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.
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   

2.
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   

3.
Molecular and Crystal Structure of Bis[chloro(μ‐phenylimido)(η5‐pentamethylcyclopentadienyl)tantalum(IV)](Ta–Ta), [{TaCl(μ‐NPh)Cp*}2] Despite the steric hindrance of the central atom in [TaCl2(NPh)Cp*] (Ph = C6H5, Cp* = η5‐C5(CH3)5), caused by the Cp* ligand, the imido‐ligand takes a change in bond structure when this educt is reduced to the binuclear complex [{TaCl(μ‐NPh)Cp*}2] in which tantalum is stabilized in the unusual oxidation state +4.  相似文献   

4.
Yttrocene‐carboxylate complex [Cp*2Y(OOCArMe)] (Cp*=C5Me5, ArMe=C6H2Me3‐2,4,6) was synthesized as a spectroscopically versatile model system for investigating the reactivity of alkylaluminum hydrides towards rare‐earth‐metal carboxylates. Equimolar reactions with bis‐neosilylaluminum hydride and dimethylaluminum hydride gave adduct complexes of the general formula [Cp*2Y(μ‐OOCArMe)(μ‐H)AlR2] (R=CH2SiMe3, Me). The use of an excess of the respective aluminum hydride led to the formation of product mixtures, from which the yttrium‐aluminum‐hydride complex [{Cp*2Y(μ‐H)AlMe2(μ‐H)AlMe2(μ‐CH3)}2] could be isolated, which features a 12‐membered‐ring structure. The adduct complexes [Cp*2Y(μ‐OOCArMe)(μ‐H)AlR2] display identical 1J(Y,H) coupling constants of 24.5 Hz for the bridging hydrido ligands and similar 89Y NMR shifts of δ=?88.1 ppm (R=CH2SiMe3) and δ=?86.3 ppm (R=Me) in the 89Y DEPT45 NMR experiments.  相似文献   

5.
Azido Derivatives of the Pentamethylcyclopentadienyl Vanadium(IV)-Fragment. Molecular Structures of the Binuclear Complexes [Cp*VCl(N3)(μ-N3)]2 and [Cp*V(N3)2(μ-N3)]2 The stepwise reaction of Cp*VCl3 with excess trimethylsilyl azide (Me3Si–N3) in solution leads to the paramagnetic, azido-bridged complexes [Cp*VCl2(μ-N3)]2 ( 3 ), [Cp*VCl(N3)(μ-N3)]2 ( 4 ) and [Cp*V(N3)2(μ-N3)]2 ( 5 ) which were characterized by their IR and mass spectra. The azide-rich binuclear complex 5 is also formed if a pentane solution of Cp*V(CO)4 is stirred in the presence of excess Me3Si–N3 in an open vessel. According to the X-ray structure analyses both 4 and 5 are centrosymmetric molecules with a planar V(N)2V four-membered ring. In the absence of free trimethylsilyl azide, solutions of 3 – 5 lose dinitrogen slowly; in the presence of traces of air, 5 is thereby converted to the diamagnetic, oxo-bridged complex [Cp*V(O)(N3)]2(μ-O) ( 6 ).  相似文献   

6.
Chalcogen Derivatives of the Halfsandwich Tungsten(V) Complexes Cp*WCl4 and Cp*WCl4(PMe3). X‐Ray Crystal Structure Analyses of anti ‐[Cp*W(Se)(μ‐Se)]2 and Cp*W(S)2(OMe) The chalcogenation of Cp*WCl4 ( 1 ) by E(SiMe3)2 (E = S, Se) and Te(SiMe2tBu)2 in chloroform solution leads to dimeric products of the type anti‐[Cp*W(E)(μ‐E)]2 (E = S ( 3 a ), Se ( 3 b ) and Te ( 3 c )). An X‐ray structure determination of 3 b indicates a centrosymmetric molecule containing a planar W(μ‐Se)2W ring, the W–W distance (297.9(1) pm) corresponds to a single bond. In the presence of air the two terminal chalcogenido ligands (E) in 3 a – c are stepwise replaced by oxido ligands (O) to give [Cp*W(O)(μ‐E)]2 (E = S ( 5 a ), Se ( 5 b ) and Te ( 5 c )) in quantitative yields. The reaction of Cp*WCl4 with H2S or ammonium polysulfide, (NH4)2Sx (x ∼ 10), leads to Cp*W(S)2Cl ( 6 a ); the corresponding methoxy derivative, Cp*W(S)2OCH3 ( 9 a ), has been characterized by an X‐ray structure analysis. On the other hand, the reaction of Cp*WCl4(PMe3) ( 2 ) with sodium tetrasulfide, Na2S4, in dimethylformamide solution gives a mixture of mononuclear Cp*W(S)(S2)Cl ( 8 a ), dinuclear [Cp*W(S)(μ‐S)]2 ( 3 a ) and a trinuclear side‐product of composition Cp*2W3S7 ( 13 a ). Terminal sulfido ligands are replaced by terminal oxido ligands in solution in the presence of oxygen. Thus, 6 a is stepwise converted into Cp*W(O)(S)Cl ( 10 a ) and CpW(O)2Cl ( 12 a ), whereas 8 a gives Cp*W(O)(S2)Cl ( 11 a ) and 13 a leads to Cp*2W3(O)S6 ( 14 a ). The disulfido complexes 8 a and 11 a are desulfurized by triphenylphosphane to give 6 a and 10 a . The new complexes have been characterized by their IR and NMR spectra and by mass spectrometry.  相似文献   

