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 218 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.
As an emerging member of endohedral fullerenes, metal cyanide clusterfullerenes (CYCF) are unique in terms of the encapsulation of a monometallic cluster. To date the reported carbon cages of CYCFs are limited to C82 and C76, and little is known about the chemical reactivity of CYCFs. Herein, two isomers of the first C84‐based CYCFs, YCN@C84, were isolated as trifluoromethyl derivatives, including YCN@C84(23)(CF3)18 and three isomers of YCN@C84(13)(CF3)16, which are based on a unique chiral C 2‐C84(13) cage. As a common feature of the CF3 addition patterns, the YCN@C84(CF3)16/18 compounds are stabilized by the formation of isolated C=C bonds and benzenoid rings on the carbon cages. The interplay between the endohedral YCN cluster and the exhohedral CF3 addends was unveiled according to single‐crystal X‐ray diffraction studies, thus offering new insight into the chemical reactivity of CYCFs.  相似文献   

3.
As an emerging member of endohedral fullerenes, metal cyanide clusterfullerenes (CYCF) are unique in terms of the encapsulation of a monometallic cluster. To date the reported carbon cages of CYCFs are limited to C82 and C76, and little is known about the chemical reactivity of CYCFs. Herein, two isomers of the first C84‐based CYCFs, YCN@C84, were isolated as trifluoromethyl derivatives, including YCN@C84(23)(CF3)18 and three isomers of YCN@C84(13)(CF3)16, which are based on a unique chiral C 2‐C84(13) cage. As a common feature of the CF3 addition patterns, the YCN@C84(CF3)16/18 compounds are stabilized by the formation of isolated C=C bonds and benzenoid rings on the carbon cages. The interplay between the endohedral YCN cluster and the exhohedral CF3 addends was unveiled according to single‐crystal X‐ray diffraction studies, thus offering new insight into the chemical reactivity of CYCFs.  相似文献   

4.
Trifluoromethylation of higher fullerene mixtures with CF3I was performed in ampoules at 400 to 420 and 550 to 560 °C. HPLC separation followed by crystal growth and X‐ray diffraction studies allowed the structure elucidation of nine CF3 derivatives of D2‐C84 (isomer 22). Molecular structures of two isomers of C84(22)(CF3)12, two isomers of C84(22)(CF3)14, four isomers of C84(22)(CF3)16, and one isomer of C84(22)(CF3)20 were discussed in terms of their addition patterns and relative formation energies. DFT calculations were also used to predict the most stable molecular structures of lower CF3 derivatives, C84(22)(CF3)2–10. It was found that the addition of CF3 groups to C84(22) is governed by two rules: additions can only occur at para positions of C6(CF3)2 hexagons and no additions can occur at triple‐hexagon‐junction positions on the fullerene cage.  相似文献   

5.
Trifluoromethylation of a higher fullerene mixture with CF3I was performed in ampoules at 550 °C. HPLC separation followed by crystal growth and X‐ray diffraction study resulted in the structure elucidation of nine CF3 derivatives of D2d‐C84 (isomer 23). The molecular structures of C84(23)(CF3)4, C84(23)(CF3)8, C84(23)(CF3)10, C84(23)(CF3)12, two isomers of C84(23)(CF3)14, two isomers of C84(23)(CF3)16, and C84(23)(CF3)18 were discussed in terms of their addition patterns and the relative formation energies. Extensive theoretical DFT calculations were performed to identify the most stable molecular structures. It was found that the addition of CF3 groups to the C84(23) fullerene is governed by two main rules: no additions in positions of triple hexagon junctions and predominantly para additions in C6(CF3)2 hexagons on the fullerene cage. The only exception with an isolated CF3 group in C84(23)(CF3)12 is discussed in more detail.  相似文献   

6.
《化学:亚洲杂志》2018,13(16):2027-2030
High‐temperature trifluoromethylation of fullerene C76 chlorination products followed by HPLC separation of C76(CF3)n derivatives resulted in the isolation and X‐ray structural characterization of thirteen C76(1)(CF3)n compounds including nine new isomers such as one isomer of C76(1)(CF3)10, two C76(1)(CF3)12, three C76(1)(CF3)14, one C76(1)(CF3)16, and two isomers of C76(1)(CF3)18. Depending on their addition patterns, C76(1)(CF3)n isomers are divided into three subgroups and discussed in terms of trifluoromethylation pathways and relative formation energies.  相似文献   

