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.
High‐temperature gas‐phase, solvent‐ and catalyst‐free reaction of naphthalene with an excess of RFI reagent (RF?CF3, C2F5, n‐C3F7, and n‐C4F9) was used for the first time to produce a series of highly perfluoroalkylated naphthalene products NAPH(RF)n with n=2–5. Four 95+ % pure 1,3,5,7‐NAPH(RF)4 with RF?CF3, C2F5, n‐C3F7, and n‐C4F9 were isolated using a simple chromatography‐free procedure. These new compounds were fully characterized by 19F and 1H NMR spectroscopy, X‐ray crystallography (for RF?CF3 and C2F5), atmospheric‐pressure chemical ionization mass spectrometry, and cyclic and square‐wave voltammetry. DFT calculations confirm that the proposed synthesis yields the most stable isomers that have not been accessed by alternative preparation techniques.  相似文献   

2.
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   

3.
Reported herein is the use of S‐perfluoroalkyl sulfilimino iminiums as a new source of RF radicals under visible‐light photoredox catalysis (RF=CF3, C4F9, CF2Br, CFCl2). These shelf‐stable perfluoroalkyl reagents, readily prepared on gram scale from the corresponding sulfoxide using a one‐pot procedure, allow the efficient photoredox‐induced oxyperfluoroalkylation of various alkenes using fac‐Ir(ppy)3 as the photocatalyst. Importantly, spin‐trapping/electron paramagnetic resonance experiments were carried out to characterize all the radical intermediates involved in this radical/cationic process.  相似文献   

4.
An unprecedented silver‐catalyzed formal insertion of arynes into Rf?I (Rf=CF3, C2F5) bonds has been developed. This protocol provides easy access to various ortho‐perfluoroalkyl iodoarenes under mild conditions. In this insertion reaction, an ionic atom‐transfer reaction of RfI occurs, and a silver‐mediated metathesis process is involved in the efficient transfer of the electropositive iodine atom.  相似文献   

5.
A convenient synthesis of the 1,3‐dihydro‐1,3,3‐tris(perfluoroalkyl)isobenzofuran‐1‐ols 3a , b was elaborated starting from commercially available phthaloyl dichloride and trimethyl(perfluoroalkyl)silanes (Me3SiRf) 1a , b (Rf=CF3, C2F5) in the presence of a fluoride source (Schemes 1 and 3). In a reaction analogous to alkyl Grignard reagents, double chloride substitution by two perfluoroalkyl groups and subsequent addition of one perfluoroalkyl group with concomitant ring closure led to this new class of compounds (Scheme 2). The syntheses of the alcohols and some alcoholates, as well as of the corresponding trimethylsilyl ethers are described. A combination of special 1D and 2D NMR experiments allowed the assignment of all atoms of the new compounds. The solid‐state structure of 1,3‐dihydro‐1,3,3‐tris(trifluoromethyl)isobenzofuran‐1‐ol ( 3a ) was elucidated by X‐ray diffraction methods.  相似文献   

6.
Perfluoroalkanesulfonyl chlorides [RFSO2Cl ; RF  CF3, C2F3, C4F9], decompose thermally to give the corresponding perfluoroalkyl chlorides with evolution of SO2. The latter retards the reaction, but it is catalysed by copper which also inhibits the SO2 effect. 2-methyl-2-nitrosopropane traps the perfluoroalkyl free radicals. In the presence of a perfluoroalkyl iodide [R′FI ≠ R′F≠RF], other products, RFI and RFCl, are obtained. A free radical chain-mechanism is then suggested.On the other hand, perfluorobutanesulfonyl fluoride is very stable thermally.  相似文献   

