|
1. |
Synthesis and crystal structure of the new organic conductors (TMTTF)2(C6H2N3O8), ET2(C6H2N3O8) and ET2(C6H2N3O7)(THF) with picrate and styphnate anions |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 125-127
Georgii G. Abashev,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) Synthesis and crystal structure of the new organic conductors (TMTTF)2(C6H2N3O8), ET2(C6H2N3O8) and ET2(C6H2N3O7)(THF) with picrate and styphnate anions Georgii G. Abashev,*a Olga N. Kazheva,b Oleg A. Dyachenko,b Victor V. Gritsenko,b Aleksei G. Tenishev,a Kazukuni Nishimurac and Gunzi Saitoc a Institute of Natural Sciences, Perm State University, 614600 Perm, Russian Federation.E-mail: gabashev@psu.ru b Institute of Problems of Chemical Physics, Russian Academy of Sciences, 142432 Chernogolovka, Moscow Region, Russian Federation. E-mail: doa@icp.ac.ru c Division of Chemistry, Graduate School of Science, Kyoto University, Kyoto, 6068502, Japan. E-mail: saito@kuchem.kyoto-u.ac.jp 10.1070/MC2001v011n04ABEH001402 The following new organic conductors with the anions of styphnic acid (trinitroresorcinol) and picric acid were synthesised: (TMTTF)2(C6H2N3O8) 1 (TMTTF is tetramethyltetrathiafulvalene), ET2(C6H2N3O8) 2 and ET2(C6H2N3O7)(THF) 3 [ET is bis(ethylenedithio)tetrathiafulvalene].While searching new organic conductors and superconductors among radical cation salts and charge-transfer complexes based on tetrathiafulvalene and its derivatives,1 one has two opportunities to influence their structure.The first one lies in the modifying of the TTF moiety, and the second one, in the introducing of new anions. Organic anions and their influence on the salt structure and electrophysical properties have not been adequately investigated. Thus, there was declared the preparation of organic radical cation salts, which incorporated such organic anions as fluorine containing moieties,2,3 oxalate,4 acetate,4 maleate,4 fumarate, 4 cyanoform,5,6 penta- or tetracyanoallyl,6 hexacyanotrimethylenemethanide, 6 cyananilate7 and tris(dicyanomethylene)- cyclopropanediide.8,9 There are only a few examples describing salts with NO2 groups in the structure of their anions [e.g., 2,4,7-trinitro-9-(dicyanomethylene)fluorene9,10].We synthesised some salts of bis(ethylenedithio)tetrathiafulvalene (ET) and tetramethyltetrathiafulvalene (TMTTF) with the anions of picric and styphnic (trinitroresorcinol) acids11 by electrocrystallization. The compounds prepared have the stoichiometry 2:1 and are semiconductors. Simultaneously and independently of us, the synthesis of a salt with the presumable composition (ET)2(C6H2N3O7)2(H2O)x was declared.12 The salt (TMTTF)2(C6H2N3O8) 1 crystallises as thin black needles of 4.5 mm length from a dichloroethane solution after electrochemical oxidation of TMTTF in the presence of tetraethylammonium trinitroresorcinate.¢Ó The crystal structure of salt 1 (Figure 1) is formed by one-dimensional nondimerised stacks of cation radicals (TMTTF)+1/2 (Figure 2) and trinitroresorcinol anions (C6H2N3O8).disposed one after another along the b-axis of the crystal. TMTTF stacks and anions alternate along [101] direction of the unit cell. Each of the donor molecule stacks is surrounded by the chains of anions on four sides. This ¡®chesslike¡� mode of cation.anion packing has been earlier observed in the molecular conductor (BETS)2Bi2Cl8,15 and it is similar to those observed in (TTF)2[NiS4C4H4]16 and (TMTSF)2(azaTCNQ)17 [BETS is bis(ethylenedithio)tetraselenafulvalene, TTF is tetrathiafulvalene, TMTSF is tetramethyltetraselenafulvalene and TCNQ is 7,7',8,8'-tetracyanoquinodimethane]. 0 a b c A B C Figure 1 Crystal structure of (TMTTF)2(C6H2N3O8) 1.¢Ó Crystal data for 1: C26H26N3O8S8, M = 746.98, monoclinic, space group P21/n, a = 12.634(6) A b = 14.343(7) A, c = 18.550(9) A, b = 105.95(4)¡Æ, V = 3232(3) A3, Z = 4, dcalc = 1.572 g cm.3. The experiment was carried out on a KM-4 automated diffractometer (Kuma Diffraction, Poland) with graphite-monochromated MoK¥á radiation using the w/2q scanning technique (6718 reflections).The crystal structure was solved by direct methods and subsequent Fourier syntheses using the SHELX-8613 and SHELXL-9314 program packages. C(5) C(6) C(15) C(15') C(5) C(6) C(20) C(20') C(5) C(6) C(15) C(15') C(5) C(6) C(20) C(20') A B C A B C A A Figure 2 Packing of conducting layers in the (a) 1, (b) 2 and (c) 3 salts. (a) (b) (c)Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) TMTTF cation radicals are characterised by different types of symmetry. One of the cation radicals is in the common position and has no local elements of symmetry (cation radical A), and the two others (B and C) are centrosymmetric (Figure 1). The trinitroresorcinol anions (C6H2N3O8). alternate along the b-direction of the crystal. One of the NO2 groups displaces from the benzene ring plane (the dihedral angle between the planes of the central six-membered ring and the NO2 group is equal to 76.8¡Æ).All other atoms of the anion lie in the plane of a benzene moiety. There are no shortened intrastack S¡�¡�¡�S contacts in TMTTF. At 300 K, the conductivity of 1 is s300 = 2.1¡¿10.3 S cm.1. The temperature dependence of the conductivity has an activation character.Two sections can be recognised corresponding to the two different values of activation energy EA1 = 0.12 eV (280. 400 K) and EA2 = 0.094 eV (200.280 K). In the phase transition region, the conductivity jumps. The ET styphnate single crystals of ET2(C6H2N3O8) 2¢Ô were prepared for ten days by electrochemical oxidation at ambient temperature in a THF solution at 1 ¥ìA cm.2 current density.The conductivity was measured by a d.c. four-probe technique; it has an activation character, s300 = 4.7¡¿10.2 S cm.1 and EA = 0.097 eV. The crystal structure of salt 2 is presented in Figure 3. The ET cation radicals form layers (Figure 2) separated by strongly disordered styphnate anions, where one of the NO2 groups displaces from the benzene-ring plane.Electrochemical crystallization of ET gave salt ET2(C6H2N3O7)- (THF) 3, elemental analysis of which indicated that a molecule of THF was incorporated in this salt. The stoichiometry was found to be 2:1:1.¡× In complex 3, the ET molecules form a two-dimensional segregated layer in the bc-plane, which is sandwiched by the anion layers. The stacking pattern of the ET molecules is of ¥á''-type (Figure 2).Picrate anions and THF molecules in the anion layer are orientationally disordered into two sites. This salt shows a semiconductive behaviour (Figure 4). The conductivity has an activation character: sRT = 0.3 S cm.1, EA = 57 meV. The temperature dependence of the magnetic susceptibility of 3 can be expressed by the alternate chain model18 (|J| = 143 K, a = 0.419) in the range over 100 K.The magnetic susceptibility at room temperature is 8.7¡¿10.4 emu mol.1, which is high and comparable to Mott insulators of ET materials. The crystal structure of the salt is presented in Figure 5. Here, as well as in the two previous salts, one of the nitro groups displaces from the ring plane in contrast to the 1:1 complex of hexanitrostylbene (HNS) with TTF,19 where all nitro groups lie in the plane of the corresponding benzene rings.Thus, new conducting salts of TMTTF and ET with picrate and styphnate anions were electrochemically synthesised. The incorporation of organic anions into the structure of radical cation salts of the TTF series can be promising for the synthesis of new conducting materials.This work was supported by the Russian Foundation for Basic Research (grant nos. 99-03-32872 and 00-03-32809) and by a Grant for Research for Future from JSPS. References 1 T. Ishiguro, K. Yamaji and G. Saito, Organic Superconductors, 2nd edn., Springer, Berlin, 1998. 2 U. Geiser, J. A. Schlueter, H. H. Wang, A. M. Kini, J. M. Williams, P. P. Sche, H. L. Zakowicz, M. L. VanZile and J.D. Dudek, J. Am. Chem. Soc., 1996, 118, 9996. ¢Ô Crystal data for 2: C26H18N3O8S16, M = 1013.39, triclinic, space group P1 (no. 2), a = 8.422(1) A, b = 21.761(2) A, c = 23.331(2) A, a = 80.96(0)¡Æ, b = 84.44(1)¡Æ, g = 89.99(1)¡Æ, V = 4202. Z = 4. The experiment was carried out on a Mac Science DIP-2020K diffractometer with graphitemonochromated MoK¥á radiation (5615 reflections).The crystal structure was solved by direct methods and subsequent Fourier syntheses using the SHELXL-9314 program packages. b c Figure 3 The crystal structure of ET2(C6H2N3O8) 2. 103 102 101 100 0.004 0.006 0.008 0.010 Resistivity of (ET)2(picrate)/§Ù cm T.1/K.1 Figure 4 The temperature dependences of conductivity for the single crystal of salt 3. sRT = 0.3S cm.1 Ea = 57 meV heating cooling ¡× Crystal data for 3: C30H26N3O8S16, M = 1069.50, monoclinic, space group C2/c (no. 15), a = 43.152(3) A, b = 4.2160(3) A, c = 23.256(1) A, b = 99.29(0)¡Æ, V = 4175.44(40) A3, Z = 4. The experiment was carried out on a Mac Science DIP-2020K diffractometer with graphite-monochromated MoK¥á radiation (3464 reflections). The crystal structure was solved by direct methods and subsequent Fourier syntheses using the SHELXL-9314 program packages.Atomic coordinates, bond lengths, bond angles and thermal parameters for 1.3 have been deposited at the Cambridge Crystallographic Data Centre (CCDC). For details, see ¡®Notice to Authors¡�, Mendeleev Commun., Issue 1, 2001. Any request to the CCDC for data should quote the full literature citation and the reference number 1135/90.a c Figure 5 The crystal structure of ET2(C6H2N3O7)(THF) 3.Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) 3 H. H. Wang, U. Geiser, M. E. Kelly, M. L. VanZile, A. J. Skulan, J. M. Williams, J. A. Schlueter, A. M. Kini and S. A. Sirchio, Mol. Cryst. Liq. Cryst., 1996, 284, 427. 4 P. Kathirgamanathan, S. A. Mucklejohn and D.R. Rosseinsky, J. Chem. Soc., Chem. Commun., 1979, 86. 5 M. A. Beno, H. H. Wang, L. Soderholm, K. D. Carlson, L. N. Hall, L. Nunez, H. Rummens, B. Andersen, J. A. Schluter, J. M. Williams, M.-H. Whangbo and M. Evain, Inorg. Chem., 1989, 28, 150. 6 (a) H. Yamochi, T. Tsuji, G. Saito, T. Suzuki, T. Miyashi and C. Kabuto, Synth. Met., 1988, 27, A479; (b) H. Yamochi, C. Tada, S. Sekizaki, G.Saito, M. Kusunoki and K. Sakaguchi, Mol. Cryst. Liq. Cryst., 1996, 284, 379. 7 Md. B. Zaman, Y. Morita, J. Toyoda, H. Yamochi, G. Saito, N. Yoneyama, T. Enoki and K. Nakasuji, Chem. Lett., 1997, 729. 8 T. Fukunaga, M. D. Gordon and P. J. Krusic, J. Am. Chem. Soc., 1976, 98, 611. 9 S. Horiuchi, H. Yamochi, G. Saito, K. Sakaguchi and M. Kusunoki, J. Am. Chem. Soc., 1996, 118, 8064. 10 O. Drozdova, H. Yamochi, K. Yakushi, M. Uruichi, S. Horiuchi and G. Saito, Proceedings of the Synth. Met. ICSM 2000, Austria, 2000. 11 G. G. Abashev, E. V. Shklyaeva and A. G. Tenishev, ISCOM-99, Oxford, 1999, p. 5. 12 C. Rodrigues, E. V. Lopes, M. Almeida and R. T. Henriques, ISCOM- 99, Oxford, 1999, p. 108. 13 G. M. Sheldrick, SHELX 86, Program for Crystal Structure Determination, University of Göttingen, Germany, 1986. 14 G. M. Sheldrick, SHELXL 93, Program for the Refinement of Crystal Structures, University of Göttingen, Germany, 1993. 15 N. D. Kushch, O. A. Dyachenko, V. V. Gritsenko, P. Cassoux, C. Faulman, A. Kobayashi and H. Kobayashi, J. Chem. Soc., Dalton Trans., 1998, 683. 16 P. Cassoux, L. Interrante and J. Kasper, Mol. Cryst. Liq. Cryst., 1982, 81, 293. 17 H. Urayama, T. Inabe, T. Mori, Y. Maruyama and G. Saito, Bull. Chem. Soc. Jpn., 1988, 61, 1831. 18 J. W. Hall, W. E. Marsh, R. R. Weller and W. E. Hatfield, Inorg. Chem., 1981, 20, 1033. 19 M. Fourmigue, K. Boubekeur, P. Batail, J. Renouard and G. Jacob, New J. Chem., 1998, 845. Received: 28th November 2000; Com. 00/17
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
2. |
The first metal complex with avic-dihydroxyamine and its oxidised derivative |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 127-128
Sergei V. Fokin,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) The first metal complex with a vic-dihydroxyamine and its oxidised derivative Sergei V. Fokin, Galina V. Romanenko and Victor I. Ovcharenko* International Tomography Centre, Siberian Branch of the Russian Academy of Sciences, 630090 Novosibirsk, Russian Federation. Fax: +7 3832 33 1399; e-mail: ovchar@tomo.nsc.ru 10.1070/MC2001v011n04ABEH001460 The NiII complex with a vic-dihydroxyamine has been synthesised, characterised by X-ray analysis and oxidised to unusual bischelate containing fully dehydrogenated hydroxyamine groups. 2,3-Dihydroxyamino-2,3-dimethylbutane 1 is widely used for syntheses of nitronylnitroxides.1,2 However, vic-dihydroxyamines were never investigated as ligands. Here, we report on the synthesis and structure of the first NiII complex with vic-dihydroxyamine, which may be oxidised to a bischelate with fully dehydrogenated hydroxyamine groups.We found that the interaction between aqueous NiCl2 and 1 forms a finely dispersed yellow precipitate of 2 (Scheme 1), which is sparingly soluble in water and organic solvents.¢Ó The same product crystallised from the reaction mixture on the addition of a base (Na2CO3 or NaOH).Compound 2 is stable in the solid state but gradually decomposes in solution during two to three days. The single-crystal study¢Ô showed that the structure of the coordination site in 2 is a nearly regular square formed by Ni2+ and the N atoms of two bidentately coordinated molecules of 1, one of which is deprotonated (Figure 1).