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 640 毫秒
1.
Linked bis(β-diketiminato) rare-earth metal complexes based on the ethylene-bridged ligand [C2H4(BDIDClP)2]H2 [DClP = 2,6-Cl2C6H3] and the cyclohexyl-bridged ligands [Cy(BDIAr)2]H2 [Ar = PMP (= p-MeOC6H4), Mes (= 2,4,6-Me3C6H2), DIPP (= 2,6-iPr2C6H3)] were prepared via amine elimination starting from [Ln{N(SiMe3)2}3] (Ln = La, Y). The three cyclohexyl-bridged complexes [{(R,R)-Cy(BDIMes)2}YN(SiMe3)2] ((R,R)-3), [{(R,R)-Cy(BDIMes)2}LaN(SiMe3)2] ((R,R)-4), and [{(R,R)-Cy(BDIDIPP)2}LaN(SiMe3)2] ((R,R)-5) were obtained enantiomerically pure. The X-ray crystal structure analysis of the racemic variants of 3 and 4 revealed a distorted square pyramidal coordination geometry around the rare-earth metal, in which the amido ligand occupies the apical position and the two linked β-diketiminato moieties form the basis. The two aromatic substituents adopt a transoid arrangement and both β-diketiminato moieties are bound in a η5 coordination mode with close Ln?C contacts. Due to the smaller ionic radius of yttrium vs. lanthanum, the front side of the yttrium complex 3 is sterically more hindered than in the lanthanum complex 4, but there is much more empty coordination space on the rear side, which may rationalize the observed differences in selectivity of 3 in comparison to 4. The catalytic efficiency of the β-diketiminato complexes was strongly affected by steric factors such as ionic radius of the metal and the steric bulk of the aryl substituents, which is an indication for highly steric encumbered catalytic species. The complexes displayed good to moderate catalytic activity in the hydroamination/cyclization of aminoalkenes depending on the steric hindrance around the metal center. The sterically most demanding diisopropylphenyl-substituted complex (R,R)-5 displayed significantly higher enantioselectivities (up to 76% ee), but lower catalytic activity in comparison to the sterically more open mesityl-substituted complex (R,R)-4. The smaller yttrium metal center in complex (R,R)-3 led to reduced activity as well as a reversal in enantioselectivity, which may be rationalized by a change of the approach of the alkene moiety to the Ln-amido bond in the cyclization transition state.  相似文献   

2.
A series of novel 2,6-bis(imino)pyridyl iron complexes {2,6-(2-X-4-Y-5-ZC6H2NCCH3)2C5H3N}FeCl2 (X = Cl, Y = CH3, Z = H (2); X = Br, Y = CH3, Z = H (3); X = F, Y = H, Z = CH3 (4); X = Cl, Y = H, Z = CH3 (5); X = Cl, Y = F (7)) have been synthesized and characterized with elemental analysis and IR. These iron coordinative complexes, activated with methylaluminoxane (MAO), lead to highly active ethylene oligomerization (>107 g/mol Fe h) and the products are mostly linear α-olefins (>90%). The catalytic activities and product properties depend on the substituents on aryl rings and the reaction conditions. As reaction temperature increases, the catalytic activities decrease rapidly and more low-molar-mass products are produced. The product distributions are almost independent of the Al/Fe molar ratio, but the catalytic activities change in different trends when the ortho-substituents on the aryl rings are different. The other three complexes have also been synthesized for comparison to investigate the steric hindrance and electronic effect on the properties of complexes. The complex with adaptable steric hindrance and electronic properties exhibits the highest catalytic activities.  相似文献   

