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 31 毫秒
1.
1,2-Digermacyclobut-3-enes were prepared by the treatment of Z-α,β-bis(chlorodialkylgermyl)ethenes with Na metal in boiling toluene and their structures were fully established by spectroscopic methods coupled with X-ray crystallography. In the presence of an appropriate catalyst, 1,2-digermacyclobut-3-enes reacted smoothly with alkynes to give the corresponding insertion products, 1,4-digermacyclohexa-2,5-dienes in moderate to good yields. Conventional complexes, such as [Pd(PPh3)4] and [Pt(PPh3)4], serve as efficient catalysts. The mechanism of the insertion reaction of alkynes into the germanium-germanium bond of 1,2-digermacyclobut-3-enes is discussed in terms of a key intermediate, a 1,4-digerma-2-buten-1,4-diylpalladium.  相似文献   

2.
Thienylmercury(II)chloride reacts with [Pd(PPh3)2Cl2], [Pd(PPh3)4] and [Pt(PPh3)4] to afford new compounds containing a metal-2-thienyl linkage. The compound [Pd(PPh3)2(2-C4H3S)Cl] probably has trans stereochemistry.2-Bromothiophen undergoes oxidative addition with [Pd(PPh3)4] and [Pt(PPh3)4], probably via a radical mechanism. With [Pd(CO)(PPh3)3], a carbonyl inserted product is obtained. The bromo-metal(II) complexes have trans stereochemistry. The course of the reaction between 3-methyl-2-bromothiophen and Pd(PPh3)4 is more complex. Thus, there is evidence of some cis bromopalladium(II) compounds amongst the products, also there is good evidence to support the view that some isomerisation of 3-methyl-2-thienyl to 4-methyl-2-thienyl occurs during the reaction, thus giving greater molar quantities of [Pd(PPh3)2(4-CH3-2-C4H2S)Br] than can be accounted for from any initial 4-methyl-2-bromothiophen impurity.The metallation of the thiophen ring, probably in the 4-position, with palladium(II) is described for 3-theylidene-4-methylaniline.  相似文献   

3.
Bis(triphenylphosphine)Palladium Complexes with Sulfur Oxide Ligands New examples in the series of sulfur oxide complexes of the type (PPh3)2Pd(SnOm) (n = 1,2; m 1–4) were found by the synthesis of (PPh3)2Pd(SO) and (PPh3)2Pd(S2O3. The SO complex is obtained by the reaction of Pd(PPh3)4 or (PPh3)2Pd(RCCR) (R=COOMe) and thiirane-S-oxide. The thiosulfato complex (PPh3)2Pd(S2O3) is formed from (PPh3)2Pd(SO) and SO2 or, alternatively, from (PPh3)3Pd(SO2) and C2H4SO. Both SO und SO2 complexes can be oxidized to the corresponding sulfato compound (PPh3)2)Pd(SO4). The SO complex is used as a SO-source for the formation of 3,4-dimethyldihydrothiophene-S-oxide from 2,3-dimethyl-1,3-butadiene.  相似文献   

4.
By unambiguous methods, (Z)- and (E)-2, 3-dimethyl(1, 1, 1, 4, 4, 4-2H6)but-2-enes ( 3 ) were synthesized and transformed to the epoxides 4 with 3-chloroperbenzoic acids. Both the isotopomeric olefins and the epoxides are detected separately by 1H-NMR at 400 MHz. Epoxidation of (Z)- 3 with [RhICl(PPh3)3]/cumene hydroperoxide resulted in a 1: 1 mixture of (Z)- and (E)- 4 , while reaction of (Z)- 3 with [FeIII(tpp)]Cl/PhIO gave only (Z)- 4 (tpp = tetraphenylporphyrin).  相似文献   