7.
A spectroelectrochemical study of the two isostructural asymmetric perfluoroalkyl derivatives C1‐7,24‐C70(CF3)2 and C1‐7,24‐C70(C2F5)2 is presented. Reversible formation of their stable monoanion radicals is monitored by cyclic voltammetry and by in situ ESR‐Vis‐NIR spectroelectrochemistry. The ESR spectrum of the C70(CF3)2?. radical is a 1:3:3:1 quartet with a 19F hyperfine coupling constant (a(F)) of 0.323(4) G, demonstrating that the unpaired spin is coupled to only one of the two CF3 groups. The 13C satellites are assigned to specific carbon atoms. The ESR spectrum of the C70(C2F5)2?. radical is an apparent octet with an apparent a(F) value of 0.83(2) G. DFT calculations suggest that this pattern is due to the superposition of spectra for four nearly isoenergetic C70(C2F5)2?. conformers. Time‐dependent DFT calculations suggest that the NIR band at 1090 nm exhibited by both C70(Rf)2?. radical anions is assigned to the SOMO→LUMO+3 transition. The analogous NIR band exhibited by the closed‐shell C70(CF3)22? dianion was blue‐shifted to 1000 nm.  相似文献   

8.
Crystals of the title compound, C4H8N5+·C2F3O2, are built up of singly protonated 2,4‐diamino‐6‐methyl‐1,3,5‐triazin‐1‐ium cations and trifluoroacetate anions. The CF3 group of the anion is disordered. The oppositely charged ions interact via almost linear N—H...O hydrogen bonds, forming a CF3COO...C4H8N5+ unit. Two units related by an inversion centre interact through a pair of N—H...N hydrogen bonds, forming planar (CF3COO...C4H8N5+...C4H8N5+·CF3COO) aggregates that are linked by a pair of N—H...O hydrogen bonds into chains running along the c axis.  相似文献   

9.
In reactions with perfluoroalkylsulfenyl chlorides (RfSCl; Rf = F3, C2F5, n-C3F7, n-C4F9) and perfluoroalkyl disulfides (RfSSRf′; Rf = Rf′ = CF3, Rf = CF3, Rf′ = C2F5) at 25°, chlorine monofluoride acts primarily as a chlorinating and fluorinating reagent to give the corresponding perfluoroalkylsulfur chloride tetrafluorides, RfSF4Cl, in good yields. However, small amounts of perfluoroalkylsulfur pentafluorides, RfSF5, are also obtained. A mixture of the cis and trans isomers of bis(trifluoromethyl)sulfur tetrafluoride and of trifluoromethyl pentafluoroethylsulfur tetrafluoride has been formed by the reaction of the corresponding bis(perfluoroalkyl) sulfides and chlorine monofluoride. The new perfluoroalkylsulfur chloride tetrafluorides are colorless, unpleasant smelling liquids. The infrared, mass and 19F NMR spectral data, as well as thermodynamic and elementary analysis data, are given for the new compounds.  相似文献   

10.
A series of heteroligated (salicylaldiminato)(β‐enaminoketonato)titanium complexes [3‐But‐2‐OC6H3CH = N(C6F5)] [PhN = C(R1)CHC(R2)O]TiCl2 [ 3a : R1 = CF3, R2 = tBu; 3b : R1 = Me, R2 = CF3; 3c : R1 = CF3, R2 = Ph; 3d : R1 = CF3, R2 = C6H4Ph(p ); 3e : R1 = CF3, R2 = C6H4Ph(o ); 3f : R = CF3, R2 = C6H4Cl(p ); 3g : R1 = CF3; R2 = C6H3Cl2(2,5); 3h : R1 = CF3, R2 = C6H4Me(p )] were investigated as catalysts for ethylene (co)polymerization. In the presence of modified methylaluminoxane as a cocatalyst, these complexes showed activities about 50%–1000% and 10%–100% higher than their corresponding bis(β‐enaminoketonato) titanium complexes for ethylene homo‐ and ethylene/1‐hexene copolymerization, respectively. They produced high or moderate molecular weight copolymers with 1‐hexene incorporations about 10%–200% higher than their homoligated counterpart pentafluorinated FI‐Ti complex. Among them, complex 3b displayed the highest activity [2.06 × 106 g/molTi?h], affording copolymers with the highest 1‐hexene incorporations of 34.8 mol% under mild conditions. Moreover, catalyst 3h with electron‐donating group not only exhibited much higher 1‐hexene incorporations (9.0 mol% vs. 3.2 mol%) than pentafluorinated FI‐Ti complex but also generated copolymers with similar narrow molecular weight distributions (M w/M n = 1.20–1.26). When the 1‐hexene concentration in the feed was about 2.0 mol/L and the hexene incorporation of resultant polymer was about 9.0 mol%, a quasi‐living copolymerization behavior could be achieved. 1H and 13C NMR spectroscopic analysis of their resulting copolymers demonstrated the possible copolymerization mechanism, which was related with the chain initiation, monomer insertion style, chain transfer and termination during the polymerization process. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 2787–2797  相似文献   