¡× The chloride ion is not involved in coordination.The cation has two nonequivalent intramolecular H-bonds between the coordinated ligands, because of which 2 is related to classical metal dioximates. However, the Ni.N bond in 2 is much longer (1.92 A) than that in metal dioximates (~1.87 A), and 2 is much less stable in solution, where it is readily decomposed by dilute acids.An interesting property of 2 is its ability to be oxidised to unusual complex 3, where the ligand is the product of complete dehydrogenation of hydroxyamine groups. Dehydrogenation occurs readily in the oxidative system 2.PbO2.C6H6.H2O.NaOH, whereupon 3 may be isolated from the benzene extract. Compound 3 is stable as a solid but much less stable in solutions ¢Ó [Ni(H3L)H4L]Cl 2.A mixture of powdered Ni(H2O)6Cl2 (0.45 g) and 1¡�H2SO4¡�H2O6 (1 g) was dissolved in 15 ml of water, the solution was filtered, and 10 ml of aqueous Na2CO3 (0.5 g) was added to the solution. The reaction mixture was allowed to stand at room temperature. Yellow crystals suitable for X-ray diffraction analysis formed in a day. They were filtered off, washed with cold water and ethanol and dried in air.Yield 58%. Tdecomp. = 186 ¡ÆC. Found (%): C, 37.1; H, 7.7; N, 14.3; Ni, 14.9; Cl, 8.5. Calc. for NiC12H31N4O4Cl (%): C, 37.0; H, 8.0; N, 14.4; Ni, 15.0; Cl, 9.1. NiL2 3. Benzene (50 ml) was added to 2 (0.4 g), and solid NaOH (0.3 g) and water (5 ml) were added in sequence with vigorous stirring. After 3 min, PbO2 (3 g) was added to the mixture. The colour of the benzene layer deepened to black green.The reaction mixture was additionally stirred for 1 h, the organic layer was separated and dried with CaCl2. Then, the mixture was filtered, and the filtrate was evaporated to dryness on a rotary evaporator. Yield 43%. When stored in normal conditions, the solid compound is stable. A toluene solution of 3 saturated at room temperature was allowed to stand overnight at .30 ¡ÆC to give single crystals suitable for X-ray diffraction analysis.Tdecomp. = 126.128 ¡ÆC. 1H NMR (C6D6) d: 0.81 (s, Me). The electronic absorption spectrum [EtOH, lmax/nm (e)]: 364 (18840), 560 (1920), 927 (6860). Found (%): C, 41.9; H, 7.2; N, 16.1. Calc. for NiC12H24N4O4 (%): C, 41.5; H, 7.0; N, 16.1. MS, m/z: 346.11474 (M+, calc. 346.11509). ¢Ô Cambridge Structural Database does not contain any information concerning metal complexes with vic-dihydroxyamines. NHOH NHOH + Ni(H2O)6Cl2 Na2CO3 H2O N N H H O O Ni N N H H O O H H H Cl PbO2 C6H6.H2O NaOH N N O O Ni N N O O 1 2 3 Scheme 1 ¡× Crystal data for 2: C12H31ClN4NiO4, M = 389.57, at 293 K crystals are orthorhombic, space group Pbca, a = 13.380(3), b = 12.825(3), c = 21.077(4) A, V = 3616.8(13) A3, Z = 8, dcalc = 1.431 g cm.3, m(MoK¥á) = = 1.242 mm.1, 2394 reflections were collected (2394 unique) on a Bruker AXS P4, (MoK¥á, graphite monochromator, q/2q scan, 1.93 < q < 24.91¡Æ, empirical absorption correction).The structure was solved by the program SIR97 and refined by the full-matrix least-square technique in an anisotropic approximation for all non-hydrogen atoms.Positions of all hydrogen atoms were located in a difference Fourier map and then refined in isotropic approximation. The final R indexes are R1 = 0.0767, wR2 = = 0.1084, for 2394 unique Ihkl > 2s(I), GOOF = 0.970. All calculations were carried out using SHELX97 program. Crystal data for 3: C12H24N4NiO4, M = 347.06, at 293 K crystals are monoclinic, space group P21/n, a = 6.7692(9), b = 9.853(2), c = = 11.583(2) A, b = 98.45(1)¡Æ, V = 764.2(2) A3, Z = 2, dcalc = 1.508 g cm.3, m(MoK¥á) = 1.291 mm.1, 1228 reflections were measured on a Bruker AXS P4 four-circle automated diffractometer (MoK¥á, graphite monochromator, q/2q scan, 3.68 < q < 24.96¡Æ).The structure was solved by the program SIR97 and refined by the full-matrix least-square technique in an anisotropic approximation for all non-hydrogen atoms.Positions of all hydrogen atoms were located in a difference Fourier map and then refined in an isotropic approximation. The final R indexes are R1 = 0.0316 and wR2 = 0.0480 for 1133 unique Ihkl > 2s(I), GOOF = 0.765. Atomic coordinates, bond lengths, bond angles and thermal parameters have been deposited at the Cambridge Crystallographic Data Centre (CCDC).For details, see ¡®Notice to Authors¡�, Mendeleev Commun., Issue 1, 2001. Any request to the CCDC for data should quote the full literature citation and the reference number 1135/87. H(4) H(8) H(3) C(11) H(5) H(7) C(12) C(1) H(14) H(6) C(22) H(13) H(12) C(2) H(10) H(11) H(9) C(21) N(2) O(2) H(16) O(3) H(18) N(3) N(1) H(2) O(1) H(1) Ni H(15) H(25) H(19) C(3) H(20) H(22) H(24) C(32) H(23) C(31) H(21) H(27) C(4) C(42) H(29) H(28) H(30) C(41) N(4) H(26) H(31) Cl O(4) H(32) Figure 1 Molecular structure of 2.Selected bond lengths (A): Ni.N(1) 1.897(3), Ni.N(3) 1.920(6), Ni.N(2) 1.931(6), Ni.N(4) 1.935(6), O(1).N(1) 1.449(8), N(2).O(2) 1.395(8), O(3).N(3) 1.457(8), N(4).O(4) 1.396(8), N.C 1.506(9).1.520(9), O(1)¡�¡�¡�O(4) 2.504(8), O(2)¡�¡�¡�O(3) 2.510(8); selected bond angles (¡Æ): N(1).Ni.N(2) 84.0(3), N(3).Ni.N(4) 84.7(3).Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) (e.g., in benzene, toluene, chloroform and ethanol). When stored for a few days, deep green solutions of 3 are gradually decolourised, and 3,3,4,4-tetramethyl-1,2-diazetine-1,2-dioxide precipitates. Noteworthy, the oxidation of free 1 by Pb4+ compounds leads to acetoxime.3 Formally, 3 may be regarded as a complex with the previously unknown nitrosohydroxyamine radical anion .O.¡�N.CMe2.CMe2.N=O (Scheme 1).Since 3 is diamagnetic, probably, because of very strong antiferromagnetic exchange interactions between the unpaired electrons of nitroxyl groups, the presence of a radical fragment may be detected using an approach developed for the metal complexes with ¥á-hydroxyaminooximes. 4 This approach consist in the elimination of a ligand from the coordination sphere. Indeed, the addition of an equimolar amount of dimethylglyoxime to a solution of 3 leads to a short-lived quintiplet with g = 2.006 and aN = 0.70 mT, which is typical of the EPR spectra of nitronylnitroxides. This indicates strong spin density delocalization in the .N.¡�NNi2+N=O fragment occurring via the metal ion and actually leading to the equivalence of N.O groups in 3.The e¡� of N.O groups is also indicated by the results of X-ray diffraction analysis of 3. In the molecule of 3, the square environment of the metal, which is responsible for the low-spin configuration of NiII, is formed by four N atoms (Figure 2).The Ni.N bond lengths are similar [1.827(3) and 1.839(3) A], also indicating strong delocalization and uniform electron density distribution in the {O.N.Ni.N.O} fragment. In contrast to 2, dimethylglyoximate complexes and metal bischelates with ¥á-hydroxyaminooximes,5 the molecular structure of 3 contains no intramolecular hydrogen bonds. Nevertheless, the coordination node is planar with much shorter Ni.N distances in 3, as compared to 2 or classical NiII dioximates.It is reasonable to assume that this is a consequence of strong delocalization and electron density conjugation in .O.¡�NNi2+.N=O fragments. In summary, the interaction of vic-dihydroxyamine 1 with NiII gives rise to bischelate 2, which contains two nonequivalent intramolecular H-bonds between the coordinated ligands.Compound 2 may be oxidised to unusual bischelate 3 whose structure contains the fragments of 1 with fully dehydrogenated hydroxyamine groups. Note that, the square-planar environment of the central atom was retained in dehydrogenated product 3, even though a molecule of 3 has no intramolecular H-bonds in contrast to 2.This work was supported in part by the U.S. Civilian and Development Foundation (grant no. REC-008) and the Russian Foundation for Basic Research (grant nos. 00-03-32987 and 00-03-04006). References 1 J. H. Osiecki and E. F. Ullman, J. Am. Chem. Soc., 1968, 90, 1078. 2 E. F. Ullman, J. H. Osiecki, D. G. B. Boocock and R. Darcy, J. Am.Chem. Soc., 1972, 94, 7049. 3 G. V. Shustov, N. B. Tavakalyan, L. L. Shustova, A. P. Pleshkova and R. G. Kostyanovskii, Izv. Akad. Nauk SSSR, Ser. Khim., 1982, 364 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1982, 31, 330). 4 V. N. Kirichenko, S. V. Larionov, I. A. Mikhailov, E. G. Boguslavskii and L. B. Volodarskii, Zh. Neorg. Khim., 1984, 29, 2835 (Russ. J. Inorg. Chem., 1984, 29, 1624). 5 E. O. Schlemper and R. K.Murmann, Inorg. Chem., 1983, 22, 1077. 6 V. I. Ovcharenko, S. V. Fokin, G. V. Romanenko, I. V. Korobkov and P. Rey, Izv. Akad. Nauk, Ser. Khim., 1999, 1539 (Russ. Chem. Bull., 1999, 48, 1519). O(2) N(2) Ni C(21) N(1) O(1) H(21B) H(21C) H(21A) C(2) C(1) H(12C) H(12A) H(12B) C(12) C(11) H(11A) H(11B) H(11C) C(22) H(22A) H(22B) H(22C) Figure 2 Molecular structure of 3. Selected bond lengths (A): Ni.N(1) 1.827(3), Ni.N(2) 1.839(3), N(1).O(1) 1.230(3), N(2).O(2) 1.218(3), N(1). C(1) 1.543(4), N(2).C(2) 1.531(4), O(1)¡�¡�¡�O(2') 2.574(3); selected bond angles (¡Æ): N(1).Ni.N(2) 83.9(1). Received: 12th April 2001; Co
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
3. |
Dissolution of uranium oxides in supercritical carbon dioxide containing tri-n-butyl phosphate and thenoyltrifluoroacetone |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 129-130
Trofim I. Trofimov,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) Dissolution of uranium oxides in supercritical carbon dioxide containing tri-n-butyl phosphate and thenoyltrifluoroacetone Trofim I. Trofimov,*a Maksim D. Samsonov,a Su C. Lee,b Boris F. Myasoedova and Chien M. Waib a V. I. Vernadsky Institute of Geochemistry and Analytical Chemistry, Russian Academy of Sciences, 117975 Moscow, Russian Federation.Fax: +7 095 938 2054 b Department of Chemistry, University of Idaho, Moscow, ID 83844, USA. Fax: +1 208 885 6173 10.1070/MC2001v011n04ABEH001468 Milligram amounts of uranium dioxide can be quantitatively dissolved in supercritical carbon dioxide containing a complex of tri-n-butyl phosphate (TBP) with nitric acid and separated from thorium(IV). The quantitative dissolution of milligram amounts of solid uranium trioxide in supercritical carbon dioxide containing thenoyltrifluoroacetone (TTA) and TBP was performed using ultrasonication.The separation of uranium(VI) and cerium(IV) in the test system was demonstrated. Supercritical fluid extraction (SFE) of metals from liquid and solid materials using environmentally friendly supercritical carbon dioxide containing a suitable ligand is a very promising method for chemical processes.1,2 SFE provides several advantages over conventional solvent extraction, including minimisation of hazardous liquid wastes.Since Laintz et al.3 demonstrated the possibility of copper chelate extraction with supercritical CO2, this method was applied to the extraction of actinides,4–7 caesium8 and strontium.9 Uranium and lanthanides can be directly extracted from their solid oxides using such ligands as TTA and TBP.4,10 Recent studies also demonstrated the possibility of lanthanide extraction from their oxides using supercritical CO2 containing the TBP–HNO3 complex.11 However, the extraction efficiency of all elements did not exceed 50%.Thus, it was of interest to develop a procedure for the quantitative extraction of uranium from its oxides simultaneously with its separation from other metals.Such a procedure can be used in a nuclear fuel cycle. Figure 1 shows a schematic diagram of the set-up employed for SFE. A syringe pump was used to deliver liquid CO2 through a pre-heating coil to the extraction system placed in a chromatographic oven for heating the system to a required temperature.Metal oxides (UO3, UO2, U3O8, CeO2, La2O3 and ThO2) were placed in a 3.5 ml extraction cell. The saturation of supercritical CO2 with ligands (TTA or the TBP–HNO3 complex) was performed in a 10.4 ml ligand cell connected upstream of the extraction cell. Pure TBP was injected into the system through a T-end joint using an HPLC pump.A flow rate of TBP injected into the system was about 0.02 ml min–1, which corresponded to its concentration in supercritical CO2 of about 5 vol.%. Extracted metal complexes were collected in a trap solution (chloroform) through the restrictors made of a capillary tube of deactivated fused-silica 25 cm in length and 50 µm in internal diameter. Uranium was back extracted from the trap solutions with 50% nitric acid in the case of TTA and TBP as the ligands or with 0.1 M (NH4)2CO3 in the case of the TBP–HNO3 complex.Uranium was determined by spectrophotometry with Arsenazo I.12 The data obtained were consistent with the ICP-MS data to within 10%. Cerium and thorium were determined by only ICP-MS. The mass balance on uranium in all runs was close to 100±7%.The extraction was performed in static, dynamic and combined modes (a static mode followed by a dynamic mode). The time of the static mode was about 10 min in all runs, and that of the dynamic one was changed from 15 to 60 min. The majority of the extractions were carried out at 60 °C and 150 atm. The above SFE conditions were found previously10 to be optimal for the UO3– TTA–TBP system.The mechanism of processes in the test system is given below: It is well known that ultrasonication can accelerate heterogeneous processes. We used an ultrasonic cleaner with a heater (model FS30, Fisher Scientific, USA) with a working frequency of 44–48 kHz (Figure 1). Table 1 shows the effect of ultrasonication on the SFE of uranium from its oxides with supercritical Figure 1 Schematic diagram of the experimental system for the dissolution of uranium oxides in supercritical carbon doxide: (1) CO2 cylinder; (2) syringe pump; (3) oven; (4) HPLC pump; (5) test-tube with TBP; (6-1)–(6-6) valves; (7) collection system; (8) restrictor; (9) fluid preheating coil; (10) extraction vessel; (11) ligand cell; (12) restrictor heater; (13) ultrasonic cleaner; (14) T-joint and (15) filter. 