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 31 毫秒
1.
Summary The [2.2]paracyclophane cluster, Ru6C(CO)14( 3- 2 2 2-C16H16) (1), undergoes reaction with Me3NO and triphenylphosphine to yield Ru6C(CO)13( 3- 2 2 2-C16H16)(PPh3) (2), which may also be produced from (1) by thermolysis with PPh3 in THF. Compound (2) has been fully characterized in solution by spectroscopy and in the solid state by a single crystal X-ray diffraction analysis at 277 K, and its structure is compared with that of the parent cluster, (1). Using the same synthetic procedures, the tricyclohexylphosphine analogue, Ru6C(CO)13( 3- 2 2 2-C16H16)(PCy3) (3), has also been prepared and characterized spectroscopically. A comparison of the chemical shifts of the 577-01 protons in the 1H-n.m.r. spectra of compounds (1)–(3) together with a variety of other [2.2]paracyclophane and benzene clusters has been made.  相似文献   

2.
The effects of aprotic inert media on the electronic absorption spectra of aromatic nitro compounds p-NO2C6H4R were used as evidence for the linear correlation between the slope s a of the solvatochromism equations max = 0 + s a* and the dipole moments of the molecules in their ground electronic state g. A linear correlation was established between 0 and the first ionization potential of subunits C6H5R. A new approach to estimating the dipole moment of electronically excited molecules (e) for molecules like p-NO2C6H4R on the basis of the correlation e = rg was proposed.  相似文献   

3.
The electron density distribution and atomic displacements were analyzed based on the results of precision low-temperature X-ray diffraction studies of a series of isostructural (Pnma, Z = 4) mixed metallocenes (5-C5H5)M(5-C7H7) (M = Ti, V, or Cr) and (5-C5H5)Ti(8-C8H8). The barriers to rotation of the cyclic ligands were evaluated based on rms libration amplitudes. Analysis of the deformation electron density demonstrated that the character of the M--(-ligand) chemical bond depends substantially both on the nature of the metal atom and the size of the ligand. Lowering of the local symmetry of the (5-C5H5)M(5-C7H7) complexes to CS leads to distortion of the cylindrical symmetry of the electron density distribution observed in vanadocene (5-C5H5)2V and titanocene (5-C5H5)Ti(8-C8H8).  相似文献   

4.
A series of are necyc lope ntadienyl complexes,i. e., [Ru(5-c5R5)(6- are ne)]+ (1, R= H, arene = C6H6; 2, R = Me, arme = C6H6; 3, R = H, arctic = C6H3Me3; 4, R = Me, arene = C6H3Me3; 5, R = H, arene = C6Me6; 6, R = Me, arene = C6Me6) was studied by cyclic voltammetry. These compounds are capable of both oxidation and reduction. The reduction potential values depend on the number of methyl groups in the complex. Reduction of benzene complexes I and 2 by sodium amalgam in THF leads to the formation of decomplexation products, the addition of hydrogen to benzene, and dimerization of the benzene ligands. Both chemical and electrochemical reductions of mesitylene complexes3 and4 result in dimeric products [(5-C5R5)Ru(-5;5-Me3H3C6H3Me3)Ru(5-C5R5)] (14, R = H; 15, R = Me). The action of sodium amalgam on compound5 gives products of hydrogen addition to both hexamethylbenzene (17) and cyclopentadienyl (18) ligands along with the major product, the dimer [5-C5H5)Ru(-5; 5-Me6C6C6Me6)Ru(5-C5H5)] (16). In contrast to5, its permcthylated analog 6 is only capable of adding hydrogen to the hexamethylbenzene ligand.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1691–1697, July, 1996.  相似文献   

