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.
The preparation and characterisation of the Group IVA neopentyls (Me3CCH2)4M (M = Ti, Zr, or Hf) is described. Spectroscopic data (IR, Raman, mass, 1H NMR, and PE) are provided; IR and Raman bands have been assigned by comparison with results on Group IVB analogues (M = Ge or Sn). MC4 stretching vibrations fall in the range 540–485 cm?1, and bending modes at 240–283 cm?1. Thermal decomposition gives neopentane as the sole detectable product; qualitatively, stability increases in the order M = Ti < Zr < Hf and for R4M : R = Me ? Me3CCH2 ≈ Me3SiCH2. (Me3CCH2)4Ti is aerobically oxidised in benzene to give (Me3CCH2O)4Ti  相似文献   

2.
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   

3.
The main Group IVB organometallic halides, R3MX (M = Si, Ge, Sn, Pb; R = CH3, C6H5) react with crotylmagnesium bromide in ether and THF, to give the corresponding organometallic allylic compounds. Generally, mixtures of the two primary (Z + E) and the secondary isomers are obtained.  相似文献   

4.
Compounds of the type CH2CHCH2MR3 and E-PhCHCHMR3 (M  Si, Ge, Sn, Pb) were allowed to react with a series of heteroatom-centered radicals (PhY ·, Y = S, Se, Te, derived from PhYYPh) and carbon-centered radicals ((CH3)2CH · derived from (CH3)2CHHgCl). We report that alkenylplumbanes and, under forcing conditions, alkenylgermanes undergo SH2 or SH2′ substitution of the metal by chain mechanism analogous to those previously reported for alkenylstannanes. Alkenylsilanes are unreactive.Based solely upon product yields, the following trends were observed: The reactivity of the alkenylmetals follow the order metal = Pb > Sn > Ge (> Si). The allylmetals were more reactive then the β- metallostyrenes toward the reactants employed in this study. The chalcogen series PhYYPh exhibits the reactivity order Y = S > Se > Te.  相似文献   

5.
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   

6.
Quantum chemical DFT calculations at the BP86/TZ2P level have been carried out for the complex [HSi(SiH2NH)3Ti–Co(CO)4], which is a model for the experimentally observed compound [MeSi{SiMe2N(4-MeC6H4)}3Ti–Co(CO)4] and for the series of model systems [(H2N)3M–M′(CO)4] (M = Ti, Zr, Hf; M′ = Co, Rh, Ir). The Ti–Co bond in [HSi(SiH2NH)3Ti–Co(CO)4] has a theoretically predicted BDE of D e = 59.3 kcal/mol. The bonding analysis suggests that the titanium atom carries a large positive charge, while the cobalt atom is nearly neutral. The covalent and electrostatic contributions to the Ti–Co attraction have similar strength. The Ti–Co bond can be classified as a polar single bond, which has only little π contribution. Calculations of the model compound (H2N)3Ti–Co(CO)4 show that the rotation of the amino groups has a very large influence on the length and on the strength of the Ti–Co bond. The M–M′ bond in the series [(H2N)3M–M′(CO)4] becomes clearly stronger with Ti < Zr < Hf, while the differences between the bond strengths due to change of the atoms M′ are much smaller. The strongest M–M′ bond is predicted for [(H2N)3Hf–Ir(CO)4].  相似文献   

7.
Organodihydridoelement anions of germanium and tin were reacted with metallocene dichlorides of Group 4 metals Ti, Zr and Hf. The germate anion [Ar*GeH2] reacts with hafnocene dichloride under formation of the substitution product [Cp2Hf(GeH2Ar*)2]. Reaction of the organodihydridostannate with metallocene dichlorides affords the reduction products [Cp2M(SnHAr*)2] (M=Ti, Zr, Hf). Abstraction of a hydride substituent from the titanium bis(hydridoorganostannylene) complex results in formation of cation [Cp2M(SnAr*)(SnHAr*)]+ exhibiting a short Ti–Sn interaction. (Ar*=2,6-Trip2C6H3, Trip=2,4,6-triisopropylphenyl).  相似文献   

