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 93 毫秒
1.
Synthesis and Spectroscopic Characterisation of some Pentacarbonyltungsten(0) Complexes with Mono‐ and Bicyclic Phosphirane Ligands: Crystal Structure of [{(Me3Si)2HCPC(H)H–C(H)Ph}W(CO)5] The tungsten(0) complex [{(Me3Si)2HCPC(Ph)=N}W(CO)5] ( 1 ) reacts upon heating with alkene derivatives 2 , 6 , 8 , and 10 in toluene to form benzonitrile and the complexes [{(Me3Si)2HCPC(R1,R2)–C(R3,R4}W(CO)5] ( 4 , 7 a , b , 9 a , b , 11 a , b ) ( 4 (trans): R1,R3 = Ph, R2,R4 = H, 7 a , b (cis, meso and rac): R1,R3 = Ph, R2,R4 = H, 9 a , b (RR und SS): R1 = Ph, R2,R3,R4 = H, 11 a , b : R1=R3 = (CH2)4, R2,R4 = H). Spectroscopic and mass spectrometric data are discussed. The structure of the complex 9 a was determined by X‐ray single crystal structure analysis showing characteristic data for the phosphirane ring such as a narrow angle at phosphorus (49,2(2)°), different P–C distances (P–C(6) 182,1(5) and P–C(7) 185,2(4) pm) and 152,9(6) pm for the basal C–C bond.  相似文献   

2.
The reaction of [(ArN)2MoCl2] · DME (Ar = 2,6‐i‐Pr2C6H3) ( 1 ) with lithium amidinates or guanidinates resulted in molybdenum(VI) complexes [(ArN)2MoCl{N(R1)C(R2)N(R1)}] (R1 = Cy (cyclohexyl), R2 = Me ( 2 ); R1 = Cy, R2 = N(i‐Pr)2 ( 3 ); R1 = Cy, R2 = N(SiMe3)2 ( 4 ); R1 = SiMe3, R2 = C6H5 ( 5 )) with five coordinated molybdenum atoms. Methylation of these compounds was exemplified by the reactions of 2 and 3 with MeLi affording the corresponding methylates [(ArN)2MoMe{N(R1)C(R2)N(R1)}] (R1 = Cy, R2 = Me ( 6 ); R1 = Cy, R2 = N(i‐Pr)2 ( 7 )). The analogous reaction of 1 with bulky [N(SiMe3)C(C6H5)C(SiMe3)2]Li · THF did not give the corresponding metathesis product, but a Schiff base adduct [(ArN)2MoCl2] · [NH=C(C6H5)CH(SiMe3)2] ( 8 ) in low yield. The molecular structures of 7 and 8 are established by the X‐ray single crystal structural analysis.  相似文献   

3.
Several compounds of the general formula R1R2C(OH)CH2OR containing different geminal group systems have been investigated in proton and carbon magnetic resonances at variable temperatures. The role of steric factors, viz. of the size of substituents and the distances between different parts of the molecules, have been discussed.  相似文献   

4.
The diiron ynamine complex [Fe2(CO)7{μ-CR)C(NEt2)}] (1:R=Me,2:R = C3H5.3:R=SiMe3.4:R = Ph) reacts at room temperature with diphenyldiazomethane Ph2CN2, in hexane to yield complexes [Fe2(CO)6{C(R)C(NEt2)N (NCPh2)] (5a:R=Me,6a:R=C3H5.7a R=SiMe3.8a:R=Ph) resulting from the insertion of the terminal nitrogen atom into the Fe=C carbene bond. Insertion the second nitrogen atom and formation of compounds [Fe2(CO)6zμ-C(R)C(NEt2)NN(CPh2)}] (5b:R=Me,6b:R=C3H5,7b:R=SiMe3,8b:R=Ph) is observed when compounds5a-5a are treated in refluxing hexane. Transformation of compoundsa tob is also obtained at room temperature within a few days. All compounds were identified by their1H NMR spectra. Compounds6a, 7a, 8a, and8b were characterized by single crystal X-ray diffraction analyses. Crystal data: for6a: space group = P21/n,a=12.853(1) A,b=24.800(7) A,c=8.947(6) A,β=99.29(3)°,Z=4, 2227 rellectionsR=0,038; for7a: space group=Pl,a=ll.483(4) A,b=14.975(4) A,c = 17.890(8) A,α = 82.80(3)°,β=94.29(7)°,γ=85.42(2),Z = 4, 5888 reflectionR = 0.035: for8a: space group = Pcab.a = 31.023(8) A.b=20.137(1) A.c=9.686(2) A.Z=8. 1651 reflections,R=0.071; for8b: space group=P21/n,a=21.459(4),b=10,100(3) A,c=28,439(8) A,ß=103.86(4)°,Z=8. 2431 reflections.R=0.057.  相似文献   