7.
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   

8.
High‐temperature trifluoromethylation of fullerene C78 followed by HPLC separation of C78(CF3)n derivatives resulted in the isolation and X‐ray structural characterization of 15 compounds, that is, two C78(1)(CF3)10, three C78(1)(CF3)12, four C78(1)(CF3)14, and five C78(1)(CF3)16 isomers as well as one isomer of C78(1)(CF3)18. The addition patterns of the C78(1)(CF3)n molecules are discussed in terms of trifluoromethylation pathways and relative formation energies.  相似文献   

9.
New experimental results on perfluoroalkylation of C60 and C70 with the use of RfI (Rf = CF3, C2F5, n-C3F7, n-C4F9, and n-C6F13), along with a critical overview of the existing synthetic methods, are presented. For the selected new fullerene (Rf)n compounds we report spectroscopic, electrochemical and structural data, including improved crystallographic data for the isomers of C70(C2F5)10 and C60(C2F5)10, and the first X-ray structural data for the dodecasubstituted perfluoethylated C70 fullerene, C70(C2F5)12, which possesses unprecedented addition pattern.  相似文献   

10.
High‐temperature trifluoromethylation of a C90 isomeric mixture with CF3I followed by HPLC separation of C90(CF3)n isomers resulted in the isolation of several individual C90(CF3)14?18 compounds. Single crystal X‐ray diffraction with the use of synchrotron radiation resulted in the structure determination of C90(30)(CF3)14, C90(35)(CF3)16/18, and C90(45)(CF3)16/18. Their addition patterns are discussed and compared with the known isomers C90(30)(CF3)18 and C90(35)(CF3)14, respectively. The presence of the most stable C90 isomer, C90(45), in the fullerene soot has been confirmed for the first time.  相似文献   

11.
Thermal decarbonylation of the acyl compounds [Mn(CO)5(CORF)] (RF=CF3, CHF2, CH2CF3, CF2CH3) yielded the corresponding alkyl derivatives [Mn(CO)5(RF)], some of which have not been previously reported. The compounds were fully characterized by analytical and spectroscopic methods and by several single-crystal X-ray diffraction studies. The solution-phase IR characterization in the CO stretching region, with the assistance of DFT calculations, has allowed the assignment of several weak bands to vibrations of the [Mn(12CO)4(eq-13CO)(RF)] and [Mn(12CO)4(ax-13CO)(RF)] isotopomers and a ranking of the RF donor power in the order CF3<CHF2<CH2CF3≈CF2CH3. The homolytic Mn−RF bond cleavage in [Mn(CO)5(RF)] at various temperatures under saturation conditions with trapping of the generated RF radicals by excess tris(trimethylsilyl)silane yielded activation parameters ΔH and ΔS that are believed to represent close estimates of the homolytic bond dissociation thermodynamic parameters. These values are in close agreement with those calculated in a recent DFT study (J. Organomet. Chem. 2018 , 864, 12–18). The ability of these complexes to undergo homolytic Mn−RF bond cleavage was further demonstrated by the observation that [Mn(CO)5(CF3)] (the compound with the strongest Mn−RF bond) initiated the radical polymerization of vinylidene fluoride (CH2=CF2) to produce poly(vinylidene fluoride) in good yields by either thermal (100 °C) or photochemical (UV or visible light) activation.  相似文献   

12.
We report an efficient and scalable synthesis of azidotrifluoromethane (CF3N3) and longer perfluorocarbon‐chain analogues (RFN3; RF=C2F5, n C3F7, n C8F17), which enables the direct insertion of CF3 and perfluoroalkyl groups into triazole ring systems. The azidoperfluoroalkanes show good reactivity with terminal alkynes in copper(I)‐catalyzed azide–alkyne cycloaddition (CuAAC), giving access to rare and stable N ‐perfluoroalkyl triazoles. Azidoperfluoroalkanes are thermally stable and the efficiency of their preparation should be attractive for discovery programs.  相似文献   