7.
Reactions of Group 4 metallocene alkyne complexes [Cp′2M(η2‐Me3SiC2SiMe3)] ( 1 : M=Zr, Cp′=Cp*=η5‐pentamethylcyclopentadienyl; 2 a : M=Ti, Cp′=Cp*, and 2 b : M=Ti, Cp′2=rac‐(ebthi)=rac‐1,2‐ethylene‐1,1′‐bis(η5‐tetrahydroindenyl)) with diphenylacetonitrile (Ph2CHCN) and of the seven‐membered zirconacyclocumulene 3 with phenylacetonitrile (PhCH2CN) were investigated. Different compounds were obtained depending on the metal, the cyclopentadienyl ligand and the reaction temperature. In the first step, Ph2CHCN coordinated to 1 to form [Cp*2Zr(η2‐Me3SiC2SiMe3)(NCCHPh2)] ( 4 ). Higher temperatures led to elimination of the alkyne, coordination of a second Ph2CHCN and transformation of the nitriles to a keteniminate and an imine ligand in [Cp*2Zr(NC2Ph2)(NCHCHPh2)] ( 5 ). The conversion of 4 to 5 was monitored by using 1H NMR spectroscopy. The analogue titanocene complex 2 a eliminated the alkyne first, which led directly to [Cp*2Ti(NC2Ph2)2] ( 6 ) with two keteniminate ligands. In contrast, the reaction of 2 b with diphenylacetonitrile involved a formal coupling of the nitriles to obtain the unusual four‐membered titanacycle 7 . An unexpected six‐membered fused zirconaheterocycle ( 8 ) resulted from the reaction of 3 with PhCH2CN. The molecular structures of complexes 4 , 5 , 6 , 7 and 8 were determined by X‐ray crystallography.  相似文献   

8.
In the crystal structure of the title complex, [Zn(N3)2(C6H8N6)]n or [Zn(N3)2(bte)]n, where bte is μ‐1,2‐bis(1,2,4‐triazol‐1‐yl)­ethane, each Zn atom is pentacoordinated in a distorted trigonal‐bipyramidal coordination environment involving two N atoms from two bte ligands and three N atoms from three azide ligands. The Zn atoms are bridged by μ‐1,1‐azide groups and bte ligands around a centre of inversion, forming an infinite one‐dimensional chain containing both four‐membered Zn(μ‐1,1‐N3)2Zn and 18‐membered Zn(gauche‐bte)2Zn rings.  相似文献   

9.
Reaction of DyCl3 with two equivalents of NaN(SiMe3)2 in THF yielded {Dy(μ‐Cl)[N(SiMe3)2]2(THF)}2 ( 1 ). X‐ray crystal structure analysis revealed that 1 is a centrosymmetric dimer with asymmetrically bridging chloride ligands. The metal coordination arrangement can be best described as distorted trigonal bipyramid. The bond lengths of Ln–Cl and Ln–N showed a decreasing trend with the contraction of the size of Ln3+. Treatment of N,N‐bis(pyrrolyl‐α‐methyl)‐N‐methylamine (H2dpma) with 1 and known compound {Yb(μ‐Cl)[N(SiMe3)2]2(THF)}2, respectively, led to the formations of [Dy(μ‐Cl)(dpma)(THF)2]2 ( 2 ) and {Yb(μ‐Cl)[N(SiMe3)2]2(THF)}2 ( 3 ). Compounds 2 and 3 were fully characterized by single‐crystal X‐ray crystallography, elemental analysis, and 1H NMR spectroscopy. Structure determination indicated that 2 and 3 exhibit as centrosymmetric dimers with asymmetrically bridging chloride ligands. One pot reactions involving LnCl3 (Ln = Dy and Yb), LiN(SiMe3)2, and H2dpma were explored and desired products 2 and 3 were not yielded, which indicated that 1 and {Yb(μ‐Cl)[N(SiMe3)2]2(THF)}2 are the demanding precursors to synthesize Dysprosium and Ytterbium complexes supported by dpma2– ligand. Compounds 2 and 3 are the first reported lanthanide complexes chelated by dpma2– ligand.  相似文献   