3.
The complex [Rh(CO)2Cl]2 reacts with two molar equivalent of pyridine carboxylic acids ligands Py-2-COOH(a), Py-3-COOH(b) and Py-4-COOH(c) to yield rhodium(I) dicarbonyl chelate complex [Rh(CO)2(L/)](1a) {L/ = η2-(N,O) coordinated Py-2-COO(a/)} and non-chelate complexes [Rh(CO)2ClL//](1b,c) {L// = η1-(N) coordinated Py-3-COOH(b), Py-4-COOH(c)}. The complexes 1 undergo oxidative addition (OA) reactions with different electrophiles such as CH3I, C2H5I, C6H5CH2Cl and I2 to give penta coordinated Rh(III) complexes of the types [Rh(CO)(CORn)XL/], {n = 1,2,3; R1 = CH3(2a); R2 = C2H5(3a); X = I and R3 = CH2C6H5 (4a); X = Cl}, [Rh(CO)I2L/](5a), [Rh(CO)(CORn)ClXL//] {R1 = CH3(6b,c); R2 = C2H5(7b,c); X = I and R3 = CH2C6H5 (8b,c); X = Cl} and [Rh(CO)ClI2L//](9b,c). The complexes have been characterized by elemental analysis, IR and 1H NMR spectroscopy. Kinetic data for the reaction of 1a–b with CH3I indicate a first order reaction. The catalytic activity of 1a–c for the carbonylation of methanol to acetic acid and its ester is evaluated and a higher turn over number (TON = 810–1094) is obtained compared with that of the well-known commercial species [Rh(CO)2I2] (TON = 653) at mild reaction conditions (temperature 130 ± 5 °C, pressure 35 ± 5 bar).  相似文献   

4.
The reaction of organoaluminum compounds containing O,C,O or N,C,N chelating (so called pincer) ligands [2,6-(YCH2)2C6H3]AliBu2 (Y = MeO 1, tBuO 2, Me2N 3) with R3SnOH (R = Ph or Me) gives tetraorganotin complexes [2,6-(YCH2)2C6H3]SnR3 (Y = MeO, R = Ph 4, Y = MeO, R = Me 5; Y = tBuO, R = Ph 6, Y = tBuO, R = Me 7; Y = Me2N, R = Ph 8, Y = Me2N, R = Me 9) as the result of migration of O,C,O or N,C,N pincer ligands from aluminum to tin atom. Reaction of 1 and 2 with (nBu3Sn)2O proceeded in similar fashion resulting in 10 and 11 ([2,6-(YCH2)2C6H3]SnnBu3, Y = MeO 10; Y = tBuO 11) in mixture with nBu3SniBu. The reaction 1 and 3 with 2 equiv. of Ph3SiOH followed another reaction path and ([2,6-(YCH2)2C6H3]Al(OSiPh3)2, Y = MeO 12, Me2N 13) were observed as the products of alkane elimination. The organotin derivatives 411 were characterized by the help of elemental analysis, ESI-MS technique, 1H, 13C, 119Sn NMR spectroscopy and in the case 6 and 8 by single crystal X-ray diffraction (XRD). Compounds 12 and 13 were identified using elemental analysis,1H, 13C, 29Si NMR and IR spectroscopy.  相似文献   

5.
Photoreaction of diaminosubstituted-phosphiteborane, BH3P(NMeCH2)2(OMe) with a methyl molybdenum complex, (η5-C5R5)Mo(CO)3Me (R5 = Me5, Me4H, H5) yielded a phosphiteboryl molybdenum complex, (η5-C5R5)Mo(CO)3BH2{P(NMeCH2)2(OMe)} (R5 = Me5: 2, Me4H: 3, H5: 4). In the reaction of 2 with MeI, the Mo–B bond was activated to give (η5-C5Me5)Mo(CO)3Me, in the reaction with PMe3, the B–P bond was activated to give (η5-C5Me5)Mo(CO)3(BH2PMe3). Complex 2 in solution was gradually converted into (η5-C5Me5)MoH(CO)2{P(NMeCH2)2(OMe)} (8) via the B–H bond activation of 2. Structures of 2, 3, and 8 were determined by single crystal X-ray diffraction studies.  相似文献   