5.
Deprotonation of the aminophosphanes Ph2PN(H)R 1a – 1h [R = tBu ( 1a ), 1‐adamantyl ( 1b ), iPr ( 1c ), CPh3 ( 1d ), Ph ( 1e ), 2,4,6‐Me3C6H2 (Mes) ( 1f ), 2,4,6‐tBu3C6H2 (Mes*) ( 1g ), 2,6‐iPr2C6H3 (DIPP) ( 1h )], followed by reactions of the phosphanylamide salts Li[Ph2PNR] 2a , 2b , 2g , and 2h with the P‐chlorophosphaalkene (Me3Si)2C=PCl, and of 2a – 2g with (iPrMe2Si)2C=PCl, gave the isolable P‐phosphanylamino phosphaalkenes (Me3Si)2C=PN(R)PPh2 3a , 3b , 3g , and (iPrMe2Si)2C=PN(R)PPh2 4a – 4g . 31P NMR spectra, supported by X‐ray structure determinations, reveal that in compounds 2a , 2b , 3a , and 3b , with bulky N‐alkyl groups the Si2C=P–N–P skeleton is non‐planar (orthogonal conformation), whereas 3g , 3h , and 4g with bulky N‐aryl groups exhibit planar conformations of the Si2C=P–N–P skeleton. Solid 3g and 4g exhibit cisoid orientation of the planar C=P–N–C units (planar I) but in solid 3h the transoid rotamer is present (planar II). From 3g , 4d , and 4g mixtures of rotamers were detected in solution by pairs of 31P NMR patterns ( 3h : line broadening).  相似文献   

6.
The 13C chemical shifts of the unsaturated carbons were measured in 31 cis and trans pairs of β-substituted enones R1? C(1)O? C(2)H?C(3)H? R2. In these polarized ethylenes the chemical shifts of the olefinic carbons are simply related by the equation δct+A. The steric and electronic effects introduced by the R1 and R2 substituents influence the chemical shifts of C-2 and C-3 in both isomers. It is shown that the sign and magnitude of the intercept A mainly reflect the π-charge electronic density changes which arise in the cis isomer and are transmitted via the π-framework. The effect of the steric interaction on the chemical shift of C-3 in the cis isomers is postulated to be related to the symmetry of the substituents. Therefore, the differential shielding of C-3 is indicative of the conformational structure of the cis molecule.  相似文献   

7.
The reactions of ten metastable immonium ions of general structure R1R2C?NH+C4H9 (R1 = H, R2 = CH3, C2H5; R1 = R2 = CH3) are reported and discussed. Elimination of C4H8 is usually the dominant fragmentation pathway. This process gives rise to a Gaussian metastable peak; it is interpreted in terms of a mechanism involving ion-neutral complexes containing incipient butyl) cations. Metastable immonium ions ontaining an isobutyl group are unique in undergoing a minor amount of imine (R1R2C?NH) loss. This decomposition route, which also produces a Gaussian metastable peak, decreases in importance as the basicity of the imine increases. The correlation between imine loss and the presence of an isobutyl group is rationalized by the rearrangement of the appropriate ion-neutral complexes in which there are isobutyl cations to the isomeric complexes containing the thermodynamically more stable tert-butyl cations. A sizeable amount of a third reaction, expulsion of C3H6, is observed for metastable n-C4H9 +NH?CR1R2 ions; in contrast to C4H8 and R1R2C?NH loss, C3H6 elimination occurs with a large kinetic energy release (40–48 kJ mol?1) and is evidenced by a dish-topped metastable peak. This process is explained using a two-step mechanism involving a 1,5-hydride shift, followed by cleavage of the resultant secondary open-chain cations, CH3CH+ CH2CH2NHCHR1R2.  相似文献   