8.
A new series of copper(II) mononuclear and copper(II)–metal(II) binuclear complexes [(H2L)Cu] ? H2O, [CuLM] ? nH2O, and [Cu(H2L)M(OAc)2] ? nH2O, n = 1–2, M = Co(II), Ni(II), Cu(II), or Zn(II), and L is the anion of dipyridylglyoxal bis(2-hydroxybenzoyl hydrazone), H4L, were synthesized and characterized. Elemental analyses, molar conductivities, and FT-IR spectra support the formulation of these complexes. IR data suggest that H4L is dibasic tetradentate in [(H2L)Cu] ? H2O and [Cu(H2L)M(OAc)2] ? nH2O but tetrabasic hexadentate in [CuLM] ? nH2O (n = 1–2). Thermal studies indicate that waters are of crystallization and the complexes are thermally stable to 347–402°C depending upon the nature of the complex. Magnetic moment values indicate magnetic exchange interaction between Cu(II) and M(II) centers in binuclear complexes. The electronic spectral data show that d–d transitions of CuN2O2 in the mononuclear complex are blue shifted in binuclear complexes in the sequences: Cu–Cu > Cu–Ni > Cu–Co > Cu–Zn, suggesting that the binuclear complexes [CuLM] ? nH2O are more planar than the mononuclear complex. The structures of complexes were optimized through molecular mechanics applying MM +force field coupled with molecular dynamics simulation. [(H2L)Cu] ? nH2O, [CuLM] ? nH2O, and the free ligand were screened for antimicrobial activities on some Gram-positive and Gram-negative bacterial species. The free ligand is inactive against all studied bacteria. The screening data showed that [CuLCu] ? H2O > [(H2L)Cu] ? H2O > [CuLZn] ? H2O > [CuLNi] ? 2H2O ≈ [CuLCo] ? H2O in order of biological activity. The data are discussed in terms of their compositions and structures.  相似文献   

9.
Abstract

Full geometry optimizations were carried out on the singlet and triplet states of β-substituted divalent five-membered rings XC4H3M (X? ?NH2, ?OH, ?CH3 ?H, ?CH3, ?Br, ?Cl, ?F, ?CF3, and ?NO2; M?C, Si, and Ge) by the B3LYP method by using 6-311++G** basis set. The thermal energy gaps, ΔEt–s; enthalpy gaps, ΔHt–s; and Gibbs free energy gaps, ΔGt–s, between the singlet (s) and triplet (t) states of the above structures were calculated by using the GAUSSIAN 03 program. The ΔGt–s of XC4H3C was changed in the order: X? ?Cl > ?Br > ?CH3 > ?H > ?CF3 > ?F > ?NO2 > ?OH > ?NH2. The changes of ΔGt–s for XC4H3Si and XC4H3Ge were in the order: X? ?NH2 > OH > F > Cl > Br > CH3 > H > CF3 > NO2. The geometrical parameters, including bond lengths (R), bond angles (A), dihedral angles (D), natural bonding orbital (NBO) charge at atoms, HOMO and LUMO, and dipole moments, were presented and discussed.

Supplemental materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements to view the free supplemental file.

GRAPHICAL ABSTRACT   相似文献   

10.
175, 181Hafnium(IV) was extracted by HDBP in 2-ethylhexanol from 1–10M solutions of HClO4, HCl and HNO3, and 1–8M H2SO4. As with low polar organic phase diluents, the acidity dependence of the distribution ratio of Hf, D, passes through a minimum for HClO4, HCl, and H2SO4 whereas only an increase of D can be observed with increasing HNO3 concentration. From the slope analysis the following complexes were found to be extracted (HDBP=HA): HfA4 at <4M HClO4 and <5M HCl, lg Kextr=9, HfX4(HA)4 (X=ClO 4 , Cl or NO 3 ) at >5M HClO4, >7M HCl and 1–10M HNO3, Hf(SO4)A2(HA)3–4 at <3M H2SO4, and Hf(SO4)2 (HA)4 at >6M H2SO4. Coextraction of sulphate with hafnium from H2SO4 solutions was evidenced in experiments with macro concentrations of Hf(IV) and35SO 4 2− . Part XX: Coll. Czech. Chem. Commun., 40 (1975) 3617.  相似文献   

