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 140 毫秒
1.
Reactions of PdRR′(η1-dppm)2 (R = R′= C6F5 or C6Cl5; R = C6F5, R′= Cl; dppm = Ph2PCH2PPh2) with the gold derivatives ClAu(tht), C6F5Au(tht), (C6F5)3Au(tht) or O3ClOAuPPh3 (tht = tetrahydrothiophen) in appropriate ratios yield the bi- or tri-nuclear complexes PdRR′(dppm)2AuCl, PdRR′(dppm)2Au(C6F5); PdRR′(dppm)2Au(C6F5)3; PdRR′(dppmAuCl)2; PdRR′(dppmAuC6F5)2; PdRR′[dppmAu(C6F5)3]2, [PdRR′(dppm)2Au]X (X = ClO4 or BPh4); [PPh3Au(dppm)Pd(C6F5)2(dppm)AuCl]ClO4 or [PPh3 Au(dppm)Pd(C6F5)2(dppm)Au(C6F5)3]ClO4. The structure of trans-Pd(C6F5)2[dppmAu(C6F5)]2 has been determined by X-ray diffraction.  相似文献   

2.
The synthesis of the germylene phosphane adduct (C2F5)2Ge?PMe3 is described. Starting from (C2F5)3GeH in an excess of PMe3, heating was applied, whereupon reductive elimination of C2F5H occurred. The molecular structure was ascertained by X‐ray diffraction and compared with information obtained by quantum chemical methods. The ligand properties were derived by studying the IR spectrum of the nickel(0) complex [Ni(CO)3{Ge(C2F5)2(PMe3)}] in the CO region. (C2F5)2Ge?PMe3 turned out to be a π‐accepting ligand comparable to PMe3, in terms of Tolman's electronic parameter. Furthermore a [2+4] cycloaddition reaction with 2,3‐dimethyl‐1,3‐butadiene, and σ‐bond insertion reactions were recorded. Activation of the C?Cl bond in dichloromethane gives rise to the formation of the phosphonium ylide complex [(C2F5)2Cl2Ge‐CH2PMe3], which was fully characterized by X‐ray diffraction.  相似文献   

3.
Two calixarene‐based bis‐alkynyl‐bridged AuI isonitrile complexes with two different crown ether pendants, [{calix[4]arene‐(OCH2CONH‐C6H4C≡C)2}{Au(CNR)}2] (R=benzo[15]crown‐5 ( 1 ); R=benzo[18]crown‐6 ( 2 )), together with their related crown‐free analogue 3 (R=C6H3(OMe)2‐3,4) and a mononuclear gold(I) complex 4 with benzo[15]crown‐5 pendant, have been designed and synthesized, and their photophysical properties have been studied. The X‐ray structure of the ligand, calix[4]arene‐(OCH2CONH‐C6H4C?CH)2 has been determined. The cation‐binding properties of these complexes with various metal ions have been studied using UV/Vis, emission, 1H NMR, and ESI‐MS techniques, and DFT calculations. A new low‐energy emission band associated with Au???Au interaction could be switched on upon formation of the metal ion‐bound adduct in a sandwich fashion.  相似文献   

4.
Structures of New Bis(pentafluorophenyl)halogeno Mercurates [{Hg(C6F5)2}3(μ‐X)] (X = Cl, Br, I) From the reactions of [PNP]Cl or [PPh4]Y (Y = Br, I) with Hg(C6F5)2 crystals of the composition [Cat][{Hg(C6F5)2}3X] (Cat = PNP, X = Cl ( 1 ); Cat = PPh4, X = Br ( 2 ), I ( 3 )) are formed. 1 crystallizes in the triclinic space group P1¯, 2 and 3 crystallize isotypically in the monoclinic space group C2/c. In the crystals the halide anions are surrounded by three Hg(C6F5)2 molecules. The reaction of [PPh4]Br with Hg(C6F5)2 under slightly changed conditions gives the compound [PPh4]2[{Hg(C6F5)2}3(μ‐Br)][{Hg(C6F5)2}2(μ‐Br)] ( 4 ).  相似文献   

5.
When C6Cl5AuAsPh3 reacts with halogens, oxidation of the gold(I) complex and formation of X2Au(C6Cl5)AsPh3 (X = Cl, Br, I) take place. However, when C6F5AuAsPh3 reacts with halogens, oxidation is only observed in the case of Cl2, whilst I2 (totally) and Br2 (partially) split the AuC bond. This behaviour is contrary to that observed with C6F5AuPPh3 and halogens, where the tendency to split the AuC bond decreases in the sequence Cl>B>I.  相似文献   

