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 15 毫秒
1.
DFT calculations using MPWB1K method with COSMO continuum solvation model have been carried out to quantify the trans influence of various X ligands (EX) in [PtIICl3X]n− complexes as well as the mutual trans influence of two X and Y ligands (EXY) in [PtIICl2XY]n− complexes. A quantitative structure energy relationship (QSER) is derived for predicting the EXY using EX and EY and this relationship showed a strong similarity to a QSER derived for predicting EXY of [PdIICl2XY]n− complexes. Quantification of the contributions of EX and EXY to the bond dissociation energy of the ligand X (BDEX) in complexes of the type [MIIX(Y)X′(Y′)] (M = Pd, Pt) is also achieved. The BDEX of any ligand X in these complexes can be predicted using the equations, viz. BDEX(Pd) = 1.196EX − 0.603EXY − 0.118EX’Y’ + 0.442DX + 15.169 for Pd(II) complexes and BDEX(Pt) = 1.420EX − 0.741EXY − 0.125EX’Y’ + 0.498DX + 13.852 for Pt(II) complexes, where DX corresponds to the bond dissociation energy of X in [MIICl3X]n− complexes. These expressions suggest that the mutual trans influence from X and Y is more dominant than the mutual trans influence from X′ and Y′ and both factors contribute significantly to the weakening of M-X bond. We also obtained a strong linear relationship between EX and the electron density ρ(r) at the bond critical point of M-Cl bond trans to the X in [MIICl3X]n− and this allows us to express the BDEX(Pd) and BDEX(Pt) in terms of only the ρ(r) and DX. We have demonstrated that using a database comprising of DX and the ρ(r), the bond dissociation energy of X in complexes of the type [MIIX(Y)X′(Y′)] can be predicted.  相似文献   

2.
C-Si reductive elimination from Pt(R)(SiPh3)(PMe3)2 (R=Me, Pr) was theoretically studied with the density functional theory. For comparisons with the experiment, substitution of PMe3 with diphenylacetylene was taken into account. The calculated activation barriers in the C-Si elimination step after the ligand exchange were 22.0 and 28.9 kcal mol−1 for R=Me and Pr, respectively, which explains the reactivity difference reported experimentally. In order to analyze the energy difference, we optimized transition states of several model complexes, and examined the influence of the steric repulsion between R and the other ligands. Comparisons of the geometries and the barrier heights reveal that the steric repulsion and the Si-alkyl bond energy are important factors controlling the reaction rate.  相似文献   

3.
A systematic search of the Cambridge structural database was undertaken to quantify the trans influence in square planar and octahedral transition metal compounds. For square planar geometry, d8 metal centers were studied, while octahedral searches focused on low-spin d6 complexes. Two probe ligands (PL) were used to measure the effect of the trans ligand (TL), chloride, and triphenylphosphine (PPh3). For the TLs O=CX2, NR3, pyridine, and Cl? (X?=?any non-metal, R?=?H or hydrocarbon), the effects on the metal–probe ligand (M–PL) distance were statistically equal and were taken as essentially no trans influence. The other ligands studied showed significant decrease in the mean M–PL bond order, relative to the above ligands: SR2?=?0.941; S=CX2?=?0.887; PPh3?=?0.825; phenyl?=?0.743; CR3?=?0.719; hydride?=?0.685. Some variation in the trans influence is observed, based on the geometry of the metal center and the PL. In general, electron-donating, σ-bonding ligands lead to a larger trans influence, but π-bonding effects can also be important, particularly when the probe ligand also has π-bonding properties.  相似文献   

4.
Electronic and ESR spectra of the complexes [Cu(II)(tpaso)4][Cu(I)Cl2]2,[Cu(tpaso)4](NO3)2 and [Cu(tpaso)4](ClO  相似文献   

5.
The vibrational spectra, both i.r. and Raman, of some N-monosubstituted oxamide complexes with Pd(II) and their deuterated derivatives are reported, using ligands which have the general form NH2COCONHR where R = H, CH3, CH2CH3, CH2CH2CH3, NH2, CH2CH2NH2 or CH2CH2OH. The article also reports on the thermal analysis of the compounds, mainly based on thermogravimetric measurements.  相似文献   