5.
Reactions of the trans-PdCl2(PPh3)2 precursor with furan-2-carbaldehyde thiosemicarbazone (Hftsc) and thiophene-2-carbaldehyde thiosemicarbazone (Httsc), in 1:1 molar ratios in the presence of Et3N base, removed one Cl and one PPh3 group from the PdII center, and yielded the complexes [Pd(η2-N3,S-ftsc)(PPh3)Cl] (1) and [Pd(η2-N3,S-ttsc)(PPh3)Cl] (2), respectively. However, when a 1:2 molar ratio (M:L) was used, both Cl and PPh3 ligands were removed, yielding the complexes trans-[Pd(η2-N3,S-ftsc)2] (3) and trans-[Pd(η2-N3,S-ttsc)2] (4). Complexes 14 have been characterized with the help of analytical data, spectroscopic techniques (IR, 1H and 31P NMR) and single crystal X-ray crystallography. The thiosemicarbazone ligands behave as uninegative N3,S-chelating ligands in complexes 14. In contrast, pyrrole-2-carbaldehyde thiosemicarbazone (H2ptsc) and salicylaldehyde thiosemicarbazone (H2stsc) invariably formed the complexes [Pd(η3-N4,N3,S-ptsc)(PPh3)] (5) and [Pd(η3–O, N3,S-stsc)(PPh3)] (6), respectively, and the ligands acted as binegative tridentate donors (N4, N3, S, 5; O, N3, S, 6).  相似文献   

6.
Reaction between Os(SiCl3)Cl(CO)(PPh3)2 and five equivalents of MeLi produces a colourless intermediate, tentatively formulated as the lithium salt of the six-coordinate, dimethyl, trimethylsilyl-containing complex anion, Li[Os(SiMe3)(Me)2(CO)(PPh3)2]. Reaction of this material with ethanol releases methane and gives the red, coordinatively unsaturated methyl, trimethylsilyl-containing complex, Os(SiMe3)(Me)(CO)(PPh3)2 (1). An alternative synthesis of 1 is to add one equivalent of MeLi to Os(SiMe3)Cl(CO)(PPh3)2, which in turn is obtained by adding three equivalents of MeLi to Os(SiCl3)Cl(CO)(PPh3)2. Treatment of 1 with p-tolyl lithium, again gives a colourless intermediate which may be Li[Os(SiMe3)(Me)(p-tolyl)(CO)(PPh3)2], and reaction with ethanol gives the red complex, Os(SiMe3)(p-tolyl)(CO)(PPh3)2 (3). Complexes 1 and 3 are readily carbonylated to Os(SiMe3)(Me)(CO)2(PPh3)2 (2) and Os(SiMe3)(p-tolyl)(CO)2(PPh3)2 (4), respectively. Heating Os(SiMe3)Cl(CO)(PPh3)2 in molten triphenylphosphine results only in loss of the trimethylsilyl ligand and formation of the previously known complex containing an ortho-metallated triphenylphosphine ligand, Os(κ2(C,P)-C6H4PPh2)Cl(CO)(PPh3)2. In contrast, heating the five-coordinate osmium-methyl complex, Os(SiMe3)(Me)(CO)(PPh3)2 (1), in the presence of triphenylphosphine results mainly, not in tetramethylsilane elimination, but in ortho-silylation as well as ortho-metallation of different triphenylphosphine ligands giving, Os(κ2(Si,P)-SiMe2C6H4PPh2)(κ2(C,P)-C6H4PPh2)(CO)(PPh3) (5). A byproduct of this reaction is the non-silicon containing di-ortho-metallated complex, Os(κ2(C,P)-C6H4PPh2)2(CO)(PPh3) (6). A similar reaction occurs when Os(SiMe3)(Me)(CO)(PPh3)2 (1) is heated in the presence of tri(N-pyrrolyl)phosphine producing Os(κ2(Si,P)-SiMe2C6H4PPh2)(κ2(C,P)-C6H4PPh2)(CO)[P(NC4H4)3] (7) but a better synthesis of 7 is to treat 5 directly with tri(N-pyrrolyl)phosphine. Heating the six-coordinate complex, Os(SiMe3)(Me)(CO)2(PPh3)2 (2), gives two complexes both containing ortho-metallated triphenylphosphine, one with loss of the trimethylsilyl ligand, giving the known complex, Os(κ2(C,P)-C6H4PPh2)H(CO)2(PPh3), and the other with retention of the trimethylsilyl ligand, giving Os(SiMe3)(κ2(C,P)-C6H4PPh2)(CO)2(PPh3) (8). Crystal structure determinations for 5, 6, 7 and 8 have been obtained.  相似文献   