13.
Perfluoroalkenyl phosphonates were formed along with Me3SiF using CF3CF=CF2, CF3CH=CF2, F5SCF=CF2 or F5SCH=CF2 and silylated phosphites, (R1O)2POSiMe3 (R1=Et, SiMe3). This straightforward method could be extended to perfluorobutadienes CF2=C(RF)C(RF)=CF2 (RF F=F, CF3). The formation of CF3C(=O)P(=O)(OSiMe3)2 and further reactions to yield bisphosphonates will be described. Acetylphosphonates, R2C(=O)P(=O)(OSiMe3)2 (R2=CH3, CF3) reacted with the ketimine, CH3C(=NiPr)Ph to give α-hydroxy-γ-imino phosphonates. Trifluoroacetylphenol and 2,6-bis(trifluoracetyl)-4-methyl-phenol have been proven to be versatile precursors for α-and γ-hydroxy phosphonates. Intermediates in these reactions were found to be cyclic λ5σ5P species.  相似文献   

14.
The carbon cage of buckminsterfullerene Ih-C60, which obeys the Isolated-Pentagon Rule (IPR), can be transformed to non-IPR cages in the course of high-temperature chlorination of C60 or C60Cl30 with SbCl5. The non-IPR chloro derivatives were isolated chromatographically (HPLC) and characterized crystallographically as 1809C60Cl16, 1810C60Cl24, and 1805C60Cl24, which contain, respectively two, four, and four pairs of fused pentagons in the carbon cage. High-temperature trifluoromethylation of the chlorination products with CF3I afforded a non-IPR CF3 derivative, 1807C60(CF3)12, which contains four pairs of fused pentagons in the carbon cage. Addition patterns of non-IPR chloro and CF3 derivatives were compared and discussed in terms of the formation of stabilizing local substructures on fullerene cages. A detailed scheme of the experimentally confirmed non-IPR C60 isomers obtained by Stone–Wales cage transformations is presented.  相似文献   

15.
Synthesis and Properties of Tetrakis(Perfluoroalkyl)Tellurium Te(Rf)4 (Rf = CF3, C2F5, C3F7, C4F9) Te(CF3)4 is obtained from the reaction of Te(CF3)Cl2 with Cd(CF3)2 complexes as a complex with e. g. CH3CN, DMF. It is a light and temperature sensitive hydrolysable liquid. The reaction with fluorides yields the complex anion [Te(CF3)4F]?, with fluoride ion acceptors the complex cation [Te(CF3)3]+. With traces of water an acidic solution is formed. Te(CF3)4 acts as a trifluoromethylation reagent. The reaction with XeF2 gives hints for the formation of Ye(CF3)4F2. Properties and NMR spectra are discussed. The much more stable complexes of Te(Rf)4 (Rf = C2F5, C3F7, C4F9) are formed from the reaction of TeCl4 with the corresponding Cd(Rf)2 complexes.  相似文献   

16.
A series of previously unknown asymmetrical fluorinated bis(aryl)bromonium, alkenyl(aryl)bromonium, and alkynyl(aryl)bromonium salts was prepared by reactions of C6F5BrF2 or 4-CF3C6H4BrF2 with aryl group transfer reagents Ar′SiF3 (Ar′ = C6F5, 4-FC6H4, C6H5) or perfluoroorganyl group transfer reagents RF′BF2 (RF = C6F5, trans-CF3CFCF, C3F7C≡C) preferentially in weakly coordinating solvents (CCl3F, CCl2FCClF2, CH2Cl2, CF3CH2CHF2 (PFP), CF3CH2CF2CH3 (PFB)). The presence of the base MeCN and the influence of the adducts RF′BF2·NCMe (RF = C6F5, CF3C≡C) on reactions aside to bromonium salt formation are discussed. Reactions of C6F5BrF2 with AlkF′BF2 in PFP gave mainly C6F5Br and AlkF′F (AlkF′ = C6F13, C6F13CH2CH2), presumably, deriving from the unstable salts [C6F5(AlkF′)Br]Y (Y = [AlkF′BF3]). Prototypical reactivities of selected bromonium salts were investigated with the nucleophile I-and the electrophile H+. [4-CF3C6H4(C6F5)Br][BF4] showed the conversion into 4-CF3C6H4Br and C6F5I when reacted with [Bu4N]I in MeCN. Perfluoroalkynylbromonium salts [CnF2n+1C≡C(RF)Br][BF4] slowly added HF when dissolved in aHF and formed [Z-CnF2n+1CFCH(RF)Br][BF4].  相似文献   

