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.
The dimesitylpropargylphosphanes mes2P?CH2?C≡C?R 6 a (R=H), 6 b (R=CH3), 6 c (R=SiMe3) and the allene mes2P?C(CH3)=C=CH2 ( 8 ) were reacted with Piers’ borane, HB(C6F5)2. Compound 6 a gave mes2PCH2CH=CH(B(C6F5)2] ( 9 a ). In contrast, addition of HB(C6F5)2 to 6 b and 6 c gave mixtures of 9 b (R=CH3) and 9 c (R=SiMe3) with the regioisomers mes2P?CH2?C[B(C6F5)2]=CRH 2 b (R=CH3) and 2 c (R=SiMe3), respectively. Compounds 2 b , c underwent rapid phosphane/borane (P/B) frustrated Lewis pair (FLP) reactions under mild conditions. Compound 2 c reacted with nitric oxide (NO) to give the persistent FLP NO radical 11 . The systems 2 b , c cleaved dihydrogen at room temperature to give the respective phosphonium/hydridoborate products 13 b , c . Compound 13 c transferred the H+/H? pair to a small series of enamines. Compound 13 c was also a metal‐free catalyst (5 mol %) for the hydrogenation of the enamines. The allene 8 reacted with B(C6F5)3 to give the zwitterionic phosphonium/borate 17 . The ‐PPh2‐substituted mes2P‐propargyl system 6 d underwent a typical 1,2‐P/B‐addition reaction to the C≡C triple bond to form the phosphetium/borate zwitterion 20 . Several products were characterized by X‐ray diffraction.  相似文献   

2.
1-hydroperfluoroalkynes (RFCCH, RF  C4F9, C6F13, C8F17) were prepared from the corresponding alkenes by a multistep process involving bromination, dehydrobromination and debromination reactions. The influence of the perfluorinated chain's length on chemical reactivity is illustrated by the important changes in experimental procedure, with respect to the lower terms, needed for the bromination and debromination steps. Physical data are reported for compounds RFCBr2CH2Br, RFCBrCHBr and RFCCH.  相似文献   

3.
Hydrosilylation of fluorinated olefins with polyhydromethylsiloxane (PHMS) in the presence of a platinum catalyst was investigated to synthesize fluorosilicone having highly fluorinated alkyl side chains (Rf; CnF2n+1? ). The hydrosilylation of 3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10‐heptadecafluoro‐1‐decene (C8F17CH?CH2) ( 1 ) with poly(dimethylsiloxane‐co‐hydromethylsiloxane) {(CH3)3SiO[? (H)CH3SiO? ]8[? (CH3)2 SiO? ]18Si(CH3)3} ( 4 ) converted the hydrogen bonded to silicons into the 3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10‐heptadecafluorodecyl group or fluorine bonded to silicons in the ratio of about 52:48, and the formation of the byproduct C7F15CF?CHCH3 ( 8 ) was observed. The hydrosilylation of 7,7,8,8,9,9,10,10,11,11,12,12,13,13,14,14,14‐heptadecafluoro‐4‐oxa‐1‐tetradecene (C8F17CH2CH2OCH2CH?CH2) ( 2 ) with 4 converted the hydrogen bonded to silicons into the 7,7,8,8,9,9,10,10,11,11,12,12,13,13,14,14,14‐heptadecafluoro‐4‐oxa‐tetradocyl group bonded to silicons, but an excess amount of 2 was required to complete the reaction because the isomerization of 2 occurred in part to form C8F17CH2CH2OCH?CHCH3 ( 9 ). The hydrosilylation of 4,4,5,5,6,6,7,7,8,8,9,9, 10,10,11,11,11‐heptadecafluoro‐1‐undecene (C8F17CH2CH?CH2) ( 3 ) with 4 converted the hydrogen bonded to silicons into the 4,4,5,5,6,6,7,7,8,8,9,9,10,10,11,11‐heptadecafluoroundecyl group bonded to silicons. This type of fluorinated olefin was successfully applied to the hydrosilylation with other PHMS's that involved a homopolymer of PHMS and a cyclic PHMS. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 3120–3128, 2002  相似文献   