7.
Three different routes have been investigated for the preparation of 6-aryl-N-(1-arylethyl)thienopyrimidin-4-amines. First the possibilities of selective Suzuki reactions on 6-bromo-4-chlorothienopyrimidine were investigated. The preference for mono arylation at C-6 could be increased, in the case of Pd(PPh3)4 catalysis, by reducing the water content of the reaction, or by using less electron rich Pd-ligands. The highest selectivity was obtained with Pd(OAc)2 or Pd2(dba)3, while reactions with the more electron rich Pd(PPh3)4 and especially XPhos gave a lower mono- to dicoupled product ratio. Secondly, two alternative strategies avoiding this selectivity issue were tested. Suzuki reaction on C-6 of 6-bromothienopyrimidin-4(3H)-one (three examples) proceeded in 70-89% yield using Pd(PPh3)4 in dioxane/water. Similar conditions on 4-amino-6-bromo-thienopyrimidine (eight examples) gave 67-95% yield. The reaction could be performed with boronic acids containing nonprotected phenolic groups in the ortho, meta and para positions. By prolonging the reaction time, coupling with sterically crowded arylboronic acids was also efficient. Diarylation of 6-bromo-4-chlorothienopyrimidine gave the corresponding 4,6-diarylated derivatives in 71-80% yield depending on the nature of the arylboronic acid.  相似文献   

8.
The palladium-catalyzed substitution of alkyl 4,6-di-O-acetyl-α-d-erythro-hex-2-eno-pyranosides using NaN3 as the nucleophile gave predominantly the corresponding alkyl 2-azido-2,3,4-trideoxy-α-d-threo-hex-2-enopyranosides in the presence of Pd(PPh3)4. However, alkyl 6-O-acetyl-4-azido-2,3,4-trideoxy-α-d-erythro-hex-2-enopyranosides were obtained as the major products using Pd(PPh3)4 as the catalyst in the presence of dppb as the added ligand. Conversely, alkyl 6-O-(tert-butyldimethylsilyl)-4-O-methoxycarbonyl-2,3-dideoxy-α-d-hex-2-enopyranosides gave exclusively alkyl 4-azido-6-O-(tert-butyldimethylsilyl)-2,3,4-trideoxy-α-d-erythro-hex-2-enopyranosides in the presence of Pd2(dba)3/PPh3 as the catalyst and Me3SiN3 as the nucleophile. The bis-hydroxylation followed by hydrogenation of ethyl 4-azido-2,3,4-trideoxy-α-d-erythro-hex-2-enopyranoside afforded the corresponding 4-amino-α-d-mannopyranoside, when propyl 2-azido-2,3,4-trideoxy-α-d-threo-hex-3-enopyranoside gave the 2-amino-α-d-altropyranoside under the same conditions.  相似文献   

9.
Unprecedented silyl‐phosphino‐carbene complexes of uranium(IV) are presented, where before all covalent actinide–carbon double bonds were stabilised by phosphorus(V) substituents or restricted to matrix isolation experiments. Conversion of [U(BIPMTMS)(Cl)(μ‐Cl)2Li(THF)2] ( 1 , BIPMTMS=C(PPh2NSiMe3)2) into [U(BIPMTMS)(Cl){CH(Ph)(SiMe3)}] ( 2 ), and addition of [Li{CH(SiMe3)(PPh2)}(THF)]/Me2NCH2CH2NMe2 (TMEDA) gave [U{C(SiMe3)(PPh2)}(BIPMTMS)(μ‐Cl)Li(TMEDA)(μ‐TMEDA)0.5]2 ( 3 ) by α‐hydrogen abstraction. Addition of 2,2,2‐cryptand or two equivalents of 4‐N,N‐dimethylaminopyridine (DMAP) to 3 gave [U{C(SiMe3)(PPh2)}(BIPMTMS)(Cl)][Li(2,2,2‐cryptand)] ( 4 ) or [U{C(SiMe3)(PPh2)}(BIPMTMS)(DMAP)2] ( 5 ). The characterisation data for 3 – 5 suggest that whilst there is evidence for 3‐centre P?C?U π‐bonding character, the U=C double bond component is dominant in each case. These U=C bonds are the closest to a true uranium alkylidene yet outside of matrix isolation experiments.  相似文献   