8.
In this work, the design, synthesis, and single-molecule conductance of ethynyl- and butadiynyl-ruthenium molecular wires with thioether anchor groups [RS=n-C6H13S, p-tert-Bu−C6H4S), trans-{RS−(C≡C)n}2Ru(dppe)2 (n=1 ( 1R ), 2 ( 2R ); dppe: 1,2-bis(diphenylphosphino)ethane) and trans-(n-C6H13S−C≡C)2Ru{P(OMe)3}4 3hex ] are reported. Scanning tunneling microscope break-junction study has revealed conductance of the organometallic molecular wires with the thioacetylene backbones higher than that of the related organometallic wires having arylethynylruthenium linkages with the sulfur anchor groups, trans-{p-MeS−C6H4-(C≡C)n}2Ru(phosphine)4 4 n (n=1, 2) and trans-(Th−C≡C)2Ru(phosphine)4 5 (Th=3-thienyl). It should be noted that the molecular junctions constructed from the butadiynyl wire 2R , trans-{ Au −RS−(C≡C)2}2Ru(dppe)2 ( Au : gold metal electrode), show conductance comparable to that of the covalently linked polyynyl wire with the similar molecular length, trans-{ Au −(C≡C)3}2Ru(dppe)2 63 . The DFT non-equilibrium Green's function (NEGF) study supports the highly conducting nature of the thioacetylene molecular wires through HOMO orbitals.  相似文献   

9.
Functionally conjugated enynes, H2CC(R1)CCCR2R3OS(O)Me, undergo 1,5-substitution with alkylsilver(I) reagents, RAGG ☆ 3 LiBr. The purity of the produced alkylated butatrienes, RCH2C(R1)CCCR2R3 depends on the nature of R in RAg ☆ 3 LiBr and on the substituents R1, R2 and R3 in the substrate.  相似文献   

10.
The reaction of Ru3(CO)12 with MeO2C(H)C=C=C(H)CO2 Me has yielded two isomeric productsanti-Ru2(CO)6[μ-η 3-η 1-MeO2C(H)CCC(H)CO2Me],1 in 70% yield andsyn-Ru2(CO)6[μ-η 3-η 1-MeO2C(H)CCC(H)CO2Me],2 in 5% yield. Both compounds were characterized by single crystal X-ray diffraction analysis. Both products are diruthenium complexes with bridging di(carboxylate)allene ligands in which the oxygen atom of the carbonyl group of one of the carboxylate groupings is coordinated to one of the metal atoms. Compound1 isomerizes partially to2 at 68°C. Crystal Data for1: space group=P21/n,a=11.131(1) Å,b=10.228(2) Å,c=15.978(2) Å,β=102.01(1)°,Z=4, 1653 reflections,R=0.025; for2: space group=P $\bar 1$ ,a=9.340(1) Å,b=14.925(4) Å,c=6.778(2) Å,α=99-02(2)°,β=104 62(2)°,γ=94.58(2)°,Z=2, 1857 reflections,R=0.027.  相似文献   

