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.
Proton-NMR. spectra of amino- and hydroxypyrimidines including biologically important bases have been measured in four solvents: CF3COOH, CF3COOH? SO2, FSO3H and FSO3H? SbF5–SO2 at 27° and ?55°C. In CF3COOH mono-cations are formed, whereas in FSO3H and FSO3H? SbF5–SO2 double protonation occurs. In each case the structures of the protonated species are derived from the chemical shifts of CH, NH and OH protons and proton-proton spin coupling constants. A combination of the measurements described leads to a complete assignment of all proton resonances of the protonated pyrimidines. This approach is also recommended for the structural determination of heterocyclic compounds.  相似文献   

2.
Hydrates of Fluorosulfuric Acid: The Melting Diagram of the System FSO3H? H2O and the Crystal Structure of the Monohydrate, (H3O)FSO3 The system FSO3H? H2O has been investigated by difference thermal analysis. By adhering to exclusively lower temperatures in the course of preparation and manipulation of the measuring samples, it was possible to avoid hydrolysis of the S? F bond and to obtain a quasi-binary melting diagram. This reveals the existence of four crystalline hydrates FSO3H · nH2O with n = 1, 2, 3, and 4, melting at –12, –53 (dec.), –46, and ?63°C (dec.), respectively. The structure of the monohydrate has been determined (crystal system orthorhombic, space group Pnma, Z = 4 formula units per unit cell; lattice constants a = 8.055, b = 6.465, c = 7.459 Å at ?50°C; final R value with 515 independent observed MoKα diffractometer data at 0.043). The result is a typical oxonium salt, (H3O)FSO3, characterized by strong hydrogen bonds only of the kind O? H…?O and with interesting relations to the isosteric, dimorphic compound (H3O)ClO4.  相似文献   

3.
Parent dihydropyrene 1 and 2,7-di-tert-butyldihydropyrene 3 are monoprotonated with FSO3H/SO2CIF to give their persistent monoarenium ions 1H + and 3H + by the attack of proton at C-3 (peri to the ethano-bridge not at C-1 as previously suggested). Dihydropyrene 3 is diprotonated in FSO3H.SbF5 1∶1 “Magic Acid”R/SO2CIF to give the symmetrical dication 3H 2 +2, similar diprotonation of 1 with “Magic Acid”/SO2CIF or with FSO3H.SbF5 (4∶1)/SO2CIF gave the diprotonated species 1H 2 +2 in a mixture. NMR characteristics of the mono- and dications are discussed. On raising temperature or on prolonged cold storage, 1H + and 3H + are converted to their corresponding pyrenium cations (2H + and 4H +). Formation of 2H + from 1 is more rapid than conversion of 3 to 4H +. Parent pyrenium cation was independently generated by protonation with FSO3H/SO2CIF. When a mixture of 2 and 1 is reacted with FSO3H/SO2CIF (dry ice/acetone temperature) only 2H + is seen in the NMR (concomitant presence of the radical cation 1 ++ is inferred from EPR). Similar protonation of a mixture of 6-chlorochrysene 5 and 1 with FSO3H/SO2CIF leads to NMR observation of 1H + (with concomitant presence of 5 ++); on raising temperature 1H + is converted to 2H +. The nature of the paramagnetic radical cation (RC) present in the arenium ion samples influences the position, and resolution of the NMR spectra. This approach may prove useful in NMR studies of large polycyclic aromatic hydrocarbons PAHs where concomitant RC formation greatly diminishes the quality of the NMR spectra.  相似文献   

4.
Using an approximation method, the dissociation constants of HJ, FSO3H and CF3SO3H in glacial acetic acid could be derived from conductivity measurements: $$K_{diss}^{HJ} = 10^{ - 5,8_, } K_{diss}^{FSO_3 H} = 10^{ - 6,1} and K_{diss}^{CF_3 SO_3 H} = 10^{ - 4,7} $$ These values show CF3SO3H to be the strongest known acid in glacial acetic acid, HJ to be slightly weaker than HBr and FSO3H intermediate between HJ and H2SO4.  相似文献   