11.
Perfluoroalkylation of a higher fullerene mixture with CF3I or C2F5I, followed by HPLC separation of CF3 and C2F5 derivatives, resulted in the isolation of several C84(RF)n (n=12, 16) compounds. Single‐crystal X‐ray crystallography with the use of synchrotron radiation allowed structure elucidation of eight C84(RF)n compounds containing six different C84 cages (the number of the C84 isomer is given in parentheses): C84 (23)(C2F5)12 ( I ), C84 (22)(CF3)16 ( II ), C84 (22)(C2F5)12 ( III ), C84 (11)(C2F5)12 ( IV ), C84 (16)(C2F5)12 ( V ), C84 (4)(CF3)12 ( VI with toluene and VII with hexane as solvate molecules), and C84 (18)(C2F5)12 ( VIII ). Whereas some connectivity patterns of C84 isomers (22, 23, 11) had previously been unambiguously confirmed by different methods, derivatives of C84 isomers numbers 4, 16, and 18 have been investigated crystallographically for the first time, thus providing direct proof of the connectivity patterns of rare C84 isomers. General aspects of the addition of RF groups to C84 cages are discussed in terms of the preferred positions in the pentagons under the formation of chains, pairs, and isolated RF groups.  相似文献   

12.
Syntheses and Properties of Perfluoroorgano Esters of the Diethyldithiocarbamic Acid, (C2H5)2NC(S)SRf (Rf = CF3, C2F5, i‐C3F7, n‐C4F9, C6F5) Tetraethylthiuram disulfide reacts under different conditions with perfluoroorgano silver(I), AgRf, and perfluoroorgano cadmium compounds, Cd(Rf)2, to give the corresponding perfluoroorgano esters of diethyldithiocarbamic acid, (C2H5)2NC(S)SRf (Rf = CF3, C2F5, i‐C3F7, n‐C4F9, C6F5), and metal diethyldithiocarbamates, AgSC(S)N(C2H5)2 and Cd[SC(S)N(C2H5)2]2. The mechanisms of the reactions with AgRf and Cd(Rf)2 are discussed.  相似文献   

13.
Perfluoroalkanesulfonic anhydrides [(RFSO2)2O; RF=CF3,C2F5,C4F9], mixed with the parent acid, decompose thermally to give the corresponding perfluoroalkyl perfluoroalkanesulfonates (RFSO3RF) with liberation of SO2. If the perfluoroalkyl moieties in the anhydride and the acid are different, a mixture of symmetric and unsymmetric esters is obtained. An ionic bimolecular mechanism is deduced from the results, and a new easy synthesis of the symmetric perfluorinated sulfonic esters is proposed.  相似文献   

14.
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   

15.
The salts 3‐[(2,2,3,3‐tetrafluoropropoxy)methyl]pyridinium saccharinate, C9H10F4NO+·C7H4NO3S, (1), and 3‐[(2,2,3,3,3‐pentafluoropropoxy)methyl]pyridinium saccharinate, C9H9F5NO+·C7H4NO3S, (2), i.e. saccharinate (or 1,1‐dioxo‐1λ6,2‐benzothiazol‐3‐olate) salts of pyridinium with –CH2OCH2CF2CF2H and –CH2OCH2CF2CF3meta substituents, respectively, were investigated crystallographically in order to compare their fluorine‐related weak interactions in the solid state. Both salts demonstrate a stable synthon formed by the pyridinium cation and the saccharinate anion, in which a seven‐membered ring reveals a double hydrogen‐bonding pattern. The twist between the pyridinium plane and the saccharinate plane in (2) is 21.26 (8)° and that in (1) is 8.03 (6)°. Both salts also show stacks of alternating cation–anion π‐interactions. The layer distances, calculated from the centroid of the saccharinate plane to the neighbouring pyridinium planes, above and below, are 3.406 (2) and 3.517 (2) Å in (1), and 3.409 (3) and 3.458 (3) Å in (2).  相似文献   

