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.
Density functional calculations at the BP86/TZ2P level were carried out to understand the ligand properties of the 16‐valence‐electron(VE) Group 14 complexes [(PMe3)2Cl2M(E)] ( 1ME ) and the 18‐VE Group 14 complexes [(PMe3)2(CO)2M(E)] ( 2ME ; M=Fe, Ru, Os; E=C, Si, Ge, Sn) in complexation with W(CO)5. Calculations were also carried out for the complexes (CO)5W–EO. The complexes [(PMe3)2Cl2M(E)] and [(PMe3)2(CO)2M(E)] bind strongly to W(CO)5 yielding the adducts 1ME–W(CO)5 and 2ME–W(CO)5 , which have C2v equilibrium geometries. The bond strengths of the heavier Group 14 ligands 1ME (E=Si–Sn) are uniformly larger, by about 6–7 kcal mol?1, than those of the respective EO ligand in (CO)5W‐EO, while the carbon complexes 1MC–W(CO)5 have comparable bond dissociation energies (BDE) to CO. The heavier 18‐VE ligands 2ME (E=Si–Sn) are about 23–25 kcal mol?1 more strongly bonded than the associated EO ligand, while the BDE of 2MC is about 17–21 kcal mol?1 larger than that of CO. Analysis of the bonding with an energy‐decomposition scheme reveals that 1ME is isolobal with EO and that the nature of the bonding in 1ME–W(CO)5 is very similar to that in (CO)5W–EO. The ligands 1ME are slightly weaker π acceptors than EO while the π‐acceptor strength of 2ME is even lower.  相似文献   

2.
A comparison was made to investigate the structures and bonding of nickel complex that carry tetrylone and tetrylene ligands [(CO)2Ni‐{E(PH3)2}] ( Ni1E ) and [(CO)2Ni‐{NHEMe}] ( Ni2E ) (E = C to Pb) using quantum chemical calculations at the BP86 level with various basis sets (SVP, TZVPP, TZ2P+). The nature of the Ni–E bonds was analyzed with charge‐ and energy decomposition methods. The structures of tetrylone complexes Ni1E exhibit an interesting trend with the ligands E(PH3)2 are bonded in a tilted orientation relative to the fragment Ni(CO)2. In contrast, the calculated equilibrium structures of complexes Ni2E exhibit the NHEMe ligands (E = C to Sn) bonded in a head‐on way to the Ni(CO)2 fragment, while the bending angle gives the strongest side‐on bonded ligand NHPbMe when E = Pb. The interesting trend of the bond dissociation energy (BDE) is observed for the tetrylone, which has the same trend BDEs compared with tetrylene complexes. The EDA‐NOCV results indicate that the tetrylone ligands {E(PH3)2} in complexes are similar to the tetrylene ligands NHEMe as strong σ‐donors and weak π‐acceptors. The BDEs calculated for the Ni–E bonds in Ni1E and Ni2E show that the effect of bulky ligands may obscure the intrinsic Ni–E bond strength. The bonding analysis shows that the tetrylone ligands in Ni1E may act in a similar fashion to the tetrylene ligands in Ni2E . All complexes Ni1E and Ni2E are suitable targets for synthesis.  相似文献   

3.
The equilibrium geometries and bond dissociation energies of 16‐valence‐electron(VE) complexes [(PMe3)2Cl2M(E)] and 18‐VE complexes [(PMe3)2(CO)2M(E)] with M=Fe, Ru, Os and E=C, Si, Ge, Sn were calculated by using density functional theory at the BP86/TZ2P level. The nature of the M? E bond was analyzed with the NBO charge decomposition analysis and the EDA energy‐decomposition analysis. The theoretical results predict that the heavier Group 14 complexes [(PMe3)2Cl2M(E)] and [(PMe3)2(CO)2M(E)] with E=Si, Ge, Sn have C2v equilibrium geometries in which the PMe3 ligands are in the axial positions. The complexes have strong M? E bonds which are slightly stronger in the 16‐VE species 1ME than in the 18‐VE complexes 2ME . The calculated bond dissociation energies show that the M? E bonds become weaker in both series in the order C>Si>Ge>Sn; the bond strength increases in the order Fe<Ru<Os for 1ME , whereas a U‐shaped trend Ru<Os<Fe is found for 2ME . The M? E bonding analysis suggests that the 16‐VE complexes 1ME have two electron‐sharing bonds with σ and π symmetry and one donor–acceptor π bond like the carbon complex. Thus, the bonding situation is intermediate between a typical Fischer complex and a Schrock complex. In contrast, the 18‐VE complexes 2ME have donor–acceptor bonds, as suggested by the Dewar–Chatt–Duncanson model, with one M←E σ donor bond and two M→E π‐acceptor bonds, which are not degenerate. The shape of the frontier orbitals reveals that the HOMO?2 σ MO and the LUMO and LUMO+1 π* MOs of 1ME are very similar to the frontier orbitals of CO.  相似文献   

