首页 | 官方网站   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.
Electron impact fragmentation patterns were obtained for 1-dephenylphospha-3,5-bis(perfluoro-n-heptyl)-2,4,6-triazine, 1-diphenylphospha-3,5-bis[C3F7(OCF(CF3)CF2)xOCF(CF3]-2,4,6-triazine (x=1 and 2) and their pentafluorophenyl analogues. In each instance the ion R2PN2C+ (RC6H5 or C6F5) constituted the base peak. Based on the observed metastable peaks, fragmentation pathways leading to the formation of this and other significant ions are postulated and similarities to s-triazine and diphenyl phosphazene trimer breakdown patterns are discussed.  相似文献   

2.
A series of monophospha-s-triazines, namely l-diphenylphospha-3,5-bis(perfluoro-n-heptyl)-2,4,6-triazine, 1-diphenylphospha-3,5-bis[C3F7-[OCF(CF3)CF2]xOCF(CF3)]-2,4,6-triazine (x = 1 and 2), and the penta-fluorophenyl-substituted analogues were prepared, in yields of 50-75%, from the respective imidoylamidines and trichlorophosphoranes. The physical properties of the corresponding phenyl and pentafluorophenyl monophospha-s-triazines did not differ significantly; the perfluoroalkyl-substituted materials were low melting solids whereas the perfluoroalkyl-ether-containing compositions were liquids. Preliminary degradative studies showed these compounds to be thermally and oxidatively stable. The l-diphenylphospha-3,5-bis[C3F7OCF(CF3)CF2OCF(CF3)]-2,4,6-triazine was found to be an effective anti-corrosion additive for perfluoroalkylether type fluids.  相似文献   

3.
The push–pull character of a series of donor–bithienyl–acceptor compounds has been tuned by adopting triphenylamine or 1,1,7,7‐tetramethyljulolidine as a donor and B(2,6‐Me2‐4‐RC6H2)2 (R=Me, C6F5 or 3,5‐(CF3)2C6H3) or B[2,4,6‐(CF3)3C6H2]2 as an acceptor. Ir‐catalyzed C?H borylation was utilized in the derivatization of the boryl acceptors and the tetramethyljulolidine donor. The donor and acceptor strengths were evaluated by electrochemical and photophysical measurements. In solution, the compound with the strongest acceptor, B[2,4,6‐(CF3)3C6H2]2 ((FMes)2B), has strongly quenched emission, while all other compounds show efficient green to red (ΦF=0.80–1.00) or near‐IR (NIR; ΦF=0.27–0.48) emission, depending on solvent. Notably, this study presents the first examples of efficient NIR emission from three‐coordinate boron compounds. Efficient solid‐state red emission was observed for some derivatives, and interesting aggregation‐induced emission of the (FMes)2B‐containing compound was studied. Moreover, each compound showed a strong and clearly visible response to fluoride addition, with either a large emission‐color change or turn‐on fluorescence.  相似文献   

4.
Kinetically stabilized congeners of carbenes, R2C, possessing six valence electrons (four bonding electrons and two non‐bonding electrons) have been restricted to Group 14 elements, R2E (E=Si, Ge, Sn, Pb; R=alkyl or aryl) whereas isoelectronic Group 15 cations, divalent species of type [R2E]+ (E=P, As, Sb, Bi; R=alkyl or aryl), were unknown. Herein, we report the first two examples, namely the bismuthenium ion [(2,6‐Mes2C6H3)2Bi][BArF4] ( 1 ; Mes=2,4,6‐Me3C6H2, ArF=3,5‐(CF3)2C6H3) and the stibenium ion [(2,6‐Mes2C6H3)2Sb][B(C6F5)4] ( 2 ), which were obtained by using a combination of bulky meta‐terphenyl substituents and weakly coordinating anions.  相似文献   

5.
Reductive late-stage functionalization of gibberellic acid is reported using three fluoroarylborane Lewis acids; (B(C6F5)3, B(3,5-C6H3(CF3)2), and B(2,4,6-C6H2F3)3) in combination with a tertiary silane and a borane (HBCat) reductant. In each case, C–O bond activation occurs, and different products are obtained depending on the reductant and catalyst employed.  相似文献   