4.
Syntheses and Properties of Bis(perfluoroalkyl)zinc Compounds The conditions for the syntheses of bis(perfluoroalkyl)zinc compounds Zn(Rf)2 · 2 D (Rf = C2F5, n‐C3F7, i‐C3F7, n‐C4F9, n‐C6F13, n‐C7F15, and n‐C8F17; D = CH3CN, tetrahydrofurane, dimethylsulfoxide) are described. Mass spectra, thermal decompositions, 19F‐ and 13C‐NMR spectra are discussed.  相似文献   

5.
Perfluoroalkyl- or nonafluoro-tert-butoxy-alkyl-substituted enantiopure amines having the structure PhCHCH3(NR1R2) [R1 = H, CH3; R2 = (CH2)3C8F17, (CH2)2OC(CF3)3; R1 = R2 = (CH2)3C8F17, (CH2)2OC(CF3)3] are obtained in high yields, when (S)-(−)-1-phenylethylamine is reacted with readily accessible alkylating reagents or fluorous 2° amines (R1 = H; R2 = (CH2)3C8F17, (CH2)2OC(CF3)3) are methylated in a Leuckart-Wallach reaction. The solubility patterns of these novel chiral amines and their hydrochlorides are qualitatively described for a broad spectrum of solvents and the fluorous partition coefficients of the free bases are determined by GC. A novel method for the resolution of enantiomers is disclosed here, which involves the use a half-equivalent of the selected resolving agent in solvent water that displays low solubility for the crystalline diastereomeric salt(s) formed even at temperatures near to its boiling point. Compound (S)-(−)-PhCHCH3[NH(CH2)3C8F17] is found to satisfy all the latter conditions and successfully used for the heat facilitated resolution of the title racemic acid. The circular dichroism (CD) spectra of six novel fluorous (S)-(−)-1-phenylethylamine derivatives are measured in ethanol, trifluoroethanol and hexafluoropropan-2-ol and discussed in detail.  相似文献   

6.
The alkenyl substituted phenoxy–imine complexes [2‐C3H5‐6‐(2, 3, 5, 6‐C6F4H‐N?CH)C6H3O]2TiCl2 (C3H5=? CH2? CH?CH2 or ? CH?CH? CH3) are synthesized and characterized by 1H NMR, 13C NMR, and elemental analysis. When activated by MAO, they show high activity for the polymerization of ethylene to UHMWPE under different conditions (temperatures and polymerization time). Most of the resulting polymers have high molecular weights (>1.0 × 106 g·mol?1) and high melting points as well as crystallinity. To clarify the effect of the alkenyl group on the catalytic performance and the resultant polymer microstructure, the corresponding saturated complexes of type [2‐C3H7?6‐(2, 3, 5, 6‐C6F4H‐N?CH)C6H3O]2TiCl2 where C3H7 = –CH2? CH2? CH3 or ? CH(CH3)2 were synthesized and tested as catalysts in ethylene polymerization under the same reaction conditions. The microstructure and morphologies of these two species of PE samples were fully compared by the analysis of 13C NMR, GPC, DSC, and SEM. As a result, the allyl substituted complex show the highest activity to prepare the highest molecular weight polyethylene of all the catalysts. An interesting feature of the UHMWPE produced by these four catalysts is that they contain only a few short‐chain branches (mainly methyl, isobutyl and 2‐methylhexyl branches) in a low amount (<2.7 branches/1000 C). © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 3808–3818  相似文献   

