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.
A series of zinc(II) silylenes was prepared by using the silylene {PhC(NtBu)2}(C5Me5)Si. Whereas reaction of the silylene with ZnX2 (X=Cl, I) gave the halide‐bridged dimers [{PhC(NtBu)2}(C5Me5)SiZnX(μ‐X)]2, with ZnR2 (R=Ph, Et, C6F5) as reagent the monomers [{PhC(NtBu)2}(C5Me5)SiZnR2] were obtained. The stability of the complexes and the Zn?Si bond lengths clearly depend on the substitution pattern of the zinc atom. Electron‐withdrawing groups stabilize these adducts, whereas electron‐donating groups destabilize them. This could be rationalized by quantum chemical calculations. Two different bonding modes in these molecules were identified, which are responsible for the differences in reactivity: 1) strong polar Zn?Si single bonds with short Zn?Si distances, Zn?Si force constants close to that of a classical single bond, and strong binding energy (ca. 2.39 Å, 1.33 mdyn Å?1, and 200 kJ mol?1), which suggest an ion pair consisting of a silyl cation with a Zn?Si single bond; 2) relatively weak donor–acceptor Zn?Si bonds with long Zn?Si distances, low Zn?Si force constants, and weak binding energy (ca. 2.49 Å, 0.89 mdyn Å?1, and 115 kJ mol?1), which can be interpreted as a silylene–zinc adduct.  相似文献   

2.
The Element-Nitrogen Double Bond in Cations of Cyclic Bis(amino)phospha-, -arsa-, -stiba-, and bismuthines Bis(amino)sila-phospha, -arsa-, -stiba-, and bismaetidines which bear a positive charge and incorporate a formally two valent element of group V are obtained from the corresponding bis(amino)elementchlorides by transfer of the chloride anion to the Lewis acids AlCl3, GaCl3, and InCl3. X-ray structure analyses on the compounds Me2Si(NtBu)2P+AlCl4- ( 2a ), Me2Si(NtBu)2Sb+AlCl4- ( 2c ), and Me2Si(NtBu)2Bi+AlCl4 -( 2d ) reveal that the electron lack at the element(V) can be compensated by two different bonding mechanisms. In the case of the phosphorus derivative 2a the electronic balance is accomplished by intramolecular backbonding from the neighbouring nitrogen atoms (mean N-P+ = 163.3 pm). In the antimony and bismuth derivatives ( 2c and 2d ) the chlorine atoms of the AlCl4 anions coordinate to the unsaturated element(V) in an intermolecular manner (mean: Sb·Cl = 305, Bi ?Cl = 309 pm). The N? Si? N group which is identical in all molecules, may be used as a probe for the electronic balance within the ring systems. The bond lengths and angles vary dramatically with respect to the electron acceptor properties of the element to which the group is bonded. 2a forms orthorhombic crystals, space group Pnma, Z = 8 (a = 3023.7(9), b = 1001.0(3), c = 1414.6(5) pm), and 2c and 2d are isotypic, again orthorhombic, space group Pbca with Z = 8 ( 2c a = 2030.8(8), b = 1193.1(4), c = 1777.1(6) pm; 2d : a = 2025.9(8), b = 1198.0(4), c = 1761.3(6) pm).  相似文献   

3.
The reaction of monomeric [(TptBu,Me)LuMe2] (TptBu,Me=tris(3‐Me‐5‐tBu‐pyrazolyl)borate) with primary aliphatic amines H2NR (R=tBu, Ad=adamantyl) led to lutetium methyl primary amide complexes [(TptBu,Me)LuMe(NHR)], the solid‐state structures of which were determined by XRD analyses. The mixed methyl/tetramethylaluminate compounds [(TptBu,Me)LnMe({μ2‐Me}AlMe3)] (Ln=Y, Ho) reacted selectively and in high yield with H2NR, according to methane elimination, to afford heterobimetallic complexes: [(TptBu,Me)Ln({μ2‐Me}AlMe2)(μ2‐NR)] (Ln=Y, Ho). X‐ray structure analyses revealed that the monomeric alkylaluminum‐supported imide complexes were isostructural, featuring bridging methyl and imido ligands. Deeper insight into the fluxional behavior in solution was gained by 1H and 13C NMR spectroscopic studies at variable temperatures and 1H–89Y HSQC NMR spectroscopy. Treatment of [(TptBu,Me)LnMe(AlMe4)] with H2NtBu gave dimethyl compounds [(TptBu,Me)LnMe2] as minor side products for the mid‐sized metals yttrium and holmium and in high yield for the smaller lutetium. Preparative‐scale amounts of complexes [(TptBu,Me)LnMe2] (Ln=Y, Ho, Lu) were made accessible through aluminate cleavage of [(TptBu,Me)LnMe(AlMe4)] with N,N,N′,N′‐tetramethylethylenediamine (tmeda). The solid‐state structures of [(TptBu,Me)HoMe(AlMe4)] and [(TptBu,Me)HoMe2] were analyzed by XRD.  相似文献   