16.
Nine organotin fluorocarboxylates RnSnO2CRf (n = 3, R = Bu, Rf = CF3, C2F5, C3F7, C7F15; R = Et, Rf = CF3, C2F5; R = Me, Rf = C2F5; n = 2, R = Me, Rf = CF3) have been synthesized; key examples have been used to deposit fluorine‐doped SnO2 thin films by atmospheric pressure chemical vapour deposition. Et3SnO2CC2F5, in particular, gives high‐quality films with fast deposition rates despite adopting a polymeric, carboxylate‐bridged structure in the solid state, as determined by X‐ray crystallography. Gas‐phase electron diffraction on the model compound Me3SnO2CC2F5 shows that accessible conformations do not allow contact between tin and fluorine, and that direct transfer is therefore unlikely to be part of the mechanism for fluorine incorporation in SnO2 films. The structure of Me2Sn(O2CCF3)2(H2O) has also been determined and adopts a trans‐Me2SnO3 coordination sphere about tin in which each carboxylate group is monodentate. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

17.
Molybdenum(VI) bis(imido) complexes [Mo(NtBu)2(LR)2] (R=H 1 a ; R=CF3 1 b ) combined with B(C6F5)3 ( 1 a /B(C6F5)3, 1 b /B(C6F5)3) exhibit a frustrated Lewis pair (FLP) character that can heterolytically split H−H, Si−H and O−H bonds. Cleavage of H2 and Et3SiH affords ion pairs [Mo(NtBu)(NHtBu)(LR)2][HB(C6F5)3] (R=H 2 a ; R=CF3 2 b ) composed of a Mo(VI) amido imido cation and a hydridoborate anion, while reaction with H2O leads to [Mo(NtBu)(NHtBu)(LR)2][(HO)B(C6F5)3] (R=H 3 a ; R=CF3 3 b ). Ion pairs 2 a and 2 b are catalysts for the hydrosilylation of aldehydes with triethylsilane, with 2 b being more active than 2 a . Mechanistic elucidation revealed insertion of the aldehyde into the B−H bond of [HB(C6F5)3]. We were able to isolate and fully characterize, including by single-crystal X-ray diffraction analysis, the inserted products Mo(NtBu)(NHtBu)(LR)2][{PhCH2O}B(C6F5)3] (R=H 4 a ; R=CF3 4 b ). Catalysis occurs at [HB(C6F5)3] while [Mo(NtBu)(NHtBu)(LR)2]+ (R=H or CF3) act as the cationic counterions. However, the striking difference in reactivity gives ample evidence that molybdenum cations behave as weakly coordinating cations (WCC).  相似文献   

18.
The title compound, C5H12NO2+·C2F3O2? or BET+·CF3COO? [BET is tri­methyl­glycine (betaine); IUPAC: 1‐carboxy‐N,N,N‐tri­methyl­methanaminium inner salt], contains pairs of bet­ainium and tri­fluoro­acetate ions forming a dimer bridged by a strong hydrogen bond between the carboxyl and carboxyl­ate groups of the two ions. The molecular symmetry of the cation is close to Cs, with protonation occurring at the carboxy O atom positioned anti to the N atom. The tri­fluoro­acetate anions are disordered over two positions. In one, the conformation of the CF3 group is staggered with respect to the carboxyl­ate group, in the other, it is close to an eclipsed conformation. The sole hydrogen bond present in the structure is the strong O—H?O bond between the anion and the cation.  相似文献   

19.
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   

20.
Reactions of NN-Dihaloperfluoroalkaneamines with Sulfur and Sulfur Derivatives Reactions of NN-Dihaloperfluoroalkaneamines RfNX2 (Rf = CF3, C2F5; X = Cl, Br) with S8, S4N4 and A = SX2 (A = RfN, O) are described. The products isolated are: Sulfurdihalideimides RfNSX2 (Rf = CF3, C2F5; X = Cl, Br), Sulfurdiimides RfNSNRf and Bis(sulfurdiimido)sulfides (RfNSN)2S(Rf = CF3, C2F5). Thionylimides RfNSO were not obtained in preparative quantities.  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