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 15 毫秒
1.
Phosphinous acid or phosphane oxide? Both tautomers of (C2F5)2POH (see picture) are found in the neat liquid, whereas only phosphinous acid is present in the gas phase or in solution. The synthesis starting from (C2F5)3PF2, the thermodynamics of the tautomerization, and the detection and thermodynamics of the cis‐/trans‐P? OH rotamers of the acid are described.

  相似文献   


2.
The electronic properties of organyl element compounds are strongly influenced by the electronic characteristics of the organic substituents. The bonding of two CF3 groups to a phosphorus atom effects a drastically decreased basicity. That is the phosphorus atom is the least basic centre in the compound (CF3)2POH. This compound, synthesized in 1960 by Burg and Griffiths, is the only known example of a phosphinous acid, although there should be a general interest in this class of compounds. However, only a few investigations have been reported which may be explained by the tedious and risky synthesis. In this paper a safe one step and high yield synthesis of (CF3)2POH is described. The compound (C6F5)2POH, originally claimed as a phosphinous acid, is proved to exist at room temperature exclusively in the tautomeric oxide form. (C6F5)2P(O)H crystallizes in the triclinic space group (no. 2) with a 992.9(1) pm; b 1501.9(2) pm; c 1539.4(2) pm; α 117.48(1)°; β 100.39(1)°; γ 96.02(1)° and Z 6.Quantum chemical investigations prove the electron withdrawing effect of s-triazinyl groups (1,3,5-triazin-4-yl derivatives) to be much stronger than that of pentafluorophenyl groups. Quantum chemical calculations at the B3PW91/6-311G(3d,p) level of theory predict for the bis(s-triazinyl) derivative (C3N3H2)2POH the phosphinous acid isomer to be favored by ΔEZP = 22 kJ/mol in relation to the corresponding phosphane oxide isomer. The phosphinous acid (CF3)2POH (Cs symmetry) is favored at the same level of theory by about ΔEZP = 14 kJ/mol compared with the phosphane oxide structure (Cs symmetry).  相似文献   

3.
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   

4.
The ESR spectra of radical anions formed by reduction of α-diketones RC(O)C(O)CF3 (R=(CF3)2CF, C6F5, (CF3)3C) with metals (Li, Na, K, Mg, Cd, Zn, Hg, In, and TI) in THF were studied. For R=(CF3)2CF and C6F5, the radical anions are formed ascis-isomers, whereas for R=(CF3)3C,trans-isomers are obtained. Line broadening due to solvation and desolvation of the cation is observed in the latter case. The reduction of α-diketone (CF3)2CFC(O)C(O)CF3 with Group II metals (Mg, Cd, Zn) results in the formation of radical pairs. Translated fromIzvestiya Akadmii Nauk. Seriya Khimicheskaya, No. 11, pp. 2228–2231, November, 1998.  相似文献   

5.
The reaction of the sterically shielded phosphane derivative, dichlorodiethylaminophosphane, Cl2PNEt2, with an excess of a mixture of 2,6‐bis(trifluoromethyl) and 2,4‐bis(trifluoromethyl)phenyl lithium gives bis[2,4‐bis(trifluoromethyl)phenyl]diethylaminophosphane, [2,4‐(CF3)2C6H3]2PNEt2, in 72 % yield as a colourless solid, while 2,6‐bis(trifluoromethyl)phenyl lithium remains unchanged in solution. The amino derivative crystallizes in the monoclinic space group P21/c (a 869.2(1), b 1857.4(1), c 1357.6(1) pm, β 100.57(4)°, Z = 4). Treatment of [2,4‐(CF3)2C6H3]2PNEt2 in CHCl3 solution with conc. HCl allows the synthesis of [2,4‐(CF3)2C6H3)]2PCl. [2,4‐(CF3)2C6H3]2PCl reacts with H2O in THF solution with quantitative formation of the corresponding secondary phosphane oxide. To obtain bis[2,4‐bis(trifluoromethyl)phenyl]phosphinic acid, [2,4‐(CF3)2C6H3]2P(O)OH, quantitatively, a CHCl3 solution of [2,4‐(CF3)2C6H3]2P(O)H, has to be stirred in an NO2 atmosphere. The phosphinic acid crystallizes is the triclinic space group (a 754.2(1), b 927.6(2), c 1305.5(2) pm, α 85.11(2)°, β 75.45(1)°, γ 79.99(2)°, Z = 2). From the reaction of the phosphinic acid with either elemental sodium or with cyanide salts, the corresponding phosphinate salts are obtained in an almost quantitatively yield.  相似文献   

