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 953 毫秒
1.
Reactions of acyl iodides RCOI (R = Me, Ph) with organosilicon compounds involve cleavage of the Si-O-C and Si-O-Si fragments. Acetyl iodide reacts with alkyl(alkoxy)silanes with evolution of heat, and cleavage of the Si-O bond results in the formation of oligo-or polysiloxanes, alkyl iodides, and alkyl acetates. 1,3-Diacetoxytetramethyldisiloxane is formed in the reaction of acetyl iodide with dimethoxy(dimethyl)silane. Acyl iodides readily react with 1-ethoxysilatrane to give 1-acyloxysilatranes as a result of cleavage of the C-O bond. The reaction of acetyl iodide with hexaethyldisiloxane yields triethylsilyl acetate and triethyliodosilane, while in the reaction with octamethyltrisiloxane iodo(trimethyl)silane and dimethyl(trimethylsiloxy)silyl acetate are obtained.  相似文献   

2.
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   

3.
Reaction of acyl iodides RCOI (R = Me, Ph) with triorganylsilanes R′2R″SiH in toluene gives 50–60% of the corresponding triorganyliodosilanes R′2R″SiI. Triethylsilane reacts with the same acyl iodides under solvent-free conditions to afford the corresponding aldehyde and triethyliodosilane as primary products. Triethyliodosilane undergoes subsequent transformations into hexaethyldisiloxane and triethyl(acyloxy)silane Et3SiOCOR (R = Me, Ph). Reactions of acyl iodides RCOI (R = Me, Ph) with triphenylgermane in the absence of a solvent lead to formation of iodo(triphenyl)germane in more than 90% yield.  相似文献   

4.
The hydrolysis reactions of 8 different N-bearing alkoxy-silane coupling agents, namely: 3-cyanopropyl triethoxy silane (CPES), triethoxy-3-(2-imidazolin-1-yl) propyl silane (IZPES), and amino silanes, 3-aminopropyl triethoxy silane (APES), 3-aminopropyl trimethoxy silane (APMS), 3-(2-aminoethylamino)propyl trimethoxysilane (DAMS), 3-[2-(2-aminoethylamino)ethylamino] propyl trimethoxysilane (TAMS), 4-amino-3,3-dibutyl trimethoxy silane (ADBMS) and trimethoxy [3-(phenylamino)propyl] silane (PAPMS) were carried out in ethanol/water 80/20 (w/w) solutions in acidic media and followed in situ by 1H-, 13C- and 29Si-NMR spectroscopy. Acidic conditions were selected in order to enhance the formation of silanol and to slow down the self condensation reactions of the hydrolyzed functions. 29Si NMR spectroscopy revealed the formation of intermediate species, particularly the solvolysis of γ-amino silanes by reaction exchange with the alcoholic solvent.  相似文献   

5.
The introduction of the organosilicon substituent into the α‐position of an amino group results in cardinal change of the amine reactivity irrespective of the coordination state of silicon. Amines R2NCH2SiX3 [R = Me, Et, PhCH2, CH2SiX3; SiX3 = SiMe3, Si(OEt)3, Si(OCH2CH2)3N] easily react with AgNO3, to give the corresponding ammonium salts (R2NH+ CH2SiX3)·NO3?. At the same time, Ag(I) is reduced to Ag(0). The interaction of N‐methyl‐N,N‐bis(silatranylmethyl)amine with AgNO3 has been investigated by EPR spectroscopy. It was proven that the reaction involved a single electron transfer stage with the formation of cation radical of this amine. A mechanism of the reaction is proposed. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

