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 62 毫秒
1.
Several methods for the synthesis of the Pd38(CO)28L12 cluster (L = PEt3) by treatment of Pd10(CO)12L6 with CF3COOH-Me3NO, CF3COOH-H2O2, Pd(OAc)2-Me3NO, and Pd2(dba)3 mixtures (dba is dibenzylideneacetone) were proposed. The tri-n-butylphosphine analog, Pd38(CO)28(PBu3)12, was synthesized by the reaction of Pd10(CO)14(PBu3)4 with Me3NO. The reaction of Pd4(CO)5L4 with Pd2(dba)3 yields clusters with an icosahedral packing of the metal atoms, Pd34(CO)24L12 and Pd16(CO)13L9.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 167–170, January, 1995.  相似文献   

2.

Structural isomerism of the Pd4(L)4(RCO2)4 (L = CO, CH2, NO; R = CC13, CF3, CH3) complexes was studied in the framework of the density functional theory (DFT). Among the Pd4(CO)4(RCO2)4 and Pd4(CH2)4(RCO2)4 complexes the most stable were the isomers with alternate coordination of pairs of carbonyl and carboxylate ligands on the sides of a planar rectangular metal core. The isomers with the pairwise coordination of NO/RCO2 on one side of the metal core are the most stable between the Pd4(NO)4(RCO2)4 complexes. The features of mutual coordination of ligands in polynuclear complexes of palladium are clarified using the obtained results.

  相似文献   

3.
Palladium(O) carbonyl complexes, Pd(CO)(PPh3)3 Pd3(CO)3(PPh3)3 and Pd3(CO)3(PPh3)4, can conveniently be prepared by the reaction of (PPh3)2PdCl2 with carbon monoxide at room temperature in methanol/amine systems involving primary and secondary amines such as diethylamine and cyclohexylamine. These carbonyl complexes are interconvertible under suitable conditions. On the other hand, use of tertiary amine such as triethylamine and tri-n-butylamine in place of the above amines give selectively a carbomethoxy complex (PPh3)2PdCl(COOCH3).  相似文献   

4.
Interaction of the tetra-nuclear phosphinecarbonyl complex of zero-valent palladium, Pd4(CO)5(PEt3)4 (I), with a four-fold excess of 8-(α-bromomercuryethyl)quinoline gives a new neutral heteronuclear cluster Pd4Hg2Br2(CO)4(PEt3)4 (IV). The structure of (IV) was determined by an X-ray analysis. The molecule of (IV) contains a “butterfly” Pd42-CO)4(PEt3)4 moiety, whose triangular “wings” are capped by Hg atoms bonded also to bromo-ligands. The metal Hg2Pd4 polyhedron consists of two heteroatomic Pd3Hg tetrahedra with a common PdPd edge. The IR- and 31P-NMR spectra of (IV) were also studied.  相似文献   