6.
The reaction of alkynyldifluoroboranes RC≡CBF2 (R = (CH3)3C, CF3, (CF3)2CF) with organyliodine difluoride R′IF2 bearing electron‐withdrawing polyfluoroorganyl groups R′ = C6F5, (CF3)2CFCF=CF, C4F9, and CF3CH2 leads to the corresponding alkynyl(organyl)iodonium salts [(RC≡C)(R′)I][BF4]. This approach uses a widely applicable method as demonstrated for a representative series of polyfluorinated aryl‐, alkenyl‐, and alkyliodine difluorides. Generally, these syntheses proceed with good yields and deliver pure iodonium salts. The distinct electrophilic nature of their [(RC≡C)(R′)I]+ cations is deduced from multinuclear magnetic resonance data. Within the series of new iodonium salts [CF3C≡C(C4F9)I][BF4] is an intrinsic unstable one and decomposed forming CF3C≡CI and C4F10.  相似文献   

7.
2‐X‐1, 2‐Difluoroalk‐1‐enylxenon(II) salts were prepared by the reaction of XeF2 with XCF=CFBF2 (X = F, trans‐H, cis‐Cl, trans‐Cl, cis‐CF3, cis‐C2F5) but no organoxenon(II) compounds were obtained when the trans‐isomers of boranes, trans‐XCF=CFBF2 (X = CF3, C4F9, C4H9, Et3Si), were used under similar conditions.  相似文献   

8.
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   

9.
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   

10.
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   

11.
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   

12.
N‐sulfinylacylamides R‐C(=O)‐N=S=O react with (CF3)2BNMe2 ( 1 ) to form, by [2+4] cycloaddition, six‐membered rings cyclo‐(CF3)2B‐NMe2‐S(=O)‐N=C(R)‐O for R = Me ( 2 ), t‐Bu ( 3 ), C6H5 ( 4 ), and p‐CH3C6H4 ( 5 ) while N‐sulfinylcarbamic acid esters R‐O‐C(=O)‐N=S=O react with 1 to yield mixtures of six‐membered (cyclo‐(CF3)2B‐NMe2‐S(=O)‐N=C(OR)‐O) and four‐membered rings (cyclo‐(CF3)2B‐NMe2‐S(=O)‐N(C=O)OR) for R = Me ( 6 and 9 ), Et ( 7 and 10 ), and C6H5 ( 8 and 11 ). The structure of 5 has been determined by X‐ray diffraction.  相似文献   

13.
Chiral, bidentate phosphane ligands, so-called PP ligands, are most frequently synthesized by reacting chiral ditosylates with diarylphosphanide ions.To use the P(CF3)2 ion in nucleophilic substitution reactions, it is necessary to reduce the negative hyperconjugation, which is associated with a C-F activation. For this reason we synthesized different bis(trifluoromethyl)phosphanido complexes of mercury, silver and tungsten and investigated their use in nucleophilic substitution reactions. The most reactive compound, resulting from this study so far, is the pentacarbonylbis(trifluoromethyl)phosphanidotungstenate, [W{P(CF3)2}(CO)5], which exhibits nearly the same bonding situation for the P(CF3)2 unit as in the free P(CF3)2 ion. For use in synthesis of bis(trifluoromethyl)phosphane derivatives, Lewis acids are desirable, which stabilize the P(CF3)2 ion by an intermediary formation of a donor acceptor adduct and can be split off after the synthesis of bis(trifluoromethyl)phosphane derivatives, as could be achieved using the extremely weak Lewis acids, CS2 and acetone. These results could be established in the synthesis of the first example of an chiral, bidentate bis(trifluoromethyl)phosphane derivative.To synthesize a chiral, bidentate bis(pentafluorophenyl)phosphane derivative, a different synthetic strategy is necessary that does not involve the P(C6F5)2 ion, which decomposes even at low temperature. The implementation of functional P(CN)2 ions leads to the synthesis of functional, chiral bidentate dicyanophosphane derivatives which finally can be transformed into the corresponding bis(pentafluorophenyl)phosphane derivatives.  相似文献   