10.
On the Reactivity of Titanocene Complexes [Ti(Cp′)22‐Me3SiC≡CSiMe3)] (Cp′ = Cp, Cp*) towards Benzenedicarboxylic Acids Titanocene complexes [Ti(Cp′)2(BTMSA)] ( 1a , Cp′ = Cp = η5‐C5H5; 1b , Cp′ = Cp* = η5‐C5Me5; BTMSA = Me3SiC≡CSiMe3) were found to react with iodine and methyl iodide yielding [Ti(Cp′)2(μ‐I)2] ( 2a / b ; a refers to Cp′ = Cp and b to Cp′ = Cp*), [Ti(Cp′)2I2] ( 3a / b ) and [Ti(Cp′)2(Me)I] ( 4a / b ), respectively. In contrast to 2a , complex 2b proved to be highly moisture sensitive yielding with cleavage of HCp* [{Ti(Cp*)I}2(μ‐O)] ( 7 ). The corresponding reactions of 1a / b with p‐cresol and thiophenol resulted in the formation of [Ti(Cp′)2{O(p‐Tol)}2] ( 5a / b ) and [Ti(Cp′)2(SPh)2] ( 6a / b ), respectively. Reactions of 1a and 1b with 1,n‐benzenedicarboxylic acids (n = 2–4) resulted in the formation of dinuclear titanium(III) complexes of the type [{Ti(Cp′)2}2{μ‐1,n‐(O2C)2C6H4}] (n = 2, 8a / b ; n = 3, 9a / b ; n = 4, 10a / b ). All complexes were fully characterized analytically and spectroscopically. Furthermore, complexes 7 , 8b , 9a ·THF, 10a / b were also be characterized by single‐crystal X‐ray diffraction analyses.  相似文献   

11.
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   

12.
The reactions of Cp*M(PMe3)Cl2 (M = Rh ( 1a ), Ir ( 1b )) with (NEt4)2[WS4] led to the heterodimetallic sulfido‐bridged complexes Cp*M(PMe3)[(μ‐S)2WS2] (M = Rh ( 2a ), Ir ( 2b )), whereas the dimers [Cp*MCl(μ‐Cl)]2 (M = Rh ( 4a ), Ir ( 4b )) reacted with (NEt4)2[WS4) to give the known trinuclear compounds [Cp*M(Cl)]2(μ‐WS4) (M = Rh ( 5a ), Ir ( 5b )). Hydrolysis of the terminal W=S bonds converts 2a, b into Cp*M(PMe3)[(μ‐S)2WO2] (M = Rh ( 3a ), Ir ( 3b )). Salts of a heterodimetallic anion, A[CpMo(I)(NO)(WS4)] ( 6 ) (A+ = NEt4+, NPh4+) were obtained by reactions of [CpMo(NO)I2]2 with tetrathiotungstates, A2[WS4]. The complexes were characterized by IR and NMR (1H, 13C, 31P) spectroscopy, and the X‐ray crystallographic structure of Cp*Rh(PMe3)[(μ‐S)2WS2] ( 2a ) has been determined. The bond lengths and angles in the coordinations spheres of Rh and W in 2a (Rh···W 288.5(1) pm) are compared with related complexes containing terminal [WS42—] chelate ligands.  相似文献   