4.
A combined experimental and theoretical study on the main‐group tricarbonyls [B(CO)3] in solid noble‐gas matrices and [C(CO)3]+ in the gas phase is presented. The molecules are identified by comparing the experimental and theoretical IR spectra and the vibrational shifts of nuclear isotopes. Quantum chemical ab initio studies suggest that the two isoelectronic species possess a tilted η11‐CO)‐bonded carbonyl ligand, which serves as an unprecedented one‐electron donor ligand. Thus, the central atoms in both complexes still retain an 8‐electron configuration. A thorough analysis of the bonding situation gives quantitative information about the donor and acceptor properties of the different carbonyl ligands. The linearly bonded CO ligands are classical two‐electron donors that display classical σ‐donation and π‐back‐donation following the Dewar–Chatt–Duncanson model. The tilted CO ligand is a formal one‐electron donor that is bonded by σ‐donation and π‐back‐donation that involves the singly occupied orbital of the radical fragments [B(CO)2] and [C(CO)2]+.  相似文献   

5.
Quantum chemical calculations at the BP86 level with various basis sets (SVP, TZVPP, and TZ2P+) were carried out for the Fe(CO)4 of group‐13 half‐sandwich ECp* [Fe(CO)4ECp*] ( Fe4‐E ) (E = B to Tl). The chemical bonding of the Fe(CO)4ECp* bond was analyzed with charge‐ and energy decomposition methods. The calculated equilibrium structures of complexes Fe4‐E show that the ligands ECp* are bonded in an end‐on way to the fragment Fe(CO)4 in Fe4‐E with E = B to Ga. The compound Fe4‐In has a distorted end‐on ligand InCp*. In contrast, Fe4‐Tl has a side‐on bonded ligand TlCp*. The calculated bond dissociation energies (BDEs) suggest that the bond in the iron group‐13 half‐sandwich complexes Fe4‐E decreases from Fe4‐B to Fe4‐Tl . Natural bond orbital (NBO) analysis of the bonding situation reveals that the Fe(CO)4ECp* donation in Fe4‐E comes from the σ lone‐pair orbital of ECp*. Bonding analysis indicates that the ligand ECp* in complexes are strong σ donors and the NOCV pairs of the bonding show small π‐back donation from the Fe(CO)4 to the ECp* ligands.  相似文献   

6.
The carbodiphosphorane CO2 adduct O2CC(PPh3)2 ( 1a ) reacts with [(CO)5W(THF)] and [(CO)3W(NCEt)3] to produce the complexes [(CO)5W{η1‐O2CC(PPh3)2}] ( 2 ) and [(CO)4W{η2‐O2CC(PPh3)2}] ( 3 ), respectively. Whereas in 2 the betain‐like ligand is coordinated at the tungsten atom in a monodentate manner, in 3 it acts as a chelating ligand with formation of a WO2C four‐membered ring. As a by‐product during the reaction with the acetonitrile adduct also some crystals of the hydrolysis product [HC(PPh3)2]2[W6O19] · 3C2H4Cl2 (4 · 3C2H4Cl2) were isolated. All compounds could be characterized by X‐ray analyses and the usual spectroscopic methods.  相似文献   