6.
The neutral compounds [Pt(bzq)(CN)(CNR)] (R=tBu ( 1 ), Xyl ( 2 ), 2‐Np ( 3 ); bzq= benzoquinolate, Xyl=2,6‐dimethylphenyl, 2‐Np=2‐napthyl) were isolated as the pure isomers with a trans‐Cbzq,CNR configuration, as confirmed by 13C{1H} NMR spectroscopy in the isotopically marked [Pt(bzq)(13CN)(CNR)] (R=tBu ( 1′ ), Xyl ( 2′ ), 2‐Np ( 3′ )) derivatives (δ13CCN≈110 ppm; 1J(Pt,13C)≈1425 Hz]. By contrast, complex [Pt(bzq)(C≡CPh)(CNXyl)] ( 4 ) with a trans‐Nbzq,CNR configuration, has been selectively isolated from [Pt(bzq)Cl(CNXyl)] (trans‐Nbzq,CNR) using Sonogashira conditions. X‐ray diffraction studies reveal that while 1 adopts a columnar‐stacked chain structure with Pt–Pt distances of 3.371(1) Å and significant π???π interactions (3.262 Å), complex 2 forms dimers supported only by short Pt???Pt (3.370(1) Å) interactions. In complex 4 the packing is directed by weak bzq???Xyl and bzq???C≡E (C, N) interactions. In solid state at room temperature, compounds 1 and 2 both show a bright red emission (?=42.1 % 1 , 57.6 % 2 ). Luminescence properties in the solid state at 77 K and concentration‐dependent emission studies in CH2Cl2 at 298 K and at 77 K are also reported for 1 , 1·CHCl3 , 2 , 2' , 2·CHCl3 , 3 , 4 .  相似文献   

7.
Metallophilic Au???Ag interactions exist in concentrated solutions as well as in the 2D polymeric solid‐state structure of [AuAg3(C6F5)(CF3CO2)3(CH2PPh3)]n (see picture for the asymmetric unit) according to EXAFS and pulsed‐gradient spin‐echo NMR spectroscopy measurements. Calculations support both the emission assignments and the existence of the metallophilic interactions in solution.

  相似文献   


8.
Structures of the l,3,5-Trisilacyclohexane-Iron Dicarbonyl-cyclopentadienyl Complexes and C3H6Si3Cl5Fe(CO)2πcp and C3H6Si3Cl4(Fe(CO2)πcp)2 Trisilapentachlorocyclo-hexyl-dicarbonylcyclopentadienyliron C3H6Si3Cl5Fe(CO)2πcp 1 and Trisilatetrachlorocyclohexyl-bis(dicarboncyclopentadienyliron)C3H6Si3Cl4(Fe(CO)2πcp)2 2 are 1,3,5-Trisilacyclohexane complexes substituted by dicarbonylcyclopentadienyliron at one and two silicon atoms of the six-membered ring, respectively. The crystal and molecular structures were determined from single crystals ( 1 ; space group P21/a (No. 14); a = 1100.5 pm; b = 2033.9 pm; c = 843.3pm; β = 98.58°; Z = 4; MoKα-radiation; 3142h k l; R = 0.036. 2 ; space group P1 ; (No. 2); a = 1231.1 pm; b = 1267.3 pm; c = 1045.9 pm; α = 113.23°; β = 83.93°; γ = 115.00°; Z = 2; Mokα-radiation; 4196 h k 1; R = 0.065). In both complexes the six-membered rings of the carbosilane ligands are in skew-boat conformation. The bond lengths Fe? Si are 226.4 pm and 228.1 pm, respectively. The distances Si? C and Si? Cl are 186 pm and 206 pm in 1 and 187 pm and 209 pm in 2 . Their different lengths depend on the position in the ligand system and can be explained with the concept of bond orders.  相似文献   

9.
Low‐temperature (200 K) protonation of [Mo(CO)(Cp*)H(PMe3)2] ( 1 ) by Et2O ? HBF4 gives a different result depending on a subtle solvent change: The dihydrogen complex [Mo(CO)(Cp*)(η2‐H2)(PMe3)2]+ ( 2 ) is obtained in THF, whereas the tautomeric classical dihydride [Mo(CO)(Cp*)(H)2(PMe3)2]+ ( 3 ) is the only observable product in dichloromethane. Both products were fully characterised (νCO IR; 1H, 31P, 13C NMR spectroscopies) at low temperature; they lose H2 upon warming to 230 K at approximately the same rate (ca. 10?3 s?1), with no detection of the non‐classical form in CD2Cl2, to generate [Mo(CO)(Cp*)(FBF3)(PMe3)2] ( 4 ). The latter also slowly decomposes at ambient temperature. One of the decomposition products was crystallised and identified by X‐ray crystallography as [Mo(CO)(Cp*)(FH???FBF3)(PMe3)2] ( 5 ), which features a neutral HF ligand coordinated to the transition metal through the F atom and to the BF4? anion through a hydrogen bond. The reason for the switch in relative stability between 2 and 3 was probed by DFT calculations based on the B3LYP and M05‐2X functionals, with inclusion of anion and solvent effects by the conductor‐like polarisable continuum model and by explicit consideration of the solvent molecules. Calculations at the MP4(SDQ) and CCSD(T) levels were also carried out for calibration. The calculations reveal the key role of non‐covalent anion–solvent interactions, which modulate the anion–cation interaction ultimately altering the energetic balance between the two isomeric forms.  相似文献   