5.
This research was an outgrowth of previous reactions with [Pd13Ni13(CO)34]4? which produced a tetragonal crystal form of Pd23(CO)20(PEt3)10 (1) that has the same cuboctahedral-based Pd23 framework with an identical number of PEt3 ligands but two fewer CO ligands than the monoclinic crystal form of Pd23(CO)22(PEt3)10 (3) originally reported from reactions with Pd10(CO)12(PEt3)6. A subsequent investigation presented herein to establish whether the carbonyl capacity is influenced by the nature of the phosphine ligands has led to syntheses of Pd23(CO) x (PR3)10 [R3=Et3 (1), Bu n 3 (4), and Me2Ph (5)] with 20 CO ligands (x=20) from corresponding Pd10(CO)12(PR3)6 precursors either by deligation with Pd(OAc)2, CF3CO2H, Ni(1,5-COD)2, [NMe4]2[Ni6(CO)12], or HCO2H or by spontaneous enlargement; yields varied from 15 to 79%. Although attempts to obtain the original Pd23(CO)22(PEt3)10 (3) were unsuccessful, a highly significant outcome was the isolation (one time) of another monoclinic crystal form possessing the triethylphosphine Pd23(CO) x (PEt3)10 cluster with 21 COs (2). Both the compositions and atomic arrangements for each of five Pd23 clusters [1a (solvated); 1b (unsolvated); 2, 4, and 5] were unambiguously established from low-temperature single-crystal CCD X-ray crystallographic determinations in accordance with their nearly identical IR carbonyl frequencies. Solution 31P{1H} NMR spectra of 1 and 4 at room temperature displayed three distinct signals with expected integral ratios of 2/4/4 that are consistent with the solid-state structures of Pd23(CO)20(PR3)10 [R3=Et3 (1), Bu n 3 (4)] remaining intact in solution. The metal-core geometries of all of these Pd23(CO) x (PR3)10 clusters, including the thermodynamically stable ones with 20 CO ligands and the kinetic products with additional CO ligands (x=21, 22), are essentially the same. The common Pd23 core may be best described as possessing a centered hexacapped cuboctahedral Pd19 kernel (alternatively denoted as a centered ν2 Pd19 octahedron) with four edge-connected exopolyhedral wingtip Pd(exo) atoms that reduce the pseudo metal-core symmetry from Oh to D2h. The 10 PR3 ligands are linked to the six tetracapped Pd(cap) and four edge-capped wingtip Pd(exo) atoms; the latter four Pd(exo) atoms are each composed of four trigonal-planar Pd(μ2-CO)2(PR3) units. These crystallographic results provide compelling geometrical evidence for a heretofore unknown stereochemical example involving variable carbonyl ligation (x=20, 21, 22) of a close-packed nanosized Pd n (CO) x (PR3) y cluster (in this case with identical PEt3 ligands) without significant changes being induced in either the overall metal-core architecture or steric dispositions of the same number of PR3 ligands. These experimental findings have particular relevance to the long-standing Muetterties cluster/surface science analogy in showing that the different number (as well as different modes) of carbonyl ligations observed in these large metal carbonyl clusters are directly related to pressure-induced dissociative/nondissociative migratory coverages in CO chemisorptions on metal surfaces. The observed expanded capacity of CO coordination on the same Pd23 polyhedron without notable changes in geometry is no doubt a consequence of its virtually nanosized metal-core architecture; distances between outermost centrosymmetrically related pairs of Pd(cap) and Pd(exo) atoms in the Pd23 framework are 0.8 and 0.9 nm, respectively. An electrochemical (CV) study revealed that 1 undergoes one quasi-reversible two-electron reduction to 1 2? (E1/2=?0.91 V) and two consecutive quasi-reversible one-electron oxidations to 1/1 + at E1/2=0.08 V and 1 +/1 2+ at E1/2=0.32 V (THF; Ag/AgCl as reference electrode). A stereochemical/electronic analysis with the isostructural Au2Pd21(CO)20(PEt3)10 analogue (9) and resulting implications are given.  相似文献   

6.
The formation of Pd(II)-containing and mixed Pd(II),Cu(II), Pd(II),Fe(III), and Pd(II),V(V) complexes with heteropolyanion PW9O9– 34was studied using 31P, 183W, 51V NMR, visible UV and IR spectroscopy, and the differentiating dissolution methods. In an aqueous solution and at optimal pH (3.7), the monometallic complexes [Pd3(PW9O34)2]12–and [Pd3(PW9O34)2Pd n O x H y ] q(n av= 3), the bimetallic complexes [Pd2Cu(PW9O34)2]12–, [Pd2Fe(PW9O34)2]11–, and [PdFe2(PW9O34)2]10–, and a mixture of the [Pd3(PW9O34)2Pd n O x H y ] q(n av 10) + [(VO)3(PW9O34)2]9–complexes are formed. The title complexes were isolated from solution as Cs+solid salts belonging to the same [M3(PW9O34)2] structural type.  相似文献   