7.
Addition of Cationic Lewis Acids [M′Ln]+ (M′Ln = Fe(CO)2Cp, Fe(CO)(PPh3)Cp, Ru(PPh3)2Cp, Re(CO)5, Pt(PPh3)2, W(CO)3Cp to the Anionic Thiocarbonyl Complexes [HB(pz)3(OC)2M(CS)] (M = Mo, W; pz = 3,5‐dimethylpyrazol‐1‐yl) Adducts from Organometallic Lewis Acids [Fe(CO)2Cp]+, [Fe(CO)(PPh3)Cp]+, [Ru(PPh3)2Cp]+, [Re(CO)5]+, [ Pt(PPh3)2]+, [W(CO)3Cp]+ and the anionic thiocarbonyl complexes [HB(pz)3(OC)2M(CS)] (M = Mo, W) have been prepared. Their spectroscopic data indicate that the addition of the cations occurs at the sulphur atom to give end‐to‐end thiocarbonyl bridged complexes [HB(pz)3(OC)2MCSM′Ln].  相似文献   

8.
Complexes Containing Antimony Ligands: [tBu2(Cl)SbW(CO)5], [tBu2(OH)SbW(CO)5], O[SbPh2W(CO)5]2, E[SbMe2W(CO)5]2 (E = Se, Te), cis‐[(Me2SbSeSbMe2)2Cr(CO)4] Syntheses of [tBu2(Cl)SbW(CO)5] ( 1 ), [tBu2(OH)SbW(CO)5] ( 2 ), O[SbPh2W(CO)5]2 ( 3 ), Se[SbMe2W(CO)5]2 ( 4 ), cis‐[(Me2SbSeSbMe2)2Cr(CO)4] ( 5 ) Te[SbMe2W(CO)5]2 ( 6 ) and crystal structures of 1 – 5 are reported.  相似文献   

9.
Aromatized cationic [(PNN)Re(π acid)(O)2]+ ( 1 ) and dearomatized neutral [(PNN*)Re(π acid)(O)2] ( 2 ) complexes (where π acid=CO ( a ), tBuNC ( b ), or (2,6‐Me2)PhNC ( c )), possessing both π‐donor and π‐acceptor ligands, have been synthesized and fully characterized. Reaction of [(PNN)Re(O)2]+ ( 4 ) with lithiumhexamethyldisilazide (LiHMDS) yield the dearomatized [(PNN*)Re(O)2] ( 3 ). Complexes 1 and 2 are prepared from the reaction of 4 and 3 , respectively, with CO or isocyanides. Single‐crystal X‐ray structures of 1 a and 1 b show the expected trans‐dioxo structure, in which the oxo ligands occupy the axial positions and the π‐acidic ligand occupies the equatorial plane in an overall octahedral geometry about the rhenium(V) center. DFT studies revealed the stability of complexes 1 and 2 arises from a π‐backbonding interaction between the dxy orbital of rhenium, the π orbital of the oxo ligands, and the π* orbital of CO/isocyanide.  相似文献   

10.
Vaska‐type complexes, i.e. trans‐[RhX(CO)(PPh3)2] (X is a halogen or pseudohalogen), undergo a range of reactions and exhibit considerable catalytic activity. The electron density on the RhI atom in these complexes plays an important role in their reactivity. Many cyanotrihydridoborate (BH3CN) complexes of Group 6–8 transition metals have been synthesized and structurally characterized, an exception being the rhodium(I) complex. Carbonyl(cyanotrihydridoborato‐κN)bis(triphenylphosphine‐κP)rhodium(I), [Rh(NCBH3)(CO)(C18H15P)2], was prepared by the metathesis reaction of sodium cyanotrihydridoborate with trans‐[RhCl(CO)(PPh3)2], and was characterized by single‐crystal X‐ray diffraction analysis and IR, 1H, 13C and 11B NMR spectroscopy. The X‐ray diffraction data indicate that the cyanotrihydridoborate ligand coordinates to the RhI atom through the N atom in a trans position with respect to the carbonyl ligand; this was also confirmed by the IR and NMR data. The carbonyl stretching frequency ν(CO) and the carbonyl carbon 1JC–Rh and 1JC–P coupling constants of the Cipso atoms of the triphenylphosphine groups reflect the diminished electron density on the central RhI atom compared to the parent trans‐[RhCl(CO)(PPh3)2] complex.  相似文献   