10.
Double chloride abstraction of Cp*AsCl2 gives the dicationic arsenic species [(η5‐Cp*)As(tol)][B(C6F5)4]2 ( 2 ) (tol=toluene). This species is shown to exhibit Lewis super acidity by the Gutmann–Beckett test and by fluoride abstraction from [NBu4][SbF6]. Species 2 participates in the FLP activation of THF affording [(η2‐Cp*)AsO(CH2)4(THF)][B(C6F5)4]2 ( 5 ). The reaction of 2 with PMe3 or dppe generates [(Me3P)2As][B(C6F5)4] ( 6 ) and [(σ‐Cp*)PMe3][B(C6F5)4] ( 7 ), or [(dppe)As][B(C6F5)4] ( 8 ) and [(dppe)(σ‐Cp*)2][B(C6F5)4]2 ( 9 ), respectively, through a facile cleavage of C?As bonds, thus showcasing unusual reactivity of this unique As‐containing compound.  相似文献   

11.
Syntheses of Oxovanadium(V) Halide Complexes Stabilized with Tripodal Oxygen Ligands LR = [η5‐(C5H5)Co{PR2(O)}3], R = OMe, OEt The sodium salts of the tripodal oxygen ligands LR = [η5‐(C5H5)Co{PR2(O)}3] (R = OMe, OEt) react with the oxovanadium halides V(O)F3 and V(O)Cl3 to yield deep red compounds of the type [V(O)X2LR]. Halide exchange reactions with [V(O)Cl2LOMe] und [V(O)F2LOMe] aiming at the preparation of the analogous bromide complex [V(O)Br2LOMe] led to the isomer [VO(LOMe)2][V(O)Br4]. The crystal structure of [V(O)Cl2LOMe] has been determined by single crystal x‐ray diffraction. The compound crystallizes in the monoclinic space group P21/n with a = 9.6332(8), b = 15.0312(11) and c = 15.3742(12)Å, β = 100.181(8)°. The coordination around vanadium is distorted octahedral.  相似文献   

12.
The resonance character of Cu/Ag/Au bonding is investigated in B???M?X (M=Cu, Ag, Au; X=F, Cl, Br, CH3, CF3; B=CO, H2O, H2S, C2H2, C2H4) complexes. The natural bond orbital/natural resonance theory results strongly support the general resonance‐type three‐center/four‐electron (3c/4e) picture of Cu/Ag/Au bonding, B:M?X?B+?M:X?, which mainly arises from hyperconjugation interactions. On the basis of such resonance‐type bonding mechanisms, the ligand effects in the more strongly bound OC???M?X series are analyzed, and distinct competition between CO and the axial ligand X is observed. This competitive bonding picture directly explains why CO in OC???Au?CF3 can be readily replaced by a number of other ligands. Additionally, conservation of the bond order indicates that the idealized relationship bB???M+bMX=1 should be suitably generalized for intermolecular bonding, especially if there is additional partial multiple bonding at one end of the 3c/4e hyperbonded triad.  相似文献   

13.
Te(C6F5)4 was prepared from the reactions of TeCl4 or Te(C6F5)2Cl2 with Grignard reagents or AgC6F5 in moderate to good yields. Substitution reactions with Me3SiX (X = Cl, Br, OSO2CF3), with equimolar amounts of Br2, with AgNO3 and with H[BF4] or BF3·OEt2 yielded the Te(C6F5)3X derivatives (X = Cl, Br, OSO2CF3, NO3, BF4). Oxidation reactions of Cd, Hg, and Pd0 complexes led to Te(C6F5)2 and the corresponding bis(pentafluorophenyl) derivatives M(C6F5)2 (M = Cd, Hg, Pd) and with InBr to In(C6F5)2Br. From very slow hydrolysis of Te(C6F5)4 the oxide Te(C6F5)2O was prepared. The thermal decomposition, the NMR and mass spectra of the partially new compounds are discussed. The crystal structures of Te(C6F5)3Br (monoclinic, P21/a, Z = 4), [Te(C6F5)3][OSO2CF3] (monoclinic, P21/n, Z = 16) and [Te(C6F5)2O]2 (triclinic, P1¯, Z = 2) were determined.  相似文献   