6.
7.
The reactions of a series of Pd(II) methyl compounds of general formula LPd(NCCH(3))CH(3), where L is a bulky phenoxydiazene or phenoxyaldimine ligand with the polar olefin acrylonitrile (AN), are reported. The compounds react with an excess of AN to give the products of 2,1 insertion into the Pd-Me bond, yielding dimers and/or trimers which feature bridging alpha-cyano groups. The reactions were studied by low temperature (1)H NMR spectroscopy, revealing an initial formation of compounds featuring N-bound AN, which isomerized to an (unobserved) pi-bound species that rapidly underwent 2,1 insertion into the Pd-Me bond. Intermediate oligomeric complexes retaining a Pd-Me function were observed at low [AN] in these reactions. Under pseudo first-order conditions, k(obs) values of 8.5 x 10(-5) to 2.68 x 10(-3) M(-1) (-22 degrees C to 10 degrees C, 100 equiv of AN) and activation parameters of DeltaH++ = 14.4(5) kcal mol(-1) and DeltaS++ = -19(5) eu were obtained in one case. Comparison of the overall rates of insertion between two LPd(NCCH(3))CH(3), differing in the overall charge on the supporting ligand L, showed that the complex bearing a negatively charged ligand reacts with AN twice as fast as one with no anionic charge. The rates of insertion in both of these complexes are significantly faster than reported rates for analogous reactions in cationic Pd(II) derivatives, indicating that increasing the negative charge on the complex enhances the rate of AN insertion. These results provide fundamental mechanistic insights into a crucial reaction for incorporation of polar comonomers into alpha olefins via a coordination polymerization mechanism.  相似文献   

8.
A series of 2-iminopyrrole ligand precursors with increasing bulkiness [HNC4H3C(R)=N-2,6-R'2C6H3] (R = R' = H, 1a; R = Me, R'= H, 1b; R = H, R' = Me, 1c; R = R' = Me, 1d; R = H, R' = iPr, 1e; R = Me, R' = iPr, 1f) were synthesized and deprotonated with NaH to give the corresponding iminopyrrolyl sodium salts 2a-f. A set of homoleptic bis-ligand Co(II) complexes of the type [Co(kappa2N,N'-NC4H3C(R)=N-2,6-R'2C6H3)2] (R = R'= H, 3a; R = Me, R'= H, 3b; R = H, R' = Me, 3c; R = R' = Me, 3d; R = H, R' = iPr, 3e; R = Me, R' = iPr, 3f) was prepared by reaction of CoCl2 with the corresponding iminopyrrolyl sodium salts 2a-f. The new complexes were characterized by elemental analysis, magnetic susceptibility measurements, in powder and in solution, UV/vis/NIR, and, in some cases, X-ray crystallography. According to X-ray diffraction and magnetic measurements, the Co complexes 3a-e proved to be tetrahedral, which is the preferred geometry for Co(II) compounds. However, a square planar geometry is observed in the case of 3f, as determined by several characterization techniques. In this case, DFT calculations suggest the square planar geometry is slightly more stable than the tetrahedral one probably due to a combination of steric and electronic reasons.  相似文献   

9.
10.
11.
《Polyhedron》1986,5(6):1183-1190
The Pd(II) and Rh(I) complexes of tetra-acetylethane [H2dahd (3,4-diacetyl-2,4-hexadiene-2,5-diol)] with O,O′-bonded chelates, represented as [M2(O2,O2-dahd)(L2)2][X]m {M = Pd, L2 = (PPh3)2 or bdpe [1,2-bis(diphenylphosphino)ethane], X = BF4 or PF6, m = 2; M = Rh, L = Co, m = 0}, have recently been prepared. [Pd2(O2,O2-dahd)-(PPh3)4][PF6]2 reacts with the potentially bidentate 1,10-phenanthroline (phen) to give the five-coordinate complex [Pd(PPh3)(phen)2][PF6]2 and [Pd(O1,O1-dahd)(phen)]n, the latter of which is rather insoluble in organic solvents. [Pd(O1, O1-dahd)(phen)]n in CH2Cl2 readily transforms to a monomer complex [Pd(C3, O′-dahd)phen)]. These anomalous Pd(II) and Rh(I) complexes of the tetra-acetylethane dianion have been characterized from elemental analyses, conductance, IR, 1H and 13C NMR spectroscopy, magnetic susceptibility and ESR spectroscopy.  相似文献   

12.
Upon addition of base the complex enPd(OH2)22+ rapidly dimerizes to form a binuclear dihydroxo bridged dimer with no significant mononuclear hydroxo complexes at 10 mM concentration. The corresponding Pt(II) complex titrates two equivs base with pKa 5.8 and 7.6 to give a dihydroxo complex which is stable indefinitely. Addition of only one equiv of base to yield near neutral solutions slowly produces the Pt(II) dihydroxo bridged dimer. The hydroxo groups of dienPtOH+, dienPdOH+ and enPt(OH)2 are effective nucleophiles for ester hydrolysis and their reactivities are in accord with a suggested dependence on basicity. The antitumor complexes of cis dichloro Pt(II) amines are converted to aquo and hydroxo complexes of Pt(II) amines. Due to the low ambient chloride ion concentration within a cell, aquo complexes are present in significant amounts and are most likely to be the reactive species. The reactivity of cis Pt(II) amine antitumor complexes is reduced by Cl, OH and dihydroxy bridged dimer substituents, the formation of which is time dependent.  相似文献   