5.
The viscosities of dilute solutions of a number of tetraalkylammonium and alkali metal halides, tetraphenylarsonium chloride, sodium tetraphenylborate, and tetrabutylammonium tetrabutylborate, as well as several nonelectrolytes have been measured in the high dielectric constant solvent N-methylacetamide (NMA) at 35 and 55°C. The relative viscosities were fitted to the extended Jones-Dole equation, = 1 + AC1/2 + BC + DC2. The pattern of behavior of the B coefficients is roughly similar to that observed in H2O. However, the small ions have exceptionally large B values in this solvent due to strong solvation effects, while the large organic ions do not display the sharp crossing of the Einstein law, B=2.5 V, characteristic in H2O of hydrophobic interaction. The D coefficients roughly parallel the B behavior and display remarkably regular ionic trends. This suggests that they arise largely from hydrodynamic origins. Nonelectrolytes have small or negative B coefficients showing that the Einstein law is not applicable at the molecular level and that nonelectrolytes are poor models for structurally similar ions. A simple mixture law is presented as an alternative to the Einstein law to explain the B coefficients.  相似文献   

6.
Summary [RuCl2(CO)2] n reacts with the Schiff base 1-acetylferrocenethiosemicarbazone, [Fe(-Cp)(-C5H4MeC=NN-HCSNH2)] to give [Fe(-Cp)(-C5H4MeC=NN-HCSNH2)RuCl2(CO)2] and with 1-acetylferrocenesemicarbazone [Fe(-Cp)(-C5H4MeC=NN-HCSNH2)] to give [Fe(-Cp)(-C5H4MeC=NN-HCSNH2)RuCl2-(CO) 2]. Spectroscopic data indicate that the Schiff bases act as bidentate ligands and coordinate to ruthenium via the hydrazinic N and either the S or O atoms, respectively, giving stable heterobimetallic complexes, which have been characterized by i.r. and 1H-n.m.r. spectroscopies, and elemental analyses.Part of this work was presented at the First International Conference in Chemistry and its applications in Doha, Qatar, 1993.  相似文献   

7.
The redox potentials of new Cr, Mn, and Fe polynuclear ladder complexes, (5-Cp)Fe(CO)2(1,5-C5H4)Fe(CO)2(1,5-C5H4)Mn(CO)3, (5-Cp)Fe(CO)2(1,5-C5H4)Mn(CO)3, (5-Cp)Fe(CO)2(1,6-Ph)Cr(CO)3, (5-Cp)Fe(CO)2(1,5-C5H4)Fe(CO)2CH2Ph, (5-Cp)Fe(CO)2(1,6-CH2Ph)Cr(CO)3, were measured and the mechanism of their electrochemical oxidation and reduction was suggested. It was shown that the - or -bonds of the bridging ligand can be cleaved selectively by applying cathodic or anodic potentials, respectively. On the basis of the obtained electrochemical data, a mechanism is suggested for the rearrangement observed when the complexes are metallated by butyllithium.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 362–366, February, 1995.This work was carried out with financial support from the Russian Foundation for Basic Research (Project No 94-03-08628a).  相似文献   