14.
Reaction of TlR2X, TlX3 and [TlX4? with RLi ( R = C6F5 or C6Cl5) leads to derivatives containing anions of the types [TlR4]?, [TlR2R′2]? or [TlR6]3?. Reactions of TlCl3 with [TlR4]? lead to [(μ-Cl)(TlR2Cl)2]? (R = C6F5) or [TlRCl3]? (R = C6Cl5) while addition of X? (X = Br? or SCN?) to Tl(C6Cl5)3 gives [Tl- (C6Cl5)3X]?. All the novel anions were isolated as salts of bulky cations (Me4N, Bu4N, PPN or Ph3BzP).  相似文献   

15.
Seven E[Cu(OR)2] copper(I) complexes (E=K+, {K(18C6)}+ (18C6=[18]crown‐6), or Ph4P+; R=C4F9, CPhMeF2, and CMeMeF2) have been prepared and their reactivity with O2 studied. The K[Cu(OR)2] species react with O2 in a copper‐concentration‐dependent manner such that 2:1 and 3:1 Cu/O2 adducts are observed manometrically at ?78 °C. Analogous reactivity with O2 is not observed with the {K(18C6)}+ or Ph4P+ derivatives. Solution conductivity data demonstrate that these K[Cu(OR)2] complexes do not behave as 1:1 electrolytes in solution. The K+ ions induce aggregation of multiple [Cu(OR)2]? units through K???F/O interactions and thereby effect irreversible O2 reduction by multiple Cu centers. Bond valence analyses for the potassium cations confirm the dominance of the fluorine interactions in the coordination spheres of K+ ions. Intramolecular hydroxylation of ligand aryl and alkyl C? H bonds is observed. Nucleophilic reactivity with CO2 is observed for the oxygenated Cu complexes and a CuII carbonate has been isolated and characterized.  相似文献   

16.
New Syntheses and Crystal Structures of Bis(fluorophenyl) Mercury, Hg(Rf)2 (Rf = C6F5, 2, 3, 4, 6‐F4C6H, 2, 3, 5, 6‐F4C6H, 2, 4, 6‐F3C6H2, 2, 6‐F2C6H3) Bis(fluorophenyl) mercury compounds, Hg(Rf)2 (Rf = C6F5, C6HF4, C6H2F3, C6H3F2), are prepared in good yields by the reactions of HgF2 with Me3SiRf. The crystal structures of Hg(2, 3, 4, 6‐F4C6H)2 (monoclinic, P21/n), Hg(2, 3, 5, 6‐F4C6H)2 (monoclinic, C2/m), Hg(2, 4, 6‐F3C6H2)2 (monoclinic, P21/c) and Hg(2, 6‐F2C6H3)2 (triclinic, P1) are described.  相似文献   

17.
Reduction of various pentafluorophenylnickel(II) complexes in the presence of phosphines gives unstable nickel(I) compounds but Ni(C6F5)(CO)2(PPh3)2 is isolated in the presence of CO. Similar NiR(CO)2(PPh3)2 (R = C6F5,C6Cl5, 2,3,5,6-C6Cl4H) are obtained by reaction of the halogenonickel(I) complex with MgRBr or LiR. Reduction of NiX2L2 in the presence of acetylenes gives [NiXL2]2(μ-PhCCR) (R = H, X = Cl and R = Ph, X = Cl, Br) when L = P-n-Bu3 but only NiX(PPh3)3 are recovered when L = PPh3. No reaction with the alkyne is observed for [NiX(PPh3)2]n but [NiCl(PPh3)]n reacts with RCCR′ to give paramagnetic NiCl(PPh3)(CRCR′) (R = Ph, R′= H, COOEt), diamagnetic [NiCl(PPh3)]2(μ-PhCCPh) and cyclotrimerization when R = R′ = COOMe. Chemical and structural behaviour of the new nickel(I) complexes is described.  相似文献   

