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 78 毫秒
1.
Addition of PR3 (R=Ph or OPh) to [Cu(η2‐Me6C6)2][PF6] results in the formation of [(η6‐Me6C6)Cu(PR3)][PF6], the first copper–arene complexes to feature an unsupported η6 arene interaction. A DFT analysis reveals that the preference for the η6 binding mode is enforced by the steric clash between the methyl groups of the arene ligand and the phenyl rings of the phosphine co‐ligand.  相似文献   

2.
Water-soluble functionalized bis(phosphine) ligands L (ah) of the general formula CH2(CH2PR2)2, where for a: R = (CH2)6OH; bg: R = (CH2)nP(O)(OEt)2, n = 2–6 and n = 8; h: R = (CH2)3NH2 ( Scheme 1), have been prepared photochemically by hydrophosphination of the corresponding 1-alkenes with H2P(CH2)3PH2. Water-soluble palladium complexes cis-[Pd(L)(OAc)2] (18) were obtained by the reaction of Pd(OAc)2 with the ligands ah in a 1:1 mixture of dichloromethane:acetonitrile. The water-soluble phosphine ligands and their palladium complexes were characterized by IR, 1H and 31P NMR. A crystallographic study of complex 1 shows that the Pd(II) ion has a square planar coordination sphere in which the acetate ligands and the diphosphine ligand deviate by less than 0.12 Å from ideal planar.  相似文献   

3.
In an effort to find simple and common single-source precursors for palladium sulfide nanostructures, palladium(II) complexes, [Pd(S2X)2] (X = COMe (1), COiPr (2)) and η3-allylpalladium complexes with xanthate ligands, [(η3-CH2C(CH3)CR2)Pd(S2X)] (R = H, X = COMe (3); R = H, X = COEt (4); R = H, X = COiPr (5); R = CH3, X = COMe (6)), have been investigated. The crystal structures of [Pd(S2X)2] (X = COMe (1), CoiPr (2)) and [(η3-CH2C(CH3)CH2)Pd(S2COMe)] (3) have been established by single crystal X-ray diffraction analysis. The complexes, 1, 2 and 3 all contain a square planar palladium(II) centre. In the allyl complex 3, this is defined by the two sulfurs of the xanthate and the outer carbons of the 2-methylallyl ligand, while in the complexes, 1 and 2 it is defined by the four sulfur atoms of the xanthate ligand. Thermogravimetric studies have been carried out to evaluate the thermal stability of η3-allylpalladium(II) analogues. The complexes are useful precursors for the growth of nanocrystals of PdS either by furnace decomposition or solvothermolysis in dioctyl ether. The solvothermal decomposition of complexes in dioctyl ether gives a new metastable phase of PdS which can be transformed to the more stable tetragonal phase at 320 °C. The nanocrystals obtained have been characterized by PXRD, SEM, TEM and EDX.  相似文献   

4.
In this account, we focus on results from our laboratory to illustrate recent developments in various fields of organometallic chemistry. Studies on hemilabile P,N donor ligands and on the ion-pair behaviour of cationic Pd(II) complexes have led to the full characterization of complexes with η1-allyl ligands. This still rare bonding mode for the allyl ligand in palladium chemistry allows facile insertion of CO into the Pd-C σ-bond, in contrast to the situation in related η3-allyl Pd(II) complexes. In order to develop new homogeneous catalysts for the selective dimerization and oligomerization of ethylene, a range of Ni(II) complexes have been prepared with new chelating P,N ligands where P represents a phosphine, phosphinite or phosphonite donor group and N a pyridine or oxazoline moiety. Finally, we shall examine bottom-up approaches to the formation of new nanomaterials of magnetic or catalytic interest by covalent anchoring of metal complexes and clusters into mesoporous materials using functional phosphine or alkyne ligands containing an alkoxysilyl group.  相似文献   