11.
In the crystal networks of N,N′‐bis(2‐chlorobenzyl)‐N′′‐(2,6‐difluorobenzoyl)phosphoric triamide, C21H18Cl2F2N3O2P, (I), N‐(2,6‐difluorobenzoyl)‐N′,N′′‐bis(4‐methoxybenzyl)phosphoric triamide, C23H24F2N3O4P, (II), and N‐(2‐chloro‐2,2‐difluoroacetyl)‐N′,N′′‐bis(4‐methylphenyl)phosphoric triamide, C16H17ClF2N3O2P, (III), C=O...H—NC(O)NHP(O) and P=O...H—Namide hydrogen bonds are responsible for the aggregation of the molecules. This is the opposite result from that commonly observed for carbacylamidophosphates, which show a tendency for the phosphoryl group, rather than the carbonyl counterpart, to form hydrogen bonds with the NH group of the C(O)NHP(O) skeleton. This hydrogen‐bond pattern leads to cyclic R22(10) motifs in (I)–(III), different from those found for all previously reported compounds of the general formula RC(O)NHP(O)[NR1R2]2 with the syn orientation of P=O versus NH [R22(8)], and also from those commonly observed for RC(O)NHP(O)[NHR1]2 [a sequence of alternate R22(8) and R22(12) motifs]. In these cases, the R22(8) and R22(12) graph sets are formed through similar kinds of hydrogen bond, i.e. a pair of P=O...H—NC(O)NHP(O) hydrogen bonds for the former and two C=O...H—Namide hydrogen bonds for the latter. This article also reviews 102 similar structures deposited in the Cambridge Structural Database and with the International Union of Crystallography, with the aim of comparing hydrogen‐bond strengths in the above‐mentioned cyclic motifs. This analysis shows that the strongest N—H...O hydrogen bonds exist in the R22(8) rings of some molecules. The phosphoryl and carbonyl groups in each of compounds (I)–(III) are anti with respect to each other and the P atoms are in a tetrahedral coordination environment. In the crystal structures, adjacent molecules are linked via the above‐mentioned hydrogen bonds in a linear arrangement, parallel to [010] for (I) and (III) and parallel to [100] for (II). Formation of the NC(O)NHP(O)—H...O=C instead of the NC(O)NHP(O)—H...O=P hydrogen bond is reflected in the higher NC(O)NHP(O)—H vibrational frequencies for these molecules compared with previously reported analogous compounds.  相似文献   

12.
Quantum chemical insights into normal Pd‐C2(NHCR) and abnormal Pd‐C5(aNHCR) bonding, dominated by dispersion interactions in N‐hetereocyclic carbene complexes [PdCl2(NHCR)2] ( I , R = H; II , R = Ph; III , R = Mes (2,4,6‐trimethyl)phenyl)) and [PdCl2(NHCR)(aNHCR] ( IV , R = H; V , R = Ph; VI , R = Mes) have been investigated at DFT and DFT‐D3(BJ) level of theory with particular emphasis on the effects of the noncovalent interactions on the structures and the nature of Pd‐C2(NHCR) and Pd‐C5(aNHCR) bonds. The optimized geometries are good agreement with the experimental values. The Pd‐C bonds are essentially single bond. Hirshfeld charge distributions indicate that the abnormal aNHCR carbene ligand is relatively better electron donor than the normal NHCR carbene ligand. The C2 atom has larger %s contribution along Pd‐C2 bond than the C5 atom along Pd‐C5 bond. As a consequence the Pd‐C2(NHCR) bonds are relative stronger than the Pd‐C5(aNHCR) bonds. Thus, the results of natural hybrid orbital analysis support the key point of the present study. Calculations predict that for bulky substituent (R = Ph, Mes) at carbene, the Pd‐C2(NHCR) bond is stronger than Pd‐C5(aNHCR) bond due to large dispersion energy in [PdCl2(NHCR)2] than in [PdCl2(NHCR)(aNHCR)]. However, in case of non‐bulky substituent with small and almost equal contribution of dispersion energy, the Pd‐C2(NHCR) bond is relative weaker than Pd‐C5(aNHCR) bond. The bond dissociation energies are dependent on the R substituent, the DFT functional and the inclusion of dispersion interactions. Major point of this study is that the abnormal aNHCs are not always strongly bonded with metal center than the normal NHCs. Effects of dispersion interaction of substituent at nitrogen atoms of carbene ligand are found to play a crucial role on estimation of relative bonding strengths of the normal and abnormal aNHCs with metal center. © 2016 Wiley Periodicals, Inc.  相似文献   