18.
A representative series of diphosphine monophosphonium salts [1‐Ph2P(C10H6)‐8‐PRPh2]+X ( 2 b : R = H, X = CF3SO3; 4 : R = Me, X = CF3SO3; 5 : R = C6H5CH2 = Bn, X = Br) has been prepared by treatment of 1,8‐bis(diphenylphosphino)naphthalene (dppn, 1 ) with stoichiometric amounts of HSO3CF3 or CH3SO3CF3 in CH2Cl2 at +20 °C and with C6H5CH2Br in toluene at +80 °C. Their X‐ray crystal structures show that there is no evidence for dative P → P+ interactions. Instead, steric repulsion deflects the substituent groups to opposite faces of the naphthalene plane [splay angles: +11.4° ( 2 b ), +13.6° ( 4 ); +16.7° ( 5 )]. In solution 2 b , 4 , and 5 were dynamic according to 31P, 13C, and 1H NMR spectroscopy. The fluxionality of 2 b involves rapid intramolecular proton exchange between the two phosphorus atoms, which slows down at low temperature, whereas the dynamic behaviour of 4 and 5 is interpreted in terms of hindered rotation of the bulky RPh2P+ groups (R = Me or Bn) about the P–C(naphthyl) bond. Treatment of 1,8‐bis(diphenylphosphoryl)naphthalene (dppnO2, 6 ) with HSO3CF3 gave the protonated bis(phosphine oxide), as the triflate salt, dppnO2H+ CF3SO3 ( 7 ). The X‐ray structure analysis of 7 revealed a highly strained molecule (P1…P2 365.5 pm) in which the P=O bonds point to the same face of the naphthalene plane to accommodate the proton. All isolated compounds were characterised by a combination of 31P, 1H, and 13C NMR spectroscopy, IR spectroscopy ( 7 ), mass spectrometry and elemental analysis.  相似文献   

19.
Co(CH3)(PMe3)4 forms 100 % regioselectively with (2‐(2‐diphenylphosphanyl)phenyl)‐1,3‐dioxalane and 2‐diphenylphosphanyl‐pyridine, by elimination of methane, the four‐membered metallacycles Co{(C3O2HC6H3)P(C6H5)2}(PMe3)3 ( 1 ) and Co{(CNC4H3)P(C6H5)2}(PMe3)3 ( 4 ). The regioselectivity is independent of the steric requirement of the ortho substituent in the 2‐diphenylphosphanylaryl‐ligands. Oxidative addition with iodomethane transforms 1 and 4 into octahedral, diamagnetic low‐spin d6 complexes Co(CH3)I‐{(C3O2HC6H3)P(C6H5)2}(PMe3)2 ( 2 ) and Co(CH3)I‐{(CNC4H3)P(C6H5)2}(PMe3)2 ( 5 ). Under an atmosphere of carbon monoxide, insertion into the Co‐C bond results in ring expansion by forming the new assembled phosphanylbenzoyl complexes Co{(C4O3HC6H3)‐P(C6H5)2}CO(PMe3)2 ( 3 ) and Co{(OCNC4H3)P(C6H5)2}CO(PMe3)2 ( 6 ). The three different types of cobaltacycles are supported by X‐ray diffraction of 1 , 3 , 5 and 6 .  相似文献   

20.
The cyclopentadienylcobalt(I) compounds C5H5Co(PMe3)P(OR)3 (R = Me, Et, Pri) and C5H5Co(C2H4)L (L = PMe3, P(OMe)3, CO) are prepared by ligand substitution starting from C5H5Co(PMe3)2 and C5H5Co(C2H4)2. Whereas the reaction of C5H5Co(PMe3)P(OMe)3 with CH2Br2 mainly gives [C5H5CoBr(PMe3)P(OMe)3]Br, the dihalogenocobalt(III) complexes C5H5CoX2(PMe3) (X = Br, I) are obtained from C5H5Co(CO)PMe3 and CH2X2. Treatment of C5H5Co(CO)PMe3 or C5H5Co(C2H4)PMe3 with CH2ClI at low temperatures produces a mixture of C5H5CoCH2Cl(PMe3)I and C5H5CoCl(PMe3)I, which can be separated due to their different solubilities. The same reaction in the presence of ligand L gives the carbenoidcobalt(III) compounds [C5H5CoCH2Cl(PMe3)L]PF6 in nearly quantitative yields. If NEt3 is used as the Lewis base, the ylide complexes [C5H5Co(CH2PMe3)(PMe3)X]PF6 (X = Br, I) are obtained. The PF6 salts of the dications [C5H5Co(CH2PMe3)(PMe3)L]2+ (L = PMe3, P(OMe)3, CNMe) and [C5H5Co(CH2PMe3)(P(OMe)3)2]2+ are prepared either from [C5H5Co(CH2PMe3)(PMe3)X]+ and L, or more directly from C5H5Co(CO)PMe3, CH2X2 and PMe3 or P(OMe)3, respectively. The synthesis of C5H5CoCH2OMe(PMe3)I is also described.  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