10.
N,N-Dimethylneopentylamine reacts with Pd(MeCO2)2 to give a novel trinuclear cyclopalladated complex [Me2NCH2CMe2CH2Pd(μ-MeCO2)2Pd(μ-MeCO2)2PdCH2CMe2CH2NMe2]?-0.5C6H6 (I). The reaction of I with PPh3 affords both trans-[Pd(MeCO2)2(PPh3)2] (II) and [Pd(CH2CMe2CH2NMe2)(MeCO2)(PPh3)] (III). The reaction of III with LiCl yields a mononuclear cyclopalladated complex, [Pd(CH2CMe2CH2NMe2)Cl(PPh3)] (IV).  相似文献   

11.
Treatment of pentafluorophenyltrimethylsilane (I) and cyanomethyltrimethylsilane (II) with enolizable ketones in the presence of a catalytic amount of potassium cyanide-18-crown-6 complex gave the corresponding trimethylsilyl enol ethers. The same dehydrogenative silylation of acetylacetone and benzoylacetone with silane I was extended to the preparation of 2,4-bis(trimethylsiloxy)-1,3-pentadiene and 1-phenyl-1,3 -bis(trimethylsiloxy)-1,3-butadiene, respectively. The dehydrogenative silylation of acetylacetone and benzoylacetone with dimethylbis(pentafluorophenyl)silane under the same conditions affords novel heterocyles 5-methylene-2,6-dioxa-1-silacylohex-3-enes. In the reaction studied the silylating ability of the silanes increases in the order Me3SiCN ? Me2Si(CN)2 < Me3SiCH2CN < Me3SiC6F5 ? Me2Si(C6F5)2. On the other hand, potassium cyanide-18-crown-6 complex catalyzed the addition of silane I or II to a carbonyl group of non-enolizable compounds such as benzaldehyde, crotonaldehyde, and methyl(triethylgermyl)ketene.  相似文献   

12.
A dramatic improvement of the catalytic activity was observed when a phosphine was added in allylic alkylation reactions catalyzed by (NHC)Pd(η3-C3H5)Cl complexes. Consequently, several palladium complexes, generated in situ from different NHC-silver complexes, [Pd(η3-C3H5)Cl]2 and PPh3, were tested in this reaction to evaluate their potential. High reaction rates and conversions could be obtained with this catalytic system in the alkylation of allylic acetates with dimethylmalonate, particularly under biphasic conditions using water/dichloromethane and KOH 1 M as the base. These conditions are experimentally more convenient and gave higher reaction rates than the classical anhydrous conditions (NaH/THF). In this system, the phosphine is essential since no conversion was obtained when it is not present. The steric hindrance of the carbene ligand has a great influence on the activity and the stability of the catalytic system. The best NHC ligands for this reaction are either 1-mesityl-3-methyl-imidazol-2-ylidene or 1-(2,6-diisopropylphenyl)-3-methyl-imidazol-2-ylidene which are less bulky among the NHC tested. These two ligands led in 5 min to a complete conversion at 20 °C. The Pd-catalyzed allylic amination reaction using (E)-1,3-diphenylprop-3-en-yl acetate and benzylamine was also tested with (NHC)(PPh3)Pd complexes and under the biphasic conditions. This reaction was found to be slower than the alkylation with dimethylmalonate but a complete conversion could be reached in 6 h at 20 °C using K2CO3 1 M as the base. NMR experiments indicated that mixed (NHC)(PPh3)Pd complexes are formed in situ but their structure could not be established exactly.  相似文献   