7.
Bis(N, N′‐dialkyldithiocarbamato)antimony(III) alkylenedithiophosphates of the type [R2NCS2]2 SbS(S)POGO [where NR2 = N(CH3)2, N(C2H5)2 and N(CH2)4; G = ? CH2? C(C2H5)2? CH2? , ? CH2? C(CH3)2? CH2? , ? CH(CH3)? CH(CH3)? and ? C(CH3)2? C(CH3)2? ] were synthesized and characterized by physico‐chemical, spectral [UV, IR and NMR (1H, 13C and 31P)] and thermal (TG, DTA and DSC) analysis. The TG decomposition analysis step of the complex indicated the formation of Sb2S3 as the final product. The first endothermic peak in DSC indicated the melting point of the complexes. These complexes were screened for their antimicrobial activities using the disk diffusion method. All the complexes showed good activity as antibacterial and antifungal agents on some selected bacterial and fungal strains, which increased on increasing the concentration. Chloroamphenicol and terbinafin were used as standards for comparison. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

8.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

9.
Polysulfonyl Amines. VII. Aliphatic Trisulfonyl Amines The compounds N(SO2R1)2(SO2R2) with R1 = R2 = CH3 ( 2a ), R1 = R2 = C2H5 ( 2b ) and R1 = CH3, R2 = C2H5 ( 2c ) are prepared by cleavage of aminostannanes (CH3)3SnN(SO2R1)2 with sulfonyl chlorides R2SO2Cl. A simple synthesis of 2a from AgN(SO2CH3)2 and CH3SO2Cl is described. From the vibrational spectra of 2a , evidence is obtained for a planar NS3 group in this compound. X-ray structure determinations of 2b and HN(SO2C2H5)2 ( 3 ) are reported. In 2b , the NS3 group is approximately planar (S? N? S bond angles 119.0 ± 0.6°, sum of bond angles at N 356.9°); the S? N bond lengths of ca. 173 pm indicate a bond order of 1. In compound 3 , the nitrogen atom has a planar coordination (S? N? S angle 125.3°, sum of bond angles at N 359.3°), the S? N bond lengths of ca. 165 pm correlate with a bond order of 1.3? 1.4.  相似文献   

10.
The hydrodeboration of the (fluoroorgano)trifluoroborates K [RFBF3] [RF = C6F5, XCF=CF (X = F, cis‐ and trans‐Cl, C3F7O, cis‐C2F5, trans‐C4F9, ‐C4H9) and C6F13] and of the organotrifluoroborates K [RBF3] (R = C6H5, cis‐ and trans‐C4H9CH=CH, C4H9 and C8H17) with CH3CO2H (100 %), CF3CO2H (100 %), aqueous HF and anhydrous HF was investigated. In the alkenyltrifluoroborates K [R'CF=CFBF3] the formal replacement of BF3 by a proton occurred stereospecifically under retention of the configuration. The 19F NMR spectra of K [RFBF3] in acids indicate strong interactions of the BF3 group with protons or acid molecules.  相似文献   

11.
The most prominent ion in the mass spectra of C6F5CH2X (X ? H, Br, CH:CH2, COCl, and CH2C6F5) is C7F5H2+, formulated as the pentafluorotropylium cation. This ion is also found, in an amount comparable to the parent ion, in the spectrum of (C6F5)2CH2. The heptafluorotropylium cation is found similarly in the spectrum of C6F5CF3. The mass spectra of (C6F5)2CHBr and [(C6H5)2CH]2 exhibit an ion C13F10H+ as the base peak, which is probably a pentafluorophenylpentafluorotropylium cation. The alcohol (C6F5)2CHOH shows loss of C6F5, followed by 2H, as a major breakdown pathway. The mode of formation, and the subsequent fragmentation, of the major ions in these spectra, are discussed.  相似文献   