5.
The energies of gas-phase solvation of the FSO3H 2 + , and ClSO3H 2 + cations and FSO 3 ? , ClSO 3 ? anions by one molecule of the respective acid (FSO3H or ClSO3H) have been calculated by a quantum chemical method taking into account electron correlation in terms of MP2 theory with a 6-311++G(d,p) basis set. The energy of additional solvation of the resulting complexes by liquid acid has been estimated within the continuum model of solvation by the IPCM method, with the acid modeled as a continuum with a large dielectric constant. The calculated data have provided a quantitative estimate for the energy of self-ionization of liquid FSO3H and ClSO3H acids. The similarly calculated energy of solvation of protons in 100% fluorosulfonic and chlorosulfonic acids is lower than the heat of hydration of the proton in aqueous solution by 23.4 and 24.5 kcal/mol, respectively. These quantum chemical data explain why 100% liquid fluorosulfonic and chlorosulfonic acids exhibit the properties of a superacid.  相似文献   

6.
13C-NMR. spectra of pterin, xanthopterin, isoxanthopterin, leucopterin, lumazine and of the model compounds isocytosine and desamino-isocytosine have been measured as anions and cations in 1 M NaOD, CF3COOH, H2SO4 and FSO3H solutions. The spectra were analysed by means of heteronuclear double resonance, with the aid of non-decoupled spectra, and by spectral comparison. The results are interpreted in terms of the ionisation state of the pteridines in the four solvents and are compared with those obtained from 1H-NMR. spectroscopy.  相似文献   

7.
Primary diazoketones, R? CO? CHN2, are O-protonated in HF? SbF5? SO2 or FSO3H? SbF5? SO2 at ?60°, as observed by NMR. The OH-proton resonates at 9.3–9.6 δ and is coupled with H? C 1 (J = 1–2.5 Hz). Secondary diazoketones, R? CO? C(N2)? R, when protonated, give an OH-singlet at 8.85 δ. The assignments are corroborated by use of deuterated diazoketones, R? CO? CDN2, or deuterated acid, FSO3D. Primary diazoketones react with FSO3H at ?60° to ?15°, giving products assigned the fluorosulfate structure, R? CO? CH2? OSO2F; they do not exchange H? C 1 with solvent before or during decomposition. Intermediate C-protonated diazonium ions and α-oxo-carbonium ions (vinyl carbonium ions) have not been identified. 3-Diazo-4-methyl-2-pentanone (VIII) reacts with FSO3H at ?15°, eliminating N2 and giving protonated mesityl oxide by a strictly intramolecular hydride shift.  相似文献   

8.
Preparation, Structure, and Reactions of N,N-Difluorosulfonylfluoridamide FSO2NF2 is obtained in high yield by fluorination of FSO2NH2 at room temperature. The reaction of FSO2NF2 with Et2NH yields FSO2NEt2 and HNF2. The structure of FSO2NF2 was determined by electron diffraction analysis.  相似文献   

9.
Protonation of acylferrocenes (FeCOR) in FSO3H-SO2CIF(SO2) solution was studied by PMR spectroscopy. The site of protonation is found to be at the carbonyl oxygen atom. Temperature dependent PMR spectra of protonated acylferrocenes FeCROH+ (R = CH3, C2H5, C6H5, OCH3) were observed indicating intermolecular hydrogen exchange with the acid solvent system. In addition the PMR spectra of acylferrocenes in FSO3 H-SO2 CIF(SO2) were found to be dependent upon the acid concentration.  相似文献   

10.
The reactions of 9-methyl-, 9-ethyl- and 9-propyl-anthracene with CF3CO2[2H] in C[2H]Cl3 and with FSO3[2H] in SO2ClF have been investigated. Using 4 equivalents of CF3CO2[2H] at 50° 1H-2H exchange was observed only for the 10-H and the side-chain α-hydrogens, and on using 8 and 12 equivalents at 50° also for the aromatic α-hydrogens. Treatment of the substrates with FSO3[2H] at -60° leads to the stable 9-alkyl-[10-2H]-10-anthracenium ions. On warming up to ?25° a slow 1H-2H exchange of only the 10-1H of these anthracenium ions is observed. A mechanism for the 1H-2H exchange of the aromatic and the side-chain α-hydrogens of the 9-alkylanthracenes is presented.  相似文献   

11.
By the reaction of FSO2N?PCl3 with perfluorpropionic acid FSO2NHC(O)C2F5 is formed, which yields FSO2N?C(Cl)? C2F5 (I) with PCl5. The chlorine atom in (I) could be replaced by the substituents NH2 (II) and N(C2H5)2 (III). FSO2N?C(Cl)? CF3 reacts with AgOCN, AgSCN, unhydrous HF and 2,3-dimethylbutadiene. FSO2N(CH3)? C(O)F reacts with elemental fluorine under exchange of a proton against a fluorine atom to give FSO2N(CH2F)? C(O)F, which liberates at room temperature COF2 and trimerises to form 1,3,5-Tris-fluorosulfonyl-s-triazine (VIII). The amides FSO2N?C(CH3)NH2 and FSO2N?C(CF3)NH2 react with SF4 in the presence of NaF to yield the iminosulfur difluorides FSO2N?C(CH3)? NSF2 (IX) and FSO2N?C(CF3)? NSF2 (X)  相似文献   