6.
Hitherto unknown 2,4,6-tris(trifluoromethyl)benzyl alcohol ( 3 ) was synthesized in 41 % yield by treatment of freshly prepared RFLi ( 2 ) with paraformaldehyde (RF = 2,4,6-tris(trifluoromethyl)phenyl). According to an X-ray diffraction study the crystal structure of 3 consists of S6 symmetric cyclic hexamers [2,4,6-(CF3)3C6H2CH2OH]6. Deprotonation of 3 with NaN(SiMe3)2 in toluene afforded the unsolvated sodium alkoxide derivative RFCH2ONa ( 4 ). Homoleptic lanthanide alkoxides of the type Ln(OCH2RF)3 (Ln = Nd ( 5 ), Sm ( 6 ), Yb ( 7 )) were made by treatment of Ln(C5H5)3 with three equivalents of 3 . Similar reactions in a 1:1 molar ratio afforded the bis(cyclopentadienyl)lanthanide alkoxide derivatives (C5H5)2Ln(OCH2RF) (Ln = Nd ( 8 ), Sm ( 9 ), Yb ( 10 )).  相似文献   

7.
The lithium complexes [(WCA-NHC)Li(toluene)] of anionic N-heterocyclic carbenes with a weakly coordinating anionic borate moiety (WCA-NHC) reacted with iodine, bromine, or CCl4 to afford the zwitterionic 2-halogenoimidazolium borates (WCA-NHC)X (X=I, Br, Cl; WCA=B(C6F5)3, B{3,5-C6H3(CF3)2}3; NHC=IDipp=1,3-bis(2,6-diisopropylphenyl)imidazolin-2-ylidene, or NHC=IMes=1,3-bis(2,4,6-trimethylphenyl)imidazolin-2-ylidene). The iodine derivative (WCA-IDipp)I (WCA=B(C6F5)3) formed several complexes of the type (WCA-IDipp)I ⋅ L (L=C6H5Cl, C6H5Me, CH3CN, THF, ONMe3), revealing its ability to act as an efficient halogen bond donor, which was also exploited for the preparation of hypervalent bis(carbene)iodine(I) complexes of the type [(WCA-IDipp)I(NHC)] and [PPh4][(WCA-IDipp)I(WCA-NHC)] (NHC=IDipp, IMes). The corresponding bromine complex [PPh4][(WCA-IDipp)2Br] was isolated as a rare example of a hypervalent (10-Br-2) system. DFT calculations reveal that London dispersion contributes significantly to the stability of the bis(carbene)halogen(I) complexes, and the bonding was further analyzed by quantum theory of atoms in molecules (QTAIM) analysis.  相似文献   

8.
The title compound {2‐[3,5‐bis(trifluoromethyl)‐1H‐pyrazol‐1‐ylmethyl]‐6‐(3,5‐dimethyl‐1H‐pyrazol‐1‐ylmethyl)pyridine}methylpalladium(II) tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate, [Pd(C18H18F6N5)][B(C8H3F6)4], crystallizes as discrete cations and anions. The cation possesses a pseudo‐twofold axis about which positional disorder of the tridentate ligand is exhibited. The four substituents on the two pyrazole rings exhibit CH3/CF3 disorder, while all other atoms are ordered. Thus, this disorder can be conveniently described `locally' as compositional, while `globally' for the entire tridentate ligand it is positional. The anion also exhibits typical rotational positional disorder in three of the CF3 groups. All disordered CF3 groups were modeled with idealized C3v geometry.  相似文献   

9.
Trimethylamine‐bis(trifluoromethyl)boranes R(CF3)2B · NMe3 (R = cis/trans‐CF3CF=CF ( 1/2 ), HC≡C ( 3 ), H2C=CH ( 4 ), C2H5 ( 5 ), C6H5CH2 ( 6 ), C6F5 ( 7 ), C6H5 ( 8 )) react with NEt3 × 3 HF depending on the nature of R at 155–200 °C under replacement of the trimethylamine ligand to form the corresponding fluoro‐bis(trifluoromethyl)borates [R(CF3)2BF] ( 1 a/2 a – 8 a ). The structures of 7 , K[C6H5CH2(CF3)2BF] ( K‐6 a ), and K[C6H5(CF3)2BF] ( K‐8 a ) have been investigated by single‐crystal X‐ray diffraction. In 7 the CF3 groups make short repulsive contacts with NMe3 and C6F5 entities – the B–CF3 bonds being unusually long. The B–F bond lengths of K‐6 a and K‐8 a (1.446(3) and 1.452(2) Å, respectively) are long for a fluoroborate.  相似文献   