6.
The reactions of iodo(trimethyl)silane with N,N-dimethylformamide and N,N-dimethylacetamide Me2NCOR (R = H, Me) at a molar ratio of 1: 2 involved mainly cleavage of the N-C(=O) bond with formation of up to 80% of N,N-dimethyltrimethylsilylamine Me3SiNMe2 and the corresponding acyl iodide RCOI. In the reaction with N,N-dimethylformamide, formyl iodide HCOI was detected for the first time by gas chromatography-mass spectrometry. The contribution of Me-N bond cleavage, leading to N-methyl-N-trimethylsilyl derivative Me(Me3Si)NCOR and methyl iodide was considerably smaller. Another by-product was the corresponding N-methyl imide MeN(COR)2 formed by reaction of the initial amide with acyl iodide. The primary intermediate in the reaction of iodo(trimethyl)silane with DMF and DMA is quaternary ammonium salt [Me2(Me3Si)N+COR] I which decomposes via dissociation of the N-CO and N-Me bonds.  相似文献   

7.
Novel organylthio(alkoxy)silanes (I, II, III and XII) and organylthio(diethylamino)silanes (IV, V) are described. They were prepared by treating lithium or lead thiolates with the corresponding chlorosilanes or by cleavage of dimethylbis(diethylamino)silane with thiols. Phenylthiosilanes (Me3SiSPh, III and XIII) furthermore can be obtained by reaction of chlorosilanes with benzenethiol in the presence of tertiary amines. The SiS bond of Me3SiSPh is cleaved by chlorosilanes like Me2Si(NEt2)Cl or Me2Si(OPr)Cl. This reaction is a convenient route to prepare compounds I and IV. The physical and chemical properties of the novel compounds were investigated.  相似文献   

8.
Reaction of acyl iodides RCOI (R = Me, Ph) was studied with acyclic and cyclic ethers (Et2O, MeCHCH2(O), ClCH2CHCH2(O), THF, O(CH2CH2)2O, EtOCH2CH2OH, EtOCH = CH2, PhOEt]. The reaction occurred with the rupture of one or two CO bonds furnishing the corresponding iodides and esters.  相似文献   

9.
Compounds of Silicon. 140. Sterical Overloaded Compounds of Silicon. 24. Disupersilylsilanes R*2SiX2, Disupersilyldisilanes R*2XSi–SiX3, and Tetrasupersilyltetrasilanes R*2XSi–SiX2–SiX2–SiXR*2 Supersilylsilanes R*2SiX2, disupersilyldisilanes R*2XSi–SiX3 and tetrasupersilyltetrasilanes R*2XSi–SiX2–SiX2–SiXR*2 (R* = supersilyl = SitBu3; X = H, Me, Ph, Hal, OH, OTf) are prepared in organic solvents (i) by reactions of supersilylhalosilanes R*X2SiHal with supersilyl sodium NaR* (Hal/R* exchange), (ii) by reactions of halosilanes X3SiHal with silanides NaSiXR*2 (Hal/SiXR*2 exchange), (iii) by dehalogenations of disupersilylhalodisilanes R*2XSi–SiX2Hal with Na, (iv) by insertions of supersilylsilylenes R*XSi into the NaSi‐bond of supersilylsodium NaR*, (v) by reactions of disupersilylated halosilanes and ‐disilanes R*2XSiHal and R*2XSi–SiX2Hal with H (Hal/H exchange), (vi) by reactions of the title silanes (X = H) with halogens Hal2 (H/Hal exchange), (vii) by reactions of the title silanes (X = Hal) first with Na (Hal/Na exchange), then with agents for protonation (Na/H exchange) or halogenation (Na/Hal exchange), (viii) by reactions of the title silanes (X = Hal) with nucleophiles like F, H2O (Hal/F or Hal/OH exchange) or (ix) by reactions of the title silanes (X = H) with strong acids like HOTf (H/OTf exchange). The colorless compounds are characterized by IR, NMR and X‐ray structure analyses (structures of R*2SiX2 with X = H, F, Cl and R*2HSi–SiHX–SiHX–SiHR*2 with X = H, Br). They may thermolize under formation of silylenes (e. g. R*2SiX2 → R*X + R*SiX) and are normally stable for hydrolysis. For other reactions confer preparation of the title silanes (i–ix).  相似文献   