11.
Quantum chemical calculations using density functional theory at the BP86/TZ2P level have been carried out to determine the geometries and stabilities of Group 13 adducts [(PMe3)(EH3)] and [(PMe3)2(E2Hn)] (E=B–In; n=4, 2, 0). The optimized geometries exhibit, in most cases, similar features to those of related adducts [(NHCMe)(EH3)] and [(NHCMe)2(E2Hn)] with a few exceptions that can be explained by the different donor strengths of the ligands. The calculations show that the carbene ligand L=NHCMe (:C(NMeCH)2) is a significantly stronger donor than L=PMe3. The equilibrium geometries of [L(EH3)] possess, in all cases, a pyramidal structure, whereas the complexes [L2(E2H4)] always have an antiperiplanar arrangement of the ligands L. The phosphine ligands in [(PMe3)2(B2H2)], which has Cs symmetry, are in the same plane as the B2H2 moiety, whereas the heavier homologues [(PMe3)2(E2H2)] (E=Al, Ga, In) have Ci symmetry in which the ligands bind side‐on to the E2H2 acceptor. This is in contrast to the [(NHCMe)2(E2H2)] adducts for which the NHCMe donor always binds in the same plane as E2H2 except for the indium complex [(NHCMe)2(In2H2)], which exhibits side‐on bonding. The boron complexes [L2(B2)] (L=PMe3 and NHCMe) possess a linear arrangement of the LBBL moiety, which has a B?B triple bond. The heavier homologues [L2(E2)] have antiperiplanar arrangements of the LEEL moieties, except for [(PMe3)2(In2)], which has a twisted structure in which the PInInP torsion angle is 123.0°. The structural features of the complexes [L(EH3)] and [L2(E2Hn)] can be explained in terms of donor–acceptor interactions between the donors L and the acceptors EH3 and E2Hn, which have been analyzed quantitatively by using the energy decomposition analysis (EDA) method. The calculations predict that the hydrogenation reaction of the dimeric magnesium(I) compound L′MgMgL′ with the complexes [L(EH3)] is energetically more favorable for L=PMe3 than for NHCMe.  相似文献   

12.
In 1‐adamantyl‐2,8,9‐trioxa‐5‐aza‐1‐germabicyclo­[3.3.3]undecane or 1‐adamantylgermatrane, [Ge(C10H15)(C6H12NO3)], (I), and (2,8,9‐trioxa‐5‐aza‐1‐germabicyclo­[3.3.3]undecan‐1‐yl)methyl N‐cyclo­hexyl­carbamate or [(germatran‐1‐yl)meth­yl] N‐cyclo­hexyl­carbamate, [Ge(C6H12NO3)(C8H14NO2)], (II), the Ge atoms are characterized by trigonal–bypiramidal configurations. The Ge⋯N distances [2.266 (3) and 2.206 (3) Å in (I) and (II), respectively] are among the longest observed in germatranes. The significant distortion of the apical N—Ge—C angle in (II) is caused by crystal packing effects.  相似文献   

13.
Only a few cyclooctatetraene dianion (COT) π‐complexes of lanthanides have been crystallographically characterized. This first single‐crystal X‐ray diffraction characterization of a scandium(III) COT chloride complex, namely di‐μ‐chlorido‐bis[(η8‐cyclooctatetraene)(tetrahydrofuran‐κO )scandium(III)], [Sc2(C8H8)2Cl2(C4H8O)2] or [Sc(COT)Cl(THF)]2 (THF is tetrahydrofuran), (1), reveals a dimeric molecular structure with symmetric chloride bridges [average Sc—Cl = 2.5972 (7) Å] and a η8‐bound COT ligand. The COT ring is planar, with an average C—C bond length of 1.399 (3) Å. The Sc—C bond lengths range from 2.417 (2) to 2.438 (2) Å [average 2.427 (2) Å]. Direct comparison of (1) with the known lanthanide (Ln) analogues (La, Ce, Pr, Nd, and Sm) illustrates the effect of metal‐ion (M ) size on molecular structure. Overall, the M —Cl, M —O, and M —C bond lengths in (1) are the shortest in the series. In addition, only one THF molecule completes the coordination environment of the small ScIII ion, in contrast to the previously reported dinuclear Ln–COT–Cl complexes, which all have two bound THF molecules per metal atom.  相似文献   