11.
Oxidative Addition Reactions of Complexed Phosphine and Arsine at Platinum(0) Complexes The reactions of E(SiMe3)3 with [W(CO)5thf] yield after alcoholysis in a one‐pot reaction the complexes [(CO)5WEH3] ( 1 ) (E = P( a ), As( b )). This procedure circumvents the direct use of gaseous phosphine and arsine, respectively. Complexes 1 react with [(C2H4)Pt(PPh3)2] in an oxidative addition type reaction to form the heterometallic complexes [(PPh3)2Pt(H){(μ‐EH2)W(CO)5}] (E = P( 2 ), As( 3 )). Complexes 2 and 3 were obtained as mixtures of their cis‐ and trans‐isomers, in which the contents of the trans‐isomer dominates. The products were comprehensively spectroscopically characterised, and the structures of the trans‐complexes 2 and 3 were determined by X‐ray crystal structure analysis.  相似文献   

12.
Carbonyls' 2π orbital populations, [2π], in W(CO)5L {L = PPh3, PPh2Me, PPhMe2} have been determined by NMR spin‐lattice relaxation techniques. Experimental values of axial [2π], compared with those reported for PMe3 and P(OMe)3, reveal that PMe3 is a slightly better π‐acid than PPh3. Through space interactions between carbonyl and phenyl groups are insignificant since values of [2π] do not vary significantly in the series of phosphines, going from PMep3 to PPh3. Natural bond orbital (NBO) studies indicate that π‐accepting capabilities for these phosphines are primarily governed by the nature of P‐C anti‐bonding, σ*P‐C. Compared with PPh3, the better π‐accepting σ*P‐C, as well as the better s‐donating lone‐pair LP(P), in PMe3 can both be explained by the higher extents of rehybridization of the coordinated phosphorus atom. Based on this rehybridization argument, the NBO predicted order of increasing π‐acidic strengths PPh3 < PPh2Me < PPhMe2 < PMe3, which cannot be clearly distinguished by NMR experiments, is ascribed to the same NBO trend of σ‐donating capabilities in a synergistic manner. Effects of coordination on P‐Y (Y = C, O, F) bonding strengths in phosphines (or phosphites) are depending on two conflicting effects: rehybridization of LP(P) and the hyperconjugative‐like dπ → σ*P‐Y back‐donation.  相似文献   

13.
Mixed‐ligands hydride complexes [RuHCl(CO)(PPh3)2{P(OR)3}] ( 2 ) (R = Me, Et) were prepared by allowing [RuHCl(CO)(PPh3)3] ( 1 ) to react with an excess of phosphites P(OR)3 in refluxing benzene. Treatment of hydrides 2 first with triflic acid and next with an excess of hydrazine afforded hydrazine complexes [RuCl(CO)(κ1‐NH2NHR1)(PPh3)2{P(OR)3}]BPh4 ( 3 , 4 ) (R1 = H, CH3). Diethylcyanamide derivatives [RuCl(CO)(N≡CNEt2)(PPh3)2{P(OR)3}]BPh4 ( 5 ) were also prepared by reacting 2 first with HOTf and then with N≡CNEt2. The complexes were characterized spectroscopically and by X‐ray crystal structure determination of [RuHCl(CO)(PPh3)2{P(OEt)3}] ( 2b ).  相似文献   

14.
The betain‐like compound S2CC(PPh3)2 ( 1 ), which is obtained from CS2 and the double ylide C(PPh3)2, reacts with [Co2(CO)8] and [Mn2(CO)10] in THF to afford the salt‐like complexes [Co{S2CC(PPh3)2}3][Co(CO)4]3 ( 2 ) and [(CO)4Mn{S2CC(PPh3)2}][Mn(CO)5] ( 3 ), respectively, in good yields. At both d6 cations 1 acts as a chelating ligand. Disproportionation reactions from formal Co0 into CoIII and Co?I and from Mn0 into MnI and Mn?I occurred with the removal of four or one carbonyl groups, respectively. The crystal structures of 2· 5.5THF and 3· 2THF are reported, which show a shortening of the C–C bond in the ligand upon complex formation. The compounds are further characterized by 31P NMR and IR spectroscopy.  相似文献   