10.
The new cyclic silanes [(C(5)H(3)N)(CH(2)O)(2)SiMe(2)](2) (1) and (C(5)H(3)N)(CH(2)CPh(2)O)(2)SiMe(2) (2) containing 16-membered and 10-membered rings, respectively, were prepared by the condensation reaction of Me(2)SiCl(2) with an appropriate pyridine diol in the presence of Et(3)N. X-ray studies show that the dimeric formulation for 1 represents a tetracoordinate cyclic silane, whereas 2 has a geometry halfway from a tetrahedron toward a trigonal bipyramid (TBP) as a result of Si-N(ax) donor action. (29)Si and (1)H NMR indicate retention of the coordination geometry for 2 in solution that undergoes rapid Si-N cleavage and ring rearrangement. In comparison with other silanes containing five- and six-membered rings that exhibit nitrogen or oxygen coordination, the presence of larger rings, as in 2 and related silanes having sulfur coordination, indicates that retention of donor action persists, thus largely ruling out ring size as a dominant factor controlling the possibility of donor action at silicon. The dimeric silane 1 crystallizes in the triclinic space group P&onemacr; with a = 6.347(3) ?, b = 12.455(4) ?, c = 14.289(5) ?, alpha = 101.63(3) degrees, beta = 102.99(3) degrees, gamma = 104.71(3) degrees, and Z = 2. The cyclic silane 2 crystallizes in the triclinic space group P&onemacr; with a = 9.733(4) ?, b = 10.938(2) ?, c = 14.312(3) ?, alpha = 89.03(2) degrees, beta = 74.59(3) degrees, gamma = 79.24(3) degrees, and Z = 2. The final conventional unweighted residuals are 0.040 (1) and 0.039 (2).  相似文献   

11.
Reactions of acyl iodides RCOI (R=Me, Ph) with glycine, β-alanine, and γ-aminobutyric acid were investigated. The reaction proceeded easily at room temperature without solvent involving both functional groups H2N and COOH. The prevalence of one of the reaction directions depends on the acidity of the amino acid. The more acidic glycine (pКa 2.4) reacts with RCOI affording mainly N-acylated product, whereas β-alanine (pК a 3.60) and especially γ-aminobutyric acid (pКa 4.06) are predominantly involved into exchange iodination furnishing the corresponding aminoacyl iodides.  相似文献   

12.
The reactions of diethylaminoethynyl(trimethyl)silane (1), bis(diethylaminoethynyl)methylsilane (2), diethylaminoethynyl(trimethyl)germane (3), dimethylaminoethynyl(triethyl)germane (4), diethylaminoethynyl(trimethyl)stannane (5) and methyl(phenyl)aminoethynyl(trimethyl)stannane (6) with trialkylboranes [BEt3 (7b), BPr3 (7c), BiPr3 (7d) and 9‐alkyl‐9‐borabicyclo[3.3.1]nonanes 9‐Me‐9‐BBN (8a) and 9‐Et‐9‐BBN (8b)] were studied. The alkynes 1 and 2 did not react even with boiling BEt3, whereas the reactions of 3–6 afforded mainly novel enamines [(E)‐1‐amino‐1‐trialkylgermyl‐2‐dialkylboryl‐alkenes (9, 10), (E)‐1‐diethylamino‐1‐trimethylstannyl‐2‐dialkylboryl‐alkenes (11, 12), (E)‐1‐methyl(phenyl)amino‐1‐trimethylstannyl‐2‐dialkylboryl‐alkenes (13, 14)]. This particular stereochemistry is unusual for products from 1,1‐organoboration reactions, indicating a special influence of the amino group. The starting materials and products were characterized by multinuclear magnetic resonance spectroscopy (1H, 11B, 13C, 15N, 29Si, 119Sn NMR). Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