6.
A chiral bidentate phosphoramidite (5a) was synthesized from Shibasaki’s linked-(R)-BINOL and P(NMe2)3 as a new ligand for rhodium(I)-catalyzed asymmetric 1,4-addition of arylboronic acids to α,β-unsaturated carbonyl compounds. The effects of 5a and Feringa’s monodentate phosphoramidite (4, R1, R2 = Et) on the yields and enantioselectivities were fully investigated. The reaction was significantly accelerated in the presence of a base such as KOH and Et3N, allowing the reaction to be completed at the lower temperatures than 50 °C. The addition to cyclic enones such as 2-cyclopentenone, 2-cyclohexenone and 2-cycloheptenone at 50 °C in the presence of an [Rh(coe)2Cl]2-4 (R1, R2 = Et) complex resulted in enantioselectivities up to 98%, though it was less effective for acyclic enones (0–70% ee). On the other hand, a complex between [Rh(nbd)2]BF4 and 5a completed the addition to cyclic enones within 2 h at room temperature in the presence of Et3N with 86–99% yields and 96–99.8% ee. This catalyst was also effective for acyclic enones, resulting in 62–98% yields and 66–94% ee. The 1,4-additions of arylboronic acids to unsaturated lactones and acyclic esters with rhodium(I)-phosphoramidites complexes were also investigated.  相似文献   

7.
The phosphorus ylides Ph3PCHC(O)C6H4R (R = 4-Me 1a, 4-Br 1b) react with PdCl2 in equimolar ratios to give the C,C-orthopalladated [Pd{CHP(C6H4)Ph2CO-C6H4-R)}(μ-Cl)]2 (R = 4-Me 2a, 4-Br 2b) which react with NaClO4/dppe, NaClO4/dppm, py and PPh3 to give the mononuclear derivatives [Pd{CH{P(C6H4)Ph2}COC6H4-R}(dppe-P,P′)[(ClO4) (R = 4-Me 3a, 4-Br 3b), [Pd{CH{P(C6H4)Ph2}COC6H4-R}(dppm-P,P′)[(ClO4 ( (R = 4-Me 4a, 4-Br 4b), [Pd{CH{P(C6H4)Ph2}COC6H4-R}Cl(L)] (L = py, R = 4-Me 5a, 4-Br 5b, L = PPh3, R = 4-Me 6a, 4-Br 6b). The C, C-metalated chelate are demonstrated by an X-ray diffraction study of 3a and 4a. Characterization of the obtained compounds was also performed by elemental analysis, IR, 1H, 31P, and 13C NMR.  相似文献   

8.
Terminal alkynes (HCCR) (R=COOMe, CH2OH) insert into the metal-carbyne bond of the diiron complexes [Fe2{μ-CN(Me)(R)}(μ-CO)(CO)(NCMe)(Cp)2][SO3CF3] (R=Xyl, 1a; CH2Ph, 1b; Me, 1c; Xyl=2,6-Me2C6H3), affording the corresponding μ-vinyliminium complexes [Fe2{μ-σ:η3-C(R)CHCN(Me)(R)}(μ-CO)(CO)(Cp)2][SO3CF3] (R=Xyl, R=COOMe, 2; R=CH2Ph, R=COOMe, 3; R=Me, R=COOMe, 4; R=Xyl, R=CH2OH, 5; R=Me, R=CH2OH, 6). The insertion is regiospecific and C-C bond formation selectively occurs between the carbyne carbon and the CH moiety of the alkyne. Disubstituted alkynes (RCCR) also insert into the metal-carbyne bond leading to the formation of [Fe2{μ-σ:η3-C(R)C(R)CN(Me)(R)}(μ-CO)(CO)(Cp)2][SO3CF3] (R=Me, R=Xyl, 8; R=Et, R=Xyl, 9; R=COOMe, R=Xyl, 10; R=COOMe, R=CH2Ph, 11; R=COOMe, R=Me, 12). Complexes 2, 3, 5, 8, 9 and 11, in which the iminium nitrogen is unsymmetrically substituted, give rise to E and/or Z isomers. When iminium substituents are Me and Xyl, the NMR and structural investigations (X-ray structure analysis of 2 and 8) indicate that complexes obtained from terminal alkynes preferentially adopt the E configuration, whereas those derived from internal alkynes are exclusively Z. In complexes 8 and 9, trans and cis isomers have been observed, by NMR spectroscopy, and the structures of trans-8 and cis-8 have been determined by X-ray diffraction studies. Trans to cis isomerization occurs upon heating in THF at reflux temperature. In contrast to the case of HCCR, the insertion of 2-hexyne is not regiospecific: both [Fe2{μ-σ:η3-C(CH2CH2CH3)C(Me)CN(Me)(R)}(μ-CO)(CO)(Cp)2][SO3CF3] (R=Xyl, 13; R=Me, 15) and [Fe2{μ-σ:η3-C(Me)C(CH2CH2CH3)CN(Me)(R)}(μ-CO)(CO)(Cp)2][SO3CF3] (R=Xyl, 14, R=Me, 16) are obtained and these compounds are present in solution as a mixture of cis and trans isomers, with predominance of the former.  相似文献   