10.
The bis(silyl)triazene compound 2,6‐(Me3Si)2‐4‐Me‐1‐(N?N? NC4H8)C6H2 ( 4 ) was synthesized by double lithiation/silylation of 2,6‐Br2‐4‐Me‐1‐(N?N? NC4H8)C6H2 ( 1 ). Furthermore, 2,6‐bis[3,5‐(CF3)2‐C6H3]‐4‐Me‐C6H2‐1‐(N?N? NC4H8)C6H2 derivative 6 can be easily synthesized by a C,C‐bond formation reaction of 1 with the corresponding aryl‐Grignard reagent, i.e., 3,5‐bis[(trifluoromethyl)phenyl]magnesium bromide. Reactions of compound 4 with KI and 6 with I2 afforded in good yields novel phenyl derivatives, 2,6‐(Me3Si)2‐4‐MeC6H2? I and 2,6‐bis[3,5‐(CF3)2? C6H3]‐4‐MeC6H2? I ( 5 and 7 , resp.). On the other hand, the analogous m‐terphenyl 1,3‐diphenylbenzene compound 2,6‐bis[3,5‐(CF3)2? C6H3]C6H3? I ( 8 ) could be obtained in moderate yield from the reaction of (2,6‐dichlorophenyl)lithium and 2 equiv. of aryl‐Grignard reagent, followed by the reaction with I2. Different attempts to introduce the tBu (Me3C) or neophyl (PhC(Me)2CH2) substituents in the central ring were unsuccessful. All the compounds were fully characterized by elemental analysis, melting point, IR and NMR spectroscopy. The structure of compound 6 was corroborated by single‐crystal X‐ray diffraction measurements.  相似文献   

11.
Reactions of CrCl3(thf)3 with bis(imino)pyridines gave a series of {bis(imino)pyridine}chromium(III) trichloride complexes, {2,6‐(RN?CMe)2C5H3N}CrCl3 [R = C6HPr2‐2,6 ( 1 ), C6H3Et2‐2,6 ( 2 ), C6H3Me2‐2,6 ( 3 ), C6H2Me3‐2,4,6 ( 4 ), C6H3Me2‐3,5 ( 5 ), C6H5 ( 6 ), cyclohexyl ( 7 ), 2‐methyl‐1‐naphthyl ( 8 ), C6H3F2‐2,6 ( 9 ), C6H3Br2‐2,6 ( 10 ), C6F5 ( 11 )]. Pseudo‐octahedral geometries of 6 , 10 , and 11 were revealed by X‐ray crystallography. The complexes having bulky substituents such as 1 – 4 showed high activity for ethylene polymerization in combination with modified methylaluminoxane (MMAO) to give linear polyethylenes. In sharp contrast, the pentafluorophenyl complex 11 /modified methylaluminoxane system was found to be moderately active for ethylene homopolymerization to give moderately branched polyethylene with only ethyl branches. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 3368–3375, 2005  相似文献   