14.
The crystal structures of the title compounds, [Mo{(C4H8NO)2P(C2F5)}(CO)5], (1a), and [Mo{(C5H10N)2P(C2F5)}(CO)5], (2a), were determined as part of a larger project that focuses on the synthesis and coordination chemistry of phosphane ligands possessing moderate (electroneutral, i.e. neither electron‐rich nor electron‐deficient) electronic characteristics. Both complexes feature a slightly distorted octahedral geometry at the metal center, due to the electronic and steric repulsions between two of the four equatorial CO groups and the pentafluoroethyl group attached to the phosphane ligand. Bond length and angle data for (1a) and (2a) support the conclusion that the free phosphane ligands are electroneutral. For complex (1a), the Mo—P, Mo—Cax and Mo—Ceq(ave) bond lengths are 2.5063 (5), 2.018 (2) and 2.048 (2) Å, respectively, and for complex (2a) these values are 2.5274 (5), 2.009 (3) and 2.050 (3) Å, respectively. Geometric data for (1a) and (2a) are compared with similar data reported for analogous Mo(CO)5 complexes.  相似文献   

15.
The complexes (η-C5Me5)2Rh2(μ-CO) {μ-η22-C(O)CRCR} are obtained from reactions between (η-C5Me5)2Rh2(CO)2 and the alkynes RCCR (R  CF3, CO2Me, or Ph) at 25°C. The molecular geometry of the complex with R  CF3 has been established by X-ray diffraction; the bridging 'ene-one' unit adopts a μ-η22 conformation. Other complexes isolated from these reactions include (η-C5Me5)Rh(C6R6) (R  CF3, CO2Me), (η-C5Me)2Rh2(C4R4) (R  CO2Me) and (η-C5Me5)2Rh2(CO2C2R2) (R  Ph). The reaction between (η-C5Me5)2Rh2(CO)2 and C6F5CCC6F5 gives (η-C5Me5)2Rh2(CO)2(C6F5C2C6F5). Mononuclear complexes such as (η-C5Me5)Co(C4R4CO) are the major products isolated from reactions between (η-C5Me5)2CO2(CO)2 and alkynes at 25°C.  相似文献   

16.
17.
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   

18.
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   

19.
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   

20.
The synthesis of the germylene phosphane adduct (C2F5)2Ge?PMe3 is described. Starting from (C2F5)3GeH in an excess of PMe3, heating was applied, whereupon reductive elimination of C2F5H occurred. The molecular structure was ascertained by X‐ray diffraction and compared with information obtained by quantum chemical methods. The ligand properties were derived by studying the IR spectrum of the nickel(0) complex [Ni(CO)3{Ge(C2F5)2(PMe3)}] in the CO region. (C2F5)2Ge?PMe3 turned out to be a π‐accepting ligand comparable to PMe3, in terms of Tolman's electronic parameter. Furthermore a [2+4] cycloaddition reaction with 2,3‐dimethyl‐1,3‐butadiene, and σ‐bond insertion reactions were recorded. Activation of the C?Cl bond in dichloromethane gives rise to the formation of the phosphonium ylide complex [(C2F5)2Cl2Ge‐CH2PMe3], which was fully characterized by X‐ray diffraction.  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