4.
Abstract

Chemistry of twocoordinated phosphorus compounds has been rapidly developed in last years. Parallel with synthesis of new such compounds and investigating of their chemistry research of their spatial and electron structure assumes ever greater importanse. Electrical methods are very informative in this aspect and also permit to determine unknown polaritirs of PII-bonds, that have great self-dependent interest. We have been studed for the first time the acyclic PII-compounds containing P=N, P=C, P=P, P=As, P=Sb - double bonds. Phosphaazenes RIP=NR2, where RI: (Me3Si)2N, 2,4,6-tBu3C6H2; R2: Me3Si, tBu, 2,4,6-tBu3C6H2 have small polarity 1.5–2.0 D, wich describes by one set of parameteres. The main role in stabilization of -P=N-fragment having E-configuration belongs to sterical factors.  相似文献   

5.
The synthesis of a 1,2,3,4-tetramethylcyclopentadienyl (Cp4) substituted four-membered N-heterocyclic silylene [{PhC(NtBu)2}Si(C5Me4H)] is reported first. Then, selected reactions with transition metal and a calcium precursor are shown. The proton of the Cp4-unit is labile. This results in two different reaction pathways: (1) deprotonation and (2) rearrangement reactions. Deprotonation was achieved by the reaction of [{PhC(NtBu)2}Si(C5Me4H)] with suitable zinc precursors. Rearrangement to [{PhC(NtBu)2}(C5Me4)SiH], featuring a formally tetravalent silicon R2C Created by potrace 1.16, written by Peter Selinger 2001-2019 Si(R′)–H unit, was observed when the proton of the Cp4 ring was shifted from the Cp4-ring to the silylene in the presence of a Lewis acid. This allows for the coordination of the Cp4-ring to a calcium compound. Furthermore, upon reaction with transition metal dimers [MCl(cod)]2 (M = Rh, Ir; cod = 1,5-cyclooctadiene) the proton stays at the Cp4-ring and the silylene reacts as a sigma donor, which breaks the dimeric structure of the precursors.

A cyclopentadienyl functionalized silylene or its derivatives can be coordinated in all three forms: silylene (A), anion (B), and sila fulvene (C).  相似文献   

6.
The syntheses of the transition metal complexes cis‐[(4‐tBu‐2,6‐{P(O)(OiPr)2}2C6H2SnCl)2MX2] ( 1 , M=Pd, X=Cl; 2 , M=Pd, X=Br; 3 , M=Pd, X=I; 4 , M=Pt, X=Cl), cis‐[{2,6‐(Me2NCH2)2C6H3SnCl}2MX2] ( 5 , M=Pd, X=I; 6 , M=Pt, X=Cl), trans‐[{2,6‐(Me2NCH2)2C6H3SnI}2PtI2] ( 7 ) and trans‐[(4‐tBu‐2,6‐{P(O)(OiPr)2}2 C6H2SnCl)PdI2]2 ( 8 ) are reported. Also reported is the serendipitous formation of the unprecedented complexes trans‐[(4‐tBu‐2,6‐{P(O)(OiPr)2}2C6H2SnCl)2 Pt(SnCl3)2] ( 10 ) and [(4‐tBu‐2,6‐{P(O) (OiPr)2}2C6H2SnCl)3Pt(SnCl3)2] ( 11 ). The compounds were characterised by elemental analyses, 1H, 13C, 31P, 119Sn and 195Pt NMR spectroscopy, single‐crystal X‐ray diffraction analysis, UV/Vis spectroscopy and, in the cases of compounds 1 , 3 and 4 , also by Mössbauer spectroscopy. All the compounds show the tin atoms in a distorted trigonal‐bipyramidal environment. The Mössbauer spectra suggest the tin atoms to be present in the oxidation state III. The kinetic lability of the complexes was studied by redistribution reactions between compounds 1 and 3 as well as between 1 and cis‐[{2,6‐(Me2NCH2)2C6H3SnCl}2PdCl2]. DFT calculations provided insights into both the bonding situation of the compounds and the energy difference between the cis and trans isomers. The latter is influenced by the donor strength of the pincer‐type ligands.  相似文献   