12.
The protonation of o-methoxy- and o-hydroxy-benzaldehydes in FSO3H? SbF5? SO2 solution was investigated by 1H NMR spectroscopy. The formation of the Z-carbonyl protonated molecule is explained by intramolecular hydrogen bonding.  相似文献   

13.
K. Nishiyama  K. Hata  T. Sato 《Tetrahedron》1975,31(3):239-244
[2.2]Metacyclophane (1) undergoes a variety of reactions according to the reagents and conditions. These include (1) substitution (path a), (2) transannular dehydrogenation (path b and c), (3) cycloisomerization (path d) and (4) transannular hydrogenation. A brief summary of these reactions is presented.The diversity of the reactions of 1 is further explored using benzoyl peroxide (BPO), cupric chloride, aluminum chloride, other metal salts, H2SO4 and FSO3H. With BPO or cupric chloride, one-electron transfer mechanism is postulated. This involves a tautomeric ion pair formed by the intramolecular arylation with an aryl cation radical. A supporting evidence in favor of the mechanism is presented from experiments using various metal salts in different solvents.On the contrary, the reaction with aluminum chloride gives decahydropyrene (8) and octahydropyrene (11) together with cycloisomerization product 5 and dehydrogenation products 2, 3 and 4. When treated with AlCl3HCl 1 gives similar products as above but the product ratios are quite different. The major product is 5 but only a trace amount of 8 is formed. The reaction with H2SO4 or FSO3H also produce 2, 4, 5 and 11. Some mechanistic evidence in favor of the disproportionation reaction is presented.  相似文献   

14.
The reaction of NO+ with o-toluidinechromium tricarbonyl has been studied. Diazotization (attack on N) competes with NO+ attack on the metal and decarbonylation. The Cr(CO)3-complexed diazonium ion is unstable and dediazoniates even at low temperature. The dediazoniation mechanism is predominantly homolytic. Competing heterolytic dediazoniation is observed in highly ionizing, low nucleophilicity solvents such as CF3SO3H (TfOH), FSO3H and CF3CH2OH (TFE).  相似文献   

15.
The homo-l, 4 adduct obtained from difluorocarbene and bicyclo [2.2.1]hepta-2, 5-diene ( 1 ) was treated successively with HCl, FSO3H and SbF5 in SO2ClF at low temperature. The protic acids underwent electrophilic addition to the cyclopropane part of 1 , giving the corresponding derivatives. However, in FSO3H at ? 50°, protonation of the gem-difluoro grouping also occurred to give the 2-fluoro-6-fluorosulfonylbicyclo [3.2.1]oct-2-en-3-yl cation. The reaction of 1 with SbF5 at ?78° led initially to the formation of the 2-fluorobicyclo [3.2.1 ]octa-2, 6-dien-4-yl cation, which rearranged to 4-fluorotricyclo [5.1.0.05,8]oct-3-en-2-yl cation at ?40°. These rearrangements are discussed in the light of those expected for C8H8F square pyramidal cations.  相似文献   