13.
Photolysis of acyl iodides RCOI (R = Me, Me2CH, Ph) under UV irradiation in toluene environment for 20–55 h proved to be a simple and efficient method of preparation of symmetrical α-diketones RCOCOR. In contrast, the photolysis under the same conditions of acyl iodides RCOI [R = Me(CH2)3, Me3C] did not lead to the formation of the corresponding diacyls, and the reaction products were unexpected 1,1-bis(4-methylphenyl)pentane and a mixture of isomeric 3- and 4-methyl(tert-butyl)benzenes respectively. The most probable mechanism of their formation is the primary photochemical acylation of toluene in the aromatic ring followed by the photochemical reduction of the arising butyl 4-methylphenyl ketone in the case of the valeroyl iodide or the photochemical Norrish type I cleavage of isomeric 3- and 4-methylphenyl (tert-butyl) ketones in event of the pivaloyl iodide. In the photolysis of acetyl iodide (R = Me) in benzene or toluene alongside the diacetyl formation polyarylation process was observed of acylated and iodinated into the aromatic ring solvents with the formation of polymeric products with semiconductor and paramagnetic properties.  相似文献   

14.
Fifteen previously unknown piperidinosilanes of the type R4–m Si[(CH2) n NC5H10] m (m=1–3; n=0–3). have been synthesized by the reaction of piperidine with diethylaminotrimethylsilane, bis(diethylamino)dimethylsilane, chlorobis(dimethylamino)methylsilane, trichloro(methyl)silane, trialkyl(chloromethyl)silanes, dialkoxy(alkyl)(chloromethyl)silanes, trialkylvinylsilanes, trialkyl(3-chloropropyl)silanes, and 3-chloropropyl(diethoxy)methylsilane. The piperidinosilanes (n>0) have been converted into the corresponding hydrochlorides and methiodides.For part VIII, see [1].  相似文献   

15.
A bimetallic system of Pd/CuF2, catalytic in Pd and stoichiometric in Cu, is very efficient and selective for the coupling of fairly hindered aryl silanes with aryl, anisyl, phenylaldehyde, p‐cyanophenyl, p‐nitrophenyl, or pyridyl iodides of conventional size. The reaction involves the activation of the silane by CuII, followed by disproportionation and transmetalation from the CuI(aryl) to PdII, upon which coupling takes place. CuIII formed during disproportionation is reduced to CuI(aryl) by excess aryl silane, so that the CuF2 system is fully converted into CuI(aryl) and used in the coupling. Moreover, no extra source of fluoride is needed. Interesting size selectivity towards coupling is found in competitive reactions of hindered aryl silanes. Easily accessible [PdCl2(IDM)(AsPh3)] (IDM = 1,3‐dimethylimidazol‐2‐ylidene) is by far the best catalyst, and the isolated products are essentially free from As or Pd (<1 ppm). The mechanistic aspects of the process have been experimentally examined and discussed.  相似文献   

16.
A Pd(dba)2–P(OEt)3 combination allowed the silastannation of arylacetylenes, 1‐hexyne or propargyl alcohols with tributyl(trimethylsilyl)stannane to take place at room temperature, producing (Z)‐2‐silyl‐1‐stannyl‐1‐substituted ethenes in high yields. Novel silyl(stannyl)ethenes were fully characterized by 1H‐, 13C‐, 29Si‐ and 119Sn‐NMR as well as infrared and mass analyses. Treatment of a series of (Z)‐1‐aryl‐2‐silyl‐1‐stannylethenes and (Z)‐1‐(3‐pyridyl)‐2‐silyl‐1‐stannylethene with hydrochloric acid or hydroiodic acid in the presence of tetraethylammonium chloride (TEACl) or tetrabutylammonium iodide (TBAI) led to the exclusive formation of (E)‐trimethyl(2‐arylethenyl)silanes with high stereoselectivity. A similar reaction of (Z)‐1‐(2‐anisyl)‐2‐silyl‐1‐stannylethene also produced E‐type trimethyl[2‐(2‐anisyl)ethenyl]silane, while (Z)‐trimethyl [2‐(2‐pyridyl)ethenyl]silane was produced exclusively from (Z)‐1‐(2‐pyridyl)‐2‐silyl‐1‐stannylethene. Protodestannylation of (Z)‐1‐[hydroxy(phenyl)methyl]‐2‐silyl‐1‐stannylethene with trifluoroacetic acid took place via the β‐elimination of hydroxystannane, providing trimethyl(3‐phenylpropa‐1,2‐dienyl)silane quite easily. The destannylation products were also fully characterized. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