9.
The RuC bond of the bis(iminophosphorano)methandiide-based ruthenium(II) carbene complexes [Ru(η6-p-cymene)(κ2-C,N-C[P{NP(O)(OR)2}Ph2]2)] (R = Et (1), Ph (2)) undergoes a C–C coupling process with isocyanides to afford ketenimine derivatives [Ru(η6-p-cymene)(κ3-C,C,N-C(CNR′)[P{NP(O)(OR)2}Ph2]2)] (R = Et, R′ = Bz (3a), 2,6-C6H3Me2 (3b), Cy (3c); R = Ph, R′ = Bz (4a), 2,6-C6H3Me2 (4b), Cy (4c)). Compounds 34ac represent the first examples of ketenimine–ruthenium complexes reported to date. Protonation of 34a with HBF4 · Et2O takes place selectively at the ketenimine nitrogen atom yielding the cationic derivatives [Ru(η6-p-cymene)(κ3-C,C,N-C(CNHBz)[P{NP(O)(OR)2}Ph2]2)][BF4] (R = Et (5a), Ph (6a)).  相似文献   

10.
Reactions of copper(I) halides with a series of thiosemicarbazones, namely, benzaldehyde thiosemicarbazone (R1R2CN–NH–C(S)–NH2, R1 = Ph, R2 = H; Hbtsc), 2-benzoylpyridine thiosemicarbazone (R1 = Ph, R2 = py; Hbpytsc), and acetone thiosemicarbazone (R1 = R2 = Me; Hactsc), in the presence of PPh3 has formed dimeric complexes, viz. sulfur bridged [Cu2(μ-S-Hbtsc)2Br2(PPh3)2]·2H2O (1), iodo-bridged [Cu2(μ-I)21-S-Hbtsc)2(PPh3)2] (2), and heterobridged [Cu23-S,N3-Hactsc)(η1-Br)(μ-Br)(PPh3)2] (3), as well as mononuclear complexes [CuX(η1-S-Hbpytsc)(PPh3)2]·CH3CN (X = Br, 4; Cl, 5). Complexes 1, 2, 4 and 5 involve thiosemicarbazone ligands in η1-S bonding mode while in compound 3, ligand acts in N3, S-chelation-cum-S-bridging mode (μ3-S,N3 mode). The intermolecular interactions such as, N2H?X, HN1H?X (X = S, Br, Cl), CH?π interactions lead to 2D networks. All the complexes have been characterized with the help of elemental analyses, IR, 1H, and 31P NMR spectroscopy, and single crystal X-ray crystallography. The role of a solvent in alteration of nuclearity and bonding modes of complexes has been highlighted.  相似文献   