16.
Polysulfonyl Amines. LXXII. Triphenylcarbenium and Triphenylphosphonium Di(fluorosulfonyl)amides: Two Crystal Structures with Ordered (FSO2)2N? Sites Treatment of HN(SO2F)2 in CH2Cl2 with Ph3P, Ph3PO or collidine (=B) affords the compounds Ph3PH[(FSO2)2N]? ( 3 ), Ph3PO · HN(SO2F)2, and BH[(FSO2)2N]? ( 7 ). The carbenium salt Ph3C[(FSO2)2N]? ( 5 ), obtained by metathesis of Ph3CBr with [(C6H6)AgN(SO2F)2] in CH2Cl2, crystallizes from chloroform/petroleum ether as a monosolvate Ph3C[(FSO2)2N]? · CHCl3 ( 6 ). In presence of a sterically hindered base, viz. collidine, 5 is a suitable reagent for the tritylation of molecules containing weakly activated H atoms (e. g.: MeCN → Ph3CCH2CN, acetone → tritylacetone; co-product: 7 ). The crystal structures of the ionic solids 3 (monoclinic, space group P21/n) and 6 (monoclinic, P21/c) were determined by X-ray diffraction at ?130°C; the structure refinements were not impaired by the notorious tendency of the (FSO2)2N moiety towards crystallographic disorder. As in the known structure of the tetraphenylarsonium salt, the anion of 3 and 6 adopts a staggered conformation of approximately C2 symmetry (averages of all values: S? N? S 121.4°, N? S 156.2, S? O 141.6, S? F 156.6 pm). The crystal packing of 6 displays a three-centre C? H(…?O)2 hydrogen bond between the CHCl3 molecule and two oxygen atoms of a single anion, resulting in a six-membered ring [R12(6) pattern; H …? O 234 and 262 pm]. The crystal of 3 contains one-dimensional arrays of alternating cations and anions connected by a three-centre P? H(…?O)2 bond [C(6) pattern; H …? O 237 and 254 pm]. The Ph3C cation of 6 is propeller-shaped, with three coplanar central bonds (mean C? C 144.5 pm) and interplanar angles of 52.7, 56.4 and 60.1° between the phenyl groups.  相似文献   

17.
《Tetrahedron letters》2003,44(35):6651-6655
Among several sulfonic acids studied (MeSO3H, p-TsOH, H2SO4, ClSO3H, FSO3H), the scarcely used chlorosulfonic acid showed to be an efficient agent for electrophilic olefin cyclizations with internal nucleophilic termination, in a similar manner that is well-established with fluorosulfonic acid. Its availability, lower price and relatively lesser handling problems makes ClSO3H an advantageous cyclizing agent particularly for high-scale applications. The stereochemical outcome of these cyclizations has been rationalized.  相似文献   

18.
A reaction between 2, 8-dichloro-4, 10-dinitro-5, 11-dehydro-5H, 11H-benzotriazolo[2, 1-a]-benzotriazole 8 and sodium azide in dimethyl sulfoxide produced 3, 9-diazido-4, 10-dinitro-5, 11-dehydro-5H, 11H-benzotriazolo [2, 1-a]benzotriazole 10 rather than the 2.8-diazido isomer 9 expected by direct displacement. Thermolytic elimination of nitrogen (2 moles) converted the dinitro diazide 10 to 3,4,9,10-bisfuroxano-5, 11-dehydro-5H, 11H-benzotriazolo[2, 1-a]benzotriazole 11 that was subsequently nitrated to give the 2,8-dinitro derivative 12 . Similar nitration converted the dinitro diazide 9 to the trinitro 15 and tetranitro 14 derivatives: thermolysis of the latter gave 1,2,7,8-bisfuroxano-4, 10-dinitro-5, 11-dehydro-5H, 11H-benzotriazolo[2, 1-a]-benzotriazole 16 . Nitration (100% HNO3, CF3SO3H) converted compound 16 to the 3,4,10-trinitro derivative 17 , whereas a similar nitration (100% HNO3, FSO3H) gave the title compound BTBB, an insensitive high-energy, high-density (d 2.03 g/cc) molecule. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
The triphenylallenyl cation (8), generated from triphenylpropynol (7) and FSO3H, reacts with alkenes to give the allyl cations 12, which may be deprotonated to yield the methyleneccylobutenes 14. Alternatively, 12 can be converted into the 2-vinyl-indenes 13 via two subsequent electrocyclic reactions.  相似文献   

20.
The gas phase structure of trifluoroethylene sultone, ( 1 ) (3,4,4-trifluoro-1,2-oxathietane-2,2-dioxide) was determined by gas electron diffraction, and the four-membered ring was found to be planar. The following ring parameters (ra distances and ∠α angles with 3σ uncertainties) were derived in the electron diffraction analysis: C? O = 1.41 Å (ass.), C? C = 1.541(18) Å, S? O = 1.652(5) Å, S? C = 1.822(8) Å, S? C? C = 86.2(15)°, C? C? O = 97.1(28)°, C? O? S = 97.5(21)°, and O? S? C = 79.1(8)°. New spectral data (IR, NMR) of 1 , its acyclic isomer FSO2CFHC(O)F ( 2 ), and the related anhydride, FSO2OSO2CFHC(O)F ( 3 ), are reported. New esters containing the fluorosulfonyl function, FSO2CFHC(O)OCH2CF3 ( 4 ), FSO2CFHC(O)OCH2CH = CH2 ( 5 ), and (FSO2CFHC(O)OCH2CH? CH2? )n ( 6 ) have been prepared and characterized.  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