13.
Previously unknown N,N-bis[ethoxy(methyl)silylmethyl]amines MeN[CH2SiMem(OEt)3-m ]2 (m = 0-2) were synthesized. According to UV spectral data, only MeN[CH2SiMe2(OEt)]2 form hydrogen bond with phenol in a heptane solution. The amines with m = 0 and 1 fail to forms hydrogen bond with phenol [under the same conditions, N-(triethoxysilylmethyl)dimethylamine Me2NCH2Si(OEt)3 forms a strong hydrogen bond with phenol]. All the amines (m = 0-2) enter transetherification with phenol to give compounds of the general formula MeN[CH2SiMem m(OPh) n (OEt)3-m-n]2 (m = 0-2, n = 1-3). Refluxing of N,N-bis[ethoxy(methyl) silylmethyl]amines with excess phenol results in cleavage of the Si-C bond by phenol, providing phenoxysilanes MemmSi(OPh)4-m (m = 0-2) and trimethylamine.  相似文献   

14.
Syntheses and single crystal X-ray diffraction studies of the η1-halogenophosphaalkene complexes trans-[RhCl(PPh3)21-PXC(SiMe3)2}] (X = F, Cl) are reported.  相似文献   

15.
《Polyhedron》1986,5(6):1183-1190
The Pd(II) and Rh(I) complexes of tetra-acetylethane [H2dahd (3,4-diacetyl-2,4-hexadiene-2,5-diol)] with O,O′-bonded chelates, represented as [M2(O2,O2-dahd)(L2)2][X]m {M = Pd, L2 = (PPh3)2 or bdpe [1,2-bis(diphenylphosphino)ethane], X = BF4 or PF6, m = 2; M = Rh, L = Co, m = 0}, have recently been prepared. [Pd2(O2,O2-dahd)-(PPh3)4][PF6]2 reacts with the potentially bidentate 1,10-phenanthroline (phen) to give the five-coordinate complex [Pd(PPh3)(phen)2][PF6]2 and [Pd(O1,O1-dahd)(phen)]n, the latter of which is rather insoluble in organic solvents. [Pd(O1, O1-dahd)(phen)]n in CH2Cl2 readily transforms to a monomer complex [Pd(C3, O′-dahd)phen)]. These anomalous Pd(II) and Rh(I) complexes of the tetra-acetylethane dianion have been characterized from elemental analyses, conductance, IR, 1H and 13C NMR spectroscopy, magnetic susceptibility and ESR spectroscopy.  相似文献   

16.
The reactions of Pt(PPH3)4 and Pt(C2H4)(PPh3)2 with CH2ClI have been investigated. The product of the reaction of Pt(PPh3)4 with CH2ClI is the cationic ylide complex cis-[Pt(CH2PPh3)Cl(PPh3)2][I], whereas the reaction of Pt(C2H4)-(PPh3)2 gives the oxidative addition product Pt(CH2Cl)I(PPh3)2. Reaction of cis- or trans-Pt(CH2Cl)I(PPh3)2] with PPh3 gives the complex cis-[Pt(CH2PPh3)-Cl(PPh3)2][I]. The structures of the complexes cis-[Pt(CH2PPh3X(PPh3)2][I] (where X = Cl or I) have been determined by X-ray crystallography. Both complexes crystalize in the monoclinic space group P21/n. For X = Cl a 1388.6(7), b 2026.7(10), c 1823.9(9) pm, β 96.51(2)° and R converged to 0.075 for 3542 observed reflections; structural parameters Pt-Cl 240(1), Pt-C(3) 212(2), Pt-P(2) (trans to Cl) 235(1) and Pt-P(1) (trans to CH2PPh3) 233(1) pm; Cl-Pt-C(3) 86.9(5), C(3)-Pt-P(2) 91.8(5), P(2)-Pt-P(1) 97.0(2) and P(1)-Pt-Cl 85.1(2)°. For X = I, a 1379.4(7), b 2044.4(10), c 1840.0(9) pm, β 96.09(2)° and R converged to 0.071 for 4333 observed reflections; structural parameters Pt-I 266(1), Pt-C(3) 212(2), Pt-P(2) (trans to I) 226(1) and Pt-P(1) (trans to CH2PPh3 233(1) pm; I-Pt-C(3) 87.2(5), C(3)-Pt-P(2) 91.5(5), P(2)-Pt-P(1) 96.5(2) and P(1)-Pt-I 85.6(1)°. Some other complexes of the type cis-[Pt(CH2PPh3)X(PPh3)2]Y are also described.  相似文献   