13.
《Tetrahedron》2019,75(24):3239-3247
An enantioselective palladium-catalyzed C(sp2)-H carbamoylation for the preparation of chiral isoindolines was described for the first time. With chiral monophosphorus ligand (R)-AntPhos as the ligand, a series of chiral isoindolines were prepared from diarylmethyl carbamoyl chlorides in excellent yields and enantioselectivities with the palladium loading as low as 1 mol%. Initial mechanistic studies indicated the asymmetric cyclization catalyzed a palladium species with a single chiral monophosphorus ligand.  相似文献   

14.
Tris(trimethylsilyl)silyllithium ( 3 ) reacted with aldehydes and ketones (molar ratio 2 : 1) according to a modified Peterson mechanism under formation of transient silenes, which were immediately trapped by excess 3 to give the organolithium derivatives (Me3Si)3SiSi(SiMe3)2C(Li)R1R2 ( 7 ). Hydrolysis of 7 afforded the alkylpolysilanes (Me3Si)3SiSi(SiMe3)2CHR1R2 ( 8 ). Depending on the substituents R1 and R2, 7 proved to be rather unstable in THF solution and underwent a rapid rearrangement, involving a 1,3‐Si,C‐trimethylsilyl migration, resulting in the formation of the lithium silanides (Me3Si)2Si(Li)Si(SiMe3)2C(SiMe3)R1R2 ( 9 ), which were hydrolized during the aqueous workup to give the H‐silanes (Me3Si)2Si(H)Si(SiMe3)2C(SiMe3)R1R2 ( 10 ). Reaction of 9 with chlorotrimethylsilane produced the 1‐trimethylsilylalkylpolysilanes (Me3Si)3SiSi(SiMe3)2C(SiMe3)R1R2 ( 11 ). The structures of the products described were elucidated by comprehensive spectral analyses. The results of X‐ray crystal structure analyses, performed for 8 l (R1 = H, R2 = 2,4,6‐(MeO)3C6H2), 10 d (R1 = H, R2 = Mes) and 11 d (R1 = H, R2 = Mes) are discussed and confirm the expected extreme sterical congestion of the molecules.  相似文献   

15.
Copolymerization of fullerene (C60) with methyl methacrylate (MMA) was carried out using triphenylbismuthonium ylide (abbreviated as Ylide) as a novel initiator in dioxan at 60°C for 4 h in a dilatometer under a nitrogen atmosphere. The reaction follows ideal kinetics: Rp∝ [Ylide]0.5[C60]?1.0[MMA]1.0. The rate of polymerization increases with an increase in concentration of initiator and MMA. However, it decreases with increasing concentration of fullerene due to the radical scavenging effect of fullerene. The overall activation energy of copolymerization was estimated to be 57 KJ mol?1. The fullerene-MMA copolymers (C60-MMA) were characterized by FTIR, UV–Vis, NMR and GPC analyses.  相似文献   

16.
Reactions of acyl iodides R1COI (R1=Me, Ph) with trialkyl(alkynyl)silanes,-germanes, and stannanes (R2C≡CMR 3 3 ; M=Si, Ge, Sn) were studied. Acyl iodides reacted with the germanium and tin derivatives with cleavage of the M-Csp bond and formation of the corresponding trialkyl(iodo)germanes and-stannanes R 3 3 MI (M=Ge, Sn) and alkynyl ketones R1C(O)C≡CR2 and R1C(O)C≡CC(O)R1. By contrast, the reaction of acetyl iodide with ethynyl(trimethyl)silane gave only a small amount of 1,2-diiodovinyl(trimethyl) silance as a result of iodine addition at the triple bond. Bis(trimethylsilyl)ethyne failed to react with acetyl iodide.  相似文献   