14.
The reaction mechanism of the elimination of CH3EH3 from the platinum complexes cis‐[Pt(CH3) · (EH3)(PH3)2] (E = Si, Ge) in the presence of acetylene has been studied using gradient‐corrected DFT calculations at the B3LYP level. The reaction proceeds in two steps. The first step is the formation of the acetylene complex [Pt(CH3)(HCCH)(EH3)(PH3)] which occurs in a associative/dissociate pathway via the five‐coordinated intermediate [Pt(CH3)(HCCH)(EH3)(PH3)2]. The rate‐determining step is the elimination of CH3EH3 via a four‐coordinated transition state. The alternative mechanism via direct dissociation from the five‐coordinated intermediates has higher activation barriers. The calculated activation energies of the model reactions are in good agreement with experimental results. The silyl complex has a lower barrier for the elimination reaction than the germyl complex. The calculated transition states show that the reason for the lower barrier is the strength of the nascending C–Si bond, which is higher than the C–Ge bond. The results are in agreement with the postulated mechanism of Ozawa et al. (Organometallics, 1998 , 17, 1018).  相似文献   

15.
The hydrometallation of the iminoboranes XB(NtBu) ( 1 a : X = tBu; 1 b : X = tBu(Me3Si)N) with Cp2ZrHCl and Cp2HfHCl gives products of the type X–BH=N(tBu)–MCp2–Cl ( 7 a , b : M = Zr; 8 a , b : M = Hf). There is a B–H–M (3c2e) bond interaction. The BN multiple bond of the iminoboranes is more or less side‐on bound to the metal. Hence, iminoboranes again turn out to behave as analogues of alkynes.  相似文献   

16.
The first thermally robust Ge II −Sn II compound 1 and the structurally characterized SnII-SnII analogue 2 , which maintain their structural integrity in solution, were obtained by treating MAr2 (M=Ge, Sn; Ar=2,6-(Me2N)2C6H3) with Sn[1,8-(NR2)2C10H6] (R=CH2tBu). On the basis of structural and spectroscopic data, the M−Sn bond is regarded as the interaction of a MAr2 donor with an Sn[1,8-(NR2)2C10H6] acceptor.  相似文献   

17.
Eight Cs‐symmetric complexes, R1R2C(Cp)(Flu)MCl2 [R1 = R2 = CH3CH2CH2, M = Zr (1), Hf (2); R1 = R2 = p? CH3OC6H4, M = Zr (3), Hf (4); R1 = p? tBuC6H4, R2 = Ph, M = Zr (5), Hf (6); R1 = R2 = p? tBuC6H4, M = Zr (7); R1 = R2 = PhCH2, M = Zr (8)] have been synthesized and characterized. Zirconocenes all showed the same high catalytic activities in ethylene polymerization as complex Ph2C(Cp)(Flu)ZrCl2 (9). However, in the propylene polymerization, the catalytic activities decreased in the order 5 ≈ 9 > 7 > 8. Introduction of tBu decreased the activities, probably due to the bulk steric hindrance. The polypropylene produced by 5 and 7 with tBu substituent showed a higher molecular weight (Mη) than that produced by 9. The 13C NMR spectrum revealed the polymers from 7 and 8 to have shorter average syndiotactic block length than polymer produced by 9. It was noted that [mm] stereodefect of polypropylene by 8 could not be observed from 13C NMR, which showed that the benzyl on bridge carbon 8 prevented chain epimerization and enatiofacial misinsertion in polymerization. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