12.
Metal Complexes of Biologically Important Ligands, CLVII [1] Halfsandwich Complexes of Isocyanoacetylamino acid esters and of Isocyanoacetyldi‐ and tripeptide esters (?Isocyanopeptides”?) N‐Isocyanoacetyl‐amino acid esters CNCH2C(O) NHCH(R)CO2CH3 (R = CH3, CH(CH3)2, CH2CH(CH3)2, CH2C6H5) and N‐isocyanoacetyl‐di‐ and tripeptide esters CNCH2C(O)NHCH(R1)C(O)NHCH(R2)CO2C2H5 and CNCH2C(O)NHCH(R1)C(O)NHCH (R2)C(O)NHCH(R3)CO2CH3 (R1 = R2 = R3 = CH2C6H5, R2 = H, CH2C6H5) are available by condensation of potassium isocyanoacetate with amino acid esters or peptide esters. These isocyanides form with chloro‐bridged complexes [(arene)M(Cl)(μ‐Cl)]2 (arene = Cp*, p‐cymene, M = Ir, Rh, Ru) in the presence of Ag[BF4] or Ag[CF3SO3] the cationic halfsandwich complexes [(arene)M(isocyanide)3]+X? (X = BF4, CF3SO3).  相似文献   

13.
Cis- and trans-1-Phosphabicyclo[4.4.0]decane A mixture of cis-( 5a ) and trans-1-phosphabicyclo [4.4.0] decane 5b has been prepared by free-radical cyclization of (CH2 = CH? CH2? CH2)2CH? PH2 10 . The isomers could be separated in a pure state. Stereostructures have been assigned by 13C n.m.r. at 153—302 K. Equilibration of 5a and 5b by u.v. irradiation gave ?G°35 ≈? 0 kJ ° mol?1 · Activation parameters for ring inversion of “cis” stereoisomer 5a and its “cis” P-sulfid 17a are found to be ΔG° = 41.9 kJ · mol?1 and 39.7 kJ · mol?1, respectively. Treatment of 5a and 5b with H2O2, sulfur, selenium, HSO3F, CH3I, CS2, and Ni(CO)4, respectively, yield the corresponding derivatives. 1H, 13C, 31P, 77Se n.m.r. and i.r. data are reported.  相似文献   

14.
Free radical addition of an F-alkyl iodide (RFI) to an alkenol or ester, followed by appropriate reduction is an efficient method for preparing the corresponding F-alkyl-alkanols of the homologous series, RF(CH2)n?OH. When n = 2,4 or higher, the two steps take place smoothly. The 1,2,3-substituted systems RFCH2CHYCH2Z, however, are susceptible to surprising difficulties. Reduction of RFCH2CHICH2ON to RF(CH2)3OH by hydrogen and catalyst (strong base acid acceptor), can be done either in one step or via RFCHCHCH2OH; however, dehydrohalogenation may also give the epoxide, and reduction in this case leads to the secondary alcohol, RFCH2CH(CH3)OH. By contrast, reduction of RFCH2CHICH2OAc by tributyltin hydride or with hydrogen over palladium (diethylamine acid acceptor) goes smoothly. Zinc and acid reduction of RFCH2CHICH2OAc gives elimination to RFCH2CHCH2; even RFCHCICH2OH gives RFCHCCH2 besides RFCHCHCH2OH. RFCHCICH2CH2OH, however, with zinc and acid is reduced cleanly to RFCHCHCH2CH2OH.  相似文献   

15.
This paper reports the regioselective synthesis of new trifluoromethylated lipid derivatives, namely, 1-(5-hydroxy-5-trifluoromethyl-3-alkyl-4,5-dihydro-1H-pyrazol-1-yl)alkan-1-ones, through cyclocondensation reactions between a series of fatty hydrazides (palmitoyl, stearoyl, and oleoyl hydrazides) obtained from fatty acids from renewable resources (1,1,1-trifluoro-4-alkoxy-3-alken-2-ones [F3CC(O)CH?C(R1)OR, where R1?=?H and R?Et; R1?=?–(CH2)6CH3, –(CH2)6CH3, –(CH2)8CH3, –(CH2)9CH3, –(CH2)10CH3, –(CH2)12CH3, –(CH2)2Ph], and R?Me). Experimental observations showed that the lipophilic characteristic of 5-hydroxy-5-trifluoromethyl-4,5-dihydro-1H-pyrazoles (5–7) prevent the acid catalyzed dehydration to aromatization of 1H-pyrazole ring, although in some cyclocondensations a proportion of the aromatic derivative 1-(5-trifluoromethyl-3-alkyl-1H-pyrazol-1-yl)alkan-1-one was obtained. All products were characterized using multinuclear (1H, 13C, 19F) NMR spectroscopy.  相似文献   