8.
The electron distributions and bonding in Ru3(CO)9( 3- 2, 2, 2-C6H6) and Ru3(CO)9( 3- 2, 2, 2-C60) are examined via electronic structure calculations in order to compare the nature of ligation of benzene and buckminsterfullerene to the common Ru3(CO)9 inorganic cluster. A fragment orbital approach, which is aided by the relatively high symmetry that these molecules possess, reveals important features of the electronic structures of these two systems. Reported crystal structures show that both benzene and C60 are geometrically distorted when bound to the metal cluster fragment, and our ab initio calculations indicate that the energies of these distortions are similar. The experimental Ru–Cfullerene bond lengths are shorter than the corresponding Ru–Cbenzene distances and the Ru–Ru bond lengths are longer in the fullerene-bound cluster than for the benzene-ligated cluster. Also, the carbonyl stretching frequencies are slightly higher for Ru3(CO)9( 3- 2, 2, 2-C60) than for Ru3(CO)9( 3- 2, 2, 2-C6H6). As a whole, these observations suggest that electron density is being pulled away from the metal centers and CO ligands to form stronger Ru–Cfullerene than Ru–Cbenzene bonds. Fenske-Hall molecular orbital calculations show that an important interaction is donation of electron density in the metal–metal bonds to empty orbitals of C60 and C6H6. Bonds to the metal cluster that result from this interaction are the second highest occupied orbitals of both systems. A larger amount of density is donated to C60 than to C6H6, thus accounting for the longer metal–metal bonds in the fullerene-bound cluster. The principal metal–arene bonding modes are the same in both systems, but the more band-like electronic structure of the fullerene (i.e., the greater number density of donor and acceptor orbitals in a given energy region) as compared to C6H6 permits a greater degree of electron flow and stronger bonding between the Ru3(CO)9 and C60 fragments. Of significance to the reduction chemistry of M3(CO)9( 3- 2, 2, 2-C60) molecules, the HOMO is largely localized on the metal–carbonyl fragment and the LUMO is largely localized on the C60 portion of the molecule. The localized C60 character of the LUMO is consistent with the similarity of the first two reductions of this class of molecules to the first two reductions of free C60. The set of orbitals above the LUMO shows partial delocalization (in an antibonding sense) to the metal fragment, thus accounting for the relative ease of the third reduction of this class of molecules compared to the third reduction of free C60.  相似文献   

9.
Interaction of [Ru(-C 6 H 6 )Cl 2 ] 2 with indenyl- or fluorenyllithium in THF gives, together with cationic benzene complexes [Ru( 5 -C 9 H 7 )(-C 6 H 6 )]+ and [Ru( 5 -C 13 H 9 )(-C 6 H 6 )]+, the neutral cyclohexadienyl derivatives Ru( 5 -C 9 H 6 -C 9 H 7 ) and Ru( 5 -C 13 H 9 )( 5 -C 6 H 6 -C 13 H 9 ), respectively. Interaction of the cyclohexadienyl complexes with Al 2 O 3 , Ph 3 C+, and CF 3 CO 2 H has been studied. Reaction of Ru( 5 -C 13 H 9 )( 5 -C 6 H 7 ) with CF 3 CO 2 H in the presence of an arene yields cationic cyclohexadienylarene complexes: [Ru( 5 -C 13 H 9 )( 6 -arene)]+ (arene=C 6 H 6 or 1,3,5-Me 3 C 6 H 3 ).A. N. Nesmeyanov Institute of Organoelemental Compounds, Russian Academy of Sciences, 117813 Moscow. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 3, pp. 699–706, March, 1992.  相似文献   

10.
In this article we review the synthesis, reactivity, and characterization of a number of clusters bearing the [2.2] paracyclophane ligand with nuclearities ranging from two to eight. Particular attention is focused on the different coordination modes that paracyclophane adopts; these being µ1- 6, µ2- 3 : 3, µ3- 1 : 2 : 2, and µ3- 2 : 2 : 2. Structural modifications which take place within the ring system on bonding in these various modes are also discussed.  相似文献   

11.
Summary Bimetallic chiral complexes [(5-C5H4CHMePh)(5-C5H5) Nb(CO)ML n ] with ML n =Co(CO)4 and HFe(CO)4 have been prepared and characterized. The lability of the metal-metal bond explains the lack of stereostability of these complexes. The two diastereoisomeric forms of polymetallic chiral complexes [(5- C5H5CHMePh)(5-C5H5)Nb(CO)[HFe3(CO)11] have been synthetized, separated and their stereostability has been investigated. All the complexes were characterized by elemental analysis i.r. and1H n.m.r. spectroscopy.  相似文献   