7.
In reactions with transition metal compounds, tBu2P? P?P(X)tBu2 (X = Br, Me) acts mainly as a precursor of the tBu2P? P ligand, whereas tBu(Me3Si)P? P?P(Me)tBu2 acts as a precursor of the (Me3Si)P?PtBu ligand. Up to now, only Pt(0) d10 ML2 metal centres were found to be able to stabilize the tBu2P? P group in ‘pure form’ by means of η2‐coordination (side on). Several compounds of the [{η2 ? tBu2P? P}Pt(PR3)2] type were sufficiently stable to be isolated and characterized; however, not all of them gave single crystals suitable for X‐ray structure determinations. The X‐ray structures of these compounds and of [{µ ? (1,2:2 ? η ? tBu2P? P)Pt(PR3)2} {M(CO)5}] strongly suggest the ethene‐like form of 1,1‐di‐tert‐butyldiphosphene in these complexes. Such a form is also in agreement with RI DFT calculations with SVP basis for free tBu2P? P. However, in trapping experiments with cyclic olefins and cyclic dienes tBu2P? P exhibits, to some extent, electrophilic ‘singlet carbene’ properties. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

8.
Treatment of [{Me2C6H(CH2PtBu2)2}Rh(η1‐N2)] ( 1a ) with molecular oxygen (O2) resulted in almost quantitative formation of the dioxygen adduct [{Me2C6H(CH2PtBu2)2}Rh(η2‐O2)] ( 2a ). An X‐ray diffraction study of 2a revealed the shortest O? O bond reported for Rh? O2 complexes, indicating the formation of a RhI? O2 adduct, rather than a cyclic RhIII η2‐peroxo complex. The coordination of the O2 ligand in 2a was shown to be reversible. Treatment of 2a with CO gas yielded almost quantitatively the corresponding carbonyl complex [{Me2C6H(CH2PtBu2)2}Rh(CO)] ( 3a ). Surprisingly, treatment of the structurally very similar pincer complex [{C6H3(CH2PiPr2)2)}Rh(η1‐N2)] ( 1b ) with O2 led to partial decomposition, with no dioxygen adduct being observed.  相似文献   

9.
Diimido, Imido Oxo, Dioxo, and Imido Alkylidene Halfsandwich Compounds via Selective Hydrolysis and α—H Abstraction in Molybdenum(VI) and Tungsten(VI) Organyl Complexes Organometal imides [(η5‐C5R5)M(NR′)2Ph] (M = Mo, W, R = H, Me, R′ = Mes, tBu) 4 — 8 can be prepared by reaction of halfsandwich complexes [(η5‐C5R5)M(NR′)2Cl] with phenyl lithium in good yields. Starting from phenyl complexes 4 — 8 as well as from previously described methyl compounds [(η5‐C5Me5)M(NtBu)2Me] (M = Mo, W), reactions with aqueous HCl lead to imido(oxo) methyl and phenyl complexes [(η5‐C5Me5)M(NtBu)(O)(R)] M = Mo, R = Me ( 9 ), Ph ( 10 ); M = W, R = Ph ( 11 ) and dioxo complexes [(η5‐C5Me5)M(O)2(CH3)] M = Mo ( 12 ), M = W ( 13 ). Hydrolysis of organometal imides with conservation of M‐C σ and π bonds is in fact an attractive synthetic alternative for the synthesis of organometal oxides with respect to known strategies based on the oxidative decarbonylation of low valent alkyl CO and NO complexes. In a similar manner, protolysis of [(η5‐C5H5)W(NtBu)2(CH3)] and [(η5‐C5Me5)Mo(NtBu)2(CH3)] by HCl gas leads to [(η5‐C5H5)W(NtBu)Cl2(CH3)] 14 und [(η5‐C5Me5)Mo(NtBu)Cl2(CH3)] 15 with conservation of the M‐C bonds. The inert character of the relatively non‐polar M‐C σ bonds with respect to protolysis offers a strategy for the synthesis of methyl chloro complexes not accessible by partial methylation of [(η5‐C5R5)M(NR′)Cl3] with MeLi. As pure substances only trimethyl compounds [(η5‐C5R5)M(NtBu)(CH3)3] 16 ‐ 18 , M = Mo, W, R = H, Me, are isolated. Imido(benzylidene) complexes [(η5‐C5Me5)M(NtBu)(CHPh)(CH2Ph)] M = Mo ( 19 ), W ( 20 ) are generated by alkylation of [(η5‐C5Me5)M(NtBu)Cl3] with PhCH2MgCl via α‐H abstraction. Based on nmr data a trend of decreasing donor capability of the ligands [NtBu]2— > [O]2— > [CHR]2— ? 2 [CH3] > 2 [Cl] emerges.  相似文献   