16.
Bis(fluorbenzoyloxy)methyl phosphane oxides CH3P(O)[OC(O)R]2 [R = C6H42F (1), C6H43F (2), C6H44F (3), C6H32,6F2 (4), C6H2,3,5,6F4 (5)] were prepared by treating silver salts of carboxylic acids AgOC(O)R with CH3P(O)C?2 (IR-, 1H-, 19?F-and 31P{1H}-NMR-data). The mixed anhydrides 1–5 show unusual thermal stability at room temperature. Stability against hydrolysis decreases with increasing number of fluorine-atoms. The reaction of R′P(O)C?2 [R′ = CH3, C6H5, (CH3)3C] with MIOC(O)RF [RF = CF3, C2F5, C6F5; MI = AgI, NaI T?I] was investigated.  相似文献   

17.
Treatment of perfluoro-n-octanonitrile with phenylphosphine gave tetraphenyltetraphosphine and a spectrum of reduction and interaction products. Fifteen compounds were identified. The imine, (RfC7F15) RfCHNH, and the amine, RfCH2NH2, were the primary reduction products. Secondary phosphorus-free products, some formed following ammonia evolution, were the following: RfCHNCH2Rf, RfCH2CH(NH2)Rf, RfC(NH)NCRf(NH2), RfCH2NHCRf(NH), (RfCN)3, RfCHNCRfNCRf(NH), RfCH2NCRfNHCH2Rf, and RfCH2NCRfNHCRf(NH). Only three phosphorus-containing materials were definitely identified: RfCH(NH2)P(C6H5)H, RfCH[P(C6H5)H]NCHRf, and RfC(NH)P(C6H5)CRf(NH). Depending on reaction conditions, specific phosphorus-containing compounds could be preferentially produced. All the structure assignments are based solely on mass spectral breakdown patterns, since pure compounds were not isolated.  相似文献   

18.
The acidity constants of both Z and E conformational isomers of five N-nitroso-N-alkyl-α-amino acids, ON? N(R1)? CH(R2)? COOH, are determined by the observation of selected pH titrated 1H NMR signals. For two glycine derivatives (1, R1?CH3, R2?H, ON? Sar; 2, R1?C2H5, R2?H, ON? EtGly) and two alanine derivatives (3, R1?CH3, R2?CH3, ON? MeAla; 4, R1?C2H5, R2?CH3, ON? EtAla) the E isomers appear to be stronger acids than the Z while for the third alanine derivative (5, R1?n-C3H7, R2?CH3, ON? PrAla) the opposite is observed. These results, also including anisotropy effects associated with the N? NO group, are discussed in terms of conformations. A 7-membered ring conformation with an ? NO…HOOC? intramolecular hydrogen bond is proposed to be statistically important in the Z isomers of 1, 2, 3 and, to a lesser extent, 4.  相似文献   