5.
Summary.  Palladium(II) complexes of the general formula PdCl2 (PR3)2 with PR3 = { P(OPh)3}, P(O-4-MeC6H4)3, P(O-2-MeC6H4)3, and PPh2(OBu) were reduced by NEt3 in chloroform or benzene to Pd(0) complexes Pd(PR3)4 and Pd(PR3)x(NEt3) 4−x . The same reaction performed in the presence of air gave CH3CHO or CH3CH2CHO when NPr3 was used instead of NEt3. Pd(P(OPh)3)4 reacted with benzyl bromide affording the oxidative addition product cis-PdBr(CH2Ph)(P(OPh)3)2. The reaction of PdCl2(P(OPh)3)2 with benzyl bromide was observed only in the presence of NEt3, and a dimeric complex of [PdBr(CH2Ph)(P(OPh)3)]2 was identified as the reaction product. Both benzyl complexes reacted fast with CO (1 atm) to form acyl complexes exhibiting ν(CO) bands at 1709 and 1650 cm−1.  相似文献   

6.
The room temperature syntheses of new chelating acyl palladium(II) complexes, [Pd(μ-Cl)(C(O)C9H6N)]2 and [Pd(μ-Cl)(C(O)C6H4N(CH3)2]2, derived from quinoline-8-carbaldehyde and 2-(dimethylamino)banzaldehyde are described. These chloro bridged dimers may be cleaved with neutral phosphine and nitrogen ligands, L, to give the monomeric [PdCl(C(O)C9H6N)L] and [PdCl(C(O)C6H4N(CH3)2)L] compounds. 1H-, 13C- and 31P-NMR data for the new complexes are reported.  相似文献   

7.
Abstract

Dinuclear Pd(II) halides that contain bridging π-conjugated groups, trans,trans-[(PR3)2(X)Pd–Y–Pd(X)(PR3)2] (X?=?Br; YH2 = terpyridine, fluorenone, benzil, benzthiadiazole), were prepared by the oxidative addition of corresponding dihalo π-conjugated reagents to [Pd(styrene)(PR3)2]. Similar reactions involving dihalobenzil, dihalobithiophene, or dihaloterthiophene afforded dinuclear Pt(II) halides containing bridging π-conjugated groups. Additionally, when the dihalosilole derivatives {2,5-dibromo-1,1-dimethyl (or diphenyl)-3,4-diphenylsilole} reacted with [Pd(styrene)(PR3)2], mono or dinuclear Pd(II) complexes bearing a dimethyl (or diphenyl)-3,4-diphenylsilole group were obtained. π-Conjugation extension reactions of dinuclear bithiophene-bridged Pd(II) halides with HC≡C–R {R?=?SiPh3, C(O)OMe} in the presence of CuI and HNEt2 led to the unexpected formation of bis(acetylide) Pd(II) complexes of the form, [Pd(C≡C–R)2(PR3)2] and bithiophene. In contrast, treatment of the dinuclear Pd(II) halides with two equiv of organic isocyanide resulted in isocyanide insertion into the Pd???C bonds to afford π-conjugation-extended dinuclear Pd(II) compounds bearing a π-conjugated moiety.  相似文献   

8.
The complexes (η5-C5H5)Pd(η1-C5H5)PR3 which are prepared from [Cl(PR3)-Pd]2(μ-OCOCH3)2 and TlC5H5 are fluxional in solution. According to the 1H and 13C NMR spectra at various temperatures, two dynamic processes occur. The process with the higher activation energy is a π/σ (η51) exchange of the two different cyclopentadienyl ligands, whereas the second one with the lower activation energy presumably is a metallotropic rearrangement (1,2-shift). The coalescence temperature for the η51 exchange depends on the size of the phosphine. The X-ray structural analysis of (C5H5)2PdPPri3 proves that it exists as a “frozen” η5 + η1 structure in the crystal with the palladium approximately in a square-planar coordination. The η5-bonded cyclopentadienyl ring shows some unusual bonding patterns which are obviously electronic in nature. EHT-MO calculations for (η5-C5H5)PdCH3(PH3) indicate that in this model system alternating CC distances in the ring and a stronger bond of the metal to one of the five carbon atoms of the C5H5 ligand are to be expected. The calculations suggest that in similar complexes possessing a six-electron donor ligand like C5H5? and a metal fragment which is isolobal to PdCH3(PH3)+, analogous distortions should be observed. Some reactions of the compounds (η5-C5H5)Pd(η1-C5H5)PR3 are described.  相似文献   