13.
The electronic properties of reactive and ancillary ligands have a large impact on the rate and scope of reductive elimination reactions. The purpose of this review is to compare and discuss published data on the effect of ligand electronic properties on the rates and scope of reductive eliminations from palladium(II). An understanding of these effects is important because reductive elimination from palladium(II) is the product-forming step of a variety of catalytic processes. The scope of this review will encompass the effect of the electron-donating abilities of alkyl, aryl, amido, alkoxo, thiolato, and phosphido groups on the rate of reductive elimination, the relative importance of inductive and resonance effects on the rate of reductive elimination, the relative sensitivity of the different classes of reductive eliminations to electronic perturbations, and the effect of the differences in electronic properties between the two aryl groups of biaryl complexes undergoing reductive elimination. In addition, this review will include the effects of electronic properties of the ancillary ligands on the rate of reductive eliminations from palladium(II). The effect of the overall electron-donating ability of ancillary ligands and the effect of the relative orientation of ancillary ligands to the two reactive ligands on the rate of reductive elimination will be discussed. Where appropriate, electronic effects on reductive elimination from complexes of other metals are described.  相似文献   

14.
A computational study of square planar organometallic complexes formed by the ligand 2,2′-bipyridine and all its possible difluoro derivatives in analogous position of the aromatic rings (n,n′-difluoro-2,2′-bipyridine, where n = n′ = 3, 4, 5, and 6) and two M–X2 (M = Pd and Pt and X = F, Cl, Br, and I) fragments has been carried out amounting to a total of 80 complexes. Relationships have been found between the chiral distinction energy and the different chemical moieties present. Using the statistical Free-Wilson method, the relative energies between the diastereomeric complexes have been correlated with the position of the substituent, the counter anion, and the central metal cation.  相似文献   

15.
Platinum(II) and palladium(II) complexes of the trithiacrown [9]aneS(3) containing a range of Group 15 donors are reviewed. These complexes have the general formula [M([9]aneS(3))(L(2))](n+) where L represents at least one Group 15 donor. Complexes involving pnictogens, with the exception of bismuth, are observed. The complexes generally have an elongated square pyramidal geometry with a long distance interaction to the third sulphur of the [9]aneS(3) which forms the apex of the square pyramid. This axial metal-sulphur distance is quite sensitive to the donor properties of L. Poorer donors such as Sb and As ligands show short axial distances whereas the better N donor ligands show longer distances. Pt(II) complexes of the formula [Pt([9]aneS(3))(EPh(3))(2)](2+) (E = P, As, Sb) show a considerable distortion towards a trigonal bipyramidal geometry due to intramolecular π-π interactions. Over seventy of these types of complexes have been crystallographically characterized and are discussed in this article. Other unique features of the complexes, including NMR spectroscopy, redox chemistry, and electronic spectroscopy, are also discussed.  相似文献   

16.
Masayuki Wasa 《Tetrahedron》2010,66(26):4811-19601
Protocols to effect β-arylation of sp3 C-H bonds via Pd(II)/(IV) and Pd(0)/(II) catalytic cycles have been achieved using a newly developed monodentate CONHC6F5 directing group. These reactions provide an unprecedented means to functionalize sp3 C-H bonds in aliphatic carboxylic acid-derived substrates.  相似文献   