10.
Deprotonation of aminophosphaalkenes (RMe2Si)2C?PN(H)(R′) (R=Me, iPr; R′=tBu, 1‐adamantyl (1‐Ada), 2,4,6‐tBu3C6H2 (Mes*)) followed by reactions of the corresponding Li salts Li[(RMe2Si)2C?P(M)(R′)] with one equivalent of the corresponding P‐chlorophosphaalkenes (RMe2Si)2C?PCl provides bisphosphaalkenes (2,4‐diphospha‐3‐azapentadienes) [(RMe2Si)2C?P]2NR′. The thermally unstable tert‐butyliminobisphosphaalkene [(Me3Si)2C?P]2NtBu ( 4 a ) undergoes isomerisation reactions by Me3Si‐group migration that lead to mixtures of four‐membered heterocyles, but in the presence of an excess amount of (Me3Si)2C?PCl, 4 a furnishes an azatriphosphabicyclohexene C3(SiMe3)5P3NtBu ( 5 ) that gave red single crystals. Compound 5 contains a diphosphirane ring condensed with an azatriphospholene system that exhibits an endocylic P?C double bond and an exocyclic ylidic P(+)? C(?)(SiMe3)2 unit. Using the bulkier iPrMe2Si substituents at three‐coordinated carbon leads to slightly enhanced thermal stability of 2,4‐diphospha‐3‐azapentadienes [(iPrMe2Si)2C?P]2NR′ (R′=tBu: 4 b ; R′=1‐Ada: 8 ). According to a low‐temperature crystal‐structure determination, 8 adopts a non‐planar structure with two distinctly differently oriented P?C sites, but 31P NMR spectra in solution exhibit singlet signals. 31P NMR spectra also reveal that bulky Mes* groups (Mes*=2,4,6‐tBu3C6H2) at the central imino function lead to mixtures of symmetric and unsymmetric rotamers, thus implying hindered rotation around the P? N bonds in persistent compounds [(RMe2Si)2C?P]2NMes* ( 11 a , 11 b ). DFT calculations for the parent molecule [(H3Si)2C?P]2NCH3 suggest that the non‐planar distortion of compound 8 will have steric grounds.  相似文献   

11.
Two potassium–dialkyl–TMP–zincate bases [(pmdeta)K(μ‐Et)(μ‐tmp)Zn(Et)] ( 1 ) (PMDETA=N,N,N′,N′′,N′′‐pentamethyldiethylenetriamine, TMP=2,2,6,6‐tetramethylpiperidide), and [(pmdeta)K(μ‐nBu)(μ‐tmp)Zn(nBu)] ( 2 ), have been synthesized by a simple co‐complexation procedure. Treatment of 1 with a series of substituted 4‐R‐pyridines (R=Me2N, H, Et, iPr, tBu, and Ph) gave 2‐zincated products of the general formula [{2‐Zn(Et)2‐μ‐4‐R‐C5H3N}2 ? 2{K(pmdeta)}] ( 3 – 8 , respectively) in isolated crystalline yields of 53, 16, 7, 23, 67, and 51 %, respectively; the treatment of 2 with 4‐tBu‐pyridine gave [{2‐Zn(nBu)2‐μ‐4‐tBu‐C5H3N}2 ? 2{K(pmdeta)}] ( 9 ) in an isolated crystalline yield of 58 %. Single‐crystal X‐ray crystallographic and NMR spectroscopic characterization of 3 – 9 revealed a novel structural motif consisting of a dianionic dihydroanthracene‐like tricyclic ring system with a central diazadicarbadizinca (ZnCN)2 ring, face‐capped on either side by PMDETA‐wrapped K+ cations. All the new metalated pyridine complexes share this dimeric arrangement. As determined by NMR spectroscopic investigations of the reaction filtrates, those solutions producing 3 , 7 , 8 , and 9 appear to be essentially clean reactions, in contrast to those producing 4 , 5 , and 6 , which also contain laterally zincated coproducts. In all of these metalation reactions, the potassium–zincate base acts as an amido transfer agent with a subsequent ligand‐exchange mechanism (amido replacing alkyl) inhibited by the coordinative saturation, and thus, low Lewis acidity of the 4‐coordinate Zn centers in these dimeric molecules. Studies on analogous trialkyl–zincate reagents in the absence and presence of stoichiometric or substoichiometric amounts of TMP(H) established the importance of Zn? N bonds for efficient zincation.  相似文献   