13.
Metallacyclic complex [(Me2N)3Ta(η2‐CH2SiMe2NSiMe3)] ( 3 ) undergoes C?H activation in its reaction with H3SiPh to afford a Ta/μ‐alkylidene/hydride complex, [(Me2N)2{(Me3Si)2N}Ta(μ‐H)2(μ‐C‐η2‐CHSiMe2NSiMe3)Ta(NMe2)2] ( 4 ). Deuterium‐labeling studies with [D3]SiPh show H–D exchange between the Ta?D ?Ta unit and all methyl groups in [(Me2N)2{(Me3Si)2N}Ta(μ‐D)2(μ‐C‐η2‐CHSiMe2NSiMe3)Ta(NMe2)2] ([D2]‐ 4 ) to give the partially deuterated complex [Dn]‐ 4 . In addition, 4 undergoes β‐H abstraction between a hydride and an NMe2 ligand and forms a new complex [(Me2N){(Me3Si)2N}Ta(μ‐H)(μ‐N‐η2‐C,N‐CH2NMe)(μ‐C‐η2‐C,N‐CHSiMe2NSiMe3)Ta(NMe2)2] ( 5 ) with a cyclometalated, η2‐imine ligand. These results indicate that there are two simultaneous processes in [Dn]‐ 4 : 1) H–D exchange through σ‐bond metathesis, and 2) H?D elimination through β‐H abstraction (to give [Dn]‐ 5 ). Both 4 and 5 have been characterized by single‐crystal X‐ray diffraction studies.  相似文献   

14.
In the title compound, [Cu(CN)(C4H5N3)]n or [Cu(μ‐CN)(μ‐PyzNH2)]n (PyzNH2 is 2‐aminopyrazine), the CuI center is tetrahedrally coordinated by two cyanide and two PyzNH2 ligands. The CuI–cyano links give rise to [Cu–CN] chains running along the c axis, which are bridged by bidentate PyzNH2 ligands. The three‐dimensional framework can be described as being formed by two interpenetrated three‐dimensional honeycomb‐like networks, both made of 26‐membered rings of composition [Cu6(μ‐CN)2(μ‐PyzNH2)4].  相似文献   

15.
Thermolysis of [Cp*Ru(PPh2(CH2)PPh2)BH2(L2)] 1 (Cp*=η5‐C5Me5; L=C7H4NS2), with terminal alkynes led to the formation of η4‐σ,π‐borataallyl complexes [Cp*Ru(μ‐H)B{R‐C=CH2}(L)2] ( 2 a – c ) and η2‐vinylborane complexes [Cp*Ru(R‐C=CH2)BH(L)2] ( 3 a – c ) ( 2 a , 3 a : R=Ph; 2 b , 3 b : R=COOCH3; 2 c , 3 c : R=p‐CH3‐C6H4; L=C7H4NS2) through hydroboration reaction. Ruthenium and the HBCC unit of the vinylborane moiety in 2 a – c are linked by a unique η4‐interaction. Conversions of 1 into 3 a – c proceed through the formation of intermediates 2 a – c . Furthermore, in an attempt to expand the library of these novel complexes, chemistry of σ‐borane complex [Cp*RuCO(μ‐H)BH2L] 4 (L=C7H4NS2) was investigated with both internal and terminal alkynes. Interestingly, under photolytic conditions, 4 reacts with methyl propiolate to generate the η4‐σ,π‐borataallyl complexes [Cp*Ru(μ‐H)BH{R‐C=CH2}(L)] 5 and [Cp*Ru(μ‐H)BH{HC=CH‐R}(L)] 6 (R=COOCH3; L=C7H4NS2) by Markovnikov and anti‐Markovnikov hydroboration. In an extension, photolysis of 4 in the presence of dimethyl acetylenedicarboxylate yielded η4‐σ,π‐borataallyl complex [Cp*Ru(μ‐H)BH{R‐C=CH‐R}(L)] 7 (R=COOCH3; L=C7H4NS2). An agostic interaction was also found to be present in 2 a – c and 5 – 7 , which is rare among the borataallyl complexes. All the new compounds have been characterized in solution by IR, 1H, 11B, 13C NMR spectroscopy, mass spectrometry and the structural types were unequivocally established by crystallographic analysis of 2 b , 3 a – c and 5 – 7 . DFT calculations were performed to evaluate possible bonding and electronic structures of the new compounds.  相似文献   

16.
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   

17.
Unexpected Reduction of [Cp*TaCl4(PH2R)] (R = But, Cy, Ad, Ph, 2,4,6‐Me3C6H2; Cp* = C5Me5) by Reaction with DBU – Molecular Structure of [(DBU)H][Cp*TaCl4] (DBU = 1,8‐diazabicyclo[5.4.0]undec‐7‐ene) [Cp*TaCl4(PH2R)] (R = But, Cy, Ad, Ph, 2,4,6‐Me3C6H2 (Mes); Cp* = C5Me5) react with DBU in an internal redox reaction with formation of [(DBU)H][Cp*TaCl4] ( 1 ) (DBU = 1,8‐diazabicyclo[5.4.0]undec‐7‐ene) and the corresponding diphosphane (P2H2R2) or decomposition products thereof. 1 was characterised spectroscopically and by crystal structure determination. In the solid state, hydrogen bonding between the (DBU)H cation and one chloro ligand of the anion is observed.  相似文献   