7.
Complexation of 1,4‐phenylenebis(methylene) diisonicotinate, L1 , with cis‐protected PdII components, [Pd( L′ )(NO3)2], in an equimolar ratio yielded binuclear complexes, 1 a – d of [Pd2( L′ )2( L1 )2](NO3)4 formulation where L′ stands for ethylenediamine (en), tetramethylethylenediamine (tmeda), 2,2′‐bipyridine (bpy), and phenanthroline (phen). The combination of 4,4′‐bipyridine, L2 , with the cis‐protected PdII units is known to yield molecular squares, 2 a – d . However, 2 b – d coexist with the corresponding molecular triangles, 3 b – d . Combination of an equivalent each of the ligands L1 and L2 with two equivalents of cis‐protected PdII components in DMSO resulted in the D ‐shaped heteroligated complexes [Pd2( L′ )2( L1 )( L2 )](NO3)4, 4 a – d . Two units of the D ‐shaped complexes interlock, in a concentration dependent fashion, to form the corresponding [2]catenanes [Pd2( L′ )2( L1 )( L2 )]2(NO3)8, 5 a – d under aqueous conditions. Crystal structures of the macrocycle [Pd2(tmeda)2( L1 )( L2 )](PF6)4, 4 b′′ , and the catenane [Pd2(bpy)2( L1 )( L2 )]2(NO3)8, 5 c , provide unequivocal support for the proposed molecular architectures.  相似文献   

8.
This work demonstrates a new nonconventional ligand design, imidazole/pyridine‐based nonsymmetrical ditopic ligands ( 1 and 1 S ), to construct a dynamic open coordination cage from nonsymmetrical building blocks. Upon complex formation with Pd2+ at a 1:4 molar ratio, 1 and 1 S initially form mononuclear PdL4 complexes (Pd2+( 1 )4 and Pd2+( 1 S )4) without formation of a cage. The PdL4 complexes undergo a stoichiometrically controlled structural transition to Pd2L4 open cages ((Pd2+)2( 1 )4 and (Pd2+)2( 1 S )4) capable of anion binding, leading to turn‐on anion binding. The structural transitions between the Pd2L4 open cage and the PdL4 complex are reversible. Thus, stoichiometric addition (2 equiv) of free 1 S to the (Pd2+)2( 1 S )4 open cage holding a guest anion ((Pd2+)2( 1 S )4?G?) enables the structural transition to the Pd2+( 1 S )4 complex, which does not have a cage and thus causes the release of the guest anion (Pd2+( 1 S )4+G?).  相似文献   

9.
    
The reaction of the tetranuclear cluster Pd4(CO)4(OOCCF3)4 witho-nitrosotoluene afforded the Pd11-containing complex [o-(NO)(CH2)C6H4]2Pd2(μ-OOCCF3)2. The elimination of CO2 and the formation of organic products of transformation of tolylnitrene species (azotoluene, ditolylamine, and tolylisocyanate) were observed in the course of the reaction. The title complex was characterized by IR and1H NMR spectroscopy. Its structure was established by X-ray diffraction analysis. It was suggested that the reaction proceeds through intermediate formation of nitrene complexes. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 147–150, January, 2000.  相似文献   

10.
The reaction of the tetranuclear cluster Pd4(CO)4(OOCCF3)4 witho-nitrosotoluene afforded the Pd11-containing complex [o-(NO)(CH2)C6H4]2Pd2(μ-OOCCF3)2. The elimination of CO2 and the formation of organic products of transformation of tolylnitrene species (azotoluene, ditolylamine, and tolylisocyanate) were observed in the course of the reaction. The title complex was characterized by IR and1H NMR spectroscopy. Its structure was established by X-ray diffraction analysis. It was suggested that the reaction proceeds through intermediate formation of nitrene complexes. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 147–150, January, 2000.  相似文献   

11.
The pyrolysis of the isonitrile substituted complexes Os3(CO)12?x(CNR)x (R = But, x = 1,2) in refluxing octane has been studied. From these pyrolysis reactions and from the reaction of Os3(CO)12 with ButNC in refluxing octane the series of hexanuclear complexes Os6(CO)18?x(CNBut)x (x = 1?5) has been isolated. The pyrolysis of Os3(CO)11(CNBun) also leads to the formation of higher nuclearity clusters and evidence is presented that one product of the reaction is Os6(CO)17(CNBun)2. Possible structures for these isonitrile substituted hexanuclear complexes are discussed in the light of the known structures of Os6(CO)16(CNBut)2 and Os6(CO)18(CNC6H4Me)2.  相似文献   