1 2 3 4 5 6-1 6-2 6-3 6-4 6-5 6-6 7 8 9 10 11 12 13 14 15 CO2 Table 1 Dissolution of uranium oxides in supercritical CO2 containing TTA and TBP using ultrasonication. Oxide U added/mg U trapped/mg (%) UO3 22.9 8.5±0.3 (37.0a) aWithout ultrasonication. UO3 25.2 24.0±0.6 (96.5) U3O8 30.9 0.7±0.1 (2.3) UO2 23.9 1.2±0.2 (5.0) Table 2 Separation of UVI and CeIV using supercritical CO2 containing TTA and TBP under ultrasonication. Oxide Metal added/mg Metal trapped/mg (%) UO3 23.0 21.3±0.4 (92.7) CeO2 13.6 0.4±0.1 (3.1) Table 3 Extraction of U, Th and La from the their oxides using supercritical CO2 containing the TBP–HNO3 complex.Oxide Metal added Metal trapped/mg (%) UO2 22.6 21.7±0.4 (96) UO2 51.6 51.6±0.6 (100) ThO2 25.1 0.2±0.1 (<0.1) La2O3 25.1 25.1±0.5 (100) Mass transport of TTA and TBP dissolved in UO3(solid) + 2TTA(sf) ® UO2(TTA)2·H2O(solid) UO2(TTA)2·H2O(solid) + TBP(sf) ® UO2(TTA)2·TBP(solid) + H2O(sf) UO2(TTA)2·TBP(solid) + supercritical CO2 ® UO2(TTA)2·TBP(sf) Mass transport of UO2(TTA)2·TBP(sf) in supercritical CO2 supercritical CO2 to UO3 reaction site from the extraction cell (1) (2) (3) (4) (5)Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) CO2 containing TTA and TBP.As can be seen, ultrasound allows uranium to be quantitatively extracted from UO3 with the above system. Unfortunately, ultrasonication did not improve the dissolution of both UO2 and U3O8 in supercritical CO2 containing TTA and TBP. This fact may be explained by structural differences between UO2, U3O8 and UO3 (UO2, face-centered; U3O8, orthorhombic; UO3, octahedral), as well as by steric hindrances in the UIV complexation with TTA and TBP.The effect of ultrasound can be attributed to the cleaning of the oxide surface by removing the complex formed. As a result, the reaction with TTA takes place more effectively. Table 2 shows that the suggested system can be successfully applied to the separation of uranium and cerium in the SFE from a mixture of UO3 and CeO2.The molar ratio [U]:[Ce] in the trap solution is about 30 times higher than that in the starting mixture. The effect of ultrasound was maximum in UO3, where uranium is in an oxidation state of 6+. We attempted to dissolve UO2 and U3O8 after a preliminary treatment with H2O2 directly in the extraction cell.For this purpose, 0.2 ml of 30% H2O2 was introduced into the extraction vessel containing either UO2 or U3O8. The mixture was heated at 90 °C for 2 h and evacuated to oxidise uranium to UVI and to evaporate the aqueous phase before SFE. About 50% UO2 was dissolved under the above conditions. However, we failed to obtain a positive result for U3O8.Apparently, conditions for the quantitative dissolution of UO2 may be found in the further investigation of the system, which may be used for uranium extraction from spent nuclear fuel, which is known to consist mainly of UO2. The system based on supercritical CO2 containing the TBP– HNO3 complex was successfully applied to the quantitative extraction of uranium from UO2.This system was recently used for Nd and Gd extraction from Nd2O3 and Gd2O3.11 To obtain the TBP–HNO3 complex, a 100% TBP solution was treated with concentrated HNO3. After separating phases by centrifuging, 2 ml of the complex obtained were placed in the ligand cell, where supercritical CO2 was saturated with the TBP–HNO3 complex for 20 min. Then, it was introduced into the extraction cell containing the oxides. SFE was conducted using the combined mode as described above for the TTA–TBP system. The freshly prepared TBP–HNO3 complex was used in all runs.As can be seen in Figure 2, uranium was quantitatively extracted from UO2. Table 3 shows the SFE of uranium, lanthanum and thorium from their oxides with supercritical CO2 containing TBP–HNO3 complex.The order of metal extraction from their oxides is UO2 » La2O3 >> ThO2; hence, uranium and thorium, as well as lanthanum and thorium, may be easily separated in the extraction from a mixture of their oxides (UO2 and ThO2, La2O3 and ThO2) with supercritical CO2 containing the TBP– HNO3 complex. Thus, milligram amounts of uranium can be quantitatively extracted from UO2 with supercritical CO2 containing the TBP– HNO3 complex, as well as from UO3 with supercritical CO2 containing TTA and TBP under ultrasonication. The enhanced dissolution of UO3 by means of ultrasound is, probably, caused by removing the UO2(TTA)2·H2O complex from the oxide surface, hence facilitating the complexation process.The results demonstrate that SFE is very promising for the reprocessing of spent nuclear fuel.This work was supported by British Nuclear Fuel Ltd. (BNFL), contract no. A80153. References 1 C. M. Wai, F. Hunt, M. Ji and X. Chen, J. Chem. Educ., 1998, 75, 1641. 2 N. G. Smart, C.M.Wai and C. L. Phelps, Chemistry in Britain, 1998, 34 (8), 34. 3 K. E. Laintz, C. M. Wai, C. R. Yonker and R. D. Smith, Anal. Chem., 1992, 64, 2875. 4 C. L. Phelps, N. G. Smart and C.M.Wai, J. Chem. Educ., 1996, 73, 1163. 5 Yuche Lin, N. G. Smart and C. M. Wai, Trends Anal. Chem., 1995, 14, 123. 6 Yuche Lin, N. G. Smart and C. M. Wai, Environ. Sci. Technol., 1995, 29, 2706. 7 Yuche Lin, C. M. Wai, F. M. Jean and R. D. Brauer, Environ. Sci. Technol., 1994, 28, 1190. 8 C. M. Wai, Y. M. Kulyako and B. F. Myasoedov, Mendeleev Commun., 1999, 180. 9 C. M. Wai, Y. M. Kulyako, H.-K. Yak, X. Chen and S.-J. Lee, Chem. Commun., 1999, 2533. 10 C. L. Phelps, PhD Thesis, University of Idaho, Department of Chemistry, 1997. 11 O. Tomioka, Y. Enokida and I. Yamamoto, J. Nucl. Sci. Technol., 1998, 35, 515 12 J. S. Fritz and M. Johnson-Richard, Anal. Chim. Acta, 1959, 20, 164. 100 80 60 40 20 0 15 30 45 60 t/min U extracted (%) Figure 2 SFE of uranium from UO2 with supercritical CO2 containing the TBP–HNO3 complex. Received: 27th April 2001; Com. 01/1794
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
4. |
The electronic structure, chemical bonding and ionic conductivity of Li6MoN4and Li6WN4 |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 131-132
Veronika M. Zainullina,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) The electronic structure, chemical bonding and ionic conductivity of Li6MoN4 and Li6WN4 Veronika M. Zainullina,*a Vladlen P. Zhukova and Vladlen H. Tammb a Institute of Solid State Chemistry, Urals Branch of the Russian Academy of Sciences, 620219 Ekaterinburg, Russian Federation. Fax: +7 3432 74 4495, e-mail: Veronika@ihim.uran.ru b Institute of the High-Temperature Electrochemistry, Urals Branch of the Russian Academy of Sciences, 620219 Ekaterinburg, Russian Federation 10.1070/MC2001v011n04ABEH001447 The calculations of the electronic structure, chemical bonding and energy of transition for lithium ions in Li6MoN4 and Li6WN4 have been carried out using the LMTO method; a possible model of lithium transport for these crystals has been suggested.The binary nitrides of lithium and transition metals (Cr, Mo and W) possess a high ionic conductivity and stability against melted lithium. Therefore, they are promising materials for lithium cells. However, the transport properties of these compounds have been studied insufficiently. For a better understanding of ionic conductivity, we evaluated the electronic structure and chemical bonding for Li6MoN4 and Li6WN4 by the first-principle linear muffin-tin orbital method in the tight-binding representation (LMTO-TB)1 and the semiempirical extended Hückel method (EH).2 The results of calculations have been used to analyse the mechanism of lithium ion migration in the anti-fluorite structure of these nitrides.The calculations of the electronic structure and total energy of Li6MoN4 and Li6WN4 crystals have been carried out by the LMTOTB method. We have employed a tetragonal unit cell (space group P42/nmc, Z = 2) with 20 atoms per cell: Li12M2N6E10, where M = Mo, W; E are empty spheres. The LMTO-ASA method has a higher precision for closely packed crystals; therefore, in our calculations additional spheres (empty spheres) with an s-orbital basis have been introduced into interstitial positions.These spheres were located at the octahedral and tetrahedral interstitial positions. The experimental lattice constants a = 6.673 and 6.679 Å, c = 4.925 and 4.927 Å for Li6MoN4 and Li6WN4, respectively, have been used.3 The optimised lattice constant for pure Li6MoN4 was 4.1% bigger than the experimental value.The valence 2s-, 2p-states of lithium and nitrogen; the ns-, np-, (n – 1)d-states of Mo, W, with n = 5, 6 and the s-states of empty spheres E were included in the atomic orbital basis used to construct the Bloch functions of crystals. The calculations were fulfilled for 128 k-vectors in the first Brillouin zone (30 k-vectors in the irreducible wedge).We found that the electronic structures of Li6MoN4 and Li6WN4 are close to each other. The total and partial densities of states for a Li6MoN4 phase are presented in Figure 1. The separation of the electronic energy spectrum into four zones is observed. A low-energy band A at about –15.0 eV was attributed to the 2s-states of nitrogen. The next band B is a band of hybrid N 2p and Mo 4d-states with some contributions of Li 2s-, 2p-states.The calculations of the overlap population of crystalline orbitals for the Mo–N bond showed that the antibonding partner of the band B is located at the bottom of the conductivity zone, which is a band with an energy of ~1.6 eV. The valence states in the range from –5.2 to –2.6 eV (band C) are hybridised N 2p and 2s-, 2p-states of Li, with the admixtures of Mo 4d-states.The presence of a forbidden gap confirms a semiconducting character of conductivity detected experimentally for similar phases.4 The main characteristics of the electronic energy spectrum for Li6MoN4 and Li6WN4 phases are presented in Table 1. The ab initio band calculations of the electronic structure for binary lithium nitrides were not studied previously.However, these results agree in the main features with the band structure for LiN3 obtained by Blaha and co-authors5 using a linearised augmented plane wave (LAPW) method. There are differences in the positions of the N-2s and N-2p bands and in their widths. The band appears, which consists of hybrid N 2p and Mo 4d-states. We also calculated the indices of chemical bonding for the considered compounds using the extended Hückel approach (Table 1).The averaged overlap population of crystalline orbitals is rather high for the Mo–N and W–N bonds. The Li–N interactions are characterised by a high degree of ionicity and an insignificant contribution of covalency. The low covalency of the Li–N bond corresponds to the high lithium mobility in the anti-fluorite structure.The investigations of the energy of defect formation using the ab initio LMTO calculations allowed us to offer a model of 30 20 10 20 10 20 10 20 10 0 –15 –10 –5 0 5 E/eV A B CE F total 2s,2p-N 2s,2p-Li 4d-Mo N(E)/eV–1 Figure 1 The (a) total and (b)–(d) partial densities of states [N(E)] for a Li6MoN4 crystal. (a) (b) (c) (d) Table 1 The characteristics of the electronic structure and chemical bonding in the crystals of Li6MoN4 and Li6WN4.Characteristic Li6MoN4 Li6WN4 Width and centre of band A/eV 1.45; –14.62 1.79; –14.80 Width and centre of band B/eV 1.21; –6.07 1.68; –5.79 Width and centre of band C/eV 2.66; –3.95 2.30; –3.63 Width of a forbidden gap/eV 2.87 3.42 Averaged bond populations by Mulliken M–N, M=Mo, W Li–N 0.837 0.047 0.849 0.046 O T N Li2 Li1 Mo Li1 Li2 Li2 Li2 Figure 2 The structure of a perfect crystal of Li6MoN4.Arrows show possible transitions for lithium ions into octahedral interstitial position (O) and through structure tetravacancies (T).Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) the ionic transport in Li6MoN4 and Li6WN4.The transport of the lithium ions is possible through the octahedral and tetrahedral interstitial positions in the anti-fluorite structure of these phases. The scheme of possible transport of the two crystallographically nonequivalent lithium atoms (Li1, Li2) is given in Figure 2. It is known3 that the lithium atoms Li1 have no tetravacancies in the nearest environment. Therefore, the Frenkel defect formation energy was calculated for the transport of Li2 into a tetraposition only.This energy was determined as the difference between the total energy of a perfect crystal (when a lithium atom is in a normal position) and the energy of a defect crystal (the lithium atom is located in octahedral or tetrahedral interstitial positions). The calculated energies are presented in Table 2.They show that the jump of lithium ions into a tetrahedral position requires an energy smaller by about 1.3 eV, than the jump of a lithium ion into octahedral positions. For both phases, the energy of transition of Li2 from a normal position into an octahedral position is a little lower than that of Li1. These energies, averaged over Li1 and Li2 atoms, within the limit of the errors of the method (0.1 eV) are practically identical for the Li6MoN4 and Li6WN4 crystals. Thus, the ab initio calculations of the electronic structure and energy of transition allow us to suggest a possible mechanism of lithium transport for the above crystals.The migration of lithium ions is more probable through the tetrahedral position, whereas the migration through the octahedral interstitial positions should be excluded.References 1 (a) O.-K. Andersen and O. Jepsen, Phys. Rev. Lett., 1984, 53, 2571; (b) O.-K. Andersen, Z. Pawlowska and O. Jepsen, Phys. Rev. B, 1986, 34, 5253. 2 M.-H. Whangbo and R. Hoffmann, J. Am. Chem. Soc., 1978, 100, 6093. 3 A. Gudat, S. Haag, R. Kniep and A. Rabenau, Z. Naturforsch., 1990, 45b, 111. 4 N. N. Batalov, O. V. Zheltonozhko, S. N. Zarembo, T. M. Akhmetzyanov, O. V. Volkova, G. V. Zelutin, V. P. Obrosov and V. K. Tamm, Elektrokhimiya, 1995, 31, 394 (Russ. J. Electrochem., 1995, 31, 356). 5 P. Blaha, J. Redinger and K. Schwarz, Z. Phys. B. Condenced Matter, 1984, 57, 273. Table 2 The energy of transition for Li1 (.Eoct1) and Li2 (.Eoct2) into an octahedral position and that for Li2 into a tetrahedral interstitial position (.Etetr2) in Li6MoN4 and Li6WN4 crystals. Crystal .Eoct1/eV .Eoct2/eV .Eaverage .Etetr2/eV Li6MoN4 4.77 4.47 4.62 3.16 Li6WN4 4.99 4.21 4.60 3.37 Received: 5th March 2001; Com. 01/1773
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
5. |
Low-energy barrier B4ring puckering rearrangement of 1,6-diaza-closo-hexaborane: anab initiostudy |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 132-134