11.
The X‐ray structure of the title compound [Pd(Fmes)2(tmeda)] (Fmes=2,4,6‐tris(trifluoromethyl)phenyl; tmeda=N,N,N′,N′‐tetramethylethylenediamine) shows the existence of uncommon C? H???F? C hydrogen‐bond interactions between methyl groups of the TMEDA ligand and ortho‐CF3 groups of the Fmes ligand. The 19F NMR spectra in CD2Cl2 at very low temperature (157 K) detect restricted rotation for the two ortho‐CF3 groups involved in hydrogen bonding, which might suggest that the hydrogen bond is responsible for this hindrance to rotation. However, a theoretical study of the hydrogen‐bond energy shows that it is too weak (about 7 kJ mol?1) to account for the rotational barrier observed (ΔH=26.8 kJ mol?1), and it is the steric hindrance associated with the puckering of the TMEDA ligand that should be held responsible for most of the rotational barrier. At higher temperatures the rotation becomes fast, which requires that the hydrogen bond is continuously being split up and restored and exists only intermittently, following the pulse of the conformational changes of TMEDA.  相似文献   

12.
A triruthenium μ-alkyl complex, (Cp1Ru)3(μ-η2-HCHCH2R)(μ-CO)23- CO) (2a, R = Ph; 2b, R = tBu, Cp1 = η5-C5Me5), which contains a two-electron and three-center interaction among Ru, C, and H atoms, has been synthesized by the reaction of a perpendicularly coordinated 1-alkyne complex, {Cp1Ru(μ-H)}3322(⊥)-RCCH) (1a; R = Ph, 1b; R = tBu), with carbon monoxide. A diffraction study for 2b clearly represented the bridging neohexyl group on one Ru–Ru edge. This μ-alkyl group exhibited dynamic behavior resulting in site-exchange of the α-hydrogen atoms between the terminal and bridging positions, which was synchronized with the migration of the μ-alkyl groups between the two ruthenium atoms. The agostic C–H bond was readily cleaved upon pyrolysis. Whereas the μ-phenethylidene intermediate resulting from the σ-C–H bond cleavage has never been observed, a μ3-phenethylidyne complex, {Cp1Ru(μ-CO)}33-CCH2Ph) (7a), and a μ3-methylidyne complex, {Cp1Ru(μ-CO)}33-CH) (8), were obtained by the successive C–H/C–H and C–H/C–C bond cleavages at the μ-alkyl moiety, respectively.  相似文献   

13.
《Comptes Rendus Chimie》2016,19(3):320-332
1,3-dipolar cycloaddition of diaryldiazomethanes Ar2CN2 across Cl3C–CHN–CO2Et 1 yields Δ3-1,2,4-triazolines 2. Thermolysis of 2 leads, via transient azomethine ylides 3, to diaryldichloroazabutadienes [Ar(Ar')CN–CHCCl2] 4. Treatment of 4a (Ar = Ar' = C6H5) and 4c (Ar = Ar' = p-ClC6H4) with NaSR in DMF yields 2-azabutadienes [Ar2CN–C(H)C(SR)2] 5. In contrast, nucleophilic attack of NaStBu on 4 affords azadienic dithioethers [Ar2CN–C(StBu)C(H)(StBu)] (7a Ar = C6H5; 7b Ar' = p-ClC6H4). The reaction of 4a with NaSEt conducted in neat EtSH produces [Ph2CN–C(H)(SEt)–CCl2H] 8, which after dehydrochloration by NaOMe and subsequent addition of NaSEt is converted to [Ph2CN–C(SEt)C(H)(SEt)] 7c. Upon the reaction of 4c with NaSiPr, the intermediate dithioether [(p-ClC6H4)2CN–CHC(SiPr)2] 5k is converted to tetrakisthioether [(p-iPrSC6H4)2CN–CHC(SiPr)2] 6. Treatment of 4a with the sodium salt of piperidine leads to [Ph2CN–CHC(NC5H10)2] 10. The coordination of 6 on CuBr affords the macrocyclic dinuclear Cu(I) complex 11. The crystal structures of 5i, 7a,b, 10 and 11 have been determined by X-ray diffraction.  相似文献   