12.
Herein, new complexes containing the [Ph2PCH2S(NtBu)3]? anion are presented, supplying three imido nitrogen atoms and a remote phosphorus atom as potential donor sites to main group and transition‐metal cations. The lithiated complex [(tmeda)Li{(NtBu)3SCH2PPh2}] ( 1 ) is an excellent starting material in transmetalation reactions. Herein, the transition‐metal complexes [M{(NtBu)3SCH2PPh2}2] (M=Mn ( 2 ), Ni ( 3 ), Zn ( 4 )) were synthesized and structurally characterized. Their isotypical molecules show SN2 chelation and no employment of the adjacent phosphorus atom in coordination. The third pendent imido group is always twisted toward the vacant face of the tetrahedrally coordinated sulfur atom.  相似文献   

13.
The potential energy surfaces of both neutral and dianionic SnC2P2R2 (R=H, tBu) ring systems have been explored at the B3PW91/LANL2DZ (Sn) and 6‐311+G* (other atoms) level. In the neutral isomers the global minimum is a nido structure in which a 1,2‐diphosphocyclobutadiene ring (1,2‐DPCB) is capped by the Sn. Interestingly, the structure established by X‐ray diffraction analysis, for R=tBu, is a 1,3‐DPCB ring capped by Sn and it is 2.4 kcal mol?1 higher in energy than the 1,2‐DPCB ring isomer. This is possibly related to the kinetic stability of the 1,3‐DPCB ring, which might originate from the synthetic precursor ZrCp2tBu2C2P2. In the case of the dianionic isomers we observe only a 6π‐electron aromatic structure as the global minimum, similarly to the cases of our previously reported results with other types of heterodiphospholes. 1 , 4 , 19 The existence of large numbers of cluster‐type isomers in neutral and 6π‐planar structures in the dianions SnC2P2R22? (R=H, tBu) is due to 3D aromaticity in neutral clusters and to 2D π aromaticity of the dianionic rings. Relative energies of positional isomers mainly depend on: 1) the valency and coordination number of the Sn centre, 2) individual bond strengths, and 3) the steric effect of tBu groups. A comparison of neutral stannadiphospholes with other structurally related C5H5+ analogues indicates that Sn might be a better isolobal analogue to P+ than to BH or CH+. The variation in global minima in these C5H5+ analogues is due to characteristic features such as 1) the different valencies of C, B, P and Sn, 2) the electron deficiency of B, 3) weaker pπ–pπ bonding by P and Sn atoms, and 4) the tendency of electropositive elements to donate electrons to nido clusters. Unlike the C5H5+ systems, all C5H5? analogues have 6π‐planar aromatic structures as global minima. The differences in the relative ordering of the positional isomers and ligating properties are significant and depend on 1) the nature of the π orbitals involved, and 2) effective overlap of orbitals.  相似文献   

14.
The generation of heavier double‐bond systems without by‐ or side‐product formation is of considerable importance for their application in synthesis. Peripheral functional groups in such alkene homologues are promising in this regard owing to their inherent mobility. Depending on the steric demand of the N‐alkyl substituent R, the reaction of disilenide Ar2Si?Si(Ar)Li (Ar=2,4,6‐iPr3C6H2) with ClP(NR2)2 either affords the phosphinodisilene Ar2Si?Si(Ar)P(NR2)2 (for R=iPr) or P‐amino functionalized phosphasilenes Ar2(R2N)Si? Si(Ar)?P(NR2) (for R=Et, Me) by 1,3‐migration of one of the amino groups. In case of R=Me, upon addition of one equivalent of tert‐butylisonitrile a second amino group shift occurs to yield the 1‐aza‐3‐phosphaallene Ar2(R2N)Si? Si(NR2)(Ar)? P?C?NtBu with pronounced ylidic character. All new compounds were fully characterized by multinuclear NMR spectroscopy as well as single‐crystal X‐ray diffraction and DFT calculations in selected cases.  相似文献   