12.
The sodium salt (CF3)2 NONa+ (I) [from (CF3)2NOH + NaH in Et2O], is an alternative bis(trifluoromethyl)amino-oxylating agent to the adduct (CF3)2NOH.CsF (III). With pentafluoropyridine it affords 4-X.C5F4N (II) + 2,4-X2.C5F3N (IV), [X = (CF3)2NO]. It has been used to obtain a number of new bis(trifluoromethyl)amino-oxy-compounds; i.e. the following conversions have been effected: perfluoro-(4-isopropylpyridine)→ 2-X.C5F3N.CF(CF3)2-4 (V) + 2,6-X2.C5F2N.CF(CF3)2-4 (VI); 3-chlorotetrafluoropyridine → 4-X.C5F3N.Cl-3 (VII) and 2-X.C5F3N.Cl-5 (VIII) (not separated) + 2,4-X2.C5F2N.Cl-5 (IX), 3,5-dichlorotrifluoropyridine → 2- (XI) and 4-X.C5F2N.Cl2-3,5 (X) (not separated) + 2,4-X2.C5FN.Cl2-3,5 (XII); and perfluorotoluene → 4-X.C6F4.CF3-1 (XIII). Hexafluorobenzene resisted attack by (CF3)2NONa under the conditions used with these aromatic substrates (ca 20 °C). Static pyrolysis (125 °C) of 4-[bis(trifluoromethyl)amino-oxy]tetrafluoropyridine (II) gave a mixture of 6-bis(trifluoromethyl)amino]tetrafluoro-4-azacyclohexa-2, 4-dienone (XV) and 4-[bis(trifluoromethyl)amino]tetrafluoro-4-azacyclohexa-2,5-dienone (XVI). The 13C chemical shifts, assigned by analysis of 19F-coupled and 19F broad-band decoupled 13C n.m.r. spectra, are in accord with a +M effect similar to that of fluorine for a (CF3)2NO- substituent in the 2- and 4- positions of a polyfluoropyridine and a slightly smaller -I effect; the steric effect of (CF3)2NO on the shifts is less than that of chlorine. In contrast, a ring carbon carrying a (CF3)2CF- substituent is markedly shielded compared with one carrying fluorine, presumably by a steric effect.  相似文献   

13.
The frustrated Lewis pair (FLP)‐catalyzed hydrogenation and deuteration of N‐benzylidene‐tert‐butylamine ( 2 ) was kinetically investigated by using the three boranes B(C6F5)3 ( 1 ), B(2,4,6‐F3‐C6H2)3 ( 4 ), and B(2,6‐F2‐C6H3)3 ( 5 ) and the free activation energies for the H2 activation by FLP were determined. Reactions catalyzed by the weaker Lewis acids 4 and 5 displayed autoinductive catalysis arising from a higher free activation energy (2 kcal mol?1) for the H2 activation by the imine compared to the amine. Surprisingly, the imine reduction using D2 proceeded with higher rates. This phenomenon is unprecedented for FLP and resulted from a primary inverse equilibrium isotope effect.  相似文献   

14.
We describe the reactivity of two linkage isomers of a boryl-phosphaethynolate, [B]OCP and [B]PCO (where [B]=N,N’-bis(2,6-diisopropylphenyl)-2,3-dihydro-1H-1,3,2-diazaboryl), towards tris- (pentafluorophenyl)borane (BCF). These reactions afforded three constitutional isomers all of which contain a phosphaalkene core. [B]OCP reacts with BCF through a 1,2 carboboration reaction to afford a novel phosphaalkene, E-[B]O{(C6F5)2B}C=P(C6F5), which subsequently undergoes a rearrangement process involving migration of both the boryloxy and pentafluorophenyl substituents to afford Z-{(C6F5)2B}(C6F5)C=PO[B]. By contrast, [B]PCO undergoes a 1,3-carboboration process accompanied by migration of the N,N’-bis(2,6-diisopropylphenyl)-2,3-dihydro-1H-1,3,2-diazaboryl to the carbon centre.  相似文献   