12.
Platinum group metal chalcogenides find extensive applications in catalysis and in the electronic industry. To develop an efficient low temperature clean preparation of these materials, molecular routes have been explored. Thus the chemistry of mononuclear organochalcogenolates of the type [M(ER’)2(PR3)2], binuclear benzylselenolates, [M2Cl2(μ-SeBz)2(PR3)2], allylpalladium complexes [Pd2(μ-ER)23-C4H7)2] and palladium/platinum sulphido/selenido-bridged complexes, [M2(μ-E)2L4] (M = Pd or Pt; E = S, Se or Te; L = tertiary phosphine ligand) has been investigated. All the complexes have been characterized by elemental analysis, NMR (1H,31P,77Se,195Pt) spectroscopy and in some cases by X-ray diffraction. The thermal behaviour of these complexes has been studied by TGA. The pyrolysis of allylpalladium complexes in refluxing xylene yields Pd4E as established by analysis and XRD patterns.  相似文献   

13.
New palladium nitrosyl carboxylate complexes Pd8(CO)4−m(NO)m(NO2)4(RCO2)8 (m = 2, 4) were obtained by the treatment of palladium carbonyl carboxylates clusters cyclo-Pdn(μ-CO)n(μ-RCO2)n (n = 6) (1) with gaseous nitrogen monoxide. These complexes are the products of CO substitution in early described Pd8(CO)4(NO2)4(RCO2)8 clusters. By adding an excess of corresponding acid to reaction mixture Pd4(CO)2(NO)(RCO2)5 complexes were obtained, their structures were determined by X-ray diffraction analysis. These clusters are intermediate products of transformation of 6-nuclear initial clusters into various 8-nuclear complexes. This fact demonstrates that carboxylate ligands can be used as stabilizers for intermediate unstable polynuclear palladium compounds.  相似文献   

14.
The reactions of LiHB(C2H5)3 and LiDB(C2H5)3 with Re2(CO)10, Ir4(CO)12, Os3(CO)12, Ru3(CO)12 and Rh4(CO)12 have been studied by 1H, 2H and 13C NMR techniques. Results suggest the formation of formaldehyde and methanol in these systems, as well as the existence of previously unreported formyl complexes. A 2H isotope effect is noted in the apparent increase in stability of cluster formyl complexes.  相似文献   

15.
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   

16.
It has been shown for the first time that the reaction of bi-valent tin acetyl-acetonate with palladium carbonylphosphine clusters, Pd4(CO)5(PPh3)4 (I), Pd4(CO)5(PEt3)4 (II) and Pd3(CO)3(PPh3)4 (III), results in the formation of heterometal pentanuclear clusters of general formula Pd3Sn2(acac)4(CO)2(PR3)3; R  Ph (IV), Et (V). X-ray analysis of Pd3Sn2(acac)4(CO)2(PPh3)3 at 20°C (λ(Mo), 4396 reflections, space group P21/n, Z = 4, R = 0.037) shows that IV in the form of the crystalline hydrate, Pd3Sn2(acac)4(CO)2(PPh3)3 · χH2O (χ ∼ 1), contains a distorted “propeller”-shaped Pd3Sn2 metal frame with PdSn distances of 2.679–2.721(1) Å; two short PdPd bonds, 2.708 and 2.720(1) Å, bridged by μ2-CO ligands, and an elongated central Pd(1)Pd(2) bond of 2.798 Å. Sn atoms have distorted octahedral coordination, the dihedral angles formed by Pd3 moieties and two Pd2Sn triangles are 127.6 and 106.5°; and the angle between Pd2Sn moieties is 126.0°.  相似文献   