15.
Amination of the C‐isopropyldimethylsilyl P‐chlorophosphaalkene (iPrMe2Si)2C=PCl ( 1 ) leads to the P‐aminophosphaalkenes (iPrMe2Si)2C=PN(R)R′ (R, R′ = Me ( 2 ), R = H, R′ = nPr ( 3 ), R = H, R′ = iPr ( 4 ), R = H, R′ = tBu ( 5 ), R = H, R′ = 1‐Ada ( 6 ), R = H, R′ = CPh3 ( 7 ), R = H, R′ = Ph ( 8 ), R = H, RR′ = 2,6‐iPr2Ph (= DIP) ( 10 ), R = H, R′ = 2,4,6‐Me3Ph (= Mes) ( 11 ), R = H, R′ = 2,4,6‐tBu3Ph (= Mes*)] ( 12 ), R = H, R′ = SiMe3 ( 13 ), and R, R′ = SiMe2Ph (1 4 ). 31P‐NMR spectra confirm that phosphaalkenes 2 – 7 and 10 – 14 are monomeric in solution; the structures of 7 , 10 , and 12 were determined by X‐ray crystallography. Freshly prepared (iPrMe2Si)2C=PN(H)Ph ( 8 ) is a monomer that dimerizes with (N→C) proton migration within several hours to the stable diazadiphosphetidine [(iPrMe2Si)2CHPNPh]2 ( 9 ). NMR‐scale reactions of deprotonated 5 and 13 with tBuiPrPCl provide by P–P bond formation the P‐phosphanyl iminophosphoranes [(iPrMe2Si)2C=](RN=)PPtBu(iPr) [R = tBu ( 15 ), R = Me3Si ( 17 )]. Deprotonated 5 and Me3GeCl deliver by N–Ge bond formation the aminophosphaalkene (iPrMe2Si)2C=PN(tBu)GeMe3 ( 20 ), which with elemental selenium 5 undergoes (N→C) proton migration to form the alkyl(imino)(seleno)phosphorane [(iPrMe2Si)2CH](tBuN=)P=Se ( 21 ), which is a selenium‐bridged cyclic dimer in the solid state.  相似文献   

16.
Treatment of R2SiCl2 (R = Me, Ph) with 2‐aminopyridine in the presence of NEt3 led to the formation of the bis(N‐2‐pyridylamino)silanes R2Si{NH(2‐Py)}2, which were isolated as pale yellow solids. Crystal structure analyses revealed that both compounds exhibit tetrahedrally coordinated silicon atoms which are linked to two 2‐pyridylamido moieties and two organyl groups (Me or Ph). As a result of intermolecular hydrogen bonding between the NH groups and the pyridyl N atoms the R2Si{NH(2‐Py)}2 molecules are catenated in the solid state. Treatment of R2Si{NH(2‐Py)}2 with nBuLi afforded the corresponding amides R2Si{NLi(2‐Py)}2, which were subsequently reacted with MCl2 (M = Sn, Pb) to give the dinuclear silylamides [{R2Si(N‐2‐Py)2M}2]. Both the tin and the lead derivatives exhibit closely related molecular structures, in which the tin (or lead) atoms are linked to two amido N atoms and a pyridyl N atom in a distorted trigonal bipyramidal coordination mode.  相似文献   

17.
Although series of N1, N1‐dimethyl‐N2‐arylformamidines and of 1,1,3,3‐tetraalkyl‐2‐arylguanidines are structurally analogous and similar electron‐ionization mass spectral fragmentation may be expected, they display important differences in the favored routes of fragmentation and consequently in substituent effects on ion abundances. In the case of formamidines, the cyclization‐elimination process (initiated by nucleophilic attack of the N‐amino atom on the 2‐position of the phenyl ring) and formation of the cyclic benzimidazolium [M‐H]+ ions dominates, whereas the loss of the NR2 group is more favored for guanidines. In order to gain information on the most probable structures of the principal fragments, quantum‐chemical calculations were performed on a selected set. A good linear relation between log{I[M‐H]+I [M]+?} and σR+ constants of substituent at para position in the phenyl ring occurs solely for formamidines (r = 0.989). In the case of guanidines, this relation is not significant (r = 0.659). A good linear relation is found between log{I[M‐NMe2]+/I [M]+?} and σp+ constants (r = 0.993). Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