Ruslan M. Minyaev,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) Low-energy barrier B4 ring puckering rearrangement of 1,6-diaza-closo-hexaborane: an ab initio study Ruslan M. Minyaev,* Vladimir I. Minkin, Tatyana N. Gribanova and Andrei G. Starikov Institute of Physical and Organic Chemistry, Rostov State University, 344090 Rostov-on-Don, Russian Federation. Fax: +7 8632 434 5667; e-mail: minyaev@ipoc.rsu.ru 10.1070/MC2001v011n04ABEH001475 The ab initio [MP2(fu)/6-311+G**] and DFT (B3LYP/6-311+G**) calculations predict stable structures of closo-diazaboranes 1,6-N2B4H4 and 1,2-N2B4H4, with the low-energy barrier B4 ring puckering rearrangement occurring in the 1,6-N2B4H4 stable structure. According to both experimental1.3 and computational data,4,5 closo-dicarborane 1,6-C2B4H6 1, which is isoelectronic to closoborane B6H6 2. 2, has a stable D4h-symmetry structure and is energetically preferable than its 1,2-isomer 3. It may be expected that similar stable structures are also characteristic of diaza-closo-boranes 1,6-N2B4H4 4 and 1,2-N2B4H4 5 isoelectronic to 1 and 3, respectively, and that 4 is more stable than its isomer 5. Indeed, early preliminary PRDDO calculations6 on N2B4H4 showed structure 4 to be more stable than 5, although the distorted trigonal prism to be predicted the most stable structure. More recent ab initio calculations7,8 also showed that 1,6-isomer 4 is more stable than 1,2-5.However, in both cases it was found that structure 4 does not correspond to a minimum on the N2B4H4 potential-energy surface (PES) and it was not studied the distortion directions from the D4h structure of 4. In this work, we performed ab initio [MP2(fu)/6-311+G**] and density functional theory (B3LYP/6-311+G**) calculations9,10 on compounds 4 and 5.For comparison, we also calculated the structures of closo-dicarboranes 1 and 3 at the same level of approximation. In agreement with published data,7,8 our ab initio calculations revealed that the structure of 4 of D4h symmetry corresponds to a saddle point rather than a minimum on the PES N2B4H4 and is the transition state for the low-energy barrier of the B4 ring puckering rearrangement 6a 4 6b. At the same time, 1,2- isomer 5, much like as its isoelectronic analogue 3, has a stable structure of C2v symmetry and is energetically less favourable than 1,6-isomer 6.According to the calculations, the structures of 1, 3 and 5, 6 correspond to minima (l = 0; hereafter, l designates the number of negative hessian eigenvalues) on the PESs of C2B4H6 and N2B4H4, respectively. The calculated geometric and energy parameters of these structures and the saddle point for the structure of 4 are depicted in Figures 1 and 2 and listed in Table 1. As can be seen in Table 1 and Figure 1, the calculated geometric characteristics of closo-dicarboranes 1 and 3 are in good agreement with the gas-phase experimental data1.3 and those obtained in previous theoretical studies.4,5 All calculated bond lengths are in BH BH CH BH HB CH BH BH BH BH HB BH 2.CH BH BH CH HB BH 1, D4h 2, Oh 3, C2v BH BH N BH HB N 4, D4h N BH BH N HB BH 5, C2v 1.627 1.624 1.635¡¾0.004 1.633¡¾0.004 1.716 1.712 1.725¡¾0.012 1.720¡¾0.004 1.735 1.733 1.752¡¾0.005 1.716 1.706 1.721¡¾0.015 1.624 1.629 1.605¡¾0.005 1.540 1.543 1.540¡¾0.005 1, D4h 3, C2v MP2 DFT Experimental } Figure 1 Geometry parameters of structures 1 and 3 calculated by ab initio (MP2/6-311+G**) and DFT (B3LYP/6-311+G**) methods.Experimental data for 1 are taken from ref. 1 (upper numbers) and from ref. 2 (lower numbers) and for 3 from ref. 3. The bond lengths and angles are indicated in angstrom units and degrees, respectively. C C C C Table 1 Results of ab initio [MP2(fu)/6-311+G**] and DFT (B3LYP/6- 311+G**) calculations for the structures of 1, 3.7.a aEtot (in a.u.) and .E are the total and relative energies (1 a.u. = 627.5095 kcal mol.1); l is the number of the negative hessian eigenvalues; ZPE (in a.u.) is the harmonic zero-point correction; .EZPE (in kcal mol.1) is the relative energy including harmonic zero-point correction; w1 (in cm.1) is the smallest or imaginary harmonic vibration frequency. bResults correspond to a slope point with 5 A distance from N2 to the BB bond.Structure Method Etot l .E ZPE .EZPE w1 1, D4h MP2 DFT .178.784605 .179.284851 00 00 0.086648 0.086128 00 421 380 3, C2v MP2 DFT .178.769941 .179.270983 00 9.2 8.7 0.086734 0.086115 9.2 8.7 433 395 4, D4h MP2 DFT .210.822065 .211.348393 11 1.0 4.2 0.060956 0.060449 1.3 4.6 i184 i268 5, C2v MP2 DFT .210.804263 .211.333654 00 12.2 13.5 0.061988 0.061182 13.1 14.3 314 253 6, D2d MP2 DFT .210.823731 .211.355047 00 00 0.060607 0.059728 00 231 300 7, Cs MP2 DFTb .210.757200 .211.294522 0 slope 41.7 37.9 0.056079 . 38.9 . 17 . B4H4, Td MP2 DFT .101.584553 .101.409193 00 .. 0.050657 0.049508 .. 617 609 N2, D¡Íh MP2 DFT .109.346230 .109.559694 00 .. 0.005570 0.004906 .. 2445 2151 BH BH N BH HB N 4, D4h BH BH N BH HB N 6a, D2d BH BH N BH HB N 6b, D2dMendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) the range of the experimental values accounted for experimental errors. 1,6-Dicarbo-closo-hexaborane 1, 1,6-C2B4H6 was found to be more stable than 1,2-isomer 3 by 9.2 kcal mol.1 at the MP2 level and by 8.7 kcal mol.1 at the DFT level. These values are consistent with the previous estimation (9.5 kcal mol.1) obtained at the MP2/6-31G** level.4 No experimental data on the heats of formation of 1 and 3 are currently available.The stable structure of 1,6-diaza-closo-hexaborane 6 has D2d symmetry with two short (MP2, 1.500 and DFT, 1.454 A) and two long (MP2, 1.750 and DFT, 1.807 A) BN bonds. The basal B4 ring has a boat conformation; the B.B bond lengths are equal to 1.699 (MP2) and 1.454 (DFT) A. This value is very close to those of the basal B.B bonds in 1 and 3.Planarization of the B4 basal cycle, 6 ¢ç 4, results in equalization of all the BN bonds and shortening of the B.B bonds. The structure of 4 is the true transition state structure (l = 1, this identification of stationary point agrees with the result of Jemmis and Subramanian8 but disagrees with McKee¡�s7 results l = 3) for the puckering rearrangement 6a 4 6b with the energy barrier as low as 1.0 (MP2) or 4.2 (DFT) kcal mol.1.Accounting for zero-point energy (ZPE) does almost not change the energy barrier. The tendency of the D4h structure of 1,6-diaza-closo-hexaborane 4 to the D4h ¢ç D2d distortion is explained by the orbital interaction diagram (Figure 3), which shows that this distortion leads to slightly lowering the energy level of the bonding 1e orbitals of the D2d cluster.Although the D4h structure 4 satisfies to the 10e electrons rule formulated for the stable bipyramidal structures of main-group element clusters,11 the orbital interaction providing for the stabilization of structures of this type, namely, mixing in the antibonding combination of p-orbitals of apical centers and eg orbitals of the basal cycle, is weakened in 4, as compared to that in its carbon analogue 1.This is due to a widened energy gap between these orbitals in 4 caused by a greater electronegativity of nitrogen, which also results in less diffuse p-orbitals and their smaller overlap with eg orbitals of the basal cycle. As congeneric 1,2-dicarbo-closo-hexaborane 3, 1,2-diaza-closohexaborane 5 has a stable C2v structure with a planar basal boron ring.It contains BN bonds of two types: short [1.550 (MP2), 1.536 (DFT) A] and long [1.639 (MP2), 1.632 (DFT) A]. This diazaborane is by 12.2 (MP2) and 13.5 kcal mol.1 (DFT) energy disfavoured as compared to 1,6-isomer 6. Note that whereas for dicarboranes 1 and 3 and 1,2-diaza-closo-hexaborane 5 the results of MP2 calculations are consistent with those of the DFT method, for 1,6-diaza-closo-hexaborane 6 the bond lengths predicted by MP2 and DFT methods notably differ (~0.05 A).The system 5 can be considered as a tight complex resulted from the interaction of dinitrogen with borane B4H4. In this context, a question arises whether N2 and B4H4 can form a stable pre-reaction colex subsequently convertible to 5.No such a complex has been found by DFT calculations: the interaction between N2 and B4H4 was repulsive at any distances. This finding is consistent with the conclusion that DFT methods do not correctly describe longrange interactions.13 At the same time, MP2 calculations predict the appearance of stable complex 7 stabilised by induced dipole. dipole interactions between its components.The complex is 1.1 kcal mol.1 stabilised relative to separated components (no account is done for the superposition error). Such a weak interaction does not affect the geometric parameters of N2 and B4H4 moieties in complex 7 as compared to separated molecules. Complex 7 is 41.7 (at MP2 level) or 37.9 kcal mol.1 (at DFT level) less stable than 1,2-isomer 5. In conclusion, the MP2 and DFT calculations on hypothetical diaza-closo-boranes 5 and 6 indicate that these compounds, which are isoelectronic to dicarbo-closo-hexaboranes 1 and 3, respectively, possess stable highly symmetric structures.Compound 6 was predicted to be susceptible to undergo the lowenergy barrier B4 ring puckering rearrangement 6a 4 6b. This work was supported by the Russian Foundation for Basic Research (grant nos. 01-03-32546 and 00-15-97320). References 1 V. S. Mastryukov, O. V. Dorofeeva, L. V. Vilkov, A. F. Zhigach, V. T. Laptev and A. B. Petrunin, J. Chem. Soc., Chem. Commun., 1973, 276. 2 E. A. McNeill, K. L. Gallaher, F. R. Scholer and S. H. Bauer, Inorg. Chem., 1973, 12, 2108. 3 R. A. Beadet and R. L. Poyntner, J. Chem. Phys., 1970, 53, 1899. 4 M.L. McKee, J. Am. Chem. Soc., 1992, 114, 879. 5 M.Buhl and P. von R. Schleyer, J. Am. Chem. Soc., 1992, 114, 477. 6 T. A. Halgren, I. M. Pepperber and W. N. Lipscomb, J. Am. Chem. Soc., 1975, 97, 1249. 7 M. J. McKee, J. Phys. Chem., 1991, 95, 9273. 8 E. D. Jemmis and G. Subramanian, J. Phys. Chem., 1994, 98, 9222. 9 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A.Robb, J. R. Cheeseman, V. G. Zakrzewski, J. A. Montgomery, Jr., R. E. Stratmann, J. C. Burant, S. Dapprich, J. M. Millam, A. D. Daniels, K. N. Kudin, M. C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B. Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G. A. Petersson, P. Y. Ayala, Q. Cui, K.Morokuma, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J.Cioslowski, J. V. Ortiz, A. G. Baboul, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. Gomperts, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M.W.Wong, J. L. Andres, C. Gonzalez, M. Head-Gordon, E. S. Replogle and J. A. Pople, Gaussian 98, Revision A.9, Gaussian, Inc., Pittsburgh PA, 1998.MP2 DFT 1.500 1.454 1.750 1.807 1.699 1.737 N N 6, D2d (l = 0) 4, D4h (l = 1) 5, C2v (l = 0) 7, Cs (l = 0) 1.609 1.598 1.663 1.658 1.637 1.619 1.773 1.785 1.550 1.536 1.668 1.663 N N 1.639 1.632 1.692 1.679 1.120 1.096 164.6 3.426 ¡Í 1.691 1.680 1.7(00) 1.119 1.096 1.097632(20) 3.426 ¡Í MP2 DFT Experimental Td D¡Íh Figure 2 Geometry parameters of structures 5.7 and borane B4H4 and dinitrogen calculated by ab initio (MP2/6-311+G**) and DFT (B3LYP/ 6-311+G**) methods.Experimental data for B4H4 are given for B4Cl4 12 and for N2 are taken from ref. 14. The bond lengths and angles are indicated in angstrom units and degrees, respectively. N N N N N N 3 2 .14 .15 2eg 2e eg 1eg 1e ¥�g D4h D2d Figure 3 Diagram of formation of bonding molecular orbitals in 4 and 6. E/eVMendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) 10 M.W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. J. Su, T. L. Windus, M. Dupuis and J. A. Montgomery, J. Comput. Chem., 1993, 14, 1347. 11 V. I. Minkin, R. M. Minyaev and Yu. A. Zhdanov, Nonclassical Structures of Organic Compounds, Mir, Moscow, 1987. 12 J. A. Morrison, Chem. Rev., 1991, 91, 35. 13 K. Müller-Dethiefs and P. Hobza, Chem. Rev., 2000, 100, 143. 14 R. J. Butcher and W. J. Jones, J. Chem. Soc., Faraday Trans. 2, 1974, 560. Received: 10th May 2001; Com. 01/18
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
6. |
Crystal properties ofN-alkyl-substituted glycolurils as the precursors of chiral drugs |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 134-136
Remir G. Kostyanovsky,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) Crystal properties of N-alkyl-substituted glycolurils as the precursors of chiral drugs Remir G. Kostyanovsky,*a Konstantin A. Lyssenko,b Angelina N. Kravchenko,c Oleg V. Lebedev,c Gul¡�nara K. Kadorkinaa and Vasilii R. Kostyanovskya a N. N. Semenov Institute of Chemical Physics, Russian Academy of Sciences, 117977 Moscow, Russian Federation.Fax: +7 095 938 2156; e-mail: kost@center.chph.ras.ru b A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 117813 Moscow, Russian Federation. Fax: +7 095 135 5085; e-mail: kostya@xray.ineos.ac.ru c N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 119991 Moscow, Russian Federation. Fax: +7 095 135 5328 10.1070/MC2001v011n04ABEH001469 2,6-Dimethylglycoluril 1 crystallises to form a conglomerate (space group P21) and co-crystallises with isomeric 2,8-dimethylglycoluril 2; 2,6-diethylglycoluril A is the best precursor in the synthesis of chiral drugs.The chemistry of glycolurils is progressing in different directions including the construction of self-assembling molecular entities such as clips, capsules and supramolecular coordination-bonded systems based on cucurbiturils,1,2 as well as the preparation of pharmaceuticals.3.5 The psychotropic drugs Mebicar and Albicar are well-known tranquilizers and antidepressants3.6 (Scheme 1).Albicar crystallises in a centrosymmetric space group (P21/a);4 therefore, its resolution into enantiomers is difficult.However, its potential precursor, chiral 2,6-diethylglycoluril A (which can be easily separated from achiral isomer B), forms a conglomerate (space group P41212)5 and thus smoothly undergoes spontaneous resolution.1 According to published data on the glycoluril structures, 1,4,5,9.11 there are two another examples of conglomerates, i.e., 2,6-dinitro-4,8-diacetylglycoluril (space group P212121)9 and compound 3 (space group P1).4 The latter can be considered as a synthetic drug precursor. In this work, we examined the spontaneous resolution of chiral glycolurils 1 and 3.Using a known method,12 the isomers of 1 and 2 were obtained in the ratio 1.8:1 (1H NMR data); they were purified by repeated crystallization12 and characterised.¢Ó An unambiguous structural 1H NMR test is the non-equivalency of the 1-CH and 5-CH protons in 2; the signal of the latter can be easily assigned by additional triplet splitting of the doublet on the HN-4 and HN-6 protons. Note that these protons are readily exchanged; therefore, the spin coupling constants 3J can be observed in only dry aprotic solvents.Difficulties in the resolution and purification12 of isomers 1 and 2 are caused by easy co-crystallization. The composition of a co-crystal grown from H2O under slow self-evaporation [1 (R)-1 + 1 (S)-1 + 2 2 + 5 H2O] was established by 1H NMR¢Ó and X-ray¢Ô methods (Figures 1 and 2).Note that co-crystallization does not occur in the case of a mixture of A with B. The crystallization of this mixture from H2O leads to the two groups of crystals5, namely, large tetragonal A and thin lamellate monoclinic B (space group P21/c).5 The separate crystallization from supersaturated aqueous solutions was also observed for the mixtures of 1 with B and 2 with A (2 days at 20 ¡ÆC) to form the crystals of pure B and 2, respectively (both compounds were identified by 1H NMR spectra).¢Ó By crystallization of pure isomer 1 from H2O under slow selfevaporation (3.5 days), the fine needle-shaped crystals suitable for an X-ray study¢Ô were grown (Figure 3). They exhibit a chiral space group similarly to the analogue A.1 Thus, the fourth conglomerate has been found in the series of glycolurils.4,5,9 The optical activity was detected in individual crystals of 1 (up to 1.5 mg).The positive Cotton effect at 202 nm was observed in the CD spectrum of (+)-1; by analogy with diethyl analog A,1 this fact permits assigning its absolute configuration R-(+)-1.Many attempts to resolve glycoluril 3 spontaneously have failed. The crystallization from ethyl acetate (under conditions of mono- N N O N R2 R1 N O R4 H H R3 1 2 3 4 5 6 7 8 Mebicar: R1 = R2 = R3 = R4 = Me Albicar: R1 = R3 = Me, R2 = R4 = Et A: R1 = R3 = Et, R2 = R4 = H B: R1 = R4 = Et, R2 = R3 = H 1: R1 = R3 = Me, R2 = R4 = H 2: R1 = R4 = Me, R2 = R3 = H 3: R1 = R2 = Me, R3 = Et, R4 = H Scheme 1 ¢Ó Characteristics and spectroscopic data. 1: mp 268.270 ¡ÆC (H2O). 