17.
Binuclear Nickel(0) Alkyne Coordination Compounds – Correlation between Ligand Periphery and Supramolecular Structure Reaction of Ni(cdt: 1,5,9-cyclododecatriene) with functionalized alkynes and subsequent reaction with ethylenediamines gives binuclear compounds of the type (diamine)Ni(μ-alkyne)Ni(alkyne). Compounds with alkyne-diols (N?N)Ni2(HOR1R2C? C?C? CR1R2OH)2 show supramolecular structures in which two identical intramolecular and one intermolecular hydrogen bonds are realized. 1 and 2 (chelate ligand in each case N,N,N′,N′-tetramethylethylenediamine, TMEDA, in 1 R1 = R2 = Me, in 2 R1 = R2 = Et) polymer-like chains are built up by connecting the binuclear units. Via two intermolecular hydrogen bonds per organometallic unit in 1 and via one intermoleculare hydrogen bond in 2 the chains are connected to give double chains. By substitution of one methyl group of TMEDA by hydrogen ( 3 : R1 = R2 = Me) a polymerlike network is produced by connecting the polymer-like chains. In compound 4 in which one of the methyl groups of TMEDA is substituted by CH2CH2NMe2 the polymer-like chains remain unconnected. In 5 (diamine = TMEDA, alkyne = (CH3)3C? C?C? CMe2OH) one intermolecular hydrogen bond per organometallic unit is observed forming again polymer-like chains that are independent of each other.  相似文献   

18.
Phosphanediyl Transfer from Inversely Polarized Phosphaalkenes R1P=C(NMe2)2 (R1 = tBu, Cy, Ph, H) onto Phosphenium Complexes [(η5‐C5H5)(CO)2M=P(R2)R3] (R2 = R3 = Ph; R2 = tBu, R3 = H; R2 = Ph, R3 = N(SiMe3)2) Reaction of the freshly prepared phosphenium tungsten complex [(η5‐C5H5)(CO)2W=PPh2] ( 3 ) with the inversely polarized phosphaalkenes RP=C(NMe2)2 ( 1 ) ( a : R = tBu; b : Cy; c : Ph) led to the η2‐diphosphanyl complexes ( 9a‐c ) which were isolated by column chromatography as yellow crystals in 24‐30 % yield. Similarly, phosphenium complexes [(η5‐C5H5)(CO)2M=P(H)tBu] (M = W ( 6 ); Mo ( 8 )) were converted into (M = W ( 11 ); Mo ( 12 )) by the formal abstraction of the phosphanediyl [PtBu] from 1a . Treatment of [(η5‐C5H5)(CO)2W=P(Ph)N(SiMe3)2] ( 4 ) with HP=C(NMe2)2 ( 1d ) gave rise to the formation of yellow crystalline ( 10 ). The products were characterized by elemental analyses and spectra (IR, 1H, 13C‐, 31P‐NMR, MS). The molecular structure of compound 10 was elucidated by an X‐ray diffraction analysis.  相似文献   

19.
β-Alkoxyalkylmercury(II) acetates have been symmetrised in situ with alkaline sodium stannite to afford good yields of bis(β-alkoxyalkyl)mercurials, [R1R2C(OR)CHR3]2 Hg.  相似文献   

20.
Separation and Absolute Configuration of the C(8)-Epimeric (app-E)-Neochromes (Trollichromes) and -Dinochromes The C(8′)-epimers of (all-E)-neochrome were separated by HPLC and carefully characterized. The faster eluted isomer, m.p. 197.8–198.3°, is shown to have structure 3 ((3S,5R,6R,3′S,5′R,8′R)-5′,8′-epoxy-6,7-dodehydro-5,6,5′,8′-tetrahydro-β,β-carotene-3,5,3′-triol). To the other isomer, m.p. 195-195.5°, we assign structure 6 , ((3S,5R,6R,3′S,5′R,8′R)-5′,8′-epoxy-6,7-didehydro-5,6,5′,8′-tetrahydro-β,β-carotene-3,5,3′-triol). The already known epimeric dinochromes (= 3-O-acetylneochromes) can now be formulated as 4 and 5 , (‘epimer 1’ and its trimethylsilyl ether) and 7 and 8 , (‘epimer 2’ and its trimethylsilyl ether), respectively.  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