 ( o5d (BF4)2), [( m6 )2(μ‐bpa)](BF4)2 ( o6b (BF4)2), [( m7 )2(μ‐bpy)](BF4)2 ( o7a (BF4)2), [( m7 )2(μ‐bpa)](BF4)2 ( o7b (BF4)2)), one 2:3 oligomer ([{( m2 )(bpy)}2(μ‐bpy)](BF4)2 ( o2a (BF4)2)), and five 1:1 polymers ([( m2 )(μ‐bpe)] n (BF4 ) n ( p2c (BF4 ) n ), [( m2 )(μ‐tmp)] n (BF4 ) n ( p2d (BF4 ) n ), [( m3 )(μ‐bpy)] n (BF4 ) n ( p3a (BF4 ) n ), [( m3 )(μ‐tmp)] n (BF4 ) n ( p3d (BF4 ) n ), [( m7 )(μ‐tmp)] n (BF4 ) n ( p7d (BF4 ) n )) were obtained as single crystals, and their structures were determined by X‐ray crystallography. Both experimental and theoretical results support the presence of two oligomeric species, [{Cu2(μ‐dppm)2(μ‐O2CR)}2(μ‐NN)]2+ and [{Cu2(μ‐dppm)2(μ‐O2CR)(NN)}2(μ‐NN)]2+), in dynamic equilibrium. The oligomers (such as o3d (BF4)2) can serve as seeds to induce the formation of soluble coordination polymers as crystals (such as p3d (BF4)n ).  相似文献   

18.
Tantalum complexes [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(CH2NMe2)?CH)py}] ( 4 ) and [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(CH2NH2)?CH)py}] ( 5 ), which contain modified alkoxide pincer ligands, were synthesized from the reactions of [TaCp*Me{κ3N,O,O‐(OCH2)(OCH)py}] (Cp*=η5‐C5Me5) with HC?CCH2NMe2 and HC?CCH2NH2, respectively. The reactions of [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(Ph)?CH)py}] ( 2 ) and [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(SiMe3)?CH)py}] ( 3 ) with triflic acid (1:2 molar ratio) rendered the corresponding bis‐triflate derivatives [TaCp*(OTf)23N,O,O‐(OCH2)(OCHC(Ph)?CH2)py}] ( 6 ) and [TaCp*(OTf)23N,O,O‐(OCH2)(OCHC(SiMe3)?CH2)py}] ( 7 ), respectively. Complex 4 reacted with triflic acid in a 1:2 molar ratio to selectively yield the water‐soluble cationic complex [TaCp*(OTf){κ4C,N,O,O‐(OCH2)(OCHC(CH2NHMe2)?CH)py}]OTf ( 8 ). Compound 8 reacted with water to afford the hydrolyzed complex [TaCp*(OH)(H2O){κ3N,O,O‐(OCH2)(OCHC(CH2NHMe2)?CH2)py}](OTf)2 ( 9 ). Protonation of compound 8 with triflic acid gave the new tantalum compound [TaCp*(OTf){κ4C,N,O,O‐(OCH2)(HOCHC(CH2NHMe2)?CH)py}](OTf)2 ( 10 ), which afforded the corresponding protonolysis derivative [TaCp*(OTf)23N,O,O‐(OCH2)(HOCHC(CH2NHMe2)?CH2)py}](OTf) ( 11 ) in solution. Complex 8 reacted with CNtBu and potassium 2‐isocyanoacetate to give the corresponding iminoacyl derivatives 12 and 13 , respectively. The molecular structures of complexes 5 , 7 , and 10 were established by single‐crystal X‐ray diffraction studies.  相似文献   

19.
Treating [Cp*V(μ‐Cl)2]3 (Cp* = C5Me5) and [(2,6‐i‐Pr2C6H3N)2MoMe2], respectively, with Me3SnF afforded the title compounds [Cp*V(μ‐F)2]4 ( 1 ) and [(2,6‐i‐Pr2C6H3N)2MoF2] · THF ( 2 ). 1 has a tetrameric structure, in which four V atoms can be regarded as being arranged at the vertices of a distorted tetrahedron, with four long edges bridged by one F atom and each of the other two short edges bridged by two F atoms with a mean V–F bond length of 2.00 Å. A hydrolyzed product of 2 , [(2,6‐i‐Pr2C6H3N)6Mo43‐F)2Me2(μ‐O)4] ( 3 ) was characterized by elemental analyses and X‐ray single crystal study. The X‐ray diffraction analysis reveals that 3 has a unique tetranuclear structure, containing two five and two six coordinated Mo atoms connecting each other by four μ‐O and two μ3‐F atoms. The geometries around the two Mo atoms can be described having distorted trigonal bipyramidal and distorted octahedral coordination spheres, respectively. The Mo–(μ‐O) bond lengths are 1.813 Å (average) for five coordinated Mo atoms and 2.030 Å (average) for those of six coordinated, respectively, indicating an additional π bonding between five coordinated Mo atoms and the μ‐O atoms. The Mo–(μ3‐F) distances range from 2.291 to 2.352 Å.  相似文献   

20.
The compounds tert‐butylarsenium(III) tri‐μ‐chlorido‐bis[trichloridotitanium(IV)], (C4H12As)[Ti2Cl9] or [tBuAsH3][Ti2(μ‐Cl)3Cl6], (II), and bis[bromidotriphenylarsenium(V)] di‐μ‐bromido‐μ‐oxido‐bis[tribromidotitanium(IV)], (C18H15AsBr)2[Ti2Br8O] or [Ph3AsBr]2[Ti2(μ‐O)(μ‐Br)2Br6], (III), were obtained unexpectedly from the reaction of simple arsane ligands with TiIV halides, with (II) lying on a mirror plane in the unit cell of the space group Pbcm. Both compounds contain a completely novel ion, with [tBuAsH3]+ constituting the first structurally characterized example of a primary arsenium cation. The oxide‐bridged titanium‐containing [Ti2(μ‐O)(μ‐Br)2Br6]2− dianion in (III) is also novel, while the bromidotriphenylarsenium(V) cation is structurally characterized for only the second time.  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