17.
The direct cyclopalladation of 3-methoxyimino-2-(4-chlorophenyl)-3H-indole (1a) and 3-methoxyimino-2-phenyl-3H-indole (1b) results in the regioselective activation of the ortho σ[C(sp2, phenyl)-H] bond affording (μ-OAc)2[Pd{κ2-C,N-C6H3-4R-1-(C8H4N-3′-NOMe)}]2 (2) {R = Cl (2a) or H (2b)} that contain a central “Pd(μ-OAc)2Pd” core. Compounds 2a and 2b reacted with triphenylphosphine (in a molar ratio PPh3:2 = 2) giving [Pd{κ2-C,N-C6H3-4R-1-(C8H4N-3′-NOMe)}(OAc)(PPh3)] (3) {R = Cl (3a) or H (3b)}. Treatment of 2a or 2b with a slight excess of LiCl in acetone produced the metathesis of the bridging ligands and the formation of (μ-Cl)2[Pd{κ2-C,N-C6H3-4R-1-(C8H4N-3′-NOMe)}]2 (4) {R = Cl (4a) or H (4b)} with a central “Pd(μ-Cl)2Pd” moiety. The reactions of 4a or 4b with deuterated pyridine (py-d5) or triphenylphosphine gave the monomeric derivatives [Pd{κ2-C,N-C6H3-4R-1-(C8H4N-3′-NOMe)}Cl(L)] with R = Cl or H and L = py-d5 (5) or PPh3 (6). The crystal structure of 6b·1/2CH2Cl2 confirmed the mode of binding of the ligand, the nature of the metallated carbon atom and a trans-arrangement of the phosphine ligand and the heterocyclic nitrogen. Theoretical calculations on the free ligands are also reported and have allowed the rationalization of the regioselectivity of the cyclopalladation process.  相似文献   

18.
Reaction of benzene solutions of arylacetylenes with 1 equiv. of chloroacetone and 2 equiv. of Et3N, using a mixture of (PPh3)4 Pd and CuJ as catalyst, affords 1,4-diaryl-butadiynes in very good yields. Under similar reaction conditions aliphatic 1-alkynes yield mixtures of simmetrically disubstituted 1,4-dialkyl-1,3-butadiynes and of 3-alkyl-4-(1-alkynyl)-hexa-1,5-diyn-3-enes.  相似文献   

19.
The kinetics of cyclohexene hydrocarbomethoxylation catalyzed by the Pd(PPh3)2Cl2-PPh3-p-toluenesulfonic acid (TSA) is reported. The reaction is first-order with respect to cyclohexene and TSA and of order 0.5 with respect to Pd(PPh3)2Cl2. The reaction rate as a function of CO pressure or methanol or PPh3 concentration passes through an extremum. The chloride anion inhibits the reaction. A mechanism involving cationic hydride complexes as intermediates is suggested. A rate equation is set up by the quasi-steady-state treatment of experimental data.  相似文献   

20.
The reactions of the dichloropermethylsilanes Cl(SiMe2)nCl (n = 1?6) with dilithium phenylphosphide yield a series of novel heterocyclic phosphasilanes. For n = 4, 5 and 6, the reaction leads to the corresponding 5-, 6- and 7-membered cyclic monophosphapolysilanes PhP(SiMe2)n, but when n = 3, a polymeric material of probable formula [PhP(SiMe2)3]n is formed. For n = 2, ring closure again occurs, to yield the 6-membered P2Si4 ring compound [PhP(SiMe2)2]2. With dimethyldichlorosilane, cyclization results in the dimeric phosphasilane (PhPSiMe2)2 at ?40°C, and the corresponding trimeric derivative (PhPSiMe2)3 at +40°C. These two ring sizes exist in an equilibrium (PhPSiMe2)2 ? (PhPSiMe2)3, the dimer being stable at room temperature, but being converted into the trimer above 150°C. The 1H, 13C, 29Si and 31P NMR parameters are reported for all the compounds, and the chemical shifts and coupling constants interpreted in terms of the molecular and electronic structures of the different ring sizes.  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