15.
The reaction of xenonbis(trifluoroacetate) and trifluoromethanesulfonic acid (triflic acid) gave the new, highly reactive unsymmetrical xenon-oxo species CF3COOXeOSO2CF3. Benzene derivates, containing electron withdrawing substituents such as -F, -CF3, -Cl or -NO2 were electrophilic attacked by this intermediate to yield arylxenon trifluoromethanesulfonates. Via this one-pot synthesis trifluoromethanesulfonates with the cations [Xe(2,4,6-F3C6H2)]+, [Xe(2-F-5-NO2C6H3)]+, [Xe(2-F-5-CF3C6H3)]+ and [Xe(3,5-(CF3)2C6H3)]+ were prepared. All compounds were characterized by their NMR, mass, and vibrational spectra. Additionally, several new arylxenon trifluoromethanesulfonates were detected by 129Xe-NMR spectroscopy as products of the reaction of 1,3-F2C6H4 and further deactivated benzenes with xenontrifluoroacetate trifluoromethane sulfonate. Fluoro substituents in ortho position to xenon significantly increase the thermal stability of the arylxenon trifluoromethanesulfonates obtained. The molecular structure of [Xe(2,6-F2C6H3)][OSO2CF3] was determined by single crystal diffraction methods. The arylxenon unit is weakly coordinated by one oxygen atom of the CF3SO3 anion. The salt crystallizes in the triclinic space group P1 , a = 880.9(3) pm, b = 1093.9(5) pm, c = 1209.8(5) pm, α = 89.04(4)°, β = 74.23(3)°, γ = 86.03(3)°, Z = 4.  相似文献   

16.
The molecules ArFXeF (ArF=C6F5, 2,4,6-C6H2F3) with a more polar Xe-F bond than XeF2 are versatile starting materials for substitution reactions. Fluorine-aryl substitutions with Cd(ArF)2, C6F5SiMe3/[F], and C6F5SiF3 formed symmetric and/or asymmetric diarylxenon compounds. Applying C6F5BF2, with a higher F-affinity than the corresponding aryltrifluorosilane, in contrast gave the salt [RXe] [ArFBF3]. Using the alkenyl and alkyl compounds CF2=CFSiMe3/[F], CF3SiMe3/[F], and Cd(CF3)2 in reactions with C6F5XeF, the perfluoroalkenyl or -alkyl transfer reagents were consumed without observing C6F5XeCF=CF2 or C6F5XeCF3 but the formation of Xe(C6F5)2 (dismutation product) and in the latter case C6F5CF3 (coupling product), gave hints of the desired intermediates.  相似文献   

17.
Herein, we present the formation of transient radical ion pairs (RIPs) by single-electron transfer (SET) in phosphine−quinone systems and explore their potential for the activation of C−H bonds. PMes3 (Mes=2,4,6-Me3C6H2) reacts with DDQ (2,3-dichloro-5,6-dicyano-1,4-benzoquinone) with formation of the P−O bonded zwitterionic adduct Mes3P−DDQ ( 1 ), while the reaction with the sterically more crowded PTip3 (Tip=2,4,6-iPr3C6H2) afforded C−H bond activation product Tip2P(H)(2-[CMe2(DDQ)]-4,6-iPr2-C6H2) ( 2 ). UV/Vis and EPR spectroscopic studies showed that the latter reaction proceeds via initial SET, forming RIP [PTip3]⋅+[DDQ]⋅, and subsequent homolytic C−H bond activation, which was supported by DFT calculations. The isolation of analogous products, Tip2P(H)(2-[CMe2{TCQ−B(C6F5)3}]-4,6-iPr2-C6H2) ( 4 , TCQ=tetrachloro-1,4-benzoquinone) and Tip2P(H)(2-[CMe2{oQtBu−B(C6F5)3}]-4,6-iPr2-C6H2) ( 8 , oQtBu=3,5-di-tert-butyl-1,2-benzoquinone), from reactions of PTip3 with Lewis-acid activated quinones, TCQ−B(C6F5)3 and oQtBu−B(C6F5)3, respectively, further supports the proposed radical mechanism. As such, this study presents key mechanistic insights into the homolytic C−H bond activation by the synergistic action of radical ion pairs.  相似文献   