1H NMR ([2H6]DMSO) d: 2.62 (s, 6H, 2Me), 5.05 (s, 2H, 2HC), 7.40 (br. s, 2H, 2HN); (CD3OD): 2.76 (s, 6H, 2MeN), 5.25 (s, 2H, 2HC). R-(+)-1: mp 335 ¡ÆC (charred), [a]578 +49.5¡Æ, [a]546 +55.0¡Æ (c 0.1, H2O).CD spectrum in H2O, cell 1 mm (c 3.3¡¿10.3 mol dm.3), .e (lmax/nm): +2.04 (202). 2: mp 298.300 ¡ÆC (H2O). 1H NMR ([2H6]DMSO) d: 2.78 (s, 6H, 2Me), 5.11 (d, 1H, 1-CH, 3J 8.2 Hz), 5.18 (dt, 1H, 5-CH, 3J 8.2 Hz, 3JHCNH 1.8 Hz), 7.3 (br. s, 2H, 2HN); (CD3OD): 2.93 (s, 6H, 2Me), 5.24 (d, 1H, HC, 3J 8.4 Hz), 5.34 (d, 1H, HC, 3J 8.4 Hz). 1 + 2, co-crystal: mp 254 ¡ÆC (H2O). 1H NMR ([2H6]DMSO) d: 2.60 (s, 6H, 2Me), 2.80 (s, 6H, 2Me), 5.10 (s, 2H, 2HC), 5.11 (d, 1H, 1-CH, 3J 8.2 Hz), 5.17 (br. s, 1H, 5-CH, 3J 8.2 Hz), 7.35 (br. s, 2H, 2HN), 7.50 (br. s, 2H, 2HN); (CD3OD): 2.76 (s, 6H, 2Me), 2.93 (s, 6H, 2Me), 5.24 (d, 1H, 1Hc), 5.25 (s, 2H, 2Hc), 5.34 (d, 1H, 1HC, 3J 8.4 Hz). 3: mp 148.149 ¡ÆC (AcOEt). 1H NMR ([2H6]DMSO) d: 1.13 (t, 3H, MeCH2, 3J 7.15 Hz), 2.68 (s, 3H, MeN), 2.82 (s, 3H, MeN), 3.25 (m, 2H, CH2N, ABX3 spectrum, .n 56.0 Hz, 2Jab .14.3 Hz, 3Jax = 3Jbx = = 7.15 Hz), 5.05 (dd, 1H, 1-CH, 3J1,5 8.2 Hz, 3J1,8 1.6 Hz), 5.16 (d, 1H, 5-CH, 3J 8.2 Hz), 7.5 (br.s, 1H, HN); (CDCl3): 1.17 (t, 3H, MeCH2, 3J 7.1 Hz), 2.79 (s, 3H, MeN), 2.94 (s, 3H, MeN), 3.34 (ddq, 2H, CH2N, ABX3 spectrum, .n 120.0 Hz, 3Jax = 3Jbx = 7.1 Hz, 2Jab .14.3 Hz), 5.00 (d, 1H, 1HC, 3J 8.4 Hz), 5.05 (d, 1H, 1HC, 3J 8.4 Hz), 7.23 (br.s, 1H, HN). 13C NMR (CDCl3) d: 13.0 (qt, MeCH2, 1J 127.0 Hz, 2J 3.1 Hz), 27.8 (q, MeN, 1J 138.4 Hz), 30.3 (q, MeN, 1J 138.0 Hz), 36.9 (tq, CH2N, 1J 138.0 Hz, 2J 4.5 Hz), 66.0 (d, 1J 167.0 Hz), 71.4 (d, CH, 1J 166.0 Hz), 158.4 (sept, 7-CO, 3J 3.0 Hz), 160.1 (br. s, 3-CO).B: mp 240 ¡ÆC (H2O). 1H NMR ([2H6]DMSO) d: 1.06 (t, 6H, 2Me, 3J 6.9 Hz), 3.20 (m, 4H, 2CH2N, ABX3 spectrum, .n ¡í 60 Hz, 2Jab .14.0 Hz, 3Jax = 3Jbx = 6.9 Hz), 5.17 (dt, 1H, 5-CH, 3J1,5 8.4 Hz, 3JHCNH 2.2 Hz), 5.36 (d, 1H, 1-CH, 3J 8.4 Hz), 7.38 (br. s, 2H, 2HN). Figure 1 Tetramers in a co-crystal of 1 + 2. The N¡�¡�¡�O distances are N(8)¡�¡�¡�O(1') 2.940(2) A, O(2)¡�¡�¡�N(4) 2.842(2) A, N(6')¡�¡�¡�O(2'A) 2.980(2) A, O(1'A)¡�¡�¡�N(8A) 2.955(2) A, N(4'A)¡�¡�¡�O(2A) 2.842(2) A.O(2WA) H(4N) C(10) N(6) N(4) C(3) O(1) H(2WA) H(2WB) O(2W) N(2) C(9) C(1) C(5) C(7) N(8) H(8N) O(2) H(1WA) O(1W) H(1WB) H(4'N) N(4') C(5') O(1') C(3') C(9') C(7') C(1') C(10') N(8') O(2') N(6') H(6'N) N(2') O(2'A) N(6'A) H(6'A) N(4'A) H(4'A) O(2A) O(1'A) H(8NA) N(8A)Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) crystal growth for X-ray diffraction study of 34) gives largesized crystals (up to 20.30 mg). However, no optical activity was detected for any individual crystal selected in several tens of our experiments. This can be explained by epitaxy phenomena (cf. refs. 13.18), and the fact that a piece of the crystal suitable for X-ray analysis was cut out from such a splice and used in the experiments.4 Many of 2-mono-R- (R = Me, Et, Prn) and 2,4,6-tri-R-substituted (R = Me, Et) glycolurils do not give wellformed crystals.Based on the results, we can state that the enantiomers of A1 obtained by spontaneous resolution are the precursors of choice for the synthesis of chiral drugs such as Albicar and its analogues.The basic geometric parameters of of isomer, and actually do not differ from those described earlier.1.4,10.12 The angle between the root-mean-square planes of five-membered rings in 1 and 1 + 2 varies in a range of 118.7.120.3¡Æ. Analysis of the crystal packing show that (similarly to A1) the molecules in 1 are combined with N.H¡�¡�¡�O=C bonds into a homochiral three-dimensional H-bonded framework formed by two helices of the molecules directed along the crystallographic axis a [H-bond N(4).H(4N)¡�¡�¡�O(1')] and along the axis b [H-bond N(8).H(8N)¡�¡�¡�O(2')] (Figure 3).It is interesting that in a molecule of 3,7-diazabicyclo[3.3.1]nonane-2,6-dione [which is similar to 1 in terms of potential donors and acceptors of protons, and for which the formation of a conglomerate (space group P212121) is also observed] the two perpendicular H-bonded helices combine bicycles into a homochiral layer.19 Like 1, achiral monohydrate 2 is crystallised in a chiral space group of P212121;10 however, the molecules of 2 are combined into a helix by an eight-membered H-ring rather than an H-bond, as it takes place in 1 and A.Thus, 1 and 2, which crystallise individually in chiral space groups, form a centrosymmetric cocrystal. The principal distinction of its crystal packing is the absence of an infinite N.H¡�¡�¡�O bonded structure (Figure 1). The basic structure unit in 1 + 2 is an H-bonded heterochiral tetramer, in which the enantiomers of 1 are not connected (Figure 2). Two solvate molecules of H2O in the structure 1 + 2 play different roles.The molecule O(1w) links additionally 1 and 2, whereas O(2w) combines tetramers into heterochiral zigzag chains similar to those observed, for example, in the racemic crystals of substituted diazabicyclo[3.3.1]nonanes.19 The associated solvate molecules of H2O combine these chains into a three-dimensional H-bonded framework (Figure 2).This work was supported by the Russian Foundation for Basic Research (grant nos. 98-03-04119, 00-03-81187 Bel and 00-03- 32738), INTAS (grant no. 99-00157) and the Russian Academy of Sciences. References 1 R. G. Kostyanovsky, K. A. Lyssenko, G. K. Kadorkina, O. V. Lebedev, A. N. Kravchenko and V. R. Kostyanovsky, Mendeleev Commun., 1998, 231. 2 K. E. Pryor and J.Rebek, Org. Lett., 1999, 1, 39. 3 O. V. Lebedev, L. I. Khmel¡�nitskii, L. V. Epishina, L. I. Suvorova, I. V. Zaikonnikova, I. E. Zimakova, S. V. Kirshin, A. M. Karpov, V. S. Chudnovskii, M. V. Povstyanoi and V. A. Eres¡�ko, in Tselenapravlennyi poisk novykh neirotropnykh preparatov (Purposeful Search for New Neurotropic Medicines), Zinatne, Riga, 1983, p. 81 (in Russian). 4 V. Z.Pletnev, I. Yu. Mikhailova, A. N. Sobolev, N. M. Galitskii, A. I. Verenich, L. I. Khmel¡�nitskii, O. V. Lebedev, A. N. Kravchenko and L. I. Suvorova, Bioorg. Khim., 1993, 19, 671 (Russ. J. Bioorg. Chem., 1993, 19, 371). 5 E. B. Shamuratov, A. S. Batsanov, Yu. T. Struchkov, A. Yu. Tsivadze, M. G. Tsintadze, L. I. Kmel¡�nitskii, Yu. A. Simonov, A. A. Dvorkin, O. V. Lebedev and T.B. Markova, Khim. Geterotsilk. Soedin., 1991, 937 [Chem. Heterocycl. Compd. (Engl. Transl.), 1991, 27, 745]. 6 I. V. Svitan¡�ko, I. L. Zyryanov, M. I. Kumskov, L. I. Khmel¡�nitskii, L. I. Suvorova, A. N. Kravchenko, T. B. Markova, O. L. Lebedev, G. A. Orekhova and S. V. Belova, Mendeleev Commun., 1995, 49. 7 A. N. Kravchenko, O. V. Lebedev and E. Yu. Maksareva, Mendeleev Commun., 2000, 27. 8 G. A. Gazieva, A. N. Kravchenko, K. Yu. Chegaev, Yu. A. Strelenko and O. V. Lebedev, Mendeleev Commun., 2000, 28. 9 J. Boileau, E. Wimmer, M. Pierrot, A. Baldy and R. Gallo, Acta Crystallogr., Sect. C., 1985, 41, 1680. 10 M. O. Dekaprilevich, L. I. Suvorova and L. I. Khmel¡�nitskii, Acta Crystallogr., Sect. C., 1994, 50, 2056. 11 N. Li, S. Maluendes, R. H. Blessing, M. Dupuis, G.R. Moss and G. T. DeTitta, J. Am. Chem. Soc., 1994, 116, 6494. 12 J. Nematolahi and R. Ketcham, J. Org. Chem., 1963, 28, 2378. ¢Ô Crystallographic data for 1 and 1 + 2: crystals of 1 (C6H10N4O2) are orthorhombic, space group P212121, Z = 4, a = 4.421(2) A, b = 7.950(3) A, c = 22.473(7) A, V = 789.8(5) A3, M = 170.18, dcalc = 1.431 g cm.3, m(MoK¥á) = 1.11 cm.1, F(000) = 360; crystals of 1 + 2 (C12H22N8O6) are triclinic, Z = 2, space group P1, a = 5.771(4) A, b = 11.114(7) A, c = = 14.053(9) A, a = 106.52(5)¡Æ, b= 98.51(5)¡Æ, g = 96.16(5)¡Æ, V = 843.9(9) A3, M = 376.39, dcalc = 1.481 g cm.3, m(MoK¥á) = 1.21 cm.1, F(000) = 360. Intensities of 1154 (for 1) and 4767 (for 1 + 2) reflections were measured on a Siemens P3 diffractometer [l(MoK¥á) = 0.71072 A, q/2q-scans, 2q < 56¡Æ] at 298 and 153 K, respectively; 1154 and 3971 independent reflections were used in the further refinement.The structures were solved by a direct method and refined by the full-matrix least-squares technique against F2 in the anisotropic.isotropic approximation. Hydrogen atoms were located from the Fourier synthesis and refined in the isotropic approximation.The riding model was used for the methyl hydrogen atoms in 1. The refinement converged to wR2 = 0.1029 and GOF = = 0.954 for all independent reflections [R1 = 0.0343 was calculated against F for 679 observed reflections with I > 2s(I)] for the structure of 1 and to wR2 = 0.1725 and GOF = 1.080 for all independent reflections [R1 = 0.0562 was calculated against F for 3484 observed reflections with I > 2s(I)] for the structure of 1 + 2.All calculations were performed using the SHELXTL PLUS 5.0 program. Atomic coordinates, bond lengths, bond angles and thermal parameters have been deposited at the Cambridge Crystallographic Data Centre (CCDC). For details, see ¡®Notice to Authors¡�, Mendeleev Commun., Issue 1, 2001. Any request to the CCDC for data should quote the full literature citation and the reference number 1135/91.Figure 2 Projection of the crystal structure of 1 + 2 on the crystallographic plane bc. 0a b c Figure 3 Three-dimensional H-bonded framework in the crystal structure of 1. The N¡�¡�¡�O distances are N(4)¡�¡�¡�O(1') 2.809(3) A, N(8)¡�¡�¡�O(2) 2.848(3) A. N(4) H(4N) O(1) H(4N) N(4) N(8) H(8N) O(2) C(7) H(8N) N(8) N(6) C(10) C(5) C(1) C(9) N(2) C(3) N(4) H(4N) O(1) O(1) H(4N) N(4) N(8) H(8N) O(2) O(2) N(8) H(8N) N(8) H(8N) O(2) O(2) N(8) H(8N)Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) 13 S. Furberg and O. Hassel, Acta Chem. Scand., 1950, 4, 1020. 14 B. S. Green and M. Knossow, Science, 1981, 214, 795. 15 R. J. Davey, S. N. Black, L. J. Williams, D. McEwan and D. E. Sadler, J. Cryst. Growth, 1990, 102, 97. 16 G. A. Potter, C. Garcia, R. McCague, B. Adger and A. Collet, Angew. Chem., Int. Ed. Engl., 1996, 35, 1666. 17 M. Berfeld, D. Zbaida, L. Leiserovitz and M. Lahav, Adv. Mater., 1999, 11, 328. 18 S. Beilles, P. Cardinael, E. Ndzie, S. Petit and G. Coquerel, J. Chem. Eng. Sci., 2001, 56, 2281. 19 R. G. Kostyanovsky, K. A. Lyssenko, Yu. I. El’natanov, O. N. Krutius, I. A. Bronzova, Yu. A. Strelenko and V. R. Kostyanovsky, Mendeleev C
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
7. |
Generation of the metallonium cations [1,2-(CH2)2C5Me3OsC5Me5]2+, [1,1'-(CH2C5Me4)2Os]2+and [1,2-(CH2)2C5Me3Os(1'-CH2C5Me4)]3+in the CF3SO3H-O2system |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 137-138
Margarita I. Rybinskaya,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) Generation of the metallonium cations [1,2-(CH2)2C5Me3OsC5Me5]2+, [1,1'-(CH2C5Me4)2Os]2+ and [1,2-(CH2)2C5Me3Os(1'-CH2C5Me4)]3+ in the CF3SO3H.O2 system Margarita I. Rybinskaya, Alla A. Kamyshova,* Arkadii Z. Kreindlin and Pavel V. Petrovskii A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 117813 Moscow, Russian Federation.Fax: +7 095 135 5085; e-mail: kreindlin@ineos.ac.ru 10.1070/MC2001v011n04ABEH001456 The title dications and trication were generated from decamethylosmocene and [C5Me5OsC5Me4CH2]+, respectively, by the interaction with dioxygen in CF3SO3H solutions. Previously,1.3 we found that strong protic acids can promote the oxidation of methyl substituents in the decamethylmetallocenes (C5Me5)2M (M = Ru or Os) under certain conditions. Decamethylruthenocene can be oxidised by molecular oxygen in the superacid CF3SO3H to form mono-, di- or even trications.In this work, we examined this reaction by the example of decamethylosmocene 1 and salt¢Ó [C5Me5OsC5Me4CH2]+BF4 . 2 as the product of its one-electron oxidation. The behaviour of compound 1 in CF3SO3H (98%, Fluka) in an inert atmosphere of argon was preliminarily studied by 1H NMR spectroscopy.It was found that, in contrast to (C5Me5)2Ru, which forms the monohydride [(C5Me5)2RuH]+ under these conditions, decamethylosmocene 1 affords dihydride [(C5Me5)2OsH2]2+ 3. The monohydride [(C5Me5)2OsH]+ is usually formed in weaker acids.1 In the case of monocation 2 in a CF3SO3H solution in an inert atmosphere, dication [C5Me5Os(H)C5Me4CH2]2+ 4 was formed as the product of monoprotonation.The reactions of 1 and 2 with oxygen in CF3SO3H solutions were performed in NMR tubes (0.7 ml) with bubbling oxygen¢Ô at regular intervals, as described previously3 for the reaction with (C5Me5)2Ru. The course of the reaction was monitored by 1H NMR spectroscopy.The formation of dihydride 3 (from 1) and protonated monocation 4 (from 2) was detected at the first step. Previously described5,6 dications [1,2-(CH2)2C5Me3OsC5Me5]2+ 5 and [1,1'-(CH2C5Me4)2Os]2+ 6, as well as a new species that corresponds to trication [1,2-(CH2)2C5Me3Os(1'-CH2C5Me4)]3+ 7, were identified among the reaction products. The completion of the reaction was judged from the disappearance of dihydride 3 or dication 4 in the oxidation of 1 or 2, respectively.A comparison between the ratios of products obtained in the oxidation of 1 demonstrates that dications 5 (62%) and 6 (34%) were primarily formed, and the portion of trication 7 was as low as 4%. The yield of trication 7 cannot be increased by further passing of oxygen (Scheme 1).In the case of oxidation of compound 2 (Scheme 2), trication 7 was the main reaction product (81% on a basis of identified products), whereas dications 5 and 6 were formed in minor amounts (19%) (5:6 ~ 1.4:1) (Scheme 1). The structures of the products were supported by 1H and 13C NMR data.