15.
Reactions of the oxorhenium(V) complexes [ReOX3(PPh3)2] (X = Cl, Br) with the N‐heterocyclic carbene (NHC) 1,3,4‐triphenyl‐1,2,4‐triazol‐5‐ylidene (LPh) under mild conditions and in the presence of MeOH or water give [ReOX2(Y)(PPh3)(LPh)] complexes (X = Cl, Br; Y = OMe, OH). Attempted reactions of the carbene precursor 5‐methoxy‐1,3,4‐triphenyl‐4,5‐dihydro‐1H‐1,2,4‐triazole ( 1 ) with [ReOCl3(PPh3)2] or [NBu4][ReOCl4] in boiling xylene resulted in protonation of the intermediately formed carbene and decomposition products such as [HLPh][ReOCl4(OPPh3)], [HLPh][ReOCl4(OH2)] or [HLPh][ReO4] were isolated. The neutral [ReOX2(Y)(PPh3)(HLPh)] complexes are purple, airstable solids. The bulky NHC ligands coordinate monodentate and in cis‐position to PPh3. The relatively long Re–C bond lengths of approximate 2.1Å indicate metal‐carbon single bonds.  相似文献   

16.
1H‐1, 3‐Benzazaphospholes react with M(CO)5(THF) (M = Cr, Mo, W) to give thermally and relatively air stable η1‐(1H‐1, 3‐Benzazaphosphole‐P)M(CO)5 complexes. The 1H‐ and 13C‐NMR‐data are in accordance with the preservation of the phosphaaromatic π‐system of the ligand. The strong upfield 31P coordination shift, particularly of the Mo and W complexes, forms a contrast to the downfield‐shifts of phosphine‐M(CO)5 complexes and classifies benzazaphospholes as weak donor but efficient acceptor ligands. Nickelocene reacts as organometallic species with metalation of the NH‐function. The resulting ambident 1, 3‐benzazaphospholide anions prefer a μ2‐coordination of the η5‐CpNi‐fragment at phosphorus to coordination at nitrogen or a η3‐heteroallyl‐η5‐CpNi‐semisandwich structure. This is shown by characteristic NMR data and the crystal structure analysis of a η5‐CpNi‐benzazaphospholide. The latter is a P‐bridging dimer with a planar Ni2P2 ring and trans‐configuration of the two planar heterocyclic phosphido ligands arranged perpendicular to the four‐membered ring.  相似文献   

17.
Syntheses, Structure and Reactivity of η3‐1,2‐Diphosphaallyl Complexes and [{(η5‐C5H5)(CO)2W–Co(CO)3}{μ‐AsCH(SiMe3)2}(μ‐CO)] Reaction of ClP=C(SiMe2iPr)2 ( 3 ) with Na[Mo(CO)35‐C5H5)] afforded the phosphavinylidene complex [(η5‐C5H5)(CO)2Mo=P=C(SiMe2iPr)2] ( 4 ) which in situ was converted into the η1‐1,2‐diphosphaallyl complex [η5‐(C5H5)(CO)2Mo{η3tBuPPC(SiMe2iPr)2] ( 6 ) by treatment with the phosphaalkene tBuP=C(NMe2)2. The chloroarsanyl complexes [(η5‐C5H5)(CO)3M–As(Cl)CH(SiMe3)2] [where M = Mo ( 9 ); M = W ( 10 )] resulted from the reaction of Na[M(CO)35‐C5H5)] (M = Mo, W) with Cl2AsCH(SiMe3)2. The tungsten derivative 10 and Na[Co(CO)4] underwent reaction to give the dinuclear μ‐arsinidene complex [(η5‐C5H5)(CO)2W–Co(CO)3{μ‐AsCH(SiMe3)2}(μ‐CO)] ( 11 ). Treatment of [(η5‐C5H5)(CO)2Mo{η3tBuPPC(SiMe3)2}] ( 1 ) with an equimolar amount of ethereal HBF4 gave rise to a 85/15 mixture of the saline complexes [(η5‐C5H5)(CO)2Mo{η2tBu(H)P–P(F)CH(SiMe3)2}]BF4 ( 18 ) and [Cp(CO)2Mo{F2PCH(SiMe3)2}(tBuPH2)]BF4 ( 19 ) by HF‐addition to the PC bond of the η3‐diphosphaallyl ligand and subsequent protonation ( 18 ) and/or scission of the PP bond by the acid ( 19 ). Consistently 19 was the sole product when 1 was allowed to react with an excess of ethereal HBF4. The products 6 , 9 , 10 , 11 , 18 and 19 were characterized by means of spectroscopy (IR, 1H‐, 13C{1H}‐, 31P{1H}‐NMR, MS). Moreover, the molecular structures of 6 , 11 and 18 were determined by X‐ray diffraction analysis.  相似文献   