19.
Inhaltsübersicht. Die Titelverbindungen R2N–CS–S–N[Si(CH3)3]2 mit Ii = CH3 bzw. CH(CH3)2 kristallisieren orthorhombisch bzw. monoklin: Gitterkonstanten für R = CH3 (bei ?165°C) a = 8,397(4) Å, b = 11,917(4) Å, c = 31,966 (11) Å, Pbca (Nr. 61), Z = 8. R = CH(CH3)2 (bei ?80°C) a =13,183(3) Å, b = 10,873(11) Å, c = 14,865(2) Å, β = 105,86(2)° P21/n (Nr. 14), Z = 4. Die Kristallstrukturen wurden unter Verwendung von 4227 bzw. 3 433 symmetrieunabhängigen Reflexen (gemessen bei ?165 bzw. ?80 °C) bestimmt und bis auf Zuverlässigkeitsfaktoren von R = 0,081 bzw. 0,082 verfeinert (Rw = 0,084 bzw. 0,114). Bei beiden Verbindungen ist der C2N–CS–S–N-Teil des Moleküls nahezu planar. Zwischen dem Thiocarbonyl-S-Atom und dem N-Atom der silylierten Aminogruppe bestehen Wechselwirkungen. On Chalcogenolates. 194. S-Bis (trimethylsilyl) amino Esters of Dithiocarbamic Acids. 3. Crystal and Molecular Structure of the Methyl and i-Propyl Derivative The title compounds R2N–CS–S–N[Si(CH3)3]2 with R = CH3 and CH(CH3)2, respectively, crystallize orthorhombic and monoclinic, resp.; cell dimensions and space group see “Inhaltsübersicht”. The structures of both compounds have been determined from single crystal X-ray data measured at ?165°C and ?80°C, resp., and refined to R's of 0.081 and 0.082, resp., (Rw = 0.084 and 0.114, resp.) using 4227 and 3433, resp., independent reflections. In both compounds the C2N–CS–S–N core of the molecule is nearly plane. Between the thiocarbonyl sulfur atom and the nitrogen atom of the amino group interactions exist. In Fortführung unserer Untersuchungen [1, 2] über N, N-Dialkyldithiocarbamidsäure-S-bis(trimethylsilyl)aminoester R2N–CS–S–N[Si(CH3)3]2 haben wir die Kristall- und Molekülstrukturen der Verbindungen mit R = CH3 und CH(CH3)2 bestimmt. Dabei sollte untersucht werden, welchen Einfluß sterisch anspruchsvollere Alkylgruppen (R = CH3 → CH(CH3)2) auf die Molekülgeo-metrie haben. Eine strukturchemische Charakterisierung dieser Verbindungs-klasse ist bis jetzt noch nicht erfolgt; vgl. die Literaturzusammenstellung bei [3].  相似文献   

20.
Reaction of phosphorus trichloride with tert-butanol and fluoroalcohols gave bis(fluoroalkyl) phosphites (RFO)2P(O)H in 42-89% yield, where RF=HCF2CH2, H(CF2)2CH2, H(CF2)4CH2, CF3CH2, C2F5CH2, C3F7CH2, (CF3)2CH, (FCH2)2CH, CF3(CH3)2C, (CF3)2CH3C, CF3CH2CH2, C4F9CH2CH2 and C6F13CH2CH2. Treatment of these with chlorine in dichloromethane gave the bis(fluoroalkyl) phosphorochloridates (RFO)2P(O)Cl in 49-96% yield. The chloridate (CF3CH2O)2P(O)Cl was isolated in much lower yield from the interaction of thionyl chloride with bis(trifluoroethyl) phosphite. Heating the latter in dichloromethane with potassium fluoride and a catalytic amount of trifluoroacetic acid gave the corresponding fluoridate (CF3CH2O)2P(O)F in 84% yield. Treatment of bis(trifluoroethyl) phosphite with bromine or iodine gave the bromidate (CF3CH2O)2P(O)Br and iodidate (CF3CH2O)2P(O)I in 51 and 46% yield, respectively. The iodidate is the first dialkyl phosphoroiodidate to have been isolated and characterised properly—its discovery lags behind the first isolation of a dialkyl phosphorochloridate by over 130 years. The fluoroalkyl phosphoryl compounds are generally more stable than known unfluorinated counterparts.  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