¡× Thus, the 1H NMR spectrum of dihydride 3 exhibits two singlet signals, and the ratio between the integral intensities of methyl protons and the signal of an OsH proton is equal to 15:1.All signals in the spectrum of dication 4 are broadened (.n1/2 ¡í 22.24 Hz for the signals of Me groups of the C5Me4 ring and the CH2 group, and .n1/2 ¡í 6.7 Hz for the signals of Me groups of the C5Me5 ring and the signal of OsH). Similarly to [C5Me5Re(H)C5Me4CH2]+,7 the protons of the CH2 group are non-equivalent, as well as the protons of Me groups of the C5Me4CH2 ring (four signals of 3H).The 1H NMR spectra of dications 5 and 6 are consistent with the spectra of dications obtained by the protonation of corresponding dicarbinols.5,6 The assignment of signals in the spectrum of trication 7 presented no special problems because all of the signals exhibited the same behaviour as those of the trication [1,2-(CH2)2C5Me3Ru(1'-CH2- C5Me4)]3+.3 Thus, the spectrum contained three signals of the same intensity (2H) from three CH2 groups, and the chemical shifts of these signals [d = 5.04 (d), 5.66 (d) and 5.49 (s) ppm] are almost equal to the values published for the Ru-containing trication.3 The appearance of signals due to three CH2 groups as two doublets and a singlet indicates that trication 7 exhibits a plane of symmetry and does not contain a C5Me5 ring.The equality of .dAB differences for trication 7 (0.62 ppm) and 1,2-dication 5 (0.62 ppm) is indicative of the 1,2-arrangement of CH2 groups in the C5Me3(CH2)2 ring of trication 7. The 13C NMR spectra of all complexes also support the structures. Thus, the carbon atoms of two 1,2-CH2 groups of trication 7 exhibit a chemical shift of 71.32 ppm in the 13C NMR spectrum and appear as a triplet with 1JCH = 172 Hz. The carbon atom of the 1'-CH2 group gives an upfield triplet (d = 65.75 ppm, 1JCH 157 Hz).Note that two CH2 groups in the Ru-containing trication exhibit a chemical shift of 88.57 ppm (1JCH = 172 Hz), whereas the chemical shift and 1JCH of the third 1'-CH2 group are consistent with the corresponding values for the 1'-CH2 group in trication 7.It is known8 that the great difference between the electronegativities of transition metal complexes and O2 induces the formation of bridging or nonbridging oxo compounds. Com- ¢Ó Complex 2 was prepared from C5Me5OsC5Me4CH2OH using HBF4 etherate.4 ¢Ô A solution of ~0.1 mmol of complex 1 or ~0.02 mmol of complex 2 in CF3SO3H (~3 mmol) was placed in a tube. Oxygen (~1 l) was bubbled through the solution for 2.3 h at ambient temperature. Os Os H H 2+ CF3SO3H, O2 20 ¡ÆC Os 2+ Os 2+ 1 3 5 6 Scheme 1 ¡× The NMR spectra were measured on a Bruker AMX-400 spectrometer (400.13 and 100.61 MHz for 1H and 13C, respectively).An external standard was used for CF3SO3H solutions (d C6D5H 7.25 and 127.96 ppm for 1H and 13C, respectively). 3: 1H NMR, d: 2.68 (s, 30H, C5Me5), .14.58 (s, 2H, OsH). 13C NMR, d: 8.77 (¥ã-Me), 106.32 (CCp). 4: 1H NMR, d: 2.57 (s, 15H, C5Me5), 2.17, 2.31, 2.51 and 2.98 (4s, 4¡¿3H, ¥á- and ¥â-Me), 5.17, 5.79 (2s, 2¡¿1H, CH2), .15.40 (s, 1H, OsH). 13C NMR, d: 9.29 (¥ã-Me), 8.55 (¥á-Me), 9.35 (¥â-Me), 66.50 (CH2), 107.06 (C1), 96.39, 102.74, 108.4 (CCp), 105.53 (¥ã-CCp). 5: 1H NMR, d: 2.57 (s, 15H, C5Me5), 2.14 (s, 6H, ¥á-Me), 2.57 (s, 3H, ¥â-Me), 4.90 and 5.52 (2d, 2¡¿2H, CH2 AB, 2Jgem HH 2.7 Hz); cf. ref. 5. 13C NMR, d: 10.00 (¥ã-Me), 9.41 (¥á-Me), 10.23 (¥â-Me), 70.92 (CH2, 1JCH 171 Hz), 134.66 (C1), 105.53, 115.06 (CCp), 107.58 (¥ã-CCp); cf. ref. 5. 6: 1H NMR, d: 2.19, 2.31, 2.41 and 2.83 (4¡¿6H, ¥á,¥á',¥â,¥â'-Me), 5.43, 5.87 (2d, 2¡¿2H, CH2); cf.ref. 6. 13C NMR, d: 6.83, 8.64 (¥á,¥á'-Me), 8.89, 10.90 (¥â,¥â'-Me), 73.31 (CH2, 1JCH 170 Hz), 99.86, 105.71, 111.51, 116.07, 116.53 (CCp). 7: 1HNMR, d: 2.19, 2.50, 2.71 and 3.03 (3¡¿6H, 3H, ¥á,¥á',¥â,¥â'-Me), 5.04, 5.66 (2d, 2¡¿2H, CH2 AB, 2Jgem HH 2.3 Hz), 5.49 (s, 2H, CH2). 13CNMR, d: 8.67, 8.71, 9.28 (3¡¿2Me), 9.47 (Me), 65.75 (CH2, 1JCH 157 Hz), 71.32 (2CH2, 1JCH 172 Hz), 93.20, 102.70, 106.93, 108.45, 116.80, 135.12 (CCp).Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) plexes having a free coordination site let O2 bind the metal in end-on and edge-on modes. The unprecedented mode of C–H activation of permethyl ligands (for example, ç6-arene) implies a monoelectronic transfer from an organometallic complex to O2 followed by the deprotonation of O2 · –, a ligand activated by the cationic metal moiety:9 The transformation of O2 into the radical anion is due to a low redox potential of this passage (E1/2 = –0.7 V/SCE).8 Now, we can only say that dication [C5Me5Os(H)C5Me4CH2]2+ 4 is an obligatory synthon on the way to trication 7 because the protonation of [C5Me5OsC5Me4CH2]+BF4 – 2 and the formation of trication 7 in the CF3CO2H–O2 system do not take place.Thus, trication 7 can be generated by the oxidation of monocation 2 with oxygen in CF3SO3H. At the same time, the oxidation of osmocene 1 with O2 in CF3SO3H can be considered as a method for generating dications 5 and 6. This work was supported by the Russian Foundation for Basic Research (grant no. 00-03-32894). References 1 A. A. Kamyshova, A. Z. Kreindlin, M. I. Rybinskaya and P. V. Petrovskii, Izv. Akad. Nauk, Ser. Khim., 1999, 587 (Russ. Chem. Bull., 1999, 48, 581). 2 A. A. Kamyshova, A. Z. Kreindlin, M. I. Rybinskaya and P. V. Petrovskii, Izv. Akad. Nauk, Ser. Khim., 2000, 517 (Russ. Chem. Bull., 2000, 49, 520). 3 M. I. Rybinskaya, A. A. Kamyshova, A. Z.Kreindlin and P. V. Petrovskii, Mendeleev Commun., 2000, 85. 4 A. Z. Kreindlin, P. V. Petrovskii and M. I. Rybinskaya, Izv. Akad. Nauk SSSR, Ser. Khim., 1987, 1620 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1987, 36, 1489). 5 A. Z. Kreindlin, E. I. Fedin, P. V. Petrovskii, M. I. Rybinskaya, R. M. Minyaev and R. Hoffmann, Organometallics, 1991, 1206. 6 M. I. Rybinskaya, A. Z. Kreindlin, P. V. Petrovskii, R. M. Minyaev and R. Hoffmann, Organometallics, 1994, 3903. 7 F. G. N. Cloke, J. P. Day, J. C. Green, C. P. Morley and A. C. Swain, J. Chem. Soc., Dalton Trans., 1991, 789. 8 L. I. Simandi, Catalytic Activation of Dioxygen by Metal Complexes, Kluwer, Dordrecht, 1992. 9 D. Astruc, J.-R. Hamon, E. Roman and P. Michaud, J. Am. Chem. Soc., 1981, 103, 7502. Os 2 CF3SO3H, O2 20 °C Os 4 H 2+ Os 7 3+ Scheme 2 MRH + O2 MRH+O2 · – MR – HO2 Received: 28th March 2001; Com. 01/1782
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
8. |
3,3'-Bi(6,8-dialkyl-2,4-dioxa-7-thia-6,8-diazabicyclo[3.3.0]octane 7,7-dioxides) as new heterocyclic system derivatives |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 138-140
Galina A. Gazieva,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) 3,3'-Bi(6,8-dialkyl-2,4-dioxa-7-thia-6,8-diazabicyclo[3.3.0]octane 7,7-dioxides) as new heterocyclic system derivatives Galina A. Gazieva,*a Angelina N. Kravchenko,a Oleg V. Lebedev,a Konstantin A. Lyssenko,b Marina O. Dekaprilevich,a Vladimir M. Men’shov,a Yurii A. Strelenkoa and Nina N. Makhovaa a N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 119991 Moscow, Russian Federation.Fax: +7 095 135 5328; e-mail: galina_ioc@mail.ru b A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 117813 Moscow, Russian Federation. Fax: +7 095 135 5085; e-mail: kostya@xrlab.ineos.ac.ru 10.1070/MC2001v011n04ABEH001465 The title compounds were synthesised by condensation of N,N'-dialkylsulfamides with glyoxal and structurally characterised by X-ray diffraction analysis.Cyclic sulfamides are biologically active substances.1,2 The reactions of sulfamides and N-alkylsulfamides with 1,2-dicarbonyl compounds are widely used in the synthesis of 1,2,5-thiadiazole (thiadiazoline) 1,1-dioxides.1,2 Here, we studied the condensation of N,N'-dialkylsulfamides 1 with glyoxal and synthesised 3,3'-bi(6,8-dialkyl-2,4-dioxa-7- thia-6,8-diazabicyclo[3.3.0]octane 7,7-dioxides) 2, the derivatives of the new [1,3]dioxolano[4,5-c][1,2,5]thiadiazolidine heterocyclic system (Scheme 1).The use of a trimeric dihydrate of glyoxal instead of aqueous glyoxal solution in the condensation with 1 resulted in an increase in the yield of compounds 2 from 5–11 to 54–85%.We cannot decide between the structures of 2 and 2' based on the data of elemental analysis, mass spectrometry and 1H NMR spectroscopy.† By analogy with trimeric glyoxal dihydrate3 3 and the product of the condensation of N,N'-dimethylurea with glyoxal4 4, the structure of 2' should be attributed to the compounds. On the other hand, the mass spectra of all the prepared compounds exhibit intense peaks corresponding to the halved molecular weight. This fact counts in favour of structure 2.It is also well known that the oligomerisation of glyoxal in aqueous solutions results in not only dioxane but also dioxolane structures.5,6 A mass-spectrometric study of trimeric glyoxal dihydrate and compound 4, which was synthesised according to a published procedure,4 showed that their mass spectra, as well as the mass spectra of products 2, exhibited peaks corresponding to the halved molecular weights.Thus, we can believe that these compounds can exist as dioxolane structures 3' and 4'. To independently determine the structure of compound 2g, a single crystal of 2g was examined by X-ray diffraction analysis.‡ It was found that the reaction product is a compound in which two bicyclic units are connected by carbon–carbon bond (Figure 1) rather than a triply annelated tetracycle.In the crystal of 2g a molecule occupies a special position – the centre of symmetry lying in the middle of the C(3)–C(3') bond. The bicyclic fragment is constructed of cis-combined thiadiazolidine and dioxolane rings with both hydrogen atoms at C(3) and C(3') orientated inside the ‘half-open books’ of the bicycles.The thiadiazolidine ring is characterised by the conformation of flattening envelope with the deviation of the N(6) atom by 0.24 Å. The dioxolane ring is characterised by a twist conformation with the C(3) and O(2) atoms shifted by 0.37 and –0.18 Å, respectively. The relative configurations of the asymmetric centres in a bicyclic fragment are C(5) – S, C(1) – R and of the pseudo asymmetric atom C(3) – s.An analysis of the crystal packing of 2g demonstrated that shortened contacts between H(1) and H(5) atoms and oxygen atoms of sulfo group occur in the crystal. These C–H···O contacts combine molecules to form corrugated layers parallel to crystallographic plane ab (Figure 2).Compounds 2 bear four asymmetric carbon atoms each (1, 5, 1' and 5') in bicyclic units (Figure 1). When the substituents at nitrogen atoms are identical, the molecule of 2 has a plane of symmetry and are meso-forms (2a,b,d–f). When the substituents at nitrogen atoms are structurally or configurationally different (methyl or ethyl in 2c; 1Ror 1S-methylpropyl in 2g) the products are obtained as the mixtures of diastereomeric meso-forms and racemates. 3,3'-Bi{6,8- 2 2 2 2 +2 +2 2+ 2+ 2 2 2 2 1 1 1 1 2 2 0H 0H 0H 0H 2 2 2 2 1 26 1 1 62 1 5 5 5 5 Scheme 1 Reagents and conditions: i, conc. HCl, 35–40 °C, stirring, 1–1.5 h. R1 NH O2S NH R2 2 O H O H 3 O O N O2S N R1 R2 O O N SO2 N R1 R2 i 1a–g 2a–g a R1 = R2 = Me b R1 = R2 = Et c R1 = Me, R2 = Et d R1 = R2 = Pr e R1 = R2 = Pri f R1 = R2 = Bu g R1 = R2 = Bus O(6) S(1) O(5) C(9) N(6) C(10) N(8) C(5) H(5) H(1) C(1) O(4) O(2) H(3) C(3) C(3') Figure 1 The general view of a molecule of 2g.The 1-methylpropyl substituents are omitted for clarity. 2 2 2 2 +2 +2 2+ 2+ 2 2 2 2 1 1 1 1 2 2 0H 0H 0H 0H Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) di(1-methylpropyl)-2,4-dioxa-7-thia-6,8-diazabicyclo[3.3.0]- octane 7,7-dioxide} 2g has four additional chiral centres in the substituents at nitrogen atoms. The C[3(3')] atoms are pseudoasymmetric. Theoretically, compound 2g can form 24 stereoisomers: the s, s and r, r configurations of C(3) and C(3') atoms can give 7 isomers each (3 racemates and 4 meso-forms) and the r, s combination of configurations of these atoms can give 10 isomers (6 racemates and 4 meso-forms).All of these isomers are constructed of the following four stereochemically different blocks: For example, the isomers with the s, s configurations of C(3) and C(3') atoms can be presented as follows: In the 1H NMR¢Ó spectra, the protons of the CH.CH fragment of blocks A and B exhibit two different singlets (d 5.83, 5.88), and those of blocks C and D exhibit the AB system with the constants JAB 11.7 Hz (dA 5.80, dB 5.91).Product 2g was separated into six fractions by HPLC. The 1H ¢Ó All new compounds gave satisfactory elemental analysis data. 1H NMR spectra were recorded on a Bruker AM 300 spectrometer in a [2H6]DMSO solution. Chemical shifts were measured with reference to residual protons of a [2H6]DMSO solvent (d 2.50 ppm).Mass spectra were measured on an MS 30 spectrometer. 2a: yield 81.85%, mp 253.254 ¡ÆC (decomp.). 1H NMR, d: 2.88 (s, 12H, 4NMe), 5.03 (s, 2H, CHCH), 5.70 (s, 4H, 2CHCH). MS, m/z (%): 193 (26) [1/2 M+], 148 (36), 83 (42). 2b: yield 54.56%, mp 219.221 ¡ÆC (decomp.). 1HNMR, d: 1.30 (t, 12H, 4Me), 3.25.3.34 (m, 8H, 4NCH2), 5.05 (s, 2H, CHCH), 5.77 (s, 4H, 2CHCH).MS, m/z (%): 441 (1), 221 (96) [1/2 M+], 193 (63), 176 (92). 2c: yield 65.67%, mp 228.230 ¡ÆC (decomp.). 1H NMR, d: 1.21 (t, 6H, 2Me), 2.79 (s, 6H, 2NMe), 3.21.3.30 (m, 4H, 2NCH2), 5.01 (s, 2H, CHCH), 5.73, 5.88 (m, 4H, 2CHCH, AB, JAB 5.8 Hz). MS, m/z (%): 414 (1) [M+], 221 (21), 207 (100) [1/2 M+], 179 (72), 162 (91). 2d: yield 72.74%, mp 173.175 ¡ÆC (decomp.). 1H NMR, d: 0.98 (t, 12H, 4Me), 1.56.1.68 (m, 8H, 4CH2), 3.06.3.23 (m, 8H, 4NCH2), 5.03 (s, 2H, CHCH), 5.80 (s, 4H, 2CHCH). MS, m/z (%): 249 (87) [1/2 M+], 221 (47), 204 (70) 179 (72), 111 (100). 2e: yield 54.57%, mp 229.231 ¡ÆC (decomp.). 1H NMR, d: 1.30 (d, 24H, 8Me), 3.71.3.88 (m, 4H, 4NCH), 5.09 (s, 2H, CHCH), 5.88 (s, 4H, 2CHCH). MS, m/z (%): 498 (1) [M+], 249 (61) [1/2 M+], 204 (63). 2f: yield 71.72%, mp 180.182 ¡ÆC (decomp.). 1H NMR, d: 0.93 (t, 12H, 4Me), 1.30.1.44 (m, 8H, 4CH2), 1.56.1.68 (m, 8H, 4CH2), 3.11. 3.25 (m, 8H, 4NCH2), 5.00 (s, 2H, CHCH), 5.75 (s, 4H, 2CHCH). MS, m/z (%): 277 (100) [1/2 M+], 232 (81), 205 (52), 137 (42). A mixture of 2g stereoisomers: yield 69.71%, mp 232.234 ¡ÆC (decomp.). 1H NMR, d: 0.94 (t, 6H, 2Me), 0.96 (t, 6H, 2Me), 1.23.1.37 (m, 12H, 4Me), 1.50.1.72 (m, 4H, 2CH2), 1.68.1.85 (m, 4H, 2CH2), 3.50.3.62 (m, 4H, NCH), 5.04 (s, 2H, CHCH), 5.79.5.92 (m, 4H, 2CHCH).MS, m/z (%): 553 (0.5), 277 (37) [1/2 M+], 232 (95), 219 (29), 139 (100). The diastereomers of 2c and 2g were separated on a Bruker LC21 liquid chromatograph at room temperature [column: Silica IBM. Eluent: light petroleum.ethyl acetate (85:15). Flow rate: 1 ml min.1.Solvent: CH2Cl2]. Fraction 1 obtained by the HPLC separation of 2g: 1H NMR, d: 0.85 (t, 6H, 2Me), 0.87 (t, 6H, 2Me), 1.29 (d, 12H, 4Me), 1.58.1.66 (m, 4H, 2CH2), 1.66.1.74 (m, 4H, 2CH2), 3.51.3.59 (m, 4H, NCH), 5.06 (s, 2H, CHCH), 5.83 (s, 4H, 2CHCH). Fraction 2 obtained by the HPLC separation of 2g: 1H NMR, d: 0.77 (t, 6H, 2Me), 0.79 (t, 6H, 2Me), 1.23.1.37 (m, 12H, 4Me), 1.50.1.80 (m, 8H, 4CH2), 3.49.3.61 (m, 4H, NCH), 5.06 (s, 2H, CHCH), 5.83 (s, 2H, CHCH), 5.80, 5.91 (m, 2H, CHCH, AB-system, JAB 11.7 Hz).Fraction 3 obtained by the HPLC separation of 2g: 1H NMR, d: 0.85 (t, 6H, 2Me), 0.87 (t, 6H, 2Me), 1.20.1.31 (m, 12H, 4Me), 1.48.1.78 (m, 8H, 4CH2), 3.49.3.61 (m, 4H, NCH), 5.06 (s, 2H, CHCH), 5.80, 5.91 (m, 4H, 2CHCH, AB-system, JAB 11.7 Hz).Fraction 4 obtained by the HPLC separation of 2g: 1H NMR, d: 0.85 (t, 6H, 2Me), 0.87 (t, 6H, 2Me), 1.20.1.31 (m, 12H, 4Me), 1.48.1.78 (m, 8H, 4CH2), 3.49.3.61 (m, 4H, NCH), 5.06 (s, 2H, CHCH), 5.83 (s, 1H, CHCH), 5.88 (s, 1H, CHCH), 5.80, 5.91 (m, 2H, CHCH, AB-system, JAB 11.7 Hz). Fraction 5 obtained by the HPLC separation of 2g: 1H NMR, d: 0.85 (t, 6H, 2Me), 0.87 (t, 6H, 2Me), 1.23.1.37 (m, 12H, 4Me), 1.48.1.76 (m, 8H, 4CH2), 3.49.3.61 (m, 4H, NCH), 5.04, 5.08 (m, 2H, CHCH, AB-system, JAB 1.4 Hz), 5.88 (s, 2H, CHCH), 5.80, 5.91 (m, 2H, CHCH, AB-system, JAB 11.7 Hz).Fraction 6 obtained by the HPLC separation of 2g: 1H NMR, d: 0.85 (t, 6H, 2Me), 0.87 (t, 6H, 2Me), 1.23.1.37 (m, 12H, 4Me), 1.48.1.76 (m, 8H, 4CH2), 3.49.3.61 (m, 4H, NCH), 5.06 (s, 2H, CHCH), 5.88 (s, 2H, CHCH), 5.80, 5.91 (m, 2H, CHCH, AB-system, JAB 11.7 Hz). 