17.
Reductive condensation of Pd(OAc)2 in dioxane in the presence of CO and PR3 (R = Et, Bun) with addition of CF3COOH leads to the formation of decanuclear Pd103-CO)42-CO)8(PBun3)6 (I) and Pd10(CO)14(PBun3) (II) at Pd(OAc)2:PR3 molar ratios of 1:4–1:10 and 1:1.5–1:2.5, respectively. The use of CH3COOH instead of CF3COOH results in tetranuclear clusters Pd4(CO)5(PR3)4 (III) and Pd42-CO)6(PBun3) (IV). I ? III and III → IV transformations occur in organic media. The structures of I (space group P21/n, Z = λMo, 12125 independent reflections, R = 0.047) and IV (Pz:3, Z = λMo, 3254 reflections, R = 0.098) were established by X-Ray diffractions analysis. Cluster I is a 10-vertex Pd10 polyhedron, an octahedron with four unsymmetrically centered non-adjacent faces. The average PdPd distances in the octahedron are 2.825 Å, in the eight short Pdoct.Pdcap. bonds with the “equatorial” Pd atoms of the inner octahedron, bridged by the μ2-CO ligands, are 2.709 Å, and in the four elongated (without bridging CO groups) bonds with the apical Pd atoms of the octahedron are 3.300–3.422 Å. The PBun3 ligands are coordinated to the apical Pd atoms and the capping atoms (PdP 2.291–2.324 Å). Cluster IV is tetrahedral, with the CO ligands symmetrically bridged; PdPd 2.778–2.817; PdP 2.232–2.291; PdC 2.06 Å (average).  相似文献   

18.

Abstract  

The reactivity of the trinuclear palladium cluster [Pd3(dppm)3(CO)] n+ (dppm = bis(diphenylphosphinomethane); n = 2, 1) towards F was investigated by electrochemical and spectroscopic methods. The reaction depends on the charge of the cluster. The chemical reduction of the cluster dication is observed in the presence of F generating the paramagnetic monocationic cluster. Spin-trapping experiments with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) provided evidence for the radical F as an intermediate. In a similar manner to the dication, the monocationic cluster [Pd3(dppm)3(CO)]+ is also reduced, but in a slower process, by the F ion to produce [Pd3(dppm)3(CO)]0. Additionally, the alkyne cluster adducts [Pd3(dppm)3(CO)(RCCR)] n+ (n = 2, 1; R = CO2Me) are also reactive towards F. Particularly, the dication adduct leads to a metastable fluoride adduct [Pd3(dppm)3(CO)(RCCR)(F)]+. The electroreductive behavior of this adduct involves electron-transfer steps and F exchange equilibriums, for which digital simulation enables the extraction of the thermodynamic parameters (standard potentials and equilibrium constants). Concurrently, the monocation adduct [Pd3(dppm)3(CO)(RCCR)]+ with F, leads to a disproponation generating 0.5 equiv. of [Pd3(dppm)3(CO)(RCCR)(F)]+ and 0.5 equiv. of [Pd3(dppm)3(CO)(RCCR)]0. The former slowly evolves to [Pd3(dppm)3(RCCR)(F)]+, which was described by X-ray diffraction method.  相似文献   

19.
The reaction of Os3(CO)10(NCMe)2 (1) with an excess of acenaphthylene at room temperature provided the complex Os3(CO)10(μ-H)(μ-η2-C12H7) (2). Compound 2 contains a σ-π coordinated acenaphthyl ligand bridging an edge of the cluster. Compound 2 was converted to the complex Os3(CO)9(μ-H)232-C12H6) (3) when heated to reflux in a cyclohexane solution. Compound 3 contains a triply bridging acenaphthyne ligand. Compound 3 reacts with acenaphthylene again at 160 °C to yield four new cluster complexes: Os4(CO)12422-C12H6) (4); Os2(CO)6(μ-η4-C24H12) (5); Os3(CO)9(μ-H)(μ34-C24H13) (6); and Os2(CO)5(μ-η4-C24H12)(η2-C12H8) (7). All compounds were characterized crystallographically. Compound 4 is a butterfly cluster of four osmium atoms bridged by a single acenaphthyne ligand. Compounds 5 and 7 are dinuclear osmium clusters containing metallacycles formed by the coupling of two equivalents of acenaphthyne. Compound 6 is a triosmium cluster formed by the coupling of an acenaphthyne ligand to an acenapthyl group that is coordinated to the cluster through a combination of σ and π-bonding.  相似文献   