12.
Reaction of Ru4(CO)13(3-PPh) (1) with the 1,3,5-hexatriyne Me3SiCCCCC CSiMe3 under mild thermal conditions affords initially Ru4(CO)10(-CO)2{4-1,1,2-P(Ph)C(CCSiMe3)C(CCSiMe3) (2), via the facile formation of a P–C bond in a manner similar to that demonstrated previously with alkynes and diynes. The 62-CVE cluster 2 readily decarbonylates to give crystallographically characterised Ru4(CO)10(-CO)(4-PPh){4-1,1,2,2-Me3SiCCC2CCSiMe3} (3). Attempts to further incorporate the pendant alkyne moieties in 3 into the Ru4 coordination environment were partially successful with Ru4(CO)10(4-PPh)(4-1,1,3,3-RC4R') (4, R/R'=SiMe3/CCSiMe3) being formed as a minor product together with the unusual toluene coordinated species Ru4(CO)7(6-C6H5Me)(4-PPh)(4-1,1,3,3-Me3SiC4CCSiMe4) (5). Cluster 3 reacts with an excess of Me3SiCCCCCCSiMe3 to give the open chain cluster Ru4(CO)9(3-PPh){4-2,2,4,4,-C4(CCSiMe3)(SiMe3)C4(CCSiMe3)3} (6).  相似文献   

13.
Redox properties of mono- and binuclear -complexes of Cr with fluoranthene with the composition of (6-C16H10)Cr(6-C6H6), (6-C16H10)Cr(CO)3 and (-6,6-C16H10)Cr2(6-C6H6)(CO)3 are studied by cyclic voltammetry. Relations between half-wave potentials of redox processes and coordination sites of fragments Cr(6-C6H6)- and Cr(CO)3 with the ligand and their nature are found.  相似文献   

14.
The viscosities of most alkali and tetraalkylammonium halides have been measured in water at 25°C. The relative viscosities can be fitted, up to 1M, with the relation r =1+A c1/2+B c+D 2. TheA term depends on long-range coulombic forces, andB is a function of the size and hydration of the solute. When combined with partial-molal-volume data, the difference B –0.0025V° is mostly a measure of the solute-solvent interactions. IonicB are obtained if the tetraethylammonium ion is assumed to obey Einstein's law. TheD parameter depends on higher terms of the long-range coulombic forces, on higher terms of the hydrodynamic effect, and on structural solute-solute interactions. As such, it cannot be interpreted unambiguously.  相似文献   

15.
The viscosities of dilute solutions of a number of tetraalkylammonium and alkali metal halides, tetraphenylarsonium chloride, sodium tetraphenylborate, tetrabutylammonium tetrabutylborate, water, and 3,3-diethylpentane have been measured in the high-dielectric constant solvent, ethylene carbonate (EC) at 40°C. Crude values of the apparent molar volumes of these solutes have also been obtained. Relative viscosities were fitted to the extended Jones-Dole equation, r=17#x002B;A c 1/2+B C+D c 2.The pattern of the B coefficients is strikingly similar to that previously observed in the high dielectric constant, linear-chain hydrogen-bonded solvent, N-methylacetamide (NMA). Ionic values for v and B have been obtained using a variety of splitting techniques. Alkali metal ions have large B coefficients indicating strong cation solvation with the normal order Li>Na>K>Cs. Small anions have positive but much smaller B values than in NMA. The observed order does suggest, however, a small degree of anion solvation. Large organic ions do not display the sharp crossing of the Einstein law,D =2.5v, uniquely characteristic in H2O of hydrophobic interaction. The two non-electrolytes have negative B coefficients showing that the Einstein law is not valid at the molecular level and that hydrocarbons are not good models for their isoelectronic tetraalkylammonium ion counterparts. An empirical modification of the Einstein law to account for the finite size of the solvent molecules is discussed. As in NMA the D coefficients are roughly linear in the square of B suggesting that they arise from hydrodynamic origins.  相似文献   