18.
The reactions of the Mannich reagent Et3SiOCH2NMe2 ( 1 ) with a variety of anilines (mono-substituted RC6H4NH2, R=H, 4-CN, 4-NO2, 4-Ph, 4-Me, 4-MeO, 4-Me2N; di-substituted R2C6H3NH2, R2=3,5-(CH3)2, 3,5-(CF3)2; tri-substituted R3C6H2NH2, R3=3,5-Me2-4-Br and a “super bulky” aniline (Ar*NH2) [Ar*=2,6-bis(diphenylmethyl)-4-tert-butylphenyl]) led to the formation of a range of products dependent upon the substituent. With electron-withdrawing substituents, previously unknown diamines, RC6H4NH(CH2NMe2) [R=CN ( 2 a ), NO2 ( 2 b )] and R2C6H3NH(CH2NMe2) [R2=3,5-(CF3)2 ( 2 c) ] were formed. Further reaction of 2 a , b , c with 1 yielded the corresponding triamines RC6H4N(CH2NMe2)2 (R=CN ( 3 a ), NO2 ( 3 b ) and R2C6H3N(CH2NMe2)2, R2=3,5-(CF3)2 ( 3 c ). The new polyamines were characterized by NMR spectroscopy, and for 2 a , 2 c , and 3 c , by single crystal XRD. In the case of electron-donating groups, R=4-OMe, 4-NMe2, 4-Me, 3,5-Me2, 3,5-Me2-4-Br, and for R=4-Ph, the reactions with 1 immediately led to the formation of the related 1,3,5-triazines, R=4-MeO ( 5 a ), 4-Me2N ( 5 b ), 4-Me ( 5 c ), 3,5-Me2 ( 5 d ), 3,5-Me2-4-Br ( 5 e ), 4-Ph ( 5 f ), 4-Cl ( 5 g ). The “super bulky” aniline rapidly produced a single product, namely the corresponding imine Ar*N=CH2 ( 4 ) which was also characterized by single crystal XRD. Imine 4 is both thermally and oxidatively stable. All reactions are very fast, thus based upon the presence of Si we are tempted to denote the reactions of 1 as examples of “Silick” chemistry.  相似文献   

19.
The synthesis and full characterization of α-silylated (α-SiCPs; 1 – 7 ) and α-germylated (α-GeCPs; 11 – 13 ) phosphorus ylides bearing one chloride substituent R3PC(R1)E(Cl)R22 (R=Ph; R1=Me, Et, Ph; R2=Me, Et, iPr, Mes; E=Si, Ge) is presented. The molecular structures were determined by X-ray diffraction studies. The title compounds were applied in halide abstraction studies in order to access cationic species. The reaction of Ph3PC(Me)Si(Cl)Me2 ( 1 ) with Na[B(C6F5)4] furnished the dimeric phosphonium-like dication [Ph3PC(Me)SiMe2]2[B(C6F5)4]2 ( 8 ). The highly reactive, mesityl- or iPr-substituted cationic species [Ph3PC(Me)SiMes2][B(C6F5)4] ( 9 ) and [Ph3PC(Et)SiiPr2][B(C6F5)4] ( 10 ) could be characterized by NMR spectroscopy. Carrying out the halide abstraction reaction in the sterically demanding ether iPr2O afforded the protonated α-SiCP [Ph3PCH(Et)Si(Cl)iPr2][B(C6F5)4] ( 6 dec ) by sodium-mediated basic ether decomposition, whereas successfully synthesized [Ph3PC(Et)SiiPr2][B(C6F5)4] ( 10 ) readily cleaves the F−C bond in fluorobenzene. Thus, the ambiphilic character of α-SiCPs is clearly demonstrated. The less reactive germanium analogue [Ph3PC(Me)GeMes2][B{3,5-(CF3)2C6H3}4] ( 14 ) was obtained by treating 11 with Na[B{3,5-(CF3)2C6H3}4] and fully characterized including by X-ray diffraction analysis. Structural parameters indicate a strong CYlide−Ge interaction with high double bond character, and consequently the C−E (E=Si, Ge) bonds in 9 , 10 and 14 were analyzed with NBO and AIM methods.  相似文献   

20.
Arylvanadium(III) Compounds. III. Preparation and Properties of Triaryl Vanadium Complexes The synthesis of triarylvanadium compounds, VR3 (R = C6H5; 2,6-(CH3)2C6H3; 2,4,6-(CH3)3C6H2; (CH3)5C6) is investigated. Only the compounds V[2,6-(CH3)2C6H3]3 and V[2,4,6-(CH3)3C6H2]3 (crystallized with tetrahydrofuran) are obtained. The complexes V(C6H5)(dipy)2 · THF, V[2,4,6-(CH3)3C6H2]3 · Do (Do = py, dipy) are described too.  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-23

京公网安备 11010802026262号