3: MS, m/z (%): 135 (1), 117 (1), 105 (42) [1/2 M+], 88 (13), 77 (8), 60 (93), 58 (98). 4: yield 18.20%, mp 236.238 ¡ÆC (decomp.) [lit.,4 237.239 ¡ÆC (decomp.)]. 1H NMR, d: 2.71 (s, 12H, 4NMe), 4.77 (s, 2H, CHCH), 5.60 (s, 4H, 2CHCH). MS, m/z (%): 313 (0.5) [M+], 157 (46) [1/2 M+], 157 (45), 129 (47), 112 (100).¢Ô Crystallographic data for 2g: crystals of C22H42N4O8S2 are orthorhombic at 110 K, space group Pbca, a = 12.6834(9) A, b = 11.5267(9) A, c = 19.548(1) A, V = 2857.9(4) A3, Z = 4, M = 554.72, dcalc = 1.289 g cm.3, m(MoK¥á) = 2.35 cm.1, F(000) = 1192. Intensities of 15138 reflections were measured on a Smart 1000 CCD diffractometer at 110 K [l(MoKa) = = 0.71072 A, w-scans with 0.3¡Æ step in w and 20 s per frame exposure, 2q < 50¡Æ], and 2517 independent reflections (Rint=0.0704) were used in further refinement.The structure was solved by a direct method and refined by the full-matrix least-squares technique against F2 in the anisotropic .isotropic approximation. The analysis of the Fourier synthesis has revealed that both 1-methylpropyl fragments are disordered.This disorder is the result of superposition of these fragments with different conformations along the N.C bond. The positions of the hydrogen atoms were calculated from the geometrical point of view. The refinement converged to wR2 = 0.1866 and GOF = 0.890 for all independent reflections [R1 = 0.0695 was calculated against F for 1348 observed reflections with I > 2s(I)].All calculations were performed using SHELXTL PLUS 5.0 on IBM PC AT. Atomic coordinates, bond lengths, bond angles and thermal parameters have been deposited at the Cambridge Crystallographic Data Centre (CCDC). For details, see ¡®Notice to Authors¡�, Mendeleev Commun., Issue 1, 2001. Any request to the CCDC for data should quote the full literature citation and the reference number 1135/94.Figure 2 The C.H¡�¡�¡�O bonded corrugated layers. The 1-methylpropyl groups are omitted for clarity. The parametrs of the C.H¡�¡�¡�O contacts are H(5')¡�¡�¡� O(5) 2.21 A, C(5').H(5').O(5) 152¡Æ, C(5')¡�¡�¡�O(5) 3.193(3) A, H(1')¡�¡�¡�O(6) 2.45 A, C(1').H(1').O(6) 165¡Æ, C(1')¡�¡�¡�O(6) 3.497(3) A. O(6) O(5) H(5') C(5') H(1') C(1') O O N S N O O H H H Et Me H Et Me R S R S A O O N S N O O H H H Et Me H Et Me S R R S B O O N S N O O H H H Et Me H Et Me R R R S C O O N S N O O H H H Et Me H Et Me S S R S D AS AS BS BS AS BS meso1 meso2 meso3 AS CS AS DS racemate1 BS CS BS DS racemate2 CS CS DS DS racemate3 CS DS meso4Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) NMR spectroscopy data of the fractions suggest that both 2g isomers with s, s configurations of atoms C(3) and C(3') and at least the isomers with s, r configurations were present (fraction 5†).The C(3) and C(3') atoms of 3,3'-bi(N-ethyl-N'-methyl-2,4- dioxa-7-thia-6,8-diazabicyclo[3.3.0]octane 7,7-dioxide) 2c are asymmetric. Theoretically two meso-forms and four racemates of 2c may exist.Compound 2c gave two HPLC peaks. In the 1H NMR spectra, the meso forms and racemates were not resolved. Thus, a simple one-pot method was proposed for the synthesis of 3,3'-bi(6,8-dialkyl-2,4-dioxa-7-thia-6,8-diazabicyclo- [3.3.0]octane 7,7-dioxides) 2 (a derivative of the new heterocyclic system [1,3]dioxolano[4,5-c][1,2,5]thiadiazoline) from the accessible components N,N'-dialkylsulfamides and trimeric glyoxal dihydrate. The structure of the new compound was found by X-ray diffraction analysis. This work was supported by INTAS (grant no. 99-0157) and the Russian Foundation for Basic Research (grant no. 00-15- 97359). References 1 V. J. Aran, P. Goya and C. Ochoa, Adv. Heterocycl. Chem., 1988, 44, 81. 2 G. A. Gazieva, A. N. Kravchenko and O. V. Lebedev, Usp. Khim., 2000, 69, 239 (Russ. Chem. Rev., 2000, 69, 221). 3 H. Raudnintz, Chem. Ind., 1944, 327. 4 R. H. Barker, S. L. Vail and G. B. Barcelo, J. Heterocycl. Chem., 1966, 3, 354. 5 E.B.Whipple, J. Am. Chem. Soc., 1970, 92, 7183. 6 I. Ya. Slonim, B. M. Arshava and V. N. Klyuchnikov, Sovrem. Aspekty YaMR Spektrosk. Polim., 1994, 53 (in Russian). R
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
9. |
Bis-aziridinomethanes: synthesis, structure and properties |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 141-143
Remir G. Kostyanovsky,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) Bis-aziridinomethanes: synthesis, structure and properties Remir G. Kostyanovsky,*a Vasilii R. Kostyanovsky,a Boris B. Averkiev,b Konstantin A. Lyssenkob and Pavel E. Dormova a N. N. Semenov Institute of Chemical Physics, Russian Academy of Sciences, 117977 Moscow, Russian Federation. Fax: +7 095 938 2156; e-mail: kost@center.chph.ras.ru b A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 117813 Moscow, Russian Federation. Fax: +7 095 135 5085; e-mail: kostya@xray.ineos.ac.ru 10.1070/MC2001v011n04ABEH001471 Bis-aziridinomethanes were prepared by the reaction of Me2NCH(OMe)2 with aziridines and characterised by X-ray diffraction analysis and NMR spectroscopy.The incorporation of a nitrogen atom into a three-membered ring excludes (a) the ¥á-aminoalkylation reaction1 and (b) amide conjugation. 2 These basic rules result from decreasing the p-character of the nitrogen lone pair (lp) and hence its electron-donating properties.1(e).(f),2(a),(d),(e),(g).(i),3 Limitations (a) and (b) are most pronounced in the chemistry of bis-aziridinomethanes (BAMs). These compounds cannot be prepared by typical syntheses of usual aminales,4 such as reactions of aziridine with aldehydes, as well as 1-aziridinocarbinol or 1-alkoxymethylaziridines.1 Bis-aziridinomethoxymethane also does not react with aziridine even under severe conditions1(e) [limitation (a)].In accordance with limitation (b), the first BAMs were obtained by a mild addition of aziridine to activated acetylenes (bis-adducts A),5 whereas usual amines give only mono-adducts, which are amide vinylogs with deactivated double bonds.The simplest BAM B,6(a) as well as aziridinoform C6(b).(d) and derivatives D,6(e) was synthesised by the alkylation of aziridine under the action of CH2Cl2, CHCl3 and a corresponding gem-dichloroalkane either in the presence of bases6(c).(e) or through potassium ethyleneamide. 6(a),(b) The first and second steps of alkylation are determined by the presence of the easier leaving group Cl. (as compared with X. = OH, OR). In addition, the synthesis of diazaquadricyclane E containing a BAM fragment was reported.7 Finally, it was found that 5,5-dimethoxytetrachlorocyclopenta- 1,3-diene reacts with aziridine under mild conditions with formation of BAM F.8 This reaction results from the allylic activation of MeO groups at both steps of the reaction.The reaction of 2-cyanoaziridine with ketones9 (Scheme 1) is inconsistent with rule (a). There are contradictory data9 on the structure of the products which possess a high biological activity; they were described as bis-cyanoaziridinoalkanes.9(c) The structure of 1 was strictly confirmed; however, data on the relative configuration are ambiguous. It can be suggested that the key intermediates in the synthesis of 1 are the structures I1 and I2.The latter is a product of the transformation of I1 by Pinner reaction. The aziridinomethylating action of I2 is determined by the presence of the easily leaving iminoyloxy group as in the case of the above intermediate 1-chloromethylaziridines.The Chapman rearrangement of I2 into a corresponding lactam,9(b) as well as the transformation of the adduct of 2-cyanoaziridine (I2-type) with substituted cyclohexanone into compound G by intramolecular aziridinomethylation at the OH group, is well known.11 On the basis of the above analysis, we proposed a new efficient way for the preparation of BAM 2.It consists in the dimethylaminomethylenation of aziridine with dimethylformamide dimethylacetal (Scheme 1).¢Ó This reaction is a result of strong electron donation of Me2N, which increases the mobility of MeO groups (like allyl activation of MeO groups in the synthesis of BAM F). A special feature of the NMR spectra of 2 is that all protons and carbons of the aziridine ring are non-equivalent¢Ó [cf.ref. 1(e)]. BAM 1 was extensively studied as a biologically active compound10 for examining the nitrogen inversion and dynamic effects in BAMs, as well as for determining the relative configuration. Compound 1 was found to crystallise in two modifications, with mp 157.5 (from acetone) and 172.5 ¡ÆC (from MeOH.Et2O), both having identical NMR spectra, which correspond to one diastereomer.Thus, BAM 1 is formed diastereoselectively. In the 1H NMR spectrum in an aprotic solvent (CDCl3, 18 ¡ÆC), a considerable broadening of the upfield signal from the A-Me group and signals from all aziridine ring protons (particularly, Ha, Hb and Ha') are observed. In contrast, at 60 ¡ÆC, the broadening disappeared and the signals due to amide protons Ha, Hs shifted upfield (0.1 and 0.3 ppm).This temperature dependence of the spectrum points to the presence of hindered rotation, nitrogen inversion in the 2-cyanoaziridine moiety, as well as an intramolecular H-bond in the 2-carbamoylaziridine fragment of 1. The monocrystals suitable for an X-ray study¢Ô were grown only from the higher melting modification of 1 (not described earlier).According to the X-ray analysis, compound 1 is crystallised as a racemate (space group P1), and asymmetric centres in 1 [C(2) and C(7)] have identical configurations in contrast to published data9(a) [Figure 1(a).] In the crystal of 1, shortened contacts are observed. Namely, H¡�¡�¡�H contacts (2.16 and 2.22 A) formed by the methyl group H3C(5) with H(7) and H(8B) atoms, and H(2)¡�¡�¡�N(3) (2.38 A) ¢Ó Characteristics and spectroscopic data. 1H and 13C NMR spectra were measured at 400.13 and 100.61 MHz, respectively. 1:9(b) mp 172.5 ¡ÆC (MeOH.Et2O) and 157.5 ¡ÆC (acetone) [cf. refs. 9(a),(b)]. 1HNMR (CDCl3 at 50 ¡ÆC) d: 1.11 (br. s, 3H, A-Me), 1.19 (s, 3H, B-Me), 1.80 (dd, 1H, HC, 3Jtrans 2.9 Hz, 2Jgem 1.2 Hz), 2.06 (dd, 1H, HB, 3Jcis 6.8 Hz, 2Jgem 1.2 Hz), 2.11 (dd, 1H, HC' , 3Jcis 2.9 Hz, 2Jgem 1.3 Hz), 2.15 (dd, 1H, HB' , 3Jcis 6.3 Hz, 2Jgem 1.3 Hz), 2.40 (ddd, 1H, HA, 3Jcis 6.8 Hz, 3Jtrans 2.9 Hz, 4JAs 0.9 Hz), 2.52 (dd, HA', 3Jcis 6.8 Hz, 3Jtrans 2.9 Hz), 5.32 (br.s, 1H, Hs), 6.08 (br. s, 1H, Ha). 2: A mixture of aziridine and dimethylformamide dimethylacetal in a molar ratio of 2:1 was kept at 20 ¡ÆC for 10.12 h, evaporated and distilled over sodium metal in vacuo, bp 36.5.37 ¡ÆC (1 torr), yield 65.75%. 1H NMR (C6D6) d: 1.10, 1.19, 1.47 and 1.64 (m, 8H, ring protons, ABCD spectrum, .nAB 43.0 Hz, nCD 35.5 Hz, 3JAB cis 5.4 Hz, 3JCD cis 7.1 Hz, 3JAD trans = 3JBC trans = 3.8Hz, 2JAC gem = 2JBD gem = 0.7 Hz), 1.77 (s, 1H, HC), 2.55 (s, 6H, Me2N). 13C NMR (C6D6) d: 22.0 (ddm) and 24.6 (ddm) (ring carbons, 1J 165.0 Hz, 1J 175.0 Hz), 39.9 (qq, MeN, 1J 133.3 Hz, 3J 4.0 Hz), 105.2 (dm, CH, 1J 153.0 Hz).¢Ô Crystallographic data for 1: at 293 K, crystals of C9H14N4O are triclinic, space group P1, a = 6.309(2) A, b = 6.791(2) A, c = 13.082(2) A, a = 75.27(2)¡Æ, b = 89.20(2)¡Æ, g = 83.34(2)¡Æ, V = 538.3(2) A3, Z = 2, d = = 1.198 g cm.3, m = 0.83 cm.1, F(000) = 208.Intensities of 3384 reflections were measured with an Enraf Nonius CAD-4 diffractometer at 293 K [l(MoK¥á) = 0.710712 A, graphite monochromator, q/2q-scans, 2q < 60¡Æ], and 3117 independent reflections (Rint = 0.0290) were used in a further refinement. The structure was solved by the direct method and refined by full-matrix least squares against F2 in the anisotropic approximation for non-hydrogen atoms.All the hydrogen atoms were located from the electron density difference synthesis and included in the refinement in an isotropic approximation. The refinement converged to wR2 = 0.1316 and GOF = 1.008 for all independent reflection [R1 = = 0.0433 was calculated against F for 2300 observed reflections with I > 2s(I)]. All calculations were performed using SHELXTL PLUS 5.0 program.Atomic coordinates, bond lengths, bond angles and thermal parameters have been deposited at the Cambridge Crystallographic Data Centre (CCDC). For details, see ¡®Notice excl;�, Mendeleev Commun., Issue 1, 2001. Any request to the CCDC for data should quote the full literature citation and the reference number 1135/92.Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125.164) [Figure 1(a)]. Note that these protons exhibited the most broadened signals in the 1H NMR spectrum [A-Me, Ha, Hb, Ha' (Scheme 1)]. In the crystal of 1, molecules take part in the formation of intermolecular hydrogen bonds. Being bonded by the inversion centre in the crystal, the molecules of 1 form dimers due to the amide H-bonds.These dimers are combined into chains directed along the axis c by the H-bonds involving the nitrile group CN¡�¡�¡�H.N [Figure 1(b)]. The mutual arrangements of aziridine rings in relation to the fragment Me2C(4) are different. The angles between the ring planes N(2)C(2)C(3) and N(3)C(7)C(8) relative to the plane C(5)C(4)C(6) are 4.3 and 67.4¡Æ, respectively. Evidently, the above shortened contacts are responsible for increasing the angle N(3)C(4)C(5) up to 115.7(1)¡Æ [the other CCN angles at C(4) are in a range of 105.6.106.2¡Æ], as well as for flattening the N(3) atom as compared with N(2) (the sums of bond angles are 301.4 and 297.6¡Æ, respectively). The observed steric hindrance in 1 is probably caused by the stabilization of a conformation with the antiperiplanar orientation of lp of the N(2) atom in relation to the C(4).N(3) bond [pseudotorsion angle lp.N(2).C(4).N(3) is 173¡Æ], i.e., by the anomeric effect nN(2) ¢ç ¥ò*C(4)N(3).The pseudotorsion angle lp- N(3).C(4).N(2) is equal to 45¡Æ; this excludes its participation in the anomeric interaction with the C(4).N(2) bond. The simultaneous antiperiplanar orientations of lp-N(2) and lp-N(3) in relation to the corresponding CN bonds are sterically hindered and result in more shortened contacts H¡�¡�¡�H. In spite of anomeric interaction, the C(4).N(2) and C(4).N(3) bond lengths are practically the same.Such a situation is typical of aziridines1(g),6(d) and it is in agreement with recent high level calculations.12 The best indicator of anomeric effects is the NCN angles, which are quite close to each other in 1 [110.6(1)¡Æ] and CH2(NH2)2 (113¡Æ).12 It should be noted that anomeric effects in aziridines were confirmed by theoretical and experimental investigations in gas, liquid and solid states.For tris-aziridinomethane, the calculated and experimental (GED) lp-NCN, NCN angles are 166.3¡Æ, 112.6¡Æ and 172.6¡Æ, 114¡Æ, respectively.6(d) For 1-methoxymethylaziridine, the lp-NCO, NCO angles are 174.2¡Æ, 113.2¡Æ and 161¡Æ, 113.4¡Æ, respectively.1(g) In the latter case, the anomeric effect reveals itself in a considerable decrease in the nitrogen inversion barrier due to stabilization of the planar transition state.1(g) It can be demonstrated by comparing 1-dimethylaminomethylaziridine and its 2,2-dimethyl analogue.13 The transformation of the former compound into methyl iodide leads to a great decrease in the nitrogen inversion barrier (by 3.5 kcal mol.1) due to the anomeric effect n(N) ¢ç ¥ò*(C.N+), whereas upon a similar transformation of the latter compound the inversion barrier remained unchanged due to steric hindrances for the anomeric effect.Finally, according to an X-ray database, the anomeric effect is also observed in 1-alkoxyaziridine (+)-G11 where it is strengthened by steric strain which is favourable to lengthening the C.O bond [bond lengths: N(1).C(3) 1.418 A, C(3).O(1) 1.458 A; bond angles: lp-N(1). C(3).O(1) 178¡Æ, N(1).C(3).O(1) 112¡Æ].Apparently, the properties of (+)-G and its precursor, an intermediate of the I2 type, are responsible for the enantioselectivity of 2-cyanoaziridine hydration catalysed by a chiral substituted cyclohexanone (ee > 99%).11 N N CCH2Y X 2 X = H, CO2Me Y = CN, CO2Me A N CH2 2 B N CH 3 C Space group P63 Z = 6 N C CCl3 2 N PN 3 D Space group P21/n Z = 4 N R' R' R' R O O O O Me Me N 2 Cl Cl Cl Cl F R = CO2But R' = But E Space group P1 Z = 2 NH + Me2CO NC N NC C Me Me OH N O HN Me Me I1 I2 NH NC N NC Ha' Hc' Hb' Me-B Me-A N Hc Hb Ha N O Ha Hs 1 Me2NCH(OMe)2 NH Me2NCHN MeO NH N C H Me2N N HA HC HB HD 2 CN Me O Me Me N NH2 O (+)-G Space group P21 Z = 2 Scheme 1 O(1) C(9) N(4) H(4B) H(4A) H(6B) C(6) H(6C) H(6A) C(4) H(8A) C(8) C(7) H(7) H(5B) H(8B) C(5) H(5A) H(5C) N(2) H(3A) C(3) H(3B) C(2) H(2) C(1) N(1) O(1A) H(4B) N(4A) H(4BA) O(1) N(4) H(4A) N(1B) N(1) H(4AB) N(4B) Figure 1 (a) The general view of 1.Selected bond lengths (A): C(1).N(1) 1.140(2), C(1).C(2) 1.442(2), N(2).C(3) 1.446(2), N(2).C(2) 1.459(2), N(2).C(4) 1.479(1), C(2).C(3) 1.483(2), N(3).C(7) 1.454(1), N(3).C(8) 1.463(2), N(3).C(4) 1.473(1), N(4).C(9) 1.318(2); selected bond angles (¡Æ): N(3).C(4).N(2) 110.6(1), N(3).C(4).C(5) 115.7(1), N(2).C(4).C(5) 105.6(1), N(3).C(4).C(6) 106.1(1), N(2).C(4).C(6) 106.0(1), C(5).C(4).C(6) 112.5(1). The parameters of intramolecular H-bonds: N(3)¡�¡�¡�H(2) 2.38 A, N(3)¡�¡�¡�C(2) 2.752(1) A, C(2).H(2).N(3) 99¡Æ, N(3).H(4A) 2.40 A, N(3)¡�¡�¡�N(4) 2.800(1) A, N(4).H(4A).N(3) 109¡Æ. (b) The formation of Hbonded chains directed along crystallographic axis c. The parameters of H-bonds: O(1)¡�¡�¡�H(4B) 2.04 A, N(4)¡�¡�¡�O(1) 2.908(2) A, O(1).H(4B).N(4) 172(2)¡Æ; N(1)¡�¡�¡�H(4A) 2.52 A, N(1)¡�¡�¡�N(4) 3.268(2) A, N(1).H(4A).N(4) 146(1)¡Æ. (a) (b) 0a b c N(3)Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) This work was supported by the Russian Foundation for Basic Research (grant nos. 98-03-04119 and 00-03-81187 Bel) and the Russian Academy of Sciences. References 1 (a) R.G. Kostyanovsky, Dokl. Akad. Nauk SSSR, 1960, 135, 853 [Dokl. Chem. (Engl. Transl.), 1960, 1363]; (b) R. G. Kostyanovsky, Dokl. Akad. Nauk SSSR, 1961, 139, 877 [Dokl. Chem. (Engl. Transl.), 1961, 1237]; (c) R. G. Kostyanovsky, O. A. Yuzhakova and V. F. Bystrov, Izv. Akad. Nauk SSSR, Otdel. Khim. Nauk, 1962, 1666 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1962, 11, 1576); (d) R.G. Kostyanovsky and V. F. Bystrov, Izv. Akad. Nauk SSSR, Otdel. Khim. Nauk, 1962, 1488 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1962, 11, 1404); (e) R. G. Kostyanovsky and O. A. Pan’shin, Izv. Akad. Nauk SSSR, Ser. Khim., 1965, 740 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1965, 14, 721); (f) R. G. Kostyanovsky, R. K. Alekperov, G. K. Kadorkina, and I. I. Chervin, Izv. Akad. Nauk SSSR, Ser.Khim., 1987, 2523 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1987, 36, 2343); (g) S. V. Varlamov, G. K. Kadorkina and R. G. Kostyanovsky, Khim. Geterotsikl. Soedin., 1988, 390 [Chem. Heterocycl. Compd. (Engl. Transl.), 1988, 24, 320]; (h) I. F. Shishkov, L. V. Khristenko, L. V. Vilkov, M. Dakkouri, G. K. Kadorkina, P. E. Dormov and R. G. Kostyanovsky, Mendeleev Commun., 2000, 217. 2 (a) R. G. Kostyanovsky and V. F. Bystrov, Dokl. Akad. Nauk SSSR, 1963, 148, 839 [Dokl. Chem. (Engl. Transl.), 1963, 97]; (b) R. P. Shibaeva, L. O. Atovmyan and R. G. Kostyanovsky, Dokl. Akad. Nauk SSSR, 1967, 175, 586 (in Russian); (c) L. V. Vilkov, I. I. Nazarenko and R. G. Kostyanovsky, Zh. Strukt. Khim., 1968, 9, 1075 [J. Struct. Chem. (Engl. Transl.), 1968, 9, 960]; (d) R.G. Kostyanovsky, K. S. Zakharov and M. Zaripova, Tetrahedron Lett., 1974, 4207; (e) T. Z. Papoyan, I. I. Chervin and R. G. Kostyanovsky, Izv. Akad. Nauk SSSR, Ser. Khim., 1978, 1530 (in Russian); (f) O. A. Dyachenko, L. O. Atovmyan, S. M. Aldoshin, A. E. Polyakov and R. G. Kostyanovsky, J. Chem. Soc., Chem. Commun., 1976, 50; (g) G. V. Shustov, G. K. Kadorkina, S. V. Varlamov, A.V. Kachanov, R. G. Kostyanovsky and A. Rauk, J. Am. Chem. Soc., 1992, 114, 1616; (h) G. V. Shustov, A. V. Kachanov, G. K. Kadorkina, R. G. Kostyanovsky and A. Raukand A. Rauk, J. Org. Chem., 2000, 65, 3612. 3 V. F. Bystrov, O. A. Yuzhakova and R. G. Kostyanovsky, Dokl. Akad. Nauk SSSR, 1962, 147, 843 [Dokl. Chem. (Engl. Transl.), 1962, 1049]. 4 R. G. Kostyanovsky and O. A. Pan’shin, Izv. Akad. Nauk SSSR, Otdel. Khim. Nauk, 1963, 182 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1963, 12, 164). 5 (a) R. G. Kostyanovsky and O. A. Yuzhakova, Dokl. Akad. Nauk SSSR, 1964, 159, 142 [Dokl. Chem. (Engl. Transl.), 1964, 1152]; (b) Yu. I. El’natanov and R. G. Kostyanovsky, Izv. Akad. Nauk SSSR, Ser. Khim., 1988, 1858 (Bull. Acad.Sci. USSR, Div. Chem. Sci., 1988, 37, 1661). 6 (a) R. G. Kostyanovsky and O. A. Pan’shin, Izv. Akad. Nauk SSSR, Ser. Khim., 1965, 567 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1965, 14, 553); (b) R. G. Kostyanovsky, Yu. I. El’natanov, and Kh. Khafizov, Izv. Akad. Nauk SSSR, Ser. Khim., 1970, 1918 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1970, 19, 1815); (c) W. Funke, Liebigs Ann. Chem., 1969, 725, 15; (d) V.P. Novikov, M. Dakkouri, A. V. Golubinskii, M. V. Popik, L. V. Vilkov, P. E. Dormov, K. A. Lyssenko and R. G. Kostyanovsky, Mendeleev Commun., 2000, 103; (e) L. I. Zhdanova, V. N. Biyushkin, I. E. Boldeskul, V. Ya. Semenii and T. I. Malinovskii, Dokl. Akad. Nauk SSSR, 1986, 287, 635 (in Russian). 7 G. Michels, R. Mynott and M. Regitz, Chem. Ber., 1988, 121, 357. 8 O. G. Nabiev, M. A. Shachgeldiev, I. I. Chervin and R. G. Kostyanovsky, Izv. Akad. Nauk SSSR, Ser. Khim., 1985, 715 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1985, 34, 656). 9 (a) K. F. Koehler, H. Zaddach, G. K. Kadorkina, V. N. Voznesenskii, I. I. Chervin and R. G. Kostyanovsky, Izv. Akad. Nauk SSSR, Ser. Khim., 1993, 2136 (Russ. Chem. Bull., 1993, 42, 2049); (b) K. Jahnisch, E. Schmitz and E. Gründemann, J. Prakt. Chem., 1979, 321, 712; (c) W. Kampe, M. Thiel, E. Fauland, U. Bicker and G. Habold, USSR Pat. no. 673167, Bull. Izobret., 1979, no. 25, 234 (in Russian). 10 (a) R. G. Kostyanovsky, P. E. Dormov, P. Trapenzieris, B. Strumfs, G. K. Kadorkina, I. I. Chervin and I. Ya. Kalvin’s, Mendeleev Commun., 1999, 26; (b) R. G. Kostyanovsky, K. A. Lyssenko, A. N. Kravchenko, O. V. Lebedev, G. K. Kadorkina and V. R. Kostyanovsky, Mendeleev Commun., 2001, 134. 11 K. Jahnisch, E. Grundemann, A. Kunath and M. Ramm, Liebigs Ann. Chem., 1994, 881. 12 L. Carballeira and I. Perez-Juste, J. Phys. Chem. A, 2000, 104, 9362. 13 V. F. Rudchenko, S. M. Ignatov, I. I. Chervin, V. S. Nosova and R. G. Kostyanovsky, Izv. Akad. Nauk SSSR, Ser. Khim., 1986, 1153 (Bull. Acad. Sci. USSR, Div. Chem. Sci., 1986, 35, 1045). Received: 4th May 2001; Com. 01/1797
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
10. |
Condensation of oxazolidines with 2-hydroxybenzaldehydes |
|
Mendeleev Communications,
Volume 11,
Issue 4,
2001,
Page 143-144
Boris F. Kukharev,
Preview
|
|
摘要:
Mendeleev Communications Electronic Version, Issue 4, 2001 (pp. 125–164) Condensation of oxazolidines with 2-hydroxybenzaldehydes Boris F. Kukharev,* Valerii K. Stankevich, Galina R. Klimenko and Victor V. Bayandin A. E. Favorsky Irkutsk Institute of Chemistry, Siberian Branch of the Russian Academy of Sciences, 664033 Irkutsk, Russian Federation. Fax: +7 3952 39 6046; e-mail: admin@irioch.irk.ru 10.1070/MC2001v011n04ABEH001467 The reaction of N-unsubstituted oxazolidines with 2-hydroxybenzaldehydes resulted in 2,3,5,10b–tetrahydrooxazole[3,2-c][1,3]- benzoxazines.It is known that cyclic O,N-acetal derivatives, oxazolidines, can be involved in the exchange reaction with aldehydes to give new oxazolidines and aldehydes.1 Reactions of oxazolidines with phenols result in N-(2–hydroxyethyl)aminomethylphenols.2 Reactions of N-unsubstituted oxazolidines with aldehydes containing phenyl hydroxyl groups have not been investigated. We found that salicylic and 5-bromosalicylic aldehydes react with 4,4-dimethyl- and 5-methyloxazolidines to give 2,3,5,10btetrahydrooxazole[ 3,2-c][1,3]benzoxazine derivatives 3a–d.The reaction took place under reflux of an equimolecular mixture of an oxazolidine and a corresponding aldehyde in benzene followed by the azeotropic distillation of water.† An attempt to involve the simplest oxazolidine in the reaction was unsuccessful.Probably, this was due to the high rate of oxazolidine trimerization to stable N,N',N''-tris(2-hydroxyethyl)- perhydro-1,3,5-triazine.3 It is likely that oxazolidine, which is a typical secondary amine, forms semi-acetal 4 at the first stage; then, the oxazolidine ring is decomposed with a phenyl hydroxyl group to give benzoxazine 5, which transforms to oxazolobenzoxazine 3 (Scheme 2).The structure of the compounds obtained was confirmed by 1H NMR spectroscopy and elemental analysis. Compounds 3c and 3d (mixtures of diastereomers) were obtained when oxazolidine 1b having an asymmetric carbon atom was involved in the reaction.In 1H NMR spectra, the proton signals of both the methyl groups and the OCNH fragment of the oxazolidine ring of a minor isomer were upfield shifted in comparison with those in the major isomer. The ratio between diastereomers calculated from the integral intensity of the proton signals was approximately 1:1.1 for compounds 3c and 3d.References 1 K. D. Petrov and O. K. Gosteva, in Sbornik statei po obshchei khimii (Collected Articles on General Chemistry), AN SSSR, Moscow– Leningrad, 1953, vol. 2, p. 1352 (in Russian). 2 (a) K. D. Petrov, O. K. Gosteva and V. I. Pukhova, Zh. Obshch. Khim., 1957, 27, 3218 [J. Gen. Chem. USSR (Engl. Transl.), 1957, 27, 1300]; (b) R. A. Fairhurst, H.Heaney, G. Papageorgiou, R. F. Wilkins and S. C. Eyley, Tetrahedron Lett., 1989, 30, 1433. 3 M. Riehl and P. A. Laurent, Bull. Soc. Chim. Fr., 1969, 1223. † A mixture of oxazolidine 1a,b (0.1 mol), aldehyde 2a,b (0.1 mol) and benzene (100 ml) was refluxed with a Dean-Stark condenser trap until water isolation was ceased. The mixture was distilled in a vacuum to give 2,3,5,10b-tetrahydrooxazole[3,2-c][1,3]benzoxazines 3a–d. 1H NMR spectra were measured at 400.13 MHz in CDCl3 solution, standard TMS. 3a: yield 59%, bp 129–131 °C (6 torr), nD 20 1.5466, d4 20 1.1132. 1H NMR, d: 1.24 (s, 3H, Me), 1.31 (s, 3H, Me), 3.67 (d, 1H, OCHAHB, 2JAB 7.8 Hz), 3.71 (d, 1H, OCHAHB, 2JAB 7.8 Hz), 4.85 (d, 1H, NCHAHB, 2JAB 11.0 Hz), 4.93 (d, 1H, NCHAHB, 2JAB 11.0 Hz), 5.87 (s, 1H, OCHN), 6.78 (dd, 1H, 7-HC, 3J7–8 8.2 Hz, 4J7–9 1.1 Hz), 6.94 (ddd, 1H, 9-HC, 3J8–9 7.4 Hz, 3J9–10 7.6 Hz, 4J7–9 1.1 Hz), 7.17 (ddd, 1H, 8-HC, 3J7–8 8.2 Hz, 3J8–9 7.4 Hz, 4J8–10 1.7 Hz), 7.26 (dd, 1H, 10-HC, 3J9–10 7.6 Hz, 4J8–10 1.7 Hz).Found (%): C, 70.03; H, 7.51; N, 6.53. Calc. for C12H15NO2 (%): C, 70.22; H, 7.37; N, 6.82. 3b: yield 65%, bp 170–172 °C (3 torr), mp 43–44 °C. 1HNMR, d: 1.22 (s, 3H, Me), 1.27 (s, 3H, Me), 3.66 (d, 1H, OCHAHB, 2JAB 7.9 Hz), 3.71 (d, 1H, OCHAHB, 2JAB 7.9 Hz), 4.87 (d, 1H, NCHAHB, 2JAB 11.3 Hz), 4.91 (d, 1H, NCHAHB, 2JAB 11.3 Hz), 5.83 (s, 1H, OCHN), 6.64 (d, 1H, 7-HC, 3J7–8 8.8 Hz), 7.25 (dd, 1H, 8-HC, 4J8–10 2.4 Hz, 3J7–8 8.8 Hz), 7.37 (d, 1H, 10-HC, 4J8–10 2.4 Hz). Found (%): C, 50.94; H, 5.06, Br, 28.03; N, 4.77. Calc.for C12H14BrNO2 (%): C, 50.72; H, 4.97; Br, 28.12; N, 4.93. 3c: yield 67%, bp 120–123 °C (4 torr), nD 20 1.5465, d4 20 1.1335. 1H NMR, d: (major) 1.28 (d, 3H, Me, 3J 6.1 Hz), 2.88 (dd, 1H, NCHAHBCHX, 2JAB 9.6 Hz, 3JAX 6.8 Hz), 3.40 (dd, 1H, NCHAHBCHX, 2JAB 9.6 Hz, 3JBX 7.0 Hz), 4.11 (m, 1H, NCHAHBCHXMe), 4.74 (d, 1H, NCHAHBO, 2JAB 9.7 Hz), 4.82 (d, 1H, NCHAHBO, 2JAB 9.7 Hz), 5.75 (s, 1H, OCHN), 6.82 (m, 1H, 7-HC), 6.95 (m, 1H, 9-HC), 7.18 (m, 1H, 8-HC), 7.27 (m, 1H, 10-HC); (minor): 1.22 (d, 3H, Me, 3J 6.1 Hz), 2.80 (t, 1H, NCHAHBCHX, 2JAB 8.9 Hz, 3JAX 8.9 Hz), 3.32 (dd, 1H, NCHAHBCHX, 2JAB 8.9 Hz, 3JBX 5.9 Hz), 4.28 (m, 1H, NCHAHBCHXMe), 4.71 (d, 1H, NCHAHBO, 2JAB 9.6 Hz), 4.76 (d, 1H, NCHAHBO, 2JAB 9.6 Hz), 5.61 (s, 1H, OCHN), 6.82 (m, 1H, 7-HC), 6.95 (m, 1H, 9-HC), 7.18 (m, 1H, 8-HC), 7.27 (m, 1H, 10-HC).Found (%): C, 69.17; H, 6.97; N, 7.11. Calc. for C11H13NO2 (%): C, 69.09; H, 6.85; N, 7.32. O NH R1 R1 R 1a R = H, R1 = Me CHO OH R2 1b R = Me, R1 = H 2a R2 = H 2b R2 = Br O R2 N O R1 R1 R 3a R = R2 = H, R1 = Me 3b R = H, R1 = Me, R2 = Br 3c R = Me, R1 = R2 = H 3d R = Me, R1 = H, R2 = Br Scheme 1 – H2O 1a + 2a OH HO N O O HO N OH – H2O 3a 5 4 Scheme 2 Received: 27th April 2001; Com. 01/1793 3d: yield 57%, bp 158–160 °C (3 torr), mp 61–62 °C. 1H NMR, d: (major): 1.27 (d, 3H, Me, 3J 6.2 Hz), 2.85 (dd, 1H, NCHAHBCHX, 2JAB 9.5 Hz, 3JAX 6.5 Hz), 3.36 (dd, 1H, NCHAHBCHX, 2JAB 9.5 Hz, 3JBX 7.0 Hz), 4.09 (m, 1H, NCHAHBCHXMe), 4.76 (d, 1H, NCHAHBO, 2JAB 10.2 Hz), 4.81 (d, 1H, NCHAHBO, 2JAB 10.2 Hz), 5.72 (s, 1H, OCHN), 6.67 (d, 1H, 7-HC, 3J7–8 8.9 Hz), 7.24 (dd, 1H, 8-HC, 4J8–10 1.8 Hz, 3J7–8 8.9 Hz), 7.37 (d, 1H, 10-HC, 4J8–10 1.8 Hz); (minor): 1.20 (d, 3H, Me, 3J 6.2 Hz), 2.79 (dd, 1H, NCHAHBCHX, 2JAB 8.9 Hz, 3JAX 8.5 Hz), 3.29 (dd, 1H, NCHAHBCHX, 2JAB 8.9 Hz, 3JBX 5.6 Hz), 4.29 (m, 1H, NCHAHBCHXMe), 4.73 (d, 1H, NCHAHBO, 2JAB 9.9 Hz), 4.75 (d, 1H, NCHAHBO, 2JAB 9.9 Hz), 5.61 (s, 1H, OCHN), 6.70 (d, 1H, 7-HC, 3J7–8 8.9 Hz), 7.24 (dd, 1H, 8-HC, 4J8–10 1.8 Hz, 3J7–8 8.9 Hz), 7.38 (d, 1H, 10-HC, 4J8–10 1.8 Hz). Found (%): C, 48.74; H, 4.59; Br, 29.17; N, 5.02. Calc. for C11H12BrNO2 (%): C, 48.91; H, 4.48; Br, 29.58; N, 5.19.
ISSN:0959-9436
出版商:RSC
年代:2001
数据来源: RSC
|
|