9.
The dinuclear hydroxo complex [{Pd(μ-OH)(Phox)}2] (I) (Phox = 2-(2-oxazolinyl)phenyl) reacts in a 1:2 molar ratio with several imidate ligands to yield new cyclometallated palladium complexes [{Pd(μ-NCO)(Phox)}2] containing asymmetric imidate –NCO– bridging units. [–NCO– = succinimidate (succ) (1), phtalimidate (phtal) (2), maleimidate (mal) (3), 2,3-dibromomaleimidate (2,3-diBrmal) (4) and glutarimidate (glut) (5)]. The reaction of these complexes with tertiary phosphines provides novel mononuclear N-bonded imidate derivatives of the general formula [Pd(imidate)(Phox)(PR3)] [R = Ph (a), 4-F–C6H4 (b) or CH2CH2CN (c)]. The new complexes were characterized by partial elemental analyses and spectroscopic methods (IR, FAB, 1H, 13C and 31P). The single-crystal structures of compounds 4, 4a and 5a have been established.  相似文献   

10.
Supported Organometallic Complexes. IV. Structural Investigations on Neutral and Cationic (Ether-phosphane)palladium(II) Complexes . Reaction of the (ether-phosphane) ligands PhP(R)CH2—D ( 2a?c ) [D=CH2OCH3: R=Ph ( a ), (CH2)3Si(OCH3)3 ( b ), (CH2)3SiO3/2 ( b ′); D= R=(CH2)3Si(OCH3)3 ( c ), (CH2)3SiO3/2 ( c ′)] with Cl2Pd(COD) ( 1 ) results in the formation of Cl2Pd(P — O)2 ( 3a?c ). Cleavage of Cl? from 3 with AgSbF6 yields the cationic, monochelated complexes [ClPd(P — O)(P ∩ O)]+ ( 4 a—c ) (P — O: η1-P-coordinated; P ∩ O: η2-O ∩ P-coordinated). 4 a crystallizes in the monoclinic space group P21/c with the lattice constants a=1 062,4(2), b=1 912,2(4) und c=1 635,5(3) pm, β=101,22(3)° and Z=4 (R=0,0341; Rw=0,033). With water 3 b, c and 4 b, c are subjected to polycondensation to give the supported complexes 3 b′, c′, 4 b′, c ′. The structure 3 b′, c′, 4 b′, c ′ is investigated by comparison of their 31P CP MAS data with the 31P{1H} NMR spectra of their soluble precursors 3 b, c, 4 b, c . 13C CP MAS NMR spectra of 3 b′, c ′ and 4 b′, c ′ indicate η1-P- and η2-O ∩ P-coordination of the ligands. The polysiloxane network of 4 b ′ was inspected by contact time variation of the 29Si CP MAS NMR spectra and it appeared that 77% of the Si—O units are crosslinked, corresponding to a ratio T4:T3:T1=67:100:10.  相似文献   

11.
Reactions of pyrimidine‐2‐thione (HpymS) with PdII/PtIV salts in the presence of triphenyl phosphine and bis(diphenylphosphino)alkanes, Ph2P‐(CH2)m‐PPh2 (m = 1, 2) have yielded two types of complexes, viz. a) [M(η2‐N, S‐ pymS)(η1‐S‐ pymS)(PPh3)] (M = Pd, 1 ; Pt, 2 ), and (b) [M(η1‐S‐pymS)2(L‐L)] {L‐L, M = dppm (m = 1) Pd, 3 ; Pt, 4 ; dppe (m = 2), Pd, 5 ; Pt, 6 }. Complexes have been characterized by elemental analysis (C, H, N), NMR spectroscopy (1H, 13C, 31P), and single crystal X‐ray crystallography ( 1 , 2 , 4 , and 5 ). Complexes 1 and 2 have terminal η1‐S and chelating η2‐N, S‐modes of pymS, while other Pd/Pt complexes have only terminal η1‐S modes. The solution state 31P NMR spectral data reveal dynamic equilibrium for the complexes 3 , 5 and 6 , whereas the complexes 1 , 2 and 4 are static in solution state.  相似文献   