17.
The structural and thermodynamic characteristics of SiX4·Py and SiX4·2Py adducts (X = H, F, Cl, Br) were calculated by ab initio and DFT methods (RHF and B3LYP). The resulting data were used to estimate for the first time the enthalpies of sublimation of trans-SiX4·2Py complexes. The distortion energies of the donor and acceptor fragments and the energies of the Si-N bonds in the 1:1 and 1:2 halide complexes were calculated. The high distortion energy makes thermodynamically unfavorable equatorial monopyridine adducts with Si-N bond energies of 150-200 kJ/mol. In trans 1:2 complexes, pyridine acts as a weaker donor than ammonia with respect to silicon tetrahalides.  相似文献   

18.
The hydrolysis of four alkoxy-silane coupling agents, 3-methacryloxypropyl trimethoxy silane (MPMS), 3-mercaptopropyl trimethoxy silane (MRPMS), octyl triethoxy silane (OES) and 3-aminopropyl triethoxy silane (APES) was carried out in an ethanol/water 80/20 (w/w) solution under acidic, alkaline and neutral conditions and followed by 1H, 13C and 29Si NMR spectroscopy. It was found that the kinetic rate of the hydrolysis of the silanes under neutral conditions was very low, except for APES, which displayed the fastest reaction speed. The addition of TEA catalyzed both silane hydrolysis and self condensation reactions. Acidic conditions enhanced the hydrolysis and the ensuing silanol entities were quite stable. In fact, these conditions slowed down the rate of the self condensation reactions, as deduced from in situ 1H and 13C NMR. Thanks to in situ 29Si NMR spectroscopy, the nature of the intermediary species versus reaction time was established.  相似文献   

19.
Reactions of acetyl iodide with dialkyl and dialkenyl sulfides RSR (R = Et, Bu, CH2=CH, CH2=CHCH2) and with disulfides RSSR (R = Pr, C6H13, PhCH2) were studied. Dialkyl sulfides reacted with MeCOI to give the corresponding alkyl ethanethioates and alkyl iodides as a result of cleavage of the S-C bond. The reactions of acetyl iodide with divinyl and diallyl sulfides involved addition across the double bond and subsequent polymerization of 1-alkenylsulfanyl-2(3)-iodoalkyl methyl ketones. Dialkyl disulfides RSSR (R = Pr, C6H13) and dibenzyl disulfide reacted with acetyl iodide via cleavage of the S-S bond to produce the corresponding ethanethioates and organylsulfenyl iodides. The latter underwent disproportionation to form the initial disulfide and molecular iodine.  相似文献   

20.
New 1-organyl-2-azasilatran-3-ones have been synthesized via the reaction of trifunctional silanes RSiX3 (R = Et, Pr, Ph, CH2=CH, or ClCH2; X = Cl or Me2N) with N,N-bis(2-hydroxyethyl)glycinamides (HOCH2CH2)2NCH2C(O)NHR (R = H or Me) and their N′,O-trimethylsilyl derivatives. The obtained products can be hydrolyzed to give the corresponding organylsilanetriols. Lithiation of 1-methyl- and 1-phenyl-2-azasilatran-3-ones with n-butyllithium or their reduction with lithium aluminum hydride leads to the products of splitting of the atrane backbone RSiBu3 and RSiH3 (R = Me or Ph), 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