16.
Reaction of [(3-C4H7)2Rh(CH3CN)2]PF6(3-C4H7 = -methallyl) with [n-Bu4N](VO3) gives a new 3-allyl cluster [n-Bu4N]2[{(3-C4H7)2Rh}2 (V4O12)] (I) which is readily converted into a diene cluster, [n-Bu4N]2 [{(4-C8H14)Rh}2(V4O12)] (II) (C8H14=2,5-dimethyl-1,5-hexadiene) by reacting with CO or P(OEt)3;I andII have been characterized crystallographically.  相似文献   

17.
The reaction of mercury(II) trifluoroacetate with the hexafluorophosphate of the 6-aniline-5cyclopentadienyliron(II) cation under reflux in dry ethanol gives rise to N-mono- and N,N-disubstituted mercury-containing salts of this cation. The same mercury-containing salts have been synthesized by the action of mercury(II) trifluoroacetate on the deprotonation product of the (6-aniline)(5-cyclopentadienyl)iron(II) cation. Direct mercuration of the [6-(N,N-dimethylaniline)](5-cyclopentadienyl)iron(II) cation into the para position of the benzene ring of the arene ligand has been performed. The reactivity of the compounds obtained has been studied.  相似文献   

18.
Oxidative dehydrodimerization of some phenylvinylidene complexes of manganese is studied by cyclic voltammetry. In the case of (5-C5H5)(CO)2Mn=C=C(H)Ph, the process occurs as the homolysis of the C–H bond in the radical cation of {(5-C5H5)(CO)2Mn=C=C(H)Ph} and the dimerization of intermediate -phenylethinyl cation [(5-C5H5)(CO)2Mn–CC–Ph]+ to a binuclear dication of bis-carbine type (5-C5H5)(CO)2Mn+C– C(Ph)=C(Ph)–CMn+(CO)2(5-C5H5). The reduction of the latter leads to binuclear bis-vinylidene complex (5-C5H5)(CO)2Mn=C=C(Ph)–C(Ph)=C=Mn(CO)2(5-C5H5). Oxidative dehydrodimerization of complexes (5-C5R5)(CO)(L)Mn=C=C(H)Ph (R = H, L = PPh3; R = Me, L = CO) occurs through the immediate C–C coupling of radical cations {(5-C5R5)(CO)(L)Mn=C=C(H)Ph} and yields binuclear dication bis-carbine complexes (5-C5R5)(CO)(L)Mn+C–C(H)(Ph)–C(H)(Ph)–CMn+(CO)(L)(5-C5R5), whose reduction leads to neutral compounds (5-C5H5)(CO)2Mn=C=C(Ph)–C(Ph)=C=Mn(CO)(L)(5-C5H5). Complex (5-C5H5)(CO)2Mn=C=C(Ph)–C(Ph)=C=Mn(CO)2(5-C5H5) undergoes the oxidation-induced nucleophilic addition of water, forming cyclic bis-carbene product with a bridge heterocyclic ligand (-3,4-diphenyl-2,5-dihydro-2,5-diylidene)-bis-(5-cyclopentadienyldicarbonyl manganese).  相似文献   

19.
The quantum-chemical DFT calculations of the Cp2Zn structure confirm the conclusion made earlier from the vibrational spectra that the sandwich structure (5-C5H5)2Zn (A) is not energetically favorable and more favorable are the close in energy -structure (5-C5H5)(1-C5H5)Zn (B) and -structure (1-C5H5)2Zn (C). The vibrational spectra of structures B and C with the DFT-derived force fields were calculated. A comparison of the calculated spectra of the isolated Cp2Zn molecules with the experimental data gives no way of deciding between the B and C structures. It is most likely that the molecule is nonrigid and experiences a strong influence from the nearest environment in solution or in the crystalline state.  相似文献   

20.
The solid-state structure of the triple-decker salt [Cp*Fe(-5:5-C4Me4P)RuCp*] · CF3SO3 shows orientational disorder for the pseudosymmetric cations. A chemically related compound was used to define a restrained structure model. Comparison of different refinement strategies proves that this restrained model is superior to an unrestrained treatment.  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