12.
The reaction of α‐keto‐stabilized diphosphine ylides [Ph2P(CH2)nPPh2═C(H)C(O)C6H4p‐CN] (n = 1 (Y1); n = 2 (Y2)) with dibromo(1,5‐cyclooctadiene) palladium(II)/platinum(II) complexes, [Pd/PtBr2(cod)], in equimolar ratio gave the new cyclometalated Pd(II) and Pt(II) complexes [Br2Pd(κ2‐Y1)] ( 1 ), [Br2Pt(κ2‐Y1)] ( 2 ), [Br2Pd(κ2‐Y2)] ( 3 ) and [Br2Pt(κ2‐Y2)] ( 4 ). These compounds were screened in a search for novel antibacterial agents and characterized successfully using Fourier transfer infrared and NMR (1H, 13C and 31P) spectroscopic methods. Also, the structures of complexes 1 and 2 were characterized using X‐ray crystallography. The results showed that the P,C‐chelated complexes 1 and 2 have structures consisting of five‐membered rings, while 3 and 4 have six‐membered rings, formed by coordination of the ligand through the phosphine group and the ylidic carbon atom to the metal centre. Also, a theoretical study of the structures of complexes 1 – 4 was conducted at the BP86/def2‐SVP level of theory. The nature of metal–ligand bonds in the complexes was investigated using energy decomposition analyses (EDA) and extended transition state combined with natural orbitals for chemical valence analyses. The results of EDA confirmed that the main portions of ΔEint, about 57–58%, in the complexes are allocated to ΔEelstat.  相似文献   

13.
A convenient synthesis of the easily handled, air stable methallyloxyphosphonium salt [CH2C(Me)CH2O-P(NMe2)3]+[(3,5-(CF3)2-C6H3)4B]5 is described. The utility of this reagent in the generation and stabilization of cationic η3-methallyl palladium complexes through oxidative addition reactions is illustrated by the preparation of the stable salts of [(η3-C4H7)Pd(NN)]+[(3,5-(CF3)2-C6H3)4B] from Pd2(dba)3. The molecular structure of one of them has been determined by a single-crystal X-ray diffraction.  相似文献   

14.
Several trans-hydridomethylbis(phosphine)-platinum(II) and -palladium(II) complexes have been made by the reaction: trans-M(H)Cl(PR3)2 + CH3MgBr → trans-M(CH3)(PR3)2 + MgClBr and their structures determined by 1H NMR and IR spectroscopy. The complexes in which M  Pt and R  Cy (cyclohexyl) or i-Pr (isopropyl) are very stable in the solid state and in solution, while the compounds in which M  Pt, R  Et (ethyl) and M  Pd, R  i-Pr slowly decompose either in the solid state or in solution. The compound in which M  Pd and R  Cy was not isolated but was identified in solution.  相似文献   

15.
Thallium [1-(p-tolylimino)-2-methylpropyl]cyclopentadienide, Tl[C5H4C(=NC6H4CH3)CH(CH3)2], was prepared and treatment of the salt with [{PdCl2(PREt2)}2] (R = Ph and Et) yielded mononuclear palladium(II) complexes, [Pd{η5-C5H4C(=NC6H4CH3)CH(CH3)2}Cl(PREt2)], with an imidoyl-substituted η5-cyclopentadienyl group. In addition, [Pd(η5-C5H4-COY)Cl(PPhEt2)] (Y = CH3 and OCH3) complexes were obtained from the sodium salts of their substituted cyclopentadienyl groups. These new compounds were characterized by means of 1H and 13C NMR and IR spectroscopy.  相似文献   