18.
The polyfluorinated title compounds, [M Cl2(C16H16F4N2O2)] or [4,4′‐(HCF2CH2OCH2)2‐2,2′‐bpy]M Cl2 [M = Pd, ( 1 ), and M = Pt, ( 2 )], have –C(Hα)2OC(Hβ)2CF2H side chains with H‐atom donors at the α and β sites. The structures of ( 1 ) and ( 2 ) are isomorphous, with the nearly planar (bpy)M Cl2 molecules stacked in columns. Within one column, π‐dimer pairs alternate between a π‐dimer pair reinforced with C—H…Cl hydrogen bonds (α,α) and a π‐dimer pair reinforced with C—Hβ…F(—C) interactions (abbreviated as C—Hβ…F—C,C—Hβ…F—C). The compounds [4,4′‐(CF3CH2OCH2)2‐2,2′‐bpy]M Cl2 [M = Pd, ( 3 ), and M = Pt, ( 4 )] have been reported to be isomorphous [Lu et al. (2012). J. Fluorine Chem. 137 , 54–56], yet with disorder in the fluorous regions. The molecules of ( 3 ) [or ( 4 )] also form similar stacks, but with alternating π‐dimer pairs between the (α,β; α,β) and (β,β) forms. Through (C—)H…Cl hydrogen‐bond interactions, one molecule of ( 1 ) [or ( 2 )] is expanded into an aggregate of two inversion‐related π‐dimer pairs, one pair in the (α,α) form and the other pair in the (C—Hβ…F—C,C—Hβ…F—C) form, with the plane normals making an interplanar angle of 58.24 (3)°. Due to the demands of maintaining a high coordination number around the metal‐bound Cl atoms in molecule ( 1 ) [or ( 2 )], the ponytails of molecule ( 1 ) [or ( 2 )] bend outward; in contrast, the ponytails of molecule ( 3 ) [or ( 4 )] bend inward.  相似文献   

19.
The diaminosilylene tBuNCH2CH2NtBuSi: reacted with the diaminogermylenes RNCH2CH2NRGe: R = 2,6-Me2C6H3, iPr, by silylene insertion into one of the Ge–N bonds to furnish the aminosilylgermylenes 8 , R = 2,6-Me2C6H3, and 9 , R = iPr. The X-ray structure analyses of these compounds revealed that 8 remains monomeric in the crystal with weak Ge … Ge interactions to the germanium atom of a neighbouring germylene molecule, whereas 9 dimerizes to give the strongly twisted (E)-1,2-diamino-1,2-disilyldigermene (E)- 10 with a long Ge–Ge double bond of 246 pm and a large trans-bent angle of 47.3°.  相似文献   

20.
In this study, the electronic structures and optical properties of a cyclometalated Pt(II) complex (M1) and a series of derivatives (M1–F, M1–CF3, and M1–CN) with electron-withdrawing substituents (–F, –CF3, and –CN) at the carbazole moiety were theoretically investigated by density functional theory and time-dependent density functional theory. The calculation results reveal that these Pt complexes display deep red phosphorescence emission above Λ = 640 nm. When the 3MLCT/π → π* to triplet metal-centered 3MC/d–d state decay mechanism is taken into consideration, the nonradiative decay rate constant (knr) decreased in the order M1 > M1–CF3 > M1–F > M1–CN. The <T1|HSOC|Sm> and kr values of M1-F are similar with those of M1, however the Knr rate ofM1-F is larger than that of M1. M1–F is expected to have improved quantum yields. Moreover, through the analyses of the HOMO/LUMO level and triplet energy, it is found that the introduction of –F and –CN substituents in M1 results in efficient energy transfer from the host material 4,4′-N,N′-dicarbazole-biphenyl to these complexes. In view of the electroluminescent applications in organic light-emitting diodes, M1–F can serve as efficient deep-red guest materials with improved electron injection and transport ability.  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