18.
Element-Organic Amine/Imine Compounds, XXXI. - Cyclometallation with N-tert-Butyl-Phosphorus-Nitrogen Iridium Complexes The interaction of R1R2N–PNR3 ( 1 ) (R1  SiMe3, tBu, iC3H7; R2  R3  SiMe3, tBu) with [M(COD)(μ-Cl)]2 ( 2 ), M  Rh, Ir, affords the amino(imino)phosphane complexes 3 , whose PN bond adds methanol with formation of the diamidophosphite complexes 4 . Already below 0°C the iridium compounds of 4 undergo cyclometallation of a tBu methyl group (R2) with formation of the hydrido-iridium metallaheterocycles 5 . The structures of 4b and 5a are elucidated by X-ray analyses.  相似文献   

19.
Abstract. The cyclopentadienyl‐substituted iron‐bismuth complexes [{Cp(CO)2Fe}BiCl2] ( 1 ), [{Cp(CO)2Fe}BiBr2] ( 2 ), [{Cp′′(CO)2Fe}BiBr2] ( 3 ) and [{Cp*(CO)2Fe}BiBr2] ( 4 ) were prepared with high yields starting from [Cpx(CO)2Fe]2 [Cpx = C5H5 (Cp), C5H3‐1, 3‐tBu2 (Cp′′), C5Me5 (Cp*)] and the corresponding bismuth halides. The single crystal X‐ray structure analyses of compounds 2 – 4 are reported. Comparison of their solubility demonstrates that the steric hindrance in this type of compounds is only slightly higher for compound 3 compared with compound 2 but significantly lower compared with the Cp* derivative 4 . Compounds 1 – 4 react with nucleophililic reagents such as KOtBu, NaOCH2CH2OCH3, and NaOSiMe3 as well as with water in the presence of an amine to give a mixture of [{Cpx(CO)2Fe}BiX] (X = Cl, Br) and [{Cpx(CO)2Fe}3Bi]. In case of a reaction with nBu4NCl and DMAP (dimethylaminopyridine) no such dismutation is observed. Instead the complexes [{Cp(CO)2Fe}BiBr2(DMAP)2] ( 5 ), [NnBu4]2[{{Cp(CO)2Fe}BiBr3}2] ( 6 ) and [NnBu4]2[{{Cp(CO)2Fe}BiCl3}2] ( 7 ) were isolated and characterized by single‐crystal X‐ray diffraction.  相似文献   

20.
Reactions of the beryllium dihalide complexes [BeX2(OEt2)2] (X=Br or I) with N,N,N′,N′‐tetramethylethylenediamine (TMEDA), a series of diazabutadienes, or bis(diphenylphosphino)methylene (DPPM) have yielded the chelated complexes, [BeX2(TMEDA)], [BeX2{(RN=CH)2}] (R=tBu, mesityl (Mes), 2,6‐diethylphenyl (Dep) or 2,6‐diisopropylphenyl (Dip)), and the non‐chelated system, [BeI21P‐DPPM)2]. Reactions of lithium or potassium salts of a variety of β‐diketiminates have given both three‐coordinate complexes, [{HC(RCNAr)2}BeX] (R=H or Me; Ar=Mes, Dep or Dip; X=Br or I); and four‐coordinate systems, [{HC(MeCNPh)2}BeBr(OEt2)] and [{HC(MeCNDip)(MeCNC2H4NMe2}BeI]. Alkali metal salts of ketiminate, guanidinate, boryl/phosphinosilyl amide, or terphenyl ligands, lead to dimeric [{BeI{μ‐[(OCMe)(DipNCMe)]CH}}2], and monomeric [{iPr2NC(NMes)2}BeI(OEt2)], [κ2N,P‐{(HCNDip)2B}(PPh2SiMe2)NBeI(OEt2)] and [{C6H3Ph2‐2,6}BeBr(OEt2)], respectively. Compound [{HC(MeCNPh)2}BeBr(OEt2)] undergoes a Schlenk redistribution reaction in solution, affording the homoleptic complex, [{HC(MeCNPh)2}2Be]. The majority of the prepared complexes have been characterized by X‐ray crystallography and multi‐nuclear NMR spectroscopy. The structures and stability of the complexes are discussed, as is their potential for use as precursors in poorly developed low oxidation state beryllium chemistry.  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