20.
A series of the octapalladium chains supported by meso-Ph2PCH2P(Ph)CH2P(Ph)CH2PPh2 (meso-dpmppm) ligands, [Pd8(meso-dpmppm)4(L)2](BF4)4 (L=none ( 1 ), solvents: CH3CN ( 2 a ), dmf ( 2 b ), dmso ( 2 c ), RN≡C: R=Xyl ( 3 a ), Mes ( 3 b ), Dip ( 3 c ), tBu ( 3 d ), Cy ( 3 e ), CH3(CH2)7 ( 3 f ), CH3(CH2)11 ( 3 g ), CH3(CH2)17 ( 3 h )) and [Pd8(meso-dpmppm)4(X)2](BF4)2 (X=Cl ( 4 a ), N3 ( 4 b ), CN ( 4 c ), SCN ( 4 d )), were synthesized by using 2 a as a stable good precursor, and characterized by spectroscopic (IR, 1H and 31P NMR, UV-vis-NIR, ESI-MS) measurements and X-ray crystallographic analyses (for 1 , 2 a , b , 3 a , b , e , f , 4 a – d ). On the basis of DFT calculations on the X-ray determined structure of 2 b ( [2b-Pd8]4+ ) and the optimized models [Pd8(meso-Ph2PCH2P(H)CH2P(H)CH2PH2)4(CH3CN)2]4+ ( [Pd8Ph8]4+ ) and [Pd8(meso-H2PCH2P(H)CH2P(H)CH2PH2)4(CH3CN)2]4+ ( [Pd8H8]4+ ), with and without empirically calculating dispersion force stabilization energy (B3LYP-D3, B3LYP), the formation energy between the two Pd4 fragments is assumed to involve mainly noncovalent interactions (ca. −70 kcal/mol) with four sets of interligand C−H/π interactions and Pd⋅⋅⋅Pd metallophilic one, while electron shared covalent interactions are almost canceled out within the Pd8 chain. All the compounds isolated are stable in solution and exhibit characteristic absorption at ∼900 nm, which is assignable to a spin allowed HOMO to LUMO transition, and shows temperature dependent intensity change with variable absorption coefficients presumably due to coupling with some thermal vibrations. The structures and electronic states of the Pd8 chains are found finely tunable by varying the terminal capping ligands. In particular, theoretical calculations elucidated that the HOMO-LUMO energy gap is systematically related to the central Pd−Pd distance (2.7319(6)–2.7575(6) Å) by two ways with neutral ligands L ( 1 , 2 , 3 ) and with anionic ligands X ( 4 ), which are reflected on the NIR absorption energy of 867–954 nm. The isocyanide terminated Pd8 complexes ( 3 ) further reacted with excess of RNC (6 eq) to afford the Pd4 complexes, [Pd4(meso-dpmppm)2(RNC)2](BF4)2 ( 13 ), and the cyclic voltammograms of 2 a (L=CH3CN), 3 , and 13 (R=Xyl, Mes, tBu, Cy) demonstrated wide range redox behaviors from 2{Pd4}4+ to 2{Pd4}0 through 2{Pd4}2+↔{Pd8}4+, {Pd8}3+, and {Pd8}2+ strings. The oxidized complexes, [Pd4(meso-dpmppm)2(RNC)3](BF4)4 ( 16 ), were characterized by X-ray analyses, and the two-electron reduced chain of [Pd8(meso-dpmppm)4](BF4)2 ( 7 ) was analyzed by spectroscopic and electrochemical techniques and DFT calculations. Reactions of 2 a with 1 equiv. of aromatic linear bisisocyanide (BI) in CH2Cl2 deposited insoluble coordination polymers, {[Pd8(meso-dpmppm)4(BI)](BF4)4}n ( 5 ), and interestingly, they were soluble in acetonitrile, 31P{1H} and 1H DOSY NMR spectra as well as SAXS curves suggesting that the coordination polymers may exist in acetonitrile as dynamically 1D self-assembled coordination polymers comprising ca. 50 units of the Pd8 rod averaged within the timescale.  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