14.
The violet ruthenium complex [(η5-C5Me5)Ru(η5-C3B2Me4R1)] (2a, R1 = Me) reacts with terminal alkynes R2CCH to give yellow 4-borataborepine compounds [(η5-C5Me5)Ru{η7-(MeC)3(R1B)2(R2C2H)}] (4c, R1 = Me, R2 = Ph; 4d, R1 = Me, R2 = SiMe3; 4e, R1 = Me, R2 = H). The insertion of alkynes into the folded C3B2 heterocycle of 2a causes some steric hindrance, which yields with elimination of the distant boranediyl group the corresponding boratabenzene complexes 5 as byproducts. The analogous reactions with internal alkynes R2CCR2 proceed slowly and afford predominantly the boratabenzene complexes [(η5-C5Me5)Ru{η6-(MeC)3(MeB)(R2C)2}] (5f, R2 = Et, 5g, R2 = p-tolyl), respectively. In the latter case, three byproducts are formed: methylboronic acid and 1,2,3,4-tetra-p-tolyl-1,3-butadiene (9) due to hydrolysis of the postulated 2,3,4,5-tetra-p-tolyl-1-methylborole (10) and unexpectedly, the cationic triple-decker complex [{(η5-C5Me5)Ru}2{μ,η7-(MeC)3(MeB)2(CH)2}]Cl (11) having two separated CH groups. The new compounds were characterized by NMR, MS, and single-crystal X-ray studies of 4c, 5f, 9 and 11.  相似文献   

15.
Reactivity of a hydrido(hydrosilylene)tungsten complex, Cp1(CO)2(H)WSi(H)[C(SiMe3)3] (1), toward oxiranes was investigated. Treatment of 1 with racemic mono-substituted oxiranes with a substituent R (R = Ph, vinyl, tBu, or nBu) at room temperature produced dihydrido(vinyloxysilyl)tungsten complexes, (E)- and/or (Z)-Cp1(CO)2(H)2W{Si(H)(OCHCHR)[C(SiMe3)3]} [(E/Z)-2: R = Ph, (E)-3: R = vinyl, (E)-4: R = tBu, (E/Z)-5: R = nBu] in high yields via regioselective ring-opening of oxiranes. When the substituent R on oxirane was relatively large, (E)-isomers (2, 3, and 4) were obtained predominantly (87–97%), while the substituent was a relatively small nBu group, an approximately 1:1 mixture of (E)- and (Z)-isomers [(E/Z)-5] was obtained. Reaction of 1 with 2,2-dimethyloxirane afforded the corresponding complex, Cp1(CO)2(H)2W{Si(H)(OCHCMe2)[C(SiMe3)3]} (6), quantitatively. A reaction mechanism is also discussed.  相似文献   

16.
《Comptes Rendus Chimie》2003,6(2):189-191
Chiral ammonium cations of the type [R1R2R3R4N]+, new templates for a new family of bimetallic oxalate bridged networks. This work deals with the obtention of a new family of bimetallic [R1R2R3R4N][MnCr(C2O4)3] oxalate bridged 2D networks templated by chiral ammonium cations of the type [R1R2R3R4N]+, in which R stands for n-alkyl groups from methyl to dodecyl. To cite this article: M. Gruselle et al., C. R. Chimie 6 (2003).  相似文献   

17.
The synthesis of a new bidentate anilide ligand and four uranium amide complexes utilizing the ligand are reported. The secondary aniline HN[R]ArMeL (R = C(CD3)2CH3, ArMeL = 2-NMe2-5-MeC6H3) is prepared by condensation of H2NArMeL and acetone-d6 followed by alkylation of the resulting imine with MeLi. The ligand precursors (Et2O)Li(N[R]ArMeL) and K(N[R]ArMeL) are prepared through deprotonation of HN[R]ArMeL with n-BuLi and KH, respectively. Treatment of UI3(THF)4 with (Et2O)Li(N[R]ArMeL) (2 equiv) provides the uranium(III) -ate complex Li[I2U(N[R]ArMeL)2] (Li[1]), while treatment of UI3 with three equiv. of K(N[R]ArMeL) provides the neutral uranium(III) complex U(N[R]ArMeL)3 (2). Both uranium(III) complexes are susceptible to 1e oxidation, as is demonstrated by the syntheses of the uranium(IV) derivatives I2U(N[R]ArMeL)2 (1) and [U(N[R]ArMeL)3][OTf] ([2][OTf]; OTf = CF3SO3). The spectroscopic and X-ray structural characterization of all four uranium complexes is described. The structures of 2 and [2][OTf] exhibit a large degree of steric pressure about the uranium center, effectively preventing the [2]+ ion from achieving a seven-coordinate structure.  相似文献   