17.
A series of platinum-based sensitizers of the general type Pt(NN)(SS), where NN is 4,4'-dicarboxy-2,2'-bipyridine (dcbpy) or 4,7-dicarboxy-1,10-phenanthroline (dcphen) and SS is ethyl-2-cyano-3,3-dimercaptoacrylate (ecda), quinoxaline-2,3-dithiolate (qdt), 1,2-benzenedithiolate (bdt), or 3,4-toluenedithiolate (tdt), that have various ground-state oxidation potentials has been synthesized and anchored to nanocrystalline titanium dioxide electrodes for light-to-electricity conversion in regenerative photoelectrochemical cells with an I(-)/I(-)(3) acetonitrile electrolyte. The intense mixed-Pt/dithiolate-to-diimine charge-transfer absorption bands in this series could be tuned from 440 to 580 nm by choosing appropriate dithiolate ligands, and the highest occupied molecular orbitals varied by more than 500 mV. Spectrophotometric titration of the Pt(dcphen)(bdt) complex exhibits a ground-state pK(a) value of 3.2 +/- 0.1, which can be assigned to the protonation of the carboxylate group of the dcphen ligand. Binding of Pt(dcbpy)(qdt) to porous nanostructured TiO(2) films was analyzed using the Langmuir adsorption isotherm model, yielding an adsorption equilibrium constant of 4 x 10(5) M(-1). The amount of dye adsorbed at the surface of TiO(2) films was 9.5 x 10(-8) mol/cm(2), which is ca. 50% lower than the full monolayer coverage. The resulting complexes efficiently sensitized TiO(2) over a notably broad spectral range and showed an open-circuit potential of ca. 600 mV with an impressive fill factor of > 0.70, making them attractive candidates for solar energy conversion applications. The visible spectra of the 3,4-toluenedithiol-based sensitizers showed an enhanced red response, but the lower photocurrent efficiency observed for these sensitizers stems in part from a sluggish halide oxidation rate and a fast recombination of injected electrons with the oxidized dye.  相似文献   

18.
The reaction of Ni(COD)(2)(COD = 1,5-cyclooctadiene) with triethylphosphine and pentafluoropyridine in hexane has been shown previously to yield trans-[NiF(2-C(5)NF(4))(PEt(3))(2)](1a) with a preference for reaction at the 2-position of the heteroaromatic. The corresponding reaction with 2,3,5,6-tetrafluoropyridine was shown to yield trans-[NiF(2-C(5)NF(3)H)(PEt(3))(2)](1b). In this paper, we show that reaction of Ni(COD)(2) with triethylphosphine and pentafluoropyridine in THF yields a mixture of 1a and 1b. Competition reactions of Ni(COD)(2) with triethylphosphine in the presence of mixtures of heteroaromatics in hexane reveal a kinetic preference of k(pentafluoropyridine):k(2,3,5,6-tetrafluoropyridine)= 5.4:1. Treatment of 1a and 1b with Me(3)SiN(3) affords trans-[Ni(N(3))(2-C(5)NF(4))(PEt(3))(2)](2a) and trans-[Ni(N(3))(2-C(5)NHF(3))(PEt(3))(2)](2b), respectively. The complex trans-[Ni(NCO)(2-C(5)NHF(3))(PEt(3))(2)](3b) is obtained on reaction of with Me(3)SiNCO and by photolysis of under CO, while trans-[Ni(eta(1)-C [triple bond CPh)(2-C(5)NF(4))(PEt(3))(2)](4a) is obtained by reaction of phenylacetylene with 1a. Addition of KCN, KI and NaOAc to complex 1a affords trans-[Ni(X)(2-C(5)NF(4))(PEt(3))(2)](5a X = CN, 6a X = I, 7a X = OAc), respectively. The PEt(3) groups of complex are readily replaced by addition of 1,2-bis(dicyclohexylphosphino)ethane (dcpe) to produce [NiF(2-C(5)F(4)N)(dcpe)](8a). Addition of dcpe to trans-[Ni(OTf)(2-C(5)F(4)N)(PEt(3))(2)](10a), however, yields the salt [Ni(2-C(5)F(4)N)(dcpe)(PEt(3))](OTf)(9a) by substitution of only one PEt(3) and displacement of the triflate ligand. The structures of 2b, 4a, 7a and 8a were determined by X-ray crystallography. The influence of different ancillary ligands on the bond lengths and angles of square-planar nickel structures with polyfluoropyridyl ligands is analysed.  相似文献   

19.
Summary Electron spin resonance studies have been carried out onN-ethylimidazole andN-propylimidazole 4 : 1 complexes of CuX2 salts (X = ClO 4 , NO 3 , , Br or Cl) in their polycrystalline and undiluted form at 295 K and 77 K. Cupric ion hyperfine structural resolution is observed for all the complexes at 295 K and the spectra are characteristic of a CuN4 chromophore with axial symmetry. In complexes involving ClO , Br and Cl the anions are nonbonding, whereas those with the NO 3 anion are weakly bonding. The electronic and e.s.r. spectral data have been correlated. The resolution of Cu2+ ion hyperfine structure in these complexes is attributed to a decrease in the dipolar interaction and has been observed for the first time since the first resolution reported in 1954 for CuN4 coordination with square planar symmetry for ,,,-tetraphenylporphyrincopper (CuTPP).  相似文献   

20.
A diastereomer pair of left- and right-handed square planar Pd(II) complexes was synthesized from a single chiral ligand as kinetic and thermodynamic products. Helicity inversion between the diastereomers occurred rapidly under thermal and microwave irradiation conditions.  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