16.
Two phosphine ligands of [Pd(PPh3)4] were substituted by π(C?S) coordination of 4‐bromodithiobenzoic acid methyl ester resulting in complex 1 . The same ester, after alkylation, afforded the dicationic complex bis(μ‐methanethiolato)tetrakis(triphenylphosphine)dipalladium(2+) bis(tetrafluoroborate) ( 2 ) from the same palladium source. A related thiolato‐bridged complex, bis(μ‐methanethiolato)bis(1‐methylpyridin‐2(1H)‐ylidene)bis(triphenylphosphine)dipalladium(2+) bis(tetrafluoroborate) ( 4 ) and the trinuclear cluster tris(μ‐methanethiolato)tris(triphenylphosphine)tripalladium(+)(3Pd? Pd) ( 5 ) resulted from treatment of a known cationic pyridinylidene complex with MeSLi. The double oxidative substitution reaction of [Pd(PPh3)4] with 1,5‐dichloro‐9,10‐anthraquinone afforded trans‐dichloro[μ‐(9,10‐dihydro‐9,10‐dioxoanthracene‐1,5‐diyl)]tetrakis(triphenylphosphine)dipalladium ( 6 ). Some of these complexes could be fully characterized by 1H‐, 13C‐, and 31P‐NMR spectroscopy, mass spectrometry, and elemental analysis. The crystal and molecular structures of all of them, and of trans‐bis(1,3‐dihydro‐1,3‐dimethyl‐2H‐imidazol‐2‐ylidene)diiodopalladium ( 3 ), were determined by single‐crystal X‐ray diffraction.  相似文献   

17.
Carbonylation of the palladium complexes [PdCH3(PP′)Cl] (PP′ = 1a, 1b, 1c, 1d, 1e) and [PdCH3(PP′)(CH3CN)](OTf) was investigated by means of high-pressure NMR with the determination of the half-life times t1/2. The results were rationalized on the basis of the electronic properties of the diphosphines and the nature of the solvento ligand in the first coordination sphere. The crystal structures of the complexes [Pd(1b)Cl2] and [Pd(1b)(H2O)2](OTf)2 are described (1b = 1-(diphenylphosphinomethyl)-2-[bis(3- trifluoromethylphenyl)phosphinomethyl]benzene).  相似文献   

18.
Hydrocarbon‐bridged Metal Complexes. L Dicarbonyl Cyclopentadienyl Pyridoyl Iron Complexes as Ligands Dicarbonyl‐cyclopentadienyl‐2‐ and 3‐pyridoyl‐iron (L1, L2) and 2,6‐dicarbonyl‐pyridine‐bis(dicarbonyl‐cyclopentadienyl‐iron) (L3) function as ligands in metal complexes and the N,O‐chelates [(OC)4M(L1)] (M = Mo, W, 8 a, b ) and [(Ph3P)2Cu(L1)]+BF4 ( 9 ) were prepared. Monodentate coordination of L1 and L2 through the pyridine N‐atom occurs in the palladium(II) complexes [Cl2Pd(PnBu3)(L1)] ( 10 ), [Cl2Pd(PnBu3)(L2)] ( 11 ) and [Cl2Pd(L2)2] ( 12 ). Ligand L3 forms the O,N,O‐bis(chelate) [Cl2Zn(L3)] ( 13 ). The crystal and molecular structures of L1, 8 b (M = W), 9–11 and 13 were determined by X‐ray diffraction.  相似文献   