18.
The compounds [o-C6-H4CH2E]2Sn-W(CO)5, (E = NMe2 (1) or PPh2 (2)) have been prepared by reaction of o-LiC6H4CH2E with pentacarbonyltungsten tin(II) chloride (CO)5WSnCl2. The complexes were characterized by 13C, 31P, and 119Sn NMR spectroscopy and by X-ray diffraction analyses. 1 crystallizes monoclinically in the space group C/c (no. 15) with a 1310.2(4), b 1552.1(4), c 1202.9(4) pm, β 90.11(4)°, and Z = 4. 2 crystallizes monoclinically in the space goup P21/n (no. 14) with a 2108.1(4), b 1707.7(4), c 1283.7(3) pm, β 97.47(2)° and Z = 4. The structures were refined to final R values of 0.029 and 0.039, respectively.The SnW bond distances of 274.9 and 276.2 pm are very similar in both complexes. The Sn atoms are penta-coordinated by 2C, 2N and W in 1 and by 2C, 2P and W in 2. The penta-coordination comprises one SnW and two SnC single bonds, and either a SnN (in 1) or a SnP bond (in 2) of bond order 0.45. In the stannyl group of 1 the SnN bond distances both are identical by symmetry (256.4 pm), whereas the two SnP bond lengths of 2 differ to some extent (283.1 and 301.2 pm).  相似文献   

19.
A second-kind asymmetric transformation involving N-atom inversion has been observed at 20° for 1(S) - α - carboxyethyl - 3,3 - bis(trifluoromethyl)diaziridine 1 and its methyl ester 2. X-Ray data for the diastereomer (1S, 2S, α-S), 1A (which is thermodynamically preferred in the crystalline phase), 1H NMR spectra of ethyl ester 3-15N(1) and 3-15N(2), CD spectra of 1A,B, 2A,B, potassium salt 4A,B and semiempirical calculations (MINDO/3 and INDO) for 1A, show that the stereospecificity of crystallization of the diastereomer A is due to the higher energy of the crystal lattice of the diastereomer (1R, 2R, α-S), 1B because of hindered charge compensation as well as to the hindrance by the CF3, groups to intermolecular H-bonds. According to semiempirical calculations, the stability of 3,3 - bis(trifluoromethyl) - diaziridines (TFD) to the action ofel w-orbital energies and depolarization of the C-N bonds due to hyperconjugation and the inductive effect of the CF3,-groups. The steric effect of these groups is the reason of the low configurational stability of TFD compared with the 3,3-dimethyl analogues.  相似文献   

20.
Charge density studies of chemical bonds for two iron complexes, [(NO)Fe(S,S-C6H4)2] [PPN] (1), where PPN = N(Pph3)2 and Fe3(NO)3(S,S-C6H4)3 (2) are investigated in terms of the topological properties at bond critical points based on the ‘atoms in molecule’ theory. The one electron reduction form (1R) of complex 1 and the one electron oxidation form (2O) of complex 2 are also included for comparison. The X-ray absorption spectroscopy of Fe K- and LIII,II-edges, as well as the N/S K-edge are applied to verify the illustration in the variation of the electronic structures. Based on the ρc, ?2ρc, and Hb values among the compound studied, Fe-S/N can be regarded as polarized covalent bond, and Fe-N bonds show stronger covalent character than that of the Fe–S bond, which is believed to be a highly polarized covalent bond.  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