17.
Pentafluoroethyl derivatives of [60]fullerene C60(C2F5)n (n = 6, 8, and 10) were synthesized by the reaction of C60 with C2F5I in glass ampoules at 380–440 °C. Isomers of composition C60(C2F5)6 (one isomer), C60(C2F5)8 (five isomers), and C60(C2F5)10 (two isomers) were isolated by chromatographic separation. Their molecular structures were established by X-ray diffraction. The relative stabilities of isomers were compared by density functional theory calculations. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 881–887, May, 2007.  相似文献   

18.
The nascent vibrational energy distributions of the HF? formed in the reactions of a series of partially fluorinated alkanes (RFH; RF = CH2F, CHF2, CF3, C2F5, C3F7, and C7F15) with electronically excited oxygen atoms O(21D2) have been determined by measuring the appearance times of stimulated emissions from various vibration–rotation transitions in a grating-tuned optical cavity. The vibrational energy contents of the HF formed in these reactions were found to be considerably greater than statistically expected. These reactions are believed to occur via vibrationally excited short-lived α;-fluorinated alcohols (RFOH?), formed by insertion of the O(21D2) atoms into C? H bonds. The observation of nonstatistical energy partitioning in the above reactions is in clear contrast to the result obtained from the O(21D2) + CF3CH3 reaction that produces the β-fluorinated alcohol CF3CH2OH, from which the HF product carries a near statistical vibrational energy distribution. A mechanism for HF? formation in these very exothermic reactions is presented.  相似文献   

19.
Sodium dithionite-initiated addition of RFI (1) [RF=1a, Cl(CF2)4; 1b, Cl(CF2)6; 1c, Cl(CF2)6; 1d, n-C6F13; 1e, n-C3F17] to the properly protected glucose derivatives 2, 9, 12 and 15 gave polyfluoroalkylated glucose derivatives 7, 11, 14 and 17 respectively in good yields after reduction and deprotection. These compounds are strongly surface active.  相似文献   

20.
This minireview updates non-exhaustive recent strategies of synthesis of original fluorosurfactants potentially non-bioaccumulable. Various strategies have been focused on (i) the preparation of CF3–X–(CH2)n–SO3Na (with X = O, C6H4O or N(CF3) and n = 8–12), (ii) the oligomerization of hexafluoropropylene oxide (HFPO) to further synthesize oligo(HFPO)–CF(CF3)CO–RH (where RH stands for an hydrophilic chain); (iii) the telomerization of vinylidene fluoride (VDF) with 1-iodopentafluoroethane or 1-iodononafluorobutane to produce CnF2n+1–(VDF)2–CH2CO2R (n = 2 or 4, R = H or NH4), (iv) the radical telomerization of 3,3,3-trifluoropropene (TFP) with isoperfluoropropyliodide or diethyl hydrogenophosphonate to prepare (CF3)2CF(TFP)x–RH or CF3–CH2–CH2–(TFP)y–P(O)(OH)2, and (v) the radical cotelomerization of VDF and TFP, or their controlled radical copolymerization in the presence of (CF3)2CFI or a fluorinated xanthate. In most cases, the surface tensions versus the surfactant concentrations have been assessed. These above strategies led to various highly fluorinated (but yet not perfluorinated) telomers whose chemical changes enabled to obtain original surfactants as novel alternatives to perfluorooctanoic acid (PFOA), ammonium perfluorooctanoate (APFO), or perfluorooctylsulfonic acid (PFOS) regarded as bioaccumulable, persistent, and toxic.  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