19.
We report the synthesis and characterization of a novel 4‐(dimethylamino)pyridinium‐substituted η3‐cycloheptatrienide–Pd complex which is free of halide ligands. Diacetonitrile{η3‐[4‐(dimethylamino)pyridinium‐1‐yl]cycloheptatrienido}palladium(II) bis(tetrafluoroborate), [Pd(C2H3N)2(C14H16N2)](BF4)2, was prepared by the exchange of two bromide ligands for noncoordinating anions, which results in the empty coordination sites being occupied by acetonitrile ligands. As described previously, exchange of only one bromide leads to a dimeric complex, di‐μ‐bromido‐bis({η3‐[4‐(dimethylamino)pyridinium‐1‐yl]cycloheptatrienido}palladium(II)) bis(tetrafluoroborate) acetonitrile disolvate, [Pd2Br2(C14H16N2)2](BF4)2·2CH3CN, with bridging bromide ligands, and the crystal structure of this compound is also reported here. The structures of the cycloheptatrienide ligands of both complexes are analogous to the dibromide derivative, showing the allyl bond in the β‐position with respect to the pyridinium substituent. This indicates that, unlike a previous interpretation, the main reason for the formation of the β‐isomer cannot be internal hydrogen bonding between the cationic substituents and bromide ligands.  相似文献   

20.
Pincer‐type palladium complexes are among the most active Heck catalysts. Due to their exceptionally high thermal stability and the fact that they contain PdII centers, controversial PdII/PdIV cycles have been often proposed as potential catalytic mechanisms. However, pincer‐type PdIV intermediates have never been experimentally observed, and computational studies to support the proposed PdII/PdIV mechanisms with pincer‐type catalysts have never been carried out. In this computational study the feasibility of potential catalytic cycles involving PdIV intermediates was explored. Density functional calculations were performed on experimentally applied aminophosphine‐, phosphine‐, and phosphite‐based pincer‐type Heck catalysts with styrene and phenyl bromide as substrates and (E)‐stilbene as coupling product. The potential‐energy surfaces were calculated in dimethylformamide (DMF) as solvent and demonstrate that PdII/PdIV mechanisms are thermally accessible and thus a true alternative to formation of palladium nanoparticles. Initial reaction steps of the lowest energy path of the catalytic cycle of the Heck reaction include dissociation of the chloride ligands from the neutral pincer complexes [{2,6‐C6H3(XPR2)2}Pd(Cl)] [X=NH, R=piperidinyl ( 1 a ); X=O, R=piperidinyl ( 1 b ); X=O, R=iPr ( 1 c ); X=CH2, R=iPr ( 1 d )] to yield cationic, three‐coordinate, T‐shaped 14e? palladium intermediates of type [{2,6‐C6H3(XPR2)2}Pd]+ ( 2 ). An alternative reaction path to generate complexes of type 2 (relevant for electron‐poor pincer complexes) includes initial coordination of styrene to 1 to yield styrene adducts [{2,6‐C6H3(XPR2)2}Pd(Cl)(CH2?CHPh)] ( 4 ) and consecutive dissociation of the chloride ligand to yield cationic square‐planar styrene complexes [{2,6‐C6H3(XPR2)2}Pd(CH2?CHPh)]+ ( 6 ) and styrene. Cationic styrene adducts of type 6 were additionally found to be the resting states of the catalytic reaction. However, oxidative addition of phenyl bromide to 2 result in pentacoordinate PdIV complexes of type [{2,6‐C6H3(XPR2)2}Pd(Br)(C6H5)]+ ( 11 ), which subsequently coordinate styrene (in trans position relative to the phenyl unit of the pincer cores) to yield hexacoordinate phenyl styrene complexes [{2,6‐C6H3(XPR2)2}Pd(Br)(C6H5)(CH2?CHPh)]+ ( 12 ). Migration of the phenyl ligand to the olefinic bond gives cationic, pentacoordinate phenylethenyl complexes [{2,6‐C6H3(XPR2)2}Pd(Br)(CHPhCH2Ph)]+ ( 13 ). Subsequent β‐hydride elimination induces direct HBr liberation to yield cationic, square‐planar (E)‐stilbene complexes with general formula [{2,6‐C6H3(XPR2)2}Pd(CHPh?CHPh)]+ ( 14 ). Subsequent liberation of (E)‐stilbene closes the catalytic cycle.  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