18.
Syntheses and Structures of η1‐Phosphaallyl, η1‐Arsaallyl, and η1‐Stibaallyl Iron Complexes [(η5‐C5Me5)(CO)2Fe–E(SiMe3)C(OSiMe3)=CPh2] (E = P, As, Sb) The reaction of equimolar amounts of [(η5‐C5Me5)(CO)2Fe–E(SiMe3)2] ( 1 a : E = P; 1 b : As; 1 c : Sb) and diphenylketene afforded the η1‐phosphaallyl‐, η1‐arsaallyl‐, and η1‐stibaallyl complexes [(η5‐C5Me5)(CO)2Fe–E(SiMe3)C(OSiMe3)=CPh2] ( 2 a : E = P; 2 b : As; 2 c : Sb). The molecular structures of 2 b and 2 c were elucidated by single crystal X‐ray analyses.  相似文献   

19.
The reaction of Na2[Fe(CO)4] with Br2CF2 in n‐pentane generates a mixture of the compounds (CO)3Fe(μ‐CO)3–n(μ‐CF2)nFe(CO)3 ( 2 , n = 2; 3 , n = 1) in low yields with 3 as the main product. 3 is obtained free from 2 by reacting Br2CF2 with Na2[Fe2(CO)8]. The non‐isolable monomeric complex (CO)4Fe=CF2 ( 1 ) can probably considered as the precursor for 2 . 3 reacts with PPh3 with replacement of two CO ligands to form Fe2(CO)6(μ‐CF2)(PPh3)2 ( 4 ). The complexes 2 – 4 were characterized by single crystal X‐ray diffraction. While the structure of 2 is strictly similar to that of Fe2(CO)9, the structure of 3 can better be described as a resulting from superposition of the two enantiomers 3 a and 3 b with two semibridging CO groups. Quantum chemical DFT calculations for the series (CO)3Fe(μCO)3–n(μ‐CF2)nFe(CO)3 (n = 0, 1, 2, 3) as well as for the corresponding (μ‐CH2) derivatives indicate that the progressively larger σ donor and π acceptor properties for the bridging ligands, in the order CO < CF2 < CH2, favor a stronger Fe–Fe bond.  相似文献   

20.
The dithiocarbene complex W(CO)5[C(SCH3)2 reacts with tertiary phosphines, PPh2CH3, PPh(CH3)2, P(C2H5)3 and P(OCH3)3 to form the phosphorane complexes W(CO)5[CH3S)2C-PR3] and with HPPh2 to form the phosphine complex W(CO)5[PPh2[CH(SCH3)2]. Kinetic studies of both types of reactions show that their rates are first order each in W(CO)5[C(SCH3)2] and in the phosphorus ligand. A mechanism involving rate determining phosphorus attack at the carbene carbon followed by rapid rearrangement to the product is consistent with this rate law. Rate constants for the reactions increase with increasing nucleophilicities of the phosphines: P(OCH3)3 < PPh2H < PPh2CH3 ? PPh(CH3)2 < P(C2H5)3. The ΔH values decrease (P(OCH3)3 > PPh2H > PPh2(CH3) > PPh(CH3)2 > P(C2H5)3) as the nucleophilicities of the phosphines increase. The ΔS values (≈-30 e.u.) remain essentially constant for all the reactions. The cyclic dithiorcarbenes W(CO)5[CS(CH2)nS], wheren- 3 or 4, react with PPh2(CH3) to form the cyclic phosphorane complexes, W(CO)5[S(CH2)nSC-PPh2(CH3)]. The 6- and 7- membered cyclic dithiocarbenes also react with PPh2H to form the phosphine complexes, W(CO)5 {PPh2- [CS(CH2)nS(H)]}.  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