18.
In the title compound, azido‐2κN‐bis­[μ‐(1η5:2κP)‐di­phenyl­phosphino­cyclo­penta­dienyl][2(η5)‐penta­methyl­cyclo­penta­di­enyl]­iron(III)­rhodium(III) hexa­fluoro­phosphate, [{Rh(C10H15)(N3)}{Fe(μ‐C17H14P)2}]PF6 or [FeRh(C10H15)(μ‐C17H14P)2(N3)]PF6, the coordination sphere of RhIII can be described as pseudo‐tetrahedral, composed of two P atoms from a 1,1′‐bis­(di­phenyl­phosphino)­ferrocene (dppf) ligand, an azido N atom and the centroid of the ring of a C5Me5 (Cp*) ligand. The two cyclo­penta­dienyl rings in the dppf moiety adopt an eclipsed conformation. The Rh⋯Fe distance is 4.340 (2) Å.  相似文献   

19.
The reactions of (R)‐ and (S)‐4‐(1‐carboxyethoxy)benzoic acid (H2CBA) with 1,3‐bis(2‐methyl‐1H‐imidazol‐1‐yl)benzene (1,3‐BMIB) ligands afforded a pair of homochiral coordination polymers (CPs), namely, poly[[[μ‐1,3‐bis(2‐methyl‐1H‐imidazol‐1‐yl)benzene][μ‐(S)‐4‐(1‐carboxylatoethoxy)benzoato]zinc(II)] monohydrate], {[Zn(C10H8O5)(C14H14N4)]·H2O}n or {[Zn{(S)‐CBA}(1,3‐BMIB)]·H2O}n ( 1‐L ), and poly[[[μ‐1,3‐bis(2‐methyl‐1H‐imidazol‐1‐yl)benzene][μ‐(R)‐4‐(1‐carboxylatoethoxy)benzoato]zinc(II)] monohydrate] ( 1‐D ). Three kinds of helical chains exist in compounds 1‐D and 1‐L , which are constructed from ZnII atoms, 1,3‐BMIB ligands and/or CBA2? ligands. When the as‐synthesized crystals of 1‐L and 1‐D were further heated in the mother liquor or air, poly[[μ‐1,3‐bis(2‐methyl‐1H‐imidazol‐1‐yl)benzene][μ‐(S)‐4‐(1‐carboxylatoethoxy)benzoato]zinc(II)], [Zn(C10H8O5)(C14H14N4)]n or [Zn{(S)‐CBA}(1,3‐BMIB)]n ( 2‐L ), and poly[[μ‐1,3‐bis(2‐methyl‐1H‐imidazol‐1‐yl)benzene][μ‐(R)‐4‐(1‐carboxylatoethoxy)benzoato]zinc(II)] ( 2‐D ) were obtained, respectively. The single‐crystal structure analysis revealed that 2‐L and 2‐D only contained one type of helical chain formed by ZnII atoms and 1,3‐BMIB and CBA2? ligands, which indicated that the helical chains were reconstructed though solid‐to‐solid transformation. This result not only means the realization of helical transformation, but also gives a feasible strategy to build homochiral CPs.  相似文献   

20.
A series of bis‐pentamethylcyclopentadienyl‐supported Dy complexes containing different ancillary ligands were synthesized and characterized. Magnetic studies showed that 1 Dy [Cp*2DyCl(THF)], 1 Dy’ [Cp*2DyCl2K(THF)]n, 2 Dy [Cp*2DyBr(THF)], 3 Dy [Cp*2DyI(THF)] and 4 Dy [Cp*2DyTp] (Tp=hydrotris(1‐pyrazolyl)borate) were single‐ion magnets (SIMs). The 1D dysprosium chain 1 Dy’ exhibited a hysteresis at up to 5 K. Furthermore, 3 Dy featured the highest energy barrier (419 cm?1) among the complexes. The effects of ancillary ligands on single‐ion magnetic properties were studied by experimental, ab initio calculations and electrostatic analysis methods in detail. These results demonstrated that the QTM rate was strongly dependent on the ancillary ligands and that a weak equatorial ligand field could be beneficial for constructing Dy‐SIMs.  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